You are on page 1of 876

Joseph L.

Doob

Classical
Potential Theory
and Its Probabilistic
Counterpart

Springer
Classics in Mathematics
Joseph L. Doob Classical Potential Theory
and Its Probabilistic Counterpart
Springer
Berlin
Heidelberg
New York
Barcelona
Hong Kong
London
Milan
Paris
Singapore
Tokyo
Joseph L. Doob

Classical Potential Theory


and Its Probabilistic
Counterpart

Reprint of the 1984 Edition

Springer
Joseph L. Doob
University of Illinois
Department of Mathematics
101 West Windsor Road #1104
Urbana, IL 61802
USA
e-mail: doob@math.uiuc.edu

Originally published as Vol. 262 of the


Grundlehren der mathematischen Wissenschaften

Cataloging-in-Publication Data applied for

Die Deutsche Bibliothek - CIP-Einheitsaufnahme

Doob, Joseph L.:


Classical potential theoryand its probabilistic counterpart / J. L. Doob. - Reprint of the 1984 ed. - Berlin;
Heidelberg; New York; Barcelona; Hong Kong; London; Milan; Paris; Singapore; Tokyo: Springer, 2001
(Classics in mathematics)
ISBN 3-540-41206-9

Mathematics Subject Classification (2000): 31-XX, 60J45

ISSN 1431-0821
ISBN 3-540-41206-9 Springer-Verlag Berlin Heidelberg New York

This work is subject to copyright. AB rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting,
reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or
parts thereof is permitted only under the provisions of the German Copyright Law of September 9,1965, in its
current version, and permission for use must always be obtained from Springer-Verlag. Violations are liable
for prosecution under the German Copyright Law.

Springer-Verlag Berlin Heidelberg New York


a member of BertelsmannSpringer Science+Business Media GmbH

O Springer-Verlag Berlin Heidelberg 2001


Printed in Germany

The use of general descriptive names, registered names, trademarks etc. in this publication does not imply,
even in the absence of a specific statement, that such names are exempt from the relevant protective laws and
regulations and therefore free for general use.

Printed on acid-free paper SPIN 10786705 41 /3142ck-5 4 3210


J. L. Doob

Classical Potential Theory


and Its Probabilistic
Counterpart

Springer-Verlag
New York Berlin Heidelberg Tokyo
J. L. Doob
Department of Mathematics
University of Illinois
Urbana, IL 61801
U.S.A.

AMS Subject Classifications: 31-XX, 60J45

Library of Congress Cataloging in Publication Data


Doob, Joseph L.
Classical potential theory and its probabilistic counterpart.
(Grundlehren der mathematischen Wissenschaften; 262)
Bibliography: p.
1. Potential, Theory of. 2. Harmonic functions. 3. Martingales
(Mathematics) I. Title. 11. Series.
QA404.7.D66 1983 515.7 83-12446

J 1984 by Springer-Verlag New York Inc.


All rights reserved. No part of this book may be translated or reproduced in any form
without written consent from Springer-Verlag, 175 Fifth Avenue, New York, New
York 10010, U.S.A.

Typeset by Asco Trade Typesetting Ltd., Hong Kong.


Printed and bound by R. R. Donnelley & Sons, Harrisonburg, Virginia.
Printed in the United States of America.

987654321
ISBN 0-387-90881-1 Springer-Verlag New York Berlin Heidelberg Tokyo
Contents

Introduction xxi

Notation and Conventions xxv

Part I
Classical and Parabolic Potential Theory

Chapter I
Introduction to the Mathematical Background of Classical Potential
Theory ..................................................... 3
1. The Context of Green's Identity .................................... 3
2. Function Averages ............................................... 4
3. Harmonic Functions .............................................. 4
4. Maximum-Minimum Theorem for Harmonic Functions ............... 5
5. The Fundamental Kernel for R" and Its Potentials .................... 6
6. Gauss Integral Theorem ........................................... 7
7. The Smoothness of Potentials; The Poisson Equation .................. 8
8. Harmonic Measure and the Riesz Decomposition ..................... II

Chapter 11
Basic Properties of Harmonic, Subharmonic, and Superharmonic
Functions ................................................... 14
I. The Green Function of a Ball; The Poisson Integral ................... 14
2. Harnack's Inequality ............................................. 16
3. Convergence of Directed Sets of Harmonic Functions ................. 17
4. Harmonic, Subharmonic, and Superharmonic Functions ............... 18
5. Minimum Theorem for Superharmonic Functions ..................... 20
6. Application of the Operation Tg .................................... 20
7. Characterization of Superharmonic Functions in Terms of Harmonic
Functions ....................................................... 22
8. Differentiable Superharmonic Functions ............................. 23
9. Application of Jensen's Inequality .................................. 23
10. Superharmonic Functions on an Annulus ............................ 24
Il. Examples ....................................................... 25
12. The Kelvin Transformation (N ; 2) ................................. 26
Vi Contents

13. Greenian Sets .................................................... 27


14. The L' (µ8 ) and D(µ8 _) Classes of Harmonic Functions on a Ball B: The
Riesz Herglotz Theorem .......................................... 27
15 The Fatou Boundary Limit Theorem ................................ 31
16. Minimal Harmonic Functions ...................................... 33

Chapter III
Infima of Families of Superharmonic Functions .................. 35
I. Least Superharmonic Majorant (LM) and Greatest Subharmonic
Minorant (GM) .................................................. 35
2. Generalization of Theorem I ....................................... 36
3 Fundamental Convergence Theorem (Preliminary Version) ............. 37
4. The Reduction Operation ......................................... 38
5. Reduction Properties ............................................. 41
6. A Smallness Property of Reductions on Compact Sets ................. 42
7 The Natural (Pointwise) Order Decomposition for Positive Superharmonic
Functions . ..................................................... 43

Chapter IV
Potentials on Special Open Sets ................................ 45
1 Special Open Sets, and Potentials on Them ........................... 45
2. Examples ....................................................... 47
3 A Fundamental Smallness Property of Potentials ..................... 48
4. Increasing Sequences of Potentials .................................. 49
5. Smoothing of a Potential .......................................... 49
6. Uniqueness of the Measure Determining a Potential ................... 50
7. Riesz Measure Associated with a Superharmonic Function ............. 51
8 Riesz Decomposition Theorem ..................................... 52
9. Counterpart for Superharmonic Functions on R' of the Riesz
Decomposition .................................................. 53
10. An Approximation Theorem ....................................... 55

Chapter V
Polar Sets and Their Applications .............................. 57
1. Definition ....................................................... 57
2. Superharmonic Functions Associated with a Polar Set ................. 58
3. Countable Unions of Polar Sets .................................... 59
4. Properties of Polar Sets ........................................... 59
5. Extension of a Superharmonic Function ............................. 60
6. Greenian Sets in R2 as the Complements of Nonpolar Sets ............. 63
7. Superharmonic Function Minimum Theorem (Extension of
Theorem 11.5) .................................................... 63
8 Evans- Vasilesco Theorem ......................................... 64
9. Approximation of a Potential by Continuous Potentials ................ 66
10 The Domination Principle ......................................... 67
11 The Infinity Set of a Potential and the Riesz Measure .................. 68
Contents vii

Chapter VI
The Fundamental Convergence Theorem and the Reduction
Operation ................................................... 70
1. The Fundamental Convergence Theorem ............................ 70
2. Inner Polar versus Polar Sets ....................................... 71
3. Properties of the Reduction Operation .............................. 74
4. Proofs of the Reduction Properties .................................. 77
5. Reductions and Capacities ......................................... 84

Chapter VII
Green Functions ............................................. 85
1. Definition of the Green Function GD ................................ 85
2. Extremal Property of GD .......................................... 87
3. Boundedness Properties of GD ...................................... 88
4. Further Properties of GD .......................................... 90
5. The Potential GDµ of a Measure µ .................................. 92
6. Increasing Sequences of Open Sets and the Corresponding Green Function
Sequences ....................................................... 94
7. The Existence of GD versus the Greenian Character of D ............... 94
8. From Special to Greenian Sets ..................................... 95
9. Approximation Lemma ........................................... 95
10. The Function GD(-, DID-1t, as a Minimal Harmonic Function............ 96

Chapter VIII
The Dirichlet Problem for Relative Harmonic Functions ........... 98
1. Relative Harmonic, Superharmonic, and Subharmonic Functions ....... 98
2. The PWB Method ................................................ 99
3. Examples ....................................................... 104
4. Continuous Boundary Functions on the Euclidean Boundary (h =_ 1) .... 106
5. h-Harmonic Measure Null Sets ..................................... 108
6. Properties of PWB"Solutions ...................................... 110
7. Proofs for Section 6 .............................................. III
8. h-Harmonic Measure ............................................. 114
9. h-Resolutive Boundaries ........................................... 118
10. Relations between Reductions and Dirichlet Solutions ................. 122
11. Generalization of the Operator rB and Application to GM" ............. 123
12. Barriers ......................................................... 124
13. h-Barriers and Boundary Point h-Regularity .......................... 126
14. Barriers and Euclidean Boundary Point Regularity .................... 127
15. The Geometrical Significance of Regularity (Euclidean Boundary, h = 1) . 128
16. Continuation of Section 13 ........................................ 130
17. h-Harmonic Measure µD as a Function of D .......................... 131
18. The Extension GD- of GD and the Harmonic Average G8 011 9) When
D c B .......................................................... 132
19. Modification of Section 18 for D = 682 .............................. 136
20. Interpretation of 0D as a Green Function with Pole oo (N = 2) .......... 139
21. Variant of the OperatortB ......................................... 140
Viii Contents

Chapter IX
Lattices and Related Classes of Functions ....................... 141
1. Introduction ..................................................... 141
2. LMD u for an h-Subharmonic Function a ............................ 141
D(Akl)-)

3. The Class ................................................ 142


4. The Class L"(µn-)(P ? I) ......................................... 144
5. The Lattices (Si, 5) and (S`, E) ................................... 145
6. The Vector Lattice (S. <) ......................................... 146
7. The Vector Lattice S . ............................... ........... 148
8. The Vector Lattice SD ............................................. 149
9. The Vector Lattice Sqb ........................................... 150
10. The Vector Lattice S. ............................................. 151
1 A Refinement of the Riesz Decomposition ........................... 152
12. Lattices of h-Harmonic Functions on a Ball .......................... 152

Chapter X
The Sweeping Operation ...................................... 155
1. Sweeping Context and Terminology ................................. 155
2. Relation between Harmonic Measure and the Sweeping Kernel ......... 157
3. Sweeping Symmetry Theorem ...................................... 158
4. Kernel Property of Sr; ............................................. 158
5. Swept Measures and Functions ..................................... 160
6. Some Properties of Si; ............................................. 161
7. Poles of a Positive Harmonic Function .............................. 163
8 Relative Harmonic Measure on a Polar Set .......................... 164

Chapter Xl
The Fine Topology ........................................... 166
I. Definitions and Basic Properties .................................... 166
2. A Thinness Criterion ............................................. 168
3. Conditions That I,' e Af ............................................ 169
4. An Internal Limit Theorem ........................................ 171
5 Extension of the Fine Topology to P" v { oo I ......................... 175
6. The Fine Topology Derived Set of a Subset of RN ..................... 177
7. Application to the Fundamental Convergence Theorem and to Reductions. 177
8 Fine Topology Limits and Euclidean Topology Limits ................. 178
9. Fine Topology Limits and Euclidean Topology Limits (Continued) ...... 179
10. Identification of Af in Terms of a Special Function u` ................. 180
If. Quasi-Lindelof Property .......................................... 180
12. Regularity in Terms of the Fine Topology ........................... 181
13. The Euclidean Boundary Set of Thinness of a Greenian Set ............. 182
14. The Support of a Swept Measure ................................... 183
15. Characterization 183
16. A Special Reduction .............................................. 184
17. The Fine Interior of a Set of Constancy of a Superharmonic Function ... 184
18. The Support of a Swept Measure (Continuation of Section 14) .......... 185
19. Superharmonic Functions on Fine-Open Sets ......................... 187
20. A Generalized Reduction .......................................... 187
Contents ix

21. Limits of Superharmonic Functions at Irregular Boundary Points of Their


Domains ........................................................ 190
22. The Limit Harmonic Measure fPD .................................. 191
23. Extension of the Domination Principle .............................. 194

Chapter XII
The Martin Boundary ........................................ 195
1. Motivation ...................................................... 195
2. The Martin Functions ............................................ 196
3. The Martin Space ................................................ 197
4. Preliminary Representations of Positive Harmonic Functions and Their
Reductions ...................................................... 199
5. Minimal Harmonic Functions and Their Poles ....................... 200
6. Extension of Lemma 4 ............................................ 201
7. The Set of Nonminimal Martin Boundary Points ..................... 202
8. Reductions on the Set of Minimal Martin Boundary Points ............ 203
9. The Martin Representation ........................................ 204
10. Resolutivity of the Martin Boundary ................................ 207
1. Minimal Thinness at a Martin Boundary Point ....................... 208
12. The Minimal-Fine Topology ....................................... 210
13. First Martin Boundary Counterpart of Theorem XI.4(c) and (d) ........ 213
14. Second Martin Boundary Counterpart of Theorem XI.4(c) ............. 213
15. Minimal-Fine Topology Limits and Martin Topology Limits at a Minimal
Martin Boundary Point ........................................... 215
16. Minimal-Fine Topology Limits and Martin Topology Limits at a Minimal
Martin Boundary Point (Continued) ................................ 216
17. Minimal-Fine Martin Boundary Limit Functions ..................... 216
18. The Fine Boundary Function of a Potential .......................... 218
19. The Fatou Boundary Limit Theorem for the Martin Space ............. 219
20. Classical versus Minimal-Fine Topology Boundary Limit Theorems for
Relative Superharmonic Functions on a Ball in R" .................... 221
21. Nontangential and Minimal-Fine Limits at a Half-space Boundary ...... 222
22. Normal Boundary Limits for a Half-space ........................... 223
23. Boundary Limit Function (Minimal-Fine and Normal) of a Potential on a
Half-space ...................................................... 225

Chapter XIII
Classical Energy and Capacity ................................. 226
1. Physical Context ................................................. 226
2. Measures and Their Energies ...................................... 227
3. Charges and Their Energies ........................................ 228
4. Inequalities between Potentials, and the Corresponding Energy
Inequalities ...................................................... 229
5. The Function 230
6. Classical Evaluation of Energy; Hilbert Space Methods ................ 231
7. The Energy Functional (Relative to an Arbitrary Greenian Subset D of
RN) ............................................................. 233
8. Alternative Proofs of Theorem 7(b') ................................ 235
X Contents

9. Sharpening of Lemma 4 ........................................... 237


10. The Classical Capacity Function ................................... 237
11. Inner and Outer Capacities (Notation of Section 10) ................... 240
12. Extremal Property Characterizations of Equilibrium Potentials (Notation
of Section 10) .................................................... 241
13. Expressions for C(A) ............................................. 243
14. The Gauss Minimum Problems and Their Relation to Reductions ....... 244
15. Dependence of C' on D ........................................... 247
16. Energy Relative to R2 ............................................. 248
17. The Wiener Thinness Criterion ..................................... 249
18. The Robin Constant and Equilibrium Measures Relative to l 2 (N = 2) .. 251

Chapter XIV
One-Dimensional Potential Theory ............................. 256
I. Introduction ..................................................... 256
2. Harmonic, Superharmonic, and Subharmonic Functions ............... 256
3. Convergence Theorems ........................................... 256
4. Smoothness Properties of Superharmonic and Subharmonic Functions.. . 257
5. The Dirichlet Problem (Euclidean Boundary) ......................... 257
6. Green Functions ...................... .......................... 258
7. Potentials of Measures ............................................ 259
8. Identification of the Measure Defining a Potential ..................... 259
9. Riesz Decomposition ............................................. 260
10. The Martin Boundary ............................................ 261

Chapter XV
Parabolic Potential Theory: Basic Facts ......................... 262
I. Conventions ..................................................... 262
2. The Parabolic and Coparabolic Operators ........................... 263
3. Coparabolic Polynomials .......................................... 264
4. The Parabolic Green Function oflt8................................. 266
5. Maximum-Minimum Parabolic Function Theorem .................... 267
6. Application of Green's Theorem ................................... 269
7. The Parabolic Green Function of a Smooth Domain; The Riesz Decom-
position and Parabolic Measure (Formal Treatment) .................. 270
8. The Green Function of an Interval .............................. ... 272
9. Parabolic Measure for an Interval .................................. 273
10. Parabolic Averages ............................................... 275
I. Harnack's Theorems in the Parabolic Context ........................ 276
12. Superparabolic Functions ......................................... 277
13. Superparabolic Function Minimum Theorem ........................ 279
14. The Operation iB and the Defining Average Properties of Superparabolic
Functions ....................................................... 280
5 Superparabolic and Parabolic Functions on a Cylinder ................ 281
16. The Appall Transformation ........................................ 282
17. Extensions of a Parabolic Function Defined on a Cylinder ............. 283
Contents xi

Chapter XVI
Subparabolic, Superparabolic, and Parabolic Functions on a Slab ... 285
1. The Parabolic Poisson Integral for a Slab ............................ 285
2. A Generalized Superparabolic Function Inequality .................... 287
3. A Criterion of a Subparabolic Function Supremum ................... 288
4. A Boundary Limit Criterion for the Identically Vanishing of a Positive
Parabolic Function ............................................... 288
5. A Condition that a Positive Parabolic Function Be Representable by a
Poisson Integral .................................................. 290
6. The and D(jie_) Classes of Parabolic Functions on a Slab ...... 290
7. The Parabolic Boundary Limit Theorem ............................. 292
8. Minimal Parabolic Functions on a Slab ............................. 293

Chapter XVII
Parabolic Potential Theory (Continued) ......................... 295
1. Greatest Minorants and Least Majorants ............................ 295
2. The Parabolic Fundamental Convergence Theorem (Preliminary Version)
and the Reduction Operation ...................................... 295
3. The Parabolic Context Reduction Operations ........................ 296
4. The Parabolic Green Function ..................................... 298
5. Potentials ....................................................... 300
6. The Smoothness of Potentials ...................................... 303
7. Riesz Decomposition Theorem ..................................... 305
8. Parabolic-Polar Sets .............................................. 305
9. The Parabolic-Fine Topology ...................................... 308
10. Semipolar Sets ................................................... 309
It. Preliminary List of Reduction Properties ............................ 310
12. A Criterion of Parabolic Thinness .................................. 313
13. The Parabolic Fundamental Convergence Theorem ................... 314
14. Applications of the Fundamental Convergence Theorem to Reductions
and to Green Functions ........................................... 316
15. Applications of the Fundamental Convergence Theorem to the Parabolic-
Fine Topology ................................................... 317
16. Parabolic-Reduction Properties .................................... 317
17. Proofs of the Reduction Properties in Section 16 ...................... 320
18. The Classical Context Green Function in Terms of the Parabolic Context
Green Function (N > 1) ........................................... 326
19. The Quasi-Lindelof Property ....................................... 328

Chapter XVIII
The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets ... 329
1. Relativization of the Parabolic Context; The PWB Method in this
Context ......................................................... 329
2. h-Parabolic Measure .............................................. 332
3. Parabolic Barriers ................................................ 333
4. Relations between the Classical Dirichlet Problem and the Parabolic
Context Dirichlet Problem ......................................... 334
5. Classical Reductions in the Parabolic Context ........................ 335
xii Contents

6. Parabolic Regularity of Boundary Points ............................ 337


7. Parabolic Regularity in Terms of the Fine Topology ................... 341
8. Sweeping in the Parabolic Context .................................. 341
9. The Extension GD of OD and the Parabolic Average to( ,GB when
b c h .......................................................... 343
10. Conditions that E *f ............................................ 345
1I. Parabolic- and Coparabolic-Polar Sets .............................. 347
12. Parabolic- and Coparabolic-Semipolar Sets .......................... 348
13. The Support of a Swept Measure ................................... 350
14. An Internal Limit Theorem; The Coparabolic-Fine Topology Smoothness
of Superparabolic Functions ....................................... 351
15. Application to a Version of the Parabolic Context Fatou Boundary Limit
Theorem on a Slab ............................................... 357
16. The Parabolic Context Domination Principle ......................... 358
17. Limits of Superparabolic Functions at Parabolic-Irregular Boundary
Points of Their Domains .......................................... 358
18. Martin Flat Point Set Pairs ........................................ 361
19. Lattices and Related Classes of Functions in the Parabolic Context ...... 361

Chapter XIX
The Martin Boundary in the Parabolic Context ................... 363
Introduction ..................................................... 363
2. The Martin Functions of Martin Point Set and Measure Set Pairs ....... 364
3. The Martin Space DM ............................................ 366
4. Preparatory Material for the Parabolic Context Martin Representation
Theorem ........................................................ 367
5. Minimal Parabolic Functions and Their Poles ........................ 369
6. The Set of Nonminimal Martin Boundary Points ..................... 370
7. The Martin Representation in the Parabolic Context .................. 371
8 Martin Boundary of a Slab D = R' x ]0, b[ with 0 < d S + op ......... 371
9. Martin Boundaries for the Lower Half-space of A' and for Al .......... 374
10. The Martin Boundary of ,6 = ]0, +oo[ x ] - oo, 6[ ................... 375
11. PWB" Solutions on DM ........................................... 377
12. The Minimal-Fine Topology in the Parabolic Context ................. 377
13. Boundary Counterpart of Theorem XVIII.l4(f) .................. ... 379
14. The Vanishing of Potentials on OMD ................................ 381
15. The Parabolic Context Fatou Boundary Limit Theorem on Martin Spaces 381

Part 2
Probabilistic Counterpart of Part I

Chapter I
Fundamental Concepts of Probability ........................... 387
1. Adapted Families of Functions on Measurable Spaces ................. 387
2. Progressive Measurability ............................... ......... 388
3. Random Variables ............................................... 390
Contents xiii

4. Conditional Expectations .......................................... 391


5. Conditional Expectation Continuity Theorem ........................ 393
6. Fatou's Lemma for Conditional Expectations ........................ 396
7. Dominated Convergence Theorem for Conditional Expectations ........ 397
8. Stochastic Processes, "Evanescent," "Indistinguishable," "Standard Modi-
fication," "Nearly ................................................. 398
9. The Hitting of Sets and Progressive Measurability .................... 401
10. Canonical Processes and Finite-Dimensional Distributions ............. 402
11. Choice of the Basic Probability Space ............................... 404
12. The Hitting of Sets by a Right Continuous Process .................... 405
13. Measurability versus Progressive Measurability of Stochastic Processes .. 407
14. Predictable Families of Functions .................................. 410

Chapter II
Optional Times and Associated Concepts ........................ 413
1. The Context of Optional Times .................................... 413
2. Optional Time Properties (Continuous Parameter Context) ............. 415
3. Process Functions at Optional Times ................................ 417
4. Hitting and Entry Times .......................................... 419
5. Application to Continuity Properties of Sample Functions ............. 421
6. Continuation of Section 5 ......................................... 423
7. Predictable Optional Times ........................................ 423
8. Section Theorems ................................................ 425
9. The Graph of a Predictable Time and the Entry Time of a Predictable
Set ............................................................. 426
10. Semipolar Subsets of 1R x 0 ...................................... 427
11. The Classes D and L° of Stochastic Processes ......................... 428
12. Decomposition of Optional Times; Accessible and Totally Inaccessible
Optional Times .................................................. 429

Chapter III
Elements of Martingale Theory ................................ 432
1. Definitions ...................................................... 432
2. Examples ....................................................... 433
3. Elementary Properties (Arbitrary Simply Ordered Parameter Set) ....... 435
4. The Parameter Set in Martingale Theory ............................ 437
5. Convergence of Supermartingale Families ........................... 437
6. Optional Sampling Theorem (Bounded Optional Times) ............... 438
7. Optional Sampling Theorem for Right Closed Processes ............... 440
8. Optional Stopping ................................................ 442
9. Maximal Inequalities ............................................. 442
10. Conditional Maximal Inequalities .................................. 444
1. An L° Inequality for Submartingale Suprema ......................... 444
12. Crossings ....................................................... 445
13. Forward Convergence in the L' Bounded Case ....................... 450
14. Convergence of a Uniformly Integrable Martingale ................... 451
15. Forward Convergence of a Right Closable Supermartingale ............ 453
16. Backward Convergence of a Martingale ............................. 454
xiv Contents

17. Backward Convergence of a Supermartingale ......................... 455


18. The r Operator .................................................. 455
19. The Natural Order Decomposition Theorem for Supermartingales ...... 457
20. The Operators LM and GM ....................................... 458
21. Supermartingale Potentials and the Riesz Decomposition .............. 459
22. Potential Theory Reductions in a Discrete Parameter Probability
Context ......................................................... 459
23. Application to the Crossing Inequalities ............................. 461

Chapter I V
Basic Properties of Continuous Parameter Supermartingales ........ 463
1. Continuity Properties ............................................. 463
2. Optional Sampling of Uniformly Integrable Continuous Parameter
Martingales ..................................................... 468
3 Optional Sampling and Convergence of Continuous Parameter
Supermartingales ................................................. 470
4. Increasing Sequences of Supermartingales ........................... 473
5. Probability Version of the Fundamental Convergence Theorem of Potential
Theory ......................................................... 476
6. Quasi-Bounded Positive Supermartingales; Generation of Supermartingale
Potentials by Increasing Processes ............................ ..... 480
7. Natural versus Predictable Increasing Processes (1= 71' or R+) ......... 483
8. Generation of Supermartingale Potentials by Increasing Processes in the
Discrete Parameter Case .......................................... 488
9. An Inequality for Predictable Increasing Processes .................... 489
10. Generation of Supermartingale Potentials by Increasing Processes for
Arbitrary Parameter Sets .......................................... 490
11 Generation of Supermartingale Potentials by Increasing Processes in the
Continuous Parameter Case: The Meyer Decomposition ............... 493
12. Meyer Decomposition of a Submartingale ........................... 495
13 Role of the Measure Associated with a Supermartingale;
The Supermartingale Domination Principle .......................... 496
14. The Operators r, LM, and GM in the Continuous Parameter Context.... 500
15. Potential Theory on a8' x 52 ....................................... 501
16. The Fine Topology of R * x f) ..................................... 502
17 Potential Theory Reductions in a Continuous Parameter Probability
Context ...................................................... 504
18. Reduction Properties ............................................. 505
19. Proofs of the Reduction Properties in Section 18 ...................... 509
20. Evaluation of Reductions ......................... .......... .... 513
21. The Energy of a Supermartingale Potential ............... .... ...... 515
22. The Subtraction of a Supermartingale Discontinuity ................... 516
23. Supermartingale Decompositions and Discontinuities .......... .... . 518

Chapter V
Lattices and Related Classes of Stochastic Processes ............... 520
I. Conventions; The Essential Order .................................. 520
2. LM when ; Is a Submartingale ........................ 521
Contents Xv

3. Uniformly Integrable Positive Submartingales ........................ 523


4. L° Bounded Stochastic Processes (p >_ 1) ............................ 524
5. The Lattices ('St, 5), ('S+, 5), (Si, 5), (S+, 5) ..................... 525
6. The Vector Lattices ('S, <) and (S, <) .............................. 528
7. The Vector Lattices ('S.,<) and (S,,,, <) ............................ 529
8. The Vector Lattices ('Si, <) and (Sr, <) ............................. 530
9. The Vector Lattices ('Sqb, <) and (Sqb, <) ........................... 531
10. The Vector Lattices ('Sr, <) and (Sr, <) ............................. 532
11. The Orthogonal Decompositions 'S,,, = 'Smgb + 'S., and S. = S,,,qb + S,,,, . 533
12. Local Martingales and Singular Supermartingale Potentials in (S, <) .... 534
13. Quasimartingales (Continuous Parameter Context) .................... 535

Chapter VI
Markov Processes ............................................ 539
1. The Markov Property ............................................. 539
2. Choice of Filtration .............................................. 544
3. Integral Parameter Markov Processes with Stationary Transition Proba-
bilities .......................................................... 545
4. Application of Martingale Theory to Discrete Parameter Markov
Processes ....................................................... 547
5. Continuous Parameter Markov Processes with Stationary Transition
Probabilities ..................................................... 550
6. Specialization to Right Continuous Processes ........................ 552
7. Continuous Parameter Markov Processes: Lifetimes and Trap Points .... 554
8. Right Continuity of Markov Process Filtrations; A Zero-One (0-1) Law.. 556
9. Strong Markov Property .......................................... 557
10. Probabilistic Potential Theory; Excessive Functions ................... 560
11. Excessive Functions and Supermartingales ........................... 564
12. Excessive Functions and the Hitting Times of Analytic Sets (Notation and
Hypotheses of Section 11) ......................................... 565
13. Conditioned Markov Processes ..................................... 566
14. Tied Down Markov Processes ...................................... 567
15. Killed Markov Processes .......................................... 568

Chapter VII
Brownian Motion ............................................ 570
1. Processes with Independent Increments and State Space I8" ............ 570
2. Brownian Motion ................................................ 572
3. Continuity of Brownian Paths ...................................... 576
4. Brownian Motion Filtrations ...................................... 578
5. Elementary Properties of the Brownian Transition Density and Brownian
Motion ......................................................... 581
6. The Zero-One Law for Brownian Motion ............................ 583
7. Tied Down Brownian Motion ...................................... 586
8. Andre Reflection Principle ......................................... 587
9. Brownian Motion in an Open Set (N z I) ............................ 589
10. Space-Time Brownian Motion in an Open Set ........................ 592
11. Brownian Motion in an Interval .................................... 594
Xvi Contents

12. Probabilistic Evaluation of Parabolic Measure for an Interval .......... 595


13. Probabilistic Significance of the Heat Equation and Its Dual ............ 596

Chapter VIII
TheIto Integral .............................................. 599
1. Notation .......... ............................................. 599
2. The Size of r. ................................................... 601
3. Properties of the Ito Integral ....................................... 602
4. The Stochastic Integral for an Integrand Process in ro ................. 605
5. The Stochastic Integral for an Integrand Process in r .................. 606
6. Proofs of the Properties in Section 3 ................................ 607
7. Extension to Vector-Valued and Complex-Valued Integrands ........... 611
8 Martingales Relative to Brownian Motion Filtrations ................. 612
9. A Change of Variables ............................................ 615
10. The Role of Brownian Motion Increments ........................... 618
11 (N = I) Computation of the Ito Integral by Riemann-Stieltjes Sums ..... 620
12. It6's Lemma ..................................................... 621
13. The Composition of the Basic Functions of Potential Theory with Brownian
Motion . ....................................................... 625
14. The Composition of an Analytic Function with Brownian Motion ....... 626

Chapter IX
Brownian Motion and Martingale Theory ....................... 627
I Elementary Martingale Applications ................................ 627
2. Coparabolic Polynomials and Martingale Theory ..................... 630
3. Superharmonic and Harmonic Functions on R' and Supermartingales and
Martingales ....... ................................. ........... 632
4 Hitting of an f. Set ....... ....................................... 635
5. The Hitting of a Set by Brownian Motion .................... ....... 636
6 Superharmonic Functions, Excessive for Brownian Motion ......... . ... 637
7 Preliminary Treatment of the Composition of a Superharmonic Function
with Brownian Motion; A Probabilistic Fatou Boundary Limit Theorem. 641
8 Excessive and Invariant Functions for Brownian Motion ............... 645
9. Application to Hitting Probabilities and to Parabolicity of Transition
Densities. .. . .................................. ....... 647
10. (N = 2) The Hitting of Nonpolar Sets by Brownian Motion ........ ... 648
11. Continuity of the Composition of a Function with Brownian Motion .... 649
12. Continuity of Superharmonic Functions on Brownian Motion .......... 650
13 Preliminary Probabilistic Solution of the Classical Dirichlet Problem .... 651
14 Probabilistic Evaluation of Reductions .................. .........
. 653
15 Probabilistic Description of the Fine Topology ....................... 656
16 x-Excessive Functions for Brownian Motion and Their Composition with
Brownian Motions .. .................................... ....... 659
17. Brownian Motion Transition Functions as Green Functions; The Corre-
sponding Backward and Forward Parabolic Equations ................ 661
18. Excessive Measures for Brownian Motion ........................... 663
19. Nearly Borel Sets for Brownian Motion ............................. 666
20. Brownian Motion into a Set from an Irregular Boundary Point ......... 666
Contents Xvii

Chapter X
Conditional Brownian Motion ................................. 668
I. Definition ....................................................... 668
2. h-Brownian Motion in Terms of Brownian Motion .................... 671
3. Contexts for (2. 1) ................................................ 676
4. Asymptotic Character of h-Brownian Paths at Their Lifetimes .......... 677
5. h-Brownian Motion from an Infinity of h ............................ 680
6. Brownian Motion under Time Reversal ............................. 682
7. Preliminary Probabilistic Solution of the Dirichlet Problem for h-Harmonic
Functions; h-Brownian Motion Hitting Probabilities and the
Corresponding Generalized Reductions .............................. 684
8. Probabilistic Boundary Limit and Internal Limit Theorems for Ratios of
Strictly Positive Superharmonic Functions ........................... 688
9. Conditional Brownian Motion in a Ball ............................. 691
10. Conditional Brownian Motion Last Hitting Distributions; The Capacitary
Distribution of a Set in Terms of a Last Hitting Distribution ........... 693
11. The Tail a Algebra of a Conditional Brownian Motion ................ 694
12. Conditional Space-Time Brownian Motion .......................... 699
13. [Space-Time] Brownian Motion in [iv] ii1;N with Parameter Set R ....... 700

Part 3

Chapter I
Lattices in Classical Potential Theory and Martingale Theory ....... 705
1. Correspondence between Classical Potential Theory and Martingale
Theory ......................................................... 705
2. Relations between Decomposition Components of S in Potential Theory
and Martingale Theory ........................................... 706
3. The Classes L° and D ............................................. 706
4. PWB-Related Conditions on h-Harmonic Functions and on Martingales . 707
5. Class D Property versus Quasi-Boundedness ......................... 708
6. A Condition for Quasi-Boundedness ................................ 709
7. Singularity of an Element of S : .................................... 710
8. The Singular Component of an Element of S4 ........................ 711
9. The Class S .. ................................................... 712
10. The Class S...................................................... 714
11. Lattice Theoretic Analysis of the Composition of an h-Superharmonic
Function with an h-Brownian Motion ............................... 715
12. A Decomposition of Sm, (Potential Theory Context) ................... 716
13. Continuation of Section I l ........................................ 717

Chapter II
Brownian Motion and the PWB Method ........................ 719
1. Context of the Problem ........................................... 719
2. Probabilistic Analysis of the PWB Method ........................... 720
xviii Contents

3 PW B' Examples .. .............................................. 723


4. Tail a Algebras in the PWB'' Context ................................ 725

Chapter III
Brownian Motion on the Martin Space .......................... 727
1. The Structure of Brownian Motion on the Martin Space ............... 727
2. Brownian Motions from Martin Boundary Points (Notation of Section I) 728
3. The Zero-One Law at a Minimal Martin Boundary Point and the
Probabilistic Formulation of the Minimal-Fine Topology (Notation of
Section I) ......... ............................................ 730
4. The Probabilistic Fatou Theorem on the Martin Space ................. 732
5. Probabilistic Approach to Theorem I.XI.4(c) and Its Boundary
Counterparts .................................................... 733
6. Martin Representation of Harmonic Functions in the Parabolic Context . 735

Appendixes

Appendix I
Analytic Sets ................................................ 741
1. Pavings and Algebras of Sets ....................................... 741
2. Suslin Schemes .................................................. 741
3 Sets Analytic over a Product Paving ................................ 742
4. Analytic Extensions versus a Algebra Extensions of Pavings ............ 743
5 Projection Characterization %f (V) .................................. 743
6. The Operation s9(.a1) ............................................. 744
7. Projections of Sets in Product Pavings ............................... 744
8 Extension of a Measurability Concept to the Analytic Operation Context. 745
9. The G, Sets of a Complete Metric Space ............................. 745
10. Polish Spaces .................................................... 746
11 The Baire Null Space ............................................. 746
12. Analytic Sets .................................................... 747
13. Analytic Subsets of Polish Spaces ................................... 748

Appendix 11
Capacity Theory ............................................. 750
1. Choquet Capacities ............................................... 750
2. Sierpinski Lemma ................................................ 750
3. Choquet Capacity Theorem ........................................ 751
4. Lusin's Theorem ................................................. 751
5. A Fundamental Example of a Choquet Capacity ...................... 752
6. Strongly Subadditive Set Functions ................................. 752
7. Generation of a Choquet Capacity by a Positive Strongly Subadditive Set
Function ........................................................ 753
8. Topological Precapacities ......................................... 755
9. Universally Measurable Sets ..... . ................................. 756
Contents XiX

Appendix III
Lattice Theory ............................................... 758
1. Introduction ..................................................... 758
2. Lattice Definitions ............................................... 758
3. Cones .......................................................... 758
4. The Specific Order Generated by a Cone ............................ 759
5. Vector Lattices .................................................. 760
6. Decomposition Property of a Vector Lattice ......................... 762
7. Orthogonality in a Vector Lattice ................................... 762
8. Bands in a Vector Lattice .......................................... 762
9. Projections on Bands ............................................. 763
10.The Orthogonal Complement of a Set ............................... 764
H. The Band Generated by a Single Element ............................ 764
12. Order Convergence ............................................... 765
13. Order Convergence on a Linearly Ordered Set ........................ 766

Appendix IV
Lattice Theoretic Concepts in Measure Theory ................... 767
1. Lattices of Set Algebras ........................................... 767
2. Measurable Spaces and Measurable Functions ....................... 767
3. Composition of Functions ......................................... 768
4. The Measure Lattice of a Measurable Space .......................... 769
5. The a Finite Measure Lattice of a Measurable Space (Notation of Section 4) 771
6. The Hahn and Jordan Decompositions .............................. 772
7. The Vector Lattice 4l ............................................ 772
8. Absolute Continuity and Singularity ................................ 773
9. Lattices of Measurable Functions on a Measure Space ................. 774
10. Order Convergence of Families of Measurable Functions .............. 775
11. Measures on Polish Spaces ........................................ 777
12. Derivates of Measures ............................................ 778

Appendix V
Uniform Integrability ........................................ 779

Appendix VI
Kernels and Transition Functions .............................. 781
1. Kernels ......................................................... 781
2. Universally Measurable Extension of a Kernel ........................ 782
3. Transition Functions ............................................. 782

Appendix VII
Integral Limit Theorems ...................................... 785
1. An Elementary Limit Theorem ... ................................. 785
2. Ratio Integral Limit Theorems ..................................... 786
3. A One-Dimensional Ratio Integral Limit Theorem .................... 786
4. A Ratio Integral Limit Theorem Involving Convex Variational Derivates. 788
xx Contents

Appendix VIII
Lower Semicontinuous Functions .............................. 791
I The Lower Semicontinuous Smoothing of a Function ............ .. . 791
2. Suprema of Families of Lower Semicontinuous Functions... ... ...... 791
3 Choquet Topological Lemma ............ ......... .. .. .. .. 792

Historical Notes 793


Part I 793
Part 2 . ............. .. 806
Part 3 .. .. ........... 815
Appendixes 816

Bibliography . .... ... ..................................


... 819
Notation Index ....... ..................................
... 827
Index .. . ............ .................................... 829
Introduction

Potential theory and certain aspects of probability theory are intimately


related, perhaps most obviously in that the transition function determining
a Markov process can be used to define the Green function of a potential
theory. Thus it is possible to define and develop many potential theoretic
concepts probabilistically, a procedure potential theorists observe with jaun-
diced eyes in view of the fact that now as in the past their subject provides
the motivation for much of Markov process theory. However that may be
it is clear that certain concepts in potential theory correspond closely to
concepts in probability theory, specifically to concepts in martingale theory.
For example, superharmonic functions correspond to supermartingales.
More specifically: the Fatou type boundary limit theorems in potential
theory correspond to supermartingale convergence theorems, the limit
properties of monotone sequences of superharmonic functions correspond
surprisingly closely to limit properties of monotone sequences of super-
martingales; certain positive superharmonic functions [supermartingales]
are called "potentials," have associated measures in their respective theories
and are subject to domination principles (inequalities) involving the supports
of those measures; in each theory there is a reduction operation whose
properties are the same in the two theories and these reductions induce
sweeping (balayage) of the measures associated with potentials, and.so on.
The purpose of this book is to develop this correspondence between
potential theory and probability theory by examining in detail classical
potential theory, that is, the potential theory of Laplace's equation, together
with the corresponding probability theory, that is, martingale theory. The
joining link which makes this correspondence especially perspicuous is the
Brownian motion process, so this process is studied as needed. In order to
carry through this program it is necessary to study parabolic potential theory,
that is, the potential theory of the heat equation, and the corresponding
process of space time Brownian motion. No knowledge of potential theory
is presupposed but it is assumed that the reader is familiar with basic
probability concepts through conditional expectations. The necessary lattice
theory, analytic set theory and capacity theory are covered in the Appendixes.
Thus this book on the one hand contains an introduction to classical
and parabolic potential theory and on the other hand contains an introduc-
xxii Introduction

tion to martingale theory, including a smattering of the general theory of


stochastic processes and of Markov process theory. There is cross referencing
between the nonprobabilistic and probabilistic aspects of the work, and the
linking of classical and parabolic potential theory with martingale theory, by
Brownian motion and space time Brownian motion, is examined in depth.
One natural criticism of this project is that there is no reason to treat the
very special potential theories of the Laplace and heat equations rather than
general axiomatic potential theory. Another criticism is that there is no
reason to treat potential theory other than as a special subhead of Markov
process theory. In the author's opinion, however, classical potential theory
is too important to serve merely as a source of illustrations of axiomatic
potential theory, which theory in turn is too important in its own right to be
left to the probabilists. To learn potential theory from probability is like
learning algebraic geometry without the geometry.
It would be quite impossible to cover all those parts of modernized
classical potential theory which are relevant to the purpose of this book.
Thus there are striking gaps. For example the treatment of energy is skimpy,
and Dirichlet spaces and the concept of bounded mean oscillation are not
even mentioned in the text. The emphasis is on the Dirichlet problem and
related topics; these are treated in considerable depth. The treatments of
classical and parabolic potential theories are sometimes separated, some-
times together, but the notation is designed to exhibit the parallelism of
the two theories: dots in the notation distinguish parabolic from classical
concepts, thereby muddling eyes but saving brains. And the martingale
theory notation is designed to point out to readers the corresponding
potential theory notation.
Only the part of Markov process theory needed for the relevant discussion
of Brownian motion and conditional Brownian motion is covered. In this
book a stochastic process is a specified family of random variables, frequently
coupled with a filtration to which the family is adapted, but the measure
space of the process is left unspecified and there is no translation operator.
Thus in a discussion of Brownian motion from a varying initial point the
measure space on which the process is defined may vary with the initial
point. This definition of a process may not be best for general Markov
process theory but is convenient in the special context of this book; it
implies for example that no matter how or on what measure space a process
is defined, if it has the properties of a Brownian motion (continuous sample
functions and the correct distributions of independent increments) then it
is a Brownian motion. In a traditional song, a child finds an object which
looks smells and tastes like a peanut so the child concludes that the object
is a peanut. As stochastic processes are sometimes defined, with special
properties demanded of the measure space on which the process random
variables are defined, this simple logic is invalid. However the point of view
of this book makes it essential in discussing Brownian motion to prove
certain invariance properties, for example that two Brownian motion pro-
Introduction xxiii

cesses in N space, with a common initial point and variance parameter, hav
the same probability of hitting an analytic set. This fact is not trivial and
such questions are treated.
There is nothing very novel in this book. Potential theorists may find the
treatment of reductions on boundary sets of interest, as well as the use of
iterated reductions to obtain limit theorems. Correspondingly, probabilists
may find the new supermartingale crossing inequalities and the technique
of iterated reductions of supermartingales of interest. A new domination
principle for supermartingales illustrates the fact that classical potential
theory still suggests interesting probability results.
The author thanks Bruce Hajek, Naresh Jain and John Taylor for helpful
comments on various chapters and, finally, thanks his typist: usually faith-
ful, sometimes accurate.
Notation and Conventions

RN is N dimensional Euclidean space, R = R', and R+ is the set [0, + oo[ of


positive reals. A is the set [ - oo, + co] of extended reals and R+ is the set
[0, + oo] of positive extended reals.
lZ is the set of integers, Z+ is the set 0, 1, 2, ... , and Z, is the set 0, ... , n.
The boundary of an unbounded subset of R' contains the adjoined point
oo of the one point compactification of RN unless some other compactifica-
tion has been specified. This boundary relative to the one point compactifica-
tion of RN will be called the Euclidean boundary of the set.
If is a point of RN and A is a subset of RN the distance between and
A is written I - Al.
S) is the ball, in whatever metric space is specified, of center and
radius S, specifically in RN : (1: In - I < S}.
IN refers to N dimensional Lebesgue measure.
If A and B are subsets of a space the set of points in A but not in B is
denoted by A - B.
"Positive" means " z 0" and monotone concepts are to be taken in the
wide sense, so that for example a constant function from R into R is both
monotone increasing and monotone decreasing.
If D is an open subset of R' the notation refers to the class of
functions from D into R which are continuous together with their derivatives
of order S k.
Limit concepts for a function fat a point do not involve the value of f
at the point. Thus limq_4 f(ry) = a means that f is near a in small deleted
neighborhoods of .
The notation for a sequence frequently uses a dot for the index set; unless
otherwise identified the index set is V, so that A. = {A,, A,, . . . }.
The set on which a function f satisfies some set S of conditions is frequently
denoted by {S}. Thus if f is a function from R into R the positivity set off
is If z0}.
The book is divided into three Parts. Section 1.11.3 is Section 3 of Chapter
11 of Part 1; in any Part, Section II.3 is Section 3 of Chapter II of that part;
in any Chapter, Section 3 is Section 3 of that Chapter, and so on.
Part 1
Classical and Parabolic Potential
Theory
(dimensionality number N > I in Chapters I-XIII)
Chapter I

Introduction to the Mathematical Background


of Classical Potential Theory

1. The Context of Green's Identity


In this chapter some of the mathematical ideas of classical potential theory
are introduced, under simplifying assumptions. The basic space is Euclidean
N space RN. For a ball 6) in R'
(N)-I

IN _1(aB(b, a)) = TrN5N-1,


nN = 27[',2r
2

IN(B(, b)) = aN

An application of the Lebesgue dominated convergence theorem shows that


if u is a locally IN integrable function on an open subset D of RN and if S > 0,
the function

I-+ u d1N
Jsq,a)

is continuous on the set I - ODI > S}.


A bounded open set for which Green's identity (1.2) is true will be called
smooth. Since a ball is obviously smooth in this sense, as is the set between
two concentric balls, and since these sets are the only smooth sets for which
the details of this chapter will be used later, a precise description of smooth
sets is omitted. The Laplacian of a function u, that is, the sum of the unmixed
second partial derivatives of u, will be denoted as usual by Au. If B is a
smooth open subset of RN, if D. is the partial derivation operator on OD
in the direction of the exterior normal, and if u and v are functions in C(z'(B),
Green's identity is

(uAv - vAu)dlN = f (uDov - vDou)dlN_1, (1.2)


SB Jas
which for v - 1 reduces to
4 1 I Introduction to the Mathematical Background of Classical Potential Theory

f Audi, = f (1.3)
B JaBB

The right side of (I.3) is the flux of the vector field grad u out through OB.

2. Function Averages
The unweighted averages of a function u over b) and over b) will
be denoted by L(u, , b) and A (u, , b), respectively; that is

udIN_t, Ni udlN, (2.1)


rtNIN-1 I aetr.al
R'6 Bts.at

assuming that the necessary measurability and integrability conditions are


satisfied. If u is defined, 1N measurable and locally IN integrable on an open
subset D of RN, the function A(u, , b) is defined and continuous on the set
14:14 - oDI > b} (see Section 1). In this case according to Fubini's theorem
the average L(u, , r) is defined for 11 almost every r in the interval 10,61
and

Nb-N f (2.2)
0
Define
((Nexp[-(I - r)-1] if 0 < r < I
0 ifrz1,
choosing CN so that aNlo rN-1yN(r)11(dr) = 1, and under the preceding hy-
potheses on u, define Aau for b > 0 by

Aau(f) = b-N fN YN a
a
= nN f fl - aDl > b).

The function Aau is infinitely differentiable, Aa I =- 1, and if u is continuous


lim8..,o Aau = u locally uniformly on D.

3. Harmonic Functions
A harmonic function is defined as a (finite-valued) continuous function u,
defined on a nonempty open subset of RN, satisfying

L(u, , b) if b) c D. (3.1)
4. Maximum-Minimum Theorem for Harmonic Functions 5

The stated continuity condition is stronger than necessary. In fact, if it is


weakened to /N measurability and local integrability, (2.4) implies that
A3(u, ) = for - ODI > a, so that u is infinitely differentiable and
I

therefore harmonic, and we have incidentally shown that harmonic functions


are infinitely differentiable. The class of harmonic functions on D is trivially
linear.
If u is harmonic on D, (2.2) yields

>b). (3.2)

Conversely (3.2) implies (3.1) if u is IN measurable and locally integrable


because under (3.2) the function u must be continuous and (2.2) then yields
(3.1).
Equation (1.3) with B = b) reduces to

Au d1N = nNS"-' b L(u, , S). (3.3)


J
Hence a function u with continuous second partial derivatives is harmonic
if and only if u satisfies Laplace's equation Au = 0, and it follows that
harmonicity is a local property. If u is harmonic all its partial derivatives
are harmonic because they satisfy Laplace's equation also. According to
(1.3) we can conclude that if u is harmonic on D the flux of its gradient out
of any smooth subdomain of D vanishes, and conversely, if u has continuous
second partial derivatives and if the flux of its gradient out of every ball
B with closure in D vanishes, then u is harmonic. (It is easy to see that it is
sufficient here if u has continuous first partial derivatives.)
N = 2. The real and imaginary parts of a (complex) analytic function
f = u + iv are harmonic. In fact, u and v satisfy Laplace's equation because
they satisfy the Cauchy-Riemann equations. Alternatively, the Cauchy
integral formula yields (3.1).

4. Maximum-Minimum Theorem for Harmonic Functions


Recall our convention that in the absence of another stated topology the
space RN with its usual Euclidean topology is supposed and is compactified
by a point at infinity, denoted by oo. This point oo is not included in IF"
but is included in the boundary of every unbounded subset of R".

Theorem. Let u be harmonic on the open subset D of R".


(a) If D is connected and if u attains its supremum or infunum at a point
of D then u is identically constant.
(b) The supremum and infunum of u are limits of u along sequences of
points approaching D.
6 1. 1. Introduction to the Mathematical Background of Classical Potential Theory

(c) If u has a continuous extension to D u 8D, the supremum and infunwn


of the extension are attained on 8D.

A typical implication of this theorem is the fact that a harmonic function


on D with limit 0 at every boundary point must vanish identically. To prove
(a), suppose that u attains, say, its infimum a at a point of D. Since (con-
tinuity of u) the set {ry: u(q) > a} is open, the harmonic function average
property (3.2) implies that u is identically a on 6) for b < I - 8DI.
It follows that the set of points of D at which u = a is open. Since continuity
of u implies that this set is also closed relative to D, this set is D if D is con-
nected. Thus part (a) of the theorem is true and parts (b) and (c) follow
easily.

5. The Fundamental Kernel for R' and Its Potentials


If ft=C12']a,b[ the function u defined by on the domain
{ : a < I I < b} is harmonic if and only if (denoting I I by r)

Au=./"(r)+N-

I f(r)=0. (5.1)
r
Thus if G is defined by

log I ifN = 2,
tl) =
I -II (5.2)

_1I N 2 ifN>2,
I

the function is harmonic on I8N - The function kernel G is the


fundamental kernel of classical potential theory. For D = 62N we shall
sometimes write Gp instead of G when N > 2 to match later notation for
Green's functions. _
If I - ryI < r < s, equation (1.3) with B = and u =
q) reduces to
Yd
0 = R's N-1
FF
dSL(u, S,s) - nNrN-' , r) (5.3)

so that sN-' d/ds L(u. , s) does not depend on s and since

limy 1._ +,, L(u, , s)/G(0, C) = 1,

logy ifN=2
L(G(', q), , r) = (I - 7I < r). (5.4)
r2-N if N > 2
6. Gauss Integral Theorem 7

On the other hand since G(-, 17) is harmonic on RN - {g}, the harmonic
function average property yields

L(G(', g), , r) = G( , g) when R - g l ? r. (5.5)

(The fact that this evaluation is valid when I - gl = r follows from an


easy continuity argument.)
If p is a measure of Borel subsets of RN the function Gp defined by

J (5.6)
RN

is the potential of p. We shall discuss the convergence of this integral later.


It is clear however that if p(R') < + oo, the integral converges absolutely
at every point not in the closed support A of p and thereby defines a con-
tinuous function on RN - A. The function is harmonic on this domain
because it has the harmonic function average property there.

6. Gauss Integral Theorem


Let D be a smooth domain, and let p be a signed measure supported by a
compact set A not meeting LID. Then

h(A n D) = --1 f Do
D
(6.1)

where
2n ifN=2,
nN- (6.2)
((N-2)aN ifN>2.
If A does not meet D the function Gp is harmonic on a neighborhood of
D so (6.1) is true because according to Section 3 the flux out of D of the
gradient of a harmonic function vanishes. The potential of the projection
of p on R' - D is covered by this remark so from now on we suppose that
A c D. If (6.1) is true for one choice of D, it is true for every smooth domain
D, containing D if D, - D is smooth because Gp is harmonic on a neigh-
borhood of D, - D so the flux out of D, - D of the gradient of Gp vanishes.
Observe that (6.1) is trivially true if D is a ball with center and if A =
and therefore, in view of the remark in the last sentence, (6.1) is true whenever
A is a singleton,

f D
n)lN-i (dg) I if e D. (6.3)

Integrating (6.3) with respect top yields (6.1).


8 1.1 Introduction to the Mathematical Background of Classical Potential Theory

See Section 7 for an extension of this theorem allowing the support of


y to meet OD.

7. The Smoothness of Potentials; The Poisson Equation

In the following theorem the coordinates of a point in RN are denoted by


"", ... , 'N). We shall use the inequalities

SNIA- lilt-x Nz2.


I a2Ga«)
(7.1)

Theorem. Let p be a signed measure on RN given by an IN measurable density


f relative to 1N, du = fd1N.
(a) If f is bounded and vanishes near oo then u = Gp is in class C"'t(RN)
and
au(k) _ r q)
a to - J RN (7.2)

(b) If in (a) there is an open set D on which f is continuous and satisfies


a Holder condition,

If(s) -f(7)I s constI - qI° 0 <p S 1, x D,

then uID is in class Cf21(D), satisfies the Poisson equation du = -ir4f, and

a2u( ) h)
lfi [f(7) -f()]lN(d11) - N a D. (7.3)
a_(oaej) = JaN

Here bf is the usual Kronecker symbol ; that is, (bif) is the identity matrix.

Proof of (a). Since is locally IN integrable, the integral for Gpconverges


absolutely under the hypotheses of (a). Define G 11k1 as G A log k when N = 2
and as G A kN-2 when N > 2. Then G'")p is continuous and

const !20-k k if N = 2,
_Galul <
IG,u
ifN>2.

Hence lim,1. G'"p = Gp uniformly on RN, and we conclude that u = Gµ


is continuous. In view of (7.1) the integral in (7.2) is absolutely convergent.
7. The Smoothness of Potentials; The Poisson Equation 9

Round off the integrand in (7.2) when it exceeds k""` in absolute value
to obtain a continuous finite-valued integrand and repeat the reasoning
proving the continuity of Gp to show that the right side of (7.2) defines a
continuous function of . This function must be

Proof of (b). In view of (7.1) and the Holder condition satisfied by Jon D
the integrand in (7.3) for n in a neighborhood of in D is majorized in
absolute value by constl - nI-"+"; so the integral converges absolutely.
Let /3 be a function from R+ into R+ satisfying the following conditions:

$eC"'(R+), 05fl51,
(0 ifOSr<,
/3(r)=1
I

and define

f
R"
a
(;,)n)1(n)fl(I' -
ni)1"(dn)

for a > 0. Then

au(k)
a a a;
I J t c.= I

so lima_, a) = uniformly on R". The function a) is in


C"'(R"). To prove (b) it is sufficient to show that, on D, lim, .0 a4,/a4'h'
is the right side of (7.3) and that the convergence is uniform on every compact
subset of D. To show this, write in the form

aft gj) [1(n) -I()]P


I_ -« nI r"(dn)
fRN

+1() rl (- ) IN*)
iN a'a'J) a

I 8G(, n) a nI
tit -
att, [!fin) -n
I
)]fl,

IJ
1"(dn)
+ a Jp" - nI
('RN aG(, ), (14 (J) - ncn
_aL) -171

+ J ) I __171
1"(dn)

and denote the four terms on the right by I, It, III, IV, respectively. Observe
that each integrand vanishes for n outside B(4, a).
10 1.1. Introduction to the Mathematical Background of Classical Potential Theory

The difference between I and the integral on the right side of (7.3) is at
most const Jo r"!, (dr) for a < I - BDI ; so when a - 0, the term I has the
integral on the right side of (7.3) as a limit uniformly on compact subsets
of D.
If i # j in II, the integral over a) vanishes because the integrand is
odd in nt". If i = j in II, the integral becomes

-(N - 2) J f(IX' -
(i
t e.Q)
(7.4)
0+N(N-2)
f
JB(.a) I - ql at

if N > 2 and a < I - aDl. The second integral does not depend on the
choice of i, and so it has as value the average of its values for i = 1, ... , N.
We conclude that the sum in (7.4) vanishes. The corresponding argument
when N = 2 shows that the integral II also vanishes in this case.
If of < I - ODl,

11111 < const


at f rp,' (r)/1(dr) 5 const a";
so when a 0, there is uniform convergence to 0 on compact subsets of D.
To evaluate IV observe that when i # j this integral over a) vanishes
because the integrand is odd in i'" - nt". If i = j and if N > 2 then

IV =
-fad) (N - 2) n"h2Q, (1 nl)
at
L4.0
The integral is the same for all i, and it is equal to the average of its values
for i = 1, ..., N. Hence

IV= - ,()1i(dr) I(')


aN 0
n"l)(N-2)Ja
f N

If N = 2 the corresponding evaluation of IV yields


and the proof of the theorem is now complete. 0
Extension. If v is a signed measure on I8" for which the integral defining
Gv converges absolutely and if the projection of v on some bounded open
set D is determined by a bounded density g relative to IN, then (Gv)1oe Ct"(D)
because if f = g1D and du = fdIN, then Theorem 7 is applicable to Gp
which differs on D from Gv by a harmonic function. The same argument
shows that if g is continuous on D and satisfies a Holder condition there,
then (Gv)1D(-C"'(D) and AGv = -i g there.
8. Harmonic Measure and the Riesz Decomposition 11

Extension of the Gauss Integral Theorem

Under the hypotheses of (b) the flux evaluation (1.3) yields, whenever B
is a smooth domain with closure in D,

p(B) _ - in"f AGItdl" _ - 1 f D,Gpol"_1. (7.5)


I n" ae

This evaluation generalizes the Gauss Integral Theorem (Section 6) in that


the support of the measure u is allowed to meet OB.

8. Harmonic Measure and the Riesz Decomposition

Let D be an open bounded subset of R". Suppose that JI E C'2)(D), and


suppose that Au is bounded on D. Define a signed measure y on D by du =
- Au dlN/n,. Theorem 7 implies that A(u - Gµ) = 0, so that u = (u - Gµ) +
G,u is the sum of a harmonic function and the restriction to D of the potential
Gp. This fact will now be proved in a slightly different version and developed
further.
If D is a smooth open subset of Rr, if c D, and if uEC12)(D),
an application of Green's identity to the pair of functions [u, on
the smooth open set D - B(i, S) yields

uI"-MW -
aatt.a)

= - (8.1)
J8D

q)

Since is bounded, the first integral on the left is majorized in absolute


value by const 61log 61 if N = 2 and by const 62 if N > 2; so this integral
tends to 0 when b -+ 0. The second integral on the left is equal to - n'" L(u, , S)
for N Z 2 and therefore has limit -ir u(g) when 6 - 0. Thus when b - 0
in (8.1), we find

u() = 1 [G(f, q)D u -


JaD
(8.2)
1 G( ,q)Au(l)I"(thl)
12 1.1. Introduction to the Mathematical Background of Classical Potential Theory

This representation of u is another version of that obtained above be-


cause the first term in (8.2) defines a function harmonic on D and the second
term is the potential of the signed measure with density -Au/nN. This
representation of u remains valid if is replaced by u(t, ),
where (i) e Ct21(D) for each in D and is harmonic on D. Suppose
that can be chosen to satisfy (i) and also (ii) on OD.
In this case the restriction to D of the difference GD(, ) = G(, ) - u(, )
is called the Green function of D with pole t, and GD is called the Green
function of D. The function u(, ), if there is such a function, is uniquely
determined because the difference between two such functions is harmonic
on D with limit 0 at every boundary point and therefore vanishes identically
(maximum-minimum theorem for harmonic functions). Recapitulating,
GD(, ) is to satisfy the following conditions:
(i') GD is defined on D x D and GD(, ) = + oo.
(ii') GD(, ) is harmonic on D - {}, and GD(, ) - G(, ) is harmonic
on D if defined suitably at .
(iii') GD(, ) has limit 0 at every point of aD.
(iv') GD(, ) - G(, ) if defined suitably at and, if defined as -G(, )
on aD, is in Ct2j(D).
An application of the harmonic function maximum-minimum theorem
to GD(, ) on D - {} shows that GD s 0. We shall show in Section 11.1
that a ball has a Green function in this classical sense. The existence of a
Green function satisfying conditions (i')-(iv') is a restriction on the smooth-
ness of aD, but in Chapter VII we shall define a Green function much more
loosely, keeping (i') and (ii'), weakening (iii'), and dropping (iv'). It will
be shown that every nonempty open subset of RN for N > 2 and every
not-too-large nonempty open subset of R2 (for example, every nonempty
bounded open set) has a unique Green function in the looser sense. The
present discussion shows what led to the more general definition and will
suggest theorems to be proved.
Suppose then that D is smooth and that GD exists satisfying (i')-(iv').
If v is a measure on D, the function GDV on D is called the Green potential
of v. Using GD instead of G in (8.2) and defining GD(, ) as 0 on aD reduces
(8.2) to

u(S) = n--
f'Du(I)DeGD(,)lN-10I) - NJDGD(,J)Au(1)lN(d+l)
(8.3)

In particular

I = - N J DGD(, )lN-1(d 1). (8.4)


D
8. Harmonic Measure and the Riesz Decomposition 13

Define the measure of Borel subsets of aD by

A) _ - JDGD(i.)lN_l(dn). (8.5)
N

Then ND(., ) is a positive measure because the normal derivative in the


integrand is negative, and aD) = 1. The measure is called the
harmonic measure relative to . The representation (8.3) can now be written

u(0 = JaD
f GD), dv = -AudlN (8.6)

In our later general treatment we shall prove the following facts:


(a) GD is symmetric and continuous on D x D(= + o0 on the diagonal).
(b) The function A) is harmonic on D for every A, and more
generally the function J8Df(r1)LD( an) is harmonic on D for every
bounded Borel measurable function f on OD.
These facts will be trivially verifiable when D is a ball (see Section 11.1),
the only case when they will be used explicitly before the general discussion.
We therefore proceed without proving facts (a) and (b) in the present
discussion. If u is harmonic on D the second term in (8.6) drops out; so

u() = J

u a weighted average of the values of u


on OD. In particular, if D is a ball with center , this averaging reduces to
the defining average property of harmonic functions.
More generally, if Au 5 0 and u > 0, (8.6) exhibits u as the sum of a
positive harmonic function and the Green potential of a positive measure.
One of the principal aims of the general theory is to generalize this result
(the Riesz decomposition, Section IV.8) by dropping the smoothness con-
ditions on u, D, and GD.
Chapter II

Basic Properties of Harmonic, Subharmonic,


and Superharmonic Functions

1. The Green Function of a Ball; The Poisson Integral


Let B = B denote by ' the image of under inversion in
B. That is, ' is on the ray from o through , and 1 - u 1 1 ' - X01 = 62.
To simplify the notation take o = 0. Then GB, as defined by

7) = forN=2 (1.1)
712-N
712-N
62-N ifS
(=1712-N -
= 0)
- - forN>2
with the understanding that oo, satisfies items (i')-(iv') of
Section 1.8, so that harmonic measure for B is given by
I,
D K(7Ni1N-1(d7), (1.2)

where 1N- 1 here refers to surface area on aB and

SN-2S2 - 1412 (1.3)

Hence, according to Section 1.8, if u is harmonic on a neighthorbood of B,

u( )= 7r b u(,)s2-1FFSI21N-I(d1)= nNSIN_, J u(7)K(7,t)IN-1(d7). (1.4)


N f
oB 17 - y l dB

The function K(7, ) is harmonic on RN - (7) because GB(-, 7) is, and K(7, )
is normalized to be I at the origin. The function GB is symmetric, positive
(= + oo when the arguments are equal), and increases with S. Moreover
1. The Green Function of a Ball; The Poisson Integral 15

+ oo if N = 2,
lim 17) _ (1.5)
lira
6-00 q) if N>2.
Since ry) is harmonic on B - {ry}, all the partial derivatives of this
function are also harmonic there so that in particular the harmonic measure
density in (1.2) defines a harmonic function of on B. It follows that if µ
is a finite measure of Borel subsets of aB, the Poisson integral (PI)

J (dq), (1.6)
a8

defines a harmonic function on B. It should cause no confusion if we write


PI(B,f) instead of PI(B,p) when t(dry) = j(ry)IN_1(dry)l(nNS" '). In later
applications the ball may not have the origin as center, and the Poisson
formula is modified accordingly. Moreover, if f is defined on a superset
S of the boundary of a ball B, we shall write PI(B,f) instead of PI(B, f ae)

Theorem. If f is Lebesgue integrable on aB, u = PI (B, f) is harmonic on B


and

lim sup lim sup (CeaB). (1.7)


4-4 4-C
teB 4EOB

Combining (1.7) with the corresponding inequality for inferior limits, it follows
that u has limitf(C) at C if f is continuous at C.

We have already noted that u is harmonic on B. Inequality (1.7) is a


special case of Theorem I of Appendix VII with X = aB and n -+ oo replaced
by C. The key fact is that for ry outside a neighborhood of C and on aB
the integrand K(ry, ) is at most const (62 - and this majorant tends
to 0 when in B tends to C.
Integral of K. The fact that JB K(ry, )IN(dC) = 1N(B) will be needed below.
To derive this evaluation observe that the value a = a(q) of this integral
does not in fact depend on ry; so

a = L(a, D, b) = I(B, 1)lN(dq) = IN(B).


fB P

Solution of the Dirichlet Problem for a Ball

If f is a finite continuous function on the boundary of a ball B, the function


u = PI(B,f) is harmonic on B and, according to Theorem 1 has limit f(C)
at every boundary point C. There is only one harmonic function on B with
16 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

this boundary limit property because the difference between two such func-
tions is harmonic on B with limit 0 at every boundary point and therefore
vanishes identically in view of the harmonic function maximum minimum
theorem. The function u is thus the unique solution of the classical first
boundary (Dirichlet) problem for harmonic functions on B. A generalized
form of this problem for the relevant class of open subsets of R' is treated
in Chapter VIII.
If u is a finite-valued function defined and continuous on the closure of
a ball B and harmonic on B, u = PI(B, u) on B because u is the unique
solution of the Dirichlet problem for the boundary function ulde. This result
weakens the conditions under which (1.3) was derived using the general
theory in Section 1.8.
If u is any Borel measurable function on an open subset D of R', if B
is a ball with closure in D, and if the restriction of u to 8B is IN_, integrable,
we define

(PI(B, u) on B,
tBu (1.8)
=1 u on D-B.
If u is upper or lower semicontinuous, Theorem I implies that tBu has the
same property.

2. Harnack's Inequality
Let A be a compact subset of the open connected subset D of RN. There is a
function (A, D)"c(A, D) such that if u is harmonic and strictly positive on D,

u()
< c(A, D) [(c, ry) e A X A]. ( 2. 1)
u(ri)

If b) c D, the representation u = b), u) yields

S2 - a2 6N_2 < 62 - a2 6N-2


(S + a)N u( o) (S - a)N
fl -Q 5a<b). (2.2)

Thus

u(S) < + a\N (2.3)


x B(So,a)].
u(r)-tbd-a
Harnack's inequality is therefore true if A c S). More generally, it
follows that Harnack's inequality is true if the compact subset A of D can
be covered by finitely many balls B,, ..., B5, each with closure in D, in
3. Convergence of Directed Sets of Harmonic Functions 17

such a way that B;,, n Bt # 0 for j = 1, ... , k - 1. Since D is connected,


every compact subset A of D can be covered in this way.

Application to Lower-Bounded Harmonic Functions on R"

A lower-bounded harmonic function v on R' must be identically constant


because the function u = v - inf v + 1 is a strictly positive harmonic function
on R" and (2.2) yields u(O) when b -' +co. (See Section 13 for the
extension of this result, when N = 2, to lower-bounded superharmonic
functions.) In particular, it follows that a bounded analytic function on the
plane is identically constant (Liouville's theorem).

Application to Local Properties of Families of Harmonic Functions

Harnack's inequality implies that a family u, of positive harmonic functions


on a connected open set D is locally uniformly bounded if bounded at a single
point. Moreover, an application of (2.3) with a near 0 shows that the family
u, + I and therefore also the family u. is equicontinuous at each point and
thus uniformly continuous on each compact subset of D. Trivially, more
generally, a locally uniformly bounded family t of harmonic functions on
an open set D is uniformly continuous on each compact subset of D. This
result can be strengthened as follows. Let f" be the family of partial deriva-
tives of some specified order of the members of r. These derivatives are
harmonic functions. Let B be a ball with closure in D. The representation
u = PI (B, u) can be differentiated to show that t' is uniformly bounded in
each smaller concentric ball, and therefore t' is locally uniformly bounded.
This same reasoning shows that F is locally uniformly continuous, as already
derived using Harnack's inequality.

3. Convergence of Directed Sets of Harmonic Functions


The following results are stated for directed sets (nets) of functions rather
than sequences because in potential theory convergence problems commonly
arise in the context of directed sets. In this section D is an open connected
subset of R' and u, is a directed set of harmonic functions on D.

(a) Harnack's Convergence Theorem. If u. is directed upward, with limit


u, there is locally uniform convergence on D to u, and u is either identically
+ oo or a harmonic function.

To prove the theorem, observe that if A is a compact subset of D and if


So E A, then by Harnack's inequality (2.1) if f a so that up Z u it follows
that
18 1 11 Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

c(A, D)"' sup [up(s) - 5


4eA (3.1)
c(A, D) inf [up(s) -

u is finite at a point, choose o as that point. Then (3.1) implies that u is


finite valued on D and that u, converges uniformly on A, so converges locally
uniformly on D. The limit function u is harmonic because it is continuous
and has the harmonic function average property. On the other hand, if u is
infinite valued at some point, choose o as that point, and then (3.1) implies
that u, converges uniformly on A, so locally uniformly, to +oo.
Observation. Whether u is identically + oo or harmonic in this convergence
theorem, some increasing sequence in u. also has limit u according to
Theorem 2 of Appendix VIII.
(b) If u, is locally bounded and converges to a function u, then the
convergence is locally uniform, the limit function u is harmonic, and if D is
any partial derivation operator, lirrme Du, = Du locally uniformly.
The fact that the convergence is locally uniform follows from the local
equicontinuity of u, (Section 2), and the last assertion follows on applying
the operator D to the limit equation limo PI(B, ua) = PI(B, u) with B a ball
with closure in D. If u, is a locally bounded sequence of harmonic functions
it is locally equicontinuous; so (by Ascoli's theorem) a subsequence converges
locally uniformly.

4. Harmonic, Subharmonic, and Superharmonic Functions


A function u from an open subset D of IN into ] - x, + oo] is called
superharmonic if
(a) u is lower semicontinuous.
(b) u is not identically + oo in any open connected component of D.
(c) z L(u, , 6) if b) c D.
Since u is necessarily locally lower bounded, the integral involved in (c)
is well defined and not equal to - oo. Applying (c) and 1(2.2) we find that
(c') A (u, , S) > - oo if 6) c D.
Hence finiteness of u at a point implies finiteness 1N almost everywhere
in 6) when 6) c D, and therefore a covering argument shows that
u is finite 1N almost everywhere on D and that u is locally 1N integrable. In
particular A(u, , b) is finite in (c') even when x. We shall show
in Section 6(a) that L(u, , 6) in (c) is finite even when ao, and
we shall show in Section 6(c) that (c') can replace (c) in the definitiop of
superharmonicity.
If u, and u2 are superharmonic functions on D and if c, and c2 are positive
constants, the functions c,u,, c,u, + c2u2, and u, A u2 are superharmonic
4. Harmonic, Subharmonic, and Superharmonic Functions 19

on D. Since superharmonic functions are lower semicontinuous, the limit


of an upward-directed family of superharmonic functions on D is the limit
of an increasing sequence of functions in the family (Theorem 2 of Appendix
VIII). The limit function u is superharmonic if condition (b) is satisfied.
A subiamonic function is defined as the negative of a superharmonic
function. A function is harmonic if and only if it is both subharmonic and
superharmonic. It is unfortunate that it is natural in pure potential theory
to consider superharmonic rather than subharmonic functions but that the
applications of potential theory are likely to involve subharmonic functions
rather than superharmonic functions.

Positive Integral Operations on Superharmonic Functions

Let D be a nonempty open subset of R", and let R(D) be the class of Borel
subsets of D. Let (up, ig a 1) be a family of superharmonic functions on D,
indexed by a set 1. If (1, .F, A) is a finite measure space and if the function
is measurable from (D x 1, R(D) x ,F) into (R,R(R)), then
the function u' = f r up A(dd) satisfies the superharmonic function average
inequality,

L(u', 4, b) = L uQA(dfl), , 6) = f L(up, t, b)A(d$) S


\Jr // Jr
(I - aDI > b),
if the double integral involved converges absolutely. Thus u' is superharmonic
if the superharmonic function finiteness and lower semicontinuity conditions
are satisfied.

EXAMPLES. Let b be a strictly positive number so small that

-aDj >b)
is not empty. Then if u is a superharmonic function on D the function
u' = L(u, , b) on Da is a special case of the preceding integral operation,
with u( + P), fl e aB(0, 6) and with A the normalized surface area
of B(0,6). In this case an application of Fatou's lemma shows that u' is
lower semicontinuous. We shall show in Section 6(a) that L(u, , 6) is finite
valued, and it follows that L(u, , 6) is superharmonic on D. We leave to the
reader the verification of the fact that if u is a superharmonic function on
D, the functions A (u, , 6) and Adu are also special cases of the above integral
operation. Since (from Section 2) A (u, , 6) is finite valued and continuous
and Aau is infinitely differentiable, these two functions are superharmonic
on D6.
For u superharmonic on D the three functions L(u, , 6), A (u, , b), and
Aau are all majorized by u, and it will be shown in Section 6(f) that these
20 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

functions increase monotonely to u when S -+ 0, in the sense that when


So > 0 and b < 60, there is monotone convergence on D60 to u when 6 -+ 0.

5. Minimum Theorem for Superharmonic Functions


Theorem. Let u be superharmonic on an open subset D of R'.
(a) If D is connected and if u attains its infunum at a point , then u is
identically constant.
(b) The infimum of u is the limit of u along a sequence of points approaching
OD.
(c) If u has a lower semicontinuous extension to D u 3D, the infunum of
the extension is attained on aD.

A typical implication of this theorem is the fact that a superharmonic


function on D with a positive inferior limit at every boundary point must
be positive. This theorem is the generalization for superharmonic functions
of the harmonic function maximum minimum theorem (see Section 1.4).
The proof of Theorem 5(a) is precisely the same as that given of Theorem
1.4(a) involving the infimum of u because that proof required only lower
semicontinuity of u and the superharmonic function average inequality.
Parts (b) and (c) follow easily from (a).

6. Application of the Operation TB


This operation, defined in Section 1, will be used to derive various important
properties of superharmonic functions.
(a) If u is superharmonic on D and if B is a ball with closure in D, then
u is IN _ 1 integrable on aB and r9u 5 u. In fact, since u is lower semicontinuous,
it is bounded below on iB, and there is an increasing sequence f of finite-
valued continuous functions on aB with limit uka. Then PI(B,f) is an
increasing sequence of harmonic functions on B with limit PI(B, u), and
an application of the superharmonic function minimum theorem to
u - PI(B, on B shows that the difference is positive there. Hence (n - oo)
u is IN _I integrable on aB and TBU 5 u.
(b) In (a) if B = B(g, S) and if u is harmonic on B.
In fact, under the stated conditions, so that the restriction to
B of u - r,u is positive superharmonic, vanishes at , and therefore vanishes
identically.
(c) In the definition of superharmonicity in Section 4, (c) can be replaced
by (c'). We have already seen that superharmonic functions satisfy (c').
Conversely, if u satisfies Section 4(a), (b), (c') the reasoning leading to the
superharmonic function minimum theorem and thereby to (a) above remains
6. Application of the Operation rB 21

valid, and then if B = b), the inequality TBU() S u(i;) is condition


Section 4(c).
(d) In the definition of a superharmonic function in Section 4 condition (c)
[or equivalently (c')] there need be supposed true only for sufficiently small
b, depending on , because the reasoning leading to (a) of the present section
used only this weakened condition, and (a) implies Section 4(c) when
B= b).
(e) In (a) the function rBu is superharmonic on D because according to
Section 1 this function is lower semicontinuous, and it is trivial that for
i; in D the inequality z L(TBu, g, b) is valid for sufficiently small b.
More generally, if B is a ball with closure in D, if u is superharmonic on D,
and if v is a function defined and superharmonic on B, with v z TBU on B,
then the function v' (equal on B to v A u and on D - B to u) lies between
u and TBU from which it follows easily that v' is superharmonic on D.
(f) If u is superharmonic on D and if e D, the functions b - L(u,, b),
b i-+ A (u, ia, bland b i--+ Aau() are monotone decreasing for 0 < b < 1 - OD I,
with limit when b -+ 0. To prove the monotoneity of L(u, c, -) observe
that if B; = B(c, b,) with b, < b2 < I - aD1, then the inequality TB2TB,u 5
TB, u reduces at 1; to L(u, , b2) < L(u, s, b,). Next apply the lower semicon-
tinuity of u and Fatou's lemma to derive the inequality

S lim L(u, l;, b) 5


a-0
and thereby complete the proof of (f) for L(u, c, ). The corresponding
results for and follow from 1(2.2) and 1(2.4). The latter
results imply that the relation u >- v or u = v, if satisifed IN almost every-
where on their domain of definition D by superharmonic functions u and v,
is satisfied everywhere on D.
The preceding results imply the truth of a slight strengthening of the
lower semicontinuity property of a superharmonic function u, namely,

u(S) = liminfu(,) = lim infu(q), (6.1)


rC nee
q*A

where A is an arbitrary iN null set. We shall see in Section XI.1 that the
natural class of sets A making (6.1) true is the class of sets thin at in the
sense of the fine topology.
Observation. If b > 0 and if 06 is a function on the interval [0, b], Borel
measurable and 11 almost everywhere strictly positive, with 160 Cd(r) dr = 1,
then every superharmonic function u on an open set D obviously satisfies
the inequality

u() z f d 1e(r)L(u, , r) dr (6.2)


0
22 111. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

when b) c D. Now suppose that So > 0 and that ¢a is defined and


satisfies the conditions imposed on ie above, for 0 < b < bo. Then if u is a
function on the open set D, satisfying the superharmonic function defining
conditions (a) and (b) of Section 4, together with (6.2) for sufficiently small
b, depending on , the function is superharmonic. This assertion has already
been proved for 4a(r) = Nb-l r"-' , in which case the right-hand side of (6.2)
reduces to A (u, , b), and the proof needs no change in the general case.

EXAMPLE (The Fundamental Kernel and the Green Function of a Ball). If the
fundamental kernel G is defined on R" x R", by 1(5.2) the function is
for each point harmonic on R" - and is superharmonic on R". In fact
we have noted in Section 1.5 that this function is harmonic on R" -
Moreover this function is continuous on R", and in view of the local nature of
superharmonicity proved in (d) above we need only observe, to prove that
the function is superharmonic on R", that the superharmonic function
average inequality is trivially satisfied at . Similarly, if B is a ball, the Green
function GB is defined on B x B by (1.1), and for each point t of B, GB(4, )
is harmonic on B - and superharmonic on D.

7. Characterization of Superharmonic Functions in Terms of


Harmonic Functions
It is important to characterize superharmonicity intrinsically, without the
use of balls. Let u be a lower semicontinuous function from the open subset
D of R" into ] - oo, + oo], not identically + oe on any open connected
component of D. Consider the following property of u: if Do is a relatively
compact open subset of D and if v is a function defined and harmonic on a
neighborhood of b, with u - v -: 0 on OD,, then u - v > 0 on Do. This
condition is necessary and sufficient for u to be superharmonic and justifies
the name "superharmonic." In fact a superharmonic function on D has this
property according to the superharmonic function minimum theorem.
Conversely, if u has this property, let B = B(4,6) be a ball with closure in D.
It is enough to prove that L(u, a:, b) and even, since u is lower semi-
continuous, to prove that z L(f, , b) for every finite-valued continuous
function f defined on aB and majorized by u. For such a function f the
function v = PI(B,f) is harmonic on B with a continuous extension to h
obtained by setting v = f on B. The difference u - v is lower semicontinuous
on B and positive on aB, so that ifs > 0, then u - v z -E near 6B (for
example, on aDo, for Do a slightly smaller concentric ball). The given
condition implies that u - v z -E on Do and therefore u - v z -E on B, in
particular at ; that is, L(f, 4, b) - E. Since E can be arbitrarily small,
the desired inequality is true.
Whenever harmonic functions can be defined, for example, on a Riemann
surface by Laplace's equation, superharmonic functions can be defined using
the intrinsic condition of this section.
9. Application of Jensen's Inequality 23

8. Differentiable Superharmonic Functions


Theorem. If D is an open subset of R' and if u E V ')(D), then u is superharmonic
if and only if Au 5 0.

Since linear functions and products of two different coordinate functions


are harmonic, the Taylor expansion of u about a point of D yields
z
L(u,, S) = u() + 0(62), (8.1)
2N

so that Au :!g 0 when u is superharmonic. Conversely if Au :r. 0 define


Then A(u - Ev) < 0 when e > 0, so that by (8.1) with u - Ev
instead of u,

L(u - Ev, , S) < (8.2)

for sufficiently small b. Then u - ev is superharmonic; so (8.2) is valid


whenever - aDI > 6, and (8.2) becomes the superharmonic function
1

inequality for u when c -' 0.

Approximation of a Superharmonic Function by Differentiable


Superharmonic Functions

If u is a superharmonic function on the open subset D of R", the function


Aau (see Section 1.2) is defined and infinitely differentiable on the set
{ a D: I - 3DI > S}. According to Section 4 this function is superharmonic
and majorized by u, and according to Section 6(f), for each point of D
the function b < 1 - aDl. and
lima.0 Aau = u. This approximation result will be improved in Section I V.10.

9. Application of Jensen's Inequality


This inequality implies that if 0 is a convex function

0[L(u, , S)] 5 L(O(u), , b), (9.1)

whenever u is Lebesgue measurable on 6) and all integrals involved


exist. This inequality has the following consequences:
(a) If u is harmonic and ¢ is convex, O(u) is subharmonic.
(b) If u is subharmonic and ¢ is convex and monotone increasing,
¢(u) is subharmonic.
Thus if u is harmonic or is positive and subharmonic, JuI" is subharmonic
24 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

when p,2! 1. If u is positive and superharmonic, u° is superharmonic when


0<p<1.

10. Superharmonic Functions on an Annulus


Theorem. Suppose that 0 5 a < b. let D_ a < I I < b} be an annulus in
98', and let u be a function from D into A.
(a) If u has the form u
u is a concave function of 0) (that is, a concave function of
- log I if N = 2 and of 14 1'- N if N > 2). In particular, u is harmonic
if and only if u is a linear function of 0).
(b) If u is a superharmonic [harmonic] function on D, the function
'-+ L(u, 0.1 1) is also superharmonic [harmonic] on D, and therefore
the function S i--+ L(u, 0, S) on ]a, b[ is a concave [linear] function of
-log 6ifN=2and of (2-NifN>2.
Observe that u is a concave function of - log I I if and only if u is a concave
function of log I -

Proof of (a). Since we can replace u by AQu and then let o(-+ 0, we can suppose
that f in (a) is infinitely differentiable. The following evaluation of Au makes
(a) trivial.
N-I
Au(k) =f'(I0 + P101
Ici

d2 ds2
if N=2 an d s= - log IC I, ( 10 . 1 )

(N - 2)2 d2f(II)
x Iy12N-2 if N > 2 an d s __1C12-N .

ds2

Proof of (b). We prove (b) for u superharmonic; the proof for u harmonic
is easier and is left to the reader. Since u is locally lower bounded we can
assume in the proof of (b), at the expense of increasing a and decreasing b,
that u is lower bounded on D. In addition suppose first that u is bounded
on D. Since a space rotation around the origin preserves superharmonicity,
integrating over all rotations (see the remark in Section 4 on positive integral
operations on superharmonic functions) yields the fact that the function
F-+ L(u, 0, ICI) is superharmonic on D, as asserted in (b). If u is not bounded
and n e Z', the function u A n is a bounded superharmonic function on D;
so the function - L(u A n, 0,1 I) is superharmonic on D and when n co,
we find that the function i-- L(u, 0, I I) is either superharmonic or identically
+ oo. The latter case is excluded because u is locally !N integrable so (by
Fubini's theorem) L(u, 0, b) < + oo for !1 almost every 6 in ] a, b[.
!i. Examples 25

The Minimum Function of a Superharmonic Function

Let D be as in Theorem 10, let u be a superharmonic function on D, and


define

m(b) = min {u(): ICI = b}.

We now show that the function is superharmonic on D; that is,


the function 6 on ]a, b[ is a concave function of -logb if N = 2
and of bZ-' if N> 2. First observe that the function is lower
semicontinuous on D because u is. Next observe that if y is an arbitrary
rotation of D about the origin, the rotated function - is super-
harmonic on D, and the infimum of the class of all these rotated functions
is the function Since an elementary argument shows that the
infimum of any locally lower bounded family of superharmonic functions
on an open set is superharmonic if this infimum is lower semicontinuous,
we conclude that the function r m(l' ) is superharmonic, as asserted.

11. Examples
(a) Suppose that N = 2 and let f be a not identically vanishing analytic
function on the connected open set D. The real and imaginary parts of f
are harmonic because they satisfy Laplace's equation, alternatively because
the Cauchy integral formula applied to a ball yields the harmonic function
average property for f. Taking absolute values in this average relation we
conclude that If I is subharmonic and therefore (Section 9) that If I° is
subharmonic when p z 1. Actually If I° is subharmonic when p > 0 because
if I° = I f °I is the absolute value of an analytic function in a neighborhood
of a nonzero of f, and it is trivial that the subharmonic function average
property is satisfied at a zero off. Similar reasoning shows that log if I =
Re (log f) is subharmonic when defined as - oo at a zero off and is harmonic
on the nonzero set off. Since If 1n for p > 0 is a monotone increasing convex
function of log If I, we have again that If I° is subharmonic.
(b) If N = 2 and if f is analytic on B(0, b) and does not vanish identically,
and if p > 0, the function If I° is subharmonic so
l zx

If(reie)I°& (r<b) (11.1)


0

is an increasing function which is a convex function of logr. Define M(r) =


maxlsj=,If(z)I and let C. be a sequence dense on the unit circle {IzI = 1}.
Then if u (z) = log I define v = uo v v up to get an increasing
sequence v, of subharmonic functions with limit the function z i-s v(z) =
26 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

log M(Iz!). Since r i-. L(v,,, 0, r) is a convex function of log r, the same is true
of the function r f--p log M(r). See a slightly different derivation of this same
property of in the context of the minimum function of a superharmonic
function in Section 10. Alternatively, the continuous function v is subhar-
monic because it has the subharmonic function average property and there-
fore the function

r i-- log M(r) = L(v, 0, r)

is a convex function of log r. This property of is known as the Hadamard


three circle theorem.

12. The Kelvin Transformation (N:2! 2)

Inversion in a sphere b) is the transformation of the one-point


compactification RN u { co } onto itself which takes the point into the
point S' on the same ray through o, with I - o! dal = b2, under the
convention that the points o and ao are interchanged. This transformation
is its own inverse. From now on to simplify the notation we take 4o = 0.
Let D be an open subset of RN, and let D' be the set of finite points of the
image of D under inversion in M(0, 6). If u is a function on D define u' on
D' by Then if ueC'Z'(D) and if A' is the Laplacian in the
variable 4',
!S'IN+z

Au = A 'IN-z64
).
(
Thus the function

vW) = u W)(!S'I)N-2

is harmonic on D' when u is harmonic on D. The characterization (Section


7) of superharmonicity in terms of harmonicity shows that v is superhar-
monic on D' if u is superharmonic on D. The function transformation from
u into v is called the Kelvin transformation (relative to the inversion sphere).
The Kelvin transformation is its own inverse.
If N = 2, if rp is an analytic not identically constant function on a con-
nected open set D, and if then Au(4) = I¢'I2(Au)(0). Thus u(4)
is harmonic on D if u is harmonic on ¢(D), and as in the preceding discussion
superharmonicity is also preserved. If 0 is replaced by the same argument
is applicable, and this case includes inversion in a circle, already discussed.
14. Classes of Harmonic Functions on a Ball 27

13. Greenian Sets


It is useful to have a special name, "Greenian set," for an open subset of
R" which supports a positive nonconstant superharmonic function. If N > 2,
every nonempty open subset of R" is Greenian because G(0, ) is positive
and superharmonic on R". If N = 2 and D is not dense in R2, the set D is
Greenian because if is an inner point of R2 - D, the function c -
is positive and harmonic on D for large c. In particular, if D is any open
disconnected subset of R2, each open connected component of D is Greenian,
as is D itself. On the other hand, R2 is not Greenian. To prove that a positive
superharmonic function u on R2 must be identically constant, we can replace
u by A,u and thereby suppose that u is infinitely differentiable. Then if
D = B(0, S) in 1(8.2) we find, with the help of 1(1.3),

u(O) L(u, 0, b) -1 `D log - I Au(ry)l2 (dn), (13.1)


2n In

and Au S 0. When b -p + co, this equation becomes impossible unless


Au - 0; so u must be harmonic and, according to Section 2, must therefore
be identically constant. [Alternatively, the fact that u must be harmonic
follows from the fact that (Section 10) the function b F- L(u, , b) is a positive
concave decreasing function of log S for 0 < S < + oo and so must be
identically constant.]
Application. Liouville's classical theorem that a bounded holomorphic
function f on the complex plane is identically constant is a special case of
the fact that the plane is not Greenian. We need only observe that the real
and imaginary parts of the function f are bounded harmonic functions and
therefore can be made positive by addition of suitable constants.
Observe that the trace on R' of the image of a Greenian subset of RN
under inversion in a sphere is Greenian.

14. The L' (µB_) and D(PB_) Classes of Harmonic Functions


on a Ball B; The Riesz-Herglotz Theorem
In a classification to be given in Chapter IX the harmonic functions charac-
terized in part (a) of the following theorem will be the harmonic members
of a class of functions denoted by L'(p .) and the harmonic functions
characterized in part (b) will be the harmonic members of a class of functions
denoted by D(pB_). The theorem will be a model that suggests corresponding
characterizations in the context of relative harmonic and parabolic functions
defined on wide classes of open sets and in the context of martingales. (See
Chapter I of Part 3.)
28 1.11 Basic Properties of Harmonic. Subharmonic, and Superharmonic Functions

Theorem. Let u be a harmonic function on B = B(0, b).


(a) L'(uB_) harmonic functions. The following conditions on u are
equivalent.
(al) u = PI(B, M,) for some signed measure M. on 8B.
(a2) u is the difference between two positive harmonic functions.
(a3) Jul has a harmonic magorant.
(a4) sup,,6L(Iul,0,r) < +oc.
Furthermore the map u - M. is a one-to-one, linear, order preserving
map from the class of L`(ua_) harmonic junctions onto the class of
signed measures on 8B, with

Gm (i) IN fdMa (14.1)


L b Nb
"-, J aB

for every continuous function f on M.


(b) D(p3_) harmonic functions. The following more restrictive conditions
on u are equivalent.
(bI) u = PI(B,f) for some IN-, measurable and integrablefunction
f on O B.
(b2) There is a uniform integrability test function (Appendix V)
4i for which ((IuI) has a harmonic majorant.
(b3) The family {ulatrto.,,, IN-0 < r < b} of paired func-
tions and measures is uniformly integrable.
(b4) (If u > 0) u is the limit of an increasing sequence of bounded
positive harmonic functions.
Furthermore the map u i- f is a one-to-one, linear, order-preserving
map from the class of D(uB_) harmonic functions on B onto the class
L'(aB,IN t),

and D(lui) in (b2) are subharmonic functions; so


L(l ul. 0, ) and L(b(Iul ), 0, ) are monotone increasing functions on ]0, b[.
In (a) the signed measure M. is the zero measure if and only if u = 0, and
M. z 0 if and only if u z 0. In (b) the functionJ vanishes IN_, almost every-
where if and only if u = 0, and f, z 0 IN_, almost everywhere if and only
if u z 0. The functionf will be identified in Section 15 as the nontangential
boundary limit function of u.

Nomenclature. The fact that a positive harmonic function on a ball has a


Poisson-Stielijes representation (al) with M. z 0 is usually called the
Herglotz theorem, although Herglotz himself referred to Riesz. To avoid
confusion this theorem will be called the Riesz-Herglotz theorem in this
book.
(al) . (a2) If M. = u, - u2 is a Jordan decomposition of M. into the
difference between two (positive) measures, the equation
14. Classes of Harmonic Functions on a Ball 29

u = PI(B,u,) - PI(B,u2)

expresses u as the difference between two positive harmonic functions.


(a2) . (a3) If u = u, - u2 with u, positive and harmonic, the function
Jul has the harmonic majorant u, + u2.
(a3) (a4) If Jul has the harmonic majorant v and if O < r < S, then

L(l u1, 0, r) < L(v, 0, r) = v(0).

(a4) (a 1) If 0 < r < S, the function -4 is harmonic on a


neighborhood of B and so is given on B by a Poisson integral,

(dq)
u = PI(B, u,), u (!4) I xNbN-'.
N-' (14.2)
S

The total variation of p, is

J
ju (a
)I'n ,-
dNdh)=L(lul,0,r),

and therefore, in view of the hypothesis of (a4), there is a strictly monotone


increasing sequence r, with limit S for which the sequence u,, of signed
measures has a vague limit, denote this limit by M. When r - S along this
sequence in (14.2), the Poisson representation becomes u = PI (B, M.). To
show that this representation of u on B determines M. uniquely, we show
that any representation u = PI(B, u) with p a signed measure on aB implies
that vague lim,.su, = u when u, is defined by (14.2). That is, we show (14.1):

limf fdu,=f fdu


'-a ae ae

whenever f is continuous on aB. Using the fact that J i7 - r/Sl = l - rq/Sl


for and q on 3B, an interchange of orders of integration yields

Jfd=e as
PI(Bf) C) u(dq). (14.3)

By Theorem 1 the right-hand side has limit Jd8 fdu when r - S, as was to
be proved.
The relation between L`(uB_) harmonic functions on B and signed mea-
sures on aB is obviously linear and positivity preserving. If the space of signed
measures on aB is ordered by setting u, 5 P2 when the difference u, - u2 is
positive, the relation u . M. is order preserving. Since the space of signed
measures on aB is a conditionally complete vector lattice (Appendix IV.7),
30 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

the space L' (µB_) of harmonic functions is also a conditionally complete


vector lattice (in the order determined by pointwise inequality). The vector
lattice isomorphism deduced here has an exact analog for harmonic func-
tions on Greenian sets provided with their Martin boundaries (Section
XII.9).
(bl) =:-. (b2) Choose a uniform integrability test function 0 for which
f eBO(I f I) d!N _, < + oc. Then by Jensen's inequality

4)(IuI) = ()[IPI(B,.ff)I] 5 PI(B, 1(I f I)). (14.4)

The last term is a harmonic majorant of 4)(IuJ).


(b2) . (b3) Let v be a harmonic majorant of (D(Iul). Then

L(O(Iul),0,r) 5 L(v, 0, r) = v(0);

so (b3) is true.
(b3) . (bl) Since (b3) is stronger than (a4), u = PI(B, M.), and we have
proved above that the signed measure u, tends to M. (vague convergence)
when r S. Now the uniform integrability described in (b3) is equivalent to
uniform integrability of the family i < r < S} of functions on
aB relative to the measure IN_,. Hence (Appendix V) there is a strictly
monotone increasing sequence r, with limit S for which the sequence

q- u(r.nlb)
has a weak limit function f on aB in the sense that

lim u (rte 9(17)IN-i(dn) = J f(rl)9(q)IN-r(dt) (14.5)


dB \ aB

for every bounded IN_, measurable function g on aB. This limit relation
implies (take g continuous) that the sequence u,., already known to be
vaguely convergent to M., is vaguely convergent to the signed measure
fdlN_,/(7r,6N-'). That is, M. is absolutely continuous relative to IN_, and
dM = fd1N-t/(7rNS"-'). Thus (bl) is true and f. is uniquely determined up
to IN_, null sets because M. is uniquely determined. The map u I-+f has
the stated properties because the map u F- M. has the corresponding
properties.
(bl) (b4) Under (bl), if u > 0, u = PI(B,f) with f. z 0, according to
what we have proved, and u = A n) represents u as the limit
of an increasing sequence of positive bounded harmonic functions.
(b4) (bl) If u is bounded positive and harmonic on B, condition (b2)
is satisfied, so (b I) is satisfied. Moreover, if u, is an increasing sequence of
positive bounded harmonic functions on B with limit u, the sequence f,4
15. The Fatoi. Boundary Limit Theorem 31

is an increasing sequence (up to 1N_, null sets) with limit some function f,
and the equation u = PI(B, f ) becomes in the limit u = PI(B, f ).
Observation. If u e D(µo_) in the theorem, we have proved that lim,..ap, _
M. (vague convergence of signed measures on OB), where p, is the signed
measure defined by (14.2). This convergence, or alternatively, a slight varia-
tion of the convergence proof, shows that if v, is the signed measure on B
supported by aB(0, r) and defined by

v,(dry) = (r = 1q1),
N

then lim,.av, = M. (vague convergence of signed measures on B).

15. The Fatou Boundary Limit Theorem


Let B = B(0, 6) in RN. A "Stolz domain" in B with vertex C on aB is defined
as the intersection of B with an open cone of revolution with vertex 1, axis
of rotation the ray from C through the origin, half-angle < n/2. A "deleted
nontangential neighborhood" of C is defined as a subset A of B with the
property that if S is a Stolz domain with vertex C, then S r B(C, e) c A for
sufficiently small s, depending on S. The class of deleted nontangential
neighborhoods of C is closed under finite intersections. A function w on B
is said to have radial limit q at C if lim,.,, w(rC) = q. The function is said to
have nontangential limit q at C if lime., q whenever the approach to
C is in a Stolz domain with vertex C, that is, if w has limit q at C along the
filter of deleted nontangential neighborhoods of C.
A signed measure p on the Borel subsets of OB can be considered as a
signed measure of Borel subsets of R', supported by aB. In particular,
(N - I)-dimensional "area" on OB when so extended to R" will be denoted
by C,v. If u and v are signed measures on R" supported by aB, the symmetric
variational derivate dp/dv (see Appendix IV. 12) can thus be defined as the
symmetric derivate for measures on R" or equivalently in the obvious way
for measures on OB.

Theorem. If u e L' (p _) and if h is a strictly positive harmonic function on B


then if Z e aB,

(nontang) Sim u() = M'(C) (symmetric variational derivate) (15.1)

whenever both the indicated derivate and the symmetric derivate dl'N/dM,,(C)
exist.
32 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

According to this theorem the limit in (15.1) exists at M,, almost every
boundary point C.
Observation. In the classical version of this theorem, as proved by Fatou,
h =_ 1, so that M,, is a normalized Lebesgue measure lN_1/(1rN6N-1) on sB.

In particular the theorem implies that if f is an IN_, measurable and integrable


function on aB then the harmonic function PI(B, f) has nontangential
boundary limit f(C) at IN_, almost every point C of B.

Reduction of the Theorem to a Special Case


Suppose that it is known that u/h has radial limit 0 at a boundary point C
whenever
(RI) uz0,
(R2) The symmetric derivate exists and is 0 at
(R3) The symmetric derivate dlN_1(dMa exists and is finite at C.
It will now be shown that then the theorem follows. In the first place, when
u z 0 and u/h has radial limit 0 at C then u/h also has nontangential limit 0
there. In fact let S be a Stolz domain with vertex C and let S, be a second
Stolz domain with vertex C but with a larger vertex angle than S. Choose p
with 0 < p < 1 and let B,,[B1,,] be the ball with center pC, internally tangent
to aS[aS,]. According to the Harnack inequality for balls as applied to u
and h in B,o there is a strictly positive constant c with the property that

u( )
5 u(PC)
eB,,.
ch(PC)'

The constant c does not depend on p because the constant in Harnack's


inequality is unaffected by a similarity transformation of I. It follows that
u/h has limit 0 on approach to C in S and therefore on nontangential approach.
In the second place we need only remark that if the derivate value in (15.1)
is a, to reduce the general case to that of positive u and a = 0 we can replace
M. by IM. - aM,,j, that is, replace u by the corresponding vector lattice
maximum (u - ah) Y (ah - u) furnished by Theorem 14.
To finish the proof of the theorem, we prove that under (Rl)-(R3) the
function u/h has radial limit 0 at C. Let a (9) and a,,(6) be, respectively, the
M. and M,, measure of the closed spherical cap on aB cut off by the closed
cone of revolution with vertex the origin, axis the radius to , half-angle
0 > 0, and define a (0) = a,,(0) = 0. The functions a and a,, are monotone
increasing on [0, n] and right continuous except perhaps at 0, and a,(6) > 0
when 0 > 0 in view of (R3). To simplify the notation take 6 = 1. Define

1 I -s 2
(15.2)
K:(e) = nN [(I - s)2 + 4ssin'(0/2)]x12
16. Minimal Harmonic Functions 33

so that

u(se) = JK(O)dcc(8), h(sC ) = J0K3(0)d0 (15.3)

The function K, is a decreasing function on [0, n],

sin' (a/2)01-2
o liminfKs(e)_N-2 lt(dO) = Jo !1(d9) _ +op (a > 0),
I S-1 K.(a) sin' (0/2)
(15.4)

and according to hypotheses (R2) and (R3),


ON-1
lim ato) = 0, lim / < +oo. (15.5)
0-0 ap(e) 8-0 ap(0)

According to Appendix VII.3 with f 1, p = N - 1, and n -+ oo replaced


by s -+ 1, relations (15.4) and (15.5) imply that Gm,.t u(sC)/h(st) = 0, as was
to be proved.

16. Minimal Harmonic Functions


A harmonic function u on an open connected subset D of RN is called
minimal if u is positive and if every harmonic function v on D satisfying the
inequality 0 S v 5 u is a constant multiple of u. If u is minimal, the positive
multiples of u are also minimal. If to e D and if r is the class of positive
harmonic functions on D with value I at 0, the class r is convex, and the
minimal members of r are the extreme elements of this convex set. In fact,
if u is minimal and in r and if u is not extremal, then u has a representation

u=put+qu2, p+q=1, 0<p<1, (16.1)

with u1 and u2 distinct elements of r. But then put and qu2 must be multiples
of u with values p and q, respectively, at 0, so u = u1 = u2, contrary to
hypothesis. Conversely, if u is an extremal element of t and if v is harmonic
on D with 0 5 v 5 u, then either (a) v vanishes at o and so vanishes iden-
tically or (b) u - v vanishes at o and so vanishes identically or

(c) u = v( o)1v( o)I + MW - V( o)] {U( o)}.

) - v(
The two functions in the braces are in r and so must be identical. Thus v
is a constant multiple of u in all three cases.
34 1.11. Basic Properties of Harmonic, Subharmonic, and Superharmonic Functions

EXAMPLE. Let D = B(0, 6) and let K be the normalized Poisson kernel density
function,
a2 Ibl2

K(,7, ) = 6N-2 q e aB(O, 6), e B(0, b). (16.2)

The class of minimal harmonic functions with value 1 at the origin is


{K(,, ), q e aD}. In fact K(q, ) is minimal because the Riesz-Herglotz
measure of this function is supported by the singleton {q} and the corre-
spondence between positive harmonic functions and their Riesz-Herglotz
measures is order preserving; so any positive harmonic function majorized
by K(q, ) also has Riesz-Herglotz measure supported by {q}. Conversely,
suppose that u is minimal with Riesz-Herglotz measure p. Unless p is
supported by a singleton {q} in which case u is a multiple of K(q, -),,u is the
sum of two nonidentically vanishing measures, say u = ,u, +021 supported
by disjoint sets, so that µ, is not a multiple of u, and µ, is then the Riesz-
Herglotz measure of a positive harmonic function majorized by u but not
a multiple of u. The point is that the minimal measures on aB(0, 6) in the
obvious definition must correspond to the minimal harmonic functions, and
the minimal measures are those supported by singletons.
Chapter III

Infima of Families of Superharmonic


Functions

1. Least Superharmonic Majorant (LM) and Greatest


Subharmonic Minorant (GM)
If F is a class of extended real-valued functions on a set D, a function v
on D will be called a majorant [minorant] of IF if v z u [v :s u] for every
u in r. If D is an open subset of R', the least superharmonic majorant
[greatest subharmonic minorant] of f, if such a function exists, will be
denoted by LMDf [GMDr], or by LMDU [GMDu] if f = {u} is a singleton.

Theorem. If a superharmonic function u on D has a subharmonic minorant,


then GMDu exists and is harmonic.

Let B. be a sequence of (not necessarily distinct) balls with closures in


D and with the property that if i; is in D, some neighborhood of is contained
in infinitely many of the balls. Define u = TB. rBau. Let u be the limit
of the decreasing sequence u. of superharmonic functions on D. If v is a
subharmonic minorant of u,

Un > rBn ... TBOV Z V,

and if is in D, some subsequence of u., in some ball with center , is a


decreasing sequence of harmonic functions. It follows (Hamack convergence
theorem) that u is harmonic and majorizes v, and therefore u = GMDU.
Without the hypothesis that u has a subharmonic minorant the limit
function um is either identically - oo or harmonic on each connected open
component of D and, if harmonic on D, is GMDU.
Observation (a). If D is a ball or is 68', a simple choice of B is an in-
creasing sequence of balls with closures in D and with union D. Choice of
B as an increasing sequence of open relatively compact subsets of D with
union D will be possible for every open set D once ie is defined (in Section
VIII.11) whenever B is an open relatively compact subset of D.
Observation (b). As the proof of the theorem shows,
36 1.111 Infima of Families of Superharmonic Functions

GMD(c1u1 + c2u2) = cIGMDUI + c2GMDU2

for superharmonic functions u1, u2 and positive constants c1, c2.

EXAMPLE (a). Let u be superharmonic on D and let B = S) have closure


in D. Then GMB(ujB) = PI(B,u). We already know that there is inequality
(z) here because PI(B, u) = (tBU)IB 5 u. If B. = S - 1/n), the minorant
in question is u. = The function u. is harmonic on B and
majorizes PI(B, u), and there is equality at 1: because

L(u, (1.1)
R-M
)
Hence (harmonic function maximum-minimum theorem) there is equality
on B, as asserted.
EXAMPLE (b). 0. In fact, if D has a Green function GD in
the sense of Section 1.8 then GD(l;, ) is a positive superharmonic function
on D with limit 0 at every boundary point. The function is
positive and harmonic and also has limit 0 at every boundary point. It
follows (harmonic function maximum-minimum theorem) that this minorant
vanishes identically. Exactly the same proof, in which oo is the only boundary
point, shows that 0 when N > 2. In Section VII.1 the Green
function Go will be defined for every Greenian set D. The property
0 remains true and in fact is very nearly a defining property
of GD.

2. Generalization of Theorem I
Theorem 1 is included in the following theorem but was proved separately
because of the importance of its constructive proof.

Theorem. If a class r of superharmonic functions on an open subset D of RN


has a subharmonic minorant, then GM IF exists and is harmonic.

Let F. be the class of subharmonic minorants of F. The class ro contains


u, v u2 with u t , u2 and is therefore directed upward in the order of pointwise
inequality, with limit u', a function majorized by r. We prove the theorem
by showing that u' is harmonic. Let B be a ball with closure in D and suppose
that v e ro. For every function u in F, v S tBV 5 r8u 5 u. Thus rBv a ro,
and the supremum of ro is the same as that of {rBv: ve ro}. This class on B
is an upward-directed family of harmonic functions majorized by r on B;
so u' is harmonic on B (Harnack convergence theorem) and therefore on
D. as was to be proved.
3. Fundamental Convergence Theorem (Preliminary Version) 37

3. Fundamental Convergence Theorem (Preliminary Version)


If u is a function from a topological space into IR, we denote (Appendix
VIII. 1) by u the lower semicontinuous smoothing of u, that is, the maximal
lower semicontinuous minorant of u:

u( ) = u() A Iiminfu(q). (3.1)


rc
In the following theorem the trivial inequality u 5 u is stated for complete-
ness and to facilitate reference.

Theorem. Let r : { u a e 1) be a family of superharmonic functions on an


open subset of 08N, locally uniformly bounded below, and define the lower
envelope (infimum) u of r by u 5 u,

+
lim nfu(q) (3.2)
11-4

and

(a) u is superharmonic.
(b) u = u on each open set on which u is superharmonic.
(c) = u IN abnost everywhere.
(d) There is a countable subset of f whose lower envelope has the same
lower semiconlinuous smoothing u.

Observation (1). The hypotheses of the theorem are not effectively


weakened by the added assumption that u. is directed downward because
if all finite minima uQ A A u,k are adjoined to the family, the enlarged
family of superharmonic functions is directed downward and has the same
lower envelope u, and (d) is true for the original family if true for the enlarged
family. The added assumption justifies the nomenclature "convergence
theorem."
Observation (2). According to Choquet's topological lemma (Appendix
VIII.3) Theorem 3(d) is true, and we can even assume that r is a decreasing
sequence. Furthermore, since the theorem is local, it can be assumed that
u is bounded below on its domain, say u z c, and replacing ua by ua - c,
it can be assumed that the members of f are positive.
In accordance with these observations the following proof assumes that
u, is a decreasing sequence of positive superharmonic functions. Apply
11(6.1) to find

u lim inf u (q) Z lim inf u(q),


R-C 7"C
38 1.111. Infima of Families of Superharmonic Functions

so that is at least equal to the last term on the right, and therefore
(3.1) implies (3.2). Since u is superharmonic,

A(u,,, , S) 5 u.(4), (3.3)

so that

A(u, ,, b) 5 A(u, , S) 5 (3.4)

Moreover continuity of A(u, , S) (Section 1.2) implies that

A (u, , S) 5 A(u, , b) 5

u is the largest lower semicontinuous function majorized by u.


Inequality (3.5) implies that u is superharmonic. Finally u = u IN almost
everywhere on D because when S- 0 inequality (3.5) becomes u 5 u 5
u 5 u IN almost everywhere on D, by a standard derivation theorem.

Application to GMDf'

Let r : { u,, or e 1) be a family of superharmonic functions on an open subset


D of R' and suppose that r has a subharmonic minorant. Then GMDr
exists; so r is locally lower bounded. Define u = infu3. According to
Theorem 3, the function u is superharmonic, and we now show that GMDI- _
GMDU. It is trivial that GMDr z GMDu and in the other direction GMDr
is a continuous minorant of u and is therefore a minorant of u by definition
of u; so GMDrS GMDu.

4. The Reduction Operation


Let D be an open subset of RN coupled with a boundary OD provided by a
metric compactification. Let A be a subset of D v 8D, let v be a positive
superharmonic function on D, and let r be the class of positive superhar-
monic functions u on D majorizing v on A n D and near A n.3D in the
sense that to each function u corresponds a neighborhood Ao of A n 8D
in the compact space D u 8D such that {u z v) (A u A0) n D. If u, and
u2 are in r, their minimum u, A u2 is also in r; so r is directed downward
in the order of pointwise inequality. The reduction RA of v on A is defined
as the infimum, that is, the limit, of the directed set r. Since v c- r , this
reduction is also the infimum and limit of the class of functions v A u with
u in r, that is, of the class of positive superharmonic functions on D equal
to v on A n D and near A n W. Thus R" 5 v, with equality on A n D. The
4. The Reduction Operation 39

set D is part of the context but is omitted from the notation. In most ap-
plications A e D. According to the Fundamental Convergence Theorem
in the preceding section, the lower semicontinuous smoothed reduction
R" is majorized by RA, is superharmonic, and coincides IN almost everywhere
on D with R. (According to the more refined version of this theorem in
Section VI.1, the set {R"v > RA
+V
} is not only IN null but polar, a more re-
strictive characterization to be defined in Section V.1, and it will be proved
in Section VIA that R" = RA on D - A.) The notation 8 VOA will sometimes
be used instead of R".
Obviously R" and R increase when A or v increases. Moreover the
reduction operation is subadditive in the sense that

B5 RB + RA + RB, (4.1)

and the corresponding inequality (4.Ism) for smoothed reductions is also


true. To prove (4.1) we observe that if u" [v"] majorizes u [v] on A n D and
near A n aD and if u8 [v8] majorizes u [v] on B n D and near B n OD, then
u" + u8 + v" + vs majorizes u + v on (A u B) n D and near (A u B) n aD.
To prove (4.Ism), observe that the two sides are superharmonic functions
and that the inequality is true 1N almost everywhere on D and therefore
everywhere on D.
If D - A is not empty and if B is a ball with closure in D - A, the function
r8u is in F whenever u is in r and rsu < u; so R" on B is the limit of the
downward-directed family of harmonic functions. Hence
(by the Harnack convergence theorem) R" is harmonic on B and so is
harmonic on D - A and equal to R" there. In particular, when A c OD,
the reduction RA is harmonic on D and R = R
Since v on A n D, it follows that R4 = RA = V on the interior
of A n D. Hence, if A is open in D u aD, the smoothed reduction RA is a
positive superharmonic function majorizing v on A n D and near A n OD
and majorized by RA ; so R" = RA on D. More generally we shall show in
Section 5(d) that R" = RA whenever A n D is open.

Ex iipu (a). Let D be a ball and let A be a relatively compact subset of D.


Then RA (and therefore also RA)
V
+V
has limit 0 at every ball boundary point.
It is instructive to prove this result directly here, although the result also
follows naturally at a later stage from the remark that when A is relatively
compact in D, the function R" is (Section IV.8) the potential G0µ of a
measure with compact support in the ball and as such (Section IV. I) has
limit 0 at every ball boundary point. To prove the stated result, let be
the center of D, let A 1 be an open neighborhood of A and be relatively
compact in D, and let B be an open neighborhood of A 1 and be relatively
compact in D. Then R"' = v on A; so R. 5 R. and R.", is harmonic on a
40 1.111. Infima of Families of Superharmonic Functions

neighborhood of aB. Choose a constant c so large that on 8B


and therefore on a neighborhood of OB. The function u equal to R. A- on B
and equal to RA' A on D - B is superharmonic on D, majorizes
R", and is a minorant of on D - B; so R' 5 u 5 near
8D, and therefore RA has limit 0 at every boundary point of D. A similar
argument shows that when N > 2, D = R', and A is a relatively compact
subset of D, then RA has limit 0 at the point oo.
EXAMPLE (b). Let D be a ball, let A = {} c D be a singleton, and let v =- 1.
Then RA(s) = 1, but since majorizes v on A whenever e > 0,
it follows that R' = 0 on D - {} and the smoothed reduction RA vanishes
identically.
EXAMPLE (c). Let D be a ball with center the origin, and let v be a positive
harmonic function on D with Riesz-Herglotz representation (Section 1I.14)

v=J K(,7, (4.2)


2D

We shall now prove

RA = R = fA K(q, )Mv(dq) (4.3)

for every Borel boundary subset A. The first equality in (4.3) is trivial
because RA is harmonic. Denote the integral on the right by vA. If u is a
positive superharmonic function on D that majorizes v near A and if A0
is a closed subset of A, the difference u - vAo is superharmonic on D and
positive near Ao. Moreover the fact that lime{ 4) = 0 uniformly on
Ao when C E 8D - Ao implies that vAa has boundary limit 0 on 8D - Ao.
Hence u - vAo has a positive inferior limit at every boundary point of D,
and the superharmonic function minimum theorem implies that u - vA0 Z 0;
so RA z vAo. If A, is an open boundary superset of A, then limt_{ 0
uniformly on DD - A, when C E A ; so v - vA has boundary limit 0 at every
point of A. Hence, if e > 0 the function VA, + e majorizes v near A, and
it follows that vA, 2 RA. Thus vAo 5 RA 5 vA,, and this inequality implies
the desired equality (4.3).
EXAMPLE (d). Let D be a Greenian subset of R', let B be an open relatively
compact subset of D, and let v be a positive superharmonic function on D,
finite and continuous at each point of 8B. Then R°-B = GMBv on B. In fact,
on the one hand, R" is harmonic on B and is majorized by v on B; so
5 GMBv on B. On the other hand, if u is a positive superharmonic
RD-B

function on D, with u >- v on D - B, and if vo is a harmonic function on


B and is majorized by v on B, then u - vo is superharmonic on B and has
a positive inferior limit at every point of 8B and so is positive on B, and it
follows that R" 2 GMBv on B.
5. Reduction Properties 41

According to Section VIII.18, the hypothesis in Example (d) that v be


finite valued and continuous at each point of BB is stronger than necessary.
It is sufficient, for example, if v is finite valued on B. More specifically,
according to Theorem VIII.18(c), in which V1 B) can be identified with
R°_B
R°-B, the evaluation = GMBv on B is true if and only if, in the ter-
minology to be developed below, the set of irregular boundary points of
B is a null set for the Riesz measure associated with v.

5. Reduction Properties

The following list of reduction properties will be useful, and expanded,


in later chapters. The reductions are relative to a Greenian subset D of OB",
and no restrictions not specifically stated are imposed on the positive super-
harmonic functions on D to be reduced or on the subsets of D u OD on which
these functions are reduced.
(a) R" in terms of reductions on subsets of D.

inf { R": A n 8D c B, B open in D u 8D}


= inf{R,"-"": A ndD c B, B open in DuaD}.
These evaluations follow from the fact that if u is a superharmonic function
on D majorizing v on A n D and near A n 3D, then there is a neighborhood
B of A n aD so small that u also majorizes v near B n aD. The smoothed
version of (5.1) is

RA =inf{Rv"B:An ODcB,Bopen inDu0D}


(5. l sm)
= inf{Rv"' °: Ana D c B, B open in DuaD}.

To prove (5.Ism), observe that in view of (5.1) and the Fundamental Con-
vergence Theorem the right sides are superharmonic functions majorized
by R" and therefore by Ro , and the reverse inequalities are trivial.
If A n OD is compact, the sets B in (5.1) and (5.1 sm) need only run through
a decreasing sequence of relatively compact open neighborhoods of A n OD
with that set as intersection.
(b) If p is a positive superharmonic function on D with GMDp = 0, if
A c OD, and if v is a positive superharmonic function on D, then R," P = R.
In particular, RD = 0. This property follows from the inequality

R"SRS P:5. RA + RA

because R", a positive harmonic function majorized by p, must vanish


identically.
42 1.111. Infima of Families of Superharmonic Functions

Observation. In Section IV.8 the positive superharmonic functions p


with GMDp = 0 will be identified with the superharmonic potentials GDµ.
(c) R" in terms of separate reductions on subsets of D and of 8D. Recall
(Section 4) that R"naD is harmonic. We now prove that if v' = v - R"11
then

RA
RA,
= RAIMD +Rd AnD (5.2)
V

and also prove the corresponding inequality (5.2sm) for smoothed reduc-
tions. In particular, if GMDV = 0 then RV = RAID and R " = RAnD. In fact,
if vt is a positive superharmonic function on D, majorizing v near A r) 6D
and if v2 ij a positive superharmonic function on D, majorizing v' on A n D,
then vt + v2 >_ R", and therefore
RV naD + RV n°
RA,. (5.3)

Conversely, if v3 is a positive superharmonic function on D, majorizing


v on A n D and near A n 8D, the positive superharmonic function v3 - RAnaD
majorizes v' on A n D; so
RVnD s RV
- R"naD
(5.4)

and this inequality combines with (5.3) to yield (5.2). Equality (5.2) implies
that the corresponding equality for smoothed reductions, an equality be-
tween two superharmonic functions, is true IN almost everywhere on D
and therefore is true everywhere on D.
(d) RV = R" whenever A n D is open. If A is open, this property was
pointed out in gection 4, and the more general property follows from (5.2)
and (5.2sm).
(e) If v is finite valued and continuous at each point of An D, then

RV = inf {Rv : A c B, B open in D u 8D}. (5.5)

In fact, if u is a positive superharmonic function on D which majorizes


v on A n D and near A n 8D, then for n z I the function (1 + 1 /n)u majorizes
v on a neighborhood of A n D and on the trace on D of a neighborhood of
A n 8D and therefore majorizes RB for some open superset B of A.

6. A Smallness Property of Reductions on Compact Sets


Theorem. If D is a Greenian subset of GIN, if A is a relatively compact subset
of D, and if v is a positive superharmonic function on D, then GMDRA = 0.

By hypothesis there is a nonconstant positive superharmonic function


u on D. It can be supposed that D is connected and, replacing u by utn if
7. The Natural Order Decomposition for Positive Superharmonic Functions 43

necessary, that u is not harmonic. Replacing u by u - GMDU if necessary,


it can also be supposed that GMDU = 0. If A, is an open relatively compact
subset of D containing A, then v, = RA, = R"1 is a positive superharmonic
function on D, harmonic on D - A, , and is equal to v on A ; so v, and v
have the same reduction on A. Replacing v by v, if necessary, it can be
supposed that v is harmonic on D - A,. Thus, if B is an open relatively
compact subset of D containing A,, the function v is harmonic on a neigh-
borhood of OB. Choose c so large that cu > v on B. This inequality will
then hold on a neighborhood of aB, so that (cu) A V = v on this neigh-
borhood. Define u, = v on B and u, = (cu) A v on D - B to get a positive
superharmonic function on D majorizing v on A and therefore majorizing
R" on D. Hence

GMDRA 5 R" = R' 5 cu

on D - B, and (by the subharmonic function maximum theorem) therefore


GMDRA 5 cu on D. Since GMDU = 0, it follows that GMDRn = 0, as was
to be proved.

7. The Natural (Pointwise) Order Decomposition for Positive


Superharmonic Functions

Theorem. If u, u, , and u2 are positive superharmonic functions on an open


subset D of RN, with u 5 u, + u2, then there are positive superharmonic
functions u1, u2 on D for which

U u2 5 u2, u = ui + u2. (7.1)

Let u, be the smoothed infimum (given by the Fundamental Convergence


Theorem) of the class of functions v, positive and superharmonic on D and
satisfying the inequality u 5 v + u2. Then u 5 u, + u2 IN almost everywhere
on D and therefore everywhere on D. Let u2 be the smoothed infimum of the
class of functions v, positive and superharmonic on D and satisfying the
inequality u 5 u, + v. Then u:5 ui + u2 on D, and if v is a positive super-
harmonic function on D for which u:!5; ui + v [u 5 v + u'2], it follows that
v z u2 [v ui]. In particular (choose v = u), it follows that u z u, and
u ,-a u2. To prove the theorem, it will be shown that u = u, + u2. Define
(u - u2) as the indicated difference whenever +oo and define

(u-u'2)()= liminf [u(q) - u2(q)]

whenever u - u2 is thereby lower semicontinuous


at each infinity of u2. and we shall now prove that u - u2 is also lower
44 I.M. Infima of Families of Superharmonic Functions

semicontinuous at the other points of D. Let B be a ball with closure in D.


Then

U=tgu+U - r9U<u'1+[r5u2+U-zBu] (7.2)

on B, and the bracketed function v is superharmonic on B and majorizes


teu2 there. Hence [Section II.6(e)] the function (u2 A v),B extended to D
by u2 is a positive superharmonic function v' on D, and u S u, + v'. It
follows that v' z u2 ; that is, rBU2 + u - rBu Z u2 on B. Thus

tB(u - U2) 5 (u - U2) (7.3)

on B. If B has center , if oo, and if B shrinks to , the value


rB(u - teu'2(l) tends to (u - and since the left
side of (7.3) is continuous (harmonic) on B we conclude that u - u2 is lower
semicontinuous at , as desired, and therefore is lower semicontinuous at
every point of D. Furthermore (7.3) implies that the function u - u2 satisfies
the superharmonic function average inequality; so this function is super-
harmonic. The relations (u - u2):5 u, and u = (u - u2) + u2 are valid
except possibly at the !,, null set of infinities of u2 and are therefore valid
everywhere on D. It follows that u - u2 = ui and u = u, + u2, as was to be
proved.
Chapter IV

Potentials on Special Open Sets

1. Special Open Sets, and Potentials on Them


Throughout this book a special open subset of R" is either a ball in R" or
R" itself, but the latter only when N > 2. The Green function GD for D a
ball was defined in Section 11. 1. The Green function GD for D = R" with
N > 2 is defined as G. If p is a measure on a special open set D, define the
function GDµ on D by

GDµ()= J f GD(c,q)p(dq). (1.1)


D

An application of Fatou's lemma shows that GDµ is lower semicontinuous,


and GDP satisfies the superharmonic function inequality because q) does
for each ry. Hence GDµ is either superharmonic or identically + 00 on D.
The function GDµ is called the potential or sometimes the Green potential
of µ. In Chapter VII a Green function GD will be defined for every Greenian
open set and will be shown to enjoy many of the properties of the Green
function of a special open set. For example, if p is a measure on the Greenian
set D, the function GDµ will be shown to be either superharmonic or iden-
tically + o0 on each open connected component of D.
In Chapters IV-VI many theorems will be stated for potentials on
Greenian sets but proved only for potentials on special sets. This is done
because the results for these special sets will be used to develop many of the
general results of classical potential theory that are needed before a general
Green function can be comfortably defined and its properties studied. It
will be obvious that circular reasoning will not be invoked, and in Sections
VII.5 and VII.8 it will be pointed out that the proofs in the preceding chapters
for special sets are also applicable when the sets are Greenian.
In this book potential in the present classical context always means a
function GDµ with D Greenian or (rarely and then always pointed out
explicitly) Gµ when N = 2 and the open set involved is the plane. The
existence of the latter potential is discussed later in this section.
Suppose again that D is special open and consider a potential GDµ. It is
46 I.IV. Potentials on Special Open Sets

clear that GDp is superharmonic if y has compact support in D because this


potential is finite off that support. More generally GDp is superharmonic if
p(D) < +oo because if pt is the projection of p on a ball B with closure in
D and if 142 is the projection of p on D - B, then GDP = GDpt + GDµ2,
GDPI is superharmonic because pt has compact support in D, and GDP2 is
superharmonic because this potential is finite on B. Moreover, if GDP is
superharmonic and if Do is an open p null subset of D, the potential GDp
is harmonic on Do because this potential is Borel measurable and satisfies
the harmonic function average property on D0.
If p has compact support in the special open set D, the Lebesgue dominated
convergence theorem when applied in (1.1) yields the fact that the potential
GDP has limit 0 at every boundary point of D, because the function q)
has this property and is uniformly bounded in a neighborhood of 8D when
q is restricted to a compact subset of D. More generally, if D is a Greenian
subset of R' and if p is a measure on D with GDP superharmonic, it will be
shown that in various senses the potential GDp, in particular G,,(-, 17) for
fixed q in D, has boundary limit function 0. See, for example, Section VII.4,
VIII. 11, XII.19, XII.23, and 2.X.8.

Potentials Gp When N = 2

The study of potentials Gp when N = 2 is complicated by the fact that G is


then not a positive function. Observe first that for an arbitrary measure p
on R2 and m > 0 the integrand in

um(b) = J logI - h1 -tp(d1), (1.2)


o.m) L0,M)

is lower bounded for each ; so uum(t;) is uniquely defined (< + oo). If we


write in the form

um() = log(nl - qJ p(B(O,m)) loge, n > m > 0,


J 8(o.m1
(1.3)

the integrand is positive for in B(0, n - m). The reasoning used above in
discussing potentials of measures on balls and on 18" when N > 2 when
applied to the function defined by the integral in (1.3) shows that um is a
superharmonic function on B(0, n - m) and is harmonic on each p null open
subset of B(0, n - m). Since n can be chosen arbitrarily large, the function
um is a superharmonic function on 082 and is harmonic on 082 - B(0, m)
and on each open p null subset of B(0, m). In particular, if p has compact
support and if m is so large that B(0, m) contains this support, it follows
that Gp is superharmonic and is harmonic on the complement of the
2. Examples 47

closed support of u. Going back to an arbitrary choice of µ, define


A = Um=, {un = +oo). Then A is IN null since u,,, is superharmonic, and
in fact A is what will be described in Section V.1 as a polar set. Observe that
for k > 0 and m > k + l the function un is harmonic and negative
on B(0, k). In view of the Harnack Convergence Theorem (Section 1I.3) we
conclude that either (a) Gp = - oo on B(0, k) - A and Gp is not uniquely
defined on A or (b) Gu is uniquely defined and superharmonic on B(0, k).
Since k can be chosen arbitrarily large, either (a) Gp = - oo on 08Z - A
and Gp is not uniquely defined on A or (b) Gp is uniquely defined and super-
harmonic on 982. Thus Gµ is superharmonic if and only if

log 17Iµ(th1) < +oo; (1.4)

the condition µ(982) < + co is necessary but not sufficient for (1.4). See
Section 9 for a further analysis of superharmonic functions on 982.

Dependence of Go and GDp on D

If D, and D2 are special open subsets of 98" with D, c D2, it follows from
the maximum-minimum theorem for harmonic functions that GD, S GD2 on
D, x D, because if cED,, the function if properly
defined at is harmonic on D, with positive limit at every boundary point.
In this sense the function Dr-'G, is monotone increasing. In particular,
GB(0,,) increases as r increases, and (direct calculation) hm, GB(o,,, is either
identically + co or G on 98N x98" according as N = 2 or N > 2. If y is a not
identically vanishing measure on 98" and if p, is the projection of U on B(0, r),
then lim,_m GB,o,,)p, is identically + co if N = 2 and is Gu if N > 2. What-
ever the dimensionality, if u is a finite measure on the ball B, the potentials
GBp and Gu are superharmonic and differ on B by a harmonic function.

2. Examples
Let p be a unit mass distributed uniformly on OB(0, S). Then

-N+ 2 on B(O,6),

t1l,
-N+2 2
on B(0, S)RN - (N > 2), (2.1)
GB(o.a,µ=Gp-a-N+2 onB(0,a)ifa>6.

In fact the potential Gp must be a function of the distance I q I from the origin,
must be superharmonic, and must be harmonic on RN - 5B(0, 6), and there-
fore (according to Section 11.10) must be a concave function of lq1-N+2, and
48 1.1 V. Potentials on Special Open Sets

must be a linear function of IqI-N+2 on each open component of R' -


aB(O, S). Finally, this potential is 6-1+2 at the origin and has limit 0 at oo.
Hence the first two lines of (2.1) are correct. If a > S, the difference Gµ -
Ge(o.a)F1 is a function of IriI, defined and harmonic on B(0, a), with limit
Gp = a-N+2 at the boundary. Hence (maximum-minimum theorem for har-
monic functions) this difference is identically a-N+2 ; so the third line of (2. 1)
is correct.
Similarly

log b on B(0, S),


Gµ(q) _
- log I q on R2 - B(0, S) (N = 2), (2.2)
G8j0,a,p = Gµ + loga on B(0,a) if a > S.

If µ,a is a unit mass distributed uniformly on S), then

L(G(ri, -), , S) = Gp,a(q), (2.3)

and it follows from (2.1) and (2.2) that for N z 2 the function of q)
defined by (2.3) is continuous on fl8N x R'. If a > S and if e B(0, a - S), the
difference G4u - Ge<0,6tµfa = h is harmonic on B(O, a); so

L(G8(0 )(q, ), , S) = Gp,e(q) - (2.4)

It follows that the function of q) defined by (2.4) is continuous on


B(0, a - S) x B(0, a).

3. A Fundamental Smallness Property of Potentials

Theorem. If D is a Greenian subset of R' and if u = GDp is a superharmonic


potential on D, then GMDu = 0.

If D is special, choose a sequence B. of balls with closures in D and with


the property that if is in D, some neighborhood of is contained in infinitely
many of the balls. Then (Section 111. 1) T8n - T8 p j GMDU. Moreover

TB....
ID

and the integrand decreases to 0 when n- oo because (Section Ill.1)


q) = 0. Hence GMDU = 0, as was to be proved. Observe that
when D is special, the simplest choice of B is an increasing sequence of
concentric balls with union D. The proof as stated will be valid for arbitrary
5. Smoothing of a Potential 49

Greenian D once GD has been defined in Section Vll.l. The converse of


Theorem 3 is also true: it will be seen in Section 8 that for u positive and
superharmonic on a Greenian set D the condition GMDU = 0 is necessary
and sufficient for u to be a potential on D.

4. Increasing Sequences of Potentials


Theorem. If D is a Greenian subset of R', if µ, is a sequence of measures sup-
ported by a compact subset A of D, if vague and if GDµt S
GDµ2S ,then
We defer the proof for Greenian sets and suppose that D is special. Since
sup.µa(A) < +oo, the function u = lim,,.,m GDµ is finite on D - A and is
therefore superharmonic on D. Apply Fubini's theorem and the fact that
q) i--. L(GD(q, ), , r) is continuous (Section 2) to deduce

L(u, , r) = lim
1.0L(GDPM, , r) = L(GD(n, ), i, r)µ(dn) = L(GDP, , r). (4.1)
J

When r -. 0, we find that u = GDµ, as was to be proved.

5. Smoothing of a Potential
Let µ be a measure on OWN with compact support A. An elementary calcula-
tion shows that A1(Gµ) = Gµ1, where pQ is absolutely continuous relative to
iN with infinitely differentiable density

(S) = a-N
J pN
}'N(Ib

a
I /lµl) .1)
(5.1)

Moreover µ1(RN) = µ(A), the measure µa is supported by Al < a),


and lima., µa = µ (vague convergence). Finally (from Section 11.6) Gµ1
increases when a decreases and lim/..o Gµ1= Gµ. According to Theorem 1.7,
the density (5.1) is equal to -A(Gµa)lr4N.
If B is a ball containing A and if a < IA - OBI, the function A1(GBµ) is
defined and superharmonic on I - OBI > a) and is harmonic and
I

equal to GBµ on I i - AI > a, Ia; - OBI > a). Define Aa(GBM) on


( a B: I - OBI S a) as GBµ to obtain an infinitely differentiable super-
harmonic function on B. Since GBµ differs from Gp on B by a harmonic
function the results in the preceding paragraph yield the fact that A1(GBµ) =
GBµ1, where µ1, the same measure as above, has density -A(GBµa)/a4 =
- A(Gµ1)/rrN relative to IN, and GBµa increases to the limit GBµ when a 10.
50 LIV Potentials on Special Open Sets

6. Uniqueness of the Measure Determining a Potential


Theorem. Let D be a Greenian subset of 68"(N > 2) or D = 682, and let p and
v be measures on D for which GDp and GDv are superharmonic. Suppose that
A is an open subset of D and that there is a function h defined and harmonic on
A for which GDp = GDv + h on A. Then the projections of p and v on A are
identical.

The proof will be given for D = 68" and requires only trivial changes for
D a ball but the remarks validating the proof for a general Greenian set are
deferred to Section VII.8. It is sufficient to prove the result for A relatively
compact in 68", and therefore it can be supposed, replacing u and v by their
projections on a ball containing A if necessary, that p and v have compact
supports. According to Section 5, the functions A,(Gp), A,(Gv) are infi-
nitely differentiable potentials of measures p v, with respective densities
-AA,(Gp)/n;v, -L A,(Gv)/ah relative to I Since these densities are equal
on the set ; e A : - aAI > a), the projections of p, and v, on this set are
I

identical, and since p, and v, have vague limits p and v when a -' 0, the
projections of p and v on A are identical, as was to be proved.

Extension of Theorem 6 to Charges

In the following extension of Theorem 6 we consider charges p on D (see


Appendix IV.7) with the following property: p has positive and negative
variations p' and -p-, respectively, with the property that GDp* and GDp-
are superharmonic. As usual p* and p- are finite on compact subsets of D,
but if D * 682 both p* (D) and p- (D) may be + ac, sop need not be a signed
measure. The potential GDp is defined as GDp* - GDp- on the subset of D
on which at least one of the potentials GDp`, GDp- is finite. If v is a second
such charge on D, if A is an open subset of D and if h is a function defined
and harmonic on A, suppose that GDp = GDv + h on A in the sense that
GD(p* + v) = GD(p- + v*) + h on A. According to Theorem 6, if this
condition is satisfied, it follows that p* + v =p + v* on subsets of A.
Since the four measures involved here are finite valued on compact subsets
of D, we conclude that p = v on the compact subsets of A, so these measures
have identical projections on A. Thus we have extended Theorem 6 to cover
potentials of charges as restricted above.
In particular, if GDp is harmonic on the open subset A of D, so that we can
take v as the zero charge, we conclude that A is p null. As a second particular
case suppose that GDp* and GDp- are finite valued on an open subset A of
D and that (GDp)IA is in the class C11(A). Let v be the charge on A yvith
density - (LtGDp)/nN relative to /,. If AO is an open relatively compact subset
of A and if vo is the projection of v on Ao then (Section 1.7) (GDVO)IA is in the
class C'21(A), and A(GDp - GDvO) = 0 on AO; so GDp differs from GDvo on
A0 by a harmonic function and it follows that p = vO on subsets of A0 and
7. Riesz Measure Associated with a Supcrharmonic Function 51

so on subsets of A; that is, dp = -A(GDp)/7[N d1N on A. The hypothesis that


(GDµ)J4 E C" )(A) can be weakened, for example, to the condition GDp
c C'2)(A) by applying the result just obtained to a smoothing A,(GDIz),
but we shall not need this refinement.

7. Riesz Measure Associated with a Superharmonic Function


Theorem. If u is a superharmonic function on an open subset D of RN(N z 2),
there is a unique associated measure p on D with the property that if A is an
open relatively compact subset of D, and if pe is the projection of u on A, there
is a superharmonic function hA on D, harmonic on A, with

u=Gµ,,+hA (7.1)

on D.

The uniqueness of µA is obvious from Theorem 6, and incidentally this


uniqueness implies that the measure determining a potential is the measure
associated with the potential. If the theorem is true with h,, supposed only
defined and harmonic on A and satisfying (7.1) on A, this function can be
extended to a superharmonic function on D which satisfies (7.1) on D. In
fact, if B is an open superset of A, relatively compact in D, and if v is the
projection of Pa on B - A, then h4 = Gv + hB on A. Thus h,, has a super-
harmonic extension Gv + hB to B, and when h,, is so extended, u = Gµ,, + h4
on B. In this way hA can be extended in step* to D. Finally, it is sufficient to
prove that if A is a ball with closure in D, there is a measure µA defined on A
and a harmonic function h,, defined on A for which u = Gµ,, + h4 on A. In fact,
if two such balls overlap the parts of the two associated measures on the
intersection of the balls must be identical by Theorem 6, and the desired
measure on D is obtained by piecing together the measures associated with
the balls.
To prove the italicized statement we first replace D by a subset if necessary
to make u bounded from below on D, and adding a constant if necessary, we
can even suppose that u is positive on D. Since only the restriction of u to A
is relevant to the desired result, we can replace u on D by R1, a positive super-
harmonic function on D, equal to u on A, harmonic on D - A. The proof
now proceeds as follows. Let B be a ball concentric with and larger than A,
with closure in D. Choose a so small that a is strictly exceeded by the dif-
ference between the radii of A and B and by IB - aDI. The function A,u is
superharmonic and infinitely differentiable on a neighborhood of B, equal
to u, and harmonic on a neighborhood of 8B, and 6ma.,o AQu = u on D.
According to 1(8.6),

A,u = GBµa + PI(B, u) (7.2)


52 I.IV. Potentials on Special Open Sets

on B, where µ, is the measure on B with density - AA,u/tr, relative to 1,'. The


value of µ,(B) can be found by applying the Gauss integral theorem (inte-
grate over 8B), and since A,u = u on B, µ,(B) does not depend on a. Thus as
a 0 along a suitable sequence p,, has a vague limit measure µo supported by
4. Applying Theorem 4, we find that if PA is the projection of µo on A, then

(G8µ0 + PI(B,u) on B,
u=
Gµ,t + h,, on A,

where h4 is harmonic on A, and we have used the fact that GBµA and Gµ4
differ on A by a harmonic function.
Observation. It follows from Theorem 6 that the map u F-' p from super-
harmonic functions into their associated measures is additive (u1 Fpi and
u2 r+p2 imply that u1 + u2 µ2) and positive homogeneous
implies that cu E-- cp for c a positive constant).

8. Riesz Decomposition Theorem


Theorem. If D is a Greenian subset of RR'(N 2), if u is a superharmonie func-
tion on D with associated measure p, and if u has a subharmonic minorant, then

u=G°µ+GM°u. (8.1)

In the most common application u is supposed positive, in which case the


harmonic function GM°u is also positive. We defer the general Greenian
case to Section VII.8 and treat here only special sets D. If D is a ball, let B be
a smaller concentric ball. In the notation of Theorem 7, u = Gµ5 + hi, =
G°µB + h8, where h'S is superharmonic on D and harmonic on B. Now
according to Section 111.1, the minorant GM°v depends additively on the
superharmonic function v, and according to Section 3, this minorant vanishes
if v is a potential. Hence GMnu = GM°h'B 5 h8. When B increases to D, we
conclude that G°µ is superharmonic, that h8 decreases to a harmonic func-
tion h on D, and that u = G°µ + h. Then h = GM°u, and the proof is com-
plete. The proof for R" when N > 2 is essentially the same.
The following assertions about a positive superharmonic function u on D
are easy consequences of the Riesz theorem.
(a) The function u is a potential GDP if and only if GM°u = 0, equiva-
lently, when D = b), if and only if lim,.a L(u, , r) = lim,_a
0 (or when D = RN with N > 2, if and only if this limit
is 0 when r -, + oo).
(b) If u is majorized by a superharmonic potential u is itself a potential.
(c) Special case of (b). If u is a superharmonic potential, then R' is also,
for every choice of A; in particular, 0 for every choice of 8D.
9. Counterpart for Superharmonic Functions on R2 of the Riesz Decomposition 53

(d) The smoothed reduction Ru" is a potential if A is a relatively compact


subset of D (because GMDRA = 0 according to Theorem 111.6).

Observe that a superharmonic function u on D is the sum of a potential


and a harmonic function if and only if u has a subharmonic minorant.

9. Counterpart for Superharmonic Functions on R2 of the


Riesz Decomposition
Let u be a superharmonic function on R' and let b be a strictly positive
number. Then (Section II.4) the function L(u, , b) is superharmonic on R2.
For fixed in R2 the function

bH L(u, , b) - L(u, , 1) = m(u,


, b)
- log b

is a positive increasing function on ] 1, +oo[ in view of the superharmonic


function average inequality and the fact (Section 11.10) that the function
L(u, l;, ) is a concave function of log b. Hence (Section 11.4) the function
m(u) = lim&-m m(u, , b) is either identically + oo or a positive superharmonic
function, and in the latter case it is identically constant because (Section IL 13)
the plane is not Greenian.

Theorem. (a) I fu is a superharmonic function on 682, then u is harmonic if and


only if m(u) = 0. _
(b) If µ is a measure on R2 and if g(b) = µ(B(0. b)), then Gp is super-
harmonic on R2 if and only if

log6dg(b) < +oo. (9.1)

(c) If µ is a measure on 682 satisfying (9.1) and if u = Gp, then

µ(68Z) = m(u) < + oo, li m [L(u, 0, b) + g(b) log b] = 0,


b (9.2)

and (9.2) is also true iJ'g(b) is replaced by p(68').


(d) Let u be a superharmonic Junction on 682 with associated Riesz mea-
sure p and define gas in (b). If

lim inf [L(u, 0, b) + g(b) log b] < + oo, (9.3)


6-OD

then Gµ is superharmonic and u = Gp + h, with h a harmonic function on 682


given by
54 l IV. Potentials on Special Open Sets

lim [L(u, , b) + µ(I82) log b]. (9.4)


a-m

Proof. (a) If u is harmonic on 682 the function m(u, , b) vanishes identically;


so m(u) = 0. Conversely, if u is superharmonic on 12 and m(u) = 0, then
since m(u, c, ) is a positive increasing function on the interval ] I, + 00 [, it
follows that this function vanishes on the interval ; that is, L(u, t, ) is a
constant function there. Since L(u, , ) is a decreasing concave function of
log b for b > 0, it follows that L(u, , ) is identically constant; that is, u
satisfies the harmonic function average equality and so is harmonic.
(b) and (c) The condition (9.1) that Gµ be superharmonic was derived
in Section 1 [see (1.4)] in a trivially different form. If (9.1) is satisfied, the
evaluation of in Section 2 or Section 1.5 yields

L(G/, 0, b) = -µ(B(0, b)) log b - log 1171 µ(d7)


RZ-B(O.b)
(9.5)

= -g(b)logb - f logsdg(s).
>a.+ml

The function g is positive, monotone increasing, and right continuous, with


limit µ(I82) at + oo. Thus (9.2) is true. Moreover (9.1) implies that

lim [µ(682) - g(b)] log b = 0; (9.6)


6-M

so g(b) can be replaced in (9.2) by µ(I82).


(d) Under (d) let µ, be the projection of µ on B(0,6). According to
Theorem 7, there is a superharmonic function h8 on 682, harmonic on B(0, b),
such that

u = Gµa + ha. (9.7)

Now (9.5) with µ replaced by µa yields the equality L(Gµ6, 0, b) =


- g(b -) log b. Furthermore the function r '-. L(ha, 0, r) is continuous and
equal to ha(0) for r < b; so L(h6, 0, b) = ha(0). Hence

L(u, 0, b) = -g(b-) log b + ha(0). (9.8)

For fixed q the function b i-. Gµa(q) is monotone decreasing on the interval
] 1 + f , 1, + co [, so the function b --. ha(p7) is increasing on this interval.
Hence (Harnack convergence theorem) either ha = + 00 on 68Z or
lima_. ha = h is a harmonic function on 682. In view of (9.8) the first case
is excluded by (9.3). Hence in the limit (9.7) yields the fact that Gµ is super-
harmonic and u = Gp + h. Moreover from (9.8)
10. An Approximation Theorem 55

h(O) = lim [L(u, 0, b) + g(b) log b],


6-OD

and as noted in the proof of (c), the value g(b) can be replaced by p(R')
in this limit relation. Thus (9.4) is true for 4 = 0 and therefore for all
since rather than the origin can be chosen as the reference point in this
discussion.

10. An Approximation Theorem

The following theorem is an example of the application of the reduction and


related operations.

Theorem. Every superharmonic function u on an open subset D of 68" is the


limit of an increasing sequence u, of superharmonic junctions with the following
properties :
(a) u is upper bounded.
(b) infpu, = infDu.
(c) u is infinitely differentiable.
(d) [At the possible sacrifice of (a)] if u is harmonic outside a compact
subset A of D, then u,, = u outside a compact neighborhood A. of A with
A A.
a A of D, we can choose u, to
satisfy (b)-(d) as follows. Choose a sequence a, in 68+ satisfying

J!A - BDI > ao>at> ,


R-M

and define u = A,Mu on the set B,, = I - AI < The function u is


superharmonic and is equal to u near BB,,. Define u = u on D - B. to obtain
a sequence u, satisfying (b)-(d), with A =
To find a sequence u, satisfying (a)-(c) for general superharmonic u
suppose first that infD u = 0 and let D. be an increasing sequence of open
relatively compact subsets of D with union D. Define v = R°na positive
superharmonic function on D, bounded by n, equal to u A n on D,,, and
harmonic on D - D,,. Now proceed as suggested by the method used in the
preceding paragraph. Define at, = ID - cD1/3, except that a = 1/n if
D = 68", and define u,, = Aa,,(v,. ) on the set B. = {g: 1 - D.1 < 2a.). The
function u is superharmonic, majorized by v,,, and equal to v,, near BB,,.
Define u = v on D - B. to obtain a sequence u, satisfying (a)-(c).
If infD u = ft > - oo, find a sequence u, satisfying (a)-(c) by adding f to
each member of a sequence satisfying these conditions for the function u - IJ.
56 1. IV. Potentials on Special Open Sets

If u is not lower bounded, modify the approximation procedure as follows.


Choose D. as above and also to satisfy D. c D.+1, and define v. by

v. = inf {v: v superharmonic on D, v u n n on Den v(U aD2j11 (10.1)


n JJ)

so that the lower semicontinuous smoothing + is majorized by u A n,


coincides with u A n on D2, and (Fundamental Convergence Theorem) is
superharmonic on D and harmonic on U. (D21 2 - D2). Define at. as a
decreasing sequence of strictly positive numbers satisfying

a. < IaD2n+2 - D2n+1I A IaD2n+1 - D2nl, lima.=0,


n-00

and define u. by

AQn+n
on Den+1,
_ (10.2)
A,k+n on Dzk+1 - D2k_1 for k > n.

The sequence u, has the desired properties (a)-(c).

Specialization to Positive Superharmonic Functions and Potentials

I f D in Theorem 10 is Greenian and if u is a positive superharmonic function


on D, each approximation u. can be chosen to satisfy (a)-(c) and also to be
the potential of a measure with compact support. If u is itself the potential
of a measure with compact support A, each approximation u. can be chosen
to satisfy (a)-(d) with un the potential of a measure supported by A. [notation
of (d)].
It is convenient to suppose first that u is the potential of a measure with
compact support A. Then u is harmonic on D - A. Furthermore u is bounded
outside each neighborhood of A because the function (, 17) is a
bounded function for q restricted to A and restricted to the complement
of a neighborhood of A. (This property of the Green function is trivial for
D special and will be proved in the general case.) The first paragraph of the
proof of Theorem 10 furnishes a sequence u. with the properties (a)-(d);
un is a potential because [Section 8(b)] u. is majorized by the potential u; the
Riesz measure associated with u,, is supported by A. because u. is harmonic
on D - A.. Thus the assertion for u when u is the potential of a measure
with compact support in D is true. For u superharmonic and positive the
proof of Theorem 10 for the case inf u = 0 provides a sequence u, with the
desired properties. In fact each function u. is a potential because by compact-
ness of D. the function v. = R° is a potential [Section 8(d)], and the
positive superharmonic function u. is a potential because it is majorized by
vn .
Chapter V

Polar Sets and Their Applications

1. Definition
A polar subset of RN is a set to each point of which corresponds an open
neighborhood of the point that carries a superharmonic function equal to
+ oo at each point of the set in the neighborhood. An inner polar set is a
set whose compact subsets are polar. It will be shown in Section VI.2 that
an analytic inner polar set is polar. If a set is (inner) polar its Kelvin trans-
forms are also.
In particular, the set of infinities of a superharmonic function is a polar
subset of its domain. Conversely, it will be shown (Theorem 2) that a polar
set is always a subset of the set of infinities of a single superharmonic function
defined on RN.
The polar sets are the negligible sets of classical potential theory. An
assumption about points of R' true except for the points of an [inner]
polar set is said to be true [inner] quasi everywhere. A subset of an [inner]
polar set is [inner] polar. A singleton is polar because is super-
harmonic on RN and equal to + oo at . Although the point ao is not in
RN, that point is considered a Euclidean boundary point of every unbounded
set. In a context allowing oo in the domain of harmonic and superharmonic
functions, this point is polar for N = 2 but not for N > 2.
Since a superharmonic function on an open subset of R' is IN integrable
on every closed ball in its domain, and since every polar set A can be covered
by a countable number of open sets, each carrying a positive superharmonic
function with value + cc on the part of A in its domain, a polar set has
IN measure 0. It follows that an IN measurable inner polar set also has IN
measure 0.
If u and v are superharmonic functions on an open subset of RN and
if u = v inner quasi everywhere, or if u >_ v inner quasi everywhere, then
the same relation holds IN almost everywhere and therefore [Section 11.6(f)]
everywhere.
58 I.V. Polar Sets and Their Applications

2. Superharmonic Functions Associated with a Polar Set


Theorem. If A is a polar subset of R', there is a function superharmonic on
and identically + oo on A. This function can be chosen to be the potential
Gp of a measure u with µ(I8") finite and to be finite at any preassigned point
ofpN-A.
To prove the theorem suppose that e I8" - A and apply the Lindelof
covering theorem to cover A by balls Bo, B,, ... so small that l: is not in
any ball closure and that to each ball Bk corresponds a function uk defined
and superharmonic on an open neighborhood of Bk and identically + oo
on B. n A. Let µk be the projection on Bk of the Riesz measure associated
with uk ; choose a strictly positive constant ck so small that ck uk(Bk) < 2-k,
that 2-k, and if N = 2, that ck J, log IhI 2-k. The super-
harmonic potential G Eo ckpk is + oo on A and finite at .
Observation (a). Since the set of infinities of a superharmonic function
v is the G6 set no {v > n}, every polar set is a subset of a G6 polar set.
Observation (b). Since a superharmonic function is 1,-, integrable on
every ball boundary in its domain, a polar set meets a ball boundary in
an 'N_, null set.
Observation (c). The complement of a closed polar subset A of RN is
connected. To see this, let B be an open connected component of R' - A,
let u be a superharmonic function on RN, identically + ao on A, and define
v = + oo on B and v = u on R' - B. The function v satisfies the conditions
for a function to be superharmonic on I8" except for the finiteness condition.
Hence v is either identically + o0 or superharmonic. Both alternatives are
impossible unless R' - A is connected.
Observation (d). If D is an open subset of R", the set I8N - D is polar
if and only if the finite part a°D = I8" n aD of the boundary is polar, and
then a°D = I8" - D. In fact, if a°D is polar, its complement is connected
and everywhere dense and so is equal to D, and therefore R" - D = a°D
is polar. Conversely, if R" - D is polar, the set D is everywhere dense; so
a°D = R" - D is polar.
Extension. If D is a Greenian subset of R' and if A is a polar subset of D,
then there is a positive function superharmonic on D and identically + oo on A.
This function can be chosen to be the potential GDµ of a measure p with u(D)
finite and to be f inite at any preassigned point of D - A.
If D is special, the proof of Theorem 2 for N > 2 with RN replaced by
D and G replaced by G. is valid in the present context. This same proof
will be valid in the case of general Greenian D once (Section VIII) GD
has been defined.
4. Properties of Polar Sets 59

3. Countable Unions of Polar Sets


Theorem. A countable union of polar sets is polar. In particular, an inner
polar F, set is polar.

In fact, if AO, A 1, ... are polar, if e IR" - Ua Ak, and if µk is a measure


on R' with Gµk = +oo on Ak and (if N = 2)
with f; log ryI µk(dry) < 2-k, then the superharmonic potential G(Eo µk) is
+ooonU0Ak.
Since singletons are polar, this theorem implies that countable sets are
polar. For example, suppose that A is a countable dense subset of P", and
let u be a superharmonic function on R", equal to + oo on A. The set of
infinities of u is a polar dense GG set and therefore is not countable.
It will be shown in Section VI.2 that an analytic inner polar set is polar.
According to Theorem 3, if u and v are superharmonic functions on an
open subset of R" and if v S u inner quasi everywhere on an F, set A, then
this inequality must hold quasi everywhere on A because the set

{v>u}nA=U{v>r}n{rZu}nA (r rational)

is an F. inner polar set and is therefore polar.

4. Properties of Polar Sets

Theorem. The following conditions on a subset A of a connected Greenian


set D are equivalent:
(a) A is polar.
(b) There is a superharmonic potential u = GDµ with µ(D) < + oo and
u=+ooonA.
(c) If u is superharmonic and strictly positive on D, R' =0.
(d) There is a strictly positive superharmonic function u on D for which
has a zero.

In the following argument D is special. We defer the general Greenian


case to Section VII.8.

Proof. (a) (b) Has already been proved for D = R". The proof for a
ball is similar.
(b) . (c) Let u be a positive superharmonic function on D, identically
+ oo on A. Then (for n z 1) u/n z u on A ; so 5 u/n, and therefore
0 at every point where u is finite. Hence the positive superharmonic
60 I. V. Polar Sets and Their Applications

function R" has a zero and accordingly must vanish identically (super-
harmonic function minimum theorem).
(c) . (d) The result follows because R' = IN almost everywhere.
(d) (a) The result follows because if 0, there is a function
v positive superharmonic on D, Z u on A, 5 2-" at , so that Eo v is
superharmonic on D, identically + oo on A.

5. Extension of a Superharmonic Function


Theorem. Let D be an open connected subset of RN, let A be a polar subset
of D, and let u be an extended real-valued function on D - A satisfying the
following conditions: - oo < u 5 + oo ; u ; + oo ; u is locally lower bounded;
u is lower semicontinuous; if e D - A, then u(l;) Z L(u, , 6) for sufficiently
small b, depending on . Then u has a unique superharmonic extension u' to
D, determined on A by
liminfu(p). (5.1)
n-4

In this theorem when 6) c D, the function u is defined IN_, almost


everywhere on 6), and u is IN_, measurable and lower bounded on
this set; so L(u, , 6) is well defined. In the simplest case A is relatively
closed in D, so that the hypotheses make u superharmonic on D - A, but
these hypotheses allow A to be dense in D. A superharmonic extension
u' of u satisfies (5.1) because of the strengthening of the lower semicontinuity
property of superharmonic functions in 11(6.1). If the theorem is true
locally, it is true as stated; so we can assume that D is a ball, that u is not
identically + oo, and (Section 4) that there is a positive superharmonic
function v on D, identically + oc on A but finite at some point of D - A
at which u is finite. When e > 0, the function u + ev, if defined as +oo
on A. is superharmonic on D, and according to the Fundamental Conver-
gence Theorem (Section 111.3), the function uo = lim,..o (u + ev) has lower
semicontinuous smoothing uo superharmonic on D, equal IN almost every-
where on D to uo. If e D - A, the given inequality z L(u, , 6) implies
[see 1(2.2)] that u(g) >- A(u, , 6) (for sufficiently small 6). Applying lower
semicontinuity of u, we find that lim infb_o A (u, , 6) >- Hence

lim A(u, , 6) ( e D - A). (5.2)


6-0

Moreover u = uo = uo 1N almost everywhere on D. Therefore

lim A(uo, , 6) = uo() c D - A); (5.3)


&-0

so uo is the desired superharmonic extension of u.


5. Extension of a Superharmonic Function 61

Special Case

If in the theorem u is bounded and if L(u, , b) for in D - A and


sufficiently small b depending on 4, then u has a unique harmonic extension
to D. In fact, under these stronger hypotheses if u' is the bounded super-
harmonic extension of u provided by Theorem 5 then when ED - A and
S is sufficiently small the function u' - is positive and superharmonic
on S), vanishes at 4, and therefore vanishes identically on b). Thus
u must be continuous on D - A, and Theorem 9 can now be applied to -u
to provide a subharmonic extension u" of u to D. The function u' - u"
vanishes identically on D because this difference is superharmonic on D
and coincides !N almost everywhere with the harmonic function 0; so u'
is the desired harmonic extension of u.

Application to Analytic Functions

If A is a polar subset of the open subset D of R2, if A is closed in D, and


if f is a bounded analytic function on D - A, the real and imaginary parts
off have harmonic extensions to D according to the preceding paragraph.
The extension to D off obtained in this way is analytic because the Cauchy-
Riemann equations are satisfied. If A is a singleton this result reduces to
Cauchy's classical theorem on isolated singularities of bounded analytic
functions.

Application to Isolated Singularities of Harmonic Functions

Let u be a function defined and harmonic on an open deleted neighborhood


B of a point i; and satisfying

lim if N > 2.
n-4

Then Theorem 5 implies that there is a constant c such that the function

ifN=2,
ifN>2
has a harmonic extension to B u We give the proof for N = 2. According
to (5.4) there is a constant ct for which the function u + ct log I is - I

positive on a deleted neighborhood of ; so this harmonic function has a


superharmonic extension to B u { }. Since the Riesz measure associated
62 1. V. Polar Sets and Their Applications

with the extension must be supported by there is a constant c such that


the function u + clog I - J has a superharmonic extension u' to B u { }
and that the Riesz measure associated with u' vanishes identically. Hence
u' is harmonic. Observe that by way of inversion in a sphere of center the
result just obtained implies that if v is a function defined and harmonic on
a deleted open neighborhood of the point co of 68" and if

lim inf V01) > - 00 if N = 2,


v-m logInI (5.5)

liminfv(,)>
q-m
-oo ifN> 2,

then there is a constant c such that the function

ifN=2,
(v-c)1-1'-2 ifN>2
has a finite limit at the point co.

Application to a Generalized Liouville Theorem (N = 2)

According to Section 11.2 a lower-bounded harmonic function on 68^` for


N a 2 must be identically constant. The preceding application of Theorem 5
implies when N = 2 that a harmonic function on R2 which satisfies (5.5)
must be identically constant because if c is chosen so that v - c log I I has
a finite limit at the point co, then v must have a limit a at the point 00, finite
or infinite according as c = 0 or c # 0. In view of the maximum-minimum
theorem for harmonic functions a is finite and v =_ a.

Application to the Greatest Subharmonic Minorants of Superharmonic


Functions

Let D be an open subset of 68^' and suppose that u is a superharmonic function


on D for which GMDu exists. Let A be a subset of D, closed relative to D,
and define Do = D - A. Then GMD (uiDo) z GMDu on Do. We now show
that there is equality if A is polar and is null for the Riesz measure associated
with u. We can assume, replacing u by u - GMDu if necessary, that GMDu =
0, and it is then sufficient to prove that if h is a function defined and harmonic
on Do, with 0 5 h S u, it follows that h = 0. According to Theorem 5, such
a function h has a positive superharmonic extension h, from Do to D and
similarly u - h has a positive superharmonic extension v from Do to D. Since
u = h, + v on Do, it follows that u = h, + v on D and therefore A must be
null for the Riesz measure associated with ht. Hence h, is positive and
7. Superharmonic Function Minimum Theorem (Extension of Theorem 11. 5) 63

harmonic on D and is majorized there by u so h, = 0, and therefore h = 0,


as was to be proved.

6. Greenian Sets in Ii2 as the Complements of Nonpolar Sets


Theorem. A nonempty open subset D of 682 is Greenian if and only if 682 - D
is not polar, equivalently, if and only if R2 n OD is not polar.

If R2 - D is polar (equivalently, according to Section 2, if 682 n aD is


polar) D is not Greenian because a positive superharmonic function on D
can be extended to be a positive superharmonic function on R2 and therefore
(Theorem 11.13) is a constant function. Conversely, suppose that R2 n aD
is not polar. If D is bounded, it was noted in Section 11. 13 that D is Greenian.
If D is unbounded, let A be the intersection of 682 - D with a closed disk
so large that A is not polar. Let B be a disk containing A and define v on
682-Aby
RI on B - A (reduction relative to B),
l
0 onR2-B.
In view of the fact that v is harmonic and positive on B - A with limit 0
at every boundary point of B [Section 111.4, Example (a)], the function v
is subharmonic; so 1 - v is positive and superharmonic on D but not identi-
cally constant there because v = 0 on D - B but (Theorem 4) v > 0 on
D n B. Hence D is Greenian.

7. Superharmonic Function Minimum Theorem (Extension of


Theorem 11.5)
The obvious application of the following theorem to harmonic functions is
left to the reader.

Theorem. Let D be a Greenian subset of RN and let u be a lower bounded


superharmonic function on D. Suppose that there is a constant c such that
Um inf,,_tu(i) z c at quasi every finite point C of aD, and also at i; = oo if
N> 2 and D is unbounded. Then u Z c.

Observation (a). We are only excluding trivia by the hypothesis that D


be Greenian because if N > 2, all nonempty open subsets of R" are Greenian
and if N = 2, a lower-bounded superharmonic function on a non-Greenian
open set is identically constant. (Moreover, according to Theorem 6, a
64 I.V. Polar Sets and Their Applications

condition on quasi every finite boundary point of a non-Greenian open


subset of R2 is necessarily satisfied vacuously.)
Observation (b). Since the set A of finite boundary points of D at which
u has inferior limit < c is a countable union of compact sets,

A= U )CeiD:ICIsn,liminfu(ry)Sc-I
n
,
R=1 9'C

the set A is polar if it is inner polar. Thus Theorem 7 is true if "quasi every"
is replaced by "inner quasi every."
To prove the theorem, observe that according to Theorem 2 if N > 2
there is a positive superharmonic function v on R", identically + 00 on A ;
if N = 2 and D is bounded, there is a positive superharmonic function v on
a ball containing D, identically + oo on A. In either case if e > 0 the function
(u + sv),D is superharmonic on D with inferior limit z c at every point of
OD, including oo if N > 2 and D is unbounded. Hence u + vv z c on D by
the superharmonic function minimum theorem of Section 11.5, and therefore
u z c quasi everywhere on D and so everywhere on D. If N = 2 and D is
unbounded, we can suppose that lim inf,.,, u(PI) z c because the plane can
be inverted in a circle with center a finite boundary point of D not in A,
so that the transformed superharmonic function on the image of D has this
property. Thus if c' < c, there is a disk B so large that u z c' on D - B. The
part of the theorem already proved yields the inequality u z c' on D n B
from which it follows that u z c on D, as was to be proved.

Application to Analytic Function Theory

Let f be a bounded complex-valued function defined and analytic on a


Greenian subset D of the plane; that is, the complement of D is nonpolar. The
function If I is subharmonic; so it follows from Theorem 7 applied to - If I
that if lim sup,.-=I f(z')I 5 c at quasi every finite boundary point of D, then
If I <conD.

8. Evans-Vasilesco Theorem

Theorem. Let D be either R2 or a Greenian subset of R' (N z 2), and let


GD,u be a superharmonic potential on D. Then if A is a closed (in D) support
of µ, continuity of (GDp)J4 at a point of A implies continuity of GDit at .

The triviality of the extension of this theorem from D special open to D


general Greenian will be explained in Section VII.8. Accordingly we assume
8. Evans-Vasilesco Theorem 65

here that D is either R2 or a special open subset of R" (N ,':t 2). We can
suppose that p has compact support because the GD potential of the projec-
tion of u on A - b) is harmonic and therefore continuous on D n S)
for every S > 0. Under the hypothesis of compactness of A, the restriction
(Gp)ID is superharmonic and is the sum of GDu and a function harmonic on
D (Section IV.1); so it follows that under the hypotheses of Theorem 8, the
function (Gp)IA is continuous at and is finite there, and it is sufficient to
prove that Gp is continuous at . Suppose first that N > 2. If pt8 is the
projection of i on finiteness of implies that 0 and
that limj,0Gp 0. Furthermore, since G(µ - p4') is continuous on
S), the function (Gp4l)IA must be continuous at . If e > 0 and if S is
sufficiently small, depending on e, Gµ480) < e. For such a value of b conti-
nuity of (Gp46 )IA at implies that (Gp J)IA < e in some neighborhood of ,
and hence this inequality is true for all smaller values of S. It follows that

lim sup Gµ4b(C') = 0. (8.1)


6-0 4'e An (C,b)

If C is a point of D, let C' be a point of A at minimum distance from C so


that if n e A,

!z'-nI51C'-CI+IC -n1521C-nl. (8.2)

The oscillation of Gµ at can be majorized as follows:

5 sup Gpl"(C)
Ce B((.a(2)
(8.3)
-5;2'-' sup Gp'6(C')
Cs B(4.a/2)

Now C' e S) n A when C e S/2) because I - CI < 6/2; so the right


side of (8.3) has limit 0 when S - 0, by (8.1). That is, Gµ is continuous at ,
as was to be proved. When N = 2, the only change needed in the preceding
argument is that in (8.3) it should be supposed that S < I to ensure positivity
of the potentials involved, and the last term in (8.3) should be replaced by

sup G, ({') + S)) log 2,


CE B({.8/2)

which tends to 0 with S.

Observation. If in the theorem it is supposed that (GDµ)IA is finite valued


and continuous, it follows that GDu is finite valued and continuous on D.
If D is special open and if p has compact support in D, the function GDp is
66 I.V. Polar Sets and Their Applications

then necessarily bounded because this potential has limit 0 at every boundary
point of D. For arbitrary Greenian D the boundedness of a continuous
finite-valued potential GDp for p of compact support in D follows from the
boundedness properties of GD to be proved in Section VII.5.

9. Approximation of a Potential by Continuous Potentials


Theorem. Let D be either 682 or a Greenian subset of R' (N z 2) and let p
be a measure on D. Suppose that GDµ is superharmonic and that GDP < + oo
on some support A of D. There is then a sequence p. of measures on D for
which µ is supported by a compact subset of A, GDµ is finite valued and
continuous (bounded if D o 682), and p = p.; so GDP = L°. o GDµfl.

Recall our convention that "supported by A" means that A is p measur-


able and that the complement of A is p null. The generality of the statement
of the theorem is convenient for reference, but there is no loss of generality
in supposing that A is a Borel set because there is always a Borel support
of p that is a subset of A. Apply Lusin's theorem to find a sequence A. of
disjoint compact subsets of A with the property that p(A - U; o 0
and that (GDp)JA. is bounded and continuous. If p is the projection of p
on A,,, the continuous function (GDP)IAM is the sum of the restrictions to A.
of the lower semicontinuous functions GDµ and GD(p - pa). Hence these
restrictions are continuous, and therefore (by the observation in Section 8)
GDp is bounded and continuous on D unless D = R2, in which case
GDp = Gp is at least finite valued and continuous.

Corollary. If B is a compact (but see the extension below) nonpolar subset


of 68', there is a measure v on R' supported by B for which if N = 2, the
function Gv is finite valued, continuous, and not identically 0 and if N 2,
the function GDv is strictly positive, continuous, and bounded whenever D is
a connected Greenian superset of'B.

If D is a Greenian superset of B, the smoothed reduction (relative to D)


RB is the potential of a measure p supported by B [Section IV.8(d)], and
according to Theorem 9, there is a measure v supported by B with GDv as
described. In particular, if N = 2, the potential Gv is then finite valued and
continuous on 682.

Extension of the Corollary to Analytic Sets B

It will be proved in Section VI.2 that a nonpolar analytic subset of R" has
a nonpolar compact subset. This fact will imply that the above Corollary
is true if B is an analytic nonpolar set.
10. The Domination Principle 67

10. The Domination Principle


The following theorem will be referred to as the domination principle,
a designation sometimes used more restrictively.

Theorem. Let D be a Greenian subset of R', let p be a measure on D, and let


v be a positive superharmonic function of D. Then each of the following three
conditions implies that GDµ 5 v:
(a) GDP < + co p almost everywhere and GDp 5 v inner quasi everywhere
on some Borel support of p.
(b) The inequalities GDp < +oo and GDp 5 v are true p almost every-
where.
(c) Polar sets are p null and GDP 5 v p almost everywhere.

It is not obvious that conditions (a)-(c) are equivalent, and we shall not
need this equivalence in the proof of Theorem 10, but according to Theorem
11, both conditions (a) and (b) imply that polar sets are p null and thereby
that (a)-(c) are equivalent. It is a defect of Theorem 10 that polar sets must
be y null, but it will be shown in Section XI.23 that GDp 5 v if

lim sup G°p(q) 5 1, (10.1)


n-t V01)
GDph)<+ao

for p almost every . In fact the condition in Theorem XI.23 is considerably


weaker than (10.1). If p is a probability measure supported by a singleton
so that GDP = and if v = the condition (10.1) but
not Theorem 10 is applicable to show that GDp 5 V.
It is sufficient to show that GDp 5 v when condition (a) is satisfied. In
fact, if B is an arbitrary Borel support of p, the set B n {GDP < + oo, GDP 5 v}
is, under either (b) or (c), a support of p on which GDP < + oo and GDp 5 v;
so condition (a) is satisfied. To prove that condition (a) implies that GDp 5 v
we assume that D is special, deferring to Section VII.8 the explanation of
why this specialization is trivial. Let A be a Borel support of p on which
GDP < + oo and GDp 5 v inner quasi everywhere. Since (by Theorem 9) the
function u = GDP is the limit of an increasing sequence of bounded con-
tinuous potentials of measures supported by compact subsets of A, we can
suppose that A is compact and that u is bounded and continuous. If e A
and if 5 v(c), then

lim inf v(q) uO = lim u(p). (10.2)


rC rC

Thus the restriction to D - A of v - u is superharmonic, is lower bounded,


and has positive inferior limit at inner quasi every point of A n a(D - A).
68 1. V. Polar Sets and Their Applications

Moreover every limiting value of this restriction at a point of 8D is positive


because (Section IV.1) u has limit 0 at every such point. (If D is not special,
it will be seen in Section VII.5 that u has limit 0 at quasi every point of 8D,
including the point oo if D is unbounded and N > 2.) It follows (Theorem 7)
that v ;-> u on D - A, so v z u inner quasi everywhere on D and hence
(Section 1) everywhere on D, as was to be proved.

Special Hypotheses on r

If v in Theorem 10 is a constant function, the theorem is sometimes called


the principle of the maximum. If v is the sum of a constant function and a
potential, the theorem is sometimes called the complete principle of the
maximum.

The Domination Principle for Potentials Gp When N = 2

If N = 2, if Gp is the superharmonic potential of a measure y on R2, and if


there is a constant c such that Gp < c p almost everywhere then the method
of proof of Theorem 10 shows that Gp < con R2.Observe that if c is replaced
by a harmonic function h in this hypothesis then any further hypothesis on
h leading to the conclusion that Gp S h on R2 would imply (according to
Section IV.9) that lim infi,i_m oo and that therefore (Section
5) h is identically constant.

11. The Infinity Set of a Potential and the Riesz Measure


Theorem. If D is a Greenian subset of RN, if A is a polar subset of D, and if
u = GDp is the potential of a measure on D, then u = + ao p almost everywhere
on A.

(The phraseology of the theorem requires that p be a completed measure.)


We can assume that D is connected. If u is not superharmonic then u - + co,
so the theorem becomes trivial. If u is superharmonic the conclusion of
the theorem can be restated in the following form: if u is finite valued on A
then p(A) = 0. It is this assertion that will be proved. The triviality of the
extension from special to Greenian sets will be explained in Section VII.8.
Suppose then that D is special and that u = GDp is finite valued on the polar
set A. Since the theorem is local it can be assumed that A is relatively compact
in D. It can also be assumed that A is a Borel set because there is a polar
Ga superset of A and the intersection of this superset with the set of finiteness
of u is a Borel set. If A. = { eA: u :!g n} and if p is the projection of y on
A then an application of the domination principle to the pair up = GDp,,,
v =- n shows that u < n on D. Let GDv be a superharmonic potential identi-
11. The Infinity Set of a Potential and the Riesz Measure 69

cally + oc on A, with v(D) < + oo. Apply Fubini's theorem and the symmetry
of GD to derive

nv(D) z f GDp,dv= f
D ,JD

It follows that µ(A,) = 0; so µ(A) = 0, as was to be proved.

Application to Gu When N = 2

If N = 2, if p is a measure on R 2, and if u = Gp is superharmonic, the con-


clusion of the theorem remains true. It is sufficient to prove that u = + 00
p almost everywhere on any bounded polar set A. Let D be a disk containing
A and let µD be the projection of p on D. According to Theorem 11, GDPD =
+ oo µD almost everywhere on A. Since GDµD - Gp is harmonic on D if
defined suitably on the set of common infinities of GDPD and Gp, it follows
that Gp = + oo p almost everywhere on A, as asserted.
Chapter VI

The Fundamental Convergence Theorem and


the Reduction Operation

1. The Fundamental Convergence Theorem


Theorem. Let r : {u a e I } be a family of superharmonic functions defined on
an open subset of R', locally uniformly bounded below, and define the lower
envelope u by u

inf

u
u u u
u
a countable subset of r whose lower envelope has the same
lower semicontinuous smoothing u.
Conversely, if A is a polar subset of a Greenian subset D of RM, there is a
decreasing sequence v, of positive superharmonic functions on D with limit v
such that v > v on A.

The direct part of the present theorem is identical with Theorem 111.3
except that Theorem IIi.3(c) allows a larger exceptional set than Theorem
1(c). Thus there remains only the proof of Theorem l (c) and of the converse
part of Theorem 1. Since Theorem 1(c) is a local assertion it can be assumed
in its proof that the functions are defined on a ball D, and in view of the
discussion in Section 111.3 it can be assumed that r is a decreasing sequence
of positive superharmonic functions on D. The limit function need only be
analyzed on a strictly smaller concentric ball B, with u replaced by
(reduction relative to D). This reduction, equal to its lower semicontinuous
smoothing because B is open, is a superharmonic potential Gals. (Section
IV.8) and is equal to u on B, so the replacement is legitimate. The measure
µ is supported by B. On D - B the sequence u, is a locally uniformly con-
vergent sequence of harmonic functions; so the sequences of partial deriva-
tives are also locally uniformly convergent on D - B (Theorem 11.3).
2. Inner Polar versus Polar Sets 71

Evaluation of µ (D) by the Gauss Integral Theorem shows that µ,(D) is a


convergent sequence. Going to a subsequence if necessary, it can be assumed
that the sequence p, is vaguely convergent to a measure p supported by B.
If c is a strictly positive constant, the function GD A c is continuous and

lim U. = u Z lim f Ac I GD(-, 1) A c p(dq), (1.2)


°-'m JD JD

so that u Z GDµ. Let v be a measure having compact support in D, with GDv


finite valued and continuous. Then

`GDvdµ < +oo.


f. GDP.dv = 0

The equality

f udv = lim f lim `GDvdµ = f GDVdµ = GDµdv, (1.3)


D
X-°D,JD 4~00 JD D D

combined with the inequality u Z GDµ, implies that u = GDµ at v almost every
point of D. Since # is the maximal lower semicontinuous minorant of u, it
follows that u z u 2t GDµ with equality v almost everywhere. Thus, if A =
(u > it has now been shown that v(A) = 0 whenever GDv is finite valued
V},

and continuous. In view of Corollary V.9 this fact means that every compact
subset of A is polar; that is, the set A is a Bore] inner polar set. Theorem 1(c)
follows from the next theorem, whose proof uses the partial result just
obtained.
Conversely, suppose that A is a polar subset of the Greenian subset D of
R', and let r be the class of positive superharmonic functions on D equal at
least to I on A. The infimum of the class is (reduction relative to D) RA,
and (Theorem VI.4) RA m 0. According to Theorem 1(d), there is a sequence
in r whose lower semicontinuous smoothed infimum vanishes identically. If
v is the minimum of the first n members of this sequence, v, has the properties
stated in the converse of Theorem 1.

2. Inner Polar versus Polar Sets


Theorem. An analytic inner polar subset of R' is polar.

It is obviously sufficient to prove the theorem for bounded sets so we


shall consider subsets of a Greenian set D, say a ball. Reductions below are
relative to D. The proof will be in several steps. Some of the preliminary
results proved below are more general than needed; this is to avoid later
repetition when the results will be strengthened.
72 1.V1. The Fundamental Convergence Theorem and the Reduction Operation

(a) If A is a compact subset of D and if v is a finite-valued positive contin-


uous superharmonic function on D, then R" = v quasi everywhere on A. In
fact, according to the part of Theorem I actually proved in Section 1, the set
ao

An{RA <v}=UAn{RA 5(1 - l/n)v}

is inner polar. Since each set in the union is compact, the union is polar.
(b) If v is a finite-valued positive superharmonic function on D and if
e D, the set functions and are strongly subadditive on the class
of compact subsets of D. To prove this, consider the strong subadditivity
inequality

R"' +RAn85R +RB (2.1)

and the corresponding inequality (2.1sm) for smoothed reductions. In-


equality (2.1) is trivially satisfied on A n B. On one of the sets but not on the
other, say on A - B, (2.1) reduces to the inequality v + RV" S v + R.
Thus (2.1) is satisfied on A u B, and hence (2.lsm), an inequality between
superharmonic potentials, is satisfied inner quasi everywhere on A u B, a
support of the measures associated with these potentials. It follows from the
domination principle that (2.1sm) is true on D. Since (2.lsm) is true, (2.1)
is true on the open set D - (A u B) on which all the reductions are harmonic
and equal to their smoothings. We have already verified (2.1) on A U B.
The foregoing proof depended on the domination principle, not available
in a form strong enough for this proof in parabolic potential theory. A proof
of reduction operator strong subadditivity not depending on the domination
principle is given in Section 4 [proof of Section 3(j)].
(c) If v is a ftnite-valued positive continuous superharmonic function on D
and if A, is an increasing sequence of compact subsets of D with compact union
A (e D), then

lim R"" = RA , (2.2)

and the corresponding equation (2.2sm) for smoothed reductions is also true.
Under these hypotheses R" , = v quasi everywhere on A because
RA A. = v quasi everywhere on A,,, and the two superharmonic potentials
lim,,.-m R". and RA are therefore equal quasi everywhere on the common
support A of their associated Riesz measures. It follows from the domination
principle that these potentials are identical. The functions R," and
Ro are trivially equal on A and are equal on D - A (on which they are
harmonic and equal to their smoothings) because (2.2sm) is true.
(d) If v is afinite-valued positive continuous superharmonic function on D
2. Inner Polar versus Polar Sets 73

and if A. is a decreasing sequence of compact subsets of D with intersection A,


then (2.2) is true. In fact; if B is an open neighborhood of A the set B must
also be a neighborhood of A. for sufficiently large n, so lim,,..,,, Ro " S R.
and since v is finite valued and continuous, it follows from III(5.5) that (2.2)
is true with equality replaced by the inequality :!9. Since the reverse inequality
is trivial, (2.2) is true.
(e) If v is afinite-valued positive continuous superharmonic function on D
and if E D, the set function R;, is a Choquet capacity on D relative to the
class of compact subsets of D. According to (b)-(d) and the monotoneity of
the reduction operator and under the stated hypotheses on v, the restriction
of the set function to the class of compact subsets of D is a topological
precapacity. This restriction therefore (Appendix 11.8) has an extension to
a Choquet capacity on D relative to the class of compact subsets of
D, for which I(i, F) = R" (l;) when F is compact and

B) = sup {R" (1; ): F c B, F compact} (B open), (2.3)

A) = inf {I(l;, B): A B, B open} (A arbitrary). (2.4)

To prove (e) we show that A) = Ro (c) for every subset A of D. Let F°


be an increasing sequence of relatively compact open subsets of the open
subset B of D, with union B, and define F = F° . Then Roy = R"°, and B)
= lim,,..,, R'"
+11
is a positive superharmonic function, identically v on B. Hence
B) z R,,, and since the reverse inequality is trivial, it follows that B)
= RB when B is open. Since both Ro(d) and satisfy (2.4), the set
functions R.(%) and 1(l;, ) are identical.
(f) Proof of the theorem. Let v be any finite-valued strictly positive
continuous superharmonic function on D, say v =- 1. Suppose that A is an
inner polar analytic subset of D and choose c in D - A. The set A is capacit-
able for the Choquet capacity Ri(d); that is,

sup {Rn (): F c A, Fcompact}. (2.5)

According to Theorem V.4, the right side of (2.5) is 0, and the consequent
vanishing of the left side implies that A is polar. o

Characterization of a Nonpolar Set

According to Theorem 2 every analytic nonpolar subset of R has a compact


nonpolar subset. Hence Corollary V.9, which asserts the existence of a not
identically 0 continuous finite-valued potential supported by a nonpolar
compact subset B of R', remains valid if B is supposed only nonpolar
analytic.
74 1. V1. The Fundamental Convergence Theorem and the Reduction Operation

3. Properties of the Reduction Operation


In the following D is a Greenian subset of R", provided with a boundary 8D
by a metric compactification. Reductions of positive superharmonic func-
tions u, v, ... on subsets of D u 8D are considered, and no further hypotheses
not stated explicitly are imposed on these functions and sets. The following
list of properties includes for the reader's convenience some properties
already discussed in Sections 111.4 and 111.5. Reduction properties linked
directly to notions of capacity are treated in Section 5.
Although it may seem logical and efficient to prove Theorem 3 and
related theorems by methods also applicable to the parabolic potential
theory treated in Chapters XV to XIX, it is unfortunately true that the most
natural methods in the present context are not all applicable to parabolic
potential theory. For this reason the methods used here will be those specially
adapted to the present theory, and the parabolic counterparts of the reduc-
tion properties listed below and proved in the next section will be proved by
different methods and in a different order, although those methods can also
be used in the present context.
(a) R" = u on A n D; RA = R = v on the interior of A n D; R. = RA
when A n D is open; Rn is harmonic on D - A and is equal to RA
there; RA lim in4-c RV (n). [See also Chapter 111(5.1),(5. Ism),
(5.2).]
(b) R" 5 R" on D, with equality on D - A and quasi everywhere on
An D.
(c) If At and A2 differ by a polar subset of D, then RA' = Rn2 on
D-(A,uA2)and R "'=R"2 on D.
(d) Ro = inf { Ro A0 : A - An a polar subset of D) = inf {u: u 0, u super-
harmonic on D, u Z v near A n 8D and quasi everywhere on
AnD).
(e) If A. is an increasing sequence of subsets of D with union A and if
v, is an increasing sequence of positive superharmonic functions on
D with superharmonic limit v, then

lim RA = R",
n-w

and the corresponding equation (3.Ism) for smoothed reductions is


also true. If v = v for all n, then (3.1) and (3.1 sm) are true for A. an
increasing sequence of subsets of D u 8D.

Observation. The following example shows that the last assertion of (e) is
false without the hypothesis that v = v for all n. Let D be a ball, let u be a
minimal harmonic function on D corresponding to some boundary point C,
that is, v is a strictly positive multiple of the Poisson kernel for the boundary
3. Properties of the Reduction Operation 75

point C, and define v = u A n. An application of Section IV.8(a) shows that


the function v is a potential, because this function is positive, bounded, and
superharmonic and has limit 0 at each boundary point except C. Finally,
R,'° = Rv° = v, RODVR
=+V.
RID = 0, and v = v, contrary to (3.1).
(f) If v = Yo v is a sum of positive superharmonic functions on D and
is superharmonic, then
00

R"= Rte, (3.2)


0

and the corresponding equation (3.2sm) for smoothed reductions is


also true.
(g) If A c 8D and if v, is a decreasing sequence of positive super-
harmonic functions on D with limit v, then R A = R+.
(h) If A c B,

OOVO"OB = OOVOBOA = Ov0

(i) If v'= R"


+V
A Re,
+ then
RAUB + RAIuB = RA
+ RB, (3.4)

and the corresponding equation (3.4sm) for smoothed reductions is


also true.
(j) The set functions Ro and R' are countably strongly subadditive.
(k) If v is finite valued and if e D, the set function is a Choquet
capacity on D u aD relative to the class of compact subsets of
D V OD.
(1) If A is analytic

Rn = sup (R": F c A, F compact}, (3.5)

and the corresponding equation (3.5sm) for smoothed reductions is


also true.
(m) The equality

Rv = inf {RB: A c B, B open} (3.6)

is true if either (i) v is finite valued and continuous at each point of


A n D (for example, if v is arbitrary and A c aD) or (ii) if v is
finite valued and A n D is analytic. Moreover (3.6) implies

R"
+u
= [inf {RB
+v
: A c B, B open}].. (3.7)
76 I.VI The Fundamental Convergence Theorem and the Reduction Operation

(n) if u and v are bounded then

sup JR - RAI S sup Ju - v1, (3.8)


D D

and the corresponding inequality (3.8sm) for smoothed reductions


is also true.
(o) Denote DuQc-, d Qu0`, Qc= , respectively, by uc,, uc,c=+ .

(o,) Let v, v', v", h' be functions from D into l8+ with v" = v' + h', and
suppose that v, v", and h' are positive superharmonic functions. Define
A={v:v'}and B={v-v"}.Then
hA + hABA + hABABA + . ' 5 v" (3.9)

and

hAB + hA8AB + ... 5 v8 5 V A V" (3.10)

(02) Let v and h be positive superharmonic functions on D, let a and b


be numbers with 0 5 a < b, and define A = {v:5 ah} and B = {v;_> bh}.
Then

V nb bh)
h,48 5 hABAB 5 (b) hAB+ hABABAB 5 (b) 2hAB+ .... (3.11)

so

V A (bh)
hAB h ARAB + S (3.12)
b-a
Furthermore

e A (ah) (a a\2
hBA 5 b hBABA 5 I b) hBA+ hBABABA 5 (b f hBA, ... ; (3.13)

so

Ub (ah).
hBA + hBABA + .. 5 (3.14)

Observation. Since hB 5 h', it is trivial that (3.9) is true with the left side
replaced by hsA + hBABA + . We shall see in Section XI.4 that the inequal-
ities under (o) yield limit properties of superharmonic functions and of ratios
of superharmonic functions. See Sections 2.111.12 and 2.111.22 for the
probability counterparts of these inequalities.
4. Proofs of the Reduction Properties 77

4. Proofs of the Reduction Properties


Proof of (a). See Sections III.4, III.5(d), VI. l.

Proof of (b). According to the Fundamental Convergence Theorem in


Section 1, R 5 R, and there is equality quasi everywhere on D. If E D - A,
let u be a positive superharmonic function on D, identically + oo on the polar
set A n { R ^D < Rn nD} but finite at (see Theorem V.2). Then v S R,"^ D + cu
on A when s > 0 so Ro and therefore
R ,D(). So there must be equality; that is, R" = R" on D - A when A c D.
R
If A n 8D is not empty the expressions 111(5.2) and III(5.2sm) for reductions
and smoothed reductions in terms of those on A n D and A n 8D separately
yield the equality of RA and R " on D - A. c3

Proof of (c). Let be in D - (A1 u A2), and (Theorem V.2) let u, be a


positive superharmonic function on D, identically + oo on the polar set
(A 1 - A 2) u (A 2 - A 1) but finite at . If u is a positive superharmonic
function on D, majorizing v on A, n D and near A, n 8D, then for every
e > 0 the function u + eu1 majorizes v on A2 n D and near A2 n D. Hence
u(g) + Z RA ,2(& so RA,, (t) Z RA and by symmetry the reverse
inequality is true so RA, = RA z on D - (A, Q A2). It then follows that the
superharmonic functions R"'
+V
and R"z+V
are equal quasi everywhere on D and
therefore everywhere on D.

Proof of (d). The first line of (d) and the equality of the two infima are
immediate consequences of (b) and (c).

Proof of (e). Under the hypotheses of (e) the reduction R increases with
n. Define u = Ro n, and observe that u is superharmonic and that
u = v quasi everywhere on A because un = v quasi everywhere on A. It
follows from (d) that u z RA , and since the reverse inequality is trivial,
(3.lsm) is true. Equality (3.1) is trivial on A and is also true on D - A
because on that set R"n = R"
+v, and Rv = R"
+v
according to (b). In proving the
second assertion of (e) we shall suppose that D is connected to avoid irr-
evant notational complexity. Assume then that v = v for all n, choose a
point in D, and let u be a positive superharmonic function on D, majorizing
v near A. n 8D, with

un(c) 5 R fr 2-".
The function

vk=u+(un_R"^^dD),
k>0, (4.1)
n=k
78 I. V1 The Fundamental Convergence Theorem and the Reduction Operation

is positive and superharmonic on D because the sum is the limit of an


increasing sequence of positive superharmonic functions and is finite at .
Moreover vk Z u Z v quasi everywhere on A n D because RIM = v quasi
everywhere on A. n D, and

vk z RV a° + (un - Ro" a°) z U.

when n z k ; so vk majorizes v near A n 8D, and therefore vk z R. A. When


k oo, we find that u z RA quasi everywhere on D and therefore everywhere
on D. Since the reverse inequality is trivial, (3.1sm) is true, and as in the first
case of (e), it follows that (3.1) is true.

Proof of (f). Suppose first that there are only two summands in (3.2); so
the equalities

Ru R", + RV, (4.2)

and (4.2sm) are to be proved. By subadditivity (4.2sm) is true for "S"; so


(natural order decomposition) there are positive superharmonic functions
u1, u2 on D, satisfying
RA+v =ut+u2, (4.3)

If A is an open subset of D, then R"


+v,
= v, on A and R"
+v, +vy
= v, + v2 on A
so u; = v, on A, and it follows that u; 2, RBA. So (4.2sm) is true. If now A is
i
still an open subset of D and if v, and v2 are supposed finite valued and
continuous, equation 111(5.5) gives the value of a reduction on an arbitrary
set in terms of reductions on open supersets, and it follows from (4.2sm)
[the same as (4.2) for open sets] that (4.2) is true for an arbitrary subset A
of D. Since a positive superharmonic function on D is the limit of an increas-
ing sequence of finite-valued positive continuous superharmonic functions,
(e) implies that (4.2) is true for an arbitrary subset A of D with no restriction
on v v2, and then (4.2sm) must also be true with this generality because the
two sides of (4.2sm) are superharmonic functions and are equal quasi every-
where on D. The evaluations in Section 111.5 of reductions and their smooth-
ings in terms of reductions and their smoothings on subsets of D show that
(4.2) and (4.2sm) are true with no restrictions on the sets or functions. Thus
(3.2) and (3.2sm) are true for two and therefore any finite number of sum-
mands. If there are infinitely many summands, write r" for R'. Then
k

n
A=
Rv o
A
Rvk +RAn
4. Proofs of the Reduction Properties 79

and since R" 5 F+ , vR, it follows that (3.2sm) is true quasi everywhere and
therefore everywhere on D. Equation (3.2) is then true on D - A, and this
equation is trivial on A r D.

Proof of (g). Define h = GMDv and define u = R a positive har


monic function. Obviously

05 1ini
N_ QD
RV _t,=u-R;.
The difference u - Rw is a positive harmonic minorant of v - h for all n,
and therefore u - R,", 5 v - h ; so u = R; by definition of h. Finally, u = RA
because R ,, is a positive harmonic minorant of h and so vanishes
identically+

Proof of (h). In view of the trivial fact that the smoothed successive reduc-
tions of v on A and B in either order lie between and BvO", it is
sufficient to prove idempotence, that is, to prove that DBvO"DA = Qv0A. The
proof will be carried through in several steps.
(h,) If A c D, the desired idempotence is a consequence of (d) because
the condition on u in (d) is unchanged if v is replaced by OvOA.
(h2) If A c 8D, the following argument yields idempotence. Use
Choquet's topological lemma to find a decreasing sequence v, of
positive superharmonic functions on D, each majorizing v near A,
with sequence limit O v D". Replacing v by V. A v if necessary, it can be
supposed that v = v near A so that QV JA. It now follows
from (g) that

Jim vJA = D DvV"D". (4.4)


n-m
(h3) For arbitrary A,
OVOAnD DOVOADAndD = ODOAnJD
DOVOADAnD = (4.5)

These inequalities follow, respectively, from


OVBAnD = DOVOAnDDAnD 5 OOVOADA'D OVOAnD,
5
(4.6)
OVOAn#D = D DVBAndDDAndD < D OVOADAndD 5 OVOAnW

Finally, to prove (h), apply 111(5.2) and (f) to write D0vO"D" in the form

D OVOADA=
DOVOADAIW+D OVOAD"nD
- DOVOADAnODAnD
80 1 V1. The Fundamental Convergence Theorem and the Reduction Operation

and then apply (h,) and 111 (5.2) to show that the right-hand side is OvOA

Proof of (i). Observe that quasi everywhere on (A u B) n D,


jv0A
V + v' = + llvrr

so that
p v OA,H = OvOA DAv. +

(4.7)

= pupA + Ov 8.

Thus (3.4sm) is true, so (3.4) is true off A u B, and the latter equation is
trivial on (A u B) n D.

Proof of (j). We have derived a weakened version of (j) in Section 2. Instead


of going on from this result, we observe that (3.4sm) implies strong sub-
additivity of R* because if v' is defined as in Section 3(i), it follows that
v' = v quasi everywhere on A n B; so
RAUB RAnB = RAn B,
+u Z +V' +v

and therefore (3.4sm) implies


RA.B + RA
+p + RH,
(4.8)
+p +p +O

the strong subadditivity inequality. This inequality is trivial on A U B for


the unsmoothed reductions and is true for these reductions elsewhere on D
because it is true for the smoothed reductions. That is, both R;, and R, are
strongly subadditive. Countable strong subadditivity follows from strong
subadditivity by an application of (e).

Proof of (k). (This proof involves the domination principle and thereby
the Green function, so at this stage the arguments are relevant only for
special open sets D, but the extension of the domination principle to all
Greenian sets will be seen to be trivial once GD has been defined for every
Greenian set D in Chapter VII.) We shall not use in the following proof
the partial result derived in the course of proving Theorem 2 that if v is
continuous as well as finite valued, is a Choquet capacity on D relative
to the class of compact subsets of D. To prove (k) in the present context
observe that all the capacity properties have been verified for except
the property that

lim R' n = R- (4.9)


4. Proofs of the Reduction Properties 81

whenever A. is a decreasing sequence of compact subsets of D u aD with


intersection A. Define u = R" The function u is superharmonic, and
u = v quasi everywhere on A n D.
Proof of (4.9) when A. c D. In this case u and R" A are finite-valued
potentials on D (Section IV.8) whose associated Riesz measures are sup-
ported by A, and both potentials are equal to v quasi everywhere on A.
According to the domination principle these potentials are therefore
identical. Since R". is harmonic on D - A., the function u is harmonic on
D - A, so u = u on that set. Since R4 = R"- on D - A. and RA = R" on
D - A, equation (4.9) is true on D - A, and this equation is trivially satisfied
on A.
Proof of (4.9) when v is a potential. If B is an open neighborhood of
A n OD the sequence A. - B is a decreasing sequence of compact subsets of
D with limit A - B; so by what we have just shown, lim, RA.-B = RA-B

By subadditivity of R;,,
RBm-B
+ RAY^B 2t RBA-B + RA m^B Z lim R" (k 5 m). (4.10)
R-OD

Hence

RA-B + RB Z lim R (4.11)

and as B shrinks to A n OD, the left side becomes RAnD + RAID in view of
III(5.1). The second reduction vanishes identically because v is a potential,
so we find that

RA RBn° z lim RAE,


4-m

and the reverse inequality is trivial.


Proof of (4.9) in the general case. In view of the Riesz decomposition of
a positive superharmonic function v and of the reduction additivity property
Section 3(f), it is sufficient to prove (4.9) separately for v a potential and
for v harmonic. The proof for v a potential was just given. The proof for v
finite valued and continuous, and A c D, was given in Section 2 under
part (d) of the proof of Theorem 2, and the proof is valid for A a compact
subset of D u 8D. Thus (4.9) is true for v harmonic. The proof of (4.9) and
thereby that of (k) is now complete.

Proof of (1). Case 1. If v is finite valued, (3.5) follows from (k) since analytic
subsets of D u 8D are capacitable for the Choquet capacity
Case 2. If A c D but if v is not necessarily finite valued, apply Section 3(e)
to derive
82 1 VI. The Fundamental Convergence Theorem and the Reduction Operation

R,A = supRA'. = sup (sup Rl,,n :Fc A, Fcompact}


nzo n2

= sup {R,',,,: F c A, Fcompact, n > 0) = sup {Rp : F c A, Fcompact};

so (3.5) is true.
Case 3. In the general case write uF for Rp ^ d ° and observe that according
to Section 111.5(c),
OF =uF - Rv^ F=VF-RV^°+RvF°. (4.12)

From now on F is to be a compact subset of A. The class of these sets ordered


by inclusion is directed so each term on the right in (4.12) defines a directed
set whose limit we now evaluate. Let h be the harmonic component of the
Riesz decomposition of v. According to Section II1.5, uF = RndD; so by
Case I above

lim tF- RA^8°


h = RA^a°
v
= VA-
Ft

Since the class of sets F includes the compact subsets of A r D, Case 2 above
shows that limFt R1,D = RV"'. The same argument shows that if F' is a
compact subset of A, then

lim RfnD z lim RFnD


uF. = RAnD
VF. (4.13)
Ft VF Ft

Now Fi- uF is an upward-directed set of harmonic functions with limit


vA ; so vA is the limit of an increasing sequence {vFn, n c- Z+ }, and therefore
letting F' in (4.13) run through this sequence, we find from Section 3(e) that
limFt RFD > R0 . There must be equality because the reverse inequality
is trivial, and (4.12) now yields
limRF=vA-RA^D+RAD=R"
Ft

so (3.5) is true. Equation (3.5sm) is true on D - A because equations (3.5)


and (3.5sm) are identical on D - A, and this equality on D - A implies
(3.5sm) on A n D by the following argument. If e A, the left side of (3.5sm)
is unchanged according to Section 3(c) when A is replaced by A - The
right side is also unchanged in view of Section 3(c) and Section 3(e) [replace
Fin (3.5sm) by F less each of a sequence of balls of center and radii tending
to 0]. Thus (3.5sm) is reduced to (3.5).

Proof of (m). Assertion (i) has already been proved in Section III(e). To
prove (ii) observe that if v is finite valued on D, then in view of the properties
(J) and (k) the restriction of the set function to the class of compact
subsets of D v OD is a topological precapacity. This topological precapacity
4. Proofs of the Reduction Properties 83

generates a Choquet capacity on D v 3D relative to the paving of


compact subsets of D v OD, as described in Appendix 11.8. More specifically,
1(g, ) = on the class of compact sets, and when A is open

A) = sup F): Fc A, Fcompact}


= sup (R','(& Fc A, Fcompact}. (4.14)

Hence A) = RA(s) when A is open, in view of Section 3(e) (since A is a


countable union of compact sets) or Section 3(l). Finally, for arbitrary A,

A) = inf B: B D A, B open)
= inf {RB(A): B c A, B open} = RA(s)).

If A is an capacitable set, in particular, if A is analytic, (4.14) is true


and combined with (3.5) yields A) = R1.1(4) and thereby yields (3.6).
According to property (a), each reduction on the right in (3.6) is equal to
its smoothing; so an application of the Fundamental Convergence Theorem
to (3.6) yields (3.7). o

Proof of (n). If c = supo1u - vj, so that v s u + c, we find that Ro 5 c,


and interchanging u and v yields the other half of (3.8). This argument
applied to smoothed reductions yields (3.8sm). o

Proof of (ot). Obviously ue 5 v on D, and therefore h' + v8 5 v" on A.


Hence h; + veA 5 v"; so hABA + VBABA 5 UBA, .... It follows that (3.9) is
true. To prove (3.10), observe first that vB 5 v" and V" B:5 v; so vB 5 v A v".
Finally, take the smoothed reduction onto B of both sides of (3.9) to find
that (3.10) is true. o
Proof of (02). To prove (3.11) and (3.12), observe first that hAB 5 he 5 v/b
and that vA 5 ahA. Hence

hABA 5 h8A 5 b (b) hA (4.16)


5

from which it follows that hABAB 5 (a/b)hAB. Iterate and sum to derive (3.11)
and (3.12) with v instead of u A (bh). Relations (3.11) and (3.12) are true as
written because if v is replaced by V A (bh), the sets A and B are unchanged.
To prove (3.13) and (3.14), observe that the inequality between the second
and fourth terms in (4.16) implies

hBABA 5 ()hAaA S ()h.


84 I V1 The Fundamental Convergence Theorem and the Reduction Operation

Iterate and sum to show that the left side of (3.14) is at most bh5A/(b - a).
Since (4.16) implies that bhBA < v and bheA 5 ah, (3.13) and (3.14) are true.
The reader is invited to derive (3.12) and (3.14) from (ot).

5. Reductions and Capacities


Let D be a Greenian subset of R', provided with a boundary SD by a metric
compactification, let t be a point of D, and let v be a positive finite-valued
superharmonic function on D.

Relation between the Capacity R,; and Topological Precapacities

According to Section 3(k), the set function is a Choquet capacity on


D u 8D relative to the class r of compact subsets of D u 3D. This fact
combined with Section 3(j) implies that the restriction of to r is a
topological precapacity. Let be the Choquet capacity on D U 8D rela-
tive to r generated by this topological precapacity. We have seen in the
course of proving Section 3(m) that 5 with equality on the
class of -)-capacitable sets and on the class of all boundary subsets, and
that there is equality for all sets if v is continuous as well as finite valued.

Finiteness of v Is Necessary in Section 3(k)

Let D be a ball of radius 1, let be the center of D, for n > 1, let A. be a


closed concentric ball of radius I/n, and define v = Then A. is a
decreasing sequence of compact subsets of D with intersection A = { }, and
RA. = RA is false because RA = v for all n even though RA vanishes
except at . Thus the finiteness of v is necessary in Section 3(k). Observe
that with D, A, v, as just defined, (3.6) is false because the left-hand side
of (3.6) is a function vanishing except at whereas the right-hand side is v.

Is Not a Choquet Capacity on D Relative to the Class ro of


Compact Subsets of D

Choose v continuous and choose a compact subset A of D containing in


such a way that RA If A. is a decreasing sequence of compact
neighborhoods of A with intersection A, then RAn(c) = RA (g) for all
n, and therefore even if v is finite valued and+vcontinuous, the +v
set function
is not a topological precapacity on r'o and is not a Choquet
capacity on D relative to f0.
Chapter VII

Green Functions

1. Definition of the Green Function GD


Let D be a nonempty open subset of R. If N > 2 or if N = 2 and D is
bounded, the function is lower bounded for each point of D; so
exists (Section 111. 1). If N = 2, if D is unbounded, and if
has a subharmonic minorant on D for some in D, then the minorant
exists for every in D. In fact is bounded below
outside each neighborhood of , and G(b', ) is bounded below on each
compact neighborhood of so that if exists,

Zc+ zc+

for some constant c depending on ' and .


If exists for in D, define by

G(4, ) - (1.1)

The function GD on D x D is called the Green function of D, and is


called the Green function of D with pole . The latter function is.positive
superharmonic, harmonic on D - with 0.
It will be shown in Section 7 that D has a Green function if and only if
D is Greenian. All we have proved so far is that D has a Green function if
N > 2 or if N = 2 and if D is bounded. If N = 2 and if D is not everywhere
dense in IB2, the Green function GD exists because if E D and if i; is an inner
point of RZ - D, the restriction to D of G(e;, ) has G(C, ) - c as a harmonic
minorant on D for sufficiently large c. If D is not connected, it follows that
each open component of D has a Green function, and it is trivial to verify
that Go exists and that is defined for in the open component Do of
D by

GD('
- (GD 0
on Do
) 0 on D - D0.
86 I.VII. Green Functions

The existence of the Green function GB of an open set B implies the


existence of GD whenever D is a nonempty open subset of B and also implies
that Go < GB on D x D. Moreover G can be obtained from GB just as GD is
obtained from G:

To see this, note that if e D, then

GMDG(, )
(1.3)
= GQ,') - GMD[GB(, ) + GMBG(, )],

and by linearity of the GM operator (Section 111. 1)

GD(, ) = G(, ) - GMDGB(, ) - GMBG(,'), (1.4)

which yields (1.2). In particular, if B - D is polar then GD = G. on D x D


in view of the corresponding result for greatest subharmonic minorants in
Section V.5. Conversely, if D is an open subset of R" for which Go exists
and if B is an open superset of D with B - D polar, then GB also exists. In
fact, if E D, the function GMDG(, ) is locally upper bounded relative to
B and so (Section V.5) has a subharmonic extension to B, and this extension
is a subharmonic minorant of G(, ) on B; so GMBG(, ) exists.

Evaluation of GD by Solving the First Boundary Value Problem

If D is an open subset of R", bounded if N = 2, and if for in D there is a


harmonic function u(, ) on D with limit G(g, n) at each point n of BD (limit
0 at the point oo if N > 2 and D is unbounded), then an application of the
superharmonic function minimum theorem and the harmonic function
maximum-minimum theorem shows that u(, ) = GMDG(, ), so that

GD(, ) = G(, ') - u(, '). (1.5)

Thus the Green function of a smooth open set as defined in Section 1.8 and
evaluated for a ball in Section 11. 1 is the Green function in the present sense,
and GD = G when N > 2 and D = R". We shall discuss the solution of the
first boundary value problem for harmonic functions on Greenian subsets
of R" in Chapter VIII. Let u(, ) be the solution in the generalized sense of
Chapter VIII, that is, the "PWB" (Perron-Wiener-Brelot) solution, of this
first boundary value problem on a Greenian set D for the boundary function
2. Extremal Property of GI) 87

ff = 91 D; if D is unbounded, define f4(oo) as - oo or 0 according as


N = 2 or N > 2. According to Theorem VIII.18, the evaluation (1.5) is
correct with this interpretation of u if N > 2 or if N = 2 and D is bounded.
If N = 2 and D is unbounded, (1.5) is correct (Theorem VIII. 19) if and only
if D does not certain too much of a neighborhood of the point oo in a sense
made precise in the statement of Theorem VIII. 19.

2. Extremal Property of GD
Theorem. Let D be an open subset of RN and let be a point of D. If GD exists,
then

inf {u z 0: u = G(i, ) + h with h defined and harmonic on D}


= inf {u z 0: u = h with h defined and
superharmonic on D} (2.1)

= inf {u z 0: u superharmonic on D, u z
on some neighborhood of }
u(g)
= inf {u Z 0: u superharmonic on D, lim inf z 1 I.
-c q)

In the following proof we denote the jth class on the right in (2.1) by r,.
These classes are not empty if Go exists, but if GD does not exist, these classes
are empty because they all involve the existence of positive nonconstant
superharmonic functions on D, and we shall see in Section 7 that Go exists
if and only if D is Greenian.

Proof for r2. If GD exists and if uE 1-2, then -h 5 on D; so -h 5


and therefore S u. On the other hand, r2.

Proof that r, v r3 u r4 c r2. If GD exists and if u e r4, let p be the Riesz


measure associated with u. If c < 1, the function u - is positive and
superharmonic on an open deleted neighborhood B of . Hence (Theorem
V.5) u has a superharmonic extension to B v { 1; sop { } z c, and therefore
p { } z l ; so u e r2 . Thus r4 c r2 . The inclusions r, c r2 and r3 c r4 are
trivial. o

Proof for r, . r3, r4. If u e IF, v r3 v r4, then we have just shown that u C- 1-2.,
and we have already seen it follows that u z Conversely, e
r,nr2nr4and (1 every e>0.
88 I.VII. Green Functions

3. Boundedness Properties of GD
Theorem. Let D be an open subset of R" with a Green function GD, and let
be a point of D.
(a) If B is an open neighborhood of s in D, relatively compact in D, and if
v is a positive superharmonic function defined on an open superset
of D - B, with v z GD(, ) on a neighborhood of v
on D - B.
(b) RBD,t , = R"Die = GD(i;, ) (reductions relative to D) whenever B is

a neighborhood of
(c) If v is a strictly positive superharmonic function on D, then outside
each neighborhood of , GD(tf , ) < const v. In particular (v a 1),
GD(,, ) is bounded outside each neighborhood of , and if , is a point
in the same open connected component of D as , then outside each
neighborhood of , GD(, ) 5 const GD(,, ).
(d) Let B, be a compact subset of D and let BZ be D less a neighborhood
of B, . Then GD is bounded on Bt X B2.
Proof of (a). Define

`GD(, ) on B,
v, =
Sl GD(, ) n v on D - B.

The function v, is superharmonic on D, and G(1;,)-v,, defined as 0


at , is a subharmonic minorant of G(l:, ) and is therefore majorized by
GMDG(, ); that is, v, Z GD(, ) on D - B.
Proof of (b). The function RB is a positive superharmonic minorant
of GD(t, ) on D, equal to GD(, ) on a neighborhood of 4, and therefore by
(a) must majorize GD(, ) outside this neighborhood. Hence these two func-
tions are identical. Since the unsmoothed reduction lies between GD(, ) and
the smoothed reduction, the three functions are identical on D.
Proof of (c) and (d). In (c) we can suppose that the neighborhood in
question is open and relatively compact in D. Then if v is a strictly positive
superharmonic function on D, the function cv majorizes GD(, ) on a
neighborhood of 8B for sufficiently large c; so (a) can be applied to yield
(c). The particular cases of (c) follow trivially, and (d) follows from the first
of these cases by an application of the Heine-Borel theorem.

The Set D.: {ri: GD(, ry) > a)


(Note. S is fixed throughout the discussion.) If a > 0, the set D. is an open
subset of D containing . No open connected component of D. not containing
3. Boundedness Properties of Go 89

is relatively compact in D. but D,, is itself connected and relatively compact


in D if a is sufficiently large. In fact, if B is a relatively compact in D open
connected component of D. not containing , then the function is
harmonic on B, with boundary limit a at every boundary point of B, and so
must be identically a on B, an impossibility. On the other hand, if /3 is the
supremum of on the complement of a compact neighborhood of
in D and if a > /3, then the set D. is relatively compact in the interior of this
neighborhood and so also relatively compact in D and is connected by what
was just proved.

Observation (N = 2). According to Theorem 3, the function GD(, ) =


G(, ) - u(, ) is bounded outside each neighborhood of . It will now be
shown that if and 2 are in D and if c is an arbitrary constant, the minorants

GMD[G(,, ) A G(2, )], GMD[G(l;,, ) A c] (3.1)

exist and that the differences

G(1, ) A G(2,) - GMD[G(1,, ) A G(2, )],


(3.2)
G(1, ) A c - GMD[G(1, ) n c]

are bounded outside any neighborhood of , and 2, outside any neigh-


borhood of ,, respectively. These facts imply that if >q is a finite boundary
point of D, the minorants in (3.1) are bounded on the trace on D of a compact
neighborhood of q in 882, a fact needed later.
Since G(,, -) - G(2, ) is bounded outside any neighborhood of , and
2 and since GD(,, ) is bounded outside any neighborhood of i, the dif-
ference u(1,) - u(2, ) is bounded outside any neighborhood A of ,
and 2. Choose a compact neighborhood A of these two points, so that
1u( 1, ) - u(2, )1 5 b on D - A, where b depends on A, , , and b2. This
inequality must hold on all of D in view of the maximum theorem for sub-
harmonic functions. Then

G(1, ) A G(2, ) 2 GMD[u(1, ) A u(2, )] 2 u(,, ) - b; (3.3)

so the first minorant in (3.1) exists. The harmonic function u(,, ) is bounded
above by c, = sup,,D G(, , ) < + oc ; so

G(1, ) A c, 2 u(S,, ) A c, = u(, ,

and it follows that for any constant c,

G(c,,)Ac2G(y,,)Ac,-IC -c,I 2u(,,)-Ic-c,1; (3.4)


90 I.VII. Green Functions

so the second minorant in (3.1) exists. Finally, the first difference in (3.2)
is at most

b = GD(bl, ) + b, (3.5)

and similarly the second is at most const; so these differences


are bounded outside each neighborhood of ,, 2 or each neighborhood
of ,, respectively.

4. Further Properties of GD

As already remarked we shall prove in Section 7 that the open sets with
Green functions are the Greenian sets. In the following theorem the boundary
involved is the Euclidean boundary.

Theorem. If D is an open subset of Il" for which there is a Green function,


the Green function GD has the following properties:
(a) Go is continuous and symmetric on D x D (= + oo on the diagonal).
(b) For quasi every finite point S of 6D and also for C = oo if N > 2
and D is unbounded, lim,._t rl) = 0 for every in D.
(c) If e D, the function GD(c, ) has a positive extension GD to
(F8", uniquely characterized by the following properties:
(c 1) Go 0 on R' - D and at quasi every finite point of 8D.
(c2) GD is subharmonic on 68" - {ld}.
(c3) If c is a finite boundary point of D,

GD (, = lim sup q); (4.1)


Dan;

if N > 2 and if D is unbounded, lim,,-,o Go ri) = 0.

Observation (1). If D is connected, the finite points ( of 8D for which


GD C) = 0 on D, equivalently, by lower semicontinuity of subharmonic
functions the boundary points { for which lim,, GD(c, ti) = 0 for all in
D, will be seen in Section VIII.14 to be the finite regular boundary points
of D relative to the Dirichlet problem for harmonic functions on D.
Observation (2). We shall prove (b) as a consequence of (c). Alternatively
it is possible to prove (b) after the Dirichlet problem for harmonic functions
on D has been treated, by showing that quasi every finite boundary point
of D is regular and that the regular boundary points are the points at which
has limit 0. The extension G() can then be obtained by an
application of the extension Theorem V.5 to - on R' - {l;}.
4. Further Properties of Go 91

Proof of (a). We have seen in Section III.1 that if GD exists, then


GMDG(4, ) can be obtained as follows. A sequence B, of balls is chosen
with closures in D and with the property that each point of D has a neigh-
borhood which lies in B. for infinitely many values of n. If T aJ is the operator
II(1.8) but applied here to functions on R', and if for 4 e D we define

t8.... TB.G(4, ), u,( ,

then n e Z* } and n e Z') are decreasing sequences of super-


harmonic functions on R" and D, respectively. The limit of the second
sequence is We shall deal with the first sequence later. Since G is
continuous on D x D, the functions uo, ut, ... are successively continuous,
so u is upper semicontinuous. Let o be any point of D and choose B. in
such a way that o e Bo and that there is an open neighborhood B e Bo
of o which is a subset either of B. or of D - B for each n. A glance at the
formula for shows that for each point ry of D the function q)
is harmonic on B, and the same reasoning shows that ut ry), U2(',17), .. .
are harmonic on B. It follows that u(-, j1) is harmonic on B, and since u
is independent of the choice of B,, the function is harmonic on D,
a minorant of q). By definition of u it follows that q) 5 u(17, -), and
reversing arguments in this inequality yields the symmetry of u on D x D.
Hence GD is symmetric on D x D.
To prove continuity of GD, we prove that u is continuous. Let B' and
B" be balls with closures in D, and express u on B' x B" using the Poisson
integral on the boundary of each ball. If tLe balls either are the same or
have disjoint closures, this representation of u shows that u is continuous
on B' x B". Thus u is continuous on D x D. C3

Proof of (b) and (c). The sequence n e Z) is a decreasing sequence


of superharmonic functions on R", all equal to on R" - D. Moreover
the sequence is locally uniformly bounded below on R" - 8D. In fact
z u on R" - D. We now show that
this sequence is uniformly bounded below on a neighborhood of each
finite boundary point of D. This fact is trivial if N > 2 because u;, is then
positive and is trivial if N = 2 and D is bounded because a lower bound of
on a neighborhood of D is also a lower bound of on the
neighborhood. If N = 2 and D is unbounded, the function is bounded
on D in a neighborhood of OD (Theorem 3), say 2- c there; so

Z GD(,') Z G(:, ) - c

on D near OD, and G(4, ) on R2 - D; so the sequence {u;(4, ),


NE Z') is locally lower bounded on a neighborhood of each finite point
of 6D, as asserted. Thus {u;,(4, ), n e Z+ } is locally uniformly lower bounded
92 I.VII. Green Functions

on R'. It now follows from the Fundamental Convergence Theorem that the
function lim,,..m is equal quasi everywhere on R" to a super-
harmonic function u More specifically,

u(, ) = on D,
u
G(, ) on R' - D and at quasi every finite
point of OD.

The function Go u satisfies the conditions in (c) and


is obviously uniquely determined by these conditions; that is, (c) is true.
By the upper semicontinuity of subharmonic functions the positive function
GD is continuous at every zero; in particular, has limit 0 at
quasi every finite point of BD. If for some point , of D the function GD(.,, )
has limit 0 at a boundary point C, then has limit 0 at C for all in
the same open connected component of D as , because (Theorem 3) outside
each neighborhood of the inequality const is valid.
Since the number of open connected components of D is countable, (b)
is true. If N > 2 and D is unbounded, the function Go(C, ) has limit 0 at
ao because GD If C is a finite boundary point of D and N z 2,
then

GD ( , C) = lim sup GD PI) (4.2)


n_t

because is subharmonic on R" - This limit relation remains


correct if I tends to C along the nonzero set of GD(C, ) and also [see 11(6. 1)]
if q avoids the polar set of points I' of BD for which of) # 0; that is,
(4. l) is true. o

Observation (3). The method of proof of (b) can be applied to show


that the function differences in (3.2) also have limit 0 at quasi every finite
point of D.

5. The Potential GDp of a Measure p


Let D be an open subset of R" for which GD exists. If p is a measure on D,
the potential GDp is defined by

GDp() = JD (5.1)

The following facts have been derived for D special but are true in general,
and the proofs for special D are applicable. A potential GDp is either super-
5. The Potential GDµ of a Measure p 93

harmonic or identically + oo on each open connected component of D


and, if superharmonic, is harmonic on everyp null open subset of D. More-
over GMDGDP = 0 if GDp is superharmonic. In particular, GDµ is super-
harmonic if p(D) < + cc.
If p has compact support in D, Theorem 3(c) implies that GDµ is bounded
outside each neighborhood of this support. Moreover in this case GDµ has
limit 0 at every Euclidean boundary point C at which lim,r-{ q) = 0
for every because one can integrate to the limit in (5.1) when p has compact
support. It follows (Section 4) that when p has compact support in D, the
potential GDµ has limit 0 at quasi every finite Euclidean boundary point
of D and also the point oo if D is unbounded and N > 2. Since a measure
p on D can be chosen, supported by a countable dense subset of D, making
GDp a superharmonic function + oo on a dense subset of D, the hypothesis
of compact support for p was necessary in the preceding sentence.
If p has compact support in D, the potential GDp differs from Gp on
D by a harmonic function. In fact, if

GDp() =
JA
fA
q)p(drl). (5.2)

The second integral defines a harmonic function on D because the function


is Borel measurable, has the harmonic function average property, and is
finite valued on a dense set. Observe that what we have proved implies that
if A is a compact subset of R', if p is a measure on R" supported by A,
and if D, and D2 are open supersets of A whose Green functions exist, then
GD, p and GD2µ differ on D, n DZ by a harmonic function.
In order to complete the proof of the Riesz Decomposition Theorem
(Section IV.8) by proving it for open sets D for which GD exists-it will
be seen in Section 7 that these sets are the Greenian sets-we need only
follow the steps of the proof for D special. According to (5.2), the conclusion
of Theorem IV.7 can be stated in the form : u = GDPA + hA, where hA is
superharmonic on D and harmonic on A. This representation of u was the
basis for the proof of the Riesz Decomposition Theorem for special open
sets, and the continuation of that proof needs no change in the present
more general context. Note again that according to that theorem, a positive
superharmonic function u on D is a potential GDµ if and only if GMDU = 0.
In view of(5.2) the Gauss Integral Theorem (Section 1.6) and also Theorem
1.7 on the smoothness of the potentials Gp of measures given by densities
are valid for the more general potentials GDp.

Application to the Function D f-- GD

Let D2 be a Greenian subset of R", let D, be a nonempty relatively compact


open subset of D2, and let p be a measure on D2 with compact support in
94 I.VII. Green Functions

D, . Then GD p = Go, p + GMD, GD=µ on D, . The minorant is [Section III


Example (d) y the restriction to D, of OGD:BD,-D, (reduction relative to D2).
This smoothed reduction must be a potential
GD on D, , where u, is a measure supported by 8D, .

6. Increasing Sequences of Open Sets and the Corresponding


Green Function Sequences
Theorem. Let D. be a monotone sequence of open subsets of l with limit D,
and suppose that GD. exists for all n.
(a) If D. T D then the increasing sequence GD has limit GD if GD exists
and has limit + oo otherwise.
(b) If D. I D and if D is open and not empty then the decreasing sequence
GD has limit GD.

Proof of (a). To avoid trivia suppose that D is connected. If is a point


of D and if u then u on D,,, and
therefore (Theorem 11.3) the limit u(l;, ) of the sequence {u n > 0}
on D is either identically - oc or harmonic, and it is easy to check that in
the latter case GMDG(t, ). Conversely, if exists, the
sequence { u n Z 0} is locally bounded below; so the limit function
is harmonic on D. Since G(1;,) - GMDG(l;, ), assertion
(a) follows.

Proof of (b). Define GD = lim,,.. GD, on D x D. For in D the function


is harmonic on D - The inequality GDS GD 5 GDS implies
that is continuous with value + oo at and that the difference
is a positive harmonic function on D if defined suitably
at g. This difference is bounded [because GDS(, ) is bounded outside a
neighborhood of 1;] and has limit 0 at quasi every Euclidean boundary
point of D, including the point oo if N > 2 and D is unbounded [because
G0 has this boundary limit property]; so this difference vanishes
identically according to the extended maximum-minimum theorem for
harmonic functions in Section V.7.

7. The Existence of GD versus the Greenian Character of D


Theorem. An open subset D of l ' is Greenian if and only if GD exists.

Since all open nonempty subsets of lE for N > 2 and all bounded non-
empty open or not connected open subsets of R2 are both Greenian and
have Green functions, only unbounded open connected subsets of l 2 re-
9. Approximation Lemma 95

main to be considered. If a set has a Green function, the set is trivially


Greenian. Conversely, suppose that D is an open unbounded Greenian
subset of l2, so that there is a positive nonconstant superharmonic function
u on D, with associated nonnull Riesz measure p. Let D. be the part of D
in B(0, n) and let p be the projection of p on D. According to Theorem 6,
unless D has a Green function, lim, Go. = + oo, but this limit relation
is impossible because (Riesz decomposition)

on D.. (7.1)

8. From Special to Greenian Sets


Many theorems have been stated for Greenian sets D but proved only for
special sets. The justification for this awkward procedure is that the pro-
perties of GD make the proofs already given for special D valid whenever
D is Greenian, that is, whenever Go exists. Some of these extensions to
Greenian D have already been checked in Sections 4 and 5. The rest are
equally easy to check.

9. Approximation Lemma
Lemma. If D is a Greenian subset of R', if 0 is a finite-valued continuous
function on D, with compact support B, and if B, is a compact neighborhood
of B in D, there is a sequence u, - v, of differences of finite-valued continuous
potentials whose associated Riesz measures are supported by Bt, with u _
v,, on D - Bt and lima.,, (u - uniformly on D.

For n sufficiently large, say n -- no, the function Atm4 is defined on a


neighborhood of Bt and vanishes off Bt. For n this large define Atm4) = 0
on D - Bt where this function is not already defined. Then {Atm¢, n z no
is a sequence of infinitely differentiable functions converging uniformly
to 0 on D. Define u as the potential of the measure with density
-((DA1m4) A 0)/niv [((AA1m4) v 0)/ni ]. Then Au o - (u - v.) = h is a
function harmonic on D, and v - u = h on D - Bt. The inequalities
v z h and u z -h are valid on D - Bt and therefore (superharmonic
function minimum theorem) also on D; so h = 0 since GMDV = GMDU _
0. That is, Atm¢ = u - v,,, and the lemma follows.

Application to the Ordering of Measures

Suppose that p and v are finite measures on a Greenian set D, with the
property that whenever u and v are finite-valued positive continuous super-
harmonic functions on D, with u Z v, then
96 I.VII Green Functions

f (u - v) dp < f (u - v) dv. (9.1)


D D

We now show that it follows that µ S v. It is sufficient to show that if 0


is a positive continuous function on D with compact support then

j4)d <f0dv.
JD

According to the lemma, the function 0 can be approximated uniformly


by a sequence u, - v, of differences between continuous potentials, equal
outside a neighborhood of the support of 0. By hypothesis

`D { [u - innf (u - v.)] - du S fD { [u - innf (u - dv,

and this inequality yields (9.2) when n -+ oo.

10. The Function OJD_{;l as a Minimal Harmonic


Function
Let D be a connected Greenian subset of 88'. We now show that if c e D,
the restriction of C) to Do = D - {i;} is a minimal harmonic function
for Do. In fact let uo be a positive harmonic function on D0, majorized
there by C). We are to show that uo must be a constant multiple of
GD(., C) on D0. The function uo has a superharmonic extension u to D
(Theorem V.5) whose associated Riesz measure is supported by {r;}, u =
c h is a positive harmonic
function on D, majorized on Do by u and therefore majorized on D by
GD(., C). Hence h = 0 and the proof is complete.
If v is a positive superharmonic function on D, with associated Riesz
measure v, then

inf v r = v({C}).
D.

To prove this, let c be the infimum in question. Then v a C) on Do,


so the difference v - C) is positive and superharmonic on Do and
therefore has a superharmonic extension vi to D. Thus v = vt
on Do, and the equality is valid at C because two superharmonic functions
equal on a deleted neighborhood of a point are equal at the point. But then
v({C}) z c, and there must be equality by definition of c, so (10.1) is true.
Furthermore
10. The Function as a Minimal Harmonic Function 97

Vol)
liminf =v({t}). (10.2)
q-t GD(1,

To prove this, observe that according to (10.1) the limit inferior a in question
is at least v({i;}). If there is strict inequality and if a > /3 > v({C}), then
v> C) on a deleted neighborhood of C and therefore (Theorem 3) on D;
so the infimum in (10.1) is z fl. Hence v((C)) z /3, contrary to hypothesis.
We shall prove (Theorem XI.4) that the limit inferior in (10.2) is a limit
in the context of the fine topology discussed in Chapter XI.
Generalization. If C is a point of a compact polar subset A of D and if
now Do is D - A, then a trivial modification of the preceding discussion
shows that the restriction of to D - A is a minimal harmonic function
for Do and that (10.1) and (10.2) are true, with the understanding that the
infimum in (10.1) may be taken over either Do or D - and that ry in
(10.2) may tend to C on either Do or D.
Chapter VIII

The Dirichlet Problem for Relative Harmonic


Functions

1. Relative Harmonic, Superharmonic, and Subharmonic


Functions
The class of relative harmonic functions is suggested by the following trivial
remark. Let (D, 2) be a measurable space, and suppose that to each point
of D is assigned some set (perhaps empty) { a e 1} of probability
measures on D. Call a function generalized harmonic if it satisfies specified
smoothness conditions and if

v() = dq) = V)
JD

for in D and a in 1,. For example, if D is an open subset of RN, if for each
g the index at represents a ball B of center with closure in D, if 1, is the
class of all such balls, and if v) is the unweighted average of v on aB,
then the class of continuous functions on D satisfying (1.1) is the class of
harmonic functions on D. Going back to the general case, suppose that h
is a strictly positive generalized harmonic function and define by

/1 A) = h(7)Y°h( . (1.2)

Then is a probability measure, and if v is a generalized harmonic


function, the function u = v/h satisfies

u(D = I

u is a function satisfying (1.3), the function v = uh is generalized


harmonic. (We omit, here and in the following, possible side conditions on
the functions.) The functions u = v/h thus satisfy the same kind of averaging
condition as the generalized harmonic functions, but with pa replaced by pa.
Theorems on generalized harmonic functions v relative to the averaging {p,}
2. The PWB Method 99

correspond to theorems on functions v/h relative to the averaging system


{µa). This remark suggests the following definition.
Let D be an open subset of R' and let h be a strictly positive harmonic
function on D. A function u = v/h will be called h-harmonic, h-superharmonic,
or h-subharmonic if v is harmonic, superharmonic, or subharmonic on D,
respectively. This definition is of interest only if h is not identically constant,
so that D is always supposed Greenian in discussing these matters. If v = Goli
is a potential, the function v/h will be called an h potential. The functions of
classical potential theory relativized by a strictly positive harmonic h play
an essential role in the study of the case h =- 1. Many properties of the
relativized functions follow trivially from the case h = 1 or can be deduced
using the proofs for that case. For example, the proof of the superharmonic
function minimum theorem translates at once into a proof of the same result
for h-superharmonic functions. The operator TB is defined by rr(v/h) = (TBv/h.
The notations GM", LM", "R., and "oul" (for u a positive h-superharmonic
function) have the obvious interpretations.
Minimal harmonic functions were defined in Section I1.16, and minimal
h-harmonic functions are defined correspondingly. The positive h-harmonic
function v/h on D is a minimal h-harmonic function if and only if v is a
minimal harmonic function. In particular, if h is minimal harmonic, the
positive constant functions are minimal h-harmonic, and conversely.
A function u is h harmonic if and only if (in the usual inner product
notation) Au = - 2(grad u, grad h)/h. If u is a function in C(2)(D) the function
is h superharmonic if and only if Au 5 - 2(grad u, grad h)lh.

Relative Superharmonic (Harmonic) Functions and the Kelvin


Transformation

Let v [h] be a superharmonic [harmonic] function on a Greenian subset D


of R", and let 0 be an inversion of R" in a sphere of center to. Define v' = v(¢)
and h' = h(4)), and let v1 [h,] be the Kelvin transform of v [h]. Then v1/hl =
v'/h', and this function is an ht-superharmonic function on 4)(D - (to)) if
h > 0, ht-harmonic if h is harmonic. According to the definition of the
Kelvin transformation, the function vt/h1 is not the Kelvin transform of
v/h unless N = 2, in which case v1 = v' and h1 = h'.

2. The PWB Method

Let D be a Greenian subset of R", provided with a boundary 8D by a metric


compactification, and let It be a strictly positive harmonic function on D.
Let f be an extended real-valued function defined on 6D. The traditional
Dirichlet (first boundary value) problem in the context of h-harmonic
functions is to find a function u which is h-harmonic on D and has limit f(O
100 l.Vlll The Dinchlet Problem for Relative Harmonic Functions

at each boundary point C. (In the usual formulation of such a problem the
boundary function f is supposed finite valued and continuous.) Since a
function on OD that is the boundary limit of a function on D is continuous,
this Dirichlet problem cannot have a solution unless f is continuous. On the
other hand, the following example shows that boundary function finiteness
and continuity may not be enough to ensure the existence of a solution.

EXAMPLE (a). Let D be a ball less the center, let h =_ 1, and let dD be the
Euclidean boundary. Let f = 0 at the ball center and f = 1 on the rest of
the boundary. Then f is continuous on OD. The Dirichlet problem on D
for this boundary function has no solution because if u were a solution,
then 0 < u S 1, and (Section V.5) u would be harmonic on the ball if defined
as 0 at the center. But then u would attain its infimum 0 at the center and
therefore would vanish identically, contrary to the hypothesis of limit I at
the ball boundary.

The following example illustrates the influence of the choice of the


relativizing function h,

EXAMPLE (b). Dirichlet problem for a ball. Let D be a ball, and let 8D be
the Euclidean boundary. According to Theorem II.1, the Poisson integral
PI(D,f) solves the Dirichlet problem with h - I for any finite-valued con-
tinuous boundary function f If h is minimal harmonic however, that is
(Section IL 16), if h is a multiple of K(C, ) for some boundary point (, the
Dirichlet problem for h-harmonic functions with a specified bounded func-
tion f cannot have a solution unless f is identically constant. In fact a solu-
tion would be an h-harmonic function bounded by the bound off and so
would be identically constant because h is minimal harmonic.

The method of attacking the Dirichlet problem will be one devised by


Perron, developed by Wiener, perfected by Brelot, now called the PWB
method. In general the PWB method assigns a "PWB solution" to certain
"resolutive" boundary functions. A finite-valued continuous boundary func-
tionf for which the traditional Dirichlet problem has a solution of is resolu-
tive with PWB solution uf, but the PWB method goes further. In fact the
class of resolutive boundary functions is the class of integrable boundary
functions relative to a certain measure, and one task of the theory is to show
that the PWB solution for a resolutive boundary function has that boundary
function as a boundary limit function in some reasonable sense. In the
context of h-harmonic functions we shall introduce h into the notation and
denote by Hf the PWB' solution determined by an h-resolutive boundary
function f. Although the method itself provides a not unreasonable justifica-
tion for describing Hf as a solution corresponding to f, we shall show that
Hf has f as a boundary limit function along certain paths.
2. The PWB Method 101

The PWB method starts with a Greenian set D provided with a boundary
OD by a metric compactification and assigns to each extended real-valued
function f on BD an upper and a lower PWB" class of function on D. A
function v on aD is in the upper [lower] PWB" class if on each open con-
nected component of D this function is either identically +oo [ - oo] or
h-superharmonic [h-subharmonic], is bounded below [above], and satisfies
the inequality

lim inf v(q) [lim sup v(q) <f(o]


rc n-{

for every boundary point C. The functions in the upper PWB' class for f are
the negatives of the functions in the lower PWB' class for -f. An application
of the minimum theorem for h-superharmonic functions yields the fact that
every function in the upper PWB" class for f is a majorant of every function
in the lower class. If u and v are in the upper class, then u A v is also. Thus
the upper class is directed downward, and dually the lower class is directed
upward. If v is an h-superharmonic function in the upper class and if B
is a ball with B c D, then rBv is also in the upper class and is a minorant
of v. Thus the infimum Hf of the upper class is, in each such ball B, the
limit of a downward-directed family of h-harmonic functions unless each
member of the upper class is identically + oo in B. In each open connected
component of D the function Hf is therefore either the constant function
+ oo, the constant function - oo, or an h-harmonic function. The supremum
H f of the lower class must also have this property, and H f < fl f'. The
function Hf [Hf] is called the lower [upper] PWB" solution for f It is
immediate that Hip = R; /h = "R; for A a subset of 3D. If the upper and
lower solution are identical and h-harmonic, they are denoted by H, f will
be called h-resolutive, or simply resolutive when h =- 1, and H f* will be called
the PWB" solution for f. A boundary of a Greenian set D will be called
h-resolutive, or simply resolutive when h = 1, if every finite-valued continuous
boundary function is h-resolutive, equivalently according to Section 6 below,
if the bounded Borel measurable boundary functions are h-resolutive. If a
boundary is h-resolutive for every h, the boundary will be called universally
resolutive.

Disconnected versus Connected Greenian Sets

The set D has not been supposed connected in the preceding discussion.
If D is not connected, let Do, D, , ... be the open connected components
of D, and let 3Dj be the boundary of Dj relative to D u 3D, where aD has
been determined by a metric compactification of D. Then OD = Uo 3Dj.
If f is a function on aD let fj be the restriction off to 3Dj. A function u on
D is in the upper [lower] PWBh class for the boundary function f if and
102 I.VIII. The Dirichlet Problem for Relative Harmonic Functions

only if for each j the restriction of u to Dj is in the upper [lower] PWB"


class on Dj for the boundary function fj. Thus the restriction to Dj of fl f"
[H f ] is the upper [lower] PWB" solution on Dj for the boundary function
f j, and f is an h-resolutive boundary function for D if and only if for all j
the boundary function f of Dj is an h-resolutive boundary function for Dj.
Let D,,, D,=, ... be the components of D on which Hf is not identically
+ oo J or j z 1, choose a point In (to be held fast below) in D,j, and choose
e > 0. Then there is a function u on D, in the upper PWB" class for f, such
thatforj >- 1,

Hf is harmonic on D.,,
5 (2.1)
ifHf ooon

(A dual assertion is true for lower PWB" solutions.) To see this, observe
that by definition of upper PWB" classes there is a function uj in the upper
PWB" class on D for f such that (2.1) is satisfied by uj. On each set D,, define
u = uj and define u = + oo on each remaining component of D.

Internal Resolutivity

It is trivial that the one-point boundary of a Greenian set D is universally


resolutive and that the PWB" solutions are the constant functions. At the
other extreme, a boundary of D will be called internally h-resolutive if every
bounded h-harmonic function is the PWB" solution of some boundary
function, and a boundary will be called universally internally resolutive if the
boundary is internally h-resolutive for every strictly positive harmonic h.
Observe that the characterization "h-resolutive" is not applied to any
proper subset of a boundary.

EXAMPLE (c) (The Classical Solution). If f is a finite-valued continuous


boundary function for which the traditional first boundary value problem
has a solution, that is, there is an h-harmonic function u on D with boundary
limit function f, then f is h-resolutive, and H' = u because u is in both upper
and lower PWB" classes; so u = Hf = Hf. Thus [see Example (b) above]
the Euclidean boundary of a ball is resolutive. This ball boundary will be
shown to be universally resolutive in Section 9 and will be shown to be
universally internally resolutive in Section IX.12.

h-Regularity of Boundary Points

A boundary point i; of D will be called h-regular, or regular when h = 1,


if whenever f is a finite-valued continuous boundary function
2. The PWB Method 103

lim HH(7) = urn Hf(ry) =f(0.


4

If every boundary point is h-regular, the boundary will be called h-regular,


or regular when h = 1. An h-regular boundary is h-resolutive because when-
ever f is a finite-valued continuous boundary function Hf = Hf in view of
the h-harmonic maximum-minimum theorem. For each such function f the
PWB" solution H f" for an h-regular boundary is the solution of the traditional
Dirichlet problem. Conversely, if this problem has a solution for every
finite-valued continuous f, the boundary is h-resolutive and h-regular.

Elementary Properties of Upper and Lower Solutions

It is trivial that fl! f = -Hg, that H f and H f increase with f, that H f = cH f


and H f = cff when c is a positive constant, and that Hf+. = Hf + c and
Hf+, = Hf + c when c is an arbitrary constant. Moreover, if e > 0 and if f
and g are finite-valued boundary functions whose upper and lower solutions
are finite-valued, and if If - g < e, then lH f - Ha : e and IHf` - H, 'l e
because

Hf - e=Hf, SHa <Hf+,=Hf +e


and the same inequalities are valid for the lower solutions. It follows that if
in addition f is h-resolutive,

Hf-e<Ha<HB<Hf+e,
and we conclude that if f is a uniformly convergent sequence of finite-valued
h-resolutive boundary functions with limit g, then g is h-resolutive, and the
sequence Hf is uniformly convergent with limit H.

The Set of h-Irregular Boundary Points

This set is an F. set, that is, a countable union of compact boundary subsets.
In fact for an arbitrary function 0 from D into A the boundary function
C'- urn sup{-{,ID 0() is upper semicontinuous, so iff is a bounded boundary
function, the boundary set

A(f, e) = { C BD: lim sup) f lim inf ze


( rc 4-c

is compact. If f is a sequence of boundary functions dense in C(D), the set


of irregular boundary points is U "so A (f., 2"), an F. set.
104 I.Vlll The Dinchlet Problem for Relative Harmonic Functions

The PWB Method and the Kelvin Transform

Let D be a Greenian subset of 68N provided with a boundary DD by a metric


compactification of D, and let 0 be an inversion of pN in a sphere with
center not in D. Then D' = 0(D) is Greenian. The D u aD metric induces
a metric on D', and the completion of D' in this metric is a metric homeo-
morph of D u aD and thereby defines M. Let h be a strictly positive
harmonic function on D with Kelvin transform h, on D'. A function f on
aD induces a corresponding function f, on aD'. If v/h is in the upper PWB"
class on D for f and if v, is the Kelvin transform of v, then v,/h, is in the
upper PWB"I class on D for f,, and the corresponding assertion is valid for
the lower classes. Thus the PWB method on D is easily translated into that
on D'. For example, f is h-resolutive on OD if and only if f, is h,-resolutive
on aD', and a point of OD is h-regular if and only if its image on OD' is h,
regular. Observe that if OD is the Euclidean boundary, then aD' = 0(aD)
is also the Euclidean boundary. It will be proved in Section 14 that in this
Euclidean boundary context a point c of aD is not only regular (that is,
h-regular with h = 1) if and only if is an h,-regular point of OD' (with
h, the Kelvin transform of the constant function 1) but that is regular if
and only if is regular, under the additional hypothesis that when N > 2
neither boundary point is the point oo.

3. Examples

EXAMPLE (a) (Euclidean Boundary). Let u be an h-superharmonic lower-


bounded function on the Greenian subset D of I{8N, and suppose that u has
a finite or infinite limit at every point of aD, thereby defining a continuous
boundary function f. Then GMDu exists, and if this minorant is bounded
above, it follows that f is h-resolutive with Hi = GMDu. In fact under these
hypotheses, u is in the upper and GMDu is in the lower PWB" class for f;
so GMDu 5 Hf 5 Hf < u, and therefore by definition of GMDu the first
three terms of this inequality are equal.

Application: Green Functions and the Dirichlet Problem

Suppose that h = 1 and that D is arbitrary if N > 2 but bounded if N = 2,


choose in D and define u = -),,O, f = with f(oo) = 0 if N > 2
and D is unbounded. Then f is bounded so GMDu is bounded (harmonic
function maximum-minimum theorem). We conclude that f is resolutive
with PWB solution and that Hf. This eval-
uation of the Green function in terms of a Dirichlet solution is also correct
if N = 2 and D is unbounded whenever D is sufficiently sparse near the
point oo. More precisely, as we shall show in Section 19, this evaluation is
3. Examples 105

correct if and only if oo is a regular boundary point of D, equivalently


(Section XI.12) if and only if D is not a deleted fine neighborhood of oo.

EXAMPLE (b) (Euclidean Boundary, h =- 1). Let u, and u2 be superharmonic


functions on the Greenian subset D of I8", and suppose that GMDu; = u;
exists. Furthermore suppose that u = u1 - u2 is well defined near OD and
has a finite limit at every point of 8D, determining a boundary function f.
Then f is bounded and continuous and we now show (i) that f is resolutive
with Hf = u1 - u2. If, in addition, u1 and u2 have superharmonic extensions
to a neighborhood of I' n D, if these extensions have finite limits at every
point of 8D, and if u; - u, is bounded outside some compact subset of D,
we shall show (ii) that lim,l..t Hf(q) = f(C) at quasi every finite boundary
point C of D and also at the point oo if N > 2 and D is unbounded. (These
conditions are convenient to verify but are more stringent than necessary.)
The function u1 - u'2 is superharmonic with limit inferior !f(C) at every
boundary point C of D. This difference is therefore lower bounded and is
in the upper PWB class for f; so u1 >- u2 + Hf, and therefore u1 >- u2 + Hf.
Interchanging u1 with u2 yields the inequality u2 z u1 - Hf and we conclude
that Hf = Hf = u, - u2, that is, (i) is true. In Section VI 1.4 the Fundamental
Convergence Theorem was used to show that the function GD(, ) = G(, )
- GMDG(, ) has limit 0 at quasi every finite point of 3D. Under the
hypotheses of (ii) the same reasoning shows that u, - u,' has limit 0 at quasi
every finite point of 8D and therefore that Hf has limit f(C) at quasi every
finite boundary point C. To prove that Hf has limit f(oo) at oo when N > 2
and D is unbounded, we apply a barrier argument, to be developed more
generally in Section 14. Define u() _ I12-" for in D, choose b > f(oo),
choose a neighborhood of oo so small that f 5 b in the neighborhood, and
then choose n so large that b + nu 2:f on 8D outside this neighborhood.
The restriction to D of the function b + nu is in the upper PWB class on D
for f and has limit bat oo. Hence lim sup4-. Hf() 5 b; so this limit superior
is at most f(oo). Since the same reasoning is applicable to -f, the function
Hf has limit f(oo) at oo, as was to be proved.
Special case (b'). Let D be bounded and let f be the restriction to 8D
of a polynomial u. Then u can be written as the difference between two
polynomials superharmonic on a heighborhood of D:

u(C) = [u(C) - c1C12] - (- cIC12) (3.1)

for sufficiently large c. The assertions (b)(i) and (b)(ii) are therefore appli-
cable to polynomial boundary functions when D is bounded.
Special case (b"). According to Section VII.3, the hypotheses of (b) (i) and
(b) (ii) are satisfied by u1= G(,, ) A c1jD for , in D and - oo < ci 5 + oo.
Special case (b"). If N > 2 or if N = 2 and D is bounded, the hypotheses
of (b)(i) and (b)(ii) are satisfied by u, = G(s, )JD' u2 0, for each point
in D.
106 I. VIII. The Dirichlet Problem for Relative Harmonic Functions

4. Continuous Boundary Functions on the Euclidean


Boundary (h = 1)

Theorem. The Euclidean boundary of a Greenian set D is resolutive. Moreover


quasi every finite point of this boundary is regular, and oo is a regular Euclidean
boundary point whenever D is unbounded and N > 2.

This theorem, together with the fact that the PWB method yields the
classical solution whenever there is one, justifies this method. In the follow-
ing, C(OD) denotes the class of finite-valued continuous functions on OD,
metrized by the supremum norm. In view of the fact (Section 2) that the
limit f of a uniformly convergent sequence f of finite-valued resolutive
boundary functions is resolutive and that Hf is uniformly convergent to
Hf on D, it is sufficient to prove that there is a countable dense subset
r: iga,aaI} of C(aD) with the property that for each index value a, the
function ga is resolutive, and H,, has limit ga(C) at quasi every finite boundary
point C, as well as at i; = oo when N > 2 and D is unbounded. In fact the
Euclidean boundary is then resolutive, and if A. is the exceptional polar
boundary subset for ga, the finite boundary points not in UaeiAa are
regular.
In view of Section 3, Example (b'), when D is bounded F can be taken
as any countable dense subset of C(aD) consisting of restrictions to aD of
polynomials. The following proof is applicable to both unbounded and
bounded sets D.
If N > 2, let t, be the class of positive superharmonic functions on Ol;"
which are finite valued and continuous with finite limit at oo. Define u(oo)
for u in Ti as the limit of u at oo. Then IF, contains the positive constant
functions and u A v, and au + by are in IF, if u and v are and if a and b are
positive constants. Let F. be the class of differences u, - u2 with ui in 1-1,
so that 1'2 is a vector lattice in the order determined by pointwise inequality.
The set 173 of restrictions to OD of the members of F2 is a vector lattice of
finite continuous boundary functions, which contains the constant functions
and separates DD because c A -)Ian (defined as 0 at oo) is in 1-3 for an
arbitrary positive constant c and an arbitrary point . The set 173 is dense
in C(aD) (Stone-Weierstrass theorem), and in view of Section 3, Example
(b) any countable subset of l'3 dense in C(OD) can serve as the desired set F.
If N = 2, let o be a point of R' and define Ti as the class of functions u
satisfying the following two conditions:
(a) u is a finite-valued continuous superharmonic function on R2.
(b) There is a strictly positive constant a and a constant fi, both depend-
ing on u, such that u is identically fi outside some bounded
set which may depend on u.
4. Continuous Boundary Functions on the Euclidean Boundary (h a I) 107

When u is in r,, the function cu is also in r, when c is a strictly positive


constant, and u + c is in r, when c is an arbitrary constant. When u, and
u2 are in r, , the functions u, + u2 and u, n u2 are in r,. Let r2 be the class
of differences u, - u2 for which u, and u2 are in r, and have the same
multiplier a in (b). Then u, - u2 is constant outside some bounded set.
Define u, - u2 at oo as this constant value. The set r2 is a lattice in the
order determined by pointwise inequality. If u is in this lattice, cu and u + c
are also in the lattice, for every constant c. If I'3 is the class of restrictions
to aD of the members of r2, T3 is a lattice of finite continuous boundary
functions which contains cf and f + c with f Moreover r3 separates aD.
To see this, let , and 2 be distinct points of aD, and let B be a ball containing
o and either , or t;2, say ,, but not the other. The boundary function

f= IA 18

is then in r3. If f has the same value at t, as at 2 decrease the radius of B


to increase f at , but not at b2. Thus r3 separates aD and therefore (Stone-
Weierstrass theorem) is dense in C(aD). In view of Section 3 Example (b)
any countable subset of r3 dense in C(0D) will serve as the desired set r.

Partial Generalization to PWB" Solutions

Suppose that D is Greenian, that h is a strictly positive harmonic function


on D and that B is an open relatively compact subset of D. If f is a finite-
valued continuous function on DB, a function u is in the upper [lower] PWB"
class on B for the boundary function f if and only if the function uh on B
is in the upper [lower] PWB class on B for the boundary function Jh18B
It follows that f is h-resolutive, with Hi = Hf"/h, and that a boundary point
of B is h-regular if and only if the point is regular. Thus quasi every boundary
point of B is h-regular.
This partial generalization of Theorem 4 will be completed in Section 8
where it will be shown that if D is given an h-resolutive boundary by a
metric compactification, if B is an open subset of D, and if M is the boundary
of B relative to D v aD, then aB is h-resolutive. This fact can be deduced now
when h = I and OD is the one-point boundary provided by the Alexandrov
compactification of D. In fact, in this specialization if f is a finite continuous
function on tB, let f be the function on the Euclidean boundary a'B of B
defined as! on OD n 8'B and (if B is not relatively compact in D) as f(aB n aD)
on the rest of a'B. The function f is finite and continuous and therefore
resolutive on B for a'B. Moreover the upper and lower PWB classes on B
for f on a'B are the same as those classes for f on B. Hencef is resolutive,
as was to be proved.
108 I. Vill. The Dirichlet Problem for Relative Harmonic Functions

5. h-Harmonic Measure Null Sets


Let D be a Greenian subset of R' provided with a boundary by a metric
compactification, and let A be a boundary subset. If the upper solution
H;A "R; = R" /h = RA/h) vanishes identically on D, the set A will be
called an h-harmonic measure null set, or harmonic measure null set if h =_ 1.
In the following discussion D,, D2, ... are the open connected compo-
nents of D and j is a point of Dj, held fast throughout. The proofs are given
for infinitely many components and the simplifications to be made when
there are only finitely many are left to the reader.
(a) A countable union A = Uo A. of h-harmonic measure null sets is
h-harmonic measure null.
If e > 0 and if (Section 2) u" is an h-superharmonic function in the upper
PWB" class on D for the boundary function 'Aft, with e2-'-' for
ail j, then the function v = J:', u" is in the upper PWB" class for the boundary
function IA, and H;A = 0 because 5 v(lj) < e for all j.
(b) A boundary subset A is h-harmonic measure null if and only if there
is a positive h-superharmonic function u on D with limit + oo at every point
of A.
If A is h-harmonic measure null, define A. = A and observe that then the
positive h-superharmonic function v in (a) has limit +oo at every point of A.
Conversely, if a function u as described in (b) exists, then for every e > 0
the function ru is in the upper PWB" class for the boundary function IA;
so H A = 0 except possibly on the polar set of infinities of u, and so H; = 0
on D.
(c) If f is a positive boundary function and if Hf = 0, then the set
If > 0} is h-harmonic measure null.
If n >_ I and if f is the indicator function on BD of the set { f > 1/n},
then

H"n"= Hn"";
0=Nf z

so the set { f > I /n} is h-harmonic measure null. Hence Ul f > 1 /n} _
{ f > 0) is h-harmonic measure null.
(d) If f is a boundary function for which H f < + oo, then the set
A = { f = + oo } is h-harmonic measure null.
Choose (Section 2) an h-superharmonic function u on D in the upper
PWB" class for the boundary function f. The function u is positive and has
limit + oo at every point of A so A is h-harmonic measure null according
to (b).
(e) Euclidean boundary, h = 1. A polar subset A of OD is a harmonic
measure null set.
If N > 2, let u be a positive superharmonic function on R", identically
+ oo on A. Then v = ulD has limit + oo at every point of A ; so according to
S. h-Harmonic Measure Null Sets 109

(b), the set A is h-harmonic measure null. If N = 2 and if C is a finite boundary


point of D, choose a ball B of center r; so small that 682 - (D u B) is not
polar, that is, so that D' = D v B is Greenian. Then if A n B is not empty,
there is a positive superharmonic function u on D', identically + 00 on
A n B; so UID has limit + oo at every point of A n B. Thus according to (b),
A is locally harmonic measure null, and it follows that A is a countable
union of harmonic measure null sets and is therefore itself harmonic measure
null.
The converse of (e) is false : a harmonic measure null subset of a Euclidean
boundary aD need not be polar. For example, if D is a ball in R' it will be
seen in Section 9 that the harmonic measure null subsets of the Euclidean
boundary are the IN_, null boundary subsets, and it is not difficult to find
examples of IN_, null subsets of a sphere which are not polar. The following
is however a near converse to (e).
(f) Let D be an arbitrary nonempty open subset of R', and let A be a
nonpolar proper subset of D, closed relative to D. Provide D with a boundary
by a metric compactification, and let a(D - A) be the boundary of the
Greenian set D - A in this compactification. Then the set A n a(D - A) is
not harmonic measure null relative to D - A. In fact, if A n a(D - A) is
harmonic measure null, then by (b) there is a positive superharmonic
function u on D - A with limit + oo at each point of A n a(D - A). If u is
extended to D by defining u = + co on A, the resulting function is super-
harmonic on D with value + co on A ; so A is polar, contrary to hypothesis.

h-Harmonic Measure Null Sets and the Kelvin Transformation

We use here the notation of the discussion in Section 2 of the PWB method
and the Kelvin transformation. It is clear from that discussion that an
h-harmonic measure null subset of aD is transformed under an inversion
into an h,-harmonic measure null subset of OD'. We shall use the following
additional fact when h = 1. If OD and aD' are Euclidean boundaries and if
A is a harmonic measure null subset of aD, then ¢(A) is a harmonic measure
null subset of OD', under the additional hypothesis when N > 2 that all
points of cb(A) are finite. This fact follows from the criterion for h-harmonic
measure null sets in (b) because if v is a positive superharmonic function
on D with limit + oo at every point of A, then the Kelvin transform of v
is a positive superharmonic function on D' with limit + oo at every finite
point of ¢(A), every point of O(A) if N = 2.

EXAMPLE (a) (Euclidean Boundary, h =- 1). If N Z 2 and if C is a finite


boundary point of D, the singleton (C) is polar and therefore is harmonic
measure null. If n = 2, the singleton { oo } is harmonic measure null for every
unbounded Greenian set D because the point oo can be made finite by an
inversion of the plane relative to a circle centered at another boundary point
110 I.VIII. The Dinchlet Problem for Relative Harmonic Functions

and we have just seen that such a transformation of the plane preserves the
harmonic measure null property. If N> 2, the singleton {oo} may not be
a harmonic measure null set for an unbounded Greenian set. For example,
if D = R' and if u is in the upper PWB class on RN for the boundary function
11,,), that is, if u is superharmonic and lower bounded on RN, with inferior
limit z I at oo, then u >- 1 by the superharmonic function minimum theorem.

EXAMPLE (b) (Euclidean Boundary). A polar subset of 8D need not be an


h-harmonic measure null set for all h. For example, let D = B(0, b), and let
h be a minimal harmonic function on D corresponding to the boundary
point C (Section 1I.16), say

I 1z

h()= 6, - SIN
IS -

Let g be the indicator function of A = OD - Since the function e/h is in


the upper PWBh class on D for g whenever c > 0, the set A is h-harmonic
measure null; so is not, because aD is not an h-harmonic measure null
set.

6. Properties of PWB' Solutions


The following properties of PWBh solution will be needed in addition to
those derived in Section 2. Proofs are given in the next section. The functions
f and g below are boundary functions.
(a) If f = g up to an h-harmonic measure null set, then H f = He and
Hh=Hh
(b) Hf < + oo if and only if H f n < + oo.
(c) If Hi < + oo, if h." < + oo, and if f + g is defined arbitrarily on the
h-harmonic measure null set, where f or g is - oo and the other is
+oo, then Hf+g :!g Hf + Ho.
(d) The class of h-resolutive boundary functions contains the constant
functions and is a vector lattice (that is, the class is linear and
contains f v g and f A g when it contains f and g). Moreover, if
f and g are h-resolutive, then

H.hf+sB = aHJ + PH,, Hj,, 6 = LM ,(H f v Hg),


h h h
A Hg).

(e) If f, is a monotone increasing sequence of boundary functions with


limit f and if fl fo > - oo, then H! = lim,,..,p H H. Moreover H f = H j
if each function f. is h-resolutive.
7. Proofs for Section 6 111

(f) If f is an h-resolutive boundary function, there are Borel measurable


h-resolutive boundary functions f, and f2 such that f, 5 f 5 f2,
H f = Hf = H fl,, and f, =f = f2 up to an h-harmonic measure null
set. In particular, if A is a boundary subset with h-resolutive bound-
ary indicator function f = 1A, there are Borel boundary subsets A,
and A 2 such that the functions f, = 1,,, andf2 = 1A, are h-resolutive,
that A, c A c A 2, that H f, = H f' = H f,, and that A2 - A, is
h-harmonic measure null.
(g) If f is h-resolutive, if Do is an open subset of D, and if fo is defined
on (Do as f on OD, n 8D and as Hr' on D n BD,, then f0 is h-
resolutive for Do and the PWB' solution on Do forfo is the restriction
of H4 to Do.
(h) For every subset A of 8D,

H" = inf {Hi8: 8D B A, B open in aD},


(6.1)
H A= sup {H;8: B c A, B compact}.

For each point of D the set function A t--. H is a Choquet


capacity relative to the class of compact boundary subsets.
h-Resolutive Boundaries. Property (e) and its dual for decreasing sequences
imply that if 8D is h-resolutive, all bounded Borel measurable boundary
functions are h-resolutive. This result will be extended in Section 8.

7. Proofs for Section 6


Since the proofs do not depend on the choice of h they will be given for
h I to simplify notation. The point is that the assertions in Section 6 are
valid for the PWB method in a very general context.

Proof of (a). Let A be the harmonic measure null set in question. It is enough
to prove the second equality, and (by symmetry) it is even enough to prove
that Hf 5 H. If v is in the upper PWB class for g and if v, is a positive
superharmonic function on D with limit + oo at every point of A, then
v + ev, is in the upper PWB class for f whenever e > 0; so Hf 5 Hg quasi
everywhere on D and therefore everywhere on D. o

Proof of (b). If Hf < + oo, there is a lower-bounded superharmonic function


v on D. say v z c, in the upper PWB class for f. Then v + Icl is in the upper
PWB class for f v 0; so Hf < +co. Conversely, if +oo, then
Hf<Hf,,o<+oc. o
Proof of (c). Redefine f and g to be finite on the harmonic measure null
set on which either function has the value + oo. Then f + g is well defined,
1 12 1. VIII. The Dirichiet Problem for Relative Harmonic Functions

and if u [v] is in the upper PWB class for f [g], the function u + v is in
the upper PWB class for f + g; so Hf,9 < u + v and (c) follows. o

Proof of (d). The fact that the class of resolutive boundary functions contains
the finite constant functions, that the class is linear, and that the map f - Hf
is linear on this class follows from the PWB properties listed in Section 2
together with (c) and its dual for lower solutions. We need not discuss the
resolutivity of both f v g and f n g because f n g = -[(-f) v(-g)].
Moreover in treating f v g we can choose g = 0 because

fvg=[(f-g)v0]+g.
We assume therefore that f is resolutive and prove that f v 0 is resolutive,
with Hf o = LMD(HI v 0). Choose a point g; in each open connected
component of D, and choose v in the upper PWB class for j with

vn(bJ) G 2-"

for all j. The series Fk°D (v - Hf) of positive superharmonic functions has
a superharmonic sum because the sum is finite at a point of each open
connected component of D. The positive function

LMD(Hf v 0) + Z (v - Hf)
k

is therefore superharmonic. Moreover this function majorizes vk and there-


fore is in the upper PWB class for f v 0. When k - co, it follows that
LMD(Hf v 0) quasi everywhere on D and therefore everywhere on D.
On the other hand, Hf and 0; so z LMD(Hf v 0)
and (d) follows. o

Proof of (e). Let f be a monotone increasing sequence of boundary func-


tions with limit f. The first assertion in (e) is trivially true in any open con-
nected component of D on which some Hf. is identically + co. We therefore
decrease D if necessary and assume that Hfn < + oo on D for all n. Choose
a point k in each open connected component of D, and choose um in the
upper PWB class for fm with um((k) < 2-m for all k. The series
Lo (um - Hfm) of positive superharmonic functions has a superharmonic
sum because the sum is finite at a point in each open connected component
of D. The superharmonic function v defined by
m
V. = lim Hfm + Y_ (um - HIm)
M-00 R+1
7. Proofs for Section 6 113

is a majorant of ur, form > n, and v is therefore in the upper PWB class for
f, with

lim Hf S 5 lira lim H1


11W "-gym R-OD

Thus the first assertion of (e) is true at each point bk and therefore everywhere
on D in view of the trivial inequality Hf z lim,, H. Under the hypotheses
of the second assertion,

Hf > Hf z lim Hf. = Hf

so there is equality throught, and the proof is complete.

Proof of (f). Suppose first that f is an arbitrary boundary function. Choose


a point Sk in each open connected component of D, and let v, be a decreasing
sequence of members of the upper PWB class for f with HM.-. V,(4) =
for all k. There is such a sequence v, because the upper PWB class
for f is directed downward. Define gq(C) = lim info..., v (ry) for i e BD. Then
g, is a decreasing sequence of lower semicontinuous boundary functions
with Borel measurable limit function f2 z f; so H1 2 Hf, and there is
equality on the set ,. Since an upper PWB solution for f is either harmonic
or identically + oo or identically - o0 on each connected open component
of D. we conclude that Hf, = Hf. A dual argument (or apply this result to
-,f) yields a Borel measureable boundary function f, with f, 5 f and
Hf. = Hf. In particular, if f is PWB resolutive,

Hf=Hf, SHf. SHfSHf, SHfz=Hf;


so f, and f2 are resolutive, and the equation Hf2 - Hf, = 0 together with
the inequality f, 5 f 5 f2 implies (Section 5) that f, =f = f2 up to a harmonic
measure null set, as was to be proved. If f = 1,, is the resolutive indicator
function of a set A, we can replace f2 [ f,] by the indicator function of the
set A 2 = { f2 = 1) [A, = (f, = 1) ] to obtain the second part of (f).

Proof of (g). If v is in the upper PWB class on D for f, the restriction of v


to Do is in the upper PWB class on Do for fo. Hence the upper PWB solution
Offfo on Do for fo satisfies the inequality oHfa 5 Hf on Do. Similarly, oHfo Z
Hf on Do so (g) is true.

Proof of (h). Since H,A = I - H,aD_A, only the first equality in (6.1) need
be proved. A function v on D is in the upper PWB class for lA if and only if
on each open connected component of D the function v is either identically
1 14 I VI11 The Dirichlet Problem for Relative Harmonic Functions

+ oc or positive superharmonic with lim inf,,...z v(i) z I when e A. For such


a function v, 0 < a < 1 implies that lim in some neighbor-
hood B of A relative to aD, so that v/a is in the upper PWB class for 1,
and H, a < v/a. Since a can be chosen arbitrarily close to 1, the infimum in
(6.1) is at most fit A* The inequality in the other direction is trivial. The set
function -. H,A( ) = R; is a Choquet capacity of boundary subsets
relative to the class of compact boundary subsets according to Section
VI.3(k). [In the present context we have already proved all the desired
properties of this set function except that if A is a decreasing sequence of
compact boundary subsets with intersection A, then H,A^ = H,A, and
this fact is elementary in view of the first equation in (6.1).]

8. h-Harmonic Measure

Let D be a Greenian subset of R', coupled with a boundary OD provided


by a metric compactification of D, and let h be a strictly positive harmonic
function on D. In view of Sections 6(d) and 6(e) the class of boundary subsets
whose indicator boundary functions are h-resolutive is a a algebra, and if
we define A) = H;A for each set A in this a algebra, the set function
µ"D(S. ) is a probability measure for each point of D. The sets of this a
algebra will be called the uD measurable sets, pD will be called h-harmonic
measure, and µ'W, ) will be called h-harmonic measure relative to . The h
will be omitted when h = 1. Observe that a boundary subset A is an h-
harmonic measure null set in the sense of Section 5 if and only if the set is
pD measurable and A) -= 0, equivalently, if and only if the function
A) has a zero in each open connected component of D. According to
Section 6(f), a boundary subset A is pD measurable if and only if there are
Bore] µD measurable boundary subsets A, and A2 such that A, c A c A2
and A2 - A, is µ"D null. Thus for i; in D, 4(S, ) is the completion of the
restriction of to the µD measurable Borel boundary subsets if D is
connected. If f is a boundary function, we write for JDf(,)4(', dn)
when this integral is defined. Under this convention A) can also be
written If f is defined on a superset of aD, the notation 4D(-,f) is
to be interpreted as fj.D). It is trivial that for f a linear combination of
indicator functions of µ"D measurable sets, the function f) is h-harmonic.
Since a bounded uD measurable boundary function is the limit of a uniformly
convergent sequence of such linear combinations, the function f) is
h-harmonic if f is bounded and µD measurable. If f is a pD measurable
boundary function,

A(',IfI)= lim11D(',I fI n n) < +x,


N--W
8. h-Harmonic Measure 115

and (Section 11.3) on each open connected component of D the limit is either
identically + oo or is h-harmonic. If this limit is h-harmonic on D, the
function f will be called µD integrable; the class of such functions f will be
denoted by L'(AD). Then feL'(4D) if and only if f v 0 and f n 0 are in this
class and, if so,

.U'(-,f) = 4D(-,f V 0) + 4D(-,f A 0).

Let Do, D,, ... be the open connected components of D, and let aDk be
the boundary of Dk relative to D u aD. Let h be a strictly positive super-
harmonic function on D, and denote by hk the restriction of h to Dk. Let f
be a function on aD, and denote by f the restriction off to aDk. Then f is
an h-resolutive boundary function for D if and only iff is h-resolutive for
Dk for all k; f is µD measurable if and only iffk is µ"Dk measurable for all k;
f e L' (,uD) if and only if f e L' (p'k for all k. When A is a µD measurable
subset of aD, the function p" k(-, A r aDk) is the restriction of 4(-, A) to D.

Theorem. If D is a Greenian subset of R" coupled with a boundary aD provided


by a metric compactification of D and if h is a strictly positive harmonic
function on D, then a boundary function f is h-resolutive if and only if f e L' (µ4D),
equivalently, if and only if f is µD measurable with both H f and H f finite valued,
and then

H1 = µD(-,f ). (8.1)

Proof That h-Resolutive Boundary Functions Are µD Measurable

Let Co be the class of continuous functions from R into R with limit 0 at


± oo and let f be a finite-valued h-resolutive boundary function. The class
of functions 0 in Co for which ¢(f) is h-resolutive is a vector lattice which
is closed under uniform convergence and which separates points of R
because the class includes the function xf- (1 - Ix - n j) v 0 for n e Z'.
Hence (Weierstrass approximation theorem) the class is Co. Since the class
of functions from R into R for which i(f) is h-resolutive is a class closed
under bounded monotone convergence [Section 6(e)] and includes Co, this
class includes every bounded Borel measurable function. When ' is the
indicator function of an interval in R, we find that f is µp measurable.
Finally, if f is an arbitrary h-resolutive boundary function, redefine f as 0 at
its infinities to find an h-resolutive boundary function differing from f on a
set of h-harmonic measure 0 and thereby to conclude that f is µD measurable.

Proof of Theorem 8. If f is the indicator function of a µ4D measurable boundary


subset, then f is h-resolutive, and (8.1) becomes the definition of µ4D. More
generally (8.1) is therefore true if f is a finite linear combination of indicator
116 1. VIII. The Dirichlet Problem for Relative Harmonic Functions

functions of uD measurable boundary subsets, that is, if f is an h-resolutive


boundary function taking on only finitely many values, all finite. Denote
this class of boundary functions by IF. Apply the dominated convergence
theorem and Section 6(e) to show that if f is a positive and h-resolutive
[µD measurable and integrable] boundary function, then f, as the limit up
to an h-harmonic measure null set of an increasing sequence of positive
functions in r, is in L`(,uD) [is h-resolutive], and (8.1) is true. This result
applied to f v 0 and (-f) v 0 shows that an arbitrary boundary function
f is h-resolutive if and only if it is in L' (,uD) and that then (8.1) is true. Finally
we prove that if the boundary function f is µD measurable, with Hf and Hf,
finite valued, then f e L' (µD). According to Section 6(b), Hf o < + oo ; so in
view of Section 6(e) and what we have just proved,

µD('>fv0)=JimAD(',(fv0)nn)=limHf o) ,,=Hfvo
Hence f v 0eL'(p'D). Similarly f A 0eLI(jD); so feL1(µD), as was to be
proved.

h-Regularity in Terms of h-Harmonic Measure

If aD is h-resolutive, the condition for h-regularity of a boundary point


becomes

limµD(C, ') = 8

(vague convergence of measures on aD), where 8t is the probability measure


supported by {C); equivalently, t is h-regular if and only if

lim A) = 1
t--4

whenever A is the trace on aD of a neighborhood of C.

EXAMPLE (a) (Relation between uB and µg for B Relatively Compact in D).


If D is Greenian, if h is strictly positive and harmonic on D, and if B is an
open relatively compact subset of D, it was shown in Section 4 that aB is
h-resolutive with Hf = Hfh/h. It follows that

d7) =
µ8(i, a B). (8.2)

EXAMPLE (b) (Extension of Example (a) to Non-relatively Compact B). Let


D be Greenian, provided with a boundary OD by a metric compactification,
and let B be an open subset of D, with boundary aB relative to D u aD.
Let h be a strictly positive harmonic function on D. Then the class of h-
8. h-Harmonic Measure 117

resolutive boundary functions on aB includes the indicator functions of the


Borel subsets of D n aB, and for such a subset A and point in B the value
A) does not depend on the choice of aD. Furthermore (8.2) is true for
q c- D n aB when pB is defined using the Euclidean boundary of B. To prove
these assertions, let A be a compact subset of D n aB, define! as the indicator
function of A on 8B, and define f' as the indicator function of A on the
Euclidean boundary of B. A function u is in the upper PWB' class on B for f
on aB if and only if uh is in the upper PWB class on B for f 'h on the Euclidean
boundary. In view of the fact that Euclidean boundaries are resolutive
(Theorem 4), it follows that Hf = Hf"lh, where the notation refers to
Dirichlet solutions on B. If e > 0 and if uh is in the lower PWB class on B
forf'h on the Euclidean boundary, then u - s/h is in the lower PWB" class
on B for f on 8B. It follows that Hf z Hf,,/h. This conclusion implies the
truth of the assertions made above.
ExA.M1 a (b) (Continued). An assertion made in Section 4 can now be
strengthened and proved. Suppose in Example (b) that OD is h-resolutive.
Then aB is h-resolutive for B, and if A is a Borel subset of OD,

A) = A n eB) + f µD(17, d17) ( a B). (8.3)


nae

In view of Example (b) and Section 6(h) it is sufficient to prove that for A
a compact subset of aD the function 1,,,,W on 8B is an h-resolutive boundary
function for B and that (8.3) is true. Observe that if v2 [u2] is in the upper
[lower] PWB" class on D for the boundary function 1 A on OD and if v, [u, ]
is in the lower [upper] PWB" class on B for the boundary function defined as
A) on D n 8B and as 0 elsewhere on 6B, the difference v2 - v, [u2 - u
on B is in the upper [lower] PWB" class for the boundary function IA,,',
on B. It follows that 1 A,,,1B is an h-resolutive boundary function on aB and
that (8.3) is true.

See Section 3.11.3(d) for the simple (after the necessary foundations have
been laid) probabilistic derivation of (8.3).

Application to the h-Superharmonic Function Inequality

Let u be an h-superharmonic function on D, and let B be an open relatively


compact subset of D. Then the restriction of u to B is in the upper PWBh
class for the boundary function uIaB; so u >- u). A more delicate result
is the following. Suppose that u is superharmonic and lower bounded on
D, and define a lower semicontinuous boundary function f on aD by

f(o = lim inf u().


118 I.VIII. The Dirichlet Problem for Relative Harmonic Functions

Then if OD is h-resolutive, we show that f is h-resolutive and that

u > P'D'(-,f). (8.4)

In fact, for n e Z+ the boundary function f n n is h-resolutive [Section 6(d)],


and u is in the upper PWB" class for this boundary function; so

u z Hf n= PD(-, f A n)

and (8.4) follows. The function f is h-resolutive because it is in L1(p"D).

EXAMPLE (c). Let D, be a Greenian subset of I8", let D be an open subset


of D, ,and let be a point of D. We consider PWB" solutions on Do = D -
with 8Do the Euclidean boundary and h = GD Let f be the function
lies on OD, We show that f is a PWB" resolutive boundary function and
has the PWB" solution

on Do. (8.5)
GD, ( ,

In particular, if D = D, it follows that µ""D,,(-, { }) = I on Do. Let u be the


restriction to Do of the function on the right in (8.5). Then u is h-harmonic
and is in the upper PWB" class on Do for the boundary function f The
Green function has boundary limit 0 at quasi every finite point of
aD and at the point oo if N > 2 and D is unbounded (Theorem VII.4), and
therefore at quasi every finite point of aDo = aD v and at oo if N > 2
and D is unbounded. The polar exceptional set A is harmonic measure null
(Section 5), so according to Section 5, there is a positive subharmonic func-
tion w on D with limit + oo at every point of A. Hence if e > 0, the h-sub-
harmonic function

e(l + w)

on D. is bounded above because (Section l) GD < GD,, and this h-subhar-


monic function has limit superior :5f at every point of aDo. This function
is therefore in the lower PWB" class for f; so (e -. 0) the lower PWB" solution
for f is at least u. Hence u is the PWB" solution for f, as stated in (8.5).

9. h-Resolutive Boundaries
Section 6 implies that a boundary is h-resolutive if and only if the Bore]
boundary subsets have h-resolutive indicator functions, equivalently, if and
only if the compact boundary subsets have h-resolutive indicator functions.
9. h-Resolutive Boundaries 119

The following theorem gives a useful criterion for this h-resolutivity


condition.

Theorem. A boundary of a Greenian set is h-resolutive if and only if the set


function A -. R' /h is additive on the class of compact boundary subsets.

The equality HA = RA/h was pointed out in Section 2. If 8D is h-resolutive


A i- H, = A) for A compact. To complete the proof of the theorem,
it will be shown that conversely if there is the stated additivity, then the
boundary function 'A is h-resolutive whenever A is compact. Let B be a
compact subset of 8D - A. By hypothesis
"
Hh,A + HIB = H1A.e.
and when B increases to 8D - A through a sequence of compact sets, this
equation becomes, in view of Section 6(e),

H,A+ H 10D-A = 1

so that H A = H H. That is, IA is h-resolutive, as was to be proved.


Observation. For an arbitrary boundary the set function A F-. H (A
compact) is subadditive; so there is additivity for all l; in D if and only if
there is additivity for a point in each open connected component of D.

Application to Balls

The Euclidean boundary of a ball B is universally resolutive. In fact, if h is


a strictly positive harmonic function on B and if M" is the Riesz-Herglotz
measure for h, then (Section III.4)

"Ri A = h" = HiA = f K(C,


A

for every Borel boundary subset A. Hence the set function in Theorem 9 is
additive. so the Euclidean ball boundary is universally resolutive. Moreover
equation (9.1) implies that

)
PhW,dC) = K(C, (9.2)
h( )

so that Hf = PI(B, fdM")/h. In particular (h - 1), the class of harmonic


measure null boundary subsets is the class of IN_, null boundary subsets;
the class of resolutive boundary functions, that is, the class L' (µe), is the
class of IN_, measurable and integrable boundary functions; and Hf =
120 I. VIII The Dirichlet Problem for Relative Harmonic Functions

PI(B, f ). The class of PWB solutions Hf is thus [Theorem 11.14(b)] the class
D(p8_) which includes all bounded harmonic functions on B. Hence the
Euclidean ball boundary is internally resolutive. The generalization of
Theorem 11.14 to h-harmonic functions in Section IX.12 will make this
reasoning applicable to show that the Euclidean ball boundary is universally
internally resolutive. The following example exhibits h-resolutivity and
internal h-resolutivity in an extreme case. If C is a Euclidean ball boundary
point and if h = K(C, ), then (9.2) implies that 1. In this case
every boundary function f finite at a is h-resolutive, and Hf = f(C); the
class of PWB' solutions is the class of finite constant functions. Since h is
minimal (Section 11.16), a bounded h-harmonic function is necessarily a
constant function; so we have proved that the Euclidean boundary is
h-resolutive and internally h-resolutive for this special choice of h.

Application to Half-spaces

Denote by dd the Nth coordinate of the point of URN and define D =


dd > 0}. For q in D let q* be the reflection of q in the boundary hyperplane
of D. Then GD is given by

logl-q*l
ifN=2
f - nl
(9.3)
IS - nl2-N - is - q*l2-N if N>2

because, as so defined, is harmonic on D - with the right


singularity at and has limit 0 at every boundary point of D, including oo.
In view of 1(8.5) it is to be expected that (Euclidean boundary) is
given by - augmented possibly by a contribution
from the singleton { oo }. Evaluation of this normal derivative leads to the
density function in (9.4). The 'N_I integral over URN-I of this density is I ; so
the natural conclusion is that dq) is given by

2dr
N'N-I(dn), =0. (9.4)
nNl - nl
In fact this evaluation of µD is correct because it is easily checked that if f
is a finite continuous function on OD, the function PD(-J), as defined using
(9.4). is harmonic on D with boundary limit function f. The function pD( J)
is called the Poisson integral off, just as in the ball case. Just as in the ball
case, to each boundary point q of D corresponds a minimal positive harmonic
function K(q, ) on D, a constant multiple of the harmonic measure density
in (9.4) for that value of q, with a special provision for q = oo. More specif-
9. h-Resolutive Boundaries 121

ically, if o is a point of D with duo = I and if K is normalized to make


K(-, 0) = 1, then

1x I tl - Col x if 00
K(q, 117

dS if q = 00.

It is shown that each function K(q, ) is minimal harmonic by showing that


there is a Riesz-Herglotz-type representation of an arbitrary positive
harmonic function on D by means of a unique measure M. on aD:

u() = K(C, (9.6)


JdD

The proof follows that in the ball case and is omitted. Just as in the ball
case, it is shown that the Euclidean boundary is universally resolutive and
that (9.2) is true in the present context.
Some of these results are easily reduced to the ball case by means of an
inversion in a sphere taking D into a ball. It will be shown in Chapter XII
that if D is an arbitrary Greenian subset of R', there is a universally resolu-
tive and universally internally resolutive boundary aMD, the Martin bound-
ary, and a function K on 3MD x D, such that K(q, -) is a minimal positive
harmonic function on D when q is in a certain subset ?4'D of eD, and that
to each positive harmonic function u on D corresponds a unique measure
on aMD, supported by a;'D, for which the counterparts of the Riesz-
Herglotz-type representation (9.6) (known as the Martin representation in
this general context) and of (9.2) are valid. The Martin boundary reduces
to the Euclidean boundary if D is a ball or half-space, in which cases a;'D =
aTMD.

The PWB Method and the Fatou Boundary Limit Theorem

According to Theorem I1.15, if h is a strictly positive harmonic function on


a ball B, the Dirichlet problem solution HI = PI(B, fdMM)/h has nontangen-
tial limit f(%) at M,, almost every boundary point C. This fact is at least a
partial justification of the PWB method. Furthermore, if D is an arbitrary
Greenian subset of RN, provided with a boundary by a metric compactifica-
tion, and if f is an h-resolutive boundary function, then (Theorem 3.11.2) H f"
has f as a boundary limit function along h-Brownian paths from a point of
D to the boundary in the sense that almost every such path tends to some
boundary subset A depending on the path (A is a singleton if the boundary
is h-resolutive), f is constant on each set A, and Hf has limit f(A) along the
path.
122 1 VIII The Dinchlet Problem for Relative Harmonic Functions

10. Relations between Reductions and Dirichlet Solutions


Let D be a Greenian set provided with a boundary 8D by a metric com-
pactification of D, and let B be an open subset of D with boundary 8B
relative to D u D. Let h be a strictly positive harmonic function on D,
and let u be an h-superharmonic function on D. Define a function f on
8B by setting J'= u on D n 8B and (if B is not relatively compact in D)
f = 0 on OD n BB. Observe that a function on B in the upper or lower PWB"
class for f will be in the same class for any other choice of dD and corre-
sponding choices of aB and off. The simplest choice for 8D in this context
is the Alexandrov compactification one-point boundary.
(a) If u z 0, the function f is an h-resolutive boundary function for B,
with PWB" solution the restriction to B of "R°-B and

{tB(h,U) ID)
u(S) ? N"9(4,u1D) _ (EB). (10.1)

If B, and B2 are open subsets of D with B, (-_ B2, then

NB,( ,u1D) ? (DEB,). (10.2)

In proving these assertions we assume, as we can without loss of generality,


that DD is the one-point boundary. Since this boundary of D is trivially
h-resolutive, it follows [Section 8, continuation of Example (b)] that 8B is
h-resolutive. If u' is a positive h-superharmonic function on D and majorizes
u on D - B, then the function u'lB is in the upper PWB" class on B for f;
so "R °-B z Hf. Since 0 < Hf 5 Hf , we conclude (Theorem 8) that f is
h-resolutive with Hf < 'R° B. In the other direction, if u' is now a function
on B in the upper PWB" class on B for the boundary function f, then the
function u' A (u1,) is also in this class and, when extended to D by u, is a
positive h-superharmonic function on D majorizing u on D - B and there-
fore majorizing "R°-B on D. Hence Hf = Hf z on B. so there is
"Ro_n

equality on B; that is, in terms of h-harmonic measure, "RD-B

on B. The inequality in (10.1) and the inequality (10.2) are now both trivial.
The equality in (10.1) follows from the relation between h-harmonic measure
and harmonic measure on D n 8B established in Section 8, Example (b).
(b) The positivity hypothesis imposed on u in (a) was made only to
allow the use of reductions. We now drop this hypothesis on u but suppose
that B is an open relatively compact subset of D. Then a trivial modification
of the proof in (a) shows that the restriction to B of the infimum of the
class I- of h-superharmonic functions on D majorizing con D - B is U),
the PWB" solution on B for the boundary function f = ulaB. Finally, (10.2)
remains true in the present context, and in this context u1D = U on B. If
B, and B2 are balls of center S and if h = 1. the inequality (10.2) reduces
11. Generalization of the Operator TB' and Application to GM` 123

to the fact that the function r) is a decreasing function for

11. Generalization of the Operator TB and Application to GM"


If u is an h-superharmonic function on an open subset D of R' and if B
is an open relatively compact subset of D, we define rru as the smoothed
infimum of the class of h-superharmonic functions on D, majorizing u on
D - B, so that tBu is h-superharmonic on D, equal (Section 10) on B to the
PWB" solution and equal to u quasi everywhere on D - B, in
particular, equal to u on the interior of D - B. When B is a ball, this defini-
tion agrees on D - d B with the Section 11. 1 definition of zeu and the Section 1
definition of rBu, and therefore the definitions agree everywhere on D
because two h-superharmonic functions equal 1, almost everywhere are
equal everywhere. Obviously, in all cases >a,u a tBzu when Bt c B2. If u
is h-subharmonic, 4Bu is defined as -TB( -u).
In view of Section 111.1 [especially Observation (a)] as generalized
trivially to allow arbitrary h and of the present extended definition of rB,
if D is a Greenian set, if B is an increasing sequence of open relatively
compact subsets of D with union D, and if u is an h-superharmonic function
on D, the limit of the decreasing sequence rB u of h-superharmonic functions
is, on each open connected component of D, either identically - 00 or
h-harmonic and, if h-harmonic on D, is GMpu. That is,

GMDu = iim tB,,u = limp u) (11.1)

if u has an h-subharmonic minorant. In particular, an h-potential u is


characterized among the positive h-superharmonic functions on D by the
condition lim, p u) _- 0.
A related result is the following, in which u is supposed positive to make
reduction notation possible. Let D, h, and B. be as in the preceding paragraph,
let u be a positive h-superharmonic function on D, define a bounary OD for
D by a metric compactification, and write D for D u aD. Then if A c D,

GMo''R. = lim
yM "R"-B
+U (11.2)

We prove this for h = I to avoid irrelevant notational complexities, and


we use the alternative reduction notation because iterated reductions will be
needed. Observe that
D"-Bft
DD-B,= Db-N QUBA-B",
D OuO" D Duo" z D Duo" =
and we have just proved that the limit (n - oo) on the left is GM°OuIA; so in
(11.2) the left side is at least equal to the right side. In the other direction,
124 I. VIII. The Dirichlet Problem for Relative Harmonic Functions

0°_1n
Du0"_B,
D DUD" D°-B' 5 + ll DUD"-B^ 0D-B1'
Duo"-B.,
5 + 0 DUD""B'" pD-H.

When n -+ oo, this inequality yields


DUD"-B"
GM0DuD" 5 Jim + GMDDUD"-,
M-OD

and the last term vanishes because A n B. is relatively compact in D; so


DUD"' " is a potential. Hence in (11.2) the left side is at most equal to the
right side so there is equality.

A Local Property of T'

If u is an h-subharmonic or h-superharmonic function and if B. is a decreasing


sequence of open relatively compact subsets of D with intersection , then
In fact the proof for h = 1 and B. a sequence of balls
of center [see Section II.6(f) for this result in a slightly different context]
is applicable in the general case.

12. Barriers
Let D be a Greenian subset of 18", coupled with a boundary OD provided
by a metric compactification, let h be a strictly positive harmonic function
on D, and let C be a point of aD. Usable conditions that C be h-regular are
most easily formulated in terms of "h-barriers" (see Section 13). A strictly
positive h-superharmonic function u on D will be called an h-barrier for D
at { if limn-, U(q) = 0 and if (*) info-Bu > 0 whenever B is a neighborhood
of If the condition (*) is omitted, u will be called a weak h-barrier for D
at By an easy application of the h-superharmonic function minimum
theorem, (*) is true if and only if Iim infn.E u(q) > 0 for in OD - {C }.
As usual, h will be omitted from the notation and the nomenclature when
h-1.
Local Nature of the Existence of an h-Barrier

Let Do be an open subset of D with boundary that relative to the topology


of D u aD, and let ho be the restriction of h to Do. Then if C is a common
boundary point of Do and D, the restriction to Do of an h-barrier for D at
is an ho-barrier for Dp at C. Conversely, if some neighborhood of C has
the same trace on Do as on D, the existence of an ho barrier uo for Do at C
implies the existence of an h-barrier for D at C. In fact, let B be the trace
12. Barriers 125

on Do of some open neighborhood of? so small that D n O B = Do n O B # 0,


and set a = infD0_Buo. Then a > 0, and if u is defined as a A ua on D n B
and as a on D - B, the function u is an h-barrier for D at C. Thus the existence
of an h-barrier for D at C is a local property of D near C, depending on h.

Lemma (Euclidean Boundary, h =- 1). If there is a weak barrier at a boundary


point {, there is a barrier at the point.

If N > 2 every unbounded open subset D of R' has a barrier at the point
ao, namely, the restriction of the function G(0, ) to D. An unbounded open
subset D of B2 has a [weak] barrier at the point ao if and only if the image
of D under an inversion in a circle has a [weak] barrier at the circle center.
Hence we can assume in the following proof that the boundary point C in
question is finite. In view of the fact that the existence of a barrier at C is
a local property of D, it is sufficient to show that there is a barrier on the
trace on D of an open neighborhood of C, so that D can be supposed bounded.
Suppose then that D has diameter 6 < + oo and that u is a weak barrier for
D at the boundary point C. Define O(PI) = J o7 - CI and f = 01,D. The function
46 is subharmonic, and OID is in the lower PWB class on D for f. Hence
Hf z ¢ on D, and it will be shown that Hf is a barrier for D at C by showing
that H. has limit 0 at C. Fix r > 0, let B = B(C, r), let A be a compact subset
of D n aB, and let 0 be the indicator function of (aB - A) n D on t?B,
so that PI(B, qi) is harmonic on B with limit I at every point of aB - A.
If uo is in the lower PWB class on D for f, the function uo on B n D is at most
r if 8B does not meet D and (by the maximum theorem for subharmonic
functions on B n D) is at most r + Su/infA u + SPI(B, 4i) if aB does meet D.
Thus in both cases

8u
Hf < r + + (Pl(B, +i) (12.1)
inf u
A

on B n D. The sum on the right has limit r + RI (B, 4/) (C) at C, and this
limit is at most 2r if A is sufficiently large. Since r is arbitrary, the function
Hf has limit 0 at C, as was to be proved.

EXAMPLE (a). If h is a strictly positive harmonic function on D with a finite


continuous strictly positive extension to b, then if u is a barrier for D at
a boundary point C, the function u/h is an h-barrier for D at C.

EXAMPLE (b) (Euclidean Boundary, h = 1). If D is unbounded and N > 2,


we have already noted that the function G(0, ) is a barrier for D at the
point oo.

EXAMPLE (c) (Poincare) (Euclidean Boundary, h = 1). If C is a finite boundary


point of D with the property that some closed ball meets D at C but at no
126 I VIII. The Dirichlet Problem for Relative Harmonic Functions

other point the restriction to D of the function C) - with the


ball center, is a barrier for D at C.

We shall show in Section 15 that it is sufficient for the existence of a barrier


at C if in this Poincare criterion the ball is replaced by a cone with vertex C.

EXAMPLE (d) (N = 2, Euclidean Boundary, h = 1). If is a finite boundary


point of D and if there is a simple continuous arc with initial point C, in
CBZ - D except for C, then there is a barrier for D at C. In fact, if B is a ball
of center C so small that the arc hits 8B, let Bo be B less the part of the arc
from 1; to the first hit of aB. Then Bo is simply connected; so the function
log(- - C) has a single-valued analytic branch 0 in Bo. The restriction to
B n D of the real part of 1/0 is a barrier for B n D at i ; so there is a barrier
for D at C.

13. h-Barriers and Boundary Point h-Regularity


Theorem. Let D be a Greenian subset of OB", coupled with a boundary 8D
provided by a metric compactification. If there is an h-barrier for D at the
boundary point ( and if f is an upper bounded boundary function, then

lim ;PHI(,) 5 lim Sup f(,) vf(C); (13.1)

in particular, if f is bounded and is continuous at C, then

lim HH(q) = lim HI(,) = f(c), (13.2)


r{ n-C

and therefore C is h-regular.

The second assertion follows from the first applied to f and -f. To prove
the first assertion, let u be an h-barrier at C. Let b be any number strictly
larger than the right side of (13.1), let B be a neighborhood of C so small
that f S b in the neighborhood, and let b be the infimum of u outside D n B.
Choose n so large that b + no exceeds the supremum of f. The function
b + nu is in the upper PWB' class on D for f and has limit b at C. Hence the
left side of (13.1) is at most b; so the theorem is true.
Extension. Since a change off on a set of h-harmonic measure 0 does
not change HI or H f, the point , can tend to C on the right side of (13. 1)
on the complement of such a boundary set, and f(C) can be omitted on the
right hand side if {C} is h-harmonic null, as is true when h - I and 8D is
the Euclidean boundary, unless N > 2 and C = oo. This extension of Theorem
13 reduces to Theorem 11. 1 when D is a ball, h 1, and 8D is the Euclidean
boundary.
14. Barriers and Euclidean Boundary Point Regularity 127

14. Barriers and Euclidean Boundary Point Regularity

Theorem. (Euclidean Boundary, h = 1). A boundary point is regular if and


only if there is a barrier at the point.

If there is a barrier at a boundary point, the point is regular by Theorem 13.


Conversely, suppose that the point C is a regular boundary point of the
Greenian set D. If N > 2 and if S = oo, there is a barrier, exhibited in Section
12, Example (b). If N = 2 and C = ao, make C finite by an inversion in a
sphere with center a finite boundary point. Thus C can be supposed finite
in proving the existence of a barrier there.
If D is bounded, define f = I - C1 ia, and u = I - CIID. Then u is subhar-
monic and is in the lower PWB class on D for f ; so Hf Z u, and Hf is a barrier
at C because (regularity of C) Hf has limit 0 at C. If D may not be bounded, the
cases N > 2 and N = 2 will be treated separately. If N > 2, define f and u as
the restrictions to 8D and D, respectively, of E; n-3 [n - G(C, ) A n], with
f(oo) = ET n-Z if oo a D. The function f is a finite-valued continuous posi-
tive boundary function vanishing at C and only there, and the function u is
a continuous positive subharmonic function in the lower PWB class for f;
so Hf z u. Moreover Hf has limit 0 at C because C is regular. Hence Hf is
a barrier at C. If N = 2, let B be a ball of center C so small that Dt = D U B
is Greenian. The function Zr n-3 [n - GD,(C, ) A n] is a positive bounded
subharmonic function on D1. Define u as the restriction to D of this function,
and define f at each point ?l of OD by f(q) = lim sup;._ Then f is positive,
bounded, upper semicontinuous, and vanishes at C and only there. The
function u is in the lower PWB class for f on D; so Hf z u. Moreover Hf
has limit 0 at C because C is regular. Hence Hf is a barrier at C.
Application (a) : Local Property of Regularity. (Euclidean Boundary,
h = 1.) Since the existence of a barrier at a boundary point is a local property
of a Greenian set D near the point, regularity at a point of the Euclidean
boundary is also a local property of D near the point.
Application (b) : Regularity of a Boundary Point of a Disconnected Set.
(Euclidean Boundary, h - 1.) Let D be a Greenian subset of R', let C be a
boundary point of D, and let D1, D2, ... be the open connected components
of D with boundary point C. Then C is a regular boundary point of D if and
only if C is a regular boundary point of each set Dk. In one direction, if C is
a regular boundary point of D, the restriction to Dk of a barrier for D at C
is a barrier for Dk. Conversely, if C is a regular boundary point of each set
D. and if uk is a weak barrier for Dk at C, then the function u defined on D
by setting u = 2-k(uk A 1) on Dk is a weak barrier for D at C.
Application (c) : Regularity of a Boundary Point in Terms of the Green
Function. (Euclidean Boundary, h =_ 1.) A boundary point C of a Greenian
set D is regular if and only if q) = O for some (equivalently every)
128 1.VIII The Dinchlet Problem for Relative Harmonic Functions

point of each open connected component of D with boundary point C. In view


of application (b) we can assume in the proof that D is connected. If the
condition on GD is satisfied for a single point (, the function is a
weak barrier at (; so ( is regular. If N > 2 or if N = 2 and if D is bounded,
the expression for GD((, ) in terms of a Dirichlet solution (Section 3) implies
the truth of the converse, for all . The following proof of the converse is
valid in all cases. If ( is regular and if u is a barrier for D at (, let ( be a point
of D, and let B, be the set c}, where c is a constant chosen so
large that B, is relatively compact in D. Such a choice is possible because
GD((, ) is bounded in a neighborhood of OD. Choose n so large that nu > c
on B,. Then (Section VII.3) GD((, ) S nu outside BB; so has limit 0
at C.

Application (d): Relative Boundaries. If D is a Greenian subset of R',


provided with a boundary t3D by a metric compactification, if h is a strictly
positive harmonic function on D, and if B is an open subset of D with relative
boundary 8B in D v aD, then quasi every point of t3B in D is h-regular
because if u is a local barrier for B at a boundary point of B in D, then u/h
is a local h-barrier for B at the point and (Section 12) can be extended to B
to be an h-barrier there.
Application (e) : The Kelvin Transformation and Regularity. (Euclidean
Boundary, h = 1.) We use here the notation of the discussion in Section 2
of the PWB method and the Kelvin transformation. If ( is a regular finite
boundary point of D, then its image under inversion in a sphere is a regular
boundary point of D' because if u is a barrier for D at (, the Kelvin transform
of u is a barrier for D' at O(() except possibly when N > 2, and the inversion
sphere has center C. In this case, however, the image of ( is the point oo,
which is a regular boundary point of every unbounded Greenian set when
N > 2 (Theorem 4). If N > 2 and if oo is a (necessarily regular) boundary
point of D, the image of oo under an inversion may or may not be a regular
boundary point of D'. If N = 2 and if oo is a regular boundary point of D,
the image of oD under an inversion is a regular boundary point of D' because
the Kelvin transform of a barrier for D at co is a barrier for D' at the image
of co.

15. The Geometrical Significance of Regularity (Euclidean


Boundary, h = 1)

If S is a boundary point of D with the property that some neighborhood of


( meets OD in a harmonic measure null set, the point ( is irregular because
a PWB solution Hf is not affected by a change off on such a set. This fact
together with Poincar&'s criterion [Section 12, Example (c)] for the existence
15. The Geometrical Significance of Regularity (Euclidean Boundary, h =- 1) 129

of a barrier suggests that regularity of C, equivalently, the existence of a


barrier at C, amounts to the requirement that D not fill up "too much" of
a neighborhood of C. Thus an isolated finite boundary point C is irregular
because {C} is polar and therefore harmonic measure null, and when N = 2,
the point oo (if an isolated point of OD) is irregular because D can be mapped
into a Greenian set by an inversion leaving harmonic measure invariant
and taking oo into a finite boundary point. The point oo is exceptional in
that when N > 2, the harmonic measure {oo}) may be strictly positive
and, in fact, is identically I when D = l'. (Recall that when N > 2, the
point oo is a regular boundary point of every unbounded Greenian set.)
It will be shown in Section XI.12 that a finite boundary point C is regular if
and only if C is a limit point of 68" - Din the fine topology, and this criterion
will be stated probabilistically in Section 2.IX. 15.

Poincare-Zaremba Regularity Criterion

Poincare's regularity criterion [Section 12, Example (c)] was improved to


the following: if C is a finite boundary point of D with the property that
some open solid cone of revolution with vertex C does not meet D in a
neighborhood of C, there is a barrier at C (so C is a regular boundary point).
In view of the discussion of barriers in Section 12, it is sufficient to prove
that if A is a closed cone of revolution with vertex the origin and if D =
B(0, 1) - A, then D has a barrier at the origin. Define ¢() = and f = ¢iaD.
Since is subharmonic on D, 41D is in the lower PWB class on D for f;
so Hf z 01D, and we show that Hf is a barrier at the origin by showing that
Hf has limit 0 there. Let Do be the part of D at distance <# from the origin.
Observe that Hf < l on D and that Hf has limit f at every point of 8D - {0}
because (Poincare criterion) every point of 8D - {0} is regular. By the
harmonic function maximum theorem a = sup,, Hf < 1. The function
i-. (a v is a bounded harmonic function on Do and has
a negative limit at every point of OD0 - {0}. Hence (Section V.7) by the
extended harmonic function maximum theorem Hf() - (a v 5 0,
and therefore

lim sup Hf( ) < (a v 1) lim sup Hf( );


4-0 4-0
so the superior limit is 0, as was to be proved.

Strengthened Poincare-Zaremba Criterion

In the preceding proof, if A, is an (N - 1)-dimensional hyperplane containing


the axis of the cone A, A n At is a flattened cone, and the preceding proof
with A replaced by A n A, is valid with trivial changes. Thus a finite boundary
point of an open subset of R' is regular if some flattened cone in this sense
130 I Vlll. The Dinchlet Problem for Relative Harmonic Functions

has vertex C and does not meet D in a neighborhood of C. The reader is


invited to strengthed this criterion still further by weakening the conditions
on At.

Regularity of Classical Boundaries: The Lebesgue Spine

Most of the open sets used in classical analysis have regular boundaries
because there are Poincare-Zaremba barriers at their boundary points. On
the other hand, if the excluded cone is sharpened into a cusp that is suffic-
iently sharp, a barrier may no longer exist at the vertex. For example,
suppose that N = 3, denote a point of 683 by its coordinates t'l, t2', t31,
let A be the closed line segment with endpoints the origin and the point
(1, 0, 0), and let u be the measure supported by A and determined by p(dd) =
""/t(ddt"). The potential Gp has value I at the origin, value +oo elsewhere
on A, and limit I at the origin along the negative 't) axis. The set D =
{ # 0: Gp(l:) < 2} is a solid of revolution about the fit" axis, includes the
negative fit" axis except for the origin, and has an exponential cusp at the
origin; D is the Lebesgue spine. We now show that the origin is an irregular
boundary point of D. If f is the boundary function equal to 2 at the finite
boundary points of D and equal to 0 at oo, then f is resolutive because the
Euclidean boundary is always resolutive. Moreover Hf is the restriction of
Gp to D because this restriction is a bounded harmonic function that has
the prescribed limit at every boundary point with the exception of the origin
and {0} is a harmonic measure null set. The origin is not a regular boundary
point of D because Hf has limit I at the origin along the negative l'I axis.

16. Continuation of Section 13


Let D be a Greenian subset of R', provided with a boundary aD by a metric
compactification, and let It be a strictly positive harmonic function on D.
(a) If OD is h-resolutive and if C is an h-regular boundary point with
1, then there is a weak h-barrier at C. In fact, if A. is the part
of the boundary at distance z2-" from C, the function Eo 2-"po(-,A") is a
weak h-barrier at C.
(b) If OD is h-resolutive and if C is a boundary point with u ,(-, {C}) = 1,
then C is necessarily h-regular, but there may be no weak h-barrier at C.
For example, if h = I and if OD = {C} is the one-point boundary, then it is
trivial that aD is resolutive and thatpo(-, {S}) = 1. In this case there can in
general be no weak barrier u for D at C because such a function would be a
strictly positive superharmonic function on D with limit 0 at every Euclidean
boundary point of D. Then u would be a weak barrier for D at every Euclidean
boundary point; so the Euclidean boundary would be regular, and this
regularity would be a restriction on D.
17. h-Harmonic Measure po as a Function of D 131

17. h-Harmonic Measure µo as a Function of D


Theorem. Let D be a Greenian subset of l provided with a boundary aD by
a metric compactifcation, let h be a strictly positive harmonic function on D,
and suppose that 8D is h-resolutive. Boundaries of subsets of D are to be
relative to D u OD.
(a) If Do is an open subset of D, then OD, is h-resolutive and

e,A) (17.1)

on Do whenever A is a Bore! subset of 8Do n 8D.


(b) If D. is an increasing sequence of open subsets of D with union D,
then for in D,

(17.2)

(vague convergence on D u aD). Moreover, if e D. and if A is a


Borel subset of OD. n aD, then

A) A) S ... < RD( , A) (17.3)

The h-resolutivity of aDo was proved in Section 8. Inequality (17.1) is a


trivial consequence of (8.3), and (17.3) is simply a repeated application of
(17.1). To prove (17.2), it will be shown that if f is a f inite-valued continuous
function on D u (D, then

(17.4)

If u is in the upper PWB" class on D forf0D and is bounded, extend u to u'


on D u OD by setting u'(C) = lim inf,l_t u(n) for C E OD. Then if e > 0, the
function u' + e - f is lower semicontinuous on D u aD, strictly positive on
OD, and therefore also strictly positive on W. for sufficiently large n. Hence
for sufficiently large n the restriction to D. of u + e is in the upper PWB"
class on D. for the boundary function I,BD; so ju%(-, f) S u + e on D,,. It
follows that lim sup.-. ,u' f) S f), and this inequality together with
the corresponding inequality for -f yields (17.2).
Application. In Theorem 17 suppose that It = I and that OD is the Eucli-
dean boundary. Let D; be an increasing sequence of subsets of D U (D, open
relative to D u 8D, define D. = D n D., and suppose that Uo D. = D.
Suppose that A 1 is a subset of D u OD, open relative to D u aD, and that
Al c D for sufficiently large n. Then if A is a Borel subset of Al n 8D and
if eeD,
132 I. VIII. The Dirichlet Problem for Relative Harmonic Functions

li m A) = A). (17.5)

To see this, define A2 = At n 8D, and observe that if n is so large that e Dn


and A t c D,;, then A 1) = pDn(4, A2) and At) = A2). In view
of the vague convergence in Theorem 17(b),

liminfpD( uD( ,At)=uD( ,A2), (17.6)


n-.O "

and in view of (17.3) with A = A2, inequality (17.6) implies that (17.5) is
true when A = A2. Now by the monotoneity in (17.3),

lim A) _< ID(b, A), li m A2 - A)<_ A2 - A) (17.7)

for A c A2, and since the sum of these two inequalities yields an equality,
there must be equality in each; that is, (17.5) is true.

18. The Extension Go of GD and the Harmonic Average


GB (17, )) When D c B
In the following theorem D and B are Greenian subsets of IF" with D c B,
and boundaries are Euclidean. Recall from Section VII.4 that for in D
the function GD is an extension of to R", subharmonic on
Q2" - {4}, vanishing on I8" - D and quasi everywhere on R' n 8D, with

GD (, ) = lim sup q) (!; a I8" n OD). (18.1)


Daq-(
In particular, if C is a finite regular boundary point of D, the function
is continuous at with value 0 there; if D is not bounded, this
function has limit 0 at co when N > 2 and also when N = 2 if oo is a regular
boundary point of D.
In discussing GB (n, )) when D is unbounded and N = 2, the value
assigned to GB (q, oo) is irrelevant because (Section 5) the singleton (co) is
PD null. When N > 2 and D is unbounded, the singleton {oo} may not be
µD null, but we define GB (s7, co) = 0, corresponding to the fact that GB (q, )
has limit 0 at oo.
If u is a superharmonic function on a superset of D, we write GMDu
instead of GMD(uID).

Theorem. (a) For each q in B the function GB (11,911D, defined as 0 at oo


if D is unbounded, is a resolutive boundary function.
(b) If * ED, then

GL 4) - Ge (7,')) (q a B); (18.2)


18. The Extension Go of Gp and the Harmonic Average po G; (q. .)) When D c B 133

that is,

GG q) = GB(S, q) - GB(q, )1 B) (I a B). (18.2')

(c) If v = GBV is a superharmonic potential on Band if v' is the projection


of v on the set of irregular boundary points of D in B, then on D,

GMDV = v1B) + Go v'. (18.3)

If D is relatively compact in B, (18.3) is true whenever v is a superharmonic


function on B and v is its associated Riesz measure.

Observation (1). If q E D, (18.2) becomes

q) = q) - GB(q, -) lB) [( , r) e D x D], (18.4)

and we thereby have found the important symmetry relation

GB(q, ) I B) = PD('i, GB(S, ) lB) [(c, q) e D x D]. (18.5)

Note that we can take B = R' here when N > 2. With this specialization
(18.4) was derived in Section 3, application of Example (a). It will be shown
in Section 19 that (18.4) and (18.5) are true when N = 2 and B = l 2 if and
only if 112 - D is not too sparse near oo, more precisely, if and only if D is
bounded or if unbounded D has oo as a regular boundary point.
Observation (2). If we write for the restriction of the measure
to the class of Borel subsets of B, then according to Theorem 18,

+oo x B]; (18.6)

so for fixed in D the potential is a finite-valued superharmonic


function on B. Thus Theorem 18(b) implies that the measure on It" -
associated with the superharmonic function on this set is the
restriction of the harmonic measure to the class of Borel subsets
of I2" -

Proof of (a). For q in B define f on 8D as GD(q, ) on B n 8D and as 0 on


aB n 8D. Since G8 (q, ) = 0 at quasi every point of aB, assertion (a) states
that the function j, is a resolutive boundary function. This fact was proved
in Section 10 (but observe that the notation here reverses the roles of D and
Bin Section 10). o

Proof of (b). (When q e B - 8D.) If q c- B, the function GB(q, -)JD is in the


upper PWB class on D for the boundary function ft. If q e B - aD, the
134 I V111. The Dirichlet Problem for Relative Harmonic Functions

function GMDGB(q, ) is in the lower PWB class on D for fq whenfq is in-


creased to + oo on the polar set of irregular boundary points of D. Since a
change of boundary function on a polar set does not change PWB solutions,

GMDGB(q, ) 5 AD(., Ge (V)) 5 G0(,91 (7EB-OD), (18.7)

and since the middle term in (18.7) is harmonic, the last term can be replaced
by GMDGB(q, ) to yield the equality

GMDGB(1,') = PD(', G;07,9) (qeB-6D). (18.8)

Under this equality equation (18.2) with q in D reduces to the expression


for GD in terms of GB derived in Section VIII [see V1I(1.2)]. If qeB - D,
(18.2) reduces to

0= i) Ge (q, )) ( a D),

and since GB(-, q) is harmonic on D, this equation is a trivial consequence


of (18.8). Thus we have proved (18.2) for P7 in B - 8D.
We now prove

OGB(q,')QB-D( ) = x B]. (18.9)

We shall need this symmetry relation which is a special case of a fundamental


symmetry relation to be derived in Section X.3. For and q in D equation
(18.9) reduces to (18.5) [see Section 10(a)], and (18.5) is true because, as
just proved, (b) is true when q is in D. Thus, if q) e D x (B - D), equation
(18.9) is true if D is enlarged to D u B(q, I /n) with 11 - aDJ < I In. When
n - oo, it follows [from Section VI.3(e)] that (18.9) is true if D is replaced
by D u {q} and therefore is true with no replacement because [by Section
VI.3(c)] a smoothed reduction is unchanged if the target set is changed
by a polar set.
(b) (When q e B n aD.) Suppose first that q is a regular boundary point
of D in B. With this choice of q, (18.2') reduces to

GB(, , ) I B) = q), (18.10)

or equivalently [Section 10(a)],


p
DGB(q,.)18_D(S) = GB(z, q).

Now in view of the symmetry of the left side of (18.10), proved above, and
of the special lower semicontinuity property I1(6.1) of superharmonic func-
tions,
18. The Extension GD of GD and the Harmonic Average pD Ge (q. )) When D c B 135

OGB(n, -) pB-DO = p GB(, ) pB °(q)


pB-D(C),
= lim inf pB-D(() A liminf
D3{^9 8-D3{-q
(18.12)

and if convenient, C can be allowed to tend to q on B less an arbitrary IN


null set. The first limit inferior in (18.12) can be written in the form
lim infD C-,I D(C, I B), and since C is a regular boundary point of D,
the limit inferior is actually a limit and is GB(., q). Since OGB(, ) pe-° =
GB(., -) quasi everywhere on B - D, the second limit inferior in (18.12)
is q) unless some neighborhood of q meets B - D in an IN null set,
in which case we can ignore this limit inferior. Hence (18.11) is true. We have
now proved that (b) is true if q either is in B - aD or is a regular boundary
point of Din B; so (b) is true for quasi every point q of B. Since the two sides
of (18.2) are equal when q = and define subharmonic functions of it on
B - {.;}, these two sides are equal for all q in B, and the proof of (b) is
complete. o

Proof of (c). If v = GBv, apply (18.2) to find that

GDv'=GDv=v-Io(',vIB), (18.13)

and then an application of the linear operation GMD to (18.13) yields (18.3).
If D is relatively compact in B and if v is superharmonic on B, it can be
supposed in proving (18.3) that v is lower bounded on B, after decreasing
B if necessary. In view of the Riesz decomposition it is then sufficient to
prove (18.3) separately for v a potential, in which case the proof has just
been given, and for v harmonic, in which case (18.3) is trivial. o

Application to the Vanishing of h-Potentials at the Boundaries of Their


Domains

Let D be a Greenian subset of R', let h be a strictly positive harmonic


function on D, let u = GDv/h be an h-superharmonic h-potential on D,
and define DD = {u > c}, for c > 0. Then DD is an open subset of D, and we
now prove that if DD is not empty, then (Euclidean boundaries)

Define v = GDv. We can assume that u S c + 1, that is, v 5 (c + 1)h, be-


cause DD is unaltered if we replace u by u A (c + 1). The measure v vanishes
on polar sets because v is finite valued; so (Theorem 18)

GMD V = (18.14)
136 1.VIll. The Dirichlet Problem for Relative Harmonic Functions

Now on the one hand (GMD v)/h = hGMDDv on D, and on the other hand
(Section 10(a)) the evaluation µ4,(, ul p) = vl p)h combined with (18.14)
yields

"GMDDu = ul p). (18.15)

The left side of (18.15) is a majorant of the constant function c and u 5 c


on D n OD, by lower semicontinuity of h-potentials, so (18.15) implies that
c5 D n 8D,) and we conclude that there is equality here and therefore
that dD n 0, as asserted.

19. Modification of Section 18 for D = R2


In Section 18 the set B is Greenian and so cannot be chosen to be RN unless
N > 2. In this section N = 2, the set D is a Greenian subset of R2, and the
work in Section 18 is adapted to the choice B = R2. As in Section 18 the
topology defining boundaries is the Euclidean topology. We adopt the
convention that oo) = - oo. The counterpart of the Section 18 set
a B r OD is the empty set if D is bounded and the singleton (00) if D is
unbounded. Recall from Section 5, Example (a), that when D is unbounded,
the singleton {oo} is a µp null boundary set. According to the following
theorem, an awkward new term appears in the counterpart of (18.2).

Theorem. (a) For each n in R2 the function G(n, -) is a resolutive boundary


function; that is, this function is in L'(0).
(b) If e D, then

GD n) = n) - pD(S, G(n, ')) + 0D(') (n a 022), (19.1)

where ¢D is a positive harmonic function on D, defined in (19.10), and


(bl) OD has limit 0 at every finite regular boundary point of D.
(b2) Op is bounded on bounded sets.
(b3) (pp = 0 if D is bounded.
(b4) if D is unbounded and connected, OD - 0 if and only if ae is a regular
boundary point of D.
(c) If v = Gv is a superharmonic potential on R2 and if v' is the projection
of v on the set of finite irregular boundary points of D, then on D,

GMDv = PD(-1 V) + Go v' - v(R2)OD. (19.2)

If D is bounded, (19.2) is true with Op 0 whenever v is a superharmonic


function on D and v is the associated measure.
19. Modification of Section 18 for D = R2 137

Observation (1). If q e D, then (19.1) becomes

GD(4, q) = G(4, q) - YD(', G(q, )) + [(c, q) E D x D]. (19.3)

According to (b3) and (b4), the following two important implications of


(19.3) are true if and only if OD = 0, that is, if and only if D is bounded or
is unbounded with regular boundary point oo :

q) = q) - G(q, )) q) E D x D], (19.4)

G(q, )) = pD(q, q) e D x D]. (19.5)

According to Section 18, Observation (1), both these relations are true when
N > 2 with no restriction on the nonempty open subset D of R". Equation
(19.4) is a natural approach to finding the Green function GD, and Theorem
19 exhibits the conditions under which it is valid, that is, under which
G(q, )) when N = 2. If there is no restriction on D, the
difference between left and right sides of (19.5) is 4D(S) - OD(q)
Observation (2). If C is a finite regular boundary point of D or is an
inner point of R2 - D, equation (19.1) yields

C) = G(g, )) - 4 D(4). (19.6)

Observation (3). According to Theorem 19, the integral

G(q, )) = G t) e D x R2] (19.7)

is well defined and finite; so for fixed in D the potential is a


finite-valued superharmonic function on R2. [There is a slight abuse of
language here because when D is unbounded, is not a measure of
subsets of R2, but we have already noted that oo) = 0 in the present
context.] In view of (19.7) Theorem 19(b) implies that the Riesz measure
on R2 - {} associated with the superharmonic function on that
set is

Proof of (a). For q e R2 - D and, if D is bounded, for q e 6D. If we define


f = G(,, )iaD, the function f is resolutive under the stated restrictions on
q and D because Euclidean boundaries are resolutive and

sup f, < + oo if q e R2 - aD,


- co < GM
G(q, ),D < + co if q e aD, D bounded.
(19.8)
138 1 VIII. The Dinchlet Problem for Relative Harmonic Functions

The function q - G(q, )) = f) is the potential for the kernel G


of the measure pn(C, ) and is superharmonic because it is finite on R2 - 8D.
Proof of (19.1) when D is bounded. If D is bounded and if B = B(0, 8)
contains D, then

GD (C, ) = GB(C, ) - PD(C, GB(l,')) (19.9)

according to Theorem 18. Furthermore according to the formula for GB


in 11(1.1),

GB(C, q) = G(C, q) + ha(C, q),

where ha is a symmetric function on B x B and ha(C, ) has a harmonic


extension to a neighborhood of B. If this evaluation of GB is substituted
in (19.9), we obtain (19.1) with OD = 0.
Proof of (a) and (19.1) when D is unbounded. If D is unbounded, let B;
be an increasing sequence of relatively compact open subsets of R2 with
union 982 and define B. = B. n D. If we write (19.1) for B,,, let n ao, and
take into account the fact that PD(C,ft) > - oo, we find (19.1) first for quasi
every q and then for all q, with

OD(S) = - lim f G(q, dC); (19.10)

for example, (19.10) can be particularized to

AD(C) = li m [p8.(C, D n OB.) log n] [B = D n B(0, n)]. (19.11)

Thus (a) and (19.1) are now completely proved.

Proof of (bl)-(b3). If C is a finite regular boundary point of D, fix q in D


and observe that since ft is upper bounded, the PWB solution G(q, ))
has superior limit 5 G(q, C) at C (Theorem 13). Moreover GD(q, ) has limit 0
at C; so (19.1) implies that

0 z G(C, q) - lim sup ID(, G(q, )) + lim sup WD(S),


r-c a-c

and so (bl) is true. The function OD is bounded on bounded sets, that is,
(b2) is true, because in (19.1) for fixed q in D the function G(q, ) is bounded
above on OD and bounded below on bounded sets, and q) is bounded
outside each neighborhood of q. Finally we have already proved (19.1) with
OB = 0 when D is bounded; that is, (b3) is true.
20. Interpretation of 0D as a Green Function with Pole oo (N = 2) 139

Proof of (b4). Observe first that with no hypotheses on D,

1im sup q) - G(q, ))] = li%!up f,,) log 71


! D(' ,''S) S 0
D9il-ao w I - ql
(19.12)

in view of Fatou's lemma, because the integrand has limit 0 when q - o0


and is at most

log 1 + I{ -I 5 log(1 + I -I) if I - ql > 1. (19.13)


C I - nl l
Here the right side of (19.13) defines a function of in L'(t0), according
to Theorem 19(a). If ¢D - 0, inequality (19.12) combined with (19.1) shows
that has limit 0 at oo ; so (Section 14) oo is a regular boundary point
of D. To prove the converse, apply in (19.1) the evaluation of limd.. L(u, 0, S)
with u a potential (kernel G) on 682 to find

lim L(GD (, ), 0, b) (19.14)


e-M
In particular, if oo is a regular boundary point of D, the function Go
has limit 0 at oo; so the left side of (19.14) is 0, and OD - 0, as was to be
proved.

Proof of (c). See the proof of Theorem 18(c).

20. Interpretation of OD as a Green Function with Pole oo


(N = 2)
If D in Section 19 is a deleted neighborhood of the point oo, this point is
an irregular boundary point of D, and the limit equation (19.12) simplifies to

lim PI) - G(q, ))] = 0. (20.1)


qm
Equation (19.1) now yields

lim q) = 0D(0- (20.2)


'-m
It is natural to write the limit on the left as oo) and to think of oo)
as the Green function of D with pole ao. See Section XIII.18 for further
remarks on this Green function.
140 I. VIII. The Dirichlet Problem for Relative Harmonic Functions

21. Variant of the Operator TB


If u is a superharmonic function on an open subset D of RN (N z 2) and if B
is an open relatively compact subset of D, we have defined TBu in Section 11
as the smoothed infimum of the class of superharmonic functions on D
majorizing u on D - B. Under this definition TBU = u) on B and TBu = u
quasi everywhere on D - B. Define TBu as the superharmonic function on
D equal quasi everywhere on D - B to u and equal on B to GMBu. The
construction of GMBu in Section III.1 in terms of a decreasing sequence of
superharmonic functions shows that TBu exists. Since TBu 5 u on B, it follows
that TBu S rBu. According to Theorem 18(c) (in which the roles of B and D
are reversed), TBU = rBu if and only if the set of irregular Euclidean boundary
points of B is a null set for the Riesz measure associated with u; so there is
equality if 8B is regular. In the general case both TBu and TBu decrease when
B increases. Moreover, if Bt and B2 are open relatively compact subsets
of D, with Bt c B2, then

TB,u Z TB,u Z TB2u -- TB2u. (21.1)

For many purposes we can use tB and tB interchangeably. For example, if


B varies through the open relatively compact subsets of D, then (21.1) shows
that r8u and TBu as B varies have a common infimum; the infimum is GMDU
if this harmonic minorant exists.
Chapter IX

Lattices and Related Classes of Functions

1. Introduction
In this chapter certain function classes that arise naturally in potential
theory will be discussed. These classes, the corresponding identically named
classes in parabolic potential theory (Section XVIII.19) and in stochastic
process theory (Chapter V of Part 2), are discussed together in Chapter I
of Part 3.
Throughout this chapter D is a Greenian subset of R', N z 2, and h is
a strictly positive harmonic function on D. Both D and h are held fast
throughout the chapter.

2. LMDu for an h-Subharmonic Function u


Suppose that D is connected and that u is an h-subharmonic function on D.
According to Section VIII.l 1 (where the discussion is in the dual context,
in which u is h-superharmonic), if A and B are open relatively compact
subsets of D with A c B, then u5 TAU 5 rq'u, and if B is an increasing
sequence of open relatively compact subsets of D with union D, then tB,u
is an increasing sequence of h-subharmonic functions, and either LMDu exists
and is given by tu, an h-harmonic function, or this limit is identically
+oo. An equivalent formulation is: the set of functions teu for B ranging
through the class of open relatively compact subsets of D is directed upward
and either LMDu exists and is given by

LMDu = sup { u) : B open relatively compact subset of D}, (2.1)

or else the supremum in (2.1) is identically +oo. [We adopt the convention
here and in similar contexts below that the domain of µ8(-, u) is B and that
the supremum of a set of functions at a point is the supremum at the point
of the values of those functions defined there.] Observe that the supremum
(directed limit) in question is unchanged if the sets B are all supposed to
contain a specified compact subset of D; that is, the analysis relates to the
properties of u near (any choice of) the boundary of D.
142 1. IX. Lattices and Related Classes of Functions

If u is positive, the existence of LMDu is equivalent to the L' boundedness


of the class

{ [ulaB, µe(, )] : B open relatively compact subset of D} (2.2)

of functions coupled with the indicated measures, for some, equivalently


each, point of D. Alternatively the set B can range through a nested
sequence B, with union D, as above. In particular, suppose that u is positive
and that the class (2.2) is uniformly integrable for some point in D, that is,
that there is a uniform integrability test function fi for which

sup c(u)): B open relatively compact subset of D} < + oo. (2.3)

Then the class (2.2) is L' bounded; so LMDu exists, and since ((u) is a
positive h-subharmonic function, the supremum in (2.3) is
Thus in this case LMDF(u) exists, and the supremum in (2.3) is finite for
all ; that is, the class (2.2) is uniformly integrable for each point in D.
Furthermore the h-harmonic function LMDu satisfies the same uniform
integrability condition as u, with the same test function 0; (2.3) is true with
u replaced by LMDu. In fact we now show, under the hypothesis that (2.3)
is true as written, that

sup b(LMDu)): B open relatively compact subset of D}


(2.4)
= LMD(O[LMDu]) = LMDt(u).

To see this, let B and B' be open relatively compact subsets of D. Then as B'
varies,
r 1l
[LMDu]) = µe -0 [suP µa4(', u)J (suPP.(.G(u)))
\\ B'
(2.5)
= LMD(D(u)) = LMDt(u)O < +oo.

Take the supremum as B varies to find that the first two terms in (2.4) are
majorized by the third. The reverse inequality is trivial.

3. The Class D(4D_ )


(See the corresponding stochastic process class in Section 2.11.11. This class
is the linear class of real-valued Borel measurable functions u on D for which
if is in D and if B is an increasing sequence of open relatively compact
subsets of D with union D, then the sequence

(3.1)
3. The Class D (t4-) 143

of coupled functions and measures is uniformly integrable. It follows from


the definition of uniform integrability that a Borel measurable function u
on D is in D(µ4_) if and only if, for each point t in B, pB(, Jul) < + oo
whenever B is an open relatively compact subset of D and there is a compact
subset A = A4 of D for which the family (2.2) of coupled functions and
measures, under the added restriction that A c B, is uniformly integrable.
This condition is satisfied if and only if there is a uniform integrability test
function D = '{ such that

sup { lt"B(1, (Dfl u1)) : A c B, B open relatively compact subset of D} < + oo.
(3.2)

The notation D- is to remind the reader that not h-harmonic measure on


some boundary of D but h-harmonic measures on the Euclidean boundaries
of open relatively compact subsets of D are involved. When D and At are
specified and there is no danger of confusion, we shall sometimes write D
instead of D(µo_). Since a function u is in D(pD_) if
and only if uh is in D(µD_). (See Section 11. 14 for the D class of harmonic
functions on a ball.)
When the function u on D is h-subharmonic the situation is simpler
because then the map is monotone increasing. In particular,
if u is positive and h-subharmonic, the supremum in (3.2) is independent
of the choice of A, and (Section 2) if (3.2) is true, the left side of (3.2) is
It is thus natural that the case of most interest in potential
theory is that in which Jul is h-subharmonic, in particular, if u is h-harmonic.

Theorem. If u is a Borel measurable real-valued function on the connected


Greenian set D and if I uI is an h-subharmonic function on D, the following
conditions are equivalent :
(a) ueD(µo_).
(b) The family (2.2) of coupled functions and measures is uniformly inte-
grable for every point in D.
(c) Condition (b) is satisfied for a single point .
(d) There is a uniform integrability test function 0 such that the h-
subharmonic function ((jui) has an h-harmonic majorant.
(e) (If u is h-harmonic) u = u1 - u2, where ut is a positive h-harmonic
function in D(µo_).
Moreover, if u satisfies (a)-(c) and if (V satisfies (d), then LM'Dju' e D(µD_),

LMDN(LMoJul) = LMo D(Jul), (3.3)

and each function u, in (e) can be chosen so that the h-subharmonic function
'(u;) has an h-harmonic majorant.
144 1.1X. Lattices and Related Classes of Functions

If Jul here is identified with u in Section 2, this theorem follows at once


from the discussion in Section 2, except for the assertion involving (e). To
prove that an h-harmonic function u in D(µ4_) has the representation (e),
observe that since Jul is h-subharmonic and LM4IuJa D(µ4_), it follows that

u = LM4Iul - (LM4Jul - u)

is the desired representation; with this choice of u1, equality (3.3 shows that
(D(u;) has LM"(D(Iu1) as an h-harmonic majorant. Conversely, if u = ut - u2
with u; positive h-harmonic and in D(µ4_ ), then I uI 5 ut + u2 ; so u e D(µ4_ ).
It will be shown in Section 3.1.9 that an h-harmonic function on a Greenian
subset D of tll' is in D(µ4_) if and only if the function is quasi bounded (a
property defined in Section 9). More detailed results on the class of harmonic
functions in D(µo_) for D a ball were obtained in Section 11.14, and these
results will be extended to h-harmonic functions on a ball in Section 12, to
h-harmonic functions on a Greenian set in Section XII.9.

4. The Class LP(pD_) (p >_ I )


This class is the linear class of extended real-valued Borel measurable func-
tions u on D for which if is in D and if B is an increasing sequence of open
relatively compact subsets of D with union D, then sup._,, JulP) < + eo.
It follows forp > 1 that LP(µ4_) c D(µD_), because the function s'-4s' is
a uniform integrability test function. A Borel measurable function u on D
is in LP(µ4_) if and only if, for each point 1; in D and each open relatively
compact subset B of D with in B, µB( , uI P) < + co and there is a compact
subset A = A, of D for which

sup {µB(1;, Jul"): A c B, Ban open relatively compact subset of D} < + eo


(4.1)

When D, h, and p are specified and there is no danger of confusion, we


shall sometimes write LP instead of LP(p4_ ). A function u is in LP(µ4_) if
and only if uh is in LP(po_). As in Section 3, the case of most interest in
potential theory is that in which Jul is h-subharmonic, in particular, if u is
h-harmonic. The following theorem is in part a specialization of Theorem 3.

Theorem. If u is a Bore! measurable real-valued function on the connected


Greenian set D and if Jul is h-subharmonic, the following conditions are
equivalent :
(a) ueL1'(µ4_).
(b) The family (2.2) of functions coupled with the indicated measures is
LP bounded for every point l: in D.
(c) Condition (b) is satisfied for a single point .
5. The Lattices (S2, 5) and (S', 5) 145

(d) The function IuIP has an h-harmonic majorant.


(e) (If u is h-harmonic) u = ul - u2, with u; positive, h-harmonic, and in
LP(1UD-)
If u satisfies (a)-(d), then LMolul e L'(4_), and in fact

LMD(LMDIuI)P = LMDIuIP.

(4.2)

The proof of this theorem is left to the reader because the theorem follows
easily from the discussion in Section 2. Observe that part (e) of the present
theorem is slightly stronger than Theorem 3(e). In fact, in Theorem 3(e)
it is not asserted that if u = u! - u2 and if d>(u1) and'(u2) have h-harmonic
majorants, then c(Iul) has an h harmonic majorant, although the counter-
part of this assertion is contained in Theorem 4(e) with ((s) = sP. How-
ever, for 0(s) = sP and p z 1 this assertion is true because then (ut + u2)P 5
2P-'(ui + u2P).
In Theorem 4, as in Theorem 3, the condition (b) involves a set B in (2.2)
increasing to D, but as in Theorem 3, it is sufficient if the set B runs through
a nested sequence with union D as described in Section 2. If D is a ball, it
is natural to choose B to increase through balls concentric with B so that if
D = B(0, S), the subharmonic function Jul on D is in LP(z _) if and only if
sup,L(I u I P, 0, r) = 1im,t,,L(l u I P, 0, r) < + oo. The class of harmonic functions
on B(0, S) in the class L' was discussed in Section 11. 14.

EXAMPLE. Let u be a harmonic function on the Greenian set D. Then the func-
tion u2 is subharmonic with associated Riesz measure d A = (Au2/trN) dIN,
so that A(D) is a multiple of the Dirichlet integral of u,

).(D)=2 f Igradul2d/Nlx,,S +oo.


D

We now prove that ueL2(1LD_) if and only if GD). is superharmonic. In the


one direction if u e L2(JD_), that is, if LMDU2 = v exists, then GMD(v - u2) =
0; so v - u2 is a potential and necessarily v - u2 = GD2 by direct calculation
of Due. Conversely, if GDA is superharmonic, the function u2 + GD). is a
majorant of u2 and is harmonic because i(u2 + GD).) = 0; so uEL2(AD_).
In particular, if the Dirichlet integral of u is finite, that is, if ).(D) < + oo,
then GD). is superharmonic; so ueL2(pD_).

5. The Lattices (St, <) and (S+, <-)

(See the corresponding martingale theory lattices in Section 2.V.5.)


The Lattice (S, <). Denote in this way the class of h-superharmonic
functions on D with positive h-superharmonic majorants, ordered by point-
146 1.1X. Lattices and Related Classes of Functions

wise inequality in the notation 2-, A, v. If r c Si and if r has an


h-superharmonic minorant, then according to the Fundamental Conver-
gence Theorem (Section VI.1) as trivially adapted to h-superharmonic
functions, the lower semicontinuous smoothing of the pointwise infimum
of r is h-superharmonic, and this function is obviously the (St, <) infimum
A F. If r c St and if r has an h-superharmonic majorant, then V r exists
and is the (St, <) infimum of the class of h-superharmonic majorants of F.
Thus (Si, 5) is a conditionally complete lattice. The pointwise order will
sometimes be called the essential order to match the essential order of
stochastic process theory defined in Section 2.1.8.
Let r be a subset of St. According to the Fundamental Convergence
Theorem, if r has an h-superharmonic minorant, some countable subset of
r has the same (S', 5) infimum as F. If r has an h-superharmonic majorant,
some countable subset of r has the same (S±, 5) supremum as r. In fact,
if IF is directed upward, v r is the pointwise supremum, and the assertion
for pointwise suprema was proved in Section 11.4; if r is not directed upward,
apply the result in the directed case to the set of (Si, 5) suprema of finite
subsets of r.
The Lattice (S+, <). The sublattice (S+, 5) of (Si, 5) is the class of
positive h-superharmonic functions on D in the pointwise order.

6. The Vector Lattice (S, -<)


(See the corresponding martingale theory lattice in Section 2.V.6.) The set
S+ has the unique subtraction property, that is, for u,, u2, u2 in S+ the
equality u, + u2 = U1 + u'2 implies that u2 = uz. This property, trivial for
finite-valued functions, is true in the general case because u2 = u'2 on the
set of finiteness of u, ; so u2 = uz quasi everywhere and therefore everywhere.
Thus S+ is a cone as defined in Appendix 111.3 and therefore determines a
specific order on itself. The specific order symbols will be ::, Y, A,
and S+ in the specific order will be denoted by (S+, ). Define S = S+ - S+
so that each member of S can be identified with a function u, - u2 with u;
in S+; this difference is well defined off the polar set of common infinities
of u, and u2. Order S by the specific order with positive cone S+ (Appendix
111.4) to obtain a partially ordered vector space (S,

Theorem. (a) The space (S, is a conditionally complete rector lattice.


In (b)-(d) let r be a subset of S with a specific order majorant.
(b) Y r is the specific order supremum of a countable subset of F.
(c) If r' is the class of specific order majorants of r, then V r P.
(d) If r is directed upward in the specific order, then Yr = V F.

The duals of (b)-(d), involving specific order infima, are obtained by


replacing r by - r. Since r' is directed downward (specific order) in (c),
the dual of (d) implies that A F' = Ar' = Yr.
6. The Vector Lattice (S, <) 147

It will be convenient in the following proof to use a generalized reduction :


the function v in R' will not be required to be superharmonic on the specified
Greenian set D but may be an arbitrary extended positive-valued function;
no other change is made in the reduction definition. In the applications to
be made, A = D and it will be obvious that v has a positive superharmonic
majorant on D so that R° will exist and be superharmonic on D. In the
following proof r and r' are as defined in the statement of the theorem.
Proof that A r' e r' if r (-_ S+. If u is in r and u' is in r', there is a function
v in S+ for which u + v = u'. Let u' run through a sequence in r' with
essential order infimum A r' to find that u < A F.
Proof that A r' = Ar' = Yr if r c S+. Let u' be in r', and define

+00 if u'() = + ao,


(u' -A if

u' S A r' + RD quasi everywhere on D and therefore everywhere on


D. According to the natural order decomposition theorem, there are func-
tions v and v, in S' such that v 5 A r', v, 5 Rm, and u' = v + v,. But then
v, quasi everywhere on D; so v, z Rm, and there must be equality; so

u'=v+RD. (6.1)

Now let u be a member of r. Since u' and A r' are in r', there are members
v2 and v3 of S+ such that

U' = U + V21 Ar'=u+v3. (6.2)

Then v2 < v3 + Re ; so repeating an argument just used, there is a member


w of S+ such that

v2=w+Rm. (6.3)

Then (6.2) implies that u' = u + w + RD; so v = u + w from (6.1). If we fix


u', the function v is also fixed; so u < v for all u, and it follows that v e r.
Hence v z A r', and there must be equality because by its definition, v satis-
fies the reverse inequality. Thus u' = Ar' + Rm from (6.1), and since u' is an
arbitrary member of r', it follows that A r' < r'. Since we have already
proved that A F c- r', we now deduce that A r' = R r; so A r' = Y r by
definition of r.

Proof of (a). The fact that Yr = kr' exists shows that (S+, <) is a condi-
tionally complete lattice and therefore that (S, <) is a conditionally complete
vector lattice. o
148 LIX. Lattices and Related Classes of Functions

Proof of (b) and (d). In proving (b) and (d) we can assume that r is directed
upward in the specific order, at the possible expense of replacing r by the
set of specific order suprema of finite subsets of r. We can then also assume
that r c S+, at the possible expense of choosing some member uo of r and
then replacing r by {u - uo : u e r, u uo}. Under these hypotheses r is also
directed upward and bounded in the essential order so that yr = u' exists
(Section 5) and is the pointwise supremum of r and in fact is the pointwise
supremum of a sequence u, in r. Choose any member u of r and choose
vo, v 1, ... successively in r to satisfy vo = u, (v Y u z 0.
The sequence v, is a specific order increasing sequence with pointwise limit
u'. There is a member w,,, of S+ such that u + wi1 = v,,, and if v' a r,, there
is a member of S' such that v., + W.2 = v'. Hence

u+limwi1=u',
n-00
u'+(limwi2)
_'O
=v

so r u' f", and we conclude that u' = Yr = Yo u,,. Thus Theorem 6(b)
and (d) are true.

Proof of (c). Let u' be in r"' and define

+00 if u oo,
(u' - V T)O if U '(D < + oo.
Then u' = v t + V r + RD quasi everywhere on D, and therefore
u' S v r + RD everywhere on D; so repeating the reasoning already used
twice above, we find that there is a function w in S+ such that

w5 Vl-, u'=w+RD. (6.4)

Furthermore by definition of r, if u is in r, there is a function w, in S+


such that u' = u + w,. Then w, 5 vp quasi everywhere on D; so w1 5 R°
and
u+RD<U+w1
It follows that u 5 w for all u in r; so v r S w, and by (6.4) there must be
equality. The second equality in (6.4) thus becomes u' = v r + R+ ; so
VF u', as was to be proved. a

7. The Vector Lattice


(See the corresponding martingale theory vector lattice in Section 2.V.7.) An
h-superharmonic function specific order majorized by an h-harmonic func-
tion is itself h-harmonic and if r is a set of positive h-harmonic functions
8. The Vector Lattices, 149

with Yr = u, then u is a specific order majorant of each member of r; so u


is positive and h-superharmonic, and GMDu is a specific order majorant of
I-. Hence u = GMDu and is h-harmonic. It follows that if S' is the cone of
positive h-harmonic functions on D and if S. = S - S,' , then S. is a band
in S. The essential and specific orders coincide on S,,. If r c 5,,,

YF = LMDr, AI' = GMDr

in the sense that if one side of an equation exists, the other side exists and
there is equality.
If D is a ball, it was shown in Section [1.14 by means of the Riesz-Herglotz
representation theorem that S. is lattice isomorphic to the conditionally
complete vector lattice of finite signed measures on 8D. For Greenian D a
corresponding result will be proved in Section XII.9 by means of the Martin
boundary and the Martin representation theorem.
According to Section 4, an h-harmonic function is in S. if and only if the
function is in L'(pD_).

8. The Vector Lattice SP

(See the corresponding martingale theory vector lattice in Section 2.V.8.) An


element of S' specific order majorized by an h-superharmonic h-potential
GDu/h of a measure is itself such a potential. If r is a set of such potentials
with Y I = u, then u is a specific order majorant of each element of r ; so u
is a positive h-superharmonic function, and linearity of the GMD operation
implies that the h-potential u-GMDu is also a specific order majorant of t
and therefore must be u; that is, u is an h-potential. The class Sp is therefore
a cone satisfying the conditions (Appendix 111.8) implying that the set
So=SP -Sp isabandinS.
S. and the Corresponding Lattice of Charges. Let M' be the set of measures
on D and let M = M' - M', the set of charges on D. The set M ordered by
the positive cone M' is a conditionally complete vector lattice. Let Ma be
the set of measures in M' whose h-potentials are h-superharmonic and
define MP = My - MD M. Then MP is a conditionally complete vector lattice,
a sublattice of M. The map p i-4 GDµ/h is a one-to-one order-preserving map
from MD onto SP inducing a one-to-one linear order-preserving map from
MP onto Sp.
SP = S.. In fact, on the one hand, if u e SD and v e S,; , then u A v = 0
because this lattice infunum is h-harmonic as a specific order minorant of
the h-harmonic function v and is a minorant in both specific and pointwise
orders of the h-potential u. Hence SP c S ; so SP c S. Conversely, if
u e (S;)', then the two relations GMDu E S ; ;and GMDu S u imply that
150 I.IX. Lattices and Related Classes of Functions

0=uAGMpu=GMDu;
that is, u e Sa . It follows that S' c SD, and so there is equality, as was to be
proved.

9. The Vector Lattice Sqb


(See the corresponding martingale theory vector lattice in Section 2.V.9.)
The class Sqb is the class of elements u in S' which satisfy the following
equivalent conditions:
(a)The function u is the specific order supremum of a set of bounded
elements of S'.
(b) The function u is the limit of a specific order increasing sequence of
bounded elements of S' ; that is, u is the sum of a series of bounded
elements of S'.
Observe that if "specific order" in (a) were replaced by "essential order,"
the resulting class of elements u would be S' itself. The class S, , is a cone
satisfying the conditions (Appendix 111.8) implying that the set SQb =
Sq, - Sqb is a band in S; the members of this band are called quasi bounded.
The bands Smgb = S. n Sqb and SDQb = Si, n SQb. In view of Section 8 these
two bands are orthogonal and SQb = S,,,Qb + SpQb. The band S,,,Qb is the band
in S. generated by the constant function 1. It will be shown in Section 3.1.9
that an h-superharmonic potential GDµ/h is quasi bounded if and only if u
vanishes on polar sets, equivalently, if and only if ueD(µo_).

EXAMPLE (Quasi Bounded h-harmonic Functions and Dirichlet Solutions).


If D is provided with a boundary aD by a metric compactification, the PWB'
solutions for this boundary are quasi bounded. To see this, observe that if
u is a positive PWB' solution, say u = Hf, then the relation

u = Jim Hf,,,,= limp

exhibits u as the limit of a specific order increasing sequence of positive


bounded h-harmonic functions. Conversely, if the boundary is internally
h-resolutive, that is (Section VIII.2), if every bounded h-harmonic func-
tion u is a PW B" solution, u = H f' = f ), then every quasi-bounded
h-harmonic function on D is a PWB" solution. It will be shown in Section
3.1.5 that S,,,Qb = S. n D(µo_). This class includes the PWBh solutions what-
ever the choice of aD, and the inclusion becomes equality when the boundary
is internally h-resolutive. We have already proved in Section VIII.9 that if
D is a ball, its Euclidean boundary is universally resolutive and is internally
resolutive, and we remarked there that the argument will be generalized in
10. The Vector Lattice S, 151

Section 12 of the present chapter to show that this boundary is universally


internally resolutive.
Roughly, a boundary is internally h-resolutive if it has enough points and
is h-resolutive if it does not have too many. For example (h = 1), if D is a disk
less a radius, the Euclidean boundary is resolutive (Theorem VIII.4) but is
not internally resolutive because a bounded harmonic function on D with a
limit at every boundary point except that there are different limits at the two
sides of the deleted radius is not a PWB solution. Adding boundary points by
ramification to separate the two sides of the deleted radius makes the boun-
dary internally resolutive. On the other hand, if h is the minimal harmonic
function corresponding to a boundary point C of the disk and if the disk
boundary is ramified so that C explodes into a set containing more than one
point, the new boundary of the disk is h-internally resolutive because quasi-
bounded h-harmonic functions are identically constant, but the new disk
boundary may not be h-resolutive.

10. The Vector Lattice S,,

(See the corresponding martingale theory vector lattice in Section 2.V.10.)


Define S, = SQS, the class of singular elements of S. and define
5,,,,=S.nSs, Sp,=SP nSs, Ss S,nS+,
Sn S S,,,,, SP, are bands, S,,,, 1 So and S, = S,,,, + S1,,. A
function u in S+ is singular if and only if 0 i. the only bounded function in
S+ which is a specific order minorant of u.
S,,,,. It will be shown in Section XII.9 that an h-harmonic function u = v/h
is singular if and only if in the Martin representation (Section XII.9) of v and
h, which generalizes the Riesz-Herglotz representation (Section 11.14), the
representing measure M is singular relative to the representing measure M..
We prove now that for a function u in Sm to be singular, it is sufficient that
u A c e So for some strictly positive constant c and necessary that u A C e Sp
for every strictly positive constant c. If U E S,,, then GM' (u A c) = u A c = 0
for every positive c because u A c is a bounded positive specific order minor-
ant of u. Hence u is an h-potential. Conversely, if u E Sm , if U A c is an h-
potential for some strictly positive c, and if v is a bounded member of S'
with supDV = a, then

(u)Av=GM[(u) n V] ,, aGMD(u n c) = 0.

so (a/c)u 1 v, and therefore u 1 V. Hence uE S,',.


S,, . It will be shown in Section 3.1.10 that an h-superharmonic h-potential
GDp/h is singular if and only if p is supported by a polar set.
152 I. IX. Lattices and Related Classes of Functions

11. A Refinement of the Riesz Decomposition


The Riesz decomposition of a positive superharmonic function can be
expressed in vector lattice language in the form S. 1 SP, S = S. + Sp. We
have now found a refinement of this decomposition: the four bands Smgb,
S", Spqb, Sp, are orthogonal and

S. = Smgb + Smp, Sp = Spgb + Sps, S = S,,b + Sms + SP,b + S,-

Thus if u e S and if umgb, ... are the respective projections of u on the bands
Smgb, ... , we have derived a unique decomposition

U = Umgb + U,,,, + Upgb + Ups,

in which all functions on the right are positive in the specific order if u is a
positive h-superharmonic function.

12. Lattices of h-Harmonic Functions on a Ball


Let B be a ball and let h be a strictly positive harmonic function on B. If
u = v/h is an h-harmonic function on B, then ueS. = S. n L'(µ4_) if
and only if v is the difference between two positive harmonic functions,
that is, if and only if there is a Riesz-Herglotz measure M. on aB (Euclidean
boundary) such that v = PI(B, M ). It follows from Theorem 11. 14 that the
map u - PI(B, is a linear one-to-one order-preserving map from
S. onto the vector lattice of signed measures on B. The classes S,,,ab, Sm
D(µ8_ ), L"(µ4_) of h-harmonic functions will now be described in terms of
the corresponding classes of Riesz-Herglotz measures.
The Classes Smgb and Sm,. In the map u = v/h i-. MM the constant h-
harmonic function I corresponds to Mb. The band Smgb generated by the
constant function I therefore corresponds to the band generated by Mh,
which is (Appendix IV.8) the class of signed measures absolutely continuous
relative to Mh. Thus uc-Smgb if and only if u = PI(B,OdM,J/h for some
Mb measurable and integrable function 0 (= dM/dMb), that is, if and only
if (Section VIII.9) u is the PWB" solution for some boundary function 0.
According to Theorem I1.15, the function 0 is the nontangential boundary
limit function of u. The band of singular h-harmonic functions corresponds
to the band of signed measures on aB lattice orthogonal to Mb, that is
(Appendix IV.8), the class of signed measures singular relative to Ma.

The Classes L' (µ8_) and D(µ8_) : Theorem 11.14 for General h. As already
noted, the L' (µ8_) class of h-harmonic functions is the class Sm. For h - 1
the classes L'(µ8_) and D(µ8_) of h-harmonic functions were characterized
in Theorem 11.14. For general h this theorem becomes the following.
12. Lattices of h-Harmomc Functions on a Ball 153

Theorem. Let u = v/h be an h-harmonic function on the ball B = B(0, S).


(a) [L1(1A4_) h-harmonic functions.] The following conditions on u are
equivalent:
(al) u = PI(B, for some signed measure M. on OB.
(a2) u is the difference between two positive h-harmonic functions.
(a3) Jul has an h-harmonic majorant.
(a4) sup,<eµa(O,)(O,JuJ) = sup,<aL(julh,0,r)/h < +oo.
Furthermore the map u i-* M is a one-to-one linear order preserving
map from the L1(µ4_) class of h-harmonic functions onto the class
of signed measures on OB.
(b) [D(µ8_) h-harmonic functions.] The following more restrictive con-
ditions on u are equivalent:
(b1) u = PI(B,f dM,J/h for some M,, measurable and integrable
function f on OB.
(b2) There is a uniform integrability test function 0 for which
0(lul) has an h-harmonic majorant.
(b3) The family {ulaa(o..),PB(o.,)(O, -), 0 < r < S } of paired functions
and measures is uniformly integrable.
(b4) u e
Furthermore the map is a one-to-one linear order preserving
map from the class of D(µ4_) h-harmonic functions on B onto the
class L1(aB, Mb), and dM = f dMb.

The proof follows that of Theorem I1.14 and is omitted. Parts of the
theorem merely repeat results already discussed at the beginning of this
section as implications of Theorem 1I.14.
Universal Internal Resolutivity of eB. We have already shown in Section
VIII.9 that aB is universally resolutive and that the class of PWB" solutions
is the class of h-harmonic functions given by (bl) above, which is the class
of D(µ4_) h-harmonic functions. Since this class includes the bounded
h-harmonic functions [criterion (b2) above], it follows that dB is universally
internally resolutive.
The Class L'(µ4_) for p > 1. We now show that ue S. n LP(µ4_) if and
only if u = PI(B, ¢dM,,)/h for some function 0 in L°(OB, Mb). Observe
first that if u has this form, then

PI(B, 101 PdM,,)


lulP <
h

by Jensen's inequality; so JuIP is majorized by an h-harmonic function and


therefore is in L"(µ4_). Conversely, if uESm n LP(µ8_) for p > 1, it follows
that ue D(µ4_); so u = v/h with dM, = u = PI(B,f dMh)/h, and it
will now be proved that f,PP is M integrable. It can be assumed that f
154 I. IX. Lattices and Related Classes of Functions

is Borel measurable. According to Theorem 1I.14, if f is a continuous func-


tion on OB, then

lim f f fdMM = f ff dM,,; (12.1)


rT6
L N JOB J@B

so if I /p + 1/q = 1, and if v,/h is an h-harmonic majorant of lulP, the absolute


value of the integral average on the left in (12.1) for fixed r is at most

flq)ah() 1N (N ,(d)l tro (12.2)


L s
in ( I lrNSN-1 J 8B ' bJ 7CNSN-I J

and the second factor in (12.2) is v,(0) "; so


11q
dM,,1 < v, (0)'/P (12.3)
Jdsff LJdB If i9 dMs]

Since this inequality is true for continuous f, it is true for f bounded and
Borel measurable. Substitute in (12.3) the choice f = l f l"-' sgn f , where
l fvl < n and f = 0 elsewhere to find

l f.l P dM,, S v, (0),

which imples that l f lP is M,, integrable, as was to be proved.


It is instructive to attempt an alternative proof of this integrability, based
on h-harmonic measure. In the first place

P8(o.r,(0, lul Ph) ti h


= PB(o."(0, lulp) 5 40"'(0, ud = ut(0) (12.4)
h(0)

if u, is an h-harmonic majorant of lul". Next try to put this inequality in a


context in which Fatou's lemma is applicable when r - 5. For example, if
h = 1, (12.4) implies
P /x- 1 (dq)
u(r1) < u,(0), (12.5)
1. JtK hN-1

and therefore Fatou's lemma yields the desired integrability of the


1 _, almost everywhere limit lim,_d l u(rry/v)lP. This application of Fatou's
lemma depends on the fact that when h - 1, the h-harmonic average
µ8,o.,,(0, l ul P) can be written as an integral average over a measure space
that does not depend on r. It will be seen that such a representation of this
harmonic average is possible in a probabilistic approach (Section 2.[X.13)
even without the hypothesis that B is a ball.
Chapter X

The Sweeping Operation

1. Sweeping Context and Terminology

Throughout this chapter D is a Greenian subset of 68", coupled with a


boundary dD provided by a metric compactification when a boundary is
relevant, and the boundary of a subset of D u 8D is that relative to the
compactification. In most of the discussion the nature of the boundary is
irrelevant, and 8D can be taken, for example, as the Euclidean boundary
or one-point boundary of D.
If A c D and if u = GDµ is the superharmonic potential of a measure
µ on D, the function OuD" is a positive superharmonic function majorized
by u and is therefore also the potential of a measure; this measure will be
denoted by OµQ". The map µHOeD" is called the balayage, that is, the sweep-
ing, of µ onto A, and DµDA is called the swept measure or, with an unfortunate
grammatical ambiguity, the sweeping of µ. The reduction operation prop-
erties lead to sweeping operation properties, some of which will now be
listed. The notation is as above: p is a measure on D with a superharmonic
potential, and all sets on which µ is swept are subsets of D.
(a) GDOµO" 5 GDµ with equality quasi everywhere on A [cf. Section
VI.3(b)].
(b) D PO", = Dµ D"2 if Al differs from A2 by a polar set [cf. Section VI.3(c)].
In particular, if A is polar, Dµ0" = 0.
(c) If A c B,

[cf. Section VI.3(h)]. In particular, the sweeping operation is


idempotent.
(d) OpD" is supported by in D [cf. Section VI.3(a)].
(e) If A is a Borel support of y and if GDµ < + o0 on A, then y =Dµ8",
equivalently, GDµ = GDDPD". In fact, on the one hand, GDOPO" 5 GDµ
by (a), and on the other hand, GDµ = GDDPD" < + o0 on A, a support
for p; so (by the domination principle) GDµ 5 GDDPOA.
156 I.X The Sweeping Operation

(f) If B is a µ null open subset of A, then 0µ0"(B) = 0 [cf. Section


VI.3(a)]. In particular, strengthening (d), if A is µ null the swept
measure 0µQ" is supported by D n 8A.
(g) If A is a Borel subset of D, the measure p can be written as the sum
µt + µ2 + µ3 of the projections of p, respectively, on the sets A n
{GDP < + oo }, A n {GDP = + oo}, D - A.
In view of the reduction additivity property of Section VI.3(f), 0µ0A _
0µt0'' + bµ26" + 0µ30A. According to (d)-(f), the sweeping operation of
µ onto A leaves p, invariant and sweeps 113 into a measure supported by
D n aA. The measure 0µ20A is supported by A n D. More detailed informa-
tion on 0µ 0A will be furnished by Theorems 5 and XI. 18. In the most common
classical application of sweeping, A = D - B with B a ball with closure in
D. In this case the fact that the operation GDuI-.+OGDPOD_8

= te[GDp]
transfers the component of µ on B to a measure supported by the set
aB = D n a(D - B) is particularly clear.

Characterization of 0µ0A (Refined in Section XI. 15)

If A is a subset of D, closed in D, and if µ is a measure on D with a finite-


valued potential GDµ, the swept measure v = 0µ0A is uniquely determined
by the following properties:

(SI) v is supported by A.
(S2) GDV = GDµ quasi everywhere on A.

These properties of 0µ0A have already been noted. Conversely, if a measure


v has these properties, then v = 0µ0A by the domination principle because
GDV = GDOpOA quasi everywhere on a common support A of v and OpOA,
and both potentials are finite valued on A.

The Sweeping Kernel 80

If E D, the probability measure on D supported by will be denoted by


The sweeping of this measure onto A, denoted by do is deter-
mined by its defining equation

6GD( , _) 6A(,) = G 6A(, GD(1, (, E D). (1.2)


Jl)

In the following we shall write GD(, q) for OGD(, )DA(n). It will be shown
in Section 3 that Go is symmetric. The proof that Bo (, B) is Borel measurable
2. Relation between Harmonic Measure and the Sweeping Kernel 157

when B is a Borel set, so that So is a kernel, and the proof that GD is Borel
measurable on D x D will be given in Section 4 and so cannot be used at
the present stage. If B is open and relatively compact in D and if v = fl 1 IB
then v is a potential, say v = GDV, and integration with respect to v(dq)
in (1.2) yields 1 z So v) z 6D A(-, B). It follows that Eo D) S 1. It will
be seen in Section 5 [see (5.1')] that So D) = Q 11A for every set A. Ac-
cording to property (f) above, if is not an interior point of A, the swept
measure Sp is supported by D n A. On the other hand, if is an interior
point of A then Go according to Section VII.3(b), and there-
fore bE(l;, ) = SD(t, ); that is, {c)) = 1.

2. Relation between Harmonic Measure and the Sweeping


Kernel

Theorem. If D is a Greenian subset of R"', if A is a subset of D, closed relative


to D, and if E D - A, then bo( , ) = on the Borel subsets of
Dn8A.

According to (1.2) and the relation between reductions and PWB solutions
derived in Section VIII.10,

PD-A(17, ID) = d"D( , GD(q, ')) [(4, q) e D x (D - A)]. (2.1)

The harmonic measure symmetry property VIII(18.5), in which D and B


correspond respectively to D - A and D here, implies that (2.1) with both
and q in D - A can be written in the form

I^D-A(b, GD(q.') ID) = bD( . GD(q, ')) [( , q) a (D - A) x (D - A)]. (2.2)

Moreover, in the present context xVIII(/1y8.2') becomes

GD(4, 0 = PD-A(b, GD(g, ')1D) ( a D) (2.3)

whenever C is a regular boundary point of D - A in D or is an inner point


of A; so (2.3) is true for quasi every { in A. Now for fixed in D - A each
side of (2.2) is the value at q of the GD potential of a measure on D. These
potentials are equal on D - A according to (2.2) and are equal quasi every-
where on A according to (2.3). Hence these potentials are equal on D;
so the measures (restricted to subsets of D) and bAV, ) generating
the potentials are the same, as was to be proved.
158 I. X. The Sweeping Operation

3. Sweeping Symmetry Theorem


Theorem. If D is a Greenian subset of R" and if A is a subset of D, then

q)eD x D]; (3.1)

that is Gp is symmetric on D x D.

This fundamental theorem is a deep generalization of the theorem that


the Green function GD is symmetric, and reduces to the latter theorem when
A = D. Suppose f i r s t that A is closed in D. For and q in D - A equation (3.1)
is equivalent to the assertion that the function GD(q, )) is
symmetric on (D - A) x (D - A), a fact noted in Section VIII.18 and just
applied in Section 2. If at least one of the points , q is in the set A, then (3.1)
is true for the pair q) when A is replaced by

A. = A - I /n) u B(tt, l/n)],

and therefore (n -- oo) equation (3.1) is true when A is replaced by


A- u {q}], in view of the reduction property Section VI.3(e). Finally,
(3.1) is true as stated for A closed in D since [Section VI.3(c)] smoothed
reductions on A are identical with smoothed reductions on A less two
points. Apply Section VI.3(e) again to find that (3.1) is true if A is an F.
set, in particular, if A is open. For an arbitrary set A it follows that (3.1)
is true when A is replaced by an arbitrary open superset of A and there-
fore [Section VI. 3(m)] that (3.1) is true for unsmoothed reductions if neither
nor q is in A. Since [Section VI.3(b)] reductions on A are equal on D - A
to their smoothings, (3.1) is true as stated if neither nor q is in A. Finally,
if at least one of these points is in A, apply (3.1) to A less these two points.
The smoothed reductions are unchanged, and we deduce that (3.1) is true
for the pair q)

4. Kernel Property of SD
Lemma. bD is a kernel.

All that needs to be proved is that B) is Borel measurable whenever


B is a Borel set. Consider the linear class r of functions of the form f = u - v,
where u and v are bounded continuous potentials on D, the measures
associated with u and v have compact support, and f has compact support.
Define f" = lul" - 6vl", and observe that f" is determined by f because
if f has two representations, f = u - v = y' - v', then u + v' = u' + v; so
finallyDu1"-IvQ"=lul"-Bv'I".If eD
4. Kernel Property of do 159

the Functional f i- Lt(f) = f is linear on r. Let r be the class of uniform


limits of sequences in r, so that (by Lemma VII.9) I' includes the class of
continuous functions on D with compact support. Iffe t with approximating
sequence f, in r, then [from Section VI.3(n)]

sDptL4(fn-fnJI sup If. -fml; (4.1)

so the sequence L,(f,) is convergent. Since two approximating sequences for


f can be combined into one, the limit of Lf(f,) is independent of the approx-
imating sequence and is when fe r. Define LS(f) = lim, LS(fn)
to get a linear functional on r. The functional Lr is positive because if f in
r is positive and if f = u - v is an approximating sequence in r for f,
with infD(un - vn) = -en, it follows that un + E. Z vn and therefore that
OunOA + en Z llvnOA; so

L4(f) = lim L4(fn) >_ - lim en = 0. (4.2)


n-ao n-W

Apply Section VI.3(n) again to obtain, for f in t with approximating


sequence f = u, - v,,

L.(f)I = lim IL.(f)I = urn IIunDA - kVjA I s lim supIII = SUP If 1; (4.3)
n-m n-m n-m D D

that is, the linear functional LS has bound 1. Thus (Riesz representation
theorem) there is a measure on D with doA(r;, D) 5 1 for which
L4(f) = ODA(F, f) whenever f is continuous on D and has compact support.
Since this equation shows that f) is Borel measurable, SDA is a kernel.
Now suppose that u = GDp is the potential of a measure p with compact
support S, and let D. be an increasing sequence of open relatively compact
subsets of D with union D, for which S c Do and B. c Dn+t . According to
Theorem IV.10, there is an increasing sequence u, of continuous potentials
with limit u for which u = u on D - Do, and there is an increasing sequence
un of continuous potentials with limit tDnu for which u k = TD.u = u on
D - Dn+t. Then u. - u,. is in r, so
A A
(4.4)

Moreover u is a potential; so (from Section V111. 11) limn-m TDnu = 0, and


therefore

lira sup ou,,,,8 A 5 Jim sup U. 5 lim TDnu = 0. (4.5)


n-W n-M n-m

Hence (4.4) yields

8uj" = 6DA(',u) (4.6)


160 1.X. The Sweeping operation

In particular, when u = we find with the help of sweeping symmetry


that 6DA satisfies the defining equation for bD ; so SD = SD" and is a kernel.

5. Swept Measures and Functions


Theorem. Let D be a Greenian subset of I8", and let A be a subset of D. Let
u = GDµ be a superharmonic potential, and let v be a positive superharmonic
function on D with associated Riesz measure v. Then

Quo"=GDOruIA=GDp=5D(',u), (5.1)

0v0" = .5D(-,v), (5,1')

JDv4d = fD v dQNQ".
(5.3)
D

In particular, if v is a superharmonic potential, (5.3) can be continued:

f QvQ"dµ = f VA'UQA
= ID
(5.3')
D D ID ID

I n the course of proving Lemma 4 we proved (4.6); that is, the first and
fourth terms of (5. 1) are equal if µ has compact support. Since every positive
superharmonic function is the limit of an increasing sequence of potentials
of measures with compact supports (Section IV. 10), the first and fourth
terms of (5.1) are equal, and (5.1') is true. The first term is (5.1) is equal to
the second by definition of Qµ0". The third term is equal to the fourth
because

6D N' u) = J
D
f D
GDOI, C)µ(dd,

and the iterated integral becomes GDµ on reversal of the order of integration.
Equation (5.2) follows immediately from (5.1); (5.3) and (5.3') follow from
(5.1), (5.1'), (5.2), and the fact that J Qv0"lA = 0vO"

Generalization of Q p Q"

It is convenient to define Qµ0" by (5.2) when µ is a measure on D even when


GDµ is not superharmonic. Under this extended definition, (5.3) remains
true.
6. Some Properties of bn 161

Application to the Representation of a Reduction. Let D be a Greenian sub-


set of R', and let A be a subset of D. Then if v is a positive superharmonic
function on D and if vA is the Riesz measure associated with DvOA (reduction
relative to D),

OvDA=GDVA+h, vA=DvADA, h=GMDOvIA=ihOA. (5.4)

The first equation in (5.4), with the stated identification of h, is simply


the Riesz decomposition of OvOA. Since the smoothed reduction operation
is idempotent, (5.4) leads to

OVDA = GDOVAUA + OhOA = G0A. + h', (5.5)

where A is a measure on D and

h' = GMAh JA :!s; Oh OA :!g h.

A comparison between (5.4) and (5.5) shows that h' = h and A = O VADA = VA,
so that the second and third equation in (5.4) are true.

ExAMPUS. If A is relatively compact in D, then [VIA is a potential; so h = 0.


If A is sufficiently large, say A = D, then D V O A = v, and (5.4) becomes the
Riesz decomposition of v.

6. Some Properties of SD
(a) Subadditivity of 6L. This set function is subadditive in the sense that

bD B 5 bD + Sn (6.1)

for all subsets A and B of D. In fact [Section VI.3(i)]

5DAvB(,, v) Dv0A
+5 A OvD ) = c5 v) + v) (6.2)

whenever v is positive and superharmonic on D; so if f is the difference


between two finite-valued positive superharmonic functions on D, with
f -e 0, it follows that

s' s (6.3)

and this inequality implies (6.1) in view of the approximation application


in Section VII.9.
(b) For each point of D and each subset A of D the measure SD(, )
vanishes on polar sets not containing . In fact, if B is a polar subset of D
162 I. X. The Sweeping Operation

not containing , there is (by Theorem V.2) a positive superharmonic


function on D, finite at but identically + o0 on B, and the inequality

+ 00 > v( ) 2t Dv[ 4( ) = 6D(S, v)

implies that B) = 0.
(c) A zero-one law. Let be a point of D, let A be a subset of D, let B
be a neighborhood of , and let v be a positive superharmonic function on
D. Then either the equalities 6D^1(, I and DvDA^B()
= v() are true
for all B and all v or SD^B(, {}) = 0 for all B, in which case whenever
oo, the function Dv0A11 tends to 0 on D when B shrinks to . The

property that 1 will be given a topological interpretation in

Section XI.3, where incidentally it will also be shown that this property is
independent of the choice of D containing . Thus this property is a local
property of A near . To prove (c), we first observe that, using the notation
of (c), bD^B(
, decreases when B increases because when B, and B2 are
neighborhoods of with B, c B2 subadditivity of bn implies

bD2( , { }) 65"(S, M) + 6 D

and the second term on the right is 0 because the measure SD= Bo(, ) is
supported by the closure of B2 - B, . Next observe that since J V O A B =
DvOA
QA^B
it follows that

D^B(, {}) = JD' {})bB( dq)(6.4)


and since the measure 6'(ry, ) vanishes on polar sets not containing q,
equation (6.4) reduces to

bD^B(, { }) = D( })bD '8( , {l}). (6.5)

When B = D, we conclude that must be either 0 or 1. In the first


case (6.5) implies that SD^B(, {}) = 0 for all B. On the other hand, if
8D(, { }) = 1, then monotoneity of as B varies implies that
8D^B(, { }) = I for all B.
If 1, then IV I A-8g) = SD-B( z , v) = vO. If SD( , { }) = 0,
let B be a sequence of balls of center , relatively compact in D, with radii
tending monotonely to 0. The sequence DvOA^B is a decreasing sequence of
potentials on D with limit a positive function vo, harmonic on D -
The restriction of vo to D - has a positive superharmonic extension v,
to D, necessarily of the form const. and the constant is 0 if as we
suppose from now on v 1() < Hence vo = 0 on
D- so
7. Poles of a Positive Harmonic Function 163

hmOvlA' '(7)=

The last term on the right is 0 by hypothesis when ry = and vanishes at all
other points ry because according to (b) above, the measure c5(,) vanishes
on polar sets not containing q.

7. Poles of a Positive Harmonic Function


Let D be a connected Greenian subset of R", coupled with a boundary
provided by a metric compactification, and let A be a boundary subset. If
h is a positive harmonic function on D, if Ra = ch (as is true, for example,
if h is minimal), and if h * 0, then c is either 0 or 1 because the smoothed
reduction operation is idempotent [Section VI.3(h)]. The boundary point
C is said to be a pole of h if Rti = h. The set of poles of h is obviously closed.
If C is a pole of h, then C is also a pole of every positive harmonic minorant
of h. In fact, if h, is such a minorant, the equality

OhiO"'+Oh -h,O1°=h'°=h

implies that the reduction operations performed on the left have not de-
creased ht or h - h,, so h, = Oh, Out.

Poles of a Minimal Harmonic Function

If h is minimal, if C is a pole of It, and if A is a boundary subset containing


C, then

hzRh - RIo=h;
so R," = h. On the other hand, if C is not a pole of h, there is a positive super-
harmonic function v on D majorizing h near C and strictly less than It at
some point of D. It follows that RtiA 5 v < h at some point of D; so RA 0
if A is a sufficiently small boundary neighborhood of C. Furthermore, if B
is a boundary subset containing no poles of It, then Rh =- 0 because B can
be covered by a countable collection A. of boundary neighborhoods of
nonpoles, with R,"i =- 0; so R':!9 EJR,'J =_ 0. This result implies that there
is at least one pole of a minimal harmonic function h because Rh° = h. It
is left to the reader to check that the boundary is h-resolutive if and only
if h has only one pole and that if C is this pole, the singleton JC) hash-harmonic
measure 1. If It has more than one pole, the indicator function of the set of
poles is h-resolutive, and the set of poles has h-harmonic measure 1.
164 1. X. The Sweeping Operation

8. Relative Harmonic Measure on a Polar Set


Let D be a connected Greenian subset of I8", let A be a compact polar
subset of D, and let h = GDvh be the potential of a not identically zero measure
supported by A. Define Do = D - A and aDo as the boundary of Do relative
to the one-point compactification of D. Since A is nowhere dense, 8Do
consists of A and the Alexandrov boundary point at of D. Let ho be the
restriction of h to Do. Recall (from Section VII. 1) that GDo is the restriction
of GD to Do x Do. We now prove that 8Do is an h-resolutive boundary for
Do, that p {a}) 0, and that for n in A and in Do

h( ). (8.1)

Reductions below not otherwise described are relative to Do. If F is an open


relatively compact subset of D containing A, the set Do - F is a deleted
neighborhood of a relative to Do u aDo, a minorant of ho and is
QhoQDo-F is

harmonic and bounded on F, and QhoI D0-F/ho is in the upper PWB"o class
on Do for the boundary function 1M. Moreover (by Theorem V.5) the
function an extension to D, harmonic on F and majorized by
Qh0QDo-F has

h. When F increases, tending to D, this extension decreases, tending to a


harmonic minorant of the potential h, that is, tending to the zero function.
Thus {a} is an ho-harmonic measure null set. Next we prove that the
evaluation

J Gd , n)v"(dn)
A

implies that if C is a compact subset of A, then

Qhr = JG(.tl)v(dll). (8 .2)

If Bo is the trace on Do of a neighborhood B of C,

QhoOBO( ) = bno( ,

8.3)
= (
IG.. ,q) QBO (S)vh( dn)
fA

If smoothed reductions relative to D are denoted by Q QD, an easy argument


involving the superharmonic function extension theorem (Theorem V.5)
shows that for , in D
8. Relative Harmonic Measure on a Polar Set 165

on B. Then for in B0

DhoOB0( ) = (8.4)
J

Now [by Theorem VII.3(b)] when PI is in C, the integrand is equal to n),


and on the other hand [Section VL3(m)], when q is in A-C and B shrinks
to C, the integrand tends to which vanishes because C is
polar. Thus (8.2) is true, and since Ih0Dd/ho is the upper PWBkQ solution on
Do for the boundary function lc, we have proved that this upper solution
is an additive function of C. Since we can ignore the ho-harmonic measure
null singleton {a}, it follows from Theorem VIII.9 that Do is ho resolutive.
Finally (8.1) is now equivalent to (8.2).
Chapter XI

The Fine Topology

1. Definitions and Basic Properties


The topology of a topological space is the class of open subsets of the space.
If T, and T2 are topologies on a space, T, is said to be finer than T2 (and
then T2 is said to be coarser than T,) if T2 c Ti. For any family of extended
real-valued functions on a space there is a coarsest topology making every
member of the family continuous, namely, the intersection of all the
topologies doing this. The fine topology of classical potential theory is
defined as the coarsest topology on lf8" making continuous every super-
harmonic function on R'. It is easy to verify that the fine and Euclidean
topologies coincide when N = I (see Chapter XIV for classical potential
theory on t8), and we suppose from now on in this chapter that N > 1.
Concepts relative to the fine topology will be distinguished by an "f," for
example, f lim sup, afA. From now on any otherwise unqualified topological
concept will refer to the Euclidean topology. Since the fine topology is
defined intrinsically in terms of superharmonic functions, it is not surprising
that this topology plays a fundamental role in classical potential theory.
If u is a superharmonic function, the function u A c is also, for every
constant c, and it follows that the fine topology is the coarsest topology
making continuous every upper-bounded superharmonic function on 68".
Since is fine continuous and since c) is a ball with center
and an arbitrary radius depending on c, balls are fine open, and therefore
the fine topology is at least as fine as the Euclidean-topology. Since there are
discontinuous superharmonic functions, the fine topology is strictly finer
than the Euclidean topology.
Let D be a nonempty open subset of IZ". Then the restriction to D of the
fine topology is the coarsest topology making continuous every super-
harmonic function with domain D. To prove this, it is sufficient to prove that
if u is superharmonic on D, then u is fine continuous, and it is even sufficient
to prove that u is fine continuous on every open relatively compact subset
Do of D. Now if µ is the measure associated with the restriction to Do of u,
the potential Gp is superharmonic on !fF", differs from u on Do by a contin-
uous (harmonic) function, and is fine continuous on R'. Hence u is fine
continuous on Do, as was to be proved.
1. Definitions and Basic Properties 167

It will sometimes be useful to extend the fine topology to R' v { ao } by


defining a set to be a deleted fine neighborhood of co if after inversion in a
sphere the image of the set is a deleted fine neighborhood of the sphere
center (see Section 5). This criterion will be seen to be unaffected by the
choice of inversion sphere, and in fact it will be seen that under an inversion
the fine topology is invariant.
The fine derived set of a set A, that is, the (fine-closed) set of fine limit
points of A, will be denoted by Af. It will be shown (in Theorem 6) that
Af = (Af)f ; that is, Af is fine perfect.
It is immediate that one base for the class of fine neighborhoods of a
finite point is the class of sets of the form

n {neB: uj(q) < cj},


t

where B is a ball containing , u; is an upper-bounded superharmonic


function on B, vanishing at t, and each constant cj is strictly positive. The
neighborhood (1.1) is fine open. It follows that another base for the class
of fine neighborhoods of is the somewhat more convenient class of sets
of the form

n u;(n) < c}, (1.2)


t

where B is a ball containing , u; is a bounded superharmonic function on


a neighborhood of B, vanishing at , and c is a strictly positive constant.
The neighborhood (1.2) is fine closed and is Euclidean topology compact. It
follows that a point is a fine limit point of a set A if and only if is a fine
limit point of every open superset of A.
A point of R' is a fine limit point of a set A if and only if is a Euclidean
limit point of A and if each superharmonic function u defined on an open
neighborhood D of has as a cluster value at along A. For example,
according to 11(6.1), the complement of an IN null set is fine everywhere
dense.
A subset A of RN is said to be thin at a finite point c if is not a fine
limit point of A, that is, if is not in Af. According to the above remarks
on fine neighborhoods, if A is thin at , there is an open superset of A which
is also thin at . The corresponding remarks for = co are left to the reader.

The Baire Property of the Fine Topology

The intersection of a sequence of fine-open fine dense sets is fine dense. To


prove this assertion, observe that the fine topology has a basis of Euclidean
compact sets so that if B, is a sequence of fine-open fine dense sets and if B
168 1. Xl. The Fine Topology

is a nonempty fine-open set there is a sequence B; of compact sets with non-


empty fine interiors such that Bo c Bo n B and that is a subset of the
fine interior of B;, n B. for all n. Then
W
0#nB, cnB,, B
0 0

so no B. is fine dense.

2. A Thinness Criterion
Theorem. If a set A has finite limit point , then A is thin at S if and only if
there is a superharmonic function u defined on an open neighborhood of 4
such that

liminfu(,) >

A is thin at and if D is a Greenian set containing , there is a potential


u = GDv satisfying (2.1) for which v has compact support and the left side of
(2.1)is+oo.
If is a fine limit point of A and if u is superharmonic on an open neighbor-
hood of , u has fine limit at and is lower semicontinuous; so (2.1) is
impossible. Conversely, if is a limit point of A and if A is thin at , some
fine neighborhood (1.2) of does not meet A except possibly at The
functions u = L' uj and minjs uj are superharmonic on B with minjsn
0, and

u>c+(n- l)minuj
jsn

on B; so

lim inf u(,) z c > 0.


A3rS

To prove the second assertion of the theorem, let u satisfy (2.1), and suppose
first that b is bounded and contained in the domain of u. Then u is bounded
below and can be made positive on D by addition of a suitable constant
without affecting (2.1). Let B, be the ball of center and radius r < 8Dl
and let u, be the projection on B, of the measure associated with u. Then
5 u differs from GDp, by a function harmonic on
B,; so if b (5 + co) is the strictly positive difference between the left and
right sides of (2.1),
3. Conditions That e A' 169

lim infGDµ,(q) = S z S.
Aa4-4

Choose r = r" so small that 2-". The superharmonic potential


GD(100 µ") has value 5 2 at and has limit + oo at along A as desired,
and the associated measure has compact support in D. If b is not bounded
and is in the domain of u, choose a ball B containing , with Bin the domain
of u. Then the potential Gev of some measure v with compact support in B
is finite at and has limit + oo at along A. The potential GDv has the same
property because GDv differs from GBv by a function harmonic on B.
Observation. If N > 2, the set D can be chosen to be l8" in the theorem.
If N = 2, the proof yields a superharmonic potential u = Gv on R" for which
the left side of (2.1) is + co.

Application to Polar Sets and Reductions

A polar set Ao has no fine limit points because according to Theorem V.2,
if Ao is polar and if e Al, there is a superharmonic function defined on an
open neighborhood of , finite at , and identically + oo on the part of A 0
in a deleted neighborhood of ; so (2.1) is satisfied, contradicting the
hypothesis that E Ao. (In the fine topology of R" u { oo } as defined in
Section 5 a polar set cannot have oo as a fine limit point; this assertion is
reduced by an inversion in a sphere to the one just treated.) The converse
result that Al = 0 implies that Ao is polar will be proved in Section 6. The
fact that a polar set has no fine limit point implies that if A is an arbitrary
subset of a Greenian set D, then (reduction relative to D) OVOA = v on Af An D.
In fact this equality is true quasi everywhere on A, that is, everywhere on
A - Ao for some polar set AO; so the fine continuity of superharmonic
functions implies that u = OvQA on (A - Al) = Af

3. Conditions That t e Af
Let D be a Greenian subset of R", let A be a subset of D, let be a point
of D, and let v be a positive superharmonic function on D. All reductions
will be relative to D. We shall write lim81, to mean the limit as B, a neighbor-
hood of , shrinks to , that is, as its diameter tends to 0. Define vo =
lim8J, Ov0A^1, that is, vo is the infimum of the indicated class of smoothed
reductions. The positive function vo is harmonic on D - so (by Theorem
V.5) the restriction of vo to D - has a superharmonic extension to D,
and (by the Riesz decomposition) this extension must have the form
cG R,') + h, where c - 0 and h is a positive harmonic function on D. Since
this superharmonic function is majorized by a potential BvrB for B relatively
compact in D, the function h must vanish identically. The inequality vo 5 u
170 I.M. The Fine Topology

implies that c = 0 if oo and more generally c = 0 if OvO "^B(,) < +00


for some B. According to the following theorem, if v is the Riesz measure
associated with v, then either c = v({ }) or c = 0.

Theorem. If e Af, then


(a) for every neighborhood B of .

(b) 1.

(c) Gii(, ) =
If 0 Af, then
(a') limBl4 Dv0"nB = 0 on D - and the limit is 0 at if +00.
(b') 6A(4, J })=0.
(c') Gp( , ) # GD(4, ) on 0}, the open connected component
of D containing

Observation (1). The fact that when B is a neighborhood of , the relations


e Af and e (A n B)f are equivalent implies that properties (c) and (c) are
true under the stated conditions if A is replaced by A n B, and B need not
appear in (a). Property (a) was phrased using B to contrast with (a'), in
which B is essential.

Observation (2). We have already noted in Section X.6 the fact that the
only possible values of are 0 and 1. We now see that the value is
0 if and only if A is thin at .

Proof. (a)-(c) Suppose that e Af. Then e (A n B)f, and as pointed out
in Section 2, it follows that v(); so (a) is true, and in particular,
GD(,) D; that is (sweeping symmetry), (c) is true. Finally,
(c) ca (b) by definition of bD.
(a')-(c') Suppose that c e D - Af We can suppose that is a (Euclidean
topology) limit point of A, adjoining a countable set with limit point to
A if necessary to achieve this. The adjunction does not change smoothed
reductions on A and does not change 61. Then (by Theorem 2) there is a
positive superharmonic function u on D, finite at but with limit + 00 at
along A. If t > 0 and if B is so small that u > 1/E on A n (B - {%}), it follows
that

SDnB(S, { } ) S (7"D-B(S, D) = O I O (S) 5


cDuDAnB(
) S EU(S) (3.1)

Hence D 0, and it follows from Section X.6(c) that


0, as asserted in Theorem 3(b'). Theorem 3(a') now also follows
from Section X.6(c), and Theorem 3(c') follows from the definition of So.
0
4. An Internal Limit Theorem 171

Application. Let A be a Borel subset of O", and let be a point at which


A is thin. Then if A, = A n t3B(t, r),

lim l"-t (Ar) = 0.


r-.0 rN-t

In fact, in the first place according to Theorem 3 with reductions relative to


D=

lim O I DAr(S) < ltm I oA,8(0.2r)( ) = 0.


r-.o r+0

In the second place the rth smoothed reduction on the left obviously major-
izes the smoothed reduction at relative to r) of the function 1 onto
the boundary set A,; the value of the latter smoothed reduction is IN_t(A,)/
aNr'-t according to 111(4.3).
This application of Theorem 3 implies that a solid cone of revolution in
R" is not thin at its vertex. According to the criterion of Theorem 12 for
Dirichlet problem regularity of a boundary point in terms of the fine topol-
ogy, the nonthinness of a solid cone of revolution at its vertex is equivalent
to the Poincare-Zaremba criterion for Dirichlet problem regularity of a
boundary point (Section VIII.15).
Strengthening of (a'). The following slight strengthening of (a') will be
needed below: if e Ar and if v() < + oo, then

lliim 0 0. (a")

To see this, observe that this limit is limBl, 0 vrr), that according to the
remarks at the beginning of this section, lim84,8vOB = 0 on D - and
that according to (b'), 0.

4. An Internal Limit Theorem


If D is an open subset of R", if v is a positive superharmonic function on
D, and if h is a strictly positive superharmonic function on D, then u = v/h
is defined in the obvious way on the set of points at which at least one of the
functions v, h is finite valued. As so defined, the function u is fine continuous.
The following theorem provides information on the character of u near a
point of the set of common infinities of v and h. Let v and v,, be, respectively,
the Riesz measures associated with v and h. Recall our convention that
dv ldvh is the Radon-Nikodym derivative of the absolutely continuous
component of v with respect to vh. The singular component of v with respect
to vh will be denoted by v.
1 72 I. XI. The Fine Topology

Theorem. (a) The function u has a fine limit u*(C) at quasi every and vp + v,,
almost every point C of D. Moreover u* < + oo quasi everywhere and v,,
almost everywhere on D, and u* = + oo vv almost everywhere on {C: v(C) _
+oo}.
(b) At v,, almost every point C of the set {C: h(C) = + oo}, equivalently, at
v,, almost every point C of an arbitrary polar subset of D,

dvn
u*(C) = (C) (4.1)

(c) In particular, if C E D,

f lim v(q)r = vo({C }) = inf v y (4.2)


7-c GD(n, C) D- c S)

(d) Let C be a point of D, and let F be a subset of D with (Euclidean


topology) limit point C. If F is thin at C, there is a superharmonic potential v
on D for which

v(n) _
+ao. (4.3)
F191 4GD(7,C) -

Conversely, if there is a positive superharmonic function v on D - {C} satisfy-


ing (4.3), then F is thin at C.

Observation. In the first term of (4.2) and in (4.3) GD can be replaced by


G without altering the validity of these equations because 1im,,..t GD(n, C)/
G(n,C) = 1.
Since the theorem is local, we can suppose in the proof that D is connected.

Proof of (c). The special result (4.2), in which h = C) and so v,, is the
unit measure supported by {C}, is so much easier to prove than the more
general case (4.1) that we prove (4.2) separately, as follows. We have already
seen (Section VII.IO) that the last two terms in (4.2) are equal. In proving
(4.2) it can be supposed, replacing v by v - C) if necessary, that
these two terms vanish; under this hypothesis, unless (4.2) is true, there is a
strictly positive number b such that the set B = {n: v(n) > bGD(n, C)} i3 not
thin at C. Apply Theorem 3(c) to find that
1v08
vZ z
contrary to the hypothesis that C) has infimum 0. Hence (4.2) is true.
13
4. An Internal Limit Theorem 173

Proof of (d). If v is a positive superharmonic function on D - {C} and if


(4.3) is true, then since v has a positive superharmonic extension to D
(Theorem V.5), the set Fmust be thin at C, in view of (4.2). Thus the converse
half of (d) is true. In the other direction, if F is thin at C, choose n(96 C) in
D. Then [by Theorem 3(a')] there is a decreasing sequence B. of neighbor-
hoods of C. shrinking to C, with
G
C) < 2-".

There is therefore a positive superharmonic function v on D, majorizing


C) on F n B. with v (q) < 2-". After replacing v by V. A 0) if
necessary, we can suppose that v is a potential on D. The potential v = Eo v
is at least (n + 1) on Fn B. and therefore satisfies (4.3).

Proof of (a). If 0 5 a < b, define

A = S B = { : Z bh(p)}. (4.4)

According to Section V1.3(o),

V A bh
hAB + hABAB + ... G (4.5)
b-a
Now
GD(V,.bA
hA = U' VA Z UGDVkUA -

and bp( , { }) = 1 when e Af n D; so if VhA is the projection of vh on Af n D,


it follows that hA GDVhA. Similarly

hAB Z OGDVhAOB - GD(VhAaD)+

so if vhc is the projection of vh on C = Af n Bf n D, that is, if vhc is the projec-


tion of vhA on Bf n D, it follows that hAB Z GDVhc. Similarly all the other
summands in (4.5) majorize GDvhc, and we conclude that vhc - 0. Since every
point of D at which u* does not exist is in C for some rational pair a and b,
we conclude that u* exists vh almost everywhere. Apply this result to h/v to
show that u* exists v almost everywhere. As noted at the beginning of this
section, the function u is fine continuous at every point at which it is defined ;
that is, u* exists and is equal to u quasi everywhere on D. Moreover u* < + 00
quasi everywhere on D because u* = u = + oo can be true only on the polar
infinity set of v. The rest of the last sentence of (a) will be proved in the
course of proving (b).
174 1.X1. The Fine Topology

Proof of (b). We show first that if F is a Borel polar subset of D and if


v,(F) = 0, then u* = 0 v,, almost everywhere on F. It is sufficient to prove
this for F compact. Under this hypothesis on F, we can assume that v and
h are potentials because if v and h are replaced by their reductions on some
open neighborhood of F, relatively compact in D, then v, h, v,,, and va are
unchanged on this neighborhood, and v and h become potentials. Define B
by (4.4). Then (reductions relative to D)

v > 6vD8 z bDhOB = bGD(VhSD)

Since I when e Bf, we can continue this inequality to find that


if vM is the projection of vh, on BJ n F, then v ,-a bGDVV,. Now GD_F is the
restriction to (D - F) x (D - F) of GD, and the restriction to D - F of
bGDvv is harmonic and is majorized by a potential, the restriction to D - F
of v. Hence GDv; = 0 on D - F; so G0v = 0 on D, and therefore v,,(Bf n F) =
0; that is, f lim sup,,.., u(q) 5 b for vh, almost every point of F for all b>0.
Hence u* = 0 v,, almost everywhere on F, as asserted. If this result is applied
to h/v with F the trace of a v,, null support for vv on the set of infinities of v,
we find that u* = + oo 4 almost everywhere on F, as asserted in (a).
There remains the proof of (4.1). Let ¢ be the Radon-Nikodym deriva-
tive in (4.1). It is sufficient to show that if Fis a vv null Borel polar subset
of D, then u* = 0 v,, almost everywhere on F. By what we have just proved,
if vw is the projection of v,, on F,

f lim h(q) - GD Vti (q) = 0


,-.c NO
at vw almost every point C of F; so in discussing u* on F we can assume that
F supports vh and that h = GDvh. Define

F. = 4,(i) 5 a}, (S,

v, 5 ah; so f lim,, va(ry)/h(q) 5 a at v1, almost every point C of F.


Moreover, by what we have proved above, f lim,,..., [v(q) - va(ry)]/h(q) = 0
at v,, almost every point i, of F. n F. Hence u* 5 a v,, almost everywhere on
F that is, u* 5 a vv almost everywhere on F where 0 5 a, and therefore
u* 5 ¢ v,, almost everywhere on F. A similar argument gives the reverse
inequality; so u* = 0 vw almost everywhere on F, as was to be proved.
Incidentally it now follows that u* < + oo v,, almost everywhere on D, as
asserted in (a), because u* = u < + oo quasi everywhere on D, and we have
just proved that on the remaining polar set, u* = which is finite vw
almost everywhere.
S. Extension of the Fine Topology to RN u { co } 175

5. Extension of the Fine Topology to IAN v { oo }


Theorem. Let A be a subset of 081, and let ¢ be an inversion of 081 in a sphere.
(a) Either (al) ¢(A) is thin at 4,(co) for every inversion 0, or (a2) 4(A)
is thin at 4,(oo) for no inversion tp.
(b) An unbounded set A is under case (al) if and only if there is a positive
superharmonic function u on some deleted neighborhood of the point
oo with the property that

lim u(q) _ =- 4 -0 0 (N= 2) , (5 . 1)


A 37-m log 1111

lim u(ri) = + oo (N > 2). (5.2)


A3q-m

Moreover, under case (a 1) with N > 2, there is a positive superharmon-


icfunction on 981 satisfying (5.2).
(c) If C and d(C) are finite, 4,(A) is thin at 4,(C) if and only if A is thin
at

Proof of (a) and (b). (For N > 2). To avoid trivialities we assume that A is
unbounded. Let 0 be an inversion in aB(C, b). Then 4,(A) has limit point C,
and we prove first that if ¢(A) is thin at C, there is a positive superharmonic
function u on 08' satisfying (5.2). In fact by Theorem 4(b) there is a positive
superharmonic function v on D = R N such that (4.1) is true, and therefore
the function

bz1
G(C, 0)

has limit + oo at the point oo along A. The function on the right is a positive
superharmonic function on 08' - {C} and in fact is a multiple of the Kelvin
transform of v. This function has a positive superharmonic extension u to
081 (Theorem V.5), and u is the desired function. Conversely, if there is a
positive superharmonic function u on some open deleted neighborhood of
the point ao satisfying (5.2) and if 0 is an inversion in aB(C, b), we prove
that 4,(A) is thin at C. The Kelvin transform v of u is a multiple of u(4,)G(C, ),
defined positive and superharmonic on a deleted open neighborhood B of C,
and has limit +oo at along ¢(A); so [by Theorem 4(b)] the set
A is thin at C, as was to be proved. Thus (a) and (b) of the theorem are true
forN>2. o
Proof of (c). (For N z 2). Under the hypothesis of (c), the set A is thin at
C if and only if (Theorem 2) there is a superharmonic function u defined on
176 I . XI. The Fine Topology

a neighborhood of C such that (2.1) is true. The Kelvin transform of u under


an inversion 0 is a positive superharmonic function defined on a neighbor-
hood of O(C) and satisfies the version of condition (2.1) at 4(C) if and only
if u satisfies (2.1). Hence (c) is true.

The proof of Theorem 5(a) and (b) for N = 2 follows the proof for N > 2
and is left to the reader.
Extension of the Fine Topology to R" u { oo }. We make the definition that
a subset A, of R" u {ao} does not have fine limit point co, and we describe
A, as thin at oo if A = A, n R" is subsumed under case (al) of Theorem 5.
The fine topology is thereby extended to R" u {oo}, and according to
Theorem 5, an inversion is a fine topology homeomorphism of I2" u {ao}
onto itself.
The Limit Relation (4.1) at the Point co. Let v be a positive superharmonic
function defined on a deleted neighborhood of a finite point C. Then (by
Theorem V.5) if v(C) is defined as lim inf,,..{ v(q), the extended function v is
positive and superharmonic on a neighborhood of C, and v satisfies the fine
limit relation (4.1). Trivially, v also satisfies the fine continuity relation

f lim v(q) = lim inf v(q).


o-C 11-C

If u is a positive superharmonic function defined on a deleted neighborhood


of the point oo, then inversion of R" in a sphere yields a context to which
(4.1) and (5.3) can be applied to yield the following.

If u is a positive superharmonic function on R" - B(0, 6), then

f"_'O
lim u(g) = lim inf u(q) (< + oo) if N = 2,
I,- ID
(5.4)
flimu(q)IgIN-2 = liminfu(,i)I,I" 2 if N> 2,

f,-M u(g)
limlog61171 - inf u(q) if N = 2,
lel>alogBl'1I (5.5)
flim u(,) = inf u(q)[l - (6117 l)2-"] 1 if N > 2.
'rm hI>a

In fact, if R" is inverted in B(0, 1), the Kelvin transform of u has domain
of definition B(0, 1/6) - {0} and is given there by q i-, v(g) = u(glgl-2)I,I2-1.
The relation (5.3) with t = 0 yields (5.4), and (4.1) with C = 0 and D =
B(0, 1/6) yields (5.5). The Green function GD is evaluated in Section I1.1.
If the domain of definition of u is an arbitrary Greenian set that is a deleted
neighborhood of the point oo, Example (c) of Section VIII.8 can be used
to evalute the right-hand side of (5.5).
7. Application to the Fundamental Convergence Theorem and to Reductions 177

6. The Fine Topology Derived Set of a Subset of R"


Theorem. A polar subset of R' has no fine limit point. Conversely, a subset
of R" with no fine limit point is polar. If A is an arbitrary subset of R", the
set Af is a fine perfect Gs set including quasi every point of A.

The converse is true whether or not a point ao is adjoined to R".


It was pointed out in Section 2 that a polar set has no fine limit point.
Since Theorem 6 is a local theorem, it can be assumed in proving the last
assertion that A is a subset of some Greenian set D. (We can take D = fib"
if N > 2.) Reductions below are relative to D. Let B, be an enumeration of
the balls with closures in D and with rational centers and rational radii.
According to Theorem 3, the Euclidean Fe set
U
j=0
is D - Af. Since each set in this union meets A in a polar set, the set D - Af
is an F. set meeting A in a polar set. In particular, A is polar if Af = 0.
The set Af is trivially fine closed, and is fine perfect because if were a fine
isolated point of Af, that is, if some deleted fine neighborhood B of
contained no point of Af, then the set B n A would contain none of its
fine limit points and thus would be polar, and could not be in Af, contrary
to hypothesis.

The Fine Boundary

If A is fine perfect in RN, that is, if A = Af n R", its fine boundary 8fA =
Af n (R" - A)f is a Euclidean Ga set, and therefore its fine interior A - 8fA
is a Borel set, the difference between two Ga sets. If A is analytic, its fine
closure in R" is analytic, and its fine interior A - (R" - A)r is analytic; so
its fine boundary is universally measurable. In particular, if A is a Borel set,
its fine interior and fine boundary are also Borel sets according to this
argument.

7. Application to the Fundamental Convergence Theorem and


to Reductions
In the Fundamental Convergence Theorem the lower envelope u of the
specified family of superharmonic functions is equal quasi everywhere
to its lower semicontinuous superharmonic smoothing u, and
lim inf, -, u(q). Hence u has a fine limit at every point,

u() = lim of u(q) = flim u(q).


n
178 1. X1. The Fine Topology

If D is a Greenian subset of 08", if A is a subset of D, and if v is a positive


superharmonic function on D,
OVBA = UV0AfnD. (7.1)

In fact the first two terms are equal because, on the one hand, A C A U
(Af n D) and, on the other hand, a superharmonic function on D majorizing
v on A necessarily majorizes v on Af n D. The first term is equal to the third
because A differs from A n Af by a polar set. The fourth term is equal to
the others because A n Af c Af c A u Af.

8. Fine Topology Limits and Euclidean Topology Limits


A function defined on a deleted Euclidean topology neighborhood of a
point and with a fine topology limit at the point need not have a Euclidean
topology limit there. For example, the indicator function of a countable
dense subset of R' has fine limit 0 at every point but has no Euclidean
topology limit at any point. Nevertheless the following lemma makes trivial
the proof of Theorem 9, which establishes a surprisingly close relation be-
tween the two kinds of limit.

Lemma. Let be a point of 98N'v {oo}, and let A. be a decreasing sequence of


subsets of I8" not containing . Suppose that each set Ak is a deleted fine
neighborhood of c [has fine limit point ]. Then there is a set A with the property
that for each k the part of A in a sufficiently small neighborhood of 1; is in Ak
and that A is a deleted fine neighborhood of [has fine limit point f].

It can be assumed that is finite. Define B(S) = S). Suppose that


each set Ak is a deleted fine neighborhood of , and define Fk = B(1) -
(Ak u The set Fk is thin at , and to show one part of the lemma, it
is sufficient to show that there is a set F, thin at , with the property that
for each k the part of Fk in a sufficiently small neighborhood of is a subset
of F. If Lot all but a finite number of values of k the point is not a limit
point of Fk, we can take Fas the union of those sets Fk for which is a limit
point of Fk. Otherwise, it is no restriction to assume that is a limit point
of Fk for all k, and we can apply Theorem 2 to find a positive superharmonic
function uk on B(1) with uk() < 2-k and with limit +oo at along Fk. The
function u = Io uk is positive and superharmonic on B(1) with 2 and
with limit + oo at along each set Fk. Choose rk so that I = ro > rl > ,
limk-. rk = 0, and so that u >_ k on Fk n B(rk). If F= Um o (Fk n B(rk)), the
function u has limit + oo at along F; so F is thin at , and Fk n B(rk) c F
as desired. To prove the second assertion of the lemma, observe that if each
set Ak has as fine limit point, then
9. Fine Topology Limits and Euclidean Topology Limits (Continued) 179

1= 1 IAk.e(,) -a(.n({)
r- o

for s > 0, according to Theorem 3 and Section VI.3(e). Hence it is possible


to choose r0 = I > r, > ... successively in such a way that r=0
and that the smoothed reduction on the right with s = rk and r = rk+t is at
least #. Define
w
A=
0

so that Ak r (B(rk) - B(rk+t)) c A. Then U 1 oA^&'k)() Z #. If follows (by


Theorem 3) that A is not thin at , and since A n B(rk) c Ak, the proof is
complete.

9. Fine Topology Limits and Euclidean Topology Limits


(Continued)
If u is a function from a deleted Euclidean topology neighborhood of a
point of R' into a topological space S, the fine cluster set of u at is defined
as n u(B)-, where B ranges through the class of deleted fine neighborhoods
of . In particular, if S is metric a is a fine cluster value of u at if and only
if f lim infq_, dist (u(rl), a) = 0.

Theorem. A function u from a deleted Euclidean neighborhood of into a


metric space has fine limit [fire cluster value] a at if and only if u has limit
a at along a subset of R' that is a deleted fine neighborhood of [that is not
thin at ].
These conditions are trivially sufficient. Conversely, if u has fine limit a
at , that is, if for n > 0 the set of points

A. _ q: dist (u(q), a) < n}

is a deleted fine neighborhood of , the set A of Lemma 8 is a set along which


u has limit at at . The cluster value assertion is treated similarly.
If the domain of u merely has as a fine limit point, the cluster value
definition and Theorem 9 have obvious rephrasings.

Continuity of a Superharmonic Function

If v is a function defined and superharmonic on an open neighborhood of


a point and if v has associated measure v, the function v is lower semi-
180 1. XI. The Fine Topology

continuous by definition. In addition the following properties have now been


proved or are now trivial.
(a) The function v is fine continuous at , and therefore (by Theorem 9)
there is a fine neighborhood A of such that VIA has (Euclidean topology)
limit at .
(b) If v(c) = + oo, then by lower semicontinuity v is continuous at ,
but v cannot become infinite too fast. More precisely (Theorem 4), there is a
fine neighborhood A of with the property that has a (Euclidean
topology) finite limit v({c}) at . If D is a Greenian set containing , G here
can be replaced by GD.

10. Identification of Af in Terms of a Special Function u"


Lemma. If D is a Greenian subset of U8^', there is abounded continuous potential
u' = GDu with the property that, for every subset A of D,

(reduction relative to D).

Let B. be a sequence of balls with closures in D, forming a basis for the


topology of D, and define
W
uE2-n01Is
0

The summands are continuous potentials because 010k" = on


D - B. according to Section VIII.10, and the points of 8B are regular
boundary points of D - B,,. The function u' is therefore a bounded contin-
uous potential on D and if A c D

If E Af n D, the left side of (10.2) is trivially u'(S) (Section 3). If E D - Af


and if e B,,, then 01 Q8.Q A( ) is arbitrarily small for small
B. [Section 3(a")]. Hence the right side of (10.2) is strictly less than
as was to be proved.

11. Quasi-Lindelof Property


An arbitrary union of open subsets of a second countable Hausdorff space
is equal to some countable subunion (Lindelof property). The fine topology
has a slightly weaker property (quasi-Lindelof property) as follows.
12. Regularity in Terms of the Fine Topology 181

Theorem. An arbitrary union of fine-open subsets of R" differs by a polar set


from some countable subunion.

We shall prove the corresponding complementary assertion : an arbitrary


intersection n7E,A7 = A of fine-closed subsets of R" differs by a polar set
from some countable subintersection. It is sufficient to prove this assertion
for subsets of some Greenian set D. Reductions below are relative to D.
Let u' be a superharmonic potential on D with the properties described in
Lemma 10. According to the Fundamental Convergence Theorem, the
family A. has a countable subfamily {A7, ate J) such that

rinfOu'i"al = (infOu'OA
`aEf +
261 +

To prove the theorem, we define B = n7E./ A. and prove that Bf = Af.


Observe that

Qu"OB 5 5 U' (11.1)


7 8.1 7E1
+ +

for a e 1, and we conclude that Bf c Aa c A. for a e 1. Hence Bf c A, and


therefore (Bf)f = Bf c Af. Since A c B, it follows that Bf = Af ; so B - A
is polar, as was to be proved.

Application to the Approximation of a Fine-Open Set by (Euclidean


Topology) Compact Subsets
If A is a fine-open subset of R" to each point of A corresponds (Sectionl),
a Euclidean topology compact subset of A which is a fine neighborhood of
the point. The union of these compact sets covers A, and therefore (quasi-
Lindelof property) some countable subunion covers A up to a polar set.

12. Regularity in Terms of the Fine Topology


Theorem. A finite Euclidean boundary point C of a Greenian subset D of R"
(or the point oo if N = 2 and D is unbounded) is regular if and only if
e(R"-D)f.
Recall that if N > 2, the point oo is a regular boundary point of every
unbounded open set. Since the case of an infinite boundary point when
N = 2 can be reduced to that of a finite boundary point by an inversion,
only finite boundary points will be considered in the following proof. Sup-
pose then that t; is a finite Euclidean boundary point of D, and let B be a
ball of center C, so small if N = 2 that B u D is still Greenian. According
to Section VIII. 10, the restriction to D of Q 1 Q B-D (reduction relative to B v D)
182 I.M. The Fine Topology

is B n aD). If C is a regular boundary point of D, it follows that


this smoothed reduction has limit I at C on approach along D. Moreover
91 B-D = I quasi everywhere on B - D; so [111-D has limit I on approach
to C excluding a polar set. Hence (fine continuity of the smoothed reduction)
Q1
8-D(C) = 1. This smoothed reduction is majorized by the smoothed reduc-
tion on B - D relative to a Greenian superset D, of B u D; so the latter
smoothed reduction is also I at C. Since this is true for arbitrarily small B,
it follows (from Theorem 3) that C e (R" - D)f . Conversely, if ( e (R" - D)f,
then 0 I o8-D(C) = l (reduction relative to B u D) by fine continuity; so the
lower semicontinuity of superharmonic functions implies that this smoothed
reduction is continuous at C. Hence B n OD) has limit I at C for every
B, and this condition implies regularity of C (Section VIII.8).
Observation (a). Regularity of a finite Euclidean boundary point C re-
quires only that C be in (R" - D)f but quasi every finite point of OD is not
only regular and therefore in (R' - D)f but is even in (OD)f (Section 6).
Observation (b). Theorem 12 shows that every condition that a point
be a fine limit point of a set is a condition for Euclidean boundary point
regularity. For example (Theorem 3), a finite Euclidean boundary point C
of D is regular if and only if whenever D' is a Greenian superset of D contain-
ing C. it follows that Ga .°({, ) = GD.(C, ). In fact, since regularity is a local
property, D' need not be a superset of D (but of course must contain 0).
It is natural as a counterpart of Theorem 12 to investigate the set of
points of aD in D! The basic result is the following.

13. The Euclidean Boundary Set of Thinness of a Greenian


Set
Theorem. A Greenian subset D of RN` is thin at p° almost no finite Euclidean
boundary point of D.

If N>2define D,=R";if N=2defineD,=DuB,where Bis aball


with center a finite boundary point of D and is so small that D U B is
Greenian. It is sufficient to prove that D is thin at pD almost no Euclidean
boundary point of D in D1. Let ul be the positive superharmonic function
described in Lemma 10 but associated with D, rather than D, and define
u = Qu'Q° (reduction relative to D,). Also define u = u = 0 on OD n OD,.
To prove the theorem, it is sufficient to prove that u = u at PD almost every
point of D, n OD, equivalently, since u 5 u , to prove that u) = U 0)
on D. Now (by Theorem V.11) the Riesz measures associated with u and
u vanish on polar sets because u and u are bounded; so [by Theorem
VIII.18(c)] u) = GMDU and PD(', u0) = GMDUL' on D. These harmonic
minorants are equal because u = u+ on D; so the harmonic averages are
also equal, as was to be proved.
15. Characterization of 0µQ A 183

14. The Support of a Swept Measure


Theorem. Let D be a Greenian subset of R N, let A be a subset of D, and let
be a point of D.
(a) If l; eAJ then bD(,
(b) If µ is a measure on D, then the measure 8µQ" [in particular, the
measure l (, )] is supported by Af n D.
(c) If A is a Borel set, if A is fine dense in itself, and if µ is a measure
supported by A, then y = Dµ0"
(d) If v is a positive superharmonic function on D, the Riesz measure
associated with Dv0" is supported by Af n D.

Observation. We shall sharpen (b) in Section 18 by proving that a is


supported by the fine boundary of A relative to D whenever e D - Af.

Proof. (a) If a Af then bD(, { }) = I according to Theorem 3(b).


(b) If u is a positive superharmonic function on D, then

Du 'D" = J Du'D"0" = (14.1)

so if u' has the properties described in Lemma 10, the set

{u'>Du'D"}=D-Af
is aD(, ) null for every ; that is, SD( , ) is supported by AfnD. The
evaluation of Dµ0" in X(5.2) shows that this swept measure is supported by
AfnD.
(c) The evaluation of 0µD" in X(5.2) yields (b).
(d) If v" is the Riesz measure associated with OvO", then VA = OVA DA
according to Section X.5; so (d) follows from (b). o

15. Characterization of
D be a Greenian subset of RN, let A be a subset of D, and let
µ be a measure on D with superharmonic potential GDµ. Then the swept
measure Oµ0A is characterized uniquely by the following properties:
(a) DµO" is supported by AfnD.
(b) GoDpO" = GDµ on Af n D.

We have already proved that Oµ0" has these two properties. Conversely,
if a measure v on D has these properties, then v = OvO" by Theorem 14;
so since is supported by Af n D,
184 l.Xl. The Fine Topology

GDV =
/
5D(.,GDV) = SD(.,GD,) = GDOPOA,

and it follows that v = OµOA


Observation. This theorem extends and makes more precise the charac-
terization of OµOA for A closed in D made in Section X.1. In comparing the
two characterizations when A is closed in D, observe that in Theorem 15(b)
the equation holds everywhere on Af n D if it is known to hold quasi every-
where on this set because Af is fine perfect.

16. A Special Reduction


The following lemma is the first step in a delicate analysis of the support
of a swept measure.

Lemma. Let D be a Greenian subset of R', let h be a strictly positive harmonic


function on D, and let u = GDp/h be an h-superharmonic h potential on D.
Then for even' positive constant c (< + oo)
hpunlusc) = U A C. (16.1)

Observe that if v = GDllv,lllla rephrasing of (16.1) is


P)vsch)
= V A (ch), (16.1')

valid for every superharmonic potential v. To avoid trivialities, it will be


assumed in the proof of (16.1) that 0 < c < +oo. It is sufficient to prove
equality of the first and third terms in (16.1) because this equality with
c(1 - 1 /n) instead of c yields equality of the second and third terms when
n + oo. Now
hlUplusc) = h]U A Cp)uc) S U A C. (16.2)

Furthermore there is equality in (16.2) quasi everywhere on the closed in D


set {u S c}, which is a support of the measures associated with the finite-
valued h-potentials in (16.2). Hence (by the domination principle) there is
equality in (16.2); so the lemma is true.

17. The Fine Interior of a Set of Constancy of a


Superharmonic Function
Corollary. Let D be an open subset of R', let v be a superharmonic function
on D with associated Riesz measure v, and let c be a constant. Then the fuze
interior of the set {v = c} is v null.
18. The Support of a Swept Measure (Continuation of Section 14) 185

This corollary is a strengthening of the result that the Euclidean interior


of the set {v = c} is v null because v is harmonic on this interior, and the
corollary suggests that Euclidean open sets and fine-open sets play similar
roles relative to superharmonic functions. This idea will be discussed further
in Section 19.
Since the corollary is local, it is no restriction in its proof to assume
that D is a ball on which v is lower bounded. Without loss of generality we
can then assume that v is a potential, v = GDv, because we can first add a
constant to v to make v positive and then replace v by its smoothed reduction
on a strictly smaller ball than D, concentric with D. The resulting function
v is a potential on the ball D and on the smaller ball differs by a constant
from the original function. To prove the corollary in this context, let A be
the fine interior of the set {v = c}. Lemma 16 applied to the function v A C
yields Dv A cOt""1 = v A C. Since (by Theorem 14) the Riesz measure vv
associated with this smoothed reduction is supported by D n {v < cy, which
contains no point of A, it follows that vv(A) = 0. Next apply Lemma 16
to derive

V A C, QvjtaS`f = VC, (17.1)

so that

0 = vv(A) = J A) (17.2)
D

Now A c {v < c}f, and therefore [by Theorem 3(b)] the integrand in (17.2)
is 1 when e A ; so "(A) = 0, as was to be proved.

Pseudogeneralization of the Corollary to h-Harmonic Functions

If h is a strictly positive harmonic function on D, the function u = v/h is


h-superharmonic, and the fine interior of the set {u = c} = {v - ch = 0} is
v null according to the corollary because the superharmonic function v - ch
has the same associated measure as v. Thus the corollary is not effectively
more general when stated for h-superharmonic functions.

18. The Support of a Swept Measure (Continuation of


Section 14)
If A is a subset of RN, we denote the set of fine limit points of A, the fine
interior of A, the fine boundary of A, and the Euclidean boundary of A by
Af, Af', afA, and 8A, respectively, we leave to the reader the proof that
afAf c 8f (A . Af) c VA. Recall from Section 6that Af, Af', and afAf are
Borel sets.
186 1.X1. The Fine Topology

According to Theorem 14, the measure is supported by Af n D


for all in D, and if v is a positive superharmonic function on D, the Riesz
measure associated with the smoothed reduction DvCA = 6D'(', V) is also
supported by Af n D. The following theorem sharpens the more elementary
result (given in Section X.1) that is supported by D n aA when
e D - A by putting the latter result into the context of the fine topology
and thereby also sharpens Theorem 14.

Theorem. Let D be a Greenian subset of R v, and let A be a subset of D.


(a) If e D n Af, then 1. If e D - AJ the measure
is supported by D n afAf.
(b) If v is a positive superharmonic function on D, the Riesz measure vA
associated with [VIA is supported by D n Af. In particular, if v is
harmonic on a Euclidean open superset of A, the measure vA is supported
by D n afAf.

Proof of (a). The first assertion of (a) is Theorem 3(b). To prove the second
assertion, let be a point of D - Af, and observe that the restriction v to
D- of the function is superharmonic with associated
Riesz measure the restriction of to the class of measurable sub-
sets of D - The set {v = 0) is a Borel superset of the set Af n (D -
and it follows from Corollary 17 that the fine interior of this superset is
5 null. Since it is already known (Theorem 14) that SD(,) is supported
by D n Af, it follows that this measure is supported by D n afAf, as was
to be proved.

Proof of (b). The first assertion of (b) is included in Theorem 14(c). To


prove the second assertion, observe that if v is harmonic on the open superset
B of A, then the set { pvQA = v} n B is the set on which the function IvIA - v,
superharmonic on B, vanishes. The fine interior of this set includes Af' and
is null for the Riesz measure associated with CvDA - v on B. Hence vA is
supported by D n afAf.

Extension of Corollary 17. If u, and u, are superharmonic functions on an


open subset of R', if p, is the Riesz measure associated with u,, and if A =
{u, = u2 }, then p, = µ, on the Borel subsets of the fine interior of A.

Since this extension of Corollary 17 (like the corollary) is local, it is no


restriction in the proof to assume that u, and u2 are defined on a ball D on
which u, and u2 are lower bounded and even (after addition of a sufficiently
large constant) positive. Thus we shall suppose that u, is a positive super-
)ujDD-'(x) =
harmonic function on a ball D. Now o\ for
quasi every point of D - A and therefore (fine continuity f superharmonic
functions) also for in D and a fine boundary point of D - A, that is,
20. A Generalized Reduction 187

a fine boundary point of A. For a fine interior point of A the measure


aD-"(t, ) is supported by the trace in D of the fine boundary of A, on which
u1 = U2; so Du10D-" + u2 = u1 + lu2OD-" Since (by Theorem 18) the mea-
sure associated with a reduction on D - A vanishes on Borel subsets of the
fine interior of A, we conclude that pl = p2 on these subsets.

19. Superharmonic Functions on Fine-Open Sets


Let D be a Greenian subset of R' and let B be an open relatively compact
subset of D. If e B, Theorem 13 implies that the measure is supported
by MB. If v is a superharmonic function on D then

v() ;-> bD-B(,V), (19.1)

and there is equality if v is harmonic. Suppose now that B more generally


is a fine-open relatively compact subset of D and that e B. Then
is no longer defined, but (by Theorem X.5) if v is a positive superharmonic
function on D,
-f SD-1(S' V). (19.2)

It follows easily that

v() z 6DD-8( ,v) (19.3)

whenever v is a superharmonic function on D. In particular, there is equality


in (19.3) when v is harmonic because (19.3) in then true for both v and - v.
Observe that the measure is supported by D r M(D - B) = MB,
according to Theorem 18.
Thus the superharmonic function inequality and the harmonic function
equality can be generalized to be valid on fine-open sets, and this suggests
that much of classical potential theory can be extended by replacing Euclid-
ean open sets by fine-open sets. The next step is to develop a fine potential
theory based on generalizations of superharmonic and harmonic functions
to fine superharmonic and harmonic functions, defined on fine-open sets.
We omit further discussion in this direction.

20. A Generalized Reduction


If D is a Greenian subset of R", if A is a subset of D, and if v is a positive
superharmonic function on D, we have defined the reduction Rv relative to
D as the infimum of the class of positive superharmonic functions on D
188 I. XI. The Fine Topology

majorizing v on A, equivalently, as the infimum of the class of superharmonic


functions on D majorizing v1A on D. In this section we replace VIA by a
function on D that is arbitrary except for boundedness and smoothness
conditions.
Let D be a Greenian subset R', let f be a locally lower bounded function
from D into A majorized by some superharmonic function on D, and let
Rf be the smoothed infimum of the class of superharmonic functions on
D majorizing f. According to the Fundamental Convergence Theorem, Rf
is superharmonic and Rf z f quasi everywhere on D; so Rf is the infmum
of the class of superharmonic functions on D majorizingf quasi everywhere
on D. Let u f be the Riesz measure associated with R.. Recall that when C
is a subset of R', the notation Cf refers to the set of fine limit points of C.

Theorem. In the above context,


(a) For c > 0 the measure p f is supported by the set

{Rf= +oo}u n {f5Rf5f+e< +oo}f.


L>o

(b) If f is fine upper semicontinuous, the measure M f is supported by the


set {Rf=f}f.

This theorem generalizes Theorem 18(b). In fact, suppose that A is a


subset of D, nonpolar in each open connected component of D to avoid
trivialities, and suppose that v is a strictly positive superharmonic function
on D. Define f = 1,,v; so Rf = Rr , and suppose first that A is fine closed in
D. Then f is fine upper semicontinuous, and according to the present theorem,
pf is supported by the set {Rf =J}1 and so is supported by {Rf =1) because
this set is fine closed in D and so is supported by Af because Rf >f = 0 on
D - A. If A is not fine closed in D, we can replace A here by the set Af n D,
fine closed in D. because by fine continuity of superharmonic functions any
superharmonic function majorizing v on A majorizes v on Af n D; so Rf
and R" are not changed by this set change. The result for fine-closed sets
shows that µ f is supported by Af n D as stated in Theorem 18(b).

Proof (a) To prove Theorem 20, suppose first that f z 0, define

Ata={f<R1<f+e<+oo}vJ>--

for r > 0, b > 0, and observe that the function

A
1+eaRf+l+rbRf
20. A Generalized Reduction 189

is superharmonic on D, is majorized by Rf, is equal to Rf quasi everywhere


on Ath, and majorizes f quasi everywhere on D - A,, because the first
summand does. Hence this superharmonic function coincides with Rf; so
Rf =Q Rf D A . According to Theorem 18, it follows that RJ is supported by
A for all strictly positive a and S. Now

up to a polar set; so

f-{RJZ},
1RfZ- I
and when b - 0, the set on the right decreases to { Rf = + oo }. Thus Theorem
20(a) has now been proved for f z 0.
To reduce the general case to the case of positive f, observe that if B is
a compact subset of D and if Do is an open neighborhood of B, relatively
compact in D, then the restriction of Rf to Do is the smoothed infimum
of the class of superharmonic functions majorizing R f on Do - B and
majorizing f on B. Without loss of generality in the discussion of of we
can suppose, adding a constant to f if necessary, that the restriction to Do
off is positive. It now follows from what we have already proved that for
every B the projection of u f on B is supported by Af,6 n Do for all strictly
positive E and S; so µf is supported by A' for all strictly positive a and 6,
and the rest of (a) follows as in the positive case.
(b) If f is fine upper semicontinuous, the set

{Rf Sf+e < +oo}v{fZ


E}
differs from A. by a polar set and is fine closed in D; so

Au=AucAu,
and part (a) of the theorem implies, when c - 0, that the measure µf is
supported by the union of {Rf =A and the polar set {R <f). Since (by
Theorem V.11) Rf = + oo at u r almost every point of a polar set, the set
{Rf <A must be µf null; so part (b) of the theorem is true. o

Special Positivity Case

Suppose in the theorem that f is positive (and as in the theorem is majorized


on D by some superharmonic function), and define A = (f > 0}. Then Rf
190 1.X1. The Fine Topology

is a positive superharmonic function, and it is easy to check that Rf = BR fQ".


We conclude from Theorem 14 that the measure associated with Rf is
supported by Of* n D.

21. Limits of Superharmonic Functions at Irregular Boundary


Points of Their Domains
If u is a lower-bounded superharmonic function defined on a deleted open
neighborhood of a point C of RN, then (by Theorem V.5) u can be extended
to be superharmonic on the neighborhood; so u has a fine topology limit
(5 + oo) at . The following theorem shows that in this result the neighbor-
hood of C need be merely a fine neighborhood. Recall (from Theorem 12)
that a Euclidean boundary point of an open set is irregular if and only if
the set is a deleted fine neighborhood of the point.

Theorem. Let u be a lower-bounded superharmonic function on an open


subset D of RN, and let C be an irregular Euclidean boundary point of D.
Then f (denoted below by fu(C)) exists (5 + oo).

Proof. If N > 2, the point C must be finite because (by Theorem VIII.4)
oo is a regular boundary point of every unbounded open set. If N = 2 and
if C = oo, an inversion of the plane in a finite boundary point reduces the
theorem to one with C finite. Thus we can assume that C is finite and therefore
also that D is bounded. If C is an isolated boundary point of D, u has a super-
harmonic extension to D u (Cl and so has a fine limit at C, the value at C
of the extension, namely, lim inf, If C is not an isolated boundary
point of D, the fact that C is irregular implies that RN - D is thin at C and
therefore (by Theorem 2) that there is a positive superharmonic function v
on an open neighborhood of D, finite at C, with limit + co at C along RN - D.
Define c = liminf,.t(u + vID)(). If c = +oo, then u has fine limit +oo at
C because v has the finite fine limit v(C) at C along D. If c < + oo, the function
u1 = (u + vID) A (c + 1) is superharmonic on D, is majorized by c + 1, and
has limit c + I at every boundary point of D in some open neighborhood
B of C except at C itself. Extend ul to D u (B - {iC}) by setting ut = c + 1
on B - (D u {C}). Then ul is superharmonic and lower bounded on a deleted
neighborhood of C, and as noted above, such a function has a fine limit at C,
necessarily the inferior limit of the function at C. Thus

flimui(C) = liminfut(C) = c.

Hence u + VID has fine limit c at C; so u has fine limit c - v(C) at C.


Observation. The preceding proof shows that if the fine limit of u in
Theorem 21 is finite and if v is a superharmonic function on an open neigh-
22. The Limit Harmonic Measure Jup 191

borhood of C, finite at C, with limit + co at C along R" - D, then u + v has


a superharmonic extension to the union of D and an open neighborhood
of C.

Application to Harmonic Measure

Let D be a Greenian subset of I8" with an irregular Euclidean boundary


point C. Iff is a finite-valued continuous function on the Euclidean boundary
of D, then f) = Hf is a bounded harmonic function on D and thus has
a finite fine limit at C. Iff is a sequence of such continuous functions on D,
dense in C(OD), then an application of Theorem 21 combined with Lemma
8 shows that there is a subset A of D, a deleted fine neighborhood of C,
such that every function f) has a limit at C along A. It follows that
lime 3 4-c pD(C, -) exists in the sense of vague convergence of measures on
OD. Denote this limit by fµD(C, ). In the next section we shall show that
this limiting harmonic measure has many of the properties of ordinary
harmonic measure.

Modifications of Theorem 21

If in Theorem 21 it is supposed only that u is lower bounded on a deleted


fine neighborhood in D of C, say u > a on such a set, the theorem can be
applied as originally stated to u on {u > a) to verify that the original conclu-
sion remains valid. Moreover a variation of the argument of the theorem
shows that if u/G(C, ) is lower bounded on D or even merely lower bounded
on a deleted fine neighborhood in D of C, then this ratio has a finite fine
limit at C. The special case in which C is an isolated boundary point of D
and u is positive on D is covered by Theorem 4(c).

22. The Limit Harmonic Measure fun


In this section boundaries of subsets of R" are relative to the Euclidean
topology. Let D be a Greenian subset of R", and let C be an irregular bound-
ary point of D; that is, let D be a deleted fine neighborhood of C. Then (from
Section 21) litnC-4P (C, ) exists in the sense of vague convergence of mea-
sures on OD when C tends to C along some subset of D that is a deleted fine
neighborhood of C. In the following t is the class of open subsets of D that
are deleted fine neighborhoods of C.
(a) If A is a Borel subset of OD and if Be r", then

fuD(C, A) = fue(C, A n OB) + L'OB uDO1 , A)fpB(C, dry). (22.1)


192 1. X1. The Fine Topology

According to VIII(8.3), this equation is true when C is a point of B.


Equation VIII(8.3) is equivalent to

ND( ,f) = PB(S,JB) + J (22.2)


D naB

for all finite-valued positive continuous functions f on BD, with fB = f on


aB n aD and fB = 0 elsewhere on aB. The function is harmonic on
B and so (by Theorem 21) has a fine limit at C. The function equal to f)
on D n aB and equal to 0 elsewhere on aB is lower semicontinuous. Hence
if we take fine limits in (22.2) when - C, we find

f1D(C,J) Z f 1im PD01,f) UB(C, (22.3)


J OnaB

and if this inequality is combined with the corresponding inequality for


-f + maxD f, we find that there is equality in (22.3). Since fB is upper semi-
continuous on tB, it follows that

fuX,f) fuB(C,fB) + J d n)' (22.4)


DnaB

Finally, this inequality combined with the corresponding inequality for


-f + maxD f yields equality in (22.4); that is.

fPD(C,f) ='uB(C,fB) + d>'1). (22.5)


J DnoB
Since this equality is true for positive continuous f on OD, it is true for
arbitrary one-side-bounded Borel measurable f on aB, and therefore (22.1)
is true.
(b) If A is a harmonic measure null subset of aD, then fpD(Z, A) = 0.
We prove this in two steps, first proving it when C is not in A by proving
it for A compact and not containing C. This result is implied by (22.1) if B
is so small that A n aB = 0. Second, we prove that fuD(C, {C}) = 0. To
prove this, we use the fact that R' - D is thin at C to find a positive super-
harmonic function v on the union of D with an open neighborhood of C,
with v(C) < + ao and with v having limit + oo at C along RN - D. Define the
lower semicontinuous function v' on aD be setting v'(q) = lim infD. c-e v( ),
and define v" as v' on aD - {C} but v"(C) = + oo. Then v" is also lower
semicontinuous. If e > 0,

fPD(S, {S}) !5 fUD(S, ev') -5 lim

inf ev') 5 ev(O.


S-{
22. The Limit Harmonic Measure JµD 193

Here we have used the fact that v" = v' off the harmonic measure null set
{C} and that v is in the upper PWB class on D for the boundary function ev'.
Inequality (22.6) implies that f1D(C, {C}) = 0, as asserted.
From now on when u is a function in D, and Be f, the averages pa(C, u)
and fµ8(0, u) will be used under the convention that in these averages u(PI)
is to be taken as lim infs j4.,u(C) when q e aB n aD. If u is superharmonic,
this convention means that the function to be integrated on aB is lower
semicontinuous.
(c) If u is a lower-bounded superharmonic function on D and if B, and
B2 are in f with B, c B2, then

f u(C) ? fAB' (C, u) Z fµs,(C, u), (22.7)

and fu(C) = sup {µ8(C, u): Be f}.


Recall the convention just made for u on aB, n aD, and observe that
(from Section VIII.8) u z µs u) on B, . (Here are below the convention
on the values of u assigned at boundary points of the sets involved may
not be the same as those in referenced inequalities, but the differences are
in favorable directions.) On approach to C along a suitable deleted fine
neighborhood of C this inequality yields the first inequality in (22.7). Similarly
uZ u) on D; so (22.5) with f replaced by u yields the inequality rILD(C, u)
5 fµ8(0, u) and thereby yields the second inequality in (22.7) if the pair B,
D is replaced by the pair B1, B2. Finally, if B= {CeD: u(C) > a} with
a < fu(C), then fµ8(C, u) z a; so the last assertion of (c) is true. It is trivial
that the supremum in this assertion is not decreased by the additional
condition on B that aB n aD = {C}.
(d) The function is lower semicontinuous on the space of posi-
tive superharmonic functions on D in the topology of pointwise convergence.
In fact, if u, is a convergent sequence in this space, with limit u, and if
a < fu(C), choose B in r in such a way that aB n OD = {C} and that fue(C, u)
> a. Then apply Fatou's lemma to obtain

lim inffun(C) z lim inf fu8(C, un) Z fus(C, u) > a.


ft-.0 n-m

This inequality implies the stated lower semicontinuity.


(e) By definition of regularity, if q is a regular boundary point of D,
then limC . µD(C, ) in the sense of vague convergence is the unit measure
supported by {q}, whereas if q is an irregular boundary point of D, then
f lim ) = fuD(q, ). That is, in terms of Dirichlet solutions if f is a
finite-valued continuous function on OD, then limt.,n Hf(C) = f(q) if q is a
regular boundary point, but f lim4-,, Hf(C) = 'PAID if q is an irregular
boundary point.
(f) If C as above is an irregular boundary point of D and if q e D, then
according to Theorem 21, the function has a fine limit at C; we
denote this fine limit by fGD(C, q). In view of the extension Go of GD defined
in Section VII.4, not only does this fine limit exist, but
194 1 X1. The Fine Topology

IGD(C, 1) = lira sup GD(b, p), (22.8)


S»;

and VIII(18.2) implies that IGD(C, ) is harmonic on D, as is also easily proved


directly. When D is connected, the superior limit in (22.8) must be strictly
positive; that is, fGD(C, ) must be strictly positive, in view of the criterion
[Section VIII.14, Application (c)] in terms of GD that a boundary point of
D be irregular.

23. Extension of the Domination Principle


Recall (from Section 4) that if h and v are strictly positive superharmonic
functions on a Greenian subset D of R' with respective associated measures
vi, and v, if u = v/h wherever either h or v is finite valued, and if u*(C) is
defined as the fine limit of u at C wherever this fine limit exists, then u* is
defined quasi everywhere and vk + v almost everywhere on D. The following
extension of the domination principle (Theorem V.10), unlike Theorem V.10,
does not suppose that polar sets are v,, null.

't'heorem. Let h = GDvi, be a potential on a Greenian subset D of R", and let


v be a positive superharmonic function on D. Then h 5 v if (fine limit function
notation) (v/h)* z I vA almost everywhere.

To prove this theorem, which obviously includes Theorem V. 10, let B be


a Borel support of vi, on which (v/h)* is defined and z 1. If 0 < c < 1, the
set A = {v z ch} is a deleted fine neighborhood of every point of B; so
B c Af, and therefore bD(C, {C}) = I when C e B. Hence

v z 11v0" z cDhD" = cGD(vi,bn) = ch;

so vzh.
Chapter XII

The Martin Boundary

1. Motivation
Let D be an open subset of R'. If D is a ball, its Euclidean boundary is so
well adapted to it from a potential theoretic point of view that the following
statements are true.
(a) The class of minimal harmonic functions on D is in a one-to-one
correspondence with the Euclidean boundary points, q a K(ij, ) (see Section
11.16), and every positive harmonic function u on D is representable as an
integral J8D K(ry, )M (dq) with M. a uniquely defined measure on the ball
boundary. The lattices of positive harmonic functions on D and of measures
on aD are thereby in a one-to-one order-preserving correspondence, and
the vector lattices of differences between positive harmonic functions on
the ball and signed measures on the ball boundary are linearly and order
isomorphic.
(b) The Euclidean boundary of the ball is universally resolutive and
universally internally resolutive (Section IX.12).
(c) Every boundary point is regular, so that has limit 0 at
every boundary point. Moreover (up to a multiplicative constant) the normal
derivative of at the boundary is the density relative to /N_, of har-
monic measure.
(d) The Fatou theorem and its generalizations (Section 11.15) assign
to each positive h-harmonic function, and therefore to each h-harmonic
function in L'(µo_) on the ball, a boundary function. In particular, the
Fatou boundary function of a PWB" solution Hf is equal M almost every-
where to f
The Martin boundary of a Greenian subset D of R" is defined in such
a way that it has some of these properties. Observe, however, that other
compactifications may be better adapted to other properties. For example,
the Kuramochi boundary is specially well adapted to the second boundary
value problem.
The only restriction to be imposed on D is that it be Greenian and con-
nected. It will be necessary to generalize two concepts playing a central
role in the ball case: special approach to the boundary (radial or non-
196 1.Xtl. The Martin Boundary

tangential) and the normal derivative of the Green function at the boundary.
The fine topology will be extended to the Martin boundary, and approach
to a Martin boundary point in this topology will replace both radial and
nontangential approach to the boundary in the ball case. If D = B(0, r)
and if I e D, then const. (r - j') when j>ij --i r so that if v is a
measure on D with compact support, const. (r - Iryf) near the
boundary. This fact suggests that the normal derivative of GD(i;, ) at the
boundary point C should be replaced by

lim h)
rc GDv(ry)

when c is a Martin boundary point, if this limit exists, for some convenient
measure v. It was Martin's idea to define a boundary in such a way that this
limit does exist and in fact to use the existence of this limit to define a
boundary.

Topological Conventions

Recall that throughout this book in dealing with R" and its subsets, if no
topology is specified, the topology of RN is to be understood as the usual
topology, sometimes identified as the Euclidean topology, compactified by
a point at infinity when boundaries of unbounded sets are in question.

2. The Martin Functions


Let D be a connected Greenian subset of R', let v be a not identically 0
measure on D with compact support a, c D, and define

K,(17, ) = GD(n, )
GDV(q)

for , and in D except that is left undefined when co.


The function K, will be called the Martin function based on v, and K will
sometimes be written for K, when the choice of v is irrelevant. If v is the
probability measure supported by a singleton so that K,(rl,
GD(rl, )/GD(rl, a), the Martin function will be said to be based on o.
For h a point of finiteness of Gov the function K,(r1, ) is a strictly positive
harmonic function on D - {, }, and for in D the function K is con-
tinuous on D - a,. In view of the properties of GD listed in Section VIL3,
if A is a compact subset of D and if A, is a neighborhood of A u or, the
functions K, and 1/K,, are bounded on (D - A,) x A. Moreover
3. The Martin Space 197

f K,(n, Ov(dd) = 1 (2.2)


,

if GDv(n) < + OC.

3. The Martin Space


Theorem. Let D be a connected Greenian subset of RN. There is a unique up
to homeomorphisms metrizable compactifcation DM of D with the following
properties.
(a) Each Martin function K,, on (D - a continuous extension
(also denoted by to (DM - x D.
(b) K,,(gl, ) = K,(n2, ) if and only if ni = n2
The boundary aMD = DM - D obtained in this way will be called the
Martin boundary, and any metric for DM compatible with its topology will
be called a Martin metric. The function K. is now defined on its original
domain augmented by 0MD x D.

Observation. According to the theorem, if n, is a sequence of points of


D with limit n in aMD and if K,, is a Martin function, then K,,(n,,, ) =
K,(n, ) is a strictly positive harmonic function satisfying (2.2) because in
any open relatively compact subset of D the above limit is the limit of a
uniformly bounded sequence of positive harmonic functions satisfying
(2.2). Moreover, because of (b) and the convergence remarks in Section 2,
a sequence n, in D with no limit point in D converges to a boundary point
if and only if lim,,..,, K,,(n,,, ) exists; the limit function is then K,(n, ) with
n= n,,. Thus the existence assertion of the theorem implies that
each point n of DM - a,, is associated with a function K,(n, ), harmonic
on D if n E 3MD, harmonic except for an infinity at n otherwise, and the
class of associated functions under the topology of pointwise convergence
is homeomorphic with DM - a,. This homeomorphism implies the unique-
ness of a compactification satisfying (a) and (b) above. Roughly, the Martin
space DM is the compactified space of the set of Green functions {GD(n, ),
n e D} normalized to be conditionally compact under pointwise convergence.
Observe that if K, and K are Martin functions, then

Kti(n, ) [17 c- D - (a, u o,.)]. (3.1)


J C)u(dZ)
oN

Hence for n in D - (a u a,,) the function KM(n, ) associated with the point
n and the Martin function K,, is a constant multiple of the function K,.(n, )
198 1.X11. The Martin Boundary

associated with q and K.. The choice of reference measure for a Martin
function is a matter of convenience, although the most common choice
is one with a, a singleton, that is, with K, based on a reference point.

Proof of theorem. Let K. be a Martin function, and suppose that v has been
chosen to make Gov finite valued and continuous. Let f be a strictly positive
continuous function on D, IN integrable over D, and define

d,(C, q) = [I A IK,(C, C) - K,(q, C)I)f(C)IN(dC). (3.2)


fo

T he function d, is a metric for D, endowing D with the Euclidean topology:


so if q, is a Cauchy sequence ford, and if some subsequence q,, has Euclidean
topology limit q in D, then lim, ,, d,(q, qa.) = 0. It follows that q, is con-
vergent to q in the d, metric and therefore also in the Euclidean metric.
If q, is a sequence of points in D and if no subsequence of q, has a Euclidean
topology limit in D, then according to Section 2, if B is an open relatively
compact subset of D, the sequence n i-. K,(,, ) on B is a uniformly bounded
sequence of harmonic functions if finitely many values of n are omitted.
It follows from the convergence properties of harmonic function sequences
(Section 11.3) that the sequence n i-4 K,(q,,, -) has a subsequence j i -+ K,(q, , -)
converging locally uniformly on D to a harmonic function h. This subse-
quence is a Cauchy sequence for d,. In particular, if q, is a Cauchy sequence
ford, and if the pair (C, q) in (3.2) is replaced by (q_, q,J ), we find when j - oo
that the integral

[ I n IK,(q,.,C) - h(OI1fC)lN(dC)
fo
tends to 0 when m -+ oo. We conclude that every convergent subsequence
of the sequence n - K,(q,,, ) has limit h, so that lim, K,(q,,, ) = h locally
uniformly on D. o
This discussion shows that if D is completed in the d, metric, the resulting
space DM is a compact metric space and that each point C adjoined to D
in the completion can be identified with a positive harmonic function, which
we denote by K,(C, ), identified by the fact that q -+ C in the DM metric if
and only if K,(q, ) -+ K,(C, -) locally uniformly on D. Moreover, integrating
to the limit shows that (2.2) is satisfied when q = C. The conditions (a)
and (b) of Theorem 3 are satisfied by K, and therefore also satisfied by an
arbitrary Martin function K,, in view of (3.1).

EXAMPLE (Balls and Half-Spaces). Let D = B(0, S), and let v be a unit mea-
sure supported by the origin. Then K,(q, 0) = I for q 0 0 and
4 Preliminary Representations of Positive Harmonic Functions and Their Reductions 199

S-Inl
ifN=2
G.(n, 0) (InI-'b),
N-2
SN_, (S - Inl) if N > 2

so that when n tends to { on the Euclidean boundary, the value K,(n, )


tends to a multiple [S-t if N = 2, (N - 2)6'-N if N > 2] of the inner normal
derivative of at C, and the Martin function K becomes the Poisson
kernel density denoted by K in Section 11. 1. Thus the Martin boundary of
a ball is the Euclidean boundary. Similarly, if D is a half-space, the Martin
function based on a suitably chosen reference point becomes the Poisson
kernel density denoted by K in Section VIII.9, and again the Martin boundary
is the Euclidean boundary. Alternatively the Martin boundary of a half-
space can be derived from that of a ball since a ball can be mapped onto
a half-space by an inversion. The same reasoning, based on the Riemann
mapping theorem, shows that the Martin boundary of any plane Jordan
domain is its Euclidean boundary and that more generally the Martin
boundary of an arbitrary simply connected plane Greenian open set other
than 682 is the Caratheodory prime end boundary.

4. Preliminary Representations of Positive Harmonic


Functions and Their Reductions
Lemma. Let D be a Greenian subset of 68" with Martin function K,. If u is a
positive harmonic function on D and if A is a subset of D with (8A) n ar = 0,
there is a measure on DM, supported by the boundary of A relative to DM,
for which (reduction relative to D)

DUD" = KV({, Z. t(DM) = lu8" dv. (4.1)


J DM fD

In particular, there is a measure A. (= supported by aMD for which

u=f K,,(C,) (d{), .(aMD) = udv. (4.2)


"MD JD

A measure A. on 3MD satisfying (4.2) must also satisfy

ouOF = OKr(t,')O'(4) (4.3)


J aMD

for every Bore! subset F of 61D.


200 l.Xlt. The Martin Boundary

Observation. In general the measure J.,, is not uniquely determined by


u in (4.2), but it will be seen in Section 9 that A. can be chosen to be supported
by the set of minimal Martin boundary points (to be defined in Section 5)
and that if so chosen, A. is uniquely determined.
If A is relatively compact in D, the function Quo" is a potential, say GDA',
A' is supported by OA, and (4.1) is true with A(dC) = GDv(C);i (d1). If A is
not relatively compact in D, let B. be an increasing sequence of open subsets
of D, relatively compact in D and with union D, so large that o c B0.
According to what has just been proved, the potential Ju6B"^" can be repre-
sented in the form
Me""" = K,. ((,)A(d(), (4.4)
JOM

where A. is a measure supported by a(B" n A) and

(4.5)
JDM

The sequence A is a bounded sequence of measures on the compact space


DM. If A. is the limit of a vaguely convergent subsequence, it is trivial that
i is supported by the boundary of A relative to DM and that (4.4) yields
(4.1).
We shall only need (4.3) when Fis compact, but this equation has general
interest and so we prove it as stated. In view of the properties of the reduction
operation the class of sets F satisfying (4.3) includes limits of monotone
sequences of its members; so it is sufficient to prove (4.3) for compact
boundary subsets F. If F is such a set, let F; be a decreasing sequence of
compact neighborhoods of Fin DM with intersection F, and set F. = D n F,.
Then [from Section 111.5(a)] lim, urr" = our. Now

bn"(', u) = J KY((, (4.6)

and (4.6) yields (4.3) when n -+ co.

5. Minimal Harmonic Functions and Their Poles


Let D be a Greenian subset of IB" with Martin function K (= K,,). A Martin
boundary point (is called minimal if its associated function K({, ) is a
minimal harmonic function for one and therefore for every choice of the
Martin function K. The set of minimal boundary points is called the minimal
Martin boundary and is denoted by 8;"D. In the following theorem K is an
arbitrary Martin function for D. Recall from Section X.7 that a Martin
boundary point C is said to be a pole of a positive harmonic function u if
6. Extension of Lemma 4 201

OuO141 = u and that then C is also a pole of every positive harmonic minorant
of U.

7lteorem. (a) Every ( 0) minimal harmonic function on D has a unique


pole on aMD. If a Marlin boundary point C is the pole of a positive (0 0)
harmonic function u, then C is the only pole of u, u = const. K(C, ), and C
is a minimal boundary point. In particular, if n is a minimal Martin boundary
point, the function K(q, ) has pole p.
(b) If C is a minimal Martin boundary point and if A is a set of minimal
Martin boundary points, then O K(C, ) 0A is either K(C, ) or 0 according as
CeAorCOA.

Proof of (a). According to Section X.7, whatever the boundary assigned


to D by a metric compactification, a minimal harmonic function on D has
at least one pole. Now let u be a strictly positive harmonic function on D
with pole C. We can assume in proving (a) that K = K and that u has been
normalized so that Jo u dv = 1. Apply Lemma 4 with A the trace on D of
a neighborhood of C to derive

u = HA = J K(7,')2UA(d4), AUA(DM) = 1. (5.1)


DM

The measure 2gA is supported by the boundary of A relative to DM ; so


when A shrinks to C, the measure tends (vague convergence) to the
probability measure supported by the singleton {C}, and therefore (5.1)
becomes u = K(C,'). Thus C is uniquely determined by u. Moreover, as
already recalled at the beginning of this section, if v is a positive harmonic
minorant of u, then v also has pole C. Hence v = const. K(C,'), and therefore
u is a minimal function; that is, C is a minimal boundary point. Finally,
if PI is a minimal Martin boundary point, then according to what we have
just proved, K(q, ) has a pole C and K(rl, ) = const. K(C, ). The constant
must be l;soq=C. o
Proof of (b). If C is a minimal Martin boundary point, the function K(C, )
has pole C according to (a) ; so surely 8 K(C, ) IA = K(C, -)if Cc-A. Furthermore
the set B = 8MD - {C} is open in aMD and contains no pole of K(C, ); so
(from Section X.7) Q K({, ) UB = 0, and therefore O K(C, ) QA - 0 if A (-- B. o

6. Extension of Lemma 4
Lemma. (Context of Lemma 4). If F is a compact subset of aMD, there is a
measure A. supported by F such that

BUOF = JK(c.)A(dc) = IF (6.1)


202 LXII. The Martin Boundary

Let F; be a decreasing sequence of compact neighborhoods of F in DM


with intersection F, and set F. = D n F,;. Then [Section III.5(a)] lim, OuOF"
= and according to Lemma 4, there is a measure Y. supported by the
boundary of F. relative to DM such that

µ"(D" ') <_ Judy.

If AF is the vague limit of a convergent subsequence of p,, the measure


),F is supported by F, and the first equation in (6.1) is true. Next apply the
operator SD" to the first equality to obtain

IIU F OF. =
U aD"(',
F

Jr
which yields the second equation in (6.1) when n -, oo in view of the idem-
potency of the smoothed reduction operation.
Application. A Martin boundary point C is nonminimal if and only if
OKy(i;, ) Qic1 - 0. In fact, according to Lemma 6 with u = K,,(C, ) and F = {C},

K,(C,') O' =

c is either 1 or 0 because the smoothed reduction operation is idem-


potent. Since (by Theorem 5) the condition c = 1 characterizes minimality,
the condition c = 0 must characterize nonminimality.
The following theorem strengthens this result.

7. The Set of Nonminimal Martin Boundary Points


Theorem. The set of nonminimal Martin boundary points of a connected
Greenian subset D of ll' is an F, set that is h-harmonic measure null for every
strictly positive harmonic function h on D.

Let h be a strictly positive harmonic function on D, let Co be a point of D,


and let K be a Martin function for D. If B is an open subset of DM, define

if = jCeamD: S K(CCo)t.

Since a smoothed reduction on B is obtained by applying the operator SDnD,


an application of Fatou's lemma shows that B' is compact. Moreover, if A
8. Reductions on the Set of Minimal Martin Boundary Points 203

is a compact subset of the F, set B' n B,

K(C2Co) (CeA). (7.1)


OK(C,')QA('o) 5 OK(C,.)i8(bo)

An application of the representation (6.1) to OhOA,

OhOA(Co) = JK(C. o).AA(d) = DK(C, ) OA( o)2hA(dC)


JA
(7.2)
5 h A(Co)
2 '

shows that Oho" _- 0; that is, A is h-harmonic measure null. The application
in Section 6 implies that the points of B' n B are not minimal and that when
B runs through the open sets of a countable topological base of DM, the
class of F. h-harmonic measure null sets B' n B covers the set of nonminimal
Martin boundary points. The theorem follows.

8. Reductions on the Set of Minimal Martin Boundary Points


Lemma. If h is a strictly positive harmonic function on the connected Greenian
subset D of U2", then

h = OhO4"1 = sup { Oho": A c a rD, A compact}. (8.1)

It is sufficient to show that the first and third terms in (8.1) are equal
because the second term ties between them. There is [from Section VIII.5(b)]
a positive h-superharmonic function u on D that has limit + oo at every point
of the h-harmonic measure null set of nonminimal Martin boundary points.
Define

A,= limyinfu(n) 5 n}, A' = UA (nEZ').


q4 0

The set A. is a compact set of minimal boundary points, and since [from
Section V1.3(e)] Jim,,..,, Oho"_ = OhoA" it will suffice to prove that OhO"' = h.
Since u has limit + oo at every point of the set B = 3MD - A', the set B is
h-harmonic measure null [Section VIII.5(b)]; so OhO5 = 0, and by sub-
additivity of the set function OhO',
h = OhVA'vB S OhoA' + OhOB = Oho'' 5 h;

so OhO` = At, as was to be proved.


204 I.XII. The Martin Boundary

9. The Martin Representation


Let D be a connected Greenian subset of U8". It will be convenient to expand
the vector lattice notation in Chapter IX by introducing into the notation
the relativizing strictly positive harmonic function : S, Sm, ... will be written
'S, when h is the relativizing function. For example, "Sqb is the
class of quasi-bounded h-harmonic functions on D.
If K, is a Martin function for D and if A,, is a signed measure on DMD, the
function u defined by (4.2) is harmonic on D because u is continuous and has
the harmonic function average property. In view of the Jordan decomposi-
tion of 1.,, the function u is in the class 'Sm. Conversely, according to Lemma
4, a positive harmonic function on D, and therefore also a harmonic function
in 'Sm, has a representation (4.2) in terms of a not necessarily uniquely
defined signed measure on DMD. The following Martin Representation
Theorem details among other things the relation between harmonic func-
tions in 'Sm and their unique Martin representing signed measures on D','D.

Theorem. Let D be a connected Greenian subset of R, let K be a Martin func-


tion for D, and let h be a strictly positive harmonic function on D.
(a) To each function v in 'Sm corresponds a unique finite-valued signed
measure M. on PD, supported by the minimal Martin boundary
a"D, positive if v is, and .satisfying

v= f K(C, (9.1)
JaMD

(b) For given K the correspondence is an isomorphism between


the vector lattice Sm and the vector lattice of signed measures on DMD
supported by D.
(c) A function v/h in "Sm is in ""S,Mb['S,,,,] if and only if M,, is absolutely
continuous [singular] relative to M,,. In the quasi-bounded case

v= (9.2)
!MD

See Section 10 for the relation between Martin representing signed mea-
sures and harmonic measures.

Uniqueness proof Suppose that v is a positive harmonic function on D and


that there is a measure M. supported by DrD for which (9.1) is satisfied.
Then according to Lemma 4 and Theorem 5(b), if A is a Borel boundary
subset,
9. The Martin Representation 205

J OK(C,') Mo(de) = J K(C, (9.3)


aMD A

Hence f D I v I" dv = M(A); so the measure M. is uniquely determined by v.


If v e IS,,, and if v has two representing signed measures, M and M., then the
function 0 has the representing signed measure M,, - M. The two positive
measures whose difference is M. - M (Jordan decomposition) are therefore
representing measures for the same positive harmonic function on D and so
must be identical according to what we have just proved ; that is, My = M.

Proof of (a). Let v be a positive harmonic function on D, and let A be a com-


pact subset of the minimal boundary, so that according to Lemma 6, there is
a measure A and satisfying

OvOA = JK(c.)(dc)(9.4)
The measure AVA was just shown to be uniquely determined. If B is a compact
subset of A, the equality vD" OB = 0v 0B implies by Lemma 4 and Theorem
5(b) that

JOK(1)AA(dC) = J K(C,
B

Hence AvB = AVA on the Borel subsets of B. Let A. be an increasing sequence


of compact subsets of the minimal Martin boundary, chosen so that (Lemma
8) Jim,,.., VBAn = v at some point of D, which implies (by the Harnack con-
vergence theorem) that this limit relation is true locally uniformly on D.
According to what we have just proved, A,,A = Av,,n*, on the Borel subsets
of A. The increasing sequence A,,A, of measures has limit M, a measure
(Appendix IV.4) of Borel sets supported by the minimal Martin boundary.
If A in (9.4) runs through the sequence A,, this equation becomes (9.1) in the
limit.

Proof of (b). The uniqueness property has already been proved. The relation
v/h «-. M. is obviously linear and is specific order preserving because v z 0 if
and only if M z 0. The vector lattices in question are therefore isomorphic.
13

Proof of (c). The assertions of (c) follow from the vector lattice isomorphism
just derived. On the one hand, ASb is the subband of "5,,, generated by the
function 1, and "S,,,, is the orthogonal complement of "Sb in''Sm; that is,
'S., is the subband of AS,,, orthogonal to the function 1. Equivalently,''S,b
206 1.XII. The Martin Boundary

is the subband of''Sm consisting of the class of functions v/h with v in the
subband of'Sm generated by h, and 'S.,, is the subband of 'S. consisting of
the class of functions v/h with v in the subband of 'S. orthogonal to h. On
the other hand, it then follows from (b) that v/h is a quasi-bounded h-har-
monic function if and only if M,, is in the band generated by Mh of signed
measures (charges) on BMD supported by 8,"D, that is, if and only if M is
absolutely continuous relative to M,,. Furthermore v/h is a singular h-har-
monic function if and only if M 1 M", that is, if and only if Mn is singular
relative to M".

The Martin Representation of a Minimal Harmonic Function

If u is a not identically 0 minimal harmonic function on D, then u = cK(q, )


for some uniquely determined minimal Martin boundary point p. In fact,
more generally we now show that if u is a not identically vanishing minimal
harmonic function on D and if

u=J K(C,
'UD

for some measure A,, on 0 'D, then A. must be supported by a uniquely deter-
mined singleton {p}. To prove this, observe that if A is a Borel subset of
3MD, then JA K(C, -).t (dd) is a positive harmonic function majorized by u
and therefore proportional to u. If now C is in the compact support of A. and
if A is the trace on 8'D of an open Martin topology neighborhood of C and
shrinks to C, it follows that

K(C, -) = c(C)u, I = c(C) f udv.


D

Hence is a constant function on the support of A,,, K(C, -) is the same


function for all C in this support, and therefore A. is supported by a singleton
{ry}. The point ry is minimal and therefore (Theorem 9) uniquely determined
by u. In particular, if u = K(p', ) for some minimal boundary point rj ,
it follows that the point ri must be n'.

The Notation M,

Let v be a positive superharmonic function on a Greenian subset D of


1", and let v, be the harmonic component of the Riesz decomposition
of v. For a given Martin function K the function v, determines a unique
measure M,,, on 8;"D, and we define M,, = Ma,. A glance at (3.1) as extended
to the Martin boundary shows that M,, and any other measure on the minimal
Martin boundary induced by a different choice of K are mutually absolutely
10. Resolutivity of the Martin Boundary 207

continuous. Slightly more generally, if ve'S, we define M. as the Martin


representing signed measure on O 'D of the 'S. component of v.

10. Resolutivity of the Martin Boundary


Theorem. The Martin boundary is universally internally resolutive and uni-
versally resolutive. If K is a Martin function and It is strictly positive and
harmonic on D, dC) = K(C, g)M"(dC)/h(C). An h-harmonic function u =
v/h is a PWB h solution if and only if it is quasi bounded, equivalently, if and
only if M. is absolutely continuous relative to M, and then

u = Hf = f f(C)K(C, -) . (10.1)
JJMD
hM (d , f = dM,,

'J
Let h be a strictly positive harmonic function on D, and let A be a closed
subset of CMD. Apply (4.3) and Theorem 5 to derive

=1L = f O K((, .)DA


M hdC) = fKM'0. (10.2)
aUD

The function A i- HiAA is therefore additive and (Section VIII.9) h-resolutivity


of 8"D follows, and also the evaluation dC) = K(C, )M"(dOlh. Ac-
cording to Section IX.9, every PWB' solution is quasi bounded. Conversely,
if u = v/h is a quasi-bounded h-harmonic function, equivalently (Theorem
9), if M is absolutely continuous relative to M", and if f = dM/dM",

v= K(C ,-)1(C)M"(dC) = uD(-,f)h, (10.3)


IMD

so that u is the PWB" solution for the boundary function f; that is, (10.1)
is true. Thus the Martin boundary is universally internally resolutive as
well as universally resolutive, and the proof of the theorem is complete.
Intrinsic Definition of h-Harmonic Measure. If u = v/h is h-harmonic on D,
if

05u:9 1, uA(1-u)=GMo[un(I-u)]=0, (10.4)

that is, if for f =

0 < f:5 1, f n (I - f) = 0 M" almost everywhere (10.4')

then f coincides M" almost everywhere on O"D with the indicator function
of a Borel set A, and u = uD(-, A). Conversely, if u is the h-harmonic measure
208 I.XI1. The Martin Boundary

of a Borel boundary subset A, the reverse argument shows that (10.4) is


satisfied.

Relations between Martin Representing Measures and Harmonic


Measures

According to Theorem 10, the measures j4(C, ) and Mh are mutually


absolutely continuous for all C in D; that is, a boundary subset is µD null
if and only if it is Mh null. In fact the equality 4D(C, dC) = K(C, C)Mh(dC)/h(C)
implies more: a boundary function is Mh measurable and integrable if and
only if it is lzD(C, ) measurable and integrable for every (equivalently a
single) point S of D.
Special Case: h Is Minimal. If C E 8;'D and if h = K(C, ), then 0 , ) = Mh
for every point C of D, and this measure is the unit measure supported by {C).

11. Minimal Thinness at a Martin Boundary Point


Theorem. Let D be a Greenian subset of R", let K be a Martin function for D,
let A he a subset of D, and let C be a minimal Martin boundary point of D.
(a) The following conditions are equivalent:
(al) RK(c..) = K(C, )
(a2) RKI;B, = K(C, -)for every Martin topology neighborhood B of C.
(b) The following conditions are equivalent:
(bl)
(b2) inf {RA B': B is a Martin topology neighborhood of C} = 0.
(b3) RA is a potential.
(c) If B is a Martin topology neighborhood of C, then
(c l) R.1 = K(C, ),
D-B
(c2) R,,( ,.) # K(C, )
Each condition (al), (a2), (bl), (cl), (c2) is satisfied if and only if it is
satisfied using the corresponding smoothed reduction.

The set A is said to be minimal thin at C if the conditions (b) are satisfied.
The last assertion of the theorem is trivial, and the proofs will be phrased
accordingly. Observe that in view of the application in Section 6 if C is a
nonminimal Martin boundary point, condition (b2) is satisfied because
the indicated infimum is
The proof of the theorem will be carried through in several steps, num-
bered for convenience in reference.

Proof. Step I. Proof that RR({ is either K(C, ) or a potential. Since K(C, )
is minimal, the Riesz decomposition of RKI{ must have the form
1 I . Minimal Thinness at a Martin Boundary Point 209

v + cK(C,

where v is a potential and c is a positive constant. If we use the fact that


the smoothed reduction operation is idempotent, we find that

RA
+K(C. ) - R A + cR A
+V +K(c. (11.2)

and since a function majorizes its smoothed reduction, it follows that the
terms on the right in (11.1) and (11.2) are pairwise equal. Hence either
c = 0 and R"
+K((, )
is a potential or c > 0 and R"
+K({. )
= K(4, ), in which case
c=Iandv=0.
Step 2. Proof of (c2). Without loss of generality we can assume that B
is so small that the compact support of the measure on which K is based
does not meet B. According to Lemma 4, there is a measure A on DM, sup-
ported by a(D - B) (boundary relative to DM), such that

RD-B = fDM K(n, )A(dp).


J

If there were equality in (c2), the integral would define a harmonic function
on D; so the measure A would be supported by OmD. However, according
to Section 9, such an integral representation of a minimal harmonic function
K(C, ) is possible only if A is supported by {C}, contrary to the definition
of B. Hence there cannot be equality in (c2).
Step 3. Proof that (a 1) ca (a2). The implication (a2) (a 1) is trivial. To
prove the reverse implication observe that if (al) is true and (a2) is false,
then R"^B is a potential for sufficiently small B by Step 1, and R'-' is a
potential according to Step 2, because this smoothed reduction is a positive
superharmonic function majorized by the potential R'-,'. Hence by set
subadditivity of reductions is a potential, contrary to hypothesis.
Step 4. Proof of (c 1). Assertion (c 1) is trivially true when B = D and
therefore true for arbitrary B by (a), which we have just proved. Alternatively
(cl) is true because [by Theorem 5(a)] C is a pole of K(C, ).
Step S. Proof that (bl) (b2). The function Rg(GB) is harmonic and
positive on D - B. Let B shrink to C, say along a sequence of balls of center
C and radii tending to 0 (in terms of some Martin space metric). Then the
limit of the corresponding sequence of reductions in (b2) is the indicated
infimum and is a positive harmonic function, majorized by RK"(S ). Since
this smoothed reduction is a potential by Step 1, the harmonic function
vanishes identically, as was to be proved.
Step 6. Proof that (b2) . (b3) . (b1). These implications follow trivially
from Step 1 and the equivalence of (al) and (a2).
210 1 X[[. The Martin Boundary

The proof of the theorem is now complete, and we turn to the definition
of the minimal-fine topology of DM.

12. The Minimal-Fine Topology


Let D be a connected Greenian subset of R", let K be a Martin function
for D, and let C be a minimal Martin boundary point. The class MT(O of
subsets of D minimal thin at C has the following properties.
(P1) Every subset of a set in MT(h) is itself in MT(C) (trivial).
(P2) A finite union of sets in MT(C) is in MT(C), because the set function
A )-4 is subadditive.
(P3) If A E MT(C), then the union A' of A and its set of fine limit points
in D, as defined in Section XT. 1, is in MT(h), because according
to Section XI.7,

RA = RA'
+K({..).

(P4) Every set in MT(C) has an open superset in MT(i;), because ac-
cording to Section 111.5(e),

RK(i.) =inf{RK((.):Bopen, AcBcD).

The Minimal-Fine Topology

We define the minimal-fine topology of DM by the following conventions:


A point C of D is a minimal-fine limit point of a set A if A r D is not
thin at C.
A point C of a'('fD is a minimal-fine limit point of a set A if A n D is
not minimal thin at C.
Each nonminimal Martin boundary point is a minimal-fine isolated
point of DM.
The minimal-fine topology of DM has as relative topology on D the
fine topology already defined on D in Section XI. I. According to Theorem
1 l (c), if B is a Martin topology neighborhood of the minimal Martin
boundary point C, then B n D is a deleted minimal-fine neighborhood of
C, and D - B does not have C as a minimal-fine limit point. Thus the minimal-
fine topology of DM is a (Hausdorff) topology finer than the Martin topology
of DM.
For some choices of D, for example, when D is a ball (Section 3). the
Martin space DM can be identified with the Euclidean closure of D. Observe
that for such a choice of D if C is a boundary point and if A is a subset of
12. The Minimal-Fine Topology 211

D, then thinness of A at C in the fine topology of R' need not be equivalent


to minimal thinness of A at C. For example, if D = B(0, S), then every
boundary point is a minimal Martin boundary point, and if B is a ball
internally tangent to BD at C, then D - B is minimal thin at C but is not
thin at C. In fact, in this case K(C, ) is a constant multiple of the function
62 - 112
1{ 01

(Section 11. 16), and D - B is the locus of the inequality K(C, -) 5 c for some
strictly positive constant c. Since we shall prove [equation (12.3)] that the
minimal-fine limit of K(C, ) at C is + oo, the set D - B is minimal thin at C.
The set D - B is not thin at C because it contains the trace on a neighborhood
of C of an open cone with vertex C (see Section XI.3). We shall use the notation
"mf lim" for minimal-fine limits.

EXAMPLE (a). Let D be a Greenian subset of RN, let C be a point of D, and


define Do = D - {C}. Then (from Section VII.1) GD° is the restriction of
GD to Do x Do, and (from Section VII.10) the restriction of GD(C, ) to Do
is a minimal harmonic function on Do. We conclude that the point C can
be identified with a minimal Martin boundary point of Do, the pole of the
restriction of GD(C, ) to Do. The Martin topology of Do coincides on D
with the Euclidean and Martin topologies of D. Thus the Martin space
Do can be identified with D". Finally, the minimal-thin topology of Do
on a Martin neighborhood of C is identical with the fine topology of DM
and of R" on that set. Hence minimal-fine limit concepts on Do at C coincide
with fine limit concepts on D at C. More generally, a trivial refinement of
this reasoning shows that if A is a closed relative to D polar subset of D
and if Do = D - A, the Martin space Do can be identified with D" by
identifying each point C of A with a point of BMDo, the pole of the restriction
to Do of GD(C, ); minimal-fine limits on Do at C coincide with fine limits
on D at C. Finally, suppose that v is a positive superharmonic function on
D whose Riesz measure v is supported by A. The function v is harmonic
on Do and thus has a Martin representation there in terms of a measure
Mo° on O "Do = A v arD. Choose a Martin function K° for Do based on a
point o in Do. Then it is clear that for C in A,

GD(C,Q
K°(C,C)Mo°(dC) = Mo°(dC) = GD(4,Ov(dO,
GD(4o, 0

so that Mo°(dC) = GD(40,C)v(dC) for C in A. It is important that Mo,, and v


are mutually absolutely continuous on A.
Minimal--Fine Limits at an Isolated Boundary Point. Example (a) implies
that to each theorem on minimal-fine limits at a minimal Martin boundary
212 1.XII. The Martin Boundary

point of a Greenian set D corresponds a theorem on fine limits at a point


of D. For example, the fact that (Section XI.1) a set thin at a point of D
has an open superset thin at the point corresponds to the fact (P4) that a
set minimal thin at a minimal Martin boundary point of D has an open
superset minimal thin at the point. The fact that [Theorem XI.4(a)] if v is
a positive superharmonic function on D and if C is in D, then v/GD(C, )
has fine limit the value infD_it, v/GD(C, -) at C corresponds to the fact that
if Z is a minimal Martin boundary point of D and if K is a Martin function
for D, then v/K(i;, -) has fine limit the value infD v/K(C, -) at C. The latter
result is proved in Section 13, and a dual result is proved in Section 14.

EXAMPLE (b). Denote by dt the Nth coordinate of the point of R', and
define D = { : dc > 0}. Then (from Section 3) DM is the closure of D in
the one-point compactification of RN ; so the Martin boundary is the
Euclidean boundary. If u is a positive superharmonic function on D, if
c > 0, if u, = u(/c), if A is a subset of D, and if cA has the obvious meaning,
then RA In particular, if u() = L-Ndd, that is, if u is a minimal
harmonic function on D with pole the origin, then uc = cN-'u and

RYA(D = C' _NRY C

If A is relatively compact in D and not polar, then Rr is a nonzero potential,


and the evaluation of Go in Section VIII.9 shows that RA(cC)
c 0, with ¢ a strictly positive finite-valued function. It follows that
if c is an arbitrary sequence of strictly positive numbers with limit 0 and if
B is the intersection of Uo (c"A) = A' with a Euclidean neighborhood of
the origin, then

RB z RM"A(C) >
2

for sufficiently large n. Hence A' is not minimal thin at the origin.

A trivial example of the application of Example (b) shows that if N = 2,


no initial segment of a ray from the origin into D is minimal thin at the origin.
An analogous argument shows that if C' is a boundary point of a disk D',
then no initial segment of a ray from C' into D' is minimal thin at C'.
In view of Example (b) if a function from the upper half-space of R'
into a Hausdorff space has both a minimal-fine and a nontangential limit
at a boundary point, the two limits must be the same. An analogous argument
leads to the same conclusion if D is a ball.
14. Second Martin Boundary Counterpart of Theorem X1.4(c) 213

13. First Martin Boundary Counterpart of Theorem XI.4(c)


and (d)
Theorem. Let D be a connected Greenian subset of R', let K be a Martin
function for D, and let C be a minimal Martin boundary point of D.
(a) If v is a positive superharmonic function on D, then

v(7) = 1Df v = )
K((,7) K((,') M-(")

(b) Let A be a subset of D with Martin topology limit point C. If A is


minimal thin at C, there is a positive superharmonic function v on D
for which

lim v(7) - +cc. (13.2)


A.o-K K(C,7)

Conversely, if there is a positive superharmonic function v on D


satisfying (13.2), then A is minimal thin at C.

We shall see (in Section 19) that Theorem 13(a) is a special case of the
Fatou boundary limit theorem for a Martin space.
To prove Theorem 13, translate the proof of Theorem XI.4(c), (d) into
the present context, replacing GD(C, ) in that proof by K(C, ) and "thin"
by "minimal thin." See Example (a) in Section 12 for a discussion of the
relation between theorems on limits at a minimal Martin boundary point and
limits at a point of D.
Application. If v is a potential in (a), we find that the minimal-fine limit
is 0. If v =_ I in (a), we find that

mf lim K(C, PI) = sup K(µ, ). (13.3)


q-c D

14. Second Martin Boundary Counterpart of Theorem XI.4(c)


Theorem. If D is a connected Greenian subset of RN, if v is a strictly positive
superharmonic function on D, if C e BD, and if e D, then

0 < mf lim v(7) = lim inf v(7) S + oo. (14.1)


7) n-c 7)

Observe that (14.1) is trivial if the indicated inferior limit is + oo. We can
therefore ignore this case and prove the equality in (14.1) by showing that
214 1.X11. The Martin Boundary

whenever c is a finite number strictly larger than the inferior limit in (14.1),
it follows that the set A = {n: v(n) z cGD(4, n)) is minimal thin at C. Let K
be a Martin function for D based on the point . By definition of the smoothed
reduction on A the inequality v z cOGD(4, .) DA is valid on D; so in view of
sweeping symmetry, if n e D -

v(n)
(14.2)
GD((, h n) n)'

and therefore since OK(n, ) r() = bo ((, K(n, -)), Fatou's lemma is appli-
cable when n C in (14.2) and yields

cOK(S, )OA( ) = cbn"( , K(C, )) 5 lim of Gn(i)n) < c. (14.3)


I-C

Hence OK(C, I = K(i(, ); so (from Section 11) the set A is minimal


thin at (, as was to be shown in proving the equality in (14.1). To show that
the minimal-fine limit in (14.1) is strictly positive, it can be assumed that v
is a strictly positive potential GDV, after replacing v if necessary by its re-
duction on a ball relatively compact in D. Under this hypothesis,

lim of GD(S)) fD K((, (')v(d(') > 0,


q-c

as was to be proved.
Special case: v =_ 1. If GD((, ) has minimal-fine limit 0 at i;, as we shall
prove (Section 18) is true at µD almost every minimal Martin boundary
point (, it follows from Theorem 14 that has limit 0 at Con approach
to C in the Martin topology. [Incidentally, this application of Theorem 14
to the function exhibits the fact that the minimal-fine limit + oo
cannot be excluded in (14.1).] This vanishing of the Green function at
the Martin boundary (to be extended by relativization in Section 18) is
one indication that the Martin boundary is well adapted to classical potential
theory.

Relation between Theorems 13, 14, and XI.4

The fact that the Laplacian is a self-adjoint differential operator leads to


the symmetry of the Green function GD, absent in the potential theory
generated by a non-self-adjoint differential operator. In such a theory the
counterpart of Theorem XI.4 splits into two theorems. See Section XVIII.14
for the versions of Theorem XI.4 in the potential theory corresponding
to the heat equation and its adjoint. The self-dual character of classical
15. Limits at a Minimal Martin Boundary Point 215

potential theory is lost at the Martin boundary of a Greenian domain,


however, and in fact Theorems 13 and 14 are dual to each other. This is
suggested by the fact that the proof of Theorem 14, unlike the proof of
Theorem 13, uses the sweeping symmetry of classical potential theory.
In the probabilistic versions of these theorems (Theorem 3.111.5) Theorem
13 states that the function v/K(C, ) has the indicated limit at C along almost
every Brownian path conditioned to go from a point of D to C, whereas
Theorem 14 states that v/GD(1;, ) has the indicated limit at C along almost
every Brownian path conditioned to go from C to a point of D. If the two
points of D here are taken to be the same, these conditional Brownian paths
to C can be identified with those from C; so the same limit concept is involved,
corresponding to minimal-fine limits at C. In particular, in classical potential
theory if C is a point of a Greenian set D, then C can be identified with a
minimal Martin boundary point of D - {C} [Section 12, Example (a)], the
corresponding minimal harmonic function is a multiple of GD(C, ), and
Theorems 13 and 14 coalesce to Theorem XI.4(c).

15. Minimal-Fine Topology Limits and Martin Topology


Limits at a Minimal Martin Boundary Point
Let D be a connected Greenian subset of I8", let K be a Martin function for
D, and let C be a minimal Martin boundary point. The following lemma is
the analog of Lemma XI.8 in the present context.

Lemma. Let A be a decreasing sequence of subsets of D, and suppose that each


set Ak is a deleted minimal-fine neighborhood of C [has minimal-fine limit point
C]. Then there is a subset A of D with the property that for each k the part of
A in a sufficiently small Martin topology neighborhood of C is in Ak and that A
is a deleted minimal-fine neighborhood of C [has minimal-fine limit point C].

The proof is similar to that of Lemma XI.8; so only the unbracketed


assertion will be proved. Suppose that each set Ak is a deleted minimal-fine
neighborhood of C, and define Fk = D - Ak. The set Fk is minimal thin at C,
and it is sufficient to show that there is a set F, minimal thin at C, with the
property that for each k the part of FA, in a sufficiently small neighborhood of
C is a subset of F. If for all but a finite number of values of k the point C
is not a limit point of Fk, we can take F as the union of those sets Fk for
which C is not a limit point of Fk. Otherwise, it is no restriction to assume that
C is a limit point of F. for all k. Let B. be the intersection with D of a ball
(in terms of some metric on DM) of center C and radius r. Let Co be a point
of D, and applying Theorem 13, let uk = vk/K(C, ) for keZL+ be a positive
K(C, )-superharmonic function on D with u(C0) < 2-k and with limit + 00
at C on approach along Fk. The function u = Eo uk is positive and K(C, )-
216 LXII The Martin Boundary

superharmonic on D with limit + oo at C on approach along each set Fk.


Choose rk so that limk-m rk = 0 and that u Z k on Fk n B,k. If

F= U(FknB,k),
0

the function u has limit +oo at C along F; so [by Theorem 13(b)] F is


minimal thin at C, and Fk n B,k c F, as desired.

16. Minimal-Fine Topology Limits and Martin Topology


Limits at a Minimal Martin Boundary Point (Continued)
Theorem. Let D be a connected Greenian subset of RN, and let C be a minimal
Martin boundary point of D. A function u from the trace on D of a Martin
topology neighborhood of C into a metric space has minimal-fine limit [minimal-
fine cluster value] a at C if and only if u has limit a at C on approach along
some subset of D that is a deleted minimal-fine neighborhood of C [is not
minimal thin at C].

The proof is that of Theorem XI.9 with trivial changes corresponding to


the change of context. Observe that if the range space of u is E8, then
mf lim sup,_, is a minimal-fine topology cluster value of u at C, and
therefore u has this value as a limit on approach to C along some subset of D
that is not minimal thin at C; that is, this subset has C as a minimal-fine
topology limit point.

17. Minimal-Fine Martin Boundary Limit Functions


Let D be a connected Greenian subset of D. If A is a subset of D, denote by
A'"f the set of minimal-fine limit points of A in DM. Recall (Section 12) that
Dmfn aMD = D.

Lemma. (a) If A is a subset of D, the set A'"f is a Martin topology Ga subset


of DM.
(b) If u is a function from D into 08, the function lim sups-t u(l; )
on the minimal Martin boundary arD is Borel measurable (Martin topology
of DM).
(c) If u is a Junction from D into a compact metric space, the set of
Martin boundary points at which u has a minimal-fine limit is a Borel set,
and the limit function on this set is Borel measurable (Martin topology of
DM).
17. Minimal-Fine Martin Boundary Limit Functions 217

Proof of (a). According to Theorem XI.6, the set Amf n D is a Euclidean Gs


set, and this set is therefore a G8 set in the Martin topology. To prove that
A'f n aMD is a Ga set, let K be a Martin function for D, and let o be a point
of D. If B is a subset of D, let B' be the class of Martin boundary points C
satisfying the inequality OK(C, ) 5 K(l, X0)/2. Since the smoothed re-
duction on the left is K(C, )), Fatou's lemma implies that the set B" is
compact. Let B. be the sequence of traces on D of the sets of a countable
topological base for DM. If A is a subset of D, the set Uo (A n Be)' is an F.
set, the set of Martin boundary points that are not minimal-fine limit points
of A.

Proof of (b). Assertion (b) follows from (a) because


l
{Ea'D: s c}= n {eD: c+ n}mf n0 MD. (17.1)
C-C

Proof of (c). According to (b), the boundary set on which u has a minimal-
fine limit is a Borel set if u is extended real valued, because the set in question
is the set on which the minimal-fine superior and inferior limits of u are equal.
Moreover (b) implies that the limit function on this set is Borel measurable.
If the range space S of u is compact metric and if 0. is a sequence of functions
dense in C(S) in the metric of uniform convergence, then the map
{O (t)/sups n e V+} is a one-to-one bicontinuous map of S onto a com-
pact subset of the compact metric space [0,11 z; so an application of (a) and
(b) to each function 0.(u) yields (c).

EXAMPLE. If h is a strictly positive harmonic function on the connected


Greenian subset D of 08" and if A is a subset of D, then
A)
GMD(''RI) = GMh(R = ND(-. Amf n aMD). (17.2)

Observe that the first equality is trivial and that the second equality asserts
that the function h1D(-, Amf n aMD) is the harmonic component of RA (Riesz
h
decomposition). To prove this assertion, apply the kernel operator 6D to the
Martin representation of h to find

+h h K(g, -) Mh(dd)
LD +RtC.'1 Amf 0 M D
(17.3)
Mh(dC).
+ R' I
aMD-Amt
218 I. X11. The Martin Boundary

The first integral after the second equality sign is equal to Af n aMD)
according to Section 10. According to Theorem 11, the function RA is
a potential whenever C is not minimal or is not in Am!, so the last integral
is a potential. Thus (17.3) displays the Riesz decomposition of R' and
thereby yields the second equality in (17.2).

Observe that according to this example, the function Oh Q" is a potential


if and only if A'f n aMD is y 4 null.
Application. If D is a connected Greenian subset of R, if h is a strictly
positive harmonic function on D, if u is a positive h-superharmonic function
on D, and if

B2 _ {(earD: mflimsupu(ry) > c},


r[
then u > cµ4(-, B2). We can suppose in the proof that c > 0. If A, a D:
a}, then according to the preceding example and the relation (Section
VIII.2) between reductions and h-harmonic measure, if 0 < a < c,

u>"RA8>a"R A0 (17.4)

Let a tend to c to obtain the desired inequality.

18. The Fine Boundary Function of a Potential


Theorem. If D is a connected Greenian subset of I3" a superharmonic h-
potential, u = G0p/h has minimal-fine limit 0 at µ4 almost every (equivalently
Mh almost every) point of 3MD.

Define AL = {Cea"D: mflimsup,-t > s}. Then according to Section 17


(Application), u >_ sµ4(-, A), and this is impossible unless A, is 1<4 null,
because GMDu = 0. Hence the theorem is true.
Application. According to Theorem 18, the function has
minimal-fine limit 0 at p4 almost every point of tMD, and it then follows
from Theorem 14 that

lim 0 (µ4 a.e. C in alD). (18.1)


,,-c h(n)

lim GD('q) = 0. (18.1')


R-c ,l)
19. The Fatou Boundary Limit Theorem for the Martin Space 219

19. The Fatou Boundary Limit Theorem for the Martin Space
Let D be a connected Greenian subset of R. In this section the vector
lattice notation of Chapter IX will be used, as further developed in Section
9 of the present chapter. For example, when It is a strictly positive harmonic
function on D, we denote by i5+ the cone of positive h-superharmonic
functions on D. Let K be a Martin function for D. There is then (from Section
9) an isomorphism v -+ M between ' S1, and the vector lattice of signed
measures on tMD supported by the minimal Martin boundary aMD. When
v IF IS, it will be convenient to denote by M the Martin representing signed
measure of the component of v in 'S.; when this signed measure
ve'S+,

M is the Martin representing measure of the harmonic component of v


in its Riesz decomposition. Recall that dMpjdM,, under our conventions is
the Radon-Nikodym derivative of the absolutely continuous component
of M, relative to M" and that (Section 10) "M" almost everywhere" for h
in 'S' is equivalent to " µo almost everywhere"; both M" and µo are sup-
ported by a'D.

Theorem. If u = v/h is in 'S, then

mflimu(i) = dM°°(%) (M,, a.e. { on tMD). (19.1)


e--4 dM,,

In particular :
(a) This boundary limit function vanishes M,, almost surely if u e "So
(b) If u OS., this boundary limit function vanishes M" almost surely
if and only if u e "S,,,,.
(c) If u is a PWB" solution, that is, if u = µo(-,f) for some h-resolutive
boundary function f, then f = M" almost everywhere on
0MD; so f is the minimal-fine boundary limit function of u up to an
M" null set.

According to Section 10, the Martin boundary is universally resolutive


and universally internally resolutive, and the class of PWB" solutions is
IS,,. In view of this fact and of the decomposition ''S = "Sa + "S + "S,,
together with the vector lattice isomorphism between "S,, and the vector
lattice of signed measures on 3MD derived in Section 9 by means of the
Martin representation theorem, it will be sufficient to prove that if ue
"Sv v'S.., then u has minimal-fine limit 0 M" almost everywhere on aMD
and that if u = f ), then u has minimal-fine limit f(O at M" almost every
point C of 8MD. Theorem 18 implies that a function in 'SP has minimal-fine
limit 0 M" almost everywhere on 0MD and therefore that the same is true
for u in IS,. because according to Section IX.10, if u e "S,, then u A I E "SD .
220 l.Xll. The Martin Boundary

Finally, if u = "Hf =jut(-J), let u, and u2 be in the lower and upper PWB"
classes, respectively, on D for f. Then (by the application in Section 17) when
e > 0,

u2 - u1 z mflimsuP[u2(,) - u1(1)] Z
`` rS 8

J(.,{c: mflimsupu() -f() z (19.2)


z r{ JJ
eµo(.,I C: J(C) - mfliminfu(q) z r
l rC J

Since the left side of this inequality can be made arbitrarily small at any
point of D, it follows that mf u(q) = J(C) for p , almost every boundary
point C, equivalently, Ml, almost every boundary point C, as was to be proved.
Special cases. If h = K(C, ) is minimal, Theorem 19 states that v/K(C, )
has minimal-fine limit M"({C}) at C, as already proved in Section 13. If A
is a Borel boundary subset, the function pt(-, A) is the PWBh solution for
the boundary function 1A and therefore has minimal-fine boundary limit
function Mh almost everywhere equal to 1A.

Application to the Internal Fine Limit Theorem of Section XI.4

Let v and h be positive superharmonic functions on D with respective Riesz


measures v" and v,,, and define u = v/h at the points of D that are not in the
polar set of infinities common to v and h. By definition of the fine topology
the function u is fine continuous at every point at which it is defined. Ac-
cording to Theorem X1.4,

f lim u(q) - dv" (C) (19.3)


rC dv,,

at v,, almost every point C of the polar set of infinities of h, and we now show
that this limit result is a consequence of Theorem 19. Let F be a compact
subset of the set of infinities of h. To prove (19.3), it is sufficient to prove
that u has fine limit dv"/dv,, at v,, almost every point of F. In view of the Martin
compactification of D less a compact polar set [Section 12, Example (a)],
if v;, and v,; are the projections of vh on F and D - F, respectively, Theorem
19 implies that

f lim v(,) = (C), f lim


GDV,, (q) = dva (C)
GDv;(rl) dvA GDv;(q) dvp
20. Boundary Limit Theorems for Relative Superharmonic Functions on a Ball 221

at v almost every point t of F. Since the second Radon-Nikodym derivative


vanishes at v; almost every point of F, equivalently, at vh almost every point
of F, the first Radon-Nikodym derivative is equal to at vh almost
every point of F, equation (19.3) follows when h is a potential; the general
case then follows easily.

20. Classical versus Minimal-Fine Topology Boundary Limit


Theorems for Relative Superharmonic Functions on a
Ball in 08"

Recall that the Martin boundary of a ball is the Euclidean boundary. We


have proved two boundary limit theorems for a positive h-harmonic function
u on a ball D. According to Theorem 11. 15 and Theorem 19, at µD almost
every point of 6D the function u has a finite limit both in the nontangential
and minimal-fine topologies, necessarily the same limit (Section 12). The
limit function is a certain Radon-Nikodym derivative. Observe, however,
that Theorem 19 is more general in that Theorem I1.15 is for h-harmonic
functions whereas Theorem 19 is applicable to h-superharmonic functions.
Furthermore Theorem 19 is applicable to functions on an arbitrary con-
nected Greenian set D, but in this section we shall always assume that D is
a ball.
Classical-type approach to a ball boundary is not always applicable in
the general context of Theorem 19. In fact the following example shows that
Theorem 19 is false for D a ball under nontangential rather than minimal-
fine approach to the boundary.

EXAMPLE (a). Let v = GDp be a superharmonic potential with p supported


by a countable dense subset of the ball D, so that the h-superharmonic func-
tion u = v/h has nontangential limit superior + co at every boundary point
of D. Furthermore it is easy to choose p in this example in such a way,
depending on h, that u has nontangential (even radial) limit inferior 0 at
every boundary point of D.

The following discussion shows that radial approach to a ball boundary


allows a wider class of functions u in Fatou-type boundary limit theorems
than nontangential approach but does not allow as wide a class as minimal-
fine boundary approach. Let v = GDp be a superharmonic potential on the
ball D. According to Theorem 19, the function v has minimal-fine limit 0 at
uD almost every (equivalently, 'N_t almost every) boundary point. For N = 2
Littlewood showed that v has limit 0 along the radius to IN_t almost every
ball boundary point, and his result was later extended to N > 2. In view of
the Riesz decomposition theorem it follows that for h = I and D a ball
222 I.XII. The Martin Boundary

Theorem 19 is true for radial as well as minimal-fine approach to the bound-


ary. The following example shows that this assertion is false for general h.

EXAMPLE (b). Let h be a strictly positive minimal harmonic function on a


ball D, corresponding to a boundary point . Then (from Section VIII.9) uD
is supported by the singleton {C}, and Theorem 19 states that a positive
h-superharmonic function u = v/h has a finite minimal-fine limit at C. This
limit is info u according to Theorem 13 and is 0 (by Theorem 19) if v = GDv
is a potential. Since [Section 12, Example (b)] the radius to C is not minimal
thin at C, the function u has radial limit inferior 0 at C. However, if v is
supported by a countable dense subset of this radius, the function u has
radial limit superior + oo at C.

Thus the fine topology is better adapted to Fatou-type boundary limit


theorems for a ball than nontangential and radial approach topologies. The
difference is that the latter approach topologies are adapted to the ball rather
than to the properties of the functions involved. To show the relation
between Theorem 19 for a ball and the corresponding results for nontangen-
tial and radial approach to a ball boundary, it will be shown in the next
sections that a positive h-harmonic function has a nontangential limit fi at
any ball boundary point at which there is a minimal-fine limit Ii and that a
superharmonic potential has radial limit 0 at 1,4_1 almost every boundary
point at which there is minimal-fine limit 0. Thus Theorem 19 for a ball
implies the truth of the corresponding partial results described above for
nontangential and radial boundary approach. The converse is false. More
exactly, it will be convenient to prove the stated implications for a half-space
rather than a ball. Since half-spaces and balls can be mapped onto each
other by inversions, the assertions on nontangential convergence to bound-
aries correspond exactly in the two contexts. The assertion on radial conver-
gence to a ball boundary corresponds very nearly to the corresponding
assertion on convergence to a half-space boundary along normal lines.

21. Nontangential and Minimal-Fine Limits at a Half-space


Boundary
Let dd be the last coordinate of the point of R', and let D be the half-space
{d, > 0).

Theorem. If u = v/h is a strictly positive h-harmonic function on the half-space


D and if u has minimal-fine limit fi at the boundary point i;, then u has non-
tangential limit ft at C.

The converse theorem can be shown to be false. Theorem 21 implies


Theorem I1.15, that u has nontangential limit dM,,/dM, at Mh almost every
22. Normal Boundary Limits for a Half-space 223

boundary point, in view of the minimal-fine topology Fatou theorem


(Theorem 19) and the correspondence between ball and half-space given by
an inversion.
We can assume in the following proof that the specified boundary point
{ of D: {dd > 0) is the origin. Let . be a sequence in D with nontangential
limit the origin, and suppose that fi'. To prove Theorem 21,
it will be shown that fi' = P. Assume from now on that 0 < fi < + 00. The
modifications to be made in the argument if fi is 0 or + oo will be obvious.
It can be assumed that the ray from the origin through tends to a limit
ray L as n ao. Let ;, be the point of L at minimum distance from c,,, let
A be a ball, closure in D, center o, let Ao be a smaller concentric ball, and
define c = The ball c,A has center ;,, and ec.A;, for sufficiently
large n. Finally (by the Harnack inequality), there is a strictly positive
constant y depending only on the ratio of the radii of A and Ao such that
1 /y 5 u(q)/u() 5 y for both and >l in Since Uo (c Ao) is not minimal-
thin at the origin [Section 12, Example (b)], there is a sequence 17. along
which u has limit fi, with q E Hence 1/7 5 fl/li' S y. Since the Harnack
constant y can be made arbitrarily near I by choosing Ao sufficiently small,
it follows that fi = fi', as was to be proved.

22. Normal Boundary Limits for a Half-space


Define the half-space D as in Section 21. A boundary point C will be called
a normal limit point of a subset A of D if C is a limit point of the part of A
on the normal to 8D through C. A function on D will be said to have a
normal limit at t if it has a limit at C along the normal.

Lemma. If A is an (open) subset of the half-space D, IN-, almost every normal


boundary limit point of A is a minimal-fine limit point of A.

Warning: "Open" was enclosed in parentheses in this statement because


(Section 23) the lemma is true for an arbitrary subset A of D.
It is sufficient to prove the lemma for a bounded open set A. A careful
application of the Vitali covering theorem shows that there is a subset Ak of
A with the following properties: Ak is a countable union of closed (N - I)-
dimensional intervals, each on a hyperplane dd = const < 1/k), each
with edges parallel to the coordinate axes; these intervals have disjoint
projections on 8D; the projection on 6D of Ak covers IN_, almost every point
of the projection of A n do < l/k}. Define the potential vk by

,)lN-t(dn)
(22.1)
vk(0 = fAk G dd

The remainder of the proof assumes that N > 2. The argument is similar
when N = 2. We first prove that there is a constant c, depending only on N
224 I.XII. The Martin Boundary

such that uk 5 c, for all k and choices of Ak. To see this, let be a point of
D, and define

D= hi(s) = sup n)
s),
qe DrOD, d,,

The evaluation (see Section VIII.9)


s-(N-2) - (s2 +
d. d
shows that the left side decreases when d,, increases, from which it follows
that

2(N - 2)ds-" if s > d4,


cur(s) = s-iN-2) - (2ds - s)-(x-2) (22.2)
if s 5 d{,
d4 -s

so that is monotone decreasing. Thus

rk() < O(s)do(s) with 4.(s) = IN_, (A n D,). (22.3)


Jo

Integration by parts, together with the fact that 4(s) 5 nN-,s"-`/(N - 1),
yields the majorant n,_, fo 4,(s)sN-2 ds of the right side of (22.3), and in
view of (22.2) the integral f u cur(s)sN-2 ds is convergent with value c, indepen-
dent of 4, A, k, Ak. If v; is defined using only finitely many of the intervals
in Ak, v,/c, is the potential of a measure supported by A,, and vv/c, 5 1.
Hence (domination principle) vk 5 c, o- 1o-", and therefore vk 5 c, o- 1 Q". To
bound vk from below, observe that when k oo, the integrand in (22.1)
tends to the normal derivative of at the boundary, that is [VIII(9.4)],
to it pD(b, dn)/IN_,(dn), uniformly for in any bounded set, so that (22.1)
yields, if A" is the set of normal limit points of A on 3D,

lim inf vk Z A").


k-m

Thus A") 5 c2 o- 1o-" for some constant c2. Denote by A' the set of
minimal-fine limit points of A on D. According to Section 17, GMDo- 1o-" =
A'), so that A'):5 A'). Since the harmonic measure of a
Borel measurable boundary set B has minimal-fine boundary limit function
1,, up to an !N_, null set, it follows that A" c A' up to an IN_, null set, as
was to be proved.
23. Boundary Limit Function of a Potential on a Half-space 225

23. Boundary Limit Function (Minimal-Fine and Normal) of


a Potential on a Half-space
Theorem. Let D be a half-space of R'.
(a) A superharmonic potential on D has normal limit 0 at 1N_, almost
every point of BD.
(b) if A is a subset of D, 'Ni _almost every normal boundary limit point
of A is a minimal-fine boundary limit point of A.

Proof of (a). To prove (a), it is sufficient to show that for u a superharmonic


potential on D and e > 0 the open set Ae = {u > e} has IN_, almost no normal
boundary limit point, and in view of Lemma 22 this follows from the fact
(Theorem 18) that u has minimal-fine limit 0 at µD (equivalently, IN-t)
almost every boundary point, so that IN_, almost no boundary point is a
minimal-fine limit point of A,

Observation. We have now proved that a positive superharmonic function


on a half-space of 1' has a limit at IN_, almost every boundary point on
both minimal-fine and normal approach to the boundary point and that the
boundary limit functions in the two approaches are equal up to IN_, null
sets. In fact (Riesz decomposition) it was sufficient to prove the theorem
separately for u harmonic and u a potential. The harmonic case was covered
in Section 19, in which case nontangential boundary approach was admis-
sible; the potential case, in which the limit vanishes 1N_, almost everywhere
on OD, is covered in (a) of the present theorem.

Proof of (b). Let An [A'] be the set of normal [minimal-fine] boundary


limit points of A. The function 018" is a positive superharmonic function
on D, equal to I quasi everywhere on A. Let Ao be the polar subset of A
on which 818" < 1, and let vo be a superharmonic potential on D, identically
+ oo on Ao. According to (a), the function vo has normal limit 0 at IN_,
almost every point of BD; so 1N_, almost no point of aD is a normal limit
point of Ao, and we shall therefore assume from now on that Ao is empty.
In view of the Riesz decomposition theorem and (17.2),

_
818A = v + PD(-, A'), (23.1)

where v is a potential on D. The smoothed reduction on the left side of (23. 1)


has normal boundary cluster value I at every point of A" and therefore by
(a) and the above observation has normal and minimal-fine limit 1 at 'Ni
every point of A". The function on the right side of (23. 1) has minimal-
fine boundary limit function 1A. up to an IN_, null set; so A" a A' up to an
1N_, null set, as was to be proved.
Chapter XIII

Classical Energy and Capacity

1. Physical Context
Consider a distribution of positive and negative electric charges on IV and the
electrostatic potential induced by this charge. By definition of a conductor,
if A is a connected conducting body in R', the charge on A distributes itself
in such a way that the net effect is that of an all-positive or all-negative
charge. and the distribution on A is in equilibrium in the sense that the
restriction to A of the potential of the charge distribution in I8' is a constant
function.
Let D be an open subset of R with a conducting smooth boundary, and
suppose that the boundary is grounded. The significance of grounding is that
if a positive charge p is imposed on D, an induced negative charge - p*
appears on W. and the potential G(p - p*) is identically 0 on OD. Thus, if
p is a unit positive charge at in D, the restriction to D of the potential
G(p - p*) is identified with the Green function and p* is identified
with the sweeping of p onto aD (relative to R'); that is, p* is identified with
the harmonic measure p the measure p* is
identified with the sweeping of p onto aD and that G(p - p*) = GDp on D.
In view of this physical context the existence of a mathematical version of
the Green function of a reasonable set D was obvious long before there was
a rigorous existence proof, and the sweeping of a measure was a natural
concept to formalize.
Now suppose that a connected conducting body is introduced into D and
given a positive charge p. This charge necessarily distributes itself in such a
way that G(p - p*) = GDp is constant on A. Such a charge p, that is, such
a measure, is called an equilibrium charge (or measure, or distribution), and
the corresponding potential GDp is called an equilibrium potential for A. Two
equilibrium potentials for A are proportional, as are their potentials, and
if the potential on A of an equilibrium measure has the constant value I the
equilibrium measure [potential] is called the capacitary measure [potential].
In view of this physical context it was clear to Gauss that there must be a
capacitary distribution in a suitable mathematical context for any reasonable
pair A and D. The pair (A. OD) is a condenser in the physical context, and
the capacity of this condenser is defined as
2. Measures and Their Energies 227

Total charge of an equilibrium distribution on A


Value on A of the corresponding potential
= total charge of the capacitary distribution on A.

The mathematical model of this physical context has already been dis-
cussed, at least in part. The set D is supposed Greenian. The Green function
GD has already been defined, and has been shown to have limit 0
at every regular boundary point of D. If p is a measure on D, the function
GDµ is the mathematical version of the electrostatic potential generated by
a distribution p of electric charge. More generally p will sometimes be allowed
to be a suitably restricted signed measure or charge in the sense of Appendix
IV.7. The energy of a charge and the mutual energy of a pair of charges will
be defined, following mathematical tradition, as twice the values assigned
by physicists. Equilibrium distributions will be derived, and the capacity
of a subset A of D will be defined by (1.1) when A is analytic.

2. Measures and Their Energies


If D is a Greenian subset of RN, the mutual energy [p, v] of a pair of measures
on D is defined by

[µ, v] =f GDµdv fG(. 1)µ(d)v(drl), (2.1)


= J
and the energy lip 11' of a measure p on D is defined as [µ, µ]. The form
is symmetric,

[µ, v] = [v, µ], (2.2)


and this symmetry is sometimes dignified by the name reciprocity law. The
symmetry of the Green function GD is a special case.

The Space 8'

The space of measures on D of finite energy will be denoted by 8'. It is


trivial that a positive constant multiple of an element of 8' is in 8', but it
is a much deeper fact that (Theorem 7) the sum of two elements in 8' is in
8', equivalently, that the mutual energy of two measures in 8' is finite.

EXAMPLE (a). If p is a measure on D and if A is a polar subset of D that is


not p null, then lip 11 = +oc because (Theorem V.11) GDµ = +oo p almost
everywhere on A.
Thus a measure of finite energy vanishes on polar sets. Conversely, if A is
an analytic subset of a Greenian set D and is null for every measure on D
228 1. XIII. Classical Energy and Capacity

of finite energy supported by A, then A is polar in view of Corollary V.9


and the fact (Theorem VI.2) that an analytic nonpolar set has a compact
nonpolar subset. It will be shown in Section 3.1.9 that a superharmonic
potential GDµ is quasi bounded if and only if u vanishes on polar sets. It
was noted in Section V.10 that the domination principle as applied to
potentials of measures p vanishing on polar sets has a simple form: if GDµ
is majorized u almost everywhere on D by a positive superharmonic function
v, then GDu < v on D.

EXAMPLE (b). Let be a point of D, choose a > 0, and define B =


a), A = D n aB (Euclidean boundary), and u = SD(I, ). Since has
limit 0 at quasi every point of aD, aB - A is polar and thus is a null
set; that is, 1. (This fact is a special case of the application in
Section VIII.18.) The measure y is supported by A, and a on A;
so

lip 02 = aba(:, A) = A) = a. (2.4)

In particular, if D = B= fl) with fi < b, then

if N = 2,

{G(S,') > R2-N - 62-N} if N > 2;


so the energy of the uniform distribution on aB of total value I is log (S/f)
if N = 2 and is R2-1 - 62-N if N > 2. The latter energy decreases to Y{r2-N,

the energy of u relative to RN, when b - oo.

EXAMPLE (c). The general case of the preceding example is the following.
Let D be a Greenian subset of RN, let B be an open subset of D, let be a
point of B, and define A = D n aB (Euclidean boundary) and u = Sp(, ).
Then
IIull2 = 6A (S, GD(c,')) = GD(S. NB(S, 1D) (2.6)

3. Charges and Their Energies


A charge (Radon measure) ,u on the open subset D of R' is (see Appendix
IV.3) a pair (ut,u2) of measures on D under the identification (u1,u2) _
(vt, v2) when pt + v2 = u2 + v1 and the operations

(PI, 142) + (v1, v2) = (101 + v1, uz + v2),


(cut, cue) ifc-0, (3.1)
c(ut, uz) _
_
(-cue, -cut) ifc < 0.
4. Inequalities between Potentials, and the Corresponding Energy Inequalities 229

A charge has a minimal representation whose component measures have


disjoint supports. A charge is called positive if its minimal representing pair
has the form (p,, 0) and is then identified with the measure p, .
Let N be the space of pairs (f,,f2) of positive extended real-valued
functions under the equivalence relation (f, ,f2) = (g,, g2) when f, + g2 =
f2 + g, and the operations (3.1) interpreted in the present context. The
potential GDµ of a charge p is defined as the point (GDµ,, GDP2) of N and
is informally identified with the function GDp, - GDp2 where this difference
is defined. The only charges of interest are those charges p the components
of whose minimal representing pair (p, , p2) have superharmonic potentials,
in which case GDp, - GDp2 is defined quasi everywhere on D. The potential
GDp of a positive charge (p, , 0) is identified with the function GDp,.
The space of charges that have representations with both components in
d'' will be denoted by d, and for elements of if only representations with
components in 8' will be used. If p: (p,, 142) and v: (v,, v2) are charges in
8, their mutual energy is defined following the definition in the positive
case,

[p, v] = [p., vt] + [p2, v2] - [PI, v2] - [p2, vt],

and the energy of u is defined as [p, p]. To justify this definition, it must be
proved that the mutual energy of a pair of elements in 8' is finite (see
Theorem 7), and until this fact is proved, the definition (3.2) will be used
only when the finiteness of the summands is obvious from the context. The
form is symmetric (reciprocity law). It will be proved that the energy
of a charge in 8 is positive, and the energy will then be written This
notation will be used before positivity is proved whenever positivity is
obvious from the context.

4. Inequalities between Potentials, and the Corresponding


Energy Inequalities

Lemma. (a) The inequalities GDµ' 5 GDV and GDP" S GDv" for potentials
of measures imply that [y', p"] 5 [v', v"].
(b) The inequality GDµ 5 GDv for potentials of measures implies that
((p((2 5 ((v((2 and that y = v if these energies are finite and equal.
(c) Let GDp, and GDv, be increasing [decreasing] sequences of potentials
of measures, with respective [smoothed] limits GDp and GDv. Then

lim,,..m [Ati., v.] = [p, v]

in the increasing case and, if [pa, vo] is finite, also in the decreasing case.
230 I .X111. Classical Energy and Capacity

Proof of (a). Apply the symmetry of GD to obtain

[i'f'] = GDµ' dµ" < f GDv' dp" = JGoKdv'


fD D D
(4.1)

5 f GDv" dv' _ [v', v "].


D

Proof of (b). If p' = p" = p and v' = v" = v in (a), then l1µ11 s 11v11, and with
this specialization if the energies, that is, the integrals in (4.1), are finite,
equality in (4.1) implies that GDµ = GDv both p almost everywhere and v
almost everywhere, so that GDµ = GDv by the domination principle for
potentials of measures of finite energy (Section 2), and therefore µ = v.

Proof of (c). In the increasing case in view of (a) the sequence [µ,, v.] is an
increasing sequence with limit at most [µ, v]. The limit is [p, v] because for
every m,

lim [µ Z lim inf f GDµ dv = lim inf GDv dµ,,,


n+m n-aD
o n-.0J, D
= f GD, dµm = J GDPm dv,
D D

and the last integral can be made arbitrarily near [p, v] by choosing m large.
The decreasing case is treated similarly, with the help of the fact that polar
sets are A null if A is a measure of finite energy.

5. The Function D G0µ

(a) Let D. be an increasing sequence of open subsets of W4 with union a


Greenian set D. Since GDS 5 GDM+t on D. x D., it follows that if µ and v are
measures on D and if µ is the projection of u on D, then

lim f G dv;
e-m JD D

that is, mutual energy relative to D. increases to mutual energy relative to


D. Observe that we have not supposed that the integrals in (5.1) are finite
valued but that once we have proved (Section 7) that when the measures µ
and v are in tf + for D, their mutual energy for D is finite valued, it will follow
trivially that (5.1) is valid for p and vin 4f for D. although the sequence GD,p,
need then not be monotone.
(b) Let D2 be a Greenian subset of 08', let Dt be a relatively compact
6. Classical Evaluation of Energy; Hilbert Space Methods 231

open subset of D2, and let p be a measure on D2 with compact support in


D, . Energies, reductions, etc., relative to D, will be distinguished by the
subscript i: 8.0;, ffi, .... According to Section VII.5,

GD,µ = GD p + GD,µt (5.2)

on D, , where p, = 0µ0°=_°j is supported by OD, . Equation (5.2) is obviously


also valid for a charge p supported by a compact subset of D, , with µ, a
charge supported by OD,. Since GD,µ, is bounded on the compact support
of p,, equation (5.2) implies that p e 62 if and only if µ E 4, . (We are abusing
language slightly here in considering p both as a measure on D, and as a
measure on D2.) If we assume that all the following mutual energies are well
defined, we can combine (5.2) with X(5.3') as applied to the measure compo-
nents of the charges to find that if p and v are in S2 with compact supports
in D,, then
[p,v] I = [p - µt+V - VIJ2, (5.3)

where v, is defined in terms of v in the way µ, was defined in terms of µ.


Thus, if all the mutual energies involved are well defined (and according to
Theorem 7, this is true if µ and v are in 82), the map µ'--. µ - µ, is a linear
mutual-energy-preserving map from a subset of 8, into d'2.

6. Classical Evaluation of Energy; Hilbert Space Methods


If D is an open subset of R" and if u and v are functions from D to the extended
reals, define

2(u, v) _ f (grad u, grad v>&", em(u) = 2(u, u) (6.1)


D

whenever the integrals are meaningful. The value 2(u) is the Dirichlet
integral of u. If u = GDµ and v = GDv are the potentials of charges in if, on
the Greenian set D, and are in class Cj2'(D), then nH dv = -Av d1N (Section
1.7), so that

[µ,v] _ -n uAvd1N. (6.2)


D

If D is bounded and smooth enough for the application of Green's


identity and associated theorems and if the potentials u and v when defined
as 0 on 3D are in class Ct21(D), (6.2) yields

u) 2(u)
[ p, v] = 2(n' [ µ 1<] = (6.3)
232 1. XI If. Classical Energy and Capacity

If D = RN with N z 3 and if u = Gp and v = Gv are potentials in class


C'2 (RN) of charges with compact support, then if B is a ball containing the
compact support of v,

f Gp dv = , lf u &v dlN
RN RN R^

_ f <grad u, grad v> dIN - 1 f


Na RN,1a

where Dov is the directional derivative in the direction of the exterior


normal. For large distances r from the origin, Igradul and Igradvi are
majorized by const r' -N, and Jul is majorized by const r2°N. Hence the
integral defining Q(u, v) converges absolutely, and when B increases to RN,
the last integral in (6.4) drops out; so (6.4) becomes (6.3). If D = R2, the
preceding analysis is valid through (6.4) but must be modified thereafter. If
is a charge on R2 with compact support A,

ICI
tm
141-- t
141 log -ii 0

and

grad GA() + 2I l212 ((A) 15 const Itl-2

for large I I. It follows that if y and v satisfy the additional condition µ(R2) _
v(R2) = 0, then the integral defining -Q(u, v) converges absolutely, and (6.4)
becomes (6.3) when B increases to R2.
Returning to the hypotheses of the first paragraph, if u = Goµ as described
there but if v is harmonic and has an extension to D in class Ct21(D), then
Green's identity [Section 1(1.2)] becomes

f vAud1N = faD vD oudlN_,, (6.5)


n

and integration by parts on r3D yields 2(u, v) = 0.


This discussion suggests that Hilbert space methods are appropriate to
the study of charges on D using the inner product [p, v] and to functions
on D using the inner product 9(u, v). The following discussion sketches a
basis for such a study. We suppose that D is connected.
(a) Charges. It will be shown in Section 7 that 9 is a linear space, that the
energy of a charge in 9 is positive, and that this energy vanishes if and only
if the charge is the zero charge (0, 0). Thus 9 is a space which, coupled with
7. The Energy Functional (Relative to an Arbitrary Greenian Subset D of RJ') 233

the bilinear form is a pre-Hilbert space, that is, a space in which all
the Hilbert space axioms except that of completeness are satisfied. It is known
that the space 6 of charges is not complete but that its subset 6 + of measures
of finite energy is complete.
(b) Functions. Let u, be a sequence of infinitely differentiable functions on
D for which 2(u - ur,) = 0. It is known that
there is then a function u from D into A with the following properties: grad u
exists IN almost everywhere on D, with 2(u) < + oo ; lim . 2(u -
u is the quasi everywhere limit of a subsequence of u.; u is fine continuous
quasi everywhere on D. A function u obtained in this way is called a "BLD"
function (Beppo-Levi-Deny). The space of BLD functions is the natural space
of functions on which to base the study of orthogonality and related topics
involving the Dirichlet integral. The harmonic BLD functions are the har-
monic functions with finite Dirichlet integrals. If the functions in the approx-
imating sequence u, have compact supports, u is said to be of potential type.
It has been proved that the class of BLD functions of potential type includes
the potentials of charges in 6, in particular, the superharmonic potentials of
measures of finite energy. If two BLD functions are identified when the
restriction of their difference to the complement of a polar set is a constant
function, the space of BLD functions coupled with the inner product 2(-, -)
is a Hilbert space in which the classes of harmonic BLD functions and of
functions of potential type are orthogonal subspaces with direct sum the
whole space. This fact generalizes our application of (6.5) according to which
if u is a superharmonic potential on D and v is harmonic on D, if D is suffi-
ciently smooth, and if u and v have sufficiently smooth extensions to D, then
2(u, v) = 0.
In view of (6.3) as suitably generalized to the BLD context, the pre-
Hilbert space it of charges can be immersed in the Hilbert space of BLD
functions by identifying a charge p in 6 with the function (nN)-"ZGop. The
Riesz decomposition of a positive superharmonic function into the sum of a
positive harmonic function and the potential of a measure can be inter-
preted, if the given function is a BLD function, as the sum of its orthogonal
projections on the subspaces of BLD harmonic functions and of functions
of potential type.
The details of this Hilbert space approach will not be carried out in this
book, and the stated results will not be needed with one exception. We shall
need the fact that the energy of a charge in 6 is positive. This fact will be
proved in the next section.

7. The Energy Functional (Relative to an Arbitrary Greenian


Subset D of RN)
Theorem. (a) The energy of a charge in 6 is positive and finite.
(b) If p and v are in 6, their mutual energy exists and
234 1. X111. Classical Energy and Capacity

I[µ,v]ISllµ11llv1l. (7.1)

(c) A charge in of is the zero charge if and only if its energy vanishes.
(d) There is equality in (7.1) if and only if µ and v are proportional.

Observation. Theorem 7 implies that 8 is a linear space, a pre-Hilbert space


when coupled with the inner product J. (See the discussion in Section 6.)
Assertion (b) for µ and v in 8+ will be denoted by (b+), and we write
(a) (sm), (b) (sm), and so on to refer to the smooth context in which the
charges involved have component measures which have compact supports
in D and have infinitely differentiable potentials. If µ and v are in 8+(sm)
their mutual energy obviously exists and is finite. In the following proof of
Theorem 7 all the assertions are relative to a specified Greenian set D unless
otherwise qualified.

Proof. Step 1. Proof that (b+) (a) (b) and that (b+)(sm) (a)(sm)
(b)(sm). The following argument without the smoothness condition is also
applicable in the smooth context. Let µ : (µ1,µ2) and v be in 8. Under (b+)

[u,µ] = 11µt112 - 2[µ,,s02] + 11102il2 z (11µ[1I -1111211)2; (7.2)

so (b) (a). Under (a) the equality in (7.2) implies that [µt 11021 < + oo,
which implies in turn that 8 is a linear space and that for ce R,

0 511µ+cvll2=l1µ112+2c[µ,v]+c211v112 (7.3)

If 1lviI = 0, inequality (7.3) is impossible unless [µ,v] = 0; if 1lvll > 0, in-


equality (7.3) with c = - [µ, v] 11°11-2 yields (7.1). Hence (a) (b).
Step 2. Proof that (b+) (sm) = (a) n (b). According to Section IV.10 a
superharmonic potential of a measure on D is the limit of an increasing
sequence of infinitely differentiable potentials of measures with compact
supports in D. Thus, if p and v are measures in 8+, there are sequences p,
and v, of measures in 8+(sm) for which GDµ, and Go v, are increasing se-
quences with respective limits GDµ and GDV. Moreover, if (b+) (sm) is true,

[p,,, vJ s IlA-11 llvvll,

and inequality (7.1) now follows from Lemma 4. Hence (b+) is true and
therefore (a) and (b) are true by Step 1.
Step 3. Proof that (a) and (b) are true when D is a ball. According to
Step 2, it is sufficient to show that (b+)(sm) is true when D is a ball. When
µ is in 8(sm) relative to a ball D, the expression II(1.1) for GD shows that
GDp, when defined as 0 on 6D, is infinitely differentiable on D. The context
is therefore that of Section 6, and the evaluation (6.3) of [µ, v] shows that
the inequality (b+)(sm) follows from Schwarz's inequality.
8. Alternative Proofs of Theorem 7(b') 235

Step 4. Proof of (a) and (b) when D is bounded. According to Steps I


and 2, it is sufficient to prove that (a)(sm) is true when D is bounded. Let
D2 be a ball containing D. If µ is a charge in B(sm) relative to D, then µ
is also in 0(sm) relative to D2, and according to Section 5(b), the energy
[p, p] relative to D is the energy relative to D2 of a certain charge supported
by 8D, and (by Step 3) this energy is positive.
Step S. Proof of (a) and (b). According to Step 1, it is sufficient to prove
(b+). Let D. be an increasing sequence of nonempty open bounded subsets
of D with union D. If µ and v are in 6+ relative to D and if µ and v are the
projections of these measures on D., then (by Step 4) [p5, v,J S
(energies relative to D,.), and (from Section 5) when n oo, this inequality
yields (7.1).
Step 6. Proof of (c) and (d). Suppose that p : (µ 1, ,42) E 49 and that 11011 = 0.
It follows from (7.1) that [µ, v] = 0 whenever v is in 9, in particular, when
v is a uniform distribution on the boundary of a ball with closure in D.
With this choice of v the vanishing of [p, v] implies the equality of spherical
averages of GD p l and GDN2 ; that is,

L(GDµt, , 6) = L(GDP2, , b)

when S) a D. When b -' 0, it follows that GDpI = GDµ2 ; so PI = 112,


and p must be the zero charge. If the charges in (7.1) are proportional, there
is obviously equality in (7.1). Conversely, if there is equality in (7.1), the
charges are trivially proportional if either has zero energy, that is, if either
is the zero charge; otherwise, equality in (7.1) implies that
11.2v112
110 - 10, V]11V = 0.

Sop= [p, v] 11v11-2v, and the proof is complete.

8. Alternative Proofs of Theorem 7(b*)

The following proofs of the key inequality (7.1 +) are unnecessary but
instructive.
(a) Heat equation potential theoretic-probabilistic proof of (7.1 +). Sup-
pose that there is a positive Borel measurable function (t, , rl) r- bD(t, 4,17)
from ]0, + oo [ x D x D into R+ with the property that bD(t, , ) is symme-
tric, that

bD(s + t, , h) = J f bD(s, , ObD(t, {, 1)IN(dC), (8.1)


D

and that bD is related to the Green function GD by


236 1. XIII. Classical Energy and Capacity

n) = c I b0(t, , n)1 (dt) (8.2)

for some positive constant c. If p is a measure on D, denote f DbD(t, , n) p(dn)


by bD(t, , p). Then

[p, v] = c J
o"
11 (dt)
L
(2, , u) bv(2, ,) , (8.3)

and Schwarz's inequality yields (7.1 +).


Every Greenian subset D of RN has a function bD with the stated proper-
ties. In fact we shall see that if IS = D x R and if (; is the heat equation
Green function of I), then bD defined by

bD(s - t, , n) = s), (n, t))

has the desired properties. This function bD will also be identified with the
transition density relative to IN of Brownian motion in D (transition from
to n in time t). See Chapter XVII for a discussion of OD, Section 2.VII.9
for the Brownian motion transition density bD, and Section 2.IX. 17 for the
identification of the potential theory bD with the probability bD.
This method of proving (7.1 +) can be applied without recourse to heat
equation potential theory or probability as follows. Define a function b on
]0, +oo[ x R by

b(t,) _ (2at)-1112 eap -IIZ.


2t
Then if N z 3, the function bRN defined by baN(t, , n) = b(t, - n) has the
desired properties for D = RN. Hence (7.1 +) is true for D = RN, and with
the help of Section 5 it follows that (7.1 +) is true for an arbitrary open
subset of RN. If N z I and D is a half-space, denote by of the reflection in
8D of a point n in D. Then the function bD defined by bD(t, , n) = b(t, - n)
- b(t, - n') has the desired properties, and with the help of Section 5 it
follows that (7.1 +) is true first for an arbitrary nonempty open subset of a
half-space and then for an arbitrary Greenian subset of RN.
(b) Proof of (7.1 +) when N z 3 by a splitting method. This method
uses the fact that there is a constant CN for which

I _ 1N(A)
(s.a
)
I- n1"-2
- CN RN I - CIN-' In - CIN-` *

To verify (8.5), observe that the integral defines a function of I


- n I. Denote
this function by 0. change the integration variable to C' = CIA - nL', and
thereby find that O(I - nl) = I - 171-1+2 0(l). Hence (8.4) is true with
vN = 1 /¢(1). Apply (8.4) to derive
10. The Classical Capacity Function 237

)[J IS - hI-N+1 ju(t)] CI-N+l v(dZ)


J, ( 8.5)
Ifn I -
[µ, V] = cN f
RN D

and apply Schwarz's inequality to obtain (7.1 +) for D = RN. Proceed as in (a).

9. Sharpening of Lemma 4

The following theorem is needed in Section 10.

Theorem. If GDµ, is an increasing sequence of potentials of measures and if


sup,I I N I I < +oo, then lim,, GDµ is the potential ofa measure p in 8+, and

llm11µJ=IIµII, llmIIµ-N,.II=0.

(It is easy to see that the second limit result implies the first.) If u is the
limit of the sequence of potentials and if v is a measure on D,

udv = li m [i1., v] s Sup IIµ.II IIVII. (9.2)


SD

In particular, let be a point of D, let B be an open relatively compact


subset of D containing , and choose v = pB( , ). Then [from Section 2,
Example (c)] (9.2) becomes

PB(S, u) < (9.3)


su
x2t

and the right side has limit 0 when B increases to D. Hence (Section VIII. 11)
u is a potential, say u = GDµ, and (9.1) follows from Lemma 4.

10. The Classical Capacity Function

Let D be a Greenian subset of RN, let f, be the class of those subsets A of


D for which Q 10A (reduction relative to D) is a potential, and for A in f,
let 2A be the measure associated with I I J A ; that is, 10 A = GD2A. The measure
A is called the capacitary measure of A, and GDAA is called the capacitary
potential of A. If c > 0, the measure ea.,, [potential cGDAA] is called an
equilibrium measure [equilibrium potential] of A. These concepts are relative
to D. According to Section XII. 17, A e f,, if and only if when D is provided
with its Martin boundary, the set AJn a"D is µD null. It is unnecessary to
use this characterization of fp to prove the elementary facts that a subset
238 1. Xiii. Classical Energy and Capacity

of a set in rp is in r,, that every polar subset B of D is in F, with A, = 0,


and that relatively compact subsets of D are in f,,. Since GDAA 5 1, polar
sets are AA null (Theorem V.11). If A and B are in f,, and if A differs from
B by a polar set, then 2A = AB. Finally, we show that if A is in rp, some
open superset of A is also in f,,. In fact, if B, = {O l0A > 1}, an open set
which covers quasi every point of A, let u be a superharmonic potential on
D, identically + oo on the polar set A - B, , and define B = B, u {u > 11.
Then B is an open superset of A and is in rp because 116' is majorized by
the potential 2 111 0 A+ U.
For A in f, the measure .A is supported by D n aA (Euclidean boundary)
and even (Corollary XI.18) by D n afA1. Recall (Section XI.6) that for an
arbitrary subset A of R' the sets Af and afAf are Borel sets. Since D 1 Q A = 1
on D n afAf we conclude that

IIAAII2= O1QAd1A=2A(D) (Aerr). (10.1)


Ond fAf

If A is a subset of D not in fr,, we do not define AA, but it is convenient


to define I12A II2 = + oc to obtain a set function II2.112 defined on every subset
of D.

Theorem. (a) The set function II2.112 is a countably strongly subadditive


Choquet capacity on D relative to the paving of compact sets, and for A in r'p

inf {II28112 : A c B, B open} = inf {2B(D): A c B, B open}. (10.2)

(b) If A is II2.112 capacitable, then

II2AII 2 = sup {2(F): F c A, Fcompact, A supported by F, GDA < I on D}.


(10.3)

(c) A subset A of D is polar if and only if II AA II2 = 0.

Observation. Assertions (a) and (c) together are equivalent to the asser-
tion that the restriction of II2.112 to the class of compact subsets of D is a
topological precapacity and that II2.112 is the Choquet capacity generated
(Appendix 11.8) by this topological precapacity.

Proof of (a). The set function II2.112 is obviously an increasing set func-
tion, and if A is an increasing sequence of sets with union A (c D), then
lim,,..," II2AII2 because this limit equation is trivial if the limit is
+ oo and the limit equation follows from Theorem 9 otherwise. Next let B
be an open relatively compact subset of D, and let A be a compact subset of
B. Then AA is supported by A, GDAA = 1 quasi everywhere on A, GD2B = 1
10. The Classical Capacity Function 239

on Af and [Section VI.3(b)] GDAA = R' on D - A and so surely on the


compact support of 2B. Since polar sets are 2A null, it follows that

112A 112 =
GDAA d2A = GRAB d2A = GDAA d i8 = f Ri dAB. (10.4)
J Ji) fD JD

Since (Theorem VI.5) the set function R, (c) is a topological precapacity on


the class of compact subsets of D, it follows that if A. is a monotone sequence
of compact subsets of B with limit the compact subset A of B, then
limw_. IIAA,I12 = 1124112. The proof that 112,112 is a Choquet capacity relative
to the paving of compact sets is now complete.
To prove (10.2), observe first that the second equality follows from (10.1)
and that the first equality is trivial when II2A112 = +00. If 1124112 < +oo, the
set A must be in fo, and we have already noted that then Io contains open
supersets of A. Furthermore, according to Section VI.3(m),

0104=[inf{p10B:A cB,Bopen}]+, (10.5)

and therefore according to the Fundamental Convergence Theorem, there


is a decreasing sequence B. of open supersets of A such that

DIIA=ClimDI p"] (10.6)


w ID
+

The first equality in (10.2) now follows from Lemma 4(c). What we have
proved implies that the restriction of the capacity 112,112 to the class of
compact subsets of D is a topological precapacity generating 112,112 ; so this
capacity is countably strongly subadditive.

Proof of (b). If F is a compact subset of D, if A is a measure supported by F,


and if GDA 5 1 on D, then we can apply (10.1) and Lemma 4 to obtain

A(F) 5 SF GDAFd2 S IIAF112 = 2F(F). (10.7)

Moreover, if A is 1P2.112 capacitable, then by definition of capacitability

1124112 = sup{112F112: Fc A,Fcompact}


(10.8)
= sup {2F(F): F c A, Fcompact}.

The combination of (10.7) and (10.8) yields (b).

Proof of (c). It is sufficient to prove (c) for relatively compact subsets A


of an arbitrary relatively compact open subset B of D, and we apply (10.4).
240 1. XIII. Classical Energy and Capacity

If A is polar, then RtA = 0 quasi everywhere on D (Theorem V.4); so (10.4)


implies that IIAA112 = 0. Conversely, if IIA4112 = 0, then equality of the first
and third terms in (10.4) implies that AA(D) = 0; so !2' vanishes identically,
and therefore (by Theorem V.4) A is polar.

Unique Characterization of AA

If A E r,, then a measure A on D is AA if and only if A is supported by Af


and GDi : 1 with equality quasi everywhere on A. In fact, in view of Theorem
XI.14 the measure AA satisfies these conditions. Conversely, if A satisfies
these conditions and if A' is the projection of A on a compact subset of Af
then (domination principle) GDA' 5 GDAA; so GDA S GDAA. Interchanging A
and AA yields the reverse inequality; so A = AA. Observe that this argument
also shows that if A is closed relative to D in the Euclidean topology and
if A is in r,, then A = AA if and only if A is supported by A and GDA S I
with equality quasi everywhere on A.

11. Inner and Outer Capacities (Notation of Section 10)


Let A be a subset of D, and define

C*(A) = sup{11AFII2: Fc A,Fcompact}, C*(A) = 11A4112. (11.1)

Recall that with this definition C *(A) < + oo implies that A E r,; so Rl is
a potential, GDAA, that AA is supported by afAf and that

C*(A) _ JJAA11' = AA(DnafAf).

On the other hand, there are sets A in TD with AA(D) = C*(A) = +oo.
The values C*(A) and C*(A) are called, respectively, the inner and outer
capacities of A (relative to D). Thus C*(A) is the infimum of the inner
capacities of the open supersets of A. The set A is 11 A. 112 capacitable, that is,
C* capacitable, if and only if C*(A) = C*(A), and then C(A), called the
capacity of A (relative to D), is defined as the common value of the inner
and outer capacities of A.
If A is a subset of D with finite outer capacity and if e > 0, there are an
open superset AE of A and a compact subset A, of A such that

C* (A) < C(AE) + e, C(AE) < C*(A) + e,

and there exist a G6 superset A" of A and an F. subset A' of A such that

C(A') = C*(A) 5 C*(A) = C(A").


12. Extremal Property Characterizations of Equilibrium Potentials 241

12. Extremal Property Characterizations of Equilibrium


Potentials (Notation of Section 10)
The canonical equilibrium measure A,*. We have used the capacitary measure
AA as a canonical equilibrium measure. A second choice, which is sometimes
convenient when A is known to be capacitable with 0 < C(A) < + oo, is .i;,
defined by A,* = 2"/C(A). Then

4 I with equality quasi


Z,*i (D) = 1, I I AA II Z = C(A) , GDAA*
everywhere on A.

A Special Class of Subsets of D

In discussing capacitary measures AA relative to D it will be convenient to


restrict A somewhat. Observe (Theorem XI.6) that for an arbitrary subset
A of D the set A - Af is polar and so has capacity 0 and is null for every
measure that has finite energy or (Theorem V.11) whose potential is finite
valued. Furthermore (Theorem XI.6) the set Af is a Borel set and so is
capacitable. Thus, if we restriqt A to be in rp so that AA is defined and further
restrict A to be fine closed in D, that is, D n Af c A, then the representation
A = (A - Af) u (D n Af) shows that A is capacitable and is measurable for
every measure A on A of finite energy or with a finite-valued potential. This
restriction on A does not restrict the class of capacitary measures because
(Section XI. 7) C 10A = I 10D A-'' Furthermore the capacitary measure A" under
this restriction on A is supported by A and in fact (by Theorem XI. 18) is
even supported by D n dfA! and C(A) = AA(A) according to (10.1). In
particular, C(A) < + oo if A is also relatively compact (Euclidean topology)
in D. We shall use the fact that if A is a measure on D, vanishing on polar
sets and supported by the fine closed in D set A, then

[A, )A] = = A(A). (12.1)

In (a)-(c) below we characterize AA and A,; by maximal and minimal


properties of measures A on D satisfying the stated side conditions under
the hypothesis
A e r,, D n Af c A, 0< C(A) <+ oo.
(a) Side condition: G°2 S 1; A is supported by A. Under this side
condition
(al) Iku 2 5 C(A).
(a2) A(A) 5 C(A).
Equality in either inequality implies that A = AA.
242 1. X111. Classical Energy and Capacity

In fact (special domination principle in Section 2), GDA 5 GDAA; so


(by Lemma 4) (al) is true, and an application of Schwarz's inequality in
(12.1) shows that

A(A) 5 IIAAII2 = C(A); (12.2)

so (a2) is true. Equality in (a2) implies equality in (al) according to (12.2),


and [by Lemma 4(b)] equality in (al) implies that A = AA.
(b) Side condition: A(A) = 1; A is supported by A. Under this side
condition,
(bl) z 1/C(A).
(b2) z 1/C(A).
Equality in one of these inequalities implies that A = AA.
Define 1.' = C(A)A. The measure A' is supported by A, and A'(A) = C(A).
If (bl) is false, then GDA' < 1 and A' = AA according to (a), a contradiction.
If there is equality in (bI), this reasoning shows that A = A,',. Inequality (b2)
follows from Schwarz's inequality applied to (12.1), and (by Theorem 7)
there is equality in this application of Schwarz's inequality if and only if
A = const AA, that is, A = AA.
(c) Side condition: GA z I quasi everywhere on A. Under this side
condition, IIAII2 z C(A), and equality implies that A = AA.
According to the domination principle, GDA z GDAA, and therefore
(by Lemma 4) IIAII2 ? IIAA112 = C(A), and equality implies that A = A(A).

Observation. Under the additional condition that A is supported by A, the


side condition (c) implies that A(A) z C(A) because

C(A) = ).A(A) < f GDA dAA = f GDA, dA = A(A). (12.3)


A A

(d) Side condition : 11211 < + oc ; A is supported by A. Under this side


condition,

f (GD;. - 2) dA z -C(A), (12.4)


JA

and equality implies that A = AA. In fact (12.1) implies that A(A) < + oo
and that

fA (G
DA-2)dA=IIA112-2i.(A)=jj2-AA112-C(A). (12.5)
13. Expressions for C(A) 243

13. Expressions for C(A)


A slight variation of the discussion in Section 12 yields the following theorem.

Theorem. Let D be a Greenian subset of RN, and let A be an analytic subset


of D of finite capacity. Then

C(A) = sup { f GDAdal: A supported by A, GDA 5 11


,1A )
= sup {A(A): A supported by A, GDA 5 1)
I
inf {sup A supported by A, A(A) = 1)
CsD (13.1)

inf! f GDAdi: A supported by A, A(A) = 11


I. JA

= inf f GDA dA: GD). z 1 quasi everywhere on Al


. )
= inf {A(A): A supported by A, GD). z 1 quasi everywhere on A}

These expressions for C(A) are so easily obtained that we shall prove
only the first. If A is supported by A and if GDA 5 1, then (domination
principle) GD. S GDAA; so (by Lemma 4) IIAII2 s IIAAII2 = C(A); that is, each
integral on the right in the first line of (13.1) is at most C(A). Furthermore
C(A) is the supremum of the capacities of the compact subsets of A; so if
A = AF in (13.1), with F a compact subset of A, the integral can be made
arbitrarily near C(A).

Capacity of a Ball
If N > 2, if D = R", and if A is a ball of radius S, a uniform distribution
on aA of total value 1 has potential identically S-"2 on A according to
Section IV.2. Hence any uniform distribution on 8A is an equilibrium distri-
bution for aA and for A. Both have capacity 0-2. Moreover, since C(A)
is the supremum of the capacities of the compact subsets of A, C(A) = 6N-2
also, and any uniform distribution on 8A is an equilibrium measure for A.
If N z 2, if D is a ball of radius fi, and if A is a concentric ball of radius
a < fi, a uniform distribution on 8A is an equilibrium distribution for A, A,
and aA, and

_ f(lo)-'
a
if N = 2,
C(A) = C(A) = C(BA) = (13.2)
l2_tl _
R2-N)-l if N > 2.
244 I. Xl 11. Classical Energy and Capacity

14. The Gauss Minimum Problems and Their Relation to


Reductions
Let D be a connected Greenian subset of D. Energies and reductions below
are relative to D, and charges are on D. Let A be a nonpolar Borel subset
of D, and let f be a Borel measurable function from A into Ul;, not vanishing
quasi everywhere, with If I majorized quasi everywhere on A by the restric-
tion to A of the potential GD,1 of some measure A of finite energy supported
by A. Observe that if A is fine closed in D, the words "supported by A"
are no restriction on f because if A is a measure on D of finite energy and if
If 1:5 GDa. quasi everywhere on A, then If 15 OGD).DA = GD)l quasi
everywhere on A, the measure II).0A is supported by A (Theorem XI.14), and
IIAII < +oo (Lemma 4).
Let I'A be the linear class of charges v supported by A, of finite energy,
with 1A I f I dlvl < + oo. This class contains A. We consider two modifications
of a problem studied in 1840 by Gauss (see-Historical Notes to this chapter)
under hypotheses suited to his era.

(GI) Minimize f GDvdv=IIvII2forVEr'Awith j'fav= 1.


A

( G2) Minimize [(GDv - 2f)dv = IIvIIZ - 2 fA fdv for ver'A.

L et r,; be the class of positive charges, that is, of measures, in rA, and let
problems (G1+) and (G2) be, respectively, (G1) and (G2) with f'A replaced
by 17A. We shall treat problems (G1+) and (G2+) only when f z 0. We shall
write that a charge solves one of the above problems if the charge minimizes
the relevant integral under the specified side conditions.
(a) A charge p solves problem (G2) if and only if

[µ, v] = ffdv (14.1)

for every v in r'A, equivalently, if and only if GDµ = f quasi everywhere on A. If


so,

f (GDµ-2f)dµ=-1Iµ112=- Jfdµ<0, (14.2)


A A

and for v in r'A,

(GDv-2f)dv= -11µ1I2+IIv-µI12: (14.3)


1.

.so µ is the only charge solving (G2). Furthermore µ/f A f dp solves problem (G 1).
In fact µ solves (G2) if and only if whenever c e 18 and v e FA, the integral in
14. The Gauss Minimum Problems and Their Relauon to Reductions 245

(G2) with v replaced by u + cv has its minimum value for fixed v when c = 0.
A condition necessary and sufficient for this is (14.1). Equation (14.1)
implies that Ilµllz = jAfdµ and implies the evaluations (14.2) and (14.3)
except for the strict inequality in (14.2), which is a consequence of the
evaluation of G0µ in terms off to be made next. Equation (14.1) with v
replaced by its projection on an arbitrary Borel subset of A yields the in-
equality G0µ = f v almost everywhere on A for every v, and therefore (Section
2) G0µ =f quasi everywhere on A. Conversely, the latter condition implies
(14.1) because (from Section 2) polar sets are Ivl null for charges v of finite
energy. Finally, if v e r'A and if JA f dv = 1, equation (14.1) implies that
IlµllllvIIzI;so
µ 2
llV1l2 z IIpli-2 =
jfd
and therefore µ/jAfdµ solves problem (G1).
(b) A charge p solves problem (GI) if and only if Ilpll > 0 and µllµll-2
solves problem (G2), equivalently, if and only if llpll > 0 and G0µ = llpll2f
quasi everywhere on A. There can be only one such charge. In fact, if p solves
problem (G1), then the side condition implies that Ilpll > 0, and the min-
imizing property of p implies that for v in 17A,

J(Gnv - 2f)& z (JA1)2hph2 - 2 fdv


A

_
_111111-2
+
(111411.14
- Ilpll-t)2
with equality when v = pllpll-2 Thus the latter charge solves problem (G2),
and the rest of (b) is now trivial.
(a+) If f z 0, a measure u solves problem (G2+) if and only if

[p, v] z f fdv, Ilpll2 = ffd14 (14.1+)


A

for every v in r,, equivalently, if and only if G0µ ;2: f quasi everywhere on A
with equality µ almost everywhere on A. If so, (14.2) is true, and for v in r,;
the identity

j(Gov_2f)dv= -Ilpll2+llv-PII2+2(IM, v]-


f fdv)(14.3+)
A

shows that p is the only measure solving (G2+). Furthermore µ/jAfdµ solves
problem (G1+). In fact, if p solves problem (G2') and if ce R , the integral
246 1. XIII. Classical Energy and Capacity

in (G2') with v replaced by cµ has its minimum value when c = 1, and the
integral in (G2') with v replaced by µ + cv with c in R+ and v in ri has its
minimum value for fixed v when c = 0. These two conditions are satisfied
if and only if (14.1') is true. Conversely (14.1') implies that y solves problem
(G2+) and is thereby uniquely determined because under (14.1') in the
identity (14.3') the right-hand side z -11µ1I2 with equality only when v = p.
Equation (14.1') with v replaced by its projection on an arbitrary Borel
subset of A implies that G6µ 2:f v almost everywhere on A; so GDµ ;->f
quasi everywhere on A. There is equality p almost everywhere on A in
view of the equality in (14.1' ). Conversely, these two conditions on y imply
that (14.1 +) is true. Finally, if p solves (G2' ), then µ/ f A f dµ solves (G 1') by
the same proof as that of the corresponding implication in (a).
(b') /f f z 0, a measure p solves problem (G 1') if and only if l1µ II > 0 and
µ11µI1-2 solves problem (G2'), equivalently, if and only if l1µ11 > 0 and GDP Z
IIpII2fquasi everywhere on A with equality p almost everywhere on A. There
can be only one such measure. The proof is left to the reader.
(c) Suppose that f z 0, and define f as 0 on D - A. Then if µ solves
problem (G2'), GDP = R . (See Section XI.20 for the reduction involved
here.) Conversely, if A is fine closed in D, then R is the potential of a measure
p solving problem (G2'). In fact, if u solves problem (G2'), then GDµ = f <
+ oo p almost everywhere on A; so if v is a positive superharmonic function
on D that majorizes f quasi everywhere on D, the domination principle
states that GDµ 5 v. Since one choice of v is G6µ, it follows that GDµ = Rf.
Conversely, R f quasi everywhere on A, and R < GDA; so R is the
potential of a measure p of finite energy (Lemma 4). To prove that p solves
problem (G2+), we need only verify that A supports p, and according to
Section X1.20 (special positivity case), µ is supported by cdfAf - D, a subset
of A since A c Af by hypothesis.
(d) Classical balayage. Suppose that A is a Borel subset of D fine closed
in D and that f is equal quasi everywhere on A to the restriction to A of the
potential GDµ' of some measure µ' of finite energy not necessarily supported
by A. We have seen at the beginning of this section that then f satisfies the
conditions imposed throughout. In the present context, problem (Gl)
becomes the problem of minimizing 11v112 for all charges v of finite energy
supported by A with [p'. v] = 1, and problem (G2) becomes the problem of
minimizing 11 v112 - 2 [p', v] for all charges v of finite energy supported by A.
The measure 8 p, [A solves (G2) because

[G6µ'" =f
quasi everywhere on A and (by Theorem XI.14) the measure iµ'0" is sup-
ported by A.
The linearity properties of (G2) solutions noted in (a) imply that if f is
15. Dependence of C* on D 247

equal quasi everywhere on the fine closed in D set A to the restriction to A


of the potential of a charge u' of finite energy, then problem (G2) has a
solution: in fact, if p' = p'+ - µ'- (Hahn decomposition), then µ'+8A -
p'- @A solves (G2).
In particular, if A is a fine-closed Borel relatively compact subset of D
and if f - I on A, f is the restriction to A of the potential of a measure of
finite energy; for example, f is the restriction to A of the reduction of the
function 1, reduced onto an open superset of A relatively compact in D. In
this special case problem (G 1) is the problem of minimizing JJvll2 for charges
v of finite energy supported by A with v(A) = 1, and problem (G2) is the
problem of minimizing 11v112 - 2v(A) for all charges of finite energy sup-
ported by A. The capacitary measure of A solves problem (G2) in this
context.
The Frostman approach (see Historical Notes). If A is compact and if f is
positive (< +oo) and continuous, it is easy to see that problems (G1+) and
(G2+) have solutions. In fact, if p. is a minimizing sequence of measures for
(G 1+), then p.(A) is a bounded sequence, and any measure that is the vague
limit of a subsequence of p, solves (G I+).

15. Dependence of C* on D
If D, and D2 are Greenian subsets of R' with D, c D2, we have noted in
Section VII. I that GD, 5 GD2. If outer capacity relative to D, is denoted by
C7, it will now be shown that C,* >- CZ on subsets of D, and that to a
compact subset B of D, corresponds a constant a = a(B) such that C,* 5 aC2
on subsets of B. In view of the relations between inner and outer capacities,
the corresponding inequalities are true for inner capacities, and it is sufficient
to prove these inequalities on compact sets, for which C,* reduces to the
capacity function C. If A is a compact subset of D, with capacitary measure
2;A relative to D;, then I >: GD2A2A > GD,A2A on D,; so by (13.1)

C(A) = sup {A(A): A supported by A, GD,A 5 1) z A2A(A) = C2(A), (15.1)

as asserted. To prove the second assertion, observe that if B is a compact


subset of D, and a = sup8 K 8 GD,/GD, (the ratio is defined as 1 on the diagonal)
and if A is a compact subset of B, then GD,22A 5 "GL), )2A on D, ; so GD, (rA2A)
z I quasi everywhere on A. Therefore by (13.1)

C, (A) = inf (A(A): A supported by A, GD, 2 z I quasi everywhere on A)


5 0(i.2A(A) = aC2(A), (15.2)

as was to be proved.
248 I. X111. Classical Energy and Capacity

16. Energy Relative to R2


In this section we consider potentials Gµ when N = 2. The Green function
G is bounded below on compact subsets of 082 x 082, and we have seen in
Section IV.I that Gp is superharmonic on 082 whenever µ is a measure on
082 with compact support. Hence, if p and v are measures on 082 with compact
support, the integrals I R2 Gp dv and 11112 Gv dµ are meaningful and equal.
Their value, the mutual energy of the pair of measures, will be denoted by
[µ, v].
Let 9 in the present context be the class of charges µ: (µt,µ2) with the
following properties (in which we suppose that pt and lee are minimal
components) :

E1. µt and µ2 have compact support.


E2. Gp; dui < + cc, i = 1, 2.
R2
E3. µ(O82) = 0.

Then the following two assertions are true.


(a) If µ and v are measures in B', then [p, v] < + oo.
This assertion follows from the fact that if D is a ball containing the
compact supports of p and v the potentials Gµ [Gv] and Goµ [Gov] differ
on D by harmonic functions; so (a) can be reduced to the corresponding
assertion for potentials relative to D, covered by Theorem 7.
In view of (a), if p and v are charges in 8', their mutual energy [µ, v]
defined formally by (3.2) in terms of their minimal components is meaningful
and finite valued. The evaluation (3.2) is independent of the choice of
components of the charges as long as all the mutual energy integrals are
finite valued. The class 9 is now seen to be a linear class.
We have not yet used the condition E3 in this discussion, but this condi-
tion is essential in the following. In view of the evaluation (6.3) of mutual
energy in terms of the Dirichlet integral when the charges involved are in B
and have E'21(R2) potentials, we can expect charges in tf to have positive
energy, and in fact we now prove the following statement.
(b) Theorem 7 is true for D = R2 and charges in 8.
To prove that [p, µ] Z 0 for µ in 9°, let A be a compact support for p, of
diameter a, and let B be a ball containing A, with center in A, of radius
ft > a. Define

u( > h) = GM pB(q, G(c, )) (16.1)


B

Band

- log (fl + a) < u 5 - log (fl - a). (16.2)


17. The Wiener Thinness Criterion 249

In view of the positivity of energy relative to B (Theorem 7), if p is a charge


in d (for R2) with minimal representation (p ,' p2)'

Z2pi(A)2log.
JR2 AA

Since the last term has limit 0 when i4 - oo, it follows that the energy [p,p]
of p relative to R2 is positive.
Schwarz's inequality (7.1) in the present context follows easily from
positivity of energy [see (7.3)]. The rest of the proof of Theorem 7 in the
present context follows that of Theorem 7 with one modification. To prove
that [p, v] = 0 whenever v is in 8 implies that u is the zero charge, we cannot
choose v as a uniform distribution on a sphere since this distribution is not
in .9. Instead choose S, > S > 0, and for in R2 let v be the charge supported
by b) v S,), equal on the larger sphere to the uniform distribu-
tion of a unit mass and equal on the smaller sphere to the uniform distribution
of a negative unit mass. Define u = GDp. The equality [p, v] = 0 implies that
L(u, , d) = L(u, , b,), and the condition E3 implies that lima,-. L(u, , S,)
= 0. It follows that L(u, , b) = 0 and therefore (S - 0) that u - 0. Hence p
is the zero charge, as was to be proved.

17. The Wiener Thinness Criterion


Let D be a Greenian subset of R"(N z 2), let A be a subset of D, and let
be a point of D. Let C* be the outer capacity !Section 11) defined on subsets
of D, relative to D. Let a e ] 1, + oo [, n e 7+, and define

B"={n: A"=Ar, B", A"=UA,,. (17.1)


"

Let k be an integer so large that Uk BB c D and, if N = 2, so large that this


union has diameter less than 1. In the following k is fixed, and it is to be
understood that A. and A" are considered only for n >_ k. Nothing in the
following theorem or its proof would have to be changed if one of the
inequalities in (17.1) is changed to be a strict inequality. Observe that there
is a constant c so large that

0<G( ' ) < on Ak. (17.2)


c

Theorem. The set A is thin at if and only if the following equivalent conditions
(with reductions relative to D) are satisfied:

(a) Ya"C*(A") < +oc.


0
250 I. XIII. Classical Energy and Capaaty

m
(b) YR;n() < +co.
0

Observe that since R;" is harmonic on a neighborhood of , it follows


that Ri "( ) = R;"(i). Since A. is a relatively compact subset of D, the
smoothed reduction Rl A. is a potential GDAAn, and [by (10.1)] C*(A") =
n.4"(D). Since AA" is supported by An,

a"C*(A") _< R1 "() = a""C*(A,) (17.3)

Hence conditions (a) and (b) are equivalent. In view of Theorem X1.3 these
conditions imply that A is thin at because under (b)

lim R;"(i) 5 lim 0, (17.4)

where we have used the countable subadditivity of R; [Theorem VI.3(j)].


Conversely, if A is thin at , we shall show that the sum in (b) is finite when
the index n is even; the proof for n odd is the same. Define A' = U:
A and therefore thin at c. If the part of A' in some neighborhood
of is polar, it is trivial that the series in (b) with n even converges. If A'
meets every neighborhood of in a nonpolar set, we use the fact that accord-
ing to Theorem XI.2, there is a positive superharmonic function u0 on D,
finite at , with limit + oo at i along A'. The function RA ' is a positive super-
0
harmonic function on D, finite at , with limit + oo at along A' less a polar
set, and R is a potential GDp because A' is relatively compact in D. The
measure p is supported by A', and 0 because +oe. Let µ2n
and pi,. be the projections of p on A2,, and D - A 2", respectively. An elemen-
tary calculation shows that there is a constant c' such that for all n k,

G(C, q) 5 q) for 'i a A' - A 2", C e A2n. (17.5)

It follows from the definition of GD and the discussion in Section VII.3 that
liml4-also [GD(C, q)/G(g, q)] = I when C and q are restricted to be in a com-
pact subset of D. Hence for C and q in a compact subset B of D there is
a constant c" = c"(B) such that GD < c"G on B x B. (If N > 2, the stronger
relation GD < G is valid on D x D.) In view of this inequality for B = Ak
and (17.2).

GD(s, q) < cc"cGD(S, q) for q e A' - A2n, C E A2n, n > k.

Hence for n > k,

GDu2" 5 const GDµ4n(S) 5 const u(i)


18, The Robin Constant and Equilibrium Measures Relative to R2 (N = 2) 251

on AZn, and the constant does not depend on n. It follows that GDµ2n Z 1
quasi everywhere on AZn for sufficiently large n; so Ri 2n 5 GDN2n quasi
everywhere on A2i for sufficiently large n, and therefore this reduction
inequality holds everywhere on D (domination principle) for sufficiently
large n. Hence, if m is sufficiently large,

Ri S Y_ GDp2nS) S GDP() < +w. (17.6)


"am "am Ram

Thus the sum in Theorem 17(b) over the terms with even n converges, as
was to be proved.
Observation. In view of (17.2) the theorem is true, and the proof requires
only trivial modification, if G in (17.1) is replaced by GD.

18. The Robin Constant and Equilibrium Measures Relative


to R2 (N = 2)
Since we shall not use the results of this section, detailed proofs will be
omitted. The material is presented as an interesting and important appli-
cation not readily available elsewhere.

Superharmonic and Harmonic Functions on Neighborhoods of oo

A function defined on an open neighborhood of the point oo of R2 (that is,


on an open set including this point) is said to be superharmonic [harmonic]
there if the function is superharmonic [harmonic] on the deleted neigh-
borhood and if the Kelvin transform of the function under an inversion is
superharmonic [harmonic] on an open neighborhood of the image of oo.
We shall use obvious consequences of theorems on superharmonic and
harmonic functions defined on open subsets of R2 even when oo is allowed
in the domains of the functions. For example, if u is a positive superharmonic
function defined on an open deleted neighborhood of oo, then u has a
superharmonic extension to the full neighborhood.

The Green Function of an Open Neighborhood of oo and the Robin


Constant of Its Complement

Let A be a compact nonpolar subset of R2, and define D = R2 - A. The


Green function Go has a continuous extension to (D i {x}) x (D u { 00)),
and we shall denote by GD(oo, ) the restriction to D of this extension with
first argument fixed at oo. The function GD(oo, ) is called the Green function
of D with pole oo. This function is positive and harmonic on D, is bounded
252 1. X111. Classical Energy and Capacity

outside each neighborhood of oo, and has limit 0 at quasi every finite
(Euclidean) boundary point of D, and the difference has
a finite limit r(A) at co. [If this difference is defined as r(A) at oo, the dif-
ference becomes harmonic on a neighborhood of oo.] The value r(A) is
called the Robin constant of A. Obviously r(A) is invariant under rotation
and translation of R2. The value a-"" is called the logarithmic capacity
of A. Although the logarithmic capacity has the advantage of positivity,
we shall see that the key set function in this context is - r(-).

EXAMPLE. If A = B(O, S), then GD(oo, ) = log (J 1/b), r(A) = -logo, and the
logarithmic capacity of A is therefore 0.
If o a A, then [all functions in (18.1) are defined on D]

GD(oo, ) = inf {u > 0: u = log 1 + h, h superharmonic}


= inf {u z 0: u = log I + h, h harmonic}
(18.1)
= inf {u z 0: u superharmonic,
u Z log I I on a deleted neighborhood of co}.

If D is not connected the function GD(oo, ) vanishes identically on every


open connected component of D except the component D. containing a
deleted neighborhood of oo.

Canonical Equilibrium Potential of a Compact Set

The function GD(oo, ) can be extended to a positive subharmonic function


G, (oo, ) on R2, vanishing quasi everywhere on A. The measure A, on R2
associated with the superharmonic function is supported by A,
in fact by the Euclidean boundary of Dc,, and even by We can write
GD (co, ) = -CAA + h, where his a harmonic function on R2. In view of the
form of GD near oo, it follows that AA(A) = I and [see IV(9.2)] that h has
limit r(A) at oo. An application of the harmonic function maximum-
minimum theorem to It shows that this function is identically r(A): so

Go (oo, ) = -GAA + r(A). (18.2)

Conversely, suppose that at is a constant, that A is a measure supported by


A, that A(A) = I, and that GA < at with equality quasi everywhere on A.
Then we now show that at = r(A) and ). = AA. Observe first that A and AA
vanish on polar sets (Section V.11), and we can therefore ignore the subsets
of A on which GA # a or GAA # r(A) in the following evaluations:

a = fA GA dA.A = fA GAA dpi = r(A).


18. The Robin Constant and Equilibrium Measures Relative to R2 (N = 2) 253

The domination principle now implies that GA = GAA ; so A = A.A. The mea-
sure AA is a canonical equilibrium measure for A in the present context.

Right and Left Continuity of the Function - r(-)

The function - is an increasing function on the class of nonpolar compact


subsets of R2. Now let A. be a monotone sequence of compact nonpolar
sets, with limit A. In view of Theorem VII.6, if A. is an increasing sequence
and if A is compact, then r(A) if A. is a decreasing sequence,
then lim,, r(A) A is nonpolar and lim,,... oo if A is
polar. If A is compact and polar, we therefore define its Robin constant
r(A) to be + oo and its logarithmic capacity to be 0.

Strong Subadditivity of - r(-)

Let B be a nonpolar compact subset of R2, and define D = RZ - B. If A


is an arbitrary compact subset of D, let vA be the reduction relative to D
of GD(oo, ) on A. This reduction is harmonic and bounded on a deleted
open neighborhood of the point oo; so (Section V.5) vA has an extension
harmonic on the full neighborhood and therefore has a finite limit vA(oo)
at oo. The restriction of the function GD(oo, ) - vA to D - A is GD_A(oo, ).
Observe that the set function A r+vA is strongly subadditive [Section
VI.3(j)]. It follows that the set function A i--. vA(oo) = -r(A u B) + r(B)
and so also the set function A i- - r(A v B) are strongly subadditive on
the class of compact subsets of R2 - B. When B shrinks to a point, we find
that is strongly subadditive on the class of compact (including polar
compact) subsets of R2. Unfortunately the logarithmic capacity function
is not strongly subadditive, in fact, not even subadditive.

Application of Section VIII.19

According to Section VIII. 19, if D is a deleted open neighborhood of OD,

R) = 7) - on('), (18.3)

where OD is a positive harmonic function on D, bounded on bounded subsets


of D, with boundary limit 0 at quasi every Euclidean boundary point of D.
Since the term on the left in (18.3) and the second term on the right both,
for fixed., define functions of bounded near oo, the function 4D must have
the form OD = log I I + h, where h is harmonic on D and bounded on a
deleted neighborhood of oo ; equivalently, OD = GD(oo, ) + h', where h'
is harmonic on D, h' is bounded on a deleted neighborhood of oo, and h'
has limit 0 at quasi every Euclidean boundary point of D. Moreover h' is
254 I. XIII. Classical Energy and Capacity

bounded on bounded subsets of D because GD(co, ) is, and it follows that


h' is bounded; so (Section V.7) h' vanishes identically. Thus WD = GD(oo, )
and
q) - GD(cc, ), (18.4)

or in view of the definition of GD,

GD(00, ) = IlDIS, G(q,')) - (18.5)

for every point q of D.

Extension of to the Class of Analytic Subsets of R2

The set function - r(-) very nearly satisfies the conditions defining a topo-
logical precapacity but is not positive. Let A. be a compact nonpolar subset
of R2, and consider the set function r(A0) defined on the class of
compact supersets of A0. This set function is a positive strongly subadditive
set function, and it is easy to modify the topological precapacity extension
theorem, using its methods to extend - r(-) + r(Ao), and therefore
to the class of analytic supersets of A0. The value -r(A) obtained in this
way does not depend on the choice of the compact nonpolar subset A0 of A.
Define -r(A) = - oo if A is polar. Since (from Section VI.2) every analytic
nonpolar set has a compact nonpolar subset, we have now defined - r(-) on
the class of analytic subsets of R2. The set function - is now an increasing
set function and is regular in the sense that

-r(A) = sup { -r(B): B c A, B compact}


(18.6)
= inf { - r(B) : A c B, B open}.

The logarithmic capacity set function a-'*) is then also regular in this sense
but as already noted is not strongly subadditive, in fact, not even sub-
additive.

Equilibrium Measures of Analytic Subsets of R2

Let A be an analytic nonpolar subset of R2, and let A' be a compact nonpolar
subset of A. Define D' = R2 - A' and D = R2 - A. The positive function

(smoothed reduction relative to D') is subharmonic on D and can be ex-


tended to be a positive subharmonic function on R2, 0 at quasi every point
of A. Denote this extension by GD (oo, ). Suppose first that A is bounded.
18. The Robin Constant and Equilibrium Measures Relative to R2 (N = 2) 255

Then Go (oo, ) is harmonic on a deleted open neighborhood of ao, and if


AA is the Riesz measure associated with - GD (oo, ), this measure has compact
support, A(RI) = 1, and there is a harmonic function h on RZ such that
Gp (oo, ) = -GAA + h. It can be shown that h is identically r(A). Thus
(18.2) is true, and GDAA s r(A) with equality quasi everywhere on A. If
A is not bounded, this reasoning must be expanded. In the first place, since
GD.(oo, ) is a minimal harmonic function on D' (Section VII. 10), we must
suppose (by Theorem XI.3) that A is thin at oo to ensure that
is not identically GD.(ao, ). This condition on A can be shown to be necessary
and sufficient for A to have finite logarithmic capacity, and then if AA is
the Riesz measure associated with -Go (oo, ), it can be shown that AA(RZ) =
1, that (18.2) is true, and that therefore again GAA S r(A) with equality
quasi everywhere on A.
Chapter XIV

One-Dimensional Potential Theory

1. Introduction
The one-dimensional version of classical potential theory is so special that
its discussion has been deferred to this chapter, and much of this theory is
so elementary that it will be left to the reader to formulate and justify. A
ball in R with center is an open interval with midpoint 1;, and the averages
L(u, . S), A(u, , S), and Asu can play the same role when N = 1 as when
N > 1, but more direct methods are sometimes clearer.
Since an open subset of I8 is a countable union of disjoint intervals, it is
usually possible to consider functions defined on intervals.

2. Harmonic, Superharmonic, and Subharmonic Functions


A function u defined on a nonempty open subset D of I8 will be called
harmonic (superharmonic, subharmonic) if on each component interval of D
the function u is finite valued and linear (concave, convex, respectively).
Such a function is necessarily continuous.
Every nonempty open subset of I8 except I8 itself supports a positive
not identically constant superharmonic function. In the terminology of
the multidimensional case F8 is the only non-Greenian open subset of F8
aside from the empty set. In agreement with the definition of a polar set
when N > 1, only the empty set is defined as polar when N = 1. That is,
the only negligible set in one-dimensional classical potential theory is the
empty set.

3. Convergence Theorems
It is trivial that the supremum of an upper directed family of harmonic
[superharmonic] functions on an open subset D of F8 is harmonic [super-
harmonic] if finite at a point of each component interval of D. It is just as
trivial that the infimum of any family of superharmonic functions on D
is superharmonic if finite valued. Thus the Fundamental Convergence
Theorem of classical potential theory is both true and trivial when N = 1.
5. The Dirichlet Problem (Euclidean Boundary) 257

If u is superharmonic or subharmonic on D and if B is an open subinterval


of D, relatively compact in D, the function TBu is defined as the function
equal to u on D - B and linear on B. If u is superharmonic, the function
TB is also superharmonic and is majorized by u.
If u is superharmonic on D and has a subharmonic minorant, the function
GMpu exists as in the case N > I and is harmonic. The proof for N > 1
is applicable to the case N = 1. A superharmonic function u on a finite
interval D has a subharmonic minorant if and only if the limit of u at each
endpoint of D is finite, and in that case GMpu is the restriction to D of the
linear function equal at each endpoint of D to the limit of u at that endpoint.
A superharmonic function u on the infinite interval ]a, +oo( with a finite
has a subharmonic minorant if and only if the limit of u at a is finite and the
limit a of u, at + oo is not - oo. In that case GMpu is the restriction to D
of the linear function with value at a the limit of u there and with slope a.

4. Smoothness Properties of Superharmonic and Subharmonic


Functions
We omit the elementary proofs of the following properties of a convex
(that is, subharmonic) function u defined on an open interval D.
(a) u is continuous, is absolutely continuous on each compact subinter-
val of D, and has finite or infinite limits at the endpoints of D.
(b) u has a right [left] derivate u, [u4] at each point of D.
(c) The function u, [ul] is monotone increasing and right [left] con-
tinuous.
(d) ui S u and there is equality except at a countable subset of D. At
a point of equality the derivative u' (= u; = u;) exists.
In one dimension the Laplacian of a function u is the second derivate
u". A function u in C12 (D) is subharmonic if and only if Au Z 0.

5. The Dirichlet Problem (Euclidean Boundary)


If D is a finite interval ]a, b[ and if f is a boundary function, it is trivial
that the PWB method in the present context yields a solution if and only
if f is finite valued and that for finite-valued f,

Hf=PD(',f), PD(S,{a})=b-a' (5.1)

On the other hand, if the interval D has one finite and one infinite endpoint,
the PWB method yields a solution if and only if the specified boundary
258 I.XIV. One-Dimensional Potential Theory

function f is finite valued, say at the finite endpoint b of D. The PWB


solution Hf is then identically fJ, that is, the harmonic measure of {b} is
identically 1. If D is not connected but is a proper subset of I8, the PWB
method is applied separately to each interval component of D, or equiva-
lently, to D itself, but the Dirichlet problem is not treated for D = 11 because
Id is not Greenian.
The Dirichlet problem for conditionally harmonic functions is left to
the reader.

6. Green Functions
When N = I and D is a finite interval ]a, b[, it is natural to define GD as a
function on D x D with the property that is harmonic on D -
and superharmonic on D, with limit 0 at each endpoint of D. These con-
ditions determine up to a multiplicative positive constant. Let d1
[d2] be the derivative of to the left [right] of . If now uE C121(b),
integration by parts yields

j d11 = u(a)dl - u(b)d2 - d2). (6.1)


a

To simplify this formula, we specify GD completely by prescribing d1 - d2 =


1, so that

if ry5 ,

(6.2)
a)
if11z ,
b -a
and then (6.1) becomes

U(0 = u) - 16 -)u" d11. (6.3)


a

The representation (6.3) should be compared with the corresponding repre-


sentation 1(8.6). Thus if uaC/2'(D) and if u is superharmonic and positive,
u is the sum of G MDu and the potential of the measure - Au d11. This example
of the Riesz decomposition will be extended to the general case in Section 9.
If D is the interval ]a, + oo [ with a finite, we let b tend to + oo in (6.2)
and thereby are led to the definition

a) A (q - a). (6.4)

If D = ] - oo, b[ with b finite, we define GD correspondingly. If D is a union


of two or more disjoint nonempty open intervals, p) is defined as
8. Identification of the Measure Defining a Potential 259

q) for , q in the same component interval Do of D and defined as 0


for , I in different component intervals. Thus GD is now defined for every
Greenian set D.

7. Potentials of Measures
If D is Greenian and if µ is a measure of Borel subsets of D, then the potential
GDµ is superharmonic, that is, concave, on each component interval of D
on which the potential is finite valued, because q) is concave for each
q. We now show that GMDU = 0 when u = GDµ is a superharmonic potential.
It will follow from the Riesz decomposition (Section 9) that conversely a
positive superharmonic function u with GMDU = 0 is the potential of a
measure. We can assume in the following that D is an interval ]a, b[. Let BB
be an increasing sequence of subintervals of D with compact closures in D
and union D. Then ra.GDu is a decreasing sequence of superharmonic
functions. Moreover, if e B,,, then

TB.(GDµ)(b) = (7.1)

Here n i-1 p5 (, q)) is a decreasing sequence, for n so large that e B,,,


with limit 0. Hence T8 GDµ = 0. It follows that u = GDµ has limit 0
at each finite endpoint of D and that if D has an infinite endpoint, the limit
of right and left derivatives u; and ui at that endpoint is 0. Hence (see Section
4) GMDU = 0. Alternatively [see the N-dimensional context in Section 111. 1,
Observation (a)), the sequence TB u can be proved directly to have limit
GMDU.

8. Identification of the Measure Defining a Potential


Theorem. If u = GDµ is a superharmonic potential, then dµ = du, in the sense
that

u,(/J) - t4(a) _ -µ(]a, fl]) for ]a, P] c D. (8.1)

In proving the theorem we can suppose that D is an interval ]a, b[ with


at least one finite endpoint. If both endpoints are finite, define functions
0, and cbb by

0a(0 = f(i -

4b(S) fb(b - q)µ(dq), a < < b.


J4
260 1. XIV. One-Dimensional Potential Theory

Here fs means the integral over ]a, fi]. The functions ¢, and ¢b are finite
valued, monotone increasing, and right continuous and satisfy

f (b - l) d4a(q) = f (q - a) dob(q),
z
e s
a<a<#<b, (8.3)

and

u( ) =
(b -)Wa(S) - ( - 4)
b-a ( 8 .

Apply (8.3) and (8.4) to find

u;() _ _M) b-a


+O,( ) (8 . 5)

first when p({i;}) = 0 and then (right continuity) for all . Then for a <
a<#<b

f (q - a)p(dq) + f." (b -
(8.6)

as was to be proved. If one endpoint, say b, is not finite, define 0. as in


(8.2), but define 4b(S) = - p(]l;, + oo [). Then (8.3), (8.4), (8.5) are replaced
by

4a0) - 4a0) =
a
' (1 - a) d4b(ti), (8.3')

0a(S) - ( - a)4b(c), (8.4')

u.() = -4(); (8.5')

so in this case also (8.1) is true.

9. Riesz Decomposition
Theorem. If u is a positive superharmonic function on an open proper subset
D of t, then

u = GMou + Goat, (9.1)

where du = du, in the sense of Theorem 8.


10 The Martin Boundary 261

We can assume that D is an interval ]a,b[ with at least one finite end-
point. In fact, however, we shall assume that both endpoints are finite,
leaving the other case to the reader. Equation (8.6) is trivial, and when
a j a and ig T b, we find that 0. and 4 as defined by (8.2) are finite valued,
monotone increasing, and right continuous and satisfy (8.3). Then (8.5)
is true up to an additive constant; so (8.4) is true up to a linear term; that is,
u = GDp + v, where v is linear. Since the GMD operation is linear, just as
in the multidimensional case, and since this operation on a superharmonic
potential yields 0, it follows that v = GMDu, and the proof of the theorem is
complete.

10. The Martin Boundary


Martin boundary theory in one dimension is rather trivial. For example,
if D = ]0, + oo [ and if o c- D, we define the Martin function with reference
point o in the obvious way.

K(q,) =
bo ^n q

We then find (q -+ 0, + oo) that the Martin space can be identified with
9+ ; the Martin boundary point 0 is associated with the minimal harmonic
function K(0, ) __ I; the Martin boundary point + oo is associated with
the minimal harmonic function f- K(+ oo, ) = /o.
Chapter XV

Parabolic Potential Theory : Basic Facts

1. Conventions

The potential theory based on the Laplace operator, developed in the


preceding chapters, will be called classical potential theory below. The
potential theory based on the heat operator A and its adjoint A, called
parabolic potential theory, will be developed in Chapters XV to XIX. Con-
cepts that are parabolic counterparts of classical concepts will be distin-
guished by dots or asterisks, depending on whether the concepts are related
to A or to A. Just as the domains of classical potential theory are subsets
of R", the domains of parabolic potential theory are subsets of "space time"
R"+' which we denote in this context by AN. Here N;?! 1, and the case
N = I is not exceptional. A point = s) of AN has space coordinate
in 18" and time coordinate s = ord (the ordinate of ), a point of R. The
point q' : (q, t) will be said to be [strictly] below 4: ( , s) if t 5 s [t < s].
If is a point of an open subset D of AN, the set of points of D [strictly]
below relative to b is the set of points of D that are endpoints of continuous
[strictly] downward-directed arcs from . That is, n is [strictly] below
relative to D if and only if there is a continuous function f from [0, 1] into
b for which f(0) f(1) and ordf is a [strictly] decreasing function.
The upper [lower] half-space of AN is the set {ord > 0) [ lord 4 < 0} ] and
the abscissa hyperplane is the set {ord c = 0). The boundary of a subset of A"
relative to the one-point compactification of A' will be called the Euclidean
boundary, and boundary will mean this boundary unless a different one is
specified.
Let b be an interval in AN, b = ]a,,b,[ x . x ]a",b"[ x ]s,,s2[.
The set of boundary points with ordinate value s2 will be called the upper
boundary, the set of boundary points with ordinate value s, will be called
the lower boundary, and the closure of the rest of the boundary will be
called the lateral boundary.
2. The Parabolic and Coparabolic Operators 263

2. The Parabolic and Coparabolic Operators


In the following discussion the Laplace operator A acting on a function
= s) - t ) defined on an open subset of R' is to act only on the space
variable . Choose a strictly positive number a (fixed throughout the discus-
sion). Define the parabolic operator A and the coparabolic operator A,
operating on sufficiently smooth functions defined on open subsets of C?8", by

, s)
eu(, s) = 2 Ari(,, s) -
as
(2.1)

s) = 2 s) + alias, s)

Parabolic potential theory is based on the pair b, A and is similar in many


respects to classical potential theory, but the fact that both b and A are
involved means that two theories dual to each other must be considered
simultaneously.
A function iu from an open subset of 18" into P. in class V" there and
also in class Ct2' relative to the space variable, and satisfying the heat equation
Ori = 0 will be called parabolic; a solution (satisfying the same smoothness
conditions) of the adjoint equation Au = 0 will be called coparabolic. A
function ti is coparabolic if and only if the function s) r-
u is a function on an open subset D of R', if d = D x68,
and if ti(i) = for = s) in d, the function is is parabolic (equivalently,
coparabolic) on 15 if and only if u is harmonic on D.

EXAMPLE (a). If y is an N-dimensional vector, the function

s) H exp [<'i> + 62I7I2 2SJ

is a positive parabolic function on d8". Recall that in contrast with this


example there is no nonconstant positive harmonic function on 18".

Minimal Parabolic Functions

A positive parabolic function is on an open subset 6 of ft' will be called


minimal if every positive parabolic function on d majorized by u is a constant
multiple of is. If 1) is 9" or is the lower half-space, it will be shown in Section
XVI.8 that the restrictions to b of the functions in Example (a) are minimal
and in fact are the only parabolic minimal functions up to constant multiples.
264 I. XV. Parabolic Potential Theory: Basic Facts

EXAMPLE (b). If ii is parabolic on AN and is of the form


the functions f and g satisfy the equations -(r) + [(N - 1)/r] f (r) = cf(r)
for r z 0 and g' = cazg/2 on R, and we thereby find the parabolic function
ti defined on A' by

(CI V), _
u(S) = eo'a12 r
2 I)(N2)/2J(i.JII)ec2cs/2
N/2) -
2 (N- zuz
p
1(N-z),2(\CI, I)e°=Cs(2

where c is an arbitrary constant, Jk is the Bessel function of the first kind,


and !k is the modified Bessel function. The parabolic function u/I-(N/2) can
also be obtained by taking the parabolic function in Example (a) and
averaging it over the values of y with 1712 = c. This representation of ti/r(N/2)
is an example of the fact that every positive parabolic function on ftN is an
integral over the set of minimal parabolic functions.

3. Coparabolic Polynomials
Define the Hermite polynomial Hm...N on RN by
am,+ +M N e-1,11=
H.,. "N(,I) = z
(m) Z 0)
an(N)mN

so that if y is a vector in IBN, the Taylor expansion

e In+rIZ = e In1' C'


ao
j' y
(1)m, ..
y
(N)IN
m,. (3.2)
n=0 m,+..`+mN=n ml ( ... MN!

yields
m (INn, (NV"N
e z<r.o> I712 _ Y- y m1!
Hm,...mN(t1). (3.3)
nOm,+...+mN=n t ... MN!
N!

According to (3.1),

11m,..mN = elnl' N
(-2n(I)e-171=)
brl an(I )m, . . . a'e)'"N
(3.4)
= -2mjHm....mN, m. =mj-blf ifmj>0,
and this partial derivative vanishes if m1= 0; so (3.4) is valid in all cases
when interpreted reasonably. Repeated integration by parts leads to
3. Coparabolic Polynomials 265

e-...rN(7)Hn,..
fIRI'H. "N(7)1N(d7)
RN
m, ! if m = n,,
=10 otherwise.

Thus the sequence of Hermite polynomials is an orthogonal sequence rela-


tive to the measure exp (- 7I2)1N(rlq), and (3.3) is the corresponding Fourier
series of the left side.
The function jyj2a2t/2) is coparabolic. Define the
space-time Hermite polynomial ...mN, homogeneous of degree ml +
+ MN in the variables q"', ... , (11,1112 , by the Taylor expansion

e<r.n>-Irl=02hl2 = C' )
,(IV", ... y (N)"N
H . (7, t).
(... MN! m
.
, . MN
n==Om,+...+mN=" ml

The polynomial Hm, ...mN is coparabolic because it is obtained by repeated


differentiation of the coparabolic function on the left side of (3.6). If y and
q in (3.3) are replaced by (621/2)"2y and -(2621)-112 q, respectively, the left
sides of (3.3) and (3.6) become identical; so
2/2
Hm.mN(q, t) = (!_) Hm,...mN(-(2621)-1127)
= m)
(n
2 _ 2

_ (-721)"eXP2a2t an(I)MI .On


aq(Nr,Nexp (3.7)
2721

711Vn, .. 7(N)MN + , . .

The term written on the last line is the only term not involving t. The relation
(3.4) becomes

(Hm,...mN - m; - by ifmj>0,
(3.8)
mi= to ifmj=0,
and (3.5) becomes

e-Iel=nc'`Hr,...-N(7+t)H",..."N(n, t)1Nd7)
I
(a2t)n+N/2(2n)NJ2mI ! ... MN! if m, = n.,
I 0 otherwise.

For each value of t the sequence of space-time Hermite polynomials is an


orthgononal sequence in the space variables, and (3.6) is the corresponding
Fourier expansion of the left side.
266 I. XV. Parabolic Potential Theory: Basic Facts

A coparabolic polynomial f (# 0) must contain a term not involving t


because

M, t) = E r"rf,(7), 0:5 n t < ... < n,,


rI
where f; is a not identically 0 polynomial in the components of 7, and f
cannot satisfy the heat equation dual unless nt = 0. In view of (3.7) some
linear combination H of space-time Hermite polynomials can be chosen
with the same terms not involving t as f, and therefore f - H vanishes
identically. Thus every coparabolic polynomial is a linear combination of
the space-time Hermite polynomials.

EXAMPLE. If N = 1, the successive Hermite and space-time Hermite poly-


nomials are

Ho= 1, Ht(11)= -211, H2 H3(7)= -8113+ 1211, ...,

Ho = 1, R, (q, t) = 7, H,(7, t) = 7z - c21, 113(11, t) = 73 - 311o-'t, ... .

4. The Parabolic Green Function of AN

Define the function t on 1'8' by


_
W
(2nazt)-"1z exp 2 1ht if t > 0,
Al 7) = (4.1)

0 ift50.
Note that the time variable is placed first in this notation, as appropriate
to the probability interpretation to be given later. The parabolic Green
function 6 of 9' is defined on 1'8' x A' (N z 1) by

n) = O((, S), (7, t)) = AS - t, - 0). (4.2)

In more detail,
(a) The function h) is the Green function with pole n for the heat
equation. This function is positive, parabolic on A' - {q}, and
vanishes below q and in the limit at the point oo.
(a*) The function G( , ) is the Green function with pole for the adjoint
equation. This function is positive, coparabolic on A' - { }, and
vanishes above and in the limit at the point oo.
5. Maximum-Minimum Parabolic Function Theorem 267

Observe that when N > 2, there is an intimate connection between the


parabolic Green function of iY and the classical Green function of R':
o
,s),(n, t))l, (dt) = J 1(s, - n)lt(ds)
o (4.3)
zz z n)
n12-N = aNG(S,
= aNI S -
with
_ T(N/2 - 1)
aN - 2Q27rN/2

This relation between G and 6 will be generalized to Greenian subsets of


R" in Section XVII.18.
The following inequality will be used below. Let D(k) be a (possibly mixed)
partial derivation operator of order k on space variables, and let fi be a
positive number. Then if s) and

nI 2
ID(k>Q,7)I I - III S c(fl,o)(s - (s > t). (4.4)
- t)
t)-(N+k-eu2exp - 4oI2(s -

This inequality will be proved for k = 1. The proof in the general case
involves more notation but no additional ideas. For k = 1 and s > I

acMIT- 3C1
IC - oil* 5 (2a) "'a " _(I r 20
(4.5)

_ (2x)-Nao-"-:(f - naexp -I I1KIf - MI: r+iuiexP -IC - 7I)1

and the right side is majorized by the right side of (4.4) when k = 1.
If p is a measure on A", the functions Gµ and FRO defined by

O(, ,G(,1) = I O(, l)p(d) (4.6)


JUN .r RN

will be called respectively the potential and copotential of µ on d2;". These


definitions will be generalized to potentials and copotentials of measures
on an open subset 6 of A" with parabolic Green function Oo.

5. Maximum-Minimum Parabolic Function Theorem


This theorem will follow from the superparabolic function minimum theorem
(Section 13), but the following direct proof is instructive. Since the function
n') is parabolic on b = 9" - {q} and takes on its minimum value at
268 I.XV. Parabolic Potential Theory: Basic Facts

every point of b below 4, the maximum-minimum parabolic function


theorem is necessarily weaker than its harmonic function counterpart.

Theorem. Let l; be a point of an open subset D of ft N , and let D(d) be the set
of points of D below C relative to D. If ti is a parabolic function on ,6 and if t (e)
is the supremum or infimum of the restriction of u to D(C), then u = ti(C) on
bg).
Observation. This theorem when applied to the part of D strictly below
an arbitrary horizontal hyperplane implies that if m is the supremum of ti,
there is a sequence 0. of boundary points of D with ordinate values strictly
less than supo ord ry such that

lim lim sup u(C) = M.


n-.D p,f_Rn

Thus if b is an interval, there is a sequence in ,6 converging to a point of the


union of lower and lateral boundaries of ,6 such that ti tends to m along this
sequence. If u is parabolic on a neighborhood of the closure of the interval
A the restriction of ti to b attains its maximum and minimum on the union
of lower and lateral boundaries.
In proving the theorem we need consider only suprema, and it is sufficient
to prove the following result. Let S be a closed line segment in D, not ortho-
gonal to the ordinate axis, and let m be the supremum of ti on the set of points
of b below the highest point of S. We prove that ri < m on S if this inequality
is true at the lowest point of S. Let : s) be the upper endpoint of S,
and to simplify the notation, let the origin be the lower endpoint of S.
If u(0) < in, choose r so that 0 < r < I, so that the set

Do={(ry,t):I,1- Stl <r,0<t<s}

has closure in D, and so that mt = supl.l., ti(ry, 0) < m. Define S = 8(q, t)


= r2 - Iq - tt/s12 and define ti on 9' by

6(q, t) = m - (m - mt)S2e-", (5.1)

where a is a positive constant to be chosen below. Then if and q have jth


coordinates 'J' and ry'j', respectively,

Av(,1, t) = (m - mt)e °` {_2 + 26 [2N + 2)

+ ?I(J) - y S't - (S( J))]


2 tt - 2r2 }
6. Application of Green's Theorem 269

As a function of 6 the quantity in braces has maximum valuel1


N
c2(N + 2) + 2 (,7(i) - (t't/s)(1/s) 2
J
-4Q2r2 + '-` a (5.3)

Choose a so that this maximum value is strictly negative on a neighborhood


of 60, on which therefore b(t - u) < 0. On the lower boundary of Do, that
is, on its lower face, v - ii z m - (m - m,) - m, = 0, and on the lateral
boundary of Do we fine that 6 - ti >- m - m = 0. Now if the minimum
value of 6 - u on Do is attained at a point (ry, t) either in the interior of Do
or in the interior of the upper face of Do, then a(ti - a)/at < 0 at the point;
so A(6 - u) < 0 (Laplacian applied to the space coordinates) there. Since
(6 - t) has a local minimum at ,, the latter inequality is impossible,
and it follows that the restriction of v - ii to Do attains its minimum value
at a point of the union of the lateral boundary and lower face; so 6 - u z 0.
In particular, this inequality on the segment S becomes

ti 11 S m - (m - m,)re-°' < m (O S t 5 s), (5.4)

and the proof is complete.

6. Application of Green's Theorem


Let b be an open subset of 14". Suppose that for t in the interval [t', t"]
the set 6(t) = {tl: (ry, t)eb} is a nonempty connected subset of d8" smooth
enough for the application of Green's theorem in N dimensions. Denote by
ob(t) the boundary of b(t) relative to the hyperplane {ord = t}, and denote
by D. the directional derivative operator at a point of OD(t) in the direction
of the outward normal to OD(t). Define 6(t, , t2) =b n (R" x III, t2 [) for
t' S t, < 12 S 1". Denote by 15. the directional derivative operator at a point
of 10'60 1, t2) in the direction of the outward normal. The angle between
this normal and the upward-directed ordinate axis will be denoted by y.
In the integrals below the differential element will always refer to Lebesgue
measure of the indicated dimensionality on the indicated set. Thus f o - d!"+,
means integration over D with respect to (N + 1)-dimensional measure and
fati - dl" means integration over aD with respect to N-dimensional "surface"
measure.
If N > I and if the functions ti and v are defined on the closure of D and
are C(2) in the space variables and C(') in space time on D,

JD(a A6 - ti Ati) dl" = f (e v%ti) dl"_, . (6.1)


(s) Jab(s)
270 1 XV. Parabolic Potential Theory Basic Facts

Equivalently,
f
(uAv - vbu)d1N = f
as
)dIN + 2 vDpu)dlN_t. (6.2)
JD:) U(s) rt (s)

Integrating with respect to s yields, if t) is sufficiently smooth,

(tiOri - uAu)d/N+t = tivcosydIN


f6cl 1, r.0
2
+2 vDpu)sinydlN.
dv

If D. is replaced by 1% in the last integral, sin y should be replaced by sin' y.


In particular, if v as 1, this equation reduces to
2
f badlN+l = - f ucosydiN+ Dotisinyd1N. (6.4)
2 IY
The right side of (6.4) can be described as the heat flow of u out of b(tt, t2).
This flow vanishes if ti is parabolic.
Equation (6.4) is valid when N = 1, in which case dl, on the right is the
differential of arc length.
If, is a point of t), the heat flow of qo) out of £(t,, t2), with t, and
t2 chosen so that qo is in this set, is the same as that out of an interval contain-
ing qo and relatively compact in 150,42). The heat flow of qo) out of an
interval containing qo is I by direct computation.

7. The Parabolic Green Function of a Smooth Domain; The


Riesz Decomposition and Parabolic Measure
(Formal Treatment)
Continuing the discussion in Section 6, choose a > 0, s) in 6, and
apply (6.3) with t2 = s and v = O((5, s + a), ) to obtain

f tip(( , s + a), ) cos y dIN = - J s + co, ) Au dIN+t

- a ' [rilSoli(( , s + (z), ) - s + a), )ISoti] sin' y d1N. (7.1)


2
fffi(f " S)

The dot replacing a variable refers to the integration variable, and the normal
derivative is with respect to this variable. Apply Theorem I of Appendix VII
to find when a 0 that the part of the integral on the left over 6(s) has
7. The Parabolic Green Function of a Smooth Domain 271

limit u(); so (7.1) yields

u() = - G( , ) Du d1N+t - cos y d1N


Lfl.s) L(I " a) 7.2)
QZ
ufi G sin2ydl
JD(t,.s)

In particular, if u is parabolic, the first integral on the right vanishes, and


the representation (7.2) then shows that a parabolic function (and therefore
also a coparabolic function) is infinitely differentiable and is analytic in its
space variables for each fixed ordinate value. In view of the corresponding
development in classical potential theory it will be natural (see Section 12)
to define superparabolic functions in such a way that a sufficiently smooth
function a is superparabolic if and only if Au < 0. Under such a definition,
if u is smooth and superparabolic, the representation of u in (7.2) exhibits
ti as the sum of the potential of the positive measure with density -eu
and a parabolic function.
Following the reasoning in the classical context, we next observe (cf.
Section 1.8) that the work leading to (7.2) can be carried through when G
is replaced by a function Go defined on D x D and enjoying the following
properties, stated for b = I)(tt, t2).
(a) For in D the function G(, ) - Ga(, ) is coparabolic on
D; for q in D the function q) is parabolic on D.
(b) has limit 0 at every point of 0,6 with ordinate strictly between
tt and ord and has value 0 at all points of b above . _
(c) The function 4(4, ) can be extended to be of class C()(,6) and of
class C 2)(D) in the space variables.
If the argument leading to (7.2) is carried through with GD instead of G,
(7.2) becomes

u(6 -)au dlN+t + J uOd(4, ) dlN


10,S)

- 26 ti1JfOD(S, ) sine y d1N.


'U(r,.s)

The function Go is unique if it exists because if ift defined on D(t1, t2) is the
difference between two functions with the properties (a)-(c), for fixed the
function is coparabolic on D(tt, t2) with boundary function 0 except
possibly at the points of D(tt). The maximum-minimum theorem for para-
bolic functions as dualized for coparabolic functions and applied to Vi
implies that 4 vanishes identically.
The restriction of to the set of points of D strictly below is
coparabolic, and by the maximum-minimum theorem it follows that
272 I.XV. Parabolic Potential Theory: Basic Facts

0, so that 5 0 in (7.3). When ti = 1 in (7.3), the equa-


tion reduces to
z /

dlN - Z sin' 7 d1N. (7.4)


D(r,) AD(r,,i)

Thus, if ti is parabolic, (7.3) exhibits ti on D(t, , tZ) as a weighted average of


its values on the boundary. This weighting is the analog in the parabolic
context of harmonic measure and will therefore be called parabolic measure.
Observe that the parabolic measure relative to 4 assigns value 0 to the part
of the boundary above 4. This property will be proved in the general case
(Section XVIII.2) when parabolic measure is defined on the boundary of an
arbitrary open subset D of 688.

8. The Green Function of an Interval


Suppose that N = I and that B is the infinite strip ]a, b[ x R, and define
c = b - a. The Green function of h in the present context, like the classical
Green function of a ball, is found by the use of reflections in the boundary,
the method of images. Consider the series

Y G(s - t, 2nc - + q), s), q=(h.t). (8.1)


_ -oD

If <2kc,ifInj >k,andifs>t,
-(2lnl - k)'C' < a(s - 1)1/2
G(s - t.2nc - + q) 5 [2na'(s - t)]-u' exp cznz
2a2(s - t)
(8.2)

Thus after dropping a finite number of summands, the series (8.1) and
similarly the series of partial derivatives of each order with respect to , q,
s, t converge uniformly on bounded subsets of 68 x A. For fixed q in
h each term of the series (8.1) except e (s - t, - + q), the term with n = 0,
defines a parabolic [coparabolic] function of [q] on B. Similarly, after
dropping a finite number of summands, the series

(8.3)
n=-m

and each derived series converge uniformly on bounded subsets of I l x68.


The sum for fixed q [] in B defines a parabolic [coparabolic] function of
[ry'] on B. The function of (, ry') on B x h defined by
9. Parabolic Measure for an Interval 273

W
4(a, [B(s - t, 2nc - + q) - G(s - 1, 2nc + 2a - - q)]
n=-00
(8.4)

is, for S fixed in Al, a function on h which differs from the function
- q) by a function coparabolic on h, and t(a, b) (4, ) has
rl i-+ 41(s - t,
limit 0 at every finite point of the boundary and vanishes when t z s. It is
natural to accept 0(a, b) as GB, the Green function of B. Now choose t, < t2
and define I) = ]a, b[ x ]t, t2 [. Let ¢(a, b, ti, t2) be the restriction of 4(a, b)
to'b x D. For in,6 the function 0(a, b, t1, t2) (4, ) differs from the function
iI '- G(s - t, q - ) by a function coparabolic on 13, has limit 0 at every
lateral boundary point of f), and vanishes when t >- s. Hence we accept
0(a, b, t,, 12) as the Green function GD of t). These definitions of GB and GD
are in agreement with the definition of the Green function of an arbitrary
open subset of 9' to be given in Section XVII.4.
x]aN,bN[x]t,,12[,andifcj=bj-aj,
the Green function GD is defined by

f Y [G(s - t, 2ncj - N + q0)


j=, n=-m
- As - t, 2ncj + 2aj - (j) - q(J))] (8.5)

for and _00-'5t1<12:9+00-


Observe that the Green function GD is the restriction to 1) x 6 of the Green
function of ] a, , b, [ x x ]aN, bN [ x ] t; , t2 [ for - o0 5 t; 5 t, and
12 5 6 5 + 00. An application of the coparabolic function minimum theo-
rem to GD( , ) on 1) - { } shows that GD( , n) is strictly positive when
ord r' < ord and vanishes otherwise.

9. Parabolic Measure for an Interval


The Green function of an interval 6 is given by (8.5). Suppose for the rest
of this section that the interval is finite, that is, that - 00 < t, < t2 < + 00.
Since Section 7 is applicable to a finite interval, the parabolic measure µD
can be written explicitly. In fact, for t in b the measure µD( , ) is supported
by the part of ad strictly below , and this measure is absolutely continuous
relative to IN with finite continuous density GD( , ) on the lower boundary
and finite continuous density
az

for 7 on the part of the lateral boundary with jth coordinate bj and with
density the negative of this derivative for fl on the part of the lateral boundary
274 1. XV. Parabolic Potential Theory: Basic Facts

with jth coordinate a;. This statement is to be understood to mean that on


the !N null set of points common to the lateral and lower boundaries where
the statement gives more than one possible value for the density, either value
can be used. For example, if N = 1 and if b = ]a, b[ x 111, 12 [, c = b - a,
S = (, s). and (q, t), the density is given by

GD( , (q, t, )) = [b(s - t, , 2nc + , - ) - b(s - t 1, 2nc + 2a - - q)],


n=-¢
(9.1)

Y ifq=b,
Q2 S in=-oo
x
2
-Y
S to
00

if q=a.

According to Section 7, if u is a parabolic function on a neighborhood


of D, then u = u) on D. Furthermore, if f is an IN measurable and
integrable function on the union of the lateral and lower boundaries of b,
the function AD(', f) is parabolic on D, by direct differentiation. The function
f) will sometimes be called the parabolic Poisson integral off and will
accordingly sometimes be denoted by PI(D,f ). If b, is an interval obtained
from b by raising the upper boundary, that is, by increasing t2, and if f
is extended to the lateral and lower boundaries of D, by defining f as 0
at the added lateral boundary points, then PI(D, f) = PI(D,, f) on b; so
PI(D, f) has a parabolic extension to D,. It follows that PI(D, f) has a
limit at every inner point of the upper boundary of D. For fixed q on the
lateral or lower boundary of b the parabolic measure density has limit 0
when 4 tends to a point of the lateral or lower boundary of D other than
il. Apply Theorem I of Appendix VII to see that if n is a lateral or lower
boundary point of D, then
llm sup f) 5 lim sup f()

and thereby to see that if j is finite valued and continuous, the function
f) has boundary limit f(n) at every point ,j of the lateral or lower
boundary. The inner points of the upper boundary of ,6 are to be considered
as irregular boundary points for the first boundary value problem for para-
bolic functions. (See Chapter XVIII for a discussion of the first boundary
value problem in the parabolic context.)
The Operation r8. If ti is a Borel measurable function on an open subset
D of AN, if h is an interval with closure in D, and if the restriction of ti to
the lower and lateral boundaries of h is !N integrable, define

PI(B, u) on h
u onD-B
10. Parabolic Averages 275

except that if 1j is an inner point of the upper boundary of A, define

ieu(q) = lim PI(E, u)().


ordf<ordq

If u is upper or lower semicontinuous, the above remarks imply that ijAu


has the same property except possibly at the inner points of the upper
boundary of h and that iii1i is parabolic on h with a finite limit from below
at each inner point of the upper boundary of A.

10. Parabolic Averages


Let : s) be a point of A', and when b > 0, let 6) be the interval

in The interval E(, 6) will play the same role in the study of parabolic
potential theory as the ball B(t, b) in the classical theory. In the following
we take N = 1. The added complications in the general case are merely
notational, and the results will be valid, and applied, for all N. Define the
function f, on the one-dimensional interval [- 1, 1] by

[e(1, 4n + ) - 1(1,4n - 2 - )], (10.1)

and define f2 on [-1, 0] by


I Go

f2(s) _ -S Y (4n + 1)G(-s,4n + 1). (10.2)

According to (9.1), the functions I- f,( /b)/b and s"j2(s/b2)/b2 are


respectively the densities relative to 1, of µ&o.d)(0, ) on the lower and lateral
boundaries of B(0, 6). If ti is a Borel measurable function on a$(rj, 6) for
which the following integrals exist, we define L(u, ry', b) by

L(u, ij, 6) = ua, a)(h, u) = d 62)ff (


f-6 b
(10.3)
o
s
/'(ds)
+ [li(q - b, s) + 07 + b, S)]f2 2) 2
-d II

In order to treat the parabolic analogs of the volume averages in Section


1.2, let 0 be a positive Borel measurable function on R vanishing on
) 1, + oo [, and consider the integral
276 1. XV. Parabolic Potential Theory: Basic Facts

4)(r)L(u,0,r8)l1(dr)_ J u( )f( (10.4)


f.'O

where
2
I ifs < _ 2,
flO
f2(z) ifs> -.
When ¢ = I on [0, 1], we denote the value in (10.4) by A(6, 0, b) and define
A(u, q, 6) = A(ti(, + -),0,6). When ¢(r) = cexp [r z(1 - rz)''] for 0 < r <
1 and ¢(0) = 0(1) = 0 and c is chosen so that Ja4(r)l1(dr) = 1, we denote
the value in (10.4) by A,(ti, 0) and define Aa(ei, q) = A5(ei(q + ), 0). If ti is a
Borel measurable function defined on an open subset D of f(8, the values
A(u, q, 6) and A,(6,'1) are defined whenever B(q, 6) c D, that is, whenever
az + a° < I ry' - aD12, if a is locally 12 integrable on D. Under the latter
condition, 4(6, , b) is continuous, and A,(ti, ) is infmitely differentiable.
Application. It is trivial from the definition that if u is parabolic on D, then

u(O) = L(u, q, S) = 4(u, q, b) = A5(u, q) (10.5)

for b so small that B(q, b) c D. Conversely, if ti is Borel measurable and


locally lz integrable and if u(q) is equal to the third (or fourth) term in (10.5)
whenever B(q, b) c D, then ti is parabolic. In fact then ii is finite valued and
continuous, and if h is an interval with closure in D, the difference ti =
u - PI(B, u) has the same average property as u in h, and a trivial argument
shows that therefore ti satisfies the maximum-minimum parabolic function
Theorem 5. Hence ti = 0 since ti has limit 0 at every lateral or lower boundary
point of B. Finally, if a is supposed Borel measurable on D, locally 1t inte-
grable on lines parallel to a coordinate axis, and either bounded on one side
or locally 1z integrable and also if 6(0) = L(ti, q, 6) whenever B(q, b) c D,
then u is parabolic because the hypotheses imply that u(q) is equal to the
third and fourth terms in (10.5) as well as the first. These criteria for para-
bolicity will be weakened to be local in Section 14.

11. Harnack's Theorems in the Parabolic Context


(a) Convergence Theorem. If u, is an upward-directed family of parabolic func-
tions on b and if the limit function a is finite at a point ry', then u is parabolic on
the open set D0 of points of ,6 strictly below i relative to D, and the convergence
is locally uniform on D0.
12. Superparabolic Functions 277

In fact we can suppose that ti. is an increasing sequence (Theorem 2 of


Appendix VIII). If b) c D, integration to the limit yields the equality
A(ti, , 6); so taking 4 = , , we find that ti is 1N+t integrable on E(il, S).
A covering argument shows that ei is locally IN,., integrable on Do, and there-
fore according to the application in Section 10, the function u is parabolic on
b,. Dini's theorem implies that the convergence is locally uniform on D0.

(b) Inequality Theorem. Let b be an open subset of l'8N, and let A be a measure
on D with minimal closed in D support S. Let A be a compact subset of D for
which to each point of A there corresponds a point in S over the first point rela-
tive to D. Then there is a constant c depending only on D, S, A, ,i such that if
ti is a positive parabolic function on b,

max ti 5 c fs ci dpi.
A

In fact, if there is no such number c, then for some choice of D, S, A, )


there corresponds to each positive integer n a positive parabolic function
iu on D such that maxAu z 1 but J. ti converges
almost everywhere on S, and the set of points of convergence is therefore
dense in S. It follows from the above convergence theorem that Eo u con-
verges uniformly on A; so the sequence u, converges uniformly to 0 on A,
contrary to hypothesis.
Special Case. If A is supported by a singleton Harnack's inequality
states that to each compact subset A of D, all of whose points are strictly
below relative to b, corresponds a constant c, depending only on D, 4,A,
such that if u is a positive parabolic function on D, then
max ti < (11.2)
A

12. Superparabolic Functions


A function ii from an open subset D of A' into ] - oo, + oo] is called super-
parabolic if
(a) ti is lower semicontinuous.
(b) to is finite on a dense subset of D.
(c) ti(p) L (ti, , S) if E( , S) c D.
Just as in the classical context (Section II.4), ti is locally bounded below
and z A(ti, , S) when 6) c D, and it follows that a superparabolic
function is locally IN+t integrable and therefore is finite IN,, almost every-
where on its domain. If ut and u2 are superparabolic functions and if ct and
c2 are positive constants, then ctut + c2u2 is superparabolic.
278 1. X V. Parabolic Potential Theory: Basic Facts

A subparabolic function is defined as the negative of a superparabolic


function, and it follows that a function is parabolic if and only if it is both
superparabolic and subparabolic. A cosuperparabolic function is defined as
a function on an open set ,6 for which the function (4, s) -s) is super-
parabolic on the reflection of D in the abscissa hyperplane, and a cosub-
parabolic function is defined as the negative of a cosuperparabolic function.

Smooth Superparabolic Functions

If su is a C/" function on an open subset of ll' and is C12' in the space variables,
then ti is superparabolic if and only if Ati < 0, in view of (7.3) with D an
interval.

Application of Jensen's Inequality

The application of Jensen's inequality in the classical context (Section 11.9)


is carried through in exactly the same way in the present context. For
example, a convex function of a parabolic function is subparabolic.

EXAMPLE (a). The function ti: s is superparabolic on [f' if


c < 1(a2N), subparabolic if c z 1/(Q2N), and parabolic if c = 1 /(o2N), be-
cause Ati = r2cN - 1.

EXAMPLE (b). If f is a monotone increasing left continuous function on IR and


if u(q, 1) = f(t) for (ii, t) in IAN, the function ti is superparabolic on W. If in
addition lim,_m f(t) = a > - ac, then the function u is the sum of the con-
stant parabolic function a and of a superparabolic potential on AN, the
potential Ou of the product measure p = IN x v, with dv = df.

Convergence of Families of Superparabolic and Parabolic Functions

In view of the parabolic Poisson integral for an interval it is clear that if D


is an open subset of A', if kc- and if ti. is a locally uniformly bounded
family of parabolic functions on D, the family of partial derivatives of these
functions of order Sk is also a locally uniformly bounded family of para-
bolic functions. It follows that a, is locally uniformly equicontinuous.
Furthermore, if a sequence of parabolic functions on D converges locally
uniformly, the corresponding sequence of partial derivatives of any pre-.
scribed order also converges locally uniformly to the corresponding partial
derivative of the limit function. Thus the limit function is also parabolic. An
application of Ascoli's theorem shows that a locally uniformly bounded
sequence of parabolic functions has a locally uniformly convergent sub-
sequence.
13. Superparabolic Function Minimum Theorem 279

Let u, be an upward-directed family of superparabolic functions on D


with limit u. Since u is an upward-directed family of lower semicontinuous
functions, there is (by Theorem 2 of Appendix VIII) an increasing sequence
in the family with limit ti. The function to has the superparabolic average
property and is therefore superparabolic if it is finite on a dense subset of
D; the latter condition is satisfied if every point of D is below, relative to D
a point of finiteness of ti.
Positive Integral Operations on Superparabolic Functions. Such operations
yield superparabolic functions and yield parabolic functions if the given
functions are parabolic. (See Section II.4 for the argument in the classical
context.) For example, if ti is superparabolic on an open set D, the functions
L(ti, , h), A(ii, , a), and Adu are superparabolic on their domains of definition.

13. Superparabolic Function Minimum Theorem


Theorem. Let u be a superparabolic function on an open subset b of if .

(a) If ti attains its infimum at a point of b, then u is identically that in-


fimum on the set of points below relative to D.
(b) The infimum of u is the limit of ti along some sequence of points tending
to aD.
(c) Any lower semicontinuous extension of ti to D attains its infimum on
the boundary.

Assertion (a) is an easy consequence of the fact that ti is lower semicon-


tinuous and that z A(ti, , S) whenever S) c D. Assertions (b) and
(c) follow from (a). The dual version of this theorem for subparabolic func-
tions will be called the subparabolic maximum theorem. Observe that Theorem
13 includes Theorem 5.

Application to Functions on Slabs

A nonempty open subset of A' that is either a half-space bounded by a


horizontal hyperplane or is the intersection of two such half-spaces will be
called a slab. If a slab f) has a lower hyperplane boundary and if a is a super-
parabolic function on D, then the infimum of ti is the limit of ti along some
sequence tending either to the point oo or to a point of the lower hyperplane
boundary. To show this we show that if is the part of b strictly below
and if is the infimum of u on then there is a sequence i, in
tending to the point oo or to a point of the lower boundary of b along which
u tends to According to the superparabolic minimum theorem, there is
a sequence 0. in tending to a point n of ab(4) along which u tends to
m(4). Unless q is on the upper boundary of we are done. If rl is on the
280 1.XV. Parabolic Potential Theory: Basic Facts

upper boundary, then (by the lower semicontinuity of ti) u(q) :9 &;(S), and a
trivial variation of the proof of the superparabolic minimum theorem shows
that u is identically m(S) on D(S). Hence there is a sequence rl, with the
desired properties.

14. The Operation ig and the Defining Average Properties of


Superparabolic Functions

We follow the reasoning of Section 11.6 to derive the corresponding results


in the present context.
(a) If a is superparabolic on D and if B is an interval with closure in D,
then ti is locally IN integrable on the intersection with b of any hyperplane
parallel to a coordinate hyperplane, ijti is parabolic on h, and i8ti 5 6. In
fact, by lower semicontinuity and local lower boundedness of 6 there is an
increasing sequence f. of finite continuous functions of aD with limit u there.
Then PI(B, f) is an increasing sequence of parabolic functions on h with
limit l l(h, «), and an application of the superparabolic minimum theorem
to u-PI(B, j) on h shows that this difference is positive on B. Hence (from
Section 11) ieti is parabolic on h and iBti <_ u. Let S be a point of B. The
foregoing proof shows that ti is IN integrable on any Borel subset of t9B on
which the derivative of µ8(S, ) with respect to IN has a strictly positive in-
timum. Trivial adjustments of and h now show that u is IN integrable on
Oh; in fact it is locally 1N integrable on the intersection with b of any hyper-
plane parallel to a coordinate hyperplane. Since rotations around a vertical
axis preserve parabolicity, a is locally IN integrable on the intersection with
D of any hyperplane parallel to the ordinate axis.
(b) In (a) if i5iu = 6 at a point S of h, then there is equality at the points of
h below S because ih - i8ti is a positive superparabolic function on B.
(c) In the definition of superparabolic function in Section 12, condition
(c) can be replaced by

(c') u(S) z A(ti. S, 6)

or
(c")
u(S) ? Aau(S)

[in both cases for 6 so small that B(S, S) c D].


In fact it is trivial that superparabolic functions satisfy (c') and (c"), and
the proof of the converse follows the corresponding proof in the classical
context [Section 11.6(c)].
We can now proceed precisely as in Section 11.6 (so proofs will be omitted)
to obtain the following results.
15. Superparabolic and Parabolic Functions on a Cylinder 281

(d) In the definition of superparabolic functions in Section 12, condition


(c) [or (c') or (c")] need only be supposed true locally, that is, for sufficiently
small S, depending on .
(e) In (a) the function iiti is superparabolic on D.
(f) If ti is superparabolic on D and if e D, the functions S u-. L(ti, , S),
4, S), S - ASu() are monotone decreasing, with limit u(e) when
6-+0,
It follows that the relation ti Z v or ti = v, if satisfied IN+, almost every-
where on their domain of definition D by superparabolic functions ti and
v, is satisfied everywhere on I). Moreover, if u is superparabolic, then

{
ti(p) = liminfti(q) = liminf
rc
qEB

where 1) is an arbitrary IN+, null set.


(g) If it is a lower semicontinuous function from an open subset D of
ft' into ] - oo, + oo], finite on a dense subset of D, then a is superparabolic
if and only if whenever D. is an open relatively compact subset of D and v
is parabolic on an open neighborhood of Do, with it - v > 0 on MDn, then
ti-vz0on D0.

Approximation of a Superparabolic Function by Infinitely Differentiable


Superparabolic Functions

According to (f), if it is superparabolic on an open set D, this function is


the limit on each open relatively compact subset of D of an increasing
sequence n 1) of infinitely differentiable superparabolic functions.
See the strengthening of this result in Section XVI1.7(e).

EXAMPLE (The Green Function G of l'8N). For fixed q in u1N the function
q) is parabolic on ffBN - {q} and therefore satisfies the parabolic function
average equality there. Since q) is lower semicontinuous on iW and
satisfies the superparabolic function average inequality at q, this function
is superparabolic on AN. Dually, for fixed S in A' the function is
coparabolic on AN - and cosuperparabolic on V.

15. Superparabolic and Parabolic Functions on a Cylinder


Let D be an open nonempty subset of RN(N > 1), and define I) = D x R.
The following results (a) and (b) are rather trivial but key results in the
relations between [super] harmonic and [super] parabolic functions.
(a) Let u bea function on D, and define iu on D by setting s) =
Then ti is [super] parabolic if and only if u is [super] harmonic. If u is of
282 1 XV. Parabolic Potential Theory: Basic Facts

class 02', the assertion is trivial because Au = Au. In the general case we
need only discuss superharmonic functions u and superparabolic functions
u. If u is superharmonic, u is (Section IV.10) the limit of an increasing se-
quence of infinitely differentiable superharmonic functions; so u is the limit
of an increasing sequence of superparabolic functions and therefore is super-
parabolic. Conversely, if ti is superparabolic, let Do be an open relatively
compact subset of D. For sufficiently small b the function .46u is defined
on Do x R, is infinitely differentiable and superparabolic, and depends only
on the space coordinate, 46u(c, s) = u6(t). Hence u6 is superharmonic on
Do, and u = lim6.o u6 is also superharmonic on Do and therefore is super-
harmonic on D.
(b) Suppose that D is connected, let o be a positive superparabolic
function on D, and define v on D by the following positive integral operation
on the family of time translates of u:

V(O = f t)11(dt) = f s + t)11(dt). (15.1)

The function v is lower semicontinuous (Fatou's lemma); so considered as


a function on D, v is superparabolic if finite at points with arbitrarily large
ordinate values, as is true unless v =- + oo. According to (a) above it follows
that v is either identically + oo or is superharmonic on D. In particular, if
t3 is parabolic, the function v is either identically + co or harmonic on D.
For example, we shall show that the Green function 66 of D (parabolic
context) has the form s), (?I, t)) = GD(s - t, , q), where for f ixed rl in
D the function s) i-+LD(s, , q) is a positive parabolic function on D -
{ (ry, 0) ), vanishing if and only if s < 0, and is superparabolic on D; according
to XVII(18.2), integration of GD yields GD, the Green function of D in the
classical context,

eD(t, &, 1)1t NO =

a positive constant.

16. The Appell Transformation


If a is a nonzero constant, the map

q: 01. t)F-+Th = rah


t

takes the upper [lower] half-space of 1N in a one-to-one way onto the lower
[upper] half-space, and 7`1rq = (- ail/t, -a2/t). Define
17. Extensions of a Parabolic Function Defined on a Cylinder 283

_ 2

bo(q) = (2na2IzI)-"f2exp
2121 /
for q in off" less the abscissa hyperplane. If u is a function with domain a
subset B of ft" not meeting the abscissa hyperplane, define Td on TO by
Tu(q) = bo(rj)u(r'i,), so that formally

a2 a
ATti(q) = bo(q)(Au)(T-1q) = T(bti)(q)

on TB. We conclude that if u is parabolic on h, then Ta is parabolic on TB,


and if u is superparabolic on h and in class C(2)(B), then Tti is superparabolic
on TB. A trivial approximation argument then shows that the C'2'(B)
hypothesis is unnecessary. The transformation T is known as the Appell
transformation.

EXAMPLE. If y e 08" and at = -1 /a2, the Appell transformation of the restric-


tion of the function y) to the upper half-space of k" is the
function

of-.(2a) eXp < Y, 7 > +


a2IYI2t

on the lower half-space. This parabolic function [without the factor (2n)-"]
was noted in Section 2 and will be seen in Section XVI.8 to play the same
role in the lower half-space for Poisson-integral-type representations and
minimal parabolic functions that y) plays in the upper half-space.

17. Extensions of a Parabolic Function Defined on a Cylinder


Lemma. Let D be an open nonempty subset of l8" (N,-:! 1), let D = D x ]a, b[
with - oo S a < b < + oo, and let v be a positive parabolic function on D.
Then if a < b' < b and if d' is a positive parabolic function on D' = D x ]a, b'[,
majorized there by v, the function v' has a positive parabolic extension v" to
!, majorized there by v.

Observation. The lemma does not assert that the extension is unique. In the
following proof we shall find the maximum extension.
Let r be the class of subparabolic minorants of v on b which are majorized
on D' by v' and define v" = sup {u: u E r}. If a < b" < b' and if ci is defined
on b by

(v' on D x ]a, b"],


ti 1
0 onDx]b",b[,
284 I. XV. Parabolic Potential Theory: Basic Facts

then t E r. Hence 6" is an extension of v'. The class r contains tit v u2 if


it contains 61 and tie, and r contains the majorant iBti of ti if tier and if
B is an interval relatively compact in D. Thus ti"IB is the limit of an upward-
directed set of parabolic minorants of ti,B; so ti" is parabolic and is the
desired extension of v'.

Application to Minimal Parabolic Functions

Let b and D' be as in the lemma, and let v be a minimal parabolic function
on D. Then 6[6, is minimal on D'. In fact, if v' is a positive parabolic minorant
of ti1D., the extension of u' to D provided by the lemma must be proportional
to d; so t' is proportional to vio..
Special Case: Extension of a Bounded Parabolic Function. Let '6 and D' be as
in the lemma. Then an arbitrary bounded parabolic function v' defined on
D' has a parabolic extension to D with the same infimum and supremum
there as i has on b'. In proving this we can assume that v' has infunum 0.
Let y = supb v'. Then the function v - y on b is a parabolic majorant of
v' on L5'; so v' has a positive parabolic extension to D, majorized there by y.
If D is a finite interval in IB", the preceding result follows easily from our
discussion of the parabolic context Poisson integral.
Chapter XVI

Subparabolic, Superparabolic, and Parabolic


Functions on a Slab

1. The Parabolic Poisson Integral for a Slab


If 6 is the slab 68" x ]0, S[, with 0 < S 5 + co, the restriction to 6 x 6 of
( satisfies the rather vague description of the Green function (;D given in
Section XV.7 for smooth regions. It is therefore to be expected from XV(7.3)
that the upper boundary of 6 if b < + oo is a parabolic measure null set
and that parabolic measure on the lower boundary is given by

uD(4, d7) = o(s, - 7)1N(d7) = (7, 0))IN(d7) [= s)],

so that if ti is parabolic on 6 with boundary function fin some suitable


sense on the lower boundary and if u is appropriately restricted, then

u() = J I(s, - 7)f(7)IN(d7) [= s)]. (1.1)


R"
Such representations will be derived below. Moreover we shall see that if
p is a suitably restricted charge on R", the Poisson integral

}11(i), Jer(s. - 7)lt(d7) [= s)]


RN

defines a parabolic function on A The integral in (1.1) will be denoted by


61(i), f) and the Poisson integral will be modified in the obvious way
when the lower boundary of L) is not the abscissa hyperplane.

Theorem. If f is a Lebesgue measurable function on 6l" and if for some S in


]0, + oo]

I 1, (1.2)
f RN
, IN(d7) < +oo

whenever S' < S, then the function ti defined on 15 = 68" x ]0, b[ by (1.1) is
parabolic. Moreover,
tim sup Jim sup f(7) (C a RN) (1.3)
6 - C-(G. o) RNa r{
286 I XVI. Subparabolic, Superparabolic, and Parabolic Functions on a Slab

and

lira sup ii(c) < lim scup f(ry). (1.4)


D94--ro neRN.l71_+x

On combining (1.3) with the corresponding inequality for inferior limits


we find that u has limitf(C) at (C, 0) if f is continuous at C, and the correspond-
ing specialization of (1.4) is valid. The fact that ti is defined and parabolic on
1) is trivial. Inequality (1.3) is a special case of Theorem I of Appendix VII
because if 0 < 5' < 6 and = s),

lim
(C-0)
"'W'0
uniformly for ry outside an arbitrary open neighborhood A of C in R'; so if
e > 0 and if is sufficiently near (t;, 0),

f.N-A
6(s, - ry)I f(tl)I1N(dd) <- e fRNG(b''ry)1f(ry)I1N(dry). (1.5)

Thus the left side of (1.5) tends to 0 when -+ (C, 0), and the same limit
relation holds when f is replaced by the constant function 1, as required for
the application of Theorem I of Appendix VII. Inequality (1.4) is also a
special case of Theorem I of Appendix VII because

t-N12e-a/t N/2` N1N/2a-N121


5 (t>0,a>0);
2e
J

so if A is an open neighborhood of the point co of t<&N,

s- N11
[(N )N12
I e(s, - tl)I f(ry)IlN(dn) <
RN-A N-A e

Hence the integral on the left and the same integral with f replaced by the
constant function 1 tend to 0 when oo in 1), as required for the applica-
tion of Theorem 1 of Appendix VII.
Extension. According to the theorem as applied to -f if limj,j_.' f(PI)
+oc, then limb, e_. u(p) = +oo also. It will be useful to sharpen this result,
as stated in the following extension of the theorem. If (1.2) is true whenever
b' < b and if, for some tg > 0,

liminf f(i)exp(-flt,I2) > 0,


Inl--

then for 0 < at < /f, 61 < (2ft2)-', and L1 = 68N x ] 0, b1 [ it follows that
2. A Generalized Supcrparabolic Function Inequality 287

lim +oo.

Note that fi 5 (2a26)-' because of (1.2). To prove this extension of


Theorem 1, observe that by hypothesis there is a constant c > 0 such that
f(n) z cexp(/lIgI2) for sufficiently large I4I4 say for Ini -f r. Then

liminf z liminf c f 1(s, - PI)exp(f I i 2 - aI I2)IN(dn)


D, D, 3;-+m
RN
(1.7)

because the part of the integral over {In1 5 r} has limit 0 when ICI - +oo.
The value of this integral is

(1
1 - 2fa s

which tends to + oo when I I - + oo with in D, . The proof of the extension


is complete.

2. A Generalized Superparabolic Function Inequality


Lemma. Let ti be a positive superparabolic function on the slab 1) = RN x
]0, S[, and define f(g) = lim inf.,., ti(C, s) for C in RN. Then ti z 01(D, f).

Let Dk be the slab RN x ] 1/k, S[ fork so large that k > 1/S. Let f. be an
increasing sequence of positive functions on IBN, finite valued and continuous
with compact support and with limit the lower semicontinuous function
ti(-, 1/k). According to Theorem 1, the function P1(Dk, is a positive para-
bolic function on Dk with limit f (j) at each point (i;,1 /k) of the lower
boundary of Dk and limit 0 at the point oo; so ii - P1(Dk, z 0 on Dk by
the minimum theorem for superparabolic functions on a slab (Section
XV.13). The lemma follows when n -+ oc and then k - oo.
Application (a). It follows trivially from Lemma 2 that

s') z f - s, ' - )ti(4, s)lN(dd), 0 < s < s' < S, (2.1)


RN

and therefore ifs' < t < S,

6(t - s', n - ')u(c', s')lN(dd') z f e (t - s, n - c)u( s)lN(dc). (2.2)


fRN JRN
,

Thus the parabolic average relative to (n, t) of the positive superparabolic


function ti on the hyperplane of constant ordinate value s' (<t) defines an
288 I XV1. Subparabolic, Superparabolic, and Parabolic Functions on a Slab

increasing function of s' on ]0, t[. This is an example of the fact that roughly
(see Section VIII. 10 for the classical counterpart) the parabolic average of a
superparabolic function over a set boundary decreases as the set increases.
According to Theorem 5 below, when ti in Lemma 2 is parabolic and positive,
the parabolic average on the left side of (2.2) is constant, equal to ti(,1, t), as
s' varies.
Application (b). If ti is bounded and parabolic on the above slab ,6 and if
u has normal limit f(C) = lim,-o u((, s) at IN almost every lower boundary
point 0), then an application of Lemma 2 to the function u + supb ti and
- ti + sup,, a shows that to = PI(D, f). According to Section 5, the bounded-
ness of ie can be replaced here by positivity if f(C) = limy-,t, o) for all ' in
R". Theorem 6(b) gives necessary and sufficient conditions that a parabolic
function on a slab be representable as the Poisson integral of a function.

3. A Criterion of a Subparabolic Function Supremum


Lemma. If 0 < 6 5 + oo, if a > 0, if ti is subparabolic on the slab D = 68" x
] 0, 6[, and if

ti(p) S 1)], limsupti(n) 5 0 (,s)] (3.1)


j-^(4, o)

for all in b and in R.V, then ti 50 on D.

In fact, if D, = R" x ]0, (8w')-'[, if f(q) = exp[2a(InJ2 + 1)], and if


ti = PI(D, , f ), apply the extension in Section 1 with ji = 2a and fit =
(8atr2)` to find that

lim 1)] = +oo.


3 4-OD

If e > 0, it now follows from the subparabolic maximum theorem applied to


u - eti in D r),6, = R" X ]O, 6 A 61 [ that ti :5 eti on ,6 n I, . Hence 6(4):5 0
when ord < 6 A b, . If 6 5 b, , there is nothing more to prove. If b > b, ,
iteration of the preceding reasoning shows that ti(4) 5 0 when ord 4 < b A
(261),... ; so u 5 0 on D.

4. A Boundary Limit Criterion for the Identically Vanishing


of a Positive Parabolic Function
Lemma. Let u be a positive parabolic function on the slab R" x ]0, 6[, and
suppose that o) ti(() = O for every S in R". Then a = 0.

It can be assumed that (a) ti(c, s) = ti(Rc, s) whenever R is a rotation of


R' about the origin and that (b) the function is an increasing function.
4. Criterion for the Identically Vanishing of a Positive Parabolic Function 289

To see that (a) can be assumed, observe that for any choice of rotation R the
function (g, s) -- s) is parabolic; so if s) is replaced by the average
of u(-, s) over the sphere in R' of radius I I and center the origin, the new
function will be parabolic and positive and have limit 0 at every point of the
abscissa hyperplane. It is sufficient to prove the lemma for this new function.
To see that (b) can be assumed, we show that the function

ti:
Jo

is parabolic on the given slab. This function satisfies (b), and it is sufficient
to prove the lemma for this function, which satisfies (a) if ei does. If a > 0,
define t by

s) = Ja(.r)1i(dr). 0<a<s<6.

Then &.(i, s) = so v is the locally uniform limit of the increasing


sequence (vt1,,, n >- 1 If, with lim, but, = 0 locally uniformly. It follows
that ri is parabolic [for example, apply XV(7.2) to ut, and go to the limit,
n oo].
Thus we now have f(Ii1,s) with monotone increasing for
every value of r z 0. In view of this monotoneity and of the parabolic func-
tion maximum-minimum theorem the maximum of ti on the cylinder (0 <
ord . < t, ICI 5 r) must be attained on the top of the lateral boundary, that
is, at a point with r, ord = t, and it follows that s) is an increasing
function for each value of s > 0. According to Lemma 2,
r2
_2a2(s

t)]-N127rN
[2aa2(s - r"-t f(r, t)exp
fo""
- t)!t(dr) Sf(O,s), (4.1)
0<t<s<6,
and therefore if r > 1,
fW
r
[21ra2(s - t)]-N'2nNf(r, t) exp a2 11(da) :9 f(0, s), (4.2)
2az(s __0

so that ifs is fixed,

t) = f(r, 1) 5 const (1 + r) exp 2es,


r2

or
r = 141, t 5 (1 - s)s. (4.3)

It now follows from Lemma 3 that u 5 0 on the slab R" x ]0, (1 - e)s[; so
ii = 0 on this slab and so on R" x ]0, S[.
290 I XVI. Subparabolic, Supcrparabolic, and Parabolic Functions on a Slab

5. A Condition that a Positive Parabolic Function Be


Representable by a Poisson Integral
Theorem. If 0 < b < + co, if ti is a positive parabolic function on 1) = 08x x
D,6[, and if lima . o) u( ) = f(r;) < + oo exists for all t' in Rx, then u =
I(Af).
The boundary limit function f is necessarily continuous since there is a
limit at every point of the abscissa hyperplane. Furthermore ti - P1(/5, f) 0
by Lemma 2, and the difference has limit 0 at every point of the abscissa
hyperplane according to Theorem I. Hence (Lemma 4) the difference
vanishes identically.

6. The L' (µs_) and D(/E_) Classes of Parabolic Functions on


a Slab
Theorem 11. 14 for harmonic functions on a ball has the following analog for
parabolic functions on a slab. Recall that if p is a charge with minimal Jordan
decomposition jt' - Y. we denote the absolute variation measure µ' + µ-
by Isl. In the following theorem if 0 < 6 5 +co, we denote by 6(0, 6) the
slab R' x ]0, b[.
Theorem. Let ti he a parabolic function on the slab h = B(O, 6).
(a) L' (µa_) parabolic functions. The following conditions on u are
equivalent:
(al) ti = PI(B, for some charge N,; on R'for which
_ 2
0 < s < b. (6.1)
2a 1 N,;I (dtl) < +oo,
JRNSexp 11

(a2) ti is the difference between two positive parabolic functions.


(a3) al has a parabolic majorant.
(a4) supo<,<, ± G(s - t. S - ry)la(ry, t)llx(d,t) < + oo for every
point (, s) of B.
The map a'-- N is a one-to-one linear order-preserving map from the
class of parabolic functions satisfying these conditions onto the vector lattice
of charges on R' satisfying (6.1).
(b) D(µa_) parabolic functions. The following more restrictive conditions
on a are equivalent :
(b1) a = PI(B, fjfor some 'N measurable functionh on R' satisfying

II,;(q)llx(17) < + cc, 0 < s < 6. (6.2)


fR^exp 2a Z
s
6. The L'(p,_) and D(p,-) Classes of Parabolic Functions on a Slab 291

(b2) If 0 < s < b, there is a uniform integrability test function Of or


which the restriction to B(0,s) of D,(ju1) has a parabolic
majorant.
(b3) For every point s) in h the family

{t r [u(ry, t), G(s - t, - q)l"(dry)], 0 < t < s)

ofpaired functions and measures on R" is uniformly integrable.


(b4) If u > 0, ti is the limit of an increasing sequence of bounded
positive parabolic functions.
The map u "f, is a one-to-one linear order-preserving map from the
class of parabolic functions satisfying these conditions onto the vector
lattice L' (R", l"), and dRj = f dl".

As will be noted in Section XVIII.19, the notation introduced in Chapter


IX for various classes of functions in classical potential theory is readily
adapted to the context of parabolic potential theory. These classes together
with their martingale theory counterparts are discussed in Chapter I of Part
3. In this spirit the parabolic functions in Theorem 6(a) are the L'(µ6_)
parabolic functions, those in Theorem 6(b) are the D(pA_) parabolic func-
tions, and Theorem 6(b4) asserts that the positive D(µii_) parabolic functions
are quasi bounded. A uniform notation will be used in Chapter I of Part 3
both for classes of functions in the classical and parabolic potential theory
contexts and for classes of stochastic processes in the martingale theory con-
text. The assertion of Theorem 6(b4) in all three contexts is covered by
Theorem 3.1.5.
To prove Theorem 6, define PI(B(s, t), f) on $(s, t) = R x ]s, t[ by

PI(A(s' t), f)( ', s') = 6(- t, ' -


JS'
R"

Observe that Itil in (a3) and m,(Icil) in (b2) are positive subparabolic functions
with parabolic majorants. It follows that for fixed s > 0 and fixed s') in
h(0, s) the parabolic averages

PI(B(s, t), s'), PI(B(s, t), S)

define decreasing functions of t on the interval ]0,s'[. To see this, for


example, for the first parabolic average, note that if v is a parabolic majorant
of Itil on h, then (by Theorem 5) PI(E(s, t), t)) = ti on E(s, t); so (from
Section 2) the function

is an increasing function on ]0,s'[. With the help of this monotoneity result


the proof of Theorem 6 becomes so close to that of Theorem 11.14 in its ideas
292 1. XVI. Subparabolic. Superparabolic, and Parabolic Functions on a Slab

that the details will be omitted. Choose 6' with 0 < d' < 6, and for 0 < r <
define charges t, and j on R' by
z
p,(dq) = ti(q, r)IN(d)l), ft:(dtl) = u(q, r) exp 1

2a2q r lN(dq)

It is convenient to consider j as a measure on the space I' compactified by


a point at infinity, with the infinite singleton ft; null. Instead of showing
directly that µ, tends to a limit measure N,; when r -+ 0, it is easier to show
that under (a3) the total variation of y is bounded independently of r, that
(vague)lim,.oA; = µ' exists, and that (al) is true with
12

tI (dq) = exp 2116, N (dq)

Finally observe that the supremum in (a4) defines a parabolic majorant of


Iul on B. Further details of the proof of Theorem 6 are left to the reader.

7. The Parabolic Boundary Limit Theorem


A function f on a slab I8N x ] 0, 6[ is said to have normal limit q at the
boundary point C : (C, 0) if lim,._o f(C, s) = q. The function is said to have
parabolic limit q at (C, 0) if limf_S q whenever : s) - C in a para-
boloid of revolution with vertex C and opening upward, that is, whenever
C -+C with liminff.;sI C - CL-z
> 0. In more sophisticated language, if a
subset A of the upper half-space is called a deleted coparabolic neighborhood
of C whenever for some a > 0 the set A contains the intersection of the
paraboloid {s > ajC - C12} with some Euclidean neighborhood of r, then f
has parabolic limit q at if f has limit q along the filter of deleted coparabolic
neighborhoods.

Theorem. Let B = IAN x ] 0, b[ be a slab in OWN, let 6 be a parabolic function


in L (fiB_ ), and let h be a strictly positive parabolic function on B. Then if
C is a point of IBN at which the convex variational derivates dlN/d?4 and dIV,;ldN,
both exist, the function ti/4 has parabolic limit (dN,;/dN1,)(C) at (C, 0).

Observe that the stated conditions are satisfied at Nh almost every point
of IBN. In the more common version of this theorem h - 1; so Nh = IN.
In particular, if u e D(Aj_) so that ti = 01(h, f) for some function f on R', the
theorem states that ti has parabolic limit f(C) at IN almost every slab boundary
point (C, 0).
Theorem 3 of Appendix VII can be applied to prove a modified version
of Theorem 7, namely, that under the hypothesis that dlN/dfh and d!1 /dI'4
8. Minimal Parabolic Functions on a Slab 293

both exist as symmetric derivates at , the function ti/h has normal limit
(dN,Jd!)(S) at (C,0). The proof of this modification follows the proof of
the corresponding result (Theorem I1.15) in the harmonic function context.
In the latter context the Harnack inequality made it possible to go from
normal approach to nontangential approach, but this step does not seem
possible for general h in the parabolic context. We therefore prove Theorem
7 as an application of Theorem 4 of Appendix VII. In the latter theorem if
we set

2
K(s, r) = exp Za2.s, 6(s) = const s12, so(a) = 2,

we obtain Theorem 7.
Application. We have proved in Section 2 that if ti is a bounded parabolic
function on the slab h and if ti has normal limit J ) at IN almost every point
(C, 0), then u = PI(B, f ). According to Theorems 6 and 7, the weaker hy-
pothesis that the parabolic function 6 on h is in D(pa_) implies that the
parabolic limit, say J(o, exists at IN almost every boundary point (C, 0) and
that 6 = l'I(B, f ).

8. Minimal Parabolic Functions on a Slab


(a) D = RN x ]0, 6[, 0 < b 5 + co. In view of the Riesz-Herglotz-type
representation (Theorem 6) of a positive parabolic function on b, such a
function is minimal if and only if it is a positive multiple of the function
( , s) i--. e (s, - C) on b for some point of R'.
(b) D = I8N x ] - oo, 0[, the lower half-space. In view of the Appell
transformation which takes the upper half-space of l' into the lower
half-space, a parabolic function ti on the lower half-space is minimal if
and only if it is the Appell transform of a minimal parabolic function for
the upper half-space, that is (Section XV.16), if and only if ti is a positive
multiple of the function

(.s)i - exp L(7, + aZ1yJ2S1 (8.1)

on the lower half-space, for some point y of Q$N. In particular (y = 0), the
positive constant functions are minimal on the lower half-space. It follows
(Liouville-type theorem) that a bounded parabolic function on the lower
half-space, and therefore surely a bounded parabolic function on ft', is a
constant function. Furthermore the representation Theorem 6 transferred
to the lower half-space by an Appell transformation yields a representation
for a positive parabolic function u on the lower half-space:
294 1.XVI. Subparabolic, Superparabolic, and Parabolic Functions on a Slab

+ [<p. > + a2IYI2] N(dy), (8.2)

where f is a measure on R' for which

exp(-ajyI2)Na(dy) < +oo (8.3)


fR'
for all a > 0. We conclude that every positive parabolic function on the
lower half-space is either strictly positive or identically 0 and is monotone
increasing in the ordinate variable. We leave to the reader the full formula-
tions of Theorems 6 and 7 in the lower half-space context.
(c) 6 =AN. It follows easily from (b) that the minimal positive para-
bolic functions on A' are the positive multiples of the functions (8.1), now
considered on ft', and that every positive parabolic function a on IAN has
a representation of the form (8.2) with (8.3) true for all real a. Every positive
parabolic function on l'' is either strictly positive or identically 0 and is
monotone increasing in the ordinate variable.
Chapter XVII

Parabolic Potential Theory (Continued)

1. Greatest Minorants and Least Majorants


If D is a nonempty open subset of off" and if r is a class of functions on D,
the greatest subparabolic minorant [least superparabolic majorant] of r,
if there is one, is denoted by GMDT [LMT]. For example, if r is a class
of superparabolic functions and if r has a subparabolic minorant then
OM j)I- exists and is parabolic. The proof is a translation of that of Theorem
111.2. The corresponding notation in the coparabolic context is OMDr and
LMDr.

EXAMPLE. Let b be either 119 (N z 1) or an interval in AN. Then


0 for every point in D. In fact, say for
when b is an interval, the parabolic minorant in question is positive, is
majorized by and so has limit 0 at every lateral and lower boundary
point of D. This minorant therefore vanishes identically, according to the
parabolic function maximum-minimum theorem. More generally it will
follow from the Riesz decomposition of a positive superparabolic function
on a nonempty open subset b of fr that the parabolic potential of a measure
on b if finite on a dense subset of D is superparabolic on l) and has greatest
subparabolic minorant 0.

2. The Parabolic Fundamental Convergence Theorem


(Preliminary Version) and the Reduction Operation
The proof of the following counterpart of the first version of the Fundamental
Convergence Theorem (Theorem 111.3) in the classical context follows the
proof of Theorem 111.3 and is therefore omitted.

Theorem. Let T : { tiQ, a e l} be a family of superparabolic functions on an


open subset of d8", locally uniformly bounded below, and define
u 5 u,
lim infii(, ), (2.1)
rt
296 I XVII Parabolic Potential Theory (Continued)

and

(a) ti is superparabolic.
(b) ri on each open set on which ti is superparabolic.
(c) u IN,, almost everywhere.
(d) There is a countable subfamily of F whose infrmum has smoothing j.

Application : The Natural Order Decomposition Theorem

As application of this simple version of the Fundamental Convergence


Theorem in the parabolic context we remark that the classical context
Natural Order Decomposition Theorem (Theorem 111.7) translates directly
into the parabolic context: If ri, ut, u2 are positive superparabolic functions
on D with u < iut + u2, then there are positive superparabolic functions
5 u2, u = u; + r '. The classical context
ti; , ti4 on D for which ti; S iu t , t
proof requires only trivial changes. Observe that this decomposition and
its proof are also valid for relative superharmonic and superparabolic
functions. Alternatively the decomposition theorem for superharmonic and
superparabolic functions implies trivially the decomposition theorem in
the relative contexts.

3. The Parabolic Context Reduction Operations


If D is a nonempty open subset of f1N coupled with a boundary aD provided
by a metric compactification, if A c D u aD, and if ti is a positive super-
parabolic [cosuperparabolic] function on b, the superparabolic [cosuper-
parabolic] reduction of ti on .4, denoted by R" [R" ], is the infimum of the
class of positive superparabolic [cosuperparabolic] functions on b which
majorize ri on A n b and near A n 3D, in the sense that each function in
the class is to majorize o both on A n D and on the trace on D of some
neighborhood of ,4 n D. The smoothed reduction +v R" ["]
+
is superparabolic
[cosuperparabolic] according to Theorem 2 and will sometimes be denoted
by ti0" [16 -4]. As in the classical context, it is trivial that A6.4 is the inrimum
of the class of positive superparabolic functions on b which are equal to
ti on A n,6 and near A n D. Thus R" = ti on A r),6, and obviously R" S
on b.
Let C be a point of D, let D4 be the set of points of b strictly below
relative to A let ti be a positive superparabolic function on D, and let vS
be the restriction of ti to Db. Then if .4 c D, we now prove that Rv (reduction
relative to D) is equal on Di to the reduction relative to Di of tii on A n bt.
(This fact implies the truth of the corresponding statement for smoothed
reductions.) The point is that roughly the reduction of ti below a point
depends only on t below that point and on the part of the target set below
that point. To prove the assertion, we need only remark that on the one
3. The Parabolic Context Reduction Operations 297

hand if ti is a positive superparabolic function on D which majorizes ti on


A, then the restriction of ti to Dt majorizes t64 on A r),64 and on the other
hand if u' is a positive superparabolic function on D{ which majorizes ti{
on .4 n Dc and if ti is a positive superparabolic function on ,6 which majorizes
v on A, then the function ti" equal to ti on b - D4 and to u n ti' on b4 is a
positive superparabolic function on D majorizing ti on A with ti" 5 ti on
band ti'"5a'on Ds.
The fact that a smoothed reduction R' in the classical context is equal
quasi everywhere on A n D to v and that this smoothed reduction is un-
changed when A n D is changed by a polar set is considerably weakened
in the parabolic context. In fact it will be shown that a smoothed reduction
A" is in general equal to to on A only up to a parabolic-semipolar set (to be
defined in Section 10) and may be changed if A is changed by a parabolic-
semipolar set. This weakening entails that some of the proofs of properties
in the classical context cannot be used to prove the corresponding properties
in the parabolic context and that it is necessary to change the order of the
derivation of the properties common to the two contexts.
Just as in the classical context (Section III.4), and k" increase when
A or v increases, the reduction and smoothed reduction operations are
subadditive, the function A-4 is parabolic on b - A and equal to As there,
and A- = when A is open. Furthermore (see Section 111.5)

R" = inf{A" ,,"B: Ana D c A, J) open in DuaD}


(3.1)
= inf A n 0,6 c A, h open in D u aD}

and

t'`=inf{J 4"B:AnabcE,J3openinbuaJ}
(3 . 1 sm)
= inf { A BjnD : A n aD c $, f3 open in b u aD}.

The counterparts of the other properties listed in Section 111.5 will be listed
below in Section 16.

EXAMPLE (a). Let .4 be the open upper half-space of A', let A0 be an arbitrary
subset of the abscissa hyperplane, and let v be a positive superparabolic
function on ft. The parabolic reduction (relative to A") and smoothed
reduction of ti on A u An are trivially v on A u Ao and 0 on the lower half-
space. Hence (by lower semicontinuity of superparabolic functions) the
smoothed reduction is ti on the upper half-space and 0 otherwise. The
reduction is 0 on off" - (A u,40) because according to Section 16(f) the
reduction and smoothed reduction are identical off the reduction set.

EXAMPLE (b). If A is a horizontal plane in 1'8", any positive superparabolic


function 6 on A' with ii z I on 4 satisfies the same inequality above 4
298 I.XVII Parabolic Potential Theory (Continued)

(Lemma XVI.2), and it follows that, for reductions relative to A , D 10A is


equal to I strictly above A and equal to 0 elsewhere. Thus I 1 D'` J A = 0, and
so, unlike the situation in the classical context as given in Section VI.3(h),
the smoothed parabolic reduction operation is not always idempotent.
[However, according to Section 16(i) this operation is idempotent if A n b
is parabolic-fine open.]

4. The Parabolic Green Function


Let b be a nonempty open subset of A'(N Z 1), and let be points of f).
The parabolic [coparabolic] Green function with pole n [] is defined on
b by

[G1( ,')-C M,O(


(4.1)

It will be shown in this section that, corresponding to the symmetry of the


Green function in the classical context, OD = Go. Thus the notation Gd is
unnecessary and will not be used in later sections. The function (D will be
called the parabolic Green function.
As defined by (4.1), the function tj) is positive and superparabolic
on 15, is parabolic on 15 - {i }, and differs from tj) by a continuous
function, and p) = 0. Conversely, these conditions uniquely deter-
mine ri). The corresponding dual remarks for Go( , ) are omitted. The
definition of the parabolic Green function OD is the counterpart of the
definition of the classical Green function GD, but no side condition on the
domain, depending on the dimensionality, is necessary in the present context
because 6 is positive for N z 1. Note that (Section 1) the present definition
of Og agrees with that given in Sections XV.4, XV.8, and XVI. I when 15 is
AN, an interval or a slab. In particular, OD = 6 when 6 = AN. The properties
assinged to O in the smooth region context discussed in Section XV.7
make there the coparabolic Green function with pole , by an easy
application of the coparabolic maximum theorem to (notation of
Section XV.7). This is as it should be because, as we shall now show, lip =
GD.
Proof that t o = GD. Define 6(-,4) = rj), and let h, be a sequence
of intervals with closures in 15 and with the property that each point of 15
has a neighborhood which lies in h. for infinitely many values of n. Define

7) = t8.... t)ib

to obtain a decreasing sequence ti,(-, ri) of superparabolic functions on I)


with limit ti(-, q) (cf. the corresponding discussion for the classical context
in Section VII.4). Let 'm be a point of 15. The sequence h, can be chosen in
4. The Parabolic Green Function 299

such a way that 4o e $o and that there is a neighborhood of 00 which is either


a subset of h, or of 6 - 1) for each n. For in 6 - {qo} the function O( , )
is coparabolic on a neighborhood of 0, so the functions 60(x, ), ... are also.
In fact each of these functions after the first is an integral average of its
predecessor, averaged over values of the first argument. Thus is
coparabolic on I) and therefore u(4, ) 5 GMDO(, ). Define as the
right side of this inequality so that u 5 L. If we had begun with u instead of
ri and carried through the dual argument, we would have obtained the
reverse inequality, and it follows that u = u; that is,

4)() = GMrid(4, )(4), OD = GD (4.2)

The counterpart of the proof in Section VII .4 that u is continuous on


D x D proves that ti is continuous on 6 x 1). We shall use the fact (cf.
Theorem VII.3) that 6D(4, ) and 4) are bounded outside neighborhoods
of their poles.
We leave to the reader the easy translation of the extremal properties of
Green functions in the classical context (Theorem VII.2) into the parabolic
context.

Relativization of the Green Function

The reasoning used in the discussion of the Green function can be relativized
with no change in detail, just as in the classical context (Section VII.1) to
use an arbitrary nonempty open subset of A' as a reference set instead of
A". That is, if 6 and a are nonempty open subsets of dl" with I) c f, then

4) = 4) - 4), OD( ,)_


(4.1')
on 6 and

GM 4)(4) = GMb -)w- (4.2')

EXAMPLE. Suppose that D is a half-space of RN, define 1) = D x R, and if


ry' a El, denote by 4' the reflection of 4 in ab. Then

4) = 4) - Q, h) (4.3)

because an application of the superparabolic minimum theorem shows that


4') = on El. More generally, if j = 1, . . . , N successively and
if for each j the set Dj is either R or a half-space of R, define
N

ri
i=1
D) x R, 6, = Df x R.
300 1. XVII. Parabolic Potential Theory (Continued)

Then the function

(4, 0 = s), (h'u, .. , nlN1, t)) fin( , q)

is the product of the Green functions for N = I of Dt , ... , D" written in the
respective variables ( ( o h ), s), (,7(l), t)), ... , s), (ry1"1, t))

Strict Positivity Set of (;o

n) > 0 if and only if q is in the set bi of points of b strictly below


relative to D. To see this, observe first that the inequality 0 S OD 5 6 implies
that C;o(, ry') = 0 if ord ry' ord . Next observe that (coparabolic maximum-
minimum theorem) if il) = 0 and if q' is strictly above p relative to D,
then fj') = 0. Since 0 at points arbitrarily close to and
strictly below g, it follows that q) > 0 whenever n e Dt. If now we fix
and define u(r) = Gp( , q) when' a AA and define ti(q) = 0 when , e b -
A. then ti is cosuperparabolic on D and 66(4, ) - u is a positive coparabolic
function on D. It follows that this difference vanishes identically; so Dr is
the strict positivity set of ), as asserted.

Extensions and Contractions of Green Functions

If is a point of b, then 6os is the restriction of 6j5 to Dz x Di. In fact this


restriction has the required properties except perhaps the property that
0 when e Dt (or the equivalent dual property). Now if h is
a positive coparabolic function on D{ and is a minorant of v(, ), then
h = 0 at the points of D; with ordinate values zord 4; so if h is extended
to D by defining h = 0 on D - D{, the extended function is a positive copara-
bolic minorant of C;D( , ) on D and so vanishes identically. Hence h =- 0 on
b, so 6m 0, as required. A similar dual argument shows that the
Green function of the set to of points of D with ordinate values strictly
greater than ord is the restriction of Gp to Da x to.
An observation in the reverse direction will be useful. Suppose that is a
finite Euclidean boundary point of b, that every point of D is strictly below
and that the part of some open neighborhood a of strictly below is in
D. Then for, in b the function O q) has the extension 66,j(-, ry'), parabolic
onDvh-{n}.

5. Potentials
If ,6 is a nonempty open subset of dA' and if p is a measure on D, the functions

n)li(dh), A6,6 =
Jd Jd
5. Potentials 301

are, respectively, the (Green) potential and copotential of A. Since the


properties of copotentials follow trivially from those of potentials, we shall
consider only the latter unless the interplay between the two is involved. The
potential 06p is lower semicontinuous (Fatou's lemma) and has the super-
parabolic function average property (Fubini's theorem) so C,p is super-
parabolic on D if to each point n of b corresponds a point of finiteness of
(p above n relative to D. If superparabolic, the potential O,,p is parabolic
off the closed support of A. If Fi(D) < + oo, the counterpart of an argument
in Section IV. I shows that 6,6A is superparabolic on D. In particular suppose
that p has compact support A in D, and let a be a neighborhood of A. Since
t;D is bounded on the set (D - E) x A it follows that the superparabolic
potential (`. p is bounded on 15 - $ and has limit 0 at every boundary point
r; of D for which limi_i OD(i,,) = 0 when , e A.
If 6,611 is a superparabolic potential then 6Md(`iDp = 0 by the counter-
part of the proof of the corresponding classical context result (Section IV.3).

EXAMPLE (a). Let u be the indicator function of the upper half-space A of


f". Then u is a superparabolic function on D = Ift" and is the potential of
the measure IN on the abscissa hyperplane. This potential is discontinuous
and vanishes on the support of its measure. Thus the domination principle
(Theorem V.10) is false in the parabolic context. A trivial variation of this
example in which the measure has compact support shows that the Evans-
Vasilesco theorem (Section V.8) is also false in the parabolic context.
Versions of the domination principle adapted to the parabolic context will
be proved in Section XVIII.16.

EXAMPLE (b). Define L5 = R" x ]0, b[ and D' = R" x ] - oo, S[ with 0 < 6
S + co. Recall (Section 4) that ( [OD.] is the restriction of G to b x D
[D' x b']. Let 6 be a positive superparabolic function on D with associated
Riesz measure i and extend o to v' on D' by setting v' = 0 on D' - D. Then v'
is superparabolic and v is the projection on b of the Riesz measure v' associ-
ated with v'. For example, if 6 = + co and v =- 1, the function v' is the potential
ti on ft' discussed in Example (a). Now suppose that v is a potential, v =
(GDv = G`D.v on D. Since G,y.v = 0 on D' -,6, it follows that v' = OD.v, and
this representation shows that D' - D is v' null. On the other hand, if v
is parabolic there is (by Theorem XVI.6) a measure Na on the abscissa
hyperplane such that v' = Thus in this case NN is the projection of V
on the abscissa hyperplane. We have now proved that whatever the choice
of v on D, the extension v' to D' is a potential. One way of phrasing the fact
(Theorem XVI.6) that a positive parabolic function on the slab b is given by
a Poisson-Stieltjes integral is to state that every such function when extended
by 0 to 15' becomes a potential. This fact suggests that one way of deriving
the Poisson-Stieltjes representation is to prove directly that the extended
function v' is a potential; it is trivial that the corresponding Riesz measure
must be supported by the abscissa hyperplane.
302 I.XVII. Parabolic Potential Theory (Continued)

ExAMPL (c). Consider the following potential (p in ft', for which p is


supported by the surface {(n, t): t = -In14, Inl < 1),

fj-,0111<1) exp21 (s+ Ini4) Clnlx+112 exp2a21In12J 1. (d,7)

u(,s) _ (S + I'iI4)Nj2 (5.1)

ifs> -1,
0
if s5 -1.
Obviously s) = + oo when s > 0; so ti is not superparabolic. On the
other hand, if fi is any constant the function # A v is superparabolic. The
evaluation

(5.2)
2oZIn14 /f
f11111<11

shows that is a finite-valued function of Isl. We now show that


lims-o 0) = + oo, for which it is sufficient to show that 0) has limit
+ oo along a ray in R" to the origin. Let g be on this ray from now on, and
let 0 be the angle between the rays from the origin to and n. Then for
ICI < 1,

z JIltl<hl<ZISU (exp0(10 Inl-N+uz/x(dn), 312 z 1 (5.3)


Inl 2Q
`
{lWxR 6)

Since the integrand for fixed ICI is a function of Inl, there is a constant c
such that
zlt1r-"2expal41z11(dr)
z c2-'lzl 1/2exp
ti( ,0) z c f141 (5.4)
16 l2

a nd therefore 6(-, 0) has limit + oo at the origin, as stated.

Thinness

Recall that a superparabolic function has limit inferior at a point equal to


its value at that point. In Example (a) the value ti(0) is strictly less than the
limit of ti at 0 along the upper half-space. In this sense the upper half-space
of AN is thin at the origin. The abscissa hyperplane is thin at the origin in
this same sense because in Example (c) the superparabolic function # A ii
with jJ > ti(0, 0) is strictly less at the origin than its limit along the abscissa
6. The Smoothness of Potentials 303

hyperplane. This concept of thinness, the counterpart of the classical concept,


will be given a topological interpretation (the parabolic-fine topology) in
Section 9.

Dependence of dop on b

(Cf. Sections IV.1 and VII.5.) If b, c t3Z, the difference dD= - 66, on
15, x 6, is a positive function, parabolic in the first argument and copara-
bolic in the second. If p is a measure on b, whose potential (,D=p is super-
parabolic, then the difference GD= p - rro, p if defined suitably at the common
infinities of the two potentials is positive and parabolic on 1, .

6. The Smoothness of Potentials


The following theorem is the parabolic context analog of Theorem 1.7. In
view of the dependence of lop on 6 described at the end of Section 5 the
smoothness conclusions of the theorem are applicable in their obvious
adaptations to potentials 66p as well as Op.

Theorem. Suppose that dp = f dl"+, , where f is 1"+, measurable and is sup-


ported by a slab R' x [a, + oo [ with a > - oo.
(a) If f is bounded, the potential ti = Gp is finite and continuous on
and has continuous first partial derivatives with respect to the space
variables, given by formal derivation of the integral defining ii.
(b) If in (a) f is continuous and satisfies a uniform Lipschitz condition of
exponent fi, 0 < fi 5 1, in the space variables,

I f(n, s) s) I s const In - I0, (6.1)

then 6 has continuous second partial derivatives in the space variables and a
continuous first partial derivative in the ordinate variable, given by
021i(S)
= 026(S,h
0 [An, t) -f( , t)]1"+, (dn)

s) (6.2)

(6.3)
aas )= f',N as

Hence dti = -f
304 1.XV11. Parabolic Potential Theory (Continued)

Proof of (a). Since (n, t))lN(dtl) 5 1, the potential u is bounded under


the hypotheses of (a). To prove continuity, observe that if lib is defined by

ub( ) = f dt f (n, (6.4)

the function lib is continuous (dominated convergence theorem), and

u - bl -
ti al < III dt=6su PIII; (6.5)

so limb_a lib = ii uniformly on ftN, and ti is continuous. Formal differentiation


of ti yields

d.
ati( ) =
__
nU1

I
aZ S-t
fON
(6.6)

The last integral in (6.6) converges absolutely because if t < s,


(it,)112
i n'"I (s - t)-`/N(dn) = Q(s - t)-'n (6.7)
RN

The last integral in (6.6) defines a continuous function of r; because if the


slab Rx x [s - b, s[ is excluded from the domain of integration, continuity
becomes trivial, and the error in excluding this slab is [by (6.7)] at most
o26 1l2 sup I f 1. Hence (6.6) is true. o

Proof of (b). Under the hypotheses of (b) the integral on the right in (6.2) is
absolutely convergent, in view of the majorant of the integrand provided by
XV(4.4) with k = 2. Moreover the integral on the right in (6.2) defines a
continuous function of by an argument following that used in proving
continuity of the two integrals in the proof of (a). Since the function n p-+
n')/a is odd about , the first equality in (6.6) can be written in the
form

au(S) = (6.8)
a"10 a"(4 (dn).
S J R.N SS

If the integral on the right in (6.2) is integrated in ''' over an interval but
evaluated by first integrating inn, then "', then t, the result is the difference
between the values of at the endpoints of the interval. It follows
that (6.2) is correct. The absolute convergence of the integral in (6.3) and
the continuity of the function of thereby defined are proved just as the
8. Parabolic-Polar Sets 305

corresponding assertions for the other integrals were proved. Moreover

a( )= J G( s (6.9)
q" co

and differentiation in (6.9) yields (6.3). 0

7. Riesz Decomposition Theorem


Theorem. If 15 is an arbitrary nonempty open subset of 9" and if ii is a super-
parabolic function on .6, there is a unique measure p on I) with the following
properties.
(a) If A is an open nonempty relatively compact subset of I) and if µ8 is
the projection of p on there is a superparabolic function h8 on 6,
parabolic on h, with ci = (pe + h8.
(b) If in addition a has a subparabolic minorant, then u has the representa-
tion u = Orp + (`iMpci.

The proof is a translation into the present context of the corresponding


classical Riesz theorem for superharmonic functions (Theorems IV.7 and
IV.8) and will therefore be omitted.
The Riesz decomposition leads at once to the following facts for a positive
superparabolic function a on D (cf. their counterparts for the classical
context in Sections IV.8 and IV.10).
(a) The function ii is a potential (;Dp if and only if OMDa = 0.
(b) If a is majorized by a superparabolic potential, then a is itself a
potential.
(c) Special case of (b). If a is a superparabolic potential, As is also, for
every A; in particular, J = 0 for every choice of 8L).
(d) The smoothed reduction A is a potential if A is a relatively compact
subset of A.
(e) The function a is the limit of an increasing sequence of bounded
infinitely differentiable potentials of measures with compact
supports.
The formulation of the full counterpart of Theorem IV. 10 is left to the
reader; this counterpart is a slight extension of (e).

8. Parabolic-Polar Sets
A parabolic-polar subset of dt" (N z 1) is defined as a subset A satisfying the
following equivalent conditions:
306 1. XVII. Parabolic Potential Theory (Continued)

(a) To each point of A corresponds an open neighborhood of the point


which carries a superparabolic function identically + oo on the part
of A in that neighborhood.
(b) If D is an open superset of A, there is a function superparabolic on
D and identically + oo on A. This function can be chosen to be the
potential GDµ of a finite measure.
Observe that in the discussion of parabolic-polar sets the case N = I is
not exceptional. The equivalence of (a) and (b) is proved just as in the
classical context (Sections V.1 and V.2). As in the classical context, a polar
set is a subset of a GG polar set. A set all of whose compact subsets are
parabolic polar will be called inner parabolic-polar.

Counterparts of Theorems V.3 to V.S.

A countable union of parabolic-polar sets is parabolic polar. If A is parabolic


polar, the smoothed reduction of a positive superparabolic function on A
vanishes identically; conversely, if the smoothed reduction of some strictly
positive superparabolic function on a set A vanishes identically, then A is
parabolic polar. The qualification "strictly positive" is necessary in this
converse because, for example, if A is the abscissa hyperplane and
lord >oi(), then (reduction relative to ft') tiO" = 0 even though A is not
parabolic polar. If A is parabolic polar and is a point not in A, there is a
positive superparabolic function finite at but identically +oo on A. This
result implies, as in the classical context [Section VI.4(c)], that if ti is a
positive superparabolic function on an open subset D of I(81P and if At and
A2 are subsets of b differing by a parabolic-polar set, then the reductions
of a on 4, and A2 coincide off the symmetric difference of these sets, and
the smoothed reductions of ti on 4, and A 2 coincide on D. The classical
context extension Theorem V.5 translates directly into the parabolic context
along with its applications. For example, if ti is subparabolic on D and if A
is a closed in b parabolic-polar set, null for the measure associated with ti,
then LM ti relative to D and LM a relative to b - A are equal on b - A. It
follows that Go = GD_,A on (D - A) x (D - A). The extended superharmon-
ic function minimum theorem (Theorem V.7) becomes (for all N): If ti is a
lower-bounded superparabolic function on an open subset D of 11 and
if at parabolic quasi every finite point of aD and also at = oo if D is
unbounded, lim inf,,.i u(q) c, then iu > c on D.

Minimality of r;)

In the classical context it was proved in Section VII. 10 as an application of


an extension theorem that the restriction of GD(C, ) to D - {C} is minimal
harmonic, and it then followed easily that for v positive and superharmonic
B. Parabolic-Polar Sets 307

on D, with associated Riesz measure v,

v
inf = v({t;}). (8.1)
D-g) GD(S, )

The corresponding argument in the present context shows that the restriction
of to D - {C} is minimal parabolic and that if r is positive and
superparabolic on D, with associated Riesz measure v,

inf U(n) = v({ }). (8.2)


h:GDla.c)>o) 6,6(4,

Similarly the restriction of GD(C, ) to b - is minimal coparabolic, and


the dual of (8.2) has the obvious formulation.

EXAMPLE (a) (Classical-polar versus parabolic-polar sets). Let A be a subset


of URN with projection A on the abscissa hyperplane. If A is classical polar,
then A is parabolic polar, and in fact A x R is parabolic polar because a
superharmonic function on R' identically + oo on A can be considered as a
superparabolic function on 118" (Section XV.15) and as such is identically
+ oo on A x P. In particular, when N > I a line in ft' parallel to the ordinate
axis is parabolic polar. Conversely, it will be proved in Section XVIII.11
that A is classical polar whenever A x P is parabolic polar.

EXAMPLE (b). If N Z 1, a singleton is parabolic polar. In fact, if N > 1, this


follows from Example (a). To prove the assertion for N Z 1, suppose that
the singleton is {0}, and let p be the measure supported by the set of points
J(0,-2 -k) , k z 1) with p { (0, -2-k) } = k-'. The superparabolic potential
Gµ is +oo at the origin; so {0} is parabolic polar. It follows that countable
subsets of libN are parabolic polar.

EXAMPLE (c). Let A be an IN measurable subset of R, with l r(A) > 0. The


set A x {0} is not parabolic polar in L N because a positive superparabolic
function on tjN, identically +oo on A x {0}. would be identically +oo on
the upper half-space (Lemma XVI.2).

Coparabolic-Polar Sets

A coparabolic-polar subset of 9' is defined as a set satisfying conditions (a)


and (b) at the beginning of this section with "superparabolic" replaced by
"cosuperparabolic." That is, a set A is coparabolic polar if and only if its
reflection in the abscissa hyperplane is parabolic polar. The application of
this section to coparabolic-polar sets is obvious. It will be proved in Section
XVIII.11 that a set is parabolic polar if and only if it is coparabolic polar.
308 1. XVII. Parabolic Potential Theory (Continued)

9. The Parabolic-Fine Topology


It will be convenient to introduce the fine topology in the parabolic context at
a much earlier stage than in the classical context. In the following discussion
otherwise unspecified topological concepts refer to the Euclidean topology.
In the present context there are two fine topologies, the parabolic and the
coparabolic. All concepts relative to the parabolic-fine topology are trivially
translatable into the corresponding concepts relative to the coparabolic-fine
topology by a reflection of U8" in the abscissa hyperplane; so no separate
discussion of the coparabolic-fine topology is given.
The parabolic-fine topology of AN (N a 1) is defined as the coarsest
topology of 68' making every superparabolic function (equivalently, every
upper-bounded superparabolic function) on 9" continuous. Since linear
functions of the space coordinate functions are parabolic and since the
ordinate function
r,, is superparabolic, it follows that every Euclidean open
interval in Id" is parabolic-fine open. Hence the parabolic-fine topology is
at least as fine as the Euclidean topology, and in fact it is finer because there
are superparabolic functions discontinuous in the Euclidean topology; for
example, the indicator function of the (open) upper half-space is superpar-
abolic and thus is parabolic-fine continuous. The parabolic-fine continuity
of this function implies that no point of the abscissa hyperplane is a parabolic-
fine limit point of the upper half-space; that is, the closure of the lower
half-space is a parabolic-fine neighborhood of each point of the abscissa
hyperplane. This result can be strengthened as follows. Let v be the potential
on AN defined in Section 5, Example (c), for which ti = + ao on the upper
half-space but u < + oo otherwise and u has limit + oo at the origin along
the abscissa hyperplane. If fi > 6(0), the function v A $ is superparabolic and
at the origin is strictly less than its limit there along the closure of the upper
half-space. Since t3 A fi is parabolic-fine continuous, it follows that the origin
is not a parabolic-fine limit point of the closure of the upper half-space;
that is, the lower half-space is a deleted parabolic-fine neighborhood of the
origin and so also of every point of the abscissa hyperplane. This fact implies
that the abscissa hyperplane, in fact each hyperplane parallel to it, has no
parabolic-fine limit point even though the hyperplane is not parabolic-polar.
Recall however that according to Theorem XI.6, every nonpolar set in the
classical context has a fine limit point. A more elegant example showing
that the lower half-space is a deleted parabolic-fine neighborhood of each
point of the abscissa hyperplane is furnished by the Green function G.
According to the dual of Theorem XVIII.l4(f) (whose proof depends on
an analysis of reductions not available at this stage), for fixed in (IN,

pf lim 6(4, + co.


nom{

Thus the set {il: 6(4, n) > c) is a parabolic-fine deleted neighborhood of


for every constant c.
10. Semipolar Sets 309

A set will be said to be "parabolic thin" at a point if the point is not a


parabolic-fine limit point of the set. Limit concepts relative to the parabolic-
fine topology will be distinguished by the prefix pf, for example, pf lim sup.
The set of parabolic-fine limit points of a set A will be denoted by ADf. An
asterisk will be added for the corresponding coparabolic concept : p*f lim sup,
AA'f.
The following properties of the parabolic-fine topology are derived in
essentially the same way as the corresponding properties of the classical
fine topology and the proofs are therefore omitted.
(a) The restriction of the parabolic-fine topology to an open subset 1)
of A' is the coarsest topology making continuous every superpar-
abolic function with domain D.
(b) The parabolic-f ine topology has a basis of (Euclidean) compact sets.
(c) The parabolic-fine topology has the Baire property that the intersec-
tion of a sequence of parabolic-fine open parabolic-fine dense sets is
parabolic-fine dense.
(d) If A has limit point 4 then A is parabolic thin at if and only if
there is a superparabolic function ti defined on a neighborhood of
4 such that

lim inf u(q) > (9.1)


A a 4-t

A point is a parabolic-fine limit point of a set A if and only if is a Euclidean


limit point of A and if each superparabolic function ii defined on an open
neighborhood of has ti(4) as a cluster value at along A. For example,
according to XV(14.1) and the example in this section, if A is the complement
of an 1N+, null set, the part of A strictly below a point of ll' has that point
as parabolic-fine limit point. Hence the space AN has no parabolic-fine
isolated point, and it follows that a nonempty parabolic-fine open subset
A of 9' is parabolic-fine dense in itself, A c A'f. If ,4 is a set parabolic thin
at and if D is an arbitrary open neighborhood of , there is a potential
ti = 6» N satisfying (9.1) for which p has compact support in b and the left
side of (9.1) is +oo. Thus a parabolic-polar set has no parabolic-fine limit
point. As already noted, the converse is false.

10. Semipolar Sets


It was proved in Section XI.6 that in the classical context a set is polar if
and only if it has no fine limit point, but it was pointed out in Section 9
that although a parabolic-polar set has no parabolic-fine limit point, the
abscissa hyperplane of Il' is not parabolic-polar even though it has no
parabolic-fine limit point. The following definition is therefore natural. A
subset 4 of 118N will be called parabolic-semipolar [coparabolic-semipolar] if
310 I.XVII. Parabolic Potential Theory (Continued)

A is a countable union of sets each of which has no parabolic-[coparabolic-]


fine limit point. If "parabolic" ["coparabolic"] is omitted here, that is, if
the definition is applied in the classical context, a semipolar set is polar so
the terminology "semipolar" is not used in the classical context. It will be
shown in Section XVIII.12 that a parabolic-semipolar set is necessarily
coparabolic semipolar, and conversely, and that such a set is IN,, null.
Trivially a countable union of parabolic-semipolar sets is parabolic semi-
polar. Thus a countable union of horizontal hyperplanes is parabolic semi-
polar. Such a union may have parabolic-fine limit points however.
A parabolic-semipolar set A is the union of countably many parabolic-
fine nowhere dense parabolic-fine closed sets; that is, the complement ofA
is the intersection of a sequence of parabolic-fine open parabolic-fine dense
sets, and (Baire property) the complement of A is therefore parabolic-fine
everywhere dense. It follows that if u and ti are superparabolic functions
on some open set band if u = ti (or ti 5 v) up to a parabolic-semipolar set,
then the relation is true everywhere on D.

11. Preliminary List of Reduction Properties


Let I) be an open subset of A', coupled with a boundary aD provided by a
metric compactification of D, and let A be a subset of ,6 v aD. In this section
we prove certain basic reduction properties, some under restrictions to be
removed or weakened later (see Section 16). We shall use repeatedly the
fact that in view of the preliminary version of the Fundamental Convergence
Theorem in Section 2 an equality or inequality between unsmoothed reduc-
tions relative to D is true IN,, almost everywhere on D for the smoothed
reductions and is therefore true everywhere on D for the latter.
(a) If v is a positive superparabolic function on D, finite valued and
continuous at each point of A r),6, then

R"=inf{RB:AcE,Al open in 6v at}. (11.1)

See the proof of the corresponding fact in the classical context in Section
111.5(e).
(b) If A and h are open subsets of D with A c h and if v is a positive
superparabolic function on b, then

OllO"QB = (11.2)

Just as in the classical context (Section VI.4), in view of the fact that the
smoothed successive reductions of v on A and h in either order lie between
Q 16lAOA and QtoO", it is sufficient to prove that (11.2) is true when A = B.
To prove this, we need only observe that since flvV" = ti on A a parabolic
function a on b majorizes 6 on ,4 if and only if u majorizes Q6" on A.
1 1. Preliminary List of Reduction Properties 311

(c) If u and t are positive superparabolic functions on b and if A is an


open subset of D, then

R"+ RA +A.4 (11.3)

See the proof of the corresponding fact in the classical context in Section
VI.4. Property (c) will be extended in Section 16(g) to cover countable sums
of positive superparabolic functions and arbitrary subsets A of b u aD.
(d) If v is a finite-valued continuous positive superparabolic function on
D, the set functions A F, R- and A - !r' are strongly subadditive on the
class of subsets of D u ab.
Property (d) will be extended in Section 16(k) where the restriction that
v be finite valued and continuous will be dropped.. To prove (d), we first
choose open subsets A and h of D, set v' = Its A Ite, and prove

AAuB + AAvB = RA + RB. (11.4)

The classical context proof of this reduction equality in Section VI.4 needs
no change except simplification : "quasi everywhere" is to be replaced by
"everywhere." Moreover, when A and h are open, R"$ " = v on A n B, so
AA ,j Thus (11.4) implies the validity of the strong subadditivity
inequality
R ire + R +RO (11.5)

when A and B are open subsets of D. In view of (a) this inequality is valid for
arbitrary subsets of D u aD, and the corresponding inequality for smoothed
reductions, true IN+t almost everywhere on D, must be true everywhere on D.
(e) If A, is an increasing sequence of subsets of D with union A and if
v is an increasing sequence of finite-valued positive continuous superpar-
abolic functions on b with finite-valued continuous limit 6, then

Iim R"R = I. (11.6)


_a
and the corresponding equation (I 1.6sm) for smoothed reductions is also
true. [See Section 16(e) for a stronger property.]
It is sufficient to prove (11.6). This limit relation is trivial on any open
connected component of b on which v vanishes identically; so we suppose
from now on that ti is strictly positive. First suppose that each set A
is open. Then (11.6) coincides with (I I.6sm) and the limit function v' _
is superparabolic on b, equal to v on A. It follows that 6' z R".
k
The reverse inequality is trivial so (11.6) is true for open sets. In the general
case choose in D, c > 0, a > 1, define = A, n If, < and apply (a) to
obtain an open superset B; of Aj satisfying
312 1. XVII. Parabolic Potential Theory (Continued)

AP(e) < 2 c.

The sequence A° is an increasing sequence of sets with union A, and we


define f? = Uo B, A. Apply the strong subadditivity property of reductions
to deduce

5 5 Rv"()+2E. (11.7)
0 0

Since (11.6) is true when the sets involved are open, (11.7) yields (n oo)

IZ () < I ) 5 lim R"1(c) + 2E 5 a lim R"rt ( ) + 2e.

Hence

A-41(6:5 lira R"^(), (11.8)


ft-M

and the reverse inequality is trivial.


(f) RV = ti on ADf n b; if ti is finite valued and continuous, then R," _
IZ.4 onD - AandAA = A =v onA"Ic b.
0 +6
If ti is a positive superparabolic function on D, majorizing ti on A r),6
and near A nab, then (by the parabolic-fine continuity of ti and v) ti also
majorizes t; on 4Pf; so R" = t on APf r),6. If c e b - A, let 1). be a decreasing
sequence of open neighborhoods of with intersection { }, and define
A^ = D - B Then A-4- is parabolic on h,,; so R ^( ) = R ^( ), and (n - co)
property (e) yields, under the stated conditions on ti,

k4(6 = $"(6- (11.9)

There remains the proof that (11.9) is true when e A°f n A. To see this,
observe that for such a point we have now proved at least that R"
Moreover the reduction value on the left is equal to because
e (A - { 5 })°f and the value on the right is equal to R"() because (Section 8)
a smoothed reduction is unaffected by a parabolic-polar change of the target
set. Thus (11.9) is true for the present choice of c.
The proof of (f) is based on (e) and therefore all but the first assertion
of (f) requires that v be finite valued and continuous. It will be proved
[statement in Section 16(e), proof in Section 17] that the conclusion of (e)
remains true for arbitrary positive superparabolic v, and it follows that the
conclusions of (f), restated in Section 16, remain true for arbitrary positive
superparabolic v.
12. A Criterion of Parabolic Thinness 313

12. A Criterion of Parabolic Thinness


The following lemma will be needed in the proof (Section 13) of the Fun-
damental Convergence Theorem in the parabolic context.

Lemma. Let b be a nonempty open subset of A", let v be a positive continuous


finite-valued superparabolic function on D, let be a point of D, and let A be
a subset of D that is parabolic thin at . Then

lim 0 (B a neighborhood of). (12.1)


144

Recall from Section 11(f) that the reduction in (12.1) is for every
choice of h if A is not parabolic thin at .
We can assume in the proof of the lemma that 0A. In fact, if A contains
then replacing A by A - { } does not affect the hypothesis that A is
parabolic thin at 4 and does not affect (12.1) because (Section 8) a smoothed
reduction on a set is unchanged if the set is decreased by a parabolic-polar
set. The lemma is trivial if 0; so we assume strict positivity below.
Since A is parabolic thin at , there is (from Section 9) a positive super-
parabolic function ti on D, majorized by v at and with limit + oo at
along A. If h is so small that v < on h, then
D,;O."(4)
hti a(
+ oc > z ) z inf u
U
"B sup ti
A'B

Z inf u vOAnB(S)s

from which inequality the lemma follows.


Just as in the classical context [Section XI.3(a")], we shall need a slight
variation of (12.1): if A is parabolic thin at t , then

lim 0. (12.1')
a"
Since sweeping has not yet been treated in the present context, the method
of proof in the classical context is not applicable. To prove (12.1'), choose
e > 0, let ut be a positive superparabolic function on D with tit(s) _ +oo,
and let Bo be a neighborhood of . Then
au(S)
1

oA ao (12.2)
1 fit sup t3.
infii t
314 I.XV11. Parabolic Potential Theory (Continued)

Apply (12.1) to find ho so small that the first term in the second inequality
is at most t;/2, and with this choice of h, observe that the second term is
at most E/2 if h is sufficiently small, since u, is continuous at .

Parabolic-Fine Limits and Cluster Values

A function ti from a deleted neighborhood of a point of A' into a metric


space has parabolic-fine limit [parabolic-fine cluster value] a at if and
only if ti has limit a at along a subset of A' which is a deleted parabolic-
fine neighborhood of [is not parabolic thin at r;]. The proof is a translation
into the parabolic context of that of Theorem XI.9, using the obvious
parabolic context version of Lemma XI.8.

13. The Parabolic Fundamental Convergence Theorem


Theorem. Let f : { ua, a e ! } be a family of superparabolic functions on an open
subset D of IB", locally uniformly bounded below, and define ti = inf.E t 6,
Then u < ti,

ti(p) = liminfti(q) = pflimti(q), (13.1)


r{ n-4

and

(a) u is superparabolic,
(b) u = u on each open set on which ei is superparabolic,
(c) ti = u except on a parabolic-semipolar set,
(d) there is a countable subfamily of r whose infmum has smoothing u.
Conversely, i/',4 is a parabolic-semipolar subset of D, there is a decreasing
sequence u, of positive superparabolic,functions on D with limit ti such that
>onA.
+

Assertions (a), (b), (d) and the first equation in (13.1) are contained in
Theorem 2. In view of (d) it will be assumed in the proof of the direct half
of the theorem that ti, is a sequence of parabolic functions, and it can even
be assumed, as in the classical context (Section 111.3), that these functions
are positive. We now choose r, and r2 with r, < r2, define

A, {ti > r2, r, > ti},

choose in D, and show that the set A = A, ,2 is parabolic thin at . If is


not a limit point of A, the assertion is trivial. If C is a limit point of A, then on
a sufficiently small neighborhood h of l; we have i+ > u(C) - (r2 - r,)/2 by
13, The Parabolic Fundamental Convergence Theorem 315

lower semicontinuity of ; so

ti Z u> +r2 2 r'= c (13 2)

on A n B. Hence (reduction relative to b) u ck ^B and to Z ch ne. Since


u(C) < c, we conclude that 1 > and therefore in view of Section
11(f) we conclude that A is parabolic thin at every point i; of D. Since each
function u, is parabolic-fine continuous, the function u is parabolic-fine
upper semicontinuous; so

ti(C) z pf lim sup ti(q) z pf lim inf ti(q) z u(O). (13.3)


y-( q-

If there were a point a in b at which the set;ond and fourth terms in (13.3)
were unequal, there would be two numbers, r, and r2, strictly between the
values of these terms at e, with r, < r2. Since is parabolic-fine continuous,
u
the set A,,,2 would not be parabolic thin at C, and consequently there can
be no such point C. Thus (13.1) is true. Finally (c) is true because

{u > u} = U ,4 (r, , r2 rational). (13.4)

Conversely, suppose that ,4 is a parabolic-semipolar subset of D. To prove


the converse half of the theorem, write A = Uo Ak with Akfn D = 0, and
let h be a sequence of open subsets of D forming a basis for the Euclidean
topology of D. Define

A,,,,,= 4knhmn{Rlk^Bm < l}

so that (Lemma 12)

Akn, C (AA- < A;km}, U Ate, = .4k.


M-0

Apply (d) to find a decreasing sequence 6,b of positive superparabolic


functions, each at most 1 and equal to I on A,,,,,, with limit t9k,,,m for which
AA_ The sequence ti, defined by
+a = A
n= tjk,,,,,2- k-m
k.m

has the properties described in the converse half of Theorem 14. In fact
ti is a monotone decreasing sequence of positive superparabolic functions
with
316 I. XVII. Parabolic Potential Theory (Continued)

6 2_*_1

so up to a parabolic-semipolar set and therefore every-


where on D. Hence v > v on A.

14. Applications of the Fundamental Convergence Theorem


to Reductions and to Green Functions

Application to Reductions

According to the parabolic Fundamental Convergence Theorem, a parabolic


reduction R" satisfies the relation A" = A- up to a parabolic semipolar set.
In view of the parabolic-fine continuity of superparabolic functions, A. = ti
on (A u.,40) 0 ,6, and this inequality combined with (13.1) implies that
R" = Ir,"; = v on A"'r b; the latter fact was proved under stronger hypoth-
eses on v in Section I1(f).

EXAMPLE. If D = iV and if A = ord z 0), then

A" _ tl on A, t on the upper half-space,


"
10 elsewhere, 0 elsewhere.

Thus in this example if t; is strictly positive, the set {A" > AA) is the abscissa
hyperplane, which is parabolic semipolar but is not parabolic polar.

Application to Green Functions

The counterpart of the argument in Section VII.4 showing by way of


the classical context Fundamental Convergence Theorem that the classical
context Green function with a given pole has limit 0 at quasi every finite
Euclidean boundary point shows in the parabolic context that d b(-, 0)
[O has limit 0 at every finite Euclidean boundary point except possibly
for those boundary points in some parabolic-semipolar [coparabolic-
semipolar] set. (It will be shown in Section XVIII.12 that a subset of A" is
parabolic semipolar if and only if it is coparabolic semipolar.) This boundary
limit result may be vacuous, however. For example, if b is the upper half-
space, the finite part of the Euclidean boundary is the abscissa hyperplane
which is both parabolic-semipolar and coparabolic-semipolar. Actually in
this case since 0 when ord 5 ord 4, it is trivial that b(-, q) has
limit 0 at every point of the bounding hyperplane.
16. Parabolic-Reduction Properties 317

15. Applications of the Fundamental Convergence Theorem


to the Parabolic-Fine Topology
Application to the Smoothness of a Parabolic-Semipolar Set

A parabolic-semipolar set is a subset of a Borel semipolar set which is a


countable union of Borel sets each of which has no parabolic-fine limit point.
In fact on the one hand according to Theorem 13 (converse assertion), a
parabolic-semipolar set is a subset of the exceptional set {u > u} in some
application of the theorem, and on the other hand according to the proof
of Theorem 13, the exceptional set {u > } in each application is a subset
of a countable union (13.4) of Borel sets each of which has no parabolic-
fine limit point.

Character of Av1

The counterpart of the argument (Section XI.6) that in the classical context
the set Af is a Euclidean Gd set and that A - Af is polar yields in the par-
abolic context that 4Pf is a Euclidean Ga set and that A - 4Pf is parabolic
semipolar. Although the set Af is fine perfect, the set 4Pf need not be par-
abolic-fine perfect. For example, if A = U i {ord = - I/n}, the set 4Pf is
the abscissa hyperplane, which has no parabolic-fine limit point.

Borel Measurability of the Parabolic-Fine Limit Superior Function

If ti is a function from an open subset of A' into A, the function i-.


pf lim supo.t a (il) is Borel measurable (and therefore the corresponding
inferior limit function is also Borel measurable) because if a e R, the set
Pf
{ :pflimsup u(q)>a}=
R--c
U :ri(d)za+
, =1 n

is a countable union of Ga sets.

16. Parabolic-Reduction Properties


The list of reduction properties in this section includes for completeness
some already discussed. The reductions are relative to an open subset I)
of A", provided with a boundary 8L) by a metric compactification. The
sets on which positive superparabolic functions are reduced are subsets of
,6 v a1), and no further hypotheses not stated explicitly are imposed on
either functions or sets. Proofs are given in Section 17 and consist merely
318 l XVII. Parabolic Potential Theory (Continued)

of the reference to the proof in the classical context when the latter proof
requires only translation into the present context. We stress that every prop-
erty in the following list is a property of reductions in the classical context
also, in which "(super)parabolic" is to be interpreted as "(super)harmonic"
and "semipolar" as "polar." Some of the proofs given in the present context
are unnecessarily indirect for the classical context but have the advantage
that they are applicable in many general contexts.
(a) If o'=u-R""6,then
RA = RA ab + RAnD (16.1)
v

and the corresponding equation (16. I sm) for smoothed reductions is also
true.
(b) R"SIi'"S6onA,
R" = v on (A u A°J) n D,
It" = RA = d on A°f n D, in particular on the parabolic-fine interior
of AnD.
R" = A-4 on b when A n 1) is parabolic-fine open.
It" is parabolic on D - A and equal to t there.
j "( ) = lim inf . R"(p) = pf lim,_S R"(ry').
(c) [See also (3.1) and (3.lsm).] If c is finite valued on A n D, then

R" = inf {! : A c h, h n D is parabolic-fine open,


(16.2)
B contains a neighborhood of A n 8D1.

If in addition v is continuous at each point of A n,6, the set B in (16.2)


can be restricted to be open in D v oD.
(d) If A, and AZ differ by a parabolic-polar subset B of b, then R" ' = Ro
onD-Band
(e) If A, is an increasing sequence of subsets of D with union A and if
v, is an increasing sequence of positive superparabolic functions on b with
superparabolic limit v, then
lim R"R =A-4 , (16.3)

and the corresponding equation (16.3sm) for smoothed reductions is also


true. If v = v for all n, then (16.3) and (16.3sm) are true for A, an increasing
sequence of subsets of D v 8D.
(f) R" 5 R" on ,6 with equality on ,6 - A, and also equality on A n D
up to a parabolic-semipolar set.
(g) If ti = Ea v is a superparabolic sum of positive superparabolic
functions on D, then
16. Parabolic-Reduction Properties 319

(16.4)
0

and the corresponding equation (16.4sm) for unsmoothed reductions is also


true.
(h) If ,4 c aD and if v, is a decreasing sequence of positive superparabolic
functions on b with limit v, then R=i .
(i) If A c R c D u aD and if A r),6 is parabolic-fine open, then
1 O61"QB = 1v0". If in addition the set Rn D is parabolic-fine open, then
mOB .=H
(j) If A n D and $ n D are parabolic-fine open and if v' = R? A R8,
then
R'4"8 + R''6. "B = R" + J. (16.5)
V V U

(k) The set functions !Y and :. are countably strongly subadditive on


the class of subsets of b u aD.
(1) (Strengthening of the assertion in (f) that A' = R'' on D - A.)
If e > 0 and if C is a compact subset of b - A, there is a positive super-
parabolic function ti on b, equal to v on ,4 n b and near ,4 n aD and satisfying
the inequality to 5 A" + e on C.
(m) If A is a relatively compact subset of D, then 6MDAA = 0; that is,
A' is a potential.
(n) If v is a potential on D, then RA = RAn° and AA = A'6.
(o) If v is finite valued and continuous, then for every point of ,6 the
set function is a Choquet capacity on DuaD relative to the class of
compact subsets of b u aD.
Observation. In the classical context (Section VI.5) the set function
was shown to be a Choquet capacity on D u aD relative to the class of
compact subsets of D u aD when v is finite valued. In the present context
finite valuedness of v is not sufficient for the validity of (o) according to
the following example. Let A be the closure of a ball in R", and let A be
the subset A x [0, I/n] of 9". Then A, is a decreasing sequence of compact
subsets of f8" with intersection A = A x {0}. Let v be the indicator function
of the upper half-space. Then (reductions relative to D = AN) the reduction
R vanishes identically, but R"n > 0 on the upper half-space because
according to Lemma XVI.2, the smoothed reduction R"' is at least equal
to the parabolic Poisson integral on the upper half-space with boundary
function the indicator function of A.
(p) If ,4 is an analytic subset of D u 0b, then

R" = sup {i : F c A, 'compact}, (16.6)


320 I. XVII. Parabolic Potential Theory (Continued)

and the corresponding equation (16.6sm) for smoothed reductions is also


true.
(q) If u and f, are bounded, then

sup IA- A< supIu - tit, (16.7)


D D

and the corresponding inequality (16.7sm) for smoothed reductions is also


true.
(r) There is a bounded continuous superparabolic potential u* on D
for which, for each subset A of b,

Av1nb={A =u*}. (16.8)

(s) Parabolic context counterpart of Section VI.3(o). To avoid repeti-


tion of complicated inequalities, we refrain from writing this property
explicitly. It is the set of reduction inequalities obtained by translating
Section VI.3(o) into the parabolic context; that is, "superharmonic" is
replaced by "superparabolic," and dots are inserted over set and reduction
symbols as required by the parabolic context.

17. Proofs of the Reduction Properties in Section 16


Proof of (a). See the proof of the corresponding property in the classical
context in Section 111.5(c). 0

Proof of (b). The properties listed under (b) have already been proved
except for the identification of A; with 9' when A n D is parabolic-fine
open. If A is a parabolic-fine open subset of D, the function is a positive
superparabolic function equal to v on A according to the third line of (b);
so f2" z A , and the reverse inequality is listed on the first line of (b). For
general A with A n 1) parabolic-fine open, combine property (a) with the
fifth assertion in (b) and the special case just considered to obtain the stated
identification. o

Proof of (c). The second assertion was proved in Section 11(a) by referral
back to the proof in the classical context. Since superparabolic functions
are parabolic-fine continuous, the same proof is applicable to prove the
first assertion. o

Proof of (d). See Section 8. 0

Proof of (e). If equation (16.3) is true, then (16.3sm) is true up to a parabolic


semipolar subset of 6 and therefore is true everywhere on 15 because both
17. Proofs of the Reduction Properties in Section 16 321

sides of (16.3sm) are superparabolic functions. In the following we therefore


consider only (16.3). The first assertion of (e) was proved in Section l 1(e) for
ti^ and ti finite valued and continuous, using Section 11(a), that is, using the
second assertion of (c) of the present section. This proof is applicable as long
as 6 is finite valued on ,4 if we use the first assertion of (c) and replace the open
sets in the Section 1 l (e) proof by parabolic-fine open sets. For arbitrary
ti,, ti, A^, A with A c D it follows that

Jim
+.1 = Aniv<+.Dt, (17.1)

and therefore (16.3) is true, in view of (d), except possibly on the parabolic-
polar set An (6 = +co), on which, however, (16.3) is trivial. To prove the
second assertion of (e) suppose first that A^ c 8D. Let i;, be a sequence
dense in b, and choose a positive superparabolic function ti,, on b, majorizing
ti near A^ n aD and satisfying

P'4-(4,) + 2-", j S n.

The function
go
tik=limn"R+
ROOD
Y
n=k

is positive and is superparabolic on b because the indicated limit is parabolic


on b, and the sum is a sum of positive superparabolic functions and is finite
at each point ,. The function 1im".m A-4 majorizes for all n, so uk z ti,
for n z k. It follows that rik z kv , and when k -+ oo, we find that lim^..
z up to a parabolic-semipolar subset of D and therefore everywhere on
D since both sides of this equation are parabolic functions on D. Since the
reverse inequality is trivial, (16.3) is true when A c aD and t, = ti for all n.
According to the first assertion of (e), equation (16.3) is true when A c D
with no restriction on v,. Now apply (a) to find
A n6 A n6 ,
A A nB6
+ ZR A dD
+ Ii,r
where 6. = v - If"""a6 and v' = ti - A4"oD.

Hence
limb""zkna6+A-nD=A..
R-00

Since the reverse inequality is trivial, (16.3) is true, as was to be proved. o

Proof of (f). We have already applied the Fundamental Convergence


Theorem to find that A-64 = A' up to a parabolic-semipolar subset of D.
322 1 XVII. Parabolic Potential Theory (Continued)

Equality on b - A was proved in Section 11(f) for ti finite valued and


continuous but it was pointed out there (in an observation following the
proof) that the present property (e) implies equality on D - A without this
restriction.

Proof of (g). To prove (g) for two summands, that is, to prove
A
V,+V =1Z" +A2 (17.2)

and the corresponding equation for smoothed reductions, observe that (17.2)
appears as (11.3) in Section 1 I (c) for A an open subset of 1) but that the
proof of this special case was there referred back to the proof of the corre-
sponding special case in the classical context. That proof [Section VI.4(f)]
translates trivially into the present context for ,4 a parabolic-fine open subset
of D, with no restriction on t%,, o,. In view of the evaluation in (3.1) of a
reduction in terms of reductions on subsets of D, (17.2) is true whenever
An D is parabolic-fine open. It then follows from (c) that (17.2) is true
with no restriction on A if 6, and v2 are finite valued on A n D. Hence
with no restriction on v, and i'2 and with Ao = 4n {v, + i2 < + oo },

ltd 12V - + A4

According to (d), these reductions are equal to the corresponding reductions


on A, that is, to the reductions in (17.2), except possibly on the parabolic-
polar set ,4 - Ao on which in fact (17.2) is trivial. Hence (17.2) is true; that is,
(16.4) is true for two summands and therefore for finitely many summands.
Hence for finitely many summands (16.4sm) is true up to a parabolic-
semipolar subset of D and is therefore true everywhere on D because each
side of (16.4) is a superparabolic function. The extension to infinitely many
summands is effected as in the classical context [Section VI.4(f)].

Proof of (h). The proof follows that of the corresponding classical property
in Section VI.4(g).

Proof of W. The proof follows that of the corresponding classical property


in Section VI.4(h) but observe that in the argument there for (h,) no extra
hypothesis on the smoothness of A was used, whereas the hypothesis that
A n D is parabolic-fine open plays an essential role in proving the parabolic
counterpart of (h,).

Proof of (j). This property was proved under added restrictions (or rather
referred back to its classical counterpart) in Section 11(d). The method of
proof referred to is applicable in the present context whenever the sets
A n D and . n D are parabolic-fine open.
17. Proofs of the Reduction Properties in Section 16 323

Proof of (k). Equation (16.5) implies the strong subadditivity of the set
function A on the class of sets A with A n D parabolic-fine open
[the argument in Section 11(d) for A n D open is applicable when A n 6 is
merely fine open]. An application of (c) then shows that the strong sub-
additivity inequality (11.5) is true whenever v is finite valued on (A v h) n D;
so (11.5) is true with no restriction if A and $ are replaced by A n {v < + oo }
and h n {v < + oo }, respectively. Hence by (d) the inequality (11.5) is true
as written except possibly on the set (A i A) n {v = +co}, and (11.5) is
therefore true on b because the inequality is trivially true on (A u h) r-%,6.
The strong subadditivity inequality for smoothed reductions is true on D
because the inequality is true up to a parabolic-semipolar subset of b and
the two sides of this inequality are superparabolic functions. Properties (e)
and (g) imply that there must be countable strong subadditivity when there
is strong subadditivity.

Proof of (l ). Denote by 4k the set of points of D n A n {v < + oo } at distance


z 1/k from C. The function It"k"'""an) is continuous, in fact parabolic, on
the neighborhood {deb: ( - CI < 1/k} of C. The downward-directed
family

{ A6: Ak u (A n aD) c E,1) r),6 parabolic-fine open,


(1 7.3)
A a neighborhood of A r),6) le

of continuous functions on e, with limit according to (c), is


uniformly convergent (Dini's theorem). Thus if E > 0 and k z 1, there is
a choice Ek of $ in (17.3) so small that
.tgk < jtAkv(AnOD)
G + 2-k-'E

on C. Fork z 1 let uk be a positive superparabolic function on b, identically


2_k-'E
+oo on the set {4: 14 - Cl z 1/k, v(4) = +oo} and at most on C.
(For example, if tik' is a positive superparabolic function on D, identically
+ oo on the set {v = + oo }, let tir be the reduction of uR on the set {14 - C I >
1/k}. Then ti; is continuous on the compact set e, and we define u, =
2 -k-' ez /sup.ti; .) By countable strong subadditivity (k)

E
Ui°Bk 5 s RA +
e u 2 2

on C. The function

A LUk]
t
J
324 1. XVII. Parabolic Potential Theory (Continued)

is superparabolic on D, equal to v on A n D and near A n 3D, and is at


most R" + s on e, as desired.

Proofs of (m) and (n). The proofs follow those of the corresponding classical
properties in Sections 111.6 and 111.5, respectively. Note however that the
counterpart ti of the function u in 111.6 should be superparabolic, but not
parabolic on any open subset of D; choose for example the potential of a
measure supported by a countable dense subset of D.

Proof of (o). In view of (e) all that remains to be proved is that


R6 whenever v is finite valued and continuous and A. is a decreasing
sequence of compact subsets of D u aD with intersection A. If A c $ and
if h is open, then A. c R for sufficiently large n so that RB z lim,,..,, Ru",
and therefore in view of (c) R" z lim,,..,, R"4. The latter inequality is actually
an equality because the reverse inequality is obviously true.

Proof of (p). Suppose first that v is a finite-valued continuous positive


superparabolic function on D. In this case (16.6) is true because according
to (o) for each point of D the set function is a Choquet capacity
relative to the paving of compact subsets of D u ab. For this choice of d
the R'6(4) capacitability of the analytic set A implies (16.6). Equation (16.6sm)
follows on b - A because by (f) the equations (16.1) and (16.6sm) are
identical on b - A; equation (16.6sm) on A is deduced by the following
argument. According to (d), when e A the left side of (I6.6sm) is unchanged
when A is replaced by A - the right side is also unchanged in view of
(d) and (e) [if 4 e tin (16.6sm), replace tin (16.6sm) by t less each member
of a sequence of balls of center and radii tending to 0]. Thus (16.6sm) is
reduced to (16.6), and we have proved (p) when v is finite valued and con-
tinuous, in particular, when ti is parabolic. Next we prove (p) when v is a
potential. Let v, be an increasing sequence of finite-valued continuous
potentials on D with limit v. According to (n) R"
6 = and R" = A"-16'
+V
-

so we can apply (e) and the fact that (p) has been shown to be true for v
to deduce

R,", =AA-6=sup{R$°:neZ+}
= sup { R': n e l+, r' c A n D, t compact)
= sup { I : C c A n b, ' compact).

Since the last supremum is at most the supremum in (16.6) which is itself
majorized by the left side of (16.6), equation (16.6) is true. The same argument
involving smoothed reductions yields (16.6 sm) when v is a potential. We
have now proved (p) when ti is either parabolic or a potential so (p) is true
as stated in view of (g) and the Riesz decomposition of a positive super-
parabolic function.
17. Proofs of the Reduction Properties in Section 16 325

Proof of (q). The proof follows that of the corresponding classical property
in Section VI.4(n). o

Proof of (r). The proof must be more than a translation of the proof of
Lemma XI.10 into the parabolic context because the parabolic Dirichlet
problem has not yet been treated. The basic method of the proof of Lemma
XI.10 will be used however. Let D be a nonempty open subset of RN, let I
be an interval with closure in D, and let a [b] be the ordinate value on the
lower [upper] face of I. Define s) on D by

s-a
s) =
b-a
0

Then ti is superparabolic, and we now show that (reduction relative to D)


the function R' is continuous. The function is equal to the smoothed reduc-
tion according to the fourth property in (b). The function is lower semicon-
tinuous, and E. 5 ti; so k. is continuous at a point if this inequality is
actually an equality at that point. Since R is parabolic and therefore contin-
uous on D - 8I, we need only prove that R,, = ti on 81. On the lower face
of j, 0 5 R 5 ti = 0 so there is equality on this face. If is a boundary
point of j not on this face, it will be proved in the next paragraph that I
is not parabolic thin at 4; so since k is parabolic-fine continuous and equal
to ti on I, there is equality at . (Or, apply the second assertion of (b).) Thus
R; is continuous. Let I, be a sequence of intervals with closure in D forming
a basic for the topology of b, let tip be defined for Ii as ti was for I, and set

(17.4)

The function ti' satisfies the conditions demanded in (r) by a translation


into the present context of the corresponding proof in the classical case in
Section XI.10.
There remains the proof that each point 4 not on the lower face of I
is in Pf. Since (from Section 9) each point of the abscissa hyperplane is a
parabolic-fine limit point of the lower half-space, it follows that Pf contains
the interior points of the upper face. Finally suppose that 4 is a point of a
lateral face, not on the lower face. If N = 1, let 1, be the reflection of j in
the bounding side of I containing t . A trivial symmetry argument shows
that I, is parabolic thin at if and only if I is. If both are parabolic thin at
, their union is also. However, the set I v I, contains all but an 12 null set
of the part of some Euclidean neighborhood of strictly below , and so
[Section 9(d)] I v it cannot be parabolic thin at . Hence e M. The similar
treatment of the case N > 1 is left to the reader. o
326 1.XVI1. Parabolic Potential Theory (Continued)

Proof of (s). The proof follows that of the corresponding inequalities in the
classical context in Section VI.4(o).

18. The Classical Context Green Function in Terms of the


Parabolic Context Green Function (N z 1)
Let D be a nonempty open subset of R", and define f) = D x It The para-
bolic context Green function 6r, is then invariant under translations of the
ordinate axis,

CD((, S), (1, t)) = G`DW, 0), (h, t - s)).

It follows that for fixed and t the function (tl, s) '- 6D((4, s), (tl, 1)) is super-
parabolic on b and is parabolic on D - t)} with the canonical isolated
singularity at t); so s), (q, t)) z OD((n, s), t)). Repeat this inter-
change of space variables to derive equality here, that is, to find that the
function (4, i) i-. CD((4, s), (ii, t)) is symmetric. Define the function GD on
RxDxDby
GD(t, , I) = OD((, 0), (17, - t)). (18.1)

If D = R" or more generally if D is R" less a closed classical context polar


set, then (from Section 8, Example (a)] D is A" less a closed parabolic-polar
set; so (from Section 8) CD = 6 on 15 and GD(t, ,,) = e (t, - ri). Whatever
the choice of D, the function 6D(1, -, -) is symmetric on D x D, and the func-
tion 4D(-, -, ry) is superparabolic on D, parabolic on D - ((q, 0)). Moreover
Go(t, , ry) 5 6(t. - q). The function GD will be identified in Section 2.IX.17
with the transition density function of Brownian motion in D.

Theorem. If D is a Greenian subset of R",

(18.2)
a" =
f(N/2- I) if N > 2, a2 =
1 2
at = a2 .
2n""2 a2 2na2'

We first verify (18.2) in three special cases. If N > 2 and D = R", then
= AN, 6D = 0, and (18.2) follows by direct evaluation. Incidentally it also
follows that for any choice of D the integrals in (18.2) converge when N > 2
and ' q. If N = 2 and if D = D+ is a half-plane, let ry be the reflection
of ry in OD'. Then (from Section VIII.9)
-n
GD. 4) = log l (18.3)
18. Classical Green Function vs Parabolic Green Function 327

On the other hand, according to (4.3), the function /,,. is given by

J(2aQ2t)-t (ex'2
2a2t
-ex p -14-g*IZ\
2a2t
ift>0
r,D, (t, x,11) = )
(18.4)
10 if t<0.
Integration yields (18.2) in this case also. If N = 1, let D = D' be the half-
line ]0. + oo[. In this case q) = A q according to Section XIV.6, and
according to (4.3), the function fD is given by (18.4) except that the first
factor on the right has exponent -3 instead of -1. In this case q = -q
Again (18.2) can be verified by direct integration.
To prove (18.2) in the general case for N > 2, denote by the value
of the equal integrals on the left. For fixed q the function s) - g0(S, q)
is superparabolic on D because the superparabolic function inequality is
satisfied; so q) is superharmonic on D. The corresponding argument
shows that q) is harmonic on D - {q}. Furthermore the first integral
in (18.2) is the value at of the superparabolic potential of the measure
1, on the line parallel to the ordinate axis through (q, 0). Hence the function
has no positive harmonic minorant other than 0. That is, 9D(., q) is
the superharmonic potential of a measure supported by {q}, and it follows
that q) = q) for some positive constant c. Now

4) = q) - u( , q), S = (i, s), q = (q, t), (18.5)

with h) a positive parabolic function on D, and integrating this equality


with respect to 1, over the line through (q, 0) parallel to the ordinate axis
yields

q) = aNG(S,,) - J u( , (q, t))1 t (di). (18.6)

The integral on the right defines a harmonic function of on D, and it


follows that c = aN, as was to be proved.
If N = 2, it is enough to prove (18.2) for a bounded set D because when
D is an increasing sequence of bounded open subsets of D with union D,
it follows that lim, and d = 6D according to Theorem
VII.6 and its parabolic counterpart. Let D+ be a half-plane containing D,
and define b' = D+ x R. Although (18.5) is valid when N = 2, we cannot
integrate as in the case N > 2 because

f W O(( ,s),(,,t))1(di)= +oc

when N < 3. We therefore replace (18.5) by a relative equation (see Section


XVII.4)
328 1 XVII. Parabolic Potential Theory (Continued)

o = 0D (c, i) - n), (18.7)

where ry') is a positive parabolic function. The proof continues as in the


case N > 2, using the truth of the theorem when D = D. When N = 1,
there remains only the proof of (18.2) for D a finite interval, and this is
carried through just as in the case of bounded open sets when N = 2.
Observation. According to Theorem 18, if D is Greenian, if p is a measure
on D, and if p is the product measure p x 1, on 1) = D x R, then s)
aNGDp(4 )

19. The Quasi-Lindelof Property


Theorem. An arbitrary union of parabolic fine open sets differs by a parabolic-
semipolar set from some countable subunion.

The proof follows that of the corresponding property in the classical


context (Theorem Xl. 11).
Chapter XVIII

The Parabolic Dirichlet Problem, Sweeping,


and Exceptional Sets

1. Relativization of the Parabolic Context; The PWB Method


in this Context
Let b be a nonempty open subset of ft', and let h be a strictly positive
parabolic function on D. A function v/h on D will be called h-parabolic,
b-superparabolic, or h-subparabolic if o is parabolic, superparabolic, or sub-
parabolic, respectively. The notation will be parallel to that in the classical
context, with h omitted when h =- 1. Thus 6m k' "1j. , it, H, ... need no
further identification. In the dual context in which h is coparabolic we write
6M4, ltp , t$, 1Y f , ... .
The PWB method for solving the first boundary value (Dirichlet) problem
for relative harmonic functions translates directly into the parabolic context.
As in the classical context, the boundary is that obtained by a metric com-
pactification of the given open subset D of ft'; the boundary of D relative
to the one-point Alexandrov compactification of off" will be called the
Euclidean boundary. Upper and lower PWB" solutions H f and ft f4 are defined
on D corresponding to a specified boundary function f, and if these solutions
are equal and parabolic, the function f is parabolic h-resolutive with PWB"
solution Hf = M4.. The definitions of h-resolutive boundaries, h-regular
boundary points, h-parabolic measure boundary subsets, and so on are
translations of the corresponding definitions in the classical context. The
properties of h-harmonic measure null sets and of PWB" solutions derived
in Sections VIII.5 to VIII.7 go over into the parabolic context with trivial
changes in the derivations. The relations between PWB" solutions and
reductions (Section VIII.2) are preserved in the present context:

when d is a boundary subset.

ExAmPLE (a) (Euclidean boundary). If e D and if A = {il e,015: ord ry' z


ord }, then H 0. In fact, if _ (, s), if _ (, s - I /n), and if u is
330 1 XVIIi. The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

the indicator function of the set f0 a b: ord it > s - 1/n}, the function u is
in the upper PWB" class on D for the boundary function 1.4; so H 0
when n is so large that a D. Hence H,., (o = 0.

EXAMPLE (b). Let D be a Greenian subset of R coupled with a boundary OD


provided by a metric compactification of D, define D = D x I8, and define
aD as the boundary of D provided by the compactif ication (D V OD) x 9,
where 08 is the two-point compactification of R. A trivial adaptation of
Example (a) shows that the upper boundary (DvaD) x { + oo} is an
h-parabolic measure zero subset of aD for every choice of strictly positive
parabolic function h on D. Alternatively this follows from the fact that the
function ti: s) i-* e' is a positive h-superparabolic function on D for every
positive superparabolic function h on D and u has limit +oo at the upper
boundary of D. The role of the lower boundary of D is more delicate. We
show that the lower boundary is parabolic measure null if D is on one side
of a hyperplane of I8". It is sufficient to exhibit a sequence ti, of positive
parabolic functions on D with limit 0 and with ti (1') = 0 when ord ry' S -n,
and it is therefore sufficient to exhibit such a sequence when D is the half-
space on one side of a hyperplane of R'; we can even assume that the
hyperplane is a coordinate hyperplane and that D is the set on which the
corresponding coordinate function is strictly positive. Finally we need only
consider the case N = 1 because a sequence ti, in this case induces one in
the general case. The sequence ti, on b = 0} (N = 1) defined by

I - (s + n) Jexp2it/i(d2)
(2/n)''22
°`
ifs > -n
tt(, s) _ (s + n)
1 ifs< -n
has the stated properties. (We have simplified the notation by taking a = 1.)
It is not difficult to show, although not necessary for present purposes, that
u (t, s) is the parabolic measure of {(0, t): t 5 -n}, the Euclidean boundary
subset relative to s).

EXAMPLE (c) (Euclidean boundary, h =- 1). Let b be an interval in W.


According to Example (a), the upper boundary of D is a parabolic measure
null set. We show that aD is parabolic resolutive by showing that every
finite-valued continuous boundary function f is parabolic resolutive with
Hf = PI(D, f ). Define f, on aD as f except that f, = info f on the upper
boundary. Then Hf = Hf because f = f, up to a parabolic measure null
set. Since Pl(D, f) is in the upper PWB class on D for f,, it follows that
l l(D,f) >- Hf, = Hf and similarly PI(D,f) 5 Hf; so PI(D,f) = Hf, as
asserted.

EXAMPLE (d) (Euclidean boundary, h = 1). Let D be a slab 08'v x ]0,6[


with 0 < 6 < + oo. According to Example (a), the finite part of the upper
1. Relativization of the Parabolic Context; The PWB Method in This Context 331

slab boundary is a parabolic measure null set. We show that the boundary
is parabolic resolutive with Hf = PI(D, jo) when f is a finite-valued contin-
uous boundary function and To prove these assertions,
observe first that the function fo has limit f(oo) at oo and (from Section
XVI.1) if a is either oo or a point of the lower slab boundary the function
PI(D, fo) has limit f(C) at t . When b = + oo, the function PI(D, fo) is in both
the lower and upper PWB classes on D for f and so can be identified with
fff ; when b < + oo, the method used in Example (b) shows that Hf =
PI(D, fo)

EXAMPLE (e) (Euclidean boundary). (Cf. Section VIII.3.) Let u be h-super-


parabolic on D, and suppose that ti has a finite or infinite limit at every
boundary point, thereby defining a boundary function f Suppose also that
u is lower bounded and that is upper bounded. Then f is parabolic
h-resolutive with Hl = OMoti because ti is in the upper and 6M4-ti is in the
lower PWB' class on b for f. Moreover if h =- 1, if ti has a superparabolic
extension to a neighborhood of d8" r D, and if this extension has a finite
limit at every point of 815, then lim, ft. (q) = whenever the boundary
point a is not in some at most parabolic-semipolar set. In fact OMDU can
be obtained as the restriction to D of the limit of a decreasing sequence of
locally-lower bounded superparabolic functions on a neighborhood of
V - D, each equal to ti outside a compact subset of b, depending on the
function. [See the corresponding argument in Section XVII.4 as applied to
analyze OMDG(-, it), or see the general discussion of GMD in the classical
context in Section 111.1.] According to the parabolic Fundamental Conver-
gence Theorem, the limit function U' majorizes its smoothing, which is
parabolic on D, and the two functions are equal up to a parabolic-semipolar
set. At a nonexceptional boundary point C of D the function OMDU has
limit ti'(C) on approach from within D because u' is lower semicon-
tinuous, u' S U, and U has this limit.

Application to the Parabolic Resolutivity of the Euclidean Boundary

The Euclidean boundary of an open subset of d8" is parabolic resolutive,


and the set of its finite parabolic irregular boundary points is parabolic-
semipolar. In fact this assertion follows from a trivial modification of the
proof for N > 2 of Theorem VIII .4, the counterpart in the classical context
of this assertion. Observe that the set of finite parabolic irregular boundary
points is parabolic-semipolar rather than parabolic-polar.

Application to a Representation of Parabolic Green Functions (Euclidean


Boundary, h = 1)

If 1 is a point of D, the restriction to aD of 6) (defined as 0 at oo) can


be taken as f in Example (e), and it follows that ry') = ry') - Hf.
332 I.XVIII. The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

More generally. if h is an open subset of b, if it a h, and if f is the restriction


to ah of q) (defined as 0 on ah n ab), then OA(', rj) is t;p(-, q) less the
PWB solution for h with the boundary function f.

2. h-Parabolic Measure
Let D be a nonempty open subset of A' coupled with a boundary provided
by a metric compactification, and let h be a strictly positive parabolic
function on D. The development of h-parabolic measure follows that of
h-harmonic measure; so the details will be omitted. If A is a subset of aD
with an h-parabolic-resolutive indicator function, we define fip(-, A) = H''a,
call A a EiD measurable set, and call µo(, A) the h -parabolic measure of A
relative to . The set function µv(, -) is a probability measure, and the
completion of the restriction of this measure to the class of At measurable
Borel sets is an extension of this measure. The h-parabolic measure null
sets defined above are the sets A for which µD(-, A) vanishes identically. A
boundary function measurable with respect to the a algebra of Nn measurable
sets will be called µ$ measurable, and L'(µo) is defined as the class of µ11
measurable functions from aD into R, with two functions identified when
they are equal un almost everywhere (that is, up to an h-parabolic measure
null set) for which ! f is is 4D(, -) integrable for every point 4 of D. A function
f from aD into A is h-parabolic resolutive if and only if f is in L' (ti"n), and
if so, then Hf =
The work in Sections VlII.8 to VIII.10 goes through in the present
context except that even if D is connected, it is not true that a positive para-
bolic function on D which vanishes at a point must vanish identically; so
the L' class for depends on 4; the class of 1o measurable sets and
the class L' (14) were defined taking this fact into account.
In the present context the notation tH of Section VIII. II becomes ie.
Thus if v' is an h-superparabolic function on b and if h is an open relatively
compact subset of D, the function ikti is h-superparabolic on D and is
h-parabolic on h, and &:9 S v, with equality on D - h up to a parabolic-
semipolar subset of Oh (Euclidean boundary). In particular, if ti z 0, then
t8v = h4U-$.

EXAMPLE (a). If b is an interval, we have seen in Section 1 that the Euclidean


boundary is parabolic resolutive, that the upper boundary is a parabolic
measure null set, and that µo(-, f) = PI(D, f) so that parabolic measure is
given by the densities evaluated in Section XV.9. The inner points of the
upper boundary are parabolic irregular, but all other boundary points are
parabolic regular according to Section XV.9.

EXAMPLE (b). If D = U8' x x ]0, d[ is a slab we have seen in Section I that


the Euclidean boundary is parabolic resolutive, that the abscissa hyperplane
3. Parabolic Barriers 333

part of the boundary has parabolic measure 1, and that

G(s, b, q)/N(d77) [= s)]

for (ry, 0) on the abscissa hyperplane. The points of the abscissa hyperplane
and the point ao are parabolic regular, but all other boundary points are
parabolic irregular. It is easy to see (cf. the classical context for D a ball in
Section VIII.9) that the Euclidean boundary is parabolic universally resolu-
tive, that the abscissa hyperplane has h-parabolic measure l for every choice
of h, and if (Riesz-Herglotz type representation, Theorem XVI.6), h =
PI(D,! .), then

µn(4, AO = AS- , h) s)]

for (rl, 0) on the abscissa hyperplane.

The Support of µD (Euclidean Boundary)

If c e b, define DS = try' e,6: ord ry < ord ). We have seen in Section 1,


Example (a), that µD(, ) is supported by the part of aD strictly below .
Moreover, if A is a Borel subset of eD strictly below , then the function
A) is the restriction to D4 of p A) because if v [u] is in the upper
[lower] PWB class on b for the boundary function 1,; on a D, the restrictions
of these functions to D4 are in the corresponding PWB classes on DS for the
restriction of 1,; to oAt. This fact has the following useful implication, in
which the roles of A and D are reversed. Let now D be an open subset of
A', and let be a boundary point of D. Suppose that every point of D is
strictly below 4 and that the part of some open neighborhood h of strictly
below lies in D. Then if A is a Borel subset of 8D, the function A)
is defined and parabolic on b and has the parabolic extension µD A) to
Dvh.

3. Parabolic Barriers
The classical context definitions of weak h-barrier and h-barrier translate
directly into the parabolic context, as does the proof that if there is an
h-barrier locally at a boundary point, then there is an h-barrier defined on
the whole open set in question. The classical context Bouligand theorem
(Section VIII. 12) is also true in the present context: If there is a weak para-
bolic barrier at a Euclidean boundary point a of an open subset D of 1'(N,
then there is a parabolic barrier there. To see this, adapt the classical context
proof as follows. In view of the fact that the existence of a parabolic barrier
334 I.XVIII. The Parabolic Dirichict Problem, Sweeping, and Exceptional Sets

at C is a local property it is sufficient to show that there is a parabolic barrier


on the trace on b of an open neighborhood of e ; so b can be supposed
bounded. Suppose then that D has diameter S < + oo and that ti is a weak
parabolic barrier for b at . For notational simplicity assume that C is the
origin. Define 4) : it = (ry, t) tll2 + t2 on b, choosing c so large that ¢ is
subparabolic. The function 4) is in the lower PWB class on ,6 for the boundary
function f: q = (ry, t) i- cI,12 + t2 on D. Hence Hf z 4), and it will be shown
that Rf is a parabolic barrier for b at the origin by showing that Hf has
limit 0 there. Let h be an interval containing the origin, of diameter r, let
A be a compact subset of b n 6h, and let t/i be the indicator function of
(oil - A) n,6 on 6h, so that Aj&, i,1i) is parabolic on h with limit I at every
point of aB - A not on the upper boundary. Since the upper boundary of h
has parabolic measure 0, there is a positive superparabolic function v on h
with v(0) < + oo and with limit + oo at every point of the upper boundary.
If uo is in the lower PWB class on ,6 for f, apply the maximum theorem for
subparabolic functions to find that uo on B n D is at most (c + I)r2 if Oh
does not meet D and is at most
t!
(c + l )r2 + (c + 1)6' r + NB(., W)J + rv
LLinf,4 ti

if Oh does meet D. Thus in both cases Hf is majorized on B n D by the


sum (3.1). This sum has limit

(c + l)r2 + (c + l)62pB(0,,i) + rti(0)

at the origin, and this limit is at most 2(c + 1)r2 + r6(0) if A is sufficiently
large. Since r can be made arbitrarily small, the function Hf has limit 0 at
the origin, as was to be proved.
Theorems VIII.13 and VIII.14 on the relations between regularity and
barriers translate directly into the parabolic context.

4. Relations between the Classical Dirichlet Problem and the


Parabolic Context Dirichlet Problem
Let D be a Greenian subset of R' coupled with a boundary OD by a metric
compactification of D. Define D = D x R, and define 0,6 as the boundary
provided by the compactification (D u aD) x R, where R is the two-point
compactification of R. Let h be a strictly positive harmonic function on D,
and define the parabolic function h on D by setting s) = We now
consider the Dirichlet problem for h-harmonic functions on D and for h-
parabolic functions on D. Recall [Section 1, Example (b)] that the upper
boundary (D v OD) x { + oo } of D is an h-parabolic measure null set. We
suppose that the lower boundary of D is also h-parabolic measure null. This
5. Classical Reductions in the Parabolic Context 335

is true if D is not too large, for example [Section 1, Example (b)], if D is


on one side of a hyperplane in R' and h - 1. Let f be an extended real-valued
function on OD, and define f on Of) by setting f( ), except that no
restriction is imposed on f on the upper and lower boundaries of D. This
looseness is convenient, and the actual definition off on the upper and lower
boundaries does not affect the PWB" solutions for/. If u is an h-superhar-
monic function in the upper PWBh class on D for f, the function ti: s) i-
u( ) is in the upper PWB class on D for f if f is defined as - co on the upper
and lower boundaries of D. Hence tl fs) 5 H and similarlyff s) z
We conclude that f is h-parabolic resolutive on b whenever f is
h-harmonic resolutive on D, and then H f(g, s) = H In particular, if A
is a p' measurable subset of 3D, the product set A = A x R is j4 measurable
and JD ((g, s), A) = pD( , A). Conversely, if f is Borel measurable and PWBh
resolutive and if 8D is h-resolutive, then f is Borel measurable, and we show
it is PWB" resolutive. In fact there is nothing to prove if f is bounded, and
in the general case

PhD(-. I f I) = lim .u
M-m
I f I A n) = lim µn(', I f I n n) = µD(', I f I) < + oo.

Suppose in this paragraph that OD is PWB' resolutive and that aD is


PWB' resolutive. This hypothesis is satisfied if h = I and aD is the Euclidean
boundary. (Note that BI) is not then the Euclidean boundary, but the PWB
resolutivity of aD follows from that of the Euclidean boundary of D and
the fact that the upper and lower boundaries of D are parabolic measure
null.) If (g, s) is a finite h-regular point of 0,6, the point is an h-regular
point of OD. In fact a bounded h-resolutive function f on OD, continuous
at , defines a bounded function jon aD as above (setting f - 0 on the upper
and lower boundaries). The function f is h-resolutive, bounded, and contin-
uous at so Hf has limit at that is, Hf has limit at .
If h - I and if is a finite regular point of OD, then each finite point s)
of aD is regular because a weak barrier for D at when considered as a
function on D is a weak barrier for D at s).

5. Classical Reductions in the Parabolic Context


Let D be a Greenian subset of R', coupled with a boundary provided by a
metric compactification, define D x R, and compactify D as in Section 4 to
obtain M. Suppose that the lower boundary of b is parabolic measure null;
we have seen in Section 1 that this condition is satisfied if D is on one side
of a hyperplane of RN. Let v be a positive superharmonic function on D,
and define the superharmonic function a on D by setting Let
A be a subset of D, and define ,4 = A x R. We now show that if A is analytic,
then
336 1 XVIII The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

R" (S, s) = R"O, (5.1)

A s) = R O. (5. l sm)

If A is a compact subset of D, then (from Section VIII.10) R" on D - A


is the PWB solution for the set D - A with boundary function equal to 0
on OD r a(D - A) and equal to v on A n a(D - A). Similarly R" on D - A
is the PWB solution for the set D - A with boundary function equal to 0
on aD n a(D - A) and equal to ti on A n a(D - A). In view of the relation
(Section 4) between the Dirichlet problems for harmonic and for parabolic
functions it follows that (5.1) is true for in D - A ; this equation is trivial
for 4 in A. If A is an analytic subset of D, then A is an analytic subset of b;
so [from Section XVII.16(p)] the value s) can be approximated arbi-
trarily closely from below by R' s) with E a compact subset of A ; if F
is the projection of f on the abscissa hyperplane, the approximation is
improved if F is replaced by F x R. Since F is a compact subset of D and
since [from Section VI.3(l)] can be approximated arbitrarily closely
by RB(A) with B a compact subset of A, it follows that (5.1) is true. Equations
(5.1) and (5.1 sm) are identical when ED - A and the left [right] sides of
these equations are equal up to a parabolic-semipolar subset of ,6 [parabolic-
polar subset of D]. Since a classical-polar subset of D is the projection of a
parabolic-polar subset of b (because a positive superharmonic function on
D identically + oo on a set B can be identified with a positive superparabolic
function on b identically +oo on B x R), equation (5.Ism) is true up to a
parabolic-semipolar subset of D and therefore is true on D.

Observation. There are three possible approaches to the properties of


reductions in the classical context.
(1) These properties can be proved using the specific classical context.
Such proofs were used in the preceding chapters. Unfortunately some of the
most natural proofs cannot be applied in the parabolic context because of
the weaker version of the Fundamental Convergence Theorem in the para-
bolic context and (a related fact) because the domination principle is false
in the parabolic context.
(2) The reduction properties in the parabolic context can be proved,
and then either it can be noted that these proofs are valid in the classical
context or
(3) 1t can be noted that in view of (5.1) and (5.1sm) the properties of
reductions in the classical context can usually be read off from those in the
parabolic context, or at least be deduced from them.
The choice (1) adopted in this book is inefficient and repetitious but was
made because the proofs in the classical context are thereby clearer and more
natural than the more generally usable proofs in choices (2) and (3).
6. Parabolic Regularity of Boundary Points 337

6. Parabolic Regularity of Boundary Points


In the following examples b is a nonempty open subset of d8' provided with
its Euclidean boundary. In this section barrier means "parabolic context
barrier." Since the existence of a weak barrier for b at a boundary point
is a property of b in a neighborhood of o, one can prove that a barrier
for b at o exists and thereby prove that o is parabolic regular by exhibiting
a weak barrier for the part of D in some neighborhood of o.

EXAMPLE (a). If every point of b in some neighborhood of the finite bound-


ary point o is either above the horizontal hyperplane through o or on one
side of some other hyperplane through 0, then 0 is a parabolic-regular
boundary point. In fact, to construct a local weak barrier for D at , define
0 by

where 4' = 0 is the equation of a hyperplane through o. If the hyperplane


is not parallel to the ordinate axis, it can be supposed that a = 1. Under this
hypothesis the half-space strictly above the hyperplane is the set {4, > 01,
and the restriction of 0 to the part of 1) in an open neighborhood of o is a
local weak barrier at o if every point of D in that neighborhood is above
the hyperplane. If every point of D in some neighborhood of o is below the
hyperplane and if a is a strictly positive number, the function 1i = 1 - exp a4'
is strictly positive strictly below the hyperplane and for sufficiently large a
is superparabolic there; so the restriction of ti to the part of D in some open
neighborhood of o is a local barrier at o. Finally, if the hyperplane is
parallel to the ordinate axis, then a = 0. If the part of D in some open
neighborhood of 0 is on one side of this hyperplane, then the restriction
of either 0 or - ¢ to this part of b is a local barrier at
Observe that if the complement of D includes the part in a neighborhood
of 90 of some not horizontal hyperplane through o, then if D contains points
arbitrarily near So on both sides of this hyperplane there is a local weak
barrier for the part of D on each side of the hyperplane and therefore a local
barrier for b itself at o. Such combinations will not be mentioned further.
Example (a) implies that the lower and lateral boundary points of an
interval are parabolic regular. Conversely, our discussion of the parabolic
Poisson integral for an interval can be used to derive the conclusions of
Example (a) involving hyperplanes parallel or perpendicular to the ordinate
axis.
EXAMPLE (b). If o : so) is a boundary point of D with the property that
the boundary in some neighborhood of So consists of points with ordinate
values 2!s0 and if D contains points arbitrarily near o with ordinate values
<so, then o is a parabolic-irregular boundary point of D. In fact (Section 1)
338 1 XVIII The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

some boundary neighborhood. of o has parabolic measure 0 relative to


points of D near but below o; therefore the boundary function values
assigned near o are irrelevant to the Dirichlet solution at points of b near
but below 0. For example, if N = I and if b is the interior of the simple
polygon with vertices (-1,0), (1,0), (1, 1), (0,#), and (-1, 1), the boundary
point (0, `-2) is parabolic irregular. Observe that for this choice of D there is
no increasing sequence of open sets with parabolic-regular boundaries and
with union D, but this lack of parabolic-regular approximation is only a
minor inconvenience. Theorem VIII.17 on the approximation of harmonic
measure by harmonic measures of boundaries of subsets translates directly
into the parabolic context.

EXAMPLE (c). [This example strengthens (a) except when the boundary
hyperplane in (a) is horizontal.] Suppose that o: is a boundary
point of D, and suppose that there is a ball of center : with S # o
such that the ball closure meets D in o but in no other point. Then o is a
parabolic-regular boundary point of D. To see this, let 6 be the ball radius,
and define u on D by

6-P - (6.1)

where p is a strictly positive number to be chosen below. Then ti > 0, u has


limit 0 at 50, to has a strictly positive lower bound outside each neighborhood
of ';o, and if, = (sl, t),
z

du(n) = pit - I-p-4 NZ (n -


X12

2
-
Then Oti < 0 sufficiently near 0 if p is large so there is a local barrier at 0.

EXAMPLE (d) (Iterated logarithm criterion for parabolic regularity). [This


example strengthens (a) when the boundary hyperplane in (a) is horizontal.]
Let I) be the set below the abscissa hyperplane defined by the inequalities

1
12 < 2a21sj logl log lsjj, -1 < s < O. (6.3)

We show that the origin is a parabolic-regular boundary point of D by


showing that the function c defined by D by
ISIIN+2 /2z Y
t'(S, s) = 110011 -' [1 - (S, S)],
(6.4)
I.sl-112eXp
-1 12
ti( ,s) =
gas
6. Parabolic Regularity of Boundary Points 339

is a barrier for b at the origin. The inequality (6.3) can be written in the form

s) < IsII log ISII .

which makes it clear that ti > 0 on b and that ti has limit 0 at the origin.
Furthermore, using the fact that ti is parabolic, we find that

bu(ss)= -N2 Slog Is1I-'ti(.,s)<0; (6.5)

so ti is a barrier.

Application. This example shows that the top point of a ball in 11N is
parabolic regular. The other boundary points of a ball are parabolic regular
according to Example (a).

EXAMPLE (e). (Iterated logarithm criterion for parabolic regularity, con-


tinued). Let D, be the set defined by the inequalities
112 < 2Q21s1cloglloglsi1, 0 > s > s', s' > -e-t. (6.6)

The value of s' will be chosen so near 0 that certain inequalities below will
be true. It will now be shown, in contrast with Example (d) in which c = 1,
that the origin is a parabolic-irregular boundary point of D, when c > 1.
Define the function ti on AN x ]s', 0[ by
2

u(,s) = Ilog Isll '-`exp ki li2 - (loglloglsII)-', (6.7)

where 0 < k < I and c > 0. Then

s) > I log ISI I ' ` - (logl log IS11)-' > 0

ifs' is sufficiently near 0. Furthermore

k14---12 Nk 1 +e
bu(ss) _ -1s1 'IloglsiI '-`exp
2
Iloglsll (6.8)

+ (k - k2)1SI2 + I loglsil -k1412


exp
2o2 1S1 2a21s1
(log l log lsll )2

Choose s' so near 0 that (I + c) I log ls' lI < Nk/2. The function u is then
superparabolic if
340 I.XVIII. The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

Nk < (k -
Ilog lsll ,
exp
-k.
2Q IsI
(6.9)
20 IsI (log I log IsII )z

This inequality will be satisfied if

Nk < (k - (6.10)
2a2Isl

On the other hand, if (6. 10) is not satisfied at s), the second term on the
right in (6.9) is at least

Iloglsll` exp -Nk


(log Iloglsll)z I -k'
which is at least Nk on D, ifs' is sufficiently near 0, depending on s and k.
Thus for any choice of a and k the function v is superparabolic and positive
on A when s' is sufficiently small. Now choose c = (I + e)/k, and define
the continuous function f on 0,6, as 1 at the origin and as ti at the other
boundary points. The PWB solution Ht on D, for f is the same as the PWB
solution for a boundary function differing from f only at a single point;
therefore Hf = Hf, when f =f, except at the origin, where f, is defined as 0.
Since ti is in the upper PWB class on D, for f,, we conclude that Hf s d
and therefore
lim Hf(0, s) 5 lim v(0, s) = 0.
s-'o S-0

Since f is continuous with value 1 at the origin, the origin is a parabolic-


irregular boundary point for D. Moreover c can be made arbitrarily near I
by choosing a near 0 and k near 1. It follows that for any choice of c > I and
any choice of s' the origin is a parabolic-irregular boundary point of 4.

Parabolic Balls

A parabolic ball in 9' of radius b and center (gyp, so) is defined as the set

6(so - (6.11)

Observe that the center of the parabolic ball is a boundary point. The set
(6.11) lies strictly below the center and is a solid of revolution with vertical
axis through the center. The highest and lowest points of the boundary are,
respectively, the center and (gyp, so - b). Example (a) shows that every
boundary point of a parabolic ball except possibly the center is parabolic
regular, and Example (c) shows that every boundary point of a parabolic
8. Sweeping in the Parabolic Context 341

ball except possibly the center and the lowest boundary point is parabolic
regular. The center is a parabolic-irregular boundary point according to
Example (e) because the set b, of that example is, for an arbitrary value of
c > I and an arbitrary 6 > 0, included in the parabolic ball of center the
origin and radius S when s' is sufficiently near 0.

7. Parabolic Regularity in Terms of the Fine Topology


Theorem. A finite Euclidean boundary point ( of an open subset D of pN is
parabolic regular if and only if C E (AN - D)p

This theorem is the parabolic context version of Theorem XI. 12, and the
proof is omitted because the only change needed in the proof of that theorem
to make it applicable in the present context is to change "polar" to "parabolic
semipolar."

EXAMPLE. If D is a parabolic ball (Section 6), the center is a parabolic-


irregular boundary point; so D is a deleted parabolic-fine neighborhood of
the center. Similarly the other examples of parabolic-regular and parabolic-
irregular boundary points in Section 6 have simple interpretations in the
parabolic-fine topology.

8. Sweeping in the Parabolic Context


The sweeping operation defined in the classical context in Section X.1
becomes a pair of linked operations in the present context. If b is a non-
empty open subset of ftN, A c D, and if GDµ [zGD] is a superparabolic
[cosuperparabolic] potential, the smoothed reduction %A_4 CuAAOA]
is also a potential, namely that of a "swept/measure" u0" [iµ0"]

Gvu6" = JAQ-4 = µ0"'Vp. (8.1)

The swept measures are supported by A n D, and it will be shown in Section


13 that if A is a Borel set, then Oµ0" [Oµ0"] is supported by the in general
smaller parabolic-fine [coparabolic-fine] closure of A in D. Since GojO"
is unaffected when A is changed by a parabolic-polar set, N 0" is also un-
affected by such a change in A. Dually Ou0A is unaffected when A is changed
by a coparabolic-polar set. It will be shown in Section 1 I that a set is parabolic
polar if and only if it is coparabolic polar.
The notation of Section X.1 is adapted to the present context by writing
So(5, ) for the probability measure on D supported by and by defining
bb and b6 by
342 1 XVII1. The Parabolic DirichletFt Problem, Sweeping, and Exceptional Sets

sp\>') _ dd(S.')DA, Vp(S,') = (8.2)

These swept measures are determined uniquely by

GD(', J)UA(S) _ xx r. 7))


,')DA(il) = p(S. VU(', (8.3)

and the reasoning in the classical case (Section X.1) shows that do 5 1 and
36<1.
The discussion proceeds as in the classical context, very slightly com-
plicated by the existence of two kernels instead of one, and the details are
left to the reader. It is first shown that if 1) is an open subset of b and if
is in 1), then

S'-B(S,') = (8.4)

on the Bore] subsets of ,6 n 8$. Next the classical context symmetry argument
is adapted, yielding here for an arbitrary subset A of D,

(8.5)

that is.

(8.6)

The common value in (8.5) will be denoted by il). It is then shown that
5 and 36 are kernels and that if r [ti] is a positive superparabolic [cosuper-

parabolic] function,
wQA
= D(', I ), Jd(., i). (8.7)

In particular, if fi is a measure on b,
fi
p` hA(',GDN) = GD.h A ` GDµ
(8.8)
NGDQA = 3-1(',µG0) = µ6o.

Moreover

(8.9)

More precisely the equations in (8.9) are correct according to our definitions
if the potentials involved are superparabolic or cosuperparabolic as the case
may be. if then Ei11 A and IµoA are defined by (8.9) for measures p on D for
which the swept measures are not already defined, every equation in (8.8)
is true. Finally a trivial integration yields
9. The Extension d, of do and the Parabolic Average da ii)) when b c h 343

JIDAdv = f +co (8.10)


D JD

first for a measure i' and a parabolic potential 6,6i and then by a limit
procedure for v and an arbitrary positive superparabolic function ti with
associated Riesz measure p. The formulation of the dual of (8.10) is left to
the reader.

Subadditivity of A F-+ b14

This subadditivity, and the dual subadditivity, are shown by a slight refine-
ment of the proof in the classical context (SectionX.6).

9. The Extension Go of Go and the Parabolic Average


GB (.,, )) when b c B
Boundaries in this section are Euclidean. Let b be a nonempty open subset
of AN, and let p be a point of D. We extend ryj) to tl§N to obtain a function
GD with the following properties:
(a) 6 J5 -
tj) is a positive subparabolic function on 9'
(b) C, i)) = 0 on D, and G(, tj) = 0 at a finite point of aD if
the point is parabolic regular.
Observe that such an extension must be unique because two such extensions
would be parabolic-fine continuous and equal up to a parabolic-fine nowhere
dense set and therefore would be identical. To define recall (from
Section XVII.4) that the function is the restriction to D of the
limit n) of a decreasing sequence n F--+ ie. igaG(, q) of superparabolic
functions on 9N . each set Bk is an interval with closure in D. Then ti(-, PI)
= G(-, il) on A"' - D, rj) = j) up to a parabolic-semipolar set, ti(-, rl)
is superparabolic on R', and we define Gp p) on AN as G(, j) - ti(, ry').
According to the parabolic context Fundamental Convergence Theorem,
the function tj) has the stated properties (a) and (b) except possibly the
vanishing at the set of parabolic-regular boundary points of D. If is a
parabolic-regular boundary point of b, the function q) has limit 0 at r
on approach along b in view of the evaluation of a Green function in terms
of a Dirichlet solution. The function GD (, tj) has limit 0 at on approach
along UA' - D if a parabolic-semipolar set is excluded. Thus the parabolic-
fine continuous function GD (%q) has limit 0 at on approach along a
parabolic-fine dense set, so Gp (, il) = 0. Dually, for in D there
is an extension GD (S, ) of GD(S, ), positive and cosubparabolic on
III' - vanishing on A' - b and vanishing at a finite boundary
point of D if the point is coparabolic regular. Thus we have extended
6D to 0 x AN) v (RN x D).
344 1 XVIII The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

We can now state the parabolic context version of Theorem VIII.18.


Observe that there is no exceptional value of N in this context because O,, is
positive for all values of N.

Theorem. Let D and B be nonempty open subsets of A' with 1) a B.


(a) For , e B [ a B] the function O (', q)lari [GB -)job], defined as 0
on aB n aD, is a [co] parabolic-resolutive boundary function.
(b) The function G, has the representations

Ga ( , q) = q) - i<ti(S, Ge (', q)) [(4, q) e b x A], (9.1)

Gn(S. il) = q) - GB (,')) [(4.7)e B X M. (9.1")

(c) If c = G, i, is a superparabolic potential on h and if v' is the projection


of v on the set of parabolic-irregular boundary points of b in B, then

C;Moc = ib(', 6l i) + G1 v'. (9.2)

If b is relatively compact in h, then (9.2) is true whenever a is a


superparabolic function on h and v is its associated Riesz measure.

We leave to the reader the dual statement of (c), for cosuperparabolic


functions.
Observation (1). If and q are points of D, Theorem 9(b) yields the
important symmetry relation

Ga( ) I a) (9.3)

in which we can take B = 9" if desired.


Observation (2). If the write µo(, )la and Ab(q, )ia for the respective
restrictions of the indicated measures to the class of Borel subsets of h,
then

6 (', 0) = (n) < + x q) e D x h], (9.4)

i o(q, C [GaNo(q, -)JAI (6 < + o0 x D]; (9.4*)

so the respective cosuperparabolic and superparabolic potentials on the


right are finite valued. According to Theorem 9(b), the measure on A - } }
[@8 - {q}] associated with the cosuperparabolic [superparabolic] function
on this set is the restriction of the Borel measure
to the class of Borel subsets of ((8" - [ftN - {q}].

Proof of (a). Since Euclidean boundaries are parabolic resolutive, the para-
bolic resolutivity of the boundary function f;,, defined as GB q) on h n aD
10. Conditions that e 4,f 345

and as 0 elsewhere on eD, reduces to the j integrability of this boundary


function. If qe B - aD, the function f,, is bounded and so is p,, integrable. If
q e h n ab, then G,&, q)16 is in the upper PWB class on D for fj ; so again
((see Section VIII. 10) for the justification in the classical context) f, is pD
integrable. The coparabolic resolutivity of the other boundary function in
(a) is derived by a dual argument.

Proof of (b). Let be in D. If it is also in b, equation (9.1) reduces to the


evaluation of Op in terms of OB already derived in Section 1. If ry c -B* - D,
the left side of (9.1) vanishes, and f, is a bounded function on aD with PWB
solution q)ia. Hence (9.1) is true for q in h - D and therefore for ry' in
h - ab. Next suppose that q is a coparabolic-regular boundary point of D
in h, so that (9.1) reduces to

(9.5)

or equivalently (see Section VIII. 10 for the expression of a Dirichlet solution


in terms of a smoothed reduction in the classical context; no change is needed
in the present context) in terms of reductions relative to h,

(9.6)

Furthermore on h - b up to a coparabolic-semipolar
set. Since ) is coparabolic-fine continuous and since a coparabolic-
semipolar set is coparabolic-fine nowhere dense, we conclude that (9.6) is
true if q is a coparabolic-fine limit point of B - D, as q is because q is co-
parabolic regular. We have now proved that for fixed equation (9.1) is true
except possibly when q is in the coparabolic-fine nowhere dense set of
coparabolic-irregular boundary points of b in B. Since both sides of this
equation define coparabolic-fine continuous functions of il, equation (9.1)
is true for all q. Equation (9.1 *) can be proved by the dual argument or can
be reduced to (9.1) by a reflection of 11N in the abscissa hyperplane.

Proof of (c). See the proof of Theorem VII1.18(c).

The application of Theorem VIII.18 to the vanishing of h-potentials at


the boundaries of their domains, as detailed in Section VIII.18, goes over
into the present context with no change.

10. Conditions that E.4Pf


The following is the parabolic context counterpart of Theorem X1.3. Let b
be a nonempty open subset of C:8", let be a point of D, and let ti be a positive
superparabolic function on D. All reductions will be relative to D. We shall
write lima j, to mean the limit as h, a neighborhood of , shrinks to c.
346 1 XVIII The Parabolic Dinchlet Problem, Sweeping, and Exceptional Sets

Theorem. If E *f, then


(a) v °v8() = v(5) for every neighborhood h of ,

(b) D
{}) = 1,
(c) GD (, ') _ v( ).
If 0,41f, then
(a') urn B1
ijQA^B
= 0 on D - { }, and the limit is 0 at if 6(6 < + oo,
(b') 6p(,{})=0,
(c') C (,) # Gp( ) on the .vet {(;D( ) > 0}.

Theorem 10(a) is already known (Section XVI1.16(b)) and implies the


truth of Theorem 10(b) and (c). In proving assertions (a) and (c') we can
[Section XVII.9(b)] enlarge A to be open but still parabolic-thin at ; so it
is sufficient to prove these assertions for A open. Under this condition on
A the proof of Theorem XI.3(a')-(c'), the classical counterpart of Theorem
10(a')-(c'), gives the latter when translated into the parabolic context. Thus
there remains only the proof of Theorem 10(b') without the hypothesis that
A is open. If we take v = I and apply Theorem 10(a'), we find

0 = lim l [''^B() = lim by^B(, (10.1)


B4 B44 )4 4

Now in the classical context we have proved [Section X.6(a)] that the set
function SD is subadditive on the class of subsets of D, and this led to a proof
that increases when the neighborhood B of decreases. This
reasoning will now be refined to the parabolic context. In the first place
recall that [Section XVI1.16(j)] if A and h are open subsets of D and if v;
is a finite-valued continuous superparabolic function on D, then
"B + A-h =
66jA n 1'i[B IUi[B, (10.2)

so if v1 < r2,

O.1vB <
A+ B - JU1 B. (10.3)

In view of approximation of smoothed reductions on sets by reductions on


open supersets (10.3) must be valid for arbitrary subsets A and B of D,
equivalently,
bn

Finally the counterpart of the classical context reasoning in Section


X.6(a) shows that this inequality implies the subadditivity inequality
by"B 5 by + Jp. If the pair A. h is replaced here by the pair 4 n B,
1 1. Parabolic- and Coparabolic-Polar Sets 347

A - h and if $ isza neighborhood of , we find that

5 (c , {t;}) S aDnB(L, {t;}) + d 8(S, {4}) = pn8(S,


/SS jjY
{ }),

where we have used the fact that the measure is supported by the
closure of A - b. Thus SD^B( in (10.1) increases as $ decreases; so
0, and the proof of Theorem 10(b') is now complete.

11. Parabolic- and Coparabolic-Polar Sets


Theorem. The following six conditions on a subset A of a nonempty open
subset D of R" are equivalent :
(a) A is parabolic polar. (a*) A is coparabolic polar.
(b) SD vanishes identically. (b*) 3D vanishes identically.
(c) [(c*)] (If A is Borel) the only measure v on b supported by A and
with (DV [v(;D] bounded is the null measure.

Observation. In connection with (c) and (c*) recall (from Section XVII.8)
that a parabolic-polar set has a parabolic-polar Borel superset; the dual
assertion for coparabolic-polar sets follows trivially. In view of this observa-
tion it is no restriction on generality in the following proof to assume that
A is a Borel set. Furthermore, if q is a point of A', the superparabolic
potential O(-, q) is the finite-valued potential on dl' of a nonnull measure
supported by the polar singleton {ry'}. Hence "bounded" in (c) and (c*)
cannot be replaced by "finite-valued."

Proof. (a) -(a*) If A is parabolic polar, then (from Section XVII.8)


q) = 0 for every point q in D; so OD (, ) = 0 for every point in b,
and therefore A is coparabolic polar. The dual argument yields the reverse
implication.
(a), (a*)-(b), (b*) The evaluation (8.3) of SodD and 3466 shows that
(b) and (b*) are equivalent to each other and to the pair (a), (a*).
(a), (a*)-(c), (c*) If ,4 is a parabolic-polar and so also a coparabolic-
polar Borel set, if v is a measure on D supported by A, if (iDv is a bounded
potential, and if AtD is the cosuperparabolic potential of a finite measure
and is identically + oo on A, then

+oo > f 'i0vdA = f ACDdv = +oo


D D

unless i = 0. Hence (a) (c). On the other hand, if A is not parabolic polar,
the superparabolic function j 1 " does not vanish identically; so [from
Section XVII.16(p)] the set 4 has a compact subset h for which the function
348 I.XV111. The Parabolic Dinchlet Problem, Sweeping, and Exceptional Sets

1 [s does not vanish identically. This function is a bounded superparabolic


potential with associated Riesz measure supported by A. Since (c) implies
that this potential vanishes identically, (c) (a). Thus (a) p(c); so dually
(a*) p (c*), and since (a) = (as), the proof of Theorem 11 is complete. 0

Application to the Projection of a Parabolic-Polar Set

According to Section XVII.8, Example (a), if A is a classical-polar subset


of 68", then A x68 is a parabolic-polar subset of G`8". Conversely, we now
prove that if A = A x R is a parabolic-polar subset of f8", then A is a
classical-polar subset of P". It is sufficient to prove this assertion for A
bounded and A a G° set, in which case A is also a Ga set. Hence (from the
application in Section VI.2) if A is not polar and if D is a Greenian superset
of A, there is a measure tt on D supported by A and making G0µ strictly
positive bounded and continuous. However (from Section XVII.18), G0µ
considered as a function on D = D x P is a bounded superparabolic poten-
tial, GDµ = GDp with µ = (p x /,)/a" supported by A. Hence A is not
parabolic polar, contrary to hypothesis.

12. Parabolic- and Coparabolic-Semipolar Sets


Theorem. The following four conditions on a subset A of a nonempty open
subset D of A8" are equivalent :
(a) A is parabolic semipolar. (a*) ,4 is coparabolic semipolar.
(b) [(b*)] (If A is Borel) the only measure v on D supported by A and
with GDv [vOD] finite valued and continuous is the null measure.

Observation (a). In connection with (b) and (b*) recall (from Section
XVII.15) that a parabolic-semipolar set has a parabolic-semipolar Borel
superset: the dual assertion for coparabolic-semipolar sets follows trivially.
In view of this observation it is no restriction on generality in the following
proof to assume that A is a Borel set.
Observation (b). As the following proof shows, "finite-valued" in (b)
and (b*) can be replaced by "bounded."
Observation (c). Condition (b) of the theorem implies that A is IN,, null
because if A is a bounded Borel /"+, nonnull subset of 9" and if v is the
projection of l"+, on A, then (by Theorem XVII.6) the potential GDv is
bounded and continuous.

Proof. (a) =::-(b*) Let v bea measure on Dsupported by a compact parabolic-


semipolar subset 4 of b, and suppose that vOD is finite valued and contin-
12. Parabolic- and Coparabolic-Semipolar Sets 349

uous. Recall that a set is exceptional for the parabolic context Fundamental
Convergence Theorem in the sense of the converse statement of Theorem
XVI1.13 if and only if the set is parabolic semipolar. Now the proof of the
classical Fundamental Convergence Theorem (Theorem VI.1) when trans-
lated into the present context shows that a measure v supported by a compact
subset A of D is necessarily the null measure if (1) A is exceptional for the
parabolic context Fundamental Convergence Theorem and if (2) vOD is
finite valued and continuous. Hence v is null in the present context. More
generally suppose that i is a measure supported by a Borel parabolic-
semipolar subset A of D, with VGp finite valued and continuous, and let Ao
be a compact subset of A. If vo is the projection of v on Ao, then doO, and
(v - vo)G, are finite valued and lower semicontinuous with a continuous
sum; so VIGD is finite valued and continuous, and we have just proved that
therefore ivo is null. Since this is true for all A0, we conclude that v is null, as
was to be proved.
(b*) (a*) It is sufficient to prove that a Borel subset A of b is co-
parabolic semipolar if every finite-valued continuous potential id,, of a
measure v supported by A vanishes identically. Actually we shall only use
this implication when vCo is bounded and continuous. Let * u = *µGp be
a bounded continuous cosuperparabolic potential satisfying the dual prop-
erty of that satisfied by ti* in Section XVII.16(r). The cosuperparabolic
potential *p0"Go is lower semicontinuous, < *ti, and continuous
at every point where there is equality, in particular, at every point of A°f n A
If µ, is the projection of * µ 0 ' on a compact subset A, of A°'f n A, the
potentials µ,Gp and µ,)G, are lower semicontinuous with con-
tinuous sum at each point of A,, and so both are continuous at such a
point. Since µ, GD is coparabolic and therefore continuous on b - A, , it
follows that µ,G6 is bounded and continuous on D; so (b*) implies that
µ, = 0. Thus vanishes on compact subsets of 4°'f n A and therefore
vanishes on this set; that is,

SD(4, A°*f n A) *µ(d4) = 0.


f6

Thus the positive superparabolic function 4°'f n,4) vanishes u almost


everywhere on D, certainly on a dense subset of 15, and therefore vanishes
identically on D. It follows (by Theorem 11) that the set A°f n'4 is parabolic
polar and so coparabolic polar, and therefore the union A of the coparabolic-
polar set 4"f n,4 and the coparabolic-semipolar set A - A°*f is coparabolic
semipolar, as was to be proved.
(a*) (b) _ (a) These implications are dual to the already proved
implications (a) = (b*) (a*) and are therefore true. The set of these
implications implies the truth of Theorem 12.
350 I.XVIII. The Parabolic Dinchlet Problem, Sweeping, and Exceptional Sets

13. The Support of a Swept Measure


Let 1) = A", and let A be the abscissa hyperplane. Then A°J = 4°'f = 0,
and the measures So( , ) and are supported by A. The situation in
the classical context (Section XI.14 and XI.18) is quite different in that the
sweeping operation is self-dual and that li (g, ) is supported by A°f. Further-
more, although A°f and 4°'f are Borel sets, the sets A u Aaf and A u A°'f,
which are involved in Theorem 13 below and its dual, are not necessarily
Borel sets. For example, if A is an arbitrary subset of the abscissa hyperplane
of AN, then ,4 u Aof = A u A°'f = A. For this reason in the following counter-
part of Theorem XI.14 the set A is supposed Borel. Recall that if 6
the measure associated with u) = BtiD" is not Qµ0" but

Theorem. Let D be a nonempty open subset of A', let A be a Borel subset


of b, and let be a point of D.
(a) I f e*f, then {i }) = I.
(b) If is a measure on D, then the measure Oft" [in particular, the
measure 5 , )] is supported by (A u A"f) n D.
(c) If A is parabolic-fine dense in itself and if N is supported by A, then
µJµ"
Proof. (a) If eAAf, then c5 (, I according to Theorem 10(b).
(b) Suppose that v and v' are positive superparabolic functions on D
and that v' < v, with equality on A. Then

A
° = = 6) = o'): (13.1)

so for every point in D the measure b is supported by the set {ti = v'}.
Now let v be a decreasing sequence of positive superparabolic functions
on D chosen (Choquet topological lemma) so that limn..,o v = v. has
smoothing DvQ". After replacing v by v A v if necessary, we can suppose
that v < z with equality on A. It follows that the measure is supported
by {v = v) for all n and therefore is supported by the intersection (ti,, = v)
of these sets. Thus S is supported by the set A u [{v. = v) n (b - A)].
Suppose that C is a compact subset of D - A. According to Section
XVII.16(l), the sequence v, can be chosen in such a way that um = ivO" on
C. Hence, if S is fixed, the projection on D - A of the measure is
supported by the set ti} n (D - A). Finally, if v is chosen as the
function ti' defined in Section XVII. 16(r), it follows that 8",(, ) is supported
by (A u A°f) n D. The evaluation of in (8.9) shows that this swept
measure is also supported by (A u A"') n 1).
(c) Follows trivially from Theorem 13(b).
14. Internal Limit Theorem; Smoothness of Superparabolic Functions 351

14. An Internal Limit Theorem; The Coparabolic-Fine


Topology Smoothness of Superparabolic Functions
According to Sections II and 12 a subset of C'8" is, respectively, coparabolic
polar or coparabolic semipolar if and only if it is parabolic polar or parabolic
semipolar; we shall write "parabolic" in both cases from now on. This
self-duality is less deep than the relations between parabolic and coparabolic
concepts displayed in the following parabolic context counterpart of
Theorem XI.4. If b is an open subset of ll' and if i and It are positive super-
parabolic functions on b, the function ti = i/h is defined in the obvious way
when the ratio is neither 0/0 nor + oo/+ oo. The function ti is thereby defined
parabolic quasi everywhere on the strict positivity set of h and is also defined
on the i strict positivity subset of the h zero set. We shall see that it is import-
ant to allow h to vanish. Let v,; and vg be respectively the Riesz measures
associated with i and h. The singular component of v6 relative to vi, will be
denoted by i,6, in particular, vL = v,; if h is parabolic. If 0 is a function defined
on a subset of D and if e D, def ine 0*(C) = p*f lim,i_ O(q) if this limit
exists.

The Zero Set of h


If is in the zero set Z of h, then every point of b below relative to b is
also in Z, and in particular, if it is the horizontal hyperplane containing C,
the set Z contains the open in it connected component of it n b containing
C. It follows that the trace in D of the Euclidean boundary of Z consists of
countably many such components and that if C is a point of such a boundary
component, h > 0 on the part strictly above C of a sufficiently small Euclidean
neighborhood of C.

Theorem. (a) The function u* is defined


(al) vc + vi almost everywhere on ,6 and
(a2) Parabolic quasi everywhere on the strict positivity set of i + lt.
(b) u* < + oo vi, almost everywhere on D and also parabolic quasi every-
where on the strict positivity set of h.
(c) Let v,;i and vii be the respective projections of i6 and vti on the zero
set Z of h. Then at vti almost every point of Z,
u« = dv, =
(14.1)
dvi, dvi,i

(d) Let t be a parabolic-polar subset of D. Then


(dl) (14.1) is true Ji, almost everywhere on t.
(d2) h * = + oc vk almost everywhere on F.
(d3) 0 = + oo 0 almost everywhere on P.
352 I.XVIII The Parabolic Dinchlet Problem. Sweeping, and Exceptional Sets

(e) The function u* is coparabolic fuse continuous, and the set where
u : ci is parabolic semipolar.
(f) In particular, if (E D,

p*flim v(n) = inf v(n) (14.2)


a--c C) 16:anc4.6>01 C)

(g) Let be a point of D, let F be a subset of ,6 with Euclidean topology


limit point S, and suppose that ord it > ord C when l a F. Then if t is copara-
bolic thin at C, there is a superparabolic potential v on D, strictly positive on
a Euclidean neighborhood of C, for which

FlimcG°(n)
_ +oo. (14.3)
C)

Conversely, if there is a positive superparabolic function v on D, satisfying


(14.3), the set P is coparabolic thin at C.

Observation. Since ti and h are positive superparabolic functions but are


otherwise unrestricted, (d2) implies that v* = + oo v,; almost everywhere on
F. Other obvious implications of Theorem 14 can be obtained by trivial
manipulations. For example, (b) implies, after an interchange of v and h,
that u* > 0 v,, almost everywhere on D. Note that (d2) is the parabolic
context counterpart of Theorem V.11.

Proof of (f). Assertion (f) is a special case both of (c) and (dl). Its direct
proof is a translation of that of Theorem XI.4(c), but we give the translation
because a direct proof is so much easier than the proof of the general case
and because the direct proof illustrates the adaptation of the Theorem XI.4(c)
proof technique to the present context. The equality of the first and second
terms on the right side of (14.2) was pointed out in Section XVII.8. It can
be supposed, replacing v by v - 6,;({C })GD(, C), that v,;((C) ) = 0. Under
this condition, unless (14.2) is true, there is a strictly positive number b
such that the set h = 10: v(ri) > bGD(rl,C)} is not coparabolic thin at C.
Apply the dual version of Theorem 10 to obtain

v(,) >_ v (q) 2- bGv(, )OB(q) = bGa(1, )l B() = bOD(f, t).

contrary to the hypothesis that 616,6(-, 1) has infimum 0.

Proof of (g). The proof is the counterpart of that of Theorem XI.4(d) and
is omitted.

Proof of (al). If 0 < a < b, define

A = {6< ah}, B = {v z bh}, G = .4p*f n $°*f n 6, , (14.4)


14 Internal Limit Theorem; Smoothness of Superparabolic Functions 353

and denote by v,,t the projection of vw on e. Following the proof of Theorem


XI.4(a), we find that in the reduction notation of Section XVII.16(s) as
translated from the classical context [see Section VI.3(o)] the smoothed
reductions h,;g, h,iaAa, ... all majorize ODvk. According to the parabolic
context version of VI(3.12), the sum of these smoothed reductions is at most
[e A (bh)]/(b - a); so v. = 0, that is, v,,(O) = 0. Now according to the
analysis of the zero set Z of h given at the beginning of this section, a point
of D must be either a Euclidean interior point of k or a coparabolic-fine
interior point of b - Z. The Euclidean interior of k is vi, null and every
coparabolic-fine interior point of D - 2 at which 0 does not exist is in e
for some rational pair a, b. It follows that t exists v,, almost everywhere
on D. Apply this result to h/c to find that u* also exists v,; almost everywhere
on D.

Proof of (a2). In the context of Theorem XI.4 the fact that u* exists quasi
everywhere on D is a triviality, and in fact u* = u quasi everywhere on D,
but in the present context assertion (a2) that u* exists parabolic quasi every-
where on the strict positivity set of h + v is by no means trivial. Since (a2)
is a local assertion, we can assume that h is bounded below on D by a strictly
positive number fl. Let ho = C'DVtio be a potential majorized by fi, with mea-
sure i'Fo supported by the set e defined in (14.4). Then

hOAA+hoABAB+ ... <hAB+hAB.IB+ ... <.. A (bh) +

b-a
and the reasoning showing that each summand of the second series majorizes
Ovvbc can be applied to ho and shows that each summand of the first series
majorizes dbv4.e, where vtioe is the projection of iota on e. Hence vho(C) = 0;
that is, the zero measure is the only measure supported by a whose potential
is bounded. It follows (by Theorem 11) that the set a is parabolic polar for
all pairs a, b with 0 < a < b, and this implies that u* exists parabolic quasi
everywhere on the strict positivity set of h. Apply this result to h/ti to com-
plete the proof of (a2).

Proof of (b). Write Bb for h defined by (14.4) and observe that the infinity
set of u* is included in 11- = nb o Brf. If v1i8m is the projection of iv,, on
A., then

rz z bOD(vhSD) (14.5)

because pb(t;, I when e Brf. Hence vt,(B,o) = 0; so u* < +oo vti


almost everywhere on D. To show that u* < + oo parabolic quasi everywhere
on the strict positivity set of It, we can suppose, localizing the context, that
It has a strictly positive lower bound $ on D. Let ho = 6dvto be a potential
on b majorized by # with vti. supported by Then (14.5) with h replaced
354 1 XVIII The Parabolic Dinchiet Problem, Sweeping, and Exceptional Sets

by ho shows that vtia(B.) = 0; so the zero measure is the only measure


supported by Bm whose potential is bounded, and therefore B,, is parabolic
polar. Thus tit < + oo parabolic quasi everywhere on the strict positivity
set of h.

Proof of (c). Since the Euclidean interior of Z is vi, null, we need only consider
the Euclidean boundary of Z in proving (c). In view of the properties of
this boundary, discussed at the beginning of this section, it is sufficient to
prove that (14.1) is true v;, almost everywhere on an arbitrary compact
subset t of this boundary lying on a horizontal hyperplane. We can suppose
that a and h are potentials, after replacing these functions by their reductions
on an open neighborhood of t, relatively compact in D. We first prove that
if v,,(F) = 0, then zit = 0 v,, almost everywhere on F. Define B by (14.4). Then

6 > 6; B Z bjhfB = b6b(v,,AD) bcpi'h,

where is the projection of vi, on B°'f n F. Consider the subset Do of b


strictly above the hyperplane of F. The Green function 66. is the restriction
to Do x Do of O,, (Section XVII.4); therefore the restriction of ti to Do is a
potential, and this potential majorizes the restriction to Dn of b4,pvh. This
restriction is a parabolic function and so vh = 0, that is, vt,(BD'f n F) = 0.
So ptf lim sup,,-4 u(q) 5 b for v;, almost every point of F, and therefore
tit = 0 almost everywhere on t, as asserted. It follows from this result
that (h has coparabolic-fine limit 0 v;, almost everywhere on F
and a corresponding result holds for ti. Only the projections of the measures
on Z are relevant. The proof of the first equality in (14.1) now follows that
of XI(4.1) and is therefore omitted.

Proof of (d). Following the proof of Theorem XI.4(b), it is proved first that
if t is a Borel parabolic-polar subset of D and if 0, then tit = 0 ivi,
almost everywhere on F. Since the proof follows closely that of Theorem
XI.4(b), it is omitted. Apply this result to 1/h to find that (d2) is true, and
apply the same result to c/h on the trace of a vg null support of the projection
of 4 on t to find that (d3) is true. The proof of (d 1) follows that of Theorem
XI.4(b) and so is omitted.

Proof of (e). It is trivial from the definition of tit that this function is copara-
bolic-fine continuous. To prove that zit = ti up to a parabolic-semipolar set,
it is sufficient to prove that if E > 0, the set

F = (q: zit is defined at ,, larctan ti(q) - arctan zit(zj)l > E}

is parabolic semipolar. Since (from Section XVI1.15) F- F°'f is parabolic


semipolar, it is sufficient to show that F" is parabolic-semipolar. Actually,
if zit is defined at i;, then
14. Internal Limit Theorem. Smoothness of Superparabolic Functions 355

p*flimiarctan ti(ry') - arctan ti'(tj)j = 0


'I-K

so a cannot be in E " ". Hence t " is even parabolic polar so t is parabolic


semipolar.

Application to the Fundamental Convergence Theorem, Reductions, and


the Fine Topologies

(a) Application to the Fundamental Convergence Theorem. In that theorem


(Section XVII.13) a locally lower bounded family f' of superparabolic func-
tions on an open subset b of 68" is given, with pointwise infimum ti. Accord-
ing to that theorem, ti has a parabolic-fine limit at every point of D,

ti(p) = liminfu(ry') = pflimu(ry'), (14.6)


o-t d-f

and a = 6 up to a parabolic-semipolar set. We shall now add to this conclu-


sion by proving that, in the notation of this section, u' is defined and equal
to u' parabolic quasi everywhere on b and that ti = u = ti up to a parabolic-
semipolar set. Suppose first that r is a decreasing sequence ti.. According
to Theorem 14 there is a parabolic-semipolar set A such that the coparabolic
limit function tip exists and is equal to ti for all n at every point of D - A.
The function u is therefore coparabolic-fine upper semicontinuous at each
point of D - A, so for 4 in D less some parabolic-semipolar set

p* f lim inf ti(il) < p*fliim supti(rl) 5 tu(b).

Thus up to a coparabolic-semipolar subset of D the function u' exists and


is equal to ti. To prove that u* exists parabolic quasi everywhere on D,
choose a, a', and b with 0 < a < a' < b, and define

A = {ti <- a}, A;, = {ti < a'}, B = {ti z b}, B = {ti z b}.

Since the problem is local and the functions are locally lower bounded, we
can suppose as usual that they are positive. Apply the reduction property
in Section XVII.16(s) to find

u A b (14.7)
lA;e + IA;,BA^B + ... < IA,B, + lA;B,,A;B,, + ... <
b - a'

Now A; is an increasing sequence of sets with union a superset A" of A.


It follows from repeated application of the reduction property in Section
XVII.16(e) that (14.7) yields (n - oo)
356 I.XVIII. The Parabolic Dinchlet Problem, Sweeping, and Exceptional Sets

I AB +1 ABAB
unb (14.8)
A B A BA B
b-a
The parabolic quasi everywhere existence of u* now follows as in the proof
of Theorem 14(a2). Further u < u ; so u < u' where both functions are
defined, that is, parabolic quasi everywhere on D. Since u = u = u' up to a
parabolic-semipolar set, the functions ri* and ti*, which are continuous
functions in the coparabolic-fine topology, are equal on a dense-in-this-
topology subset of their domains and so are equal where both are defined,
that is, are equal parabolic quasi everywhere. Thus the proof is complete
when f is a decreasing sequence. In the general case according to the Funda-
mental Convergence Theorem, more specifically according to Choquet's
topological lemma (Appendix VIII.3), there is a sequence u, in r with point-
wise infimum rig, such that ti = u. We can suppose that ti. is a decreasing
sequence, after replacing u by do A A u if necessary. Apply the result
for F a decreasing sequence to ti. to find that the inequality u 5 u S tu,o
implies

p*fliminfu(q) 5 p*flimsupti(q) 5 ri,*o( ) = 0(t)


R

for parabolic quasi every point of b; so u* is defined and equal to ;i*


parabolic quasi everywhere on D. Hence u = u* = u up to a parabolic-
semipolar set, and the proof of the application of Theorem 14 to the Funda-
mental Convergence Theorem is complete.
(b) Application to Reductions. Apply the preceding result to find that if
t is a positive superparabolic function on b and A is an arbitrary subset of
D, then R. ( ) = p*f limi_t Re(q) if is not in some parabolic-semipolar
set. Since RA = a on A, we conclude that RC = o` on 4''f less a parabolic-
semipolar set, and since o* = i up to a parabolic-semipolar set, we find that
R = c on A°'f less a parabolic-semipolar set. In contrast recall [from
Section XVII.16(b)] that this equality is true everywhere on 4°f.
(c) Application to the Fine Topologies. In particular, in application (b)
if D = AN and if t is the function u* with the properties listed in Section
XVII.16(r), we conclude that 4°'f c A°f up to a parabolic-semipolar set,
and since the dual result reverses the inclusion, it follows that 4°'f = A°f
up to a parabolic-semipolar set. Equivalently, the parabolic-fine and copara-
bolic-fine closures of A differ by a parabolic-semipolar set.
(d) Application to Parabolic-Fine Lower Semicontinuous Functions. Let f
be an arbitrary parabolic-fine lower semicontinuous funtion from an open
subset b of AN into R, and let f be the coparabolic-fine lower semicontinuous
smoothing of f, jo(b) = A f(q). We show that the set
15. Application to Fatou Boundary Limit Theorem on a Stab 357

(f > fo } is parabolic semipolar. It is sufficient to show that for arbitrary a in


R the set { fo < a < f } is parabolic semipolar. By definition of fo each point
of { fo < a} is in the complement of the coparabolic-fine interior h of the
parabolic-fine open set f f > a). Hence { fo < a < f } c (f > a} - $. The set
difference is parabolic semipolar according to application (c). In particular,
if f is a parabolic-fine continuous function, apply the preceding result to
both f and -f to show that the set of coparabolic-fine discontinuities of a
parabolic-fine continuous function is parabolic semipolar.

15. Application to a Version of the Parabolic Context Fatou


Boundary Limit Theorem on a Slab
Suppose that D = R' x ]0, b[ with 0 < S < + oo, and recall that GD is the
restriction of 6 to D x D. Let v and h be positive superparabolic functions
on D with respective Riesz measures v and v;,. If we extend v and h to func-
tions v' and h', respectively, by defining v' = h' = 0 on the closed lower
half-space of R, the extended functions are positive superparabolic func-
tions of D' = R" x ] - oo, S[. The measure v,, [vh] is the projection on D
of the Riesz measure v6, [vh.] associated with o' [h']. We have seen (Section
XVII.5 Example (b)) that v' and h' are potentials on D' and that if v [h]
is a potential on D, then the abscissa hyperplane is vd. [vh.] null. Denote
by Re [Nh] the projection of v6. [vh.] on the abscissa hyperplane so that the
parabolic component of v [h] in its Riesz decomposition is the restriction
of &&' [ORA] to D.

Theorem. With the preceding definitions

p*f lim Vi(n) =


dN
(C) (15.1)
a- h(ti) dNh
at Nh almost every point a of the abscissa hyperplane.

Observation. If h is a potential, we have already remarked that Nh is the


zero measure so the theorem is vacuous in this case. If a is a potential,
k6 is the zero measure so the limit in (15.1) is 0 at Nh almost every point of
the abscissa hyperplane, as would be expected. If both a and h are parabolic,
Theorem XVI.7 states that (15.1) is true for approach in the parabolic sense
as defined in Section XVI.7. Such a restatement of (15.1) in terms of a geo-
metrically simple approach to { is not possible in the general case, however.
For example, if N = 1 and h =- 1, a potential a can be defined for which
e/h does not have a limit at any point of the abscissa axis on approach
along the line normal to the axis.
To prove the theorem we need only apply Theorem 14(c) to the ratio
e'/h' on D'.
358 1 XVIII The Parabolic Dinchlet Problem, Sweeping, and Exceptional Sets

16. The Parabolic Context Domination Principle


As noted in Section XVII.5, the classical context domination principle is
not valid in the parabolic context. The following theorem, stated in the fine
limit notation of Theorem 14, is the parabolic version of Theorem XI.23,
the extended classical domination principle.
Theorem. Let h = Gov,, be a potential on an open subset D of and let u
be a positive superparabolic junction on D. Then h 5 ti on D if (v/h)* z 1
v, almost everywhere.

For the convenience of the reader we also state a weaker result, the
parabolic context version of Theorem V.10, the classical context domination
principle: h 5 r under each of the following three conditions.
(a) h* < + oo vh almost everywhere, and h* 5 ti* parabolic quasi every-
where on some Borel support of v,,.
(b) The inequalities h* < + oo and h* 5 v* are true v,, almost everywhere.
(c) Parabolic polar sets are vh null, and h* S ti* v,, almost everywhere.
Observation. Each of these three conditions implies the validity of the
condition of Theorem 16, namely, that (ti/h)* z 1 v, almost everywhere. Since
[by Theorem I4(d2)] h* = + oo v,; almost everywhere on each parabolic-
polar subset of A conditions (a) and (b) as well as (c) require (but Theorem
16 does not) that parabolic-polar sets are v,, null. A useful although stronger
condition than (c) is
(c') Parabolic-semipolar sets are vh null, and h 5 a vh almost everywhere.
This condition is stronger than (c) because iu = u4' and r = ti* up to parabolic-
semipolar sets.
The proof of Theorem 16 is merely a translation into the parabolic context
of the proof of Theorem XI.23. That is, if B is a Borel support of vh on which
(r/h)* is defined and z 1, and if 0 < c < 1, the set A = {ti z ch} is a deleted
coparabolic-fine neighborhood of every point of B; so h c A", and there-
for cfip(, { 5 }) = I when e B. Hence

U OVOA Z c CGD(vhOD) = Ch,

so t: _ h.

17. Limits of Superparabolic Functions at Parabolic-Irregular


Boundary Points of Their Domains
We sketch here the parabolic context versions of the classical context results
in Sections X1.21 and XI.22. The boundaries of subsets of A' are relative
to the Euclidean topology. The proofs are direct translations of the classical
context proofs and so are omitted.
17. Limits of Superparabobc Functions at Parabolic-Irregular Boundary Points 359

Let d be an open subset of j", let be a finite parabolic-irregular bound-


ary point of d, that is, t) is a parabolic-fine deleted neighborhood of ,
and let t be the class of open subsets of 6 which are parabolic-fine deleted
neighborhoods of C. The class t is ordered by inclusion. The basic result
for present purposes is that each lower-bounded superparabolic function u
on 6 has a parabolic-fine limit (S + oo) at {. This limit will be denoted by
°fu(C). It then follows that µD( , ) has a limit measure v'µo(C, ) in the sense
of vague convergence of measures on oD when tends to along some
deleted parabolic-fine neighborhood of C. The relation XI(22.1) between
fµD and fµ8 for B in r translates directly into the parabolic context. Further-
more, if A is a parabolic measure null subset of 8f), then A is also v'io(C, )
null. If ti is a lower-bounded superparabolic function on I), the function
h t into R is monotone decreasing with supremum
Here the obvious convention corresponding to that in the classical context
(Section XI.22) is made for the values assigned to ti on 8h n 8A The function
ui-+°Iti(b is lower semicontinuous on the space of positive superparabolic
functions on 15 in the topology of pointwise convergence.
In particular, if rq e b, the positive superparabolic function ry) has a
parabolic-fine limit at C; we denote this parabolic-fine limit by °ffD(g,i]).
Then

the function °fOD(C, ) is coparabolic on A

Martin Point Set Pairs

It will be useful to formulate a strengthened connectedness hypothesis


adapted to the present context. In view of the application to be made to
parabolic context Martin boundaries we shall call a pair (C,1)) consisting
of a point C of I'8' and an open subset '60f III" a Martin point set pair if the
following two conditions are satisfied:
MPS(1) The point { is a parabolic-irregular boundary point of A
MPS(2) If 1 E l), there is a deleted parabolic-fine neighborhood of C
whose points are all strictly above s; relative to 1).
Observe that MPS(2) implies that 1) is connected and that every point of
b is strictly below C. We now show under MPS(1) that MPS(2) is equivalent
to

MPS(2') °IOD(C, ) > 0 on !.


In fact under MPS(2) if 4E! the set a, of strict positivity of is a
deleted parabolic fine neighborhood of C. If DJt;D({, q) = 0 there is a smaller
deleted parabolic fine neighborhood . of e along which has limit 0,
and h can be chosen to be open because ((-, 4) is continuous. Then C is a
parabolic regular boundary point of A because q)iA is a weak parabolic
360 1.XVII1. The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

barrier for h at (. In view of the parabolic fine topology characterization


of parabolic regularity it follows that (e(d2" - B)Pf and therefore that
( e (W " - D)Pf. . Hence ( is a parabolic regular boundary point of D, contrary
to hypothesis. Thus MPS(2) MPS(2'). Conversely MPS(2') implies that
when h e b the function tj) is strictly positive on a deleted parabolic fine
neighborhood of ( and therefore (by the discussion in Section XVII.4 of the
strict positivity set of (D) implies MPS(2).

Application to the Zero Set of a Positive Superparabolic Function


[Martin Point Set Pair ((, ID)]

If ii is a positive superparabolic function on D and if Pfe (Z) = 0, then ti == 0.


In fact the zero set A of ti is closed in D, and e A implies by the super-
parabolic minimum theorem that Dr c A. Furthermore, if ti is defined on
b as ld_A, the function ti is superparabolic. Hence t has a parabolic-fine
limit at (. Since

0= it) = Pfa8(C, ti) for h in 1`,

it follows that Pfti(C) = 0, and therefore A is a deleted parabolic-fine neighbor-


hood of (; so by MPS(2) A = D, as was to be proved. This result is a gener-
alization of the superparabolic minimum theorem.

Application to the Harnack Inequality [Martin Point Set Pair (C,.6)]

To each compact subset A of D corresponds a constant c depending only


on (, 6, ,4 such that if ti is a positive parabolic function on D, then

maxti 5 cPfti((). (17.1)

In fact, if there is no such constant c, then there corresponds to each positive


integer n a positive parabolic function u" on D such that max,;ti" Z 1,
although P%(6:5 2-". Define u = Eo u". Then if a > 0, the function
r A a = lim"-m (Eo a,,,) A a is superparabolic on D with

Pf (u n a) = Pf lim (u n a) (%) < 1;

so pf lim4_4 ti(p) exists and is at most 1. The finiteness set of ti is therefore a


set including a deleted parabolic-fine neighborhood of {, and (Harnack
inequality applied to each function ti,) if ti(() < + oo, then u < +oo on D{.
Hence the condition MPS(2) implies that u is finite valued on b and so is
parabolic there. The convergence is locally uniform (Dini's theorem) so the
sequence u, converges uniformly to 0 on A, contrary to hypothesis. The
proof of this generalization of Harnack's inequality is now complete.
19. Lattices and Related Classes of Functions in the Parabolic Context 361

18. Martin Flat Point Set Pairs

Let I) be a nonempty open subset of A, and let C be a point of AN linked


to 6 by the following conditions :
MFPS(1) The part strictly below ( of some neighborhood of C is in 1).
MFPS(2) If e 15, there is a continuous function 0 from the interval
[0,1] into 1) u such that ord 0 is a strictly monotone
decreasing function with 0(0) = C and 0(1) = .
According to MFPS(2), every point of 1) is strictly below . Let h be an
interval containing e so small that every point of 1) strictly below a is in A,
and define b, = 1) u 1). According to Section XVII.9, every point of h with
the same ordinate value as C is a parabolic-irregular boundary point of 6.
It is now clear that is a Martin point set pair; we shall characterize
the pair as flat in view of the geometry of 1) near e. The following analysis
shows that the flatness trivializes the study of °fCD(e, -) and 'fuD(C, ).
Define h and 15, as in the preceding paragraph. According to Section
XV.17, a bounded parabolic function defined on h r 1) has a parabolic
extension to 1). Thus, if tj a 1), the superparabolic function d i&, n), which is
bounded and parabolic on 1) outside an arbitrary neighborhood of , has
a limit in the Euclidean topology at C, a limit already denoted by °fOD({, tq).
In the present context even this identification of the parabolic-fine limit
"ft;D(e,,') as a Euclidean topology limit can be simplified and made more
concrete because (from Section XVII.4) OD is the restriction to 1) x 1) of
dD, ; so 6D, is an extension of to 1S, . Moreover (Section XVII.4),
when e b, , the function GD, is strictly positive at the points of t),
strictly below relative to b:; so °f6D(e, ) = OD, (C, ) > 0 on 6. If j is a
bounded continuous function on at), the bounded parabolic function j)
on 6 has a parabolic extension to 1), according to Section XV.17; so the
measure tends to a limit measure °fgo(C, ) in the sense of vague
convergence when 4 -+ C in the Euclidean topology. More specifically,
according to Section 2, if A is a Borel subset of a1), the function A)
is the restriction to d of j D, A); so ofµD(C, A) = ,D, (C, A).

19. Lattices and Related Classes of Functions in the Parabolic


Context
The classes of functions defined in the classical context in Chapter IX have
parabolic potential theory counterparts. In the latter context we fix an
arbitrary nonempty open subset 1) of dl" (N;?: 1) and an arbitrary strictly
positive parabolic function h on ,6 and go on very nearly as in Chapter IX,
replacing "h-harmonic measure" by "h-parabolic measure" and so on. The
change involves a slight complication left to the reader to take into account
362 1.XVIII. The Parabolic Dirichlet Problem, Sweeping, and Exceptional Sets

as needed, due to the fact that h-parabolic measure relative to a point of b


vanishes on sets above the point.
The change needed to apply these concepts to the coparabolic context
will always be obvious.

EXAMPLE. The class L°(,4_) is the class of extended real-valued Borel mea-
surable functions on ,6 for which, if is in ,6 and if h, is an increasing sequence
of open relatively compact subsets of b with union b then 4 Itil°)
< + oo. This is the exact counterpart of the classical context definition in
Section IX.4, but observe that, in contrast to that context, even if D is
connected and 161° is h-subparabolic, the fact that this condition is satisfied
for one point of b does not imply that the condition is satisfied for every
point of D.
Chapter XIX

The Martin Boundary in the Parabolic Context

1. Introduction
In discussing the parabolic context Martin boundary of an open subset d of
ft' for N z I we first make the obvious remark that there are necessarily two
boundaries, one adapted to the operator d and superparabolic potentials,
the other adapted to the operator A and cosuperparabolic potentials. The
first is called the exit boundary; the second is called the entrance boundary.
These dual contexts are interchanged by a reflection of Et VV in the abscissa
hyperplane. We shall treat the exit boundary but shall omit the word "exit"
unless both boundaries are involved. The following remarks are offered to
orient the reader to the new features that arise in parabolic context Martin
boundary theory.
Let 1) be a nonempty open subset of kN. Throughout this chapter it will
denote the set of points of 1) strictly below 4 relative to d. Suppose that a
Martin function k based on a point i; of 6 is defined in the natural way,

q)
Ko, ) =
6D(c, 0)

Since (Section XVII.4) ( (C, ry') > 0 if and only if ,i e,64 the function K has
domain 6 x A. For q in di the function K(q, ) is superparabolic on 1),
parabolic on 1) - {ry'}, with value I at C. The classical Martin boundary treat-
ment suggests that the Martin boundary points to be introduced should
correspond to limit functions of the family {IC(q, ), rj a fit} when q in ,6t tends
to the Euclidean boundary 8b of b. This procedure will only yield Martin
boundary points of 6 below C; more precisely this procedure will yield a
boundary for I). A second new complication is the one-sided nature of
Harnack's parabolic context inequality, which is relied on to bound K(h, )
as n tends to a!. In fact this inequality, based on the normalization K(4, C) =
1, only bounds K(ry', ) strictly below C. Thus a limit function of K(, ) as
it - 8d cannot be defined at C so the normalization of such a function at C
requires a new formulation. This problem does not arise in the classical
context.
These difficulties will be at least partially surmounted in two ways, with-
364 I.XIX. The Martin Boundary in the Parabolic Context

out renouncing the Martin approach. One way is based on a Martin point
set pair, as defined in Section XVIII.17; the other way is based on a Martin
measure set pair, to be defined in the next section.

2. The Martin Functions of Martin Point Set and Measure Set


Pairs
The Martin Function of a Martin Point Set Pair

The Martin function k of a Martin point set pair (C,15) is defined on b x b


by

K(ii' 6 = o (2.1)
PdD(b' o

For fixed q the function K(q, -) is superparabolic on 15, parabolic on ,6 -


and has the normalization "K(r), a;) = 1, that is, K(ry', -) has parabolic-fine
limit I at . This normalization displays a limitation of the approach to
Martin boundaries by way of Martin point set pairs: every positive super-
parabolic function u to be considered on b will have to satisfy the condition
°fti(t) < + oo. In particular, the minimal parabolic functions on D and the
positive parabolic functions on b represented by the parabolic context
Martin representation will have to satisfy this condition.
If D' is an arbitrary nonempty open subset of A" and if e b', the pair
b) is a flat Martin point set pair to which Martin boundary theory based
on Martin point set pairs can be applied. In some cases we shall see that we
can then vary to obtain a Martin-type boundary for d itself and a Martin-
type representation of an arbitrary positive parabolic function on D'.

Martin Measure Set Pairs and Their Martin Functions

Let b be a nonempty open subset of ", and let v be a measure on D. Then


the pair (v, b) will be called a Martin measure set pair if the following two
conditions are satisfied :
MMS(1) The copotential v0D is finite valued and continuous on the
complement of a compact subset o, of 15.
MMS(2) If e b, there is a point strictly above relative to D such
that no neighborhood of ' is v null; that is, 4' is in the minimal
closed in ,6 support of v.
Observe that condition MMS(2) implies that v cannot have compact sup-
port in D. It is trivial that to every nonempty open subset ,6 of ft' corresponds
a measure v such that (v, D) is a Martin measure set pair, and in fact measures
2. The Martin Functions of Martin Point Set and Measure Set Pairs 365

can be chosen for which v( is finite valued and continuous on b (so Q,, = 0)
and D is itself the smallest closed in b support of v. Since C,,(4, ) > 0 on
AI condition MMS(2) implies that v( > 0 on D. The Martin function K
for the pair (iv, b) is defined on D x D by

K(q, ) _
D(' , q)
(2.2)
v,(

(= 0 at the infinities of For fixed q with id M) < +oo the function


K(q, -) is superparabolic on D, parabolic on D - {ry' }, and has the normaliza-
tion f DK(q, )dv =L This normalization displays a limitation of the approach
to Martin boundaries by way of measure set pairs: every positive super-
parabolic function ti to be considered on D will have to satisfy the condition
fti u dv < + oo. In particular, the minimal parabolic functions on D and the
positive parabolic functions on b represented by the parabolic context
Martin representation will have to satisfy this condition. Every positive
parabolic function on D satisfies this condition, however, for a suitable
choice of v, depending on the specified function. Observe that for a Martin
measure set pair (i,,6) the function u f v u dv from the space of positive
superparabolic functions on D, with the topology of pointwise convergence,
into A' is lower semicontinuous (Fatou's lemma), as is (Section XVIII.17)
the function u --. °fr (r) for ({, b) a Martin point set pair. The fact that there
is lower semicontinuity here rather than continuity as in the corresponding
classical context is a complicating feature of the parabolic context Martin
boundary constructions.
Although the approaches to the parabolic context Martin boundary by
way either of Martin point pairs or set pairs involve restrictions on the classes
of positive parabolic functions representable by the Martin representation,
in certain cases (for example, if the domain ,6 involved is a slab), we shall see
that the Martin representation leads to a representation of every positive
parabolic function on D.

Generalization of Martin Point and Measure Set Pairs

It is clear that if D is an arbitrary nonempty open subset of AN and if we


assign a measure v on the union of D with the set of parabolic-irregular
boundary points of b, and if we suppose that v has suitable properties, we
can combine the Martin boundary developments based on Martin point set
and measure set pairs. We shall not need this refinement, however.

Admissible Superparabolic Functions

If (C,,6) [(v,,6)] is a Martin point [measure] set pair we shall call a positive
superparabolic function u on 6 admissible if °Jti(C) < + oo [16 u dv < + J.
366 1.XIX. The Martin Boundary in the Parabolic Context

If h is an admissible strictly positive parabolic function on D, a function


ti/h on b will be called admissible if ti is the difference between two admissible
positive superparabolic functions on D.

3. The Martin Space DM


Theorem. If (c', b) [(v,,6)] is a Martin Point [measure] set pair there is a
unique up to homeomorphisms metrizable compactification DM of D with the
following properties:
(a) The Martin function k has an extension (also denoted by K) to
DM x D[(DM - o;,) x b] for which, when C e D, the function
is finite valued and continuous on DM - [D - o;].
(b) K(n t , )= IC(rl2 , ) if and only if n t = i12
The set 8MD = DM - D will be called the Martin boundary of (C', D) or
(v, b) as the case may be. This theorem implies (see the discussion in the
corresponding classical context of Theorem XII.3) that K(il, ) is a positive
parabolic function on D when i is a Martin boundary point. The proof
follows that of Theorem XII.3, but observe that DM is defined in terms of a
specified pair (S', D) or (v,,6) so k is uniquely determined. Thus ,6 itself has
not been assigned a Martin boundary. Let f be a strictly positive continuous
function on D, lv,.t integrable over D, and for h and ii2 in D define

d(i1 1 i2) = I [I A )I]l( )lx+t(d )


0

to obtain a metric on D compatible with the Euclidean topology of D. The


proof of the theorem then follows that of Theorem XII.3 [with a minor
variation for (i,,6) when 6i # 0] and will be omitted. The appropriate
versions of the Harnack inequality have been proved for (c', b) in Section
XVIII.17 and for (i,,6) in Section XV.11, and in view of these versions
C e a'D implies that IC(t , ) = K(C, ) locally uniformly on b and that
a sequence ry'. of points of b converges to a Martin boundary point C if and
only if ) = K(C, ) locally uniformly on D. The lower semi-
continuity property of our normalizations (Section 2) implies that for a in
amb

(C',D): -fk ,C') s I;


(3.1)
(v,D): JI(C.)dv 1.

We shall need more general inequalities. Suppose that A is a finite mea-


sure on DM. supported by DM - fr; in the (i,15) case. Then the function
4. Preparatory Material for the Parabolic Context Martin Representation Theorem 367

JpM K(C, )A(dC) is superparabolic on 1), parabolic if A is supported by


eMb, and we now show that

b): Pf(J K(C, )A(dC)) (0 rrg(C, C')A(dC) 5 , (bm)


= L.-
(3.2)

(v, b): f i(d4) k(C, C)A(dC) 5 . (,6m),


D J DM
with equality if and only if A is supported by the set of points C for which
there is equality in the corresponding inequality under (3.1). Inequality (3.2)
for (t, b), with equality under the stated condition, follows trivially from
(3.1) for (v,1)). To analyze (3.2) for (C',,6) observe (see Section XVIII.17,
but note that C there is replaced here by C') that if h is a Euclidean open sub-
set of t) with every Euclidean boundary point except C' in 1) and if h is a
parabolic-fine deleted neighborhood of C', then

PfNe (C', f k(C, -)A(dC)) = s z(bM). (3.3)


DM pM

If we order the class of sets 1} by inclusion, then (from Section XVIII.17)


when h shrinks to {, the first term of (3.3) increases to the first term of (3.2)
for (C', tl) and the integrand in the second term of (3.3) increases to Pf'(C, C').
Thus (3.2) for (C',,6) is true if integration of the second term of (3.3) to the
limit when h shrinks to C can be justified; on the latter point it need only be
pointed out that (Fatou's lemma) the function C' Pfµ$(C', K(C, )) is lower
semicontinuous. Finally, it is trivial that there is equality in (3.2) for (C', b)
if A is supported by the set of points C for which there is equality in (3.1) for
(C',1)).

The Zero Set of

Since (from Section XVII.4) i) > 0 if and only if it is strictly below


relative to 6, it follows that if 7, is a sequence of points in d with limit a
Martin boundary point C and if the numerical sequence ord, . has limit a,
then K(C. ) = 0 when ord C 5 at.

4. Preparatory Material for the Parabolic Context Martin


Representation Theorem
In this section we shall derive the parabolic context counterpart of Lemma
XII.4. The latter lemma was the key lemma in the derivation of the Martin
representation theorem, and it will be clear from the following proof that the
reasoning in the derivation of that theorem can be translated without diffi-
368 1 XIX. The Martin Boundary in the Parabolic Context

culty into the parabolic context. We shall therefore omit the proofs of the
counterparts of the other results in Sections XII.4 to XII.8 leading to the
Martin representation theorem, but we shall state the basic results in the
parabolic context for ease in later reference.

Lemma (Parabolic Counterpart of Lemma XII.4). Let (c','6) [(f,,,6)] be a


Martin point [measure] set pair with Martin function K. If ti is a positive
admissible parabolic function on ,6 and if A is a subset of b [with t n aA = 0]
there is a measure A, on DM, supported by the boundary of A relative to D"',
for which (reduction relative to b)

QuOA = (4.1)
DM

and

D): 1 uV (C'):
(4.2)

(v, D): AA(DM)=

d'
fib

In particular, there is a measure A,;(= A,;o) supported by aMD for which

,i = f IC(e, )2;,(dC) (4.3)


J BM0

and

(c', D): PJu( ');


(4.4)

(',D): Judy.

A measure ,, on aMD satisfying (4.3) must also satisfy

(4.5)
?MD

for every Bore! subset t of aMD.

We treat first the (z',15) case. If A is relatively compact in A then jtiD" is


the potential GDA' of a measure j.' supported by 0,4; so

Q A = ODA = J a(dd) (4.6)


r'A
5. Minimal Parabolic Functions and Their Poles 369

and by (3.2)

PfO, A(C") = fk(C, )A(dd) _ ,(aA).


P
f.4

If A is not relatively compact in b, let A be the intersection of A with the nth


member of an increasing sequence of relatively compact open subsets of
b with union D. According to what has just been proved, the potential
WA. can be represented in the form

(4.7)
f,5.1

where 1 is a measure supported by aA and

2n(bm) = PfNOAflg) 5 "fW (O- (4.8)

The sequence 2 is a bounded sequence of measures on the compact space


D". If A is the limit of a vaguely convergent subsequence, then 2 is supported
by the boundary of A relative to D", and (4.7) yields (4.1); equality (4.2) for
W, D) follows from (4.8) and the lower semicontinuity of the operator
ti -Pfv(e'), operating in this case on an increasing sequence of positive super-
parabolic functions. The proof that (4.3) implies (4.4) is carried through by a
direct translation of the corresponding implication proof in Theorem XII.4.
In the (i,D) case A(dd) in (4.6) is and we go on as in the
(r', D) case.

5. Minimal Parabolic Functions and Their Poles


Poles of Parabolic Functions

If ,6 is a nonempty open subset of 9" provided with a boundary by a metric


compactification, a boundary point is said to be a Pole of a positive para-
bolic function h on D if (reduction relative to D) Uh1(0 = h. If h is minimal
parabolic on D and is not identically 0, either is a pole of h or bhOul = 0
because the parabolic reduction operation on a boundary subset is idem-
potent. The elementary properties of poles in the classical context (Section
X.7) are valid in the present context with no change in derivation.

Minimal Martin Boundary Points

A Martin boundary point of ((,b) [(i,,,6)] will be called minimal if its


associated parabolic function k(C, ) is minimal with
370 I.XIX The Martin Boundary in the Parabolic Context

h=I f K(C, ') dv = (5.1)


I D 11

and the set of minimal boundary points will be denoted by 8;'D. Observe
that the counterpart of (5.1) in the classical context is trivially true for each
function associated with a Martin boundary point but that in the present
context one Martin boundary point, say Co, associated with the identically
vanishing function, is associated with a minimal function but is not a minimal
boundary point. According to the following counterpart of Theorem XII.5,
however, every minimal parabolic function which is admissible and not
identically 0 is a constant multiple of for some uniquely determined
minimal Martin boundary point C. The proof of this theorem follows that of
Theorem XII.5 and is omitted.

Theorem (Parabolic Counterpart of Theorem XII.5). Let (r', D) [(i,,6)] be


a Martin point [measure] set pair with Martin function K determining a Martin
boundary aMD.
(a) Every minimal parabolic function on ,6 has a pole on 8"'D. If a Martin
boundary point is the pole of a positive *0 admissible parabolic
function ti, then is the only pole of ti, ti = const K(C, ), and is a
minimal Martin boundary point. In particular, if ri is a minimal Martin
boundary point, the function k(,, ) has pole ry'.
(b) IfC is a minimal Martin boundary point and if A is a set of minimal
Martin boundary points, then is either K(C, ) or 0 according

6. The Set of Nonminimal Martin Boundary Points


We leave to the reader the easy formulation and proof of the parabolic
counterpart of Lemma XII.6. In the parabolic context Theorem XII.7
becomes the following.

Theorem. If b is a nonempty open subset of 98', then in the context of either


Martin point or Martin measure set pairs for D the set A of nonminimal
Martin boundary points is an F, set, and for every positive admissible parabolic
function h on b (reduction relative to D), hQ" 0. In particular, if 4 is strictly
positive the set A is h -parabolic measure null.

The proof follows that of Theorem XII.7.


We also leave to the reader the application of Theorem 6 to prove the
parabolic context counterpart of Lemma XII.8. The embarrassing possibility
in this context that h may not be strictly positive can be treated by proving
the desired result first for h + E with c > 0 and then letting a tend to 0.
8. Martin Boundary of a Slab b = R' x ]0, d[ with 0 < b 5 + co 371

7. The Martin Representation in the Parabolic Context


Let D be a nonempty open subset of d8". We shall use the notation corres-
ponding to that of the classical context Martin representation theorem
(Theorem XII.9). Thus "Sm refers to the class of differences v/h = (v1 - v2)/h,
where v, is a positive parabolic function on D and h is a strictly positive
parabolic function on D, omitted from the notation if h =- 1. The following
is a parabolic version of the classical Martin representation theorem.
Theorem. Let (C', D) [(i,,6)] be a Martin point [measure] set pair determining
a Martin function k and Martin boundary a"'D.
(a) To each admissible function v in!;. corresponds a unique finite-valued
signed measure MM on aMD, supported by the minimal Martin boundary
amD, positive if v is, and satisfying

v =f vfv(C') Jvdv].
aMb C=
(7.1)

(b) If h is a strictly positive admissible parabolic function on D, the


correspondence v/h .-. M,; is an isomorphism between the vector lattice
of admissible functions in 45,,, and the vector lattice of finite signed
measures on a",6 supported by aMD.
(c) An admissible function v/h in "5,,, is quasi bounded or singular if and
only if k6 is, respectively, absolutely continuous or singular relative
to A. In the quasi-bounded case

v= K(C, ) M;, Mi(dC) (7.2)


J aM»

The proof follows that of Theorem XII.9 with slight variations.

8. Martin Boundary of a Slab D = RN x ] 0, 8[ with


0<S<+oo
Boundary for a Martin Point Set Pair
Suppose first that b < + oo. Let C' be a point of R", and define b).
Then (C',,6) is a Martin point set pair, and if k is the Martin function for
the pair,

KQ,C)=
As-t,C-?1)
6n(C,0=6(s-t,C-q), tt(S - t, C' - q)'
=(C,s), n=(n,t).
372 I.XIX. The Martin Boundary in the Parabolic Context

The limit functions of k are as follows:


(a) If = ((, 0) is a point of the abscissa hyperplane,

lim K(n' ) =
G(s, D
4- 0101 0,
(b) If q = (n, t) e D and t b with no restriction on the varying of
'i or if 1?11 --+ + oo with no restriction on the varying of 1, then
lim,,. K(il, ) = 0.
The Martin space of (r', b) is therefore D compactified by adjunction of the
abscissa hyperplane and a single point 0: K((C, 0), 4) = e (s, - C/G(b, ' - C),
K(0,) = 0; n in D tends to (C, 0) in the Martin topology if and only if
ry' 0) in the Euclidean topology; q in b tends to 0 in the Martin topo-
logy under the conditions in (b). The point 0 is a nonminimal Martin boun-
dary point. It is trivial that every point or no point of the abscissa hyperplane
is a minimal Martin boundary point so every point of the abscissa hyperplane
is minimal. In the present context the Martin representation theorem states
that if ti is a positive parabolic function on D with 6) < + co, then
there is a unique finite measure Ma on the abscissa hyperplane such that

u() = J K(((, 0), )Mo(dC) (8.2)


RN

and

b) = MO(RN); (8.3)

that is [define Rj,(dd) = M (dd)/t (b, ' - c)], there is a unique measure 1V;,
on 1 N such that

u() = J 41(s, - [ _ (4,s)] (8.4)


RN

and

rfu(', b) = b(b, ' - )No(dd). (8.5)


J RN

Conversely, the Martin representation theorem states that if A, is a finite


measure on R' [if R. is a measure on RN making the integral in (8.5) finite],
then (8.2) [(8.4)] defines a positive parabolic function on D with °fti(c', b)
given by (8.3) [(8.5)].
Observe that for it determined in this way if ' is replaced by another point
8. Martin Boundary of a Slab b = R'`' x ]0, S[ with 0 < & 5 + oo 373

c", with a) < + oo, then k is changed but the representing measure
N;, in (8.4) is unchanged. Observe also that if N,; is an arbitrary measure on
R" making u as defined by (8.4) finite at points of D with ordinate values
arbitrarily near b, then ti is necessarily parabolic on D. Moreover we have
seen that (8.5) is then true if either side is known to be finite so (8.5) is true
in general.
Finally we show that if 0 < 8 5 + oo and if ti is an arbitrary positive
parabolic function on D = RN x ]0, S[, then there is a unique measure N;,
on RN for which (8.4) is true, and (8.5) is true for every point ' of R" when
d < + cc, and that conversely, if N;, is a measure on R" for which u as defined
by (8.4) is finite valued on D, then ti is a positive parabolic function on 1) and,
if b < + oo, (8.5) is true for every point ' of R". All that remains to be
proved is the existence and uniqueness of Na satisfying (8.4) when a is
specified. To prove this, choose S' < 6 and apply the preceding work to the
Martin point set pair ((c', S'), R x ]0, 6'[) with arbitrary '. The restriction
of 6 to R' x ]0, b'[ determines a unique measure N;, for which (8.4) is true
for s < 6', and the uniqueness of N;, shows that this measure depends neither
on ' nor on 6'. Hence (8.4) is valid for s < S as desired. Observe that we have
now found a new proof of the representation part of Theorem XVI.6(a) and
in addition have derived (8.5) in the positive case when b < + oo. As already
remarked in Section XVI.8, the representation (8.4) shows that a positive
parabolic function on b = I8" x ]0, S[ is minimal if and only if the function
is a positive multiple of the function s) r-- 41(s, - C) on D for some point
C of R".

Boundary for a Martin Measure Set Pair

If 0 < 6 5 + oc and if b = R" x ]0, 6[, choose a measure v for a Martin


measure set pair (v, D) with v supported by a set whose intersection with each
slab R' x ]0,6'[ with b' < 6 is compact. We can also suppose that v( is
cosuperparabolic on RN and therefore coparabolic on a neighborhood of the
abscissa hyperplane. Since vOD is the restriction to D of 'O, the function
V6ip has the strictly positive limit 6O(C) at every point e = (C, 0) of the abscissa
hyperplane. Hence the Martin function k for (v,,6) has a limit at C:

limK(n ,'c) _ t (s C - C) =
4-{

Furthermore the function k(q, ) tends to 0 when p' tends to any Euclidea
boundary point of 1) not on the abscissa hyperplane. Thus the Martin bound-
ary for (v, b) is the abscissa hyperplane and a point 0, as in the Martin point
set context. The class of positive parabolic functions on D given by the
Martin representation is now limited by the side condition ID a dv < + oo.
Just as in the Martin point set pair case, this form of the Martin representa-
374 I.XIX. The Martin Boundary in the Parabolic Context

tion can be applied to derive a representation of an arbitrary positive para-


bolic function on D.

9. Martin Boundaries for the Lower Half-space of A' and for


AN

Martin Boundaries for the Lower Half-space of 9

Let b be this lower half-space, and let ' = (c', 0) be a point of the abscissa
hyperplane. Then (c', D) is a Martin point set pair. According to Section
XVII.4, the Green function GD is the restriction of G to D x D. The possible
limit functions of K(q, ) when ry' - OD are the following [with tT _ (, s),
_ (n, 01:
(a) If t -. - or, and q/t -' - a2 y, then

limK(ry, c = exp < - y, '>. (9.1)

(b) If In I(1 - t)-' -' + oo, then /t(j,) 0.

The Martin boundary therefore consists of a point y corresponding to each


point y of 08", with K(y, ) given by the right side of (9.1), and a point 0 with
/t(0, ) - 0. Since the context here reduces to that in Section 8 under an
Appell transformation taking the lower into the upper half-plane, we go no
further except to remark that b has many parabolic irregular boundary
points but that the upper half-space has a quite different configuration,
which suggests the interest of a more careful analysis than we have given of
regularity at co in the parabolic context.

Martin Boundaries for U8'

Martin point set pairs cannot assign a boundary to Q;8", but if a measure v is
chosen on Ll to make (v, A") a Martin measure set pair (for example, if v is
chosen in such a way that vG is cosuperparabolic on d8", that v is supported
by a set whose intersection with the slab R' x ] - oo, a[ is bounded for each
a, and that the smallest closed support of v has points with arbitrarily large
ordinate values), then we obtain a Martin boundary for (v, A'). This bound-
ary consists of a point 0 corresponding to the identically vanishing parabolic
function and of points y obtained like the corresponding Martin boundary
points of the lower half-space obtained above.
It was shown in Section XVI.8 as an application of the Appell transforma-
tion to the known properties of parabolic functions on a slab that a positive
parabolic function on 9" is minimal if and only if it is a positive multiple of
a function -+exp(<y, > + a2lyl2s/2) and that a positive parabolic
10. The Martin Boundary of t5 = ]0, + oo [ x ] - oo, b[ 375

function on @8" can be represented in a unique way as an integral XVI(8.2)


over the class of minimal parabolic functions. The reader is invited to derive
these results as applications of the Martin boundary theory developed in this
section.

10. The Martin Boundary of b= ]0, +[ x ] - oo, S [


This Martin boundary has features not present in those discussed in Sections
8 and 9. To simplify the notation, we have taken N = 1, and we leave to
the reader the formulations for N > 1. Suppose first that S < + co. Let '
be a strictly positive number, and define 4' = (c', b). Then (c', I)) is a Martin
point set pair, and if K is the Martin function for this pair,

6b(4,0=As-t, 7)-Os
S), n = (7, t).
e(s - - 7) - G(s - t, + 7) (10.1)
K(q, 4) =
B(b-7)-G(b-t,+7)
Corresponding to each point T of R we introduce the symbol T' and for in
the right half-plane ]0, + co [ x I8 define
_ z
ifs > r
K(r )= a(2n)112(s - T) 3/2 exp 2az(s r) (10.2)
0 if sST.

Corresponding to each point y of - U8+ we introduce the symbol y" and for
in the right half-plane define

}
zazs
(e -Y4 - &4) exp if y < 0,
(10.3)
if y=0.
The functions K(r', ) and K(},", ) are positive and parabolic on the right
half-plane. The limit functions of k are the following:
(a) When q - i = (0, r) with r < S,

lim K(tq, ) = K(r', ) on b. (10.4)


n-r K(r', ')
(b) Whent - -ocand 7/t-.a', 50,

lim K(7', ) = K(` ') on D. (10.5)


376 I.XIX. The Martin Boundary in the Parabolic Context

Observe that

K(T'' ) _ = K(0",
lim for e D.
K(0',
(c) If n = (q,t)ed and either (cl) I-y +oo and q/(1 + jtl)-i +oo or
(c2) t - 6 with no restriction on the varying of ?I, then lim K(i1, ) - 0.

The Martin space of (c', /)) is therefore D compactified by adjunction of


the set {0} x ] - oc, 6[ identified with the set {i : - ao < T < S}, the set
{! ": ; E - P+ }, and a single point 0: k(T', ) and K(y", ) are defined respec-
tively by the right sides of (10.4) and (10.5); K(0, ) 0; y in b tends to i or
or 0 in the Martin topology if and only if the respective conditions (a) or
(b) or (c) are satisfied. Next we show that every Martin boundary point r' is
minimal. According to Theorem 5, it is sufficient to show that each function
K(T', ) has pole r'. Let A be the set of all Martin boundary points except the
point T'. If c > 0, the function identically equal to t on £ is positive and
superparabolic and majorizes K(T', ) on a deleted Martin neighborhood of
each point of A. Hence (reductions relative to the Martin space) K(T', ) QA
-0;so
k(T', k(T', ) OiMD < QK(T', ) Q (c 1 + QK(T',) Q A

= QK(T'. ) QI`) 5 K(T', ).

and therefore there is equality here; so k(T', ) has pole T'. Thus T' is minimal,
and similarly every boundary point ;" is minimal. Hence 0 is the only non-
minimal Martin boundary point.
In view of the corresponding discussion of the Martin representation for
positive parabolic functions on a slab in Section 8 we omit details in the
following. In the present context the Martin representation theorem states
that if a is a positive parabolic function on 6 with 6) < + oo, then
there is a unique measure 1V on ] - cc, n[ and a unique measure on - 98+
such that

u(6 = k(T', )N;:(dt) + o K(;', Nj," (d'1) (10.6)


J d

and [recall that w. 6)]

Pfu(b = J b K(r',C')N,(dT) + o K(r",S")N (dy). (10.7)

Conversely, measures N.' and Nj, making the integrals on the right in (10.7)
finite define a parabolic function ii by (10.6), and (10.7) is then true. For
12. The Minimal-Fine Topology in the Parabolic Context 377

example, if N;; = 1, on ] - oo, S[ and k4" = 0, then ti =- 1. More generally, if


measures N4 and Na" define a finite-valued function it by way of (10.6), then
it is positive and parabolic, and (10.7) is true, but the two sides may be + oo.
Finally, if 0 < b S + oc and if it is a positive parabolic function on D =
]0, + oo [ x ] - oo, S[, then (see the corresponding discussion in Section 8)
there is a unique measure Ni; on ] - oo, S [ and a unique measure Nj," on - 9t+
such that (10.6) is true and (10.7) is true when S < + Co. A positive parabolic
function on D is minimal if and only if it is a positive multiple of a function
K(r', -) or K(y", ).
The discussion of Martin boundaries for the sets in this section in the
context of Martin measure set pairs is left to the reader.

11. PWB4 Solutions on D"

Let (C', b) [(v, D)) be a Martin point [measure] set pair determining a
Martin function K and Martin boundary aD. In this context we define
universal resolutivity to mean that a",6 is h-resolutive for every strictly posi-
tive admissible parabolic function h, and universal internal resolutivity is to
mean that for every such choice of h every bounded h-parabolic function is
a PWB" solution. Under these conventions the classical Martin boundary
resolutivity theorem (Theorem XII.10) goes over directly into the parabolic
context. The parabolic version is stated for the record, but the proof follows
that of Theorem XII.10 and is omitted.

Theorem. The Martin boundary of a Martin point or Martin measure set pair
is universally internally resolutive and universally resolutive, with

h()
(for h strictly positive and admissible). An admissible h-parabolic function
it = ti/h is a ISWB" solution if and only if it is quasi bounded, equivalently, if
and only if k6 is absolutely continuous relative to k6, and then

ti = Hf =
JiMo) f= (11.2)

Let (r,,,6) [(v, D)] be a Martin point [measure] set pair defining a Martin
function k and Martin boundary aMD.
378 I.XIX. The Martin Boundary in the Parabolic Context

Parabolic Minimal Thinness

Theorem XII.11, which led to the definition of minimal thinness in the


classical context, is true when translated into the parabolic context, and the
proof of Theorem XII.11 goes over without change except for the proof of
Step 1 which involves idempotency of the smoothed reduction operator, a
property not valid in general in the parabolic context. We therefore now
prove the parabolic counterpart of this step; that is, we prove that if is a
minimal Martin boundary point and if A is a subset of D, then !K((, ) 0" is
either or a potential. If A is open, the parabolic reduction operation
on A is idempotent, so the proof of step I in Section X1 I. I I is applicable. In
the general case we need only prove that if there is a point 4 of D such that
B fC(, ) < then K(g, ) 0" is a potential. If there is such a point
apply Section XVII.II(a) on parabolic reductions to find a decreasing
sequence h of open subsets of D for which

lim
R-W
K(, ) "() _ K(, ) 'O < K(, )
Since the assertion to be proved is true when A is open, the smoothed reduc-
tion JK(1', ) OB" must be a potential for sufficiently large n. Hence lk(Z, ) p"
is majorized by a potential and so is itself a potential, as was to be proved.
The subset A of b is said to be coparabolic minimal thin at the minimal
Martin boundary point t if the conditions of the parabolic context counter-
part of Theorem XII.11(b) are satisfied, that is, if the following equivalent
conditions are satisfied:

X(4 # K(, ) on the set {IC(, ) > 0}.


inf { A'' e : E is a Martin topology neighborhood of 0.
A; is a potential.

The first condition is satisfied if and only if it is satisfied using the correspond-
ing smoothed reduction. In particular [cf. Section XII.12, Example (a)], if
!) is a deleted Euclidean neighborhood of a finite point , the point C can be
identified with a minimal Martin boundary point of b - {C}, with associated
parabolic function a multiple of GD(, C), and according to the dual of
Theorem XVIII. 10. a subset A of D - { } is coparabolic minimal thin at t
if and only if A is coparabolic thin at .

Coparabolic Minimal-Fine Limits at a Minimal Martin Boundary Point

We shall not need and it may be confusing to define a coparabolic minimal-


fine topology on a Martin space, but the coparabolic minimal-thinness
definition makes it possible to define limit concepts at a minimal Martin
13. Boundary Counterpart of Theorem XVII1.14(f) 379

boundary point. In fact it follows easily from the above criteria that the
union of two subsets of D coparabolic minimal thin at a minimal Martin
boundary point is also coparabolic minimal thin at the point. Hence, if we
call a subset A of D a deleted coparabolic minimal-fine neighborhood of the
point when D - A is coparabolic minimal thin at the point, the class (filter)
of these deleted neighborhoods defines a limit theory. Limit concepts for this
filter will be distinguished by the prefix p*mf. A subset of b which is a deleted
coparabolic minimal-fine neighborhood of a minimal Martin boundary
point has a closed-in-,6 subset with the same property. Lemma XII.15 goes
over into the present context and thereby leads to the parabolic context
counterpart of Theorem XI1.16. Thus, if is a minimal Martin boundary
point, a function u from the trace on D of a Martin topology neighborhood
of C into a metric space has coparabolic minimal-fine limit [coparabolic
minimal-fine cluster value] a at C if and only if u has limit a at C on approach
along some subset of D which is a deleted coparabolic minimal-fine neighbor-
hood of C [is not coparabolic minimal thin at C].

Coparabolic Minimal Thinness at Martin Boundary Points of a Slab


D=1F" x ]0,S[with 0<S<+oo
According to Section 8, if ' a R', the Martin boundary for the point set pair
6),6) includes the abscissa hyperplane. If k is the Martin function and
if C = (C, 0) is a point of the abscissa hyperplane, the function k(C, ) is a
multiple of the function s) -G(s, - C) = s), (C, 0)). It follows from
the criteria of Theorem XVIII. 10 for parabolic-fine limit points of a set, as
dualized for the coparabolic-fine topology, that a subset of D is coparabolic
thin at C if and only if the subset is coparabolic minimal thin relative to D
at C.

Coparabolic Minimal Thinness at Martin Boundary Points of


D = ]0, + oo [ x ] - oo, S[ with S < + oo

According to Section 10, if ' is a strictly positive number, the Martin


boundary for the point set pair ((c', S),D) includes the left boundary
{0} x ] - oo, S[. If k is the Martin function and if i = (0,,r) with t < S, the
corresponding Martin boundary point is denoted by t', and K(t', ) is a
multiple of the function K(i ,-) defined by (10.2). It is trivial that the set
]0, + oo[ x ] - oo, t] is coparabolic minimal thin relative to D at T'.

13. Boundary Counterpart of Theorem XVIII.14(f)


Theorem. Let (c', D) [(i,,6)] be a Martin point [measure] set pair determining
a Martin junction K and Martin boundary 8"r6, and let C be a minimal Martin
boundary point.
380 I.XIX. The Martin Boundary in the Parabolic Context

(a) If v is a positive superparabolic function on b, then

p'mflim v('l) = inf gi(n) =

ti is a positive cosuoerparabolic function on £ and if e,6 with


K(t, ) > 0, then O , ) > 0 on a deleted Martin topology neigh-
borhood of C and

v(n)
p*mflim , ++
= lim inft ) :5+00. (13.2)
R VU(, i) q-+(

Theorem 13(a) is the parabolic counterpart of Theorem XII.13 and is


the boundary counterpart of Theorem XVIII. 14(f). The proof follows that
of Theorem XVIII.14(f) and so will be omitted. Theorem 13(b) is the para-
bolic counterpart of Theorem XII.14. It is stated for a cosuperparabolic
function because the dual theorem for a superparabolic function involves
the Martin entrance boundary, that is, the Martin boundary dual to the
one we have studied. The proof of Theorem 13(b) is slightly more com-
plicated than that of Theorem XII. 14 and so will be given. We can assume
that the inferior limit in (13.2) is finite, and we prove that if c is a finite
number strictly larger than this inferior limit, then the set

A = {tjeb: ti(n) z

is coparabolic minimal thin at C. Under the hypotheses of (b) the function


> 0 on a deleted Martin topology neighborhood of and

4) <
"D(b, h) VD(+h)

When n , an application of Fatou's lemma yields c l k' ) "( ) <


and therefore A is coparabolic minimal thin at {. Thus the coparabolic
minimal-thin limit in (13.2) exists and has the indicated value.
Observation. The limit in (13.2) can be 0 even when ti does not vanish iden-
tically, but if ti is strictly positive, this limit is not 0, by the counterpart of
the argument in the proof of Theorem XII.14 making the corresponding
limit strictly positive.
15. The Parabolic Context Fatou Boundary Limit Theorem on Martin Spaces 381

14. The Vanishing of Potentials on (3Mb


Using the notation of the preceding sections, Lemma XII.17 on the character
of the set of (classical context) minimal-fine limit points of a subset of D and
the set of minimal-fine limit values of a function on D goes over into the
parabolic context without difficulty.

Theorem (Counterpart of Theorem XII.18). If h is a strictly positive parabolic


function on b, then each superparabolic h -potential ti = GDjt/h has coparabolic
minimal fine limit 0 at j4 almost every (equivalently, M' almost every) point
of ^6.
The proof follows that of Theorem XII.18.

15. The Parabolic Context Fatou Boundary Limit Theorem on


Martin Spaces
(We use the notation of Section 7.) The following theorem is stated for ease
of reference, but the proof is omitted because it follows that in the classical
context (Theorem XI1.19.)

Theorem. If h is a strictly positive admissible parabolic function on D and if


ti = ti/h is in "S` and is admissible, then

dAfV()
p*mfJimti(1) = (15.1)
a-c dM"

for M,, almost every point of aMD. In particular,


(a) This limit function vanishes Al,, almost surely if tiE"SA;
(b) If ti c- "Sm, the boundary function vanishes A ,h almost surely if and only
if ti E''SMS ;
(c) f) jor some Mg integrable
I f t i is a PW B" solution, that is, i f u = p
boundary function f, then f = ddc./dA;, Mk almost everywhere on
a'b; therefore f is the coparabolic minimal fate boundary limit func-
tion of e up to an Mg null set.

Extension to Nonstrictly Positive h

If h is not strictly positive, apply Theorem 15 to the ratio 1/(h + 1) to find


that this ratio has a finite coparabolic minimal-fine limit M;, + Mt almost
everywhere on the Martin boundary and so certainly M,, almost everywhere
there. Hence h has a strictly positive (5 + oo) coparabolic minimal-fine limit
382 I.XIX. The Martin Boundary in the Parabolic Context

M;, almost everywhere on the Martin boundary. If Theorem 15 is applied to


v/(h + I) and to h/(h + 1), it follows that v/h has a finite coparabolic minimal-
fine limit Mk almost everywhere on the Martin boundary, in fact the limit
dMo/dMh. The key point in understanding the situation is a rephrasing of
part of one fact just derived: the zero set of h is coparabolic minimal thin
at M;, almost every Martin boundary point. Thus we can ignore this zero
set in the present discussion. The necessary rephrasing of Theorem 15(a)-(c)
when h may vanish is left to the reader.

Application to Functions on a Slab I8" x ]0, 6[ with 0 < b < + oo

Let h be a strictly positive parabolic function on the slab determined using


(8.4) by a measure N;, on I. Let t be a positive superparabolic function on
the slab with parabolic component determined by the measure Ni,. Let
= (c',s') be a point of finiteness of v, and set t) = RN x ]O,s'[. Then
15) is a Martin point set pair to which Theorem 15 can be applied. In
view of the relation (Section 12) between coparabolic thinness and coparabol-
ic minimal thinness at a point (C,0) and in view of the relation between the
measures f , Re and the corresponding measures in the Martin representa-
tions of h and the parabolic component of v, we conclude from Theorem 15
that ti = v/h has coparabolic-fine limit df4/dN;, N4 almost everywhere on the
abscissa hyperplane. Thus we have derived Theorem XVIII.15 as a special
case of the Fatou boundary limit theorem in the parabolic Martin boundary
context. Observe that if t is parabolic, we have proved (using Theorem XVI.7)
that v/h has parabolic limit dNo/dNh N;, almost everywhere on the abscissa
hyperplane. On considering the corresponding situation in the classical
context (Section XII.20) it is natural to conjecture that if v is a superparabolic
potential on the slab, then v/h has limit 0 Nh almost everywhere on the
abscissa hyperplane on approach along normal lines. This conjecture is false
(see Historical Notes) even when h = 1, in which case N;, reduces to 1". Thus
coparabolic minimal-fine boundary approach is the most appropriate bound-
ary approach in treating Fatou boundary limit theorems, just as the minimal
fine topology approach was the most appropriate in the classical context.

Application to Functions on the Set ]0, + oo [ x ] - oo, 6[ with 6 5 + oo

Let h be a strictly positive parabolic function on this set, determined using


(10.6) by a measure N;; on ] - oo, S[ and a measure k; on - R+. Let ti be a
positive superparabolic function on this set with positive parabolic compo-
nent in the Riesz decomposition determined by measures R. and R,,'. Let
'_ s') be a point of finiteness of V and set 6 = ]0, + oo [ x ] - co, s'[.
Then b) is a Martin point set pair to which Theorem 15 can be applied.
The measures N ; ; and f do not depend on the choice of ', and if s' is
increased to s" < 6, the measures Nj and N,; do not change on ]-oo,s'[.
15. The Parabolic Context Fatou Boundary Limit Theorem on Martin Spaces 383

Theorem 15 in this context, applied only to the left boundary {0} x j - oo, d[,
asserts that ti/h has coparabolic minimal-fine limit dN/dN/, at NA' almost
every point of the left boundary. See Historical Notes for the details on this
result for other than coparabolic minimal-fine approach to the left boundary.
Observe that coparabolic minimal-fine approach to a point (0, r) is signif-
icant only for approach by way of points with ordinate values strictly greater
than r because (from Section 12) the set ]0, + oo [ x ] - oo, r] is coparabolic
minimal thin at (0, r). If h is not strictly positive, there is a number ro such
that 0 if and only if r > To. The boundary set {T': r 5 To} is then
N/, null, and the discussion needs only trivial changes.
Part 2
Probabilistic Counterpart of Part 1
Chapter I

Fundamental Concepts of Probability

1. Adapted Families of Functions on Measurable Spaces


(a) Filtrations of a Measurable Space. Let (fl, F) be a measurable space,
and let (I, 5) be a linearly ordered set. A filtration of (f2, F) is a map t - F(t)
from I into the class of sub a algebras of IF, increasing in the sense that
s 5 t implies that F(s) a .9r(t). The triple (0, F, is called a filtered
measurable space.
If (52, .F, fl-)) is a filtered measurable space and if the index set I is an
interval in l ordered by 5, define F+(t) = n,,,F(s) for t in I to get a
filtration with F(t) c .F+(t) for tin I. The filtration is said to
be right continuous if F(-) = In particular, is necessarily
right continuous.
(b) Families of Functions. Let {x(t), t e I) be a family, indexed by a pa-
rameter set I, of functions from a space f2 into a "state space" S2'. The value
of the function x(t) at the point w of 0 will be denoted by x(t, w); the function
x(t) will also be denoted by x(t, ). The function w) : t r- x(t, CO) is called
the sample function determined by co. The family x(-) of functions on 0 is
described as continuous, or right continuous, or ... if the sample functions
are all continuous, or right continuous, or .... for a state space and pa-
rameter set structured to make the description meaningful.
(c) Adapted Families of Functions. If {x(t), tell is a family of measurable
functions from a filtered measure space (S2, into a measurable
space (12', -W') the family is said to be adapted to if for each tin the
parameter set I the function x(t) is measurable from the measurable space
(0,9(t)) into the measurable space (0%9'). The notation
will always imply adaptedness unless the contrary is stated. If { x(t), t e 1)
is a family of measurable functions from a measurable space (Q, F) into
a measurable space and if I is linearly ordered, define ,9r0(t) =
,9r{x(s),s 5 t} for t in I (see Appendix IV.1 for the notation used here)
to obtain a filtration making an adapted family. The
filtration go(-) is minimal for adaptedness in the sense that if (F(t), to l)
is a filtration with respect to which is adapted, then #0(t) c
F(t)fortin1.
388 2.1. Fundamental Concepts of Probability

Filtered measurable spaces and their adapted families of functions pro-


vide a mathematical formalism modeling certain physical ideas. A measur-
able space (Q, 37) is a mathematical model of the set of possible events in
some physical context, together with a distinguished class of compound
events. If I is a subset of R, a filtration {.fi(t), t e 11 of (Q, F) is a mathe-
matical model for the flow of events in time. Each pair (t, w) represents a
possible outcome of an experiment at time t, and .fi(t) represents the class
of compound events observable before or at time t. The value x(t, w) of a
function x(t, -) at co models the value of some observable at the outcome
(t, w), and the function x(t, ) itself is therefore incorporated in F (t) in the
sense that this function is supposed .F(t) measurable; that is,
is an adapted process.

The Measurable Sets of a Topological Measurable Space

If a measurable space is given as a topological space, the a algebra of measur-


able sets will always be the a algebra of Borel subsets of the space unless
some other a algebra is specified. In particular, the state space R means
the measurable space (R, .(R)).

2. Progressive Measurability
If I is a subinterval or singleton of R, we denote by(/) the class of Bore]
subsets of /. If I is a singleton, R (l) consists of that singleton and the empty
set. A family {x(t), t e 1) of functions from a measurable space (5),.x)
into a measurable space (0', .I') is called measurable if the function (t, (0) r+
x(t, w) is measurable from the measurable space (I x Q,. (I) x .F) into
the measurable space (0',.F'). This definition will frequently not be strong
enough, however, when we deal with an adapted family {x(t),.F(t), teR+}.
In fact in dealing with an adapted family one frequently wishes to deal with
a subfamily of the form {x(t), .F(t), I < c} and wishes the analysis of this
subfamily, in particular, the measurability of the subfamily, to depend
only on properties involving parameter values in the interval [0, c]. Un-
fortunately even if 9 = YE R+ .fi(t), the measurability of the original family
does not imply that the subfamilies of the stated form are measurable.
The following definition is formulated precisely to provide this measur-
ability.
Suppose then that {x(t), ,fi(t), t n R+ } is an adapted family of functions. A
subset A of R+ x 0 will be called progressively measurable if A n ([0, c] x Q)
e 9([0, c]) x F(c) when c >_ 0. The class of progressively measurable sets
is a a algebra. The adapted family is called progressively measurable if
the function is measurable from the space R+ x 0 coupled with the
2. Progressive Measurability 389

or algebra of progressively measurable sets into (a', .5'), equivalently, if


for c z 0 the restriction of the function x
c] x ([0, x .5(c)) into (0', 5'). (In this phraseology the
condition that a set A be progressively measurable is equivalent to the
condition that the adapted family { le(t, ), .fi(t), t e 68+ } with state space
(R+,.V(R+)) be progressively measurable. Observe that even without the
prior hypothesis of adaptedness the progressive measurability condition
implies adaptedness. We leave to the reader the easy proof that if {x(t),
.fi(t), t e 18+ } is progressively measurable, then for b > 0 the family {x(b + t),
5(b + t), to R+} is also progressively measurable. It is immediate that if
is progressively measurable and if f is a measurable function
from the state space (f2', .') into the measurable space then
If is also progressively measurable, with state space

Alternative Characterization of Progressive Measurability

In the preceding context define a new filtration by setting .5-(0) = .5(0)


and .5- (t) = Y<, 5(s) for t > 0, and consider the following conditions
on the family {x(t), to R+} of functions from (Q,JF) into (0', Jr):
(a) The family is adapted to
(a') is right continuous.
(b) For c > 0 the restriction of the family to the set [0, c[ is
R([0, c[) x F- (c) measurable.
Observe that the pair (a'), (b) implies (a). We now show that
is progressively measurable if and only if (a) and (b) are satisfied. To prove
that (a) and (b) together imply progressive measurability, we need only
observe that the subsets [0, c[ x 0 and {c} x 0 of [0, c] x 0 have the
latter set as union and are in R Q0, c]) x .5(c) and that under (a) and (b)
the function is measurable on both subsets and so also on their union,
relative to .([0,c]) x JIF(c). Conversely, progressive measurability implies
(a), and another piecing together of measurable functions shows that pro-
gressive measurability also implies (b).

EXAMPLE. Let (x(t), .fi(t), t e R+ } be an adapted family whose state space


is Polish. Then the family is progressively measurable if the sample functions
are all right continuous or all left continuous. For example, in the right
continuous case define

n / if(j - 1) nC <t5ic,
n
1 5j 5n,
(2.1)
x (0, w) = x(0, 0
390 2.1. Fundamental Concepts of Probability

when c > 0. Then the restriction of x x F(c)


measurable and x so the progressive measurability
condition is satisfied when c > 0. The condition when c = 0 is satisfied
because the family is adapted. The left continuous case is treated similarly.
Observe that if whenever e > 0 the restriction of the family to the
set [0, c] is 4[0, c] x ,F(c + e) measurable, then (b) is satisfied. In fact
this condition implies that for 0 < S < c the restriction of this family to
[0,C - 6] is

91([O, c - S]) x ., (c - 21 ..([0, c[) x F-(c

measurable, and this yields (b). In other words the adapted process
is progressively measurable if the process {x(t), F(t + e), to I8+} is
progressively measurable for every e > 0.

3. Random Variables

A probability space is a triple (S2, JF, P) consisting of a measurable space


(U..x) and a measure P on F with P{O} = 1. It will frequently be un-
necessary to specify 0 or F, but it will always be supposed even though
not mentioned that P is complete, that is, that subsets of P null sets in F
are also in F.
A random variable is a measurable function from a probability space
into a measurable "state" space. If the state space is topological, its measur-
able sets unless otherwise specified will be the Borel sets. The distribution
of a random variable x is defined as the measure y on the class of measurable
sets of the state space induced by x: u(A) = P{xeA}. The integral of a
real-valued random variable x will be denoted by E{x}, and the integral of
x over a set A will sometimes be denoted by E{x; Al.
Sometimes the context is expanded to allow measures P other than
probability measures, even to allow P{i2} = + oc, but if so, the expansion
is mentioned explicitly.
We follow the common abuse of notation in measure theory, in which
the same notation is used for equivalence classes of measurable functions
under the equivalence relation of almost everywhere (a.e.) equality, or
"almost sure" (a.s.) equality in the slang of probability, as for their members.
The proper interpretation will be clear from the context. Thus the essential
infimum ess info z, of a family of random variables (Appendix IV.9) is
determined only up to a null set and properly speaking is an equivalence
class, but the language and the notation will refer either to the equivalence
class or to one of its members, as indicated by the context. The members of
such an equivalence class will sometimes be called versions of the class.
4. Conditional Expectations 391

4. Conditional Expectations
If x is a real integrable random variable on a probability space or a real
random variable bounded on one side, and if St is a a algebra of measurable
sets, the conditional expectation E{xIS } is the uniquely determined up
to null sets random variable which is Sr measurable and satisfies

E{E{xjS=};A} = E{x;A} for AeSr. (4.1)

The notation E{xlF } always refers to any one of the possible choices
of the conditional expectation rather than to the equivalence class. Unless
the contrary is stated, it will always be assumed in discussing this conditional
expectation that x is integrable. The Radon-Nikodym theorem ensures
the existence and uniqueness up to null sets of conditional expectations,
and the following six properties follow easily from the defining properties.
(a) E{ yjF } = y a.s. if y is 9 measurable, in particular, if y is a constant
function.
(b) If y is bounded and 9 measurable, E{xylS } = yE{xjSr } a.s.
(c) E{x + yjSr } = E{x!F } + E{ yjSF } a.s.
(d) x5 y a.s. implies that E{xjg } 5 E{yjF } a.s. and almost sure
equality in the first inequality implies almost sure equality in the second.
These properties imply that the map x"E{xjS } is a positive linear
idempotent transformation from L' into L', of norm 1.
(el) If Ae.Sand if x=ya.s.on A,then E{xjS=}=E{yf.F}a.s.on A.
(e2) If A E Sr r I and if the a algebras Sr and I have the same trace
on A (that is, if a subset of A is in 9 if and only if the subset is in !fl, then
E{xj5} = E{xlg} a.s. on A.
(f) If S= c 4, then

E{E{xlSt}14} = E{E{xl§}I5} = E{xI.4r} as. (4.2)

The next properties are less immediate, and their proofs will be sketched.
(g) If x is independent of Sz (that is, if x is independent of every Sr
measurable random variable), E{xIF } = E{x} a.s.
In fact, if y is the indicator function of a set A in Sr, the random variables
x, y are independent and

E{x; A} = E{xy} = E{x} E{ y} = E{E{x} ; A} as. (4.3)

(h) Jensen's inequality for conditional expectations. If f is a convex


function and if x and fix) are integrable, then

f[E{xIS=}] S E{ f(,c)I5} 'a.s. (4.4)

In fact, if z a + b for all ,


392 2.1. Fundamental Concepts of Probability

E{ f(x)IF} z aE{xl.*T} + b a.s., (4.5)

and since a convex function is the supremum of all the linear functions
it majorizes, and even the supremum of a properly chosen countable subset
of those linear functions, (4.5) implies (4.4). If 3F contains only the space
and the empty set, (4.4) reduces to Jensen's inequality.
According to (h), if pal, the transformation takes LP
into itself. In particular, let 9R be the space of square integrable F measur-
able random variables. The transformation maps LZ onto 0, and it is
easy to verify that x - E{xl.F } is orthogonal to 0, so that the trans-
formation acting on LZ is the orthogonal projection onto R.
(i) If x is integrable, the family of all conditional expectations of x is
uniformly integrable.
In proving this it can be assumed that x is positive (or replace x by lxl),
and then if we write x, for E{xl.F } and if c is a strictly positive constant,

E{ x,, ; x, > c} = E{x; x, > c}, P{x,, > c} 5


E xs = Mx
c c
(4.6)

The second relation ensures that, for sufficiently large c, P{x, > c} is
uniformly small as .F varies and the left side of the first relation must then
also be uniformly small for large c as F varies, and this property is precisely
the uniform integrability property.

Alternative Proof

The following less direct proof displays a technique which has many appli-
cations. Since x is integrable, this function constitutes a uniformly integrable
class; so there is a uniform integrability test function 0 (Appendix V) for
which E{b(I xl) } < + oo. Apply Jensen's inequality (h) above for conditional
expectations to obtain

E{1(E{Ixl IF})) s E(E{D(Ixl)IF} = E{'D(Ixl)}.

Since the left side is uniformly bounded as .F varies, the family of conditional
expectations of x is uniformly integrable.
More generally a trivial extension of either proof shows that the class
of conditional expectations of the random variables of a uniformly inte-
grable family is uniformly integrable.
(j) If {x(t), t e J } is a family of integrable random variables and is
directed upward neglecting null sets, then (whether or not ess sup,. I x(t) is
integrable)
5. Conditional Expectation Continuity Theorem 393

ess sup E{x(t)lF } = E {ess sup x(t)IF} a.s. (4.7)


rel rcl 111

This fact for integrals, that is, in the present context for expectations,
is proved in Appendix IV.9. Since the family (E{x(t)I5 }, t e 1} is directed
upward neglecting null sets, the result for integrals yields, when A E .F,

E {esssuPE{x(t)Is } ; A } = sup E{x(t); A} = E {esEsup x(t); A} ,


(4.8)

and this equation is equivalent to (4.7) if we choose the essential supremum


on the left, as we can, to be .F measurable.
The corresponding result for downward-directed families is deduced
by replacing x(t) by -x(t).
Specialization. If x is an integrable random variable and if is a family
of random variables, is defined as In particular,
if is a finite set of random variables, say x(1), ... , x(n), it follows
(Appendix IV.3) that there is a Borel measurable function 0 on R" such that

E{xlx(l),... ,x(n)} = 4,[x(1), ... ,x(n)] a.s. (4.9)

5. Conditional Expectation Continuity Theorem


Theorem. Let {,9=(n), n e l+ } [ {F(n), n e -V}] be a filtration of a proba-
bility space (0, .F, P), and define F( + oo) =Yo .fi(n) [JF(- oo) =
n°_. 5(n)]. Then if x is an integrable random variable on f2, it follows that
in both the almost everywhere and Lt senses

lim E{xl.fln)} = E{xlfl+oo)}


ft-M
(5 . 1)

in the forward case and

lim E{xIF(n)} = E{xl#'(-oo)} (5.2)

in the backward case.

This theorem has applications in surprisingly many apparently unrelated


contexts. It is a martingale convergence theorem and as such will be derived
again (Sections 111.14 and 111.17), but an independent proof will be given
here to avert suspicion of circular reasoning. To simplify the notation,
define x(n) = E{xl S(n)} for n in 1+ or in -Z+ according to the context.
394 2.1. Fundamental Concepts of Probability

To prove (5.1), choose integers j < k < ! < m and real numbers a < b,
and define

nf{n21:x(n)Zb},l
v=i{min
[j, k, !; r] = x(n) S a; v =
/snsk J
(5.3)

[j, k;1, m] = 5min x(n) 5 a; max x(n) z b = U [j, k, !; r],


jSnSk 15n5m r=t

[a; b] _ {liminfx(n) < a, lim sup x(n) z b } .


01-Go

Then x(r) z b on [ j, k, !; r]. so

f x dP= J x(r)dPZ f J,k.l;rl bdP


Ij.k;l.w,1 lJ,k.l;r1
(5.4)

=bP{[j,k;l,m]}.
Observe that if "5 a" had been " < a" or " Z b" had been "> b" in the
above definitions, the change would carry over throughout with no change
in the reasoning, and a trivial continuity argument for varying a and b
shows that here and in further reasoning below the inequalities allowing
equality imply those prescribing strict inequality and conversely. In view
of this fact (5.4) yields, when successively m -+ + oo, ! -+ + oo, k --+ + oo,
j-4 +oo,

xdP> bP{[a;b]}. (5.5)

If the triplex, a, b is replaced by - x, -b, -a in (5.5), this inequality becomes

xdP < aP{[a;b]}, (5.6)


Lb,
and we conclude that P{[a;b]} = 0. The equality

diverges} = U [a; b]
a<b
a,b rational

implies that there is almost everywhere convergence, say lim, x(n) on


x'(+ o o), and since according to Section 4(i) the sequence is uniformly
integrable, the limit function x'(+oo) is integrable and there is L' conver-
gence (Appendix V). Finally, x'(+oo) can be chosen F(+oo) measurable,
5. Conditional Expectation Continuity Theorem 395

and if A is in the algebra Uo F(n) which generates (+ co), then Jn x dP =


JA x(n) dP for sufficiently large n so in view of the L' convergence of

J IxdP= f x'(+oc)dP.
A Jn

The equality must also be true for A in .IF(+oo), so x'(+co) = x(+oo)


almost surely, as was to be proved.
The proof of (5.2) is similar to that of (5.1) but is easier. If a < b and
m < n, the same argument as that in the preceding proof shows that if v is
defined as in (5.3), if

[!, m] = {x(- cc) 5 a, max x(n) z b},


iSnen'

and if [a; b] is defined as in (5.3) except that n - oo, then

xdP Z bP{[!,m]}, (5.4')


Jil.m]

so that if successively I -+ - oo, m -+ - oo,

xdP z bP{[a; b]}. (5.5')


J'la bl

Replacing x, a, b by -x, -b, -a as in the preceding proof, we find that


P{[a;b]} = 0, which implies that the uniformly integrable sequence is
almost surely and L' convergent to some random variable x'(-oc). This
random variable can be chosen .J"(- oo) measurable. Finally, if A c -.F(- co )

fX( - oo) dP = f'X dP= f'x(n) dP


n

f or al l finite n and when n -' - oo, the L' convergence of implies that
these three integrals are also equal to J x'(- co) dP; so x'(- co) = x(- co)
almost surely, as was to be proved.

Generalization of Theorem 5 to Linearly Ordered Index Sets

Let (1, 5) be a linearly ordered set and let {I(t), t e I} be a monotone family
of a algebras of measurable subsets of a probability space. Suppose that I
does not have a last element and denote by l(oo) either'. r f(t) or nfe I I(t)
according as is monotone increasing or decreasing. If x is an integrable
random variable, we shall now show, generalizing Theorem 5, that
396 2.1. Fundamental Concepts of Probability

limE{xj4(t)} = E{xlf(cc)} (5.7)


(t

both as an almost everywhere essential limit and an L' limit. We shall use
the notation x(t) = E{x14(t)} for tely {oo}.
Case (a) : I is countable. In this case there is according to Appendix
Theorem IV. 10 an increasing sequence t, in I for which the essential limits
inferior and superior of along I are, respectively, the same as along t..
The generalization of Theorem 5 for countable I thus follows from Theorem
5. Moreover the limit in (5.7) can be taken as an ordinary almost everywhere
pointwise limit because I is countable.
Case (b) : General case. If I has a countable cofinal subset, there is even
(Appendix 111.13) a countable cofinal subset J of I such that has the
same essential limits inferior and superior, respectively, along J as along 1.
According to case (a), there is almost everywhere convergence along J, and
it follows that there is essential order convergence along I. The identification
of the limit and the L' convergence are left to the reader. If I has no countable
cofinal subset, we prove a stronger result than (5.7): there is a point to of I
such that for t > to

P{x(t) = x(to)} = 1. (5.8)

To see this when is increasing, observe that (Appendix IV.2) there is a


countable subset J of 1 such that x(oo) is YFj G(t) measurable, and we leave
to the reader the proof that (5.8) is true with to any upper order bound of J.
The following argument is more instructive and shows that to exists satisfying
(5.8) whether is increasing or decreasing. An order-preserving map 0
will be exhibited in Section 111.4, involving only elementary inequalities,
taking I into a subset I' of R, with the property that ¢(s) = 4)(t) if and only
if P{x(s) = x(t) } = 1. The set I' must have a last element t' because I' has a
cofinal sequence s;, and if s, is a sequence in I with 4)(s) = s.' and if t is an
order upper bound of s,, then 4)(t) = t' is the last element of 1'. Each element
to of 46-(t') satisfies (5.8).

6. Fatou's Lemma for Conditional Expectations

Lemma. Define 1, as in the generalization of Theorem 5, and let {x(t),


t E I } be a family of positive extended real-valued random variables. Then

essliminfE{x(t)J4(t)} z E{essliminfx(t)1f(oo)} a.s. (6.1)

In particular, if I = Z+,
7. Dominated Convergence Theorem for Conditional Expectations 397

liminfE{x(n)I'4(n)} z E{liminfx(n)Il(oo)} a.s. (6.1')

For every k in Z+ and sin I the left side of (6.1) is almost surely at least

essliminfE{k n essinfx(r)I5(t)} = E{k n es2infx(r)I#(co)} a.s. (6.2)


it r,> S s

Inequality (6.1) follows from the fact that the right side of (6.2) increases as
k and s increase and has the essential limit E{essliminf,rx(t)JI(co)}
according to (4.7).

7. Dominated Convergence Theorem for Conditional


Expectations
Theorem. Let {9(n), n e Z + } be a monotone sequence of a algebras of mea-
surable subsets of a probability space, and denote by V(oo) either z 4(n)
or n.,4(n) according as W(-) is monotone increasing or decreasing. Let
{x(n), n e Z + } be a sequence of random variables on the space, with almost
sure limit x(oo) and Ix(n)j) < + oo. Then

lim E{x(n)I4(n)} = E{x(oo)I3(oo)} a.s. (7.1)

To prove this theorem, apply Fatou's lemma for conditional expectations


to the sequences

{Msuy Ix(m)I + x(n), neZ+}, {sup Ix(m)I - x(n), neZ+}.

Application. Let Jr be a a algebra of measurable sets of a probability


space. Let x and v be random variables on the space with x independent of
y measurable with respect to F. Let j be a bounded Borel measurable
function on 682, and define f, on 68 by f, (?I) = E{ f(.r, q)}. Then f, is Borel
measurable, and E{ f(x,,v)I.fl = f, (y) almost surely. To prove this asser-
tion, let r be the class of bounded Borel measurable functions f for which
this assertion is true. The class r includes the functions of the form q) r--#
g and h bounded Borel measurable functions on R because in
this case f, (i) = E{g(x) }h(ry), and by Section 4(b) and (g)

E{g(x)h(y)I.,F} = E{g(x)I,}h(y) =fl(y) a.e. (7.2)

The class r is linear and is closed under bounded pointwise convergence in


view of the dominated convergence theorem for conditional expectations.
Hence r is the class of bounded Borel measurable functions on 682, as was
to be proved.
398 2.1. Fundamental Concepts of Probability

8. Stochastic Processes, "Evanescent," "Indistinguishable,"


"Standard Modification," "Nearly"
Let (S2, .,F, P) be a probability space, and let I be an arbitrary set (to be used
as the parameter set of a family of random variables). A subset A of I x d2
is called evanescent if there is a set A in F for which P{A} = 0 and A C I x A.
A subset of an evanescent set is evanescent, and a countable union of evanes-
cent sets is evanescent. For many purposes evanescent subsets of I x 0 are
counterparts of the polar subsets of the state space in classical potential
theory, and we shall accordingly describe a relation true on I x 0 up to an
evanescent set as true quasi everywhere.
A stochastic process on (Q, 337, P, 1) is a family of random variables
{x(t), t e I } with a common state space, defined on (0, S=, P), with parameter
set 1. The random variable x(t) of this process has the value x(t, (.0) at the
point w of S2, and the process thus defines a function from I x .0 into the
state space of the process, that is, the state space of the random variables.
with the property that each function x(t) = x(t, ) is measurable. Two
processes and with common probability space, parameter set, and
state space are called indistinguishable if quasi everywhere on
I x f2, that is, if

P{w: x(t,w) = y(t,(9) for all t} = 1. (8.1)

Two subsets of I x S2 are called indistinguishable if their indicator functions


determine indistinguishable processes, that is, if the sets differ by an evanes-
cent set. Two processes and with common probability space,
parameter set, and state space are said to be standard modifications of each
other if

P{w: x(t,w) =Y(t,(0)) = 1 (8.2)

for all t. The adjective "standard" is sometimes omitted.


Sums of processes, functions of a process, and so on are defined in the
obvious way in terms of the random variables of the processes.
The concept of an adapted family {x(t), .fi(t), t e 1) of functions on a
measurable space was defined in Section 1. If the measurable space is a
probability space, the family becomes an adapted stochastic process,
and the above notation always implies adaptedness unless the contrary
is stated.
A stochastic process indistinguishable from one having a specified char-
acter will sometimes be said to have nearly that character, and the corre-
sponding description will be used for subsets of I x 0. Thus a nearly pro-
gressively measurable process is a process indistinguishable from a progres-
sively measurable process.
8. "Evanescent," "Indistinguishable," "Standard Modification," "Nearly" 399

Normalizing Hypotheses on Filtrations

Let {.f(t), t e 1 } be a filtration of a measure space, and suppose that I has a


first point, denoted by 0. If the given measure is complete, as is our conven-
tion for probability measure spaces, the hypothesis that .F(0) contains the
null sets implies that the restriction of the given measure to each a algebra
.f(t) is complete, but the converse is false. In practically every investigation
involving a filtration the parameter set has a first element, and it is possible
without loss of generality to suppose that the first a algebra contains the
null sets, at the cost of enlarging each a algebra .fi(t) to the a algebra gener-
ated by .fi(t) and the null sets. If the parameter set is it is usually also
OB+,

possible without loss of generality to replace by the right continuous


filtration

Progressively Measurable Stochastic Processes

If {x(t), .f(t), t e I8+ } is a progressively measurable stochastic process every


sample function w) is necessarily Borel measurable. If this process is only
nearly progressively measurable, it need not be progressively measurable, but
almost every sample function is Borel measurable. For example, define a real-
valued process with w) arbitrary for w in some null set of a probability
space, define co) = 0 otherwise, and define .fi(t) for all t as the or algebra
consisting of the null sets and their complements. The process fl-)) is
then indistinguishable from the identically 0 progressively measurable pro-
cess but is progressively measurable only in trivial cases. It is nevertheless
true that most theorems involving progressively measurable processes are
valid with unimportant modifications for nearly progressively measurable
processes.

Almost Surely [Right] [[Left]] Continuous Processes and Progressive


Measurability

If {x(t), t e R + } is a stochastic process with a topological state space the


process will be called [almost surely] continuous or right continuous or left
continuous if [almost every] sample function has the stated property. If
to R+} is almost surely [right] [[left]] continuous and adapted
and if .F(0) contains the null sets the sample function (0) can be redefined,
say by setting x(t, (o) = x(0, (o) for all t when w is in the exceptional null set,
to obtain a [right] [[left]] continuous process adapted to Jr(-) and indis-
tinguishable from the given process. If in addition the state space is Polish,
this modified process is progressively measurable according to Section 2.
It would be consistent with our conventions to call an almost surely right
or left continuous process nearly right or left continuous, but the almost
surely terminology is the customary one.
400 2.1. Fundamental Concepts of Probability

Stochastic Processes in the Essential Order

The class of extended real-valued processes with a common probability


space and parameter set is ordered by the essential order in which is a
majorant of x(), denoted by 5 or z if P{x(t) 5 y(t)} = 1
for all t. This order is thus an ordering of equivalence classes in which two
processes are identified if each is a standard modification of the other. The
usual inexact language in which a "process" may actually be an equivalence
class will be used when there is no danger of confusion.

Essential Order Infima and Suprema

If r : {.r,(-),ac-J} is a family of stochastic processes with a common


probability space, common parameter set, and extended real-valued
random variables, the essential order infimum of the family, denoted by
ess inf,E j or ess inf r, is the process whose tth random variable is
ess in f , x,(t) [chosen to be .F(t) measurable if r is specified as a family
of processes adapted to a common filtration The process ess sup r is
defined and denoted dually. These processes are uniquely determined up to
standard modifications and therefore are unique in the essential order.

LP Bounded and Uniformly Integrable Processes

A stochastic process is said to be L° bounded if the family of its random


variables is L" bounded and is said to be uniformly integrable if the family of
its random variables is uniformly integrable.

Sample Function Integrals

If {x(t), .fi(t), t E 98' } is a progressively measurable stochastic process with


state space the extended reals, the process sample functions are Borel mea-
surable. If the process random variables are positive, the sample function
integrals are well defined, that is, meaningful, and if y(t) = fo x(s)1, (ds), that
is, if for each point to of the probability space on which the process is
defined y(t. to) = It x(s, w)1, (ds), then the process { y(-),.F(-)) is progres-
sively measurable (Fubini's theorem). This integral process is also well
defined and progressively measurable if instead of supposing that the random
variables are positive, it is supposed that the process sample functions
are 1, absolutely integrable over finite intervals. If the process {x(-), is
not supposed progressively measurable but merely indistinguishable from a
progressively measurable process, the above remarks are modified in the
obvious way.
9. The Hitting of Sets and Progressive Measurability 401

9. The Hitting of Sets and Progressive Measurability


Theorem. Let (f ,.F, P; .fi(t), t e R) be a filtered probability space, suppose
that the restriction of P to .fi(t) is complete for all t, and let c be a strictly
positive constant.
(a) If A" is a subset of R x 0, analytic over the class of progressively
measurable sets, then the projection of A" n ([0, c] x (2) on 0 is in
3F(c).
Let be a progressively measurable stochastic process
on (S2, P, 5), with an arbitrary state space (X, .1).
(b) If A' is a subset of P+ x X analytic over i(1) x T. then

(co: (t,x(t,w))eA' for some t 5 c}e.F(c). (9.1)

(c) If A is a subset of X analytic over L", then

(co; x(t, co) e A for some t S c} E .5(c). (9.2)

Assertion (c) is a special case of (b) because in (b) the set A' can be chosen
as [0, c] x A. Assertion (b) is a special case of (a) by the following argument.
The map (t, co) t-- (t, x(t, (o)) from the measurable space R' x S2 coupled with
the class of progressively measurable sets into the measurable space (R x X,
R(68`) x L") is measurable because the inverse image of any product set
[0, b] x Xo with X. in LI is a progressively measurable set. Hence (Appendix
Theorem 1.8) the inverse image of any set A' analytic over R(R) x .I" is a
set A" analytic over the class of progressively measurable sets. Finally the
projection of A" n ([0, c] x i2) is then the set in (9.1); so (a) implies (b). To
prove (a), observe that by Theorem 7 of Appendix I the projection of A" on
i2 is analytic over 5(c) and therefore is in 5(c) by Lusin's theorem (Theorem
4 of Appendix II because the restriction of P to JW(c) is a complete measure.

Variant of the Theorem

In (a) the projection of A" n ([0, c[ x S2) on i2 and in (b) and (c) the sets in
(9.1) and (9.2) with "t < c" replaced by "t < c" are all in w- (c) = Y«.f(t).
The method of proof of the theorem yields this variant, or the variant can be
deduced by an application of the theorem with c replaced by c - 1 In, n z 1.

Generalization

If in (b) and (c) the process is supposed merely indistinguishable from a


progressively measurable process, the conclusion remains valid if 5(0) con-
tains the P null sets. This added condition implies the validity of the condi-
402 2.1. Fundamental Concepts of Probability

tion already imposed, that the restriction of P to .f(t) be complete, because


P as always is itself supposed complete.

Change of Origin

The analysis of the hitting of a set in the parameter interval [0, c] or [0, c[ is
valid with trivial changes for hitting in an interval [b, c] or [b, c[ with
0<b<c.

10. Canonical Processes and Finite-Dimensional Distributions


Let (X, a?") be a measurable space, let I be a space and let n be a subset of the
space X' of functions to from I into X. Denote to(t) by z(t, (h), and define
-0 = f {x(t), t E Pj ; that is, .0 is the a algebra generated by the algebra
U .f { _Y(t I )..... z(t") }, where the union is over all finite subsets t, of I. If
P is a probability measure on .I, extended by completion to a complete
probability measure on A' 31', {z(t), t e I } is a stochastic process on
(6, _4"*, F). A process defined in this way, by means of a measure on a subset
of the function space X', will be called canonical.
If {x(t), t e I } is a stochastic process with state space (X, 37) on a prob-
ability space (S), ,F, P), a finite-dimensional distribution of the process is the
distribution of a finite set of the process random variables,

Ar-.p(t1, ...,t,,;A)=P{[x(t,), ...,x(t,J]EA} (AE.("). (10.1)

An associated canonical process with the same parameter set, state space,
and finite-dimensional distributions can be defined as follows. Set 6 = X'.
The map 0: is measurable from (Q,9) into (6, A) because

..... (t")]EA} _ {[x(t1). ...,x(t")]EA}E.F (A c- 37").

(10.2)

Define I' on A as the measure induced by P; that is, P{R} = P{0-' (A) }. If
as usual P is supposed complete, it follows that F; so 0 is a
measurable map from (Q. S) into (Sy2, . *). Note that F may be strictly
larger than 4 1(. ) so that there may be sample function properties defin-
ing measurable subsets of 0 but not of n. Sample function properties involv-
ing uncountably many parameter values are obvious candidates.
If I is a topological space and if (X, ff ) is a topological space with 5t the
a algebra generated by the open subsets of X, suppose that in the pre-
ceding paragraph is a continuous process. Then we can choose an associated
continuous canonical process by setting 1 = n,, the space of continuous
functions from I into X, a subset of X'. The map 0 above is now from f2 into
(g2,, and P induces a measure on this function space. The family of
10. Canonical Processes and Finite-Dimensional Distributions 403

coordinate variables on 6c is the canonical continuous process associated


with If is an almost surely continuous process, that is, if there is a
measurable subset t2c of ( with P{f2 } = I whose corresponding sample
functions are continuous, the preceding procedure applied to C2 instead of
0 yields a canonical continuous process associated with If I is ordered,
with a first point a, and if the distribution of x(a) is supported by some
measurable set A, it is sometimes convenient to restrict the associated canoni-
cal process further by defining 6. as the class of continuous functions from
I into X with initial value in A. If l is an interval of R and if is an almost
surely right continuous process, the preceding discussion for continuous
processes has the obvious modification in which L, is replaced by the space
of right continuous functions from I into X.
The class of finite-dimensional distributions of a stochastic process
satisfies two consistency conditions: (a) if A EX"

p(t,, ...,t"+1;A x X)=p(tt, ...,t"; A) (10.3)

and (b) if t; , ... , t ; , is a permutation oft, , ... , t", the measurep(t,, ... , t" ; )
becomes the measure p(r', ... , t ; ) when the coordinates of X" undergo
the same permutation.

Kolmogorov's Construction of Canonical Processes (see the Historical


Notes to this Section)

Let (X, X) be a Polish space coupled with its or algebra of Borel sets. Then
(X", 3-") is for n z I also a Polish space coupled with its a algebra of Borel
sets, and therefore (Appendix IV. 11) every measure on '" is inner regular.
Kolmogorov showed that for an arbitrary parameter set I, state space
(X,.-), and arbitrary finite-dimensional distributions for this choice of
parameter set and state space, satisfying (a) and (b) above, there is a canonical
stochastic process with L' = X' and the given finite-dimensional distri-
butions. (Kolmogorov had X = R, but the above inner regularity property
was all his proof needed.) His method of proof was to define P on the
algebra U5{' ......i(Q) by
I'{ [.Y(t,)...... Y(t"))EA} = p(t,, ...,t.,A) (AE.9C") (10.4)

and then to show that P was countably additive. The extension of P to A


followed by completion makes P the measure on function space inducing
the desired process

Specification of a Stochastic Process and Analysis of Its Sample Functions

A stochastic process is usually specified (i) by its state space, parameter


set, and finite-dimensional distributions, and sometimes in addition (ii) by
404 2.1. Fundamental Concepts of Probability

such further properties as continuity, progressive measurability, and so on.


In discussing process sample functions all probabilities should be derived
from these specifications, and if there is a choice of probability measure
space, process sample function probabilities should not depend on the
choice. In analyzing the meaning of this admonition suppose first that there
are no specifications under (ii) above. Let (0, ,F, P) be a probability space,
and let {x(t), t E 1 } be a process on this space with arbitrary state space
(X, .fl and specified finite-dimensional distributions If tl, ... , t.
are parameter values, if AE. , and if A = {[x(11), . . . then
P{A} is determined by the finite-dimensional distributions; so these dis-
tributions determine P on the algebra of all such sets A and so determine
P on the a algebra F' = F {x(t), t c- I } generated by this algebra. Finally
the finite-dimensional distributions determine P on the class of F sets
obtained by completing the restriction of P to .'. That is to say in the
notation of the beginning of this section with 6 = X', the finite-dimensional
distributions determine P on -' ( ). A set A not in ¢ '(A*) but defined
b conditions on sample functions, that is, A = 0-'(A) with A not in
', may be measurable for some choices of (i2, .9r, P) and with the
specified finite-dimensional distributions, but if so, P{A} may depend on
the choice. (See Section 11 for an example of this possibility.)
This discussion is modified in the obvious way if there are specifications
prescribed under (ii) above. For example, suppose that I is an interval of
R, that (X, X) = (18, .(18)), and that the process x(-) is almost surely con-
tinuous. Define the associated continuous canonical process, setting L =
6c. The a algebra A is now a a algebra of subsets of 6,, and new sample
function properties can be analyzed. For example, it is now true that if
I' is a countable dense subset of 1 and if

A = {su?o 5 1 then R = 5 1} E.A and, up to a null set,

( 1)

= {su1xti <
fE
_{supx(t5 1} E '( );
111 111

so P{¢-'(R)} is defined and equal to P{R} no matter what the choice of


(0, 9, P) and the process as long as the process has the given finite-
dimensional distributions and is almost surely continuous.

11. Choice of the Basic Probability Space


In each example is a stochastic process with parameter set I and state
space (X, X) on the probability space P).

EXAMPLE (a). Let I = Z', (X, ") = (18, .(18)), and let x(0), x(1), ... be
mutually independent random variables. The theorem that the probability
12. The Hitting of Sets by a Right Continuous Process 405

of convergence of Eo x(n) is either 0 or I does not depend for its validity


on the choice of probability space, and the value of this probability, 0 or
I as the case may be, depends only on the distributions of the random
variables, that is, on the finite-dimensional distributions of the process

EXAMPLE (b). Let I = 0 = X = 08+, let X = R(I), let .F be the a algebra


of subsets of i2 measurable for the completion P of a probability measure
on R(lV ), and suppose that all singletons are P null. Define .fi(t) = F
for t e R+, and define

l ift=w,
x(t, w) _
0 ift#w, x ff8+.

The processes and are progressively measurable.


Moreover these two processes have the same finite-dimensional distributions,
and in fact is a standard modification of Since no process
sample function is continuous and every process sample function is
continuous, an assertion involving the value of the probability of a sample
function continuity property for a progressively measurable process must
involve specification of the process beyond its finite-dimensional distri-
butions. The hypotheses of Theorem 9(b) and (c) are satisfied by both
and so for each process that theorem, whose
conclusion is trivial for these processes, asserts that if A is an analytic subset
of X = R', the probability that a sample function ever takes on a value in
A, that is, the probability that a sample path hits A, is well defined. If A
is the set R - (0), this hitting probability is I for the process and 0 for
the process. Thus in Theorem 9(b) and (c) if the finite-dimensional
distributions of are specified, the probability of hitting A depends on
which progressively measurable process with those finite-dimensional dis-
tributions is chosen.

12. The Hitting of Sets by a Right Continuous Process


ExAMPLE (a). If {x(t), re 08+ } is a right continuous process with a topological
state space and if c > 0, the probability that an path hits an open set B
at some time 5 c is the probability of the set

U {x(t) e B} = U {x(rc) e B}. (12.1)


tSr 05r51
r rational

Since only countably many parameter values are involved on the right-hand
side the finite-dimensional distributions of the process determine the desired
hitting probability. Thus for a right continuous (or almost surely right
continuous) process the probability of hitting the open set B by time c
406 2.1 Fundamental Concepts of Probability

does not depend on the choice of almost surely right continuous process
with the given finite-dimensional distributions.

EXAMPLE (b) (The hitting of an F, set by a continuous process). If {x(t),


t E R+ } is a continuous process with a metric state space, if A is a closed
subset of the state space, and if B. for n >_ I is the open set of points at dis-
tance < 1/n from A, the probability of hitting A by time c is the probability
of hitting every B. by time c, and this probability is that of the set

U (x(t)EA)= n U {x(rc)EB,}. (12.2)


f5[ n=1 OS,5l
r rational

Again only countably many parameter values are involved on the right-hand
side. If A is a countable union of closed sets, the set A is hit by time c if and
only if some summand is. Thus for a continuous (or almost surely continuous)
process with a metric state space the probability of hitting an F. set by time
c does not depend on the choice of almost surely continuous process with the
given finite-dimensional distributions.

These examples should be compared with the Section 11 examples which


show that two progressively measurable processes with state space the line
and identical finite-dimensional distributions may have quite different prob-
abilities for hitting an open set. The following discussion shows that the
hypothesis of right continuity of a process prevents such a circumstance
for the hitting of analytic subsets of a Polish state space.
Let (X,. ) be a Polish space together with its Borel subsets, let {x(t),
.F(t), t E I } be a right continuous adapted process with (X, 9C) as state
process, on the probability measure space (0,,F,P), and let be the
associated right continuous canonical process with the same state space and
finite-dimensional distributions, on the probability measure space (CL A, P).
As usual it is supposed that all process measures are complete. Suppose
that each a algebra has been enlarged if necessary to make .-F(0) contain
the null sets. Define 34 (t) as the a algebra generated by {z(s),s < t} and
the P null sets. The processes and {.i( ), . () } are progressively
measurable according to Section 2; thus (by Theorem 9) if A' is a subset
of W x X analytic over -4(I&) x I, if c > 0, if

A = {w: (t, x(t, (!))) E A' for some t < c}, (12.3)

and if A is defined similarly for then A c ,F(c) and A e.?(c). Moreover


since A = t-1(A) (notation of Section 10), it follows that P{A} = P{A}.
Thus P{A} depends on the finite-dimensional distributions and the right
continuity hypothesis but not otherwise on the choice of P) and
In particular, if A is an analytic subset of X and A' = [0, c] x A, it follows
that the probability of the set
13. Measurability versus Progressive Measurablity of Stochastic Processes 407

{w:.x(t,w)eA for some t < c} (12.4)

depends only on the finite-dimensional distributions and the right continuity


hypothesis.
If the process is supposed only almost surely right continuous, we
can omit the P null subset of C corresponding to non-right continuous
sample functions to get the same results as in the right continuous case.
If in (12.3) and (12.4) the restriction on t is t < c instead oft < c, all of
the above results remain valid, in fact can be slightly sharpened since the
sets in (12.3) and (12.4) are then in F - (c).

13. Measurability versus Progressive Measurablity of


Stochastic Processes
Let (12, .F, P) be a probability space. We shall need a definition of con-
vergence in measure for random variables from this space into IB when the
random variables may have the values ± oo with strictly positive probability.
In this context we shall say that a sequence x, of random variables converges
in measure to x if the sequence arctan x, converges in measure to x in the
usual sense, equivalently, if

lim E{arctan x - arctan 0.

The metric

d(x, ') = E(I arctan x - arctan y } (13.1)

on the space S of extended real-valued random variables has the property


that a sequence converges in the metric if and only if the sequence converges
in measure in the above sense. This metric, together with the identification
of two random variables whenever they are equal almost everywhere, makes
S a complete metric space. Observe that if a sequence in S converges in
the metric (13.1), then some subsequence of the corresponding random
variables converges almost everywhere, in the topology of A. There is
therefore a sequence c. of strictly positive numbers such that if d(x,
x x almost everywhere. We use here the standard abuse of
language, leaving it to the reader to decide when a symbol represents a
function on S2 and when it represents a point of S, that is, an equivalence
class of functions. Observe that if we wished to treat only random variables
bounded in modulus by 1, the L' metric could be used instead of the metric
(13.1). In fact one way to carry out the work of this section is to replace
each random variable x by (2/it) arctan x and thereby make all random
variables bounded in modulus by 1.
408 2.1. Fundamental Concepts of Probability

In the present context a process {x(t), t e R') with state space 68 defines
a function fi'x(.) : t r- x(t) from IB+ into S. This map taking a process into a
function is one to one and onto if two processes are identified when they
are standard modifications of each other.

Theorem. Let (12, . , P) be a probability space, and let {x(t), t e l l } be a


process on the space, with state space A.
(a) The process has a measurable standard modification if and only if
the function fl., is Bore! measurable.
Suppose that the probability space is provided with a right con-
tinuous filtration (.F (t), t e R+ } for which .5(0) contains the null sets.
(b) If the process is adapted and measurable, it has a pro-
gressively measurable standard modification.
(c) If x 12 relative to the completion of the
measure 1, x P, there is a progressively measurable process such
that

P{x(t) = x'(t)} = I for 1, almost every t. (13.2)

Proof of (a). (i) If is a measurable process, then fix(.) is Bore! measurable.


Since each measurable process is determined by a function (t, w))-+
x(t,w) and since pointwise convergence of a sequence ne+}
implies pointwise convergence of the corresponding sequence { fx1(.)' n e 71+ },
it is sufficient to show that if has the form
n
x(t, w) _ Y_ ct 1AJ(t)1 At(w)
I

with At, ... , A. disjoint Borel subsets of I8+ and Ate., then #,(., is Borel
measurable. This Borel measurability is trivial.
(ii) If fix).) is Borel measurable, then has a measurable standard modi-
fication. To show this, it is sufficient to show that the class of functions
from OB+ into S corresponding to processes which have measurable standard
modifications includes the continuous functions and is closed under con-
vergence. Suppose first that fx,.) is continuous. Define

Ok(t) =.12-k for (j - l)2-k < t < j2-k (j= 1,2 ....).

Since fl., is uniformly continuous on compact intervals, there are posi-


tive integers k, < k2 < ... such that (see the beginning of this section)
d[x(s),x(t)] < c if Is - tI < 2-k and s v t < n; so for each fixed tin O8+

lim x[4k.(t)] = x(t) a.s. (13.3)


13. Measurability versus Progressive Measurability of Stochastic Processes 409

If we define x'(t) = lim sup, x[4kn(t)] the process is a measurable


standard modification of Next suppose that n e Z } is a sequence
of processes on (52,,x', P) with the property that each function is Borel
measurable, that has a measurable standard modification, and that
limn- flan,.) = P.,.) exists in the S metric on 08+ ; that is, limn-m x(t)
exists in measure for all t. We can suppose that has been chosen to be
measurable. Define by

min {k: d [x(t), xk(t)] <

The set {t: mn(t) =j) is a Borel subset of R, and for each t

d[x(t), xmn(t)] < en;

so

urn x(t) a.s. (13.4)

and The process defined by

(1, w) -s xj(t, (j)) lAnj(1)

is measurable because it is the product of two measurable processes, so


summing over j we find that the process is measurable. The process
lim is therefore a measurable standard modification of
and the proof of (a) is complete. o

Proof of (b). There is no loss of generality in proving (b) if we suppose,


after replacing by that S 1, and we shall assume
that all processes are bounded in this way throughout the proof of (b).
To each process then corresponds a process

x '(t) = E{x(t)I.F(t)}.

Recall our convention that x°(t) is a version of the indicated conditional


expectation and that is therefore determined only up to a standard
modification. The process is adapted and is a standard modification
of if the latter process is adapted. To prove (b), it is sufficient to show
that whenever is measurable, in particular, whenever /3X,., is a Borel
measurable function, the process has a progressively measurable
standard modification. It is therefore sufficient to show (i) that such a
modification exists for if it exists when /x,.) is continuous and (ii) that
such a modification exists for if it exists for each process of the sequence
and /3r,.,.
410 2.1. Fundamental Concepts of Probability

Proof of (b)(i). If fist.) is continuous, the conditional expectation con-


tinuity theorem can be applied to (13.3) to obtain, in view of the right
continuity of

lim E{x(t) (t}} = x°(t). a.s. (13.5)

When 2-"^ < e, the nth term on the left defines a process adapted to and
progressively measurable relative to the filtration ,f( + e). Hence if the
limit in (13.5) is replaced by the limit superior, the resulting process is
progressively measurable relative to .IF(- + e) for all strictly positive e and
so (Section 1.2) is progressively measurable relative to and this process
is a standard modification of
Proof of (b)(ii). Apply the conditional expectation continuity theorem
to (13.4) to obtain

lira E{x.,(,)(t)1F x°(t) a.s., (13.6)

and repeat the reasoning just used, thereby showing that the process defined
when the limit in (13.6) is replaced by the limit superior is a progressively
measurable standard modification of The proof of (b) is now complete.

Proof of (c). If x is a function


(t, ,,)--. x1(t, w) on e8+ x f? which is measurable relative to -4(R) x F,
that is, for which is a measurable process, such that
x P almost everywhere. Hence (Fubini)

P(x(t) = x,(t)} = 1 (for 1, a.e. t).

According to (b). the process x, has a progressively measurable standard


modification and this is the process demanded by (c).

14. Predictable Families of Functions


Let (i2, ; fi(t), t e 1 } be a filtered measurable space. The concept of a
predictable function family strengthens that of an adapted family in a way
suggested by the name.

Discrete Parameter Case, I = Z'

A sequence {x(n); F(n), n c- Z' } of functions from Q into a measurable


space is called predictable (relative to if x(0) is .F(0) measurable and
if x(n) is F (n - 1) measurable for n > 0.
14 Predictable Families of Functions 411

Continuous Parameter Case, I = R'

Consider the class of adapted families {x(t); .fi(t), t e R' } of functions from
0 into R for which every sample function w) is left continuous, that is,
for which the family is left continuous. (Left continuity at 0 is vacuously
satisfied by every function on R'.) The smallest a algebra of subsets of
R' x Il making the function (t, w) t -x(t, w) measurable for every such
left continuous family is called the predictable a algebra, and the sets
in this or algebra are called predictable. A family { y(t), t e R' } of functions
from fl into an arbitrary measurable space is called predictable if the function
(1, w) i-+y(t, co) on the space R' x 0 coupled with its predictable a algebra
is measurable. A predictable family is necessarily adapted to the given
filtration. Since left continuity of a function family with state space (R, R(R))
implies progressive measurability (Section 2), the predictable sets are pro-
gressively measurable, and a predictable function family is progressively
measurable. The predictability definition implies that if is predictable,
if a > 0, and if x,(t) = x((t - a) v 0), then the family is also predictable.

EXAMPLE (Continuous parameter case). By definition of predictability a left


continuous adapted family with state space (R, M(R)) is predictable. More
generally a left continuous adapted family with state space a metric space
coupled with its Borel sets is predictable. In fact, whenever f is a continuous
function from the state space into R, the family (f is predictable,
and therefore the set {(t,(o): x(t,w)eA} is a predictable set whenever A
is a Bore] subset of the state space; so is predictable.

Generators of the Predictable a Algebra

If 0 5 a < b 5 + x and if A e F(a), the subset ]a, b] x A of R' x () if


b < + x or ]a, + oo [ x A if b = + x is a predictable set, as is the product
set {0} x An when A,E.F(0). The class of finite unions of sets in the class
r of product sets of these two types is an algebra, and we now show that
the a algebra I generated by r is the predictable a algebra. To see this,
observe first that if x(t) = 1,,l(t)4 with 0 an JF(0) measurable function
from 0 into R, or if 0 5 a < b and if x(r) = 11,.bl(t)o with 0 an .f(a) measur-
able function from fl into R, then the family is 4 measurable. Second,
observe that if { is a left continuous adapted family of functions
with state space (R,-V(R)) and if
4,'

y;,(t) = y(0)I1oi(t) + Y y((i - 1)2-")11(J-1)2-".J2 -"10),

i=t

then the family y.(-) is If measurable because it has just been shown that
each summand family is. Finally lim, so is I measurable,
412 2.1. Fundamental Concepts of Probability

and therefore I is the predictable a algebra, as asserted. The class of sets


analytic over the paving r therefore (Theorem 4 of Appendix I) includes
the predictable a algebra (so is the same as the class of analytic sets over
the predictable a algebra).
For some purposes it is convenient to replace the generating family r
by a generating family r' involving compact parameter intervals. Define
r as the class of product sets [a, b] x A with 0 5 a < b, A e.F(0) if a = 0
and AE U,<,.f(t) if a> 0. The a algebra generated by r' includes the
predictable a algebra because

]a, b] x A= U a+ 1, bl x A [a < b, A e .te(a)],


=t L n JJ
(14.1)
{0} x A= n x0,11 x A
n=t L n

Here the sets on the left are in r and those on the right are in the a algebra
generated by r'. Conversely, just as trivial a representation of r' sets in
terms of predictable sets shows that the a algebra of predictable sets includes
r'. Thus the a algebra generated by r' is the predictable a algebra. The
representation (14.1) shows that the class of analytic sets over r' includes
the predictable sets. We leave to the reader the similar trivial argument
showing that if 0 5 a < b 5 + oc, the predictable a algebra is the a algebra
generated by the class r" of product sets [a, b[ x A with A e ,T(0) if a = 0
and Ae U, .(t) if a > 0. The class of finite unions of these product sets
is an algebra. The class of analytic sets over r" includes the predictable sets.
The predictable a algebra is the smallest a algebra making the function
(t, (n) (o) measurable whenever is an adapted continuous
family with state space (l8,.i(R)). To prove this, we need show only that if
G" is a a algebra of subsets of R' x Q making every such family measurable,
then G" contains r. This inclusion follows from the fact that the indicator
function of a compact interval of I8' is the limit of a decreasing sequence
of continuous functions on hr vanishing outside an arbitrarily small neigh-
borhood of the interval.
Chapter II

Optional Times and Associated Concepts

1. The Context of Optional Times


Let (Q, 4r ; .fi(t), t e 1) be a filtered measurable space. If I does not have a
last element, extend I to I u { + oo }, where + oo is a new element ordered
after every element of I, and define .F(+oo) as any sub a algebra of S
containing Y<+m.f(t). The choice of .5(+ oo) within these limits is usually
irrelevant. If I has a last element, that element will be denoted by +oo in
this section. The most common choices off are the set 1+ (discrete parameter
case) and the set R + (continuous parameter case). The index set I is thought
of as representing a set of time points, and then represents the class
of events up to time t. The problems of defining what is meant by a random
time T corresponding to the arrival time of an event whose arrival is deter-
mined by preceding events and of defining the class .5(T) of preceding events
are solved by the following definitions.

Definition. An optional rime (also called a stopping time or a Markov time)


T is a function from Q into I u { + co } satisfying

{T:5 c} e 9(c) (1.1)

for c < + oo. This relation is trivially true for c = + co.

Definition. If T is optional, denote by .9(T) the a algebra of subsets A of


t for which
An{TSc}e.flc) (1.2)

for c 5 +oo. Equivalently, .9(T) is the a algebra of sets A in .F(+ oo) for
which (1.2) is true when c < + oo.

A function T from D into a countable subset of 1 u {+ oo} is an optional


time if and only if
{ T = c} e.flc) (1.ld)

for c in the range of T, and if T is optional, A e 9r(T) if and only if


414 2.11 Optional Times and Associated Concepts

An {T= c}e*(c) (1.2d)

for c in the range of T. In particular, an identically constant function T - t


on S1 with value in I u { + oo } is an optional time and in this case *(T), as
defined above, reduces to the specified a algebra *(t).

*(T) in the Continuous Parameter Case

If I = IR+, the conditions

IT < c} e F(c) (1.1')

AnIT <c}e*(c) (1.2')

for all finite c are equivalent to (T :!g c} a*(c+) and A n IT 5 c} a *(c+),


respectively, for all finite c. That is, (1.1') is the condition that T is optional
for the family and (1.2') together with Ae*(+oo) is the condition
that Ac 9 +(T). If T is optional, then T + e is optional for positive a and
(T) c F (T + e). Define *(T+) = f 5,o. (T+ E). Then A c- 9(T+) if
and only if A n {T < c} e*+(c). It follows that .4t(T+) = *+(T).

Stochastic Intervals (Parameter Set IB+)

If S and T are optional with S < T, the stochastic interval QS, TD, a subset of
I8' x 0, is defined by

[S, TI = { (t, w) : S(w) < t < T(co), t c - I.

In particular, we write ITI for the interval [T, TI, the graph of T (but
observe that this graph is a subset of 68+ x 0; if T =_ + oo, this graph is the
empty set). The open stochastic interval IS, TQ and the half-open ones are
defined in the obvious way.
A stochastic interval of the form IS, TI is a predictable set because each
function w) is left continuous. The classes r and r' of generators of
the predictable a algebra, as defined in Section 1.14, can be described in
terms of stochastic intervals. For example, r consists of the graphs JSD with
S = 0 on an * (0) set and S = + oo elsewhere and of the stochastic intervals
IS, T] with S = a on an *(a) set A and S = + oo elsewhere, T = b on A
and T = + oo elsewhere.

Optional Times on a Filtered Probability Space

Suppose that the given measurable space is a probability space, and that
each a algebra .(t) of the filtrations contains the null sets. Then if T is
optional, * (T) also contains the null sets, and if Tt and T2 are functions
2. Optional Time Properties (Continuous Parameter Context) 415

from 92 into I u { + oc } which differ only on a null set, Tt is optional if and


only if T2 is. Moreover if both are optional, then ,y (Tt) = .F(TZ). Thus
under this hypothesis on we can ignore null sets in discussing simple
relations between optional times and the a algebras they determine.

2. Optional Time Properties (Continuous Parameter Context)


In this section the commonly used properties of optional times and their
associated a algebras on a filtered measurable space l E 68 +) are
listed, and proofs of those properties not immediate consequences of the
definitions are given. The corresponding properties in the discrete parameter
context of the parameter set 71+ will be obvious.
(a) If S is optional, S is .F(S) measurable.
(b) If S is optional and if T is a random variable for which T >_ S and
T is .F(S) measurable, then T is optional.

EXAMPLE (bl). If S is optional and a is a positive constant, then S + a is


optional.

EXAMPLE (b2). If S is optional define [S]" by

j2-" when (j - 1)2-":5 S <j2-n,


[S]. +oo when S= +cc.
Then [S]" >_ S, with strict inequality where S is finite valued, and the
sequence [S]. is a monotone decreasing sequence of optional times with
limit S.

(c) If S and T are optional, then S v T and S A Tare optional.


(d) If S and T are optional and S< T, then F(S) c .F(T); if S < T,
then ,F(S+) c .F(T).
In fact, if AE.F(S) and S < T, then

An{T<c}=[An{S<c}]n{T<c}E,y(c); (2.1)

so AE.01'(T). Thus the first assertion of (d) is true. To prove the second
assertion, replace IS < c} by IS < c} in (2.1).
(e) If S and T are optional .F(S) n.F(T) = F(S A T). In particular,
{S<c}E.F(SAc).
(f) If S and T are optional, the sets IS < T}, IS :5 T}, IS = T} are in
,F(S A T).
In fact, since
{S<T}n{SAT<c}= U {S<rc}n{T>rc} (2.2)
o<rst
r rational
416 2.11. Optional Times and Associated Concepts

and IT > rc} =Q - IT :5 rc} e,(c), it follows that IS < T } e,(S AT).
Interchanging S and T, we conclude that the set {T < S} and therefore also
the complementary set IS ;-> T} are in ,(S A 7). Finally

IS =T}={SST}n{TSS}e,(SA T).
(g) If S and T are optional and if A e,(S), then

AnIS 5 T}e,(SA T).


In fact A n IS 5 T} n IS A T 5 c} e,(c) because this intersection is the
same as A n (S:5 T} n {S:5 c} in which expression both A and IS 5 T}
are in ,(S). Property (g) generalizes the defining relation of ,(S). The
same proof shows that IS 5 T} in (g) can be replaced by IS < T} or IS = T).
(h) Let S be optional and finite valued, and let T be a positive (5 + oo)
random variable. Then optionality of S + T implies optionality of T relative
to ,(S + ). Conversely, optionality of T relative to ,(S + ) implies
optionality of S + T for fl-) if F(-) is right continuous.
In fact, if S + T is optional for ,(),

IT Sc}=IS +TSS+c}e,(S+c) (2.3)

by (f); therefore T is optional for .F(S + ). Conversely, if T is optional


for F(S + ),
{S+T<c}= U {S<a,T<b}. (2.4)
a+b«
a,b rational

For each pair (a, b) the brace set on the right is in ,(c) because by hypothesis
IT < b} e.F(S + b), that is,

{T<b}n{S+b<a+b}e,(a+b)c,(c). (2.5)

Hence S + T is optional for if this family is right continuous.


According to property (h) if is right continuous and S is finite valued
and optional and ,S(t) is defined as ,(S + t), then optionality of T for
implies that S + T is optional for and ,(S + T) =,s(T).
(i) Let { T,,, n e Z 'I be a sequence of optional times.
(i1) supZ. T is optional.
(i2) infZ. T is optional
(i3) lim inf, T. and lim sup,, T. are optional for.F
(i4) If lima . T. = T = infZ. T , then ,+(T)=no- ,+(T.)
(i5) If T. T T [T0 I T] and To = T for sufficiently large n = n(w), then
T is optional and F(T) = yo ,(T.) v m = no .F(T,J].
3. Process Functions at Optional Times 417

Only (i4) and (i5) need comment. In (i4) we need only prove that
no c .f+(T), and this inclusion follows from the fact that if
AEno .F+(T"), then

An{T<c} = UAn{T" <c}E.F(c).


0

To prove (3), observe that if T. T T, then Yo .F(T") c #(T) trivially,


and the inclusion is true in the other direction because if AE.F(T), then
A n {T = T"} E F(T") by (g) and summing over n yields A E Yo g(T1). In
(6) if T. I T, the limit is optional because {T 5 c) = Uo {T" 5
Moreover in this case no .f(T") 9(T), and the inclusion is true in the
other direction because if AE A0 F(T"), then

An {T5c}=U[An{T"Sc}]E.flc)
0

so that A E
(j) Suppose that { fo(r), t E I } is a filtration of a probability space

(52,F, P), that .Fo(+oo) is a a algebra satisfying YR..4ro(t) c ,f(+oo)


c .F, that F(t) is the a algebra generated by .Fo(t) and the null sets, that T is
optional, and that A E flT). Then there are an .moo optional time
To and a set Ao E .moo (To) such that T = To almost surely and that A differs
from A0 by a null set.
For j = 0, 1, ... , + oo define M"t as an .Fo(j 2-") set differing from the
.F(j2-") set {[T]" = j2-"} by a null set, and define S. = j2-" on M", -
Uk<tM"r. Then S. is an optional time, S. is a decreasing sequence,
and the limit To is an .go optional time equal almost surely to T. If
A E 9(T) and if T = + oo everywhere on A, let Ao be a set in go(+ co)
differing from A by a null set. The set I TO = + co } n An is then in .F0' (To)
and differs from A by a null set. If AEF(T) and if T < +oo everywhere on
A, define the fl-) optional time T' by setting T' = Ton A and T' = + oo
elsewhere. We have shown above that there is an .9o optional time To
equal almost surely to T'. The set { TT = To < + oo } is in .F (To) and differs
from A by a null set. If A is an arbitrary member of 9(T), apply the two
results just obtained to the sets A n { T = + oo } and A n { T < + oo } to
complete the proof of (j).

3. Process Functions at Optional Times


If {x([), t E I } is a family indexed by I of functions on a space Q and if T
is a function from S2 into a set including I, we write x(T) for the function
to -x(T(w), co), defined on the set (c o: T(cu) e 1).
418 2.11. optional Times and Associated Concepts

Theorem. If {x(t)..F(t), tE l8+ } is a progressively measurable family of func-


tions and if T is optional, then x(T) is.f(T) measurable.

According to our conventions, the function x(T) is defined on the set


{ T < + x, } which [Section 2(a)] is ,(T) measurable. To prove the theorem,
let (S ',.F') be the state space of the family and suppose that A'E''.
The theorem asserts that { T < + oo, .r(T) E A", E .F (T), that is, for c >_ 0.

(3.1)

Let F, be the class of intersections with { T5 c} of .f(c) sets, and consider


the three measurable spaces

(;T<([0.c] x52.:-f([0.c])x. (c)), (a',. ').


The condition (3.1) is that the restriction to T< c} of x(T) be measurable
as a function from the first to the third space. This restriction is the composi-
tion of the two functions cu i-+(T(cu). (u) and (1, a0 i-* x(t, (u). respectively.
from the first space to the second and the second space to the third. The
first function is measurable because by optionally of T the inverse image of
a product set [a, b] x A is in.F(c) when a < h < c and A e .F(c). The second
function is measurable by definition of progressive measurability. Hence
the composed function is measurable.
Special Case. If ;x(t). t E R' } is a measurable family of functions on the
measurable space (52, . ) and if T is a measurable function from 11 into A',
then Theorem 3 with .t(t) = t for all t implies that x(T) is measurable.
Observation. If the process {x(t), .fi(t), t e L'8+ i on a probability space is
nearly progressively measurable and if .F(0) contains the null sets, then
Theorem 3 implies that x(T) is measurable whenever T is an optional time.

Predictability of a Family of Functions at a Predicted Time (Continuous


Parameter Context. Arbitrary State Space)

If ;.a( is a predictable function family and if S. is an increasing


sequence of optional times with limit S, then x(S)I,s<,, is a Y;
measurable function. To prove this assertion. Let I be a compact interval
of P. and suppose first that the state space is (1,. (I)). The assertion is
trivially true for this state space if is a left continuous family. Since the
property to be proved is invariant under convergence of function families.
the assertion is true in general for this state space. The assertion is therefore
true for an arbitrary state space because whenever 0 is a measurable function
from the state space into (1. (1)), the function di[x(S)]I,s, +x, is Y,
measurable, and so x(S)I,s<,_, is.
4. Hitting and Entry Times 419

4. Hitting and Entry Times


Let (Q, t c- 1) be a filtered measurable space, with l either 7L + or R'.

If A" is a subset of I x S2, the hitting time of A" (by t e I }) is defined as

inf{t > 0: (t,w))EA"} (tel). (4.1)

Here and below the infimum of the empty set is defined as +oo. If
{x(t), t c- I) is a family of functions on SZ with arbitrary state space (X,.X)
and if A' is a subset of I x X, the hitting time of A' (by { [t, x(t)], t E 1) ) is
defined as

inf{t>0: [1,x(t))eA'} (te1). (4.2)

Finally, if A is a subset of X, the hitting time of A (by {x(t), t e l }) is defined as

inf{1 > 0: x(t,w)eA} (tel). (4.3)

In each case the entry time of the set in question is defined in the same way
as the hitting time except that "t > 0" is replaced by "t >_ 0." Observe that
If A' = I x A, the hitting and entry times of A' reduce to those of A and
that if A" = { (t, uw): [t, x(t, (o)] a A", the hitting and entry times of A"
reduce to those of A'.
The analysis of measurability of hitting and entry times is trivial when
I= ll+. For example, if A e" , it is clear that the entry and hitting times of
A by are measurable and in fact are optional for From now on
we shall therefore assume that I = H. Observe that if T' [T"] is the entry
[hitting] time of A" and if c > 0, then

{T' < c} = {a): (t,w)EA" for some t < c},

{T" < c} = U 10): (t, (0) E A" for some t Ern, cl },

and the corresponding observation is valid for the entry and hitting time
of A' and A. The evaluations in (4.4) and the corresponding ones for A' and
A make it easy to deduce optionality of entry and hitting times in a suitable
context, as follows.

Theorem. Let (f),f,P;.f(t),teH+) be a filtered probability space, and


suppose that for each t the restriction of'P to .f (1) is a complete measure.
(a) If A" is a subset of H+ x S2, analytic over the class of progressively
measurable sets, the entry and hitting times of A" are optional for
420 2.11. Optional Times and Associated Concepts

(b) Let be a progressively measurable process on the proba-


bility space, with an arbitrary state space (X, fl, let A' be a subset
of 68+ x X, analytic over V(I8+) x X, and let A be a subset of X
analytic over X. Then the entry and hitting times of A' and A by
{ [t, x(t)], t e OB + } and {x(t), t e R+ }, respectively, are optional for
In particular, if the state space is Polish and if the process is
right continuous, then the distributions of these times depend on A',
A, and the finite-dimensional distributions of together with the
assumption of right continuity but do not depend on the choice of the
probability space.

To prove the theorem, recall that the sets on the right in (4.4) and the
corresponding ones for A' and A were shown in Section 1.9 to be in PF(c);
so the entry and hitting times are optional for The last assertion of
(b) is trivial because the probability of hitting and entry in the stated con-
texts, by time c, does not depend on the choice of probability space according
to Section 1.9.

Generalization

For n < + oo let T. be the nth entry time of A"; that is,

inf it > 0: (t, w) a A" for at least n values of t}.

Then T, is the entry time of A" and under the hypotheses of (a) is optional
for If n < + on and if T. is optional, the stochastic interval T,,, + oo Q
is progressively measurable. Hence is the entry time of the progressively
measurable set A" n ] T,,, + oo Q and so is optional. It follows that T, , T2,. . .

and therefore also T,,, = lim, T. are optional. The corresponding assertion
is true for A' and A. Further, if .f(0) contains the null sets, part (a) of the
theorem as just generalized is valid if A" differs by an evanescent set from
one as described in (a), and part (b) of the theorem as just generalized is
valid if is indistinguishable from a progressively measurable
process.

Last Hitting Times

We use the notation introduced at the beginning of this section. The last
hitting time of A" by { (t, w), t e 1) is defined as

S" = sup It > 0: (1,w)eA"} (4.5)

under the convention that this supremum is 0 of the set in question is empty.
The last hitting times of A' and A, in the respective contexts of (4.2) and
S. Application to Continuity Properties of Sample Functions 421

(4.3), are defined in the obvious way. If c > 0, let S(c) be the hitting time
of A" by { (c + t, w), t e R ) . The function S" is not optional except in
trivial cases, but {S" > c} = {S(c) < +oo} so that measurability problems
for last hitting times can be reduced to dual problems for hitting times.

5. Application to Continuity Properties of Sample Functions


Let {x(t), .fi(t), t e R+ } be a process on (S2, .F, P) with a Polish state space.
Let be a metric for the state space, define the strict right oscillation
0,* Q, w) of the sample function w) at t by

O!(t, w) = lim sup d[x(rt, w), x(r2, w)],


a-o t<,,.?,< t+a

and define the right oscillation by

O,1 (t, (v) = lim sup d[x(t, w), x(r, w)].


1<.<t+6

If is extended real valued, define

z,(t, w) = lim sup x(r, (o),


'ti
and define x, (t, to) as the corresponding limit inferior. (Recall that according
to our conventions r > t in these superior and inferior limits.) We shall not
need an analysis of the left limit properties of sample functions.

Theorem. Let {x(t), .fi(t), t e R+ } be a progressively measurable process with


a Polish state space. Suppose that is right continuous and that the restric-
tion of P to each a algebra .fi(t) is complete.
(a) The processes and are progressively
measurable. _
(b) If the state space is (Fl, M(R)), the processes and
are progressively measurable.

We prove the first assertion of (a) but omit the similar proofs of the other
parts of the theorem.
Fix c > 0, a > 0, and define .f(c -) = Yb<C 5(b), 9V = -4([0, c[) x
.F(c -), T(1, to) = inf {se [t, c[: O!(s,w) Z a}, for te[0,c[. Then

{(s,w): O!(s,w) z a,0 <s < c) = {(s,to): T(s,w) = s}. (5.1)

To prove that is progressively measurable it is sufficient to


prove that the function is I measurable, because then the set in (5.1)
is in W. Define
422 2.11. Optional Times and Associated Concepts

A(a,m) = }(s.r,,r2,co): d[x(r,,to),x(r2,w)] >_ x.


0<s<r,,r2 <(s+ I/m) A c}.
Then A(a, m) e. ([0, x f(c -). According to Theorem 5 of Appendix
I. the projection A'(a, m) of A(a, m) on (s, co) space is analytic over 1. as is
therefore also the set nk , nT=, A'(a - 1/k, m), which is the set in (5.1).
Project the latter set onto co space to see that for each value of s the function
T(s, ) is.t (c -) measurable. The desired measurability of follows from
the fact that co) is monotone increasing and left continuous, so that
n
T(t.w) = lim T((j - I)c/n, (o)1[ui
n-ac j=t

and each sum on the right is IN measurable.


Observation. One can derive from Theorem 5 the measurability of the 0
subsets corresponding to the sets of sample functions which have right limits
on an interval, or are right continuous there, or have oscillation at most some
special value there, etc. Observe, however, that in this discussion there has
been no assurance that another progressively measurable process with the
same finite-dimensional distributions will assign the same probabilities as
the given process to the subsets of its probability space discussed above.
For example, there has been no assurance that the finite-dimensional distri-
butions of the process are independent of the choice of (0,.x, P) with
the given finite-dimensional distributions. In fact examples in Section
1.1 1 show that such an assurance would be false. If however, is given as
a right continuous process and if either the family is already assigned
and has the desired properties or .f(t) is defined as the a algebra generated
by the null sets and.r (x(s),s < t}, then all the work in this section is appli-
cable in this stricter context, and (Section I.12) the probabilities mentioned
above will be the same for any other choice of with the given finite-
dimensional distributions as long as the process is right continuous, or even
almost surely right continuous.

6. Continuation of Section 5

Theorem. Let {x(t),. (t),teR`} he a progressively measurable process on


P) with a Polish state space. Suppose that is right continuous and
that the restriction offP to each a algebra F (t) is complete.
(a) If whenever T is a hounded optional time almost every sample
function has a right limit at T, then almost a very sample function
has a right limit at every parameter value.
7. Predictable Optional Times 423

(b) If whenever T is a bounded optional time almost every sample


function is right continuous at T, then the process is almost surely
right continuous.

We use the notation of Theorem 5. If e > 0, let TE be the entry time of the
set

{t,w): O*(t,w) z e}.

Then T is optional and (T (w), (o) is in this set when T(w) < + oo : so the
hypothesis of (a) implies that T A n = n almost surely for every n > 0, and
(a) follows. The proof of (b) is similar.

7. Predictable Optional Times


If (0,.F; .fi(t), t e R+) is a filtered measurable space, an optional time T will
be called predictable if there is an increasing sequence T of optional times
such that

T,ST, Tp<T on{T>O}, (7.1)

The sequence T is said to announce T. If there is a complete probability


measure P on .F and if F(0) contains the P null sets then if each condition in
(7.1) is true merely P almost surely, it follows that T is predictable according
to the following argument. Let A be the null set where at least one of the
conditions in (7.1) fails to hold. Then A e F(0), and if we redefine T. on A
as (T - 2-") v 0 or T A n, according as T is finite or not, the modified op-
tional times are still optional, and (7.1) is now satisfied. Thus under the
stated hypotheses on P and 9(0), an optional time equal almost surely to a
predictable time is itself predictable.
We leave to the reader the proof that the maximum and minimum of a
finite number of predictable times is predictable. The following properties
are slightly deeper. Let T be a sequence of predictable times, and let T,, be
a sequence announcing T,,.
(a) The optional time T. is predictable because it is announced
by the sequence

Imax T,,, ne/4I


I. /s" JJ

(b) The optional time T= T, is predictable if Ua {T = T} =f)


up to a null set because if 0,((o) = inf {k: Tb,(w) < T(w) }, the
sequence S. defined by
424 2 11 Optional Times and Associated Concepts

S.= Tmn l f$, =ml


m=0

announces T up to a null set.

The (7 Algebra .f(T- )

If T is a predictable optional time and if S. and T. are sequences of optional


times announcing T, then

IimS.ATm=Tm, U{S.AT.=TmS2
n=o

so

Y f(S A Tm)=.lt'(Tm)
-0 n=0

by Section 2(i5). Hence z) Ym=o F(1 ), and there must be


equality because S. and T can be interchanged. does not
depend on the choice of the announcing sequence T, and we denote this a
algebra by .f(T-). According to Section 3, if S is a predictable optional
time and if is a predictable process, then Sks<+n, is.f(S- ) measurable.

Predictable Filtrations

A filtration ;.f (t), t E R', of a probability space will be said to be predictable


if whenever T is a predictable time, t(T) = f(T-). Observe that this
condition implies that f (+ ac.) must have been defined as YF , .fi(t).

EXAMPLE A predictable process defined by a predictable optional time. If


T is an optional time, the process defined by setting AT(t) = I[T. wl(t)
(= 0 when T = + oc) is a right continuous adapted process. In particular.
if T is predictable and if T. is an increasing sequence of optional times
announcing T, the process IjTR_,Q,lO is adapted and left con-
tinuous, therefore predictable, and AT'n(t) = AT(t); so is
predictable

8. Section Theorems
Let (S ..f, P) be a probability space, and let A be a subset of R' x Q.
A section theorem is a theorem stating that under certain conditions on A
there is a function T with useful properties, from Q into A', such that a
8. Section Theorems 425

significant part of the graph of T lies in A. In applications to probability


it is desirable that T be optional relative to a given filtration of the probability
space.
The following context will be useful. Define an outer measure P" on Q by

infP{M: M A,MeF}.
Let A be the class of finite unions of subsets of R+ x Q of the form C x A
with C a compact subset of R+ and A in F. If A c R+ x fl, let n(A) be the
projection of .4 on 0, and define 1(A) = According to Appendix
11.5, the set function I is a Choquet capacity on R+ x 0 relative to A, and

d(I) R(R+) x F, n(sd(I)) = 31;


I(A) = P{n(A)} if Aesd(1).

Theorem. Let {Q, .F, P; .fi(t), t e R+ } be a filtered probability space. Suppose


that is right continuous and that ,F(0) contains the null sets.
(a) If Ae.V(R+) x F, there is an Jr measurable function T from 0 into
R+ for which

[T] c A, P{T < +oo} = P{zr(A)}. (8.2)

(b) If A is predictable and if e > 0, there is a predictable time T for which

[T]cA, P{T<+co}+s2P{n(A)}. (8.3)

Proof of (a). According to the Choquet capacity theorem, the set A is I


capacitable; so there is an A,, set 1), (that is, 1), is a countable intersection
of A sets) for which $, c A and 1($,) > 1(A) - 2-'; equivalently, P{n($,) }
> P{n(A)} - 2'. Similarly since the set A2 = A - (R+ x n(A,)) is in
9(R+) x 9, there is an Aa and E2 for which a2 c A2 and P{n(E2)} >
P{n(A)} - 2-2. Continuing in this way we find sequences A, =,4,,4,, .. .
and $,, B2, ... of subsets of A, in R(R+) x 9 such that for n z 1,

An+t = An - (R+ x n(E1)). P{n(E,)} > P{n(A,J} - 2-".

The set h = U'I$ is a subset of A, is in -4(R+) x ,F, P{n(E)} = P{tr(A)},


and by definition of A the set {t: (t, co) a h) is compact for every w. The
entry time T of $ is measurable because

{T5c}=n{An([0,c] xQ)}esd(.F),
and T obviously satisfies the other conditions prescribed in (a). o
426 2.11. Optional Times and Associated Concepts

Observation. If A is supposed merely in d(.$), only the first step in the


above proof remains valid, but since the number 2' in the first step can be
replaced by an arbitrary strictly positive s, there is an .F measurable function
Ton S2, the entry time of $t, for which (8.3) is true.

Proof of (b). Observe first that if A is a set in the algebra r'" generated by
r" (defined in Section 1.14), then (b) is trivial because the entry time T of
A is predictable and satisfies (8.3) with s = 0. Second, suppose that A e r;,
that is, A = n,,,4. is a countable intersection of I"" sets. Then the entry
time of A is predictable because it is the supremum of the sequence of entry
times of A1, A2, ... , and again (8.3) is satisfied with c = 0. The general case
is reduced to this special case as follows. According to (a), if ,4 is predictable,
there is an 9 measurable function S from D into A' for which QST c A and
P{S < +oo} = P{n(A)}. Define

1(E) = P{(0: (S(wU),(9)eB} (8.4)

for Ee r-. The set function 2 is a measure on the algebra r"" and therefore
(Hahn-Kolmogorov theorem) has a unique measure extension to the a
algebra generated by r-, that is, to the predictable a algebra, and (8.4)
is valid for the extension. The extended measure 2 is supported by .4, with
2(A) = P{n(A)}. According to the full statement of the Hahn-Kolmogorov
theorem, there is a set h. in r;' such that h. c A and i(h.,) >: P{a(A)} - E.
Let T be the entry time of h,. According to the result in the special case of
(b) already treated, QTD c h, and P{T < +oo} = P{rt(E,)); so PIT < +oo}
z 2(hd z P{n(A)} - c, and T therefore satisfies (8.3).

9. The Graph of a Predictable Time and the Entry Time of a


Predictable Set

Theorem. Let (0, .F, P; .fi(t), t e R+) be a filtered probability space, Suppose
that is right continuous and that F(0) contains the null sets.
(a) An optional time is predictable if and only if its graph is predictable.
(b) The intersection of a predictable set with the graph of its entry time
is a predictable set. In particular, the entry time of a predictable
set is predictable if the entry time graph is included in the set.

Proof of (a). If T is an optional time announced by T , then

QT]=(n T,,,+ooi-iT,+ooi)u({0} x {T=0}).


o J
10. Semipolar Subsets of R' x 0 427

The sets ] T,,, + oo [ and ] T, + oo [ are predictable because their indicator


functions define adapted left continuous processes. Hence [T] is a pre-
dictable set. Conversely, suppose that T is an optional time whose graph
is a predictable set. Then [Theorem 8(b)] for n z 1 there is a predictable
time T. such that P{ T. * T) < 1 In and that T. = + oo where T. # T. Hence
Um IT. = T} = n up to a null set so [from Section 7, property (b)] it
follows that T is predictable. 0

Proof of (b). If a set A is predictable and if T is its entry time, the stochastic
interval IT, + oo [ is predictable because its indicator function defines an
adapted left continuous process. Hence the set A - ] T, + oo [ = A n [ TI
is predictable.

Generalization

Let Q, 5, P, be as in Theorem 9. Let A" be a nearly predictable subset


of R' x 0, and let T. be the nth entry time of A",

T (w) = inf It e R' : (t, W) E A" for at least n values of t},

and suppose that the graph of T. is up to an evanescent set included in A".


Then we show that T,, T2, ..., TW = T. are predictable optional
times. We can suppose that A" is predictable and therefore progressively
measurable and apply Section 4 (Generalization) to find that these functions
are optional times. We need only show in addition that each time T. with
finite n is predictable. According to Theorem 9 the optional time T1 is
predictable. For n > 1 the stochastic interval ] +ooQ is predictable
because its indicator function defines a left continuous adapted process;
so T, is the entry time of the predictable set A" n ] 1, + oo [, and the
hypotheses of Theorem 9 are satisfied so T. is predictable.

10. Semipolar Subsets of ff8+ x S2


Let (S2, P; .fi(t), t e R') be a filtered probability space. Suppose that
.4r(-) is right continuous and that .4r(O) contains the null sets.
In discussing potential theory on R' x Q we shall need appropriate
definitions of small sets, the counterparts of the polar and semipolar sets
of classical and parabolic potential theory. The counterpart of a polar set
that we adopt in the present context is an evanescent set, and we have
already (Section 1.8) defined "quasi everywhere on R' x 0" accordingly.
The following remark suggests the definition to be given of a semipolar
subset of R' x 0.
It is trivial that if T is optional and if [T] is evanescent, then P{ T < + oc }
= 0. Conversely, consider the class of subsets 4 of R' x S2 with the property
428 2.11. Optional Times and Associated Concepts

that PI T < + cc } = 0 whenever T is optional and QT] c A. If A satisfies


this restriction and is sufficiently smooth, for example (by Theorem 9),
if A is predictable, then A is evanescent. The negation of this property of
A leads to the following definition.
A subset A of f8+ x S2 will be called semipolar if there is a sequence T
of optional times such that A c Ua QT up to an evanescent set. Subsets of
semipolar sets are semipolar, and countable unions of semipolar sets are
semipolar. If A is semipolar, then the set of values of i with (t,w) in A is
countable for almost every co. Conversely, if A is sufficiently smooth, for
example if A is predictable, it has been shown that A is semipolar when this
countability condition is satisfied, but we shall not need this result.

11. The Classes D and LP of Stochastic Processes

In this section the concepts are relative to a specified filtered probability


space (S2, f , P. .F (t), t e 1) ; 1 is an arbitrary linearly ordered parameter set.
The processes involved are defined on this space, adapted, and have state
space lf8

The Class D

This class of processes is the class of processes for which the family

x(T) : T optional, countably valued with values in 1; (11.1)

of random variables is uniformly integrable. A class D process is uniformly


integrable but a uniformly integrable process need not be in the class D.
The family (11.1) is uniformly integrable if and only if there is a uniform
integrability test function b and a constant c such that

E{D[jx(T)j]} 5 c (11.2)

for all Tin (11.1). Observe that if a point +oo and a a algebra .9f(+co)
are introduced as in Section 1 and if x(+ co) is an arbitrary F(+ oo) measur-
able and integrable random variable, then if the original process was in D
relative to the original filtration the augmented process is in D relative to the
augmented filtration. In fact the family (11.1) augmented by the random
variable x(+ oo) is uniformly integrable so there is a uniform integrability
test function 4) and a constant c such that (11.2) is true for T as in (11.1)
and also so that E{1)[x(+oo)I]} 5 c. The augmented process, in which T
is allowed to have the value + co, then satisfies (11.2) with c replaced by 2c.
12. Accessible and Totally Inaccessible Optional Times 429

In particular, if 1= p+, if is an almost surely right continuous class


D process, and if .F(O) contains the null sets, then the family (11.1) is uni-
formly integrable even if Tis allowed to be an arbitrary finite-valued optional
time S. To see this observe that if (11.2) is true for T as in (l 1.1), then (11.2)
is true for [S] [see Section 2, Example (b2), for the definition of this dis-
cretized optional time]. When n co, Fatou's lemma implies that (11.2)
is true for T = S. Finally, if x(+ oo) is an .F(+ co) measurable and integrable
random variable, we find as above that the augmented process is in D if the
original process was, and if so, the augmented family (11.1) with T now an
arbitrary optional time is uniformly integrable.

The Class LP (p z 1)

This class of processes is the class of processes for which

sup {E{ Jx(T)IP} : T optional, countably valued with values in I} < + co


(11.3)

Observe that LP c D when p > 1. Adjoin a pair {x(+ co), .F(+ co) } to each
process as in the discussion of the class D, but under the additional
condition of finiteness of E{Ix(+co)IP}. Follow that discussion to show
that (11.3) is true for the augmented process if true for the original process
and that then in the continuous parameter context (11.3) is true with Tan
arbitrary (< + co) optional time.

12. Decomposition of Optional Times; Accessible and Totally


Inaccessible Optional Times
In this section the reference space (0, S=, P; .F(t), t e R+) is a probability
space provided with a right continuous filtration for which .9=(0) contains
the null sets.

Totally Inaccessible Optional Times

(The reader is warned that the now accepted definition of these optional
times differs slightly from the original version.) An optional time T is said
to be totally inaccessible if P{S = T < + co } = 0 whenever S is a predictable
optional time. In particular, the distribution function of T on R+ is con-
tinuous; that is, P{T = c} = 0 for every finite constant c. Observe that this
definition only involves the properties of Ton (T < + oo 1; total inaccessi-
bility is a property of T where T is finite valued. A striking token of this
fact is that an optional time T is both predictable and totally inaccessible
if and only if T = + co almost surely. Observe also that predictability of
430 2.11. Optional Times and Associated Concepts

an optional time is defined relative to a measurable space provided with a


filtration, but total inaccessibility involves a measure space.

A Characterization of Total Inaccessibility

An optional time T is totally inaccessible if and only if T > 0 almost surely


and if whenever S. is an increasing sequence of optional times with limit S,
the increasing sequence IS. z T, T < + oo} of sets has union IS z T, T <
+ x }. In fact for an arbitrary optional time the latter set includes the union,
and the difference between the two is the ,(S) set

A={Sn<S=T<+cc,foralln}.
If S;; = S. on the F (Sn) set {Sn < S < + oo } and S = + oo elsewhere on
S2, then the sequence n i-+ S. A n is an increasing sequence of optional times
whose limit, announced by this sequence, is predictable and equal to T on
A. Hence A is null if T is totally inaccessible. Conversely, if A is null for all
sequences S,, it follows that P{S = T < + oo } = 0 whenever S is a predict-
able optional time not almost surely 0; so T (by hypothesis almost surely
strictly positive) is totally inaccessible.

Decomposition of an Optional Time

Let T be an optional time, and choose a sequence S. of predictable optional


times maximizing P{ Uo { T = S. < + co } } for all such sequences. Let a
be the maximum probability.
Special Case : a = 0. In this case and only in this case T is totally in-
accessible.
Special Case: x = 1. In this case T is called accessible and necessarily

QT]=UIISnD
n

up to an evanescent set. Here the union can be replaced by a union


Uo QS, 'J of disjoint graphs of predictable optional times by the usual
procedure: define
n-1

A
on S2-An.
12 Accessible and Totally Inaccessible Optional Times 431

Thus to an accessible optional time Tcorresponds a sequence S; of predict-


able optional times and a sequence A. of disjoint subsets of 0 such that for
all n

on A., onO-A,,, and PlUAm}=1.


0

Special Case: 0 < a < 1. In this case define the accessible optional time
T' and the totally inaccessible optional time T" by

T(w) if T(w) = oo,


+ oo otherwise;
T(w) if -(w) _ +oo,
T" =
+ oo otherwise.

These optional times are uniquely determined up to null sets, and there is
a subset A of Q such that T' = Ton A and T" = Ton 0 - A. Thus QT] =
QT'D u QT"] up to an evanescent set. The optional times ' and T" are
called, respectively, the accessible and totally inaccessible components of T.

EXAMPLE (Predictable process defined by an accessible optional time). It


was pointed out in Section 7 that if T is a predictable optional time, the right
continuous process = AT(.)
is predictable. In view of the de-
composition of an accessible optional time noted above, A'(-) is nearly
predictable if T is accessible. In fact, if up to an evanescent set QT] =
Uo [T ] (disjoint union) with T. predictable, then AT is indistinguishable
from Eo AT-().
Chapter III

Elements of Martingale Theory

1. Definitions
Let (f2, .9, P; ,fi(t), t e I) be a filtered probability space, and let 90)
be a process on this space, with state space (R,M(R)). The process is called a
supermartingale if the process random variables are integrable and if the
supermartingale inequality

x(s) z E{x(t)JF(s)} a.s. ifs < t (1.1)

is satisfied. The exceptional null set may depend on s and t. If I is a set of


consecutive integers, inequality (1.1) for t = s + 1 implies (1.1) for all pairs
s, t with s < t. If the inequality is reversed the process iscalled a submartingale,
and if there is equality in (1.1) the process is called a martin gale. The martin-
gale definition is sometimes also applied to complex-valued or vector-
valued random variables, but in this book the state space will always be
(R. M(R)) unless some other state space is specified. Martingale theory
refers to the mathematics of supermartingales and submartingales as well
as martingales.
A supermartingale is a mathematical model for the fortune of a player
of a game in which the player has fortune x(s) at time s and given the past
up to and including s the player expects his fortune at the later time t to be
at most x(s). Thus a supermartingale represents an unfavorable game; in
this context a martingale represents a fair game, and a submartingale
represents a favorable game.
The negative of a supermartingale is a submartingale, and an adapted
process is a martingale if and only if it is both a supermartingale and a
submartingale. In any one of the three cases the process is said to be left
[right] closed if I has a first [last] element and is said to be left [right]
closable if a first [last] element can be adjoined to I, together with an ap-
propriate a algebra, to get an enlarged process which is left [right] closed
and of the same type as the given one. If a process is right closed, say with
last parameter element o c, the a algebra floc) is not involved in the
appropriate version of (I.1) and is restricted only by the condition that
the process is adapted. A martingale is always left closable by the pair
2. Examples 433

[c, (0, Q)], where c is the common expected value of the process random
variables.

EXAMPLE. Let I= Z in increasing order, let .F(n) consist, for each n, of


the empty set and the whole probability space, and consider the process
all of whose sample functions are identically 1. This process is a martingale
and becomes a right closed martingale if the constant function 1 is adjoined
at the end or a right closed supermartingale if the constant function 0 is
adjoined at the end.

Choice of Filtration

Let 9(-)) be a supermartingale. The following remarks are made for


the supermartingale case but are valid with the obvious changes in the
other two cases. Define .mo(t) = .F{x(s), s5 t). Then .mo(t) a 9(t) and
go(-)) is a supermartingale because ifs < t,

x(s) = E{x(s)IAo(s)} z E{E{x(t)I F(s)}IAo(s)} = E{x(t)l41a(s)} a.s.


(1.2)

Thus is the minimal filtration for which is a super-


martingale. The larger the a algebras of the filtration, the more one knows
about the process, more precisely, the more (1.1) implies. It is sometimes
convenient to suppose that each a algebra .fi(t) contains all the null sets.
If this inclusion is not already true, S=(t) can be replaced by F1 (1), the
smallest a algebra containing .fi(t) and the null sets, to obtain a filtration
for which is a supermartingale and contains the
null sets.

Omission of the Filtration in Martingale Theory Notation

If a process is described as a supermartingale, martingale, or sub-


martingale without specification of the reference filtration, it is to be under-
stood that the reference filtration is or.,F1 as defined in the preceding
paragraph.

2. Examples
EXAMPLE (a). If is a filtration (arbitrary linearly ordered parameter
set I) of a probability space, if x is an integrable random variable, and if
x(t) = E{xJ.flt)}, then the process is a martingale. Every right
closable martingale is of this type. Suppose in this example that T is a
countably valued optional time with values in 1. Then we shall show that
434 2 111 Elements of Martingale Theory

x(T) = E{xl.F(T)} a.s., (2.1)

and if S is a second such optional time, we shall show that

E{x(S)j.f(T)I = E{.x(T)j.F(S)} =x(S n T) as. (2.2)

In the most common application of these results the parameter set I


is 0. I ..... + ro in the indicated order, and then optional times are neces-
sarily countably valued. A continuous parameter version of (2.1) and (2.2)
will be derived in Section IV.2.
Since it is sufficient to prove (2.1) for x v 0 and (- x) v 0, we can assume
in proving (2.1) that x is positive. If A e.F(T) and if A, = A n ; T = r;,
then A,E.F(r); so x(T) is.F(T) measurable, and

fxdP IA,x(r)dP=n,x(T)dP. (2.3)


=
Sum in (2.3) over the countably many values of r taken on by T to find that
x(T) is integrable. The equality (2.3) for all r is an integrated version of
(2.1). In the proof of (2.2) we shall use freely the properties in Section 11.2
of optional times and the a algebras they determine. The translation of
these properties into the present context of countably valued optional
times is trivial. If S >_ T [S :s: T], then F(T) c .F(S) [F(S) c flT)];
so (2.2) reduces to the special case 1(4.2). We now reduce the general case
to this special one. Observe that the set {S:!9 T} and its complement are
in .F(S A T) and that x(S A T) is F(S A T) (c .F(T)) measurable; so

E{.v(S)I.y(T)j = E{ l,st.T(x(S A T)j.F(T), + E{ 1(s,T(x(S v T)I,F(T)1


= I(s<T,x(S A T) + lts>T(E{E{xj.F(S v T)}I.f(T)
= I (s<TIX(S A T) + 1,s>7-,x(T) = x(S A T) a.s.

Interchange S and T to complete the proof of (2.2).

EXAMPLE (b). Let v(l ), v(2), ... be mutually independent integrable random
variables, and let Y(n) = E, f(j), .f(n) = . { < nJ. Then O}
is a submartingale if E;v( j); >- 0 for j > I and is a martingale if these
expectations vanish.

EXAMPLE (c) [Continuous parameter version of Example (b)]. Let


t e R+ } be a stochastic process with state space ll and independent incre-
ments; that is, 0 < t, < . < tk implies that the increments _x(t2) - x(tl),
.... x(tj - x(tw _, )are mutually independent. Define fi(t) = ,F { x(s) - x(0),
S< t,. Then if every increment x(t) - x(s) with t > s is integrable and has a
positive expectation, the process x(0), r is a submartingale.
3. Elementary Properties (Arbitrary Simply Ordered Parameter Set) 435

If these expectations all vanish, the process is a martingale. We shall see


that Brownian motion is a special case of this example. To clarify the ideas
involved here and in Example (b), we prove the submartingale assertion.
Using the fact that x(t) - x(s) is independent of .ms(s) when t > s, we find

E{x(t) - x(0)1.fls)} = E{x(s) - x(0)1.fls)} + E{x(t) - x(s)j.(s)}


= x(s) - x(0) + E{x(t) - x(s)} Z x(s) - x(0) a.s.,

as was to be proved.

EXAMPLE (d). Let 0 be the interval [0, 1], let .F be the class of Borel subsets
of Q, and let P be Lebesgue measure on F. Define a stochastic process
by

52" on [0, 2-"],


fln) = ,F {x(1), ... , x(n) }, n > 0.
10 on ]2-", l],

The process {x(-), ,F(-)) is a martingale and

E{x(n)} = 1, lim x(n) = 0 a.s.


R-W

This process cannot be uniformly integrable because the sequence is


not L' convergent to 0, although it is almost surely convergent to 0. The
process is a right closable supermartingale, right closable by the constant
random variable 0, for example. The process is not a right closable sub-
martingale and hence not a right closable martingale because if x right
closes this submartingale, then E{xl,F(n)} z x(n) almost surely, and this
inequality implies uniform integrability of because [Section 1.4(i)]
the family of conditional expectations of an integrable random variable is
uniformly integrable.

3. Elementary Properties (Arbitrary Simply Ordered


Parameter Set)
Proofs are omitted if trivial.
(a) If is a supermartingale, the function is
decreasing, and if s 5 t, 5 t2,

E{x(t2)J.F(s)} 5 E{x(t,)J.fls)} a.s. (3.1)

The proof of the second assertion is simply a manipulation of conditional


expectations :
436 2.111. Elements of Martingale Theory

E(x(t2)I,F(s)} = E{E{x(12)IF(t0}IF(s)} s E{x(tl)I.9T(s)) as.


(3.2)

(b) If for each n e l the process is a positive super-


martingale and ifs < t, then

lim inf x (s) Z linm inf E{x (t) I.fls)) Z E {liminfxn(t) IF(s) } a.s.,
(3.3)

so that the process {lim inf,... is a supermartingale if its random


variables are integrable.
(c) If is a submartingale and if f is an increasing convex
function with E{ f [x(t)] } < + oo for all t, the process {f
is a submartingale. If the first process is a martingale, f need not be increasing
for the conclusion to hold. The analogs of these properties in classical
potential theory are in Section 1.11.9. To prove the first one, for example,
apply Jensen's inequality for conditional expectations to deduce, when
s<t,
f [x(s)] < f [E{x(t)l.F(s)}] s E{ f [x(t)]I.F(s)) a.e., (3.4)

and this is the desired submartingale inequality.


(d) If {x(t),.F(t); t e II is a supermartingale, define c = sup,.I E{x(t)}.
If the supermartingale is left closed with smallest parameter value a, then
E{x(a)} = c; so left closability of the supermartingale implies finiteness
of c. Conversely, if c is finite, the pair [c, (0, Q)] left closes the super-
martingale.
(e) If is a right closed positive submartingale with largest
parameter value fl, then this process is uniformly integrable [Section I.4(i)]
because each random variable of the process is majorized by a conditional
expectation of x(IJ). Positivity is a necessary hypothesis here. In particular,
a right closed martingale is uniformly integrable because
F(-)) is a positive right closed submartingale. In the other direction, if
{x(t), .F(t); t e I } is a uniformly integrable submartingale, it is right closable
according to the following argument. If A e .ms(s), the function t'--. E{x(t); Al
is an increasing function for t z s. Define O(A) as the supremum (limit as
t increases) of this function. The set function 0 as so defined on the set
algebra UE,.F(s) is additive, and in view of the uniform integrability
hypothesis
lim 4(A) = 0. (3.5)
PIN-0

Hence 0 is countably additive and thus has a countably additive extension


to the a algebra Y,.f(s), and this extension is absolutely continuous
5. Convergence of Supermartingale Families 437

relative to ¢. If x = d¢JdP, the pair [x, l(E f.fls)] right closes the given
submartingale. A trivial variation of this argument shows that a uniformly
integrable martingale is right closable; the converse martingale result was
noted above.
(f) A supermartingale which majorizes a uniformly integrable sub-
martingale is right closable. In fact, if is a supermartingale
which majorizes the uniformly integrable submartingale we
use the notation of (e) and show that the pair just obtained which closes
also closes For Ae.f(s) and t s,
E{ y(s); A} z E{y(t); A} ;2t E{x(t); A} -+ 4,(A) (t T),

and the inequality here between first and last terms is the integrated form
of the desired almost sure supermartingale inequality y(s) Z E{xl.F(s)}.

4. The Parameter Set in Martingale Theory


If {x(t), .F(t), t e I } is a submartingale and b is a constant, the process
b] v 0 = is a submartingale [Section 3(c)] ; so s < t implies
that E{yb(t)} z E{yb(s)}. If there is equality here for some triple s, t, b,
then the pair [yb(s),,fls)], [yb(t),F(t)] is a martingale; so

f [x(t) - b] v OdP = f [x(s) - b] v OdP = 0, (4.1)


,J{Kr)5bl ,Jid,)56}

and therefore x(t) 5 b almost everywhere where x(s) 5 b. If b, is a sequence


dense in l and if E{ybJ(t)} = E{ybj(s)} for all j, it follows that x(t) 5 x(s)
almost surely, and therefore there is almost sure equality because E{x(t)} z
E{x(s)}. Choose cJ > 0 so small that Eo cC(1 + JbjI) < +oo. The process
Eo is a submartingale with the property that is
an increasing function and that the equality E{ y(s) } = E { y(t) } , implies
that x(s) = x(t) almost surely. If the parameter set I is mapped into O2 by
¢: then ¢ is monotone increasing and 4,(s) = 4,(t) if and only
if x(s) = x(t) almost surely. Thus it is not an essential restriction to suppose
that the ordered parameter set in martingale theory is a subset of 68 ordered
by inequality.

5. Convergence of Supermartingale Families


If r is a family of supermartingales adapted to then ess inf IF (defined
in Section 1.8) is also a supermartingale if its random variables have finite
expectations because if s and t are parameter values with s < t,

E{(essinfr)(t)I.F(s)} 5 x(s) a.s. (5.1)


438 2.111. Elements of Martingale Theory

for each process x(-) in r, and therefore the right side can be replaced by
(ess inf r)(s) to obtain the supermartingale inequality. In the other direction,
if r is a family of supermartingales which is directed upward, the process
ess sup r is a supermartingale if its random variables have finite expectations
in view of the monotone convergence theorem for conditional expectations.
If the processes in this upward-directed set r are martingales, this reasoning
shows that the process ess sup r is a martingale.
If r is an upward-directed family of supermartingales (essential order
of processes) with essential limit a supermartingale, there is an increasing
sequence {x n e Z+ } in r with the property that ess sup r = supne Z+
up to a standard modification. To see this, it can be supposed (Section 4)
that the parameter set of the process x'(-) = ess sup r is a subset of 11 ordered
by inequality. The monotone decreasing function is the su-
premum of the upward-directed family of decreasing functions tr+E{x(t)}
for in r. Hence there is a sequence {x n e Z+ }, increasing in the
essential order, for which

lim E{x (t)} = E{x'(t)}


-m

for every parameter value t, and it follows that, for every parameter value
t, x.(1) = x'(t) almost surely, as was to be proved. If the hypothesis
that the essential limit of r is a supermartingale, equivalently, that the
random variables of this process are integrable, is not satisfied, the conclusion
still holds, as can be seen by applying the result just obtained to the family
{x(-) n k: for each positive integer k.
Similarly, but somewhat more easily, if r is a family of supermartingales
and if ess inf r is a supermartingale, that is, if the random variables of this
process are integrable, there is a countable subset r, of r for which ess inf r
= inf r, up to a standard modification. See the Fundamental Convergence
Theorem for supermartingales (Theorem IV.5) for a more complete study
of the essential infimum of a family of right continuous supermartingales.

6. Optional Sampling Theorem (Bounded Optional Times)


The idea that if a game is unfavorable to a player, it looks unfavorable to
him at random times chosen without foreknowledge suggests the following
integral parameter theorem.

Theorem. Let {x(n),5(n),nE7L+} be a supermartingale. If T is a bounded


optional time, then x(T) is integrable. If S and T are bounded optional times
with S:5 T, then

x(S) z E{x(T)I.5(S)} a.s., (6.1)


6. Optional sampling Thcurcm (Buunded Optional rimes) 439

and there is almost sure equality if the process is a martincff(fl

If T is bounded by k, then
k
Eflx(T)I} S YE{jx(j)I} < 4-co.
0

If S< T, if Ae.F(S), and if A1= A n {S = j}, then A/e.F(j), if A31 =


j. Hence the supermartingale in-
AJ n {T > i}, then A 1 c- 9(i) when i
equality yields
k T-1
[x(T) - x(S)]dP = Y Y [x(i + 1) - .x(i)] dP
Jn 1=0 Al i=1
(6.2)
k

zi
i
I

n.J,
[x(i + 1) - x(i)] dP < 0,

and there is equality if the given process is a martingale. Thus the integrated
version of the desired supermartingale inequality or martingale equality
is true.
Observation. According to this theorem if T. is an increasing sequence
of bounded optional times the process {x(T,), .I(T.) } is a supermartingale,
or a martingale if the original process is a martingale.

Application to Right Closable Submartingales and the Class D

Let {x(t), .r (t), t e 1) be a positive right closable, equivalently, uniformly


integrable according to Section 3(e), submartingale, right closable by a
positive random variable x. In view of Theorem 6 as applied to -,x,
x(7)< E{xj.F(T)} a.s. (6.6)

if T is optional, taking on finitely many values, and therefore (6.6) is true


if T is optional and countably valued. Hence is a class D process be-
cause [Section 1.4(i)] the family of conditional expectations of x is uniformly
integrable. Thus there is a uniform integrability test function 'D and a
constant c such that

E{b[x(T)]} < c (6.7)

whenever T is a countably valued optional time. In particular, if I = OB+,


if is almost surely right continuous, and if S is an arbitrary optional
time, (6.7) can be applied to [S] as defined in Section II, Example (b2),
and (n -+ co) (6.7) is therefore true for T = S. Thus this submartingale is
440 2.111. Elements of Martingale Theory

in the class D in the strong sense that the family {x(T): T optional} is
uniformly integrable. In particular, a right closable, equivalently, uniformly
integrable, martingale { y(t), .fi(t), t e I } is a class D process [set x(t) _
I '(')I]; if I = 68+ and if the martingale is almost surely right continuous,
the family {y(T): T optional} is uniformly integrable.

7. Optional Sampling Theorem for Right Closed Processes


If the process in Theorem 6 is right closed, the optional times need not be
bounded according to the following theorem.

Theorem. If {x(n), ,fln), n e Z` } } is a supermartingale and if T (< + x)


is optional, then x(T) is integrable. If S (< + ac) is also optional with S:5 T,
then (6.1) is true, and there is equality if the process is a martingale.

If the process is a martingale, x(T) is integrable and x(T) =


E{x(+x)J,F(T)} according to Section 2; so (6.1) with equality reduces to

x(S) = E{E{x(+x)I,F(T)}I.F(S)} a.s. (7.1)

which is true because : (S) e .F(T). If the process is a supermartingale,

x(n) = [x(n) - E{x(+oc)j,F(n)}] + E{x(+x)I.F(n),' a.s. (7.2)

so that is the sum of a positive supermartingale with last element 0 and a


martingale, both relative to Thus it is sufficient to prove the theorem
for a positive supermartingale with last element 0. Under this hypothesis
define T. = T A n, S. = S A n, and observe that according to Theorem 6
E{x(0)} _> E{x(T)}. that E{x(0)} >_ E{x(T)}. Thus
x(T) is integrable. Again according to Theorem 6, if m < n,

X(Sm) Z E{x(T;)IF(Sm)} a.s., (7.3)

and when n - x this inequality becomes

x(SS) > E{X(T)I.F(Sm)} a.s. (7.4)

by Fatou's lemma for conditional expectations. Now A e F(S) implies that


Am = A r {S < m} e,F(Sm); so (7.4) implies that

Jx(S)dP jx(T)dP. (7.5)


f" m
This inequality when m oo yields the integrated version of (6.1).
7. Optional Sampling Theorem for Right Closed Processes 441

EXAMPLE. If (x(n),.F(n),ne7L+} is a positive supermartingale, the theorem


is applicable because the process can be closed on the right by defining
x(+oo) = 0. This is the most used example. More generally the theorem
is applicable according to Section 3(f) to a supermartingale {x(n), .f(n),
n e 7L+ } which majorizes a uniformly integrable submartingale.

Generalization. Let {x(t),.f(t), t e 1 } be a supermartingale with an arbi-


trary linearly ordered parameter set I for which there is a first and last
element. If T is a countably valued optional time (with values in 1), then we
shall now prove that x(T) is integrable. Moreover, if S is also a countably
valued optional time with S < T, then we shall prove that (6.1) is true and
that there is almost sure equality if the process is a martingale. The proof
follows that of Theorem 7. Denote by 0 the first element of I and by oo
the last element of /. Then just as in the proof of Theorem 7 the martingale
case is covered by Section 2, and only the case of a positive supermartingale
with last random variable 0 needs further examination. Let r, be the sequence
of parameter values taken on by S and T, define
n
on U {T = r;},
i=o
T. _
oc on U {T = rj},
j=n+I

and define S similarly in terms of S. Then Sn and T, are optional, take


on only finitely many values, and Sn < Tn. Theorem 6 is applicable to S.
and T n because only the parameter values r 0 . . . . . rn, oo are involved. An

application of Theorem 6 to the pair (0, Tn) of optional times shows that
E{x(T,) } < E{ x(0) 1, so when n - + co, we find that x(T) is integrable. An
application of Theorem 6 to the pair (Sn, Tn) shows that

x(T,) dP = fs" x(T,)dP < f x(S,) dP f x(S) dP


IS=ri ) =rj) JISn=rjl S=rll

so (n-+x)
f x(T) dP < x(S) dP,
S=rfl

and this inequality is an integrated version of (6.1).


442 2.111. Elements of Martingale Theory

8. Optional Stopping
Let { 4i (t), t e 1 } be a filtration of a probability space, with I either Z' or 68',
and let T be an optional time. If (x(-), 9(-)) is a stochastic process on the
probability space, the process {x(T A t), t E I } is described as the process
stopped at T.
Suppose that I = Z' and that is a supermartingale [martin-
gale]. According to Theorem 6, the stopped process {x(T A n),F(T A n),
n e l' ) is also a supermartingale [martingale]. It is sometimes important
that even the process {x(T A n), .fi(n), n e Z' } is a supermartingale [mar-
tingale]. To prove this result, suppose that A e F(n). Then

E{x(T A n); An {T 5 n}} = E{x(T A (n + 1)); An {T5 n}} (8.1)

because the integrands are the same on the integration domain, and

E{x(T A n); A n {T > n)) z E{x(T A (n + 1)); A n {T > n} }, (8.2)

with equality in the martingale case, because the integrands are x(n) and
x(n + 1), respectively, on the integration domain; so the supermartingale
inequality [martingale equality] is applicable. Adding (8.1) and (8.2) we
obtain an integrated version of the desired supermartingale inequality, or of
the martingale equality in the martingale case.

9. Maximal Inequalities
Theorem. (a) If x(0), ... , x(n) is a submartingale and a is an arbitrary real
number,

aP{maxx(j) z a} S E{x(n); maxx(j) >- a} < E{x(n) v 0), (9.1)


JSn j5n

aP{minx(j) 5 a) z E{x(0)} - E{x(n); minx(j) > a)


J5n j5n

E{x(0)} - E{x(n) v 0}.

(b) If x(0), ... , is a positive supermartingale and a > 0,

aP{maxx(j) >- a) 5 E{x(0)). (9.2)


Js,,

To prove the first inequality in (9.1), define T = n A min j j: x(j) -> a)


Then T is optional for the submartingale; so (Theorem 6) the ordered pair
[x(T),x(n)] is a submartingale. The submartingale inequality E{x(T)) 5
E{x(n) } yields
9. Maximal Inequalities 443

E{x(n)} > aP{maxx(j) >- a} + E{x(n); maxx(j) <a}, (9.3)


jsn jSn

which implies the first inequality in (9.1). To prove the second inequality,
define S = n n min { j: x(j) < a}. Then S is optional; so the ordered pair
[x(0), x(S)] is a submartingale, and the submartingale inequality E{x(0) } <
E{x(S)} yields

E{x(0)} < aP{minx(j) <a} + E{x(n); minx(j) > a}, (9.4)


j sn j sn

which implies the second line in (9.1). To prove (9.2), apply (9.4) to
Observe that in (9.1) and (9.2) if ">-a" and "Sa" are replaced by the
corresponding strict inequalities, the resulting versions of (9.1) and (9.2) are
apparently weaker than the old versions but actually imply them by a trivial
continuity argument.
If the parameter set of the submartingale or positive supermartingale in
question is any linearly ordered countable set A, the inequalities (9.1) or (9.2)
as the case may be are valid for every finite set x(to), ... , x(tn), where t, is a
finite ordered subset of A. It follows that

aP{supx(t) a} < supE{x(t) v 0}


lEA tEA
a submartingale]
aP{inf x(t) < a} >_ inf E{x(t) } - sup E{x(t) v 01
tEA tEA tEA
(9.1')

aP{supx(t) > a} < sup E{x(t)} a positive supermartingale].


tEA t,A
(9.2')

If the process is left closed, by x(a),

inf E{x(t) } = E{x(a) } a submartingale]


IEA

sup E{x(t) } = E{x(a) } a supermartingale].


tEA

If the process is right closed, by x(b),

sup E{x(t) v 01 = E{x(b) v 0} a submartingale].

Moreover in this right closed case the derivation of the first inequality in
(9.1') yields the in general smaller right-hand side E{x(b); sup,EAx(t) Z a}.
If the parameter set of the submartingale or supermartingale is uncount-
able, (9.1') and (9.2') remain true if "sup" and "inf" on the left are replaced
444 2.111. Elements of Martingale Theory

by "ess sup" and "essinf" respectively, because these essential bounds are
equal almost everywhere to ordinary bounds over suitably chosen countable
parameter subsets. On the other hand, if the process parameter set A is an
interval and if the process is almost surely right continuous, then .(9.1') and
(9.2') are correct as written since replacing A in these inequalities by a count-
able dense subset including the right-hand endpoint of A if A is right closed
changes neither the left nor the right side.

10. Conditional Maximal Inequalities


The inequalities in Section 9 can be made more precise by conversion to
conditional inequalities. For example, let {x(j), .'(j), 0 5 j 5 n} be a sub-
martingale, and suppose that A e 5(0). An application of the first inequality
in (9.1) to the submartingale { 1,x(j), 9(j), 05 j5 n j leads to the integrated
version of

aP{maxx(j) z al.F(0)} < E{x(n)zI. (0)} S E{x(n) v 05(0)} a.s.,


jSn
(10.1)

where z is the indicator function of the set {max;s x(j) >_ a}. The other
inequalities in Section 9 can be extended similarly. Such extensions to con-
ditional inequalities will be omitted from now on.

11. An L" Inequality for Submartingale Suprema


Theorem. If {x(t), t e 1 } is a positive right closed submartingale, closed by the
random variable x, if A is a countable subset of I, if p > 1, and if 1 /p + 1 /q = 1,
then

E{supx(t)P} < qPE{xP}.


tEA

According to Section 9,

aP{supx(t) >_ a} 5 E{x; supx(t) >_ a), (11.2)


teA teA

and it is therefore sufficient to drop the martingale theory context and to


show that if x and y are positive random variables satisfying

aP{y>_a}5E{x;y>_a}, (11.3)

then
12. Crossings 445

E{ yP} < qPE{xP}. (11.4)

Now
y(w) a

E{ yP} = p f dP f ap-' da = p P{ y z a}aP-' da


n o o

y'Iw)
5p Lip-2da xdP=p f xdP f aP-2da (11.5)
I Jn Jo
= qE{xyP-'} 5 gE'IP{xP}E'w{ yP}

If E{ yP} < + oo, this inequality yields (11.4). Otherwise, replace y by y A k,


for which (11.3) is still true, thereby deduce (11.4) for y n k, and let k - + oo.
A common application is to processes with parameter set 0, 1, ... , +oo
ordered as indicated. In this case one chooses A = 1. If I is uncountable,
(11.1) is true with A replaced by 1 and "sup" replaced by "ess sup" because
there is a countable subset to of I with the property that ess sup, E, x(t) _
sup,E,ox(t) almost everywhere. If I is a right closed interval of R and if the
given process is supposed almost surely right continuous, then (11.1) is true
with A = I because if A is a dense subset of I including the right-hand
endpoint, then sup,,, x(t) = sup, E A x(t) almost everywhere.

Modification of Theorem 11 if the Submartingale Is Not Right Closed

Observe that in Theorem 11 the process .F(-)) is a submartingale


if x(t)P is integrable for all t; so the function t)-+E{x(t)"} is monotone in-
creasing. If sup,E, E{x(t)P} = lima E{x(t)P} < + oo, the process x() is
uniformly integrable because the pth power is a uniform integrability test
function when p > 1. Hence [Section 3(e)] the submartingale {x(-), .F(-)) is
right closable. The closability argument in Section 3(e) can be interpreted to
imply that x(t) has a weak limit x in L'(Q, Y,E, P) when t increases and
that x closes the submartingale We shall see in Theorem 15 that if
I= Z+, it is even true that x(n) = x exists almost surely and in L.
Then by Fatou's lemma E{xP} < +oo; so by (11.1) the submartingale
is also uniformly integrable and therefore is right closed by
xP; moreover (dominated convergence theorem) x(n) = x in the LP
sense. We leave to the reader the easy extension of these results to other
parameter sets.

12. Crossings
Let f, g, h be functions from a linearly ordered set I into R, with g 5 h. The
number Dn [ f; g, h] of downcrossings of [g, h] by f is defined as the su-
premum of the values of the positive integer k for which there exist t, <
446 2.111. Elements of Martingak Theory

< t2k in I satisfyingf(t) h(tt) when j is odd and f(it) < g(t) when j is even.
Upcrossings are defined in the obvious dual way. Observe that limt_,o j = f
implies that

Dn[f;g,h] 5 inff 2`', h - 2-"'] (12.1)


.ea 'gym

when g and h are finite valued.

Theorem. Let n e 1L ', and let be adapted processes on a filtered


probability space (0, 9, P; .f (j), 0 < j < n). Suppose that z 0 and that
and are positive supermartingales relative to Then

E{[x"(n) - x'(-), x"(-)]) < E{x(0) A x"(0)}. (12.2)

In particular, if x(-) and are positive supermartingales relative to and


if a, b are numbers with 0 < a < b,

E{x(0) A [by(0]
b-a
and in fact with the convention that 00 = I

E{Y(n); k} <
(a) E{x(0) b[bY(0)]'r
(k z 1).
(12.4)

If is a (not necessarily positive) supermartingale and if


-oo<a<b<+z,
E{x(0) A b} - E{x(n) A b}
(12.5)
b-a
Observation. The inequalities in Section I.VI.3(o) limiting the oscillation
of a superharmonic function by means of iterated reductions correspond to
supermartingale downcrossing inequalities. In fact it will be shown in Section
23 that the exact counterpart of the reduction procedure in Section I.VI.3(o)
leads to (12.2) and (12.3). Observe also that if 0, the left side of (12.3)
is E{ y(n) Dn a, b] 1; the corresponding remark is of course ap-
plicable to (12.4). Finally observe that the right sides of (12.2-5) are ma-
jorized, respectively, by E{x"(0)}, bE{y(0)}/(b - a), (a/b)k-'E{y(0)},
bl(b - a), which do not involve
To prove the theorem, define S. and T by setting S0 = TO = 0 and

S, = min (j >- 0: x(j) >t x"(j)},


12. croIng
447

_ min { j > Tk_, : x(j) x"(j)} (k odd, 3),


(12.6)
Sk
- {min{j> T_1 :- r(j)- < x'(j) } (keven,Z2) ,

Tk= SkAn.

As usual the minimum of the empty set of numbers is defined as + oo. The
sequence T. is an increasing sequence of optional times for Jr(-).

Proof of (12.2). In this proof we abbreviate Dn[x(-); x'(-), x"(-)] to Dn. In


the first place if j is odd,

E{x(Tj) - x(Tj+,)} = E{x(Tj) - x(Tj+,); Tj < n} (12.7)

> E{x"(Tj+1) - x(T,+1); Tj < n}

E{x" T j+ 1_ E J j+1}
2 2

- E{x(n);Tj <n,Dn <7 2 111

E{x"(n)-x'(n);Dn>7 21}

- E{x(n); Dn =7 2 1 }.

In the second place (supermartingale inequality) E{x(Tk) - x(Tk+,) } Z 0


for all k. Hence we can drop the summands with even j in the first sum below
to obtain

E{x(0)} _ E{x(T,) - x(Tj+,)} + E{x(n)}


j=o
n

E{x'(n) - x'(n); Dn z k} (12.8)


k=1

Y E{k[x"(n) - x'(n)];Dn = k} = E{[x"(n) - x'(n)]Dn}.


k=1

If is replaced in this inequality by A the number of down-


crossings is unchanged, and (12.8) yields (12.2).

Proof of (12.3) and (12.4). Inequality (12.3) is a special case of (12.2). Alter-
natively (12.3) can be obtained by summing (12.4). To prove (12.4), define
S, and T by (12.6) with by(-) and ay(-). Observe that for
k z 1,
448 2.111. Elements of Martingale Theory

E{y(S2k+2);S2k+2 5 n) 5 E{y(T2k+2); S2k+1 5 n}

5 E{y(S k+1 2k+1 5 n} <- E{x(S2k+l);bS2k+1 5 n}


(12.9)
E{x(T2k+1); S2k 5 n} E{(S2k); S2k 5 n}
b
5 b
a
5 E{ y(S2k); S2k 5 n}
b

and so fork z 1, under the convention 00 = 1, needed when a = 0,

a k- I
E{ y(SZk); S215 5 n} < ()E(Y(S2);S2
b
5 n}

a
5 ()k_lE{y(T2);St 5 n} 5 ()'E{Y(Si);si 5 n}

5 E{x(S1);S1 < n} (a)kE{x(0)}


(12.10)
b b
(a)5

Furthermore fork z 1.

E{ y(n); ay(-), by(-)] z k} = E{ y(n); S2k 5 n}

SE{y(S2k);S2k5n}5(b)
b
(12.11)

If is replaced by A [by(-)], the number of downcrossings is un-


changed and (12.11) becomes (12.4).

Proof of (12.5). To prove (12.5), define S. and T by (12.6) with x"(-) = b and
.x a. Since T. = n,

x(0) - x(n) _ [x(TT) - x(Tf+1)], (12.12)


1=0

and (supermartingale inequality) each bracket has a positive expectation. We


shall minorize each bracket with odd j. On the set Ak = a, b] = k}
each of the first k brackets in (12.12) with odd j is z b - a, and of the later
brackets with odd j only [x(T2k+1) - x(T2k+2)] can be nonzero, and if so,

x(T2k+l) z b, T2k+2 = n, x(n) > a.


12. Crossings

Thus on Ak,

C(T2k+t) - V(T2kl2) = x(T2.,,J - x(n) j h - x(n) (12.13)

and therefore

E{x(0) - x(n)} z (b - a, b]) + E{ [b - x(n)] A 0]. (12.14)

If is replaced by A b, the number of downcrossings is unchanged


and (12.14) yields (12.5).

Strict Downcrossings

If downcrossings had been defined using strict inequalities, f > h and f < g,
the resulting apparently weaker versions of the inequalities of Theorem 12
would imply the present versions by a trivial continuity argument.

Adaptation to Infinite Parameter Sets

If the parameter set of the theorem is replaced by a countably infinite linearly


ordered set 1, the inequalities obtained have obvious extensions. For example,
(12.5) becomes

sup E{x(t) A b} - inf E{x(t) A b}


b-Qef (12.15)

and (12.15) is also valid if the parameter set I is an interval of IB and the
process is almost surely right continuous. In fact under this hypothesis
the two sides of (12.15) are unchanged, for strict downcrossings and therefore
for downcrossings, if I is replaced by a countable dense subset including
each endpoint of I in I. In the context of (12.15) if the supermartingale
is left closed, by a random variable x(a), the supremum in (12.15) is
E{x(a) A b}. If this supermartingale is right closed, by a random variable
x(fl), the infimum in (12.15) is E{x(fl) A b}.

The Role of Downcrossing Inequalities

The importance of the downcrossing concept lies in the following three


facts. Let f be either (i) a sequence of numbers, or (ii) a function from an
interval I into R, or (iii) a function from a dense subset of an interval I
into R. Then the number of downcrossings Dn[f; a, b] by f of an interval
[a, b] is finite for every interval if and only if, respectively, (i) f is convergent
to a (possibly infinite) limit; (ii) f has right and left (possibly infinite) limits
450 2.111. Elements of Martingale Theory

at every point of 1 and at the endpoints of 1 not in I; (iii) f has an extension


to I which has the property stated in (ii). Moreover in all three contexts
it is sufficient if Dn [ f ;a, b] < + oo for all intervals [a, b] with endpoints
in a countable dense subset of R. The relative downcrossing function
of a pair of positive super-
martingales has not yet proved useful in probability theory, but its potential
theory counterpart was useful in Part 1 of this book. In fact we shall see
in Section 23 that Theorem 12 can be obtained by a reduction method, and
it will then be clear that Theorem 12 is the exact counterpart of a set of
inequalities [Sections I.VI.3(o) and 1.XVII.16(s)] for iterated reductions.
The latter inequalities were essential in proving Theorems 1.XI.4 and
I.XVIII.14 on the fine topology limit properties of the ratio of two positive
superharmonic, respectively, superparabolic functions.

13. Forward Convergence in the L' Bounded Case


The fundamental martingale theory convergence theorems are the following
forward one for the parameter set 0, 1, ... and backward one (Theorem 17)
f o r the parameter set ... , -1, 0.

Theorem. Let {x(n), .f(n), n e l' } be an L ` bounded supermartingale mar-


tingale or .submartingale. Then limn.., x(n) exists (finite) almost surely.

It is sufficient to consider the supermartingale case. In that case if r, < r2,


Theorem 12 implies that r, , r2] } < + oo so that almost no
sample sequence has infinitely many downcrossings over the interval [r, , r2].
Thus the summands on the right in the relation

{Iim sup x(n) > lim inf x(n)l


n+eo n_ao
(13.1)
cU =
{lir?Px(n)> r2 > r, > n-M
linfx(n)j
r rational

are null sets so limn, x(n) exists almost surely, and the limit is almost
surely finite because

El lim x(n)
r.O
(13.2)

The hypotheses of the theorem are satisfied if the process is a positive


supermartingale (the most natural case for potential theory) or more
generally if the process is a right closable supermartingale because if x(+ oo)
right closes the supermartingale
14. Convergence of a Uniformly Integrable Martingale 451

E{ lx(n)I } = E{x(n) } - 2E{x(n) A 0} < E{x(0) } - 2E{ x(+ ao) A 0}.


(13.3)

This inequality shows that is L' bounded if and only if 0 E{x(n) A 0)


> -oo. See Section 15 for further discussion of the convergence of a right
closable supermartingale.

Generalization to Arbitrary Parameter Sets

Let {x(t), t E 1 } be an L' bounded supermartingale martingale or sub-


martingale with an arbitrary linearly ordered parameter set I except that
I has no last element. If I is countable, the method of proof of Theorem 13
is applicable to yield the fact that lim,t x(t) exists and is integrable. If I
is not countable, ess lima x(t) exists and is integrable because we can assume
(Section 4) that I is a subset of R ordered by in which case I has a cofinal
sequence and (Appendix Theorem IV. 10) the stated result follows from the
convergence result for countable I.

Continuous Parameter Case

Suppose that the process is a supermartingale martingale or submar-


tingale with parameter set a subinterval I of R whose right-hand endpoint
a (< + oo) is not in 1. If the process is L' bounded, we have just seen that
exists almost surely when t tends to a along the rationals, and
this fact implies, if the process is known to be almost surely right continuous,
that this limit exists almost surely when t tends to a unrestrictedly. This is
the typical application of the L' bounded martingale theory convergence
theorem in the continuous parameter case, and it will be shown in Chapter
IV that the hypothesis of almost sure right continuity is not very restrictive.

14. Convergence of a Uniformly Integrable Martingale


Let (0, F, P, .f(n), n e Z) be a filtered probability space, and define
.f (+ oo) = Y0 .f(n). Recall from Section 3 that a martingale is uniformly
integrable if and only if it is right closable, so that the most general uniformly
integrable martingale relative to the above unextended filtration is a sequence
{x(n), F(n), n E Z+ } with x(n) = E{xl.f (n) }, where x closes the martingale.
In this context define x(+oo) = E{xl.fi(+oo)}. If p >_ 1, then (submar-
tingale inequality if the expectations are finite) E{Ix(0)I°} <E{lx(1)l°} <
5 E{ l xl p}, so that the martingale is L' bounded (as is implied by the
given uniform integrability without invoking martingale theory), and in
general if the last expectation is finite, the martingale is L° bounded. The
following martingale theorem, in which p > 1 and the preceding notation
452 2.111. Elements of Martingale Theory

is used, contains the forward conditional expectation continuity theorem


proved directly in Section I.S.

Theorem. If {x(n), (n), n E 7L+ } is a uniformly integrable martingale right


closable by x, then lim, x(n) = x(+ oc) almost surely, and also in the L°
metric whenever xc L° [equivalent) v, for p >I whenever E{Ix(n)I P}
< + oo ]. The process

[x(0),F(0)], . . . , [x(+co),.f(+00)], [x,.F] (14.1)

is a martingale.

Since is L' bounded, Theorem 13 is applicable, and under the given


uniform integrability hypothesis the almost everywhere convergence assured
by Theorem 13 implies L' convergence. Furthermore, if AE,F(n), then
(martingale equality)

E{x(m); A} = E{x(n); A} = E{x; A} if m Z n, (14.2)

and when m - cc. this equation yields, in view of the L' convergence of

E{ lim x(m); Al = E{x; A}. (14.3)


M_m

This equality holds for A in the algebra Uo ,f(n) and therefore holds in
the generated a algebra fl+ oo) ; so this limit random variable is x(+ oo).
It is trivial that the process (14.1) is a martingale. If p > I and if
E{Ix(n)IP} < +co, then (Fatou's lemma) E{lx(+oo)IP} < +oo. If
p> I and if E{Ix(+oo)IP} < +co, then (Theorem 11 applied to
lx(n)IP} < +oo. So Ix(n) - x(+oo)IP} < +co, and
therefore (dominated convergence) the sequence converges to x(+oo)
in the LP metric.

Extension to Other Parameter Sets

Suppose that is a uniformly integrable martingale, right closable by


x, for which the parameter set I is either (a) a countable dense subset of
an open interval ]0, a[ or (b) the interval ]0, a[. Theorem 14 (and its proof
with trivial modifications) remains valid in both these contexts when suitably
rephrased. More specifically, in (a) lim,tax(t) exists almost surely and in
L'; in (b) esslim,t.x(t) exists, and the limit also exists in V. Moreover
in (b) if the process is almost surely right continuous, the essential limit is
also an almost sure limit. In both cases the limit is almost surely
E{xtY,E,.F(t)}. The extension when xeLP is obvious.
15. Forward Convergence of a Right Closable Supermartingale 453

Application to Approximation Theory

Let {x(t), t E I } be an arbitrary infinite collection of random variables on


some probability space, and if J c ],define f(J) = f{x(t), t EJ}. Then
if x is an f(I) measurable and integrable random variable, we shall now
show that there is an increasing sequence J, of finite subsets of I such that

limE{xl.F(J,)}=E{x19lUJnl}=x a.s. and in L'. (14.4)

Since (Section I.4) E{xl.f(J,)} is equal almost surely to a Bore] measurable


function of the finitely many random variables with indices in in, equation
(14.4) exhibits a canonical way to approximate x by functions of finitely
many of the random variables. In fact let x be as stated, and (Appendix
IV.2) choose J countable and so large that x is f(J) measurable. If Jn in
(14.4) is chosen to be the set of first n members of J in some enumeration,
(14.4) reduces to a special case of the conditional expectation continuity
theorem, that is, a special case of Theorem 14. Observe that if x is an arbitrary
integrable random variable on the given probability space, (14.4) should
be replaced by

lim E{xj.f(Jn) } = E{xI.f I U Jn l} = E{xj.f(!)} a.s. and in L'. (14.5)


rm \o f)))

15. Forward Convergence of a Right Closable


Supermartingale
Theorem. If {x(n), .f(n), n E 7L+ } is a right closable supermartingale,
limn_,, x(n) = x(+ oo) exists (finite) almost surely, and if x right closes the
supermartingale the process, (14.1) is a supermartingale.

It was pointed out in Section 13 that Theorem 13 is applicable to right


closable supermartingales. Hence only the second assertion of the theorem
requires proof. Since the process is the sum of a positive supermartingale
and a right closable martingale according to

x(n) = [x(n) - E{xI.F(n)}] + E{xj.F(n)}, (15.1)

and since the corresponding right closure result for martingales was proved
in Section 14, it can be supposed from now on that the given supermartingale
is positive. Apply Fatou's lemma for conditional expectations to get

E{_r(+oo)I,F(n)} < liminfE{x(m)j.f(n)} < x(n) a.s., (15.2)


454 2.111. Elements of Martingale Theory

which shows that.r(+ x) coupled with F(+ oo) = Yo .f(n) right closes the
given supermartingale. Finally

E;x4.f(+x) limE{xj,f(n)} 5limx(n)=.r(+x) a.s., (15.3)

and therefore (14.1) is a supermartingale in the present context.


As the negative of Example (d) in Section 2 shows, an L' bounded
supermartingale need not be right closable.

Extension to Other Parameter Sets

The rephrasing of Theorem 15 for a process with a parameter set which is


either an interval ]0, a[ or a countable dense subset of this interval is similar
to the corresponding rephrasing in Sections 13 and 14.

16. Backward Convergence of a Martingale

The fact that the parameter set Z .... - I. 0 has a last element makes
:

backward martingale theory convergence theorems stronger than forward


ones It will be convenient to treat the martingale case first.

Theorem (Backward Half of the Conditional Expectation Continuity Theo-


rem Already Proved Directly in Section 1.5). Suppose that {x(n),_-fln),
nE/ ; isamartingale,that is,x(n) = E;x(0)1.f(n), a.s. Then limn-_, x(n) _
E; r(0)in" , as. and in L' norm.

The process is uniformly integrable because it consists of conditional


expectations of a random variable x(O) [Section 1.4(i)]. Just as in the proof
of Theorem 13 the downcrossing inequality, easier here because the process
is right closed, together with Fatou's lemma implies that there is an almost
sure limit which is integrable. To identify this limit, which we denote by
.V(- x ), choose A in F( - x) = n°_ ,, f(n). The martingale property and
the uniform integrability of the process yield

1x(_dP= lim JA v(n)dP


1^v(0)0. (16.1)

and this equality identifies x(- x) as the stated conditional expectation.


The convergence is L' convergence because the process is uniformly in-
tegrable.
The rephrasing of Theorem 16 for other parameter sets follows that
of the forward convergence theorems and is left to the reader.
18. The r Operator 455

17. Backward Convergence of a Supermartingale


Theorem. Suppose that {x(n), ,`i=(n), n E z- } is a supermartingale. Then
(a) E{x(0)} 5 E{x(-1)} 5 . -' 15 + oo.
(b) limN, x(n) = x(- oo) exists almost surely; x(- 00) A 0 is in-
tegrable; - oo < x(- oo) 5 + oo almost surely.
(c) If l < + oo, then x(- oo) is almost surely finite and left closes the
supermartingale, the supermartingale is uniformly integrable, and
the convergence in (b) is in L' as well as almost sure.

The example x(n) _ - n shows that the limit in (b) may be identically + x.
Part (a) is trivial and is stated only for orientation. Just as in the proof of
Theorem 13 but more easily here because the process is right closed, the
downcrossing inequality implies that there is an almost sure backward
limit x(- oo). Apply Fatou's lemma and the supermartingale inequality
to obtain

E{x(-oo) A 0) Z lim E{x(n) A 0) -a E{x(0) A 01, (17.1)

which implies that x(- o0) A 0 is integrable. In view of Theorem 16 and


the fact that the process can be written as the suet of a positive supermar-
tingale and a uniformly integrable martingale according to

x(n) = [x(n) - E{x(0)j.F(n)}] + E{x(0)1.fln)}, (17.2)

it is sufficient to prove the present theorem for a positive supermartingale,


and positivity will be assumed from now on. If I < +oo, the process is L'
bounded; so x(- oo) is integrable. Finally, according to Fatou's lemma,
E{x(-oo)) 51, whereas for every c the process A c is a bounded
supermartingale; so for all n,

E{x(-00) A c} = lim E{x(m) n c} Z E(x(n) A c). (17.3)

This inequality yields E(x(-oo)) Z E{x(n)} and thereby E{x(-oo)} = 1.


Thus there is equality in Fatou's lemma which implies uniform integrability
of the process.
The rephrasing of Theorem 17 for other parameter sets is left to the
reader (see Section 14).

18. The T Operator


This probabilistic operator is the counterpart of the r operator in classical
potential theory. Let be a stochastic process with state space
(68, R(R)) and with an arbitrary linearly ordered parameter set I. It is
456 2.111. Elements of Martingale Theory

supposed that the process random variables are integrable. Let S and T
be countably valued optional times with values in 1 and with S:5 T. Suppose
that x(T) is integrable, and define the process writing zsTx(t)
instead Of TSTX(') (t), by

_ x(t) on IS > t} v IT < t},


TSTY(t) -1E{x(T)IF(t)} (18.1)
on {S5 t 5 T}.

Observe that

TSTx(t) = E(x(T v t)I.flt)} a.s. on {S:5 t). (18.2)

The process is adapted to and TSTx(T) = x(T) almost surely


by Section 2, Example (a). Furthermore the process is a martingale
between times S and T in the sense that the process

{TSTX((S V 1) A T), .F((S V t) A T), t e 1}

is a martingale; in fact

TSTx((S v t) A T) = E{x(T)LF((S v 1) A T)} a.s.

In the most important applications is a supermartingale or


submartingale. Suppose, for example, that this process is a supermartingale.
If T is order bounded above and below, it follows from the supermartingale
optional sampling theorem that x(T) is integrable and that P{TSTx(t) <
x(t) } = 1 for all t, that is, 5 in the essential order. Furthermore
the process is then also a supermartingale (and a martingale
between times S and Tin the above sense). In fact, if s < t, then

E{tSTx(t)IF(s)} < E{x(t)I.F(s)} 5 x(s) = TSTx(s) a.e. on IS > s}

and, almost everywhere on {S < s},

E{TSTx(t)I.F(s)} = E{1{sss)TSTx(t)LF(s))
= E{l1sss1E{x(T v t)I.flt)}I.fls)}
= E{x(T v t)I.F(s)}
= E{E{x(T v t)I.F(T v s)}I.F(s)} _< E{x(T v s)I.9r(s)}
= TSTX(s)

Define TT by (18.1) with all references to S deleted. If T is order bounded


above and below, and if is a supermartingale [submartingale],
then is also one, is an essential order minorant [majorant] of x(),
19. The Natural Order Decomposition Theorem for Supermartingales 457

and is a martingale up to time Tin the sense that the process

{trx(T A t),.F(T A t),tE1}

is a martingale. Observe that if I has a first element a.


The operation on processes is the martingale theory coun-
terpart of the classical potential theory operation u i-- rBU on functions,
as defined in Section 1.11. 1 for B a ball and in Section I.VIII.I I for B an
open set (in both cases relatively compact in the domain of u). In most
probabilistic applications I = R+, the processes involved are almost surely
right continuous, and the optional times S and T need not be countably
valued. The adaptation to this context will be given in Section IV. 14.

19. The Natural Order Decomposition Theorem for


Supermartingales
In the following theorem and its proof all the supermartingales have the
same (arbitrary) parameter set and filtration.

Theorem. Let be positive supermartingales, and suppose that


(essential order) 5 x1 that is, for each parameter value the
indicated inequality is true almost surely. Then there are positive super-
martingales for which (essential order) 5 and
X10 +
The proof is a translation of that of Theorem 1.111.7, a particularly easy
example of the translation of potential theoretic reasoning into the corre-
sponding probabilistic context. Choose x; as the essential order infimum
of the class of positive supermartingales for which 5
and then choose as the essential order infimum of the class of positive
supermartingales for which x(-) :g x(-) + Then S x,(-) + x2
x(-) 2 and for any parameter value t,

r X(')] < X'10 +


for parameter
It is easy to check that the bracketed process is the same as x'2
values z t and is a positive supermartingale. Hence this process majorizes
and it follows that

z
If this inequality is evaluated at a parameter value s < t, we find that
is a positive supermartingale, majorized (essential order) by
impossible unless there is equality. The proof is complete.
458 2111. Elements of Martingale- Theory

See Section V.5 for Theorem 19 as modified for right continuous pro-
cesses in the continuous parameter case.

20. The Operators LM and GM


Let r: be a class of stochastic processes with a common
linearly ordered parameter set, on a common probability space, and adapted
to a common filtration If in the essential order of processes adapted
to there is a least supermartingale majorant [greatest submartingale
minorant] of r, this majorant [minorant] will be denoted by LMr [GM r].
Strictly speaking LMr, for example, is an equivalence class, and we shall
sometimes describe its members as its versions. The notation LMr can
refer either to the equivalence class or to one of its versions, but the intended
meaning will always be clear from the context.

Theorem. If r is a class of supermartingales and has an essential order sub-


martingale minorant, then GMr exists and is a martingale.

The proof of this theorem is formally identical with that of its classical
counterpart Theorem 1.111.2. The class ro of submartingale essential order
minorants of r contains v with and is therefore
directed upward in the essential order with essential order supremum (limit),
say an essential order minorant of r. We prove the theorem
by showing that is a martingale. For every parameter value t,
every in r, every in ro,

s 5 t,x(-) S

in the essential order. Thus and the essential order supremum


of ro is the same as that of e ro}. For every t the latter class
on the parameter set <_ t is an upward-directed set of martingales; so
is a martingale.
Special Case (Counterpart of Theorem 1.111.1). If is a super-
martingale with a submartingale essential order minorant, then the greatest
submartingale minorant of will be denoted by the tth random
variable of this process will be denoted by GMx(t), and can be
obtained as follows. For each parameter value t the process is a
supermartingale essential order minorant of and is a martingale up
to the parameter value t. As t increases, the supermartingale decreases, and
the essential order infimum and limit of this decreasing family is
The proof is formally identical with that of Theorem 1.III.1 and is left to
the reader. According to Section 5, there is an increasing sequence t. of
parameter values such that up to a standard
modification.
22. Potential Theory Reductions in a Discrete Parameter Probability Context 459

21. Supermartingale Potentials and the Riesz Decomposition


The special case in the preceding section yields the following result. If
is a supermartingale (arbitrary linearly ordered parameter set I)
and if this process has an essential order submartingale minorant, then
{GM is a martingale and GM x(t) = ess inf,,, f r,x(t). Moreover

GM (21.1)

where { .F(-)} is a positive supermartingale with the following properties:


(a) GMy(t) = 0 almost surely, for each t.
(b) inf,E1E{y(t)} = 0.
(c) ess lim,t t,y(t) = 0 almost surely, for each t.
Since the function t f--+E{ y(t) } is monotone decreasing, the infimum in (b)
is actually the limit as t increases. Each of these three properties of a positive
supermartingale implies the other two, and a positive supermartingale with
these properties will be called a supermartingale potential. Observe that in
the present context if the parameter set has a last element fi, a positive super-
martingale is a potential if and only if x(fl) = 0 almost surely.
Recall that in classical potential theory a positive superharmonic func-
tion u on a connected Greenian set D is a potential if and only if GMDu = 0,
equivalently (for a point of D), if and only if u) = 0 when B
ranges through the open relatively compact subsets of D containing ,
equivalently, if and only if limetD rBu = limstp µn(-, u) = 0 when B increases
through the open relatively compact subsets of D. These three classical con-
text conditions are respective counterparts of conditions (a)-(c) above. The
decomposition (21.1) is the counterpart of the Riesz decomposition in
classical potential theory and will accordingly also be called the Riesz
decomposition.

22. Potential Theory Reductions in a Discrete Parameter


Probability Context
(See Section 2.IV.17 for the corresponding continuous parameter theory.)
Let (12, JF, P; JF(j), j e l+) be a filtered probability space. Let {z(.), .FO I
be a positive process, and denote by r the class of supermartingales
.fl-)) majorizing z(-) in the essential order. It is supposed that I- is not
empty. The minimum of two supermartingales in r is in r; so r is directed
downward. Denote the positive supermartingale ess inf r by and
define R=,,,(+oo) = z(+oo) = 0. In reduction theory we are dealing with
equivalence classes of adapted stochastic processes under standard modifica-
tion; so is to mean either the appropriate equivalence class or one of
its members, as indicated by the context.
460 2.111. Elements of Martingale Theory

Theorem. Under the stated conditions

R=,.,(j)=esssup {E{z(T)j.-flj)}: T optional, >_j (22.1)

To prove the theorem, define y(j) as the right side of (22.1), and observe
(set T - j) that y(j) >- z(j) almost surely. Next observe that for fixed j the
class of conditional expectations in (22.1) is directed upward. In fact if S and
T are optional and 2j. define U = T or U = S according as the inequality

E{z(T)j.flj)} z E{z(S)jf(j)} (22.2)

is true or false. Then U is optional and E{z(U)j.(j)} is almost surely the


maximum of the conditional expectations in (22.2). Next observe that
f( ); is a supermartingale: so In fact, if j>0 and if the
optional time sequence S. is chosen with Sk z j for all k and is also chosen to
make the sequence monotone increasing with limit t'(j). then

EjT(I)I.f(j- I)} = limE{E{z(Sk)IF(j)}I


k-m
(j- 1)}
= lim E{z(S5)j9(j - 1) } 5 T(j - I) a.s.
k-M

Finally, if r and if T is optional and >-j, then

z'(j) ? E{z'(T)I/ (j)} ? E{z(T)j. (j)', a.s.:

so :'(j) >_ r(j) almost surely; that is, essinf r. as asserted.


We shall not pursue the general case far enough to obtain a probability
counterpart to Theorem I.XI.20 but proceed at once to the probability
counterpart of classical potential theory reductions.

Application to Reductions

Let A be a subset of Z' x C for which the process j -+ IA(j, ) is adapted to


.fl-). and let If be a positive supermartingale. Define R A.,, also to
be denoted by llA , as R.(.,,0,. Then according to Theorem 22. if we

define I&(+ c)=0.


R.,(j)=esssup {F.{z(T)l,t(T)j.(j)},Toptional, a.s. (22.3)

and we show further that if T; is the entry time >-j of A, that is,

T (w) = min {k >_ j: (k, w) E A },


23. Application to the Crossing Inequalities 461

thin

R= .,(j) = E{z(T)I(j)} as. (22.4)

To prove (22.4), observe that if T is optional and zj, then

IA(T) 5 1JT A T) < IA(T;),

and observe that IA(T A Tj) = 0 when T < Tj; so

E{z(T)1A(T)I.I(j)} <E{E{z(T)I,F(T A Tj)}I,t(T A T)I,f(j)} (22.5)


< E{z(T A T)IA(T A Tj)I-flj)} < E{z(T)Iflj)} a.s.

Thus the supremum in (22.4) is attained when T = T j, as asserted.

Specialization to the Parameter Set 7L


If the parameter set is 1, instead of Z', the corresponding analysis is easily
carried through, or we can define ,F(j) = f (n) and z(j) = z(n) for j > it to
reduce the context to that of the parameter set V. Equations (22. I ), (22.4),
and (22.5) remain true under the convention that all processes are extended
to vanish at the parameter value + ce.

23. Application to the Crossing Inequalities


We suppose that we are in the context of the preceding section with parameter
set Z., and we shall use the notation ze,('), ze,e2('). . respectively, for

OcI, 0 Q el..... Let be adapted processes on the


given filtered probability space. Suppose that 0 and that
and are positive supermartingales; that is, we are in the
context of Theorem 12. In Section 12 we defined Dn as the
number of times an sample sequence proceeds from above to below
Define

A = Ici,(o): Z 1 , ,x(j,(o) S x'(j,w)}.


B= {(j,(u): je7L,; ,x(j,w) z x"(j,(u)).

It is easy to check using (22.1) that (almost surely)

yie(O) > E{y'(n); Dn[x('); x'(-), x"(-)] Z 1IF(0)};


(23.1)
°A 8Aa(0) Z E{}' (n); x'(,),x"(,)] Z 21-F(O));
462 2.111. Elements of Martingale Theory

etc., with equality if is a martingale. Now the iterated reduction in-


equalities in Section 1.VI.3(o) were proved in the context of classical poten-
tial theory using elementary reduction properties valid in the present context.
Hence (almost surely)

.v (') + ABA(') + YAaAEA(') + .5 (23.2)

and

Yee(') +YAaAe(.) + 5 S A (23.3)

In view of (23.1) inequality (23.3) implies

E{y'(n) S x(0) n x"(0) a.s., (23.4)

which yields (12.2) on integration. (See the remarks in Section 10 on con-


ditional inequalities.) In particular let a and b be positive constants with
a < b, let { be a positive supermartingale, and set ay(-),
in (23.3) to obtain the almost sure inequality

A by(')
Y.4$(') + + ... S (23.5)
b-a
The corresponding particularization of (23.4) can be obtained by the same
substitution or by summation and integration in (23.5). The martingale
theory translation of I . VI(3.11) is the sequence

b by(-)
5 ()YAe. ... (23.6)

of almost sure inequalities. On integration these yield (12.4). The point is


that the classical context iterated reduction inequalities in 1.VI.6(o) remain
valid when translated into the martingale theory context, and that the
translated inequalities yield, on integration, the crossing inequalities proved
directly in Section 12.
Chapter IV

Basic Properties of Continuous Parameter


Supermartingales

1. Continuity Properties
The continuity properties of continuous parameter supermartingales derived
in this section are of course also valid for continuous parameter sub-
martingales and martingales. Recall that a process is said to be [almost
surely] right continuous if [almost] every sample function is right continuous
and the process is said to be [almost surely] right continuous with left limits
if [almost] every sample function is right continuous and has a left limit
at every point.
The usual smoothness hypothesis to be imposed on a continuous para-
meter supermartingale is almost sure right continuity. Moreover the
following remark shows that it is possible to change the reference filtration
of an almost surely right continuous supermartingale to obtain a new one
to which the supermartingale is adapted, which is right continuous, and for
which each a algebra contains the null sets. Let {x(t), .F(t), t e R+ } be an
almost surely right continuous supermartingale, and define as the
a algebra generated by .F(t) and the null sets. Then is a super-
martingale. Moreover the filtration is right continuous, .F,+(0) con-
tains the null sets, and (x(-),.Fl" is a supermartingale because ifs' < t,

E{x(t)j.3F,(s')} 5 x(s') a.s.,

and when s' Is sequentially, we obtain the supermartingale inequality for


the process in view of the conditional expectation continuity
theorem.
In the following theorem convergence in measure has the usual definition,
not the special definition for R valued random variables needed in Section
1.13.

Theorem. Let {x(t), F(t), t e R' } be a supermartingale, and suppose that


.r(0) contains the null sets. Let B [B+] be the set of values oft in R' for which
it isfalse that lim,.., x(s) = x(t) [lim,l, x(s) = x(t)] in the sense of convergence
in measure, and let A be a countable dense subset of R+.
464 2.IV. Basic Properties of Continuous Parameter Supermartingales

(a) The set B is countable.


(b) The restriction to A of almost every sample function is bounded
on the trace on A of every compact interval, has left and right limits
at every point of R', and fort in R+ - B',

P { lim x(s) =
e 3slr
x(t) = 1.
J

(c) If t e B, define y(t) = x(t) ; if t e R' - B+, define y(t) = x(s)


where this limit exists, and define y(t) arbitrarily where this limit does
not exist. For all t define x(t) = lim4asl,x(s) where this limit exists,
and define x(t) arbitrarily where this limit does not exist. Then
(c l) The processes {y(-), .F +(-))are supermartingales,
martingales if is a martingale.
(c2) Almost every sample function and almost every sample
function are bounded on compact intervals of R+
(c3) Almost every sample function has right and left limits at
every point of R+ and is right continuous at every point of
R+ - B+. Almost every sample function is right continuous
and has a left limit at every point of R+. The processes
are L' bounded on compact intervals
ofR+.
(c4) The process is a standard modification of indistin-
guishable from if and only if B+ is empty, as is true if
is a martingale and

Proof of (a) and (b). If to > 0, the restriction to A n [0, to] of almost every
process sample function is bounded, according to the inequalities
111(9.1') applied to - x(-). If such a restriction does not have a left and right
limit at every point of [0, to[, there must be a pair of rational numbers
rt , r2 with rt < r2 such that the sample function restriction to A n [0, to[
has infinitely many downcrossings of [r1 , r2]. Since the downcrossing in-
equality 111(12.5) yields
E{x(0) A r2 - x(to) A r2)
5 (1.1)
r2 - r t
the number of downcrossings is almost surely finite simultaneously for
all rational pairs rt, r2 and values of to. Therefore the restriction to A of
almost every sample function has one-sided limits x(t-) and x(t+) for
every t. Moreover, since

E{!x(t)j} = E{x(t)} - 2E{x(t) A 01 < E{x(0)} - 2E(x(to) A 0) (t s to),


(1.2)
1. Continuity Properties 465

an application of Fatou's lemma shows that these one-sided limits are


integrable and that the processes are Lt bounded on
compact intervals of W. Now consider the space of almost surely finite-
valued random variables, identifying two random variables if they are equal
almost surely, and define the distance between the random variables x and y
as E{ I x - Y J A 1). Convergence of a sequence in this metric is equivalent
to convergence in measure of the corresponding sequence of random
variables. Let 4(t) be the point of this metric space corresponding to the
random variable x(t). Observe that if At is a second countable dense subset
of 68+, the results proved above for A when applied to A u At show that the
left- and right-hand sample function limits are almost surely the same for
A as for A t . Among other things, this fact implies that when s tends sequen-
tially and strictly monotonely to t, the random variable x(s) tends almost
surely to x(t-) or x(t+) depending on the direction of approach; so the
function ¢ on R has left and right limits at all points. If e > 0, the set of
points t at which the oscillation of 0 is at least a is closed and can have no
finite limit point; so this set is finite in each compact interval, and it follows
that the set of discontinuities of 0, which is the set B of the theorem, is
countable. Moreover the definitions of B and B+ can now be strengthened:
tefR+ - B [teER+ - B+] if and only if whenever A is a countable subset
of R with [left] limit point t, it follows that

A lim x(s) = x(t) [Alim x(s) = x(t)] a.s. 13

Proof of (cl)-(c4). If is a martingale, if .F(-) is right continuous,


and if 0:5 s' < s < t, the almost sure equality x(s') = E{x(t)jF(s')} com-
bined with the conditional expectation continuity theorem yields x(s+) =
E(x(t)j.fls)} = x(s) almost surely when s' tends sequentially and mono-
tonely to s. Thus in this case B+ is empty. To prove the rest of (cl)-(c4),
observe that the process is a standard modification of and is
adapted to because .F(0) contains the null sets. Hence is a
supermartingale, a martingale if is a martingale. If we choose
A to be a superset of B, an application of (a) and (b) shows that (c l)-(c4) are
true for y(-) and that is an almost surely right continuous
adapted process with left limits. If B+ is empty, x(-) is adapted to
Furthermore P{x(t) = x(t)} = 1 if and only if teR+ - B. To prove that
is a supermartingale, which implies that this process is a mar-
tingale if is [apply the first result to suppose that
05s<s'<t<t'<tp,andobserve that
E{x(t') - E{x(to)I_1F(t')}IJr(s')} + E(x(to)j.fls )} < x(s') a.s. (1.3)

When t' j t sequentially and s' j s sequentially, apply Fatou's lemma for
conditional expectations together with the conditional expectation con-
tinuity theorem to derive
466 2.IV. Basic Properties of Continuous Parameter Supermartingales

E{x(t) - E{x(to)I,F+(t)}I.F+(s)} + s x(s) a.s.

Hence E{x(t)J.F' (s)} < x(s) almost surely; so the process {xO, +O} is a
supermartingale, and the proof of (cl)-(c4) for is now complete.

The Right Continuous Case

In most applications the process in Theorem I is given as an


almost surely right continuous supermartingale. In this context Theorem I
implies that the process almost surely has finite left limits, that the processes
and are L' bounded on compact intervals of R+, and that there is
an at most countable subset B of R' such that P{x(t) = x(t-)} = I if and
only if t e I8' - B. We stress that even if B is empty, the process is not neces-
sarily almost surely continuous.

Semipolar Sets and Supermartingale Smoothing

Recall that in the Fundamental Convergence Theorem (Theorem IN]. 1) of


classical potential theory a basic operation was the lower semicontinuous
smoothing of the infimum of a locally lower bounded family of super-
harmonic functions. This smoothing yielded a superharmonic function equal
quasi everywhere to the infimum function. The corresponding smoothing in
the parabolic context (Section 1.XVII.13) yielded a superparabolic function
equal to the infimum function up to a semipolar set. In these contexts the
infimum function satisfies the average property of the function class con-
sidered, the class of superharmonic or superparabolic functions as the case
may be, but does not satisfy the lower semicontinuity property of the func-
tions in the class except in trivial cases. The lower semicontinuous smoothing
replaces the infimum function by a function in the class equal to the infimum
function off a small set. The analogous smoothing operation in the prob-
ability context is the smoothing of into in Theorem 1. This will be
seen in the context of the probabilistic Fundamental Convergence Theorem
(Theorem 5 below). The following application of Theorem 1 is an illustration
of this smoothing.

The Set of Discontinuities of a Supermartingale

Let R be a filtered probability space with right


continuous and F(O) containing the null sets. Let be a nearly
progressively measurable supermartingale on the space, and suppose that
almost no sample function has an oscillatory discontinuity; that is, the
I. Continuity Properties 467

process almost surely has right and left limits. Define x(t,w) = x(t+,w)
[x(t, (o) = x(t - , w)] if this right [left] limit exists, and define x(t, w)
[x(t, (o)] arbitrarily otherwise. Then we show that x(, ) = x(, ) = x(, )
up to a semipolar subset of R` x Q. Since an evanescent set is semipolar, we
lose no generality if we adjust the process if necessary on a null set of Q to
make the process progressively measurable with no sample function having
an oscillatory discontinuity, and we shall suppose that this has been done.
Define

Ak={ (t,(o):Ix(t,(0)-x(t,w)Izk}.

The process {x(), ,F(-)) is progressively measurable because it is right con-


tinuous (Section 1.2); so the process { Jx() - x()I, f()} is progressively
measurable, and therefore Ak is a progressively measurable subset of
R+ x Q. Since the sample function x(, w) has no oscillatory discontinuity.
the set of values of t with (t, co) a Ak has no finite limit point; if k and n are
strictly positive integers, let Tk be the nth value of tin this set (with Tk = + oe
if there is no nth value). Then (Section II.4) Ti. is the nth entry time of the
progressively measurable set Ak ; so Tkn is optional. Furthermore

{x(,)#x(,)}_ UAk= U [TIA;


k=1 k.n=l

so the set on the left is semipolar as asserted, and similarly the set {x(,) #
x(, ) } is semipolar.
Returning to the original hypotheses of this application of Theorem 1,
observe that one choice of x(, ), in fact the choice we shall always use below
whenever x() almost surely has right limits, is x(t) = lim inf.,,, x(s). This
choice makes {x(), ()} progressively measurable if fl-) is right con-
tinuous even if {`x()()} is not progressively measurable.

Application of Theorem 1 to a Conditional Expectation Process


x(t) = E(zl.flt)}
Let (Q, .9r, P; .f(t), t e I8') be a probability measure space with a right con-
tinuous filtration, let z be an integrable random variable on the space, and
define a martingale {x(), .F()} by setting x(t) = E{zlf(t)}. Right con-
tinuity of the filtration implies that the set B4 of Theorem I is empty. Accord-
ing to Theorem 1, the process x() is a standard modification of x() and is
almost surely right continuous. In other words the conditional expectation
defining x(t) can be chosen to make x() an almost surely right continuous
process, or even right continuous if ..(O) contains the null sets.
468 2.IV Basic Properties of Continuous Parameter Supermartingales

Application to the Decomposition of a Right Closable Supermartingale

Let { x(t), F(t); t e R' } be a right closable almost surely right continuous
supermartingale, and suppose that is right continuous. If x right closes
the supermartingale and if E{xlF(t)} is chosen properly, the equation

x(t) = [x(t) - E{xj.F(t)}] + E{xj.9r(t)} (1.4)

exhibits as the sum of a positive almost surely right continuous super-


martingale and an almost surely right continuous uniformly integrable mar-
tingale. The discrete parameter version of this decomposition was used in
Section 111. 15. Going farther, the representation

E{xjF(t)} = E{x v 015f(t)} - E{(-x) v 0I.9F(t)}, (1.5)

with suitable choices of the conditional expectations, exhibits the martingale


component in (1.5) as the difference between two positive almost surely right
continuous uniformly integrable martingales.

2. Optional Sampling of Uniformly Integrable Continuous


Parameter Martingales
Recall [Section III.3(e)] that a martingale is uniformly integrable if and only
if it is right closable. The following theorem is a continuous parameter ver-
sion of the results obtained in Section 111.2, Example (a), and is proved using
those results.

Theorem. Let (fl, .9, P; .9r(t), t e Ut') be a filtered probability space. If


9(-)) is an almost surely right continuous uniformly integrable martin-
gale, closable, say by {x(+ oo ), F(+ cc) ',,and if S and Tare optional (S + oo),
then

x(T) = E{.r(+ oo)I.9=(T)} a.s. (2.1)

and

E{x(S)jJr(T)} = E{x(T)j9(S)} = x(S A T) a.s. (2.2)

Observe that we have not supposed that is right continuous. To


avoid trivia, enlarge the filtration if necessary to make F(0) contain the null
sets. Define [T)., an optional time for and also for the discrete filtration
(. 9 Q 2 - " ) , j = 0, 1, ... , + oo ), as in Section 11.2, and define [S]" correspond-
ingly. According to Section 111.2, Example (a),
2. Optional Sampling of Uniformly Integrable Continuous Parameter Martingales 469

x([S]") = E{x(+oc)j.f([S]")} a.s. (2.3)

and

E{x([S]")Ic([T]")} = x([S]n A [T]") a.s. (2.4)

The integrated version of (2.4) is

fx([S]n)dP= Jx([s]n A [T])dP (2.5)

for Ae.F([T]"). Hence (2.5) is true for A in the smaller a algebra .F(T).
When n -+ oo, we can integrate to the limit on the left side of (2.5) because
the sequence of integrands, a sequence of conditional expectations of x(+ oc)
according to (2.3), is uniformly integrable [Section 1.4(i)]. Since the same
argument is applicable to the right side of (2.5), equation (2.5) implies that
x(S) and x(S A T) are integrable and that

VJx(s)dP = A T)dP, Ae(T). (2.6)

Furthermore x(S A T) is .F(S A T) measurable (Section 11.3) and so -,F(T)


measurable. Equation (2.6) is the integrated version of the almost everywhere
equality between the first and third terms of (2.2). Interchange S and T to
derive the almost everywhere equality between the second and third terms.
Set S = +oo in (2.2) to obtain (2.1).

Application of Theorem 2 to Totally Inaccessible Optional Times

Let T be a totally inaccessible optional time, and define

a(t) = 1(r,+.Dt(t)1
a"(t) = E{a((k + l)2-")j.F(t)} on I"k = [k2-",(k + 1)2-"[, ke7L+.

Here we use the versions of the conditional expectations making an


almost surely right continuous process. Observe that I z a"(t) z an+1(t) z
a(t) almost surely, simultaneously for all t. We now show that almost surely
lim"... a"(t) = a(t) uniformly on I. To see this, choose E > 0, and define

T"=inf{t: a"(t)-a(t)>E}.
Then T is an increasing sequence of optional times. Let T. = lim"_. T.
Since a"(t) = a(t) = I for sufficiently large t almost everywhere on
{ T < + oo }, and for t in I", the evaluation
470 2.IV. Basic Properties of Continuous Parameter Supermartmgales

an(t) = P{T < (k + l)2-"I.flt)}

implies that lim,..,, an(t) = a(t) = 0 almost everywhere on {T = +oo}


(by the conditional expectation continuity theorem), it is sufficient to prove
that T. = + oo almost surely. Now

E{an(Tn)llr +co)} _ E{a((k+ 1)2-")I.W(T")}dP


k =O flTne Ink)

_ Y_ a((k + 1)2-")dP
k_OJ {TneInk)
2.7)

<Y
cc

k_0
f a(Tn + 2-') dP
Tne/nk)

= E{a(T,, + 2-")1(Tn<+m)} = P{T < T. + 2-" < +oc}


5P{TSToo +2-",T<+oo}.
Moreover by definition of

Iim E{a(Tn) I lim P{T < Tn, T < + oo }, (2.8)

and by definition of T.

E{[an(Tf) - a(Tn)]l)rn<+m)} > eP{Tn < +oo} z eP{T,o < +oo}. (2.9)

Thus (2.7)-(2.9) lead to

P{T<T, T<+oo}-limP{TST.,T<+oo}ZiP{T,,<+oo}.
2.10)

Finally the total inaccessibility of T implies that the limit in (2.10) is


P{T S T., T < +oo} and thereby implies that P{T,, < +oo} = 0, as was
to be proved.

3. Optional Sampling and Convergence of Continuous


Parameter Supermartingales
Theorem. Let {x(t), .fi(t), t e R+ } be a right continuous right closable super-
martingale. Then
(a) lim,_. x(t) = x(+ oo) exists a.s. and right closes the supermartingale.
(b) If T(5 + oo) is optional, it follows that x(T) is measurable and
integrable.
3. Optional Sampling and Convergence of Continuous Parameter Supermartingales 471

(c) If S and Tare optional with S < T:5 + x, it follows that

E{x(T)j.1(S)} <x(S) a.s. (3.1)

Observation (a). Almost sure right continuity of instead of right con-


tinuity, is sufficient for the proof of (a), but under this weakened hypothesis
x(T) may not be measurable, and so (b) and (c) may fail. If ,y (0) contains the
null sets, however, the theorem is true with no change in proof even if is
only almost surely right continuous because (Section 1.8) is indistin-
guishable from a right continuous process adapted to If the supermar-
tingale is almost surely right continuous and if ,F(0) does not contain the
null sets, each a algebra .F(t) can be replaced by the a algebra generated by
,F(t), and the null sets to get an enlarged family of a algebras for which the
preceding argument becomes applicable.

Observation (b). If the process in the theorem is not right closable but if
S and T are bounded, say T:5 c, the process restricted to the parameter
interval [0, c[ is right closable, and the theorem is applicable with a trivial
reformulation to the process on this parameter interval. In particular, and
we state the following for martingales as well as for supermartingales since
the supermartingale result can be applied to both and in the
martingale case, if {x(t), .fi(t), t e 118+ } is a right continuous supermartingale
[martingale] and if T is optional, the process {x(T A t), .5(T A t), t e I8+ },
the process stopped at T, is a supermartingale [martingale]. A slight
refinement of the proof below (see Section 111.8 for the discrete parameter
case) yields the stronger result that {x(T A t), '(t), t e I8+ } is a supermar-
tingale [martingale].
Since the given supermartingale is right closable, the almost sure limit
x(+ oc) exists (Section 111.15). and the reasoning used in the discrete parame-
ter case in that section shows that x(+ cc) right closes the supermartingale.
Since the process is right continuous and therefore progressively mea-
surable, the function x(T) is measurable. The process {x(j2-"),S (j2-"),
j = 0, 1, ... , o--, } is a supermartingale and [T] as defined in Section 11.2 is
optional for the indicated discrete a algebra family. Hence (Theorem 111.7)

x([S].)
E{x([T].) - E{x(+x) (3.2)
a.s.

In view of Fatou's lemma for conditional expectations this inequality yields


(n -i x)
x(S) > E{.r(T)-E{v(+a.)jy'(T)}Ly+(S)} +E{x(+x)j, '(S)} a.s.
(3.3)
472 2.1V. Basic Properties of Continuous Parameter Supermarungales

If S vanishes identically (3.3) implies that x(T) is integrable. For general


S inequality (3.3) yields

x(S) -
E{ -I .F(S)} performed on both sides of this inequality
yields (c).

Extension of Theorem 3 Involving Left Limits

Under the hypotheses of Theorem 3 we have seen that the process x(-) has
almost sure left limits. Suppose that Sand Tare as described in that theorem
and that T is an increasing sequence of optional times with limit T and with
T. = Ton IS = T}. Define x-(T) = x(T,J. We shall now prove that
the ordered triple

[X(S),. (S)]. [x(T),.F(T)]


0

is a supermartingale, In particular, if T. = T for all n, this result is a re-


writing of part of Theorem 3. At the other extreme, if P{T = T > 0} = 0
for all n, which implies that T is a predictable optional time announced by
T, the assertion is that the ordered triple

[x(S),.f(S)], [x(T-), F(T-)], [x(T), F(T)]


is a supermartingale. [See the definition of ,f(T-) in Section 11.7.] More-
over if the given process is a right closable martingale, these triples are
martingales.
It must be proved that x-(T) is integrable and

(a) E{x-(T)jflS)} S x(S),


(b) E j(T) I Y.F(T.)l 5 x-(T) a.s.
(3.4)

IX o

According to Theorem 3, 5 almost surely, and when


n - oo this inequality becomes (3.4b) by virtue of the conditional expectation
continuity theorem. In view of the decomposition (1.4) of x(-) into the sum
of a positive supermartingale and a uniformly integrable martingale it is
sufficient to prove that (3.4a) is true for a positive supermartingale and also
for a uniformly integrable martingale. If x(-) L* 0 and if c is a positive con-
stant, the following supermartingale inequalities for

E{x(T 5 x(S) a.s.


4. Increasing Sequences of Supermartingales 473

yield when n -+ oo the fact that x-(T) is integrable and (3.4a) is true. If
x(-) is a uniformly integrable martingale, it is a class D martingale (Section
111.6); so the martingale equality

[Ac-.(S)]
ii
fro m Theorem 3 yields the integrated form of (3.4a) when n - oo.

4. Increasing Sequences of Supermartingales


Theorem. Let {x.(t), .fi(t), t e R+ } be an increasing sequence of almost surely
right continuous supermartingales, with limit process
(a) If E{x(0) } < + oo, the process fl-)) is a supermartingale.
(b) The process is almost surely right continuous with left limits,
and if T is the hitting time by (t,(o) of the set {(t,w): x(t,w) < +oo},
then for t > T almost every sample function is finite valued and
has finite left limits.

Assertion (a) is trivial. In proving (b) it can be assumed (Section 1),


enlarging each a algebra .fi(t) if necessary, that is right continuous,
that 9(0) contains the null sets, and that each process is right con-
tinuous. It is sufficient to prove the theorem for x.(-) right closable because
it is sufficient to derive the conclusions for each parameter interval [0, b[.
We can nevertheless continue to use the parameter interval R. Furthermore
it can be assumed that each process is positive because (Section 1)
is the sum of a right continuous martingale yo(-) and a positive right
continuous supermartingale, and we can replace x.(-) by yo(-) to
get an increasing sequence of positive right continuous supermartingales.
The truth of (b) for this sequence implies its truth for the given sequence.
Proof of almost sure right continuity of x(-). It is sufficient to prove
almost sure right continuity for a bounded process since this result
can be applied to the sequence A c. To prove almost sure right con-
tinuity in this case, under the simplifying assumptions developed above,
observe first that is progressively measurable since this process
is the limit of a sequence of progressively measurable processes. Hence
(Section 11.6) it is sufficient to prove that whenever T is a finite-valued
optional time, almost every x(-) sample function is right continuous at T.
In fact in the present context it is sufficient to prove that almost every
sample function is right continuous at 0 because this result can be applied
to the increasing sequence {x.(T + t), F(T + t), 1E R+} of right continuous
supermartingales to get almost surely right continuous at T. To prove
almost sure right continuity at 0, observe that since x is right continuous,
474 2.1V Basic Properties of Continuous Parameter Supermartingales

the sample functions are right lower semicontinuous. Furthermore the


restriction of the supermartingale to the parameter set of strictly positive
rationals almost surely has a right limit y at the origin according to Section
111. 17, and

x(0) < lim inf x(t) < y a.s. (4.1)


r-0

On the other hand, the supermartingale inequality E{x(0)} > E{x(t)}


implies

E{x(0)} > E{y}. (4.2)

It follows that

x(0) = lim inf x(t) = y a.s.

and the right lower semicontinuity of then implies that

lim sup x(t) = lim sup x(r) = y as. (r rational).


r0
Thus is almost surely right continuous at 0, and hence as pointed out
above, is almost surely right continuous. Almost sure right continuity
implies (Section 1) almost sure left limits at all points.
Proof of (b). To finish the proof of (b), it is sufficient to prove (b'): If
a and fi are strictly positive numbers, then for almost every co in the set
{x(a) < /3} the values x(t, w) and x(t+, co) are finite for t > of. To reduce
this fact to what we have already proved, define z.(1) for t > 0 as xx(a + t)
on
Z.({x(a) < /3} and as 0 elsewhere. Apply (a) to the increasing sequence
1),.9r(a + t), t e Q8+ } of supermartingales to find that z ac-
cording to what we have proved already is an almost surely right continuous
supermartingale and that therefore almost every sample function is bounded
on compact intervals. The remaining part of (b) follows.

Elementary Proof of Almost Sure Right Continuity of

The proof given above of almost sure right continuity of is natural,


but it is of interest to exhibit an elementary proof that does not involve
the relatively deep theorems on hitting probabilities. The following proof
does not even need the supermartingale convergence theorems. We can
assume as in the proof already given that is right continuous, that F(0)
contains the null sets, that each process x is positive and right continuous,
and that is bounded. Define x(+ oo) = x (+ oo) = 0. Observe first
4. Increasing Sequences of Supermartingales 475

that if S. is a sequence of optional times and if A = S. = 0}, then


Ae.,F(O), and an application of the right lower semicontinuity of
sample functions and of the supermartingale inequality at optional times
[valid for because it is valid for yields

E{x(0);A} Ejliminfx(S,);Ar <liminfE{x(S,);A) <E{x(0);A}

from which it follows in view of sample function right lower semicontinuity


that

x(O) (4.3)

almost everywhere on A. Next choose e > 0 and define

x (t) - x(0) > -21, T = lim sup T,,.


I
n-cO

The time T is optional because up to a null set

T. z c} = sup I x (r) - x(0)I < 2 (r rational, c > 0)

and therefore T is also optional. Furthermore T is almost surely strictly


positive because otherwise, there is a set of strictly positive probability
on which lim, T. = 0 and
E
I x.(T.) - x(0)I z 2 (4.4)

for sufficiently large n. Now under (4.4) either x


x(0) >- e/2 for sufficiently large n, which leads to lim inf, x(0)
e/2 and thereby contradicts (4.3) or under (4.4), for every k,
x x(0) < -e/2 for infinitely many values of n, in which case xk(0) -
x(0) < -e/2, which is impossible for large k. Thus Tis almost surely strictly
positive. Furthermore I x (t) - x(0)I < e/2 on [0, and so almost surely
1x(l) - x(0)I < e/2 on [0, T[, and if we set

r(t) = lim sup Ix(s) - x(t)I,


6-0 r<s<,+b

then almost surely r(t) < e on [0, T[. The class 1'e of almost surely strictly
positive optional times T for which almost surely r(t) < e on [0, T[ is not
empty according to what we have just proved, is directed upward, and is
476 2.IV. Basic Properties of Continuous Parameter Supermartingales

closed under countable suprema; so r, contains its own essential supremum


T. This is impossible unless ' = + oo almost surely because the above
proof that r, is not empty when applied to the process x(T' A k + ) shows
that there is a member of r' almost surely strictly larger than ' A k for
every k. Thus for every a the right oscillation of almost every sample
function is at most a at all points, and the proof is complete.

Generalization to Upward-Directed Supermartingale Families

Observe that Theorem 4 can be applied to an upward-directed family of


supermartingales (essential order) since according to Section 111.5 the es-
sential order supremum of the family is a sequential supremum.

5. Probability Version of the Fundamental Convergence


Theorem of Potential Theory

The Right Continuous Smoothing of a Supermartingale

Let {x(t), 9r(t),tER+} be a supermartingale with almost sure right limits;


that is, we suppose that almost every sample function has a right limit
at every parameter value. According to Section 1, the process also has
almost sure left limits. From now on in this context we define

x(t) = lim inf x(s), (5.1)


alr

thus making the definition of slightly more specific than that in Section
1. According to Section 1, if flO) contains the null sets and if is right
continuous, the process is an almost surely right continuous
supermartingale and up to a semipolar subset of R+ x 0;
in particular. P{x(t) = x(t)} = I up to a countable set of values of t. The
(almost surely) right continuous smoothing plays the same role for
as the lower semicontinuous smoothing of a function plays in classical
and parabolic potential theory, according to the theorem of this section.
The martingale theory concept corresponding to that of a locally lower
bounded family of superharmonic functions is a locally lower bounded
family of supermartingales. Although local lower boundedness is not diffi-
cult to formulate in the martingale theory context, we shall restrict ourselves
to families of positive supermartingales to simplify the discussion.

Theorem. Let a e f } be a family of positive almost surely right


continuous supermartingale.s on a filtered probability space
5. Probability Version of the Fundamental Convergence Theorem of Potential Theory 477

(0, 9, P;.(t), t e 68+).


It is supposed that is right continuous and that 9(0) contains the null sets.
(a) There is then a process determined up to indistinguish-
ability by the following conditions.
(al) fl-)) is nearly progressively measurable.
(a2)If a eJ, then 5 x 0.
(a3)If (y(-), fl-)) is a process satisfying (a]) and (a2), then S
x Q.
(b) The process in (a) is a supermartingale with almost sure
right and left limits. If is the (almost surely) right continuous
smoothing of then up to an evanescent subset of Q2+ x S2

+(t) = x(t+), x(t-) = x(t-), x(t) S x(t), (5.2)

and there is equality in the inequality up to a semipolar subset of


R+ x Q.
(c) For some countable subset Jo of J (Jo = J ijJ is countable) the process
in (a).
(d) Conversely, if 4 is a semipolar subset of R+ x 11, there is a decreasing
sequence {x n e Z+ } of right continuous positive supermar-
tingales such that if inf. E1+ then d e up
to an evanescent set.

The notation esss infaE f If satisfies the conditions under


(a), then every process indistinguishable from also satisfies these con-
ditions. The equivalence class of these processes under indistinguishability
and in the usual notational abuse the individual members of this class will
be denoted by ess* inf,E f The equivalence class of right continuous
smoothings and its individual members will be denoted by N,E,
when lattice concepts in martingale theory are discussed systematically in
Chapter V. A right continuous smoothing is up to indistinguishability
the maximum almost surely right continuous supermartingale majorized
in the essential order by every supermartingale

Theorem 5 as a Convergence Theorem

There is no loss of generality in supposing that the given supermartingale


family is directed downward (pointwise order) because if the given family
is enlarged to the family of minima of finite subsets, the enlarged family
is directed downward and this enlargement does not change the scope of the
theorem's statement. The process in the theorem is a version of the
essential order limit (and infimum) of the downward-directed family. In
particular, if J is countable, the enlarged supermartingale family is also
478 2.1 V. Basic Properties of Continuous Parameter Supermarungales

countable and without loss of generality can be replaced by a decreasing


sequence of its members. (If the given family has a smallest member, the
decreasing sequence has all but a finite number of its members this smallest
member of the original family.)

Deletion of "Almost Surely"

Without loss of generality we can suppose that every process is right


continuous because can be made so by a change on a null set (depending
on a) and this change does not affect the theorem. We shall suppose through-
out the proof that this change has been made.

Proof of (a)-(c) for J countable. As observed above there is no loss of gener-


ality in this case if we suppose that the family of supermartingales is a
decreasing sequence {x n e 1' } of positive right continuous super-
martingales. The process x is a supermartingale (Section
111.5), progressively measurable because each process is right
continuous and therefore progressively measurable. The other conditions
under (a) are trivially satisfied. Observe that E{x (0)} 5 E{xo(0)} and there-
by conclude that the expected number of downcrossings of an interval by
is bounded because this expected number for x is bounded by a
number independent of n. Hence has almost sure left and right limits.
To prove (5.2) observe that the first equation is the definition of x(t), that
the second equation follows from the first and that the inequality follows
from the right continuity of x.(-), which implies right upper semicontinuity
of Finally there is equality up to a semipolar set in this inequality
according to Section 1. o

Proof of (a)-(c) for J uncountable. Choose any countable subset Jo of J, and


define inf., j. According to the preceding discussion with count-
able J, the process -fl-)) is a progressively measurable supermartingale
and (b) is true. To prove (a)-(c), it suffices to show that Jo can be chosen
so large that (a2) is true. First choose Jo so large (Section 111.5) that
is a version of ess inf,, j that is, for each t the random variable x(t)
is a version of ess inf., f x,(t). Then for a in J it follows that 5
almost surely on the set of positive rationals; so S almost surely
on R' in view of sample function right continuity of The exceptional
null set depends on a. With this preliminary choice of Jo define

M = { (t, w) : w) and w) have left and right limits at t and


x(t+,w) = X(1, (j)) = x(t+,w), x(t-,(0) = x(t-,(0)),
{(t, w): w) is continuous at t},
5. Probability Version of the Fundamental Convergence Theorem of Potential Theory 479

Bd = M n {(t, co): co) is discontinuous at t}.

The set (08+ x S) - M is evanescent. When (t, (o) a E , the sample functions
co) and w) are continuous at t and equal there; so for each a in J,
x(, ) = x(, ) < x2(, ) quasi everywhere on A . The problem is to enlarge
J, to ensure that also x(, ) < x.(, ) quasi everywhere on Bd. If J' is a
countable superset of Jo and if x'() = inf8E J x.(), then x'() = x() almost
surely on the positive rationals; so the almost surely right continuous
processes x() and x'(-) are indistinguishable. Moreover and Bd are
changed by at most evanescent sets when J. is increased to Y. According
to the discussion of the discontinuities of a supermartingale at the end of
Section 1, there is a sequence T of optional times the union of whose graphs
covers h, Choose J' so large that infaE,,..xa(T") is a version of ess inffE (T")
for all n. Then for each a in J, x'(,) 5 x.(, ) quasi everywhere on U8+ x S2;
the exceptional evanescent set depends on a. The countable set J' is the final
choice of Jn and has the properties described in (c).

Proof of (d). We can suppose that A c Uo QSkj for some sequence of


optional times. For k and n in 1+ define

2-k on Q0, Sk + 2-"Q,


xk"(, ) x"(t) _ Y_ xk"(t).
0 on QSk + 2 ", + oo Q

The process {x"(), F(-)) is a positive right continuous supermartingale


because each summand process is. The sequence {.x"(), ,1F(), n e 71+ } is a
decreasing sequence with limit process {x(), .f () } for which

so (d) is true.

A Special Property of ess* inf2E, xa

Let y() be a version of this infimum process. Since y() is equal quasi every-
where to the infimum of a countable set of almost surely right continuous
positive supermartingales, equivalently, equal quasi everywhere to the limit
of a decreasing sequence of almost surely right continuous positive super-
martingales, the fact that the optional sampling Theorem 5 is applicable
to an almost surely right continuous positive supermartingale implies that
this theorem is applicable to y().
480 2.1 V. Basic Properties of Continuous Parameter Supermartingales

6. Quasi-Bounded Positive Supermartingales; Generation of


Supermartingale Potentials by Increasing Processes
Let I be a linearly ordered set, and let {.fi(t), to 1} be a filtration of a prob-
ability space.

Quasi-Bounded Positive Supermartingales

A positive supermartingale {x( ), is called quasi-bounded if it is the


sum of a series of bounded (adapted to positive supermartingales.

Increasing Processes

From now on in this section we shall suppose that I has a first point, denoted
by 0, but no last point. A process {A(t), t e 11, not necessarily adapted to
#'(-), with state space (18+,.(R )), will be called an increasing process if
the following conditions are satisfied.
(a) A(0) = 0, P{A(s) S A(t)} = 1 when s S t;
(b) E{A(t)} < +oo for tel.
Let +co" denote a point not in I, and define A(+oo) = ess sup,. , A (t).
If is Lt bounded, that is, if E{A(+oo)} < +oo, define (uniquely up
to a standard modification) a supermartingale {x(t), .F(t), t e 1) by

x(t) = E{A(+oo) - (6.1)

This supermartingale is a supermartingale potential because

limE{x(t)} = limE{A(+co) - A(t)} = 0,

and this supermartingale potential is said to be generated by A(-). Observe


that if T is a countably valued optional time with values in I, an application
of Section 111.2, Example (a), shows that

x(T) < E{A(+oo)j.r(T)} a.s;

so [Section 1.4(i)] the family

{x(T): T optional, countably valued, with values in 1}

is uniformly integrable; that is, the process is in the class D


defined in Section II.11. Moreover, if k is a positive integer, the process
defined by
6. Quasi-Bounded Positive Supermartingales and Increasing Processes 481

Ak(t) = A(t) n (k + 1) - A(t) n k, to 1,

is also an L' bounded increasing process, generating a supermartingale


potential which can be chosen to be bounded by 1, and
Eo up to a standard modification. Thus a supermartingale potential
generated by an L' bounded increasing process is quasi-bounded and in
the class D. Conversely, we shall prove in Section 10 that every class D or
quasi-bounded supermartingale potential can be generated by an L' bounded
increasing process. The corresponding results for the discrete and continuous
parameter cases are proved, respectively, in Sections 8 and 11.

Adapted Increasing Processes

If is adapted to (6.1) can be written in the form

x(t) = E{A(+oo)I.F(t)} - A(t) a.s. (6.1')

Observe that if and are L' bounded increasing processes adapted


to they generate the same supermartingale potential up to a standard
modification if and only if for each parameter value t,

E{A(+oo) - A(t)l..(t)} = E{B(+oo) - B(t)l,F(t)} a.s.,

that is, if and only if the difference process is a martingale.


In particular, suppose that I is either Z+ or R+ and that f A(-), is an
adapted L' bounded increasing process generating a supermartingale
If I = R+, it is supposed that is right continuous and that is almost
surely right continuous; so we can suppose that is also almost surely
right continuous. Under these hypotheses if x(+ oo) is defined as 0, equality
(6.1') can be strengthened.
In fact, if T (S + oo) is an arbitrary optional time,

x(T) = E{A(+oo)I.F(t)} - A(T) a.s. (6.1")

in view of Section 111.2, Example (a), when I = Z+ and Theorem 2 when


1= R+.

The Support of the Measure Defined by an Increasing Process

If (A(-), R'(-)) is an increasing adapted process with I either Z+ or 18+ and


under the hypotheses of the end of the preceding paragraph, define the
(closed if I = R+) support of the measure A(dt), more precisely the support
of the measure d,A(t, co), by
482 2 IV. Basic Properties of Continuous Parameter Supermartingalcs

1= 71+ : A = { (n, w): n > O, A(n, w) - A(n -I, w) > 0},


1=R+:A= (1 I w!>0},
`n n J
where we set A(t) = 0 when t < 0. In the continuous parameter case, the
nth set on the right is nearly progressively measurable relative to the filtration
.F( + 1/n) so A is nearly progressively measurable relative to this filtration
for all n, and therefore (Section 1.2) A is nearly progressively measurable
relative to -fl-).

Natural Increasing Processes with I = 7L+ or R+

We now suppose that 0is a filtered probability space


P;
with I either 7L+ or R, that .F(0) contains the null sets, and if I = lR, that
.fl-) is right continuous. Observe that if ! = R+ and if A(-) is an increasing
almost surely right continuous process, then almost all its sample functions
are monotone increasing. In this context if { f(t), t E R+} is a stochastic
process with state space R and if H c: R+, the integral f H f(t) dA(t) is defined
as the integral f x f (t, (o) drA(t, w) at each point w for which co) is mono-
tone increasing and fl-, w) is integrable in the usual sense over H with
respect to the measure A(dt, w). When I is either 7L+ or R+, the problem of
finding a unique increasing process generating a specified supermar-
tingale potential leads to a condition on which we shall now formulate.
Let be a bounded martingale, almost surely right continuous
if 1= R+, and define y(+ oo) = lim,-,D y(t); this limit exists almost surely.
An adapted increasing L' bounded process almost surely right
continuous if ! = R+, is called natural if

I= 7 l + : E 1) - y(k)] [A (k + 1) - A(k)]} = 0
k-0
(6.2)

/ = R+ : I [Y(t) - Y(t - )] dA(t)1= 0


E o )

for every choice of It will be shown in Section 7 that this somewhat


awkward condition on is satisfied if and only if is nearly predictable.
In the following examples I= R, and is adapted, L' bounded,
almost surely right continuous.

EXAMPLE (a). The process A is natural if almost surely continuous because


(6.2,1= R+) is then trivial because almost every sample function in
this equation has at most countably many discontinuities.

EXAMPLE (b). If T is a predictable optional time and if AT(t) = llT,+,01(0,


then AT (.) is natural. In fact condition (6.2, 1 = R+) for AT reduces to
7. Natural versus Predictable Increasing Processes (I = Z+ or tt+) 483

E{ [y(T) - y(T-)] Itr<+ml} = 0, (6.3)

and if T is an increasing sequence of optional times announcing T, the


martingale equality E{ y(T A n) - y(TT A n)} = O leads to (6.3) when n - oo.

7. Natural versus Predictable Increasing Processes (I = 71+ or


0V)
In this section (0, P; ,f(t), t e I) is the reference space, a filtered prob-
ability space with I either Z+ or ff8+, .F(O) contains the null sets, and, if
I = R+, is right continuous. In the following theorem it is supposed
without explicit mention that the martingales and increasing processes
involved are almost surely right continuous when I= 02+. If is an
increasing process A(+ oo) is defined as lim,..,, A(t). If z is an integrable
random variable the martingale can be chosen to be right con-
tinuous [by Section 1], and we define E{zl.5( -)} as the left limit process
of this martingale. This left limit process is an almost surely left continuous
martingale and is therefore nearly predictable.

'theorem. Let be an adapted L'-bounded increasing process.


(a]) If { is a bounded martingale and if y(+ oo) is defined as
lim,_ y(t), then

1= 7L+ : E{ y(k + 1)[A(k + 1) - A(k)]}


k-0
= E{y(+oo)A(+oo)}, (7.1)

1= R : E{ fW y(t)dA(t)} = E{y(+oo)A(+oo)}.
0

(a2) The process is natural if and only if whenever a is a constant


in I and z is a bounded f(a) measurable random variable

I = 1+ : E{ y E{zIF(k + 1)} [A(k + 1) - A(k)]}


k-0
= E{zA(a + 1)), (7.2)

1= R+ : Elf E{zl,,F(t-)} dA(t)} = E{zA(a)}.


o.a,

(a3) The process is natural if and only if it is nearly predictable.


Let and (B(-), F(-)) be adapted L' bounded increasing pro-
cesses for which (A(-) - .F(-)) is a martingale.
484 2.IV. Basic Properties of Continuous Parameter Supermartingales

(b 1) Iffy(-), () } is a bounded predictable process then

1=l': E{> y(k+l)[A(k+1)-A(k)]}


k-0

= EI Y y(k + 1)[B(k + 1) - B(k)]} (7.3)

1= R: E{ f Y(t)dA(t)} = E{Jy(t)dB(t)}.
(o
(b2) If and are predictable then these processes are indis-
tinguishable.

Proof of (al) when 1= 7L+. Since is a martingale in (al),

E{[y(k + 1) - y(k)]A(k)} = E{A(k)E{ y(k + 1) - y(k)j.flk)} } = 0;

so

E y(k + 1)[A(k + 1) - A(k)]1


k-o )

= E{ [y(k + 1)A(k + 1) - y(k)A(k)] i

= E{y(+oo)A(+oo)}. 0

Proof of (a2) when I = r+. See the exactly parallel proof below for the case
!=. o
Proof of (a3) when I = V. If is predictable apply (7.1,1= V+) to both
and the increasing process {A(k + 1) - A(1); ,F(k), k e Z+ } to derive
(6.2, 1 = V+). Conversely, suppose that (A(-), .F(-)) is an adapted L'
bounded increasing process and that (6.2,1= 7L+) is satisfied for every
choice of To show that then is nearly predictable, it is sufficient
to observe that A(0) is -,F(0) measurable and to prove that A(1) is also .F(0)
measurable, because this result can then be applied to the increasing process
{ A(k + I) - A (l), .flk + 1), k e g+ } which satisfies the counterpart of (6.2,
1 = V+) for the filtration {,F(k + 1),keg+} to show that A(2) is .mi(l)
measurable, and so on. To prove that A (1) is 9(0) measurable, let y be a
bounded F(1) measurable random variable and define as the martingale
E{ yj.4r(0)}, y, y, ... Then (6.2,1= V+) yields

E{[y - E{yI.F(0)}]A(1)} = 0. (7.4)

Hence
7. Natural versus Predictable Increasing Processes (I = Z' or R') 485

E{yA(1)} = E(E{yj.F(0)}A(1)} = E{E{E{yj.F(0)}A(1)IF(0)}}


= E{yE{A(1)I.F(0)}}.

Since y can be the indicator function of an arbitrary .f(1) set, it follows that

A(l) = E{A(1)J.fll)} = E{A(1)J.F(0)} a.s.;

so A(1) is F(0) measurable, as was to be proved.

Proof of (a l) for I = R'. If { y(t), t e R+ } is a bounded almost surely right


continuous martingale and if is the process defined by

y^(0) = y(0), y^(t) _ Y y((k + 1)2_ ) ljk2-^.Ik+t)2-^s(t) if t > 0,


k=0

then (7.1,1= Z +) as applied to the martingale { y(k2-°), .F(k2-^), k e Z')


and the increasing process {A(k2-"), k e Z+ } yields

E{ y((k + 1)2 ")[A((k + 1)2-") - A(k2-^)]j = EI f m y^(t)dA(1)}


)) o

= E{ y(+ oo) A(+ oo)}.


(7.5)

Since lim^.. y^(t) = y(t) almost surely, simultaneously for all t, (7.5) yields
(7.1,1= R+) by the Lebesgue dominated convergence theorem. o

Proof of (a2) for I= R+. If is natural apply both (6.2,1= R+) and
(7.1,1= R+) with y(t) = E(zl,flt)} to derive

E{ f 0E{zj.F(t-)}dA(t)}=E{ f 'E{zl.F(t)}dA(t)j=E{xA(+oo)}
0" )

and (7.2,1 = R+) now follows from the fact that E{zl#(t-)} = z almost
surely when t > a. Conversely, if (7.2,1= R+) is true and if is a bounded
almost surely right continuous martingale, apply (7.2,1= R+) with z = y(a)
and then let a - + cc to derive the condition (6.2,1= R+) that be
natural.

Proof of (a3) for I = R +. According to Section 6, if either A(-) is almost surely


continuous or if there is a predictable optional time T such that A(t) =
cltr,+ml(t), then is natural. Thus to show that is natural if nearly
predictable, it is sufficient to show that (dropping evanescent sets) if is
predictable, right continuous with left limits, and if all its sample functions
486 2 IV. Basic Properties of Continuous Parameter Supermartingales

are increasing, then can be written as a countable sum of processes of


these two types. Choose fl > 0 and define To = 0 and

T",1=inf{t>T":A(t)-A(t-)2! P), n>0.


The process A(- -) is predictable by left continuity; so the process A(-) -
A( -) is predictable, and therefore the set {(t,w): A(t,w) - A(1 -,co) z fi}
Is predictable. For n > I the nth entry time of this set is T". Hence (Section
11.9) T. is predictable; so (Section 11.7) the process AT"(-) = lrT".+, is
predictable. The process

Y
"=o

is a predictable increasing process for which the sample function discon-


tinuities of magnitude z fl are decreased by ft. First choose jJ = 1, thereby
decreasing jumps >- I by 1, then apply the same procedure, but with fi = #,
to the above difference process, then with fl = I to the new difference process,
and so on, to obtain an evaluation of as a sum of the desired type.
Conversely, suppose that is natural. We show first that does not
charge any totally inaccessible optional time T, that is, A(T) = A(T-)
almost everywhere where T < + oo. To see this, define
]k2-",(k + I)2-"], and observe that
E{A(T) - A(T-); T < +oo}

= lim [A((k + 1)2-") - A(k2-")] dP;


"~ k=0 ITs
(7.6)

so it is sufficient to show that the right side vanishes. The sum in (7.6) is
equal to

E{ IITS(k+1)2-"1[A((k + 1)2-") - A(k2-")]


k=0
IITSk2-";[A((k + 1)2-") - A(k2-")]},
-
which by (7.2, / = R) is equal to

k EJ P{ T< (k + 1)2 dA(t)J-k EI f


!"5 J
(7.7)

If we define the interval I"k and the processesand as in Section 2, the


first sum in (7.7) can be written in the form E{f'o a"(t-)dA(t)}, and this
7 Natural versus Predictable Increasing Processes (I = Z' or R') 487

expectation, by the uniform convergence of the sequence proved in


Section 2. has limit (n oo)

E{ f, a(t-)dA(1)} = E{ I Ilr.+xt(I)dA(t)} (7.8)


J Jo 1

If now T. is defined by

(k + 1)2-" when Te /,;k,


T" +00 when T = + oo,

then T is a decreasing sequence of optional times with limit T and T. > T


almost everywhere where Tis finite valued and not equal to a dyadic rational
number. Since the distribution function on R of a totally inaccessible
optional time is continuous. T. > T almost surely where T is finite. The
second sum in (7.7) is equal to
'
E! If llr +a,t(r)dA(r)},

which has limit (n - x,) the right side of (7.8); so does not charge T. as
asserted.

In the proof of (a3) it was shown that if nearly predictable can be


expressed as a countable sum of increasing processes of which one is almost
surely continuous and the others have the form with T and
predictable. That reasoning in the present context, in which is to be
proved nearly predictable and in which we have just proved that charges
no totally inaccessible optional time, leads to the same expression for
except that as first defined each optional time Tis accessible. In fact we need
only point out that the discussion in the proof of (a3) leads to optional times
T charged by and therefore without totally inaccessible components.
Since is nearly predictable if T is accessible (Section 11.12), is
nearly predictable if natural.

Proof (bl) If and / = /', then

E{ '(k + 1)[x(k + 1) - x(k)] } =

E{ i'(k + 1)E{x(k + 1) - x(k)I,flk) } } = 0

because v(k + 1) is .F(k) measurable and {x(-),.fl-)) is a martingale. Hence


(b I) is true when I = l' . If I = R' and if { is a bounded left
continuous process, define
488 2 IV Basic Properties of Continuous Parameter Supermartingales

1,(0) =1'(0), y"(t) = E y(k2-")t]k2-".(k+111-n1(t) if t > 0,


k=0

so that the process { v"(k2-"), ,F(k2-"), k e 71 + } is predictable and therefore


by what we have just proved

E{ If zl;,(t)dA(t)}=E{J"y"(t)dB(t)}. (7.9)
0 o

When n -' oc this equation yields (7.3,1= R+). Thus the latter equation is
true if is bounded and left continuous, and since the class of bounded
predictable processes on O+ is the smallest class of bounded processes closed
under bounded monotone convergence and containing the left continuous
adapted processes, (b1) is true.
(b2) We give the proof for 1= OB+; the proof for I = 7L+ is similar but
less deep. Suppose first that A(+ oo) and B(+ oc) are bounded, and combine
(7.1,1= ff8+) with (7.3,1= IB+) to find that E{ y(+ oc) [A(+ oo) - B(+ oo)] }
= 0 whenever (y(-), .F(-) I is a bounded almost surely right continuous pre-
dictable martingale. If now we choose we find that
A(+ or,) = B(+oc) almost surely. Hence (martingale equality) for each
value of t it follows that A(t) = B(t) almost surely; so by the almost sure
right continuity of and these processes must be indistinguishable.
In the general case we need only find an increasing sequence S, of optional
times for which lim"..,,S. = +oc almost surely and for which A (S.) and
B(S") are almost surely bounded for each n. In fact, if the result for bounded
processes is applied to the pair A(S" A ), B(S" A ), it follows that these
processes are indistinguishable so and are. To find a sequence S.,
observe that for n > 0 the entry time T. of the predictable set

{(t,w): A(t,w) + B(t,w) >- n}

is predictable (Theorem 11.9), and if T", is a sequence of optional times


announcing T", then A(T.) + B(T",) < n; so we can choose

S"=T1.V... VT"".

8. Generation of Supermartingale Potentials by Increasing


Processes in the Discrete Parameter Case
Theorem. Every supermartingale potential {x(n), .f(n), n e71 + } is a quasi-
bounded class D supermartingale and is generated by a unique up to standard
modification, that is, unique up to indistinguishability, increasing predictable
L' bounded process.
9. An Inequality for Predictable Increasing Processes 489

In view of Section 6 it is necessary only to show that there is a unique


generating process as described. Given a supermartingale potential {x(n),
.fi(n), n e 71 + } define an increasing predictable process by

A(O) = 0, A(n) _ Y [x(m) - E{x(m + l )j. '(m)}] if n > 0, (8.1)


M=O

and define A(+co) = sup, A(n). Then E{A(+oo)} = E{x(0)} and

E{A(+oo) - A(n)Ifln)} = Y E{[x(m) - E(x(m + 1)I9(m)}]I f(n)}


mn
= x(n)a.s.; (8.2)

so is generated by Moreover, if is any super-


martingale potential generated by an increasing predictable process
it is immediate that

x(m) - E{x(m + 1)I.9(m)} = A(m + 1) - A(m) a.s. (8.3)

from which (8.1) follows as an almost sure equality; so is uniquely


determined up to a standard modification. [Uniqueness here also follows
from Theorem 7(b).]

Decomposition of a Submartingale

If {A(n), 9(n), n e 71 + } is a predictable increasing process and if { y(n), .fi(n),


ne71+} is a martingale, the process is a submartingale.
Conversely, if {x(n),9(n),ne71+} is a submartingale and if A (n) is defined
by (8.1) with replaced by then is a predictable in-
creasing process and where is a martingale.
It is left to the reader to check that is Lt bounded if and only if and
are. This decomposition of a submartingale makes it possible to deduce
the almost sure convergence of an L' bounded discrete parameter submar-
tingale from the almost sure convergence of an L' bounded discrete parame-
ter martingale.

9. An Inequality for Predictable Increasing Processes


Lemma. Let {x(n), .F(n), n e 71 + } be a supermartingale potential generated by
the predictable increasing process A (n).
Let c be a strictly positive constant, and define the optional time Tc by

+oo if A(+oo)Sc
T`-{inf{j:A(j+1)>c} ifA(+oo)>c.
490 2.1V. Basic Properties of Continuous Parameter Supermartingales

Then

A(+co)dP53 fA(+o)>c/2) x(Tn)dP. (9.1)


J A(+m)>c)
By definition of T, A(T,) 5 c and {A(+oo) > c) _ {T, < +co} eF(TT)
so that

cP{A(+oo) > c} 5 2 I [A(+oo) - A(TcJ2)]dP


(9.2)
52I x(7 12) dP,

and this inequality combines with the supermartingale inequality to yield

f A(+oo)dP5 I x(Tc)dP+cP{A(+oo)>c}
A(+ao)>c) (A(+ao)>c)

,3f x(T,,2) dP,


A(+ao)>c/2)

as was to be proved.

10. Generation of Supermartingale Potentials by Increasing


Processes for Arbitrary Parameter Sets
Let (0, .F, P; .fi(t), t e I) be a filtered probability space. It is supposed that
the linearly ordered parameter set I has a first but not a last point.

Theorem. The following conditions on a supermartingale potential {x( ),


are equivalent:
(a) The process is generated by an increasing L' bounded adapted process
for each tin I,

x(t) = E{A(+co) - A(t)I9r(t)}


(10.1)
= E{A(+oo)jF(t)} - A(t) a.s.

(b) The process is in the class D.


(c) The process is quasi-bounded.

Observe that when I= Z+, this theorem does not contain Theorem 8
because in that theorem the increasing process is predictable; in fact this
10. Generation of Supermartingale Potentials by Increasing Processes 491

predictability determines the increasing process in Theorem 8 up to a stan-


dard modification. Uniqueness cannot be asserted in the present theorem.
Moreover, according to Theorem 8, a supermartingale potential for the
parameter set Z+ is necessarily quasi-bounded and in the class D, whereas
there is an example in Section IX.6 below for the parameter set I8+ of a
continuous supermartingale potential which has neither property.
In the following proof we denote'(` I .fi(t) by F(+ co).

Proof. (a) (b) and (a) (c) See Section 6 where these implications are
derived without the adaptedness hypothesis on
(b) (a) Assume first that I is a bounded subset of G2+ and includes 0.
The random variable x(t) can be identified with a point of Lt (Q, .F(+ oo), P).
This function has a left [right] limit at each point of the closure !of I which
is a limit point of I from the left [right] according to the supermartingale
convergence theorems and the condition that the given process is in D,
so is uniformly integrable. According to an elementary theorem on functions
from a subset of R into a metric space, the set Io of points t of !for which
two of the members x(t -), x(t), x(t+) of Lt (12, 9(+ co), P) exist but are not
the same is countable. Let J be a countable subset of I, dense in 1, including
0, !o, and every point of ! which is not a bilateral limit point of 1, and let J
be a finite subset of J, including 0, with Jo c Jt c , Uo J. = J. The
restriction of to J. is a positive supermartingale to which Theorem 8 as
trivialized for finite parameter sets can be applied. (Theorem 8 is stated for
the parameter set Z+ but can be applied to a finite parameter set, say 0, . . . , k,
by defining the process random variables to vanish identically at parameter
values strictly greater than k.) According to Theorem 8, there is a process
{A (t), t e for which A (0) = 0, A&) is increasing and is predictable, and
if we define co) as the value of A.(-) at the maximum parameter value
in J,,,

x(t) = Aa(t) a.s., teJ,,. (10.2)

Moreover for c >r( 0,

2f 5 2E{x(0)}, (10.3)

and according to` Lemma 9, there is an optional time T (depending on n


and c) with values in J,,, for which

3 x(T)dP. (10.4)
f1An(+oo)>e) fA,(+oo)>c/2j

In view of (10.3), limr.,m P{A (+ co) > c/2} = 0 uniformly as n varies; so by


492 2.IV. Basic Properties of Continuous Parameter Supermartingales

the class D property of x(-) the right side of (10.4) has limit 0 when c - oo,
uniformly as n varies, and it follows that the sequence A,(+ oo) is uniformly
integrable. Hence some subsequence A.(+oo) converges weakly in L'(f ,
.F(+oo),P) to a limit random variable A(+oo), and consequently for each
parameter value t in I the sequence E{A, (+oo)19(t)} converges weakly in
L'(1,.4z(t),P) to E{A(+oo)I.F(t)}. From (10.2) if teJ, the sequence A,(t)
also converges weakly in L' (i2, .fi(t), P) to a positive F(t) measurable
random variable A(t), with A(0) = 0. Furthermore (10.1) is true for t in J,
and A(-) is an increasing process on J. By definition of J each point t in
I - J is a bilateral limit point of J, and x(t) = x(t+) = x(t-) almost surely
if the one-sided limits are defined either as L' limits or as sequential almost
sure limits. Hence for t in I - J,

x(t) = E((A(+oo)j.f(t+)} - A(t+)


(10.5)
=E{A(+oo).F(s)A(t-) as.
l $<f I-
On taking expectations here it follows that A (I +) = A(t-) almost surely
when t is in I - J; so if A(t) is defined when t is in I - J as supJ9,<, A(s),
the process {A(t). F(t), to 1) is an adapted increasing process, and for t in I

x(t) = E{A(+oo)jJv(t+)} - A(t) as.


Taking conditional expectations with respect to.F(t) yields (10.1). Moreover,
when t tends to the supremum of I along J, we find that

0 = E{A(+oo)I.01"(+co)} - limA(t) = A(+oo) - ess supA(t) a.s.


tt $@I

(b) (a) (in the general case) According to Section 111.4, there is an
order-preserving map from I into B such that 4)(s) = (¢(t) if and
only if x(s) = x(t) almost surely. We can choose 0 to be bounded and
to take the initial point of I into 0. Define 1= 4)(I). If ieI, define .#(i) =
Y.0-y) .fi(t), and define i(t) as x(t) for some t in 0`(t); the choice of t
does not matter. Ifs and i are in I ((and if s < i, then ))when ¢(t)

E{i(i)j,#(s)} = E{x(t)I Y 3'(s)} a.s.


l( Se '(3) ))

Here s:5 t; so by the conditional expectation continuity theorem the right


hand side is almost surely at most fr(s). Thus the process {z(i),. (i), i c- I } is
a positive supermartingale. It follows from the special case of the assertion
(b) (a) already proved that there is an increasing adapted process {A(i),
F(i ), ieI } for which, defining 4(+ oo) as in Section 6,

z(I) = E{A(+oo)IA(i)} - A(i) a.s., ieI.


1 1. Meyer Decomposition of a Supermartingale Potential 493

Define A(+oo) = A(+oo), A'(t) = A(i) for tee-'(i). Then if te4-'(i),

x(1) = E{A(+oo)I,*(i)} - A'(t) a.s.

Apply E{-IF(t)} to both sides to obtain

x(t) = E{A(+oo)IF(t)} - E{A'(t)IF(t)} a.s.

If A(t) = E{A'(t)I,F(t)}, the increasing adapted process {A(t),F(t),tel}


has the desired properties.
(c) (a) Under (c), Eo with { a positive super-
martingale which is bounded and therefore in the class D. Hence is
generated by some increasing adapted process fl-)); so Eo
generates that is, (a) is true.

11. Generation of Supermartingale Potentials by Increasing


Processes in the Continuous Parameter Case: The Meyer
Decomposition
Let (f2, F, P; F(t), t e R+) be a filtered probability space for which fl-) is
right continuous and F(O) contains the null sets.

Theorem. The following conditions on the almost surely right continuous


supermartingale potential F(-)) are equivalent:
(a) The process is generated by an increasing adapted L' bounded process
teR+ and A(+oo) = sup,ER. A(t),

x(t) = E{A(+oo)IF(t)} - A(t) a.s.

(b) The process is in the class D.


(c) The process is quasi-bounded.
If these conditions are satisfied the process can be chosen to be
almost surely right continuous and predictable, and is uniquely determined up
to indistinguishability by these conditions.

If and are increasing processes adapted to and generate the


same supermartingale, then the process is a martingale
(Section 6). Thus in view of Sections 6 and 7 to prove the theorem, all we
need prove is that in the present context there is an almost surely right
continuous predictable L' bounded increasing process that generates
Define F(+ oo) = Y,ER F(t), and define F(t) = F(O) for t < 0. Let J be
the set of dyadic points of R+, and let J. be the set {m2-",meZ' }. Then
according to Theorem 8, there is a process F(t), t e for which A ()
494 2.1V Basic Properties of Continuous Parameter Supermartingales

is increasing, L' bounded, and predictable, and (10.2) is true with A"(+ oo) =
sup,E,"A"(t). It follows as in the proof of Theorem 10 that the sequence
is uniformly integrable and that therefore some subsequence
A,,(+ oc) converges weakly in L' (i2, F(+ oc), P) to a limit random variable
A(+oo). When teJ, the sequence E{A,_(+oo)I.f(t)} therefore converges
weakly in L'(0,. (t), P) to E{A(+oo)I.F(t)}, and so (10.2) implies that the
sequence A, (t) converges weakly in P) to an .F(t) measurable
random variable A(t). In particular, the sequence A, (0) converges trivially
to A (0) = 0. Equation (11.1) is true for t e J, and (t T + oo)

0 = E{A(+oo)I.F(+oo)} - supA(t) = A(+oc) - supA(t) as. (11.2)


IE.t tEJ

Observe that in view of the right continuity of and the conditional


expectation continuity theorem the process as we have defined it on J
is almost surely right continuous. If now we apply Theorem 2 to choose a
version of E{A(+oo)IF(t)} for each t to obtain an almost surely right
continuous process and define A(t) for t e I8+ - J by (11.1), the process
is an almost surely right continuous increasing L' bounded process generat-
ing Finally we prove that is a natural increasing process.
According to Theorem 7, it will follow that {A(), is nearly predict-
able; this process can be trivially adjusted to be predictable if desired. Let
be a bounded almost surely right continuous martingale, and
define y(+ oo) = lim,y(t). Then is a natural increasing
process because

y(s-)dA(s)}
EI If.
= lim Y E{ y(k2-") [A((k + 1)2-m) - A(k2-n)])
k=0

= lim Y E{y(k2-")E{A((k + 1)2-") - A(k2-")I.F(k2-")}}


"-0p k=0

= lim Y E{y(k2-")E{x(k2-n) - x((k + 1)2-")IF(k2-")}}


"_0° k=0 (11.3)

= lim Y E{y(k2-")E{A"((k + 1)2-") - A"(k2-")I..(k2-")}}


-co k=0

= lim Y E{y(k2-")[A"((k + 1)2-") - A"(k2-4)]}


n-. k=0
= lim E{y(+cc)A"(+oo)} = E{y(+co)A(+oo)},

where we have used the discrete parameter case of Theorem 7 to derive the
last line, and in the last line n oo along the sequence Incidentally this
proof shows that there is convergence in (11.3) when n oc unrestrictedly.
12. Meyer Decomposition of a Submartingale 495

It is easily deduced that the original sequence oo) converges weakly to


A(+ oo), but we shall not need this fact.

Uniqueness Observation

If the supermartingale Y'(-)) satisfies the conditions of Theorem II and


if there are an almost surely right continuous martingale {x'(-), fl-)) and
an almost surely right continuous predictable increasing L' bounded process
for which then is the unique up to
indistinguishability generator of In fact the process
is a martingale; so (by Theorem 7) the processes and are
indistinguishable.

Predictable Supermartingale Potentials

If the supermartingale (x(-), F(-)) in Theorem 1I is predictable then the


martingale component of the Meyer decomposition must also be predictable
because the increasing component {A is. We shall show (Section 23)
that an almost surely right continuous predictable martingale is almost surely
continuous, and it follows that the supermartingale sample function jumps
are balanced out by the increasing process sample function jumps; that is,
for P almost every w,

x(t -, w) - x(t, w) = A(t, w) -A (t-,w)


simultaneously for all t. Trivially more generally this assertion is true if
is indistinguishable from a predictable process. In particular if
the given supermartingale potential is almost surely continuous it follows
that the increasing process component A is also almost surely continuous.

Meyer Decomposition for an Arbitrary Parameter Interval [0, c[.

If c > 0, the adaptation of the Meyer decomposition theorem to the param-


eter interval [0, c[ is accomplished trivially by the map t' = (2c/n) arctan t
taking I2+ into [0, c[.

12. Meyer Decomposition of a Submartingale


Let (1, .., P; .fi(t), t e R') be a filtered probability space for which is
right continuous and F(0) contains the null sets. Let (A(-), 9(-)) be an
almost surely right continuous adapted increasing process. The restriction
of this process to a parameter interval [0, c[ is in the class D because if T
is optional and if T < c, then A(T) is majorized by the integrable random
variable A(c). Let be an almost surely right continuous mar-
496 2.1 V. Basic Properties of Continuous Parameter Supermartingales

tingale. The restriction of this process to [0, c[ is also in the class D according
to Section 111.6. The sum process { is a submartingale
whose restriction to any parameter interval [0, c[is in the class D. Conversely,
we shall now show that an almost surely right continuous submartingale
If whose restriction to each parameter interval [0, c[is in the class D
can be written as a sum { y(-) + with { y(-), -F(-)) an almost surely
right continuous martingale and an almost surely right contin-
uous predictable increasing process; moreover both components are unique-
ly determined up to indistinguishability. Let be the restriction of to
the parameter interval [0, c[. This submartingale is in the class D by hypoth-
esis, and the right closable martingale

{E{z(c)j.F(t)},.F(t),0 5 t < c}

is also in the class D (Section 111.6). The version of the conditional expecta-
tion in the process

{E{z(c)I.1F(t)} - zjt),.F(t), 0 5 t < c}

can be chosen to make the difference process a positive almost surely right
continuous class D supermartingale. The supermartingale potential compo-
nent of the Riesz decomposition of the difference process is then also in the
class D and so is generated by an almost surely right continuous predictable
increasing L' bounded process {Ajt),.f(t), 0 5 t < c}. That is, we can
write on the interval [0, c[, where is an
almost surely right continuous martingale on the parameter interval [0, c[.
In view of the uniqueness property of the Meyer decomposition we can
choose an almost surely right continuous predictable increasing process
5() } whose restriction to the parameter interval [0, c[ is indistin-
guishable from and therefore is a martingale on
the parameter interval R+.

EXAMPLE. If { B(t), fi(t), 1E I8+ } is an almost surely right continuous adapted


increasing process, this process is a submartingale so there is an almost surely
right continuous predictable increasing process {B'(t),.F(t),tn68+} such
that is a martingale; is determined uniquely up to
indistingui shabili ty.

13. Role of the Measure Associated with a Supermartingale;


The Supermartingale Domination Principle
Let (i2, 9, P; .fi(t), t e I8+) be a filtered probability space for which is
right continuous and 9(0) contains the null sets. Let be a class
D positive almost surely right continuous supermartingale with associated
13. The Supermartingale Domination Principle 497

(Meyer decomposition) nearly predictable almost surely right continuous


L' bounded increasing process so that is the sum of an almost
surely right continuous martingale and a class D almost surely right contin-
uous potential,

x(t) = E{x(+oo -)I.F(t)} + E{A(+oo)jF(t)} - A(t) a.s. (13.1)

In this context is the martingale theory counterpart of a positive super-


harmonic function on an open subset of R', and is also the martingale
theory counterpart of a positive superparabolic function on an open subset
of A". The counterpart of the Riesz measure associated with such a super-
harmonic or superparabolic function is the measure A(dr). As a trivial
illustration of this fact recall that a superharmonic [superparabolic] function
is harmonic [parabolic] off the closed support of the associated Riesz
measure. One probabilistic counterpart of this property is the following.
Let S and T be finite optional times with S 5 T, and suppose that A(T) =
A(S) almost surely. Then the process x(-) is a martingale between S and T;
more precisely, the process

{x((S+1) n T);.F((S+1) A T),tcR+}


is a martingale. In fact

x((S + t) A T) = E(x(+oo-)J.F((S + t) A T)}


+ E{A(+oo)jJ7((S+ t) n T)} - A(S) a.s.

Next consider the domination principle in classical and parabolic poten-


tial theory. First consider the following example. Let 12 be a singleton {w},
let JF be the a algebra consisting of f2 and 0, let F(t) = .F forte IB+, define
P(O) = I = I - P{Q}, let be the indicator function of the interval
[0, 1 [, and define yo(-, co) _- 0. Then and yo(-) are class D right continuous
supermartingale potentials. One version of the predictable increasing process
associated with is the indicator function of the interval [1, +oo[;
so yo(-) on the support {(1,w)} of the measure B(dt), although the
processes and yo(-) are not indistinguishable. Thus the obvious counter-
part of the classical domination principle (Theorem 1.V.10) is invalid. On
the other hand, the parabolic domination principle (Theorem I.XVIII.16)
suggests that not y(i) and yo(t) but y(t-) and yo(t-) should be compared
on the support of B(dt), and in fact yo(t-) < y(t-) on this support. The
statement of the supermartingale domination principle suggested by this
example and by Theorem I.XVIII.16 is the following, in which we return
to the context and notation of the beginning of this section.

Theorem. Let be a class D almost surely right continuous super-


martingale potential with associated measure A (di) of closed support A. Let
498 2.1 V. Basic Properties of Continuous Parameter Supermartingales

{) O, be a positive almost surely right continuous supermartingale.


Suppose that x(t -, w) < y(t -, (9) quasi everywhere on

{(t,(0)EA: A(t,(o) > A(t-, w)}

and that x(t, w) < y(t, (o) quasi everywhere on

I (t,(9)EA: A(t,w) = A(t-,w)}.

Then quasi everywhere on 118+ x 0; that is, 5


(We have ignored the evanescent set on which one of the processes
involved is not right continuous with left limits.)
Set x(+ co) = 0, y(+ oc) = y(+ oc - ). For t e l8 + let T(t) be the first entry
time >-t of A by Since A is progressively measurable, T(t) is optional.
Choose t a point of almost sure continuity of that is, a point of continuity
of the monotone function This choice of t excludes an at most
countable set, and in view of the almost sure right continuity of and
it is therefore sufficient to prove that x(t) < y(t) almost surely. Define,
for the chosen value t.

T, (T(t) on {A(T(t)) > A(T(t)-)j,


+ oe elsewhere,

T = I(T(t) on {A(T(t)) = A(T(t)-)},


+ x elsewhere.

Then T' and T" are optional, t < T' < + oo, t S T" < + co, and

{T'=T"=+x}={T(t)=+oc}.
(Null sets are ignored throughout.) The optional time T' can have no totally
inaccessible component because [see the proof of Theorem 7(a3)] does
not charge any totally inaccessible optional time. Hence T' is accessible.
Let S be a predictable time whose graph is a subset of that of T', and let S.
be a sequence of optional times announcing S. It is no restriction to suppose
that So >- t. The almost sure supermartingale inequality

y(t) > E{y(S A T")13 (t)}t

yields (n - x)

y(t) z E{y(S-)Ifs<+Mi + y(T")IIT +=EJ'r(t)}


(13.2)
a.s.
13. Role of the Measure Associated with a Supermartingale; The Supermartingale
Domination Principle 499

Now

E{x((T'-)IIs<+00)

=E { lim x(S,) lls<+aOII_F(t)


In-a0 I (13.3)

= E{lim E{A(+oo) - A(S,)IF(SS))

E{[E{A(+oo)I YF(Si} lls<+,pllg(t) } as.


= 0 - A(T'-)J 111

Each optional time S1 is Yu .A(SS) measurable; so S is also, and therefore


the random variable l Is<+,,, can be put inside the inner conditional expecta-
tion operation in the last line of (13.3). Hence

E{x(T'-)l1s<+mll_1F(t)} = E{[A(+oo) - A(T'-)]lls<+mllF(t)} a.s.


(13.4)
Now A(T-) = A(t) when T' < +oo; so

E{x(T'-)lis<+m1IF(t)} = E{[A(+oo) -

A similar but easier argument shows that

fi(t)} = E{[A(+oo) - a.s. (13.6)

Furthermore IS < + cc } c {T' < + oo }, and S can be chosen to make the


probability of the difference arbitrarily small. Using this fact, we can replace
S by T' in (13.2) and (13.5) so (13.2) yields

y(t) z E{[A(+oo) - I.F(t)} a.s. (13.7)

Since A(+oo) = A(t) on the set {T' = T" = +oo}, the right side of (13.7) is
almost surely x(t), and the proof is now complete.
Special Case: Almost Sure Continuity of The parabolic context
domination principle (Theorem 1.XVIII.16) is the exact translation of the
classical domination principle (Theorem 1.V.10) into the parabolic context
if the Riesz measure corresponding to the given superparabolic potential
vanishes on parabolic semipolar sets; that is, under this condition coparabol-
ic-fine limits are not involved. Corresponding to this fact, in the present
Theorem 13 if A (di) almost surely vanishes on semipolar subsets of Fl+ x i2,
that is, if
P{ JBdtA(t,co) = 0} = 1
500 2.1V Basic Properties of Continuous Parameter Supermartingales

whenever h is semipolar, then is almost surely continuous [apply


Section 6 on the discontinuity set of a supermartingale to -A(-)]; so the
inequality hypothesis of Theorem 13 reduces to the inequality S
quasi everywhere on A ; that is, left limits are not involved.

14. The Operators T, LM, and GM in the Continuous


Parameter Context
Let (0, .F, P; .F(t), t e R `) be a filtered probability space for which is
right continuous and ,F(0) contains the null sets.
The Operator T. Let be an almost surely right continuous
process, let S and T be optional, with SS T < + oo, and suppose that x(T)
is integrable. For example, according to Section 3, if the process is a super-
martingale, then x(T) is integrable if T is bounded, or if T is not bounded
and is right closable. Following Section 111. 18 in which, however, the
parameter set was an arbitrary linearly ordered set and the optional times
were countably valued, define

X(t) on {S > t} u {T < t}, (14.1)


tsrx(t) = {E{x(T)JF(t)) on {S:!9 t 5 T}.

Here the conditional expectations are to be chosen to make rSTx(-) almost


surely right continuous (Section 2). Obviously rSTx(T) = x(T) almost surely.
Suppose now that is a supermartingale. It follows from Theorem
3 that x(T) is integrable and that 5 (essential order); so by almost
sure right continuity this inequality is true quasi everywhere on R` x 0.
The argument in Section 111.18 translated into the present context shows
that -fl')) is a supermartingale, a martingale on QS, T]J, in the
sense that the process

{tSrx((S V t) A T),.F((S v t) A T), to R' }

is a martingale. The dual results for a submartingale are clear.


Observe that if is a positive supermartingale and if we define
x(+ oo) = 0, then we need not restrict S and T to be finite valued in the
preceding discussion.

The Operators LM and GM

It will be sufficient to discuss GM. Suppose that is an almost


surely right continuous supermartingale which has an essential order sub-
martingale minorant. Then (from Section III.20) one version of the mar-
15. Potential Theory on R' x ) 501

tingale is For each positive integer k the sequence


n -. .F(t), t < k} is, for n >_ k, a decreasing sequence of almost
surely right continuous martingales. An application of Theorem 4 shows
that one version of is an almost surely right continuous martingale,
and in the present continuous parameter context we shall accept only such
versions. More generally, if r is a class of almost surely right continuous
supermartingales adapted to for which GM r exists, we now show that
GMr has an almost surely right continuous version. Such a version is the
only one we shall accept; it can be modified trivially to yield a right contin-
uous version, if desired. To prove the assertion, observe first that we can
assume that r is directed downward, at the price of adjoining to r the finite
minima of its members; the enlargement of r does not change GMr. The
class {GM r} is a downward-directed class of almost surely right
continuous martingales with essential order infimum and limit GMr, and
according to Section 111.5, there is a decreasing sequence of these mar-
tingales whose limit is a version of GMr. An application of Theorem 4
shows that this version is almost surely right continuous.

15. Potential Theory on R' x 0


Let (0, .f, P; ff(t), t E R+) be a filtered probability space for which fl-) is
right continuous and F(0) contains the null sets. Many of the results we
have obtained suggest that one can construct a theory on tl + x S2 which is
formally parallel in many ways to parabolic potential theory and therefore
also to classical potential theory, in which almost surely right continuous
supermartingales play the role of superparabolic functions. The nomencla-
ture, for example, Fundamental Convergence Theorem, GM and r opera-
tors, has been chosen with this in mind. In later chapters when we compose
superparabolic functions with space-time Brownian motion to get super-
martingales, it will become clear that the relation between a suitably con-
structed potential theory on 08+ x fI and parabolic potential theory is more
than just a formal resemblance.
Probability theory and potential theory developed independently of each
other aside from interrelations between random walks and Brownian motion
with solutions of parabolic and elliptic differential equations, until on the
one hand axiomatics of generalized potential theory were devised and on the
other hand the theory of Markov stochastic processes led to a generalized
potential theory in which potential theoretic ideas were defined in terms of
probabilistic ones. In both theories generalized versions of superparabolic
functions lay at the base of the subject, and under suitable conditions it was
shown that the axiomatic potential theory could be generated by Markov
process theory. It then became possible to prove theorems in axiomatic
potential theory by probabilistic methods. We shall discuss in later chapters
the Brownian motion process which generates classical potential theory and
502 2.IV. Basic Properties of Continuous Parameter Supermartmgales

the space-time Brownian motion process which generates parabolic potential


theory.
Although probabilistic methods have frequently been used to derive
potential theoretic results, derivations in the reverse direction have been
rare. To illustrate the possibilities and to give insight into the whole subject,
we shall use the ideas and methods of potential theory to analyze reductions
in the context of potential theory on 98' x Q. We shall need a few properties
of the fine topology on this space.

16. The Fine Topology of R x i2


Let M. .F, P;.f(t), to R') be a filtered probability space for which is
right continuous and .r(0) contains the null sets. The fine topology of
R' x fi is by definition the smallest topology on R' x 0 making every
right continuous supermartingale into a (fine topology) contin-
uous process in the sense that the function (1, w) -' x(t, w) is to be continuous
in the fine topology. The set of fine limit points of a set A will be denoted by
A.
EXAMPLE (a). If A is a P null subset of 0 and if x(-, -) is the indicator function
of A = R' x A, the process is a right continuous martingale;
so A is both fine open and fine closed.

EXAMPLE (b). Let T be optional and define as the indicator function


of the stochastic interval 10, nJ. Then {x(-),.F(-)} is a right continuous
supermartingale; so both stochastic intervals TO, T[ and I T, + oo Q are fine
open and fine closed. If S and Tare optional with S < T, we conclude that
the stochastic interval IS, TQ = 10, 71 n IS, + oo Q is fine open and fine
closed. If we choose T. = T + I/n, we find that the stochastic interval
IS, TJ = f i, IS, T,[ is fine closed. In particular, the graph [Sj is fine closed.
If S is optional, if A c .F(S), and if S,, = S on A but S = + oo otherwise,
the stochastic interval QSA, TAI = (R' x A) n IS, TQ is fine open and fine
closed. Observe that the intersection of this stochastic interval with another
one (R' x A) n QS', T'[ of the same type is

(R' xAnA')nISvS',TAT'Q,
also of this type, and that S v S' is identically constant if S and S' are.

A Base for the Fine Topology. The class of sets (stochastic intervals)
(R' x A) n [a, TQ with a a positive constant and A c F(a) is a base for the
fine topology because if { is a right continuous process, not
necessarily a supermartingale, and if T, is the entry time of the set
16. The Fine Topology of R' x S2 503

{ (s, w) : s > t, x(s, (0) S c),

then

{(t, w): x(t,w) > c} = U. (R x {co: x(t,w) > c)) r) it, T,Q.
IER

If {wo} is a P null singleton and if toe 08', the set

(R+ x {wo}) n Qto, to + e[

is a fine neighborhood of (to, wo) for every e > 0, and the set of these fine
neighborhoods for all E > 0 is a base for the fine neighborhoods of (to, wo).
Hence a subset A of l8' x f2 has (to, wo) as a fine limit point if and only
if A contains points (t,,, wo) with t > to, t = to. Thus if all 0 single-
tons are P null the fine topology is the direct product of the discrete topology
on fl and the one sided Euclidean topology on l8+ in which a base for the
topology is the class of left closed intervals.

Evanescent Sets and the Fine Topology

The fine closure of an evanescent set is evanescent because an evanescent


set is a subset of an evanescent set of the form R+ x A with P{A} = 0 and
[from Example (a) above] this product set is fine closed.

Almost Thinness at an Optional Time

If T is optional, call a subset A of [I x i2 almost thin at T if the set [TD n Af


is evanescent. Then an evanescent set A is almost thin at every optional time.
A progressively measurable set A almost thin at every optional time T is
semipolar if all Sl singletons are null, according to the following argument.
If T, is the entry time of A, then IT,I c A up to an evanescent set, and the
hitting time T2 of A after T that is, the second entry time of A, is almost
surely strictly larger than T, on { T, < + oo }. Continuing in this way, we find
an increasing sequence T. of optional times such that T. < almost
everywhere on {T < +oo}, and c A up to an evanescent set. Going
on by transfinite induction, the next optional time is the first entry time
>_ T,,, .... If a is a countable ordinal for which T. has been defined,
either T,+, = + ac almost surely, in which case A is semipolar because
Ac IT up to an evanescent set, or P{T,+, > T,} > 0; so

E{arctan T,+,} > E{arctan T,}.

It follows that the second possibility can occur only countably many times,
504 2 I V Basic Properties of Continuous Parameter Supermartingales

that is, there is some countable ordinal a with T,+, = + x almost surely,
and we have just seen that then A must be semipolar.

17. Potential Theory Reductions in a Continuous Parameter


Probability Context
(See Section 111.22 for the corresponding discrete parameter theory.)
Let (S .. , P: f (t), t e 08 +) be a filtered probability space for which 9(-)
is right continuous and .y(0) contains the null sets. All processes below are
adapted processes on this space. Let be a positive process, and denote
by r the class of almost surely right continuous supermartingale pointwise
majorants of It is supposed that r is not empty. The class r' is directed
downward, and we define R_,., = ess inf r (= either an equivalence class
under identification of indistinguishable processes or a member of the class,
as indicated by the context). Observe that the equivalence class R,(.) is
unchanged if r is defined as the class of almost surely right continuous
supermartingales which are quasi everywhere majorants of or if "almost
surely" is omitted here. Define R_i.,(+x) = :(+ x) = 0. According to the
Fundamental Convergence Theorem (Theorem 5), almost every sample
function of the supermartingale R_,., has right and left limits at all points,
and if R") is defined as the right smoothing of R2(.), then R=V) < R,(., quasi
everywhere. For i c- R + define

r(t)=esssup{E{z(T)Jf(t)}: Toptional, >_ t); (17.1)

this random variable is determined only up to a null set. The method of


proof of Theorem 111.22 shows that is a positive supermartingale
which in the essential order is both a majorant of and a minorant of r.
Thus no matter how v(t) is chosen, R=,., > in the essential order;
that is, P; R_,.1(t) ? Y(i) > z(t) J = I for all t. According to Section 1 there is
a countable parameter set B such that r(t) can be chosen for each t in such
a way that almost every sample function has right and left limits at all
points, and is right continuous except possibly at points of B. Choose Y(1)
in this way. With this choice R_ quasi everywhere. If
Y, (t) is defined by the right side of (17.1) except that the optional time T is
supposed strictly greater than t, then r+ is a supermartingale, r+ (t) < r(t)
almost surely, and t' J t sequentially implies that

t'

(t) = vu +) and thereby have :5y(-) quasi every-


where. In particular, if is almost surely right lower semicontinuous.
Y, quasi everywhere; so and therefore
quasi everywhere. We conclude that and are indistinguishable
18. Reduction Properties 505

almost surely right continuous supermartingales. Unfortunately we shall


not be able to make this lower semicontinuity hypothesis in the following
application to reductions, in which the process corresponding to above
is the product where is now an almost surely right continuous
supermartingale and A is a subset of 08+ x Q.

Application to Reductions (on the Above Filtered Probability Space)

Let A be a subset of R+ x 0 for which t - IA (t) is a process adapted to


and let be a positive almost surely right continuous supermartingale.
Define the reduction R (.) as the supermartingale R=,.)IA,.). The right smooth-
ing of the reduction will be denoted by RA() or by According to
Theorem 5, the process RA.) is determined up to indistinguishability by the
following conditions:
The reduction is a progressively measurable process.
Whenever is in the class f of almost surely right continuous
positive supermartingales quasi everywhere on A, the inequality
z is true quasi everywhere on 08+ x fl.
If is a process satisfying the preceding conditions, then Z RA.

quasi everywhere on 68+ x Q.


Furthermore a reduced process is a supermartingale with almost sure
right and left limits, coinciding quasi everywhere with the respective right
and left limits of its right smoothing R=( and Rz 5 quasi
everywhere, with equality up to a semipolar set. Finally, a suitably chosen
decreasing sequence of members of r' has as limit a version of the reduction.
We shall see that many reduction properties in the present context are
close counterparts of reduction properties in the context of classical and
parabolic potential theory. In one respect the present context is simpler than
those earlier ones. The fact that superharmonic and parabolic functions can
have infinities complicates their theory, whereas almost every sample func-
tion of an almost surely right continuous supermartingale is finite valued
with finite left limits.

18. Reduction Properties


The context is that of Section 17. We consider reductions of a positive almost
surely right continuous supermartingale on subsets of 68+ x Q.
Reduction properties in this context are counterparts of reduction properties
in parabolic potential theory as listed in Sections 1.XVII.I I and 1.XVII.16
except that for simplicity we have not introduced counterparts of reductions
on boundary sets. The following properties are listed as far as possible in
the order of their parabolic counterparts, and their counterparts are always
506 21 V. Basic Properties of Continuous Parameter Supermartingales

referred to. Proofs are given or referred to in Section 19. We leave to the
reader the translation of the listed properties to the context of a countable
parameter set.
(a) [See Section 1.XVII.16(b).]

RA (t+) = Rz .,(t+), RA. (t-) = R A.,(t-) q.e.;

z(t) z Ri.)t) z R i.,(t+) q.e.;


Ra., = R= ., up to a semipolar set;

R" .) = Ra., q.e. if A differs from h by an evanescent set.

(b) [See Section I.XVII.16(b).]

q.e.on AnAf;
R".) = R" =z(-) q.e. on ,4f, in particular, q.e. on the fine interior of A;
R" = RA(.) q.e. when A is fine open.

(c) [See Section 1.XVII.16(b).] If S and T are optional times with


S 5 T and if IS, T[ r),4 is evanescent, then R; ., = R .) quasi everywhere on
QS, TQ, and Ra is an almost surely right continuous martingale on [S, T]
in the sense that the process

(R=.,((S v t) A T),.. ((S v t) A T),teUB+} (18.1)

is an almost surely right continuous martingale.


(d) [See Sections 1.XVII.11(a) and 1.XVII.16(c).]

R (.) = ess'inf (R8 ): A c , E progressively measurable and fine open} q.e.;


(18.2)

one version of the right-hand side is the infimum when a runs through some
decreasing sequence of fine-open supersets of A. In particular, if all almost
surely right continuous supermartingales on the given space are almost
surely lower semicontinuous, then whenever is almost surely continuous,
the set 1) in (18.2) can be further restricted to have open t cross sections for
fixed co.
(e) [See Sections 1.XVII.11(b) and 1.XVII.16(i).] If A c $ and if A is
fine open, then the supermartingales OB and are indistinguish-
able. If 1) is also fine open, these supermartingales are indistinguishable from
flz(i 0 P".
18. Reduction Properties 507

(f) [See Section 1.XVII.16(f ).] If A is indistinguishable from a progres-


sively measurable set, then R ) = R".) quasi everywhere on (R+ x S2) - A.
(g) [See Sections 1.XVII.11(d) and 1.XVII.16(k).] The set functions
A i R 1.1 and A -4 R' are strongly subadditive in the sense that quasi
everywhere on R+ x 0 both

R :8 + Ri :)B < R= + Ra., (18.3)

and the corresponding inequality (18.sm) for smoothed reductions are true.
The exceptional evanescent set depends on A, B. [See (i) for countable
strong subadditivity.]
(h) [See Sections 1.XVII.11(e) and 1.XVII.16(e).] If A, is an increasing
sequence of subsets of Q8+ x f2 with union A and if is an
increasing sequence of almost surely right continuous positive supermartin-
gales with limit then quasi everywhere on I8+ x 0 both

lim RTry.) = Rz .) (18.4)


n-M

and the corresponding equation (18.4sm) for smoothed reductions are true.
(i) [See Section 1.XVII.16(k).] The set functions A)- Ra.) and A) -+R )
are countably strongly subadditive in the sense that if A. and B are sequences
of subsets of R+ x 0 with An c Bn, then quasi everywhere on R+ x S both

RU-1" + R' An< RU'


0
RU'-"- +
0
8n
R:r) (18.5)

and the corresponding inequality (18.5sm) for smoothed reductions are true.
(j) [See Section 1.XVII.16(g).] If n e 7L+ } is a sequence of almost
surely right continuous positive supermartingales, if Eo (almost
surely right continuous by Theorem 4) is a supermartingale, that is, if
Eo E{zn(0) } < + oo, then quasi everywhere on 68+ x D both

R1r R./
(18.6)
0

and the corresponding equation (I 8.6sm) for smoothed reductions are true.
(k) [See Section 1.XVII.16(m).] If S is a finite optional time and if
A = [0, S], then GM R".) = 0 quasi everywhere (equivalently, R=) is a
supermartingale potential) if either S is bounded or S is unrestricted and
is uniformly integrable.
(1) [See Section 1.XVII.16(o).] Suppose that (S2, , has the
property that every almost surely right continuous supermartingale is almost
surely lower semicontinuous, and suppose that is an almost surely
continuous class D positive supermartingale. Then the set function 4F-* R",.,
508 2 IV. Basic Properties of Continuous Parameter Supermartingales

is a Choquet capacity on R' x Q relative to the class I', of progressively


measurable sets whose t cross sections for fixed co are compact, in the sense
that
AcB=t,-0:5R,.,5Re., q.e. (18.7')

A IA ---- R("iIR(., q.e. (18.7")

R A^>IR;., q.e. (18.7'")

(m) [See Section 1.XVII.16(p).] Suppose that (f2, has the


property that every almost surely right continuous supermartingale is almost
surely lower semicontinuous, and suppose that is a positive almost
surely continuous supermartingale. Let r, be the class of predictable product
sets defined in Section 1.14, and let r, be the class of finite unions of r' sets.
Then if A is analytic over the class of predictable sets, there is an increasing
sequence E. of r,, subsets of ,4 for which quasi everywhere on R' x Q both

limR=' =R'4 (18.8)

and the corresponding equation (I8.8.sm) for smoothed reductions are true.
(n) [See Section 1.XVII.16(q).] If and are bounded positive
almost surely right continuous supermartingales, then

sup' IRA., - R (.) 1 5 sup' I, (18.9)

where sup' means the supremum up to evanescent sets, and the corresponding
inequality (18.9sm) is true for smoothed reductions.
(o) [See Section 1.XVII.16(s).] Let {x( ), () } and { y( ), be
almost surely right continuous positive supermartingales, let a and b be
positive numbers with 0 5 a < b, and define

A = ((f, (9): x(t,w) 5 ay(t,w)}, B= {(t,w): x(t, w) z by(t,w)}.

Denote r. . . theiterated reduc-


tion and downcrossing inequalities 1I1(23.2)-111(23.6) are true in the present
context, as are also the other martingale theory translations of the classical
context reduction inequalities in Section 1.VI.3(o). Moreover these in-
equalities are true for the parameter restricted to an interval. In particular
if a, b] refers to downcrossings on the parameter interval
[0, t], then (almost surely) for t e R',

[b. (0)]
E{ 5 x(0)b (18.10)
19. Proofs of the Reduction Properties in Section 18 509

19. Proofs of the Reduction Properties in Section 18


(a) The properties listed in Section 18(a) were all derived in Section 17.
(b) The first two lines in Section 18(b) follow easily from two facts:
(i) There is a decreasing sequence of right continuous supermartingales
equal to on A and so also on Af, and tending quasi everywhere
to RA =).).
(ii) Each fine neighborhood of a point (to, (DO) contains a set of the form
[to, to + e] x (co.) with e > 0.

To prove the third line, observe that R".) = R .) quasi everywhere on A;


so R"
+=(-)
is a version of the unsmoothed reduction.
(c) To prove Section 18(c), observe (see Section 14) that if T is a finite-
valued optional time and if (x.(-), n e 1+ } is a decreasing sequence of almost
surely right continuous positive supermartingales with quasi everywhere
limit R A.), then Tsrx is an almost surely right continuous positive martin-
gale on IS, majorizes quasi everywhere on A, and
x.(-) quasi everywhere. Hence TSTR .) is indistinguishable
from R.) and is (by Theorem 4) an almost surely right continuous martingale
on QS, T] and so is equal quasi everywhere on [S, T to its right smoothing.
If Tis not finite valued, the same argument is applicable without any change
if we define R-4. )(+oo) = 0 and allow T to be infinite valued in
the definition of Tsr. The needed properties of the operator TST remain 'valid
under this extension.
(d) The first assertion of Section 18(d), in which R4. is indistinguishable
from RB.), follows from the fact that, assuming as we can in this proof that
is right continuous, if is a right continuous positive supermartingale
essential order majorant of on A and if c > 0, then e is an essen-
tial order majorant of on a progressively measurable fine-open neighbor-
hood of A. If every almost surely right continuous supermartingale on the
probability space is almost surely lower semicontinuous "fine open" in the
preceding argument can be replaced by "open t cross sections for fixed w."
The right-hand side of (18.2) is a countable infimum according to Theorem 5.
(e) See the proof of Section 18(e) in the parabolic context [Section
I.XVII.11(b)].
(f) To prove the equality of reduction and smoothed reduction quasi
everywhere on A = (R' x f2) - A, observe first that the strict inequality set
is not only semipolar but (see the proof of the Fundamental Convergence
Theorem) is, up to an evanescent set, a countable union Uo [S. T of optional
time graphs. It is to be shown that, for each value of n, the set A =
{w: [S.((9), w] eh) is null. This set, the projection on f2 of the set [S D r )A,
is in because, for c z 0, the set A n {S S c}, the projection
on f Z of [S ] n h n ([0, c] x fa), is in .F(c). Define an optional time S
by setting S, = S on A and S = + oo elsewhere. If T. is the time of
510 21V Basic Properties of Continuous Parameter Supermartingales

(r, (o) entry into the nearly progressively measurable set [S ] r).4, then
{ T. = S.' < + co } c Af up to a null set ; it then follows from (b) that
T. > S almost everywhere on A. Since the set [S,;, Tn[ n A is evanescent
it follows from (c) that R= Rz almost everywhere on A; this
equality implies that A is null.
(g) To prove strong subadditivity of the reduction operator and its
smoothing, suppose first that A and h are fine open, in which case (18.3sm)
reduces to (18.3), and define z'(-) = R= .) A RB.). Then [see the proof of the
corresponding potential theory equality I.VI(3.4) and 1.XVII(11.4)]

RZ iB + Ri t.e = R A.) + RB.) q.e. (19.1)

This equality implies inequality (18.3) for fine-open A and $ and in view
of (18.2) thereby implies (18.3) for all A and E. Inequality (18.3sm) follows
trivially from (18.3).
(h) The proof of (18.4) follows the proof of the parabolic potential
theory equation 1.XVII(11.6). If each set An is fine open, (I8.4sm) reduces
to (18.4) and follows from the fact that on the one hand (quasi everywhere)
the limit on the left side of (18.4) exists and is majorized by the right side
and on the other hand the limit process is almost surely right continuous
(Theorem 4) and (quasi everywhere) majorizes on A and so majorizes
the right side on I8+ x Q. If the sets involved are not fine open, let T be a
finite valued positive Y20JF'(r) measurable function, and let be the
indicator function of the set 5 k}. Then

E{R=j.,.)(T)xk(T)} S E{RB.,(T)xk(T)} < k.

Choose e > 0, at > 1, define

An = An n

and apply (18.2) to find a fine-open superset $; of A satisfying

E{R=()(T)xk(T)} + 2-'e. (19.2)

The sequence A° is an increasing sequence with union A, and we define


A = Uo hj A. By strong subadditivity of reductions

E{Rs Uo81(T)xk(T)} S E{R ")(T)xk(T)} +

-Ej R'4J(T)xt(T)} (19.3)


0 111

5 E(R")(T)xk(T)} + 2e.
19. Proofs of the Reduction Properties in Section 18 511

Since (18.4) is true when the sets involved are fine open, (19.3) yields (when
n-4oo)

E{R ".)(T)xk(T)} 5 E{R".)(T)xk(T)} S li m E{R ")(T)xk(T)) + 2e


R-W
(19.4)
5 a lint E{ R=.1.)(T)xk(T)} + 2e.
-m
Hence

E{R A.)(T)xk(T)} S lint (19.5)


n-W

and the reverse inequality is trivial. Since there is monotone convergence


of integrands in (19.5), we conclude that for every k

!im R=1.)(T)xk(T) = R=.)(flxk(T) as.


The set

Ck: (R'4.)(-, li m

has the property that its intersection with the graph of Thas a null projection
on 0, whatever the choice of T. It follows (from Theorem 11.8) that Ok is
evanescent and that therefore (18.4) is true. Then (18.4sm) is true up to a
semipolar set and since both sides of this equation are almost surely right
continuous processes the two sides must be indistinguishable.
(i) Inequality (18.5) and (18.5sm) follow from strong subadditivity
combined with (18.4) and (18.4sm).
(j) To prove (18.6) for two summands, it is sufficient to give the proof
for A fine open, in which case the proof of the classical. context counterpart
I . VI (4.2) translates easily into one valid in the present context. It then follows
from (18.2) that (18.6) is true for arbitrary A and finitely many summands
and therefore by (18.4) for countably many summands. Finally (18.6sm)
has now been shown to be true up to a semipolar set, and since both sides
of (18.6sm) are almost surely right continuous processes, the two sides must
be indistinguishable.
(k) If A = Q0, SD and if A = QS, + oo Q, it is trivial that R-4., = 0 quasi
everywhere on A. Hence the martingale GMR= vanishes quasi everywhere
if S is bounded. If S is not bounded

limE{R= (t)} = lira [(S> z(t)dP, (19.6)


. ri

and this limit is 0 if is uniformly integrable; so R'. is a supermartingale


potential.
512 2 IV Basic Properties of Continuous Parameter Supermartingales

(I) Assertion (18.7') is trivial, and (18.7") is included in Section 18(h).


To prove (18.7"'), suppose first that A = 0. Then if T. is the entry time
of A,,, the sequence T, is increasing. U" { T = + oo } = D almost surely.
and by Section 18(c) the process R= .) is an almost surely right continuous
process on f 0, T,11. Furthermore, if we define z(+ tx) = 0.

RV) <

quasi everywhere on this set. Hence the submartingale maximal inequality


applied to the martingale on the right, which can be supposed right con-
tinuous, yields
aP{sup R"")(t) a} < E{z(T );. (19.7)
r In
Since T. _ + -j-. for sufficiently large n, almost surely, the class D property
of implies that the expectation in (19.7) has limit 0 as n x . Hence

lim sup RA °)(t) = 0 a.s.,


n'-'STS
and therefore (18.7"') is true when A = 0. In the general case let h be a
progressively measurable superset of A with open cross sections for each
fixed w. Then by subadditivity of the reduction and by (18.7')
A, g

A a decreasing sequence to which the special case just treated


is applicable. We conclude that lim, R=^ S RB., quasi everywhere, and
Section 18.(1) follows, in view of Section 18(d).
(m) The class rp is closed under finite unions and intersections, and the
sets in rp have compact cross sections for fixed u,. Furthermore (Section I.14)
the class of analytic sets over rp includes the predictable sets. Suppose first
that is bounded. It follows from Section 18(1) that the set function
Ar--s RZ .,
is a Choquet capacity on R' x 0 relative to rp in the sense stated.
Hence, if T is an optional time and if we define R A.,(+oo) = 0, the set
function is a Choquet capacity on U8+ x Q relative to T'p.
Now choose A analytic over I'p and therefore capacitable. Choose an
increasing sequence F. of countable intersections of rp sets satisfying

A. c A. lim E{R".)(T)}. (19.8)

Define F= U' I. In view of the monotoneity of the integrand sequence


in (19.8) and in view of Section 18(h),

lim R'i)(T) = R=i.,(T) = R" .,(T) a.s. (19.9)


20. Evaluation of Reductions 513

An easy diagonal procedure shows that E, can be chosen to make (19.9)


valid almost surely simultaneously for an arbitrary prescribed countable set
of optional times T, and we choose ; for the following countable set of
optional times:

T is identically a positive rational number, for each such number.


T = S,,, where S. is chosen so that the semipolar set {R"
+:(
< R A.)} is a
subset of Uo up to an evanescent set.

With t so chosen the processes R= and R-4. are indistinguishable, since


they are almost surely right continuous and are equal almost surely at the
rational parameter values. Thus (18.8sm) is true. Moreover quasi everywhere
on R+ x S2,
F F A
RA) = R=( S R=(.) 5 R=(.),

and there is equality throughout except possibly on Uo QS I. Since t. was


defined to make R= .) = R= .) quasi everywhere on this union of graphs, we
have proved that R=., R-4.) quasi everywhere; that is, (18.8) is true. Without
the boundedness condition on that we have imposed in this proof (18.8)
and (18.8sm) are true with replaced by A n with n e l+, and the
desired equations then follow when n - oo, using Section 18(h).
(n) The proof of (18.9) is omitted because this proof is the exact counter-
part of that of the corresponding classical inequality I.VI(3.8).
(o) The proof is obtained by a translation into martingale theory of the
corresponding classical context reduction inequalities in Section I.VI.3(o).
The discrete parameter argument is in Section 111.23.

20. Evaluation of Reductions

Ex&Htr[.E (a). Let A be a fine-open (Section 16) stochastic interval QS, TQ


with S < T. Then (Section 18(b)) R1 = R I.) quasi everywhere. If is a
positive almost surely right continuous supermartingale majorizing on
A, then under the convention that all processes are 0 at the parameter value
+oo,

y(t) z E{y(S)j.F(t)} z E{z(S)15(t)} a.s. on {S > t}. (20.1)

The process defined by


514 2 W. Basic Properties of Continuous Parameter Supermartingales

E{z(S)j.9'(t)} if t < S
zo(t) = z(t) if S:5 t < T
0 if +oo>t>T
is a positive right continuous supermartingale if the conditional expectations
are defined suitably (Theorem 2), is equal to z(-) on A, and is a quasi every-
where minorant of v(-). Hence is a version of RA.).

EXAMPLE (b). If A is as in Example (a) and B = QS, T}, then any positive
right continuous supermartingale majorizing on h majorizes R z(.)

quasi everywhere and satisfies (20.1). Moreover for s > 0 the process
go is also a positive right continuous supermartingale majorizing
z(-) on B. It follows that quasi everywhere

E{z(S)j.f(t)} ift<S,
R".,(t) = z(t) if S<t<T,t < +oo,
0 if+oo>t> T,
RB =R-4 quasi everywhere, and RB < RB., on QTD, a semipolar set.

EXAMPLE (c). If 0 < a < h and if A e 9(0) when a = 0 but A E


when a > 0. then [a, b] x A = RS, Tl with S = a on A and S = + co other-
wise, T = h on A and T = + oo otherwise, so the reduction evaluation in
Example (b) is applicable.

Theorem. Suppose that (S2, . , j F(-), P) has the property that every almost
.surely right continuous supermartingale is almost surely lower semicontinuous,
and .suppose that is an almost .surely continuous positive supermartingale.
Let A he predictable , and let T,' [T,] he the entry [hitting] time ol'A n Q t, + cc
Then for each t > 0

RA.,(t) = E{z(T,')j.(t)} a.s. (20.2)

R" (t) = E{z(T,)j.f(t); a.s. (20.2sm)

It is sufficient to prove this theorem for bounded since (20.2) and


(20.2sm) are true for if true for A n when ne7 Z. Moreover the truth
of (20.2) implies that of (20.2sm) by the following argument. Since RA is
the smoothing of R= ,, there is a countable subset of O8+ such that for t not
in this set R" (t) = R".,(t) up to a null set which depends on t. Now the
process T, is the smoothing of T. T, = limsy, T,. so there is a countable
subset of 08' such that for t not in this set T, = T,' up to a null set which
depends on t. Thus, if (20.2) is true, it follows that (20.2sm) is also true for t
21. The Energy of a Supermartingale Potential 515

not in some countable set. Finally if s > 0 and if t decreases sequentially to s,


not hitting the exceptional countable set, (20.2) yields (20.2sm) for s, in view
of the almost sure right continuity of the left side and the dominated con-
vergence theorem for conditional expectations as applied on the right.
We proceed to prove (20.2) for bounded. Let A be a sequence of sets
for which (20.2) is true. To prove (20.2) for arbitrary predictable A, it is
sufficient to prove the following two assertions.
(i) If Ao c A, c .. , and U',4,, = A, then (20.2) is true.
(ii) If A o A, , and no A = A, and if each set A is progressively
measurable and has compact t cross sections for fixed w, then (20.2)
is true.
In fact, if F, is the class of finite unions of the sets of the type considered in
Example (c), it follows from (ii) that (20.2) is true for countable intersections
of ro sets, and therefore by (i) combined with Section 18(m) it follows that
(20.2) is true for every predictable set A. To prove (i) and (ii), let S be the
entry time of A n QT, + co Q. In (i) So, >- S1, >_ , and T,'.
On the one hand, lim, R") = R i.) quasi everywhere by Section 18(h).
On the other hand (dominated convergence theorem for conditional
expectations),

lim E{z(T,)I.f(t)} a.s. (20.3)


n-M

so (20.2) for A becomes (20.2) for A when n - oo. In (ii) So, < S < ,
and T,. On the one hand, R -4-H = R".) by (18.7-). On the
other hand, (20.3) is true; so again (20.2) for A becomes (20.2) for A when
n
Observation. The reduction R'. ) is identified in (20.2) only up to a stan-
dard modification. In fact, however, we have defined R'.) as a very special
standard modification, unique up to an evanescent set.

21. The Energy of a Supermartingale Potential


Let be an almost surely right continuous increasing predictable
L' bounded process generating the almost surely right continuous super-
martingale potential In view of the classical definition (Section
I.XVII.2) of the energy of a measure it is natural to define the energy of
by f o x(t) dA(t), but the standard more useful definition is

A()11Z = [x(t) + x(t-)] dA(t), (21.1)


J0
516 2 IV Basic Properties of Continuous Parameter Supers artingales

and we shall now show that if E{A(+co)2} < +oo, then according to this
definition,

ZE{A(+ao)2}. (21.2)

Suppose first that is bounded. Choose the conditional expectation


y(t) = E{A(+ co)I,9(t)} in such a way that { is a right continuous
martingale. Then

jx(1±)dA(t)= f y(t±)dA(t) - f A(t±)dA(t). (21.3)


0 0

Since is bounded, the processes and are also bounded; so


(by Theorem 7)

E1 f y(t±)dA(t)y = E{y(+oo)A(+oo)} = E{A(+ao)2}. (21.4)

An application of Fubini's theorem to the product measure d,A(t, (o) x


d,A(s, (o) yields

E{ (A(r±)dA(r) = ZE{A(+oo)2 + 2 f"' [A(rt) - A(t+)]dA(t)}

and thereby yields (21.2). If A(+ oo) is not bounded, let (Section 7) S. be an
increasing sequence of finite optional times for which S. = +oo and
for which A is bounded for each n. Then (21.2) for the increasing process
A(S A ) yields (21.2) for when n -* oo.

22. The Subtraction of a Supermartingale Discontinuity


Let {x(t), .*(t), t e 68' } be a right closable almost surely right continuous
supermartingale, suppose that .F(0) contains the null sets, and define
F(+o o) = Y,ER .IF(t). According to Section 1I1.15 the random variable
x(+ oo) = lim,-m x(t) right closes the supermartingale. In the following
discussion will always be the reference filtration but will not always
be mentioned explicitly. Almost every sample function necessarily has
a left limit at every strictly positive parameter value. The left limit process,
defined as x(O) at the parameter value 0 and defined arbitrarily at strictly
positive parameter values when the left limit does not exist, will be denoted
by x( -). Let T be a predictable optional time and define

x(T-) - x(T) on {T< +oo},


10 on {T= +oo},
22. The Subtraction of a Supermartingate Discontinuity 517

and for to R+ define y(t) = Jlirst) (=0 on the set (T= +oo}). Let T be a
sequence of optional times predicting T. Then
(a) the process is an almost surely right continuous super-
martingale, almost surely continuous at T I (r< +,,i, right closable by x(+ oo) + J,
and E{J} 5 E{ x(0) - x(+oo)}.
In fact according to Section 3 (extension of Theorem 3), x(T-) is inte-
grable, and the following supermartingale inequalities are true:

E{x(+oo)} S E{x(T)} 5 E{x(T-)} S E{x(0)}. (22,1)

Hence J is integrable, and E{J} satisfies the stated inequality. To prove that
is a supermartingale observe that is adapted to and that
satisfies the supermartingale inequality if and only if 0 5 s < I
implies that

E{x(t) + Jl),<rs,)I F(s)} 5 x(s) a.s. (22.2)

The sequence (T v s) A t is an increasing sequence of optional times with


limit (T v s) A t. Define x = lim,,... x((T v s) A t). According to Section 3
(extension of Theorem 3), the ordered triple

[X(s), F(s)], i, Y.,F((T v s) A t)]. [x((T v s) A t), 33'((T v s) A 1)]


L 0

is a supermartingale, and it follows that

x(s) Z E{. I.F(s)} z E{x((T v s) A t)I.F(s)) E{x(t)IF(s)} a.s. (22.3)

Hence

E{z - x((T v s) A t)I.T(s)} 5 x(s) - E{x(t)I.F(s)} a.s. (22.4)

and this inequality leads to (22.2) because . - x((T v s) A t) = J1),<rsr)


almost surely. Thus is a supermartingale. An elementary calcula-
tion shows that this supermartingale is right closed by the random variable
x(+oo) - IJI and therefore is right closed by the supermartingale limit
x(+oo)+J.
Finally we prove
(b) If x(T) is Yo measurable, for example (Section 11.3), if
is nearly predictable, it follows that J Z 0 almost surely and that the process
is nearly predictable.
The almost sure inequality 5 x(T,J yields (n cc)
x(T) 5 x(T-) almost surely; that is, J;?: 0 almost surely. We prove that
is nearly predictable by proving that each of the two processes with
respective tth random variables
518 2 IV. Basic Properties of Continuous Parameter Supermartingales

x(T-) I{TS,i x(T) l I Ts r1 (22.5)

(defined as 0 on IT= + oo }) is nearly predictable. The processes with


respective tth random variables

11T.<q (22.6)

are nearly predictable because they are adapted and left continuous; so their
almost sure limit processes (n -+ oo) are nearly predictable. The first almost
sure limit process is that with tth random variable the first in (22.5). The
second almost sure limit process is that with tth random variable the second
in (22.5). Thus is nearly predictable, as asserted.

23. Supermartingale Decompositions and Discontinuities


All processes in this section have parameter set R', and as in Section 22 the
left limit process of an almost surely right continuous supermartingale
will be denoted by Recall that almost sure lower semicontinuity of
means that x(t-) >- x(t) almost surely, simultaneously for all t.

Theorem. Let be a right continuous filtration of a probability space, and


suppose that ,(0) contains the null sets.
(a) A nearly predictable almost surely right continuous supermartingale
{x( ), 9rO} is almost surely lower semicontinuous, almost surely
continuous if the process is a martingale.
If
(b) in (a) is right closable it can be written as a sum
of processes adapted to .f nearly predictable, for which
(bl) Almost every sample function is monotone increasing, constant
except for the - x(-) sample function jumps, and x'(0) = 0.
(b2) is an almost surely continuous supermartingale and is right
closable.
(c) If every. IF(-) optional time is predictable and if .F(-) is predictable,
then every almost surely right continuous supermartingale
is nearly predictable and therefore is almost surely lower semicontin-
uous, almost surely continuous if the process is a martingale.

Proof of (a). Since it is sufficient to prove (a) for x( A n) for all n > 0, we
can assume that is a right closable supermartingale, or a right closable
martingale if the process is a martingale. The process {x( - ),() } is
almost surely left continuous and therefore nearly predictable so the process
x( - ), is nearly predictable, and it follows that when c > 0,
the set
H, = {(t,(o). x(t,uw) - x(t-,w) z c}
23. Supermartingale Decompositions and Discontinuities 519

is a nearly predictable set. Furthermore the graph of the entry time T of H.


is in H, up to an evanescent set ; so (by Theorem 11.9) T is nearly predictable.
Hence x(T-) ;?.t x(T) almost everywhere where T < + oo according to
Section 22(b), and this is impossible unless T = + oo almost surely. We
conclude that the supermartingale assertion of (a) is true. If is a (right
closable) martingale, this result applied to and shows that the
martingale assertion of (a) is true.
It will be convenient to prove (c) before (b).

Proof of (c). Suppose first that is right closable as a supermartingale.


If n E Z+ and k e 1, define Tok = 0, and if T.-, k is already defined, define
Tnk = +oo if T. -I k = +oo and

Tnk = inf {t > Tn_, k: 2k < Jx(t) - x(t-)I 5 2k+' } otherwise.

Then Tnk is optional, therefore predictable by hypothesis, and therefore by


Section 22(b) x(Tnk-) > x(Tnk) when n > 0 and Tnk < +oo. Define Jnk =
x(Tnk -) - x(Tnk) when Tnk < + co and Jnk = 0 otherwise, and define a right
continuous process by

xnk(t) = JkIITks0 (=0 when Tnk = +oo) for I e P+.

Since is right closable it is right closable by lim, x(t) which we denote


by x(+ oo ). According to Section 22, Jnk Z 0 almost surely, the process
is nearly predictable [the reference filtration is here and below], the
process is a right closable (by x(+oo) + Jnk) almost surely right
continuous supermartingale, and E{Jnk} < E{x(0) - x(+oo)}. If the jump
processes are added successively to and if we define (summing over all
n and k) J = EJnk and x'(t) = Exnk, then E{J} s E{x(0) - x(+oo)}. It
follows that J and x'(t) are almost surely finite. The process is nearly
predictable, and its sample functions are almost all right continuous except
for positive jumps Jnk at Tnk. The process is almost surely
continuous and therefore is nearly predictable, and we conclude that
(supposed right closable) is nearly predictable. The supermartingale
is right closable by x(+oo) + J. If is not right closable, this result
implies that x( A n) is nearly predictable for all strictly positive n; so
is nearly predictable.

Proof of (b). In the proof of (c) a decomposition of was obtained satisfy-


ing (bl) and (b2), under the hypothesis that was a right closable almost
surely continuous supermartingale satisfying certain conditions on the
reference filtration A glance at the proof shows that wherever these
conditions on are used, near predictability of suffices.
Chapter V

Lattices and Related Classes of Stochastic


Processes

1. Conventions; The Essential Order


In this chapter certain stochastic process classes which arise naturally in
martingale theory will be discussed. These classes and their relations with
the identically named classes in Chapter IX of Part I will be studied in later
chapters. See Appendix III for the lattice theory to be used.
All stochastic process concepts in this chapter are relative to a specified
filtered probability space (0_F, P; .fi(t), t e I), where I is a linearly ordered
set, arbitrary unless specifically limited. As always the probability measure
is supposed complete; in addition it is supposed that each a algebra .fi(t)
contains the null sets. The stochastic processes considered are adapted to
and have state space (A, .(68)).
Recall from Section 1.8 that in the essential order stochastic processes
are grouped into equivalence classes by identifying processes which are
standard modifications of each other and a process is an essential order
majorant of if

P(y(t) z x(t)) = 1, tel.


Recall further that the essential order infimum ess inf r of a set r of stochas-
tic processes, that is, the essential order infimum of their equivalence classes,
can be obtained as follows. If we write e r to mean that the equivalence
class containing is in r, in other words that is a version of an equiva-
lence class in r (see the remark on the abuse of notation in Section 1.1),
then ess inf r is the equivalence class consisting of all the versions of

(essinfx(t),tel}.
x(. r

Recall that the essential infimum of a family of random variables is deter-


mined only up to a null set. The class of processes in the essential order is a
complete lattice for I a singleton and therefore for arbitrary /.
2. LM when Is a Submartingale 521

The Continuous Parameter Context

This is the context in which I = R', is right continuous, .F(0) contains


the null sets, and the processes to be classified will be almost surely right
continuous. Recall that two almost surely right continuous processes on the
parameter set R' which are standard modifications of each other are
indistinguishable, that is, equal quasi everywhere on 0 x R'.

2. when Is a Submartingale
(See Section 1.IX.2 for the potential theory counterpart of this section.)
Let be a submartingale with an arbitrary linearly ordered
parameter set I having a first element. If S and T are countably valued
optional times, upper bounded in I, with S:5 T, then 5 5 TTX(')
in the essential order. According to the dual of the Special case in Section
111.20, either has no essential order supermartingale majorant and
lim,tE{x(s)} = +oo, or has an essential order supermartingale major-
ant, LM exists, the process { LM is a martingale,

LM x(-) = esslim es (2.1)

up to a standard modification, and

E{LM x(t)} = lim E{x(s)} = su?E{x(s)}. (2.1')


st SC-

(As usual we write LM x(t) for [LM (t).) In view of one of the forms
of the submartingale sampling theorem (Theorem 111.7) the supremum and
directed limit

sup (E{x(S) } : S optional, countably valued, bounded) (2.2)

is equal to the supremum in (2.1'). Thus LM if and only if the set


of expectations of the random variable class

{x(S): S optional, countably valued, bounded} (2.3)

is bounded.
If is a positive submartingale, the existence of LM is
equivalent to the L' boundedness of the random variable class (2.3) and
in this case a trivial argument shows that S in (2.3) need not be bounded in
1. Moreover the hypothesis that I has a first element can be dropped because
in any case the process is left closable by the random variable 0 coupled with
the trivial a algebra (0, i2). In particular, suppose that the positive sub-
522 2.V. Lattices and Related Classes of Stochastic Processes

martingale is uniformly integrable; that is, suppose that there is a uniform


integrability test function 0 for which

sup E{m[x(s)] } < + oo. (2.4)


BE

In this case LM exists, and, since {'F[x(-)] is a positive submar-


tingale, (2.4) is true if and only if LM exists. We now show that then
LMt[LM also exists and

LM 'F[LM LM (2.5)

up to a standard modification. To see this observe that up to standard


modifications

E{4[LM x(s)]J.'(-)} = E{'F[ess limE{x(s')I,F(s)}]I.


S't
S E{ess lim E{t[x(s')]J _fls)} IF(-)}
+' T

= E{

Take the essential limit as s increases to find that the left-hand side of (2.5)
is an essential order minorant of the right-hand side. The reverse order
relation is trivial.

Continuous Parameter Context

Recall from Section IV. 14 that in the continuous parameter context (defined
in Section 1) if {x(-), is an almost surely right continuous submartingale
for which LM x(-) exists, this martingale majorant can be chosen to be right
continuous; the notation LM x(-) will always refer to an almost surely right
continuous version. Moreover, if _fl-)) is a positive almost surely right
continuous submartingale, not only is the existence of LM equivalent to
the finiteness of

sup (E{x(S)}: S optional (< +oo) countably valued}, (2.6)

as already stated in the general context, but the restriction that S be countably
valued is unnecessary. In fact if the supremum in (2.6) is c, let T be an arbi-
trary optional time except that T< +oo. Then if [T], is defined as in
Section 11.2, Example (b2), it follows from Fatou's lemma that

E{x(T)} S liminfE{x([T]n)} 5 c.
n_ .O

If an integrable random variable x(+oo) can be adjoined to right close the


submartingale, as is possible if the submartingale is uniformly integrable
3 Uniformly Integrable Positive Submartingalcs 523

[Section II1.3(e)], then this argument is valid for an arbitrary optional time
T < + x : so in this case S in (2.6) can be an entirely unrestricted optional
time.

3. Uniformly Integrable Positive Submartingales

In Section 1.iX.3 a class D(,4 _) of functions on a Greenian subset of 118'


was defined, and Theorem 1.IX.3 treated functions u for which u e D(µ"1)_ ).
The applications in view were to positive h-subharmonic functions and to
h-harmonic functions. The probabilistic counterparts of class D(µD_) func-
tions on a specified set are the class D stochastic processes on a probability
space, relative to a given parameter set and filtration. as defined in Section
11.11. The following theorem is the probabilistic counterpart of Theorem
i.IX.3, and the applications in view are to positive submartingales and to
martingales. All the processes in the theorem are defined on the same filtered
probability space.

Theorem. Let 1, be a stochastic process with state space


and an arbitrary linear/' ordered parameter set. /J { j x(-) 1, ,9r is a submar-
tingale, the following conditions are equivalent :
(a) a D.
(b) is uniformly integrable.
(c) There is a uniform integrabilitp test Junction N for which the submar-
tingale has a martingale essential order majorant.
(d) (If {x( is a martingale) xt(') - x2('), where
is a positive martingale, in D, and therefore wti/brmlp integrable.
(e) The submartingale { I 1, 9(-)' is right closable.
Moreover, if satisfies (a) and (b), and if O satin/ies (c). then the martingale
{LM is uniformly integrable,

LM (D[LM LM (3.1)

up to a standard modification, and, if is a martingale. each process


in (d) can be chosen so that the submartingale {m[x;( )], t O} has a
martingale essential order majorant.

If here is identified with in Section 111.2 and Section 2. the


present theorem follows; if {x( ),() } is a martingale, one representation
of with the properties stated in (e) and the final assertion of the theorem is

LM [LM (3.2)

(See the corresponding discussion in the proof of Theorem I.IX.3.)


524 2 V. Latuces and Related Classes of Stochastic Processes

Theorem 3 versus Theorem 1.1X.3. The parallelism between these two


theorems is obvious except that no potential theory counterpart of Theorem
3(e) has been suggested. To find such a counterpart, suppose that Jul is a
class D(pD_) h-subharmonic function on a connected Greenian subset D
of R", as in Theorem 1.IX.3, and define v = LMplul, a class D(4_) h-
harmonic function according to that theorem. The function v - Jul is an
h-potential because

GM,(v - Jul) = v - LMoIui = 0;


so (Theorem 1.XII.18) the function v - Jul has u' almost everywhere
minimal-fine boundary limit 0 on the Martin boundary 0"D. According to
Theorem I.XII.19, a quasi-bounded h-harmonic function on D is the PWB"
solution Hj for some h-resolutive boundary function f on a' D, and Hf has
f as its p' almost everywhere minimal-fine Martin boundary limit function.
Moreover we shall show (Theorem 3.1.5) that an h-harmonic function on
D is quasi bounded if and only if it is in D(/D_ ). Hence there is an h-resolutive
function f on a' D which is the µD almost everywhere minimal-fine boundary
limit function of both v and Jul, and Jul S Hf = v. The µo measurable and
integral Martin boundary functions f,, that is, the h-resolutive Martin
boundary functions f,, for which f, z f are potential theory counterparts
of the random variables which close the submartingale { in
Theorem 3(e).

Continuous Parameter Context

In this context, as already noted in Section 2, when in Theorem 3 is


supposed almost surely right continuous, all the least majorants in the
theorem can be supposed almost surely right continuous and in view of
(3.2) each process x;(-) in Theorem 3(e) can be chosen to be almost surely
right continuous and to satisfy the last assertion of the theorem.

4. L" Bounded Stochastic Processes (p >_ 1)


The following theorem is in the context of Theorem 3 and when p > I is
merely a specialization of that theorem. Recall that a process {x(t), t e 1) is
called L" bounded if sup,E,E{ix(t)I"} < +oo.

Theorem. Let be a stochastic process with state space (R, 91(R))


and an arbitrary linearly ordered parameter set, and suppose that p z 1. If
fl-)) is a submartingale, the following conditions are equivalent:
(a) is L" bounded.
(b) The submartingale has a martingale essential order
majorant.
5. The Lattices ('S=, 5), ('S', 5). (S=, 5), (S'. 5) 525

(c) (If is a martingale); where


F(-)) is a positive LP bounded martingale.
(d) (If p> 1); the submartingale is right closable by a
random variable in LP.
Moreover, if p z I and if satisfies (a) and (b), then the martingale
is LP bounded,

(4.1)

up to a stochastic modification, and if (x(-), F(-)) is a martingale, each process


in (c) can be chosen so that the submartingale {x f has a martingale
essential order majorant.

The proof is left to the reader because whatever is not already covered
by Theorem 3 with D(s) = sP follows easily from the discussion in Section 2.
Observe that assertion (c) of the present theorem with p > 1 is slightly
stronger than Theorem 3(d) with 0(s) = sP. In fact in Theorem 3(d) it is
not asserted that if and if has a martingale
essential order majorant, then has a martingale essential order
majorant. However, for cD(s) = sP and p z I this assertion is true because
then
I x(')I P S [xl(') + X2(')]' S 2P-'[x1(') +

Continuous Parameter Context


In this context (see the corresponding remarks in Section 3) if is almost
surely right continuous, all the other processes involved in Theorem 4 can
also be supposed almost surely right continuous.
Observation (Recall the definition of LP in Section 11.11.). It is trivial that a
process in LP is LP bounded. Conversely (context of Theorem 4) if is
LP bounded with p > I and if is a submartingale, it follows
that a LP. In fact it is sufficient to prove that if T is an optional time with
finitely many values then E{Ix(T)IP} 5 sup,.,E{Ix(t)I"}, and this inequality
is true because if s is the maximum value of T in the parameter set order
then E{Ix(T)IP} S E{Ix(s)IP).

5. The Lattices ('Si, 5), ('S+, 5), (Si, 5), (S+1 5)


(See the corresponding potential theory lattices in Section 1.IX.5.)

The Lattice CS', :9)


We assume, as described in Section 1, a specified filtered probability space
with an arbitrary linearly ordered parameter set. Denote by ('Si, 5) the
526 2 V Lattices and Related Classes of Stochastic Processes

lattice, in the essential order, of those stochastic process equivalence classes


under standard modification which contain supermartingales having positive
supermartingale essential order majorants. Recall the essential order nota-
tion 5, z, =, V, A. Let r be a subset of 'S'. If the set ro of supermar-
tingales in the r equivalence classes has an essential order minorant in 'Si,
then (from Section 111.5) the equivalence class containing the versions of
ess inf ro is Ar. It follows that ('Si, 5) is a conditionally complete lattice,
but observe that if r is a subset of 'Si with an essential order majorant in
'Si, then Vr is not necessarily the equivalence class containing the essential
suprema of the class of supermartingales in the equivalence classes of r but
is the essential order infimum of the class of essential order 'S' majorants
of IF.
The careful language used above is correct but inconvenient, and we shall
frequently follow the usual incorrect but convenient abuse of language in
which, for example, ro above is identified with r and Ar is described as
the essential infimtun of r even though r is not a set of processes but a set
of equivalence classes and ess inf ro is not an equivalence class but a process.
If r c 'Si and if r has an essential order minorant in 'Si, then (Section
111.5) Ar = Art for some countable subset rt of r. If r c 'St and if r
has an essential order majorant in 'S', then Vr = Vrt for some countable
subset rt of r. In fact, if r is directed upward in the essential order, then
v r = ess sup r, and the assertion was proved in Section III.5 for essential
suprema. If r is not directed upward, apply this result in the directed case
to the set of 'S' suprema of finite subsets of F.
In the following, we sometimes prime V and A when referring to 'S*.

The Lattice ('S', 5)


The sublattice ('S, 5) of ('S5, S) consists of the 'St equivalence classes
containing positive supermartingales.

The Continuous Parameter Context: (Si, 5), (S', 5)


Observe that if l = Wan equivalence class in 'St contains a right continuous
supermartingale if the class contains an almost surely right continuous
supermartingale and that two almost surely right continuous supermar-
tingales in the same equivalence class are indistinguishable, that is, are equal
quasi everywhere on U8' x 0, or in our other terminology are equal up to
an evanescent subset of this product space. If define

x(t) = liminfx(t) (r rational). (5.1)

Then (Section IV.1) the process is an almost surely right continuous


supermartingale, and
5. The Lattices ('S*, 5), ('S', S), (St, 5), (S', 5) 527

P{x(1) = x(1)} = 1

except possibly for a countable set of values of t. Moreover the limit inferior
in (5.1) is an almost sure limit for each t. Let St be the set of equivalence
classes of almost surely right continuous supermartingales, under the relation
of indistinguishability. Then St can be imbedded in 'St (in the present
continuous parameter context) in an obvious way. If r is a subset of St
and if we denote by 'r the subset of 'S± consisting of the equivalence classes
of the latter set which contain those of r, we have noted above that if AT
exists there is a sequence {x n e 7L+ } in r such that the process
determines AT. The process is almost surely right upper
semicontinuous so 5 in the essential order, and there may be strict
inequality. The process is in the equivalence class of the maximal
essential order St minorant of IF. Thus, if we denote by (St, S) the set St
in the essential order, this set becomes a conditionally complete lattice, for
which we shall use the order symbols :5. 2--, V, A, but the St order infima
and suprema are not inherited from the natural imbedding in ('St, 5). The
argument just given together with the analysis of ('St, 5) shows that if
r c St then the (St, 5) infimum [supremum] of r, if it exists, is the
infimum [supremum] of a countable subset of r. If r is countable and if ro
is a set of supermartingales consisting of one member from each equivalence
class in r, then the equivalence class nr has as one member the process
for the pointwise infimum of 170, and (Theorem IV.4) if r is directed
upward and is bounded above in (S', --5), the equivalence class VF has as
one member the pointwise supremum of rt,.

The Lattice (S+, S)

In the continuous parameter context the sublattice (S+, 5) of (St, S)


consists of the St equivalence classes containing positive right continuous
supermartingales.

The Natural Decomposition in the Continuous Parameter Context

The natural decomposition theorem (Section 111. 19) is valid in the continuous
parameter context, in which all the supermartingales in the theorem are
almost surely right continuous. In fact, if are almost surely
right continuous positive supermartingales and if up to
an evanescent set, then according to the natural decomposition theorem of
Section I11.19 there are positive supermartingales x, such that

5 x&), x, 0 + x'2()
528 2.V. Lattices and Related Classes of Stochastic Processes

up to a standard modification, and it follows that

X;(') x,('), x(') = +'(') + q.e.

6. The Vector Lattices ('S, -<) and (S, -<)


(See the corresponding potential theory vector lattice in Section 1.IX.6.)

The Vector Lattice ('S, -<)

The set 'S+ is a cone as defined in Appendix 111.3 and therefore defines a
specific order on itself for which we use the order symbols Y, A, and
'S+ in the specific order will be denoted by ('S+, S). Define'S = 'S+ - 'S+,

so that each member of 'S can be identified with an equivalence class of


differences between two positive supermartingales. Each
random variable x, (t) - x2(t) is well defined off the probability null set of
common infinities of x, (t) and x2(t). If'S is ordered by the specific order with
positive cone 'S+, we obtain a partially ordered vector space ('S,

Theorem. (a) The space ('S, ) is a conditionally complete vector lattice.


[In (b), (c), (d) let r be a subset of'S with a specific order majorant.]
(b) yr is the specific order supremum of a countable subset of r.
(c) If r' is the class of specific order majorants of IF, then Vr, -< r'.
(d) If r is directed upward in the specific order, then yr = W.

The duals of (b), (c), (d), involving infima, are obtained by replacing IT
by - r. Since r' is directed downward in the specific order in (c), the dual
of (d) implies that Ar' = Ar' = yr.
The reader will observe that this theorem has precisely the same statement
as Theorem 1.IX.6, although S in that theorem does not have the same
meaning as 'S here. The point is that the contexts of the two theorems are
quite different but the order properties in the two contexts are identical.
The proof of Theorem 6 is simply a translation of that of Theorem 1.IX.6
into the present context. For example, to prove that Ar' = Ar' = yr if
r c 'S+, we can follow the proof of this assertion in Section I.IX.6. That is,
we now interpret the members u, v, ¢, ... of S+, r, r' in that section as
positive supermartingales and interpret R° there as a generalized probabil-
istic reduction, namely, as the equivalence class of essential infima of the
set of positive supermartingale essential majorants of The details of
the translated proof are left to the reader.

Continuous Parameter Context: The Vector Lattice (S, --<)


The set S+ is a cone and therefore defines a specific order on itself, for
which we use the order symbols ::, Y, A, and S+ in the specific order
7. The Vector Lattices ('S.,5) and (S., S) 529

will be denoted by (S+,

If and are positive almost surely right


continuous supermartingales, we describe as a specific minorant of
and write if this relation holds between the equivalence classes
determined by the processes. The corresponding significance is given to
Y and other abuses of notation involving processes and equivalence
classes.
Define S = S+ - S+, and denote by (S, :!9) the vector space S ordered
by the specific order with positive cone S. Then S can be identified with a
subset of 'S. Moreover the following relations between ('S, ) and (S+, )
are true.
(rl) If and are in S+, then in ('S,:!5) implies that
in (S, In fact by hypothesis there is a in 'S+ such that
is a stochastic modification of and therefore is a
stochastic modification of and in fact is indistinguishable from so
in (S, )
(r2) If and are in S+, then in (S, if and only if
in ('S, ::5). In fact "if" is trivial, and "only if" follows from (rl).
(r3) If E'S and if is almost surely right continuous, then ES
because (up to a standard modification) by hypothesis xt
with e'S+ ; so xt with x+1(-) c- S+.
If r c S+ we can identify r with a subset r of'S+. If is in the equiv-
alence class AT then in ('S, by (rl). Moreover, if is in
S+ and if t in (S, then S I in ('S, so S in
('S,:::5), and therefore in (S, So k t exists and is the equiv-
alence class in S+ containing Hence (S+, ) is a conditionally complete
vector lattice. We leave to the reader the verification that Theorem 6(b)-(d)
holds for (S, -<).

Intrinsic Definition of S

It will be shown in Section 13 that the processes in S can be characterized


without involving supermartingales.

7. The Vector Lattices ('S., ) and (S., -<)


(See the corresponding potential theory vector lattice in Section 1.IX.7.)

The Vector Lattice ('S., -<)


A supermartingale specific order majorized by a martingale is itself a martin-
gale. If IF is a set of positive martingales with YI = then is a specific
order majorant of each member of r, so is a positive supermartingale,
and since is also a specific order majorant of r, it follows that
530 2.V Lattices and Related Classes of Stochastic Processes

and so this process is a martingale. Hence (Appendix 111.8)


if 'S' is the cone in 'S whose equivalence classes contain the positive martin-
gales, the set 'S. = 'S - 'S is a band in CS,=0. The restrictions to 'S;
of the essential and specific orders coincide. If I c 'S then, with some
abuse of the notation LM and GM,

Yr=LMr, Ar=GMr
in the sense that if one side of an equation exists, the other side also exists
and there is equality. According to Section 4, a martingale is in an 'S.
equivalence class if and only if the martingale is L' bounded.

Continuous Parameter Context: The Vector Lattice (S. -<)

In the context of (S, an almost surely right continuous supermartingale


specific order majorized by an almost surely right continuous martingale is
itself a martingale. If r is a set of positive almost surely right continuous
martingales with yr = then as above it follows that x(-) is a martingale.
Thus if S ; is the cone in S whose equivalence classes contain the positive
martingales, the set S. = S.' - S.' is a band in (S, S) and as such is a
conditionally complete sublattice of (S, -<).
Observe that in the continuous parameter context if x(-) is an L' bounded
almost surely right continuous martingale, then x(-) a Sm because up to a
standard modification xt with e S ; ; so up to an
evanescent set x(-) = x x2(.), with e S.*.

8. The Vector Lattices ('Si, and (Se, -<)


(See the corresponding potential theory vector lattice in Section 1.IX.8.)

The Vector Lattice ('Sr, <)

Recall that in Section 111.21 we defined a supermartingale potential as a


positive supermartingale {x(-),.F(-)} with inf,E,E(x(t)} = 0, equivalently
with 0 up to a standard modification. A positive supermartingale
specific order majorized by a supermartingale potential is itself one and if
r is a set of supermartingale potentials with Yr = then x(') must be
a positive supermartingale for which is also a specific order
majorant of r and therefore must also be a version of Yr so 0
and is a potential. It follows that if'SP is the cone in 'S whose equivalence
classes contain the supermartingale potentials then the set 'Sp = 'S; - 'So
is a band in ('S, ) and as such is a conditionally complete vector sublattice
of('S,-<).
9. The Vector Lattices ('S,6. ) and (S, , S) 531

Continuous Parameter Context: The Vector Lattice (Sr,

In this context observe that if is a supermartingale potential, then


is also a supermartingale potential and conversely. The reasoning used above
with trivial adaptation to the context of S shows that if SP is the cone in
S whose equivalence classes contain the right continuous supermartingale
potentials, then the set Sp = SP - S, is a band in (S, -<) and as such is a
conditionally complete vector sublattice of (S, -<).

Theorem. 'SP = 'S; and Sp = S..

The proof of these relations in the potential theory context (Section 1.IX.8)
is applicable in the present context.

9. The Vector Lattices ('Sqb, and (Sqb, )

(See the corresponding potential theory vector lattice in Section 1.IX.9.)

The Vector Lattice ('Sb, -<)

The class'Sgb is defined as the subset of'S+ whose equivalence classes contain
the quasi-bounded positive supermartingales, that is (Section IV.6), contain
the supermartingales x(-) which satisfy the following equivalent conditions,
in which all processes are adapted to the specified filtration.
(a) The process is the specific order essential supremtun of a set of
bounded positive supermartingales.
(b) The process is the limit of a specific order increasing sequence
of bounded positive supermartingales; that is, is the sum of a
series of bounded positive supermartingales.
The class 'Sqb is a cone which satisfies the conditions (Appendix 111.8)
implying that the set 'Sqb = 'Sqb - 'Sqb is a band in ('S, and as such is
a conditionally complete vector lattice. The equivalence classes in this band,
and also their stochastic process members, are called quasi bounded. Observe
that if in (a) and (b) above is a martingale, then the bounded positive
supermartingales in (a) and (b) must also be martingales because they are
specific order minorants of

The Bands 'S,,,qb = 'S. n 'Sqb and 'Spgb = 'Sp r 'Sb

In view of Section 8 these two bands are mutually orthogonal, and 'Sqb =
'S,,,qb + 'Spgb. The band 'S,,,qb is the band in 'S generated by the equivalence
class of the process all of whose random variables are identically 1. According
532 2.V Lattices and Related Classes of Stochastic Processes

to Theorem IV.10, 'Spgb = 'S, n D if the parameter set has a first point but
not a last point.

Continuous Parameter Context: The Vector Lattice (Sqb, :!)

In this context a quasi-bounded positive supermartingale which is almost


surely right continuous and which is the sum Eo x of a series of bounded
positive supermartingales is the sum of a series of bounded positive almost
surely right continuous supermartingales. In fact the sum Eo of bounded
positive almost surely right continuous supermartingales is almost surely
right continuous (Theorem IV.4), and except for a countable parameter set,
P{x(t) = Eo x (t)} = 1. Hence by almost sure right continuity x() =
Eo up to an evanescent set. With the help of this result the reasoning
at the beginning of this section, with trivial adaptation to the context of S,
shows that if SQ6 is the cone in S whose equivalence classes contain the
supermartingales which satisfy the equivalent conditions (a) and (b) above
with all the positive bounded supermartingales involved supposed almost
surely right continuous, then the set Sqb = Sqb - Sqb is a band in S and as
such is a conditionally complete vector sublattice of S. Define S,,,qb =
S. n Sqb and S,,b = S, n Sqb to obtain orthogonal bands with vector sum
Sqb. According to Theorem IV. 11, Sab = S, n D.
In the following theorem UI denotes the class of uniformly integrable
processes in the given context or, with the usual abuse of language, the set
of equivalence classes in 'S or S whose members are uniformly integrable.

Theorem. 'S,,,yb = 'S. n UI and Sqb = S. n UI.

It is sufficient to prove the first equality, and [by Theorem 3(d)] even
sufficient to prove 'S n UI. In this context if is a uniformly
integrable positive martingale, it is right closable [Section 11I.3(e)] by some
random variable x; so x(t) = E{xj.'F(t)} almost surely. If x is the bounded
martingale defined by x (t) = E{x A nI.F(t)}, then lim,,... x (t), t c Ig+
defines a martingale in the equivalence class of x(-); so x(-) is quasi bounded.
Conversely, if x(-) is a quasi-bounded martingale, so that there is a sequence
of bounded martingales, { y.(-), n e Z'), such that x(-) = Eo y.(-), then y
is right closable, say by a random variable yy, and must be right closable
by Eo so is uniformly integrable.

10. The Vector Lattices ('Ss, ) and (Ss, -<)


(See the corresponding potential theory vector lattice in Section 1.IX.10.)
Define 'S, = 'S b relative to ('S, -,), and define 'S; = 'S, n 'S'. Define
S, = Sqb relative to (S, :!5), and define S, = S, n S' S. The equivalence classes
1. The Orthogonal Decompositions 'S,, = 'Siiijb + 'S,,, and S. = S,,,,, + S 533

in the bands 'S, and S, and the processes they contain are called singular.
A process in 'S+ [S'] is singular if and only if every bounded 'S+ [S']
specific order minorant of the process is a standard modification of [indis-
tinguishable from] the identically zero process. We leave to the reader the
verification of the fact that in the continuous parameter context if is
an almost surely right continuous supermartingale, then e'S; if and
only if a S; and that e'S, if and only if a S,. Thus in the con-
tinuous parameter context the equivalence classes in S, [S,] can be identified
with those in'S; ['S] which contain almost surely right continuous processes.
We shall denote by and 'S
S, respective bands
and of singular supermartingale potentials are defined correspond-
ingly. Thus we now have an orthogonal decomposition of each of the lattices
'S and S into four bands:

'S = 'S,,,qb + 'S,,,, + 'Spgb + 'Sp S = S,,,qb + S,,,, + Spgb + Sp,. (10.1)

11. The Orthogonal Decompositions 'Sm = 'Sqb + 'Sms and


Sm=Sgb+Sm,
In this section we restrict outselves to the continuous parameter context
except in the last paragraph where we show how the work can be modified
to be applicable to the ('S, context. Suppose then (continuous parameter
context) that is an almost surely right continuous positive martingale,
that is, e S.+. Then (Section 111. 13) lim,..,, x(t) = x(+ oo) exists almost
surely.
Case (a) : a UI. In thise case, equivalently (Theorem 3) if a D,
equivalently (Theorem 9) if a Sqb, this martingale is right closable by the
random variable x(+oo) (Section Ill. 14); that is, x(t) = E{x(+oo)I.$(t)}
almost surely. Thus x(+oo) determines up to an evanescent set.
Case (b): We show that if and only if x(+co) = 0 almost
surely. If x(+oo) = 0 almost surely, then every bounded specific order
minorant of in S+ has almost sure limit 0 at + oo ; so according to
Case (a), the minorant is indistinguishable from the zero process. Hence
c -S.+. Conversely, if e S ;,, then for every positive constant c,
E{x(+o0) A Iim E(x(s) A S a.s.

and the martingale on the left is bounded and so is indistinguishable from


the zero process. It follows that E{x(+ oo) A c} = 0 for every c; so x(+ oo)
= 0 almost surely.
Case (c) : General case. In all cases we can write in the form

x() = E{x(+co)j..()} + [x() - E{x(+oo)I.F()}] (11.1)


534 2.V. Lattices and Related Classes of Stochastic Processes

and choose the conditional expectations in such a way (Section IV. I) that
is an almost surely right continuous martingale. This
conditional expectation martingale comes under Case (a). The bracketed
difference in (1 1.1) is a martingale with almost sure limit 0 at + oD and so
comes under Case (b). Thus (11.1) exhibits as the sum of its quasi-
bounded and singular components.
If S,,,, then esslim,tx(t) exists (Section 111.13), and Cases (a)- (c)
go through as in the continuous parameter context.

12. Local Martingales and Singular Supermartingale


Potentials in (S, -<)

Local Martingale (Continuous Parameter Context)

A process in this context is called a local martingale if there is an increas-


ing sequence of finite optional times with almost sure limit + oo such that
each process

{x(T A t),.'F(t),teiFe } (12.1)

is an almost surely right continuous martingale. For example (take T. - n),


an almost surely right continuous martingale is a local martingale. Observe
that if T. in (12.1) is replaced by T A n, that is, if the almost surely right
continuous martingale (12.1) is stopped at t = n, then the stopped process
is a martingale (Section IV.3). uniformly integrable because it is right closed
by x(7 A n). Hence it is no restriction on a local martingale if T. in (12.1)
is supposed bounded and if each martingale (12.1) is supposed uniformly
i ntegra ble.
It is trivial (calculate expectations or use the fact that S, 1 S.) that an
almost surely right continuous supermartingale potential which is a martin-
gale is indistinguishable from the identically zero process. As the following
theorem shows the situation is quite different if "martingale" is replaced
here by "local martingale."

Theorem. A process in Sn is singular if and only it it is a local martingale.

If Sn, and nd7L' . define

A infIt En':x(t)>_n},
Then T, = + oo because almost every almost surely right continuous
supermartingale sample function is bounded on compact intervals. Observe
that )e S+ with r,.x(t) < n for t < T. and that with r(t) < n
13. Quasimartingales (Continuous Parameter Context) 535

for t < T. and y(t) = 0 for t Z T. Thus x(), and is bounded;


so 0 quasi everywhere, that is,

x(T A t) = x(t) = (12.2)

almost everywhere on the set {T. > t}. Since the first and third terms in
(12.2) are trivially equal almost everywhere on the set {T 5 t) [because
this set is in .F(t) and the function x(T A t) is .fi(t) measurable], the process
(12.1) is a martingale. Hence is a local martingale. Conversely, suppose
that e S, and that is a local martingale, so that there is an increasing
sequence T. of finite optional times with almost sure limit + oo such that
each process (12.1) is a martingale. If is in S+ and is a bounded specific
order minorant of then z(T A -)S x(T A ); so z(T A ) is a martin-
gale with a bound independent of n. The martingale equality E{z(0)} =
E{z(T A t)} yields E{z(0)} = E{z(t)} in the limit, and therefore
Since it follows that 0 quasi everywhere; so is singular,
as was to be proved.

Adaptation to the Parameter Set Z+

Under the obvious definition of a local martingale with parameter set I+


a trivial adaptation of the preceding proof shows that the counterpart of
Theorem 12 for the parameter set Z+ is true.

13. Quasimartingales (Continuous Parameter Context)

In the continuous parameter context (see Section 1) quasimartingales have


been defined variously. We shall call an almost surely right continuous
adapted process a quasimartingale if it is L' bounded and if
there is a constant c for which 0 = t0 < tt < -,, +oo implies that

YE{IE{x(tk) - x(tk+t)I,'(tk)}I) 5 c. (13.1)


0

Observe that, for a given probability space and filtration, the class of quasi-
martingales is linear and includes L' bounded almost surely right continuous
martingales (set c = 0) and positive almost surely right continuous super-
martingales (set c = E{x(0) } ). Hence the members of S are quasimartingales.
According to the following theorem, the converse is also true.

Theorem. An adapted almost surely right continuous Lt bounded process


is a quasimartingale if and only if it is the difference between two
almost surely right continuous positive supermartingales; that is, a S.
536 2.V. Lattices and Related Classes of Stochastic Processes

We have already noted that the members of S are quasimartingales.


Conversely, suppose that {.X( ), ()} is a quasimartingale satisfying (13.1).
A trivial iterated conditional expectation argument shows that if k ?:j, the
kth term of the sum in (13.1) majorizes the same term with .F(tk) replaced
by .,F(tj). Hence the definition

y(tj) = Y I E(x(tk) - x(tk+t)IF(tj))I (13.2)


k2j

is meaningful, and E{y(0)} < c; the series converges both almost surely
and in L'. Furthermore the L' convergence of this series implies that
L' limk-m E{x(tk)I.IF(tj)) exists. Since this is true for all j, so that tj can be
any number in IB+, and since two sequences t, can be combined into a single
one, this L' limit property implies that

L' lim
t-m
E{x(t)I.IF(s)} = x.(s)

exists for all s. Moreover Y.(s) is .ms(s) measurable because JF(0) contains
the null sets, the process is L' bounded because

E{ Ixm(s)I } = E { I L' lim E(x(t)I.f(s)} I } < lirm E{ IE{x(t)I.F(s)} I}


(13.3)
< su
,zgE{Ix(t)I} < +oo,
and {xm(t,), .F(t,) } is a martingale because if st < s2 and if A E.1(sl), then

r x,,,(si)dP = `n L' lim E{x(t)I5(s2)} dP = lire fn E{x(t)I.IF(s2)} dP

= lim f E{x(t)I'F(si)}dPL' li mE{x(t)I.(sl)}dP


n n

=
1

Let Q be the set of positive dyadic rational numbers. According to Section


IV.1, the right limit process tE R+} defined by

X., (I) = lim xm(s)


Q 31t

(t) is defined arbitrarily when this right limit does not exist] is an almost
surely right continuous martingale and is obviously L' bounded. Moreover
(Section 4) the process is the difference between two positive almost
surely right continuous martingales. At the price of replacing by
[which does not change the sum in (13.1) or y(t) in (13.2)] we can
therefore suppose from now on that L' lim,-m E{x(t)I.fls)} = 0. If the
13. Quasimartingales (Continuous Parameter Context) 537

absolute value signs in (13.3) are omitted, the sum is

x(t) - Lt lim E{x(tk)I.F(tt)} = x(tt) a.s.,


k-.D

and therefore

P{y(t) z x(t)} = 1, (13.4)

Furthermore, if j > i, manipulation of conditional expectations yields


1-I
y(ti) - E{ y(t;)I-,`(t,)} z Y I E{x(tk) - x(tk+t)I ,F(ti)}
k=i (13.5)
z E{x(ti) - x(tt)I,F(ti)} a.s.

According to (13.5), the process { y(t,) - .F(t.) } is a supermartingale.


Thus the positive processes

{Y(t),.f(t.)}, {Y(t.) -
are supermartingales. We have already noted that E{ y(0)} 5 c. The reader
can readily verify that adding points to t replaces by a supermartingale
which majorizes on the original sequence t.. Let {y"(j2-"),.F(j2-"),
j e Z+ } be, for n e Z+, the version of when t, = { j 2-", j e Z+ } . For tin Q
and n so large that y"(t) is defined, y"(t) increases almost surely as n increases.
Define to obtain positive supermartingales

{z(t),."(t), teQ), {z(t) - x(t),.F(t), teQ}.

Apply Section IV.1 again to find that the positive right limit processes
z( - are almost surely right continuous supermartingales
relative to on the parameter set R+. The desired representation of
is x() = z() - [z() - x()}.
Observation. Theorem 13 asserts that the quasimartingale x() has a
representation x() = x,() - x2(), where x.() is a positive almost surely
right continuous supermartingale. There is some interest in minimizing x, ()
and x2(). An indirect but interesting way is to observe that the class of pairs
(x,(), x2()) with the stated properties has the property that if (x'1(), x2())
and (x,,xz) are in the class, then (x1() A x;(), x2() A x2()) is in the class.
Thus the set of first components and the set of second components are both
directed downward, and it is not difficult to show that the properly smoothed
version of the pair of essential infima of these two sets is the desired minimal
pair. A more direct method suggested by the classical discussion of the
positive and negative variations of a function of bounded variation is to
538 2.V. Lattices and Related Classes of Stochastic Processes

replace the discussion of as defined by (13.2) by the processes defined


by the two series

Y [E{x(Ik) - x(tk+l)I',F(I,)} v 0],


k2j
Y [E{x(tk) - x(tk+l)I'f(t)}
- kzj A 0].

The Parameter Set 7L+

A process {x(n), ,fi(n), n e Z+ } is called a quasimartingale if it is L' bounded


and if

YE{IE{x(k) - x(k + 1)1.F(k)}I} < +w.


0

An adapted process is a quasimartingale if and only if it is the difference


between two positive supermartingales. The proof follows that in the
continuous parameter case, with obvious simplifications.
Chapter VI

Markov Processes

1. The Markov Property


Let be a stochastic process with state space (X, 9t") on a filtered
probability space (f2, .F, P; .fi(t), t E 1). The process is called a Markov
process if when s < 1 and A E ", then

P{x(t)EAI.,F(s)} = P{x(t)EAIx(s)} a.s. (1.1)

Define W(s) = .F{x(r), r >_ s}. It will now be shown that (1.1) implies

E{zl,F(s)} = E{zJx(s)} a.s. (1.1')

whenever z is a function from 0 into I8 which is #(s) measurable and is


either integrable or positive. If s < t and z = with A in (, equation
(1.1') reduces to (1.1). The validity of (1.1) or the equivalent (1.1') will be
referred to as the Markov property. To prove that (1.1') is true under (1.1) it
is sufficient according to the usual approximation procedure to prove (1.1')
for z the indicator function of a set in V(s). Since this a algebra is generated
by the algebra of finite disjoint unions of sets of the form {x(tt) e Aj, j < n},
with to = s < < t and A; in 1, it is enough to prove (1.1') for z =
zo z,,, where zj = ¢;[x(t;)] and 4i is a bounded measurable function from
(X,.f") into (l ,-R(R)). Equation (1.1') is trivial when n = 0. When n = 1,

E{zoz, f(s)} =z0E{z,Jf(s)} a.s.. (1.2)

and if z, = 1,[x(11)], the right side of (1.2) becomes, using (1.1),

zoP{x(t,)EAl. (s)} = zoP{x(t,)EAlx(s)} = E{zoz1Ix(s)} a.s., (1.3)

so that (1.1') is true for z = zoz, when z has this special form. Equation (1.1')
for z, = 0, [x(t,)] then follows using the usual approximation procedure.
We now proceed by induction. If (1.1') is true for z = zo zk with an
arbitrary choice of to = s < t, < < tk and functions 0a, . . , Ok, for some
.

k>_ 1.
540 2.VI. Markov Processes

E{zo ... zk+,IF(s)} = zoE{E{z, zk+,L-1F(t,)}I.F(s)}

= zoE{E{z, ... zk+,1x(t,)}I. (s)}

= zoE{E{z, zk+, Ix(t,)}I x(s)} (1.4)

= zoE{E{z, zk+jI.f(t,)}Ix(s)}

= E{zo zk+,Ix(s)} a.s.,

as was to be proved.
A manipulation of conditional probabilities which will be left to the
reader shows that if the process parameter set is a set of consecutive integers,
(1.1) is true in general if true for t = s + 1.
The Markov property can be reformulated: is Markovian if
and only if the past and future are independent, given the present, or, in
precise form, if A e .ms(s) and M e v(s), then

P{AnMJx(s)} = P{Ajx(s))P{MIx(s)} a.s., (1.5)

or equivalently, if y is F(s) measurable and z is W(s) measurable and both


are positive or both are bounded,

E{yzlx(s)} = E{yjx(s)}E{zjx(s)} as. (1.5')

To derive (1.5') from the Markov property, suppose that y and z are measur-
able, as described, and bounded. Using (1.1'),

E{yzl,F(s)} = yE{ziF(s)} = yE{zlx(s)} a.s.; (1.6)

so performing the operation E{ - Ix(s) } on the first and third terms yields
(1.5'). Conversely, under (1.5')

E{yzIx(s)} = E{yE{zlx(s)}I x(s)} a.s.; (1.7)

so

E{yz} = E{yE{zjx(s)}}. (1.8)

If now y is the indicator function of a set A in F(s), equation (1.8) becomes

IzdP= j'E{zlx(s)}dP. (1.9)


n n

which is the integrated version of (1.1').


1. The Markov Property 541

The symmetry of (1.5') implies that if lx(-), .F(-)) is Markovian, the


process is Markovian when the parameter order is reversed.
Roughly, a Markov process under time reversal is a Markov process.

The Markov Property for Processes with Topological State Spaces

According to the discussion in this section, the Markov property can be


stated in the following form : an adapted process is Markovian
if and only if

E{ f [x(t)] I.I(s) } = Elf [x(t)] I x(s) } a.s. (1.10)

when s < t and f is an arbitrary bounded measurable function from the


state space into R. We leave it to the reader to verify that if the state space
is a Polish space coupled with its Borel sets, the process is Markovian if and
only if (1.10) is satisfied whenever f is bounded and continuous, and in
particular, if the state space is locally compact and second countable, it is
sufficient if (1.10) is satisfied whenever./'is continuous with compact support.

Initial Distribution and Transition Function of a Markov Process

Let (x(t),.F(t), to 1) be a Markov process with measurable state space


(X, .). If I has a first point to, the distribution of x(to) is called the initial
distribution of the process. If there is a stochastic transition function q with
parameter set I (Appendix VI.3) such that for s and t in I with s < t and for
A in X,

P{x(t)eA.f(s)} = q(s, x(s); t, A) a.s., (1.11)

then is said to have transition function q. This equation implies, when


the operation E{ -Ix(s) } is applied to both sides, that the right side is almost
surely P{x(t) E A I x(s) } . Thus (1.11) implies that -fl-)) is Markovian.
Recall that by definition of transition function with parameter set I the
Chapman-Kolmogorov equation

q(t, q; u, A)q(s, ; t, dn) (s<t<u) (1.12)


x

is satisfied. Observe that in the following equation for iterated conditional


expectations (see Section 1.4),

E{P{x(u)EAIF(t)}I.y(s)} a.s. (s < t < u), (1.13)

conditioning by r (s) and .fi(t) can be replaced, respectively, by conditioning


542 2.V1. Markov Processes

by x(s) and x(t), in view of (1.1'), and that as so modified (1.13) can be
written in the form

q(s, x(s); u, A) = f q(t, rl; u, A)q(s, x(s); t, dq) a.s., (1.14)


x

which is (1.12) up to the "a.s." in (1.14). Thus the Chapman-Kolmogorov


equation amounts to a property of conditional expectations combined with
the Markov property.

Absolute Probability Function of a Markov Process


If q is a stochastic transition function with parameter set I and state space
(X, i"), an absolute probability function for q is defined as a function (t, A) i-
p(t, A) from I x .2" into [0,1 ] for which p(t, ) is a probability measure and

p(t, A) = f q(s, ; t, A)p(s, dd) (s < t). (1.15)

If If is a Markov process with transition function q and parameter


set !, the function (t, A) - p(t, A) defined by

p(t,A) = P{x(t)eA} [(t,A)eI x 1] (1.16)

is an absolute probability function for q and is called the absolute probability


function of the Markov process. This absolute probability function and the
transition function together determine the finite-dimensional distributions
of the process: if t, < . < t are parameter values,

P{x(tt)eAjj:<n}=JA,
fA2
(1.17)
J q(ta-, S"-, ; tn, dcn).
A
In particular, if ! has a first point to, and if v is the distribution of x(to),

p(1, A) = q(to, ; t, A) v(dd) (t > to). (1.18)


JA

Recall (Section 1.10) that to a prescribed state space and prescribed finite-
dimensional distributions correspond a stochastic process with those finite-
dimensional distributions if the state space satisfies a certain weak condition
(for example, if the state space is a Polish space coupled with its Borel sets)
and if the finite-dimensional distributions are consistent. Thus, under the
I. The Markov Property 543

stated restriction on the state space, if I has a first point, to each specified
initial distribution and stochastic transition function with parameter set I
correspond a Markov process with the specified initial distribution and
transition function; if / has no first point, to each stochastic transition
function q and absolute probability function for q correspond a Markov
process with the specified transition and absolute probability functions. The
finite-dimensional distributions of the process are determined by (1.17), and
the absolute probability function is determined by the initial distribution if
1 has a first point.

Notation Involving the Initial Distribution


When I has a first point and the Markov process context requires the
identification of the initial distribution, two systems of notation will be used.
We shall sometimes identify the initial distribution v by writing P,, and E,,
for probabilities and expectations, specializing to P. and EE when v is
supported by { 1; in the latter case the process will be said to have initial
point or to be a process from . Alternatively, we may denote probabilities
and expectations by P and E but identify the initial distribution by a sub-
script in the process notation, writing for the process and specializing
to if the process has initial point .

The Stationary Context


If I = Z+ and if there is a stochastic kernel A) F- A) for which

q(s, c; s + 1, A) = A) (Se Z').

then the transition function value q(s, ; s + t, A) does not depend on s for
any value oft = 1, 2, ..., and in fact according to the Chapman-Kolmo-
gorov equation, (s, A) t-4 q(s, ; s + t, A) is the tth kernel iterate of In
this case is said to have stationary probabilities and to have
transition function p. If 1= R+ and if q is stationary (Appendix V1.3), that
is, if there is a stationary continuous parameter transition function (t, , A) F-.
p(t, , A) for which

t,A) (seI8+.0 < teR+),

then {x( ), is said to have stationary transition probabilities and


to have stationary transition function p. In this context recall that the
Chapman-Kolomogorov equation becomes

p(s + t, , A) = p(t, q. A)p(s, , dn) (0 < s. t) (1.19)


Jx
544 2.V1. Markov Processes

and equation (1.15) linking absolute probability and transition functions


becomes

µ(s + t, A) = f p(t, , A)p(s, dd) (s z 0, t > 0); (1.20)


x

here s = 0 yields the version of (1.18) in the present context, since u(0, ) is
the initial distribution.

2. Choice of Filtration
Let {x(-), .F(-)) be a Markov process with an arbitrary linearly ordered
parameter set

mo(t) = .flx(s), s s t},


mo(t) = a algebra generated by .mo(t) and the null sets,
.moo =

F'(t) = a algebra generated by .4z(t) and the null sets.

The larger the a algebras of the filtration the more significant is the
assertion that has the Markov property. The minimal choice of
filtration to which is adapted is operating on both sides of
(1.1) with E{-I9o(s)} shows that is Markovian. In the other
direction it is trivial that is Markovian. Thus

.lo(t) c F(t) c F'(t), . (t) c F'(t), (2.1)

and the filtrations AO(-), some of which may be identical,


all make Markovian. The following lemma will be used in Section 7.

Lemma. If .f(t) c .tea for some parameter value t, then F'(t) c .ryro(t).

To prove the lemma, consider the class r of bounded random variables


y for which

E{ylflt)} = E{yl. o} a.s. (2.2)

If y, is bounded and .mo(t) measurable, and if y2 is bounded and is mea-


surable with respect to the a algebra .9` generated by the null sets and
.F{x(s), s >- t}, then YIY2 e I- because

E{y,Y2I. (t)} =YiE(Y2IF(t)) =YiE{YsIx(t)} =Y1E{Y2I9o(t)}


(2.3)
= E{YIY2I.'o(t)} a.s.
3. Integral Parameter Markov Processes with Stationary Transition Probabilities 545

Since r is a linear class which is closed under bounded almost everywhere


convergence, r includes the bounded random variables measurable with
respect to .moo (t) Y.F' ; that is, r includes the bounded .F, measurable random
variables. In particular, if y is bounded and ,F'(t) measurable, the left side
of (2.2) is equal almost surely to y, and (2.2) therefore implies that y is
.mo(t) measurable; that is, .F'(t) a .mo(t), as was to be proved.

3. Integral Parameter Markov Processes with Stationary


Transition Probabilities
Let (X, 3l) be a measurable space, let p be a probability measure on X, and
let A) F- p(s, A) be a stochastic kernel with state space (X, X) (Appendix
VI.1). Let {x(n), .fi(n), n e Z+ } be a Markov process with initial distribution
p and stationary transition probabilities, with transition function p (Section
1) so that

PP,{x(A)EA} = µ(A), 1)eAIF(n)} =p(x(n),A) a.s., (3.1)

Pa{x(j)eAjj5 n} = j'(do)dt)
,
jP(.i ,(3.2)
and i n particular,

P{{x(j)EAj,j 5 n} = lAp() fA, (3.2')


f.n
Abusing language somewhat we shall describe this Markov process from
now on as a Markov process with stationary transition function p. The
evaluation (3.2) determines the finite-dimensional distributions of the process
and thereby determines P. on .Or{x(n), n E Z+ }, because the latter a algebra
is generated by the algebra Un o.F{x(m), m 5 n}. Conversely, as noted in
Section 1, to a specified measurable state space, initial distribution, and
transition function corresponds a Markov process {.i(n), .fi(n), n e Z+ } with
the specified initial distribution and transition function, under a slight
restriction on the state space. Here the random variables of the process are
defined on the space ' of functions from Z+ into X, i(n) is the function
value at n, and (3.2) is now a definition. Since the parameter space is Z +,
the Tulcea generalization of the Kolmogorov extension theorem shows that
no restriction need be imposed on the measurable states ace. We shall
sketch the construction of the process. Each point w of is a sequence g
t , ...) of points of X, i(n, d)) = c,,, and .4 (n) = JJIF{z(0), ... , z(n) },
.4g = Y i(n) = {x(n), n e Z } . The measure P is defined on I as the
extension to .I of the measure on finite-dimensional product sets determined
by setting P{z(j) E Aj, j 5 n} equal to the right side of (3.2). (The Tulcea
546 2.V1. Markov Processes

proof that the extension exists is omitted.) This construction has the advan-
tage that the space (!ft, A) is defined without reference top orp, an advantage
illustrated by the fact that if Jr contains the singletons, the function g F-.
{R} for A in is 3C measurable and for arbitrary p

PN{R} = f (3.3)
X

because these assertions are true when A is a finite-dimensional product set.


If is an integral parameter Markov process on a probability
space (S2, JF, with the above state space and transition function, the map
0: w [x 0, w), x(l, w), ... ] is measurable from (S2, .fl into (L'2, and
P,, = 0' (P,) in the sense that ¢-' (,fi(n)) = .F{x(j), j < n}} c 5(n), -`(
.F{x(j),je7L+} c 9, and P{R} for Ac.?. In view of
these facts it is usually correct as well as convenient to prove theorems for
discrete parameter Markov processes by proving them for the special ones
as just defined on (n, A).

Arbitrary Initial Measures

Observe that the preceding discussion remains valid if p(X) # 1, which


implies that P{S2} 1. To avoid pathology, we shall always assume that y
is the sum of a sequence of finite measures, but we allow p to be infinite
valued. All that is needed is the acceptance of the idea that in probability
theory the measure on the space on which random variables are defined
need not be a probability measure; that is, P{Q) need not be 1! Whenever
P may not be a probability measure, this fact will always be mentioned,
however.

Substochastic Transition Functions

If p is substochastic, the state space (X. fl can be enlarged (Appendix VI.I )


to a state space (X,,X-) by adjoining an absorbing state a, a "trap", to
obtain a stochastic transition function extending p to (Xe', do). The adjoined
point a is absorbing in the sense that if a sample path reaches the point, the
path stays there from then on. The first time Tthat a is reached is an optional
time, the "lifetime" of the process, equal to + oo for a path that never
reaches a. If convenient, the state a can now be dropped, so that to any
initial distribution corresponds a process for which

P{x(n) is defined} = P{x(n)eX} < I.

In discussing processes with substochastic transition functions it is always


to be understood in writing x(n) when the trap point has been dropped that
4. Application of Martingale Theory to Discrete Parameter Markov Processes 547

there is an unwritten convention limiting the context to the set on which


x(n) is defined, the set on which n is strictly less than the process lifetime.

Theorem (Strong Markov Property in the Discrete Parameter Context). Let


{x(n), .f(n), n e7L+ } be a Markov process with stationary transition function
p, and let S be afinite-valued optional time. Then {x(S + n), .'F(S + n), n e IL+ }
is a Markov process with transition function p and initial distribution the
distribution of x(S).

Since S + n is optional as well as S, all that must be proved is that if


A e. ", then

P{x(S+ l)eAJ.F(S)} =p(x(S),A) as. (3.4)

Since {S = n} e.F(S), the left side of (3.4) is equal almost everywhere on


{S = n} to P{x(n + 1) e A jx(n) }, which itself is almost surely p(x(n), A), as
was to be proved.
If S is identically constant in (3.4), this equation is equivalent to the
Markov property of as stated in terms of a transition function.
The fact that this relation holds when S is optional is called the strong
Markov property.

Extension of Theorem 3

A slightly more general version of (3.4) is obtained by observing that the


proof of (3.4) also shows more generally that if S is not necessarily finite
valued, then (3.4) is true almost everywhere on the set {S < + oo }. Still more
generally, if p is substochastic, if has lifetime S', if A C- X, and if S < S'
almost surely, then the proof of (3.4) shows that this relation is true almost
everywhere on the set {S < S'}. A still more general version of Theorem 3
in which S + I is replaced by a random variable measurable relative to .F(S)
is useful in the continuous parameter context and in that context will be
proved in Section 9. The reader is invited to formulate and prove the counter-
part of Theorem 9 in the present context.

4. Application of Martingale Theory to Discrete Parameter


Markov Processes
Let (X,.") be a measurable state space, and let {x(n),ne7L+} be a Markov
process with this state space, from a point of X. It is supposed that the
process has stationary transition probabilities, with a transition function p.
Call a function u from X into A superharmonic [harmonic] [ [subharmonic] ]
relative to p if u is measurable, if
. I uI) < + oo, and if u) < u
548 2.V1. Markov Processes

[p(-, u) = u] [ [p(-, u) Z u] ]. A trivial computation shows that the composed


process {u[x(n)],n >_ k) is a supermartingale [martingale] [[submar-
tingale]] relative to with ,fi(n) _ 3F , . . . , x } if u is superharmonic
[harmonic] [[subharmonic]] and if E{ju[x(n)]j) is finite for nzk.
Observe that this expectation is finite when n = 0 if a is finite at the initial
point of the process and that this expectation is finite by hypothesis when
n = 1, and therefore if u is superharmonic and positive, this expectation
will be finite when n > 1. According to the martingale theory convergence
Theorem 111.13, we conclude that u[x(n)] exists and is finite almost
surely if u is positive and superharmonic.

EXAMPLE (Boundary limit theorems in potential theory). Let D be an open


Greenian subset of RN, and for each point of D let be an open rela-
tively compact subset of D containing . Define

A) = MDISI(b A).

Here µD14, is harmonic measure; so this measure is supported by It is


supposed that depends on in way that A) is Borel measur-
able when A is a Bore] subset of D. The classical superharmonic, harmonic,
subharmonic functions are, respectively, superharmonic, harmonic, sub-
harmonic relative to this transition function. The same procedure in the
context of parabolic potential theory makes the superparabolic, parabolic,
subparabolic functions, respectively, superharmonic, harmonic, subharmon-
ic relative to a transition function defined in terms of parabolic measure.
Observe that in this case the sample sequences move in the direction of
decreasing ordinate values.

We return to the classical context, and to simplify the situation we make


the hypothesis that D is bounded. We shall choose in two ways. First
let D. be an increasing sequence of open relatively compact subsets of D
with union D and with D c D.,,. Define Dk for k = min {n: E
Then p as defined above is a stochastic transition function, and the corres-
ponding Markov process from a point x(0) of D determines sample
sequences which run through successive boundaries of D. toward 8D.
Furthermore each coordinate function u of RN is harmonic and bounded on
D; so is a bounded martingale, and it follows that u[x(n)]
exists almost surely. Hence x(n) exists almost surely; that is, almost
every sequence path converges to a boundary point of D. Next we apply
this technique to a different choice of This time let e be a strictly
positive number, and choose C A (] - 8DI/2)), so that
µa4,0, -), that is, is the uniform probability distribution on To
prove that with this choice of function A) is Borel measurable when A
is a Borel subset of D, it is sufficient to prove that f) is continuous when
.f is continuous and bounded on D, and this continuity of f) follows from
4. Application of Martingale Theory to Discrete Parameter Markov Processes 549

the fact that is the unweighted average off on and that the
radius of varies continuously with . With this choice of the fact
that x(oo) exists almost surely is proved as before, and the limit
must be on OD because

Ix(n) 2 aDI
I x(n + 1) - x(n)I = e n a.s.

The distribution of x(oo) is a measure pD[x(0), ] on OD. It will be shown


later that for both choices of this measure is the harmonic measure
PD[x(0), ], and in fact we shall show much more than this, but now we
verify this assertion only for aD regular. Let 0 be a continuous function
from OD into R. Then H,, is a bounded harmonic function on D with bound-
ary limit function 0; so is a bounded martingale closed (Theorem
111. 14) by its almost sure limit H,[x(oo)]. Hence

MD[x(o), O] = E{o[x(oo)] { = Hm[x(0)] = uD[x(0). m],

and since 0 is arbitrary, µD = µD, as asserted.


For both choices of if u is a positive superharmonic function on D,
the composed process is a positive supermartingale except that the
parameter value 0 should be omitted if u[x(0)] = +eo. Thus according to
the martingale theory convergence Theorem 111.1 3, a positive superharmonic
function has a finite limit on almost every path to the boundary; this is
a Fatou-type boundary limit theorem. Observe that for the second choice of
when a is small, the paths are sequences whose successive points
are close together, and this fact suggests that the above results are valid for
suitable continuous paths instead of sequences. This is true: what we have
proved is valid when properly interpreted for Brownian motion continuous
paths from a point of D to the boundary. In fact for both of the above
choices of D(5) an sequence can be identified with a sequence of points
tending to a boundary point of D along a Brownian motion path to the
point. The probabilistic evaluation of harmonic measure as a Brownian
motion hitting distribution, the boundary limit theorem for positive super-
harmonic functions, and the supermartingale significance of the composition
of a positive supermartingale with Brownian motion are parts of Theorem
IX.7. The identification of harmonic measure with a Brownian motion
hitting distribution is made in Theorem IX.13(a). The latter identification
carries with it the fact that the hitting distribution on is µD(4)0, )
which implies that for each choice of D(4) above a sequence with the
assigned distribution can be obtained by choosing x(1) as the first hitting
place on (D(x(0)) by a Brownian motion path from x(0), x(2) as the first
hitting place thereafter on (D(x(1)) and so on.
In the parabolic context the discussion is changed only in that "harmonic"
is replaced by "parabolic," and "Brownian motion" by "space-time
550 2.VI. Markov Processes

Brownian motion." If h [h] is a strictly positive harmonic [parabolic] func-


tion, harmonic [parabolic] measure can be replaced by h-harmonic [h-
parabolic] measure throughout the discussion to derive corresponding
results in relative contexts. The continuous path processes for which the
results are valid are the conditional [space-time] Brownian motion processes
discussed in Chapter X, where the counterparts of the classical context
results are proved.

5. Continuous Parameter Markov Processes with Stationary


Transition Probabilities
Let (X, I) be a measurable space, let P be a probability measure on X, and
let (t, , A) "p(1, g, A) be a stochastic continuous parameter transition func-
tion (Appendix VI.3). Let {x(t),F(1),te98+} be a Markov process with
initial distribution u and stationary transition probabilities, specifically with
transition function p (Section 1), so that p satisfies the Chapman-Kolmo-
gorov equation : for A e X, s > 0, t > 0,

p(s + t, , A) = J P(1, 7, A)P(s, , d01), (5.1)


X

andfors-0,1>0,
Pµ{x(0)EA}=µ(A), a.s. (5.2)

We shall describe this Markov process from now on as a Markov process

AA
with stationary transition function p. If 0 = to < .. < 1, and A;E3-,

EA,,j <(n'}

=J o) I P(t1odi) ... J(z_ (5.3)

and in particular,

PS(x(t) EAj,(j<n}

1A ()J ... j(iff- (5.3')

The evaluation (5.3) determines the finite-dimensional distributions of


and thereby determines P on 9 {x(t), to 68+}. Observe that the left side of
(5.3') defines an 1 measurable function of whose integral with respect to
N is the left side of (5.3).
Throughout the rest of this section the measurable state space will be
5. Continuous Parameter Markov Processes with Stationary Transition Probabilities 551

Polish, coupled with its Borel subsets. Given a Polish state space (X, X), a
probability measure y on 37, and a stochastic transition function p on (X. X ),
there is a canonical Markov process with state space (X, X), initial distribu-
tion p, and transition function p. We sketch the definition of this process
(cf. Section 1.10). The space 6 is the space of functions w from R' into X,
and .'(t, (b) is the value of w at t. Define A(t) = F{z(s), sy,5 t}, A =
The measure P. on A is determined by defining Yµ{z(tf)eA;,
j 5 n} by the right side of (5.3), and this definition is extended to .A using
Kolmo$orov's theorem. The measure P. is then completed, thereby enlarging
to ,#,,, and finally ,.(t) is defined as the cr algebra generated by .1(t) and
.

the IN null sets. The process is the desired canonical process on


the probability space Observe that if is a Markov
process with this state space (X, X) and transition function p, on a probability
space (S2µ, Pµ), and if ¢ maps S2, into h by 0(cu) = w = w), then in
view of our convention that process probability measures are complete,

4-t(1) c ,F, 4-1 (1(t)) c .4=(t),

and in addition ¢ -t (.µ(t)) c .F(t) if.4t(0) contains the Pµ null sets. Moreover

Pµ{c-I(A)) _ Pµ{R} (ReA,,). (5.4)

These relations all follow from the fact that for t and A. as above, if R =
{z(t)eAt, '_-5n)' then and (5.4) is true for this set R.
If R e , the function is x measurable, that is, Borel measur-
able, and
/"µ{R} = f (R)µ(d ) (5.5)
X

because (5.3) is valid for µ, and therefore the assertion is true for in ,F{s(ti),
j:!5 n), hence true for A in'n' the algebra U {x(tt), j 5 n} (where the union
is over all finite parameter sets), and finally the assertion is true for A in
I. From the general theory (Appendix V1.2) we know that if for each the
measure P is restricted to the class &(.*) of universally measurable sets
over A or even to the possibly larger class IPr (.*) of universally measurable
sets over,* relative to the kernel A) -+ PA(R) on T x 1, then the function
P.{R} is universally measurable on X, and (5.5) remains true. In expectation
language the function is dr'[*(f)] measurable if z is the indicator
function of a set in I and therefore by the usual approximation
procedure if z is .A [c&! ( )] measurable and is positive or absolutely rµ
integrable, then `,{f} is .1[W(X)] measurable and

Cµ{x} = f Es{z}µ(d). (5.6)


x
552 2.V1. Markov Processes

Observe that (5.4) can now be interpreted as an integrated version of

PP,{4-'(A)Ix(0)} = P:(o,{R} a.s. (AeW¢(A)), (5.7)

and the right side of (1.1') can be expressed in the form It,(s) {z}.
In view of (5.4) and the corresponding equation for expectations the
relations (5.5) and (5.6) can be applied to noncanonical processes to deduce
that

jP{4)_I(A)}/2(d)> EN{4-'(. )} = fX
(5.8)

where A and z are as above and ¢-'(X) = z(4). Here PP is the probability
measure, and E{ is the corresponding expectation operator for an arbitrary
Markov process with state space (X, -T), transition function p, initial point
, on a probability space (52,, ..;, Pc), and the set S24 may vary with .

6. Specialization to Right Continuous Processes


Suppose in Section 5 that the transition function p has the property that for
each point in the Polish state space X there is a probability space ((h, .414, Pe)
on which there is a right continuous process with initial point and transition
function p. We can then define the associated right continuous process, in
which 6 is the class of right continuous functions from R+ into X, z(t), A,
.1(t), 46 are defined as in Section 5 (with this change in the definition of 15,
and P {R} for A in .A is defined by

P4(A) = P4 {4)-1(A)). (6.1)

The measure .6, is then completed, thereby enlarging .0 to A,. This definition
of implies that I satisfies (5.2) (with u in that equation supported by
{ } ), and it follows as in Section 5 that the function r-+ p{ {R) is X measur-
able when Re. A. The measure P,, defined by (5.4) on A makes
a right continuous Markov process with initial distribution p and transition
function p. We have thus proved that there is such a process for every U.
It is now clear that everything in Section 5 is true in the present context in
which the members of 15 are right continuous. What is new is that the
canonical processes are now right continuous, and this fact induces new
process properties. We shall derive properties of the hitting of analytic sets
in the present context, keeping the notation of Section 5 but using the fact
that the canonical processes are now necessarily progressively measurable.
Suppose that A is an analytic subset of X. We have shown in Section 1.9
that
6. Specialization to Right Continuous Processes 553

R = 6:1(s, 6) e A for some s 5 t} eA. (t). (6.2)

It now follows from (5.4) that for any almost surely right continuous
Markov process on a probability space P) with state space (X, d"),
initial distribution p, and transition function p the probability of hitting A
at some time 5 t, that is, the probability that the entry time T of A is at
most t, is P{T 5 t} = J6,{&-' (A)) so that this probability does not depend on
the choice of probability space. If µ is supported by {r;}, this probability
is f{&'(A)} and is universally measurable as a function of . Finally the
probability for general p is the integral The corre-
sponding remarks are valid for hitting and entry time distributions. We
conclude that such probabilities and expectations as the following define
universally measurable functions of the initial point .
(a) P,{sup,E,u[x(t)] > c} for I an interval and u a Borel measurable
function from X into A.
(b) E, (sup,. t u[x(r)] } if [notation of (a)] u -a 0 or if this expectation
with u replaced by Jul is finite for all .
(c) P{{limsup,tru[x(t)] > c} for u as in (a) and for T the hitting time
of an analytic subset of X by
(d) E E {lim sup,tT u[x(t)] } if [notation of (c)] u 2 O OT if this expectation
with u replaced by Jul is finite for all .
If for every point of the state space there is a continuous process with
initial point and the given transition function p, then all the work in this
section can also be carried through with Sl the space of continuous functions
from R+ into X.

The Hitting of an F. Set

We now change some of the notation in the right continuous context we are
treating. Let A be an analytic subset of X, let I be a compact parameter
interval, let be a right continuous Markov process from g with the
given transition function p, and recall that the space on which is defined
may depend on c. Consider the function

u--. P {x,(t) e A for some t in 11. (6.3)

According to the discussion we have given, the function (6.3) is universally


measurable. We shall now show, under further hypotheses imposed on A
and the transition function that the function (6.3) is even Borel measurable
and that the proof of this Borel measurability does not need the theory of
analytic sets. This fact means that some later results in this book can be
obtained without the (unpalatable to some readers) invocation of analytic
sets. Abstention from analytic set theory makes stochastic process theory
554 2 VI. Markov Processes

much less elegant and sadly incomplete, however. Observe that if is as


above and if !' is a countable subset of parameter values, the function

BHP{x,(t)eA for some tin I'} (6.4)

is Borel measurable. In particular, if A is open and if !' is a countable dense


subset of I including both endpoints, the functions in (6.3) and (6.4) are
identical. Now suppose that for each the process can be supposed
continuous. In this context we shall now show that the function (6.3) is
Borel measurable when A is an F, set. It is sufficient to prove this Borel
measurability when A is closed. Let B. be for n z I the open set of points
at distance < 1/n from A. Then A is hit by in the parameter interval I
if and only if B. is hit by in the parameter interval I for every n. Hence
the function (6.3) is the limit as n -+ o0 of (6.3) with A replaced by B.; so
the function (6.3) is Borel measurable.
Observe that if above is supposed merely almost surely right contin-
uous rather than right continuous, or almost surely continuous rather than
continuous, then the conclusions remain valid and the reasoning is un-
changed aside from the acknowledgement of exceptional null sets.

7. Continuous Parameter Markov Processes: Lifetimes and


Trap Points
Recall that a function defined on a subset A of a measurable space is said
to be measurable if A is measurable and if the function is measurable with
respect to the class of measurable subsets of A.
Let (t, , A) -p(t, , A) be a substochastic stationary transition function
with state space (X, .Y). The Chapman-Kolmogorov equation implies that
the function i i-+p(t, , X) is monotone decreasing on ]0, + oc [ for all c. We
assume throughout this discussion that lim,.o p(t, , X) = 1, which implies
that the function X) is right continuous on ]0, + x [ for all S. Gener-
alizing Section 5 slightly, define a Markov process with state space (X,. "),
parameter set R',

initial probability distribution µ, and stationary sub-


stochastic transition function p as an adapted process {x(t),.F (t), t e 08+ }
satisfying (for 0< s < t and A e 1)

P{x(0)eA} =µ(A), P{x(t)eAI F(s)} = p(t - s, x(s), A) a.s. (7.1)

The process random variables need not be defined on the whole space; so
"a.s." means "almost everywhere where x(s) is defined." In any expression
involving a random variable of the process the added condition that the
random variable is defined is to be understood and will sometimes be written
explicitly. Thus the set {x(t) e X } is the domain of definition of x(t). With
7. Continuous Parameter Markov Processes. Lifetimes and Trap Points 555

this convention (5.3) remains valid. According to (7.1) and the Chapman-
Kolmogorov equation, the random variable x(O) is defined almost surely,
and the function t c - P{x(t) e X), equal to Jxp(t, c, X)p(dc) when t > 0, is
monotone decreasing and right continuous on R+. If 0 < s < t, the random
variable x(s) is defined almost everywhere where x(t) is defined because

P{x(s)eX,x(t)eX}= `p(dc) J'(t


X x (7.2)

= f P{x(t)eX}.
X

If there is a positive random variable S (S + oo) such that x(t) is defined if


and only if t < S, the random variable S is called the process lifetime. Integral
parameter process lifetimes were discussed in Section 3. The existence of a
process lifetime is not much of a restriction. In fact without the hypothesis
of the existence of a process lifetime define S by

S((v) = sup {r: r is rational, x(r)eX}

so that S is almost surely strictly positive and

P{x(t)EX} = P{S z t} = P{S > t}.

Define y(t) on the set IS > t} [which differs from the domain of x(t) by
a null set] as x(t) and leave y(t) undefined otherwise. Then the process
{ is Markovian with transition function p, initial distribution p,
lifetime S and is a standard modification of
If is a Markov process with substochastic transition function
p and lifetime T, enlarge the state space by adjoining a single point a and
extend p to be stochastic on the enlarged space, making a an absorbing
point (Appendix VI.3). If now x(t) is defined as a fort > T, the extended
process is Markovian with transition function the extension of p. Thus
results for Markov processes with stochastic transition functions can be
applied to those with substochastic transition functions.
Going in the opposite direction, if p is a substochastic transition function
with Polish state space X and if p is a probability distribution on the state
space, there is a Markov process with transition function p and initial
distribution p according to the following argument. Adjoin a point a to X
as an isolated point of X u a, and extend p by making a an absorbing point.
There is then a Markov process with the extended state space,
the extended transition function, and the given initial distribution (supported
by X). If i(t) is replaced by its restriction x(t) to the set {i(t) e X }, the process
.0(-)) will have the properties described at the beginning of this section
and, if desired, can be modified to have a lifetime, as discussed above.
556 2.V1. Markov Processes

A a Algebra Equality for Almost Surely Continuous Markov Processes

We shall need the following intuitively obvious fact. Let be an almost


surely continuous Markov process from a point . It is supposed that the
state space is a Polish locally compact but not compact space. Let S be the
process lifetime. A trap point is adjoined as described above. Let { be the
adjoined point of the Alexandrov one-point compactification of the state
space. It is supposed that x(S-) = C almost surely. Fort in R+ let flt) be
the a algebra generated by the null sets of the x(-) probability space and
.F {x(s), s 5 t}, and define -9 = YE R. ,fi(t). Let B. bean increasing sequence
of open relatively compact subsets of the state space, containing , with
union the state space, and let S. be the hitting time of 8B1 by x(-). Then
S. = S almost surely. Define 9 = Y.Ez+ _*(S.). Then we show that
F =,#. The inclusion I c F is known from Section I.12. To prove the
reverse inclusion, we prove that x(t) is I measurable for each t. In doing
this it is no restriction to identify the process trap point with . Then we
need only observe that x(t A S.) is F(t A c 9(S.) c 9 measurable and
that x(t) = liM.-. x(t n almost surely.

8. Right Continuity of Markov Process Filtrations;


A Zero-One (0-1) Law

In some contexts the desirable right continuity of Markov process filtrations


is intrinsic, as illustrated by the following theorem.

Theorem. Let {x(-), bean almost surely right continuous Markov process
with Polish state space and with transition function p. Suppose that whenever
t > 0 and f is a bounded continuous function on the state space, the function
p(t, , f) is continuous, or at least the process p(t, x(-), f) is almost surely right
continuous. Then
(a) The process .F '(-))is Markovian with transition function p.
(b) If for all t z 0 the a algebra .fi(t) is generated by the null sets and
.F{x(s),s < t}, then

Observation. If the state space is locally compact and second countable,


it will be clear from the proof of the theorem that f in the hypotheses need
only be continuous with compact support.
The theorem and the following proof are valid even when p is substo-
chastic. For f as in the theorem the Markov property in the form

Elf [x(s' + t)] I.fls') } = p(t, x(s'), f) a.s. (t > 0) (8.1)

implies when s' I s > 0 sequentially that


9. Strong Markov Property 557

E{ f [x(s + t)]j.F+(s)} = p(t, x(s),j) a.s. (8.2)

in view of the hypotheses of the theorem and the dominated convergence


theorem for conditional expectations. Thus (a) is true. Under the added
hypothesis of (b), Lemma 2 implies that ,F+(t) a .fi(t); the reverse inclusion
is trivial ; so (b) is true.
A 0- 1 Law. Suppose under the hypotheses of (b) in the theorem that the
Markov process is one from an initial point. In this case the a algebra
F(0) = .F(0+) consists of the null sets and their complements. In intuitive
phraseology this evaluation of F(0+) states that every sample function
asymptotic property as the parameter value tends to 0 is either almost surely
true or almost surely false. For example, if T is the hitting time by of
an analytic subset of the state space, then P{T = 0} must be either 0 or I.
See Section VII.6 for a discussion of this 0-1 law and its application in the
context of Brownian motion.

9. Strong Markov Property


The following continuous parameter version of the strong Markov property
indicates possible strengthening of the discrete parameter version in Section
3, but we shall not need such a strengthening. Recall (from Sections II.1 and
11.2) that if T is optional for a filtration then .F+(T) _ ,F(T+) and
that if S is an .F(T) measurable random variable and if S z T, then S is
optional. Recall also that if is a process with parameter set R and state
space the extended reals, and if 0 5 S S + oo, then the notation x(S) refers
to the function defined on the set {S < + oo) as x(S) and defined as 0 else-
where. We also adopt the convention that if p is a continuous parameter
transition function, thenp(0, , A) = 1A for A a measurable state space subset.

Theorem. Let p be a continuous parameter substochastic transition function


with a Polish state space, and suppose that p has the following properties :
(a) If is a point of the state space, there is a right continuous Markov
process with transition function p and initial point .
(b) If f is a bounded continuous function on the state space (or merely
continuous with compact support if the state space is locally compact
and second countable) and if t > 0, the function p(t, -J) is continuous,
or at least the process { p(t, xr(s), f ), s e R + } is almost surely right
continuous for each process in (a).
Then if (x(-), .F(-)) is a right continuous Markov process with transition
function p, if S and T are optional times with S z T, if S is .F(T) measurable,
and if A is a Borel subset of the state space,

P{x(S)eAI -,F(T+)} =p(S- T,x(T),A) a.e. on {S< +oo}. (9.1)


558 2.V1. Markov Processes

In particular, when S = T + t with t a positive constant,

P{x(T+ t)eA13Z(T+)} = p(t,x(T),A) a.e. on IT < +x}. (9.2)

Observation (1). Condition (a) implies that for f as in (b) the function
f) is right continuous on 68' because if is chosen as in (a), the
function t i- p(t, S, f) = E{ f [x1(t)] } is right continuous.
Observation (2). If.fr(t) _ 9(T + t), the theorem implies that the process
{x(T + ), .F '(-))
T is a Markov process on { T < + oo } with transition func-
tion p and initial distribution the distribution of x(T). To see this, first note
that the process in question is adapted because (Section 11.3) the function
x(T + t) is 9(T + t) measurable, and second note that if to Z 0 and if the
pair (T, T + t) in (9.2) is replaced by (T + to, T + to + t), then (9.2) becomes
the Markov property for the translated process; that is,

P{x(T + to + t)eAI.IFT(to)} = p(t,x(T + to), A) a.e. on {T < +o }.


(9.3)

Observation (3). If S = t v Twith t a strictly positive constant, (9.1) yields

P{x(t)eAI.9 (T+)} = p(t - T,x(T),A) a.e. on {T< t}. (9.4)

Observation (4). Hypothesis (a) is not weakened if is supposed only


almost surely right continuous because then there is also a right continuous
process with the other stated properties. In the same vein the following is
an alternative conclusion to the theorem: if is an almost surely
right continuous Markov process with transition function p, if .F(0) contains
the null sets, if S and Tare optional times with S z T, if S is .F(T) measur-
able, and if A is a Borel subset of the state space, then (9.1) is true. In fact
under these hypotheses the process is indistinguishable from a
right continuous process also adapted to and the theorem can be
applied to the latter process.
In proving the theorem it is no restriction (Section 7) to assume that p
is stochastic. We can also assume without loss of generality that _'F(0) con-
tains the null sets because (Section 1.8) fl-) can be extended if necessary to
make this so. Furthermore it can be assumed that Jr(-) is right continuous
because (by Theorem 8) the process 9' is Markovian with transi-
tion function p.
It will be convenient to prove (9.1) in the equivalent form

E{ f[x(S)]J.-F(T)} = p(S- T, x(T),f) a.e. on IS < +oo} (9.5)

when .f is as described in (b).


9. Strong Markov Property

The Special Case (9.2)

If S = T + t, equation (9.2) will be proved by proving first that

E{.f[x([T]" + t)]I' ([T].)} =p(t,x([T]n),f) a.e. on {[T]" < +oo


(9.6)

That is we prove that for A in . ([T]")

J f[x([T]n+t)]dP= p(t,x([T]n),f)dP. (9.6')


Iv [Tln<+ao) J An([T1n<+ac[

It is sufficient to prove (9.6') with the integration set replaced by


A n { [T]" = j2-"}, a set in ,f(j2-"), and with this replacement [T]" can
be replaced in the integrands by j2-" so that (9.6') becomes an immediate
consequence of the Markov property of In particular, (9.6') is valid
when A is in F(T), and with this choice (9.6') becomes the integrated
version of (9.2) when n - oc.

The Special Case T = 0

If z;" is the indicator function of the set { [S]" = j2-"},

E{f[x([S]")]j. (0)} = E z,E{f[x(j2-")]jF(0)) (9.7)


po
= p([S]",x(0),f) a.e. on {S < +cc
When n -+ cc, the left side of (9.7) becomes E{ f [x(S)] I ,f(0)} almost surely
because is right continuous, and the right side has almost sure limit
p(S, x(0), f) in view of Observation (1). Thus (9.5) is true when T = 0.

General Case

Apply the first special case to find that the process {x(T + ), -,FT(-) I is a
right continuous Markov process on { T < + oo } with transition function p.
Next we apply the second special case. If T is finite valued, the fact that S
and T are optional for implies (according to Section 11.2(h), but the
notation there is different) that S - T is optional for .FT(-) = .F(T + ).
Since S - T is T(0) measurable, the second special case can be applied to
{x(T+ with the pair (0,S - T) of optional times, to show that
(9.5) is true. If T is not necessarily finite valued, observe that S A n is
.F(T A n) measurable for all n and apply (9.5), but with the pair (S A n,
560 2.VI. Markov Processes

T A n) of optional times, to find that for AeflT), so that

{SSn}=An{TSn}nIS 5n}e.f(n)n.9r(T)=,F(TAn),

j'f[x(S)] dP = JP(S - T, x(S),f) dP.

When n - oc this equation becomes the integrated version of (9.5).

10. Probabilistic Potential Theory; Excessive Functions


Let p be a continuous parameter substochastic transition function with
measurable state space (X, T). G. A. Hunt showed how a far-reaching
generalization of classical potential theory can be obtained by defining the
analogs of the classical concepts in terms of p. In this theory the role of
positive superharmonic functions is taken by the a-excessive functions. Here
a is a positive parameter, and an a-excessive function u is by definition a
measurable function from X into R+ which satisfies

(a) e-°'p(t, , u) 5 u,
(10.1)
(b) limp(t, , u) = u.
r-+0

An application of the Chapman-Kolmogorov equation shows that (10.1a)


alone implies that the function i r-- e-°'p(t, , u) is monotone decreasing, and
it then follows readily that the limit function lim,_o p(t, , u) is necessarily
a-excessive.

ExAMPLEs. The identically vanishing function is a-excessive for all at.


A positive constant function u = c satisfies (10.1a) for all a. Hence, if
a z 0 and if 0 < c < + oo, the function u is a-excessive if and only if
lim,_o p(t, , X) = 1, and if c = + co, the function u is a-excessive if and only
if lim,_o p(t, , X) > 0. If p is the transition function for Brownian motion in
a connected Greenian subset of R", it will be shown (Sections IX.6 and IX.8)
that a positive function is 0-excessive if and only if it is identically +oo or
is superharmonic. Corresponding results will be proved in Section IX.16
for a-excessive functions (a > 0).

The Hunt theory has been presented in varying degrees of generality. For
example, the theory includes classical potential theory on Greenian subsets
of R' with regular boundaries if the following hypotheses are imposed.
Hl. The state space is a locally compact Hausdorff space satisfying the
second separability condition, coupled with its Borel sets (but
10. Probabilistic Potential Theory; Excessive Functions 561

sometimes it is necessary for adequate generality to make 1 the


class of universally measurable sets over the Borel sets).
H2. The transformation -J) takes the Banach space of contin-
uous functions f on X with compact support into itself, and this
transformation has the identity as strong limit when t - 0.
We shall discuss various aspects of Hunt theory below, as needed. A
0-excessive function is called excessive. An a-excessive function is also
fl-excessive for fi > a. If pa(t, , ) = e `p(t, , ), then pQ is a substochastic
continuous parameter transition function, and a function is fl-excessive
relative to pa if and only if the function is (a + fl)-excessive relative top. If
p is made stochastic by adjunction of a trap point to the state space, as
described in Section 7 an a-excessive function before the trap is adjoined
remains so if defined as 0 at the trap point. A simple argument shows that
the limit of an increasing sequence of a-excessive functions is a-excessive.
If it is supposed that p(-, , A) is R(R') x X measurable when A is in X,
then the role of a Green function is taken by the kernel Gz with state space
(X, X) and positive parameter a, defined by

A) = f"O e A)!1(dt). (10.2)

If f is a positive measurable function on X its a -potential is the measurable


function on X defined by

jetP(tf)li(dO. (10.3)
0

This a-potential is a-excessive because

J e°'p(s,ff)lt(ds) 5 (10.4)

and there is equality in the limit when t - 0.


It will be shown in Section IX.17 that if the state space is a Greenian
subset D of RN coupled with its Borel subsets and if p is the transition
function of the Brownian motion process in D, then

d7) = const 7)1N(d7)

Thus the Hunt potential of a function f becomes the classical potential of


the measure f dlN up to a multiplicative constant. A corresponding result
will be proved relating parabolic potential theory will space-time Brownian
motion.
The analysis of a-excessive functions and a-potentials for the transition
562 2.VI. Markov Processes

function p is the same as the analysis of excessive functions and 0-potentials


for the transition function p,.

Invariant Excessive Functions

If u is not only a-excessive but

e-°'p(l, , u) = u (10.5)

for all t > 0, then u is called invariant a-excessive. For example, if v is a-


excessive, the function lim, a "p(t, , v) is invariant a-excessive. If u is
a-excessive and if (10.5) is true for some strictly positive t, then u is invariant
a-excessive. In fact on the one hand (10.5) is true for 21 when true for i
because if true for t, then

px(2t, , u) = pa(t, .pa(t, , u)) =pa(t, , u) = u,

and on the other hand by monotoneity of the function u) equation


(lb.5) is true for t < .s if true for s.
The concept of an a-excessive function links Hunt potential theory with
martingale theory because the condition (10.1a) is precisely the condition
that if is a Markov process with transition function p. then the
process {e-a`u[x(t)], ,fi(t), t e l `} is a supermartingale on the parameter set
for which the process random variables have finite expectations. The condi-
tion (10.5) for an invariant a-excessive function u is the condition that this
supermartingale be a martingale on this parameter set.
The following lemma will be useful.

Lemma. If a is a-excessive and if


1/n
u=n ' e-'sp(.s, , u A n)!1(ds), n e Z',
o
then the sequence u is an increasing sequence of bounded a-excessive functions,
with limit u.

At the expense of replacing p by p, we can assume that x = 0. Obviously


un<UAnand
In
p(t, , un) = n p(s + t, , u A n)11(ds). (10.6)
0

It is trivial that u A n satisfies (10. la); so the function u A n) is mono-


tone decreasing and furthermore (10.6) implies that this function is contin-
10. Probabilistic Potential Theory; Excessive Functions 563

uous. Hence the evaluation (10.6) implies that un is excessive. Moreover the
monotoneity of u A m) implies that the sequence
1(n

n i-s n p(s, , u A m)11(ds)


0

is monotone increasing for fixed m; so the sequence u, is a monotone in-


creasing sequence. Finally,
un
uz limunzlimn Am)!1(ds)=u Am
n-ao n-m o

for all m; so u, has limit u as desired.

Excessive Measures

The dual of an a-excessive function for a transition function p with state


space (X, T) is an a-excessive measure, defined as a measure on which
satisfies, for A e T,

(a) a-2' I p(t, , A) u(dd) µ(A),


x

(b) lim p(t, f, Aµ(dd) = p(A)


1-0
f
X

Condition (b) is not usually imposed because in most contexts it follows


from (a). Further conditions, such as (if X is topological) finiteness of p
on compact sets, are usually imposed. An application of the Chapman-
Kolmogorov equation shows that (10.7)(a) alone implies that the function
on the left side of (10.7)(a) defines a decreasing function of t. The limit
measure when t -s 0 necessarily satisfies both conditions in (10.7). A 0-
excessive measure is called excessive. If u is not only a-excessive but satisfies,
for A e fit",

e-a, {,
p(t, , A)µ(dd) = p(A) (10.8)
x

for all t > 0, then p is called invariant a-excessive.

The Choice of State Space

In the most common Markov process setting the state space for the transition
function p is a topological space coupled with some a algebra of subsets
564 2.VI. Markov Processes

including the Borel subsets. For example, the state space of Brownian motion
in an open subset D of R" can be taken as the a algebra of Bores subsets
of D or perhaps more naturally as the a algebra of 1. measurable subsets of
D (the last because the transition distributions are absolutely continuous
relative to 1"). At first glance it may seem that whatever the context, it makes
no difference which of the possible a algebras is chosen. But in fact the choice
is involved in the definition of excessive functions and measures. In the
Brownian motion case it makes no difference in the sense that (Sections
IX.6, 8, 18) whichever choice is made of the three possibilities mentioned,
a function u is excessive if and only if on each connected component of D
this function is either identically + oo or positive and superharmonic, and
a measure is excessive if and only if it is the indefinite integral of an excessive
function. For space-time Brownian motion in an open subset b of h",
however, the situation is different. It will be seen (Section IX.18) that
coupling 6 with its a algebra of Borel subsets is less suitable than coupling
with the much larger a algebra of those subsets of 15 which meet every
hyperplane orthogonal to the ordinate axis in a Bore) set. (Replacing "Borel"
here by "1" measurable" would not change the classes of excessive functions
and measures.)
When the context is not otherwise unambiguous, we shall use the language
".' measurable excessive function" and "excessive measure on " to indi-
cate the choice of X.

11. Excessive Functions and Supermartingales


In probabilistic potential theory the state space is usually topological, and
enough hypotheses (for example, H I and H2 in Section 10) are imposed on
the given state space and transition function to ensure the following.
(a) For any initial distribution there is a right continuous Markov
process on some probability space, with the specified
initial distribution and transition function.
(b) If u is a-excessive, the process is almost surely right
continuous.
If u is a bounded a-excessive function, the process {e-°`u[x(t)],.F(t),
t e R') is therefore an almost surely right continuous supermartingale, under
the convention that u[x(t)] is defined as 0 at times at least equal to the
process lifetime. Since (Lemma 10) every a-excessive function u is the limit
of an increasing sequence of bounded a-excessive functions, Theorem IV.4
implies that for any a-excessive u the process is almost surely right
continuous with left limits. The process {e`u[x(t)], 5(t), teR } is a
supermartingale on the parameter set for which the process random variables
have finite expectations, and if the expectation is finite at time tr,, the process
is a supermartingale on the parameter interval [to, + co[.
12. Excessive Functions and the Hitting Times of Analytic Sets 565

12. Excessive Functions and the Hitting Times of Analytic


Sets (Notation and Hypotheses of Section 11)
Let p be a substochastic transition function, let A be an analytic subset of
the (Polish) state space, and let be an almost surely right
continuous Markov process with transition function p and initial point .
The probability space on which the process is defined may depend on .
Let T{ [Tj be the hitting time of A by x,(-), [x,(t + )]. Then lim,_o T,4 = TS
almost surely. Let a e R+, let u be an a-excessive function on the state space,
and define

E{e''Ttu[x{(T{)]}, (12.1)

with the convention that the integrand is 0 where TT = + oo. According to


Section 6, the function v is universally measurable. If u is bounded, the
process {e-"u[x4(t)], t e R+ } is a supermartingale, and the supermar-
tingale inequality (Theorem IV.3) at times 0, TT, t + T,, yields

z T,{)] }
= e-'"E{E{e-°T'tu[x,(t + T,4)]j.F(t)}} = e °`E{v[x,(t)]} (12.2)

= e`p(t, , v).
Moreover, when t - 0 the first expectation tends to v is a-excessive
and is majorized by u when u is bounded. If u is not bounded, u is the limit
of an increasing sequence of bounded a-excessive functions (Lemma 10);
so v is the limit of an increasing sequence of bounded a-excessive functions
majorized by u. Hence v is a-excessive and is majorized by u.
If a = 0 and u =_ 1, the value is the probability that an path
hits A at a strictly positive time, and p(t, , v) is the probability that an
path hits A at some time > t. Thus the condition that v be excessive, namely
that p(t, , v) increase to as t decreases to 0, has a simple probabilistic
interpretation in this special case.
Suppose that a > 0, and define the substochastic transition function pa
by p (t, , A) = e-"p(t, , A). Suppose that u =_ I so that E{e-aT }.
Let S be a positive random variable defined on the same probability space
as x,(-), independent of with distribution density (relative to /t )
t - ae-" on R. Then the probability that an path hits A at a strictly
positive time but before time S is Equivalently, v(D is the probability
that a path from of an almost surely right continuous Markov process
with transition function pa hits A at a strictly positive time.
566 2.V1 Markov Processes

13. Conditioned Markov Processes


Let (X, X) be a measurable space, let p be a stationary substochastic transi-
tion function with state space (X, X), and let h be a strictly positive excessive
function for p. The hypothesis of strict positivity is unnecessary but avoids
some technicalities and is satisfied in the applications to be made. Define
X"={h<+oo}and

" P(t, , if SEX",


(13.1)
P (t, > d4)
10 if,EX-X".
Inequality (10.la) for excessive functions implies

(13.2)

and also implies that p" is a stationary substochastic transition function


with state space (X, X). Moreover, if v is excessive for p, the function v/h
redefined as 0 on X - X" is excessive for p". Conversely, if u is excessive
for p", then u = 0 on X - X", and the function vo defined as uh on X" and
+oo otherwise satisfies the excessive function inequality (10.1a) relative to
p. If v is defined as lim,.,o p(t, , vo), then v is excessive forp, and u = v/h on X".
Let be a Markov process from e X" with transition function
p, lifetime SS, .F,(s) = F {x;(r), r g s}, on a probability space (S2. F, P).
Choose t > 0, and define a measure P" on A Q) by

P"{A} =
E{h[ (j]; A}
for Ae.94(t). (13.3)

Here as before we adopt the convention that the integral of a function over
a set means the integral over the part of the set on which the function is
defined. Thus in (13.3) without this convention A would be replaced A n
(Sc > t}. In particular,

P"{x4(t)eX} = P"{x{(t)c- X"} =p(t,( ,h). (13.4)


h

A Markov process with transition function p" is called an h -path process.


If is an h-path process from , with lifetime S{, the right hand term in
(13.4) is the probability that Sr z t. More generally, if 0:!9 t 1 < < t = t,
the P" distribution of x,(t,), ... is the distribution of x2(t,), ... ,
In particular, if X is a topological space and if is almost surely
[right] continuous under P, this discussion shows that there is a Markov
process on X from , with transition function p" which is almost surely [right]
continuous strictly before time t (on paths for which t is less than the path
14. Tied Down Markov Processes 567

lifetime), and it follows easily that there is an almost surely [right J continuous
process with initial point and transition function p'.
In probabilistic potential theory hypotheses are imposed on the state
space and transition function p ensuring that for any in the state space there
is a right continuous Markov process from with transition function p
and that for such a process the composed process is almost surely
right continuous whenever u is excessive. In this theory it then follows as
sketched above that h-path processes x'(-) from points of finiteness of h can
be chosen to be right continuous. Moreover, if u is excessive for p, the fact
that is almost surely right continuous and therefore almost surely
right continuous for the measure P" suggests that is also almost
surely right continuous. Finally the fact that P"{xr(t)eX - X"} = 0 sug-
gests that almost no h-path from ever hits X - X". Proofs of these state-
ments and conjectures will be given in more detail in the contexts of Brownian
motion and space-time Brownian motion.

14. Tied Down Markov Processes


Let p be a substochastic transition function with state space (X,.f ). Suppose
that there is a measure I on :1 with respect to which p(s, , ) for s > 0 is
absolutely continuous with strictly positive density p'(s, , ) and that p'
satisfies (identically) the Chapman('-Kolmogorov equation for densities:

p'(s + t,, C) = J P (s, , q;P (t, q, (14.1)


X

for and q in X. The hypothesis of strict positivity of p' will be satisfied in the
applications to be made and avoids technicalities. This hypothesis implies
that if h is excessive for p, then 11h = + oo } = 0 unless h is identically + oo.
Choose t > 0, and let , q be any (not necessarily distinct) points of X.
If is a Markov process with initial point and transition function p,
if we consider in X, that is, before transition to a trap point, and if
0 < s, < ... < s < t, the formally calculated joint density of x,(s,) _
C ... , C. given that xs(t) = q is

P'(si, C,)P'(sz - s1.C1,C2) ... p'(s - (t - S.,C.,17) (14.2)


P"(t, 4, q)

This joint density can be interpreted as that of a Markov process on the


parameter interval [0, t[ with initial value and nonstationary transition
density (from C, at time s, to C2 at time s2 with 0 5 s, < sz < t)

p'(sz -s,.Ci,Cz)P,(t- s,,C,,17), (14.3)


568 2 Vt. Markov Processes

The .joint density (14.2) is also that of a Markov process reversed in time on
the parameter interval ]0, t] with initial value q at time t and nonstationary
transition density (from S2 at time .s2 to C, at time s, with 0 < s, < s2 < t)

P(s2-s1,s1 C2) (14.4)


p'(s2, 4, (2)

Probabilities on [0, t] can be obtained from probabilities by


integrating out q using as measure the distribution p(t, , ) of xti(t). If h is an
excessive function for p and is strictly positive, replacing p by p" corresponds
to replacing p'(s, (,, 42) by p'(s, 4,, 42)h(C2)/h(C1). but the densities in (14.3)
and (14.4) are unchanged. Thus the probabilities for an h-path process
on [0, t] for 4 in X' can be obtained from probabilities by integrating
out t using as measure the distribution p"(t, g, ) of
This discussion suggests that many h-path properties, that is, properties
of an h-path process are independent of h. For example, if has
continuous sample functions, it is plausible that for I almost every q the
process fixed at q at time t can be chosen to have continuous sample functions,
and therefore (as in fact already proved in Section 13) every h-path process
can be so chosen.

15. Killed Markov Processes

Let (t, 4, A) r. p(t, S. A) be a continuous parameter stationary transition func-


tion with Polish state space X, and suppose that for every point S of X there
is a right continuous Markov process with transition function p and
initial point c. Let A be an analytic subset of X, and let T. and L4 be, respec-
tively, the hitting time and last hitting time of A by The function h
defined by

h(5) = P;T4 < +,m fl = P(L4 > 0} (15.1)

is excessive (Section 12). If A is closed and if 4 e X - A, the process


killed at time T., that is, the restriction of to T4 > t }, is Markovian with
state space X - A and transition function p, given by

p,(r,q,B)=P,x,,(t)eB,Try> t;. (15.2)

Next suppose that A is not necessarily closed but that h is strictly positive,
and define p"(t, S, dq) = p(t, 4, dq)h(q)/h(S). Consider killed at L, except
that the parameter value 0 is dropped; that is, consider x 4C defined for each
t > 0 as the restriction to ; L4 > t} of If 0 < t, < .. < t.,
15. Killed Markov Processes 569

P{x4L,(tf)Edgt,j 5 n) = P{xr(t1)Edryt,j 5 n;LS > (15.3)


= p(t , dtl,) ... p(t - nn-1,

This evaluation has a simple interpretation: is Markovian with

P{xCL,(t) E drt} = p(t, , dry)h(n) (15.4)

and stationary transition function p". That is, probabilities can be


computed by computing probabilities for a Markov process from , with
transition function p", and then multiplying the probabilities obtained by the
constant Thus killing a Markov process at the last hitting time of a set
leads to a conditioned Markov process in the sense of Section 13. The process
killed at L, in the present context is trivially right continuous, verifying a
property of conditioned Markov processes derived in Section 13. Observe
that p" may be substochastic even if p is stochastic and that the distribution
of xeL;(t) may not be a probability measure:

P { xcL{(t) e X } = p(t, c, h) ; (15.5)

when t tends to 0, the right-hand side tends (increasing) to the limit h().
Chapter VII

Brownian Motion

1. Processes with Independent Increments and State Space 9


(a) Discrete Parameter Case

Let y(0), y(1), ... be mutually independent random variables with distribu-
tions qo, q1, ... , respectively, on R", and define x(n) = 2:oy(j), F(n) =
.F {y(0), ... , y(n)} = 9 {x(0), ... ,x(n)}. If the transition function p with
state space R" is defined by A) = (A - ), where A is a Borel subset
of R" and A - is the translation of A by the process is
Markovian with initial distribution qo and successive transition functions
po, p...... In fact according to the application in Section 1.7,
P{x(n + 1)eAJ.f(n)) = P{x(n) + y(n + 1)eAJ.fln)}
= x(n)) = A) a.s.

If 0 5 n1 < < nk, the increments x(n2) - x(n1), ... , x(nk) - x(nk_ 1) are
independent random variables, and these increments are independent of x(0).
If yo, yl , ... is an arbitrary sequence of random variables with values in
R" and if x _ Yo y,, the joint distribution of yo, y, , ... determines that of
x0, x1, ... , and conversely. Hence two different distributions of yo, y,, .. .
do not induce the same distribution of xo, x, . .... In the context of the
preceding paragraph this one-to-one correspondence between distributions
implies that if, for n Z 0, q is a probability measure on R' and if p is defined
by qn+1(A - ), then any Markov process determined by the
initial distribution qo and these transition functions is a sequence of succes-
sive sums of independent random variables, x = Yo y,, for which yy has
distribution q;.

(b) Continuous Parameter Case

Let {x(t), to R+ } be a stochastic process with state space R" and with
independent increments; that is, 0 S t1 < tk implies that the increments
1. Processes with Independent Increments and State Space 68^ 571

x(12) - X(tl), ... , X(Ik) - x(tk_1) are mutually independent. If s < t and if
q(s, t, ) is the distribution of x(t) - x(s), the fact that the distribution of the
sum of two independent random variables with state space R' is the convolu-
tion of their distributions implies that for 0 < s, < s2 < s3,

q(sl,s3,A)= J q(s2,s3,A-7)q(s,,.s2,dry). (1.2)


ggN

Suppose, strengthening slightly the condition of independent increments,


that 0 < t, < < tk implies that the random variables

x(0), x(11) - X(0), ... , X(tk) - X(tk-1) (1.3)

are mutually independent, so that x(0) is independent of the set of process


increments. If now .fi(t) = .F{x(s), s < t), the argument used in the discrete
parameter case shows that is a Markov process with initial
distribution that of x(0) and transition function given by

P{x(t)eAI .f(s)} = q(s, t,A - ) a. s. (1.4)

Equation (1.2) is the Chapman-Kolmogorov equation for the Markov


process transition function. This process is a very special type of Markov
process in that as is obvious from the discussion, or as follows readily from
(1.4), for selre the set of increments {x(t) - x(s): t > s} is independent of
the a algebra .ms(s). This independence is equivalent to the mutual indepen-
dence of the random variables (1.3) for all k and choices of t ... , tk.
Conversely, if is a Markov process with state space l ' and
(Section VI. 1) stochastic transition function (s, , t, A) i-+q(s, ; t, A) and if
this transition function can be written in the form (s, , t, A)'--p q(s, t, A - ),
where (1.2) is true, then the process has independent increments
in the strengthened sense that for se R+ the set of increments {x(t).- x(s):
t > s} is independent of F(s); moreover for 0 < s < t the increment x(t) -
x(s) has distribution q(s, t, ). Finally, if v is a probability measure on R(If8"),
if q(s, t, ) is a probability measure on R(R"), and if (1.2) is satisfied, then
there is (Section VI. 1) a canonical Markov process with initial distribution
v and transition function (s, , t, A) 1-o q(s, t, A - ).
We now summarize the preceding discussion in the special case of contin-
uous parameter processes with stationary independent increments. Let {x(t),
i e R } be a stochastic process with state space I8" for which 0 < t, < tk
implies that the random variables (1.3) are mutually independent. Suppose
also (increment stationarity) that for 0 < s, < s2 the distribution q(s,, s2, )
of x(s2) - x(s,) depends only on s2 - s, . If we denote this distribution by
q(s2 - S1, ), the family {q(s, ), s > 0} of probability distributions on 68"
satisfies the convolution equation
572 2.V11. Brownian Motion

q(s + r, A) = J A - )q(s, dd). (1.5)


N
gqN

If we define .F(t) _ . (x(s),s < t}, the process is a Markov


process with a stationary transition function given by

p(t, 5, A) = q(t. A - ), (1.6)

and for seR4 the set of increments {x(t) - x(s): t > s} is independent of
F(s). Equation (1.5) is the Chapman-Kolmogorov equation for the transi-
tion function p. Conversely, suppose that { x(t),.fi(t), t e t+ } is a Markov
process with stationary transition function (t, , A) i- p(t, 5, A) and suppose
that p has the form (1.6). Then the random variables (1.3) are mutually
independent, and for 0 < s < t the increment x(t) - x(s) has distribution
q(t - .s. ). Finally to any family {q(s, ), s > 0} of probability distributions
on RN satisfying (1.5) along with a further probability distribution v on R'
corresponds a Markov process with initial distribution v and
transition function (s, , t, A) i - q(t - s, A - ).

2. Brownian Motion
Define f(t,rt) by I.XV(4.1). but in discussing transition densities it will
sometimes be more intuitive to write ((t, , ry) for 4(1,17 - ). The variance
parameter a2 is fixed throughout. A stochastic process is called an
N-dimensional Brownian motion or a Brownian motion in RN if the following
four conditions are satisfied :
BM 1. The parameter set is R'. The state space is RN.
BM2. The process has stationary independent increments. If 0 < to, <
r + r, the distribution of the increment w(tt, + t) - w(to) has
density t(t, ) relative to ly.
BM3. The random variable w(0) is independent of the set of process
increments.
BM4. The process is almost surely continuous.
A Brownian motion in ' can also be defined as a Markov process
{ w(t), 9(t). t e R+ } satisfying the following three conditions:
BMI*. BMI.
BM2*. The (stationary) stochastic transition function has density
, (t, 5, q) = 15 (t, q - ) relative to /N.
BM3*. BM4.
Observe that BM 1 *- BM2* imply that for fixed s > 0 the set of increments
x(t) - x(s): r > s} is independent of .'(s).
The two Brownian motion definitions are equivalent in the following
2. Brownian Motion 573

sense. According to Section 1, if is a Markov process satisfying


BM 1 *-BM3*, then conditions BM 1-BM4 are also satisfied, and conversely,
if BM 1-BM4 are satisfied by a process and if F(t) is defined as F(w(s),
S:5 t), then satisfies BMI*-BM3*.
Throughout this book a "Brownian motion" specified as and
not otherwise delimited will mean a Markov process satisfying BM I *-BM3*
with no further unstated hypotheses imposed on the filtration
All sample functions of a Brownian motion can be made continuous
without affecting BMI-BM3 or BMI*-BM2* by redefining the discontin-
uous sample functions to make them continuous. Observe that if is a
Brownian motion from a point, then the Ncomponent processes are mutually
independent one-dimensional Brownian motions with the same variance
parameter as
If is a Brownian motion with variance parameter a2 and if to > 0,
the process {w(to + t), t e R+} is a Brownian motion with the same variance
parameter as and with initial distribution the distribution of w(to). The
process {w(t/a2), to R+} is a Brownian motion with variance parameter I;
the processes {[w(t) - w(0)]/a, to R+} and {w(t/a2) - w(0), t e R+} are both
Brownian motions from the origin with variance parameter 1.

Brownian Motion with the Parameter Set R

A Brownian motion with the parameter set R is defined as a process (w(t),


,fi(t), t e R} which satisfies BMI *-BM3* except that the parameter set is R
instead of R+ and that it is no longer supposed that the filtered measure
space on which the process is defined is a probability space or even that the
probability value assigned to the whole space is finite. More precisely it is
supposed that there is for each t in R a measure p(t, ) of Borel sets, finite for
compact sets, with 0 < p(t, R") 5 + oo, which is the distribution of w(t), and
probabilities are then defined as usual for a Markov process, in terms of p
and the transition density assigned in BM2*. Then for s e R, t > 0, A E R(R),

J P(s, di) I t (t, , P(s + t, A). (2.1)


R" JJA

Hence R") is identically some strictly positive constant fi (5 + oo). If s


is replaced in (2.1) by so - t, we find that ,a-"t'"'2 1"(A) z p(so, A), and
(t -+ + oo) it follows that fi = + oo. Thus the context imposes the condition
that the absolute probability distribution p(sqq, ) be an infinite-valued
measure for all so. If we defneti(ry, t) for (n, t) a R" by

ti(n, t) = J a(a, , n)P(t - a, dt) (2.2)


R"
574 2 Vii. Brownian Motion

with a > 0, it follows from (2.1) and the Chapman-Kolmogorov equation


that ti does not depend on the choice of a. The function ti(-, t) is a version
of the density dp(t, )/dIN of the w(t) distribution and

u(4, s)1N(d4) = + co, s),1(t, , n)1N(d4) = u(n, s + t) (2.3)


JaN RN

when t > 0. Since the function (n, PI) = 6(t, - tr) is parabolic on
9N - the function ti, which is Borel measurable and is finite valued
on a dense subset of ftN, has the parabolic average property and so is parabol-
ic on fit'. Under our hypotheses when seR the process {w(s + t), teR+} is
a Brownian motion with initial distribution of IN density s). The distribu-
tion of each increment x(t2) - x(t,) with t, # tZ is an infinite-valued
measure. Since the process is Markovian, the reversed time process is also
Markovian, and a version of the reverse transition density relative to IN is
easily calculated: the density of a transition from at time s to r at time
t<sis
ti(,i, t)G(s - t, 17, )
(2.4)

(Recall from Section I.XVI.8 that a positive parabolic function on f l' is


either strictly positive or identically zero.) In terms of ordinary probabilities
a version of the distribution of {w(so + t), t c- R') conditioned by w(so) = o
is that of a Brownian motion from go, and the distribution of {w(so - t),
t e R+ } under the same condition is that of a Markov process from o with
transition density determined by (2.4). The special case of stationarity is
important: if p(t,-) = IN for all t, that is, if ti = 1, the process forward and
backward transition densities are the same. In this case the processes
{ w(so + t), t e f+ } and { w(so - t), t E R+ } are both Brownian motions with
initial distribution 1N.

Existence of Brownian Motions (Parameter Set R+)

The conditions BM I -BM2 are conditions for a Markov process with state
space RN and with a specified transition function. Hence (Section VI.5) such
a process exists for an arbitrary choice of initial probability distribution µ,
for example, as a canonical process on the space of all functions from P+
into RN. A useful additional fact is that since the distribution of the process
in question is the integral with respect to p(dd) of the distribution of the
process with initial point , it follows that there is such a process when p is
an arbitrary measure on R', finite on compact sets, and not identically 0.
In Section 3 we shall show that any process satisfying has a
standard modification satisfying BM3*. Thus Brownian motions with
parameter set R+ exist.
2. Brownian Motion 575

Existence of Brownian Motions (Parameter Set R)

We show that if ti is an arbitrary not identically vanishing positive parabolic


function on <f8", then there is a Brownian motion with parameter interval R
having ti as its absolute probability density function. Observe first that u
satisfies (2.3) by direct calculation, using the representation of a in terms of
minimal parabolic functions given in Section 1.XVI.8. Observe secondly that
if we impose the condition w(0) = o in addition to the other conditions to
be satisfied, the processes {w(t), t E R+ } and { w(t), - t E R' } are to be
mutually independent Markov processes whose distributions we have
discussed above. Hence (Section VI.5) under the added condition at the
parameter value 0 a process {w(t), t E R) exists as a canonical process on the
space of all functions from R into R". The measure so defined on this
function space depends on the choice of o, and we integrate this measure
with respect to u( o, 0)!"(ddo) to get a measure on the function space, and
thereby a stochastic process, which has the desired properties aside from
sample function continuity. According to Section 3, we can obtain sample
function continuity by choosing an appropriate standard modification of the
process obtained in this way.

Extension of Brownian Motion for a Finite Interval


Suppose that b > 0, that I = [0, b], and that { w(t), t e 1) is a process which
satisfies the Brownian motion defining conditions BM I-BM4 and BM I -
BM3 except that R' is replaced by I as parameter set. Then the process
may not be the restriction to I of a Brownian motion process for the para-
meter set R+, but if 0 < a < b, we now exhibit a Brownian motion {w(t),
t e R') such that w'(t) = w(t) for t 5 a: choose any function f, strictly
increasing and continuous, mapping [a, + oo[ onto [a, b[, and define

w(t) if 05t5a,
w"(t) = w(a) + (t - a)U2[W [f(t)] - w(a)]
[f(t) - al 112
ifa<t<+ o0.

Space-Time Brownian Motion

Let o : so) be a point of E:8", and let be a Brownian motion in R"


from o. The process

{tG(t),tER'} = {[w(t),so - t],tER+}


with state space l' is called a space-time Brownian motion from So. This
process has stationary independent increments and is Markovian relative
to the same filtration as Space-time Brownian motion with an arbitrary
576 2.VII. Brownian Motion

initial distribution is defined in the obvious way and has these same proper-
ties. In this definition space-time Brownian motion moves downward in
4.8", that is, in the direction of decreasing ordinate values. The dual motion
moving upward will be called space-cotime Brownian motion. It will be
convenient to introduce the following notation. If A is a subset of A',
define A, = {q: (q, t) EA} and A, = A, x {t}. The transition probability func-
tion (r, s), A) F--, p(r, s), A) for space-time Brownian motion has the
property that p(r, s), ) is a measure supported by As_r, and this measure
when considered as a measure on A,_, is absolutely continuous relative to
IN, with density 8(r, , ). Define

(s - 1, , q), = S), 1 = (q, t)

so that f il) governs the transition density for time s - t when s > t,

p(r, , A) = (q, s - r))iv( ),


As-r

and C( , q) = 0 when s < t. This context suggests that an appropriate state


space for the transition function p is 98" coupled with the a algebra of those
sets meeting every hyperplane orthogonal to the ordinate axis in a Borel
set. (Nothing in the discussion would need any change, however, if "Borel"
here is replaced by "t,v measurable.")

3. Continuity of Brownian Paths

In this section it will be shown that a process satisfying conditions BMI-


BM3, equivalently, BMI* and BM2*, has a standard modification satis-
fying BM4. Thus Brownian motions exist. It is sufficient to treat the one-
dimensional case.
(a) If x is a Gaussian random variable with mean 0 and variance a 2
and if a > 0, then
z
P{x Z a} < a exp 2az (3.1)

In fact the left side of (3.1) is


-2z
(2n)-/2
o
exp 2 d <a
z Q

us
exp d, (3.2)

and the right side is equal to the right side of (3.1).


Recall that a random variable y is said to have a symmetric distribution
if P{y >- a} = P{ y < -a} for all a (equivalently, for all positive co. A finite
3. Continuity of Brownian Paths 577

sum of symmetrically distributed mutually independent random variables


is itself symmetrically distributed.
(b) If y(1), . .. , y(n) are mutually independent symmetrically distributed
random variables and if x(k) = E; y(i), then

P { max x(k) >- a } < 2P{x(n) >- a}. (3.3)


k5n

Let T= min { j: x(j) >- a}. Then the set {T=j} is independent of
x(n) - x(j), and since the latter random variable is symmetrically distributed,
n n
P{maxx(k)>-a}=>P{T=j} <2>P{T=j,x(n)-x(j)>-0)
k!gn 1 i
(3.4)
< 2P{x(n) > a}.

(c) If N = 1 and if satisfies the defining conditions BM I -BM3 of


a Brownian motion,

P (sup [x(r) - x(0)] >- at } < 2P{x(t) - x(0) >- a} (r rational).


r5I

(3.5)

According to (b) this inequality is true if the supremum on the left side
is replaced by the supremum for only finitely many values of r:5 t. The
inequality is therefore true as stated except possibly for the countably
many values of a which are discontinuities of the distribution function of
the supremum on the left. The inequality is true without exception because
both sides of (3.5) define left continuous functions of a.
By symmetry, if x(r) - x(0) is replaced in (3.5) by its absolute value,
the right side of (3.5) should be doubled.

Theorem. If is a process with the finite-dimensional distributions of a


Brownian motion, it has a standard modification with continuous sample
junctions.

We can assume in the proof that the variance parameter is 1 and that
N = 1. Applying (c) above, if m > 0, n > 0 and if r and s are rational,
mn-1
P{ sup Ix(s) - x(r) I > 2a J< Y P { s0<s-u/p Ix(s) - x(j/n)I >a l
0<s-r< 11n j=0 n52/re
S5m

= mnP { sup Ix(s) - x(0)I >- a J< 4mnP{x(2/n) - x(0) >- a} (3.6)
s 52/n

< 8mn'/Za-' exp -a2n


4
578 2.VII. Brownian Motion

Since the last term has limit 0 when n oo, the restriction to the rationals
of almost every sample function is uniformly continuous on [O,m[. In
other words the restriction to the rationals of co) for w not in some null
set coincides with a continuous function on R'. Since

IimE{[x(t) - x(s)]2} = 0,

x, is a standard modification of and has continuous sample functions


if is defined, say, to be identically 0 on the exceptional null w set.

Canonical Brownian Motion in I8"

According to our conventions a Brownian motion in R' is any stochastic


process satisfying certain distribution and continuity conditions. For some
purposes it is convenient to have a canonical Brownian motion, however,
for a specified initial distribution and variance parameter. Suppose then
that an initial distribution p and variance parameter a2 are specified. Let f2
be the space of all continuous functions tb from 68' into R', define v(t, w)
= co(t) and define .A = {w t), t e R'). Then (Section 1.10) a probability
measure p can be defined on . making a Brownian motion with initial
distribution p and variance parameter a2. The process will be described
as the canonical Brownian motion for the specified u and a2. In the language
of Section I.10 is the canonical process associated with every Brownian
motion with the specified initial distribution and variance parameter. In
accordance with our conventions, P is to be completed. The measure Y is
a measure on the space of continuous functions and is commonly described
as "Wiener measure" because Wiener was the first to construct this measure
rigorously. It is a matter of taste and convenience whether in discussing a
canonical Brownian motion from a point of R" we further restrict 6 by
the condition dh(0) = 5: the choice will be specified when it matters. We
shall not limit ourselves to canonical Brownian motions because noncanon-
ical ones arise in many contexts. The distributions of the interesting functions
defined on a Brownian motion with given initial distribution and variance
parameter do not depend on the choice of Brownian motion. For example,
the probability that a Brownian path from a point hits a specified analytic
set when the parameter varies in a specified interval may depend on the
variance parameter but (Section 1.12) not on the choice of Brownian motion.

4. Brownian Motion Filtrations


According to Section VI.9, a Brownian motion has the strong
Markov property if F(0) contains the null sets of the probability space, and
this condition is unnecessary if is continuous instead of merely almost
4. Brownian Motion Filtrations 579

surely continuous. If f, (t) is the a algebra generated by ,F(t) and the null
sets, the process is still a Brownian motion. Since s < s' < t
implies that w(t) - w(s') is independent of the random variable
w(t) - w(s) is independent of .F; (s), and this independence implies that
the process J w(-), () } is a Brownian motion. Thus there is no loss of
generality in assuming that for a Brownian motion F(-)) the a algebra
.5(0) contains the null sets of the probability space and fl-) = that
is, is right continuous. In particular, if (w(-), .F(-)) is a Brownian motion
with .fi(t) = .5{w(s), .s < t} and if .5,(t) is the a algebra generated by 5(t)
and the null sets, then { .IF, is a Brownian motion, and (by Theorem
VI.8) is right continuous.
In view of the properties of Brownian motions the entry and hitting times
of analytic subsets of the state space are optional for a Brownian motion
{ whenever. (0) contains the null sets of the probability space and
is right continuous. The space-time Brownian motions have the strong
Markov property and the preceding entry and hitting time property under
the stated conditions on the Brownian motions involved.

Theorem. Let .fl-)} be a Brownian motion in RN with ,fi(t) generated


by S{w(s), s >_ t} and the null sets. Define 5(+co) = YER fi(t).
(a) Every optional time is predictable.
(b) is predictable; in fact, if T is an increasing sequence of optional
times with limit T then .5(T) = Yo 5(7,).
(c) Every almost surely right continuous supermartingale is
nearly predictable, almost surely lower semicontinuous, almost surely
continuous if the process is a martingale.

Observation (1). In view of the martingale continuity properties proved


in Section IV. 1, assertion (c) for martingales is equivalent to the assertion
that every martingale {x(.), with the filtration determined by a
Brownian motion as stated in the theorem has an almost surely continuous
standard modification.
Observation (2). In (c) for supermartingales if the supermartingale is
right closable, Theorem IV.23(c) together with assertions (a) and (b) of
the present theorem implies that the sample function discontinuities
can be separated out to make a jump process.

Proof of (c) in the martingale case. A trivial argument shows that it is sufficient
to prove that if x is an integrable .5(+oo) measurable random variable,
then x(t) = E(xI,flt) } can be defined for t < + oo to make a continuous
process. In proving (s) we shall use repeatedly the fact that if x. is a sequence
of random variables with L' limit x and if (*) is true for x and its martingale
(x.(-), fl-)), then is true for x and In fact, going to a subsequence
if necessary, it can be assumed that E{Ix,,,., - x.1} < 2-" which implies
580 2.VI1. Brownian Motion

[submartingale maximal inequality applied to I x,,()I]

P{suplx"+l(t)-x"(t)I zn-2} <n22-"; (4.1)


r2

so (Bore] Cantelli theorem) for almost every w, lim, x"(t, co) exists
uniformly in t. The limit process is an almost surely continuous version
of and can then be modified to be continuous. To prove (*), let fo,
f, be continuous functions on lt' with limit 0 at + oo and define 0 by
0(a, r) = E{fl (z + r)}, for z an N-dimensional Gaussian random variable
whose components are mutually independent with mean 0 and variance a,
and for r in R. The function 4' is continuous. If 0 < t1 < t2 and if x =
fo[W(0)]f1[w(12) - w(tt)], then

fo[W(0 ma2(12 - ti), 0] if/ < tl


E{xl.f(t)} = fo[Hw(0)]4,[o2(t2 - t), w(t) - lv(tl)] if tl < t < t2 (4.2)

fo[W(0)]4,[0, w(12) - w(11)] = x if t > t2,

neglecting null sets. Thus (*) is true for this choice of x. More generally an
only slightly less trivial calculation shows that (*) is true if
"
X =fo[w(0)] F1 J[W(ti) - w(ti-r )]
1

where f, is continuous on R' with limit 0 at oo and 0 S to < .. < t". It


follows (Stone-Weierstrass theorem) that (*) is true for the random variable

x =f [W(0), w(tt) - w(to), ... , w(t") - w(t"-1)]

for f continuous on 08"'"+t) with limit 0 at oo and therefore if x is an


integrable random variable which is a Borel measurable function of w(0),
w(t 1) - w(to), ... , w(t") - equivalently, if x is an integrable random
variable which is a Borel measurable function of tv(0), iv(to), ... , w(t").
To finish the proof, we need only remark that according to Section 111. 14
every F(+ co) measurable and integrable function x can be approximated
arbitrarily closely in L' by random variables of this special type.

Proof of (a). To prove that an optional time T is predictable, it suffices to


prove that T A b is predictable for every positive constant b, and so it suffices
to prove that every bounded optional time T is predictable. Define

x(t)=E{T- TA 'I (')}=E(TI-flt)}-Tnt, 0St5+oo. (4.3)

The process is a supermartingale, and in view of what we have


just proved the conditional expectations in (4.3) can be chosen to make
5 Elementary Properties of the Brownian Transition Density and Brownian Motion 581

this supermartingale almost surely continuous. Moreover (Theorem IV.2)


x(T) = 0 almost surely. For n z I define T. = inf {t: x(t) 5 l/n} to obtain
an increasing sequence of optional times for which almost surely T.5 T,
and T. < Ton {T > 0} because 1/n almost everywhere on this set.
Furthermore, if T' = Tn, then almost everywhere on IT > 0}

0=x(T') =E{T- T A T'jF(T')};


so T = T' almost surely. Thus T, announces T if we ignore exceptional null
sets which can be eliminated by trivial modifications of T, .

Proof of (b). To prove that each set in F(T) is in Yo .F(T ), let x be the
indicator function of a set in .4=(T), and recall that we have proved above
that we can define x(t) = E{xI.F(t)} fort 5 +oo to make a continuous
process. Apply the conditional expectation continuity theorem and the
martingale optional sampling Theorem IV.2 to obtain

E{xlY.,F(T,)} = lim E{xl.F(T,)} = Iim x(T,) = x a.s.; (4.4)


p n-m n-M

so x is Yo .F(T,) measurable, as was to be proved.

Proof of (c) in the supermartingale case. According to Theorem IV.23.


assertion (c) is implied by (a) and (b).

5. Elementary Properties of the Brownian Transition Density


and Brownian Motion
(a) The transition density (t, , ry) - t (t, , rl) = 41(1, q - ) satisfies the
heat equation

and in fact G(s - t, , q) = s), (,, t)). This relation between Brownian
motion transition density and the parabolic Green function of ft N will be
extended to the corresponding relation between the transition density for
Brownian motion in an open set D and the parabolic Green function of
D x R in Section IX.17. As we shall see in cases related to the Dirichlet
problem for parabolic functions many Brownian motion probabilities can be
evaluated in terms of solutions of the heat equation on the relevant domains
with the appropriate boundary conditions or, if there is stationarity in time,
can be evaluated in terms of solutions of Laplace's equation with appropriate
boundary conditions.
582 2.VII. Brownian Motion

(b) A token of the intimate relation between Brownian motion and


classical potential theory is the fact that [see I .XVII(18.2)] to G(t, , ry)1, (dt)
is a constant multiple of the Green function of i8" when N > 2. This relation
between Brownian transition densities and Green functions will be extended
to transition densities for Brownian motions in Green domains in Section
IX. 17.
(c) A trivial calculation shows that if is a Brownian motion, then
lim,..,o P{I w(t)I > c} = I for every c > 0, that is, lim,. Iw(t)) = +oo in
measure.
(d) The part of almost every Brownian path corresponding to a specified
parameter interval has infinite length. To prove this, it is sufficient to show
that for one-dimensional Brownian motion from 0 with variance para-
meter l the part of almost every Brownian sample function corresponding to
the parameter interval [0, b] is of unbounded variation, for every b > 0. To
see this, define s = E' I w(jb/n) - w((j - 1)b/n)l, and note

{el -Iw
E"xp n

Since the distribution of n'12jw(6/n)j is independent of n, njw(b/n)I =


+ao in measure, so that lim,,. E{exp(-sp)} = 0, and since s increases
with n to the sample function total variation on [0, b] when n oo along
the integral powers of 2, almost every sample function has infinite total
variation on [0, b], as asserted.
(e) (N = 1) Let be a Brownian motion from the origin, with
variance parameter a2. Then E{w(s)w(t)) = a2(s A t) because if t > s,

E{w(s)w(t)) = E{ [w(t) - w(s)]w(s)} + E{w(s)2} = a2s. (5.3)

Conversely, if w(-) is a process with Gaussian finite-dimensional distributions


having zero means and if E{w(s)w(t)} = a2(s A t) for some strictly positive
constant a2, has the finite-dimensional distributions of a Brownian
motion from the origin with variance parameter a2.
(f) If is a Brownian motion from the origin, define

t > 0,
Y(t) =
0 ift=0.
An application of (e) to the one-dimensional component processes of
shows that has the same finite-dimensional distributions as More-
over the process sample functions are almost all continuous on ]0, + oo [,
and lim,.o y(t) = 0 a.s. if t -+ 0 along a countable dense set because has
this property. Hence this limit relation is true when t - 0 unrestrictedly. Thus
is Brownian motion, and the relation between and shows that
6. The Zero-One Law for Brownian Motion 583

properties of Brownian motion for large parameter values can be translated


into properties for small parameter values and conversely. For example,
continuity of Brownian motion at 0 implies that lim,.o tw(1 /t) = 0 almost
surely, that is,

w(t)
lim = 0 a.s. (5.4)
t-.m I

This result is a version of the strong law of large numbers in the continuous
parameter context, as applied to Brownian motion.

6. The Zero-One Law for Brownian Motion


If the sets of a a algebra of measurable sets of a measure space are all null
sets or the complements of null sets the a algebra is called trivial.

Theorem. If is a Brownian motion from a point, the a algebras 9, and F2


generated, respectively, by the null sets and

n .F{w(s), s< t}, n .97 {w(s), s >_ t} (6.1)


t t>o

are trivial.

It can be assumed that the initial point of the Brownian motion is the
origin. If A c- f, and if t > 0, the set A is independent of the class of dif-
ferences {w(s2) - w(s,), t < s, < s2) and therefore is independent of the a
algebra .f{w(s2) - w(s,), 0 < s, < s2} which, since w(0+) = 0 almost sure-
ly, together with the class of null sets generates a a algebra including .F,. .
Thus A is independent of itself and so has probability either 0 or 1. Apply
this result to the Brownian motion { tw(1 /t), t e I8+ } [see Section 5(f)] to
derive the triviality of .F2.
Observation. Since a Brownian motion process satisfies the conditions of
Theorem VI.8(b) after the associated filtration is enlarged if necessary
to have .`1v(0) contain the null sets, it follows from Section VI.8 that for a
Brownian motion from a point the a algebra .F, is trivial. This proof was
not used above because a direct proof is more elementary and just as short.

The Zero-One Law for Space-Time Brownian Motion

In view of the definition of space-time Brownian motion in terms of Brownian


motion the direct translation of Theorem 6 into the space-time Brownian
motion context is true.
584 2.V11. Brownian Motion

Applications. In the following applications is a Brownian motion in


R" from the origin, and TA is the hitting time of a set A by
(a) N = I. According to Section 5(c), lim,.., Iw(t)I = +co in measure:
so if c > 0. it follows that P { w (t) > c} > 1/4 for sufficiently large t. Hence

P; w(n) > c for infinitely many values of n in 7L' } > y.

Since the condition in the braces defines an .F set, the probability in question
must be 0 or 1. and so is 1. and therefore lim sup,- w(t) = + oo almost surely
because c is arbitrary. Similarly or by symmetry the inferior limit is almost
surely - x. Thus almost every path hits each point of R at arbitrarily
large parameter values. In particular, almost every path hits 0 at ar-
bitrarily large parameter values and therefore also [Section 5(f)] hits 0 at
arbitrarily small strictly positive parameter values. We shall see in Section
IX.5 that when N > I, almost no Brownian path in R" ever hits a specified
point at a strictly positive parameter value.
(b) According to Theorem 6, if A is an analytic subset of R". the value
of P; T" = 0} must be either 0 or 1. The following examples exploit this fact.
The dichotomy suggests that a topology of R" [A'] can be defined by making
a point a limit point of a set A if the hitting time of every analytic superset A,
of A by a [space-time] Brownian motion from the point is almost surely I.
We shall see in Section IX.15 that this can be done (even with A, open in
the Euclidean topology) and that the topology so defined is the [parabolic]
fine topology already defined in Section I.XI.I [Section I.XVII.9]. From
now on in this section the remarks relating probability results to fine topology
results will always be for the classical context, but the corresponding remarks
for the parabolic context will also be true.
(c) Let A be an open right circular cone in R" with vertex the origin.
Then P{ w(t) E A; is independent oft > 0 and is proportional to the cone
central angle. Hence, if t. is a sequence of strictly positive parameter values
with limit 0.

P; E A for infinitely many values of n} (6.2)

is strictly positive, and it follows from (b) that P{T" = 0} = 1. Actually


Theorem 6 implies the stronger result that the probability in (6.2) is 1. We
conclude that if r(t,o,) is the point of B(0, 1) hit by the ray from the origin
through w(t, (o), then for almost every co the sequence y(1,, w) is dense in
B(0, 1).

Cluster Values along Arcs


Let u be a function from a metric space D into a metric space D' with distance
function d'. Recall that if 0 is a function from an interval ]0, i5[ into D, with
C = O(]0. (Jr), and if the limit 0(0 +) exists, then the closed possibly empty
6. The Zero-One Law for Brownian Motion 585

set n,>o u(4(]0, t[))- is the "cluster set," that is, the set of limiting values,
of u at ¢(0+) along C. A point q' is in this cluster set if and only if

lim infd'(q', 0.
r_o

In (d)-(f) below, u is a Borel measurable function from W into a Polish


space D' with metric d'.
(d) If D' = dB, we weaken the hypotheses on u by demanding only that
the set Ar = {u > c} be analytic for all c in R. Under this hypothesis

{ lim sup u [w(t)] > b } = U { T"r = 0} a .f, (r rational) ;


r-+o r> b

so lim sup,.o u[w(t)] is .ft measurable and therefore almost surely constant.
In particular, if A is an analytic subset of D and if u = 1", the set A, is analytic
for all c and we find that P{T" = 0) is 0 or 1, as already observed in (b).
This special case illustrates the fact that Bore] measurability of u is sometimes
too strong a hypothesis. If u is supposed Borel measurable, both the above
limit superior and the corresponding limit inferior are almost surely con-
stant. We shall see in Section IX. 15 that the constant values here are, respec-
tively, the fine topology superior and inferior limits of u at the origin.
(e) A point q' of D' is a cluster value of u either along almost no or along
almost every path back to the origin because according to (d) the
probability
P{liminfd'(q',u[w(t)]) = 0)
r-o

is either 0 or 1. Let A' be the set of points q' for which this probability is 1.
We shall see in Section IX.15 that q' is in A' if and only if q' is a fine topology
limit value of u at the origin.
(f) The set A' defined in (e) is closed and is the cluster set of u along
almost every path back to the origin and if D' is compact,

P{lim d'(A', w(t)) = 01 = 1. (6.3)


r-o

In fact A' is obviously closed so if q' a D' - A', there is a neighborhood B' of
of so small that if B = u-' (B'), then P{ TB' = 0) = 0. The set D' - A' is
therefore a countable union Uj Bj of open sets Bj with this property. It
follows that neglecting a null set of paths, the cluster set of u along each
path back to the origin is a subset of A'. Furthermore, if A o' is a countable
dense subset of A', the cluster set of u on almost every path back to the
origin includes Ao and therefore includes A' and so is A'. The set A' may be
empty, but if D is compact, A' cannot be empty, and (6.3) is true because if
A" is an open neighborhood of A', a finite subunion of U; B; covers D' - A".
586 2.VII. Brownian Motion

(g) According to Theorem 6 the value of P{lim,.,c u[w(t)] exists) is


either 0 or 1. If this probability is 1, then the limit is almost surely constant,
that is, a single point of D'. In fact the set A' discussed in (f) is a singleton
in this case: so this assertion follows from (f) if D' is compact. If D' is not
compact, the assertion follows from the fact that the limit random variable
is measurable from a trivial o algebra into a separable metric space and
therefore is almost surely constant. It will be shown in Section IX. 15 that
lim,. u[w(t)] = of almost surely if and only if f limf.,0 u() = ry'.

7. Tied Down Brownian Motion


Let be a Brownian motion on ir' from and let n be any point of R'
(possibly ). If 0 < s, < < s < I the joint density of ww(s1), ... ,
given that w{(t) = 01 is

,CR-1 ,CJ (7.1)

This Gaussian joint density is that of a Markov process on the para-


meter interval [0, t], with initial value , value >I at t, and transition density
(from t;, at time s, to S2 at time s2, with s, < s2 < t)

01 - S21 C21 0 (7.2)


"(S2 - S1, C1, C2)
C2)
C(t-st,C1,PW

The joint density (7.1) is also that of a Markov process reversed in time on
[0, t] with initial value, at time t and value at time 0 and transition density
(from C2 at time S2 to C, at time s, , with 0 < s, < s2)

s1, Ct, C2)


Osi, , CJt'(s2 - (7.3)

Either way the N component processes are mutually independent.


When N = 1. the transition density (7.2) is Gaussian with mean and
variance, respectively,

C1 (r - s2) + 17(s2 - si) Q2[(s2 s2)]


r-S, t-S,
The random variable w; (s) is (for 0 < s:5 t) Gaussian. with mean and
variance, respectively,

tr,(1 -s)+qs o 2s(t-s)


I t
8. Andre Reflection Principle 587

The joint distribution of ww,,(st), ww,,(s2) is Gaussian with means and var-
iances as just evaluated and covariance u2st(t - s2)/t.
Suppose that N = I and that t is specified, and consider the process
defined by y(s) = wo(s) - swo(t)/t, 0S s < t. The process is almost
surely continuous, and its finite-dimensional joint distributions are Gaussian
with zero means and

- s2)
E{Y(st)Y(s2)} = a2s1(t (7.4)

If y*(s) = ww,(s) - s) + qs]/t, 0 < s < t, the and processes


have the same finite-dimensional joint distributions, and it follows that there
is a choice of that is, a process with the specified finite-dimensional
distributions, with continuous sample functions having values at time 0
and q at time 1. In the following the notation will refer to such a
process on RN except that the continuity condition may be weakened to
almost sure continuity. This process is sometimes called a Brownian bridge.
Denote expectations and probabilities for w,, by If 0 is a
bounded Borel measurable function on (RN)" and if 05 s, < < s" < t,
define x, = 4[we(st), ... , ww(s")] and x,,, = 4[xwn(st), ..., ww,,(s")]. These
are random variables defined on the probability spaces of and
processes respectively. Then

E4',, {x,,} = E4 x4 G(1 - S., w (sn), q) (7.5)

The reader is invited to generalize this equality to a larger class of pairs


(xc, x4',,).
By definition the finite-dimensional distributions when time is
reversed become the distributions. In particular, and
probabilities are the same for events that are invariant under time reversal,
for example, the event: a sample path enters a specified set at some time in
[0, t]. That is, Pc',, = Pq, on the class of such symmetric events.

8. Andre Reflection Principle


Let be a Brownian motion in R from the origin. Let a be a
strictly positive constant, and let T be the hitting time of of by Then
according to the strong Markov property of Brownian motion [see VI(9.4)]
if A is a Bore] subset of R and if A' is the reflection of A in the point a,

P{w(t)eAl.,F(T)} = $z(t - T, a, 1)11(dn) = Jt(t - T,a,q)11(dp)


(8.1)
= P{w(t)eA'J.F(T)} a.e. on {T < t}.
588 2.VII. Brownian Motion

In particular, if A = [a, +oo[, the value of each integral is Z . Ignore the


last two terms in (8.1) in this case and integrate over {T :!g t} to obtain

P{x(t)>_a,T<r}=P T<1 =P{sup,s,w(s)>-a}


2 2

Since P{w(t) >- a, T > t} = 0, we find that

P{ sup w(s) Z a} = 2P{w(t) z a}; (8.2)


s5I
that is, there is equality in (3.5). Since the left side of (8.2) is P{T < t},
the distribution of T is absolutely continuous relative to IN, with density
given by
z
(Density of the distribution of T) = (2no2)-t'2at-312
exp - -- . (8.3)

To derive a refinement of (8.2), suppose that A in (8.1) is a Borel subset of


] - oo, a[ and integrate over (T< t} to obtain

P{w(t)eA, T< t} = P{w(t)eA', T< t}. (8.4)

Since the term on the right is trivially equal to P{w(t)eA'}, we have proved

P {sup w(s) > a, w(t) a drl ) = G(t, 2a - n)h (dn) (n < a), (8.5)
,sl
which implies that

P { sup w(s) < a, w(t) a do } = [e (t, n) - 6(t, 2a - n)] 1 t (dn) (n < a). (8.6)
s51

Since the distributions involved here are continuous, "<a" and "<a" are
interchangeable in (8.6), and the corresponding remark is valid in the
preceding equations.

Intuitive Approach: The Andre Reflection Principle

The preceding derivations are formally satisfactory but give less insight than
the following derivation of corresponding results in RN, results that could
of course also have been derived by the above methods. Let D be a half-
space of R" whose closure does not contain the origin, and let n' [A'] be
the reflection of a point n [subset A] of RN in the bounding half-plane it
of D. Let { -fl-)) be a Brownian motion in R' from the origin, and let
T be the hitting time of it by The strong Markov property of Brownian
motion implies that the process {w(T + ),.F(T + )} is a Brownian motion
9. Brownian Motion in an Open Set (N .a: 1) 589

independent of .F(T); equivalently, the process { w(T + ) - w(T),.F(T + ) }


is a Brownian motion from the origin, independent of.F(T). The distribution
of the latter Brownian motion is unaffected if the sample paths are reflected
in a translation of it containing the origin. (Here we use the fact that the
distribution of Brownian motion from the origin is spherically symmetric
about the origin and that each coordinate process of this Brownian motion is
a Brownian motion in R from the origin, independent of the other coordinate
processes.) Hence (8.4) is true with the present interpretation of and A',
and it follows that

P{T < t, w(t)edq} = e(t,ti')IN(dq) (rleD), (8.7)

which implies that

P{T> t,w(t)edrl} = [11(t, P7) - t(t,tl')]IN(drl) (neD). (8.8)

If N = 1 and it = {a}, these results reduce to (8.4)-(8.6). Observe that in


the present context if a is the distance from the origin to it, then (8.3) follows
because to verify (8.3) we can assume that n is the hyperplane = a},
where 'N' is the Nth coordinate of , and if wl')(t) is the Nth coordinate
random variable of w(t), then (8.2) follows for and T is now the
hitting time of {a} by We shall interpret (8.8) in Section 9 as the
distribution at time t of "Brownian motion in D" from the origin.

9. Brownian Motion in an Open Set (N z 1)


Let D be a nonempty open subset of RN. For each point of RN denote
by a Brownian motion in RN from , and let SS be the hitting
time of OD (Euclidean boundary) by Let { be a Brownian
motion in RN with initial distribution supported by D, and let S be the
hitting time of OD by Denote by the process killed at S, so
that is a Markov process with state space D, initial distribution
that of and substochastic transition function p given by

p(t, , A) = P{ww(t) a A, t < SC}. (9.1)

An almost surely continuous Markov process 9(-)) with state space D


and transition function given by (9.1) will be called a Brownian motion in
D ; the process { ,F(')) is a natural example. If 4(t) is the a algebra
generated by the null sets and .F{z(s), s 5 t}, then is right continuous
according to Theorem VI.8. Depending on the context, it may or may not
be useful to make the transition function stochastic by adjoining a trap point
to D.
Since the right side of (9.1) is at most P{ws(t) a A}, the measure p(t, , )
590 2.VII. Brownian Motion

is absolutely continuous relative to IN, determined by a density 'D(t, t, ). A


unique choice of this density will now be made. Let , n be points of D,
and let fit, , q) be the probability that a Brownian bridge from to
rj never hits M. Define GD as suggested by the definition of the Brownian
bridge and the definition of f(t, , p) for 1 < 0:

(t,. 1)1(t,, ri)


f D(t,, 1) = 10 (9 2)
if t < 0.

In particular, ((t, , q) = ((t, , ry) when D = IBN. By the symmetry remark


closing Section 7 the function 4D(t, , ) is symmetric. Equation (7.5) yields

(9.3)
Stf
for i > 0, and the equation is trivially true for r < 0. The function GD is Bore)
measurable on I8 x D x D. The expectation on the right in (9.3) can be
evaluated when 0 < s < t using the strong Markov property of Brownian
motion :

E{Eills(4ss,G(t (Ss)}}
(
= E j l Iss<sl J 1(t - s, C, rl)P(s - SS, ww(SS), C)IN(dC)J (9.4)
l RN ))

= E{ l,ss<Slf(I - SS, WS(SS), q) }

so that (9.3) reduces to

/'DO, , n) = r(t, , 1) - E{L(t - SS, ws(SS), 1) }. (9.5)

This evaluation is valid on I8 x D x D. Since (ww(SS) - ql I aD - >7I when


SS < + xc, the integrand in (9.5) is bounded uniformly for all t, in D, and
n in a compact subset of D. Hence GD(t, 4, ) is continuous on D and is
obviously the only continuous density for the transition probability of
Brownian motion in D. The Chapman-Kolmogorov density equation

/D(s + t. . S ) = s)/N(dq) (S > 0, t > 0) (9.6)


D

is true for all , C in D in view of the following facts:


(a) The left side varies continuously with C.
(b) The right side varies continuously with C, because the integrand does
and the integrand is majorized by the integrable function
fD(s, , )(2nvzt)-Niz
9. Brownian Motion in an Open Set (N 2 I) 591

(c) The equation is true for IN almost every C when is fixed because
GD is a transition density for Brownian motion in D.
In view of the inequality GD 5 G it follows from (9.6) that GD(t, , ) is a
continuous bounded function on D x D for each value of t. Moreover,
according to the analysis of the integrand in (9.5), the expectation on the
right in (9.5) is for fixed q a bounded function of t) on D x ]0, + oo [.

Potential Theory Significance of eD

Let D be a nonempty open subset of R", define D = D x R, and let


(ID be the parabolic Green function of D. The function ((c, s), (q, t)) F_.
GD((, s), (q, 1)) is a function of s - t, , q, and in Section I.XVII.18 a func-
tion GD was defined by

GD(s - t, , q) = ObW, s), (q, t)). (9.7)

In Section IX.17 we shall identify GD as defined by (9.7) with the Brownian


transition density function GD defined in this section. The probabilistic
evaluation of parabolic measure to be made in Section IX. 13 makes the
expectation in (9.5) the solution Hf of the parabolic Dirichlet problem on
b with f the restriction to ad (Euclidean boundary) of e(-, , q). In this light
(9.5) can be interpreted as the Dirichlet solution construction of (ID described
in Section I.XVIII.I.

EXAMPLE. Let D be a half-space of RN, define b = D x R, and if ,i E U,


denote by q' the reflection on 1 in aD. Then [Section 1.XVII(4.3)] rl) =
Q, q) - O(, q'); that is, if GD satisfies (9.7), if s) and t), and
if q' is the reflection in R" of q in OD,

GD(t, c, q) = G(t, , q) - G(t, a, q') = G(t, q - ) - G(t, q' - ). (9.8)

Observe that according to (8.8), in which = 0, equation (9.8) is true with


the probability interpretation of GD as a Brownian transition density function.

Canonical Brownian Motion in an Open Set

Let D be a nonempty open subset of R". Adjoin a point a to D, defining


the topology of D u a so that D v a is the one-point compactification of D.
Let t) be the class of all functions m from R+ into D u a with the following
properties :

cu(s) = a implies that w(t) = a for t > s.


592 2.VII. Brownian Motion

The function tv is continuous on the interval [0, . (d)) [, where S(to) =


inf {s > 0: to(s) = 8} and there is a left limit (Euclidean topology of RN) at
Si(tu) when 9((o) < + oo.
Define w(t, d)) = tu(t), fi(t) = .F {w(s),s 5 t},, .F{w(s),sc- Rr}. If P
is the completion of a measure on , under which .4 is a Brownian
motion in D, the process will be said to be a canonical Brownian
motion in D. If D = RN this definition yields a canonical Brownian motion
in RN as already defined in Section 3. One way to obtain a canonical
Brownian motion in D is to operate on a canonical Brownian motion in RN
as follows. If S is the hitting time of the Euclidean boundary of D by the
Brownian motion, change every sample function value to 0 at every para-
meter value at least equal to S.

10. Space-Time Brownian Motion in an Open Set


If D is an open subset of AN, a space-[co]time Brownian motion in D
is a space-[co]time Brownian motion with initial distribution supported
by D, killed at the hitting time of DD. The killed process is Markovian with
substochastic transition function p given by

P{wt(t)EA,t < 94}, 4= (10.1)

where w is a space-[co]time Brownian motion from and 94 is the hitting


time of ab. From now on only the space-time Brownian motion is considered
unless the contrary is stated. We shall use the space-time notation introduced
in Section 2. The measure p(t, 4, ) is supported by D,_ and the set function
A s - p(t, , A x is - t}) for A a Borel subset of D5_, is majorized by

Af- f A

so the first set function is absolutely continuous relative to IN and is therefore


the integral of a density. A unique choice of this density will now be made.
Following the reasoning in the Brownian motion case, let : s) and, : (q, t)
be points of b, and if s > t let n) be the probability that a Brownian
bridge w has the property that the space-time bridge
r],0 < r 5 t}

from to q never hits 3D. Define

JOs 07)I(4, 0 ifs > t,


(10.2)
0 if s<t.
10. Space-Time Brownian Motion in an Open Set 593

This definition of the transition density (from to ' in time s - t) leads to


the evaluation

In( , n) = I(s - i , . ,,7)- E{I(s - t - (10.3)

(See the corresponding transition density derivation for Brownian motion


in Section 9). The density q -+fD( j) is continuous, and the Chapman-
Kolmogorov equation takes the form
ss
ID(SI , '3) = J S2) 1(S2+
YY

(Si,.Si). (10.4)
nsx

satisfied identically for s, and S3 in b with .s, > s2 >


The condition that an extended real-valued positive universally measur-
able function ii on D be excessive for space-time Brownian motion in D is
the validity of the inequality

I 5 t}(SI) (sI > s2) (10.5)

together with equality in the limit when .s2 T s, . If in addition there is equality
in (10.5) whenever .s, > s2, the function ti is invariant for space-time Brown-
ian motion.
The transition density (o of space-cotime Brownian motion is defined
by dualizing the preceding discussion and satisfies the equation

In(n, )=fn( (10.6)

in view of the remarks on time reversal in a Brownian bridge at the end of


Section 7.

The Function and I n( 1)


Let D be a nonempty open subset of 08" (N z 1), and define D = D x R.
Equation (9.6) combined with the inequality In 5 f. implies that for fixed
, in D the function is excessive for space-time Brownian motion
in b, invariant excessive for space-time Brownian motion in b - '(?1. 0)),
continuous and 0 at the points of D - {(q, 0)) with ordinate values <0.
We shall show in Section IX.17 that In(s - t, S, ry) = O, s), (n, t)). Even
without this identification the fact that ry) is superparabolic on D and
parabolic on D - {(ry, 0) } follows from the fact to be proved in Section IX.8
that a space-time Brownian motion excessive function is superparabolic on
the set of points strictly below any point of finiteness of the function and that
a space-time Brownian motion invariant excessive function is parabolic on
the set of points strictly below any point of finiteness of the function. More
594 2.V11. Brownian Motion

generally, if b is an arbitrary nonempty open subset of ft' and if 4 =


7 = (q, t) are points of b, equation (10.4) together with the inequality
i1) < /(,s - t. - p) implies that the function r) is excessive for
space-time Brownian motion in b, invariant excessive for space-time
Brownian motion on D - {tj}, continuous and 0 at the points of D - {i1}
below ry'. We shall show in Section 1X.17 that U = G0.

11. Brownian Motion in an Interval


Let I be the one-dimensional interval ]a, h[, and let be a point of 1. Let
it-(-) be a one-dimensional Brownian motion from the origin, so that
S + it-(-) is a Brownian motion from . Fix t > 0, and consider the following
events. Each parameter value t, below depends on the sample function.
QaO': it-,(-) first meets of at a and does so before time t.
Oho': iv,(-) first meets 81 at b and does so before time t.
ON There exists it with 0 < t, < t and w,(t,) = h.
QahQ: There exist t,, t, with 0 < t, < t2 < t and ww(t,) = a, w,(12) = b.
QhahQ: There exist t, I. 12, 13 with 0 < t, < 12 < 13 < t and ww(i,) = h.
w4(12) = a, µ',(t3) = b
and so on. Then
ObO' = ObO - OaO' n OhO,

OaO' n ObO = OabO - ObO' n OabO.

Oho' n QabO = ObabO - QaO' n QhahQ,

so that
OhO' = OhO - QabO v QhahQ - QabahO u . (11.2)

The probabilities of these events are easily calculated using the reflection
principle. For example. P;OabO} is the probability that a path reaches
h by time t after having hit a, or, reflecting the path in a at the hitting time
of a, and settings=h - a
P {QahQ} = P{minwjs)
S<I
<a - c} = PI S5/
min tv(.s)<a- e -
(11.3)
= 2P{w(t) > - h + 2c}.

Computing the probabilities of the sets in (11.2) in this way yields

P{QhQ'; = 2 (P{w(t) > 2ne + h - S, - P{w(t) > (2n + 2)c - b + 5}].


(11.4)
12. Probabilistic Evaluation of Parabolic Measure for an Interval 595

Just as (8.2) was strengthened into (8.5), so the condition w,(t) a dry can be
imposed on each term of (11.2) to obtain

00

P{QbQ', v4(t)1=dq} = [6(t, 2nc + h - - + I q - bj)


(11.5)
(gE18)

and similarly

P{QaO',ww(t)Edq} =Y[f.(t,2nc - a + + Ia- q{)


0 (11.6)
- G(t, (2n + 2)c + a - + I a - q!)]l, (dq) (q E 68)

The probability (density) that a Brownian path from will be at n in 1 in


time t without having left I before that time is v 1(t, , q) less the sums in
(11.5) and (11.6), and trivial manipulation yields

r(t, S, q) _ Y [t (t, 2nc - + q) - G(t, 2nc + 2a - - q)] (q E 1). (11.7)

The function Gr is the transition density of Brownian motion in 1, defined


in Section 9, because G,(t, , ) is continuous on I.
More generally, if D is the interval I, x ... x I. with I, = ]a,, b,[, it is
trivial that the transition density of Brownian motion on D, that is, the
probability (density) GD(t, , q) that a Brownian path from : (S11) ... , YIN))
will be at q: (q(", ... , ,(N) in D at time t, without having left D before that
time is
N
"DO, , q) = nG,i(t, Ii) qli)). (11.8)

If b = D x 18 and if t;D is the parabolic Green function of D, a comparison


of (11.8) with 1.XV(8.5) verifies for D the relation between the transition
density GD and the Green function OD announced in Section 9; that is, the
transition density e D satisfies (9.7).

12. Probabilistic Evaluation of Parabolic Measure for an


Interval
Let D be an interval of R', and let ] t, , t2 [ be an interval of U8 so that D =
D x ]t,, t2[ is an interval of 9BN. It will now be shown that the distribution
of the first hitting point of aD (Euclidean boundary) by a space-time Brown-
ian motion from 5 = s) in D is the parabolic measure po(, ). The discus-
596 2.VI1. Brownian Motion

sion will be given for N = 1 to simplify the notation. Suppose then that
D = ]a,b[.

Hitting Distribution on the Lower Boundary of ,6


According to Section 11, the distribution has density (relative to 1t)

11-GD(s- 660,(11,1[)),

and according to 1.XV(9.1), this density is that of parabolic measure on


the lower boundary.

Hitting Distribution on the Lateral Boundary of D


Only the hitting distribution on the segment of the lateral boundary with
abscissa value b will be considered. Let Tb be the first time a Brownian path
from reaches b, considering only those paths reaching b before a, so that
in the notation of Section II

P{QbO'} = P{Tb < t}.

Differentiating in (11.4) yields the Tb distribution density:

t"- Y (2nc + b - )e(t, 2nc + b -


t -1
The first hit of the segment of the lateral boundary with absicca value b is
at the point (b,s - Tb); so (12.1) can be interpreted to yield the distribution
density of the hitting point, and comparison with equation I.XV(9.1) shows
that this density is the density of parabolic measure, as asserted.
This identification of the parabolic measure µD(4, ) with the distribution
of the first hitting point of oD by a space-time Brownian motion from
will be made for arbitrary open subsets D of E` 1N in Section IX.13. The
identification gives an intuitive interpretation of the fact (Section I.XVIII.1)
that the parabolic measure of a boundary subset relative to a reference
point vanishes if the set is above the reference point.

13. Probabilistic Significance of the Heat Equation and Its Dual


If ,6 is an open subset of A", we have defined a function 6D which determines
the space-time Brownian motion transition function, and we have remarked
that the identification oh = dD will be made. In particular, if 1) = D x62
for D an open subset of 62", this identification reduces to GD(s - t, , r1)
= D({ , s), (q, t)). Since (Section 1.XVIII.1) OD(-, (r1, t)) is equal to (q, t))
less the parabolic function Dirichlet solution on ,6 for the boundary function
13 Probabilistic Significance of the Heat Equation and Its Dual 597

(q, t))Iao (=0 at oo), it is sometimes convenient to evaluate GD, in


particular, to evaluate GD, by solving the relevant Dirichlet problem. We
have already (Section 9, Example) verified the correctness of this evaluation
of GD when D is a half-space and (Section 11) when D is an interval. In these
examples the essential point is that the differential equation method of images
which leads to the Green function is the counterpart of the probabilistic
Andre reflection principle which leads to the Brownian motion transition
function.
In the physical context the function (q, t) F-.s), (q, t)) is the temper-
ature at time t at the point q due to a heat source atactivated at time s.
At every time t the boundary OD, of the N-dimensional set {n: (j7, 1) c-,61
is held at temperature 0. Alternatively, the context is that of a substance
diffusing from a point source at , starting at time s. At every time t the
boundary OD, is an absorbing barrier, and Go((, s), (q, t)) is the density of
the diffusing substance at time tat the point q. In these contexts the derivation
of the heat equation is local, and the various physical boundary conditions
lead to solutions of the heat equation with corresponding mathematical
boundary conditions. For example, suppose that almost every space-time
Brownian path from a point of b hits eD, as will be true if D is bounded.
We would expect from the localization just noted that for a sufficiently
smooth boundary and for a sufficiently smooth boundary function f the
expected value off at the first hitting place of a space-time Brownian path
from _ s) would define a parabolic function ti of . Moreover, since
for 4 near aD it is plausible that paths from 4 are likely to hit ab soon,
it seems plausible that ti has boundary limit function ff Thus it is plausible
that ti as just defined probabilistically is the PWB solution of the Dirichlet
problem for parabolic functions whenever f is resolutive. In particular, if
D = D x R" with D a Greenian subset of R" and if (q, t) H f (q, t) is a function
of the space variable, say An, t) = f(q), the function ti will be a function of
the space variable, say t u is harmonic, and it is plausible
that u is the PWB solution of the Dirichlet problem for harmonic functions
whenever f is resolutive in this context.
The principal tool we shall use in discussing these matters is martingale
theory, and the evaluations of Dirichlet solutions will turn out to be applica-
tions of the martingale equality. To show how martingale theory arises in
a natural way in this context, consider the following problem. Let be
a Brownian motion in R" from a point. What functions u on R' have the
property that {ti[t, (t), t], t E 68+ } is a martingale, in some local sense at
least? If 6 > 0, a formal application of Taylor's theorem leads to

0 = E{{u[w(t + 6),t + 6]Iw(s), s < t} - u[w(t), t]

=b
2 etil + 0(62) (13.1)

= 6&i + 0(62),
59t 2 VII Brownian Motion

where the partial derivatives are evaluated at [w(t), r]. Thus it is plausible
that ti must he coparaholic and that conversely the process tt[tr( ), ] is a
martingale in some local sense if iu is coparabolic. Equivalently, u com-
pounded with space-time [cotime] Brownian motion is a martingale in some
local sense if and only if ii is parabolic [coparabolic]. In particular (if u is
a function of the space variable), a function u on R" composed with Brownian
motion is a martingale in some local sense if and only if u is harmonic.
The results suggested in this section will be formulated rigorously and
proved in later chapters
Chapter VIII

The Ito Integral

1. Notation
Let be a Brownian motion in IJB, defined on some probability
space (Q, .F, P). It is supposed that is right continuous and that ,F(0)
contains the null sets. The Ito integral !{, ¢ dw will be defined for stochastic
processes in the space IF of not necessarily adapted to F(-) real processes
{ Ib(t), t c l } with the following property: there is a progressively measurable
process depending on for which

P{ a,: f jk(s,ctw)j2ds < +ao} = I (1.1)


0

for all finite t and


P{4(t) = ar(t)} = 1 (1.2)

for /, almost every t. In this chapter ds refers to Lebesgue measure /, . For


economy in later references absolute value signs are used in (1.1) and similar
contexts because the present discussion will be extended in Section 7 to cover
vector processes and processes with complex state spaces. Observe that a
process indistinguishable from one in r is itself in r. Let /, x P be the
completed indicated product measure on R' x 0, defined on the completion
of R (R') X . relative to this product measure. According to Section 1.13,
if {4 (), is an extended real-valued adapted process and if the function
is /, x P measurable, then this process is in IF if (1.1) is satisfied.
We metrize r by identifying two members and of r whenever
(1.2) is true for /, almost every t, thereby making r into a space of equivalence
classes, and by defining (abbreviating the notation)

r dist (0, 0) _ 2-kE J I A


012
ds lizl . (1.3)
k=1 1 Ifo k 10 _ )J}}

Here and are progressively measurable, and we are adopting the


usual convention : the right-hand side of (1.3) is the distance between the
2.Vl1I. The Ito Integral

equivalence classes containing and vi(-). In the following the reader is


asked to judge from the context whether notation like refers to an
individual process or to an equivalence class. It is also left to the reader to
verify the fact, which we shall not use, that r is complete in its metric.
The subset of r for which (1.1) is true with t = + oo will be denoted by
i, and the subset of T for which

E! fo lcI2ds1 < +ao


metric

will be denoted by r'2. A for f sometimes preferable to that induced


by the r metric is the F metric in which (1.3) is strengthened to
ltrz).
I dist(0,t)=El A I0-IpI2ds{[J
(1.4)

A r2 metric sometimes preferable to that induced by the r or r metric is


the L2 metric in which (1.3) and (1.4) are strengthened to

)=
r'2dist(.0,Eu21jo. 1.0 -,I2ds}. (1.5)

The spaces f and r2 are complete in their metrics. There is convergence of


a sequence n eV) to in the r metric if and only if

plim f (1.6)
n- o

for all finite t, and there is convergence in the I' metric if and only if (1.6)
is true for t = + oc. Here 0.(-) and are chosen to be progressively
measurable members of their equivalence classes, in t or r as required.
There are sometimes formal advantages in having members of r defined
at the parameter value + oo. This can be effected for our purposes by defining
¢(+oo) arbitrarily, by defining ffl(+oo) = F, and leaving the r, F and r2
metrics unchanged.
Observe that if 0 is a bounded member of r, then is in I' or i'
or r'2 if is. A useful special case is ii(-) = 1sT, defined as the indicator
function of the stochastic interval IS, TD.
As usual the definition of an integral is first given for a simple linear
class of integrands. In the present context this class, a subclass of 1'2, is the
class r'o each of whose equivalence classes contains a process defined
as follows. There is a finite set 0 < t, < < tk < + oo and corresponding
bounded random variables f, , ._f, f, such that f is F(:) measurable and
that
2. The Size of ro 601

0 if tSt,,
0(1) = if t;_, < 1:5 1,, 2 < j < k, (1.7)
0 if t > tk.

A function 0 in ro has many representations (1.7) because additional parti-


tion points can be adjoined without changing adjusting f as required.
If and in ro are to be considered simultaneously, partition points
can be adjoined if necessary to obtain representations of and with
the same partition. Using this fact it is obvious that ro is an algebra.
Observation on predictability of Ito Integral integrands. The process
defined by (1.7) is adapted, left continuous, and therefore predictable. The
following lemma states that ro is dense in the sets r, r, r2, and a trivial
adaptation of the proof shows that each equivalence class of each of these
sets contains a predictable member. We shall not use this fact.

2. The Size of ro
Lemma. The set ro is dense in r, r, and F. in the metrics of these spaces.

It is sufficient to consider progressively measurable members of these


spaces. Suppose first that is a progressively measurable member of r,
and define

>V(t) ifl41(t)lSnandt5n,
0 otherwise.

Then lim,, . 44') in the r metric. Next define

ift< m
<t5 J, 1 <j5nm,
ift>n.
The process is progressively measurable, I4'. 1 < n, a ro, and
lim,,,-. in the r metric because there is bounded 1, almost
everywhere sample function convergence. This completes the proof of the
lemma for the r metric and the proof of the lemma is concluded by the
observation that for 4(') in r or 1-2 the convergence assertions in the above
proof are also valid in the metric of the space f or r2 as the case may be.
602 2.V11I. The Ito Integral

Refinement. A trivial computation involving the Schwarz inequality shows


that

a.s.
0 0

Thus an arbitrary element of r, r, or r2 is the limit, in the metric of its


space, of a sequence {¢ n z 0} in I'0 satisfying almost surely

JII2ds fI2ds (2.1)


0 0

simultaneously for all t.

3. Properties of the Ito Integral


If 0 05t oo if E r], the integral
x(t) = J 0dw will be defined uniquely up to a null set. Moreover it will be
shown that a version of x(t) can be chosen for each tin such a way that the
map 0(-) - is a linear continuous map from IF [I'] [[r2]] into a metric
space r' [r] [[I-Z]] of processes defined as follows. The space I" is the
space of almost surely continuous adapted processes metrized
into a complete metric space (in which indistinguishable processes are
identified with each other) by

r' dist Y 2-kE{ I A sup Ix(t) - y(t)I }. (3.1)


1 rsk

The space F" is the subset of r' for which almost surely lim, x(t) = x(+ oo)
exists and is finite. The I" metric is defined by strengthening (3.1) to

I' dist E{ I A sup IX(f) - y(t) 1). (3.2)


rc+m
The space r2 is the subset of I"' for which E{sup,<+,, lx(I)III < +oc. The
I'Z metric is defined by strengthening (3.2) to

rZ E"2{ ,<+
sup I x(t) - y(t)I2}. (3.3)
OD

The spaces r', 1", and rZ are complete in their metrics. When 4'(')e r, the
integral !o 0 dw = x(t) will be defined for 0 5 t < + oo [or 0 5 t 5 + oo if
¢(.)e r] and will have the following properties.
(a) x(0) = 0 almost surely, and a version of x(t) can be chosen for each
t in such a way that E F. In the following it will be assumed that
is so chosen.
3. Properties of the Ito Integral 603

(b) The map is linear and continuous from r into r, i into


r', and I'2 into r; in terms of the metrics of the spaces involved.
(c) If T is a finite optional time, then

lOT0dw = x(T) a.s. (3.4)


f"'
The upper limit oo is legitimate here because the integrand process is in I'.
Moreover, if F, then (3.4) is true for an arbitrary optional time T.
Observe that if T is an arbitrary optional time the process is in F
and is identified with 0 in the t metric; therefore if S and T are optional,
with S< T, it follows that los + Isr =for under the t identifications. Hence
if is in r and if T is finite valued,

I5TOdw = x(T) - x(S) a.s. (3.5)


f"'o

If is in F, the optional times need not be finite valued. It will sometimes


be convenient to use the notation is 4dw for the left side of (3.5).
(d) Let S and T be optional times with S< T < + oo. If f is a bounded
.F(S) measurable random variable and if is in I', then
fT
f T fO dw = f 0 dw a.s. (3.6)
S is
If is in F, the optional times S and T need not be finite valued.
(e) The process is a local martingale when is in F.
(f) Let S and T be optional times with S < T < + oo. Suppose that
is in t and that

E{ If Icbl2ds} < +oo. (3.7)

Then

4dwl'I-a2E{Js I4,I2dc}
EIl
and the process {x(T A (S + t)), F (S + t), 0 < t < + 001 is an almost surely
continuous martingale. If and in t satisfy (3.7), then

Elf T.Odw f'0 dw} = a2E IJs .P0 ds}.


[The integral IT 0 dw is defined in (3.8) and (3.9) even if T is infinite valued
because (3.7) implies that the process is in F.]
604 2.VIII. The Ito Integral

For in r define

f.1012dvf, 0:!9 t < + cD (3.10)


2

and allow t = + co when a F. The process is almost surely


continuous.
(g) For in r the process is a positive supermartingale
on the parameter interval [0, +oo], with y,(O) = I almost surely.
(h) Ifb>Oand6>0,then
r 2 r1012
D

PIsup I 4 A 6+ ds 5 2e-ija. (3.11)


r5D Z
(i) If cel andifp> I then for 0(-) in r

E{Iym(t)I`} <Ew-,)1r{exp(cppcp 11)


LL
2 f(3.12)
limCP(cP - 1)= 1.
(i P- I (3.13)

For 0 in F the process { is a local martingale.


If d e F and if for some b 5 + oo,

E{expr JI.0I2ds]I +c0, (3.14)


2 o <

then the process {ye(t), (t), t S b} is a martingale.


(1) If )'a I8 and if F2, then

E{ (t) f" ¢ dw} = a2Y JE{,' (s)O(s)} ds (3.15)

a nd

y,(t) - 1 = Y f r yr dw a.s. (3.16)


Jo

[More generally for in r

ym(t) - I = JcbY.dw a.s.,


0
4. The Stochastic Integral for an Integrand Process in F 605

as will be seen in Section 12, Example (b), by means of Ito's formula, but
the special case (3.16) will be needed before Ito's formula is derived.]

4. The Stochastic Integral for an Integrand Process in I'o


For ¢ in r'o, specified by (1.7), the stochastic integral is defined in the obvious
way :

x(t) = f O dm- = Y./ _t [w(t; A t) - w(tt_ { A 1)], 0 5 t s + oo. (4.1)


2

Under this definition x(0) = 0, x(t) is .(t) measurable, and is a contin-


uous process, but we allow as other versions of the integral process every
process indistinguishable from that in (4.1). Furthermore, if to = 0, then
almost surely

E{x(tt) I.F(tt_ 1) }
0 = x(0) ifj = 1, (4.2)

x(tl-1) + ft_,E{w(t) - w'(tt-t)I (tt-t)} = x(tt_,) ifj > I.

Hence is a martingale; so if 0< s < t and if s, t are in it


follows that E{x(t)I.fls)} = x(s) almost surely. Since any pair of parameter
values can be adjoined to the set it follows that is a martingale.
This martingale is L2 bounded, hence has orthogonal increments, and

E{Ix(+oo)I2} =62EI f °°I4(s)I2ds}. (4.3)


0

If q() is defined by (1.7) and by (3.10)

E{ ym(r)I'F(i _1)}
I = y,,(0) a.s. ifj = 1,
y,,(tt-1)E{exp [I--t (K'(tt) (4.4)

- w(tt_1))]I_F(tt_1)) exp 1 2 (tt - tt t)fJ tJ a.s. ifj > 1.

Now f_, and w(l) - w(tt_,) are mutually independent, and whenever z
is a normally distributed random variable with mean 0 and variance a,
E{exp(az)} = exp(a2(x/2). Hence the right side of (4.4) reduces to y'y(tt_,).
Thus { is a martingale, and it follows repeating an argument
just used that {y,,(-). F(-)) is a martingale.
606 2.VIII. The Ito Integral

5. The Stochastic Integral for an Integrand Process in r


Suppose that c f,. Then (Section 4) the process { y, defined
by (3.10) is a positive martingale, and if c5 > 0 and b > 0, the submartingale
maximal inequality 111(9.1') yields

PJsup I 1 +?z <c-iia.


(5.1)
r_h ,Iii 6 2 o rsh

If 0 is replaced by - c1 and the resulting inequality is combined with (5.1),


we find

P{supl+o4dwI>_ <2e
j+ 2 JJ42ds} (5.2)

or, if (h is replaced by (k/o2,

P,sup 4) dw >0+ f h tae 1 5 2e- ua. (5.3)


rsh I.(') 262

We conclude that if ne Z' } is a sequence in 1' with limit 0 in the r


metric, then the sequence { J' 4 dtv, n e Z i has limit 0 in the r, metric.
That is, the linear map 0 -+J / dw from r0 in the f metric into the
metric space r' is uniformly continuous. It follows that there is a unique
uniformly continuous linear extension of this map, from the r metric closure
of r,, that is, from r, into F'. We accept this extension as the definition
of the integral O ¢ dw; this integral then satisfies Section 3(h).
Furthermore the map q5"Ja 4 dw takes r into F'. and if we apply
the reasoning just used but this time with the f metric on F00 and the T'
metric on the image of CO, we find, setting b = + oe in (5.2), that the map
is uniformly continuous. Hence this map is uniformly continuous from T
into F'.
Finally we consider r, in the IF2 metric. Since 110 0 dwl is a submartingale
for 4 in f;,, the L2 maximal inequality 111(11.1) for the continuous parameter
context is applicable and yields

E{sup I f odidwl2} <4o2E{ f -IQil2dr1, (5.4)

and it follows that the map is uniformly continuous from 172


into F.

Localization of the Stochastic Integral


We have defined IT 0 dtr for T optional as Jo lor4 dw. If 0 is defined initially
only on X10, T1, define 0 as 0 on 08+ x Q - ] 0, TI, and if the extension is
in F. define Jo 0 dw using the extended integrand.
6. Proofs of the Properties in Section 3 607

6. Proofs of the Properties in Section 3


Section 3(a), (b), (h) have already been proved. We prove the remaining
properties in the order (cdfegikjl).

Proof of Section 3(c). (cl) In order to prove (3.4) either for in r' with
T < + o0 or for in r with T :g + oc, it is actually sufficient to prove (3.4)
for T bounded. In fact, if (3.4) is true for T bounded, then for general T,

x(T A n) a.s. (6.1)


f."'

Now when either is in r and T < +oo or is in r and T:!,- +oo, it is


true that limx(T A n) = x(T) almost surely and the integrand process
in (6.1) tends to the process in the r metric so that
ar, o0

IOTO dW- 107'nn0d"'


plt°° IJ 0 0

<- p lim sup


"-w 12!0
dw - I dw' I = 0. (6.2)
0 0

Thus (3.4) is true. From now on in the proof of (3.4) we shall suppose that
T is bounded and that 4(() is in F.
(c2) If (3.4) is true for the optional time it is true for T because
loT(')4(') in the F metric and lim, x(T)
almost surely.
(c3) Proof of (3.4) for in r,. According to (c2), it is sufficient to
prove (3.4) with T replaced by At the possible cost of modifying the
representation (1.7) of we can suppose that t. includes the values
{j2-",je7L+} in the interval [O,tk]. Finally by linearity of the map
1o 0 dw it is then sufficient to prove (3.4) for given by (1.7) with k = 2.
That is, we now suppose for f the indicator function of the .(t,) set
{ [T] > t, } that
(0 ift<tl,
jl j if t, < t < 1Z, (6.3)
0 ift>l2.
This process is in 1,0, and the definition of the integral for an integrand
process in r70 makes (3.4) trivial in this case.
(c4) To prove (3.4) for arbitrary in F, let n e Z') be a sequence
in F0 with limit in the F metric, and define $o 0 dw. According
to (c3),

J
608 2.VIlI. The Ito Integral

and (3.4) follows because in view of the boundedness of T,

lira
n-n; IoT(')On(') = lor(')O(')

in the r metric; so

plimJx(T)-xn(T)I<plimsupl f1,,Odw - f lor4ndwl=0. (6.4)


o ,Jo

Pro of of Section 3(d). We can suppose, redefining 0(t) as 0 for t >- T if


necessary, that T = + oo and that is in f. In view of the metric properties
of the map 0 dw it is sufficient to prove (3.6) with S replaced by
[S]n and with in To, given by (1.7), and adding points to t in (1.7) if
necessary, we can suppose that t includes the values {j2-",jeZ+} in the
interval [0. tk]. Finally in view of the linearity of the above map we can even
suppose that k = 2. The process is now in ro and has a represen-
tation of the form (1.7) with k = 2, and the computation verifying (3.6) is
trivial.

Proof of Section 3(f). Suppose first that S=_ 0, T=- + oo and that (3.7) is
true. Then is in r2. Let be a sequence in r0 with limit
in the r2 metric, and define xn(t) = f o 0. dw. Then lim"-,,
in the r2 metric; in particular,

lim E{ Jx(t) - x"(t)12} = 0 (t < +00).

Since the process {xn(t), ,f(t), t < + oo } is a martingale, it follows that the
limit process is a martingale. Moreover

E{ Jx(+oo)12} r
= a2E I0I2dsl
Jo J)

because this equation is valid for and 4"(-). In the general case of
Section 3(f) if we apply what we have just proved to we find
[see (3.5)] that

{x(Tnt)-x(Snt),.I'(/),t<+OD }={ f 15TOdw,.y(t),t<+oo} (6.5)


0

is a martingale and that (3.8) is satisfied. Equation (3.9) is obtained from (3.8)
as usual by polarization. Since the martingale (6.5) is almost surely contin-
uous and right closed, the process remains a martingale when t is replaced
by S + t; so the process {x(T n (S + t)) - x(S),.F(S + t), t < + ao} is a
martingale; equivalently, {x(T A (S + t)),.F(S + t), t < + x j is a martin-
gale.
6. Proofs of the Properties in Section 3 609

Proof of Section 3(e). We prove this property using Section 3(f), just proved.
For in t and a > 0 define

T,=infjt: fo 'I0I2dsZa}. (6.6)


l 2
Then T. is optional, and (a2/2) S.- IZ ds < a. Hence according to Section
3(f), the process {x(T, A t), ,fi(t), t 5 + oo } is a right closed and therefore
uniformly integrable martingale. Since lim,_m T, = + oo almost surely, it
follows that is a local martingale.

Proof of Section 3(g). Let be in r, and let ne Z+} be a sequence


in f0 with limit in the I- metric. Then

li m f f cbdw
o Jo

in the r' metric; so p li1n,, ym1(t) = y(t) for each t. The process { y,,.(-),
.fl-)J is a positive martingale, and an application of Fatou's lemma for
conditional expectations shows that { is a supermartingale.

Proof of Section 3(i). If ce IE, and if p > 1, then

a2
y4(t)`=y,, (t)tweXp[cCoc - 1) 2 J I#12dS]

(6.7)
o

fromwhich the inequality (3.12) follows by an application of Holder's in-


equality and Section 3(g). The limit relation (3.13) is trivially true for p as
stated.

Proof of Section 3(k). If b < +oo, replace by to obtain an


equivalent context in which we can suppose that b = + oo and a r72. We
are to prove that (ym(t), ,f(t), t 5 + oo } is a martingale; equivalently, since
this process is an (almost surely continuous) supermartingale with
E{ym(0)} = 1, we are to prove that E{ym(+oo)} = 1. According to Section
2, there is a sequence in ro with limit in the r2 metric
and satisfying

I0.I2dSs jI#Ids nEZ+.

(6.8)

It follows from Section 3(h) that for every real at,

p lim sup I Yam(s) - Y:m.(s)I = 0.


R-M x2O
610 2.VII1. The Ito Integral

According to Section 4, the process { yam"(t), F(t), t 5 + cc } is a martingale,


with E{yamn(t)} = 1, and by Section 3(i) if 0 < a < 1, the constants c and p
can be chosen so that c > 1, p > 1, and cp(cp - 1)a2/(p - 1) < 1, and then
in view of (3.12) and (6.8)

sup E{ I yamn(t)I`} = fi. < + co.


n,(2O

Hence
E{I yam(t)I`} S lim E{Iyam"(t)i`} :5
n-W

E{yam(t)} = urn E{yam"(t)} = 1.

Thus { yam(t), .F(t), t 5 + oo } is an L` bounded uniformly integrable almost


surely continuous martingale. Now define an optional time S"(5 + oo) by

S"=infit: Jo0dw-az IoldIzds<

The martingale sampling theorem implies that E{ yam(Sj)} = 1, that is,

t=I ym(+oo)dP+ I y;m(Sn)dP


(sn=+aoy sn<+m))
z
= y;m(+oo)dP+ f expr(a 21vz f dP.
Is,,,=+m Sn<+co L` l ,Jo J
(6.9)

In view of the definition of S. the function a --. is an increasing


function of a when S. = + oo ; furthermore a -- a - a2/2 is an increasing
function when 0 < a < 1. Hence when a j 1 in (6.9), we find
z fSn
1= r y4,(+ oo) dP + e expr2 I0Izds]dP. (6.10)
%=+W, ffs.< + W) L o J

The sequence S. is an increasing sequence with limn. P{Sn = + oo } = 1


because almost every sample function is bounded. Hence when n - oo,
the first integral on the right tends to E{ym(+oo)}. The second term on the
right is at most
2
e-"E JeXp ror2 I¢Iz dsl 1
o

so when n oo, (6.10) yields E{ym(+oo)} = 1, as was to be proved.


7. Extension to Vector-Valued and Complex-Valued Integrands 611

Proof of Section 3(j). If T. is defined by (6.6) and if is in r, then the


process I OT.(-) satisfies the hypotheses of Section 3(k) because
az Jt

2 I0loriIzds5a
0

for all t. The process { yy(T, A ), .F(')) is therefore a martingale, so


f y#(-), ,F(-)) is a local martingale. o

Proof of Section 3(l). It is sufficient to prove (3.15) for in I'o and specified
by (1.7) with k = 2. In this case when t, < t S tz, the left side of (3.15) can
be put in the form

E{f. exp [Ywll) - Yzz2 t E{[w(t) - w(te)]exp[Y(w(t) - w(ti))]lF(te)}}.(6.1)111)

l
In view of the fact that
E{ze°=} = aae°''1z

when z is a normally distributed random variable with mean 0 and variance


a, the value of the expectation in (6.11) is yaz E{ yr(t,) f, } (t - t,), as it should
be according to (3.15). The verification of (3.15) for other values oft is now
trivial. Equation (3.16), an easy consequence of the Ito formula to be proved
in Section 12, can be proved at this stage using (3.15) by computing (and
finding it to be 0) the expectation of the absolute value of the square of the
difference between the two sides of (3.16). o

7. Extension to Vector-Valued and Complex-Valued


Integrands
Real Vector-valued Integrands
Suppose that { is a Brownian motion in R", with right
continuous and .(0) containing the null sets. If fi is a vector with N com-
ponents P "), . . . . fl ',, define 1#12 = as usual. Let I' be the space of
processes with state space R" whose N component processes are in the space
I' as defined in the one-dimensional case in Section 1, and define the r metric
by (1.3) with the above interpretation of the notation I 1. The corresponding
definitions of I- and r2 are made. For ¢(-) in r define
N t

O dw = 4ot dw°°.
o i o
612 2.VIII. The Itb Integral

The spaces I-', r', r2 defined in Section 2 are used below, with the same
state space R. It is left to the reader to check that at most minor modifications
of the proofs in Section 6 show that (a)-(1) in Section 3 are valid for N,-:t I
under the following conventions. In (3.9) the integrand ¢4/ is to be interpreted
as the inner product E; Ou'Viu1. In Section 3(l) y is to be interpreted as a
vector; the right side of (3.15) is to be interpreted as the inner product of
the vector y and the vector-valued integral; (3.16) is to be interpreted
correspondingly.

Complex Vector-valued Integrands


If the component processes of the integrand processes above have the
complex plane as state space, the usual conventions are to be made. That
is,'fl2 in (3.9) 0 is to be replaced by +', and 4 is then the
Hermitian symmetric inner product E; is to be replaced by 4, in
(3.15) and (3.16). Observe, however, that y`(t) is no longer necessarily a
real-valued random variable; so Section 3(g) must be dropped. Furthermore
in (3.11) the first term on the right side of the inequality should be changed
from h to 26. No other changes are necessary in Section 3(a)-(1). It is easy
to deduce from Section 3(g) and the form of that is a
supermartingale.

8. Martingales Relative to Brownian Motion Filtrations


If is a Brownian motion in R" and if the process
{!o 46 is an L2 bounded almost surely continuous martingale (com-
plex state space) whose random variables have zero expectation. It is a
remarkable fact that the converse assertion is true if .fl-) is specified ap-
propriately, as asserted in the following theorem.
Theorem. Suppose that is a Brownian motion in R^`, from the
origin, and that for all t in R' the a algebra .fi(t) is generated by f { w(s), s S 1)
and the null sets.
(a) 1f z is a (complex-valued) Ysc li .ms(s) measurable square integrable
random variable, there is a process in I-2 for which

z = E{z} + f ¢ dw a.s.
0

(b) If is an L2 bounded abnost surely continuous martingale,


there is a process in I'2 for which

z(t) = z(0) + J4'dW a.s. (8.1)


8. Martingales Relative to Brownian Motion Filtrations 613

In (b) it would be no more general to suppose that is almost surely


right continuous rather than almost surely continuous because Theorem
VII.4 implies that an almost surely right continuous martingale relative to a
Brownian motion filtration as defined in Theorem 8 is almost surely
continuous.
Observe that if 0 < /3 < + oo and if z in (a) is .F(p) measurable, the
upper limit of integration in (a) can be reduced to P. (Apply the operator
E{ - I.F(/3)} to both sides of the equation for z.) Part (b) is applicable to an
L2 bounded almost surely right continuous martingale on a
finite parameter interval [0, fl[ or [0, /3] because such a martingale can be
extended to one on the parameter interval I+ to which Theorem 8 is ap-
plicable by setting z(t) = lims_8z(s) for t >- /3 in the first case, z(t) = z(/3)
for i > /3 in the second case.
Parts (a) and (b) of the theorem are equivalent. In fact if (a) is true, then
under the hypotheses of (b) right close the martingale by setting z(+ oo)
= lim,_, z(t) and apply (a) to z(+oo) to obtain, using the fact that .F(O) is
trivial,

z(i) = E{z(+x)j.F(t)} = E{z(0)} + E{ f ^« dwI .f(t)}

= z(0) + j'bdw a.s.


o
0

If (b) is true then under the hypotheses of (a) define as the almost surely
continuous martingale and then apply (b) and the condi-
tional expectation continuity theorem to find that

z = lim z(t) = E{z} + lim idw = E{z} + f"' 0dw a.s.

To To prove (a), suppose to simplify the notation that N = I and that z is a


real random variable. Observe that the class (1' 0 dw: 4 e r'2 } (N = I, real
o
context) of random variables is a closed linear subspace of the L2 space of
real random variables, so it is sufficient to prove that if z is YSER..y(s)
measurable and is square integrable with E{z} = 0, and if z is orthogonal to
this subspace, then z = 0 almost surely.
(al) We prove first that if z(t) = E{zI.z(t)} and if oer2i then
{z( )f o dw, is a martingale. By the orthogonality hypothesis if
0 < s < t, if f is a bounded F (s) measurable function, and if 0 is in r'2,
then in view of Section 3(d)

0= E{:f4Idw} dwl = E{z(t)f f 0 dw};


= E{zf i 0
614 2.VI11. The Ito Integral

so

i
0 = E{z(I) jf" dwl.F(s)} = E{z(t) J0dwI(s)}_ z(s) Es 0 dw a.s.

This is the desired martingale equality. In particular, it now follows from


(3.16) that for every constant y the process is a martingale
with E{z(t)v.,(t)} = 0. Observe that ¢ and therefore also y can be complex
here.
(a2) If ;, and y2 are complex constants and if 0 < s, < s2. manipulate
conditional expectations to derive the first following equality and apply (al )
to derive the second:

F_{} Gc,)t' ,( 2)z} = E{y,(st)Y,,(s2)Z(s2)} = E{vrjst)vr,(st)z(s,)}


(8.2)
= E{z(s, )y,,+v2(s,) } exp (a2yt yes, ).

According to (a I). the last expectation vanishes: so (8.2) yields

E{zexp [y, w(.s,) + 72 w(s2)] } = 0. (8.3)


It follows that
E{ f [w(s, )] f 2 [w(s2)] } = 0 (8.4)

whenever f is an exponential polynomial of the form E;,=, with


a real and therefore that (8.4) is true whenever fj is a continuous periodic
function on R. Since any continuous bounded function on R is the point-
wise limit of a bounded sequence of continuous periodic functions (whose
periods may become infinite), (8.4) is true whenever f, and f 2 are continuous
and bounded on R. It then follows that (8.4) is true when f, and .12 are
the indicator functions of open subintervals of R, and we conclude that
E;zjtv(s, ). tv(s2) } = 0 almost surely. Replace the left side of (8.2) by

m
Ir{znv,.(Sj) ,0<St <S2 < ...
ll

and proceed as above to find that E{zlw(s,)} = 0 almost surely for every
finite subset .s of R+ and therefore for a countable dense subset of R'
(conditional expectation continuity theorem). Hence by the almost sure
continuity of Brownian motion E{zlw(s),se R) = 0 almost surely, equiv-
alently F_ ; z Y E a .f(s)) = 0 almost surely, and since z is by hypothesis
YEtt..f(s) measurable, this vanishing conditional expectation is almost
surely
9. A Change of Variables 615

9. A Change of Variables
Review of the Inverses of Monotone Functions

Let f be a function from R+ into B+ satisfying the following condition.


M. f is monotone increasing, right continuous, and link_. f(t) = + 00.
Define f(t) = inf {r: f(r) > t} for to R+. Then f satisfies M, and f(t) 5 r
if and only if f(r + s) > t whenever s > 0. Moreover f =,f Finally f [ f(r)] _
r if r is not a point of a constancy interval off, and Jj f(r)] = r if r is not a
point of a constancy interval off that is, if r is not a discontinuity point off.

Increasing Families of Optional Times and Their Inverses

Let { flt), t e R+} be a filtration of a probability space, right continuous


and with ."(0) containing the null sets. Let IS,, t e R+} be a right continuous
process whose random variables are finite-valued optional times for 'fl-)
and whose sample functions satisfy condition M. Define an inverse process
S. by the above procedure, so that
m
S,(cu) = inf {r: S,(w) > t}, {S, 5 r) =n {Sr+im > t}. (9.1)

The process IS., -fl-)) is an adapted process whose sample functions satisfy
condition M. The filtration is right continuous, and 9(0)
contains the null sets. Moreover {.S 5 r} e 9(r) in view of (9.1) and the
right continuity of Hence each random variable S, is optional for
Furthermore .fi(t) c .*(S,) for all t; that is, if A e9(t), then A n {S, 5 a} e
F(a) for all a, equivalently,

Ann{S,+,M>t}n{S, 5P}e.,F(P) (9.2)

for all at and P. In fact (9.2) is trivial if t > fi, whereas if t 5 fi, all the sets
in (9.2) are in .f(fi).

A Change of Variables in the Ito Integral

Let be a one-dimensional Brownian motion with variance


parameter o2, let {0(-), 9(-)) be a progressively measurable process in the
class r (Section 1), and suppose that {S, t e R } is a process as above whose
sample functions satisfy
616 2.VIII. The 116 Integral

JS,lo(s)12ds
(9.3)
0

for all tin IB+. Define


rosI
iv(t) = 0 dw,
J

and observe that E{w(t)2} = att.

Theorem. The process { iv( is a one-dimensional Brownian motion with


variance parameter a2.

Define J o 0 dt+w. To prove that {ti (),() } is a Brownian motion


with variance parameter a2, it is sufficient to show that for a < f3 the condi-
tional distribution of x(SS) - x(S,) given .f(S,) is Gaussian with mean 0
and variance a2(fl - a). Thus in view of the characteristic function of the
normal distribution it is sufficient to show that for each complex constant y
2
E ex [x S xS .F S ex a2(2 - a) a.s.; (9.4)

that is, it is sufficient to show that, in the notation of (3.10),

E{ yrm(SS)I. (SS) } _ yrm(S,) as.


In other words the problem is to show that the process { is a
martingale. Now according to Section 3(k), y4los", the process

{yy( {YY4,(Sp A '),.F(')}

is a martingale, right closable by y, ,,(S,,), and therefore uniformly integrable.


Hence (martingale sampling theorem) the martingale equality at times t = S.
and t = + oo is satisfied; that is, the process { ,F(-)) is a martingale,
as was to be proved.

Extension to N > I

If is a Brownian motion in 1 N with variance parameter a2 and


if ¢ is a progressively measurable process with state space 08 N, in the (vector)
space t, satisfying (9.3), the theorem remains true with no change in proof:
the process { is a one-dimensional Brownian motion with variance
parameter a2.
Observe that the theorem remains true if null exceptional sets are allowed,
that is, if only almost every sample function satisfies condition M and if (9.3)
9. A Change of Variables 617

is only true almost everywhere for each t [which implies that (9.3) is true
almost surely simultaneously for all t.J
JA(t)
Application (a). If N z 1 and if 101 = I in Theorem 9, then S, = t, _
.F(t), and w(t) = Yo 0 dw. We leave it to the reader to show that in this case

`, 41 div = Jo (i(,4,) dw a.s. (9.5)


Jo 0

whenever * e t (state space R) and (0¢) is the scalar multiple of the vector
0 by 0. More generally, if M(t) is for each t in R' an N x N orthogonal
matrix-valued random variable and if the component processes of are
in r, then the vector process {Jo M dw, is an N-dimensional Brownian
motion with parameter value o2.
Application (b). (N,2: 1) Let be a real vector process in r, and
define
00(t, w)
if 100(t'(0)l 0- 0'

(N-u2 .. , N-112) if 0.

w(t) = J' I0dw.


0

Then 1; so according to Application (a), the process is a


one-dimensional Brownian motion with variance parameter o2 and

JIoId=Jodw
r r

a.s.
o, 0

Thus every vector stochastic integral can be represented as a one-dimensional


stochastic integral with a positive integrand.
Application (c). (N z 1) Let be a real vector process in r, define
1; 0 dw, and define T. by (6.6). Suppose that T. is almost surely finite
valued for all a. Then according to Theorem 9, the process
is a Brownian motion. In other words the integral process is a Brownian
motion under a change of time. If 7, is not almost surely finite valued for
all a, that is, if to 10 12 ds is not almost surely + co, define

(fi(t) if t:5 k,
k(t) =
(1,...,1) ift>k,
Tka = inf it :
I
rkl2 ds Z a}, xk(t) = Jo, Ok dw.
2 J 0r
618 2.VI11. The Ito Integral

According to these definitions and Theorem 9, the process


is a Brownian motion and coincides with on the set {Ts < k} for
parameter values < a.

10. The Role of Brownian Motion Increments


Let be a Brownian motion in RN. Fix t > 0, define
Alt) = Wla)(f2-nt) - 1V(a)((I - 1)2 "t),
omitting the superscript when N = 1, and for f a complex-valued function
on R' x R define

Jj,, =f[wV(j2 "t],

fu- t )n if (j - 1)2t < S 5 j2 -"t,) -> 1?


fn (s) =
IfTw(O), 0] ifs=0.

Lemma. When f is continuous and finite valued

Jim 2" az r f [w(s), s] ds if a = fl,


p Y fUOj''Ago) = Jo
10.1
"~m j-1 0 if a # fl. ( )

If f is bounded these limits are also L2 limits.

To simplify the notation, take a = 1. Assume first that jf < c. Then


z" 2
f[w(s),s]ds - Y_ fu-11n2-"t

= lim E 1I r [.Jttv(s). s] -.f (s)] dsl2 J. (10.2)


J

limtEi If[w(s),s]-f(s)l2dr}=0.
-W o

Thus to prove (10.1) for a = fi in the L2 limit sense for f bounded it is suffi-
cient to prove

: fu_t)n(0;n'z - 2 "t)12 }=o.


lim E11j= (10.3)
10. The Role of Brownian Motion Increments 619

When the indicated square is multiplied out, each term has the form
tah - 2 n t)(Akn
f(j-1)nik-1)n(ejn (a)Z
- 2-RI), j <- k.k
If j < k, the last factor is a random variable with expectation 0 and is inde-
pendent of the other factors. Hence the expectation of the term vanishes.
The terms with j = k yield
1 z"

E Ifj-I)nI2(Ajn): - 2-nt)2l < (10.4)


j=1 f
so (10.3) is true. If f is not bounded, define g, as any bounded continuous
function on R" x R +, equal to f when If 15 m. Then

lim f gm [w(s), s] ds = JJ1w(s)s]ds a.s.


"_00
o o

The sum on the left side of (10.1) for a = fi differs from the sum with f
replaced by gm on a set of small probability when m is large. It follows that
(10.1) for a = fi is true as stated.
To prove (10.1) for a: fi with the limit in the L2 sense for If I bounded
by c, observe that (for o = 1)
11f n n
2 2
(a) (B) 21S 2 (a11 (1)'j (10.5)
E1 Y_ jU-I)nejn ejn = Y EIIjU 1),,J ejn ejn
l j=1 1 j=1

22"`

< C2 L f E{Aj(.)2Ar.)11
=2 -n,.212 ~ 0 (a
j=1

The proof for unbounded f follows that when a = J.


Observation. If N = I and J'= 1, the proof of (10.1) for a = fi, dropping
some now unnecessary details, yields
2n f = 2-"+1a412.
Y A - a21 (10.6)
j=1

An application of the Borel Cantelli theorem yields Levy's theorem, that is,
2n
lim Y A;n = o2t a.s. (10.7)
n_ j=1
620 2.VIII. The Ito Integral

11. (N = 1) Computation of the Ito Integral by Riemann-


Stieltjes Sums
Let be a progressively measurable process in r, and choose b > 0 and a
partition n: to=0<t, < .. [0,b].Define

6(n) = maxls (t1 - t1-,)

and

0(0) if t = 0,
¢R(t) _ 0(t1-,) if t1_, < t 5 t1, J 5 n,
0(t) if t > b.

Then is in r if is almost surely finite valued on it, and

J0dw = j 4(tj-()[w(t1) - W(t1-1)1

Theorem. (a) If the restriction to [0, b] of almost every sample function


is bounded and 1, almost everywhere continuous, then
b b
plim &xdw= 0dw. (11.2)
6(°)-O fo fo

(b) If E{ Ic(t)I2) < + co fort 5 b and if

lim Efl4(t) - O(s)I2) = 0, (11.3)


S-0
oss.rsb

then (11.2) is true as an L2 limit.

Proof. (a) Under the hypotheses of (a) the usual reasoning in the discussion
of the Riemann integral yields
b
lim 1¢R-0I2ds=0 a.s.;
6(m)- O
fo
so lim j(,.o r dist (¢R, 0) = 0, and since the map 0 dw is continuous
in the (T, n pair of metrics, it follows that (11.2) is true and even that

(11.4)
rs o
12. Ito's Lemma 621

(b) Apply Section 3(f) to find

f 1.0. dW 121 =
IIJ¢d/+ - UzEll I.0 -.0.12T
E o o
(11.5)
5 azb sup E{14(t) - 4(s)Iz}.
It-zIsb(.1
Oss,tsb

Therefore (b) is true, and in fact (11.4) is true as an LZ limit because the
map Jo 0 dw is continuous in the (I'z, I'z) pair of metrics. o

Integration by Parts
If is in t and if the restriction to the interval [0, b] of almost every
sample function is right continuous and of bounded variation, then

f Odw = w(b)4,(b) - w(a)4(a) - f bwdo a.s., (11.6)


o Jo

where for almost every point co of the basic measure space the integral on
the right is the Riemann-Stieltjes integral J w(s, w) d,4(s, w). In fact the
hypotheses of Theorem 11 (a) are satisfied, and the sum in (11.1) is equal to
n
w(b)O(b) - w(a)O(a) - Y w(t) [O(t) - 0(t;-1)], (11.7)
!=1

which almost surely yields the right side of (11.6) when S(ic) -+ 0.

12. Ito's Lemma


Let { 9(-)) be a Brownian motion in R', and let (q, t) -. u(q, t) be a
continuous function on RN x R' for which the derivatives

_ au au
u,=anu,' azu
uO at ' uri=aq(;,aqu,

exist and are continuous. Before proving Ito's lemma we prove an important
special case: simultaneously for all t e R',
N r a2 N f0t
u[w(t), t] - u[w(0), 0] = uo ds + Y, 1 ua dw(a)(s) + 2 Y u. ds
o a=1 o a=1

or N r
(12.1)
= Au ds + f ua dw(a)(s) a.s.
a=1 Jo
622 2.VI11. The Ito Integral

The argument in each integrand is [w(s), s]. This relation is commonly


written in the form

du = du dt + Y ua dw(a', (12.1')
a=1

and from now on relations like (12.1) will sometimes be written in the
corresponding differential form. It is sufficient to prove (12.1) to be valid
for fixed t because each side of (12.1) is the tth random variable of an almost
surely continuous process. Fix t and denotej2-"t by sj". so that
2n
u[w(t),I] - u[w(0),0] = Y {u[w('sjn),sjn] - u[W(sjn),S(J-I)n]j
j=1
(1221
2"
+ E {u[W(sjn)'S1/- lln] - u[w(Sti- lln) S(i- 1)n] If -
j=I

The jth term in the first sum is u0[w(sj"), sj"]2-"t up to an error which for
each continuous Brownian motion sample function is uniformly o(2-") as j
varies and n -+ x. Hence the first sum in (12.2) has the first integral in (12.1)
as almost sure limit when n - oo. The jth term in the second sum in (12.2)
is, if ej(n) = Nrlal(sjn) - W(a1(S(j-1).)'

N N
ua9[w(s(j-I,),s(j-!),,]A1")ej(n)

2 (12.3)
a-I a.p=1

up to an error which for each continuous Brownian motion sample function


is uniformly n(Ea=1 A('j.)') as j varies and n - oc,. According to Lemma 10 and
Theorem 1 1 , when n -+ x, the sum over j = 1, ... , 2" of the terms in (12.3)
has as limit in measure the sum of the second and third sums in (12.1).
Finally the summed error o(EaN=, 11" Aj ) is o(1) for almost every Brownian
motion sample function because the inside sum has almost sure limit cr2t
when n -" r_ by Levy's result (10.7).

Localization

If, more specially, u is a function of class C(21 on 1 N and we consider u


composed with equation (12.1) reduces to
0,2 N
u[N (t)] - u[w(0)] = Duds + Y f') uadw(a)(s)
2 0 x=1

More generally, if u is only defined and of class L(21 on an open subset D


of I8N and if P{It'(0)eD} = 1, let S be the hitting time of 8D by Then
12. Ito's Lemma 623

(12.1) is almost surely valid for t < S in the sense that if Do is an arbitrary
open relatively compact subset of D and if S0 is the hitting time of aD0 by
then (12.4) is almost surely true on {w(0)E Do} if i is replaced by t A So.
To see this, let u0 be a L(2)(R') extension of the restriction of u to Do. Then
an application of (12.4) to u0 with t replaced by t n So gives the desired
result.
We now turn to Ito's lemma. Let u and be defined as at the beginning
of this section. Let M be a strictly positive integer, and for I < m 5 M and
1 < n < N let and Wm be processes in r except that the integrability
requirement for Y'm is weakened: it is supposed only that almost every
sample function is integrable over finite intervals. Define
x"'(.)]
by
N r ,{{,

x(m)(t) = x(m)(O) Wmndw(n) +


+ Y, 0md1,
n=1 J0 fo
where x(m)(0) is an arbitrary flO) measurable random variable.

Ito's Lemma. Under the stated conditions, for u: t I-+u[x(t), t],


2 M
du = u0 dt + Ymum dx(m) + Y dx(m) dx("), (12.5)
m=1 2 m.n=1

where dx(m) dx(") is to be interpreted as the formal product under the convention

(dt)2 = dt dw(") = 0, dw(m) dw(n) = a2bmn di

so that

dx(m) dx(") = a2 j
a=1
0. W,,//,naA

To simplify the notation, we give the proof for M = N = 1. That is,


dx(t) = 0 dw(t) + /i dt, u = u[x(t), t], and (12.5) reduces to
z
du = uo+u1O+ 2u11O2)dt+u14.dw. (12.6)

It is sufficient to prove the lemma for ¢ and i/i in I70, given by representations
of the form (1.7) with a common partition :

0(t)=''(t)=0 if t<t,,
00) =4-1' 0(t)=gg_1 if t;_, <t<t;,25j5k,
0(t) _ P(t) = 0 if t > tk.
624 2. VIII. The Ito Integral

In the interval [0, tl] the evaluation (12.6) is trivial. In the interval ]t1, t2]

u[x(t), t] = u[ff, [w(t) - w(t,)] + g1 (t - t'), 1].


A glance at the proof of the special case of Ito's lemma already treated
shows that it is applicable with the help of Section 3(d) to yield

du=uodt+u1J1dw+u,g,dt+ 2

which agrees with (12.6). The verification of (12.6) on the remaining inter-
vals ]t;_,, t;] is similar, and the verification on ]tk. +oo[ is trivial.

The Algebra of Integrals (N = 1)

In view of Ito's lemma the integrals of the form

J¢dw+ tyds
0

with 0 and t' as described above form an algebra. In fact, if dx; = 4., dw +
t1 dt,

d(x,x2) = x1 dx2 + x2dx1 + (¢1O2)dr

= (x142 +x2,1)''+ (xt42 +x2411 +4'102)dt.

EXAMPLE (a) (N = 1). If n is a strictly positive integer

d(w") = nw"-' dw + n(n - 1)wn-2 dt.

EXAMPLE (b) (N = 1). If 0 is in r and if is defined by (3.10), then


dy4, = ¢y4, dw. In particular, as already proved by a different method in
Section 6, dyY = 'y.. dw for constant y. Thus the process y plays the role
of the exponential function in the stochastic calculus based on the Ito
integral.

EXAMPLE (c) (N z 1). Recall (Section 1.XV.3) that the space-time Hermite
polynomials are coparabolic and satisfy I.XV(3.8). Define ft.,_. N; as
H,,,11 .... , with

m;-8i; ifmm>0,
m;
to if m1=0.
13. The Composition of the Basic Functions of Potential Theory with Brownian Motion 625

Then
N
mN[w(t),1] = Y
!=t

Thus in the stochastic calculus based on the Ito integral the process
{Jim ..,mN[w(t),t],teR+} plays the role of the product of powers mt,...,mN
of the coordinate variables.

13. The Composition of the Basic Functions of Potential


Theory with Brownian Motion
Let u be a CZ/ function from IV into R. If is a Brownian motion
in RN from a point , Ito's lemma yields
CT2

u[w(t)]-u()=JO' <grad u,dw>+2 f studs a.s. (13.1)

{u[w(t)] - 2 JAudc;F(t)teR*} (13.2)

is a local martingale and is a martingale if

E{ f0lgradul2dsj<+oo (13.3)

for all t > 0. In particular, if u is harmonic, the process is a


martingale if (13.3) is satisfied, for example, if u is a harmonic polynomial.
More generally, if Au :5. 0, that is, if u is a C" superharmonic function, and
if (13.3) is satisfied, the process is the sum of a martingale and an
adapted process with decreasing sample functions and is therefore a super-
martingale if its random variables are integrable. If u is a C'Z' function from
an open subset D of pN into R, let D0 be an open relatively compact subset
of D, choose in Do, and let So be the hitting time of r3D0 by iv(-). There is
a Ct21 function u' on ', with compact support, extending u1bo, and if the
preceding argument is applied to u', we find that the process in (13.2) with
u replaced by u' is a martingale. This process, stopped at time So, is therefore
a martingale, and in particular if u is [super]harmonic on D, the process
{u[w(So n t 5 + oo) is a [super]martingale. In the superharmonic
context the supermartingale inequality at times 0, + oo yields the inequality
Z E{u[w(S0)]), with equality in the harmonic function context. Hence
626 2.VIII. The 10 Integral

the distribution of w(S0) is harmonic measure on 0D0 relative to . The


corresponding discussion in the parabolic context is left to the reader.
The point of the preceding discussion is that the composition of a suitably
restricted [super] parabolic function with space-time Brownian motion is a
[super] martingale and the composition of a suitably restricted [super]har-
monic function with Brownian motion is a [super] martingale. The condition
(13.3) and the associated regularity conditions are unnecessarily strong,
however. In Chapter IX we shall treat the problems of this section more
directly and in more detail, without the use of the Ito integral, and eliminate
superfluous hypotheses. This elimination can of course also be effected
using the results of this section, but each of the two approaches is too
important to omit.

14. The Composition of an Analytic Function with Brownian


Motion
Suppose that N = 2 and that { is a Brownian motion from a point
of R2, and write z(1) = w"'(t) + iw'21(t). Let D be an open subset of the
complex plane, and let Do be a relatively compact open subset of D. Suppose
that z(O) is in Do and that f is a complex-valued regular analytic function
on D. Then if So is the hitting time of tDo by Ito's lemma yields (see
the remarks on localization in Section 12)

df[z(t)] = f [z(t)] dz, t 5 So,

and { Jjz(So n t)] ; .F(t), t < + oo } is a martingale. An adaptation of Appli-


cation (b) in Section 9 shows that if we define

TQ=inf t JIrz(s2= a ,
fsoIJ
0
[Z(s)]I2ds,
0

then Jjz(T A So)] is a Brownian motion on the complex plane, killed at


time S. That is, f maps Brownian paths into Brownian paths with a new
time scale.
Chapter IX

Brownian Motion and Martingale Theory

The applications of the Ito integral in Sections VIII.12 to VIII.14 exhibit


aspects of the intimate relation between Brownian motion and martingale
theory. In the following we shall go from simple examples of this relation
to an analysis by means of martingale theory of the composition of the basic
functions of the potential theory for Laplace's equation [the heat equation]
with Brownian motion [space-time Brownian motion]. This will be effected
by a direct method without the use of the Ito integral, but there will be a
slight repetition of some of the most elementary topics in Chapter VIII.

1. Elementary Martingale Applications


Let fl-)) be a Brownian motion on 18". The following processes are
martingales relative to if the initial distribution of the Brownian motion
is chosen to make the process random variables have finite expectations,
for example, if the initial distribution is supported by a singleton.
M1
M2 { Iw(t) - , 2
- Na2 t, to 11+}, q arbitrary
M3 In)
The first process is a vector martingale (in the obvious definition) be-
cause for s < t,

E{w(t) - w(s)L (s)} = E{w(t) - w(s)} = 0 a.s. (1.1)

since w(t) - w(s) is independent of F(s) for s < t. The second process is a
martingale when N = I and therefore when N Z I because s < t implies that

E{[w(t) - q]21 F(s)}


= E{ [w(t) - w(s)]21F(s)} + 2[w(s) - q]E{w(t) - w(s)L°f(s)} + [w(s) - q]2
= a2(t - s) + [w(s) - t]]2 a.s. (1.2)

The third process is a martingale (for any complex vector ') because if s < t
628 2.1X. Brownian Motion and Martingale Theory

E{exp < y, w(t) - w(s)> I.F(s) } = E{exp <y, w(t) - v(s)> }


(1.3)
NZ
= exp a2(t - s) 2 a.s.

Conversely, if t ()} is a stochastic process for which M3 is a martin-


gale whenever the components of y are purely imaginary, then the process
has the Brownian motion finite-dimensional distribu-
tions with variance constant a2. In fact, if s < t. the martingale equality
(with y = ifJ and /3 real)

E{exPli<,tv(t))+a2t 2'jIF(s)}=expli<.w(s)>+a2.s 1JJ a.s.

(1.4)

yields

N
E{expi «, v(t) - w(s)> 19(s) If = exp [_2(( - s) ?
2' a.s. (1.5)

The difference vv(t) - iv(s) is therefore independent of ,ms(s), and it follows


that has independent increments. Moreover (1.5) implies that w(t) -
tv(s) is normally distributed with independent components, each of mean 0
and variance a2(t - s), as was to be proved.
Application of M 1. Let { be a Brownian motion on IfBN from ,
let D be a bounded open subset of lV containing , and let SS be the hitting
time of OD, almost surely finite because [Section VlI.5(c)] lim,_q, lww(t)I =
+ oo in measure. The function S{ is optional for (Section 11.4), and
therefore (Section IV.3) the stopped process {tv{(S{ A t), .3rs(t), t e 08+ } is a
martingale. Equating expectations of the random variables of the stopped
process at times 0 and t yields the equation = E{ ww(S{ A t) }, which becomes

= E{ws(S{)} (1.6)

when t -+ oc. In particular, if N = I and if D is the interval ]a, b[, (1.6)


becomes

= bP{ww(SS) = b} + aP{wt(SS) = a}, (1.7)

which implies

P{w4(S4)=b}=6-a. (1.8)
I. Elementary Martingale Applications 629

Thus the probability that a Brownian path from in ]a,b[ reaches the
boundary first at b is the solution of Laplace's equation on ]a, b[ with
boundary limit I at b and 0 at a. In other words in this elementary case the
distribution of at the first hitting place on the boundary is harmonic
measure on the boundary relative to the initial point S. This evaluation of
harmonic measure will be extended to the Borel boundary subsets of an
arbitrary Greenian subset D of R' in Section 13. That is, if trr() is an N-
dimensional Brownian motion from S in D and if S4 is the hitting time of
aD by try( ), it will be shown that S4 is almost surely finite when N = 2 and
that under the convention tr,(+ x) = oo when N > 2 the harmonic measure
evaluation (I.8) generalizes to

PIw,(SS)EAi

From this it follows that if f is a resolutive boundary function, the PWB


solution Hf is given by

Hf(S) = (1.9)

Corresponding results will be proved for PWB' solutions. The evaluation


(1.8) can also be derived by solving the appropriate differential equation
problem. The rigorous argument is the following, once it has been shown that
the left side of (1.8) defines a harmonic function of S, that is, a linear function.
Let be a Brownian motion from 0, so that S + is a Brownian motion
from S and the measure space does not depend on . A So + iv(-) path (with
a < 5o < b) which hits b before a will do so if the path is translated by
S - So With S > So : so the function -- P ps,4(S4) = b} is an increasing func-
tion of S. Since almost every is,(-) path is strictly positive at some parameter
values arbitrarily near 0, this increasing function has limit I at b, and a
similar argument shows that the limit is 0 at a. Hence the function must be
given by the right side of (1.8). This trivial example illustrates some of the
problems involved in a rigorous justification of the evaluation of Brownian
motion probabilities.
Application of M2. Let D be a bounded open subset of R", and define
Sy as above. Equating expectations of the random variables of M2 with
q= 0 at times 0 and S{ A t yields

0 = E{ I w4(S4 A 1)12 - Na2(S4 A i)}, (1.10)

and therefore
No2E{S } = Eflws(S4)J2}. (1.11)

In the special case N = 1. D = ]a, b[ this equation reduces to


630 2.IX. Brownian Motion and Martingale Theory

a2E{Se} = ( - a)(b - ), (1.12)

in view of the distribution of wf(SE) already evaluated. Thus the function


i-+E{Sf} is the solution of the equation a2 Au = -2 with boundary limit
function 0. For N > I and bounded D the martingale equality (1.11) becomes,
if we accept (1.8') (to be proved in Section 13),

a2NE(S4)=f
p
- J
aD
(1.12')

As in the case N = I, the function - E{S4} is the solution of the equation


a2NAu = -2 with the boundary limit function 0 in the sense that
a2NE{S4} + is the harmonic function which is the PWB solution of the
harmonic function Dirichlet problem with boundary function >f F--.Ir1I2.
Application of M3. Let N = 1, let be a Brownian motion from the
origin, choose a > 0, and let T be the hitting time of {a} by The density
of the distribution of T is given by VII (8.3) and will now be derived again
using the martingale M3. The optional time T is almost surely finite because
according to the application of MI above the probability that a Brownian
path from 0 hits a before -c (with c > 0) is c/(c + a), which has limit 1
when c + oo. Equating expectations of the random variables of M3 at
times 0 and T A t yields

T2 A tl
1 = E{exp [YW(TA t) - y2a2 (1.13)
D.
Take y real and strictly positive, so that the integrand is at most exp (ya).
When t - + oo (1.13) becomes

I = E{exp Lya - y2a2 2J }, (1.14)

that is

E exp [_Y2a22]}=exP(_Ya). (1.15)

This equation specifies the distribution of Tby way of its Laplace transform.

2. Coparabolic Polynomials and Martingale Theory


It was shown in Section VIII.13 that the composition of a [co] parabolic
polynomial with space-[co] time Brownian motion from a point is a martin-
gale. We shall establish this result again in this section without using the
2. Coparabolic Polynomials and Martingale Theory 631

Ito integral and incidentally gain insight into the space-time Hermite poly-
nomials. Since every space-time coparabolic polynomial is a linear combina-
tion of space-time Hermite polynomials, it is sufficient to consider these as
defined in Section 1.XV.3 by
ao
(1)m, (N)mN

n,, ...mN(7),
n=o m,+- +mN=n m(1 ... mNl
Y_ Y_

(2.1)

and to compose these polynomials with space-cotime Brownian motion from


the origin. That is, we consider

[w(t), 1],.f(t), t e R+}

with a Brownian motion in l from the origin, and we shall


prove that the process { Hm ... mN .F(-)) is a martingale.
To prove this fact, define yr(t) = exp [<y, w(t)> - jyj2a21/2] as in VIII
(3.10) so that

(1)m (N) N
yr(t) _
} m1(... MN!
H.,...mN[u'(t)] (2.2)
N
and by 1.XV(3.9) (if n = m1 + + mN)

E{Rrn rN[i'(t)]Hnl.. nN[t;(t)]}


1(a2t)n+NI2(27r)N/2m11 ... m,! if m =
(2.3)
i0 otherwise.

For fixed y and t the series in (2.2) is a series of orthogonal random variables
with mean square limit sum yr(t). Thus integration to the limit with respect
to the probability measure is legitimate. Now the process
is a Brownian motion with variance parameter j'I2 a2 ; so (Section 1, Example
M3) the profess { is a martingale. The validity of the martingale
equality for this process is an identity in y and implies the validity of the
martingale equality for each process {Hm and each of
these processes is therefore a martingale. The L2 martingale maximal in-
equality (Theorem III.11) together with (2.2) implies that for b > 0
(N)N
j ,(1)m
ao
E sup' yr(t) - Hm, mN [H (t)]
)Sb n=k m,+ +mN=n m 1 ... MN!

5 4(2na2t)N12 (ldZa2t)n.
n=k n
632 2.1X. Brownian Motion and Martingale Theory

This inequality implies that almost surely the partial sums in (2.2) converge
uniformly for t < b, ft 1 < b, for each pair b, b, .

Harmonic Polynomials

if u is a polynomial on R' and is harmonic, u considered as a function on


9' is a coparabolic (and parabolic) polynomial, and therefore the composi-
tion of u with a Brownian motion from a point of W4 is a martingale. This
result is generalized in the next section, as is the preceding result for copara-
bolic polynomials.

3. Superharmonic and Harmonic Functions on R" and


Supermartingales and Martingales

The following theorem for superharmonic functions on IRN will be extended


in Section 7 to apply to superharmonic functions on Greenian sets but is
proved separately because its proof not only is easy but indicates without
technical complications why the relation between classical potential theory
and martingale theory is so close.

Theorem. Let i _101 (-) } be a Brownian motion in RN from . Let u be a


superharmonic_junction on IRN, and suppose that either (a) u is bounded below
or (b) u satisfies for all t > 0 the integrahilit condition

JN i(t,
(3.1)
(27Za21)-N12 nN j exp -r L(Iu+,r,S)rN-' dr < +oo.
2 721
Ju

Then if +oc,. the process is a supermartingale, a


martingale if a is harmonic. If u(t) = +oo, the assertion remains true for the
parameter interval ]0, + xx F.

If a is a harmonic polynomial, (3.1) is satisfied. Thus this theorem includes


the polynomial case noted in Section 2. If u is harmonic, case (a) is trivial
because a one-sided bounded harmonic function on RN is identically constant
according to Section 1.11.2.
To treat case (a), suppose that u is lower bounded and apply the super-
harmonic function inequality L(u, r. -) < u(-) to the equality in (3.1) with
lul replaced by u to obtain

f(t,-,tc) (3.2)
3. Superharmonic and Harmonic Functions on R" 633

which is precisely the supermartingale inequality, so that is


a supermartingale over the set of parameter values t making E{u[w(t)] }
finite. If u(C) < + oo, the inequality (3.2) at C yields E{u[w(t)] } = t (t, C, u) <
+oo; so is a supermartingale on the parameter interval R+. If
oc, this argument fails, but the following inequality shows that
E{u[w(t)] } < + co for t > 0 so that {u[w(t)],.f(t), t > 0) is a super-
martingale:

(3.3)
_ (27ra21)-NIA (u, 1,C) + L(u, 1,C) < +oc.

Here we have used the monotoneity of the function L(u, , C).


To treat case (b), suppose that u is superharmonic and satisfies (3.1),
which is simply the statement that Eflu[w(t)] 1} < + oo. Then (3.2) is again
valid; so is a supermartingale as stated, with the parameter value 0
omitted when u(C) = +oo. Finally, if u is harmonic and satisfies (3.1), this
discussion is applicable to both u and - u ; so is a martingale.

Application to Potentials

If N > 2, a superharmonic potential Gp of a measure µ satisfies the lower


boundedness condition (a). If N = 2, a superharmonic potential Gp satisfies
(3.1). In fact µ(R2) < + oc because G1t is superharmonic, and the following
inequalities (3.4) and (3.5) show that (3.1) is satisfied.

ff
,1 R=
At, C, q)12WO filly-0:5 log I 1 - CI -5 (dC)
I

(3.4)
<- (27ra21)-1 f µ(dC) f log 11 - CI-'12(d7)
,1 a2 Jlle clstl
(4a2t)-1µ(R2)
_ < +oc.

Use this inequality and set a = (7ra2t/2)112 to derive

fa2
logMn-Clµ(dC)

-C112(di1)+(4a2t-1u(R2)

f p(dC)JRz 1n

< f',q(dC) logIfRI


R
634 2.1 X. Brownian Motion and Martingale Theory

<I P(110 log ?'(t.s.q)(I q - I + I - sl)/2(dt1)] + (4a2()-tp(R2)

.R {fit =

p(Q82)
= Jlog [a + I - CI ]u(d() + (4721)-'

< [log(2a) + (4x21) -']µ(Q82) + , log(2IS - SI)u(dr). (3.5)


J
04-0-21

the last line is finite because the potential of the projection of u on the
exterior of B(5, a) is harmonic on that disk and so is finite at g.

Parabolic Context

Theorem 3 is easily translated into the parabolic context. Let


be a space-time Brownian motion in ft' from 5 = s). Let u be a super-
parabolic function on a half-space {ord it < c1 of 9' containing , and
suppose that either (a) u is bounded below or (b) a satisfies (for all t > 0)

/(t. S, q) I u(q. s - t)IIN(dq) < + x . (3.6)


J_R'V

Then if u(5) < +x,, the process is a supermartingale, a


martingale if ti is parabolic. If ti(p) = + x, the assertion remains true for
the parameter interval ]0, + xf. The proof follows that of Theorem 3.

Application to the Hitting of a Point by a Brownian Path

We have seen in Section VI1.6 that almost every path of a Brownian motion
in 11 hits every point of R at arbitrarily large parameter values. When N > I
we shall now show that almost no path of a Brownian motion in Q8' hits a
specified point q of Q8" at a strictly positive parameter value. It is sufficient
to consider Brownian motions from a point of Q8N. The function u = G(q. )
is superharmonic on 118" with value + x at q. This function is positive when
N > 2 and satisfies the integrability condition (3.1) for every point (q not
excluded) when N = 2. If is a Brownian motion from f. the
process ; is therefore a supermartingale except that the param-
eter value 0 is to be excluded if S = q. This supermartingale is almost surely
continuous, so (Section IV. I) almost every sample function is finite valued.
Hence almost no path ever hits q except when = q and the parameter
value is 0. It follows that if N > I and if A is a countable subset of 68",
almost no path of a Brownian motion ever hits A at a strictly positive param-
eter value. This result will be extended to polar sets A in Section 5.
4. Hitting of an F, Set 635

4. Hitting of an F. Set
Let A be a subset of R", and let A) be the probability (if this probability
is defined) that a Brownian motion from hits A at a strictly positive time.
The function A) is defined and Borel measurable if A is an F. set (Section
VI.6), defined and universally measurable if A is an analytic set according to
the same section. We suppose that A is an F. set in the present section to
show how far one can go using relatively elementary methods. See Section 9
for A analytic; such generality will not be needed in the application to be
made in Section 5.

Lemma. If A is an F. subset of R" for N Z 2, the function A) is super-


harmonic, harmonic on R" - A.

[If N = 1, the function A) is identically I when A is not empty.] If


N > 1, let B be a ball in R" with center , and let TT be the hitting time of
aB by the Brownian motion from . Since Brownian motion is rota-
tionally symmetric about its initial point, the distribution of ww(T{) is the
uniform distribution pa(i , ) (harmonic measure) on B. To show that A)
is harmonic on R" - A, suppose that B c R" - A. Apply the strong Markov
property of Brownian motion to find that the probability that hits A,
which is the same as the probability that hits A after time TE, is given by

A) = E{u[ww(TS)] } = µ8(i;, u). (4.1)

Thus A) has the harmonic function average property on R" - A and


hence is harmonic there. To show that u is superharmonic on R' observe
that if A is an F. set, the set A - B is also an F. set, and A - B) is harmonic
on B according to what has just been proved. Since A - B) 5 A) and
there is equality in the limit when B shrinks to its center because (Section 3)
0, the function A) is lower semicontinuous at . Moreover
u(i;, A) is at least equal to the probability that a path hits A after time
T,; so in the present context (4.1) is modified to the superharmonic function
inequality
A) z E{u[ww(TT)] } = u). (4.2)

The function A) is therefore superharmonic on R", as was to be proved.

Parabolic Context

Let A) be the probability, if this probability is defined, that a space-time


Brownian motion from hits the set ,4 at a strictly positive time. In view of
the fact (Section VII.12) that parabolic measure on an interval boundary
636 2.1X. Brownian Motion and Martingale Theory

relative to a point of the interval is the hitting distribution on the boundary


of a space-time Brownian motion from the reference point, the method of
proof of Lemma 4 with the ball B replaced by an interval shows that when
A is an F,, subset of A' for N 1, the function ti(-, 4) is superparabolic on
AN, parabolic on AN - A. In particular, if A = A x R with A an F, subset of
RN, the set A is an F, subset of ft N, and s), A) = A) for all s.

5. The Hitting of a Set by Brownian Motion


Theorem. Let be a Brownian motion on RN.
(a) If A is a polar subset of RN, almost no path meets A at a strictly
positive parameter value.
(b) If N > 2, lim,_. w(t) = ao almost surely. Hence almost every
Brownian path from a point of a Greenian subset D of RN either hits
a finite boundary point of D or lies in D and has the boundary point
w as a limit.
(c) If N = 2, almost every path hits every disk at arbitrarily large
parameter values.

Observation (N = 2). According to (a), almost no Brownian path from a


point of an open subset D of R2 hits OD if D is not Greenian. Conversely
(to be proved in Section 10), almost every Brownian path from a point of D
hits OD in a finite boundary point if D is Greenian.

Proof of (a). If A is a polar subset of RN and N > 2, there is a positive super-


harmonic function u on R', identically + co on A (Theorem 1.V.2). Choose
6 > 0. According to Theorem 3, if is a Brownian motion from , the
process {u[w{(t)],t z S} is a supermartingale, and according to the sub-
martingale maximal inequality (Section 111.9) applied to the negative of this
supermartingale, the restriction to the rationale in [6, 1/6) of almost every
sample function is bounded. Hence the same assertion is true if
is replaced by a Brownian motion with an arbitrary initial distribu-
tion. Since u is lower semicontinuous, this function is continuous at each of
its infinities, and it follows that almost no path can hit A at a strictly
positive parameter value. If N = 2, it will be convenient and sufficient to
assume that A is bounded. According to Theorem 1.V.2, there is a measure
u whose potential u = Gp is identically + oo on A, and a glance at the proof
shows that u can be chosen with compact support if A is bounded. In view
of the application to potentials in Section 3 the function u satisfies (3.1), and
the proof of (a) now goes through as in the case N > 2.

Proof of (b). It is sufficient to consider a Brownian motion from a


I'iJ-1+2
point . If N > 2 and u(PI) = the function u satisfies the hypotheses
6. Superharmonic Functions, Excessive for Brownian Motion 637

of Theorem 3 so that is a positive almost surely continuous super-


martingale, omitting the parameter value 0 if = 0. Hence (Section II1.15)
lim,., u[w4(t)] exists and is finite almost surely; that is, lim,_m Iww(t)I exists
almost surely, and this limit is + oo because (Section VII.5(c)) the limit is
+ oo in measure.

Proof of (c). Let B be a disk, and let B) be the probability that a Brownian
motion wt(-) from ever hits B. According to Section 4, the positive function
B) is superharmonic on 982, and it follows (Section 1.I1.13) that B)
is identically constant, necessarily identically 1 because this function is 1 on
B. Thus almost every Brownian path from hits B at a strictly positive time.
Ifs > 0, the process {wf(s + t), t e 98+ } is a Brownian motion with initial
distribution the distribution of wt(s), and so almost every Brownian path
hits B after time s; that is, almost every Brownian path from a point of 982
hits B at arbitrarily large parameter values and therefore hits every disk
with rational radius and rational center at arbitrarily large parameter values.
Part (c) follows.

Part (c) implies that the cluster set of almost every plane Brownian path
as the parameter value becomes infinite is the whole plane. According to
Theorem 10 below, (c) is true with "disk" replaced by "analytic nonpolar
set."
Theorem 5 exhibits the special nature of dimensionality 2 in classical
potential theory. Dimensionality I is also special, for example, in that only
the empty set of R is polar; so Theorem 5(a) becomes trivial when N = 1.
Theorem 5(c) is almost as trivial because almost every Brownian path is
unbounded positively and negatively. On the other hand the cases N = 1, 2
are not special for the parabolic version of Theorem 5.

Parabolic Context

Theorem 5(a) is true in the parabolic context, that is, almost no space-time
Brownian path meets a parabolic-polar set at a strictly positive time. The
proof of Theorem 5(a) is applicable, and simpler in the parabolic context
in that the proof given above for N > 2 is valid in the parabolic context for
N z 1. The parabolic version of Theorem 5(b) is trivial: if is a space-
time Brownian motion lim,_. ord w(t) = - oo. See Section 15 for the hitting
of a parabolic-semipolar set by space-time Brownian motion.

6. Superharmonic Functions, Excessive for Brownian Motion


Theorem. If u is a positive superharmonic function on the open subset D of
RN, then u is excessive jor 6D, and GD(t, , u) < + oo for t > 0.
638 2.1X. Brownian Motion and Martingale Theory

See Section 8 for the converse theorem. The condition that u be excessive is

(a) GD(t, , u) 5 u,
(6.1)
(b) lim GD(t, , u) = 11,
t-o

and we show first that (6.1b) and the finiteness of GO, , u) follow from
(6.1a). For each point of D let be a Brownian motion in D
from , with lifetime S (Section VII.9); for example, this process can be a
Brownian motion in U3 killed at the hitting time SS of 8D. We can assume
that contains the null sets and that is right continuous. Since u
is lower semicontinuous and positive and is almost surely continuous,
Fatou's lemma yields

(6.2)
t-o t-o

and therefore (6.1 a) implies (6.1 b). The finiteness of GD(t, 5, u) follows from
(6.1 a) if co. If oe, let B be a ball with center S and closure
in D. The function tBu is positive and superharmonic on D, finite on B,
equal to u on D - B. Apply (6.1a) to tBU to find that

GD(t, , GD(t, , tBU) + GD(t, , u l B) < G(t, , u l B) < + ao (6.3)

in view of the fact that u is IN integrable over B. Thus to prove the theorem,
it is sufficient to prove (6.1a).
In the following proof we refine an example given in Section VIA. Let
Bu(ry) be the ball with center p in D and radius e A (In - 8DI/2), in particular,
of radius a if D = pN. Choose a point q at which u is finite, and define

To = 0, inf{t > w,,(i)eBjwjT,J)}, n > 0.

The function T, is optional for because T, is the hitting time by the


almost surely continuous process { of a closed set. The function
T2 - T, is the hitting time of a closed set by the almost surely continuous
process

([ww(Ti + ), w+',,(T1)], -,F,,(T1 + ')}

with state space D x D. Hence T2 - T, is optional for .,F,,(T, + ), and


therefore (Section II.2h) T2 is optional for Proceeding by induction
we find that T. is optional for for all n. In view of the strong Markov
property the discrete parameter process {t',,( T ), .F,t(T ), n e Z+) is a Markov
process with state space D and stationary transition function q, where
6. Superharmonic Functions, Excessive for Brownian Motion 639

q(q, ) is the uniform probability distribution on tB,(q). Since q(q, u) =


L(u,q,e A (Iq - aDI/2)), the process {u[w,(T)],.F,(T )} is a supermar-
tingale. Obviously lim,,..m T. = S, almost surely. Choose i > 0 and define

{min {n: T. t} if S, > t


k--
1 +00 otherwise.

Then k is optional for and lim,-o Tk = t almost everywhere where


S, > t. If u[w,(T+m)] is defined as 0, the supermartingale inequality at times
0 and k (optional sampling theorem) yields

u(q) Z E{u[w,(Tk)] }. (6.4)

When e -+ 0, this inequality yields, in view of the lower semicontinuity of u


and Fatou's lemma,

u(q) E{u[w,(t)]; S, > t} = r.D(t, q, u); (6.5)

so (6.1a) is true, as was to be proved.

Probabilistic Interpretation of Theorem 6

In probability language Theorem 6 states the following. Let u be a positive


superharmonic function on an open subset D of R", let be a point of D,
let be a Brownian motion in R" from , and let SS be the hitting
time of 8D (Euclidean boundary) by Then the process

{U[w (t)]I}s4>r) ; {(t),tER+}

is a supermartingale except that the parameter value 0 is to be omitted if


u(4) = +oo. Equivalently, if is a Brownian motion in D from
, with lifetime S4, in which case wi(t) has domain of definition {SS > t}
when t > 0, we can rephrase the result by stating that

{u[w{(t)]; t E R+ }

is a supermartingale if defined as 0 for t SS, except that the parameter


value 0 is to be omitted if u(t) = +co.

Parabolic Context
Theorem 6 and its probabilistic interpretation and proof translate directly
into the parabolic context except that balls in the proof of Theorem 6 are
to be replaced by intervals.
640 2.IX. Brownian Motion and Martingale Theory

EXAMPLE. Let { .t be a Brownian motion in 08' from the origin, let


So be a point of 08' other than the origin, and define u(s) = I - oI-'. Then
u is a positive superharmonic function on 081, and (by Theorem 3 or Theorem
6) the process is an almost surely continuous supermartin-
gale, trivially right closable by 0. This supermartingale has the following
properties.
(a) is L2 bounded (and therefore uniformly integrable because
the function ri-'r2 is a uniform integrability test function on 08+). In fact,
if B = B(to, and if e(t, ) = 6(t, 0, ), then

J (W(fIE R'-S)
dP = J
v -B
e(t, ) I -
f uz[w(t)]dPS f f(t. 2 - oI 13 (d )= 3I 2
JIw(1)EBI JB \ \ /
(b) lim, u[w(t)] = 0 almost surely because [by Theorem 5(b)]
lim,_,,w(t) = co almost surely.
(c) Choose 6 < and let Td be the hitting time of S) by w(-).
Then, under the convention that u[w(+ec)] = 0, the process

{u[w(T, A t)];.F(t),te08+} (6.7)

is a martingale, trivially bounded and therefore uniformly integrable. In


fact the process (6.7) is a supermartingale because it can be obtained by
stopping at time T6 the supermartingale right closed by 0, and the
process (6.7) is a submartingale because 2/6 - u[w(Ta A )] is a supermar-
tingale on the parameter set 08+. The latter process is a supermartingale
because if is w(-) killed at T,,, to obtain a Brownian motion in 08' -
B(o, 6/2) and if u' is the restriction of it to 08' - 6/2), then the compo-
sition 2/6 - of the positive harmonic function 2/8 - uwith
is a supermartingale which stopped at time Ta yields 216 - u[w(Ta A )].
[Assertion (c) is a consequence of Theorem 7(c) below because u' is a
bounded harmonic function on 083 - B(o, 6) and is continuous on the
closure of its domain, but it seemed more natural to prove (c) at this stage.]
(d) E{u[w(Ta)] } because this is the martingale equality for the
process (6.7) at times 0, + oo.
(e) {u[w(t)];,f(t), te08+} is not a martingale because if this uniformly
integrable process were a martingale it would be right closable by its limit
0 at i = + co (Section Ill. 14) and would therefore almost surely vanish
identically.
(f) The process in (e) is a supermartingale potential because it is uni-
formly integrable; so there is L' convergence to 0 in (b).
(g) The process in (e) is a singular potential because Ta = + co
almost surely since (Section 3 or Theorem 5) almost no w(-) path hits o,
7. Composition of a Superharmonic Function with Brownian Motion 641

and therefore the process in (e) is a potential which in view of (c) is a local
martingale and as such (Theorem V.12) is singular.
(h) The process in (e) is not in the class D. In fact, if this process were
in D, we could go to the limit (h -+ 0) in (d) to find, since

lim P{u[w(TT)] = 0} = 1,

the false evaluation 0.

7. Preliminary Treatment of the Composition of a


Superharmonic Function with Brownian Motion;
A Probabilistic Fatou Boundary Limit Theorem

Theorem 6 makes it possible to analyze by martingale theory the composi-


tion of a superharmonic function with Brownian motion. "Preliminary" in
the section title refers to the fact that although in Theorem 12 below it will
be shown that almost every sample function of such a composition is contin-
uous, only right continuity will be proved in this section. Specifically, the
process in (7.1) below is not only almost surely right continuous, as
proved in this section, but is in fact almost surely continuous, and the process
in (7.1) is almost surely continuous except for a possible jump discon-
tinuity at the parameter value S. Theorem 7 is also preliminary in that
(Theorem X.8) it is valid in the more general context of relative superhar-
monic functions and conditional Brownian motion.
In Section 1.XII.19 it was proved that if h is a strictly positive harmonic
function on a Greenian subset D of R N and if v is a positive superharmonic
function on D, then v/h has a minimal-fine limit at Mh almost every point of
the Martin boundary of D. We shall see (Theorem 3.111.4) that an equivalent
probabilistic formulation of this theorem is Theorem X.8, which does not
involve a boundary and which states that v/h has a limit along certain condi-
tional Brownian paths. When h = 1, this result is part (a) of the following
theorem.

Theorem. Let D be a Greenian subset of R", let be a point of D, and let


be a Brownian motion from . It is supposed that 9,(0) contains
the null sets. Define

S4 =sup{t>0:w{(s)eDfors<t},
and let u be a positive superharmonic function on D with potential, singular
harmonic, and quasi-bounded harmonic components. respectively, up, urns, and
u,,,gb, so that u = up + urns + urngb. Then
642 2.IX. Brownian Motion and Martingale Theory

(a) lim,ts4 u[wt(t)] exists (finite) almost surely.


Define 3rs(+ co) = YIER` .FS(t),

x{(t) = xs(t) = u[ww(t)] if t < SS;


x'(t)
q = 0, X4(1) =limu [ wf(s)] ifS4 S 1<+co. (71)
stse

(b) The processes A,(-)) and A,(-)) are almost surely right
continuous supermartingales except that the parameter value 0 is to
be omitted if +oo. (The second process will be shown in
Section II to be almost surely continuous.)
(c) If u = up or u = u,,,s, the limit in (a) vanishes almost surely. I fu = u,,,qb,
the process A,(-)) is a uniformly integrable martingale.
(d) u,,,qb() = E{lim,tssu[wt(t)]}.

Observation. When SS 5 t 5 + oc, the random variable xs(t) can be de-


fined arbitrarily on the null set on which the limit in (a) does not exist. Note
that if u is necessarily continuous at %; so
and are almost surely continuous at the parameter value 0 in this case,
even though this parameter value must be excluded in the supermartingale
assertion.

The Optional Time SS


This random variable is the hitting time of the Euclidean boundary by
The equivalent definition in the theorem was formulated to stress that the
theorem does not involve any specific boundary of D. When N > 2, the
hypothesis that D is Greenian reduces to the hypothesis that D is nonempty
and open. In this case SS may be infinite valued with strictly positive proba-
bility and, in fact, is almost surely + oo if D = R". When N = 2, the hypo-
thesis that D is Greenian, that is, (Theorem I.V.6) that RZ - D is not polar
and that D is not empty, is made to avoid trivialities. In fact, if D is not
empty and not Greenian, then a positive superharmonic function on D is
necessarily identically constant; so the theorem is trivially true. Finally
according to Section 10 below, SS < + co almost surely when N = 2 and D
is Greenian.

Proof of (a) and (b). The conditional expectation supermartingale inequality


for is precisely (6.1a) except when the parameter value + co is involved,
in which case the inequality becomes trivial because xS(+oo) = 0 almost
surely. Moreover according to Theorem 6,

E{xr(t) } = GD(t, t, u) < + 00

when t > 0. It follows that the process is a supermartingale if


7. Composition of a Superharmonic Function with Brownian Motion 643

the parameter value 0 is omitted when oo. Throughout the follow-


ing proof we assume that co ; the modifications to be made in the
arguments when u(g) = + oo will be obvious. Let u be an increasing sequence
of finite-valued positive continuous superharmonic functions on D with
limit u (Theorem 1.IV.10). Denote by the primed process (7.1) with u
replaced by u,,. According to what we have just proved, the process
is a supermartingale, almost surely right continuous because
is almost surely continuous. Thus the process is the limit
of an increasing sequence of almost surely right continuous supermartin-
gales, lim, x;,,(t), and is therefore almost surely right continuous
(Theorem IV.4). Since an almost surely right continuous supermartingale
almost surely has finite left limits (Theorem IV.1), it follows that (a) is true.
Thus can be defined by (7.1) and is almost surely right continuous. Let
D be an increasing sequence of open relatively compact subsets of D with
union D and with in Do, and let be the hitting time of OD,, by
The process stopped at that is, the process

{x;,S(t n S 4); . 4(1),1 c - R },

is an almost surely right continuous supermartingale (Theorem IV.3). Hence

E{x;,,(t A E{x;,,(t A n as. (7.2)

when 0 < s < 1. Furthermore, for each parameter value t, x;,4(1 A S.4)=
x,(1) almost surely. Apply Fatou's lemma for conditional expectations
in (7.2) when n -a oo to find that the almost surely right continuous process
is a supermartingale. (According to Theorem I I below, this
process is actually almost surely continuous.)

Proof of (c) and (d). Denote the expectation in (d) by The supermar-
tingale inequality for at the pair of times 0, + oo yields u Z u'. The
reader is invited to prove, without analytic set theory, that u' is Borel mea-
surable. It follows from the general theory in Section VI.6 that u' is univer-
sally measurable, which is all we shall need. If 6) c D and if T is the
hitting time of S) by the strong Markov property of Brownian
motion implies that w,(T+ ) is a Brownian motion with initial distribution
the distribution of w,(T), the uniform distribution on 6) in view of the
spherical symmetry of a Brownian motion about its initial point. Hence

E {E{lim u[w,(t)] I.F,(T) } } = L(u', , 6).


ItS4 JJJ

Since u' (< u) is finite IN almost everywhere on D, this function is harmonic;


that is, u' is a harmonic minorant of u. If u = up is a potential, then GMD u = 0
(Section 1.IV.3); so u = 0, and the limit in (a) vanishes almost surely. If
u= is a singular harmonic function,then for every constant c the function
644 2.IX. Brownian Motion and Martingale Theory

u A c is a potential (Section 1.IX.10), and therefore again the limit in (a)


vanishes almost surely. If u is bounded, say u 5 a, the inequality u z u' can
be applied to both u and a - u to find that u = u'. Similarly the supermar-
tingale inequality can be applied to both x,(-) and a - to find that
is a martingale. If u = umgb is a quasi-bounded harmonic func-
tion, the limit of an increasing sequence u of positive bounded harmonic
functions on D, then the fact that the process in (7.1) with u
replaced by u is a martingale implies that (n -+ oo) the unprimed process
in (7.1) is a martingale, necessarily uniformly integrable because it is right
closed [Section 111.3(e)]. The equation in (d) is the martingale equality for
times 0, + co. o

The Vanishing of a Superharmonic Potential at the Boundary of Its


Domain

This vanishing can be phrased in various ways. For example, in view of the
Riesz decomposition of a positive superharmonic function u on a Greenian
set D the function is a potential if and only if GMDU = 0; equivalently
(Section 1.VIIl.1I), when B is an increasing sequence of open relatively
compact subsets of D with union D, the positive superharmonic function u
is a potential if and only if

lira re,,u = lira u) = 0.


-m ft-M

That is, u is a potential if and only if u has limit 0 at the boundary in


the L' sense relative to harmonic measure. According to Theorem 13
below, if T,, is the hitting time of BB by a Brownian motion from
in Bp, then µ8n(, u) = so u is a potential if and only if
E{u[%,,(T, )]) = 0. This condition is again an L' convergence condi-
tion. (It is sufficient in the preceding discussion if the L' condition is satisfied
for a single point in each open connected component of D.) According to
Theorem 7(c), a superharmonic potential u has limit 0 along almost every
path to the boundary. This condition on a positive superharmonic
function u, if not reinforced by some kind of L' convergence condition, is
not sufficient to make u a potential, however. For example, consider for
D = B(0,1) the Poisson kernel function corresponding to the boundary
point C,
l_ z

U(O

The function u is positive and harmonic on D and has limit 0 at every bound-
ary point except i ; so although u is not a potential, u has limit 0 along almost
every Brownian path from a point of D to the boundary.
8. Excessive and Invariant Functions for Brownian Motion 645

Parabolic Context

Theorem 7 and its proof translate directly into the parabolic context.
Observe, however, that "right continuity" will not be strengthened below
to "continuity" in the parabolic context. That is, if u is a positive super-
parabolic function on an open subset D of AN, then the parabolic context
analog of the process is almost surely right continuous but not neces-
sarily almost surely continuous, as shown by an example in Section 12.

8. Excessive and Invariant Functions for Brownian Motion


Theorem. /f D is an open subset of RN and iJ' u is an excessive [invariant
excessive] function for ttD, then on each open connected component of D the
function u is either superharmonic [harmonic] or identically + oo.

See Section 6 for the (partial) converse theorem. It can be assumed in the
proof that D is connected. Suppose first that u is excessive continuous and
bounded on D, and let be a point of D. Then if is a Brownian
motion in D from and if is defined by (7.1), the process
is an almost surely right continuous supermartingale. If b) c D and if
T(S) is the hitting time of 6) by the Brownian motion, the supermartin-
gale inequality for times 0 and T(S) yields the superharmonic function
inequality u(a) >_ L(u, , c) because ws(T(cS)) is uniformly distributed on
(S) (spherical symmetry of Brownian motion about its initial point).
Hence u is superharmonic. In the general case define
,In

un = n 4D(t, , u A n)lt (dt). (8.1)


0

According to Section VI.10 the function un is excessive for 6D and u, is an


increasing sequence with limit u. Moreover un < n, and in view of the
continuity properties of 6D (Section VII.9) the function un is continuous.
According to the special case of Theorem 8 just treated, the function un
is superharmonic, and it follows that u, as the limit of an increasing sequence
of superharmonic functions, is either superharmonic or identically + or,
If u is invariant excessive and not identically + oo, so that u is superharmonic
and positive, we know (by Theorem 6) that BD(t, , u) < + oo for t > 0. Hence
u = 6(t, , u) is finite valued, and the process {xj(t), .JF4(t), t E R+ } is an almost
surely right continuous martingale. Note that the parameter set does not
include the point +oo. The martingale equality for this process at times 0
and t A T(h) is

u() = f u[w,(T(c5))] dP + J u[w,(t)] dP. (8.2)


(T(a)s) (T(b)>q
646 2.IX. Brownian Motion and Martingale Theory

The second integral is at most

r2r)-Nl2 I
D(t, . p)IN(dry) < (27r

and therefore tends to 0 when t -+ + oo. The first integral in (8.2) increases
to L(u. , <i) when t + oo. and since u is locally IN integrable, it follows
that u is harmonic, as asserted.

EXAMPLE (a). Let u be a positive superharmonic function on the Greenian


set D. Then u is excessive for ID; so the function u) is monotone
decreasing. The function uo = lim,_, tD(r, -, u) is an invariant excessive
function (Section V1.10) and is therefore a harmonic minorant of u. A
probabilistic interpretation of uo will be given in Section X.1.

Parabolic Context

Theorem 8 in the parabolic context becomes: If D is an open subset of QAN


and if ti is an excessive [invariant excessive] function for in, then ri is super-
parabolic [parabolic] on D if each point of D is below (relative to D) some
point of Initeness of u. The proof follows that of Theorem 8 except that balls
in 68' are replaced by intervals in A'. For example, (8.1) becomes in the
present context

t I1 a") nn'In.tE n n]IN(thl). (8.1')


1 in l)1

and it, is continuous on D. Observe that even if a is not superparabolic on


D, the function a is lower semicontinuous there and is superparabolic on
every open set strictly under (relative to b) a point of finiteness of a. When
a is invariant excessive for tD, the function a is parabolic on every open set
strictly under (relative to b) a point of finiteness of ti.

EXAMPLE (b). If D = A', the function a on D defined as + oo or 0 according


as the ordinate is strictly positive or not is parabolic excessive but is not
superparabolic.

EXAMPLE (c). Let be a Brownian motion in IIi' with parameter set 118,
and let u(-, s) be the distribution density of v(s) relative to IN, as defined in
Section V11.2. Equation VII(2.3) satisfied by a shows that a is invariant
excessive for space-time Brownian motion; so a is parabolic on 6FN. Con-
versely, direct calculation shows that a minimal positive parabolic function
on RJ' . given by I.XVI(8.I ). is invariant excessive for space-time Brownian
9. Application to Hitting Probabilities and to Parabolicity of Transition Densities 647

motion. In view of the integral representation I.XVI(8.2) of an arbitrary


positive parabolic function in terms of minimal ones it follows that every
positive parabolic function on ERN is invariant excessive for space-time
Brownian motion.

9. Application to Hitting Probabilities and to Parabolicity of


Transition Densities

(a) Hitting Probabilities. Let D be a Greenian subset of R'. and let A be


an analytic subset of D. Let be a Brownian motion in D from , and
let A) be the probability that a path hits A at a strictly positive
time; that is, A) is the probability that an unkilled Brownian path from
hits A before it hits D. Then the function A) is superharmonic on D
and harmonic on D - A. It was shown in Section 4 how this fact can be
proved when D = R' and A is an F. set. The method of proof is applicable
in the general case, but we remark that if A is an F. set, the theory of analytic
sets need not be invoked. At the present stage a second proof that A)
is superharmonic can be given by noting that (Section VI.12) this function
is excessive for Brownian motion and so is superharmonic by Theorem 8.
It will be shown in Section 14 that A) = R'. The translation of this
discussion to the parabolic context is trivial: "superharmonic" becomes
"superparabolic" and so on.
(b) Parabolicity of Transition Densities. According to Section VII.10, if D
is a nonempty open subset of R', if D = D x IR, and if n is a point of D.
then the function q) is excessive for space-time Brownian motion in
D, invariant excessive for space-time Brownian motion in D - ; (q, 0) and
continuous and 0 at the points of D - {(ri,0)} with ordinate values <0. It
follows that ry) is superparabolic on b and parabolic on D - {(p, 0)}.
More generally. if D is a nonempty open subset of 9" and if it is a point of D,
the fact (Section VII.l0) that ri) is excessive for space-time Brownian
motion in D, invariant excessive for space-time Brownian motion in ,6 - {7},
and continuous and 0 at the points of ,6 - {il} with ordinate values <ord il,
implies that n) is superparabolic on D and parabolic on D - {n}.
Now apply the Riesz decomposition to find that

n) = ry') + 6M, (9.1)

Since [notation of VII(10.3)] j4'4(1'4) - rij >- I8D - ry'1, the last term in
VII(10.3) defines a bounded function of g. It follows that c = I and that the
last term in VII(10.3) defines a parabolic function of in D. In Section 17
we shall show that t`j = G6, that is, the last term in (9.1) vanishes identically.
648 2.IX. Brownian Motion and Martingale Theory

10. (N = 2). The Hitting of Nonpolar Sets by Brownian


Motion
The following theorem strengthens Theorem 5(c).

Theorem. (a) Let be a Brownian motion in R2, and let A be a plane set
which is not inner polar (for example, A may be analytic and nonpolar). Then
almost every path hits A at arbitrarily large parameter values.
(b) If a plane open set D is Greenian, almost every Brownian path from a
point of D hits the boundary. If D is not Greenian, almost no Brotvnian path
from a point of D hits the boundary.

Proof of (a). It can be assumed that vv(-) = ;t is a Brownian motion from


some point l:. It can also be assumed that A is compact and nonpolar. Then
its complement D is Greenian (Theorem 1.V.6). Choose any point q in the
same open connected component of D as l:. The function u = G0(ry, ) A 1
is a positive nonconstant continuous superharmonic function on D. Accord-
ing to Section 6, if S4 is the hitting time of 8D, the process {u[wc(t)]
t e R' } is a supermartingale. The sample functions of this supermartingale
are almost all right continuous, even continuous except for a possible jump
at the parameter value S4 if S4 < + oo. Such a supermartingale has almost
surely a limit as the parameter becomes infinite (Section 111.13). If
P{S, = + co } > 0. there is a iv,(-) path with the following properties:
(i) The path never meets A.
(ii) u has a limit c on the path as the parameter becomes infinite.
(iii) The path meets every disk at arbitrarily large parameter values.
But then D is connected, and in view of the continuity of u this function
must be identically c, a contradiction. Hence Sc is almost surely finite valued;
that is, almost every iv,(-) path hits A. The proof of (a) is concluded in the
same way as the corresponding part of the proof of Theorem 5(c). o

Proof of(b). If D is Greenian, 082 - D is not polar (Theorem I.V.6); so by


(a) of the present theorem almost every Brownian path from a point of
D hits the boundary. Conversely, if D is not Greenian, 082 - D is polar;
so by Theorem 5(a) almost no Brownian path from a point of D hits the
boundary. o

The Distribution of wc(SS) and Harmonic Measure (N >- 2)

Let D be a Greenian subset of 08'v, let be a point of D, let be a Brownian


motion from and let Sc be the hitting time of 8D (Euclidean boundary).
When N = 2, this optional time is almost surely finite according to Theorem
10. When N > 2, xc(t) = ao almost surely, and it is convenient to
11. Continuity of the Composition of a Function with Brownian Motion 649

define wi(+oo) = oo. It will be shown in Section 13 that for N z 2 the


distribution of wr(SS) is the harmonic measure

11. Continuity of the Composition of a Function with


Brownian Motion
In this section all Brownian motions have the same variance parameter.

Theorem. Let u be a Bore! measurable extended real-valued function on an


open subset D of 6YN, and suppose that whenever is a Brownian motion
from in D with SS the hitting time of aD, the function is for
almost every to right continuous on the interval [0, Se(w)[. Then whenever
is a Brownian motion in IBN, the function w)] is for almost every
co continuous on the intervals on which it is defined.

According to Section VI.6, if A is a probability distribution on D and if


is a Brownian motion with initial distribution .t, then almost every
process sample function is almost surely right continuous up to the
hitting time of OD because this is true by hypothesis when A is supported
by a singleton. Slightly more generally, if A is a measure on R' and if AD
is the projection of A. on D, then if 2D(D) > 0, we can consider a Brownian
motion w with initial distribution AD (perhaps not a probability distri-
bution but the generalization to allow this case is trivial) and deduce that
almost every sample function of this process is right continuous up to the
hitting time of aD. Now if r > 0, the process wA(r + ) is a Brownian motion
with initial distribution the distribution of w,t(r). It follows from what we
have just shown that almost every u[wA(r + )] process sample function for
which wa(r)e D, and considered only up to the hitting time by u[w,t(r + )]
of aD, is right continuous. The exceptional PA null set depends on r, but the
union of these null sets as r ranges through the positive rational numbers is a
PA null sets, and if co is not in this null set, the sample function t r--. u[w,t(t, w)],
defined on the set of points t with wA(t, w) in D, is right continuous. This
argument needs no change if A is not a probability measure and in particular
is valid if 2 = IN (see Section VII.2), a choice making the Brownian motion
stationary. With this choice suppose that b > 0, and consider the process
w',k(-) defined by

wx(b - t)
wA(0) + w2(t) - wA(b) if t > b.

This process is a Brownian motion; so almost every sample function


is right continuous where defined. In particular, almost every u[w,t(b - )]
sample function is right continuous, where defined, on the parameter interval
650 2.IX. Brownian Motion and Martingale Theory

[0, h [. Hence (under the present hypothesis that n. = IN) for every h > 0
almost every process sample function is continuous, where defined,
on the parameter interval [O,b[ and therefore on R. Referring back to
Section VI.6 again, it follows that for IN almost every in R' this sample
function property holds for when I is supported by and therefore
holds for when ), is absolutely continuous relative to 1N. If now
is a Brownian motion in RN with an arbitrary initial distribution and
if r > 0, the process w(r + ) is a Brownian motion with an initial distribution
absolutely continuous relative to IN. The theorem follows for parameter
values >_ r and therefore as stated.

12. Continuity of Superharmonic Functions on Brownian


Motion
Theorem. if'u is a superharmonic function on an open subset D of R v and i/
is a Brownian motion on RN. then almost ever' sample function
(where defined) is finite valued and continuous except for a possible infinity
at the parameter value 0.

It is sufficient to prove the theorem for D a bounded open relatively


compact subset of an open set D,, with u superharmonic on a neighborhood
of D. Adding a constant to u if necessary, it can be assumed that u is positive
on D. According to Theorem 7, the hypotheses of Theorem I 1 are satisfied
by u. Hence the sample functions are almost surely continuous
where defined. These sample functions are almost surely finite valued except
possibly at the parameter value 0 because (Theorem 5) almost no Brownian
path hits the polar set of infinities of u at a strictly positive time.

Parabolic Context
According to Section 7, if s is the parabolic context version of the process
defined in (7.1), the process is an almost surely right continuous
supermartingale except that the parameter value 0 is to be omitted if
x.. It follows that except possibly at the parameter value 0, the x )
sample functions are almost surely finite valued with finite left limits. It is
left to the reader to check that more generally if a is any superparabolic
function on an open subset of A" and if is a space-time Brownian
motion on 68N, then except possibly at the parameter value 0, almost every
sample function (where defined) is finite valued and right continuous
with finite left limits. As the following example shows these sample functions
need not be continuous.

EXAMPLE. Let f' be a monotone increasing left continuous function on H.


and define a on A" by s) =1(s). The function ri is superparabolic and
13. Preliminary Probabilistic Solution of the Classical Dirichlet Problem 651

if is a Brownian motion on il' the sample functions will have


jumps if f has discontinuities at values of the argument strictly less than
ord i(0).

13. Preliminary Probabilistic Solution of the Classical


Dirichlet Problem
The following theorem is a special case of the much deeper Theorem 3.11.2
but will be needed before it is possible to prove the latter. Recall that the
Euclidean boundary of a Greenian subset of R' is PWB resolutive. Although
we have not and shall not define a Brownian motion w(-) at the parameter
value + oo, we shall occasionally have formulas containing w(S), where S is
a possibly infinite valued random variable; if so, w(+ co) is to be interpreted
as oo.

Theorem. Let D be a Greenian subset of [R", let be a Brownian


motion from in D, and let SS be the hitting time of the Euclidean boundary
8D by
(a) The distribution of ww(SS) on dD is the harmonic measure PD(', ).
(b) Let f he a PWB resolutive boundary function. Then

E{f[wt(S{)]), (13.1)

1m Hf[wc(s)] = f[we(S;)] a.s., (13.2)


StS4

and if xt(t) is defined by

Hf[w{(t)] if 0:5 t < SC,


xr(t) = f[wj(SS)]
if S4:5 t 5 + ac,

then the process {x,(t), 3r{(t), 0 5 t 5 + oo) is an almost surely con-


tinuous uniformly integrable martingale.

Observation. The evaluation (13.1) is the martingale equality for the


process at times 0, + oo, equivalently, at times 0, S. The representation
Hf = Hf as the difference between two positive harmonic
functions makes clear by way of Theorem 7 why the limit in (13.2) exists. We
shall not use this argument, however. Equation (13.2) is an elegant justifica-
tion of the PWB method of solving the Dirichlet problem, in that the solution
Hf has the desired boundary limit function on approach along almost every
Brownian path. Equation (13.2) will be extended to an arbitrary PWB"
resolutive boundary function, defined on a not necessarily PWB'' resolutive
boundary, in Theorem 3.11.2.
652 2.1X. Brownian Motion and Martingale Theory

Proof It is sufficient to prove that (13.1) is true when f is a bounded


Borel measurable function, say If I <_ c. Let v be a function on D in the upper
PWB class for f: Then v + e is a positive superharmonic function on D and
has inferior limit at least f(q) + c at each boundary point rl. Hence Theorem
7(d) with u = v + c yields the inequality (v + E{ f[ws(SS)] + c};
that is, E { f[w4(SS)]}. By definition of Hf we conclude that z
E.{,f[w,( SS)] }, and (13.1) follows on applying this inequality to both f and
-1.
Proof of(b). Equation (13.1) for PWB resolutive f follows from (a) since
HJ = f ). To prove the rest of (b) we first prove that for each t >_ 0

x4(t) = E{ f[wt(SS)]I55(t)} a.s. (13.3)

This equation is trivial on the set IS, 5 11. On the complementary set
S > t; the Markov property of Brownian motion implies that the right-
hand side of (13.3) is almost surely E{ f [ws(S4)I w4(t) }. Since the process
is a Brownian motion with initial distribution the distribution of
this conditional expectation is the expectation of f at the first hit of DD
by a Brownian motion from v4(t). In view of (a) this expectation is almost
surely H f[ws(t)]. Thus (13.3) is true and shows that is an almost
surely right continuous martingale, uniformly integrable because it is right
closed, a family of conditional expectations of a given random variable.
Since this martingale is almost surely right continuous, it must almost surely
have left limits at all points, so the limit in (13.2) exists. (Alternatively apply
Theorem 7 as in the above observation to obtain the existence of this limit).
To identify the limit as .f [w4(S4)], we can apply Theorem VII.4 which states
that an almost surely right continuous martingale relative to a suitably
defined filtration of a Brownian motion is almost surely continuous, but we
can also give the following more elementary direct proof. Observe that in
the notation of the proof of Theorem 7 it is clear that ws(Sk4) is Y" # (S.4)
measurable for all k: so lim, _ wt(Sks) = ws(S4), and therefore also f [w4(S4)]
are Y; measurable. It now follows from this measurability and the
conditional expectation continuity theorem that

lim E{ f[ww(S4)]I Yo . =f[wc(S4)] a.s.,


(13.4)

that is. f[ww(SS)] almost surely, and (b) is now com-


pletely proved.

Parabolic Context
Theorem 13 translates directly into the parabolic context with no change in
proof.
14. Probabilistic Evaluation of Reductions 653

EXAMPLE. Let A be an analytic subset of I V, let S, be the hitting time of A by


a Brownian motion from , and define i(5, s) = P{S, < s) for s > 0. Then
s) is the probability that a space-time Brownian motion in b = R' x
]0, +oo[ from hits A = A x ]0, +oo[ during the process lifetime at a
strictly positive time. It follows (Section 9) that ti is superparabolic on b,
parabolic on D - A. In particular, if A is closed define Do = 68" - A,
Do = Do x ] 0, + oo [. The parabolic version of Theorem 13 applied to the
restriction ti of u to Do shows that tto = Do x ] 0, + x[).

14. Probabilistic Evaluation of Reductions


Let D be a Greenian subset of R", let be a point of D, let be a Brownian
motion process from , and denote by T'A [TA] the entry [hitting] time of a
set A by

Theorem. If A is an analytic subset of D and if v is a positive superharmonic


function on D, then (reductions relative to D)

E{v[w,(T't)]; T'. < T-°} (14.1)

E{v[w,(T )]; TA < (14.1sm)

that is,

P{w,(TA)eB.T4 (14.2)

for every Bore! set B. In particular,

R1()=P{T'4 (14.3)

R () = P{T4 < T4°}. (14.3sm)

Observation (a). Equations (14.1) and (14.1sm) make obvious many of


the reduction properties which are not at all obvious from their potential
theoretic context, for example, the countable additivity of the function
vi-.R" [Section 1.VI.3(f)].
Observation (b). The proof of (14.1) and (14.Ism) for compact A given
below involves neither analytic set theory nor capacity theory, and since
these equations are true for A an F. set if true for A compact, it follows that
the proof of these equations for A an F. set involves neither analytic set
theory nor capacity theory.
Observation (c). The right side of (14.3sm) defines a superharmonic func-
tion of according to Section 9. More generally the function defined by the
right side of (14.1sm) is excessive for Brownian motion (Section VI.12) and
654 2.IX. Brownian Motion and Marungale Theory

is therefore superharmonic or identically + co on each open connected com-


ponent of D. Thus Theorem 14 can be considered as the identification of
certain probabilistically defined superharmonic functions as reductions.
It is sufficient to prove (14.1) and (14.1sm) for bounded v (and the
boundedness assumption is made tacitly throughout the following proof)
because v is the limit of an increasing sequence n '-+ v A n of bounded positive
superharmonic functions.
Proof of (14.1) for A compact. If e D - A, then RA TT" _
T", and if A is compact, we have seen (Section 1. VIII.10) that RA on D - A
is the PWB solution on D - A for the boundary function v on A n a(D - A)
and 0 elsewhere. According to Theorem 13 this PWB solution is precisely
the right side of (14.1) on D - A. On A the two sides of (14.1) are both
Proof of (14.1) and (14.1 sm) for analytic A. Fix , v and write ¢(A) for
the right side of (14.1) when A is analytic. The function 0 is an increasing
set function because if is the process v redefined as 0 for parameter
values > T, °, then is an almost surely right continuous supermartingale
and O(A) = E{x(T{")}; so the fact that O (B) >_ 4(A) when A c B is the
supermartingale inequality for at the pair of times T{', Tc". According
to Section 1.VI.5, if 4o(A) is defined as the left side of (14.1), 0o is a Choquet
capacity on D, that generated by a topological precapacity, the restriction
of 0o to the class F of compact subsets of D. Since 0 is an increasing set
function, equal to 0. on r, and since A. T A implies that O (A.) j 4(A), it
follows first that when A is open, ¢o(A) = O(A), second that when A is
analytic,
¢o (A) = sup { Oo(B) : B c A, B compact)
=sup {¢(B): B c A, B compact} < O(A),

and finally when A is analytic,

¢o(A) = inf {40(B): B A, B open} = inf {¢(B): B c A, B open} >t 4(A).

Hence 0 = 0o for A analytic, as asserted in (14.1). Equation (14.1sm) is


true because (14.1) with A replaced by A - becomes (14.1sm) when
r-p0.
If (14.1 sm) is applied to v = G°(p, ), the right side of (14.1 sm) becomes
G°i.(ry), where ). is the distribution of w,(T,) for T{ < T4°, so that (14.Ism)
becomes the equation defining the measure Sp (ry, ). Hence (14.2) is true.

Classical Reductions Composed with Brownian Motion

Let D be a Greenian subset of IBN, and let be a Brownian


motion in D from , with lifetime S. It is supposed that JF4(0) contains
the null sets. Let v be a positive superharmonic function on D. Then
14. Probabilistic Evaluation of Reductions 655

lim u[ww(s)]
SfSS

exists almost surely, and we define v [wf(1)] for SS < t < + oo as this limit,
so that (Section 7) fir( )} is an almost surely continuous super-
martingale. Let A be a Borel subset of D, and define ,4 = { (t, a)): wt(t, w) a A}.
The set A is nearly predictable because is a nearly predictable process.
In the classical context R' is a positive superharmonic function on D, and
we define

RA[w,(t)] = lim RA [w,(s)]

for S4:5 t < + oo. The supermartingale reduction 1 is an almost


+vlw

surely continuous supermartingale, and Theorem 14 combines with Theo-


rem IV.20 to show that this supermartingale is indistinguishable from

Generalization. Let A and B be analytic subsets of D, and let T{ .B be


the hitting time by of B after hitting A but before hitting 8D,

TT B=inf{t: TT <t <T4°,w,(t)eB}.

Then using the notation Q for smoothed reductions,

E{c[wr(TA.B)]; T,-' < T4°}, (14.4)


0

and in particular,

Q I DBlA() = P{TT .B < T4a°}. (14.5)

To prove (14.4), observe that if is the process killed at T{°, then


(strong Markov property) z(TT + ) is a Brownian motion process killed
at To' - TT with initial distribution that of z(T{ ), of total value 5 1. The
hitting time of B by this process is T{ B - TA. Thus the right side of (14.4) is

E{E{v[ww(T, .B)]:T, .B < T4°Iw4(7 )}} = E{QcQB[z(TA)]} = Q


(14.6)

Parabolic Context. Theorem 14 and the generalization translate directly


into the parabolic context, together with their proofs. In particular, in the
obvious notation 5 , ) = P(,v (t,4) e h; T? < 7'°}.
656 2 IX Brownian Motion and Martingale Theory

15. Probabilistic Description of the Fine Topology


Let be a point of l", let be a Brownian motion from . and for any
subset A of R' denote by T" the hitting time of A by

Theorem. For to he a fine limit point of A it is necessary that

P{ TB = 01 = 1 (15.1)

whenever B is an anal vtic .superset of A, and it is .sufficient that (15. I) be true


B is an open superset of'A.

According to this theorem, if A is analytic, the point is a fine limit


point of A if and only if (15.1) is true when B = A ; a Borel set A is a fine
neighborhood of l; if and only if almost every path lies in A for some
parameter interval [0. t[ with t = t(vi) > 0. The value t can be taken as the
hitting time of RN - A.
In proving Theorem 15 we can assume that A is a subset of a Greenian
set D. Since (Theorem 1.X1.3) is a fine limit point of an analytic set A
if and only if J I the theorem for A analytic follows from (14.2)
with B = It 1. Since (Section 1.X1.1) l; is a fine limit point of a set A if and
only if is a fine limit point of every open superset of A the theorem follows.

Application to Fine Limits and Cluster Values

Let u he a Borel measurable function from an open subset D of R into a


complete separable metric space D', let be a point of D, and let K,,(-) be a
Brownian motion in Fl' from . An application of Theorem 15 to the analysis
of limit values of u along Brownian paths in Section VII.6(d)-(f) yields the
following results. If D' = U8, then

lim sup u[ii (t)] = f lim sup u(q) a.s.


1-4) q-4
along with the corresponding relation for the inferior limit. For general D'
a point q' of D' is a fine cluster value of a at if and only if q' is a cluster
value of u along almost every path back to , and u has fine limit q'
at c if and only if lim,_, u[tv,(t)] = q' almost surely. Recall that a point
if is a cluster value of u along either almost no or almost every path
back to and that P{lim,.0u[ii,4(t)] exists} is either 0 or 1, and in the latter
case this limit is almost surely constant, that is, a single point q'.

Parabolic-Fine Topology

In the context of the parabolic-fine topology the counterpart of Theorem


15 with Brownian motion replaced by space-time Brownian motion is true.
15. Probabilistic Description of the Fine Topology 657

and the proof is an immediate translation of the proof of Theorem 15 into


the parabolic context. The application of Theorem 15 to fine limits and
cluster values goes over into the parabolic context with no change.

Application to the Support of a Swept Measure (Notation of Theorem 14)

According to Theorem 14, the distribution of ww(T,) for Tt < T4° is


SD Observe that on the one hand for TA < T{° the process wc(TT + )
is a Brownian motion with initial distribution that of ww(T,) and on the
other hand for T; < 7,' and ww(TT) not in A the set of strictly positive
values of t with ww(TA + 1) in A must have limit point 0. Hence according
to Theorem 15 and the strong Markov property of Brownian motion, the
distribution of w{(Tt) for Ti < T, °, that is, the measure by is supported
by the trace on D of the fine closure A v Af of A. This reasoning in the
parabolic context yields the corresponding result for the support of a swept
measure in that context. These results have already been derived non-
probabilistically in Sections I.XI.14 and 1.XVIII.13. In the classical context
but not in the parabolic context we can go further. In fact (classical context)
almost no path hits the polar set A - Af; so SD must be supported
by Af as proved nonprobabilistically in Section 1.X1.14. A simple time
reversal argument shows that ww(T,) is almost never a fine interior point
of A if e D - Af; so for such a point the distribution Sp is supported
by D r BfA1, as was proved nonprobabilistically in Section 1.XI.18.

The Hitting of a Parabolic-Semipolar Set by a Space-Time Brownian


Motion

Let be a space-time Brownian motion in AN, and let A be a parabolic-


semipolar subset of d8". We now show that for almost every tw(-) path the set
of parameter values t with w(t) e,4 is at most countable. In view of the struc-
ture of a parabolic-semipolar set it is sufficient to prove the assertion for A a
Borel set parabolic-thin at every point of 9'. Moreover it is sufficient to
consider only hits of A during the parameter interval ]0, c[ for arbitrary
c > 0. Let T, be the hitting time of .4 by If is a space-time Brownian
motion from a point , then P{ T, = 0} is either 0 or I and must be 0 because
A is parabolic-thin at . It follows that P{T, = 01 = 0 with no restriction
on the distribution of w(0). Define T, = T; n c. The process w(T, + ) is
a space-time Brownian motion; so if Tz is the hitting time of A by this
process, we conclude that P{Tz = 0) = 0. Hence w(T, + TT)eA almost
everywhere where T, + TT < + oo. Define T2 = (T + TT) A c. Continuing
in this way we find an increasing sequence T of optional times such that
almost surely T. 5 c; T. < c implies that and T. < for
t < To = lim,,.,m T. the inclusion w(t) e A is true if and only if t = T, or
T2, .... Here fi is the first transfinite ordinal. We can continue by transfinite
658 2.IX Brownian Motion and Martingale Theory

induction, defining T,+, < T,,+2 < . Observe that E{T} = E{T,+, } if
and only if T, = c almost surely; so there is a countable ordinal ), such that
T. = c almost surely, and it follows that almost every path meets A
at most countably often in the parameter interval ]0, c[, as was to be proved.
Since (Section I.XVIII.12) a parabolic-semipolar set is also coparabolic
semipolar, a parabolic-semipolar set is a countable union of Bore! sets
which are both parabolic thin and coparabolic thin at every point of AN,
so we could have supposed in the preceding proof that A has this property.
Under this hypothesis on A it follows easily that P{ U , { T. = c } } = I in
the preceding proof: so transfinite induction could have been avoided.

The Iterated Logarithm Law

Let be a Brownian motion in l from the origin. Then

IK'(t)I
limosup (2a 2tlogllogtl)1!2 I a.s. (15.2)
r- =

In fact a slightly stronger assertion is true: when c > 1.

Itr(')I2 < 2ca 21 log I log 1 (15.3)

almost surely for all sufficiently small strictly positive values of t, depending
on the Brownian path. but when c = 1, there are almost surely arbitrarily
small strictly positive values oft for which there is equality in (15.3). This
result is a restatement in probability language of the significance of Section
1.XVIII.6. Examples (d) and (e). To see this, observe that on the one hand
according to Example (d) when c = 1, the open set

!)_ <2ca2IsllogllogIsll,-I <s<0

has the origin as a parabolic-regular boundary point, and therefore (Theorem


I.XVIII.7) D is not a deleted parabolic-fine neighborhood of the origin.
Hence almost every space-time Brownian path from the origin hits 9" - D
at arbitrarily small strictly positive parameter values; so the limit superior
in (15.2) is almost surely at least 1. On the other hand when c > 1, according
to Example (e). the set b has the origin as a parabolic-irregular boundary
point, and therefore (Theorem 1.XVIII.7) D is a deleted parabolic-fine
neighborhood of the origin; so almost every space-time Brownian path
from the origin lies in b for all sufficiently small strictly positive parameter
values, and thus (15.3) is true almost surely for all sufficiently small strictly
positive values of t, depending on the path. Hence the limit superior in
(15.2) is almost surely at most c and so must be almost surely I.
16. Composition of a-Excessive Functions with Brownian Motions 659

16. a-Excessive Functions for Brownian Motion and Their


Composition with Brownian Motions
Let D be a connected Greenian subset of 68N. Throughout this section a
is a fixed positive number, and if u is a function from D into l , the notation
u refers to the function on b = D x 01 defined by r (n, t) = e"u(q). Define
an operator Aa with domain the 0" class of functions on D by

Aau= Z Au - au. (16.1)

The condition that a universally measurable function u from D into 08+


be a-excessive for Brownian motion in D is

e-°`bD(1, , u) 5 u

with equality in the limit when i - 0, and it follows that u is a-excessive


if and only if ti is excessive for space-time Brownian motion in D. Hence
(parabolic context version of Theorem 8) ri is superparabolic if finite on
a dense set, and we can derive the following assertions (al) and (a2).
(al) If u is a-excessive, then u is lower semicontinuous and either
u = + oe or u < + oo quasi everywhere on D.
The finiteness assertion follows from the fact (Section I.XVIII.1l) that
a subset A of D is polar if the set A x R is parabolic polar.
(a2) If u is a positive function from D into R of class 01i, then u is
a-excessive if and only if A°u < 0, and if u is a-excessive, the Riesz measure
associated with u (Theorem I.XVII.6) has 1N+t density

1) = -e'A°u(?I)
at (PI, t).
If u is an arbitrary not identically +oo a-excessive function, the decom-
position iu = UUp + cimgb + u,,,s of the positive superparabolic function ti into
its potential, quasi-bounded parabolic, and singular parabolic components
is invariant under a translation (q, 1) r (q, t + c). Hence, if ti is any one of
the above three components,

6(n, t) = e-°`ti(,, t + c) = eaY(q, 0),

and we write tip(p, 0) = up(n), umgb(q, 0) = Umgb(q), 0) = u,,,s(q). Then


0 and A°um, = 0. Moreover the function (q, t) r-* e'u p(ry) is a super-
parabolic potential on b, and the above translation argument shows that
the Riesz measure associated with this potential is a product measure of
the form (dt) with u a measure on D. If u is of class V2), we have
already evaluated p: A(dq) = -A'u(q)1N(dn).
660 2.IX. Brownian Motion and Martingale Theory

Now let be a point of D, let {w be a Brownian motion from


and let SS be the hitting time by of the Euclidean boundary of D.
The definition of SS in Section 7 shows that actually no specific boundary
is involved here. In the following u is an arbitrary not identically +00 a-
excessive function on D. According to the parabolic version of Theorem 7
applied to u, almost every sample function, for parameter values
< SS, is right continuous and is finite except when +oc and the
parameter value is 0. Apply Theorem II to find that "right continuous"
here can be replaced by "continuous" and that for an arbitrary Brownian
motion in IBN almost every u[w0] sample function is continuous when
the path is in D and is finite valued except when w(0)eD, u[w(0)] =
+ oo, and the parameter value is 0. Theorem 7 applied to ti yields the results
(a3)-(a6).

(0) urn e-"u[ww(t)]


ttS4
exists (< + oo) almost surely.

Define ,v,(+ oc) = Y E os+ S(s),

vS(t) = xS(t) = e'u[ t' (t)] if t < SS;


x{(t) = 0, xS(t) = lim e',u[ww(s)] if SS 5 t < +00.
5?SS

(a4) The processes {xS( ), S( )} and {x{( ), are almost surely


continuous supermartingales except that the parameter value 0 is to be
omitted if oo and that will have a jump at t = SS when SS <
+oo unless 0.
(a5) If u = up or u = u,,,3, the limit in (0) vanishes almost surely. If
u = umgb, the process is a uniformly integrable martingale.
(a6) u() ? un,gb() = E im e-"u[wS(t)]e
I
EXAMPLE. If u is the sum of a convergent series of positive bounded C'2)
solutions on D of the equation A 'u = 0, we now show that u = lt,,,gb; so

W.)=El 1ftS4m e-"u[ww(t)] }. (16.2)

It is sufficient to prove this assertion when u is bounded. (Observe that


boundedness of u does not imply boundedness of ti.) Define

s) = E{n A lim [e21`1u[ww(t)]]


IfS4 I.
The parabolic context version of Theorem 7 applied to the positive super-
parabolic function u A n on b implies that ti is the quasi-bounded parabolic
17. Brownian Motion Transition Functions as Green Functions 661

component of u n n. Since u is an increasing sequence of bounded parabolic


functions with limit ti, the function ti is quasi bounded.
It is left to the reader to show that if f is a positive Borel measurable
function on the Euclidean boundary of D and if we define u on D by

E{e- 4f [w4(SS)] r

with the convention that a-' = 0, then u is either identically + 00 or a


0121 solution of the equation A 'u = 0 and to show that if f is bounded and
is continuous at a regular Euclidean boundary point S of D and if SS is
almost surely finite, then u has limit f(C) at . This result, which can be
considerably strengthened, illustrates the possibility of analyzing the
Dirichlet problem for solutions of the equation 0°u = 0 by means of the
corresponding analysis of the Dirichlet problem for parabolic functions.
The reader is also invited to generalize the whole discussion of this section
by studying the linear space of differences u, - u2 of functions on D for
which u, and tie are positive and superparabolic on D. The above restriction
to positive functions will thereby be dropped.

a-Excessive Functions for Space-Time Brownian Motion

If u is an arbitrary a-excessive function for space-time Brownian motion


in an open subset b of AN, that is, if the function t : iI = (y, t) f-+ a `u(q) is
excessive for space-time Brownian motion on D, the properties of excessive
functions in the parabolic context are immediately applicable to ti,.

17. Brownian Motion Transition Functions as Green


Functions; The Corresponding Backward and Forward
Parabolic Equations
Let D be an open subset of A'. Throughout this section s), ry = (rl, t),
and these points are in D. Recall that o as defined in Section VI1.10 is a
density determining the transition function of space-time Brownian motion.
In particular (Section VII.9), when D = D x R with D a Greenian subset
of I18v, space-time Brownian motion on D is governed by a transition density
f"t, with fo(t, 5,,) = f,;((c, s), (11, s - t)). Furthermore for this choice of b
the notation GD was also introduced in Section 1.XVII.18 but was defined
there to satisfy (b) of the following theorem. According to (b), these two
interpretations of t` are equal.

Theorem. Let b he a nonempty open subset of C8", and let tri be the transition
density of space-time Brownian motion in D.
662 2.IX. Brownian Motion and Martingale Theory

(a) n) = 0-
(b) in particular, if D = D x OB with D a Greenian subset of j N and if
I'D is the transition density of Brownian motion in D, then (,(t.
61)W- s),(q,s- t)).

Proof. (a) Assertion (a) is true if b = IFB' because then both sides of the
equality in (a) reduce to f(s - t, , n). In the general case (, is given by
VII(10.3), and according to the probabilistic evaluation of parabolic measure
in Section 13, equation VII(10.3) means that for fixed n', p(5,>7) is equal
to/'(.v - t, , n) less the parabolic Dirichlet solution on D at for the boundary
function i( - t. , n) (defined as 0 at the point ao if b is unbounded). Since
Gr, can also be expressed in terms off in this way (Section I.XVII1.1). (a)
is true.
Assertion (b) is a specialization of (a) requiring no separate proof. 0

Observation (a). The transition density /o satisfies two differential equa-


tions: i) is parabolic on b - {q }, that is, q) = 0 there; r°D( , )
is coparabolic on b - 4, that is, iq) = 0 there. These equations are
called, respectively, the backward and forward equations of the space-time
Brownian motion because they refer, respectively, to initial and later posi-
tions. If D = D x OB with D a Greenian subset of 08N the transition density
ft, also satisfies two differential equations: defines a parabolic
function of (5, t) and one of (n, t) for i; 96 n. This formulation obscures the
difference between the backward and forward equations, however. The
point is that fD(s - 1. , n) is parabolic in s) (backward equation) and
coparabolic in (n, t) (forward equation) for s) # (n, t).
Observation (h). It was proved in Section I.XVII.18 that if b = D x ff8
with D a Greenian subset of D$N, then

fD(t, S, n)lt (dt) = aNGD(S, n) (17.1)

for a,% as specified in that section. We shall now sketch how (17.1) can be
derived from the probabilistic definition of iD in Section V11.9 without
using the potential theory derivation. Observe first that for N z 3 equality
(17.1) follows from the evaluation of iD in VII(9.5). In fact (17.1) is true
if D = R', in which case iD = i, and on integration V[I(9.5) becomes

iD(t,s.n)lt(dt) = p. (17.2)

in view of the fact that has distribution µD(5, ). The bracketed quantity
was identified with n) in Section 1.VIII.3. When N = 2, first let D
he a half-plane D. We have evaluated GDa in 1.XVII(18.3) and in VII(9.8)
18. Excessive Measures for Brownian Motion 663

we have evaluated the Brownian transition density from the origin as initial
point for a half-plane not containing the origin. This evaluation is easily
adapted to give the expression 1.XVII(18.4) for the transition density of
Brownian motion in a half-plane. Equation (17.1) can now be verified for
D = D. When D c D, the evaluation

t°D(t, , q) = Cn+(t, "1) - E{tD+(t - SS, x'w(SS), q)}, (17.3)

derived as VII(9.5) was, is then integrated to prove that (17.1) is true for
D c D+, and therefore surely if D is an arbitrary nonempty open bounded
plane set. Hence, applying this result to each member of an increasing se-
quence of bounded open subsets of an arbitrary Greenian set D, with union
D. yields (17.1) in the general case when N = 2. The case N = I is treated
similarly. Observe that the above argument when N Z!: 3 is simpler than the
argument given in Section I .XVII.18 because parabolic measure is available
in the present context.

Application to Energy

In Section 1.X111.8 it was pointed out that the evaluation of GD given in


Theorem 17 can be used to prove that the energy of a charge is positive.

18. Excessive Measures for Brownian. Motion


The symmetry in classical potential theory, expressed. for example, by the
facts that the Laplace operator s is formally self-adjoint and that f.D(t, , )
and are symmetric functions, suggests that the excessive measures
for Brownian motion are in some sense the same as their duals, the excessive
functions for Brownian motion. The following theorem shows that this
identification of f.D-excessive functions with GD-excessive measures is in
fact correct in the most natural sense.

Theorem. If D is a Greenian subset of R', a (totally unrestricted) measure N


of Borel subsets of D is a 4D-excessive measure if and only if there is a tD-
excessive function u such that for every Bore! subset A of D

µ(A) = I u d1N. (18.1)


JA

Observe that this theorem means that on a connected open component


of D either the measure p is the indefinite integral of a positive superharmonic
function u, or (u = + co) p(A) is either 0 or + oo according as A is or is not
IN null.
664 2.IX. Brownian Motion and Martingale Theory

In the following proof of the theorem we assume to avoid trivialities


that D is connected. Recall that by definition a measure k of Borel subsets
of D, not necessarily finite on compact sets, is called excessive if and only if

J I oD(t, , n)!N(dn) u(A) (t > 0) (18.2)


f.D A

whenever A is a Borel subset of D, with equality in the limit when t - 0.


If p is a GD-excessive measure, this limiting equality implies that µ(A) = 0
whenever IN(A) = 0. If p is GD-excessive and if u(A) = + oc whenever
!N(A) > 0, then p is given by (18.1) with u the identically +oo excessive
function and conversely (18.1) with this excessive function yields an excessive
measure. On the other hand, if p is GD-excessive and if p(A) < + oo for some
not IN null Borel subset A of D, the finiteness of the right side of (18.2) along
with the continuity and strict positivity of tD for t > 0 implies that p is finite
valued on compact sets. Thus in this case p is absolutely continuous relative
to IN, given by (18.1) with u a Lebesgue measurable function finite IN almost
everywhere on D. Then for fixed t > 0,

n)IN(dd) 5 u(n) (IN a.e. q), (18.3)


i'D

which in view of the Chapman- Kolmogorov equation and the symmetry


of GD(t, , ) implies that

+ t, )IN(d) 5 I u()GD(s, , n)1N(d) (18.4)


ID

for all s > 0, t > 0, n e D. Hence for fixed the left side of (18.3) increases
as t decreases. Define u'(n) as the limit of this left side when t -+ 0. Then
u' S u !N almost everywhere on D, and there must be equality IN almost
everywhere because the left side of (18.2) has limit u(A) when t - 0. The
function u is uniquely defined up to an IN null set, and we now replace u
by u'. This change does not affect the left side of (18.3); that is, we now have
u' = u, and u is obviously excessive. Conversely, if u is a t -excessive function
and is not identically + oc, that is, if u is a positive superharmonic function
on D, then u is locally IN integrable, and using again the symmetry of GD(t, , ),
we find that

u()GD(t, , 0!N(d) 5 u(n) (18.5)

with equality in the (monotone) limit when t -- 0. Integrating 1A - IN(dn)


yields the fact that the measure p given by (18.1) is GD-excessive.
Is. Exmsivc Measures for Browntan Motion 665

Space-Time (Parabolic) Context. Let D be a nonempty open subset of


tfd". We use the notation introduced in Section Y11.2: if A is a subset of G'8",
define A, = {n: (n, 1) e,4) and A, = A, x {t}. Furthermore let be the a
algebra of subsets A of b for which A r),6, is a Borel set for all t. A measure
µ on the a algebra 2 determines a measure t) of Borel subsets of D,
by way of µ(A, t) = µ(A x It)). Conversely, if for each t there is a measure
t) of Borel subsets of D a measure 4 on 2 can be defined by setting

u(A) = E µ(A,, t) (18.6)

with the obvious convention for uncountable sums. In the following "mea-
sure on !P" refers to a measure defined in this way, but we make no hypotheses
on the finiteness of ti or of t) on any class of sets. A measure µ on I
is Go-excessive if and only if whenever A e b,

Dp(dd, s) J Vin(( , s), (n, t))lx(dR) 5 µ(A,, t) (s > t), (18.7)


s A,

and there is equality in the limit when t T .s. A 4-excessive measure is a


measure on !' satisfying the obvious dual condition. The counterpart of
Theorem 18 in the parabolic context can now be stated : a measure µ on 2
is a 4- [16-1 excessive measure if and only if there is a 6D- [GD-] excessive
function ti such that for A e

µ(A,, t) = I t)lx0l). (18.8)


A,

In the one direction let ti be a Go-excessive function, so that 0 < ti < + m,


u is 2 measurable, and

s), (n, t))u(n, t)lv(dn) s s) (s > t) (18.9)


f',
with equality in the limit when t T s. Define a measure µ on 2 by (18.8)
and (18.6). In view of the duality relation VII(10.6) between GD and e
the inequality (18.9) implies that for AE,

I u(dn, t) J ri((7, t), (S, v))lx(d) 5 µ(A,, s) (18.10)


D, A3

with equality in the (monotone) limit when t T s; so u is 4-excessive. Con-


versely, if 4 is a tD-excessive measure, then following the argument for the
classical case as given in the first part of this section it is not difficult to see,
but more so than in the classical case, and we leave the details to the reader,
that there is a 6,5-excessive function ti satisfying (18.8).
666 2.IX Brownian Motion and Martingale Theory

The dual assertion for G.-excessive functions and lo-excessive measures


is derived in the same way or reduced to the one just proved by a reflection
of A' in the abscissa hyperplane.

19. Nearly Borel Sets for Brownian Motion


In a general context a set A in the (topological) state space of a stochastic
transition function p is called nearly Bore! if for every initial distribution
it on the state space and an almost surely right continuous Markov process
with transition function p there are Bore] subsets A and A of the
state space for which A;, c A c A and for which almost no path hits
A',' - A', at any strictly positive time. The subtleties of this definition are
needed in more general contexts but are quite unnecessary in the Brownian
motion context according to the following theorem.

Theorem. A set A is nearly Bore! for Brownian motion in an open subset of


R' if and only if A differs from a Bore! set by a polar set.

We can suppose that we are dealing with Brownian motions in a connected


Greenian set D. If A is nearly Borel and if p is supported by the singleton
the difference A - A. is a Borel set whose hitting time by a Brownian
motion from S is almost surely + x so that by Theorem 14 the smoothed
reduction of the function I on the difference set is identically 0. Hence
(Theorem I.V.4) this difference set is polar; so A differs from A', by a polar
set. Conversely, if A differs from a Borel set by a polar set, it can be supposed
that A = Ao u A,, where A is a Borel set and A, is polar. and if so, the sets
A, and A can be chosen, respectively, as A and the union of AO with any
polar G,, superset of A, .
Near/v Borel Sets for Space-Time Brownian Motion. The parabolic con-
text counterpart of Theorem 19 is a direct translation.

20. Brownian Motion into a Set from an Irregular Boundary


Point
Let D be a Greenian subset of I8'v, and let aD be the Euclidean boundary.
Let 5 be a finite irregular boundary point of D. let be a Brownian motion
from C, and let S be the hitting time of aD by is,(-). According to the fine
topology criterion of irregularity of a boundary point (Theorem 1.XVIII.7)
and the corresponding probabilistic criterion by way of Theorem 15, the
random variable S is almost surely strictly positive. Let be the process
killed at time S. Then { tr'(t), t > 0} is a Markov process with state space
D and transition density (D. Now let v be a positive superharmonic function
20. Brownian Motion into a Set from an Irregular Boundary Point 667

on D. A trivial adaptation of Theorem 8 shows that the process {v[w'(t)],


t > 0}, if defined as 0 for t >_ S. is almost surely continuous for t < S and
is a positive supermartingale if the expectations of the process random
variables are finite. If we replace v by v A c if necessary to ensure finiteness
of these expectations and then let c tend to + oc, we find by the backward
supermartingale convergence theorem that v has a finite or infinite limit
along almost every path back to C, that is, v has a fine limit < + or,
at C. We have thereby obtained a probabilistic proof of the limit result in
Section I.XI.21; a similar argument yields the corresponding parabolic
context theorem in Section I.XVIII.17.
For fixed ry in D the function = s) i-+GD(s, , ry) on D x a8 is super-
parabolic and positive and is bounded on s > so} for each strictly
positive so ; so for t > 0 and rt E D the limit pf limS.tt, 0) f.D(t + s, . q) exists
and is finite. Denote this limit by GD(t, C, ?I). Equivalently, there is a deleted
parabolic-fine neighborhood A of (Z, 0) such that

lira GD(t + s,, q) = GD(t, C, ry). (20.1)


. 3h-(4.0)

Here A depends on it = (?I, t), but in view of the parabolic version of Lemma
1.X1.8 we can choose A so that (20.1) is true for , in a countable dense
subset of D x ]0, +oo[. The parabolic context Harnack theorem now
implies that (20.1) is true for all 7 in this product set and that the convergence
is locally uniform so the function h --+GD(t, (, q) is parabolic on this product
set. Finally we show that GD(t, C") is the distribution density of w'(1). To
see this, observe that if f is a continuous function on D with compact support,

lim
S-0 D
GD(t - s, w'(s), i) = I
D
GD(t a.s.
(20.2)

in view of the fine limit result we have just obtained. The integral on the
left is a version of E{ f [w'(t)]Jw'(s)}, and in view of the Markov property
of Brownian motion this conditional expectation defines a martingale for
0 < s < t, t fixed, and an application of the conditional expectation con-
tinuity theorem and the 0-1 law of Brownian motion shows that when s -+ 0
sequentially the almost sure limit of this conditional expectation is
E{ f [w'(t)] }. This fact combined with (20.2) shows that GD(t, C, ) is the
stated density. The corresponding result in the parabolic context is left to
the reader.
Chapter X

Conditional Brownian Motion

1. Definition
Let D be an open subset of RN, and recall that a Brownian motion in D
was defined in Section VI I.9 as an almost surely continuous Markov process
with state space D and transition density Go. The probability space on which
the process is defined may or may not be rich enough to extend the process
to be a Brownian motion in RN.
We now define a Brownian motion in D conditioned by an arbitrary
strictly positive superharmonic function h on D. If D is not Greenian, for
example, if N = 2 and 982 - D is polar, the definition yields nothing new
because then h is necessarily a constant function and conditional Brownian
motion reduces to Brownian motion. We therefore suppose below that D
is Greenian. Following Section VI.13, define D" = {h < + cc), and define

h(i7) if ( , 7) a D" x D
(1.1)
0 if ( ,7)e(D - D") x D.
Observe that the fact that h is GD-excessive implies that GD(t, , ) = 0 IN
almost everywhere on D - D" when e D". An h-Brownian motion in D
is a Markov process with state space D, transition density dD, and initial
distribution supported by D. After proving in Section 2 that there is always
a continuous process satisfying these conditions, we shall add to the definition
the condition that the process be almost surely continuous. It will be shown
that then almost no sample path leaves D". Thus a 1-Brownian motion in
D is what we have defined as a Brownian motion in D.

The Notation P" and E"

In previous chapters we have used the notation P and E generically: when-


ever a probability space was encountered, P and E referred to probabilities
and expectations, respectively, on the space, and if several probability
spaces were encountered simultaneously, this notation P and E was used
for every space. In the first sections of the present chapter this notation
1. Definition 669

could be confusing because we shall deal with h-Brownian motion and


Brownian motion simultaneously. To avoid confusion, we shall use P and
E for Brownian motion but Ph and Eh for h-Brownian motion. (If h is
identically constant, h-Brownian motion reduces to Brownian motion, and
so the superscript will be omitted.) After Section 8 of this chapter, however,
and throughout later chapters the special notation Ph and Eh will be used
only when necessary to avoid ambiguity. Ordinarily the notation for the
processes involved, for example, for h-Brownian motion, will warn
the reader that P and E on the space refer to h-Brownian motion
probabilities and expectations.

Existence and Lifetime of h-Brownian Motion

According to Markov process theory, if e Dh, the transition density ell,


determines an h-Brownian motion from . The distribution of the process
lifetime SS is determined by

=GD(t'''h)
Ph{SS > t} Ph{Se = +ao} = lim (1.2)
h() ' 1^aC NO
The function ho = lim,-,o GD(t, , h) is an invariant excessive harmonic
minorant of h [Section IX.8, Example (a)]. Thus if h is a potential, the
function ho vanishes identically, so almost every h-Brownian path has a
finite lifetime. If h is a strictly positive superharmonic function on D which
is IN integrable over D, for example, if D is bounded and h is a bounded
function, then
(21t02)-N/2

P h {St-
hmt_N/2
h
= + ao} < f 0; (1.3)
n

so under these hypotheses almost every h-Brownian path has a finite lifetime.
If h is minimal harmonic on D, then ho = ch for some constant c, and since
ho is invariant excessive for eD,

ho = GD(t, -, ho) = CCD(t, -, h) - cho,

so that either c = 1, in which case almost every h-Brownian path has an


infinite lifetime, or c = 0, in which case almost every Brownian path has a
finite lifetime.

t -Excessive Functions

To avoid trivialities, we assume in this paragraph that D is connected. We


use (Sections IX.6 and IX.8) the identification of GD-excessive functions
with the positive superharmonic functions together with the identically
670 2.X. Conditional Brownian Motion

+oe function. According to Section VI.13, if v is a positive superharmonic


function on D, the function v/h (defined as 0 on D - D'') is (D-excessive,
and conversely, if u is a (o-excessive function, then u = 0 on D - D' and
either u - + x on D" or there is a positive superharmonic function v on
D such that u = v/h on D"".

Supermartingales Defined by Composing fp-Excessive Functions with


h-Brownian Motion

Let v and h be strictly positive superharmonic functions on D, and let


tt with lifetime S{ be an h-Brownian motion from a point of D". Define
u = v/h on D", the values of u on D - D' will be irrelevant. In view of the
preceding paragraph the process {u[ww (t)], t e R }, under the convention
that a[w (t)] = 0 for t z S,, is a supermartingale if x, that is,
if x, just as in the case h =- 1. Furthermore, as in the case It = 1,
the fact that (Theorem IX.6) (D(t, , v) < +x when t > 0, even when
x, implies that then also t (t, S, u) < + oc ; so if u(c) = + oo, the process
{u[w' (t)], 0 < t < + oo If is a supermartingale. We shall see in Section 2
that these supermartingales are almost surely continuous aside from a possi-
ble jump discontinuity at the parameter value S{, under the condition that
is chosen (as we shall see is possible) to be almost surely continuous.

h-Brownian Motion When h Is Harmonic

In particular, if h is a strictly positive harmonic function on D, the h-Brownian


motions play the same role for h-harmonic and h-superharmonic functions
that Brownian motions play for harmonic and superharmonic functions.
For example, when h is harmonic, a fo-excessive function is, on each con-
nected open component of D, either identically + x or h-superharmonic.
The fact that almost every path of a Brownian motion from a point
of D [define w{(+ oo) = oc if N > 2] hits the (Euclidean) boundary implies
that almost every path of a Brownian motion in D tends to a boundary
point at the path lifetime, but the latter property does not hold for h-
Brownian motion in D for general harmonic h. We shall show, however,
in Section 4 that when h is harmonic almost every h-Brownian path from a
point of D tends to D. although the path cluster set on 8D may not be a
singleton.

Canonical Conditional Brownian Motion

Let D and h be as above. In Section VII.9 we defined a canonical Brownian


motion in D, on a space 6. A canonical h-Brownian motion is defined
similarly, but there is one essential difference: when h a 1, almost every
2. h-Brownian Motion in Terms of Brownian Motion 671

path of an h-Brownian motion in D tends to a point of the Euclidean


boundary of D at the path lifetime, but this is not true for all choices of h.
Hence we must change the definition of el in that we drop the condition
in Section VII.9 that the function oh is to have a left limit in the Euclidean
topology of R at the lifetime . (6*h). When canonical h-Brownian motion is
discussed with no restriction on h, we shall mean the present definition even
though the case h e I is not excluded. In the next section we shall show that
h-Brownian motions exist for every choice of h and are almost surely con-
tinuous up to their lifetimes. If is such an h-Brownian motion in D,
on a space f2, the existence of a canonical h-Brownian motion in D with the
same initial distribution (supported by D") follows easily by mapping. In
fact the map w F- (o) = th takes a point w of fl into a point u, of 6 (we
assume that the processes have a common trap point adjoined to D), and
the iv(-) measure on 92 thereby induces a measure on 6 under which the
coordinate function process is the desired canonical h-Brownian motion.

Parabolic Context

The translation of the discussion in this section into the parabolic context
is left to the reader. Observe, however, that in the parabolic context an
example in Section IX.12 shows that the composition of a superparabolic
function with space-time Brownian motion, almost surely right continuous
with left limits, may have a dense discontinuity set which is the same for
every sample function.

2. h-Brownian Motion in Terms of Brownian Motion


In this section we derive relations between h-Brownian motion and Brownian
motion in order to reduce h-Brownian motion properties to corresponding
Brownian motion properties. Throughout the section D is a Greenian subset
of 68", h is a strictly positive superharmonic function on D, and is a point
of D". The process is a Brownian motion in D from with
lifetime S defined on a filtered probability space (S2, P), and the
process is an h-Brownian motion in D from with lifetime
S" defined on a filtered probability space (S2", f" Ph) Observe that

{w(t)eD} = {S> t), {w"(t)eD} = {S"> t}.

A trap state is supposed adjoined to D to make the transition functions


stochastic, and h is defined as 0 at the trap state.
We wish to prove the following, in various contexts. Let T be an fl-)
optional time on S2, and suppose that A e f(T). Let T" be an optional
time on (2", and suppose that A"e.F"(T"). We wish to show that if T and
672 2.X. Conditional Brownian Motion

A are defined in the same way in terms of sample functions as T" and
A" are defined in terms of sample functions, then

P"{A"n{S">T"}}= h[w(T)]
LIS>T)
Although we shall only need very special cases of (2. 1), its intuitive meaning
clarifies the subject so we prove (2.1) under very general hypotheses.
Case (a) of (2.1): T - t for some strictly positive constant t. Define a
measure Q` on the measurable space (el,.9 (t)) by

Q`(A)= h[w(t)]h (2.2)


J. 1s>11

to obtain a measure space on which the process {w(s),s < tis almost
surely continuous. Suppose that 0 < i t < < t" = i and that A E $(D").
Define

Mu = { [w(t1), .... w(tn)] eA}. M,", [w"(t, ), .... ti'"(t")] E A }.


(2.3)

and observe that M,, c IS > t{ and Mo a {S" > 1j. In view of the fact
that D - D" is polar and therefore IN null we can evaluate Q`(M0) in terms
of

Q'(M.) = I' ...f rO(t.S>SO) ... D(trt - to-l, n I, nNN(dJ) ...


A
(2.4)

Thus

Q`(M) = P"{M"} (2.5)

for M = M and M" = M"; so (2.1) is true when T - t, T" - t, A = M0,


and A" = M. It follows that (2.1) is true when T -= t and T" m t if A [A"] is
in { w(.s), .s < t}> ,) [,F { w"(s), s < '}1s">] and if A and A" are defined
in the same way in terms of process sample functions. More precisely, let
6 be the space of functions u) from [0, t] into D, and define .z(s, (h) = w(s).
If Me. {r"(.s)..s < t}, define M as the inverse image of M under the map
cur-. (0)110., from {S > t} into 6, and define M" as the inverse image
of M under the map from {S" > t} into 6. Then (2.1)
is true with T = T" - t, A = M, and A" = M". In fact this assertion for the
paired sets is true for M in the algebra of subsets of i of the form {
... , tV(tn)] e A } with t. and A as in (2.3) and therefore for Mo in the a algebra
. t} generated by this algebra. We leave it to the reader to verify
2. h-Brownian Motion in Terms of Brownian Motion 673

the fact, which we shall not need, that for the class of pairs (M, M")
obtained in this way the class of sets M is the class of subsets of {S > t} in
the a algebra .F{w(s),s S t} and the class of sets M" is the class of subsets
of {S" > t) in the a algebra .{w"(s),s 5 t}.
Almost Sure Continuity of h-Brownian Motion. Equation (2.5) states
that the almost surely continuous process { w(s), s 5 t} under Qt has the
same finite-dimensional distributions as the process {w"(s), s 5 It
follows that the latter process has an almost surely continuous standard
modification and that therefore every h-Brownian motion in D from a
point of D", and more generally every h-Brownian motion in D with initial
distribution supported by D", has an almost surely continuous standard
modification up to the process lifetime. In more detail, what we have shown
implies that when t > 0 the restriction to the rationals in [0, t] of almost
every sample function is uniformly continuous and that for fixed s

P"{ lim w"(r) # w"(s), s < S"} = 0 (r rational).


r_5

Hence, if w'(s) is defined on C h for s < S" as the limit at s of w"(-) along the
rationals, w'(-) is the required almost surely continuous standard modifi-
cation of w"(-) and is itself an h-Brownian motion from . We refine our
preliminary definition of h-Brownian motion accordingly: from now on
an h-Brownian motion in D is any almost surely continuous up to (but not
including) the process lifetime process in D with transition density GD and
with initial distribution supported by D". In particular we assume from
now on in this section that w"(-) is almost surely continuous up to but not
including the process lifetime.

Killed h-Brownian Motion

Let B be an F, subset of D, and let T [T"] be the hitting time of B by


Then if (M, M") is a pair linked as described above, the pair
(M n {T > t}, M" n {T" > t}) is a linked pair in the same sense because of the
simple expression for the probability of hitting an F, set (cf. Section VI.6).
Hence Case (a) of (2.1) yields

P"{M" n {S" A T" > t} h[xt(t) ] (2.6)

In particular, if Do is an open proper subset of D containing , then (2.6)


with B = D - Do shows that w"(-) killed at T" has the same finite-dimensional
distributions as hIDo Brownian motion from , up to process lifetimes, that
is w"(-) killed at T " and then provided with a trap state is an hi00 Brownian
motion in Do from .
Observe that since the almost surely continuous process {w(s),s 5 t}us}
674 2.X. Conditional Brownian Motion

under Q` and the almost surely continuous process {w"(s),s S t}ilsh,t) have
the same finite-dimensional distributions, it follows that these processes
have the same probabilities of hitting an analytic subset B of D during a
specified parameter interval (Section I.12), and more generally the same
argument shows that the same assertion is true if the processes are restricted
to sets M and M", respectively, linked as described above. That is, if T [T"]
is the hitting time of B by then (2.6) remains true.
Representation of an Arbitrary h-Brownian Motion in terms of a Brownian
Motion. For t > 0 define a measure Q` on 3F'(1) by

h()
Q'"(M") = (2.7)
JMh(Sb>rlh[w(t)]

and note that QV'(M") = 0 if and only if P" { M" n {S" > t)) = 0. The process
{w"(s),s 5 1}1(s">,) under Q` has the same finite-dimensional distributions
as the restricted Brownian motion process jw(s),s S t}ils>ti. We can now
complete a full circle in the discussion : the integral
h["'"(t)]dQ'h

>

JMhSh>r)

whose value is P"{ M" n {S" > t} } when expresses wh(-) process
probabilities to time t under the side condition S" > tin terms of a Brownian
motion in D under the corresponding side condition. The important point
is that this Brownian motion and the h-Brownian motion involve the same
probability space and the same random variables; only the measures are
different. In particular, if u is a function on D, then for parameter
values < t is not only a function of under P" but also under Q",
and we can thereby deduce properties of u on h-Brownian motion from the
known properties of u on Brownian motion.
The following assertions are immediate consequences of the representa-
tion of h-Brownian motion in terms of Brownian motion that we have
obtained.
(a) If A is a polar subset of D, almost no path of an h-Brownian motion
in D hits A at a strictly positive parameter value. In particular, almost no
such path hits D - D".
(b) A superharmonic function r on D composed with h-Brownian
motion in D yields a process almost surely continuous up to the h-Brownian
motion lifetime with finite-valued sample functions except possibly at the
parameter value 0.
We have already pointed out in Section 1 that if v is a positive super-
harmonic function on D and if u = v/h (= 0 where h = +oo), then
is a supermartingale if defined as 0 for parameter values z S" and with the
understanding that the parameter value 0 is to be omitted if oo.
2. h-Brownian Motion in Terms of Brownian Motion 675

We now see from (a) and (b) that this supermartingale is almost surely
continuous except for a possible discontinuity at S". For almost every
sample function the supermartingale right limit exists at S" trivially and
is 0; the left limit at S' must exist (finite) almost surely because almost
surely right continuous supermartingales almost surely have finite left limits
at all parameter values, including the parameter value + oo if the super-
martingale is positive.
(c) h-Brownian motion filtrations and the strong Markov property.
Theorems VI.8 and VI.9 are not directly applicable to an h-Brownian motion
unless h is finite valued, in which case D" = D, but the proofs of these
theorems are applicable to the h-Brownian motion context. Hence, if ."(t)
is generated by the null sets and .F{w"(s), s 5 t}, it follows that is
right continuous. Moreover, even without this special choice of filtration,
the strong Markov property VI(9.l) holds for These assertions are
of course true not only if x,"(0) _ but for an arbitrary initial distribution
supported by D".
(d) The zero-one law for h-Brownian motion. Let -Ft be the a algebra
generated by nr,o F{tv"(s, s 5 t) and the null sets. Then [assuming as
above that tv"(-) is an h-Brownian motion in D from in D"] .fit is trivial;
that is, it consists of the null sets and their complements. Two proofs of
the zero-one law for Brownian motion were given in Section VII.6; the
second one is applicable to h-Brownian motion for all h. The zero-one law
implies that if B is an arbitrary analytic subset of D and if T is the hitting
time of B by w"(-), then the probability P{T = 0) must be either 0 or 1.
The representation of h-Brownian motion in terms of Brownian motion
shows that this probability does not depend on the choice of h, and it follows
(Theorem IX.15) that this probability is if and only if is a fine limit
1

point of A. Roughly, the initial character of h-Brownian motion from a


point does not depend on the choice of h. Because of this fact, assertions
(c)-(f) of Section VII.6 are valid for h-Brownian motion for all h.

h-Brownian Motion at Path Lifetimes [Notation D, , as


Above]

If is a Brownian motion in 1V from and if So is the hitting time of the


Euclidean boundary aD by wo(-), then almost every path killed at
So tends to wo(S0) at the parameter value So. It follows that almost every
path tends to a point of OD at the path lifetime, but it does not follow
from the relation between Brownian motion and h-Brownian motion that
almost every path tends to a point of aD at the path lifetime. In fact,
when h is a potential, we shall show in Section 4 that almost every h-Brownian
path in D tends to a point of D at the path lifetime, and we shall find the
distribution of this limit point. When h is harmonic, it will be shown that
almost every h-Brownian path in D tends to aD but not necessarily to a
point of aD, at the path lifetime.
676 2.X. Conditional Brownian Motion

3. Contexts for (2.1)


As already explained in Section 2, the first problem in finding a proper
context for (2.1) is to find appropriate linked pairs (T, T') of optional
times and linked pairs (A, A") of sets. The point is that T" should depend on
in the same way T depends on and A" should be a set depending
in the same way on up to time T" as A depends on up to time T.
In Section 2 the situation was simplified by choosing T" and T to be the
same constant function. In this section T" and Twill be nontrivially optional.
Case (b) of (2.1): and are canonical, T = T", A = M. Let
I be a canonical h-Brownian motion in D from in D, on the
probability function space (&,P, p") When h = 1 , we omit the super-
script. Then and the process lifetimes S', . are identical.
Define

do(t) _{H(s).s < t} _"(t), A,(+ -0) = Y .o(t) = A011(+ 00).


r>0

The basic filtration in this case is A o'(-). Let t bean arbitrary A,'(-) optional
time and let A be an AV) set. We make the obvious choices of fi" and
R", namely t" = t and R" = A, and prove (2.1) in this context. Observe
that, with the optional time notation defined in Section 11.2, Example (b),
and using the fact that
j2_"
A . {9" > = [fi"]n} eAo(j2-"),

(2.6) yields

p"{R" > Y P"{A" {S'" > j2 = [fi"]n} }


j=o

_
f h[K'(j2_-)]d Pzz
Y_
J=O n"r,{$>j2-n=jf).) NO
(3.1)

=J
n "nIJ`>ftt,)

When n oc, the left side of (3.1) tends to p" {tt" n { 9" > th) }. The right
side is The sequence ..., h[w([t jt)],
ordered as written is a supermartingale, left closable by h[w(t)] and
(Theorem 111.17) uniformly integrable; so there is L' as well as almost
everywhere convergence to the limit h[w(t)]. Hence we can integrate to
the limit (n - oc) in (3.1) to obtain (2.1) in the present context.
Application of Case (b). If is a Brownian
[h-Brownian] motion in D from on the probability space ((),.F, P)
4. Asymptotic Character of h-Brownian Paths at Their Lifetimes 677

[(S2'' " P")] with lifetime S [S"], the problem of finding large classes of
pairs (T, T"), (A, A") for which (2.1) is true is partially solved by a reduction to
case (b). In fact, if we make the convention that all processes involved have a
common trap, the maps co i- w(-, w) = th, w" H w take 0 and
i2" into L = ?", the space of a canonical Brownian motion in D from .
In the notation of case (b), an optional time ton 1) has as inverse images
under these maps an optional time T on S2 and an optional time T" on 0"
if we suppose, as we can, that and are right continuous. A set
A in .00'(t) has as inverse images a set A in and a set A" in .F"(T")
and (2.1) is true as written if T, A, P. and A" are obtained in this way be-
cause (2.1) is true in the context of case (b). For example, if t is the hitting
time by of a closed subset of D, then T [T"] is equal almost surely to
the hitting time of this set by

Generalization of Case (b) for (2.1)

(Not used below.) It is natural to try to extend Case (b) and thereby the
application just noted to allow for the P and P" null sets. This generalization
is needed, for example, if T is to be the hitting time by of an arbitrary
analytic or even merely Borel subset of D. More specifically define .O(t)
[A"(t)] as the a algebra generated by Ao(t) and the P [P"] null sets. Ac-
cording to Section 2, the filtrations and ^-) are right continuous.
If t is an . optional time and if Ae.*(t), there are [from Section
11.2(j)] an optional time fio and an Ao(1o) optional time Ao such that,
up to a P null set, to = t and AO = A. Now according to case (b) of (2. 1),
an o (to) subset of {9 > to} is P null if and only if it is P" null. Hence,
up to a P" null set, t = t and Ao = A. Thus if we define T = 11, T" = to,
A = A, A" = AO, equation (2.1) is true in the canonical process context,
generalizing case (b). We could equally well have started with an
optional time t" and an .?"(t") set A".

4. Asymptotic Character of h-Brownian Paths at Their


Lifetimes
Theorem. Let D be a Greenian subset of R"', let h be a strictly positive super-
harmonic function on D, and let be a point of D".
(a) If h is harmonic, almost every h-Brownian path from g tends to 8D
at the path lifetime.
(b) If h = G0µ is a superharmonic potential, almost every h-Brownian
path from c has a finite lifetime and tends to a point C of D at its
lifetime. The distribution of C is given by dC) _
If A is polar, A n D") = 0.
678 2.X. Conditional Brownian Motion

Observation (1). Part (a) of the theorem does not state that when h is
harmonic, almost every h-Brownian path from tends to a point of aD,
and in fact it will be proved in Section 3.11.2 that if aD is obtained by a
metric compactification of D, then almost every h-Brownian path from
tends to a point of OD if and only if OD is h-resolutive, and in that case the
distribution of the path limit point on aD is OD We can already verify
this result in a special case, when h = I and is the Euclidean boundary,
resolutive according to Theorem I .VIII.4. In fact, if a Brownian motion in
08" from 6 in D is killed at the hitting time of the Euclidean boundary OD
to obtain a Brownian motion in D, almost every path of the Brownian
motion in D obtained (without loss of generality for the present purpose)
in this way tends to a point of OD at the path lifetime, namely to the point
at which the original Brownian path first hits aD. The distribution of this
hitting point on aD is according to Theorem IX.13.
Observation (2). We shall prove a stronger result than Theorem 4(b):
h-Brownian paths from can be obtained in this case by first choosing
the asymptotic endpoint for a path according to the probability distribution
and then choosing a 0)-Brownian path from (necessarily
tending to 5 at the path lifetime). It follows, as stated in (b), that if A is polar.
v((', A n D") = 0 because (Theorem I .V.11) h = GDµ = + co p almost every-
where onA;sop(AnD")=0.
Proof of (a). Let be an h-Brownian motion in D from 5. with lifetime
S", let Do be an open relatively compact subset of D containing , and let
T" be the hitting time of aD by The superscript h is omitted throughout
when h = 1. It is trivial that P{S > T } = 1. According to (2.1) under case
(b) as applied in Section 3, if h is harmonic,

dP E{h w(T)]}
P"; S" > T"} = h[w(T)]h) h()
;s>ri
(4.1)

I<no( h)
= I.
h()
Thus almost every path hits aDo. To prove (a), we prove that if Cis a
compact subset of D and if now T" is the first hitting time by of C
after hitting aDo, then P"{S" > T"} is arbitrarily small if D0 is sufficiently
large. With this definition of T" (2.1) yields

Csup h} P{S > T}


P"{S"> T"} <( C

Now the probability on the right is the probability that a Brownian path
in D from f hits C after hitting aD0 and this probability is arbitrarily small
4. Asymptotic Character of h-Brownian Paths at Their Lifetimes 679

for sufficiently large Do since almost every Brownian path in D from


tends to the boundary at the path lifetime. Hence (a) is true.

Proof of (b). We have already seen in Section 1 that when h is a potential


almost every h-Brownian path has a finite lifetime. If Do is an open relatively
compact subset of D containing and if T' is the hitting time of OD, by
then (2.1) yields

ItDo(S,h
P" {S" > T") =
h(C)

as in the proof of (a). Since his now a potential, the right side of this equality
can be made arbitrarily small by choosing Do sufficiently large (Section
I.VIII.I I); so almost every h-Brownian path from has closure a compact
subset of D. Now consider a particular case, h = C) for some point
C of D other than c. We show that in this case almost every path tends
to C at the path lifetime. To see this, define D' = D - {C}, and denote by
h' the restriction of h to D'. The function h' is harmonic on D', and h'-
Brownian motion in D' from can be identified with h-Brownian motion
in D from g; so it follows from (a) that almost every path tends to
OD' = {C}u aD at the path lifetime. Since we have just proved that almost
every path has closure a compact subset of D, it follows that almost
every h-Brownian path from tends to C at the path lifetime. More generally,
if h = GDµ,

G (t, , 7) = J 1 D(f, , 7) GD(, dC), (4.2)


D
, C)]

and if 0 < t, < . . . < t,,, the joint density of v"(t,), ... , Who.) is
v(S,dy

1,b11S2)...4D(tn-
f -D(t1+S,S1)GD(t2 - tn-1,Sn-1, n)GD(SYYM. O
C)
D GD(b, 0
(4.3)

This expression shows that an h-Brownian path from can be obtained by


first choosing the asymptotic path endpoint C according to the probability
distribution dd) and then choosing a C)-Brownian path from ,
as stated in Observation (2) above. The formal statement of this construction
in the language of conditional probabilities is left to the reader. Part (b)
of the theorem follows from this construction.

h-Brownian Motion for General h

Let h be a strictly positive superharmonic function on D, so that (Riesz


decomposition) his the sum of a positive harmonic function h, and a potential
680 2.X. Conditional Brownian Motion

h, . The evaluation

ht (t, , n) + 7)
(n (t, , tl) = (4.4)

of fD in terms of fo and 60= can be interpreted as follows: to obtain an


h-Brownian path from [when x], choose an h;-Brownian path
from with probability Thus the probability that an h-Brownian
path in D from tends to a point of D at the path lifetime is h2()/h(S),
and the distribution of the asymptotic endpoint is determined as described
in (b), whereas the probability that an h-Brownian path in D from tends
to the boundary of D at the path lifetime is h, ()/h(S). In the latter case the
path may or may not almost surely tend to a boundary point, depending
on h and the choice of boundary, but the probability that an h-Brownian
path in D from has some asymptotic property at the boundary is the product
of times the probability that an h,-Brownian path in D from
has this property.

5. h-Brownian Motion from an Infinity of h


If D is a Greenian subset of RN, if h is a strictly positive superharmonic
function on D, and if oo, an h-Brownian motion in D from o
is a process t with the following properties (all densities are relative
to IN).
h-BM(l) is an almost surely continuous process from o with
state space D, except that if the transition density is not stochastic and if a
trap 0 is adjoined to D to obtain a stochastic transition probability, there is
a discontinuity at the time of transition to a.
h-BM (2) w o(t) has distribution density function Gp(t, o, ) when t > 0.
h-BM(3) w0(-). is Markovian with transition density function Go on D;
equivalently, under h-BM(2), the reverse transition density (transition from
q at time t to at time s. with s < t) is

t' 'D(S, o, S)eD(t - S, S, 11)


(5.1)

Suppose now that all probabilities are multipled by This change does
not affect the conditional densities in h-BM(3) but replaces h-BM(2) by
h-BM(2') wz(t) has distribution density function GD(t, 5o, )h when
t > 0. The fact that the measure space on which the process is defined now
has measure which may not be I causes no difficulty. From now on
an h-Brownian motion from o will mean as before a process satisfying
h-BM(1)-(3) when +x, but satisfying h-BM(l), (2'), (3) when
+oo. In the latter case the measure space on which the process is
5. h-Brownian Motion from an Infinity of h 681

defined necessarily has measure + w, and in fact

P" {w (0) _ o} = P"{wwa(1) = a} = +oo

when t > 0. We have not yet shown that an h-Brownian motion from an
infmity of h exists, and we now proceed to do so. We shall construct the
desired process on the space 0 of all functions from R' into D v (a) with
value o at the parameter value 0. Observe first that if a process exists,
satisfying either h-BM(l)-(3) or h-BM(l), (2'), (3), then if 'eD and if
s' > 0, and whether is finite or not, the process conditioned by
wwo(s) has the following properties: the process on the parameter
set [0, s'] is independent of the process on the parameter set [s', + oo
the process on the first parameter set is Markovian, with reverse transition
density (5.1); the process on the second parameter set is an h-Brownian
motion from '. Now a process can be constructed on the function
space Q with the finite-dimensional distributions of the so-conditioned
process. In fact the standard procedure of Kolmogorov can be used, in
which a measure P. is assigned to 0 making the family of coordinate func-
tions {x(t), to R+} have the specified finite-dimensional distributions for
strictly positive parameter values, with x(0) = o. The class of measurable
sets is the smallest class making every coordinate function measurable.
The probability assigned to Q itself in this way is 1. If now ' is given the
distribution on D with density v'D(s', o, )h, the probability measure P.
on t) becomes a measure

P. = J 'D(s',

with P,".(S2) the latter value is finite according to Theorem


IX.6. The sequence {P',,' n z I) of measures on 0 is an increasing sequence
with the property that if S is a measurable subset of Sl and if S' = S n
{x(1/n)eD}, then PiM{S'} = Pig,,,{S'} when n > m. The limit of this in-
creasing sequence of measures is a measure P" on Q for which P"{Q} _
5 + oc and for which the process is Markovian and satisfies
h-BM(2'), (3). The measure P," is the restriction of P" to the class of measur-
able subsets of {x(s')cD}. The measure P" can be expressed in terms of
Brownian motion following Section 2 without the measure normalization
used there. That is, if {w is a Brownian motion in D from
with lifetime S, ifs' > 0, and if A e.F(s'), then the measure

J h[wro(s')]dP
AnfS>Y

assigns a distribution to {woo(s), s:5 s'} which has the same finite-dimensional
distributions as {x(t), t S s') on the set {x(s') a D} under P. It follows,
682 2.X. Conditional Brownian Motion

as in the corresponding discussion in Section 2, that has a stochastic


modification which is continuous except at the time of transition
to 0; w .satisfies h-BM(1), (2'), (3) as desired.
EXAMPLE. Let h = Then an h-Brownian motion in D fromo
is a process whose paths have finite lifetimes and tend back to o at these
lifetimes.

6. Brownian Motion under Time Reversal


Let D be a Greenian subset of RN, and let It be a strictly positive harmonic
function on D. Let w4"(-) be an h-Brownian motion in D from , with lifetime
S4'. It is tempting to conjecture that this process reversed in time, so that
the paths tend to at their lifetimes, is a motion. The
following discussion shows that this conjecture is correct if suitably inter-
preted.
Denote by * the Alexandrov point in the one-point compactification
of D, and define

wsti (t) =
C ift<0
ift>S. h

The extended process is a Markov process on the parameter interval R,


with nonstationary transition function, and the process reversed in time
is therefore also Markovian. This fact is the motivation for the following
discussion. It will be useful to reverse the time from a random origin. Let
z be a positive random variable, independent of the process w{ with the
distribution 1, . More precisely replace the space .0 on which the given random
variables are defined by the product space Q x R, let z be the second co-
ordinate function of this space, and define the measure Q on this product
space as the completed product measure of the given 0 measure P" and 1, .
The random variable w2(t) becomes a function, also denoted by wf(t), on
the product space. Define x(l) for t e R as w{ (z - t), so that x(t) is a function
defined on a space of infinite measure. An elementary calculation justified
by Fubini's theorem shows that the restriction to D of the distribution of
x(s) is absolutely continuous relative to IN with

S)h(C)
Density at S in D of the x(s) distribution = (6.1)

Furthermore for i > 0 the conditional distribution of x(s + t), given x(s)
in D - can be chosen to be absolutely continuous on D for each value
of x(s), with

(Density J.r(s)),x,, °(t, q, Q


Gn(S. q)
(6.2)
6. Brownian Motion under Time Reversal 683

More generally the conditional density in (6.2) is unaltered if the condition


x(s) = n is replaced by the more restrictive condition

x(st) = nt, .. , x(s,,) = %, x(s) = n

for s, < . . < s < s and nje D. Thus for any choice of initial parameter

value to the restriction of the process x(to + ) to the parameter set I8+ and
to the tZ x R set {x(to) a D} = {0 5 z - to < S,*) is a Markov process with
initial distribution specified by the density (6.1) (which involves h) and
transition density (6.2) (which does not involve h) as long as the x(to + )
paths lie in D. More precisely, if the process is killed when the paths reach
, so that the process lifetime is z - to, the so-restricted and killed process
x(to + ) is a motion, with initial distribution of density
(6.1). Furthermore an easy adaptation of Theorem VI.9 (strong Markov
property) to the present context shows that if A is an analytic relatively
compact subset of D and if T' is the hitting time of A by then the process
x(T' + ), restricted to the set { T' < + oo } and killed when the paths reach
, is a motion process. More precisely, if we define a
filtration of 12 x R by setting ..'(t) for t in I8 to be the a algebra
generated by the P x 1, null sets and.F{x(s), - cc < s:5 t}, then the process
{x(T' + 0,.47'(T' + t), to R+}, restricted and killed as just described, is a
motion process. Going back to the original space f2,
define L' as the last hitting time of A by and let be the filtration
of Li obtained by defining .F(t) to be the a algebra of subsets of f2 generated
by the P null sets and .F{w'(L" - s),s 5 t}. Then what we have proved
yields in this context that the process { ws(L" - t), .fi(t), t e I8+ } killed at
time L is a motion, with initial distribution the dis-
tribution of w{ (L{). The distribution of w{ (L) will be evaluated in Section 10.

EXAMPLE (a). Let D. be an increasing sequence of open relatively compact


subsets of D with union D, let be a point of Do, and let L4 be the last
hitting time of D. by Then we have proved that the process ws(L' - ),
killed at time L', is a motion. Observe that the initial
distribution of this process is supported by 8D and that L{, is an increasing
sequence, with limit SS, the lifetime of the wS process. We now show that
if SS is almost surely finite then the process w{ (SS - ), with lifetime S,",
is a motion on the open parameter interval ]0, + oo [.
In fact, if f is a positive continuous function on D, with compact support,
and if 0 < s < t, then

E" {(f [w4 "(L" - t)] I'(s17)J(07) (6.3)


= J GD(t
D
- s, w2 (L{ - s), n) GD(,
wC(L{e
s))1,v(dn) a.s.

(Markov property). When n oc, Fatou's lemma yields the inequality


684 2.X. Conditional Brownian Motion

E"{ f t)] ms(s)} > E"{ f [ws(S{ - s)} a.s., (6.4)

and there is actually almost sure equality because the expectation of each
side is Eh { f [ww (SS - t)] }. It follows that the process vh (S, - ) is a
Brownian motion, as asserted. In particular, suppose that D is provided
with a boundary OD by a metric compactification with the property that
lim,tshw4'(t) exists almost surely. Define w,(S,) as this limit where the
limit exists, and define w,h(Sh) arbitrarily elsewhere. This convergence con-
dition is satisfied if the boundary is the Euclidean boundary and h = I or
at least if h has a strictly positive harmonic extension to a neighborhood of
D. When D is connected, we shall show that (Theorem 3.111.2) that this
convergence condition is satisfied if and only if the boundary is h-resolutive.
If this convergence condition is satisfied, the process { W4 (S4' - t), t e OB+ },
killed at time SS, is an almost surely continuous Markov process whose
initial distribution is supported by aD. We shall prove in Section 3.11I.2
that this initial distribution is the h-harmonic measure Observe
that, whether or not a boundary is introduced, if w,(-) is an h-Brownian
motion in D with initial distribution ). and lifetime Sx, and if S," is almost
surely finite valued, then the process {wx(S,,h - t), t > 0}, killed at time
S;", is a GDJ.-Brownian motion, which can be extended to the parameter
set R with initial distribution supported by aD if a boundary is introduced
with the above stated properties. We shall describe the wx(S,", - ) process as
a GD,t-Brownian motion whether or not the parameter set is extended to R'.

EXAMPLE (b). Let o and S, be distinct points of D, and define B = D -


If we define h as the restriction to B of the preceding example is
applicable to B and h. Observe that (Section 4) almost every h-Brownian
motion path in B from So has a finite lifetime L and tends to , at the path
lifetime. According to Example (a), the GD(c, , )-Brownian motion from
o reversed at time L is a GD(a;0, )-Brownian motion. Thus, for example,
in investigating limits of functions at , along Brownian motion paths from
, we have seen (Section 2) that we can replace the Brownian motion by
a motion from , to 5o, and we now see that the paths
of this conditional Brownian motion can be identified with the paths of
a motion from o to ,.

7. Preliminary Probabilistic Solution of the Dirichlet Problem


for h-Harmonic Functions; h-Brownian Motion Hitting
Probabilities and the Corresponding Generalized
Reductions
In this section is an h-Brownian motion in a Greenian subset D of
R', from the point of D"; has lifetime S2, and the entry [hitting]
time by µ 2() of a subset A of D is T'4"[TT"]. When h - 1, the superscript I is
dropped from the notation.
7. Dirichlet Problem, Hitting Probabilities, Generalized Reductions 685

Probabilistic Solution of the Dirichlet Problem for h-Harmonic Functions


Suppose here that D is bounded and that h not only is harmonic but has a
strictly positive harmonic extension to an open neighborhood D, of D.
According to Section 2, an h-Brownian motion in D can be identified with
an (extended h)-Brownian motion in D, killed at the hitting time of aD.
Therefore almost every path tends to a point w,'(S2 -) of aD at the
path lifetime. In particular, if h = 1, then according to Theorem 1X.13 the
harmonic measure is the distribution on aD of tv,(S -). We shall
now prove the corresponding fact for µo(S, ). The representations of h-
Brownian motion probabilities in terms of Brownian motion probabilities
and of h-harmonic measure in terms of harmonic measure (Section 1.VIII.8)
yield for B a Borel subset of OD,

1 Bh)
I %,'(S,'-)E B) = f h[w,(S,-)] dP -_ = IUD( . B).
i Jiw4(S4-)FB) h() (7.1)

That is, extending tri\'ially the meaning of "hitting," the h-harmonic measure
is the hitting distribution of )v'(-) on OD, as was to be proved. Further-
more Theorem IX. 13 can now be translated directly into a theorem on h-
Brownian motion, with no change of proof, but recall that D here is bounded
and that h here has a harmonic extension to an open neighborhood of D.
The general case, for arbitrary Greenian D and strictly positive harmonic
h on D is treated in Section 3.11.2 and is more delicate because even the
Euclidean boundary of a Greenian set is not necessarily h-resolutive for
every strictly positive harmonic h.

Hitting Probabilities for h-Brownian Motion (h Not Necessarily


Harmonic)
Fix A and define
(P''{T4A <S2}
u()- 0

Since u is e,-excessive (Section VI.12), u = v/h on D' for some positive super-
harmonic function v on D. Here v:!5; h on D'; 'so v < h on D. In particular,
if h is harmonic, the function u is h-superharmonic, and the proof given in
Section IX.9 for the case h = I shows that u is h-harmonic on D - A. The
following discussion for general h puts hitting probabilities into the context
of reductions.

h-Brownian Motion and Generalized Reductions


Let D be provided with a boundary OD by a metric compactification, and let
A be a subset of D u aD. Let h be a strictly positive superharmonic function
686 2.X. Conditional Brownian Motion

on D, and if u ( qtt + x on Dh) is a XD-excessive function, define the reduction


hR. as the infimum of the class of 4'-excessive functions which majorize u
both on A n D and near A n OD. This reduction has already been defined in
Section 1. VI 11.1 when h is harmonic, and trivially 'R. = R.A. According to
Section I. every 4D'-excessive function u vanishes on D - Dh, and there is a
superharmonic function, which in this section we shall denote by [uh]. such
that u = [uh]/h on Dh. Obviously hRA = u = 0 on D - Dh, the reduction hRA
is not changed if A is replaced by A n (D' u aD), and
RAr(DhuPD(
hRY A tdhlh on Dh. (7.2)

The most natural smoothing of 'R.A is the Go-excessive function equal to 0


on D - Dh and to
RAn(DhudD1
+tuhl
h

on Dh. and in this section we shall abuse notation by denoting this smoothing
by hRA even though this /p-excessive function is not necessarily the lower
semicontinuous smoothing of 'Ru . With this definition, hRA < 'R. and there
is equality quasi everywhere on D, in particular [Section 1.VI.3(b)], there is
equality on D - A.
Recall from Section IX. 14 the reduction evaluations

RA(S) = E:z,[us(7:A)] I. (7.3)

Ejt [w,(T )]} (7,3sm)

for r a positive superharmonic function on D and A an analytic subset of D.


The conditions TEA < T, and T! < T`'D in Section IX.14 are unnecessary
here because we have defined as a Brownian motion in D and r is by
convention set equal to 0 at the process trap point. Observe that

E{r[x'(TA)] } = TA < TAD}

when c . We shall now generalize (7.3) and (7.3sm) to the context


of conditional Brownian motion and shall also provide D with a boundary
and allow A to contain boundary points. A few preliminary remarks will be
needed. In the first place observe that if as above u is a Gp-excessive function
on D, then (defined as 0 at parameter values -Sr) is an almost
surely right continuous supermartingale (Sections 1 and 2) and so has almost
sure left limits. In the second place recall that if h, is the (Riesz decomposi-
tion) harmonic component of h, then is the probability that a
7. Dirichlet Problem, Hitting Probabilities, Generalized Reductions 687

path tends to 8D at the path lifetime Sr. Such a path does not necessarily
tend to a single boundary point, however, and for each point w" of the
probability space on which is defined we denote by rr(w") the compact
cluster set on 8D of a path w") which tends to 8D at the path lifetime.
To avoid messy typography, we write I$(A) for the indicator function of the
boundary subset rr n A and write l" for the indicator function of D".

Theorem. If u is a Go-excessive Junction ( + oo on D") and if A is an analytic


subset of D u aD, then for e D",

"R: E" {u[wr(TT"")] l iT{""<s{l


(7.4)
+ Ir(A) llr{ "xshl lint u[wt(t)] },
s''

"RA
E"{u[wr(T"")] 1, "< I

(7.4sm)
+ Ir(A) ltrh"as41 ltm u[w{(t)] }.

Observe that in particular for e D"

P"{T{"" < Sr or rh A # 0}. (7.5)

Replacing TT"" here by TtA yields the smoothed reduction.


Since the equalities "R. "RA () and Ti"" = T A hold when e
D" - A and since (7.4) is trivial when e A, it follows that (7.4sm) z (7.4).
Conversely, (7.4) = (7.4sm) because if e D" n A, an application of (7.4) to
A - { } yields (7.4sm). From now on we shall assume that u is bounded
since the theorem for bounded u can be applied to u A (nl") to yield the
general result when n -* + oo.
Proof when A e D. When A c D Equations (7.4) and (7.4sm) reduce to

"RA(s) = E"{u[wr(Ti"")]}, (7.6)

E D",
"RA E"{u[iv(Tr")]}, (7.6sm)

which are merely reformulations of (7.3) and (7.3sm) by way of (7.2) and the
relation between Brownian motion and h-Brownian motion probabilities
discussed in Sections 2 and 3.
Proof when A n 8D is a countable union of compact sets. If A n 8D is
compact, let B, be a decreasing sequence of open subsets of D u OD with
intersection A n OD, and define A. = (A u n D. By definition of the
reduction operation

lim"RA^="RY.
n-'an
688 2.X Conditional Brownian Motion

Furthermore (7.4) is true when A is replaced by the subset A. of D, and

lim En{u[wV (Ti"".)]}


rm
is equal to the right side of (7.4) by the dominated convergence theorem.
Since (7.4) is true when A n 3D is compact, this equation is true when A n OD
is a countable union of compact sets. In fact, if A. is an increasing sequence
of analytic subsets of D u OD with union A, if A n D = Ao n D, and if each
set A. n 8D is compact, then limn-,,"RA n = "Rµ by Section 1.VI.3(e) for
and if A. replaces A on the right side of (7.4), integration to the limit
is permissible because only I,(An) actually changes when n changes, and
I,'(A,) is a monotone increasing sequence with limit I,(A).

Proof in the General Case

Fix SaD", write A=BuC, where B=AnD and C=An8D, and


consider the function for a fixed analytic subset B of D and a
varying compact boundary subset C. This set function is strongly subadditive
according to Section 1.VI.3(j) for Rjh1. If C. is a monotone sequence of
compact boundary subsets with limit C, then

lim "Re"cn = lim R °nlc" = RAl = nR'


n-m n-. /1 h

on D". In fact this equation was proved in the preceding paragraph in the
increasing case, and this equation is true in the decreasing case by definition
of the reduction operation. Thus each side of (7.4) defines a topological
precapacity on OD for fixed A n D and fixed S in D". Each side is equal to
the extension (Appendix 11.8) of its precapacity to a Choquet capacity for A
analytic because as proved above for each side the value on an F. boundary
subset is the supremum of the values on compact subsets. Hence (7.4) is
true.

Parabolic Context
Theorem 7 and its proof can be translated directly into the parabolic context.

8. Probabilistic Boundary Limit and Internal Limit Theorems


for Ratios of Strictly Positive Superharmonic Functions
Let h and t' be strictly positive superharmonic functions on a connected
Greenian subset D of I8", and let v" and v be the respective associated Riesz
measures. According to Theorem I.XI.4, the function u = v/{t has a fine
8. Boundary Limit and Internal Limit Theorems 689

limit at quasi every and vb + vv almost every point of D. According to


Theorem 1.XII.19, if h is harmonic, the function u has a minimal-fine limit
at Mh almost every Martin boundary point of D. In this section we shall
give probabilistic versions of these results. No boundary of D will be involved.
The basic theorem is the following, the counterpart for conditional Brownian
motion of Theorem IX.7 for Brownian motion. Theorem 8(a) asserts that u
has a limit along almost every h-Brownian path at the path lifetime. This
result will be applied in the present section, for h a potential in which case
the paths tend at their lifetimes to points of D, to derive the internal limit
Theorem 1.XII.19. It will be seen in Section 3.111.4 that when his harmonic,
in which case the paths tend at their lifetimes to the one-point Alexandrov
boundary, Theorem 8(a) yields the existence of the minimal-fine boundary
limit function derived in Theorem 1.XII.19.

Theorem. Let h and v be strictly positive superharmonic functions on a Greenian


subset D of R , and let {w be an h-Brownian motion in D from a
point of Dh = {h < +oo}, with lifetime Sh. It is supposed that Fh(O)
contains the null sets. Then if u = v/h,
(a) lim,ts#u[w{(t)] exists (finite) almost surely.
Define .Fh(+ oo) = 1/E a' "W"(1),

x'h(t) = X14(t) = u[w2(t)] if t < SS ;


(8.1)
x'h(t) = 0, xh(t) _ im u[w4'(s)] if s4 < t < +oo.

(b) The processes {xh( ), ()} and are almost surely


continuous supermartingales except that the parameter value 0 is to be
omitted if +co and that may have a jump discontinuity
at the parameter value S2.
In (c)-(e) it is supposed that h is harmonic and that the h -potential,
singular h-harmonic, and quasi-bounded h-harmonic components of u
(Section 1.IX.11) are denoted by up, u,,,,, and uqb, respectively.
(c) I fu = un or u = the limit in (a) vanishes almost surely. If'u = umgb,
the process is a uniformly integrable martingale.
(d) Eh{lim,tsku[w4h(s)]}.
(e) If h is a minimal harmonic function the limit in (a) is almost surely
infD u.

Observation. If h is harmonic, almost every path tends to the


Alexandrov one-point boundary of D at the path lifetime; so (a) can be
interpreted as a boundary limit result for every choice of boundary for D.
When S, S t <_ + oo, the random variable x4(t) can be defined arbitrarily on
the null set on which the limit in (a) does not exist. Note that if
u is necessarily continuous at ; so and are almost
690 2.X. Conditional Brownian Motion

surely continuous at the parameter value 0 in this case even though this
parameter value must be excluded in the supermartingale assertions.
The proof of (a)-(d) follows that of Theorem IX.7 (the special case
h = 1) and is therefore omitted. To prove (e), observe that u = up + u,,,, + uqb ;
so in view of (c) it is sufficient to show that u,,,Qb is identically constant, and
this constancy follows trivially from the minimality of h.

Extension to Lower-Bounded h-harmonic Functions


If h is harmonic, in Theorem 8 the hypotheses on v can be weakened : instead
of positivity it need only be supposed that v Z ch for some constant c; that
is, it need only be supposed that u is lower bounded. This case can be reduced
to that of the theorem by replacing v by v - ch.
composed process is a martingale.

EXAMPLE. Suppose that D is a Greenian subset of R', that v is a positive


superharmonic function on D, and that C ED, and define h = GD(C, ). Since
GD(C, ) is minimal harmonic on D - (Section 1.VII.10) and since
GD(C, )-Brownian paths from a point of D - {C} almost all tend to C at the
path lifetimes, the function u = v/GD(t;, ) has infou as limit along almost
every GD(C, )-Brownian path from any point of D - {C } to C. According to
the symmetry result in Section 6, the function u has this limit at C along
almost every conditional Brownian path from C to a second point of D;
equivalently (Section 2), u has this limit at C along almost every path of a
Brownian motion with initial point C. That is (Section IX.15), f lim,,-; u(rl) _
infD u, as already proved nonprobabilistically [see Theorem I.XI.4(c)].

Application of Theorem 8 to Derive an Interior Limit Theorem


As noted at the beginning of this section, it was shown in Section I.XI.4
that v/h has a fine limit at vh almost every infinity of h. To derive a probabil-
istic version of this result, observe first that the harmonic component
of h in its Riesz decomposition does not affect the truth of this assertion;
so we can assume that h is a potential, h = GDVh. Now according to Section
4, almost every path tends to a point C of D at the path lifetime, the
distribution of C is and the conditional distribution of
given the asymptotic endpoint C is that of GD(C, )-Brownian motion
from . Hence according to the preceding example, at
almost every point C of D, equivalently, at v,, almost every point C of D, there
is a finite number c = c(S) such that the function v/h has limit C along almost
every GD(C, )-Brownian path from an arbitrary point of D - {C } to C, that
is, v/h has fine limit cat C. This limit result is trivial unless v(C) = h(C) = ,1- cc
because v and h are continuous in the fine topology. Thus we have proved
probabilistically that v/h has a fine limit at vb almost every point of D, but
further analysis, for example that in Section 1.XI.4, is necessary to identify
the limit as at vh almost every infinity of h.
9 Conditional Brownian Motion in a Ball 691

Comparison of the Nonprobabilistic and Probabilistic Fatou Boundary


Limit Theorems

I f h is harmonic, Theorem 8 states that u has a finite limit along almost every
h-Brownian path from a point of D to the one-point boundary of D. On the
other hand Theorem I.XII.19 states that u has a minimal-fine limit at MI,
almost every Martin boundary point of D. These two results will be seen to
be equivalent in Section 3.111.4 by means of the following reasoning. Let K
be a Martin function. It will be shown that almost every path tends to
a minimal Martin boundary point at the path lifetime and that the distribu-
tion of ww (SS -) is the harmonic measure p' on the Martin boundary.
In particular, it follows that if C is a minimal Martin boundary point, almost
every K(C, )-Brownian path from c has limit C at the path lifetime. Further-
more it will be shown that for arbitrary strictly positive harmonic h the
distribution of can be constructed by choosing a minimal Martin
boundary point C with the distribution dC) and then choosing K(C, )-
Brownian paths from . More precisely the conditional distribution of w20
given w, (S2 -) = C is the distribution of K(C. )-Brownian motion from . It
follows from Theorem 8 that for almost every minimal boundary
point C the positive h-superharmonic function u has a limitf(C) along almost
every K(C, )-Brownian path from to C. It will be shown that the valuef(C)
does not depend on the path, does not depend on , and is in fact the minimal-
fine limit at C whose existence is asserted in Theorem XII.19 of Part 1. The
equivalence between the latter theorem and the present theorem in the con-
text of the Martin boundary lies in the facts to be proved in Chapter Ill of
Part 3 that conditional Brownian motion on the Martin space can be gen-
erated as described above and that a function has a minimal-fine limit fi at
a minimal Martin boundary point S if and only if the function has limit #
along almost every K(C, )-Brownian path from a point of D to C. Precise
statements will be given in Chapter III of Part 3. In Section 9 it will be seen
that the preceding reasoning is easily carried through when D is a ball, in
which case the Martin boundary is the Euclidean boundary. It was shown
in Sections l.XI1.19 to 1.XII.23 how the fine topology Fatou theorem yields
the classical one for a ball or half-space, in which the classical boundary
approach is nontangential or normal.

9. Conditional Brownian Motion in a Ball


Let D = B(0, b) in RN, and let K be the ball Poisson kernel,

K(C, n) = aN-Z IC - InIZ ICI = S, 1171 < o.

The function K(C, ) is minimal harmonic on D (Section 1.11. 16) with value 1
at the origin and limit 0 at every ball boundary point except C. Moreover
692 2.X. Conditional Brownian Motion

(Section 1.11. 1) f D K(C. q)/N(dq) < + o o. Fix C and consider a K(C, )-Brownian
motion from a point 5 of D. According to Section 1, the lifetime of the
process is almost surely finite because

/(C. ,t) = I ,q)K(r,)lx(d)


JD
(9.2)

< (2no2 t)-N/2 f K(C. q)IN(dq) 0


n

when t - + ,x . More generally, if h is an arbitrary strictly positive harmonic


function with Riesz- Herglotz representation

h(1) = f K(C, q)MN(dC), (9.3)


n

the lifetime of an h-Brownian motion from s is almost surely finite because


the probability that the lifetime is at least t is f,,D1(C,C, t)M,,(dC)/h(C), which
has limit 0 when t -+ + x. Since the lifetime of an h-Brownian motion when
h is a potential is also almost surely finite (Section 1), it follows using the
decomposition of conditional Brownian motion in Section 4 that every
conditional Brownian motion on a ball has an almost surely finite lifetime.
If C is a ball boundary point, almost every K(C. )-Brownian path from C
tends to the boundary at the path lifetime (Theorem 4), and the function
l /K(C, ) has a finite limit along almost every such path (Theorem 8). It
follows that almost every K(C, )-Brownian path has limit C at the path
lifetime. Again let h be an arbitrary strictly positive harmonic function on D
with Riesz- Herglotz measure M,,, and recall (front Section 1.V111.9) that
/z,",(5, dC) = K(C, C)Mh(dO/h(C). If 0 < t, < . . . < t and if is an h-
Brownian motion in D from C with lifetime V. then the joint density of
u (11 ), .... in D relative to /N is

r = K(C. ). (9.4)

This expression shows that h-Brownian motion paths from 5 can be obtained
by first choosing the asymptotic path endpoint C according to the distribu-
tion and then choosing K(C, )-Brownian paths from c to C. Thus
+(S"-) exists almost surely and has the distribution pD(C, as already
proved (Theorem IX. 13) when D is an arbitrary Greenian set, h = 1, and 8D
is the Euclidean boundary. In Section 3.11.2 this evaluation of µ will be
extended to every pair (h, OD) for which OD is h-resolutive, in fact extended
in a natural sense to every pair (h, OD).
If h = h, + h2 is an arbitrary strictly positive superharmonic function on
the ball with h, positive harmonic and hz a potential (Riesz decomposition),
10. Last Hitting Distributions: Capacitary Distributions 693

the structure of an h-Brownian motion from S with lifetime S,' is now


clear from the decomposition in Section 4. [It is supposed that is
finite.] In intuitive language a path either is an ht-Brownian path
[probability tending to a boundary point w (S{ -) or is an
h2-Brownian path [probability h2(i)lh(i)] tending to an interior point
and the distribution of the asymptotic path endpoint has been
found in both cases, respectively, in this section and in Section 4. The
extension to the case when +oo is left to the reader.
It will be seen in Section 3.111.1 that conditional Brownian motion on an
arbitrary Greenian set has a similar structure if the set is provided with the
Martin boundary. The process lifetime need not be almost surely finite,
however.

The Probabilistic Fatou Theorem for a Ball

Recall that for a ball the Fatou theorems involving radial and nontangential
approach to the boundary, and the relations between those theorems and
those involving the minimal-fine topology boundary approach, have already
been discussed in Sections 1.11.15, I.XII.19 to 1.XII.23. According to
Theorem 8, if v and It are strictly positive superharmonic functions on a ball
D and if w is an h-Brownian motion in D from , with lifetime S{, then
(almost surely) the left limit vlh[(S, -)] exists and is finite. Now suppose
that h is harmonic. In view of the structure of h-Brownian motion Theorem 8
in this case is equivalent to the statement that at M" almost every ball
boundary point C, that is, at every boundary point up to an h-harmonic
measure null set, the function v/h has a finite limit along almost every
K(C, )-Brownian path from to . The fact that the existence of a limit in
this sense at is equivalent to the existence of a minimal-fine limit at C will
be proved in Section 3.111.3.

10. Conditional Brownian Motion Last Hitting Distributions;


The Capacitary Distribution of a Set in Terms of a Last
Hitting Distribution
Let D be a Greenian subset of 68^`, let It be a strictly positive superharmonic
function on D, let D" _ h < +oc}, and let be an h-Brownian motion
in D, with lifetime Si', from in D". Let A be an analytic subset of D. Recall
(7.5), according to which that is, is the probability that
aw path hits A at a strictly positive time; equivalently, if L" is the last
hitting time of A by

R"()
hO = P{Lf > 0}.
694 2.X Conditional Brownian Motion

Hence almost no path hits A arbitrarily near 1W if and only if when B


is an open relatively compact subset of D, the value Rh can be made
arbitrarily small by choosing B sufficiently large. According to Section
I.VI[I.I I. this condition is satisfied if and only if Rn is a potential, as we
suppose from now on, Rti = GDA . Under this hypothesis is well
defined aside from the points of the probability space for which
1_ = V. in which case we define wh(S, - ), almost surely a point
of D. (Recall from Section I our convention on the generic use of P and E.)

Theorem. For RBA = G,,41 as Just described the distribution oJ' it, (Lh) on
1) - ; ; i.A giren h.r

AA(dn)
(10.1)

and P;ii (Lh) PAL' = 0; = I - Rn (5)/h(S).

According to Section V1.15, on the parameter set ]0, + r z[ the


process killed at L" becomes an motion from except that the
Rh -Brownian motion probabilities are to be multiplied by Rh
According to Section 4, the distribution of R'-Brownian motion at the path
lifetime is G (E, n)'.;(dn)jRh and (10.1) follows. The second assertion is
now trivial

Application to Capacitary Distributions

Since R' = Gni..a is the capacitary potential of A with respect to D. we


find from (10. I) that (h = 1) the distribution of on D - is
G,,((". Thus the last hitting distribution of A by a Brownian motion
determines the capacitary distribution aA in a very simple way. The first
hitting distribution of A by ;t leads (Section 1X.14) to the sweeping
kernel i)' and, in particular, leads to harmonic measure: the last hitting
distribution leads to capacitary measure.

11. The Tail 6 Algebra of a Conditional Brownian Motion


Let D be a connected Greenian subset of R', let h be a strictly positive
harmonic function on D, and let be an h-Brownian motion in D from
with lifetime S. Let D be the one-point compactification of D, and make
the adjoined point a trap for so that is this adjoined point for
S'' < i < + i Let f fi(t) be the 'a algebra of subsets of the tt ) probability
.
11. The Tail a Algebra of a Conditional Brownian Motion 695

space generated by the null sets and .F{wa(s), s S t}. In Section VII.6 it was
supposed that h _= 1 and the asymptotic properties of paths and of
functions on these paths, as the parameter value tends to 0, were investigated.
In Section 2 of the present chapter it was shown that the results in Section
VII.6 do not depend on the choice of h. In this section we investigate the
dual questions, in which the parameter value tends to SS instead of 0.
In Sections VII.6 and in Section 2 of the present chapter the key was the
initial a algebra, denoted by .fit in Section VII.6. In this section the corres-
ponding role is taken by the tail a algebra of sets, determined by w{ as the
parameter tends to Si'. The natural definition of this a algebra I,' is the
following. If B c D, let Spa be the hitting time of 8B by 4 0. Define I,' as
the a algebra generated by the null sets and the a algebra

(1 { t), to 68+} : Ba , B open, relatively compact in D}. (11.1)

This intersection is unchanged if B is restricted to an increasing sequence of


open relatively compact subsets of D with union D. Observe that I{ is a a
algebra of subsets of the probability space, a space which may vary
with . Nevertheless the or algebra (11.1) is defined by conditions on
sample functions which are meaningful for all .
If u is a function from D into A, define

uS = limsup u[µ(t)]. (11.2)


,tsc
Let Fj' be the smallest a algebra of subsets of the probability space
containing the null sets and for which u, is measurable whenever u is Borel
measurable. The a algebra cj' is generated by the null sets and the algebra
of sets of the form

{[u14, ...,u"jeA'} (11.3)

for n - 1, A' a Borel subset of fl", and u; a Borel measurable function from
D into A. A trivial map shows that 08 can be replaced here and in the defini-
tion of !§j" by an arbitrary compact subinterval of R.
The following results (a)-(c) on the tail a algebra and associated concepts
will be used in later chapters. A characterization of the tail a algebra in terms
of a suitable boundary of D is given in Section 3.11.4.
(a) Let u be a function from D into 08 with the property that the set
{u > c} is analytic for all real c, define ut by (11.2), and define E{ u,}
whenever this expectation is meaningful. Then
(al) u, is !§' measurable.
(a2) If u is lower bounded, the function u' is either identically + 00 or
h-harmonic and quasi bounded.
(a3) If in (a2) the function u' is h-harmonic, then
696 2.X. Conditional Brownian Motion

lint u'[w,(t)] = u, a.s. (11.4)


'ts{
for each in D.
(a4) If u is an arbitrary Borel measurable function from D into A, then
unless E{ ju{j} = + co for all in D, the function u' is h-harmonic
and quasi-bounded, and (11.4) is true.

Application. If u is the indicator function of an analytic subset of D and if


L', is the last hitting time of A by then (a) implies that the function
P{Lhc = S,) is h-harmonic and that almost surely lim,tsh u' [ww (t)]
exists and is the indicator function of the set {L' = S,). The measurability-
type hypothesis imposed on u in (al)-(a3) was made weaker than Borel
measurability to allow for this application.

Proof of (al). Let B. be an increasing sequence of open relatively compact


subsets of D containing , with union D, and define

m, a) = hits (D - Bm) n {u > al).

Then up to a w () null set m, a) a w4 (S,Bm + t), t e U8+ } and up to a


null set

{us>a}= U n Am,a+k
k=I m=0

so u, is 114, measurable.

Proof of (a2). Suppose first that u is bounded. Then the strong Markov
property of conditional Brownian motion yields

E{u4l.g2(SfBn)} = u'[w,h(S B")] a.s. (11.5)

Hence (Section 7)

E{u{} = E{u'[ww(Shen)]I _ en( ,u'); (11.6)

so u' is h-harmonic on B. for all n and therefore is h-harmonic on D. If u is


lower bounded but not necessarily upper bounded, this result applied to
u A n implies that the function n n} is h-harmonic; so (Section
1.II.3) the function u', the limit of an increasing sequence of bounded
h-harmonic functions, is either identically + 00 or h-harmonic and quasi
bounded.

Proof of (a3). We can suppose that u is positive. Then by (a2) either u' - + 00
or u' is h-harmonic and quasi bounded. In the latter case (11.5) is applicable.
11. The Tail a Algebra of a Conditional Brownian Motion 697

,Ea.
Note that u, is I,,, measurable, that W, C and that according to
Section VI.7 the a algebra on the right here is Y.e Z. - 4 ( S 4 so when
114

n x in (11.5), the conditional expectation continuity theorem yields

Y f"(S4B^)}=limu"[w"(S''B^)] as. (11.7)

Finally lim,t'I u'[tvw(t)] exists almost surely according to Theorem 8, and


this limit is almost surely uS according to (11.7). Thus (a3) is true.

Proof of (a4). If u is Borel measurable, the preceding work is applicable with


minor changes when u, is defined by (11.2) with the inferior rather than the
superior limit. On applying these results appropriately to u v 0 and (- u) v 0
it follows that the function S -- E{ jus } on D is either identically + oo or
h-harmonic and quasi bounded, as is then also the function i-' E{u,}, and
that in the latter case (11.4) is true. The proof of (a) is now complete.

(b) s{ _". Assertion (al) implies c Conversely, fix in


D and a set A, e 1.R,. We shall now show that then A, 0;V1j'. Define B as
in the proof of (a 1) except that now we also suppose that B. c B,,, for all n.
According to the Markov property of conditional Brownian motion,

P{A4j.f(SeB^)i = P{A{Jw (S,B^)} a.s., (11.8)

and (Section 1.4) one version of the conditional probability on the right has
the form 46^[w (SnB^)], where 0^ is a Borel measurable function from aB^
into [0, 1]. We apply the conditional expectation continuity theorem, again
using the fact that

yrER',2 (t) = yMER,


1$ B^).

and find that

lA = lim a.s. (11.9)

If u is the function on D defined for all n as 0^ on O B. and defined as 0 else-


where on D, this function is Borel measurable from D into [0, 1] and

1', = limsup4[w{(t)] a.s.


rts{

Thus A, e `o h, as was to be proved.

Observation. According to (a) and (b), it is no restriction on if u in


(11.2) and the following definition of c{' is bounded and h-harmonic.
698 2.X. Conditional Brownian Motion

(c) If h is minimal harmonic, the following assertions are true.


(cl) The tail a algebra Iris trivial for every in D. _
(c2) If u is an arbitrary Borel measurable function from D into I2, there
is a (not necessarily finite) constant c such that P{u, = c} = 1 for
every in D.
In (c3)-(c5) v is a Borel measurable function from D into a Polish space D',
and d' is a metric on D' compatible with the D' topology.
(c3) Let n' be a point of D', and define

M{(ry') = {q' is a cluster value of v[w{ (t)] when t T S4 h).

Then Mh(q') a ( , and the function P{M,'(q') } is either identically 0


or identically 1.
(c4) Let A' be the set of points tI' for which the function P{M"(q')} is
identically 1. The set A' is closed and is the cluster set of almost
every sample function as t T S2. If D' is compact,

P{ lid'(A', v[ (t)]) = 0} = I
ttsh4

for every c.
(c5) The function -- P{lim,t5h v[wf (t)] exists a.s.) is either identically
0 or identically 1, and in the latter case there is a point q' of D'
such that lim,tsh v[w2 (t)] = PI' almost surely, for every point of D.

Proof of (c 1) and (c2). We shall prove (c2) which implies (c 1). It can be
supposed, replacing u by arctan u if necessary, that u is bounded. The func-
tion " E{u,} is h-harmonic according to (a2). Since u' is bounded
and h is minimal, the function u' is identically constant, u' - c, and (11.7)
then yields u, = c almost surely, as was to be proved.

Proof of (c3). A point tl' is a cluster value of v[,,' (t, a))] when t T S{ if and
only if

lim infd'(q', v[tti{(t, co)]) = 0;


it S41

so if -d'(ry', assertions (a) and (c2) combine to imply (c3).

Proof of (c4). The proof of Section VII.6(e) with obvious changes to the
present context yields (c4).

Proof of (c5). Let A, be the probability space subset for which v{ =


lim,tshv[nw(t)] exists. Then A,e19{ (Section 11.5); so P{A} is either 0 or 1.
12. Conditional Space-Time Brownian Motion 699

The function is h-harmonic by the same reasoning [see (11.6)]


making u' an h-harmonic function. Hence the function is either
identically 0 or identically 1. In the second case fix and observe that v, is
94h measurable; so v{ is almost surely a point n of D', that is,

lim s v[w (t)]) = 0 a.s.


ltsl
According to (c2), it follows that this limit relation is true for all ; so q'
does not depend on . The proof of (c5) is complete. 0

The Role of Analytic Set Theory

Although (a)-(c) as treated above involve analytic set theory where, as in


(a) and (c2), functions are discussed under very weak hypotheses, the proofs
of (b) and (c2) do not involve this theory. Analytic set theory can be avoided
in the other parts of (a) and (c) if appropriate restrictions are imposed
on the functions under discussion. For example, in (al) if u is supposed
lower semicontinuous, the set {u > c} is open, and therefore (Section VI.6)
measurability problems for hitting become trivial.

12. Conditional Space-Time Brownian Motion


Let I) be an open nonempty subset of U'B", and let h be a positive super-
parabolic function on 15. The definition and treatment of h space-time
Brownian motion in b follow their counterparts for conditional Brownian
motion in a Greenian subset of IB" except that h is allowed to have zeros.
We define the transition measure from a zero of h to be identically 0. Define
154 = {0 < Ih < + oo }. We restrict ourselves to the following remarks. An h
space-time Brownian motion from a point of 14 is an almost surely
continuous process whose sample functions never leave iY' and proceed
downward, that is, in the direction of decreasing ordinate values. If ti is a
positive superparabolic function on D, the process (v/h)[w is an almost
surely right continuous process with left limits at all parameter values and at
the path lifetime and is a supermartingale under the same conventions as in
the conditional Brownian motion context. If it is parabolic, almost every
tv{ path tends to the one-point boundary of b at the path lifetime, and
h-parabolic measure µ8(i, ) on the Euclidean boundary of an open relatively
compact subset i of I), containing , is the hitting distribution on this
boundary. If h = t) for some point it of 15 and if Oo(4, 0) > 0, then
almost every wf path has lifetime ord 4 - ord ri and tends to 1 at the path
lifetime. In general if h = d6µ is a superparabolic potential, almost every h
space-time Brownian path from a point of I'' has a finite lifetime and tends
at the path lifetime to a point it whose distribution is i(, q)µ(d1)/h()
700 2.X. Conditional Brownian Motion

The h space-cotime Brownian motion processes are defined in the obvious


dual way. When q) > 0, almost every 66(-, q) space-time Brownian
path from tends down to q at the path lifetime and almost every
space-cotime Brownian path from q tends up to at the path lifetime.
Moreover a reversal of time in one of these processes yields the other process.
Theorem 8 on the composition with conditional Brownian motion of the
ratio of two positive superharmonic functions translates directly into the
parabolic context, and its statement is left to the reader. Observe, however,
that the sample functions of the composed process, up to the conditional
Brownian motion lifetimes, are almost all right continuous but not necessar-
ily continuous, as already noted in Section IX.12 for the special case of the
composition with Brownian motion of a superparabolic function. The
parabolic counterpart of Theorem 8 can be applied to derive the existence
of a coparabolic-fine limit function for the ratio of two positive super-
parabolic functions (see Theorem 1 XVIII.14 for an exact statement, and
see the classical context argument in Section 8).

13. [Space-Time] Brownian Motion in [W'] R N with


Parameter Set II8

All measure densities in this section are relative to IN.

Brownian motion in R' with parameter set R

Let {w(t), tE R} be a Brownian motion in R' with parameter set R. Then


(Sections VII.2 and IX.8) there is a positive parabolic function ti on A"
such that for s in R the random variable w(s) has distribution density i*, s),
with integral + o0 over RN. The process has reverse time transition
density (transition from at time s ton at time t < s)

u(q, t)G(s - t, t, s).

In particular, suppose that u is a minimal parabolic function on AN, so that


(Section I.XVI.8) ii has the form

s) = exp 1<Y,
+ a' 171'sJ (13.1)

with } a point of R". Then the above reverse transition density becomes
G(s - t, , rl - a2(s - t)y). This is the transition density of a Brownian
motion with drift: conditioned by fixing w(0) = , the process

{w(-t) +azty,tER+}
13. [Space-Time] Brownian Motion in [d8"] RN with Parameter Set R 701

is a Brownian motion in R' from . According to VII(5.4), it follows that


for the conditional and therefore also for the original process,

w(t)
lim = aZ; a.s. (13.2)
f-'-aC t

Observe that if y # 0, then lim,-_. w(t) = co almost surely. If y = 0, then


(Theorem IX.5) this convergence to oo is true if and only if N > 2. If y = 0
and N = 2 [N = 1], almost every w(-) sample path hits every disk [point]
at arbitrarily large negative parameter values. If ti is an arbitrary positive
parabolic function on RN, then (Section XVI.8) there is a measure A, on RN
such that

o expL<i',
and we conclude that if a is the absolute probability density function for
vi,(-), then lim,__, [w(t)/t] exists almost surely and has distribution cr2N (dy).
Observe that the only condition imposed on 1Vj is XVI(8.3); so N;(18N) need
not be finite. In particular, if R. is the unit measure supported by the origin,
the function ti is given by (13.1) with y = 0, and so u = 1. This is the stationary
case. The choice of ti is the choice of the distribution family s . s)IN(d);
this family is called the entrance law of the Brownian motion.

Conditional Brownian Motion in RN with Parameter Set R

In the study of h-Brownian motion in IRN with parameter set H the function
h is a strictly positive superharmonic function on RN, and the transition
density (from at time s to at time t > s) is given by

(13.4)

and it is trivial that the absolute probability density function ti must be


replaced by tih. Observe that if h is harmonic we obtain nothing new because
(Section 1.11.2) every positive harmonic function on RN is identically con-
stant. Moreover (Section 1.11.13 for N = 2; the result is trivial when N = 1)
every positive superharmonic function on RN is identically constant when
N < 3. Hence in the following we suppose that N > 2 and that h is a potential.
The reverse transition density for (13.4) of an h-Brownian motion w"(-) with
parameter set H is unchanged by the change from a and t to tih and e' aside
from the fact that this density is not defined when either or t) is in the polar
and therefore IN null set of infinities of h. In particular, suppose that ii is
given by (13.1) and that h = for some point C of RN. Then as t increases
702 2 X Conditional Brownian Motion

from - x to the finite process death time, every w"e) sample path comes in
from the point oo, along a direction determined by ; if 7 # 0, and tends to
at the process death time. For general ti and general It the process is made
up of such processes. Thus in general lim,__. [w(t)/t] is a random variable
with distribution and (Theorem 4) if h = Gp and if h(c) < +oc,
the conditional distribution for iv(s) = of the left limit of at the
process death time is Roughly, the choice of « determines
at the beginning of the process, and the choice of h determines the
process at the end.

[Conditional] Space-Time Brownian Motion in l with Parameter Set R

A space-time Brownian motion in A' with parameter set 11 is a process of


the form ((w(t), to - t), to 68}, where to is an arbitrary constant and w(-) is
a Brownian motion in R" with parameter set R. A positive parabolic func-
tion a is then the absolute probability density function for as explained
above. We now study h space-time Brownian motion in R" with parameter
set R, necessarily replacing i by tih. In particular, suppose that ti and h are
both minimal parabolic functions on A', given by (3.1) with y ='to, Ti,
respectively. Then the process ((w(t), a - t)', t e R) becomes a space-time
process with the original time component. The space component is a process
with state space R" and is Markovian with stationary transition density
f(t,, ry - a2ty,); so the process { w'(t) - a2ty, , to l8+ } conditioned by
w'(0) _o is a Brownian motion from o. The reverse transition density
for arbitrary h is the same as that for h 1. Thus lim,.._, [w'(t)/t] = To
almost surely and lim,_. [w'(t)/t] = y, almost surely. The character of
for general i4 and general parabolic h can be deduced at once from this special
case, and the case when h is superparabolic is handled similarly.
Part 3
Chapter I

Lattices in Classical Potential Theory and


Martingale Theory

1. Correspondence between Classical Potential Theory and


Martingale Theory
Submartingales martingales and supermartingales are analogs in the context
of martingale theory of subharmonic harmonic and superharmonic func-
tions in the context of classical potential theory. The correspondence between
these two contexts has two aspects. In the first place many of the manipula-
tions of supermartingales correspond exactly to manipulations of super-
harmonic functions. This has been exhibited in previous chapters by the
common choice of nomenclature, for example, D, S, Sm, LM, GM, T,, R:.
In the second place under appropriate hypotheses the composition of a
superharmonic function with Brownian motion is a supermartingale; for
example, see Section 2.IX.7. In this chapter lattice aspects of classical
potential theory and martingale theory will be developed simultaneously.
Throughout this chapter in the classical potential theory context D is a
fixed connected Greenian subset of R", and h is a fixed strictly positive
harmonic function on D. The notation S', etc., will always refer to classes
of positive h-superharmonic functions, etc., on D. The martingale theory
context is the continuous parameter context as described in Section 2.IV.1;
that is, processes with parameter set R', on a complete probability measure
space (S2, ., P) provided with a right continuous filtration are treated.
It is supposed that .F(0) contains the null sets. Thus S+, .. refer to the class
of positive almost surely right continuous supermartingales on

(si, . , P), ... .

Classes L°, D, S, and so on have been defined in both classical potential


theory and martingale theory contexts, and it will frequently be possible
in the analysis to use the same proof in both contexts, by means of the
obvious translation in which "positive h-superharmonic function" becomes
"positive almost surely right continuous supermartingale" and operations
like GM, . . . are interpreted according to the context under consideration.
It will be convenient in both contexts to abbreviate the notation for the
intersection of lattices by writing S,,,q,, for S. n Sqb, Sps for S, n S, and so on.
706 3.1. Lattices in Classical Potential Theory and Martingale Theory

It is trivial that if u is a strictly positive h-superharmonic function, then


uh is a strictly positive superharmonic function, and conversely, and that u
is in a class U. D, S, etc., relative to h if and only if uh is in the corresponding
class for h as 1. Moreover, when u is a positive h-superharmonic function,
RW,/h =hRA and R F/h =hRu . Thus, although the results in this chapter are
stated for general h to facilitate reference, proofs can be given for h as 1
without loss of generality.
The Parabolic Context. We have defined the classes mentioned above
both in the classical potential theory and the parabolic contexts. In the
latter the Greenian subset D of I8' is replaced by a nonempty open subset
D of 1+8", h is replaced by a strictly positive parabolic function h on D, and
h-superharmonic functions are replaced by h-superparabolic functions. In
most but not all of the classical potential theory discussion theorems and
their proofs will be seen to be directly translatable into the parabolic context,
as noted in Section 1.XVIII.19.
The Martingale Theory Lattices 'S, etc. We have defined and discussed
these lattices in Chapter V of Part 2 under very general hypotheses. The
most useful case is that with parameter set Z', and the reader should have
no difficulty in finding the counterparts in this context of the martingale
theory continuous context results in the present chapter. Some of the results
are not applicable to all linearly ordered parameter sets, however.

2. Relations between Decomposition Components of S in


Potential Theory and Martingale Theory
In both the classical potential theory (Chapter IX of Part 1) and martingale
theory (Chapter V of Part 2) contexts we have proved that Smgb, Sm,, Spqb,
Scs are mutually orthogonal bands in S. with vector sum S. The potential
theory proofs are applicable without change to the parabolic context.
In the martingale theory context we have proved (Theorems 2N.3, 2.V.9)
that S. n D = Smgb = S. n UI. The equality S. n D = Smar, in the potential
theory contexts will be proved in Section 5.
In the martingale theory context we have proved (Theorem 2.IV.11)
that SD n D = S,, b. This equality will be proved in the potential theory
contexts in Section 9.

3. The Classes LP and D


See Sections 1.IX.3, 1.IX.4, 2.I1.11, and 2.V.4. In the classical potential
theory context we denote these classes more specifically by L°(µ4_) and
D(µ4_) when there is danger of ambiguity. In classical potential theory, in
parabolic potential theory. and in martingale theory S c L' and for p > I :

The set S n L° is a vector sublattice of (S, :S) but not a band; S n LP c D.


4. PWB-Related Conditions on h-Harmonic Functions and on Martingales 707

4. PWB-Related Conditions on h-Harmonic Functions and on


Martingales
Lemma (Classical Potential Theory Context). If the h-harmonic function u
on D is in D(µo_ ), then to each E > 0 and in D correspond a lower-bounded
h-harmonic pointwise majorant u,{ in D(µ4_) and an upper-bounded h-harmonic
pointwise minorant u6, in D(4_) with u 5 E. Conversely, ifu is an
h-harmonic function and if to each e > 0 and in D correspond a lower-bounded
h-superharmonic pointwise majorant u,t and an upper-bounded h-subharmonic
point wise minorant u,, with u is in D(v_ ).

(Martingale Theory Context). If eDn that is, if .F(-)) is an


almost surely right continuous uniformly integrable martingale, then to each
E > 0 correspond a lower-bounded martingale essential majorant x;(-) of x(-)
in D and an upper-bounded martingale essential minorant xa(-) in D with
E{xt(t) - x,(t)} < E. Conversely, if x(-) is an almost surely right continuous
martingale and if to each E > 0 correspond a lower-bounded supermartingale
essential majorant x'() and an upper-bounded submartingale essential minorant
with

sup E{x'(t) - 5e
reR

then x(-) is in D.

Proof in the martingale theory context. If is in D n S. apply Section


2.V.2 to find

E{LM[x(-) v n](s)} - E{x(s)} = sup E{x(t) v n} - E{x(s)}. (4.1)


reR

The difference on the right is constant in s and is at most c/2 for sufficiently
small (negative) n because x(-) is in D. The process x;(-) is defined as
LM v n] for such a choice of n, and the process x;(-) is defined dually.
Conversely, if x' and xl(-) exist as in the converse assertion, with x,"(-) z
- K,, and 5 K1, where K, > 0, and if b > 0 is chosen so that

Pflx(t)I = b} = 0,

then

E(lx(t)I ; jx(t)j ;>-- b} 5 E{x; (t) - x,(t)} - E{xL (t); x(t) < -b}
+ E{x;(t); x(t) > b} (4.2)
5 E + K`P{jx(t)j > b}.
It follows that sup,,, Efl x(t)j} < + oo and if K is this supremum, the right
708 3.1. Lattices in Classical Potential Theory and Martingale Theory

side of (4.2) is at most e + KK/b < 2E when b is sufficiently large. Hence


a D.
The proof in the potential theory context is a translation of that just given.
Observation. The conditions of this lemma have not received much atten-
tion in martingale theory, but they are fundamental in potential theory. In
fact by their very definition the PWB' solutions of the Dirichlet problem are
in D(µ D_ ), according to Lemma 4, for any choice of boundary. Conversely,
if the boundary is internally h resolutive (Section 1.VIII.2), there corresponds
to every h-harmonic function u in S,,,qb [ = S. n D(µo_) according to the
next section] a PWB" resolutive boundary function f with Hf = is.

Parabolic Context

Lemma 4 and its proof need no change in the parabolic context.

5. Class D Property versus Quasi-Boundedness


Theorem. In the classical and parabolic potential theory contexts and in the
martingale theory context

S. n D = S,,,qb. (5.1)

Recall that in the martingale theory context we have proved (Theorems


2.V.3 and 2.V.9) that S. n D = S,,,qb = S. n UI; so (5.1) need only be dis-
cussed in the potential theory contexts. Actually Lemma 4 yields (5.1) in both
contexts. We give the proof of (5.1) in the classical potential theory context
in the notation of Lemma 4, with h = I to simplify the typography. The proof
in the martingale theory context, as based on Lemma 4, is essentially the
same. If u e S, n D, the bounded positive harmonic function v 0)
is a pointwise majorant of u,, and minorant of u; so LM(u, v 0) < C.
Thus u is the supremum of its positive bounded harmonic minorants and
as such is in S; qb. Conversely, if ueS ;qb, it is trivial from Lemma 4 that
ueD.

Application to L° for p > I

In the following discussion we compare certain results for S.n L° in


martingale and potential theory contexts. Suppose that is a
uniformly integrable martingale (continuous parameter context), that is,
e S. n D. According to the continuous parameter version of Theorem
2.1I1.14 there is an almost sure limit x(+ oo) = lim,... x(t), and x(t) =
E{x(+ co)I,F(t) } almost surely, for all t. Moreover according to Theorem
6. A Condition for Quasi-Boundedness 709

2.111.14, E{ Ix(+ oo) IP} < + oo if and only if is L° bounded, equivalently


(by Section 2. V.4) if and only if a LP. Under the latter condition it
follows easily from Section 2.V.2 that

a.s.

The following are the classical potential theory counterparts of these


martingale theory results. Let D be a connected Greenian subset of R',
let h be a strictly positive harmonic function on D, and let u = v/h be a
positive h-harmonic function on D. According to Theorem 1.X11.10 the
Martin boundary is universally resolutive and universally internally resolu-
tive. Hence in the Martin space context the results in the present and preced-
ing sections imply that the class of PWB' solutions is S. n D(µ4_) = Smgb
According to Theorem 1.XII.19, if u = Hf then f is the h-harmonic measure
almost sure boundary limit function on approach to 8"D in the minimal
fine topology, and u = u' f ). Moreover it follows easily from Section
1.1X.2 (the classical potential theory counterpart of Section 2.V.2) that
J ]P) < + oo if and only if u e LP( 4_ ), and in that event
I I f IP) _
LMh uIP.

6. A Condition for Quasi-Boundedness


Lemma (Classical Potential Theory Context). If u E Sqb and u e S+, then

lim 0 q.e. (6.1)


C-M

(Martingale Theory Context). If e SQb, e S+, and A, = { c l,


then

lim 0 q.e. (n eV ). (6.2)


n-m

Observe that (6.1) and (6.2) are identical with the same equations for
unsmoothed reductions because {u > c} is an open subset of 1 N and A, is a
fine-open subset of Ili3+ x 0. In (6.1) the reduction decreases when c increases.
In (6.2) we can only assert that c < d implies that D"d < Oz(-) "c quasi
everywhere, and that is why in (6.2) the limit is sequential. The set lL+ in
(6.2) can of course be replaced by any other unbounded increasing sequence.

Proof in the classical potential theory context. (The martingale theory proof
is a direct translation.) If u = E0 u; with u; in S+ and bounded, then

hkrlv>`s
<_ > (6.3)
710 3.1. Lattices in Classical Potential Theory and Martingale Theory

and (dominated convergence) it is therefore sufficient to show that


t>`i = 0 for all j at the points of finiteness of v. If u; is bounded
by aJ, this desired limit relation follows from

lu al w
' c c

Observation. The relation (6.1) can also be written in the form

lim QuhO" = 0 q.e. (6.1')


C_W

Parabolic Context

The lemma and its proof need no change in the parabolic potential theory
context.

7. Singularity of an Element of S,'

Theorem (Classical Potential Theory Context). For a function u in S, to be


singular it is sufficient that u A c E Sp for some strictly positive constant c and
necessary that u A C E SD for every stric t/v positive constant c.
(Martingale Theory Context). (a) For a process in S,n to be singular
it is sufficient that A cESD for some strictly positive constant c and
necessary that A c C SP for every strictly positive constant c.
(b) The process in Sm is singular if and only ij' lim,_,, z(t) = 0
almost surely.

The proof of Theorem 7 in the classical potential theory context was


given in Section 1.1X.10, and the martingale proof of (a) is a direct transla-
tion of that proof. The martingale assertion (b) was proved in Section 2N. 1I .
A classical potential theory counterpart of the martingale theory assertion
(b) can be formulated in two ways as follows. (Classical potential theory
context) :
(bl) An h-harmonic function u in S, is singular if and only if (Theorem
I.XIl.l9) its minimal-fine boundary limit function on the Martin
boundary of D vanishes Mh, almost everywhere, that is, up to an h-
harmonic measure null set.
(b2) An h-harmonic function u in S, is singular if and only if (Theorem
2.X.8) when is an /t-Brownian motion in D from , with life-
time S,. lim,thu[ t' (t)] = 0 almost surely.
8. The Singular Component of an Element of S ` 711

Parabolic Context

Theorem 7 and its proof need no change in the parabolic context. The
parabolic criterion corresponding to (b2) does not require a new proof and
is of course stated in terms of the limit of a positive h-parabolic function
along h space-time Brownian paths.

8. The Singular Component of an Element of S+


Theorem (Classical Potential Theory Context). If u E S+ and has singular
component us, then

lim hDu0{u'cl = us q.e. (8.1)


C_00

In particular, u in S+ is singular if and only Y 'u = hOuOD> for every (equiv-


alently some) strictly positive constant c.
(Martingale Theory Context). If and has singular component
and if Ac = c}, then

lim q.e. (ne7L+). (8.2)

In particular, in S+ is singular if and only if quasi everywhere


for every (equivalently some) strictly positive constant c.

Observation. Equations (8.1) and (8.2) are equivalent to the same equa-
tions with unsmoothed reductions. See the observation in Section 6 on the
corresponding point for Lemma 6, and on the explanation of the sequential
convergence (n -+ co) in (8.2) rather than the unrestricted approach (c -* oo)
in (8.1).

Proof in the classical potential theory context. (The martingale theory proof
is a direct translation.) Suppose first that uESs . By Section I.Vl.3(i) as
adapted to reductions in the h-superharmonic context,
hOuO,u>C,
u + hOu0(u5c) (8.3)

According to the vector lattice decomposition theorem (Theorem 6 of


Appendix III), there are members u, and u2 of S+, specific order minorants,
respectively, of the first and second terms on the right, with sum u. Since SS
is a band the functions u, and u2 must be singular, and since u2 is bounded
by c, u2 E Ss n Sqb ; so u2 = 0. Thus u = u, , and u is the first term on the
right in (8.3), as was to be proved. For general u in S+ write u = uqb + us,
712 3.1. Lattices in Classical Potential Theory and Martingale Theory

the sum of the quasi-bounded and singular components of u. Then [from


Section 1.VI.3(f)]
114(u>c) = IUgb](r>CI + IUAI->cl

The first term on the right has limit 0 quasi everywhere when c - oo according
to Lemma 6, and the second term on the right is us for every c because
b1Us0(Us>ci <
Us = us,

so Theorem 8 is true.

Parabolic Context
Theorem 8 and its proof need no change in the parabolic context.

9. The Class Spqb


Theorem (Classical Potential Theory Context). The following conditions on a
function u = G0µ/h in SP are equivalent :
PT(a) ueSpgb.
PT(b) uED(pp-).
PT(c) lim, 14w>0 = 0 q.e.
PT(d) p vanishes on polar sets.
(Martingale Theory Context) The following conditions on a process in
SP are equivalent:
MT(a)
MT(b)
MT(c) If A, = (z(-,-)> c}, then

lim 0 q.e.
R-M

Observation. PT(c) and MT(c) are equivalent to the same equations with
unsmoothed reductions. See the observation in Section 6 on the corre-
sponding point for Theorem 6 and on the explanation of the sequential
convergence (n - cc) in MT(c) rather than the unrestricted approach (c - oo)
in PT(c).

Proof in the classical potential theory context. In clarification of PT(d) recall


that (Theorem 1.V.11) a superharmonic potential G0µ has the value + oo at
p almost every point of a polar set. Hence p vanishes on polar sets if and
only if µ {G0µ = + oo} = 0. To simplify the notation in the following proof,
we take h =_ 1.
9. The Class S,0 713

PT(a)=:47(b) If u = Eo ujwith each summand in SP and bounded, let


be a point of D, and let Bo be an open relatively compact subset of D
containing . To show that ueD(µD_), we show that the family of pairs

{(u, µB(, )): B - B0, B open and relatively compact in D} (9.1)

is uniformly integrable. Observe that according to the generalized super-


harmonic function average property in Section 1.VIII.10, the harmonic
average u) increases when B decreases, and therefore attains its max-
imum value when B = Bo. Now define vn = E', uj, choose s > 0, choose n
so large that v,) < e/2, and define c = supD Eo uu. When B is a set in
(9.1), the same superharmonic function average inequality yields v,) <
E/2, and if A is a Borel subset of 8B, so small that µB(;, A) < e/(2c), then

PB(S, UlA) S I1B(S, Vn) + µB (uJlA) < E. (9.2)


no

Hence the family (9.1) is uniformly integrable.


PT(b) PT(d) Suppose that for some it in S+ condition (b) is true but
condition (d) is false. Replacing µ by its projection on a suitable compact
set if necessary, it can be supposed in deriving a contradiction that µ is
supported by a compact polar set A. Let B be an open relatively compact
subset of D containing A, let be a point of B - A, and let A. be a decreasing
sequence of compact neighborhoods of A, subsets of B, with intersection A.
The boundary of B. = B - A, consists of dB and a subset C. of OA,,. Since u
is harmonic on a neighborhood of B,,

(9.3)

To show that the second term on the right has limit 0 when n oe, note
first that (Section 1.V.4) there is a function v positive and superharmonic
on B, with v = +oo on A and v z I on C. for large n;
so lim sup.- µB.0, C,) 5 and since for every S > 0 the function Sv
satisfies the same conditions as v, it follows that µ8n0, 0. Hence
(class D property) limn_.µB.0, ulcn) = 0; so

uG) = li µB.(4, Ul aB) 5 sup U. (9.4)


m
Thus u is bounded on B - A, and therefore (Section 1.V.5) u has a unique
superharmonic extension to B, and this superharmonic extension is har-
monic, contrary to fact. It follows that PT(b) PT(d).
PT(d) PT(a) If u = GDµ is a superharmonic potential for which µ
vanishes on polar sets, define A. = in S u < n + 1), and let µ be the projec-
tion of µ on An. Then p = Eo µ because y {u = + oo } = 0, and according to
the domination principle, GDµ 5 n + 1. Thus the representation u =
Eo GDµ exhibits u as a quasi-bounded potential.
714 3 1. Lattices in Classical Potential Theory and Martingale Theory

PT(a) eo PT(c) See Theorem 8.


MT(a) MT(b) If with each summand in SP and
bounded, we show that the family {z(T): Toptional} is uniformly integrable.
[Define :(+ x) = 0.] Define E. , choose e > 0, choose n so
large that E{yy(0)} < e/2, and let c be an upper bound of I" Then if T
is optional, the supermartingale inequality yields

E z(T)} <E{z(0)), E{y,,(T)} <_2e ;

so if P{A; < el(2c),

fz(r)dP < E{y,(T)} + f iz;(T)dP<t. (9.5)


A n()

Hence the family {z(T): T optional} is uniformly integrable, that is, D.


MT(b) - MT(a) See Theorem 2.IV. 11.
MT(a)' MT(c) See Theorem 8.
The proof of the theorem is now complete.

Observation (Discrete Parameter Case). In the martingale theory context


SP = Spgb if the parameter set is 7L+ according to Theorem 2.IV.8.

Parabolic Context. The proofs in the classical potential theory context


that PT(c) p PT(a) PT(b) PT(d) need no change in the parabolic
context. The above proof that PT(d) PT(a), however, used the domina-
tion principle which is not valid in the parabolic potential theory context
according to Section I.XVIl.5. Now define
PT(d') u vanishes on semipolar sets.
Then PT(d') ea PT(d) in the classical potential theory context because semi-
polar sets are polar in that context and under PT(d') the domination prin-
ciple is valid in the parabolic potential theory context (Section I.XVIII.16):
so PT(d') PT(a) in the latter context. Thus the parabolic potential theory
version of Theorem 9 is slightly weaker than the stated classical potential
theory version. It is false that PT(a) PT(d') because if 6 = U8N and if µ
is supported by the abscissa hyperplane on which µ = IN, then is supported
by a semipolar set, but Gpµ E Spq,,.

10. The Class Sps

Theorem (Classical Potential Theory Context). The following conditions on a


function u = G0p/h in SP art, equivalent:
11. Composition of an h-Superharmonic Function with an h-Brownian Motion 715

PT(a) ueS;,.
PT(b) blur'"' = u for every (equivalently some) strictlYpositive constant
c.
PT(c) p is supported by a polar set.

(Martingale Theory Context). The following conditions on a .supermar-


tingale in SP are equivalent:
MT(a) a SD5.
MT(b) If A r = c}. then quasi everywhere for
ever (equivalently some) strictly positive constant c.
MT(c) is a local martingale.

PT(a) p PT(b) and MT(a) <o MT(b) according to Theorem 8.


MT(a) p MT(c) according to Theorem 2. V. 12.
PT(a).' PT(c) (Proof for It =- 1). We use the notation in Section 1.IX.8,
in which MP is the set of measures on D whose potentials are superharmonic.
The vector lattice M. = MP - MD is isomorphic to the vector lattice (Se, !)
of classical potential theory under the correspondence G0:t. The class
Mob of measures in MD vanishing on polar sets corresponds to the class
Sib according to Theorem 9; so the class Mo, of positive components of the
class orthogonal to Mob - that is, the class of measures in MD
supported by polar sets, corresponds to the class Sps, since S, = S. That is,
PT(a) = PT(c), and the proof of the theorem is now complete.

Parabolic Context

We shall need a further condition:


PT(c') p is supported by a semipolar set.
Then PT(c') p PT(c) in the classical context because in that context semi-
polar sets are polar. We leave to the reader the easy verification, modifying
the discussion in the classical potential theory context, that in the parabolic
context PT(a) p PT(b) PT(c').

11. Lattice Theoretic Analysis of the Composition of an


h-Superharmonic Function with an h-Brownian Motion
In this section we use the notation of Section 2.X.8 but restrict the context
slightly by assuming that h is harmonic. Thus u is a positive h-superharmonic
function on D to which the lattice theoretic discussion of this section is
applicable. The restriction to the parameter set R+ of the process
defined in 2.X(8.1) will be denoted by 040. If u() < +oc, the latter
process is an almost surely right continuous supermartingale to which the
716 3.1. Lattices in Classical Potential Theory and Martingale Theory

lattice theoretic discussion is applicable. We use the notation S, Sqb, etc., in


both potential theory and martingale theory contexts.
(a) If u e S; qb for every point of Jinlteness of u. This
assertion follows trivially from Theorem 2.X.8.
(b) If is a martingale for some (equivalently every) , then uES,,,gb
In fact the martingale equality at times 0, + oo yields equality at S in Theorem
2.X.8(d), and therefore there is equality on D. Observe that according to (d)
below the present assertion (b) becomes false if is replaced in the
hypothesis by 040.
(c) If u e SP +, then eSD for every point 5 of fniteness of u. To see
this recall that by 2.X(1.2) if we write iv,(-) for a Brownian motion in D
from and denote the process lifetime by SS, then

E{cx,h(t)} = i t}. (11.1)


IIS4>11 h(5)

Now by hypothesis uh is a superharmonic potential; so (Section 2.X.1) the


probability on the right has limit 0 when t oo, and (c) follows. Observe
that E SD implies that x',(+ oo) = 0 almost surely, but that the converse
implication is false, although the latter condition is necessary and sufficient
that 40 be in So S. The conclusion of (c) would thus be considerably weak-
ened if were replaced by in the conclusion. This weakening is
exhibited explicitly in (d) to be proved next.
(d) IJ u E then E S, ,for every S. [Recall that under the hypoth-
esis on u. P{x4(+oo) = 0} = I for all .] To prove (d), it is sufficient to
prove that if is a bounded supermartingale in S` with parameter set 68+
majorized in the specific order by then is a zero process up to a
standard modification, equivalently, E{y(0) } = 0. Now if D is an increasing
sequence of open relatively compact subsets of D with union D, containing
S. and if SS, is the hitting time of aD, by then the process o :() stopped
at time S4 is a martingale. the process stopped at the same time is a
supermartingale, and in view of the order relation there is
another process which added to yields 0x4() and which when stopped at
SS is a positive supermartingale. It follows that stopped at S", is a
martingale and therefore that

E{y(0)} = E{y(S )} E{ r {S -)} < E{ox{(S' -)} = E{xt(+oo)} = 0,


(11.2)
as was to be proved.

12. A Decomposition of SS (Potential Theory Context)


As in the previous sections h is a strictly positive harmonic function on the
connected Greenian subset D of R'. Let S ;. be the cone of functions u in
SL for which GD(t, , u) = u for all t > 0, that is, the cone of Gn-invariant
13. Continuation of Section 11 717

excessive positive singular h-harmonic functions on D, and let S , f be the


cone of functions u in S.*. for which lim,_m Go(t, , u) = 0. It is easily checked
that if u is in one of these classes, its positive specific order minorants are
in the same class and that a countable sum of functions in one of these
classes is also in that class if the sum is in S+. The classes S,,,. = S;,. -
S ;, , and S-f = S+ f - S, , f are therefore orthogonal subbands of If u
is in S,;,, we have seen in Section 2.IX.8, Example (a) (for h =_ I and the
general case follows trivially), that 0* (t, -, u) = uo is an h-harmonic
minorant of u, in S, . Hence lim,_m GD(t, -, u - uo) = 0, that is, u - uo e
S,',f, and therefore S., = S,,,,,, + S,,,,f.
The classes S,,,,... and S,,,,f have simple probabilistic characterizations. The
condition that u e is the condition that u is in S,R, and that the process
ox{(-) in Section I 1 is a martingale for some (equivalently every) , equiv-
alently and more intuitively, that the process defined in 2.X(8. l) is a
martingale for some (equivalently every) . A second probabilistic inter-
pretation is that (Section 2.X.1) uES;,,m [ueS, ,f] if and only if the lifetime
of uh-Brownian motion in D from some (equivalently every) 4 is almost
surely + oc [almost surely finite]. This characterization inspires the notation.

13. Continuation of Section I 1


Section 11(d) is made more precise by (e) and (f) below.
(e) If u e S; ,. , then e S ;,. In fact, as already noted in Section 12,
the process is a martingale, and this process is singular according to
Section 11(d).
(f) If u e S ;, f, then o a S', because the equation

lim Lo(t, , u) = lim E{of(t) } = 0

shows that a Sp , and this process is singular according to Section 11(d).


Section I I (c) is made more precise by (g) and (h) below.
(g) If ueSpyb, then In view of Section 1 I(c) this assertion
is trivial.
(h) I f u e Sp then ox a S;, for every for which oo. If
u(O < + oo, then by Section l 1(c) the process is in SP +. To prove that
the process is singular, suppose first that the measure associated with the
superharmonic function uh has compact polar support A. (According to
Section 10, this measure is supported by a polar set.) Define Do = D - A,
and let uo be the restriction of u to Do. Since almost no w20 path from
the point meets A, the process for uo on Do analogous to x'(-) for u
on D can be taken as this same process The fact that may not be
in Do will not affect the following reasoning. The function uo on Do is
singular as well as h-harmonic because a bounded positive h-harmonic
minorant u, of uo has an h-harmonic extension to D by a trivial generaliza-
718 3.1. Lattices in Classical Potential Theory and Martingale Theory

tion of Theorem 1.V.5, and the extension of u, is a bounded li-harmonic


minorant of a and therefore vanishes identically. Thus 0.vh()ESs by (e); so
as stated. If. however, p does not have compact support. the
measure is the sum of a sequence of measures with disjoint compact polar
supports, so as was to be proved.

EXAMPLE. Let N = 2. let D be the upper half-plane, and let h = I. The


ordinate function u on D is a singular positive harmonic function, singular
because any bounded harmonic minor-ant of u has limit 0 at every finite
boundary point of D and therefore (Theorem L V.7) vanishes identically. (It
is easy to prove that u is minimal, but this fact will not be needed.) The
positive u-harmonic function 1/u has a finite limit along almost every
u-Brownian path from a point S to oD so almost every such path tends to
oo. Direct calculation shows that u is subsumed under (e) so that if is
a Brownian motion in D from with lifetime S,, the process with
h = I is a singular martingale. The martingale property amounts to the
statement that a one-dimensional Brownian motion from a point a > 0
killed when it hits the origin is a martingale, a fact easily verified by direct
computation.
Chapter 11

Brownian Motion and the PWB Method

1. Context of the Problem


Let D be an open nonempty subset of R', coupled with a boundary 8D
provided by a metric compactification. To avoid trivial complications, we
shall assume that D is connected; if D is disconnected, the results are applic-
able to each open connected component of D. Let h be a strictly positive
harmonic function on D. The PWB method of attacking the first boundary
value (Dirichlet) problem for h-harmonic functions on D was detailed in
Chapter VIII of Part 1. Recall that the a algebra of µ4 measurable boundary
subsets is the a algebra of boundary subsets A for which the boundary in-
dicator function 1, is h-resolutive and that HA. The class of
Borel boundary subsets for which 1, is h-resolutive is a a algebra, and for
each point of D the restriction of µ4(c , ) to this a algebra, on completion.
is the measure µ"D( , ) on the a algebra of µhn measurable sets. The class of
µ4 measurable boundary functions f which are µ4c, ) integrable does not
depend on and is the class of h-resolutive boundary functions, and Hf =
This class of boundary functions will also be described as the class
of µ4 integrable boundary functions.
The PWB method solves the Dirichlet problem when the boundary is
h-regular and the given boundary function is finite valued and continuous,
but no relation other than the h-resolutivity itself was given in Chapter VIII
of Part I to connect an h-resolutive boundary function f on a not necessarily
h-resolutive boundary with the solution Hf . In Section 2 it will be shown that
in the general case h-harmonic measure has a probabilistic evaluation and
that Hf has f as a boundary limit function along h-Brownian paths. The
solution Hf, which (by Lemma 1.4) is necessarily in the class D(µ4_). when
composed appropriately with h-Brownian motion defines a uniformly inte-
grable martingale, and the evaluation Hf = becomes the martingale
equality for the parameter values 0, +x.
Throughout the discussion in Section 2 1, w -Ff (-)) is an h-Brownian
motion in D from , with lifetime Si'. With no loss of generality (Section
2.X.2) it can be assumed that .y is right continuous and that F,'(0) con-
tains the null sets. Since h is harmonic, almost every path tends to 8D
at the path lifetime (Section 2.X.4). Let I (u)) be the cluster set of ww w) at
720 3.11. Brownian Motion and the PWB Method

the parameter value S,. For almost every to this cluster set is a compact
boundary subset. In particular, if h = I and if aD is the Euclidean boundary,
the process iv,'(-) can be identified with a Brownian motion in R' killed
at the hitting time of OD, and r!(())) is then almost surely a singleton,
{ :'[S4 ((o) - ] , . Similarly, if D is bounded, if aD is the Euclidean boundary,
and if h has a harmonic extension h' to an open neighborhood D' of D, then
can he identified with an h'-Brownian motion in D' killed at the hitting
time of (3D. and again r"((,)) is almost surely a singleton, {w2 [S, (o))-];.
According to the following theorem, whatever the choice of (D, the dis-
tribution of 1 is the h-harmonic measure an h-resolutive function
/ is for almost every (j) a constant function on the cluster set q(w), and Hf
has the prescribed value of./ as limit along almost every path to the
boundary. Furthermore aD is h-resolutive if and only if r,' is almost surely a
singleton, that is, if and only if the left limit r, (SS -) exists almost surely.
Thus the PWB method generalizes the classical concept of solution of the
Dirichlet problem but nevertheless leads to solutions which have the pre-
scribed boundary limit function in a reasonable sense.

The Parabolic Context


The methods used in this chapter are applicable in the parabolic context, and
the statements of the results in that context are left to the reader.

2. Probabilistic Analysis of the PWB Method


Theorem. (a) For every boundary subset A

HA(5)="Ri(S)= R (5) (=P{T," nA 0 }if A is analytic) (2.1)


)

and

P{T4 c A; if aD - A is analytic. (2.2)

(b) 1/ A is pD measurable. then

A) = P;I", c A} = n A 96 Qf ). (2.3)

U j is a pll measurable boundary function then, for every S in D, j is identically


constant on almost every cluster .set V. Conversely ij f is a Borel measurable
boundary Junction and U. for some in D, f is identically constant on almost
every cluster .set r4", then J'is µv measurable.
(c) A boundaryfunction.f is h resolutive and only if/ is µD measurable
and integrable, and in that case
2. Probabilistic Analysis of the PWB Method 721

Hf () = µD( ,f) = E{f(ry) }, (2.4)

lim Hf[wh(t)] = f(r2) a.s.,


. (2.5)
S141

and the process {x2 f defined by

Hf [w" (t)] if t < Sf


x ti (t) (2 . 6)
f(ry) if SS < t < +oU

is a un(ormly integrable almost surely continuous martingale.


(d) The boundary is h-resolutive if and only if, for each point in D
(equivalently, for a single point l; in D), r{ is almost surely a singleton, that is,
if and only if the left limit w4 (Sh_ ) almost surely exists in the D u aD metric.
The distribution of ww (Sf -) is then p'

Observation. The equality Hf () = E{ f(ry)} is the martingale equality


for at the pair of parameter values 0, + co.

'
Proof of (a). The equality H =''R; = RAA /h was pointed out in Section
1.VIII.2. The last equality in (2.1) is a special case of Theorem 2.X.7. An
application of (2.1) to OD - A yields (2.2).

Proof'of (b). If A is pD measurable, that is, if 1e is an h-resolutive boundary


function, then if A is a Borel set, the equality of the terms in (2.1) with those
in (2.2) yields (2.3). Since every µD null set is a subset of a Borel uD null set
and since the a algebra of 1tD measurable sets is the completion of the restric-
tion of this a algebra to the class of its Borel sets, it follows first that (2.3) is
trivial when A is a µD null set and next that (2.3) is true when A is 1c4 mea-
surable. Conversely, suppose that for some point and Borel boundary
subset A the last equation in (2.3) is true. Then the positive harmonic function
Ht -H'', vanishes at and therefore vanishes identically; that is, the set A
is i4 measurable. Thus the assertions in (b) for a boundary function f are
true when f is the indicator function of a set, and are therefore true without
this restriction.

Proof of (c). The h-resolutivity of a boundary function f implies the µD


measurability off and therefore implies that for each point of D the func-
tion f is identically constant on almost every cluster set rs ; that is, J (r,) is
almost surely uniquely defined. The first equality in (2.4) was obtained in
Section 1.VIII.8, and the last equality follows from the probabilistic evalua-
tion of uD in (2.3). The following more direct proof of (2.4) also proves (2.5).
If u2 [ut] is in the upper [lower] PWB' class for the h-resolutive boundary
function.f, the h-superharmonic functions u2 + c, c - ut, u2 - Hf are posi-
tive if the constant c is sufficiently large. Hence (Theorem 2.10.8) each
722 3.11. Brownian Motion and the PWB Method

function u2, u,, Hf has a finite limit along almost every it`(-) path to the
boundary, and

u, () 5 E{ lim u, [w (t)] } < E{ lim Hf [w (t)] } < E{ lim u2[W (t)] } u2( ).
(2.7)
Moreover

lim u, [w (t)] < inf f(g) < sup f(C) 5 lira u2 [w (t)] as. (2.8)
fsE CEr Erg
S*41

Since u, and u2 can be chosen to make u, arbitrarily small, f must


be constant on almost every cluster set I'2, and (2.4) and (2.5) are true.
The martingale assertion in (c) follows from Theorem 2.X.8 and Sec-
tions 1.4 and 1.5 but will be clearer from the following direct proof. Denote
the 40 process in (2.6) by x2(f; ). All supermartingales and martingales
below are relative to the filtration (). If f is an h-resolutive boundary
function, Theorem 2.X.8 implies that the following processes are almost
surely continuous positive supermartingales:

x(IfI A n);')
(2.9)

The fact that the last two of these processes are supermartingales implies that
x( I f l n n: ) is a martingale and therefore (n -* oo) that x{(If 1, ) is a mar-
tingale, that is, that (c) is true when f is positive. Since the first two processes
in (2.9) are now known to be almost surely continuous martingales, the
process x(./; ) is also, as was to be proved.

Proof of (d). If the boundary is h-resolutive, that is, if the Borel boundary
subsets are µD measurable, chooses > 0, and let A, , ... , A. be compact
boundary subsets of diameter at most r and with union D. Since (2.3)
implies that r,' is almost surely a subset of any A, it intersects, the diameter of
rh is almost surely at most E. Hence r,' is almost surely a singleton; that is,
the left limit tt>2(SS -) almost surely exists. Conversely, if r is almost surely
a singleton for some point , the set function A F- is an additive
function of Borel boundary sets by (2.1); so the boundary is h-resolutive
(Section 1.VIII.9).
Special Case: It Is Minimal. If h is minimal, then for every boundary subset
A the h-harmonic function ff' = Re/h is either identically 1 or identically
0. In fact by minimality H is a constant function, so that RA = ch, and
iterating the reduction operation [Section I.VI.3(h)] yields ch = cRA = c2h;
so c = 0 or 1. For every boundary point C if R;; ' - 0, then (since R,; 1 is the
infimum of the class of reductions on boundary neighborhoods of C) RA = 0
3. PWB" Examples 723

for A a sufficiently small boundary neighborhood of C. The set B of boundary


points { with Rit} = 0 is therefore open, and Rf,, = 0. Thus the set B' =
OD - B is a compact set with the property that IF,' = B' almost surely, for
each . The class of µ"n measurable boundary subsets consists of all boundary
subsets which are either supersets of if or subsets of B. A boundary function
f is h-resolutive if and only if it is constant (finite) on B' and the PWBh
solutions are the constant functions. In particular, a boundary is h-resolutive
if and only if if is a singleton {C} which supports every h-harmonic measure
14(x, ), equivalently, if and only if for all in D almost every path
tends to C at the path lifetime. See Section 2.X.9 for this situation when D
is a ball.

The Role of Capacity Theory

A close examination of the analysis reveals that the theory of analytic sets
and Choquet capacities is not needed in the derivation of Theorem 2 if it is
supposed that the left limit wf (Sf -) exists almost surely for some (equiva-
lently every) point of D or, going in the other direction, if the boundary is
h-resolutive. This hypothesis is satisfied if aD is the Euclidean boundary and
if either h - I or D is a relatively compact open subset of an open set on
which h has a positive harmonic extension.

3. PWB' Examples
EXAMPLE (a). The Alexandrov one-point boundary of a Greenian set D is
trivially universally resolutive. From the point of view of Theorem 2 universal
resolutivity corresponds to the fact that for strictly positive harmonic h on
D almost every h-Brownian path from a point of D tends to the one-point
boundary at the path lifetime.

EXAMPLE (b). If D is a ball, its Euclidean boundary is universally resolutive


according to Section 1.VIII.9; this universal resolutivity corresponds to the
fact deduced in Section 2.X.9 that for strictly positive harmonic h on a ball
D almost every h-Brownian path from a point,of D tends to a point of the
Euclidean boundary at the path lifetime.

EXAMPLE (c). Let N = 2, let D be a disk, let { be a boundary point, and let
h = K(C, ) be the minimal harmonic function corresponding to C (Section
1.II.1). For in D let be the angle between the rays from { to and to
the disk center, - n/2 < 0 < n/2. Then 0 is harmonic on D. Define a
distance function on D x D by

d(4,7)=I4-7I+I4(4)-* (7)I. (3.1)


724 3.11. Brownian Motion and the PWB Method

If D is provided with a boundary a'D by completion under this distance


function, the Euclidean boundary point S is ramified into a compact set A
of points corresponding to the directions of approach to S. In this example
h is minimal: so the quasi-bounded h-harmonic functions, in particular the
PWB' solutions, are the constant functions, and the Euclidean boundary
singleton ; y; has h-harmonic measure identically 1. Almost every h-Brownian
path from a point c of D has limit r in the Euclidean metric. We now inves-
tigate the cluster sets of these paths on a'D, that is. the boundary cluster sets
in the d metric defined by (3.1). Recall from Section 1.XII.12 that if ( is
identified with a minimal Martin boundary point of D, no ray from r into D
is minimal thin at r. We shall show in Section 111.3 that a Borel subset B of a
Greenian subset of R'5 is not minimal thin at a minimal Martin boundary
point if and only if almost every conditional Brownian path from a point of
the Greenian set to that boundary point hits B arbitrarily near the boundary
point. In the present context, using the notation relative to r7'D. it follows
that r is almost surely the set A. According to Theorem 2, a function f on
)'D is h-resolutive if and only if f is identically constant (# ± x) on A, and
if so, the PWB' solution for f is identically the constant,f(A). The ramified
boundary a'D has too many points to be h-resolutive.

EXAMPLE (d). Let B be an open subset of a Greenian set D, let D be provided


with a boundary OD by a metric compactification, and let aB be the boundary
of B relative to this compactification. Then aB depends on the choice of
aD unless B is relatively compact in D. According to Theorem 2. aB is h-
resolutive if and only if almost every h-Brownian path in D from a point of B
either hits D r, aB or tends to a point of aD n aB at the path lifetime. Thus it
is sufficient for h-resolutivity of aB that OD be h-resolutive. This simple
resolutivity argument should be compared with the nonprobabilistic argu-
ment in Section 1.VIll.8, Example (b). The relation 1.VIII(8.3) connecting
Nn and pe is an immediate consequence of the strong Markov property of
conditional Brownian motion.

EXAMPLE (e). Suppose that the metric compactification D = D u OD of a


Greenian set D has the following properties. There is a family ; q5;, i e 1; on
D, indexed by some set 1. for which
(e,) Each function 4,; has a continuous extension (also denoted by q,)
to D.
(e2) If q e aD, a sequence q, in D has limit q if and only if
O;(q) for all is that is, the family 0, separates aD.
(e3) Under these conditions if also for some strictly positive harmonic
function h on D lim,1s¢;[x{(t)] exists almost surely simultaneously
for all i whenever uh(') is an h-Brownian motion from ( with lifetime
S'', it follows that the left limit ich(Sh-) exists almost surely: so aD
is h-resolutive. Since one standard technique of defining a boundary
4. Tail a Algebras in the PWB` Context 725

for a set D is to choose a compactification of D which makes each


function of a specified family of functions on D have a continuous
extension to the compactification, this example is frequently en-
countered. For example, if D is a Greenian subset of 18" and if o is
a point of D, the Martin compactification of D (Chapter XII of
Part 1) can be defined as the space satisfying (ei) and (e2) for the
family of functions {04, eD} defined by 0S = c0)-
Moreover, if convenient, we can restrict to a countable dense
subset of D - A direct argument in Section 111.5 will show
that (e3) is satisfied for every h and therefore that the Martin bound-
ary is universally resolutive, but in fact this universal resolutivity
was derived in Section 1.XII.10 from which it follows (Theorem 2)
that almost every h-Brownian path from a point converges to a
Martin boundary point and that the distribution of the limit
we (Sc -) is

a Algebras in the PWB' Context


Let D be a connected Greenian subset of R, coupled with a boundary OD
provided by a metric compactification, let h be a strictly positive harmonic
function on D, and let w{ be an h-Brownian motion in D from , with life-
time S'. Let I' be the ww process tail o algebra, and let r, (m) be the
4(-, (0) cluster set on aD. Recall that by definition the o algebra I{ contains
the 40 null sets. Throughout this section and ww are fixed.

Theorem. If OD is internally h-resolutive, IN{ consists [modulo ww null sets]


of the class of sets of the form

{o>: r4'(o)) e A } for Borel po measurable boundary subsets A. (4.1)

In particular, if OD is also h-resolutive, 5,1 consists [modulo null sets] of


the sets of the form

{ w : K S (SS - , o)) E A) for Bore! boundary subsets A. (4.2)

Observation. Since a µD measurable boundary subset differs from some


Borel and pD measurable boundary subset by a µD null set, and since the h-
harmonic measure pD is determined by the distribution of r'2 through (2.3),
"Borel" can be omitted in (4.1) and replaced by "µD measurable" in (4.2).
Theorem 4 implies that if OD is internally h-resolutive, then a w; process
random variable z is Vh measurable if and only if there is a uD and Borel
measurable function g on OD such that z = g(r) almost surely, and if aD is
also h-resolutive, r 4 can be replaced here by w' (S4 - ).
726 3.11. Brownian Motion and the PWB Method

Special Case: h Is Minimal. If h is minimal, every bounded h-harmonic


function is identically constant; so every choice of boundary for D is inter-
nally h-resolutive. The a algebra'., and the a algebra of µD measurable sets
are both trivial in this case; that is, their sets are all of measure 0 or 1. If A
is a Borel boundary subset of h-harmonic measure I. the a algebra of pt
measurable boundary subsets is generated by A and the pt null sets. If
A e Sp'.' and if A is not r" null, then T' is the a algebra generated by A and
the null sets. In particular, if OD is also h-resolutive, there is a point S
of aD such that 1. In this case all subsets of 8D are µD mea-
surable.

Proof o/'Theorem 4. If A is µD measurable, that is, if the boundary function


J = 1,, is h-resolutive, then (by Theorem 2)

lim Hf [w (r)] =l(r) a.s.,


its
and it follows (Section 2.X.1 I) that f(rs) is ¶ measurable, that is,
{w: q(w) c A}esh. Conversely, if Ac-14', it was shown in Section 2.X.1 l
that there is a bounded h-harmonic function u such that 1im,tsh 1

almost surely. Since the boundary is internally h-resolutive, u = Hf for some


and Bore] measurable function f, and u has almost sure boundary limit
function f along tr{ O paths; that is, f(rs) = 1, almost surely. Hence there is
a pD and Borel measurable boundary subset A such that f = l,+ up to a µD
null set, and A c A} up to a null set. If 3D is also h-
resolutive, every Borel measurable boundary subset is pD measurable, and
a µD measurable boundary subset is a Borel set up to a pD null set ; so the last
assertion of the theorem is trivial.
Chapter III

Brownian Motion on the Martin Space

1. The Structure of Brownian Motion on the Martin Space


Let D be a connected Greenian subset of 08", let K be a Martin function for
D, let h be a strictly positive superharmonic function on D, and let
Y,'(-)) be an h-Brownian motion in D from with lifetime SS. For A a subset
of D let SS" and Lh", respectively, be the hitting and last hitting times of A
by According to Theorem I.XII.10, if h is harmonic, the Martin
boundary is h-resolutive and ph dC) = K(C, where Mh is the
Martin representing measure of h corresponding to K. According to Theorem
11.2, the left limit w!(SS-) exists almost surely and has distribution
supported (Section I.XII.7) by the minimal Martin boundary carD. In
particular, if C is a minimal Martin boundary point and if h = K(C, ), then
I ; so vt'$(SS-) = C almost surely. With this choice of h we shall

sometimes write S44", Le", respectively, for ws SS", L.


The analysis in Section 2.X.9 of h-Brownian motion in a ball compac-
tified by its Euclidean boundary is applicable to an arbitrary connected
Greenian set D compactified by its Martin boundary, with no change except
for allowance for the possible presence of nonminimal Martin boundary
points. That is, if h is harmonic, an h-Brownian path in D from S is obtained
by first choosing the minimal asymptotic path endpoint C on ??,MD according
to the distribution pD(, ) and then choosing a K(C, )-Brownian motion
path from to C. More formally, one version of the conditional distribution
of for ws(SS -) = C is the distribution of K(C, )-Brownian motion from
. If h = GDp is a potential and if h(g) < + x, then (from Section 2.X.4)
almost surely SS is finite, and the left limit rs(SS-) exists and is in D;
wf (SS -) has distribution GD(C, O)p(dC)/h(C). With this choice of h an h-
Brownian motion path can be obtained by first choosing the asymptotic path
endpoint C in D according to the distribution and then
choosing a C)-Brownian motion path from to C; that is, one version
of the conditional distribution of for r{(SS-) _ C is the distribution
of C)-Brownian motion from . The case h(C) _ + oo is treated in
Section 2.X.5. If h = h, + h2 is an arbitrary strictly positive superharmonic
function on D, with h, positive harmonic and h2 a potential (Riesz decom-
position), h-Brownian motion from , when +x. is h,-Brownian
728 3.111. Brownian Motion on the Martin Space

motion from S with probability and is h2-Brownian motion from


with probability

Attainable and Unattainable Martin Boundary Points

According to Section 2.X.1, if C is a minimal Martin boundary point, the


function c - P { SS = + x) is either identically 0 or identically 1. The point
s is called attainable in the first case and unattainable in the second. For
example, if D is a ball (Section 2.X.9), its Martin (i.e., Euclidean) boundary
points are all minimal and attainable. If D is a half-space its Martin (i.e.,
Euclidean) boundary points are all minimal, and the finite boundary points
are attainable, but the infinite boundary point is not attainable. If D = RN
with N > 2, the minimal harmonic functions are the positive constant func-
tions, and the point ao is the only Martin boundary point, is minimal, and
is unattainable.

2. Brownian Motions from Martin Boundary Points


(Notation of Section 1)
Brownian Motions from Attainable Boundary Sets

Let h be a strictly positive harmonic function on D whose Martin represent-


ing measure is supported by the set of attainable minimal Martin boundary
points. Then for each point g of D the h-harmonic measure µDW, dC) =
K(C, is also supported by this set. Let wx be an h-Brownian
motion in D with initial distribution A and lifetime Sx ; this lifetime is neces-
sarily almost surely finite. We define wx(Sx) as limas" wx(t), a limit which
according to Theorem 11.2 exists almost surely and has distribution
According to Section 2.X.6. the process {w,'(Sx - t), to R+}
with lifetime S, is a GDA-Brownian motion, and obviously A is the distribution
of path endpoints (t Sr).

Special Case: Brownian Motion from an Attainable Martin Boundary


Point to a Point of D

Let i be an attainable minimal Martin boundary point, and let be a point


of D. Then [particular case of the preceding paragraph with h = K(C, ) and
with A supported by on the one hand the process {w{(t), to R1 } is a
K(C, ) Brownian motion from , and almost every path tends to s at the path
lifetime SS. whereas on the other hand the process {ws(SS - t),teR } is a
motion from C, and almost every path tends to at the
path lifetime S. From now on it will be convenient to write wz (t) instead of
t) and denote by t the motion transition density.
2. Brownian Motions from Martin Boundary Points (Notation of Section 1) 729

We now fix and S and discuss the absolute probability distributions of the
N4 process: it, A) F--. $w(i) e A} = p(i, A) for A ranging through the Borcl
subsets of D - Since

f1x(di) J'(r - s, q', OAS, dq') = p(t, A). 0 < s < t, (2.1)
A D

it follows that p(t, ) is absolutely continuous relative to /N and that one


choice of the Radon-Nikodym derivative dp(t, )/dlN at q is

PD(1,C,1)= 1,d(t-s,7',q)P(s,dq'). (2.2)

Observe that the value of the integral on the right does not depend on the
choice of s for 0 < s < t because t satisfies the Chapman-Kolmogorov
equation; so (2.2) defines C. ) on ] 0, + co[ x (D - {s}). The function
p,(t, , ) is the IN density of the distribution of is (t) in D, and the function
p, is that is, ¢ = C, is parabolic, on
]0, + oc [ x (D - { }) in view of the following properties of 0 on this
product set:
4) is Borel measurable.
4) is locally IA+t integrable.
¢ has the parabolic function average property locally, because fD
(- s, q, ) is parabolic on the domain of 0.
It is natural to conjecture that when of - C suitably, the transition density
t (t, q', q) tends to p j(t, S, q), and we now derive one version of such a limit
relation : for each strictly positive t,

lim4,4(t - s, tsS(s), q) =Pi(t, S, q). (2.3)

To see this, observe that the function (t l', t') ' -, q) is GD(5, )-parabolic
on (D - x ] 0, + oo[ and, in fact, is invariant excessive for
space-time Brownian motion on this set. Moreover (2.2) implies that

E{GD(t - s, wti (s), q) } = PL(t, 0 < s < t. (2.4)

Hence the process 14f (t - s, w e(s), q), 0 < s < t} is a martingale. This mar-
tingale is almost surely continuous and so (from Section 2.111.16) has an
almost sure limit as the parameter tends to 0. This limit is a random variable
measurable with respect to the K(C, )-Brownian motion tail a algebra and
so (from Section II.4) is almost surely identically constant. Since every mar-
tingale is uniformly integrable on a right closed set of parameter values
[Section 2.111.3(e)], this constant must be the common expectation in (2.4);
that is, (2.3) is true.
730 3.111. Brownian Motion on the Martin Space

Brownian Motion from an Unattainable Minimal Martin Boundary Point


to a Point of D

If ( is an unattainable minimal Martin boundary point of D, that is (Section


2.X.1), if the function K((, ) is an invariant excessive function for Brownian
motion in D, let be a point of D, and let D. be an increasing sequence of
open relatively compact subsets of D, containing , with union D. Let L{ be
the last hitting time of D. by The function OK((, ) 0°`' (reduction relative
to D) is the potential of a measure supported by 8D,,, OK(C, ) Dn =
According to Section 2.X.10, the distribution of ,(Lc,) is

GD(S, ,1).un(d7)/K(S, L),

and the process {wl'(LC4. - t), to 6F+) with lifetime L4,. is a


motion with initial distribution the distribution of Observe that
L; = + oc almost surely and that lim,,. almost surely.

EXAMPLE. Denote by d{ the Nth coordinate of the point of R', and let D
be the half-space {dd > 0}. The Green function GD was evaluated in Section
1.VIII.9, and in Section 1.XII.3 the Martin boundary was found to be the
Euclidean boundary. All boundary points are minimal, and the finite bound-
ary points are attainable. By direct calculation, if ( is a finite boundary
point and if a2 = 1,

lim t (t - s, rl', 7) = lim


G(t - s, 7',, 7)
GD(t, 7')

d°I
dr2ntC12G D ( , if N = 2 ,

d,,I(-(IN
dd(N - 2) (2n)"r2 t(x+2Y2 ' 7) if N > 2.

According to the results in this section, this limit is p4(t, (,, ). If now tends
to a second finite boundary point C,, the transition density b, tends to a
limit transition density, that of K((,, )-Brownian motion, and p,(t, C, )
tends to a limit density which can be described as the absolute probability
density at time t of a Brownian motion from ( to (, .

3. The Zero-One Law at a Minimal Martin Boundary Point


and the Probabilistic Formulation of the Minimal-Fine
Topology (Notation of Section 1)
If C is a minimal Martin boundary point, almost every path tends to
at the path lifetime, and (Section 2.X.11) the tail a algebra of this process is
trivial. As detailed in Section 2.X. I I for h-Brownian motion with h a minimal
3. Probabilistic Formulation of the Minimal-Fine Topology 731

harmonic function, the asymptotic properties 2.VII.6(b)-(g) relating to set


and function properties in the neighborhood of the initial point of a Brownian
motion correspond to set and function properties near the lifetime of the
h-Brownian motion, that is, in the present context, near C. For example,
Section 2.VII.6(b) corresponds to (b*): if A is an analytic subset of D, the
function -. P{ Lf" = S4) is either identically 0 or identically 1.
The probabilistic formulation of the fine topology was given in Section
2.IX.15 [see also Section 2.X.2(d)]. The following theorem describes in
probabilistic language the minimal-fine topology at a minimal Martin
boundary point, defined nonprobabilistically in Section 1.XII.12.

Theorem. Let C be a minimal Martin boundary point of D. For C to be a minimal-


fine limit point of the subset A of D it is necessary that for every point C of D,

P{L;R=SC} = 1 (3.1)

whenever B is an analytic superset of A, and it is sufficient that (3.1) be true for


a single point C of D whenever B is an open superset of A.

Thus if A is analytic, the point C is a minimal-fine limit point of A if and


only if (3.1) is true when B = A. A subset A of D for which D - A is analytic,
for example, a Borel subset A of D, is a deleted minimal-fine neighborhood
of C if and only if almost every path lies in A during some parameter
interval ] t, S [ with t = t(w) < Si ; the value oft can be taken as the last
hitting time of D - A. In both these statements it is necessary that the con-
dition be satisfied for each point C of D and sufficient that the condition be
satisfied for a single point C of D.

Proof of Theorem 3. An equivalent version of (3.1) is

P{lim sup h[x (t)] = 1) = 1. (3.1')


ttsi

According to Section 2.X.11, this condition is satisfied when A is analytic


either for every point of D or for no point C of D. If B is the trace on D
of a neighborhood of then (Section 2.X.7) when A is analytic,

K(C, C)

Since the left side is I for all B if and only if C is a minimal-fine limit point
of A and the right side is 1 for all B if and only if (3.1) is satisfied, the theorem
is true when A is analytic. Since C is a minimal-fine limit point of a subset A
of D if and only if C is a minimal-fine limit point of every open superset of A
(Section I.XII.12), the theorem is true for every set A.
732 3.111. Brownian Motion on the Martin Space

Application to Function Limits

According to Theorem 3 and the discussion at the beginning of this section,


a Borel measurable function from D into a Polish space D' has minimal-fine
limit [minimal-fine cluster value] tl' at the minimal Martin boundary point
C if and only if lim,tsc u[w4(1)] = s has cluster value tl' at 4 along
wf(-) paths] almost surely, for each point 4 of D, equivalently, for some point
4 of D. In particular, if D' = R,

limsupu[ww(t)] = mflimsupu(q) a.s.


ITSIC

for each point 4 of D, and the corresponding equation for inferior limits is
also true.

4. The Probabilistic Fatou Theorem on the Martin Space


Let h be a strictly positive harmonic function on a Greenian subset D of
R", let v be a positive superharmonic function on D, and define u = v/h.
According to Theorem 2.X.8,

u4 = (4.1)
,ts
exists almost surely. In view of the structure of h-Brownian motion discussed
in Section 1, the existence of this almost sure limit means that for pD almost
every Martin boundary point 4 (which we can suppose minimal) the function
u has a limit along almost every K(4, )-Brownian path to 4 from g. According
to Section 2.X.11(c), this limit is for each C almost surely a number f(4) which
does not depend on 4. In view of Section 3 we have proved that

mf lim f(4)
rC

exists for pD almost every 4, a result we have already proved nonprobabil-


istically (see Theorem 1.XII.19). Observe that the boundary limit function
was identified in Theorem 1.XII.19 as dM/dM,,, where M,, is the
Martin representing measure corresponding to h [to the harmonic compo-
nent of v in its Riesz decomposition]. The following argument shows how
fcan beidentified in the present context, without invoking Theorem 1 .XI 1.19.
It is sufficient to identify f for a an h-potential, u a singular h-harmonic func-
tion, and u a quasi-bounded h-harmonic function since (Section 1.IX.11) u
is the sum of its components of these types. According to Theorem 2.X.8(c)
as now interpreted, the limit function f vanishes pD almost surely if u is an
h-potential or is a singular h-harmonic function. According to Theorem
5. Probabilistic Approach to Theorem I.XI.4(c) and Its Boundary Counterparts 733

2.X.8(d), E{ur} = u°,(g, f) when u is quasi bounded and h-harmonic,


in view of the expression we have found in Theorem 11.2 for h-harmonic
measure, and f here is according to Theorem 1.XII.10. Thus we
have derived Theorem I.XII.19 probabilistically.

5. Probabilistic Approach to Theorem 1.XI.4(c) and Its


Boundary Counterparts
Throughout this section v is a positive superharmonic function on a con-
nected Greenian subset D of 1 N with N > 1. We have already (Sections
1.XII.13 and 1.XI1.14) proved the Martin boundary counterparts of
Theorem I .XI.4(c) which states that if C is a point of D, the function C)
has a fine limit at C. The first boundary counterpart (Theorem 1.XII.13) of
Theorem 1.XI.4(c) asserts that if K is a Martin function for D and if C is a
minimal Martin boundary point of D, then the function v/K(i;, ) has a
minimal-fine limit at C. In view of the role of a point s of D as a minimal
Martin boundary point of D - {() [Section I.XII.12, Example (a)] Theorem
1.XII.13 includes Theorem I.XI.4(c). As noted in Section I.XII.19 the
boundary limit theorem (Theorem 1.XII.13) is a special case of the Fatou
boundary limit theorem for the Martin space of D; so in view of the discus-
sion in Section 4 there is no need to discuss this case probabilistically here
except to remark that almost every K(C, )-Brownian path from a point of D
to the Martin boundary tends to C, and so the probabilistic Fatou boundary
limit theorem becomes localized at C in this case. The second Martin bound-
ary counterpart (Theorem 1.XII.14) of Theorem I.XI.4(c) asserts that if
v 0 0, the function v/GD(5, ) has a strictly positive perhaps infinite minimal-
fine limit at each minimal Martin boundary point C, and the probabilistic
approach to this result will now be discussed.
It is sufficient to prove the existence and strict positivity of this minimal-
fine limit for v/GD(c, ) bounded, at the expense of replacing v if necessary by
vA for a positive constant c. In the following we therefore suppose
that 0 < c. Let h be a strictly positive harmonic function on D,
and let w (-) be an h-Brownian motion in D from with lifetime S,. The
function h will be chosen to be K(C, ) below, but it is instructive and does not
complicate the preliminary work to allow the stated generality of h. If Sh is
almost surely finite, that is, if the Martin representing measure of h is sup-
ported by the set of attainable minimal Martin boundary points, the process
{ tv4 (S4 - t), t e O8+ If with lifetime SS is a motion with
initial distribution (Section 2). Hence the restriction to D - of
the function v/GD(C, ) is excessive for this process; so the composition of
v/GD(5, ) with this process, on the parameter set ]0, + x[, defined as 0 for
parameter values z S,, is a positive bounded almost surely right continuous
supermartingale and as such has an almost sure limit x(0+), which we
denote by x(0). The process {x(t), to 68+} is a supermartingale, and E{x(0)}
734 3.111. Brownian Motion on the Martin Space

> 0 because v W 0. In particular, if h = K(l, ) for some attainable minimal


Martin boundary point C, the initial distribution of wv (SS - -) is
supported by "Cl. so has a limit at t along almost every
Brownian path from C to ; equivalently (Section 3), has a minimal-
fine limit at C, equal to E{x(0)} and therefore strictly positive, as was to be
proved. Observe that without the hypothesis that h = K(C, ) we obtain only
the weaker result that for almost every minimal Martin boundary
point b the function has a minimal-fine limit at C. In view of this
fact we shall assume that h = K(C, ) in discussing limits at unattainable
minimal Martin boundary points. Suppose then that S is such a point, and
let D. be an increasing sequence of open relatively compact subsets of D with
union D, suppose that 5 EDo, let LCn be the last hitting time of D. by w'(-), and
define the K(C, )-potential GDin/K(C, -) by
Dn
RKu
K(C,

Then (Theorem 2.X.7) GD2n/K(C, ) at , is the probability that a K((, -)-


Brownian path from q ever hits Dn. According to Section 2.X.10, the distri-
bution of w (LC,,) is q)dn(dq)/K((,,), and as discussed in Section 2, the
process {ww(LC, - 1), t E R') is a motion with lifetime
L and with initial distribution the distribution of The function
is excessive for motion on D - so the
process

- R+
{xn(t), tE R+ } = l- ( 4t1L(j.
C
Cn
)'
tE

(set equal to 0 at times z is a bounded almost surely right continuous


supermartingale. The supermartingale downcrossing inequality is applicable
to x and we find that the expected number of downcrossings of an interval
[r,,r2] by is at most c/(r2 - r,). That is, the expected number of up-
crossings, defined in the obvious way, of [r1, r2] by the process

( v[ws(t)]
1 O<t<{n
Lf
tt''(t))' }
is at most c/(r2 - r,). Since this is true for all n, the same assertion is true for
0 < t < Se, and we deduce just as in the discussion of supermartingale conver-
gence that the function has a limit at i along almost every 40 path;
equivalently (Section 3), the function has a minimal-fine limit at C.
Observe that E{xn(0)} > 0 because v 0 0, and observe also that L' 1 - LCn
is the hitting time of D. by so by the supermartingale inequality at
optional times E{xn+, (0) } Z E{x (0)}, and therefore limn..,E{xn(0) } > 0.
6. Martin Representation of Harmonic Functions in the Parabolic Context 735

Since this limit is the almost sure limit of at ( along paths as


well as the minimal-fine limit of the function at S, this minimal-fine limit
must be strictly positive (S + ao).

Parabolic Context

The parabolic context version ofTheorem 1.XI.4(c) is Theorem 1.XVIII.14(f)


together with its dual. Theorem 1.XIX.13 is, together with its dual, the
Martin boundary limit counterpart of the latter theorem. The probabilistic
approach to these Martin boundary limit theorems is left to the reader.

6. Martin Representation of Harmonic Functions in the


Parabolic Context
Let D be a Greenian subset of IAN (N Z 1), define I) = D x 98, f)+ = D x
]0, + oo [, and let h be a positive harmonic function on D. Since h is bD-
excessive, the function h) is a finite-valued monotone decreasing
function on ]0, + oo[. The function h) is parabolic on b+ because
q) is parabolic there for q in D; SO h) is obtained by an integral
operation on parabolic functions. The fact that h is /D-excessive implies that
limr40 /D(t, , h) = h(c) and by Dini's theorem this monotone convergence
is locally uniform on D. Thus if 40, , h) is defined on D as /D(t, , h) for
t > 0 and ash for t < 0, the function h) is continuous on D and in fact
is parabolic there because this function is already known to be parabolic on
b' and has the parabolic function average property on b - D+. In the fol-
lowing we write 6 (t, , h) for dtD(t, c, h)/dt. The function /D is thereby
defined, negative, and parabolic on D, identically 0 on D - D. This func-
tion vanishes identically if and only if h is /D invariant excessive.
(a) The function h) is 6D invariant excessive on D+ ; that is,

-f D
S, 1)/v(t, q, h)IN(dn) _ -/D(s + t, , h) (s > O, t > 0). (6.1)

In proving this it will be convenient to denote by (6.1)5 the relation (6.1)


with "=" replaced by "<-." Inequality (6.1)' is true because -lo(, ,h) is
positive and parabolic and so is /o-excessive. If both sides of (6.1)5 are
integrated with respect to I,(dt) over a compact subinterval of ]0, +oo[, we
obtain an equality in view of the Chapman-Kolmogorov equation satisfied
by /D. It follows that for fixed s) equation (6.1) is true for /t almost every
1. Furthermore the right side of (6.1)5 defines a parabolic function of s)
on D+, as does the left side, which is obtained by applying an integral
operator to parabolic functions. The difference between the right and left
sides thus defines a positive parabolic function on D+, necessarily vanishing
736 3.111. Brownian Motion on the Martin Space

at all points below a zero. Hence, if s' > 0, equation (6.1) is true for (5, s) E
D x ]0, s'] if t is not in some 1r null set. When s' runs through the strictly
positive integers, we find that there is an 1t null set A such that (6.1) is true
for (C, s) a D+ if and only if t is not in A. However, fort not in A and for
strictly positive s and 6,

-GD(S + s + t, C, h) _- J('D D
GD(b, C, )GD(s, , t1)GD(t, 1, h)1N(d)7)1N(dd)

_ - J GD(s.C,S)Gn(s + t, ,h)IN(dd)

(b+s+t,c,h).
Hence there is equality throughout; so .s + t is not in A, and it follows that
A is empty.
(b) Either h) = 0 (and this is true if and only if his an invariant
excessive harmonic function) or h) < 0 on D+ because the positive
parabolic function - h) vanishes identically below each zero and (6.1)
implies that this function vanishes identically above each 0 in 6.
(c) The function Go(-, , h) determines h uniquely up to an invariant
excessive harmonic function. Since [Section 2.IX.8, Example (a)] the
function

lim GD(t, , h) = ho (6.2)

is an invariant excessive harmonic function majorized by h, we obtain the


same class of functions t h) if we restrict h by supposing that ho = 0.
(d) We have proved that if h is a positive harmonic function on D, the
positive parabolic function u = - t h) on b has the following properties:
(dl)ti=0on D-D'.
(d2) ir(C, s)1 t (ds) < + oo.
0

(d3) f GD(s, C, r7)ti(q, t)1N(dil) = 4(C, s + t) (s > 0,1 > 0).


D

Conversely, we now prove that if ti is a positive parabolic function on D with


these three properties, then ti can be obtained in this way; that is, there is a
positive harmonic function h on D such that ti = h). In fact the
integral in (d2) defines a positive harmonic function h on D (Section 1.XV.15)
and

4D(t,5,h)= J u(C,s)1t(ds);
6 Martin Representation of Harmonic Functions in the Parabolic Context 737

so -n h) = u. Observe that this procedure yields a function h for which


ho in (6.2) vanishes identically.
(e) Probabilistic significance of -t h). If h is a strictly positive
harmonic function on D and if ho in (6.2) vanishes identically, then according
to Section 2.X. I the restriction of h)/h() to 68+ is the density
relative to 1, of the distribution of the lifetime of h-Brownian motion in D
from .
(f) The positive harmonic function h on D is minimal harmonic if and
only if the parabolic function -, is minimal parabolic on D. (It is
trivial that a positive parabolic function on b, vanishing identically off b,
is minimal parabolic on D if and only if it is minimal parabolic on b'.) In
fact, if h is minimal harmonic on D and if u, is a positive parabolic minorant
of h) on D, then ti, satisfies (d 1) and (d2). The function ti, also
satisfies (d3) because both u, and u, are positive parabolic
functions and so satisfy (6.1)5 and their sum satisfies (6.1), and so iu, must
also satisfy (6.1). Thus according to (d) there is a positive harmonic function
h, on D such that

u, = -(' ,h1), /t, = f.""

Hence

h, < Ja(.s)li(ds)=h;
n

so h, = ch, and therefore u, = - t h,) = - h) on b' and


therefore on D. Thus -n , , h) is minimal parabolic on D. Conversely, if
is minimal parabolic on D and if h, is a positive harmonic
minorant of h, then

h t) + fD(s. c, h - h t) = ",
h) = - J'0(z.h)l1(dt).

Differentiate to find that h,) < h); so there is a constant c


such that 6 h,) = h). Integration over 08+ yields h, = ch, and it
follows that h is minimal harmonic on D.
(g) If h is a minimal harmonic function on D but is not GD invariant
excessive, then ho in (6.2) vanishes identically. In this case define the parabolic
function h on b by setting and if aEQB, define
-f"D(s - a, , h) to obtain a set of minimal parabolic functions on D. More-
over (from Section I .XV.17), if b c- 118, the restriction of ha to D x ] - co, b [
is minimal parabolic on this capped cylinder. Aside from normalizations the
representation

h= f hal, (da)
738 3 111 Brownian Motion on the Martin Space

is the Martin representation of h on b in terms of minimal parabolic


functions. It is trivial that an ha space-time Brownian motion in b from
a point with s > a has almost sure lifetime s - a. As noted in (e)
the lifetime of h-Brownian motion in D from n has 1, distribution density
equivalently, the lifetime of h space-time Brownian
motion in 1) from s0) has this distribution density. This space-time
Brownian motion conditioned by fixing the lifetime to he .s - a becomes h,
space-time Brownian motion.

EXAMPLE. If N = I and if D = ]0, +aj[, there are two minimal harmonic


functions on D up to multiplicative constants: the function h =_ I and the
function "h(5) = s The first of these leads to the minimal parabolic
functions I.XIX(IO.2); the second is In invariant excessive.
Appendixes
Appendix I

Analytic Sets

1. Pavings and Algebras of Sets


If X is a space, a paving of X is any class .( of subsets of X which includes the
empty set. If 1' is a paving, ff, [YJ denotes the class of countable unions
[intersections] of .1 sets. If a paving includes the complements and the
finite [countable] unions of its sets, the paving is called a [a] algebra. The
smallest [a] algebra containing a paving T, that is, the intersection of the
collection of all [a] algebras containing.', is called the [a] algebra generated
by Y. If :1 is an algebra, the a algebra generated by . is the smallest collec-
tion of sets containing .I and closed with respect to countable monotone
unions and intersections.
If (X, ,T) and Y, :1) are paved spaces, the product paved space is the
product space X x Y together with the product paving I x >,t4' consisting of
the class of product sets A x B with A in .I and B in V. If I and :' are a
algebras, f X ?y denotes the product a algebra, that is, the a algebra gener-
ated by the product paving.

2. Suslin Schemes
We use the notation N = 71+ x71+ x . Let (Y, ?/) be a paved space. A
Suslin scheme is a map Y, from the set of finite sequences of points of 71'
into #, (n, , ... , nk) f-s Yn nk
E V. To each point n, = (n, , n2, ...) of N
then corresponds the intersection Yn1 n Y-1-2 n , and the uncountable
union U. E N n Y,' n - ) is called the nucleus of the Suslin scheme
and is said to be a set analytic over 4. The class of these nuclei will be denoted
by If A e ' and if Y n... -'k = A for all k and n I,..., nk, the nucleus is
A; soYY c Furthermore s.4(J),, = d(J) because if Ate.d(J), say

Aj = U (Yn n Yin, n2 n ... ) Y,,.. ..nkE-N.


(2.1)
ncN

let at be a one-to-one map from 71+ onto 71+ x71' and define

Yn,.....nk = YS(n,)n,...nk.
742 A I Analytic Sets

Then U ; Al is the nucleus of Y. Somewhat less easily sa9(J), = .d(?'). To


prove this. suppose that A; is given by (2.1). Then no Al is an uncountable
union whose general summand can be obtained by choosing in succession the
first two sets from the union for A, , then the first from the union for A 2, then
the third from the union for A,, then the second from the union for A2. then
the first from the union for A3, and so on in the usual diagonal procedure, so
that (remembering that the subscripts are dummies)

I 1 A/ = U n Yin.., n YZn' n Yin, n,, n Y2n,n, n ...


41 tN

The left side is the nucleus of Y; defined by

Y = Yin,, Y,,;nz = Yin'-

'-The class although closed under countable unions and intersections,


is not necessarily a a algebra because it is not necessarily closed under
complementation.

3. Sets Analytic over a Product Paving

Theorem. Let (X, .,Y') and (Y, -Y) be paved spaces. Then x d(,) c
x fl. If SEX and AE,vI(.1' x ?%), then {q: (5,tj)EAj

To prove the first assertion, suppose that A ef. Using the obvious
notation, it is clear that

A x .d(0#) c .W(!T x 1'). (3.1)

Similarly, if Ac.W(g4),

.4(57') x A c .s1(; ' x ^t'), (3.2)

and this is the desired inclusion relation. To prove the second assertion,
suppose that 4E.91(-T x N), say
,i = U (X x Y 7)n ...

The section of A for given is the nucleus of for

Yn, ... "k if a Xn, . -'I'


Zn, nk(0 =
10 otherwise.
5. Projccuon Charactcrication . V(Y) 743

4. Analytic Extensions versus a Algebra Extensions of Pavings

Theorem. IJ'the complement of each set in the paving J is in sd(ft sd(-#)


contains the a algebra generated by J.

The algebra .alo(2') generated by J is the class of finite unions of finite


intersections of J sets and their complements. Hence .ado(') c Since
sal (aJ) = .sal (CJ)s = sd (fts, the class d(') is a monotone class. According to
a standard theorem, sd(V) must therefore contain the a algebra generated
by,W,(?J), which is the same as that generated by'.
For example, if ' is the class of closed [open] subsets of a metric space
Y, sd(xJ) contains the open [closed] subsets and therefore contains the a
algebra of Bore] subsets.

5. Projection Characterization of sl (O.M)

The following characterization of sd(,Y) is important for the application of


the Suslin operation to stochastic processes.

Theorem. Let (Y, ,.W) be a paved space, and let (X, I') be a topological space
paved by its class of compact subsets. Then the projection on Y of a set in
.al(; ' x ') is in .W(W). For a suitable choice of (X,,1), and some suitable
choices are compact metric, sd(4') is the class of projections on Y of the sets in
(X x V),a
Suppose that the paved spaces are as described in the first assertion and
that A E sd(.1 x -Y),

A = U (Xn, x Y.,)n(X.,., x Y.,.) n ... Xn,. nkEX, Yn,...nkE:'

n Xn n Y.'n, n...

n ... nk is defined by

` Yn, ... nk
if Xn r) ... n 'n nk # 0
10 otherwise,

the projection of A on Y is the nucleus of A and is therefore in sd(J), as


stated in the first assertion of the theorem. To prove the second assertion,
define X as the space of all sequences n, of extended (that is, < + co) positive
integers, metrized to be compact by defining the distance between m, and n,
to be
744 A.I. Analytic Sets

m
E 2-'larctan mj - arctan nj!.
i=o

If n, . ... . nk are in 7L let X., , , .,,k be the compact set of points of X with first
k coordinates nt, ... nk. The nucleus of the Suslin scheme Y, on Y is then
the projection on Y of

U R. nX
IIE RI
00

=I I U (X.,.. nk X Y.'....') E \.' x oCo,


0

where for each k the union in the second expression is over all finite-valued
k-tuples nt . . . ,nk.

6. The Operation sl(sd)


Theorem. sal[.sal(M)] =

The proof in Section 2 that d(QJ) = sal(V), = sal(4t1)a can be extended to


prove this theorem, but the following proof is less tedious. A set in sd [,sal(QJ)]
is the projection on Y of a set in [.t x for a suitable paved space
(X, X), as described in Theorem 5. Since (from Sections 2 and 3)

[X X Sd(Y)]ca c [.sal(.1 x &)]"a = dff x (W)

and since (by Theorem 5) the projection on Y of a set in the class on the right
is in the present theorem is true.
Application. The first assertion in Theorem 3 can now be strengthened.
In fact
.al[.W(-6C) x .sal(f)] = d(. ' x N)

because Theorems 3 and 6 combine to yield

.sd[dff) x c sd['4(X x V)] = Cal(.' ' x 6#).

7. Projections of Sets in Product Pavings


Theorem. Let (X, 5t") be a locally compact second countable Hausdorff space
paved by the class of its compact sets. If (Y, ''J) is a measurable space paved by
its class of measurable sets, the projection on Y of a set in sd [sal (5C) x 6Y] is in
vv*).
9. The G, Sets of a Complete Metric Space 745

In view of Theorem 5 it need only be pointed out that the application in


Section 6 implies that sd(X x 91) = sd[sd(1) x V].
Application. In particular, if R(X) is the class of Borel subsets of X in
Theorem 7, the projection on Y of a set in sd[_V(X) x 6J] is in sd(IJ).

8. Extension of a Measurability Concept to the Analytic


Operation Context
Theorem. If (X, I), (Y, ,W) are paved spaces and if f is a function from X into
Y with f `(IN) c d", then f -' (sd(QJ)) c szT(°.,l").

This theorem is a trivial consequence of the fact that the inverse image
under f of an arbitrary intersection [union] of sets is the intersection [union]
of their inverse images.

9. The Ga Sets of a Complete Metric Space


The following standard topological theorem is proved for orientation.

Theorem. (a) A Gs of a complete metric space is homeomorphic with a com-


plete metric space.
(b) A set which is a subset of a complete metric space and is homeomorphic
with a complete metric space is a G.
(c) A complete separable metric space is homeomorphic with a Gb in a
compact metric space.

Note that since a homeomorph of a separable space is separable, (a) and


(b) are trivially specialized to separable spaces. Throughout the following
proof, X is a complete metric space with distance function d.

Proof. (a) If X, = no O" with O" an open subset of X, define a new distance
function d, on X, by

dt( Y2-" A
0

to get a complete metric space (X,, d,) homeomorphic with (X,, d).
(b) Let f map X homeomorphically into a complete metric space X'.
Let a(') be the oscillation of the inverse function at ' in the closure off(X),
and define a as 0 off the closure off(X). Then a is upper semicontinuous on
X' and-vanishes on f(X). Suppose that ' is in the closure of f(X) and that
0. Let B;, be the ball of center ' and radius I/n, so that f -' (B") is
746 A.I. Analytic Sets

open and shrinks (n - oo) to a point of X. Necessarily ' =


Let A. be the set of points of X' at distance <1/n from f(X). Then
n' [A.n(a < 1/n)] = f(X) is a G,,, as was to be proved.
(c) Replacing the distance function don X by d A I if necessary, it can

-
be supposed that d 5 1. Let , be a sequence dense in X. Then

z 1)

is a map of X into the space X' of sequences b of positive reals S 1. If the


distance between b. and b; in X' is defined as Ep 2`1b. - b.1, the space X'
is compact, and f is a homeomorphism. According to (b),f(X) is a G. o

10. Polish Spaces


A Polish space is defined as a Hausdorff space homeomorphic with a com-
plete separable metric space. Applying Theorem 9 and the fact that a
homeomorph of a separable set is separable, the following two assertions
are trivial.
(a) A Ga in a Polish space is a Polish space. In particular, every open
subset (or G5) of I8" is a Polish space.
(b) A Polish space can be defined as any homeomorph of a Ga of a
compact metric space or of a Ga of a complete separable metric
space.
If X and Y are complete metric spaces, the product space topology can be
metrized to make X x Y complete metric, and this product space is sepa-
rable if X and Y are. If X and Y are Polish, the product space with product
topology is Polish.
If X is a Hausdorff space which is locally compact and second countable,
X is an open subset of its one-point compactification. Since the latter is
metrizable, X is Polish.

11. The Baire Null Space


This space, denoted by BN below, is the metric space of sequences (nj, j Z 1)
of strictly positive integers with

dist(m., n,) = (first j with mj # nj)-' if m, 96n,

This space is complete and separable. It is zero dimensional because there is


a topological base of clopen sets, the sets obtained by fixing the f i r s t k co-
ordinates, k = 1, 2, ... .
According to Theorem 9, the set of irrationals in [0, 1], a G8 in the com-
12. Analytic Sets 747

plete metric space [0, 1], is homeomorphic with a complete metric space. In
this case the image can be taken as the Baire null space, and the map can be
written explicitly by means of the continued fraction representation of an
irrational number x in ]0, 1 [,

x n EBN.
nt +
nZ+...
It will be convenient to denote by BN°° the space of sequences {n;,j Z 1) for
nn a strictly positive integer or + oo, with metric
a)
dist(m,, n,) = Y 2-ijarctan mt - arctan nnl.
I

The space BN°° is compact metric, and the subset of this space with finite
coordinates is a G6 homeomorphic with BN.

Theorem. Every Polish space is a continuous image of BN.

Let X be Polish, and assume without loss of generality that X is complete


separable metric. Choose closed nonempty sets X,, X2, . .. of diameters < 1
with union X. If Xnl , .. nk has been chosen, choose Xn, ... nk t . Xn,... nk2,
closed nonempty sets of diameters < I /(k + 1) with union X,... "k. The
desired map from BN into X is the one taking n. into the single point in
Xn, nXn , n2 n .

12. Analytic Sets


A set will be called analytic if it is a subset of a metrizable space and is
analytic over the class of closed subsets of the space. The class of analytic
subsets of a metrizable space includes the class of its Borel sets (Section 4).
This inclusion is strict in all interesting contexts, for example, if the space is
RN

Theorem. The following conditions on a subset A of a Polish space are


equivalent.
(a) A is analytic.
(b) A is the continuous image of a Polish space.
(c) A is the continuous image of a G6 of a compact metric space.
(d) A is the continuous image of BN.
(e) A is the continuous image of the set of irrationals in [0, 1].

In view of Theorems 9 and I 1 only (a). (d) needs proof.


748 A.I. Analytic Sets

Proo/. (a) (d) Under (a) there is a complete metric space X such that the
set A is the nucleus of a Suslin scheme X.: n, , .... nk,-
X nk a closed subset of X. It will be convenient, and irrelevant to the
construction of analytic sets, to restrict n, , n2, ... to be strictly positive. For
j and k strictly positive integers choose B;k a closed subset of X, of diameter
< Ilk. with B;k = X. Then

A = U(B , nXe) (12.1)

where the union is over all sequences n, , n2, ... of strictly positive integers.
The kth set in parentheses is closed and has diameter <I/k. Let
x(a, h) be a one-to-one map of the set of pairs of strictly positive integers onto
the set of strictly positive integers, and define

Xaln,. n.) = n
Xaln,. n.laln,. nal = (Bn, n n (B,,2 n

to get a Suslin scheme X; for which, given n,, the set nk is closed, has
diameter < Ilk, decreases ask increases, and either is empty for sufficiently
large k or shrinks to a singleton f(n,). The function f fis thereby defined on a
subset of BN, is continuous, and maps its domain onto A. To extend f to
BN, choose some point of A, and if is not empty, choose a point
k in this set. If f(n,) is not defined, either X is empty, in which case

define 1(n) = ', or there is a maximal k >_ l for which is not empty,
in which case define./'(n,) = S k. Then J'is a continuous map from BN
onto A. (Alternatively, it is easy to see that the original domain off is closed
and so is itself a Polish space and as such is the continuous image of BN,
which makes A the continuous image of BN.)
(d) = (a) If (d) is true A = f(BN), where f'is continuous from BN into
the Polish space containing A. Let Y ... k be the set of points of BN with
first k coordinates n, . . . h 1 . This set is closed and has diameter 11(k + 1).
. .

Let X ,, be the closure of /(Y k). If n, a BN, .

Yn, nk 1 In.}. Xn, ..nk 1f(n.) (k - or).


and therefore A is the nucleus of X. and is analytic, as was to be proved. o

13. Analytic Subsets of Polish Spaces


Theorem. Let X and Y he Polish spaces.
(a) Il A [B] is an analytic subset of'X [Y], A x B is an analytic subset of
X X Y.
13. Analytic Subsets of Polish Spaces 749

(b) If A is an analytic subset of X x Y and iJ' E X, the set J? J: ry) E'4)


is an analytic subset of Y.
(c) The projection on Y of an analytic subset of .Y x Y is analytic.
(d) The image on Y of an analytic subset of X under a Bore! measurable
map from X into Y is analytic.

Parts (a) and (b) are special cases of Theorem 3. To prove (c), observe
that the projection map is continuous; so in view of Theorem 12(b) the
projection on Y of an analytic subset of X x Y is the continuous image of a
continuous image of a Polish space and is therefore analytic. To prove (d),
let f be a Borel measurable function from X into Y. The graph off is a Borel
and therefore analytic subset of X x Y. Furthermore, if A is an analytic
subset of X, A x Y is an analytic subset of X x Y; so the intersection A of
the graph off with A x Y is analytic, and its projection on Y, that is to say
J(A), is therefore analytic by part (c), as was to be proved.
Appendix II

Capacity Theory

1. Choquet Capacities
If (X, . ') is a paved space for which l is closed under finite unions and inter-
sections, a Choquet capacity on (X,.'), that is, on X relative to X, is a func-
tion / from the class of subsets of X into k with the following properties:
(a) / is increasing; that is, X, X2 implies that 1(X,) < 1(X,).
(b) X. T X,,, implies that
I X, implies that I(Xj.
A subset A of X is called capacitable [relative to (X,'1',1)] if

1(A) = sup {1(B): B c A, BEfa}. (1.1)

It is trivial that the union of a monotone increasing sequence of capacitable


sets is capacitable.

2. Sierpinski Lemma
Lemma. Let (X, .:'1') be a paved space, let A he the nucleus of a Suclin scheme X.
over .r', and let h, , h2, ... he a sequence in P. Define

Ak = I I X' n X^,nz n . . .

n :ni56i
<k (2.1)
M
Bk = U Xn, n ... n X., nk B=nB'.

ni < bi
'Sk

Then AkcA.AkcBk,B' DB2D ,and BeA.


Only the assertion that B c A is not trivial. To prove this relation. suppose
( n ,.. nk) with ni < bi for
that 6 E B, so that for every k there is a k-tuple (n,
which E Xn n n X n ...nk. Call such a k-tuple admissible. Then admis-
( n ,..nk_,). Since the
sibility of (n,, ... , nk) implies admissibility of (n,
4. Lusin's Theorem 751

first integer of an admissible k-tuple is at most b,, some integer m, must


be the first integer of an admissible k-tuple for infinitely many values of k.
Similarly, for some m2 S b2 the pair must be the first two integers
of an admissible k-tuple for infinitely many values of k, and so on, so that
there is a sequence m, with e X., n X,,,,., n . c A, and therefore B c A,
as was to be proved.
Observe that Be Xa if X is closed under finite unions and intersections.
The set B increases when members of the sequence b, increase, and it is
intuitively reasonable that in some useful sense B can be made close to A
by choosing each bt sufficiently large. The following Choquet Capacity
Theorem offers a precise version of this closeness.

3. Choquet Capacity Theorem


Theorem. Let (X, X) be a paved space for which X is closed under finite unions
and intersections. If I is a Choquet capacity on (X, X), the sets of ,rd(1) are
capacitable.

If A e.4(fl and 1(A) = - oo, the set A is trivially capacitable because


the empty set is in 2. If I(A) > - oo, choose a < I(A). It will be proved,
in the notation of Sierpinski's lemma, that I(B) z a if b, is chosen suitably,
and in view of property (c) of a Choquet capacity it is sufficient to choose b,
to make I(B") Z a for all k and therefore sufficient to choose b, to make
I(A') > a for all k. To make such a choice, choose b, so large [property (b)
of capacity] that I(A') > a, and if b,, . . , bk_, have been chosen so that
.

I(Ak-') > a, choose bk so large that I(Ak) > a, again using property (b).

4. Lusin's Theorem
Theorem. If (X, X, p) is a complete measure space for which p is a countable
sum of finite measures, for example, if p it a finite, then .rd(1) = T.

It is sufficient to prove the theorem for p a complete finite measure.


Define the outer measure of an arbitrary subset A of X by

p*(A) = inf {p(B): A c Be X }.

Then p is a Choquet capacity relative to (X, i'). In fact the defining property
Section 1(a) is trivial, (b) is a standard property of outer measures defined
in this way, and (c) is a standard property of finite measures. According to
the Choquet capacity theorem, the sets of .W(X) are capacitable, and in the
present context capacitability and measurability are equivalent.
752 .Capacity Theory
Al!.

5. A Fundamental Example of a Choquet Capacity


Let (D, .F, P) be a complete finite measure space, and define an outer
measure P' on the class of subsets of i2 by

inf {P(B): B D A, Be .F }.

Let I be the class of finite unions of subsets of R+ x 0 of the form C x A,


where C is a compact subset of R+ and A e .F. Then A is a paving of R+ x i2,
closed under finite unions and intersections. If A c R+ x i2, let x(A) be the
projection of A on i2, and define I(A) = Observe that

.d(A) = s4(9(R+) x .F) z) R(R) x ,F,

that sV(.F) = F, and (by Theorem 1.5) that I(A) = P(R(A)) if


A e d (F). An important result in capacity theory is that I so defined is a
Choquet capacity on R+ x i2 relative to A, that is, that I satisfies Section
I (a)-(c). Condition (a) that I is monotone increasing is obviously satisfied.
The left continuity condition (b) is satisfied because every outer measure
defined like P* in terms of a measure satisfies this condition. To prove that
the right continuity condition (c) is satisfied, suppose that S, is a decreasing
sequence in A with intersection S. Then the point co is in if and only
if the compact set (t: is not empty; so x(S,) is a decreasing
sequence in 3F with intersection n(S). Hence

lim I(Sn) = lim P(n(S)) = 1(9);

that is, (c) is satisfied.


Application. It is trivial that since I is a Choquet capacity relative to A,
this set function is also a Choquet capacity relative to an arbitrary subpaving
of F closed under finite unions and intersections. The Choquet capacity
theorem applied to I with various choices of such subpavings of A can be
applied (Section 2.11.8) to show that certain subsets of R+ x .0 contain
significant parts of the graphs of suitably defined functions from 0 into R.

6. Strongly Subadditive Set Functions


Let (X, .f) be a paved space for which :i<" is closed under finite unions and
intersections. A function I from £1 into either ] - oo, + oo] or [- oo, + oo[
is called strongly subadditive on (X, X) if
(a) I is increasing; that is, A c B implies that 1(A) S I(B).
(b) I(AuB)+I(AnB)5I(A)+I(B).
If I is strongly subadditive and a is a constant, I A a and I + a are also
strongly subadditive. Under (a), condition (b) is equivalent to
7. Gcncration of a Choquct Capacity by a Positive Strongly Subadditive Set Function 753

(c) For an arbitrary strictly positive integer n when AI, .... A and
Bl, ...,Bnare in !and AicBBforj<n,

I(Ue,)+1(At)<>1(Bt)+/(UA). (6.1)
J t i t

In fact under (a) and under (c) with n = 2 inequality (6.1) yields (b) when
A, = A n B, A2 = A, B, = B, B2 = A. Conversely, if (a) and (b) are true,
(c) is proved as follows. If A = B, and B = A I u B2, conditions (a) and (b)
yield

1(B, v B2) + 1(A,) < 1(B,) + 1(A, v B2). (6.2)

If A = A I u A2 and B = B2, conditions (a) and (b) yield

I(AI u B,) + I(A2) < 1(AI u A2) + I(B2). (6.3)

These two inequalities combine to yield (6.1) for n = 2, and the general case
is proved by induction. If (6.1) is true for a countable sequence (involving
countable unions and sums) of pairs (As, BB) in X with Aj c B,, the set
function I is called countably strongly subadditive.

7. Generation of a Choquet Capacity by a Positive Strongly


Subadditive Set Function

Let I be a positive strongly subadditive set function on (X, .1), and let f be
a subclass of T,, closed under finite intersections and countable unions.
Define the function A F-+ 1 *(A) from the class of all subsets of X to 68' by

1*(A)=sup{I(B):BcA,Bed'} ifAeif', (7.1)

and for an arbitrary subset A of X define

1*(A) = inf{1*(B): B A.BeI (7.2)

In (7.2) it is to be understood as usual that the right side is to be + oo if ..f'


contains no superset of A. Equations (7.1) and (7.2) are consistent for A
in Y. Note that /* need not be equal to / on .!. For example, if ! is empty,
1 *(A) _ + oc for every A.

Theorem. In the preceding context if


(a) 1(A4) -*1(A) lihen A4 and A are in! and A41 A
(b) jA
754 All. Capacity Theory

then 1 * is an extension of I, does not depend on the choice of- f, and is a Choquet
capacity on X relative to X, countably strongly subadditive on the class of all
subsets of X. If I *(A) < + oo, the set A is (X, .Y) capacitable if and only if to
each e > 0 correspond sets A. in .Q'a and A; in X satisfying

A; e A C A", I *(A") < 1 *(A;) + E. (7.3)

It is trivial from (b) with A, = AZ = that 1 * = I on X. The proof


that J* is a countably strongly subadditive Choquet capacity will be carried
through in four steps.
(i) A. T A with A. in I implies that I*(A) because if A. _
Uk=o A,,,, with in I, if B. = Um,kS Amk, and if B is in "' and is a subset
of A. then

lim lim I(B n B) = 1(B). (7.4)

The limit on the left is therefore at least (and trivially at most) I*(A).
(ii) 1 * is strongly subadditive on the class of all subsets of X. In fact,
since the inequality in Section 6(b) is true for A and B in d", increasing
sequences of pairs (A, B) yield the same inequality for sets in Y. If A and B
are arbitrary and if either P (A) or P (B) is infinite, this strong subadditivity
inequality is trivially true for P. If both values are finite and A e A' a X,
BcB'e.T.then
I *(A v B) + /*(A n B) < 1 *(A' u B') + !*(A' n B') < !*(A') + 1 *(B'), (7.5)

and (ii) is true because the right side can be made arbitrarily close to
!*(A) + I*(B).
(iii) A. T A implies that P (A.) -4 !*(A). This conclusion is trivial if
+oc. If this limit is finite and if t > 0, choose A;, in 1 in
such a way that A. D A,,, 1*(A,,) < 2-"e. By strong subadditivity
[see (6.1)]

(7.6)
1 J \1 JJ

Hence

A;,) = lim 1 * (U A;,l 5 lim I *(A,) + E,


A. (7.7)
1 *(A) 5 1 *
ff P. `, JJ
m-ao

so that limm_. 1 *(Am) is at least (and trivially at most) P (A).


(iv) P is is count ably strongly subadditive because 1 * satisfies (6.1) for arbi-
trary sets A, c B,, and this inequality yields countable strong subadditivity
in view of (iii) when n -+ oo.
8. Topological Precapacities 755

The conditions (a), (b), and (i)-(iv) imply that I* is a countably strongly
subadditive Choquet capacity relative to (X, 1). The last assertion of the
theorem is trivial. There remains the proof that I* does not depend on the
choice of 9i . Let 1, be the Choquet capacity obtained with the choice
9C = X and let 1* be the Choquet capacity obtained with some other
choice. It is assumed that (b) in the theorem is true for both choices. It is
trivial that 1; 5 P and that there is equality on 9C and therefore also on X..
Now if A is an arbitrary subset of X,

1,(A)=inf{1;(B): B=3 A,Be9C,} =inf{t*(B): BnA,BeXC,} zP(A),

and therefore 1; = P, as asserted.

8. Topological Precapacities
Let (X, 9C) be a locally compact second countable space together with its
class of compact subsets. Suppose that I is a function from £ into l with
the following properties:
(a) I is strongly subadditive on T.
(b) If A. is a monotone sequence of compact sets with compact limit A,
then 1(A).
Then I will be called a topological precapacity on X. Theorem 7 can be applied
with t the class of open subsets of X if 1*(A) = I(A) for A compact, and
this is true because if B. is a decreasing sequence of relatively compact
open subsets of X with no B = no B = A, then in view of (b) and the
fact that 1(A) 5 5
1(A) 5 lim 5 lim I(B,) = 1(A). (8.1)
M-10 n-.o

Thus a topological precapacity I has an extension to a countably strongly


subadditive Choquet capacity 1* relative to the class of compact sets and

1*(A) = sup {1(B): Bcompact, B c A} (A open), (8.2)

I *(A) = inf {1 *(B): B open, A c B} (A arbitrary). (8.3)

The class of capacitable sets includes the analytic subsets of X. A set A with
1*(A) finite is capacitable if and only if to each s > 0 correspond a compact
subset A; of A and an open superset A, of A such that 1 *(A,') < 1(A + C.
Letting a run through a sequence with limit 0, it follows that A with 1 *(A)
finite is capacitable if and only if there is an F. subset A' of A and a Ga
superset A" of A such that 1*(A') = 1*(A) = 1*(A"). This condition is
trivially necessary and sufficient for capacitability of A when 1 *(A) = + oo.
756 A. If. Capacity Theory

Note, however, that if A has compact subsets with arbitrarily large values
of 1, then 1 *(A) = + oo, and A is capacitable no matter how complicated
its structure.

EXAMPLE. Let (X, 3r) be R together with its compact subsets, and define
1(A) for A compact as 0 or I according as A is empty or not. Then I is a
topological precapacity. Let . be the class of open subsets of R. The exten-
sion l* is l on every nonempty set, and all sets are l* capacitable. Observe
that l* has the property that there is a decreasing sequence A. of capacitable
sets for which

1*(0 -X
0

Application to a Uniqueness Result

Let I be a function from the class of all subsets of a locally compact second
countable space into 9'. Suppose that I is monotone increasing, that the
restriction of I to the class of compact sets is a topological precapacity, and
that if B is open

I(B) = sup {I(A): A c B, A compact}.

It follows that the extension of this restriction to /* coincides with / on the


class of capacitable sets, in particular, on the analytic sets.

9. Universally Measurable Sets


Let (X, _Y) be a measurable space, and let p be a finite measure on T. The
class d,, of p measurable sets obtained by completing p is the a algebra of
sets generated by 9C' and the subsets of p null sets. The intersection W(T) =
nN -n is a a algebra, the class of universally measurable sets. It is easy to
verify that ''(X) is unchanged if y in this definition is allowed to be a finite
or more generally if p is allowed to be a countable sum of finite measures.
Sincelt'(I() c X'N for every finite measurep, it follows that !(4!( )) c II(T),
and the reverse inclusion is trivial.
Apply Lusin's theorem to find that for every finite measure p on 5C,

'1!(.-T) c sa1('Wff )) C d(xN) CT"

so that
!(?) c A(V(I)) c V(.fl,
that is, .c1(?l( )) = 1(.").
9. Universally Measurable Sets 757

If f is a map from the measurable space (X, 37) into the measurable space
(Y,,Y) and if f -' (,Y) c f that is, if f is measurable, a measure p on X trans-
forms into a measure Nf on IN by way of pf(A) =,u [f -'(A)]. Moreover, if µ
is complete, this relation remains valid for A in the domain of the completion
of pr. In particular, this relation is valid for A in V(Y) and c:
*(X); that is, f is measurable from (X,,&(. ')) into (Y, W(G')).
Appendix III

Lattice Theory

1. Introduction
The following is an outline of the lattice theory used in this book. Elementary
proofs are omitted or merely sketched. Vector spaces are always over the
reals. The reader is warned that the nomenclature of this subject has not
been standardized and that therefore the definitions given below are not
universally accepted.

2. Lattice Definitions
A lattice is a class of objects ordered by a transitive reflexive binary relation
for which every pair x, y of the objects has a unique order supremum x Y y
and infimum x A y. The lattice will be called complete if every subset has
a supremum and infimum. If every upper-bounded subset has a supremum,
then every lower-bounded subset has an infimum, the supremum of the lower
bounds of the subset, and if every lower-bounded subset has an infimum,
then every upper-bounded subset has a supremum, the infimum of the upper
bounds of the subset. If these suprema and infima exist, the lattice will be
called conditionally complete.
A lattice will be said to have the countability property if each subset which
has a supremum [infimum] contains a countable subset with the same
supremum [infimum].
A subset of a lattice is called a sublattice if x Y y and x A y are in the
subset whenever x and y are.

3. Cones
A cone is a set closed under a commutative addition operation (x + y) and
multiplication by positive constants (cx) satisfying the usual associative and
distributive laws together with

x+y=0=x=y=0.
x + J. = x + y' . y = y' (unique subtraction).
4. The Specific Order Generated by a Cone 759

When x + y = z, the element y will be written z - x. According to this


definition a cone .K in a vector space 4Y is a convex set with the property
that xe.N' implies that cxe.N' if and only if c Z 0 or x = 0. The cone is
said to generate .!! if .,t = .N' - .N', that is, if every element in ./1 is the
difference between two elements of X. An arbitrary cone .K can be immersed
in a vector space.!! generated by .K, as follows. The elements of .,IY are the
pairs x: (x, , x2) of elements of A' under the identification (XI, x2) _ (Yi,Y2)
when x, + Y2 = x2 + y,. This identification is a transitive relation because
there is unique subtraction in A. Addition of elements of .4" and multiplica-
tion by real constants are defined by

(x1,x2)+ (Y1,Y2) _ (x1 +Yt,x2 +Y2)

(cx, , cx2) if c Z. 0,
c(x, , x2) _
((-c)x2, (-c)x,) if c < 0.
The cone . t' is identified with the class of elements of .9 having representa-
tions of the form (x, 0), and (x, , x2) is also written x, - x2 when there is no
danger of confusion.

4. The Specific Order Generated by a Cone


A cone can be ordered by the convention that x y if there is a cone element
z such that x + z = y. This order, called the specific order, is transitive and
reflexive, and x y implies that cx < cy when c 2! 0 and that x + z y + z
for all cone elements z. Moreover if x + z y + z for some cone element z,
then x y. A supremum or infimum of a set of cone elements is necessarily
unique.
If in the specific order every pair of cone elements has a supremum
[infimum], then every pair also has an infimum [supremum] and the cone
is therefore a lattice. To see this, suppose for example that every pair x, y
has a supremum x Y y. Then it will now be shown that x A y exists and

xYY+xAy=,r+Y. (4.1)

Since .r + y >-.x and x + Y >-y . it follows that x + y xYy. Define


z=x+y-xYy.Then

z z x. On the other hand, if z' x and z' y, then

_'+xYy=(z'+x)Y(z'+Y)-< ,r +Y;
so z' -< z, and it follows that z = x,A v, as asserted.
760 All. Lattice Theory

If .4' is a vector space and if .N' is a cone in #, the space A! can be given
a transitive reflexive order by setting x < y if y - x e . Y. The order thereby
assigned to .4' is the specific order, and .N' becomes the set .4' K+ of positive
elements of .,K, that is, the set of elements x 0. This ordering of .,' is
called the specific order relative to X. It is trivial that in this order the pair
of relations x< y and y< x implies that x = y. Conversely, let .,K be a
vector space with a transitive reflexive order in which this implication is
valid and in which x< y implies both that cx < cy for c e 68+ and that
x + z < y + z for z e.,&. The set .&+ of elements ::0 is then a cone, and the
.4' order is the specific order relative to .,#+.

5. Vector Lattices

A "vector lattice" .4' is defined as a vector space which is a lattice in the


specific order determined by some cone .,/l+ as described in Section 4. The
cone .4'+ is a sublattice of J( and has the property that to each element xo
of .' corresponds an element x (for example, xo Y 0) of .4'+ such that
x x0. Conversely, if .At is a cone in a vector space .,!!, if .K is a lattice
in its specific order, and if to each element xo of ..K corresponds an element
x in .A' such that x - xo is in .At, then .4' is a lattice in the specific order
induced by X. To see this, observe that if x, and x2 are in .', there must
be an element x in A' such that x - x, and x - x2 are both in .K, and it is
easy to verify that the desired supremum x, Y x2 is given by

x-[(x-x,)A(x-x2)]
and that then the desired infimum is given by -[(-x,) Y (-x2)]. A
vector lattice .41' is [conditionally] complete if and only if its positive cone
is [conditionally] complete in its specific order. For example, if .,K+ is
conditionally complete, a subset {xa,aeI} of .. t' with an upper bound x
has a supremum, namely x -,&, , (x - x.).
If I( is a vector lattice,

z+(xYy)=(z+x)Y(z+y), z+(.xAy)=(z+x)A(z+),) (5.1)

and

(-x) ,k (-y) = -(x Yy). (5.2)

The property (5.2) can be used to deduce properties involving suprema from
dual properties involving infima and conversely. We shall use the notation

.r+=xY0, x-=(-x)Y 0, IxI =xY(-x). (5.3)


5. Vector Lailicc3 761

Applying (5.1) and (5.2),

x+y - X.

that is, (4.1) is true for vector lattices. We rewrite this result for later reference
together with some useful relations easily deducible from it.

xYy+xAy=x+y, x=x+-x-, jcxj =Icl 1xI, x+Ax-=0,


1xI = x+ + x- = x+ Y X-' Ix +),1 1xI + I yl.
(5.4)

The second relation implies that + - ..ll+. If x = x, - x2 with


xi b, then x, x+ and x2 x-.
Associative Law. If is a subset of the vector lattice .#, if I is
partitioned arbitrarily, I = U Oc j Ip, and if yE,P. x, = yy exists for every fi
in J, then YE,x, exists if and only if Y/.,, ys exists, in which case the two
suprema are the same. This fact is trivial, as is the corresponding dual one
for infima.
Distributive Law. Let {x,, at e I } be a subset of .,# for which Y.. Ix, exists.
It follows that Y.. l (x, A y) exists for y in .,9 and that

Y (x, A y) = (Y x,) Ay. (5.5)


aEl aEI /
Dually, the existence of A,E, I.Y. implies that A,E, (x8 Y y) exists and that

It A(x,Yy)_(AxaYy. (5.6)
;Cl SEl )

Only (5.5) will be proved. The right side of (5.5) majorizes xa A y for every a,
and conversely, if x x. A y for every a, then

Yy;
Ky
aEl J

so

Yxa+y-[(Yx,)Yy]=(Yx))Ay,
and therefore (5.5) is true.

-All
762 A.M. Lattice Theory

6. Decomposition Property of a Vector Lattice


Theorem. If x, y, y2 are in the positive cone of a vector lattice and if
X < y1 + y2, there are positive elements x1 and x2 satisfying

x1::5y1, x2:! y2, x1+x2=x. (6.1)

In fact, if x1 is defined as x A y1 and if x2 = x - x1, then

x2=(x-y1)Y0,
so that x2 is positive and is majorized by y2.
This decomposition theorem implies that if y1, y2, y are positive elements
of a vector lattice, then

(y1 + y2) A y y1 )' y + y2 A y. (6.2)

In fact the left side is majorized by y1 + Y2; so according to the decomposi-


tion theorem, there are positive elements x1, x2 satisfying

x1 + x2 = (Yi + Y2) A Y.

Then x, -< y; so x; y, A y, and it follows that x1 + x2 y1 A y + y2 A y,


which yields (6.2).

7. Orthogonality in a Vector Lattice


Elements x and y of a vector lattice N are said to be orthogonal if 1xI A Iy1
= 0, a relation written x 1 y. Two subsets of .A' are said to be orthogonal
if every element of one is orthogonal to every element of the other. The set
of elements of .4' orthogonal to a set .K is denoted by .N'1. According to
(5.4), x 1 y if and only if IxI + lyl = 1xI Y IA.

8. Bands in a Vector Lattice


A band in a vector lattice ..K is defined as a vector sublattice .A' satisfying
the following two conditions:
(a) If a subset of _41' has a supremum in ..K, the supremum is in .K.
(b) If jxj :! jyi and if ye.N', then xc-- .
Condition (a) implies the validity of the dual condition for infima. If a
cone W of positive elements of ..K satisfies conditions (a) and (b), then
W - 16 is a band in .4'. If ,.4''o is a subset of .,K, the intersection of all the
bands containing No is a band, the band generated by No.
9. Projections on Bands 763

Theorem. If X is a subset of a vector lattice the set X' is a band.

To prove linearity of X', observe that if x e X' and c e R, then cx a Xl


because when y is in X,

I cxI A IYI (Icl + I)(IxI A IYI) = 0,


and if x, and x2 are in Xl, their sum is in Xl because (6.2) implies that
when y is in X,

Ix,+X21Alyl_IxtIAIYI+Ix2IAlyl=o.
The band condition (b) is obviously satisfied by X'; so if x and y are in
Xl, the elements lxl, lyl, lxi + IyI, x Y y, x k y are also in Xl, which is
thereby seen to be a sublattice of the given lattice..K. To finish the proof, it
will be shown that if {x., a E 1) is an arbitrary subset of X 1 with a supremum
Y, E, x, = x, then x E X 1. Suppose first that x. has a smallest element, say
xs = A,E,x Then YE,(x, - x5) = x - x,,, and if y is in X, an applica-
tion of the distributive law yields

I a6l
y (X. - IyI=QY[(xa-xp)A IYI]=0;

so x - xp, and therefore x are in A. In the general case choose any index
value fi, and replace x, by xQ = x, Y x, for all a. The set x; has a smallest
element ; so its supremum x is in X1, as was to be proved.

9. Projections on Bands
If X is a band in a conditionally complete vector lattice i, define the map
nX from .11 onto X by
n.-x= 1 / x A y (9.1)
YEJ,

for x 0, and for arbitrary x define n x = n,,-x+ - it x-. The map n,,,
reduces to the identity on X and tr,, x x for x 0.

Theorem. If X is a band in the conditionally complete vector lattice ..K, then


X" l = X, and ..K is the direct swn of X and .K' :for all x e 4', x = x, + x2 ,
where x, = ,t x e X and x2 = n,,1 E X 1. The map it, is linear, idempotent
and order preserving.

The map n, is obviously order preserving on ..dl+. The linearity of ny,,


to be proved below will therefore imply that n,,, is order preserving on ..9.
764 A.I11 Lattice Theory

If ye.K and if x 0, then IyI A (x- n, x) + it, x x, and since the left
side is a positive element of A, it follows that operating with it, on both
sides shows that the left side is majorized by it, x. Thus 0 IyI A (x - a, x)
0. The indicated infimum is therefore 0; that is, x - it, xe.,4%l, and x can
be written in the form x = x, + x2 with x, = 7r, x e . V' and x2 = x - tr , x C-
A". This representation of x as the sum of an element of ,'V and one of
.N'1 is trivially extendible to not necessarily positive elements x. Since such
a representation must be unique, ..# is the direct sum of .A" and /V l ,,y 11 =
At, and the map x i--* x, = it, x must be linear. Replace X by rV' above to
find that x2=rr,,x.

10. The Orthogonal Complement of a Set


Lemma. Let .iV be a subset of a conditionally complete vector lattice, and let
.N' be the band generated by .A,. Then No = ,,ti'l.

According to Theorem 9, .it''il = X. Now .A' c . Y" implies that .A'' c


.N'o and therefore that X," c .JV" = A°. In the other direction, since
, it follows that AN'o c . V ' ; so X," is a band containing A, and
. N ' o IAA'
must therefore include *', and so there is equality. Finally .A = .N'0'-1 =
.4", as was to be proved.

11. The Band Generated by a Single Element


Let z be a positive element of the conditionally complete vector lattice h',
and denote by .N' the band generated by z. Call an element y of l! bounded
(relative to z) if lyI -< cz for some constant c. All bounded elements are in
,N' so any supremum of a set of bounded elements is in .N". If B' is the class
of suprema of sets of positive bounded elements, B' - B' is a band, which
must be .ti'; so B' = ,-t%' . Actually every element of .A'' is the supremum
of an increasing sequence of positive bounded elements of .,!!, and in fact
we shall show that

a.,x=Y xA nz (xe.,!!'). (11.1)


nEZ'

To prove (11.1), define nx as the supremum on the right, so that it is a


map from -#' into . V' and rt x -< x. For k Z 0,

xAkz=n(xAkz)<rz2x:! nx; (11.2)

so (k oo) nx rr2 x rzx, and it follows that it = 7<2. The map it is additive
because on' the one hand using (6.2)
12. Order Convergence 765

n(x + y) nx + ny (11.3)

an d on t he other h an d

xAkz+ Y [yA(n-k)z]=xAkz+ny;
.Gz
(11.4)

so n(x + y) nx + ny, and there must be equality. Moreover x - nxe.N'1


because n(x - nx) = nx - n2x = 0, and ny = 0 implies that y 1 z, equiv-
alently (by Lemma 10), ye.h' . It follows that nr(x - nx) = 0; that is,
n,,x = n,,,ax = nx, as was to be proved.

12. Order Convergence


Let I be a directed set; that is, I is a set ordered by a transitive reflexive
binary relation 5 with the property that each pair of elements has an order
upper bound. A subset of I is called cofinal if each element of I precedes
some element of the subset. The set I has a cofinal countable set if and only
if I has a cofinal increasing sequence. Suppose that is a function from I
into a complete lattice L. The inferior and superior limits of are defined
by

liminf x(t) = Y k x(s),


t 1
lim sup x(t) = A Y x(s). (12.1)
id S2i it t6i fat

Then

lim inf x(t) 5 lim sup x(t), (12.2)


it it
and if these two values are equal, is said to have this common value as
limit. This discussion becomes trivial if I has a last element because then
both sides of (12.2) become the value of x(-) at the last element.

The Monotone Case

Suppose that the function is monotone increasing [decreasing]. Then


obviously has limit Y,E,x(t) [A,.,x(t)]. In particular, suppose that
is monotone increasing and that L has the countability property (Section
2). Define x = lim, t x(t). Then we now show that there is an increasing
sequence t in I such that lim, x(t) = x whenever t is an
upper bound of To see this, let t; be a sequence in I for which x(t') =
Y'. I x(t). Define to = to ; if to, ... , t have been defined, define ti+, in I so
that t t has the stated properties.
766 A.M. Lattice Theory

Moreover any increasing sequence s, in I with s >_ t for all n has these
same properties. Thus if 1 has a cofinal increasing sequence, we can increase
t. above if necessary to make t, cofinal in addition to the above stated
properties.

EXAMPLE. (Downward- and upward-directed sets). Let L be a partially


ordered set, and let L, be a subset of L. If L, when given the [reverse of the]
order inherited from L is a directed set, L is said to be [downward] upward
directed. If L is a lattice and if L, has the property that x Y y [x A y] is in
L, when x and y are, then La is a directed set in the order and is
directed upward [downward]. In both cases the identity map from Lo onto
itself is a monotone function from Lo into L; so the order convergence
concept is applicable.

13. Order Convergence on a Linearly Ordered Set


Theorem. Let be a function from a linearly ordered set I into a complete
lattice satisfying the countability condition.
(a) If I has no cofinal increasing sequence, there is an element s of I such
that r z s implies that

fkx(t) = liminfx(t), Yx(t) = limsupx(t). (13.1)


I2r it I-r it

(b) If I has a cofinal increasing sequence, there is a cofinal countable


subset J of 1 such that if is the restriction of to J, then

lim?nfx,(t) = lim1nfx(t), limsupx,(t) = lim1upx(t). (13.2)

In view of the discussion of the monotone case in Section 12 there is an


increasing sequence t, in I, cofinal if 1 has a cofinal increasing sequence, such
that

lim ,& x(t) = lim infx(t), lim Y x(t) = lim sup x(t)
A"W 12!1, it n-. ,,2,n it
and that ifs is an upper bound of t,, then (13.1) is true. If I has no cofinal
increasing sequence, there are (infinitely many) such points s, and we are
done. If I has a cofinal increasing sequence, the sequence t, is cofinal, and
we define 1 = {t: t 5 t < Let I be a countable subset of I., chosen
so that has the same infimum and supremum on I as on I,,. The set
J = Uo I satisfies (13.2).
Appendix IV

Lattice Theoretic Concepts in Measure Theory

1. Lattices of Set Algebras


Recall that a [a] algebra of subsets of a set X is a nonempty class of subsets
containing complements, finite [countable] unions, and finite [countable]
intersections of its members. The intersection of an arbitrary collection of
[a] algebras is a [a] algebra. The smallest [a] algebra containing a class IF
of subsets of X, that is, the intersection of the collection of all [a] algebras
containing r, is the [a] algebra generated by T. In particular, the generated
[a] algebra is called countably generated if r is countable. If to each point a
of some set I corresponds a collection .9a of subsets of X, we denote by
or {UaE,.a} the a algebra generated by UaE,.F. It will
frequently be true in this book that each collection .tea is a a algebra and that
.F is directed upwards in the sense that if a and fi are in I, there is a point
y in I such that .tea u .Ffl c .F,. In this case UQE,.Fa is an algebra but need
not be a a algebra.
The class 2' of a algebras of subsets of X ordered by inclusion (c) is a
complete lattice in which if {.y, a E 1} is a collection of these a algebras,

Y ,F. = aE I}, ,& ,a = n -,-wa.

QE1 ail a E I

Observe that to each set A in Y.E, .tea corresponds a countable subset J of I,


depending on A, such that A E Y. E,, .tea, because the class of sets A with this
property is a a algebra containing Ua.,.a but contained in and therefore
equal to Y E, .Fa. Observe also that the class of sub a algebras of a given a
algebra is a complete sublattice of Y. We omit the corresponding remarks
on the lattice of algebras of subsets of X.

2. Measurable Spaces and Measurable Functions


Recall that a measurable space is a set X coupled with a distinguished a
algebra .9t" of its subsets. In this book if X is endowed with a topology, then
unless some other choice is specified the class 9t" will be the class R (T) of
768 A IV Lattice Theoretic Concepts in Measure Theory

Borel subsets of X, that is, the a algebra generated by the class of open
subsets of X. If (X, X) and (Y, 1) are measurable spaces, a function 0 from
X into Y is defined as measurable if 4 '(S() c .Q', and for this it is sufficient
that O'(A) E- 5C for each set A in a subclass of N which generates the a
algebra V. If 01 is a measurable function from (X,1) into (Y,?J) and if 02
is a measurable function from (Y,g) into a third measurable space (Z,.'),
then the composed function ¢2(¢,) is measurable from (X, X) into (Z,.7).
If X is a set, if .tea is a a algebra of subsets of X for each point a of some
index set 1, and if ¢ is a measurable function from (X, Y., I 9a) into a
measurable space (Y, 1), then if 'J is countably generated, there is a count-
able subset J of / such that 0 is measurable from (X, YE J .tea) into (Y, 9). In
fact it is sufficient to show that for each set A in a countable generating class
for Y the set 0-'(A) is in the a algebra generated by a countable subclass of
.. ., and we have already remarked in Section 1 that this is true.
Suppose that to each point t of a set I corresponds a measurable space
(X, i). The product a algebra X, 1, is defined as the a algebra of subsets
of the product space x,,, X, generated by the class of product sets x,,, A,
with A, = X, for every value of t with one possible exception and for that
value A, e..G,. The product measurable space is defined as (x,E, X, X,, , X,).
If (X, X) is a measurable space and if (X, Y,) is a replica of (X,.7) for every
t in a set 1, we denote the product measurable space by (X, V), or by
(X`, ik) if 1 has cardinality k.

3. Composition of Functions
If y is a function from a space X into a measurable space (Y, 'J), the smallest
a algebra F of subsets of X making _y measurable from (X, .F) into (Y, ')
will be denoted by .f {y}. That is, ..{y} = y-'(N) is the class of subsets of
X of the form l y e A} with A in I'. If Qi is a measurable function from (Y, IN)
into a measurable space (Z, .'), the composed function z = 4)(y) is measur-
able from (X,.F{y}) into (Z,-7). Conversely, if z is a measurable function
from (X, .F{ y}) into (Z, .'), it is a useful fact that z must have this form
4)(y) under suitable restrictions on the measurable space (Z, .7). The follow-
ing lemma, stated in the preceding context, gives one such restriction.

Lemma. Let (Z, .7) be a Polish space coupled with its class of Borel sets. Then
if z is a measurable function from (X, 9 (y)) into (Z, -7), there is a measurable
function 0 from (Y, I/) into (Z, 7) such that z = 0(y).

We can suppose that Z has been assigned a metric making the space
complete and separable. To prove the lemma, suppose first that z has as
range a sequence C. of points of Z. Then z-' ({,) e.F{y} ; so there is a (perhaps
not uniquely determined) set B; in ' such that z-'(C;) = y-'(B,). If we
replace Bo, B, , ... by Bo, B, - Bo, B, - (Bp u B,), ... if necessary, we can
4. The Measure Lattice of a Measurable Space 769

suppose that the sequence B. is disjoint, and we define 0 on Y as r;, on B,


and 0 _- const (an arbitrary point of Z) on the remaining set Y - U so B.
This function 4) has the desired properties. For arbitrary z there is a conver-
gent sequence z, of countably valued measurable functions from (X,.F{y})
into (Z, 3°) with limit z. For example, for each positive integer n choose a
partition B"o, B",, ... of Z into disjoint nonempty Borel sets of diameter
S 2-", choose a point Cj in B. j, and define z" on X as C", on z-' (B"J). Then
by the lemma in the special case just treated there is a measurable function
¢" from (Y, 'W) into (Z, .) such that z" = ¢"(y), and it follows that the
sequence ¢, is convergent on the range of y. Since the convergence set of
0, is in 31, if we redefine 4" = const on the divergence set (the same point of
Z for all n), the sequence is convergent on Y to a function 0 with the
required properties.
Application to Product Spaces. (Notation of the Beginning of This Section.)
If k is a strictly positive integer and ifyk = [y( I ), . . . , y(k)] is a function from
X into Yk, then .F{yk}, usually written .F {y(1), ... , y(k)}, becomes the class
of subsets of X of the form { [y(1), ... , y(k)] e Ak} with A k c- c8fk. Observe
that .F{yk} is the smallest a algebra .F of subsets of X making every function
y(i) measurable from (X, i) into (Y, J). More generally, if I is an arbitrary
nonempty set and if { y(t), t E I) is a family of functions from X into Y,
indexed by I, let y(t, ) be the value at in X of the function y(t). The func-
tion y,: - y ( , ) is a measurable function from (X,.F{y,}) into (Y', IN'),
and .Fly,), usually written ,F{y(t),tel}, can also be described as the
smallest a algebra 9 of subsets of X making every function y(t) measurable
from (X, .F) into (Y, 6Y); that is,
F{y(t),IE1) = y.9t{y(t)}.
(El

According to Section 1, a set A in . { y(t), t E 1 } is in F{ y(t), t E J } for some


countable subset J of 1, depending on A. It follows that if Y is countably
generated, a measurable function y from (X, F { y(t), t E l }) into (Y, V) is
measurable from (X, . { y(t), t E J }) into (Y, Y) for some countable subset J
of I depending on y.

4. The Measure Lattice of a Measurable Space


Let (X, 97) be a measurable space, and let K+ be the class of measures on :1,
ordered by setting p v if p 5 v on i, that is, if p(A) 5 v(A) for every A in
T. From now on all subsets of X mentioned are in 97, and this qualification
will no longer be stated. The space K+ in this order is a lattice with

(p Y v)(A) = sup { p(B) + v(A - B): B c A}


(4 1)
(p A v)(A) = inf {p(B) + v(A - B): B c A).
770 A.W. Lattice Theoretic Concepts in Measure Theory

nJ
The only nontrivial part of this assertion is the countable additivity of the
set functions defined in (4.1). If A(A) is the right-hand side of the first line
and if A = U 'o AJ (disjoint union),

A(A)=sup{E [p(Bj)+v(Aj-By):BcA,Bj:BAJ}
(4.2)
M 00

_ Y sup { p(Bj) + v(Aj - B): BJ c Aj} _ Y A(A).


j=o J=o

Thus A is a measure, and an analogous argument shows that the second line
of (4.1) defines a measure.
The class ...r has the identically vanishing measure as order infimum and
the measure assigning + oo to each nonempty set as order supremum. More-
over the lattice /l+ is a complete lattice. In fact we now prove that an
arbitrary subset { pa, a e 1) of ..K+ has an order supremum (and therefore
also an order infimum). In proving that the supremum exists it can be
supposed that p, contains every supremum of finitely many of its elements
because adjoining these suprema does not change the conditions for an
overall supremum. Define p on . by p(A) = sup,E f p,(A). It is trivial that
p= j pQ once it is shown that p is a measure. If A = Uo Aj (disjoint
union),
no

p,(A) _ Y µa(Aj) 5 Y u(A) (4.3)


J=0 j=o

for all a; so p is countably subadditive. Finite additivity is trivial and implies


n m n
p(A) _ p(A,) + p AJ z p(A). (4.4)
J=0 j=n+1 j=0

so (n -+ co) p is countably superadditive as well as subadditive and therefore


is a measure.

Support of a Measure

A support of a measure p on (X, ') is a set with the property that its comple-
ment is p null. In some contexts uniqueness can be attained by a further
condition on a support. For example, if p is a measure of Borel subsets of
an open subset X of G8", finite on compact subsets, then p has a smallest
support closed relative to X.

Orthogonal Measures

If p and v are measures on (X, 1), they are orthogonal if p A v = 0, that is,
if (p A v) (X) = 0. This condition is satisfied if and only if the measures have
5. The a Finite Measure Lattice of a Measurable Space (Notation of Section 4) 771

disjoint supports. In fact the latter condition is obviously sufficient for


orthogonality. Conversely, if the measures are orthogonal, it follows from
(4.1) with A = X that for each positive integer n there is a set R such that
p(Bn) + v(X - Bn) < 2-n. If A = lim sup.-. B. (= n fl , Um=n B,n), then
p(A) = v(X - A) = 0; sop and v have the respective supports X - A and A.

Almost Everywhere-Almost Surely

If ).E.K+, it will sometimes be convenient to follow probability usage and


write "A almost surely" instead of "A almost everywhere."

5. The a Finite Measure Lattice of a Measurable Space


(Notation of Section 4)
Let .#' be the class of a finite measures on (X, .2"), in the order inherited
from !!+. Then . KQ is a cone, ordered in its specific order, and is a condi-
tionally complete lattice, a sublattice of

Theorem. The lattice K,+ has the countability property; that is if { p2, ae I} is
an upper-bounded subset of .,ll;, there is a countable subset J of I such that
YYEJp2 - YQEIp3

It can be assumed in the proof that p, contains every supremum of finitely


many of its elements since these suprema can be adjoined to p. if necessary
to achieve this, without altering the generality of the context. It can also be
assumed that p,(X) is bounded because if v is an upper bound of p, in
there is a disjoint sequence A. of sets of finite v measure, with union X, so
that p,(A;) 5 v(A) < + oo, and if the theorem is known to be true for the
projections on A, of the measures involved, the theorem follows as stated.
Under these assumptions choose index values a, a2, ... in such a way that
&2 5 p2, <- and that limn-. p2,(X) = sup2E, p2(X). Define p = Yn2, pen
that is, p(A) = limn-. p2n(A) for every set A in Jr. Then p 5 YSEIpa In the
other direction observe first that for all a,

p Y P. = (Y /
n21
lean f " i 2= Y (pen Y P2)
n21

Hence by definition of p,

lim (pan Y p2)(X) = p(X)


(p Y P.)(X) = n_ .D

The measure y Y p2 - p has value 0 on X and is therefore the zero measure;


that is, p 2- p2, and so p ;?! Y,, l u,. Hence there is equality, and we have
proved that the supremum in question is the supremum of the sequence p2 .
772 A.IV. Lattice Theoretic Concepts in Measure Theory

6. The Hahn and Jordan Decompositions


Let % be a signed measure on a measurable space (X, X). In the following all
sets mentioned are in I. A Hahn decomposition of X for A is a partition of
X into a positivity set A+ on whose subsets A > 0 and a negativity set A- =
X - A+ on whose subsets ). < 0. The signed measure A can be expressed as
the difference between two measures (Jordan decomposition). For example,

A. = p - v. p(A) = A(A n A' ), v(A) = -A(A n A- ).

This particular Jordan decomposition is minimal in the sense that p < p' and
v < v' whenever p' - v' is a Jordan decomposition of A. Conversely, if u and
v are measures on X, their difference p - v need not be a signed measure,
but if both p and v are a finite, there is an increasing sequence B. of sets
with union A 'such that both u and v are finite on every set BB ; so the restric-
tions of p and r to the class of subsets of B, are finite. and we can therefore
build a set A' from the positivity sets of these restrictions such that p > v
on the subsets of A+ and p < v on the subsets of A- = X - A'. In this way
we have found what amounts to a Hahn decomposition of X for p - v and
thereby found a Jordan decomposition of this difference, although the
difference may not be defined on all T sets. In Section 7 such differences will
be rigorously defined and ordered to obtain a vector lattice.

7. The Vector Lattice #.


According to Sections 111.3 to 111.5, the cone ,4Y can be identified with the
positive cone of a conditionally complete vector lattice ,,4, consisting of
pairs (p,. p2) of members of.A//; , with the convention that (p,, p2) = (v,, v2)
when p, + v, = p2 + vi. and with the ordering ( P , when
p, + v2 t 92 +v , . Each member p of A has a minimal representation
(p ` , p-) in which the component measures are orthogonal. that is. have
disjoint supports, and are minimal in the sense that if (p, , p2) is a representa-
tion of p, then p, Z p+ and p, >- p-. If p+ or p- is a finite measure, then p
can be identified with the signed measure p+ - p-. A set A is a support for
the element p = (p, , p2) of . KQ if there is a representation of p with both
components supported by A, equivalently, if A is a support for the measure
p, + 42. It follows that two elements of ..* are orthogonal if and only if
they have disjoint supports.

Charges

Whenever we deal with an extended real-valued countably additive set


function on a measurable space (X, 9C), it is to be understood when X has a
8. Absolute Continuity and Singularity 773

topology that . ' is either the class M(.') of Borel subsets of X or the perhaps
partial completion of .(9t") relative to some class of measures. Suppose now
that X is a locally compact second countable Hausdorff space. (In this book
X will in fact usually be an open subset of IB".) In this context it is to be
understood in the absence of a contrary statement that the set functions are
finite valued on compact sets. A member y of .,lf, for which 1µ I = µ+ +,u-
is finite valued on compact sets will be called a charge. If either µ+ or µ- is
finite valued, the charge can be identified with the signed measure µ+ -,U-,
in particular with the measure µ+ if µ- = 0. We shall not use the term
"Radon measure" in this book, in which "measure" always implies positivity.
The class of charges is a sublattice of ..lf, and is itself a conditionally complete
vector lattice with the countability property.
In the following sections it is to be understood that the measures and
charges discussed are defined on some measurable space, specified further
as needed by the context.

8. Absolute Continuity and Singularity


If A e.,#,+, an element µ of sV will be called absolutely continuous relative to
A if y has a representation (µ,,µZ) for which A null sets are also p, + µ2 null
sets; that is, in the usual terminology the measures µ, and P2 are absolutely
continuous relative to A. If so, µ, and 92 can be taken as the minimal choices
µ, = µ+, 42 = p. For any acceptable choice (µ,,µ2) there are unique up
to A null sets positive finite-valued measurable functions, which we shall
denote by dµ,/dt, dµ2/dl, for which

µ.(A) = d2di.

Je

Conversely, if f, andf2 are positive A almost everywhere finite-valued mea-


surable functions and if µ; is defined by

µ.(A) = J JdA, (8.1)


A

then (p,,µ2) is a A absolutely continuous member of K,. We shall write


dµ; = f dal for (8.1). If v = (v,, v2) with dv; = g; di is a second A absolutely
continuous member of -#,, then
dµ.
µYv=(p',µ') with dµ', = (ft + g2) v (f2 + 91) dA, = (f2 + 92)dA,

µ A v = (µa, µ') withdµi =(f, +g2) n (f2+g,)dl.

In particular, if it and v are positive, we can choosef2 g2 = 0.


774 A.IV. Lattice Theoretic Concepts in Measure Theory

An element p of A'7 will be called singular relative to A if p 1 A, that is,


if p and A have a pair of disjoint supports. The Lebesgue decomposition
theorem implies that each element of W. has a unique representation as the
sum of a A absolutely continuous element and a A-singular element of M,.
According to Section III.I 1 the positive elements in the band generated by
A are the elements of .,Ko of the form Yo (v A nA), with v in The
measure v k nA is majorized by nA and is therefore A absolutely continuous,
as is the limit measure y obtained when n oo. Conversely, if p is a A abso-
lutely continuous measure, the measure p A nA has the Radon-Nikodym
density (dp/d1) A n; so p = Yo (p A nA), and p is in the band generated by
:. Hence the band generated by A is the set of A absolutely continuous mem-
bers of .,#. and the Lebesgue decomposition is precisely the decomposition
of -+Y, into the orthogonal bands of A absolutely continuous and A-singular
elements. If (X, X) is a locally compact second countable Hausdorff space
coupled with its class of Borel sets, the lattice .,K, can be replaced in this
discussion by the lattice of charges.

9. Lattices of Measurable Functions on a Measure Space


Every concept in this section is relative to a specified complete measure A
in .,Il; . That is, we are discussing functions on a complete a finite measure
space (X. T, A). We conform to the usual custom that in discussing a function
from a measurable space into i "measurable" means that the range mea-
surable space is (IB,R(R)). Let .1( be the class of measurable functions
from (X, .() into A with the convention that a function need be defined only
A almost everywhere and that two functions are identified if they are equal
A almost everywhere. Strictly speaking A* is a space of equivalence classes
of functions, but we shall adhere to the usual convenient carelessness of
analysts in which "function" and "member of .,K * " may be either a function
or an equivalence class, as indicated by the context. The space M* is a
lattice under the order 5 to be interpreted as follows: if f and g are mea-
surable functions and if f and g are the respective equivalence classes con-
taining these functions, the inequality f S g means that f 5 g up to a A null
set. Similarly f v g [f A g] is the equivalence class containing f v g [f A g].
As already noted, we shall blur the distinction between function and equiva-
lence class.
The subclass KQ of A almost everywhere finite members of #* is a
vector sublattice of A' *. The set of positive members of A' * is not a cone.
but the set A; of positive members of .# is a cone and .!!; + -
0

Theorem. The lattice A' * is complete, and .,# is conditionally complete. Both
lattices have the countability property that the supremum [infunum] of a set,
iJ it exists, is the supremum [infimum] of a countable subset.
10. Order Convergence of Families of Measurable Functions 775

Since the transformation ft.-. arctan f is order preserving from &* into
..U; , it is sufficient to prove the assertion of the theorem for .,# and there-
fore for .,/!; ' . Each element f of . K; ' determines an element p f of .,#, by
way of du f = f d.1, and the map f " p f is an order isomorphism between . K;
and the lattice of A absolutely continuous elements of .11 . The theorem now
follows from the conditional completeness of #Q and Theorem 5.

Essential Order Terminology

The order of .I(* is called the essential order. If { fa, a E 1) is a subset of


..K*, the elements Vas,fa and ^e,fa are denoted, respectively, by
esssup.E,f. and ess inf,E,f.. Thus according to Theorem 9, there is a count-
table subset J of I such that supp5 jff zfa for every a neglecting a A null
set depending on a. If/is directed, ess lim supat f, is defined (Section III.12)
as

ess inf [ess sup fa],


PEI a2p

and the essential limit infimum is defined dually. A function f is called the
essential limit off if both the essential limit superior and essential limit
inferior are equal A almost everywhere to f.

10. Order Convergence of Families of Measurable Functions


Theorem. Let I be a linearly ordered set with no last element and with a cofinal
increasing sequence. Let be a function from I into the space -&, based
on a complete a finite measure space (X, X, A). Then there is a cojinal in-
creasing sequence I. for which

lim inf x(t) lim sup x(t)


R_.G It R-.0 tt
(10.1)

This theorem with I an open interval of R is useful in stochastic process


theory. To prove the theorem, assume [if necessary, replacing the given
measure A by a finite measure with the same null sets and replacing x(t) by
arctan x(t)] that A is finite and that 5 const ( < + oo). Let t; be a
cofinal increasing sequence in I, define I.= f1 :1 5 t S and choose
t;,o, ... , in I. in such a way that the functions minjsa. x(t.) and
maxJSa. x(Q are, respectively, at L' distance at most 2 -" from ess inf,.,^ x(t)
and ess sup, E, x(t). Then

lim [minx(t.j) - essinfx(t)] = lim [esssppx(t) - maxx(t;,)] = 0 a.s.


,,-.aG jaa t6I, R- W 9eI jsa
776 A.1 V. Lattice Theoretic Concepts in Measure Theory

because the sums of the integrals of the bracketed A almost everywhere posi-
tive functions converge. Hence (10.1) is true with t, the set {1,,j: n >_ 0, j < a j
arranged in increasing order.

Generalization

Let S be a compact metric space, let S be the space of measurable functions


from (X, .1', A) into (S, .4(S)). and let be a function from 1 into S. Theorem
10 can be applied to x(t) =f [y(t)] for J 'a real finite-valued continuous
function on S. In fact a slight extension of the proof of the theorem yields
the fact that there is a cofinal increasing sequence t, in I for which

limsupj'[ y(t)]
n-al rt

simultaneously for every function f in a countable dense subset of C(S), and


therefore simultaneously for every function in C(S). In particular, if I is
itself countable, "ess" can be omitted in (10.2), and it is then easy to conclude
that, neglecting a ;! null set, .v(t) has the same cluster values when t increases
through t, as through 1. This result is important in discussing stochastic
process separability. If it is known that exists A almost surely
whenever t is a cofinal increasing sequence in I, it follows that ess lim, t x(t)
exists and is A almost surely this sequential limit, which obviously does not
depend, neglecting null sets, on the choice of t,. If I is countable, it follows
that lim,t x(t) exists almost surely.

Fatou's Lemma and the Lebesgue Dominated Convergence Theorem

Let ,x(t), te!} be an upper-directed (in the essential order) family of mea-
surable functions from the a finite measure space (X, fit", A) into OB. The
ordering of the functions defines an ordering of I, and

esslimx(t) = esssupx(t) = supx(t) a.s.,


it Icl tE.I

where (by Theorem 9) J is some countable subset of I and (from Section


111.12) we can choose J to be an increasing sequence t,, cofinal if 1 has a
countable cofinal set, so that

esssupx(t) = lim a.s.


idI a-m

[If 1 does not have a countable cofinal set, this result is trivial because
according to Theorem 111.12 there is a point t' of I such that x(t) = x(t')
almost surely whenever t >_ t'.] If ess inf,E, x(t) >_ 0, then
1. Measures on Polish Spaces 777

fx ess sup x(t) dA = lim fx x(t) d A.


1cl it
Now suppose that is a function from a directed set I into the space
of measurable functions from (X, X, A) into R. Then if ess inf c, fp >- 0, the
preceding monotone limit result yields Fatou's lemma in the present context:

f ess lim inffp dA = lim f ess inf fp dA. < lim inf fX f dA.
X
pt at x /Za at

D rop the positivity hypothesis and apply the usual argument using Fatou's
lemma to prove the Lebesgue dominated convergence theorem in the present
context: if ess limpt fp =f exists and if ess suppE, fpl is A integrable, then

f fdA=lim f fpdA.
X
Pt x

11. Measures on Polish Spaces


Whenever a measure on a Polish space X is discussed in this book, it is to be
understood that the domain of the measure is mod" = -V(X) or, if so stated,
the completion of M(X) with respect to a specified measure or family of
measures. It is an important property of Polish spaces that a finite measure
A on 9t" (and therefore also a a finite measure on T) is inner regular in the
sense that for A in i

A(A) = sup {A(B): B c A, B compact}. (11.1)

We sketch the proof for A a finite measure. In the first place (11.1) is
true if "compact" is replaced by "closed" because the class of Borel sets A
for which (11.1) as so modified is true includes the closed sets and is closed
under countable unions and intersections, and is therefore .4(X). Thus it is
sufficient to verify (11. I) as stated for closed sets A. Let A be a closed set,
and let a be a strictly positive number. In some metric for X let A. be a
sequence of nonempty closed subsets of A, of diameter < 1, with union A.
Define Bo = A and B, = Ui=o At, where n is so large that A(B,) > A(Bo) - e.
Continue by induction: if Bk has been defined as a closed subset of Bk_,
with A(Bk) > A(A) - t, we can use the preceding argument to find Bk+,, a
closed subset of Bk, the union of finitely many closed sets of diameter
< 1/(k + 1), with A(Bk+,) > A(A) - E. The set B = nk o Bk is a closed set
which can be covered by finitely many sets of diameter < 1 /k for every k > 0;
so B is compact. Moreover A(B) z A(A) - E. Hence (11.1) is true.
This theorem implies that for any a finite Borel measure p on a Polish
778 A.IV Lattice Theoretic Concepts in Measure Theory

space the class of continuous functions with compact support on the space
is a dense subset of L' (µ).

12. Derivates of Measures


Let p and v be charges on (E8" with v >_ 0, and lets be a point of RN. If S is
a closed convex subset of R' containing i;, let d'(S) be the distance from 5
to OS, and let d"(S) be the diameter of S. A number q will be called the
convex derivate at s of p with respect to v if whenever S. is a sequence of
closed convex sets containing C for which

d"(Sn) > 0, lim 0, lim inf d, , > 0. (12.1)


n-a

it follows that v(Sn) > 0 for large n and that

lim u(Sn) q (12.2)


n- 7 V(S1) -

In particular, if (12.2) is required only when each set S. is a closed ball with
center 5, the derivate is called the symmetric derivate. The convex derivate
exists at v almost every point C, and the derivate function is v almost surely
the Radon Nikodym derivative dp/dv. The latter notation will also be used
for the convex and symmetric derivates, with the intended interpretation
always specified.
If there is a finite number q such that

0,
d1/`v
d

where the derivate is in the convex [symmetric] sense, then (dµ/dv)(i() = q


in the same sense, and we shall then say that the derivate is q in the convex
[symmetric] variational sense. We outline the proof, standard when N = 1
and v = /, , that dp/dv exists v almost everywhere on U8" in the convex varia-
tional sense. If the function dp/dv (convex derivate) is denoted by J'and if q
is any real number,

dtp - qvl
dv
° 1.I-gj (12.3)

t' almost surely; so (12.3) is true v almost surely simultaneously for all
rational q, and it follows from an easy continuity argument that (12.3) is
true almost surely simultaneously for all q. If C is not in the exceptional set
for the last statement, set q = J(C) to complete the proof.
Appendix V

Uniform Integrability

This Appendix. consisting of only one section, lists for easy reference what
is needed in this book of the uniform integrability concept.
It will be convenient to call a function 1 from R* to 68+ a uniform
integrabilityy test function if b is monotone increasing and convex and if
lim,_m [O(s)/s] = + x.
Let (X,.£',A) be a measure space with i.(X) < +x. All functions on X
in the discussion below are measurable from (X, I") into (l . (OB)) unless
specified otherwise. Recall that a family , j,, I E I) of such functions is said
to be uniformly integrable (relative to i.) if

limsup lJ,Idi.=0.
n-x tel
f1UP"I
The following uniform integrability properties are used in this book. Except
in (d), all functions of the family are in L'(;.).
(a) The family.f is uniformly integrable if and only if it is L' bounded
and if the set function family ; A F-. f A Jj; J(1i., t E i) is uniformly absolutely
continuous relative to i. ; more formally J; is uniformly integrable if and
only if

sup I Jill di.<+x, and lim sup I


ill)-O te/ ,A
di.=0.
'CI ,

(b) The family], is uniformly integrable if and only if there is a uniform


integrability test function 0 such that sup,.If, 0(I J,I) rli < + X.
(c) A i, almost everywhere convergent sequence of functions on X is L'
convergent if and only if the sequence is uniformly integrable.
(d) 1fJ; is a sequence of positive functions on X, then (by Fatou's lemma)

lim inf I f di. >_ J5l1n-x

In particular, if limn-. fn = f i. almost surely, if each fn is integrable, and if


780 A.V. Uniform Integrability

lim f f" dal = fXfdA < + oc,


"-m x
then the sequence f is uniformly integrable.
(e) A uniformly integrable sequence f, on X is weakly sequentially com-
pact; that is, there is a subsequence fa. and a A integrable function f on X
such that

lim fX f,"g di = lx fg dl
"-aD

for every bounded function g on X.

Generalization

Let (X, ') be a measurable space, and let I be an arbitrary nonempty set.
Suppose that for each t in I there is a function f, on X and a finite measure
A, on 1'. The family is said to be uniformly integrable relative to A. if

sup,u,(X) < + co, and lim supj If I dA, = 0.


re/ "-'00 iEl IU4I>")

The obvious translations of (a) and (b) into the present context are valid.
Appendix VI

Kernels and Transition Functions

1. Kernels
Let (X, £C') and (Y, 4.1) be measurable spaces. A kernel from the first space
into the second is defined as a function from X x Y into G'8+ with the follow-
ing properties:
(KI) A) is X measurable when AE*.
(K2) p(l;, ) is a measure on W when e X.
The kernel is called substochastic if I and stochastic if p(-, V) __ 1.
If (X, X) = (Y, W), the kernel is said to have state space (X, 1).

Kernels Given by Densities


If q is a measurable function from (X x Y, T X'3/) into (98+,.(l)) and if d
is a measure on Y, define

A) = q)).(dq) ( e X, A E 'J)
JA

to obtain a kernel p with density q relative to A.

Notation
If p is a kernel from (X, f) into (Y, IN) and if f is a measurable function from
(Y, IN) into (U8,.V(fi)), we write p(, f) for j,. dry) when this integral is
well defined, for example, when f >_ 0 and f is Y measurable. If p has density
q relative to a specified measure, we also write f) for p(-, f ).

Extension of a Substochastic Kernel to a Stochastic Kernel

Let p be a substochastic kernel with state space (X, 5C'). This kernel can be
extended to be stochastic as follows. Adjoin a trap point (also called a
cemetery) 8 to obtain an enlarged space X' = Xu {8}, and define F as the
782 A.V1. Kernels and Transition Functions

a algebra of subsets of X' generated by d and the singleton {e}. A stochastic


kernel p' with state space (X'. X') is then determined by setting p' = p on
Xx.1"and
I if EX,
p' {8}) =
1 ifg = a.

This kernel is the desired extension of p.

2. Universally Measurable Extension of a Kernel


If p is a kernel from (X..'1.") into ( Y,4) and if the function
X) is bounded,
we can extend each measure by partial completion to the a algebra
4'(V) of universally measurable sets over V. Under this extension p becomes
a kernel from (X, V(d" )) into (Y, Pl(w1)). To see this, let u be the completion
of a finite measure on .f, suppose that A e'W(NJ), define a measure v on .( by
!x p(', )p(dd), and extend v by partial completion to the a algebra
*(IN). Since A is v measurable, there are sets A 1, A 2 in ' with A, c A c A2
and v(A 1) = v(A 2). This equality implies that A,) = A) = A2) it
almost everywhere on X; that is, p measurable for every t, as was
to be proved. This argument shows that the definition of v(A) for A in i?I
remains valid for A in ' /(4').
It is important for the application to probability that we have actually
proved more than was asserted. In fact let be the a algebra of those
subsets of Y which, for every choice of u, are in the completion of the measure
v defined above. This a algebra includes *(ft and its sets will be called the
universally measurable sets over ' relative to p. What we have proved is that
if is extended by partial completion to this a algebra, then A) is
universally measurable, and the definition of v remains valid.

3. Transition Functions
Let I be a linearly ordered set, let (X, ') be a measurable space, and suppose
that to each pair (r, s) of points of I with r < s corresponds a stochastic
kernel p(r, , s, ) with state space (X, 9C') such that for A in *' and r1 < r2 < r3
in I the Chapman-Kolmogorov equation

p(r1, , r2, A) = I p(r2,17, r3, A)p(rt , - r2, dtl) (3.1)


X

is satisfied. Such a function p is called a Markov transition Junction, or simply


a transition Junction. In some contexts it is necessary to allow the kernels to
be substochastic or even infinite valued, but any divergence from the above
3. Transition Functions 783

definition will be specified explicitly. The parameter set is usually either the
set of strictly positive integers discrete parameter case or the interval ]0, + 00[
(continuous parameter case).

Discrete Parameter Stationary Case

Let (X, _f) be a measurable space, let p(1, , ) be a kernel with this state
space, and define inductively a sequence of kernels with this state space by

p(n + 1, , A) = I p(1, it. A)p(n, , dq), n > 1. (3.2)


x

Then

p(r + s, , A) = J p(r, ry, A)p(s, , dq). (3.3)


x

Ifp(l, , ) is a stochastic kernel, the function (m, , n, A) -. p(n - m, , A) for


0 < m < n is a transition function with parameter set the set of strictly
positive integers, but in this context it should cause no confusion if the
function (n, S, A) i-. p(n, S, A) for n > 0 is called a stationary transition func-
tion or temporally homogeneous transition function and if (3.3) is identified
as the relevant Chapman-Kolmogorov equation.
If the given kernel is substochastic, (3.3) yields a substochastic
transition function p. Let p'(1, , ) be the extension of the substochastic
kernel p(1, , ) to be stochastic by way of a trap point. Then the inductive
definition (3.2) applied to p'(1, , ) leads to a (stochastic) transition function
p', and p'(n, ) is for each strictly positive integer n the extension of p(n, , )
to be stochastic by way of the same trap point.

Continuous Parameter Stationary Case

Let (X, ") be a measurable space, and let p be a function from ]0, + cc [ x
X x .£ into R with the following properties:
(a) For each t in ]0, + o t,[ the function p(t, , ) is a kernel with state
space (X, X).
(b) The Chapman-Kolmogorov equation (3.3) (with r and s now in
]0, +oo[) is satisfied.
If each kernel p(t, , ) is stochastic, the function (r, , .s, A) r-. p(s - r, , A)
for 0 < r < s is a transition function with parameter set ]0, + oo[, but as in
the discrete parameter case the function p itself is called a stationary or
temporally homogeneous transition function. If the kernels are allowed to
be substochastic, in which case p is a substochastic transition function, a
784 A.VI. Kernels and Transition Functions

single trap point can be adjoined to X to make each kernel p(t, ,) stochastic
and thereby extend p to be a (stochastic) transition function.

Universally Measurable Extensions


In the discrete parameter stationary case it is trivial that if p(l, , X) in (3.2)
is bounded and if p (I, , ) is the universally measurable extension ofp(l, , ),
that is (Section 2), the extension of this kernel to the state space (X, 'W (52)),
then the nth iterate p (n, , ) of p(1, , ) in is
p is the extension of the state
the
be proved. this if
for t p (t, , ) is the the kernel
state space (X, is
by p,,. To verify (3.3) forp with A in V(52'), fix r, s, and , and choose
At and A2 in .t to satisfy

A, c A c A 2, p(r + s, , A t) = p(r + s, , A2).

Then

p(r+s..A)=p(r+s,.A1)= f jP(s ,, A1)p(r, , d,), i= 1, 2;


x
(3.4)

so (3.4) is valid when A, is replaced in the integrand by A, as was to be proved.


In this context if we strengthen the hypotheses on p by supposing that for A
in 52' the function p(t, -,A) is R(]0, + oo[) X X measurable, then if f is a
Lebesgue measurable function from 18+ into 98+, the function

(,A)-' f(t)p(t,,A)!t(dt)
0

is a kernel with state space (X, i) and so (Section 2) when considered on


X x x!(i) is a kernel with that state space, and therefore the above integral
defines a W(52) measurable function for fixed A in *(X). Functions defined
by integrals of this form are basic in probabilistic potential theory.
Appendix VII

Integral Limit Theorems

1. An Elementary Limit Theorem


Let X be a metric space, let A be a completed measure of Borel subsets of
X, and let k, be a sequence of A measurable functions from X into A. In
many contexts there is a point C of X with the property that A({t}) = 0 and
that for f a suitably restricted function from X into 08, continuous at C,

lim (1.1)
'-' Jx
The following theorem gives conditions that such a limit relation be true.
The conditions are phrased for ease in application rather than for maximal
generality.

Theorem. Suppose that k, and f satisfy the following conditions:


(a) Each function k is positive, with Ix k dA = 1.
(b) The function f is A measurable, and if A is an open neighborhood of
then Jx_e f I dA = JX_Ak dA = 0.
Under these conditions

lim sup ` kjdA < lim sup

If in addition to (b) the function f is continuous at C, then (1.1) is true.

If the right side of (1.2) is +oo, inequality (1.2) becomes trivial. Thus to
prove (1.2), which implies the last assertion of the theorem, it is sufficient to
show that if f _< a < + oo on a deleted open neighborhood A of C, then the
left side of (1.2) is at most a. We can suppose, adding a constant to f if
necessary, that a Z 0. Then

lim sup I k,,,jdA < x+ li m sup k.I f I dA = a,


x x_A

as was to be proved.
786 A.VII. Integral Limit Theorems

2. Ratio Integral Limit Theorems


For each positive integer n let k(n, ) be a positive Borel measurable function
on QRN. Let H and U be, respectively, a measure and a charge on QRN. Define

a = I' k(n, ?l)U(dq), b = Jkn?l)H(d1l). (2.1)


N

In many contexts k(n, q) is defined in such a way that

lim °" _
Al
a. b dH (0). (2.2)

We shall derive conditions on k(n, ) for (2.2) to be true. The following


remark will be useful. Suppose that H is fixed and that 0 < b < + x for
all n. The charge U will be allowed to vary in some specified linear class r
of charges, including H and including I UI with U, for which

k(n, q)I UI(dq) < + x.

Suppose it has been proved that (2.2) is true under the following special
conditions. t1 e F, U > 0. and the derivate in (2.2) is the symmetric derivate
and vanishes. It then follows that (2.2) is true for U in f whenever the
derivate on the right exists as a symmetric variational derivate (see Section
IV.12 for the derivate definitions). In fact if q is the value of this derivate.

J' k(n,rl)IU-gHI(dq).

and by hypothesis I i'% - qH I E F. and by definition of the symmetric varia-


tional derivate the symmetric derivate d I U - qHI /dH exists and vanishes at
the origin.
The preceding remark is of course also valid if the derivate in question is
the convex instead of the symmetric derivate.
In some contexts ratio integral limit theorems of the type described are
easily reduced to rather simple ratio integral limit theorems in one dimen-
sion. Such a one-dimensional ratio integral limit theorem is proved in the
next section.

3. A One-Dimensional Ratio Integral Limit Theorem


The following ratio integral limit theorem will be used to derive the desired
ratio limit theorem for the Poisson integral for harmonic functions on a ball.
Let k. be a sequence of positive Bore] measurable functions on R', and let
3. A One-Dimensional Ratio Integral Limit Theorem 787

a, $ be monotone increasing functions on 68+, vanishing at 0, strictly positive


on ]0, + oo [, and right continuous except possibly at 0. Define

a"= fo k" da, b"= f k"dfl. (3.1)


Jo

The following theorem stating conditions which imply that

lima-"=0
w b"

is phrased for ease in application rather than generality.

Theorem. Suppose that a", b" as defined by (3.1) are finite for all n and that
the following conditions, in which p is some strictly positive integer, are satisfied.
(a) k" is positive and monotone decreasing and has a continuous derivative
k;,.
(b) If a > 0,
fa
lim infk"k"(a) ' 1t (dr) = + oo.
o "°c
(c) There is a strictly positive function f on 68+ such that the function k"/f
is monotone decreasing for large n and that fo fda < +oo.

,in P
(d) liro P(r) 0, soup fi(r) < + oo.

Then (3.2) is true.

To prove the theorem, choose t strictly positive and so small that the
limit superior in (d) is strictly less than I/E. If a is so small that cc <- E3 on
[0, a], then
fa fa fa
k" da = a(a)k"(a) - ak;, d1 t < a(a)k"(a) - E f3k;, d1 t
0 0 o
(3.3)

= k"(a)[a(a) - E3(a)] + E fak"dfl < eb".


0

If a is so small that ErP < 1(r) for r in [0, a], then


ra ra ra

(3.4)

= k"(a) [#(a) - Ea"] + Ep Jka(r)rt'_1l1(dr) > E f " k"(r)rP-' l t (dr).


0
788 A.VII. Integral Limit Theorems

Combining these two inequalities, we find that for sufficiently small a,

j k dot
lim sup b" < e + lim sup
"-°°
o f " 1, (dr)

The numerator in the fraction on the right is at most

r" f f da 5 jfdrx.

[Incidentally this inequality shows that condition (c) implies the finiteness
of a..] In view of condition (b) the limit superior in (3.5) is 0, and therefore
(3.2) is true.

4. A Ratio Integral Limit Theorem Involving Convex


Variational Derivates
In the following theorem, phrased to be applicable to the desired ratio
boundary limit theorem for parabolic functions on a half-space, K(s, ) is
for each strictly positive s a positive Borel measurable function on R,
U is a charge on RN, H is a measure on 1 N, and u and h are defined on
RN x ]0, + oo[ by

u(, s) = JK(s. l - 7I) U(d7), s) = l - n!) H(d7). (4.1)


JRN

Theorem. Suppose that the following conditions are satisfied.


(a) For each s > 0 the function u, with U replaced by I U1, and the function
h are locally bounded on RN.
(b) For each s > 0 the function K(s, ) is a positive monotone decreasing
function on R+, with a continuous derivative, and K(s, 0) = 1.
(c) There is a strictly positive junction on ]0, + oo [ satisfying

lim b(s) = 0, lim inf K(s, b(s)) > 0.


r0
(d) If a > 0, there is a strictly positive number so = so(a) for which,
uniformly for r z a,

L0 K(s, r) _ 0.
K(so r)b(s)N -
4. A Ratio Integral Limit Theorem Involving Convex Variational Derivates 789

Then if at a point r; of 08N the derivate dlN/dH exists as a finite convex


derivate and dU/dH exists as a finite convex variational derivate, and
if 0 < c < 1, it follows that

u(g,s) dU (r)
lim sup = 0. (4.2)
s~o It-Slsca(s) h(l:,s) dH

Observe that (Section IV. 12) dU/dH and d1N/dH exist and are finite as convex
variational derivates at H almost every point C of RN.
For fixed H the class of charges U satisfying (a) is linear, contains I U1
with U, and contains H. Hence according to the argument in Section 2, it
is sufficient to prove that (4.2) is true for C when U is positive and both
dlN/dH and dU/dH exist as finite convex derivates with (dU/dH)(C) = 0. To
simplify the notation, we assume that C = 0. For r > 0 define r) and
(3(l;, r), respectively, as the U and H measures of r), and define a(l;, 0) _
0) = 0. By definition of the convex derivate

r).
lim sup cr = 0,
r-o W, r)
rN
limsup
r~o
I < cr < +00.
r)

To obtain a minorant of h, observe that

s) >t LAS)] K(s, r) d r) z K(s, b(s)) S(s)) (4.4)

so that

K(s r)da(i,r) < a(,S(s)) (4.5)


/;(l:, s) ,6(s))*
fro.6(s))

If a > 0 and if s is so small that b(s) < a,

K(s, r) da(l;, r) = K(s, a) - K(s, 8(s))


(4.6)

K'(s, r)l, (dr).


)a(s).al

If now e > 0 and if a is so small that a(g, r) < r) when I :!5; cr and 1

r S a, then the right side of (4.6) is majorized when K'a is replaced by eK'Q,
and integration by parts yields
790 A VII. Integral Limit Theorems

K(s, r) r) < K(s, a) a) - e fl(g, a)]


I
K(s, b(s)) [a(S, b(s)) - b(s))]

+e K(s, r) r)
>aw).a1

<

Thus for sufficiently small a and I I < cb(s),

da(, r) <
K(s. r) + (4.8)
s) K(s, d(s))

Finally, if so is chosen as in condition (d) of the theorem, if e' > 0, and then
if s is sufficiently small,

l[
N
K(s, r) ha(S, r) < e' K(so, r) r) b(s)
I M.+ao[
S)
h
<
K(s, b(s))

Now u//i is the sum of the integrals on the left in (4.5), (4.8), and (4.9). The
first of these integrals tends to 0 with s when I I < cb(s) by (4.3). The second
can be made arbitrarily small in the limit (s 0) by choosing a small and then
choosing a small. For such a choice of a the right side of (4.9) is arbitrarily
small with c' when I 1 <- cb(s). Hence Theorem 4 is true.
Appendix VIII

Lower Semicontinuous Functions

1. The Lower Semicontinuous Smoothing of a Function


Let H be a Hausdorff space, and if S E H, let N(1;) be the set of neighborhoods
of . Recall that a function u from H into k is lower semicontinuous if and
only if the set {u > c} is open for c in It If u is not necessarily lower semi-
continuous, the lower function a of u, defined by

sup inf u(,) = A lim infu(ry), (1.1)


qeA q--{

is lower semicontinuous, is a minorant of u, and majorizes every lower


semicontinuous minorant of u. The function u will be called the lower
semicontinuous smoothing of u in this book. Finally recall that the supremum
u of a family {up, f e 1 } of lower semicontinuous functions is lower semi-
continuous because
{u > c} = U {up > c} (1.2)
PC/

and the sets on the right are open.

2. Suprema of Families of Lower Semicontinuous Functions


Theorem. Let { u6. PE 1 } be a family of lower semicontinuous functions from a
second countable Hausdorff space into R. and if J c I, define u' = supp, j up.
Then u' is lower .semicontinuous, and there is a countable subset J of 1 such
that tt' = u'. In particular, if u, is directed upward, there is an increasing
sequence up with limit u'.

We have already noted in Section 1 that u' is lower semicontinuous.


Since the domain of definition of the functions is second countable, (1.2)
implies that there is a countable subset J of I such that

{u' > c} U {up > c}


PC I
792 A.VIII. Lower Semicontinuous Functions

simultaneously for every rational c and therefore for all c. Then u' is lower
semicontinuous, u' < u', and {u' > c} = {u' > c} for all c. Hence {u' z c} _
{u' >_ c} also, and subtraction yields {u' = c} = {u' = c}; so u' = u'. The
second assertion of the theorem is left to the reader.

3. Choquet Topological Lemma


Lemma. Let {up, 13 e I } be a family of functions from a second countable
Hausdorff space into U8, and if J c I, define u' = infs. j up. Then there is a
countable subset J of I such that u' = u'. In particular, if I is directed down-
ward, there is a decreasing sequence up. with limit v such that v = u'.

Let H be the given second countable Hausdorff space, and let H, be a


sequence of open subsets of H containing each set of a countable base for
the topology of H infinitely often. After replacing up by arctan up if neces-
sary, we can assume that the family u, is uniformly bounded. For n z 1
choose in H. to satisfy u'(,) 1/n. There is a point P.
in I for which 5 1/n. Set J = f,. Then

inf u'(q) 5 inf u'(ry) + ?


"EH nen n

and it follows that u' 5 u'. The reverse inequality is trivial; so there is
equality. The second assertion of the lemma is left to the reader.
Historical Notes

The modernization of classical potential theory during the last fifty years
has proceeded for the most part in small steps, many of which were more
and more rigorous derivations in more and more general contexts of facts
stated years earlier. Thus it is certain that the following notes, which give
reference to the "originators" of the more important advances, will appear
to knowledgeable readers to be inaccurate and unfair. The historical notes
to the probability discussion are only slightly more reliable. It is trivial for
research mathematicians to discover mistakes in historical accounts of their
own subjects, especially when the accounts rely on publication dates and on
the references in research papers. Thus no one should have much confidence
in a detailed history of the development of a mathematical theory. The
difficulty is compounded by the fact that many theorems are proved by
several authors in more or less overlapping formulations and that (see the
many papers on the Martin boundary) authors are frequently unwilling to
read their colleagues' papers carefully enough to state for the record the
exact differences between the scopes of their papers. Thus the reader is
requested to read the following notes with a grain of salt and a lump of
tolerance.

Part 1
See Burkhardt and Meyer [1, 1900] and Lichtenstein [1, 1919] for reviews
of nineteenth and early twentieth century potential theory, and see Kellogg
[2, 1929] for potential theoryjust before its renaissance. See Brelot [11, 1952]
and [17, 1972] for more modern reviews. The books of Tsuji [1, 1959].
Helms [1, 1969], Brelot [15, 1969], Constantinescu and Cornea [1, 1972],
Landkof [l, 1972], and Wermer [1, 1974] on potential theory contain much
of the material in Part I but are organized differently with emphasis on
different aspects of the subject. The books of Murali Rao [4, 1977], Port
and Stone [1, 1978]. and Chung [2, 1981] derive their potential theory as
a by-product of probability.
794 Historical Notes

Chapter 1.1

It would be both futile and difficult to give a detailed historical background


to the work in Chapter I of Part 1, but note that the development in Section 8
culminating in (8.3) follows [Green I. 1828].

Chapter 1.11

,Section /. What was essentially the Poisson integral for N = 2 was


displayed by Poisson in 1820. In [Poisson 1, 1823]. he displayed his integral
for N = 3 in its present form and tried to prove the boundary limit part of
Theorem I for continuous boundary functions. Schwarz [1, 1872] gave the
first rigorous proof of Theorem I (for continuous boundary functions).
Sections 2 and 3. See Harnack [ 1, 1887] for his inequality and convergence
theorem, the latter of course only for monotone sequences. The fact that a
one-sided bounded harmonic function on R' is identically constant is due
to Bocher [I, 1903].
Section 4. F. Riesz [2, 1926; 3, 1930] inaugurated the systematic study of
superharmonic and subharmonic functions; the properties and applications
in this chapter are almost all due to him. See [Radii 1, 1937] for early
references and further details on these functions.
Section 10. Theorem 10 on the concavity of spherical averages of super-
harmonic functions was proved in [Riesz 2, 1926] in the dual form for
convexity of spherical averages of subharmonic functions.
Section 12. Lord Kelvin (= W. Thomson) introduced his transformation
in [Thomson I, 1847].
Section 14. The theorem that a positive harmonic function on a ball is
the Poisson integral of a measure has been called the "Herglotz theorem"
with a reference to [Herglotz 1, 1911 ], but this theorem was proved first
by F. Riesz [1, 1911]. and in his own paper Herglotz refers to the Riesz
paper for this result.
Section /5. The essential content of Theorem 15 is that u/h has nontan-
gential limit function almost everywhere (Mb measure) on the ball
boundary. Here the derivate dM /c/M,, has different possible definitions, and
the exceptional M,, null boundary subset depends on the definition chosen.
No distinction is made between these definitions in the following remarks.
Fatou [1, 1906] proved Theorem 15 for the case N = 2, h =_ I, and proofs
were given later for N > 2 with h = I by other mathematicians. The theorem
was proved by Doob [ 12. 1959] in the general case.
Section 16. Minimal harmonic functions were introduced into potential
theory by Martin [1. 1941].
Pan 1 795

Chapter 1.111

Section 1. F. Riesz [2, 1926] introduced least harmonic majorants of


subharmonic functions.
Section 3. See these notes to Chapter VI for the history of the Fundamental
Convergence Theorem.
Section 4. The reduction concept developed from the dual concept of
"extremization." If D is a Greenian subset of l", if u is a negative sub-
harmonic function on D, and if A is a subset of D, the extremization of u
on A was defined as the supremum of the class of negative subharmonic
functions majorized quasi everywhere on D - A by u. See Brelot [8, 1945]
for further details and references to earlier related work by him and A. F.
Monna. Brelot [13, 1956] extended work of Martin [1, 1941] by treating
reductions of positive superharmonic functions on certain subsets of D.
The treatment of reductions in this book extends the standard treatment by
allowing reductions on an arbitrary subset of the closure of D in an arbitrary
compactification.
Section 7. The natural order decomposition is apparently due to Moko-
bodzki. See Mokobodzki and Sibony [1, 1968] for the relation between the
natural order decomposition theorem and additivity of reductions in a very
general context.

Chapter l .I V

Cartan's paper [2, 1945] on potentials on special sets had an important


influence in the modernization of potential theory. See references below to
key results in this paper.
Sections 7 and 8. F. Riesz [3, 1930] derived the existence of the measure
associated with a superharmonic function and proved what is now known
as the Riesz Decomposition Theorem.

Chapter IN
Section 1. Brelot [5, 1941] introduced polar sets to take the place of sets
of inner capacity 0 (called inner polar sets in this book) as the negligible
sets of classical potential theory. Cartan proved in [2, 1945], although he
announced the result in [1, 1942], that the polar sets are the sets of capacity
0 in the capacity definition of Chapter 1.XIII.
Section 2. Choquet [2, 1957] showed, in a trivially different context, that
if A is a polar Ga subset of a Greenian subset D of I", then there is a measure
p supported by A such that A is the set of infinities of GDu. G. C. Evans had
proved the existence of such a potential for A polar and compact.
796 Historical Notes

Section 5. Theorem 5 is usually stated for A a set closed in D. but the


formulation in Theorem 5 is sometimes more convenient. The fact that a
bounded harmonic function on an open set D less a compact polar subset
has a harmonic extension to D was proved by Bouligand [1, 1926]. Vasilesco
[2. 1937] proved Theorem 5 for A compact, Brelot [5, 1941], for A closed in
D. The fact that a one sided-bounded harmonic function defined on an open
deleted neighborhood of a point must have the form cG(c, ) + h with h
harmonic on the full neighborhood goes back at least as far as [Bocher
1, 1903]. The less general fact that a bounded harmonic function defined
on an open deleted neighborhood of a point can be extended to be harmonic
on the full neighborhood was proved by Schwarz [l, 1872].
Section 6. Szego [1, 1924] proved that in RZ the complement of a compact
set A has a Green function if and only if the Robin constant of A (see Section
XII1.18 for the Robin constant) is strictly positive. This result is equivalent
to Theorem 6.
Section 8. The Evans- Vasilesco theorem references are Evans [2, 1935]
and Vasilesco [ I, 1935].
Section 10. The principle of the maximum is sometimes called the Maria-
Frostman principle of the maximum because Maria [ 1, 1934] and Frostman
[1. 1935] proved this special case of the domination principle (with D = UR"
and N > 2).

Chapter I . V1

Section 1. Szpilrajn [1, 1933] proved the Fundamental Convergence


Theorem in the weak version of Theorem 111.3 in the context of a locally L'
convergent sequence of superharmonic functions. Radd [1. 1937] remarked
that Szpilrajn's result implied Theorem 111.3 in the context of a monotone
decreasing locally lower bounded sequence of superharmonic functions.
(Both mathematicians actually phrased their results in the dual context, for
sequences of subharmonic functions.) Brelot [2. 1938] strengthened the
Szpilrajn Rado result by proving that the exceptional set ;u < u; is inner
polar. Cartan [2, 1945], but he announced the result in [1, 1942], proved
that this exceptional set is polar.
Section 2. In a series of notes in the Paris Comptes Rendus leading up to
his definitive "Theory of capacities" [1, 1955] Choquet analyzed set func-
tions satisfying strong subadditivity and related conditions and obtained
his capacitability theorem (Theorem 3 of Appendix II), which makes easy
the proof that an analytic inner polar set is polar.
Section 3. The reduction properties listed go back to Brelot's early work
except for the properties involving sets meeting OD and for property (o)
which corresponds to certain crossing inequalities in martingale theory
(see Section 2.111.22).
Part 1 797

Chapter 1.VII

Writers who describe GD as "the Green's function" should be condemned to


differentiate the Lebesgue's measure using the Radon-Nikodym's theorem.
Section 1. The older mathematical literature contains many attempts to
show that open sets of various types have Green functions. The modern
approach removes the difficulty by generalizing the definition of a Green
function. This trick shifts the problem to that of showing that the Green
function has certain desired properties, but some of the old smoothness
demands have disappeared. For example, it is no longer necessary to define
harmonic measure by means of a Green function, as was done in Section 1.8.

Chapter 1.VIII

See Vasilesco [3, 1938] for a survey of progress in the Dirichlet problem
just before Brelot's [3, 1939] definitive study of what is now called the
PWB method.
Section 1. The fact that relativization of a family of functions having an
average property (1.1) yields a family with a corresponding average property
(1.3) was pointed out in [Doob 9, 1958]. In a Markov process probabilistic
context this relativization corresponds to a modification of transition prob-
abilities (Section 2.VI.13) conditioning paths to go where the conditioning
function is large. The analysis of the Dirichlet problem on a Greenian subset
of R" provided with its Martin boundary in [Brelot 13, 1956] was essentially,
although not explicitly, for relative harmonic functions.
The Dirichlet problem for harmonic functions has been attacked in several
ways, but it was finally seen that each way involves two distinct problems.
(i) The first problem is to find, for a specified suitably restricted boundary
function f, certainly for an arbitrary finite-valued continuous boundary
function f, a perhaps generalized "solution" u1. The function u f is to be
harmonic, and if the given domain and the given boundary function are
sufficiently smooth, the function of is to have boundary limit function f;
that is, of is the solution of the Dirichlet problem in the classical sense.
(ii) If of exists and if a domain boundary point i; is specified, the second
problem is to find conditions on the domain and boundary function ensuring
that of has limit f(C) at C. The simple Zaremba [1, 1911] example [Section 2,
Example (a)] and the more discouraging Lebesgue [1, 1912] spine example
(Section 15) showed that the Dirichlet problem may not be solvable in the
classical sense even when the boundary is the Euclidean boundary and
the boundary function is finite valued and continuous. Poincar6 [1, 1890;
2, 1899], however, had shown that for a sufficiently smooth Euclidean
boundary the Dirichlet problem has a solution in the classical sense for
every finite-valued continuous boundary function. The problem of finding
Dirichlet solutions in more general contexts was solved by generalizing the
798 Historical Notes

definition of "Dirichlet solution." The PWB method of finding (generalized)


Dirichlet solutions is a development of a method devised by Perron [ 1, 1923]
(independently and almost simultaneously by Remak [I, 1924]) and per-
fected by Brelot [3, 1939]. Wiener [2, 1924; 3, 1924; 4, 1925] was the first
to assign a generalized Dirichlet solution to an arbitrary finite-valued
continuous Euclidean boundary function. He used two methods of which
one was an elaboration of the Perron method, but he did not obtain Brelot's
later definitive results because a miscalculation in [4, 1925] gave him a
supposed counterexample which prevented him from proceeding to the class
of bounded Borel measurable boundary functions. He showed, however,
that for the Euclidean boundary the finite-valued continuous boundary
functions are resolutive.
Lebesgue was perhaps the first to see clearly the distinction between
problems (i) and (ii) above; so it is not surprising that he coined the term
"regular boundary point." Kellogg [1, 1928] for N = 2 and Evans [1. 1933]
for N >_ 2 showed that every subset of strictly positive capacity of the
Euclidean boundary contains a regular boundary point; that is, the set of
irregular boundary points is inner polar. This fact implies that the F. set
of irregular Euclidean boundary points is polar.
Sections 12 to 14. Poincare [1, 1890] used the idea of what is now called
a barrier. Lebesgue [2, 1912] coined the term and used the existence of
barriers for a regular region to prove the convergence to the Dirichlet solution
(for a finite-valued continuous boundary function) of a certain sequence of
approximants. Lebesgue [3, 1924] found that for the existence of a barrier
at a Euclidean boundary point it is necessary and sufficient that the point
be regular. Bouligand [1, 1926] showed that it is sufficient for regularity
that the barrier be a weak one and proved the related result that a Euclidean
boundary point is regular if and only if the Green function of
the given domain D has limit 0 at the boundary point for every pole c
(equivalently, a single pole if D is connected).
Section 15. In the years when potential theory was finally conquering
the Dirichlet problem the cone regularity condition was ascribed to both
Poincare and Zaremba. Only Clio now remembers why Zaremba was later
given the sole credit; the condition will be given both names in this book.
Section 18. The extension of GD in Section 18 was analyzed by Brelot
[6, 1944]. See [Brelot 1, 1938], [M. Riesz 1, 1938], [Frostman 2, 1938],
and [Vallee Poussin 4, 1938] for earlier discussions of the relation between
GMor (for r superharmonic on a neighborhood of D) and µo(-, c), of the
kernel 6', and of related questions.
Section 19. The proof of Theorem 19(b4) follows that in [Murali Rao
3, 1974]. The result was proved by Brelot [6, 1944] in the course of his
treatment of potential theory on R' extended by a point at infinity.
Part I 799

Chapter I.IX

Everything in this chapter is in the folk lore of potential theory. and much
of it has appeared explicitly.
Sections 9 and 10. The classes of quasi-bounded and singular harmonic
functions were first discussed by Parreau [I, 1951].

Chapter I.X

The concept of sweeping (balayage) goes back at least as far as Gauss


[1, 1840]. There are two approaches to the subject. one by operations
on functions, now treated by Brelot's reduction technique, the other by
orthogonality-minimization techniques originated by Gauss but first made
rigorous by Frostman [1, 1935; 2, 1938] and Cartan [2, 1945; 3, 1946].
The first approach is stressed in this book. The following outline of Poincare's
method of solving the Dirichlet problem [1, 1890; 2, 1899] shows why his
method is identified with sweeping, but see <below> Vallee Poussin's remark
about Poincare's work. Let D be a bounded open subset of 08^, with a
smooth boundary. For it a C'-' function on a neighborhood of D. with
Au < 0 (so that u is superharmonic on a neighborhood of D). Poincare
evaluated what we now identify as GM Du by the method of proof of Theorem
III.1 and proved that under his smoothness hypotheses on D the function
GMDU was the desired Dirichlet solution for the boundary function UIPD.
If Au is not supposed strictly negative, Poincare reduced the problem to
the case already treated by writing it as the difference between two functions
with the above properties of it. Using this solution and an approximation
procedure Poincare solved the Dirichlet problem for an arbitrary continuous
boundary function. Thus the basis of Poincare's method was the construction
of GMDU for u superharmonic on D. His construction, in the language of
the proof of Theorem 111.1. was to apply operators of the form to with B
a ball relatively compact in D. For it superharmonic on D the operation
U F - TBu sweeps the Riesz measure for u out of B (observe that tau = R°-"
when u is positive on D. so that reduction notation is applicable), but
Poincare did not go on to analyze this aspect of his method. Vallee Poussin
in a series of papers including [l. 1932; 2, 1937: 3, 1938. 4. 1938] developed
sweeping further and in fact in [3. 1938] he asserted that he had coined the
word "balayage" and that (in spite of his own title for [I, 1932]) Poincare's
work had nothing to do with sweeping. Vallee Poussin concentrated on the
swept measures: so it was he who developed and applied the interpretation
of the sweeping kernel as harmonic measure. He applied sweeping to analyze
capacity, the role of sets of inner capacity 0, and many other topics, not
always including full details, obtaining among other results one equivalent
to Theorem XI.13. The full use of the reduction technique on functions had
to wait for the final form of the Fundamental Convergence Theorem, which
800 Historical Notes

was quickly applied to reductions by Brelot [8, 1945]. Most of Chapter X


first appeared in the latter paper.

Chapter 1.X1

See [Brelot 16, 1971] for further elaboration of the fine topology and its
applications.
Section 1. Brelot [4, 1940] defined thinness of a set at a point by X1(2.1).
Cartan pointed out in a letter to him that thinness could be interpreted by
means of what is now called the fine topology. For some years thereafter,
however, the fine topology was merely a tool for phrasing results elegantly;
in fact topological language was not commonly used in discussions involving
thinness. Cartan [3, 1946] noted that the fine topology is the restriction of a
certain topology of measures to the class of unit measures supported by
singletons. According to Constantinescu and Cornea [1, 1972]. the Baire
property of the fine topology was noted by Cornea in 1966.
Section 4. The existence of the fine limits in Theorem 4(a) was proved by
Doob [8. 1957] by probabilistic and [11, 1959] by nonprobabilistic methods.
The present proof is new. The identification of these limits could easily
(see the application in Section XII.19) have been made in these earlier
papers but was not.
Section 5. See [Brelot 6, 1944] for classical potential theory extended
to If8" u { oo ; .

Section 6. Theorem 6 is due to Brelot [7. 1944].


Section 9. Theorem 9 is due to Cartan [see Deny 1, 1950, p. 171 ] for
fine limits and to Doob [11, 1959] for fine cluster values (in the Martin
boundary context).
Section 11. Theorem I I is due to Doob [17, 1966].
Section 12. Theorem 12 is due to Brelot [4, 1940].
Sections 14 and 15. Theorems 14 and 15 are due to Brelot [8, 1945].
Section 18. Theorem 18 is due to Brelot [10, 1948], but see also Vallee
Poussin 4. 1938].
Section 19. See [Fuglede 1, 1972] for a discussion of classical potential
theory based on fine superharmonic and fine harmonic functions.
Sections 21 and 22. The material in these sections is taken from [Brelot
9, 1946].
Section 23. See the note to Section XVIII.16 on the parabolic context
domination principle.
Pan I 801

Chapter l.Xll

Sections 1 to 9. Martin [1, 1941] defined what is now called the Martin
boundary and proved what is now called the Martin representation theorem
(Theorem 9). By now the Martin boundary of a set D on which harmonic
functions in some generalized sense are defined is a rather vague concept.
Roughly any set of points (, to each of which corresponds a minimal har-
monic function K(C, ), which is adjoined to D along with possible other
points to form a topological space and which then gives as a representation
of an arbitrary positive harmonic function u an integral J K(C, )MM(dd) is
called a Martin boundary for D. The treatment in Sections 1 to 9 follows
Martin as modernized by Brelot [13, 1956] and put into the context of
h-harmonic functions. The significance for its Martin representation of quasi
boundedness of a harmonic function was pointed out by Parreau [1, 1951 ].
Section 10. Theorem 10 is due to Brelot [ 13, 1956]. Heins [1, 1959] defined
a "generalized harmonic measure" on a Riemann surface D as a positive
harmonic function u for which 0:5 u< 1 and GMD[U A (I - u)] = 0.
According to the remark "Intrinsic definition of h-harmonic measure,"
such a function u is actually the harmonic measure of a Martin boundary
subset. (The Martin boundary discussion is applicable to Riemann surfaces.)
Sections 11 to 14. Naim [1, 1957] defined the minimal-fine topology on
a Martin space (in a more satisfying way than that in Section 12, where her
elegant approach is considerably abbreviated) and proved Theorems 11,
13, and 14. Theorems 13(a) and 14 had already been proved in different
terminology for D a half-space by Lelong [1, 1949].
Section 16. NaIm [1, 1957] proved the part of Theorem 16 involving
minimal-fine limits. The cluster set part of Theorem 16 is taken from [Doob
11, 1959].
Section 18. Theorem 18 is due to Naim [1, 1957].
Section 19. Theorem 19 was proved by Doob [8, 1957] using probabilistic
and [11, 1959] nonprobabilistic methods. This theorem has been generalized
to axiomatic potential theory contexts by Gowrisankaran (to Brelot har-
monic spaces) and by Sibony (to a context including Bauer harmonic spaces).
See [Taylor 1, 1980] for a simple version of Sibony's result with references
to previous work.
Section 21. Theorem 21 was proved by Brelot and Doob [1, 1963] along
with further theorems on the relations between nontangential and minimal-
fine boundary limits of functions defined on a half-space. See the references
there to work of Constantinescu and Cornea and others on the relations
between nontangential and minimal-fine cluster values of functions defined
on a ball and see [Brossard 1, 1978] for such relations derived in probabilistic
language.
802 Historical Notes

Sections 22 and 23. Theorem 23(a) for n = 2, in the disk rather than the
half-plane context, was proved by Littlewood [1, 1928] long before the fine
topology was invented. His result was generalized to the N-dimensional
context by Privalov [1, 1938]. The proofs of Lemma 22 and Theorem 23
are taken (corrected) from [Doob 16, 1965], an unreadably garbled paper.
Fine topology confusion. Let D be a half-space, let be a finite boundary
point of D, and let A be a subset of D. The relations between thinness of A
at C, minimal thinness of A at C, and the relations between these two concepts
when A is in a cone with vertex S are not obvious and have sometimes been
misstated in the literature. The situation has been complicated by the habit
of Doob and some other authors of omitting the qualifier "minimal" in
discussing minimal-fine limits and cluster values at Martin boundary points.
Jackson [1, 1970] surveyed this situation and made the necessary corrections.

Chapter I.XIII

Section 1. In a pioneering paper Gauss [1, 1840] considered measures v


supported by a smooth surface A in 083. For f a given function on A he
studied the properties of v chosen to minimize JA (Gv - 2f) dv for specified
v(A); the measure v was by hypothesis given by a smooth density relative
to surface area. He took it for granted that a minimizing measure p existed
and showed that then Gp - f is identically constant on the support of p.
In particular (f = const), Gauss thereby obtained an equilibrium distribu-
tion for A and, if f is the restriction to A of the potential of a measure A,
Gauss obtained the sweeping of i onto A. Finally Gauss showed how to
manipulate his result to solve the Dirichlet problem for the domain bounded
by A. His work was of course not rigorous but offered ideas not fully digested
for a century. Careful formulations suggested by some of his ideas are
presented in Section 14. The first rigorous derivation of an equilibrium
distribution for an arbitrary compact subset of 1183 was carried through by
Frostman [1, 1935], whose basic tool was a modernization of Gauss's
technique of minimization of the energy integral. The capacity of an arbi-
trary compact subset A of 083 was defined by Wiener [2, 1924] by solving
the (generalized) Dirichlet problem for the unbounded open connected
component of 083 - A with boundary function I on 3A and 0 at the point
oo. This solution can be identified with the restriction to D of the potential
Gp of a measure on A, and Wiener defined the capacity of A as p(A). [Since
Gp is in fact the smoothed reduction RA (relative to 083), the potential Gp
is the capacitary potential of A.] This capacity definition was extended by
Vallee Poussin [1, 1932] who defined the capacity of an arbitrary subset A
of 083 as the supremum of v(H3) for measures v supported by compact subsets
of A with potentials Gv 5 1, and he showed that this definition coincided
with Wiener's for compact sets. The Vallee Poussin capacity is the inner
capacity as defined in Section 11.
Part 1 803

The problem of finding an equilibrium distribution for a set is sometimes


called the Robin problem in view of [Robin 1, 1886] in which Robin showed
that the problem for a smooth surface in I83 is equivalent to the problem of
solving a certain integral equation.
Section 6. Cartan [2, 1945] showed that the space f+ is complete. (He
worked with what we have called "special" domains.) See [Brelot 12,
1953-1954] for a discussion of BLD functions.
Sections 7 and 8. Frostman [1, 1935; 2, 1938] and M. Riesz [1, 1938]
gave the first rigorous proofs of Theorem 7.
Section 10. Choquet was led to his theory of capacities by potential theory
problems. He proved Theorem 10 in [1, 1955].
Section 17. The Wiener thinness criterion was derived in [Wiener 3, 1924]
in the context of boundary point regularity for the Dirichlet problem. Brelot
[4, 1940] put it into the equivalent thinness context.
Section 18. See [Choquet 3, 1958] for an approach to logarithmic capacity
somewhat different from that in Section 18, but observe that he does not
use the standard terminology. H. Jackson kindly showed the author an
example demonstrating that the logarithmic capacity set function is not
subadditive.

Chapter 1.XIV

The material in this chapter is well known and mostly elementary. It is a


fine example of material that is easy to develop, useful to have available for
reference, and for which it is as difficult as it is pointless to find the original
source.

Chapter 1.XV

Section 1. Parabolic potential theory has not been treated systematically


in the literature. Petrowsky [2, 1934-1935] is a valuable source for the
Dirichlet problem. Tychonoff [1, 1935] has many useful results. Widder's
book [2. 1975] is useful for the solutions of the heat equation on a slab.
Some work in axiomatic potential theory has included applications to
parabolic potential theory. See, for example, [Constantinescu and Cornea
1, 1972], a book which covers contexts far more general than parabolic
potential theory with explicit applications to parabolic potential theory.
Probabilists have invoked parabolic potential theory as a convenient tool
in studying Brownian motion in a potential theoretic context. See, for
example. [Doob 6. 1955]. but the reader should be warned that the statement
804 Historical Notes

there that for an open set bounded by hyperplanes the Euclidean boundary
is regular for the heat equation is false.
Section 5. The proof of Theorem 5 is taken from [Tychonoff 1, 1935].
Section 11. See [Moser 1, 1964] for a more precise Harnack-type inequal-
ity than that given in Section 11.
Section 16. The Appell transformation dates back to [Appell 1, 1892].

Chapter I.XVI

Sections 1 to 6. The essential result in these sections is the equivalence of


Theorem 6 parts (al) and (a2), due to Widder [l, 1944]. The preliminary
results leading up to Theorem 6 are mostly taken from Widder's paper and
[Tychonoff 1, 1935].
Section 7. Theorem 7 with h = I is old as a special case of a general
integral limit theorem; see, for example, [Titchmarsh 1, 1937]. Theorem 7
as stated was first proved in [Doob 13. 1960] along with a general ratio
integral limit theorem. See these notes to Appendix VII for references to
such ratio integral limit theorems. See [Hartman and Wintner 1, 1950] for
a Fatou boundary limit theorem for parabolic functions on an interval.

Chapter 1. XVII

Many results in Chapters XVII and XVIII of Part I can be obtained by


translating classical results and their proofs into the parabolic context. Such
results will not be referenced in these notes. Some of the classical reduction
property proofs are no longer valid in the parabolic context, however. Where
essentially different proofs are given, they are due to Boboc, Constantinescu
or Cornea, separately or together, and have been taken from [Constantinescu
and Cornea 1, 1972], which contains proofs valid in a very general context
along with detailed references.
Section 13. The direct half of Theorem 13 (in a general axiomatic context)
is due to Brelot [14, 1962].
Section 18. Theorem 18 is due to Hunt [1, 1956], who proved it in a
probabilistic context.

Chapter I.XVIII

Section 1. Sternberg [1, 1929] was apparently the first to observe that the
Perron method of attacking the Dirichlet problem could be applied in the
parabolic context, although he restricted himself to functions on very special
domains.
Part I 805

Section 4. Babuka and Vyborny [1, 1962] showed that a finite Euclidean
boundary point of a subset D of RN is regular for Laplace's equation if
and only if every Euclidean boundary point s) of D x R is regular for
the heat equation.
Section 6. The iterated logarithm criterion of Example (d) is due to
Petrowsky [2, 1934-1935] who knew that this criterion corresponded to the
Khintchine iterated logarithm law for Brownian motion (Section 2.IX. 15).
Sections 11 and 12. Meyer [ 1, 1962] showed that in a very general potential
theoretic context, defined probabilistically, a set is polar [semipolar] if and
only if it is copolar [cosemipolar].
Section 14. Theorem 14 and Applications (a) and (b) appear to be new.
Applications (c) and (d) in a more general (probabilistic) context are exercises
in Blumenthal and Getoor's [1, 1968] and are proved in their [2, 1970].
Section 15. See [Doob 13, 1960] for a probabilistic version of Theorem 15,
and see Section XIX.15 for a proof of Theorem 15 by Martin boundary
technique. It is natural to conjecture that when h - 1, then (15.1) is true as
a normal limit (that is, as a limit along the normal line to the boundary
point) as well as a coparabolic-fine limit, but in fact Kaufman and Wu
[1, 1982] have exhibited a measure v on the upper half-space D of I such
that lim sup,_0 GD s) = + oo for 0 < c < 1.
According to Theorem 15, a positive parabolic function v on a slab
UFN x ] 0, 6[ has coparabolic-fine limit function AMIN at IN almost every
point of the abscissa hyperplane. Koranyi and Taylor [1, 1983] have shown
that this result can be used to prove Theorem XVI.7 in the special case
h =_ 1, but their method fails for general h. That is, the reasoning used in
proving Theorem XII.21 to go from minimal-fine boundary limits to non-
tangential boundary limits does not seem to have a counterpart in the
parabolic context.
Section 16. In a general axiomatic setting including parabolic potential
theory Janssen [2, 1975] proved that (notation of Theorem 16) h < ii if
suph..4 v(il)/h(,) >_ 1 for Jt, almost every . This result becomes
Theorem 16 with the help of Theorem 14. Blumenthal and Getoor state that
h 5 6 under condition (c') in a more general context but with ti = const as
an exercise in their [1, 1968] and give the proof in their [2, 1970].

Chapter 1.XIX

Various versions of a parabolic context Martin boundary have appeared in


the literature but that given here, especially that relative to a Martin point
set pair, seems closest to Martin's ideas and the easiest for calculational
purposes. See Janssen [1, 1971], for example, for the Martin boundary
obtained by convex set methods in a context including parabolic potential
theory.
806 Historical Notes

Section 15. In the following discussion of functions on the right half-plane


]0, + oo [ x R we shall write that a function on this set has a [one-sided]
parabolic limit a at a boundary point (0, r) if the function has the limit a as
-.(0,r) with s - z < fl2S2] r - sl < for each strictly
positive choice of Y2 with f, < Y2] P. According to Section 10, if h is a
strictly positive parabolic function on the right half-plane, it can be written
in the form

where N,; is a uniquely determined measure on Q8 and Nti is a uniquely deter-


mined measure on - 68+. Let a be a positive superparabolic function on the
right half-plane with parabolic component (Riesz decomposition) deter-
mined by measures N,; and According to Section 15, the function o/h
has coparabolic-fine limit dN;IA4 at N,; almost every point of the ordinate
axis. This result should be compared with the following results. Kemper
[1, 1972] proved that if it - 1, that is, if N,; = /, and Nh - 0, and if t% is
parabolic, then c/h has the stated limit at 1, almost every point of the ordinate
axis in the sense of parabolic approach to the point. On the other hand,
Kaufman and Wu [ 1, 1982] have exhibited an example of a positive parabolic
function h on the right half-plane specified by a continuous singular relative
to 1, measure Nh and by Nti = 0 such that lim info 0 for all S.
Thus 1/h does not give the limit d1,/dNh along normal approach to the
ordinate boundary, but Kaufman and Wu showed that if r and h are positive
and parabolic on the right half-plane, then c/h has the limit dN,;/dNh at Nh;
almost every point of the ordinate axis in the sense of one-sided parabolic
approach to the point. On the other hand, if c is a superparabolic potential
on the right half-plane, Wu [1, 1979] has shown that 0 for
1, almost every s. One-sided parabolic approach does not yield this zero
limit, for example, if c is the superparabolic potential of a measure assigning
a strictly positive value to each point of a countable dense subset of the right
half-plane.

Part 2
The most useful reference books for Part 2 are [Meyer 6, 1966], [Dellacherie
and Meyer 1. 1975; 2. 1980]. and [Dellacherie 2. 1972]. For Markov pro-
cesses see [Dynkin 1. 1960; 2, 1965] and [Blumenthal and Getoor I, 1968].
[Meyer 8, 1968] contains a useful summary of key ideas in the general theory
of stochastic processes, including classification of optional times, classifica-
tion of types of stochastic process measurability, and a discussion of section
theorems. The books of K. M. Rao [4, 1977]. Port and Stone [1, 1978], and
Chung [2, 1981] are useful references for the relations between probability
and classical potential theory.
Pan 2 807

Chapter 2.1

Section 1. The concept of a filtered measure space always underlay


probabilistic analysis involving the Markov property, and the concept was
finally formalized in [Doob 4, 1953] to deal with martingale theory. In the
schools of Dynkin and Meyer the concept has become the base of all analysis
involving Markov process theory, martingale theory, and stochastic integra-
tion. The Dynkin approach to Markov processes in his [2, 1965] leads to a
further refinement of a filtered space which is not needed in this book.
Section 2. Dynkin defined process measurability, in the context of Markov
processes, in his [1, 1960] essentially as what is now called progressive
measurability. Meyer defined and applied progressive measurability in [2,
1962-1963]. This measurability was rediscovered and given its present name
in [Chung and Doob 1, 1965], written in ignorance of the earlier work.
Section 5. Let ..., xo, x,, ... be mutually independent random vari-
ables, each uniformly distributed on the interval [0, 1], and define 9(n) =
9f. In what amounts to this context Theorem 5 was proved by
Jessen [1, 1934] and for n-. +co by Levy [1, 1935], who supposed x in
Theorem 5 to be the indicator function of a set. It is clear in Levy's discus-
sion, however, that he does not use the special independence context, and
in fact this special context is not presupposed in his treatment of the same
convergence result in his [2, 1937]. Theorem 5 as stated is in [Doob 1, 1940].
Section 9. The first rigorous treatment in a general continuous parameter
context of the hitting of a set by a process was given by Hunt [2, 1957].
Meyer [2, 1962-1963] was the first to discuss the hitting of progressively
measurable sets.
Section 10. Kolmogorov [1, 1933] gave the measure theoretic basis for
probability, including the definitions of expectation and conditional expecta-
tion and a construction of measures on infinite-dimensional product spaces.
The latter construction is what is needed for canonical processes.
Section 13. For variations of Theorem 13 see [Chung and Doob 1, 1965],
[Meyer 6, 1966], and [Dellacherie and Meyer 1, 1975].
Section 14. See [Dellacherie 2, 1972] for a full discussion of the important
a algebras of subsets of 68' x S2, first defined by Meyer. See [Meyer 6. 1966;
8, 1968] for earlier discussions.

Chapter 2.11

Section 1. Like the Moliere character who spoke in prose without realizing
it, mathematicians had dealt with optional times for centuries without finding
it necessary to define them precisely. Precise definitions were finally intro-
duced, along with a algebra filtrations, to answer the needs of Markov
808 Historical Notes

process and martingale theories. Meyer [6, 1966; 8, 1968] first classified
optional times.
Section 4. Hunt [2, 1957] was the first to show that in an appropriate
context the hitting time of an analytic set is optional. Doob [5, 1954] had
shown that the hitting time by Brownian motion of a capacitable set is
optional, but in 1954 it was not known that all Borel sets are capacitable
in his context. Meyer [2, 1962-1963] proved that the hitting time of a
progressively measurable set is optional.
Section 7. Observe that predictable optional times are defined slightly
differently in [Dellacherie and Meyer 1, 1975].
Section 8. Section theorems were introduced into probability theory by
Meyer [2, 1962-1963], more successfully in [6, 1966]. See Dellacherie [1,
1972], from which the proof of Theorem 8 is taken, for an elegant unified
approach to them.
Section 9. The proof of Theorem 9 is taken from the correction sheet in
[Dellacherie and Meyer 1, 1975].
Section 10. Dellacherie [1, 1969] proved that, in the notation of this
section, a predictable subset A of If8' x n is semipolar if for almost every
w the set {t: (t,w)EA} is countable.

Chapter 2.III

Before martingales had been formally christened, Levy [1, 1935; 2, 1937],
Bernstein [1, 1937], and other mathematicians had analyzed some of their
properties in special contexts; usually the martingales in question arose as
partial sums n -. Eo y) of a sequence y, of random variables under the condi-
tion E{ y,I ,, . y'1-i } = 0 so that the sums arose as generalizations of
sums of independent random variables with zero means. Ville [1, 1939]
defined a martingale very nearly as a positive martingale is now defined but
tied it to a sequence of independent random variables under analysis. His
fundamental tool, a fact he proved, was that almost every sample sequence
ofa positive martingale sequence is bounded (see Theorem 9). Doob [1, 1940]
discussed martingales and proved the basic martingale convergence theorems
under the name "family of random variables with the property E." ("E"
was chosen not as the initial letter of "expectation" but as the first letter in
the alphabet following D.) Under the respective names "semimartingale"
and "lower semimartingale," submartingales and supermartingales were
introduced in [Snell 1, 1952] and [Doob 4, 1953]. This obviously inappro-
priate nomenclature was chosen under the malign influence of the noise
level of radio's SUPERman program, a favorite supper-time program of
Doob's son during the writing of [Doob 4, 1953]. For further work in
martingale theory see [Neveu 1, 1972] (discrete parameter context) and
[Dellacherie and Meyer 2, 1980] (continuous parameter context).
Part 2 809

Section 4. The trivial but useful map in this section appears to be new.
Section 6. Theorem 6 and its generalizations to various forms of optional
sampling in Chapters III and IV are adapted from [boob 1, 1940; 4, 1953]
with improvements from [Meyer 6, 1966].
Section 9. Theorem 9 is taken from [Doob 4, 1953]. In the martingale
case or the Imartingalel case Theorem 9(a) was proved by Levy [2. 1937].
Bernstein [l, 1937], and Ville [1, 1939].
Section /1. Theorem I I is taken from [Doob 4, 1953].
Section 12. Inequality (12.2) is new, as are (12.3) and (12.4) except when
1. Crossings (implicit in [Doob 1. 1940]) were formally introduced
into martingale theory in Doob [3, 1951], where a variant of (12.3) was
obtained for a martingale and 1. Snell in [I, 1952] extended
crossing inequalities to submartingales and supermartingales. Dubins [1.
1966] obtained a variant of (12.4) with 1.

Sections 13 to 17. These results are taken from [Doob 4, 1953]. In the
martingale case Theorems 13 and 17 were in [Doob I. 1940].
Section 19. This may be the first time that the natural decomposition
theorem appears in print in the martingale theory context, but the result is
merely a special application of the general theory
Section 22. The First application of reduction theory to martingale theory
was made by Snell [I, 1952]. See Neveu [I, 1972] for further discussion of
related problems. The infimum of the class of supermartingales majorizing
a given process under appropriate side conditions is called the Snell envelope
of the given process.

Chapter 2.I V

Section 1. The study of the continuity properties of continuous parameter


martingale sample functions was initiated in [Doob 1, 1940]. completed in
[Doob 3, 1951], and extended to supermartingales in [Doob 4. 1953].
Perhaps the first sophisticated use of optional times appeared in the first
of these papers to obtain martingale right continuity properties (without
crossing inequalities, unknown at that time).
Sections 2 and 3. Theorems 2 and 3 are adapted from [Doob 4, 1953]
with improvements in formulation from [Meyer 6, 1966).
Section 4. Theorem 4, the counterpart of, but much deeper than, the
classical potential theory rather elementary fact (Section 1.11.4) that the
limit of an increasing sequence of superharmonic functions on a connected
open subset of R" is either superharmonic or identically + 00, is due to
Meyer [6, 1966]. The elementary right continuity proof is new.
810 Historical Notes

Section 5. This theorem appears to be new, but see related work of


Mertens [1, 1972]. As Mertens' work indicates, the choice of standard
continuous parameter supermartingales need not be the almost surely right
continuous ones. For some purposes the right continuity hypothesis can be
advantageously replaced by a combination of an appropriate type of mea-
surability together with invariance of the supermartingale inequality under
optional sampling.

Sections 6 to 12. Doob [4, 1953] proved the rather trivial discrete param-
eter decomposition Theorem 8 and after searching for several years found a
mathematician who could prove the continuous parameter version Theorem
I 1 [Meyer 4, 1962; 5, 1963]. See the appendix to Meyer [6, 1966] for earlier
decomposition results due to Volkonski, Sur, and Meyer which are for
supermartingales associated with Markov processes. Perhaps as tribute to
Doob's persistent search for this holy grail, the Meyer decomposition is
sometimes inappropriately called the Doob-Meyer or even the Doob decom-
position. Meyer proved Theorem 11 with the monotone increasing process
natural in the sense of Section 7. The equivalence between natural and
nearly predictable [Theorem 7(a3)] was proved by Doleans [1, 1967].
Rao's [1, 1969] proof of Theorem 11 is followed in the text. See [Dellacherie
and Meyer 2, 1980] for extensive applications of the Meyer decomposition.

Section 13. The supermartingale domination principle appears to be new.


This example of the potential theory of supermartingales indicates how
closely such a theory parallels classical potential theory. Unfortunately, it
was impractical to include the elegant Airault-Follmer [1, 1974] results or
the Azema [1, 1972] and Azema-Jeulin [1, 1976] results in this direction.
Among many other things these papers contain a counterpart for super-
martingales of conditional Markov processes. The most glaring probabilistic
potential theory omission from this book, however, is that of additive
functionals of Markov processes (see, for example, [Dynkin 2, 1963] and
[Blumenthal and Getoor 1, 1968]). The author has three convincing excuses
for this omission, of increasing cogency: (1) the existence of these books,
(2) lack of space, (3) ignorance of the subject.

Sections 17 to 20. Reduction theory in this form seems to be new, but see
also related work involving Snell envelopes by Mertens [1, 1972], Azema
[1, 1972], and Azema and Jeulin [1, 1976]. Results in the latter paper
together with Theorem 13 suggest that in some contexts R i.) in Section 17
should be defined as the infimum of the class of almost surely right contin-
uous positive supermartingales whose left limit processes majorize that of
on A. This definition is also suggested by the awkward extra hypotheses
in Section 18(m).

Section 21. This energy treatment is due to Meyer [3, 1962-1963].


Part 2 811

Chapter 2.V

Sections 2 and 4. Krickeberg [1, 1956] introduced lattice ideas into


martingale theory. In particular, he discussed the LM operator and proved
Theorem 4(c) for p = 1, the "Krickeberg decomposition" of a martingale.
See [Dellacherie and Meyer 2, 1980] for further elaboration of this decom-
position.
Section 12. Local martingales were introduced into martingale theory by
Ito and S. Watanabe [1, 1965].
Section 13. Quasimartingales were introduced by Fisk [1, 1965]. Theorem
13 is due to K. M. Rao [2, 1969]. See his paper and that of Dellacherie and
Meyer [2, 1980] for further discussion of the decomposition of a quasimar-
tingale. See also [Follmer 1, 1973] for a correspondence, unfortunately
omitted from this book, between quasimartingales relative to a filtered
probability space and measures on the a algebra of predictable sets deter-
mined by the filtration.

Chapter 2.VI

This chapter treats only those aspects of Markov process theory needed in
this book. Readers interested in going further into probabilistic potential
theory, which is based on Markov processes, may consult [Dynkin 2, 1965],
[Blumenthal and Getoor 1, 1968], and [Chung 2, 1981]. Knowledgeable
readers will observe that translation operators are not mentioned in this
chapter. They are not needed in this book.
Section 1. A. A. Markov introduced what are now called Markov pro-
cesses into probability theory in [1, 1906].
Section 3. The strong Markov property in the discrete parameter case is
trivial to prove, but the realization that such a property is relevant and
requires proof had to await the sophistication not present before the property
was needed in the technically more difficult continuous parameter case. Even
now rather than explicitly recalling and applying this property, it is frequently
easier, when the discrete parameter strong Markov property is relevant in a
computation, to make the computation detailed enough to avoid explicit
statement and use of the property.
Section 4. Kakutani [3, 1945] defined random walks on domains of R2
as in the example in this section. He proved, omitting some details, what
amounts to the fact that the distribution of x(n) in the example tends to
harmonic measure relative to the initial point.
Section 6. Hunt [2, 1957] was the first to consider in depth the hitting
probabilities of sets by Markov process trajectories.
812 Historical Notes

Section 8. See Meyer [7, 1967] for a more complete analysis of the
filtrations generated by Markov processes. The 0-1 law in this section
is known as the Blumenthal 0-1 law in view of [Blumenthal 1, 1957].
Blumenthal's paper is probably the first paper to treat a Markov process
explicitly as a process adapted to a general filtration and having the Markov
property relative to the filtration.
Section 9. The first version of the strong Markov property in a continuous
parameter context was proved by Doob [2, 1945] in the context of a count-
able state space. Hunt [1, 1956] proved a version for processes with inde-
pendent increments. The first generally applicable versions were proved
independently by Dynkin and Yushkevich [1, 1956] and Blumenthal [1,
1957].
Sections 10 to 12. Hunt [2, 1957] introduced excessive functions and
measures into Markov process theory as part of his fundamental develop-
ment [1, 1957; 2, 1957; 3, 1958] of probabilistic potential theory based on
Markov processes.
Section 13. Conditioned Markov processes were introduced in [Doob 8,
1957] in the context of Brownian motion. See [Dynkin 2, 1963] for condi-
tioned Markov processes in the context of multiplicative linear functionals.
Section 15. See [Meyer, Smythe, and Walsh 1, 1972] for a deep discussion
of Markov processes killed at various kinds of random times.

Chapter 2.VII

The Brownian motion process is the link between classical potential theory
and martingale theory, and only those Brownian motion properties relevant
to this linkage are discussed in this book. For more information on Brownian
motion see, for example, [Levy 4, 1948], [Ciesielski 1, 1966], [Freedman
1, 1971], [Ito and McKean 1, 1974], [Knight 1, 1981], and, especially for
the physical significance of Brownian motion, [Nelson 1, 1967]. The process
takes its name from the English botanist Brown who in 1827 observed
irregular motion of pollen particles in a liquid. Einstein in 1905 obtained
from physical considerations the fact that the mean square displacement of
a Brownian particle is a multiple of the displacement time and evaluated
this multiple. Bachelier [1, 1900; 2, 1901] and later papers derived many
distributions involving Brownian motion considered as a limit of random
walks and saw the connection with the heat equation. The first rigorous
construction of a Brownian motion process was given by Wiener [1, 1923],
but his work remained unknown to or at least ignored by other probabilists
for about 15 years. Thus probabilists were at a disadvantage in treating
sample functions of Brownian motion; although they realized that in some
sense these sample functions were continuous, it was not clear how to
Part 2 813

formulate this property without Wiener's work. It was intuitively clear,


however, that many distributions involving Brownian motion from a point
of a domain involved a process whose paths were ordinary Brownian motion
paths until they hit the boundary of the domain, where they were absorbed
or reflected in some way so that the transition probabilities involved, like
the transition densities of Brownian motion in 08", were governed by the
heat equation in the domain. The conduct of the paths at the boundary
then determined corresponding heat equation boundary conditions.
Sections 2 and 3. The definition of a Brownian motion process used here
implies that one example is the canonical example: a measure is assigned to
the space of continuous functions from R into I8" making the coordinate
functions random variables with the properties BM I -BM4. This measure is
known as Wiener measure, and Brownian motion is sometimes defined as this
canonical example. Under this definition a process satisfying BM I -BM4 is
not necessarily a Brownian motion.
Section 4. Theorem 4 is a special case of a theorem of Meyer [7, 1967].
The proof in the text follows that of Chung and Walsh [2, 1974].
Section 6. The zero-one law for Brownian motion can be considered as a
special case of a general zero-one law of Kolmogorov [1, 1933].
Section 8. The Andre reflection principle goes back to [Andre 1, 1887],
where a reflection idea is used in a ballot counting problem. The distributions
in Section 8 were derived by Bachelier (see the above notes to this chapter),
but his work was ignored and rediscovered by many others.
Section 9. Hunt [1, 1956] derived the basic properties of the transition
density for Brownian motion in an open set.
Section 11. The transition density for Brownian motion in an interval
was found by Bachelier (see the above notes to this chapter).

Chapter 2.VIII

The integral J, f dw with f a function from I into U8 and a Brownian


motion was first discussed by Wiener [1, 1923]. Ito [1, 1944] allowed the
integrand to depend on the Brownian paths, thereby inaugurating a far-
reaching further development. See McKean [1, 1969] for many applications
of Ito's integral, and see [Dellacherie and Meyer 2, 1980] for stochastic
integrals with differentials more general than Brownian motion differentials.
Section 3. The problem of finding conditions that Y. as defined by (3.10)
be a martingale is a special case of a problem proposed by Girsanov in 1960
in which is a local martingale. Various sufficient conditions have been
found; for example, Novikov [1, 1972] found a condition reducing to (k)
in our context.
814 Historical Notes

Section 8. Theorem 8 is a consequence of Ito's [3, 1951] construction of


a complete orthonormal system on a Brownian motion measure space. See
also [Kunita and S. Watanabe 1, 1967] for a systematic study of general
stochastic integrals in L2 contexts. The proof in Section 8 is taken from
[Parthasarathy 1, 1978]. Dudley [1, 1977] proved that every finite-valued
.F w(t), sE IlB+; measurable function can be expressed as an Ito integral with
integrand in r*.
Section 10. Levy [3, 1940] proved that if t, is a dense sequence in [0, b]
and if t4"', ... , are t,,, ... , t" in increasing order, then

lim i [w(ti"') - w(tj"'1)]2 = a2 b a.s.

See [Doob 10. 1953] for a martingale proof of this limit theorem.
Section 12. It6's lemma, also called Ito's formula, was proved in [2,195 1].
Section 14. The fact that the composition of an analytic function with
plane Brownian motion is Brownian motion with a new time scale is due
to Levy [4, 1948].

Chapter 2.IX

The martingale theory properties of the composition of superharmonic and


harmonic functions with Brownian motion were treated in [Doob 5, 1954],
and the corresponding discussion in the parabolic context was given in
[Doob 6, 1955]. Nonroutine results in Chapter IX, aside from results
specifically attributed otherwise, are taken from these papers, although most
of the proofs are different and some of the results are slightly refined.
Section 5. Levy [3, 1940] proved Theorem 5(c) and stated Theorem 5(a)
for A a singleton. Kakutani [1, 1944] proved Theorem 5(b) and [2, 1944]
sketched a proof of Theorem 5(a) for N = 2.
Section 10. Theorem 10(a) was indicated without proof details in
[Kakutani 2, 1944].
Section 11. Theorem I 1 in its classical and parabolic versions [Doob 5,
1954; 6, 1955] was generalized by Hunt [2, 1957], Dynkin [2, 1963], and
Meyer [7, 1967] to yield various continuity properties of the composition
of excessive functions with their corresponding Markov processes.
Section 13. The idea that Theorem 13(a) and its parabolic counterpart
must be true is old, but a rigorous proof had to await the rigorous develop-
ment of harmonic measure and of Brownian motion. Courant, Friedrichs,
and Lewy [1, 1928] obtained harmonic function Dirichlet solutions for
smooth domains by solving a corresponding problem for difference equa-
Part 3 815

tions and going to the obvious limit. They refer to the probabilistic inter-
pretation of their difference equation. Petrowsky [1, 1933-1934] gave a
similar discussion, and Khintchine [1, 1933] in a corresponding discussion
for both harmonic and parabolic functions mentions the Brownian motion
interpretation of the limiting case. Such ideas were common at the time,
but Wiener's [1, 1923] treatment of Brownian motion was unknown or
unappreciated, and, for example, Khintchine's Brownian motion interpreta-
tion was therefore somewhat artificial. Kakutani [2, 1944; 3, 1945] stated
Theorem 3(a) with indications of a proof.
Section 14. Theorem 14 (in a much more general context) is due to Hunt
[2, 1957].
Section 15. The dichotomy in Theorem 5 in the classical and parabolic
contexts [Doob 5, 1954; 6, 1955] has been generalized to define fine and
cofine topologies in the probabilistic potential theory of Markov processes.
Khintchine [ 1, 1933] proved the iterated logarithm law for Brownian motion,
aside from the fact that he could not quite define the probabilities involved
without a rigorous definition of Brownian motion.
Section 17. Theorem 17 is part of the folk lore of the subject but is new
as stated.

Chapter 2.X
Section 6. See [Chung and Walsh 1, 1969] for a general discussion of
reversing a Markov process.
Section 8. The anplication of martingale theory to derive a probability
version of Theorem 1.XI.4 is taken from [Doob 8, 1957]. See [Weil 1, 1969]
and [Airault 1, 1973] for such work in a more general context. Airault shows
that the specification of the limit in Theorem I.XI.4(b) is possible because
if the Riesz measure associated with a positive superharmonic function h is
supported by a polar set, the lifetime of an h Brownian motion process is
predictable. Unfortunately the stated hypotheses underlying most probabil-
istic analysis of this sort, including the above references, are too strong to
cover the parabolic context.
Section 10. The evaluation of the capacitary distribution in terms of last
hitting distributions (in a more general context) is due to Chung [1, 1973].

Part 3
Chapter 3.1

The lattice theoretic results in this chapter are mostly routine or in the
folklore; so few references to their origin will be given.
816 Historical Notes

Section 4. Lemma 4, the key to the connection between uniform inte-


grability and the PWB method, is taken from [Doob 7, 1956], where the
potential theory version is proved in an axiomatic potential theory context.
Section 8. See [Arsove and Leutwiler 1, 1974] for a general approach to
Theorem 8.
Section 12. Lamb [1, 1971] pointed out the decomposition S., =
+Smsf

Chapter 3.11

Section 2. The delicate parts of Theorem 2 not treated in previous chapters


are taken from [Doob 7, 1956; 9, 1958].

Chapter 3.1I1

Section 1. See [Kunita and T. Watanabe 1, 1965. 2. 1966], [Meyer 9,


1968], and [Dynkin 3, 1969; 4, 1971] for probability-potential theory
approaches to the Martin space and process paths on it, carried out in
Markov process contexts at various levels of generality. The reader will
observe that this chapter is rather skimpy and that a parabolic context
counterpart is missing. The point is that the author is a firm believer in
finite-valued stopping times.
Section 4. Observe that, as pointed out in the notes to Section XI.4, the
probabilistic proof of the Fatou boundary limit theorem for the Martin
space preceded the nonprobabilistic proof.
Section 6. The example is taken from [Doob 10, 1958], which contains
several examples of conditional Brownian motions in a half-space, involving
paths to and from boundary points.

Appendixes
Appendixes I and II

The operation constructing a nucleus from a Suslin scheme was devised by


Suslin in [1, 1917]. Apparently Lusin coined the name "analytic sets."
Choquet created his capacity theory in [1, 1955; 4, 1959]. Theorem 11.5,
which makes analytic set theory conveniently applicable in many probabil-
istic contexts, is taken from [Meyer 6, 1966]. See [Dellacherie and Meyer
1, 1975; 2, 1980] for more on analytic sets and capacities and their application
to probabilistic potential theory.
Appendixes 817

Appendixes III to VI

The material in these appendixes, except possibly for Section 12 of Appendix


IV for which see [Besicovitch 1, 1946], is traditional and routine and is
assembled for the reader's convenience. [Peressini 1, 1967] is a useful source
for a reader who wants enough but not too much vector lattice theory.

Appendix VII

See [Doob 13, 1960; 14, 1960-1961] for ratio integral limit theorems; those
in Appendix VII are adaptations to fit the needs of this book.
Bibliography

Helen Airault: [1] "Theoreme de Fatou et frontiere de Martin." J. Funct. Anal. 12 (1973),
418-455.
Hekne Airault and Hans Follmer: [I] "Relative densities of semimartingales." Inventions
Math. 27 (1974), 299-327.
Desire Andre: [I] Solution directe d'un probleme resolu par M. Bertrand. C.R. Acad. Sci.
Paris 105 (1887). 436-437.
P. Appell: [I] Sur ('equation a=z/ax' - Oz/ay = 0 et la theorie du chaleur. J. Math. Pures.
Appl. (4) 8 (1892), 187-216.
Maynard Arsove and Heinz Leutwiler: [ I ] Quasi bounded and singular functions. Trans. Amer.
Math. Soc. 189 (1974), 275-302.
Jacques Azema: [1] Quelques applications de la theorie generale des processus. 1. Inventions
Math. 18 (1972), 293-336.
J. Azema and T. Jeulin: [I] Precisions sur la measure de Follmer. Ann. Inst. Henri Poincare
12 (1976), 257-283.
1. Babuska and R. Vyborn': [1] Reguliire and stabile Randpunkte fur das Problem der
Wiirmeleitungsgleichung. Ann. Polon. Math. 12 (1962), 91-104.
Louis Bachelier: [1] Theorie de la speculation. Ann. Sci. Ecole Norm. Sup. (3) 17(1900), 21-86.
[2] Theone mathematique dujeu. Ann. Sci. Ecole Norm. Sup. (3) 18 (1901),143-210.
Serge Bernstein: [I] On some transformations of the Chebychev inequality. (Russian) Dokl.
Akad. Nauk SSSR 17 (1937), 275-277.
A. S. Besicovitch: [I] "A general form of the covering principle and relative differentiation
of additive functions II." Proc. Cambridge Phil. Soc. 42 (1946), 1-10.
R. M. Blumenthal: [1] An extended Markov property. Trans. Amer. Math. Soc. 85 (1957),
52-72.
R. M. Blumenthal and R. K. Getoor: [I] Markou Processes and Potential Theory. New York,
Academic, 1968.
[2] "Dual processes and potential theory." Proc. Twelfth Biennial Sem. Can. Math. Congr.
1970, 137-156.
Maxime Bocher: [I] "Singular points of functions which satisfy partial differential equations
of the elliptic type." Bull. Amer. Math. Soc. 9 (1903), 455-465.
Georges Bouligand: [1] "Sur le probleme de Dirichlet." Ann. Soc. Polon. Math. 4 (1926),
59-112.
M. Brelot: [1] "Fonctions sous-harmoniques et balayage 1, 11." Acad. Roy. Belgique, Bull. Cl.
Sci. (5) 24 (1938), 301-312, 421-436.
[2] "Sur Ie potentiel et les suites de fonctions surharmoniques." C. R. Acad. Sci. Paris 207
(1938), 836-838.
[3] "Families de Perron et probleme de Dirichlet." Acta Litt. Sci. Szeged 9 (1939), 133-153.
4 "Points irreguliers et transformations continues en theorie du potentiel." J. Math. Pures.
Appl. 19 (1940), 319-337.
[5] "Sur la theorie autonome des fonctions sousharmoniques." Bull. Sci. Math. 65 (1941),
72-98.
[6] "Sur le role du point it l'infini dans la theorie des fonctions harmoniques." Ann. Sci.
Ecole Norm. Sup. 61 (1944), 301-332.
[7] "Sur les ensembles effiles." Bull. Sci. Math. 68 (1944), 12-36.
[8] "Minorantes sous-harmoniques, extremales et capacites." J. Math. Pures. Appl. 24 (1945),
1-32.
820 Bibliography

[9] "Etude generale des fonctions harmoniques ou surhaimoniques positive au voismage


d'un point-frontiere irregulier." Ann. Univ. Grenoble 22 (1946), 201-219.
[10] "Quelques propnetes et applications du balayage." C R. Acad. Sci. Paris 227 (1948),
19-21
[I I ] "La theone moderne du potentiel." Ann. Inst. Fourier Grenoble 4 (1952), 113-140(1954).
[ 12] "Etude et extensions du pnncipe de Dinchlet." Ann. Inst. Fourier Grenoble 5 (1953-1954),
371-419
[13] "Le probleme de Dinchlet Axiomatique et frontiere de Martin " J Math Pures. APP/
35 (1956), 297-335
[14] "Quelques propnetes et applications nouvelles de I'effilement." Sem. (Brelot-Choquet-
Deny) Theorie du Potentiel6 (1961-1962), 1-27-1-40, 1962.
[15] Elements de la Theorie Classique du Potentiel. 4th ed. Centre du Documentation Univer-
sitaire Paris, 1969
[16] On Topologies and Boundaries in Potential Theory Lecture Notes in Mathematics 175.
Berlin. Spnnger-Verlag, 1971.
[17] "Les etapes et les aspects multiples de la theone du potentiel," L'Enseignment Math
Ser 11 18 (1972), 1-36.
M. Brelot and J L. Doob. [1] "Limites angulaires et limites fines." Ann. Inst. Fourier Grenoble
13 (1963), 395-415
Jean Brossard [1] Comportement "non-tangentiel" et comportment "Brownien" des fonc-
tions harmoniques dans un demi-espace. Demonstration probabiliste d'un theoreme de
Calderon et Stein. Sem Prob. XII 1976/77, Lecture Notes in Mathematics 649. Berlin,
Spnnger-Verlag, 1978, pp 378-397.
Heinrich Burkhardt and W. Franz Meyer: [1] "Potentialtheone (Theone der Laplace-Poisson-
schen Differentialgleichung)." Enzyc. Math. Wiss. lIA7b (1900), 464-503.
Henn Cartan [1] "Capacite exterieure et suites convergentes." C. R. Acad. Set. Parts 214
(1942), 944-946
[2] "Theone du potentiel newtomen. energie, capacite, suites de potentiels." Bull. So(. Math
Fr 73 (1945), 74-106.
[3] "Theone generale du balayage en potentiel newtonien " Ann. Inst. Fourier Grenoble 22
(1946). 221-280.
Gustave Choquet. [1] "Theory of capacities." Ann. Inst. Fourier Grenoble 5 (1953- 1954),
131 -295 (1955).
[2] " Potentiels sur un ensemble de capacite nulle Suites de potentiels " C R Acad Sci.
Parts 244 (1957), 1707-1710
[3] " Capacitabilite en potentiel loganthmique." Acad Rov Belg. Bull. Cl Sci (5) 4411958),
321 326
[4] "Forme abstraite du theoreme de capacitabilite " Ann. Inst. Fourier Grenoble 9 (1959),
83-89
Kai Lai Chung [1] "Probabilistic approach in potential theory to the equilibrium problem."
Ann Inst Fourier Grenoble 23/3 (1973), 313-322
[2] Lectures from Markov Processes to Brownian Morton. Berlin, Spnnger-Verlag, 1981
Kai Lai Chung and J L. Doob- [I] "Fields, optionality and measurability." Amer J Math
87 (1965), 397-424.
Kai Lai Chung and John B Walsh [ I ] "To reverse a Markov process." Acta Math. 123 (1969).
225 251
[2] -Meyer's theorem on predictability " Ztschr Wahrscheinlichkettstheorie verw Geb 29
(1974),253-256
Z Ciesielski [ I ] Lectures on Brownian motion, heat conduction and potential theory. Aarhus.
Denmark Aarhus Univ., 1966.
Corneliu Constantinescu and Aurel Cornea. [1] Potential theory of harmonic spaces Berlin.
Spnnger-Verlag, 1972.
R Courant, K Fnedrichs, and H. Lewy: [1] "Uber die partiellen Differenzengleichungen der
mathematischen Physik." Math Ann. 100 (1928), 32-74
Claude Dellacherie [1] Ensembles aleatoires 1. 11. Sem. Prob. III /967 8, Lecture Notes in
Mathematics 88 Berlin, Spnnger-Verlag, 1969.
[2] Capacites et Processus Stochastiques. Erg. Math. u ihrer Grenzgebiete 67 Berlin,
Spnnger-Verlag, 1972.
Claude Dellacherie and Paul-Andre Meyer. [1] Probabilites et potentiel Chapters I IV
Act Sc, Ind. 1372 (1975), Paris, Hermann.
[2] Chapters V-VIII Act Sri. Ind 1385(1980), Pans, Hermann
Bibliography 821

Jacques Deny: [1] "Les potentiels d'hnergie finie." Acta Math. 82 (I950), 107-183.
C. Dolhans (= C. Dolhans-Dade): [1] "Processus croissant naturels et processus croissant
tres-bien-measurables." C. R. Acad. Sci. Paris 264 (1967), 874-876.
J. L. Doob: [I] "Regularity properties of certain families of chance variables." Trans. Amer.
Math. Soc. 47 (1940), 455-486.
[2] "Markoffchains-denumerable case." Trans. Amer. Math. Soc. 58 (1945), 455-473.
[3] "Continuous parameter martingales." Proc. Sec. Berkeley Symp. Math. Statistics Prob.
1950, Berkeley, 1951, pp. 269-277.
[4] Stochastic Processes. New York, Wiley, 1953.
[5] "Semimartingales and subharmonic functions." Trans. Amer. Math. Soc. 77 (1954),
86-121.
[6] "A probability approach to the heat equation." Trans. Amer. Math. Soc. 80 (1955),
216-280.
[7] "Probability methods applied to the first boundary value problem." Proc. Third Berkeley
Symp. Math. Statistics and Prob. 1954/5 Vol. 2, Berkeley, 1956, pp. 49-80.
[8] "Conditional Brownian motion and the boundary limits of harmonic functions." Bull.
Soc. Math. Fr. 85 (1957), 431-458.
[9] "Probability theory and the first boundary value problem." 111. J. Math. 2(1958),19-36.
[10] "Boundary limit theorems for a halfspace." J. Math. Pures. Appl. (9)37(1958),385-392.
(11] "A non-probabilistic proof of the relative Fatou theorem." Ann. Inst. Fourier Grenoble
9 (1959), 293-300.
[12] "A relativized Fatou theorem." Proc. Nat. Acad. Sci. USA 45 (1959),215-222.
[13] "A relative limit theorem for parabolic functions." Trans. Second Prague Conference on
Information Theory, Statistical Decision Functions, Random Processes. Prague, Czechoslo-
vak Acad. Sci., 1960, pp. 61-70.
[ 14] "Relative theorems in analysis." J. Anal. Math. 8 (1960-1961), 289-306.
[ 15] "Conformally invariant cluster value theory." Ill. J. Math. 5 (1961), 521-549.
[ 16] "Some classical function theory theorems and their modern versions." Ann. Inst. Fourier
Grenoble 15 (1965), 113-136.
[17] "Applications to analysis of a topological definition of smallness of a set." Bull. Amer.
Math. Soc. 72 (1966), 579-600.
Lester E. Dubins: [I] "A note on upcrossings of semimartingales." Ann. Math. Stat. 37(1966),
728.
R. M. Dudley: [I] "Wiener functions as 1to integrals." Ann. Prob. 5 (1977), 140-141.
E. B. Dynkin: [ I ] Foundations of the Theory of Markov Processes (translation of his 1959 Russia
book:Ocnoeanux Teopuu Mapkoeckux llpoyeccoe). Oxford. Pergamon, 1960.
[2] Markov Processes (translation of his 1963 Russian book: Mapkoeckue Ilpoyeccs,). Berlin,
Springer-Verlag, 1965.
[3] "The space of exits of a Markov process." Russ. Math. Surv. 24/4 (1969), 89-157.
[4] "Initial and final behavior of Markov process trajectories." Russ. Math. Surv. 26 (1971),
165-185.
E. B. Dynkin and A. A. Yushkevich: [1] Strong Markov processes. Theory Prob. Appl. 1(1956),
134-139.
G. C. Evans: [1] "Application of Poincare's sweeping-out process." Proc. Nat. Acad. Sci. USA
19 (1933), 457-461.
[2] "On potentials of positive mass." Trans. Amer. Math. Soc. 37 (1935), 226-253.
P. Fatou : [I] "Series trigonometriques et series de Taylor." Acta Math. 30 (1906), 335-400.
D. L. Fisk: [I] "Quasi-martingales". Trans. Amer. Math. Soc. 120 (1965), 369-389.
Hans Follmer: [1] "On the representation of semimartingales." Ann. Prob. 1 (1973), 580-
589.
David Freedman : [1] Brownian motion and diffusion. San Francisco, Holden-Day, 1971.
Otto Frostman: [1] "Potentiel d'hquilibre et capacith des ensembles avec quelques applications
it la theorie des fonctions." Meddel. Lunds Univ. Mat. Sem. 3 (1935), 1-118.
[2] "Sur le balayage des masses." Acta Litt. Sci. Univ. Szeged, Sec. Sci. Math. 9 (1938),
43-51.
Bent Fuglede: [1] Finely Harmonic Functions. Lecture Notes in Mathematics 289. Berlin,
Springer-Verlag, 1972.
C. F. Gauss: [1] Allgemeine Lehrsatze in Beziehung auf die im vehrkehrten Verhiiltnisse des
Quadrats der Entfernung wirkenden Anziehungs- and Abstossungs-Krafte. Gauss Werke 5,
pp. 197-242, 1840, Gottingen, 1867.
George Green: [1] An essay on the application of mathematical analysis to the theories of
822 Bibliography

electricity and magnetism. Nottingham 1828. Math. Papers. London, Macmillan, 1871,
pp. 9-41
A. Harnack. [ I ] Die Grundlagen der Theorie des logarithmischen Potentials and der eindeutigen
Potentialfunktion. Leipzig, Teubner, 1887.
Philip Hartman, Aurel Wintner: [I] "On the solutions of the equation of heat conduction."
Amer J Math. 72 (1950), 367-395.
Maurice Heins- [1] On the principle of harmonic measure. Comment. Math. He/v. 33 (1959),
47 - 58.
L. L. Helms [ I ] Introduction to Potential Theory. New York, Wiley, 1969
G. Herglotz : [ I ] "Uber Potenzreihen mit positivem reellen Teil im Einheitskreis." Ber. Verhandl.
Sachs Akad. Wiss. Leipzig Math.-Phys. K!. 63 (1911), 501 - 511.
G. A. Hunt: [I] "Some theorems concerning Brownian motion." Trans. Amer. Math. Soc.
81 (1956).294-319.
[2] "Markoff processes and potentials l." 111. J. Math. 1(1957), 44-93.
[3] "Markoff processes and potentials 11." III. J. Math. 1(1957).316-369.
[4] "Markoff processes and potentials Ill." 111. J. Math. 2 (1958),151-213.
Kiyosi Ito: [ 1] "Stochastic integral." Proc. Imp. Acad. Tokyo 20 (1944), 519-524.
[2] "On a formula concerning stochastic differentials." Nagoya Math. J. 3 (1951). 55 -65.
[3] "Multiple Wiener integral." J. Math. Soc. Japan 3 (1951), 157-169.
K Itd and H. P. McKean, Jr.: [I] Diffusion processes and their sample paths. Berlin, Springer-
Verlag. 1974.
K. Ito and S. Watanabe: [1] "Transformation of Markov processes by multiphcative func-
tionals." Ann. Inst. Fourier Grenoble 15, 1, (1965), 15-30.
H. L. Jackson: [I] "Some results on thin sets in a halfplane." Ann Inst. Fourier Grenoble 20
(1970). 201-218.
Klaus Janssen: [I] Martin Boundary and Hr Theory of Harmonic Spaces. Lecture Notes in
Mathematics 226. Berlin, Springer-Verlag, 1971, pp. 102- 151.
[2] "A co-fine domination principle for harmonic spaces." Math. Ztschr. 141(1975),185-191.
Barge Jessen : [ I ] "The theory of integration in a space of an infinite number of dimensions."
Acta Math. 63 (1934), 249-323.
Shizuo Kakutani: [1] "On Brownian motions in n-space." Proc. Imp. Acad. Tokyo 20 (1944),
648-652.
[2] "Two dimensional Brownian motion and harmonic functions." Proc. Insp. Acad. Tokyo
20 (1944), 706-714.
[3] "Markoff process and the Dirichlet problem." Proc. Japan Acad. 21 (1945), 227-233
(1949).
Robert Kaufman and tang-Mei Wu: [I] "Parabolic potential theory." J. Differential Equations
43 (1982). 204- 234.
0 D Kellogg: [1] "Unicith des fonctions harmoniques." C. R. Acad. Sci. Paris 187 (1928).
526-527.
[2] Foundations of Potential Theory. Berlin, Springer-Verlag. 1929.
John T. Kemper: [1] "Temperatures in several variables. Kernel functions, representations,
and parabolic boundary values." Trans. Amer. Math. Soc. 167 (1972), 243- 262.
A. Khtntchme: [I] "Asymptotische Gesetze der Wahrscheinlichkeitsrechnung." Erg. Math.
Grenzgebiete 2/4 (1933), Berlin, Springer-Verlag.
Frank B. Knight: [I] Essentials of Brownian motion and diffusion. Math. Surveys 18. Provi-
dence, Amer. Math. Soc., 1981.
A. Kolmogoroff: [1] "Grundbegriffe der Wahrschleinlichkeitsrechnung." Erg. Math. Grenzge-
hiete 2/3 (1933), Berlin. Springer-Verlag.
Adam Koranyi and J. C Taylor: [1] "Fine convergence and parabolic convergence for the
Helmholtz equation and the heat equation." 111. J. Math.
K. Krickeberg: [1] "Convergence of martingales with a directed index set." Trans. Amer.
Math. Soc. 83 (1956), 313-357.
Hiroshi Kunita and Shinzo Watanabe: [ I ] "On square integrable martingales." Nagoya Math. J.
30 (1967), 209 -245.
Hiroshi Kunita and Taken Watanabe: (1] "Markov processes and Martin boundaries Part I."
Ill. J Math. 9 (1965). 485-526.
[2] "On certain reversed processes and their applications to potential theory and boundary
theory " J. Math. Mech. 15 (1966). 393-434.
Charles W Lamb: [1] "A note on harmonic functions and martingales." Ann. Math. Stat.
Bibliography 823

42 (1971), 2044- 2049.


N S. Landkof: [1] Foundations of modern potential theory (translation from his Russian book
of 1966: OCHOBbI coepeMeHHOB TeopHH noTeuuxana) Berlin, Springer-Verlag. 1972
H. Lebesgue: [1] "Sur les cas d'impossibilite du probleme de Dirichiet." C. R. Seances Soc.
Math. Fr. (1912).
[2] "Sur Ie probleme de Dirichlet." C. R. Acad. Sci. Paris 154 (1912), 335-337.
[3] "Conditions de regularite, conditions d'irregularite, conditions d'impossibilite dans le
probleme de Dirichlet." C. R. Acad. Sci. Paris 178 (1924), 349-354.
J. Lelong: [I] "Etude au voisinage de la frontiere des fonctions surharmoniques positive dans
un demi-espace." Ann. Sci. Ecole Norm. Sup. (3) 66 (1949), 125-159.
Paul Levy: [I] "Propnetes asymptotiques des sommes de variables aleatoires enchainees"
Bull. Soc. Math. Fr. 59 (1935), 1-32.
[2] Theorie de !'Addition des Variables Aleatoires. Pans, Gauthier-Villars, 1937.
[3] "Le mouvement brownien plan." Amer. J. Math. 62 (1940), 487-550.
[4] Processes Slochastiques et -Mouvement Brownien. Paris. Gauthier-Villars, 1948 (2d. ed.,
1965).
L. Lichtenstein: [1] "Neuere Entwicklungdes Potentialtheone. Konforme Abbildung." Encykl.
Math. Wiss. 11C3 (1919), 177- 377.
J. E. Littlewood: [1] "Mathematical Notes (8): On functions subharmonic in a circle (11)."
Proc. London Math. Soc. (2) 28 (1928), 383-394.
Alfred J. Maria: [1] "The potential of a positive mass and the weight function of Wiener."
Proc. Nat. Acad. Sci. USA 20 (1934), 485-489.
A. A. Markov: [I] "Extension of the law of large numbers to dependent events" (Russian).
Bull. Soc. Phys. Math. Kazan (2) 15 (1906), 135-156.
R. S. Martin: [I] "Minimal positive harmonic functions." Trans. Amer. Math. Soc. 49 (1941),
137-172.
H. P. McKean, Jr.: [1] Stochastic Integrals. New York, Academic, 1969.
Jean-Francois Mertens: [I] "Theorie des processus stochastiques generaux. Applications aux
surmartingales." Ztschr. Wahrscheinlichkeiistheorie 22 (1972), 54-68.
Paul-Andre Meyer: [I] "Fonctions multiplicatives et additives de Markov." Ann. Inst. Fourier
Grenoble 12 (1962), 125-230.
[2] "Une presentation de la theorie des ensembles sousliniens. Applications aux processus
stochastiques." Sem. Brelot-Choquet-Deny (Theorie du potentiel) 7 (1962- 1963).
[3] "Interpretation probabiliste de la notion d'energie." Sem. Brelot-Choquet-Deny (Theorie
du Potentie!) 7 (1962-1963).
[4] "A decomposition theorem for supermartingales." Ill. J. Math. 6 (1962), 193- 205.
5] "Decomposition of supermartingales: the uniqueness theorem." Ill. J Math. 7 (1963),
1-17.
[6] Probability and Potentials. Waltham, Blaisdell, 1966.
[7] Processus de Markov. Lecture Notes in Mathematics 26. Berlin, Springer-Verlag, 1967.
[8] Guide detaille de la theorie generale des processus. Sem. Prob. It. Lecture Notes in
Mathematics 51. Berlin, Springer-Verlag, 1968.
[9] Processus de Markou: la frontiere de Martin. Lecture Notes in Mathematics 77. Berlin,
Springer-Verlag, 1968.
P. A. Meyer, R. T. Smythe, and J. B. Walsh: Birth and death of Markov processes Proc. Sixth
Berkeley Symp. Math. Stat. Prob. 1970/71 Vol. 111. Berkeley, Univ. of California, 1972,
pp. 295-305.
Gabriel Mokobodzki and Daniel Sibony: "Sur une propriete charactenstique des cones de
potentiels." C. R. Acad. Sci. Paris 266 (1968), 215-218.
Jurgen Moser: [I] "A Harnack inequality for parabolic differential equations." Comm Pure
App!. Math. 17 (1964), 101-134.
Linda Naim (= Linda Lumer-Naim): [1] "Sur le role de la frontiere de R. S. Martin dans Is
theorie du potential." Ann. Inst. Fourier Grenoble 7 (1957), 183-28 1.
Edward Nelson: [I] Dynamical Theories of Brownian Motion. Math. Notes Princeton U. Press,
Princeton, 1967.
J. Neveu: [I] Martingales a Temps Discret. Paris, Masson, 1972.
A. A. Novikov: [1] "On an identity for stochastic integrals." Theory Prob. App!. 17 (1972),
717-720.
M. Parreau: [ I ] "Sur les moyennes des fonctions harmoniques et analytiques et la classification
des surfaces de Riemann." Ann. Inst. Fourier Grenoble 3 (1951), 103-197.
824 Bibliography

K R. Parthasarathy: [I] "Square integrable martingales orthogonal to every stochastic inte-


gral." Stochastic Proc. Appl. 7 (1978). 1- 7.
Anthony L. Peressini: [I] Ordered Topological Vector Spaces. New York, Harper, 1967.
0. Perron: [I] "Eine neue Behandlung der ersten Randwertaufgabe fur Au = 0." Math.
Ztachr. 18 (1923), 42 - 54.
1 Petrowsky : [ I ] "Uber das Irrfahrtproblem." Math. Ann. 109 (1933- 1934). 425 -444.
[2] "Zur ersten Randwertaufgabe der Warmeleitungsgleichung." Comp. Math. 1(1934-1935),
383 419
H. Pomcarc [I] "Sur les equations aux derivees partielles de la physique mathematique."
Amer. J Math. 12 (1890), 211 294.
[2] Theorie du Potentiel Newronien. Paris, Gauthier-Villars. 1899.
S. D. Poisson [I] "Addition au memoire precedent et au memoire sur la maniere d'expnmer
les fonctions par les series de quantites periodiques." J. Ecole Roy. Polylechnique 19 cattier
12(18223), 145-162.
Sidney C. Port, Charles J Stone: [I] Brownian Motion and Classical Potential Theory. New
York. Academic. 1978.
N. Privalov: [I J "Boundary problems of the theory of harmonic and subharmonic functions
in space" (Russian). Mat Shornik 3(1938), 3-25.
Tibor Rado [I] "Subharmonic functions." Erg. Math. Grenzyehiete 5 (1). Berlin, Springer-
Verlag. 1937. (Reprinted New York, Chelsea, 1949.)
K. Murali Rao: [ 11 "On decomposition theorems of Meyer." Math. Scand. 24 (1969).66 78.
[2] "Quasi martingales." Math. Scand. 24(1969), 79-92.
[3] "On Green functions in ) 2 " Israel J. Math. 19 (1974). 313 328.
[4] Brownian motion and classical potential theory. Math Inst.. Aarhus Univ., 1977.
Robert Remak. [I] "Uber potentialkonvexe Funktionen." Math. Ztschr. 20(1924), 126- 130.
F. Ricsz: [ I] "Sur certains systemes singuliers d'equations integrates." Ann. Ecole Norm.
Sup. (3) 28 (1911), 33 62.
[2] "Sur les functions subharmoniques et leur rapport a la theonc du potentiel I."Acta Math.
48 (1926). 329- 343
[3]II. Actor Math. 54 (1930), 321-360.
M Riesz [ I ] " Integrales de Riemann-Liouville et potentiels." Actor Litt.Sci. Univ. Szeyed
9 (I) (1938), 1-42
G. Robin. [1] "Sur la distribution de 1'electricite a la surface des conducteurs fermes et des
conducteurs ouverts." Ann. Sri. Ecole Norm. Sup. (3) 3 (1886), Supp. I - 58.
H. A Schwarz: (1] "Zur Integration der partiellen Differentialgleichung ? u/dx2 + 81u/l i'2
= 0." J. Reme An yew. Math. 74 (1872), 218-253.
J L. Snell [ I ] "Application of martingale system theorems." Trans. Amer. Math. Soc. 73
(1952). 293 312.
M. Souslin [I] "Sur une definition des ensembles mesurables B sans nombres transfinis."
C R. Acad. Sc!. Paris 164 (1917). 88-91.
W. Sternberg: [ I ] "Uber die Gleichungder Wiirmeleitung." Math. Ann... 101 (1929), 394- 398.
G Szego [1] "Bemerkung zu einer Arbeit von Herrn M. Fekete: ber die Verteilung der
Wurzeln bet gewissen algebraischen Gleichungen mit ganzzahligen Koeffizienten." Math.
Ztschr 21 (1924), 203-208.
E Szpilrajn (= E. Marczewski): [1] "Remarques sur les fonctions sousharmoniques" Ann.
Math. (2) 34 (1933), 588-594.
J C. Taylor. [I] An elementary proof of the theorem of Fatou-Naim-Doob. /980 Sent on
Harmonic Analysis, Montreal, 1980, pp. 153-163.
William Thomson (= Lord Kelvin): "Extraits de deux lettres addressees a M. Liouville."
Jr Math Pures. Appl. (1) 12 (1847), 256-264.
E. C Titchmarsh I I ] Introduction to the Theory o/ Fourier Integrals. Oxford, Clarendon Press.
1937.
M. Tsuji. I I] Potential Theory in Modern Fwtction Theory. Tokyo, Maruzen. 1959.
A Tychonoff: [I] "Thboreme d'unicite pour 1'equation de la chaleur." Mat. Shornik 42 (1935),
199 216
C. de la Vallee Poussin: [ I ] "Extension de la methode du balayage de Poincare et probleme de
Dirichlet." Ann. Inst H. Poincare 2 (1932), 169-232.
[2] "Les nouvelles methodes de la theorie du potentiel et le problems generalise de Dirchlel "
Act. Sci. Ind. 578 (1937). Paris, Hermann.
[3] "Potentiel et probleme generalise de Dirichlet." Math. Gazette 22 (1938), 17-36.
Bibliography 825

[4] "Points irregulier. Determination des masses par les potentiels." Acad. Beiq. Bull. Cl. Sri.
(5) 24 (1938), 368-384.
Florin Vasilesco: [I] "Sur la continuite du potentiel i1 travers les masses, et Ia demonstration
d'un femme de Kellogg." C. R. Acad. Sri. Paris 200 (1935), 1173 -1174.
[2] "Sur une application des families normales de distributions de masse " C. R. Acad. Sci.
Paris 205 (1937), 212-215.
[3] La notion de point irrbgulier dans Ic probleme de Dirichlet. Act. Sci. Ind. 660 (1938),
Paris, Masson.
Jean Ville: [I] Etude Critique de la Notion de Collectij. Pans, Gauthier-Villars, 1939.
Michel Weil: [I] "Proprietes de continuite fine des fonctions coexcessives." Ztschr. Wahrsrh-
einlichkeitstheorie verve. Geb. 12 (1969), 75- 86.
John Wermer [1] Potential Theory. Lecture Notes in Mathematics 408. Berlin, Springer-Verlag,
1974.
D. V. Widder: [1] "Positive temperatures on an infinite rod." Trans. Amer. Math. Soc. 55
(1944), 85- 95.
[2] The Heat Equation. New York, Academic, 1975.
Norbert Wiener: [1] "Differential space." J. Math. Phys. Mass. Inst. Tech 2(1923),131-174.
[2] "Certain notions in potential theory." J. Math. Phys. Mass. Inst. Tech. 3(1924),24-51.
[3] "The Dirichlet problem." J. Math. Phys. Mass. Inst. Tech. 3 (1924),127-146.
[4] "Note on a paper by O. Perron." J. Math. Phys. Mass. Inst. Tech. 4 (1925), 21-32.
Jang-Mei G. W u : [ 1 ] "On parabolic measures and subparabolic functions." Trans. Amer. Math.
Soc. 25 (1979), 171-185.
S. Zaremba: [I] "Sur Ie principe de Dirichlet." Acta Math. 34 (1911), 293 316.
Notation Index

08,
68+,

68, xxiv b
Hf, n
Hf, Hrk 101
7 L , Z , 7L xxiv AD 114
00 xxiv GD 132
1 - Al xxiv WD 136
6) xxiv D(µ4_) 142
IN Xxiv LP(4_) 144
A-B xxiv St, Si 145
L(k'(D) xxiv S 146
nN 3 Sm 148
A 3 SP 149
D. 3 Sqb 150
L(u, 6) 4 Smqb, Spqb 150
A(u, , 6) 4 Ss, Sms, Sps 151
'IN 4 SD 156
4 f lim 166
G 6 Af 167
a 7 u* 172
hD(S, A) 13 u# 180
GB 14 R 188
K(q, 14
fAD 191
Pi 15 193
fGD
TB 16 K 196
L`(uB-) 27 DM, aMD 197
D(JB-) 27 a;MD 200
LM, GM 35 [u, v], 1IN11 227
u 37 9+ 227
R, 38 9 229
0v04 39 -9 (u, v), -9 (u) 231
TB 99 BLD 233
GM", LMh 99 C*, C*, C 240
hRA 99 r(A) 252
IDu0A 99 Al A 262
PWB, PWBh 100 ftN 262
82_8 Notation Index

Hm1 m5
264 E{xj.F} 391
265 QS, TP 414
Hm l m,v
266
[S]. 415
G 266
O*, O** 421

271
Dn[f;g,h] 445
274
rsr, Ti 456
TB
275
GM, LM . 458
A(u, q, c)) 276
Rz1., 459
Aa(ti, rj) 276 R^1., 460
285 Tsr, LM, GM 500
PI
st 525
LI(NB ) 290
290 'S', S±, S+ 526
GM. LM 295 'S 528
S 529
296
'Sm, Sm, 'Sp 530
GD, 296
'Sqb, 'Smgb, 'Spgb 531
6b 298 Sqb, 'ST, Ss 532
pflim 308 Sms 533
APt. At' 309 GX 561
p*flim 309 t(t, , ?) 572
rI N 320 590
fU 326
592
UN 326
r 599
GMh, GMh 329 r, rZ, ro 600
329 r', r, r2 602
hTB htA 329 1'a 604
FIj, Hf 329 Sms00 716
Smsf 717
D 332
r,, (w) 719
343
u5(t) 728
351
Pf1r, PfpD Pf(,D 359 fD 728
362
G, (t, . h) 735
LP(µp_)
363
.a1(4) 741
4+/(X) 756
K 364
41+ 769
DM, 366
b ,#.+ 771
371
W. 772
387
E{ x}, El x;A 390
Index

References are to sections, say 1.11.3 or, for the Appendixes, A.1.2. The index
covers neither the Historical Notes nor the Bibliography.

Above: a point of (8", relative to a set, I.XV.I


Abscissa hyperplane: of ft", 1.XV.1
Absolute probability function (entrance law): 2.Vl. I; of Brownian motion,
2.X.13
Adapted: family of functions, 2.I.1; stochastic process, 2.1.8
Analytic function: generalizations of Liouville's theorem, 1.11.2, 1.11.13;
-J° is subharmonic, Hadamard three circle theorem, 1.I1.11; generaliza-
tion of Cauchy's removable singularity theorem, 1.V.5; generalization of
the maximum theorem, I.V.7; composition with Brownian motion,
2.Vlll.l4
Analytic set: over a paving, A.1.2; over a product paving, A.1.3; projection
characterization of sJI(Y), A.I.5, A.I.7; d(sl), A.1.6; inverse image under
a measurable transformation, A.1.8; in a metric space, A.1.12, A.1.13;
hitting of by a progressively measurable process, 2.11.4
Andre reflection principle: 2.VII.8
Appell transformation: 1 .XV.16
Approximation: of a superharmonic function by infinitely differentiable
superharmonic functions, 1.11.8, I.IV.10; of a potential by continuous
potentials, 1.V.9; of a continuous function by potential differences,
1.V11.9; of a superparabolic function by infinitely differentiable super-
parabolic functions, 1.XV.14, I.XVII.7(e); of a random variable by its
conditional expectations, 2.1.5, 2.111.14

definition, in particular for a half-space or interval, 2.VII.9, 2.VII.11;


as a Green function, 2.Vl1.9, 2.1X.17
"D, lD: definition, 2.VII.10; i"D = Go, 2.IX.17
Backward parabolic equation : 2.1 X.17
Baire property of the fine topology: classical context, I.XI.I ; parabolic
context, I.XVII.9
Baire null space: A.1.11
Balayage: see Sweeping
830 Index

Ball
Classical context: the Green function and the Poisson integral, 1.11.1;
Riesz-Herglotz theorem, 1.11.14: L'(µ8_) and D(µB_) classes of har-
monic functions, 1.11.14, 1.1X.12; Fatou boundary limit theorem,
1.11.15; minimal harmonic functions, 1.11.16; reductions. 1.111.4;
potential of a uniform boundary distribution, I.IV.2. PWB solutions,
I.V1II.2: universal and universal internal resolutivity of the boundary,
I.V111.9, lattices of relative harmonic functions. I.IX.12; Martin
boundary. 1.X11.3: classical versus minimal fine boundary limit
theorems for relative harmonic functions, 1.X11.20: ball capacity.
1.X111.13; conditional Brownian motion in, 2.X.9
Parabolic context: regularity of the boundary, parabolic ball and the
irregularity of its highest point, I.XVlll.6
Band: of a vector lattice, A.I11.8; projection on, A.111.9; generated by a
singleton. A.111.I I
Barrier
Classical context: definition, local nature, versus weak-, at oo, I .V11l.12,
relation with boundary point regularity, I.V111.13, 1.VI1I.14, 1.VIll.16,
Poincare- Zaremba examples, 1.VIll.12, 1.Vlll.15
Parabolic context: definition, local nature, versus weak-, relation with
boundary point regularity, I.XVII1.3
Below: a point of 1W relative to a set, I .XV. I
BLD function: 1.X111.6
Boundary limit theorems
Classical context: for the Poisson integral, semicontinuous boundary
function, 1.11.1; ratio, for harmonic functions on a ball, 1.11.15, 2.X.8:
for superharmonic functions at an irregular boundary point, 1.X1.21:
ratio, for superharmonic functions on a Martin space, 1.X11.13,
1.X11.14, 1.X11.18, 1.X11.19; classical versus fine topology approach
to a ball boundary, 1.X11.20: for potentials on a half-space, 1.X11.22,
1.X11.23: ratio, for superharmonic functions, along stochastic process
paths. 2.Vl.4, 2.1X.7, 2.1X.13, 2.X.8, 3.111.4, 3.111.5: for Dirichlet
solutions, 2.1 X.13, 3.11.2
Parabolic context: for the Poisson integral on an interval or slab, semi-
continuous boundary function, I.XV.9, I.XVI.1 ; ratio, for superpara-
bolic functions on a slab, 1.XVI.7, I.XV111.I5: for superparabolic
functions at an irregular boundary point. 1.XVIII.17. ratio, for super-
parabolic functions on a Martin space, with applications to a slab and
to the fourth quadrant of 082, l.XIX.13-l.XIX.15; ratio for super-
parabolic functions, along stochastic process paths. 2.1X.7, 2.1X.13,
2.X.12, 3.11
Brownian motion (see also Conditional Brownian motion and Space-time
Brownian motion): definition, 2.VlI.2; path continuity, 2.VII.3; right
continuity and predictability of-filtrations, predictability of-optional
times, continuity properties of supermartingales relative to-filtrations,
Index 831

2.VI1.4, 2.VI1.6, 2.V111.8, 2.X.2(c); in an open set, relation between


transition densities and Green functions, 2.V11.5, 2.VI1.l1, 2.1X.9,
2.IX.17; law of large numbers for and path length of, 2.VI1.5; zero-one
law for, 2.VI1.6; tied down, the Brownian bridge, 2.VII.7; in an interval
and the evaluation of parabolic measure, 2.VIl.11, 2.V11.12; under change
of parameter, 2.V111.9: composition of a function with, 2.V1I1.13,
2.VII1.14, 2.1X.11, and, if the function is superharmonic, 2.IX.2, 2.1X.3,
2.IX.6-2.IX.8, 2.1X.11-2.IX.13, 2.IX.16, 2.X.2; excessive functions and
measures for, 2.IX.6, 2.IX.8, 2.!X.18; for large parameter values, 2.IX.5;
hitting of sets by, 2.IX.4, 2.1X.5, 2.1X.9, 2.!X.10; from an irregular
boundary point, 2.IX.20; killed, 2.X.2, 2.X.6; to and from Martin
boundary points, 3.111.1, 3.111.2

Canonical stochastic process: definition and construction, 2.1.10; Markov


case, 2.VI.3, 2.V1.5, 2.V1.6, and Brownian motion case, 2.VI1.3, 2.VI1.9
Capacitary measure and capacitary potential: see Equilibrium measure and
equilibrium potential
Capacity: Choquet-definition, A.I1.1; Choquet capacitability theorem,
A.ll.3: projection example, A.11.5; generated by a strongly subadditive
set function, in particular by a topological precapacity, A.II.7, A.11.8;
generated by the reduction operation, I.V1.2(e), 1.VI.3(k), I.VI.4(k),
I.VI.5, LXVII.16(o), I.XVII.17(o); generated by upper PWB solu-
tions, I.V1I1.6(h); classical, I.XIII.I, I.XIII.10-I.XIII.13, logarithmic,
1.X111.18
Chapman-Kolmogorov equation: A.VI.3, 2.V1.1, 2.VI.5, 2.VII.1
Charge: definition, A.IV.7; energy, l.Xlll.l, 1.X111.3; energy is positive,
1.X111.7, I.XIII.8
Choquet capacity: see Capacity
Choquet topological lemma : A. V I 1 I.3
Closability: in martingale theory, 2.111.1-2.111.3
Cluster set: of a function along Brownian and space-time Brownian paths,
2.VII.6, 2.1X.5, 2.1X.15, 2.X.2, 3.11.1-3.11.4
Condenser: definition, l.XII1.1
Conditional Brownian motion (see also Brownian motion): definition, transi-
tion density, lifetime, excessive functions for, 2.X.1; in terms of Brownian
motion, 2.X.2, 2.X.3; at path lifetimes, 2.X.4; from an infinity of the
conditioning function, 2.X.5: time reversal of, 2.X.6; and the PWB
method, 2.X.7, 3.11.1-3.11.4: last hitting distribution, 2.X.10; tail a alge-
bra, 2.X.11: parameter set R, 2.X.13; application to boundary limit
theorems, see Boundary limit theorems; in a Martin space, see Martin
boundary
Conditional expectation: properties. 2.1.4; continuity theorem for, 21.5.
Fatou lemma for, 2.1.6; dominated convergence theorem for, 2.1.7; itera-
tion of at optional times, 2.111.2(a); sample functions of the process of
conditional expectations relative to the a algebras of a filtration, 2.IV. I
832 Index

Conditional Markov process: definition, 2.V1.13; see Conditional Brownian


motion for the application to Brownian motion
Conditional inequalities: in martingale theory, 2.I11.10
Cone: definition. A.I11.3: specific order generated by, A.1I1.4
Continuity properties: of a progressively measurable process sample func-
tion, 2 11.5. 2.11.6: of a supermartingale sample function, 2.IV.l: of a
Brownian motion sample function, 2.V11.3; of the composition of a func-
tion with a Brownian motion sample function, 2.1X.11, 2.1X.12, 2.1X.16,
2.X.2
Convergence (see also Fundamental Convergence Theorem): of a directed set
of harmonic functions, Harnack's convergence theorem, 1.11.3; of an
upward directed set of superharmonic functions, 1.I1.4, 1.IV.4; of families
of superparabolic functions, I.XV.12; of families of supermartingales,
2 111 5. 2 IV.4: forward, of a supermartingale, 2.111.13-2.111.15, 2.IV.3,
backward, of a supermartingale, 2.111.16, 2.111.17
Coparabolic function (see also Parabolic function): definition. I.XV.2:
coparaholic polynomials, I.XV.3: Coparaholic polynomials and martin-
gale theory. 2.IX.2
Copotential (see also Potential): definition in the parabolic context. I.XVI1.5
Cylinder: superparabolic function on, I.XV.15, parabolic Dirichlet problem
on, I XV1lI.4

D class
Classical context (D(pn )): of h-harmonic functions on a ball. 1.11.14.
I IX 12. of functions on an open set, in particular h-harmonic and
positive h-suhharmonic functions, I.IX.3
Parabolic context (D(µn_ )): of parabolic functions on a slab, I.XVI.6; of
functions on an open set, 1.XVIII.19
Probability context: of stochastic processes, 2.11 11, includes right closed
positive submartingales, 2.111.6, 2.V.3: includes potentials generated by
increasing processes, 2.IV.6
Combined context: notation, 3.1.3: the PWB method. 3.1.4: S. n D =
Smgh. 3 1 5. SPyg, = D n S'. 3.1.9
Decomposition
See Krickeberg decomposition, LP. Rao decomposition, and Riesz-Herglotz
representation for decompositions in various contexts of a vector lattice
element into a difference between positive elements, see A.1I1.5 for the
abstract version of such a decomposition, and see A.IV for such decom-
positions in measure theory. See Lattice for decompositions of a vector
lattice into bands.
Riesz: of a superharmonic function. 1.1.8, I.IV.8, I.IX.I 1 ; of a su-
perparabolic function, I.XVI1.7: of a supermartingale, 2.111.21,
2.V.8
Natural order: for superharmonic functions, 1.111.7; for superparabolic
functions, I.XVII.2: for supermartingales, 2.111.19, 2.V.5
Index 833

-: of optional times, 2.11.12


-: of a supermartingale into continuous and jump components, 2.IV.22,
2.I V.23
-: of conditional Brownian motion, 2.X.4
Meyer: of a supermartingale, 2.IV.11; of a submartingale, 2.IV.12
d, A: definitions, 1.XV.2
Derivate: in the convex and symmetric senses, and in the convex and sym-
metric variational senses, A.IV.12
Dirichlet integral: 1.XIII.6
Dirichlet problem: see PWB method
Dominated convergence theorem: A.IV.10, for conditional expectations, 2.1.7
Domination principle: classical context, I.V.10. 1.XI.23, I.XII1.2: parabolic
context, 1.XVII.5, l.XVII1.16. martingale theory context, 2.IV.13
Downcrossing inequalities: 2.111.12, 2.111.23, 2.IV.18(o), 2.IV.19(o)

Energy: electrical, l.X1II.l ; energies and mutual energies of charges,


I.XI11.2, 1.X1I1.3; dependence of energy on the containing set, 1.XII1.5;
the Dirichlet functional.9 and Hilbert space methods, I .XII1.6; positivity
of, 1.X111.7, 1.XIII.8, 2.IX.17; Gauss minimum problems, I.XII1.14;
relative to 182, I.XIIl.16; of a supermartingale potential, 2.IV.21
Entrance law: see Absolute probability function
Entry time (see also Hitting time): definition, 2.11.4; of a predictable set,
2.11.9
Equilibrium measure and equilibrium potential: I.XIII.1, I.XIII.10. I.XII1.12;
relative to I82, 1.XI11.18
Essential order: of measurable functions, A.IV.9, 2.1.3; convergence,
A.IV.10; of stochastic processes, 2.1.8, 2.V.1, 2.V.2, 2.V.5; convergence
of supermartingale families, 2.111.5
Euclidean boundary: definition, Notation and Conventions; resolutivity and
set of regular points in the classical context, 1.VII1.4; resolutivity and
set of regular points in the parabolic context, I.XVIII.I
Evanescent set: definition, 2.1.8
Evans-Vasilesco theorem : 1 . V .8
Excessive functions and measures (see also Invariant-): definitions, 2.VI.10;
excessive functions as generators of supermartingales, 2.VI.I I ; excessive
functions defined by hitting probabilities, 2.VI.12; excessive functions
for [space-time] Brownian motion versus positive [superparabolic] super-
harmonic functions, 2.IX.6, 2.IX.8; a excessive functions for Brownian
and space-time Brownian motion, 2.IX.16; excessive measures for
Brownian and space-time Brownian motion, 2.IX.18
Extension: of a superharmonic function through a polar set, 1.V.5; of a
Green function GD to GD, 1.VII.4, I.VIII.18, I.VIII.19; of a parabolic
function defined on a truncated cylinder, 1.XV.17; and contraction of
OD, I.XVII.4; of a superparabolic function through a polar set, I.XVII.8;
of a Green function GD to GD, 1.XVIII.9
834 Index

Fatou boundary limit theorem: see Boundary limit theorems


Fatou's lemma: A.IV.10, for conditional expectations, 2.1.6
Filtration of a measurable space, 2.1.1; composition .F(T) with an
optional time T, 2.11.1 ; predictable-and F(T- ), 2.11.7; choice of for a
supermartingale. 2.111.1, 2.IV.1; choice of for a Markov process, 2 VI.2:
right continuity of. for a Markov process, 2.V1.8; right continuity and
predictability of Brownian motion filtrations, predictability of Brownian
motion optional times, continuity properties of supermartingales relative
to Brownian motion filtrations. 2.V11.4. 2.VI1.6, 2.VIII.8, 2.X.2(c). for
conditional Brownian motion, 2.X.2(c), (d)
Fine topology (see also Thinness)
Classical context: definition and basic properties, I .X1.I-1.X1.3, I.XI.5,
and, in probabilistic formulation, 2.IX.15. character of derived set,
boundary, and interior, I.XI.I, 1.XI.6:-limits versus Euclidean topol-
ogy limits. 1.X1.9; identification of a derived set by a special function
u''. I X1.10: quasi Lindelof property. 1.X1.11; relation to boundary
point regularity, 1.X1.12; -open sets as domains for superharmonic
functions, I.X1.19: minimal-fine topology, 1.X11.1 I, 1.X11.12, 1.X11.17,
and, in probabilistic formulation, 3.111.3; minimal-fine topology limits
versus Martin topology limits at the Martin boundary, I.XII.15,
I XII 16: normal boundary limit points of a half-space subset versus
minimal-fine boundary limit points of the set, 1.X11.22
Parabolic context: definition and basic properties, I.XVII.9, I.XVII.12.
I XVIII 10. and, in probabilistic formulation, 2.IX.15; character of a
derived set and its identification by a special function ta', I.XVII.15,
I XVII 16(r). I.XVII.17(r): quasi Lindelof property. I.XVII.19:
minimal-fine coparaholic limits versus Martin topology limits at the
Martin boundary. l .XIX. l2
Probability context: in Q8' x 0, 2.IV.16
Finite dimensional distributions (see also Stochastic process): of a stochastic
process. 2 1 10
First boundary value problem: see PWB method
Forward parabolic equation: for Brownian motion, 2.1X.17
Fundamental Convergence Theorem: classical context. 1.111.3, 1 VI I. I AI 7,
1 XIV 3. parabolic context. I.XVII.2. I.XVII.l3- I.XVII.l5. martingale
theormy context. 2.IV 5

Gh sets: homeomorphs of, A.1.9


Gauss Integral Theorem: 1.1.6, 1.1.7
Gauss minimum problems: 1. X 111.14
GM (see also LM)
Classical context: definition. 1.111.1. 1111.2. I.XIV.3: of a potential,
1111 1, I IV 3. 1 IV 8(a), relation with the reduction operator, 1.1114.
111 6: unaffected by deletion of polar sets, 1.V.5; relation with
1

Dinchlet solutions. I VIII 3. I Vill.I8(c): relation with the r operator,


I VIII I I
Index 835

Parabolic context (GM, OM): definition, 1.XVII.1; of a potential,


I.XVII.5; relation with Dirichlet solutions, I.XVIII.I
Martingale theory context: definition and basic properties, 2111.20,
2.111.21, 2.IV.14; of a potential, 2.111.21
Graph: of an optional time, 2.11.1;-predictability, 2.11.9
Green function
Classical context (GD): preliminary definition, 1.1.5, 1.1.8; of a ball, 1.1I.1,
definition, I.VII.!; in terms of a Dirichlet solution, 1.VII.1, I.VI11.3,
I.VII1.18, I.VHII.19; extremal property, I.V11.2; boundedness property,
1.V11.3; symmetry property, 1.V11.4; extension to GD, I.VII.4,
1.VIII.l8, 1.VIII.19; vanishing of C) at 8D, 1.V11.4, l.X1I.14,
1.X11.18; minimality of G,(-, C), 1.VII.10; (N=2) with pole oo, the
Robin constant, I.VI11.20, 1.X111.18; the case N = 1, I.XIV.6; in terms
of GD, I.XVII.18; probabilistic expression for, 2.1X.17
Parabolic context (GD, GD): preliminary definition, 1.XV.4, 1.XV.7; of an
interval, 1.XV.8; definition, examples, properties, GD = GD, I.XVII.4;
minimality of I.XV11.8; vanishing of i;) at 815, 1.XVI1.14;
in terms of a Dirichlet solution, 1.XVIII.1; extension to GD , I.XVIII.9;
probabilistic expression for, 2.IX.17
Greenian set: definition, trivial in l when N > 2, but I82 is not one, 1.1I.13;
conditions that an open subset of I8Z be Greenian, 1.V.6, 1.VII.7 in non-
probabilistic terms, and 2.IX.10 in probabilistic terms

Hadamard three circle theorem: 1.11.11


Hahn decomposition: A.IV.6
Half-space
Classical context (-of I8"): Green function, harmonic measure, Poisson
integral and Riesz-Herglotz representation, I.V1II.9; Martin boundary,
I .XII.3 ; minimal thinness, minimal-fine versus nontangential limits at
a boundary point, I XII.12, ratio boundary limit theorems for harmonic
functions, 1.X11.21 ; normal versus minimal-fine boundary limit func-
tions, 1.XII.22; boundary limit function of a potential, 1.X11.23;
Brownian motion, 2.VII.9
Parabolic context (-of C'8", see also Slab): upper and lower half-space of
ftN, 1.XV.1; Green function, l.XV1.l, I.XVI1.4; Martin boundary,
1.XIX.8, l.XIX.9
Harmonic function (see also Subharmonic function and Superharmonic func-
tion): definition, average properties, 1.1.3; maximum-minimum theorem,
1.I.4; Poisson integral, 1.1I.1; Harnack inequality, 1.11.2; Harnack con-
vergence theorem, 1.11.3; classes L", D, and the Riesz-Herglotz theorem,
1.11.14, 1.IX.3, 1.1X.4; ratio boundary limit theorem in a ball, 1.11.15;
minimal, 1.11.16, 1.X11.5; removable singularity set, 1.V.5; relative,
l.VIIl.l ; boundary poles, 1.X.7, I.XII.5; Martin representation, 1.XII.9,
3.111.6; vector lattice Sand its band decompositions, 1.1X.7, 1.IX.9-
I.IX.12, 3.1.12; first boundary value (Dirichlet) problem, see PWB
method; composition with Brownian motion, see Superharmonic function
836 Index

Harmonic measure (p'): for domains with smooth boundaries, 1.1.8; null
sets. I.Vl1I.5; of fool, LVIII.5; definition, l.V1I1.8; for ball and half-
space, I.V111.9; dependence on D, 1.VIII.17; relation with the sweeping
kernel. 1.X.2; on a polar set, 1.X.8; relative to an irregular boundary
point, I .X1.22. for the Martin space, 1.X11.10; as a conditional Brownian
motion hitting distribution, 2.1X.13, 2.X.7, 3.11.2
Harnack convergence theorem: classical context, 1.11.3, parabolic context.
I.XV.I I
Harnack inequality: classical context, 1.11.2; parabolic context, 1.XV.11.
1.XVI11.17
Hermite polynomials: 1.XV.3; composition of space-time-with space-
cotime Brownian motion, 2.IX.2
Hitting: by progressively measurable processes, and the dependence on the
choice of process with prescribed finite dimensional distributions, 2.1.9-
2.1.12; hitting, entry, and last hitting times, 2.11.4: the hitting probability
for a Markov process of an analytic set, in particular of an F. set, as a
function of the process initial point, 2.V1.6, 2.1X.4, and the excessive
function so defined, 2.VI.12; by Brownian motion, 2.VI1.6, 2.V11.8,
2.1X.1, 2.1X.4, 2.1X.9, 2.1X.15; of a parabolic semipolar set by space-time
Brownian motion, 2.1X.15: harmonic measure as a conditional Brownian
motion hitting distribution. 2.1X.13, 2.X.7, 3.11.2; capacitary distribu-
tions in terms of last hitting distributions, 2.X.10
Hunt potential theory: 2. V1.10

Increasing process: definition, support of measure generated by, 2.IV.6:


potential generated by. 2.1V.6, 2.IV.8: natural versus predictable, 2.IV.6,
2.IV.7: as the generator of the Riesz measure for a supermartingale.
2.IV.13
Independent increment process: 2.111.2(c). 2.VII.I
Indistinguishable processes: 2.1.8
Integral limit theorems: A. V I I
Internal potential theory limit theorems: classical context, I.XI.4, 1.X11.19;
parabolic context, I.XVIII.14
Interval of definition, upper, lower, lateral boundaries, I.XV.I : para-
bolic Green function, I.XV.8: parabolic measure. I.XV.9, 2.VI1.12;
Dirichlet problem, I.XVI11.1, I.XVIII.6, Brownian motion in, 2.VII.1 I
Invariant excessive functions and measures: definitions, 2. VI.10 ; for Brownian
and space-time Brownian motions, 2.VII.10, 2.1X.8
Inversion in a sphere: the Kelvin transform, 1.11.12, 1. VII I.1: image of a
polar set, IN. 1: and the PWB method, I.VI11.2: and h-harmonic measure
null sets. I.V111.5: and boundary point regularity, I.V111.14
Irregular boundary point (see also Regularity of a boundary point)
Classical context: limit of a superharmonic function, of harmonic measure,
and of a Green function at one, 1.X1.21, 1.X1.22; Brownian motion
from one, 2.IX.20
Index 837

Parabolic context: limit of a superparabolic function, of parabolic mea-


sure, and of a Green function at one, I.XVIII.17, 1.XVIII.18
Iterated logarithm: regularity criterion and Brownian motion law, I.XVIII.6,
2.IX.15
Ito integral: integrand classes, 2.VII1.1, 2.VIlI.2; properties, 2.VIII.3,
2.VIII.6; definition, 2.VIII.5, 2.VIII.7; integration by parts, 2.VIII.1I
Ito's lemma : 2. V I I1.12

Jensen inequality: applied to subharmonic and harmonic functions, 1.11.9,


applied to subparabolic and parabolic functions, 1.XV.12; applied to
conditional expectations, 2.1.4(h): applied to submartingales and mar-
tingales, 2.111.3(c)
Jordan decomposition: A.IV.6

Kelvin transform: see Inversion in a sphere


Kernel: definition, extension from substochastic to stochastic, A.VI.1;
universally measurable extension, A.VI.2
Kolmogorov construction of canonical processes: 2.1.10
Krickeberg decomposition: of a martingale, 2.V.4(c)

L° bounded: stochastic process, 2.1.8; submartingale, 2.V.4; Krickeberg


decomposition of an-martingale, 2.V.4(c)
L° class
Classical context of h-harmonic functions, 1.11.14, I.IX.4,
3.1.3
Parabolic context (L°(4_)): of h-parabolic functions, 1.XVI.6,
I.XVIII.19, 3.1.3
Last hitting time: definition, 2.11.4: Markov process killed at a, 2.VI.15;
Brownian motion killed at a, 2.X.6: capacitary distribution and-, 2.X.10
Lattice
Abstract theory: definitions, countability property. A.III.2; cone, vector
lattice, specific order, A.I11.3-A.I11.5: decomposition property of a
vector lattice, A.111.6; orthogonality, bands, projections, A.111.7-
A.M. 11; order convergence, A.II1.12, A.I11.13, A.IV.10
Application to measure theory: lattice of set algebras, A.IV.1; lattices of
measures, A.IV.4-A.IV.7; absolute continuity and singularity, A.IV.8;
lattices of measurable functions, A.IV.9; essential order and the
corresponding convergence, A.IV.10
Application to classical potential theory: St, S', I.IX.5; S, I.IX.6; Sm,
I.IX.7; SD, I.IX.8; Sgb, Spgb, Smgb, the relation between Smgb and the
PWB method, 1.IX.9; S.,, Sps, S.,, l.IX.10; refinement of the Riesz
decomposition of a positive superharmonic function, I.IX.11, I.IX.12;
S,,, , Smsf, 3.1.12
Application to parabolic potential theory: translation from classical to
parabolic context, 1.XVI I I.19
838 Index

Application to martingale theory: 'S', 'S', S. S+, 2.V.5; 'S, S, 2.V.6:


'Sm, Sm, 2.V.7; 'Sp, Sp, 2.V.8; Se,, Spgb, Smgb in the specific order, and
their primed counterparts, 2.V.9; Ss, S,,,,, Sps in the specific order,
and their primed counterparts, with orthogonal decompositions of 'S
and S, 2.V.10, 2.V.I I
Combined application to nonprobabilistic potential theory and to martingale
theory: background, 3.1.1; band relations, 3.1.2; L° and D, 3.1.3; D
and the PWB method, 3.1.4, 3.1.5; characterizations of Sqb, Ss, Sms,
Spgb, Smgb, Sps, 3.1.6-3.1.10; band identification of the composition
of an h-superharmonic function with h-Brownian motion, 3.1.11,
3.1.13
Law of large numbers: for Brownian motion, 2.VI1.5
Lebesgue decomposition: A.IV.8
Lebesgue spine : I . V 111.15
Levy Brownian motion square increment theorem: 2.V111.10
Lifetime: of a Markov process, 2.V1.3, 2.VI.7; of conditional Brownian
motion, 2.X.1. 2.X.4, 2.X.9, 3.1.12
Liouville's theorem: 1.11.2, 1.11.13, 1.V.5
LM (see also GM)
Classical potential theory: definition and existence. 1.111.1, 1.111.2: for
subharmonic functions in various classes, 1.1X.2-I.IX.4
Parabolic potential theory: definition and existence, I.XVII.I
Martingale theory: definition and existence, 2.111.20, 2.IV.14; for sub-
martingales in various classes, 2.V.2-2.V.4
Logarithmic potential on l : defined, I .IV.I . Riesz type decomposition of a
superharmonic function, I.IV.9: domination principle. I.V.l0; infinity
set. I.V I I
Lower semicontinuous functions: smoothing of a function, A.VIII.I ; supre-
mum of a family of, A.VlIl.2; Choquet topological lemma, A.VI11.3
Lusin's measurability theorem: A.11.4

Markov process: defined, Markov property, initial distribution, transition


function, absolute probability function, 2.V1.1; choice of filtration.
2.VI.2: strong Markov property, 2.V1.3, 2.V1.9: stationary stochastic and
substochastic transitions. 2.V1.3. 2.V1.5, 2.VI.6; application of martingale
theory, 2.V1.4: lifetimes and trap points, 2.V1.3, 2.VI.7; right continuity
of filtrations, 2.VI.8. conditional-, 2.V1.13; tied down-. 2.V1.14:
killed . 2.V1.15: Brownian motion as a--, 2.VII.2
Markov time: see Optional time
Martin boundary
Classical context (aMD): Martin space D'M, 1.X11.1, 1.X11.3; Martin
function K, 1.X11.2; Martin representation, I.Xll.4, 1.X11.6, 1.X11.9:
minimal harmonic functions and the set arD of their poles, the minimal
boundary, 1.X11.5, 1.X11.7, I.XII.8; resolutivity, I.XI1.10; minimal
thinness and the minimal-fine topology, I.Xll.11, 1.X11.12, I.XIl.15
Index 839

1.XII.17; attainable-points, Brownian motion to and from-points,


3.111.1, 3.111.2
Parabolic context: Martin point set pairs, flat point set pairs, measure set
pairs, 1.XVIII.17, 1.XV111.18, I.XIX.2; exit and entrance boundaries,
I.XIX.1; Martin function k and admissible superparabolic functions,
I.XIX.2; Martin space and boundary, I.XIX.3; Martin representation,
1.XIX.4, I.XIX.7, 3.II1.6; minimal boundary, 1.XIX.5, I.XIX.6; slab,
I.XIX.8, I.XIX.9; right half-slab, I.XIX.10; resolutivity, 1.XIX.11;
minimal thinness and the minimal-fine topology, 1.XIX.12
Probabilistic context: attainable -points, Brownian motion to and from-
points, 3.111.1, 3.11I.2; zero-one law at a minimal-point, 3.111.3
Martingale (see also Submartingale and Supermartingale): definition,
examples, elementary properties, closability, 2.111.1-2.111.3; optional
sampling of a uniformly integrable-, 2.111.2(a), 2.IV.2; subsets of l8 as
general parameter sets, 2.111.4; convergence if L' bounded, 2.111.13;
convergence if uniformly integrable, 2.111.14; backward convergence,
2.111.16: Krickeberg decomposition if L° bounded, 2.V.4(c); 'Sm, S. and
their decompositions, 2.V.7, 2.V.11, 3.1.2, 3.1.4-3.1.6, 3.1.12, 3.1.13;
local-, 2.V.12, 2.VIII.3(e), 2.VII1.3(j), 2.VIII.6(e), 2.VIII.6(j); relative
to a Brownian motion filtration, 2.VIII.8; generated by composing
functions with Brownian or space-time Brownian motion, 2.IX.1-2.IX.3,
2.IX.7(c), 3.11.2(c)
Maximal inequalities: in martingale theory, 2.111.9-2.111.11
Maximum: -minimum theorem for harmonic functions, 1.1.4, for parabolic
functions, I.XV.5; theorem for analytic functions, I.V.7; principle of and
complete principle of, 1.V.10
Measurable family of functions: 2.1.2
Measurable space: A.IV.2
Meyer decomposition: of a supermartingale and of a submartingale, 2.IV.11,
2.IV.12
Minimal function
Classical context: harmonic, 1.I1.16, relative harmonic, 1.VI1I.1;
on D - Jr }, I.VIl.10; Martin boundary pole of, 1.X11.5
Parabolic context: parabolic, 1.XV.2; on the upper half-space of OB',
extension of, when defined on a cylinder, 1.XV.17; on a slab, 1.XVI.8;
e) on D - 1.XVII.8
Modification: see Standard modification

Nearly: progressively measurable, predictable, and so on, 2.1.8; Borel sets


for Brownian motion, 2.1
Nontangential: deleted neighborhood filter for a ball boundary point,
1.11.15: limits versus minimal-fine limits at ball and half-space boundary
points, I.XII.12, I.XII.20, I.XII.21
Normal: boundary limit points of a half-space subset versus minimal-fine
limit points of the set, I.XII.22; boundary limit function of a classical
840 Index

context potential, 1.X11.23; boundary limit function on a slab boundary


determines a bounded parabolic function, I.XVI.2
Nucleus: of a Suslin scheme, A.1.2

Optional sampling and stopping theorems: 2.111.6-2.I1I.8, 2.IV.3


Optional time: definition and properties, 2.11.1, 2.11.2; predictable, 2.11.7,
2.11.9; decomposition, totally inaccessible and accessible-'s, 2. 11. 12
Ordinate: of a point of U8^'. I.XV. I
Orthogonality: in a vector lattice, A.111.7

Parabolic function (see also Subparabolic function and Superparabolic func-


tion): definition. 1.XV.2; maximum-minimum theorem, I.XV.5; average
properties. I.XV.I0. Harnack convergence and inequality theorems,
LXV.I I. extension of, when defined on a cylinder, 1.XV.17, Poisson
integral and the L' and Dclasses. for functions on a slab, 1.XVI.1, 1.XVI.5,
I XVI.6: ratio boundary limit theorem on a slab, I.XVI.7, I.XVIII.IS:
removable singularity set, 1.XVII.8: relative, I.XVIII. I ; minimal,
I.XIX.5. Martin representation. I.XIX.7-I.XIX.10; vector lattice Sm
and its band decompositions. 3.1.1: first boundary value (Dirichlet)
problem. see PWB method; composition with space-time Brownian
motion, see Superparabolic function
Parabolic limit: at a slab boundary point, 1.XVI.7
Parabolic measure (p h)): for domains with smooth boundaries, I.XV.7; for
an interval, I.XV.9. I.XVIII.2; definition, 1.XVIII.2: relation with the
sweeping kernel, I.XVIII.8; relative to an irregular boundary point,
I.XVIf1.17: for the Martin space, I.XIX.I 1; as a conditional space-time
Brownian motion hitting distribution, 2.V11.12, 2.IX.13. 2.X.7, parabolic
version of 3.11.2
Paving: A.1.1
Poincare-Zaremba barrier: I .VII1.12, I .VIII.15
Poisson equation: classical context, 1.1.7; parabolic context, I.XVII.6
Poisson integral: classical context, for ball and half-space, 1.11.1, I.VIIl.9.
parabolic context, for a slab, I.XVI.1, I .XVI.5, I .XVI.6. 1.XVII.5
Polar set
Classical context (in R'): definition and properties, inner-, l.V. I -1.V.4,
I .XIV.2: removable singularity set of a lower bounded superharmonic
function, I.V.5: role in characterization of Greenian subsets of OZ,
1.V.6: infinities of a potential on a-, I.V.I l: exceptional set in the
Fundamental Convergence Theorem, I.V1.I: analytic inner-is-,
I.V1.2: h-harmonic measure on. 1.VIII.5. 1.X.8; vanishing of sweeping
kernel on. 1.X.6: characterized by absence of fine limit points, I.X1.6;
null for a measure of finite energy, I.X1l1.2; not hit by Brownian
motion, 2.1X.5, 2.X.2. but analytic non polar sets are surely hit if N = 2,
hit with probability > 0 if N > 2, 2.1X. 10
Parabolic context (in AN): definition, counterparts of classical context
Index 841

properties, 1.XVII.8; criteria for, parabolic- = coparabolic-,


1.XVII1.1 I ; not hit by space-time Brownian motion, 2.1X.5
Probability context (in R' x f2): definition, 2.11.10
Pole (see also Martin boundary): boundary pole of a minimal function,
L X.7, 1.X11.5, 1.XIX.5
Polish space: definition, A.1.10; measure on, A.IV.1I
Potential
Classical context: kernel G, 1.1.5, I.IV.I, I.IV.9; of a smooth density,
1.1.7; on a special open set, 1.IV.1; vanishing of GM(-), 1.IV.3,
I.IV.8; smoothing of, by averaging, I.IV.5; generating measure,
1.IV.6, 1.IV.7; condition that a superharmonic function be a potential,
1.IV.8, 1.VI11.1 I, 1.X11.17 example; infinities of, on a polar set, I .V.11;
boundary limit function, 1.V11.5, 1.X11.18, 1.XI1.23, 2.1X.7; lattice S,,,
1.IX.8; the case N = 1, I.XIV.7
Parabolic context: kernel GD, 1.XV.4, I.XVII.5; of a smooth density,
I.XVI1.6; boundary limit function, 1.XIX.14, 2.IX.7; lattice, I .XVIII.19
Martingale theory context: definition, 2.111.21; generation of, by an
increasing process, 2.IV.6-2.IV.11: associated measure, domination
principle, 2.IV.13; potential theory on R` x S2, 2.IV.15; energy,
2.IV.21
Precapacity: topological, and the generated Choquet capacity, A.II.8
Predictable:-a algebra, and--family of functions, 2.1.14; predictability of
a function family at an announced time, 2.11.3;-optional times and-
filtrations, 2.1I.7;-process defined by an accessible optional time, 2.11.7,
2.11.12; section theorem for a-set, 2.11.8; graph of a-optional time,
2.11.9;-increasing process = natural-process, 2.IV.7; continuity prop-
erties of--supermartingales and martingales, 2.IV.23
Progressively measurable: definition, right and left continuous processes
are-, 2.1.2; hitting of sets by-processes, 2.1.9, 2.11.4, condition for
existence of a-standard modification, 2.1.13;-process at an optional
time, 2.11.3; continuity properties of-processes, 2.11.5, 2.11.6
PWB method (see also Barrier, Regularity of a boundary point, Resolutivity)
Classical context (PWB''): upper and lower classes and solutions, I.VI11.2,
I.V111.6, 1.V111.7; solutions as reductions, 1.VI11.2, 1.VII1.10; solu-
tions are quasi bounded, I.IX.9
Parabolic context (PWB''): upper and lower classes and solutions, relation
with reductions, 1.XVI I1.1. classical context as a special case of parabol-
ic context, l.XV11I.4
Probabilistic solution: 2.1X.1, 2.1X.13, 2.X.7, 3.1I.1-3.11.4

Quasi-bounded
Classical context: the classes Sqb, Smgb, Sp'gb, 1.IX.9; significance for the
PWB method of quasi-boundedness, 1.IX.9; significance of quasi-
boundedness of a harmonic function in relation to the class D property
and the Riesz-Herglotz and Martin representations, 1.IX.12, 1.X11.9
842 Index

Parabolic context : I . X V I I I.19


Martingale theory context:-positive supermartingales and their genera-
tion by increasing processes, 2.IV.6-2.IV.I l ; classes 'Sqb, Sqb and their
subclasses 'Smgb, 'S pqb, S; qb, Spgb, 2.V.9
Combined context: see Lattice
Quasi-everywhere: classical context, inner-, IN. I, probabilistic context,
2.1.8
Quasi-Lindelof property: classical context, I.XI.11; parabolic context,
I.XVII.19
Quasi-martingale: definition, Rao representation as an element of S, 2.V.13

Reciprocity law: l .X111.2


Reduction (see also Sweeping)
Classical context: definition and properties, 1.1I1.4, 1.111.5, I .VI.2-l .VI.4;
GM(-), 1.111.6; relation to capacity, I.VI.5; generalized, 1.X1.20;
probabilistic evaluation, 2.IX.14, 2.X.7
Parabolic context: definition and properties, I.XVII.3, I.XVII.I1,
I.XVII.14, 1.XVII.16, I.XVII.17; probabilistic evaluation, 2.X.7
Martingale theory context: discrete parameter case, 2.111.22; application
to crossing inequalities, 2.111.23, 2.IV.18(o), 2.IV.19(o); continuous
parameter case, 2.I V.17-2.I V.20
Regularity of a boundary point (see also Barrier)
Classical context: definition, I.VIII.2; Euclidean boundary, I.VIII.4;
Poincare and Poincare-Zaremba criteria, Lebesgue spine, I.VI1l.12,
I.VIII.15; in terms of harmonic measure. I.VIlI.8; in terms of the fine
topology, 1.X1.12
Parabolic context: Euclidean boundary, I.XVlll.l, 1.XVIlI.6; for an
interval, 1.XVIJI.2; classical context regularity in terms of parabolic
context regularity. I.XVIII.4; for a parabolic ball, I.XVIII.6: iterated
logarithm criterion, I .XVIII.6; in terms of the fine topology, I .XVIII.7
Resolutivity
Classical context: definition, universal and internal, 1.VII1.2; Euclidean
boundary. I.VIII.4: condition for-of a boundary function, 1.V111.8,
I.V111.9, 3.11.2; condition for-of a boundary, I.VI1I.9, 3.11.3(e);
universal and universal internal- of a ball boundary. I .VIII.9. I .IX.12:
of a Martin boundary, I.XII.10
Parabolic context: adaptation of the classical context definitions and
results, I.XVIII.I - I.XVIII.3, 3.11.1; universal and universal internal-
of a Martin boundary, 1.XIX.I I
Reversal: of Brownian motion. 2.X.6
Riesz decomposition: of a superharmonic function, 1.1.8, 1.IV.8, 1.IV.9.
1.1X.1 I. I.XIV.9; of a superparabolic function, I.XV.7, I.XV1I.7: of a
supermartingale, 2.111.21
Riesz measure: associated with a superharmonic function. I.IV.7; associated
with a superparabolic function. I.XVII.7; associated with a supermar-
tingale. 2.1 V.13
Index 843

Riesz-Herglotz representation: of a harmonic function on a ball, 1.11. 14


Robin constant: I .XII1.18

Sample function: definition, integrability, 2.1.8; continuity properties, 2.11.5,


2.11.6; for properties of sample functions of specific process types, see
those types
Section theorem: 2.11.8
Semipolar set
Parabolic context, in A": definition, I.XVII.10; exceptional set of the
Fundamental Convergence Theorem, I.XVII.13; criteria for, parabolic
semipolar = coparabolic semipolar, 1.XVII.15, 1.XVIII.12; hitting of
by space time Brownian motion, 2.1X.15
Probabilistic context, in I x 0: definition, 2.11.10; role in supermar-
tingale smoothing, 2.IV.I; exceptional set of the Fundamental Conver-
gence Theorem, 2.IV.5
Sierpinski lemma: on Suslin schemes, A.II.2
Singular
Classical context: classes SS , SS, Sps of positive singular functions,
respectively superharmonic, harmonic, and potentials of measures,
1.IX.10; conditions that a function in S be singular, 3.1.7; evaluation
of the singular component of a function; in S+, 3.1.8; conditions that a
function in SD be singular, 3.1.10
Parabolic context: 1.XVIII.19 and the parabolic context part of Sections
3.1.7, 3.1.8, 3.1.10
Martingale theory context: classes 'Ss , S, of positive supermartingales
with their subclasses 'S, , 'S', Sms, S ,, 2.V.10; condition that a
process in SP be singular, 2.V.12; martingale theory part of Sections
3.1.7, 3.1.8, 3.1.10
Slab in ft': definition, I.XV.13; Green function and the Poisson integral,
1.XVI.1, 1.XVI.5, I.XV1.6, 1.XVII.4; generalized superparabolic average
inequality, 1.XVI.2; L` and D classes of parabolic functions, 1,XVI.6;
ratio parabolic function boundary limit theorem, 1.XVI.7, 1.XVIII.15,
I.XIX.15; minimal parabolic function, 1.XV1.8; Martin boundary,
I.XIX.8, I.XIX.9; coparabolic minimal thinness and minimal-fine limits
at a boundary point, l.XIX.12
Space-time Brownian motion (see also Brownian motion): definition, 2.VI1.2;
properties induced by Brownian motion properties, 2.VI1.4; transition
function for-in an open set, 2.VII.10; transition function for-in an
interval, parabolic measure as a-boundary hitting distribution, 2.VII.12,
2.VII.13; composition of a superparabolic function with-, excessive
functions and measures for-, 2.1X.2, 2.IX.3, 2.IX.6-2.IX.8, 2.IX.12,
2.IX.13, 2.IX.16, 2.IX.18; probability of hitting sets by-, 2.IX.4, 2.1X.5;
from a parabolic irregular boundary point (parabolic counterpart of)
2.1X.20; conditional, 2.X.12
Special open set: definition, I.IV.1; transition from, to Greenian sets, 1.VII.8
Specific order: definition, A.II1.4
844 Index

Standard modification: of a stochastic process, 2.1.8: measurable-, 2.1.13


State space: of a random variable, 2.1.3: of space-time Brownian motion,
2.V11.2
Stochastic interval: definition, 2.1I.1; fine open, closed, 2.IV.16
Stochastic process (see individual types under their own names): definition,
adaptedness, indistinguishability, almost sure properties, 2.1.8; canonical,
2.1.10: dependence of certain probabilities on the choice of probability
space, 2.1.11, 2.1.12, 2.11.4
Stolz domain: definition, 1.11.15
Stopping time: see Optional time
Strongly subadditive set function: definition, A.11.6, generation of a Choquet
capacity by. A.I1.7. A -+ R,,,. and A R",+11
I.VI.2(b), I.VI.3(j), I.VI.4(j);
classical capacity function. I.XIII.10; negative of the Robin con-
stant, I.XII1.18; 4- R", and A I.XVII.11(d), I.XVII.16(k),
l.XV11.17(k); A , and A - k , 2.1V.18(i), 2.IV.19(i)
R,_,_",

Subharmonic function (see also Harmonic function and Superharmonic func-


tion): definition. 1.11.4, 1.XIV.2: Jensen's inequality applied to, 1.11.9;
LMU operator on, I.IX.2-1.1X.4
Submartingale (see also Martingale and Supermartingale): definition,
2.111.1 ; Jensen's inequality applied to, 2.I11.3(c); is right closable if
uniformly integrable and is uniformly integrable if positive and right
closable, 2.111.3(e), 2.111.6, 2.V.3, 2.V.4; LM(--), 2.V.2
Subparabolic function (see also Parabolic function and Superparabolic func-
tion): definition, Jensen's inequality applied to, I.XV.12; maximum
theorem for a slab domain, I.XVI.3
Superharmonic function (see also Harmonic function and Subharmonic func-
tion): definition, 1.11.4, 1.XIV.2: minimum theorem, 1.11.5, I.V.7; charac-
terization in terms of harmonic functions. 1.11.7; differentiable, 1.11.8;
approximation of by differentiable ones, 1.11.8, I.IV.5, 1.IV.10; on an
annulus, 1.11.10; Riesz measure and decomposition. condition that a
positive- be a potential, 1.1.8, I.IV.7 1.IV.9. 1.XIV.9: relative-,
I.VIII.I: smoothness properties when N = I, 1.XIV.4; is a superparaholic
function of the space coordinate of I.XV.15; composition with
Brownian motion, 2 IX.3, 2.1X.6-2.1X.8, 2.IX.12: composition of ratio
of 's with conditional Brownian motion, 2.VI11.13. 2.X.1, 2.X.8, 3.1.11,
3.1.13, 3.11.2
Supermartingale (see also Martingale and Submartingale): definition,
closability. 2.111.1 2.111.3; first treatment of limits of monotone sequences
of. 2.111.5: at optional times. 2.111.6-2.111.8. 2.IV.2, 2.IV.3: subset of R
as general parameter set. 2.111.4: maximal inequalities, 2.111.9, 2.111.10;
downcrossing inequalities, 2.111.12, 2.111.23; convergence. 2.111.13-
2.111.17, 2.IV.3; potential. 2.111.21, generated by an increasing process,
2.IV.6 2.IV.11: sample function continuity properties, 2.IV.1, in par-
ticularas influenced by the reference filtration, 2.IV.23. 2.VII.4. 2.V111.8;
limit of an increasing sequence of- 's, 2.IV.4: Fundamental Convergence
Index 845

Theorem, 2.IV.5; measure associated with a-, domination principle,


2.IV.13; decomposition into jump and continuous components, 2.IV.22,
2.IV.23; lattices of-'s, 2.V.5-2.V.1I
Superparabolic function (see also Parabolic function and Subparabolic func-
tion): smooth context Riesz decomposition, 1.XV.7; definition, 1.XV.12;
minimum theorem, I.XV.13; approximation of, by differentiable ones,
I.XV.14; on a cylinder, 1.XV.15, I.XV.17; generalized-inequality on a
slab, 1.XVI.2; Riesz measure and decomposition, I.XVII.7; relative-,
1.XVIII. I ; admissible-, I.XIX.2; composition with space-time Brownian
motion, 2.IX.3, 2.IX.6-2.IX.8, 2.[X.12, 2.IX.13
Suslin scheme: A.I.2
Sweeping of a measure
Classical context: definition, the sweeping kernel, 1.X.1, 1.X.4-1.X.6;
relation of the sweeping kernel to harmonic measure, 1.X.2; sweep-
ing symmetry, 1.X.3; support of a swept measure, I.XI.14, 1.X1.18,
2.IX. 15
Parabolic context: definition, the sweeping kernel and parabolic measure,
dualization of sweeping symmetry, I.XVIII.8; support of a swept mea-
sure, 1. X V I I I.13, 2.I X.15

Tail a algebra: of a conditional Brownian motion, 2.X.11, 3.11.4


T operator
Classical context (rhq): definition, 1.11.1, 1.XIV.3; application to super-
harmonic functions, 1.11.6; generalized, application to GM"" operator,
l.V1II.11; variant.of, I.VIII.21
Parabolic context (ihj): definition, I.XV.9; application to superparabolic
functions, 1.XV.14, I.XVIII.2
Martingale theory context (rST): definition, 2.111. 18, 2.IV.14
Thinness (see also Fine topology)
Classical context: definition, 1.XI.1; criterion, 1.XI.2, 1.XI.3; at oo,
1.X[.5; minimal, 1.XII.11; a ray from a half-space boundary point is
not minimal thin at the point, I.XII.12; Wiener criterion, I.XIII.17;
probabilistic criterion, 2.IX.15, 3.111.3
Parabolic context: definition, criterion, 1.XVII.5 Example (c), 1.XVII.9,
I.XVII.12; minimal thinness at Martin boundary points of a slab and
of the fourth quadrant of 12, I.XIX.12; probabilistic criterion, 2.IX.15
Probability theory context: at an optional time, 2.IV.16
Transition function: definition and universally measurable extension, A.VI.2,
A.VI.3
Trap: definition, A.VI.1, 2.VI.3, 2.VI.7
Tulcea construction: of a canonical Markov process, 2.VI.3

Uniform integrability: definition, test function, implications, A.V; of the


class of conditional expectations of a random variable, 2.1.4(i); of a
process, 2.1.8; of a right closed positive submartingale, and the right
846 Index

closability of a uniformly integrable submartingale, 2.111.3(e), notation


UI, quasi-bounded martingales are the uniformly integrable ones,
2.V.9: example of a uniformly integrable but not class D supermartingale,
2.IX.6
Universal measurability: definition, A.11.9; of hitting probability as a func-
tion of a Markov process initial point, 2.V1.6
Upcrossing inequalities: see Downcrossing inequalities

Version: of an equivalence class of random variables, 2.1.3

Wiener measure: defined. 2.V11.3


Wiener thinness criterion: I.Xl1l.l7

Zaremba barrier: see Poincare-Zaremba barrier


Zero-one law: for the classical and parabolic sweeping kernels, I.X.6(c),
I.XVIII.IO; at the initial point of a Markov process, in particular for
Brownian motion, 2.VI.8, 2.VI1.6, 2.X.2(d); at a minimal Martin boun-
dary point, 3.111.3
Zero set of a positive superparabolic function : l. X V I 11.14, l. X VI 11.17
CLASSICS IN MATHEMATICS

Springer-Verlag began publishing books in higher


mathematics in 1920, when the series Grundlehren
der mathematischen Wissenschaften, initially
conceived as a series of advanced textbooks, was
founded by Richard Courant. A few years later,
a new series Ergebnisse der Mathematik and ihrer
Grenzgebiete, survey reports of recent mathema-
tical research, was added.
Of over 400 books published in these series, many
have become recognized classics and remain
standard references for their subject. Springer is
reissuing a selected few of these highly successful
books in a new, inexpensive softcover edition, to
make them easily accessible to younger generations
of students and researchers.

Classical Potential Theory


and Its Probabilistic Counterpart
From the reviews:
"This huge book written in several years by one of the few mathe-
maticians able to do it, appears as a precise and impressive study
of this bothsided question that replaces, in a coherent way, with-
out being encyclopaedic, a large library of books and papers
scattered without a uniform language. Instead of summarizing
the author gives his own way of exposition with original comple-
ments. This requires no preliminary knowledge....The purpose
which the author explains in his introduction, i.e. a deep prob-
abilistic interpretation of potential theory and a link between
two great theories, appears fulfilled in a masterly manner".
M. Brelot in Metrika, 1986

"In the early 1920's, Norbert Wiener wrote significant papers


on the Dirichlet problem and on Brownian motion. Since then
there has been enormous activity in potential theory and
stochastic processes, in which both subjects have reached a high
degree of polish and their close relation has been discovered.
Here is a monumental work by Doob, one of the masters, in
which Part i develops the potential theory associated with
Laplace's equation and the heat equation, and Part 2 develops
those parts (martingales and Brownian motion) of stochastic
process theory which are closely related to Part C.
G. E. H. Reuter in Short Book Reviews, 1985

ISBN 3-540-41206-9

91111 ISSN 1431-0821


9 41206
littp://www.spritiger.de

You might also like