A First Course in Spectral Theory [1 ed.] 2022028354, 9781470466565, 9781470471927, 9781470471910

The central topic of this book is the spectral theory of bounded and unbounded self-adjoint operators on Hilbert spaces.

229 118 8MB

English Pages 472 [492] Year 2022

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Contents
Preface
Chapter 1. Measure theory
1.1. šœŽ-algebras and monotone classes
1.2. Measures and CarathĆ©odoryā€™s theorem
1.3. Borel šœŽ-algebra on the real line and related spaces
1.4. Lebesgue integration
1.5. Lebesgueā€“Stieltjes measures on ā„
1.6. Product measures
1.7. Functions on šœŽ-locally compact spaces
1.8. Regularity of measures
1.9. The Rieszā€“Markov theorem
1.10. Exercises
Chapter 2. Banach spaces
2.1. Norms and Banach spaces
2.2. The Banach space š¶(š¾)
2.3. šæ^{š‘} spaces
2.4. Bounded linear operators and uniform boundedness
2.5. Weak-* convergence and the separable Banachā€“Alaoglu theorem
2.6. Banach-space valued integration
2.7. Banach-space valued analytic functions
2.8. Exercises
Chapter 3. Hilbert spaces
3.1. Inner products
3.2. Subspaces and orthogonal projections
3.3. Direct sums of Hilbert spaces
3.4. Orthonormal sets and orthonormal bases
3.5. Weak convergence
3.6. Tensor products of Hilbert spaces
3.7. Exercises
Chapter 4. Bounded linear operators
4.1. The š¶*-algebra of bounded linear operators on ā„‹
4.2. Strong and weak operator convergence
4.3. Invertibility, spectrum, and resolvents
4.4. Polynomials of operators
4.5. Invariant subspaces and direct sums of operators
4.6. Compact operators
4.7. Exercises
Chapter 5. Bounded self-adjoint operators
5.1. A first look at self-adjoint operators
5.2. Spectral theorem for compact self-adjoint operators
5.3. Spectral measures
5.4. Spectral theorem on a cyclic subspace
5.5. Multiplication operators
5.6. Spectral theorem on the entire Hilbert space
5.7. Borel functional calculus
5.8. Spectral theorem for unitary operators
5.9. Exercises
Chapter 6. Measure decompositions
6.1. Pure point and continuous measures
6.2. Singular and absolutely continuous measures
6.3. Hausdorff measures on ā„
6.4. Matrix-valued measures
6.5. Exercises
Chapter 7. Herglotz functions
7.1. Mƶbius transformations
7.2. Schur functions and convergence
7.3. CarathƩodory functions
7.4. The Herglotz representation
7.5. Growth at infinity and tail of the measure
7.6. Half-plane Poisson kernel and Stieltjes inversion
7.7. Pointwise boundary values
7.8. Meromorphic Herglotz functions
7.9. Exponential Herglotz representation
7.10. The PhragmĆ©nā€“Lindelƶf method and asymptotic expansions
7.11. Matrix-valued Herglotz functions
7.12. Weyl matrices and Dirichlet decoupling
7.13. Exercises
Chapter 8. Unbounded self-adjoint operators
8.1. Graphs and adjoints
8.2. Resolvents and self-adjointness
8.3. Unbounded multiplication operators and direct sums
8.4. Spectral measures and the spectral theorem
8.5. Borel functional calculus
8.6. Absolutely continuous functions and derivatives on intervals
8.7. Self-adjoint extensions and symplectic forms
8.8. Exercises
Chapter 9. Consequences of the spectral theorem
9.1. Maximal spectral measure
9.2. Spectral projections
9.3. Spectral type and spectral decompositions
9.4. Ruelleā€“Amreinā€“Georgescuā€“Enss (RAGE) theorem
9.5. Essential and discrete spectrum; the min-max principle
9.6. Spectral multiplicity
9.7. Stoneā€™s theorem
9.8. Fourier transform on ā„
9.9. Abstract eigenfunction expansions
9.10. Exercises
Chapter 10. Jacobi matrices
10.1. The canonical spectral measure and Favardā€™s theorem
10.2. Unbounded Jacobi matrices
10.3. Weyl solutions and š‘š-functions
10.4. Transfer matrices and Weyl disks
10.5. Full-line Jacobi matrices
10.6. Eigenfunction expansion for full-line Jacobi matrices
10.7. The Weyl š‘€-matrix
10.8. Subordinacy theory
10.9. A Combesā€“Thomas estimate and Schnolā€™s theorem
10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map
10.11. Direct spectral theory of periodic Jacobi matrices
10.12. Exercises
Chapter 11. One-dimensional Schrƶdinger operators
11.1. An initial value problem
11.2. Fundamental solutions and transfer matrices
11.3. Schrƶdinger operators with two regular endpoints
11.4. Endpoint behavior
11.5. Self-adjointness and separated boundary conditions
11.6. Weyl solutions and Greenā€™s functions
11.7. Weyl solutions and š‘š-functions
11.8. The half-line eigenfunction expansion
11.9. Weyl disks and applications
11.10. Asymptotic behavior of š‘š-functions
11.11. The local Borgā€“Marchenko theorem
11.12. Full-line eigenfunction expansions
11.13. Subordinacy theory
11.14. Potentials bounded below in an šæĀ¹_{}š‘™š‘œš‘ sense
11.15. A Combesā€“Thomas estimate and Schnolā€™s theorem
11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map
11.17. Direct spectral theory of periodic Schrƶdinger operators
11.18. Exercises
Bibliography
Notation Index
Index
Recommend Papers

A First Course in Spectral Theory [1 ed.]
 2022028354, 9781470466565, 9781470471927, 9781470471910

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

GRADUATE STUDIES I N M AT H E M AT I C S

226

A First Course in Spectral Theory Milivoje LukicĀ“

A First Course in Spectral Theory

GRADUATE STUDIES I N M AT H E M AT I C S

226

A First Course in Spectral Theory Milivoje LukicĀ“

EDITORIAL COMMITTEE Matthew Baker Marco Gualtieri Gigliola Staļ¬ƒlani (Chair) Jeļ¬€ A. Viaclovsky Rachel Ward 2020 Mathematics Subject Classiļ¬cation. Primary 47B15, 47B25, 47B02, 47B36, 34L40, 36C05.

For additional information and updates on this book, visit www.ams.org/bookpages/gsm-226

Library of Congress Cataloging-in-Publication Data Names: LukiĀ“c, Milivoje, 1984ā€“ author. Title: A ļ¬rst course in spectral theory / Milivoje LukiĀ“ c. Description: Providence, Rhode Island : American Mathematical Society, [2022] | Series: Graduate studies in mathematics, 1065-7339 ; 226 | Includes bibliographical references and index. Identiļ¬ers: LCCN 2022028354 | ISBN 9781470466565 (hardcover) | ISBN 9781470471927 (paperback) | ISBN 9781470471910 (ebook) Subjects: LCSH: Spectral theory (Mathematics)ā€“Textbooks. | AMS: Operator theory ā€“ Special classes of linear operators ā€“ Hermitian and normal operators (spectral measures, functional calculus, etc.). | Operator theory ā€“ Special classes of linear operators ā€“ Symmetric and selfadjoint operators (unbounded). | Operator theory ā€“ Special classes of linear operators ā€“ Jacobi (tridiagonal) operators (matrices) and generalizations. | Ordinary diļ¬€erential equations ā€“ Ordinary diļ¬€erential operators ā€“ Particular operators (Dirac, one-dimensional Schrodinger, ĀØ etc.). | Partial diļ¬€erential equations ā€“ Elliptic equations and systems ā€“ SchrĀØ odinger operator. | Functional analysis ā€“ Inner product spaces and their generalizations, Hilbert spaces ā€“ Hilbert and pre-Hilbert spaces: geometry and topology (including spaces with semideļ¬nite inner product). Classiļ¬cation: LCC QC20.7.S64 L85 2022 | DDC 515/.7222ā€“dc23/eng20221013 LC record available at https://lccn.loc.gov/2022028354

Copying and reprinting. Individual readers of this publication, and nonproļ¬t libraries acting for them, are permitted to make fair use of the material, such as to copy select pages for use in teaching or research. Permission is granted to quote brief passages from this publication in reviews, provided the customary acknowledgment of the source is given. Republication, systematic copying, or multiple reproduction of any material in this publication is permitted only under license from the American Mathematical Society. Requests for permission to reuse portions of AMS publication content are handled by the Copyright Clearance Center. For more information, please visit www.ams.org/publications/pubpermissions. Send requests for translation rights and licensed reprints to [email protected].  c 2022 by the author. All rights reserved. The American Mathematical Society retains all rights except those granted to the United States Government. Printed in the United States of America. āˆž The paper used in this book is acid-free and falls within the guidelines 

established to ensure permanence and durability. Visit the AMS home page at https://www.ams.org/ 10 9 8 7 6 5 4 3 2 1

28 27 26 25 24 23

To my teachers and my students

Contents

Preface

xiii

Chapter 1. Measure theory

1

Ā§1.1. Ļƒ-algebras and monotone classes

1

Ā§1.2. Measures and CarathĀ“eodoryā€™s theorem

6

Ā§1.3. Borel Ļƒ-algebra on the real line and related spaces

10

Ā§1.4. Lebesgue integration

15

Ā§1.5. Lebesgueā€“Stieltjes measures on R

24

Ā§1.6. Product measures

30

Ā§1.7. Functions on Ļƒ-locally compact spaces

32

Ā§1.8. Regularity of measures

35

Ā§1.9. The Rieszā€“Markov theorem

38

Ā§1.10. Exercises

41

Chapter 2. Banach spaces

45

Ā§2.1. Norms and Banach spaces

45

Ā§2.2. The Banach space C(K)

48

Ā§2.3.

54

Lp

spaces

Ā§2.4. Bounded linear operators and uniform boundedness

59

Ā§2.5. Weak-āˆ— convergence and the separable Banachā€“Alaoglu theorem

65

Ā§2.6. Banach-space valued integration

68

Ā§2.7. Banach-space valued analytic functions

71

Ā§2.8. Exercises

74 vii

viii

Contents

Chapter Ā§3.1. Ā§3.2. Ā§3.3. Ā§3.4. Ā§3.5. Ā§3.6. Ā§3.7.

3. Hilbert spaces Inner products Subspaces and orthogonal projections Direct sums of Hilbert spaces Orthonormal sets and orthonormal bases Weak convergence Tensor products of Hilbert spaces Exercises

77 77 82 88 91 97 100 104

Chapter Ā§4.1. Ā§4.2. Ā§4.3. Ā§4.4. Ā§4.5. Ā§4.6. Ā§4.7.

4. Bounded linear operators The C āˆ— -algebra of bounded linear operators on H Strong and weak operator convergence Invertibility, spectrum, and resolvents Polynomials of operators Invariant subspaces and direct sums of operators Compact operators Exercises

107 107 110 113 118 119 122 125

Chapter Ā§5.1. Ā§5.2. Ā§5.3. Ā§5.4. Ā§5.5. Ā§5.6. Ā§5.7. Ā§5.8. Ā§5.9.

5. Bounded self-adjoint operators A ļ¬rst look at self-adjoint operators Spectral theorem for compact self-adjoint operators Spectral measures Spectral theorem on a cyclic subspace Multiplication operators Spectral theorem on the entire Hilbert space Borel functional calculus Spectral theorem for unitary operators Exercises

129 130 136 139 141 143 146 149 153 155

Chapter Ā§6.1. Ā§6.2. Ā§6.3. Ā§6.4. Ā§6.5.

6. Measure decompositions Pure point and continuous measures Singular and absolutely continuous measures Hausdorļ¬€ measures on R Matrix-valued measures Exercises

159 160 162 169 176 178

Chapter 7. Herglotz functions Ā§7.1. MĀØobius transformations

183 184

Contents

ix

Ā§7.2. Schur functions and convergence

188

Ā§7.3. CarathĀ“eodory functions

190

Ā§7.4. The Herglotz representation

193

Ā§7.5. Growth at inļ¬nity and tail of the measure

196

Ā§7.6. Half-plane Poisson kernel and Stieltjes inversion

199

Ā§7.7. Pointwise boundary values

204

Ā§7.8. Meromorphic Herglotz functions

210

Ā§7.9. Exponential Herglotz representation

212

Ā§7.10. The PhragmĀ“enā€“LindelĀØ of method and asymptotic expansions 215 Ā§7.11. Matrix-valued Herglotz functions

216

Ā§7.12. Weyl matrices and Dirichlet decoupling

219

Ā§7.13. Exercises

222

Chapter 8. Unbounded self-adjoint operators

227

Ā§8.1. Graphs and adjoints

228

Ā§8.2. Resolvents and self-adjointness

231

Ā§8.3. Unbounded multiplication operators and direct sums

236

Ā§8.4. Spectral measures and the spectral theorem

238

Ā§8.5. Borel functional calculus

243

Ā§8.6. Absolutely continuous functions and derivatives on intervals 247 Ā§8.7. Self-adjoint extensions and symplectic forms

253

Ā§8.8. Exercises

262

Chapter 9. Consequences of the spectral theorem

267

Ā§9.1. Maximal spectral measure

268

Ā§9.2. Spectral projections

270

Ā§9.3. Spectral type and spectral decompositions

272

Ā§9.4. Ruelleā€“Amreinā€“Georgescuā€“Enss (RAGE) theorem

275

Ā§9.5. Essential and discrete spectrum; the min-max principle

278

Ā§9.6. Spectral multiplicity

283

Ā§9.7. Stoneā€™s theorem

289

Ā§9.8. Fourier transform on R

290

Ā§9.9. Abstract eigenfunction expansions

293

Ā§9.10. Exercises

296

x

Contents

Chapter 10. Jacobi matrices

299

Ā§10.1. The canonical spectral measure and Favardā€™s theorem

300

Ā§10.2. Unbounded Jacobi matrices

305

Ā§10.3. Weyl solutions and m-functions

309

Ā§10.4. Transfer matrices and Weyl disks

313

Ā§10.5. Full-line Jacobi matrices

319

Ā§10.6. Eigenfunction expansion for full-line Jacobi matrices

322

Ā§10.7. The Weyl M -matrix

325

Ā§10.8. Subordinacy theory

328

Ā§10.9. A Combesā€“Thomas estimate and Schnolā€™s theorem

334

Ā§10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

336

Ā§10.11. Direct spectral theory of periodic Jacobi matrices

347

Ā§10.12. Exercises

352

Chapter 11. One-dimensional SchrĀØodinger operators

359

Ā§11.1. An initial value problem

361

Ā§11.2. Fundamental solutions and transfer matrices

367

Ā§11.3. SchrĀØodinger operators with two regular endpoints

373

Ā§11.4. Endpoint behavior

379

Ā§11.5. Self-adjointness and separated boundary conditions

386

Ā§11.6. Weyl solutions and Greenā€™s functions

390

Ā§11.7. Weyl solutions and m-functions

394

Ā§11.8. The half-line eigenfunction expansion

399

Ā§11.9. Weyl disks and applications

407

Ā§11.10. Asymptotic behavior of m-functions

415

Ā§11.11. The local Borgā€“Marchenko theorem

423

Ā§11.12. Full-line eigenfunction expansions

425

Ā§11.13. Subordinacy theory

429

Ā§11.14. Potentials bounded below in an

L1loc

sense

433

Ā§11.15. A Combesā€“Thomas estimate and Schnolā€™s theorem

439

Ā§11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map

443

Ā§11.17. Direct spectral theory of periodic SchrĀØ odinger operators

450

Ā§11.18. Exercises

453

Contents

xi

Bibliography

459

Notation Index

467

Index

469

Preface

Spectral theory can be viewed as a generalization of linear algebra with a focus on linear operators on inļ¬nite-dimensional spaces. However, it is a branch of mathematical analysis that has its roots in the Fourier decomposition of a periodic function into sines and cosines. Those sines and cosines are solutions of the boundary value problem āˆ’f  = Ī»f , f (0) = f (2Ļ€), f  (0) = f  (2Ļ€). In modern language, they are eigenvectors of a diļ¬€erential operator (second derivative on an interval with periodic boundary conditions), acting on a suitable space of functions (which is an inļ¬nitedimensional vector space). Modern spectral theory studies classes of recurrence and diļ¬€erential operators which are motivated by mathematical physics, orthogonal polynomials, partial diļ¬€erential equations, and integrable systems. This text is intended as a ļ¬rst course in spectral theory, with a focus on the general theory of self-adjoint operators on separable Hilbert spaces and the direct spectral theory of Jacobi matrices and one-dimensional SchrĀØ odinger operators. It has been written as a textbook for three adjacent purposes: (a) an undergraduate course on bounded self-adjoint operators, (b) a ļ¬rst course for graduate students interested in the spectral theory of bounded and unbounded self-adjoint operators, (c) a topics course on continuum one-dimensional SchrĀØodinger operators. The intended audience for this text includes beginning graduate students and advanced undergraduates, so the text is written with minimal prerequisites. It is assumed that the reader knows linear algebra and basic analysis, xiii

xiv

Preface

including basic complex analysis. In an eļ¬€ort to keep the text accessible, we avoid unnecessary abstractions and get by without topology. Measure theory is not assumed as a prerequisite; the required background in measure theory is developed in Chapter 1, including some specialized results needed for our purposes (e.g., a criterion for a subalgebra of bounded Borel functions to be the entire algebra, used for the proof of uniqueness of the Borel functional calculus for self-adjoint operators). Chapter 2 introduces Banach spaces; these are vector spaces equipped with a norm (a suitable notion of length of vectors) which are complete. This chapter is a nonstandard introduction to functional analysis shaped by a spectral theoristā€™s needs: it includes a discussion of Banach space valued integrals, Banach space valued analytic functions, and important examples of Banach spaces, without going deep into abstract Banach space theory. Chapter 3 introduces Hilbert spaces, which are a special case of Banach space equipped with an inner product (an abstract version of a dot product). The chapter includes inļ¬nite direct sums of Hilbert spaces (needed for the multiplication operator form of the spectral theorem) and tensor products. Chapter 4 describes the general structure and properties of bounded linear operators on Hilbert spaces; this provides the basic language for the remainder of the text. Chapter 5 begins the study of bounded self-adjoint operators. Selfadjoint operators can be viewed as an inļ¬nite-dimensional generalization of Hermitian matrices, and this chapter can be viewed as a generalization of diagonalizability of Hermitian matrices. Spectral measures are introduced and two central results are proved; namely, the spectral theorem and the Borel functional calculus. The spectral theorem is presented in multiplication operator form, which we ļ¬nd more useful and intuitive (we introduce and use spectral projections later in this text, but we do not use integration with respect to projection-valued measures or the historically more common approach via resolution of the identity). The Borel functional calculus is constructed using the spectral theorem. Chapter 6 presents several measure decompositions (continuous/pure point, absolutely continuous/singular, and decompositions with respect to Hausdorļ¬€ measures) and pointwise descriptions of these decompositions. This is part of the standard vocabulary of spectral theory, where continuity properties of spectral measures are of great importance. One of the goals of this text is to present the spectral theory of selfadjoint operators from the ground up as a correspondence of three main objects: self-adjoint operators, their spectral measures (which are measures on R), and Herglotz functions (which are complex-analytic functions from

Preface

xv

the upper half-plane to itself). Accordingly, Herglotz functions are introduced in Chapter 7. Through an integral representation, they are related to measures on R, and this chapter studies this correspondence. This may seem like a detour from spectral theory, but the truth is quite the opposite: although Chapter 7 doesnā€™t mention operators, we will see that it contains the hard parts of proofs of important spectral theoretic results. In Chapter 8, we study unbounded self-adjoint operators, culminating in their spectral theorem and Borel functional calculus. The presentation is independent from the bounded case, although the bounded case serves as a strong motivation. Many techniques from the bounded case have suitable analogues or restatements in the unbounded case, but there are technical complications. This chapter includes the study of symplectic forms over the complex ļ¬eld of scalars and a description of self-adjoint extensions of a symmetric operator. Chapter 9 can be read as a continuation of Chapter 5 or of Chapter 8. It describes many general consequences of the spectral theorem and the Borel functional calculus, such as spectral type, spectral multiplicity, etc., which are part of the basic language of spectral theory. It contains a study of Stoneā€™s theorem and its applications to constructing diagonalizations of diļ¬€erential operators; for instance, we provide a self-contained introduction to the Fourier transform on L2 (R) through the problem of diagonalizing the d viewed as an unbounded self-adjoint operator on R. This derivative āˆ’i dx approach is constructive and based on Stoneā€™s theorem, and it serves as a warm-up for eigenfunction expansions of SchrĀØodinger operators. Chapter 10 discusses bounded and unbounded Jacobi matrices, which are a well-studied class of self-adjoint operators corresponding to a secondorder recurrence relation on 2 (N) and 2 (Z). While they can be viewed as an extended example for general spectral theory, their connections to orthogonal polynomials and mathematical physics make them a classical subject of their own; we present some of their general properties and techniques for their study. We emphasize the correspondence with Weyl m-functions and use Weyl disks as a robust way of deriving approximation results, such as Carmonaā€™s theorem. The chapter includes subordinacy theory, eigenfunction expansions for full-line Jacobi matrices, and the Weyl M -matrix approach. Finally, we present the direct spectral theory of periodic Jacobi matrices, using the Marchenkoā€“Ostrovski map as a central object. 2

d Chapter 11 is dedicated to one-dimensional SchrĀØ odinger operators āˆ’ dx 2+ V , considered on a ļ¬nite or inļ¬nite interval, with locally integrable potentials V . The chapter starts with self-adjointness and the limit point-limit circle alternative. Eigenfunction expansions are introduced constructively, using Stoneā€™s theorem. Weyl disks are used to derive various approximation

xvi

Preface

results, including Carmonaā€™s formula and continuity of m-functions under L1loc perturbations of the potential. We also prove the local Borgā€“Marchenko theorem, asymptotic behavior of the m-functions, and Schnolā€™s theorem. We conclude this chapter with the direct spectral theory of periodic SchrĀØ odinger operators studied via the Marchenkoā€“Ostrovski map. The book can of course be read cover to cover, but various selections of the material are possible. For instance, beyond the introductory chapters, we suggest the following. ā€¢ A course on bounded self-adjoint operators can contain Chapters 4 and 5 and Section 10.1. It can continue, time permitting, with Sections 6.1ā€“6.2 and Sections 9.1ā€“9.6. ā€¢ A course on unbounded self-adjoint operators can contain Chapter 4, Sections 7.1ā€“7.5, Chapter 8, and a selection of topics from Chapters 9, 10, 11. ā€¢ A course on Jacobi or SchrĀØ odinger operators can be based on the corresponding Chapter 10 or 11. It requires Chapter 5 or Chapter 8 as a prerequisite; it is also heavily reliant on Chapters 6, 7, 9, which can be studied in preparation or in parallel with Chapter 10 or 11. Many analytical tools are developed in Chapters 6 and 7, applied in Chapter 9 to self-adjoint operators, then reļ¬ned in more specialized settings in Chapters 10 and 11. They can be studied by taking cross-sections of diļ¬€erent chapters. I would like to thank Ilia Binder, David Damanik, Ana Djurdjevac, Benjamin Eichinger, Jake Fillman, Fritz Gesztesy, Manuela Girotti, Michael Goldstein, Ethan Gwaltney, Svetlana Jitomirskaya, Ilya Marchenko, Shaan Nagy, Maria Ntekoume, Barry Simon, Selim Sukhtaiev, Chunyi Wang, Xingya Wang, Ronen Wdowinski, Bohan Wu, Chengcheng Yang, Giorgio Young, Peter Yuditskii, and Maxim Zinchenko for helpful discussions and valuable feedback which improved this book.

Chapter 1

Measure theory

The subject of this chapter is the Lebesgue theory of measures and integration. This is one of the foundations of modern analysis; compared to Riemann integration, it includes a much wider class of functions which can be integrated and has better behavior with respect to limits. It is a classical idea to measure the size of a set by a positive number. A notion of size, such as the number of elements, length, or area, is intuitively expected to be additive for disjoint sets. A key idea in Lebesgue theory is that additivity should also hold for countable disjoint families, i.e.,  āˆž āˆž   An = Ī¼(An ) Ī¼ n=1

n=1

if An āˆ© Ak = āˆ… whenever n = k. This stronger property, called Ļƒ-additivity, will be part of the deļ¬nition of a measure; it leads to good behavior of measures and integrals with respect to limits of sequences. Another fundamental question is which sets should be measured; this is captured by the notion of a Ļƒ-algebra. We will quickly specialize to the setting of Borel sets and Borel functions on metric spaces. This class is large enough to contain the sets which occur in our work, while avoiding some foundational paradoxes and topological distractions. Our choice of topics is shaped by the goals of this text; many other texts on measure theory are available [32, 81, 97, 99].

1.1. Ļƒ-algebras and monotone classes Let X be a nonempty set. Our ļ¬rst deļ¬nition describes classes of subsets of X which are closed under certain set operations. We denote by P(X) 1

2

1. Measure theory

the set of all subsets of X, and we denote complements of subsets of X by Ac = X \ A when there is no risk of confusion. Deļ¬nition 1.1. A Ļƒ-algebra on a set X is a family A āŠ‚ P(X) that obeys (a) āˆ… āˆˆ A; (b) A āˆˆ A implies Ac āˆˆ A; (c) for any sequence (Aj )āˆž j=1 such that Aj āˆˆ A for all j,

āˆž

j=1 Aj

āˆˆ A.

Some authors replace (a) by the condition X āˆˆ A; by (b), this is equivalent to our deļ¬nition, since āˆ… = X c and X = āˆ…c . The deļ¬nition has some easy By (b) and (c), if Aj āˆˆ A c  consequences. āˆž āˆž c āˆˆ A. Of course, (c) also holds for for all j āˆˆ N, then j=1 Aj = j=1 Aj ļ¬nite unions, since we can take some of the Aj to be āˆ…. Thus, A1 , A2 āˆˆ A implies A1 āˆŖA2 āˆˆ A and A1 āˆ©A2 āˆˆ A. It also implies A1 \A2 = A1 āˆ©Ac2 āˆˆ A. Informally speaking, A is closed under ļ¬nite and countable set operations. Example 1.2. For any set X, A = P(X) is a Ļƒ-algebra on X. Example 1.3. For any set X, A = {āˆ…, X} is a Ļƒ-algebra on X. It is common in mathematics to obtain a minimal set with some property by showing that there exists a set with the property, and that intersections of sets with the property also have the property. For example, the closure of a set B in a metric space X can be deļ¬ned as the intersection of all closed sets in X that contain B, because an arbitrary intersection of closed sets is closed, and the whole space is closed. We are about to make an analogous construction for Ļƒ-algebras. It is important that the following result holds for the intersection of an arbitrary (not only countable) collection of Ļƒ-algebras. Lemma 1.4. The intersection of an arbitrary nonempty collection of Ļƒ-algebras on X is a Ļƒ-algebra on X.  Proof. Let AĪ³ , Ī³ āˆˆ Ī“, be Ļƒ-algebras on X, and let A = Ī³āˆˆĪ“ AĪ³ . Since āˆ… āˆˆ AĪ³ for all Ī³, āˆ… āˆˆ A. If A āˆˆ A, then A āˆˆ AĪ³ for all Ī³, so Ac āˆˆ AĪ³ for c all j āˆˆ N, then Aj āˆˆ AĪ³ for all j and all Ī³, so āˆžĪ³, so A āˆˆ A. If Aj āˆˆ Afor āˆž A āˆˆ A for all Ī³, so  j Ī³ j=1 j=1 Aj āˆˆ A. Deļ¬nition 1.5. Let F āŠ‚ P(X). The Ļƒ-algebra generated by F is the intersection of all Ļƒ-algebras on X that contain F . Since P(X) is a Ļƒ-algebra on X which contains F , the family of Ļƒalgebras which contain F is not empty, so the intersection of this family is well deļ¬ned. This intersection is a Ļƒ-algebra by Lemma 1.4, and it is the smallest Ļƒ-algebra that contains F .

1.1. Ļƒ-algebras and monotone classes

3

In a metric space X, the topology generated by the metric d is the family TX = {A āŠ‚ X | A is open with respect to the metric d}. Not every topology is generated by a metric: topological spaces are a generalization of metric spaces and are studied in their own right. In this text, we only use topologies generated by a metric (so-called metric topologies), even though some of the theory below can be stated more generally. Deļ¬nition 1.6. Let X be a metric space. The Borel Ļƒ-algebra on X, denoted BX , is the Ļƒ-algebra generated by TX . Elements of the Borel Ļƒ-algebra are called Borel sets. Example 1.7. A metric space X is said to be discrete if every subset of X is open. One example of a discrete metric on any set X is

0 x=y d(x, y) = 1 x = y. If X is a discrete metric space, then TX = P(X), so BX = P(X). Deļ¬nition 1.8. For spaces X, Y with Ļƒ-algebras AX , AY , a function f : X ā†’ Y is called measurable if and only if B āˆˆ AY implies f āˆ’1 (B) āˆˆ AX . In particular, if X, Y are metric spaces, f : X ā†’ Y is a Borel function if B āˆˆ BY implies f āˆ’1 (B) āˆˆ BX . Proposition 1.9. If f : X ā†’ Y and g : Y ā†’ Z are Borel functions, then so is their composition g ā—¦ f : X ā†’ Z. Proof. For any B āˆˆ BZ , since g is Borel, g āˆ’1 (B) āˆˆ BY . Since f is Borel,  (g ā—¦ f )āˆ’1 (B) = f āˆ’1 (g āˆ’1 (B)) āˆˆ BX . To prove that a function is Borel, we sometimes use the notion of pushforward of a Ļƒ-algebra. This is based on the fact that the inverse image f āˆ’1 commutes with set operations. Proposition 1.10. Let f : X ā†’ Y and let A be a Ļƒ-algebra on X. Then B = {B āŠ‚ Y | f āˆ’1 (B) āˆˆ A} is a Ļƒ-algebra on Y , called the pushforward of A by f . Proof. From f āˆ’1 (āˆ…) = āˆ… āˆˆ A, we conclude āˆ… āˆˆ B. If f āˆ’1 (B) āˆˆ A, so āˆ’1 f āˆ’1 (Y \ B) =X \ f āˆ’1 (B) āˆˆA and Y \ B āˆˆ B. If f (Bj ) āˆˆ A for some sets  āˆž āˆž āˆ’1 (B ) āˆˆ A.  Bj , then f āˆ’1 j j=1 Bj = j=1 f Borel sets and Borel functions are meant to be large enough classes to include all sets and functions which we will encounter in our work. The following lemma is a ļ¬rst step in that direction.

4

1. Measure theory

Lemma 1.11. Every continuous function is Borel. Proof. By Proposition 1.10, the set S = {B | f āˆ’1 (B) āˆˆ BX } is a Ļƒ-algebra on Y . If f : X ā†’ Y is continuous, and if B is open, then f āˆ’1 (B) is open, so f āˆ’1 (B) āˆˆ BX . Thus, the Ļƒ-algebra S contains all open sets in Y . Therefore,  it contains BY , so for any B āˆˆ BY , f āˆ’1 (B) āˆˆ BX . Example 1.12. For any A āˆˆ BX , the characteristic function of the set A

1 xāˆˆA Ļ‡A (x) = (1.1) 0 x āˆˆ Ac is a Borel function. Proof. For any set B, the inverse image Ļ‡āˆ’1 A (B) is equal to one of the sets  āˆ…, A, Ac , or X. All of these are Borel sets, so Ļ‡A is a Borel function. Ļƒ-algebras, and in particular Borel Ļƒ-algebras, behave naturally with respect to restrictions and inclusions (Exercise 1.1). To prove that some property holds for all elements of the Ļƒ-algebra A generated by G, we usually introduce the set S of elements of A with that property, prove G āŠ‚ S, and prove that S is a Ļƒ-algebra. However, that can be a diļ¬ƒcult task. As our last topic in this section, we show an abstract result which reduces that task to proving an easier conditionā€”that S is closed under increasing and decreasing countable limitsā€”as long as the set G has certain algebraic properties. We need the following deļ¬nitions. Deļ¬nition 1.13. An algebra on X is a family G āŠ‚ P(X) that obeys (a) āˆ… āˆˆ G; (b) A āˆˆ G implies Ac āˆˆ G; (c) A1 , A2 āˆˆ G implies A1 āˆ© A2 āˆˆ G. This has immediate further corollaries: Any algebra contains X = āˆ…c ; any algebra is closed under ļ¬nite intersections, unions, and diļ¬€erences of sets. Every Ļƒ-algebra is an algebra, but not conversely: Example 1.14. The family G = {A āŠ‚ R | A is ļ¬nite or Ac is ļ¬nite} is an algebra, but not a Ļƒ-algebra. Deļ¬nition 1.15. A monotone class on X is a family C āŠ‚ P(X) that obeys  (a) if An āˆˆ C and An āŠ‚ An+1 for all n āˆˆ N, then nāˆˆN An āˆˆ C;  (b) if Bn āˆˆ C and Bn+1 āŠ‚ Bn for all n āˆˆ N, then nāˆˆN Bn āˆˆ C.

1.1. Ļƒ-algebras and monotone classes

5

Every Ļƒ-algebra is a monotone class, but not conversely: Example 1.16. The family C = {āˆ…, R} āˆŖ {(a, āˆž) | a āˆˆ Z} is a monotone class, but not a Ļƒ-algebra. An arbitrary intersection of monotone classes is a monotone class, and P(X) is a monotone class. Thus, for any E āŠ‚ P(X), there exists a smallest monotone class which contains E, i.e., the monotone class generated by E. Theorem 1.17 (Monotone class theorem). If G āŠ‚ P(X) is an algebra, the monotone class generated by G is equal to the Ļƒ-algebra generated by G. Proof. Denote by C the monotone class generated by G. The main step is to prove that for all E, F āˆˆ C, E \ F, F \ E, E āˆ© F āˆˆ C.

(1.2)

CE = {F āˆˆ C | E \ F, F \ E, E āˆ© F āˆˆ C}.

(1.3)

Deļ¬ne for E āˆˆ C,

This is a monotone class, since the expressions E \ F, F \ E, E āˆ© F viewed as functions of F preserve monotonicity and monotone limits. For instance, āˆž āˆž Fn āŠ‚ Fn+1 implies E \ Fn+1 āŠ‚ E \ Fn and E \ ( n=1 Fn ) = n=1 (E \ Fn ). Assume E āˆˆ G. Then (1.2) holds for F āˆˆ G, since G is an algebra and G āŠ‚ C. Thus, G āŠ‚ CE . Thus, CE is a monotone class with G āŠ‚ CE āŠ‚ C, so CE = C. Thus, (1.2) holds for all E āˆˆ G and F āˆˆ C. The conditions in (1.3) are symmetric in E, F , so (1.2) holds for all E āˆˆ C and F āˆˆ G. Now the previous argument can be repeated for any E āˆˆ C and shows CE = C. Thus, (1.3) holds for all E, F āˆˆ C. Since X āˆˆ G āŠ‚ C, (1.2) implies that for all E āˆˆ C, X \ E āˆˆ C, and that for all E, F āˆˆ C, E āˆ© F āˆˆ C, so C is an algebra.  For any An āˆˆ C, n āˆˆ N, consider Bn = nj=1 Aj āˆˆ C. This is a monotone sequence: Bn āŠ‚ Bn+1 for all n āˆˆ N. Since C is a monotone class,   An = Bn āˆˆ C, nāˆˆN

nāˆˆN

so C is a Ļƒ-algebra. Denoting by A the Ļƒ-algebra generated by G, we conclude A āŠ‚ C. Conversely, A is a monotone class and G āŠ‚ A, so C āŠ‚ A.  The monotone class theorem will be used twice in this text: in the proof of a uniqueness result for Borel measures on R in Section 1.5, and in the study of product measures in Section 1.6.

6

1. Measure theory

1.2. Measures and CarathĀ“ eodoryā€™s theorem In this section, we deļ¬ne measures, study their general properties, and give an important method for constructing measures. Deļ¬nition 1.18. A measure on a Ļƒ-algebra A is a map Ī¼ : A ā†’ [0, āˆž] with Ī¼(āˆ…) = 0, which is Ļƒ-additive, i.e., for any pairwise disjoint sets An āˆˆ A, n āˆˆ N,  āˆž āˆž   Ī¼ An = Ī¼(An ). (1.4) n=1

n=1

If X is a metric space, a measure on BX is called a Borel measure on X. The measure Ī¼ is ļ¬nite if Ī¼(X) < āˆž. It is ļ¬nite on compacts if Ī¼(K) < āˆž for every compact K āŠ‚ X. In (1.4) we are using the natural convention c + āˆž = āˆž for c āˆˆ [0, āˆž]. Explicitly, if the series in (1.4) is divergent or if any of the terms in the series are inļ¬nite, the value of the series is taken to be +āˆž. Let us see some easy examples and general properties of measures: Example 1.19. Ī¼ ā‰” 0 is the trivial measure on any Ļƒ-algebra. Example 1.20. Fix x āˆˆ X. The Dirac measure at x is the measure Ī“x on P(X) deļ¬ned by

1 xāˆˆA Ī“x (A) = 0 xāˆˆ / A. Example 1.21. Let #A denote the number of elements of a set A (if A is inļ¬nite, we write #A = āˆž). For any set X, the counting measure on P(X) is deļ¬ned by Ī¼(A) = #A. Theorem 1.22. Let Ī¼ be a measure on A. Then, for any sets in A, the following hold. (a) If n āˆˆ N and sets A1 , . . . , An are pairwise disjoint, then   n n   Ī¼ Aj = Ī¼(Aj ). j=1

j=1

(b) If S āŠ‚ T , then Ī¼(S) ā‰¤ Ī¼(T ). (c) If S āŠ‚ T and Ī¼(S) < āˆž, then Ī¼(T \ S) = Ī¼(T ) āˆ’ Ī¼(S). (d) For any sequence of sets (Bn )āˆž n=1 such that Bn āŠ‚ Bn+1 for all n āˆˆ N,  āˆž  Bn = lim Ī¼(Bn ). Ī¼ n=1

nā†’āˆž

1.2. Measures and CarathĀ“eodoryā€™s theorem

7

(e) For any sequence of sets (Cn )āˆž n=1 such that Cn+1 āŠ‚ Cn for all n āˆˆ N, if there exists k āˆˆ N such that Ī¼(Ck ) < āˆž, then āˆž  Ī¼ Cn = lim Ī¼(Cn ). nā†’āˆž

n=1

(f) For any sequence of sets (An )āˆž n=1 ,  āˆž āˆž   An ā‰¤ Ī¼(An ). Ī¼ j=1

j=1

Proof. (a) This follows from Ļƒ-additivity with Aj = āˆ… for j > n. (b) This follows by representing T as the disjoint union of S and T \ S. (c) This follows from Ī¼(T ) = Ī¼(S) + Ī¼(T \ S) by subtracting Ī¼(S). (d) Denote An = Bn \ Bnāˆ’1 for  n ā‰„ 2 and A1 = B1 . The sets An are disjoint, so for each n, Bn = nj=1 Aj implies by (a) that Ī¼(Bn ) = āˆž āˆž n j=1 Ī¼(Aj ). Since j=1 Bj = j=1 Aj , Ļƒ-additivity implies āŽž āŽ› āŽž āŽ› āˆž āˆž āˆž n     āŽ  āŽ āŽ  āŽ Bj = Ī¼ Aj = Ī¼(Aj ) = lim Ī¼(Aj ) = lim Ī¼(Bn ). Ī¼ j=1

j=1

j=1

nā†’āˆž

nā†’āˆž

j=1

(e) Applying (d) to the increasing sequence of sets Ck \ Cn gives    āˆž āˆž  Ī¼ Ck \ Cn = Ī¼ (Ck \ Cn ) = lim Ī¼(Ck \ Cn ). n=1

nā†’āˆž

n=1

Subtracting both sides from Ī¼(Ck ) completes the proof.  (f) Consider the increasing sequence of sets Bn = nj=1 Aj , with B0 = āˆ…. āˆž The āˆž sets Cn = Bn \ Bnāˆ’1 are pairwise disjoint, Cn āŠ‚ An , and n=1 An = n=1 Cn . Thus, by Ļƒ-additivity and (b),  āˆž  āˆž āˆž āˆž     An = Ī¼ Cn = Ī¼(Cn ) ā‰¤ Ī¼(An ).  Ī¼ n=1

n=1

n=1

n=1

In this theorem, ļ¬niteness appears as an assumption whenever the proof uses subtraction, because we cannot subtract āˆž. This assumption cannot be removed: for instance, for part (e), if Ī¼ is the counting measure on N and  āˆˆ N | k ā‰„ n}, then An+1 āŠ‚ An and Ī¼(An ) = āˆž for all n, but An = {k Ī¼ nāˆˆN An = Ī¼(āˆ…) = 0. The importance of Ļƒ-additivity will be evident; however, when constructing measures, Ļƒ-additivity presents a challengeā€”constructing ļ¬nitely additive maps is much easier. We will now present a robust abstract way to

8

1. Measure theory

construct measures, which will be used several times in this text. The intermediate step will be an object called an outer measure, which has weaker properties than a measure, but it is deļ¬ned on all subsets of the space X. Deļ¬nition 1.23. An outer measure on X is a map Ī¼āˆ— : P(X) ā†’ [0, āˆž] such that (a) Ī¼āˆ— (āˆ…) = 0; (b) Ī¼āˆ— (A) ā‰¤ Ī¼āˆ— (B) if A āŠ‚ B;   āˆž āˆ— (c) (Ļƒ-subadditivity) Ī¼āˆ— āˆž n=1 An ā‰¤ n=1 Ī¼ (An ) for all sets An āŠ‚ X.  A cover of A is a family of sets {EĪ³ }Ī³āˆˆĪ“ such that A āŠ‚ Ī³āˆˆĪ“ EĪ³ . The cover is called ļ¬nite or countable if Ī“ is ļ¬nite or countable, respectively. To construct an outer measure, let us ļ¬rst choose a fairly arbitrary class of elementary sets E and a weight Ļ on elementary sets (we do not call Ļ a measure, because the elementary sets in general do not form a Ļƒ-algebra and because Ļ is not required to obey any kind of additivity properties), and then deļ¬ne Ī¼āˆ— (A) as an inļ¬mum over countable covers of A: Theorem 1.24. Let E āŠ‚ P(X) with āˆ… āˆˆ E and X āˆˆ E. Let Ļ : E ā†’ [0, āˆž] be a map with Ļ(āˆ…) = 0. Deļ¬ne, for all A āŠ‚ X,

āˆž  āˆž     āˆ— Ļ(Ej )  A āŠ‚ Ej , Ej āˆˆ E āˆ€j āˆˆ N . (1.5) Ī¼ (A) = inf j=1

Then

Ī¼āˆ—

j=1

is an outer measure on X.

Proof. Since X āˆˆ E, every set A has a countable cover, so the deļ¬nition is well posed. The property Ī¼āˆ— (āˆ…) = 0 follows by taking Ej = āˆ… for all j. If A āŠ‚ B, any cover of B is also a cover of A, so the inļ¬mum deļ¬ning Ī¼āˆ— (A) is over a larger set than the inļ¬mum deļ¬ning Ī¼āˆ— (B), so Ī¼āˆ— (A) ā‰¤ Ī¼āˆ— (B).  If A = āˆž for any > 0 and n, there exists a countable cover n=1 An , āˆž āˆ— n such that {En,j }āˆž j=1 j=1 Ļ(En,j ) ā‰¤ Ī¼ (An ) + /2 . Then the countable collection {En,j }āˆž n,j=1 is a cover for A and Ī¼āˆ— (A) ā‰¤

āˆž āˆž   n=1 j=1

Ļ(En,j ) ā‰¤

āˆž 

Ī¼āˆ— (An ) + .

n=1

Since > 0 is arbitrary, this implies that Ī¼āˆ— (A) ā‰¤

āˆž

n=1 Ī¼

āˆ— (A ). n



For any E āˆˆ E, by taking the countable cover E1 = E and Ej = āˆ… for j ā‰„ 2, we conclude Ī¼āˆ— (E) ā‰¤ Ļ(E). For some choices of weights Ļ, it can happen that Ī¼āˆ— (E) < Ļ(E) (Exercise 6.8). In several important constructions, we will show manually that Ī¼āˆ— (E) = Ļ(E) for E āˆˆ E.

1.2. Measures and CarathĀ“eodoryā€™s theorem

9

The core of the outer measure approach is CarathĀ“eodoryā€™s deļ¬nition of being ā€œmeasurable with respect to an outer measureā€: Deļ¬nition 1.25. The set A āŠ‚ X is measurable with respect to Ī¼āˆ— if Ī¼āˆ— (E) = Ī¼āˆ— (E āˆ© A) + Ī¼āˆ— (E āˆ© Ac )

āˆ€E āŠ‚ X.

(1.6)

Theorem 1.26 (CarathĀ“eodory). Let Ī¼āˆ— be an outer measure on X. The family A of sets measurable with respect to Ī¼āˆ— is a Ļƒ-algebra, and the restriction Ī¼āˆ— |A is a measure on A. Proof. Using Ī¼āˆ— (āˆ…) = 0, it easily follows that āˆ… āˆˆ A. Condition (1.6) is equivalent for A and Ac , so A āˆˆ A implies Ac āˆˆ A. Consider an increasing sequence Bn āŠ‚ Bn+1 āŠ‚ X, n āˆˆN, and its limit āˆž B = j=1 Bj . With the convention B0 = āˆ…, we note B = āˆž j=1 (Bj \ Bjāˆ’1 ) and conclude by Ļƒ-additivity of Ī¼āˆ— that, for any E āŠ‚ X, āˆž  āˆ— āˆ— c āˆ— āˆ— c Ī¼āˆ— (Eāˆ©(Bj \Bjāˆ’1 )). (1.7) Ī¼ (E) ā‰¤ Ī¼ (Eāˆ©B )+Ī¼ (Eāˆ©B) ā‰¤ Ī¼ (Eāˆ©B )+ j=1

Let us prove that these inequalities sometimes turn into equalities.  Fix E āŠ‚ X, let An āˆˆ A, n āˆˆ N, and take Bn = nj=1 Aj for n āˆˆ N. Measurability of Aj with respect to Ī¼āˆ— implies c c c ) = Ī¼āˆ— (E āˆ© Bjāˆ’1 āˆ© Acj ) + Ī¼āˆ— (E āˆ© Bjāˆ’1 āˆ© Aj ), Ī¼āˆ— (E āˆ© Bjāˆ’1

which we rewrite as c c ) = Ī¼āˆ— (E āˆ© Bjc ) + Ī¼āˆ— (E āˆ© (Bj \ Bjāˆ’1 )). Ī¼āˆ— (E āˆ© Bjāˆ’1

By induction in n, this gives Ī¼āˆ— (E) = Ī¼āˆ— (E āˆ© Bnc ) +

n 

c Ī¼āˆ— (E āˆ© (Bj \ Bjāˆ’1 )).

j=1

By monotonicity of the outer measure, Ī¼āˆ— (E āˆ© Bnc ) ā‰„ Ī¼āˆ— (E āˆ© B c ) so Ī¼āˆ— (E) ā‰„ Ī¼āˆ— (E āˆ© B c ) +

n 

c Ī¼āˆ— (E āˆ© (Bj \ Bjāˆ’1 ))

j=1

for any n. Taking n ā†’ āˆž, Ī¼āˆ— (E) ā‰„ Ī¼āˆ— (E āˆ© B c ) +

āˆž 

c Ī¼āˆ— (E āˆ© (Bj \ Bjāˆ’1 )).

(1.8)

j=1

This gives an inequality in the opposite direction compared to (1.7), so it implies that all three quantities are equal, āˆž  āˆ— āˆ— c āˆ— āˆ— c c Ī¼āˆ— (Eāˆ©(Bj \Bjāˆ’1 )). (1.9) Ī¼ (E) = Ī¼ (Eāˆ©B )+Ī¼ (Eāˆ©B) = Ī¼ (Eāˆ©B )+ j=1

10

1. Measure theory

Since E is arbitrary, the ļ¬rst equality in (1.9) shows that B āˆˆ A. Thus, A is closed under countable unions, so it is a Ļƒ-algebra. If the sets An are disjoint, then Bj \ Bjāˆ’1 = Aj , so the second equality in (1.9), taken for E = B, proves that āˆž  āˆž   Ī¼āˆ— Aj = Ī¼āˆ— (Aj ). j=1

Thus,

Ī¼āˆ—

j=1

is Ļƒ-additive on A.



1.3. Borel Ļƒ-algebra on the real line and related spaces We now specialize to Borel Ļƒ-algebras on some important spaces, starting with the real line R, with the goal of obtaining useful criteria which establish that certain sets and functions are Borel. We use a topological notion: Deļ¬nition 1.27. A base U of X is a family of open sets in X such that, for every open set V and every x āˆˆ V , there exists A āˆˆ U such that x āˆˆ A āŠ‚ V . The space is said to be second countable if it has a countable base. Lemma 1.28. If U is a base of X, every open set V in X can be written as a union of elements of U ,  A. V = AāˆˆU AāŠ‚V

Proof. Denote by U the union of all A āˆˆ U with A āŠ‚ V . Obviously U āŠ‚ V . For the converse, take any x āˆˆ V . By the deļ¬nition of a base, there exists A āˆˆ U with x āˆˆ A āŠ‚ V , so x āˆˆ U . This shows V āŠ‚ U .  Lemma 1.29. U = {(a, b) | a, b āˆˆ Q, a < b} is a countable base of R. Proof. Assume V āŠ‚ R is open and x āˆˆ V . There exists > 0 such that (x āˆ’ , x + ) āŠ‚ V . By density of Q in R, there exist rational numbers a āˆˆ (x āˆ’ , x), b āˆˆ (x, x + ). Then x āˆˆ (a, b) āŠ‚ V . Thus, U is a base of R. Its countability follows from countability of Q.  A metric space is called separable if it contains a countable dense subset. In concrete situations it is useful to write down an explicit base, but second countability of a metric space is equivalent to separability (Exercise 1.2). Note how countability of the base is used in the following proof: Lemma 1.30. The Borel Ļƒ-algebra on R is the Ļƒ-algebra generated by the intervals (a, āˆž) with a āˆˆ R. Proof. Denote by A the Ļƒ-algebra generated by the family {(a, āˆž) | a āˆˆ R}. The sets (a, āˆž) are open, thus they are Borel, so A āŠ‚ BR .

1.3. Borel Ļƒ-algebra on the real line and related spaces

11

By takingcomplements, (āˆ’āˆž, a] = R \ (a, āˆž) āˆˆ A for any a. For any b, (āˆ’āˆž, b) = nāˆˆN (āˆ’āˆž, b āˆ’ 1/n] āˆˆ A. Then (a, b) = (āˆ’āˆž, b) āˆ© (a, āˆž) āˆˆ A for every a < b. Using intervals (a, b) with rational endpoints and their countable unions, by Lemma 1.29, every open set is in A. Thus, BR āŠ‚ A.  Corollary 1.31. f : X ā†’ R is a Borel function if and only if f āˆ’1 ((a, āˆž)) āˆˆ BX for all a āˆˆ R. Proof. By Proposition 1.10, the set {A āŠ‚ R | f āˆ’1 (A) āˆˆ BX } is a Ļƒ-algebra. Thus, it contains BR if and only if it contains the sets (a, āˆž) for a āˆˆ R.  Corollary 1.32. (a) If f, g : X ā†’ R are Borel, then their pointwise maximum h(x) = max{f (x), g(x)} is Borel. (b) If f : X ā†’ R is Borel, then āˆ’f is also Borel. Proof. (a) hāˆ’1 ((a, āˆž)) = f āˆ’1 ((a, āˆž)) āˆŖ g āˆ’1 ((a, āˆž)) āˆˆ BX for all a āˆˆ R. (b) The function h(x) = āˆ’x is continuous, so it is a Borel function from R to R. Thus āˆ’f = h ā—¦ f is Borel as a composition of Borel functions.  We now turn to Rn . If X, Y are metric spaces, let us deļ¬ne a metric on X Ɨ Y by d((x1 , y1 ), (x2 , y2 )) = max{dX (x1 , x2 ), dY (y1 , y2 )}.

(1.10)

By induction, this can be applied to a product of n metric spaces; for instance, this makes Rn a metric space with metric dāˆž (x, y) = max |xj āˆ’ yj |. j=1,...,n

(1.11)

Although this is not the Euclidean metric on Rn , it induces the same topology on Rn (Exercise 1.3), so for topological questions, we can use whichever metric is more practical. Metrics that generate the same topology are said to be equivalent; equivalent metrics obviously generate the same Borel Ļƒ-algebra. They also give the same notion of convergence, because convergence of sequences in a metric space can be restated in terms of the metric topology (Exercise 1.4). The metric (1.10), or any metric equivalent to it, will be called a product metric for X Ɨ Y . Lemma 1.33. If U is a base for X and V is a base for Y , the set {U Ɨ V | U āˆˆ U , V āˆˆ V} is a base for X Ɨ Y . Proof. For any open set E āŠ‚ X Ɨ Y and (x, y) āˆˆ E, there is an -ball around (x, y) contained in E. By (1.10), this -ball is of the form A Ɨ B where A, B are -balls in X, Y , respectively. In particular, A, B are open, so

12

1. Measure theory

there exist U āˆˆ U , V āˆˆ V such that x āˆˆ U āŠ‚ A and y āˆˆ V āŠ‚ B. It follows that (x, y) āˆˆ U Ɨ V āŠ‚ A Ɨ B āŠ‚ E. Conversely, since U is open in X and V is open in Y , ļ¬x (x, y) āˆˆ U Ɨ V . Ėœ) < implies x Ėœ āˆˆ U , and dY (y, yĖœ) <

There exists > 0 such that dX (x, x implies yĖœ āˆˆ U . Using (1.10), it follows that U Ɨ V contains the -ball around (x, y). Thus, U Ɨ V is open in X Ɨ Y .  Applying this inductively gives a countable base for Rn :  Corollary 1.34. U = { nj=1 (aj , bj ) | aj , bj āˆˆ Q, aj < bj āˆ€j} is a countable base for Rn . n Corollary n 1.35. The Borel Ļƒ-algebra on R is the Ļƒ-algebra generated by the sets j=1 (aj , bj ), where a1 , . . . , an , b1 , . . . , bn āˆˆ R.  Proof. Denote by A the Ļƒ-algebra generated by the sets nj=1 (aj , bj ). Since those sets are open, A āŠ‚ BRn . For the converse inclusion, by Lemma 1.34 and Lemma 1.28, any open set V āŠ‚ Rn is a countable union of sets of the n  form j=1 (aj , bj ), so any open set V is in A. Thus, BRn āŠ‚ A.

For any vector-valued function h : X ā†’ Rn , denote its components by hj = Ļ€j ā—¦ h, where Ļ€j : Rn ā†’ R denotes the projection to the jth coordinate, Ļ€j (x) = xj . This is also denoted by h = (h1 , . . . , hn ). Proposition 1.36. A vector-valued function h : X ā†’ Rn is Borel if and only if its components hj : X ā†’ R are Borel for all j = 1, 2, . . . , n. Proof. Since the projections Ļ€j are continuous, if h is Borel, then hj = Ļ€j ā—¦h is Borel for each j. Conversely, assume that h1 , . . . , hn : X ā†’ R are Borel functions. For any a1 , b1 , . . . , an , bn āˆˆ R,   n n  (aj , bj ) = hāˆ’1 hāˆ’1 j ((aj , bj )) āˆˆ BX . j=1

j=1

Thus, by Corollary 1.35 and Proposition 1.10, hāˆ’1 (B) āˆˆ BX for every B āˆˆ  BRn . Corollary 1.37. A function f : X ā†’ C is Borel if and only if Re f and Im f are Borel functions from X to R. Proof. In the identiļ¬cation of C as R2 , the absolute value metric on C corresponds to the Euclidean metric on R2 , so BC = BR2 . In that interpretation, Re f, Im f are the components of f , so the claim follows from Proposition 1.36.  Similarly, the next proof uses the identiļ¬cation of C2 with R4 .

1.3. Borel Ļƒ-algebra on the real line and related spaces

13

Proposition 1.38. If f, g : X ā†’ C are Borel, then so are f + g and f g. Proof. The functions F1 (x, y) = x + y and F2 (x, y) = xy are continuous functions from C2 to C, so they are Borel. Since f, g are Borel functions, so is h = (f, g) : X ā†’ C2 by Proposition 1.36. Thus, the functions f + g = F1 ā—¦ h  and f g = F2 ā—¦ h are Borel as compositions of Borel functions. Our next goal in this section is to consider pointwise limits of sequences of Borel functions. Here the robustness of the Borel condition (compared to Riemann integrability) becomes fully apparent. Since limits can be inļ¬nite, general results are naturally formulated on the extended real line Ė† = R āˆŖ {āˆ’āˆž, +āˆž}. R Ė† in the obvious way, every nonempty With the order relation extended to R Ė† has a least upper bound (i.e., supremum) in R, Ė† which can be subset A āŠ‚ R Ā±āˆž. This matches the common usage of sup A = +āˆž in calculus and allows Ė† sup{āˆ’āˆž} = āˆ’āˆž. In particular, for any sequence of functions fn : X ā†’ R, Ė† the following are well deļ¬ned pointwise as functions from X to R: sup fn , nāˆˆN

inf fn ,

nāˆˆN

lim sup fn = inf sup fk , nā†’āˆž

nāˆˆN kā‰„n

lim inf fn = sup inf fk . nā†’āˆž

nāˆˆN kā‰„n

Ė† by compressing R into a bounded interval We can construct a metric on R and measuring distances in the image. Such a metric will correctly capture the notion of convergence to Ā±āˆž used in calculus. Formally, let Ļ„ : R ā†’ (cāˆ’ , c+ ) be a strictly increasing bijection for some āˆ’āˆž < cāˆ’ < c+ < āˆž. Then Ļ„ and Ļ„ āˆ’1 are continuous, so U āŠ‚ R is open if and only if Ļ„ (U ) is Ė† ā†’ [cāˆ’ , c+ ] by Ļ„ (Ā±āˆž) = cĀ± and open. Let us extend Ļ„ to a function Ļ„ : R Ė† deļ¬ne a metric on R by d(x, y) = |Ļ„ (x) āˆ’ Ļ„ (y)|

Ė† āˆ€x, y āˆˆ R.

Ė† by deļ¬Since d is the pullback of the standard metric from [cāˆ’ , c+ ] to R, Ė† if and only if Ļ„ (U ) is open in [cāˆ’ , c+ ]; in other nition, a set U is open in R words, the restriction of d to R generates the standard topology on R. As Ė† is given by in the proof of Lemma 1.29, a countable base for R UĖ† = {(a, b) | a, b āˆˆ Q} āˆŖ {[āˆ’āˆž, b) | b āˆˆ Q} āˆŖ {(a, +āˆž] | a āˆˆ Q}. Using UĖ† , analogously to the proof of Lemma 1.30: Lemma 1.39. BRĖ† is the Ļƒ-algebra generated by the sets (a, āˆž], a āˆˆ R. Ė† is continuous, so any Borel function f : X ā†’ R The inclusion i : R ā†’ R Ė† iā—¦f : X ā†’ R Ė† (for a complete is also Borel when viewed as a function into R, Ė† description of R-valued Borel functions, see Exercise 1.8). The general result about sequences of Borel functions is:

14

1. Measure theory

Ė† Lemma 1.40. For any sequence of Borel functions fn : X ā†’ R, sup fn , nāˆˆN

inf fn ,

lim sup fn ,

nāˆˆN

nā†’āˆž

lim inf fn nā†’āˆž

Ė† are also Borel functions from X to R. Proof. If f (x) = supnāˆˆN fn (x), then for any a āˆˆ R, f āˆ’1 ((a, āˆž]) =

āˆž 

fnāˆ’1 ((a, āˆž]).

n=1

Thus, supnāˆˆN fn is Borel. It follows that inf nāˆˆN fn = āˆ’ supnāˆˆN (āˆ’fn ) is also a Borel function. Using those results, lim supnā†’āˆž fn = inf nāˆˆN supkā‰„n fk and  lim inf nā†’āˆž fn = supnāˆˆN inf kā‰„n fk are also Borel functions. We ļ¬nish this section with some remarks for Borel measures. Let E be a Borel set in X. For a Borel measure on X, its restriction to BE is a Borel measure on E. Conversely, a Borel measure Ī¼ on E generates a Borel measure on X by Ī½(A) = Ī¼(A āˆ© E). This idea of restricting or extending the space can motivate the following deļ¬nition. Deļ¬nition 1.41. Let Ī¼ be a Borel measure on X. (a) The measure Ī¼ is supported on a set E āˆˆ BX if Ī¼(E c ) = 0. (b) The support of Ī¼, denoted supp Ī¼, is the set of all x āˆˆ X such that for every open V containing x, Ī¼(V ) > 0. Note a linguistic subtlety: To say that Ī¼ is supported on E is not the same as saying that the support of Ī¼ is E. The measure can be supported on many diļ¬€erent sets, but its support is uniquely deļ¬ned. For instance, the Dirac measure Ī“x is supported on any set E that contains x, and supp Ī“x = {x}. The measure on R deļ¬ned by Ī¼(A) = #(A āˆ© Q) is supported on the countable set Q, and supp Ī¼ = R. The support has a useful characterization: Lemma 1.42. For any Borel measure Ī¼ on a second-countable space X, supp Ī¼ is the smallest closed set E āŠ‚ X such that Ī¼(E c ) = 0. Proof. Let U be a countable base for X. Taking complements of the definition, (supp Ī¼)c is the set of all x for which there is an open set V such that x āˆˆ V and Ī¼(V ) = 0. This is equivalent to existence of A āˆˆ U with x āˆˆ A āŠ‚ V and Ī¼(A) = 0. In other words,   V = A. (1.12) (supp Ī¼)c = V open Ī¼(V )=0

AāˆˆU Ī¼(A)=0

The second union in (1.12) is countable, so from Ī¼(A) = 0 for all A, it follows that Ī¼((supp Ī¼)c ) = 0; thus, (supp Ī¼)c is an open set of zero measure. By

1.4. Lebesgue integration

15

the ļ¬rst union in (1.12), (supp Ī¼)c is the smallest open set of zero measure. Taking complements completes the proof. 

1.4. Lebesgue integration In this section, we develop integration with respect to a Borel measure. Any set or function appearing below is implied to be Borel, and where other sets or functions are derived from it, they can be proved to be Borel by the material from the previous sections; we will keep such steps implicit. Deļ¬nition 1.43. A function s : X ā†’ C is called simple if it only takes ļ¬nitely many values, i.e., the set s(X) = {s(x) | x āˆˆ X} is ļ¬nite. For a positive simple function s, we deļ¬ne the integral of s with respect to Ī¼ as   cĪ¼(sāˆ’1 ({c})). (1.13) s dĪ¼ = cāˆˆs(X)

As all of integration, this formula is motivated by the area of a rectangle as height times base; the strength of Lebesgue integration can be traced to the fact that the bases of our rectangles are arbitrary Borel sets. Integration theory uses the convention that c Ā· āˆž = āˆž for c > 0 but 0 Ā· āˆž = 0. In formula (1.13), if s takes the value c = 0, its contribution to the integral is zero regardless of whether Ī¼(sāˆ’1 ({0})) is ļ¬nite or inļ¬nite.  A family of sets {AĪ± | Ī± āˆˆ I} is called a partition of X if X = Ī±āˆˆI AĪ± and AĪ± āˆ© AĪ² = āˆ… whenever Ī± = Ī². The deļ¬nition (1.13) can be rephrased by using characteristic functions of sets: if c1 , . . . , cn are distinct elements of [0, āˆž) and A1 , . . . , An is a partition of X, then the integral of the function s=

n 

cj Ļ‡Aj

(1.14)

j=1

is deļ¬ned by

 s dĪ¼ =

n 

cj Ī¼(Aj ).

(1.15)

j=1

It would be cumbersome to always search for this exact partition of a simple function s; fortunately, as we are about to see, this is not necessary. Lemma 1.44. (a) If A1 , . . . , An is a partition of X and c1 , . . . , cn ā‰„ 0 (not necessarily distinct), the integral of the function (1.14) is given by (1.15).  (b) If s is a positive simple function and Ī» ā‰„ 0, then (Ī»s) dĪ¼ = Ī» s dĪ¼.

16

1. Measure theory

(c) If s, t are positive simple functions,   then s + t is a positive simple function and (s + t) dĪ¼ = s dĪ¼ + t dĪ¼.  (d) If s, t are simple functions and 0 ā‰¤ s ā‰¤ t pointwise, then s dĪ¼ ā‰¤ t dĪ¼. (e) If A1 , . . . , An are any Borel subsets of X and c1 , . . . , cn āˆˆ [0, āˆž), the integral of the function (1.14) is given by (1.15). Proof. (a) If cj = ck for some j = k, using Ī¼(Aj ) + Ī¼(Ak ) = Ī¼(Aj āˆŖ Ak ) allows us to merge those two sets in the partition without aļ¬€ecting the sum; after ļ¬nitely many steps, we will end up at the partition used in (1.13). (b) This follows immediately from (1.14) and (1.15). (c) Denote by c1 , . . . , cn the values of s, by d1 , . . . , dm the values of t, and denote Aj = sāˆ’1 ({cj }), Bk = tāˆ’1 ({dk }). Then {Aj āˆ© Bk | 1 ā‰¤ j ā‰¤ n, 1 ā‰¤ k ā‰¤ m} is a partition such that s, t, s + t are constant on each set. Written in that partition, the claim to be proved reduces to the obvious equality m  n 

(cj + dk )Ī¼(Aj āˆ© Bk ) =

j=1 k=1

m  n 

cj Ī¼(Aj āˆ© Bk ) +

j=1 k=1

m  n 

dk Ī¼(Aj āˆ© Bk ).

j=1 k=1

(d) The function t āˆ’ s is positive because s ā‰¤t and is simple    because s and t are simple. Thus, by (c), t dĪ¼ = s dĪ¼ + (t āˆ’ s) dĪ¼ ā‰„ s dĪ¼.  (e) By deļ¬nition, cj Ļ‡Aj dĪ¼ = cj Ī¼(Aj ) + 0Ī¼(Acj ) = cj Ī¼(Aj ). Using (b) to take the sum of these functions completes the proof.  If f is a positive simple function, we can conclude   s dĪ¼ f dĪ¼ = sup

(1.16)

s simple 0ā‰¤sā‰¤f

  because s ā‰¤ f implies s dĪ¼ ā‰¤ f dĪ¼ and equality holds for s = f . Noting that the right-hand side makes sense even if f is not simple, we can use it to generalize the integral.  Deļ¬nition 1.45. For f : X ā†’ [0, āˆž], deļ¬ne f dĪ¼ by (1.16). Lemma 1.46. (a) If 0 ā‰¤ f ā‰¤ g, then



f dĪ¼ ā‰¤



g dĪ¼.   (b) If f ā‰„ 0 and c āˆˆ [0, āˆž), then cf dĪ¼ = c f dĪ¼. Proof. (a) Any simple function  s such that 0 ā‰¤ s ā‰¤ f also obeys 0 ā‰¤s ā‰¤ g, so the deļ¬ning supremum for g dĪ¼ is over a bigger set than that for f dĪ¼. (b) The case c = 0 is trivial. For c > 0, the simple function s obeys  0 ā‰¤ s ā‰¤ f if and only if the simple function cs obeys 0 ā‰¤ cs ā‰¤ cf .

1.4. Lebesgue integration

17

 Note that f dĪ¼ can be inļ¬nite. A trivial but   often used consequence of (a) is that if 0 ā‰¤ f ā‰¤ g and g dĪ¼ < āˆž, then f dĪ¼ < āˆž. The ļ¬rst remarkable result of integration theory is: Theorem 1.47 (Monotone convergence theorem). For any sequence fn : X ā†’ [0, āˆž] such that fn ā‰¤ fn+1 for all n āˆˆ N,   lim fn dĪ¼. fn dĪ¼ = lim nā†’āˆž

nā†’āˆž

Lemma 1.48. If s :  X ā†’ [0, āˆž) is a simple function and En are sets such that En āŠ‚ En+1 and nāˆˆN En = X, then   (1.17) sĻ‡En dĪ¼ = s dĪ¼. lim nā†’āˆž

Proof. Let s be given by (1.14). For any n, the function sĻ‡En is simple and  m  sĻ‡En dĪ¼ = cj Ī¼(Aj āˆ© En ). (1.18) j=1

By Theorem 1.22(c), Ī¼(Aj āˆ© En ) ā†’ Ī¼(Aj ) as n ā†’ āˆž. Applying this to each term of (1.18) gives (1.17).  Proof of Theorem 1.47. Denote the pointwise limit by f = limnā†’āˆž fn . Since the sequence is increasing, fn ā‰¤ fn+1 ā‰¤ f for all n. This implies that the integrals of fn have a limit and that   fn dĪ¼ ā‰¤ f dĪ¼. lim nā†’āˆž

To prove the converse inequality, ļ¬x c āˆˆ (0, 1) and a simple function s ā‰¤ f . Deļ¬ne En = {x āˆˆ X | fn (x) ā‰„ cs(x)}. Then fn ā‰„ fn Ļ‡En ā‰„ csĻ‡En implies   (1.19) fn dĪ¼ ā‰„ csĻ‡En dĪ¼.  Note that En āŠ‚ En+1 because fn ā‰¤ fn+1 . Moreover, let us show nāˆˆN En = X by verifying three cases: If f (x) = 0, then s(x) = 0 so x āˆˆ En for all n. If f (x) āˆˆ (0, āˆž), then limnā†’āˆž fn (x) = f (x) > cf (x) ā‰„ cs(x) so fn (x) ā‰„ cs(x) for large enough n. If f (x) = āˆž, then limnā†’āˆž fn (x) = āˆž > cs(x), so fn (x) ā‰„ cs(x) for large enough n. By Lemma 1.48, taking n ā†’ āˆž in (1.19),     fn dĪ¼ ā‰„ lim csĻ‡En dĪ¼ = cs dĪ¼ = c s dĪ¼. lim nā†’āˆž

nā†’āˆž

Since c āˆˆ (0, 1) is arbitrary, this also implies   fn dĪ¼ ā‰„ s dĪ¼. lim nā†’āˆž

Taking the supremum over simple functions s ā‰¤ f completes the proof. 

18

1. Measure theory

To illustrate this abstract Lebesgue integral, let us show that it includes series with positive terms as a special case, and use that to prove a rearrangement theorem for series. Example 1.49. Consider a sequence (an )āˆž n=1 with an ā‰„ 0 for all n. (a) If Ī½ denotes the counting measure on N and f (n) = an for all n, then  āˆž  aj . (1.20) f dĪ½ = j=1

(b) If Ļ€ : N ā†’ N is a bijection, then

āˆž

j=1 aj

=

āˆž

k=1 aĻ€(k) .

2, . . . , n}. For each n āˆˆ N, the function Ļ‡En f Proof. (a) Denote En =  {1, n is simple, so En f dĪ½ = j=1 f (j). Since Ļ‡En f is an increasing sequence of functions converging pointwise to f , by the monotone convergence theorem, letting n ā†’ āˆž gives (1.20). (b)  Repeating the proof n of (a) for the sets En = {Ļ€(1), Ļ€(2), . . . , Ļ€(n)} gives f dĪ½ = limnā†’āˆž k=1 aĻ€(k) . By (1.20), this completes the proof.  Instead of the deļ¬nition of the integral as a supremum, it is often useful to use an explicit sequence of simple functions which monotonically converges to f and combine this with the monotone convergence theorem. Such a sequence is constructed in the next lemma, and will be immediately used to prove additivity of the integral. Lemma 1.50. If f : X ā†’ [0, āˆž], there exist simple functions sn : X ā†’ [0, āˆž) such that sn ā‰¤ sn+1 and sn ā†’ f pointwise. Proof. A sequence satisfying these conditions is given by

2āˆ’n 2n f (x) 0 ā‰¤ f (x) < n  sn (x) = n n ā‰¤ f (x).    Lemma 1.51. If f, g : X ā†’ [0, āˆž], then (f + g) dĪ¼ = f dĪ¼ + g dĪ¼. Proof. Pick increasing sequences of simple functions such that sn ā†’ f , tn ā†’ g. Then sn + tn ā†’ f + g, so taking the limit as n ā†’ āˆž of    (sn + tn ) dĪ¼ = sn dĪ¼ + tn dĪ¼,    monotone convergence gives (f + g) dĪ¼ = f dĪ¼ + g dĪ¼.  Proposition 1.52. For any sequence of functions gn : X ā†’ [0, āˆž],    āˆž āˆž   gn dĪ¼ = gn dĪ¼. n=1

n=1

1.4. Lebesgue integration

19

Proof. This follows by monotone  convergence n applied to the increasing sen  quence fn = j=1 gj , because fn dĪ¼ = j=1 gj dĪ¼ for each n. When going beyond monotone limits, Fatouā€™s lemma will be useful: Theorem 1.53 (Fatouā€™s lemma). For any functions fn : X ā†’ [0, āˆž],   lim inf fn dĪ¼ ā‰¤ lim inf fn dĪ¼. nā†’āˆž

nā†’āˆž

Proof. Let gn = inf kā‰„n fk . Then gn is an increasing sequence of functions and limnā†’āˆž gn = lim inf kā†’āˆž fk . By monotone convergence,    lim gn dĪ¼ = lim gn dĪ¼. lim inf fn dĪ¼ = nā†’āˆž nā†’āˆž nā†’āˆž   However, gn ā‰¤ fn implies gn dĪ¼ ā‰¤ fn dĪ¼, and therefore     gn dĪ¼ = lim inf gn dĪ¼ ā‰¤ lim inf fn dĪ¼. lim nā†’āˆž

nā†’āˆž

nā†’āˆž

The integral (1.16) should be thought of as the integral of f over the entire space X. For Borel subsets E āŠ‚ X, we deļ¬ne the integral over E as   f dĪ¼ = Ļ‡E f dĪ¼. E

The following proposition provides a construction of a new measure from another measure and a multiplicative weight. Proposition 1.54. For any Borel measure Ī¼ on X and h : X ā†’ [0, āˆž], another Borel measure Ī½ on X is deļ¬ned by  Ī½(E) = h dĪ¼ āˆ€E āˆˆ BX . E

Moreover, for all g : X ā†’ [0, āˆž],   g dĪ½ = gh dĪ¼.

(1.21)

This measure is commonly described by saying dĪ½ = h dĪ¼. Proof. Clearly, Ī½(āˆ…) = 0. For any  disjoint sets āˆžEn , n āˆˆ N, Proposition 1.52 E ) = applied to gn = hĻ‡En gives Ī½ ( āˆž n=1 n n=1 Ī½(En ). Thus, Ī½ is a measure. Equality (1.21) holds for all g = Ļ‡E , so by linear combinations, it holds for all simple functions g. For an arbitrary g : X ā†’ [0,  āˆž], use simple functions sn ā‰¤ sn+1 such that sn ā†’ g. Then sn dĪ½ = sn h dĪ¼ for each n. Since sn h ā‰¤ sn+1 h, sn h ā†’ gh, applying monotone convergence to both sides of this equality gives (1.21).  Another very useful construction is the pushforward of a measure:

20

1. Measure theory

Lemma 1.55. If Ī¼ is a Borel measure on X and g : X ā†’ Y is a Borel function, the pushforward of Ī¼ by g is the Borel measure Ī½ on Y deļ¬ned by Ī½(B) = Ī¼(g āˆ’1 (B)). For any Borel function f : Y ā†’ [0, āˆž],   f dĪ½ = (f ā—¦ g) dĪ¼.

(1.22)

Proof. If the sets Bn , n āˆˆ N are disjoint, so are g āˆ’1 (Bn ), so  āˆž  āˆž  āˆž    āˆ’1 āˆ’1 Bn g (Bn ) = Ī¼(g āˆ’1 (Bn )) =Ī¼ Ī¼ g n=1

n=1

n=1

implies Ļƒ-additivity of Ī½. Also, Ī½(āˆ…) = Ī¼(āˆ…) = 0, so Ī½ is a measure. For f = Ļ‡E , (1.22) holds by deļ¬nition. By linearity, (1.22) holds for all simple functions. For simple functions sn such that sn ā‰¤ sn+1 and sn ā†’ f , the functions sn ā—¦ g are also  simple and obey sn ā—¦ g ā‰¤ sn+1 ā—¦ g and sn ā—¦ g ā†’ f ā—¦ g. Since sn dĪ½ = (sn ā—¦ g) dĪ¼, taking n ā†’ āˆž and applying monotone convergence on both sides proves (1.22).  So far, we have seen integration theory as derived from measure theory. However, sometimes we use integrals to estimate measures: Lemma 1.56 (Markovā€™s inequality). For any f : X ā†’ [0, āˆž] and c > 0,  1 f dĪ¼. Ī¼({x | f (x) ā‰„ c}) ā‰¤ c Proof. This follows from f ā‰„ cĻ‡A where A = {x | f (x) ā‰„ c}.



A property is said to hold Ī¼-almost everywhere (or ā€œĪ¼-a.e.ā€) if there is a set A such that Ī¼(A) = 0 and the property holds for all x āˆˆ Ac . Sets of measure 0 are negligible in integration theory: Proposition 1.57. Let f, g : X ā†’ [0, āˆž]. Then the following hold.  (a) If f dĪ¼ < āˆž, then f < āˆž holds Ī¼-a.e.  (b) f dĪ¼ = 0 if and only if f = 0 Ī¼-a.e.   (c) If f = g Ī¼-a.e., then f dĪ¼ = g dĪ¼. Proof. (a) By Markovā€™s inequality with c = k āˆˆ N, Ī¼({x | f (x) = āˆž}) ā‰¤ Ī¼({x | f (x) ā‰„ k}) ā‰¤

1 k

 f dĪ¼.

Taking k ā†’ āˆž proves Ī¼({x | f (x) = āˆž}) = 0. (b) Assume that f = 0 Ī¼-a.e. For every simple function 0 ā‰¤ s ā‰¤ f , s = 0  Ī¼-a.e. Thus, by deļ¬nition, s dĪ¼ = 0, so taking the supremum over simple

1.4. Lebesgue integration

21

  functions s ā‰¤ f , f dĪ¼ = 0. Conversely, if f dĪ¼ =0, then by Markovā€™s inequality, for every k āˆˆ N, Ī¼({x | f (x) ā‰„ 1/k}) ā‰¤ k f dĪ¼ = 0, so taking the union over k āˆˆ N shows Ī¼({x | f (x) > 0}) = 0. (c) The set E = {x | f (x) = g(x)} obeys Ī¼(E)  = 0. Using the dec composition f = f Ļ‡ + f Ļ‡ and (b), f dĪ¼ = f Ļ‡E c dĪ¼. Analogously, E E    g dĪ¼ = gĻ‡E c dĪ¼. Since f Ļ‡E c = gĻ‡E c , this completes the proof. Since a countable union of sets of zero measure has zero measure, it is common to impose countably many conditions that hold Ī¼-a.e. and assume that they all hold away from the same set of zero measure, like in the following theorem. In cases when f (x) = limnā†’āˆž fn (x) exists Ī¼-a.e., it is common to consider f to be deļ¬ned by that equation and to not explicitly specify the value of f on the remaining zero measure set. For example: Theorem 1.58 (Monotone convergence theorem, again). If functions fn : X ā†’ [0, āˆž] obey fn ā‰¤ fn+1 Ī¼-a.e. for all n āˆˆ N, and fn ā†’ f Ī¼-a.e., then   fn dĪ¼ = f dĪ¼. lim nā†’āˆž

Proof. Denote by E a set such that all assumptions hold on E and Ī¼(E c ) =   0. By monotone convergence, E fn dĪ¼ ā†’ E f dĪ¼. Since fn = f Ļ‡E and  f = f Ļ‡E Ī¼-a.e., the claim follows by Proposition 1.57. Ė† Let us now extend integration to real-valued Borel functions h : X ā†’ R. For such h, we denote hĀ± = max{Ā±h, 0}. Note that hĀ± ā‰„ 0, h = h+ āˆ’ hāˆ’ , and |h| = h+ + hāˆ’ . If at least one of the integrals hĀ± dĪ¼ is ļ¬nite, we deļ¬ne    h dĪ¼ = h+ dĪ¼ āˆ’ hāˆ’ dĪ¼.  For instance, if f : X ā†’ [0, āˆž] and f dĪ¼ < āˆž, then log f ā‰¤ f implies (log f )+ ā‰¤ f , so (log f )+ dĪ¼ < āˆž. Thus, log f dĪ¼ is deļ¬ned, although its value can be āˆ’āˆž. However, in most situations, we will work in the case when both hĀ± dĪ¼ are ļ¬nite, and we call such functions h integrable.    Lemma 1.59. |h| dĪ¼ < āˆž if and only if h+ dĪ¼ < āˆž and hāˆ’ dĪ¼ < āˆž. Proof. One implication follows from hĀ± ā‰¤ |h| and the other from |h| ā‰¤  h+ + hāˆ’ . Proposition 1.60. Let Ī¼ be a measure on X. (a) If c āˆˆ R and f : X ā†’ R is integrable, then cf is integrable and   (cf ) dĪ¼ = c f dĪ¼. (1.23)

22

1. Measure theory

(b) If f, g : X ā†’ R are integrable, then f + g is integrable and    (f + g) dĪ¼ = f dĪ¼ + g dĪ¼. (1.24)   Proof. (a) Integrability of cf follows from |cf | dĪ¼ = |c| |f | dĪ¼ < āˆž. Equation (1.23) follows from Lemma 1.46 with the observation that for c ā‰„ 0, (cf )Ā± = cfĀ± , and for c < 0, (cf )Ā± = (āˆ’c)fāˆ“ . (b) Integrability of h = f + g follows from the triangle inequality, since     |f + g| dĪ¼ ā‰¤ (|f | + |g|) dĪ¼ = |f | dĪ¼ + |g| dĪ¼ < āˆž. (1.25) Since h+ āˆ’ hāˆ’ = f+ āˆ’ fāˆ’ + g+ āˆ’ gāˆ’ implies h+ + fāˆ’ + gāˆ’ = hāˆ’ + f+ + g+ , additivity of integrals of positive functions implies       h+ dĪ¼ + fāˆ’ dĪ¼ + gāˆ’ dĪ¼ = hāˆ’ dĪ¼ + f+ dĪ¼ + g+ dĪ¼. Regrouping terms gives (1.24).



We will now further generalize integration to complex-valued functions. Deļ¬nition 1.61. We denote by L1 (X, dĪ¼) the set of f : X ā†’ C such that  |f | dĪ¼ < āˆž. Consistently with prior terminology, we call such functions f integrable. Lemma 1.62. f is integrable if and only if Re f and Im f are integrable. Proof. One direction follows from |Re f | ā‰¤ |f | and |Im f | ā‰¤ |f |, and the other from |f | ā‰¤ |Re f | + |Im f |.  For f āˆˆ L1 (X, dĪ¼) we deļ¬ne    f dĪ¼ = Re f dĪ¼ + i Im f dĪ¼. Lemma 1.63. (a) If c āˆˆ C and f āˆˆ L1 (X, dĪ¼), then cf āˆˆ L1 (X, dĪ¼) and (1.23) holds. (b) If f, g āˆˆ L1 (X, dĪ¼), then f + g āˆˆ L1 (X, dĪ¼) and (1.24) holds.   Proof. (a) Integrability of cf follows from |cf | dĪ¼ = |c| |f | dĪ¼ < āˆž. The equality (1.23) follows from the real-valued case by Re(cf ) = Re c Re f āˆ’ Im c Im f and Im(cf ) = Re c Im f + Im c Re f . (b) Repeating argument (1.25) shows that f + g āˆˆ L1 (X, dĪ¼). Equality (1.24) follows from Re(f +g) = Re f +Re g and Im(f +g) = Im f +Im g. 

1.4. Lebesgue integration

23

Lemma 1.64. If f āˆˆ L1 (X, dĪ¼), then       f dĪ¼ ā‰¤ |f | dĪ¼.      Proof. Pick Ļ‰ āˆˆ C such that |Ļ‰| = 1 and Ļ‰ f dĪ¼ =  f dĪ¼. Then         f dĪ¼ = Ļ‰ f dĪ¼ = Re Ļ‰f dĪ¼ = Re(Ļ‰f ) dĪ¼.    Using Re(Ļ‰f ) ā‰¤ |Ļ‰f | dĪ¼ = |f | completes the proof.



Theorem 1.65 (Dominated convergence theorem). Consider a sequence of fn āˆˆ L1 (X, dĪ¼) dominated by some g āˆˆ L1 (X, dĪ¼) in the sense that |fn (x)| ā‰¤ g(x)

(1.26)

for all n āˆˆ N and Ī¼-a.e. x. Assume that fn converge pointwise Ī¼-a.e. to a Borel function f . Then f āˆˆ L1 (X, dĪ¼),  (1.27) lim |fn āˆ’ f | dĪ¼ = 0 nā†’āˆž





and lim

nā†’āˆž

fn dĪ¼ =

f dĪ¼.

(1.28)

Proof. From (1.26), by passing to pointwise limits, it follows that |f | ā‰¤ g Ī¼-a.e., so f āˆˆ L1 (X, dĪ¼). Deļ¬ne hn = 2g āˆ’ |fn āˆ’ f | ā‰„ 0. Since hn ā†’ 2g Ī¼-a.e., by Fatouā€™s lemma,   2g dĪ¼ ā‰¤ lim inf (2g āˆ’ |fn āˆ’ f |) dĪ¼. nā†’āˆž  Since 2g and |fn āˆ’ f | are integrable, we can subtract the constant 2g dĪ¼ from both sides and multiply by āˆ’1 to obtain  lim sup |fn āˆ’ f | dĪ¼ ā‰¤ 0. nā†’āˆž

By this implies (1.27). Now (1.28) follows from  bound,  a trivial  lower   fn dĪ¼ āˆ’ f dĪ¼ ā‰¤ |fn āˆ’ f | dĪ¼.  Lebesgue integration does not include conditionally convergent integralsā€”note that f is integrable if and only if |f | is integrableā€”but this is usually not an important limitation. The connection with series from Example 1.49 motivates: Deļ¬nition 1.66. Let Ī½ denote counting measure on a set Ī“. If f : Ī“ ā†’ [0, āˆž] or f : Ī“ ā†’ C with |f | dĪ½ < āˆž, we deļ¬ne   f (j) = f dĪ½. jāˆˆJ

24

1. Measure theory

This gives a notion of summation over any set, in a way that does not include conditionally convergent sequences but is independent of ordering. The steps in the deļ¬nition of the integral, from positive to complex functions, are reļ¬‚ected in many proofs in integration theory. For instance: Proposition 1.67. Let dĪ½ = f dĪ¼ in the notation of Proposition 1.54. For any f : X ā†’ C, f āˆˆ L1 (X, dĪ½) if and only if f h āˆˆ L1 (X, dĪ¼). If this holds, then   f dĪ½ =

f h dĪ¼.

(1.29)

Proof. By Proposition 1.54, equation (1.29)   holds for positive functions. Thus, applying it to |f | shows |f | dĪ½ = |f |h dĪ¼, which proves the ļ¬rst claim.   If f is real-valued, applying (1.29) to fĀ± gives  fĀ± dĪ½ =  fĀ± h dĪ¼. Using f = f+ āˆ’ fāˆ’ and subtracting integrals gives f dĪ½ = f h dĪ¼, so (1.29) holds for real-valued functions. Likewise, if f is complex-valued, using f = Re f + i Im f and applying (1.29) to Re f, Im f shows that it holds for f . 

1.5. Lebesgueā€“Stieltjes measures on R In this section, we study measures of R. Since the Borel Ļƒ-algebra on R is generated by intervals, it is natural to try to understand a measure on R by examining how it acts on intervals. This motivates the following deļ¬nition: Deļ¬nition 1.68. Let Ī¼ be a Borel measure on R. A function Ī± : R ā†’ R is called a distribution function of Ī¼ if Ī¼((x, y]) = Ī±(y) āˆ’ Ī±(x)

āˆ€x, y āˆˆ R, x < y.

(1.30)

Example 1.69. Ļ‡[x0 ,āˆž) is a distribution function for the Dirac measure Ī“x0 . If a distribution function exists, it is determined uniquely up to an additive constant. Its existence is considered in the following lemma: Lemma 1.70. For a Borel measure Ī¼ on R, the following are equivalent: (a) Ī¼ is ļ¬nite on compacts; (b) Ī¼((x, y]) < āˆž for all x, y āˆˆ R with x < y; (c) Ī¼ has a distribution function. Proof. Every compact set K āŠ‚ R is contained in some interval (āˆ’C, C] āŠ‚ [āˆ’C, C], so (a) and (b) are equivalent. If Ī¼ has a distribution function, then Ī¼((x, y]) < āˆž for all x, y āˆˆ R by (1.30), so (iii) implies (ii). To prove that

1.5. Lebesgueā€“Stieltjes measures on R

25

(ii) implies (iii), deļ¬ne āŽ§ āŽŖ x>0 āŽØĪ¼((0, x]) Ī±(x) = 0 x=0 āŽŖ āŽ© āˆ’Ī¼((x, 0]) x < 0.

(1.31)

The property (1.30) follows from (1.31) by additivity of Ī¼, applied on a case-by-case basis depending on the signs of x, y.  By (1.30), the distribution function is an increasing function, i.e., x < y implies Ī±(x) ā‰¤ Ī±(y). We recall some properties of increasing functions: Lemma 1.71. Any increasing function Ī± : R ā†’ R has the following properties. (a) The function Ī± has one-sided limits Ī±+ (x) = lim Ī±(t),

x āˆˆ R āˆŖ {āˆ’āˆž},

Ī±āˆ’ (x) = lim Ī±(t),

x āˆˆ R āˆŖ {+āˆž},

tā†“x tā†‘x

which are themselves increasing functions of x. (b) Ī±āˆ’ (x) ā‰¤ Ī±(x) ā‰¤ Ī±+ (x) for all x āˆˆ R. Ė† with x < y. (c) Ī±+ (x) ā‰¤ Ī±āˆ’ (y) for all x, y āˆˆ R (d) (Ī±āˆ’ )+ = Ī±+ and (Ī±+ )āˆ’ = Ī±āˆ’ . Proof. (a) Let us begin by deļ¬ning Ī±+ (x) = inf Ī±(t), t>x

Ī±āˆ’ (x) = sup Ī±(t).

(1.32)

t Ī±+ (x) there exists t > x such that Ī±(t) < c, so for all y āˆˆ (x, x + t), Ī±+ (x) ā‰¤ Ī±(y) < c. This implies that Ī±+ (x) is the right limit of Ī± at x. Since taking the inļ¬mum over a larger set can only give a smaller value, Ī±+ (x) ā‰¤ Ī±+ (y) if x < y. The statements for Ī±āˆ’ follow analogously. (b) Since Ī±(x) ā‰¤ Ī±(t) for all t > x, taking the limit as t ā†“ x implies that Ī±(x) ā‰¤ Ī±+ (x). Similarly, Ī±āˆ’ (x) ā‰¤ Ī±(x). (c) Picking t āˆˆ (x, y) and using (1.32), we obtain Ī±+ (x) ā‰¤ Ī±(t) ā‰¤ Ī±āˆ’ (y). (d) The ļ¬rst claim follows from the squeeze theorem applied to Ī±+ (x) ā‰¤  Ī±āˆ’ (t) ā‰¤ Ī±(t) as t ā†“ x. The second claim is proved analogously.

26

1. Measure theory

Varying interval endpoints provides additional links between Ī¼ and Ī±: Lemma 1.72. If Ī¼ is a measure on R with a distribution function Ī±, then the following hold. (a) Ī± is right-continuous, i.e., Ī±+ (x) = Ī±(x) for all x āˆˆ R. (b) For all x, y āˆˆ R with x < y, we have Ī¼((x, y)) = Ī±āˆ’ (y) āˆ’ Ī±+ (x). (c) For any x, y āˆˆ R with x ā‰¤ y, we have Ī¼([x, y]) = Ī±+ (y) āˆ’ Ī±āˆ’ (x). Proof. (a) For any x āˆˆ R, by considering a decreasing sequence of intervals, lim (Ī±(x + 1/n) āˆ’ Ī±(x)) = lim Ī¼((x, x + 1/n]) = Ī¼(āˆ…).

nā†’āˆž

This gives

nā†’āˆž

Ī±+ (x)

āˆ’ Ī±(x) = 0, so Ī± is right-continuous.

(b) This is proved similarly by computing limnā†’āˆž Ī¼((x + 1/n, y āˆ’ 1/n]). (c) This is proved similarly by computing limnā†’āˆž Ī¼((x āˆ’ 1/n, y + 1/n]).  This discussion of distribution functions has been merely a warmup; some further calculations of this kind, which compute the measures of other intervals, singletons, and arbitrary open sets V āŠ‚ R, are left as Exercises 1.14 and 1.15. We turn instead to the ļ¬rst of two important results in this section: the construction of measures with prescribed distribution functions. Theorem 1.73. For any increasing right-continuous function Ī± : R ā†’ R, there exists a Borel measure Ī¼Ī± on R such that Ī¼Ī± ((a, b]) = Ī±(b) āˆ’ Ī±(a)

āˆ€a, b āˆˆ R, a < b.

This measure is called the Lebesgueā€“Stieltjes measure corresponding to Ī±. The proof will use CarathĀ“eodoryā€™s theorem. Consider the family Ė† a < b}, E = {āˆ…} āˆŖ {(a, b) | a, b āˆˆ R,

(1.33)

ļ¬x an arbitrary increasing function Ī± : R ā†’ R, and deļ¬ne Ļ : E ā†’ [0, āˆž] by Ļ(āˆ…) = 0,

Ļ((a, b)) = Ī±āˆ’ (b) āˆ’ Ī±+ (a)

Ė† a < b. āˆ€a, b āˆˆ R,

This weight generates an outer measure Ī¼āˆ— by (1.5). Our goal is to prove that all Borel sets are measurable with respect to Ī¼āˆ— and that, if Ī± is rightcontinuous, the resulting Borel measure has distribution function Ī±. The ļ¬rst step is to determine the outer measure of intervals. To pass from countable covers to ļ¬nite covers, it is useful to ļ¬rst consider the compact case: Lemma 1.74. For any p, q āˆˆ R with p ā‰¤ q, Ī¼āˆ— ([p, q]) = Ī±+ (q) āˆ’ Ī±āˆ’ (p).

1.5. Lebesgueā€“Stieltjes measures on R

27

Proof. [p, q] āŠ‚ (p āˆ’ , q + ) implies Ī¼āˆ— ([p, q]) ā‰¤ Ī±āˆ’ (q + ) āˆ’ Ī±+ (p āˆ’ ) for any > 0. Letting ā†’ 0, we get Ī¼āˆ— ([p, q]) ā‰¤ Ī±+ (q) āˆ’ Ī±āˆ’ (p). Conversely, consider any countable cover of [p, q] by open intervals Ij , j āˆˆ N. By compactness, this cover has a ļ¬nite subcover; among all ļ¬nite subcovers, consider one with the smallest possible number of intervals, and denote the intervals by (aj , bj ), j = 1, . . . , n. Minimality implies that aj = ak and bj = bk for j = k, otherwise one of the intervals (aj , bj ), (ak , bk ) would contain the other and could be removed from the cover. Label the intervals so that a1 < a2 < Ā· Ā· Ā· < an . Minimality further implies that b1 < b2 < Ā· Ā· Ā· < bn (otherwise (aj+1 , bj ) āŠ‚ (aj , bj+1 ) for some j). Moreover, a1 < p < b1 since p is covered and each interval intersects [p, q]. Analogously, an < q < bn . Finally, ak+1 < bk for 1 ā‰¤ k ā‰¤ n āˆ’ 1, otherwise the point bk would not be covered. Thus, Ī±+ (ak+1 ) ā‰¤ Ī±āˆ’ (bk ), so nāˆ’1 

(Ī±āˆ’ (bk ) āˆ’ Ī±+ (ak )) ā‰„

k=1

Adding

nāˆ’1 

(Ī±+ (ak+1 ) āˆ’ Ī±+ (ak )) = Ī±+ (an ) āˆ’ Ī±+ (a1 ).

k=1

Ī±āˆ’ (bn )

āˆ’ Ī±+ (an ) n  āˆ’

and using Ī±āˆ’ (bn ) ā‰„ Ī±+ (q) and Ī±+ (a1 ) ā‰¤ Ī±āˆ’ (p),

(Ī± (bk ) āˆ’ Ī±+ (ak )) ā‰„ Ī±+ (q) āˆ’ Ī±āˆ’ (p).

k=1

Thus, the sum of weights over this ļ¬nite subcover is bounded below by Ī±+ (q) āˆ’ Ī±āˆ’ (p). This lower bound then also applies to the original countable  cover, which was arbitrary, so Ī¼āˆ— ([p, q]) ā‰„ Ī±+ (q) āˆ’ Ī±āˆ’ (p). Compactness was used crucially to obtain the lower bound for the outer measure. Using the result for compact intervals, it becomes easy to compute outer measures of other intervals: Lemma 1.75. For any open interval (a, b) āŠ‚ R, Ī¼āˆ— ((a, b)) = Ļ((a, b)) = Ī±āˆ’ (b) āˆ’ Ī±+ (a). Proof. The trivial cover of (a, b) by itself shows Ī¼āˆ— ((a, b)) ā‰¤ Ī±āˆ’ (b) āˆ’ Ī±+ (a). For any compact interval [p, q] āŠ‚ (a, b), Ī¼āˆ— ((a, b)) ā‰„ Ī¼āˆ— ([p, q]) = Ī±+ (q) āˆ’ Ī±āˆ’ (p). Taking limits p ā†“ a, q ā†‘ b proves Ī¼āˆ— ((a, b)) ā‰„ Ī±āˆ’ (b) āˆ’ Ī±+ (a). Lemma 1.76. For any half-open interval (a, c] āŠ‚ R, Ī¼āˆ— ((a, c]) = Ī±+ (c) āˆ’ Ī±+ (a). Proof. For any b > c, (a, c] āŠ‚ (a, b) implies Ī¼āˆ— ((a, c]) ā‰¤ Ī¼āˆ— ((a, b)) = Ī±āˆ’ (b) āˆ’ Ī±+ (a), so taking the limit b ā†“ c gives Ī¼āˆ— ((a, c]) ā‰¤ Ī±+ (c) āˆ’ Ī±+ (a).



28

1. Measure theory

Conversely, for any p āˆˆ (a, c], [p, c] āŠ‚ (a, c] implies Ī¼āˆ— ((a, c]) ā‰„ Ī¼āˆ— ([p, c]) = Ī±+ (c) āˆ’ Ī±āˆ’ (p), and taking the limit p ā†“ a gives Ī¼āˆ— ((a, c]) ā‰„ Ī±+ (c) āˆ’ Ī±+ (a).



Lemma 1.77. For any I āˆˆ E and any c āˆˆ R, Ī¼āˆ— (I) = Ī¼āˆ— (I āˆ© (āˆ’āˆž, c]) + Ī¼āˆ— (I āˆ© (c, āˆž)).

(1.34)

Proof. If I āŠ‚ (āˆ’āˆž, c] or I āŠ‚ (c, āˆž), this is trivial. In the case when I = (a, b) intersects both (āˆ’āˆž, c] and (c, āˆž), both sides of (1.34) can be computed by Lemmas 1.75 and 1.76, so (1.34) follows from the trivial Ī±āˆ’ (b) āˆ’ Ī±+ (a) = Ī±+ (c) āˆ’ Ī±+ (a) + Ī±āˆ’ (b) āˆ’ Ī±+ (c).



Lemma 1.78. For any c āˆˆ R, (c, āˆž) is measurable with respect to Ī¼āˆ— . Proof. Consider a set E āŠ‚ R and a countable cover of E by elementary sets, {Ij }āˆž j=1 . By Lemmas 1.75 and 1.77, āˆž 

Ļ(Ij ) =

j=1

āˆž  j=1

Ī¼āˆ— (Ij āˆ© (āˆ’āˆž, c]) +

āˆž 

Ī¼āˆ— (Ij āˆ© (c, āˆž)).

j=1

The sets Ij āˆ©(āˆ’āˆž, c] cover Eāˆ©(āˆ’āˆž, c] and the sets Ij āˆ©(c, āˆž) cover Eāˆ©(c, āˆž), so by Ļƒ-subadditivity of the outer measure, āˆž 

Ļ(Ij ) ā‰„ Ī¼āˆ— (E āˆ© (āˆ’āˆž, c]) + Ī¼āˆ— (E āˆ© (c, āˆž)).

j=1

Taking the inļ¬mum over all countable covers {Ij }āˆž j=1 gives Ī¼āˆ— (E) ā‰„ Ī¼āˆ— (E āˆ© (āˆ’āˆž, c]) + Ī¼āˆ— (E āˆ© (c, āˆž)).

(1.35)

The opposite inequality holds by subadditivity. Thus, equality holds in  (1.35) for any E āŠ‚ R, so (c, āˆž) is measurable with respect to Ī¼āˆ— . Proof of Theorem 1.73. By CarathĀ“eodoryā€™s theorem, the set A of measurable sets with respect to Ī¼āˆ— is a Ļƒ-algebra and Ī¼āˆ— is a measure on A. Since A contains all intervals of the form (c, āˆž), it contains BR . Therefore, the restriction of Ī¼āˆ— to BR is a Borel measure, denoted by Ī¼Ī± . Right-continuity  of Ī± gives Ī±+ = Ī±, so by Lemma 1.76, Ī¼Ī± has distribution function Ī±. For the function Ī±(x) = x, the corresponding measure is called the (onedimensional) Lebesgue measure and is denoted by m or m1 . We warn the reader that Lebesgue measure of a set is not tightly related to its cardinality (there exist uncountable sets of zero Lebesgue measure, e.g., the middle third Cantor set), or to its topological properties (there exist sets with empty interior but positive Lebesgue measure; Exercise 1.16).

1.5. Lebesgueā€“Stieltjes measures on R

29

Integration with respect to Lebesgue measure generalizes Riemann integration (Exercise 1.17), so integration with respect to Lebesgue measure is  b commonly denoted by a f (x) dx := [a,b] f dm for f āˆˆ L1 ([a, b], dm). The construction of Lebesgueā€“Stieltjes measures is complemented by an important uniqueness result: Theorem 1.79. If two Borel measures on R have the same distribution function, they are equal. The usual strategy suggests that, for two such measures Ī¼, Ī½, we should prove that S = {E āˆˆ BR | Ī¼(E) = Ī½(E)} is a Ļƒ-algebra. However, knowing for some sets En āˆˆ BR does not allow us to compare that Ī¼(En ) = Ī½(En )  āˆž Ī¼ ( n=1 En ) with Ī½ ( āˆž n=1 En ), since we cannot compute the measures of the unions. We notice the mismatch between the conditions for a Ļƒ-algebra, which must be closed under all countable unions, and Ļƒ-additivity for a measure, which only says something for disjoint countable unions. This is precisely the kind of obstacle for which monotone classes are needed. Proof. We use the family of left-open intervals, J = {(a, b] | a āˆˆ R āˆŖ {āˆ’āˆž}, b āˆˆ R, a < b} āˆŖ {(a, +āˆž) | a āˆˆ R āˆŖ {āˆ’āˆž}}, and the family of their ļ¬nite disjoint unions, G = {āˆ…} āˆŖ {

n 

Ij | n āˆˆ N, Ij āˆˆ J for all j, Ij āˆ© Ik = āˆ… if j = k}.

j=1

Since G contains all half-lines BR , it generates the Ļƒ-algebra. But unlike the family of half-lines, the family G is an algebra. To prove this, observe that I āˆˆ J implies I c āˆˆ G and that I1 , I2 āˆˆ J implies I1 āˆ© I2 āˆˆ G; thus, E āˆˆ G implies E c āˆˆ G and E, F āˆˆ G implies E āˆ© F āˆˆ G. Fix k āˆˆ N and consider the set Ck = {E āˆˆ BR | Ī¼(E āˆ© (āˆ’k, k]) = Ī½(E āˆ© (āˆ’k, k])}. If E āˆˆ G, then E āˆ© (āˆ’k, k] is a ļ¬nite disjoint union of intervals (aj , bj ] with āˆ’k ā‰¤ aj < bj ā‰¤ k. Since Ī¼((aj , bj ]) = Ī½((aj , bj ]), by additivity, E āˆˆ Ck . in Ck and let E be Let (En )āˆž n=1  be an increasing or decreasing sequence its limit, E = En if En is increasing and E = En if En is decreasing. Since Ī¼((āˆ’k, k]) = Ī½((āˆ’k, k]) < āˆž, by dominated convergence with the dominating function Ļ‡(āˆ’k,k] , Ī¼(E āˆ© (āˆ’k, k]) = lim Ī¼(En āˆ© (āˆ’k, k]) = lim Ī½(En āˆ© (āˆ’k, k]) = Ī½(E āˆ© (āˆ’k, k]), nā†’āˆž

nā†’āˆž

so E āˆˆ Ck . Thus, Ck is a monotone class and G āŠ‚ Ck āŠ‚ BR . By the monotone class theorem, Ck = BR .

30

1. Measure theory

Thus, for all E āˆˆ BR and k āˆˆ N, Ī¼(E āˆ© (āˆ’k, k]) = Ī½(E āˆ© (āˆ’k, k]). By monotone convergence, the limit k ā†’ āˆž gives Ī¼(E) = Ī½(E). 

1.6. Product measures A metric space is called Ļƒ-compact if it can be written as a countable union of compact sets. We give a quick construction of product measures on Ļƒcompact product spaces. Theorem 1.80. If Ī¼, Ī½ are measures on Ļƒ-compact spaces X, Y and Ī¼, Ī½ are ļ¬nite on compacts, then for every Borel function f : X Ɨ Y ā†’ [0, āˆž] the following hold. (a) For every y āˆˆ Y , the function X ā†’ [0, āˆž], x ā†’ f (x, y) is Borel.  (b) The function Y ā†’ [0, āˆž], y ā†’ f (x, y) dĪ¼(x) is Borel. (c) For every x āˆˆ X, the function Y ā†’ [0, āˆž], y ā†’ f (x, y) is Borel.  (d) The function X ā†’ [0, āˆž], x ā†’ f (x, y) dĪ½(y) is Borel. (e) Iterated integrals of f are independent of order of integration, i.e.,   f (x, y) dĪ¼(x) dĪ½(y) = f (x, y) dĪ½(y) dĪ¼(x). Proof. Denote by M the class of all Borel functions f : X Ɨ Y ā†’ [0, āˆž] with the desired properties. If f, g āˆˆ M, then f + g āˆˆ M by additivity of integrals. Moreover, if fn are a pointwise increasing sequence of functions in M, then the pointwise limit f = limnā†’āˆž fn is also in M by monotone convergence and because pointwise limits of Borel functions are Borel. We will use these observations repeatedly. We call a rectangle a set R = A Ɨ B where A āˆˆ BX , B āˆˆ BY . For any rectangle R, Ļ‡R āˆˆ M by a straightforward veriļ¬cation. Consider the family of ļ¬nite disjoint unions of rectangles, āŽ§ āŽ« n āŽØ āŽ¬ G = {āˆ…} āˆŖ Rj | Rj are rectangles, Rj āˆ© Rk = āˆ… if j = k . āŽ© āŽ­ j=1

By additivity, E āˆˆ G implies Ļ‡E āˆˆ M. Note also that G is an algebra in X Ɨ Y because the intersection of two rectangles is a rectangle and the complement of a rectangle is a disjoint union of rectangles, (A Ɨ B)c = (Ac Ɨ B) āˆŖ (A Ɨ B c ) āˆŖ (Ac Ɨ B c ). Fix compacts K āŠ‚ X and L āŠ‚ Y and denote C = {F āˆˆ BXƗY | Ļ‡F Ļ‡KƗL āˆˆ M}.

1.6. Product measures

31

Since F āˆˆ G implies F āˆ© (K Ɨ L) āˆˆ G, it follows that G āŠ‚ C. Note also that F āˆˆ C implies F c āˆˆ C by subtracting from Ļ‡KƗL (there are no inļ¬nities in that subtraction, so this is an algebraic veriļ¬cation). Let us prove that C is a  monotone class. If (Fn )āˆž n=1 is an increasing sequence in C and F = āˆž F , the pointwise increasing sequence n=1 n Ļ‡Fn Ļ‡KƗL ā†’ Ļ‡F Ļ‡KƗL implies that F āˆˆ C. For a decreasing sequence (Fn )āˆž n=1 , passing to complements reduces to increasing sequences. By the monotone class theorem, C contains the Ļƒ-algebra generated by G. Since compact metric spaces are separable, so are Ļƒ-compact metric spaces. Thus, they are second countable (Exercise 1.2). By Lemma 1.33, X Ɨ Y has a countable base consisting of rectangles, so C contains all open sets, and therefore BXƗY . Since X, Y are Ļƒ-compact, they have countable covers by compacts Kn , Ln , respectively. Since ļ¬nite unions of compact sets are compact, we can assume sequences Kn , Ln to be increasing. For any F āˆˆ BXƗY and n āˆˆ N, Ļ‡F Ļ‡Kn ƗLn āˆˆ M. Taking the increasing limit as n ā†’ āˆž gives Ļ‡F āˆˆ M. By additivity, M contains all simple functions. Any positive Borel function is the pointwise limit of an increasing sequence of simple functions, so M contains all positive functions.  Deļ¬nition 1.81. In the setting of the previous theorem, the product measure Ī¼ āŠ— Ī½ is the Borel measure deļ¬ned by  (Ī¼ āŠ— Ī½)(E) = Ļ‡E (x, y) dĪ¼(x) dĪ½(y). This is indeed a measure: it is Ļƒ-additive because it is additive (by additivity of integrals) and because monotone convergence can be used to move the limit inside the iterated integrals. For instance, from Lebesgue measure m1 = m on R, we inductively deļ¬ne n-dimensional Lebesgue measure mn = mnāˆ’1 āŠ— m on BRn . Theorem 1.82 (Tonelli). Assume that X, Y are Ļƒ-compact and Ī¼, Ī½ are ļ¬nite on compacts. For any Borel function f : X Ɨ Y ā†’ [0, āˆž],    f d(Ī¼ āŠ— Ī½) = f (x, y) dĪ¼(x) dĪ½(y) = f (x, y) dĪ½(y) dĪ¼(x). (1.36) Proof. By Theorem 1.80 and the deļ¬nition of Ī¼ āŠ— Ī½, (1.36) holds if f = Ļ‡E for some set E āˆˆ BXƗY . By linearity, (1.36) holds for all simple functions. Taking an increasing limit of simple functions and using monotone convergence, we conclude that (1.36) holds for all positive Borel functions. 

32

1. Measure theory



For complex-valued f , applying Tonelliā€™s theorem to |f | gives   |f | d(Ī¼āŠ—Ī½) = |f (x, y)| dĪ¼(x) dĪ½(y) = |f (x, y)| dĪ½(y) dĪ¼(x), (1.37)

so we can check whether f is integrable by computing iterated integrals. This is often checked in order to apply the following theorem: Theorem 1.83 (Fubini). Assume that X, Y are Ļƒ-compact and Ī¼, Ī½ are ļ¬nite on compacts. For any f āˆˆ L1 (X Ɨ Y, d(Ī¼ āŠ— Ī½)) the following hold. (a) For Ī½-a.e. y, the function x ā†’ f (x, y) is in L1 (X, dĪ¼). (b) For Ī¼-a.e. x, the function y ā†’ f (x, y) is in L1 (Y, dĪ½). (c) Equation (1.36) holds, with the interpretation that the inner integrals are well deļ¬ned a.e., and ignoring the exceptional zeromeasure sets, the outer integrals give the stated value.  Proof. From (1.37), it follows that |f (x, y)| dĪ½(y) < āˆž for Ī¼-a.e. x and |f (x, y)| dĪ¼(x) < āˆž for Ī½-a.e. y. The proof follows from Tonelliā€™s theorem in the usual way, by passing from positive to real-valued and then to complex-valued functions, using linearity of integrals. 

1.7. Functions on Ļƒ-locally compact spaces In this section, we begin to use continuous functions as approximants and test functions. The main result of this section is a kind of approximation of bounded Borel functions by continuous functions, which gives a new way of proving that certain statements hold for all bounded Borel functions. In particular, we will use it for the study of the Borel functional calculus for self-adjoint operators. In order to work with continuous functions on X, we impose some topological assumptions on X. The following class suļ¬ƒces for our purposes: Deļ¬nition 1.84. A metric space X is Ļƒ-locally compact if it has compact āˆž subsets Ln āŠ‚ X such that Ln āŠ‚ int Ln+1 for all n āˆˆ N and n=1 Ln = X. Any such sequence (Ln )āˆž n=1 is called an exhaustion of X by compact sets.  Not every compact sequence Kn āŠ‚ Kn+1 with āˆž n=1 Kn = X gives an exhaustion of X by compact sets; a counterexample in X = R is given by Kn = [āˆ’n, 0] āˆŖ [1/n, n]. There exist metric spaces which are Ļƒ-compact but not Ļƒ-locally compact (Exercise 1.20). In fact, Ļƒ-local compactness can be seen as a combination of separability and a local condition (Exercise 1.21). However, many common spaces are Ļƒ-locally compact: Example 1.85. Any countable space with the discrete metric is Ļƒ-locally compact. To obtain an exhaustion by compact sets, choose an enumeration of the space X = {xn | n āˆˆ N} and set Ln = {x1 , . . . , xn }.

1.7. Functions on Ļƒ-locally compact spaces

33

Example 1.86. For any k āˆˆ N, Rk is Ļƒ-locally compact with Ln = [āˆ’n, n]k . Lemma 1.87. On any Ļƒ-compact space, the Borel Ļƒ-algebra is generated by the family of compact subsets. Proof. Denote by A the Ļƒ-algebra generated by compact subsets of the space X. Compact sets are closed, so they are Borel  sets. Thus, A āŠ‚ BX . For any closed Conversely, let Ln be compact sets such that X = āˆž n=1 Ln . F āŠ‚ X, the sets F āˆ©Ln are compact, so F āˆ©Ln āˆˆ A; thus F = āˆž n=1 (F āˆ©Ln ) āˆˆ A. Since A contains all closed sets; passing to complements, A contains all  open sets, so BX āŠ‚ A. To proceed, we need some separation facts which are easily proved in our metric space setting. Distance between points and sets is deļ¬ned by d(x, B) = inf d(x, y). yāˆˆB

This is a continuous function of x because |d(x, y) āˆ’ d(x , y)| ā‰¤ d(x, x ) implies |d(x, B) āˆ’ d(x , B)| ā‰¤ d(x, x ). Similarly, for A, B āŠ‚ X, we denote d(A, B) = inf inf d(x, y). xāˆˆA yāˆˆB

Lemma 1.88. If K is compact, V open, and K āŠ‚ V , then d(K, V c ) > 0. Proof. The function d(x, V c ) is continuous in x and strictly positive on the open set V , so it has a strictly positive minimum on the compact K.  The support of a continuous function f : X ā†’ C is deļ¬ned as supp f = {x āˆˆ X | f (x) > 0}. This is a closed set; we denote by Cc (X) the set of continuous f : X ā†’ C such that supp f is compact. Note that f āˆˆ Cc (X) if and only if there exists a compact K āŠ‚ X such that f (x) = 0 for all x āˆˆ / K, and that Cc (X) is a vector space. The following lemma separates sets by a function f āˆˆ Cc (X): Lemma 1.89. In a Ļƒ-locally compact metric space X the following hold. (a) For any compact K, there exists Ī“ > 0 such that {x āˆˆ K | d(x, K) ā‰¤ Ī“} is compact. (b) If K is compact, V open, and K āŠ‚ V , then there exists f āˆˆ Cc (X) such that Ļ‡K ā‰¤ f ā‰¤ 1 and supp f āŠ‚ V . Proof. (a) Let (Ln )āˆž n=1 be an exhaustion of X by compact sets. Since Ln āŠ‚ int Ln+1 , the sets int Ln+1 are an open cover of K. There is a ļ¬nite subcover int Lm1 , . . . , int Lmk , and taking m = max{m1 , . . . , mk } gives K āŠ‚ int Lm . Thus, for Ī“ = d(K, (int Lm )c ) > 0, the set {x āˆˆ K | d(x, K) ā‰¤ Ī“} is compact as a closed subset of the compact Lm .

34

1. Measure theory

(b) If < Ī“ and < d(K, V c ), the function f (x) = (1 āˆ’ āˆ’1 d(x, K))+ has compact support and supp f āŠ‚ V .  We now consider the family of bounded Borel functions, its algebraic properties, and a useful notion of convergence: Deļ¬nition 1.90. Denote by Bb (X) the set of bounded Borel functions from X to C. A subset M āŠ‚ Bb (X) is said to be a subalgebra of Bb (X) if it contains the constant function 1 and is closed under scalar multiplication, pointwise addition, and pointwise multiplication. M is said to be closed under pointwise convergence of uniformly bounded sequences if, for any sequence of gn āˆˆ M, such that sup sup |gn (x)| < āˆž

nāˆˆN xāˆˆX

and the limit g(x) = limnā†’āˆž gn (x) is convergent for all x āˆˆ X, it follows that g āˆˆ M. We emphasize that this notion of convergence does not correspond to any metric, and that we are not working with respect to any measure or any kind of almost-everywhere condition. This makes the current setting diļ¬€erent from, say, that in Chapter 2, where some density properties in L1 (X, dĪ¼) will be considered. This distinction will be essential when there is no a priori distinguished measure that can be used. Lemma 1.91. If M is a subalgebra of Bb (X) closed under pointwise convergence of uniformly bounded sequences, {A āˆˆ BX | Ļ‡A āˆˆ M} is a Ļƒ-algebra. Proof. Denote A = {A āˆˆ BX | Ļ‡A āˆˆ M}. Since Ļ‡āˆ… = 0 āˆˆ M, āˆ… āˆˆ A. If A āˆˆ A, then Ļ‡Ac = 1 āˆ’ Ļ‡A āˆˆ M, so Ac āˆˆ M. If A, B āˆˆ A, then Ļ‡Aāˆ©B = Ļ‡A Ļ‡B āˆˆ M , so A āˆ© B āˆˆ A. Thus, A is an algebra. For any sequence of sets An āˆˆ A, the uniformly bounded pointwise limit

shows that

āˆž

j=1 Aj

= lim Ļ‡nj=1 Aj Ļ‡āˆž j=1 Aj nā†’āˆž

āˆˆ A, so A is a Ļƒ-algebra.



Proposition 1.92. Let X be a Ļƒ-locally compact metric space and let M be a subalgebra of Bb (X). If M is closed under pointwise convergence of uniformly bounded sequences, the following are equivalent: (a) Cc (X) āŠ‚ M; (b) Ļ‡B āˆˆ M for all Borel sets B āŠ‚ X; (c) M = Bb (X). Proof. (a) =ā‡’ (b): For any compact K āŠ‚ X, the functions fn (x) = (1 āˆ’ nd(x, K))+ are uniformly bounded and converge pointwise to Ļ‡K , with

1.8. Regularity of measures

35

fn āˆˆ Cc (X) for large enough n; thus, Ļ‡K āˆˆ M. Thus, the Ļƒ-algebra {A āˆˆ BX | Ļ‡A āˆˆ M} contains all compact sets, so it contains all Borel sets. (b) =ā‡’ (c): Since M is an algebra and contains all characteristic functions of Borel sets, M contains all simple functions (functions which take ļ¬nitely many values). Any positive Borel function f bounded by C āˆˆ N is the pointwise limit of the uniformly bounded functions fn =

n C2 

k=0

k Ļ‡ k k+1 , 2n {x| 2n ā‰¤f (x)< 2n }

so M contains all positive bounded Borel functions. By linear combinations, we obtain all complex-valued bounded Borel functions, so M = Bb (X). 

(c) =ā‡’ (a): This is trivial.

Pointwise convergence does not correspond to convergence with respect to a metric, so intuition from metric spaces cannot be applied. The smallest subalgebra of Bb (X), which contains Cc (X) and is closed under pointwise convergence of uniformly bounded sequences, is not the set of limit points of Cc (X). Despite Proposition 1.92, not every bounded Borel function is a pointwise limit of a uniformly bounded sequence of continuous functions (Exercise 1.22).

1.8. Regularity of measures Since Borel sets are deļ¬ned somewhat implicitly, it is of interest to know how well they can be approximated by open and closed sets, and how well their measures can be approximated by integrals of continuous functions. Theorem 1.93. Let Ī¼ be a ļ¬nite Borel measure on a metric space X. For any Borel set E and > 0, there exist closed F and open V with F āŠ‚ E āŠ‚ V such that Ī¼(V \ F ) < . Proof. We will prove that the family A = {E āˆˆ BX | āˆ€ > 0 āˆƒF closed āˆƒV open

F āŠ‚ E āŠ‚ V, Ī¼(V \ F ) < }

is a Ļƒ-algebra. Trivially, āˆ… āˆˆ A, by taking F = V = āˆ…. If E āˆˆ A, then F āŠ‚ E āŠ‚ V gives V c āŠ‚ E c āŠ‚ F c with F c \ V c = V \ F , so E c āˆˆ A.  Let En āˆˆ A, and denote E = nāˆˆN En . For any > 0, there exist closed āŠ‚ En āŠ‚ Vn with Ī¼(Vn \ Fn ) < /2n+1 . Thus, Fn and open Vnsuch that Fn  by taking V = nāˆˆN Vn , A = nāˆˆN Fn , we have A āŠ‚ E āŠ‚ V and Ī¼(V \ A) ā‰¤

āˆž  n=1

Ī¼(Vn \ Fn ) < /2.

36

1. Measure theory

The set V is open. n The set A is not closed, but it is the increasing limit of the closed sets j=1 Fj , so using ļ¬niteness of measure, for some n āˆˆ N, āŽž āŽ› n  Ī¼ āŽA \ Fj āŽ  < /2. Thus, with F =

n

j=1 Fj ,

j=1

we have F āŠ‚ E āŠ‚ V and Ī¼(V \ F ) < , so E āˆˆ A.

If E is a closed set, the sets Vn = {x āˆˆ X | d(x, E) < 1/n} obey E āŠ‚ Vn+1 āŠ‚ Vn and (Vn \ E) = Vn \ E = E \ E = āˆ… nāˆˆN

nāˆˆN

because E is closed. Since the sequence of sets Vn \ E are decreasing and have ļ¬nite measure, this implies Ī¼(Vn \ E) ā†’ 0 as n ā†’ āˆž. Thus, choosing F = E and V = Vn for large enough n implies E āˆˆ A. Thus, A āŠ‚ BX is a  Ļƒ-algebra and contains all closed sets, so A = BX . We often must work with inļ¬nite measures; the counting measure on N and the Lebesgue measure on R are just two examples. However, inļ¬nities on compacts introduce unnatural obstacles (Exercise 1.19), so we deļ¬ne: Deļ¬nition 1.94. A Baire measure on a Ļƒ-locally compact metric space X is a Borel measure Ī¼ such that Ī¼(K) < āˆž for all compact K āŠ‚ X. Baire measures on R are precisely the Lebesgueā€“Stieltjes measures, by Theorems 1.73 and 1.79. Baire measures are usually deļ¬ned on more general spaces, on the Ļƒ-algebra generated by compact sets; by Lemma 1.87, in our level of generality, this matches our deļ¬nition. Deļ¬nition 1.95. A Borel measure Ī¼ is said to be inner regular if Ī¼(A) =

sup

Ī¼(K)

KāŠ‚A K compact

for all Borel sets A, and outer regular if Ī¼(A) = inf Ī¼(V ) V āŠƒA V open

for all Borel sets A. If Ī¼ is inner regular and outer regular, it is said to be regular. Theorem 1.96. On a Ļƒ-locally compact metric space, every Baire measure is regular. Proof. We ļ¬x an exhaustion (Ln )āˆž n=1 of X by compact sets (see Deļ¬nition 1.84), the convention L0 = āˆ…, and the ļ¬nite measures Ī¼n (E) = Ī¼(E āˆ© Ln ).

1.8. Regularity of measures

37

We decompose a Borel set A as a disjoint union of sets An = A āˆ© (int Ln \ int Lnāˆ’1 ). Fix > 0. For any n, since Ī¼n is a ļ¬nite measure, there exist closed Fn and open Vn such that Fn āŠ‚ An āŠ‚ Vn and Ī¼n (Vn \ Fn ) < /2n . Without loss of generality we can replace Vn by Vn āˆ© int Ln ; then, by the set of inclusions Fn āŠ‚ Vn āŠ‚ int Ln and the deļ¬nition of Ī¼n , we also conclude

Ī¼(Vn \ Fn ) < n . 2   Thus, deļ¬ning F = nāˆˆN Fn and V = nāˆˆN Vn gives F āŠ‚ A āŠ‚ V and Ī¼(V \ F ) < . This implies Ī¼(F ) ā‰„ Ī¼(A) āˆ’ and Ī¼(V ) ā‰¤ Ī¼(A) + . Since > 0 is arbitrary, sup Ī¼(F ) ā‰„ Ī¼(A) ā‰„ inf Ī¼(V ). V open AāŠ‚V

F closed F āŠ‚A

Since by monotone convergence limnā†’āˆž Ī¼(F āˆ© Ln ) = Ī¼(F ), using compacts K = F āˆ© Ln for large enough n shows that sup K compact KāŠ‚A

Ī¼(K) ā‰„ Ī¼(A) ā‰„ inf Ī¼(V ). V open AāŠ‚V

The opposite inequalities are trivial since K āŠ‚ A āŠ‚ V implies Ī¼(K) ā‰¤ Ī¼(A) ā‰¤ Ī¼(V ).  Lemma 1.97. If Ī¼ is a Baire measure on a Ļƒ-locally compact metric space X, then Cc (X) āŠ‚ L1 (X, dĪ¼). Proof. For f āˆˆ Cc (X), consider the compactK = supp f and the maximum  a = maxxāˆˆK |f (x)|. Then 0 ā‰¤ |f | ā‰¤ aĻ‡K , so |f | dĪ¼ ā‰¤ aĪ¼(K) < āˆž. An outer regular measure can be recovered from its values on open sets. Further, it is useful to know when a measure can be completely recovered from integrals of functions in Cc (X). For an open set V , we deļ¬ne FV = {f āˆˆ Cc (X) | 0 ā‰¤ f ā‰¤ 1, supp f āŠ‚ V }. Proposition 1.98. For any open set V ,  f dĪ¼. Ī¼(V ) = sup f āˆˆFV

(1.38)

(1.39)

Proof. For any compact F āŠ‚ V , by Lemma 1.89, there exists f āˆˆ FV such  that Ļ‡F ā‰¤ f ā‰¤ 1 and therefore Ī¼(F ) ā‰¤ f dĪ¼. Taking the supremum over compacts gives, by inner regularity,  f dĪ¼. Ī¼(V ) = sup Ī¼(F ) ā‰¤ sup F compact F āŠ‚V

f āˆˆFV

The opposite inequality follows from f ā‰¤ Ļ‡V for all f āˆˆ FV .



38

1. Measure theory

Thus, the integrals of functions f āˆˆ Cc (X) determine the measure on open sets by (1.39) and then on all Borel sets by outer regularity.

1.9. The Rieszā€“Markov theorem We have seen constructions of measures that were geometrically motivated by the concept of length on the real line, and area/volume on Rn . In more abstract situations, measures often appear because of how they act on functions rather than sets (this will be the case for spectral measures as well). In other words, instead of a measure, we usually ļ¬rst encounter a functional: Deļ¬nition 1.99. Let X be a metric space. A positive linear functional on Cc (X) is a linear map Ī› : Cc (X) ā†’ C such that f ā‰„ 0 implies Ī›(f ) ā‰„ 0. Any Baire measure Ī¼ on X generates a positive linear functional  Ī›(f ) = f dĪ¼ āˆ€f āˆˆ Cc (X), (1.40) and the goal of this section is to prove the converse: Theorem 1.100 (Rieszā€“Markov). Let X be a Ļƒ-locally compact metric space. For every positive linear functional Ī› on Cc (X), there is a unique Baire measure Ī¼ on X such that (1.40) holds. We assume throughout this section that X is a Ļƒ-locally compact metric space and Ī› is a positive linear functional. Uniqueness of Ī¼ follows from outer regularity of Ī¼ and from Proposition 1.98. Existence of Ī¼ will be proved through a series of lemmas. It uses the outer measure construction, with open sets V as elementary sets, and the weight Ļ(V ) = sup Ī›(f ), f āˆˆFV

where FV is deļ¬ned by (1.38); of course, Fāˆ… = {0} and Ļ(āˆ…) = 0. We will prove Ļƒ-subadditivity of Ļ, using the following reļ¬nement of compactness: Lemma 1.101 (Continuous partitions of unity). If K āŠ‚ X is compact, for any open cover V of K, there exists a ļ¬nite subcover V1 , . . . , Vn and functions h1 , . . . , hn āˆˆ Cc (X) such that hj ā‰„ 0, supp hj āŠ‚ Vj , and n 

hj (x) = 1

āˆ€x āˆˆ K.

(1.41)

j=1

Proof. For every y āˆˆ K, choose Vy āˆˆ V such that y āˆˆ Vy . By Lemma 1.89 applied to {y} āŠ‚ Vy , for small enough > 0, the function gy (x) = (1 āˆ’ āˆ’1 d(x, y))+ is in Cc (X) and obeys supp gy āŠ‚ Vy .

1.9. The Rieszā€“Markov theorem

39

Since gy (y) = 1, the set Uy = {x | gy (x) > 0} is open and contains y. Thus, the family {Uy | y āˆˆ K} is an open cover of K. By compactness, this cover has a ļ¬nite subcover Uy1 , . . . , Uyn . Since K āŠ‚ nj=1 Uyj , n 

āˆ€x āˆˆ K.

gyi (x) > 0

i=1

Moreover, let G(x) = d(x, K). Then the functions gy nj hj = G + k=1 gyk are well deļ¬ned, supp hj = supp gj āŠ‚ Vyj for each j, and (1.41) holds, so the  proof is complete with the ļ¬nite subcover {Vy1 , . . . , Vyn } of V. Lemma 1.102. For any open sets Vj , āˆž  āˆž   Vj ā‰¤ Ļ(Vj ). Ļ j=1

(1.42)

j=1

, let K = supp f . Since {Vj }āˆž Proof. If f āˆˆ Fāˆž j=1 are an open cover j=1 Vj of K, by Lemma n 1.101, there exists n āˆˆ N and h1 , . . . , hn āˆˆ Cc (X) such that hj ā‰„ 0, j=1 hj = 1 on K, and supp hj āŠ‚ Vj for each j. Then f = nj=1 f hj , so Ī›(f ) =

n 

Ī›(f hj ) ā‰¤

j=1

n 

Ļ(Vj ) ā‰¤

j=1

āˆž 

Ļ(Vj ).

j=1

gives (1.42). Taking the supremum over all f āˆˆ Fāˆž j=1 Vj



Lemma 1.103. For any set E āŠ‚ X, we deļ¬ne Ī¼āˆ— (E) = inf Ļ(V ). EāŠ‚V V open

(1.43)

Then Ī¼āˆ— is an outer measure on X. Proof. Obviously, Ī¼āˆ— (āˆ…) = Ļ(āˆ…) = 0 and A āŠ‚ B implies Ī¼āˆ— (A) ā‰¤ Ī¼āˆ— (B). Take any sequence of sets Ej āŠ‚ X. For any > 0, thereexist open sets Vj such that Ej āŠ‚ Vj and Ļ(Vj ) ā‰¤ Ī¼āˆ— (Ej ) + /2j . Then V = jāˆˆN Vj is open  and jāˆˆN Ej āŠ‚ V , so by Ļƒ-subadditivity of Ļ, āˆž  āˆž āˆž    Vj ā‰¤ Ļ(Vj ) ā‰¤ Ī¼āˆ— (Ej ) + . Ī¼āˆ— (E) ā‰¤ Ļ j=1

j=1

j=1

Since > 0 is arbitrary, this implies Ļƒ-subadditivity of Ī¼āˆ— . Lemma 1.104. All open sets are measurable with respect to Ī¼āˆ— .



40

1. Measure theory

Proof. Fix an open set V and arbitrary E āŠ‚ X. It suļ¬ƒces to prove that Ī¼āˆ— (E) ā‰„ Ī¼āˆ— (E āˆ© V ) + Ī¼āˆ— (E \ V )

(1.44)

because the opposite inequality follows from subadditivity. Moreover, we can assume Ī¼āˆ— (E) < āˆž, otherwise, the inequality is trivial. Take an open set U such that E āŠ‚ U , and let > 0. Since U āˆ© V is open, there exists f āˆˆ FU āˆ©V such that Ī›(f ) ā‰„ Ļ(U āˆ© V ) āˆ’ , and since U \ supp f is open, there exists g āˆˆ FU \supp f such that Ī›(g) ā‰„ Ļ(U \ supp f ) āˆ’ . Since Ļ(U āˆ© V ) ā‰„ Ī¼āˆ— (E āˆ© V ) and Ļ(U \ supp f ) ā‰„ Ī¼āˆ— (E \ V ), this gives Ī›(f ) + Ī›(g) ā‰„ Ī¼āˆ— (E āˆ© V ) + Ī¼āˆ— (E \ V ) āˆ’ 2 . Note that supp f āˆ© supp g = āˆ…, so f + g āˆˆ FU . This and additivity of Ī› imply Ļ(U ) ā‰„ Ī›(f + g) ā‰„ Ī¼āˆ— (E āˆ© V ) + Ī¼āˆ— (E \ V ) āˆ’ 2 . Since > 0 is arbitrary and U is arbitrary with E āŠ‚ U , (1.44) follows.



By CarathĀ“eodoryā€™s Theorem 1.26, the restriction of Ī¼āˆ— to BX is a Borel measure on X, which we denote Ī¼ from now on. Lemma 1.105. For any f āˆˆ Cc (X), if for some compact K and open V , Ļ‡K ā‰¤ f ā‰¤ Ļ‡V , then Ī¼(K) ā‰¤ Ī›(f ) ā‰¤ Ī¼(V ). Proof. For t āˆˆ (0, 1), deļ¬ne Vt = {x | f (x) > t}. Then g āˆˆ FVt implies tg ā‰¤ f so Ī›(g) ā‰¤ tāˆ’1 Ī›(f ). Taking the supremum over g āˆˆ FVt gives Ī¼(Vt ) ā‰¤ tāˆ’1 Ī›(f ). These sets have ļ¬nite measure; the limit as t ā†’ 1 gives Ī¼({x | f (x) ā‰„ 1}) ā‰¤ Ī›(f ). Therefore, Ī¼(K) ā‰¤ Ī›(f ). In particular, Ī¼(K) < āˆž. Let us ļ¬x g āˆˆ Cc (X) such that 0 ā‰¤ g ā‰¤ 1 and g = 1 on supp f . For any t > 0, (f āˆ’ t)+ āˆˆ FV , so Ī›((f āˆ’ t)+ ) ā‰¤ Ī¼(V ). Since f ā‰¤ tg + (f āˆ’ t)+ , Ī›(f ) ā‰¤ tĪ›(g) + Ī›((f āˆ’ t)+ ) ā‰¤ tĪ›(g) + Ī¼(V ). Since t > 0 is arbitrary, this implies Ī›(f ) ā‰¤ Ī¼(V ).



Taking V = X, we see that Ī¼(K) < āˆž for any compact K. Thus, every f āˆˆ Cc (X) is integrable with respect to Ī¼, and it remains to prove that the integral is Ī›(f ):  Lemma 1.106. For every f āˆˆ Cc (X), f dĪ¼ = Ī›(f ). Proof. By linearity of both sides, it suļ¬ƒces to prove this for f such that 0 ā‰¤ f ā‰¤ 1. Fix n āˆˆ N and deļ¬ne sets Ak = f āˆ’1 ((k/n, āˆž)) and functions ! ! kāˆ’1 k āˆ’ fāˆ’ . gk = f āˆ’ n n + +

1.10. Exercises

41

These functions obey f = nk=1 gk and Ļ‡Ak ā‰¤ ngk ā‰¤ Ļ‡Akāˆ’1 for each k. We will use this in two ways: Integrating in Ī¼ gives  Ī¼(Ak ) ā‰¤ gk dĪ¼ ā‰¤ Ī¼(Akāˆ’1 ), whereas applying Lemma 1.105 gives Ī¼(Ak ) ā‰¤ Ī›(gk ) ā‰¤ Ī¼(Akāˆ’1 ). Averaging in k and using linearity of Ī› gives  n n 1 1 Ī¼(Ak ) ā‰¤ f dĪ¼ ā‰¤ Ī¼(Akāˆ’1 ), n n k=1

1 1 Ī¼(Ak ) ā‰¤ Ī›(f ) ā‰¤ n n n

k=1

k=1 n 

Ī¼(Akāˆ’1 ).

(1.45) (1.46)

k=1

Note that An = āˆ… and A0 = supp Ī¼ has ļ¬nite measure, so (1.45) and (1.46) place the values f dĪ¼ and Ī›(f ) within the same interval of length n n n n 1 1 1 Ī¼(A0 ) 1 . Ī¼(Akāˆ’1 ) āˆ’ Ī¼(Ak ) ā‰¤ Ī¼(Akāˆ’1 ) āˆ’ Ī¼(Ak ) = n n n n n k=1 k=1 k=1 k=1    Thus,  f dĪ¼ āˆ’ Ī›(f ) ā‰¤ Ī¼(A0 )/n. Taking n ā†’ āˆž shows f dĪ¼ = Ī›(f ). 

1.10. Exercises 1.1. Let X be a metric space. View E āˆˆ BX , E = āˆ…, as a metric subspace of X. Prove that {A āˆˆ BX | A āŠ‚ E} = {B āˆ© E | B āˆˆ BX } = BE . If f : X ā†’ Y is Borel, prove that f |E : E ā†’ Y is also Borel. 1.2. For any metric space X, prove that the following are equivalent: (a) X is separable; (b) X has a countable base; (c) any base U of X contains a countable base U  āŠ‚ U . 1.3. (a) If d, dĖœ are metrics on X and there exist a, b āˆˆ (0, āˆž) such that Ėœ y) ā‰¤ bd(x, y) ad(x, y) ā‰¤ d(x, āˆ€x, y āˆˆ X, prove that d and dĖœ generate the same topology in X. (b) Prove that the metric dāˆž in Rn deļ¬ned in (1.11) and the metrics  n 1/p  p (xj āˆ’ yj ) (1.47) dp (x, y) = j=1

for p āˆˆ [1, āˆž) obey dāˆž (x, y) ā‰¤ dp (x, y) ā‰¤ n1/p dāˆž (x, y), and conclude that they generate the same topology.

42

1. Measure theory

1.4. For a sequence (xn )āˆž n=1 in a metric space (X, d) and x āˆˆ X, prove that āˆ€ > 0 āˆƒN āˆˆ N āˆ€n ā‰„ N d(xn , x) <

(1.48)

if and only if āˆ€A āˆˆ Td x āˆˆ A =ā‡’ (āˆƒN āˆˆ N āˆ€n ā‰„ N xn āˆˆ A).

(1.49)

1.5. Prove that BR is generated by the sets [a, āˆž), a āˆˆ R. 1.6. Prove that any increasing function Ī± : R ā†’ R is Borel. 1.7. Provethat BRn is the smallest Ļƒ-algebra containing all sets of the form nj=1 (aj , āˆž) with a1 , . . . , an āˆˆ R. Ė† is Borel if and only if f āˆ’1 ({+āˆž}) āˆˆ 1.8. Prove that a function f : X ā†’ R āˆ’1 BX , f ({āˆ’āˆž}) āˆˆ BX , and f |E : E ā†’ R is Borel, where E = f āˆ’1 (R). 1.9. (a) Let f : X ā†’ R āˆŖ {+āˆž} be lower semicontinuous, i.e., for every x0 āˆˆ X, f (x0 ) ā‰¤ lim inf xā†’x0 f (x). Prove that f is Borel. Hint: Prove that f āˆ’1 ((a, āˆž]) is open for any a āˆˆ R. (b) Let f : X ā†’ R āˆŖ {āˆ’āˆž} be upper semicontinuous, i.e., for every x0 āˆˆ X, f (x0 ) ā‰„ lim supxā†’x0 f (x). Prove that f is Borel. 1.10. Let {Tn }āˆž n=1 be a partition of X into Borel sets, and let fn : Tn ā†’ Y be Borel functions for n āˆˆ N. Prove that the function f : X ā†’ Y , deļ¬ned by f (x) = fn (x) for x āˆˆ Tn for all n, is a Borel function. 1.11. For a sequence of Borel functions gn : X ā†’ R, let S be the set of points x āˆˆ X such that limnā†’āˆž gn (x) exists and is ļ¬nite. Prove that S is Borel and the function g : S ā†’ R deļ¬ned by g(x) = limnā†’āˆž gn (x) is Borel. 1.12. If Ī¼j , j āˆˆ N are Borel measures and cj āˆˆ [0, āˆž), prove that Ī¼ = āˆž j=1 cj Ī¼j is a Borel measure. Justify any exchanges of limits.   1.13. Let f āˆˆ L1 (X, dĪ¼). Prove that | f dĪ¼| = |f | dĪ¼ if and only if there exists Ļ‰ āˆˆ C with |Ļ‰| = 1 such that Ļ‰f = |f | Ī¼-a.e. 1.14. For a Borel measure Ī¼ on R with a distribution function Ī±, prove the following. (a) For any x āˆˆ R, Ī¼({x}) = Ī±+ (x) āˆ’ Ī±āˆ’ (x). (b) Ī± is continuous at x if and only if Ī¼({x}) = 0. (c) For any real x < y, Ī¼([x, y)) = Ī±āˆ’ (y) āˆ’ Ī±āˆ’ (x). (d) Ī¼((āˆ’āˆž, 0]) < āˆž if and only if Ī±+ (āˆ’āˆž) is ļ¬nite. 1.15. Let Ī¼ be a Borel measure on R with a distribution function Ī±. (a) Prove that any open set V āŠ‚ R can  be written as a countable disjoint union of open intervals, V = jāˆˆJ (aj , bj ). (b) Prove that Ī¼(V ) = jāˆˆJ (Ī±āˆ’ (bj ) āˆ’ Ī±+ (aj )).

1.10. Exercises

43

1.16. Prove that the set āˆž 2 

nāˆ’1

B=

n=1 k=1

1 2k āˆ’ 1 1 2k āˆ’ 1 āˆ’ 2n+1 , + 2n+1 n n 2 2 2 2

!

obeys B = [0, 1], m(B) < 1, and m(B āˆ© [a, b]) > 0 for any 0 ā‰¤ a < b ā‰¤ 1. 1.17. Prove the following link between Riemannā€“Stieltjes and Lebesgue integrals: for increasing right-continuous Ī± and continuous f : [0, 1] ā†’ R,  nāˆ’1  lim f (k/n)(Ī±((k + 1)/n) āˆ’ Ī±(k/n)) = f dĪ¼Ī± . nā†’āˆž

k=0

āˆž

(a,b]

and cn > 0 for n āˆˆ N. If n=1 cĪ±n < āˆž for some Ī± āˆˆ (0, 1), 1.18. Let xn āˆˆ R cn prove that āˆž n=1 |xāˆ’xn | < āˆž for Lebesgue-a.e. x āˆˆ R. cĪ± n Hint: Consider the set of x āˆˆ [āˆ’k, k], where āˆž n=1 |xāˆ’xn |Ī± < āˆž. 1.19. Consider the counting measure of Q viewed as a measure on R, Ī¼(A) = #(A āˆ© Q).  Prove that Ī¼(V ) = āˆž for every nonempty open set V , and that |f | dĪ¼ = āˆž for every continuous function f except for f = 0. 1.20. Prove that the metric space Q is Ļƒ-compact, but not Ļƒ-locally compact. 1.21. A metric space X is called locally compact if for every x āˆˆ X there is an open set V such that x āˆˆ V and V is compact. Prove that a metric space is Ļƒ-locally compact if and only if it is separable and locally compact. 1.22. Let fn be a uniformly bounded sequence of continuous real-valued functions which converges pointwise to Ļ‡B . Prove that the set B is a GĪ“ set, i.e., a countable intersection of open sets.

Chapter 2

Banach spaces

Banach spaces are, simply put, complete metric vector spaces whose metric behaves in a natural way with respect to the vector space operations. We will give the precise deļ¬nition below. In this chapter, we present the basic properties of Banach spaces and consider some concrete spaces of interest. Further treatments of functional analysis include [76, 81, 97]. In this text, virtually all vector spaces are over the ļ¬eld of scalars C. For the questions we study, this is not a limitation. Just as Rn can be viewed as a subset of Cn , the objects we study can be viewed as complex valued with no loss of generality (the only exception will be the proof of the Stoneā€“ Weierstrass theorem, which is naturally ļ¬rst proved on real-valued functions and then suitably generalized to complex-valued functions). We will freely use terminology and notation inherited from linear algebra. For instance, in any vector space V , a linear combination of vectors in X āŠ‚ V is a ļ¬nite sum n j=1 Ī»j xj , where Ī»j āˆˆ C and xj āˆˆ X, and the span of X, denoted span X, is the set of all linear combinations of vectors in X.

2.1. Norms and Banach spaces Deļ¬nition 2.1. A seminorm on a vector space V is a map Ā·: V ā†’ [0, āˆž) such that, for all Ī» āˆˆ C and x, y āˆˆ V , (a) Ī»x = |Ī»| x, (b) x + y ā‰¤ x + y. A seminorm that obeys x = 0 whenever x = 0 is called a norm. Since (b) implies x āˆ’ z ā‰¤ x āˆ’ y + y āˆ’ z, any norm induces a metric d(x, y) = x āˆ’ y,

(2.1) 45

46

2. Banach spaces

so every normed vector space is a metric space, and when we use metric space terminology this refers to the induced metric (2.1). In particular, a sequence (xn )āˆž n=1 in V is Cauchy if lim

sup xm āˆ’ xn  = 0,

N ā†’āˆž n,mā‰„N

and it is convergent if for some x āˆˆ V , limnā†’āˆž xn āˆ’ x = 0. We denote convergence as always by limnā†’āˆž xn = x or xn ā†’ x, n ā†’ āˆž. Deļ¬nition 2.2. A Banach space is a normed vector space which is complete with respect to the induced metric. Example 2.3. For any n āˆˆ N, Cn is a Banach space with any of the norms " # n  # n |zj |, z2 = $ |zj |2 , zāˆž = max |zj |. z1 = j=1

j=1

1ā‰¤jā‰¤n

The proofs for Ā·1 and Ā·āˆž are the same as in Rn . The proof for Ā·2 can be done by adapting the proof from Rn , or more quickly by interpreting Cn as R2n in the standard way and noting that Ā·2 rewrites as the Euclidean norm on R2n since |zj |2 = (Re zj )2 + (Im zj )2 . Later in this chapter, Example 2.3 will be vastly generalized in the context of Lp spaces, and further examples will be discussed. Lemma 2.4. In any normed vector space, |x āˆ’ y| ā‰¤ x āˆ’ y.

(2.2)

Proof. By the triangle inequality, y ā‰¤ x + y āˆ’ x and x ā‰¤ y + x āˆ’ y. Rearranging and using y āˆ’ x = āˆ’(x āˆ’ y) = x āˆ’ y, we get āˆ’x āˆ’ y ā‰¤ x āˆ’ y ā‰¤ x āˆ’ y, which is equivalent to (2.2).



The estimate (2.2) implies one of the basic continuity observations: Lemma 2.5. For any normed vector space V , the following are continuous: (a) the norm V ā†’ [0, āˆž), x ā†’ x; (b) vector addition V Ɨ V ā†’ V , (x, y) ā†’ x + y; (c) scalar multiplication C Ɨ V ā†’ V , (Ī», x) ā†’ Ī»x. Proof. If xn ā†’ x, then by deļ¬nition, xn āˆ’ x ā†’ 0, so the inequality |xn  āˆ’ x| ā‰¤ xn āˆ’ x implies xn  ā†’ x. Thus, the norm is continuous. Continuity of addition follows similarly from (xn + yn ) āˆ’ (x + y) ā‰¤ xn āˆ’ x + yn āˆ’ y,

2.1. Norms and Banach spaces

47

and continuity of scalar multiplication from Ī»n xn āˆ’ Ī»x ā‰¤ Ī»n (xn āˆ’ x) + (Ī»n āˆ’ Ī»)x = |Ī»n |xn āˆ’ x + |Ī»n āˆ’ Ī»|x.  Series in a Banach space are again deļ¬ned as a limit of partial sums: āˆž Lemma 2.6 āˆž(Weierstrass). For a sequence (xn )n=1 in a Banach space V such that n=1 xn  < āˆž, the series āˆž 

xn = lim

n=1

is convergent in V and

N ā†’āˆž

xn

n=1

% % āˆž āˆž %  % % % ā‰¤ x xn . % n% % % n=1

Proof. Denote Sn =

N 

n=1

n

k=1 xk .

For m < n, by the triangle inequality, % % n n āˆž %  %   % % xk % ā‰¤ xk  ā‰¤ xk . Sm āˆ’ Sn  = % % % k=m+1

k=m+1

k=m+1

Since tails of convergent series converge to 0, this implies that (Sn )āˆž n=1 is a Cauchy sequence, so it is convergent. By continuity of the norm, n āˆž % %   % % xk  = xk .  % lim Sn % = lim Sn  ā‰¤ lim nā†’āˆž

nā†’āˆž

nā†’āˆž

k=1

k=1

Deļ¬nition 2.7. A nonempty set S āŠ‚ V is a subspace of a normed vector space V if x, x Ėœ āˆˆ S implies x + x Ėœ āˆˆ S and Ī» āˆˆ C, x āˆˆ S implies Ī»x āˆˆ S. S is a closed subspace of V if S is a subspace, and S is closed with respect to the induced metric from V . Any subspace of a normed vector space is also a normed vector space with the inherited norm. However, only a closed subspace of a Banach space is a Banach space with the inherited norm. Finite-dimensional subspaces of a Banach space are always closed (Exercise 2.1). Finally, we describe a general construction for obtaining a normed vector space from a vector space equipped with a seminorm. Lemma 2.8. Let V be a vector space with a seminorm Ā·. Then: (a) V0 = {x āˆˆ V | x = 0} is a vector subspace of V ; (b) if x āˆ’ y āˆˆ V0 , then x = y; (c) on the quotient vector space V /V0 , [x] := x deļ¬nes a norm.

48

2. Banach spaces

Proof. (a) Let Ī» āˆˆ C and x, y āˆˆ V0 . Then Ī»x = |Ī»|x = 0 and 0 ā‰¤ x + y ā‰¤ x + y = 0, so Ī»x, x + y āˆˆ V0 . (b) follows from (2.2). (c) By (b), [x] = x deļ¬nes a function on V /V0 . It inherits seminorm properties from the seminorm on V , and [x] = [0] implies x āˆˆ / V0 , so [x] = x = 0. 

2.2. The Banach space C(K) If X, Y are metric spaces, C(X, Y ) denotes the set of continuous functions from X to Y . We are particularly interested in the set C(K) = C(K, C) of continuous functions from a compact K to C. Compactness of K implies that every f āˆˆ C(K) is bounded and has a maximum absolute value, so f  = sup |f (x)| = max|f (x)| xāˆˆK

xāˆˆK

deļ¬nes a norm on C(K). Theorem 2.9. If K is a compact metric space, C(K) is a Banach space. Proof. Let (fn )āˆž n=1 be a Cauchy sequence in C(K). Since |fm (x)āˆ’fn (x)| ā‰¤ fm āˆ’ fn , (fn (x))āˆž n=1 is a Cauchy sequence in C for every x āˆˆ K. Thus, f (x) = limnā†’āˆž fn (x) exists pointwise. For every > 0 there exists N such that for all m, n ā‰„ N and all x āˆˆ K, |fm (x) āˆ’ fn (x)| < . Taking m ā†’ āˆž gives āˆ€ > 0 āˆƒN āˆˆ N āˆ€n ā‰„ N āˆ€x āˆˆ K |f (x) āˆ’ fn (x)| ā‰¤ .

(2.3)

Fix y āˆˆ K and > 0. Use (2.3) to choose n such that supxāˆˆK |f (x)āˆ’fn (x)| ā‰¤

. Choose Ī“ > 0 such that d(x, y) < Ī“ implies |fn (x) āˆ’ fn (y)| < . Then d(x, y) < Ī“ implies |f (x) āˆ’ f (y)| ā‰¤ |f (x) āˆ’ fn (x)| + |fn (x) āˆ’ fn (y)| + |fn (y) āˆ’ f (y)| < 3 . Since is arbitrary, f is continuous at y. Since y is arbitrary, f āˆˆ C(K).  Now (2.3) says that limnā†’āˆž f āˆ’ fn  = 0, so C(K) is complete. Convergence in C(K) is called uniform convergence. Perhaps surprisingly, uniform convergence can be characterized by pointwise convergence: Lemma 2.10. Let K be a compact metric space and let fn āˆˆ C(K), f āˆˆ C(K) be functions with the property that lim xn = x =ā‡’ lim fn (xn ) = f (x).

nā†’āˆž

nā†’āˆž

Then fn converge to f uniformly on K.

(2.4)

2.2. The Banach space C(K)

49

Proof. We deļ¬ne a function F on the compact set L = K Ɨ ({0} āˆŖ {1/n | n āˆˆ N}) by F (x, 1/n) = fn (x) and F (x, 0) = f (x), and we use the fact that most points in {0} āˆŖ {1/n | n āˆˆ N} are isolated. At a point of the form (x, 1/n), continuity of F follows from continuity of fn . At a point of the form (x, 0), continuity of F follows from continuity of f and (2.4). Since F is continuous on the compact set L, it is uniformly continuous. Uniform continuity lets us estimate F (x, 1/n)āˆ’F (x, 0) uniformly in x which precisely gives uniform convergence of fn to f .  A subset of a metric space is called precompact if its closure is compact. Bounded subsets of C(K) are not in general precompact; to formulate a criterion for precompactness, we need the following notions. Deļ¬nition 2.11. Let X be a metric space with metric d. A family F āŠ‚ C(X, C) is said to be (a) pointwise bounded if supf āˆˆF |fn (x)| < āˆž for every x āˆˆ X; (b) equicontinuous if for every x āˆˆ X and > 0, there is Ī“ > 0 such that |f (x) āˆ’ f (y)| < holds for all f āˆˆ F and all y āˆˆ K with d(x, y) < Ī“. Theorem 2.12 (Arzel` aā€“Ascoli). Let K be a compact metric space. If F āŠ‚ C(K) is pointwise bounded and equicontinuous, any sequence in F has a convergent subsequence in C(K). The proof has several steps, which we separate for independent interest and formulate in a more general setting with other applications in mind. Lemma 2.13. If X is separable and F āŠ‚ C(X, C) is pointwise bounded, then any sequence in F has a subsequence which converges pointwise on a dense subset of X. Proof. Since X is separable, it has a countable dense subset {xk | k āˆˆ N}. To construct a subsequence of a sequence (fn )āˆž n=1 in F , we use a diagonalization argument. Denote j(0, n) = n. Inductively in k āˆˆ N, since supnāˆˆN |fn (xk )| < āˆž, the sequence {j(k āˆ’ 1, n)}āˆž n=1 has a subsequence such that lim f (x ) exists. Since {j(n, n)}āˆž {j(k, n)}āˆž nā†’āˆž j(k,n) k n=1 n=k is a āˆž āˆž subsequence of {j(k, n)}n=1 , the subsequence {fj(n,n) }n=1 converges at xk for every k.  Theorem 2.14. Let fn : X ā†’ C be an equicontinuous sequence of functions on a metric space X, which converges pointwise on a dense set in X. Then the sequence converges pointwise everywhere, the pointwise limit f is continuous, and the functions have property (2.4). In particular, fn converge to f uniformly on compact subsets K āŠ‚ X.

50

2. Banach spaces

Proof. We deļ¬ne C(x) = lim

sup |fm (x) āˆ’ fn (x)|.

N ā†’āˆž m,nā‰„N

By deļ¬nition, the sequence (fn (x))āˆž n=1 is Cauchy if and only if C(x) = 0. For any x āˆˆ X and > 0, by equicontinuity, there exists Ī“ > 0 such that |fn (x) āˆ’ fn (y)| <

āˆ€n āˆˆ N āˆ€y āˆˆ BĪ“ (x),

(2.5)

where we denote BĪ“ (x) = {y āˆˆ X | d(x, y) < Ī“}. By using |fm (x) āˆ’ fn (x)| ā‰¤ |fm (x) āˆ’ fm (y)| + |fm (y) āˆ’ fn (y)| + |fn (y) āˆ’ fn (x)|, inequality (2.5) implies C(x) ā‰¤ 2 + C(y) for y āˆˆ BĪ“ (x). Since there is a dense set of y āˆˆ X such that C(y) = 0, this implies C(x) ā‰¤ 2 , and since

> 0 is arbitrary, it implies C(x) = 0. Thus, fn (x) converges pointwise. Taking n ā†’ āˆž in (2.5) shows continuity of f (x) = limnā†’āˆž fn (x). For ļ¬xed x āˆˆ X and > 0, choose N so that |fn (x) āˆ’ f (x)| < for n ā‰„ N . Combining this with (2.5) shows that |f (x) āˆ’ fn (y)| < for all n ā‰„ N and all y āˆˆ BĪ“ (x). Since > 0 is arbitrary, this proves (2.4). By Lemma 2.10, fn converge uniformly to f on compact K āŠ‚ X.



Proof of Theorem 2.12. Since K is compact, it is separable. Thus, Lemma 2.13, any sequence has a subsequence which converges pointwise some dense set. By Theorem 2.14, this subsequence converges uniformly K.

by on on 

The remainder of this section is dedicated to an important criterion for density in C(K), called the complex Stoneā€“Weierstrass theorem. The criterion uses additional algebraic structure of C(K): In addition to the vector space structure, C(K) is equipped with the binary operation of pointwise multiplication, because products of continuous functions are continuous. Moreover, pointwise multiplication in C(K) is continuous, since fn gn āˆ’ f g ā‰¤ fn gn āˆ’ g + fn āˆ’ f g. Deļ¬nition 2.15. Let F = R or F = C. A subset S āŠ‚ C(K, F) is a subalgebra of C(K, F) if it contains the constant function 1 and is closed under pointwise addition, pointwise multiplication, and scalar multiplication by Ī» āˆˆ F. S separates points if for every x = y, there is f āˆˆ S such that f (x) = f (y). Theorem 2.16 (Stoneā€“Weierstrass). Let S be a subalgebra of C(K, R). If S separates points on K and if 1 āˆˆ S, then S is dense in C(K, R).

2.2. The Banach space C(K)

51

The proof uses approximation of |Ā·| by polynomials on compact intervals: Lemma 2.17. For any a > 0 and Ī“ > 0, there is a polynomial P such that max ||y| āˆ’ P (y)| ā‰¤ Ī“.

(2.6)

yāˆˆ[āˆ’a,a]

Proof. By the binomial expansion, for x āˆˆ (āˆ’1, 1), ! āˆž  āˆš 1 Ɨ 3 Ɨ Ā· Ā· Ā· Ɨ (2k āˆ’ 3) k kāˆ’1 1/2 . 1āˆ’x=1āˆ’ ck x , ck = (āˆ’1) = 2k k! k k=1

Since ck > 0 for all k, by monotone convergence, āˆž  k=1

ck = lim xā†‘1

āˆž 

ck xk = lim(1 āˆ’

k=1

xā†‘1

āˆš

1 āˆ’ x) = 1.

āˆž

In particular, the series k=1 ck is convergent and the binomial expansion holds at x = 1 as well, by continuity of both sides. Denote Qn (x) = 1 āˆ’ nk=1 ck xk . Then for all x āˆˆ [0, 1], āˆž āˆž    āˆš  1 āˆ’ x āˆ’ Qn (x) = c k xk ā‰¤ ck . k=n+1

k=n+1

By the decay of tails of convergent series, the polynomials Qn converge to āˆš 1 āˆ’ x uniformly on [0, 1] as n ā†’ āˆž. By substituting x = 1 āˆ’ y 2 /a2 , polynomials Qn (1 āˆ’ y 2 /a2 ) converge to |y| a uniformly on y āˆˆ [āˆ’a, a]. Multiplying by a and using P (y) = aQn (1 āˆ’ y 2 /a2 ) for large enough n gives (2.6).  Separation of points enters through the following lemma, which allows us to arbitrarily prescribe values at two points: Lemma 2.18. Assume that S separates points on K and 1 āˆˆ S. Then, for any x, y āˆˆ K and f āˆˆ C(K, R), there exists x, y āˆˆ S such that hx,y (x) = f (x),

hx,y (y) = f (y).

Proof. If x = y, this is trivial: it suļ¬ƒces to take hx,y a constant function. From now on we assume x = y. Point evaluations at x, y give a map ! h(x) h ā†’ h(y) from S to R. Since S is a vector space, the image under   this map is a vector subspace of R2 . Since 1 āˆˆ S, the image contains 11 . There exists Hx,y āˆˆ S such that Hx,y (x) = Hx,y (y). Thus, the image contains two linearly independent vectors in R2 , so it contains all of R2 . In other words, by choosing a linear combination of Hx,y and 1, denoted hx,y āˆˆ S, we can  ensure that hx,y (x) = f (x) and hx,y (y) = f (y).

52

2. Banach spaces

With these ingredients we can complete the proof of the real Stoneā€“ Weierstrass theorem: Proof of Theorem 2.16. We will work with the closed subalgebra S and prove that S = C(K, R). The proof consists of several steps. The ļ¬rst step is to prove that f āˆˆ S implies |f | āˆˆ S. If f āˆˆ S, denoting a = f , by Lemma 2.17 for any > 0 there is a polynomial P such that sup ||y| āˆ’ P (y)| ā‰¤ . yāˆˆ[āˆ’a,a]

Since f takes values in [āˆ’a, a], |f | āˆ’ P ā—¦ f  = sup ||f (x)| āˆ’ P (f (x))| ā‰¤ . xāˆˆK

āˆˆ S for any k āˆˆ N, and then P ā—¦ f āˆˆ S. Since Since S is a subalgebra,

> 0 is arbitrary, |f | āˆˆ S. fk

It follows that, for any f, g āˆˆ S, f + g |f āˆ’ g| + āˆˆ S, 2 2 and similarly min(f, g) āˆˆ S. By induction in n āˆˆ N, maxima and minima of n functions in S are also in S. max(f, g) =

From now on, let us ļ¬x f āˆˆ C(K, R) and > 0. For any x, y āˆˆ K, there exists hx,y āˆˆ S such that hx,y (x) = f (x) and hx,y (y) = f (y). Let us ļ¬x x for the moment. By continuity, for every y āˆˆ K, there is an open neighborhood Uy of y on which hy > f āˆ’ . By compactness, K has a ļ¬nite subcover Uy1 , . . . , Uyn . Choosing gx = max(hx,y1 , . . . , hx,yn ) gives a function gx āˆˆ S such that gx (x) = f (x) and gx > f āˆ’ on K. By continuity, any x āˆˆ K has an open neighborhood Vx on which gx < f + . By compactness, K has a ļ¬nite subcover Vx1 , . . . , Vxm . The function F = min(gx1 , . . . , gxm ) āˆˆ S obeys f āˆ’ < F < f + for all x āˆˆ K. Since > 0 is arbitrary, this implies f āˆˆ S.



Theorem 2.19 (Complex Stoneā€“Weierstrass theorem). Let K be a compact metric space. Let S be a subalgebra of C(K) which separates points, and assume also that for any f āˆˆ S, its complex conjugate fĀÆ is also in S. Then S is a dense subset of C(K). Proof. Let SR = S āˆ© C(K, R). Clearly, 1 āˆˆ SR . If f āˆˆ S, then f āˆˆ S, so Re f = f +f 2 āˆˆ SR . For any x = y, there exists h āˆˆ S such that h(x) = h(y).

2.2. The Banach space C(K)

53

Then Re(eiĻ† h) āˆˆ S for any Ļ† āˆˆ R; the choice Ļ† = āˆ’ arg(h(x) āˆ’ h(y)) guarantees that Re(eiĻ† h(x)) = Re(eiĻ† h(y)), so SR separates points. By Theorem 2.16, the closure of SR is C(K, R). It follows that f = Re f āˆ’  i Re(if ) āˆˆ S for any f āˆˆ C(K). This criterion eļ¬€ortlessly recovers some classical approximation results: Corollary 2.20 (Weierstrassā€™s ļ¬rst theorem). For compact K āŠ‚ R, polynomials with complex coeļ¬ƒcients are dense in C(K). Proof. Polynomials with complex coeļ¬ƒcients are a subalgebra of C(K), and the complex conjugate of a polynomial is a polynomial. Polynomials separate points because f (x) = x is a polynomial and is injective on K. Thus, polynomials are dense in C(K).  If we remove the assumption K āŠ‚ R, polynomials may not be dense. An important special case is the unit circle āˆ‚D = {z āˆˆ C | |z| = 1}. Note that for any n āˆˆ Z, z ā†’ z n is a continuous function from āˆ‚D ā†’ C. Polynomials are not dense in C(āˆ‚D): for instance, the function 1/z cannot be approximated by polynomials since for any polynomial p, % %  %  2Ļ€ !% %1 % % %    dx  1 ix ix % āˆ’ p(z)% %  = 1. %  1 āˆ’ e p(e ) āˆ’ p(z) % = %z ā‰„ %z %  z 2Ļ€ 0 C(āˆ‚D) C(āˆ‚D) To obtain a dense set, one also includes negative powers of z: Corollary 2.21 (Weierstrassā€™s second theorem). The subspace of Laurent polynomials span{z n | n āˆˆ Z} is dense in C(āˆ‚D). Proof. S = span{z n | n āˆˆ Z} is a subalgebra of C(āˆ‚D) because z m z n = z m+n and z 0 = 1. S is closed under complex conjugation because z n = z āˆ’n . S separates points because f (z) = z is injective on āˆ‚D. Thus, by the complex  Stoneā€“Weierstrass theorem, span{z n | n āˆˆ Z} is dense in C(āˆ‚D). Weierstrassā€™s second theorem is often restated using the substitution z= n as a statement about density of trigonometric polynomials k=m ck eikt (where m, n āˆˆ Z, m ā‰¤ n) in the space C(T), where T = R/2Ļ€Z.

eit ,

Some other concrete density results are left to Exercises 2.2, 2.3, and 2.4. Each of these applications implies separability of the corresponding space C(K). For instance, it follows from Weierstrassā€™s ļ¬rst theorem that the set of polynomials with coeļ¬ƒcients in Q + iQ is a countable dense subset of C([a, b]), so C([a, b]) is separable. Those are special cases of a general fact: Theorem 2.22. If K is a compact metric space, then C(K) is separable.

54

2. Banach spaces

Proof. Let {xn }āˆž n=1 be a countable dense set in K, and denote by d the metric in K. Deļ¬ne for n, m āˆˆ N fn,m (x) = max(1 āˆ’ md(x, xn ), 0). This is a countable set of functions which separates points. The set of all ļ¬nite products V = {1} āˆŖ {fn1 ,m1 . . . fnk ,mk | k āˆˆ N, n1 , . . . , nk , m1 , . . . , mk āˆˆ N} is also countable and is closed under multiplication. Thus, span V obeys all the assumptions of the Stoneā€“Weierstrass theorem, so span V is dense in C(K). Any linear combination of elements of V can be approximated by one with coeļ¬ƒcients in Q + iQ. Thus, linear combinations of elements of V with coeļ¬ƒcients in Q + iQ are dense in C(K); the set of such linear combinations is countable. 

2.3. Lp spaces Fix a measure Ī¼ on X and p āˆˆ [1, āˆž]. For f : X ā†’ C, deļ¬ne !1/p  p |f | dĪ¼ , p āˆˆ [1, āˆž), f p =

(2.7)

X

f āˆž = inf{t āˆˆ [0, āˆž] | |f (x)| ā‰¤ t for Ī¼-a.e. x}.

(2.8)

First, let us prove that the inf in (2.8) is a minimum: Lemma 2.23. Let f : X ā†’ C. For Ī¼-a.e. x, |f (x)| ā‰¤ f āˆž . Proof. There exists a sequence of tn ā‰„ f āˆž such that tn ā†’ f āˆž and Ī¼({x | |f (x)| > tn }) = 0. A countable union of zero measure sets has zero  measure, so Ī¼({x | |f (x)| > f āˆž }) = 0. In other words: |f | ā‰¤ t holds Ī¼-a.e. if and only if t ā‰„ f āˆž . Corollary 2.24. f p = 0 if and only if f = 0 holds Ī¼-a.e. Proof. For p = āˆž this follows from Lemma 2.23, and for p āˆˆ [1, āˆž) from  Proposition 1.57 applied to |f |p . For any p āˆˆ [1, āˆž], deļ¬ne Lp (X, dĪ¼) = {f : X ā†’ C | f p < āˆž}. We will see that f p is a seminorm on Lp (X, dĪ¼); passing to a quotient space will then give a normed vector space Lp (X, dĪ¼). The property Ī»f p = |Ī»|f p is immediate, so it remains to prove the triangle inequality. The case p = 1 follows from    f + g1 = |f + g| dĪ¼ ā‰¤ |f | dĪ¼ + |g| dĪ¼ = f 1 + g1 .

2.3. Lp spaces

55

The case p = āˆž follows from |f (x) + g(x)| ā‰¤ |f (x)| + |g(x)| ā‰¤ f āˆž + gāˆž ,

for Ī¼-a.e. x.

Although the cases p = 1 and p = āˆž are easier, they require separate treatment so we will exclude them in the arguments below, leaving them as an exercise. If p, q āˆˆ [1, āˆž] and p1 + 1q = 1, then p, q are called conjugate exponents. Lemma 2.25 (Youngā€™s inequality). Let p, q āˆˆ (1, āˆž) be conjugate exponents. For any x, y ā‰„ 0, xp y q + . (2.9) xy ā‰¤ p q Proof. The exponential function is convex, so for all u, v āˆˆ R and t āˆˆ (0, 1), etu+(1āˆ’t)v ā‰¤ teu + (1 āˆ’ t)ev (Exercise 2.6). Using u = log xp , v = log y q , t = 1p , 1 āˆ’ t = for x, y > 0. If x = 0 or y = 0, the inequality is trivial.

1 q

proves (2.9) 

Theorem 2.26 (HĀØ olderā€™s inequality). If p, q āˆˆ (1, āˆž) are conjugate expop nents and f āˆˆ L (X, dĪ¼), g āˆˆ Lq (X, dĪ¼), then gĀÆf āˆˆ L1 (X, dĪ¼) and      gf dĪ¼ ā‰¤ f p gq . (2.10)   Proof. If f p = 0, then f = 0 Ī¼-a.e., so gĀÆf = 0 Ī¼-a.e., and the statement is trivial. If f p = 0, by dividing by f p , it suļ¬ƒces to consider the case f p = 1. Similarly, it suļ¬ƒces to consider gq = 1. By Youngā€™s inequality, |g(x)f (x)| ā‰¤

|f (x)|p |g(x)|q + , q p

so by integrating,    q p    gf dĪ¼ ā‰¤ |gf | dĪ¼ ā‰¤ g + f  = 1 + 1 = 1 = f p gq .   q p q p



Noting that for any f it is possible to choose g such that equality holds, we obtain the following corollary. + 1q = 1. For any f āˆˆ Lp (X, dĪ¼),     .  gf dĪ¼ (2.11) f p = max   gāˆˆLq (X,dĪ¼)

Corollary 2.27. Let p, q āˆˆ (1, āˆž) obey

gq =1

1 p

56

2. Banach spaces

Proof. By rescaling and using HĀØolderā€™s inequality,  it suļ¬ƒces to show that q there exists g āˆˆ L (X, dĪ¼) with gq = 1 and gĀÆf dĪ¼ = f p = 1. It is straightforward to verify this for

|f (x)|pāˆ’2 f (x) f (x) = 0 g(x) =  0 f (x) = 0. HĀØ olderā€™s inequality holds also for p = 1 and p = āˆž (see also Exercise 2.7). Corollary 2.27 characterizes f p as an extremum over linear expressions in f , which is useful for proving subadditivity of the norm: Theorem 2.28 (Minkowskiā€™s inequality). For any p āˆˆ [1, āˆž], for all f1 , f2 āˆˆ Lp (X, dĪ¼), (2.12) f1 + f2 p ā‰¤ f1 p + f2 p . Proof. The cases p = 1, p = āˆž are easy and were proved before. Let p āˆˆ (1, āˆž) and denote by q the conjugate exponent. For any f1 , f2 āˆˆ Lp (X, dĪ¼) and g āˆˆ Lq (X, dĪ¼), by the triangle inequality,              g(f1 + f2 ) dĪ¼ ā‰¤  gf1 dĪ¼ +  gf2 dĪ¼ ā‰¤ gq f1 p + gq f2 p .       X

X

X

Taking the supremum over g with gq = 1 gives (2.12).



Collecting the facts, we have proved the following: Theorem 2.29. For any p āˆˆ [1, āˆž], Ā·p is a seminorm on Lp (X, dĪ¼) and f p = 0 if and only if f (x) = 0 for Ī¼-a.e. x. Thus, with the zero norm subspace V0 = {f : X ā†’ C | f = 0 Ī¼-a.e.}, the quotient space construction in Lemma 2.8 gives the normed vector space Lp (X, dĪ¼) = Lp (X, dĪ¼)/V0 = {[f ] | f p < āˆž}. This is commonly phrased in terms of an equivalence relation, f āˆ¼ g if and only if f = g Ī¼-a.e. An element of Lp (X, dĪ¼) is an equivalence class [f ] corresponding to a Borel function f , but following standard conventions, when considering elements of Lp (X, dĪ¼), we typically do not distinguish between a Borel function and its equivalence class. For instance, since Corollary 2.27 is not aļ¬€ected by changing f or g on a zero measure set, it can be formulated as a statement for the quotient spaces: for any f āˆˆ Lp (X, dĪ¼),     .  gf dĪ¼ (2.13) f p = max  gāˆˆLq (X,dĪ¼)  gq =1

Let us consider completeness. For p = āˆž, convergence in Lāˆž (X, dĪ¼) corresponds to a ā€œĪ¼-a.e.ā€ version of uniform convergence, so by working

2.3. Lp spaces

57

away from a zero measure set, the proof of completeness of C(K) also proves that Lāˆž (X, dĪ¼) is a Banach space. It remains to consider p āˆˆ [1, āˆž). Theorem 2.30 (Rieszā€“Fischer). For p āˆˆ [1, āˆž), Lp (X, dĪ¼) is a Banach space. p Proof. Let (fn )āˆž n=1 be a Cauchy sequence in L (X, dĪ¼). By general metric āˆž space arguments, there is a subsequence (fnk )k=1 such that fnk+1 āˆ’fnk p ā‰¤ 1 . It is notationally convenient to use fn0 = 0 and to conclude 4k āˆž 

fnk āˆ’ fnkāˆ’1 p < āˆž.

k=1

Consider h(x) = āˆž k=1 |fnk (x) āˆ’ fnkāˆ’1 (x)|. By monotone convergence, p    m    p |h| dĪ¼ = lim |f (x) āˆ’ f (x)|  dĪ¼,  nk nkāˆ’1 mā†’āˆž   k=1

so taking pth roots and using Minkowskiā€™s inequality (2.12), % % m m % %  % % % % %fn āˆ’ fn % . hp = lim % |fnk āˆ’ fnkāˆ’1 |% ā‰¤ lim k kāˆ’1 p mā†’āˆž % mā†’āˆž % k=1

p

k=1

The right-hand side is a convergent series, so h āˆˆ Lp (X, dĪ¼), and h < āˆž Ī¼-a.e. By the deļ¬nition of h, this implies that for Ī¼-a.e. x, the sequence (fnk (x))āˆž k=1 is Cauchy and |fnk (x)| ā‰¤ |h(x)| for all k, so the pointwise limit f (x) = limkā†’āˆž fnk (x) exists Ī¼-a.e., and |f | ā‰¤ h. Due to |fn | ā‰¤ h and |f | ā‰¤ h, we estimate |fn āˆ’ f |p ā‰¤ 2p hp , so by dominated convergence with the dominating function 2p hp ,   lim |fnk āˆ’ f |p dĪ¼ = lim |fnk āˆ’ f |p dĪ¼ = 0, kā†’āˆž

kā†’āˆž

so fnk āˆ’ f p ā†’ 0. By general metric space arguments, since the Cauchy  sequence (fn )āˆž n=1 has a convergent subsequence, it is convergent. This proof yields an additional fact which will be useful. Convergence in p-norm does not imply pointwise convergence, and pointwise convergence does not imply convergence in p-norm; however: p Corollary 2.31. If (fn )āˆž n=1 is a sequence such that fn ā†’ f in L (X, dĪ¼) and fn ā†’ g pointwise Ī¼-a.e., then f = g Ī¼-a.e.

Proof. By the proof of the Rieszā€“Fischer theorem, there is a subsequence p (fnk )āˆž k=1 which converges both in L (X, dĪ¼) and pointwise to the same limit. That limit must be equal Ī¼-a.e. to both f and g, so f = g Ī¼-a.e. 

58

2. Banach spaces

We now start using topological properties of X to prove a density statement. We will work in the setting of Ļƒ-locally compact spaces and Baire measures on them, as deļ¬ned in Section 1.8. Theorem 2.32. Let X be a Ļƒ-locally compact metric space, let Ī¼ be a Baire measure on X, and let p āˆˆ [1, āˆž). Then Cc (X) is a dense subset of Lp (X, dĪ¼). Proof. First, we note that any f āˆˆ Cc (X) is in Lp (X, dĪ¼), because |f | ā‰¤ CĻ‡K for some C > 0 and K compact, and because Ī¼ is ļ¬nite on compacts. Denote by M the closure of Cc (X) in Lp (X, dĪ¼). Since Cc (X) is a vector subspace of Lp (X, dĪ¼), so is M. We will show M = Lp (X, dĪ¼). Consider a Borel set B āŠ‚ X with Ī¼(B) < āˆž. By regularity of Ī¼, for any > 0, there exist compact K and open V such that K āŠ‚ B āŠ‚ V and Ī¼(V \ K) < . By Lemma 1.89, there exists f āˆˆ Cc (X) with Ļ‡K ā‰¤ f ā‰¤ Ļ‡V . It follows that |f āˆ’ Ļ‡B | ā‰¤ Ļ‡V \K so f āˆ’ Ļ‡B p ā‰¤ Ī¼(V \ K) < . Since > 0 is arbitrary, Ļ‡B āˆˆ M. Thus, M contains all Borel sets B with Ī¼(B) < āˆž. By taking linear combinations, any simple function s āˆˆ Lp (X, dĪ¼) is in M. Any positive function f āˆˆ Lp (X, dĪ¼) can be approximated from below by simple functions 0 ā‰¤ sn ā‰¤ f , snā†’ f , so by dominated convergence with dominating function |f |p , we have |sn āˆ’ f |p dĪ¼ ā†’ 0 and f āˆˆ M. Since any complex-valued f can be written as a linear combination of four positive functions f = (Re f )+ āˆ’ (Re f )āˆ’ + i(Im f )+ āˆ’ i(Im f )āˆ’ , and those functions  are in Lp (X, dĪ¼) if f is, we conclude that any f āˆˆ Lp (X, dĪ¼) is in M. Combining this density result with Theorem 2.22 gives: Corollary 2.33. Let X be a Ļƒ-compact metric space, let Ī¼ be a Baire measure, and let p āˆˆ [1, āˆž). Then Lp (X, dĪ¼) is a separable Banach space.  Proof. If X = nāˆˆN Ln and sets Ln are compact, it suļ¬ƒces to take a union of countable dense sets in C(Ln ), n āˆˆ N.  The density statements above were only for p āˆˆ [1, āˆž), as the case p = āˆž is very diļ¬€erent in this regard (Exercises 2.8 and 2.9). We end by remarking upon two notationally special cases. When X is a subset of Rd for some d, if Ī¼ is chosen to be the restriction of the d-dimensional Lebesgue measure to X, we denote Lp (X) = Lp (X, dĪ¼). For any set X, if Ī¼ is chosen to be the counting measure on X, we will denote p (X) = Lp (X, dĪ¼). When X is countable (most commonly, X = N or X = Z), it is Ļƒ-locally compact with respect to the discrete metric; thus, it is a special case of the above considerations. With respect to the counting measure on X, the only zero measure set is the empty set, so the

2.4. Bounded linear operators and uniform boundedness

59

general quotient space step is not needed here, and p (X) is exactly the set of sequences (and not equivalence classes of sequences) & '  1/p p p  (X) = f : X ā†’ C | f p = |f (x)| 0. Then x < r implies T x < 1. Applying this to vectors with x = r/2 and rescaling by 2/r shows that for any x with x = 1, we have T x < 2/r. Thus, T  ā‰¤ 2/r. If T is bounded and linear, then T x āˆ’ T y = T (x āˆ’ y) ā‰¤ T x āˆ’ y, so T is continuous.



60

2. Banach spaces

For T āˆˆ L(X, Y ), the kernel and range are deļ¬ned by Ker T = {x āˆˆ X | T x = 0}, Ran T = {T x āˆˆ Y | x āˆˆ X}. Lemma 2.37. If T āˆˆ L(X, Y ), then Ker T is a closed subspace of X. Proof. Since T is continuous, if xn ā†’ x and T xn = 0, then T x = T lim xn = lim T xn = lim 0 = 0. nā†’āˆž

nā†’āˆž

nā†’āˆž



It should be noted that the subspace Ran T is not always closed in Y . Since Y is a vector space, L(X, Y ) inherits a vector space structure deļ¬ned by (S + T )x = Sx + T x,

(Ī»S)x = Ī»(Sx)

āˆ€x āˆˆ X.

Proposition 2.38. If X is a normed vector space and Y is a Banach space, then L(X, Y ) is a Banach space with the norm (2.14). Proof. It is straightforward to verify that L(X, Y ) is a vector space and that (2.14) is a norm. To show that L(X, Y ) is complete, let (Tn )āˆž n=1 be a Cauchy sequence in L(X, Y ). For any x āˆˆ X, the inequality Tn x āˆ’ Tm x ā‰¤ Tn āˆ’ Tm x shows that (Tn x)āˆž n=1 is a Cauchy sequence in Y , so it is convergent. We deļ¬ne T : X ā†’ Y by T x = lim Tn x. nā†’āˆž

Linearity of T follows from linearity of Tn . Since (Tn )āˆž n=1 is a Cauchy sequence in L(X, Y ), supn Tn  < āˆž, so T x = lim Tn x ā‰¤ lim supTn x. nā†’āˆž

nā†’āˆž

Therefore, T is bounded with norm at most lim supnā†’āˆž Tn . Similarly, (Tn āˆ’ T )x = lim (Tn āˆ’ Tm )x ā‰¤ lim supTn āˆ’ Tm x mā†’āˆž

mā†’āˆž

implies that Tn āˆ’ T  ā‰¤ lim supTn āˆ’ Tm . mā†’āˆž

Since

(Tn )āˆž n=1

is a Cauchy sequence, it follows that limnā†’āˆž Tn āˆ’T  = 0. 

Lemma 2.39. Let T1 , T2 āˆˆ L(X, Y ). If the set {x | T1 x = T2 x} is dense in X, then T1 = T2 . Proof. This set is Ker(T1 āˆ’ T2 ), which is a closed subspace of X. If it is  also dense, then Ker(T1 āˆ’ T2 ) = X, so T1 = T2 .

2.4. Bounded linear operators and uniform boundedness

61

Composition of linear operators is denoted by multiplicative notation, as in the following statement. Lemma 2.40. If T āˆˆ L(X, Y ) and S āˆˆ L(Y, Z), then ST āˆˆ L(X, Z) and ST  ā‰¤ ST . Proof. Linearity of ST follows from linearity of S and T and boundedness follows from ST x ā‰¤ ST x ā‰¤ ST x.  Deļ¬nition 2.41. The operator U āˆˆ L(X, Y ) is norm-preserving if U x = x

āˆ€x āˆˆ X.

(2.15)

U is unitary if it is norm-preserving and Ran U = Y . Remark 2.42. For readability, even in discussions that involve more than one norm, our notation for norms usually leaves that implicit. For instance, in (2.15), the norm on the left-hand side corresponds to the space Y , and the norm on the right-hand side corresponds to the space X. Lemma 2.43. If X is a Banach space, Y is a normed space, and U āˆˆ L(X, Y ) is norm-preserving, then Ker U = {0} and Ran U is a closed subspace of Y . Proof. By (2.15), U x = 0 implies x = 0, so Ker U = {0}. Assume that y is in the closure of Ran U , i.e., there exists a sequence U xn = yn ā†’ y. The sequence (U xn )āˆž n=1 is convergent, so it is a Cauchy sequence. Since U xm āˆ’ U xn  = U (xm āˆ’ xn ) = xm āˆ’ xn , the sequence (xn )āˆž n=1 is Cauchy, so it is convergent in X. Its limit x obeys U x = U lim xn = lim U xn = y, nā†’āˆž

which shows that y āˆˆ Ran U .

nā†’āˆž



The importance of unitary maps lies in the fact that they preserve all Banach space operations: since a unitary map is a linear bijection, it preserves notions from linear algebra such as linear independence, and since it preserves distances, it preserves metric space notions such as density. Unitary maps U āˆˆ L(X, Y ) are also sometimes called isometric isomorphisms, and if such U exists, X and Y are said to be isometrically isomorphic. Throughout this text, we will often work with bounded linear operators, which are initially deļ¬ned only on a dense subspace of X. In particular, some of the central results in this text are constructions of speciļ¬c unitary maps, and such constructions often start with a norm-preserving map on a dense subspace. The following procedure is therefore useful.

62

2. Banach spaces

Proposition 2.44. Let X, Y be Banach spaces, and let V be a dense subspace of X. If the linear map T : V ā†’ Y is bounded, then: (a) T can be uniquely extended to a bounded linear operator T āˆˆ L(X, Y ); (b) T  = T ; (c) {T x | x āˆˆ X} āŠ‚ {T x | x āˆˆ V }. Proof. For any x āˆˆ X and any sequence of xn āˆˆ V such that xn ā†’ x, the sequence T xn is Cauchy because T xm āˆ’ T xn  ā‰¤ T xm āˆ’ xn , so it is convergent. Moreover, the limit is independent of the choice of sequence: if Ėœn ā†’ x, then T xn āˆ’ T x Ėœn  ā‰¤ T xn āˆ’ x Ėœn  ā†’ 0. Thus, we also x Ėœn āˆˆ V , x can deļ¬ne T : X ā†’ Y by T x = limnā†’āˆž T xn for any sequence of xn āˆˆ V such that xn ā†’ x. Note that this makes (c) obvious. For x āˆˆ V , we can take xn = x for all n to conclude T x = T x, so T is an extension of T . Linearity of T follows from the linearity of T , and T x = lim T xn  ā‰¤ lim T xn  = T x nā†’āˆž

nā†’āˆž

shows that T  ā‰¤ T . The reverse inequality follows from T |V = T . If there were two extensions of T in L(X, Y ), they would be equal on the dense set V , so they would be equal by Lemma 2.39.  The specialization to norm-preserving maps has additional properties: Proposition 2.45. Let X, Y be Banach spaces, and let V be a dense subspace of X. If the linear map U : V ā†’ Y is norm-preserving, i.e., obeys (2.15) for all x āˆˆ V , then: (a) U can be uniquely extended to a bounded linear operator U āˆˆ L(X, Y ); (b) this extension is a norm-preserving map U : X ā†’ Y ; (c) {U x | x āˆˆ X} = {U x | x āˆˆ V }. Proof. Continuing from the proof of Proposition 2.44, for xn ā†’ x, xn āˆˆ V , U x = lim U xn  = lim xn  = x, nā†’āˆž

nā†’āˆž

so U is norm-preserving. The range of U contains that of U and is closed  by Lemma 2.43, so it contains {U x | x āˆˆ V }. Dense sets can also be used to study pointwise convergence of operators.

2.4. Bounded linear operators and uniform boundedness

63

Lemma 2.46. Let X, Y be Banach spaces. Consider a sequence (Tn )āˆž n=1 in L(X, Y ). If supn Tn  < āˆž and Tn converge pointwise on some dense subset of X, then Tn converge pointwise on X and T x = lim Tn x nā†’āˆž

(2.16)

deļ¬nes some T āˆˆ L(X, Y ). Moreover, T  ā‰¤ lim inf nā†’āˆž Tn . Proof. Denote M = supnāˆˆN Tn . The operators obey Tn x āˆ’ Tn y = Tn (x āˆ’ y) ā‰¤ M x āˆ’ y. Viewing this as a Banach-space valued version of equicontinuity, we can carry over an argument from Theorem 2.14. We deļ¬ne C(x) = lim

sup Tm x āˆ’ Tn x.

N ā†’āˆž m,nā‰„N

The sequence (Tn x)āˆž n=1 is Cauchy if and only if C(x) = 0. For any x, y āˆˆ X, by using Tm x āˆ’ Tn x ā‰¤ Tm x āˆ’ Tm y + Tm y āˆ’ Tn y + Tn y āˆ’ Tn x ā‰¤ M x āˆ’ y + Tm y āˆ’ Tn y + M x āˆ’ y, we obtain C(x) ā‰¤ 2M xāˆ’y+C(y). Since there is a dense set of y āˆˆ X such that C(y) = 0, this implies C(x) = 0 for all x āˆˆ X. Thus, Tn x converges for every x āˆˆ X. The map T deļ¬ned by (2.16) is linear because Tn are linear, and for any x,  T x = lim Tn x ā‰¤ lim inf Tn x. nā†’āˆž

nā†’āˆž

Further statements are left as exercises. In particular, Exercise 2.10 describes an abstract completion of a normed vector space to a Banach space, which is based on the metric space completion realized by equivalence classes of Cauchy sequences. Exercise 2.11 shows that this abstract completion is, up to a unitary, the only Banach space which contains V as a dense subset. Accordingly, it is common to call a Banach space B a completion of a normed vector space X if there is a norm-preserving map i : X ā†’ B such that Ran i is dense in B. It is also common to identify X and Ran i and think of X as a dense subset of B. For instance, the spaces Lp (X, dĪ¼) for p āˆˆ [1, āˆž) are completions of Cc (X) with respect to the Lp -norm on Cc (X). We have now reached an important general result, known as the uniform boundedness principle or the Banachā€“Steinhaus theorem. Theorem 2.47 (Uniform boundedness principle). Let X be a Banach space, and let Y be a normed vector space. If a family F āŠ‚ L(X, Y ) is pointwise bounded, i.e., āˆ€x āˆˆ X, sup T x < āˆž T āˆˆF

64

2. Banach spaces

then F is norm bounded, i.e., supT āˆˆF T  < āˆž. The proof uses an upper bound on the operator norm obtained from its values in an arbitrary ball, not necessarily centered at 0. Lemma 2.48. Let T : X ā†’ Y be a bounded linear operator between normed vector spaces. For any x āˆˆ X and any r > 0, sup T y ā‰„ T r.

(2.17)

yāˆˆX yāˆ’xā‰¤r

Proof. For all v ā‰¤ r, by the triangle inequality, 1 T v ā‰¤ (T (x + v) + T (x āˆ’ v)) ā‰¤ 2

sup T y.

yāˆˆX xāˆ’yā‰¤r

Taking the supremum over v gives an upper bound for rT .



Proof of Theorem 2.47. Assume that F is not norm bounded. Then there is a sequence of Tn āˆˆ F such that Tn  ā‰„ 4n . We construct a sequence (xn )āˆž n=1 in X inductively, by setting x0 = 0 and using Lemma 2.48 with Tn to obtain some xn in a ball of radius 3āˆ’n around xnāˆ’1 : xn āˆ’ xnāˆ’1  ā‰¤ 3āˆ’n , 2 Tn xn  > 3āˆ’n Tn . 3

(2.18) (2.19)

The factor 2/3 is added because the supremum in (2.17) may not be a maximum. By (2.18), xn converge to some x and x āˆ’ xn  ā‰¤ 12 3āˆ’n . Thus, 1 Tn x ā‰„ Tn xn  āˆ’ Tn x āˆ’ xn  > 3āˆ’n Tn  ā†’ āˆž, 6 which shows that (Tn )āˆž n=1 is not pointwise bounded.



The above discussion of L(X, Y ) is quite general and will be applied to various Banach spaces in this text. There are two common special cases. One is the case Y = C. Bounded linear operators from X to C are also called bounded linear functionals of X. The notation X āˆ— := L(X, C) is customary, and the space X āˆ— is called the dual space of X. Another is the case Y = X. The notation L(X) := L(X, X) is customary, and elements of L(X) are called bounded linear operators on X. In this special case, composition of operators can be viewed as a multiplicative operation on L(X), which provides additional algebraic structure.

2.5. Weak-āˆ— convergence and the separable Banachā€“Alaoglu theorem

65

2.5. Weak-āˆ— convergence and the separable Banachā€“Alaoglu theorem For any Banach space B, its dual space B āˆ— = L(B, C) has the induced (operator) norm which makes it a Banach space and gives a notion of norm convergence in B āˆ— . However, a weaker notion of convergence is often useful: Deļ¬nition 2.49. A sequence of Ī›n āˆˆ B āˆ— weak-āˆ— converges to Ī› āˆˆ B āˆ— if lim Ī›n x = Ī›x

nā†’āˆž

āˆ€x āˆˆ B. w

This is often denoted by w-limnā†’āˆž Ī›n = Ī› or Ī›n ā†’ Ī› as n ā†’ āˆž. Since weak-āˆ— convergence is not deļ¬ned as convergence with respect to a metric, the reader is warned not to automatically apply preconceptions about convergence taken from metric spaces. For instance, uniqueness of the weak-āˆ— limit has to be proved: w

w

Lemma 2.50. If Ī›n ā†’ Ī› and Ī›n ā†’ Ī› , then Ī› = Ī› . Proof. For all x āˆˆ B, Ī›x = limnā†’āˆž Ī›n x = Ī› x, so Ī› = Ī› .



w

Lemma 2.51. Ī›n ā†’ Ī› implies Ī›n ā†’ Ī›. Proof. This follows from |Ī›n x āˆ’ Ī›x| = |(Ī›n āˆ’ Ī›)x| ā‰¤ Ī›n āˆ’ Ī›x.



The connections between weak-āˆ— convergence and boundedness are described in the following proposition. w

Proposition 2.52. If Ī›n ā†’ Ī›, then supnāˆˆN Ī›n  < āˆž and Ī› ā‰¤ lim inf Ī›n . nā†’āˆž

Proof. Pointwise convergence implies pointwise boundedness. Thus, by the uniform boundedness principle, Ī›n are uniformly bounded. Moreover, Ī›x = lim Ī›n x ā‰¤ lim inf Ī›n x nā†’āˆž

nā†’āˆž

āˆ€x āˆˆ B.



Specializing Lemma 2.46 to functionals immediately implies the following criterion for weak-āˆ— convergence: āˆ— Proposition 2.53. If a bounded sequence (Ī›n )āˆž n=1 in B converges on a dense subset of B, then it weak-āˆ— converges to some Ī› āˆˆ B āˆ— .

We now reach the main result of this section. Bounded sequences in Banach spaces do not have to have convergent subsequences. However: Theorem 2.54 (Separable Banachā€“Alaoglu theorem). Any bounded sequence in a separable Banach space has a weak-āˆ— convergent subsequence.

66

2. Banach spaces

āˆ— Proof. Let (Ī›n )āˆž n=1 be a bounded sequence in B . For any x āˆˆ B,

sup|Ī›n x| ā‰¤ supĪ›n x < āˆž, nāˆˆN

nāˆˆN

so the sequence Ī›n is pointwise bounded. By Lemma 2.13, there is a subsequence which converges pointwise on a dense subset of B. By Proposition 2.53, this subsequence is weak-āˆ— convergent.  This property can be described as weak-āˆ— sequential compactness. Indeed, the Banachā€“Alaoglu theorem in its more general form is stated as weak-āˆ— compactness of the closed ball {Ī› āˆˆ B āˆ— | Ī› ā‰¤ r} and, with that topological reformulation, holds also for nonseparable Banach spaces. As the last abstract topic, we describe an attempt to describe weak-āˆ— convergence in terms of a metric on B āˆ— ; this leads to an imperfect but still relevant description. We use a construction from metric space theory. A semimetric on a set X is a symmetric function d : X Ɨ X ā†’ [0, āˆž) which obeys the triangle inequality; in a semimetric, d(x, y) = 0 does not necessarily imply x = y. Lemma 2.55. If dk , k āˆˆ N are semimetrics on X, then so is d(x, y) =

āˆž 

min{2āˆ’k , dk (x, y)}.

(2.20)

k=1

Proof. For ļ¬xed k, let us prove that dĖœk (x, y) = min{2āˆ’k , dk (x, y)} obeys the triangle inequality. If |dk (x, y)| ā‰„ 2āˆ’k or |dk (y, z)| ā‰„ 2āˆ’k , then dĖœk (x, z) ā‰¤ 2āˆ’k ā‰¤ dĖœk (x, y) + dĖœk (y, z). If dĖœk (x, y) < 2āˆ’k and dĖœk (y, z) < 2āˆ’k , then dĖœk (x, z) ā‰¤ dk (x, z) ā‰¤ dk (x, y) + dk (y, z) = dĖœk (x, y) + dĖœk (y, z). Since dĖœk is also symmetric, it is a semimetric. The sum (2.20) is convergent due to the upper bound by 2āˆ’k , and as a sum of symmetric functions which obey the triangle inequality, it has the same properties.  Theorem 2.56. If {xk }āˆž k=1 is a dense sequence in a separable Banach space B, then āˆž   min{2āˆ’k , |(Ī› āˆ’ Ī› )xk |} (2.21) d(Ī›, Ī› ) = k=1

Bāˆ— .

Moreover, consider an arbitrary sequence of Ī›n āˆˆ deļ¬nes a metric on B āˆ— . This sequence is weak-āˆ— convergent to Ī› āˆˆ B āˆ— if and only if it is bounded and d(Ī›n , Ī›) ā†’ 0 as n ā†’ āˆž.

2.5. Weak-āˆ— convergence and the separable Banachā€“Alaoglu theorem

67

Proof. (a) By Lemma 2.55, d is a semimetric. Assume d(Ī›, Ī› ) = 0; then Ī›xk = Ī› xk for all k, so by density of {xk }, Ī› = Ī› . Thus, d is a metric. w

(b) If Ī›n ā†’ Ī›, then Ī›n is a bounded sequence in B āˆ— . Moreover, (Ī›n āˆ’ Ī›)xk ā†’ 0 for every k, so by dominated convergence with dominating sequence 2āˆ’k applied to the counting measure on N, lim d(Ī›n , Ī›) =

nā†’āˆž

āˆž  k=1

lim min{2āˆ’k , |(Ī› āˆ’ Ī› )xk |} = 0.

nā†’āˆž

Conversely, if d(Ī›n , Ī›) ā†’ 0, then min(2āˆ’k , |(Ī›n āˆ’ Ī›)xk |) ā†’ 0 for each k, so (Ī›n āˆ’ Ī›)xk ā†’ 0 for each k. Thus, Ī›n converge to Ī› on a dense set. Using w  Lemma 2.46, since Ī›n are uniformly bounded, Ī›n ā†’ Ī›. Thus, if we restrict to a bounded subset of B āˆ— such as {Ī› āˆˆ B āˆ— | Ī› ā‰¤ r}, weak-āˆ— convergence can be interpreted as convergence with respect to a metric. However, the restriction to a bounded subset was crucial here: on the entire set B āˆ— , there is no metric which precisely gives weak-āˆ— convergence (in topological language: the topology of weak-āˆ— convergence is not metrizable) unless B is ļ¬nite dimensional. We will not prove or use that fact. The notion of weak-āˆ— convergence on B āˆ— is tied to the original Banach space B, so it is imprecise to discuss it without specifying the space B. This is nonetheless common practice for some common Banach spaces. It is also common in some cases to refer to this notion as weak convergence, although weak convergence is in general a diļ¬€erent concept. A ubiquitous special case is obtained from B = C(K), for K a compact metric space, using the correspondence from the Rieszā€“Markov theorem: Deļ¬nition 2.57. We denote by M(K) the set of ļ¬nite positive Borel measures on a compact metric space K. The measures Ī¼n āˆˆ M(K) converge weakly to Ī¼ āˆˆ M(K) if   āˆ€f āˆˆ C(K). f dĪ¼n ā†’ f dĪ¼ w

We denote this by Ī¼n ā†’ Ī¼ or w-limnā†’āˆž dĪ¼n = dĪ¼. Corollary 2.58. Any sequence of Ī¼n āˆˆ M(K) such that supnāˆˆN Ī¼n (K) < āˆž has a weakly convergent subsequence. Proof. The measures Ī¼n correspond to positive linear functionals Ī›n (f ) = f dĪ¼n on C(K). Since Ī›n  = Ī¼n (K) is uniformly bounded and C(K) is separable, by the Banachā€“Alaoglu theorem there is a subsequence such that w Ī›nl ā†’ Ī› as l ā†’ āˆž for some Ī› āˆˆ C(K)āˆ— . When f ā‰„ 0, Ī›nl f ā‰„ 0 for all l, so Ī›f ā‰„ 0. Thus Ī› is also a positive linear functional, so it is of the form w Ī›f = f dĪ¼ for a positive Borel measure Ī¼, and Ī¼nl ā†’ Ī¼, by deļ¬nition. 

68

2. Banach spaces

2.6. Banach-space valued integration In this section, we present some basics of Banach-space valued integration. Let A be a Ļƒ-algebra on X, let Ī¼ be a measure on A, and let B be a Banach space. A simple function s : X ā†’ B can be written as s=

n 

yj Ļ‡Aj

(2.22)

j=1

with Aj āˆˆ A, yj āˆˆ B. If can deļ¬ne



s dĪ¼ < āˆž, then Ī¼(Aj ) < āˆž for each j, so we  s dĪ¼ =

n 

Ī¼(Aj )yj .

(2.23)

j=1

This integral is additive and % % % %  n n %   % % % % % % s dĪ¼% = % Ī¼(Aj )yj % ā‰¤ Ī¼(Aj )yj  = s dĪ¼. % % % % j=1

j=1

The Bochner integral is developed for functions which can be approximated by simple functions: Deļ¬nition 2.59. A measurable function f : X ā†’ B is said to be Bochnerintegrable if there exists a sequence of simple functions sn : X ā†’ B such  that sn  dĪ¼ < āˆž for every n and  (2.24) lim f āˆ’ sn  dĪ¼ = 0. nā†’āˆž

For any Bochner-integrable function, its Bochner integral is deļ¬ned as   (2.25) sn dĪ¼. f dĪ¼ = lim nā†’āˆž

Since %  %    % % % sm dĪ¼ āˆ’ sn dĪ¼% ā‰¤ sm āˆ’ sn  dĪ¼ ā‰¤ f āˆ’ sm  dĪ¼ + f āˆ’ sn  dĪ¼, % %  Equation (2.24) implies that the sequence sn dĪ¼ is Cauchy in B and therefore convergent. Similar arguments prove that the limit is independent of the choice of sequence sn , so the Bochner integral (2.25) is well deļ¬ned. The setting of Lemma 2.6 can be understood as a Bochner integral with respect to counting measure on N. We note another special case, with respect to Lebesgue measure m on an interval [a, b]:

2.6. Banach-space valued integration

69

Example 2.60. Any continuous f : [a, b] ā†’ B is Bochner-integrable and, b denoting by a f (x) dx its Bochner integral, 1 f lim nā†’āˆž n nāˆ’1 j=0

j a + (b āˆ’ a) n

!



b

=

f (x) dx.

(2.26)

a

Proof. Since f is continuous, by the same proof as in the scalar case, it is uniformly continuous. Thus, the simple functions ( )! xāˆ’a 1 sn (x) = f a + n n bāˆ’a satisfy the condition (2.24), which leads to (2.26).



It is often useful to reduce to scalar objects. To do this in the Banach space setting, the key tool is the following property: Lemma 2.61. For any Banach space B, if we denote B1āˆ— = {Ī› āˆˆ B āˆ— | Ī› = 1}, then x = sup |Ī›x| Ī›āˆˆB1āˆ—

āˆ€x āˆˆ B.

(2.27)

This includes the fact that functionals separate points: if Ī›x = Ī›y for all Ī› āˆˆ B1āˆ— , then by (2.27), x āˆ’ y = 0, so x = y. The reader can proceed in two ways. One is to note that this is an immediate corollary of the Hahnā€“Banach theorem. Instead of proving this in an abstract Banach space setting, we instead observe that in all the cases of interest to us, it is easy to verify the statement manually. To verify (2.27) for some Banach space B, it is not even necessary to have a complete description of the dual space B āˆ— ; every Ī› āˆˆ B1āˆ— obeys |Ī›x| ā‰¤ x, so supĪ›āˆˆB1āˆ— Ī›x ā‰¤ x is trivial and it suļ¬ƒces to show the opposite inequality by using some elements of B1āˆ— . For instance, for Lp (X, dĪ¼) it follows from Corollary 2.27 (with the roles of p and q reversed) that any g āˆˆ Lq (X, dĪ¼) induces a Ī› āˆˆ (Lp (X, dĪ¼))āˆ— by  Ī›f = gĀÆf dĪ¼ and that Ī› = gq . Using Corollary 2.27 a second time, it follows that (2.27) holds for Lp (X, dĪ¼). Other cases will appear later: when B = H is a Hilbert space, the property (2.27) will follow from the Riesz representation theorem and when B = L(H), the property (2.27) will follow from Exercise 4.14.

70

2. Banach spaces

Since a functional can always be multiplied by eiĻ† without changing its norm, (2.27) implies x = sup Re Ī›x Ī›āˆˆB1āˆ—

āˆ€x āˆˆ B.

(2.28)

Lemma 2.62. If f : X ā†’ B is Bochner-integrable, then:   Ī› f dĪ¼ = Ī›f dĪ¼ āˆ€Ī› āˆˆ B āˆ— % %  % % % f dĪ¼% ā‰¤ f  dĪ¼. % % Proof. For a simple function (2.22), by linearity of Ī›,   n n   Ī¼(Aj )yj = Ī¼(Aj )Ī›yj = Ī›s dĪ¼. Ī› s dĪ¼ = Ī› j=1

j=1

For a sequence of simple functions sn obeying (2.24), use         Ī›f dĪ¼ āˆ’ Ī›sn dĪ¼ ā‰¤ |Ī›f āˆ’ Ī›sn | dĪ¼ ā‰¤ |Ī›(f āˆ’ sn )| dĪ¼ ā‰¤ f āˆ’ sn  dĪ¼,     so Ī›sn dĪ¼ ā†’ Ī›f dĪ¼, and the continuity of Ī› implies      sn dĪ¼ = lim Ī› sn dĪ¼ = lim Ī›sn dĪ¼ = Ī›f dĪ¼. Ī› f dĪ¼ = Ī› lim nā†’āˆž

nā†’āˆž

nā†’āˆž

To prove the second claim, we note that for any Ī› āˆˆ B1āˆ— ,            Ī› f dĪ¼ =  Ī›f dĪ¼ ā‰¤ |Ī›f | dĪ¼ ā‰¤ f  dĪ¼.     Taking the sup over Ī› āˆˆ B1āˆ— concludes the proof by (2.27).



Remark 2.63. For a diļ¬€erent approach to integration, assume that f : X ā†’ B is such that there exists I āˆˆ B so that for all Ī› āˆˆ B āˆ— , Ī›f : X ā†’ C is measurable and  Ī›I =

Ī›f dĪ¼.

Then the value I is called the Pettis integral of f . Existence of the Pettis integral is a strictly weaker notion than existence of the Bochner integral. Deļ¬nition 2.64. Let F : I ā†’ B where I āŠ‚ R is an interval and B a Banach space. The derivative of F at x0 āˆˆ int I, if it exists, is F  (x0 ) āˆˆ B such that F (x0 + h) āˆ’ F (x0 ) āˆ’ hF  (x0 ) = 0. hā†’0 |h| lim

If I is open and F is diļ¬€erentiable at every point in I, then F  is also a function from I to B. Based on this notion of derivative, C n (I, B) for n āˆˆ N āˆŖ {āˆž} is deļ¬ned analogously to the scalar case.

2.7. Banach-space valued analytic functions

71

Theorem 2.65 (Fundamental theorem of calculus). (a) If f āˆˆ C([a,  x b], B), then the function F : [a, b] ā†’ B deļ¬ned by F (x) = a f (t) dt is continuous on [a, b], diļ¬€erentiable on (a, b), and F  (x) = f (x) for all x āˆˆ (a, b). (b) If G, f āˆˆ C([a, b], B) and G (x) = f (x) for x āˆˆ (a, b), then  b f (t) dt = G(b) āˆ’ G(a). a

The proof of the fundamental theorem of calculus follows the same steps as in the scalar case, with one additional ingredient. In the scalar case, the following key step follows from the mean value theorem; in the Banach-space valued case, one uses (2.28) to reduce to the scalar case. Lemma 2.66. If g āˆˆ C([a, b], B) and g  = 0 on (a, b), then g(b) = g(a). Proof. For any Ī› āˆˆ B āˆ— , Re(Ī›g) āˆˆ C([a, b], R) and (Re Ī›g) = Re Ī›g  = 0 for x āˆˆ (a, b), so Re Ī›g(b) = Re Ī›g(a) by the mean value theorem. Since  Re Ī›(g(b) āˆ’ g(a)) = 0 for all Ī› āˆˆ B āˆ— , (2.28) implies g(b) āˆ’ g(a) = 0.

2.7. Banach-space valued analytic functions We now consider complex-analytic calculus for Banach-space valued functions. We will use the customary notation in C, Dr (z0 ) = {z āˆˆ C | |z āˆ’ z0 | < r}. If Ī© āŠ‚ C is an open set, Ī³ : [a, b] ā†’ Ī© is a *C 1 contour, and f : Ī© ā†’ B is a continuous map, then the contour integral Ī³ f (z) dz is deļ¬ned by  b + f (z) dz = Ī³  (t)f (Ī³(t)) dt. Ī³

a

Theorem 2.67. Let Ī© āŠ‚ C be an open set, let B be a Banach space, and let f : Ī© ā†’ B. The following are equivalent: (a) Holomorphicity: For every z0 āˆˆ Ī© there is a value of f  (z0 ) āˆˆ B such that f (z) āˆ’ f (z0 ) āˆ’ (z āˆ’ z0 )f  (z0 ) = 0; (2.29) lim zā†’z0 z āˆ’ z0 (b) Weak analyticity: For every Ī› āˆˆ B āˆ— , the function Ī›f : Ī© ā†’ C is analytic; (c) Cauchyā€™s integral formula: f is continuous and, for every disk Dr (z0 ) āŠ‚ Ī© and every z āˆˆ Dr (z0 ), + f (w) 1 dw; (2.30) f (z) = 2Ļ€i |wāˆ’z0 |=r w āˆ’ z

72

2. Banach spaces

(d) Local representability by power series: For every z0 āˆˆ Ī©, there is a neighborhood Dr (z0 ) āŠ‚ Ī© and coeļ¬ƒcients Fn āˆˆ B such that āˆž n n=0 r Fn  < āˆž and f (z) =

āˆž 

(z āˆ’ z0 )n Fn

āˆ€z āˆˆ Dr (z0 ).

(2.31)

n=0

Proof. (a) =ā‡’ (b): Applying arbitrary Ī› āˆˆ B āˆ— to (2.29) shows that Ī›f : Ī© ā†’ C is holomorphic with (Ī›f ) = Ī›f  . Thus, Ī›f is analytic. (b) =ā‡’ (c): We begin by proving continuity of f at an arbitrary point z0 āˆˆ Ī©. Fix r > 0 such that Dr (z0 ) āŠ‚ Ī©. For every Ī› āˆˆ B āˆ— , analyticity of Ī›f implies that    Ī›f (z) āˆ’ Ī›f (z0 )    0, Dr (z0 ) āˆ© Ī© = Dr (z0 ) \ {z0 } and the limit limzā†’z0 f (z) is convergent. If f has a removable singularity at z0 , prove that deļ¬ning f at z0 by f (z0 ) = limzā†’z0 f (z) gives an analytic function on Ī© āˆŖ {z0 }.

Chapter 3

Hilbert spaces

The distinguishing feature of Hilbert spaces is inner product, which is the n x inspired by the dot product in Rn , x Ā· y = j=1 j yj . The dot product takes a central place in Euclidean geometry, and it can be used to recover, among other things, the length of vectors, x2 = xĀ·x. When generalizing to Cn , to retain the connection with Euclidean norm, one must conjugate the entries of one of the vectors, leading to the deļ¬nition of the inner product n ĀÆj yj (it is a matter of convention whether complex on C , x, y = nj=1 x conjugation is applied on the ļ¬rst or second vector). We will consider Hilbert spaces over the ļ¬eld of scalars C, which arise as an abstract generalization of this inner product on Cn . The analystā€™s main interest in Hilbert spaces lies in inļ¬nite-dimensional settings, but given the original dot product on Rn , it should not be surprising that a geometric intuition is very useful in the study of Hilbert spaces. The reader should compare the knowledge and intuition from linear algebra with what is presented here, and note the more subtle phenomena which only appear in inļ¬nite-dimensional settings.

3.1. Inner products In this section, we will start with the general setting of sesquilinear forms and gradually specialize to the setting of inner products on Hilbert spaces. Deļ¬nition 3.1. Let V be a vector space. A sesquilinear form on V is a map Ā·, Ā· : V Ɨ V ā†’ C with the following properties: 77

78

3. Hilbert spaces

(a) Linearity in the second parameter : For all x, y, y  āˆˆ V and Ī» āˆˆ C, x, Ī»y = Ī»x, y, x, y + y   = x, y + x, y  . (b) Conjugate-linearity in the ļ¬rst parameter : For all x, x , y āˆˆ V and Ī» āˆˆ C, ĀÆ y, Ī»x, y = Ī»x, x + x , y = x, y + x , y. It is a matter of convention which of the two parameters is linear, since we could exchange the two parameters in all forms throughout the text. A sesquilinear form can be recovered by its values on the diagonal x, x with x āˆˆ V : Lemma 3.2 (Polarization identity). For all x, y āˆˆ V ,  1 Ļ‰ āˆ’1 x + Ļ‰y, x + Ļ‰y. x, y = 4

(3.1)

Ļ‰āˆˆ{1,i,āˆ’1,āˆ’i}

Proof. By sesquilinearity, for any Ļ‰ āˆˆ C with |Ļ‰| = 1, Ļ‰ āˆ’1 x + Ļ‰y, x + Ļ‰y = Ļ‰ āˆ’1 x, x + x, y + Ļ‰ āˆ’2 y, x + Ļ‰ āˆ’1 y, y. The proof is completed by summing over Ļ‰ āˆˆ {1, i, āˆ’1, āˆ’i} using

 4 k āˆˆ 4Z k Ļ‰ = 0 k āˆˆ Z \ 4Z. Ļ‰āˆˆ{1,i,āˆ’1,āˆ’i}



Corollary 3.3. If a sesquilinear form obeys x, x āˆˆ R for all x āˆˆ V , then it is conjugate symmetric, i.e., y, x = x, y

āˆ€x, y āˆˆ V.

(3.2)

Proof. Using properties of the sesquilinear form, y + Ļ‰x, y + Ļ‰x = Ļ‰(x + Ļ‰ āˆ’1 y), Ļ‰(x + Ļ‰ āˆ’1 y) = Ļ‰Ļ‰x + Ļ‰ āˆ’1 y, x + Ļ‰ āˆ’1 y. Multiplying by Ļ‰ āˆ’1 , summing over Ļ‰ āˆˆ {1, i, āˆ’1, āˆ’i}, taking the complex conjugate, and using the polarization identity gives (3.2).  We will now impose a positive deļ¬nite condition and show that it naturally turns V into a normed vector space: Deļ¬nition 3.4. Let V be a vector space. An inner product on V is a sesquilinear map Ā·, Ā· : V Ɨ V ā†’ C which is positive deļ¬nite, i.e., x, x > 0

āˆ€x āˆˆ V \ {0}.

(3.3)

3.1. Inner products

79

The induced norm of an inner product is deļ¬ned by . x = x, x.

(3.4)

Soon, we will prove that the induced norm is, indeed, a norm. Condition (3.3) is directly motivated by this. We begin by proving some other general properties of inner products. Lemma 3.5. For all x āˆˆ V and Ī» āˆˆ C, Ī»x = |Ī»|x. ĀÆ Proof. This follows from Ī»x, Ī»x = Ī»Ī»x, x.



Deļ¬nition 3.6. Vectors x1 , x2 āˆˆ V are orthogonal if x1 , x2  = 0; this is denoted x1 āŠ„ x2 . A sequence (xj )nj=1 in V is called (pairwise) orthogonal if xj āŠ„ xk whenever j = k. Lemma 3.7 (Pythagorean theorem). If x1 , . . . , xn āˆˆ V are pairwise orthogonal, then %2 % n n % %  % % xj % = xj 2 . % % % j=1

j=1

Proof. Using sesquilinearity of the inner product, xj , xk  = 0 for j = k, and xj , xk  = xj 2 , we compute %2 / % 0 n n n n  n n % %     % % xj % = xj , xk = xj , xk  = xj 2 .  % % % j=1

j=1

k=1

j=1 k=1

j=1

The Pythagorean theorem has obvious roots in Euclidean geometry. Similarly, the following lemma describes orthogonal projection of a vector x to a vector y = 0. Lemma 3.8. For any x, y āˆˆ V with y = 0, there is a unique Ī» āˆˆ C such that x āˆ’ Ī»y āŠ„ y, and it is given by Ī» = y, x/y2 . Proof. By linearity of the inner product in the second parameter, the equation y, xāˆ’Ī»y = 0 is equivalent to y, xāˆ’Ī»y, y = 0, which has the unique  solution Ī» = y, x/|y2 . Theorem 3.9 (Cauchyā€“Schwarz inequality). For all x, y āˆˆ V , |x, y| ā‰¤ xy.

(3.5)

Proof. If y = 0, (3.5) holds trivially. For y = 0, we use Ī» from Lemma 3.8 and z = x āˆ’ Ī»y. Since y, Ī»z = Ī»y, z = 0, by the Pythagorean theorem, x2 = z2 + Ī»y2 ā‰„ Ī»y2 = |Ī»|2 y2 =

|y, x|2 . y2

Multiplying by y2 completes the proof since |y, x| = |x, y|.



80

3. Hilbert spaces

Proposition 3.10. For all x, y āˆˆ V , x + y ā‰¤ x + y. Proof. By the deļ¬nition of induced norm and conjugate symmetry, x + y2 = x2 + x, y + y, x + y2 = x2 + 2 Rex, y + y2 . Estimating Rex, y ā‰¤ |x, y| ā‰¤ x y gives x + y2 ā‰¤ x2 + 2x y + y2 = (x + y)2 .



Collecting Lemma 3.5 and Proposition 3.10, we obtain: Corollary 3.11. The induced norm (3.4) is a norm. Deļ¬nition 3.12. A vector space with an inner product is called a Hilbert space if it is complete with respect to the induced norm. In particular, every Hilbert space is a Banach space with the induced norm. It is common to denote a Hilbert space by H instead of V . Example 3.13. For any n āˆˆ N, Cn is a Hilbert space with the inner product w, z =

n 

wj zj .

j=1

Proof. It is trivial that this is an inner product; the induced norm is pre cisely the norm Ā·2 encountered in Example 2.3. Example 3.14. The set of square-summable sequences āŽ§ āŽ« 1/2  āŽØ āŽ¬    |zĪ³ |2 0. Now we consider f (t) = x āˆ’ y āˆ’ tv2 ,

t āˆˆ R.

For all t, y + tv āˆˆ S so f (t) ā‰„ c2 = f (0). Thus, f has a global minimum at 0. However, by expanding in terms of inner products, we write f (t) = x āˆ’ y2 āˆ’ 2x āˆ’ y, vt + v2 t2 and compute f  (0) = 2x āˆ’ y, v > 0, which contradicts a minimum at 0.



Before deriving further properties of orthogonal projections, we will take a detour, using the projection theorem to describe the space of bounded linear functionals on H. Immediately from the deļ¬nition, for any subspaces S, T , S āŠ‚ T implies āŠ‚ S āŠ„ . Some other general properties are given in Exercise 3.9. We prove a criterion for a subspace of H, not necessarily closed, to be dense in H:

TāŠ„

Corollary 3.23. A subspace S of H is dense in H if and only if S āŠ„ = {0}. Proof. Assume S = H. Fix x āˆˆ H\S and let y be the orthogonal projection of x to S. Then y āˆˆ S so x = y. Thus, x āˆ’ y is a nonzero element of (S)āŠ„ āŠ‚ S āŠ„ . Conversely, assume z āˆˆ S āŠ„ and z = 0. Then for any y āˆˆ S, / S.  y āˆ’ z2 = y2 + z2 ā‰„ z2 , so z āˆˆ Recall that Hāˆ— denotes the set of bounded linear functionals on H, i.e., the set of bounded linear operators from H to C. On a Hilbert space, using the inner product, it is easy to generate many bounded linear functionals:

84

3. Hilbert spaces

Lemma 3.24. For any y āˆˆ H, Ī›x = y, x

(3.8)

deļ¬nes a bounded linear functional Ī› āˆˆ Hāˆ— with norm Ī› = y. Proof. The map Ī› : H ā†’ C is linear because the inner product is linear in the second argument. By the Cauchyā€“Schwarz inequality, for any x āˆˆ H, |Ī›x| = |y, x| ā‰¤ yx.

(3.9)

Thus, Ī› is bounded and Ī› ā‰¤ y. If y = 0, this gives Ī› = 0. If y = 0, equality holds in (3.9) for x = y, so Ī› ā‰„ y. The two inequalities combine to give Ī› = y.  Remarkably, this construction provides all bounded linear functionals on H: Theorem 3.25 (Riesz representation theorem). For every Ī› āˆˆ Hāˆ— , there is a unique y āˆˆ H such that (3.8) holds for all x āˆˆ H. Proof. Let Ī› āˆˆ Hāˆ— . If Ī› = 0, (3.8) holds with y = 0. Otherwise, Ker Ī› is a closed subspace of H and Ker Ī› = H, so by Corollary 3.23, there exists z āˆˆ (Ker Ī›)āŠ„ , z = 0. For any x āˆˆ H, the calculation ! Ī›x Ī›x z = Ī›x āˆ’ Ī›z = 0 Ī› xāˆ’ Ī›z Ī›z shows that x āˆ’

Ī›x Ī›z z

is in Ker Ī› and therefore orthogonal to z. This implies 1 2 Ī›x Ī›x z, x āˆ’ z, z = z, x āˆ’ z = 0. Ī›z Ī›z

Solving for Ī›x gives Ī›z z, x = Ī›x = z2

1

2 Ī›z z, x , z2

which is a representation for Ī› of the desired form. If the same functional Ī› can also be represented in the form Ī›x = Ėœ y , x, subtracting gives the zero functional (Ī› āˆ’ Ī›)(x) = y āˆ’ yĖœ, x. It follows that y āˆ’ yĖœ = Ī› āˆ’ Ī› = 0 so y = yĖœ, which proves uniqueness of y.  We now return to orthogonal projections; existence and uniqueness allow us to view orthogonal projection as a map on H: Deļ¬nition 3.26. If S is a closed subspace of H, the orthogonal projection to S is the map P : H ā†’ H deļ¬ned so that for every x āˆˆ H, P x is the unique vector such that P x āˆˆ S and x āˆ’ P x āˆˆ S āŠ„ .

3.2. Subspaces and orthogonal projections

85

Proposition 3.27. The orthogonal projection P to a closed subspace S of H has the following properties: (a) P is a bounded linear operator on H, i.e., P āˆˆ L(H); (b) Ker P = S āŠ„ and Ran P = S; (c) P  = 1 if S = {0}; (d) P P = P ; (e) x, P y = P x, y for all x, y āˆˆ H. Proof. For any x, x āˆˆ H, P x + P x āˆˆ S and (x + x ) āˆ’ (P x + P x ) = (x āˆ’ P x) + (x āˆ’ P x ) āˆˆ S āŠ„ , so by uniqueness, P (x + x ) = P x + P x . Similarly, cP x āˆˆ S and cx āˆ’ cP x = c(x āˆ’ P x) āˆˆ S āŠ„ implies P (cx) = cP x for all c āˆˆ C. Thus, P is linear. By the Pythagorean theorem, x2 = P x2 + x āˆ’ P x2 ā‰„ P x2 , so P is bounded and P  ā‰¤ 1. In particular, P āˆˆ L(H). For any x āˆˆ S, P x = x because x āˆˆ S and x āˆ’ x āˆˆ S āŠ„ . Thus, if S = {0}, P  = 1. For any x āˆˆ H, P x āˆˆ S implies P P x = P x by the above. Thus, P P = P and S āŠ‚ Ran P . Since Ran P āŠ‚ S, we conclude Ran P = S. By deļ¬nition, P x = 0 if and only if x āˆˆ S āŠ„ , so Ker P = S āŠ„ . For all x, y āˆˆ H, x āˆ’ P x āˆˆ S āŠ„ and P y āˆˆ S imply x āˆ’ P x, P y = 0, so x, P y = P x, P y. Analogously, P x, P y = P x, y, so x, P y = P x, y.  This has a dual point of view, in which orthogonal projections are deļ¬ned without reference to a subspace as operators with certain properties: Deļ¬nition 3.28. An operator P āˆˆ L(H) is called an orthogonal projection if P 2 = P and u, P v = P u, v āˆ€u, v āˆˆ H. (3.10) This deļ¬nition is compatible with earlier terminology and completes a correspondence between closed subspaces of H and orthogonal projections as a family of operators in L(H): Proposition 3.29. Let P āˆˆ L(H) be an orthogonal projection. Then Ran P is a closed subspace of H and P is the orthogonal projection to Ran P . Proof. If x āˆˆ Ran P , then x = P y for some y, so P x = P 2 y = P y = x. Conversely, if P x = x, then x āˆˆ Ran P . This proves that Ran P = Ker(I āˆ’ P ), and in particular, Ran P is closed. For any x, y āˆˆ H, (I āˆ’ P )x, P y = P (I āˆ’ P )x, y = (P āˆ’ P 2 )x, y = 0, so (I āˆ’ P )x āˆˆ S āŠ„ . Since P x āˆˆ S and x āˆ’ P x āˆˆ S āŠ„ , it follows that P is the orthogonal projection to the subspace S. 

86

3. Hilbert spaces

The projection theorem is an existence and uniqueness result, but orthogonal projection can often be computed. By Lemma 3.8, orthogonal projection to a one-dimensional subspace span{y}, where y = 0, is Px =

y, x y. y2

We can view one-dimensional subspaces Sj = span{yj } as a motivating special case for the following results: Theorem 3.30. Let S1 , . . . , Sn be mutually orthogonal closed subspaces of H and let Pj denote orthogonal projection to Sj . Then the subspace S = span

n 

Sj

j=1

is a closed and orthogonal projection to S, which is given by n  Pj x. Px =

(3.11)

j=1

Moreover, for all x āˆˆ H, n 

Pj x2 ā‰¤ x2

(3.12)

j=1

with equality if and only if x āˆˆ S; this is known as Besselā€™s inequality. Proof. Fix x āˆˆ H and denote y = nj=1 Pj x āˆˆ S. Fix k. For all j = k, Sj āŠ‚ SkāŠ„ implies Pk Pj = 0, so by properties of Pk , Pk y =

n 

Pk Pj x = Pk Pk x = Pk x.

j=1

Thus, z, x āˆ’ y = 0 if z āˆˆ Sk for some k. By linearity, z, x āˆ’ y = 0 for all z āˆˆ S, and by continuity, z, x āˆ’ y = 0 for all z āˆˆ S. Thus, y āˆˆ S āŠ‚ S and āŠ„ x āˆ’ y āˆˆ S , so P = nj=1 Pj is an orthogonal projection to S. Moreover, for x āˆˆ S, x = P x āˆˆ S, so S āŠ‚ S; thus, S = S is closed. The vectors Pj x āˆˆ Sj for j = 1, . . . , n and the vector x āˆ’ P x āˆˆ S āŠ„ are pairwise orthogonal and their sum is x, so by the Pythagorean theorem, n  Pj x2 + x āˆ’ P x2 . x2 = j=1

This implies Besselā€™s inequality, with equality if and only if x = P x.



Next, we consider the setting where there are iniļ¬nitely many closed subspaces SĪ³ , indexed by an abstract index Ī³ āˆˆ Ī“.

3.2. Subspaces and orthogonal projections

87

Theorem 3.31. Let SĪ³ , Ī³ āˆˆ Ī“, be mutually orthogonal closed subspaces of H, and let PĪ³ denote orthogonal projection to SĪ³ . Then:  (a) S = span Ī³āˆˆĪ“ SĪ³ is a closed subspace of H. (b) For any x āˆˆ H, the set {Ī³ āˆˆ Ī“ | PĪ³ x = 0} is countable. (c) For any x āˆˆ H and any injective map Ļƒ : N ā†’ Ī“ such that {Ī³ āˆˆ Ī“ | PĪ³ x = 0} āŠ‚ Ļƒ(N), the orthogonal projection of x to S is given by Px =

āˆž 

PĻƒ(j) x

(3.13)

j=1

(in particular, the series is convergent in H). (d) For all x āˆˆ H,



PĪ³ x2 ā‰¤ x2

(3.14)

Ī³āˆˆĪ“

with equality if and only if x āˆˆ S. Proof. By Theorem 3.31, for any distinct Ī³1 , . . . , Ī³n āˆˆ Ī“, n  PĪ³j x2 ā‰¤ x2 . j=1

Viewing these ļ¬nite sums as integrals of simple functions on Ī“ with respect to counting measure and taking the supremum over all ļ¬nite subsets of Ī“,  PĪ³ x2 ā‰¤ x2 . Ī³āˆˆĪ“

By Markovā€™s inequality (Lemma 1.56), for any k āˆˆ N, the set {Ī³ āˆˆ Ī“ | PĪ³ x2 ā‰„ 1/k} is ļ¬nite, so the set {Ī³ āˆˆ Ī“ | PĪ³ x = 0} is countable. By Lemma 3.16, the series in (3.13) is convergent. Denote by y the value of the series. The vector y is in S. Using linearity and continuity of PĪ³ ,

āˆž  PĪ³ x Ī³ āˆˆ Ļƒ(N) PĪ³ PĻƒ(j) x = PĪ³ y = 0 Ī³āˆˆ / Ļƒ(N). j=1 Thus, z, x āˆ’y = 0 if z āˆˆ SĪ³ for some Ī³. By linearity, z, x āˆ’ y = 0 for all z āˆˆ span Ī³āˆˆĪ“ SĪ³ , and by continuity, z, x āˆ’ y = 0 for all z āˆˆ S. Thus, y āˆˆ S and x āˆ’ y āˆˆ S āŠ„ . Thus, y is the orthogonal projection of x to S. For any n āˆˆ N, by the proof of Theorem 3.31, % %2 n n % %   % % PĻƒ(j) x2 + %x āˆ’ PĻƒ(j) x% . x2 = % % j=1

j=1

88

3. Hilbert spaces

Taking n ā†’ āˆž implies x2 =

āˆž 

PĻƒ(j) x2 + x āˆ’ P x2 ,

j=1

so (3.14) holds, with equality if and only if x = P x.



Remark 3.32. Theorem 3.31(b) shows that such a map Ļƒ exists, and part (c) shows that the right-hand side of (3.13) is independent of the choice of map Ļƒ, so we will denote it more concisely by  PĪ³ x. Px = Ī³āˆˆĪ“

3.3. Direct sums of Hilbert spaces Any closed subspace S of a Hilbert space H is also a Hilbert space, since the restriction of the inner product on H is an inner product of S and since a closed subset of a complete space is complete. By the projection theorem, any vector v can be uniquely decomposed as v = P v +(v āˆ’P v) with P v āˆˆ S, v āˆ’ P v āˆˆ S āŠ„ . With respect to this decomposition, inner products can be computed as v, w = P v, P w + v āˆ’ P v, w āˆ’ P w. Thus, the projection theorem can be viewed as a decomposition of the Hilbert space H into Hilbert spaces S and S āŠ„ . This motivates a construction which creates, from two Hilbert spaces H1 and H2 , a new Hilbert space whose vectors are formal sums (or, more formally, ordered pairs) of vectors in H1 and H2 and whose inner product is the sum of inner products in H1 and H2 . The resulting space is called the direct sum of Hilbert spaces H1 and H2 and is denoted H1 āŠ• H2 . This construction could then be iterated or generalized to the construction of a direct sum of n Hilbert spaces, H1 āŠ• H2 āŠ• Ā· Ā· Ā· āŠ• Hn . Instead of doing this, we will present a further generalization right away: the direct sum of an arbitrary (possibly inļ¬nite) family of Hilbert spaces. We will need this level of generality in order to state the spectral theorem for self-adjoint operators. Deļ¬nition 3.33. Given Hilbert spaces HĪ³ , Ī³ āˆˆ Ī“, we deļ¬ne their direct sum as the space

   3  HĪ³ = (vĪ³ )Ī³āˆˆĪ“  vĪ³ āˆˆ HĪ³ for all Ī³ āˆˆ Ī“ and vĪ³ 2 < āˆž (3.15) Ī³āˆˆĪ“

Ī³āˆˆĪ“

with addition and scalar multiplication given by (uĪ³ )Ī³āˆˆĪ“ + (vĪ³ )Ī³āˆˆĪ“ = (uĪ³ + vĪ³ )Ī³āˆˆĪ“ ,

Ī» (uĪ³ )Ī³āˆˆĪ“ = (Ī»uĪ³ )Ī³āˆˆĪ“ ,

3.3. Direct sums of Hilbert spaces

89

and inner product given by (uĪ³ )Ī³āˆˆĪ“ , (vĪ³ )Ī³āˆˆĪ“  =



uĪ³ , vĪ³ .

(3.16)

Ī³āˆˆĪ“

Note 4 that if Ī“ is ļ¬nite, the summability condition in (3.15) is trivial, and Ī³āˆˆĪ“ HĪ³ consists of arbitrary N -tuples with vĪ³ āˆˆ HĪ³ . If we view C as a Hilbert space with inner product z, w = zĀÆw and set HĪ³ = C for all Ī³, as a special case of the above construction we obtain: 4 Example 3.34. For any set Ī“, Ī³āˆˆĪ“ C = 2 (Ī“). In particular, the following theorem contains an independent proof of completeness of 2 (Ī“). Theorem 3.35. For any family of Hilbert spaces HĪ³ , Ī³ āˆˆ Ī“: 4 (a) The direct sum Ī³āˆˆĪ“ HĪ³ is a Hilbert space.

4 (b) Vectors with ļ¬nitely many nonzero entries are dense in Ī³āˆˆĪ“ HĪ³ . 4 (c) If HĪ³ are separable spaces and Ī“ is countable, Ī³āˆˆĪ“ HĪ³ is separable. 4 Proof. (a) The set H = Ī³āˆˆĪ“ HĪ³ is obviously closed under scalar multiplication, and it is closed under addition due to   uĪ³ + vĪ³ 2 ā‰¤ (2uĪ³ 2 + 2vĪ³ 2 ). Ī³āˆˆĪ“

Ī³āˆˆĪ“

Thus, H is a vector space. By the Cauchyā€“Schwarz inequality,  1/2  1/2     2 2 |uĪ³ , vĪ³ | ā‰¤ uĪ³ vĪ³  ā‰¤ uĪ³  vĪ³  < āˆž, Ī³āˆˆĪ“

Ī³āˆˆĪ“

Ī³āˆˆĪ“

Ī³āˆˆĪ“

so the sum deļ¬ning the inner product is absolutely convergent; thus, (3.16) is well deļ¬ned as a map from H Ɨ H to C. It then follows that (3.16) is an inner product on H. The induced norm is, of course,  1/2  1/2   2 vĪ³ , vĪ³  = vĪ³  . (vĪ³ )Ī³āˆˆĪ“  = Ī³āˆˆĪ“

Ī³āˆˆĪ“

The next step is to prove that H is complete. Consider a Cauchy sequence of vectors vn = (vn,Ī³ )Ī³āˆˆĪ“ āˆˆ H, n āˆˆ N. We will prove that it is convergent. Since the sequence is Cauchy, it is bounded. For any Ī³ āˆˆ Ī“, it follows from vm,Ī³ āˆ’ vn,Ī³  ā‰¤ vm āˆ’ vn  that (vn,Ī³ )āˆž n=1 is a Cauchy sequence in HĪ³ , and therefore the limit wĪ³ = lim vn,Ī³ nā†’āˆž

(3.17)

90

3. Hilbert spaces

exists for each Ī³. By Fatouā€™s lemma applied to the counting measure,   wĪ³ 2 ā‰¤ lim inf vn,Ī³ 2 = lim inf vn 2 < āˆž, Ī³āˆˆĪ“

nā†’āˆž

Ī³āˆˆĪ“

nā†’āˆž

so w = (wĪ³ )Ī³āˆˆĪ“ āˆˆ H. Since (vn )āˆž n=1 is Cauchy, for any > 0 there exists n0 such that for all m, n ā‰„ n0 , vn āˆ’ vm 2 < 2 . By Fatouā€™s lemma, for n ā‰„ n0 ,   wĪ³ āˆ’ vn,Ī³ 2 ā‰¤ lim inf vm,Ī³ āˆ’ vn,Ī³ 2 = lim inf vm āˆ’ vn 2 ā‰¤ 2 . mā†’āˆž

Ī³āˆˆĪ“

nā†’āˆž

Ī³āˆˆĪ“

This means that limnā†’āˆž w āˆ’ vn 2 = 0, that is, vn ā†’ w in H. (b) For any v = (vĪ³ )Ī³āˆˆĪ“ āˆˆ H and any > 0, the condition Ī³āˆˆĪ“ vĪ³ 2 < āˆž implies that there is a ļ¬nite A āŠ‚ Ī“ such that  vĪ³ 2 < 2 . Ī³āˆˆĪ“\A

Thus, the vector w with wĪ³ = vĪ³ for Ī³ āˆˆ A and wĪ³ = 0 for Ī³ āˆˆ / A obeys v āˆ’ w < . (c) Let DĪ³ be some dense subsets of HĪ³ . Then, for the vector w from the proof of (b), we can denote K = #A and x āˆˆ H such that xĪ³ āˆ’wĪ³  < 2 /K / A. Then x āˆ’ w < . Thus, the set for Ī³ āˆˆ A and xĪ³ = 0 for Ī³ āˆˆ  {(xĪ³ )Ī³āˆˆĪ“ | xĪ³ āˆˆ DĪ³ for all Ī³ āˆˆ A and xĪ³ = 0 for all Ī³ āˆˆ / A} D= AāŠ‚Ī“ A ļ¬nite

is dense in H. If Ī“ is countable, then the set {A āŠ‚ Ī“ | A is ļ¬nite} is countable. If HĪ³ are separable, the sets DĪ³ can be chosen to be countable. Since ļ¬nite Cartesian products of countable sets are countable, their countable union D is a countable dense subset of H.  We originally motivated the direct sum construction through orthogonal subspaces of a single Hilbert space. But we then developed it in the diļ¬€erent setup of a sum of Hilbert spaces. We now revisit this construction in the special case of mutually orthogonal closed subspaces SĪ³ of a single Hilbert space, oļ¬€ering a diļ¬€erent interpretation up to a natural isomorphism. Theorem 3.36. Let SĪ³ , Ī³ āˆˆ Ī“, be mutually orthogonal closed subspaces of the Hilbert space H, and let PĪ³ denote orthogonal projection to SĪ³ . Then the map  3 SĪ³ ā†’ SĪ³ (3.18) U : span Ī³āˆˆĪ“

Ī³āˆˆĪ“

deļ¬ned by U w = (PĪ³ w)Ī³āˆˆĪ“ is unitary, with inverse given by w =



Ī³āˆˆĪ“ PĪ³ w.

3.4. Orthonormal sets and orthonormal bases

91

 Proof. Denote by S the closure of the span of Ī³āˆˆĪ“ SĪ³ . By Theorems 3.30 and 3.31, for any w āˆˆ S,  PĪ³ w2 = w2 , Ī³āˆˆĪ“

so U is well-deļ¬ned and norm-preserving. If w āˆˆ SĪ² for some Ī² āˆˆ Ī“, then

,w Ī³ = Ī² (U w)Ī³ = 0 Ī³ = Ī², so Ran U contains all vectors with only one nonzero entry. By linearity, it contains all vectors with ļ¬nitely many nonzero entries. Since those are dense and Ran U is closed, it follows that U is surjective.  Due to the natural unitary map between them, the two spaces in (3.18)  are often conļ¬‚ated, and span Ī³āˆˆĪ“ SĪ³ is often called the direct sum of the mutually orthogonal closed subspaces SĪ³ . For instance, in this language, the projection theorem can be concisely restated as for any closed subspace S of H, S āŠ• S āŠ„ = H.

3.4. Orthonormal sets and orthonormal bases In this section, we develop the notion of an orthonormal basis of a Hilbert space H, which allows a useful representation of arbitrary vectors and a classiļ¬cation of Hilbert spaces up to unitary equivalence. Let us begin by comparing this with the situation from linear algebra. As discussed in Chapter 2, we recall that in any vector space V , the span of a subset X āŠ‚ V consists of vectors of the form n  c j xj , v= j=1

where n āˆˆ N, x1 , . . . , xn āˆˆ X and c1 , . . . , cn āˆˆ C. We emphasize that n must be ļ¬niteā€”in a general vector space, there is no notion of convergence and therefore no notion of series. Likewise, linear independence of a set X is deļ¬ned as linear independence of every ļ¬nite subset of X. A Hamel basis of V is deļ¬ned as a linearly independent set of vectors X such that span X = V . While it can be proved using Zornā€™s lemma that every vector space has a Hamel basis, for almost all purposes in analysis, that is not the useful object to consider. Instead, in a Hilbert space H, it is useful to consider sets X such that span X is dense, and to allow vectors to be represented as inļ¬nite linear combinations of basis vectors, addressing issues of convergence as they arise. In another departure from general linear algebra, we will only consider orthonormal bases.

92

3. Hilbert spaces

Deļ¬nition 3.37. Let H be a Hilbert space. (a) A set of vectors X āŠ‚ H is an orthonormal family if x = 1 for all x āˆˆ X and x, x  = 0 for all x, x āˆˆ X with x = x . (b) A set of vectors X āŠ‚ H is an orthonormal basis of H if it is an orthonormal family and span X is dense in H. A countable orthonormal family is often enumerated and written as a sequence, and we will alternate between these points of view. Example 3.38. Deļ¬ne vectors eĪ³ āˆˆ 2 (Ī“) by

1 Ī²=Ī³ (eĪ³ )Ī² = 0 Ī² = Ī³.

(3.19)

Then (eĪ³ )Ī³āˆˆĪ“ is an orthonormal basis of 2 (Ī“). Proof. These vectors form an orthonormal family. Their span is the set of vectors in 2 (Ī“) with ļ¬nitely many nonzero entries, which is dense in 2 (Ī“) by Theorem 3.35.  dx Example 3.39. {eikx | k āˆˆ Z} is an orthonormal basis for L2 ([0, 2Ļ€], 2Ļ€ ).

Proof. This is an orthonormal family because

 2Ļ€ 1 kāˆ’l =0 ilx ikx i(kāˆ’l)x dx = e e , e  = 2Ļ€ 0 k āˆ’ l āˆˆ Z \ {0}. 0 Since a Lebesgue measure gives zero weight to boundary points, the map dx dx x ā†’ eix induces an isomorphism from the space L2 ([0, 2Ļ€], 2Ļ€ ) to L2 (āˆ‚D, 2Ļ€ ), dx now denoting normalized Lebesgue measure on the unit circle āˆ‚D. with 2Ļ€ dx This space has a dense subspace C(āˆ‚D), so for any f āˆˆ L2 (āˆ‚D, 2Ļ€ ) and > 0, there exists h āˆˆ C(āˆ‚D) such that f āˆ’ h < /2. By Weierstrassā€™s second theorem (Corollary 2.21), for any h āˆˆ C(āˆ‚D), there exists a trigonometric polynomial Q such that h āˆ’ Qāˆž < /2. Thus, f āˆ’ Q2 ā‰¤ f āˆ’ h2 + h āˆ’ Q2 ā‰¤ f āˆ’ h2 + h āˆ’ Qāˆž < . dx ). Thus, span{eikx | k āˆˆ Z} is dense in L2 ([0, 2Ļ€], 2Ļ€



If a Hilbert space H has an orthonormal basis (eĪ³ )Ī³āˆˆĪ“ , applying Theorem 3.36 with one-dimensional subspaces SĪ³ = span{eĪ³ } describes the structure of the Hilbert space by a unitary correspondence with 2 (Ī“): Theorem 3.40. Let (eĪ³ )Ī³āˆˆĪ“ be an orthonormal basis for the Hilbert space H. The map U : H ā†’ 2 (Ī“) deļ¬ned by U w = (eĪ³ , w)Ī³āˆˆĪ“ ,

w āˆˆ H,

3.4. Orthonormal sets and orthonormal bases

93

is unitary and the inverse map is given by  U āˆ’1 Īŗ = ĪŗĪ³ e Ī³ , Īŗ āˆˆ 2 (Ī“).

(3.20)

Ī³āˆˆĪ“

Proof. Since projection to SĪ³ = span{eĪ³ } is PĪ³ w = eĪ³ , weĪ³ and PĪ³ w = |eĪ³ , w|, Theorem 3.36 implies that for every w āˆˆ H,  |eĪ³ , w|2 , (3.21) w2 = Ī³āˆˆĪ“

and that for any Īŗ āˆˆ 2 (Ī“), the vector w =



Ī³āˆˆĪ“ ĪŗĪ³ eĪ³

obeys U w = Īŗ.



The norm-preserving property (3.21) is called Parsevalā€™s equality. Special cases of Theorem 3.40 give unitary representations of interest, such as the Fourier series expansion: dx )ā†’ Example 3.41 (Fourier series expansion). The map F : L2 ([0, 2Ļ€], 2Ļ€ 2  (Z), deļ¬ned by  2Ļ€ dx eāˆ’inx f (x) , (F f )n = 2Ļ€ 0 is unitary, and its inverse is given by  uk eikx , (F āˆ’1 u)(x) = kāˆˆZ dx ). with the series understood as a limit in L2 ([0, 2Ļ€], 2Ļ€

Hilbert spaces H, K are said to be unitarily equivalent if there exists a unitary map U : H ā†’ K. Theorem 3.40 provides such unitary equivalences, which describe the structure of a Hilbert space. For instance, since spaces 2 ({1, . . . , n}) = Cn and 2 (N) are separable, it follows that any Hilbert space with a countable orthonormal basis is separable. Conversely, we will prove that every separable Hilbert space has a countable orthonormal basis; then Theorem 3.40 will lead to a classiļ¬cation of separable Hilbert spaces. Parts of the proof are constructive. We will need a formula for orthogonal projection to certain ļ¬nite-dimensional subspaces: Corollary 3.42. Let y1 , . . . , yn be an orthonormal sequence in H. Then the subspace S = span{y1 , . . . , yn } is closed and the orthogonal projection to S is given by n  yj , xyj . (3.22) Px = j=1

Moreover, for all x āˆˆ H, n  j=1

|yj , x|2 ā‰¤ x2 ,

(3.23)

94

3. Hilbert spaces

with equality if and only if x āˆˆ S. This is also known as Besselā€™s inequality. Proof. This is a special case of Theorem 3.30 with Sj = span{yj }.



This motivates a process for obtaining orthonormal sequences with a given span, known as the Gramā€“Schmidt process. This procedure can be expressed in several superļ¬cially diļ¬€erent ways, depending on when one deals with linear dependence and when one treats normalization. Let us address linear dependence in a preliminary step: Lemma 3.43. Any ļ¬nite or inļ¬nite sequence of vectors has a linearly independent subsequence with the same span. Proof. Starting from the sequence (xn )N n=1 , we include in the subsequence / span{xj | j ā‰¤ n āˆ’ 1}. By induction in all the elements xn such that xn āˆˆ  m, for every ļ¬nite m ā‰¤ N , span{xn | n ā‰¤ m} = span{xnk | nk ā‰¤ m}. We now describe the Gramā€“Schmidt process, formulating it as an existence and uniqueness result with an explicit solution: Proposition 3.44 (Gramā€“Schmidt process). Let (xn )N n=1 be a linearly independent sequence in H, with N ļ¬nite or āˆž, and denote V0 = {0} and Vn = span{xj | 1 ā‰¤ j ā‰¤ n} for n ā‰„ 1. Then there is a unique orthonormal sequence (yn )N n=1 such that for all n, span{yj | 1 ā‰¤ j ā‰¤ n} = Vn ,

(3.24)

and for some scalars cn > 0, xn āˆ’ cn yn āˆˆ Vnāˆ’1 . The sequence is given explicitly by a recursive formula xn āˆ’ nāˆ’1 j=1 yj , xn yj . yn = nāˆ’1 xn āˆ’ j=1 yj , xn yj 

(3.25)

(3.26)

Proof. We prove uniqueness by induction in n. The basis of induction n = 0 is trivial; in the inductive step, we assume that y1 , . . . , ynāˆ’1 are orthonormal and that span{y1 , . . . , ynāˆ’1 } = Vnāˆ’1 . The orthogonality conditions on yn āŠ„ . Together with imply that yn āŠ„ yj for all j < n, so by linearity, yn āˆˆ Vnāˆ’1 xn āˆ’ cn yn āˆˆ Vnāˆ’1 , this implies that xn āˆ’ cn yn is the orthogonal projection of xn to Vnāˆ’1 . By Corollary 3.42, this implies xn āˆ’ cn yn =

nāˆ’1 

yj , xn yj .

j=1

3.4. Orthonormal sets and orthonormal bases

95

Since xn āˆˆ / Vnāˆ’1 , cn yn = 0. Since yn is normalized and cn > 0, this implies that % % nāˆ’1 % %  % % c n = %x n āˆ’ yj , xn yj %, % % j=1

and (3.26) is the unique solution. From xn āˆ’ cn yn āˆˆ Vnāˆ’1 and the inductive  assumption, it follows that span{y1 , . . . , yn } = Vn . Corollary 3.45. Every separable Hilbert space has a countable orthonormal basis. In particular, every separable Hilbert space is unitarily equivalent to Cn for some n āˆˆ N or to 2 (N). Proof. If H is separable, it has a countable dense set. By Lemma 3.43, H has a linearly independent sequence with a dense span. Applying the Gramā€“Schmidt process gives an orthonormal basis (yj )N j=1 . It follows that 4N  there exists a unitary map from H to j=1 C. This is very close to a classiļ¬cation result for separable Hilbert spaces. It remains to prove that diļ¬€erent orthonormal bases have the same cardinality. We need two results: Theorem 3.46. If a Hilbert space H has a ļ¬nite orthonormal basis consisting of n vectors, then any orthonormal family in H has at most n vectors. Proof. Assume that (ek )nk=1 is an orthonormal basis of H and that X is an orthonormal family in H. Using Parsevalā€™s equality with respect to the orthonormal basis (ek )nk=1 and Besselā€™s inequality with respect to the orthonormal family X gives  xāˆˆX

1=

 xāˆˆX

n n   2 x = |ek , x| ā‰¤ ek 2 = n, 2

xāˆˆX k=1

which shows that X has at most n elements.

k=1



Theorem 3.47. In a separable Hilbert space, every orthonormal set is countable. Proof. Let X be an orthonormal set in H. By Parsevalā€™s equality with respect to a countable orthonormal basis (eĪ³ )Ī³āˆˆĪ“ , for each x āˆˆ X there

96

3. Hilbert spaces

exists Ī³ such that x, eĪ³  = 0. However, for each Ī³ āˆˆ Ī“, the set {x āˆˆ X | x, eĪ³  = 0} is countable by Theorem 3.31 applied to orthogonal projections to x āˆˆ X. Taking the union over Ī“ shows that the set X is countable.  Thus, in a separable Hilbert space H, every orthonormal basis has the same cardinality, which is called the dimension of H and is denoted dim H. In the ļ¬nite-dimensional case, this Hilbert space dimension matches the notion of dimension from linear algebra; however, in the inļ¬nite-dimensional case, the cardinalities are not the same (Exercise 3.15). An important special case is obtained by starting with the sequence of monomials with respect to a suitable measure. In the theory of orthogonal polynomials, measures such that supp Ī¼ is inļ¬nite are called nontrivial. This facilitates the following construction: Example 3.48. Let Ī¼ be a measure on C with all ļ¬nite moments, i.e.,  āˆ€n āˆˆ N āˆŖ {0}. |z|n dĪ¼(z) < āˆž Then z n āˆˆ L2 (C, dĪ¼) for all n āˆˆ N āˆŖ {0} and the following hold: (a) Monomials 1, z, . . . , z n are linearly independent in L2 (C, dĪ¼) if and only if supp Ī¼ consists of more than n points. (b) If Ī¼ is nontrivial, there is a unique sequence of polynomials (pn (z))āˆž n=0 such that each pn is of degree n, pn have positive leading coeļ¬ƒcients, and  (3.27) pm , pn  = pm (x)pn (x) dĪ¼(x) = Ī“m,n . (c) If Ī¼ is nontrivial and supp Ī¼ is a compact subset of R, then 2 (pn (z))āˆž n=0 is an orthonormal basis in L (R, dĪ¼). Proof. (a) Monomials 1, z, . . . , z n are linearly dependent if and only if there exists a nontrivial polynomial Q with deg Q ā‰¤ n such that Q2 = 0, i.e., Q = 0 Ī¼-a.e. Such a polynomial exists if and only if supp Ī¼ consists of at most n points. (b) The sequence (pn )āˆž n=0 is the sequence obtained by the Gramā€“Schmidt 2 process from the linearly independent sequence (z n )āˆž n=0 in L (C, dĪ¼). (c) By Weierstrassā€™s theorem, if supp Ī¼ is a compact subset of R, polynomials are dense in C(supp Ī¼), which is itself a dense subset of L2 (R, dĪ¼).  The polynomials pn are called the orthonormal polynomials for the measure Ī¼. Orthogonal polynomials for measures supported on R are closely related to Jacobi matrices, and we will revisit them in Chapter 10; see also [25, 92, 98, 107].

3.5. Weak convergence

97

For measures supported on the unit circle āˆ‚D = {z āˆˆ C | |z| = 1}, orthonormal polynomials do not usually give an orthonormal basis, but are closely related to a basis of trigonometric polynomials (Exercise 3.16). For a systematic study of orthogonal polynomials on the unit circle see [88, 89].

3.5. Weak convergence The Riesz representation theorem motivates the following deļ¬nition. Deļ¬nition 3.49. A sequence (xn )āˆž n=1 in H converges weakly to x āˆˆ H if xn , v ā†’ x, v

āˆ€v āˆˆ H. w

This is denoted w-limnā†’āˆž xn = x or xn ā†’ x. Of course, this is the special case of weak-āˆ— convergence (Section 2.5) in the setting of a Hilbert space H, written as a statement about vectors in H instead of about functionals in Hāˆ— . The next few basic properties are mostly specializations of general properties of weak-āˆ— convergence: w

w

Lemma 3.50. If xn ā†’ x and xn ā†’ y, then x = y. Proof. For all v āˆˆ H, x, v = limnā†’āˆž xn , v = y, v. Thus, x āˆ’ y, Ā· is the trivial functional on H, so x = y.  It is common to refer to convergence with respect to the Hilbert space norm as strong convergence, to distinguish it from the newly deļ¬ned weak convergence. The two are related: w

Lemma 3.51. xn ā†’ x implies xn ā†’ x. Proof. This follows from |xn , v āˆ’ x, v| = |xn āˆ’ x, v| ā‰¤ xn āˆ’ xv.  w

Lemma 3.52. If H is ļ¬nite dimensional, xn ā†’ x implies xn ā†’ x. Proof. If H has a ļ¬nite orthonormal basis {ej }N j=1 with N < āˆž, then w xn ā†’ x implies ej , xn  ā†’ ej , x for all j, so N N   ej , xn ej ā†’ ej , xej = x. xn = j=1



j=1

Accordingly, we will focus on the inļ¬nite-dimensional case from now on. In that case, weak convergence does not imply strong convergence: Example 3.53. Any orthonormal sequence converges weakly to 0, but does not converge strongly.

98

3. Hilbert spaces

Proof. Let (xn )āˆž Pythagorean then=1 be an orthonormal āˆš sequence. By the āˆž orem, for all n = m, xm āˆ’ xn  = 2. Therefore, (xn )n=1 is not a Cauchy sequence, so it is not convergent. For any v āˆˆ H, by Besselā€™s inequality, āˆž 

|xn , v|2 ā‰¤ v2 < āˆž,

n=1 w

so xn , v ā†’ 0 as n ā†’ āˆž. Thus, xn ā†’ 0.



w

This example also shows that xn ā†’ x does not imply convergence of xn  to x. The connections between weak convergence and boundedness are described in the following proposition. Proposition 3.54. w

(a) If xn ā†’ x, then supnāˆˆN xn  < āˆž. w

(b) If xn ā†’ x, then x ā‰¤ lim inf nā†’āˆž xn . w

(c) If xn ā†’ x and x ā‰„ lim supnā†’āˆž xn , then xn ā†’ x. Proof. (a) The functionals Ī›n = xn , Ā· āˆˆ Hāˆ— converge pointwise, so they are pointwise bounded. Thus, by the uniform boundedness principle, they are uniformly bounded. Since Ī›n  = xn , the sequence xn is bounded. (b) This follows from x2 = lim |xn , x| ā‰¤ lim inf xn x. nā†’āˆž

nā†’āˆž

(c) Since xn , x ā†’ x, x = x2 , starting from xn āˆ’ x2 = xn 2 + x2 āˆ’ 2 Rexn , x, we conclude lim supxn āˆ’ x2 = lim supxn 2 āˆ’ x2 ā‰¤ 0. nā†’āˆž

nā†’āˆž

Thus, xn ā†’ x.



Weak convergence does not imply convergence of norms, so it does not w w imply convergence of inner products: to ļ¬nd xn ā†’ x and yn ā†’ y such that xn , yn  does not converge to x, y, it suļ¬ƒces to take yn = xn to be an orthonormal sequence in H. The next lemma is therefore in some sense optimal. w

Lemma 3.55. If xn ā†’ x and yn ā†’ y, then xn , yn  ā†’ x, y. Proof. Since the sequence (xn )āˆž n=1 is bounded, it follows from |xn , yn āˆ’ y| ā‰¤ xn yn āˆ’ y

and yn ā†’ y

3.5. Weak convergence

99

that xn , yn āˆ’ y ā†’ 0. Weak convergence implies xn āˆ’ x, y ā†’ 0, so xn , yn  āˆ’ x, y = xn āˆ’ x, y + xn , yn āˆ’ y ā†’ 0.



Applying Lemma 2.46 to a sequence in Hāˆ— and using the Riesz representation theorem provides the following criterion for weak convergence. Lemma 3.56. If (xn )āˆž n=1 is a bounded sequence in H and there is a dense set D āŠ‚ H such that for all y āˆˆ D, limnā†’āˆž xn , y is convergent, then the sequence (xn )āˆž n=1 is weakly convergent. Applying the Banachā€“Alaoglu theorem to Hilbert spaces gives the following result, often stated as weak compactness of a closed ball in H. Theorem 3.57. In a separable Hilbert space, every bounded sequence has a weakly convergent subsequence. Similarly, a small modiļ¬cation of Theorem 2.56 and its proof show: Theorem 3.58. Let H be a separable Hilbert space, and let (ek )āˆž k=1 be an orthonormal basis of H. Then āˆž  d(x, y) = min(2āˆ’k , |ek , x āˆ’ y|) (3.28) k=1

deļ¬nes a metric on H. Moreover, let (xn )āˆž n=1 be a sequence in H, and let w x āˆˆ H. Then xn ā†’ x if and only if supxn  < āˆž

and

nāˆˆN

lim d(xn , x) = 0.

nā†’āˆž

Proof. (a) By Lemma 2.55, d is a semimetric. If d(x, y) = 0, then ek , x āˆ’ y = 0 for all k, which implies x = y because (ek )āˆž k=1 is an orthonormal basis. Thus, d is a metric. w

(b) If xn ā†’ x, then xn is a bounded sequence in H. Moreover, ek , xn āˆ’ x ā†’ 0 for every k, so by dominated convergence with dominating sequence 2āˆ’k applied to the counting measure on N, āˆž  lim min{2āˆ’k , |ek , xn āˆ’ x|} = 0. lim d(xn , x) = nā†’āˆž

k=1

nā†’āˆž

Conversely, if d(xn , x) ā†’ 0, then min(2āˆ’k , |ek , xn āˆ’ x|) ā†’ 0 for each k, so ek , xn āˆ’ x ā†’ 0 for each k, and then by linearity, v, xn  ā†’ v, x for all v āˆˆ span{ek }āˆž k=1 . Since that set is dense and the sequence xn is bounded, w  by Lemma 3.56, xn ā†’ x. In other words, on bounded sets, convergence in this metric is equivalent to weak convergence; this is only true on bounded subsets of H, and not on all of H, as seen in the following example.

100

3. Hilbert spaces

w

Example 3.59. If we take xn = nen , then d(xn , 0) = 2āˆ’n ā†’ 0, but xn ā†’  0 āˆž 1 because for the vector v = j=1 j ej āˆˆ H, we have v, xn  = 1 ā†’ 0. The restriction to bounded sets does not mean that this choice of metric is wrong; rather, it turns out there is no metric that would work. If dim H = āˆž, weak convergence on H is not metrizable, i.e., there is no metric d on H w such that d(xn , x) ā†’ 0 if and only if xn ā†’ x. This result will not be proved or needed in this text.

3.6. Tensor products of Hilbert spaces The tensor product of Hilbert spaces H, K is a new Hilbert space H āŠ— K obtained by a multiplicative construction: vectors are obtained from formal products of vectors and the inner product is a product of inner products. We describe that construction in this section, as well as a universal property which determines the tensor product uniquely and which can be more transparent than the construction itself. We begin with a motivating example. Example 3.60. Consider the Hilbert space L2 ([0, 1]2 ) = L2 ([0, 1]2 , dm2 ), where m2 denotes two-dimensional Lebesgue measure on [0, 1]2 . For f, g āˆˆ L2 ([0, 1]) = L2 ([0, 1], dm), we deļ¬ne the function f āŠ— g āˆˆ L2 ([0, 1]2 ) by (f āŠ— g)(x, y) = f (x)g(y).

(3.29)

(a) The map (f, g) ā†’ f āŠ— g is a bilinear map from L2 ([0, 1]) Ɨ L2 ([0, 1]) to L2 ([0, 1]2 ), i.e., it is linear in each parameter. (b) For all f1 , f2 , g1 , g2 āˆˆ L2 ([0, 1]), f1 āŠ— g1 , f2 āŠ— g2  = f1 , f2 g1 , g2 

(3.30)

with inner products taken in the respective Hilbert spaces. (c) span{f āŠ— g | f, g āˆˆ L2 ([0, 1])} is dense in L2 ([0, 1]2 ). (d) span{f āŠ— g | f, g āˆˆ L2 ([0, 1])} = L2 ([0, 1]2 ). Proof. (a) Bilinearity follows directly from (3.29), and f āŠ— g āˆˆ L2 ([0, 1]2 ) from Tonelliā€™s theorem:   1  1 |f (x)g(y)|2 dm2 (x, y) = |f (x)|2 dx |g(y)|2 dy < āˆž. [0,1]2

0

0

(b) By the deļ¬nition of the inner product on L2 ([0, 1]2 ),  f1 (x)g1 (y)f2 (x)g2 (y) dm2 (x, y). f1 āŠ— g1 , f2 āŠ— g2  = [0,1]2

The integrand is in L1 ([0, 1]2 ), so using Fubiniā€™s theorem to separate this as a product of single integrals gives (3.30).

3.6. Tensor products of Hilbert spaces

101

(c) span{e2Ļ€ikx āŠ— e2Ļ€ily | k, l āˆˆ Z} is a subalgebra of C([0, 1]2 ) which separates points and is closed under complex conjugation, so it is dense in C([0, 1]2 ) by the Stoneā€“Weierstrass theorem. Thus, it is dense in L2 ([0, 1]2 ). (d) The set {e2Ļ€ikx āŠ— e2Ļ€ily | k, l āˆˆ Z} is an orthonormal family by (3.30) and has a dense span by (c), so it is an orthonormal basis for L2 ([0, 1]2 ). For f, g āˆˆ L2 ([0, 1]),   |e2Ļ€inx āŠ— e2Ļ€iny , f āŠ— g| = |e2Ļ€inx , f (x)e2Ļ€iny , g(y)| ā‰¤ f g nāˆˆZ

nāˆˆZ

by the Cauchyā€“Schwarz inequality in 2 (Z). Thus, for h = f āŠ— g,  |e2Ļ€inx āŠ— e2Ļ€iny , h| < āˆž.

(3.31)

nāˆˆZ

By linearity, (3.31) then also holds for all h āˆˆ span{f āŠ— g | f, g āˆˆ L2 ([0, 1])}. However, there exist h āˆˆ L2 ([0, 1]2 ) for which (3.31) fails: it suļ¬ƒces to take  an e2Ļ€inx āŠ— e2Ļ€iny h= nāˆˆZ

with a sequence a āˆˆ (so that the sum gives an element of L2 ([0, 1]2 )) 1 and a āˆˆ /  (Z) (so that (3.31) fails). An explicit example is the vector āˆž 1 2Ļ€inx āŠ— e2Ļ€iny .  h = n=1 n e 2 (Z)

A description of a tensor product Hilbert space would not have any substance without the description of the accompanying bilinear map (f, g) ā†’ f āŠ— g. The previous example suggests the following deļ¬nition. Deļ¬nition 3.61. Let H, K, V be Hilbert spaces, and let i : H Ɨ K ā†’ V be a map with the following properties. (a) i is bilinear, i.e., for any Ī» āˆˆ C, x, x āˆˆ H, y, y  āˆˆ K, Ī»i(x, y) = i(Ī»x, y) = i(x, Ī»y), 



i(x + x , y) = i(x, y) + i(x , y), 



i(x, y + y ) = i(x, y) + i(x, y ).

(3.32) (3.33) (3.34)

(b) For all x, x āˆˆ H and y, y  āˆˆ K, i(x, y), i(x , y  ) = x, x y, y  . (c) The image of i has a dense span in V . The space V is called the tensor product of Hilbert spaces H, K and i is called the canonical bilinear map. We will now prove existence of the tensor product (by an explicit construction) and its uniqueness (up to a unitary map). The construction is diļ¬€erent from the algebraic construction of a tensor product of modules: we

102

3. Hilbert spaces

rely on the inner product structure throughout, and in order to make H āŠ— K a Hilbert space, we use an additional Hilbert space completion, which is considered in an abstract setting in Exercise 3.8. Theorem 3.62. Let H, K be Hilbert spaces. (a) There exists a Hilbert space V1 and a bilinear map i1 : H Ɨ K ā†’ V1 which obeys properties (a), (b), and (c) of Deļ¬nition 3.61. (b) If there is a Hilbert space V2 and a bilinear map i2 : H Ɨ K ā†’ V2 with the same properties, then there is a unitary map U : V1 ā†’ V2 such that U ā—¦ i1 = i2 . Proof. (a) We begin by considering the set of formal linear combinations of pairs (x, y) āˆˆ H Ɨ K, 

n  cj (xj , yj ) | n āˆˆ N, cj āˆˆ C, xj āˆˆ H, yj āˆˆ K A= j=1

(more formally, this can be presented as the set of all functions H Ɨ K ā†’ C which are equal to 0 except at ļ¬nitely many points). The set A is a vector space; on it, we deļ¬ne the sesquilinear form 0 / n m n  m      cj (xj , yj ), dk (xk , yk ) = cĀÆj dk xj , xk yj , yk . j=1

j=1 k=1

k=1

We deļ¬ne the set A0 = {v āˆˆ A | v, w = 0 for all w āˆˆ A}. v, w

w

If v āˆ’ āˆ’ āˆˆ A0 , then v, w = v, w  = v  , w , so this sesquilinear form induces a sesquilinear form on the quotient vector space A/A0 . Denote by x āŠ— y the coset of (x, y) in A/A0 . It is directly veriļ¬ed that the map i(x, y) = x āŠ— y is bilinear. For instance, if we compute ĀÆ x y, y   āˆ’ Ī»x, x y, y   = 0, Ī»(x, y) āˆ’ (Ī»x, y), (x , y  ) = Ī»x, taking linear combinations for the second parameter shows that Ī»(x, y) āˆ’ (Ī»x, y), w = 0

āˆ€w āˆˆ A,

and therefore Ī»(x āŠ— y) = (Ī»x) āŠ— y. Similar calculations show that

(3.35)

Ī»(x āŠ— y) = x āŠ— (Ī»y),

(3.36)

x āŠ— y + x āŠ— y = (x + x ) āŠ— y, 

(3.38)



(3.39)

x āŠ— y + x āŠ— y = x āŠ— (y + y ), 





(3.37)



x āŠ— y, x āŠ— y  = x, x y, y .

3.6. Tensor products of Hilbert spaces

103

Our next goal is to prove that this sesquilinear form on A/A0 is positive deļ¬nite. In order to do this, let us ļ¬rst rewrite an arbitrary vector v=

n 

cj xj āŠ— yj

j=1

in A/A0 . By the Gramā€“Schmidt process, there is an orthonormal sequence x1 , x2 , . . . , xm , m ā‰¤ n, with the same span as x1 , . . . , xn . Writing xj as linear combinations of xk and using (3.35) and (3.37), we can write v as a linear combination of vectors of the form xk āŠ— y for some y āˆˆ K. Grouping terms with the same k by using (3.36) and (3.38), we can ļ¬nally write v=

m 

xk āŠ— yk

k=1

for some

y1 , . . . , yk

āˆˆ K. Since

v, v =

xk

are orthonormal,

m  m m   xj , xk yj , yk  = yk 2 ā‰„ 0. j=1 k=1

k=1

yk

= 0 for all k, so v = 0. In conclusion, we Moreover, v, v = 0 implies have proved that the sesquilinear form on A/A0 is positive deļ¬nite. By construction, the span of the range of i is A/A0 . The vector space A/A0 is equipped with an inner product, but is not (in general) complete; denoting by V its Hilbert space completion (Exercise 3.8) completes the proof. (b) Deļ¬ne a map W : A ā†’ V2 by deļ¬ning W : (x, y) ā†’ i2 (x, y) and extending linearly. The map W preserves inner products, so Ker W = A0 . Thus, W induces a norm-preserving map U : A/A0 ā†’ V2 . Since A/A0 is a dense subset of V1 , this extends to a norm-preserving map U : V1 ā†’ V2 . The range of U is V2 because the span of the range of i2 is dense in V2 and  U ā—¦ i1 = i2 by construction. Due to this existence and uniqueness theorem, it is customary to denote the tensor product V by H āŠ— K and the values of the canonical bilinear map i by i(x, y) = x āŠ— y. For example, we can now say that L2 ([0, 1]) āŠ— L2 ([0, 1]) = L2 ([0, 1]2 ) with the canonical bilinear map (3.29). In practice, it is easier to check the deļ¬nition than to trace the explicit construction of the tensor product: Lemma 3.63. If (ej )jāˆˆJ is an orthonormal basis of H and (fk )kāˆˆK is an orthonormal basis of K, then (ej āŠ— fk )jāˆˆJ,kāˆˆK is an orthonormal basis of H āŠ— K. In particular, dim(H āŠ— K) = dim H dim K.

104

3. Hilbert spaces

Proof. The set (ej āŠ— fk )jāˆˆJ,kāˆˆK is an orthonormal set because ej āŠ— fk , ej  āŠ— fk  = ej , ej  fk , fk , and this is equal to 1 if j = j  and k = k  and zero otherwise. It suļ¬ƒces to prove that M = span{ej āŠ— fk | j āˆˆ J, k āˆˆ K} is dense in H āŠ— K. For any k āˆˆ K, bilinearity of the tensor product implies that span{ej āŠ— fk | j āˆˆ J} = {x āŠ— fk | x āˆˆ span{ej | j āˆˆ J}}, and since x āŠ— fk  = xfk , this set is dense in {x āŠ— fk | x āˆˆ H}. Thus, x āŠ— fk āˆˆ M for all x āˆˆ H and k āˆˆ K. Repeating this argument for K, we conclude x āŠ— y āˆˆ M for all x āˆˆ H and y āˆˆ K. Since M is a closed subspace,  this implies that M = H āŠ— K. In particular, it is common to say that Cm āŠ— Cn āˆ¼ = Cmn , denoting the m standard basis of C by e1 , . . . , em , the standard basis of Cn by f1 , . . . , fn , and viewing Cmn as the space of m Ɨ n matrices so that ej āŠ— fk is the matrix with a 1 in the jk entry and zeros in all other entries.

3.7. Exercises 3.1. Let S āŠ‚ [0, 2Ļ€) be a subset of positive Lebesgue measure. Prove that there exists C > 0 such that  |a + beiĪø |2 dĪø ā‰„ C(|a|2 + |b|2 ) āˆ€a, b āˆˆ C, S

and ļ¬nd the optimal constant C as a function of S. 3.2. Prove that |x, y| = xy if and only if x, y are linearly dependent. 3.3. The Gram matrix of vectors x1 , . . . , xn āˆˆ H is the n Ɨ n matrix B with entries bjk = xj , xk . Prove the following. (a) B is always positive semideļ¬nite, i.e., Ī»āˆ— BĪ» ā‰„ 0 for all Ī» āˆˆ Cn . (b) B is positive deļ¬nite, i.e., Ī»āˆ— BĪ» > 0 for all Ī» āˆˆ Cn \ {0}, if and only if vectors x1 , . . . , xn are linearly independent. 3.4. Prove that x + y = x + y if and only if x = 0 or y = Ī»x for some Ī» ā‰„ 0. in H obeying (3.6), for 3.5. Given a pairwise orthogonal sequence (xj )āˆž āˆž j=1 any bijection Ļƒ : N ā†’ N, prove that j=1 xĻƒ(j) = āˆž j=1 xj . 3.6. If X is Ļƒ-locally compact and Ī¼ is a Baire measure on X whose support contains at least two points, prove that for p āˆˆ [1, āˆž] \ {2}, the p-norm on Lp (X, Ī¼) is not the induced norm of an inner product. 3.7. Let V be a vector space, and let Ā·, Ā·,  : V Ɨ V ā†’ C be a map which is linear in the second parameter, conjugate-symmetric, and for all x āˆˆ V , x, x ā‰„ 0. . (a) Prove that x = x, x deļ¬nes a seminorm on V .

3.7. Exercises

105

(b) Let V0 = {x āˆˆ V | x = 0}. For any x āˆˆ V0 and y āˆˆ V , prove that x, y = 0. (c) Prove that Ā·, Ā·,  induces an inner product on V /V0 . Hint: Use Lemma 2.8. 3.8. Let V be a pre-Hilbert space, and let B be its Banach space completion (Exercise 2.10). Explicitly, we assume that B is a Banach space, i : V ā†’ B is a norm-preserving linear map, and Ran i is dense in B. (a) Prove that the map Ā·, Ā· : Ran i Ɨ Ran i ā†’ C deļ¬ned by i(x), i(y) = x, y can be extended uniquely to a continuous map B Ɨ B ā†’ C, which we also denote by Ā·, Ā·. (b) Prove that this extension is an inner product on B and that x, x = x2 for all x āˆˆ B, so the norm induced by Ā·, Ā· is the norm of B. (c) Conclude that B is a Hilbert space with inner product Ā·, Ā·. āŠ„

3.9. (a) If S is a subspace of H, prove that S āŠ„ = S . (b) If S is a closed subspace of H, prove that (S āŠ„ )āŠ„ = S. (c) If S is an arbitrary subspace of H, prove that (S āŠ„ )āŠ„ = S. 3.10. If P, Q are orthogonal projections on H, prove that Ran P āŠ‚ Ran Q if and only if QP = Q. 3.11. A subspace S of H is ļ¬nite-dimensional if S = span{v1 , . . . , vn } for some ļ¬nite set v1 , . . . , vn . Prove that any ļ¬nite-dimensional subspace is closed. āˆš 2 3.12. Prove that ( 2 sin(nĻ€x))āˆž n=1 is an orthonormal basis for L ([0, 1]). 3.13. For any n āˆˆ N, prove that the sequence (e2Ļ€ikĀ·x )kāˆˆZn is an orthonormal basis for L2 ([0, 1]n , dmn ) where mn denotes Lebesgue measure on [0, 1]n . 3.14. Let (eĪ³ )Ī³āˆˆĪ“ be an orthonormal basis of a Hilbert space H. Prove that Ī³āˆˆĪ“ ĪŗĪ³ eĪ³ can be interpreted as a Bochner integral (Deļ¬nition 2.59) if and only if Īŗ āˆˆ 1 (Ī“), but it can be interpreted as a Pettis integral (Remark 2.63) if and only if Īŗ āˆˆ 2 (Ī“).

106

3. Hilbert spaces

3.15. Let H be a separable, inļ¬nite-dimensional Hilbert space. Prove that any set X āŠ‚ H such that span X = H is uncountable. Hint: If X was countable, prove there would exist an orthonormal sequence (yj )āˆž j=1 with span{yj | j āˆˆ N} = H and consider āˆž āˆ’1 v = j=1 j yj . 3.16. Let Ī¼ be a probability measure on āˆ‚D = {z āˆˆ C | |z| = 1}. Assume that Ī¼ is nontrivial, i.e., supp Ī¼ is inļ¬nite. (a) Prove that applying the Gramā€“Schmidt process to the sequence 1, z, z āˆ’1 , z 2 , z āˆ’2 , . . . 2 gives an orthonormal basis {Ļ‡n }āˆž n=0 for L (āˆ‚D, dĪ¼). This is called the CMV basis following Canteroā€“Moralā€“VelĀ“ azquez [16]. āˆž (b) Denote by (Ļ•n )n=0 the result of applying Gramā€“Schmidt process to (z n )āˆž n=0 . Prove that for all n āˆˆ N āˆŖ {0},

Ļ‡2n (z) = z n Ļ•2n (1/z),

Ļ‡2n+1 (z) = z āˆ’n Ļ•2n+1 (z).

3.17. For any Hilbert space H and n āˆˆ N, construct4a canonical bilinear map that justiļ¬es the identiļ¬cation H āŠ— Cn = nj=1 H. 3.18. Consider Hilbert spaces H, K and their tensor product H āŠ— K. Prove that span{x āŠ— y | x āˆˆ H, y āˆˆ K} = H āŠ— K if and only if H and K are inļ¬nite dimensional. 3.19. If H1 , H2 , H3 are Hilbert spaces, prove that there exists a unitary map U : (H1 āŠ— H2 ) āŠ— H3 ā†’ H1 āŠ— (H2 āŠ— H3 ) such that U ((x1 āŠ—x2 )āŠ—x3 ) = x1 āŠ—(x2 āŠ—x3 ) for all xj āˆˆ Hj , j = 1, 2, 3.

Chapter 4

Bounded linear operators

In this chapter, we study bounded linear operators on a Hilbert space H. It is assumed throughout that H is separable. Composition of operators on L(H) is viewed as a multiplicative operation, with identity operator I deļ¬ned by Ix = x for all x āˆˆ H; together with the linear structure, this makes L(H) an algebra. The Hilbert space structure on H induces an additional unary operation, which associates to every operator A its adjoint operator Aāˆ— such that for all u, v āˆˆ H, (4.1) u, Av = Aāˆ— u, v. That this is well deļ¬ned and determines a unique operator Aāˆ— āˆˆ L(H) will be proved promptly in Proposition 4.2. We will then consider the resulting structure and properties of L(H). Some of the material of this chapter is preparation for a detailed study of self-adjoint operators, i.e., those with A = Aāˆ— , which will follow in later chapters.

4.1. The C āˆ— -algebra of bounded linear operators on H For operators on a Hilbert space, the norm can be characterized in terms of the inner product: Lemma 4.1. For any linear map A : H ā†’ H, A =

sup

|u, Av|.

u,vāˆˆH u=v=1

In particular, A is a bounded linear operator if and only if this supremum is ļ¬nite. 107

108

4. Bounded linear operators

Proof. By Lemma 3.24, for any v āˆˆ H, Av = sup |u, Av|. uāˆˆH u=1

Taking the supremum over all normalized v āˆˆ H completes the proof.



Proposition 4.2. Let A āˆˆ L(H). For any u āˆˆ H, there is a unique vector Aāˆ— u āˆˆ H such that (4.1) holds for all v āˆˆ H. The map u ā†’ Aāˆ— u is linear and Aāˆ—  = A. In particular, Aāˆ— is a bounded linear operator on H. Proof. For ļ¬xed u āˆˆ H, consider the linear map Ī›u : H ā†’ C deļ¬ned by Ī›u v = u, Av. Since for all v āˆˆ H, |Ī›u v| = |u, Av| ā‰¤ uAv ā‰¤ uAv, Ī›u is a bounded linear functional on H. By the Riesz representation theorem, it corresponds to a unique vector Aāˆ— u. Linearity of Aāˆ— follows from uniqueness; namely, for any Ī» āˆˆ C and u, v āˆˆ H, Aāˆ— (Ī»u), v = Ī»u, Av = Ī»u, Av = Ī»Aāˆ— u, v = Ī»Aāˆ— u, v, so Aāˆ— (Ī»u) = Ī»Aāˆ— u, and similarly, for any u, u , v āˆˆ H, Aāˆ— (u + u ), v = u + u , Av = Aāˆ— u, v + Aāˆ— u , v = Aāˆ— u + Aāˆ— u , v, so Aāˆ— (u+u ) = Aāˆ— u+Aāˆ— u . Thus, Aāˆ— is a linear operator on H. Boundedness of Aāˆ— and Aāˆ—  = A follow from Lemma 4.1.  Deļ¬nition 4.3. The adjoint of A āˆˆ L(H) is the unique operator Aāˆ— āˆˆ L(H) such that (4.1) holds for all u, v āˆˆ H. Example 4.4. Let A be a complex n Ɨ n matrix, viewed as an element of L(Cn ). Its adjoint Aāˆ— is the matrix with entries (Aāˆ— )ij = Aji . Proof. Since Aāˆ— āˆˆ L(Cn ) is uniquely determined by (4.1), it suļ¬ƒces to compute its matrix elements, which follow from (4.1) by (Aāˆ— )ij = Ī“i , Aāˆ— Ī“j  = Aāˆ— Ī“j , Ī“i  = Ī“j , AĪ“i  = Aji .



Lemma 4.5. Let U āˆˆ L(H). Then: (a) U is norm-preserving if and only if U āˆ— U = I. (b) U is unitary if and only if U āˆ— U = U U āˆ— = I. Proof. (a) By deļ¬nition, U is norm-preserving if and only if U v = v for all v āˆˆ H. By the polarization identity, this is equivalent to U v, U w = v, w for all v, w āˆˆ H and therefore equivalent to v, U āˆ— U w = v, w for all v, w āˆˆ H. By the Riesz representation theorem this is equivalent to U āˆ— U w = w for all w and then to U āˆ— U = I.

4.1. The C āˆ— -algebra of bounded linear operators on H

109

(b) If U āˆ— U = U U āˆ— = I, then U is norm-preserving by (a); moreover, U U āˆ— = I implies Ran U = H, so U is unitary. Conversely, let U be unitary. Then it is norm-preserving, so U āˆ— U = I. Moreover, any v āˆˆ H can be written in the form v = U w so U U āˆ— v = U U āˆ— U w = U w = v. Thus, U U āˆ— = I.  Example 4.6. Let S denote the shift operator on 2 (N), deļ¬ned by (Su)n = un+1 ,

u āˆˆ 2 (N).

(4.2)

n=1 n ā‰„ 2.

(4.3)

Its adjoint S āˆ— is the operator

0 (S āˆ— u)n = unāˆ’1

Note that SS āˆ— = I but S āˆ— S = I. Proof. Note that S is a bounded linear operator with S = 1, because Su = 2

āˆž 

āˆž āˆž   2 |un+1 | = |uk | ā‰¤ |uk |2 = u2

n=1

2

k=2

k=1

and equality holds for any vector with u1 = 0. From (4.2), we compute SĪ“k = Ī“kāˆ’1 if k ā‰„ 2 and SĪ“1 = 0. Thus, for u āˆˆ 2 (N), (S āˆ— u)n = Ī“n , S āˆ— u = SĪ“n , u and splitting cases gives (4.3).  Direct calculations give SS āˆ— = I and S āˆ— Sx = x āˆ’ x1 Ī“1 . Lemma 4.7. For any A āˆˆ L(H), (Ran A)āŠ„ = Ker Aāˆ— . Proof. u, Av = 0 for all v āˆˆ H is equivalent to Aāˆ— u, v = 0 for all v āˆˆ H,  so it is equivalent to Aāˆ— u = 0. Recall that any bounded linear operator is continuous, so it maps convergent sequences to convergent sequences. It also maps weakly convergent sequences to weakly convergent sequences: w

w

Lemma 4.8. Let A āˆˆ L(H). If xn ā†’ x, then Axn ā†’ Ax. Proof. For any y āˆˆ H, y, Axn  = Aāˆ— y, xn  ā†’ Aāˆ— y, x = y, Ax.



On L(H), as already noted, we interpret the composition of operators as a multiplicative operation. Viewing the adjoint as a unary operation leads to the following structure. Deļ¬nition 4.9. Let X be a Banach space equipped with a binary operation denoted multiplicatively and a unary operation āˆ— . X is a C āˆ— algebra if it is

110

4. Bounded linear operators

a Banach space, a ring, and for all a, b āˆˆ X and z āˆˆ C, the following hold. (a) ab ā‰¤ ab. (b) (a + b)āˆ— = aāˆ— + bāˆ— . (c) (za)āˆ— = zĀÆaāˆ— . (d) (ab)āˆ— = bāˆ— aāˆ— . (e) (aāˆ— )āˆ— = a. (f) If a is invertible, then so is aāˆ— and (aāˆ— )āˆ’1 = (aāˆ’1 )āˆ— . (g) aāˆ— a = a2 . If X has an identity element for multiplication, it is a C āˆ— algebra with identity. If multiplication in X is commutative, X is a commutative C āˆ— algebra. Theorem 4.10. L(H) is a C āˆ— -algebra with identity. Proof. L(H) is a Banach space by Proposition 2.38. The algebraic properties are obvious, with the multiplicative identity I. The property Aāˆ— A = A2 is proved by proving two inequalities. By Lemma 4.1, Aāˆ— A ā‰„ sup |u, Aāˆ— Au| = sup |Au, Au| = sup Au2 = A2 . uāˆˆH u=1

uāˆˆH u=1

uāˆˆH u=1

Conversely, Aāˆ— A ā‰¤ Aāˆ— A = A2 , so Aāˆ— A = A2 .



L(H) is the canonical example of a C āˆ— -algebra and is the reason why we introduce them here, but we should point out a few additional examples. In all of the following examples, multiplication is pointwise multiplication of functions and the unary operation is complex conjugation, f āˆ— (x) = f (x). Example 4.11. If K is compact, C(K) is a commutative C āˆ— -algebra with identity. Example 4.12. The set Bb (X) of bounded Borel functions on X, with the norm f  = supxāˆˆX |f (x)|, is a commutative C āˆ— -algebra with identity. Example 4.13. Lāˆž (X, dĪ¼) is a commutative C āˆ— -algebra with identity.

4.2. Strong and weak operator convergence In L(H), convergence in the operator norm is sometimes also called uniform convergence. In addition to uniform convergence, in L(H), there are notions of strong operator convergence and weak operator convergence, which are the subject of this section.

4.2. Strong and weak operator convergence

111

Deļ¬nition 4.14. A sequence of operators An āˆˆ L(H) converges strongly to A āˆˆ L(H) if for every v āˆˆ H, An v ā†’ Av as n ā†’ āˆž. We denote this by s An ā†’ A or s-lim An = A. nā†’āˆž

Deļ¬nition 4.15. A sequence of operators An āˆˆ L(H) converges weakly to A āˆˆ L(H) if for every u, v āˆˆ H, u, An v ā†’ u, Av as n ā†’ āˆž. We denote w this by An ā†’ A or w-lim An = A. nā†’āˆž

Similarly to weak convergence in H (Section 3.5), the reader is warned that strong operator convergence and weak operator convergence in L(H) are not deļ¬ned with respect to a metric, so intuitively natural properties must be veriļ¬ed. For instance: s

s

Lemma 4.16. If An ā†’ A and An ā†’ B, then A = B. Proof. This follows from Av = limnā†’āˆž An v = Bv for all v āˆˆ H. s



s

It is obvious that An ā†’ A implies An ā†’ A, and that An ā†’ A implies w An ā†’ A. Exercise 4.3 shows that all three types of operator convergence are equivalent for dim H < āˆž, and Exercise 4.4 that they are distinct when s dim H = āˆž. Moreover, Exercise 4.4 shows that An ā†’ A does not necessarily s imply Aāˆ—n ā†’ Aāˆ— . We will now focus on strong operator convergence, which will be central to our treatment of functional calculus in Chapter 5. We have already encountered strong operator convergence in Theorem 3.31; the series āˆž j=1 Pj considered there converges in the sense of strong operator convergence. It usually does not converge in norm (Exercise 4.5). Weak operator convergence will not play an important role in this text, and its properties will be left as exercises. s

Proposition 4.17. If An ā†’ A, then the sequence An is bounded and A ā‰¤ lim inf An .

(4.4)

nā†’āˆž

Proof. For every v āˆˆ H, the sequence (An v)āˆž n=1 is convergent, so it is bounded. By the uniform boundedness principle, the sequence (An )āˆž n=1 is bounded. The bound on A follows from Av = lim An v = lim inf An v ā‰¤ lim inf An v. nā†’āˆž s

nā†’āˆž s

nā†’āˆž

s

Lemma 4.18. If An ā†’ A and Bn ā†’ B, then An Bn ā†’ AB.



112

4. Bounded linear operators

Proof. For any v āˆˆ H, write An Bn v āˆ’ ABv = An (Bn āˆ’ B)v + (An āˆ’ A)Bv. Since the sequence An is bounded and Bn v ā†’ Bv, it follows that An (Bn āˆ’ B)v ā†’ 0. s

Since An ā†’ A, it follows that (An āˆ’A)Bv ā†’ 0. Adding these two statements completes the proof.  Finally, we show that for separable Hilbert spaces, strong operator convergence is metrizable on bounded subsets of L(H), by revisiting the idea of Theorems 2.56 and 3.58: Theorem 4.19. Let H be a separable Hilbert space with orthonormal basis (ek )āˆž k=1 . Then d(x, y) =

āˆž 

min(2āˆ’k , (A āˆ’ B)ek )

k=1

deļ¬nes a metric on H. Moreover, let (An )āˆž n=1 be a sequence in L(H) and s let A āˆˆ L(H). Then An ā†’ A if and only if supAn  < āˆž

and

nāˆˆN

lim d(An , A) = 0.

nā†’āˆž

Proof. (a) By Lemma 2.55, d is a semimetric. If d(A, B) = 0, then (A āˆ’ B)ek = 0 for all k, so by linearity (A āˆ’ B)v = 0 for all v āˆˆ span{ek | k āˆˆ N} and ļ¬nally by continuity (A āˆ’ B)v = 0 for all v āˆˆ H. Thus, d is a metric on L(H). s

(b) If An ā†’ A, then An is a bounded sequence in H. Moreover, (An āˆ’ A)ek ā†’ 0 for every k, so by dominated convergence with dominating sequence 2āˆ’k applied to the counting measure on N, lim d(xn , x) =

nā†’āˆž

āˆž  k=1

lim min{2āˆ’k , (An āˆ’ A)ek } = 0.

nā†’āˆž

Conversely, if d(xn , x) ā†’ 0, then min(2āˆ’k , (An āˆ’ A)ek ) ā†’ 0 for each k, so An ek ā†’ Aek for each k, and then by linearity, An v ā†’ Av for all v āˆˆ span{ek }āˆž k=1 . Since that set is dense and the sequence An is bounded, s  by Lemma 2.46, An ā†’ A. As usual, this only makes strong operator convergence metrizable on bounded subsets of L(H) and not on the entire space L(H).

4.3. Invertibility, spectrum, and resolvents

113

4.3. Invertibility, spectrum, and resolvents An operator A āˆˆ L(H) is called invertible if it has an inverse in L(H), i.e., if there exists a bounded linear operator Aāˆ’1 such that AAāˆ’1 = Aāˆ’1 A = I. Of course, the inverse has all the algebraic properties guaranteed in any ring: if A is invertible, the inverse is unique, and if A and B are invertible, then so is AB, and (AB)āˆ’1 = B āˆ’1 Aāˆ’1 . There is a simple criterion for invertibility: Lemma 4.20. An operator A āˆˆ L(H) is invertible if and only if Ran A is dense in H and Au > 0. (4.5) inf uāˆˆH u u=0

While the condition that Ran A is dense can be viewed as a weakening of surjectivity, (4.5) can be viewed as a strengthening of injectivity; this perspective will be apparent in the proof. Proof of Lemma 4.20. Denote the inļ¬mum in (4.5) by C. If A is invertible, then Ran A = H is dense. Moreover, for any u, u = Aāˆ’1 Au ā‰¤ Aāˆ’1 Au, which implies that C ā‰„ Aāˆ’1 āˆ’1 > 0. Conversely, assume that Ran A is dense and (4.5) holds. Since C > 0, it follows that Au ā‰„ Cu > 0 whenever u = 0, so Ker A = {0}. This implies injectivity of A, since Au āˆ’ Av = A(u āˆ’ v) = 0 whenever u = v. Moreover, for any convergent sequence vn ā†’ v with vn = Aun āˆˆ Ran A, 1 1 um āˆ’ un  ā‰¤ A(um āˆ’ un ) = vm āˆ’ vn , C C āˆž so (un )n=1 is a Cauchy sequence in H. Thus, (un )āˆž n=1 is convergent in H. Continuity of A implies v = limnā†’āˆž Aun = A limnā†’āˆž un , so v āˆˆ Ran A. Thus, Ran A is closed, and since it is dense, Ran A = H. Therefore, A is a bijection. Finally, C > 0 implies that for any u āˆˆ H, u ā‰¤ C1 Au, so the  inverse Aāˆ’1 is bounded. A lot of information about A can be obtained by considering the invertibility of A āˆ’ z = A āˆ’ zI for z āˆˆ C. Deļ¬nition 4.21. The spectrum of A āˆˆ L(H) is the set Ļƒ(A) = {z āˆˆ C | A āˆ’ z is not invertible}. Its complement C \ Ļƒ(A) is called the resolvent set. For z āˆˆ C \ Ļƒ(A), the inverse RA (z) = (A āˆ’ z)āˆ’1 is called the resolvent of A at z.

114

4. Bounded linear operators

We warn the reader that some sources deļ¬ne the resolvent as the inverse of z āˆ’ A, diļ¬€ering by a minus sign from our convention. Next, we recall some terminology from linear algebra: Deļ¬nition 4.22. Eigenvalues of A are values of z āˆˆ C such that Ker(A āˆ’ z) = {0}. The subspace Ker(A āˆ’ z) is called the eigenspace corresponding to z, and its nonzero elements are called eigenvectors. Of course, since Ker(A āˆ’ z) = {0} prevents invertibility of A āˆ’ z, an eigenvalue of A is always in the spectrum of A. For matrices, the converse also holds: it is a standard result in linear algebra that z is an eigenvalue of A if and only if A āˆ’ z is not invertible. Example 4.23. If A āˆˆ L(Cn ), then Ļƒ(A) is the set of eigenvalues of A. On inļ¬nite-dimensional Hilbert spaces, elements of the spectrum are not necessarily eigenvalues; we will see this in Examples 4.32 and 5.6. We now turn to some general properties of resolvents. Proposition 4.24 (The ļ¬rst resolvent identity). For any z, w āˆˆ / Ļƒ(A), (A āˆ’ z)āˆ’1 āˆ’ (A āˆ’ w)āˆ’1 = (z āˆ’ w)(A āˆ’ z)āˆ’1 (A āˆ’ w)āˆ’1 .

(4.6)

It is easy to motivate the ļ¬rst resolvent identity: it corresponds to a partial fraction decomposition 1 1 zāˆ’w āˆ’ = . xāˆ’z xāˆ’w (x āˆ’ z)(x āˆ’ w) Of course, this is not a proof of (4.6), but it is the ļ¬rst of many indications to come that one can successfully apply scalar functions to operators and expect some properties to carry over. This will be especially true later, when we focus on self-adjoint operators. Proof of Proposition 4.24. The proof is the calculation RA (z)(z āˆ’ w)RA (w) = RA (z)[(A āˆ’ w) āˆ’ (A āˆ’ z)]RA (w) = RA (z)(A āˆ’ w)RA (w) āˆ’ RA (z)(A āˆ’ z)RA (w) = RA (z) āˆ’ RA (w).



Corollary 4.25. Resolvents of A commute, i.e., for any z, w āˆˆ / Ļƒ(A), (A āˆ’ z)āˆ’1 (A āˆ’ w)āˆ’1 = (A āˆ’ w)āˆ’1 (A āˆ’ z)āˆ’1 . Proof. For z = w, this is trivial; for z = w it follows from the ļ¬rst resolvent identity by interchanging z and w and comparing the two equalities. 

4.3. Invertibility, spectrum, and resolvents

115

Just as composition of operators in L(H) is denoted multiplicatively, positive integer powers of an operator are deļ¬ned inductively by B 0 = I and B k = BB kāˆ’1 . This appears in the following expansion, which is merely the geometric series reinvented in the operator setting. Theorem 4.26 (Neumann series). If B āˆˆ L(H) and B < 1, then I āˆ’ B is invertible and the resolvent is given by the norm-convergent series (I āˆ’ B)āˆ’1 =

āˆž 

Bk .

k=0

Moreover, the norm of the resolvent is bounded by (I āˆ’ B)āˆ’1  ā‰¤

1 . 1 āˆ’ B

Proof. By Lemma 2.6, since B k  ā‰¤ Bk , the Neumann series T =

āˆž 

Bk

k=0

is norm-convergent and deļ¬nes an operator T with T  ā‰¤ 1/(1 āˆ’ B). By using telescoping series, both T (I āˆ’ B) and (I āˆ’ B)T are computed to be equal to nāˆ’1  (B k āˆ’ B k+1 ) = lim (I āˆ’ B n ) = I, lim nā†’āˆž

k=0

nā†’āˆž

where the last step uses B n  ā‰¤ Bn ā†’ 0. Thus, T = (I āˆ’ B)āˆ’1 .



This allows us to consider invertibility perturbatively and to start studying (A āˆ’ z)āˆ’1 as an L(H)-valued function of z āˆˆ C \ Ļƒ(A). We will use the discussion of analytic Banach-space valued functions (Deļ¬nition 2.68), which relied on the property (2.27); this property holds in any Banach space, but more concretely, it can be manually proved for the Banach space B = L(H) (Exercise 4.14). The next two statements are applications of the Neumann series, one by viewing A āˆ’ z as a perturbation of A āˆ’ z0 for z near z0 āˆˆ C \ Ļƒ(A), and the other by a perturbation around āˆž. Proposition 4.27. The spectrum Ļƒ(A) is a closed set in C, and the resolvent (A āˆ’ z)āˆ’1 is an L(H)-valued analytic function on z āˆˆ C \ Ļƒ(A). Proof. Fix z0 āˆˆ C \ Ļƒ(A) and denote r = RA (z0 )āˆ’1 . For any z āˆˆ C, we can write A āˆ’ z = (A āˆ’ z0 ) āˆ’ (z āˆ’ z0 ) = [I āˆ’ (z āˆ’ z0 )RA (z0 )](A āˆ’ z0 ).

(4.7)

116

4. Bounded linear operators

For z āˆˆ Dr (z0 ), we have the norm estimate (z āˆ’ z0 )RA (z0 ) < 1, so (4.7) can be inverted by using the Neumann series, āˆž  RA (z) = RA (z0 )[I āˆ’ (z āˆ’ z0 )RA (z0 )]āˆ’1 = (z āˆ’ z0 )k RA (z0 )k+1 . k=0

This shows that Dr (z0 ) āŠ‚ C \ Ļƒ(A) and that the resolvent is locally represented by a convergent power series. Thus, RA (z) is analytic on C\Ļƒ(A).  Proposition 4.28. For any A āˆˆ L(H), Ļƒ(A) is a nonempty compact subset of C and z āˆˆ Ļƒ(A) implies |z| ā‰¤ A. Proof. If |z| > A, then z āˆ’1 A < 1, so the operator Aāˆ’z = āˆ’z(I āˆ’z āˆ’1 A) can be inverted by applying the Neumann series to I āˆ’ z āˆ’1 A. Explicitly, this gives the bounded inverse āˆž  Ak (4.8) RA (z) = āˆ’z āˆ’1 (I āˆ’ z āˆ’1 A)āˆ’1 = āˆ’ z k+1 k=0

for any |z| > A, with the norm estimate 1 1 = . RA (z) ā‰¤ |z|(1 āˆ’ z āˆ’1 A) |z| āˆ’ A Thus, Aāˆ’z is invertible whenever |z| > A. In particular, Ļƒ(A) is bounded. The norm estimate also implies that RA (z) ā†’ 0 as |z| ā†’ āˆž. If Ļƒ(A) was the empty set, RA (z) would be an entire function. Since RA (z) ā†’ 0 as |z| ā†’ āˆž, by Liouvilleā€™s theorem (Proposition 2.70), this would imply RA (z) = 0, which is a contradiction with RA (z)(Aāˆ’z) = I.  The previous proof can be improved by a closer look at where the Neumann series converges, which leads to Gelfandā€™s spectral radius formula. Deļ¬nition 4.29. The spectral radius of A āˆˆ L(H) is r(A) = max |z|. zāˆˆĻƒ(A)

Theorem 4.30. For any A āˆˆ L(H), r(A) = lim An 1/n . nā†’āˆž

The proof of this result requires a lemma about subadditive sequences. Lemma 4.31. Let (xn )nāˆˆN be a sequence in [āˆ’āˆž, āˆž) such that xn+m ā‰¤ xn + xm for all n, m āˆˆ N. Then limnā†’āˆž xn /n exists in [āˆ’āˆž, āˆž) and xn xn = inf . lim nā†’āˆž n nāˆˆN n

4.3. Invertibility, spectrum, and resolvents

117

Proof. For any n āˆˆ N, by induction in k, subadditivity of the sequence implies xkn+r ā‰¤ kxn + xr . Thus, for any 0 ā‰¤ r ā‰¤ n āˆ’ 1, lim sup kā†’āˆž

xn xkn+r kxn + xr ā‰¤ lim sup = . kn + r kn + r n kā†’āˆž

Combining the subsequences (xkn+r )āˆž k=0 for r = 0, 1, . . . , n āˆ’ 1 gives lim sup mā†’āˆž

xm xn ā‰¤ . m n

Since this holds for any n āˆˆ N, it follows that xm xn ā‰¤ inf . lim sup nāˆˆN n mā†’āˆž m Trivially, lim sup ā‰„ lim inf ā‰„ inf, which completes the proof.



Proof of Theorem 4.30. Since Am+n  ā‰¤ Am An  for all m, n āˆˆ N, the sequence logAn  is subadditive, so limnā†’āˆž An 1/n = inf nāˆˆN An 1/n . For |z| > limnā†’āˆž An 1/n , we have % n %1/n % A % % < 1, lim sup % % z n+1 % nā†’āˆž so by the root test, the series (4.8) is absolutely convergent and gives the resolvent, so z āˆˆ / Ļƒ(A). Thus, r(A) ā‰¤ lim An 1/n . nā†’āˆž

(4.9)

For the opposite inequality, consider the substitution z = 1/w to expand the resolvent around āˆž: the resolvent is given by the convergent power series (A āˆ’ 1/w)

āˆ’1

=āˆ’

āˆž 

wk+1 Ak

k=0

in a punctured neighborhood of w = 0. This series has a removable singularity at w = 0 and, by Corollary 2.69, its radius of convergence is at least 1/r(A), which implies 1 lim supkā†’āˆž

Ak 1/k

ā‰„

1 . r(A)

Combining this with (4.9) completes the proof.



Example 4.32. The operators S, S āˆ— on 2 (N) given by (4.2) and (4.3) obey Ļƒ(S) = Ļƒ(S āˆ— ) = {z āˆˆ C | |z| ā‰¤ 1}. The set of eigenvalues of S is {z āˆˆ C | |z| < 1}.

118

4. Bounded linear operators

Proof. Since S = S āˆ—  = 1, Ļƒ(S), Ļƒ(S āˆ— ) āŠ‚ {z āˆˆ C | |z| ā‰¤ 1}. Note that Sv = zv if and only if vn = z nāˆ’1 v1 for all n. Such v āˆˆ 2 (N), v = 0, exist if and only if |z| < 1, so this is the set of eigenvalues of S. Since Ļƒ(S) is closed, it follows that Ļƒ(S) = {z āˆˆ C | |z| ā‰¤ 1}. Likewise, for |z| < 1, since Ker(S āˆ’ z) = (Ran(S āˆ— āˆ’ zĀÆ))āŠ„ , it follows that Ran(S āˆ— āˆ’ zĀÆ) = H, so zĀÆ āˆˆ Ļƒ(S āˆ— ). Since Ļƒ(S āˆ— ) is closed, it follows that  Ļƒ(S āˆ— ) = {z āˆˆ C | |z| ā‰¤ 1}.

4.4. Polynomials of operators Let A āˆˆ L(H). For nonnegative integers k, we have already considered the kth power of A, which is deļ¬ned inductively by A0 = I and Ak = AAkāˆ’1 . The algebra of polynomials with complex coeļ¬ƒcients is denoted C[x]; for any p āˆˆ C[x], we deļ¬ne p(A) = nk=0 ck Ak . The algebraic properties of this notion are apparent. For ļ¬xed A, the map p ā†’ p(A) is a homomorphism of algebras, i.e., it preserves linear operations, multiplication, and the multiplicative identity. It also obeys p(A)āˆ— = pĀÆ(Aāˆ— ). What about the spectrum of p(A)? Theorem 4.33 (Spectral mapping theorem for polynomials). For any A āˆˆ L(H) and p āˆˆ C[x], Ļƒ(p(A)) = {p(Ī») | Ī» āˆˆ Ļƒ(A)}. Proof. Assume that Īŗ āˆˆ / {p(Ī») | Ī» āˆˆ Ļƒ(A)}. This means that the polynomial p(x) āˆ’ Īŗ has no zeros in Ļƒ(A); thus, its factorization into linear factors is of the form n  (x āˆ’ Ī»j ), p(x) āˆ’ Īŗ = Ī± j=1

where Ī± āˆˆ C \ {0} and Ī»j āˆˆ C \ Ļƒ(A) for all j. In particular, p(A) āˆ’ Īŗ = Ī±

n 

(A āˆ’ Ī»j ).

j=1

Since all the A āˆ’ Ī»j are invertible and Ī± = 0, their product p(A) āˆ’ Īŗ is invertible. It follows that Īŗ āˆˆ / Ļƒ(p(A)). Conversely, assume that Īŗ = p(Ī») for some Ī» āˆˆ Ļƒ(A). Then p(x) āˆ’ p(Ī») is divisible by x āˆ’ Ī», and we will use the polynomial factorizations p(x) āˆ’ p(Ī») = q(x)(x āˆ’ Ī») = (x āˆ’ Ī»)q(x). By Lemma 4.20, Ī» āˆˆ Ļƒ(A) implies that A āˆ’ Ī» is not surjective or inf (A āˆ’ Ī»)u = 0.

uāˆˆH u=1

(4.10)

4.5. Invariant subspaces and direct sums of operators

119

If Aāˆ’Ī» is not surjective, then p(A)āˆ’p(Ī») = (Aāˆ’Ī»)q(A) cannot be surjective either. If (4.10) holds, then since q(A) is a bounded operator, inf (p(A) āˆ’ p(Ī»))u = inf q(A)(A āˆ’ Ī»)u = 0,

uāˆˆH u=1

uāˆˆH u=1

so p(A) āˆ’ p(Ī») is not invertible. In both cases, we have proved p(Ī») āˆˆ Ļƒ(p(A)).  We say that operators A, B āˆˆ L(H) commute if AB = BA. If two operators commute, so do their polynomials: Lemma 4.34. If AB = BA, then p(A)q(B) = q(B)p(A) for all p, q āˆˆ C[x]. Proof. The set M = {T āˆˆ L(H) | T B = BT } is closed under multiplication because if T1 , T2 āˆˆ M , then T1 T2 B = T1 BT2 = BT1 T2 . The set M contains the identity operator, and by assumption, it contains A. Thus, by induction, it contains all powers Ak . The set M is closed under linear operations, so it contains all p(A). Thus, AB = BA implies p(A)B = Bp(A) for all p āˆˆ C[x]. Applying this argument again, p(A)B = Bp(A) implies p(A)q(B) = q(B)p(A) for all q āˆˆ C[x].  The polynomial functional calculus deļ¬ned here is very robust, since it allows arbitrary A āˆˆ L(H). However, it treats polynomials as algebraic objects rather than as functions on the spectrum. This distinction is illustrated in Exercise 4.16. In the next chapter, we will deļ¬ne a substantial generalization of f (A) to bounded Borel functions f on the spectrum at the cost of specializing to self-adjoint operators A.

4.5. Invariant subspaces and direct sums of operators In this section, we will introduce direct sums of bounded operators on Hilbert spaces and, as a dual point of view, decompositions of some operators into smaller blocks. The constructions considered here generalize the notion of block diagonal matrices from linear algebra; however, we are working with operators on arbitrary Hilbert spaces, and we consider countable direct sums 4N n=1 with a ļ¬nite or inļ¬nite number of terms (N can be ļ¬nite or āˆž, where āˆž denotes countably many summands). In the general context of vector 4 spaces, for any linear maps An : Hn ā†’ Kn , one can deļ¬ne a linear map N n=1 An by   N 3 N An (vn )N (4.11) n=1 = (An vn )n=1 , n=1

120

4. Bounded linear operators

as a map from one Cartesian product of vector spaces to another. To make this a map between direct sums of Hilbert spaces, we need to ensure that ļ¬niteness of norm is preserved: Proposition 4.35. Given linear maps An : Hn ā†’ Kn between Hilbert spaces, n = 1, . . . , N , (4.11) deļ¬nes a bounded linear operator N 3

An :

N 3

n=1

Hn ā†’

n=1

N 3

Kn

n=1

if and only if each An is bounded and supn An  < āˆž. In this case, % % N % %3 % % An % = supAn . % % % n n=1

4 Proof. Assume An āˆˆ L(Hn , Kn ) and supn An  < āˆž. For v āˆˆ N n=1 Hn , %  %2 N N N % % 3   % % 2 2 v A = A v  ā‰¤ sup A  vn 2 % % n n n n % % n n=1

n=1

4N

n=1

shows that n=1 An is a map between direct sums of Hilbert spaces with 4 norm at most  N n=1 An  ā‰¤ supn An . For the converse, ļ¬x k and note that for all v with vn = 0 for all n = k, Av = Ak vk . Since such v obey v = vk , taking the supremum over normalized vk āˆˆ Hk shows that A ā‰„ Ak  for all k. Since k is arbitrary, we conclude A ā‰„ supk Ak .  4 Deļ¬nition 4.36. The operator N n=1 An is called the direct sum of operators An āˆˆ L(Hn , Kn ). 4 Lemma 4.37. If all the An are unitary, then N n=1 An is unitary. 4 Proof. If all An are unitary, then N An is norm-preserving by a direct 4N n=1 āˆ’1  calculation and has the inverse n=1 An . We now specialize to the case Kn = Hn and describe how direct sums behave with respect to adjoints and invertibility. Proposition 4.38. Let An āˆˆ L(Hn ) for n = 1, . . . , N , and let supn An  < 4N 4N āˆ— āˆ— āˆž. If A = n=1 An , then A = n=1 An . In particular, if all An are self-adjoint, their direct sum is self-adjoint. Proof. This follows from the calculation that, for any v, w āˆˆ v, Aw =

N 

vn , An wn  =

n=1

N 

4N

n=1 Hn ,

Aāˆ—n vn , wn  = (Aāˆ—n vn )N n=1 , w.

n=1



4.5. Invariant subspaces and direct sums of operators

121

Proposition 4.39. Let An āˆˆ L(Hn ) for n = 1, . . . , N , and let supn An  < 4 āˆž. For the direct sum A = N / Ļƒ(A) if and only if z āˆˆ / Ļƒ(An ) for n=1 An , z āˆˆ all n and supn (An āˆ’ z)āˆ’1  < āˆž. For such z, (A āˆ’ z)āˆ’1 =

N 3

(An āˆ’ z)āˆ’1 .

(4.12)

n=1

Proof. For v, w āˆˆ

4N

n=1 Hn ,

we have (A āˆ’ z)v = w if and only if

(An āˆ’ z)vn = wn

āˆ€n.

This system has a unique solution for every w if and only if each An āˆ’ z is a bijection. In this case, the unique solution of the system is vn = (An āˆ’ z)āˆ’1 wn . Thus, A āˆ’ z has a bounded inverse if and only if each An āˆ’ z 4N āˆ’1 is bounded; has a bounded inverse and the linear map n=1 (An āˆ’ z) moreover, in this case, (4.12) holds.  In particular, since Ļƒ(An ) āŠ‚ Ļƒ(A) for all n and Ļƒ(A) is closed, N 

Ļƒ(An ) āŠ‚ Ļƒ(A).

n=1

It is left as an exercise to show that this can be a strict inclusion if N = āˆž. Recall that direct sums of Hilbert spaces are constructed as new Hilbert spaces but can also be used for a decomposition of a Hilbert space into its subspaces. Similarly, direct sums of operators were deļ¬ned as a way to construct new operators but are often used to express a decomposition of an operator into blocks. The existence of such a decomposition depends on the existence of so-called invariant subspaces. Deļ¬nition 4.40. A subspace S āŠ‚ H is invariant for A āˆˆ L(H) if v āˆˆ S implies Av āˆˆ S. Lemma 4.41. Let S be a subspace of H which is invariant for A āˆˆ L(H). Then: (a) S is invariant for A; (b) S āŠ„ is invariant for Aāˆ— . Proof. Any v āˆˆ S can be written as a limit v = limnā†’āˆž vn with vn āˆˆ S. By continuity of A, Av = limnā†’āˆž Avn , so Avn āˆˆ S implies Av āˆˆ S. Let w āˆˆ S āŠ„ . For any v āˆˆ S, we have Av āˆˆ S, and therefore Aāˆ— w, v =  w, Av = 0. Thus, Aāˆ— w āˆˆ S āŠ„ . If S is a closed invariant subspace for A, then the restriction of A to S, denoted A|S , is a bounded linear operator on the Hilbert space S.

122

4. Bounded linear operators

Proposition 4.42. If A āˆˆ L(H) and Hn are closed invariant subspaces for 4 A such that H = N n=1 Hn , then A=

N 3

(A|Hn ).

n=1

Proof. The operators A|Hn are uniformly bounded, so their direct sum is an element of L(H). By deļ¬nition, it agrees with A on each Hn , so by linearity and continuity, the two are equal on H.  Finally, the following proposition considers norm and strong convergence for direct sums of operators: Proposition 4.43. Consider bounded operators An,k āˆˆ L(Hn ), 1 ā‰¤ n ā‰¤ N , k āˆˆ N āˆŖ {āˆž}. Then the following hold. 4 4N (a) If supn An,k āˆ’An,āˆž  ā†’ 0 as k ā†’ āˆž, then N n=1 An,k ā†’ n=1 An,āˆž . s

(b) If supn supkāˆˆN An,k  < āˆž and An,k ā†’ An,āˆž as k ā†’ āˆž for all n, 4 s 4N then N n=1 An,k ā†’ n=1 An,āˆž . Proof. (a) This follows from % % N N % %3 3 % % A āˆ’ A % n,āˆž % = supAn,k āˆ’ An,āˆž . n,k % % n n=1

n=1

(b) The operators are uniformly bounded, so it suļ¬ƒces to prove convergence on a dense set of vectors. If v = (vn )N n=1 is such that vn = 0 4 4N for all but ļ¬nitely many n, then n=1 An,k v ā†’ N n=1 An,āˆž v follows from  An,k vn ā†’ An,āˆž vn . We say that an operator A āˆˆ L(H) has an orthonormal basis of eigenvectors if there is an orthonormal basis {vn }N n=1 of H such that each vn is an eigenvector of H. Reformulating, operators with an orthonormal basis of eigenvectors are precisely those that can be represented as a direct sum of multiplication operators on one-dimensional subspaces {cvn | c āˆˆ C}. We will soon focus on self-adjoint operators, which do not always have eigenvectors; however, the direct sum formalism will still be used to decompose self-adjoint operators into simpler blocks of a standard form.

4.6. Compact operators In this section, we consider compact operators, a subclass of bounded operators with properties reminiscent of the ļ¬nite-dimensional case, and we

4.6. Compact operators

123

describe an important class of examples known as compact integral operators. In the Hilbert space setting, compact operators can be deļ¬ned by the following convergence-improving property: Deļ¬nition 4.44. An operator K āˆˆ L(H) is called compact if for every w weakly convergent sequence un ā†’ u, Kun ā†’ Ku. Recall that, if dim H < āˆž, weak convergence is equivalent to strong convergence, so every bounded operator is compact. If H is inļ¬nite dimensional, this is no longer the case: the identity operator I is not compact, w since there exist un ā†’ 0 such that un ā†’ 0. Compactness of an operator can be characterized in terms of the image of the unit ball: Proposition 4.45. Denote by B = {u āˆˆ H | u ā‰¤ 1} the closed unit ball in a Hilbert space H. For K āˆˆ L(H), the following are equivalent: (a) K is a compact operator; (b) the image K(B) = {Ku | u āˆˆ B} is a precompact subset of H; (c) the image K(B) = {Ku | u āˆˆ B} is a compact set. Proof. (a) =ā‡’ (c): For any sequence vn āˆˆ K(B), we must prove existence of a convergent subsequence in K(B). Write vn = Kun , un āˆˆ B. By Theorem 3.57, the sequence (un )āˆž n=1 has a weakly convergent subsequence w unk ā†’ u as k ā†’ āˆž, and from u ā‰¤ lim inf kā†’āˆž unk , we conclude u āˆˆ B. By compactness of K, this implies vnk = Kunk ā†’ Ku āˆˆ K(B). (c) =ā‡’ (b): This is trivial. w

(b) =ā‡’ (a): For any weakly convergent sequence un ā†’ u, we must prove Kun ā†’ Ku. Any weakly convergent sequence is bounded, so by w rescaling, we assume un  ā‰¤ 1 for all n. Since K is bounded, un ā†’ u implies w Kun ā†’ Ku by Lemma 4.8. Thus, any strongly convergent subsequence of Kun must converge to Ku; otherwise, it would weakly converge to two diļ¬€erent limits. In other words, the sequence Kun āˆˆ K(B) has only one possible limit point Ku in H. By precompactness of K(B), this implies  Kun ā†’ Ku. An operator F āˆˆ L(H) is called ļ¬nite-rank if Ran F is ļ¬nite dimensional. Corollary 4.46. Any ļ¬nite-rank operator is compact. Proof. If F is ļ¬nite-rank, then F (B) is a bounded subset of the ļ¬nitedimensional space Ran F , so F (B) is compact; thus, F is compact.  Proposition 4.47. The set of compact operators is closed in L(H).

124

4. Bounded linear operators

Proof. Let Kk be compact operators such that Kk ā†’ K as k ā†’ āˆž. For w any weakly convergent sequence un ā†’ u, Kk un ā†’ Kk u for any k, so from Kun āˆ’ Ku ā‰¤ Kun āˆ’ Kk un  + Kk un āˆ’ Kk u + Kk u āˆ’ Ku we conclude lim supKun āˆ’ Ku ā‰¤ 2Kk āˆ’ K supun . nā†’āˆž

n

Since Kk ā†’ K as k ā†’ āˆž, Kk āˆ’ K can be made arbitrarily small, so lim supnā†’āˆž Kun āˆ’ Ku = 0. Thus, Kun ā†’ Ku as n ā†’ āˆž, so K is compact.  By the previous two statements, if an operator can be approximated in norm by ļ¬nite rank operators, then it is compact. We will prove the converse in Theorem 5.25. Obviously, linear combinations of compact operators are compact. In fact, compact operators form an ideal in L(H): Lemma 4.48. Let K, A āˆˆ L(H). If K is compact, then AK and KA are compact. w

w

Proof. If un ā†’ u, then Kun ā†’ Ku so AKun ā†’ AKu. Similarly, if un ā†’ u, w then Aun ā†’ Au by Lemma 4.8, so KAun ā†’ KAu.  Lemma 4.49. For K āˆˆ L(H), K is compact if and only if K āˆ— K is compact. Proof. If K is compact, then so is K āˆ— K by the previous lemma. Conversely, w w if K āˆ— K is compact and un ā†’ u, then un āˆ’ u ā†’ 0 so K āˆ— K(un āˆ’ u) ā†’ 0. Since (un )āˆž n=1 is bounded, this implies K(un āˆ’ u)2 = un āˆ’ u, K āˆ— K(un āˆ’ u) ā†’ 0, i.e., Kun ā†’ Ku.



Lemma 4.50. For K āˆˆ L(H), K is compact if and only if K āˆ— is compact. Proof. If K āˆ— is compact, then K āˆ— K is compact, so K is compact by the previous two lemmas. Analogously, if K is compact, then K āˆ— is compact.  Finally, we describe an important family of compact operators, called compact integral operators. Integral operators on L2 (X, dĪ¼) are deļ¬ned by an integral kernel, which is a function on X Ɨ X, and it is customary to denote both the kernel and the operator by the same letter: Proposition 4.51. Let X be Ļƒ-locally compact metric space with a Baire measure Ī¼. Let K āˆˆ L2 (X Ɨ X, d(Ī¼ āŠ— Ī¼)). Then the integral operator K, deļ¬ned by  K(x, y)u(y) dĪ¼(y), (Ku)(x) = X

4.7. Exercises

125

is a compact operator on L2 (X, dĪ¼). Its adjoint is the integral operator with kernel K āˆ— (x, y) = K(y, x). Proof. By the Cauchyā€“Schwarz inequality, for any x āˆˆ X,   !1/2    2  K(x, y)u(y) dĪ¼(y) ā‰¤ u |K(x, y)| dĪ¼(y) ,   X

X

so squaring and integrating in x gives   2 2 |K(x, y)|2 dĪ¼(y) dĪ¼(x). Ku ā‰¤ u X

This shows that K is a bounded operator and !1/2   2 |K(x, y)| dĪ¼(y) dĪ¼(x) = K ā‰¤ X

(4.13)

X

X

!1/2

 |K| d(Ī¼ āŠ— Ī¼) 2

.

XƗX

Moreover, if Kn is a sequence of integral kernels converging to K in L2 (X Ɨ X, d(Ī¼ āŠ— Ī¼)), applying this norm estimate to K āˆ’ Kn shows norm convergence of operators, Kn ā†’ K in L(H). We now use density arguments to approximate K āˆˆ L2 (X ƗX, d(Ī¼āŠ—Ī¼)). Ėœ āˆˆ Cc (X Ɨ X) such that K āˆ’ K Ėœ 2 < . For any ļ¬xed > 0, there exists K Using Cartesian projections Ļ€j : X Ɨ X ā†’ X, j = 1, 2, we obtain compacts Ėœ By Stoneā€“Weierstrass, K Ėœ can be approximated by linear comĻ€j (supp K). Ėœ Ėœ g āˆˆ C(Ļ€2 (supp K)). Thus, binations of f (x)g(y) with f āˆˆ C(Ļ€1 (supp K)), 2 K can be approximated in L (X Ɨ X, d(Ī¼ āŠ— Ī¼)) by kernels of the form F (x, y) =

k 

fj (x)gj (y).

j=1

For any such kernel F , the corresponding integral operator is ļ¬nite rank: Ran F āŠ‚ span{fj | 1 ā‰¤ j ā‰¤ k}. Thus, the integral operator K can be approximated in operator norm by ļ¬nite rank operators, so it is compact. 

4.7. Exercises 4.1. For A āˆˆ L(H), prove that A = w

sup

Reu, Av.

u,vāˆˆH u=v=1

w

4.2. If An ā†’ A and An ā†’ B, prove that A = B. 4.3. If H is a ļ¬nite-dimensional Hilbert space, prove that in L(H), norm convergence, strong operator convergence, and weak operator convergence are all mutually equivalent.

126

4. Bounded linear operators

4.4. Recall the operators S, S āˆ— on 2 (N) given by (4.2) and (4.3). Prove that: (a) S n converges strongly, but not in norm, to 0 as n ā†’ āˆž. (b) (S āˆ— )n converges weakly, but not strongly, to 0 as n ā†’ āˆž. 4.5. If Pj , j āˆˆ N are orthogonal projections to mutually orthogonal sub P spaces and Ran Pj = {0} for all j, prove that the series āˆž j=1 j is not convergent in norm. w

4.6. If An ā†’ A, prove that the sequence An is bounded and (4.4) holds. w

s

w

w

w

4.7. If An ā†’ A and Bn ā†’ B, prove that An Bn ā†’ AB. 4.8. If An ā†’ A and Bn ā†’ B, prove that An Bn ā†’ AB. s

4.9. If dim H = āˆž, prove that there exist sequences such that An ā†’ A w w and Bn ā†’ B, but An Bn ā†’  AB. 4.10. Construct a sequence (An )āˆž n=1 in L(H) that obeys d(An , 0) ā†’ 0 but is not strongly convergent. 4.11. Let H be an inļ¬nite-dimensional separable Hilbert space with orthonormal basis (ej )āˆž j=1 . Deļ¬ne a metric d such that, for any sequence w āˆž (An )n=1 in L(H) and A āˆˆ L(H), An ā†’ A if and only if (An )āˆž n=1 is bounded and d(An , A) ā†’ 0 as n ā†’ āˆž. 4.12. Let H be an inļ¬nite-dimensional separable Hilbert space. For each of the following statements, determine whether it is true or false: (a) Every bounded sequence in L(H) has a strongly convergent subsequence. (b) Every bounded sequence in L(H) has a weakly convergent subsequence. 4.13. Prove the second resolvent identity: if A, B āˆˆ L(H) have resolvents at z, then RA (z) āˆ’ RB (z) = RA (z)(B āˆ’ A)RB (z). 4.14. (a) Let Ī› āˆˆ L(H)āˆ— be given by Ī›(A) = u, Av for some u, v āˆˆ H. Prove that Ī› = uv. (b) Prove (2.27) for B = L(H). Hint: Use Lemma 4.1. 4.15. Find all eigenvalues of the operator S āˆ— given by (4.3) on 2 (N). 4.16. (a) Assume that A is a diagonalizable n Ɨ n matrix, i.e., there exists a unitary V such that V āˆ’1 AV is a diagonal matrix. Use this to compute V āˆ’1 p(A)V for any polynomial p, and to prove that if p(Ī») = q(Ī») for all Ī» āˆˆ Ļƒ(A), then p(A) = q(A).

4.7. Exercises

127

(b) For the Jordan block A=

! t 1 , 0 t

ļ¬nd a polynomial p such that p(Ī») = 0 for all Ī» āˆˆ Ļƒ(A) but p(A) = 0. (c) For any n Ɨ n matrix A which is not diagonalizable, prove that there exists p āˆˆ C[x] such that p(Ī») = 0 for all Ī» āˆˆ Ļƒ(A) but p(A) = 0. 4.17. Prove that A āˆˆ L(H) is an orthogonal projection if and only if A = Aāˆ— and Ļƒ(A) āŠ‚ {0, 1}. 4.18. Give an example of bounded linear operators An on separable Hilbert spaces Hn , n āˆˆ N, such that supn An  < āˆž and  āˆž āˆž 3  An āŠ‚ Ļƒ(An ). Ļƒ n=1

n=1

4.19. Prove that there exist operators on Hilbert spaces 4āˆžHn , n āˆˆ N, such k ā†’ āˆž for all n, but that An,k ā†’ An,āˆž as n=1 An,k does not 4āˆž converge in norm to n=1 An,āˆž .

Chapter 5

Bounded self-adjoint operators

In this chapter, we consider bounded self-adjoint operators on separable Hilbert spaces: Deļ¬nition 5.1. A āˆˆ L(H) is self-adjoint if Aāˆ— = A. Self-adjoint operators are the natural generalization of Hermitian matricesā€”deļ¬ned in linear algebra as n Ɨ n matrices A such that Aij = Aji for all i, j (compare Example 4.4). A central result is that every Hermitian matrix is diagonalizable; i.e., there exists an orthonormal basis {v1 , . . . , vn } of Cn consisting of eigenvectors of A. This is called diagonalizability because if we assemble an n Ɨ n matrix V out of the eigenvectors by Vij = (vi )j and a diagonal matrix D out of the corresponding eigenvalues by Djj = Ī»j and Dij = 0 for i = j, then V is unitary and AV = V D. In other words, A is represented in the form A = V DV āˆ’1 . A commonly demonstrated ļ¬rst application in linear algebra is an eļ¬ƒcient method for computing high powers of a diagonalizable matrix using An = V D n V āˆ’1 . More fundamentally, diagonalizability leads to a classiļ¬cation of all Hermitian matrices up to unitary equivalence: two Hermitian matrices are unitarily equivalent if and only if they have the same eigenvalues with the same multiplicities. Our goal in this chapter is a generalization of the above discussion, and much more, on separable Hilbert spaces. We will prove that compact selfadjoint operators have an orthonormal basis of eigenvectors (in particular, 129

130

5. Bounded self-adjoint operators

this includes diagonalizability of Hermitian matrices), but the bulk of the chapter is dedicated to the general setting, not assuming compactness. The central result is called the spectral theorem for bounded self-adjoint operators. When dim H = āˆž, we cannot describe every self-adjoint operator A in terms of eigenvalues and eigenvectors, since there may not be any (Example 5.6). Thus, inevitably, the spectral theorem for bounded self-adjoint operators will appear diļ¬€erent from the compact case. Instead of individual eigenvectors, we will work with unitary maps, and diagonal matrices will be replaced by multiplication operators and their direct sums. We say that operators A āˆˆ L(H), B āˆˆ L(K) are unitarily equivalent if there exists a unitary U āˆˆ L(H, K) such that U AU āˆ’1 = B and denote this by A āˆ¼ = B. In that terminology, we prove that every bounded self-adjoint operator is unitarily equivalent to a direct sum of multiplication operators. One of the main applications will be to deļ¬ne in a consistent way functions of self-adjoint operators g(A) for bounded Borel functions g : Ļƒ(A) ā†’ C. This is called the Borel functional calculus for self-adjoint operators. We already know the meaning of p(A) if p is a polynomial, but the Borel functional calculus provides a vast generalization for self-adjoint operators A.

5.1. A ļ¬rst look at self-adjoint operators We begin the chapter with some general consequences of self-adjointness. Lemma 5.2. Let A be a self-adjoint operator, and let z āˆˆ C \ R. For any u āˆˆ H, (A āˆ’ z)u ā‰„ |Im z|u. (5.1) Proof. For z = x + iy, x, y āˆˆ R, (5.1) follows from the calculation (A āˆ’ z)u2 = (A āˆ’ x āˆ’ iy)u, (A āˆ’ x āˆ’ iy)u = (A āˆ’ x)u, (A āˆ’ x)u āˆ’ iy(A āˆ’ x)u, u + iyu, (A āˆ’ x)u + (āˆ’iy)(iy)u, u = (A āˆ’ x)u2 + y 2 u2 .



For z āˆˆ / R, (5.1) is a kind of strong injectivity condition on A āˆ’ z. This leads to a general result about invertibility of A āˆ’ z: Corollary 5.3. If A is self-adjoint, then Ļƒ(A) āŠ‚ R. Proof. For z āˆˆ C \ R, (5.1) implies that Ker(A āˆ’ z) = {0}. Applying (5.1) also to zĀÆ and using Proposition 4.7 gives (Ran(A āˆ’ z))āŠ„ = Ker(Aāˆ— āˆ’ zĀÆ) = Ker(A āˆ’ zĀÆ) = {0}, so Ran(A āˆ’ z) is dense. Thus, A āˆ’ z is invertible by Lemma 4.20.



5.1. A ļ¬rst look at self-adjoint operators

131

Corollary 5.3 focuses our remaining interest on invertibility of A āˆ’ Ī» for Ī» āˆˆ R. We reļ¬ne Lemma 4.20 to the setting of self-adjoint operators: Lemma 5.4 (Weylā€™s criterion). Let A be self-adjoint, and let Ī» āˆˆ R. The operator A āˆ’ Ī» is invertible if and only if inf

uāˆˆH u=0

(A āˆ’ Ī»)u > 0. u

Proof. If the inļ¬mum is strictly positive, then Ran(A āˆ’ Ī»)āŠ„ = Ker(A āˆ’ Ī») = {0}, so Ran(Aāˆ’Ī») is dense. Thus, Aāˆ’Ī» is invertible by Lemma 4.20. Conversely, if the inļ¬mum is 0, then A āˆ’ Ī» is not invertible by Lemma 4.20.  Weylā€™s criterion is usually restated in the following equivalent form. Proposition 5.5 (Weylā€™s criterion). Let A be self-adjoint and let Ī» āˆˆ R. Then Ī» āˆˆ Ļƒ(A) if and only if there exists a sequence (un )āˆž n=1 of normalized vectors such that lim (A āˆ’ Ī»)un  = 0. nā†’āˆž

The vectors un in Weylā€™s criterion are sometimes described as approximate eigenvectors for Ī» āˆˆ Ļƒ(A). This perspective is useful, because elements of the spectrum are not necessarily eigenvalues: Example 5.6. Let A be the operator on L2 ([0, 1]) = L2 ([0, 1], dx) given by (Af )(x) = xf (x). The operator A is self-adjoint and has no eigenvalues, but Ļƒ(A) = [0, 1]. Proof. The operator A is self-adjoint because for all f, g āˆˆ L2 ([0, 1]),  1  1 f (x)xg(x) dx = xf (x)g(x) dx = Af, g. f, Ag = 0

0

(xāˆ’z)āˆ’1

is bounded on [0, 1], so multiplication For z āˆˆ C\[0, 1], the function by (x āˆ’ z)āˆ’1 is a bounded operator on L2 ([0, 1]): for any f āˆˆ L2 ([0, 1]),      1  f (x) 2  1 2 1    max  |f (x)|2 dx.   x āˆ’ z  dx ā‰¤ xāˆˆ[0,1] x āˆ’ z 0 0 / Ļƒ(A). Thus, multiplication by (xāˆ’z)āˆ’1 is a bounded inverse for Aāˆ’z, so z āˆˆ Fix Ī» āˆˆ [0, 1]. For any > 0, the characteristic function f = Ļ‡[Ī»āˆ’,Ī»+] is a nonzero element of L2 ([0, 1]) and, since |(x āˆ’ Ī»)f (x)| ā‰¤ |f (x)| for all x, (A āˆ’ Ī»)f  ā‰¤ f .

132

5. Bounded self-adjoint operators

Thus, by Weylā€™s criterion, A āˆ’ Ī» does not have a bounded inverse, so Ī» āˆˆ Ļƒ(A). In conclusion, Ļƒ(A) = [0, 1]. However, Af = Ī»f implies that xf (x) = Ī»f (x) for Lebesgue-a.e. x, so f (x) = 0 for Lebesgue-a.e. x = Ī», and therefore f = 0 as an element of L2 ([0, 1]). Thus, A has no eigenvalues.  If a self-adjoint operator has any eigenvalues and eigenvectors, their properties are analogous to those for Hermitian matrices: Lemma 5.7. If z is an eigenvalue of a self-adjoint operator A, then z āˆˆ R. Proof. If z is an eigenvalue of A, then z āˆˆ Ļƒ(A), so z āˆˆ R.



Lemma 5.8. If Ī» = Īŗ are eigenvalues of a self-adjoint operator A and u, v the corresponding eigenvectors, then u āŠ„ v. ĀÆ this follows from Proof. Since Īŗ = Ī» = Ī», ĀÆ v = Ī»u, v = Au, v = u, Av = u, Īŗv = Īŗu, v. Ī»u,



We will now describe the spectral radius of a self-adjoint operator, r(A) = max |Ī»|. Ī»āˆˆĻƒ(A)

Proposition 5.9. If A is self-adjoint, then r(A) = A. We show a proof using the spectral radius formula (Theorem 4.30), and a direct proof using only the Neumann series (Theorem 4.26). Proof using the spectral radius formula. Since A is self-adjoint and L(H) is a C āˆ— -algebra, A2  = Aāˆ— A = A2 . k

k

Thus, A2  = A2 for any k āˆˆ N, so by the spectral radius formula, k

k

r(A) = lim An 1/n = lim A2 1/2 = A. nā†’āˆž

kā†’āˆž



Proof using only the Neumann series. If |Ī»| > A, then Ī»āˆ’1 A < 1, so by the Neumann series, the operator A āˆ’ Ī» = āˆ’Ī»(I āˆ’ Ī»āˆ’1 A) is invertible. Thus, r(A) ā‰¤ A. For the converse, it suļ¬ƒces to show that A or āˆ’A is in the spectrum. Since products of invertible operators are invertible, it suļ¬ƒces to show that A2 āˆ’ A2 = (A āˆ’ A)(A + A) is not invertible. By self-adjointness and the deļ¬nition of operator norm, there exist normalized vectors un such that lim un , A2 un  = lim Aun 2 = A2 .

nā†’āˆž

nā†’āˆž

(5.2)

5.1. A ļ¬rst look at self-adjoint operators

133

By elementary estimates, A2 un āˆ’ A2 un 2 = A2 un 2 āˆ’ 2A2 Reun , A2 un  + A4 un 2 ā‰¤ 2A4 un 2 āˆ’ 2A2 un , A2 un . The right-hand side converges to 0 by (5.2), so lim A2 un āˆ’ A2 un 2 = 0.

nā†’āˆž

Thus, by Weylā€™s criterion, A2 āˆ’ A2 is not invertible.



The previous proof relied on the inner product u, A2 u which, due to the square, is easy to rewrite as a square of a norm. However, it is often useful to consider the quantity u, Au. This leads us to a nontrivial improvement of Lemma 4.1 for self-adjoint operators. As a preliminary, we note that by self-adjointness and skew-symmetry of the inner product, u, Au = Au, u = u, Au, so u, Au āˆˆ R for all u āˆˆ H. Proposition 5.10. If A āˆˆ L(H) is self-adjoint, then A = sup |u, Au|. uāˆˆH u=1

Proof. Denote by C the supremum in the statement. By Lemma 4.1, C ā‰¤ A. For the converse, recall that A or āˆ’A are in the spectrum of A. For a choice of Ā± sign for which Ā±A āˆˆ Ļƒ(A), by Weylā€™s criterion, there exist normalized vectors un such that lim (A āˆ“ A)un = 0.

nā†’āˆž

This implies by the Cauchyā€“Schwarz inequality that lim un , (A āˆ“ A)un  = 0

nā†’āˆž

and ļ¬nally that lim un , Aun  = Ā± lim un , Aun  = Ā±A,

nā†’āˆž

which proves that C ā‰„ A.

nā†’āˆž



By shifting the operator by constants, we can remove the absolute value in the previous proposition.

134

5. Bounded self-adjoint operators

Proposition 5.11. If A is self-adjoint, then min Ļƒ(A) = inf u, Au,

(5.3)

max Ļƒ(A) = sup u, Au.

(5.4)

uāˆˆH u=1

uāˆˆH u=1

Proof. Applying Proposition 5.10 to A āˆ’ c for arbitrary c āˆˆ R, we see that max |x| = sup |u, (A āˆ’ c)u| = sup |u, Au āˆ’ c|.

xāˆˆĻƒ(Aāˆ’c)

u=1

u=1

By the spectral mapping theorem, Ļƒ(A āˆ’ c) = {Ī» āˆ’ c | Ī» āˆˆ Ļƒ(A)}, so this can be rewritten as max |Ī» āˆ’ c| = sup|Ī» āˆ’ c|, (5.5) Ī»āˆˆĻƒ(A)

Ī»āˆˆS

where S = {u, Au | u = 1}. Both S and Ļƒ(A) are contained in [āˆ’A, A]. Thus, for c ā‰¤ āˆ’A, all expressions |Ī» āˆ’ c| in (5.5) are equal to Ī» āˆ’ c, and (5.5) implies max Ļƒ(A) = sup S, which is (5.4). Similarly, for c ā‰„ A, (5.5) implies (5.3).  The set S from the previous proof is connected as the continuous image of the unit sphere in H, so it cannot be expected to provide any further information about Ļƒ(A) beyond (5.3) and (5.4). A more sophisticated generalization of Proposition 5.11, called the min-max principle, will be discussed later. We next deļ¬ne positivity for self-adjoint operators. This generalizes the notion of positive semi-deļ¬niteness of matrices, often encountered in the context of the second derivative test of functions of several variables. Deļ¬nition 5.12. A self-adjoint operator A is said to be positive if u, Au ā‰„ 0 for all u āˆˆ H, and we denote this by A ā‰„ 0. As an immediate corollary of (5.3), we obtain a criterion for positivity of A in terms of Ļƒ(A). Corollary 5.13. A ā‰„ 0 if and only if Ļƒ(A) āŠ‚ [0, āˆž). This notion of positivity is also used to deļ¬ne a partial order relation: Deļ¬nition 5.14. If A, B are self-adjoint operators on H, we say that A ā‰¤ B if B āˆ’ A ā‰„ 0. Lemma 5.15. The relation A ā‰¤ B is a partial order on the set of bounded self-adjoint operators on H.

5.1. A ļ¬rst look at self-adjoint operators

135

Proof. If A = 0, then v, Av = 0 for all v āˆˆ H, so A ā‰„ 0. This implies reļ¬‚exivity. If A ā‰„ 0 and āˆ’A ā‰„ 0, then v, Av = 0 for all v āˆˆ H, so A = 0 by Proposition 5.10. This implies antisymmetry. If A ā‰¤ B and B ā‰¤ C, then v, (C āˆ’ A)v = v, (C āˆ’ B)v + v, (B āˆ’ A)v ā‰„ 0 for all v āˆˆ H, so A ā‰¤ C. This implies transitivity.  Most of this chapter is devoted to the study of a single self-adjoint operator. However, we often study an operator A by approximating it by some ā€œsimplerā€ operators An , so in the rest of this section, we consider how the spectrum behaves with respect to norm and strong convergence of selfadjoint operators. The Hausdorļ¬€ distance between nonempty subsets S, T of a metric space is   dH (S, T ) = max sup inf d(x, y), sup inf d(x, y) . xāˆˆS yāˆˆT

yāˆˆT xāˆˆS

This deļ¬nes a metric on nonempty compact subsets of the metric space (Exercise 5.4). Proposition 5.16. If A, B āˆˆ L(H) are self-adjoint, then dH (Ļƒ(A), Ļƒ(B)) ā‰¤ A āˆ’ B. Proof. For z āˆˆ Ļƒ(A), from (5.16) and (B āˆ’ z)u ā‰¤ (A āˆ’ z)u + A āˆ’ Bu, we conclude that dist(z, Ļƒ(B)) ā‰¤ A āˆ’ B. Repeating the argument with the roles of A and B reversed concludes the proof.  Corollary 5.17. If An ā†’ Aāˆž , then Ļƒ(An ) ā†’ Ļƒ(Aāˆž ) in Hausdorļ¬€ distance. Thus, in case of norm convergence, the spectra of An uniquely determine the spectrum of Aāˆž . This does not hold for strong operator convergence (Exercise 5.5), but one inclusion holds: s

Proposition 5.18. For self-adjoint operators An with An ā†’ Aāˆž , Ļƒ(Aāˆž ) āŠ‚ {Ī» āˆˆ R | lim dH ({Ī»}, Ļƒ(An )) = 0}. nā†’āˆž

Proof. Let Ī» āˆˆ Ļƒ(Aāˆž ). For any > 0 there exists normalized v with (Aāˆž āˆ’ Ī»)v < . By strong operator convergence, for large enough n, (An āˆ’ Ī»)v < 2 , which implies d(Ī», Ļƒ(An )) < 2 . Since this holds for any

> 0, the claim follows.  This inclusion has an immediate corollary: Corollary 5.19. Let F be a closed subset of R. For self-adjoint operators s An with An ā†’ Aāˆž , if Ļƒ(An ) āŠ‚ F for all n, then Ļƒ(Aāˆž ) āŠ‚ F .

136

5. Bounded self-adjoint operators

5.2. Spectral theorem for compact self-adjoint operators In the introduction to this chapter, the study of self-adjoint operators was motivated by the spectral theorem for Hermitian matrices: Theorem 5.20. If A is an n Ɨ n Hermitian matrix, then there exists an orthonormal basis {v1 , . . . , vn } of Cn consisting of eigenvectors of A. As promised, we will prove a generalization of that statement: Theorem 5.21 (Spectral theorem for compact self-adjoint operators). If K is a compact self-adjoint operator on a separable Hilbert space H, then there H exists an orthonormal basis (vn )dim n=1 of H consisting of eigenvectors of K. Moreover, if dim H = āˆž and we denote by Ī»n the eigenvalues for vn , then limnā†’āˆž Ī»n = 0. The special case H = Cn recovers Theorem 5.20. For the general setting, the key lemma follows. Lemma 5.22. Let K be a compact self-adjoint operator on a nontrivial Hilbert space H. Then K or āˆ’K is an eigenvalue. Proof. By Proposition 5.10, there exists a sequence (un )āˆž n=1 of normalized vectors such that |un , Kun | ā†’ K. By considering K or āˆ’K, we can assume without loss of generality that un , Kun  ā†’ K. This sequence has a weakly convergent subsequence, which by relabelling we denote also by (un )āˆž n=1 , and we denote u = w-limnā†’āˆž un . By compactness of K, Kun ā†’ Ku, so by Lemma 3.55, un , Kun  ā†’ u, Ku. Thus, u, Ku = K. Weak convergence implies u ā‰¤ 1, so K = u, Ku ā‰¤ uKu ā‰¤ Ku2 ā‰¤ K. This implies that u = 1 and Ku = u, Ku = K, so Ku āˆ’ Ku2 = Ku2 āˆ’ 2Ku, Ku + K2 u2 = 0, so Ku = Ku.



Proof of Theorem 5.21. For any Ī» āˆˆ C, Ker(K āˆ’ Ī») is a closed subspace of H, so it has an orthonormal basis (which may be empty). Since eigenvectors corresponding to distinct eigenvalues are mutually orthogonal, the union over Ī» of those orthonormal bases is an orthonormal set in H, which we denote by {vn | n = 1, . . . , N } with N ļ¬nite or āˆž. Consider M = span{vn | n = 1, . . . , N }. In other words, M is the direct sum of subspaces Ker(K āˆ’ Ī») for Ī» āˆˆ C; of course, since H is separable, only countably many of those subspaces can be

5.2. Spectral theorem for compact self-adjoint operators

137

nontrivial. Since vn are eigenvectors, their span is an invariant subspace for K, and so is its closure M . By Lemma 4.41, M āŠ„ is also an invariant subspace for K. The restriction of K to M āŠ„ is a compact self-adjoint operator on M āŠ„ , because it obeys u, Kv = Ku, v for all u, v āˆˆ M āŠ„ and takes weakly convergent sequences to strongly convergent sequences. However, K|M āŠ„ has no eigenvectors by the construction of M . Thus, Lemma 5.22 implies M āŠ„ = {0}. Thus, M is a dense subspace of H, so (vn )N n=1 is an orthonormal basis of H. w

If H is inļ¬nite dimensional, vn ā†’ 0 implies Ī»n vn = Kvn ā†’ 0. Since  vn  = 1 for all n, this implies |Ī»n | = Ī»n vn  ā†’ 0. Expanding vectors with respect to the orthonormal basis (vn )N n=1 gives 4N N a unitary map V : H ā†’ n=1 C where V u = (vn , u)n=1 (Theorem 3.40). Since the vn are eigenvectors, V KV āˆ’1 has a particularly simple form 4 which generalizes the diagonalization of Hermitian matrices: for any f āˆˆ N n=1 C, (V KV āˆ’1 f )n = Ī»n fn . The remainder of this section is an aside, an application to arbitrary compact operators. Theorem 5.23 (Singular value decomposition). Any compact operator K on H can be represented in the form Kv =

N 

Ī¼n en , vfn ,

(5.6)

n=1 N where (en )N n=1 and (fn )n=1 are orthonormal families in H, Ī¼n > 0 for all n, and Ī¼n ā†’ 0 if N = āˆž. If N = āˆž, (5.6) denotes a norm-convergent series.

Before proving this, let us explain the convergence of a series such as (5.6). Lemma 5.24. (a) Let (ej )nj=m , (fj )nj=m be two ļ¬nite orthonormal families in H, and let Ī¼m ā‰„ Ā· Ā· Ā· ā‰„ Ī¼n > 0. Deļ¬ne a ļ¬nite rank operator F āˆˆ L(H) by Fv =

n 

Ī¼j ej , vfj .

j=m

Then F  = Ī¼m . āˆž (b) Let (ej )āˆž j=1 , (fj )j=1 be two orthonormal families in H, and let (Ī¼j )āˆž j=1 be a decreasing sequence with Ī¼j ā†’ 0 as j ā†’ āˆž. Then

138

5. Bounded self-adjoint operators

the series T =

āˆž 

Ī¼j ej , Ā·fj

j=1

is norm convergent and deļ¬nes a compact operator T with T  = Ī¼1 . Proof. (a) For any v āˆˆ H, by Besselā€™s inequality, nj=m |ej , v|2 ā‰¤ v2 , so F v2 =

n 

Ī¼j ej , vfj 2 =

j=m

n 

Ī¼2j |ej , v|2 ā‰¤ Ī¼2m v2 .

j=m

Moreover, equality holds for v = em . Thus, F  = Ī¼m . (b) Denote Fn = nj=1 Ī¼j ej , Ā·fj . By (a), Fn āˆ’Fm  = Ī¼m+1 for m < n, so the Fn form a Cauchy sequence in L(H). Its limit K is a bounded operator and, since Fn  = Ī¼1 for all n, K = Ī¼1 . Moreover, the operators Fn are ļ¬nite rank and Fn ā†’ K, so by Corollary 4.46 and Proposition 4.47, K is compact.  Proof of Theorem 5.23. The operator K āˆ— K is compact and self-adjoint. Moreover, K āˆ— K ā‰„ 0, because v, K āˆ— Kv = Kv, Kv ā‰„ 0 for all v. By the spectral theorem for compact self-adjoint operators, K āˆ— K has an orthonormal basis of eigenvectors. If we remove from that basis eigenvectors with zero eigenvalue, we obtain a basis (en )N n=1 for the subspace S = Ker(K āˆ— K)āŠ„ . In particular, K āˆ— Ken = Ī»n en , with Ī»n > 0 and Ī»n ā†’ 0 if N = āˆž. āˆš Denote Ī¼n = Ī»n and fn = Ī¼āˆ’1 n Ken . Then āˆ’1 āˆ’1 āˆ’1 āˆ— fm , fn  = Ī¼āˆ’1 m Ī¼n Kem , Ken  = Ī¼m Ī¼n em , K Ken  = Ī“mn ,

so (fn )N n=1 is an orthonormal family. By Lemma 5.24, a bounded operator B is deļ¬ned by N  Bv = Ī¼n en , vfn . n=1

Note that Ken = Ī¼n fn = Ben for each n, so Kv = Bv for all v āˆˆ S = āŠ„ āˆ— span{en }N n=1 . If v āˆˆ S , then v āˆˆ Ker(K K), so Kv2 = v, K āˆ— Kv = 0, which implies Kv = 0. Thus, Kv = 0 = Bv for v āˆˆ S āŠ„ . Since K and B  agree on S and S āŠ„ , they are equal. Theorem 5.25. The set of compact operators in L(H) is the closure of the set of ļ¬nite rank operators.

5.3. Spectral measures

139

Proof. If K is compact, it can be approximated in operator norm by ļ¬niterank operators by (5.6). Conversely, if Fn are ļ¬nite-rank they are compact by Corollary 4.46, and if Fn ā†’ K, then K is compact by Proposition 4.47. 

5.3. Spectral measures In this section, we introduce the notion of spectral measure corresponding to a vector. We denote N0 = {n āˆˆ Z | n ā‰„ 0}. Theorem 5.26. Let A be a bounded self-adjoint operator on H, and let Ļˆ āˆˆ H. There exists a unique compactly supported Borel measure Ī¼A,Ļˆ on R such that for all k āˆˆ N0 ,  (5.7) Ļˆ, Ak Ļˆ = xk dĪ¼A,Ļˆ (x). Moreover, supp Ī¼A,Ļˆ āŠ‚ Ļƒ(A). Deļ¬nition 5.27. The measure Ī¼A,Ļˆ which obeys (5.7) for all k āˆˆ N0 is called the spectral measure for the vector Ļˆ and the operator A. The spectral measure will often be denoted more concisely by Ī¼Ļˆ . The reader should be warned that the term ā€œspectral measureā€ will later also be used with related but diļ¬€erent meanings, corresponding to an operator A but perhaps not to any particular vector Ļˆ āˆˆ H.  The integrals ck = xk dĪ¼(x) for k āˆˆ N0 are called the moments of the measure Ī¼. Equation (5.7) precisely speciļ¬es the moments of the desired spectral measure. The reader should keep in mind that not every sequence of real numbers is the sequence of moments of a positive measure on R; for c0 = Ī¼(R) ā‰„ 0, and for every t āˆˆ R, c2 āˆ’ 2tc1 + t2 c0 =  instance, 2 (x āˆ’ t) dĪ¼(x) ā‰„ 0, so c21 ā‰¤ c0 c2 . Investigating this set of constraints more systematically would lead us to the so-called moment problem. Instead, we proceed directly to the proof of Theorem 5.26, which accounts for those constraints somewhat implicitly. In the proof, we will use the left-hand side of (5.7) to deļ¬ne a linear functional on polynomials. We will prove that it has a unique extension to a positive linear functional on C(Ļƒ(A)) and will use the Rieszā€“Markov theorem to obtain the spectral measure. We will use the polynomial spectral mapping theorem from Section 4.4. In a preliminary lemma, we obtain boundedness from positivity, and address a technicality: the distinction between polynomials as algebraic expressions and polynomials as elements of C(Ļƒ(A)) (this distinction is nontrivial if Ļƒ(A) is a ļ¬nite set). We denote by F[x] the set of polynomials in one variable with coeļ¬ƒcients in the ļ¬eld F.

140

5. Bounded self-adjoint operators

Lemma 5.28. Let Ī› : C[x] ā†’ C be a linear map, and let K āŠ‚ R be compact. Assume that for every p āˆˆ R[x] such that p(x) ā‰„ 0 for all x āˆˆ K, Ī›(p) ā‰„ 0. Then the following hold. (a) For all p āˆˆ C[x], |Ī›(p)| ā‰¤ Ī›(1) max|p(x)|. xāˆˆK

(5.8)

(b) If p1 (x) = p2 (x) for all x āˆˆ K, then Ī›(p1 ) = Ī›(p2 ), so Ī› gives a bounded linear functional on the subspace span{xn | n āˆˆ N0 } in C(K). (c) Ī› has a unique extension to a bounded linear functional on C(K), and that extension is a positive linear functional on C(K). Proof. (a) Denote p = maxxāˆˆK |p(x)|. For any p āˆˆ R[x], the polynomials pāˆ“p āˆˆ R[x] are nonnegative on K, so by positivity and linearity, pĪ›(1)āˆ“ Ī›(p) = Ī›(p āˆ“ p) ā‰„ 0. This can be rewritten as āˆ’pĪ›(1) ā‰¤ Ī›(p) ā‰¤ pĪ›(1). In particular, Ī›(p) āˆˆ R. Any p āˆˆ C[x] can be written as a linear combination p = Re p + i Im p by separating real and imaginary parts of coeļ¬ƒcients. Since Ī›(Re p), Ī›(Im p) āˆˆ R, from Ī›(p) = Ī›(Re p) + iĪ›(Im p) it follows that Re Ī›(p) = Ī›(Re p). Thus, |Ī›(p)| = sup Re(eiĻ† Ī›(p)) = sup Re Ī›(eiĻ† p) ā‰¤ sup Ī›(1)eiĻ† p ā‰¤ Ī›(1)p. Ļ†āˆˆR

Ļ†āˆˆR

Ļ†āˆˆR

(b) If p1 = p2 on K, applying (5.8) to p1 āˆ’p2 implies that Ī›(p1 ) = Ī›(p2 ). (c) By Weierstrassā€™s theorem (Corollary 2.20), polynomials form a dense subspace in C(K), so by Proposition 2.44, Ī› extends uniquely to a bounded linear functional on C(K), which we denote by the same letter, Ī› : C(K) ā†’ C. Any f āˆˆ C(K) can be approximated uniformly by a sequence of polynomials pn . If f is real valued, pn can also be chosen with real coeļ¬ƒcients (by taking their real parts); if f is nonnegative, pn can also be chosen nonnegative (by replacing pn with the polynomials pn āˆ’ max(0, minxāˆˆK pn (x))). These nonnegative approximants pn obey Ī›(pn ) ā‰„ 0, so f ā‰„ 0 implies Ī›(f ) = lim Ī›(pn ) ā‰„ 0. nā†’āˆž



Proof of Theorem 5.26. Let us denote K = Ļƒ(A) and, for p āˆˆ C[x], Ī›(p) = Ļˆ, p(A)Ļˆ. This deļ¬nes a linear map Ī› : C[x] ā†’ C. Since A is self-adjoint, p(A)āˆ— = p(A). If p āˆˆ R[x], this implies that p(A) is self-adjoint. If, moreover, p ā‰„ 0 on Ļƒ(A), the polynomial spectral mapping theorem (Theorem 4.33) implies

5.4. Spectral theorem on a cyclic subspace

141

Ļƒ(p(A)) āŠ‚ [0, āˆž), so p(A) is a positive operator and Ī›(p) = Ļˆ, p(A)Ļˆ ā‰„ 0. Thus, by Lemma 5.28, Ī› extends uniquely to a bounded, positive linear functional on C(K). Existence of the spectral measure follows from the Rieszā€“Markov theorem (Theorem 1.100). It remains If Ī¼1 , Ī¼2 are compactly supported mea k  kto prove uniqueness. sures and x dĪ¼1 (x) = x dĪ¼2 (x) for all k āˆˆ N0 , the corresponding bounded linear functionals are equal on polynomials, so by density of polynomials, they are equal on C(supp Ī¼1 āˆŖsupp Ī¼2 ). Thus, by the Rieszā€“Markov  theorem, Ī¼1 = Ī¼2 . Note that Ī¼A,Ļˆ (R) = Ļˆ2 by (5.7) with k = 0. Moreover, the spectral measure of a vector contains a lot of information about the values of the spectral parameter which correspond to v. This is illustrated by the following examples. Recall that Ī“Ī» denotes the Dirac measure at Ī». Example 5.29. Let v be an eigenvector of A, and let Ī» be the corresponding eigenvalue. The spectral measure of v is Ī¼A,v = v2 Ī“Ī» . Proof. From Av = Ī»v, it follows by induction that Ak v = Ī»k v for k āˆˆ N0 . Thus, by linearity, for any polynomial p, p(A)v = p(Ī»)v. Thus, 

v, p(A)v = v, p(Ī»)v = p(Ī»)v2 .

This is equal to p dĪ¼ for the choice of measure Ī¼ = v2 Ī“Ī» , so by uniqueness  of the spectral measure, Ī¼A,v = v2 Ī“Ī» . Generalizations of this example are considered in Exercises 5.9 and 5.10. We give a diļ¬€erent example: Example 5.30. Let A be the self-adjoint operator from Example 5.6 and let f āˆˆ L2 ([0, 1], dx). Then dĪ¼A,f (x) = |f (x)|2 dx. Proof. For any k āˆˆ N0 , (Ak f )(x) = xk f (x), so  1  1 f (x)xk f (x) dx = xk |f (x)|2 dx. f, Ak f  = 0

0

From this, we directly read oļ¬€ dĪ¼A,f (x) = |f (x)|2 dx.



5.4. Spectral theorem on a cyclic subspace In this section, we show that the spectral measure of the vector Ļˆ describes the behavior of the operator on a certain subspace generated by Ļˆ. Deļ¬nition 5.31. For a bounded self-adjoint operator A and Ļˆ āˆˆ H, we deļ¬ne the cyclic subspace of Ļˆ as CA (Ļˆ) = {p(A)Ļˆ | p āˆˆ C[x]}. The vector Ļˆ is said to be cyclic if CA (Ļˆ) = H.

142

5. Bounded self-adjoint operators

Lemma 5.32. The cyclic subspace CA (Ļˆ) is the smallest invariant closed subspace for A that contains Ļˆ. Proof. For any polynomial p, Ap(A) is also a polynomial in A, so the subspace {p(A)Ļˆ | p āˆˆ C[x]} is invariant for A. Thus, its closure CA (Ļˆ) is also invariant for A, by Lemma 4.41. Of course, Ļˆ = A0 Ļˆ āˆˆ CA (Ļˆ). Let M be an invariant subspace for A. If Ļˆ āˆˆ M , it follows by induction that Ak Ļˆ āˆˆ M for all k = 0, 1, 2, . . . , and then that p(A)Ļˆ āˆˆ M for all  polynomials p. Since M is closed, it follows that CA (Ļˆ) āŠ‚ M . We will describe the behavior of A on the cyclic subspace CA (Ļˆ) by using L2 (R, dĪ¼A,Ļˆ ) as a model space. More precisely, we ļ¬nd a unitary map between these spaces which encodes A by multiplication by the function x: Theorem 5.33. Let A be a bounded self-adjoint operator, and let Ī¼A,Ļˆ be the spectral measure of a vector Ļˆ āˆˆ H. Then the map U : C[x] ā†’ CA (Ļˆ) deļ¬ned by U p = p(A)Ļˆ extends uniquely to a unitary map U : L2 (R, dĪ¼A,Ļˆ ) ā†’ CA (Ļˆ) such that U 1 = Ļˆ and, for all f āˆˆ L2 (R, dĪ¼A,Ļˆ ), (U āˆ’1 AU f )(x) = xf (x).

(5.9)

Proof. Let us write Ī¼ = Ī¼A,Ļˆ . For all k, l āˆˆ N0 ,   k l k+l k+l A Ļˆ, A Ļˆ = Ļˆ, A Ļˆ = x dĪ¼(x) = xk xl dĪ¼(x). By sesquilinearity, for all polynomials p, q,  p(A)Ļˆ, q(A)Ļˆ = p(x)q(x) dĪ¼(x). In particular, U pH = p(A)ĻˆH = pL2 (R,dĪ¼) . Since polynomials are dense in L2 (R, dĪ¼) and U is norm-preserving on polynomials, U extends uniquely to a norm-preserving map, denoted by the same letter U , from L2 (R, dĪ¼) to H. The range of U is the closure of {p(A)Ļˆ | p āˆˆ C[x]}, which is precisely CA (Ļˆ). To prove (5.9), consider ļ¬rst that for any polynomial p, AU p = Ap(A)Ļˆ = U (xp(x)). Thus, (5.9) holds for all polynomials, so by density and continuity it holds  for all f āˆˆ L2 (R, dĪ¼). If the operator A has a cyclic vector Ļˆ, Theorem 5.33 immediately implies that A is unitarily conjugated into the form of a multiplication operator:

5.5. Multiplication operators

143

Theorem 5.34 (Spectral theorem for bounded self-adjoint operators with a cyclic vector). If the operator A has a cyclic vector Ļˆ, there is a unitary U : L2 (R, dĪ¼A,Ļˆ ) ā†’ H such that U 1 = Ļˆ and for all f āˆˆ L2 (R, dĪ¼A,Ļˆ ), (U āˆ’1 AU f )(x) = xf (x).

(5.10)

This result is already suļ¬ƒcient for some classes of self-adjoint operators, such as half-line Jacobi matrices (see Chapter 10). However, we will also prove a more general version of the spectral theorem, applicable for every self-adjoint operator on a separable Hilbert space. Exercise 5.16 considers an alternative description of the cyclic subspace of Ļˆ. When we discuss unbounded operators in Chapter 8, that alternative description will be used as the deļ¬nition.

5.5. Multiplication operators By Theorem 5.34, every self-adjoint operator with a cyclic vector is unitarily equivalent to a multiplication operator. Due to this, multiplication operators serve as models for self-adjoint operators, and it is worthwhile to study them systematically. Just as for diagonal matrices, many properties of multiplication operators can be explicitly computed and characterized. Although Theorem 5.34 involves a ļ¬nite spectral measure supported on a compact subset of R, with almost no additional eļ¬€ort, we will allow Baire measures on R, i.e., positive Borel measures on R which are ļ¬nite on compacts. We recall that these are precisely the Lebesgueā€“Stieltjes measures on R (see Chapter 1). Deļ¬nition 5.35. Let Ī¼ be a Baire measure on R. Let g āˆˆ Lāˆž (R, dĪ¼). The multiplication operator Tg,dĪ¼ on L2 (R, dĪ¼) is deļ¬ned by (Tg,dĪ¼ f )(x) = g(x)f (x). We will also use the notation Tg when the measure is clear from context. Conversely, we will write Tg(x),dĪ¼(x) when that is needed; for instance, the conclusion (5.10) can now be written more concisely as U āˆ’1 AU = Tx,dĪ¼A,Ļˆ (x) . Lemma 5.36. If Ī¼ is a Baire measure on R and g āˆˆ Lāˆž (R, dĪ¼), then Tg is a bounded linear operator and Tg  = gāˆž .   Proof. For any f āˆˆ L2 (R, dĪ¼), |gf |2 dĪ¼ ā‰¤ g2āˆž |f |2 dĪ¼, so Tg is bounded and Tg  ā‰¤ gāˆž . Conversely, for any C < gāˆž , the set A = {x | |g(x)| > C} has Ī¼(A) > 0. Since Ī¼ is a Baire measure, for some k āˆˆ N, f = Ļ‡Aāˆ©[āˆ’k,k]   is a nonzero element of L2 (R, dĪ¼) and |gf |2 dĪ¼ ā‰„ C 2 |f |2 dĪ¼. It follows  that Tg  ā‰„ C for any C < gāˆž , so Tg  ā‰„ gāˆž . Algebraic properties of multiplication operators are veriļ¬ed trivially:

144

5. Bounded self-adjoint operators

Lemma 5.37. The map g ā†’ Tg,dĪ¼ is a homomorphism of C āˆ— -algebras Lāˆž (R, dĪ¼) ā†’ L(L2 (R, dĪ¼)), i.e., for any g, h āˆˆ Lāˆž (R, dĪ¼) and Ī» āˆˆ C, TĪ»g = Ī»Tg ,

Tg+h = Tg + Th ,

Tgh = Tg Th ,

Tgāˆ— = TgĀÆ,

T1 = I.

Lemma 5.36 tells us, in particular, that Tg = 0 if and only if g = 0 Ī¼-a.e. Combining this principle with algebraic properties leads to further criteria: Lemma 5.38. Tg is self-adjoint if and only if g is real-valued Ī¼-a.e. Proof. Tg is self-adjoint if and only if Tg āˆ’ Tgāˆ—  = 0. By the calculation Tg āˆ’ Tgāˆ— = Tg āˆ’ TgĀÆ = Tgāˆ’ĀÆg , this is equivalent to g āˆ’ gĀÆāˆž = 0.



Invertibility of Tg āˆ’ z is related to division by g āˆ’ z, which leads us to consider whether |g āˆ’ z| has a lower bound that holds Ī¼-a.e. This leads to the notion of essential range of g with respect to Ī¼, which is deļ¬ned as 5   6 RanĪ¼ g = z āˆˆ C | Ī¼ {x | |g(x) āˆ’ z| < } > 0 for all > 0 . Properties of the essential range are collected in the following lemma. Lemma 5.39. For any Baire measure Ī¼ on R and Borel function g: (a) RanĪ¼ g is the support of the pushforward of Ī¼ by g; (b) RanĪ¼ g is the smallest closed set E āŠ‚ C such that g āˆˆ E holds Ī¼-a.e.; (c) RanĪ¼ g is bounded if and only if g āˆˆ Lāˆž (R, dĪ¼), and in that case, max{|z| | z āˆˆ RanĪ¼ g} = gāˆž ; (d) If g is continuous, then RanĪ¼ g = {g(x) | x āˆˆ supp Ī¼}. If, in addition, Ī¼ is compactly supported, then RanĪ¼ g = {g(x) | x āˆˆ supp Ī¼}. Proof. (a) Denote by Ī½ the pushforward of Ī¼ by g (Lemma 1.55). By deļ¬nition, z āˆˆ RanĪ¼ g if and only if Ī¼(g āˆ’1 (D (z)) > 0 for all > 0. Equivalently, z āˆˆ RanĪ¼ g if and only if Ī¼(g āˆ’1 (U )) > 0 for all open sets U which contain z, i.e., z āˆˆ RanĪ¼ g if and only if z āˆˆ supp Ī½. (b) This follows from (a) by Lemma 1.42. (c) If g āˆˆ Lāˆž (R, dĪ¼), then |g| ā‰¤ gāˆž holds Ī¼-a.e., so by (b), RanĪ¼ g āŠ‚ {z āˆˆ C | |z| ā‰¤ gāˆž }. This shows M ā‰¤ gāˆž where M = max{|z| | z āˆˆ RanĪ¼ g}. Conversely, since g āˆˆ RanĪ¼ g holds Ī¼-a.e., then |g| ā‰¤ M holds Ī¼-a.e., so gāˆž ā‰¤ M . (d) Denote A = {g(x) | x āˆˆ supp Ī¼}. Since A is closed and g āˆˆ A Ī¼-a.e., RanĪ¼ g āŠ‚ A. Conversely, for any open U āŠ‚ C which intersects A, g āˆ’1 (U ) is open and intersects supp Ī¼ so Ī¼(g āˆ’1 (U )) > 0. Thus, A āŠ‚ RanĪ¼ g. Since

5.5. Multiplication operators

145

A āŠ‚ RanĪ¼ g āŠ‚ A and RanĪ¼ g is closed, RanĪ¼ g = A. If Ī¼ is compactly supported, A is compact as the continuous image of a compact set.  We can now describe the spectrum and resolvent of a multiplication operator. / Ļƒ(Tg,dĪ¼ ), the resolvent Proposition 5.40. Ļƒ(Tg,dĪ¼ ) = RanĪ¼ g, and for z āˆˆ of Tg,dĪ¼ at z is (Tg,dĪ¼ āˆ’ z)āˆ’1 = T1/(gāˆ’z),dĪ¼ . 1 āˆˆ Lāˆž (R, dĪ¼), the operator T1/(gāˆ’z),dĪ¼ is bounded and inverse Proof. If gāˆ’z 1 āˆˆ / Lāˆž (R, dĪ¼), then for every > 0, to Tg,dĪ¼ āˆ’ z = Tgāˆ’z,dĪ¼ . Conversely, if gāˆ’z the set A = {x | |g(x) āˆ’ z| < } has Ī¼(A) > 0. Since Ī¼ is a Baire measure, there exists k āˆˆ N such that f = Ļ‡Aāˆ©[āˆ’k,k] is a nonzero element of L2 (R, dĪ¼) and  

|(g āˆ’ z)f |2 dĪ¼ ā‰¤ 2

|f |2 dĪ¼.

Thus, Tg āˆ’ z cannot have a bounded inverse.



We single out an important special case in which we can very explicitly describe the spectrum and compute the norm of the resolvents; this will be used in the next section to compute the norm of the resolvent for an arbitrary self-adjoint operator! Corollary 5.41. For any Baire measure Ī¼ on R, Ļƒ(Tx,dĪ¼(x) ) = supp Ī¼ and for z āˆˆ C \ supp Ī¼, % % % 1 % 1 āˆ’1 % % . = (Tx,dĪ¼(x) āˆ’ z)  = % x āˆ’ z %Lāˆž (dĪ¼) dist(z, supp Ī¼) Proof. As in the previous proof, the resolvent is equal to T(xāˆ’z)āˆ’1 ,dĪ¼(x) and exists precisely when (x āˆ’ z)āˆ’1 āˆˆ Lāˆž (R, dĪ¼). Moreover, by pushforwards, the function g(x) = (x āˆ’ z)āˆ’1 has gāˆž = 1/ dist(z, supp Ī¼).  The following proposition gives criteria for a sequence of multiplication operators to converge in norm or in the sense of strong operator convergence. Proposition 5.42. Let Ī¼ be a Baire measure on R. Consider functions gn āˆˆ Lāˆž (R, dĪ¼), n āˆˆ N āˆŖ {āˆž}. (a) If gn ā†’ gāˆž in Lāˆž (R, dĪ¼), then Tgn ā†’ Tgāˆž . (b) If gn are uniformly bounded, i.e., supgn āˆž < āˆž,

(5.11)

nāˆˆN

s

and limnā†’āˆž gn (x) = gāˆž (x) for Ī¼-a.e. x, then Tgn ā†’ Tgāˆž .

146

5. Bounded self-adjoint operators

Proof. (a) This is immediate from Tgn āˆ’ Tgāˆž  = gn āˆ’ gāˆž āˆž . (b) Denote by C the supremum in (5.11). For all f āˆˆ L2 (R, dĪ¼), by dominated convergence with dominating function 4C 2 |f |2 ,  2  lim Tgn f āˆ’ Tgāˆž f  = lim |gn f āˆ’ gāˆž f |2 dĪ¼ = 0. nā†’āˆž

nā†’āˆž

5.6. Spectral theorem on the entire Hilbert space In Theorem 5.33, we described the action of A on a cyclic subspace deļ¬ned by some vector Ļˆ āˆˆ H. In this section, we extend that to describe the action of A on the entire Hilbert space. We prove that any bounded self-adjoint operator is unitarily equivalent to a direct sum of multiplication operators, i.e., that there exists a sequence (Ī¼n )N n=1 of probability measures such that Aāˆ¼ =

N 3

Tx,dĪ¼n (x) .

n=1

4N Recall that our notation for direct sums n=1 allows N ļ¬nite or āˆž, and N = āˆž denotes a countable sum. The key ingredient in the proof is a decomposition of the Hilbert space as a direct sum of cyclic subspaces. Deļ¬nition 5.43. Let H be a separable Hilbert space and A a self-adjoint operator. A spectral basis for A is a (ļ¬nite or inļ¬nite) sequence (Ļˆn )N n=1 of normalized vectors such that H=

N 3

CA (Ļˆn ).

(5.12)

n=1

Example 5.44. If A is a compact self-adjoint operator, any orthonormal basis of eigenvectors of A is a spectral basis for A. Lemma 5.45. Any self-adjoint operator on a separable Hilbert space has a spectral basis. This sequence can be chosen to begin with an arbitrary normalized vector Ļˆ1 . To motivate the proof, recall that any cyclic subspace CA (Ļˆ) is invariant for A and that CA (Ļˆ)āŠ„ is an invariant subspace for Aāˆ— = A. This suggests an appealing, naive approach: start with some Ļˆ1 , then pick arbitrary Ļˆ2 āˆˆ CA (Ļˆ1 )āŠ„ , Ļˆ3 āˆˆ (CA (Ļˆ1 ) āŠ• CA (Ļˆ2 ))āŠ„ , and so on. However, the sequence chosen in this way may not cover the entire space H. To ensure that it does, we modify the construction so that Ļ†n āˆˆ CA (Ļˆ1 ) āŠ• Ā· Ā· Ā· āŠ• CA (Ļˆn ) for all n, where (Ļ†n )N n=1 is a ļ¬xed orthonormal basis of H. Proof. Since H is separable, it has a ļ¬nite or countable orthonormal basis N (Ļ†n )N n=1 . We will deļ¬ne (Ļˆn )n=1 inductively: let Ļˆ1 = Ļ†1 and let Ļˆn be the

5.6. Spectral theorem on the entire Hilbert space

147

orthogonal projection of Ļ†n onto Vnāˆ’1 =

nāˆ’1

CA (Ļˆk )āŠ„ .

k=1

)āŠ„

Since Ļˆn āˆˆ CA (Ļˆk for k < n, Lemma 5.32 implies that CA (Ļˆn ) āŠ‚ CA (Ļˆk )āŠ„ for k < n. Thus, the cyclic subspaces CA (Ļˆn ) are mutually orthogonal. Now the construction implies that āŠ„ = Ļˆn āˆ’ Ļ†n āˆˆ Vnāˆ’1

nāˆ’1 3

CA (Ļˆk ),

k=1

so, by induction in n, Ļ†n āˆˆ

n 3

CA (Ļˆk ).

k=1

Since (Ļ†n )N n=1 is an orthonormal basis of H, it follows that H=

N 3

CA (Ļˆn ).

n=1

The proof is completed by discarding from the sequence all vectors Ļˆn which are equal to 0 (relabeling the sequence and changing the value of N in the process) and normalizing all the nonzero vectors.  Of course, the spectral basis is not unique: Diļ¬€erent spectral bases of the same operator may not even be of the same cardinality (Exercise 5.18). Now that we know that every self-adjoint operator has a spectral basis, we can prove the spectral theorem. Theorem 5.46 (Spectral theorem for self-adjont operators). Let A be a bounded self-adjoint operator on a separable Hilbert space H, and let (Ļˆn )N n=1 be a spectral basis for A. Denoting Ī¼n = Ī¼A,Ļˆn , there exists a unitary map U:

N 3

L2 (R, dĪ¼n ) ā†’ H

(5.13)

n=1

such that U āˆ’1 AU =

N 3

Tx,dĪ¼n (x) .

(5.14)

n=1

Proof. Applying Theorem 5.33 to each Ļˆn , there exist measures Ī¼n and unitaries Un : L2 (R, dĪ¼n ) ā†’ CA (Ļˆn ) such that Unāˆ’1 AUn = Tx,dĪ¼n (x) for all n. Since A is the direct sum of its 4  restrictions to CA (Ļˆn ), (5.14) holds for the unitary U = N n=1 Un .

148

5. Bounded self-adjoint operators

The spectral representations (5.13) and (5.14) can be used to study the spectrum and resolvent of A. This provides a signiļ¬cant improvement over (5.1) and the ļ¬rst application of the spectral theorem: where before we were only able to give an upper bound on the norm of the resolvent for nonreal z, the spectral theorem allows us to compute the norm for any z āˆˆ C \ Ļƒ(A). Proposition 5.47. If A has the spectral representations (5.13) and (5.14), then Ļƒ(A) =

N 

supp Ī¼n .

n=1

Moreover, for all z āˆˆ C \ Ļƒ(A), U āˆ’1 (A āˆ’ z)āˆ’1 U =

N 3

T(xāˆ’z)āˆ’1 ,dĪ¼n (x)

n=1

and (A āˆ’ z)āˆ’1  =

1 . dist(z, Ļƒ(A))

(5.15)

Proof. The proof is based on the observation that conjugation by a unitary map does not aļ¬€ect invertibility or the norm of the inverse. Thus, it suļ¬ƒces to study the direct sum of multiplication operators. / supp Ī¼n and, in that By Corollary 5.41, Tx,dĪ¼n (x) āˆ’ z is invertible if z āˆˆ case, (Tx,dĪ¼n (x) āˆ’ z)āˆ’1  =

1 . dist(z, supp Ī¼n )

It follows that sup(Tx,dĪ¼n (x) āˆ’ z)āˆ’1  = sup n

n

1 1   , = N dist(z, supp Ī¼n ) dist z, n=1 supp Ī¼n

and the proof is then completed by Proposition 4.39.



Corollary 5.48. For any z āˆˆ C, dist(z, Ļƒ(A)) = inf (A āˆ’ z)u. uāˆˆH u=1

(5.16)

Proof. For z āˆˆ Ļƒ(A), this is precisely Weylā€™s criterion. For z āˆˆ / Ļƒ(A), it follows from (5.15). 

5.7. Borel functional calculus

149

5.7. Borel functional calculus In this section, we present a consistent way of deļ¬ning operators g(A), where A āˆˆ L(H) is self-adjoint and g : Ļƒ(A) ā†’ C is a bounded Borel function. This will vastly generalize the deļ¬nition of p(A) for polynomials p. We denote by Bb (Ļƒ(A)) the set of bounded Borel functions on Ļƒ(A), previously discussed in Section 1.7. This is a C āˆ— -algebra, with addition and multiplication deļ¬ned pointwise, with complex conjugation in the role of taking adjoints, and with the supremum norm. The Borel functional calculus will be a homomorphism of C āˆ— -algebras which preserves a certain kind of convergence of sequences. Theorem 5.49 (Borel functional calculus). For any bounded self-adjoint operator A on H, there is a unique map Ī¦A : Bb (Ļƒ(A)) ā†’ L(H) such that the following hold. (a) Ī¦A is an algebraic homomorphism, i.e., it is linear, preserves mulg ) = Ī¦A (g)āˆ— . tiplication, Ī¦A (1) = I, and Ī¦A (ĀÆ (b) If g is the identity map g(x) = x, then Ī¦A (g) = A. (c) If gk ā†’ gāˆž pointwise and sup sup |gk (x)| < āˆž, kāˆˆN xāˆˆĻƒ(A) s

then Ī¦A (gk ) ā†’ Ī¦A (gāˆž ). Existence is shown by the following construction: Lemma 5.50. Let A be a self-adjoint operator with a spectral representations (5.13) and (5.14). If the operator Ī¦(g) is deļ¬ned, for g āˆˆ Bb (Ļƒ(A)), by U āˆ’1 Ī¦(g)U =

N 3

Tg(x),dĪ¼n (x) ,

n=1

then Ī¦ has the properties (a), (b), and (c) from Theorem 5.49. Moreover, for any g āˆˆ Bb (Ļƒ(A)), Ī¦(g) = supgLāˆž (dĪ¼n ) .

(5.17)

n

Proof. Property (a) of Theorem 5.49 follows from properties of multiplication operators, and (b) follows from the spectral representation since 4N āˆ’1 AU . n=1 Tx,dĪ¼n (x) = U

150

5. Bounded self-adjoint operators

s

By Proposition 5.42, Tgk ,dĪ¼n ā†’ Tgāˆž ,dĪ¼n for each n. Since gk are uniformly bounded on Ļƒ(A) āŠƒ supp Ī¼n , by Proposition 4.43, N 3

s

Tgk ,dĪ¼n ā†’

n=1

N 3

Tgāˆž ,dĪ¼n ,

n=1 s

and conjugating by U gives Ī¦(gk ) ā†’ Ī¦(gāˆž ), which proves (iii). Finally, (5.17) follows from Ī¦(g) = supn Tg(x),dĪ¼n (x) .



The proof of uniqueness is based on a criterion for a subalgebra of Bb (X) to be equal to Bb (X) (Proposition 1.92), reļ¬ned to a compact X āŠ‚ R as follows: Proposition 5.51. Let X āŠ‚ R be compact. Let M be a subalgebra of Bb (X), i.e., closed under addition, scalar multiplication, multiplication, and 1 āˆˆ M. If M is closed under pointwise convergence of uniformly bounded sequences, the following are equivalent: (a) The function g(x) = x is in M. (b) C(X) āŠ‚ M. (c) Ļ‡B āˆˆ M for all Borel sets B. (d) M = Bb (X). Proof. (a) =ā‡’ (b): Since 1, x āˆˆ M and M is a subalgebra, M contains all polynomials. Note that M is also closed under uniform convergence; thus, by density of polynomials in C(X) (Weierstrassā€™s theorem), C(X) āŠ‚ M. (b) =ā‡’ (c) and (c) =ā‡’ (d): These follow from Proposition 1.92. (d) =ā‡’ (a): This is trivial.



Proof of Theorem 5.49. Existence was proved in Lemma 5.50. Assume that Ī¦1 , Ī¦2 obey the properties (a), (b), and (c) of Theorem 5.49 and denote M = {g āˆˆ Bb (Ļƒ(A)) | Ī¦1 (g) = Ī¦2 (g)}. This is a subalgebra of Bb (Ļƒ(A)) which contains the identity map g(x) = x and is closed under pointwise convergence of uniformly bounded sequences,  so M = Bb (Ļƒ(A)). Due to the uniqueness of the Borel functional calculus, we will write g(A) instead of Ī¦A (g) from now on. Uniqueness tells us, in particular, that the construction in Lemma 5.50 gives the same operators, regardless of the spectral representation, so we can choose a spectral representation which is convenient for a given argument. This trick will be used below. We used spectral measures to construct the functional calculus, but functional calculus can also be used to express spectral measures:

5.7. Borel functional calculus

151

Corollary 5.52. Let A be a self-adjoint operator on H. Let g āˆˆ Bb (Ļƒ(A)) and Ļˆ āˆˆ H. Then g(A)Ļˆ āˆˆ CA (Ļˆ) and  Ļˆ, g(A)Ļˆ = g(x)dĪ¼A,Ļˆ (x). Proof. Consider the spectral representation with respect to a spectral basis (Ļˆn )N n=1 which has Ļˆ as the ļ¬rst basis vector, i.e., Ļˆ1 = Ļˆ. By the construction of the unitary map U , the map U āˆ’1 maps CA (Ļˆ) to the set of vectors āˆ’1 Ļˆ = (f )N (Fn )N n n=1 with n=1 with Fn = 0 for all n = 1. In particular, U āˆ’1 N f1 = 1 and fn = 0 for n = 1. Thus, U g(A)Ļˆ = (gfn )n=1 , and gfn = 0 for n = 1 precisely means that g(A)Ļˆ āˆˆ CA (Ļˆ). Moreover, / 0 N 3 āˆ’1 āˆ’1 Ļˆ, g(A)Ļˆ = U Ļˆ, Tg(x),dĪ¼n (x) U Ļˆ n=1

= 1, Tg(x),dĪ¼1 (x) 1  = g(x) dĪ¼1 (x). This completes the proof, since Ī¼1 is the spectral measure for Ļˆ1 = Ļˆ.



Many further identities follow from properties of the functional calculus. For instance, the functional calculus includes resolvents in a natural way: Lemma 5.53. For any z āˆˆ C \ Ļƒ(A), the function g(x) = (x āˆ’ z)āˆ’1 is contained in Bb (Ļƒ(A)) and g(A) = (A āˆ’ z)āˆ’1 . Proof. Since |g(x)| ā‰¤ 1/ dist(z, Ļƒ(A)) for all x āˆˆ Ļƒ(A), g is bounded on Ļƒ(A). Since g(x)(x āˆ’ z) = (x āˆ’ z)g(x) = 1 for all x āˆˆ Ļƒ(A) and the Borel functional calculus is an algebraic homomorphism, g(A)(A āˆ’ z) = (A āˆ’ z)g(A) = I, 

so g(A) is the resolvent for A at z.

Through the formula (5.17), this gives another way to compute the norm of (A āˆ’ z)āˆ’1 as (5.15). Moreover: Corollary 5.54. For any z āˆˆ C \ Ļƒ(A) and Ļˆ āˆˆ H,  1 dĪ¼A,Ļˆ (x). Ļˆ, (A āˆ’ z)āˆ’1 Ļˆ = xāˆ’z

(5.18)

Proof. This follows from Corollary 5.52 applied to g(A) = (A āˆ’ z)āˆ’1 .



The connection (5.18) between resolvents of A and spectral measures has a central place in spectral theory, as will be seen in later chapters.

152

5. Bounded self-adjoint operators

Since the functional calculus is explicitly constructed in terms of multiplication operators, further properties are straightforward to derive; see, e.g., the spectral mapping theorem for continuous functions (Exercise 5.21), which generalizes that for polynomials. We discussed uniqueness of the entire Borel functional calculus, but sometimes a single function of A can be uniquely characterizaed by a natural āˆš set of properties. For A ā‰„ 0, the function g(x) = x is deļ¬ned on Ļƒ(A), so āˆš g(A) = A is well deļ¬ned by the Borel functional calculus. The square root lemma (Exercise 5.24) gives a set of properties which describe it uniquely. The Borel functional calculus can be used to solve the initial value problem for a function Ļˆ : R ā†’ H given by iāˆ‚t Ļˆ(t) = AĻˆ(t),

Ļˆ(0) = Ļˆ0 .

(5.19)

This has the physical interpretation as a time-independent SchrĀØodinger equation. Formally, it resembles the scalar initial value problem if  = Ī»f , f (0) = f0 , which has the solution f (t) = eāˆ’iĪ»t f0 . This motivates: Lemma 5.55. If A is a bounded self-adjoint operator on H, then U (t) = eāˆ’itA are unitary operators for all t āˆˆ R. They obey U (t + s) = U (t)U (s) for all t, s āˆˆ R and U (0) = I. As a function of t, U (t) is norm-diļ¬€erentiable and iU  = AU in the sense that, taking limits in L(H), U (s) āˆ’ U (t) = AU (t). (5.20) sā†’t sāˆ’t In particular, for any Ļˆ0 āˆˆ H, the family Ļˆ(t) = U (t)Ļˆ0 solves (5.19). i lim

Proof. From eāˆ’itx = 1/eāˆ’itx , we conclude U (t)āˆ— = U (t)āˆ’1 . Other properties follow from eāˆ’i(t+s)x = eāˆ’itx eāˆ’isx and eāˆ’i0x = 1. To prove diļ¬€erentiability, ļ¬x t and denote  x eāˆ’isx āˆ’ eāˆ’itx āˆ’ xeāˆ’itx = (eāˆ’iux āˆ’ eāˆ’itx ) du. f (s, x) = i sāˆ’t |s āˆ’ t| [t,s] By using a Lipschitz estimate |eāˆ’iux āˆ’ eāˆ’itx | ā‰¤ |iux āˆ’ itx| and integrating,  |(s āˆ’ t)x2 | |x| , |iux āˆ’ itx| du ā‰¤ |f (s, x)| ā‰¤ |s āˆ’ t| [t,s] 2 so by the functional calculus, % % % A2 |s āˆ’ t| % U (s) āˆ’ U (t) %ā‰¤ %i āˆ’ AU (t) , % % sāˆ’t 2 which implies the norm convergence (5.20).



5.8. Spectral theorem for unitary operators

153

5.8. Spectral theorem for unitary operators So far, unitary maps have appeared mostly as a way to communicate the equivalence of certain objects. In this section, we change the perspective and consider unitary operators W āˆˆ L(H) in their own right. We will describe elements of their spectral theory, with close parallels to self-adjoint operators. Some steps will be left as exercises. By Lemma 4.5, W āˆˆ L(H) is unitary if and only if W W āˆ— = W āˆ— W = I. In other words, W is unitary if and only if it is invertible and W āˆ’1 = W āˆ— . This can be compared with self-adjoint operators, which obey A = Aāˆ— . In order to obtain a version of the spectral theorem, we will need a decomposition of H as a direct sum of cyclic subspaces; for this, we need a notion of cyclic subspace such that both the subspace and its orthogonal complement are invariant for W . Equivalently, the cyclic subspace should be invariant for W and W āˆ— = W āˆ’1 , so we allow negative powers of W : Deļ¬nition 5.56. Let W āˆˆ L(H) be a unitary operator. The cyclic subspace generated by Ļˆ āˆˆ H is CW (Ļˆ) = span{W k Ļˆ | k āˆˆ Z}. Another diļ¬€erence from the self-adjoint case is that the role of R is replaced by the unit circle āˆ‚D = {z āˆˆ C | |z| = 1}. It is notationally convenient to parametrize z = eiĪø . The following theorem examines the notion of spectral measure; here, too, it is natural to include negative powers because trigonometric polynomials are dense in C(āˆ‚D) (Corollary 2.21): Theorem 5.57. Let W be a unitary operator on H and let Ļˆ āˆˆ H. Then there exists a unique positive Borel measure Ī¼ on āˆ‚D such that  k eikĪø dĪ¼(Īø), āˆ€k āˆˆ Z. (5.21) Ļˆ, W Ļˆ = āˆ‚D

The measure satisļ¬es Ī¼(āˆ‚D) = Ļˆ2 . Moreover, there exists a unitary operator U : L2 (āˆ‚D, dĪ¼W,Ļˆ ) ā†’ CW (Ļˆ) such that for all f āˆˆ L2 (āˆ‚D, dĪ¼), (U āˆ’1 W U f )(eiĪø ) = eiĪø f (eiĪø ).

(5.22)

Deļ¬nition 5.58. The measure that obeys (5.21) is called the spectral measure for the vector Ļˆ and operator W and is denoted by Ī¼W,Ļˆ or Ī¼Ļˆ . The proof requires the following lemma. Lemma 5.59 (FejĀ“erā€“Riesz). Let f be a Laurent polynomial, i.e., f (z) = n k k=m ck z with m, n āˆˆ Z. If f (z) ā‰„ 0 for all z āˆˆ āˆ‚D, then there exists a polynomial P such that z ). f (z) = P (z)P (1/ĀÆ

(5.23)

154

5. Bounded self-adjoint operators

Proof. Since f (z) and f (1/ĀÆ z ) are analytic functions of z which coincide on āˆ‚D, they must be equal, i.e., f (z) = f (1/ĀÆ z ).

(5.24)

Writing f in the form f (z) = z m Q(z) with Q a polynomial, we see that f can be decomposed as a product of linear factors, f (z) = az m

K 

(z āˆ’ zk )jk ,

(5.25)

k=1

where the zk are distinct and jk are their multiplicities. Substituting this zk is a zero on both sides of (5.24), we see that for every zero zk of f , 1/ĀÆ of the same multiplicity. Since f (z) has constant sign on āˆ‚D, zeros on āˆ‚D have even multiplicity. Thus, one can take P to be a constant b times the product of (z āˆ’ zk )ik , where āŽ§ āŽŖ |zk | < 1 āŽØ jk , ik = jk /2, |zk | = 1 āŽŖ āŽ© 0, |zk | > 1. For a suitable choice of b, we obtain a polynomial such that (5.23) holds.



Proof of Theorem 5.57. We deļ¬ne a linear functional Ī› on the vector space S of Laurent polynomials by 0  /  n n   ikĪø k ck e ck W Ļˆ . = Ļˆ, Ī› k=m

k=m

n ikĪø ā‰„ 0 for all Īø āˆˆ R, then by Lemma 5.59, k If k=m ck e k=m ck z = z ). Since W āˆ— = W āˆ’1 , this implies nk=m ck W k = P (W )āˆ— P (W ) P (z)P (1/ĀÆ and  n   ikĪø Ī› ck e = Ļˆ, P (W )āˆ— P (W )Ļˆ = P (W )Ļˆ, P (W )Ļˆ ā‰„ 0. n

k=m

Thus, Ī› is a positive linear functional on S. By Weierstrassā€™s second theorem, S is dense in C(āˆ‚D), so as in the proof of Lemma 5.28, Ī› extends to a positive linear functional on C(āˆ‚D). By the Rieszā€“Markov theorem,  there exists a unique positive measure Ī¼ such that Ī›(f ) = f dĪ¼ and, in particular, (5.21) holds. Applying (5.21) with k = 0 implies Ī¼(āˆ‚D) = Ļˆ2 . We deļ¬ne U on monomials by U : z k ā†’ W k Ļˆ and extend to Laurent polynomials by linearity. Viewing S as a subspace of L2 (D, dĪ¼), W is a

5.9. Exercises

155

norm-preserving map from S to CW (Ļˆ) because %  n 0 %2 / n n % %    % % ck eikĪø % = ck W k Ļˆ, ck W k Ļˆ %U % % k=m k=m k=m 0 / n n   āˆ’k k cĀÆk W ck W Ļˆ = Ļˆ,  =Ī›

k=m

k=m

n 

n 

cĀÆk eāˆ’ikĪø

k=m



ck eikĪø

k=m

2    n    ck eikĪø  dĪ¼(Īø). =    k=m

Since S is dense in L2 (āˆ‚D, dĪ¼), this means that U can be uniquely extended to a unitary map of L2 (D, dĪ¼) onto CW (Ļˆ). Finally, it is immediate that (5.22) holds for f (z) = z k , so by linearity, density of Laurent polynomials,  and boundedness of both sides, (5.22) holds for all f āˆˆ L2 (āˆ‚D, dĪ¼). By the same arguments as in the self-adjoint case, for any unitary W , the 4NHilbert space H can be written as a direct sum of cyclic subspaces H = n=1 CW (Ļˆn ). Thus, starting from Theorem 5.57, the following theorem follows by the same arguments as in the self-adjoint case. Theorem 5.60 (Spectral theorem for unitary operators). Let W āˆˆ L(H) be unitary. There exists a sequence of probability measures (dĪ¼n )N n=1 on āˆ‚D (N may be ļ¬nite or inļ¬nite) and a unitary map U:

N 3

L2 (āˆ‚D, dĪ¼n ) ā†’ H

n=1

such that for every f = (fn )N n=1 āˆˆ

4N

n=1 L

2 (āˆ‚D, dĪ¼

(5.26) n ),

(U āˆ’1 W U f )n (eiĪø ) = eiĪø fn (eiĪø ).

(5.27)

A collection of measures Ī¼n together with a unitary map U as in (5.26) and (5.27) is called a spectral representation. Some further consequences are left as exercises. In particular, the Borel functional calculus can be introduced for unitary operators (Exercise 5.27).

5.9. Exercises 5.1. Prove that ā‰¤ is not a total order unless dim H = 1. 5.2. Let P, Q be orthogonal projections on H. Prove that P ā‰¤ Q if and only if Ran P āŠ‚ Ran Q.

156

5. Bounded self-adjoint operators

5.3. If Kn āˆˆ L(H) are positive compact operators, prove that s

s

Kn ā†’ I ā‡ā‡’ Kn1/2 ā†’ I. 1/2

1/2

Hint: Use I āˆ’ Kn = (I + Kn )(I āˆ’ Kn ). 5.4. In a metric space (X, d), denote by F (X) the set of nonempty compact subsets of X. Prove that the Hausdorļ¬€ distance is a metric on F (X). 5.5. Construct a sequence of self-adjoint operators An on L2 ([0, 1], dx) s such that Ļƒ(An ) = [0, 1] for every n and An ā†’ 0 as n ā†’ āˆž. 5.6. If An , n āˆˆ NāˆŖ{āˆž} are bounded self-adjoint operators and An ā†’ Aāˆž , prove that (An āˆ’ z)āˆ’1 ā†’ (Aāˆž āˆ’ z)āˆ’1 for all z āˆˆ C \ R. Hint: Check and use An āˆ’ z = ((An āˆ’ Aāˆž )(Aāˆž āˆ’ z)āˆ’1 + I)(Aāˆž āˆ’ z). 5.7. Let A, B be compact self-adjoint operators on a separable Hilbert space H. If A and B commute, prove that there exists an orthonormal H basis (vn )dim n=1 of H such that every vn is an eigenvector of both A and B. Hint: For every eigenvalue Ī» of A, prove that Ker(A āˆ’ Ī») is an invariant subspace for B. 5.8. An operator K is called normal if KK āˆ— = K āˆ— K. Let K be a compact normal operator. Prove that there exists an orthonormal basis H (vn )dim n=1 of H such that every vn is an eigenvector of K. The corresponding eigenvalues Ī»n can be complex, but if dim H = āˆž, then limnā†’āˆž Ī»n = 0. āˆ— āˆ— Hint: Consider the operators A = K+K , B = Kāˆ’K 2 2i . 5.9. Let v1 , . . . , vn be normalized eigenvectors of a self-adjoint operator A , . . . , Ī»n . Find the corresponding to mutually distinct eigenvalues Ī»1 spectral measure of their linear combination v = nj=1 Īŗj vj . 5.10. Let (vj )āˆž j=1 be normalized eigenvectors of a self-adjoint operator A corresponding to mutually distinct eigenvalues (Ī»j )āˆž j=1 . Find the āˆž 2 < āˆž. Īŗ v , where |Īŗ | spectral measure for v = āˆž j=1 j j j=1 j 5.11. Let Ļˆ be an eigenvector of A. What is the cyclic subspace of Ļˆ? What is its dimension? 5.12. Let A be a Hermitian nƗn matrix with n distinct eigenvalues Ī»1 , . . . , Ī»n . Let v1 , . . . , vn be the corresponding eigenvectors. If Ļˆ = nj=1 Īŗj vj with all Īŗj nonzero, prove that Ļˆ is cyclic. 5.13. If A is a Hermitian n Ɨ n matrix with a cyclic vector, prove that A has n distinct eigenvalues.

5.9. Exercises

157

5.14. Let A be the self-adjoint operator from Examples 5.6 and 5.30. Prove that a vector v āˆˆ L2 ([0, 1], dx) is cyclic if and only if v(x) = 0 for Lebesgue-a.e. x āˆˆ [0, 1]. 5.15. Let A be a self-adjoint operator and let u āˆˆ H. If v āˆˆ CA (u), prove that there exists f āˆˆ L2 (R, dĪ¼) such that dĪ¼A,v = |f |2 dĪ¼A,u . Hint: For f āˆˆ L2 (R, dĪ¼), ļ¬nd the spectral measure of f with respect to Tx,dĪ¼(x) . 5.16. Prove that span{(A āˆ’ z)āˆ’1 Ļˆ | z āˆˆ C \ R} = CA (Ļˆ). Hint: To prove āŠ‚, use Theorem 5.33 to ļ¬nd (A āˆ’ z)āˆ’1 Ļˆ in CA (Ļˆ). To prove āŠƒ, use the Neumann series and extract An Ļˆ as suitable limits as z ā†’ āˆž. 5.17. Prove that the multiplication operator Tg,dĪ¼ is: (a) unitary if and only if |g(x)| = 1 for Ī¼-a.e. x; (b) a projection if and only if g(x) āˆˆ {0, 1} for Ī¼-a.e. x. 5.18. If A is a 2Ɨ2 matrix with distinct eigenvalues Ī»1 , Ī»2 , prove that A has a spectral basis (v1 , v2 ) of cardinality 2 and a spectral basis (v1 + v2 ) of cardinality 1. 5.19. If A is a self-adjoint operator with an eigenvalue/eigenvector pair AĻˆ = Ī»Ļˆ, prove that for all g āˆˆ Bb (Ļƒ(A)), g(A)Ļˆ = g(Ī»)Ļˆ. 5.20. If A is a self-adjoint operator and g āˆˆ Bb (Ļƒ(A)) such that g ā‰„ 0 on Ļƒ(A), prove that g(A) ā‰„ 0. 5.21. Prove the spectral mapping theorem for continuous functions: If A is a bounded self-adjoint operator and f āˆˆ C(Ļƒ(A)), then Ļƒ(f (A)) = {f (Ī») | Ī» āˆˆ Ļƒ(A)}. In particular, for any z āˆˆ C \ Ļƒ(A), ' & 1 | Ī» āˆˆ Ļƒ(A) . Ļƒ((A āˆ’ z)āˆ’1 ) = Ī»āˆ’z 5.22. Let A, B be bounded self-adjoint operators. If AB = BA, prove that for all g āˆˆ Bb (Ļƒ(A)) and h āˆˆ Bb (Ļƒ(B)), g(A)h(B) = h(B)g(A). Hint: Consider the set of g āˆˆ Bb (Ļƒ(A)) for which g(A) commutes with B.

158

5. Bounded self-adjoint operators

5.23. The following measurability statement is useful when considering ergodic families of operators, such as random operators. Consider selfadjoint operators AĻ‰ āˆˆ L(H) parametrized by Ļ‰ āˆˆ Ī© with Ī© a measure space. If M = sup AĻ‰  < āˆž Ļ‰āˆˆĪ©

and the map Ļ‰ ā†’ u, AĻ‰ v is measurable for every u, v āˆˆ H, prove that the map Ļ‰ ā†’ u, h(AĻ‰ )v is measurable for every u, v āˆˆ H and every h āˆˆ Bb ([āˆ’M, M ]). āˆš 5.24. Square root lemma: Let A ā‰„ 0. Prove that A is the only operator B āˆˆ L(H) which obeys B ā‰„ 0 and B 2 = A. 5.25. For any bounded self-adjoint operator A and w āˆˆ C, prove that lim (I + wA/n)n = ewA

nā†’āˆž

with the limit taken in the sense of norm-convergence. 5.26. Let W āˆˆ L(H) be a unitary operator with a spectral representation  (5.26) and (5.27). Prove that Ļƒ(W ) = N n=1 supp Ī¼n . 5.27. Borel functional calculus for unitary operators: If W āˆˆ L(H) is unitary, prove that there is a unique map Ī¦W : Bb (āˆ‚D) ā†’ L(H) such that the following hold. (a) Ī¦W is an algebraic homomorphism, i.e., it is linear, preserves g ) = Ī¦W (g)āˆ— . multiplication, Ī¦W (1) = I, and Ī¦W (ĀÆ (b) If g is the identity map g(z) = z, then Ī¦W (g) = W . (c) If gk ā†’ gāˆž pointwise and sup sup |gk (z)| < āˆž, kāˆˆN zāˆˆāˆ‚D s

then Ī¦W (gk ) ā†’ Ī¦W (gāˆž ).

k 5.28. If W āˆˆ L(H) is unitary, prove that the limit s-limnā†’āˆž n1 nāˆ’1 k=0 W exists and that it is an orthogonal projection in H. Describe its range.

Chapter 6

Measure decompositions

It is clear that there are qualitative diļ¬€erences between, e.g., Lebesgue measure on R and the counting measure on Z, nāˆˆZ Ī“n . In this chapter, we consider several such diļ¬€erences, which lead to measure decompositions with an important role in spectral theory. One of these diļ¬€erences is in how the measure acts on countable sets. To state this, recall (from Deļ¬nition 1.41) that a measure Ī½ is said to be supported on some Borel set S if Ī½(S c ) = 0. Lebesgue measure gives zero measure to singletons {x} and therefore to all countable sets, whereas the counting measure on Z is supported on the countable set Z. We can ļ¬nd the same distinction by comparing the spectral measures in Example 5.6 to those in Example 5.29 and Exercises 5.9 and 5.10. This is considered in Section 6.1. Another decomposition is based on how the measure acts on sets of zero Lebesgue measure; this is considered in Section 6.2. Hausdorļ¬€ measures have an additional parameter Ī± which represents a kind of fractal dimension and allows us to quantify intermediate behaviors between counting measures and Lebesgue measures. Hausdorļ¬€ measures and decompositions based on them are considered in Section 6.3. We will sometimes work in the setting of measures on an abstract metric space X but will often focus on Baire measures on R, i.e., Borel measures on R which are ļ¬nite on compacts. All measures are assumed to be positive unless otherwise stated, but we will sometimes generalize from ļ¬nite positive measures to the following

159

160

6. Measure decompositions

class: let us call Ī½ : BX ā†’ C a complex measure if there exist ļ¬nite positive measures Ī½1 , Ī½2 , Ī½3 , Ī½4 : BX ā†’ [0, āˆž) such that Ī½ = (Ī½1 āˆ’ Ī½2 ) + i(Ī½3 āˆ’ Ī½4 ).

(6.1)

Integration is accordingly deļ¬ned by      h dĪ½ = h dĪ½1 āˆ’ h dĪ½2 + i h dĪ½3 āˆ’ i h dĪ½4  for h āˆˆ 4j=1 L1 (dĪ½j ). It is more common to deļ¬ne complex measures as Ļƒadditive maps BX ā†’ C; that deļ¬nition is equivalent to ours (Exercise 6.1). Moreover, the representation (6.1) naturally arises in spectral theory so we can view Exercise 6.1 as an aside.

6.1. Pure point and continuous measures In this section, we consider the ļ¬rst, and simplest, decomposition of the measure, which is based on how it acts on individual points. In Section 9.3, this decomposition will be linked with the eigenvalues and eigenvectors of a self-adjoint operator. Deļ¬nition 6.1. A measure Ī¼ is said to have a point mass at x if Ī¼({x}) > 0. Deļ¬ne by P the set of point masses of Ī¼. The measure Ī¼ is a pure point measure if Ī¼(P c ) = 0. The measure Ī¼ is a continuous measure if P = āˆ…. Deļ¬nition 6.2. A Borel measure Ī¼  on X is said to be Ļƒ-ļ¬nite if there is a sequence of Fn āˆˆ BX such that X = āˆž n=1 Fn and Ī¼(Fn ) < āˆž for each n. Lemma 6.3. Any Ļƒ-ļ¬nite measure has countably many point masses. Proof. For any m, n āˆˆ N, the set Am,n = {x āˆˆ Fm | Ī¼({x}) ā‰„ 1/n} is ļ¬nite because #Am,n ā‰¤ nĪ¼(Fm ) < āˆž. Every point mass of Ī¼ is in Am,n for some m, n āˆˆ N, and their countable union is countable.  Lemma 6.4. For a Ļƒ-ļ¬nite measure Ī¼, the following are equivalent: (a) Ī¼ is a pure point measure; (b) Ī¼ is supported on some countable set S; (c) Ī¼ = Ī»āˆˆP Ī¼({Ī»})Ī“Ī» , where P is the set of point masses of Ī¼. Proof. (a) =ā‡’ (b): This is immediate from Lemma 6.3. (b) =ā‡’ (c): Let S denote a countable set such that Ī¼(S c ) = 0. For any set B, by countability of B āˆ© S,   Ī¼({Ī»}) = Ī¼({Ī»})Ī“Ī» (B). Ī¼(B) = Ī¼(B āˆ© S) = Ī»āˆˆBāˆ©S

Ī»āˆˆS

6.1. Pure point and continuous measures

161

In particular, Ī¼({x}) > 0 implies x āˆˆ S, so P āŠ‚ S. Moreover, any Ī» āˆˆ / P can be removed without aļ¬€ecting the sum, so the result follows. (c) =ā‡’ (a): This is immediate since Ī“Ī» (P c ) = 0 for all Ī» āˆˆ P .



Theorem 6.5. Any Ļƒ-ļ¬nite measure can be uniquely decomposed as a sum of a pure point measure and a continuous measure. Proof. Assume that Ī¼ = Ī¼pp + Ī¼cont with Ī¼cont continuous and Ī¼pp pure point. Then Ī¼pp ({x}) = Ī¼({x}) for all x. By Lemma 6.4(iii), this determines the pure point measure uniquely as  Ī¼({Ī»})Ī“Ī» , (6.2) Ī¼pp = Ī»āˆˆP

which proves uniqueness. To prove existence, deļ¬ne Ī¼pp by (6.2). For any Borel set B, by countability of B āˆ© P ,  Ī¼({Ī»}) = Ī¼pp (B), Ī¼(B) ā‰„ Ī¼(B āˆ© P ) = Ī»āˆˆBāˆ©P

so Ī¼cont = Ī¼ āˆ’ Ī¼pp is a positive measure. By construction, for any point x,  Ī¼cont ({x}) = Ī¼({x}) āˆ’ Ī¼pp ({x}) = 0, so Ī¼cont is a continuous measure. The above decomposition can be written in the form dĪ¼cont = Ļ‡P c dĪ¼,

dĪ¼pp = Ļ‡P dĪ¼

(this notation was deļ¬ned in Proposition 1.54), where P is the set of pure points. The Fourier transform of a ļ¬nite measure on R is deļ¬ned by  Ī¼ Ė†(k) = eāˆ’ikx dĪ¼(x) (up to diļ¬€erent conventions about factors of 2Ļ€). For measures of the form dĪ¼(x) = f (x) dx, this corresponds to the Fourier transform of the function f . Generally speaking, smoothness of a function or measure is related to the decay of its Fourier transform. In particular, presence of pure points in the measure should be related to a lack of decay of its Fourier transform. It is nonetheless remarkable that pure points of Ī¼ precisely correspond to the Cesar`o-averaged limit of |Ė† Ī¼(k)|2 : Theorem 6.6 (Wiener). For any ļ¬nite Borel measure Ī¼ on R,  T  1 |Ė† Ī¼(k)|2 dk = Ī¼({Ī»})2 . lim T ā†’āˆž 2T āˆ’T

(6.3)

Ī»āˆˆR

In particular, this limit is zero if and only if Ī¼ is a continuous measure. With the same eļ¬€ort, we will prove a more general version.

162

6. Measure decompositions

Theorem 6.7 (Wiener). For any complex Borel measures Ī¼, Ī½ on R,  T  1 lim Ī¼ Ė†(k)Ė† Ī½ (k) dk = Ī¼({Ī»})Ī½({Ī»}). (6.4) T ā†’āˆž 2T āˆ’T Ī»āˆˆR

Proof. Let us ļ¬rst assume that Ī¼, Ī½ are ļ¬nite positive measures. Using the deļ¬nition of Ī¼ Ė†(k), we obtain the iterated integral  T  T   1 1 Ī¼ Ė†(k)Ė† Ī½ (k) dk = eāˆ’ik(xāˆ’y) dĪ¼(x) dĪ½(y) dk. 2T āˆ’T 2T āˆ’T R R Since Ī¼ and Ī½ are ļ¬nite, using Fubiniā€™s theorem and integrating in k gives  T   1 Ī¼ Ė†(k)Ė† Ī½ (k) dk = sinc(T (x āˆ’ y)) dĪ¼(x) dĪ½(y), 2T āˆ’T R R with the sinc function deļ¬ned by sinc u = sin u/u for u = 0 and sinc 0 = 1. By dominated convergence with dominating function 1, we compute  T   1 Ī¼ Ė†(k)Ė† Ī½ (k) dk = Ļ‡{0} (x āˆ’ y) dĪ¼(x) dĪ½(y), lim T ā†’āˆž 2T āˆ’T R R and computing the remaining integrals gives (6.4). For complex measures Ī¼, Ī½, writing them as linear combinations of positive measures and using sesquilinearity of (6.4) proves the general case. 

6.2. Singular and absolutely continuous measures In this section, we begin to consider decompositions of one measure with respect to another. The important distinction is the following. Deļ¬nition 6.8. The measure Ī¼ is singular with respect to Ī½ if there exists S such that Ī¼(S c ) = 0 and Ī½(S) = 0. This is denoted Ī¼ āŠ„ Ī½. The measure Ī¼ is continuous with respect to Ī½ if for all measurable A, Ī½(A) = 0 implies Ī¼(A) = 0. This is denoted Ī¼ " Ī½. The question is whether a measure Ī¼ can be decomposed as a sum of an absolutely continuous and a singular measure with respect to Ī½. We begin with uniqueness: Lemma 6.9. For any two measures Ī¼, Ī½, there is at most one way to decompose Ī¼ = Ī¼ac + Ī¼s so that Ī¼ac " Ī½ and Ī¼s āŠ„ Ī½. If such a decomposition exists, it is necessarily of the form dĪ¼ac = Ļ‡S c dĪ¼, for some measurable set S.

dĪ¼s = Ļ‡S dĪ¼

(6.5)

6.2. Singular and absolutely continuous measures

163

Proof. Since Ī¼s āŠ„ Ī½, there exists S such that Ī½(S) = 0 and Ī¼s (S c ) = 0. Since Ī¼ac " Ī½, Ī½(S) = 0 implies Ī¼ac (S) = 0. Thus, for any measurable A, Ī¼s (A āˆ© S c ) = 0 and Ī¼ac (A āˆ© S) = 0, so Ī¼s (A) = Ī¼s (A āˆ© S) = Ī¼(A āˆ© S),

Ī¼ac (A) = Ī¼ac (A āˆ© S c ) = Ī¼(A āˆ© S c ).

Thus, the decomposition is necessarily of the form (6.5). If there are two such sets S, T , then Ī½(S) = Ī½(T ) = 0 and Ī¼(S āˆ© T c ) = Ī¼(T āˆ© S c ) = 0; thus, Ī¼(S#T ) = 0, and the two sets give the same decomposition.  Existence of such a decomposition can be ensured using ļ¬niteness properties of the measures and this will be considered below. To appreciate the result that will follow, note a class of continuous measures: Example 6.10. A measure Ī¼ is said to be absolutely continuous with respect to Ī½ if there exists a function f ā‰„ 0 such that dĪ¼ = f dĪ½. Every such measure is continuous with respect to Ī½.   Proof. If Ī½(A) = 0, then Ļ‡A f = 0 a.e., so Ī¼(A) = Ļ‡A f dĪ½ = 0. Continuity with respect to Ī½ does not imply absolute continuity, but it does for ļ¬nite measures as a consequence of the following theorem: Theorem 6.11 (Radonā€“Nikodym). Let Ī¼, Ī½ be ļ¬nite measures on X. There exists f āˆˆ L1 (X, dĪ½), f ā‰„ 0, and Ī¼s āŠ„ Ī½ such that dĪ¼ = f dĪ½ + dĪ¼s .

(6.6)

In particular, this is the unique decomposition into a continuous and a singular part with respect to Ī½. The function f is called the Radonā€“Nikodym derivative. Proof. We denote Ī· = Ī¼ + Ī½ and deļ¬ne a linear functional on L2 (X, dĪ·) by  Ī›(h) = h dĪ½. This functional is bounded, since 7   . |h|2 dĪ·. |Ī›(h)| ā‰¤ |h| dĪ½ ā‰¤ |h| dĪ· ā‰¤ Ī·(X) Thus, by Rieszā€™s representation theorem, the functional is of the form  Ī›(h) = hg dĪ· (6.7) for some g āˆˆ L2 (X, dĪ·). Since the functional is positive, applying it to 1 } gives functions h = Ļ‡A with Am,Ā± = {x | Ā± Im g(x) ā‰„ m  1 Ī·(Am,Ā± ) ā‰¤ Ā± Im Ļ‡A g dĪ· = Ā± Im Ī›(h) = 0, m

164

6. Measure decompositions

so Ī·(Am,Ā± ) = 0. Taking the union over m and over Ā± signs shows that g is real-valued Ī·-a.e. Moreover, subtracting hg dĪ· from (6.7) gives   h(1 āˆ’ g) dĪ½ = hg dĪ¼ āˆ€h āˆˆ L2 (X, dĪ·). (6.8) Applying this to characteristic functions of sets {x | g(x) ā‰¤ āˆ’1/m} gives !   1 1 Ī½(Cm ) ā‰¤ Ļ‡Cm (1 āˆ’ g) dĪ½ = Ļ‡Cm g dĪ¼ ā‰¤ āˆ’ Ī¼(Cm ), 0ā‰¤ 1+ m m so Ī¼(Cm ) = Ī½(Cm ) = 0 for each m; thus, g ā‰„ 0 Ī·-a.e. Analogous arguments show that g ā‰¤ 1 Ī·-a.e. Note that (6.7) already gives dĪ½ = g dĪ·. To obtain the decomposition (6.6), we deļ¬ne f : X ā†’ [0, āˆž] by f = g āˆ’1 āˆ’ 1 and write (6.8) as   1 1 dĪ¼ āˆ€h āˆˆ L2 (X, dĪ·). dĪ½ = h (6.9) h āˆ’1 1+f 1+f Let us denote S = {x | f (x) = āˆž} and dĪ¼s = Ļ‡S dĪ¼. Applying (6.8) to h = Ļ‡S gives Ī½(S) = 0, so Ī¼s āŠ„ Ī½. For any Borel set B and k āˆˆ N, denote Bk = {x āˆˆ B | f (x) ā‰¤ k}. Applying (6.9) to h = (1 + f )Ļ‡Bk āˆˆ Lāˆž (X, dĪ·) āŠ‚ L2 (X, dĪ·) and taking k ā†’ āˆž gives, by monotone convergence,   Ļ‡Bāˆ©S f dĪ½ = Ļ‡Bāˆ©S dĪ¼. Since Ī½(S) = 0, we can write this as  Ļ‡B f dĪ½ = Ī¼(B āˆ© S) = Ī¼(B) āˆ’ Ī¼s (B),

(6.10)

which precisely means that dĪ¼ āˆ’ dĪ¼s = f dĪ½. Applying (6.10) to B = X  implies f āˆˆ L1 (X, dĪ½). For the following generalization, we denote by L1loc (R, dĪ½) the set of locally integrable functions on R, L1loc (R, dĪ½) = {f : R ā†’ C | Ļ‡K f āˆˆ L1 (R, dĪ½) for all compacts K āŠ‚ R}. Theorem 6.12 (Radonā€“Nikodym). Let Ī¼, Ī½ be two Baire measures on R. There exists f āˆˆ L1loc (R, dĪ½), f ā‰„ 0, and Ī¼s āŠ„ Ī½ such that (6.6) holds. In particular, this is the unique decomposition into a continuous and a singular part with respect to Ī½. The function f is called the Radonā€“Nikodym derivative. Proof. Denote F1 = [āˆ’1, 1], Fn+1 = [āˆ’n āˆ’ 1, n + 1] \ [āˆ’n, n], apply the Radonā€“Nikodym decomposition to ļ¬nite measures Ļ‡Fn dĪ¼, Ļ‡Fn dĪ½, and sum in n. 

6.2. Singular and absolutely continuous measures

165

A generalization to Ļƒ-ļ¬nite measures is considered in Exercise 6.5. The Radonā€“Nikodym decomposition with respect to Lebesgue measure Ī½ gives the Lebesgue decomposition dĪ¼ = dĪ¼ac + dĪ¼s = f (x) dx + dĪ¼s

(6.11)

into the absolutely continuous and singular part of Ī¼. In this case it is common to omit the qualiļ¬er ā€œwith respect to Lebesgue measureā€. The importance of this decomposition has led to additional terminology: Deļ¬nition 6.13. The set Sac is said to be an essential support of the absolutely continuous part of Ī¼ if the measure Ļ‡Sac (x) dx is mutually absolutely continuous with f (x) dx. An essential support of the absolutely continuous part of Ī¼ can be obtained by Sac = {x āˆˆ R | f (x) > 0}. Of course, it is not uniquely determined; it is only determined up to symmetric diļ¬€erence with a set of Lebesgue measure zero. An essential support of the absolutely continuous part of Ī¼ determines the topological support of the absolutely continuous part of Ī¼, but the converse is false: supp Ī¼ac does not determine Sac , even up to a set of measure zero (Exercise 6.6). Thus, essential support of the absolutely continuous part of Ī¼ contains more information about the measure. The Lebesgue decomposition is often combined with the decomposition into pure point and continuous parts. Since Ī¼pp is supported on a countable set, it is part of Ī¼s , and we obtain the decomposition Ī¼ = Ī¼ac + Ī¼sc + Ī¼pp , where Ī¼sc = Ī¼s āˆ’ Ī¼pp is called the singular continuous part of Ī¼. Whereas continuity of a measure is equivalent to Cesar` o-decay of |Ė† Ī¼(k)|2 by Wienerā€™s Theorem 6.6, Fourier transforms of absolutely continuous measures decay pointwise: Lemma 6.14 (Riemannā€“Lebesgue). For any f āˆˆ L1 (R),  lim

kā†’Ā±āˆž

eāˆ’ikx f (x) dx = 0.

(6.12)

166

6. Measure decompositions

Proof. By density of Cc (R) in L1 (R), for any > 0, there exists g āˆˆ Cc (R) such that f āˆ’ g1 < . Using its modulus of continuity Ļ‰g , we can estimate           āˆ’ikx āˆ’ikx āˆ’ik(x+Ļ€/k) 2 e g(x) dx =  e g(x) dx + e g(x + Ļ€/k) dx      āˆ’ikx  (g(x) āˆ’ g(x + Ļ€/k)) dx = e ā‰¤ Ļ‰g (Ļ€/k)(diam supp g + Ļ€/k). Since g āˆˆ Cc (R) is uniformly continuous, taking k ā†’ Ā±āˆž implies  eāˆ’ikx g(x) dx = 0. lim kā†’Ā±āˆž

Together with

     āˆ’ikx āˆ’ikx  e f (x) dx āˆ’ e g(x) dx ā‰¤ f āˆ’ g1 , 

this implies that

    āˆ’ikx  f (x) dx ā‰¤ f āˆ’ g1 < . lim sup  e kā†’Ā±āˆž

Since > 0 is arbitrary, this proves (6.12).



The proof of the Radonā€“Nikodym decomposition obtains f by an existence theorem, but f can be recovered by pointwise diļ¬€erentiation: Theorem 6.15. In the setting of Theorem 6.12, the limit f (x) = lim ā†“0

Ī¼((x āˆ’ , x + )) Ī½((x āˆ’ , x + ))

(6.13)

exists for (Ī¼ + Ī½)-a.e. x and recovers the decomposition (6.6) with S = f āˆ’1 ({+āˆž}) and dĪ¼s = Ļ‡S dĪ¼. The proof requires some prerequisites. Let Ī· be a Baire measure on R and let f āˆˆ L1 (R, dĪ·). For all x āˆˆ supp Ī·, we can deļ¬ne the maximal function  Ļ‡(xāˆ’r,x+r) (t)|g(t)| dĪ·(t) . (6.14) (M g)(x) = sup Ī·((x āˆ’ r, x + r)) r>0 Lemma 6.16 (Croftā€“Garsia covering lemma). Let Ī· be a Baire measure on R, and let I1 , . . . , In be a ļ¬nite family of intervals in R. There is a disjoint subfamily Ij1 , . . . , Ijk such that n  k   Ii ā‰¤ 2 Ī·(Iji ). Ī· i=1

i=1

6.2. Singular and absolutely continuous measures

167

Proof. Note that it suļ¬ƒces to prove the statement for families of intervals such that no interval is contained in the union of all the others. Let us denote Ij = (aj , bj ). If no interval is contained in the union of all the others, then aj = ak for j = k (otherwise one of the intervals Ij , Ik would be contained in the other), so we can relabel the intervals such that a1 < a2 < Ā· Ā· Ā· < an . Similarly, b1 < b2 < Ā· Ā· Ā· < bn since Ij āŠ‚ IjĀ±1 . Moreover, bj+2 ā‰„ aj for all j since Ij+1 āŠ‚ Ij āˆŖ Ij+2 . Thus, each of the subfamilies {Ij | j odd} and {Ij | j even} is a disjoint n 1 I .  subfamily; at least one of them has total measure at least 2 Ī· j j=1 Lemma 6.17. Let Ī· be a Baire measure on R and let g āˆˆ L1 (R, dĪ·). For any c > 0,  2 Ī·({x | (M g)(x) > c}) ā‰¤ |g| dĪ· (6.15) c and Ī·({x | (M g)(x) + |g(x)| > 2c}) ā‰¤

3 c

 |g| dĪ·.

(6.16)

Proof. Let K be a compact subset of {x | (M g)(x) > c}. For any x āˆˆ K, there is an interval Ix = (x āˆ’ rx , x + rx ) such that  Ļ‡Ix (t)|g(t)| dĪ·(t) > c. (6.17) Ī·(Ix ) Since K is compact, there is a ļ¬nite subcover that we denote by I1 , . . . , In . By Lemma 6.16, the ļ¬nite subcover contains a disjoint family of intervals Ij1 , . . . , Ijk such that āŽž āŽ› n k   āŽ  āŽ Ij ā‰¤ 2 Ī·(Ijl ). Ī·(K) ā‰¤ Ī· j=1

l=1

Using (6.17) and since the intervals are disjoint, this implies 2 Ī·(K) ā‰¤ c k

l=1



2 Ļ‡Ijl (t)|g(t)| dĪ·(t) ā‰¤ c

 |g(t)| dĪ·(t).

Since Ī· is inner regular, this implies (6.15). Since (M g)(x) + |g(x)| > 2c implies (M g)(x) > c or |g(x)| > c, using (6.15) and Markovā€™s inequality applied to g gives (6.16).  Although (6.15) resembles Markovā€™s inequality and is called a weak-L1 property, the maximal function is usually not an L1 function (Exercise 6.4).

168

6. Measure decompositions

Theorem 6.18. Let Ī· be a compactly supported ļ¬nite Borel measure on R and let g āˆˆ L1 (R, dĪ·). Then for Ī·-a.e. x āˆˆ R,  Ļ‡(xāˆ’,x+) (t)|g(t) āˆ’ g(x)| dĪ·(t) lim =0 (6.18) ā†“0 Ī·((x āˆ’ , x + )) and, in particular,  Ļ‡(xāˆ’,x+) (t)g(t) dĪ·(t) = g(x). (6.19) lim ā†“0 Ī·((x āˆ’ , x + )) Proof. Let us deļ¬ne



(T g)(x) = lim sup ā†“0

Ļ‡(xāˆ’,x+) (t)|g(t) āˆ’ g(x)| dĪ·(t) . Ī·((x āˆ’ , x + ))

This is well deļ¬ned for all x āˆˆ supp Ī·. The triangle inequality implies that (T g)(x) ā‰¤ (M g)(x) + |g(x)|, and subadditivity of lim sup implies that for any f, g āˆˆ L1 (dĪ·), (T (f + g))(x) ā‰¤ (T f )(x) + (T g)(x), and therefore (T f )(x) āˆ’ (T g)(x) ā‰¤ (T (f āˆ’ g))(x) ā‰¤ (T f )(x) + (T g)(x). In particular, if f is continuous, (T f )(x) = 0 identically so (T g)(x) = (T (g āˆ’ f ))(x). By Lemma 6.16, for any c > 0,

 3 |g āˆ’ f | dĪ·. c However, the right-hand side can be made arbitrarily small by density of continuous functions in L1 (R, dĪ·). Thus, Ī·({x | (T g)(x) > 2c}) = 0. Since c > 0 is arbitrary, for Ī·-a.e. x,  Ļ‡(xāˆ’,x+) (t)|g(t) āˆ’ g(x)| dĪ·(t) lim sup ā‰¤ 0, Ī·((x āˆ’ , x + )) ā†“0 Ī·({x | (T g)(x) > 2c}) = Ī·({x | (T (g āˆ’ f ))(x) > 2c}) ā‰¤

which implies (6.19).



Theorem 6.19. Let Ī· be a Baire measure on R and let f āˆˆ L1loc (R, dĪ·). Then for Ī·-a.e. x āˆˆ R, equations (6.18) and (6.19) hold. Proof. Since the condition (6.19) is local, it suļ¬ƒces to apply the previous  result to the ļ¬nite measures Ļ‡[āˆ’n,n] dĪ·, n āˆˆ N. In the special case when Ī· is Lebesgue measure, points x at which (6.18) holds are called Lebesgue points of the function g. In that terminology, the above theorem says that almost every point is a Lebesgue point.

6.3. Hausdorļ¬€ measures on R

169

Proof of Theorem 6.15. Applying Lebesgueā€™s diļ¬€erentiation theorem to g āˆˆ L1 (R, dĪ·) and using dĪ½ = g dĪ·, for Ī·-a.e. x, Ī½((x āˆ’ , x + )) = g(x). ā†“0 Ī·((x āˆ’ , x + )) By inverting and subtracting 1, we conclude that for Ī·-a.e. x, Ī¼((x āˆ’ , x + )) = f (x). lim ā†“0 Ī½((x āˆ’ , x + )) lim



6.3. Hausdorļ¬€ measures on R With the motivation that d-dimensional volume scales as the dth power of the length, Hausdorļ¬€ measures are constructed with scaling by the Ī±th power of the length, where Ī± ā‰„ 0 is not necessarily an integer and serves as a fractal dimension parameter. Compared with the decompositions seen above, Hausdorļ¬€ measures allow a more reļ¬ned look at the singular continuous part of the measure. Hausdorļ¬€ measures are studied in geometric measure theory as measures on Rd [31] and on even more general metric spaces. We will focus on Hausdorļ¬€ measures on R, which allows some simpliļ¬cation. The construction of Hausdorļ¬€ measures uses CarathĀ“eodoryā€™s Theorem 1.26; aspects of this construction can be compared with the construction of Lebesgue measure in Section 1.5. We denote the diameter of an interval by |I| = b āˆ’ a,

I = (a, b).

For Ī± < 1, assigning weight |I|Ī± to the interval I has undesired eļ¬€ects for large intervals, so a cutoļ¬€ in allowed interval sizes is needed. Thus, we start by selecting as elementary sets the intervals of length at most Ī“, EĪ“ = {āˆ…} āˆŖ {(a, b) | a, b āˆˆ R, a < b ā‰¤ a + Ī“},

(6.20)

applying to them the weight ĻĪ± (āˆ…) = 0,

ĻĪ± (I) = |I|Ī± ,

and using countable covers of arbitrary A āŠ‚ R by elements of EĪ“ to deļ¬ne

āˆž  āˆž     hāˆ—Ī±,Ī“ (A) = inf ĻĪ± (Ij )  A āŠ‚ Ij , Ij āˆˆ EĪ“ āˆ€j āˆˆ N . (6.21) j=1

For any Ī“ > 0 and s ā‰„ 0,

j=1

hāˆ—Ī±,Ī“

is an outer measure on R by Theorem 1.24.

If Ī“1 ā‰¤ Ī“2 , we note EĪ“1 āŠ‚ EĪ“2 ; since the inļ¬mum over a smaller set is larger, we observe that hāˆ—Ī±,Ī“1 (A) ā‰„ hāˆ—Ī±,Ī“2 (A). Thus, we can deļ¬ne hāˆ—Ī± (A) = lim hāˆ—Ī±,Ī“ (A) = sup hāˆ—Ī±,Ī“ (A). Ī“ā†“0

Ī“>0

For any Ī± ā‰„ 0, this is an outer measure by abstract arguments:

170

6. Measure decompositions

Lemma 6.20. If Ī¼āˆ—Ī“ are outer measures for all Ī“, then Ī¼āˆ— (A) = sup Ī¼āˆ—Ī“ (A) Ī“

is also an outer measure. Proof. Obviously, Ī¼āˆ— (āˆ…) = 0, and A āŠ‚ B implies Ī¼āˆ—Ī“ (A) ā‰¤ Ī¼āˆ—Ī“ (B) for all Ī“, so it implies Ī¼āˆ— (A) ā‰¤ Ī¼āˆ— (B). For any sets An , n āˆˆ N, āˆž  āˆž āˆž    āˆ— āˆ— Ī¼Ī“ An ā‰¤ Ī¼Ī“ (An ) ā‰¤ Ī¼āˆ— (An ), n=1

n=1

n=1

so taking the supremum over Ī“ shows Ļƒ-subadditivity of Ī¼āˆ— .



In particular, hāˆ—Ī± is an outer measure. Let us note a special case: Lemma 6.21. Hausdorļ¬€ outer measure hāˆ—0 is the counting measure on R. Proof. If A contains n elements x1 , . . . , xn , denote Ī“0 = min{|xj āˆ’ xk | | 1 ā‰¤ j < k ā‰¤ n}. An interval of length smaller than Ī“0 contains at most one of the numbers x1 , . . . , xn , so hāˆ—0,Ī“ (A) ā‰„ n for all Ī“ < Ī“0 . It follows that hāˆ—0 (A) ā‰„ n and therefore hāˆ—0 (A) ā‰„ #A. Conversely, a ļ¬nite set A with #A = n can be  covered by n intervals of arbitrarily small length, so hāˆ—0 (A) ā‰¤ #A. Moreover, hāˆ—1 is the Lebesgue outer measure on R (Exercise 6.7). For any Ī± > 1, hāˆ—Ī± is identically zero (Exercise 6.8), so we will restrict ourselves to Ī± āˆˆ [0, 1] from now on. In order to apply CarathĀ“eodoryā€™s Theorem 1.26 and prove that outer measures hāˆ—Ī± generate Borel measures, we need the following. Lemma 6.22. For any c āˆˆ R, (c, āˆž) is measurable with respect to hāˆ—Ī± . Proof. Since hāˆ—0 is counting measure on R, it suļ¬ƒces to consider the case Ī± > 0. Consider a countable cover of a set E āŠ‚ R by intervals Ij of length at most Ī“. Separating the intervals Ij based on whether they intersect (āˆ’āˆž, c] or not, we obtain two subfamilies: the ļ¬rst is a cover of E āˆ© (āˆ’āˆž, c], and the second is a cover of E āˆ© [c + Ī“, āˆž), because the intervals have length at most Ī“. By adding the interval (c, c + Ī“) to the second subfamily, we conclude  ĻĪ± (Ij ) ā‰„ hāˆ—Ī±,Ī“ (E āˆ© (āˆ’āˆž, c]), j:Ij āˆ©(āˆ’āˆž,c]=āˆ… Ī±

Ī“ +



j:Ij āˆ©(āˆ’āˆž,c]=āˆ…

ĻĪ± (Ij ) ā‰„ hāˆ—Ī±,Ī“ (E āˆ© (c, āˆž)).

6.3. Hausdorļ¬€ measures on R

171

Adding these two inequalities gives Ī“Ī± +

āˆž 

ĻĪ± (Ij ) ā‰„ hāˆ—Ī±,Ī“ (E āˆ© (āˆ’āˆž, c]) + hāˆ—Ī±,Ī“ (E āˆ© (c, āˆž)),

j=1

and taking the inļ¬mum over all countable covers of E gives Ī“ Ī± + hāˆ—Ī±,Ī“ (E āˆ© (c, āˆž)) ā‰„ hāˆ—Ī±,Ī“ (E āˆ© (āˆ’āˆž, c]) + hāˆ—Ī±,Ī“ (E āˆ© (c, āˆž)). In the limit Ī“ ā†“ 0, this gives hāˆ—Ī± (E) ā‰„ hāˆ—Ī± (E āˆ© (āˆ’āˆž, c]) + hāˆ—Ī± (E āˆ© (c, āˆž)).

(6.22)

The opposite inequality holds by subadditivity. Thus, equality holds in (6.22) for any E āŠ‚ R, so (c, āˆž) is measurable with respect to hāˆ—Ī± .  Theorem 6.23. For any Ī± ā‰„ 0, hĪ± = hāˆ—Ī± |BR is a Borel measure on R. Proof. By CarathĀ“eodoryā€™s Theorem 1.26, the family of measurable sets with respect to hāˆ—Ī± is a Ļƒ-algebra; since it contains the half-lines (c, āˆž), it contains all Borel sets. Thus, the restriction of hāˆ—Ī± to the Borel Ļƒ-algebra  BR is a measure. Hausdorļ¬€ measures have many applications. As a ļ¬rst application, if we ļ¬x a set A and vary Ī±, we observe a certain critical value at which the value of hāˆ—Ī± (A) changes: Theorem 6.24. Let A āŠ‚ R. (a) If hāˆ—Ī± (A) < āˆž for some Ī±, then hāˆ—Ī² (A) = 0 for all Ī² > Ī±. (b) If hāˆ—Ī² (A) > 0 for some Ī², then hāˆ—Ī± (A) = āˆž for all Ī± < Ī². (c) There is a unique number d āˆˆ [0, 1] such that hāˆ—Ī± (A) = āˆž āˆ€Ī± āˆˆ [0, d),

hāˆ—Ī² (A) = 0 āˆ€Ī² āˆˆ (d, 1].

(6.23)

Proof. If 0 ā‰¤ Ī± ā‰¤ Ī² and x āˆˆ [0, Ī“], then xĪ² ā‰¤ Ī“ Ī²āˆ’Ī± xĪ± . Thus, for any interval I of length at most Ī“, ĻĪ² (I) ā‰¤ Ī“ Ī²āˆ’Ī± ĻĪ± (I). Applying this to countable covers of A by intervals of length at most Ī“ gives hāˆ—Ī²,Ī“ (A) ā‰¤ Ī“ Ī²āˆ’Ī± hāˆ—Ī±,Ī“ (A).

(6.24)

Now (a) follows by taking Ī“ ā†“ 0 in (6.24), and (b) follows by dividing (6.24) by Ī“ Ī²āˆ’Ī± and then taking Ī“ ā†“ 0. (c) For the set {0} āˆŖ {x āˆˆ [0, 1] | hāˆ—x (A) = āˆž} is nonempty; denote by d its supremum. Applying (a) and (b) concludes the proof.  Deļ¬nition 6.25. The number d with the property (6.23) is called the Hausdorļ¬€ dimension of the set A and is denoted by dimH A.

172

6. Measure decompositions

Hausdorļ¬€ dimension provides a ļ¬ne gradation in the relative thickness of sets; it can distinguish between many Cantor sets with zero Lebesgue measure (Exercise 6.11). Although every countable set has zero Hausdorļ¬€ dimension, the converse is not true (Exercise 6.10). Deļ¬nition 6.26. A setA is said to be Ļƒ-ļ¬nite with respect to Ī½ if there exist sets An with A = āˆž n=1 An and Ī½(An ) < āˆž for all n. Theorem 6.24 can be reļ¬ned to show that for Ī± < dimH A, the set A is not Ļƒ-ļ¬nite with respect to hĪ± (Exercise 6.12). We now turn to the characterization and decomposition of Baire measures with respect to Hausdorļ¬€ measures. Deļ¬nition 6.27. A Baire measure Ī¼ is said to be Ī±-continuous if Ī¼ " hĪ± , and it is said to be Ī±-singular if Ī¼ āŠ„ hĪ± . For Ī± < 1, the measures hĪ± are not Baire measures, and they are not Ļƒ-ļ¬nite. Thus, the Radonā€“Nikodym theorem does not apply; in order to decompose Ī¼ into an Ī±-continuous and an Ī±-singular part, we need a diļ¬€erent approach. We will use the upper Ī±-derivative Ī¼((x āˆ’ r, x + r)) . (6.25) DĪ¼Ī± (x) = lim sup (2r)Ī± rā†“0 Lemma 6.28. DĪ¼Ī± is a Borel function. Proof. For any r > 0, the set S = {(x, t) āˆˆ R2 | x āˆ’ r < t < x + r} is Borel. So by Tonelliā€™s theorem,  Ī¼((x āˆ’ r, x + r)) = Ļ‡S (x, t) dĪ¼(t) is a Borel function of x. Since Ī¼((x āˆ’ r, x + r))/(2r)Ī± is left-continuous in r āˆˆ (0, āˆž), and since any limit can be written as a limit along a sequence, Ī¼((x āˆ’ r, x + r)) Ī¼((x āˆ’ r, x + r)) = lim sup . Ī± nā†’āˆž rāˆˆ(0,1/n)āˆ©Q Rā†’0 rāˆˆ(0,R) (2r) (2r)Ī±

DĪ¼Ī± (x) = lim sup

The right-hand side is a Borel function by Lemma 1.40.



The function DĪ¼Ī± allows us to decompose the Baire measure Ī¼ into absolutely continuous and singular parts with respect to hĪ± : Theorem 6.29 (Rogersā€“Taylor). For a Baire measure Ī¼ on R, denote T = {x | DĪ¼Ī± (x) = āˆž} and deļ¬ne measures dĪ¼Ī±c = Ļ‡T c dĪ¼,

dĪ¼Ī±s = Ļ‡T dĪ¼.

Then Ī¼ = Ī¼Ī±c + Ī¼Ī±s , Ī¼Ī±c " hĪ± , and Ī¼Ī±s āŠ„ hĪ± .

6.3. Hausdorļ¬€ measures on R

173

The proof requires a covering theorem which allows inļ¬nite families of intervals. We call a family of sets G disjoint if for any G1 , G2 āˆˆ G, G1 = G2 implies G1 āˆ© G2 = āˆ…. Theorem 6.30 (Vitali). For any family F of open intervals whose union is bounded, there exists a countable disjoint subfamily G āŠ‚ F such that   JāŠ‚ (x āˆ’ 5r, x + 5r). (6.26) JāˆˆF

(xāˆ’r,x+r)āˆˆG

Proof. Set F1 = F . The construction is inductive: If Fn = āˆ…, we set Ln = sup{|I| | I āˆˆ Fn } and choose In = (xn āˆ’ rn , xn + rn ) āˆˆ Fn with |In | > Ln /2, then we set Fn+1 = {I āˆˆ Fn | I āˆ© In = āˆ…}. If FN = āˆ… for some N , we write LN = 0 and terminate the construction. Otherwise, we write N = āˆž. This construction gives a countable subfamily G = {In | 1 ā‰¤ n < N }. By construction, Ik āˆ© In = āˆ… if k < n. For N= boundedness of the union of disjoint intervals gives āˆž n=1 |In | < āˆž āˆž, āˆž so n=1 Ln < āˆž; thus, Ln ā†’ 0 as n ā†’ āˆž. Thus, in both cases, for any J āˆˆ F , there exists n such that Ln < |J| and therefore J āˆˆ / Fn . Choose the largest n such that J āˆˆ Fn . Then J āˆ© In = āˆ…  and |J| ā‰¤ Ln < 2|In |. Thus, J āŠ‚ (xn āˆ’ 5rn , xn + 5rn ). Proof of the Rogersā€“Taylor theorem. We will prove that for ļ¬nite, compactly supported measures Ī¼, hĪ± (T ) = 0 and for any A, hĪ± (A) = 0 =ā‡’ Ī¼(A \ T ) = 0. The general case then follows by Ļƒ-additivity using the measures Ļ‡[āˆ’n,n] dĪ¼. Fix k, Ī“ āˆˆ (0, āˆž) and denote Tk = {x | DĪ¼Ī± (x) > k}. For any x āˆˆ Tk , by deļ¬nition of DĪ¼Ī± , there exists rx < Ī“/10 such that Ī¼((x āˆ’ rx , x + rx )) ā‰„ k(2rx )Ī± . The intervals (x āˆ’ rx , x + rx ) cover the set Tk , so there is a countable disjoint subfamily G such that  (x āˆ’ 5rx , x + 5rx ). Tk āŠ‚ (xāˆ’rx ,x+rx )āˆˆG

Using the deļ¬nition of Hausdorļ¬€ measures,   5Ī± (10rx )Ī± ā‰¤ hĪ±,Ī“ (Tk ) ā‰¤ k (xāˆ’rx ,x+rx )āˆˆG

(xāˆ’rx ,x+rx )āˆˆG

Ī¼((x āˆ’ rx , x + rx )).

174

6. Measure decompositions

Since G is a disjoint family, this implies hĪ±,Ī“ (Tk ) ā‰¤

5Ī± Ī¼(R). k

Applying this with k ā†’ āˆž shows hĪ±,Ī“ (T ) = 0 for all Ī“ > 0, so hĪ± (T ) = 0. To prove the other claim, observe that DĪ¼Ī± (x) < āˆž implies Ī¼((x āˆ’ r, x + r)) 0. For any a, b > 0, using r = max{a, b} gives Ī¼((x āˆ’ a, x + b)) Ī¼((x āˆ’ r, x + r)) ā‰¤ sup . Ī± (a + b) rĪ± a,bāˆˆ(0,1] rāˆˆ(0,1] sup

Thus, if we denote Ī¼((x āˆ’ a, x + b)) (a + b)Ī± a,bāˆˆ(0,1]  and write Ek = {x | CĪ¼Ī± (x) ā‰¤ k}, then T c āŠ‚ kāˆˆN Ek . CĪ¼Ī± (x) =

sup

Take A āŠ‚ R such that hĪ± (A) = 0. Consider a countable cover of A by intervals In of length |In | ā‰¤ 1. For n such that In āˆ© Ek = āˆ…, using the point x āˆˆ In , CĪ¼Ī± (x) ā‰¤ k implies Ī¼(In ) ā‰¤ CĪ¼Ī± (x)|In |Ī± ā‰¤ k|In |Ī± . Thus Ī¼(A āˆ© Ek ) ā‰¤

 n:In āˆ©Ek =āˆ…

Ī¼(In ) ā‰¤

āˆž 

k|In |Ī± .

n=1

Taking the inļ¬mum over covers with |In | ā‰¤ 1 for all n gives Ī¼(A āˆ© Ek ) ā‰¤  khĪ±,1 (A) = 0. Taking the union over k āˆˆ N gives Ī¼(A āˆ© T c ) = 0. In the special case Ī± = 1, the decomposition Ī¼ = Ī¼Ī±c + Ī¼Ī±s is the Lebesgue decomposition, because h1 is Lebesgue measure. These decompositions characterize the behavior of parts of Ī¼ with respect to sets of zero Ī½-measure. A diļ¬€erent kind of decomposition is performed with respect to Ļƒ-ļ¬nite sets with respect to hĪ± : Deļ¬nition 6.31. The measure Ī¼ is strongly Ī±-continuous if Ī¼(A) = 0 for every set A which is Ļƒ-ļ¬nite with respect to hĪ± . The measure Ī¼ is almost Ī±-singular if there exists S which is Ļƒ-ļ¬nite with respect to hĪ± such that Ī¼(S c ) = 0.

6.3. Hausdorļ¬€ measures on R

175

A Baire measure Ī¼ has a unique decomposition into a strongly Ī±-continuous part and an almost Ī±-singular part: uniqueness is proved in Exercise 6.15 and existence in Exercise 6.16. Note that a set is Ļƒ-ļ¬nite with respect to counting measure if and only if it is countable, so the special case Ī± = 0 recovers continuous and pure point measures. Deļ¬nition 6.32. Let Ī± āˆˆ (0, 1]. A ļ¬nite measure Ī¼ on R is uniformly Ī±HĀØolder continuous (abbreviated UĪ±H) if there exists C āˆˆ (0, āˆž) such that for all x āˆˆ R and all r āˆˆ (0, 1/2], Ī¼((x āˆ’ r, x + r)) ā‰¤ C(2r)Ī± . Theorem 6.33 (Rogersā€“Taylor). A ļ¬nite measure Ī¼ on R is Ī±-continuous if and only if for every > 0, there exists a decomposition Ī¼ = Ī¼1 + Ī¼2 with Ī¼1 a UĪ±H measure and Ī¼2 a measure with Ī¼2 (R) < . Proof. Assume that Ī¼ " hĪ± . Then in the notation of Theorem 6.29 and its proof, Ī¼(T ) = 0. In particular, there exists k āˆˆ N such that Ī¼(Ekc ) < . Denote dĪ¼1 = Ļ‡Ek dĪ¼, dĪ¼2 = Ļ‡Ekc dĪ¼. Then Ī¼2 (R) < by deļ¬nition. For any x āˆˆ R and all r āˆˆ (0, 1/2], let us prove Ī¼1 ((x āˆ’ r, x + r)) = Ī¼((x āˆ’ r, x + r) āˆ© Ek ) ā‰¤ k(2r)Ī± . If (x āˆ’ r, x + r) āˆ© Ek = āˆ…, choosing a point y in the intersection, we write (x āˆ’ r, x + r) = (y āˆ’ a, y + b) with a, b āˆˆ (0, 1], so the inequality follows from the deļ¬nition of Ek . If (x āˆ’ r, x + r) āˆ© Ek = āˆ…, the inequality is trivial. Conversely, assume that Ī¼ has such a decomposition for every > 0. Any UĪ±H measure is Ī±-continuous, so Ī¼1 " hĪ± . Decomposing Ī¼2 into an Ī±continuous and an Ī±-singular part and adding to Ī¼1 , we get a decomposition of Ī¼; in particular, Ī¼Ī±s = (Ī¼2 )Ī±s . It follows that Ī¼Ī±s (R) = (Ī¼2 )Ī±s (R) < .  Since is arbitrary, it follows that Ī¼Ī±s = 0, so Ī¼ " hĪ± . Existence of subsets with ļ¬nite nonzero Hausdorļ¬€ measure is a nontrivial result, whose proof goes beyond the scope of this text. By results of Besicovitch [10] and Davies [23] (see also [31, Theorem 1.6, Theorem 5.6]): Theorem 6.34. For every Borel set A āŠ‚ R and Ī± āˆˆ [0, 1], if hĪ± (A) = āˆž, there exists a compact subset K āŠ‚ A such that 0 < hĪ± (K) < āˆž. In particular, taking dĪ¼ = Ļ‡K dhĪ± and applying Theorem 6.33 shows: Corollary 6.35. For every Borel set A āŠ‚ R and Ī± āˆˆ [0, 1], if hĪ± (A) > 0, there exists a UĪ±H measure Ī½ such that 0 < Ī½(R) < āˆž and supp Ī½ āŠ‚ A. This can be used to estimate the Hausdorļ¬€ dimension of sets (Exercise 6.17).

176

6. Measure decompositions

6.4. Matrix-valued measures We now consider matrix-valued measures Ī©, with positivity in the sense of operator order (i.e., the values are positive semideļ¬nite matrices). This generalization will occur naturally in spectral theory when we consider fullline Jacobi matrices and full-line SchrĀØodinger operators. Recall that L(Cd ) is the set of d Ɨ d matrices, and that inequalities between matrices should be interpreted in the sense of operator order. Deļ¬nition 6.36. A map Ī© : BX ā†’ L(Cd ) is a positive d Ɨ d-matrix valued measure on X if it is Ļƒ-additive, Ī©(āˆ…) = 0, and for every B āˆˆ BX , Ī©(B) ā‰„ 0. Note that Ī© cannot take inļ¬nite values, so ļ¬niteness of the measure is built into the deļ¬nition. The following lemma provides a decomposition of matrix-valued measures in the style of Radonā€“Nikodym: Lemma 6.37. If Ī© is a positive d Ɨ d-matrix valued Borel measure on X, then the following hold. (a) Tr Ī© is a ļ¬nite positive measure. (b) There exists a matrix-valued Borel function W : X ā†’ L(Cd ) such that  W d Tr Ī© (6.27) Ī©(B) = B

for any Borel set B. (c) Tr Ī©-a.e., Tr W = 1 and W ā‰„ 0. Proof. (a) For any v āˆˆ Cd , the map Ī¼v (B) = v āˆ— Ī©(B)v is Ļƒ-additive. Since Ī©(B) ā‰„ 0 implies v āˆ— Ī©(B)v ā‰„ 0, Ī¼v is a positive measure. Since Ī¼v (X) = v āˆ— Ī©(X)v āˆˆ [0, āˆž), it is a ļ¬nite positive measure. Thus, so is Tr Ī© =

d 

Ī¼ej ,

j=1

where (ej )dj=1 denotes the standard basis of Cd . (b) For any set B with Tr Ī©(B) = 0, Ī©(B) ā‰„ 0 implies Ī©(B) = 0 and then v āˆ— Ī©(B)v = 0 for all v. Thus, for any v āˆˆ Cn , Ī¼v " Tr Ī©, so by the Radonā€“Nikodym decomposition, there exists a positive function fv āˆˆ L1 (X, d Tr Ī©) such that  āˆ— fv d Tr Ī©. v Ī©(B)v = B

By the polarization identity for the sesquilinear form (u, v) ā†’ uāˆ— Ī©(B)v,   1 Ļ‰ āˆ’1 fej +Ļ‰ek d Tr Ī© eāˆ—j Ī©(B)ek = 4 B Ļ‰āˆˆ{1,i,āˆ’1,āˆ’i}

6.4. Matrix-valued measures

177

from which we read oļ¬€ functions  1 wjk = Ļ‰ āˆ’1 fej +Ļ‰ek āˆˆ L1 (X, d Tr Ī©) 4 Ļ‰āˆˆ{1,i,āˆ’1,āˆ’i}

such that (6.27) holds. In particular, Ī©jk is a complex measure. Note also that each wjk is a.e. ļ¬nite so by changes on zero measure sets, we can assume W : X ā†’ L(Cd ). (c) For each v āˆˆ Cd , we can express the positive measure Ī¼v as dĪ¼v = v āˆ— W v d Tr Ī© and conclude that v āˆ— W v ā‰„ 0 Tr Ī©-a.e. Applying this to the countable set of v āˆˆ (Q + iQ)d and taking a countable union of zero measure sets, we obtain a set E with Tr Ī©(E) = 0 such that v āˆ— W (x)v ā‰„ 0 for all v āˆˆ (Q + iQ)d and all x āˆˆ E c . Then, by continuity, v āˆ— W v ā‰„ 0 for all v āˆˆ Cd and all x āˆˆ E c . Thus, W ā‰„ 0 a.e. Summing (6.27) on the diagonal gives d Tr Ī© = Tr W d Tr Ī© so Tr W = 1 Tr Ī©-a.e.  We proved that any positive d Ɨ d matrix-valued measure is of the form dĪ© = W dĪ¼, where W = (wjk )dj,k=1 ā‰„ 0 is a matrix-valued function with Tr W = 1 and Ī¼ is a ļ¬nite positive measure. It is sometimes natural to consider objects formally given by dĪ© = W dĪ¼ in greater generality, with the following warning. If W  dĪ¼ = āˆž, the symbol dĪ© = W dĪ¼ usually does not deļ¬ne a map Ī© on the original Ļƒ-algebra, since the entries  wjk may not be elements of L1 (X, dĪ¼), and in particular, for j = k, wjk dĪ¼ may be undeļ¬ned. Nonetheless, integration with respect to W dĪ¼ is well deļ¬ned and it is the natural setting for deļ¬ning vector-valued L2 spaces: Lemma 6.38. Let Ī¼ be a positive measure on X, and let W be a positive matrix-valued function on X. Then the following hold. (a) For vector-valued Borel functions f : X ā†’ Cd , !1/2  āˆ— f  = f W f dĪ¼

(6.28)

is well deļ¬ned (i.e., the integral is nonnegative) and deļ¬nes a seminorm. (b) The relation deļ¬ned by f āˆ¼ g if f āˆ’ g = 0 is an equivalence relation. (c) The set L2 (X, Cd , W dĪ¼) of equivalence classes with f  < āˆž is a Hilbert space with the inner product  (6.29) f, g = f āˆ— W g dĪ¼.

178

6. Measure decompositions

Proof. Denote

& '  d āˆ— L (X, C , W dĪ¼) = f : X ā†’ C | f W f dĪ¼ < āˆž . 2

d

Writing W = W 1/2 W 1/2 , we note that   f āˆ— W f dĪ¼ = (W 1/2 f )āˆ— (W 1/2 f ) dĪ¼, 4 so f āˆˆ L2 (X, Cd , W dĪ¼) if and only if W 1/2 f āˆˆ dj=1 L2 (X, dĪ¼). In particular, L2 (X, Cd , W dĪ¼) is a vector space, and the map T f = W 1/2 f , T : L2 (X, Cd , W dĪ¼) ā†’

d 3

L2 (X, dĪ¼)

j=1

4d 2 is norm-preserving. By pulling back properties from j=1 L (X, dĪ¼), it follows that (6.28) deļ¬nes a norm, āˆ¼ an equivalence relation, and (6.29) an inner product on the quotient space L2 (X, Cd , W dĪ¼). sequence in L2 (X, Cd , W dĪ¼). Then Let (fn )āˆž n=1 be an arbitrary Cauchy4 d 2 (W 1/2 fn )āˆž n=1 is a Cauchy sequence in j=1 L (X, dĪ¼), so by completeness 4d it has a limit g āˆˆ j=1 L2 (X, dĪ¼) in the sense that  lim (W 1/2 fn āˆ’ g)āˆ— (W 1/2 fn āˆ’ g) dĪ¼ = 0. nā†’āˆž

By the Rieszā€“Fischer theorem, there is a subsequence (W 1/2 fnk )āˆž k=1 which converges to g pointwise a.e. If limkā†’āˆž W (x)1/2 fnk (x) = g(x), then g(x) āˆˆ Ran W (x)1/2 (note that Ran W (x)1/2 is closed for each x). Choosing f (x) āˆˆ Ran W (x) with W 1/2 f (x) = g(x) gives a factorization g = W 1/2 f such that 2 d  the Cauchy sequence (fn )āˆž n=1 converges to f in L (X, C , W dĪ¼).

6.5. Exercises 6.1. Let A be a Ļƒ-algebra on X, and let Ī¼ : A ā†’ C be a Ļƒ-additive map, i.e., āŽž āŽ› āˆž āˆž   Aj āŽ  = Ī¼(Aj ) (6.30) Ī¼āŽ j=1

j=1

for all pairwise disjoint Aj āˆˆ A (in particular, the series is always convergent). (a) The variation of Ī¼ is deļ¬ned as 

āˆž āˆž     |Ī¼(Aj )|  A = Aj , Aj āˆ© Ak = āˆ… if j = k . |Ī¼|(A) = sup j=1

j=1

Prove that |Ī¼| is a positive measure on A.

6.5. Exercises

179

(b) If |Ī¼|(E) = āˆž for some set E, prove that there is a disjoint decomposition E = A1 āˆŖ E1 such that |Ī¼(A1 )| ā‰„ 1 and |Ī¼|(E1 ) = āˆž. Hint: Reduce to the case when Ī¼ takes only real values. In this āˆž case, ļ¬nd a partition (Bn )āˆž n=1 of E such that n=1 |Ī¼(Bn )| > 2 + 2|Ī¼(E)| and group the Bn ā€™s by whether Ī¼(Bn ) is positive or negative. (c) Prove that |Ī¼|(X) < āˆž. Hint: Assume the opposite and use (b) inductively. (d) Prove that there exist (positive) ļ¬nite measures Ī¼j : A ā†’ [0, āˆž) such that Ī¼ = Ī¼1 āˆ’ Ī¼2 + i(Ī¼3 āˆ’ Ī¼4 ) and Ī¼j " |Ī¼| for all j. . Hint: If Ī¼ is real valued, use Ī¼Ā± (A) = |Ī¼|(A)Ā±Ī¼(A) 2 1 (X, d|Ī¼|) such that Ī¼(A) = (e) Prove that there exists f āˆˆ L  A f d|Ī¼| for all A āˆˆ A, and prove that |f | = 1 |Ī¼|-a.e. 6.2. For any ļ¬nite Borel measures Ī¼, Ī½ on R, prove that   1 T Ī¼ Ė†(k)Ė† Ī½ (k) dk = Ī¼({Ī»})Ī½({Ī»}). lim T ā†’āˆž T 0 Ī»āˆˆR

6.3. Prove that measures Ī¼, Ī½ on X are mutually singular if there exist sets Sn āŠ‚ X such that lim Ī¼(Sn ) = 0,

nā†’āˆž

lim Ī½(Snc ) = 0.

nā†’āˆž

āˆ’n and consider the equation Hint: Reduce āˆž to the case Ī¼(Sn ) ā‰¤ 2 āˆž S = m=1 n=m Sn .

6.4. If Ī½ denotes the Lebesgue measure on R and f āˆˆ L1 (R, dĪ½) is not 0, prove that the function M f deļ¬ned in (6.14) is not in L1 (R, dĪ½). 6.5. (a) Prove that any Ļƒ-ļ¬nite measure Ī¼ can be written in the form dĪ¼ = w dĖœ Ī¼ for some ļ¬nite measure Ī¼ Ėœ and function 0 < w < 1. (b) Prove that for any Ļƒ-ļ¬nite measures Ī¼, Ī½, there is a unique decomposition dĪ¼ = f dĪ½ + dĪ¼s , where f ā‰„ 0 and Ī¼s āŠ„ Ī½. Hint: Use (a) to reduce to ļ¬nite measures Ī¼ Ėœ, Ī½Ėœ. 6.6. Consider a Baire measure Ī¼ with decomposition (6.11) and an essential support of the absolutely continuous part of Ī¼, denoted Sac . (a) Prove that the topological support of f (x) dx is equal to the essential closure of Sac , i.e., the set {x | m(E āˆ© (x āˆ’ , x + )) > 0 for all > 0}. (b) Prove that there exists an absolutely continuous measure Ī¼ such that supp Ī¼ = [0, 1] but Ī¼ is not mutually absolutely continuous with Ļ‡[0,1] (x) dx (i.e., does not have [0, 1] as an essential support). Hint: Use Exercise 1.16.

180

6. Measure decompositions

6.7. Prove that for Ī± = 1 and any Ī“ > 0, the Hausdorļ¬€ outer measure hāˆ—1,Ī“ is the Lebesgue outer measure on R. In particular, so is hāˆ—1 . 6.8. For any Ī± > 1 and Ī“ > 0, prove that hāˆ—Ī±,Ī“ (A) = 0 for every A āŠ‚ R. In particular, hāˆ—Ī± (A) = 0. 6.9. Prove hĪ± (A) = 0 if and only if there exist open intervals In with āˆž that Ī± < āˆž such that every point x āˆˆ A belongs to inļ¬nitely |I | n=1 n many of the intervals In . 6 5 āˆž āˆž N is uncount6.10. Prove that the set A = n=1 an /n! | (an )n=1 āˆˆ {0, 1} able and has zero Hausdorļ¬€ dimension. 6.11. Fix Ī³ āˆˆ (0, 1/2). The middle- 1āˆ’Ī³ 2 Cantor set is obtained by denoting f0 (x) = Ī³x, f1 (x) = Ī³x + 1 āˆ’ Ī³, and recursively deļ¬ning C0 = [0, 1],

Cn+1 = f0 (Cn ) āˆŖ f1 (Cn ),

C=

āˆž

Cn

n=0

(the special case Ī³ = 1/3 is the middle-third Cantor set). Denote Ī± = log 2/ log(1/Ī³). (a) Prove that for any intervals I1 , I2 with |I1 |, |I2 | ā‰„

Ī³ d(I1 , I2 ) 1 āˆ’ 2Ī³

there is an interval I such that I1 āˆŖI2 āŠ‚ I and |I1 |Ī± +|I2 |Ī± ā‰„ |I|Ī± . (b) For any Ī“, > 0, prove that there exists a ļ¬nite cover of C by closed intervals I1 , . . . , In such that n  |Ij |Ī± ā‰¤ hĪ±,Ī“ (C) + , j=1

and any endpoint of any Ij is a boundary point of Cn for some n. (c) Prove that the step in (a) can be applied to the ļ¬nite cover in (b) iteratively until there is only one interval left, and conclude that hĪ±,Ī“ (C) ā‰„ 1. (d) Prove that hĪ± (C) = 1 and dimH C = Ī±. 6.12. (a) Let A āŠ‚ R. If Ī± < dimH A, prove that A is not Ļƒ-ļ¬nite with respect to hĪ± . (b) Prove that for any Ī± < 1, hĪ± is not a Ļƒ-ļ¬nite measure on R. 6.13. Let Ī¼ be a Baire measure on R, and let S be a Borel set with Ī¼(S) > 0. If for some Ī± āˆˆ [0, 1], DĪ¼Ī± (x) < āˆž for Ī¼-a.e. x āˆˆ S, prove that dimH (S) ā‰„ Ī±.

6.5. Exercises

181

6.14. Let Ī¼ be a Baire measure on R. Prove that log Ī¼((x āˆ’ r, x + r)) Ī±āˆ— (x) = lim inf rā†’0 log r (with the convention log 0 = āˆ’āˆž) is a Borel function of x āˆˆ R, and that DĪ¼Ī± (x) = 0

āˆ€Ī± < Ī±āˆ— (x),

DĪ¼Ī± (x) = āˆž

āˆ€Ī± > Ī±āˆ— (x).

6.15. For measures Ī¼, Ī½, prove that there is at most one way to decompose Ī¼ = Ī¼1 + Ī¼2 so that both of the following hold. (a) For any set A which is Ļƒ-ļ¬nite with respect to Ī½, Ī¼1 (A) = 0. (b) There exists a set S which is Ļƒ-ļ¬nite with respect to Ī½ and Ī¼2 (S c ) = 0. 6.16. Let Ī¼ be a Baire measure on R, and denote PĪ± = {x | DĪ¼Ī± (x) = 0}. (a) For the measures Ī¼sĪ±c = Ļ‡PĪ± dĪ¼,

Ī¼aĪ±s = Ļ‡PĪ±c dĪ¼,

prove that Ī¼sĪ±c is strongly Ī±-continuous, Ī¼aĪ±s is almost Ī±-singular, and Ī¼ = Ī¼sĪ±c + Ī¼aĪ±s . (b) Prove that {x | 0 < DĪ¼Ī± (x) < āˆž} is Ļƒ-ļ¬nite with respect to hĪ± and that the measure Ī¼Ī±c āˆ’ Ī¼sĪ±c can be represented in the form Ī¼Ī±c āˆ’ Ī¼sĪ±c = f dhĪ± for some Borel function f ā‰„ 0.

Ī± 6.17. Let xn āˆˆ R and cn > 0 be such that āˆž n=1 cn < āˆž for some Ī± āˆˆ (0, 1). (a) Prove that for any Ī± < Ī² and any ļ¬nite UĪ²H measure Ī¼,   āˆž cĪ±n dĪ¼(x) < āˆž. |x āˆ’ xn |Ī± n=1 āˆž cĪ± cn n (b) Prove that āˆž n=1 |xāˆ’xn | = āˆž implies n=1 |xāˆ’xn |Ī± = āˆž. (c) Using Corollary 6.35, prove that the set 

āˆž  cn =āˆž S= xāˆˆR| |x āˆ’ xn | n=1

has Hausdorļ¬€ dimension at most Ī±.

Chapter 7

Herglotz functions

We have already encountered integrals of the form  1 dĪ¼(x), f (z) = xāˆ’z

(7.1)

where Ī¼ is a ļ¬nite Borel measure on R; they appeared in the identity (5.18) relating resolvents of self-adjoint operators with the spectral measures. The function f deļ¬ned by (7.1) is sometimes called the Stieltjes transform or the 1 Im z = |xāˆ’z| Borel transform of the measure Ī¼. Since Im xāˆ’z 2 , the function f (z) maps the upper half-plane C+ = {z āˆˆ C | Im z > 0} to itself. Functions with this property have a central place in the spectral theory of self-adjoint operators, and it is beneļ¬cial to consider them starting from a general perspective. Deļ¬nition 7.1. A Herglotz function is an analytic function f : C+ ā†’ C+ . Although not all Herglotz functions are of the form (7.1), we will show that every Herglotz function has an integral representation which generalizes (7.1) in a natural way, !  x 1 āˆ’ dĪ¼(x) (7.2) f (z) = az + b + 1 + x2 R xāˆ’z with a ā‰„ 0, b āˆˆ R, and with Ī¼ a positive measure on R which obeys  1 dĪ¼(x) < āˆž. 1 + x2 R

(7.3)

Measures obeying (7.3) are said to be Poisson-ļ¬nite. To see that (7.2) is a generalization of (7.1), note that if Ī¼ is ļ¬nite, then the two terms in the 183

184

7. Herglotz functions

integrand in (7.2) are separately integrable, and the second is independent of z so it is simply an additive constant. In this case, f (z) is an aļ¬ƒne function of z plus an integral of the form (7.1). From one perspective, a description of all Herglotz functions is possible because the condition f (z) āˆˆ C+ is a form of boundedness: boundedness away from āˆ’i. This becomes more transparent when working with bounded domains, so we will begin with brief considerations of MĀØobius transformations and of certain families of functions on the unit disk D = {z āˆˆ C | |z| < 1}. We will then prove the representation (7.2) and explore the various relations between the function f and the measure Ī¼.

7.1. MĀØ obius transformations Recall the equivalence relation $ on C2 \ { u$v

0 0 } deļ¬ned by

if and only if u = Ī»v for some Ī» āˆˆ C \ {0}.

(7.4)

Ė† = C āˆŖ {āˆž} The quotient space can be identiļ¬ed with 0 the Riemann sphere wC  2 Ė† deļ¬ned by Ļ€ 1 = w1 . We by using the quotient map Ļ€ : C \ { 0 } ā†’ C w w2 2 w1  0 w1 Ė† refer to w2 = 0 as projective coordinates corresponding to w = w2 āˆˆ C. Ė† More Every 2 Ɨ 2 matrix A preserves cosets, so it induces a map on C. Ė† Ė† explicitly, the MĀØobius transformation fA : C ā†’ C is uniquely deļ¬ned by fA ā—¦ Ļ€ = Ļ€ ā—¦ A. This is often written as ! ! fA (w) w $A 1 1   1 (even if w = āˆž or fA (w) = āˆž, with the convention āˆž 1 $ 0 ) or as fA (w) =

A11 w + A12 , A21 w + A22

where Aij denote entries of the matrix A. Lemma 7.2. Let A, B be invertible 2 Ɨ 2 matrices. Then the following hold. (a) fA = id if and only if A = Ī»I for some Ī» āˆˆ C \ {0}. (b) For any invertible 2 Ɨ 2 matrices A and B, fAB = fA ā—¦ fB . Ė† (c) For any invertible 2 Ɨ 2 matrix A, the map fA is a bijection of C to itself. Proof. (a) fA = id means that every nonzero vector w āˆˆ C2 is an eigenvector of A. This is only possible if A is a multiple of the identity matrix.

7.1. MĀØobius transformations

185

(b) fA ā—¦ Ļ€ = Ļ€ ā—¦ A and fB ā—¦ Ļ€ = Ļ€ ā—¦ B imply fAB ā—¦ Ļ€ = Ļ€ ā—¦ AB = (Ļ€ ā—¦ A) ā—¦ B = fA ā—¦ Ļ€ ā—¦ B = fA ā—¦ fB ā—¦ Ļ€. Since Ļ€ is surjective, this implies fAB = fA ā—¦ fB . (c) follows from (a) and (b) using fA ā—¦ fAāˆ’1 = fAāˆ’1 ā—¦ fA = fI = id.



Due to the ļ¬rst property, when considering MĀØobius transformations, it is common to normalize the matrix by the condition det A = 1. Recall that SL(2, C) = {A āˆˆ L(C2 ) | det A = 1}. Even with that normalization, there is a remaining ambiguity: matrices A and āˆ’A have the same determinant and fA = fāˆ’A . We note also that in computational problems, it is convenient not to worry about normalization: Example 7.3 (Cayley transform). The function zāˆ’i z+i

Ī³(z) =

(7.5)

is an analytic bijection from C+ to D. Its inverse is w+1 . iw āˆ’ i

Ī³ āˆ’1 (w) = Proof. From the calculation 1 āˆ’ |Ī³(z)|2 =

|z + i|2 āˆ’ |z āˆ’ i|2 4 Im z = , 2 |z + i| |z + i|2

it follows that Ī³(z) āˆˆ D if and only if z āˆˆ C+ . The formula for Ī³ āˆ’1 follows from ! ! ! 1 āˆ’i 1 1 2 0 = .  1 i i āˆ’i 0 2 Example 7.4. For any z0 āˆˆ D, Ī³z0 (z) =

z āˆ’ z0 1 āˆ’ z0 z

= Ī³āˆ’z0 . is an analytic bijection of D to itself and Ī³zāˆ’1 0 Proof. Similarly to the previous example, this follows from the calculations 1 āˆ’ |Ī³z0 (z)|2 = and 1 āˆ’z0 āˆ’z0 1

!

1 z0 z0 1

(|1 āˆ’ |z0 |2 )(|1 āˆ’ |z|2 ) |1 āˆ’ zĀÆ0 z|2 ! =

! 1 āˆ’ |z0 |2 0 . 0 1 āˆ’ |z0 |2



186

7. Herglotz functions

These examples hint at a general property of MĀØobius transformations. Ė† any of the following regions: a disk Let us call a generalized disk in C {w āˆˆ C | |w āˆ’ c| < r} with c āˆˆ C, r > 0, the complement of its closure {w āˆˆ C | |w āˆ’ c| > r} āˆŖ {āˆž}, or a half-plane {w āˆˆ C | Re(eāˆ’iĻ† w) > t} with Ļ†, t āˆˆ R. The three cases are distinguished by whether āˆž is outside, inside, or on the boundary of the generalized disk. Ė† can be represented in projective Lemma 7.5. Every generalized disk in C coordinates in the form ! ! w1 w1 āˆ— M >0 (7.6) w2 w2 for some Hermitian 2 Ɨ 2 matrix M with det M < 0, and conversely, every set of this form is a generalized disk. Proof. The condition |w āˆ’ c| < r is equivalent to (7.6) with the matrix ! āˆ’1 c . Mc,r = cĀÆ r2 āˆ’ |c|2 The region |w āˆ’c| > r can likewise be represented by (7.6) with M = āˆ’Mc,r . The half-plane condition Re(eāˆ’iĻ† w) > t is equivalent to (7.6) with the matrix ! 0 eiĻ† Ėœ MĻ†,t = āˆ’iĻ† . e āˆ’t Ėœ Ļ†,t are 2 Ɨ 2 Hermitian matrices with The matrices Mc,r , āˆ’Mc,r , and M strictly negative determinant. For the converse, assume that M = M āˆ— and det M < 0. Note that multiplying M by a positive scalar does not aļ¬€ect the condition (7.6). By separating cases based on the sign of M11 , it is straightforward to see that Ėœ Ļ†,t .  M is a positive multiple of some matrix Mc,r , āˆ’Mc,r , or M Lemma 7.6. The generalized disk described by (7.6) is a disk in the complex plane if and only if M11 < 0. In this case, its radius r and center c are given by āˆš M12 āˆ’ det M , c=āˆ’ . (7.7) r=āˆ’ M11 M11 Proof. This follows from the previous proof by noting that the formulas (7.7) are invariant under rescaling M by a positive constant and by verifying  them for the matrix Mc,r . Proposition 7.7. MĀØ obius transformations map every generalized disk bijectively to a generalized disk.

7.1. MĀØobius transformations

187

Proof. For the generalized disk given by wāˆ— M  w > 0, its image under the MĀØobius transformation fA is characterized by the condition wāˆ— M w > 0, where M = (Aāˆ’1 )āˆ— M  Aāˆ’1 . Since M is also Hermitian and det M =  det M  /|det A|2 , the image is also a generalized disk. An important special case is upper half-plane: Example 7.8. Let J = A vector w āˆˆ C2 \ { if wāˆ— J w > 0.

! 0 i . āˆ’i 0

0 0 } corresponds to a coset Ļ€(w) =

(7.8) w1 w2

in C+ if and only

Proof. This follows from the calculation ! ! w1 2 w1 w1 āˆ— = āˆ’i(w1 w2 āˆ’ w1 w2 ) = 2 Im(w1 w2 ) = J Im . 2 w2 w2 |w2 | w2



Deļ¬nition 7.9. A matrix A is called J -expanding if and J -contracting if

Aāˆ— J A āˆ’ J ā‰„ 0

(7.9)

J āˆ’ Aāˆ— J A ā‰„ 0.

(7.10)

These notions provide the criterion for a MĀØobius transformation to map the upper half-plane into (but not necessarily onto) itself. Lemma 7.10. If A is invertible and J -contracting, then fAāˆ’1 (C+ ) āŠ‚ C+ . Proof. By (7.9), (Av)āˆ— J Av > 0 implies v āˆ— J v ā‰„ v āˆ— Aāˆ— J Av > 0, so Ļ€(Av) āˆˆ  C+ implies Ļ€(v) āˆˆ C+ . If A āˆˆ SL(2, C) is J -expanding and J -contracting, it is said to be J -unitary. For our choice of J , J -unitary matrices can be described in terms of entries: Lemma 7.11. An SL(2, C) matrix is J -unitary if and only if it has real entries. Of course, this set is denoted by SL(2, R) = {A āˆˆ SL(2, C) | Ajk āˆˆ R for all j, k}. Proof. The set G = {A āˆˆ SL(2, C) | Aāˆ— J A = J } is a group. If A āˆˆ SL(2, R), a direct calculation shows Aāˆ— J A = J . Thus, SL(2, R) āŠ‚ G. Conversely, assume that A āˆˆ SL(2, C) is J -unitary. Then fAāˆ’1 (C+ ) = C+ so fA preserves RāˆŖ{āˆž}. We separate cases based on the value of fA (āˆž).

188

7. Herglotz functions   If fA (āˆž) = āˆž, then A 01 $ 01 , which shows that ! a11 a12 . A= 0 a22

Thus, fA (z) = (a11 /a22 )z+(a12 /a22 ) is an aļ¬ƒne map on C. Since it preserves C+ , it must have a11 /a22 > 0 and a12 /a22 āˆˆ R. Together with det A = a11 a22 = 1, this implies A āˆˆ SL(2, R). If fA (āˆž) = Ī» āˆˆ R, choosing B=

!

0 1 āˆ’1 Ī»

āˆˆ SL(2, R)

gives fB ā—¦ fA = fBA which maps āˆž to āˆž and BA āˆˆ G. By the previous case, BA āˆˆ SL(2, R) so A āˆˆ SL(2, R).  Some further properties of J -contracting matrices are left as exercises. While we will only use the notion of a J -contracting matrix for the matrix (7.8) which corresponds to C+ , other choices are also of interest; most ! āˆ’1 0 notably, j = corresponds to D. 0 1

7.2. Schur functions and convergence For analytic functions on a ļ¬xed region Ī© āŠ‚ C, the natural notion of convergence is uniform convergence on compact subsets of Ī©. For instance, since every contour on Ī© has a compact image, this notion of convergence allows us to exchange limits and contour integrals. It is a remarkable fact that uniform convergence on compacts can sometimes be concluded from pointwise convergence. In this section, we will not aim for a general treatment, but we present a self-contained discussion for the following class of functions. Deļ¬nition 7.12. A Schur function is an analytic function g : D ā†’ D. By the maximum principle, if |g(z)| = 1 for some z āˆˆ D, then g is constant. Thus, any Schur function is either a unimodular constant g ā‰” eiĻ† or a function g : D ā†’ D. Lemma 7.13 (Schwarz lemma). If g is a Schur function and g(0) = 0, then g(z)/z is a Schur function. Proof. Since g(0) = 0, the function h(z) = g(z)/z has a removable singularity at 0. Thus, h is an analytic function on D, and by the maximum principle, for any r < 1, 1 1 sup |g(z)| ā‰¤ . sup |h(z)| = sup |h(z)| = r z:|z|=r r z:|z|ā‰¤r z:|z|=r Taking the limit r ā†‘ 1 completes the proof.



7.2. Schur functions and convergence

189

Theorem 7.14 (Schwarzā€“Pick theorem for D). Let g : D ā†’ D be an analytic function. Then for all z, w āˆˆ D,    g(z) āˆ’ g(w)   z āˆ’ w     . (7.11)  ā‰¤  1 āˆ’ g(w)g(z)   1 āˆ’ wz  Proof. Using the MĀØ obius transformations from Example 7.4, the inequality (7.11) can be written as |Ī³g(w) (g(z))| ā‰¤ |Ī³w (z)| āˆ’1 and Ī¶ = Ī³ (z), as |h(Ī¶)| ā‰¤ |Ī¶|. Since or, in terms of h = Ī³g(w) ā—¦ g ā—¦ Ī³w w h : D ā†’ D and h(0) = 0, this follows from the Schwarz lemma. 

The Schwarzā€“Pick theorem has an analogue for the half-plane (Exercise 7.5). In the setting of the unit disk, we use the Schwarzā€“Pick theorem to gain insight about the set of all Schur functions, viewed as a subset of C(D, C). Lemma 7.15. The set of all Schur functions is equicontinuous on D. Proof. For |z|, |w| < r < 1, for any Schur function g, we claim that |g(z) āˆ’ g(w)| ā‰¤

2 |z āˆ’ w|. 1 āˆ’ r2

If g is constant, this is trivial; otherwise, it follows from the Schwarzā€“Pick theorem together with |1 āˆ’ g(z)g(w)| ā‰¤ 2 and |1 āˆ’ zw| ā‰„ 1 āˆ’ r2 . This implies equicontinuity at any point z āˆˆ D.  Proposition 7.16. If a sequence of Schur functions converges pointwise on a dense set in D, then it converges to a Schur function uniformly on compact subsets of D. Proof. Denote the sequence by gn . Since the functions gn are equicontinuous and converge pointwise on a dense set, by Theorem 2.14, they converge uniformly on compact subsets of D to a continuous function gāˆž : D ā†’ D. Analyticity of gāˆž now follows from Moreraā€™s theorem: since any contour Ī³ in D has a compact image, by uniform convergence on the range of Ī³,   lim gn (z) dz = lim gn (z) dz = 0.  Ī³ nā†’āˆž

nā†’āˆž Ī³

From the previous results, we prove a special case of Montelā€™s theorem, a more general result in complex analysis. Theorem 7.17. Every sequence of Schur functions has a subsequence which converges to a Schur function uniformly on compact subsets of D.

190

7. Herglotz functions

Proof. By a diagonalization argument (Lemma 2.13), there is a subsequence which converges pointwise on a countable dense subset of D. By equicontinuity and Theorem 2.14, this subsequence converges uniformly on compact subsets of D.  Note that any Herglotz function f corresponds to a Schur function g = Ī³ ā—¦ f ā—¦ Ī³ āˆ’1 , where Ī³ is the Cayley transform (7.5). This generates all the Schur functions except the unimodular constants g(z) āˆˆ āˆ‚D, which correspond to f (z) = c āˆˆ Ė† ā†’ C Ė† are continuous, the R āˆŖ {āˆž}. By this conjugation, since Ī³, Ī³ āˆ’1 : C previous results immediately extend to Herglotz functions: Ė† pointCorollary 7.18. If a sequence of Herglotz functions converges (on C) wise on a dense set in C+ , then it converges uniformly on compact subsets of C+ to a Herglotz function or to a constant c āˆˆ R āˆŖ {āˆž}. Corollary 7.19. Every sequence of Herglotz functions has a subsequence which converges uniformly on compact subsets of C+ to a Herglotz function or to a constant c āˆˆ R āˆŖ {āˆž}. The set of Schur functions can be equipped with a metric which corresponds to convergence on compacts (Exercise 7.4); with such a metric, Theorem 7.17 tells us that the set of Schur functions is a compact metric space.

7.3. CarathĀ“ eodory functions The Cauchy integral formula represents an analytic function inside a region by the values on the boundary. In fact, more is true: an analytic function on D can, up to an imaginary constant, be reconstructed from the real part of its values on āˆ‚D: Proposition 7.20 (Schwarz integral formula). Let F be analytic in a neighborhood of D. Then for all z āˆˆ D,  iĪø e +z dĪø F (z) = i Im F (0) + Re F (eiĪø ) . (7.12) iĪø e āˆ’z 2Ļ€ We mention Proposition 7.20 purely for context: we will not use or directly prove it, but our goal is to describe a generalization of this formula dĪø is in which F is not necessarily analytic on the boundary, and Re F (eiĪø ) 2Ļ€ replaced by a positive measure. Proposition 7.20 will follow easily from that generalization. We will use the set M(āˆ‚D) of ļ¬nite positive Borel measures on āˆ‚D and the notion of weak convergence of measures (Deļ¬nition 2.57).

7.3. CarathĀ“eodory functions

191

Theorem 7.21. Every analytic function F : D ā†’ {z āˆˆ C | Re z ā‰„ 0} can be written uniquely in the form  iĪø e +z F (z) = iĪ² + dĻ(Īø) (7.13) eiĪø āˆ’ z for some Ī² āˆˆ R and Ļ āˆˆ M(āˆ‚D). The constant Ī² and the measure Ļ can be obtained from F by Ī² = Im F (0) and dĻ(Īø) = w-lim Re F (reiĪø ) rā†‘1

dĪø . 2Ļ€

Remark 7.22. This integral representation is often discussed in the context of CarathĀ“eodory functions, which are analytic functions F : D ā†’ {z āˆˆ C | Re z ā‰„ 0} with F (0) = 1. Since (7.13) implies F (0) = iĪ² + Ļ(āˆ‚D), CarathĀ“eodory functions correspond to the case when Ī² = 0 and Ļ is a probability measure. The proof of Theorem 7.21 requires some preliminary statements. Let us begin with the easy direction. Lemma 7.23. For any Ī² āˆˆ R and Ļ āˆˆ M(āˆ‚D), (7.13) deļ¬nes an analytic function F : D ā†’ {z āˆˆ C | Re z ā‰„ 0}. If moments of Ļ are denoted for k āˆˆ Z by  ck = eāˆ’ikĪø dĻ(Īø), then the power series representation of F around 0 is F (z) = iĪ² + c0 + 2

āˆž 

ck z k .

(7.14)

k=1

Proof. The integral kernel in (7.12) can be expanded as a geometric series as āˆž  1 + zeāˆ’iĪø eiĪø + z = =1+ 2z k eāˆ’ikĪø . (7.15) eiĪø āˆ’ z 1 āˆ’ zeāˆ’iĪø k=1

Substituting (7.15) into (7.13) gives    āˆž  2z k eāˆ’ikĪø dĻ(Īø). F (z) = iĪ² + 1+ k=1

The series and integral can be exchanged by Fubiniā€™s theorem, since    āˆž  1 + |z| Ļ(āˆ‚D) < āˆž. |2z k eāˆ’ikĪø | dĻ(Īø) = 1+ 1 āˆ’ |z| k=1

(7.16)

192

7. Herglotz functions

Integrating (7.16) term by term using the moments of Ļ gives the power series (7.14). As noted, by Fubiniā€™s theorem, the power series is convergent for z āˆˆ D; thus, it deļ¬nes an analytic function on D. It follows from Re

eiĪø + z 1 āˆ’ |z|2 = >0 eiĪø āˆ’ z |eiĪø āˆ’ z|2

that Re F (z) ā‰„ 0 for all z āˆˆ D.

(7.17) 

Lemma 7.24. If, for a sequence of measures Ļn āˆˆ M(āˆ‚D), the limits of moments  (7.18) eāˆ’ikĪø dĻn (Īø) lim nā†’āˆž

are convergent for k āˆˆ N0 , then Ļn converge weakly to some measure Ļāˆž āˆˆ M(āˆ‚D).  Proof. Observe the corresponding functionals Ī›n (h) = h dĻn on C(āˆ‚D). Using (7.18) for k = 0, we conclude that Ļn (āˆ‚D) converge; as Ļn are positive measures, Ī›n  = Ļn (āˆ‚D), so Ī›n is a bounded sequence of linear functionals. By the statement (7.18) and its complex conjugate, the limit Ī›āˆž (h) = lim Ī›n (h) nā†’āˆž

is convergent for h = eāˆ’ikĪø for all k āˆˆ Z. By linearity, it is convergent if h is a trigonometric polynomial, and by density and boundedness, it is convergent for every h āˆˆ C(āˆ‚D) by Lemma 2.46. The limit Ī›āˆž (h) is a positive linear functional on C(āˆ‚D), so it corresponds to some Ļāˆž āˆˆ M(āˆ‚D) by the Rieszā€“Markov theorem.  Proof of Theorem 7.21. For analytic F : D ā†’ {z āˆˆ C | Re z ā‰„ 0}, denote by āˆž  ak z k F (z) = k=0

its power series centered at 0. For r < 1, consider the measures dĪø dĻr = Re F (reiĪø ) . 2Ļ€ They are positive measures because Re F ā‰„ 0. By Cauchyā€™s integral theorem, the moments of the function F (reiĪø ) for k āˆˆ Z are

  2Ļ€ ak rk k ā‰„ 0 1 iĪø āˆ’ikĪø dĪø āˆ’kāˆ’1 = F (re )e F (rz)z dz = (7.19) 2Ļ€ 2Ļ€i āˆ‚D 0 k ā‰¤ āˆ’1. 0 Complex-conjugating and replacing k by āˆ’k gives

 2Ļ€ 0 kā‰„1 dĪø āˆ’ikĪø = F (reiĪø )e āˆ’k 2Ļ€ aāˆ’k r k ā‰¤ 0. 0

(7.20)

7.4. The Herglotz representation

193

Taking the average of (7.19) and (7.20) gives the moments of Re F (reiĪø ) as āŽ§ 1 k āŽŖ kā‰„1  āŽØ 2 ak r dĪø āˆ’ikĪø iĪø = Re a0 Re F (re ) e k=0 2Ļ€ āŽŖ āŽ©1 āˆ’k ĀÆāˆ’k r k ā‰¤ āˆ’1. 2a By Lemma 7.24, the measures Ļr converge weakly as r ā†‘ 1. Also, comparing these moments with Lemma 7.23 and with the Taylor expansion of F (rz) shows that  iĪø dĪø e +z Re F (reiĪø ) . F (rz) = i Im F (0) + eiĪø āˆ’ z 2Ļ€ Since F (rz) ā†’ F (z) as r ā†‘ 1, weak convergence implies (7.13) with Ī² =  Im F (0) and dĻ = w-limrā†‘1 dĻr . Beyond the existence and uniqueness of the representation (7.13), we also want to know its continuity properties. Theorem 7.25. Given analytic functions Fn : D ā†’ {z āˆˆ C | Re z ā‰„ 0}, n āˆˆ N āˆŖ {āˆž}, with representations  iĪø e +z dĻn (Īø), Fn (z) = iĪ²n + eiĪø āˆ’ z the following are equivalent: (a) Fn (z) ā†’ Fāˆž (z) for every z āˆˆ D. (b) The sequence Fn converges to Fāˆž uniformly on compact subsets of D. w

(c) Ī²n ā†’ Ī²āˆž and Ļn ā†’ Ļāˆž . Proof. (a) =ā‡’ (b): This follows from Proposition 7.16 applied to the Schur functions fn (z) = Ī³(iFn (z)), where Ī³ is the Cayley transform (7.5). (b) =ā‡’ (c): Ī²n = i Im Fn (0) ā†’ i Im Fāˆž (0) = Ī²āˆž . Cauchyā€™s integral formula applied to a circle of radius r < 1 implies that Taylor coeļ¬ƒcients of Fn converge to those of Fāˆž . Thus, moments of Ļn converge to those of Ļāˆž , w so Ļn ā†’ Ļāˆž by Lemma 7.24. iĪø

is in C(āˆ‚D), so (c) =ā‡’ (a): For every z āˆˆ D, the function eiĪø ā†’ eeiĪø +z āˆ’z this follows by the deļ¬nition of weak convergence of measures. 

7.4. The Herglotz representation In this section, we derive an integral representation for Herglotz functions. Note that the Herglotz condition Im f (z) > 0 is an open condition, but for the following result it is more natural to allow the slightly more general case Im f (z) ā‰„ 0. Of course, if f : C+ ā†’ C+ āˆŖ R obeys f (z0 ) āˆˆ R for some

194

7. Herglotz functions

z0 āˆˆ C+ , then f is constant by the maximum principle. Thus, every analytic f : C+ ā†’ C+ āˆŖ R is a Herglotz function or a real-valued constant. Theorem 7.26 (Herglotz representation, ļ¬rst form). Every analytic function f : C+ ā†’ {z āˆˆ C | Im z ā‰„ 0} has a unique representation of the form  1 + xz dĪ½(x), (7.21) f (z) = az + b + R xāˆ’z where a ā‰„ 0, b āˆˆ R, and Ī½ is a ļ¬nite positive measure on R. Proof. Recall the Cayley transform Ī³ deļ¬ned by (7.5). The function F (w) = āˆ’if (Ī³ āˆ’1 (w)) maps D to {z āˆˆ C | Re z ā‰„ 0}, so there exist Ī² āˆˆ R and Ļ āˆˆ M(āˆ‚D) such that  eiĪø + w dĻ(Īø). (7.22) F (w) = iĪ² + iĪø āˆ‚D e āˆ’ w Solving for f gives  eiĪø + Ī³(z) dĻ(Īø). f (z) = āˆ’Ī² + i iĪø āˆ‚D e āˆ’ Ī³(z) The measure Ļ may have a point mass at eiĪø = 1; separating that point mass from the rest of the integral gives  1 + Ī³(z) eiĪø + Ī³(z) Ļ({1}) + i dĻ(Īø). f (z) = āˆ’Ī² + i iĪø 1 āˆ’ Ī³(z) āˆ‚D\{1} e āˆ’ Ī³(z) Since

Ī³(x) + Ī³(z) 1 + xz 1 + Ī³(z) = z, i = , 1 āˆ’ Ī³(z) Ī³(x) āˆ’ Ī³(z) xāˆ’z the representation (7.21) is obtained by algebraic manipulations after denoting b = āˆ’Ī², a = Ļ({1}), and denoting by Ī½ the measure on R obtained as the pushforward of the measure Ļ on āˆ‚D \ {1} under the bijection Ī³ āˆ’1 : āˆ‚D \ {1} ā†’ R. Explicitly, for any Borel set B āŠ‚ R, Ī½(B) = Ļ(Ī³(B)). i

Conversely, starting from the representation (7.21), the above steps can be reversed to represent the function F (w) in the form (7.22) with Ī² = āˆ’b and Ļ = aĪ“1 + Ī½ ā—¦ Ī³ āˆ’1 , where Ī“1 denotes the Dirac measure at 1. Since the representation (7.22) is unique, so is the representation (7.21).  While the Herglotz representation in the form (7.21) follows naturally from the Cayley transform, the Herglotz representation is more commonly stated in the form stated in the introduction to this chapter. Theorem 7.27 (Herglotz representation, second form). Every analytic function f : C+ ā†’ {z āˆˆ C | Im z ā‰„ 0} has a unique representation of the form (7.2), with a ā‰„ 0, b āˆˆ R, and Ī¼ a positive measure on R which obeys (7.3).

7.4. The Herglotz representation

195

Proof. (7.2) is obtained from (7.21) by x 1 + xz 1 āˆ’ = 2 xāˆ’z 1+x (x āˆ’ z)(1 + x2 )

(7.23) 

with dĪ¼(x) = (1 + x2 )dĪ½(x).

While (7.2) is more commonly referred to as the Herglotz representation and is more natural from the perspective of generalizing (7.1) and of the Stieltjes inversion described below, the alternative form (7.21) is more convenient for other purposes; we will use them interchangeably. We are also interested in the continuity properties of the Herglotz representation. Denote C0 (R) = {h āˆˆ C(R) | lim h(x) = 0}. xā†’Ā±āˆž

Proposition 7.28. Given analytic functions fn : C+ ā†’ {z āˆˆ C | Im z ā‰„ 0}, n āˆˆ N āˆŖ {āˆž}, with Herglotz representations  1 + xz dĪ½n (x), (7.24) fn (z) = an z + bn + R xāˆ’z the following are equivalent: (a) fn (z) ā†’ fāˆž (z) for every z āˆˆ C+ ; (b) the sequence fn converges to fāˆž uniformly on compact subsets of C+ ; (c) bn ā†’ bāˆž , an + Ī½n (R) ā†’ aāˆž + Ī½āˆž (R), and   āˆ€h āˆˆ C0 (R). h dĪ½n ā†’ h dĪ½āˆž

(7.25)

Proof. By applying Theorem 7.25 to the functions  eiĪø + w āˆ’1 dĻn (Īø), Fn (w) = āˆ’ifn (Ī³ (w)) = iĪ²n + iĪø āˆ‚D e āˆ’ w it follows that (a) and (b) are mutually equivalent and equivalent to the condition that   āˆ€g āˆˆ C(āˆ‚D). g dĻn ā†’ g dĻāˆž Any g āˆˆ C(āˆ‚D) can be uniquely written as a linear combination of the constant function 1 and a function obeying g(1) = 0. Convergence for the constant function is equivalent to an +Ī½n (R) ā†’ aāˆž +Ī½āˆž (R) and convergence for functions obeying g(1) = 0 is equivalent to (7.25) with h = g ā—¦ Ī³.  Corollary 7.29. In the setting of Proposition 7.28, if fn ā†’ fāˆž , then lim sup an ā‰¤ aāˆž . nā†’āˆž

(7.26)

196

7. Herglotz functions

Proof. Fix c > 0 and a continuous function h on R such that Ļ‡[āˆ’c,c] ā‰¤ h ā‰¤ Ļ‡[āˆ’2c,2c] . Then  Ī½āˆž ([āˆ’c, c]) ā‰¤ h dĪ½āˆž = lim h dĪ½n ā‰¤ lim inf Ī½n (R). nā†’āˆž

nā†’āˆž

Since this holds for any c > 0, it follows that Ī½āˆž (R) ā‰¤ lim inf Ī½n (R). nā†’āˆž

Subtracting this from an + Ī½n (R) ā†’ aāˆž + Ī½āˆž (R) implies (7.26).



The inequality (7.26) can be strict (Exercise 7.9). Finally, let us note a specialization of Proposition 7.28. In spectral theory, we usually consider approximations of a Herglotz function with aāˆž = 0, which leads to a slight simpliļ¬cation. Corollary 7.30. In the setting of Proposition 7.28, if aāˆž = 0, the following are equivalent: (a) fn (z) ā†’ fāˆž (z) for every z āˆˆ C+ ; (b) the sequence fn converges to fāˆž uniformly on compact subsets of C+ ; (c) bn ā†’ bāˆž , an ā†’ 0, Ī½n (R) ā†’ Ī½āˆž (R), and (7.25) holds. Proof. (a) =ā‡’ (b) and (c) =ā‡’ (a) follow directly from Proposition 7.28. For (b) =ā‡’ (c), note that since aāˆž = 0 and an ā‰„ 0 for all n, (7.26) implies an ā†’ 0. Then an + Ī½n (R) ā†’ aāˆž + Ī½āˆž (R) implies Ī½n (R) ā†’ Ī½āˆž (R), and the rest follows. 

7.5. Growth at inļ¬nity and tail of the measure In this section we express the coeļ¬ƒcients a and b in the Herglotz representation in terms of the values of f ; we will see that the value of a is related to the asymptotic behavior of f at inļ¬nity. We also give a necessary and suļ¬ƒcient condition for a Herglotz function to be of the special form (7.1). We begin by noting that f (i) = ai + b +



1 + xi dĪ½(x) = b + (a + Ī½(R))i xāˆ’i

so b = Re f (i), a + Ī½(R) = Im f (i). However, isolating the value of a requires taking a limit: Proposition 7.31. If the function f is given by (7.2), then f (iy) . yā†’āˆž iy

a = lim

7.5. Growth at inļ¬nity and tail of the measure

Proof. Since

aiy+b iy

ā†’ a, it suļ¬ƒces to prove that  1 1 + ixy dĪ½(x) = 0 lim yā†’āˆž iy x āˆ’ iy

197

(7.27)

for ļ¬nite measures Ī½. The integrand converges to 0 pointwise as y ā†’ āˆž, so (7.27) follows by dominated convergence with the bound .   . 2y2  1 1 + ixy  1 + x 1 + x2 y 2   =. ā‰¤ 1,  iy x āˆ’ iy  = . 2 y x + y2 y 4 + x2 y 2 which is valid for all x āˆˆ R and y ā‰„ 1.



The limit can also be taken nontangentially (Exercise 7.11). Proposition 7.32. Let f be a Herglotz function. The function f is of the form (7.1) for some ļ¬nite positive measure Ī¼ on R if and only if there exists C < āˆž such that C āˆ€z āˆˆ C+ . (7.28) |f (z)| ā‰¤ Im z Proof. If (7.1) holds and Ī¼ is ļ¬nite, then |x āˆ’ z| ā‰„ Im z implies that  Ī¼(R) 1 dĪ¼(t) ā‰¤ , |f (z)| ā‰¤ |t āˆ’ z| Im z so (7.28) holds with C = Ī¼(R). Conversely, let f be a Herglotz function. If (7.28) holds, by Proposition 7.31, f (iy) a = lim = 0. yā†’āˆž y Since  y dĪ¼(t), Im f (x + iy) = (x āˆ’ t)2 + y 2 monotone convergence implies  y2 lim y Im f (x + iy) = lim dĪ¼(t) = Ī¼(R), yā†‘āˆž yā†‘āˆž (x āˆ’ t)2 + y 2 so (7.28) implies Ī¼(R) < āˆž. Thus, the two terms in the integrand in (7.2) are separately integrable and f is of the form  1 dĪ¼(t) f (z) = Ī² + R tāˆ’z for some Ī² āˆˆ R. Now limyā†‘āˆž f (iy) = Ī² by dominated convergence, but that limit is zero by (7.28).  These results relate the behavior of the measure at āˆž and the behavior of the Herglotz function at āˆž, as does the following result:

198

7. Herglotz functions

Proposition 7.33. Let f be a Herglotz function with the Herglotz representation (7.2). For any Ī³ āˆˆ (0, 2),   āˆž dĪ¼(x) Im f (iy) a = 0 and < āˆž ā‡ā‡’ dy < āˆž. (7.29) Ī³ yĪ³ 1 R 1 + |x| Lemma 7.34. For any Īŗ āˆˆ (āˆ’1, 1),  āˆž Ļ€/2 tĪŗ . dt = 2 1+t cos(ĪŗĻ€/2) 0

(7.30) Īŗ

z Proof. Consider the meromorphic function f (z) = 1+z 2 on C \ [0, āˆ’iāˆž) Īŗ with arg(z ) = Īŗ arg z for arg z āˆˆ (āˆ’Ļ€/2, 3Ļ€/2). Let 0 < r < R < āˆž, and consider the region Ī© = {z āˆˆ C | r < |z| < R, 0 < arg z < Ļ€}. The function f has a pole at i, so by residue calculus, the contour integral of f over āˆ‚Ī© is + f (z) dz = 2Ļ€i Resi (f ) = Ļ€eiĪŗĻ€/2 . āˆ‚Ī©

Parametrizing the contour gives  R  Ļ€  Ļ€ 1+Īŗ iĪŗĪø tĪŗ R1+Īŗ eiĪŗĪø r e iĪŗĻ€ dt+i dĪø āˆ’i dĪø = Ļ€eiĪŗĻ€/2 . (1+e ) 2 iĪø 2 iĪø 2 r 1+t 0 1 + (Re ) 0 1 + (re ) Since Īŗ āˆˆ (āˆ’1, 1), letting r ā†“ 0 and R ā†‘ āˆž gives  āˆž tĪŗ iĪŗĻ€ dt = Ļ€eiĪŗĻ€/2 , (1 + e ) 1 + t2 0 which implies (7.30).



Proof of Proposition 7.33. If a > 0, then the right-hand side is also false by Proposition 7.31, so we assume a = 0 from now on. If we decompose !   t 1 1 dĪ¼(t) + āˆ’ dĪ¼(t), f (z) = b1 + 1 + t2 (āˆ’1,1) t āˆ’ z R\(āˆ’1,1) t āˆ’ z  t where b1 = b āˆ’ (āˆ’1,1) 1+t 2 dĪ¼(t), by Proposition 7.32 the contribution from the ļ¬rst integral is O(1/ Im z), so it does not aļ¬€ect the equivalence (7.29). Thus, it suļ¬ƒces to prove the equivalence in the case !  t 1 āˆ’ dĪ¼(t), f (z) = b + t āˆ’ z 1 + t2 where supp Ī¼ āŠ‚ (āˆ’āˆž, āˆ’1] āˆŖ [1, āˆž). In that case, f (iy) is bounded for y āˆˆ (0, 1) because  7  2 2  1 + iyt  = 1+y t ā‰¤1  āˆ€y āˆˆ (0, 1), āˆ€t āˆˆ R \ (āˆ’1, 1),  t āˆ’ iy  t2 + y 2

7.6. Half-plane Poisson kernel and Stieltjes inversion

so (7.29) is equivalent to the equivalence  āˆž Im f (iy) dy < āˆž ā‡” yĪ³ 0

 R

199

1 dĪ¼(x) < āˆž. |x|Ī³

This follows from (7.30) by Tonelliā€™s theorem,   āˆž  āˆž Im f (iy) y 1āˆ’Ī³ 1 Ļ€/2 dy = dĪ¼(x) dy = dĪ¼(x).  Ī³ 2 2 y sin(Ī³Ļ€/2) R |x|Ī³ 0 0 R x +y

7.6. Half-plane Poisson kernel and Stieltjes inversion In this section, we consider ways of recovering the measure in the Herglotz representation from the function f . Instead of relying on the Cayley transform and functions on D, it is useful to take a more direct approach. Lemma 7.35. Fix a ā‰„ 0, b āˆˆ R, and a measure Ī¼ which obeys (7.3). The right-hand side of (7.2) deļ¬nes an analytic function on C \ supp Ī¼ which obeys f (ĀÆ z ) = f (z). Proof. The key step is to provide some uniform estimates on the integrand. For any 1 ā‰¤ R < āˆž, if dist(z, supp Ī¼) ā‰„ 1/R and |z| ā‰¤ R, let us prove    1 x  4R3  āˆ’ āˆ€x āˆˆ supp Ī¼. (7.31) ā‰¤  x āˆ’ z 1 + x2  1 + x2 Using (7.23), for |x| > 2R this follows from    1 + xz  1 + R|x| 2R|x|    x āˆ’ z  ā‰¤ |x| āˆ’ |z| ā‰¤ |x|/2 ā‰¤ 4R, and for x āˆˆ supp Ī¼ āˆ© [āˆ’2R, 2R] from    1 + xz  1 + |x||z| 1 + 2RR 3    x āˆ’ z  ā‰¤ |x āˆ’ z| ā‰¤ Rāˆ’1 ā‰¤ 4R . By (7.31) and (7.3), the integral in (7.2) is convergent for each z āˆˆ C \ supp Ī¼, so it deļ¬nes a function f on C \ supp Ī¼. By Moreraā€™s theorem, it suļ¬ƒces to prove that f has zero integral over any closed null-homotopic contour Ī³ in C \ supp Ī¼. For any such contour Ī³, its image Ran Ī³ is a compact subset of C\supp Ī¼ so there exists R such that dist(z, supp Ī¼) ā‰„ 1/R and |z| ā‰¤ R for all z āˆˆ Ran Ī³. Thus, (7.31) implies that Fubiniā€™s theorem can be applied as follows: !     x 1 āˆ’ f (z) dz = 0 dĪ¼(x) = 0 dz dĪ¼(x) = x āˆ’ z 1 + x2 Ī³ R Ī³ R because the integrand is holomorphic in C \ supp Ī¼.



200

7. Herglotz functions

Remark 7.36. Lemma 7.35 will be repeatedly used as part of a method to prove that all Herglotz functions have some property, usually describing the behavior of f near some part of the real line. This property will be obviously additive, i.e., if it holds for two Herglotz functions, it holds for their sum. We will choose a large enough interval [p, q] āŠ‚ R and decompose  1 f (z) = dĪ¼(x) + g(z), [p,q] x āˆ’ z where g consists of all remaining terms in (7.2): !   x x 1 āˆ’ dĪ¼(x) + dĪ¼(x). g(z) = az + b āˆ’ 2 1 + x2 [p,q] 1 + x R\[p,q] x āˆ’ z Since g corresponds to the measure Ļ‡R\[p,q] dĪ¼, by Lemma 7.35 it has an analytic extension to C \ supp(Ļ‡R\[p,q] dĪ¼) āŠƒ C+ āˆŖ (p, q) āˆŖ Cāˆ’ which obeys g(ĀÆ z ) = g(z) and, in particular, has real values on (p, q). Often the desired property is trivial for such functions g, in such cases it remains to  and 1 dĖœ Ī¼(x) with a ļ¬nite, compactly consider Herglotz functions of the form xāˆ’z supported measure dĖœ Ī¼ = Ļ‡[p,q] dĪ¼. Similarly to Proposition 7.31, a point mass in Ī¼ can be computed as a normal or nontangential limit. Note that the following nontangential limit includes as a special case the normal limit z = x0 + i , ā†“ 0: Lemma 7.37. For any Herglotz function f , any x0 āˆˆ R, and Ī“ > 0, Ī¼({x0 }) =

lim

zā†’x0 Ī“ā‰¤arg(zāˆ’x0 )ā‰¤Ļ€āˆ’Ī“

(x0 āˆ’ z)f (z).

(7.32)

Proof. The property (7.32) is additive, in the sense that if it holds for two / supp Ī¼, then f has an Herglotz functions, it holds for their sum. If {x0 } āˆˆ analytic extension at x0 and lim (x0 āˆ’ z)f (z) = 0 Ā· f (x0 ) = 0,

zā†’x0

so f has the property (7.32). By Remark 7.36, it therefore suļ¬ƒces to consider the case when Ī¼ is compactly supported and f is of the form (7.1). Thus, it remains to prove  x0 āˆ’ z dĪ¼(x) = Ī¼({x0 }). (7.33) lim zā†’x0 xāˆ’z Ī“ 0;

202

7. Herglotz functions

(c) For any Ī“ > 0,

 lim ā†“0

R\(āˆ’Ī“,Ī“)

P (s) ds = 0.

(7.35)

Proof. (a) This is immediate from the deļ¬nition. (b) It is elementary to compute  q q p 1 arctan āˆ’ arctan P (s) ds = Ļ€



p for ļ¬nite p, q. By the monotone convergence theorem, this formula holds also for p = āˆ’āˆž and q = +āˆž with the notation arctan(Ā±āˆž) = Ā±Ļ€/2. Using limā†“0 arctan y = Ļ€2 sgn y, we compute for any āˆ’āˆž ā‰¤ p < q ā‰¤ +āˆž, āŽ§ āŽŖ p (DĪ¼)(x), and ļ¬x Ī“ > 0 such that for all t āˆˆ (0, Ī“], C1 Ɨ 2t ā‰¤ Ī¼((x āˆ’ t, x + t)) ā‰¤ C2 Ɨ 2t.

(7.40)

We will prove that   P (x āˆ’ t) dĪ¼(t) ā‰¤ lim sup P (x āˆ’ t) dĪ¼(t) ā‰¤ C2 . C1 ā‰¤ lim inf ā†“0

R

(7.41)

R

ā†“0

Since Ī¼ is ļ¬nite, for any Ī“ > 0, 

P (x āˆ’ t) dĪ¼(t) ā‰¤ P (Ī“)Ī¼(R) ā†’ 0, R\(xāˆ’Ī“,x+Ī“)

ā†“ 0,

so it suļ¬ƒces to prove (7.41) with integrals over (x āˆ’ Ī“, x + Ī“). Since P is even and decreasing on [0, āˆž), the remaining integral can be written as a positive linear combination of the values of Ī¼((x āˆ’ s, x + s)) with s ā‰¤ Ī“: using Tonelliā€™s theorem, we can rewrite  P (x āˆ’ t) dĪ¼(t) (xāˆ’Ī“,x+Ī“)



Ī“

= P (Ī“)Ī¼((x āˆ’ Ī“, x + Ī“)) + 0

Ī¼((x āˆ’ t, x + t))(āˆ’P (t)) dt

and analogously, with Lebesgue measure instead of Ī¼,  Ī“  P (x āˆ’ t) dt = P (Ī“) Ɨ 2Ī“ + 2t(āˆ’P (t)) dt. (xāˆ’Ī“,x+Ī“)

0

206

7. Herglotz functions

Comparing the right-hand sides by using the inequalities (7.40), and rewriting in terms of left-hand sides, implies  (xāˆ’Ī“,x+Ī“) P (x āˆ’ t) dĪ¼(t) C1 ā‰¤  ā‰¤ C2 , (xāˆ’Ī“,x+Ī“) P (x āˆ’ t) dt and taking the limit ā†“ 0 using Lemma 7.39 completes the proof.



Theorem 7.46. If f is a Herglotz function, then the following hold: (a) the limit 1 lim Im f (x + i ) Ļ€ ā†“0 exists Lebesgue-a.e. and Ī¼-a.e. with a value in [0, āˆž]; w(x) =

(b) w(x) < āˆž for Lebesgue-a.e. x; (c) w(x) > 0 for Ī¼-a.e. x; (d) the Radonā€“Nikodym decomposition of Ī¼ with respect to Lebesgue measure is given by dĪ¼ = w dx + dĪ¼s , where dĪ¼s = Ļ‡S dĪ¼, S = wāˆ’1 ({āˆž}). Proof. This follows from the Radonā€“Nikodym decomposition of Ī¼ with respect to Lebesgue measure, diļ¬€erentiation of measures, and the fact that exists, it is equal to the normal wherever the derivative limā†“0 Ī¼((xāˆ’,x+)) (2) boundary value w(x) by Lemma 7.45.  Since the singular part of the measure is supported on the set S = this provides a kind of upper bound on the possible singular part. We emphasize that S = āˆ… does not guarantee that Ī¼s = 0 (Exercise 7.12). Moreover, the normal boundary limit does not necessarily exist for all x āˆˆ R (Exercise 7.13). wāˆ’1 ({āˆž}),

Theorem 7.46 has various generalizations; from normal limits, it can be generalized to nontangential limits. Moreover, while the normal limit of Im f is of special interest, the normal limit of Re f also exists. Proposition 7.47. Let f be a Herglotz function. Then the limit lim f (x + i ) ā†“0

exists and is ļ¬nite for Lebesgue-a.e. x āˆˆ R. āˆš āˆš Proof. Since f and i f are Herglotz functions, their imaginaryāˆšvalues have ļ¬nite normal boundary values Lebesgue-a.e. It follows that f has ļ¬nite boundary values Lebesgue-a.e. Thus, so does f . 

7.7. Pointwise boundary values

207

Corollary 7.48. Let f be a Herglotz function. Then limā†“0 f (x + i ) = 0 for Lebesgue-a.e. x āˆˆ R. Proof. Applying Proposition 7.47 to the Herglotz function āˆ’1/f , we con clude that limā†“0 (āˆ’1/f (x + i )) = āˆž for Lebesgue-a.e. x. This allows us to re-express the Lebesgue decomposition of Ī¼ in terms of boundary values of f on the closure of C+ (viewed as part of the Riemann Ė† this is useful in the context of subordinacy theory. sphere C); Corollary 7.49. If f is a Herglotz function, then (a) as a limit in the Riemann sphere, f (x + i0) := lim f (x + i ) ā†“0

exists Lebesgue-a.e. and Ī¼-a.e. with a value in C+ = C+ āˆŖRāˆŖ{āˆž}; (b) the set S  = {x āˆˆ R | f (x + i0) = āˆž} has zero Lebesgue measure; (c) the Radonā€“Nikodym decomposition of Ī¼ with respect to Lebesgue measure is given by 1 Im f (x + i0) dx + dĪ¼s , Ļ€ where dĪ¼s = Ļ‡S  dĪ¼. dĪ¼ =

Proof. By Proposition 7.47, m(S  ) = 0. Moreover, Ī¼s (S c ) = 0. Combining these statements gives Ī¼(S  \ S) = 0. Since S āŠ‚ S  , this implies Ļ‡S  dĪ¼ = Ļ‡S dĪ¼ = dĪ¼s . For Lebesgue-a.e. x āˆˆ R, f (x + i0) is ļ¬nite and w(x) = Ļ€1 f (x + i0), which implies the new characterization of the absolutely continuous part of Ī¼.  Instead of comparison with Lebesgue measure, this can be generalized to a comparison of two measures and used in conjunction with the Radonā€“ Nikodym theorem (Section 6.2): Theorem 7.50. Let f, g be Herglotz functions corresponding to measures Ī¼, Ī½ on R. Let w be the Radonā€“Nikodym derivative w(x) = lim rā†“0

Ī¼((x āˆ’ r, x + r)) , Ī½((x āˆ’ r, x + r))

which exists Ī¼ + Ī½-a.e. with w(x) āˆˆ [0, āˆž]. Then for (Ī¼ + Ī½)-a.e. x āˆˆ R, lim ā†“0

Im f (x + i ) = w(x). Im g(x + i )

208

7. Herglotz functions

Proof. For (Ī¼ + Ī½)-a.e. x, the limit lim Im(f (x + i ) + g(x + i )) ā†“0

exists and is strictly positive. Thus, by Remark 7.36, it suļ¬ƒces to prove the claim for the case when Ī¼, Ī½ are compactly supported and f, g is of the form   1 1 dĪ¼(x), g(z) = dĪ½(x). f (z) = xāˆ’z xāˆ’z Moreover, by symmetry, it suļ¬ƒces to prove that for every x āˆˆ R, lim sup ā†“0

Im f (x + i ) Ī¼((x āˆ’ r, x + r)) ā‰¤ lim sup . Im g(x + i ) Ī½((x āˆ’ r, x + r)) rā†“0

Fix a constant C > lim sup rā†“0

(7.42)

Ī¼((x āˆ’ r, x + r)) , Ī½((x āˆ’ r, x + r))

and ļ¬x Ī“ > 0 such that for all t āˆˆ (0, Ī“], Ī¼((x āˆ’ t, x + t)) ā‰¤ CĪ½((x āˆ’ t, x + t)).

(7.43)

Then as in the proof of Lemma 7.45, expressing Poisson integrals of Ī¼, Ī½ in terms of values of the measures on intervals, it follows that   P (x āˆ’ t) dĪ¼(t) ā‰¤ C P (x āˆ’ t) dĪ½(t), (xāˆ’Ī“,x+Ī“)

(xāˆ’Ī“,x+Ī“)



which proves (7.42).

In the second part of this section, we discuss the problem of using pointwise boundary behavior of f to study the Ī±-continuous and Ī±-singular parts of the measure, following del Rioā€“Jitomirskayaā€“Lastā€“Simon [26]. The Rogersā€“Taylor decomposition uses the upper Ī±-derivative DĪ¼Ī± (x) = lim sup rā†“0

Ī¼((x āˆ’ r, x + r)) . (2r)Ī±

(7.44)

More precisely, it uses the set of x where DĪ¼Ī± (x) = āˆž. We characterize this set in terms of the quantities QĪ±Ī¼ (x) = lim sup 1āˆ’Ī± Im f (x + i ), ā†“0

RĪ¼Ī± (x)

= lim sup 1āˆ’Ī± |f (x + i )|. ā†“0

Theorem 7.51. For any Herglotz function f , any Ī± āˆˆ [0, 1) and x āˆˆ R, {x āˆˆ R | DĪ¼Ī± (x) = āˆž} = {x āˆˆ R | QĪ±Ī¼ (x) = āˆž} = {x āˆˆ R | RĪ¼Ī± (x) = āˆž}.

7.7. Pointwise boundary values

209

Proof. We will prove this by proving three implications DĪ¼Ī± (x) = āˆž =ā‡’ QĪ±Ī¼ (x) = āˆž =ā‡’ RĪ¼Ī± (x) = āˆž =ā‡’ DĪ¼Ī± (x) = āˆž. Starting with the inequality

1 Ļ‡ (t) ā‰¤ 2 (xāˆ’,x+) (t āˆ’ x)2 + 2 and integrating with respect to dĪ¼(t) gives Ī¼((x āˆ’ , x + )) ā‰¤ Im f (x + i ). Multiplying by 1āˆ’Ī± and taking ā†’ 0 gives DĪ¼Ī± (x) ā‰¤ 2QĪ±Ī¼ (x), which proves the ļ¬rst implication. The trivial observation Im f (x+i ) ā‰¤ |f (x+i )| implies QĪ±Ī¼ (x) ā‰¤ RĪ¼Ī± (x), which proves the second implication. The third implication is trivial for Ī± = 0: by Lemma 7.38, RĪ¼0 (x) < āˆž for all x. Let us assume Ī± āˆˆ (0, 1) and let x obey DĪ¼Ī± (x) < āˆž. Then there exists C < āˆž such that Ī¼((x āˆ’ Ī“, x + Ī“)) ā‰¤ CĪ“ Ī±

āˆ€Ī“ āˆˆ (0, 1].

By the standard trick (Remark 7.36) we can assume that Ī¼ is supported on (x āˆ’ 1, x + 1). Then, by Tonelliā€™s theorem,  āˆž  dĪ¼(t) . = Ī¼((x āˆ’ Ļ„ (y), x + Ļ„ (y)) dy, |f (x + i )| ā‰¤ (t āˆ’ x)2 + 2 0 (xāˆ’1,x+1) where Ļ„ (y) = min{1, y āˆ’2 āˆ’ 2 }. The important thing is that this integral depends only on the value of the measure on intervals (x āˆ’ Ī“, x + Ī“), so we estimate it by comparison with the measure CĪ± Ļ‡ |t āˆ’ x|Ī±āˆ’1 dt. 2 (xāˆ’1,x+1) This measure has the property Ī½((x āˆ’ Ī“, x + Ī“)) = CĪ“ Ī± for Ī“ āˆˆ (0, 1] so   āˆž dĪ½(t) . Ī½((x āˆ’ Ļ„ (y), x + Ļ„ (y)) dy = . |f (x + i )| ā‰¤ (t āˆ’ x)2 + 2 0 (xāˆ’1,x+1) dĪ½(t) =

Multiplying by 1āˆ’Ī± and using symmetry and t = x + v gives  1  1/ Ī±āˆ’1 |t āˆ’ x|Ī±āˆ’1 v dv 1āˆ’Ī± 1āˆ’Ī± . āˆš . |f (x + i )| ā‰¤ CĪ±

dt = CĪ±

v2 + 1 (t āˆ’ x)2 + 2 0 0  āˆž Ī±āˆ’1 dv < āˆž, it follows that RĪ¼Ī± (x) < āˆž, which proves the third Since 0 vāˆšv2 +1 implication.  The exclusion of Ī± = 1 in the previous theorem was necessary; for a Herglotz function, it is possible to have convergence of Im f (i ) as ā†’ 0 and divergence of |f (i )| (Exercise 7.14).

210

7. Herglotz functions

7.8. Meromorphic Herglotz functions In spectral theory, we often encounter extensions of Herglotz functions to domains larger than C+ , with the reļ¬‚ection symmetry f (ĀÆ z ) = f (z).

(7.45)

These extensions can be analytic or even meromorphic, with possible poles on the real line. The domain of such an extension is related to the support of the measure and to the essential support of Ī¼, denoted ess supp Ī¼, deļ¬ned as the set of nonisolated points of supp Ī¼. Proposition 7.52. Let f : C+ ā†’ C+ be the Herglotz function with the Herglotz representation (7.2) for z āˆˆ C+ . Then f extends to (a) an analytic function f : C \ supp Ī¼ ā†’ C with the property (7.45); Ė† with the property (b) a meromorphic function f : C \ ess supp Ī¼ ā†’ C (7.45). Proof. (a) This was already proved as Lemma 7.35. (b) If supp Ī¼ has an isolated point Ī», the measure can be decomposed / supp Ī½. By the Herglotz representation, as dĪ¼ = Ī¼({Ī»})Ī“Ī» + dĪ½ where Ī» āˆˆ ! !  Ī» 1 x 1 āˆ’ + az + b + āˆ’ dĪ½(x). f (z) = Ī¼({Ī»}) Ī» āˆ’ z Ī»2 + 1 x āˆ’ z x2 + 1 The ļ¬rst term has a simple pole at Ī» with residue āˆ’Ī¼({Ī»}) and all other terms are analytic in a neighborhood of Ī».  Note that our proof also shows: Ė† is a Corollary 7.53. Any isolated singularity Ī» of f : C \ ess supp Ī¼ ā†’ C simple pole and the residue of f at Ī» is strictly negative. It is also important to know a kind of converse to Proposition 7.52; namely, that f cannot be extended analytically or meromorphically to any domain not contained in C \ supp Ī¼ or C \ ess supp Ī¼, respectively: Lemma 7.54. Let f have the Herglotz representation (7.2) for z āˆˆ C+ . (a) If f extends to an analytic function f : C+ āˆŖ (p, q) āˆŖ Cāˆ’ with the property (7.45), then supp Ī¼ āˆ© (p, q) = āˆ…. (b) If f extends to a meromorphic function f : C+ āˆŖ (p, q) āˆŖ Cāˆ’ with the property (7.45), then ess supp Ī¼ āˆ© (p, q) = āˆ…. Proof. (a) By (7.45), f is real valued on (p, q). Ī¼((p, q)) = 0.

By Proposition 7.43,

7.8. Meromorphic Herglotz functions

211

(b) By (a), the set supp Ī¼ āˆ© (p, q) can contain only poles of f . Poles of a meromorphic function are isolated, so supp Ī¼ āˆ© (p, q) has no accumulation points in (p, q); thus, ess supp Ī¼ āˆ© (p, q) = āˆ….  In this text, we will usually consider extensions which obey (7.45). This will avoid some complications normally associated āˆš with analytic extensions. For āˆš instance, the Herglotz function f (z) = āˆ’ āˆ’z, deļ¬ned on C+ so that arg(āˆ’ āˆ’z) āˆˆ (0, Ļ€/2), has analytic extensions to C \ [0, āˆž) and to C \ (āˆ’āˆž, 0], but only the ļ¬rst of those obeys (7.45). Deļ¬nition 7.55. We call the function f given by (7.2) on C \ supp Ī¼ an analytic Herglotz function, and we call C \ supp Ī¼ its domain of analyticity. We call the function f given by (7.2) on C \ ess supp Ī¼ an analytic Herglotz function, and we call C \ ess supp Ī¼ its domain of analyticity. Ė† be a meromorphic Herglotz function. At Proposition 7.56. Let f : Ī© ā†’ C any point x āˆˆ R āˆ© Ī© which is not a pole of f , f  (x) > 0. Proof. If f (z) is a meromorphic Herglotz function on Ī©, then so is f (z) āˆ’ f (x) and so is g(z) = āˆ’1/(f (z) āˆ’ f (x)). The function g has an isolated singularity at x. By Corollary 7.53, that singularity is a simple pole with strictly negative residue, so f (z) āˆ’ f (x) has a simple zero at x with strictly positive derivative.  Proposition 7.56 can also be proved by deriving a formula for f  (Exercise 7.15). Together, Proposition 7.56 and Corollary 7.53 describe the behavior of f on intervals (p, q) āŠ‚ R āˆ© Ī©. The function f is strictly increasing, except at poles, where it has vertical asymptotes. Qualitatively, this resembles the graph of the tangent function. Given two discrete sets A, B āŠ‚ R, we say that they strictly interlace if the following conditions hold: (a) A āˆ© B = āˆ…. (b) For any x, y āˆˆ A with x < y, there exists t āˆˆ B āˆ© (x, y). (c) For any x, y āˆˆ B with x < y, there exists t āˆˆ A āˆ© (x, y). Ė† be a meromorphic Herglotz function. On Proposition 7.57. Let f : Ī© ā†’ C any interval I āŠ‚ R āˆ© Ī©, the sets {x āˆˆ I | f (x) = u} and {x āˆˆ I | f (x) = v} strictly interlace in I for any u, v āˆˆ R āˆŖ {āˆž} with u = v. Proof. If v = āˆž, by replacing f by āˆ’1/(f āˆ’ v), we reduce to the case of v = āˆž. Likewise, by then subtracting u, we reduce to the case u = 0, so it suļ¬ƒces to prove that zeros and poles of f strictly interlace in I.

212

7. Herglotz functions

The set of poles of a nonconstant meromorphic function has no accumulation points in the domain. If p < q are two consecutive poles of f on I, then f : (p, q) ā†’ R is strictly increasing. Since poles are simple and have negative residue, lim f (x) = āˆ’āˆž, xā†“p

lim f (x) = +āˆž, xā†‘q

so f : (p, q) ā†’ R is a bijection and f has a zero in (p, q). Thus, f has a zero between any two poles on I. Applying the same argument to the meromorphic Herglotz function āˆ’1/f shows that between any two zeros of f there is a pole of f , which completes the proof.  In the previous proofs, we used the convenient observation that the maps z ā†’ āˆ’1/z and z ā†’ z āˆ’ u preserve C+ , so they preserve meromorphic Herglotz functions. The natural generality for that observation follows. Ė† is a meromorphic Herglotz function and Corollary 7.58. If f : Ī© ā†’ C A āˆˆ SL(2, R), then the function g deļ¬ned by ! ! g(z) f (z) $A 1 1 is also a meromorphic Herglotz function with the same domain Ī©. If A is upper triangular, the functions f and g have the same poles; otherwise, on any interval in Ī© āˆ© R, poles of f and g strictly interlace. Proof. Since action by A preserves C+ , it maps f to another meromorphic Herglotz function g. The result for poles follows from ! ! a11 g(z) 1 .  f (z) = āˆž ā‡ā‡’ $A ā‡ā‡’ g(z) = 1 0 a21 In spectral theory, action by a rotation matrix ! cos Ļ† āˆ’ sin Ļ† A= āˆˆ SL(2, R) sin Ļ† cos Ļ† will correspond to a change of boundary condition for a half-line SchrĀØodinger operator and poles correspond to its discrete spectrum.

7.9. Exponential Herglotz representation Let us ļ¬x the branch of log such that 0 < Im log z < Ļ€ for z āˆˆ C+ . Then, if f is a Herglotz function, so is log f . Applying the Herglotz representation to log f provides a very useful multiplicative representation for f .

7.9. Exponential Herglotz representation

213

Theorem 7.59 (Exponential Herglotz representation). Let f be a Herglotz function. Then the limit Ī¾(x) =

1 lim Im log f (x + i ) āˆˆ [0, 1] Ļ€ ā†“0

(7.46)

exists for Lebesgue-a.e. x āˆˆ R, and there exists a constant k āˆˆ R such that !  x 1 log f (z) = k + āˆ’ Ī¾(x) dx. 1 + x2 R xāˆ’z Proof. Since log f (z) is a Herglotz function, it has a Herglotz representation !  1 x dĪ¼(x). āˆ’ log f (z) = az + b + 1 + x2 R xāˆ’z Since 0 < Im log f < Ļ€, a = lim

yā†’āˆž

Im log f (iy) = 0. y

For the same reason, for any c < d and > 0,  1 d Im log f (x + i ) dx ā‰¤ d āˆ’ c, Ļ€ c so by Stieltjes inversion, 1 (Ī¼((c, d)) + Ī¼([c, d])) ā‰¤ d āˆ’ c. 2 Taking d ā†“ c implies that Ī¼ has no pure points, so Ī¼((c, d)) ā‰¤ d āˆ’ c for all c < d. Denoting Lebesgue measure by |Ā·|, this implies that Ī¼(A) ā‰¤ |A| for all open intervals, then for all open sets, and ļ¬nally for all Borel sets (by outer regularity). By the Radonā€“Nikodym theorem, dĪ¼ = Ī¾(x) dx for some Borel function Ī¾ with 0 ā‰¤ Ī¾ ā‰¤ 1. Finally, Ī¾ is reconstructed from normal boundary values of Im log f by Theorem 7.46.  Lemma 7.60. If kn ā†’ k and Ī¾n ā†’ Ī¾ pointwise Lebesgue-a.e., then fn ā†’ f uniformly on compacts. Proof. By Proposition 7.28, it suļ¬ƒces to prove that for all h āˆˆ C0 (R),   dx dx ā†’ h(x)Ī¾(x) . h(x)Ī¾n (x) 1 + x2 1 + x2 This follows from dominated convergence with the dominating function  |h(x)|/(1 + x2 ).

214

7. Herglotz functions

An important special case of Theorem 7.59 is when Ī¾ is piecewise constant. In that case, the piecewise integrals can be computed, by using the elementary calculation āˆš !  d x 1 (d āˆ’ z)/ d2 + 1 āˆš . (7.47) āˆ’ dx = ln x āˆ’ z 1 + x2 (c āˆ’ z)/ c2 + 1 c Then the integral turns into a sum, and exponentiating turns that into a product formula for f (z). We give some examples and leave others as exercises: Example 7.61. Let f be a Herglotz function with a meromorphic extension to C with the symmetry (7.45). Assume that it has zeros (Ī»n )āˆž n=1 and poles such that p < Ī» < p for all n āˆˆ N. Then f is of the form (pn )āˆž n n n+1 n=1 . āˆž  (Ī»n āˆ’ z)/ Ī»2n + 1 . (7.48) f (z) = C 2 n=1 (pn āˆ’ z)/ pn + 1 for some C > 0. Proof. The sets of zeros and poles are discrete, so pn ā†’ āˆž, Ī»n ā†’ āˆž as n ā†’ āˆž. The function f is strictly increasing between poles, so Ī¾(x) = 1 if x āˆˆ (pn , Ī»n ) for some n and Ī¾(x) = 0 otherwise. The exponential Herglotz representation can be integrated piecewise to give . āˆž  (Ī»n āˆ’ z)/ Ī»2n + 1 . ln log f (z) = k + (p āˆ’ z)/ p2n + 1 n n=1 and exponentiating gives (7.48) with C = ek > 0.



The square roots in (7.48) are z-independent but their presence ensures a convergent product; compare Exercise 7.19. Example 7.62. Let f be a Herglotz function with a meromorphic extension to C with the symmetry (7.45). Assume that it has zeros (Ī»n )āˆž n=1 and poles āˆž (pn )n=1 such that Ī»n < pn < Ī»n+1 for all n āˆˆ N. Then f is of the form 8 āˆž (Ī» āˆ’ z)/ Ī»2n+1 + 1  n+1 Ī»1 āˆ’ z . f (z) = āˆ’C . 2 Ī»1 + 1 n=1 (pn āˆ’ z)/ p2n + 1 for some C > 0. Proof. This follows by applying Example 7.61 to the meromorphic Herglotz function āˆ’1/f and a telescoping argument to rearrange the inļ¬nite product, or directly by computing the exponential Herglotz representation for f . 

7.10. The PhragmĀ“enā€“LindelĀØ of method and asymptotic expansions

215

7.10. The PhragmĀ“ enā€“LindelĀØ of method and asymptotic expansions The PhragmĀ“enā€“LindelĀØ of method is a technique for bounding the values of an analytic function in an (often unbounded) domain in C in terms of bounds on its boundary values and growth rates. We present one special case which will have important consequences for Herglotz function asymptotics. Theorem 7.63 (PhragmĀ“enā€“LindelĀØ of). Let Ī© = {z āˆˆ C | Ī± < arg z < Ī²}. ĀÆ āŠ‚ C, and If h : Ī© ā†’ C is analytic on Ī©, has a continuous extension to Ī© there exist C1 , C2 > 0 and Ī· < Ļ€/(Ī² āˆ’ Ī±) such that |h(z)| ā‰¤ C1 eC2 |z| , Ī·

(7.49)

then h is bounded on Ī© and sup|h(z)| = sup |h(z)|. zāˆˆĪ©

zāˆˆāˆ‚Ī©

Proof. By composing with a power z ā†’ eiĻ† z Īŗ for suitable Ļ† and Īŗ > 0, we can reduce to the case Ī· < 1 < Ļ€/(Ī² āˆ’ Ī±) and āˆ’Ī± = Ī² āˆˆ (0, Ļ€/2). On that domain, Re z ā‰„ |z| cos Ī² > 0 so for any > 0, the function h (z) = h(z)eāˆ’z ĀÆ with is analytic on Ī©, with a continuous extension to Ī© sup |h (z)| ā‰¤ sup |h(z)|. zāˆˆāˆ‚Ī©

Moreover, by

|eāˆ’z |

=

eāˆ’ Re z

zāˆˆāˆ‚Ī©

ā‰¤

eāˆ’ cos Ī²|z| ,

(7.49) implies that

lim h (z) = 0.

zā†’āˆž ĀÆ zāˆˆĪ©

Ė† Thus, by the maximum principle applied to the closure of Ī© in C, sup|h (z)| = sup |h (z)| ā‰¤ sup |h(z)|. zāˆˆĪ©

zāˆˆāˆ‚Ī©

zāˆˆāˆ‚Ī©

As ā†’ 0, h (z) ā†’ h(z), and the claim follows.



When a Herglotz function has an explicit nontangential asymptotic expansion and a meromorphic continuation through the negative half-line, the PhragmĀ“enā€“LindelĀØ of method can often be used to extend that expansion through the negative half-line. We formulate a criterion: Corollary 7.64. Let f be an analytic Herglotz function on C \ [c, āˆž), and z ) = g(z). let g be an analytic function on C \ [c, āˆž) with the symmetry g(ĀÆ Assume that n, Ī³ > 0 are such that for all Ī“ > 0, g(z) = O(|z|n),

z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]

216

7. Herglotz functions

and

f (z) = g(z) + O(|z|āˆ’Ī³ ), Then for all Ī“ > 0,

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“].

f (z) = g(z) + O(|z|āˆ’Ī³ ),

z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“].

Proof. By shifting z by a real constant, we can assume that c = 1. Such shifts do not aļ¬€ect the nontangential limits in the hypotheses and conclusions. We assume Ī“ āˆˆ (0, Ļ€) from now on. By symmetry, since the asymptotic behavior holds for arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“], it holds also for arg z āˆˆ [Ļ€ + Ī“, 2Ļ€ āˆ’ Ī“], so it suļ¬ƒces to extend it into the sector arg z āˆˆ [Ļ€ āˆ’ Ī“, Ļ€ + Ī“]. The Herglotz representation for f has the form  1 + xz 1 dĪ¼(x), supp Ī¼ āŠ‚ [1, āˆž). f (z) = az + b + x āˆ’ z 1 + x2 When Im z < 0 and x ā‰„ 1, x ā‰¤ |x āˆ’ z|, so  1 + |z| dĪ¼(x). |f (z)| ā‰¤ a|z| + |b| + 1 + x2 In particular, f (z) = O(|z|) as z ā†’ āˆž, arg z āˆˆ [Ļ€ āˆ’ Ī“, Ļ€ + Ī“]. Thus, the function h(z) = z Ī³ (f (z) āˆ’ g(z)) obeys all the hypotheses of Theorem 7.63 with Ī© = {z āˆˆ C | Ļ€ āˆ’ Ī“ < arg z < Ļ€ + Ī“}: In particular, it has a continuous ĀÆ with h(0) = 0, since f, g are bounded at 0. Thus, h is bounded extension to Ī© ĀÆ  on Ī©, which implies f (z) = g(z) + O(|z|āˆ’Ī³ ) for Ļ€ āˆ’ Ī“ ā‰¤ arg z ā‰¤ Ļ€ + Ī“.

7.11. Matrix-valued Herglotz functions In this section, we consider two generalizations. We begin by considering a generalization of the Herglotz representation to complex measures and then study matrix-valued Herglotz functions. Matrix-valued Herglotz functions naturally appear in spectral theory; Exercise 7.21 illustrates this. Complex measures are linear combinations with complex coeļ¬ƒcients of (positive) ļ¬nite measures. To avoid issues associated with inļ¬nite measures, here we will work with the alternative Herglotz representation  1 + xz dĪ½(x). (7.50) f (z) = az + b + R xāˆ’z In this generalization, to be able to recover the measure from the function, the function should be considered on the domain C \ R instead of C+ . Lemma 7.65. Let a, b āˆˆ C, and let Ī½ be a complex measure on R. If f is deļ¬ned on C \ R by (7.50), then f (iy) , yā†’āˆž iy

a = lim

b=

f (i) + f (āˆ’i) , 2

(7.51)

7.11. Matrix-valued Herglotz functions

217

for any x0 āˆˆ R and Ī“ > 0, (1 + x20 )Ī½({x0 }) =

lim

zā†’x0 Ī“ 0, a=

lim

zā†’āˆž Ī“ 0}; then [dĪ¼u ] = [Ļ‡S dĪ¼]. In particular, u, Ļ‡R\S (A)u = Ī¼u (R \ S) = 0, and therefore u āˆˆ Ker Ļ‡R\S (A) = Ran Ļ‡S (A). Pick w such that Ī¼w is a maximal spectral measure for A and let v = u + Ļ‡R\S (A)w.

(9.26)

Applying Ļ‡S (A) to (9.26) gives Ļ‡S (A)v = u, so u āˆˆ CA (v). The two summands in (9.26) lie in mutually orthogonal invariant subspaces Ran Ļ‡S (A) and Ran Ļ‡R\S (A), so for any Borel set B, Ī¼v (B) = v, Ļ‡B (A)v = u, Ļ‡B (A)u + w, Ļ‡B\S (A)w = Ī¼u (B) + Ī¼w (B \ S). This is equal to zero if and only if Ī¼(B āˆ© S) = Ī¼(B \ S) = 0, so if and only if Ī¼(B) = 0. Thus, [Ī¼v ] = [Ī¼].  Proposition 9.34. Let A āˆˆ L(H) be self-adjoint. There exists a spectral basis (Ļˆj )N j=1 for A such that Ī¼Ļˆj ' Ī¼Ļˆj+1 for all j < N . Proof. This proof relies on the notation and method from the proof of Lemma 5.45. As in that construction, we will start with an orthonormal basis (Ļ†j )N j=1 and deļ¬ne inductively a sequence of vectors Ļˆn , together with a decreasing sequence of subspaces Vn given by V0 = H and Vn =

n

CA (Ļˆj )āŠ„ .

j=1

However, now the Ļˆn are deļ¬ned diļ¬€erently: For any n, we ļ¬rst deļ¬ne un as the orthogonal projection of Ļ†n onto Vnāˆ’1 . Then, using Lemma 9.33, we deļ¬ne Ļˆn to be a vector such that un āˆˆ CA (Ļˆn ) and that Ī¼Ļˆn is a maximal spectral measure for A|Vnāˆ’1 .

9.6. Spectral multiplicity

285

As in the proof of Lemma 5.45, the cyclic subspaces CA (Ļˆj ) are mutually orthogonal by construction, and since un āˆˆ CA (Ļˆn ), it follows that H=

N 3

CA (Ļˆj ).

j=1

Moreover, since Vn are a decreasing sequence of subspaces and Ī¼Ļˆn are maximal spectral measures for A|Vnāˆ’1 , it follows that Ī¼Ļˆnāˆ’1 ' Ī¼Ļˆn . Finally, we note that any zero elements can be removed from the se quence (Ļˆn )N n=1 , and the remaining elements can be normalized. Lemma 9.35. If [Ī¼] = [Ī½] and g āˆˆ Bb (R), then Tg,dĪ¼ āˆ¼ = Tg,dĪ½ . Proof. By Radonā€“Nikodym, Ī½ " Ī¼ implies that dĪ½ = hdĪ¼ with h ā‰„ 0 Ī¼-a.e. This representation implies Ī½({x | h(x) = 0}) = 0, so since Ī½ ' Ī¼, we conclude Ī¼({x | h(x) = 0}) = 0. Thus, dĪ½ = hdĪ¼ with h > 0 Ī¼-a.e. The map U : L2 (dĪ¼) ā†’ L2 (dĪ½) given by U f = hāˆ’1/2 f is unitary and obeys  U āˆ’1 Tg(x),dĪ½(x) U = Tg(x),dĪ¼(x) . Lemma 9.36. If a measure dĪ½ is decomposed as a ļ¬nite or countable sum dĪ½ =

N 

dĪ½j

(9.27)

j=1

4N of mutually singular measures on R, then Tg,dĪ½ āˆ¼ = j=1 Tg,dĪ½j . 4N 2 N Proof. The map U : L2 (dĪ½) ā†’ j=1 L (dĪ½j ) given by U f = (f )j=1 is norm-preserving because of (9.27) and onto because the Ī½j are mutually 4  singular, so it is unitary. Moreover, U Tg,dĪ½ U āˆ’1 = N j=1 Tg,dĪ½j . Proof of Theorem 9.31(a). Using the spectral basis from Proposition 9.34 and adding trailing zero measures if necessary to make the sequence inļ¬nite, āˆž 3 āˆ¼ Tx,dĪ½j (x) A= j=1

for a sequence of measures obeying Ī½j+1 " Ī½j for all j. The condition Ī½j+1 " Ī½j implies existence of Borel sets Mj such that [dĪ½j+1 ] = [Ļ‡Mj dĪ½j ]. Introducing the mutually disjoint sets māˆ’1  Mi \ Mm , Sm = i=1

Sāˆž =

āˆž j=1

Mj

286

9. Consequences of the spectral theorem

allows us to represent the decreasing sequence of sets M1 , M1 āˆ© M2 , M1 āˆ© M2 āˆ© M3 , . . . as jāˆ’1  Mi = Sm i=1

māˆˆNāˆŖ{āˆž} mā‰„j

 (we use the convention 0i=1 Mi = R). Therefore, the measures Ī¼m = Ļ‡Sm dĪ¼Ļˆ1 are mutually singular by construction and < ;  Ī¼m . [Ī½j ] = māˆˆNāˆŖ{āˆž} mā‰„j

Therefore, by Lemma 9.36, Aāˆ¼ =

3

3

Tx,dĪ¼m (x) ,

jāˆˆN māˆˆNāˆŖ{āˆž} mā‰„j

which, after changing the order of summation, is the same as (9.25).



Proving uniqueness also requires some preliminary lemmas. Lemma 9.37. If A āˆ¼ = B, then for any Borel set S, A|Ran Ļ‡ (A) āˆ¼ = B|Ran Ļ‡ (B) . S

S

Proof. By the uniqueness of functional calculus, U Ļ‡S (A)U āˆ’1 = Ļ‡S (B). Since v āˆˆ Ran Ļ‡S (B) if and only if Ļ‡S (B)v = v, this implies that U is a bijection from Ran Ļ‡S (A) to Ran Ļ‡S (B). Thus, the restriction of U to  Ran Ļ‡S (A) conjugates A|Ran Ļ‡S (A) to B|Ran Ļ‡S (B) . This lemma will be applied to multiplication operators; note that if A = Tx,dĪ¼(x) , then Ran Ļ‡S (A) is naturally identiļ¬ed with L2 (Ļ‡S dĪ¼) so A|Ran Ļ‡ (A) āˆ¼ = Tx,Ļ‡ (x) dĪ¼(x) . S

S

Lemma 9.38. Assume that Ī¼, Ī½ are not the zero measure and that m n 3 3 Tx,dĪ¼(x) āˆ¼ Tx,dĪ½(x) . = i=1

i=1

Then m = n and [Ī¼] = [Ī½]. Proof. These operators have maximal spectral measures Ī¼, Ī½ respectively, so their unitary equivalence implies [Ī¼] = [Ī½]. Using Tx,dĪ¼(x) āˆ¼ = Tx,dĪ½(x) , m 3 i=1

Tx,dĪ¼(x)

āˆ¼ =

n 3 i=1

Tx,dĪ½(x)

āˆ¼ =

n 3

Tx,dĪ¼(x) .

i=1

By symmetry, it remains to show that m > n would lead to a contradiction.

9.6. Spectral multiplicity

287

4m 2 4n 2 Assume that m > n and that U : i=1 L (dĪ¼) ā†’ i=1 L (dĪ¼) is a unitary map such that  m  n 3 3 Tx,dĪ¼(x) U āˆ’1 = Tx,dĪ¼(x) . U i=1

i=1

By uniqueness of the Borel functional calculus, for any Borel set S,  m  n 3 3 TĻ‡S ,dĪ¼ U āˆ’1 = TĻ‡S ,dĪ¼ . U i=1

i=1 Cm .

Let v1 , . . . , vn+1 be an orthonormal set View v1 , . . . , vn+1 as con4in m 2 stant functions of x, so as elements of i=1 L (dĪ¼), and deļ¬ne fj = U vj āˆˆ 4 n 2 i=1 L (dĪ¼). Using unitarity of U and computing inner products in both Hilbert spaces, for any Borel set S and i, j = 1, . . . , n + 1,   āˆ— fi (x) fj (x) dĪ¼(x) = viāˆ— vj dĪ¼ = Ī“ij Ī¼(S). S

S

(x)āˆ— f

It follows that fi j (x) = Ī“ij for Ī¼-a.e. x. Thus, for Ī¼-a.e. x, the vectors f1 (x), . . . , fn+1 (x) are an orthonormal set in Cn , which is a contradiction.  Proof of Theorem 9.31(b). Since Ī¼m are mutually singular measures, there exists a partition of R into Borel sets Sm such that each Ī¼m is sup such that each Ī½ ported on Sm , and analogously a partition into sets Sm m  . By Lemma 9.37 applied to S āˆ© S  , we obtain is supported on Sm m n m 3 j=1

Tx,Ļ‡S (x)dĪ¼m (x) āˆ¼ = n

n 3

Tx,Ļ‡Sm (x)dĪ½n (x) .

(9.28)

j=1

For m = n, this implies Ļ‡Sn dĪ¼m and Ļ‡Sm dĪ½n are zero measures. Thus, each  , and each Ī½ Ī¼m is supported on Sm m is supported on Sm . Thus, by (9.28) applied to m = n, m m 3 3 āˆ¼ Tx,dĪ¼m (x) = Tx,dĪ½m (x) , j=1

which implies [Ī¼m ] = [Ī½m ] for all m.

j=1



In practice, if an operator is already represented in terms of multiplication operators, we would not retrace the above proofs (constructing a spectral basis, etc.) in order to determine the decomposition into spectral multiplicities. Instead, we may directly manipulate the operator into the form (9.25) and appeal to uniqueness. We illustrate this with two examples. Example 9.39. Let A be a self-adjoint operator on H. If A has a cyclic vector Ļˆ, then it has only multiplicity 1 spectrum: Ī¼1 = Ī¼Ļˆ and Ī¼m = 0 for all m ā‰„ 2. Conversely, if Ī¼m = 0 for all m ā‰„ 2, then A has a cyclic vector.

288

9. Consequences of the spectral theorem

Proof. Since Ļˆ is a cyclic vector, A āˆ¼ = Tx,dĪ¼Ļˆ (x) . This is already in the form (9.25) with Ī¼1 = Ī¼Ļˆ and Ī¼m = 0 for all m ā‰„ 2, so by uniqueness of this representation, these are the multiplicity m measures. Conversely, if Ī¼m = 0 for all m ā‰„ 2, then (9.25) simpliļ¬es to A āˆ¼ =

Tx,dĪ¼1 (x) . More precisely, there exists a unitary U : L2 (R, dĪ¼1 ) ā†’ H such that U āˆ’1 AU = Tx,dĪ¼1 (x) . Since the constant function 1 āˆˆ L2 (R, dĪ¼1 ) is  cyclic for Tx,dĪ¼1 (x) , U 1 āˆˆ H is cyclic for A.

Example 9.40. Denote by A the operator of multiplication by 2 cos k on dk ). Its multiplicity 2 measure is dĪ¼2 (x) = Ļ‡(āˆ’2,2) (x) dx, and L2 ([0, 2Ļ€], 2Ļ€ Ī¼n = 0 for all n = 2. Proof. By Lemma 9.36 applied to the decomposition dk dk dk = Ļ‡[0,Ļ€] (k) + Ļ‡[Ļ€,2Ļ€] (k) , 2Ļ€ 2Ļ€ 2Ļ€ A is unitarily equivalent to the direct sum of multiplications by 2 cos k on dk L2 ([(n āˆ’ 1)Ļ€, nĻ€], 2Ļ€ ), n = 1, 2. Ļ‡[0,2Ļ€] (k)

On each interval [(n āˆ’ 1)Ļ€, nĻ€], the map g(k) = 2 cos k is strictly monotone with image [āˆ’2, 2], so by a change of variables,  2  nĻ€ 1 2 dk = |f (k)| |f (g āˆ’1 (Ī»))|2 |(g āˆ’1 ) (Ī»)| dĪ». 2Ļ€ 2Ļ€ āˆ’2 (nāˆ’1)Ļ€ Therefore, with the choice of measure Ļ‡(āˆ’2,2) (Ī») Ļ‡(āˆ’2,2) (Ī») āˆ’1  |(g ) (Ī»)| dĪ» = āˆš dĪ», 2Ļ€ 2Ļ€ 4 āˆ’ Ī»2   dk the maps Un : L2 [(n āˆ’ 1)Ļ€, nĻ€], 2Ļ€ ā†’ L2 ([āˆ’2, 2], dĪ½(Ī»)) given by Un f = āˆ’1 f ā—¦ g are unitary. dĪ½(Ī») =

Thus, the unitary map U1 āŠ• U2 conjugates A to the operator TĪ»,dĪ½(Ī») āŠ• TĪ»,dĪ½(Ī») . Since dĪ½ is mutually absolutely continuous with Ļ‡(āˆ’2,2) (x) dx, Lemma 9.35 completes the proof.  Exercise 9.13 demonstrates how this notion of multiplicity generalizes the notion of multiplicity of eigenvalues. Exercise 9.14 characterizes cyclic vectors in the multiplicity 1 case, and Exercise 9.15 generalizes Example 9.39 to the case when there is a ļ¬nite spectral basis. Exercise 9.16 introduces a decomposition of the Hilbert space into multiplicity m subspaces for A, and Exercise 9.17 gives an interpretation of that decomposition in terms of the minimal number of cyclic subspaces needed to cover a subspace of the form Ļ‡S (A). We conclude this section by showing how to read oļ¬€ spectral multiplicity for multiplication operators on vector-valued L2 spaces introduced in Lemma 6.38:

9.7. Stoneā€™s theorem

289

Proposition 9.41. Let W dĪ¼ be as in Lemma 6.38. The operator A of multiplication by x on L2 (R, Cd , W (x) dĪ¼(x)) has the following properties: (a) A has maximal spectral measure Ī¼. (b) Denoting Sm = {x | rank W (x) = m}, the multiplicity m measure for A is dĪ¼m = Ļ‡Sm dĪ¼; in particular, Ī¼m = 0 for m > d. Proof. Since W ā‰„ 0, we can diagonalize W (x) = U (x)āˆ’1 D(x)U (x) with U (x) unitary and D(x) diagonal matrices, D(x) = diag(Ī»1 (x), . . . , Ī»d (x)),

Ī»1 ā‰„ Ā· Ā· Ā· ā‰„ Ī»d ā‰„ 0.

Moreover, U (x) and D(x) can be chosen as Borel functions of x, since W (x) is Borel. Thus, the map (U f )(x) = U (x)f (x) is a unitary map U : L2 (R, Cd , W dĪ¼) ā†’ L2 (R, Cd , D dĪ¼). Since D is diagonal, viewing Cd -valued functions as vectors of scalar functions gives L2 (R, Cd , D dĪ¼) =

d 3

L2 (R, Ī»k dĪ¼).

k=1

Therefore Aāˆ¼ =

d 3

Tx,Ī»k (x) dĪ¼(x) .

k=1

This representation is in the form of that in Proposition 9.34, so the claims follow as in the proof of Theorem 9.31, since rank W is the number of nonzero eigenvalues of W .  It is also common to combine the decomposition by multiplicity with the decomposition by spectral type, by decomposing the multiplicity m measures Ī¼m instead of the maximal spectral measure. For instance, if we say that the singular spectrum of some operator A has multiplicity 1, we mean that (Ī¼m )s = 0 for m ā‰„ 2.

9.7. Stoneā€™s theorem Stoneā€™s theorem expresses spectral projections in terms of resolvents. The proof will be based on functional calculus and calculations related to the Stieltjes inversion formula.

290

9. Consequences of the spectral theorem

Theorem 9.42 (Stone). Let A be a self-adjoint operator, and let c < d be real numbers. Then  d 1 1 (Ļ‡(c,d) (A) + Ļ‡[c,d] (A)) = s-lim ((A āˆ’ t āˆ’ i )āˆ’1 āˆ’ (A āˆ’ t + i )āˆ’1 ) dt. ā†“0 2Ļ€i c 2 Proof. By the Borel functional calculus,  d 1 ((A āˆ’ t āˆ’ i )āˆ’1 āˆ’ (A āˆ’ t + i )āˆ’1 ) dt = g (A), 2Ļ€i c where g are deļ¬ned by (7.36). Since, by Theorem 7.40, 0 ā‰¤ g ā‰¤ 1 and g converges pointwise to 12 (Ļ‡(c,d) + Ļ‡[c,d] ) as ā†’ 0, the corresponding multiplication operators converge strongly to  1 Ļ‡(c,d) (A) + Ļ‡[c,d] (A) .  2 Stoneā€™s theorem can be improved to norm convergence if we include a test function in Cc (R): Theorem 9.43 (Stone). If A is self-adjoint and h āˆˆ Cc (R), then  1 h(t)((A āˆ’ t āˆ’ i )āˆ’1 āˆ’ (A āˆ’ t + i )āˆ’1 ) dt. h(A) = lim ā†“0 2Ļ€i Proof. By the Borel functional calculus,  1 h(t)((A āˆ’ t āˆ’ i )āˆ’1 āˆ’ (A āˆ’ t + i )āˆ’1 ) dt = h (A), 2Ļ€i where h is deļ¬ned by (7.39). As in the proof of Proposition 7.44, the  functions h converge to h uniformly on R, so h (A) ā†’ h(A). Besides Stoneā€™s formula, other useful identities can be obtained by combining functional calculus and Herglotz functions (Exercise 9.18).

9.8. Fourier transform on R This section can be seen as a detour and an extended example. In it, we d deļ¬ned rely on the material of Chapter 8, revisiting the derivative D = āˆ’i dx 2 in (8.27) and (8.28) as a self-adjoint operator on L (R). We will show that D is diagonalized, i.e., conjugated to a multiplication operator, by a unitary operator known as the Fourier transform. The ļ¬rst step is to use resolvents and Stoneā€™s theorem to compute further functions of D. This line of argument will lead us to a derivation of the Fourier transformā€”which diagonalizes the derivative, i.e., conjugates it to a multiplication operator. This is not the standard approach to introducing the Fourier transform and proving its unitarity, but it illustrates the techniques which will soon be used for SchrĀØodinger operators.

9.8. Fourier transform on R

291

Lemma 9.44. For f āˆˆ L1 (R), the function fĖ† : R ā†’ C deļ¬ned by  1 fĖ†(k) = āˆš eāˆ’iky f (y) dy 2Ļ€ is a bounded continuous function of k āˆˆ R. Proof. Boundedness follows from the k-independent estimate  1 Ė† |f (y)| dy. |f (k)| ā‰¤ āˆš 2Ļ€ Continuity follows from dominated convergence with dominating function |f |, since eāˆ’iky is continuous in k for each y.  Lemma 9.45. For g āˆˆ L1 (R, dk), the function gĖ‡ : R ā†’ C deļ¬ned by  1 āˆš eikx g(k) dk gĖ‡(x) = 2Ļ€ is a bounded continuous function of x āˆˆ R. Proof. This follows from the previous lemma by the observation gĖ‡ = = gĀÆ.



Our goal is to prove that ā€œĖ†ā€ and ā€œĖ‡ā€ extend to unitary maps which are the Fourier transform and the inverse Fourier transform, respectively. Their relation to the operator D is found in the following key calculation. Lemma 9.46. For h āˆˆ Cc (R) and f āˆˆ L1 (R) āˆ© L2 (R), >

h(D)f = (hfĖ†). Proof. The right-hand side is well deļ¬ned, since hfĖ† āˆˆ Cc (R). By Stoneā€™s theorem, h(D) = s-lim h (D), ā†“0

 1 h(k)[RD (k + i ) āˆ’ RD (k āˆ’ i )] dk. h (D) = 2Ļ€i If f āˆˆ L1 (R) āˆ© L2 (R), then h (D)f ā†’ h(D)f in L2 (R). However, using the formula for the resolvents, we can evaluate h (D)f pointwise as where

(h (D)f )(x) !  +āˆž   x 1 i(k+i)(xāˆ’y) i(kāˆ’i)(xāˆ’y) e f (y) dy + e f (y) dy dk h(k) = 2Ļ€ āˆ’āˆž x  1 h(k)eik(xāˆ’y)āˆ’|xāˆ’y| f (y) dy dk. = 2Ļ€

292

9. Consequences of the spectral theorem

By dominated convergence with dominating function |h(k)f (y)| āˆˆ L1 (dy dk),  1 lim(h (D)f )(x) = h(k)eik(xāˆ’y) f (y) dy dk. ā†“0 2Ļ€ Integrating in y and then in k gives precisely  1 āˆš h(k)eikx fĖ†(k) dk = (hfĖ†)(x). lim(h (D)f )(x) = ā†“0 2Ļ€ >

>

Since h (D)f converges to h(D)f in L2 (R) and to (hfĖ†) pointwise, the two limits are equal, which concludes the proof.  Theorem 9.47. The map f ā†’ fĖ† on L1 (R) āˆ© L2 (R) extends to a unitary operator F : L2 (R) ā†’ L2 (R) such that F DF āˆ’1 = Tk,dk . The inverse F āˆ’1 is an extension of the map g ā†’ gĖ‡. Proof. Let us ļ¬rst note that for any f, g āˆˆ L1 (R) āˆ© L2 (R), by Fubiniā€™s theorem,    ikx f (x)Ė‡ g (x) dx = f (x)e g(k) dk dx = fĖ†(k)g(k) dk. (9.29) Combining this with Lemma 9.46, for h āˆˆ Cc (R) and f āˆˆ L1 (R) āˆ© L2 (R),  Ė† dk. f, h(D)f  = fĖ†(k)h(k)f(k) Applying this to a sequence of nonnegative functions hn āˆˆ Cc (R) which is increasing in n and converges pointwise to 1, we obtain  2 Ė† dk f, hn (D)f  = hn (k)|f(k)| s

for each n; taking limits as n ā†’ āˆž, using hn (D) ā†’ I for the left-hand side and monotone convergence for the right-hand side gives the Plancherel formula  f 2 = |fĖ†(k)|2 dk, so fĖ† āˆˆ L2 (R, dk) and the map f ā†’ fĖ† is norm-preserving. Thus, this map extends to a norm-preserving map F : L2 (R, dx) ā†’ L2 (R, dk). gĀÆ, we conclude that g ā†’ gĖ‡ also extends to Using again the observation gĖ‡ = = a norm-preserving map W : L2 (R, dk) ā†’ L2 (R, dx), ĀÆ W G = F (G). (9.30) Using continuity, Lemma 9.46 implies that for all h āˆˆ Cc (R), h(D) = W Th(k),dk F .

9.9. Abstract eigenfunction expansions

293

Using again the sequence hn ā†‘ 1, by strong convergence, we obtain I = W F , so W is onto. By (9.30), this implies that F is onto, so F is unitary and W = F āˆ’1 . Now the earlier conclusions can be stated as h(D) = F āˆ’1 Th(k),dk F = F āˆ’1 h(Tk,dk )F , for all h āˆˆ Cc (R). The set of h āˆˆ Bb (R) for which equality holds is a subalgebra closed under pointwise convergence of uniformly bounded sequences, and since it contains Cc (R), it is equal to Bb (R). Thus, D = F āˆ’1 Tk,dk F , which completes the proof.  For f āˆˆ L2 (R), its Fourier transform F f is not deļ¬ned pointwise. It is deļ¬ned as an element of L2 (R), and since F is a bounded operator and Ļ‡[āˆ’n,n] f ā†’ f as n ā†’ āˆž, F f = lim F (Ļ‡[āˆ’n,n] f ). nā†’āˆž

the Fourier transforms on the right-hand side are Since Ļ‡[āˆ’n,n] f āˆˆ deļ¬ned pointwise, and it is common to write this as  n 1 eāˆ’ikx f (x) dx, (F f )(x) = lim āˆš nā†’āˆž 2Ļ€ āˆ’n L1 (R),

emphasizing that the limit is taken in the sense of L2 (R)-convergence, rather than pointwise. For f āˆˆ L2 (R, dx), f āˆˆ D(D) if and only if F f āˆˆ D(Tk,dk ), so if and only if  k 2 |(F f )(k)|2 dk < āˆž. For n āˆˆ N, this generalizes inductively to f āˆˆ D(Dn ) if and only if  k 2n |(F f )(k)|2 dk < āˆž. For the most part, the above proofs kept a clear conceptual diļ¬€erence between L2 (R, dx) and L2 (R, dk). However, the two are of course equal as Hilbert spaces up to a notational change, and this symmetry was used in the observation (9.30) to relate the Fourier transform and the inverse Fourier transform. When we construct eigenfunction expansions for SchrĀØodinger operators, we will deļ¬ne ā€œĖ†ā€ and ā€œĖ‡ā€ in an operator-dependent way; the symmetry between ā€œĖ†ā€ and ā€œĖ‡ā€ will be lost and additional arguments will be needed.

9.9. Abstract eigenfunction expansions Stoneā€™s theorem allows us to compute operators in the Borel functional calculus, and this can be used to ļ¬nd unitary equivalences which conjugate the given self-adjoint operator A to a multiplication operator B. We will now

294

9. Consequences of the spectral theorem

describe the abstract portion of this approach, which will be applied later to obtain eigenfunction expansions of Jacobi and SchrĀØodinger operators. The operator B will be the operator of multiplication by Ī» on a Hilbert space of the form L2 (R, Cn , W (Ī») dĪ¼(Ī»)) (see Section 6.4) with W (Ī») ā‰„ 0 and Tr W (Ī») = 1 for Ī¼-a.e. Ī». In applications to so-called half-line eigenfunction expansions, we will have n = 1, W = 1. We will also denote L2loc (R, Cn , W dĪ¼) = {g : R ā†’ Cn | gĻ‡[āˆ’k,k] āˆˆ L2 (R, Cn , W dĪ¼) āˆ€k āˆˆ N}. Our goal is to prove: Theorem 9.48. Let A be a self-adjoint operator on H. Let H0 āŠ‚ H be a dense subset of H. Let Ī¼ be a Baire measure on R, and let W be an n Ɨ n matrix-valued function on R with W ā‰„ 0 and Tr W = 1 Ī¼-a.e. Let f ā†’ fĖ† be a linear map from H0 to L2loc (R, Cn , W dĪ¼) such that for all f, g āˆˆ H0 and all h āˆˆ Cc (R),  (9.31) g, h(A)f  = hĖ† g āˆ— W fĖ† dĪ¼. Denote K = L2 (R, Cn , W dĪ¼). Then the following hold. (a) The map f ā†’ fĖ† extends to a norm-preserving map U : H ā†’ K. (b) There exists a linear map U āˆ— : K ā†’ H such that U āˆ— g, f  = g, U f 

(9.32)

for all f āˆˆ H and all g āˆˆ K. This map obeys U āˆ—  ā‰¤ 1. (c) Denote by B the operator of multiplication by Ī» in K. Then h(A) = U āˆ— h(B)U

(9.33)

for all bounded continuous functions h. In particular, U āˆ— U = I. (d) U U āˆ— is an orthogonal projection in K with Ran(U U āˆ— ) = Ran U and Ker(U U āˆ— ) = Ker U āˆ— . (e) Ker U āˆ— = (Ran U )āŠ„ is a resolvent-invariant subspace for B. (f) If in addition Ker U āˆ— = {0}, then U is unitary, so (9.33) provides a unitary equivalence between A and B, (9.33) holds for any bounded Borel function h, and A = U āˆ— BU. The desired case Ker U āˆ— = {0} cannot be established by abstract arguments: for instance, if H is a proper subspace of K and U is inclusion, then U U āˆ— = I. In applications, we will always verify by hand that Ker U āˆ— = {0}, and that veriļ¬cation will use the fact that Ker U āˆ— is a resolvent-invariant subspace.

9.9. Abstract eigenfunction expansions

295

The relation (9.33) is crucial. Viewed as a property of h, note that it is not obviously multiplicative. Thus, we will require a formula which uses the functional calculus linearly: Lemma 9.49. If B is self-adjoint, its resolvents for z āˆˆ C \ R can be expressed as  āˆž RB (z)Ļ• = i eikz eāˆ’ikB Ļ• dk, z āˆˆ C+ , 0  0 eikz eāˆ’ikB Ļ• dk, z āˆˆ Cāˆ’ . RB (z)Ļ• = āˆ’i āˆ’āˆž

Proof. Let z āˆˆ C+ . Pointwise convergence of Riemann sums with uniform boundedness implies strong convergence  t n t  iztj/n āˆ’iBtj/n eikz eāˆ’ikB Ļ• dk = lim i e e Ļ• i nā†’āˆž n 0 j=1

itz āˆ’itB

= (I āˆ’ e e

)Rz (B)Ļ•.

Moreover, eitz eāˆ’itB Rz (B)Ļ• ā‰¤ eāˆ’t Im z Rz (B)Ļ• which goes to 0 as t ā†’ +āˆž. This proves the ļ¬rst formula; the second is proved analogously.  Proof of Theorem 9.48. (a) For any f āˆˆ H0 ,  h(Ī»)fĖ†(Ī»)āˆ— W (Ī»)fĖ†(Ī») dĪ¼(Ī») = f, h(H)f . Let us apply this to a sequence of hn āˆˆ Cc (R) with 0 ā‰¤ hn ā‰¤ 1 which monotonically converges to 1 everywhere. Then using monotone convergence s on the left-hand side and hn (A) ā†’ I on the right-hand side implies that  fĖ†(Ī»)āˆ— W (Ī»)fĖ†(Ī») dĪ¼(Ī») = f 2 . In particular, this proves that fĖ† āˆˆ K. Thus, the map f ā†’ fĖ† is normpreserving from H0 to K, so it extends continuously to a norm-preserving map U : H ā†’ K. (b) For any g āˆˆ K, the map f ā†’ g, U f  is a bounded linear functional and |g, U f | ā‰¤ gU f  ā‰¤ gf . By the Riesz representation theorem, there is a unique vector U āˆ— g which obeys (9.32) for all f āˆˆ H, and U āˆ— g ā‰¤ g. The map U āˆ— is linear since the right-hand side of (9.32) is skew-linear in g. (c) Equation (9.31) can now be rewritten as g, h(A)f  = U g, h(B)U f 

(9.34)

296

9. Consequences of the spectral theorem

for f, g āˆˆ H0 and h āˆˆ Cc (R). Since h(A), h(B) are bounded operators, by density of H0 in H, this holds for all f, g āˆˆ H and h āˆˆ Cc (R). For any bounded continuous h : R ā†’ C, there is a sequence of uniformly bounded approximants hn āˆˆ Cc (R) such that hn ā†’ h pointwise. Then g, hn (A)f  = U g, hn (B)U f , so using strong operator convergence on both sides, we conclude that h also obeys (9.34). Taking h = 1 gives U āˆ— U = I. (d) For any g1 , g2 āˆˆ K, g1 , U U āˆ— g2  = U āˆ— g1 , U āˆ— g2  = U U āˆ— g1 , g2 , so (U U āˆ— )āˆ— = U U āˆ— . Moreover, U U āˆ— U U āˆ— = U IU āˆ— = U U āˆ— . Thus, U U āˆ— is an orthogonal projection. From Ran(U U āˆ— U ) āŠ‚ Ran(U U āˆ— ) āŠ‚ Ran U and U U āˆ— U = U I = U , we conclude Ran(U U āˆ— ) = Ran U . Since U U āˆ— is an orthogonal projection, U U āˆ— g = 0 if and only if g, U U āˆ— g = 0, which is equivalent to U āˆ— g = 0. (e) By (d), g āˆˆ Ran U = Ran(U U āˆ— ) if and only if g, U U āˆ— g = g2 , so if and only if U āˆ— g = g. We will use this as a criterion for Ran U . By (c), we have eitA = U āˆ— eitB U . Since operators eitA and eitB are unitary on H and K, for any f āˆˆ H, f  = eitA f  = U āˆ— eitB U f  ā‰¤ eitB U f  = f . Equality must hold, which implies that eitB U f āˆˆ Ran U . This means that Ran U is invariant for eitB for any t āˆˆ R. Thus, Ker U āˆ— = (Ran U )āŠ„ is invariant for eāˆ’itB = (eitB )āˆ— by Lemma 4.41. In other words, g āˆˆ Ker U āˆ— implies U āˆ— eāˆ’itB g = 0. By Lemma 9.49, for z āˆˆ C+ ,  āˆž āˆ— āˆ’1 U (B āˆ’ z) g = i eitz U āˆ— eāˆ’itB g dt = 0, 0

so Ker U āˆ— is invariant for (B āˆ’ z)āˆ’1 for all z āˆˆ C+ . The case z āˆˆ Cāˆ’ is proved analogously, so Ker U āˆ— is resolvent-invariant for B. (f) If Ker U āˆ— = (Ran U )āŠ„ = {0}, then Ran U is dense in H. Since U is norm-preserving, this implies that Ran U = H and U is unitary. Now (9.33) holds for all bounded Borel functions by Theorem 8.44 and then A = U āˆ— BU holds by Proposition 9.13. 

9.10. Exercises 9.1. Let A be self-adjoint. If S āŠ‚ T āŠ‚ R, prove that Ļ‡S (A) ā‰¤ Ļ‡T (A) in the sense of operator order. 9.2. Let A be self-adjoint. Prove that min Ļƒ(A) = sup{E āˆˆ R | Ļ‡(āˆ’āˆž,E) (A) = 0}.

9.10. Exercises

297

9.3. Let u, v āˆˆ H. If Ī¼u and Ī¼v are mutually singular, prove that u āŠ„ v. Hint: Use S such that Ī¼A,u (S) = 0 and Ī¼A,v (S c ) = 0, and compute u, v = u, Ļ‡S (A)v + u, Ļ‡S c (A)v. 9.4. If A is self-adjoint and w = u + v, prove that Ī¼A,w " Ī¼A,u + Ī¼A,v . 9.5. If Ī¼ is a maximal spectral measure for A and S is a Borel set, prove that Ļ‡S dĪ¼ is a maximal spectral measure for the restriction of A to Ran Ļ‡S (A). 9.6. Let A, B be unbounded self-adjoint operators. Prove that the following are equivalent: (a) eikA eilB = eilB eikA for all k, l āˆˆ R. (b) RA (z)RB (w) = RB (w)RA (z) for all z, w āˆˆ C \ R. (c) f (A)g(B) = g(B)f (A) for all f, g āˆˆ Bb (R). If these conditions hold, the unbounded operators A, B are said to commute. 9.7. Let A be a self-adjoint operator on H which has an orthonormal basis āˆž of eigenvectors (vn )āˆž n=1 , with corresponding eigenvalues (Ī»n )n=1 . (a) Prove that Ļƒ(A) = {Ī»n | n āˆˆ N}. (b) If dim H = āˆž, construct a self-adjoint operator on H which has an orthonormal basis of eigenvectors and Ļƒ(A) = [0, 1]. 9.8. If A is self-adjoint, K āˆˆ L(H) relatively compact, and Ļˆ āˆˆ Hac (A), prove that lim Keāˆ’itA Ļˆ2 = 0.

tā†’āˆž

Hint: Use the Riemannā€“Lebesgue lemma and imitate the proof of the RAGE theorem. 9.9. If A is a bounded self-adjoint operator on H and dim H = āˆž, prove that Ļƒess (A) = āˆ…. 9.10. Prove a strengthening of Weylā€™s criterion: for any Ī» āˆˆ C, if V stands for the set of all orthonormal sequences v = (vn )āˆž n=1 in H, or for the w set of all normalized sequences with vn ā†’ 0, then dist(Ī», Ļƒess (A)) = inf lim inf (A āˆ’ Ī»)vn . vāˆˆV nā†’āˆž

9.11. Let A be a self-adjoint operator, and let {Ī»1 , . . . , Ī»k } āŠ‚ R be a ļ¬nite set. Prove that Ļƒess (A) āŠ‚ {Ī»1 , . . . , Ī»k } if and only if the operator k j=1 (A āˆ’ Ī»j ) is compact. 9.12. If A is a self-adjoint operator bounded below and Ī»n are deļ¬ned as in Section 9.5, prove that min Ļƒess (A) = limnā†’āˆž Ī»n (A).

298

9. Consequences of the spectral theorem

9.13. Let A be a self-adjoint operator on H and Ī¼m , and let m āˆˆ N āˆŖ {āˆž} be its multiplicity m measures. Let Ī» be an eigenvalue of A. (a) Prove that Ī¼m ({Ī»}) > 0 for exactly one value of m āˆˆ N āˆŖ {āˆž}. (b) Prove that Ī¼m ({Ī»}) > 0 if and only if dim Ker(A āˆ’ Ī») = m. 9.14. Let A be a self-adjoint operator with multiplicity 1 spectrum (i.e., denoting by Ī¼m its multiplicity m measures, Ī¼m = 0 for all m ā‰„ 2). Prove that a vector Ļˆ is cyclic for A if and only if Ī¼A,Ļˆ is mutually absolutely continuous with Ī¼1 . 9.15. Let A be a self-adjoint operator, and let Ī¼m be its multiplicity m measures. For any n āˆˆ N, prove that A has a spectral basis with at most n vectors if and only if Ī¼m = 0 for all m > n. 9.16. Let A be a self-adjoint operator on H and Ī¼m , and let m āˆˆ N āˆŖ {āˆž} be its multiplicity m measures. (a) Let Sm be any Borel set such that Ī¼m is supported on Sm and Ī¼k (Sm ) = 0 for all k = m. Prove that the subspace Hm (A) = Ran Ļ‡Sm (A) is independent of the choice of Sm . This subspace is called the multiplicity m subspace for the operator A. (b) Prove that 3 Hm (A). H= māˆˆNāˆŖ{āˆž}

9.17. Let A be a self-adjoint operator on H, let Ī¼m , m āˆˆ N āˆŖ {āˆž} be its multiplicity m measures, and let n āˆˆ N. 4n (a) Prove that j=1 Hj (A) can be written as a direct sum of n cyclic subspaces of A. (b) Assume that a Borel set S is such that Ļ‡S (A) can be written as a direct sum of n cyclic subspaces of A. Prove that Ī¼m (S) = 0 for all m > n. 9.18. Let A be a self-adjoint operator, and let Ī» āˆˆ R. Prove that s-lim i (A āˆ’ Ī» āˆ’ i )āˆ’1 = Ļ‡{Ī»} (A). ā†“0

Chapter 10

Jacobi matrices

Jacobi matrices are tridiagonal self-adjoint matrices with real diagonal entries and positive oļ¬€-diagonal entries. The simplest form is a ļ¬nite Jacobi matrix, āŽž āŽ› b1 a1 āŽŸ āŽœa1 b2 a2 āŽŸ āŽœ āŽŸ āŽœ .. .. āŽŸ āŽœ . . a2 (10.1) J =āŽœ āŽŸ āŽŸ āŽœ . . .. .. a āŽ  āŽ dāˆ’1 adāˆ’1 bd with a1 , . . . , adāˆ’1 āˆˆ (0, āˆž) and b1 , . . . , bd āˆˆ R. The elements left blank (matrix elements Jkl with |k āˆ’ l| ā‰„ 2) are implied to be 0. These are clearly Hermitian matrices, i.e., self-adjoint operators on Cd . Similarly, half-line Jacobi matrices are operators on 2 (N) given formally by the tridiagonal matrix expression āŽ› āŽž b1 a1 āŽœa1 b2 a2 āŽŸ āŽœ āŽŸ āŽœ āŽŸ a b a 2 3 3 āŽŸ J =āŽœ (10.2) āŽœ .. .. āŽŸ āŽœ āŽŸ . . a3 āŽ āŽ  .. . with an > 0 and bn āˆˆ R for all n. For operators on 2 (N), one has to be careful with matrix notation: denoting by (Ī“n )āˆž n=1 the standard basis of 2 (N), every operator J on 2 (N) corresponds to an inļ¬nite matrix of coeļ¬ƒcients Jkl = Ī“k , JĪ“l , but not every inļ¬nite matrix (Jkl )āˆž k,l=1 corresponds to a bounded linear operator. In Section 10.1 we proved that for 299

300

10. Jacobi matrices

āˆž āˆž (an )āˆž n=1 , (bn )n=1 āˆˆ  (N), (10.2) deļ¬nes a bounded self-adjoint operator with Ī“1 as a cyclic vector. From there on, we will refer to the spectral measure Ī¼ = Ī¼Ī“1 as the spectral measure corresponding to J. We will also describe an orthogonal polynomial construction which starts from Ī¼ and results in a sequence of recursion coeļ¬ƒcients an > 0, bn āˆˆ R, and prove that the two correspondences are mutually inverse. In Section 10.2 we will discuss the case of unbounded Jacobi matrices; this section can be skipped by a reader interested only in the bounded case.

After that, we will consider connections between J, Ī¼, and the corresponding Herglotz function m(z) = Ī“1 , (J āˆ’ z)āˆ’1 Ī“1 ,

(10.3)

which in this context is called the Weyl m-function or simply the m-function. Since Ī¼ is the spectral measure for the cyclic vector Ī“1 ,  1 dĪ¼(x), (10.4) m(z) = xāˆ’z and the m-function encodes spectral properties of J. Sections 10.3 and 10.4 contain useful perspectives on the Weyl function. In Section 10.5 we introduce full-line (or two-sided ) Jacobi matrices, deļ¬ned with the same tridiagonal pattern but acting on 2 (Z).

10.1. The canonical spectral measure and Favardā€™s theorem In this section we introduce the basic objects associated with bounded halfline Jacobi matrices. We start with the precise deļ¬nition: Lemma 10.1. If a, b āˆˆ āˆž (N) and an > 0, bn āˆˆ R for all n āˆˆ N, then

b1 u1 + a1 u2 n=1 (10.5) (Ju)n = anāˆ’1 unāˆ’1 + bn un + an un+1 n ā‰„ 2 deļ¬nes a bounded self-adjoint operator J on 2 (N) and J ā‰¤ 2 sup an + sup|bn |. nāˆˆN

(10.6)

nāˆˆN

Proof. Denote Ī± = supnāˆˆN an and Ī² = supnāˆˆN |bn |. By the Cauchyā€“Schwarz inequality, for any u āˆˆ 2 (N), |anāˆ’1 unāˆ’1 + bn un + an un+1 |2 ā‰¤ (Ī± + Ī² + Ī±)(Ī±|unāˆ’1 |2 + Ī²|un |2 + Ī±|un+1 |2 ). Taking the sum over n shows that Ju2 ā‰¤ (2Ī± + Ī²)2 u2 so (10.6) holds. Self-adjointness is the statement that u, Jv = Ju, v

(10.7)

10.1. The canonical spectral measure and Favardā€™s theorem

301

holds for all u, v āˆˆ 2 (N). Since

āŽ§ āŽŖ āŽØbk Ī“k , JĪ“l  = JĪ“k , Ī“l  = amin(k,l) āŽŖ āŽ© 0

k=l |k āˆ’ l| = 1 |k āˆ’ l| ā‰„ 2,

(10.7) holds for u, v āˆˆ {Ī“n | n āˆˆ N}. By sesquilinearity, (10.7) then holds for all u, v āˆˆ 2c (N) = span{Ī“n | n āˆˆ N}. Finally, by continuity, (10.7) holds for all u, v.  Another proof of (10.6) consists of decomposing J = A + B + C where each of the operators A, B, C has one nonzero diagonal and their norms are bounded by Ī±, Ī², Ī±, respectively. The estimate (10.6) has a converse, up to a multiplicative constant (Exercise 10.1). Due to their tridiagonal structure, strong and weak operator convergence of Jacobi matrices can be described very explicitly in terms of their coeļ¬ƒcients (Exercise 10.2). Lemma 10.2. For any bounded half-line Jacobi matrix J: (a) Ī“1 is a cyclic vector for J; (b) the support of its spectral measure Ī¼J,Ī“1 is an inļ¬nite set; (c) the sequence (J n Ī“1 )āˆž n=0 is linearly independent; applying to it the Gramā€“Schmidt process gives the orthonormal basis (Ī“n )āˆž n=1 . Proof. From (10.5), it follows by induction that for all n āˆˆ N, where cn =

n

J n Ī“1 āˆ’ cn Ī“n+1 āˆˆ span{Ī“k | 1 ā‰¤ k ā‰¤ n},

k=1 ak k

and therefore

span{J Ī“1 | 0 ā‰¤ k ā‰¤ n} = span{Ī“k | 1 ā‰¤ k ā‰¤ n + 1}, from which it follows that Ī“1 is cyclic. Reversing this, for each n we have n k Ī“n+1 āˆ’ cāˆ’1 n J Ī“1 āˆˆ span{J Ī“1 | 0 ā‰¤ k ā‰¤ n āˆ’ 1}.

Since (Ī“n+1 )āˆž n=0 is an orthonormal sequence and cn > 0, the claim follows by uniqueness of the Gramā€“Schmidt process. By the spectral theorem, there is a unitary map from L2 (R, dĪ¼Ī“1 ) to 2 (N), so the two Hilbert spaces have equal dimension. If supp Ī¼Ī“1 was a ļ¬nite set, the space L2 (R, dĪ¼Ī“1 ) would be ļ¬nite dimensional, leading to a contradiction.  Since Ī“1 is a cyclic vector, the spectral measure Ī¼Ī“1 is a maximal spectral measure for J. From the perspective of general spectral theory, there is no reason to prefer Ī¼J,Ī“1 over another maximal spectral measure; however, in

302

10. Jacobi matrices

the theory of Jacobi matrices, Ī¼J,Ī“1 is regarded as the canonical spectral measure corresponding to J. The proof of Lemma 10.2 also applies for ļ¬nite Jacobi matrices with appropriate restrictions of indices: Lemma 10.3. For any d Ɨ d Jacobi matrix J: (a) Ī“1 is a cyclic vector for J; (b) the support of its spectral measure Ī¼J,Ī“1 has cardinality d; (c) the sequence (J n Ī“1 )dāˆ’1 n=0 is linearly independent; applying to it the Gramā€“Schmidt process gives the orthonormal basis (Ī“n )dn=1 . The unitary map U : L2 (R, dĪ¼J,Ī“1 ) ā†’ 2 (N) provided by the spectral theorem maps the constant function 1 to Ī“1 and conjugates J to TĪ»,dĪ¼J,Ī“1 (Ī») . Thus, U maps the monomial Ī»n to the vector J n Ī“1 . Thus, the Gramā€“ Schmidt process in Lemma 10.2(c) is related to the orthonormal polynomial construction in Example 3.48. We will now develop this idea and obtain the inverse of the map J ā†’ Ī¼J,Ī“1 . An index shift by 1 is apparent already in Lemma 10.2(c) and will reappear below. This is an artifact of the standard indexing conventions for half-line Jacobi matrices; it would vanish if we regarded half-line Jacobi matrices as operators on 2 (N āˆŖ {0}). For a moment, let us forget about Jacobi matrices and work in the Hilbert space L2 (R, dĪ¼) for a compactly supported probability Borel measure Ī¼ on R. Recall that in orthogonal polynomial theory, a measure is said to be nontrivial if supp Ī¼ is not a ļ¬nite set. Then every nontrivial polynomial is nonzero in L2 (R, dĪ¼) and polynomials are dense in L2 (R, dĪ¼); in other words, the sequence (xn )āˆž n=0 is a linearly independent sequence with a dense span (see Example 3.48). Thus, applying to it the Gramā€“Schmidt process gives a sequence of orthonormal polynomials pn (x), deg pn = n,  (10.8) pm , pn  = pm (x)pn (x) dĪ¼(x) = Ī“m,n , which form an orthonormal basis in L2 (R, dĪ¼). Since Ī¼ is supported on the real line, pn (x) = pn (x) follows by induction through the Gramā€“Schmidt process, so the polynomials pn have real coeļ¬ƒcients. Thus, the complex conjugate in (10.8) can be removed. Proposition 10.4 (Jacobi recursion). Let Ī¼ be a nontrivial probability Borel measure on R with ļ¬nite moments, and let pn be its orthonormal polynomiāˆž als. Then there exist sequences (an )āˆž n=1 , (bn )n=1 with an > 0, bn āˆˆ R for all n āˆˆ N, such that xpn (x) = an pnāˆ’1 (x) + bn+1 pn (x) + an+1 pn+1 (x)

(10.9)

10.1. The canonical spectral measure and Favardā€™s theorem

303

holds for all n ā‰„ 1 and xp0 (x) = b1 p0 (x) + a1 p1 (x).

(10.10)

āˆž āˆž Moreover, if Ī¼ is compactly supported, then (an )āˆž n=1 , (bn )n=1 āˆˆ  (N).

Remark 10.5. It is standard to set the convention pāˆ’1 (x) = 0 and claim that (10.9) holds for all n ā‰„ 0, with an arbitrary value of a0 . Proof. Since xpn (x) is a polynomial of degree n+1, it is a linear combination of p0 , . . . , pn+1 . Since pn is orthogonal to all polynomials of degree ā‰¤ n āˆ’ 1, for k ā‰¤ n āˆ’ 2, pk , xpn  = xpk , pn  = 0, so xpn (x) is a linear combination of pnāˆ’1 , pn , pn+1 ; i.e., there exist coeļ¬ƒcients an+1 , bn+1 , cn+1 āˆˆ R such that xpn (x) = cn+1 pnāˆ’1 (x) + bn+1 pn (x) + an+1 pn+1 (x) (for n = 0, the term cn+1 pnāˆ’1 (x) should be ommitted). Since pn , pn+1 have positive leading coeļ¬ƒcients, it follows that an+1 > 0. Moreover, for n ā‰„ 1, cn+1 = pnāˆ’1 , xpn  = xpnāˆ’1 , pn  = an . Now assume that supp Ī¼ āŠ‚ [āˆ’C, C] for some C < āˆž. From  an = xpnāˆ’1 , pn  = xpn (x)pnāˆ’1 (x)dĪ¼(x), using |x| ā‰¤ C and using the Cauchyā€“Schwarz inequality gives an ā‰¤ Cpn pnāˆ’1  = C. Similarly, |bn | = |pnāˆ’1 , xpnāˆ’1 | ā‰¤ C so the sequences are bounded.



N Deļ¬nition 10.6. The coeļ¬ƒcients (an , bn )āˆž n=1 āˆˆ ((0, āˆž) Ɨ R) in Proposition 10.4 are called the Jacobi parameters of the measure Ī¼.

Next, we see that Jacobi parameters determine a compactly supported measure uniquely: Lemma 10.7. If two nontrivial compactly supported probability measures on R have the same Jacobi parameters, then they are equal. Ėœ have Proof. For any probability measure, pāˆ’1 = 0 and p0 = 1. If Ī¼ and Ī¼ the same Jacobi parameters, then by induction using (10.9), they have the same orthonormal polynomials. For all n,   Ī¼ pn dĪ¼ = Ī“n,0 = pn dĖœ

304

10. Jacobi matrices

since this integral can be   interpreted as the inner product of pn with 1. Then, by linearity, P dĪ¼ = P dĖœ Ī¼ for any polynomial P . Since polynomials are dense in C(supp Ī¼ āˆŖ supp Ī¼ Ėœ), it follows that Ī¼ = Ī¼ Ėœ.  Of course, the Jacobi parameters an , bn of a measure Ī¼ can be used to assemble a half-line Jacobi matrix J. By Lemma 10.7, this construction from Ī¼ to J is injective. We will now see that it is also surjective. This, and more, follows from the following lemma: Lemma 10.8. For a bounded half-line Jacobi matrix J with coeļ¬ƒcients (an , bn )āˆž n=1 , the Jacobi parameters of the spectral measure Ī¼J,Ī“1 are precisely (an , bn )āˆž n=1 . Proof. By Lemma 10.2, there is a unitary map U : L2 (R, dĪ¼) ā†’ 2 (N) which maps 1 to Ī“1 and conjugates multiplication by x with the operator J. It follows that U (P (x)) = P (J)Ī“1 for any polynomial P . Unitary maps preserve inner products, so they preserve Gramā€“Schmidt processes. In L2 (R, dĪ¼J,Ī“1 ), the Gramā€“Schmidt process on (xn )āˆž n=0 gives 2 (N), the Gramā€“Schmidt process on (J n Ī“ )āˆž , so in  (pn (x))āˆž 1 n=0 n=0 gives . However, by Lemma 10.2 this Gramā€“Schmidt process in 2 (N) (pn (J)Ī“1 )āˆž n=0 gives (Ī“n+1 )āˆž n=0 , so we conclude that for all n, pn (J)Ī“1 = Ī“n+1 . From the deļ¬nition of the Jacobi matrix, JĪ“n+1 = an Ī“n + bn+1 Ī“n+1 + an+1 Ī“n+2 , and we can now rewrite this as Jpn (J)Ī“1 = an pnāˆ’1 (J)Ī“1 + bn+1 pn (J)Ī“1 + an+1 pn+1 (J)Ī“1 . Applying U āˆ’1 to this equality gives (10.9), and concludes the proof.



In summary, the two constructions, from a Jacobi matrix to its spectral measure and from a measure to its Jacobi parameters, are bijections and they are mutually inverse: Theorem 10.9 (Favardā€™s theorem). The map J ā†’ Ī¼J,Ī“1 is a bijection between the set of bounded half-line Jacobi matrices and the set of compactly supported nontrivial probability measures on R. Its inverse is obtained by taking the Jacobi parameters of a measure and using them as coeļ¬ƒcients of the Jacobi matrix. All the arguments presented here also apply to ļ¬nite Jacobi matrices, with the appropriate range of indices (see Exercise 10.3). From now on, we will denote Ī¼ = Ī¼J,Ī“1 and always consider J and Ī¼ related as in Favardā€™s theorem.

10.2. Unbounded Jacobi matrices

305

10.2. Unbounded Jacobi matrices In this section, we consider how the matrix representation (10.2) leads to unbounded self-adjoint Jacobi matrices in the case when at least one of the āˆž sequences (an )āˆž n=1 , (bn )n=1 is unbounded. This section can be skipped by a reader who is only interested in bounded Jacobi matrices, except for a glance at some terminology (e.g., every bounded Jacobi matrix is limit point, and in later sections we will state various results for Jacobi matrices which are limit point). We will see that in the unbounded case, the matrix representation may āˆž be incomplete; depending on the coeļ¬ƒcient sequences (an )āˆž n=1 , (bn )n=1 , it may be necessary to also specify a boundary condition at āˆž in order to specify an unbounded self-adjoint Jacobi matrix. We begin by deļ¬ning a maximal Jacobi operator on 2 (N) with domain

 āˆž  2 2 |anāˆ’1 unāˆ’1 + bn un + an un+1 | < āˆž . D(Jmax ) = u āˆˆ  (N) | n=2

Its action on the domain is deļ¬ned by

b1 u1 + a1 u2 (Jmax u)n = anāˆ’1 unāˆ’1 + bn un + an un+1

n=1 n ā‰„ 2.

This operator may or may not be self-adjoint. We deļ¬ne the Wronskian of two sequences u, v as the sequence Wn (u, v) = an (un+1 vn āˆ’ un vn+1 ). We show that the Wronskian has a limit as n ā†’ āˆž and that this limit provides the obstruction to self-adjointness: Lemma 10.10. For any u, v āˆˆ D(Jmax ), lim Wn (u, v) = Jmax u, v āˆ’ u, Jmax v.

nā†’āˆž

Proof. For n āˆˆ N, by a direct calculation,

Wn (u, v) āˆ’ Wnāˆ’1 (u, v) n ā‰„ 2 (Jmax u)n vn āˆ’ un (Jmax v)n = n = 1, W1 (u, v) so summing from 1 to n gives n n   (Jmax u)j vj āˆ’ uj (Jmax v)j = Wn (u, v). j=1

j=1

by the Cauchyā€“Schwarz inequality, the Since u, v, Jmax u, Jmax v āˆˆ left-hand side converges as n ā†’ āˆž, and taking this limit concludes the proof.  2 (N),

306

10. Jacobi matrices

Accordingly, we deļ¬ne the boundary Wronskian W+āˆž (u, v) = lim Wn (u, v) nā†’āˆž

for u, v āˆˆ D(Jmax ). To search for self-adjoint restrictions of Jmax , we begin by ļ¬nding its adjoint. Recall that 2c (N) denotes the span of {Ī“n | n āˆˆ N}, i.e., the set of sequences with ļ¬nitely many nonzero entries. Theorem 10.11. The restriction J0 of Jmax to D(J0 ) = 2c (N) obeys J0āˆ— = Jmax and J0 is the restriction of Jmax to D(J0 ) = {u āˆˆ D(Jmax ) | W+āˆž (u, v) = 0 for all v āˆˆ D(Jmax )}. Proof. Assume that u, w āˆˆ 2 (N) obey u, J0 v = w, v

āˆ€w āˆˆ 2c (N).

Since any v is a linear combination of Ī“n ā€™s, this is equivalent by linearity to u, J0 Ī“n  = w, Ī“n 

āˆ€n āˆˆ N.

This is equivalent to wn = anāˆ’1 unāˆ’1 + bn un + an un+1 (with the convention u0 = 0), so it is equivalent to (u, w) āˆˆ Ī“(Jmax ). This proves that J0 is densely deļ¬ned and J0āˆ— = Jmax . As the adjoint of another operator, Jmax is closed. Since J0 is a restriction of Jmax , so is J0 . Moreover, u āˆˆ J0 = (J0āˆ— )āˆ— if and only if W+āˆž (u, v) = 0  for all v āˆˆ D(Jmax ). We deļ¬ne a self-adjoint half-line Jacobi matrix J as a self-adjoint restriction of Jmax . By general principles, this is equivalent to J being a self-adjoint extension of J0 and equivalent to J0 āŠ‚ J āŠ‚ Jmax ,

J āˆ— = J.

We will use the framework of Section 8.7 to describe such self-adjoint restrictions. Lemma 10.12. The quotient vector space D(Jmax )/D(J0 ) has dimension 0 or 2. Proof. For every n, the map D(Jmax ) ā†’ C2 , given by ! un+1 , u ā†’ an un

10.2. Unbounded Jacobi matrices

307

lets us express Wn in terms of a symplectic form on C2 , ! ! ! un+1 0 1 vn+1 Wn (u, v) = . an un an vn āˆ’1 0 ucker identity By Theorem 8.64, Wn obeys the PlĀØ Wn (v1 , v2 )Wn (v3 , v4 ) āˆ’ Wn (v1 , v3 )Wn (v2 , v4 ) + Wn (v1 , v4 )Wn (v2 , v3 ) = 0 for all v1 , v2 , v3 , v4 āˆˆ 2 (N); alternatively, this follows directly from the identity   (v1 )n+1 (v2 )n+1 (v3 )n+1 (v4 )n+1    an (v1 )n an (v2 )n an (v3 )n an (v4 )n    (v1 )n+1 (v2 )n+1 (v3 )n+1 (v4 )n+1  = 0.   an (v1 )n an (v2 )n an (v3 )n an (v4 )n  Taking n ā†’ āˆž, we conclude that for all v1 , v2 , v3 , v4 āˆˆ D(Jmax ), W+āˆž (v1 , v2 )W+āˆž (v3 , v4 ) āˆ’ W+āˆž (v1 , v3 )W+āˆž (v2 , v4 ) + W+āˆž (v1 , v4 )W+āˆž (v2 , v3 ) = 0. Thus, by Theorem 8.64, the quotient space D(Jmax )/D(J0 ) has dimension 0 or 2.  Deļ¬nition 10.13. The case J0 = Jmax is called the limit point case. The case dim(D(Jmax )/D(J0 )) = 2 is called the limit circle case. In particular, Jmax is self-adjoint if and only if we are in the limit point case. By Lemma 10.1, a bounded Jacobi matrix is self-adjoint, so W+āˆž (u, v) = 0 for all u, v āˆˆ 2 (N). Thus, every bounded Jacobi matrix is in the limit point case. More generally: āˆž Lemma 10.14. If (an )āˆž n=1 āˆˆ  (N), we are in the limit point case.

Proof. By the Cauchyā€“Schwarz inequality, for all u, v āˆˆ 2 (N), āˆž 

|Wn (u, v)| =

n=1

āˆž 

an |un+1 vn āˆ’ un vn+1 | ā‰¤ 2aāˆž u2 v2 < āˆž.

n=1

In particular, for all u, v āˆˆ 2 (N), Wn (u, v) ā†’ 0 as n ā†’ āˆž.



Jacobi matrices with an = 1 for all n are also called discrete SchrĀØodinger operators; by Lemma 10.14, discrete SchrĀØodinger operators are always in / āˆž (N). the limit point case, even if (bn )āˆž n=1 āˆˆ In the limit circle case, we see that the matrix representation does not fully describe a self-adjoint Jacobi matrix. Self-adjoint extensions J of J0 are in bijective correspondence with Lagrangian subspaces D(J) āŠ‚ D(Jmax ), and they can be parametrized by a single self-adjoint boundary condition by the results of Section 8.7:

308

10. Jacobi matrices

Corollary 10.15. In the limit circle case, for any v āˆˆ D(Jmax )\D(J0 ) such that W+āˆž (v, v) = 0, D(J) = {u āˆˆ D(Jmax ) | W+āˆž (u, v) = 0} deļ¬nes a self-adjoint restriction of Jmax . Conversely, every self-adjoint restriction of Jmax is of this form. Finally, let us establish the canonical spectral measure for an unbounded self-adjoint Jacobi matrix: Theorem 10.16. Let J be a self-adjoint half-line Jacobi matrix. The vector Ī“1 is cyclic and the spectral measure dĪ¼J,Ī“1 has all ļ¬nite moments, i.e.,  āˆ€n āˆˆ N āˆŖ {0}. |x|2n dĪ¼J,Ī“1 (x) < āˆž The Jacobi parameters of Ī¼J,Ī“1 are the coeļ¬ƒcients of the Jacobi matrix. Proof. For any v āˆˆ 2 (N) and z āˆˆ C \ R, (J āˆ’ i)āˆ’1 v, (J āˆ’ z)āˆ’1 (J āˆ’ i)āˆ’1 v = v, (J + i)āˆ’1 (J āˆ’ z)āˆ’1 (J āˆ’ i)āˆ’1 v, and by the Borel functional calculus, this can be written as   1 1 dĪ¼J,(Jāˆ’i)āˆ’1 v (x) = dĪ¼J,v (x). xāˆ’z (x + i)(x āˆ’ z)(x āˆ’ i) As a function of z, this determines the measure uniquely, so we conclude 1 dĪ¼J,v (x). dĪ¼J,(Jāˆ’i)āˆ’1 v (x) = |x āˆ’ i|2 Applying this inductively, using (J āˆ’ i)n Ī“1 āˆˆ D(J) = Ran((J āˆ’ i)āˆ’1 ), we conclude that 1 n (x). dĪ¼ dĪ¼J,Ī“1 (x) = |x āˆ’ i|2n J,(Jāˆ’i) Ī“1 In particular, for any n,   2n |x āˆ’ i| dĪ¼J,Ī“1 (x) = 1 dĪ¼J,(Jāˆ’i)n Ī“1 = (J āˆ’ i)n Ī“1 22 < āˆž, so Ī¼J,Ī“1 has ļ¬nite moments. For any n, k āˆˆ N, the functions hn,k (x) = min{k, |x|n } sgn x are bounded, so hn,k (J)Ī“1 is in the cyclic subspace of Ī“1 . By dominated convergence with dominating function x2n ,  n 2 k ā†’ āˆž, J Ī“1 āˆ’ hn,k (J)Ī“1 2 = |xn āˆ’ hn,k (x)|2 dĪ¼J,Ī“1 (x) ā†’ 0, so J n Ī“1 is in the cyclic subspace of Ī“1 . Applying the Gramā€“Schmidt process as in the proof of Lemma 10.2, this cyclic subspace contains all Ī“n , so it is equal to 2 (N).

10.3. Weyl solutions and m-functions

309

Comparing the Gramā€“Schmidt process for monomials in L2 (R, dĪ¼J,Ī“1 ) 2 with the Gramā€“Schmidt process for (J n Ī“1 )āˆž n=0 in  (N) shows, as in the bounded case, equality of Jacobi parameters and coeļ¬ƒcients of the Jacobi matrix.  Thus, we obtain once again a canonical spectral measure Ī¼ = Ī¼J,Ī“1 and the corresponding Weyl function (10.3) and (10.4). For measures with unbounded support, the Jacobi parameters may or may not uniquely determine the measure. This has a fascinating connection with the limit point/limit circle dichotomy for Jacobi matrices [2], [92, Section 3.8], [25, Section 2.4]. In the rest of this chapter, we will restrict our work to the limit point case.

10.3. Weyl solutions and m-functions In this section, we will introduce Weyl solutions for half-line Jacobi matrices, and use them to study the Weyl m-function. A sequence v is said to be a (formal) eigensolution at z āˆˆ C if it obeys the Jacobi recursion anāˆ’1 vnāˆ’1 + bn vn + an vn+1 = zvn .

(10.11)

The word ā€œformalā€ is used to emphasize that the sequence v is not required to be square-summable (and therefore not part of the Hilbert space), but we will usually omit it. When discussing eigensolutions, we will use the convention to also assume existence of a coeļ¬ƒcient a0 > 0 and assume that v0 is also deļ¬ned. Then (10.11) can be used to express v2 in terms of v0 and v1 , and so on, so an eigensolution is uniquely deļ¬ned by the values of v1 and a0 v0 . In other words, since (10.11) is a second-order recursion relation and an = 0 for all n, for any ļ¬xed z āˆˆ C, the set of eigensolutions is a two-dimensional vector space. The second-order recursion relation can be rewritten as a ļ¬rst-order system, ! ! vn vn+1 = A(an , bn ; z) , (10.12) an vn anāˆ’1 vnāˆ’1 where A(an , bn ; z) = is called the 1-step transfer matrix.

zāˆ’bn an

an

āˆ’ a1n 0

!

310

10. Jacobi matrices

We call an eigensolution nontrivial if it is not identically zero. By (10.12), if vnāˆ’1 = vn = 0 for some n, then v is trivial; in other words, a nontrivial eigensolution can never have two consecutive zeros. It would have seemed more obvious to rewrite the Jacobi recursion as the ļ¬rst-order system ! ! ! zāˆ’bn āˆ’ anāˆ’1 vn+1 vn a a n n = , vn vnāˆ’1 1 0 but this has the disadvantage that a single transfer matrix depends on two oļ¬€-diagonal Jacobi parameters. Our transfer matrices also have the useful property that det A(a, b; z) = 1. Note that the Wronskian of two sequences can also be expressed as a determinant, ! un+1 vn+1 . Wn (u, v) = det an un an vn Lemma 10.17. If u, v are eigensolutions at z, then their Wronskian is independent of n. Proof. Combining the recursions for u, v as ! ! un vn un+1 vn+1 = A(an , bn ; z) an un an vn anāˆ’1 unāˆ’1 anāˆ’1 vnāˆ’1

(10.13)

and taking determinants, the claim follows from det A(an , bn ; z) = 1.



Lemma 10.18. If u, v are eigensolutions at z, and Ī±, Ī² āˆˆ C, the following are equivalent: (a) Ī±u + Ī²v = 0. (b) For one value of n, un+1 vn+1 an un an vn

!

Ī± Ī²

! = 0.

(10.14)

(c) For all values of n, (10.14) holds. In particular, u, v are linearly independent if and only if their Wronskian is nonzero. Proof. (a) =ā‡’ (b) and (c) =ā‡’ (a) are obvious. (b) =ā‡’ (c) is proved by induction, multiplying by A(an+1 , bn+1 ; z) or by A(an , bn ; z)āˆ’1 from the left to increase or decrease the index by 1. In particular, u, v are linearly independent if and only if (10.14) has no nontrivial solutions, i.e., if and only if the Wronskian is nonzero. 

10.3. Weyl solutions and m-functions

311

Of particular importance are certain eigensolutions deļ¬ned by their behavior at +āˆž: Deļ¬nition 10.19. A Weyl solution for J at z is a nontrivial sequence Ļˆ = (Ļˆn )āˆž n=0 which obeys (10.11) for all n āˆˆ N and Ļˆ is in the domain of J. In the limit point case, this can be restated as: a Weyl solution at z is 2 an eigensolution at z such that āˆž n=0 |Ļˆn | < āˆž. Even though Ļˆ is in the domain of J and solves (10.11), in general (Ļˆn )āˆž n=1 is not really an eigenvector of the operator J, and z is not an eigenvalue, unless Ļˆ0 = 0. The Weyl solution should be thought of as a solution which obeys the boundary condition at +āˆž but may not obey the boundary condition Ļˆ0 = 0. Proposition 10.20. For any z āˆˆ C\Ļƒess (J), the set of Weyl solutions, with the trivial solution added, is one dimensional. Moreover: (a) if z āˆˆ Ļƒd (J), then for any Weyl solution Ļˆ, Ļˆ0 = 0; (b) if z āˆˆ C \ Ļƒ(J), then for any Weyl solution Ļˆ, Ļˆ0 = 0 and m(z) = āˆ’

Ļˆ1 . a0 Ļˆ0

(10.15)

Proof. If Ļˆ, ĻˆĖœ are both Weyl solutions at z, they decay as n ā†’ āˆž, and so does their Wronskian. Since the Wronskian is independent of n, it must be zero, so Ļˆ, ĻˆĖœ are linearly dependent. Thus, there cannot be two linearly independent Weyl solutions. If z āˆˆ Ļƒd (J), then there is an eigenvector of J, i.e., a nontrivial Ļˆ in the domain of J such that JĻˆ = zĻˆ. This sequence obeys (10.11) for n ā‰„ 2 and b1 Ļˆ1 + a1 Ļˆ2 = (JĻˆ)1 = zĻˆ1 , so Ļˆ can be extended to a Weyl solution at z by setting Ļˆ0 = 0. āˆ’1 If z āˆˆ C \ Ļƒ(J), we consider (Ļˆn )āˆž n=1 = Ļˆ = (J āˆ’ z) Ī“1 in the domain of J. Since (J āˆ’ z)Ļˆ = Ī“1 , this sequence obeys (10.11) for n ā‰„ 2 and

b1 Ļˆ1 + a1 Ļˆ2 = zĻˆ1 + 1. The sequence Ļˆ can therefore be extended to a Weyl solution by setting Ļˆ0 so that a0 Ļˆ0 = āˆ’1. Thus, a Weyl solution exists, and āˆ’

Ļˆ1 = Ļˆ1 = Ī“1 , Ļˆ = Ī“1 , (J āˆ’ z)āˆ’1 Ī“1  = m(z). a0 Ļˆ0



We now deļ¬ne a procedure called coeļ¬ƒcient stripping. Starting from the Jacobi matrix J with parameters (an , bn )āˆž n=1 , consider the Jacobi matrix J1 āˆž with parameters (an , bn )n=2 . In terms of the shift operator S and its adjoint

312

10. Jacobi matrices

S āˆ— from (4.2) and (4.3), the coeļ¬ƒcient stripped Jacobi matrix can be written as J1 = SJS āˆ— . In the unbounded case, let us clarify that D(J1 ) = SD(J); in particular, coeļ¬ƒcient stripping also inherits a boundary condition at +āˆž, if any. Proposition 10.21. If m1 is the m-function corresponding to J1 , then for all z āˆˆ C+ , 1 m(z) = . (10.16) b1 āˆ’ z āˆ’ a21 m1 (z) Proof. If Ļˆ is a Weyl solution for J, then SĻˆ is a Weyl solution for J1 , so m(z) = āˆ’Ļˆ1 /(a0 Ļˆ0 ) and m1 (z) = āˆ’Ļˆ2 /(a1 Ļˆ1 ). By a direct calculation, m(z) = āˆ’

Ļˆ1 1 Ļˆ1 = = , a0 Ļˆ0 (b1 āˆ’ z)Ļˆ1 + a1 Ļˆ2 b1 āˆ’ z + a21 aĻˆ1 Ļˆ2 1

which implies (10.16).



Example 10.22. The free half-line Jacobi matrix is deļ¬ned by an = 1 and bn = 0 for all n āˆˆ N. It corresponds to the m-function āˆš āˆ’z + z 2 āˆ’ 4 m(z) = 2 āˆš 2 with the branch of z āˆ’ 4 on C \ [āˆ’2, 2], which takes positive values on (2, āˆž), and to the spectral measure . 1 dĪ¼(x) = Ļ‡(āˆ’2,2) (x) 4 āˆ’ x2 dx. (10.17) 2Ļ€ In particular, Ļƒ(J) = [āˆ’2, 2] and J has purely absolutely continuous spectrum. Proof. Since in this case J = J1 , coeļ¬ƒcient stripping gives 1 . m(z) = āˆ’z āˆ’ m(z) This turns into a quadratic equation for m(z), which has two solutions corresponding to the two branches of square root. Only the branch with positive values on (2, āˆž) corresponds to a Herglotzāˆš function. Since Im m(z) 1 extends continuously to R with values 2 Ļ‡[āˆ’2,2] (x) 4 āˆ’ x2 , (10.17) follows from Proposition 7.43.  Exercise 10.4 considers a related example, and Exercise 10.7 indicates the analogue of Weyl solutions for ļ¬nite Jacobi matrices. Corollary 10.23. Ļƒess (J) = Ļƒess (J1 ). On any interval I āŠ‚ R \ Ļƒess (J), the sets Ļƒd (J) and Ļƒd (J1 ) strictly interlace.

10.4. Transfer matrices and Weyl disks

313

Proof. Since J is unitarily equivalent to the operator Tx,dĪ¼(x) of multiplication by x on L2 (R, dĪ¼), it follows that Ļƒess (J) = ess supp Ī¼. The complement of this set is the largest domain on which m(z) has a meromorphic extension which obeys m(ĀÆ z ) = m(z). By (10.16), the functions m(z) and m1 (z) have meromorphic extensions to the same regions, which implies Ļƒess (J) = Ļƒess (J1 ). By Proposition 7.57, since m1 is a meromorphic Herglotz function on C \ Ļƒess (J), the zeros and poles of m1 (z) strictly interlace on I. By (10.16),  the poles of m1 (z) are precisely the zeros of m(z). We will combine the exact calculation for the free Jacobi matrix with the result about compact perturbations: Corollary 10.24. If J is a Jacobi matrix with an ā†’ 1 and bn ā†’ 0 as n ā†’ āˆž, then Ļƒess (J) = [āˆ’2, 2]. Proof. Denote by J0 the free Jacobi matrix, which has spectrum [āˆ’2, 2]; we will show that J āˆ’ J0 is compact. Denote by Pn an orthogonal projection to span{Ī“1 , . . . , Ī“n }. Then the operator Pn (J āˆ’ J0 )Pn is ļ¬nite rank and, as in the proof of Lemma 10.1, (J āˆ’ J0 ) āˆ’ Pn (J āˆ’ J0 )Pn  ā‰¤ 2 sup ak + sup|bk |. kā‰„n

k>n

This converges to 0 as n ā†’ āˆž, so J āˆ’ J0 is the norm limit of ļ¬nite rank  operators. Thus, J āˆ’ J0 is compact, so Ļƒess (J) = Ļƒess (J0 ) = [āˆ’2, 2].

10.4. Transfer matrices and Weyl disks Of course, the ļ¬rst-order matrix recursion (10.12) can be iterated: we deļ¬ne an n-step transfer matrix by Tn (z) = A(an , bn ; z) Ā· Ā· Ā· A(a1 , b1 ; z),

(10.18)

so that for any eigensolution v at z, ! ! vn+1 v1 = Tn (z) . an vn a0 v0 By comparing (10.11) with (10.9), we note that the sequence vn = pnāˆ’1 (z) is the eigensolution with a0 v0 = 0, v1 = 1, and therefore ! ! 1 pn (z) . (10.19) = Tn (z) 0 an pnāˆ’1 (z) Lemma 10.25. The polynomials pn and pnāˆ’1 have no common zeros.

314

10. Jacobi matrices

Proof. From det A(a, b; z) = 1 we conclude det Tn (z) = 1. In particular, Tn (z) is invertible, so (10.19) implies ! ! pn (z) 0 = .  0 an pnāˆ’1 (z) Of course, (10.19) means that the right-hand side is the ļ¬rst column of Tn (z). Considering the second column leads us to introduce the second kind polynomials qn (z) as the solution of the recursion zqn (z) = an qnāˆ’1 (z) + bn+1 qn (z) + an+1 qn+1 (z) with a0 qāˆ’1 = āˆ’1, q0 = 0. By induction, for n āˆˆ N, qn is a polynomial in z of degree n āˆ’ 1. Since vn = qnāˆ’1 (z) is the eigensolution with a0 v0 = āˆ’1, v1 = 0, ! ! 0 qn (z) , = Tn (z) āˆ’1 an qnāˆ’1 (z) so we can ļ¬nally conclude Tn =

! pn āˆ’qn . an pnāˆ’1 āˆ’an qnāˆ’1

It is sometimes useful to note that since det Tn = 1, the inverse is ! āˆ’an qnāˆ’1 qn āˆ’1 . Tn = āˆ’an pnāˆ’1 pn

(10.20)

(10.21)

In terms of the projective relation (7.4) on C2 , the formula (10.15) can be written as ! ! āˆ’m(z) Ļˆ1 . $ 1 a0 Ļˆ0 In terms of MĀØ obius transformations, the coeļ¬ƒcient stripping formula (10.16) can be rewritten as ! ! āˆ’m(z) āˆ’m1 (z) $ . A(a1 , b1 ; z) 1 1 These identities suggest that it would be better to conjugate by the matrix corresponding to the MĀØ obius transformation w ā†’ āˆ’w and consider transfer matrices ! ! 1 zāˆ’b āˆ’1 0 a a ? = jA(a, b; z)j, j= . A(a, b; z) = 0 1 āˆ’a 0 These would encode the Jacobi recursion for an arbitrary eigensolution at z by the formula ! ! āˆ’vn āˆ’vn+1 ? = A(an , bn ; z) , (10.22) an vn anāˆ’1 vnāˆ’1

10.4. Transfer matrices and Weyl disks

315

and their MĀØobius transformations would precisely correspond to action by coeļ¬ƒcient stripping, ! ! m(z) m1 (z) ? A(a1 , b1 ; z) $ . 1 1 Unfortunately, this is not the standard convention; we will only use it in this section because it ļ¬ts the Weyl disk formalism. Likewise, we deļ¬ne ? n , bn ; z) Ā· Ā· Ā· A(a ? 1 , b1 ; z) = jTn (z)j. T?n (z) = A(a These transfer matrices have the J -contracting property (see Deļ¬nition 7.9): Lemma 10.26. Fix z āˆˆ C+ . ? b; z) is J -contracting. (a) For any a > 0, b āˆˆ R, the matrix A(a, (b) For any n, J āˆ’ T?n (z)āˆ— J T?n (z) = 2 Im z

nāˆ’1  k=0

! 1 0 ? T?k (z)āˆ— T (z). 0 0 k

Proof. (a) follows from the calculation ? b; z)āˆ— J A(a, ? b; z) = J āˆ’ A(a,

! 2 Im z 0 . 0 0

(10.23)

(10.24)

Applying this to a = ak+1 , b = bk+1 , multiplying from the right by T?k (z)  and from the left by T?k (z)āˆ— , and summing in k proves (b). The identity (10.23) shows that the n-step transfer matrix (10.18) is also J -contracting, since the right-hand side is positive (compare Exercise 7.1). Moreover, it provides a J -monotonicity property which will lead to a nesting property below. Deļ¬nition 10.27. For z āˆˆ C+ and n āˆˆ N āˆŖ {0}, Weyl disks are deļ¬ned by & ! ! ' w āˆ—? w āˆ— ? Ė† ā‰„0 . Dn (z) = w āˆˆ C | Tn (z) J Tn (z) 1 1 Projectively, recall that ! ! w āˆ— w J ā‰„0 ā‡ā‡’ w āˆˆ C+ , 1 1   so w āˆˆ Dn (z) if and only if T?n (z) w1 corresponds to a point in C+ . Thus, obius transformation corresponding Dn (z) is the inverse image of C+ in the MĀØ to T?n (z). This proves: Lemma 10.28. For any nontrivial eigensolution v of J at z, vn+1 v1 āˆˆ Dn (z) ā‡ā‡’ āˆ’ āˆˆ C+ . āˆ’ a0 v0 an vn

(10.25)

316

10. Jacobi matrices

It is more customary in the literature to talk about the Weyl circles āˆ‚Dn (z) rather than the disks themselves; the circles are characterized by vn+1 v1 āˆˆ āˆ‚Dn (z) ā‡ā‡’ āˆ’ āˆˆ R āˆŖ {āˆž}. (10.26) āˆ’ a0 v0 an vn Lemma 10.29. For every n āˆˆ N and z āˆˆ C+ , the Weyl disk Dn (z) is a disk in C+ and the disks are nested in the sense that for all n āˆˆ N, Dn (z) āŠ‚ Dnāˆ’1 (z).

(10.27)

Proof. From (10.23), we have the identity T?nāˆ’1 (z) J T?nāˆ’1 (z) āˆ’ T?n (z) J T?n (z) = 2 Im z T?nāˆ’1 (z) āˆ—

āˆ—

āˆ—

! 1 0 ? (z). T 0 0 nāˆ’1

This implies the inequality T?nāˆ’1 (z)āˆ— J T?nāˆ’1 (z) ā‰„ T?n (z)āˆ— J T?n (z) from which the nesting property is immediate. Ė† under Since Dn (z) is the image of a half-plane (generalized disk in C) a MĀØobius transformation, to prove that it is a disk, it suļ¬ƒces to prove āˆžāˆˆ / Dn (z). Since āˆ’ av01v0 = āˆž corresponds to the solution with a0 v0 = 0, v1 = 1, we compute z āˆ’ b1 v2 =āˆ’ 2 āˆˆ / C+ . āˆ’ a1 v1 a1 / Dn (z) for all n by the nesting property. This implies āˆž āˆˆ / D1 (z), so āˆž āˆˆ



 Since the closed disks Dn (z) are nested, their intersection nāˆˆN Dn (z) is a point or a closed disk in C+ . This dichotomy corresponds to the limit point case in the sense of Deļ¬nition 10.13: Proposition 10.30. If J is in the limit point case, for any z āˆˆ C+ , Dn (z) = {m(z)}. nāˆˆN



Proof. Let w āˆˆ nāˆˆN Dn (z) and let (vn )āˆž n=0 be the eigensolution at z obeying a0 v0 = 1, v1 = āˆ’w. Then ! ! w āˆ’vn+1 ? = , Tn (z) an vn 1    āˆ— so multiplying (10.23) on the left by w1 and on the right by w1 , we obtain ! ! ! ! nāˆ’1  w āˆ’vn+1 āˆ— āˆ’vn+1 w āˆ— J āˆ’ J |vk+1 |2 . = 2 Im z an vn an vn 1 1 k=0

10.4. Transfer matrices and Weyl disks

317

The condition w āˆˆ Dn (z) rewrites to ! ! āˆ’vn+1 āˆ— āˆ’vn+1 J ā‰„ 0, an vn an vn so we obtain the inequality 2 Im z

nāˆ’1 

|vk+1 |2 ā‰¤

k=0

w 1

!āˆ—

J

w 1

! = 2 Im w.

The upper bound is independent of n, so v āˆˆ 2 (N). Since J is limit point, this shows that v is a Weyl solution for J. It follows that w = m(z), which concludes the proof.  The previous proof shows that for the Weyl solution normalized by a0 Ļˆ0 = 1, we have 2 Im z

n  |Ļˆk |2 = 2 Im m(z) + 2iWn (Ļˆ, Ļˆ), k=1

and the limit point condition guarantees Wn (Ļˆ, Ļˆ) ā†’ 0 as n ā†’ āˆž; thus, āˆž  k=1

|Ļˆk |2 =

Im m(z) . Im z

Exercise 10.9 proves further properties of Weyl solutions and Exercises 10.10 and 10.11 explore some further properties of Weyl disks. The connection between Weyl disks and the Herglotz function provides a valuable tool for deriving certain approximants of m(z). Proposition 10.31. If J is in the limit point case, uniformly on compact subsets of C+ , qn (z) . (10.28) m(z) = lim āˆ’ nā†’āˆž pn (z) Proof. From (10.21), for every n āˆˆ N, ! ! 0 āˆ’qn (z) $ T?n (z)āˆ’1 , pn (z) 1 so 0 āˆˆ RāˆŖ{āˆž} implies āˆ’qn (z)/pn (z) āˆˆ āˆ‚Dn (z). In particular, since Dn (z) āŠ‚ C+ , we observe that āˆ’qn (z)/pn (z) is a Herglotz function. By Proposition 10.30, the diameters of Dn (z) shrink to 0 as n ā†’ āˆž, and the limit (10.28) holds pointwise for every z āˆˆ C+ . Since the functions are Herglotz, by Proposition 7.28, pointwise convergence implies uniform  convergence on compact subsets of C+ .

318

10. Jacobi matrices

This technique is very robust: the main argument was that 0 āˆˆ C+ , so āˆ’qn (z)/pn (z) āˆˆ Dn (z). Other values lead to other approximants, some of which correspond to explicitly computable measures on R. In such cases, the approximations of the m-function lead to approximations of the spectral measure. We describe one such application, known as Carmonaā€™s formula. An important feature of Carmonaā€™s formula is that it allows the study of Ī¼ through the behavior of pn (x) as n ā†’ āˆž for real values of x. Theorem 10.32 (Carmona). If J is in the limit point case, for every h āˆˆ Cc (R),   1 h dĪ¼ = lim dx. (10.29) h(x) 2 nā†’āˆž Ļ€(pn (x) + a2n p2nāˆ’1 (x)) Proof. Deļ¬ne m(n) (z) by

! ! i m(n) (z) ? . $ Tn (z) 1 1

Since i āˆˆ C+ , m(n) (z) āˆˆ Dn (z) for all z āˆˆ C+ , so as n ā†’ āˆž, m(n) (z) converge pointwise to m(z). Since m(z) has no point mass at inļ¬nity, by (n) (n) Proposition 7.28, the measures  Ī¼ corresponding to m (z) converge to dĪ¼ in the sense that h dĪ¼n ā†’ h dĪ¼ for all h āˆˆ Cc (R). It remains to compute the measures Ī¼(n) . In terms of the entries of the transfer matrix, the property det T?n = 1 lets us easily compute T?āˆ’1 and explicitly write the MĀØobius transformation n

m(n) (z) = āˆ’

qn (z) + an qnāˆ’1 (z)i . pn (z) + an pnāˆ’1 (z)i

Since pn , pnāˆ’1 have real coeļ¬ƒcients and have no common zeros, m(n) (z) extends continuously to R; the imaginary part of the boundary value is Im m(n) (x) = āˆ’

an pn (x)qnāˆ’1 (x) āˆ’ an pnāˆ’1 (x)qn (x) . pn (x)2 + a2n pnāˆ’1 (x)2

Using again det Tn = 1 gives Im m(n) (x) =

p2n (x) +

1 , a2n p2nāˆ’1 (x)

and therefore

1 dx. + a2n p2nāˆ’1 (x)) As already stated, by Proposition 7.28, this completes the proof. dĪ¼(n) (x) =

Ļ€(p2n (x)

A common variation also known as Carmonaā€™s formula is   1 dx h(x) h dĪ¼ = lim 2 2 nā†’āˆž Ļ€(an pn (x) + p2nāˆ’1 (x))



(10.30)

10.5. Full-line Jacobi matrices

319

(note the diļ¬€erent placement of an compared to (10.29)); this can be proved by a similar proof (Exercise 10.13). For some purposes, it is better to use more specialized approximations. For decaying perturbations of the free Jacobi matrix (i.e., Jacobi matrices J with an ā†’ 1, bn ā†’ 0), we know that Ļƒess (J) = [āˆ’2, 2], so particular attention is focused on determining the spectral type on [āˆ’2, 2]. In such cases, the denominator in Exercise 10.14 often has better behavior than the one in Carmonaā€™s formula, for x āˆˆ (āˆ’2, 2).

10.5. Full-line Jacobi matrices We now turn our attention to full line or two-sided Jacobi matrices, which are operators on 2 (Z) with a matrix representation āŽž āŽ› .. .. . . āŽŸ āŽœ āŽŸ āŽœ. . āŽŸ āŽœ . bāˆ’1 aāˆ’1 āŽŸ āŽœ āŽŸ āŽœ a b a āˆ’1 0 0 āŽŸ āŽœ āŽŸ. āŽœ (10.31) a0 b1 a1 J =āŽœ āŽŸ āŽŸ āŽœ a b a 1 2 2 āŽŸ āŽœ āŽœ .. .. āŽŸ āŽœ . .āŽŸ a2 āŽ  āŽ .. . The sequences of coeļ¬ƒcients an āˆˆ (0, āˆž), bn āˆˆ R are now indexed by n āˆˆ Z. +āˆž āˆž If (an )+āˆž n=āˆ’āˆž , (bn )n=āˆ’āˆž āˆˆ  (Z), then J is a bounded self-adjoint operator deļ¬ned precisely by

(Ju)n = anāˆ’1 unāˆ’1 + bn un + an un+1 for any u āˆˆ 2 (Z), by the same proof as in Lemma 10.1. If at least one of the sequences is not bounded, J will also be an unbounded operator. To motivate the choice of domain, let us denote by P+ and Pāˆ’ orthogonal projections to span{Ī“n | n > 0} and span{Ī“n | n ā‰¤ 0} and formally decompose with respect to subspaces Ran PĀ± as follows, āŽž āŽ› .. .. . āŽŸ āŽœ . āŽŸ āŽœ .. āŽŸ āŽœ . bāˆ’1 aāˆ’1 āŽŸ āŽœ āŽŸ āŽœ aāˆ’1 b0 a0 āŽŸ āŽœ āŽŸ. āŽœ J =āŽœ a0 b1 a1 āŽŸ āŽŸ āŽœ a1 b2 a2 āŽŸ āŽœ āŽœ .. .. āŽŸ āŽœ . . āŽŸ a2 āŽ  āŽ .. .

320

10. Jacobi matrices

The lower right block is a half-line Jacobi matrix J+ on Ran P+ and the upper left part is, up to a reļ¬‚ection of the real line, a half-line Jacobi matrix Jāˆ’ on Ran Pāˆ’ . The remaining a0 entries are collected into a ļ¬nite rank self-adjoint operator a0 F where F u = Ī“0 , uĪ“1 + Ī“1 , uĪ“0 , and we write J = Jāˆ’ āŠ• J+ + a0 F.

(10.32)

This motivates the following deļ¬nition: Deļ¬nition 10.33. A full-line self-adjoint Jacobi matrix J with separated boundary conditions is deļ¬ned in terms of two self-adjoint half-line Jacobi matrices JĀ± and a coeļ¬ƒcient a0 > 0 as an operator with domain D(J) = D(Jāˆ’ ) āŠ• D(J+ ), acting on elements of the domain by (10.32). Since F is bounded self-adjoint, āˆ— āˆ— (Jāˆ’ āŠ• J+ + a0 F )āˆ— = (Jāˆ’ āŠ• J+ )āˆ— + a0 F = Jāˆ’ āŠ• J+ + a0 F = Jāˆ’ āŠ• J+ + a0 F,

so the operator J deļ¬ned above is indeed self-adjoint. Since ļ¬nite rank perturbations do not change the essential spectrum, this decomposition immediately implies Ļƒess (J) = Ļƒess (J+ ) āˆŖ Ļƒess (Jāˆ’ ).

(10.33)

If JĀ± are both in the limit point case, then there are no boundary conditions, so 

 |anāˆ’1 unāˆ’1 + bn un + an un+1 |2 < āˆž . D(J) = u āˆˆ 2 (Z) | nāˆˆZ

Formal eigensolutions at z are now sequences u = (un )+āˆž n=āˆ’āˆž which solve anāˆ’1 unāˆ’1 + bn un + an un+1 = zun

āˆ€n āˆˆ Z.

There are now unique, up-to-normalization Weyl solutions Ļˆ Ā± (z) for each half-line JĀ± ; each Weyl solution can be extended uniquely as an eigensolution on Z. Example 10.34. The free full-line Jacobi matrix J is deļ¬ned by an = 1 and bn = 0 for all n āˆˆ Z. For any z āˆˆ C+ , Weyl solutions normalized by Ļˆ0Ā± (z) = 1 are given by Ā±n  āˆš z āˆ’ z2 āˆ’ 4 Ā± , n āˆˆ Z, Ļˆn (z) = 2

10.5. Full-line Jacobi matrices

with the branch of (2, āˆž).

321

āˆš z 2 āˆ’ 4 on C \ [āˆ’2, 2] which takes positive values on

Proof. The m-function for the half-line free Jacobi matrix is āˆš āˆ’z + z 2 āˆ’ 4 m(z) = . 2 Coeļ¬ƒcient stripping does not aļ¬€ect the free Jacobi matrix, so for all n, āˆ’

+ (z) Ļˆn+1

Ļˆn+ (z)

=āˆ’

+ Ļˆn+1 (z)

an Ļˆn+ (z)

= m(z).

By induction we ļ¬nd Ļˆn+ (z) = (āˆ’m(z))n . By a reļ¬‚ection we ļ¬nd Ļˆnāˆ’ (z).  Weyl solutions exist for z āˆˆ / Ļƒess (J Ā± ). Another easy observation is that, for z āˆˆ / Ļƒ(J), the Weyl solutions Ļˆ āˆ’ and Ļˆ + are linearly independent; otherwise, one nontrivial solution would obey the conditions at both endpoints, so it would be an eigenvector of the operator J. Denote the Wronskian of Ļˆ Ā± by + āˆ’ (z)Ļˆnāˆ’ (z) āˆ’ Ļˆn+1 (z)Ļˆn+ (z)], W (z) = an [Ļˆn+1

which is independent of n. Denote also the Greenā€™s function Gm,n (z) =

āˆ’ + Ļˆmin(m,n) (z)Ļˆmax(m,n) (z)

W (z)

.

This turns out to be the integral kernel for the resolvent of J (of course, the integral here is with respect to the counting measure on Z): Proposition 10.35. For any u āˆˆ 2 (Z) and z āˆˆ C \ Ļƒ(J),  Gn,m (z)un . [(J āˆ’ z)āˆ’1 u]m =

(10.34)

māˆˆZ

Proof. Fix m and denote v = (Gm,n )+āˆž n=āˆ’āˆž . Gm,n is for n ā‰„ m a multiple of Ļˆn+ , and for n ā‰¤ m a ļ¬xed multiple of Ļˆnāˆ’ . In particular, v āˆˆ D(J). The same observation shows ((J āˆ’ z)v)n = 0

āˆ€n = m.

Meanwhile, ((J āˆ’ z)v)m =

āˆ’ + (z) + b Ļˆ āˆ’ (z)Ļˆ + (z) + a Ļˆ āˆ’ (z)Ļˆ + (z) amāˆ’1 Ļˆmāˆ’1 (z)Ļˆm m m m m m m+1 , W (z)

and using the eigenfunction equation for Ļˆ āˆ’ turns this into ((J āˆ’ z)v)m =

āˆ’ + (z) + a Ļˆ āˆ’ (z)Ļˆ + (z) āˆ’am Ļˆm+1 (z)Ļˆm m m m+1 = 1. W (z)

322

10. Jacobi matrices

Thus, (J āˆ’ z)v = Ī“m , so (10.34) holds for u = Ī“m . By linearity, it holds for all u āˆˆ 2c (Z). For arbitrary u āˆˆ 2 (Z), apply the above to compactly supported vectors uĻ‡[āˆ’N,N ] . Using uĻ‡[āˆ’N,N ] ā†’ u as N ā†’ āˆž, the left-hand side of (10.34) converges in the 2 sense, so it also converges pointwise. The right-hand side also converges for each m, so the limits are equal. 

10.6. Eigenfunction expansion for full-line Jacobi matrices In the full-line case, the spectrum can be of multiplicity 2, as already seen in the following example. For this example, the reader should recall the unitary dk ) ā†’ 2 (Z) and its inverse F āˆ’1 which correspond to map F : L2 ([0, 2Ļ€], 2Ļ€ the usual Fourier series expansion (Example 3.41). Example 10.36. The free full-line Jacobi matrix is deļ¬ned on 2 (Z) by (Ju)n = unāˆ’1 + un+1 . dk ). Then F āˆ’1 JF is the operator of multiplication by 2 cos k on L2 ([0, 2Ļ€], 2Ļ€ In particular, J has a purely absolutely continuous spectrum of multiplicity 2.

Proof. Since F āˆ’1 JF and multiplication by 2 cos k are bounded operators dk ), it suļ¬ƒces to prove that they agree on the dense set of on L2 ([0, 2Ļ€], 2Ļ€ compactly supported sequences. By linearity, it suļ¬ƒces to prove that they agree on the vectors Ī“n . For this, use JĪ“n = Ī“nāˆ’1 + Ī“n+1 to compute F āˆ’1 JF eink = ei(nāˆ’1)k + ei(n+1)k = 2 cos k eink . 

The remaining claims follow from Example 9.40.

Thus, there is no hope in general for a cyclic vector. However, we will see that every Ī“n can be obtained from Ī“0 and Ī“1 using polynomials of J, and this will be a starting point in the canonical construction of a unitary map that diagonalizes the full-line Jacobi matrix. This canonical unitary map will naturally be in terms of a matrix-valued measure, which leads us to use matrix-valued analogues of multiplication operators and Herglotz functions. Let us consider eigenfunctions u0 , u1 at x, which solve anāˆ’1 unāˆ’1 + bn un + an un+1 = xun

āˆ€n āˆˆ Z,

(10.35)

and the initial conditions u10 (x) = 0,

u11 (x) = 1,

u00 (x) = 1,

u01 (x) = 0.

For any n āˆˆ Z, u0n (x) and u1n (x) are polynomials in x. The following is the analogue of the formula Ī“n = pnāˆ’1 (J)Ī“1 from the half-line case.

10.6. Eigenfunction expansion for full-line Jacobi matrices

323

Lemma 10.37. For any n āˆˆ Z, Ī“n = u1n (J)Ī“1 + u0n (J)Ī“0 .

(10.36)

Proof. For n āˆˆ Z, let us denote Ļˆn = u1n (J)Ī“1 + u0n (J)Ī“0 āˆˆ 2 (Z). By the Borel functional calculus, since uj are solutions of (10.35) for j = 0, 1, Jujn (J) = anāˆ’1 ujnāˆ’1 (J) + bn ujn (J) + an ujn+1 (J). Applying this to Ī“j and summing in j = 0, 1 implies JĻˆn = anāˆ’1 Ļˆnāˆ’1 + bn Ļˆn + an Ļˆn+1 . Since the same holds for the sequence of Ī“n , it suļ¬ƒces to note that Ļˆ0 = Ī“0 and Ļˆ1 = Ī“1 and proceed by induction in Ā±n.  This lemma can easily be used to construct a spectral basis for J with at most two elements starting from Ī“0 , Ī“1 . Already from this, it could be concluded that the spectrum of J can have multiplicity at most 2, but we are about to present a more precise spectral representation. Corresponding to the Jacobi matrix J, we deļ¬ne the 2 Ɨ 2 matrix-valued measure Ī© = (Ī©i,j )1i,j=0 by the property that for all Borel sets B,  (10.37) Ī“i , Ļ‡B (J)Ī“j  = Ļ‡B dĪ©i,j (see discussion of matrix-valued measures and corresponding L2 spaces in Ī»0  Section 6.4). For any Borel set B āŠ‚ R and any Ī» = Ī»1 āˆˆ C2 , if we denote v = 1j=0 Ī»j Ī“j , then āˆ—

Ī» Ī©(B)Ī» =

1 

Ī»i Ī“i , Ļ‡B (J)Ī“j Ī»j = v, Ļ‡B (J)v ā‰„ 0,

i,j=0

so Ī© is a positive matrix-valued measure. In particular, it can be written as dĪ© = W dĪ¼, where Ī¼ = Tr Ī© is a ļ¬nite positive measure and W is a matrix-valued function with W ā‰„ 0 and Tr W = 1 Ī¼-a.e. We will construct a unitary map U from 2 (Z) to the Hilbert space which conjugates J to a multiplication operator. We begin by introducing the map pointwise on a dense subset of 2 (Z): For f āˆˆ 2c (Z), we deļ¬ne !  u0n Ė† fn 1 . f (Ī») = un L2 (R, C2 , dĪ©),

nāˆˆZ

324

10. Jacobi matrices

Conversely, for g āˆˆ L2 (R, C2 , dĪ©), we deļ¬ne !  u0n gĖ‡n = W g dĪ¼. u1n Theorem 10.38 (Eigenfunction expansion for full-line Jacobi matrices). Let J be a full-line self-adjoint Jacobi matrix. There is a unitary map U : 2 (Z) ā†’ L2 (R, C2 , dĪ©) such that: (a) U f = fĖ† for all f āˆˆ 2c (Z); (b) U āˆ’1 g = gĖ‡ for all g āˆˆ L2c (R, C2 , dĪ©); (c) U JU āˆ’1 = TĪ»,dĪ©(Ī») . Proof. By the deļ¬nition and basic properties of the matrix-valued measure dĪ© = W dĪ¼, for any m, n āˆˆ Z and any compactly supported bounded function h : R ā†’ C, !āˆ— !  1   u0n u0m W huim ujn Wi,j dĪ¼ dĪ¼ = h 1 um u1n i,j=0

=

1 

uim (J)Ī“i , h(J)ujn (J)Ī“j 

i,j=0

/ =

1 

uim (J)Ī“i , h(J)

i=0

1 

0 ujn (J)Ī“j

j=0

= Ī“m , h(J)Ī“n . First, we apply this calculation with m = n and h = Ļ‡[āˆ’k,k] and use monotone convergence as k ā†’ āˆž to conclude that Ī“Ė†n āˆˆ L2 (R, C2 , dĪ©) for each n, and then fĖ† āˆˆ L2 (R, C2 , dĪ©) for each f āˆˆ 2c (Z). By sesquilinearity, the same calculation shows that  āˆ€f1 , f2 āˆˆ 2c (Z). hfĖ†1 W fĖ†2 dĪ¼ = f1 , h(J)f2  This allows us to apply Theorem 9.48 with A = J and B = TĪ»,dĪ©(Ī») ; we conclude that the map f ā†’ fĖ† extends to a norm-preserving map U and there exists a linear map U āˆ— : L2 (R, C2 , dĪ©) ā†’ 2 (Z) with U āˆ—  ā‰¤ 1 and U āˆ— g, f  = g, U f  for all f, g in the appropriate Hilbert spaces. For f āˆˆ 2c (Z) and g āˆˆ L2c (R, C2 , dĪ©), !   u0n āˆ— Ė† g W 1 fn dĪ¼ = g, f. un

10.7. The Weyl M -matrix

325

Formally, by placing the sum inside the integral, this looks like Ė‡ g , f ; applying this to f = gĖ‡Ļ‡[āˆ’n,n] and letting n ā†’ āˆž proves by monotone convergence Ė† for a dense set of f implies that gĖ‡ āˆˆ 2 (Z). Now the equality Ė‡ g , f  = g, f āˆ— 2 2 that U g = gĖ‡ for all g āˆˆ Lc (R, C , dĪ©). Moreover, Theorem 9.48 says that Ker U āˆ— = (Ran U )āŠ„ is a resolventinvariant subspace for multiplication by Ī». Fix g āˆˆ Ker U āˆ— . Since the subspace is resolvent-invariant, for any Ī»1 < Ī»2 , Ļ‡(Ī»1 ,Ī»2 ] g āˆˆ Ker U āˆ— . Thus, 

u0n u1n

! W Ļ‡(Ī»1 ,Ī»2 ] g dĪ¼ = 0

āˆ€n āˆˆ Z.

Evaluating this at n = 0 and n = 1 implies that  W Ļ‡(Ī»1 ,Ī»2 ] g dĪ¼ = 0, and since Ī»1 < Ī»2 are arbitrary, W g = 0 Ī¼-a.e. Thus, g āˆ— W g = 0 Ī¼-a.e., so g = 0 in L2 (R, C2 , dĪ©). This proves Ker U āˆ— = {0}, which concludes the proof by Theorem 9.48(f).  In the case of bounded self-adjoint Jacobi matrices, we could have argued more directly that U is onto by proving that its image contains every P  polynomial Q ; this proof would proceed by induction in deg P + deg Q, using the degrees of u0n , u1n . The eigenfunction expansion provides us with a canonical matrix-valued spectral measure Ī© for the self-adjoint operator J. Corollary 10.39. The measure Ī¼ = Tr Ī© is a maximal spectral measure for J. The spectral multiplicity n measures for J (see Theorem 9.31) are given by dĪ¼n = Ļ‡Sn dĪ¼, where Sn = {Ī» | rank W (Ī») = n} (in particular, Sn = 0 unless n = 1 or 2). The measure Ī¼ = Tr Ī© is regarded as the canonical spectral measure for the full-line Jacobi matrix. Note that Ī¼ = Ī¼J,Ī“0 + Ī¼J,Ī“1 , which relates to the special role Ī“0 , Ī“1 had in our construction.

10.7. The Weyl M -matrix In this section we continue studying the full-line setting from the previous sections. We have already observed that the essential spectrum of J is the union of essential spectra of JĀ± . To obtain ļ¬ner connections between

326

10. Jacobi matrices

spectral properties of J and JĀ± , we introduce the Weyl M -matrix. This is the Borel transform of the matrix-valued spectral measure,  1 M (z) = dĪ©(Ī»). Ī»āˆ’z This is a matrix-valued Herglotz function. In particular, Tr M (z) is the Borel transform of the canonical spectral measure Ī¼. Integrating each entry, we obtain a formula in terms of the Greenā€™s function, ! G0,0 (z) G0,1 (z) . M (z) = G1,0 (z) G1,1 (z) The Weyl M -matrix can be expressed in terms of the half-line m-functions m+ (z) = Ī“1 , (J+ āˆ’ z)āˆ’1 Ī“1 , māˆ’ (z) = Ī“0 , (Jāˆ’ āˆ’ z)āˆ’1 Ī“0 . Lemma 10.40.

!āˆ’1 a0 māˆ’1 āˆ’ . M= a0 māˆ’1 + In particular, the diagonal Greenā€™s function elements are given by 1 2 āˆ’ = āˆ’māˆ’1 + + a0 māˆ’ , G1,1 1 2 = āˆ’māˆ’1 āˆ’ āˆ’ + a0 m+ . G0,0

(10.38)

(10.39) (10.40)

Proof. By the second resolvent identity, (J āˆ’ z)āˆ’1 = (Jāˆ’ āŠ• J+ āˆ’ z)āˆ’1 āˆ’ a0 (J āˆ’ z)āˆ’1 F (Jāˆ’ āŠ• J+ āˆ’ z)āˆ’1 . Applying this to Ī“0 and Ī“1 and then taking the inner product with Ī“0 and Ī“1 gives four equalities: G0,0 = māˆ’ āˆ’ a0 G0,1 māˆ’ , G0,1 = āˆ’a0 G0,0 m+ , G1,0 = āˆ’a0 G1,1 māˆ’ , G1,1 = m+ āˆ’ a0 G1,0 m+ . Viewing the ļ¬rst two as a set of linear equations for G0,0 and G0,1 , and the second two as a set of linear equations for G1,0 and G1,1 , we can rewrite the systems in the combined form ! ! ! a0 G0,0 G0,1 1 0 māˆ’1 āˆ’ = , G1,0 G1,1 0 1 a0 māˆ’1 + which proves (10.38). The formulas for the diagonal Greenā€™s function follow easily. 

10.7. The Weyl M -matrix

327

This allows us to express spectral properties of J in terms of Weyl functions mĀ± , which allows us to relate spectral properties of J to the spectral properties of half-line operators JĀ± . For instance, it provides a second proof of (10.33). Namely, the functions mĀ± are Herglotz functions with region of meromorphicity C\Ļƒess (JĀ± ), so the same is true for āˆ’1/mĀ± . By (10.39) and (10.40), G0,0 and G1,1 have region of meromorphicity C\(Ļƒess (J+ )āˆŖĻƒess (Jāˆ’ )), and so does their sum. Since the Herglotz function G0,0 + G1,1 corresponds to the maximal spectral measure for J, (10.33) follows. More importantly, (10.38) provides immediate consequences of results proved in an abstract setting in Section 7.12. To relate it to that setting, we write !āˆ’1 !āˆ’1 āˆ’aāˆ’1 māˆ’1 āˆ’1 āˆ’1 āˆ’m Ėœ āˆ’1 āˆ’ āˆ’ 0 =āˆ’ , a0 M = āˆ’ āˆ’1 āˆ’1 m Ėœ+ āˆ’1 āˆ’aāˆ’1 0 m+ where āˆ’1 m Ėœ + = āˆ’aāˆ’1 0 m+ ,

m Ėœ āˆ’ = a0 māˆ’ .

This puts a0 M in the setting of Section 7.12, up to an additive constant self-adjoint matrix which does not aļ¬€ect the phenomena studied here. Corollary 10.41. The absolutely continuous spectrum of J is precisely the sum of absolutely continuous spectra of JĀ± , with multiplicities added, i.e., J|Hac (J) āˆ¼ = J+ |Hac (J+ ) āŠ• Jāˆ’ |Hac (Jāˆ’ ) . Corollary 10.42. Let Ī¼s denote the singular part of the canonical spectral measure for the full-line Jacobi matrix J. For Ī¼s -a.e. x āˆˆ R, (a) rank W (x) = 1, (b) mĀ± have normal limits which are values in R āˆŖ {āˆž}, (c) there exists Ī± = Ī±(x) āˆˆ [0, Ļ€) such that (a0 m+ (x + i0))āˆ’1 = a0 māˆ’ (x + i0) = cot Ī±(x) and W (x) =

! cos2 Ī±(x) āˆ’ cos Ī±(x) sin Ī±(x) . āˆ’ cos Ī±(x) sin Ī±(x) sin2 Ī±(x)

In other words, the singular part of the maximal spectral measure is supported on the set  {x | (a0 m+ (x + i0))āˆ’1 = a0 māˆ’ (x + i0) = cot Ī±}. S= Ī±āˆˆ[0,Ļ€)

Moreover, S has Lebesgue measure zero.

328

10. Jacobi matrices

10.8. Subordinacy theory Spectral properties of Jacobi matrices may be studied through the behavior of formal eigensolutions with real spectral parameters. The pure point spectrum of a Jacobi matrix corresponds to eigenvectors, which are simply formal eigensolutions which are square-integrable. Carmonaā€™s formula allows us to recover the measure, but it is not a pointwise criterion. Subordinacy theory was discovered by Khanā€“Pearson [54] with important developments by Gilbertā€“Pearson [41], Gilbert [40], and Jitomirskayaā€“Last [47, 48]. It uses a pointwise characterization in terms of the behavior of eigensolutions to describe the decomposition into absolutely continuous/singular spectra and, more generally, decomposition into Ī±-continuous/Ī±-singular spectra. We begin with the half-line setting. Deļ¬nition 10.43. Fix Ī» āˆˆ R. A nontrivial eigensolution u at Ī» is called subordinate (at +āˆž) if n 2 j=1 |uj | =0 (10.41) lim n 2 nā†’āˆž j=1 |vj | for some eigensolution v at Ī». It also helps to consider a continuous interpolation of 2 -norms: for L > 0, deļ¬ne u2L =

L 

|uk |2 + (L āˆ’ L)|uL+1 |2 .

k=1

Lemma 10.44. (a) If (10.41) holds for some eigensolution v, then it holds for every eigensolution v linearly independent with u. (b) If u is subordinate, it is linearly dependent with u. Then, u is a constant multiple of a real-valued eigensolution. (c) (10.41) holds if and only if uL = 0. Lā†’āˆž vL lim

(10.42)

Proof. (a) If v = Cu, the limit (10.41) is 1/C 2 . Thus, (10.41) implies that v is linearly independent with u. Now any eigensolution w can be written as w = C1 u + C2 v and, if w is linearly independent with u, then C2 = 0. By elementary estimates, 1 |w|2 ā‰„ |C2 |2 |v|2 āˆ’ |C1 |2 |u|2 . 2

10.8. Subordinacy theory

329

This implies

n n 2 |wk |2 1 2 k=1 k=1 |vk | |C ā‰„ | lim inf āˆ’ |C1 |2 = āˆž, lim inf n 2 n 2 2 nā†’āˆž nā†’āˆž |u | 2 |u | k=1 k k=1 k

and inverting completes the proof. (b) If v = u, then the limit in (10.41) is equal to 1. Thus, if u is subordinate, u must be linearly dependent with u. Thus, vectors (u1 , u0 ) and (u1 , u0 ) are linearly dependent, so u1 /u0 āˆˆ R āˆŖ {āˆž}. Thus, by a constant multiplicative factor, one can make (u1 , u0 ) āˆˆ R2 \ {0}. (c) It is a simple analysis that for any n, & ' u2n u2n+1 u2n + t|un+1 |2 = max , max 2 v2n v2n+1 tāˆˆ[0,1] v2 n + t|vn+1 | since the function of t is monotone. This implies equivalence of (10.41) and (10.42).  This narrows our focus to the question of when some real-valued eigensolution is subordinate. The following inequality relates this to the behavior of the m-function: Lemma 10.45 (Jitomirskayaā€“Last inequality). For any L > 0, deļ¬ne 1 . (10.43)

(L) = 2pĀ·āˆ’1 L qĀ·āˆ’1 L Then (L) āˆˆ (0, āˆž) and for all L > 0, āˆš āˆš pĀ·āˆ’1 L 5 + 24 5 āˆ’ 24 ā‰¤ . ā‰¤ |m(Ī» + i (L))| qĀ·āˆ’1 L |m(Ī» + i (L))|

(10.44)

Proof. Consider the Weyl solution Ļˆn (z) for z = Ī» + i , normalized by a0 Ļˆ0 = 1 so that Ļˆ1 = āˆ’m(z). We use variation of parameters to compare this to eigensolutions at Ī»: we deļ¬ne ! āˆ’1 Ļˆn+1 vn = Tn (Ī») an Ļˆn and derive

  vn āˆ’ vnāˆ’1 = Tnāˆ’1 (Ī»)āˆ’1 A(an , bn ; Ī»)āˆ’1 A(an , bn ; z) āˆ’ I Tnāˆ’1 (Ī»)vnāˆ’1

which simpliļ¬es to vn āˆ’ vnāˆ’1 = Tnāˆ’1 (Ī»)

āˆ’1

0 0 āˆ’i 0

! Ļˆn . anāˆ’1 Ļˆnāˆ’1

āˆ’m(z) , we obtain 1 ! ! nāˆ’1  āˆ’m(z) 0 āˆ’1 āˆ’ i

Tk (Ī») . Ļˆk+1 1

By telescoping and using v0 = vn =

!

k=0

330

10. Jacobi matrices

Multiplying on the left by (1, 0)Tn (Ī») gives Ļˆn+1 = āˆ’qn (Ī») āˆ’ pn (Ī»)m(z) āˆ’ i

nāˆ’1 

(pn (Ī»)qk (Ī») āˆ’ qn (Ī»)pk (Ī»)) Ļˆk+1 .

k=0

Using Cauchyā€“Schwarz twice on the right-hand side implies |Ļˆn+1 | ā‰„ |qn (Ī») + pn (Ī»)m(z)| āˆ’ |pn (Ī»)|qĀ·āˆ’1 L ĻˆL āˆ’ |qn (Ī»)|pĀ·āˆ’1 L ĻˆL . Rearranging and using the triangle inequality for Ā·L gives qĀ·āˆ’1 (Ī») + pĀ·āˆ’1 (Ī»)m(z)L ā‰¤ ĻˆL + 2 pĀ·āˆ’1 L qĀ·āˆ’1 L ĻˆL . Squaring this, combining with Ļˆ2L ā‰¤ Ļˆ2 = Im m(z)/ , and using =

(L), we obtain qĀ·āˆ’1 (Ī») + pĀ·āˆ’1 (Ī»)m(z)2L ā‰¤ 8pĀ·āˆ’1 L qĀ·āˆ’1 L |m(z)|. Using the triangle inequality in the left-hand side, this implies (qĀ·āˆ’1 (Ī»)L āˆ’ |m(z)|pĀ·āˆ’1 (Ī»)L )2 ā‰¤ 8pĀ·āˆ’1 L qĀ·āˆ’1 L |m(z)|. Dividing this by qĀ·āˆ’1 (Ī»)2L and expanding gives a quadratic inequality for Īŗ = |m(z)|pĀ·āˆ’1 (Ī»)L /qĀ·āˆ’1 (Ī»)L , which implies 5 āˆ’

āˆš

Īŗ2 āˆ’ 10Īŗ + 1 ā‰¤ 0, āˆš 24 ā‰¤ Īŗ ā‰¤ 5 + 24, completing the proof.



Subordinacy of (pnāˆ’1 (Ī»))āˆž n=1 corresponds to inļ¬nite normal boundary values of m: Theorem 10.46. Consider a half-line Jacobi matrix J in the limit point case with Weyl function m(z). The solution (pnāˆ’1 (Ī»))āˆž n=1 is subordinate if and only if (10.45) lim m(Ī» + i ) = āˆž. ā†“0

Proof. The function (L) deļ¬ned above is a continuous, strictly decreasing āˆž function of L. The sequences (pnāˆ’1 (Ī»))āˆž n=1 and (qnāˆ’1 (Ī»))n=1 are not both square-summable due to the limit point condition, so lim (L) = 0.

Lā†’āˆž

By taking L ā†’ āˆž in the Jitomirskayaā€“Last inequality, we conclude that (pnāˆ’1 (Ī»))āˆž n=1 is subordinate if and only if lim |m(Ī» + i (L))| = āˆž.

Lā†’āˆž

By properties of (L), this is equivalent to (10.45).



We will now extend this to a characterization of subordinacy of an arbitrary real solution, using a trick of varying the b1 coeļ¬ƒcient in the Jacobi matrix, which is of some independent interest.

10.8. Subordinacy theory

331

cos Ī±10.47. Fix Ī» āˆˆ R and Ī± āˆˆ R. The eigensolution at Ī» with Corollary u1  u0 $ sin Ī± is subordinate if and only if lim a0 m(Ī» + i ) = āˆ’ cot Ī±. ā†“0

Proof. We separate cases and reduce each case to Theorem 10.46. The case Ī± = 0 is Theorem 10.46, and the case Ī± = Ļ€/2 follows by applying Theorem 10.46 to the coeļ¬ƒcient-stripped matrix J1 = SJS āˆ— . For Ī± āˆˆ R \ Ļ€2 Z, note that an eigensolution (un )āˆž n=1 obeys a0 u0 + b1 u1 + a1 u2 = Ī»u1 . By grouping the ļ¬rst two terms, we conclude that the eigensolution obeys u1 /u0 = cot Ī± if and only if (b1 + a0 tan Ī±)u1 + a1 u2 = Ī»u1 , so the same eigensolution (un )āˆž n=1 corresponds to orthonormal polynomials for the modiļ¬ed Jacobi matrix J (Ī±) = J + a0 tan Ī±Ī“1 , Ā·Ī“1 . Denote by m(Ī±) its m-function. Since J and J (Ī±) have the same coeļ¬ƒcientstripped Jacobi matrix J1 , coeļ¬ƒcient stripping gives ! ! ! zāˆ’b1 1 m(z) m1 (z) a1 a1 $ 1 1 āˆ’a1 0 and

zāˆ’b1 āˆ’a0 tan Ī± a1

āˆ’a1

1 a1

0

!

! m(Ī±) (z) $ 1

! m1 (z) . 1

Combining these equations to express m(Ī±) in terms of m, we obtain ! ! ! ! zāˆ’b1 āˆ’a0 tan Ī± 1 āˆ’1 zāˆ’b1 1 m(z) m(Ī±) (z) a1 a1 a1 a1 , $ 1 1 āˆ’a1 0 āˆ’a1 0 so m(z) . m(Ī±) (z) = a0 tan Ī±m(z) + 1 In particular, lim m(Ī±) (z) = āˆž ā‡ā‡’ lim a0 m(z) = āˆ’ cot Ī±, ā†“0

ā†“0

so applying Theorem 10.46 to J (Ī±) concludes the proof.



Theorem 10.48. Let J be a half-line Jacobi matrix in the limit point case. The singular part of its canonical spectral measure Ī¼ is supported on the set S = {Ī» āˆˆ R | (pnāˆ’1 (Ī»))āˆž n=1 is subordinate},

332

10. Jacobi matrices

and the absolutely continuous part of Ī¼ mutually absolutely continuous with Ļ‡N (Ī») dĪ», where N = {Ī» āˆˆ R | there is no subordinate solution at Ī»}. Proof. The set S is precisely the set on which m(Ī» + i0) = āˆž. Moreover, Ī» āˆˆ N if and only if m(Ī»+i0) āˆˆ C+ or m(Ī»+i0) does not exist; however, the second case happens on a set of Lebesgue measure zero. Thus, the theorem follows from Corollary 7.49.  The most commonly used consequence of this is a criterion for absolutely continuous spectrum in terms of bounded eigensolutions [6, 85, 103]: Theorem 10.49. Let J be a half-line Jacobi matrix with sup an < āˆž. Let Ī» āˆˆ R. If all eigensolutions at Ī» are bounded, then there is no subordinate solution at Ī». In particular, on the set Sāˆž of such Ī», Ļ‡Sāˆž dĪ¼ is mutually absolutely continuous with Ļ‡Sāˆž (Ī») dĪ». Proof. Lt u, v be linearly independent eigensolutions at Ī». Then their Wronskian W is nonzero and |W | ā‰¤ |an | (|un ||vn+1 | + |vn ||un+1 |) ā‰¤ |an | (|un | + |un+1 |) vāˆž . This implies |un |2 + |un+1 |2 ā‰„ so

|W |2 1 |uk |2 ā‰„ . n |an |2 v2āˆž n

lim inf nā†’āˆž

Since

|W |2 , 2|an |2 v2āˆž

lim supnā†’āˆž n1

n

k=1

k=1 |vk

|2

ā‰¤ v2āˆž , this implies n |uk |2 lim inf k=1 > 0, n 2 nā†’āˆž k=1 |vk | 

so u is not subordinate.

Further criteria for absolutely continuous spectrum have been proved by Lastā€“Simon [61] with closely related work by Remling [77] (see also [92, Sections 7.3 and 7.4]). By strengthening the subordinacy assumption, we can characterize spectral decompositions with respect to Hausdorļ¬€ measures. (Note the lim inf in the following deļ¬nition.) Deļ¬nition 10.50. Fix Ī² āˆˆ (0, 1] and Ī» āˆˆ R. A nontrivial eigensolution u at Ī» is called Ī²-subordinate (at +āˆž) if lim inf nā†’āˆž

2āˆ’Ī² uL

vĪ²L

=0

(10.46)

10.8. Subordinacy theory

333

for some eigensolution v at Ī». Theorem 10.51. Let J be a half-line Jacobi matrix in the limit point case. Fix Ī² āˆˆ (0, 1). The Ī²-singular part of its spectral measure Ī¼ is supported on the set SĪ² = {Ī» āˆˆ R | (pnāˆ’1 (Ī»))āˆž n=1 is Ī²-subordinate}, and the Ī²-continuous part of Ī¼ is supported on SĪ²c . Proof. Raising (10.43) to power 1 āˆ’Ī² and using that to divide (10.44) gives āˆš āˆš 2āˆ’Ī² 5 + 24 5 āˆ’ 24 1āˆ’Ī² pĀ·āˆ’1 L ā‰¤2 . ā‰¤

(L)1āˆ’Ī² |m(Ī» + i (L))|

(x)1āˆ’Ī² |m(Ī» + i (L))| qĀ·āˆ’1 Ī² L

Taking L ā†’ āˆž proves that (pnāˆ’1 (Ī»))āˆž n=1 is Ī²-subordinate if and only if lim sup 1āˆ’Ī² |m(Ī» + i )| = āˆž. ā†“0



Now the claim follows from Theorem 6.29 and Theorem 7.51.

Subordinacy can also be used to study spectra of full-line Jacobi matrices: with obvious modiļ¬cations, we say a nontrivial eigensolution u at Ī» is subordinate at āˆ’āˆž if for some eigensolution v at Ī», āˆ’1 |uk |2 = 0. lim k=n āˆ’1 2 nā†’āˆ’āˆž k=n |vk | Denote SĪ±Ā±

=

& Ī» āˆˆ R | the solution with

u1 u0

!

' = cot Ī± is subordinate at Ā± āˆž ,

N Ā± = {Ī» āˆˆ R | there is no subordinate eigensolution at Ā± āˆž}. Note that the set S=



(SĪ±āˆ’ āˆ© SĪ±+ )

Ī±āˆˆ[0,Ļ€)

is precisely the set of Ī» āˆˆ R for which there exists a nontrivial eigensolution which is subordinate at both endpoints Ā±āˆž. Theorem 10.52. Let J be a full-line Jacobi matrix which is the limit point at Ā±āˆž. Let Ī¼ be its canonical spectral measure. (a) The singular part of Ī¼ is supported on S and has multiplicity 1. (b) Nāˆ’ āˆŖ N+ is an essential support for Ī¼ac , i.e., Ī¼ac is mutually absolutely continuous with Ļ‡Nāˆ’ āˆŖN+ (Ī») dĪ». (c) Nāˆ’ āˆ© N+ is an essential support for the multiplicity 2 part of Ī¼ac . Proof. This follows immediately from Corollary 10.41, Corollary 10.42, and Theorem 10.48. 

334

10. Jacobi matrices

10.9. A Combesā€“Thomas estimate and Schnolā€™s theorem In this section, we specialize to bounded Jacobi matrices and consider two related results about the behavior of eigensolutions on and oļ¬€ the spectrum. The ļ¬rst is that the decay properties of Weyl solutions can be signiļ¬cantly improved, from square-summability to exponential decay; estimates of this type are called Combesā€“Thomas estimates. Proposition 10.53. Let J be a bounded half-line Jacobi matrix, and let Ļˆ be a Weyl solution at z āˆˆ C \ Ļƒess (J). There exist Ī³ > 0 and C < āˆž such that for all n, |Ļˆn | ā‰¤ Ceāˆ’Ī³n . Proof. Let us ļ¬rst assume that z āˆˆ / Ļƒ(J). The key observation is that (Ļˆn ) is an eigensolution at z if and only if (un ) = (eĪ³n Ļˆn ) is an eigensolution of the operator (JĪ³ u)n = anāˆ’1 eĪ³ unāˆ’1 + bn un + an eāˆ’Ī³ un+1 . It is easy to estimate the operator norm JĪ³ āˆ’ J ā‰¤ (eĪ³ āˆ’ eāˆ’Ī³ ) sup an , nāˆˆN

so for small enough Ī³ > 0, JĪ³ āˆ’ J < dist(z, Ļƒ(J)), and therefore (JĪ³ āˆ’ J)(J āˆ’ z)āˆ’1  < dist(z, Ļƒ(J))(J āˆ’ z)āˆ’1  = 1. By Theorem 4.26, follows that JĪ³ āˆ’ z = (JĪ³ āˆ’ J) + (J āˆ’ z) = ((JĪ³ āˆ’ J)(J āˆ’ z)āˆ’1 + I)(J āˆ’ z) is invertible as a product of invertible operators. In particular, taking u = (JĪ³ āˆ’ z)āˆ’1 Ī“1 āˆˆ 2 (N), u is a bounded eigensolution of JĪ³ for n ā‰„ 2, so Ļˆn = eāˆ’Ī³n un is an exponentially decaying eigensolution of J. Thus, Ļˆ must be a Weyl solution. If z āˆˆ Ļƒ(J), then Ļˆ0 = 0, so Ļˆ1 = 0 since Ļˆ is nontrivial. Therefore, / Ļƒ(J1 ). SĻˆ is a Weyl solution for the coeļ¬ƒcient stripped matrix J1 and z āˆˆ By applying the above argument to J1 , we conclude exponential decay of SĻˆ.  The rate of exponential decay can be estimated precisely: Ī³ can be made arbitrarily close to the value of the so-called potential theoretic Greenā€™s function at z. This gives a universal inequality which is one of the foundations of Stahlā€“Totik regularity [98]. The second result is Schnolā€™s theorem. We begin with the half-line case, which characterizes the spectrum in terms of where pn (x) grows at most polynomially:

10.9. A Combesā€“Thomas estimate and Schnolā€™s theorem

335

Theorem 10.54 (Schnol). Let J be a bounded half-line Jacobi matrix. Fix Īŗ > 1/2 and denote SĪŗ = {Ī» āˆˆ C | |pn (Ī»)| = O(nĪŗ ), n ā†’ āˆž}. (a) Ī¼ is supported on SĪŗ . (b) SĪŗ āŠ‚ Ļƒ(J). (c) SĪŗ = Ļƒ(J). Proof. (a) By (10.8),   āˆž āˆž  āˆž   āˆ’2Īŗ 2 āˆ’2Īŗ 2 n |pn (Ī»)| dĪ¼(Ī») = nāˆ’2Īŗ < āˆž. n |pn (Ī»)| dĪ¼(Ī») = n=1

n=1

āˆž

By Tonelliā€™s theorem, this implies that and convergence of the series implies pn

n=1

āˆ’2Īŗ |p (Ī»)|2 < n n=1 n (Ī») = O(nĪŗ ) as n ā†’

āˆž for Ī¼-a.e. Ī», āˆž.

(b) For Ī» āˆˆ / Ļƒ(J), the Weyl solution Ļˆ obeys Ļˆ0 = 0, so it is linearly independent with the eigensolution un = pnāˆ’1 (Ī»). Thus, their Wronskian w = an (un Ļˆn+1 āˆ’ un+1 Ļˆn ) is nonzero and independent of n. By the triangle inequality, |w| = |an (un Ļˆn+1 āˆ’ un+1 Ļˆn )| ā‰¤ an Ceāˆ’Ī³n (|un | + |un+1 |). This implies that |un | + |un+1 | ā‰„ C  eĪ³n for some C  > 0 independent of n, so pn is not polynomially bounded. (c) Since Ī¼ is supported on SĪŗ , it is supported on SĪŗ āŠ‚ Ļƒ(J). Since Ļƒ(J) = supp Ī¼ is the smallest closed set on which Ī¼ is supported, this implies SĪŗ = Ļƒ(J).  The full-line Schnolā€™s theorem characterizes the spectrum of a full-line Jacobi matrix in terms of the set of values of z for which there exists a polynomially bounded eigensolution (polynomially bounded at both Ā±āˆž): Theorem 10.55. Let J be a bounded full-line Jacobi matrix, and let Ī¼ be a maximal spectral measure for J. Fix Īŗ > 1/2 and denote by SĪŗ the set of z for which there exists a nontrivial eigensolution which obeys un = O(|n|Īŗ ) as n ā†’ Ā±āˆž. Then: (a) Ī¼ is supported on SĪŗ . (b) SĪŗ āŠ‚ Ļƒ(J). (c) SĪŗ = Ļƒ(J). Proof. (a) Using the orthogonality relations !āˆ— !  u0n u0m W dĪ¼ = Ī“m , Ī“n , u1m u1n

336

10. Jacobi matrices

we conclude that 

2 āˆ’Īŗ

(1 + n )

nāˆˆZ

u0n u1n

!āˆ—

Thus, for Ī¼-a.e. Ī», 

2 āˆ’Īŗ

(1 + n )

nāˆˆZ

Choosing a nonzero vector Ī» = 

2 āˆ’Īŗ

(1 + n )

nāˆˆZ

u0n u1n !

Ī»0 Ī»1

u0n u1n

!

W

u0n u1n

W

u0n u1n

!āˆ—

dĪ¼ < āˆž. ! < āˆž.

āˆˆ C2 such that W ā‰„ Ī»Ī»āˆ— , we obtain !āˆ—

āˆ—

Ī»Ī»

u0n u1n

! < āˆž,

which implies that the nontrivial solution un = Ī»0 u0n + Ī»1 u1n obeys  (1 + n2 )āˆ’Īŗ |un |2 < āˆž. nāˆˆZ

Thus, this solution obeys un = O(|n|Īŗ ) as n ā†’ Ā±āˆž. (b) Assume that there exists a nontrivial, polynomially bounded eigensolution u at z. For z āˆˆ / Ļƒ(J), there exist Weyl solutions Ļˆ Ā± which are exponentially decaying at Ā±āˆž, respectively. As in the proof of Theorem 10.54, since u is polynomially bounded at Ā±āˆž, it must be linearly dependent with Ļˆ Ā± , so it follows that W (Ļˆ+ , Ļˆāˆ’ ) = 0. This would mean that z is an eigenvalue of J, leading to contradiction. (c) Now it follows by the same argument as in the proof of Theorem 10.54.  Schnolā€™s theorem implies an important criterion for the pure point spectrum of a Jacobi matrix (Exercise 10.17), which is used in proofs of a phenomenon called localization.

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map We now consider full-line Jacobi matrices J with q-periodic Jacobi coeļ¬ƒcients, bn+q = bn āˆ€n āˆˆ Z. an+q = an , The behavior of the eigensolutions will be determined by behavior over one period, encoded in the q-step transfer matrix Tq (z) = A(aq , bq ; z) Ā· Ā· Ā· A(a1 , b1 ; z), called in this context the monodromy matrix.

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

337

Lemma 10.56. Ī· is an eigenvalue of Tq (z) if and only if there exists a nontrivial eigensolution v such that vn+q = Ī·vn for all n āˆˆ Z. Proof. This follows from the very deļ¬nition of Tq (z) as a transfer matrix. If Ī· is an eigenvalue, choose v0 , v1 so that av01v0 is an eigenvector. Then  v1  vq+1  = Ī· aq v q a0 v0 so vn+q = Ī·vn for n = 0, 1. By forward and backward induction using q-periodicity of the Jacobi parameters, vn+q = Ī·vn for all n āˆˆ Z. The converse is similar: if v is a nontrivial   eigensolution such that  vn+q = Ī·vn for n = 0, 1, then Tq (z) av0 1v0 = Ī· av01v0 . We will need a general fact about 2 Ɨ 2 matrices with determinant 1: Lemma 10.57. Let A be a 2Ɨ2 matrix and let det A = 1. Denote t = Tr A. Then the following hold. (a) If t āˆˆ (āˆ’2, 2), the matrix A has two distinct eigenvalues both of which lie on the unit circle āˆ‚D.

āˆš tĀ± t2 āˆ’4 , 2

(b) If t āˆˆ {āˆ’2, 2}, the matrix A has a single eigenvalue t/2 of geometric multiplicity 1 or 2. (c) If t āˆˆ C \ [āˆ’2, 2], the matrix A has eigenvalues which lies in D \ {0} and the other in C \ D.

āˆš tĀ± t2 āˆ’4 , 2

one of

Proof. Solving the characteristic polynomial Ī· 2 āˆ’ Tr AĪ· + det A = 0 gives āˆš 2 the eigenvalues Ī· = tĀ± 2t āˆ’4 . From this, (b) follows immediately, and (a) follows by      t Ā± āˆš t2 āˆ’ 4 2  t Ā± iāˆš 4 āˆ’ t2 2 t2 + 4 āˆ’ t2     = 1, t āˆˆ (āˆ’2, 2).  =  =      2 2 4 Conversely, if A has an eigenvalue Ī· with |Ī·| = 1, then by det A = 1, the other zero of its characteristic polynomial is 1/Ī· = Ī·ĀÆ, so the trace is t = Ī· + Ī·ĀÆ = 2 Re Ī· āˆˆ [āˆ’2, 2]. By contraposition, in the case (c) there are no eigenvalues Ī· with |Ī·| = 1. Thus, by det A = 1, there must be two distinct  eigenvalues, one in D and the other in C \ D. The main fact proved above was that Ī· + 1/Ī· āˆˆ [āˆ’2, 2] if and only if Ī· āˆˆ āˆ‚D, which can be recognized as a standard fact about the Zhukovsky map Ī· ā†’ Ī· + 1/Ī·. Equivalently, with the substitution Ī· = eiw , this provides a fact about the cosine as a complex analytic function: cos w āˆˆ [āˆ’1, 1] if and only if w āˆˆ R. This and related basic facts (e.g., for w āˆˆ C, sin w = 0 if and only if w āˆˆ Ļ€Z) will be used below. In our setting, since det Tq (z) = 1, a central object will be the discriminant deļ¬ned by Ī”(z) = Tr Tq (z).

338

10. Jacobi matrices

The previous two lemmas indicate that the set E = {z āˆˆ C | Ī”(z) āˆˆ [āˆ’2, 2]} is relevant. For z āˆˆ E, there is a nontrivial eigensolution v such that vn+q = / E, Ī·vn for some Ī· āˆˆ āˆ‚D (in particular, a bounded eigensolution). For z āˆˆ Ā± there are nontrivial solutions v obeying vn+q = Ī· Ā±1 vnĀ± with Ī· āˆˆ D. Thus v Ā± decays exponentially at Ā±āˆž and grows exponentially at āˆ“āˆž. In particular, v Ā± are linearly independent, and any linear combination C+ v+ +Cāˆ’ vāˆ’ grows exponentially at +āˆž if C+ = 0 and at āˆ’āˆž if Cāˆ’ = 0. In particular, there is no nontrivial, polynomially bounded eigensolution at z āˆˆ / E, and it already follows from Schnolā€™s theorem that Ļƒ(J) = E. We will soon reprove this as part of a more detailed study which will describe the set E and the spectral properties of J much more precisely. Some basic properties of the discriminant are read oļ¬€ from its deļ¬nition: Lemma 10.58. The discriminant Ī” is a polynomial of degree q with real coeļ¬ƒcients and leading coeļ¬ƒcient (a1 Ā· Ā· Ā· aq )āˆ’1 . Proof. Viewed as a polynomial in z, the 1-step transfer matrix A(a, b; z) is ! ! 1/a 0 āˆ’b/a āˆ’1/a A(a, b; z) = z+ , 0 0 a 0 so the monodromy matrix is a polynomial of degree q with ! 1 1 0 q z + O(z qāˆ’1 ), z ā†’ āˆž. Tq (z) = a1 Ā· Ā· Ā· aq 0 0 Taking the trace, we conclude that Ī”(z) is a polynomial with leading term (a1 Ā· Ā· Ā· aq )āˆ’1 z q . The symmetry A(a, b; z)āˆ— = A(a, b; zĀÆ) implies that Tq (z)āˆ— = Tq (ĀÆ z ). Therefore Ī”(z) = Ī”(ĀÆ z ), and Ī” has real coeļ¬ƒcients.  To proceed further, we want to track the z-dependence of Weyl solutions and half-line m-functions. We will do this by using the closely related Marchenkoā€“Ostrovski map. The Marchenkoā€“Ostrovski map is a natural object which can be deļ¬ned for almost periodic spectral problems [49], for which there is no discriminant. However, the Marchenkoā€“Ostrovski map is not entire: we will start with an unmotivated deļ¬nition on C+ and gradually prove the properties of this map and connections with the discriminant. Denote by m+,k the m-function of the half-line Jacobi matrix with coeļ¬ƒcients (an+k , bn+k )āˆž n=1 , that is, of the k times coeļ¬ƒcient-stripped matrix J+,k = S k J+ (S āˆ— )k . In particular, m+,0 = m+ . Taking the branch of log on C+ with Im log āˆˆ (0, Ļ€), the Marchenkoā€“Ostrovski map Ī˜ is deļ¬ned on C+ by qāˆ’1 i log(ak m+,k (z)). (10.47) Ī˜(z) = āˆ’Ļ€ āˆ’ q k=0

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

339

Since Im log m+,k āˆˆ (0, Ļ€), it follows immediately that āˆ’Ļ€ < Re Ī˜(z) < 0

āˆ€z āˆˆ C+ .

This deļ¬nition (10.47) was chosen to be related to the Weyl solution at +āˆž and, as we will see, to an eigenvalue of Tq : Lemma 10.59. For all z āˆˆ C+ , Im Ī˜(z) > 0 and ! ! āˆ’m+ (z) iqĪ˜(z) āˆ’m+ (z) =e . Tq (z) 1 1

(10.48)

Proof. Consider the Weyl solution at +āˆž at energy z, denoted (Ļˆn )nāˆˆZ ; by q-periodicity of the Jacobi matrix, shifting the Weyl solution by q places gives again a Weyl solution. Since the Weyl solution is unique up to normalization, there exists Ī· = Ī·(z) āˆˆ C such that Ļˆn+q = Ī·Ļˆn for all n āˆˆ Z. Taking the product of Ļˆk+1 = āˆ’ak mk = eāˆ’iĻ€+log(ak mk (z)) Ļˆk from k = 0 to q āˆ’ 1 gives Ļˆq = Ļˆ0 eiqĪ˜ , which implies Ī·(z) = eiqĪ˜(z) . Since qāˆ’1 

|Ļˆkq+n |2 = |Ī·|2k

n=0

qāˆ’1 

|Ļˆn |2 ,

n=0

square-summability of Ļˆ at +āˆž implies |Ī·| < 1, i.e., Im Ī˜(z) > 0.



Lemma 10.60. For any z āˆˆ C \ R, Ī”(z) āˆˆ / [āˆ’2, 2]. Proof. For z āˆˆ C+ , the matrix Tq (z) has an eigenvalue eiqĪ˜(z) āˆˆ D, so z ), the same holds for by Lemma 10.57, Ī”(z) āˆˆ / [āˆ’2, 2]. Since Ī”(z) = Ī”(ĀÆ  z āˆˆ Cāˆ’ . We now relate the discriminant to Ī˜ on C+ and use this to construct analytic continuations of Ī˜ into certain simply connected regions. The proof will use the following basic fact from complex analysis. If g is a nonzero analytic function on a simply connected domain Ī©, then there exists analytic h : Ī© ā†’ C such that g = eh . Therefore, there exists an analytic branch of āˆš g = eh/2 on Ī©. Lemma 10.61. For all z āˆˆ C+ , Ī”(z) = 2 cos(qĪ˜(z)).

(10.49)

For any interval (c, d) āŠ‚ R containing no zeros of Ī”2 āˆ’ 4, Ī˜ has an analytic continuation to C+ āˆŖ (c, d) āˆŖ Cāˆ’ such that (10.49) holds.

340

10. Jacobi matrices

Proof. Since det Tq = 1 and one eigenvalue is eiqĪ˜ , the other eigenvalue is 2 2 eāˆ’iqĪ˜ , so (10.49) .follows. Then 4 āˆ’ Ī”(z) = 4 sin (qĪ˜(z)), so on C+ we can 2 ļ¬x a branch of Ī”(z) āˆ’ 4 by setting . Ī”(z)2 āˆ’ 4 = āˆ’2i sin(qĪ˜(z)), so that

iĪ” (z) . (10.50) Ī˜ (z) = . q Ī”(z)2 āˆ’ 4 Since the right-hand side has an analytic continuation to any simply connected subset Ī© of C \ {z | Ī”(z)2 = 4}, so does Ī˜ ; thus, so does Ī˜, with the analytic continuation deļ¬ned by Ī˜(z) = Ī˜(zāˆ— ) + Ī³ Ī˜ (w) dw, where zāˆ— is an arbitrary reference point and Ī³ is an arbitrary path from zāˆ— to z in Ī©.  Lemma 10.62. For any z āˆˆ C, if Ī”(z) āˆˆ (āˆ’2, 2), then Ī” (z) = 0.

Proof. For any interval (c, d) āŠ‚ R on which Ī” āˆˆ (āˆ’2, 2), consider the analytic extension of Ī˜(z) to C+ āˆŖ (c, d) āˆŖ Cāˆ’ . For z āˆˆ (c, d), Ī”(z) āˆˆ (āˆ’2, 2) implies qĪ˜(z) āˆˆ R\Ļ€Z. In particular, Ī˜ is an extended Herglotz function, so Ī˜ (z) > 0 for z āˆˆ (c, d). This implies Ī” (z) = āˆ’2q sin(qĪ˜(z))Ī˜ (z) = 0.  The behavior of the discriminant on R can now be described very explicitly. Theorem 10.63. (a) All zeros of the polynomial Ī” are simple and lie in R. (b) All zeros of Ī”2 āˆ’ 4 are real and can be listed, with multiplicity, in the form Ī»1 < Ī»2 ā‰¤ Ī»3 < Ā· Ā· Ā· < Ī»2qāˆ’2 ā‰¤ Ī»2qāˆ’1 < Ī»2q

(10.51)

(in particular, Ī»2jāˆ’1 < Ī»2j for j = 1, . . . , q). (c) Each zero of Ī”2 āˆ’ 4 is a zero of Ī” āˆ’ 2 or Ī” + 2:

2 n ā‰” 2q, 2q āˆ’ 3 (mod 4) Ī”(Ī»n ) = āˆ’2 n ā‰” 2q āˆ’ 1, 2q āˆ’ 2 (mod 4).

(10.52)

(d) Ī” has exactly one simple zero Īŗj āˆˆ [Ī»2j , Ī»2j+1 ] for each j āˆˆ {1, . . . , q āˆ’ 1} and no other zeros. (e) For each j, either Ī»2j < Īŗj < Ī»2j+1 or Ī»2j = Īŗj = Ī»2j+1 . Proof. (a) All zeros of Ī” are real and simple by Lemmas 10.60 and 10.62, so Ī”(z) has q distinct real zeros c1 < Ā· Ā· Ā· < cq . Since Ī” is a polynomial with real coeļ¬cients, Ī” has at least one zero Īŗj āˆˆ (cj , cj+1 ) for 1 ā‰¤ j ā‰¤ q āˆ’ 1. Since deg Ī” = q āˆ’ 1, those zeros are simple and Ī” has no other zeros in C.

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

341

(b) For each zero cj of Ī”, deļ¬ne by Ī»2jāˆ’1 the largest real zero of Ī”2 āˆ’ 4 smaller than cj , and by Ī»2j the smallest real zero of Ī”2 āˆ’ 4 larger than cj . This is well deļ¬ned even in the border cases j = 1, q since |Ī”(Ī»)| ā†’ āˆž as Ī» ā†’ Ā±āˆž. Thus, by construction, Ī»2jāˆ’1 < cj < Ī»2j . Moreover, since cj < Īŗj < cj+1 and |Ī”(Īŗj )| ā‰„ 2 by (a), it follows that Ī»2j ā‰¤ Īŗj ā‰¤ Ī»2j+1 for j = 1, . . . , q āˆ’ 1. The only possible case of equality among the Ī»k ā€™s is the case Ī»2j = Īŗj = Ī»2j+1 . In this case, since (Ī”2 āˆ’ 4) = 2Ī”Ī” , this shared value is at least a double zero of Ī”2 āˆ’ 4. Thus, the sequence (10.51) is a list of zeros of Ī”2 āˆ’ 4 with repetitions no higher than their algebraic multiplicity. Since deg(Ī”2 āˆ’4) = 2q, we conclude that (10.51) lists all zeros with precisely their algebraic multiplicity. Moreover, if Ī»2jāˆ’1 < Ī»2j , then Ī»2jāˆ’1 < Īŗj < Ī»2j . Since Ī”(Ī») ā†’ +āˆž as Ī» ā†’ +āˆž, (10.52) follows by backward induction in n.  It is now clear that the set E is given by E=

q 

[Ī»2jāˆ’1 , Ī»2j ].

j=1

The closed intervals [Ī»2jāˆ’1 , Ī»2j ] are called spectral bands. The open intervals (Ī»2j , Ī»2j+1 ) are spectral gaps. The jth gap is said to be open if Ī»2j < Ī»2j+1 and closed if Ī»2j = Ī»2j+1 . Since closed gaps are possible, this merely tells us that E is a disjoint union of at most q closed intervals; see Figure 10.1 for an example. If we merely know the set E, the location of closed gaps (if any) seems lost; however, we will see below that if E is the spectrum of a periodic Jacobi matrix, the number and placement of closed gaps are uniquely determined by the set E.

2 Ī»2 = Īŗ1 = Ī»3 Ī»1

Ī»6

Īŗ3

Ī»7

Ī»8

Ī»4 Īŗ2 Ī»5

āˆ’2

Figure 10.1. The discriminant on R for a 4-periodic Jacobi matrix with closed ļ¬rst gap.

Ī»

342

10. Jacobi matrices

Factorizing the polynomials Ī” and Ī”2 āˆ’ 4 gives the product formulas Ī”(z)2 āˆ’ 4 =

q  1 (z āˆ’ Ī»2jāˆ’1 )(z āˆ’ Ī»2j ), (a1 Ā· Ā· Ā· aq )2 j=1

Ī” (z) =

q a1 Ā· Ā· Ā· aq

qāˆ’1 

(z āˆ’ Īŗj ),

j=1

which imply a product formula for Ī˜ by (10.50). Lemma 10.64. For z āˆˆ C+ , " # qāˆ’1  # (z āˆ’ Īŗj )2 1  $ . Ī˜ (z) = (z āˆ’ Ī»1 )(Ī»2q āˆ’ z) (z āˆ’ Ī»2j )(z āˆ’ Ī»2j+1 )

(10.53)

j=1

This function has an analytic extension to C\E which obeys Ī˜ (ĀÆ z ) = āˆ’Ī˜ (z),  and the branch of square root is such that arg Ī˜ (Ī») = Ļ€/2 for Ī» āˆˆ (Ī»2q , āˆž).   2 Proof. The product āˆš formula for Ī˜ follows from those for Ī” and Ī” āˆ’ 4. The square root Ī”2 āˆ’ 4 extends from C+ continuously with real values on R \ E, so by the reļ¬‚ection principle, it has an analytic extension to C \ E which obeys a reļ¬‚ection symmetry. By (10.50) the analytic extension of Ī˜ to C \ E follows.

The functions mk (z) have meromorphic Herglotz extensions with asymptotic behavior mk (z) āˆ¼ āˆ’1/z as z ā†’ āˆž, so by extending both sides of (10.47) analytically through an interval of the form (C, āˆž) with large enough C, it follows that the extension of Ī˜ is purely imaginary and āˆ’iĪ˜ is strictly increasing in (C, āˆž). Thus, the analytic extension of Ī˜ obeys  arg Ī˜ = Ļ€/2 on (C, āˆž). By āˆš (10.50), this also means that our choice of branch of the square root Ī”2 āˆ’ 4 extends to C \ E with positive values on (Ī»2q , āˆž). We will consistently use that branch in what follows. By counting argument changes, for Ī» āˆˆ (Ī»2jāˆ’1 , Ī»2j ), . Im lim Ī”(Ī» + i )2 āˆ’ 4 > 0 if j ā‰” q (mod 2), ā†“0 (10.54) . Im lim Ī”(Ī» + i )2 āˆ’ 4 < 0 if j ā‰” q āˆ’ 1 (mod 2). ā†“0

This is illustrated in Figure 10.2 We also note that, by the product formula for Ī”2 āˆ’ 4 and the choice of branch, . 1 Ī”(z)2 āˆ’ 4 = z q + O(z qāˆ’1 ), z ā†’ āˆž. a1 Ā· Ā· Ā· aq

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

āˆ’i +i

+i āˆ’i

āˆ’i +i

Figure 10.2. The boundary values on E of ei arg

343

+i āˆ’i āˆš

Ī”2 āˆ’4

āˆš Ī”2 āˆ’4 = āˆš 2 . |

Ī” āˆ’4|

Note the sign change occurs even at a closed gap.

It follows from (10.53) or from Exercise 7.20 that iĪ˜ is a Herglotz function and that iĪ˜ (z) = āˆ’1/z + O(1/z 2 ) as z ā†’ āˆž. By Proposition 7.32, its Herglotz representation is of the form  1 iĪ˜ (z) = dĪ½(Ī»), (10.55) Ī»āˆ’z where Ī½ is a probability measure. By taking boundary values of (10.53), this measure can be explicitly obtained in the form 1/2    qāˆ’1 2    (Ī» āˆ’ Īŗj ) 1   dĪ». dĪ½(Ī») = Ļ‡E (Ī»)    (Ī» āˆ’ Ī»1 )(Ī» āˆ’ Ī»2q ) j=1 (Ī» āˆ’ Ī»2j )(Ī» āˆ’ Ī»2j+1 )  The measure Ī½ is called the density of states. Lemma 10.65. On C+ ,  i Ī˜(z) = āˆ’ log(a1 Ā· Ā· Ā· aq ) + i log(z āˆ’ x) dĪ½(x). q

(10.56)

Proof. Both sides of the proposed equality are analyticfunctions in C+ and, by (10.55), have equal derivatives. Thus, Ī˜(z) = c+i log(z āˆ’x) dĪ½(x) for some complex constant c. To ļ¬nd the constant c, we will compare the asymptotics as z ā†’ āˆž, using the branch of log with āˆ’Ļ€ < Im log < Ļ€. Since log(z āˆ’ x) āˆ’ log z = log(1 āˆ’ x/z) ā†’ 0 as z ā†’ āˆž uniformly in x āˆˆ E, it follows that   i log(z āˆ’ x) dĪ½(x) = i log z dĪ½(x) + o(1) = i log z + o(1) as z ā†’ āˆž, z āˆˆ C+ . Since m+,k (z) = (āˆ’1/z)(1 + o(1)) for each k, (10.47) implies that Ī˜(z) = āˆ’Ļ€ āˆ’

qāˆ’1 qāˆ’1 i i log ak āˆ’ i log(āˆ’1/z) + o(1) = āˆ’ log ak + i log z + o(1). q q k=0

k=0

Comparing these asymptotics allows us to read oļ¬€ c, and concludes the proof.  Proposition 10.66. Ī˜ extends continuously to C+ . This extension obeys the following. (a) Im Ī˜ = 0 on E.

344

10. Jacobi matrices

āˆ’Ļ€

0

Figure 10.3. Image of Ī˜(R) for a 4-periodic Jacobi matrix with closed ļ¬rst gap.

(b) Re Ī˜ = āˆ’Ļ€(q āˆ’ j)/q on [Ī»2j , Ī»2j+1 ] for j = 1, . . . , q āˆ’ 1. (c) Re Ī˜ = 0 on [Ī»2q , āˆž). (d) Re Ī˜ = āˆ’Ļ€ on (āˆ’āˆž, Ī»1 ]. This proposition describes the image of Ī˜ on R as a generalized poklygonal curve, with open gaps mapped to vertical line segments traversed up and then down; see Figure 10.3. Proof. It is already known that Ī˜ has an analytic extension through any interval (c, d) āŠ‚ R which contains no zeros of Ī”2 āˆ’ 4. Consider Ī»k , a zero of Ī”2 āˆ’ 4. By the product formula (10.53), Ī˜ (z) = O(|z āˆ’ Ī»k |āˆ’1/2 ),

z ā†’ Ī»k , z āˆˆ C+ .

By the mean value theorem, this implies that |Ī˜(z) āˆ’ Ī˜(w)| = O( 1/2 ),

z, w āˆˆ D (Ī»k ) āˆ© C+ ,

ā†’ 0,

(10.57)

and by continuity, this holds also for z, w āˆˆ D (Ī»k ) āˆ© C+ \ {Ī»k }. This implies that Ī˜ has a limit at Ī»k ; namely, for any sequence zn ā†’ Ī»k with zn āˆˆ C+ \{Ī»k }, (10.57) implies that Ī˜(zn ) is a Cauchy sequence, and (10.57) also implies that its limit is independent of the choice of zn ā†’ Ī»k . By continuity, the extension obeys Ī”(z) = 2 cos(qĪ˜(z)). Then Ī”(z) = [āˆ’2, 2] on E so Ī˜(z) āˆˆ R for z āˆˆ E. Combining this with the observation that Ī˜ is purely imaginary in gaps gives Ī˜(Ī»2j+1 ) = Ī˜(Ī»2j ). Meanwhile, at band edges (and only at band edges), qĪ˜(z) āˆˆ Ļ€Z. Since Ī˜ > 0 on band interiors, this implies that Ī˜(Ī»2j ) āˆ’ Ī˜(Ī»2jāˆ’1 ) = Ļ€q Z for each j. It follows that Ī˜(Ī»2q ) āˆ’ Ī˜(Ī»1 ) = Ļ€, and since āˆ’Ļ€ ā‰¤ Re Ī˜ ā‰¤ 0 on C+ , it follows that Ī˜(Ī»2q ) = 0. Thus, Ī˜(Ī»2j ) = Ī˜(Ī»2j+1 ) = āˆ’Ļ€(q āˆ’ j)/q for all j, which completes the proof. 

10.10. The periodic discriminant and the Marchenkoā€“Ostrovski map

āˆ’Ļ€

Ļ€

āˆ’2Ļ€

345

0

Figure 10.4. Images of analytic extensions of Ī˜ through (Ī»2q , āˆž) and through (āˆ’āˆž, Ī»1 ).

Corollary 10.67. The analytic extension of Ī˜(z) through C+ āˆŖ(Ī»2j , Ī»2j+1 )āˆŖ Cāˆ’ obeys qāˆ’j . Ī˜(ĀÆ z ) = āˆ’Ī˜(z) āˆ’ 2Ļ€ q See Figure 10.4. Proof. This follows from the reļ¬‚ection principle since Re Ī˜(z) = āˆ’ qāˆ’j q Ļ€ for  z āˆˆ (Ī»2j , Ī»2j+1 ). Corollary 10.67 shows that analytic extensions of Ī˜ through diļ¬€erent gaps diļ¬€er, but only by additive real constants. Thus, although Ī˜ does not have an analytic extension to C \ E, its imaginary part has a harmonic extension to C \ E: Corollary 10.68. The function deļ¬ned on C+ by L(z) = Im Ī˜(z) has an extension to C which is a positive harmonic function on C \ E, continuous on C, and zero on E. Moreover,  1 āˆ€z āˆˆ C, (10.58) L(z) = āˆ’ log(a1 Ā· Ā· Ā· aq ) + log|z āˆ’ x| dĪ½(x) q and L(z) = log|z| āˆ’

1 log(a1 Ā· Ā· Ā· aq ) + o(1), q

z ā†’ āˆž.

Proof. Harmonicity on C \ E and L(ĀÆ z ) = L(z) follows from Corollary 10.67. Continuity on C follows from Proposition 10.66 and symmetry, and the integral representation for L follows from Lemma 10.65 on z āˆˆ C \ E. It remains to prove that the integral representation holds also for z = Ī» āˆˆ E. By the monotone convergence theorem,       Ī» + i āˆ’ x  Ī»+iāˆ’x     dĪ½(x)  dĪ½(x) ā†’ log  L(Ī»+i)āˆ’L(Ī»+i/n) = log  Ī» + i/n āˆ’ x  Ī»āˆ’x 

346

10. Jacobi matrices

as n ā†’ āˆž. Subtracting this from L(Ī» + i) and using L(Ī» + i/n) ā†’ L(Ī») shows that the integral representation holds at Ī». The asymptotic behavior of L follows from that of Ī˜.



The function L is called the Lyapunov exponent and (10.58) is the Thouless formula. Our deļ¬nition of L in terms of Ī˜ implies that for all z āˆˆ C, eqL(z) is the norm of the largest eigenvalue of Tq (z), and that L(z) describes the exponential growth/decay rates of the eigensolutions and the exponential growth rate of transfer matrices (Exercise 10.24). This last property is usually taken as the deļ¬nition in more general settings. The distribution function N (Ī») = Ī½((āˆ’āˆž, Ī»]) is called the integrated density of states. It can be proved (Exercise 10.25) that it is up to an aļ¬ƒne substitution equal to the function Re Ī˜ on R and, as a consequence, that Ī½ gives equal weight to each band of the spectrum: 1 j = 1, . . . , q. (10.59) Ī½([Ī»2jāˆ’1 , Ī»2j ]) = , q Corollary 10.68 implies that L is a subharmonic function on C. Moreover, the properties of the Lyapunov exponent have a remarkable interpretation in the language of potential theory [4, 72]. Without introducing the terminology, we will point out that interpretation here. Corollary 10.69. The Lyapunov exponent L is equal to the potential theoĖ† \ E with the pole at āˆž. The measure retic Greenā€™s function for the domain C Ī½ is the equilibrium measure for Ī½. The logarithmic capacity of the set E is Cap E = (a1 Ā· Ā· Ā· aq )1/q . In particular, this means that the measure Ī½ is uniquely determined by the set E. Writing E as a disjoint union of closed intervals, each of those intervals may contain more than one spectral band, but (10.59) implies that the weight of each interval must be a multiple of 1/q. This presents constraints for which ļ¬nite unions of intervals can be q-periodic spectra. Moreover, (10.59) then uniquely determines the locations of closed gaps. Finally, we point out a remarkable interpretation of the Marchenkoā€“ Ostrovski map as a conformal map. An analytic map is said to be conformal if it is injective. We recall two facts from complex analysis: Lemma 10.70. If f : C+ ā†’ C is analytic and Re f  > 0 on C+ , then f is injective. Proof. For any z1 , z2 āˆˆ C+ , z1 = z2 , by the mean value theorem, f (z2 ) āˆ’ f (z1 ) = Re f  (z1 + t(z2 āˆ’ z1 )) > 0 z2 āˆ’ z1 for some t āˆˆ (0, 1), so f (z1 ) = f (z2 ). Re



10.11. Direct spectral theory of periodic Jacobi matrices

347

For any injective map f on C+ , let us denote Ī  = f (C+ ). Clearly, Ī  is connected, and the open mapping theorem implies Ī  is an open set. Lemma 10.71. Assume that a conformal map f : C+ ā†’ C extends to a Ė† Denote Ī  = f (C+ ). Then continuous map on the closure of C+ in C. f (āˆ‚C+ ) = āˆ‚Ī . Proof. It is a general fact about continuous maps and closures that f (C+ ) āŠ‚ f (C+ ). On the other hand, f (C+ ) is compact as a continuous image of a compact set; in particular, it is closed. Since it contains f (C+ ), we conclude f (C+ ) = Ī . Let z0 āˆˆ C+ and w0 = f (z0 ). Let Ī“ > 0 such that DĪ“ (z0 ) āŠ‚ C+ . The set U = f (DĪ“ (z0 )) is open, so it contains some D (w0 ). Since f is injective on C+ , it follows that for all z āˆˆ C+ \ DĪ“ (z0 ), |f (z) āˆ’ w0 | ā‰„ . By continuity, the same holds for z in the boundary of C+ . Thus, f (āˆ‚C+ ) āˆ© f (C+ ) = āˆ…. Thus, f (āˆ‚C+ ) = f (C+ ) \ f (C+ ) = Ī  \ Ī  = āˆ‚Ī .  Applying these to the Marchenkoā€“Ostrovski map Ī˜, since we know the description of Ī˜(R), we conclude: Corollary 10.72. The Marchenkoā€“Ostrovski map maps C+ bijectively to ' qāˆ’1  & qāˆ’j Ļ€ + it | 0 < t ā‰¤ hj , āˆ’ Ī  = {z āˆˆ C | āˆ’Ļ€ < Re z < 0, Im z > 0} \ q j=1

where hj = L(Īŗj ) = max{L(z) | z āˆˆ (Ī»2j , Ī»2j+1 )}. The region Ī  is called a comb domain; see [30].

10.11. Direct spectral theory of periodic Jacobi matrices We now turn to investigating the spectral properties of the Jacobi matrix J and its half-line restrictions JĀ± . This requires the study of Dirichlet eigenvalues. Let us denote the entries of the monodromy matrix by ! t11 t12 Tq = . (10.60) t21 t22 By the representation (10.20), t21 = aq pqāˆ’1 is a polynomial of degree q āˆ’ 1 with positive leading coeļ¬ƒcient. Deļ¬nition 10.73. We say z āˆˆ C is a Dirichlet eigenvalue for the periodic Jacobi matrix J if t21 (z) = 0. There are several equivalent characterizations of Dirichlet eigenvalues:

348

10. Jacobi matrices

Lemma 10.74. For any z āˆˆ C, the following are equivalent: (a) z is a Dirichlet eigenvalue.  (b) 10 is an eigenvector of Tq (z). (c) There is a nontrivial solution of the Jacobi recursion (10.11) such that v0 = vq = 0. (d) z is an eigenvalue of the ļ¬nite (q āˆ’ 1) Ɨ (q āˆ’ 1) Jacobi matrix āŽ› āŽž b1 a1 āŽœa1 b2 a2 āŽŸ āŽœ āŽŸ āŽœ āŽŸ . . . . āŽŸ. . . a2 JD = āŽœ āŽœ āŽŸ āŽœ āŽŸ . . . . āŽ . . aqāˆ’2 āŽ  aqāˆ’2 bqāˆ’1 Proof. (a) ā‡ā‡’ (b) follows from the form of the transfer matrix (10.60). (b) ā‡ā‡’ (c) follows from Tq

0.

v1 v0

! =

! vq+1 . vq

(10.61)

(c) ā‡ā‡’ (d) follows from (10.11) for n = 1, . . . , q āˆ’ 1 since a0 , aqāˆ’1 = 

Corollary 10.75. If z is a Dirichlet eigenvalue, then z āˆˆ R and Ī”(z) āˆˆ / (āˆ’2, 2). Proof. z is real because it is an eigenvalue of the Hermitian matrix JD . Since Tq (z) is upper triangular, t11 (z)t22 (z) = det Tq (z) = 1 so |Ī”(z)| = |t11 (z) + 1/t11 (z)| ā‰„ 2 by the arithmetic meanā€“geometric mean inequality.  Theorem 10.76. The m-function for J+ is given on C+ by āˆš t22 āˆ’ t11 + Ī”2 āˆ’ 4 . (10.62) m+ = 2t21 Moreover, the polynomial t21 has q āˆ’ 1 distinct simple real zeros x1 < Ā· Ā· Ā· < xqāˆ’1 and xj āˆˆ [Ī»2j , Ī»2j+1 ] for all j. Proof. Rewriting (10.48) projectively, m+ obeys āˆ’t11 m+ + t12 . āˆ’m+ = āˆ’t21 m+ + t22 This can be rewritten as a quadratic equation for m+ , whose solutions, using det Tq = 1, are found to be āˆš t22 āˆ’ t11 Ā± Ī”2 āˆ’ 4 . 2t21

10.11. Direct spectral theory of periodic Jacobi matrices

349

āˆš Since Ī”2 āˆ’ 4 is nonzero on C+ , analyticity of m+ dictates that the sign Ā± be chosen uniformly throughout C+ . We will now show that this choice of sign, and the location of zeros of t21 , are dictated by the condition that m+ is a Herglotz function. On every band (Ī»2jāˆ’1 , Ī»2j ), the boundary values of Im m+ are determined by the square root, āˆš Ā± Ī”2 āˆ’ 4 lim Im m+ (Ī» + i ) = lim Im (Ī» + i ). ā†“0 ā†“0 2t21 By established properties of Ī” and t21 , these boundary values are nonzero and have constant sign on each band interior (Ī»2jāˆ’1 , Ī»2j ). Since m+ is Herglotz, this sign must be positive on each band interior; by (10.54), this means that the sign of t21 must change between (Ī»2jāˆ’1 , Ī»2j ) and (Ī»2j+1 , Ī»2j+2 ). This means precisely that t21 has a zero in each interval [Ī»2j , Ī»2j+1 ], j = 1, . . . , q āˆ’ 1. Since deg t21 = q āˆ’ 1, it follows that all zeros xj āˆˆ [Ī»2j , Ī»2j+1 ] are simple, that there is exactly one per gap, and that t21 has no other zeros in C. Since t21 has positive leading coeļ¬ƒcient, it is positive on the rightmost band interior (Ī»2qāˆ’1 , Ī»2q ), so another consideration of the sign of Im m+ there shows that m+ is given by (10.62).  It should be noted that this proof and result hold even in the case of closed gaps. In the closed gap case, of course, Ī»2j = xj = Ī»2j+1 , but even in an open gap, it can still happen that xj = Ī»2j or xj = Ī»2j+1 . Although (10.62) has been proved on C+ , the right-hand side is already in the form of a meromorphic Herglotz function on C\E. It is therefore immediate from our study of meromorphic Herglotz functions that Ļƒess (J+ ) āŠ‚ E. A closer look at (10.62) will reveal that J+ has purely absolutely continuous spectrum on E and eigenvalues precisely at those Dirichlet eigenvalues which are not at gap edges: Theorem 10.77. The operator J+ has essential spectrum Ļƒess (J+ ) = E and discrete spectrum Ļƒd (J+ ) = {xj | 1 ā‰¤ j ā‰¤ q āˆ’ 1, |t11 (xj )| < 1}. More precisely, the spectral measure Ī¼+ is given by dĪ¼+ (Ī») = w+ (Ī») dĪ» + qāˆ’1 j=1 Īŗj Ī“xj where āŽ§āˆš āŽØ 4āˆ’Ī”(Ī»)2 Ī» āˆˆ (Ī»2jāˆ’1 , Ī»2j ) for some j |t21 (Ī»)| (10.63) w+ (Ī») = āŽ©0 else, and Īŗj > 0 if and only if |t11 (xj )| < 1.

350

10. Jacobi matrices

Proof. To describe the spectral measure Ī¼+ , we use (10.62) and consider boundary values of m+ , considering separately the interiors of bands and the remaining isolated points which may contain eigenvalues of J+ . On band interiors (Ī»2jāˆ’1 , Ī»2j ), Im m+ extends continuously from C+ with the boundary values . Ī”(Ī»)2 āˆ’ 4 Im m+ (Ī» + i0) = Im . t21 (Ī») Rewriting using a real positive square root and using Proposition 7.43, the spectral measure Ī¼+ on band interiors is given by . 4 āˆ’ Ī”(Ī»)2 dĪ». Ļ‡(Ī»2jāˆ’1 ,Ī»2j ) (Ī»)dĪ¼+ (Ī») = Ļ‡(Ī»2jāˆ’1 ,Ī»2j ) (Ī») |t21 (Ī»)| The only remaining contributions to the spectral measure can be pure points at a gap edge or at poles of the meromorphic Herglotz function m+ . While these can be investigated using (10.62), we proceed diļ¬€erently. As in the proof of Lemma 10.59, if z āˆˆ R is an eigenvalue of J+ , then the Weyl solution v must obey v0 = vq = 0, so z must be a Dirichlet eigenvalue. Moreover, if we normalize the Weyl solution by v1 = 1, then vq+1 = t11 (z) by (10.61). It follows, as in the proof of Lemma 10.59, that v is square-integrable if and  only if |t11 (z)| < 1. The explicit formula (10.63) shows that w+ is continuous and strictly positive on each band interior (Ī»2jāˆ’1 , Ī»2j ). In fact, w+ extends continuously and strictly positively through closed gaps; moreover, its asymptotics at all gap edges can be precisely described (Exercise 10.18). The point masses Īŗj can also be calculated (Exercise 10.19). Our discussion gave preferential treatment to the positive half-line and the oļ¬€-diagonal entry t21 , but the proofs of Lemma 10.59 and Theorem 10.76 can be repeated for the Weyl solution at āˆ’āˆž to describe the other eigenvalue and eigenvector of Tq (z), which correspond to māˆ’ ; we leave the details as an exercise. Lemma 10.78. If māˆ’ denotes the m-function corresponding to Jāˆ’ , then for any z āˆˆ C+ , ! ! 1 1 āˆ’iqĪ˜(z) =e . (10.64) Tq (z) āˆ’a20 māˆ’ (z) āˆ’a20 māˆ’ (z) Lemma 10.79. The entry t12 is a polynomial of degree q āˆ’ 1 with negative leading coeļ¬ƒcients. It has simple zeros yj āˆˆ [Ī»2j , Ī»2j+1 ] for j = 1, . . . , q āˆ’ 1 and no other zeros. In particular, by Lemma 10.78, we have the other solution of the quadratic equation considered in the proof of Theorem 10.76:

10.11. Direct spectral theory of periodic Jacobi matrices

351

Corollary 10.80. For z āˆˆ C+ ,

āˆš 1 t22 āˆ’ t11 āˆ’ Ī”2 āˆ’ 4 = . 2t21 a20 māˆ’

The explicit formulas for mĀ± give an important relation between their boundary values: Corollary 10.81. For all Ī» in the interior of E, 1 a20 māˆ’ (Ī» + i0)

= m+ (Ī» + i0).

(10.65)

This means that periodic Jacobi matrices are reļ¬‚ectionless; in general, a full-line Jacobi matrix is said to be reļ¬‚ectionless if (10.65) holds Lebesguea.e. on its spectrum. Theorem 10.82. The Jacobi matrix J has purely absolutely continuous spectrum on E with multiplicity 2, i.e., J āˆ¼ = TĪ»,Ļ‡E (Ī») dĪ» āŠ• TĪ»,Ļ‡E (Ī») dĪ» . Proof. From the formulas for mĀ± , we calculate G0,0 = āˆ’

a20

āˆš

t21 , Ī”2 āˆ’ 4

G1,1 = āˆš

t12 . Ī”2 āˆ’ 4

These are analytic Herglotz functions on C \ E, which again proves that Ļƒ(J) āŠ‚ E. Moreover, they have continuous extensions to band interiors (Ī»2jāˆ’1 , Ī»2j ) and at most square root singularities at gap edges, so the corresponding measures are purely absolutely continuous on (Ī»2jāˆ’1 , Ī»2j ) and there are no point masses at band edges. It follows that the maximal spectral measure for J is absolutely continuous with respect to Ļ‡E (Ī») dĪ». Thus, by Corollary 10.41 and spectral properties of JĀ± , J āˆ¼ = TĻ‡E (Ī») dĪ» āŠ•TĻ‡E (Ī») dĪ» .  At points where Ī”(z) = Ā±2, Lemma 10.57 does not provide a way to distinguish whether Tq (z) has an eigenvalue of geometric multiplicity 1 or 2. Of course, geometric multiplicity 2 means that Tq (z) = Ā±I, and  1 geometric multiplicity 1 means that Tq (z) is unitarily equivalent to Ā±1 0 Ā±1 . Remarkably, this dichotomy at gap edges is precisely linked to the open gap/closed gap dichotomy. Proposition 10.83. For Ī» āˆˆ C, the following are equivalent. (a) Ī» is a closed gap of J, i.e. Ī» = Ī»2j = Ī»2j+1 for some j āˆˆ {1, . . . , q āˆ’ 1}. (b) Ī» is a double root of Ī”2 āˆ’ 4. (c) Tq (Ī») āˆˆ {+I, āˆ’I}.

352

10. Jacobi matrices

Proof. (a) ā‡ā‡’ (b): This is known by Theorem 10.63. (a) =ā‡’ (c): If Ī» = Ī»2j = Ī»2j+1 , then xj = Ī» so t21 (Ī») = 0. Analogously, yj = Ī» so t12 (Ī») = 0. Thus, Tq (Ī») is a diagonal matrix. Now det Tq (Ī») = 1 and Tr Tq (Ī») = Ā±2 imply that Tq (Ī») = Ā±I. (c) =ā‡’ (b): Since det Tq (z) = 1 for all z, diļ¬€erentiating gives (t11 t22 āˆ’ t21 t12 ) = 0. Using the product rule and applying for z = Ī» gives Ā±(t11 (Ī») + t22 (Ī»)) = 0, which means that Ī” (Ī») = 0. Thus, Ī» is a double root of  Ī”2 āˆ’ 4. Knowing that in band interiors, mĀ± have continuous ļ¬nite nonzero extensions denoted mĀ± (Ī» + i0), we can single out the following formal eigensolutions of J. Deļ¬nition 10.84. For Ī» āˆˆ R with Ī”(Ī») āˆˆ (āˆ’2, 2), Floquet solutions v Ā± are the formal eigensolutions at Ī» which obey v1+ = āˆ’a0 m+ (Ī» + i0), v0+

v1āˆ’ 1 āˆ’ = āˆ’ a m (Ī» + i0) . v0 0 āˆ’

(10.66)

In particular, v0Ā± = 0 and v1Ā± = 0 (see also Exercise 10.21). Floquet solutions have the following skew-periodic property: Corollary 10.85. The sequence eāˆ“inĪ˜(Ī») vnĀ± is q-periodic. Here Ī˜(Ī») denotes the value of Ī˜ obtained from C+ by analytic continuation. Proof. By continuity, it follows from (10.48) and (10.64) that ! Ā± ! vq+1 v1Ā± Ā±iqĪ˜(Ī») =e , a0 v0Ā± a0 vqĀ± Ā± so eāˆ“iqĪ˜(Ī») vq+n = vnĀ± for n = 0, 1. By forward and backward induction in n, q-periodicity follows. 

Floquet solutions are related to a direct integral representation which provides another approach for spectral theoretic properties of the full-line Jacobi matrix J; see Exercises 10.22 and 10.23.

10.12. Exercises 10.1. If J is a bounded half-line Jacobi matrix, prove that supnāˆˆN an ā‰¤ J and supnāˆˆN |bn | ā‰¤ J. 10.2. This problem describes criteria for strong and weak operator convergence of a sequence of Jacobi matrices in terms of their coeļ¬ƒcients.

10.12. Exercises

353

(a) Consider half-line Jacobi matrices Jk , indexed by k āˆˆ N āˆŖ {āˆž}, s such that Jk has coeļ¬ƒcients (ak,n , bk,n )āˆž n=1 . Prove that Jk ā†’ Jāˆž as k ā†’ āˆž if and only if sup sup(ak,n + |bk,n |) < āˆž, kāˆˆN nāˆˆN

and, for each n āˆˆ N, ak,n ā†’ aāˆž,n and bk,n ā†’ bāˆž,n as k ā†’ āˆž. s w (b) Prove that Jk ā†’ Jāˆž if and only if Jk ā†’ Jāˆž . 10.3. A ļ¬nite Favardā€™s theorem: Prove that as a map on the set of d Ɨ d Jacobi matrices, J ā†’ Ī¼J,Ī“1 is a bijection with the set of probability measures on R whose support consists of exactly d points. Let J be the bounded half-line Jacobi matrix (10.2), and let (pn )āˆž n=0 be the corresponding orthonormal polynomials. (a) Prove that pd (z) = 0 if and only if z is an eigenvalue of the d Ɨ d Jacobi matrix (10.1) (note the same Jacobi parameters from J). (b) Prove that pd has d distinct real zeros. 10.4. Find the m-function and the spectral measure corresponding to the Jacobi parameters bn = 0 and

1 nā‰„2 an = āˆš 2 n = 1. Hint: You may have to apply Proposition 7.43 away from some singularities and apply Lemma 7.37 to check for the presence of point masses at the singularities. 10.5. Fix c > 0 and consider the half-line Jacobi matrix with parameters bn = 0 and

1 nā‰„2 an = c n = 1. Prove that Ļƒess (J) = [āˆ’2, 2]. For which values of c > 0 does J have nonempty discrete spectrum? 10.6. Let Ī± > 0 and Ī² āˆˆ R. Find the spectral measure of the half-line Jacobi matrix with an = Ī± and bn = Ī² for all n. 10.7. If J is a d Ɨ d Jacobi matrix, we say that a Weyl solution at z āˆˆ C is an eigensolution at C such that Ļˆd+1 = 0. Deļ¬ne the m-function by (10.3) and prove (10.15). 10.8. Let J be a half-line Jacobi matrix, and let Jn = S n J(S āˆ— )n be the n times coeļ¬ƒcient-stripped Jacobi matrix. If I āŠ‚ R is an interval such that Ļƒ(Jn ) āˆ© I contains at most k points, prove that Ļƒ(J) āˆ© I contains at most n + k points. 10.9. Let J be a half-line Jacobi matrix, and let Ļˆ(z) be a Weyl solution at z.

354

10. Jacobi matrices

(a) For z, w āˆˆ C, prove that J āˆ’ T?n (w)āˆ— J T?n (z) = āˆ’i(z āˆ’ w)

nāˆ’1  j=0

! āˆ— 1 0 ? ? Tj (w) T (z). 0 0 j

(b) For z, w āˆˆ / Ļƒess (J), prove that Wnāˆ’1 (Ļˆ(w), Ļˆ(z)) āˆ’ Wn (Ļˆ(w), Ļˆ(z)) = (z āˆ’ w)Ļˆn (w)Ļˆn (z).   from the right and Hint: Multiply the result of (a) by m(z) 1 m(w)āˆ— from the left. 1 (c) For z, w āˆˆ / Ļƒ(J), if Weyl solutions are normalized by a0 Ļˆ0 = āˆ’1, prove that m(z) āˆ’ m(w) = (z āˆ’ w)

āˆž 

Ļˆn (w)Ļˆn (z).

n=1

/ R, with Weyl solutions normalized (d) For any sequence zk ā†’ z āˆˆ Ļˆ(zk ) ā†’ Ļˆ(z) in 2 (N). by a0 Ļˆ0 = āˆ’1, prove that 2 Hint: Use (b) to express āˆž n=1 |Ļˆn (zk ) āˆ’ Ļˆn (z)| in terms of the m-function. (e) For any z āˆˆ / R, prove that m (z) =

āˆž 

Ļˆn (z)2 .

n=1

(f) Generalize (c) and (d) to z āˆˆ R \ Ļƒ(J). 10.10. For ļ¬xed n āˆˆ N and z āˆˆ C+ , do the circles āˆ‚Dnāˆ’1 (z) and āˆ‚Dn (z) have a nonempty intersection? 10.11. Improve Proposition 10.30 by proving that for any z āˆˆ C+ , int Dn (z) = {m(z)}. nāˆˆN

10.12. Prove that

 lim

nā†’āˆž

Ļ€(p2n (x) +

1 dx = 1. a2n p2nāˆ’1 (x))

)(āˆ’i/a2n ). 10.13. Prove (10.30) by using m(n) (z) = āˆ’(fTāˆ’1 n (z) 10.14. In the setting of Carmonaā€™s theorem, prove that for h āˆˆ C(R) with supp h āŠ‚ (āˆ’2, 2), āˆš   h(x) 4 āˆ’ x2 1 dx. h(x) dĪ¼(x) = lim nā†’āˆž 2Ļ€ p2n (x) āˆ’ xan pn (x)pnāˆ’1 (x) + a2n p2nāˆ’1 (x) (10.67)

10.12. Exercises

355

Hint: Consider the approximations ! ! m(n, z) m0 (z) ? $ Tn (z) , 1 1 āˆš

where m0 (z) = āˆ’z+ 2 z āˆ’4 . This result is motivated by recalling that m0 (z) is the m-function corresponding to the free Jacobi matrix, so by coeļ¬ƒcient stripping, m(n, z) is the m-function corresponding to the Jacobi matrix J (n) with coeļ¬ƒcients



ak k ā‰¤ n bk k ā‰¤ n (n) (n) ak = bk = 1 k > n, 0 k > n. 2

10.15. Compute the Weyl M -matrix for the full-line Jacobi matrix J with coeļ¬ƒcients an ā‰” 1, bn ā‰” 0 and use its normal limits on the real line to prove that J āˆ¼ = T[āˆ’2,2],dx āŠ• T[āˆ’2,2],dx . This provides a diļ¬€erent proof for Example 10.36. 10.16. Let J be a full-line Jacobi matrix. We proved that for z āˆˆ C \ Ļƒess (J), there exist exponentially decaying Weyl solutions at Ā±āˆž. This problem considers the converse. Assume that for some z āˆˆ C, there exist nontrivial eigensolutions v Ā± which decay exponentially at Ā±āˆž, i.e., there exist C, Īŗ > 0 such that |vnĀ± | ā‰¤ Ceāˆ’Ī³n for n āˆˆ N. Prove that z is an eigenvalue of J if v Ā± are linearly dependent and zāˆˆ / Ļƒ(J) otherwise. 10.17. Let J be a bounded full-line Jacobi matrix with the following property: for any z āˆˆ R and any sequence u such that Ju = zu, if un = O(|n|Īŗ ) for some Īŗ, then u āˆˆ 2 (Z). Prove that J has an orthonormal basis of eigenvectors. 10.18. Consider the spectral density w+ of the half-line periodic Jacobi matrix J+ . (a) If Ī»2j = Ī»2j+1 is a closed gap, prove that limĪ»ā†’Ī»2j w+ (Ī») exists, Ī»āˆˆR

is ļ¬nite and nonzero. Accordingly, w+ has a strictly positive continuous extension to the interior of E. (b) At any open gap edge Ī»k which is also a Dirichlet eigenvalue, prove that limĪ»ā†’Ī»k |Ī» āˆ’ Ī»k |1/2 w+ (Ī») exists, is ļ¬nite and nonzero. Ī»āˆˆE

(c) At any open gap edge Ī»k which is not a Dirichlet eigenvalue, prove that limĪ»ā†’Ī»k |Ī»āˆ’Ī»k |āˆ’1/2 w+ (Ī») exists, is ļ¬nite and nonzero. Ī»āˆˆE

10.19. For the half-line periodic Jacobi matrix J+ , if the Dirichlet eigenvalue xj is an eigenvalue of J+ , prove that Ī¼+ ({xj }) =

t22 (xj ) āˆ’ t11 (xj ) . t21 (xj )

356

10. Jacobi matrices

10.20. Borgā€™s theorem: If J is a periodic full-line Jacobi matrix with all gaps closed, i.e., such that Ļƒ(J) is a single closed interval, prove that it has constant coeļ¬ƒcients, i.e., there exist Ī± > 0 and Ī² āˆˆ R such that an = Ī± and bn = Ī² for all n āˆˆ Z. Hint: Find the half-line spectral measure Ī¼+ and compare with Exercise 10.6. 10.21. For any Ī» with Ī”(Ī») āˆˆ (āˆ’2, 2), prove that the Floquet solutions deļ¬ned by (10.66) obey vnĀ± = 0 for all n āˆˆ Z. 10.22. For t āˆˆ R \ Ļ€Z, consider the q Ɨ q matrix āŽž āŽ› a1 eāˆ’it aq b1 āŽŸ āŽœ a1 b2 a2 āŽŸ āŽœ āŽŸ āŽœ . . . . āŽŸ. . . a2 J(t) = āŽœ āŽŸ āŽœ āŽŸ āŽœ . . . . āŽ . . aqāˆ’1 āŽ  eit aq aqāˆ’1 bq (a) For t āˆˆ R \ Ļ€Z, prove that J(t) has q distinct real eigenvalues Ļ1 (t) < Ā· Ā· Ā· < Ļq (t) which are precisely the solutions Ļ of Ī”(Ļ) = 2 cos t. (b) On the intervals (0, Ļ€) and (Ļ€, 2Ļ€), prove that Ļj (t) are real analytic functions of t with nonzero derivative. (c) On the intervals (0, Ļ€) and (Ļ€, 2Ļ€), prove that there is a family of unitary q Ɨ q matrices U (t) which depends on real analytically on t and such that U (t)āˆ’1 J(t)U (t) are diagonal matrices. Hint: Relate J(t) with Floquet solutions. 10.23. Consider the Hilbert space 2

q

L ([0, 2Ļ€], dt) =

q 3

L2 ([0, 2Ļ€], dt)

j=1

viewed also as a space of square-integrable vector-valued functions [0, 2Ļ€] ā†’ Cq . (a) Prove that the mod q Fourier decomposition Fq : L2 ([0, 2Ļ€], dt)q ā†’ 2 (Z) deļ¬ned by



2Ļ€

(Fq f )kq+r =

fr (t)eāˆ’ikt dt,

k āˆˆ Z, r = 1, . . . , q,

0

is unitary. (b) Prove that, with the matrices J(t) from the previous exercise, (Fqāˆ’1 JFq f )(t) = J(t)f (t). This is described as the direct integral representation, since the right-hand side can be viewed as a pointwise (in t) multiplication

10.12. Exercises

357

by J(t) on Hilbert spaces Cq , which is similar to a direct sum construction, but is parametrized by t āˆˆ [0, 2Ļ€] with Lebesgue measure instead of a countable sum. (c) Using the unitaries U (t) from 4qthe previous exercise, prove that J is unitarily equivalent to j=1 TĻj (t),Ļ‡[0,2Ļ€] (t) dt . (d) Prove that āˆ¼ TĻ (t),Ļ‡ (t) dt = TĻ‡ (x) dx āŠ• TĻ‡ (x) dx j

[0,2Ļ€]

[Ī»2jāˆ’1 ,Ī»2j ]

[Ī»2jāˆ’1 ,Ī»2j ]

for each j and that J āˆ¼ = TĻ‡E (x) dx āŠ• TĻ‡E (x) dx . This provides another proof that J has purely absolutely continuous spectrum of multiplicity 2 with essential support E. 10.24. Let L(z) denote the Lyapunov exponent associated to the periodic Jacobi matrix J. (a) If Tn (z) denote the n-step transfer matrices associated to J+ , prove that L(z) = limnā†’āˆž n1 logTn (z). Hint: Reduce to the case where n is a multiple of q and use the spectral radius of Tq (z). (b) If z āˆˆ C+ and v is a nontrivial eigensolution at z, prove that 1 lim log|vn | = āˆ’L(z) nā†’āˆž n if v is a Weyl solution at +āˆž and 1 lim log|vn | = L(z) nā†’āˆž n if v is linearly independent from the Weyl solution at +āˆž. 10.25. If Ī˜ denotes the Marchenkoā€“Ostrovski map, prove that for all Ī» āˆˆ R, Re Ī˜(Ī») = āˆ’Ļ€Ī½((Ī», āˆž)) = āˆ’Ļ€ + Ļ€Ī½((āˆ’āˆž, Ī»]) and that Ī½ has the property (10.59). Hint: Use (10.56) to compute boundary values of Re Ī˜.

Chapter 11

One-dimensional SchrĀØ odinger operators

SchrĀØ odinger operators are operators given by the formal expression āˆ’Ī” + V acting on functions in L2 (Ī©), where Ī© is a region in Rd (with Lebesgue measure), Ī” denotes the Laplacian, and V stands for pointwise multiplication by a real-valued function V on Ī©; the function V is often called the potential. Their name, and part of the motivation for their study, comes from quantum mechanics, in which they correspond to the Hamiltonian of a particle conļ¬ned to a region Ī© with an external potential V . However, their study in the one-dimensional case dates back to the work of Sturm and Liouville in 1836 on the boundary value problem: āˆ’f  + V f = Ī»f, cos Ī±f (0) + sin Ī±f  (0) = 0, 

cos Ī²f (1) āˆ’ sin Ī²f (1) = 0.

(11.1) (11.2) (11.3)

Classically, it was common to assume that V is smooth or at least continuous, but the theory applies with only minor changes to integrable potentials, V āˆˆ L1 ([0, 1]). We will study the diļ¬€erential equation (11.1) in Sections 11.1 and 11.2. In Section 11.3, we will interpret the boundary value problem (11.1), (11.2), and (11.3) as a self-adjoint SchrĀØodinger operator H on L2 ([0, 1]). In these sections, integrability of V will ensure that solutions f and functions in the domain of H have pointwise values of f and f  , which play an important role. 359

360

11. One-dimensional SchrĀØodinger operators

The main scope of this chapter is more general: we study diļ¬€erential operators of the form d2 H =āˆ’ 2 +V dx on the Hilbert space L2 (I), on an open interval I = (āˆ’ , + ) āŠ‚ R, which can be ļ¬nite or inļ¬nite. We will assume that V is a real-valued potential on I such that V āˆˆ L1loc (I), i.e., V is integrable on every compact subinterval of I. Even for ļ¬nite endpoints, this is more general than before, since V may not be integrable in neighborhoods of the endpoints: it is only required to be integrable on compact intervals [c, d] āŠ‚ I. If an endpoint Ā± obeys Ā± āˆˆ R

and

V āˆˆ L1 ([c, d]) for [c, d] āŠ‚ I āˆŖ {Ā± },

(11.4)

it is said to be a regular endpoint (of course, the case of both regular endpoints is precisely the special case of Section 11.3). In general, endpoint behavior can be more varied; this is described by the so-called Weyl limit pointā€“limit circle alternative discussed in Section 11.4, which informs us whether a boundary condition is needed at an endpoint. After regular endpoints, the most often encountered special case is that of an inļ¬nite endpoint Ā± at which the potential is bounded below. Consistently with our choice to consider L1loc -potentials, we will study the L1loc -generalization of this condition and study inļ¬nite endpoints at which the negative part of the potential, Vāˆ’ (x) = max(0, āˆ’V (x)), is uniformly locally L1 , i.e.,  x+1 and lim sup Vāˆ’ (t) dt < āˆž. (11.5) Ā± = Ā±āˆž xā†’Ā±

x

We will see in Section 11.4 that in this case Ā± is a limit point endpoint. Using the Weyl alternative, in Section 11.5, we will describe self-adjoint SchrĀØ odinger operators with separated boundary conditions, which are the central object of the entire chapter. In Section 11.6 we will study their resolvents, introducing Weyl solutions and the Greenā€™s function. From Section 11.7 to Section 11.11, we specialize to the setting of one regular endpoint, often called the half-line setting. We introduce the mfunction and canonical spectral measure Ī¼ of the operator and construct eigenfunction expansionsā€”these are the canonical unitary maps which diagonalize one-dimensional SchrĀØodinger operators, i.e., conjugate them to multiplication operators. These eigenfunction expansions will connect us to the abstract theory of unbounded self-adjoint operators. We will also introduce Weyl disks, which provide a diļ¬€erent perspective on the limit

11.1. An initial value problem

361

pointā€“limit circle alternative; they are useful for approximations, including approximations of the spectral measure from formal eigensolutions (Carmonaā€™s theorem) and continuous dependence of the m-function on the potential. The theory in these sections should be seen as the interplay between three main objects: the SchrĀØodinger operator H, the m-function, and the spectral measure Ī¼. These three objects determine each other uniquely; we will prove that through a local Borgā€“Marchenko theorem. In Section 11.12 we study an arbitrary SchrĀØodinger operator with separated boundary conditions and construct the full-line eigenfunction expansion. Once again this will connect us to the abstract theory of unbounded self-adjoint operators. Through the notion of the Weyl M -matrix, properties of a SchrĀØodinger operator H on (āˆ’ , + ) will be related to properties of SchrĀØ odinger operators with the same potential on (āˆ’ , c) and (c, + ), both of which have one regular endpoint at c and are taken with a Dirichlet endpoint at c. Thus, the eļ¬€ort in Sections 11.7ā€“11.11 is useful for the general case. In Section 11.13 we study subordinacy theory, which is a very robust way of characterizing spectral properties of a SchrĀØodinger operator in terms of the behavior of its eigensolutions at real values of the spectral parameter. In Section 11.14, we specialize to SchrĀØodinger operators for which each endpoint is either regular or of the form (11.5), and explore their ļ¬ner properties. This starts with semiboundedness of the spectrum and includes important estimates about the pointwise behavior of eigensolutions and their derivatives; for example, this includes the result that the boundedness of eigensolutions implies an absolutely continuous spectrum. This theme is continued in Section 11.15 with a Combesā€“Thomas estimate and Schnolā€™s theorem. In Sections 11.16 and 11.17 we study periodic SchrĀØodinger operators. This classical setting can be studied in diļ¬€erent ways; we will use the Marchenkoā€“Ostrovskii map as the central object. Other texts about SchrĀØ odinger operators include [8, 19, 65, 73ā€“75, 108, 110, 111]. Sturmā€“Liouville and SchrĀØ odinger operators are sometimes considered under weaker regularity assumptions than those assumed here (see, e.g., [8, 28, 44, 83, 110]), using quasi-derivatives.

11.1. An initial value problem Our goal in this section is to study the initial value problem āˆ’f  + (V āˆ’ z)f = g,

f (0) = a,

f  (0) = b,

(11.6)

362

11. One-dimensional SchrĀØodinger operators

where a, b, z āˆˆ C and g āˆˆ L1 ([0, 1]). We will work on the unit interval [0, 1] and that aļ¬€ects some of the estimates in this section; however, all qualitative conclusions extend to an arbitrary interval [c, d] by an aļ¬ƒne substitution. By a solution of (11.6), we mean a function f which belongs to the class AC2 ([0, 1]) = {f āˆˆ AC([0, 1]) | f  āˆˆ AC([0, 1])},

(11.7)

and we interpret the diļ¬€erential equation in (11.6) as equality of L1 functions, i.e., equality Lebesgue-a.e. Properties of the space AC2 ([0, 1]) are summarized in the following lemma, whose proof is left as Exercise 11.1. Lemma 11.1. (a) AC2 ([0, 1]) is a Banach space with the norm 



f AC2 ([0,1]) = |f (0)| + |f (0)| +

1

|f  (x)| dx.

(11.8)

0

(b) For any y āˆˆ [0, 1], the point evaluations f ā†’ f (y) and f ā†’ f  (y) are bounded linear functionals on AC2 ([0, 1]). (c) f C([0,1]) ā‰¤ f AC2 ([0,1]) for all f āˆˆ AC2 ([0, 1]). (d) For any y āˆˆ [0, 1], the norm (11.8) is equivalent to the norm 



1

f  = |f (y)| + |f (y)| +

|f  (x)| dx.

0

(e) For any f āˆˆ AC2 ([0, 1]), there exists a bounded linear functional Ī› on AC2 ([0, 1]) such that Ī› = 1 and Ī›(f ) = f AC2 ([0,1]) (this implies that the space AC2 ([0, 1]) has the property (2.27)). For the study of the initial value problem (11.6), it is useful to view V as a perturbation of āˆ’āˆ‚x2 āˆ’ z, so we start with a brief look at the case V = 0, g = 0. It will be natural to use the quasi-momentum k=

āˆš āˆ’z.

For now, we use this substitution pointwise, with an arbitrary choice of square root. Later, we will focus on z āˆˆ C \ [0, āˆž), and it will be beneļ¬cial to set the analytic branch of k such that Re k > 0. In particular, āˆ’k will be a Herglotz function. We will emphasize analyticity of solutions as Banach-space valued functions (see Section 2.7).

11.1. An initial value problem

363

Proposition 11.2. (a) For any z āˆˆ C, the functions c(x, k) =

s(x, k) =

āˆž  k 2n x2n n=0 āˆž  n=0

(2n)!

= cosh(kx),

k 2n x2n+1 = (2n + 1)!



sinh(kx) k

x

z=  0 z=0

obey āˆ‚x s(x, k) = c(x, k),

āˆ‚x c(x, k) = k 2 s(x, k),

and c(0, k) = 1,

(āˆ‚x c)(0, k) = 0,

s(0, k) = 0,

(āˆ‚x s)(0, k) = 1.

(b) The maps z ā†’ s(Ā·, k), z ā†’ c(Ā·, k) are entire functions from C to AC2 ([0, 1]). (c) The initial value problem āˆ’f  āˆ’ zf = 0,

f (0) = a,

f  (0) = b

(11.9)

has a unique solution f āˆˆ AC2 ([0, 1]), given by f (x) = ac(x, k) + bs(x, k). Proof. (a) and (b): The functions c(x, k), s(x, k) are deļ¬ned as power series with even powers of k, so they are power series in z. Since % % % 1 n% 1 % x % = % n! % 2 (n āˆ’ 1)! AC ([0,1]) for n ā‰„ 1, the power series converge in AC2 ([0, 1]) for all z, so they deļ¬ne entire functions. The other properties are trivial calculations. (c): It follows from (a) that f = ac(Ā·, k) + bs(Ā·, k) is a solution. If the initial value problem had two solutions, their diļ¬€erence F would obey āˆ’F  āˆ’zF = 0, F (0) = F  (0) = 0. It suļ¬ƒces to prove that this implies F = 0, keeping in mind that our notion of solution requires equality āˆ’F  āˆ’ zF = 0 only almost everywhere. The proof is by a Gronwall-type argument. Deļ¬ne g = |F |2 + |F  |2 and note that g ā‰„ 0 and g āˆˆ AC([0, 1]). Compute g  = 2 Re(FĀÆ F  + FĀÆ  F  ) = 2 Re((1 āˆ’ z)FĀÆ F  ) ā‰¤ Cg with C = |1 āˆ’ z|. These calculations and inequalities hold Lebesgue-a.e. Now h(x) = eāˆ’Cx g(x) is also absolutely continuous and h (x) = eāˆ’Cx (g  (x) āˆ’ Cg(x)) ā‰¤ 0.

364

11. One-dimensional SchrĀØodinger operators

By its deļ¬nition, an absolutely continuous function h such that h ā‰¤ 0 a.e. obeys  x h (t) dt ā‰¤ h(0). h(x) = h(0) + 0

Thus, g(x) ā‰¤

eCx g(0)



= 0, so g = 0 and F = 0 identically.

We will now note some important estimates. There will be a duality in our estimates: we want estimates uniform on bounded sets of z, but we also want good estimates for large z. To present both eļ¬ƒciently, we use the notation |||k||| = max(1, |k|). Lemma 11.3. For all z āˆˆ C and x āˆˆ [0, 1], |c(x, k)| ā‰¤ e|Re k|x ,

(11.10)

|s(x, k)| ā‰¤ |||k|||āˆ’1 e|Re k|x .

(11.11)

Proof. By Eulerā€™s formula, |c(x, k)| ā‰¤ e|Re k|x ,

|s(x, k)| ā‰¤

e|Re k|x . |k|

Another estimate for s(Ā·, k) follows from (āˆ‚x s)(x, k) = c(x, k) and   x    c(t, k) dt ā‰¤ xe|Re k|x ā‰¤ e|Re k|x . |s(x, k)| = 



0

We now return to the initial value problem (11.6) and apply the standard idea to rewrite it as an equivalent integral equation. There is more than one way to do this; in this section, we ļ¬rst add g, and then V . Lemma 11.4. For any g āˆˆ L1 ([0, 1]), the function T g deļ¬ned by  x s(x āˆ’ t, k)g(t) dt (T g)(x) =

(11.12)

0

is in AC2 ([0, 1]) and is the unique solution of the initial value problem āˆ’f  āˆ’ zf = āˆ’g,

f (0) = 0,

f  (0) = 0.

Proof. Let f = T g. Using Fubiniā€™s theorem, it is straightforward to verify  x y c(y āˆ’ t, k)g(t) dt dy, f (x) = 0

0

which implies that f āˆˆ AC([0, 1]) and  x c(x āˆ’ t, k)g(t) dt. f  (x) = 0

11.1. An initial value problem

365

Similarly, it is proved that f  āˆˆ AC([0, 1]) and  x  f (x) = k 2 s(x āˆ’ t, k)g(t) dt + g(x) = āˆ’zf (x) + g(x), 0

which implies that f āˆˆ AC2 ([0, 1]) and f solves the initial value problem. If the initial value problem had two solutions, their diļ¬€erence would obey F (0) = F  (0) = 0, āˆ’F  āˆ’ zF = 0, so it would be zero by Proposition 11.2.  We now deļ¬ne a linear operator A from AC2 ([0, 1]) to itself by Af = T (V f ). We use A to rewrite the initial value problem (11.6) as an integral equation: Proposition 11.5. Fix a, b, z āˆˆ C, g āˆˆ L1 ([0, 1]), and V āˆˆ L1 ([0, 1]). A function f āˆˆ AC2 ([0, 1]) is a solution of (11.6) if and only if it is a solution of the integral equation f āˆ’ Af = ac(Ā·, k) + bs(Ā·, k) āˆ’ T g.

(11.13)

Proof. If f solves (11.6), then h = f āˆ’ac(Ā·, k)āˆ’bs(Ā·, k) obeys h(0) = h (0) = 0 and āˆ’h āˆ’ zh = āˆ’f  āˆ’ zf = g āˆ’ V f, so h = T (V f āˆ’ g) = Af āˆ’ T g by Lemma 11.4. Conversely, assume that f obeys (11.13). Since (Af )(0) = (Af ) (0) = 0, it follows that f (0) = a and f  (0) = b. Moreover, by Lemma 11.4, āˆ’(Af ) āˆ’ zAf = āˆ’V f and āˆ’(T g) āˆ’ zT g = āˆ’g, so āˆ’f  āˆ’ zf = āˆ’(Af ) āˆ’ zAf + (T g) + zT g = āˆ’V f + g, 

and f solves (11.6).

Since A is a Volterra-type operator, we will solve the integral equation by showing that An  ā‰¤ C n /(n āˆ’ 1)! for some C, and therefore the operator n I āˆ’ A has an inverse given by the convergent series āˆž n=0 A . Bounds on the norm of powers of A can be produced directly in C([0, 1]) or AC2 ([0, 1]). But we will soon also want some sharper estimates which follow the growth rate e|Re k|x already appearing in (11.10) and (11.11); thus, we have already stated the main estimate in that form. Lemma 11.6. If |f (x)| ā‰¤ M e|Re k|x for all x āˆˆ [0, 1], then |(An f )(x)| ā‰¤ for all x āˆˆ [0, 1].

M |||k|||āˆ’n V nL1 e|Re k|x n!

366

11. One-dimensional SchrĀØodinger operators

Proof. Denoting x = t0 and iterating (11.12) gives      n  x t1  tnāˆ’1     n |(A f )(x)| =  Ā·Ā·Ā· s(tjāˆ’1 āˆ’tj , k)V (tj ) f (tn ) dtn Ā· Ā· Ā· dt2 dt1 .  0 0  0 j=1

Applying estimate (11.11) to all the factors s and multiplying with the bound on f gives  tnāˆ’1   x  t1 n āˆ’n |Re k|x n Ā·Ā·Ā· |V (tj )| dtn Ā· Ā· Ā· dt2 dt1 . |(A f )(x)| ā‰¤ M |||k||| e 0

0

0

j=1

By using permutations of (t1 , . . . , tn ) and symmetry, the remaining n-fold x n  integral is ( 0 |V (t)| dt) /n!, and the proof is complete. Theorem 11.7. Fix a, b, z āˆˆ C, g āˆˆ L1 ([0, 1]), and V āˆˆ L1 ([0, 1]). The initial value problem (11.6) has the unique solution f=

āˆž 

An (ac(Ā·, k) + bs(Ā·, k) āˆ’ T g).

(11.14)

n=0

Proof. Applying Lemma 11.6 with M = f C([0,1]) implies that e|Re k| |||k|||āˆ’n V nL1 f C([0,1]) . n! Now (Af ) = V f āˆ’ zAf with (Af )(0) = (Af ) (0) = 0 implies that for f āˆˆ AC2 ([0, 1]) and n āˆˆ N, An f C([0,1]) ā‰¤

An f AC2 ([0,1]) ā‰¤ V L1 Anāˆ’1 f C([0,1]) + |z|An f C([0,1]) . Since f C([0,1]) ā‰¤ f AC2 ([0,1]) , this implies an operator norm estimate of the form Cn An L(AC2 ([0,1])) ā‰¤ (n āˆ’ 1)! n for some constant C independent of n āˆˆ N. Thus, the series āˆž n=0 A is 2 convergent in L(AC ([0, 1])), and it is then trivial to verify that it is the inverse of I āˆ’ A. By invertibility of I āˆ’ A, (11.13) has the unique solution f given by the convergent series (11.14).  We also want to acknowledge the analyticity of the solution in certain parameters. The following result is suļ¬ƒcient for our purposes: Corollary 11.8. Let a = a(z) and b = b(z) be analytic functions from some domain Ī© āŠ‚ C to C. Denote by fz the solution of āˆ’fz + (V āˆ’ z)fz = g,

fz (0) = a(z),

fz (0) = b(z).

Then z ā†’ fz is an analytic function from Ī© to AC2 ([0, 1]).

11.2. Fundamental solutions and transfer matrices

367

Proof. The terms of the series (11.14) are analytic in z by Lemma 2.72. Since the series converges uniformly on compact subsets of Ī©, it is analytic by Lemma 2.71.  As we worked on the space AC2 ([0, 1]) from the start, we can extract an immediate corollary. Since for any y āˆˆ [0, 1] point evaluations f ā†’ f (y) and f ā†’ f  (y) are bounded linear functionals on AC2 ([0, 1]), it follows that they too are analytic functions of z, in the setting of Corollary 11.8. Analyticity of such point evaluations will be used repeatedly. We will also need joint continuity of the solution in the potential and initial condition. Corollary 11.9. Consider convergent sequences aj ā†’ aāˆž and bj ā†’ bāˆž in C, gj ā†’ gāˆž in L1 ([0, 1]), and Vj ā†’ Vāˆž in L1 ([0, 1]). The corresponding solutions of (11.6) converge: fj ā†’ fāˆž in AC2 ([0, 1]). Proof. We ļ¬rst note that in AC2 ([0, 1]), aj c + bj s āˆ’ T gj ā†’ aāˆž c + bāˆž s āˆ’ T gāˆž . Since T : L1 ([0, 1]) ā†’ AC2 ([0, 1]) is a bounded linear operator, for every n āˆˆ N, (T Vj )n (aj c + bj s āˆ’ T gj ) ā†’ (T Vāˆž )n (aāˆž c + bāˆž s āˆ’ T gāˆž ). Thus, each term of the series solution (11.14) converges; the terms are uniformly bounded by the Volterra-type estimates above, so fj ā†’ fāˆž .  As already noted, these results can be rescaled to an arbitrary compact interval instead of [0, 1]; alternatively, this material could have been developed on a more general compact interval from the start, with small diļ¬€erences (e.g., in Lemma 11.3).

11.2. Fundamental solutions and transfer matrices Fundamental solutions are deļ¬ned as solutions u(x, z), v(x, z) of the initial value problems āˆ’āˆ‚x2 u + (V āˆ’ z)u = 0,

u(0, z) = 0,

(āˆ‚x u)(0, z) = 1,

āˆ’āˆ‚x2 v + (V āˆ’ z)v = 0,

v(0, z) = 1,

(āˆ‚x v)(0, z) = 0.

(11.15)

They are the subject of this section; their asymptotic behavior as z ā†’ āˆž will be of great importance. We begin by noting their explicit series representations, using the notation Ī”n (x) = {t āˆˆ Rn | x ā‰„ t1 ā‰„ t2 ā‰„ Ā· Ā· Ā· ā‰„ tn ā‰„ 0}.

368

11. One-dimensional SchrĀØodinger operators

Proposition 11.10. Fundamental solutions and their ļ¬rst derivatives are given by the series representations āˆž   u(x, z) = s(x, k) + s(x āˆ’ t1 , k) n=1 Ī”n (x)

Ɨ

nāˆ’1 

v(x, z) = c(x, k) + Ɨ

V (tj )s(tj āˆ’ tj+1 , k) V (tn )s(tn , k) dn t, (11.16)

j=1 āˆž  

s(x āˆ’ t1 , k)

n=1 Ī”n (x) nāˆ’1 



V (tj )s(tj āˆ’ tj+1 , k) V (tn )c(tn , k) dn t, (11.17)

(āˆ‚x u)(x, z) = c(x, k) + Ɨ



j=1 āˆž  

c(x āˆ’ t1 , k)

n=1 Ī”n (x) nāˆ’1 



V (tj )s(tj āˆ’ tj+1 , k) V (tn )s(tn , k) dn t, (11.18)

j=1

(āˆ‚x v)(x, z) = k 2 s(x, k) + Ɨ

āˆž  

c(x āˆ’ t1 , k)

n=1 Ī”n (x) nāˆ’1 



V (tj )s(tj āˆ’ tj+1 , k) V (tn )c(tn , k) dn t. (11.19)

j=1

Proof. By Proposition 11.5, fundamental solutions solve the integral equations  x s(x āˆ’ t, k)V (t)u(t, z) dt, (11.20) u(x, z) = s(x, k) + 0  x s(x āˆ’ t, k)V (t)v(t, z) dt, (11.21) v(x, z) = c(x, k) + 0

and Theorem 11.7 gives series expansions for u, v, which can be written in the forms (11.16) and (11.17). By the proof of Lemma 11.4, the derivatives of fundamental solutions obey  x c(x āˆ’ t, k)V (t)u(t, z) dt, (11.22) (āˆ‚x u)(x, z) = c(x, k) + 0  x c(x āˆ’ t, k)V (t)v(t, z) dt, (11.23) (āˆ‚x v)(x, z) = k 2 s(x, k) + 0

so substituting (11.16) and (11.17) gives (11.18) and (11.19).



11.2. Fundamental solutions and transfer matrices

369

Proposition 11.11. The series expansions in Proposition 11.10 converge uniformly on (x, z) āˆˆ [0, 1] Ɨ C and uniformly for V in bounded subsets of L1 ([0, 1]). For all z = āˆ’k 2 āˆˆ C and x āˆˆ [0, 1], |u(x, z)| ā‰¤ |||k|||āˆ’1 e|Re k|x+V L1 , |v(x, z)| ā‰¤ e|Re k|x+V L1 , |(āˆ‚x u)(x, z)| ā‰¤ e|Re k|x+V L1 , |(āˆ‚x v)(x, z)| ā‰¤ |||k|||e|Re k|x+V L1 . Proof. As in the proof of Lemma 11.6, using Lemma 11.3 and  n  1 |V (tj )| dn t = V nL1 , n! Ī”n (x) j=1

1 |||k|||āˆ’(n+1) V nL1 e|Re k|x . the nth term of (11.16) is bounded above by n! Summing from n = 0 to āˆž, using |||k|||n+1 ā‰„ |||k|||, and evaluating an exponential series, the estimate for u follows. The other estimates are proved analogously, with the diļ¬€erent powers of |||k||| originating in the application of Lemma 11.3. 

While the previous proposition only gives upper bounds on u, v, the next proposition compares u, v to s, c, by viewing V as a perturbation of āˆ’āˆ‚x2 āˆ’ z; this point of view is especially eļ¬€ective for large z. Proposition 11.12. For all z āˆˆ C and x āˆˆ [0, 1], |u(x, z) āˆ’ s(x, k)| ā‰¤ |||k|||āˆ’2 e|Re k|x+V L1 , |v(x, z) āˆ’ c(x, k)| ā‰¤ |||k|||āˆ’1 e|Re k|x+V L1 , |(āˆ‚x u)(x, z) āˆ’ c(x, k)| ā‰¤ |||k|||āˆ’1 e|Re k|x+V L1 ,   (āˆ‚x v)(x, z) āˆ’ k 2 s(x, k) ā‰¤ e|Re k|x+V L1 . Proof. This is a modiļ¬cation of the previous proof, estimating only terms for n from 1 to āˆž and using āˆž āˆž   1 1 |||k|||āˆ’nāˆ’1 V n ā‰¤ |||k|||āˆ’2 V n ā‰¤ |||k|||āˆ’2 eV L1 n! n!

n=1

n=1

to estimate |u āˆ’ s|. The other estimates are proved analogously.



The Wronskian of functions f, g āˆˆ AC2 ([0, 1]) is deļ¬ned as the absolutely continuous function (11.24) W (f, g) = f g  āˆ’ f  g.

370

11. One-dimensional SchrĀØodinger operators

The key property that makes this useful is that W (f, g) = f g  āˆ’ f  g = (āˆ’f  + V f )g āˆ’ (āˆ’g  + V g)f.

(11.25)

The Wronskian appears in considerations of self-adjointness; here is our ļ¬rst glimpse of that. Lemma 11.13. If f, g āˆˆ AC2 ([0, 1]) and āˆ’f  + V f, āˆ’g  + V g āˆˆ L2 ([0, 1]), then f, āˆ’g  + V g āˆ’ āˆ’f  + V f, g = W (fĀÆ, g)(1) āˆ’ W (fĀÆ, g)(0).

(11.26)

Proof. This follows by integrating (11.25), with f replaced by fĀÆ.



The Wronskian is also valuable when studying eigensolutions: Lemma 11.14. If f, g are two solutions of āˆ’y  +V y = zy, their Wronskian is independent of x. Moreover, W (f, g) = 0 if and only if f, g are linearly dependent. Proof. Independence of x follows from W (f, g) = f g  āˆ’ f  g = (āˆ’f  + V f )g āˆ’ (āˆ’g  + V g)f = zf g āˆ’ zgf = 0. By uniqueness of solutions, a solution of āˆ’h +V h = zh is trivial if and only if h(0) = h (0) = 0. Applying this to a linear combination h = c1 f + c2 g, we conclude that c1 f + c2 g = 0 if and only if ! ! ! c1 0 f  (0) g  (0) = . (11.27) c2 f (0) g(0) 0 Thus, f, g are linearly independent if and only if (11.27) has only the trivial  solution, i.e., if and only if f  (0)g(0) āˆ’ f (0)g  (0) = 0. In particular, since the fundamental solutions u(Ā·, z), v(Ā·, z) are eigensolutions at z, their Wronskian is independent of x; by evaluating at x = 0, we obtain W (v(Ā·, z), u(Ā·, z))(x) = W (v(Ā·, z), u(Ā·, z))(0) = 1

āˆ€x.

(11.28)

These considerations can be written in matrix form. Let us introduce the transfer matrices ! (āˆ‚x u)(x, z) (āˆ‚x v)(x, z) . T (x, z) = u(x, z) v(x, z) This is a 2 Ɨ 2 matrix-valued function, entire in z and absolutely continuous in x. The initial value problems for u, v translate to ! 0 V (x) āˆ’ z T (x, z), T (0, z) = I, āˆ‚x T (x, z) = 1 0

11.2. Fundamental solutions and transfer matrices

371

and (11.28) becomes det T (x, z) = 1. The basic property of the transfer matrix is that it describes the transfer of values from 0 to x for an arbitrary eigensolution at z: Lemma 11.15. If āˆ’f  + V f = zf , then ! ! f  (0) f  (x) = T (x, z) . f (x) f (0)

(11.29)

Proof. Any solution of āˆ’f  + V f = zf can be written as a linear combination of u, v as f = c1 u + c2 v. This implies f  = c1 āˆ‚x u + c2 āˆ‚x v, so ! ! ! ! (āˆ‚x u)(x, z) (āˆ‚x v)(x, z) c1 f  (x) . = c1 + c2 = T (x, z) c2 f (x) u(x, z) v(x, z) Evaluating at x = 0 determines constants as ! ! f  (0) c1 = , c2 f (0) 

which ļ¬nally leads to (11.29).

Transfer matrices will be indispensable for the Weyl disk formalism and various proofs. For now, note that they help us to derive a variation of parameters formula for the solution of (11.6): Lemma 11.16. The solution of the initial value problem āˆ’f  + (V āˆ’ z)f = g, is given by



f (0) = a,

f  (0) = b,

(11.30)

x

(v(x, z)u(t, z)āˆ’v(t, z)u(x, z))g(t) dt. (11.31)

f (x) = av(x, z)+bu(x, z)+ 0

f   f , (11.30) can be written as ! ! ! 0 V āˆ’z g b F+ , F (0) = . 1 0 0 a

Proof. In terms of F = F =

Using (T āˆ’1 ) = āˆ’T āˆ’1 T  T āˆ’1 , this leads to ! āˆ’1  āˆ’1 g(t) (T (t, z) F (t)) = T (t, z) , 0 Integrating from 0 to x gives T (x, z)

āˆ’1

F (x) =

T (0, z)

āˆ’1

F (0) =

! b . a

!  x ! b āˆ’1 g(t) + T (t, z) dt. a 0 0

Using det T (t, z) = 1 and the standard formula for the 2 Ɨ 2 matrix inverse, multiplying by T (x, z) and taking the second entry of the resulting identity gives (11.31). 

372

11. One-dimensional SchrĀØodinger operators

Recalling that u, v are entire AC2 ([0, 1])-valued functions of z, it is natural to ask about their derivatives in the same sense. Proposition 11.17. The AC2 ([0, 1])-valued derivatives of u, v are given by  x (v(x, z)u(t, z) āˆ’ v(t, z)u(x, z))u(t, z) dt, (11.32) (āˆ‚z u)(x, z) = 0  x (v(x, z)u(t, z) āˆ’ v(t, z)u(x, z))v(t, z) dt. (11.33) (āˆ‚z v)(x, z) = 0

Proof. Fix z āˆˆ C and consider u(Ā·, z + h) for h āˆˆ C. Since āˆ’(āˆ‚x2 u)(x, z + h) + (V (x) āˆ’ z)u(x, z + h) = hu(x, z + h), viewing u(x, z + h) as a solution of this inhomogeneous initial value problem and using (11.31) implies  x (v(x, z)u(t, z) āˆ’ v(t, z)u(x, z))u(t, z + h) dt. u(x, z + h) = u(x, z) + h 0

If we rewrite this as  x u(x, z + h) āˆ’ u(x, z) = (v(x, z)u(t, z) āˆ’ v(t, z)u(x, z))u(t, z + h) dt h 0 and use limhā†’0 u(Ā·, z +h) = u(Ā·, z), the ļ¬rst formula follows. The second formula is proved analogously. Note that it was convenient to evaluate the derivative pointwise; since we already know the derivative exists in AC2 ([0, 1]), the two must be equal.  As deļ¬ned, the transfer matrix depends only on the potential V . In later sections, it will sometimes be appropriate to incorporate a boundary condition at 0. Corresponding to the boundary condition (11.2) for some Ī± āˆˆ R, we consider the eigensolutions Ļ†(x, z), Īø(x, z) at z, with initial conditions ! ! āˆ‚x Ļ†(0, z) āˆ‚x Īø(0, z) cos Ī± āˆ’ sin Ī± āˆ’1 RĪ± = = RĪ± , sin Ī± cos Ī± Ļ†(0, z) Īø(0, z) (the special case Ī± = 0 gives Ļ† = u, Īø = v), and consider the transfer matrices ! āˆ‚x Ļ†(x, z) āˆ‚x Īø(x, z) . TĪ± (x, z) = RĪ± Ļ†(x, z) Īø(x, z) Note that TĪ± (x, z) = RĪ± T (x, z)RĪ±āˆ’1 . These obey the initial value problem ! 0 V (x) āˆ’ z TĪ± (0, z) = I. RĪ±āˆ’1 TĪ± (x, z), āˆ‚x TĪ± (x, z) = RĪ± 1 0 We conclude this section with an important lemma about the independence of values at diļ¬€erent points. Fix z āˆˆ C. For any h āˆˆ L2 ([0, 1]), let f āˆˆ AC2 ([0, 1]) be the unique solution of āˆ’f  + (V āˆ’ z)f = h,

f (0) = 0,

f  (0) = 0.

(11.34)

11.3. SchrĀØodinger operators with two regular endpoints

373

Since the solution is unique and linear in h, the map h ā†’ f is a linear operator B : L2 ([0, 1]) ā†’ AC2 ([0, 1]). In particular, the values f (1) and f  (1) and their linear combinations, are linear functionals of h āˆˆ L2 ([0, 1]). The following lemma describes the functions which correspond to these functionals in the sense of Rieszā€™s representation theorem, and we obtain important corollaries from this. Lemma 11.18. Let V āˆˆ L1 ([0, 1]) and z āˆˆ C. (a) Fix Ī±, Ī² āˆˆ C and let g āˆˆ AC2 ([0, 1]) be the solution of āˆ’g  + (V āˆ’ z)g = 0,

g(1) = Ī±,

g  (1) = Ī².

For all h āˆˆ L2 ([0, 1]), the solution f of (11.34) obeys ĀÆ (1) āˆ’ Ī± Ī²f ĀÆ f  (1) = g, h.

(11.35)

(b) For any Ī³, Ī“ āˆˆ C, there exists h āˆˆ L2 ([0, 1]) such that the solution of (11.34) obeys f (1) = Ī³, f  (1) = Ī“. (c) The set of h āˆˆ L2 ([0, 1]), for which the solution f of (11.34) obeys f (1) = f  (1) = 0, is the orthogonal complement {g | āˆ’g  + (V āˆ’ z)g = 0}āŠ„ . Proof. (a) This is immediate from (11.26). (b) The initial value problem (11.34) deļ¬nes a linear map L2 ([0, 1]) ā†’ C2 by h ā†’ (f (1), f  (1)). If this map was not onto, its range would be a proper subspace of C2 , so there would exist a choice of (Ī±, Ī²) = (0, 0) for which (11.35) is the trivial functional. This is a contradiction, because the vector g corresponding to that functional is nontrivial, as a nontrivial solution of āˆ’g  + (V āˆ’ z)g = 0. (c) f (1) = f  (1) = 0 if and only if (11.35) holds for all Ī±, Ī² āˆˆ C, so if  and only if h is orthogonal to all solutions of āˆ’g  + (V āˆ’ z)g = 0. Part (b), rescaled from [0, 1] to an arbitrary compact interval [c, d], shows that there are no hidden constraints between values of f at diļ¬€erent points, where f is an arbitrary function in the domain of a SchrĀØ odinger operator.

11.3. SchrĀØ odinger operators with two regular endpoints In this section, we will assume that V āˆˆ L1 ([0, 1]) is real-valued, and study SchrĀØ odinger operators on L2 ([0, 1]) with boundary conditions (11.2) and (11.3); this is an important special case which already illustrates the use of the methods developed above.

374

11. One-dimensional SchrĀØodinger operators

Theorem 11.19. Let V āˆˆ L1 ([0, 1]) be real-valued, Ī±, Ī² āˆˆ R, and let H be the operator on L2 ([0, 1]) deļ¬ned by D(H) = {f āˆˆ AC2 ([0, 1]) | āˆ’f  + V f āˆˆ L2 ([0, 1]), f obeys (11.2) and (11.3)} and Hf = āˆ’f  + V f . Then the following hold. (a) H is self-adjoint. (b) H has a complete orthonormal basis of eigenvectors, i.e., there is a sequence of fn āˆˆ L2 ([0, 1]) such that (fn )āˆž n=1 is an orthonormal basis of L2 ([0, 1]) and Hfn = Ī»n fn for some Ī»n āˆˆ R. (c) All eigenvalues are simple, the set of eigenvalues is discrete, and Ļƒ(H) = {Ī»n | n āˆˆ N}. (d) For all z āˆˆ C \ Ļƒ(H), (H āˆ’ z)āˆ’1 is a compact integral operator. Symmetry of the operator will follow from Lemma 11.13. We will prove further properties by deriving an explicit formula for the inverse (Hāˆ’z)āˆ’1 for zāˆˆ / Ļƒ(H) and the spectral theorem for compact self-adjoint operators. The description of (H āˆ’ z)āˆ’1 requires us to consider nontrivial eigensolutions which obey the boundary conditions at 0 or 1, respectively, so we deļ¬ne Ļˆ Ā± (x, z) as the solutions of āˆ’āˆ‚x2 Ļˆ + V Ļˆ = zĻˆ,

(11.36)

which obey Ļˆ āˆ’ (0, z) = āˆ’ sin Ī±, Ļˆ + (1, z) = āˆ’ sin Ī²,

(āˆ‚x Ļˆ āˆ’ )(0, z) = cos Ī±, (āˆ‚x Ļˆ + )(1, z) = āˆ’ cos Ī².

By Corollary 11.8, Ļˆ Ā± (Ā·, z) are entire functions of z. For Ļˆ + (Ā·, z), to conclude this from Corollary 11.8, use the linear substitution x = 1 āˆ’ t to reduce to initial conditions at 0. The notation Ļˆ Ā± (x, z) is useful for emphasizing the essential properties of Ļˆ Ā± as functions of z, but it will often be convenient to use the more compact notation ĻˆzĀ± (x) = Ļˆ Ā± (x, z). Lemma 11.20. Let z āˆˆ C. The kernel Ker(H āˆ’ z) is nontrivial if and only if W (Ļˆz+ , Ļˆzāˆ’ ) = 0. In this case, dim Ker(H āˆ’ z) = 1. Proof. An eigensolution at z must be a multiple of Ļˆzāˆ’ in order to obey the boundary condition at 0, and a multiple of Ļˆz+ in order to obey the boundary condition at 1. Thus, nontrivial eigensolutions exist if and only if Ļˆzāˆ’ , Ļˆz+ are linearly dependent, in which case they are multiples of ĻˆzĀ± . Thus, z is an eigenvalue if and only if W (Ļˆz+ , Ļˆzāˆ’ ) = 0, and every eigenspace is one dimensional. 

11.3. SchrĀØodinger operators with two regular endpoints

375

In particular, if W (Ļˆz+ , Ļˆzāˆ’ ) = 0, then z āˆˆ Ļƒ(H). Conversely: Proposition 11.21. If W (Ļˆz+ , Ļˆzāˆ’ ) = 0, then H āˆ’ z has a bounded inverse. The inverse is the compact integral operator  1 ((H āˆ’ z)āˆ’1 g)(x) = G(x, y; z)g(y) dy (11.37) 0

with the kernel G(x, y; z) =

1 Ļˆzāˆ’ (min(x, y))Ļˆz+ (max(x, y)). + W (Ļˆz , Ļˆzāˆ’ )

The kernel G is called Greenā€™s function; the idea of the proof is that G constructed in this way obeys (āˆ’āˆ‚x2 + V (x) āˆ’ z)G(x, y; z) = Ī“y (x),

(11.38)

where Ī“y denotes the Dirac delta function centered at y. The proof we will present will not use any distributional calculus, but the Heaviside function will appear. Proof. Since G(x, y; z) is jointly continuous in (x, y) āˆˆ [0, 1]2 , the righthand side of (11.37) is a compact integral operator. To prove (11.37), since H āˆ’z is injective, it suļ¬ƒces to prove that, for any g āˆˆ L2 ([0, 1]), the function  1 G(x, y; z)g(y) dy (11.39) f (x) = 0

is in D(H) and that (H āˆ’ z)f = g. For ļ¬xed y, G is absolutely continuous in x and

(Ļˆzāˆ’ ) (x)Ļˆz+ (y) x < y 1 Ɨ āˆ‚x G(x, y; z) = W (Ļˆz+ , Ļˆzāˆ’ ) Ļˆzāˆ’ (y)(Ļˆz+ ) (x) x > y.

(11.40)

The function āˆ‚x G(x, y; z) has at x = y a jump of size āˆ’1 because lim āˆ‚x G(x, y; z) āˆ’ lim āˆ‚x G(x, y; z) = xā†“y

xā†‘y

Ļˆzāˆ’ (x)(Ļˆz+ ) (x) āˆ’ (Ļˆzāˆ’ ) (x)Ļˆz+ (x) , W (Ļˆz+ , Ļˆzāˆ’ )

and the numerator is equal to āˆ’W (Ļˆz+ , Ļˆzāˆ’ ). However, denoting by h(x) = 1 2 (1 + sgn x) the Heaviside function, āˆ‚x G(x, y; z) + h(x āˆ’ y) āˆˆ AC([0, 1]) and

(Ļˆzāˆ’ ) (x)Ļˆz+ (y) x < y 1 āˆ‚x (āˆ‚x G(x, y; z) + h(x āˆ’ y)) = + āˆ’ Ɨ W (Ļˆz , Ļˆz ) Ļˆzāˆ’ (y)(Ļˆz+ ) (x) x > y, so since ĻˆzĀ± are solutions of (11.36), āˆ‚x (āˆ‚x G(x, y; z) + h(x āˆ’ y)) = (z āˆ’ V (x))G(x, y; z). Of course, this is another way of expressing (11.38).

(11.41)

376

11. One-dimensional SchrĀØodinger operators

For any s < t, multiplying (11.40) by g(y) and integrating in (x, y) āˆˆ [s, t] Ɨ [0, 1] shows by Fubiniā€™s theorem that  t 1  1 āˆ‚x G(x, y; z)g(y) dy dx = [G(x, y; z)]ts g(y) dy = f (t) āˆ’ f (s). s

0

0

Since s < t is arbitrary, this implies that f āˆˆ AC([0, 1]) and  1  āˆ‚x G(x, y; z)g(y) dy. f (x) =

(11.42)

0

By the same arguments, from (11.41) we obtain  t 1  t (V (x) āˆ’ z)f (x) dx = (z āˆ’ V (x))G(x, y; z)g(y) dy dx s s 0  1 [(āˆ‚x G(x, y; z) + h(x āˆ’ y))]x=t = x=s g(y) dy 0  t   g(y) dy, = f (t) āˆ’ f (s) + s

so

f

āˆˆ AC([0, 1]) and

f 

= (V āˆ’ z)f āˆ’ g.

Finally, using (11.39) and (11.42),  1 1 Ļˆzāˆ’ (0)Ļˆz+ (y)g(y) dy, f (0) = W (Ļˆz+ , Ļˆzāˆ’ ) 0  1 1  f (0) = (Ļˆzāˆ’ ) (0)Ļˆz+ (y)g(y) dy, W (Ļˆz+ , Ļˆzāˆ’ ) 0 so since Ļˆzāˆ’ obey the boundary condition at 0, so does f . Analogous calculations show that f obeys the boundary condition at 1, so f āˆˆ D(H).  Proof of Theorem 11.19. If f, g āˆˆ D(H), then by (11.2), (f (0), f  (0)) and (g(0), g  (0)) are both multiples of (āˆ’ sin Ī±, cos Ī±) in C2 , so    g (0) f  (0)  = 0. W (f , g)(0) =  g(0) f (0)  Similarly, W (f , g)(1) = 0, so f, Hg = Hf, g

āˆ€f, g āˆˆ D(H),

i.e., H is a symmetric operator. Thus, its eigenvalues are real by Lemma 8.16. Since all eigenvalues of H are real, the entire function z ā†’ W (Ļˆz+ , Ļˆzāˆ’ ) has zeros only on R; in particular, it is not identically zero and the set of its zeros are discrete. For z āˆˆ C with W (Ļˆz+ , Ļˆzāˆ’ ) = 0, the operator (H āˆ’ z)āˆ’1 is compact. In particular, if z is real and W (Ļˆz+ , Ļˆzāˆ’ ) = 0, then G(x, y; z) = G(y, x; z) so (H āˆ’ z)āˆ’1 is a compact self-adjoint operator. By the spectral theorem for

11.3. SchrĀØodinger operators with two regular endpoints

377

compact self-adjoint operators, it has an orthonormal basis of eigenfunctions (vn )āˆž n=1 , (H āˆ’ z)āˆ’1 vn = an vn

(11.43)

z)āˆ’1

(note an = 0 because (H āˆ’ is injective). The functions vn are also eigenfunctions of H because (11.43) is equivalent to (H āˆ’ z)vn = aāˆ’1 n vn and + z)v . Hvn = (aāˆ’1 n n Self-adjointness of H now follows from Example 8.21 applied to (Hāˆ’Ī»)āˆ’1 for some Ī» āˆˆ R \ Ļƒ(H).  The above argument was quite qualitative and relied on compactness; however, eigenvalues can be located much more precisely. Let us study d2 the locations of Dirichlet eigenvaluesā€”eigenvalues of H = āˆ’ dx 2 + V with Dirichlet boundary conditions Ī± = Ī² = 0. In this case, using the fundamental solution u, we note that u(1, z) = 0 if and only if z is a Dirichlet eigenvalue. Thus, we want to think of the entire function u(1, z) as an analogue of the characteristic polynomial. In that analogy, the following lemma guarantees equality of algebraic and geometric multiplicities. Lemma 11.22. If V is real-valued, the function u(1, z) has only simple zeros. Proof. Assume u(1, z) = 0. Then z āˆˆ R. From (11.32) it follows that  1 (v(1, z)u(t, z) āˆ’ v(t, z)u(1, z))u(t, z) dt (āˆ‚z u)(1, z) = 0  1 = v(1, z) u(t, z)2 dt. 0

1 Now v(1, z) = 0 because of (11.28), and 0 u(t, z)2 dt = 0 because u(Ā·, z) is  a nontrivial real-valued function. Thus, (āˆ‚z u)(1, z) = 0. Using the characteristic function u(1, z), we can obtain more precise information about the distribution of the eigenvalues. This will be an application of RouchĀ“eā€™s theorem, using the special case V = 0 for comparison. Since complex analytic techniques count zeros with multiplicity, Lemma 11.22 will be useful. Example 11.23. For V = 0 and Dirichlet boundary conditions at 0 and 1, the spectrum of the SchrĀØodinger operator is the set {n2 Ļ€ 2 | n āˆˆ N}. Proof. V = 0 implies u(1, z) = s(1, k), so it suļ¬ƒces to solve the equation s(1, k) = 0. This equation has solutions k = iĻ€n, n āˆˆ Z\{0}, so the Dirichlet  eigenvalues are z = āˆ’k 2 = n2 Ļ€ 2 .

378

11. One-dimensional SchrĀØodinger operators

Lemma 11.24 (Counting lemma). Consider the operator H corresponding to Dirichlet boundary conditions at 0 and 1 and a real-valued potential V āˆˆ L1 . For positive integers N > eV L1 , the operator H has exactly N eigenvalues smaller than (N + 12 )2 Ļ€ 2 . Proof. We begin by noting that the estimate āˆš āˆš 3|sinh āˆ’z| > e|Re āˆ’z|

(11.44)

holds on some curves z āˆˆ C. It holds on the parabolas ! āˆš 1 Ļ€ |Im āˆ’z| = N + 2 āˆš for N āˆˆ N, because if āˆ’z = x Ā± i(N + 12 )Ļ€, then āˆš āˆš 2|sinh āˆ’z| = 2|cosh x| = |ex + eāˆ’x | > e|x| = e|Re āˆ’z| .

(11.45)

The estimate (11.44) also holds on the parabolas āˆš |Re āˆ’z| = CĻ€ āˆš for C ā‰„ 1, since if āˆ’z = x + iy and |x| ā‰„ 1, then

(11.46)

1 1 |sinh(x + iy)|2 = sinh2 x + sin2 y ā‰„ sinh2 x > (e2|x| āˆ’ 2) > e2|x| . 4 9 Im z

āˆš |Im āˆ’z| = 52 Ļ€

āˆš |Im āˆ’z| = 32 Ļ€

āˆš |Re āˆ’z| = 1.1Ļ€

Ļ€2

4Ļ€ 2

9Ļ€ 2

Re z

Figure 11.1. Contours used in the proof of Lemma 11.24 and Dirichlet eigenvalues for V = 0.

11.4. Endpoint behavior

379

These two kinds of parabolas are illustrated in Figure11.1. The parabolas (11.45) and (11.46) intersect at two points and deļ¬ne a closed contour which encloses exactly N zeros of s(1, z) and on which (11.44) holds. If N, C > eV L1 , then on this contour, |z|1/2 > Ļ€eV L1 , so by the basic estimate for u(x, z), āˆš 1 < |z|āˆ’1/2 e|Re āˆ’z| < |s(1, k)|. 3 All the zeros are real by self-adjointness. Thus, by RouchĀ“eā€™s theorem, u(1, z) has exactly N zeros including multiplicity on the interval

|u(1, z) āˆ’ s(1, k)| ā‰¤ |z|āˆ’1 e|Re

āˆš āˆ’z|+V L1

(āˆ’C 2 Ļ€ 2 , (N + 1/2)2 Ļ€ 2 ). Since C can be arbitrarily large, u(1, z) has exactly N zeros including multiplicity on the interval (āˆ’āˆž, (N + 1/2)2 Ļ€ 2 ). Since this holds for all large enough N , the conclusion follows.



The result can also be restated in the following way. Corollary 11.25. Consider the operator H corresponding to a real-valued potential V āˆˆ L1 ([0, 1]) and Dirichlet boundary conditions at 0 and 1. The spectrum of H is bounded from below. Arranging its elements in increasing order, Ļƒ(H) = {Ī»n | n āˆˆ N}, with Ī»n < Ī»n+1 , the eigenvalues obey the asymptotics n ā†’ āˆž. Ī»n = n2 Ļ€ 2 + O(n), The asymptotic behavior of Dirichlet eigenvalues and eigenvectors can be studied much more precisely; see [71]. The nth eigenvector for the Dirichlet boundary conditions has precisely n āˆ’ 1 zeros in (0, 1); this is a special case of Sturm oscillation theory, see survey [90].

11.4. Endpoint behavior In a more general setting, V may not be integrable on the entire interval or the interval may be inļ¬nite. We will assume that V āˆˆ L1loc (I), that is, V āˆˆ L1 ([c, d]) for every compact subinterval [c, d] āŠ‚ I. To describe the amount of smoothness required from functions in the domain, we denote AC2loc (I) = {f āˆˆ ACloc (I) | f  āˆˆ ACloc (I)}.

(11.47)

Lemma 11.26. Let x0 āˆˆ I and z, Ī±, Ī² āˆˆ C. If g āˆˆ L1loc (I), then there exists a unique solution f āˆˆ AC2loc (I) of the initial value problem āˆ’f  + (V āˆ’ z)f = g,

f (x0 ) = Ī±,

f  (x0 ) = Ī².

(11.48)

380

11. One-dimensional SchrĀØodinger operators

Proof. By using Theorem 11.7 and aļ¬ƒne transformations of the interval, there is a unique solution on any compact intervals [x0 , x0 + L] āŠ‚ I and [x0 āˆ’ L, x0 ] āŠ‚ I. Solutions on overlapping intervals must match, so by using an increasing sequence of compact intervals whose union is I, the result follows.  We introduce the local domain Dloc = {f āˆˆ AC2loc (I) | āˆ’f  + V f āˆˆ L2loc (I)}. Under the stronger assumption V āˆˆ L2loc (I), this would be equivalent to Dloc = {f āˆˆ AC2loc (I) | f  āˆˆ L2loc (I)}, but we are working with the more general condition V āˆˆ L1loc (I); due to this, the local domain depends on V . Functions in Dloc need not have any integrability properties at the endpoints, so we also deļ¬ne Xāˆ’ = {f āˆˆ Dloc | āˆƒc > āˆ’ : f, āˆ’f  + V f āˆˆ L2 ((āˆ’ , c))}, X+ = {f āˆˆ Dloc | āˆƒc < + : f, āˆ’f  + V f āˆˆ L2 ((c, + ))}. Since any f āˆˆ AC2loc (I) is bounded on compact intervals, the deļ¬ning conditions in XĀ± are actually independent of the choice of c āˆˆ I. In particular, Xāˆ’ āˆ© X+ = {f āˆˆ AC2loc (I) | f āˆˆ L2 (I), āˆ’f  + V f āˆˆ L2 (I)}. The sets XĀ± encode required properties of a function near an endpoint Ā± , and the following separation property shows that these are independent of each other. Lemma 11.27. For every fĀ± āˆˆ XĀ± , there exists f āˆˆ Xāˆ’ āˆ© X+ such that f = fāˆ’ on some interval (āˆ’ , c) and f = f+ on some interval (d, + ). Proof. Fix [c, d] āŠ‚ I. By applying Lemma 11.18 on intervals [c, c+d 2 ] and c+d 2  2 [ 2 , d], there exists f āˆˆ AC ([c, d]) such that āˆ’f + V f āˆˆ L ([c, d]) and ! ! ! ! ! ! 0 f  (d) f  ( c+d fāˆ’ (c) ) f+ (d) f  (c) 2 = , , . = = 0 f (c) fāˆ’ (c) f (d) f+ (d) f ( c+d 2 ) Extend f to a function on I by setting f (x) = fāˆ’ (x) for x < c and f (x) = f+ (x) for x > d. Then f āˆˆ AC2loc (I) and f has all the required properties.  This lemma would be almost trivial if we were working with V āˆˆ L2loc (I): then we could take any f āˆˆ AC2 ([c, d]) with desired values of f and f  at c and d. The set Xāˆ’ āˆ© X+ will be a maximal domain for a SchrĀØodinger operator, but to obtain self-adjoint operators, it may be necessary to restrict this domain by boundary conditions. Here, Wronskians will play a role.

11.4. Endpoint behavior

381

The Wronskian of two functions f, g āˆˆ AC2loc (I) is the function W (f, g) = f g  āˆ’ f  g.

(11.49)

While functions in XĀ± do not necessarily have boundary values at Ā± , their Wronskians do: Proposition 11.28. (a) For any f, g āˆˆ Xāˆ’ , the limit Wāˆ’ (f, g) = lim W (f, g)(x) xā†“āˆ’

is convergent. (b) For any f, g āˆˆ X+ , the limit W+ (f, g) = lim W (f, g)(x) xā†‘+

is convergent. (c) For any f, g āˆˆ Xāˆ’ āˆ© X+ , āˆ’f  + V f, g āˆ’ f, āˆ’g  + V g = W+ (f , g) āˆ’ Wāˆ’ (f , g).

(11.50)

Proof. (a) As in Lemma 11.13, on any compact interval [c, d] āŠ‚ I,  d  d (āˆ’f  + V f )g dx āˆ’ f (āˆ’g  + V g) dx = W (f , g)(d) āˆ’ W (f , g)(c). c

c

(11.51) +V +Vg āˆˆ Since āˆ’ , d)), by the Cauchyā€“Schwarz inequality and dominated convergence, (11.51) has a ļ¬nite limit as c ā†“ āˆ’ . f, g, āˆ’f 

f, āˆ’g 

L2 ((

(b) This is analogous to (a). (c) Taking the limit of (11.51) as c ā†“ āˆ’ and d ā†‘ + gives (11.50).



In equation (11.50), the diļ¬€erence of boundary Wronskians appears as the obstruction to self-adjointness. We have to understand these better in order to describe choices of domain which lead to self-adjoint operators. For either choice of Ā± sign, the Wronskian WĀ± is an alternating bilinear map, so the framework of Section 8.7 applies. We denote āˆ— = {f āˆˆ XĀ± | WĀ± (f, g) = 0 āˆ€g āˆˆ XĀ± }. XĀ±

(11.52)

By Section 8.7, this is a vector subspace of XĀ± and WĀ± induces a symplecāˆ— . The ļ¬rst step is to estimate the tic form on the quotient space XĀ± /XĀ± dimension of this quotient: āˆ— is Lemma 11.29. At each endpoint Ā± , the quotient vector space XĀ± /XĀ± trivial or two dimensional.

382

11. One-dimensional SchrĀØodinger operators

Proof. For every x āˆˆ I, the Wronskian at x corresponds through the point evaluation ! f  (x) f ā†’ f (x) ucker identity. This follows to a symplectic form on C2 , so it obeys the PlĀØ from Theorem 8.64 or the following calculation. Starting from    f1 (x) f2 (x) f3 (x) f4 (x)   f1 (x) f2 (x) f3 (x) f4 (x)   f  (x) f  (x) f  (x) f  (x) = 0, 1 2 3 4   f1 (x) f2 (x) f3 (x) f4 (x) we obtain W (f1 , f2 )W (f3 , f4 ) āˆ’ W (f1 , f3 )W (f2 , f4 ) + W (f1 , f4 )W (f2 , f3 ) = 0. This is true at any x āˆˆ I, so by taking x ā†’ Ā± , we conclude that the bounducker identity. Thus, by Theorem 8.64, ary Wronskian WĀ± also obeys the PlĀØ āˆ— is 0 or 2.  the dimension of the quotient space XĀ± /XĀ± Deļ¬nition 11.30. At the endpoint Ā± , the potential V is the limit point if āˆ— ) = 0 and the limit circle if dim(X /X āˆ— ) = 2. dim(XĀ± /XĀ± Ā± Ā± The choice of terminology is motivated by Weyl disk formalism, which will be explained later. In the remainder of this section, we present two important special cases and explain how they ļ¬t in the limit pointā€“limit circle alternative. A regular endpoint has been deļ¬ned by (11.4). Informally speaking, regular endpoints behave just like internal points of the interval. For concreteness, let us work with āˆ’ ; of course, analogous statements hold for + . The endpoint āˆ’ is called regular for the potential V if it is a ļ¬nite endpoint (āˆ’ = āˆ’āˆž) and V āˆˆ L1 ((āˆ’ , c)) for some, and therefore all, c āˆˆ I. Proposition 11.31. Let āˆ’ be a regular endpoint of V . (a) For every f āˆˆ Xāˆ’ , the limits f (āˆ’ ) := lim f (x), xā†“āˆ’

f  (āˆ’ ) := lim f  (x) xā†“āˆ’

exist, and f extends to a function on {āˆ’ } āˆŖ I so that for d āˆˆ I, f āˆˆ AC2 ([āˆ’ , d]),

āˆ’f  + V f āˆˆ L2 ([āˆ’ , d]).

(b) For every f, g āˆˆ Xāˆ’ , Wāˆ’ (f, g) = f (āˆ’ )g  (āˆ’ ) āˆ’ f  (āˆ’ )g(āˆ’ ).

11.4. Endpoint behavior

383

(c) The map T : Xāˆ’ ā†’ C2 deļ¬ned by T : f ā†’

f  (āˆ’ ) f (āˆ’ )

!

āˆ—. has Ran T = C2 and Ker T = Xāˆ’

(d) V is the limit circle at āˆ’ . Proof. (a) For f āˆˆ Xāˆ’ , consider g = āˆ’f  + V f āˆˆ L2 ((āˆ’ , c)) āŠ‚ L1 (āˆ’ , c). By an aļ¬ƒne transformation, Theorem 11.7 can be applied on the interval [āˆ’ , c], and it provides existence of a function F āˆˆ AC2 ([āˆ’ , c]) with āˆ’F  + V F = g and F (c) = f (c), F  (c) = f  (c). By uniqueness of solutions, F = f on every interval [āˆ’ + , c] with āˆ’ < āˆ’ + < c, so the extension of f is given by F . (b) This follows from (a) by computing the limit in the deļ¬nition of Wāˆ’ . (c) The equation āˆ’h + V h = 0 has a solution h āˆˆ AC2 ([āˆ’ , c]) with any prescribed values of h(āˆ’ ) = Ī± and h (āˆ’ ) = Ī², and h extends to a function āˆ— if and only if h āˆˆ Xāˆ’ , so Ran T = C2 . Using (b), it follows that f āˆˆ Xāˆ’  f (āˆ’ ) = f (āˆ’ ) = 0. āˆ— ) = dim(X / Ker T ) = dim Ran T = 2. (d) dim(Xāˆ’ /Xāˆ’ āˆ’



For an inļ¬nite endpoint Ā± = Ā±āˆž, a diļ¬€erent perspective is needed. Standard Sobolev estimates give upper bounds on f  in terms of f and f  , and as a variation of that idea, we need an upper bound on f  in terms of f and āˆ’f  + V f . We will work under the assumption (11.5) and prove that the endpoint is in the limit point case. To avoid confusion, we state the following proposition for + ; analogous statements hold for āˆ’ . Proposition 11.32. Assume that V āˆˆ L1loc (I) and that (11.5) holds at the endpoint + = +āˆž. Then the following hold. āˆž (a) For any f āˆˆ X+ and c āˆˆ I, c |f  |2 dx < āˆž. (b) V is a limit point at +āˆž. (c) For any f āˆˆ X+ , limxā†’+āˆž f (x) = 0. (d) For any c āˆˆ (āˆ’ , +āˆž), there exists M < āˆž such that for all f āˆˆ X+ with f (c) = 0 or f  (c) = 0,   āˆž  āˆž 1 āˆž 2 2 |f | dx ā‰¤ M |f | dx + Re f (āˆ’f  + V f ) dx. (11.53) 2 c c c  x+1 The constant M depends only on supxā‰„c x Vāˆ’ (t) dt. The proof starts with a simple Sobolev estimate:

384

11. One-dimensional SchrĀØodinger operators

Lemma 11.33. If [p, q] is an interval of length 12 ā‰¤ q āˆ’ p ā‰¤ 1, such that f āˆˆ AC([p, q]) and f  āˆˆ L2 ([p, q]), then for any > 0, ! q  q 1 2  2 sup |f (x)| ā‰¤

|f (x)| dx + 2 + |f (x)|2 dx.

p p xāˆˆ[p,q] Proof. Since f 2 āˆˆ AC([p, q]) and (f 2 ) = 2f f  , for any x, y āˆˆ [p, q],     max(x,y)  q   (f 2 ) dt ā‰¤ 2|f f  | dt. |f (x)2 āˆ’ f (y)2 | =   min(x,y)  p By the Cauchyā€“Schwarz inequality, this implies   q 1 q |f  (t)|2 dt + |f (t)|2 dt. |f (x)|2 ā‰¤ |f (y)|2 +

p p

(11.54)

By the mean value theorem, there exists y āˆˆ [p, q] such that |f (y)|2 =  q 1 2  qāˆ’p p |f (t)| dt. Using that value of y in (11.54) concludes the proof. Proof of Proposition 11.32. (a) Let us ļ¬x c āˆˆ I and denote  x+1 Vāˆ’ (t) dt. C = sup xā‰„c

x

This is a ļ¬nite constant since V āˆˆ L1loc (I) and (11.5) holds. For any d > c, integration by parts implies  d  d d  2 |f | dx = f (āˆ’f  ) dx + f f  c c

c



d

=āˆ’



d

V |f | dx + 2

c

c

d f (āˆ’f  + V f ) dx + f f  c .

(11.55)

On any interval [p, q] āŠ‚ [c, āˆž) of length between 1/2 and 1, Lemma 11.33 allows us to estimate ! q  q  q  q 1 2 2  2 āˆ’ V |f | dx ā‰¤ Vāˆ’ |f | dx ā‰¤ C

|f | dx + C 2 + |f |2 dx.

p p p p (11.56) For any d ā‰„ c + 1, the interval [c, d] can be partitioned into intervals of length between 1/2 and 1. Summing over those intervals shows that (11.56) holds also for p = c, q = d. Combining with (11.55) implies  d |f  |2 dx (1 āˆ’ C ) c (11.57) ! d  d  1 2   d |f | dx + Re f (āˆ’f + V f ) dx + Re f f c . ā‰¤C 2+

c c To proceed, we need < 1/C; in fact, let us ļ¬x = 1/(2C).

11.4. Endpoint behavior

385

Let us take lim inf dā†’āˆž of both sides of (11.57). In fact, many of the terms in (11.57) have a limit as d ā†’ āˆž. The ļ¬rst term on the right-hand side converges because f āˆˆ L2 ((c, āˆž)), and the second by the Cauchyā€“Schwarz inequality since f, āˆ’f  + V f āˆˆ L2 ((c, āˆž)). Now we prove by contradiction that  lim inf Re f (x)f  (x) ā‰¤ 0. (11.58) xā†’āˆž

If this was false, that would imply (|f (x)|2 ) ā‰„ Ī“ > 0 for all x large enough, and therefore |f |2 would grow at least linearly at +āˆž, contradicting f āˆˆ L2 ((c, āˆž)). Thus, taking lim inf dā†’āˆž of (11.57) implies that  d 1 lim inf |f  |2 dx 2 dā†’āˆž c  āˆž  āˆž 2 |f | dx + Re f (āˆ’f  + V f ) dx āˆ’ Re(fĀÆ(c)f  (c)). ā‰¤ 2C(C + 1) c

c

(11.59)

āˆž The left-hand side is equal to c |f  |2 dx by nonnegativity of |f  |2 and monotone convergence, and this proves that f  āˆˆ L2 ((c, āˆž)). (b) For any f āˆˆ X+ , the functions f and f  are square-integrable on (c, āˆž). For any f, g āˆˆ X+ , the Cauchyā€“Schwarz inequality implies f g  , f  g āˆˆ L1 ((c, āˆž)). Thus, W (f, g) āˆˆ L1 ((c, āˆž)); therefore, the only possible value of the limit W+ (f, g) = limxā†’āˆž W (f, g)(x) is 0. (c) For any f āˆˆ X+ , f, f  āˆˆ L2 ((c, āˆž)) implies that  x+1  x+1 2 lim |f (t)| dt = lim |f  (t)|2 dt = 0, xā†’+āˆž x

xā†’+āˆž x

and then Lemma 11.33 implies that limxā†’+āˆž f (x) = 0. (d) This follows immediately from (11.59) using (11.58).



Part (d) is a technical estimate, which will be needed twice below from diļ¬€erent perspectives. It will be used as an upper bound for the L2 -norm of f  in the proof of a Combesā€“Thomas estimate. It will also be used as a lower d2 bound on f, Hf , once a self-adjoint choice of H = āˆ’ dx 2 +V has been ļ¬xed, and this will imply a lower bound on the spectrum. The interpretation as a lower bound on f, Hf  is particularly intuitive: if V is bounded below, d2 even in the L1 sense considered here, the self-adjoint expression āˆ’ dx 2 + V accepts a lower bound. Part (d) has been stated for an internal point c āˆˆ I, but it also holds for c = āˆ’ if that is a regular endpoint. More notably, (d) can be stated more elegantly when both endpoints are inļ¬nite and obey (11.5) (Exercise 11.5).

386

11. One-dimensional SchrĀØodinger operators

Part (c) is another conclusion which was noted for future use. It will help us to conclude that many SchrĀØ odinger operators have domains D(H) āŠ‚ Lāˆž (I), which will be an important technical ingredient in the proof of Schnolā€™s theorem.

11.5. Self-adjointness and separated boundary conditions 2

d We are now ready to turn the diļ¬€erential expression H = āˆ’ dx 2 + V into a self-adjoint operator in the general setting where I is an interval on R and V āˆˆ L1loc (I). We begin by deļ¬ning the maximal operator Hmax on L2 (I) by

D(Hmax ) = {f āˆˆ Dloc | f, āˆ’f  + V f āˆˆ L2 (I)} and Hmax f = āˆ’f  + V f. Equation (11.50) can be written as Hmax f, g āˆ’ f, Hmax g = W+ (fĀÆ, g) āˆ’ Wāˆ’ (fĀÆ, g),

(11.60)

which can be interpreted as an obstruction to self-adjointness. Indeed, we āˆ— āŠ‚ Hmax and that, if both endpoints are a limit point, will prove that Hmax Hmax is self-adjoint. Otherwise, we will construct self-adjoint restrictions of Hmax by separately restricting the domain at each limit circle endpoint. In order to separate the contributions from diļ¬€erent endpoints, we write the domain as (11.61) D(Hmax ) = Xāˆ’ āˆ© X+ , āˆ— deļ¬ned by (11.52). Recall that we denote and we will use the subspaces XĀ± by L2c (I) the set of compactly supported functions in L2 (I),

L2c (I) = {f āˆˆ L2 (I) | f Ļ‡[c,d] = f for some compact [c, d] āŠ‚ I}. Theorem 11.34. The restriction H0 of Hmax to D(H0 ) = Dloc āˆ© L2c (I) obeys the following. (a) H0 is densely deļ¬ned. (b) H0āˆ— = Hmax . āˆ— āˆ© Xāˆ— . (c) H0 is the restriction of Hmax to Xāˆ’ +

Proof. To prove that H0 is densely deļ¬ned and to ļ¬nd its adjoint, let us assume that u, v āˆˆ L2 (I) obey u, H0 f  = v, f 

āˆ€f āˆˆ D(H0 ).

(11.62)

11.5. Self-adjointness and separated boundary conditions

387

Let us temporarily ļ¬x a compact interval [c, d] āŠ‚ I. Consider any h āˆˆ L2 ([c, d]) which obeys  d gh dx = 0 (11.63) c

for all solutions of of

āˆ’g  + V

g = 0. For such h, by Lemma 11.18, the solution

āˆ’f  + V f = h,

f (c) = f  (c) = 0

obeys f (d) = f  (d) = 0. Therefore, f = 0 on I \ [c, d] so f āˆˆ D(H0 ). For such f , (11.62) becomes  d  d  u ĀÆ(āˆ’f + V f ) dx = vĀÆf dx. c

c

as any solution of āˆ’w + V w = v, and use (11.51) Introduce w āˆˆ to rewrite as  d  d  d u(āˆ’f  + V f ) dx = (āˆ’w + V w)f dx = w(āˆ’f  + V f ) dx. AC2loc (I)

c

c

c

āˆ’f  +V

f can be an arbitrary function in L2 ([c, d]) which Recalling that h = is orthogonal to all solutions of āˆ’g  + V g = 0, we have proved that  d (u āˆ’ w)h dx = 0 c

for every h which obeys (11.63) for all solutions of āˆ’g  + V g = 0. By Lemma 11.18, it follows that in L2 ([c, d]), u āˆ’ w āˆˆ {g āˆˆ AC2 ([c, d]) | āˆ’g  + V g = 0}āŠ„āŠ„ . Since any ļ¬nite-dimensional subspace is closed, this becomes u āˆ’ w āˆˆ {g āˆˆ AC2 ([c, d]) | āˆ’g  + V g = 0}. It follows that u āˆˆ AC2 ([c, d]) and āˆ’u + V u = āˆ’w + V w = v on [c, d]. Since the interval [c, d] āŠ‚ I was arbitrary, this implies that u āˆˆ AC2loc (I) and āˆ’u + V u = v on I. Since u, v āˆˆ L2 (I), we have shown that (11.62) implies (u, v) āˆˆ Ī“(Hmax ). Conversely, for any (u, v) āˆˆ Ī“(Hmax ), āˆ— āˆ© X āˆ— . This (11.62) holds because boundary Wronskians vanish for f āˆˆ Xāˆ’ + āˆ— implies that H0 is densely deļ¬ned and H0 = Hmax . āˆ— ) if and only if It now follows from (11.50) that g āˆˆ D(Hmax

W+ (g, f ) āˆ’ Wāˆ’ (g, f ) = 0

āˆ€f āˆˆ D(Hmax ).

388

11. One-dimensional SchrĀØodinger operators

āˆ— āˆ© X āˆ— , it is obvious that for all g āˆˆ X āˆ© X , W (g, f ) = 0, If f āˆˆ Xāˆ’ āˆ’ + Ā± + āˆ— ). Conversely, let f āˆˆ D(H āˆ— ). For any h āˆˆ X , use the so f āˆˆ D(Hmax + max function g āˆˆ Xāˆ’ āˆ© X+ from Lemma 11.27; then

W+ (h, f ) = W+ (g, f ) = W+ (g, f ) āˆ’ Wāˆ’ (g, f ) = 0. āˆ—. X+

Analogously, f āˆˆ Therefore, f āˆˆ āˆ— āˆ© Xāˆ— . Hmax to the domain Xāˆ’ +

āˆ—. Xāˆ’

Thus,

āˆ— Hmax

(11.64)

is the restriction of 

āˆ— = X , so In particular, if V is limit point at both endpoints, then XĀ± Ā± H0 = Hmax is self-adjoint. Otherwise, we will look for self-adjoint extensions H of H0 , which must obey āˆ— H0 = Hmax āŠ‚ H āˆ— = H āŠ‚ Hmax .

We will extensively use Section 8.7. By the results of that section, selfadjoint extensions of H0 correspond to Lagrangian subspaces for the skewsymmetric sesquilinear form (11.60). The standard procedure is to pass to the quotient vector space āˆ— āˆ— āˆ© X+ ), (Xāˆ’ āˆ© X+ )/(Xāˆ’

which turns W+ āˆ’ Wāˆ’ into a symplectic form. In our current setting, by Lemma 11.27, this decomposes into a sum of symplectic forms induced by āˆ— and X /X āˆ— . Thus, we will be especially interested in selfWĀ± on Xāˆ’ /Xāˆ’ + + adjoint restrictions H with separated boundary conditions: Deļ¬nition 11.35. A SchrĀØodinger operator on L2 (I) with separated boundary conditions is a SchrĀØodinger operator with domain D(H) = Yāˆ’ āˆ© Y+ , where YĀ± are Lagrangian subspaces of XĀ± with respect to Wronskians WĀ± . The Lagrangian property of YĀ± is explicitly written as g , f ) = 0 āˆ€g āˆˆ YĀ± }. YĀ± = {f āˆˆ XĀ± | WĀ± (ĀÆ

(11.65)

Starting with the easy case, if the endpoint Ā± is a limit point, then XĀ± = āˆ— , so XĀ± YĀ± = XĀ± . Informally speaking, at a limit point endpoint, we do not impose any boundary conditions. āˆ— ) = 2, so subspaces If the endpoint Ā± is a limit circle, then dim(XĀ± /XĀ± obeying (11.65) are one-dimensional subspaces generated by a suitable vector:

Lemma 11.36. At a limit circle endpoint Ā± , Lagrangian subspaces are subspaces of the form YĀ± = {f āˆˆ XĀ± | WĀ± (v, f ) = 0} āˆ— such that W (ĀÆ for some vector v āˆˆ XĀ± \ XĀ± Ā± v , v) = 0.

(11.66)

11.5. Self-adjointness and separated boundary conditions

389

āˆ— ) = 2, Y /X āˆ— must be one dimensional. Thus, it Proof. Since dim(XĀ± /XĀ± Ā± Ā± āˆ— . In other words, must be generated by some nontrivial vector [v] āˆˆ XĀ± /XĀ± āˆ— āˆ— YĀ± = span{v} + XĀ± , where v āˆˆ XĀ± \ XĀ± .

The symplectic complement of YĀ± with respect to WĀ± is then YĀ±āŠ„ = {f āˆˆ XĀ± | WĀ± (v, f ) = 0}. Thus, YĀ± = YĀ±āŠ„ if and only if WĀ± (v, v) = 0.  Deļ¬nition 11.37. Let Ā± be a limit circle endpoint. For any choice of āˆ— such that W (v, v) = 0, we will call the equation v āˆˆ XĀ± \ XĀ± Ā± WĀ± (v, f ) = 0

(11.67)

a self-adjoint boundary condition at Ā± . Let us also note that self-adjoint boundary conditions respect an expected complex conjugation symmetry: Lemma 11.38. For any self-adjoint boundary condition at Ā± , f āˆˆ YĀ± if and only if f āˆˆ YĀ± . āˆ— . TrivProof. By the complex conjugation symmetry of WĀ± , v āˆˆ XĀ± \ XĀ± ially, WĀ± (v, v) = 0, so v āˆˆ YĀ± ; thus, v can be used instead of v to characterize the Lagrangian subspace as

YĀ± = {f āˆˆ XĀ± | WĀ± (v, f ) = 0}. Using again the symmetry of WĀ± , we rewrite this as YĀ± = {f āˆˆ XĀ± | WĀ± (v, f ) = 0}, and comparing this with (11.66), we see f āˆˆ YĀ± if and only if f āˆˆ YĀ± .



Corollary 11.39. For any SchrĀØ odinger operator H with separated, selfadjoint boundary conditions, f āˆˆ D(H) if and only if f āˆˆ D(H). Self-adjoint boundary conditions can be written more concretely if the endpoint behavior of functions in XĀ± is well understood. Most notably: Proposition 11.40. Let Ā± be a regular endpoint. Every self-adjoint boundary condition at Ā± is of the form cos Ļ†f (Ā± ) + sin Ļ†f  (Ā± ) = 0

(11.68)

for some Ļ† āˆˆ R. Proof. At the regular endpoint Ā± , functions f āˆˆ D(Hmax ) have continuous boundary values f (Ā± ) and f  (Ā± ), so in the notation of Proposition 11.31, the Wronskian at Ā± can be evaluated as !    0 āˆ’1 (T f ). WĀ± (g, f ) = g(Ā± )f (Ā± ) āˆ’ g (Ā± )f (Ā± ) = (T g) 1 0

390

11. One-dimensional SchrĀØodinger operators

āˆ— is equivalent to T v = Thus, the condition v āˆˆ / XĀ± !  0 āˆ’1 (T v) = 0 (T vĀÆ) 1 0

0 0 , and the condition

is equivalent to v(Ā± )v  (Ā± ) āˆˆ R. Together, they are equivalent to the existence of Īŗ, Ļ† āˆˆ R such that v(Ā± ) = eiĪŗ sin Ļ† and v  (Ā± ) = eiĪŗ cos Ļ†. Accordingly, (11.67) is equivalent to (11.68). 

11.6. Weyl solutions and Greenā€™s functions Let H be a SchrĀØodinger operator with separated boundary conditions, with the domain D(H) = Yāˆ’ āˆ© Y+ , as introduced in the previous section. We will introduce the corresponding Weyl solutions and use them to describe the resolvents (H āˆ’ z)āˆ’1 for z āˆˆ C \ Ļƒ(H). Informally, as inverses of diļ¬€erential operators, it will not be surprising that resolvents are integral operators; unlike the special case of two regular endpoints in Section 11.3, their integral kernels will often not be in L2 (I Ɨ I), but the integral representations will nonetheless be convergent. Deļ¬nition 11.41. A Weyl solution at z āˆˆ C at the endpoint Ā± is a nontrivial solution of āˆ’Ļˆ  + V Ļˆ = zĻˆ such that Ļˆ āˆˆ YĀ± . We will denote Weyl solutions at z by ĻˆzĀ± (x) or Ļˆ Ā± (x, z). If V is limit point at the endpoint Ā± , then YĀ± = XĀ± , so Weyl solutions can be characterized simply as nontrivial solutions of āˆ’Ļˆ  + V Ļˆ = zĻˆ which are square-integrable in a neighborhood of Ā± . However, if V is a limit circle at the endpoint Ā± , the Weyl solution depends not only on V , but also on the boundary condition at Ā± through the requirement Ļˆ āˆˆ YĀ± . This is consistent with the usage in Section 11.3. Theorem 11.42. Consider a self-adjoint SchrĀØ odinger operator H with separated boundary conditions, D(H) = Yāˆ’ āˆ© Y+ , and z āˆˆ C \ Ļƒess (H). (a) At each endpoint Ā± , there exist Weyl solutions ĻˆzĀ± āˆˆ YĀ± . The set of Weyl solutions, together with the trivial solution, is one dimensional. Moreover, WĀ± (ĻˆzĀ± , ĻˆzĀ± ) = 0. (b) z is an eigenvalue of H if and only if W (Ļˆz+ , Ļˆzāˆ’ ) = 0. Proof. Fix the Ā± sign and assume that f1 , f2 are Weyl solutions at the endpoint Ā± . From f1 , f2 āˆˆ YĀ± , it follows that WĀ± (f1 , f2 ) = 0. Since f1 , f2 solve the same ordinary diļ¬€erential equation āˆ’f  + V f = zf , their Wronskian W (f1 , f2 ) = f1 f2 āˆ’f1 f2 is independent of x. Thus, it is constantly

11.6. Weyl solutions and Greenā€™s functions

391

zero, so f1 , f2 are linearly dependent. Thus, the set of Weyl solutions at each endpoint is at most one dimensional. By Lemma 11.38, f āˆˆ YĀ± implies f āˆˆ YĀ± , so WĀ± (f , f ) = 0. If z is in the discrete spectrum, it is an eigenvalue, so the corresponding eigenvector f āˆˆ Ker(H āˆ’z)\{0} is a Weyl solution at both endpoints. Thus, it remains to consider z āˆˆ C \ Ļƒ(H). Fix [c, d] āŠ‚ I. For any g āˆˆ L2 (I) with supp g āŠ‚ [c, d], consider f = (H āˆ’ z)āˆ’1 g āˆˆ Yāˆ’ āˆ© Y+ and evaluate at c and d to deļ¬ne linear maps Tc , Td : L2 ([c, d]) ā†’ C2 , ! ! f  (c) f  (d) , Td g = . Tc g = f (c) f (d) Any nontrivial value of Tc g āˆˆ C2 corresponds to f āˆˆ Yāˆ’ , which is nontrivial on (āˆ’ , c) and obeys āˆ’f  + V f = zf on (āˆ’ , c). In other words, on the interval (āˆ’ , c), f is an eigensolution, and extending that eigensolution to I gives a Weyl solution at āˆ’ . Analogously, any nontrivial value of Td g āˆˆ C2 leads to a Weyl solution at + . The set of Weyl solutions at each endpoint is at most one dimensional, so dim Ran Tc ā‰¤ 1 and dim Ran Td ā‰¤ 1. Our remaining goal is to show that dim Ran Tc = dim Ran Td = 1. If Ran Td = {0}, this would imply that for all g āˆˆ L2 ([c, d]), the solution of the initial value problem āˆ’f  + (V āˆ’ z)f = g,

f (d) = f  (d) = 0,

has values (f (c), f  (c)) āˆˆ Ran Tc . This leads to a contradiction since by Lemma 11.18, by varying g we can produce an arbitrary (f (c), f  (c)) āˆˆ C2 . Analogously, Ran Tc = {0} would lead to a contradiction. If the Weyl solutions Ļˆzāˆ’ and Ļˆz+ were linearly dependent, they would both be in Yāˆ’ āˆ©Y+ , so they would be in Ker(Hāˆ’z), contradicting invertibility of H āˆ’ z.  In particular, for z āˆˆ C \ Ļƒ(H), there exist Weyl solutions ĻˆzĀ± at Ā± and their Wronskian is nonzero, so we can deļ¬ne Greenā€™s function G(x, y; z) =

1 Ļˆzāˆ’ (min(x, y))Ļˆz+ (max(x, y)). + W (Ļˆz , Ļˆzāˆ’ )

(11.69)

Note that this deļ¬nition is independent of the normalization of ĻˆzĀ± . We will prove that this is the integral kernel of the resolvent. This is often formally written using Dirac delta functions as G(x, y; z) = Ī“x , (H āˆ’ z)āˆ’1 Ī“y 

392

11. One-dimensional SchrĀØodinger operators

or as (H āˆ’ z)G(Ā·, y; z) = Ī“y . We mention this only for motivation; we will not formally use distributions in the proofs. We ļ¬rst collect simple properties of Greenā€™s function into the following lemma. Let us denote as before the Heaviside function by 1 h(t) = (1 + sgn t). 2 Lemma 11.43. For any z āˆˆ C \ Ļƒ(H), the Greenā€™s function G(x, y; z) has the following properties: (a) G(x, y; z) = G(y, x; z) for all x, y āˆˆ I.  (b) For any y āˆˆ I, |G(x, y; z)|2 dx < āˆž. (c) For any y āˆˆ I, as functions of x, G(x, y; z) āˆˆ ACloc (I), āˆ‚x G(x, y; z) + h(x āˆ’ y) āˆˆ ACloc (I), and āˆ‚x (āˆ‚x G(x, y; z) + h(x āˆ’ y)) = (V (x) āˆ’ z)G(x, y; z). (d) The map y ā†’ G(x, y; z) is continuous as a function from I to L2 (I). Proof. (a) This follows immediately from (11.69). (b) This follows from the fact that ĻˆĀ± are square-integrable near Ā± , respectively. (c) This is a calculation as in the proof of Proposition 11.21. (d) For y1 , y2 āˆˆ I, assuming without loss of generality that y1 < y2 ,  |G(x, y1 ; z) āˆ’ G(x, y2 ; z)|2 dx  y1 1 |Ļˆzāˆ’ (x)|2 |Ļˆz+ (y1 ) āˆ’ Ļˆz+ (y2 )|2 dx = |W (Ļˆz+ , Ļˆzāˆ’ )|2 āˆ’  y2 1 |Ļˆzāˆ’ (y1 )Ļˆz+ (x) āˆ’ Ļˆzāˆ’ (x)Ļˆz+ (y2 )|2 dx + |W (Ļˆz+ , Ļˆzāˆ’ )|2 y1  + 1 |Ļˆz+ (x)|2 |Ļˆzāˆ’ (y1 ) āˆ’ Ļˆzāˆ’ (y2 )|2 dx. + |W (Ļˆz+ , Ļˆzāˆ’ )|2 y2 As y1 ā†‘ y2 or y2 ā†“ y1 , this converges to 0 by the square-integrability of ĻˆzĀ± at Ā± and their continuity on I. That shows that as an L2 (I)-valued function of y, G is left- and right-continuous, so it is continuous. 

11.6. Weyl solutions and Greenā€™s functions

393

Theorem 11.44. Let H be a SchrĀØ odinger operator with separated boundary conditions. For any z āˆˆ C \ Ļƒ(H) and g āˆˆ L2 (I), the value (H āˆ’ z)āˆ’1 g is given pointwise by  āˆ’1 ((H āˆ’ z) g)(x) = G(x, y; z)g(y) dy. (11.70) Proof. We begin by proving (11.70) for compactly supported g. Assume that supp g āŠ‚ [c, d] āŠ‚ I and denote the right-hand side of (11.70) by f . Compact support of g allows us to use Fubiniā€™s theorem as in the proof of Proposition 11.21 to conclude f āˆˆ AC2loc (I) and āˆ’f  + (V āˆ’ z)f = g. It remains to prove that f āˆˆ Yāˆ’ āˆ© Y+ . For this, note that for x > d,  d 1 āˆ’ + f (x) = + āˆ’ Ļˆz (y)Ļˆz (x)g(y) dy, c W (Ļˆz , Ļˆz ) which is a ļ¬xed multiple of Ļˆz+ (x), so f āˆˆ Y+ because Ļˆz+ āˆˆ Y+ . Similarly, in a neighborhood of āˆ’ , f is found to be a multiple of Ļˆzāˆ’ , so f āˆˆ D(H) and (H āˆ’ z)f = g. Now let g āˆˆ L2 (I). By the above, (11.70) holds for the functions gĻ‡[c,d] . In the double limit c ā†“ āˆ’ , d ā†‘ + , the left-hand side of (11.70) converges in the L2 (I)-sense. Meanwhile, the right-hand side of (11.70) converges for each x, because for any ļ¬xed x āˆˆ I, g āˆˆ L2 (I) and  |G(x, y; z)|2 dy < āˆž. I

Moreover, G(x, Ā·; z) is a multiple of Ļˆzāˆ’ (y) for y < x and is a multiple of Ļˆz+ (y) for y > x. If a sequence of functions converges both in the L2 (I) sense and pointwise, then the two limits are equal almost everywhere by Corollary 2.31, which shows that (11.70) holds for any g āˆˆ L2 (I) as equality of functions in L2 (I). However, both sides of (11.70) are continuous in x: the left-hand side because (H āˆ’z)āˆ’1 g āˆˆ D(H) and D(H) consists of continuous functions; and the right-hand side because y ā†’ G(x, y; z) is a continuous map from I  to L2 (I) by Lemma 11.43(d). This notion of Weyl solution generalizes (and shares notation with) the solutions ĻˆzĀ± from Section 11.3, but note a subtle diļ¬€erence between Theorem 11.44 and Proposition 11.21: In Theorem 11.44, z āˆˆ / Ļƒ(H) is an assumption, rather than a conclusion. This weakening is necessary because in the general setting, even for z āˆˆ Ļƒess (H), Weyl solutions may exist and their Wronskian may be nonzero; however, in such cases, (11.69) does not have the same interpretation as the integral kernel of the resolvent. Due to the loss of this key interpretation, the term ā€œWeyl solutionsā€ is usually only used for z āˆˆ / Ļƒess (H).

394

11. One-dimensional SchrĀØodinger operators

The explicit form of Greenā€™s function allows us to describe certain relatively compact operators relevant for the RAGE theorem (Theorem 9.23): Lemma 11.45. For any compact [c, d] āŠ‚ I, the projection P f = Ļ‡[c,d] f on L2 (I) is relatively compact with respect to H, i.e., the operator P (H āˆ’ i)āˆ’1 is compact. Proof. The operator K = P (H āˆ’ i)āˆ’1 is an integral operator with kernel K(x, y) = Ļ‡[c,d] (x)G(x, y; i). By symmetry, Lemma 11.43 gives L2 (I)-continuity of Greenā€™s function in x, so by continuity and compactness,  |G(x, y; i)|2 dy < āˆž. sup xāˆˆ[c,d] I

It follows that  





|K(x, y)|2 dx dy = I

I

|G(x, y; i)|2 dy dx < āˆž, [c,d]

I

so K is compact (Proposition 4.51).



 Thus, for any increasing sequence of intervals [cn , dn ] with [cn , dn ] = I, the analysis of Section 9.4 applies to projections Pn f = Ļ‡[cn ,dn ] f . These results have direct physical interpretations in quantum mechanics where, for instance, Pn eāˆ’itH f 2 corresponds to the probability of ļ¬nding the particle in the region [cn , dn ] at time t. In particular, RAGE Theorem 9.23 describes the dynamics of vectors in the pure point and continuous subspaces for H, and Exercise 9.8 describes a property of vectors in the absolutely continuous subspace for H.

11.7. Weyl solutions and m-functions In this section we focus on the half-line case characterized by one regular endpoint, and we change the notation a bit. We write the interval as I = (0, b), and assume that 0 is a regular endpoint with the boundary condition at 0, cos Ī±f (0) + sin Ī±f  (0) = 0.

(11.71)

The following discussion is most commonly used in the case when b = +āˆž and the potential is the limit point at b. However, the endpoint b can be ļ¬nite or inļ¬nite; it can even be a regular endpoint. We ļ¬x the behavior at b by ļ¬xing a Lagrangian subspace Y+ āŠ‚ X+ . As discussed before, this incorporates a self-adjoint boundary condition at b if H is the limit circle at b.

11.7. Weyl solutions and m-functions

395

Recall that Ļ†(x, z) = Ļ†z (x) and Īø(x, z) = Īøz (x) are solutions of the ordinary diļ¬€erential equation āˆ’f  + V f = zf, satisfying the initial conditions

! Ļ†z (0) Īøz (0) = Ļ†z (0) Īøz (0)

cos Ī± āˆ’ sin Ī± sin Ī± cos Ī±

!āˆ’1 .

(11.72)

The solutions Īøz , Ļ†z are Ī±-dependent in order to match the operator, in particular, Ļ†z obeys the boundary condition at 0, so Ļ†z is a Weyl solution at the regular endpoint 0. The solutions also obey the useful relations W (Īøz , Īøz ) = 0,

W (Īøz , Ļ†z ) = 1,

W (Ļ†z , Īøz ) = āˆ’1,

W (Ļ†z , Ļ†z ) = 0 (11.73) obtained by evaluating those Wronskians at 0. Further properties of Īøz , Ļ†z on [0, c] for any c āˆˆ (0, b) follow by an aļ¬ƒne transformation from earlier results, in particular, by Corollary 11.8: Corollary 11.46. For any c āˆˆ (0, b), Īøz and Ļ†z are entire AC2 ([0, c])-valued functions of z. For z āˆˆ C \ Ļƒess (H), let us denote simply by Ļˆz a Weyl solution for H at b. Since Ļ†z and Ļˆz are Weyl solutions at 0 and b, respectively, their Wronskian is nonzero for all z āˆˆ / Ļƒ(H). Thus, we can deļ¬ne: Deļ¬nition 11.47. The Weyl m-function associated to H is the map m : C \ Ļƒ(H) ā†’ C deļ¬ned by m(z) = āˆ’

W (Ļˆz , Īøz ) . W (Ļˆz , Ļ†z )

(11.74)

The Wronskians in this deļ¬nition are independent of x; evaluating them at 0 gives cos Ī±Ļˆz (0) āˆ’ sin Ī±Ļˆz (0) , (11.75) m(z) = sin Ī±Ļˆz (0) + cos Ī±Ļˆz (0) which can be written in the notation of MĀØobius transformations as ! ! ! m(z) cos Ī± āˆ’ sin Ī± Ļˆz (0) . $ Ļˆz (0) 1 sin Ī± cos Ī± However, the seemingly more implicit representation (11.74) in terms of Wronskians is often more convenient. We will now derive various properties of Weyl solutions and m-functions; the key property of Weyl solutions is the following. Lemma 11.48. For all z, w āˆˆ C \ Ļƒess (H), Weyl solutions obey  b Ļˆw (x)Ļˆz (x) dx = Wāˆ’ (Ļˆw , Ļˆz ). (z āˆ’ w) 0

(11.76)

396

11. One-dimensional SchrĀØodinger operators

Proof. Since Weyl solutions are in the maximal domain and Hmax Ļˆz = zĻˆz , the standard formula (11.60) gives (w āˆ’ z)Ļˆw , Ļˆz  = wĻˆw , Ļˆz  āˆ’ Ļˆw , zĻˆz  = W+ (Ļˆ w , Ļˆz ) āˆ’ Wāˆ’ (Ļˆ w , Ļˆz ). Moreover, Weyl solutions obey the boundary condition (if any) at b, i.e.,  Ļˆz , Ļˆw āˆˆ Y+ , so W+ (Ļˆw , Ļˆz ) = 0, and the claim follows. Until now, the formulas were independent of the normalization of Ļˆz , but for the remainder of this section, it will be convenient to ļ¬x the normalization W (Ļˆz , Ļ†z ) = 1

āˆ€z āˆˆ C \ Ļƒ(H).

(11.77)

This is possible because Ļ†z is a Weyl solution at 0, so it is linearly independent with Ļˆz . Lemma 11.49. With the normalization (11.77), m(z) = āˆ’W (Ļˆz , Īøz ) and Ļˆz = Īøz + m(z)Ļ†z .

(11.78)

Proof. m(z) = āˆ’W (Ļˆz , Īøz ) follows immediately from (11.74). To prove (11.77), begin by writing Ļˆz as a linear combination, Ļˆz = aĪøz +bĻ†z . Viewing the Wronskians W (Ā·, Īøz ) and W (Ā·, Ļ†z ) as nontrivial linear functionals on the two-dimensional space of eigensolutions of H, (11.73) allows us to compute W (Ļˆz , Īøz ) = aW (Īøz , Īøz ) + bW (Ļ†z , Īøz ) = āˆ’b and similarly W (Ļˆz , Ļ†z ) = a. Since we know the Wronskians, this gives us the values of a = 1 and b = m(z).  Corollary 11.50. For z, w āˆˆ C\Ļƒ(H), if the Weyl solutions are normalized by (11.77), they obey  b m(z) āˆ’ m(w) . (11.79) Ļˆw (x)Ļˆz (x) dx = zāˆ’w 0 Proof. This is immediate from Lemma 11.48 if we prove that our normalization gives Wāˆ’ (Ļˆw , Ļˆz ) = m(z)āˆ’m(w). This can be obtained by a brute force calculation using (11.72) or by noting that the reality and z-independence of the initial conditions in (11.72) imply that, similarly to (11.73), Wāˆ’ (Īøw , Īøz ) = 0,

Wāˆ’ (Īøw , Ļ†z ) = 1,

Wāˆ’ (Ļ†w , Īøz ) = āˆ’1,

Wāˆ’ (Ļ†w , Ļ†z ) = 0,

and then using bilinearity of the Wronskian to expand and compute Wāˆ’ (Ļˆ w , Ļˆz ) = Wāˆ’ (Īøw + m(w)Ļ†w , Īøz + m(z)Ļ†z ).



11.7. Weyl solutions and m-functions

397

Theorem 11.51. The function m(z) is analytic on C \ Ļƒ(H) and it obeys sgn Im m(z) = sgn Im z. In particular, it is a Herglotz function, and for z āˆˆ C \ R, if the Weyl solutions are normalized by (11.77), then  b Im m(z) = |Ļˆz |2 dx. (11.80) Im z 0 Proof. By (11.79) applied to z āˆˆ C \ R and w = z, we obtain (11.80). In particular, z āˆˆ C+ implies m(z) āˆˆ C+ . We now note a symmetry in our eigensolutions. Since Īøz , Ļ†z are deļ¬ned with real initial conditions, they obey the symmetry ĪøĀÆz = ĪøzĀÆ, Ļ†ĀÆz = Ļ†zĀÆ. Note also that Ļˆz is an eigensolution at zĀÆ which is in Y+ since Ļˆz is; thus, Ļˆz is a Weyl solution at zĀÆ. It also follows from (11.77) that W (ĻˆĀÆz , Ļ†ĀÆz ) = W (ĻˆĀÆz , Ļ†zĀÆ) = 1, so ĻˆĀÆz obeys the correct normalization and therefore ĻˆĀÆz = ĻˆzĀÆ. Finally, this implies by (11.78) that m(z) = m(ĀÆ z)

āˆ€z āˆˆ C \ Ļƒ(H).

It remains to prove that m(z) is analytic on C \ Ļƒ(H). Using the Weyl solutions Ļ†z , Ļˆz at 0, b, Greenā€™s function for H is given by G(x, y; z) = Ļ†z (min(x, y))Ļˆz (max(x, y)) = Ļ†z (min(x, y))Īøz (max(x, y)) + m(z)Ļ†z (x)Ļ†z (y).

(11.81)

Analyticity of (H āˆ’ z)āˆ’1 implies analyticity, for any f āˆˆ L2 (I), of  āˆ’1 f (x)Ļ†z (min(x, y))Īøz (max(x, y))f (y) dy dx f, (H āˆ’ z) f  =  (11.82) + m(z) f (x)Ļ†z (x)Ļ†z (y)f (y) dy dx. Fix c āˆˆ (0, b). Since Ļ†z , Īøz are analytic AC2 ([0, c])-valued functions dividing cases x < y and x > y and writing the ļ¬rst term on the right-hand side as a sum of two iterated integrals, this shows that it is entire in z for any f āˆˆ L2 (I) with supp f āŠ‚ [0, c]. Analyticity of m(z) will therefore follow from analyticity of everything else in (11.82), if we can choose for every z0 āˆˆ C \ Ļƒ(H) a function f āˆˆ L2 (I) with supp f āŠ‚ [0, c] such that !2   f (x)Ļ†z (x) dx = 0 f (x)Ļ†z (x)Ļ†z (y)f (y) dy dx = holds for z = z0 (and therefore, by continuity, in a neighborhood of z0 ). Since Ļ†z and Ļ†z are jointly continuous in z and x, in a neighborhood of any z āˆˆ C \ R it suļ¬ƒces to choose x0 āˆˆ [0, c) such that Ļ†z0 (x0 ) = 0 and  f = Ļ‡[x0 ,x0 +] for suļ¬ƒciently small > 0.

398

11. One-dimensional SchrĀØodinger operators

We conclude this section by noting some further properties of Weyl solutions and m-functions as functions of z. Theorem 11.52. If Weyl solutions are normalized by (11.77), then they are L2 (I)-continuous on C \ Ļƒ(H), i.e., for any z āˆˆ C \ Ļƒ(H),  b lim |Ļˆw āˆ’ Ļˆz |2 dx = 0. (11.83) wā†’z

0

Proof. Begin by assuming z, w āˆˆ C \ R. Expanding |Ļˆw āˆ’ Ļˆz |2 = |Ļˆz |2 + |Ļˆw |2 āˆ’ 2 Re(Ļˆw Ļˆz ) and using (11.79) to integrate gives  b m(z) āˆ’ m(w) m(z) āˆ’ m(z) m(w) āˆ’ m(w) + āˆ’ 2 Re . |Ļˆw āˆ’ Ļˆz |2 dx = zāˆ’z wāˆ’w zāˆ’w 0 (11.84) m(z) m(z)āˆ’m(z) = 0, which already As w ā†’ z, this converges to 2 ImIm z āˆ’ 2 Re zāˆ’z proves (11.83) for z āˆˆ C \ R.

By (11.78), analyticity of the m-function on C\Ļƒ(H) implies AC2 ([0, c])analyticity of the Weyl solutions for any c < b. Thus, by Fatouā€™s lemma, for any Ī» āˆˆ R \ Ļƒ(H) and z āˆˆ C \ R,   |Ļˆz āˆ’ ĻˆĪ» |2 dx ā‰¤ lim inf |Ļˆz āˆ’ Ļˆw |2 dx. wāˆˆC\R wā†’Ī»

Using (11.84), we can compute this lim inf, because m(w) ā†’ m(Ī») and m(w)āˆ’m(w) wāˆ’w ĀÆ

ā†’ m (Ī») by analyticity. Thus, for all z āˆˆ C \ R and Ī» āˆˆ R \ Ļƒ(H),



|Ļˆz āˆ’ ĻˆĪ» |2 dx ā‰¤

m(z) āˆ’ m(z) m(z) āˆ’ m(Ī») + m (Ī») āˆ’ 2 Re . z āˆ’ zĀÆ zāˆ’Ī»

(11.85)

By repeating this trick, if z ā†’ Īŗ for some Īŗ āˆˆ R \ Ļƒ(H), Īŗ = Ī», by Fatouā€™s lemma,  m(Īŗ) āˆ’ m(Ī») |ĻˆĪŗ āˆ’ ĻˆĪ» |2 dx ā‰¤ m (Īŗ) + m (Ī») āˆ’ 2 Re . (11.86) Īŗāˆ’Ī» Taking the limit of (11.85) as z ā†’ Ī» with z āˆˆ C \ R gives  |Ļˆz āˆ’ ĻˆĪ» |2 dx ā‰¤ m (Ī») + m (Ī») āˆ’ 2m (Ī») = 0, lim zāˆˆC\R zā†’Ī»

and taking the limit of (11.86) as Īŗ ā†’ Ī» with Īŗ āˆˆ R \ Ļƒ(H) gives  |ĻˆĪŗ āˆ’ ĻˆĪ» |2 dx ā‰¤ m (Ī») + m (Ī») āˆ’ 2m (Ī») = 0. lim ĪŗāˆˆR\Ļƒ(H) Īŗā†’Ī»

11.8. The half-line eigenfunction expansion

399

Together, these two conclusions show L2 (I)-continuity of Weyl solutions at Ī» āˆˆ R \ Ļƒ(H).  The L2 (I)-continuity of the Weyl solutions can be used to extract additional consequences. We prove one corollary and leave another as Exercise 11.7. Corollary 11.53. If Weyl solutions are normalized by (11.77), then for all z āˆˆ C \ Ļƒ(H),  b Ļˆz (x)2 dx = m (z). (11.87) 0

Proof. By the Cauchyā€“Schwarz inequality,  b   b   2  Ļˆz (x) dx āˆ’ Ļˆw (x)Ļˆz (x) dx ā‰¤ Ļˆz Ļˆz āˆ’ Ļˆw ,  0

so

0

L2 (I)-continuity

of Weyl solutions implies  b  b Ļˆz (x)2 dx = lim Ļˆw (x)Ļˆz (x) dx. 0

wā†’z

0

By the symmetry Ļˆw = Ļˆw , and Corollary 11.50, this limit can be computed as  b  b m(z) āˆ’ m(w) 2 = m (z).  Ļˆz (x) dx = lim Ļˆw (x)Ļˆz (x) dx = lim wā†’z wā†’z z āˆ’ w 0 0 Although we considered the Weyl m-function as a function on C \ Ļƒ(H), one can also consider its singularities at points in Ļƒd (H) and describe qualitatively and quantitatively the simple poles obtained there (Exercise 11.8). Of course, a diļ¬€erent normalization will be needed instead of (11.77), since W (Ļˆz , Ļ†z ) = 0 for z āˆˆ Ļƒd (H).

11.8. The half-line eigenfunction expansion We continue to work under the assumptions and notation of the previous section. In this section, we will construct eigenfunction expansions for SchrĀØ odinger operators with one regular endpoint. The eigenfunction expansion will be an explicit unitary operator, bearing some resemblance to the Fourier transform but based on formal eigenfunctions of H. This unitary operator will conjugate H to the operator of multiplication with respect to a canonical choice of spectral measure. We begin by introducing the measure. The Weyl m-function corresponding to H has a Herglotz representation involving a Baire measure Ī¼ on R.

400

11. One-dimensional SchrĀØodinger operators

In particular, Ī¼ is given by Stieltjes inversion: for all h āˆˆ Cc (R),   1 h(Ī») Im m(Ī» + i ) dĪ» = h(Ī») dĪ¼(Ī»). lim ā†“0 Ļ€ It is not a priori clear that this is related to the spectral properties of H, but we will see below that the resulting measure Ī¼ is a maximal spectral measure for H; in fact, we will consider this the canonical spectral measure for the operator H. The eigenfunction expansion for H will be a unitary map conjugating H to the multiplication operator TĪ»,dĪ¼(Ī») . In particular, this will be a unitary map from L2 (I) to L2 (dĪ¼). We will begin by constructing the eigenfunction expansion and its presumed inverse on dense subsets of the Hilbert spaces (note that it is not immediately obvious that these integral transforms even map into the other Hilbert space). Recall that L2c (I) denotes the set of compactly supported functions in L2 (I). Lemma 11.54. For f āˆˆ L2c (I), the function fĖ† : R ā†’ C deļ¬ned by  fĖ†(Ī») = Ļ†Ī» (x)f (x) dx (11.88) is a continuous function of Ī» āˆˆ R. Proof. If f āˆˆ L2c (I), then f āˆˆ L1 (supp f ) by the Cauchyā€“Schwarz inequality. Since Ļ†Ī» (x) is jointly continuous in Ī», x, the integral  Ė† f (Ī») = f (x)Ļ†Ī» (x) dx is uniformly convergent on compacts and deļ¬nes a continuous function fĖ†.



Lemma 11.55. For g āˆˆ L2c (dĪ¼), the function gĖ‡ : I ā†’ C deļ¬ned by  (11.89) gĖ‡(x) = Ļ†Ī» (x)g(Ī») dĪ¼(Ī») is in AC2 ([0, d]) for every d < b and  gĖ‡ (x) = Ļ†Ī» (x)g(Ī») dĪ¼(Ī»),  gĖ‡ (x) = Ļ†Ī» (x)g(Ī») dĪ¼(Ī»).

(11.90) (11.91)

Proof. Since Ļ†Ī» is uniformly bounded in the AC2 ([0, d]) norm on compact sets of Ī», it follows from Fubiniā€™s theorem that for any x1 < x2 ,   x2  Ļ†Ī» (x)g(Ī») dĪ¼(Ī») dx = (Ļ†Ī» (x2 ) āˆ’ Ļ†Ī» (x1 ))g(Ī») dĪ¼(Ī») = gĖ‡(x2 ) āˆ’ gĖ‡(x1 ), x1

11.8. The half-line eigenfunction expansion

401

which proves gĖ‡ āˆˆ ACloc ([0, b)) and (11.90). Analogously, computing   x2  g  (x1 ), Ļ†Ī» (x)g(Ī») dĪ¼(Ī») dx = (Ļ†Ī» (x2 )āˆ’Ļ†Ī» (x1 ))g(Ī») dĪ¼(Ī») = gĖ‡ (x2 )āˆ’Ė‡ x1

proves gĖ‡ āˆˆ AC2loc ([0, b)) and (11.91).



We can now state precisely the main result of this section: Theorem 11.56 (Half-line eigenfunction expansion). There exists a unitary map U : L2 (I) ā†’ L2 (R, dĪ¼(Ī»)) with the following properties. (a) U f = fĖ† for f āˆˆ L2c (I). (b) U āˆ’1 g = gĖ‡ for g āˆˆ L2c (dĪ¼). (c) U HU āˆ’1 = TĪ»,dĪ¼(Ī») . It will follow immediately from (c) that h(H) = U āˆ’1 Th,dĪ¼ U for any bounded Borel function h. In particular, as a special case when (a) and (b) apply, the theorem implies that >

h(H)f = (hfĖ†)

āˆ€h āˆˆ Cc (R), āˆ€f āˆˆ L2c (I).

(11.92)

However, this logic is backwards, because the ļ¬rst key step in the proof of Theorem 11.56 will be to prove (11.92). This will be Proposition 11.58 below, and it will be proved by using resolvents and Stoneā€™s theorem. This will then allow us to use the abstract eigenfunction expansions of Section 9.9. Accordingly, the ļ¬rst technical ingredient is the behavior of Greenā€™s function for values of z approaching the real line. Recall from (11.81) that Greenā€™s function for H is G(x, y; z) = Ļ†z (min(x, y))Īøz (max(x, y)) + m(z)Ļ†z (x)Ļ†z (y).

(11.93)

We use the concise notation f = oĖœ(g) if f = o(g) pointwise and f = O(g) uniformly in the given parameters. Lemma 11.57. For any d āˆˆ (0, b) and compact interval [Ī»1 , Ī»2 ] āŠ‚ R,

ā†“ 0,

Im G(x, y; Ī» + i ) = Ļ†(x, Ī»)Ļ†(y, Ī») Im m(Ī» + i ) + oĖœ(1), uniformly in (x, y, Ī») āˆˆ (0, d]2 Ɨ [Ī»1 , Ī»2 ]. Proof. By AC2 ([0, d])-analyticity of Ļ†z and Īøz , Ļ†(t, Ī» + i ) = Ļ†(t, Ī») + i (āˆ‚z Ļ†)(t, Ī») + O( 2 ),

ā†“0

uniformly in t āˆˆ [0, d] and Ī» āˆˆ [Ī»1 , Ī»2 ] and analogously for Īøz . Moreover, since m(z) is Herglotz, m(Ī» + i ) = O( āˆ’1 ),

ā†“ 0,

(11.94)

402

11. One-dimensional SchrĀØodinger operators

uniformly in Ī» āˆˆ [Ī»1 , Ī»2 ] by Lemma 7.38. Applying these expansions to (11.93) and using reality of fundamental solutions for real Ī», a short calculation implies Im G(x, y; Ī» + i ) = Ļ†(x, Ī»)Ļ†(y, Ī») Im m(Ī» + i ) + āˆ‚z (Ļ†(x, Ā·)Ļ†(y, Ā·))|z=Ī» Re( m(Ī» + i )) + O( ). Finally, Re( m(Ī»+i )) = oĖœ(1) as ā†“ 0 follows from (11.94) and Lemma 7.37, and this concludes the proof.  We can now prove (11.92): Proposition 11.58. For h āˆˆ Cc (R) and f āˆˆ L2c (I), h(H)f is given by  (h(H)f )(x) =

h(Ī»)Ļ†Ī» (x)Ļ†Ī» (y)f (y) dy dĪ¼(Ī»).

(11.95)

Proof. By Stoneā€™s theorem (Theorem 9.43), for any f āˆˆ L2 (I), 1 h(H)f = lim ā†“0 2Ļ€i



  h(Ī») (H āˆ’ Ī» āˆ’ i )āˆ’1 āˆ’ (H āˆ’ Ī» + i )āˆ’1 f dĪ», (11.96)

where the integral is of a continuous compactly supported L2 (I)-valued function and the limit is taken in L2 (I). For f āˆˆ L2c (I), we will evaluate this limit pointwise. Let d āˆˆ (0, b) be large enough that supp f āŠ‚ [0, d], and let [Ī»1 , Ī»2 ] āŠƒ supp h. Then, as L2 functions of x, 

  h(Ī») (H āˆ’ Ī» āˆ’ i )āˆ’1 āˆ’ (H āˆ’ Ī» + i )āˆ’1 f dĪ»  d  Ī»2 1 = h(Ī») (G(x, y; Ī» + i ) āˆ’ G(x, y; Ī» āˆ’ i )) f (y) dĪ» dy (11.97) 2Ļ€i 0 Ī»1   1 d Ī»2 h(Ī») Im G(x, y; Ī» + i )f (y) dĪ» dy. = Ļ€ 0 Ī»1

1 2Ļ€i

For any kernel K such that K(x, y; Ī», ) = oĖœ(1) as ā†“ 0 uniformly in (x, y, Ī») āˆˆ (0, d]2 Ɨ [Ī»1 , Ī»2 ], dominated convergence implies that for every x,   1 d Ī»2 h(Ī»)K(x, y; Ī», )f (y) dĪ» dy = 0. (11.98) lim ā†“0 Ļ€ 0 Ī»1

11.8. The half-line eigenfunction expansion

403

Thus, by Lemma 11.57 and since h(Ī»)Ļ†Ī» (x)Ļ†Ī» (y) āˆˆ Cc (R) (as a function of Ī»), for every x, the Stieltjes inversion implies that   1 d Ī»2 lim h(Ī») Im G(x, y; Ī» + i )f (y) dĪ» dy (11.99) ā†“0 Ļ€ 0 Ī»1   1 d Ī»2 h(Ī»)Ļ†Ī» (x)Ļ†Ī» (y)f (y) Im m(Ī» + i ) dĪ» dy = lim ā†“0 Ļ€ 0 Ī»1  d  Ī»2 h(Ī»)Ļ†Ī» (x)Ļ†Ī» (y)f (y) dĪ¼(Ī») dy. = 0

Ī»1

Thus, we have computed the limit of (11.97) pointwise; by (11.96), this is  equal to the L2 -limit h(H)f , which concludes the proof. This will allow us to apply the abstract eigenfunction expansion Theorem 9.48: our application is to denote by A = H on H = L2 (I) and to denote by B the operator of multiplication by Ī» on K = L2 (dĪ¼(Ī»)). Theorem 9.48 implies that Ran U and Ker U āˆ— are resolvent-invariant for B. It will then remain to prove that Ker U āˆ— = {0}, and as remarked near Theorem 9.48, this cannot be concluded by abstract arguments. In our setting, Ker U āˆ— = {0} will follow from resolvent-invariance of Ker U āˆ— together with the following lemma: Lemma 11.59. If g āˆˆ L2c (dĪ¼) and gĖ‡ = 0 in L2 (I), then  g(Ī») dĪ¼(Ī») = 0.

(11.100)

Proof. If g āˆˆ L2c (dĪ¼), then gĖ‡ āˆˆ AC2loc ([0, b)), so gĖ‡ = 0 in the L2 sense implies pointwise equalities gĖ‡(x) = gĖ‡ (x) = 0 for all x āˆˆ [0, b). By (11.89) and (11.90),  gĖ‡(0) = āˆ’ sin Ī± g(Ī») dĪ¼(Ī»),   gĖ‡ (0) = cos Ī± g(Ī») dĪ¼(Ī»). Since at least one of sin Ī±, cos Ī± is nonzero and gĖ‡(0) = gĖ‡ (0) = 0, (11.100) follows.  We have now collected all the ingredients for the proof of the half-line eigenfunction expansion: Proof of Theorem 11.56. By Theorem 9.48, the map f ā†’ fĖ† extends to a norm-preserving map U : L2 (I) ā†’ L2 (dĪ¼), with a map U āˆ— : L2 (dĪ¼) ā†’ L2 (I) such that (11.101) U āˆ— g, f  = g, U f 

404

11. One-dimensional SchrĀØodinger operators

for all f āˆˆ L2 (dĪ¼) and g āˆˆ L2 (dĪ¼) and h(H) = U āˆ— Th,dĪ¼ U for all bounded continuous functions h. For f āˆˆ L2c (I) and g āˆˆ L2c (dĪ¼), consider the double integral  f (x)Ļ†Ī» (x)g(Ī») (dx āŠ— dĪ¼(Ī»)). Fubiniā€™s theorem is applicable because f (x)g(Ī») is integrable and compactly supported in (x, Ī») and Ļ†Ī» (x) is bounded on compacts. Thus, we get equality of iterated integrals which simpliļ¬es using the deļ¬nitions of fĖ† and gĖ‡ to   f (x)Ė‡ g (x) dx = fĖ†(Ī»)g(Ī») dĪ¼(Ī»). (11.102) This holds for all g āˆˆ L2c (dĪ¼) and f āˆˆ L2c (I); in particular, it holds for a dense set of f āˆˆ L2 (I). Comparing with (11.101), we see that U āˆ— g = gĖ‡ for all g āˆˆ L2c (dĪ¼). Let us prove that Ker U āˆ— = {0}. Let g āˆˆ Ker U āˆ— . Since Ker U āˆ— is a resolvent-invariant subspace for TĪ»,dĪ¼(Ī») , it follows that Ļ‡(Ī»1 ,Ī»2 ] g āˆˆ Ker U āˆ— for any Ī»1 < Ī»2 . Moreover, Ļ‡(Ī»1 ,Ī»2 ] g āˆˆ L2c (dĪ¼), so by Lemma 11.59,  g(Ī»)Ļ‡(Ī»1 ,Ī»2 ] (Ī») dĪ¼(Ī») = 0 for all Ī»1 < Ī»2 . This implies that g(Ī») = 0 Ī¼-a.e. Thus, Ker U āˆ— = {0}, so by Theorem 9.48(f), U, U āˆ— are mutually inverse unitary maps, h(H) = U āˆ— Th,dĪ¼ U holds for all bounded Borel functions, and H = U āˆ— TĪ»,dĪ¼(Ī» )U .



The eigenfunction expansion provides a multiplication operator representation which is precisely of the form considered abstractly in the spectral theorem for unbounded self-adjoint operators. Thus, Theorem 11.56 allows us to apply abstract spectral theory and obtain several corollaries. The ļ¬rst, immediate corollary of Theorem 11.56 is the following. Corollary 11.60. H has simple (multiplicity 1) spectrum and Ī¼ is a maximal spectral measure for H. In particular, Ļƒ(H) = supp Ī¼ and Ļƒess (H) = ess supp Ī¼. If we emphasize the dependence on the parameter Ī± in the boundary condition at 0 and write HĪ± , mĪ± , Ī¼Ī± for the corresponding Ī±-dependent objects, (11.75) can be written in the notation of MĀØobius transformations as ! ! ! cos(Ī± āˆ’ Ī²) āˆ’ sin(Ī± āˆ’ Ī²) mĪ² (z) mĪ± (z) $ . (11.103) 1 sin(Ī± āˆ’ Ī²) cos(Ī± āˆ’ Ī²) 1

11.8. The half-line eigenfunction expansion

405

Proposition 7.57 has an immediate corollary: Corollary 11.61. The essential spectrum of HĪ± is independent of Ī±. Moreover, on any interval in R \ Ļƒess (HĪ± ), the discrete spectra of HĪ± and HĪ² strictly interlace whenever Ī± āˆ’ Ī² āˆˆ / Ļ€Z. In the special case of two regular endpoints, we can apply this twice to change the boundary condition at each endpoint; thus, the special case of Dirichlet eigenvalues (Corollary 11.25) implies the following. 2

d 1 Corollary 11.62. Consider the operator H = āˆ’ dx 2 + V with V āˆˆ L ([0, 1]) and boundary conditions (11.2) and (11.3). The spectrum of H is bounded from below. Arranging its elements in increasing order, Ļƒ(H) = {Ī»n | n āˆˆ N}, with Ī»n < Ī»n+1 , the eigenvalues obey the asymptotics

Ī»n = n2 Ļ€ 2 + O(n),

n ā†’ āˆž.

(11.104)

Proof. The case Ī± = Ī² = 0 is Corollary 11.25. By changing the boundary conditions twice, Ļƒ(HĪ±,Ī² ) strictly interlaces Ļƒ(HĪ±,0 ), which strictly interlaces Ļƒ(H0,0 ). Since interlacing preserves the property (11.104), the proof is complete.  Returning to the general setting, using pointwise boundary values of Herglotz functions, we can also study the absolutely continuous and singular spectrum. They have very diļ¬€erent dependence on the boundary condition at 0: Proposition 11.63. For Ī± āˆ’ Ī² āˆˆ / Ļ€Z, the absolutely continuous parts of Ī¼Ī± and Ī¼Ī² are mutually absolutely continuous, i.e., [(Ī¼Ī± )ac ] = [(Ī¼Ī² )ac ], and the singular parts are mutually singular, i.e., (Ī¼Ī± )s āŠ„ (Ī¼Ī² )s . Proof. By Proposition 7.47 and Theorem 7.46, the limit lim mĪ± (Ī» + i ) ā†“0

exists for Lebesgue-a.e. Ī» āˆˆ R, and (Ī¼Ī± )ac is mutually absolutely continuous with Ļ‡AĪ± (Ī») dĪ», where AĪ± = {Ī» āˆˆ R | lim mĪ± (Ī» + i ) āˆˆ C+ }. ā†“0

Since (11.103) represents mĪ± in terms of mĪ² by a MĀØobius transformation which preserves C+ , it follows that AĪ± = AĪ² .

406

11. One-dimensional SchrĀØodinger operators

Similarly, the singular part of the measure is supported in the set SĪ± = {Ī» āˆˆ R | lim mĪ± (Ī» + i ) = āˆž} ā†“0

and, using (11.103), this can be written as SĪ± = {Ī» āˆˆ R | lim m0 (Ī» + i ) = āˆ’ cot Ī±}. ā†“0

It follows that SĪ± āˆ© SĪ² = āˆ…, so Ī¼Ī± āŠ„ Ī¼Ī² .



Spectral properties of H, both qualitative and quantitative, can now be studied via the m-function and therefore via the Weyl solutions. For instance, Exercise 11.8 provides a formula for the residue of the m-function at an isolated eigenvalue Ī», which can now also be interpreted as Ī¼({Ī»}). We also note an explicitly computable example (see Exercise 11.9 for a generalization): Example 11.64. On the interval I = (0, +āˆž), the potential V ā‰” 0 is the limit circle at 0 and the limit point at āˆž. If we set the Dirichlet boundary condition at 0, the m-function is āˆš m(z) = āˆ’ āˆ’z āˆš āˆš with the branch of āˆ’z such that Re āˆ’z > 0 on C \ [0, āˆž). The spectrum is Ļƒ(H) = [0, āˆž). The spectrum is purely absolutely continuous and the canonical spectral measure is āˆš 1 (11.105) dĪ¼(Ī») = Ļ‡(0,āˆž) (Ī») Ī» dĪ». Ļ€ āˆš Proof. For z āˆˆ C \ R, let k = āˆ’z, with the branch of square root such that Re k > 0. The equation āˆ’f  = āˆ’k 2 f has linearly independent solutions eĀ±kx . Of those, the square integrable solution is eāˆ’kx , so the Weyl solution is computed by (11.75) with Ī± = 0 as m(z) = āˆ’k. Since Im m(z) āˆš extends continuously to the closed upper half-plane with values Ļ‡(0,āˆž) (Ī») Ī» on the real line, by Proposition 7.43, the spectral measure is precisely (11.105).  The unitary map U in the eigenfunction expansion is uniquely determined as the closure of the densely deļ¬ned map f ā†’ fĖ†. If f is not compactly supported, U f can still be computed by suitable approximations or test functions. We give one useful example. Informally speaking, naively computing the eigenfunction expansion (11.88) of the Dirac delta function Ī“y would give the function Ī“Ė†y (Ī») = Ļ†Ī» (y). Applying (H āˆ’ z)āˆ’1 to Ī“y should correspond to multiplying the eigenfunction expansion by (Ī» āˆ’ z)āˆ’1 , which motivates us to expect that the eigenfunction expansion maps the function Ī» (y) . This is not a rigorous argument, but it G(Ā·, y; z) to the function Ļ†Ī»āˆ’z motivates the correct formula:

11.9. Weyl disks and applications

407

Proposition 11.65. Fix z āˆˆ C \ Ļƒ(H). For any y āˆˆ (0, b), the function Ī» (y) f (x) = G(x, y; z) is mapped by U to the function (U f )(Ī») = Ļ†Ī»āˆ’z . Proof. For any g āˆˆ L2 (dĪ¼),  b āˆ’1 G(x, y; z)(U āˆ’1 g)(x) dx = ((H āˆ’ zĀÆ)āˆ’1 U āˆ’1 g)(y). U f, g = f, U g = 0

; < 2 Since (H āˆ’ zĀÆ)āˆ’1 U āˆ’1 g = U āˆ’1 g(Ī») Ī»āˆ’ĀÆ z , if g āˆˆ Lc (dĪ¼), the right-hand side can be evaluated pointwise as @ A   Ļ†Ī» (y) g(Ī») āˆ’1 g(Ī») U (y) = Ļ†Ī» (y) dĪ¼(Ī») = g(Ī») dĪ¼(Ī»). Ī» āˆ’ zĀÆ Ī» āˆ’ zĀÆ Ī» āˆ’ zĀÆ Thus, the equality 

 (U f )(Ī»)g(Ī») dĪ¼(Ī») =

Ļ†Ī» (y) g(Ī») dĪ¼(Ī») Ī» āˆ’ zĀÆ

holds for all g āˆˆ L2c (dĪ¼), so it follows that (U f )(Ī») =

Ļ†Ī» (y) Ī»āˆ’z .



In particular, note that this proves that   |Ļ†Ī» (y)|2 dĪ¼(Ī») = |G(x, y; z)|2 dx < āˆž. |Ī» āˆ’ z|2 Additional examples are given in Exercise 11.11. Sturm oscillation theory [79, 90] counts eigenvalues below the bottom of the essential spectrum in terms of the number of zeros of eigensolutions. Renormalized oscillation theory [33] counts eigenvalues in gaps of Ļƒess (H) (connected components of R \ Ļƒess (H)) in terms of the number of zeros of Wronskians of eigensolutions.

11.9. Weyl disks and applications In this section we consider another perspective on the limit pointā€“limit circle dichotomy. This is used to generate approximations of the Weyl m-function and compute its asymptotics; already in this section, we will use it to derive the Carmona formula and prove continuity with respect to the potential. As before, we denote our interval by I = (0, b), where b can be ļ¬nite or +āˆž, and assume that the real-valued potential V obeys V āˆˆ L1loc ([0, b)), i.e., L1 ([0, d]) for all d < b. We will study in detail the behavior of eigensolutions for z āˆˆ C+ .

408

11. One-dimensional SchrĀØodinger operators

Lemma 11.66. Let z āˆˆ C+ . For any nontrivial solution of āˆ’f  + V f = zf and any x āˆˆ (0, b),  x |f (t)|2 dt = iW (fĀÆ, f )(x) āˆ’ iW (fĀÆ, f )(0). (11.106) 2 Im z 0

In particular, the function āˆ’iW (fĀÆ, f )(x) = 2 Im(f (x)f  (x))

(11.107)

is a strictly decreasing, real-valued function of x. Proof. As before, starting from the calculation iW (fĀÆ, f ) = 2 Im zf fĀÆ and integrating gives (11.106). Since f is a nontrivial eigensolution, if f (y) = 0 for some y, then f  (y) = 0, so f has only isolated zeros. In particular, ĀÆ f ) = 2 Im zf fĀÆ > 0 away from a discrete set, so the function iW (fĀÆ, f ) iW (f, is strictly increasing.  In terms of the matrix J = we can write

! 0 i , āˆ’i 0

! f  (x) J , āˆ’iW (f , f )(x) = f (x)   (x) as projective which leads to a geometric interpretation: considering ff (x) Ė† as in Section 7.1 (see Example 7.8), we see that coordinates on C āˆ’iW (f , f )(x) ā‰„ 0

f  (x) f (x)

ā‡ā‡’

!āˆ—

f  (x) āˆˆ C+ = C+ āˆŖ R āˆŖ {āˆž} f (x)

with equality corresponding to R āˆŖ {āˆž}. The Weyl disk formalism will take this geometric perspective further, by linking the sign of āˆ’iW (f , f )(x) to the values of f, f  at 0. This will be accomplished by using MĀØ obius transformations corresponding to transfer matrices; we do this while incorporating the boundary condition at 0 with Ī± āˆˆ R. We recall the transfer matrices ! āˆ‚x Ļ†Ī± (x, z) āˆ‚x ĪøĪ± (x, z) , TĪ± (x, z) = RĪ± ĪøĪ± (x, z) Ļ†Ī± (x, z) where Ļ†Ī± (x, z), ĪøĪ± (x, z) are eigensolutions at z with initial conditions at 0 chosen so that TĪ± (0, z) = I. Then an arbitrary eigensolution f corresponds to an arbitrary v āˆˆ C2 by ! f  (x) (11.108) = RĪ±āˆ’1 TĪ± (x, z)v. f (x)

11.9. Weyl disks and applications

409

Deļ¬nition 11.67. For z āˆˆ C+ and x such that V āˆˆ L1 ([0, x]), the Weyl disks DĪ± (x, z) are deļ¬ned by & !āˆ— ! ' Ė† | w TĪ± (x, z)āˆ— J TĪ± (x, z) w ā‰„ 0 . DĪ± (x, z) = w āˆˆ C 1 1 Let us note the promised geometric interpretation of Weyl disks: Lemma 11.68. For any nontrivial eigensolution f at z, cos Ī±f  (0) āˆ’ sin Ī±f (0) āˆˆ DĪ± (x, z) sin Ī±f  (0) + cos Ī±f (0)

ā‡ā‡’

f  (x) āˆˆ C+ . f (x)

The left-hand side is on the boundary of DĪ± (x, z) if and only if the right-hand side is on the boundary of C+ . Proof. Using (11.108) at x = 0 and using the deļ¬nition of the Weyl disk, the left-hand side is equivalent to v āˆ— TĪ± (x, z)āˆ— J TĪ± (x, z)v ā‰„ 0. Substituting J = RĪ± J RĪ±āˆ’1 and using (11.108) again gives v āˆ— J v ā‰„ 0, which is equivalent to the right-hand side. The cases of equality are equivalent.  The Weyl circle āˆ‚DĪ± (x, z) (boundary of the Weyl disk) is naturally parametrized by self-adjoint boundary conditions at x: Example 11.69. For any Ī±, Ī² āˆˆ R, denote by mĪ±,Ī² the Weyl function d2 corresponding to the SchrĀØodinger operator H = āˆ’ dx 2 + V with boundary conditions cos Ī±f (0) + sin Ī±f  (0) = 0, 

cos Ī²f (x) āˆ’ sin Ī²f (x) = 0.

(11.109) (11.110)

For any z āˆˆ C+ , the Weyl circle is parametrized by āˆ‚DĪ± (x, z) = {mĪ±,Ī² (z) | Ī² āˆˆ R}. Proof. As stated above, cos Ī±f  (0) āˆ’ sin Ī±f (0) āˆˆ āˆ‚DĪ± (x, z) sin Ī±f  (0) + cos Ī±f (0) if and only if f  (x)/f (x) āˆˆ R āˆŖ {āˆž}. This is equivalent to f being the Weyl solution for some self-adjoint boundary condition. If f is the Weyl solution for the boundary condition (11.110), then by deļ¬nition, mĪ±,Ī² (z) =

cos Ī±f  (0) āˆ’ sin Ī±f (0) . sin Ī±f  (0) + cos Ī±f (0)



410

11. One-dimensional SchrĀØodinger operators

Statements about arbitrary eigensolutions at z can be turned into statements about transfer matrices at z and then into statements about Weyl disks. Starting from Lemma 11.66, we obtain the following. Lemma 11.70. Fix z āˆˆ C+ . (a) For any x > 0,



āˆ—

x

J āˆ’ TĪ± (x, z) J TĪ± (x, z) = 2 Im z

! 0 0 RĪ±āˆ— TĪ± (t, z) dt. 0 1 (11.111)

āˆ—

TĪ± (t, z) RĪ± 0

(b) For any x1 < x2 , TĪ± (x1 , z)āˆ— J TĪ± (x1 , z) > TĪ± (x2 , z)āˆ— J TĪ± (x2 , z)

(11.112)

in the sense of matrix (operator ) order. (c) The sets DĪ± (x, z) are disks in C+ for all x > 0. (d) The sets DĪ± (x, z) are strictly nested, i.e., DĪ± (x2 , z) āŠ‚ int DĪ± (x1 , z) whenever x1 < x2 . Proof. Multiplying (11.111) from the right by arbitrary v āˆˆ C2 and from the left by v āˆ— reduces to the correct statement (11.106). By the polarization identity, this is suļ¬ƒcient to conclude the matrix identity (11.111). Moreover,  for any v āˆˆ C2 \ { 00 },  x |f (t)|2 dt, v āˆ— (J āˆ’ TĪ± (x, z)āˆ— J TĪ± (x, z))v = 2 Im z 0

which is strictly increasing in x, implying the strict inequality (11.112). Ė† By Lemma 11.70, By general principles, these are generalized disks in C. they are strictly nested. Since T (0, z) = I, a direct calculation shows D(0, z) = C+ , this implies that D(x, z) are Euclidean disks for all x > 0 and  subsets of C+ . The statement J āˆ’TĪ± (x, z)āˆ— J TĪ± (x, z) ā‰„ 0 is J -contractivity of TĪ± (x, z), and the strict inequality (11.112) is strict J -monotonicity of this family of transfer matrices. Now let us consider the limit x ā†’ b. For ļ¬xed z āˆˆ C+ and Ī± āˆˆ R, for the decreasing family of compact disks DĪ± (x, z), the intersection DĪ± (x, z) DĪ± (b, z) := xāˆˆ(0,b)

is a point or a disk (for clarity, let us emphasize that a disk has strictly positive radius). This limiting object also has an interpretation:

11.9. Weyl disks and applications

411

Lemma 11.71. For any nontrivial eigensolution f at z, cos Ī±f  (0) āˆ’ sin Ī±f (0) āˆˆ DĪ± (b, z) sin Ī±f  (0) + cos Ī±f (0)

ā‡ā‡’

f āˆˆ X+ and āˆ’ iW+ (f , f ) ā‰„ 0.

Proof. The left-hand side holds if and only if cos Ī±f  (0) āˆ’ sin Ī±f (0) āˆˆ DĪ± (x, z) sin Ī±f  (0) + cos Ī±f (0) for all x < b, i.e., if and only if āˆ’iW (f , f )(x) ā‰„ 0 for all x < b. Since this function is strictly decreasing, this holds if and only if lim āˆ’iW (f , f )(x) ā‰„ 0.

xā†’b

On the other hand, by (11.106),



b

lim āˆ’iW (f , f )(x) = āˆ’iW (f , f )(0) āˆ’ 2 Im z

xā†’b

|f (t)|2 dt.

0

so f āˆˆ X+ ; moreover, in that Finiteness of this limit implies f āˆˆ  case, the limit can be interpreted as āˆ’iW+ (f , f ). L2 ((0, b)),

Recall that we deļ¬ned the Weyl limit pointā€“limit circle dichotomy by whether the boundary Wronskian on X+ is trivial or not. The following result explains that terminology: Theorem 11.72 (Equivalent characterizations of the limit circle case). For any z āˆˆ C+ , the following are equivalent: (a) The boundary Wronskian W+ is not trivial on X+ . (b) The set of eigensolutions at z which are in L2 ((0, b)) has dimension 2. (c) The intersection DĪ± (b, z) is a disk. Proof. (c) =ā‡’ (b): If the intersection contains two distinct points, then by Lemma 11.71, there are two linearly independent eigensolutions f1 , f2 āˆˆ L2 ((0, b)). (b) =ā‡’ (a): If there are two linearly independent eigensolutions f1 , f2 āˆˆ L2 ((0, b)), they are both in X+ and by linear independence, W+ (f1 , f2 ) = 0. Thus, W+ = 0. (a) =ā‡’ (b): Since W+ is not trivial, there are two distinct Lagrangian subspaces of X+ , denoted Y1 , Y2 , and two distinct self-adjoint SchrĀØodinger operators H1 , H2 with the same V, Ī±. Each of them has a Weyl solution at b, denoted f1 , f2 ; by the resolvent formula, f1 , f2 must be linearly independent. Every eigensolution can be expressed as a linear combination of f1 , f2 , so every eigensolution is in L2 ((0, b)).

412

11. One-dimensional SchrĀØodinger operators

 (b) =ā‡’ (c): Let us ļ¬rst characterize the intersection of disks: w āˆˆ x DĪ± (x, z) if and only if ! ! w āˆ— w āˆ— TĪ± (x, z) J TĪ± (x, z) ā‰„0 1 1 for all x > 0, and by monotonicity in x, this is true if and only if !! ! w w āˆ— āˆ— TĪ± (x, z) J TĪ± (x, z) ā‰„ 0. lim xā†’b 1 1 If all solutions are in L2 ((0, b)), we can compute this limit: the second row RĪ±āˆ’1 TĪ± (x, z) consists of functions in L2 ((0, b)), so the entries of the matrix ! ! 0 0 |Ļ†Ī± (t, z)|2 Ļ†Ī± (t, z)ĪøĪ± (t, z) āˆ— āˆ— RĪ± TĪ± (t, z) = TĪ± (t, z) RĪ± 0 1 ĪøĪ± (t, z)Ļ†Ī± (t, z) |ĪøĪ± (t, z)|2 are in L1 ((0, b)), and integrating this gives a convergent limit !  x |Ļ†Ī± (t, z)|2 Ļ†Ī± (t, z)ĪøĪ± (t, z) āˆ— dt. J āˆ’ lim (TĪ± (x, z) J TĪ± (x, z)) = xā†’b ĪøĪ± (t, z)Ļ†Ī± (t, z) |ĪøĪ± (t, z)|2 0 The limit is self-adjoint and det lim (TĪ± (x, z)āˆ— J TĪ± (x, z)) = lim det(TĪ± (x, z)āˆ— J TĪ± (x, z)) = āˆ’1, xā†’b

xā†’b

so the inequality w 1

!āˆ—

w lim (TĪ± (x, z) J TĪ± (x, z)) xā†’b 1 āˆ—

! ā‰„0

deļ¬nes a disk.



The implication (b) =ā‡’ (c) can also be proved by explicitly computing the radius (Exercise 11.12). Taking the negation of the statements in Theorem 11.72 gives equivalent characterizations of the limit point case; recall that since there is always some self-adjoint operator and a Weyl solution at z, the set of eigensolutions in L2 ((0, b)) always has dimension at least 1. Theorem 11.73 (Equivalent characterizations of the limit point case). For any z āˆˆ C+ , the following are equivalent: (a) The boundary Wronskian W+ is trivial on X+ . (b) The set of eigensolutions at z which are in L2 ((0, b)) has dimension 1. (c) The intersection DĪ± (b, z) is a point. In the limit point case, the sole point in the intersection of Weyl disks is the value of the Weyl m-function:

11.9. Weyl disks and applications

413

Proposition 11.74. If V is limit point at b, then for any z āˆˆ C+ , DĪ± (b, z) = {mĪ± (z)}. Proof. Let f be a Weyl solution at b. Then mĪ± (z) =

cos Ī±f  (0) āˆ’ sin Ī±f (0) . sin Ī±f  (0) + cos Ī±f (0)

The Weyl solution obeys f āˆˆ X+ and W+ (f , f ) = 0, so by Lemma 11.71, mĪ± (z) āˆˆ DĪ± (b, z). Since we are in the limit point case, this concludes the proof.  In the limit circle case, the limit circle is parametrized by self-adjoint boundary conditions at b (Exercise 11.13). Weyl disks provide a powerful tool for proving convergence of Herglotz functions; we present two applications in the limit point case. The ļ¬rst application is a theorem of Carmona, which allows us to study spectral measures through the behavior of eigensolutions with real energies. Theorem 11.75 (Carmona). Assume that V is regular at 0 and is a limit point at b. Fix Ī± āˆˆ R. For any h āˆˆ Cc (R),   h(Ī») (11.113) lim dĪ» = h(Ī») dĪ¼Ī± (Ī»). xā†’b Ļ€(Ļ†Ī± (x, Ī»)2 + Ļ†Ī± (x, Ī»)2 ) Proof. We deļ¬ne functions mĪ± (x, z) by ! ! i mĪ± (x, z) āˆ’1 . $ TĪ± (x, z) RĪ± 1 1

(11.114)

Since i āˆˆ C+ , by the deļ¬nition of Weyl disks, mĪ± (x, z) āˆˆ DĪ± (x, z) for all x. Since V is a limit point at b, it follows that mĪ± (x, z) ā†’ mĪ± (z) for each z āˆˆ C+ . Moreover, mĪ± (x, z) are Herglotz functions, and (11.114) can be written as iāˆ‚x ĪøĪ± (x, z) + ĪøĪ± (x, z) iāˆ‚x Ļ†Ī± (x, z) + Ļ†Ī± (x, z) (iāˆ‚x ĪøĪ± (x, z) + ĪøĪ± (x, z))(āˆ’iāˆ‚x Ļ†Ī± (x, z) + Ļ†Ī± (x, z)) . =āˆ’ |iāˆ‚x Ļ†Ī± (x, z) + Ļ†Ī± (x, z)|2

mĪ± (x, z) = āˆ’

Since Ļ†Ī± (x, z) and āˆ‚x Ļ†Ī± (x, z) are entire functions of z, real-valued on R, and have no common zeros, the denominator is continuous and nonzero on C+ āˆŖ R. Thus, mĪ± (x, z) extend continuously to R with boundary values mĪ± (x, Ī») = āˆ’

(iāˆ‚x ĪøĪ± (x, Ī») + ĪøĪ± (x, Ī»))(āˆ’iāˆ‚x Ļ†Ī± (x, Ī») + Ļ†Ī± (x, Ī»)) . (āˆ‚x Ļ†Ī± (x, Ī»))2 + Ļ†Ī± (x, Ī»)2

414

11. One-dimensional SchrĀØodinger operators

Since W (ĪøĪ± (Ā·, z), Ļ†Ī± (Ā·, z)) = 1, the imaginary part is computed to be 1 Im mĪ± (x, Ī») = , (āˆ‚x Ļ†Ī± (x, Ī»))2 + Ļ†Ī± (x, Ī»)2 so the Herglotz function mĪ± (x, z) corresponds to the measure 1 dĪ». Ļ€((āˆ‚x Ļ†Ī± (x, Ī»))2 + Ļ†Ī± (x, Ī»)2 ) Since the Herglotz functions mĪ± (x, z) converge pointwise to the Herglotz function mĪ± (z), they converge uniformly on compacts, so corresponding  measures converge to Ī¼Ī± by Proposition 7.28. Carmonaā€™s formula (11.113) is just one of many possible approximations, corresponding to a speciļ¬c choice made in (11.114). Other choices are useful in speciļ¬c situations; a variation useful in the study of decaying potentials V is given as Exercise 11.14. Our second application of Weyl disks involves continuity of the mfunction viewed as a function of the potential. Denote by mH the m-function corresponding to the operator H. Theorem 11.76. Let V āˆˆ L1loc ([0, b)) be regular at 0 and a limit point at b. d2 Fix Ī± āˆˆ R. Let H = āˆ’ dx 2 + V with boundary condition (11.2) at 0. Let Vn āˆˆ L1loc ([0, b)) be such that  c |Vn (x) āˆ’ V (x)| dx = 0 lim nā†’āˆž 0

2

d for all c < b, and let Hn = āˆ’ dx 2 + Vn with the same boundary condition (11.2) at 0 and (if Vn are limit circles at b) an arbitrary self-adjoint boundary condition at b. Then mHn (z) ā†’ mH (z) uniformly on compact subsets of z āˆˆ C+ .

Proof. By Corollary 11.9, for any real Ī²n ā†’ Ī², solutions un,Ī²n of  āˆ’fn,Ī² + Vn fn,Ī²n = zfn,Ī²n , n

fn,Ī²n (c) = cos Ī²n ,

 fn,Ī² (c) = sin Ī²n , n

converge in AC2 ([0, c]) to the solution of āˆ’fĪ² + V fĪ² = zfĪ² ,

fĪ² (c) = cos Ī²,

fĪ² (c) = sin Ī²,

so in particular, lim

nā†’āˆž

 cos Ī±fn,Ī² (0) āˆ’ sin Ī±fn,Ī²n (0) n  sin Ī±fn,Ī² (0) + cos Ī±fn,Ī²n (0) n

=

cos Ī±fĪ² (0) āˆ’ sin Ī±fĪ² (0) sin Ī±fĪ² (0) + cos Ī±fĪ² (0)

By Lemma 2.10 this implies uniform convergence in Ī², lim

nā†’āˆž

 (0) āˆ’ sin Ī±f cos Ī±fn,Ī² n,Ī² (0)  (0) + cos Ī±f sin Ī±fn,Ī² n,Ī² (0)

=

cos Ī±fĪ² (0) āˆ’ sin Ī±fĪ² (0) sin Ī±fĪ² (0) + cos Ī±fĪ² (0)

,

.

11.10. Asymptotic behavior of m-functions

415

so the Weyl circle āˆ‚DVn (x, z) converges in Hausdorļ¬€ metric dH to the Weyl circle āˆ‚DV (x, z). Since V is a limit point at b, for any > 0 there exist x < b such that the diameter of DV (x, z) is smaller than . For large enough n, dH (DVn (x, z), DV (x, z)) < , so |mHn (z)āˆ’mH (z)| < 2 . This proves that mHn converges to mH pointwise  in C+ ; uniform convergence on compacts follows from Corollary 7.18. Since the Weyl disk formalism is based on the J -monotonicity property of transfer matrices, they can be studied for any solution of an initial value problem of the form āˆ‚x T (x, z) = iJ (A(x)z + B(x))T (x, z), where A, B āˆˆ L1loc ([0, āˆž)), Tr(AJ ) = Tr(BJ ) = 0, A(x) ā‰„ 0, and B(x)āˆ— = B(x); such an initial value problem is sometimes called a Hamiltonian system. This point of view allows us to associate a Weyl function to the family of transfer matrices without reference to a self-adjoint operator (see, e.g., [11]). The special case B = 0 is the setting of de Branges canonical systems; it is particularly natural from an inverse spectral theoretic point of view, since the correspondence between trace-normalized canonical systems (Tr A = 1) and their Weyl functions is a bijection to the set of all Herglotz functions, by a deep theorem of de Branges; see [24, 78, 80].

11.10. Asymptotic behavior of m-functions We will now investigate the asymptotics of the m-functions as z ā†’ āˆž. In order to be concise and complete, it will be convenient to use the following convention: Deļ¬nition 11.77. Let P be a metric space, and let Ī© āŠ‚ C. For two Ė† we use the notation functions F, G : Ī© Ɨ P ā†’ C, F = oĖœ(|G|),

z ā†’ āˆž, z āˆˆ Ī©, uniformly in bounded subsets of P

to denote that for every bounded subset Q āŠ‚ P , lim sup sup zā†’āˆž pāˆˆQ zāˆˆĪ©

|F (z, p)| 0. In particular, āˆ’k is a Herglotz function; it is the Weyl function corresponding to the free half-line SchrĀØ odinger operator with a Dirichlet boundary condition at 0. The central result is the following: Theorem 11.78. Let V āˆˆ L1loc ([0, b)) be real-valued, b ā‰„ 1, and let H = d2 āˆ’ dx 2 + V have a Dirichlet boundary condition at 0. If V is a limit circle at b, assume an arbitrary self-adjoint boundary condition at b. Then the following hold. (a) For any Ī“ > 0,  1 eāˆ’2kt V (t) dt + oĖœ(|k|āˆ’1 ), m(z) = āˆ’k āˆ’

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“],

0

uniformly in bounded subsets of V āˆˆ L1 ([0, 1]). (b) If in addition H is semibounded, i.e., inf Ļƒ(H) > āˆ’āˆž, then  1 m(z) = āˆ’k āˆ’ eāˆ’2kt V (t) dt + o(|k|āˆ’1 ), z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]. 0

This is uniform in bounded subsets of V āˆˆ L1 ([0, 1]) with H such that inf Ļƒ(H) ā‰„ C, where C āˆˆ R. The proof of (a) consists of two parts: one is the derivation of the special case of operators on [0, 1] with a Dirichlet boundary condition at 1, and the other is an Atkinson argument which uses Weyl disks. We formulate both as lemmas below. The O( ) estimates come from explicit bounds on fundamental solutions and explicit functions like c(x, k) and s(x, k), whereas some of the o( ) estimates will come from the dominated convergence theorem and will therefore be pointwise in V āˆˆ L1 ([0, 1]). The proof of (b) uses the PhragmĀ“enā€“LindelĀØ of principle to extend to a bigger sector which includes a negative half-line. Poles of m arbitrarily far on the negative half-line would be an obstacle to such asymptotics, so this is only possible if m is analytic on C+ āˆŖ (āˆ’āˆž, C) āˆŖ Cāˆ’ for some C āˆˆ R, or equivalently, if H is semibounded. In general, the potential V outside [0, 1] can be modiļ¬ed to make inf Ļƒ(H) arbitrarily small or āˆ’āˆž; nonetheless, (b) is often applicable and useful. By Corollary 11.62, SchrĀØodinger operators

11.10. Asymptotic behavior of m-functions

417

with two regular endpoints are semibounded, and this suļ¬ƒcient criterion for semiboundedness will be generalized in Section 11.14. Lemma 11.79. For any Ī“ > 0 and V āˆˆ L1 ([0, 1]), assuming Dirichlet boundary conditions at 0 and 1,  1 eāˆ’2kt V (t) dt + oĖœ(|k|āˆ’1 ), z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“], m(z) = āˆ’k āˆ’ 0

(11.115)

uniformly in bounded subsets of V āˆˆ

L1 ([0, 1]).

Proof. By evaluating Wronskians at x = 0 and x = 1, we can express the m-function as W (Ļˆz+ , vz ) v(1, z) m(z) = āˆ’ =āˆ’ . (11.116) + u(1, z) W (Ļˆz , uz ) The key is to revisit Proposition 11.12 and its proof: by estimating only terms for n from 3 to āˆž and leaving the terms for n = 0, 1, 2 intact, we obtain asymptotic expansions for u(1, z) and v(1, z) with a higher power of |||k|||āˆ’1 . With the notation nāˆ’1    s(1āˆ’t1 , k) V (tj )s(tj āˆ’tj+1 , k) V (tn )s(tn , k) dn t, An = 2k n+1 eāˆ’k Ī”n (1)

Bn = 2k n eāˆ’k

 s(1 āˆ’ t1 , k) Ī”n (1)

j=1

nāˆ’1 



V (tj )s(tj āˆ’ tj+1 , k) V (tn )c(tn , k) dn t,

j=1

the proof of Proposition 11.12 gives  k  u(1, z) āˆ’ s(1, k) āˆ’ e A1 āˆ’  2k 2  k  v(1, z) āˆ’ c(1, k) āˆ’ e B1 āˆ’  2k

  ek A2  ā‰¤ |||k|||āˆ’4 e|Re k|+V L1 , 3 2k   ek  ā‰¤ |||k|||āˆ’3 e|Re k|x+V L1 . B 2  2 2k

In the nontangential limit z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“], the term |eāˆ’k | = eāˆ’ Re k ā‰¤ eāˆ’|k| sin(Ī“/2) decays exponentially so the series expansions for u(1, z), v(1, z) imply ! A1 A2 ek āˆ’3 1+ + 2 + O(|k| ) , u(1, z) = 2k k k ! B1 B2 ek 1+ + 2 + O(|k|āˆ’3 ) . v(1, z) = 2 k k By (11.116), this implies

! B1 āˆ’ A1 B2 āˆ’ A2 āˆ’ A1 (B1 āˆ’ A1 ) āˆ’3 + + O(|k| ) . m0,0 (z) = āˆ’k 1 + k k2 (11.117)

418

11. One-dimensional SchrĀØodinger operators

These terms can be further simpliļ¬ed. In particular,  1  1 B1 āˆ’ A1 = (1 āˆ’ eāˆ’2k(1āˆ’t) )V (t)eāˆ’2kt dt = V (t)eāˆ’2kt dt + O(eāˆ’2 Re k ). 0

0

(11.118) The second term can be rewritten more explicitly (Exercise 11.15), but for our purposes, it suļ¬ƒces to use the bounds |s(t, k)| ā‰¤

et Re k , |k|

|c(t, k)| ā‰¤ et Re k

and the pointwise limits lim

zā†’āˆž arg zāˆˆ[Ī“,2Ļ€āˆ’Ī“]

1 s(t, k) c(t, k) = lim = tk tk zā†’āˆž e /k e 2 arg zāˆˆ[Ī“,2Ļ€āˆ’Ī“]

to conclude by dominated convergence that nontangentially,  n  V (tj ) n d t. An , Bn ā†’ 2 Ī”n (1) j=1

It immediately follows that for each V , B2 āˆ’ A2 āˆ’ A1 (B1 āˆ’ A1 ) ā†’ 0. Using this and (11.118), the expansion (11.117) improves to (11.115).



Lemma 11.80. In the setting of Theorem 11.78, the radius r(z) of the Weyl disk D0 (1, z) decays exponentially as z ā†’ āˆž, r(z) =

āˆš 2|z| āˆš eāˆ’2 Re āˆ’z (1 + O(|z|āˆ’1/2 ), |Im āˆ’z|

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“],

uniformly for bounded sets of V āˆˆ L1 ([0, 1]) for any Ī“ > 0. Proof. The Weyl disk is given by ! ' & ! w w āˆ— āˆ— T (1, z) J T (1, z) ā‰„0 . D(1, z) = w | 1 1 The matrix M = T (1, z)āˆ— J T (1, z) obeys det M = āˆ’1, so the radius of the Weyl disk can be computed by Lemma 7.6 as 1 1 . (11.119) r(z) = āˆ’ = M11 |uz (1)uz (1) āˆ’ uz (1)uz (1)| Proposition 11.12 implies ek (1 + O(|k|āˆ’1 )), 2k which together imply uz (1) =

uz (1)uz (1)

=e

uz (1) =

2 Re k

ek (1 + O(|k|āˆ’1 )), 2

! 1 āˆ’2 + O(|k| ) . 4k

(11.120)

(11.121)

11.10. Asymptotic behavior of m-functions

419

We insert this into (11.119) and use   1   āˆ’ 1  = |k āˆ’ k| = Im k .  4k 4k  4|k|2 2|k|2 Since |k| = O(|Im k|) as z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“], this implies |Im k| 2 Re k 1 = e (1 + O(|k|āˆ’1 )), r(z) 2|k|2 

and inverting this completes the proof.

Proof of Theorem 11.78. (a) Denote by m0 (z) the m-function which corresponds to a Dirichlet boundary condition at 1. Then m0 (z) āˆˆ āˆ‚D0 (1, z) and m(z) āˆˆ D(1, z) so |m(z) āˆ’ m0 (z)| ā‰¤ 2r(z). Since r(z) decays exponentially as z ā†’ āˆž with arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“], the polynomial asymptotics of m0 (z) from Lemma 11.79 apply also to m(z) in the sector arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“]. (b) Since m and m0 are meromorphic Herglotz functions whose sets of poles are bounded below, by a corollary of the PhragmĀ“enā€“LindelĀØ of method (Corollary 7.64) the asymptotics of m(z) extends to the sector arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“].  If we are only interested in the leading asymptotics, we can simplify this: Corollary 11.81. In the setting of Theorem 11.78, āˆš m(z) = āˆ’ āˆ’z + oĖœ(1)

(11.122)

as z ā†’ āˆž, with the same uniformity statements as in Theorem 11.78. Proof. By dominated convergence, for any V āˆˆ L1 ([0, 1]),  1 lim eāˆ’2kt V (t) dt = 0, zā†’āˆž arg zāˆˆ[Ī“,2Ļ€āˆ’Ī“] 0

so the result follows from Theorem 11.78.



While the boundary condition at 1 was shown to have an asymptotically exponentially small contribution, the eļ¬€ect of the boundary condition at 0 is much more interesting; we leave this to Exercise 11.17. Under some continuity assumptions, the m-function asymptotics can be made more precise by a more careful analysis of integrals involving V . The multiplier eāˆ’2kt decays as t goes from 0 to 1, so this integral emphasizes the values of V (t) from small t. In particular, with some additional regularity at 0, only the value V (0) matters:

420

11. One-dimensional SchrĀØodinger operators

Corollary 11.82. In the setting of Theorem 11.78, if in addition V has a Lebesgue point at 0, then V (0) + o(|k|āˆ’1 ) 2k as z ā†’ āˆž in the same sector as before. m(z) = āˆ’k āˆ’

(11.123)

Proof. It suļ¬ƒces to prove that if V has a Lebesgue point at 0, then  1 2keāˆ’2kt V (t) dt ā†’ V (0), z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]. (11.124) 0 x Let us denote f (x) = 0 |V (t) āˆ’ V (0)| dt and f (x) . xāˆˆ(0,] x

Ļ‰( ) = sup

By the Lebesgue point condition, limā†’0 Ļ‰( ) = 0. Now we ļ¬x > 0 and note  1 |2keāˆ’2kt ||V (t) āˆ’ V (0)| dt ā‰¤ 2|k|eāˆ’2 Re k V 1 , 

which converges to 0 as k ā†’ āˆž in the sector arg k āˆˆ [Ī“/2, Ļ€ āˆ’ Ī“/2]. For the integral on [0, ], we use f āˆˆ AC([0, 1]) and integration by parts,   2|k|eāˆ’2 Re kt |V (t) āˆ’ V (0)| dt 0     āˆ’2 Re kt f (t) + 4|k| Re keāˆ’2 Re kt f (t) dt. = 2|k|e 0

Obviously, 

0

2|k|eāˆ’2 Re k f ( ) 

ā†’ 0, and the remaining integral obeys   4|k| Re keāˆ’2 Re kt f (t) dt ā‰¤ Ļ‰( ) 4|k| Re keāˆ’2 Re kt t dt.

0

0

Another integration by parts on the right-hand side solves this and gives   lim sup 2|k|eāˆ’2 Re kt |V (t) āˆ’ V (0)| dt ā‰¤ Ļ‰( ). kā†’āˆž 0 arg kāˆˆ[Ī“/2,Ļ€āˆ’Ī“/2]

We conclude that



lim sup

1

2|k|eāˆ’2 Re kt |V (t) āˆ’ V (0)| dt ā‰¤ Ļ‰( ),

kā†’āˆž 0 arg kāˆˆ[Ī“/2,Ļ€āˆ’Ī“/2]

which is arbitrarily small since > 0 is arbitrary. Thus,  1 2|k|eāˆ’2 Re kt |V (t) āˆ’ V (0)| dt = 0, lim kā†’āˆž 0 arg kāˆˆ[Ī“/2,Ļ€āˆ’Ī“/2]

which implies (11.124).



11.10. Asymptotic behavior of m-functions

421

This concludes the m-function asymptotics we will need in this chapter. We mention that Theorem 11.78 can be strengthened under stronger smoothness assumptions on V (Exercise 11.20). For instance, if V āˆˆ C n ([0, 1]) for some n āˆˆ N, there exist coeļ¬ƒcients c0 (V ), c1 (V ), . . . , cn+2 (V ) such that m(z) = āˆ’

n+2 

cj k 1āˆ’j + o(|k|āˆ’nāˆ’1)

j=0

as z ā†’ āˆž in the appropriate sector. The coeļ¬ƒcients are uniform for V in bounded subsets of C n ([0, 1]). By Corollary 11.82, we already know that c0 = 1,

c1 = 0,

c2 = V (0)/2.

To compute further terms, instead of trying to follow the calculations in the proof, the coeļ¬ƒcients cj are more naturally found by the following method. We will vary x and use the logarithmic derivative of the Weyl solution m(x, z) =

(Ļˆz+ ) (x) . Ļˆz+ (x)

For each x, the function m(x, z) obeys an expansion of the same form but with diļ¬€erent coeļ¬ƒcients, m(x, z) = āˆ’k

n+2 

cj (x)k āˆ’j + o(|k|āˆ’nāˆ’1),

arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“],

(11.125)

j=0

and the ļ¬rst few coeļ¬ƒcients are already known to be c0 = 1,

c1 = 0,

c2 (x) =

V (x) . 2

(11.126)

To obtain an eļ¬€ective way of computing coeļ¬ƒcients cj (x), let us consider their x-dependence. This relies on the following: Lemma 11.83. For any z āˆˆ C+ , m(x, z) obeys the Ricatti equation āˆ‚x m(x, z) = V (x) āˆ’ z āˆ’ m(x, z)2 . Proof. This follows by a calculation from āˆ’(Ļˆz+ ) + V Ļˆz+ = zĻˆz+ .



Theorem 11.84. For V āˆˆ C n ([0, 1]), the coeļ¬ƒcients cj (x) are locally absolutely continuous in x and are described by the recursion relation 1 1 cj (x)cnāˆ’j (x), cn (x) = cnāˆ’1 (x) āˆ’ 2 2 nāˆ’2 j=2

together with (11.126).

n ā‰„ 3,

(11.127)

422

11. One-dimensional SchrĀØodinger operators

Proof. It follows from the Ricatti equation and (11.125) that āˆ‚x m(z, x) has an asymptotic expansion of the form āˆ‚x m(x, z) = āˆ’k

n+2 

dj (x)k āˆ’j + o(|k|āˆ’nāˆ’1)

j=0

with d0 = d1 = 0 and dl (x) =

l+1 

for l ā‰„ 2.

cj (x)cl+1āˆ’j (x)

j=0

However, integrating āˆ‚x m(āˆ’k 2 , x) from 0 to a and using uniform boundedness of the error implicit in o(|k|āˆ’nāˆ’1 ) implies that n+2  a dj (x) dx k āˆ’j + o(|k|āˆ’nāˆ’1). m(a, z) āˆ’ m(0, z) = āˆ’k j=0

0

Comparing coeļ¬ƒcients with (11.125), it follows that cj (x) āˆˆ AC([0, a]) and cj = dj . Thus, the previous relations combine into cl =

l+1 

cj cl+1āˆ’j = 2cl+1 +

j=0

lāˆ’1 

cj cl+1āˆ’j .

j=2

Using (11.126), this can be rewritten as a formula for cl+1 in terms of lower order coeļ¬ƒcients, which gives the recursion (11.127).  Now the coeļ¬ƒcients cn are computable recursively using (11.126) and (11.127), e.g., c3 (x) =

V  (x) , 4

c4 (x) =

V  (x) āˆ’ V (x)2 , 8

... . 2

d Let us return to the general setting of a self-adjoint operator H = āˆ’ dx 2+ V on I = (āˆ’ , + ) with separated boundary conditions. For x āˆˆ I, denote

mĀ± (x; z) = Ā±

(ĻˆzĀ± ) (x) . ĻˆzĀ± (x)

These are the half-line m-functions corresponding to the SchrĀØodinger operators on I āˆ© [x, āˆž) and I āˆ© (āˆ’āˆž, x], respectively, with Dirichlet boundary conditions at x. In terms of them, from (11.69), the diagonal Greenā€™s function can be written as 1 . (11.128) G(x, x; z) = āˆ’ māˆ’ (x; z) + m+ (x; z) This allows us to compute their asymptotics, e.g.,

11.11. The local Borgā€“Marchenko theorem

423

Lemma 11.85. For any x āˆˆ I, the diagonal Greenā€™s function obeys the nontangential asymptotics 1 G(x, x; z) = āˆš + o(z āˆ’1 ), z ā†’ āˆž, 2 āˆ’z in the sector arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“] for any Ī“ > 0. Proof. The asymptotics in arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“] follow from (11.128) since āˆš  mĀ± (x; z) = āˆ’ āˆ’z + o(1) in the same sector.

11.11. The local Borgā€“Marchenko theorem For a SchrĀØ odinger operator with a regular endpoint at 0 and its m-function m(z), we have already seen that the leading asymptotic behavior of m(z) encodes the boundary condition at 0, and that further terms in the expansion encode the behavior of V at 0. In fact, we are about to see that the mfunction determines the entire potential uniquely: Theorem 11.86 (Borgā€“Marchenko). Let V āˆˆ L1loc ([0, b)), VĖœ āˆˆ L1loc ([0, Ėœb)). Ėœ with separated boundary conditions Consider SchrĀØ odinger operators H, H Ėœ and potentials V, V , respectively, and denote by m(z), m(z) Ėœ the correspondĖœ ing m-functions. If m = m, Ėœ then H = H. This result has a local version. We will present the local version ļ¬rst, following a short proof due to Bennewitz: Theorem 11.87 (The local Borgā€“Marchenko theorem). In the setting of Ėœ by Theorem 11.86, denote the boundary conditions at 0 for H, H cos Ī±f (0) + sin Ī±f  (0) = 0, cos Ī± Ėœ f (0) + sin Ī± Ėœ f  (0) = 0. For any d āˆˆ (0, min(b, Ėœb)), the following are equivalent: (a) Ī± = Ī± Ėœ and V = VĖœ on (0, d). (b) For every , Ī“ > 0, m(z) āˆ’ m(z) Ėœ = O(eāˆ’2(dāˆ’) Re

āˆš āˆ’z

),

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“]. (11.129)

Proof of Theorem 11.87. āˆš From the asymptotics of fundamental solutions we know that m(z) = āˆ’ āˆ’z(1 + o(1)) if Ī± = 0 and (see Exercise 11.17) m(z) = cot Ī±(1 + o(1)) if Ī± āˆˆ (0, Ļ€), so the condition (11.129) implies Ī± = Ī± Ėœ. The leading asymptotics of fundamental solutions implies āˆš āˆš 1 Ļ†(x, z) = (āˆ’ sin Ī±+cos Ī±/ āˆ’z)ex āˆ’z (1+o(1)), z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€āˆ’Ī“]. 2 (11.130)

424

11. One-dimensional SchrĀØodinger operators

Moreover, by Lemma 11.85, G(x, x; z) = Ļ†(x, z)Ļˆ(x, z) ā†’ 0 as z ā†’ āˆž in the sector arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“]. Ėœ z) ā†’ 1 as z ā†’ āˆž with arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“], so By (11.130), Ļ†(x, z)/Ļ†(x, Ėœ z) and Ļ†(x, Ėœ z)Ļˆ(x, z) converge to 0. Thus, so does their diļ¬€erence Ļ†(x, z)Ļˆ(x, Ėœ z)Īø(x, z) āˆ’ Ļ†(x, z)Īø(x, Ėœ z) + (m(z) āˆ’ m(z))Ļ†(x, Ėœ z). Ļ†(x, Ėœ z)Ļ†(x,

(11.131)

Ėœ z) and Īø(x, z) = (a) =ā‡’ (b): If V = VĖœ on (0, d), then Ļ†(x, z) = Ļ†(x, Ėœ z) for x āˆˆ (0, d), so (11.131) implies Īø(x, Ėœ z) ā†’ 0, (m(z) āˆ’ m(z))Ļ†(x, Ėœ z)Ļ†(x, Ėœ z) implies m(z) āˆ’ which by the leading asymptotics of Ļ†(x, z) and Ļ†(x, āˆš m(z) Ėœ = o(eāˆ’2x Re āˆ’z ) for all x < d. Ėœ z), (11.131) (b) =ā‡’ (a): By the leading asymptotics of Ļ†(x, z) and Ļ†(x, implies that for any x āˆˆ (0, d), Ėœ z)Īø(x, z) āˆ’ Ļ†(x, z)Īø(x, Ėœ z) ā†’ 0, F (z) = Ļ†(x,

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“].

The function F is entire and, by the estimates for fundamental solutions, obeys āˆš 1/2 z ā†’ āˆž. F (z) = O(e2x|Re āˆ’z| ) = O(e2x|z| ), Moreover, by the symmetry F (ĀÆ z ) = F (z), the function decays along any nonreal ray. By the PhragmĀ“enā€“LindelĀØ of Theorem 7.63 applied in the left and right half-planes, F is bounded on C, so it is constant by Liouvilleā€™s theorem and therefore identically equal to 0. Thus, for any z āˆˆ C \ R and x āˆˆ (0, d), Ėœ z) Īø(x, Īø(x, z) = . Ėœ z) Ļ†(x, z) Ļ†(x, Diļ¬€erentiating in x and using Ļ† Īø āˆ’Ļ†Īø = Ļ†Ėœ ĪøĖœāˆ’ Ļ†ĖœĪøĖœ = 1 implies that Ļ†2 = Ļ†Ėœ2 . Taking the logarithmic derivative in x gives Ļ†Ėœ (x, z) Ļ† (x, z) , = Ėœ z) Ļ†(x, z) Ļ†(x, and diļ¬€erentiating again implies Ļ†Ėœ (x, z) Ļ† (x, z) = , Ėœ z) Ļ†(x, z) Ļ†(x, Ėœ implies that V = VĖœ on which, since āˆ’Ļ† + V Ļ† = zĻ† and āˆ’Ļ†Ėœ + VĖœ Ļ†Ėœ = z Ļ†, (0, d).  Proof of Theorem 11.86. Theorem 11.87 applies for any d < min(b, Ėœb), Ėœ so Ī± = Ī± Ėœ and V = VĖœ on (0, c), with  c = min(b, Ėœb). In particular, H and H have the same Weyl disks D(c, z) = xāˆˆ(0,c) D(x, z).

11.12. Full-line eigenfunction expansions

425

Note that b = c if and only m(z) āˆˆ āˆ‚D(c, z), and analogously Ėœb = c if and only if m(z) Ėœ āˆˆ āˆ‚D(c, z). Thus, m = m Ėœ implies that b = c if and only if Ėœb = c. Then, by the deļ¬nition of c, we conclude b = Ėœb = c. Now the value of m=m Ėœ encodes the boundary condition, if any, at the endpoint b = Ėœb (see Ėœ Exercise 11.13), so H = H. 

11.12. Full-line eigenfunction expansions In this section, we present the general eigenfunction expansion for onedimensional SchrĀØ odinger operators with separated boundary conditions. We write the interval as I = (āˆ’ , + ). The endpoints can be ļ¬nite or inļ¬nite. If the potential is a limit circle at the endpoint Ā± , we assume that a selfadjoint boundary condition has been ļ¬xed at Ā± , i.e., vĀ± , f ) = 0 WĀ± (ĀÆ

for f āˆˆ D(H)

āˆ— such that W (ĀÆ for some vĀ± āˆˆ XĀ± \ XĀ± Ā± vĀ± , vĀ± ) = 0.

While the main motivation is the full-line SchrĀØ odinger operator corresponding to a potential on R, this expansion can be applied to any onedimensional SchrĀØ odinger operator with separated boundary conditions, and this expansion is sometimes useful even when applied to half-line problems. We ļ¬x an internal point x0 āˆˆ I and denote by Ļ†z , Īøz solutions of āˆ’f  + (V āˆ’ z)f = 0 satisfying the initial conditions ! ! 1 0 Ļ†z (x0 ) Īøz (x0 ) = . 0 1 Ļ†z (x0 ) Īøz (x0 ) In particular, W (Īøz , Ļ†z ) = 1. We could ļ¬x more general conditions in the style of (11.72), but here that does not correspond to a boundary condition of the SchrĀØodinger operator and is only an internal choice; a diļ¬€erent choice would lead to an only superļ¬cially diļ¬€erent eigenfunction expansion and is not of interest to us. Lemma 11.88. For z āˆˆ C \ R, the Weyl solutions ĻˆzĀ± have nonzero Wronskian with Ļ†z . If we normalize them so that W (ĻˆzĀ± , Ļ†z ) = 1 and deļ¬ne mĀ± (z) = Ā±W (Īøz , ĻˆzĀ± ), then ĻˆzĀ± = Īøz Ā± mĀ± (z)Ļ†z

(11.132)

z ) = mĀ± (z). and mĀ± (z) are Herglotz functions with the symmetry mĀ± (ĀÆ Proof. The claims for Ļˆz+ and m+ are immediate from the half-line setting applied on the interval [x0 , + ). The claims for Ļˆzāˆ’ and māˆ’ follow from the same results after an aļ¬ƒne transformation to reverse the interval. That reversal changes the sign of the derivative in x, which explains the Ā± sign in (11.132). 

426

11. One-dimensional SchrĀØodinger operators

In particular, mĀ± correspond naturally to the half-line SchrĀØodinger operators Hāˆ’ and H+ on (āˆ’ , x0 ) and (x0 , + ) with a Dirichlet boundary condition at x0 and boundary conditions at Ā± (if needed) inherited from H. The main object corresponding to the whole-line operator on (āˆ’ , + ) is the Weyl M -matrix, deļ¬ned on C \ R by   m m m āˆ’m āˆ’

M= This can be written as  M=

māˆ’ m+ māˆ’ +m+ māˆ’ māˆ’ +m+

+

māˆ’ +m+ 1 māˆ’ āˆ’m+ 2 māˆ’ +m+

1 āˆ’ + 2 māˆ’ +m+ 1 āˆ’māˆ’ āˆ’m+

māˆ’ māˆ’ +m+ 1 āˆ’ māˆ’ +m +



.

! 0 1/2 āˆ’ , 1/2 0

so by Lemma 7.68, it is a matrix-valued Herglotz function. It will serve as the full-line analogue of the Weyl m-function; compare the following lemma with Lemma 11.57. As in Deļ¬nition 11.77, we use notation oĖœ(. . . ) to signify both a uniform O(. . . ) and a pointwise o(. . . ). Lemma 11.89. The Weyl M -matrix is a matrix-valued Herglotz function and !āˆ— ! ĪøĪ» (y) ĪøĪ» (x) + oĖœ(1), ā†“ 0, Im M (Ī» + i ) Im G(x, y; Ī» + i ) = Ļ†Ī» (x) Ļ†Ī» (y) uniformly on (x, y, Ī») āˆˆ [c, d]2 Ɨ [Ī»1 , Ī»2 ], with compact [c, d] āŠ‚ (āˆ’ , + ) and [Ī»1 , Ī»2 ] āŠ‚ R. Proof. With ĻˆzĀ± normalized as in (11.132), denote ! !  m āˆ’ m + 1 āˆ’māˆ’ m+ māˆ’ +m+ = M1 = + 1 1 āˆ’ māˆ’m+m W (Ļˆz+ , Ļˆzāˆ’ ) + Note that 1 M = (M1 + M1 ), 2

M1 āˆ’ M1

=

māˆ’ māˆ’ +m+ 1 āˆ’ māˆ’ +m +

 .

! 0 āˆ’1 . 1 0

Using (11.69) and (11.132), we can express Greenā€™s function in the form āŽ§    āŽŖ Ļ† (x) (y) Ļ† āŽŖ z z āŽŖ , x ā‰¤ y, M1 (z) āŽŖ āŽØ Īø (x) Īøz (y) z G(x, y; z) =     āŽŖ āŽŖ Ļ†z (x) (y) Ļ† z āŽŖ āŽŖ , x ā‰„ y. M1 (z) āŽ© Īøz (x) Īøz (y) It remains to study the asymptotics of Im G(x, y; Ī» + i ). Using AC2 ([c, d])analyticity of the fundamental solutions, their reality at Ī» āˆˆ R, and the

11.12. Full-line eigenfunction expansions

427

observation that the diļ¬€erence M1 āˆ’ M1 has real entries, it follows that, with z = Ī» + i , , ! !Ļ†z (x) Ļ†z (y)  = O( ), ā†“ 0, Im (M1 (z) āˆ’ M1 (z) ) Īøz (x) Īøz (y) so there is no diļ¬€erence in asymptotics in the cases x ā‰¤ y and x ā‰„ y. In both cases, M1 or M1 can be replaced by the average M = 12 (M1 + M1 ) to conclude , ! !Ļ†z (y) Ļ†z (x) + O( ), ā†“ 0. M (z) Im G(x, y; z) = Im Īøz (x) Īøz (y) The proof now proceeds analogously to the half-line case (Lemma 11.57), using analyticity of the fundamental solutions, and estimates for matrixvalued Herglotz functions. Namely, by (7.61), M (Ī» + i ) = O( āˆ’1 ),

ā†“ 0,

uniformly for Ī» āˆˆ [Ī»1 , Ī»2 ] and (M (Ī» + i ) + M (Ī» + i )āˆ— ) = o(1) as ā†“ 0.  By the matrix-valued Stieltjes inversion (Theorem 7.67), the Weyl M matrix corresponds to a Poisson-ļ¬nite measure Ī¼ and a 2 Ɨ 2 matrix-valued function W on R such that W ā‰„ 0 and Tr W = 1 Ī¼-a.e. and   1 lim h(Ī») Im M (Ī» + i ) dĪ» = h(Ī»)W (Ī») dĪ¼(Ī») Ļ€ ā†“0 for all h āˆˆ Cc (R). Note that M = M  implies W = W  . The full-line eigenfunction expansion will conjugate H to a multiplication operator on the Hilbert space L2 (R, C2 , W dĪ¼) (see Lemma 6.38). We can now introduce the eigenfunction expansion and its presumed inverse; their basic properties are proved analogously to the half-line case, so we omit details. Lemma 11.90. For f āˆˆ L2c (I), the function fĖ† : R ā†’ C2 deļ¬ned by  ! Ļ† (x)f (x) dx fĖ†(Ī») =  Ī» ĪøĪ» (x)f (x) dx is a continuous function of Ī» āˆˆ R. Lemma 11.91. For g āˆˆ L2c (R, C2 , W dĪ¼), the function gĖ‡ : I ā†’ C deļ¬ned by !  Ļ†Ī» (x) W (Ī»)g(Ī») dĪ¼(Ī») gĖ‡(x) = ĪøĪ» (x)

428

11. One-dimensional SchrĀØodinger operators

is in AC2 ([c, d]) for every [c, d] āŠ‚ (āˆ’ , + ) and !  Ļ†Ī» (x)  W (Ī»)g(Ī») dĪ¼(Ī»), gĖ‡ (x) = ĪøĪ» (x) !  Ļ†Ī» (x) gĖ‡ (x) = W (Ī»)g(Ī») dĪ¼(Ī»). ĪøĪ» (x) In particular, if g āˆˆ L2c (R, C2 , W dĪ¼) and gĖ‡ = 0 Lebesgue-a.e., then !  0 W (Ī»)g(Ī») dĪ¼(Ī») = . 0 The main result about full-line eigenfunction expansion follows. Theorem 11.92 (Full-line eigenfunction expansion). There is a unitary map U : L2 (I) ā†’ L2 (R, C2 , W dĪ¼) such that: (a) U f = fĖ† for compactly supported f ; (b) U āˆ’1 g = gĖ‡ for compactly supported g; (c) U HU āˆ’1 is the operator of multiplication of Ī» on L2 (R, C2 , W (Ī») dĪ¼(Ī»)). We omit the details of the proof, since the remaining steps are analogous to the proof of Theorem 11.56. As an immediate consequence of Proposition 9.41 we obtain the following. Corollary 11.93. The measure Ī¼ is a maximal spectral measure for H. The multiplicity n measures for H (see Theorem 9.31) are given by dĪ¼n = Ļ‡Sn dĪ¼, where Sn = {Ī» | rank W (Ī») = n}. The spectral representation provided by Theorem 11.92 can also be used to express spectral information about H in terms of half-line Weyl functions mĀ± . This allows us to relate spectral properties of H to the spectral properties of half-line operators HĀ± . The following results are immediate consequences of results proved in an abstract setting in Section 7.12: Corollary 11.94. Ļƒess (H) = Ļƒess (H+ ) āˆŖ Ļƒess (Hāˆ’ ). Corollary 11.95. The absolutely continuous spectrum of H is precisely the sum of absolutely continuous spectra of HĀ± , with multiplicities added, i.e., H|H (H) āˆ¼ = H+ |H (H ) āŠ• Hāˆ’ |H (H ) . ac

ac

+

ac

āˆ’

Corollary 11.96. For Ī¼s -a.e. x āˆˆ R, the following hold. (a) rank W (x) = 1. (b) mĀ± have normal limits which are values in R āˆŖ {āˆž}.

11.13. Subordinacy theory

429

(c) There exists Ī± = Ī±(x) āˆˆ [0, Ļ€) such that m+ (x + i0) = āˆ’māˆ’ (x + i0) = āˆ’ cot Ī±(x) and W (x) =

! āˆ’ cos Ī±(x) sin Ī±(x) cos2 Ī±(x) . āˆ’ cos Ī±(x) sin Ī±(x) sin2 Ī±(x)

In other words, the singular part of Ī¼ is supported on the set  S= {x | m+ (x + i0) = āˆ’māˆ’ (x + i0) = āˆ’ cot Ī±}. Ī±āˆˆ[0,Ļ€)

Moreover, S has Lebesgue measure zero.

11.13. Subordinacy theory Spectral properties of SchrĀØodinger operators can be studied through the behavior of formal eigensolutions with real spectral parameters. Obviously, a pure point spectrum of a SchrĀØodinger operator corresponds to formal eigensolutions which are square-integrable. Carmonaā€™s formula is useful, but it is not a pointwise criterion. Subordinacy theory developed by Gilbertā€“Pearson [40, 41] and Jitomirskayaā€“Last [47, 48] describes the decomposition into absolutely continuous/singular spectra and, more generally, decomposition into Ī±-continuous/Ī±-singular spectra in terms of the behavior of eigensolutions. We begin with the half-line setting with a potential V āˆˆ L1loc ([0, āˆž)). Deļ¬nition 11.97. Fix Ī» āˆˆ R. A nontrivial solution f of āˆ’f  + V f = Ī»f is called subordinate (at +āˆž) if x |f (t)|2 dt =0 (11.133) lim 0x 2 xā†’āˆž 0 |g(t)| dt for some solution g of āˆ’g  + V g = Ī»g. Lemma 11.98. (a) If (11.133) holds for some eigensolution g, then it holds for every eigensolution g linearly independent with f . (b) If f is subordinate, it is linearly dependent with f ; it is a constant multiple of Ļ†Ī± for some Ī± āˆˆ R. Proof. (a) If g = Cf , the limit (11.133) is 1/C 2 . Thus, (11.133) implies that g is linearly independent with f . Now any eigensolution h can be written as h = C1 f + C2 g and, if h is linearly independent with f , then C2 = 0. By elementary estimates, 1 |h|2 ā‰„ |C2 |2 |g|2 āˆ’ |C1 |2 |f |2 . 2

430

11. One-dimensional SchrĀØodinger operators

This implies

x x 2 |h(t)|2 dt 1 2 0 0 |g(t)| dt  ā‰„ āˆ’ |C1 |2 = āˆž, |C | lim inf lim inf  x 2 x 2 dt 2 dt xā†’āˆž xā†’āˆž 2 |f (t)| |f (t)| 0 0

and inverting completes the proof. (b) If g = f , then the limit in (11.133) is equal to 1. Thus, if f is subordinate, f must be linearly dependent with f . Thus, vectors (f  (0), f (0)) and (f  (0), f (0)) are linearly dependent. This implies that f  (0)/f (0) āˆˆ R āˆŖ {āˆž}, i.e., f is a scalar multiple of the solution Ļ†Ī± for some Ī± āˆˆ R.  Thus, when looking for subordinate solutions, we are not only focused on real spectral parameters Ī» but also on real solutions: to check whether Ļ†Ī± is subordinate, it suļ¬ƒces to compare it to ĪøĪ± . This comparison can be related to values of mĪ± (z): Lemma 11.99 (Jitomirskayaā€“Last inequality). For any x > 0, deļ¬ne (x) > 0 by !āˆ’1/2  x  x 2 2 Ļ†Ī± (t, Ī») dt ĪøĪ± (t, Ī») dt . (11.134)

(x) = 4 0

For all x > 0, āˆš 5 āˆ’ 24 ā‰¤ |mĪ± (Ī» + i (x))|

0

x āˆš !1/2 Ļ†Ī± (t, Ī»)2 dt 5 + 24 0x . ā‰¤ 2 |mĪ± (Ī» + i (x))| 0 ĪøĪ± (t, Ī») dt

(11.135)

Proof. Lemma 11.16 generalizes with the same proof to use Ļ† = Ļ†Ī± (Ā·, Ī»), Īø = ĪøĪ± (Ā·, Ī») instead of fundamental solutions u, v; in particular, the Weyl solution Ļˆ at z = Ī» + i can be viewed as a solution of āˆ’Ļˆ  + (V āˆ’ Ī»)Ļˆ = i Ļˆ,

W (Ļˆ, Ļ†) = 1,

W (Ļˆ, Īø) = āˆ’m(z)

(note the Weyl solution is taken at the complex spectral parameter z but compared to eigensolutions at Ī») and expressed by variation of parameters as  x (Īø(x)Ļ†(t) āˆ’ Īø(t)Ļ†(x))Ļˆ(t) dt. Ļˆ(x) = Īø(x) + mĪ± (z)Ļ†(x) + i

8

0 x 2 0 |f (t)| dt,

using the Cauchyā€“Schwarz inWith the notation f x = equality twice on the right-hand side implies |Ļˆ(x)| ā‰„ |Īø(x) + mĪ± (z)Ļ†(x)| āˆ’ |Īø(x)|Ļ†x Ļˆx āˆ’ |Ļ†(x)|Īøx Ļˆx . Rearranging and using the triangle inequality in L2 ([0, x]) gives Īø + mĪ± (z)Ļ†x ā‰¤ Ļˆx + 2 Īøx Ļ†x Ļˆx .

11.13. Subordinacy theory

431

Squaring this, combining with Ļˆ2x ā‰¤ Ļˆ2 = Im mĪ± (z)/ , and using the choice of = (x) such that 2 Īøx Ļ†x = 1, we obtain Im mĪ± (z) ā‰¤ 8Īøx Ļ†x |mĪ± (z)|.

Using the triangle inequality on the left-hand side, this implies Īø + mĪ± (z)Ļ†2x ā‰¤ 4

(Īøx āˆ’ |mĪ± (z)|Ļ†x )2 ā‰¤ 8Īøx Ļ†x |mĪ± (z)|. Dividing this by Īø2x and expanding gives a quadratic inequality for Īŗ = |mĪ± (z)|Ļ†x /Īøx , Īŗ2 āˆ’ 10Īŗ + 1 ā‰¤ 0, āˆš āˆš  which implies 5 āˆ’ 24 ā‰¤ Īŗ ā‰¤ 5 + 24, completing the proof. The main result of subordinacy theory is that subordinacy of Ļ†Ī± corresponds to inļ¬nite normal boundary values of mĪ± : Theorem 11.100. Let H be regular at 0 and a limit point at āˆž. For any Ī» āˆˆ R, Ļ†Ī± (Ā·, Ī») is subordinate if and only if lim mĪ± (Ī» + i ) = āˆž. ā†“0

(11.136)

Proof. Obviously, the function (x) deļ¬ned above is a strictly decreasing, continuous function of x; moreover, Ļ†Ī± and ĪøĪ± are not both square-integrable in the limit point case, so lim (x) = 0. xā†’āˆž

By taking x ā†’ āˆž in the Jitomirskayaā€“Last inequality, we conclude that Ļ†Ī± is subordinate if and only if lim |mĪ± (Ī» + i (x))| = āˆž.

xā†’āˆž

By observed properties of (x), this is equivalent to (11.136).



Theorem 11.101. Let H be regular at 0 and a limit point at āˆž. The singular part of its spectral measure Ī¼Ī± is supported on the set SĪ± = {Ī» āˆˆ R | Ļ†Ī± is subordinate}, and the absolutely continuous part of the spectral measure Ī¼Ī± is mutually absolutely continuous with Ļ‡N (Ī») dĪ», where N = {Ī» āˆˆ R | there is no subordinate solution at Ī»}. Proof. The set SĪ± is precisely the set on which mĪ± (Ī» + i0) = āˆž. Moreover, Ī» āˆˆ N if and only if mĪ± (Ī» + i0) āˆˆ C+ or mĪ± (Ī» + i0) does not exist; however, the second case happens on a set of Lebesgue measure zero. Thus, the theorem follows from Corollary 7.49. 

432

11. One-dimensional SchrĀØodinger operators

Strengthening the subordinacy assumption, we will be able to characterize spectral decompositions with respect to Hausdorļ¬€ measures. Note the lim inf in the following deļ¬nition. Deļ¬nition 11.102. Fix Ī² āˆˆ (0, 1] and Ī» āˆˆ R. A nontrivial solution f of āˆ’f  + V f = Ī»f is called Ī²-subordinate (at +āˆž) if 2āˆ’Ī²  x 2 0 |f (t)| dt (11.137) lim inf  x Ī² = 0 xā†’āˆž 2 0 |g(t)| dt for some solution g of āˆ’g  + V g = Ī»g. Theorem 11.103. Let H be regular at 0 and a limit point at āˆž. Fix Ī² āˆˆ (0, 1). The Ī²-singular part of its spectral measure Ī¼Ī± is supported on the set SĪ±,Ī² = {Ī» āˆˆ R | Ļ†Ī± is Ī²-subordinate}, c . and the Ī²-continuous part of Ī¼Ī± is supported on SĪ±,Ī²

Proof. Raising (11.134) to power 1 āˆ’ Ī² and using that to divide (11.135) gives āˆš āˆš 2āˆ’Ī² 5 + 24 5 āˆ’ 24 1āˆ’Ī² Ļ†x ā‰¤2 . ā‰¤

(x)1āˆ’Ī² |mĪ± (Ī» + i (x))|

(x)1āˆ’Ī² |mĪ± (Ī» + i (x))| ĪøĪ²x Taking x ā†’ āˆž proves that Ļ† = Ļ†Ī± (Ā·, Ī») is Ī²-subordinate if and only if lim sup 1āˆ’Ī² |mĪ± (Ī» + i )| = āˆž. ā†“0

Now the claim follows from Theorem 6.29 and Theorem 7.51.



Subordinacy can also be used to study spectra of full-line SchrĀØodinger operators [40]: with obvious modiļ¬cations, we say a nontrivial eigensolution f at Ī» is subordinate at āˆ’āˆž if for some eigensolution g at Ī», 0 |f (t)|2 dt lim x0 = 0. 2 dt xā†’āˆ’āˆž |g(t)| x Denote SĪ±Ā± = {Ī» āˆˆ R | Ļ†Ī± (Ā·, Ī») is subordinate at Ā± āˆž}, N Ā± = {Ī» āˆˆ R | there is no subordinate eigensolution at Ā± āˆž}. Note that the set S=



(SĪ±āˆ’ āˆ© SĪ±+ )

Ī±āˆˆ[0,Ļ€)

is precisely the set of Ī» āˆˆ R for which there exists an eigensolution which is subordinate at both endpoints Ā±āˆž.

11.14. Potentials bounded below in an L1loc sense

433

Theorem 11.104. Let H be a SchrĀØ odinger operator on R which is limit point at Ā±āˆž. Let Ī¼ be its canonical spectral measure. (a) The singular part of Ī¼ is supported on S and has multiplicity 1. (b) Nāˆ’ āˆŖ N+ is an essential support for Ī¼ac , i.e., Ī¼ac is mutually absolutely continuous with Ļ‡Nāˆ’ āˆŖN+ (Ī») dĪ». (c) Nāˆ’ āˆ© N+ is an essential support for the multiplicity 2 part of Ī¼ac . Proof. This follows from Corollary 11.95, Corollary 11.96, and Theorem 11.100. 

11.14. Potentials bounded below in an L1loc sense In this section, we begin to specialize to potentials for which each ļ¬nite endpoint is regular, and that at each inļ¬nite endpoint,  x+1 lim sup Vāˆ’ (t) dt < āˆž. xā†’Ā±āˆž

x

(In Proposition 11.32 we showed that this condition implies a limit point endpoint.) Under this additional assumption, we will study semiboundedness, elements of the operator domain, and properties of eigenfunctions. Namely, we denote



Vāˆ’ L1

loc,unif

=

x+1

sup x:(x,x+1)āŠ‚I

Vāˆ’ (t) dt,

x

assuming from now on that I has length at least 1. Some properties of the operator domain follow immediately from properties of regular endpoints and Proposition 11.32: Corollary 11.105. If V āˆˆ L1loc (I) and Vāˆ’ L1 D(H) has the properties f āˆˆ

Lāˆž (I)

and

f

āˆˆ

loc,unif

< āˆž, then any f āˆˆ

L2 (I).

Next, we will prove that the corresponding SchrĀØ odinger operators are semibounded, with a controllable lower bound on the spectrum. We begin with the half-line setting. < āˆž. There Proposition 11.106. Let V āˆˆ L1loc ([0, āˆž)) and Vāˆ’ L1 loc,unif , such that the is a constant C, depending only on the value of Vāˆ’ L1 loc,unif following hold. (a) Consider the SchrĀØ odinger operators H0 , HĻ€/2 corresponding to this potential with a Dirichletā€“Neumann boundary condition at 0. Then Ļƒ(H0 ), Ļƒ(HĻ€/2 ) āŠ‚ [āˆ’C, āˆž).

434

11. One-dimensional SchrĀØodinger operators

(b) The Dirichlet m-function m0 obeys the asymptotic behavior, for any x > 0 and Ī“ > 0,  x m0 (z) = āˆ’k āˆ’ eāˆ’2kt V (t) dt + o(|k|āˆ’1 ), z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]. 0

In particular, m0 (z) = āˆ’k + o(1) as z ā†’ āˆž on the same sectors. (c) m0 is strictly increasing on (āˆ’āˆž, āˆ’C) and m0 (z) < 0 for z āˆˆ (āˆ’āˆž, āˆ’C). Proof. (a) Let Ī± āˆˆ {0, Ļ€/2}. By Proposition 11.32(d), there is a constant C such that for all Ļˆ āˆˆ D(HĪ± ), 1 CĻˆ, Ļˆ + Ļˆ, HĪ± Ļˆ ā‰„ Ļˆ  , Ļˆ   ā‰„ 0 2 (Ļˆ, HĪ± Ļˆ āˆˆ R because HĪ± is self-adjoint). This implies that Ļˆ, HĪ± Ļˆ ā‰„ āˆ’CĻˆ, Ļˆ

āˆ€Ļˆ āˆˆ D(HĪ± ),

so the criterion for semiboundedness (Corollary 8.42) completes the proof. (b) This follows from Theorem 11.78. (c) It follows from (a) that m0 and mĻ€/2 have analytic extensions through the interval (āˆ’āˆž, āˆ’C). Since the Neumann m-function is mĻ€/2 = āˆ’1/m0 , we conclude that m0 does not have poles or zeros on (āˆ’āˆž, āˆ’C). As a Herglotz function, m0 is strictly increasing on this interval, and by (b), limĪ»ā†’āˆ’āˆž m0 (Ī») = āˆ’āˆž. Thus, by continuity, m0 is strictly negative throughout that interval.  Now let us examine the full-line setting. Proposition 11.107. Let V be a potential on R such that V āˆˆ L1loc (R) and Vāˆ’ L1 < āˆž. There is a constant C, depending only on the value of loc,unif , such that the following hold. Vāˆ’ L1 loc,unif

(a) Consider the SchrĀØ odinger operator H corresponding to this potential. Then Ļƒ(H) āŠ‚ [āˆ’C, āˆž). (b) The diagonal Greenā€™s function G(x, x; z) obeys the asymptotic behavior, for any x > 0 and Ī“ > 0, 1 + o(z āˆ’1 ), G(x, x; z) = āˆš 2 āˆ’z

z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“].

(c) G(x, x; z) is strictly increasing on (āˆ’āˆž, min Ļƒ(H)) and is strictly positive on (āˆ’āˆž, min Ļƒ(H)).

11.14. Potentials bounded below in an L1loc sense

435

Proof. (a) We will use the properties of half-line m-functions mĀ± (x; z) which follow from Proposition 11.106. These functions have analytic extensions through (āˆ’āˆž, āˆ’C) which are strictly negative on that interval, so the functions 1 m+ māˆ’ āˆ’ , m+ + māˆ’ m+ + māˆ’ also have analytic extensions through the same interval. Since their sum is the Herglotz function Tr M which corresponds to the maximal spectral measure for H, it follows that Ļƒ(H) āŠ‚ [āˆ’C, āˆž). (b) This now follows from Lemma 11.85 and the PhragmĀ“enā€“LindelĀØ of method. (c) G(x, x; Ī») is strictly increasing on the interval (āˆ’āˆž, min Ļƒ(H)) and,  by (b), limĪ»ā†’āˆ’āˆž G(x, x; Ī») = 0. Thus, G(x, x; Ī») > 0 on this interval. Part (a) can also be proved more elegantly by using Exercise 11.5. The asymptotic behavior of eigensolutions at an inļ¬nite endpoint have important spectral consequences; we have already seen that through the context of Weyl solutions and through Carmonaā€™s formula. In such arguments, we often need input not only on the behavior of the eigensolution f , but also of its derivative f  . Our goal in this section is to show that various asymptotic properties of f extend to f  under the assumption that the negative part of V is uniformly locally L1 . The main technical estimate is adapted from work of Stolz [103]. It should be noted that all constants in it depend only on V and |z|. Lemma 11.108. Let f be a solution of āˆ’f  + V f = zf on the interval I, . and let C = |z| + Vāˆ’ L1 loc,unif

(a) Let [x, y] āŠ‚ I and assume Ļ‰ āˆˆ C, f (x) = 0, and Re[ĀÆ Ļ‰ f (t)] ā‰„ 0 for t āˆˆ [x, y]. Then Re[ĀÆ Ļ‰ f (y)] ā‰„ Re[ĀÆ Ļ‰ f (x)]+(yāˆ’x) Re[ĀÆ Ļ‰ f  (x)]āˆ’C(yāˆ’x)(yāˆ’x+1)|Ļ‰| max |f (t)|. xā‰¤tā‰¤y

(11.138) āˆš (b) Denote K = 1/ C. For any x such that [x āˆ’ K, x + K] āŠ‚ I, |f  (x)| ā‰¤ C(1 + 2K)

max

yāˆˆ[xāˆ’K,x+K]

|f (y)|.

(11.139)

8 1 . If x āˆˆ I obeys f (x) = 0, Re[f (x)f  (x)] ā‰„ (c) Denote Ī“ = āˆ’ 12 + 14 + 2C 0, and x + Ī“ āˆˆ I, then for all y āˆˆ [x, x + Ī“), |f (y)| >

|f (x)| . 2

(11.140)

436

11. One-dimensional SchrĀØodinger operators

Proof. (a) Since f āˆˆ AC2loc (I), A  t  y@   f (s)ds dt f (y) = f (x) + f (x) + x x  y = f (x) + (y āˆ’ x)f  (x) + (y āˆ’ s)f  (s)ds.

(11.141)

x

Ļ‰ f (s)] ā‰¤ |ĀÆ Ļ‰ f (s)| ā‰¤ |Ļ‰|M Denoting M = maxxā‰¤tā‰¤y |f (t)|, we have 0 ā‰¤ Re[ĀÆ for s āˆˆ [x, y]. Since f is an eigensolution, this implies A @  y  (y āˆ’ s)f (s)ds Re Ļ‰ ĀÆ x  y  y = (y āˆ’ s)V (s) Re [ĀÆ Ļ‰ f (s)] ds āˆ’ (y āˆ’ s) Re [ĀÆ Ļ‰ zf (s)] ds x x  y Vāˆ’ (s)ds āˆ’ |Ļ‰z|M (y āˆ’ x)2 ā‰„ āˆ’|Ļ‰|M (y āˆ’ x) x

ā‰„ āˆ’|Ļ‰|M (y āˆ’ x)(y āˆ’ x + 1)C, which together with (11.141) proves (11.138). (b) Without loss of generality, assume Re[f (x)f  (x)] ā‰„ 0 (the other case follows by considering f (āˆ’x)). Let M = maxxāˆ’Kā‰¤yā‰¤x+K |f (y)|. Assume that, contrary to (11.139), we have |f  (x)| > C(1 + 2K)M.

(11.142)

Denote g(y) = Re[f  (x)f (y)]. The function g is continuous, g(x) ā‰„ 0, and g  (x) = Re[f  (x)f  (x)] > 0, so g > 0 in some interval (x, x+ ). We claim that g > 0 in (x, x + K]. Assume to the contrary, that there exists y āˆˆ (x, x + K] such that g(y) = 0, and pick the smallest such y. Then g ā‰„ 0 on [x, y], so applying (a) with Ļ‰ = f  (x), we have g(y) ā‰„ g(x) + (y āˆ’ x)|f  (x)|2 āˆ’ C(y āˆ’ x)(y āˆ’ x + 1)|f  (x)|M   ā‰„ (y āˆ’ x)|f  (x)| |f  (x)| āˆ’ CM (y āˆ’ x + 1) . (11.143) Thus, by (11.142), g(y) > (y āˆ’ x)|f  (x)|CM (2K āˆ’ (y āˆ’ x)) > 0,

(11.144)

contradicting our assumption and proving g > 0 on (x, x + K]. Taking y = x + K in (11.144), we have Re[f  (x)f (x + K)] > CM K 2 |f  (x)| = M |f  (x)| ā‰„ |f  (x)f (x + K)|, which is a contradiction. Thus, the initial assumption (11.142) is wrong.

11.14. Potentials bounded below in an L1loc sense

437

(c) Assume the contrary. Then there exists y āˆˆ (x, x + Ī“) such that |f (y)| = |f (x)| 2 . Let s āˆˆ [x, y) be such that |f (s)| = max |f (t)|. tāˆˆ[x,y]

d |f (t)|2 , we have Re[f (s)f  (s)] = 0 (this is true even Since Re[f (t)f  (t)] = 12 dt if s = x, since we know a priori that Re[f (x)f  (x)] ā‰„ 0). Note also

Re[f (s)f (y)] ā‰¤ |f (s)f (y)| ā‰¤

|f (s)|2 , 2

so we may pick t āˆˆ (s, y] as the smallest t > s with Re[f (s)f (t)] =

(11.145) |f (s)|2 2 .

Using (b) with x replaced by s and y replaced by t, and with Ļ‰ = f (s) gives Re[f (s)f (t)] ā‰„ |f (s)|2 [1 āˆ’ C(t āˆ’ s)(t āˆ’ s + 1)] > |f (s)|2 [1 āˆ’ CĪ“(Ī“ + 1)] |f (s)|2 , 2 where we used t āˆ’ s ā‰¤ y āˆ’ x < Ī“. This is a contradiction with (11.145), which completes the proof.  =

This allows us to extend L2 -type estimates on f with polynomial or exponential weights to L2 -type and pointwise estimates on f and f  . We allow weighted estimates and consider weights w : I ā†’ (0, āˆž) which obey lim sup xā†’+āˆž

w(y) < āˆž. w(x) yāˆˆ[xāˆ’1,x+1] sup

(11.146)

Besides the trivial weight w = 1, the main examples to keep in mind are the polynomial weight w(x) = xĪŗ and the exponential weight w(x) = eĪŗx , for Īŗ āˆˆ R. Theorem 11.109. Let f be a solution of āˆ’f  + V f = zf on I = (c, āˆž) < āˆž. If w : I ā†’ (0, āˆž) obeys with a potential V which obeys Vāˆ’ L1 loc,unif (11.146) and  āˆž w(x)|f (x)|2 dx < āˆž, c

then



āˆž

w(x)|f  (x)|2 dx < āˆž

c

and lim

xā†’āˆž

.

w(x)|f (x)| = lim

xā†’āˆž

.

w(x)|f  (x)| = 0.

438

11. One-dimensional SchrĀØodinger operators

Proof. We use the constant Ī“ > 0 from Lemma 11.108(c). By (11.146), there exists C1 such that for all x large enough and all y āˆˆ [x āˆ’ Ī“, x + Ī“], C1āˆ’1 ā‰¤

w(y) ā‰¤ C1 . w(x)

We claim that 4C1 w(x)|f (x)| ā‰¤ Ī“



x+Ī“

2

w(y)|f (y)|2 dy. xāˆ’Ī“

For Re[f (x)f  (x)] ā‰„ 0, the claim follows from (11.140) by squaring it, multiplying by C1 w(y) ā‰„ w(x), and integrating from x to x + Ī“. The case Re[f (x)f  (x)] < 0 follows analogously, by considering f (āˆ’x). This implies that limxā†’āˆž w(x)|f (x)|2 = 0. Similarly, using (11.139), we conclude  4C1 C 2 (1 + 2K)2 x+K+Ī“  2 w(y)|f (y)|2 dy. w(x)|f (x)| ā‰¤ Ī“ xāˆ’Kāˆ’Ī“ āˆš This implies limxā†’āˆž w(x)|f  (x)|2 = 0 by square-integrability of wf , and it implies  āˆž  āˆž 8C1 C 2 (1 + 2K)2 (K + Ī“) w(x)|f  (x)|2 dx ā‰¤ w(y)|f (y)|2 dy < āˆž Ī“ c cāˆ’Kāˆ’Ī“ 

by Tonelliā€™s theorem.

In particular, for Weyl solutions, with the trivial weight w = 1, we conclude that the derivative of a Weyl solution is square-integrable and obtain pointwise decay: Corollary 11.110. Let Ļˆz+ be a Weyl solution for the endpoint + = +āˆž < āˆž. Then for c āˆˆ I, (Ļˆz+ ) āˆˆ with a potential V which obeys Vāˆ’ L1 loc,unif

L2 ((c, +āˆž)) and lim Ļˆz+ (x) = lim (Ļˆz+ ) (x) = 0.

xā†’+āˆž

xā†’+āˆž

Another application of Lemma 11.108 is to relate boundedness of eigensolutions to lack of subordinate solutions and therefore to absolutely continuous spectrum: < āˆž. Let Ī» āˆˆ R. If Lemma 11.111. Let V āˆˆ L1loc ([0, āˆž)) and Vāˆ’ L1 loc,unif all eigensolutions at Ī» are bounded, then there is no subordinate solution at Ī».

11.15. A Combesā€“Thomas estimate and Schnolā€™s theorem

439

Proof. Let f, g be linearly independent eigensolutions at Ī». Since f, g are bounded, by Lemma 11.108, g  is also bounded and there exist c, d such that !1/2 !1/2  x+d  x+d 2  2 |f (x)| ā‰¤ c |f (t)| dt , |f (x)| ā‰¤ c |f (t)| dt . xāˆ’d

xāˆ’d

Since f, g are linearly independent, their Wronskian W is nonzero. By the Cauchyā€“Schwarz inequality, !1/2  x+d     2 |f (t)| dt gāˆž + g  āˆž . |W | ā‰¤ |f ||g | + |f ||g| ā‰¤ c xāˆ’d

Squaring, integrating in x, dividing by x, and letting x ā†’ āˆž, we conclude  |W |2 1 x |f (t)|2 dt ā‰„ > 0. lim inf xā†’āˆž x 0 2dc2 (gāˆž + g  āˆž )2 x Since lim supxā†’āˆž x1 0 |g(t)|2 dt ā‰¤ g2āˆž < āˆž, dividing gives x |f (t)|2 dt > 0, lim inf 0x 2 xā†’āˆž 0 |g(t)| dt which proves that f is not subordinate.



Combining this with subordinacy theory (Theorem 11.101) gives the following criterion: < āˆž. Fix a Corollary 11.112. Let V āˆˆ L1loc ([0, āˆž)) and Vāˆ’ L1 loc,unif boundary condition at 0, and let Ī¼ denote the canonical spectral measure of the corresponding SchrĀØ odinger operator. If S denotes the set of Ī» āˆˆ R for which all eigensolutions are bounded, then Ļ‡S dĪ¼ is mutually absolutely continuous with Lebesgue measure on S. This is a very eļ¬€ective and commonly used criterion for establishing absolutely continuous spectrum; other sophisticated criteria for absolutely continuous spectrum have been proved by Lastā€“Simon [61]. A ļ¬rst application of Corollary 11.112 can be to prove that potentials V āˆˆ L1 ([0, āˆž)) give rise to absolutely continuous spectrum on (0, āˆž) (Exercise 11.23). Note that integrability of V on (0, āˆž) should be viewed as a kind of decay condition at +āˆž; SchrĀØodinger operators with decaying spectra at +āˆž are the subject of much study; see reviews [27, 56].

11.15. A Combesā€“Thomas estimate and Schnolā€™s theorem Weyl solutions are by deļ¬nition L2 -integrable near the corresponding endpoint. We will now see that this can be signiļ¬cantly improved away from the essential spectrum for potentials whose negative part is uniformly locally integrable. Estimates of this form are known as Combesā€“Thomas estimates.

440

11. One-dimensional SchrĀØodinger operators

Theorem 11.113. Let H be a SchrĀØ odinger operator on I = (āˆ’ , + ) with separated boundary conditions. Assume that (11.5) holds at + = +āˆž. If Ļˆz+ is the Weyl solution for some z āˆˆ C \ Ļƒess (H), then there exists Ī³ > 0 such that Ļˆz+ (x) = O(eāˆ’Ī³x ),

x ā†’ +āˆž.

While the result is stated for an arbitrary SchrĀØodinger operator with separated boundary conditions, the behavior of Weyl solutions at + does not depend on the potential near āˆ’ , and we will be able to reduce the proof to the case of a ļ¬nite, regular endpoint āˆ’ . Another observation is that f is a solution of āˆ’f  + V f = zf if and only if g(x) = eĪ³x f (x) is a solution of āˆ’g  + 2Ī³g  āˆ’ Ī³ 2 g + V g = zg. Exponential decay of f is equivalent to boundedness of g, which motivates the use of diļ¬€erential operators corresponding to the diļ¬€erential expression HĪ³ g = āˆ’g  + 2Ī³g  āˆ’ Ī³ 2 g + V g.

(11.147)

These are not self-adjoint for Ī³ = 0 but will still be useful in the proof. Proof. Let us ļ¬x z āˆˆ / Ļƒess (H) and the Weyl solution Ļˆ = Ļˆz+ . Let us choose c āˆˆ I such that Ļˆ(c) = 0. Then, we consider the operator H0 on [c, +āˆž) with the potential V and a Dirichlet boundary condition at c. More explicitly, writing the domain of H in terms of Lagrangian subspaces as D(H) = Yāˆ’ āˆ© Y+ , the domain of H0 is D(H0 ) = {g āˆˆ L2 ([c, āˆž)) | g = G|[c,āˆž) for some G āˆˆ Y+ and g(c) = 0}. / Ļƒess (H0 ), and Corollary 11.94 implies Ļƒess (H0 ) āŠ‚ Ļƒess (H) and therefore z āˆˆ then the assumption Ļˆ(c) = 0 ensures that z āˆˆ / Ļƒ(H0 ). From now on, we work with the operator H0 on the Hilbert space L2 ((c, āˆž)). As motivated above, we wish to consider the operator HĪ³ deļ¬ned formally by HĪ³ = eĪ³x H0 eāˆ’Ī³x . By Proposition 11.32(d), for any g āˆˆ D(H0 ), 1  2 g  ā‰¤ M g2 + g, H0 g 2 (g, H0 g is real because H0 is self-adjoint). By the Cauchyā€“Schwarz inequality and arithmetic meanā€“geometric mean inequality, there exists C such that g  2 ā‰¤ Cg2 + CH0 g2

(11.148) g

L2 ((c, āˆž)),

so we for all g āˆˆ D(H0 ). In particular, g āˆˆ D(H0 ) implies āˆˆ can rigorously deļ¬ne the operator HĪ³ by D(HĪ³ ) = D(H0 ) and (11.147).

11.15. A Combesā€“Thomas estimate and Schnolā€™s theorem

441

Combining (11.148) with the estimates g ā‰¤ (H0 āˆ’ z)āˆ’1 (H0 āˆ’ z)g and H0 g ā‰¤ (H0 āˆ’ z)g + |z|g shows that g ā‰¤ C0 (H0 āˆ’ z)g,

g   ā‰¤ C1 (H0 āˆ’ z)g

for some constants C0 , C1 independent of g āˆˆ D(H0 ). Thus, (HĪ³ āˆ’ H0 )g ā‰¤ 2Ī³g   + Ī³ 2 g ā‰¤ (2Ī³C1 + Ī³ 2 C0 )(H0 āˆ’ z)g, which ļ¬nally implies that for small enough Ī³ > 0, (HĪ³ āˆ’ H0 )(H0 āˆ’ z)āˆ’1  ā‰¤ (2Ī³C1 + Ī³ 2 C0 ) < 1. For such Ī³ > 0, the operator HĪ³ āˆ’ z = ((HĪ³ āˆ’ H0 )(H0 āˆ’ z)āˆ’1 + I)(H0 āˆ’ z) is invertible. Fix d < āˆž; by the proof of Theorem 11.42, there exists h āˆˆ L2 ([c, āˆž)) such that supp h āŠ‚ [c, d] and (H0 āˆ’ z)āˆ’1 h is a nontrivial multiple of Ļˆ on [d, āˆž). Then (HĪ³ āˆ’ z)āˆ’1 (eĪ³x h(x)) is a nontrivial multiple of eĪ³x Ļˆ(x) on [d, āˆž). However, (HĪ³ āˆ’z)āˆ’1 (eĪ³x h(x)) āˆˆ  D(H0 ) āŠ‚ Lāˆž (I), so Ļˆ(x) = O(eāˆ’Ī³x ) as x ā†’ +āˆž. The previous proof used the fact that any f āˆˆ D(H0 ) is a bounded function (Corollary 11.105). The same fact will be useful in the next two results, which closely relate the spectrum to the polynomial growth of formal eigensolutions. Theorem 11.114 (Schnol). Consider V āˆˆ L1loc ([0, āˆž)) such that (11.5) holds at + = +āˆž. Let Īŗ > 1/2 and deļ¬ne SĪŗ = {Ī» | Ļ†Ī» (x) = O(xĪŗ ),

x ā†’ āˆž}.

Then: (a) SĪŗ āŠ‚ Ļƒ(H); (b) the maximal spectral measure for H is supported on SĪŗ ; (c) SĪŗ = Ļƒ(H). Proof. (a) For Ī» āˆˆ / Ļƒ(H), there exists a Weyl solution Ļˆ(x) which is exponentially decaying at āˆž, i.e., there exists Ī³ > 0 such that Ļˆ(x) = O(eāˆ’Ī³x ) as x ā†’ āˆž. By Theorem 11.109, this implies Ļˆ  (x) = O(eāˆ’Ī³x ) and Ļ†Ī» (x) = O(xĪŗ ) implies Ļ†Ī» (x) = O(xĪŗ ), so their Wronskian obeys W (Ļˆ, Ļ†)(x) = O(xĪŗ eāˆ’Ī³x ). Since the Wronskian is independent of x, it must be 0, so Ļˆ, Ļ† are eigenvectors and Ī» is an eigenvalue of H, leading to a contradiction.

442

11. One-dimensional SchrĀØodinger operators

(b) We use eigenfunction expansions. If f (y) = G(y, x; z) = G(x, y; z) 1 Ļ†Ī» (x), and since U is unitary, for some z āˆˆ C \ R, recall that (U f )(Ī») = Ī»āˆ’z   1 dĪ¼(Ī»). |G(x, y; z)|2 dy = |Ļ†Ī» (x)|2 |Ī» āˆ’ z|2 By the integral representation for resolvents, since D(H) āŠ‚ Lāˆž (I), we see that for every f āˆˆ L2 (I),      sup  G(x, y; z)f (y) dy  < āˆž, yāˆˆI

so by the uniform boundedness principle,  sup |G(x, y; z)|2 dy < āˆž. xāˆˆI

Thus,   |Ļ†Ī» (x)|2 2 āˆ’Īŗ 2 (1 + x2 )āˆ’Īŗ dĪ¼(Ī») dx < āˆž. (1 + x ) |G(x, y; z)| dy dx = |Ī» āˆ’ z|2 By Tonelliā€™s theorem, this implies that for Ī¼-a.e. Ī»,  (1 + x2 )āˆ’Īŗ |Ļ†Ī» (x)|2 dx < āˆž, and therefore Ļ†Ī» (x) = O(|x|Īŗ ) as x ā†’ āˆž. (c) Since SĪŗ āŠ‚ Ļƒ(H), SĪŗ āŠ‚ Ļƒ(H); conversely, since Ī¼ is supported on  SĪŗ , Ļƒ(H) = supp Ī¼ āŠ‚ SĪŗ . Theorem 11.115 (Schnol). Let H be a SchrĀØ odinger operator on L2 (R) such < āˆž. Let Īŗ > 1/2 and denote by SĪŗ the set of Ī» for which that Vāˆ’ L1 loc,unif there exists a nontrivial solution of āˆ’u +V u = Ī»u such that u(x) = O(|x|Īŗ ) as x ā†’ Ā±āˆž. Then: (a) SĪŗ āŠ‚ Ļƒ(H); (b) the maximal spectral measure for H is supported on SĪŗ ; (c) SĪŗ = Ļƒ(H). Proof. (a) For Ī» āˆˆ / Ļƒ(H), there exist Weyl solutions Ļˆ Ā± , which are exponentially decaying at Ā±āˆž, respectively. If a solution is polynomially bounded, it must be a multiple of Ļˆ āˆ’ and of Ļˆ + , so it follows that W (Ļˆ+ , Ļˆāˆ’ ) = 0. This would mean that Ī» is an eigenvalue of H, leading to a contradiction. This proves SĪŗ āŠ‚ Ļƒ(H). (b) Consider W dĪ¼ as in the eigenfunction expansion. Fix y āˆˆ R and z āˆˆ C \ R. The function f (x) = G(x, y; z) is in L2 (R) and U f can be

11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map

443

computed analogously to Proposition 11.65 to give ! 1 Ļ†Ī» (y) . (U f )(Ī») = Ī» āˆ’ z ĪøĪ» (y) In particular, since U is unitary, !āˆ—   Ļ†Ī» (y) 2 |G(x, y; z)| dx = W ĪøĪ» (y)

Ļ†Ī» (y) ĪøĪ» (y)

!

1 dĪ¼(Ī»). |Ī» āˆ’ z|2

Since D(H) āŠ‚ Lāˆž (R), as in the half-line case the uniform boundedness principle implies  sup |G(x, y; z)|2 dx < āˆž, yāˆˆR

and therefore  

2 āˆ’Īŗ

(1 + x )

!āˆ— ! Ļ†Ī» (y) Ļ†Ī» (x) W (Ī») dĪ¼(Ī») dx < āˆž. ĪøĪ» (x) ĪøĪ» (y)

Since the integrand is nonnegative, by Tonelliā€™s theorem, for Ī¼-a.e. Ī», !āˆ— !  Ļ† (y) Ļ†Ī» (x) dx < āˆž. W (Ī») Ī» (1 + x2 )āˆ’Īŗ ĪøĪ» (x) ĪøĪ» (y) ! v1 āˆˆ C2 such that W ā‰„ vv āˆ— , we obtain Choosing a nonzero vector v = v2 !āˆ— !  2 āˆ’Īŗ Ļ†Ī» (x) āˆ— Ļ†Ī» (x) dx < āˆž, vv (1 + x ) ĪøĪ» (x) ĪøĪ» (x) which implies that the nontrivial solution f = v1 ĪøĪ» + v2 Ļ†Ī» obeys  (1 + x2 )āˆ’Īŗ |f (x)|2 dx < āˆž. Thus, this solution obeys f (x) = O(|x|Īŗ ) as x ā†’ Ā±āˆž. (c) Since SĪŗ āŠ‚ Ļƒ(H), SĪŗ āŠ‚ Ļƒ(H); conversely, since Ī¼ is supported on  SĪŗ , Ļƒ(H) = supp Ī¼ āŠ‚ SĪŗ .

11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map We will now consider SchrĀØ odinger operators with periodic potentials. Up to rescaling, we can assume that V is a periodic locally integrable function on R with period 1. Periodicity ensures that V is limit point at Ā±āˆž; we denote by H the corresponding SchrĀØodinger operator on L2 (R). As in the case of periodic Jacobi matrices, an important role will be played by the monodromy matrix (i.e., the transfer matrix over one period), which we will denote by T (z) = T (1, z). Since det T (z) = 1, the behavior of T (z) will be largely determined by its trace Ī”(z) = Tr T (z), called the

444

11. One-dimensional SchrĀØodinger operators

discriminant. By Lemma 10.57, the value of Ī”(z) determines the magnitude of eigenvalues of T (z), which enters the following proof. Lemma 11.116. The spectrum of H is Ļƒ(H) = {z āˆˆ C | Ī”(z) āˆˆ [āˆ’2, 2]}. In particular, for any z āˆˆ C \ R, Ī”(z) āˆˆ / [āˆ’2, 2]. Proof. For z such that Ī”(z) āˆˆ [āˆ’2, 2], T (z) has a unimodular eigenvalue, which generates a bounded eigenfunction f . Conversely, if Ī”(z) āˆˆ / [āˆ’2, 2], there exist eigenfunctions Ļˆ Ā± which are exponentially decaying at Ā±āˆž and exponentially growing at āˆ“āˆž, so any nontrivial linear combination will be exponentially growing in at least one direction. In summary, a polynomially bounded, nontrivial eigensolution exists if and only if Ī”(z) āˆˆ [āˆ’2, 2]. The claim now follows by Schnolā€™s theorem, since the set {z āˆˆ C | Ī”(z) āˆˆ [āˆ’2, 2]} is closed.  In particular, all zeros of Ī”2 āˆ’4 are on the real line. For further study, we introduce Herglotz function techniques. We deļ¬ne for z āˆˆ C+ the function  1 m+ (x; z) dx. (11.149) w(z) = 0

Since m+ is jointly continuous in (x, z) āˆˆ [0, 1] Ɨ C+ , it is locally uniformly continuous, so the integral is an analytic function of z by Moreraā€™s theorem. Lemma 11.117. For all z āˆˆ C+ ,  w(z) =

1 0

māˆ’ (x; z) dx.

(11.150)

Proof. Since G(x, x; z) āˆˆ C+ for all x, the function g(x; z) = log G(x, x; z) is well deļ¬ned on RƗC+ with Im g āˆˆ (0, Ļ€). Since W (Ļˆ + , Ļˆ āˆ’ ) is independent of x, diļ¬€erentiating gives āˆ‚x Ļˆ + Ļˆ āˆ’ + Ļˆ + āˆ‚x Ļˆ āˆ’ = m+ āˆ’ māˆ’ . Ļˆ+Ļˆāˆ’ Integrating in x from 0 to 1 and using periodicity of g shows  1 (m+ (x; z) āˆ’ māˆ’ (x; z)) dx = g(1, z) āˆ’ g(0, z) = 0, āˆ‚x g =

0

so now (11.150) follows from the deļ¬nition (11.149).



By averaging (11.149) and (11.150) and using (11.128), it also follows that !  1 1 dx. āˆ’ w(z) = 2G(x, x; z) 0

11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map

445

It is obvious that w is a Herglotz function, since the functions m+ are. Strikingly, we will soon see that w has two more Herglotz properties. Let us also introduce the Marchenkoā€“Ostrovski map Ī˜(z) = āˆ’iw(z). Lemma 11.118. For all z āˆˆ C+ , Im Ī˜(z) > 0 and ! ! m+ (z) iĪ˜(z) m+ (z) =e , T (z) 1 1 ! ! āˆ’māˆ’ (z) āˆ’iĪ˜(z) āˆ’māˆ’ (z) T (z) =e . 1 1

(11.151) (11.152)

Proof. The monodromy matrix evolves the Weyl solution from x = 0 to 1: ! ! (āˆ‚x Ļˆ + )(1, z) (āˆ‚x Ļˆ + )(0, z) = . T (z) Ļˆ + (0, z) Ļˆ + (1, z) By periodicity, the Weyl solution shifted by 1 is again a Weyl solution, so by uniqueness up to normalization, there exists Ī· āˆˆ C such that Ļˆ + (x + 1, z) = Ī·Ļˆ + (x, z). Since  1  1 (āˆ‚x Ļˆ + )(x, z) Ļˆ + (1, z) = dx = m+ (x; z) dx = w(z), ln + Ļˆ (0, z) Ļˆ + (x, z) 0 0 we conclude Ī· = ew(z) , which implies (11.151). Since Ļˆ + is square-integrable at +āˆž, |Ī·| < 1, which implies Im Ī˜(z) > 0. Equation (11.152) is proved analogously.  Lemma 11.119. (a) For all z āˆˆ C+ , Ī”(z) = 2 cos Ī˜(z).

(11.153) Ī”2 āˆ’ 4,

Ī˜ has an (b) For any interval (c, d) āŠ‚ R containing no zeros of analytic continuation to C+ āˆŖ (c, d) āˆŖ Cāˆ’ such that (11.153) holds. (c) For any z āˆˆ C, if Ī”(z) āˆˆ (āˆ’2, 2), then Ī” (z) = 0. Proof. Since T (z) has eigenvalues eĀ±iĪ˜(z) , its trace is computed as (11.153), which proves (a). The proofs of (b) and (c) are analogous to those of Lemmas 10.61 and 10.62 with q = 1.  Lemma 11.120. Zeros of Ī”2 āˆ’4 have multiplicity at most 2. At any double zero, (Ī”2 āˆ’ 4) < 0. Proof. Assume that Ī”2 āˆ’ 4 has a zero of multiplicity m at Ī», so that Ī”(Ī») = Ā±2. Then, by standard results in complex analysis, the equation Ī”(z) = Ā±2 cos t for t āˆˆ (0, ) locally has solutions Ī³j (t), j = 1, . . . , m, which  (0) lie on curves Ī³j with Ī³j (0) = Ī» and with arguments of Ī³1 (0), . . . , Ī³m

446

11. One-dimensional SchrĀØodinger operators

equispaced. Since Ī”(z) āˆˆ [āˆ’2, 2] implies z āˆˆ R, this can only happen if m ā‰¤ 2. Moreover, in the case m = 2, the two curves must lie on R, which  implies that Ī”2 āˆ’ 4 as a function on R has a local maximum at Ī». Since T (z) is the transfer matrix from x = 0 to x = 1, its entries are given in terms of the fundamental solutions u(x, z), v(x, z) on the interval [0, 1] from (11.15). In particular, Ī”(z) = v(1, z) + (āˆ‚x u)(1, z), so Propositions 11.11 and 11.12 immediately imply that |Ī”(z)| ā‰¤ 2e|Re k|+

1 0

|V (t)| dt

and |Ī”(z) āˆ’ 2c(1, k)| ā‰¤ 2|||k|||āˆ’1 e|Re k|+

1 0

āˆ€z āˆˆ C

|V (t)| dt

āˆ€z āˆˆ C.

(11.154)

This directly implies: Lemma 11.121. limĪ»ā†’āˆ’āˆž Ī”(Ī») = +āˆž. Using (11.154), we can adapt the counting lemma to prove the following: Lemma 11.122. For large enough positive integers N , Ī”2 āˆ’ 4 has exactly 2N + 1 zeros (counted with multiplicity) smaller than (N + 12 )2 Ļ€ 2 . Proof. The previous estimates imply |Ī”(z)2 āˆ’ 4c(1, k)2 | ā‰¤ 8|||k|||āˆ’1 e2|Re k|+2

1 0

|V (t)| dt

so that |(Ī”(z)2 āˆ’ 4) + 4k 2 s(1, k)2 | ā‰¤ 8|||k|||āˆ’1 e2|Re k|+2

1 0

,

|V (t)| dt

.

The function g(z) = āˆ’4k 2 s(1, k)2 = 4zs(1, k)2 is entire. It has a simple zero at 0 and double zeros at n2 Ļ€ 2 for n āˆˆ N, and no other zeros; thus, it 2 2 has 2N + 1 zeros including āˆš + 1/2) Ļ€ ). āˆš multiplicity1 on the interval (āˆ’āˆž, (N Moreover, on curves Im āˆ’z = (N + 2 )Ļ€ for N āˆˆ N and Re āˆ’z = CĻ€ for C ā‰„ 1, 4 |g(z)| > e2|Re k| . 9  1 2 01 |V (t)| dt , then on those contours, |k| > 6Ļ€e2 0 |V (t)| dt , Thus, if N, C > 6e so 8 2|Re k|+2  1 |V (t)| dt 0 e ā‰„ |(Ī”(z)2 āˆ’ 4) āˆ’ g(z)|, |g(z)| > |k| and RouchĀ“eā€™s theorem completes the proof.  It is now possible to describe the behavior of the discriminant on R and the spectrum of H:

11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map

447

Theorem 11.123. All zeros of Ī”2 āˆ’ 4 are real and can be listed, with multiplicity, as a sequence (Ī»n )āˆž n=1 such that Ī»2jāˆ’1 < Ī»2j ā‰¤ Ī»2j+1

āˆ€j āˆˆ N.

Moreover,

2 n ā‰” 1, 4 (mod 4) Ī”(Ī»n ) = āˆ’2 n ā‰” 2, 3 (mod 4),  and the periodic spectrum is E = āˆž j=1 [Ī»2jāˆ’1 , Ī»2j ].

(11.155)

Proof. The zeros of Ī”2 āˆ’ 4 divide R into intervals; counting the sign of Ī”2 āˆ’ 4 from āˆ’āˆž using Lemma 11.121, it follows that |Ī”| < 2 on the intervals (Ī»2jāˆ’1 , Ī»2j ) and |Ī”| > 2 on the intervals (Ī»2j , Ī»2j+1 ), whenever those intervals are open. Using Lemma 11.120 and Lemma 11.119(c) then determines the sign of Ī”(Ī»n ) by induction in n, which implies (11.155). 

2 Ī»1

Ī»2

Ī»3

Ī»4

Ī»5

Ī»

āˆ’2 Figure 11.2. The discriminant on R.

As a function on R, the discriminant is oscillatory toward +āˆž; see Figure 11.2. The intervals (Ī»2j , Ī»2j+1 ) are spectral gaps; the jth gap is said to be open if Ī»2j < Ī»2j+1 and closed if Ī»2j = Ī»2j+1 . For further study, we indicate another Herglotz property related to the Marchenkoā€“Ostrovski map, which comes from a relation with the diagonal Greenā€™s function: Proposition 11.124. For z āˆˆ C+ ,  1 G(x, x; z) dx. w (z) =

(11.156)

0

In particular, w is a Herglotz function, and w has an analytic extension to z ) = w (z) and C \ Ļƒ(H) which obeys w (ĀÆ 1 + o(1), w (z) = āˆš 2 āˆ’z for any Ī“ > 0.

z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]

(11.157)

448

11. One-dimensional SchrĀØodinger operators

Proof. Consider the function  āˆž āˆ‚z m+ (x; z) Ļˆzāˆ’ (x) Ļˆz+ (y)2 dy. h(x) = = āˆ’m+ (x; z) āˆ’ māˆ’ (x; z) W (Ļˆz+ , Ļˆzāˆ’ )Ļˆz+ (x) x Using

! Ļˆzāˆ’ (Ļˆzāˆ’ ) Ļˆz+ āˆ’ Ļˆzāˆ’ (Ļˆz+ ) W (Ļˆz+ , Ļˆzāˆ’ ) = = , Ļˆz+ (Ļˆz+ )2 (Ļˆz+ )2 the derivative of h is  āˆž 1 Ļˆ āˆ’ (x)Ļˆ + (x)  Ļˆz+ (y)2 dy āˆ’ z + z āˆ’ = āˆ‚z m+ (x; z) āˆ’ G(x, x; z) h (x) = + 2 Ļˆz (x) x W (Ļˆz , Ļˆz )

(the last step uses (11.87)). Since the function h is independent of nor1 malization of the Weyl solutions, it is 1-periodic, so 0 h (x) dx = 0. This implies  1  1 G(x, x; z) dx = āˆ‚z m+ (x; z) dx = w (z). 0

0

Since G(x, x; z) is Herglotz for each x, it follows that w is Herglotz. Due to joint continuity of G(x, y; z) in R Ɨ R Ɨ (C \ Ļƒ(H)), the righthand side of (11.156) deļ¬nes an analytic function on C \ Ļƒ(H) by Fubiniā€™s theorem and Moreraā€™s theorem. The conjugation symmetry of w follows from G(x, x; zĀÆ) = G(x, x; z). Finally, since the diagonal Greenā€™s function obeys the asymptotics G(x, x; z) = 2āˆš1āˆ’z + o(1) as z ā†’ āˆ’āˆž uniformly in x,  the function w obeys the same asymptotics. Theorem 11.125. All zeros of Ī” are simple and can be listed, with multiplicity, as a sequence Īŗj āˆˆ [Ī»2j , Ī»2j+1 ] with j āˆˆ N. Moreover, for each j, either Ī»2j < Īŗj < Ī»2j+1 or Ī»2j = Īŗj = Ī»2j+1 . Proof. It was already proved that there are no zeros of Ī” on the set where Ī” āˆˆ (āˆ’2, 2). Moreover, a zero of Ī”2 āˆ’ 4 of multiplicity m is also a zero of Ī” of multiplicity m āˆ’ 1, so Ī” has a simple zero at every closed gap and no zeros at open gap edges. It remains for us to consider zeros on C \ E. Since Ī” (z) = āˆ’2 sin Ī˜(z)Ī˜ (z), zeros of Ī” match those of Ī˜ and w . In particular, by Proposition 11.124, w is Herglotz, so there are no zeros on C \ R. By (11.157), w ā†’ 0 as z ā†’ āˆ’āˆž, and since w is increasing on (āˆ’āˆž, Ī»1 ), it has no zeros there. On each open gap (Ī»2j , Ī»2j+1 ), since Ī”(Ī»2j ) = Ī”(Ī»2j+1 ), there exists a zero of Ī” . By Proposition 11.124 and Proposition 7.56, w is strictly increasing there, and in particular, Ī˜ has at most one zero there, and it is simple.  Proposition 11.126. The function Ī˜, originally deļ¬ned on C+ , has a continuous extension to C+ . This extension obeys the following:

11.16. The periodic discriminant and the Marchenkoā€“Ostrovski map

0

Ļ€

2Ļ€

449

3Ļ€

Figure 11.3. Image of Ī˜(R) for a periodic SchrĀØ odinger operator.

(a) Im Ī˜ = 0 on E; (b) Re Ī˜ = 0 on (āˆ’āˆž, Ī»1 ]; (c) Re Ī˜ = jĻ€ on [Ī»2j , Ī»2j+1 ] for j āˆˆ N. This describes the image of Ī˜ on R as a generalized polygonal curve, with open gaps mapped to vertical line segments traversed up and then down; see Figure 11.3. Proof. It is known that Ī˜ has an analytic extension through any interval (c, d) āŠ‚ R which contains no zeros of Ī”2 āˆ’4, so Ī˜ has a continuous extension to C+ \ {Ī»j | j āˆˆ N}. Consider a zero Ī»k of Ī”2 āˆ’ 4. From the exponential Herglotz representation of w , it follows that Ī˜ (z) = O(|z āˆ’ Ī»k |āˆ’1/2 ),

z ā†’ Ī»k , z āˆˆ C+ .

As in the proof of Proposition 10.66, by the mean value theorem, this implies that lim Ī˜(z)

zā†’Ī»k zāˆˆC+

exists, and this completes the continuous extension of Ī˜ to C+ . Since w is real-valued on R \ E, it follows that Im Ī˜ is constant on (āˆ’āˆž, Ī»1 ] and on [Ī»2j , Ī»2j+1 ] for j āˆˆ N. On E, Ī” āˆˆ [āˆ’2, 2] implies Im Ī˜ = 0. Combining these conclusions shows Ī˜(Ī»2j ) = Ī˜(Ī»2j+1 ) for j āˆˆ N. āˆš Since m+ (x; z) =āˆšāˆ’ āˆ’z + o(1) for each x, it follows from the deļ¬nition of w that w(z) = āˆ’ āˆ’z + o(1) as z ā†’ āˆ’āˆž and therefore Im Ī˜(z) = 0 for z āˆˆ (āˆ’āˆž, Ī»1 ). Since integration over each band shows that  Ī»2j  Ī»2j Ī” (Ī») . Ī˜ (Ī») dĪ» = dĪ» = Ļ€, Ī˜(Ī»2j ) āˆ’ Ī˜(Ī»2jāˆ’1 ) = 4 āˆ’ Ī”(Ī»)2 Ī»2jāˆ’1 Ī»2jāˆ’1 the remaining conclusions follow by induction.



450

11. One-dimensional SchrĀØodinger operators

Corollary 11.127. The analytic extension of Ī˜ to C+ āˆŖ (Ī»2j , Ī»2j+1 ) āˆŖ Cāˆ’ obeys Ī˜(ĀÆ z ) = āˆ’Ī˜(z) + 2jĻ€. Proof. This follows from the reļ¬‚ection principle, since Re Ī˜ = jĻ€ on the interval (Ī»2j , Ī»2j+1 ). 

11.17. Direct spectral theory of periodic SchrĀØ odinger operators The potential V also determines SchrĀØ odinger operators on subintervals of R. We denote by HĀ± the SchrĀØodinger operators on the intervals (0, Ā±āˆž) with a Dirichlet boundary condition at 0, and denote by H1 the operator on (0, 1) with Dirichlet boundary conditions at both endpoints. We denote the entries of the monodromy matrix by ! t11 t12 T = . t21 t22 Since t21 (z) = u(1, z), where u denotes the Dirichlet solution, we will call zeros of t21 Dirichlet eigenvalues. The following statement about Dirichlet spectrum is mostly familiar from Section 11.3: Lemma 11.128. All zeros of t21 are simple. Moreover, for any z āˆˆ C, the following are equivalent: (a) z is a zero of t21 .  (b) 10 is an eigenvector of T (z).

(c) there is a nontrivial eigensolution of āˆ’f  + V f = zf such that f (0) = f (1) = 0.

(d) z is an eigenvalue of H1 . Proof. The equivalence of (a), (c), and (d) was proved in Section 11.3, and     (1,z) .  (a) ā‡ā‡’ (b) is elementary since T (z) 10 = uu(1,z) We now wish to compare the locations of Dirichlet eigenvalues to the periodic spectrum. The ļ¬rst step is the following. Corollary 11.129. If z is a Dirichlet eigenvalue, then z āˆˆ R and Ī”(z) āˆˆ / (āˆ’2, 2). Proof. z is real because it is an eigenvalue of the self-adjoint operator H1 . If t21 (z) = 0, then T (z) is lower triangular, so t11 (z)t22 (z) = det T (z) = 1. This implies |Ī”(z)| = |t11 (z) + 1/t11 (z)| ā‰„ 2 by the arithmetic meanā€“ geometric mean inequality. 

11.17. Direct spectral theory of periodic SchrĀØ odinger operators

To state the next result, let us ļ¬x the branch of . Ī”2 āˆ’ 4 = āˆ’2i sin Ī˜(z).

āˆš

451

Ī”2 āˆ’ 4 on C \ E by

Note that this branch is positive on (āˆ’āˆž, Ī»1 ). Theorem 11.130. The m-function for H+ is given on C+ by āˆš t11 āˆ’ t22 āˆ’ Ī”2 āˆ’ 4 . (11.158) m+ = 2t21 Moreover, the zeros of t21 can be listed as (Ī¼j )āˆž j=1 so that Ī¼j < Ī¼j+1 and Ī¼j āˆˆ [Ī»2j , Ī»2j+1 ] for all j āˆˆ N. Proof. Rewriting (11.151) projectively implies that t11 m+ + t12 m+ = . t21 m+ + t22 This can be rewritten as a quadratic equation for m+ , whose solutions are āˆš t11 āˆ’ t22 Ā± Ī”2 āˆ’ 4 . 2t21 āˆš Since Ī”2 āˆ’ 4 is nonzero on C+ , the Ā± sign must be chosen uniformly throughout C+ . We will determine this choice of sign, and the placement of zeros of t21 , based on the condition that m+ is Herglotz. On every band (Ī»2jāˆ’1 , Ī»2j ), the boundary values of Im m+ are given by 2 limā†“0 sin Ī˜(Ī» + i ) . ā†“0 2t21 (Ī») These boundary values are nonzero and have constant sign on the band interior (Ī»2jāˆ’1 , Ī»2j ). Since m+ , this sign must be positive on each band interior. Since sin Ī˜ changes sign between consecutive gaps, t21 (Ī») must also change sign, so it must have at least one zero in the gap closure [Ī»2j , Ī»2j+1 ] for each j. lim Im m+ (Ī» + i ) = āˆ“

By the counting lemma, Lemma 11.24, for large enough N , t21 has precisely N zeros smaller than (N + 1/2)2 Ļ€ 2 . Since that many zeros have already been found in the intervals [Ī»2j , Ī»2j+1 ] for j = 1, . . . , N , this shows that there is precisely one zero in each [Ī»2j , Ī»2j+1 ] and no zeros in (āˆ’āˆž, Ī»1 ]. Finally, on the band [Ī»1 , Ī»2 ], limā†“0 sin Ī˜(Ī» + i ) > 0 and t12 (Ī») > 0 because allāˆšzeros of t21 are greater than Ī». This implies the choice of sign  in front of Ī”2 āˆ’ 4 in (11.158). From m+ , spectral properties of H+ can be read oļ¬€: Theorem 11.131. The operator H+ has essential spectrum Ļƒess (H+ ) = E and discrete spectrum Ļƒd (H+ ) = {Ī¼j | j āˆˆ N, |t11 (Ī¼j )| < 1}.

452

11. One-dimensional SchrĀØodinger operators

More precisely, the spectral measure Ī¼+ is given by  Īŗ j Ī“ Ī¼j , dĪ¼+ (Ī») = w+ (Ī») dĪ» + jāˆˆN

where

āŽ§āˆš āŽØ 4āˆ’Ī”(Ī»)2 w+ (Ī») =

āŽ©0

|t12 (Ī»)|

Ī» āˆˆ (Ī»2jāˆ’1 , Ī»2j ) for some j āˆˆ N else,

and Īŗj > 0 if and only if |t22 (Ī¼j )| < 1. Similarly, māˆ’ can be found as the second solution of the quadratic equation: Proposition 11.132.

āˆš t22 āˆ’ t11 āˆ’ Ī”2 āˆ’ 4 . māˆ’ = 2t21

It follows, in particular, that mĀ± obey the reļ¬‚ectionless condition māˆ’ (Ī» + i0) = āˆ’m+ (Ī» + i0) for all Ī» in the interior of E. From this, it follows just as for periodic Jacobi matrices that: Theorem 11.133. The full-line periodic SchrĀØ odinger operator H has purely absolutely continuous spectrum on E with multiplicity 2, i.e., H āˆ¼ = TĪ»,Ļ‡E (Ī») dĪ»āŠ• TĪ»,Ļ‡E (Ī») dĪ» . Operators on (0, Ā±āˆž) and (0, 1) with Neumann boundary conditions, N and H N , can be related to the entry t . We leave as an denoted HĀ± 12 1 exercise to the reader the following facts, which follow the same ideas as above. Lemma 11.134. All zeros of t12 are simple. Moreover, for any z āˆˆ C, the following are equivalent: (a) z is a zero of t12 .  (b) 01 is an eigenvector of T (z).

(c) there is a nontrivial solution of āˆ’f  + V f = zf such that f  (0) = f  (1) = 0.

(d) z is an eigenvalue of H1N . Lemma 11.135. All zeros of t12 are real and can be listed in the form (Ī½j )āˆž j=0 where Ī½0 āˆˆ (āˆ’āˆž, Ī»1 ] and Ī½j āˆˆ [Ī»2j , Ī»2j+1 ] for j āˆˆ N; in particular, Ī½jāˆ’1 < Ī½j for all j āˆˆ N. Finally, we obtain equivalent characterizations of the open gapā€“closed gap dichotomy; see the discussion preceding Proposition 10.83 and its proof.

11.18. Exercises

453

Theorem 11.136. For Ī» āˆˆ C, the following are equivalent: (a) Ī» is a closed gap of H, i.e., Ī» = Ī»2j = Ī»2j+1 for some j āˆˆ N. (b) Ī» is a double root of Ī”2 āˆ’ 4. (c) T (Ī») āˆˆ {+I, āˆ’I}. The characterization through the geometric multiplicity has a direct spectral interpretation through a SchrĀØodinger operator on the interval (0, 2) with periodic boundary conditions (Exercise 11.28).

11.18. Exercises 11.1. Prove Lemma 11.1. 11.2. Prove that the initial value problem (11.6) has the unique solution given by (11.31). 11.3. Let V āˆˆ L1 ([0, 1]) and Ļ• āˆˆ R. Prove that the operator HĻ• , deļ¬ned by HĻ• f = āˆ’f  + V f with D(HĻ• ) = {f āˆˆ D(Hmax ) | f (1) = eiĻ• f (0) and f  (1) = eiĻ• f  (0)} is self-adjoint. These boundary conditions are called skew-periodic. The case Ļ• = 0 is called periodic; the case Ļ• = Ļ€, antiperiodic. 11.4. Let V āˆˆ L1 ([0, 1]). Besides the separated boundary conditions (11.2) and (11.3) with Ī±, Ī² āˆˆ R and the skew-periodic boundary conditions from the previous problem, are there any other self-adjoint choices of boundary conditions? 11.5. If V is a potential on R such that  x+1 Vāˆ’ (t) dt < āˆž, sup xāˆˆR

x

prove that there exists M < āˆž, which depends only on the value of this supremum, such that for all f āˆˆ Xāˆ’ āˆ© X+ ,   +āˆž  +āˆž 1 +āˆž  2 2 |f | dx ā‰¤ M |f | dx + f (āˆ’f  + V f ) dx. 2 āˆ’āˆž āˆ’āˆž āˆ’āˆž 11.6. If Ļˆ Ā± (x, z) are Weyl solutions for the SchrĀØodinger operator H, denote mĀ± (x, z) = Ā±

āˆ‚x Ļˆ Ā± (x, z) . Ļˆ Ā± (x, z)

Prove that āˆ‚x G(x, x; z) =

māˆ’ (x; z) āˆ’ m+ (x; z) . māˆ’ (x; z) + m+ (x; z)

(11.159)

454

11. One-dimensional SchrĀØodinger operators

11.7. In the setting of Theorem 11.52, prove that for all z, w āˆˆ C \ Ļƒ(H),  b m(z) āˆ’ m(w) , |Ļˆw āˆ’ Ļˆz |2 dx = B(z) + B(w) āˆ’ 2 Re zāˆ’w ĀÆ 0 where B is deļ¬ned by

m(z)āˆ’m(z) z āˆˆC\R zāˆ’ĀÆ z B(z) =  z āˆˆ R \ Ļƒ(H). m (z) Hint: Use Theorem 11.52. 11.8. Consider a SchrĀØodinger operator H on (0, b) with a regular endpoint at 0. Since its m-function m(z) was deļ¬ned as a function on C\Ļƒ(H), any Ī» āˆˆ Ļƒd (H) is an isolated singularity of m(z). For ļ¬xed Ī» āˆˆ Ļƒd (H), prove the following. (a) The Weyl solution ĻˆĪ» at z = Ī» is a multiple of Ļ†Ī» . (b) In some neighborhood of Ī», the Weyl solutions Ļˆz are linearly independent with Īøz , so they can be normalized by W (Ļˆz , Īøz ) = 1. With that normalization,  b |Ļˆz āˆ’ ĻˆĪ» |2 dx = 0. lim zā†’Ī» 0

(c) The m-function has a simple pole at Ī», and its residue is  b |ĻˆĪ» (x)|2 dx, ResĪ» m = āˆ’ 0

where ResĪ» m = limzā†’Ī» (z āˆ’ Ī»)m(z) denotes the residue of m at Ī». 2

d 11.9. Consider the operator H = āˆ’ dx 2 on the interval I = (0, +āˆž) with the boundary condition at 0

cos Ī±f (0) + sin Ī±f  (0) = 0. Find the m-function and the canonical spectral measure as functions of Ī±. 11.10. For any z, w āˆˆ C, prove that the transfer matrices TĪ± (x, z) have the property !  x 0 0 āˆ— āˆ— J āˆ’ TĪ± (x, w) J TĪ± (x, z) = āˆ’i(z āˆ’ w) TĪ± (t, w) RĪ± RĪ±āˆ— TĪ± (t, z) dt. 0 1 0 11.11. Let H be a SchrĀØodinger operator on (0, b) with a regular endpoint at 0. Let U denote its eigenfunction expansion as in Theorem 11.56. Let z āˆˆ C \ Ļƒ(H). Prove the following. Ļ†Ī» (y) . (a) For any y āˆˆ (0, b), if f (x) = āˆ‚y G(x, y; z), then (U f )(Ī») = Ī»āˆ’z (b) If the Weyl solution Ļˆz is normalized by W (Ļˆz , Ļ†z ) = 1, prove 1 . that (U Ļˆz )(Ī») = Ī»āˆ’z

11.18. Exercises

455

1 ([0, b)) and z āˆˆ C . Prove that the radius of the limit 11.12. Let V āˆˆ L + loc Weyl disk x DĪ± (x, z) is !āˆ’1  b 2 2 Im z |Ļ†Ī± (t, z)| dt . 0

and that V is a limit circle at b. Fix 11.13. Assume that V āˆˆ a boundary condition at 0. Recalling that any self-adjoint boundary āˆ— obeys v , f ) = 0, where v āˆˆ X+ \X+ condition at b is described by W+ (ĀÆ v , v) = 0, let us denote the corresponding m-function by mv (z). W+ (ĀÆ Prove that for any z āˆˆ C+ , the boundary of the limit disk DĪ± (b, z) āˆ— , W (ĀÆ is the set {mv (z) | v āˆˆ X+ \ X+ + v , v) = 0}. L1loc ([0, b))

11.14. Assume that V is regular at 0 and a limit point at b. Prove that for any h āˆˆ Cc ((0, āˆž)),  āˆž  āˆž Ī»1/2 h(Ī») lim dĪ» = h(Ī») dĪ¼Ī± (Ī»). xā†’b 0 Ļ€(Ī»Ļ†Ī± (x, Ī»)2 + Ļ†Ī± (x, Ī»)2 ) 0 This variant of Carmonaā€™s formula is useful for the study of decaying potentials, in combination with PrĀØ ufer variables [58, 66]. 11.15. In the setting of Theorem 11.78, prove the asymptotic behavior  x V (t)eāˆ’2kt dt m(z) = āˆ’k āˆ’ 0   1 x t1 āˆ’2kt1 e (1 āˆ’ eāˆ’2kt2 )V (t1 )V (t2 ) dt2 dt1 + O(|k|āˆ’2 ) + k 0 0 as z ā†’ āˆž in the appropriate sector. 11.16. Denote by m0,Ī² (z) the Weyl m-function for a SchrĀØodinger operator with a Dirichlet boundary condition at 0 and a Ī²-boundary condition at 1. The Atkinson argument proves that sup|m0,0 (z) āˆ’ m0,Ī² (z)| ā†’ 0,

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“].

Ī²āˆˆR

Is the same true in the limit z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“]? 11.17. For Ī± āˆˆ (0, Ļ€), let mĪ± denote the Weyl m-function corresponding to an Ī±-boundary condition at 0. Prove that as z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“], mĪ± (z) = cot Ī± +

1 cos Ī± āˆ’2 āˆ’1 k + o(k āˆ’2 ). + 2 k sin Ī± sin3 Ī±

11.18. Let HĪ±,Ī² denote the SchrĀØodinger operator on [0, 1] with V āˆˆ L1 ([0, 1]) and boundary conditions (11.2) and (11.3). Prove that 0 0,

m(z) = āˆ’k

n+2 

cj (V )k āˆ’j + oĖœ(|k|āˆ’nāˆ’1),

z ā†’ āˆž, arg z āˆˆ [Ī“, Ļ€ āˆ’ Ī“],

j=0

uniformly in bounded subsets of V āˆˆ C n ([0, 1]). (b) If in addition H is semibounded, i.e., inf Ļƒ(H) > āˆ’āˆž, then m(z) = āˆ’k

n+2 

cj (V )k āˆ’j + o(|k|āˆ’1 ),

z ā†’ āˆž, arg z āˆˆ [Ī“, 2Ļ€ āˆ’ Ī“].

j=0

This is uniform in bounded subsets of V āˆˆ C n ([0, 1]) with H such that inf Ļƒ(H) ā‰„ C, where C āˆˆ R. 11.21. Let HĪ±,Ī² denote the SchrĀØodinger operator on [0, 1] with V āˆˆ L1 ([0, 1]) and boundary conditions (11.2) and (11.3), and let mĪ±,Ī² denote its Weyl function. (a) Prove that m0,Ī² is uniquely determined by the spectra Ļƒ(H0,Ī² ) and Ļƒ(HĻ€/2,Ī² ) with Dirichlet and Neumann boundary conditions at 0. Hint: Use Example 7.62. (b) Combine with the Borgā€“Marchenko theorem to conclude that Ļƒ(H0,Ī² ) and Ļƒ(HĻ€/2,Ī² ) determine the potential V and Ī² uniquely. 11.22. Prove that each coeļ¬ƒcient cn (x) described in Theorem 11.84 is a polynomial in V (x), . . . , V (nāˆ’2) (x). If we deļ¬ne a notion of degree B B (j) = 2 + j and that deg for diļ¬€erential polynomials of V so that degV is multiplicative (informally speaking, every V counts as 2, and every derivative counts as 1), prove that every monomial in cn has degree exactly 2n.

11.18. Exercises

457

11.23. Let V āˆˆ L1 ([0, āˆž)) (this is not a typo; V is assumed integrable on the entire half-line, which is a kind of decay condition at +āˆž). Denote by T (x, z; V ) the corresponding transfer matrices. (a) Derive a ļ¬rst-order ordinary diļ¬€erential equation for S(x, z; V ) = T (x, z; 0)āˆ’1 T (x, z; V ). (b) Use it to prove that for all Ī» āˆˆ (0, āˆž), there is a convergent limit lim S(x, z; V ).

xā†’āˆž

(c) Conclude that for Ī» āˆˆ (0, āˆž), all eigensolutions are bounded and that the maximal spectral measure on (0, āˆž) is mutually absolutely continuous with Lebesgue measure. 11.24. Let H be a SchrĀØodinger operator on (0, b) with a regular endpoint at 0 and Ļ†Ī» deļ¬ned, as usual, as a nontrivial eigensolution at Ī» which obeys the boundary condition at 0. If for some Ī» āˆˆ C, Ļ†Ī» (x) grows at most subexponentially, i.e., for all > 0, Ļ†Ī» (x) = O(ex ) as x ā†’ āˆž, prove that Ī» āˆˆ Ļƒ(H). 11.25. If for some Ī» āˆˆ R, there exists a subexponentially growing eigensolution, i.e., a nontrivial eigensolution u(x) such that for every > 0, u(x) = O(ex ) as x ā†’ Ā±āˆž, prove that Ī» āˆˆ Ļƒ(H). 11.26. If H is a SchrĀØodinger operator with limit circle endpoints, prove that it has compact resolvent; in particular, Ļƒess (H) = āˆ…. Hint: Use the  behavior of eigensolutions at limit circle endpoints to prove that IƗI |G(x, y; z)|2 dx dy < āˆž. 11.27. If the SchrĀØodinger operator H on (āˆ’ , + ) is a limit circle at āˆ’ , prove that it has simple (multiplicity 1) spectrum. Hint: Use a Weyl matrix with respect to some internal point c āˆˆ (āˆ’ , + ) and apply the previous exercise to a SchrĀØodinger operator on (āˆ’ , c). 11.28. For a 1-periodic potential V , let Ī” denote the discriminant, and let H2P denote the corresponding SchrĀØodinger operator on the interval (0, 2) with periodic boundary conditions f (2) = f (0), f  (2) = f  (0). Prove that Ļƒ(H2P ) = {Ī» | Ī”(Ī»)2 āˆ’ 4 = 0} and that for Ī» āˆˆ Ļƒ(H2P ), dim Ker(H2P āˆ’ Ī») is equal to the multiplicity of Ī» as a root of Ī”2 āˆ’ 4.

Bibliography

[1] M. Aizenman and S. Warzel, Random operators: Disorder eļ¬€ects on quantum spectra and dynamics, Graduate Studies in Mathematics, vol. 168, American Mathematical Society, Providence, RI, 2015, DOI 10.1090/gsm/168. MR3364516 [2] N. I. Akhiezer, The classical moment problem and some related questions in analysis, Hafner Publishing Co., New York, 1965. Translated by N. Kemmer. MR0184042 [3] W. O. Amrein and V. Georgescu, On the characterization of bound states and scattering states in quantum mechanics, Helv. Phys. Acta 46 (1973/74), 635ā€“658. MR363267 [4] D. H. Armitage and S. J. Gardiner, Classical potential theory, Springer Monographs in Mathematics, Springer-Verlag London, Ltd., London, 2001, DOI 10.1007/978-1-4471-02335. MR1801253 [5] F. V. Atkinson, On the location of the Weyl circles, Proc. Roy. Soc. Edinburgh Sect. A 88 (1981), no. 3-4, 345ā€“356, DOI 10.1017/S0308210500020163. MR616784 [6] H. Behncke, Absolute continuity of Hamiltonians with von Neumann Wigner potentials. II, Manuscripta Math. 71 (1991), no. 2, 163ā€“181, DOI 10.1007/BF02568400. MR1101267 [7] C. Bennewitz, A proof of the local Borg-Marchenko theorem, Comm. Math. Phys. 218 (2001), no. 1, 131ā€“132, DOI 10.1007/s002200100384. MR1824201 [8] C. Bennewitz, M. Brown, and R. Weikard, Spectral and scattering theory for ordinary differential equations. Vol. I: Sturm-Liouville equations, Universitext, Springer, Cham, 2020, DOI 10.1007/978-3-030-59088-8. MR4199125 [9] G. Berkolaiko and P. Kuchment, Introduction to quantum graphs, Mathematical Surveys and Monographs, vol. 186, American Mathematical Society, Providence, RI, 2013, DOI 10.1090/surv/186. MR3013208 [10] A. S. Besicovitch, On existence of subsets of ļ¬nite measure of sets of inļ¬nite measure, Nederl. Akad. Wetensch. Proc. Ser. A. 55 = Indagationes Math. 14 (1952), 339ā€“344. MR0048540 [11] R. Bessonov, M. LukiĀ“ c, and P. Yuditskii, Reļ¬‚ectionless canonical systems, I: Arov gauge and right limits, Integral Equations Operator Theory 94 (2022), no. 1, Paper No. 4, 30, DOI 10.1007/s00020-021-02683-z. MR4360428 [12] M. Sh. Birman and M. Z. Solomjak, Spectral theory of selfadjoint operators in Hilbert space, Mathematics and its Applications (Soviet Series), D. Reidel Publishing Co., Dordrecht, 1987. Translated from the 1980 Russian original by S. KhrushchĀØ ev and V. Peller. MR1192782

459

460

Bibliography

[13] G. Borg, Eine Umkehrung der Sturm-Liouvilleschen Eigenwertaufgabe. Bestimmung der Diļ¬€erentialgleichung durch die Eigenwerte (German), Acta Math. 78 (1946), 1ā€“96, DOI 10.1007/BF02421600. MR15185 [14] G. Borg, Uniqueness theorems in the spectral theory of y  + (Ī» āˆ’ q(x))y = 0, Den 11te Skandinaviske Matematikerkongress, Trondheim, 1949, Johan Grundt Tanums Forlag, Oslo, 1952, pp. 276ā€“287. MR0058063 [15] D. Borthwick, Spectral theory: Basic concepts and applications, Graduate Texts in c Mathematics, vol. 284, Springer, Cham, [2020] 2020, DOI 10.1007/978-3-030-38002-1. MR4180682 [16] M. J. Cantero, L. Moral, and L. VelĀ“ azquez, Five-diagonal matrices and zeros of orthogonal polynomials on the unit circle, Linear Algebra Appl. 362 (2003), 29ā€“56, DOI 10.1016/S00243795(02)00457-3. MR1955452 [17] R. Carmona and J. Lacroix, Spectral theory of random SchrĀØ odinger operators, Probability and its Applications, BirkhĀØ auser Boston, Inc., Boston, MA, 1990, DOI 10.1007/978-1-46124488-2. MR1102675 [18] T. S. Chihara, An introduction to orthogonal polynomials, Mathematics and its Applications, Vol. 13, Gordon and Breach Science Publishers, New York-London-Paris, 1978. MR0481884 [19] H. L. Cycon, R. G. Froese, W. Kirsch, and B. Simon, SchrĀØ odinger operators with application to quantum mechanics and global geometry, Springer Study Edition, Texts and Monographs in Physics, Springer-Verlag, Berlin, 1987. MR883643 [20] D. Damanik, SchrĀØ odinger operators with dynamically deļ¬ned potentials, Ergodic Theory Dynam. Systems 37 (2017), no. 6, 1681ā€“1764, DOI 10.1017/etds.2015.120. MR3681983 [21] D. Damanik and J. Fillman, One-dimensional ergodic SchrĀØ odinger operators. I. General theory, Graduate Series in Mathematics, vol. 221, American Mathematical Society, Providence, RI, 2022. [22] D, Damanik and J. Fillman, One-dimensional ergodic SchrĀØ odinger operators. II. Speciļ¬c classes, in preparation. [23] R. O. Davies, Subsets of ļ¬nite measure in analytic sets, Nederl. Akad. Wetensch. Proc. Ser. A. 55 = Indagationes Math. 14 (1952), 488ā€“489. MR0053184 [24] L. de Branges, Hilbert spaces of entire functions, Prentice-Hall, Inc., Englewood Cliļ¬€s, N.J., 1968. MR0229011 [25] P. A. Deift, Orthogonal polynomials and random matrices: A Riemann-Hilbert approach, Courant Lecture Notes in Mathematics, vol. 3, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 1999. MR1677884 [26] R. del Rio, S. Jitomirskaya, Y. Last, and B. Simon, Operators with singular continuous spectrum. IV. Hausdorļ¬€ dimensions, rank one perturbations, and localization, J. Anal. Math. 69 (1996), 153ā€“200, DOI 10.1007/BF02787106. MR1428099 [27] S. A. Denisov and A. Kiselev, Spectral properties of SchrĀØ odinger operators with decaying potentials, Spectral theory and mathematical physics: a Festschrift in honor of Barry Simonā€™s 60th birthday, Proc. Sympos. Pure Math., vol. 76, Amer. Math. Soc., Providence, RI, 2007, pp. 565ā€“589, DOI 10.1090/pspum/076.2/2307748. MR2307748 [28] J. Eckhardt, F. Gesztesy, R. Nichols, and G. Teschl, Weyl-Titchmarsh theory for SturmLiouville operators with distributional potentials, Opuscula Math. 33 (2013), no. 3, 467ā€“563, DOI 10.7494/OpMath.2013.33.3.467. MR3046408 [29] V. Enss, Asymptotic completeness for quantum mechanical potential scattering. I. Short range potentials, Comm. Math. Phys. 61 (1978), no. 3, 285ā€“291. MR523013 [30] A. Eremenko and P. Yuditskii, Comb functions, Recent advances in orthogonal polynomials, special functions, and their applications, Contemp. Math., vol. 578, Amer. Math. Soc., Providence, RI, 2012, pp. 99ā€“118, DOI 10.1090/conm/578/11472. MR2964141

Bibliography

461

[31] K. J. Falconer, The geometry of fractal sets, Cambridge Tracts in Mathematics, vol. 85, Cambridge University Press, Cambridge, 1986. MR867284 [32] G. B. Folland, Real analysis: Modern techniques and their applications, Pure and Applied Mathematics (New York), John Wiley & Sons, Inc., New York, 1984. A Wiley-Interscience Publication. MR767633 [33] F. Gesztesy, B. Simon, and G. Teschl, Zeros of the Wronskian and renormalized oscillation theory, Amer. J. Math. 118 (1996), no. 3, 571ā€“594. MR1393260 [34] F. Gesztesy, Inverse spectral theory as inļ¬‚uenced by Barry Simon, Spectral theory and mathematical physics: a Festschrift in honor of Barry Simonā€™s 60th birthday, Proc. Sympos. Pure Math., vol. 76, Amer. Math. Soc., Providence, RI, 2007, pp. 741ā€“820, DOI 10.1090/pspum/076.2/2307754. MR2307754 [35] F. Gesztesy and H. Holden, Soliton equations and their algebro-geometric solutions. Vol. I: (1 + 1)-dimensional continuous models, Cambridge Studies in Advanced Mathematics, vol. 79, Cambridge University Press, Cambridge, 2003, DOI 10.1017/CBO9780511546723. MR1992536 [36] F. Gesztesy, H. Holden, J. Michor, and G. Teschl, Soliton equations and their algebrogeometric solutions. Vol. II: (1 + 1)-dimensional discrete models, Cambridge Studies in Advanced Mathematics, vol. 114, Cambridge University Press, Cambridge, 2008, DOI 10.1017/CBO9780511543203. MR2446594 [37] F. Gesztesy and B. Simon, On local Borg-Marchenko uniqueness results, Comm. Math. Phys. 211 (2000), no. 2, 273ā€“287, DOI 10.1007/s002200050812. MR1754515 [38] F. Gesztesy and E. Tsekanovskii, On matrix-valued Herglotz functions, Math. Nachr. 218 (2000), 61ā€“138, DOI 10.1002/1522-2616(200010)218:161::AID-MANA613.3.CO;2-4. MR1784638 [39] F. Gesztesy and M. Zinchenko, On spectral theory for SchrĀØ odinger operators with strongly singular potentials, Math. Nachr. 279 (2006), no. 9-10, 1041ā€“1082, DOI 10.1002/mana.200510410. MR2242965 [40] D. J. Gilbert, On subordinacy and analysis of the spectrum of SchrĀØ odinger operators with two singular endpoints, Proc. Roy. Soc. Edinburgh Sect. A 112 (1989), no. 3-4, 213ā€“229, DOI 10.1017/S0308210500018680. MR1014651 [41] D. J. Gilbert and D. B. Pearson, On subordinacy and analysis of the spectrum of onedimensional SchrĀØ odinger operators, J. Math. Anal. Appl. 128 (1987), no. 1, 30ā€“56, DOI 10.1016/0022-247X(87)90212-5. MR915965 [42] M. Harmer, Hermitian symplectic geometry and extension theory, J. Phys. A 33 (2000), no. 50, 9193ā€“9203, DOI 10.1088/0305-4470/33/50/305. MR1804888 [43] P. D. Hislop and I. M. Sigal, Introduction to spectral theory: With applications to SchrĀØ odinger operators, Applied Mathematical Sciences, vol. 113, Springer-Verlag, New York, 1996, DOI 10.1007/978-1-4612-0741-2. MR1361167 [44] R. O. Hryniv and Ya. V. Mykytyuk, 1-D SchrĀØ odinger operators with periodic singular potentials, Methods Funct. Anal. Topology 7 (2001), no. 4, 31ā€“42. MR1879483 [45] D. Hundertmark, Some bound state problems in quantum mechanics, Spectral theory and mathematical physics: a Festschrift in honor of Barry Simonā€™s 60th birthday, Proc. Sympos. Pure Math., vol. 76, Amer. Math. Soc., Providence, RI, 2007, pp. 463ā€“496, DOI 10.1090/pspum/076.1/2310215. MR2310215 [46] M. E. H. Ismail, Classical and quantum orthogonal polynomials in one variable, Encyclopedia of Mathematics and its Applications, vol. 98, Cambridge University Press, Cambridge, 2009. With two chapters by Walter Van Assche; With a foreword by Richard A. Askey; Reprint of the 2005 original. MR2542683 [47] S. Jitomirskaya and Y. Last, Power-law subordinacy and singular spectra. I. Half-line operators, Acta Math. 183 (1999), no. 2, 171ā€“189, DOI 10.1007/BF02392827. MR1738043

462

Bibliography

[48] S. Ya. Jitomirskaya and Y. Last, Power law subordinacy and singular spectra. II. Line operators, Comm. Math. Phys. 211 (2000), no. 3, 643ā€“658, DOI 10.1007/s002200050830. MR1773812 [49] R. Johnson and J. Moser, The rotation number for almost periodic potentials, Comm. Math. Phys. 84 (1982), no. 3, 403ā€“438. MR667409 [50] R. Johnson and J. Moser, Erratum: ā€œThe rotation number for almost periodic potentialsā€ [Comm. Math. Phys. 84 (1982), no. 3, 403ā€“438; MR0667409 (83h:34018)], Comm. Math. Phys. 90 (1983), no. 2, 317ā€“318. MR714441 [51] I. S. Kac, On the spectral multiplicity of a second-order diļ¬€erential operator. (Russian), Dokl. Akad. Nauk SSSR 145 (1962), 510ā€“513. MR0145375 [52] I. S. Kac, Spectral multiplicity of a second-order diļ¬€erential operator and expansion in eigenfunction (Russian), Izv. Akad. Nauk SSSR Ser. Mat. 27 (1963), 1081ā€“1112. MR0159982 [53] T. Kappeler and J. PĀØ oschel, KdV & KAM, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 45, SpringerVerlag, Berlin, 2003, DOI 10.1007/978-3-662-08054-2. MR1997070 [54] S. Khan and D. B. Pearson, Subordinacy and spectral theory for inļ¬nite matrices, Helv. Phys. Acta 65 (1992), no. 4, 505ā€“527. MR1179528 [55] S. Khan and D. B. Pearson, Subordinacy and spectral theory for inļ¬nite matrices, Helv. Phys. Acta 65 (1992), no. 4, 505ā€“527. MR1179528 [56] R. Killip, Spectral theory via sum rules, Spectral theory and mathematical physics: a Festschrift in honor of Barry Simonā€™s 60th birthday, Proc. Sympos. Pure Math., vol. 76, Amer. Math. Soc., Providence, RI, 2007, pp. 907ā€“930, DOI 10.1090/pspum/076.2/2310217. MR2310217 [57] W. Kirsch, An invitation to random SchrĀØ odinger operators (English, with English and French summaries), Random SchrĀØ odinger operators, Panor. Synth` eses, vol. 25, Soc. Math. France, Paris, 2008, pp. 1ā€“119. With an appendix by FrĀ“ edĀ“ eric Klopp. MR2509110 [58] A. Kiselev, Y. Last, and B. Simon, Modiļ¬ed PrĀØ ufer and EFGP transforms and the spectral analysis of one-dimensional SchrĀØ odinger operators, Comm. Math. Phys. 194 (1998), no. 1, 1ā€“45, DOI 10.1007/s002200050346. MR1628290 [59] D. G. Larman, Subsets of given Hausdorļ¬€ measure in connected spaces, Quart. J. Math. Oxford Ser. (2) 17 (1966), 239ā€“243, DOI 10.1093/qmath/17.1.239. MR201597 [60] Y. Last, Quantum dynamics and decompositions of singular continuous spectra, J. Funct. Anal. 142 (1996), no. 2, 406ā€“445, DOI 10.1006/jfan.1996.0155. MR1423040 [61] Y. Last and B. Simon, Eigenfunctions, transfer matrices, and absolutely continuous spectrum of one-dimensional SchrĀØ odinger operators, Invent. Math. 135 (1999), no. 2, 329ā€“367, DOI 10.1007/s002220050288. MR1666767 [62] Y. Last and B. Simon, The essential spectrum of SchrĀØ odinger, Jacobi, and CMV operators, J. Anal. Math. 98 (2006), 183ā€“220, DOI 10.1007/BF02790275. MR2254485 [63] N. Levinson, The inverse Sturm-Liouville problem, Mat. Tidsskr. B 1949 (1949), 25ā€“30. MR32067 [64] B. M. Levitan and I. S. Sargsjan, Introduction to spectral theory: selfadjoint ordinary differential operators, Translations of Mathematical Monographs, Vol. 39, American Mathematical Society, Providence, R.I., 1975. Translated from the Russian by Amiel Feinstein. MR0369797 [65] B. M. Levitan and I. S. Sargsjan, Sturm-Liouville and Dirac operators, Mathematics and its Applications (Soviet Series), vol. 59, Kluwer Academic Publishers Group, Dordrecht, 1991. Translated from the Russian, DOI 10.1007/978-94-011-3748-5. MR1136037 [66] M. Lukic, SchrĀØ odinger operators with slowly decaying Wigner-von Neumann type potentials, J. Spectr. Theory 3 (2013), no. 2, 147ā€“169, DOI 10.4171/JST/41. MR3042763

Bibliography

463

[67] V. A. Marchenko, Sturm-Liouville operators and applications, Operator Theory: Advances and Applications, vol. 22, BirkhĀØ auser Verlag, Basel, 1986. Translated from the Russian by A. Iacob, DOI 10.1007/978-3-0348-5485-6. MR897106 [68] V. A. MarĖ‡ cenko, Some questions of the theory of one-dimensional linear diļ¬€erential operators of the second order. I (Russian), Trudy Moskov. Mat. ObĖ‡sĖ‡ c. 1 (1952), 327ā€“420. MR0058064 [69] C. A. Marx and S. Jitomirskaya, Dynamics and spectral theory of quasi-periodic SchrĀØ odinger-type operators, Ergodic Theory Dynam. Systems 37 (2017), no. 8, 2353ā€“2393, DOI 10.1017/etds.2016.16. MR3719264 [70] L. Pastur and A. Figotin, Spectra of random and almost-periodic operators, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 297, Springer-Verlag, Berlin, 1992, DOI 10.1007/978-3-642-74346-7. MR1223779 [71] J. PĀØ oschel and E. Trubowitz, Inverse spectral theory, Pure and Applied Mathematics, vol. 130, Academic Press, Inc., Boston, MA, 1987. MR894477 [72] T. Ransford, Potential theory in the complex plane, London Mathematical Society Student Texts, vol. 28, Cambridge University Press, Cambridge, 1995, DOI 10.1017/CBO9780511623776. MR1334766 [73] M. Reed and B. Simon, Methods of modern mathematical physics. II. Fourier analysis, selfadjointness, Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1975. MR0493420 [74] H. Reed and B. Simon, Methods of modern mathematical physics. IV. Analysis of operators, Academic Press [Harcourt Brace Jovanovich, Publishers], New Yorkā€“London, 1978. MR0493421 [75] H. Reed and B. Simon, Methods of modern mathematical physics. III, Academic Press [Harcourt Brace Jovanovich, Publishers], New Yorkā€“London, 1979, Scattering theory. MR529429 [76] M. Reed and B. Simon, Methods of modern mathematical physics. I: Functional analysis, 2nd ed., Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1980. MR751959 [77] C. Remling, The absolutely continuous spectrum of Jacobi matrices, Ann. of Math. (2) 174 (2011), no. 1, 125ā€“171, DOI 10.4007/annals.2011.174.1.4. MR2811596 [78] C. Remling, Spectral theory of canonical systems, De Gruyter Studies in Mathematics, vol. 70, De Gruyter, Berlin, 2018. MR3890099 [79] C. Remling and K. Scarbrough, Oscillation theory and semibounded canonical systems, J. Spectr. Theory 10 (2020), no. 4, 1333ā€“1359, DOI 10.4171/jst/329. MR4192754 [80] R. Romanov, Canonical systems and de branges spaces, arXiv:1408.6022, 2014. [81] W. Rudin, Real and complex analysis, 3rd ed., McGraw-Hill Book Co., New York, 1987. MR924157 [82] D. Ruelle, A remark on bound states in potential-scattering theory (English, with Italian summary), Nuovo Cimento A (10) 61 (1969), 655ā€“662. MR246603 [83] A. M. Savchuk and A. A. Shkalikov, Sturm-Liouville operators with singular potentials (Russian, with Russian summary), Mat. Zametki 66 (1999), no. 6, 897ā€“912, DOI 10.1007/BF02674332; English transl., Math. Notes 66 (1999), no. 5-6, 741ā€“753 (2000). MR1756602 [84] J. H. Shapiro, Volterra adventures, Student Mathematical Library, vol. 85, American Mathematical Society, Providence, RI, 2018, DOI 10.1090/stml/085. MR3793153 [85] B. Simon, Bounded eigenfunctions and absolutely continuous spectra for one-dimensional SchrĀØ odinger operators, Proc. Amer. Math. Soc. 124 (1996), no. 11, 3361ā€“3369, DOI 10.1090/S0002-9939-96-03599-X. MR1350963 [86] B. Simon, A new approach to inverse spectral theory. I. Fundamental formalism, Ann. of Math. (2) 150 (1999), no. 3, 1029ā€“1057, DOI 10.2307/121061. MR1740987

464

Bibliography

[87] B. Simon, On a theorem of Kac and Gilbert, J. Funct. Anal. 223 (2005), no. 1, 109ā€“115, DOI 10.1016/j.jfa.2004.08.015. MR2139882 [88] B. Simon, Orthogonal polynomials on the unit circle. Part 1: Classical theory, American Mathematical Society Colloquium Publications, vol. 54, American Mathematical Society, Providence, RI, 2005, DOI 10.1090/coll054.1. MR2105088 [89] B. Simon, Orthogonal polynomials on the unit circle. Part 2: Spectral theory, American Mathematical Society Colloquium Publications, vol. 54, American Mathematical Society, Providence, RI, 2005, DOI 10.1090/coll/054.2/01. MR2105089 [90] B. Simon, Sturm oscillation and comparison theorems, Sturm-Liouville theory, BirkhĀØ auser, Basel, 2005, pp. 29ā€“43. MR2145076 [91] B. Simon, Trace ideals and their applications, 2nd ed., Mathematical Surveys and Monographs, vol. 120, American Mathematical Society, Providence, RI, 2005, DOI 10.1090/surv/120. MR2154153 [92] B. Simon, SzegĖ oā€™s theorem and its descendants: Spectral theory for L2 perturbations of orthogonal polynomials, M. B. Porter Lectures, Princeton University Press, Princeton, NJ, 2011. MR2743058 [93] B. Simon, Advanced complex analysis, A Comprehensive Course in Analysis, Part 2B, American Mathematical Society, Providence, RI, 2015, DOI 10.1090/simon/002.2. MR3364090 [94] B. Simon, Basic complex analysis, A Comprehensive Course in Analysis, Part 2A, American Mathematical Society, Providence, RI, 2015, DOI 10.1090/simon/002.1. MR3443339 [95] B. Simon, Harmonic analysis, A Comprehensive Course in Analysis, Part 3, American Mathematical Society, Providence, RI, 2015, DOI 10.1090/simon/003. MR3410783 [96] B. Simon, Operator theory, A Comprehensive Course in Analysis, Part 4, American Mathematical Society, Providence, RI, 2015, DOI 10.1090/simon/004. MR3364494 [97] B. Simon, Real analysis, A Comprehensive Course in Analysis, Part 1, American Mathematical Society, Providence, RI, 2015. With a 68 page companion booklet, DOI 10.1090/simon/001. MR3408971 [98] H. Stahl and V. Totik, General orthogonal polynomials, Encyclopedia of Mathematics and its Applications, vol. 43, Cambridge University Press, Cambridge, 1992, DOI 10.1017/CBO9780511759420. MR1163828 [99] E. M. Stein and R. Shakarchi, Real analysis: Measure theory, integration, and Hilbert spaces, Princeton Lectures in Analysis, vol. 3, Princeton University Press, Princeton, NJ, 2005. MR2129625 [100] T.-J. Stieltjes, Recherches sur les fractions continues (French), Ann. Fac. Sci. Toulouse Sci. Math. Sci. Phys. 8 (1894), no. 4, J1ā€“J122. MR1508159 [101] T. J. Stieltjes, Recherches sur les fractions continues [Suite et ļ¬n], Ann. Fac. Sci. Toulouse Sci. Math. Sci. Phys. 9 (1895), no. 1, A5ā€“A47. MR1508160 [102] P. Stollmann, Caught by disorder: Bound states in random media, Progress in Mathematical Physics, vol. 20, BirkhĀØ auser Boston, Inc., Boston, MA, 2001, DOI 10.1007/978-1-4612-01694. MR1935594 [103] G. Stolz, Bounded solutions and absolute continuity of Sturm-Liouville operators, J. Math. Anal. Appl. 169 (1992), no. 1, 210ā€“228, DOI 10.1016/0022-247X(92)90112-Q. MR1180682 [104] M. H. Stone, On one-parameter unitary groups in Hilbert space, Ann. of Math. (2) 33 (1932), no. 3, 643ā€“648, DOI 10.2307/1968538. MR1503079 [105] M. H. Stone, Linear transformations in Hilbert space, American Mathematical Society Colloquium Publications, vol. 15, American Mathematical Society, Providence, RI, 1990. Reprint of the 1932 original, DOI 10.1090/coll/015. MR1451877 [106] G. SzegĀØ o, Orthogonal Polynomials, American Mathematical Society Colloquium Publications, Vol. 23, American Mathematical Society, New York, 1939. MR0000077

Bibliography

465

[107] G. Teschl, Jacobi operators and completely integrable nonlinear lattices, Mathematical Surveys and Monographs, vol. 72, American Mathematical Society, Providence, RI, 2000, DOI 10.1090/surv/072. MR1711536 [108] G. Teschl, Mathematical methods in quantum mechanics: With applications to SchrĀØ odinger operators, Graduate Studies in Mathematics, vol. 99, American Mathematical Society, Providence, RI, 2009, DOI 10.1090/gsm/099. MR2499016 [109] E. C. Titchmarsh, Eigenfunction expansions associated with second-order diļ¬€erential equations. Part I, 2nd ed., Clarendon Press, Oxford, 1962. MR0176151 [110] J. Weidmann, Spectral theory of ordinary diļ¬€erential operators, Lecture Notes in Mathematics, vol. 1258, Springer-Verlag, Berlin, 1987, DOI 10.1007/BFb0077960. MR923320 [111] A. Zettl, Sturm-Liouville theory, Mathematical Surveys and Monographs, vol. 121, American Mathematical Society, Providence, RI, 2005, DOI 10.1090/surv/121. MR2170950

Notation Index

#A, the number of elements of set A, 6 , absolute continuity of one measure with respect to another, 162 āŠ„, mutual singularity of measures, 162 Ā·L1 , 433 loc,unif

|||k||| = max(1, |k|), 364 AC([a, b]), set of absolutely continuous functions on [a, b], 247 AC2loc (I) = {f āˆˆ ACloc (I) | f  āˆˆ ACloc (I)}, 379 ACloc (I), set of locally absolutely continuous functions on I, 250 Ac = X \ A, the complement of A in space X, 2 BX , the Borel Ļƒ-algebra on X, 3 Bb (X), the algebra of bounded Borel functions from X to C, 34

D(A), domain of unbounded operator A, 227 Ī“x , the Dirac measure at x, 6 Ī”, discriminant of a periodic SchrĀØ odinger operator, 444 D(x, z), Weyl disk, 409 F , Fourier transform on L2 (R), 292 obius transformation induced by fA , MĀØ matrix A, 184 fĖ†, eigenfunction expansion of f , 400, 427 fĖ†, Fourier transform of f , 291 gĖ‡, inverse eigenfunction expansion of g, 400, 427 gĖ‡, inverse Fourier transform of g, 291 G(x, y; z), Greenā€™s function, 391

C(K) = C(K, C), the space of continuous maps K ā†’ C, 48 C(K, R), the space of continuous maps K ā†’ R, 50 C0 (R), the set of continuous decaying functions on R, 195 CA (Ļˆ), cyclic subspace of vector Ļˆ, 141 Ė† = C āˆŖ {āˆž}, the Riemann sphere, 184 C C+ , the upper half-plane, 183 C[x], algebra of polynomials with complex coeļ¬ƒcients, 118

Hac , the absolutely continuous subspace for A, 274 HĪ±c , the Ī±-continuous subspace for A, 275 HĪ±s , the Ī±-singular subspace for A, 275 Hcont , the continuous subspace for A, 273 hĀ± , the positive and negative parts of a function, 21 Hpp , the pure point subspace for A, 272 Hsc , the singular continuous subspace for A, 274 Hs , the singular subspace for A, 274

D, the unit disk in C, 184

Ker, the kernel of an operator, 60

467

468

L(X, Y ), the set of bounded linear operators from X to Y , 59 Lp (X, Ī¼), 54 Lp ([a, b]), Lp space on [a, b] with respect to Lebesgue measure, 247 p (X), Lp -space with counting measure on X, 58 Lpc (I), set of compactly supported functions in Lp (I), 400 p Lloc (I), set of locally Lp functions on I, 250 m(x, z), 421 mĀ± (x, z), 422 Ī¼Ī± , a Lebesgueā€“Stieltjes measure, 26 Ī¼ āŠ— Ī½, product measure, 31 M (z), Weyl M -matrix, 219, 326, 426 oĖœ( ) asymptotic notation, 415 P(X), the set of subsets of X, 1 ĻˆzĀ± (x) = Ļˆ Ā± (x, z), Weyl solution, 390 Ran, the range of an operator, 60 RanĪ¼ g, essential range of g with respect to Ī¼, 144 Ė† = R āˆŖ {āˆ’āˆž, +āˆž}, the extended real R line, 13 Ī˜(z), Marchenkoā€“Ostrovski map, 338, 445 TX , the metric topology on metric space X, 3 W (f, g), Wronskian, 381 WĀ± (f, g), endpoint Wronskians, 381 odinger XĀ± , endpoint domains for SchrĀØ operators, 380 āˆ— , null subspaces of endpoint XĀ± domains, 381 YĀ± , endpoint domains for self-adjoint SchrĀØ odinger operators, 388

Notation Index

Index

Baire measure, 36 Banach space, 46 Banachā€“Alaoglu theorem, 65 Banachā€“Steinhaus theorem, 63 base of a metric/topological space, 10 Besselā€™s inequality, 86, 93 Borel Ļƒ-algebra, 3 Borel function, 3 Borel functional calculus, 149, 158, 243 Borel measure, 6 Borel set, 3 Borgā€™s theorem, 356 Borgā€“Marchenko theorem, 423 local, 423 bounded linear functional, 84 bounded operator, 64

Cauchyā€“Schwarz inequality, 79 Cayley transform, 185 closed graph theorem, 231 closure of an operator, 229 coeļ¬ƒcient stripping, 311 Combesā€“Thomas estimate, 334, 440 compact resolvent operator, 264 completion of a Banach space, 75 of a Hilbert space, 105 complex measure, 160 continuous spectrum, 274 convergence norm resolvent, 235 strong operator, 111, 122, 135 strong resolvent, 235 strong, in Hilbert space, 97 weak, 123 weak operator, 111 weak, in Hilbert space, 97, 99 weak-āˆ—, 65, 67, 97 counting lemma, 378, 446 counting measure, 6 cover of a set, 8 Croftā€“Garsia covering lemma, 166 cyclic subspace, 141

C āˆ— algebra, 110 CarathĀ“eodory inequality, 223 CarathĀ“eodoryā€™s theorem, 9, 40 Carmonaā€™s theorem, 318, 413 Cauchyā€™s integral formula, Banach-space valued, 71

Dirac measure, 6 direct sum of bounded operators, 120 of Hilbert spaces, 88, 89 of operators, 146 of subspaces of a Hilbert space, 90

absolutely continuous function, 247 absolutely continuous spectrum, 274 adjoint of bounded operator, 108 of direct sum of operators, 120 of matrix, 108 of unbounded operator, 230 algebra of sets, 4 Arzel` aā€“Ascoli theorem, 49

469

470

of unbounded operators, 237 of unitary operators, 120 Dirichlet eigenvalue, 347, 450 discrete spectrum, 278 discriminant of a periodic Jacobi matrix, 337 of periodic SchrĀØ odinger operator, 444 distribution function of measure, 24 dominated convergence theorem, 23 dual space, 64, 65 eigenfunction expansion, 400 for full-line Jacobi matrix, 324 eigensolution, 309 eigenvalue, 114 eigenvector, 114, 136, 141 equicontinuity, 49 essential range, 144 essential spectrum, 278 preservation under compact perturbations, 280 exhaustion by compact sets, 32 exponential Herglotz representation, 213 Fatouā€™s lemma, 19 ļ¬rst resolvent identity, 232 Floquet solutions, 352 Fourier series, 93, 105 Fourier transform, 292 Fubiniā€™s theorem, 32 function convex, 75 lower semicontinuous, 42 upper semicontinuous, 42 fundamental solution, 367 Gramā€“Schmidt process, 94 graph of an operator, 228 Greenā€™s function, 391 for a Jacobi matrix, 321 of SchrĀØ odinger operator, 375 HĀØ olderā€™s inequality, 55 Hausdorļ¬€ dimension, 171 Hausdorļ¬€ distance, 135 Hausdorļ¬€ measure, 169 Heaviside function, 392 Hellingerā€“Toeplitz theorem, 235 Herglotz function, 183 Herglotz representation, 194 Hilbert space, 80

Index

inner product, 78 on 2 (N), 80 on Cn , 80 on L2 (X, dĪ¼), 80 on quotient Hilbert space, 104 integrable function, 22 J -contracting matrix, 187 J -expanding matrix, 187 Jacobi matrix, 299 Jacobi recursion, 309 kernel, 60, 114 of integral operator, 125 Lagrangian subspace, 254 Lebesgue decomposition, 165 Lebesgue measure, 28 Lebesgueā€“Stieltjes measure, 26 limit circle, 382 limit point, 382 linear functional bounded, 64 linear relation, 228 Liouvilleā€™s theorem, Banach-space valued, 73 locally compact metric space, 43 Lyapunov exponent, 346 m-function of a Jacobi matrix, 300 Marchenkoā€“Ostrovski map, 338, 445 matrix-valued measure, 176 measurable function, 3 measure, 6 Ī±-continuous, 172 Ī±-singular, 172 absolutely continuous, 163, 165 almost Ī±-singular, 174 complex, 178 continuous, 160 continuous with respect to another measure, 162 pure point, 160 singular continuous, 165 singular with respect to another measure, 162 strongly Ī±-continuous, 174 measure class, 268 metric space discrete, 3 metric topology, 3 min-max principle, 280

Index

monodromy matrix of a periodic SchrĀØ odinger operator, 444 monotone class, 4 monotone class theorem, 5 monotone convergence theorem, 17, 21 multiplication operator, 143, 236 multiplicity m spectral measure, 284 Neumann series, 115 norm, 45 induced from inner product, 79, 80 induced metric, 45 of a linear operator, 59, 107 norm-preserving map, 61, 62 operator closable, 228 closed, 228 compact, 123, 136 densely deļ¬ned, 227 ļ¬nite rank, 123 integral, 124 inverse, 113 order, 134 positive, 134, 243 self-adjoint, 129 unbounded, 227 orthogonal complement, 82 orthogonal projection, 84ā€“87, 93, 270 orthonormal basis, 92, 136 orthonormal polynomial, 96 outer measure, 8, 39 parallelogram identity, 82 partition, 15 periodic spectrum, 447 PhragmĀ“enā€“LindelĀØ of method, 215 Poisson kernel for C+ , 201 polarization identity, 78 positive linear functional, 38 precompact subset, 49 product measure, 31 projection theorem, 82, 91 pure point spectrum, 274 pushforward of a Ļƒ-algebra, 3 Pythagorean theorem, 79, 81 Radonā€“Nikodym theorem, 163, 164 range, 60 regular endpoint, 360, 382 regular measure, 36 resolvent, 113, 115

471

of self-adjoint operator, 148, 151 of unbounded operator, 231 resolvent identity, 114 Ricatti equation, 421 Rieszā€“Fischer theorem, 57 Rieszā€“Markov theorem, 38 Schnolā€™s theorem, 335, 441, 442 Schur function, 188 Schwarz integral formula, 190 Schwarz lemma, 188 Schwarzā€“Pick theorem, 189, 223 second kind polynomials, 314 second-countability, 10 of R, 10 of Rn , 12 self-adjoint operator unbounded, 232 seminorm, 45, 47, 54 separability of C(K), 53 of Lp spaces, 58 of a Hilbert space, 95 separable metric space, 10 sesquilinear form, 77 nondegenerate, 254 positive deļ¬nite, 78 skew-symmetric, 254 symplectic, 254 shift operator, 109, 117, 126 Ļƒ-algebra, 2 generated by a set, 2 Ļƒ-locally compact space, 32 Ļƒ-additive, 6 Ļƒ-compactness, 30 simple function, 15 singular continuous spectrum, 274 singular spectrum, 274 singular value decomposition, 137 spectral basis, 146 spectral mapping theorem, 118, 157 spectral measure, 139, 153, 156 for unbounded self-adjoint operator, 238 maximal, 268 spectral multiplicity, 283 spectral projection, 270 spectral radius, 116 spectral representation, 148 spectral theorem, 136, 143, 147, 153, 155, 242 spectrum, 113, 115, 232

472

square root of positive operator, 152, 158 Stieltjes inversion, 202 Stoneā€™s theorem, 290 Stoneā€“Weierstrass theorem, 50, 52 subalgebra of Bb (R), 244 of Bb (X), 34, 150 of C(K, R) and C(K, C), 50 subordinate solution, 328, 332, 429 subspace closed, 47 cyclic, 153 invariant, 121 of Banach space, 47 resolvent-invariant, 242 support essential, 165, 210 of function, 33 support of a measure, 14 symmetric operator, 232 tensor product of Hilbert spaces, 100 Tonelliā€™s theorem, 31 transfer matrix, 313 trigonometric polynomials, 53, 92 uniform boundedness principle, 63 unitary map, 61, 93, 147 Weyl M -matrix, 219, 326, 426 Weyl disk, 409 for a Jacobi matrix, 315 Weyl solution, 311, 390 Weylā€™s criterion, 131, 148, 279 Wronskian, 305, 369 Youngā€™s inequality, 55

Index

Selected Published Titles in This Series 226 Milivoje LukiĀ“ c, A First Course in Spectral Theory, 2022 225 Jacob Bedrossian and Vlad Vicol, The Mathematical Analysis of the Incompressible Euler and Navier-Stokes Equations, 2022 223 Volodymyr Nekrashevych, Groups and Topological Dynamics, 2022 222 Michael Artin, Algebraic Geometry, 2022 221 David Damanik and Jake Fillman, One-Dimensional Ergodic SchrĀØ odinger Operators, 2022 220 Isaac Goldbring, Ultraļ¬lters Throughout Mathematics, 2022 219 Michael Joswig, Essentials of Tropical Combinatorics, 2021 218 Riccardo Benedetti, Lectures on Diļ¬€erential Topology, 2021 217 Marius Crainic, Rui Loja Fernandes, and Ioan MĖ˜ arcut Āø, Lectures on Poisson Geometry, 2021 216 Brian Osserman, A Concise Introduction to Algebraic Varieties, 2021 215 Tai-Ping Liu, Shock Waves, 2021 214 213 212 211

Ioannis Karatzas and Constantinos Kardaras, Portfolio Theory and Arbitrage, 2021 Hung Vinh Tran, Hamiltonā€“Jacobi Equations, 2021 Marcelo Viana and JosĀ“ e M. Espinar, Diļ¬€erential Equations, 2021 Mateusz Michalek and Bernd Sturmfels, Invitation to Nonlinear Algebra, 2021

210 Bruce E. Sagan, Combinatorics: The Art of Counting, 2020 209 Jessica S. Purcell, Hyperbolic Knot Theory, 2020 Ā“ Ā“ 208 Vicente MuĖœ noz, Angel GonzĀ“ alez-Prieto, and Juan Angel Rojo, Geometry and Topology of Manifolds, 2020 207 Dmitry N. Kozlov, Organized Collapse: An Introduction to Discrete Morse Theory, 2020 206 Ben Andrews, Bennett Chow, Christine Guenther, and Mat Langford, Extrinsic Geometric Flows, 2020 205 204 203 202

Mikhail Shubin, Invitation to Partial Diļ¬€erential Equations, 2020 Sarah J. Witherspoon, Hochschild Cohomology for Algebras, 2019 Dimitris Koukoulopoulos, The Distribution of Prime Numbers, 2019 Michael E. Taylor, Introduction to Complex Analysis, 2019

201 Dan A. Lee, Geometric Relativity, 2019 200 Semyon Dyatlov and Maciej Zworski, Mathematical Theory of Scattering Resonances, 2019 199 Weinan E, Tiejun Li, and Eric Vanden-Eijnden, Applied Stochastic Analysis, 2019 198 Robert L. Benedetto, Dynamics in One Non-Archimedean Variable, 2019 197 196 195 194

Walter Craig, A Course on Partial Diļ¬€erential Equations, 2018 Martin Stynes and David Stynes, Convection-Diļ¬€usion Problems, 2018 Matthias Beck and Raman Sanyal, Combinatorial Reciprocity Theorems, 2018 Seth Sullivant, Algebraic Statistics, 2018

193 192 191 190

Martin Lorenz, A Tour of Representation Theory, 2018 Tai-Peng Tsai, Lectures on Navier-Stokes Equations, 2018 Theo BĀØ uhler and Dietmar A. Salamon, Functional Analysis, 2018 Xiang-dong Hou, Lectures on Finite Fields, 2018

189 I. Martin Isaacs, Characters of Solvable Groups, 2018 188 Steven Dale Cutkosky, Introduction to Algebraic Geometry, 2018 187 John Douglas Moore, Introduction to Global Analysis, 2017

For a complete list of titles in this series, visit the AMS Bookstore at www.ams.org/bookstore/gsmseries/.

The central topic of this book is the spectral theory of bounded and unbounded self-adjoint operators on Hilbert spaces. After introducing the necessary prerequisites in measure theory and functional analysis, the exposition focuses on operator theory and especially the structure of self-adjoint operators. These can be viewed as infinite-dimensional analogues of Hermitian matrices; the infinite-dimensional setting leads to a richer theory which goes beyond eigenvalues and eigenvectors and studies self-adjoint operators in the language of spectral measures and the Borel functional calculus. The main approach to spectral theory adopted in the book is to present it as the interplay between three main classes of objects: self-adjoint operators, their spectral measures, and Herglotz functions, which are complex analytic functions mapping the upper half-plane to itself. Self-adjoint operators include many important classes of recurrence and differential operators; the later part of this book is dedicated to two of the most studied classes, Jacobi operators and one-dimensional Schrƶdinger operators. This text is intended as a course textbook or for independent reading for graduate students and advanced undergraduates. Prerequisites are linear algebra, a first course in analysis including metric spaces, and for parts of the book, basic complex analysis. Necessary results from measure theory and from the theory of Banach and Hilbert spaces are presented in the first three chapters of the book. Each chapter concludes with a number of helpful exercises.

For additional information and updates on this book, visit www.ams.org/bookpages/gsm-226

GSM/226

www.ams.org