Mass transportation problems [V.2] 9780387983523, 038798352X

This is the first comprehensive account of the theory of mass transportation problems and its applications. In volume I,

292 86 2MB

English Pages 450 Year 1998

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Mass transportation problems [V.2]
 9780387983523, 038798352X

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

Mass Transportation Problems, Volume II: Applications

Svetlozar T. Rachev Ludger Rüschendorf

Springer

To my wife Zoja and To my parents Nadezda and Todor Rachevi.

To my wife Gabi.

Svetlozar (Zari) Rachev

Preface to Volume II

The second volume of the Mass Transportation Problems is devoted to applications in a variety of fields of applied probability, queueing theory, mathematical economics, risk theory, tomography, and others. In Volume I we encompassed the general mathematical theory of mass transportation, concentrating our attention on: • the general duality theory of the transportation and transshipment problem; • explicit optimality results; • applications to minimal probability metrics, stochastic ordering, approximation and extension problems; • applications to functional analysis and mathematical economics (the Debreu theorem, utility theory, dynamical systems, choice theory, and convex and nonconvex analysis were dicsussed in this context). In Volume II we expand the scope of applications of mass transportation problems. Some of them arise from modifications of the admissible transportation plans. In fact, for applications to mathematical economics it is of interest to consider relaxations of the marginal constraints, such as upper or lower bounds on the supply and demand distributions, or additional constraints like capacity bounds for the transportation plans. In mathematical tomography the basic problem is to reconstruct the multivariate

viii

Preface to Volume II

probability distribution based on some information about the marginal distributions in a certain finite number of directions. This information may be represented by additional constraints on the support functions or distributional moments, or it may be contained in only partial information on the marginals. Thus there is a close relationship between a class of problems in mathematical tomography and the classical theory on moment problems, which again can be viewed as a relaxation on the set of constraints in mass transportation problems. We discuss in detail applications to approximation problems for stochastic processes and to rounding problems based on moment-type characteristics. A particular example will be the approximation of queueing models. The minimal metrics allow us to compare various rounding rules and to determine optimal ones from an asymptotic point of view. An important field of applications of mass transportation problems we shall consider in this second volume is to probabilistic limit theorems. This approach was introduced in the seventies by the Russian school of probability theory, headed by V.M. Zolotarev. By inherent regularity properties of probability metrics defined via certain mass transportation problems, there are streamlined proofs for central limit theorems on Banach spaces yielding sharp quantitative estimates of Berry–Esseen type for the convergence rate. The probability metric approach will be applied to general stable and operator stable limits theorems, martingale-type limit theorems, limit behavior of summability methods, and compound Poisson approximation. A particular application is to the classical problem in mathematical risk theory dealing with sharp approximation of the individual risk model by the collective risk model. The probability metric approach will also be applied to the quantitative asymptotics in rounding problems. A new field of application of probability metrics arising as solutions of mass transportation problems is the analysis of deterministic and stochastic algorithms. This research area is of increasing importance in computer science and various fields of stochastic modeling. Based on regularity properties of probability metrics, a general “contraction” method for the asymptotic analysis of algorithms has been developed. The contraction method has been applied successfully to a variety of search, sorting, and other tree algorithms. Furthermore, the recursive structure in iterated functions systems (image encoding), fractal measures, bootstrap statistics, and time series (ARCH) models has been analyzed by this method. It becomes clear that there are many interesting probabilistic applications of this method to be rigorously developed in the future. In the final chapter we consider applications to stochastic differential equations (SDEs) and to convergence of empirical measures. SDEs will be interpreted as continuous recursive structures. From this point of view we provide a detailed discussion on the approximative solution of nonlinear stochastic differential equations of McKean–Vlasov type by interactive par-

Preface to Volume II

ix

ticle systems with application to the Kac theory of chaos propagation. The probability metrics approach allows us to establish approximation results for various modifications of the diffusion system, some of them of “nontraditional” type. In a general context we establish approximation results for empirical measures and give applications to the approximation of stochastic processes. As final applications we discuss a weak approximation of SDEs of Itˆo type by a combination of the time discretization methods of Euler and Milshtein with a chance discretization based on the strong invariance (embedding) principle. This approximation is given in terms of minimal Lp -metrics and thereby based on regularity properties of the solutions of the corresponding mass transportation problem.

Preface to Volume I

The subject of this book, mass transportation problems (MTPs), concerns the optimal transfer of masses from one location to another, where the optimality depends upon the context of the problem. Mass transportation problems appear in various forms and in various areas of mathematics and have been formulated at different levels of generality. Whereas the continuous case of the transportation problem may be cast in measure-theoretic terms, the discrete case deals with optimization over generalized transportation polyhedra. Accordingly, work on these problems has developed in several separate and independent directions. The aim of this monograph is to investigate and to develop, in a systematic fashion, the Monge–Kantorovich mass transportation problem (MKP) and the Kantorovich–Rubinstein transshipment problem (KRP). We consider several modifications of these problems known as the MTP with partial knowledge of the marginals and the MTP with additional constraints (MTPA). We also discuss extensively a variety of stochastic applications. In the first volume of Mass Transportation Problems we concentrate on the general mathematical theory of mass transportation. In Volume II we expand the scope of applications of mass transportation problems. In 1781 Gaspard Monge proposed in simple prose a seemingly straightforward problem of optimization. It was destined to have wide ramifications. He began his paper on the theory of “clearings and fillings” as follows: When one must transport soil from one location to another, the custom is to give the name clearing to the volume of the soil that one

xii

Preface to Volume 1 must transport and the name filling (“remblai”) to the space that it must occupy after transfer. Since the cost of transportation of one molecule is, all other things being equal, proportional to its weight and the interval that it must travel, and consequently the total cost of transportation being proportional to the sum of the products of the molecules each multiplied by the interval traversed; given the shape and position, the clearing and the filling, it is not the same for one molecule of the clearing to be moved to one or another spot of the filling. Rather, there is a certain distribution to be made of the molecules from the clearing to the filling, by which the sum of the products of molecules by intervals travelled will be the least possible, and the cost of the total transportation will be a minimum. (Monge, (1781, p. 666)).

In mathematical language Monge proposed the following nonlinear varational problem. Given two sets A, B of equal volume, find an optimal volume-preserving map between them; the optimality is evaluated by a cost function c(x, y) representing the cost per unit mass for transporting material from x ∈ A to y ∈ B. The optimal map is the one that minimizes the total cost of transferring the mass from A to B. Monge considered this problem with cost function equal to the Euclidean distance in IRd : c(x, y) = |x − y|. Monge’s problem turned out to be the prototype for a class of problems arising in various fields such as mathematical economics, functional analysis, probability and statistics, linear and stochastic programming, differential geometry, information theory, cybernetics, and ma trix theory. The optimization function A c(x, t(x)) dx is nonlinear in the transportation function t, and moreover, the set of admissible transportations is a nonconvex set. This explains why it took a long time until even existence results for optimal solutions could be established. The first general existence result was given in 1979 by Sudakov. On the second page of his paper Monge himself had remarked that to obtain a minimum, the intervals traversed by two different molecules should not intersect. This simple observation applied to the discrete case—where there are only a finite number of molecules—leads to a “greedy” algorithm, the so-called northwest corner rule. The totality of mass transferences plans in the discrete case is a polytope that arises in the transportation problem of mathematical programming, where it is treated in specialized form as an assignment problem and in generalized form as a network-flow problem. The northwest corner rule solves transportation problems having a particular structure on the costs and is, moreover, at the heart of many seemingly different problems having an “easy” solution (cf. Hoffman (1961), Barnes and Hoffman (1985), Derigs, Goecke, and Schrader (1986), Hoffman and Veinott (1990), Olkin and Rachev (1991), and Rachev and R¨ uschendorf (1994); see also Burkard, Klinz, and Rudolf (1994) and the references therein). The Academy of Paris offered a prize for the solution of Monge’s problem, which was claimed by the differential geometer P. Appell (1884–1928), who

Preface to Volume I

xiii

established some geometric properties of optimal maps in the plane and in IR3 . But it took a long time until a real breakthrough in the transportation problem came, originating in the seminal 1942 paper of L.V. Kantorovich entitled “On the transfer of masses.” Kantorovich stated the problem in a new, abstract, and in more easily accessible setting and without knowledge of Monge’s work. Kantorovich learned of Monge’s work only later (cf. his 1948 paper). In the Kantorovich formulation of the mass transportation problem (the so-called “continuous” MTP), the initial mass (the clearing) and the final mass (the filling) can be considered as probability measures on a metric space. The essential step in this formulation is the replacement of the class of transportation map by the wider class of generalized transportation plans, that are identifiable with the convex set of all probability measures on the product space with fixed marginals. The difficult nonlinear Monge problem was thereby replaced by a linear optimization problem over an abstract convex set. This made it possible to put this problem in the framework of linear optimization theory and encouraged the development of general duality theory for the solution of the Kantorovich formulation of the transportation problem as the basic tool. Accordingly, these problems and their generalizations will be referred to as Monge–Kantorovich Mass Transportation Problems (MKPs). Kantorovich’s measure theoretic formulation made the problem accessible to various areas of the mathematical sciences and other scientific fields. Kantorovich himself received a Nobel Prize in Economics for related work in mathematical economics.(1) Here is a list of some references in the mathematical sciences: • Functional analysis: Kantorovich and Akilov (1984) • Probability theory: Fr´echet (1951), Cambanis et al. (1976), Dudley (1976, 1989), Kellerer (1984), Rachev (1991c), R¨ uschendorf (1991) • Statistics: Gini (1914, 1965), Hoeffding (1940, 1955), Kemperman (1987), Huber (1981), Bickel and Freedman (1981), R¨ uschendorf (1991) • Linear and stochastic programming: Hoffman (1961), Barnes and Hoffman (1985), Anderson and Nash (1987), Burkard, Klinz and Rudolf (1994) • Information theory and cybernetics: Wasserstein (1969), Gray et al. (1975), Gray and Ornstein (1979), Gray et al. (1980) • Matrix theory: Lorentz (1953), Marcus (1960), Olkin and Pukelsheim (1982), Givens and Shortt (1984) (1) L.V.

Kantorovich together with T.C. Koopmans received the Nobel Memorial Prize in Economic Science in 1975 for “contributions to the theory of optimum allocation of resources”; see Dudley (1989, p. 342).

xiv

Preface to Volume 1

Many practical problems arising in various scientific fields have led mathematicians to solve MKPs: e.g., in • Statistical physics: Tanaka (1978), Dobrushin (1979) • Reliability theory: Barlow and Proschan (1975), Kalashnikov and Rachev (1990), Bene˘s (1985) • Quality control: Jirina and Nedoma (1957) • Transportation: Dantzig and Ferguson (1956) • Econometrics: Shapley and Shubik (1972), Pyatt and Round (1985), Gretsky, Ostroy, and Zame (1992) • Expert systems: Perez and Jirousek (1985) • Project planning: Haneveld (1985) • Optimal models for facility location: Ermoljev, Gaivoronski, and Nedeva (1983) • Allocation policy: Rachev and Taksar (1992) • Quality usage: Rachev, Dimitrov and Khalil (1992) • Queueing theory: Rachev (1989), Anastassiou and Rachev (1992a, 1992b) There are several surveys in the vast literature about MKP, among them Rachev (1984b), Rachev and R¨ uschendorf (1990), Burkard, Klinz, and Rudolf (1994), Cuesta-Albertos, Matr´ an, Rachev, and R¨ uschendorf (1996), and Gangbo and McCann (1996) related to dual solutions and applications of MKP; Shorack and Wellner (1985, Sect. 3.6) on optimal processes; Benes and Stepan (1987, 1991) on extremal mass transportation plans; R¨ uschendorf (1981, 1991, 1991a), Kellerer (1984), Rachev (1991c) on multivariate transportation problems; Dudley (1989) on distances in the space of measures; Talagrand (1992) and Yukich (1991) on matching problems. In recent years, characterizations of the solutions of the Monge–Kantorovich problem have been given in terms of c-subgradients of generalized convex functions defined in terms of the cost functions c(x, y) (cf. Knott and Smith (1984, 1992), Brenier (1987), R¨ uschendorf and Rachev (1990), R¨ uschendorf (1991, 1991a, 1995), Cuesta-Albertos, Matr´ an, Rachev, and R¨ uschendorf (1996), and Gangbo and McCann (1996)). For the case of squared Euclidean costs c(x, y) = |x − y|2 , the generalized convexity property is equivalent to convexity, and c-subgradients are identical to the usual subgradients of convex analysis. From this characterization

Preface to Volume I

xv

a series of explicit solutions of the transportation problem could be established. It also implies that the solutions of the MKP are under continuity assumptions given by mappings. Therefore, the solutions of the “easier” MKP imply as well the existence and characterizations of solutions of the original Monge problem, and so the MKP turns out to be the fundamental formulation of the transportation problem. For this reason, we concentrate in this book on the Kantorovich-type mass tranportation problems. For a discussion of interesting analytic aspects of the Monge problem, we refer to Gangbo and McCann (1996). Another type of MTP appears in probability theory, even if it leaves the framework of probability measures as transportation plans. Its solutions are bounded measures on a product of two spaces with the difference of marginals equal to the difference of two given probability measures. It will be called the Kantorovich–Rubinstein Problem (KRP), since the first results were obtained by Kantorovich and Rubinstein (1958). In its relation to the practical task of mass transportation it is sometimes referred to as the transshipment problem; see Kemperman (1983), and Rachev and Shortt (1990). The KRP has been developed to a great extent in the Russian school of probabilists and functional analysts, in particular by V.L. Levin, A.A. Milyutin, and A.M. Vershik and their students. For metric cost functions the KRP coincides with the corresponding MKP; for general cost functions it can be reduced to the MKP for a corresponding reduced cost function. For the duality theory of the KRP a specific detailed theory with many results that are of value in themselves has been developed with wide-ranging applications to mathematical economics. For a different approach to the KRP as introduced in Dudley (1976) and as further extended in Rachev and Shortt (1990) we refer to the book of Rachev (1991c). A problem related to both MKP and KRP is the Mass Transportation Problem with Partial Knowledge of the Marginals (MTPP), which is expressed by stating finitely many moment conditions. Problems of this type were formulated and extensively studied by Rogosinski (1958), Kemperman (1983), and Kuznezova-Sholpo and Rachev (1989). Barnes and Hoffman (1985) considered mass tranportaion problems with capacity constraints on the admissible transportation plans as an example of Mass Transportation Problems with Additional Constraints (MTPA) (see Rachev (1991b) and Rachev and R¨ uschendorf (1994)). In this book we give an extensive account of the duality theory of the MKP and the KRP, including the known results on explicit constructions and characterizations of optimal solutions. In Chapters 2 and 3 we present important duality theorems for the Monge–Kantorovich problem based on work of H. Kellerer, L. R¨ uschendorf, S.T. Rachev, and D. Ramachandran.

xvi

Preface to Volume 1

In Chapters 4 and 5 we present basically work of V.L. Levin; we analyze measure-theoretic methods for infinite-dimensional linear programs developed in context with the KRP as well as applications to general utility theorems (the Debreu theorem), extension theorems, choice theory, and set-valued dynamical systems.(2) In Chapters 6 and 8 we discuss new material on applications of the MKP and the KRP to the representation of ideal metrics and on various probabilistic approximation and limit theorems. This supplements the earlier results in this direction as described in the book of Rachev (1991) on probability metrics and stochastic models. In particular, we show that probability metrics allow us to find unified proofs for central limit theorems for martingales, (operator) stable limit theorems, and to more specific problems like compound Poisson approximation or rounding problems. Chapter 7, the first chapter in the second volume, is concerned with modifications of the MKP by additional or relaxed constraints. We discuss various types of moment problems and applications to the tomography paradoxon and to the approximation of queueing systems. A wide range of applications of metrics based on the transportation problem has been established in recent years in connection with recursive stochastic equations. We discuss algorithms of informatics (sorting, searching, branching, search trees) as well as applications to the approximation of stochastic differential equations, to the propagation of the chaos property of particle systems with applications to the approximation of nonlinear PDEs, as well as to the rate of convergence of empirical measures, which is of interest for matching problems in Chapters 9 and 10. From the technical point of view, MKPs can be subdivided into the discrete and continuous cases, according to the nature of their basic spaces and to the supports of the initial and the final masses. In the discrete case, the totality of the mass transference plans is the polytope that arises in the transportation problem of mathematical programming. There is, of course, a vast literature on the transportation problem, its specialization to the assignment problem, and its generalization to network flow problems. It turns out, as will be elaborated further in the book, that the northwest corner rule in the discrete case corresponds to a closed form for the solution in the continuous case. Indeed, the discrete analogue of a result known in the continuous case provides a new result in the discrete case; and its simple proof in the discrete case provides a new proof for the continuous case, see Rachev and R¨ uschendorf (1994c) and the references therein. Another approach in the discrete linear case prefers to exploit the special structure of supplies and demands (or clearings and fillings) and permits a particularly simple combinatorial algorithm for finding an optimal solution as developed (2) These

two chapters were written following closely the notes kindly provided to us by V.L. Levin.

Preface to Volume I

xvii

by Balinski (1983), Balinski and Russakoff (1984), Balinski (1985, 1986), Goldfarb (1985), Kleinschmidt, Lee, and Schannath (1987), and Burkard, Klinz, and Rudolf (1994). MTPs may be viewed as an analogue and a unifying framework of a problem considered by probabilists at the beginning of the twentieth century: How does one measure the difference between two random quantities? Many specific contributions to the analysis of this problem have been made, including Gini’s (1914) notion of concordance, Kendall’s τ , Spearman’s ̺, the analysis of greatest possible differences by Hoeffding (1940) and others, by Fr´echet (1951, 1957), Robbins (1975), and Lai and Robbins (1976), and the generalizations of these results by Cambanis, Simons, and Stout (1976), R¨ uschendorf (1980), Tchen (1980), and Cambanis and Simons (1982). These (and others) offer piecemeal answers to basic questions that arise from different stochastic models; they give no guidance as to the question of what concept should be used where: There is no general theory underlying the diverse approaches. We refer to Kruskal (1958), Gini (1965), and Rachev (1984b, 1991c). In this book we investigate, develop, and exploit the connections between the discrete and continuous versions of the mass transportation problems (MTP) as well as study systematically the relationships between the methods and results from different versions of the MTP. The MTPs are the basis of many problems related to the question of stability of stochastic models, to the question of whether a proposed model yields a satisfactory approximation to the phenomenon under consideration, and to the problem of approximation of stochastic and deterministic algorithms. It is our belief that MTPs hold great promise in stochastic analysis as well as in mathematical analysis. The MTP is full of connections with geometry, (partial) differential equations, (generalized) convex analysis, moment problems, infinite-dimensional linear programming, measurable choice theory, and extension problems, and it has many open problems. It has a great potential for a series of applications in several scientific fields. This book grew out of joint work and lectures delivered by the authors at the Steklov Mathematical Institute, Universit¨at M¨ unster, Universit¨at Freiburg, the Ecole Polytechnique, SUNY at Stony Brook, and the University of California, Santa Barbara, over many years. Many colleagues provided helpful suggestions after reading parts of the manuscript. All chapters were rewritten several times, and preliminary versions were circulated among friends, who eliminated many inaccuracies and obscurities. We would like to thank H.G. Kellerer, V.L. Levin, M. Balinski, D. Ramachandran, G.A. Anastassiou, M. Maejima, M. Cramer, I. Olkin, M. Gelbrich, W. R¨ omisch, V. Bene˘s, L. Uckelmann, and many other friends and colleagues who encouraged us to complete the work. We are indebted to Mrs. M. Hattenbach and Ms. A. Blessing for their superb typing; the appearance of this monograph owes much to them. We are grateful to the publisher

xviii

Preface to Volume 1

and especially to J. Kimmel for support and patience. We are particularly thankful to J. Gani for his invaluable suggestions concerning improvements of this work, his help with the organization of the material, and his encouragement to continue the project. Finally, we thank the Alexander von Humboldt Foundation for its generous financial support of S.T. Rachev in 1995 and 1996, which made this joint work possible. (3)

(3) The

work of S.T. Rachev was also partially supported by NSF Grants. The joint work of the authors was supported by NATO-Grant CRG900798.

Contents to Volume II

Preface to Volume II

vii

Preface to Volume I

xi

7 Relaxed or Additional Constraints 7.1 Mass Transportation Problem with Relaxed Marginal Constraints . . . . . . . . . . 7.2 Fixed Sum of the Marginals . . . . . . . . . . . . . . 7.3 Mass Transportation Problems with Capacity Constraints . . . . . . . . . . . . . . . 7.4 Local Bounds for the Transportation Plans . . . . . 7.5 Closeness of Measure on a Finite Number of Directions . . . . . . . . . . . 7.6 Moment Problems of Stochastic Processes and Rounding Problems . . . . . . . . . . . . . . . . 7.6.1 Moment Problems and Kantorovich Radius . . . . . 7.6.2 Moment Problems Related to Rounding Proportions 7.6.3 Closeness of Random Processes with Fixed Moment Characteristics . . . . . . . . . . 7.6.4 Approximation of Queueing Systems with Prescribed Moments . . . . . . . . . . . . . . . 7.6.5 Rounding Random Numbers with Fixed Moments . .

1 . . . . . .

2 10

. . . . . .

17 36

. . .

42

. . . . . . . . .

52 54 57

. . .

62

. . . . . .

71 80

xx

Contents to Volume II

8 Probabilistic-Type Limit Theorems 8.1

85

Rate of Convergence in the CLT with Respect to Kantorovich Metric . . . . . . . . . . . .

85

8.2

Application to Stable Limit Theorems . . . . . . . . . . . 102

8.3

Summability Methods, Compound Poisson Approximation 126

8.4

Operator-Stable Limit Theorems . . . . . . . . . . . . . . 131

8.5

Proofs of the Rate of Convergence Results . . . . . . . . . 153

8.6

Ideal Metrics in the Problem of Rounding . . . . . . . . . 178

9 Mass Transportation Problems and Recursive Stochastic Equations

191

9.1

Recursive Algorithms and Contraction of Transformations . . . . . . . . . . . . . . . . . . . . . . 191

9.2

Convergence of Recursive Algorithms . . . . . . . . . . . . 204

9.2.1 Learning Algorithm . . . . . . . . . . . . . . . . . . . . . . 204 9.2.2 Branching-Type Recursion . . . . . . . . . . . . . . . . . . 206 9.2.3 Limiting Distribution of the Collision Resolution Interval . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220 9.2.4 Quicksort . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 9.2.5 Limiting Behavior of Random Maxima . . . . . . . . . . . 231 9.2.6 Random Recursion Arising in Probabilistic Modeling: Limit Laws . . . . . . . . . . . . . . . . . . . . . . . . . . 236 9.2.7 Random Recursion Arising in Probabilistic Modeling: Rate of Convergence . . . . . . . . . . . . . . . . . . . . . 248 9.3

Extensions of the Contraction Method . . . . . . . . . . . 254

9.3.1 The Number of Inversions of a Random Permutation . . . 254 9.3.2 The Number of Records . . . . . . . . . . . . . . . . . . . 257 9.3.3 Unsuccessful Searching in Binary Search Trees

. . . . . . 260

9.3.4 Successful Searching in Binary Search Trees . . . . . . . . 263 9.3.5 A Random Search Algorithm . . . . . . . . . . . . . . . . 269 9.3.6 Bucket Algorithm

. . . . . . . . . . . . . . . . . . . . . . 272

10 Stochastic Differential Equations and Empirical Measures 10.1

277

Propagation of Chaos and Contraction of Stochastic Mappings . . . . . . . . . . . . . . . . . . . . 277

10.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 277

Contents to Volume II

xxi

10.1.2 Equations with p-Norm Interacting Drifts . . . . . . . . . 279 10.1.3 A Random Number of Particles

. . . . . . . . . . . . . . 290

10.1.4 pth Mean Interactions in Time: A Non-Markovian Case . 293 10.1.5 Minimal Mean Interactions in Time

. . . . . . . . . . . . 307

10.1.6 Interactions with a Normalized Variation of the Neighbors: Relaxed Lipschitz Conditions . . . . . . . . . . . . . . . . 308 10.2

Rates of Convergence in the Kantorovich Metric . . . . . . . . . . . . . . . . . . 322

10.3

Stochastic Differential Equations . . . . . . . . . . . . . . 332

References

351

Abbreviations

395

Symbols

397

Index

409

Contents to Volume I

Preface to Volume I

vii

Preface to Volume II

xv

1 Introduction

1

1.1

Mass Transportation Problems in Probability Theory . . .

1

1.2

Specially Structured Transportation Problems . . . . . . .

21

1.3

Two Examples of the Interplay Between Continuous and Discrete MTPs . . . . . . . . . . . . . . . . . . . . . .

23

Stochastic Applications . . . . . . . . . . . . . . . . . . . .

27

1.4

2 The Monge–Kantorovich Problem 2.1

57

The Multivariate Monge–Kantorovich Problem: An Introduction . . . . . . . . . . . . . . . . . . . . . . . .

58

2.2

Primal and Dual Monge–Kantorovich Functionals . . . . .

64

2.3

Duality Theorems in a Topological Setting . . . . . . . . .

76

2.4

General Duality Theorem . . . . . . . . . . . . . . . . . .

82

2.5

Duality Theorems with Metric Cost Functions . . . . . . .

86

2.6

Dual Representation for Lp -Minimal Metrics . . . . . . . .

96

xxiv

Contents to Volume I

3 Explicit Results for the Monge–Kantorovich Problem

107

3.1

The One-Dimensional Case . . . . . . . . . . . . . . . . . 107

3.2

The Convex Case . . . . . . . . . . . . . . . . . . . . . . . 112

3.3

The General Case . . . . . . . . . . . . . . . . . . . . . . . 123

3.4

An Extension of the Kantorovich L2 -Minimal Problem . . 132

3.5

Maximum Probability of Sets, Maximum of Sums, and Stochastic Order . . . . . . . . . . . . . . . . . . . . . 144

3.6

Hoeffding–Fr´echet Bounds . . . . . . . . . . . . . . . . . . 151

3.7

Bounds for the Total Transportation Cost . . . . . . . . . 158

4 Duality Theory for Mass Transfer Problems

161

4.1

Duality in the Compact Case . . . . . . . . . . . . . . . . 161

4.2

Cost Functions with Triangle Inequality . . . . . . . . . . 172

4.3

Reduction Theorems . . . . . . . . . . . . . . . . . . . . . 190

4.4

Proofs of the Main Duality Theorems and a Discussion . . 207

4.5

Duality Theorems for Noncompact Spaces . . . . . . . . . 219

4.6

Infinite Linear Programs . . . . . . . . . . . . . . . . . . . 241

4.6.1 Duality Theory for an Abstract Scheme of Infinite-Dimensional Linear Programs and Its Application to the Mass Transfer Problem

. . . . 241

4.6.2 Duality Theorems for the Mass Transfer Problem with Given Marginals . . . . . . . . . . . . . . . . . . . . . 245 4.6.3 Duality Theorem for a Marginal Problem with Additional Constraints of Moment-Type . . . . . . . 251 4.6.4 Duality theorem for a Further Extremal Marginal Problem . . . . . . . . . . . . . . . . . . . . . . . 258 4.6.5 Duality Theorem for a Nontopological Version of the Mass Transfer Problem . . . . . . . . . . . . . . . . 265 5 Applications of the Duality Theory

275

5.1

Mass Transfer Problem with a Smooth Cost Function—Explicit Solution . . . . . . . . . . . . . . 275

5.2

Extension and Approximate Extension Theorems . . . . . 290

5.2.1 The Simplest Extension Theorem  the Case X = E(S) and X1 = E(S1 ) . . . . . . . . . . . 290

5.2.2 Approximate Extension Theorems . . . . . . . . . . . . . 292 5.2.3 Extension Theorems . . . . . . . . . . . . . . . . . . . . . 295

Contents to Volume I

xxv

5.2.4 A continuous selection theorem . . . . . . . . . . . . . . . 302 5.3

Approximation Theorems . . . . . . . . . . . . . . . . . . 306

5.4

An Application of the Duality Theory to the Strassen Theorem . . . . . . . . . . . . . . . . . . . 319

5.5

Closed Preorders and Continuous Utility Functions . . . . 322

5.5.1 Statement of the Problem and the Idea of the Duality Approach . . . . . . . . . . . . . . . . . . . 322 5.5.2 Functionally Closed Preorders . . . . . . . . . . . . . . . . 324 5.5.3 Two Generalizations of the Debreu Theorem . . . . . . . . 329 5.5.4 The Case of a Locally Compact Space . . . . . . . . . . . 335 5.5.5 Varying preorders and a universal utility theorem

. . . . 337

5.5.6 Functionally Closed Preorders and Strong Stochastic Dominance . . . . . . . . . . . . . . 341 5.6

Further Applications to Utility Theory . . . . . . . . . . . 344

5.6.1 Preferences That Admit Lipschitz or Continuous Utility Functions . . . . . . . . . . . . . . . 344 5.6.2 Application to Choice Theory in Mathematical Economics . . . . . . . . . . . . . . . . . 352 5.7

Applications to Set-Valued Dynamical Systems . . . . . . 354

5.7.1 Compact-Valued Dynamical Systems: Quasiperiodic Points . . . . . . . . . . . . . . . . . . . . . 354 5.7.2 Compact-Valued Dynamical Systems: Asymptotic Behavior of Trajectories . . . . . . . . . . . . 358 5.7.3 A Dynamic Optimization Problem . . . . . . . . . . . . . 363 5.8

Compensatory Transfers and Action Profiles . . . . . . . . 367

6 Mass Transshipment Problems and Ideal Metrics

371

6.1

Kantorovich–Rubinstein Problems with Constraints . . . . 372

6.2

Constraints on the κth Difference of Marginals . . . . . . 383

6.3

The General Case . . . . . . . . . . . . . . . . . . . . . . . 402

6.4

Minimality of Ideal Metrics . . . . . . . . . . . . . . . . . 414

References

429

Abbreviations

473

Symbols

475

Index

487

7 Modifications of the Monge–Kantorovich Problems: Transportation Problems with Relaxed or Additional Constraints

In this chapter we study modifications of the usual transportation problem by allowing additional constraints on the admissible supply—resp. demand—distributions. In particular, we consider the case that the marginal distribution function of the supply is bounded below by a d.f. F1 , while the marginal d.f. of the demand is bounded above by a d.f. F2 . We also examine transportation plans with constraints of a local type concerning the densities of the marginals, and finally, we study transportation problems with additional moment-type constraints. For the solution of these problems we make use of some methods arising in the theory of marginal and moment problems, duality theory, and stochastic ordering results. The next part is concerned with a solution of the tomography paradox. With respect to some weak metrics, two distributions are getting close if they coincide on an increasing number of directions. In the final sections we review results on the closeness of distributions under given momenttype characteristics and discuss applications to the rounding problem. Most of the results in these sections are contained in Rachev and R¨ uschendorf (1993, 1994c), Levin and Rachev (1989), Klebanov and Rachev (1995), and Anastassiou and Rachev (1992). A survey on related discrete transportation problems is given in Burkard, Klinz, and Rudolf (1994).

2

7. Relaxed or Additional Constraints

7.1 Mass Transportation Problem with Relaxed Marginal Constraints For distribution functions F1 , F2 let F(F1 , F2 ) denote the set of all d.f.s F on IR2 with marginals F1 , F2 (i.e., F (x, ∞) = F1 (x), F (∞, y) = F2 (y)). Then the transportation problem with nonnegative cost function c(x, y), x, y ∈ IR, has the form  minimize c(x, y) dF (x, y) over all F ∈ F(F1 , F2 ). (7.1.1) IR2

Usually, in the linear programming setting, F1 is viewed as the supply distribution and F2 as the demand distribution. Clearly, (7.1.1) is an infinitedimensional of the discrete transportation problem: Given ai ≥ analogue  m n 0, bj ≥ 0, i=1 ai = j=1 bj , minimize

n m  

cij xij , subject to the constraints

(7.1.2)

i=1 j=1

n  j=1

xij = ai , 1 ≤ i ≤ m,

m  i=1

xij = bj , j = 1, . . . , n, xij ≥ 0, ∀i, j.

Suppose c(x, y) (resp. (cij )) satisfies the “Monge” conditions, i.e., c is right continuous, and c(x , y  ) − c(x, y  ) − c(x , y) + c(x, y) ≤ 0

for all x ≥ x, y  ≥ y;

(7.1.3)

in the discrete case these conditions are of the form cij + ci+1,j+1 − ci,j+1 − ci+1,j ≤ 0, ∀1 ≤ i < m, 1 ≤ j < n.

(7.1.4)

Then the solution of (7.1.1), (7.1.4) is well known and based on the “northwest corner rule,” which leads to a greedy algorithm; see Hoffman (1961). For (7.1.1) the solution is given by the d.f. F ∗ , F ∗ (x, y) = min{F1 (x), F2 (y)}.

(7.1.5)

F ∗ is the upper Fr´echet bound, see (3.6.2). Recall that the Fr´echet bounds provide the following characterization of F(F1 , F2 ): F ∈ F(F1 , F2 )

if and only if

(7.1.6)

F∗ (x, y) := (F1 (x) + F2 (y) − 1)+ ≤ F (x, y) ≤ F ∗ (x, y) (here (·)+ = max(0, ·)); the lower Fr´echet bound yields the solution of the maximization problem corresponding to (7.1.1).

7.1 Mass Transportation Problem with Relaxed Marginal Constraints

3

In terms of random variables an equivalent formulation of the transportation problem is the following: minimize Ec(X, Y ),

subject to FX = F1 , FY = F2 ,

(7.1.7)

where X, Y are random variables on a rich enough (e.g., atomless) probability space (Ω, A, P ). The solutions (7.1.5), resp. (7.1.6), then can be represented as distributions of r.v.s X ∗ , Y ∗ : X ∗ = F1−1 (U ),

Y ∗ = F2−1 (U )

X ∗ = F1−1 (U ),

Y ∗ = F2−1 (1 − U )

(for (7.1.1), (7.1.5)),

(7.1.8)

resp. (for F∗ ),

(7.1.9)

where U is uniformly distributed on (0, 1), and F1−1 (u) = inf{y; F1 (y) ≥ u} is the generalized inverse of F1 . We next consider the mass transportation problem (7.1.1), but with relaxed marginal constraints. For d.f.s F1 , F2 the set H(F1 , F2 ) = {F ; F is a d.f. on IR2 with marginal d.f.s F1 ≤ F1 , F2 ≥ F2 }

(7.1.10)

of all d.f.s F with F1 (x) = F (x, ∞) ≤ F1 (x), ∀x ∈ IR1 , and F2 (y) = F (∞, y) ≥ F2 (y), ∀y ∈ IR1 . We consider the transportation problem: minimize



c(x, y) dF (x, y),

subject to F ∈ H(F1 , F2 ),

(7.1.11)

R2

or, equivalently, minimize Ec(X, Y ),

subject to FX ≤ F1 , FY ≥ F2 .

(7.1.12)

In the discrete case the problem is to minimize



cij xij ,

(7.1.13)

where for some “supplies” s1 , . . . , sn , a1 ≤ s1 , a1 +a2 ≤ s1 +s2 , . . . , and for some demands d1 , . . . , dn , b1 ≥ d1 , b1 +b2 ≥ d1 +d2 , . . . (ai , bi as in (7.1.2)). This describes a production process and a consumption process subject to some priorities (e.g., queueing priorities) with capacities s1 , . . . , sn having the following property: Every remaining free capacity at stage i of the production (resp. consumption) process can be transferred to some of the next stages i + 1, . . . , n.

4

7. Relaxed or Additional Constraints

Theorem 7.1.1 Let the cost function c(x, y) be symmetric in x, y, let c(x, y) satisfy the Monge condition (7.1.3), and let c(x, x) = 0 for all x ∈ IR. Set H ∗ (x, y) = min{F1 (x), max{F1 (y), F2 (y)}},

x, y ∈ IR.

(7.1.14)

Then (a)

H ∗ ∈ H(F1 , F2 ),

(b)

H ∗ solves the relaxed transportation problem (7.1.11),

(c)



IR2

(7.1.15)

1    c(x, y) dH (x, y) = c F1−1 (u), min F1−1 (u), F2−1 (u) du. ∗

0

Remark 7.1.2 Setting G1 (y) = max{F1 (y), F2 (y)}, we see from Theorem 7.1.1 that the relaxed transportation problem (7.1.11) is equivalent to the transportation problem (7.1.1) with marginals F1 , G1 . In terms of random variables the solution can be expressed by the joint distribution of X ∗ = F1−1 (U ) and   −1 −1 (U ) (U ), F (U ) = min F Y ∗ = G−1 2 1 1

(7.1.16)

(cf. (7.1.8)).

Proof: The Monge condition implies that we can view the function −c(x, y) as a “distribution function” corresponding to a nonnegative measure μc on IR2 . Let X, Y be any real r.v.s, and for x, y ∈ IR1 set x ∨ y = max{x, y}, x ∧ y = min{x, y}. Theorem 7.1.1 is a consequence of the following two claims. Claim 7.1.3 (Cambanis, Simons, and Stout (1976); see also Dall’Aglio (1956) for the special case c(x, y) = |x − y|p )  2Ec(X, Y ) = (P (X < x ∧ y, Y ≥ x ∨ y) IR2

+P (X ≥ x ∨ y, Y < x ∧ y))μc ( dx, dy).

(7.1.17)

For the proof of Claim 7.1.3 define the function f (x, y, w) : IR2 × Ω → IR by ⎧ ⎨ 1 if X(w) < x, y ≤ Y (w) or Y (w) < x, y ≤ X(w), f (x, y, w) = ⎩ 0 otherwise.

7.1 Mass Transportation Problem with Relaxed Marginal Constraints

5

Using Fubini’s theorem,   (Ew f (x, y, w))μc ( dx, dy). (7.1.18) Ew f (x, y, w)μc ( dx, dy) = IR2

IR2

Next, the symmetry of c(x, y) and c(x, x) = 0 yield  f (x, y, w) dμc

(7.1.19)

IR2

= − [c (Y (w), Y (w)) + c (X(w), X(w)) − c (X(w), Y (w)) = 2c (X(w), Y (w)) .

−c (Y (w), X(w))]

Clearly, Ew f (x, y, w)

(7.1.20)

= P (X < x ∧ y, Y ≥ x ∨ y) + P (X ≥ x ∨ y, Y < x ∧ y). Combining (7.1.18), (7.1.19), and (7.1.20) we obtain (7.1.17).   Claim 7.1.4 Define X ∗ = F1−1 (U ), Y ∗ = min F1−1 (U ), F2−1 (U ) . Then Ec(X ∗ , Y ∗ ) = min (Ec(X, Y ); FX ≤ F1 , FY ≥ F2 ) ,

(7.1.21)

and the value of the expectation in (7.1.21) is given by  1 ∗ ∗ max (0, F2 ((x ∧ y)−) − F1 ((x ∨ y)−)) μc ( dx, dy) Ec(X , Y ) = 2 IR2

1    c F1−1 (t), min F1−1 (t), F2−1 (t) dt. =

(7.1.22)

0

For the proof of Claim 7.1.4 let X, Y be any r.v.s with d.f.s FX ≤ F1 , FY ≥ F2 . Using Claim 7.1.3 we obtain  P (X ≥ x ∨ y, Y < x ∧ y)μc ( dx, dy) (7.1.23) 2Ec(X, Y ) ≥ IR2

=



(P (Y < x ∧ y) − P (X < x ∨ y, Y < x ∧ y)) μc ( dx, dy)

IR2





IR2

(P (Y < x ∧ y) − min {P (X < x ∨ y),

6

7. Relaxed or Additional Constraints

P (Y < x ∧ y)}) μc ( dx, dy)

=



(P (Y < x ∧ y) − P (X < x ∨ y))+ μc ( dx, dy)

IR2





(F2 ((x ∧ y)−) − F1 ((x ∨ y)−))+ μc ( dx, dy).

IR2

Next, we check that the lower bound in (7.1.23) is attained for X ∗ =   F1−1 (U ), Y ∗ = min F1−1 (U ), F2−1 (U ) . In fact, by Claim 7.1.3 using X ∗ ≥

Y ∗ and {U < F2 (z)} = F2−1 (U ) < z a.s. we get 2Ec(X ∗ , Y ∗ )  = (P (X ∗ ≥ x ∨ y, Y ∗ < x ∧ y)

(7.1.24)

IR2

=

+P (X ∗ < x ∧ y, Y ∗ ≥ x ∨ y)) μc ( dx, dy)



P (X ∗ ≥ x ∨ y, Y ∗ < x ∧ y)μc ( dx, dy)

IR2

=



IR2

=



IR2

=



    P F1−1 (U ) ≥ x ∨ y, min F1−1 (U ), F2−1 (U ) < x ∧ y μc ( dx, dy)

  P F1−1 (U ) ≥ x ∨ y, F2−1 (U ) < x ∧ y μc ( dx, dy) P (U ≥ F1 (x ∨ y), U < F2 (x ∧ y))+ μc ( dx, dy)

IR2

=



(F2 ((x ∧ y)−) − F1 ((x ∨ y)−))+ μc ( dx, dy).

IR2

Obviously, F(X ∗ ,Y ∗ ) = H ∗ ∈ H(F1 , F2 ), and the proof of Theorem 7.1.1 is complete. 2

Remark 7.1.5 The optimal coupling (7.1.16) leads to the following “greedy” algorithm for solving the finite discrete transportation problem with relaxed side conditions: minimize

n  n 

cij xij

i=1 j=1

subject to:

xij ≥ 0,

(7.1.25)

7.1 Mass Transportation Problem with Relaxed Marginal Constraints j  n  s=1 r=1

n i   r=1 s=1

xrs ≥ xrs ≤

j 

bs =: Gj ,

1 ≤ j ≤ n,

ar =: Fi ,

1 ≤ i ≤ n,

s=1

i  r=1

7

n where the sum of the “demands” s=1 bs equals the sum of the “supplies” n r=1 ar , assuming that (cij ) are symmetric, cii = 0, and c satisfies the Monge condition (7.1.4). Set =

Hi δ1

max(Fi , Gi ), 1 ≤ i ≤ n,

(7.1.26)

= H1 , δi+1 = Hi+1 − Hi , 1 ≤ i ≤ n − 1.

Then (7.1.25) is equivalent to the standard transportation problem (7.1.2) with side conditions (ai ), (δi ). 7In the following example we compare the solution of problem (7.1.25) with inequality constraints with the “greedy” solution of the standard transportation problem with equality constraints (7.1.2). For the problem with inequality constraints we first calculate the new artificial demands δj as in (7.1.26) and then apply the northwest corner rule. supply a1

Example 7.1.6 yij xij

20 10

10

20 20

20 10

10 20 20 10 10 10 10

demand b1 Gj =

j 

bs

10

30

10

40

0

10

10

40

50

90

90

100

20

40

60

90

90

100

20

20

20

30

0

10

Fi =

i 

ar

r=1

20

20

0

20

40

60

20

80

10

90

10

100

s=1

Hj = Fj ∨ Gj δ1 = H1 , δj+1 = Hj+1 − Hj

“artificial” demands

xij = solution of the standard transportation problem (7.1.2), using the classical northwest corner yij = solution of the transportation problem with relaxed side conditions

8

7. Relaxed or Additional Constraints

We next extend the solution to the nonsymmetric case. We relax the symmetry condition, assuming that for any x, y the functions c(x, ·), c(·, y) are unimodal: c(x, y1 ) ≤ c(x, y2 ) if x ≤ y1 ≤ y2 or y2 ≤ y1 ≤ x,

(7.1.27)

c(x1 , y) ≤ c(x2 , y) if x2 ≤ x1 ≤ y or y ≤ x1 ≤ x2 . Theorem 7.1.7 If c(x, x) = 0 for all x ∈ IR and c satisfies the Monge condition and the unimodality condtion (7.1.27), then the relaxed transportation problem minimize Ec(X, Y )

subject to FX ≥ F1 , FY ≤ F2

has a solution, given by the coupling   X ∗ = F1−1 (U ), Y ∗ = max F1−1 (U ), F2−1 (U )

(7.1.28)

(7.1.29)

with joint distribution

  FX ∗ ,Y ∗ (x, y) = min F1 (x), min (F1 (y), F2 (y)) ,

and the optimal value is given by

1    Ec(X ∗ , Y ∗ ) = c F1−1 (u), max F1−1 (u), F2−1 (u) du. 0

Proof: Let X, Y be r.v.s with FX ≥ F1 , FY ≤ F2 . Then by (7.1.8),   −1 Ec(X, Y ) ≥ Ec FX (U ), FY−1 (U ) . (7.1.30)

−1 −1 −1 −1 −1 = Let G(y) X (y), FY (y)). Then FX ≤ F1 , FY ≥ F2 , and G  −1= min(F  −1 max FX , FY . We now need the following

Claim 7.1.8

1 1   −1   −1 (u), G−1 (u) du. c FX c FX (u), FY−1 (u) du ≥ 0

(7.1.31)

0

−1 To show Claim 7.1.8 set (for a fixed u ∈ (0, 1)), x = FX (u), y1 = −1 −1 −1 −1 FX (u) ∨ FY (u) = G (u), and y2 = FY (u).

Case 1: x < y2 . In this case, x ≤ y1 ≤ y2 , and therefore, the unimodality condition (7.1.27) implies c(x, y2 ) ≥ c(x, y1 ). Case 2: y2 ≤ x. In this case, y1 = x, and therefore, y2 ≤ y1 = x. Again by the unimodality condition, c(x, y2 ) ≥ c(x, y1 ). So Claim 7.1.8 holds.

7.1 Mass Transportation Problem with Relaxed Marginal Constraints

9

Claim 7.1.9 The following bound holds for every coupling (X, Y ) : 1   −1 −1 c FX (u) du (u), FY−1 ∨ FX

(7.1.32)

0

1   ≥ c F1−1 (u), F2−1 (u) ∨ F1−1 (u) du. 0

−1 (u), x 2 = FY−1 (u), x1 = F1−1 (u), x2 = For the proof define x 1 = FX 1 ≤ x1 , x2 ≤ x 2 . F2−1 (u) for a fixed u. Then x

If x 1 < x 2 , then x 1 ≤ x 2 ∨ x2 ≤ x 2 . if

x 1 ≥ x 2 , then x 1 = x 1 ∨ x2 ≥ x 2 .

(7.1.33)

From (7.1.33) we obtain Claim 7.1.10

c( x1 , x 1 ∨ x2 ) ≥ c(x1 , x1 ∨ x2 ).

(7.1.34)

For the proof of Claim 7.1.10 we use the relation x1 ≥ x 1 .

1 . Then c( x1 , x2 ) = c( x1 , x 1 ∨ x2 ) ≥ c(x1 , x2 ) = Case 1: x2 > x1 > x c(x1 , x1 ∨ x2 ) by the unimodality condition. Case 2:

(a) x1 ≥ x2 ≥ x 1 . Then, trivially, c( x1 , x2 ) = c( x1 , x2 ∨ x 1 ) ≥ c(x1 , x1 ∨ x2 ) = c(x1 , x1 ) = 0.

1 ≥ x2 . Then again, c( x1 , x 1 ) = c( x1 , x 1 ∨x2 ) ≥ c(x1 , x1 ∨x2 ) = (b) x1 ≥ x c(x1 , x1 ) = 0. Claims 7.1.8, 7.1.9, and 7.1.10 imply (7.1.28).

2

Remark 7.1.11 (a) The unimodality assumption (7.1.27) is quite natural from an application point of view. Note that the transportation problem in Theorem 7.1.7 is the same as in Theorem 7.1.1 (where only the indices 1,2 have been changed). We used this to demonstrate that the solution F ∗ is not unique. Without the symmetry, resp. the unimodality condition, the solution may differ substantially. Given a right continuous function f = f (y) ≥ 0, consider the cost function c(x, y) = f (y). Then

10

7. Relaxed or Additional Constraints

c satisfies the Monge condition, and so (7.1.28) is equivalent to the following problem: minimize



f (y) dFY (y)

subject to FY ≤ F2 .

(7.1.35)

Equivalently, we are seeking a d.f. F2 ≤ F2 such that the distribution of f with respect to F2 has a minimal first moment. Obviously, the solution (7.1.31) of Theorem 7.1.7 is not a solution of (7.1.35).

(b) In the proof of Theorem 7.1.7, the assumption c(x, x) = 0 can be replaced with a weaker one, c(x, x) ≤ c(x, y) ∧ c(y, x),

for all x, y ∈ IR.

(7.1.36)

7.2 Mass Transportation Problem with Fixed Sum (Difference) of the Marginals and with Stochastically Ordered Marginals Consider a flow in a network with n nodes, i = 1, . . . , n, and let xij be the flow from  node i to node j. Assume that for all nodes k the value of  x + j xkj is fixed and equal to hk . As motivation, suppose ai = in ik n x , b = ik i k=1 k=1 xki to be the amount of workload corresponding to the outflow, resp. to the inflow, in node i. Assume that the total work capacity at node i is given by hi (in a certain time period). Then every admissible flow (xij ) should satisfy the condition hi = ai + bi ,

1 ≤ i ≤ n.

(7.2.1)

k k k Set A(k) = i=1 ai , B(k) = i=1 bi , and H(k) = i=1 hi . Then hk = A(k) + B(k) − (A(k − 1) + B(k − 1)), and (7.2.1) is equivalent to H(k) = A(k) + B(k),

1 ≤ k ≤ n.

(7.2.2)

Let cij be the transportation cost of a  unit from node i to node j. Then the problem is to minimize the total cost cij xij subject to the admissibility condition (7.2.1) and xij ≥ 0. The general formulation of this problem is the following. For two d.f.s A, B define G(x) = 21 (A(x) + B(x)). For a given cost function c(x, y),  c(x, y) dF (x, y), subject to F ∈ FA+B . (7.2.3) minimize IR2

7.2 Fixed Sum of the Marginals

11

Here FA+B is the set of all d.f.s F (x, y) with marginal d.f.s F1 , F2 satisfying F1 (x) + F2 (x) = A(x) + B(x). Consider next the special case c(x, y) = |x − y|. Let X, Y be real r.v.s with joint d.f. F . Then by the triangle inequality, E|X − Y | ≤

inf (E|X − a| + E|Y − a|).

a∈IR1

(7.2.4)

 Since E|X − a| + E|Y − a| = |x − a| d(FX + FY )(x) depends only on the sum of the marginals, (7.2.3) is the best possible improvement of (7.2.4), provided that the sum of the marginal FX + FY is known. Rachev (1984d) showed that sup {E|X − Y |p ; FX + FY = A + B} = 1   −1 G (t) − G−1 (1 − t)p dt, p ≥ 1.

(7.2.5)

0

The following result gives an explicit solution of the general problem in (7.2.3). Proposition 7.2.1 Suppose c ≥ 0 is symmetric and satisfies the Monge condition: c(x , y  ) − c(x, y  ) − c(x , y) + c(x, y) ≤ 0

∀x ≥ x, y  ≥ y.

(7.2.6)

1 = c(G−1 (u), G−1 (u)) du,

(7.2.7)

Then inf



c(x, y) dF (x, y); F ∈ FA+B



0

and sup



c(x, y) dF (x, y); F ∈ FA+B



1 = c(G−1 (u), G−1 (1−u)) du.(7.2.8) 0

The corresponding optimal pairs of r.v.s (couplings) are given by (G−1 (U ), G−1 (U )), resp. (G−1 (U ), G−1 (1 − U )). Proof: Since c is symmetric, we obtain for any F ∈ FA+B ,   1 c(x, y) dF (x, y) = (c(x, y) + c(y, x)) dF (x, y) 2  F (x, y) + F (y, x) . = c(x, y) d 2

12

7. Relaxed or Additional Constraints

(y,x) ∈ F(G, G). Consequently, we On the other hand, Fs (x, y) = F (x,y)+F 2 obtain (7.2.7), (7.2.8) by making use of (7.1.8) and (7.1.9) with F1 = F2 = G. 2

We have the following analogue of the above proposition for nonsymmetric cost functions. Proposition 7.2.2 If c(x, y) satisfies the Monge condition (7.2.6) and furthermore, x1 ≤ y ≤ x2 implies that c(x1 , x2 ) ≥ c(y, y), then inf



c(x, y) dF (x, y); F ∈ FA+B



1 c(G−1 (u), G−1 (u)) du. =

(7.2.9)

0

Proof: Applying the Monge conditions for every X, Y with FX,Y ∈ FA+B , −1 (U ), FY−1 (U )). Since FX (x)+FY (x) = 2G(x), it follows Ec(X, Y ) ≥ Ec(FX −1 −1 ∧ FY−1 ≤ G−1 ≤ FX ∨ that FX ∧ FY ≤ G ≤ FX ∨ FY , and therefore, FX −1 −1 −1 −1 −1 FY . Consequently, we have c(FX (U ), FY (U )) ≥ c(G (U ), G (U )), which proves (7.2.9). 2

Remark 7.2.3 The marginals of the class FA+B have largest and smallest elements, defined by ⎧ ⎨ 2 G(x), x < x0 , ∗ F1 (x) = ⎩ 1, x≥x , 0

and

⎧ ⎨ 2 G(x) − 1, ∗ F2 (x) = ⎩ 1,

x < x0 , x ≥ x0 ,

with x0 := inf{y; 2G(y) ≥ 1}. Note that there is no smallest d.f. in FA+B . To show this let F1 (x), F2 (x) be the marginal d.f.s of the smallest elements F ∈ FA+B and let G1 , G2 be d.f.s such that G1 (x) + G2 (x) = 2G(x). If the lower Fr´echet bounds satisfy (F1 (x) + F2 (y) − 1)+ ≤ (G1 (x) + G2 (y) − 1)+ , then F1 ≤ G1 and F2 ≤ G2 , which implies that F1 = G1 , F2 = G2 . In particular, this implies that (G−1 (U ), G−1 (1 − U )) is in the general nonsymmetric case no longer a solution to the problem of maximizing  c(x, y) dF (x, y) over the class FA+B . For example, let G be the d.f. of 1 4

4 

i=1

ε(i) . Then P1 = P (G ∗ −1

while P2 = P ((F1 )

−1

(U ),G−1 (1−U ))

(U ),(F2∗ )−1 (1−U ))

=

= 14 (ε(1,4) + ε(2,3) + ε(3,2) + ε(4,1) ), 1 2 (ε(1,4)

+ ε(2,3) ). For c1 (x, y) =

7.2 Fixed Sum of the Marginals

13

1(−∞,(3,2)] (x, y), we have EP1 c1 = 41 , EP2 c1 = 0, while for c2 = 1[(2,3),∞) , we have EP1 c1 = 14 , EP2 c2 = 12 . Note that both functions, −c1 , −c2 , are Monge functions (but are not unimodal). We next consider the case where in the network example we fix the total outflow minus the inflow of each node. This problem is known in the literature as the minimal network flow problem (cf. for example Barnes and Hoffman (1985, Section 9) or Anderson and Nash (1987)). Assume that the outflow minus the inflow of each node is fixed; i.e., the following Kirchhoff equations hold   xki = ai − bi = hi for all i, xik − k

k

or equivalently, H(k) = A(k) − B(k), with A(k) =

k 

j=1

aj , B(k) =

1 ≤ k ≤ n, k 

bj , and H(k) =

j=1

(7.2.10) k 

hj . Consider now the

j=1

general case: Let A, B be distribution functions and let FA−B be the set of all “generalized” d.f.s of finite measures on IR2 with marginals F1 , F2 satisfying F1 − F2 = A − B. We consider the following transportation problem:  minimize c(x, y) dF (x, y) subject to F ∈ FA−B , (7.2.11) with c(x, y) satisfying the Monge condition (7.2.6). To solve (7.2.11) we make use of the following dual representation (cf. (6.1.23)):   (7.2.12) c(x, y) dF (x, y); F ∈ FA−B inf   f d(A − B)(x); f (x) − f (y) ≤ c(x, y), ∀x, y . = sup We first consider a particular type of cost function. Proposition 7.2.4 Let c(x, y) = |x−y| max(1, h(|x−a|), h(|y−a|)), where h is a monotonically nondecreasing function on IR+ . Then   c(x, y) dF (x, y); F ∈ FA−B inf (7.2.13)  = max(1, h(|x − a|))|A − B|(x) dx, provided that h(|x − a|) is locally integrable.

14

7. Relaxed or Additional Constraints

Proof: We first note that the duality constraints condition f (x) − f (y) ≤ c(x, y), for all x, y, holds if and only if f is absolutely continuous and moreover, |f  (x)| ≤ max(1, h(|x − y|)) a.s. Consequently, by the dual representation (7.2.12), we obtain   c(x, y) dF (x, y); F ∈ FA−B inf    f d(A − B)(x); |f | ≤ max(1, h(|x − a|)), ∀x = sup     = sup f (x) d(A − B)(x) dx; |f | ≤ max(1, h(|x − a|)), ∀x  = max(1, h(|x − a|))|A − B|(x) dx.

2

To handle the general case set c(x, y) = |x − y|ζ(x, y)



c(x, y) i.e., ζ(x, y) = |x − y|



.

(7.2.14)

Theorem 7.2.5 Assume that for any x < t < y, ζ(t, t) ≤ ζ(x, y), ζ(x, y) = ζ(y, x). Moreover, let ζ(x, y) be right continuous in y, and also assume that t → ζ(t, t) is locally bounded. Then the optimal value in the minimization problem (7.2.11) is equal to    = ζ(t, t)|A − B|(t) dt. (7.2.15) c(x, y) dF (x, y); F ∈ FA−B inf Proof: Let F = {f ; f (x)−f (y) ≤ c(x, y), ∀x, y}, and let F ∗ = {f absolutely continuous and |f  (t)| ≤ ζ(t, t), ∀t}. Then F ⊂ F ∗ , and for f ∈ F (y) (y) ≤ ζ(x, y), and therefore, lim f (x)−f ≤ ζ(x, x). Also, we have f (x)−f |x−y| |x−y| y→x

f (y) − f (x) f (x) − f (y) = − lim ≥ − lim ζ(y, x) = −ζ(x, x). |x − y| |x − y| y→x lim

Since f is locally Lipschitz, it is absolutely continuous, so the inequalities above imply that |f  (t)| ≤ ζ(t, t) a.s. If, conversely, f ∈ F ∗ , then y  y  f (x) − f (y) = f (t) dt, and therefore, |f (x) − f (y)| ≤ |f (t)| dt ≤

y x

x

x

ζ(t, t) dt ≤ |x − y|ζ(x, y) = c(x, y). The dual representation (7.2.12) again

implies (7.2.13) (by the same arguments as in the proof of Proposition 7.2.4). 2

7.2 Fixed Sum of the Marginals

15

Next, we consider the following transportation problem with stochastically ordered marginals posed by Rogers (1992). Let F, G be real distribution functions, F ≤st G; here as usual ≤st stands for the stochastic order. Let C := {(x, y) ∈ IR2 ; x ≤ y}, and let MC (F, G) := M (F, G) ∩ {μ ∈ M 1 (IR2 , IB2 ); μ(C) = 1}

(7.2.16)

be the set of all measures with marginals F, G that are concentrated on the order cone C. The problem is to determine, for a given strictly convex function ϕ, the bound   sup ϕ(x − y)μ( dx, dy); μ ∈ MC (F, G) . (7.2.17) The motivation for problem (7.2.17) is to get a good monotone coupling of random walks (Sn ), (Sn ) with S0 = x ≥ X0 = 0, Sn ≥ Sn for all n, and Sn = Sn for all large enough n. Without the order restriction, a solution of (7.2.17) is given by the random variables X = F −1 (U ), Y = G−1 (1 − U ) for a uniform (0, 1) distributed r.v. U. It is intuitively clear that a solution of (7.2.17) should concentrate as much mass on the diagonal as possible. This is indeed true. Theorem 7.2.6 (Rogers (1992)) Each solution (X, Y ) of (7.2.17) has the property that  P (X = Y ) = |F ∧ G| = f ∧ g dm, (7.2.18) when F = f m, G = gm. There exists a solution of (7.2.17). We next characterize the optimal solutions by an order-type relation. Theorem 7.2.7 Let X, Y be r.v.s with d.f.s F, G and X ≤ Y a.s. Then (X, Y ) defines a solution of (7.2.17) iff X(ω) < X(ω  ) ≤ Y (ω) ≤ Y (ω  )

implies Y (ω  ) = Y (ω)

(7.2.19)

a.s. (for (ω, ω  ) and with respect to the product measure). Proof: If (X, Y ) is an optimal admissible coupling and if on a set of pairs (ω, ω  ) with positive measure X(ω) < X(ω  ) ≤ Y (ω) < Y (ω  ) holds, then let us define Y  (ω  ) := Y (ω), Y  (ω) := Y (ω  ) and set Y  = Y otherwise. Then Y  has d.f. G and Eϕ(X − Y  ) > Eϕ(X − Y ) because ϕ is strictly convex. Since there is essentially up to simultaneous rearrangements only one pair of r.v.s X, Y with d.f.s F, G satisfying the order relation (7.2.18), the opposite direction follows from the first part of the proof. 2

16

7. Relaxed or Additional Constraints

In terms of measures μ ∈ MC (F, G), the characterization of optimality of μ in (7.2.19) can be formulated as μ ⊗ μ ({(x1 , y1 , x2 , y2 ); x1 < x2 ≤ y1 < y2 }) = 0.

(7.2.20)

We remark that the characterization of optimal pairs in (7.2.19), resp. (7.2.20), implies the “maximal concentration on the diagonal” property in (7.2.18). For finite discrete distributions one can explicitly construct optimal pairs with the ordering property given in (7.2.19). We consider at nfirst the case 1 of  equiprobable atoms in each distribution. So let μ1 = n i=1 εai , μ2 = n 1 i=1 εbi be the measures corresponding to F, G, where a1 ≤ · · · ≤ an , n b1 ≤ · · · ≤ bn , and ai ≤ bi for all i. Problem (7.2.17) is equivalent to the following problem: Find a permutation π ∈ Υn such that n  i=1

ϕ(bi − aπ(i) ) is maximal.

(7.2.21)

Here, the maximum is considered over all permuations π ∈ Υn such that aπ(i) ≤ bi , 1 ≤ i ≤ n. Permutations with this property are called admissible permutations. An optimal admissible permutation is essentially unique (up to indices with equal values of ai ), and it is given in the following theorem. Theorem 7.2.8 Define π ∗ ∈ Υn inductively: π ∗ (1)

:= max{k ≤ n; ak ≤ b1 } (7.2.22) π ∗ (k) := max{ℓ ≤ n; ℓ ∈ {π ∗ (1), . . . , π ∗ (k − 1)}, aℓ ≤ bk }, 2 ≤ k ≤ n. Then π ∗ ∈ Υ is the optimal admissible permutation. Proof: Define on Ω = {1, . . . , n} (supplied with the uniform distribution P ) random variables X(i) := ai and Y (i) := bπ∗ (i) , 1 ≤ i ≤ n. Then X ≤ Y , since π ∗ is admissible and X, Y satisfy the order relation (7.2.19). Therefore, they are optimal couplings. Equivalently, π ∗ is the optimal admissible permutation. 2 It is clear from the construction in Theorem (7.2.8) that up to a simultaneous permutation of the probability space, an optimal pair of r.v.s is essentially unique. Remark 7.2.9 Theorem 7.2.8 can be extended to the case that μ1 = n n i=1 pi εai , μ2 = i=1 qi εbi with rational pi , qi , by representing pi , qi in the formal equiprobable case. By an approximation argument—as given in Rogers (1992)—one can approximate the optimal couplings for F, G with

7.3 Mass Transportation Problems with Capacity Constraints

17

couplings having compact support. The general case then can be approximated via the ordering criterion (7.2.19) using a truncation technique. Thus, applying Theorem 7.2.8, we are able to construct explicit approximate solutions in the general case.

7.3 Mass Transportation Problems with Capacity Constraints In this section we obtain explicit solutions of Monge–Kantorovich mass transportation problems with capacity constraints. The Hoeffding–Fr´echet inequality is extended for bivariate distribution functions having fixed marginals and satisfying additional constraints. In the discrete case, our results lead to “greedy” algorithms similar to the classical northwest corner rule. Let us start with recalling the abstract version of the MKP: Given two Borel measures μ and ν on a separable metric space S with equal total mass λ = μ(S) = ν(S) < ∞ and a measurable cost function c on S × S, find  Lc (μ, ν) = inf c(x, y)P ( dx, dy), (7.3.1)  Uc (μ, ν) = sup c(x, y)P ( dx, dy), (7.3.2) where the infimum and supremum are taken over all Borel measures P on S × S having projections (marginals) P (· × S) = μ(·),

P (S × ·) = ν(·).

(7.3.3)

As shown in Section 3.1, the explicit solutions of MKP are based on the Hoeffding–Fr´echet inequality (referred to as upper and lower Fr´echet bounds): max(0, F μ (x) + F ν (y) − λ) ≤ F P (x, y) ≤ min(F μ (x), F ν (y)),

(7.3.4)

for any P on IR2 that satisfies (7.3.3) with S = IR. (In (7.3.4) and in the sequel, F P stands for the distribution function of P .) If c is a lattice superadditive (equivalently, −c is a Monge function): c(x , y  ) + c(x, y) ≥ c(x , y) + c(x, y  )

for all x ≥ x, y  ≥ y,

(7.3.5)

then under mild moment conditions on μ and ν the explicit values of Lc and Uc were given in Section 3.1. In this section we consider two marginal problems with additional constraints on the joint distribution functions. Suppose μ and ν are two nonnegative Borel measures on IR, μ(IR) = ν(IR) = λ < ∞. Suppose c : IR2 →

18

7. Relaxed or Additional Constraints

IR is a right-continuous Monge function generating a nonnegative measure on IR2 . Let σ be a nonnegative bounded Borel measure on IR2 . (Note that the total mass of σ may be different from λ.) Problem I.

Find  maximum c(x, y)P ( dx, dy)

(7.3.6)

IR2

subject to the constraints P is a nonnegative Borel measures on IR2

(7.3.7)

with marginals μ and ν, and P ((−∞, x] × (−∞, y]) ≤ σ((−∞, x] × (−∞, y])

(7.3.8)

for all x, y ∈ IR.

Problem II.

Find  minimum c(x, y)P ( dx, dy)

(7.3.9)

IR2

subject to (7.3.7) and P ((−∞, x] × [y, ∞)) ≤ σ((−∞, x] × [y, ∞))

for all x, y ∈ IR. (7.3.10)

Problem I with discrete μ and ν was studied by Barnes and Hoffman (1985). Olkin and Rachev (1990) extended their results by completing the characterization of the “optimal feasible” P; i.e., P satisfies (7.3.7), (7.3.8) and attains the maximum in (7.3.6). This method is extended to solve Problem II as well. We start with a refinement of the Fr´echet bounds (7.3.4). We shall do this by determining the exact bounds for a d.f. F P (x, y) with marginals F μ and F ν assuming that P satisfies the constraint (7.3.8) or (7.3.10). Then we shall apply the extended Fr´echet bounds to solve Problems I and II. Whereas in the discrete case the solution of Problem I leads to the Barnes– Hoffman greedy algorithm, the solution of Problem II implies a new greedy algorithm for a transportation problem with capacity constraints (7.3.10). We begin with some notation. For two nonnegative Borel measures μ and ν on IR with equal total mass λ denote by M (μ, ν) the set of all nonnegative Borel measures on IR2 with projections μ and ν. Without loss of generality set λ = 1. Given a nonatomic probability space, the set F(A, B) of joint d.f.s F (x, y) = FX,Y (x, y) = P (X ≤ x, Y ≤ y) with fixed

7.3 Mass Transportation Problems with Capacity Constraints

19

marginals FX = A and FY = B is the set of d.f.s of the probability laws in M (μ, ν). Thus, the Fr´echet bounds (7.3.4) can be rewritten as max

F (x, y) = F ∗ (x, y) :=

min(A(x), B(y)),

(7.3.11)

max

G(x, y) = G∗ (x, y) :=

min(A(x), B(y)),

(7.3.12)

F ∈F (A,B) F ∈F (A,B)

where B(y) := ν([y, ∞)) and G(x, y) := GX,Y (x, y) := P (X ≤ x, Y ≥ y). Clearly, the laws corresponding to F ∗ and G∗ are in M (μ, ν). Furthermore, given a nonnegative bounded Borel measure σ on IR2 , set F σ (x, y) := σ((−∞, x] × (−∞, y]), Gσ (x, y) := σ((−∞, x] × [y, ∞))

(7.3.13)

F(A, B, F σ ) := {F ∈ F(A, B); F ≤ F σ }, G(A, B, Gσ ) := {GX,Y ; FX,Y ∈ F(A, B), GX,Y ≤ Gσ } . Our objective in the next two theorems is to extend the Fr´echet bounds; we shall characterize the bounds max

F ∈F (A,B,F σ )

max

G∈G(A,B,Gσ )

F (x, y) =: F(x, y),

 y), G(x, y) =: G(x,

x, y ∈ IR,

(7.3.14)

x, y ∈ IR,

(7.3.15)

and shall examine the conditions implying F ∈ F(A, B, F σ )

Theorem 7.3.1 If

 ∈ G(A, B, Gσ ). and G

F σ (x, y) ≥ max(0, A(x) + B(y) − 1),

(7.3.16)

(7.3.17)

then the maximum in (7.3.14) is attained: F(x, y) =

=

inf {F σ (t, s) + μ((t, x]) + ν((s, y])}

t≤x s≤y

(7.3.18)

inf {F σ (t, s) + μ((t, x]) + ν((s, y])} ∧ (A(x) ∧ B(y)),

t≤x s≤y

and F ∈ F(A, B, F σ ), where ∧ := min.

Remark 7.3.2 Condition (7.3.17) is necessary and sufficient for F(A, B, F σ ) = Ø, cf. Fr´echet (1951), Kellerer (1964). Remark 7.3.3 The second equality in (7.3.18) follows from the fact that F σ (t, s) = 0 for t = −∞ or s = −∞.

20

7. Relaxed or Additional Constraints

Remark 7.3.4 From (7.3.18) F is not greater than the Hoeffding–Fr´echet upper bound F ∗ (7.3.11). Remark 7.3.5 By (7.3.4) the maximum in (7.3.11) is attained for the pair X ∗ = A− (U ), Y ∗ = B − (U ), where A− is the generalized inverse of A, and U is uniformly distributed on [0, 1]. In contrast, for F given in (7.3.18),  Y ) with joint d.f. given by F is the explicit form of the optimal pair (X, not known. However, in the discrete case one can use the Barnes–Hoffman greedy algorithm to compute F. Suppose μ, ν, and σ are discrete measures, ai



bj ai

i∈M

σij

:= μ({xi }), i ∈ M = {1, 2, . . . , m}, := ν({yj }), j ∈ N = {1, 2, . . . , },  = bj = 1; j∈N σ

:= F (xi , yj ),

(7.3.19)

i ∈ M, j ∈ N.

(7.3.20)

Then F(xi , yj ) =

j i  

prs ,

(7.3.21)

r=1 s=1

where the probabilities prs are determined by the following variant of the northwest corner rule (see Hoffman (1961), Barnes and Hoffman (1985)); in fact, we set p11 pij

:=

min(a1 , b1 , σ11 ); (7.3.22) ⎧ ⎫ ⎪ ⎪ j−1 i−1 ⎨ ⎬    := min ai − pis , bj − prj , σij − prs , ⎪ ⎪ ⎭ ⎩ r≤i s≤j s=1 r=1 (r,s)=(i,j)

if prs is determined for r ≤ i < m and s ≤ j < n, and we let   j−1 i−1   pij := min ai − prj , if i = m or j = n. pis , bj − s=1

r=1

In other words, taking discrete versions of μ, ν, and σ in (7.3.19) one can apply the greedy algorithm (7.3.22) to approximate F in (7.3.18) by means of (7.3.21). Proof of Theorem 7.3.1: The proof is based on three assertions. Claim 7.3.6 (Fr´echet (1951)) The condition F σ (x, y) ≥ H− (x, y) = max(0, A(x) + B(y) − 1) is necessary and sufficient for F(A, B, F σ ) = Ø.

7.3 Mass Transportation Problems with Capacity Constraints

21

Suppose F(A, B, F σ ) = Ø. Then, by (7.3.4) H− (x, y) ≤ F (x, y) < F σ (x, y),

F ∈ F(A, B, F σ ).

(7.3.23)

On the other hand, if H− ≤ F σ , then H− ∈ F(A, B, F σ ). Claim 7.3.7 F defined by (7.3.18) has marginal d.f.s A and B and for all x, y ∈ IR, sup

F ∈F (A,B,F σ )

F (x, y) ≤ F(x, y).

(7.3.24)

For any F ∈ F(A, B, F σ ) and any t ≤ x, s ≤ y, we have F (x, y) ≤ F σ (t, s) + μ((t, x]) + ν((s, y]), which clearly implies (7.3.24). Invoking Remark (7.3.3), F(x, y) ≤ H+ (x, y) where H+ is the upper Hoeffding–Fr´echet bound, H+ (x, y) := min(A(x), B(y)). Since F ≥ H− (cf. (7.3.23), (7.3.24)), F ∈ [H− , H+ ] has marginals A and B. Theorem 7.3.1 is now a consequence of the following assertion.

Claim 7.3.8 F is a d.f.

To this end, we choose −∞ = x0 < x1 < · · · < xm−1 < xm = ∞, −∞ = y0 < y1 < · · · < yn−1 < yn = ∞ such that μ((xi−1 , xi )) < ε, ν((yn−1 , y1 )) < ε, and σ((xi−1 , xi ) × (yj−1 , yj )) < ε for all i ∈ M = {1, . . . , m} and j ∈ N = {1, . . . , n}. Set ai := μ((xi−1 , xi ]), bj := ν((yj−1 , yj ]), and σij := F σ (xi , yj ). Consider the convex polygon ⎧ ⎨  pij = ai , (7.3.25) p = (pij ) i∈M ; pij ≥ 0, pi· := j∈N ⎩ j∈N  j i    pij = bj , prs ≤ σij , for all i ∈ M, j ∈ N p·j := =



i∈M

r=1 s=1

p; pij ≥ 0, pi· = ai , p·j = bj , j = 1, . . . , n − 1,

j  s=1

j i   r=1 s=1

prs ≤ σij , i = 1, . . . , m − 1,

p·s ≤ σmj , j ∈ N,

i  r=1

pr· ≤ σin , i ∈ M



.

By the Fr´echet condition (7.3.17) (cf. Claim 7.3.6) the marginals of F σ j majorize A and B, respectively, and thus σmj ≥ s=1 p·s and σin ≥ i r=1 pr· for all j ∈ N, i ∈ M . The polygon (7.3.25) becomes  p; pij ≥ 0, pi· = ai , p·j = bj

for all i ∈ M, j ∈ N,

(7.3.26)

22

7. Relaxed or Additional Constraints j i   r=1 s=1



prs ≤ σij

for all i = 1, . . . , m − 1, j = 1, . . . , n − 1 .

Consider now the discrete analogue of F in (7.3.18): dij

:=

dij

:=

min {σrs + ar+1 + · · · + ai + bs+1 + · · · + bj },

0≤r≤i 0≤s≤j

0

(7.3.27)

if i = 0 or j = 0,

where σrs = 0 if r = 0 or s = 0. Our aim now is to show that d = (dij ) i∈M j∈N

determines a bivariate d.f. with support on X × Y, X = (xi )i∈M , Y = (yj )j∈N .

Claim 7.3.9 The greedy algorithm (7.3.22) is determined uniquely by (7.3.27); i.e., j i  

dij :=

prs ,

r=1 s=1

i ∈ N, j ∈ M.

(7.3.28)

Proof: Consider the discrete version of F (cf. (7.3.21), (7.3.25)). Let σr,s := 0 if r = 0 or s = 0, and define dij

:=

dij

=

min {σrs + ar+1 + · · · + ai + bs+1 + · · · + bj },

0≤r≤i 0≤s≤j

(7.3.29)

0 if i = 0 or j = 0.

We need to check the equality dij =

j i   r=1 s=1

prs ,

i ∈ M, j ∈ N,

(7.3.30)

where the pij ’s are determined by the greedy algorithm (7.3.22). If i = j = 1, then p11 = min(a1 , b1 , σ11 ) (cf. (7.3.22)), and by (7.3.29) d11 = min{σ11 + a1 + b1 , σ11 + a1 , σ10 + b1 , σ11 } = p11 . Suppose we have proved that d1,j−1 = p11 + · · · + p1,j−1 . Then p11 + · · · + p1j

(7.3.31)

7.3 Mass Transportation Problems with Capacity Constraints

=

j−1  s=1



p1s + min a1 −

j−1  s=1

p1s , bj , σ1j −

j−1 

p1s

s=1

23



min{a1 , bj + d1,j−1 , σ1,j } = min{a1 , b1 + · · · + bj , σ11 + b2 + · · · + bj , . . . , σ1,j−1 + bj , σi,j } = d1,j . =

These equalities hold due to (7.3.22), (7.3.31), and (7.3.29), respectively. By symmetry, di,1 = p11 + · · · + pi1 . Suppose next that drs =

s r  

for all r ≤ i, s ≤ j, (r, s) = (i, j).

pkl

k=1 l=1

(7.3.32)

Then for 1 ≤ i ≤ m − 1, 1 ≤ j ≤ n − 1,   j−1 j j i  i  i−1     prs , σij = dij , prs , bj + prs = min ai + r=1 s=1

r=1 s=1

r=1 s=1

where the equalities follow from (7.3.22) and (7.3.32). Thus dij =

j i  

prs

r=1 s=1

for all 1 ≤ i ≤ m − 1, 1 ≤ j ≤ n − 1.

(7.3.33)

Consider now the case i = m. Then,   m m−1 m−1    pr1 = pr1 + min am , b1 − pr1 r=1

r=1

=

r=1

min{am + dm−1,1 , b1 } = dm,1 ,

which follows from (7.3.22) and (7.3.33). Suppose that dm,j−1 =

j−1 m  

prs .

(7.3.34)

r=1 s=1

Then using (7.3.22), (7.3.33), and (7.3.34), for 1 ≤ j ≤ n,   j j−1 n m  m−1 m     prs = min am + prs , bj + prs r=1 s=1

r=1 s=1

=

r=1 s=1

min{σrs + ar+1 + · · · + am bs+1 + · · · + bj },

for 0 ≤ r ≤ m, 0 ≤ s ≤ j, (r, s) = (m, j); m m   r=1 s=1

prs = dm,j ,

for r = m, s = j, σm,j = F σ (∞, yj ) ≥ bj .

24

7. Relaxed or Additional Constraints

Similarly, di,n =

i  n 

r=1 s=1

prs , for all i ∈ M , which proves Claim 7.3.9.

2

The greedy algorithm (7.3.22) defines nonnegative pij ’s (cf. Barnes and Hoffman (1985, Lemma 3.2)). Define the probability P (ε) on X × X by P (ε) ((−∞, xi ], (−∞, yj ]) := dij ,

i ∈ M, j ∈ N.

(7.3.35)

Similarly, (ai )i∈M and (bi )i∈N determine probabilites μ(ε) and ν (ε) with supports X and Y , respectively. If ̺ is the Kolmogorov (uniform) distance ̺(μ, ν) := sup |F μ (x) − F ν (x)| ,

(7.3.36)

x∈IR

    then the sequences μ(ε) ε>0 and ν (ε) ε>0 are ̺-relatively compact, and thus there exists εn ↓ 0 such that     (εn ) (εn ) ̺ μ , μ → 0 and ̺ ν , ν → 0. (7.3.37) (For more facts on ̺-relative compactness cf. Rachev (1984a) and Kakosjan, Klebanov, and Rachev (1988, Sec. 2.5).) Similarly, by definition of σij := F σ (xi , yj ), σ((xi−1 , xi )×(yj−1 , yj )) < ε we have that (σij ) i∈M determines a measure σ (ε) on X × Y . Again, the i∈N  (ε)  is ̺-relatively compact. Thus, without loss of generality, family σ ε>0 we may assume that as εn → 0,   (ε )     σ n σ (εn ) (7.3.38) ̺ σ (x, y) − F (x, y) → 0. , σ = sup F x,y∈R

As inClaim 7.3.7, we conclude that P (ε) has marginals μ(ε) and ν (ε) , and  thus P (ε) ε>0 is tight. By (7.3.37), (7.3.38), (7.3.26), and (7.3.18), there ′ exists a subsequence {εn } ⊂ {εn } such that P (εn ) weakly converges to a measure P with d.f. F. The proof of Theorem 7.3.1 is now complete. 2 The next theorem provides an explicit expression for the Fr´echet type bound (7.3.15). We recall the notations (7.3.11)–(7.3.13).

Theorem 7.3.10 Suppose Gσ (x, y) := σ((−∞, x] × [y, ∞)) (cf. (7.3.13)) satisfies the condition   ◦  ◦ B(y) := ν((−∞, y)) . (7.3.39) Gσ (x, y) ≥ max 0, A(x) − B(y)

7.3 Mass Transportation Problems with Capacity Constraints

25

Then the maximum in (7.3.15) is attained, and  y) = inf {Gσ (t, s) + μ((t, x] + v([y, s))} . G(x, t≤x s≥y

(7.3.40)

Conclusions similar to those in Remarks 7.3.2–7.3.5 can be made. Here we shall only point out the greedy algorithm that can be used to approximate  We use the notations (7.3.19) again and let the optimal distribution G. λij := Gσ (xi , yj ),

i ∈ M, j ∈ N.

(7.3.41)

 has the form Then, in this discrete case, G  i , yj ) = G(x

i  n 

prs ,

(7.3.42)

r=1 s=j

where the probabilites pij are determined by the following southwest corner rule: pin

:=

pij

:=

min{a1 , bn , λ1n }; (7.3.43) ⎫ ⎧ ⎪ ⎪ i−1 n ⎨ ⎬    prs , (7.3.44) prj , λij − pis , bj − min ai − ⎪ ⎪ ⎭ ⎩ r≤i s≥j r=1 s=j+1 (r,s)=(i,j)

if i = m or j = 1.

if prs is determined for r ≤ i ≤ m − 1 and s ≥ j > 1; moreover, ⎧ ⎫ n i−1 ⎨   ⎬ pij := min ai − pis , bj − prj , if i = m or j = 1. ⎩ ⎭ s=j+1

(7.3.45)

r=1

We now give explicit solutions of the marginal problems I and II.

Theorem 7.3.11 Suppose (i)

c : IR2 → IR is a right-continuous lattice superadditive function (−c is a Monge function);

(ii)

μ and ν are two Borel nonnegative measures on IR with μ(IR) = ν(IR) = λ < ∞ and d.f.s F μ and F ν , and such that



IR

c(x, y0 )μ( dx) +



IR

c(x0 , y)ν( dy) < ∞

for some x0 , y0 ∈ IR;

(7.3.46)

26

7. Relaxed or Additional Constraints

(iii)

σ is a nonnegative bounded Borel measure on IR2 and F σ (x, y) ≥ max(0, F μ (x) + F ν (y) − λ)

for all x, y ∈ IR. (7.3.47)

Then the maximum in (7.3.6) is attained at the “optimal” measure P. P satisfies the feasibility conditions (7.3.7), (7.3.8) and is determined by 

F P (x, y) := inf {F σ (t, s) + μ((t, x]) + ν((s, y])} , t≤x s≤y

x, y ∈ IR.

(7.3.48)

Proof: We need Theorem 3.1.2 (cf. Cambanis, Simons, and Stout (1976, Theorem 1); see also Rachev (1991c, Section 7.3)). If (7.3.5) holds, then for measures P1 and P2 on IR2 with marginals μ and ν,   P1 P2 F ⇒ ≤ F c dP1 ≤ c dP2 , (7.3.49) IR2

IR2

which with an appeal to Theorem 7.3.1 yields the result.

2

Remark 7.3.12 The assumption (7.3.46) can be replaced by one of the following assumptions: 

c(x, x)(μ + ν)( dx) < ∞.

(a)

c(x, y) is symmetric, and

(b)

c(x, y) is uniformly integrable for all P with marginals μ and ν.

That (a) implies (7.3.49) follows from Cambanis, Simons, and Stout (1976); that (b) implies (7.3.49) follows from Tchen (1980, Corollary 2.1); see also Rachev (1991c, Theorem 7.3.2). Remark 7.3.13 Condition (7.3.47) guarantees that the set of feasible solutions P determined by (7.3.7), (7.3.8) is not empty. Remark 7.3.14 If F σ (x, y) ≥ min(F μ (x), F ν (y)) =: H+ (x, y), 

(7.3.50)

then F P in (7.3.48) equals H+ (cf. Remark 7.3.3). Thus, Theorem 7.3.11 (see also the next Theorem 7.3.17) can be considered as a generalization of Theorem 2 of Cambanis, Simons, and Stout (1976) and Corollary 2.2 of Tchen (1980). In this case, Hoffman’s (1962) northwest corner rule gives a greedy algorithm to determine an “optimal” measure P, provided that μ and ν have finite discrete support.

7.3 Mass Transportation Problems with Capacity Constraints

27

Remark 7.3.15 Consider the discrete version of Problem I (see (7.3.6)). Suppose c(i, j), i ∈ M, j ∈ N , is a lattice superadditive sequence c(i, j) + c(i + 1, j + 1) ≥ c(i, j + 1) + c(i + 1, j), (7.3.51) i = 1, . . . , m − 1, j = 1, . . . , n − 1. Hoffman (1961) and Barnes and Hoffman (1985) treat c(i, j) as the (negative) cost of shipping a unit commodity from origin i to destination j. Suppose the discrete measures μ and ν with supports M and N are given. Then ai = μ{i} and bj = ν{j} are interpreted as the amount of a product available at i and the amount required at destination j. Suppose the (m − 1) × (n − 1) matrix (σij ) satisfies ⎧ ⎫ i n ⎨   ⎬ σij ≥ max 0, bs , ar − (7.3.52) ⎩ ⎭ r=1

s=j+1

σij ≤ σis , σij ≤ σrj , σij + σrs ≥ σis + σrj , r ≥ i, s ≥ j.

(These conditions are related to what is called a uniformly tapered matrix; see Marshall and Olkin (1979).) Barnes and Hoffman (1985) consider the following transportation problem:  maximize c(i, j)pij (7.3.53) i∈M j∈N

subject to

j i  

pij

≥ 0, pi· = ai , p·j = bj

pij

≤ σij ,

r=1 s=1

for all i ∈ M, j ∈ N, (7.3.54)

i = 1, . . . , m − 1, j = 1, . . . , n − 1.

Clearly, (7.3.54) is a special case of Problem I. Following Barnes and Hoffman, (7.3.54) can be viewed as the capacity restrictions on the amount that can be shipped from the first i origins to the first j destinations. Theorem 7.3.11 is completed by showing that the greedy algorithm of Barnes and Hoffman (1985) for determining the solution (pij ) i∈M of (7.3.53) is also j∈N characterized by P

F (i, j) :=

j i  

prs

(7.3.55)

r=1 s=1

=

σrs

=

min {σrs + ar+1 + · · · + ai + bs+1 + · · · + bj } ,

0≤r≤i 0≤s≤j

⎧ ⎨ 0 if r = 0 or s = 0, ⎩ +∞ if r = m or s = n.

28

7. Relaxed or Additional Constraints

Remark 7.3.16 One can determine the extremal value in (7.3.6): max μ ν



F ∈F (F ,F ;F σ ) R2

c dF =





c dF P ,

IR2

(7.3.56)



where F P is given by (7.3.48). By (7.3.46) and Cambanis, Simons, and Stout (1976, p. 288, (9)), 

IR2

c dF = 

(7.3.57) μ

c(x, y0 )F ( dx) +

IR



IR

ν

c(x0 , y)F ( dy) − c(x0 , y0 ) +



B(x, y)μc ( dx, dy)

IR2 

for any bivariate d.f. F with marginals F μ and F ν . (Since F P ∈ μ ν· σ F(F  , F , F ) by Theorem 7.3.1, (7.3.57) can be used to compute the value 

c dF P . In (7.3.57) the points x0 and y0 are the same as in condition

of

(7.3.23), the measure μc is generated by c (see condition (i) in Theorem 7.3.11), and we also assume that c is a nondecreasing function in both arguments.) Finally,

(7.3.58) := B1 − B2 ⎧ ⎪ ⎪ 1 + F (x, y) − F μ (x) − F ν (y) if x0 < x, y0 < y, ⎪ ⎨ B1 (x, y) := F (x, y) if x ≤ x0 , y ≤ y0 , ⎪ ⎪ ⎪ ⎩ 0 otherwise; ⎧ ⎪ ⎪ F μ (x) − F (x, y) if x ≤ x0 , y0 ≤ y, ⎪ ⎨ B2 (x, y) := F ν (y) − F (x, y) if x0 < x, y ≤ y0 , ⎪ ⎪ ⎪ ⎩ 0 otherwise. B

Theorem 7.3.17 Suppose conditions (i) and (ii) of Theorem 7.3.11 hold, and in addition ∗

(iii )

σ is a nonnegative bounded Borel measure on IR2 satisfying

Gσ := σ ((−∞, x], [y, ∞)) ≥ max (0, F μ (x) − ν((−∞, y)))

(7.3.59)

for all x, y ∈ IR. Then, the minimum in (7.3.9) is attained at an optimal measure Q satisfying the feasibility conditions (7.3.7) and (7.3.10); Q is

7.3 Mass Transportation Problems with Capacity Constraints

29

determined by GQ (x, y) = Q((−∞, x] × [y, ∞)) = inf {Gσ (x, y) + μ((t, x]) + ν([y, s))} .

(7.3.60)

t≤x s≥y

All the Remarks 7.3.12–7.3.16 can be easily reformulated regarding Theorem 7.3.17. In particular, consider the transportation problem  minimize c(i, j)pij (7.3.61) i∈M j∈N

subject to

i  n 

pij

≥ 0, pi· = ai , p·j = bj

prs

≤ λij , i = 1, . . . , m − 1, j = 1, . . . , n − 1.

r=1 s=j



for all i ∈ M, j ∈ N, (7.3.62)



bj and c(·, ·) is a lattice superadditive sequence    j i (cf. (7.3.51)). Suppose also that λij ≥ max 0, r=1 ar − s=1 bs and for any r < i and s > j the inequalities Suppose

i∈M

ai =

j∈N

λrj ≤ λij ≥ λis , λij + λrs ≥ λis + λrj ≥ 0, r < i, s > j,

(7.3.63)

hold. Then the greedy algorithm (7.3.43)–(7.3.45) realizes the minimum in (7.3.61). Moreover, the optimal pij ’s are determined by pij = fij − fi,j+1 − fi−1,j + fi−1,j+1 , where fij

:=

min {λrs + (ar+1 + · · · + ai ) + (bj + · · · + bs−1 )}

1≤r≤i j≤s≤n



i  r=1

ar ∧

n 

bs .

(7.3.64)

s=j

The rest of this section is devoted to a generalization of the MKP with additional constraints stated in Problems I and II; see (7.3.6)–(7.3.10). The results are motivated by Hoffman and Veinott (1990), where the discrete version of the problem has been considered. We shall only state the results. The proofs are similar to those of Theorems 7.3.1 and 7.3.10 and will therefore be omitted. The abstract form of the problem is the following. Suppose that (i)

μ and ν are two nonnegative Borel measures on IR, μ(IR) = ν(IR) = λ < ∞;

30

7. Relaxed or Additional Constraints

(ii) L is a union of disjoint sublattices Li ⊂ IR2 , i ∈ S, and the projections of L on each axis equal IR; (iii) (σi )i∈S are nonnegative σ-finite Borel measures on Li . Then the problem is to find  min c dP,

(7.3.65)

L

where the minimum is subjet to the following constraints: (i)

P ’s are nonnegative Borel measures on L with marginals μ and ν;

(7.3.66)

(ii)

P (A) ≤ σi (A)

(7.3.67)

for any A = Li ∩ (−∞, x] × (−∞, y],

(x, y) ∈ Li , i ∈ S.

As before, see (7.3.1)–(7.3.3), the measures μ and ν are viewed as initial and final mass distributions, and P in (7.3.66), (7.3.67) are the (feasible) transportation plans. Here the generalization of problems I and II is that L describes the path of the transportation flow and σi ’s are capacity constraints on the cumulative supply–demand flow. Finally, c : L → R is a cost function, and therefore, the integral in (7.3.65) represents the total cost of mass transportation applying the plan P . Suppose c is subadditive on the lattice L; that is; for all x, y ∈ L, f (x) + f (y) ≥ f (x ∧ y) + f (x ∨ y). Then we shall call a feasible plan of transportation achieving the minimum in (7.3.65) an optimal measure P ∗ . As in problems I and II we start with extensions of the classical Hoeffding– Fr´echet bounds (7.3.4), assuming that P meets the constraints (7.3.66) and (7.3.67), or their alternatives:  Li (7.3.68) P is a nonnegative Borel measure on L = 

i∈S

 Li := {(x, y); (x, −y) ∈ L} with marginals μ and ν ;

P (B) ≤ σi (B) for any B = Li ∩ ((−∞, x] × [y, ∞)). The restriction on the support of P given in (7.3.66) has the form  L = Li , where S = {0, 1, . . . , s}, i∈S

(7.3.69)

7.3 Mass Transportation Problems with Capacity Constraints

31

+ − + or S = IN, and each sublattice Li is a rectangle (x− i , xi ] × (yi , yi ], where − − + − + − − + + − x− 0 = y0 = −∞, xi < xi , yi < yi , xi−1 ≤ xi ≤ xi−1 ≤ xi , yi−1 ≤ + + yi− ≤ yi−1 ≤ yi+ , x+ s = ys = ∞. Write PL (resp. PL ) to denote the class of all P ’s on L with (7.3.66) and (7.3.67) (resp. (7.3.68), (7.3.69)). Recall that for any measure P on IR2 , F P stands for the d.f. of P , and GP (x, y) = P ((−∞, x]×[y, ∞)). In the next two theorems we shall compute the bounds

F ∗ (x, y) = max F P (x, y)

(7.3.70)

G∗ (x, y) = max GP (x, y).

(7.3.71)

P ∈PL

and P ∈PL

For L = IR2 and σi = +∞, F ∗ is indeed the upper Hoeffding–Fr´echet bound H+ (x, y) = min {F μ (x), F ν (y)}

(F μ (x) := μ((−∞, x]) . (7.3.72)

On the other hand, G∗ (x, y) = min {F μ (x), Gν (y)} determines a measure with d.f.

(Gν (x) := ν([y, ∞)))

H− (x, y) = max (0, F μ (x) + F ν (y) − λ) ,

(7.3.73)

which is the lower Hoeffding–Fr´echet bound. Theorem 7.3.18 Suppose that F ∗ : L → IR is defined iteratively as follows: F ∗ (x, y) =

min

[μ((u, x]) + ν((v, y]) + F σ0 (u, v)]

(7.3.74)

min

[μ((u, x]) + ν((v, y]) + F σi (u, v)

(7.3.75)

+ −∞ 0, if u < x0 .

(7.4.9)

7.4 Local Bounds for the Transportation Plans

39

Then for any cost function c satisfying the Monge condition (7.1.3) and the unimodality condition (7.1.27) we have inf



c(x, y) dfP (x, y); P ∈ Pμμ12



1   −1 (u), F (u) du, = c Fμ−1 2 1

(7.4.10)

0

where F2 is the d.f. of the measure with density f2 with respect to μ. The optimal distribution is determined by the r.v.s X ∗ = Fμ−1 (U ), Y ∗ = F2−1 (U ). 1 Proof: Invoking the Monge condition, for any P ∈ Pμμ12 with marginals  1  −1  Fμ1 , G2 , we have c(x, y) dFP (x, y) ≥ c Fμ1 (u), G−1 2 (u) du. 0

By the definition of F2 , G2 (y) F2 (y)

≥ F2 (y) ≥ Fμ1 (y)

=

for all y ≥ x0 , for all y ≤ x0 ;

Fμ1 (y)

and

(7.4.11)

in fact, (7.4.11) implies that Fμ−1 (u) ≥ F2−1 (u) ≥ G−1 2 (u) for u > F2 (x0 ) 1 −1 −1 and F2 (u) = Fμ1 (u) for u ≤ F2 (x0 ). Our assumptions on c imply that     −1 −1 −1 c Fμ−1 (u) for all u. 2 (u) ≥ c F (u), F , G μ1 2 2 1 Remark 7.4.4 It is not difficult to extend the solution of Theorem 7.4.3 to the case μ1 (IR1 ) < 1 and to the case f1 ≥ h1 , f2 ≤ h2 for the densities (here, f1 and f2 are the marginal densities of an admissible plan P ), if we still keep the assumption (7.4.8). To see this, choose x0 as in (7.4.9), and define

f2 (x) =

where z0

y0

⎧ h2 (x)  ⎪ ⎪ ⎪ ⎪ ⎪ h2 (x) dμ(x) 1− ⎪ ⎪ ⎪ ⎨ (z0 ,∞)

⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩

0 ⎧ ⎪ ⎨ = inf x; ⎪ ⎩

⎧ ⎪ ⎨ = inf y; ⎪ ⎩

if x > z0 ,

if x = z0 and μ(z0 ) > 0,

μ(z0 )

otherwise, ⎫ ⎪  ⎬ h2 (x) dμ(x) ≤ 1 . Define next ⎪ ⎭

(x,∞)



(y,∞)

h2 (x) dμ(x) ≤



(y,∞)

h1 (x) dμ(x)

⎫ ⎪ ⎬ ⎪ ⎭

(7.4.12)

40

7. Relaxed or Additional Constraints

and ⎧ h1 (x) ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ f2 (x) ⎪ ⎪ ⎪ ⎨  f1 (x) = (h2 (x) − h1 (x)) dμ(x) ⎪ ⎪ ⎪ ⎪ ⎪ [y0 ,∞) ⎪ ⎪ ⎪ ⎪ μ(y0 ) ⎩

if x > y0 , if x < y0 , (7.4.13) if μ(y0 ) > 0.

Then for a cost function c defined as in Theorem 7.4.3, we have   c(x, y) dFP (x, y); π1 P ≥ μ1 , π2 P ≤ μ2 inf

(7.4.14)

1   = c F1−1 (u), F2−1 (u) du, 0

where Fi have densities fi with respect to μ, i = 1, 2. Let us return to the comment we made in Remark 7.3.26. Consider transportation problems with local upper bounds on the transportation plans xij ≤ μij in the discrete case, while P ≤ μ for some finite measure μ in the general case. The following framework allows us to handle quite general transportation problems. On a measurable space (X, B), let Bi ⊂ B, be sub-σ-algebras, 2 ≤ i ≤ n, Pi ∈ M 1 (X, Bi ). Further, let μ be a finite measure on (X, B), and define

Mμ := P ∈ M 1 (X, B); P/Bi = Pi , 1 ≤ i ≤ n, P ≤ μ . (7.4.15)

Ø and define the set of generalized transportation plans Assume that Mμ = with local upper bound μ as follows:    Uμ (ϕ) := inf U (ϕ0 ) + h dμ; h ≥ 0, ϕ0 + h ≥ ϕ , (7.4.16) where U (ϕ0 ) := inf

 n   i=1

fi dPi ; fi ∈ L1 (Bi , Pi ), ϕ0 ≤

n  i=1

fi



.

We view U as the dual operator for the “pure” transportation problem, and typically,   ϕ0 dP ; P ∈ M (P1 , . . . , Pn ) sup = U (ϕ0 ) (7.4.17)

7.4 Local Bounds for the Transportation Plans

41

will hold (cf. Chapter 2). The duality principle allows us to infer the corresponding “minimization” problem. Similarly, Uμ is the dual operator for the local majorized transportation problem. A linear operator S is majorized by Uμ , S ≤ Uμ if and only if S ≥ 0, S/Bi = Pi , 1 ≤ i ≤ n, and S ≤ μ. (7.4.18) Therefore, the approach developed in Chapter 2 yields the duality theorem   =: Mμ (ϕ) (7.4.19) ϕ dP ; P ∈ Mμ Uμ (ϕ) = sup for any upper semicontinuous or uniformly approximable integrable functions ϕ in the case of a compactly approximable measure space (X, Bi , Pi ) with countable topological basis. In some sense (7.4.19) gives the duality result for the general case of order restrictions as considered, for example, in Sections 3.5 and 5.5. In particular, we obtain upper bounds n Mμ (ϕ) i=1 fi dPi for any admissible system of functions (fi ) with n≤ ϕ ≤ i=1 fi .

We next consider the question of more explicit evaluations of the dual operator Uμ for the case ϕ = 1B , B ∈ B; or equivalently, we wish to establish sharp upper Fr´echet bounds in the class Mμ . Define Mμ (B) := Mμ (1B ), and assume the duality (7.4.19) for φ = 1B . Theorem 7.4.5 Mμ (B) =

sup P ∈M (P1 ,...,Pn )

P ∧ μ(B),

(7.4.20)

where P ∧ μ is the infimum in the lattice of measures. Proof: From (7.4.19), Mμ (B) = inf{μ(h) + U (ϕ); h ≥ 0, ϕ + h ≥ 1B } = inf{μ(h) + U (1B − h); 0 ≤ h ≤ 1B }.

(7.4.21)

To see the second equality in (7.4.4), take ϕ = (1B − h)+ . Thus, 0 ≤ ϕ, and it is possible to assume that h ≤ 1B . Next, we make use of the “integration” approach in Strassen (1965), Mμ (B) =

=

inf

sup

inf

⎧ 1 ⎨

0≤h≤1B P ∈M (P1 ,...,Pn )

sup

0≤h≤1B P



0

{μ(h) + P (1B − h)}

μ(h > t) dt +

1 0

(7.4.22)

⎫ ⎬ P (1B − h ≥ 1 − t) dt . ⎭

42

7. Relaxed or Additional Constraints

With Ct := {h > t} ⊂ B we see that {x; h(x) ≤ 1B (x) − 1 + t} = {x ∈ B; h(x) ≤ t} = B \ Ct . Therefore, Mμ (B) =

inf

sup

0≤h≤1B P

1 0

(μ(Ct ) + P (B \ Ct )) dt

≥ sup inf {μ(C) + P (B \ C)} = sup μ ∧ P (B). P

C⊂B

P

On the other hand, Mμ (B) = sup{P (B); P ∈ M (P1 , . . . , Pn ), P ≤ μ} = sup{P ∧ μ(B); P ∈ M (P1 , . . . , Pn ), P ≤ μ} ≤ sup P ∧ μ(B). P ∈M (P1 ,...,Pn )

2 Theorem 7.4.5 allows us to reduce the problem of the majorized Fr´echet boundsto a problem of “usual” Fr´echet bounds, but for a more complicated functional. It remains an open problem to determine more explicit formulas for Mμ (B) in the general case.

7.5 Closeness of Measure with Joint Marginals on a Finite Number of Directions In this section we follow the work of Kakosjan and Klebanov (1984), Khalfin and Klebanov (1990), Klebanov and Rachev (1995a, 1995b, 1995c), on the application of marginal problems to computer and diffraction tomography. Here, estimates of the closeness between probability measures defined on IRn that have the same marginals on a finite number of arbitrary directions will be provided. The estimates show that the probability laws get closer in a certain metric when the number of coinciding marginals increases. The results offer a solution to the computer tomography paradox stated in Gutman, Kemperman, Reeds, and Shepp (1991). We start with some historical remarks and with the statement of the problem. Let Q1 and Q2 be a pair of probabilities on IR, i.e., probability measures defined on the Borel σ-field of IR. Lorentz (1949) studied conditions for the existence of a probability density function g(·) on IR2 taking

7.5 Closeness of Measure on a Finite Number of Directions

43

only two values, 0 or 1, and having Q1 and Q2 as marginals. In his 1961 paper Kellerer generalized this result and gave necessary and sufficient conditions for the existence of a density f (·) on IR2 that satisfies the inequalities 0 ≤ f (·) ≤ 1 and has Q1 and Q2 as marginals (see also Strassen (1965) and Jacobs (1987)). Fishburn et al. (1990) were able to show that Kellerer’s and Lorentz’s conditions are equivalent; i.e., for any density 0 ≤ f ≤ 1, on IR2 there exists a density taking the values 0 and 1 only that has the same marginals. In general, similar results hold for probability densities on IRm , m ≥ 2, when the (m − 1)-dimensional marginals are prescribed. Gutmann et al. (1991) show that for any probability density 0 ≤ f ≤ 1 on IRm and for any finite number of directions, there exists a probability density taking the values 0, 1 only that has the same marginals in the chosen directions. It follows that densities having the same marginals in a finite number of arbitrary directions may differ considerably in the uniform metric between densities, which is indeed a very strong metric; recall that convergence in the uniform metric implies convergence in total variation. The goal in this section is to show that under moment-type conditions, measures having a “large” number of coinciding marginals are close to each other in the weakmetrics.(1) The method is based on techniques used in the classical moment problem. On the other hand, most of our results will make use of relationships between different probability metrics, analyzed in the monograph by Kakosjan, Klebanov, and Rachev (1988), referred to below as KKR (1988). The key idea in showing that measures with a large number of common marginals are close to each other in the weak metrics is best understood by comparing three results. The first is the theorem of Gutman et al. (1991) mentioned above. The second (see Karlin and Studden (1966, p. 265)) states that if a finite number of moments μ1 , . . . , μn of a function f , 0 ≤ f ≤ 1, are given, then there exists a function g that takes the values 0 or 1 only and possesses the moments μ1 , . . . , μn . Finally, the third result (see KKR (1988, pp. 170–197)) gives estimates of the closeness in terms of a weak metric (the so-called λ-metric) on IR for measures having a finite number of common moments. Of course, since the condition of common marginals seems to be more restrictive than the condition of equal moments, one should be able to construct a similar estimate expressed in terms of the common marginals only. Furthermore, the technique should be similar to that used here. For simplicity, let us consider the 2-dimensional case. Let θ1 , . . . , θn be n unit vectors in the plane and P1 , P2 be two probabilities on IR2 having the same marginals in the directions θ1 , . . . , θn . To estimate the distance (1) Here

weak metric stands for a metric metrizing the weak convergence in the space of probability measures on a Euclidean space.

44

7. Relaxed or Additional Constraints

between P1 and P2 , various weak metrics can be used; however, it seems that the λ-metric is the most convenient for this purpose. This metric is defined as follows (see, for example, Zolotarev (1986)): Let  ei(t,x) Pi ( dx), i = 1, 2, ϕi (t) = IR2

be the characteristic function of Pi . Then define the λ-distance between P1 and P2 as   1 λ(P1 , P2 ) = min max max |ϕ1 (t) − ϕ2 (t)|, ; (7.5.1) T >0 T t ≤T here (·, ·) is the inner product and  ·  is the Euclidean norm. Clearly, λ metrizes the weak convergence. Our first result concerns the important case where one of the probability measures considered has compact support. Lemma 7.5.1 Let θ1 , . . . , θn be n ≥ 2 unit vectors in IR2 , no two of which are collinear. Let the support of the probability P1 be a subset of the unit disk, and let the probability P2 have the same marginals as P1 in the directions θ1 , . . . , θn . Set(2) $ # n−1 . (7.5.2) s = 2 2 Then   1 2 s+1 λ(P1 , P2 ) ≤ . s!

(7.5.3)

Remark 7.5.2 We can replace the right-hand side of (7.5.3) by C/s, where 1  2  s+1 C is a constant; note that as s → ∞, 2! ∼ e/s. The difference 1  2  s+1 − es is plotted in figures 7.1 and 7.2. s! Proof of Lemma 7.5.1: The λ-metric is invariant under rotations of the coordinate system, so without loss of generality we assume that (a) the directions θj (j = 1, . . . , n) are not parallel to the axis; (b) there exists at least one pair of directions, say θj1 and θj2 , such that θj1 = (a, b), θj2 = (a, −b), where a = 0, b = 0; i.e., the vectors θj1 and θj2 are symmetric about the horizontal axis. (2) Here

and in what follows [r] denotes the integer part of the number r.

45

-0.6

-0.4

-0.2

0.0

7.5 Closeness of Measure on a Finite Number of Directions

-1.2

-1.0

-0.8

FIGURE 7.1. Plot of the difference (2/s!)1/(s+1) − e/s for s = 1, . . . , 100

20

40

60

80

100

0.0

0.001

0

-0.002

-0.001

FIGURE 7.2. Plot of the difference (2/s!)1/(s+1) − e/s for s = 10, . . . , 100

20

40

60

80

100

The law P1 has bounded support, and so, since the marginals on the directions θ1 , . . . , θn of P1 and P2 coincide, then for all j = 1, . . . , n, 

IR2

k

(x, θj ) P1 ( dx) =



(x, θj )k P2 ( dx).

(7.5.4)

IR2

To see that P2 has moments of any order, consider (7.5.4) with j = j1 , j = j2 , and x = (x1 , x2 ). Then 

(x1 a ± x2 b)k (P1 − P2 )( dx) = 0,

IR2



IR2

  (x1 a + x2 b)k + (x1 a − x2 b)k (P1 − P2 )( dx) = 0,

(7.5.5)

and all integrals are finite. If k is even, then (ax1 + bx2 )k + (ax1 − bx2 )k ≥ ak xk1 + bk xk2 , and thus (7.5.5) implies the existence of all moments of P2 of even order.

46

7. Relaxed or Additional Constraints

The next step is to show that all moments of P1 and P2 of order ≤ n − 1 agree. Set  μr,t (Pℓ ) = xr1 xt2 Pℓ ( dx), ℓ = 1, 2. IR2

Then setting θj = (uj , vj ) in (7.5.4) yields k    k ℓ=0



uℓj vjk−ℓ [μℓ,k−ℓ (P1 ) − μℓ,k−ℓ (P2 )] = 0,

j = 1, . . . , n; k ≥ 0. Now, setting zj = vj /uj in the last equation leads to k    k ℓ=0



zjk−ℓ [μℓ,k−ℓ (P1 ) − μℓ,k−ℓ (P2 )] = 0,

(7.5.6)

j = 1, . . . , n. Since no two of the directions θ1 , . . . , θn are collinear, the points z1 , . . . , z2 are distinct. Hence from (7.5.6) we find that the following polynomial of degree k of the variable z, k    k l=0



z k−ℓ [μℓ,k−ℓ (P1 ) − μℓ,k−ℓ (P2 )] ,

(7.5.7)

has n distinct roots z1 , . . . , zn . If n ≥ k + 1, then this is possible only if all coefficients of (7.5.7) are equal to zero, that is, μℓ,k−ℓ (P1 ) = μℓ,k−ℓ (P2 ), ℓ = 0, . . . , k; k = 0, . . . , n − 1. So, for any unit vector t, and k = 0, 1, . . . , n − 1,   k (t, x) P1 ( dx) = (t, x)k P2 ( dx). (7.5.8) IR2

IR2

(t)

Denote by Pℓ the marginal of Pℓ (ℓ = 1, 2) in the direction t, and by ϕℓ (τ ; t)(τ ∈ IR) its characteristic function. By assumption, the support of (t)

P1

is in the segment [−1, 1]. Then (7.5.8) is equivalent to (k)

(k)

ϕ1 (τ ; t)|τ =0 = ϕ2 (τ ; t)|τ =0 ,

k = 0, . . . , n − 1,

(7.5.9)

(k)

where ϕℓ (τ ; t) is the kth derivative of ϕℓ (τ ; t) with respect to τ (ℓ = 1, 2). A Taylor expansion now gives ϕ1 (τ ; t) − ϕ2 (τ ; t) =

(7.5.10)

s−1 (k) (s)  ϕ (0; t) − ϕ (0; t) 1

k=0

2

k!

(k)

τ

k

(s)

τ ; t) − ϕ2 ( τ ; t) s ϕ ( τ + 1 s!

7.5 Closeness of Measure on a Finite Number of Directions

47

for some τ ∈ (0, τ ). From (7.5.9), the first sum on the right-hand side of (7.5.10) is equal to zero. Since s is an even number, (s) |ϕℓ ( τ ; t)|





z

s

(t) Pℓ ( dx)

1 (t) z s P1 ( dz) ≤ 1, =

ℓ = 1, 2.

−1

IR

Thus for all τ ∈ IR,

τs |ϕ1 (τ ; t) − ϕ2 (τ ; t)| ≤ 2 . s! 1

s+1 ;(3) then Choose T = ( s! 2)

  1 2 s+1 . sup |ϕ1 (τ ; t) − ϕ2 (τ ; t)| ≤ s! |τ |≤T 2 Corollary 7.5.3 Let θ1 , . . . , θn , n ≥ 2, be directions in IR2 no two of which are collinear. Suppose that the marginals of the probabilities P1 and P2 with respect to the directions θ1 , . . . , θn have moments up to the even order k ≤ n − 1. Then the marginals of P1 and P2 with respect to any direction t have the same moments up to order k. Corollary 7.5.4 Lemma 7.5.1 still holds if we replace the assumption that P1 and P2 have coinciding marginals with respect to the directions θj (j = 1, . . . , n) with the assumption that these marginals have the same moments up to order n − 1. To prove our main result we must relax the condition that the support of P1 is compact, assuming only the existence of all moments together with Carleman’s conditions for the definiteness of the moments problem. Set  μk = sup (x, θ)k P1 ( dx), k = 0, 1, . . . , θ∈S 1 IR2

where S 1 is the unit circle, and let (s−1)/2

βs =





1

μ2j2j ,

j=1

where the number s is determined in Lemma 7.5.1; see (7.5.2). (3) This

s

choice of T is optimal, since 2 Ts! =

1 T

; see the definition (7.5.1) of λ-metric.

48

7. Relaxed or Additional Constraints

Theorem 7.5.5 Let θ1 , . . . , θn be n ≥ 2 directions in IR2 , no two of which are collinear. Suppose that the measure P1 has moments of any order. Suppose also that the marginals of the measures P1 and P2 in the directions θ1 , . . . , θn have the same moments up to order n − 1. Then there exists an absolute constant C such that(4) − 41

λ(P1 , P2 ) ≤ Cβs

(μ0 +



1/4

μ2 )

.

Proof: Let t be an arbitrary vector of the unit circle. From Corollary 7.5.3 (t) (t) we have that the marginals P1 and P2 have the same moments up to order s. From KKR (1988, p. 180) and Klebanov and Mkrtchian (1980), it follows that ⎞ ⎛ −1/4 (s−1)/2    1/4  % (t) (t) −1/(2j) ⎠ ⎝ ≤ C μ0 (t) + μ2 (t) , λ P1 , P2 μ2j (t) j=1

where μk (t) =

∞

−∞

(t)

uk Pi ( du), k = 0, . . . , s, i = 1, 2. The theorem now

follows from the obvious inequality μ2j (t) ≤ μ2j

(j = 0, 1, . . . , s/2).

2

Let us now consider the situation where the marginals of P1 and P2 in the directions θ1 , . . . , θn are not the same but are close in the metric λ. Theorem 7.5.6 Let θ1 , . . . , θn , n ≥ 2, be directions in IR2 , no two of which are collinear. Suppose that the supports of the measures P1 and P2 are in the unit disk, and that P1 and P2 have ε-coinciding marginals with respect to the directions θj (j = 1, . . . , n); i.e.,   (θj ) (θj ) λ P1 , P2 (7.5.11)   ≤ ε. := min max max |ϕ1 (τ ; θj ) − ϕ2 (τ ; θj )|, 1/T T >0

|τ |≤T

Then there exists a constant C depending on the directions θj (j = 1, . . . , n) such that for sufficiently small ε > 0, we have     1 + 1/s , (7.5.12) λ(P1 , P2 ) ≤ C 1/ ln ε   . where s = 2 n−1 2 (4) That

is, C is independent of s, P1 , and P2 .

7.5 Closeness of Measure on a Finite Number of Directions

49

Proof: Set ψj (τ ) := ϕ1 (τ ; θj ) − ϕ2 (τ ; θj ), j = 1, . . . , n. For 0 < ε ≤ 1 we have sup|τ |≤1 |ψj (τ )| ≤ ε, cf. (7.5.11). Since the supports of the measures (θj )

(θj )

are subsets of [−1, 1], for any even number k ≥ 2 we have   (k) (k) |ϕ1 (0; θj )| + |ϕ2 (0; θj )| 2 (k) sup |ψj (τ )| ≤ ≤ . (7.5.13) k k! |τ |≤1 P1

and P2

Now we apply Corollary 1.5.1 in KKR (1988), which states that there exist constants Cℓk such that (ℓ)

sup |ϕj (τ )|

|τ |≤1

≤ Cℓk



(7.5.14)

sup |ϕj (τ )|

|τ |≤1

  k−ℓ k

(k)

 k1

sup |ϕj (τ )| ,

|τ |≤1

ℓ = 0, 1, . . . , k.

Choosing k ≥ 2s, ℓ ≤ s, and applying (7.5.13), we obtain (ℓ)

sup |ϕj (τ )| ≤ Cs ε1/2 ,

ℓ = 0, 1, . . . , s; j = 1, . . . , n,

|τ |≤1

where Cs is a new constant depending on s only. In particular, (ℓ)

(ℓ)

|ϕ1 (0; θj ) − ϕ2 (0; θj )| ≤ Cs ε1/2 ,

ℓ = 0, 1, . . . , s; j = 1, . . . , n,

or equivalently,        (x, θj )k (P1 − P2 )( dx) ≤ Cs ε1/2 ,    2 

(7.5.15)

IR

k = 0, 1, . . . , s; j = 1, . . . , n. Following the notation in Lemma 7.5.1, we can rewrite (7.5.15) in the form for k = 0, . . . , s and j = 1, . . . , n,   k     k ℓ k−ℓ   uj vj [μℓ,k−ℓ (P1 ) − μℓ,k−ℓ (P2 )] ≤ Cs ε1/2 .    ℓ ℓ=0

Thus, setting

Rkj =

k    k ℓ=0



zjk−ℓ [μℓ,k−ℓ (P1 ) − μℓ,k−ℓ (P2 )] ,

(7.5.16)

k = 2, . . . , s; j = 1, . . . , n; zj = vj /uj , we obtain  1/2 , |Rkj | ≤ Cε

(7.5.17)

 depends on the directions θ1 , . . . , θn only. For any fixed k (k = where C 2, . . . , s) consider

50

7. Relaxed or Additional Constraints (k)

(i) the matrix Ak with elements aℓj = (ii) the vector Bk with elements (k)

bℓ



k ℓ−1



k−(ℓ−1)

zj

= μℓ−1,k−ℓ+1 (P1 ) − μℓ−1,k−ℓ+1 (P1 ),

, ℓ, j = 1, . . . , k+1;

ℓ = 1, . . . , k + 1;

(iii) the vector Dk with elements dj = Rkj , j = 1, . . . , k + 1. Then (7.5.16) has the form Ak Bk = Dk (k = 1, . . . , s − 1), while (7.5.17)  1/2 . The matrices Ak are invertible, and so yields Dk  ≤ Cε 1/2 Bk  ≤ A−1 , k  Dk  ≤ Cε

(7.5.18)

where the constant C depends on the directions θ1 , . . . , θn only. Inequality (7.5.18) shows that the first s − 1 moments of the two-dimensional distributions are close when ε > 0 is sufficiently small. Such an evaluation of closeness holds for the first s − 1 moments of the marginals corresponding to an arbitrary direction t; i.e.,       k  (x, t) (P1 − P2 )( dx) ≤ Cε1/2 ,     2 IR

k = 0, . . . , s − 1. Now we have

  s−1 (j)  (j)    2 ϕ (0; t) − ϕ (0; t) j 1 2 |ϕ1 (τ ; t) − ϕ2 (τ ; t) ≤  τ  + |τ |s j!  j=0  s! ≤

s−1  Cε1/2 j=0

j!

|τ |j +

2|τ |s 2|τ |s ≤ Cε1/2 e|τ | + . s! s!

&  1 '  1/2 1/2   s!  s−1 Choose T = min ln 1 + Cε , 2 .

Since t is arbitrary on the unit circle, we obtain   1   s−1 2 1/2 1/4 λ(P1 , P2 ) ≤ max C ε + Cε1/2 + , 1/T s! ≤ C [1/ ln(1/ε) + 1/s] , which proves the theorem.

2

Remark 7.5.7 The statement in Theorem 7.5.6 still holds if instead of the ε-coincidence of the marginals as in (7.5.11), we require the ε-coincidence of the moments up to order s of these marginals.

7.5 Closeness of Measure on a Finite Number of Directions

51

Theorems 7.5.5 and 7.5.6 can be generalized for probability measures defined on IRm . However, we cannot choose the directions θ1 , . . . , θn in an arbitrary way. Furthermore, to obtain the same order of precision in IRm , m > 2, corresponding to the n directions in IR2 , we need nm−1 directions. The results can be obtained by induction on the dimension m. We define next the set of directions we are going to use. Choose n ≥ 2 distinct real numbers u1 , . . . , un , all different from zero, and first construct the set of n two-dimensional vectors (1, u1 ), (1, u2 ), . . . , (1, un ). Then construct n2 three-dimensional vectors (1, uj1 , uj2 ), j1 , j2 = 1, . . . , n. Repeating this process, by the last step we have constructed a set of m-dimensional vectors (1, uj1 , uj2 , . . . , ujm−1 ),

jℓ = 1, . . . , n; ℓ = 1, . . . , m − 1.

(7.5.19)

Denote these m-dimensional vectors by θ1 , . . . θN , where N = nm−1 . These inductive arguments lead to the following extensions of Theorems 7.5.5 and 7.5.6. Theorem 7.5.8 The results in Theorems 7.5.5 and 7.5.6 still hold if we consider the measures P1 and P2 in IRm , and we choose as directions the N = nm−1 vectors in (7.5.19). Further, s = 2[(n − 1)/2]. To prove this, it is sufficient to note that instead of the m-dimensional vectors, we can first consider a pair of one-dimensional probabilities; the first component is the distribution of the inner product of the projections of the vector x and the vector θj upon the (m − 1)-dimensional subspace, while the second is the law of the last coordinate of the vector x. This allows us to decrease the dimensionality by one. To complete the proof it is sufficient to apply inductive arguments. The bounds of the deviation between probability measures with coinciding marginals offers a solution to the computer tomography paradox as stated in Gutman et al. (1991): “It implies that for any human object and corresponding projection data there exist many different reconstructions, in particular, a reconstruction consisting only of bone and air (density 1 or 0), but still having the same projection data as the original object. Related nonuniqueness results are familar in tomography and are usually ignored because CT machines seem to produce useful images. It is likely that the ‘explanation’ of this apparent paradox is that point reconstruction in tomography is impossible.” Lemma 7.5.1 shows that although the densities of the probability measures P1 and P2 (given that such densities exist) can be quite distant from each other for any “large” number of coinciding marginals, yet the measures P1 and P2 themselves are close in the weak metric λ. Khalfin and Klebanov (1990) have analyzed this paradox and obtained some bounds for the closeness of probability measures with coinciding

52

7. Relaxed or Additional Constraints

marginals for specially chosen directions for the case of uniform distance between the smoothed densities of these measures. In tomography the observations are, in fact, integrals of body densities along some straight lines. Using quadratic formulas enables us to evaluate the moments of a set of marginals; these in turn make it possible to apply the results in this section (see Remark 7.5.7) to evaluate the precision of the reconstruction for densities. The classical theory of moments makes it possible to give numerical methods for reconstructing the probability measures using the moments (see, for example, Ahiezer (1961)).

7.6 Moment Problems with Applications to Characterization of Stochastic Processes, Queueing Theory, and Rounding Problems The theory of moment has a long history, which originated in the pioneering works of Shohat and Tamarkin (1943), Hoeffding (1955), Rogosinsky (1958), Ahiezer and Krein (1962), Karlin and Studden (1966), Kemperman (1968). It was also in the 1950s and ’60s that moment theory became a separate mathematical discipline. Currently, it is appropriate to talk about the moment problems as beeing a whole range of problems with applications to many mathematical theories. We refer to the monograph of Annastassiou (1993) for a recent survey on the developments in the theory of moments. In this section we present some applications of moment theory to probabilistic-statistical models. The results presented here are due to Anastassiou and Rachev (1992). For the proofs of the theorems, which are only stated but not proved in this section, we refer to Anastassiou (1993). First we shall state results on the following five moment problems: Moment problem 1: Find  sup |x − y|p μ( dx, dy), S ⊂ IR2 , p ≥ 1, (7.6.1) μ

S

and inf μ



|x − y|p μ( dx, dy),

S ⊂ IR2 ,

(7.6.2)

S

where the supremum (resp. infimum) is taken over the set of all probability measures μ with support in S having fixed marginal moments   i x μ( dx, dy) = αi , y i μ( dx, dy) = βi , i = 1, 2, . . . , n. (7.6.3) S

S

7.6 Moment Problems of Stochastic Processes and Rounding Problems

53

Remark 7.6.1 Problem (7.6.2) with fixed marginal distributions μ(· × S) = μ1 (·),

μ(S × ·) = μ2 (·)

(7.6.4)

is indeed the Lp -Kantorovich problem on mass transportation (see Chapters 2 and 3). Moment problem 2: For given x0 ∈ IR and positive α find the Kantorovich radius sup E|X − x0 |α ,

(7.6.5)

where the supremum is over all random variables X with fixed moments EX = p and EX 2 = q. Remark 7.6.2 Problem 2 will be used in approximation of complex queueing models by means of deterministic models. Moment problem 3: Find  sup [t]c μ( dt), A = [0, a] or [0, ∞),

(7.6.6)

A

and inf



[t]c μ( dt),

A = [0, a] or [0, ∞),

(7.6.7)

A

over the set of all probability measures with support A having fixed rth moment  tr μ( dt) = dr , r > 0, dr > 0, (7.6.8) A

where for a given nonnegative x the c-rounding (0 ≤ c ≤ 1) of x is defined by ⎧ ⎨ m if m ≤ x ≤ m + c, [x]c = ⎩ m + 1 if m + c < x ≤ m + 1.

Remark 7.6.3 Moment problem 3 can be applied to the problems of rounding and apportionment; see Mosteller, Youtz, and Zahn (1967), Diaconis and Freedman (1979), Balinski and Young (1982), and Balinski and Rachev (1993). In the apportionment theory, c = 0 corresponds to the Adams method; c = 1/2 corresponds to the Webster method (or conventional rounding, or Mosteller–Youtz–Zahn “broken stick” rule of rounding); c = 1 corresponds to the Jefferson method; see Balinski and Young (1982).

54

7. Relaxed or Additional Constraints

Moment problem 4: Find (7.6.6) and (7.6.7) subject to (7.6.8) and  tμ( dt) = d1 . (7.6.9) A

We next consider some infinite-dimensional analogues of the moment problems 1 and 2. Let C[0, 1] be the space of continuous functions on [0, 1] with the usual sup-norm x, and let X (C[0, 1]) be the space of r.v.s on a nonatomic probability space (Ω, A, P ) with values in C[0, 1]. Let M be the class of all strictly increasing continuous functions f : [0, ∞] → [0, ∞], f (0) = 0, f (∞) = ∞. Finally, let T be a set of finitely many points in [0, 1], 0 ≤ t1 < t2 < · · · ≤ tN ≤ 1.

(7.6.10)

Moment problem 5: Given h, gi ∈ M(i = 1, . . . , N ) find inf Eh(X − Y ),

(7.6.11)

where the infimum is over the set of all possible joint distributions of X and Y subject to the moment constraints Egi (|X(ti )|) = ai ,

Egi (|Y (ti )|) = bi .

(7.6.12)

Remark 7.6.4 This problem can be interpreted as follows. Having observations of two random processes (more precisely, we suppose the moments (7.6.12) are known), the goal is to evaluate the minimal possible distance Eh(||X −Y ||) between the processes X and Y . We shall determine the minimum in (7.6.11) and show that essentially this minimum can be achieved.

7.6.1

Moment Problems and Kantorovich Radius

In this section we state the solutions of moment problems 1 and 2; the proofs are given in Anastassiou and Rachev (1992) and the monograph Anastassiou (1993). Let S = [a, b] × [c, d] ⊂ IR2 and ϕ(x1 , x2 ) = |x1 − x2 |p , p ≥ 1. Suppose (α, β) ∈ S, and denote by U = U (ϕ, α, β) the supremum in (7.6.1) subject to   x1 μ( dx1 , dx2 ) = α, x2 μ( dx1 , dx2 ) = β. (7.6.13) S

S

Theorem 7.6.5 The supremum U in (7.6.1) is given by U = Dδ + T,

(7.6.14)

7.6 Moment Problems of Stochastic Processes and Rounding Problems

55

where 

   |b − d|p + |a − c|p − |b − c|p + |a − d|p ,

D

:=

T

(1 − B)|b − c|p + (B + C − 1)|a − c|p + (1 − C)|a − d|p , d−β b−α , C := , := b−a d−c := max(0, 1 − B − C).

B δ

:=

Remark 7.6.6 Since ϕ is convex on any S ⊂ IR2 , then given (7.6.13),  inf ϕ dμ = |α − β|p . μ

S

Next, consider nonbounded regions S. Namely, define for b ≥ 0 the following stripes in IR2 : S1b S2b Sb

:= {(x, y); y = x + b , where 0 ≤ b ≤ b}, := {(x, y); y = x − b , where 0 ≤ b ≤ b}, := S1b ∪ S2b .

We extend Theorem 7.6.5 to this type of unbounded region. Theorem 7.6.7 Assume that 0 < p ≤ 1. (i) If S = S1b or S2b , (α, β) ∈ S, then the supremum U in (7.6.1) is equal to U := U (ϕ, α, β) = |α − β|p . (ii) If S = S b , then U = bp . (iii) Let L be the lower bound (7.6.2) subject to (7.6.13). Then if S = S1b or S = S2b or S = S b , (α, β) ∈ S, we have L := L(ϕ; α, β) = bp−1 |α − β|. Next, consider another type of stripe in IR2 : For b, γ > 0, S1b S2γ S b,γ

:= {(x, y); y = x + b , where 0 ≤ b ≤ b},

:= {(x, y); y = x − γ  , where 0 ≤ γ  ≤ γ}, := S1b ∪ S2γ .

Theorem 7.6.8 Let p ≥ 1. (i) If S = S1b , (α, β) ∈ S, then U := U (ϕ, α, β) = bp−1 (β − α).

56

7. Relaxed or Additional Constraints

(ii) If S = S2γ , (α, β) ∈ S, then U = U (ϕ, α, β) = γ p−1 (α − β). (iii) If S = S b,γ , (α, β) ∈ S, then U =

(bp − γ p )(β − α − b) + bp (b + γ) . b+γ

Next, we shall state the explicit solutions of Moment problem 2. So we will be interested in the following problem. For given x0 ∈ IR, α > 0, p ∈ IR, q > 0 (p2 ≤ q), −∞ ≤ a < b ≤ +∞, find the Kantorovich radius K

:= K(x0 ; α, p, q, a, b) :=

α

(7.6.15) 2

sup{E|X − x0 | ; X ∈ [a, b] a.s., EX = p, EX = q}.

Theorem 7.6.9 (Case (A): α ≥ 2, −∞ < a < b < +∞) Let x0 = (a + b)/2, a ≤ p ≤ b, and 0 ≤ q ≤ b2 + (a + b)(p − b). Then the Kantorovich radius K admits the following bound:  α−2 # $ b−a (a + b)2 K ≤ q − p(a + b) + . 2 4 Moreover, if there exist λ1 , λ2 ≥ 0, λ1 + λ2 ≤ 1, such that p

=

q

=

a+b b−a + (λ1 − λ2 ), 2 2 b2 − a2 (a + b)2 (b − a)2 + (λ1 − λ2 ) + (λ1 + λ2 ), 4 2 4

then K =



b−a 2

α−2 # $ (a + b)2 q − p(a + b) + . 4

The next theorem gives an analogue of Theorem 7.6.9 when the X’s in (7.6.15) have unbounded support. Theorem 7.6.10 (Case (B): 0 < α ≤ 2, a = −∞, b = +∞) For any x0 ∈ IR, p ∈ IR, q > 0, p2 ≤ q, the Kantorovich radius K is given by K = K(x0 ; α; p, q) = (q − 2x0 p + x20 )α/2 . The rest of the results in this section treat various versions of Theorems 7.6.9 and 7.6.10. Theorem 7.6.11 (Case (C): 0 < α ≤ 2, −∞ < a < b < +∞) For any x0 ∈ (a, b); p ∈ IR, p2 ≤ q, K = sup{E|X − x0 |α ; X ∈ [a, b] a.s., EX = p, EX 2 = q}. Set P = p − x0 , Q = q − 2x0 p + x20 , A(x0 ) = a − x0 , B(x0 ) = b − x0 , C(x0 ) = min(−A(x0 ), B(x0 )).

7.6 Moment Problems of Stochastic Processes and Rounding Problems

57

(i) If 0 ≤ Q ≤ C 2 (x0 ), then K = Qα/2 . (ii) If Q > C 2 (x0 ) and (A(x0 ) + B(x0 ))P − Q − A(x0 )B(x0 ) ≥ 0, then K ≤ Qα/2 . Theorem 7.6.12 (Case (D): 1 ≤ α ≤ 2, −∞ < a < b ≤ +∞) For any p ∈ IR, p2 ≤ q, a ≤ p ≤ b, set P = p − a, Q = q − 2ap + a2 , B = b − a. Suppose Q ≤ BP . Then K := sup{E|X − a|α ; X ∈ [a, b], EX = p, EX 2 = q} = p2−α Qα−1 . Theorem 7.6.13 (Case (E): 1 ≤ α ≤ 2, −∞ ≤ a < b < +∞) For any p ∈ IR, p2 ≤ q, a ≤ p ≤ b, set P = p − b, Q = q − 2bp + b2 , θ = a − b. Suppose Q ≤ θP . Then K := sup{E|X − b|α ; X ∈ [a, b], EX = p, EX 2 = q} = p2−α Qα−1 .

7.6.2

Moment Problems Related to Rounding Proportions

Here we state results on explicit solutions of Moment see (7.6.6) and (7.6.9). To this end recall the definition c ≤ 1): For any x ≥ 0, ⎧ ⎨ m, if m ≤ x < m + c, [x]c = ⎩ m + 1, if m + c ≤ x < m + 1, ⎧ ⎨ 0, if 0 ≤ x < c, = ⎩ m, if m − 1 + c < x ≤ m + c,

problems 3 and 4; of c-rounding (0 ≤

m ∈ IN ∪ {0}, m ∈ IN.

The next four theorems deal with problem 3.(5)

Theorem 7.6.14 Let c ∈ (0, 1), r > 0, 0 < a < ∞, d > 0, and U := U[·]c (a, r, d) := sup{E[X]c ; 0 ≤ X ≤ a a.s., (EX r )1/r = d}. Set n = [a]. (I) If n + c < a, n + c ≤ d ≤ a, then U = n + 1. (II) If n + c ≥ a, n − 1 + c ≤ d ≤ a, then U = n. (5) In

the sequel, the underlying probability space is assumed to be nonatomic, and thus the space of laws of nonnegative r.v.s coincides with the space of all Borel probability measures on IR+ .

58

7. Relaxed or Additional Constraints

(III) If 0 < a ≤ c, then U = 0. ln 2 (< 1), n + c < a, and 0 ≤ d ≤ n + c, then ln(1 + 1/c) U = (n + 1)dr (n + c)−r .

(IV) If 0 < r ≤

ln 2 , n + c ≥ a, and 0 ≤ d ≤ n − 1 + c, then ln(1 + 1/c) U = ndr (n − 1 + c)−r .

(V) If 0 < r ≤

(VI) If r ≥ 1 and 0 ≤ d ≤ c, then U = dr c−r . (VII) Suppose r ≥ 1. If either (a) n + c < a and k ∈ {1, . . . , n} is determined by k − 1 + c ≤ d < k + c, or (b) n + c ≥ a and k ∈ {1, . . . , n} is determined by k − 1 + c ≤ d < k + c, dr − (k − 1 + c)r then U = k + ≤ 1 − c + d. (k − c)r − (k − 1 + c)r The next theorem extends Theorem 7.6.14 to the case a = +∞. Theorem 7.6.15 Let 0 < c < 1, r > 0, d > 0, and U := U[·]c (r, d) := sup{E[X]c ; X ≥ 0 a.s., (EX r )1/r = d}. (I) If 0 < r < 1, then U = +∞. (II) If r ≥ 1 and 0 ≤ d ≤ c, then U = dr c−r . (III) Suppose r ≥ 1. Define k ∈ N by k − 1 + c ≤ d < k + c. Then U = k+

dr − (k − 1 + c)r ≤ 1 − c + d. (k + c)r − (k −+ c)r

The next two theorems are versions of the previous two; here we consider the lower bounds in c-rounding. Theorem 7.6.16 Let c ∈ (0, 1), r > 0, 0 < a < ∞, d > 0, and L := 1/r L[·]c (a, r, d) := inf{E[X]c ; 0 ≤ X ≤ a a.s., (EX r ) = d}. Set n = [a].

(I) If 0 < d ≤ c, then L = 0. (II) If c < a ≤ 1 + c and c ≤ d ≤ a, then L = (dr − cr )/(ar − cr ).

7.6 Moment Problems of Stochastic Processes and Rounding Problems

59

(III) Let 0 < r ≤ 1, n + c < a, and determine k ∈ {0, 1, . . . , n − 1} by k + c ≤ d < k + 1 + c. Then dr − (k + c)r . L = k+ (k + 1 + c)r − (k + c)r (IV) If 0 < r ≤ 1, n + c < a, and n + c ≤ d ≤ a, then L = n+

dr − (n + c)r . ar − (n + c)r

The case a = +∞ is extended as follows. Theorem 7.6.17 Let c ∈ (0, 1), r > 0, d > 0, and L := L[·]c (r, d) := inf{E[X]c : X ≥ 0 a.s., (EX r )1/r = d}. (I) If r > 1, then L = 0. (II) If 0 ≤ r ≤ 1, 0 < d ≤ c, then L = 0. (III) If r = 1, c ≤ d < ∞, then L = d − c. (IV) If 0 < r ≤ 1, define k ∈ IN ∪ {0} by k + c ≤ d < k + 1 + c. Then dr − (k + c)r L := k + . (k + 1 + c)r − (k + c)r Next we pass to Moment problem 4, see (7.6.6)–(7.6.9). For simplicity we shall consider only special cases of c-rounding. For the general case we refer to Anastassiou and Rachev (1992) and Anastassiou (1993). First consider the conventional (Webster) rounding, or MYZ-rounding, [x] := [x]1 . Theorem 7.6.18 Let a > 0, 0 < r = 1, d1 > 0, dr > 0, and U = U[·] (a, r, d1 , d2 ) := sup{E[X]; 0 ≤ X ≤ a a.s. and EX = d1 , EX r = dr }. Let θ = [a].

(I) Set Δr,a := ar +

ar − θr (d1 − θ). a−θ

Suppose that a = θ, θ ≤ d1 ≤ a, and either r > 1, dr1 ≤ dr ≤ Δr,a , or 0 < r < 1 and Δr,a ≤ dr ≤ dr1 . Then U = θ.

60

7. Relaxed or Additional Constraints

(II) Suppose 0 < θ = a and there are λ1 , λ2 ≥ 0 with λ1 + λ2 ≤ 1 and such that d1 = λ1 θ + λ2 a and dr = λ1 θr + λ2 ar . Then U =

(θr − ar )d1 + (a − θ)dr . a(θr−1 − ar−1 )

(III) Let θ ≥ 1 and suppose there exists k ∈ {0, 1, . . . , θ − 1} such that k ≤ d1 < k+1 and either r > 1 and dr,k := k r +[(k+1)r −k r ](d1 −k) ≤ dr ≤ θr−1 d1 or 0 < r < 1 and θr−1 d1 ≤ dr ≤ dr,k . Then U = d1 . For the case a = +∞ we have the following version of the above theorem. Theorem 7.6.19 Let 0 < r = 1, d1 > 0, dr > 0, and U := U[·] (r, d1 , dr ) := sup{E[X]; X ≥ 0 a.s. and EX = d1 , EX r = dr }.

Suppose there exists a nonnegative integer k such that k ≤ d1 ≤ k + 1 and either r > 1 and dr ≥ dr,k := k r + [(k + 1)r − k r ](d1 − k) or 0 < r < 1 and 0 < dr ≤ dr,k . Then U = d1 .

If we change in Theorems 7.6.18 and 7.6.19 the upper bound U to the corresponding lower bound, we obtain the following two theorems. Theorem 7.6.20 Let a > 0, 0 < r = 1, d1 > 0, dr > 0, and L := L[·] (a, r, d1 , dr ) := inf{E[X]; 0 ≤ X ≤ a a.s., EX = d1 , EX r = dr }. (I) Suppose there exist t1 , t2 , λ such that 0 ≤ t1 ≤ t2 ≤ 1, and d1 = (1 − λ)t1 + λt2 and dr = (1 − λ)tr1 + λtr2 . Then L = 0. (II) If 0 < a ≤ 1, then L = 0. (III) If 1 < a < 2 and there exist λ1 , λ2 > 0 with λ1 + λ2 ≤ 1 and such (dr − d1 ) that d1 = λ1 + λ2 a, dr = λ1 + λ2 ar , then L = . (ar − a) From now on assume that a ≥ 2, and let θ = [a]. (IV) Suppose Δθ :=

(θ − 1)(ar − a) ≤ θ. (θr − θ)

(i) If d1 = λ1 + λ2 θ, dr = λ1 + λ2 θr for some λ1 , λ2 ≥ 0, λ1 + λ2 ≤ 1, then L = Δθ . (ii) If d1 = λ1 θ+λ2 a and dr = λ1 θr +λ2 θr for some λ1 , λ2 ≥ 0, λ1 +λ2 ≤ 1, then L =

(θ − a(θ − 1))dr − (θr+1 − ar (θ − 1))d1 . θar − aθr

7.6 Moment Problems of Stochastic Processes and Rounding Problems

61

(V) Suppose Δθ > θ. (i) If d1 = λ1 +λ2 θ+λ3 a and dr = λ1 +λ2 θr +λ3 ar for some λ1 , λ2 , λ3 ≥ 0, λ1 + λ2 + λ3 = 1, then L=

(θ − 1)(θ − a + 1)(dr −1) − ((θr −1)θ − (ar −1)(θ − 1))(d1 − 1) . (θ − 1)(ar − 1) − (a − 1)(θr − 1)

(ii) If d1 = λ1 + λ2 a, dr = λ1 + λ2 ar for some λ1 , λ2 ≥ 0, λ1 + λ2 ≤ 1, θ(dr − d1 ) . then L = ar − a (VI) Suppose θ > 1 and one of the following holds. (i) r > 1 and there exists k ∈ {1, . . . , θ − 1} such that k ≤ d1 k + 1 and dr,k

:= k r + ((k + 1)r − k r ) (d1 − k) ≤ dr (θr − 1)(d1 − 1) ; ≤ Δr,θ := 1 + θ−1

(ii) 0 < r < 1 and there exists k ∈ {1, . . . , θ − 1} such that k ≤ d1 ≤ k + 1 and Δr,k ≤ dr ≤ dr,k . Then L = d1 − 1. The special case a = +∞ is treated as follows. Theorem 7.6.21 Let 0 < r = 1, d1 > 0, dr > 0, and L := L[·] (r, d1 , dr ) := inf{E[X] : X ≥ 0 a.s., EX = d1 , EX r = dr }. (I) Suppose there exist 0 ≤ t1 ≤ t2 ≤ 1, 0 ≤ λ ≤ 1 such that d1 = (1 − λ)t1 + λt2 and dr = (1 − λ)tr1 + λtr2 . Then L = 0. (II) Suppose 0 < r < 1 and either d1 = λ1 + λ2 , dr = λ1 , λ1 ≥ 0, λ1 + λ2 ≤ 1 or d1 ≥ 1, 0 < d1 ≤ 1. Then L = d1 − dr . (III) Suppose 0 < r < 1 and for some integer k, k ≤ d1 < k + 1, and 1 ≤ dr ≤ (k r + ((k + 1)r − k r ))(d1 − k). Then L = d1 − 1. (IV) Suppose r > 1 and either d1 = λ1 , dr = λ1 + λ2 , λi ≥ 0, i = 1, 2, λ1 + λ2 ≤ 1 or 0 < d1 ≤ 1 and 1 ≤ dr . Then L = 0. (V) Suppose r > 1 and for some integer k, k ≤ d1 < k + 1, and (k r + ((k + 1)r − k r ))(d1 − k) ≤ dr . Then L = d1 − 1. Similar results are valid for Adams (c = 0) and Jefferson (c = 1) rules of roundings; see Anastassiou and Rachev (1992) and Anastassiou (1993).

62

7. Relaxed or Additional Constraints

7.6.3

Closeness of Random Processes with Fixed Moment Characteristics

The moment problems we are going to consider in this section may be viewed as extensions of Moment problem 1 (on page 52) for measures μ generated by the joint distribution of random processes. Namely, let the class M, the space X (C[0, 1]), and the set T = {t1 , . . . , tN } be defined by (7.6.10). The subject of this section is the following general version of problem (7.6.10)–(7.6.12). Moment problem 6: Given h, gi,j ∈ M (i = 1, . . . , N, j = 1, . . . , n) find the set of valued Eh(||X − Y ||) subject to the moment constraints Egi,j (|X(ti )|) = ai,j ,

Egi,j (|Y (ti )|) = bi,j ,

i = 1, . . . , N.

(7.6.16)

In particular, determine the bounds inf Eh(||X − Y ||), sup Eh(||X − Y ||),

(7.6.17)

given the constraints (7.6.16). One interpretation of the problem is as follows: Suppose we observe two continuous processes X and Y , only through the “window” T . Suppose for each point of the “window” we know some moment characteristics of X and Y . The problem is to determine the possible deviations between the processes outside the window. In particular, “What is the minimal distance between X and Y with given moment information (7.6.16)?” is just a special case of Moment problem 7.6.3. We start with the case n = 1 in (7.6.16); i.e., given h, gi ∈ M(i = 1, . . . , N ) and assuming that Egi (|X(ti )|) = ai ,

Egi (|Y (ti )|) = bi ,

i = 1, . . . , N,

(7.6.18)

we are interested in the range of Eh(||X − Y ||). The solution of this problem will be given under some assumptions of the following type: Assumption A(h, g): h ◦ g −1 (t) (t ≥ 0) is a convex function (here and in the sequel, f −1 stands for the inverse of f ∈ M). Assumption B(g): g −1 (Eg(|ξ + η|)) ≤ g −1 (Eg(|ξ|)) + g −1 (E(|η|)) for any ξ, η ∈ X (R) (here, X (R) is the set of all real-valued r.v.s.). Assumption C(g): Eg(|ξ + η|) ≤ Eg(|ξ|) + Eg(|η|) for any ξ, η ∈ X (R). Assumption D(h, g): limt→∞ h(t)/g(t) = 0.

7.6 Moment Problems of Stochastic Processes and Rounding Problems

63

Remark 7.6.22 Take the most interesting case: h(t) = tp , g(t) = tq (p > 0, q > 0). Then A(h, g) ⇔

p ≥ q,

B(g) ⇔ q ≥ 1, C(g) ⇔ q ≤ 1, D(h, g) ⇔ q > p. Now let T = {0 ≤ t1 ≤ · · · ≤ tN ≤ 1} and take a = (a1 , . . . , aN ) ∈ = (b1 , . . . , bN ) ∈ IRN + and g = (g1 , . . . , gN ) ∈ M to be fixed vectors. Denote by X (T, g, a), the space of all X ∈ X (C[0, 1]) satisfying the marginal moment conditions Egi (|X(ti )|) = ai (i = 1, . . . , N ), and let IRN +,b

I{h, g, T, a, b} :=

inf{Eh(||X − Y ||); X ∈ X (T, g, a), Y ∈ X (T, g, b)}.

(7.6.19)

In the next four theorems we describe the exact range of values of Eh(||X − Y ||) under different conditions of type A–D. Theorem 7.6.23 Let A(h, gi ) and B(gi ) hold for any i = 1, 2, . . . , N . Then   (7.6.20) (i) I{h, g, T, a, b} = sup h |gi−1 (ai ) − gi−1 (bi )| ; 1≤i≤N

(ii) for any ν ≥ I{h, g, T, a, b} there exist random processes Xν ∈ X (T, g, a) and Yν ∈ X (T, g, b) such that Eh(||Xν − Yν ||) = ν.

(7.6.21)

Proof: We shall split the proof into three claims. Claim 1: I{h, g, T, a, b} ≥ sup φi (ai , bi ), where φi (ai , bi ) := h(|gi−1 (ai )− 1≤i≤N

gi−1 (bi )|). Proof of Claim 1: Let X, Y ∈ X (C[0, 1]) and ξ := g(|X(ti ) − Y (ti )|). Then, by the Jensen’s inequality and A(h, gi ), h−1 (Eh(||X − Y ||)) ≥ h−1 (Eh(|X(ti ) − Y (ti )|)) = h−1 (Eh ◦ gi−1 (ξ)) ≥ h−1 ◦ h ◦ gi−1 E(ξ) = gi−1 E(ξ). Further, by B(gi ), h ◦ gi−1 (E(ξ)) ≥ h



   gi−1 (Egi (|X(ti )|) − gi−1 (Egi (|Y (ti )|)

= h(|gi−1 (ai ) − gi−1 (bi )|),

64

7. Relaxed or Additional Constraints

which proves the claim. Claim 2: The infimum in the left-hand side of (7.6.19) is attained, and (7.6.20) holds.  Y ∈ X (C[0, 1]) to be random polygonal Proof of Claim 2: Define X, lines with vertices at points 0, t1 , . . . , tn , 1 given by ⎧ ⎪  i , ω) = g −1 (ai ), Y (ti , ω) = g −1 (bi ), i = 1, . . . , N, ⎪ X(t ⎪ ⎨  ω) = Y (0, ω) = 0, (7.6.22) X(0, if t1 > 0, ⎪ ⎪ ⎪ ⎩ X(1,  ω) = Y (1, ω) = 0 if tN < 1, ω ∈ Ω. For any ω ∈ Ω, ( ( ( (  (X(·, ω) − Y (·, ω)( = sup |gi−1 (ai ) − gi−1 (bi )|, 1≤i≤N

and hence

 − Y ||) = Eh(||X

sup φi (ai , bi ).

(7.6.23)

1≤i≤N

Further, by (7.6.22),

 ∈ X (T, g, a), X

Y ∈ X (T, g, b).

(7.6.24)

Invoking (7.6.23), (7.6.24), and Claim 1, we complete the proof of the claim. Claim 3: (ii) is satisfied. Proof of Claim 3: Let τ ∈ (0, 1), t ∈ T . Define the r.v.s Xν and Yν in X (C[0, 1]) as follows:  Xν (t) = X(t),

Yν (t) = Y (t)

for t = 0, t1 , . . . , tN , 1,

 Y are determined by (7.6.22), and where X, Xν (τ ) = h−1 (ν),

Yν (τ ) = 0.

(7.6.25)

(7.6.26)

Next, let Xν (t), Yν (t) be a random polygonal lines with vertices at 0, t1 , . . . , tN , 1 and τ . Making use of Claim 2, we have ||Xν − Yν || = h−1 (ν) ≥ and thus (7.6.21) holds.

sup |gi−1 (ai ) − gi−1 (bi )|,

1≤i≤N

2

7.6 Moment Problems of Stochastic Processes and Rounding Problems

65

Theorem 7.6.24 Let A(h, gi ) and C(gi ) hold for any i = 1, . . . , N . Then (i)

I{h, g, T, a, b} = sup h ◦ gi−1 (|ai − bi |), 1≤i≤N

(ii)

for any ν ≥ I{h, g, T, a, b} there exist Xnν ∈ X (T, g, a) and Ynν ∈ X (T, g, b) such that Eh(||Xnν − Ynν ||) → Y as n → ∞.

Proof: Claim 1: I{h, g, T, a, b} ≥ sup ϕi (ai , bi ), where ϕi (ai , bi ) := h◦gi−1 (|ai − 1≤i≤N

bi |).

Proof of Claim 1: Let X, Y ∈ X (C[0, 1]). Then, as in Claim 1 of Theorem 7.6.23, by A(h, gi ), C(gi ), and Jensen’s inequality, we have h−1 (Eh(||X − Y ||)) ≥ gi−1 E[gi (|X(ti ) − Y (ti )|)] ≥ gi−1 (|Egi (|X(ti )|) − Egi (|Y (ti )|)|) = h−1 ◦ ϕi (ai , bi ),

which proves the claim. Claim 2: For any ε > 0 there exists a pair (Xε , Yε ) ∈ X (T, g, a)×X (T, g, b) such that Eh(||Xε − Yε ||) = sup h ◦ gi−1 (|ai − bi | + ε) sup 1≤i≤N

1≤i≤N

|ai − bi | . (7.6.27) |ai − bi | + ε

Proof of Claim 2: Without loss of generality we can assume that ai > bi , i = 1, . . . , N . Let pi := (ai −bi )/(ai −bi +ε) and qi := 1−pi , i = 1, . . . , N . We rearrange the indices i so that p1 ≤ p2 ≤ · · · ≤ pN . Choose sets Ai ∈ A such that A1 ⊂ · · · ⊂ AN and P (Ai ) = pi by using the assumption of (Ω, A, P ) being a nonatomic space. More precisely, since (Ω, A, P ) is nonatomic, then for any B ∈ A and any λ ∈ [0, P (B)] there exists C = C(B, λ) ∈ A, C ⊆ B such that P (C) = λ (see Loeve (1977, p. 101)). Then the required sets Ak (k = 1, . . . , N ) are given by Ak = C(Ak+1 , Pk ),

k = 1, . . . , N

(AN +1 := Ω).

Further, for any ω ∈ Ω, define ⎧ ⎨ c := g −1 (a − b + ε), i i i i Xε (ti , ω) := ⎩ d := g −1 (b /q ), i i i i and

⎧ ⎨ 0, Yε (ti , ω) := ⎩ d, i

if ω ∈ Ai , if ω ∈ Ai .

if ω ∈ Ai , if ω ∈ Ai ,

66

7. Relaxed or Additional Constraints

We define (t, Xε (t))t∈[0,1] to be a random polygonal line with vertices (ti , Xε (ti )) and let Xε (0, ω) = 0 if ti > 0 and Xε (1, ω) = 0 if tN < 1 for any ω ∈ Ω. Analogously define the process Yε (t). Then Egi (|Xε (ti )|) = gi (ci )pi + gi (di )qi = ai , and Egi (|Yε (ti )|) = gi (di )qi = bi for any i = 1, . . . , N ; i.e., Xε ∈ X (T, g, a) and Yε ∈ X (T, g, b). Further, Eh(||Xε − Yε ||) = Eh[sup |Xε (t) − Yε (t)|],

(7.6.28)

t∈T

where

⎧ ⎨ c, i |Xε (ti , ω) − Yε (ti , ω)| = ⎩ 0,

if ω ∈ Ai ,

(7.6.29)

if ω ∈ Ai .

Since g ∈ M, then for any i = 1, . . . , N − 1, pi ≤ pi+1 ⇔ ai − bi ≤ ai+1 − bi+1 ⇔ ci ≤ ci+1 ; i.e., c1 ≤ c2 ≤ · · · ≤ cN . Hence, by (7.6.29) and A1 ⊂ A2 ⊂ · · · ⊂ AN , ⎧ ⎨ c , if ω ∈ A , N N sup |Xε (t, ω) − Yε (t, ω)| = (7.6.30) ⎩ 0, t∈T if ω ∈ A . N

Now, (7.6.28) and (7.6.30) imply that Eh(||Xε − Yε ||) = h(cN )pN , which is in fact the right-hand side of equality (7.6.27), and thus the claim is proved. Claims 1 and 2 prove the desired equality (i). Claim 3: (ii) is satisfied. Proof of Claim 3: Let τ ∈ (0, 1), τ ∈ T . Using the same notations as in Claims 1 and 2 we define Xν (ti , ω) Xν (τ, ω) Yν (τ, ω) ν > h(cN )

:= Xε (ti , ω), Yν (ti , ω) := Yε (ti , ω), ⎧ ⎨ h−1 (ν), if ω ∈ A , N = ⎩ 0, if ω ∈ AN , = =

0

ω ∈ Ω,

for any ω ∈ Ω, and ε > 0 is chosen so small that sup h ◦ gi−1 (|ai − bi | + ε).

1≤i≤N

We define the random broken lines Xν and Yν with vertices Xν (ti ), Yν (ti ), i = 1, . . . , N and Xν (0) = Yν (0) if ti > 0; Xν (1) = Yν (1) = 0 if tN < 1. Hence, as in Claim 2 we conclude that ⎧ ⎨ max(c , h−1 (ν)), if ω ∈ A , N N ||Xν (·, ω) − Yν (·, ω)|| = ⎩ 0, if ω ∈ A . N

Hence, Eh(||Xν − Yν ||) = νpN = ν aNaN−b−bNN+ε . This proves the claim.

2

7.6 Moment Problems of Stochastic Processes and Rounding Problems

67

Theorem 7.6.25 Let D(h, gi ) hold for any i = 1, . . . , N . Then I{h, g, T, a, b} = 0,

(7.6.31)

and for any ν > 0 there exist random processes Xnν ∈ X (T, g, a) and Ynν ∈ X (T, g, b) such that Eh(||Xnν − Ynν ||) → ν. Proof: Claim 1: For any n = 1, 2, . . . there exist Xn ∈ X (T, g, a) and Yn ∈ X (T, g, b) such that Eh(||Xn − Yn ||) ≤

n   i=1

ai bi h(nbi ) + h(nai ) gi (nbi ) gi (nai



. (7.6.32)

Since gi ∈ M for n large enough, say n ≥ n0 , we can define disjoint sets Ain , Bin , Cin , and such that Ain + Bin + Cin = Ω and P (Ain ) = cin :=

ai , gi (nai )

P (Bin ) = din :=

bi . gi (nbi )

Now, for any i = 1, . . . , N , n ≥ n0 , define ⎧ ⎪ ⎪ X (t , ω) = nai , Yn (ti , ω) = 0, ⎪ ⎨ n i Xn (ti , ω) = 0, Yn (ti , ω) = nbi , ⎪ ⎪ ⎪ ⎩ X (t , ω) = Y (t , ω) = 0, n i n i

if ω ∈ Ain , if ω ∈ Bin ,

(7.6.33)

if ω ∈ Cin .

Then Egi [|Xn (ti )] = gi (nai )cin = ai and Egi [|Yn (ti )|] = gi (nbi )din = bi ; i.e., X ∈ X (T, g, a) and Y ∈ X (T, g, b). Further, we define the random broken lines Xn (t), Yn (t) (t ∈ [0, 1]) in the way we already did in Theorems 7.6.23 and 7.6.24. Without loss of generality we can assume that a1 ≤ a2 ≤ · · · ≤ aN ≤ b1 ≤ b2 ≤ · · · ≤ bN . Then ||Xn − Yn || =

sup |Xn (t) − Yn (t)|, t∈T

(7.6.34)

68

7. Relaxed or Additional Constraints

=

Hence,

⎧ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩

Eh(||Xn − Yn ||) ≤

nbN

if ω ∈ BN,n ,

nbN −1 .. .

if ω ∈ BN −1,n \ BN,n , if ω ∈ B1,n \

nb1

.. . if ω ∈ A1,n \

na1

if ω ∈

0

N 

N )

j=1

h(nbj,n )dj,n +

j=1

Bj,n ,

j=1

*

N )

j=2

Aj,n ∪

N 

Bj,n ,

j=2 N )

if ω ∈ AN,n \

naN

N )

Aj,n ∪

N )

N )

+

Bj,n ,

j=1

Bj,n .

j=1

h(naj,n )cj,n ,

j=1

which proves (7.6.32) and the claim. By D(h, gi ) (i = 1, . . . , N ) it follows that the right-hand side of (7.6.32) goes to 0 as n → ∞. Hence, (7.6.31) holds true. Claim 2: (ii) is valid. ∈ (0, 1), t ∈ T , and define νn

Let τ

= 1−P

ν ,

N j=1

Cj,n

,

Xn,ν (τ ) = h−1 (νn ), Yn,ν (τ ) = 0. We define the random broken line Xn,ν (t) with vertices Xn,ν (tj ) = Xn (tj ) (see (7.6.33)) and Xn,ν (τ ) (cf. Claim 2 of Theorem 7.6.23). Following the same notations as in Claim 1, we have ⎛

h(nbi )P ⎝Bi,n \

N -

j=i+1



Bj,n ⎠ ≤ h(nbi )din → 0

as n → ∞

(7.6.35)

Bj,n ⎦⎦ ≤ h(nai )cin → 0

(7.6.36)

and ⎡



h(nai )P ⎣Ai,n \ ⎣

N -

j=i+1

Aj,n ∪

N -

j=1

⎤⎤

as n → ∞. Hence, by (7.6.34)–(7.6.36), for n large enough,

7.6 Moment Problems of Stochastic Processes and Rounding Problems

Eh(||XN,ν − YN,ν ||) =

N  i=1

+

max(νn , h(nbi ))P ⎝Bi,n \

N  i=1

=

N  j=1

+





i=1

j=i+1



max(νn , h(nai ))P ⎝Ai,n \ ⎣ ⎛

νn P ⎝Bi,n \

N 

N -



N -

j=i+1



νn P ⎝Ai,n \ ⎣



= νn P ⎝

N -

j=1

Bj,n ⎠ N -

j=i+1

Aj,n ∪





Bj,n ⎠ N -

j=i+1



69

N -

j=1

Aj,n ∪

N -

j=1

⎤⎞

Bj,n ⎦⎠

⎤⎞

Bj,n ⎦⎠

(Aj,n ∪ Bj,n )⎠ .

2

Theorem 7.6.26 For any gi ∈ M (i = 1, . . . , N ), (i)

I{0, g, a, b} (7.6.37) := inf{P (X = Y ); X ∈ X (T, g, a), y ∈ X (T, g, b)} = 0;

(ii) for any ν ∈ (0, 1) there exists a sequence (Xnν , Ynν ) ∈ X (T, g, a)× X (T, g, b) such that P (Xnν = Ynν ) → ν Proof: (i)

as n → ∞.

Let cn ∈ A and P (cn ) =

1 n.

For any i = 1, . . . , N define

⎧ ⎨ X (t , ω) := g −1 (na ), Y (t , ω) := g −1 (nb ), if ω ∈ C , n i i n i i n i i (7.6.38) ⎩ X (t , ω) := Y (t , ω) = 0, if ω ∈ Cn . n i n i

Then (7.6.38) determines the random polygonal lines Xn ∈ X (T, g, a) and Yn ∈ X (T, g, b). Since Xn (ti , ω) = Yn (ti , ω) = 0 whenever ω ∈ Cn and i = 1, . . . , N , then Xn (t, ω) = Yn (t, ω) = 0 if ω ∈ Cn and t ∈ [0, 1]. Hence, P (Xn = Yn ) ≥ P (Ω \ Cn ) =

as desired.

n−1 → 1 n

70

7. Relaxed or Additional Constraints

(ii) Let 0 < ν < 1 and τ ∈ (0, 1) \ T . Choose A ∈ A with P (A) = ν and let ⎧ ⎪ ⎪ X (τ, ω) = 1, Ynν (τ, ω) = 0, for any ω ∈ A, ⎪ ⎨ nν (7.6.39) Xnν (τ, ω) = Ynν (τ, ω) = 0, for any ω ∈ A, ⎪ ⎪ ⎪ ⎩ X (t ) = X (t ), Y (t ) = Y (t ), i = 1, . . . , N, nν

i

n

i



i

n

i

where Xn (ti ) and Yn (ti ) are given by (7.6.38). We construct the random broken lines Xnν and Xnν by using (7.6.39) (cf. Claim 3 of Theorem 7.6.23). From the implications

Xnν (·, ω) = Ynν (·, ω) ⇔



⎧ ⎪ ⎪ X (t , ω) = Ynν (ti , ω), ⎪ ⎨ nν i

i = 1, . . . , N,

⎪ ⎪ ⎪ ⎩ X (τ, ω) = Y (τ, ω) nν nν ω ∈ A ∩ Cn ,

it follows that P (Xnν = Ynν ) = P (Ω \ (A ∪ Cn )) → 1 − ν, which proves (ii), and the theorem as well.

2

As a consequence of Theorems 7.6.23–7.6.26 we obtain the following solution of Moment problem 7.6 for qi (t) = tqi (yi > 0) and h(t) = tp (P ≥ 0, with the convention t0 := I{t = 0}). Corollary 7.6.27 Let q = (q1 , . . . , qN ) (qi > 0), p ≥ 0, and

I{p, q, a, b} :=

inf {E||X − Y ||p ; X, Y ∈ X (C[0, 1]), E|X(ti )|qi = ai , E|Y (ti )|qi = bi , i = 1, . . . , N } .

Then ⎧ p   1/qi 1/qi  ⎪ sup ai − bi  , if p ≥ qi ≥ 1, i = 1, . . . , N, ⎪ ⎪ ⎪ 1≤i≤N ⎪ ⎨ p/q sup |ai − bi | i , if p ≥ qi , 0 < qi < 1, I{p, q, a, b} = (7.6.40) ⎪ 1≤i≤N ⎪ i = 1, . . . , N. ⎪ ⎪ ⎪ ⎩ 0, if 0 ≤ p ≤ qi , i = 1, . . . , N.

7.6 Moment Problems of Stochastic Processes and Rounding Problems

71

Moreover, for any ν > I{p, q, a, b} there exists a sequence (Xnν , Ynν ) ∈ X ([0, 1]) such that E||Xnν − Ynν ||p → ν

as n → ∞

and E|X(ti )|qi = ai ,

E|Y (ti )|qi = bi ,

i = 1, . . . , N.

Remark 7.6.28 Corollary 7.6.27 gives an explicit expression for I{p, q, a, b} if p and qi are subject to certain inequalities (cf. (7.6.40)) or if qi = q for all i = 1, . . . , N . The problem of an explicit description of I{p, q, a, b} for any p ≥ 0 and qi > 0 is still open.

7.6.4

Approximation of Queueing Systems with Prescribed Moments

In this section we discuss applications of Moment problem 1 (on page 52) to the problem of best approximation of a queueing system with known moment characteristics. As an example, suppose our “real” queueing system is of type G|G|1|∞ (for some acquaintance with the usual notations in queueing theory we refer to Borovkov (1984), Kalashnikov and Rachev (1990)). For this system, the sequences of nonnegative r.v.s (possibly dependent and nonidentically distributed) e = {en }n∈IN , s = {sn }n∈IN (IN = (1, 2, . . .)) are viewed as sequences of interarrival and service times. Looking at e and s as “input” of laws, we define (as the “output” flow) the sequence of waiting times w1 = 0,

wn+1 = (wn + sn − en )+ ,

n ∈ IN,

(7.6.41)

where (·)+ = max(0, ·). Since the distribution of w = {wn }n∈IN is not known, the aim is to approximate, model, or simulate the “real” system determined by the triplet (e, s, w) with a “simpler” queueing model (e∗ , s∗ , w∗ ). Assuming that the marginal distributions (the laws of ei , si ) are known, Borovkov (1984, Chapter 4) and Kalashnikov and Rachev (1990) examine different approximating models (e∗ , s∗ , w∗ ) and estimate the possible discrepancy between the “real” system (e, s, w) and the “ideal” model (e∗ , s∗ , w∗ ). Further, we shall relax the constraints “the laws of ei ’s and si ’s are known” by “certain moment characteristics of ei ’s and si ’s are fixed.” In this setup the solutions of Moment problem 1 are used in cases when the “ideal” model is not deterministic, say G|G|1|∞ but with simpler structure. We invoke Moment problem 2 (on page 53) when the approximation model has some deterministic components, like D|G|1|∞ (i.e., e∗j ’s are constants), or D|D|1|∞ (i.e., e∗j ’s and s∗j ’s are constants). Summarizing, we shall consider here the following two problems:

72

7. Relaxed or Additional Constraints

(a) Bounds for the deviation of output characteristics of two dependent queueing models. (b) Approximation of queueing systems by deterministic-type queueing models. Consider the following problem, which occurs in investigations stability of queueing models (see Kalashnikov and Rachev (1990, Chapter 5)). Suppose two queueing models of type G|G|1|∞, (e, s, w) and (e∗ , s∗ , s∗ ), with dependent characteristics are given. Here e = {en }n∈IN , s = {sn }n∈IN , w = {wn }n∈IN are, respectively, the sequences of interarrival, service, and waiting times. Assume that the components dj , sj , j ∈ IN of the “input flows” e and s are dependent and nonidentically distributed. The “output” flow is given by the sequence of waiting times (7.6.41). Suppose that the distribution of ej (resp. sj ) is concentrated on a compact interval [aj , bj ] (resp. [cj , dj ]). While this assumption is quite natural from the practical point of view, it is not used frequently in the literature, simply because it is easier to analyze queueing models with input distributions having unbounded support. We make similar assumptions for the model (e∗ , s∗ , w∗ ); in particular, it is assumed that a∗j ≤ e∗j ≤ b∗j , c∗j ≤ s∗j ≤ d∗j a.s. for all j ∈ IN. The input pairs (ej , e∗j ), (sj , s∗j ) of the two models are arbitrarily mutually dependent, the distributions of ej ’s, e∗j ’s, sj ’s, s∗j ’s are unknown. We assume that only the moments Eej = αj ,

Ee∗j = αj∗ ,

Esj = βj ,

Es∗j = βj∗

(7.6.42)

are given. Our problem to find a sharp bound for the deviation between the waiting times in both models. Let ϕk (en−k,n−1 , sn−k,n−1 ) (en−k,n−1 := (en−k , . . . , en−1 ), sn−k,n−1 := (sn−k , . . . , sn−1 )) be the waiting time for the nth arrival, assuming that the system is “free” at the moment n − k. In other words, ϕ(en−k,n−1 , sn−k,n−1 ) := max [0, en−1 − sn−1 , (en−1 − sn−1 ) + (en−2 − sn−2 ), . . . ,

(7.6.43)

(en−1 − sn−1 ) + (en−2 − sn−2 ) + · · · + (en−k − sn−k )] .

As a measure of deviation between the waiting times of the both systems we shall use δp (T ) = sup max Lp [ϕk (en,n+k−1 , sn,n+k−1 ), ϕk (e∗n,n+k−1 , s∗n,n+k−1 )], n∈IN 1≤k≤T

where p ≥ 1 and T ≥ 2 are fixed, and Lp (X, Y ) := {E|X − Y |p }1/p , p ≥ 1, X, Y ∈ X (R).

(7.6.44)

For random vectors we extend (7.6.44) as follows: Lp (X, Y ) = {E||X − Y ||p }1/p , X, Y ∈ X (RT ),

(7.6.45)

7.6 Moment Problems of Stochastic Processes and Rounding Problems

73

T where ||(x1 , . . . , xT )|| = i=1 |xj |. Since ϕk is a Lipschitz function with respect to the Minkowski norm || · ||, we have that for any k = 1, . . . , T , Lp [ϕk (en,n+k−1 , sn,n+k−1 ), ϕk (e∗n,n+k−1 , s∗n,n+k−1 )]

(7.6.46)

≤ Lp [(en,n+T −1 , sn,n+T −1 ), (e∗n,n+T −1 , sn,n+T −1 )] + Lp [(e∗n,n+T −1 , sn,n+T −1 ), (e∗n,n+T −1 , s∗n,n+T −1 )] ≤

n+T −1 j=n

  Lp (ej , e∗j ) + Lp (sj , s∗j ) .

Now we invoke Theorem 7.6.5 to obtain sharp estimates of Lp (ej , e∗j ) and Lp (sj , s∗j ). Namely, Lp (ej , e∗j ) ≤ (Dj δj + sj )1/p ,

(7.6.47)

where Dj

:= Dj (aj , bj , a∗j , bj ∗)

(7.6.48)

:= |bj − b∗j |p + |aj − a∗j |p − |bj − a∗j |p − |aj − b∗j |p ;

Tj

Bj δj

:= Tj (aj , a∗j , bj , b∗j , αj , αj∗ ) (7.6.49) ∗ p ∗ p := (1 − Bj )|bj − aj | + (Bj + Cj − 1)|aj − aj | + (1 − Cj )|aj − b∗j |p ; :=

bj − αj , bj − aj

Cj

b∗j − αj∗ := ∗ ; bj − a∗j

:= δj (aj , a∗j , bj , b∗j , αj , αj∗ ) := max(0, 1 − Bj − Cj ).

(7.6.50) (7.6.51)

Remark 7.6.29 If ej and e∗j are unknown, i.e., aj = a∗j = 0, bj = b∗j = +∞, then sup{Lp (dj , e∗j ); Eej = αj , Ee∗j = αj∗ } = ∞ (cf. Kuznezova-Sholpo and Rachev (1989)). In a similar way,  j δj + Tj )1/p , Lp (sj , s∗j ) ≤ (D

(7.6.52)

 j , δj , Tj are defined by (7.6.48)–(7.6.51), exchanging bj with dj , where D b∗j with d∗j , aj with cj , and a∗j with c∗j . In this way we have proved the following theorem. Theorem 7.6.30 For any p ≥ 1 and T = 2, 3, . . . ,    j δj + Tj )1/p . δp (T ) ≤ T sup (Dj δj + Tj )1/p + (D j≥1

(7.6.53)

74

7. Relaxed or Additional Constraints

The estimate is sharp or nearly sharp, since the inequalities (7.6.53) and (7.6.52) are the best possible bounds under the moment assumptions (7.6.42) (cf. Theorem 7.6.5), and also, the inequality (7.6.46) cannot be improved in the set of all possible input flows e, e∗ , s, s∗ . Next, we shall consider a much more general case than the single-channel models discussed above. Suppose the dynamics of a queueing system are determined by the transformation F from the set U of input flows U to the set V of output flows V . Let V0 represent the output at moment zero; V0 is assumed to be an ℓ-dimensional vector; i.e., V0 ∈ X (Rℓ ). It is quite general to assume that the input and the output flows have the form U = (V0 , U0 , U1 , . . .) and V = (V0 , V1 , . . .), where Uj ∈ X (Rk ). We endow U and V with the norms ||U ||U :=

∞ 

2−j ||Uj ||k,p + ||V0 ||ℓ,p

(7.6.54)

||V ||V :=

∞ 

2−j ||Vj ||ℓ,p ,

(7.6.55)

j=0

and

j=0

where p ≥ 1, ||Uj ||k,p

:=

(E||Uj ||pk )1/p ,

||Uj ||k

=

||(Uj , . . . , Uj || = |Uj | + · · · + |Uj |,

(1)

(k)

(1)

(k)

and ||Vj ||ℓ,p is defined in a similar way. Suppose the transformation F : U → V is determined by the set of mappings Fj : Rℓ × Rkj → Rℓ ,

j ∈ IN,

(7.6.56)

such that the output at “time” j is defined recursively: Vj = Fj (V0 , U0 , . . . , Uj−1 ).

(7.6.57)

A smoothness assumption on Fj is given by the Lipschitz condition ⎡ ⎤ j−1  (7.6.58) ||βj ||k ⎦ . ||Fj (α0 , β0 , . . . , βj−1 ||ℓ ≤ cj ⎣||α0 ||ℓ + j=0

A reasonably large number of queueing models meet conditions (7.6.56)– (7.6.58). Among them are the single-channel models G|G|1|∞, the multichannel models G|G|J|∞, and the multichannel–multiphased model (G|G|J1 ) → (G|J2 ) → · · · → (G|Jn ) (cf. Kalashnikov and Rachev (1990,

7.6 Moment Problems of Stochastic Processes and Rounding Problems

75

 Chapter 5)). By (7.6.55), (7.6.57), and (7.6.58), ||Vj ||ℓ,p ≤ cj ||V0 ||ℓ,p +  j−1 ||U || , and thus i k,p i=0 ≤ 2cj ||U ||U ⎡ ⎤ ℓ k ∞   (i) (i) ≤ 2cj ⎣ ||V0 ||ℓ,p + 2−1 ||Vj ||ℓ,p ⎦ .

||V ||V

i=1

(7.6.59)

i=1 j=0

Combining (7.6.59) with Theorem 7.6.5 gives us a sharp bound on the deviation of two queueing models V = FU, V ∗ = FU ∗ , whose dynamics are determined by (7.6.54)–(7.6.58). Theorem 7.6.31 Suppose V = FU,

U ∈ U,

V ∈U

(7.6.60)

is a queueing model satisfying (7.6.54)–(7.6.58) such that (i)

a0

(i)

cj

(i)

≤ V0

(i)

≤ Vj

i = 1, . . . , ℓ,

(i)

j = 0, 1, . . . , i = 1, . . . , k.

= L0 ,

(i)

= βj ,

a.s.,

EV0

(i)

a.s.,

EVj

≤ dj

(i)

(i)

(i)

≤ b0

In addition to model (7.6.60) consider the same type model indexed by ∗ and satisfying the above two sets of inequalities with constants indexed by ∗. Then ||V − V ∗ ||V ⎤ ⎡ k ℓ ∞       1/p 1/p (i) (i) (i)  (i) δ(i) + T(i) ⎦, + ≤ 2cj ⎣ 2−j D D0 δ0 + T0 j j j i=1 j=0

i=1

where the D’s, δ’s, and T ’s are determined by the same formula as in (7.6.58)–(7.6.60), and (i) D0 (i)

T0

(i)

δ0

 (i) D j (i) Tj (i) δj



(i) (i) (i)∗ (i)∗ a0 , b0 , a0 , b0



= Di ,   (i)∗ (i) (i) (i) (i)∗ (i)∗ , = Ti a0 , b0 , a0 , b0 , α0 , α0   (i)∗ (i) (i) (i) (i)∗ (i)∗ = δi a0 , b0 , a0 , b0 , α0 , α0 ,   (i) (i) (i)∗ (i)∗ = Dj cj , dj , cj , dj ,   (i) (i) (i)∗ (i)∗ (i) (i)∗ = Tj cj , dj , cj , dj , βj , βj ,   (i) (i) (i)∗ (i)∗ (i) (i)∗ = δj cj , dj , cj , dj , βj , βj .

76

7. Relaxed or Additional Constraints

The rest of this section deals with Problem (b) (on page 72). Suppose again that the “real” queueing system is determined by the triplet (e, s, w), where w is given by the recursive equation (7.6.41). Often in practice one models the random input characteristics by replacing their random values with constants, usually equal to the corresponding means. In doing so, it is natural to investigate the deviation between the “real” output w and the modeled (“ideal”) output w∗ . (In the sequel, all quantities related to the approximating model will have the same notations as in the “real” system but superscribed with ∗.) The deviation between w and w∗ will be expressed by the Kantorovich metric ℓp , defined here as follows: For X, Y ∈ X (IR∞ ), ℓp (X, Y ) := ℓp (P X , P Y )

(7.6.61)

d d  Y ); X,  Y ∈ X (IR∞ ), X = min{Lp (X, X, Y = Y }, p > 0,  q p     where Lp (X, Y ) := E d (X, Y ) , q = min(1, 1/p) is the Lp -metric. In the above definition, the space X (IR∞ ) consists of all random taking ∞sequences ∞ −j values in the metric space (IR , d), where d(x, y) := j=1 2 ||xj − yj ||. Since we have assumed that the underlying probability space is not atomic, the minimum in the right-hand side of (7.6.61) is equal to ⎧ ⎨  d(x, y)P ( dx, dy); P s are probabilities on IR∞ × IR∞ min ⎩ ⎫ IR∞ ×IR∞ ⎬ X Y . with fixed projections P and P ⎭

:=

(n)





(n) (n) = X1 , X2 , . . . For X  (n)  we have ℓp X , X ≥ 2−j ℓp

∈ Xp (IR∞ ), X = (X1 , X2 , . . .) ∈ Xp (IR∞ ),   (n) (n) Xj , Xj , and thus ℓp (X , X) → 0 imd

(n)

plies the weak convergence of any j-component Xj = X and E|Xj |p → E|Xj |p . Further, we consider two types of approximating queues D|G|1|∞ (i.e., e∗j are constants) and D|D|1|∞ (i.e., e∗j and s∗j are constants). Similar results can be obtained if one examines the model G|D|1|∞ (i.e., s∗j are constants) as an approximation of the “real” queue G|G|1|∞. In both queues D|G|1|∞ and G|G|1|∞, the sequences of service times ∗ s and s consist of dependent nonidentically distributed random variables. The next lemma shows that the outputs for the ideal and real models meet a lower bound of deviation if s∗ is chosen to have independent components. Let ε > 0 and X = (X1 , X2 , . . .) ∈ Xp (IR∞ ). The components of X are said to be (ℓp , ε)-independent if IND(X) := ℓp (X, X) ≤ ε,

(7.6.62)

7.6 Moment Problems of Stochastic Processes and Rounding Problems

77

d

where the X i ’s (the components if X) are independent and X i = Xi (i ∈ IN). Lemma 7.6.32 Let the approximating model be of type D|G|1|∞. Assume that the sequences e and s of the queueing model G|G|1|∞ are independent. Then 1 ℓp (w, w∗ ) 2

(7.6.63) ∗

≤ IND(s) + IND(s ) +

∞ 

2−j (ℓp (ej , e∗j ) + ℓp (sj , s∗j )).

j=1

Proof: Using the recursive equations (7.6.41) for w and w∗ , we obtain 1 1 ∗ ∗ ∗ ∗ ∗ ∗ 2 d(w, w ) ≤ d(e, e ) + d(s, s ). Hence, 2 Lp (w, w ) ≤ Lp (e, e ) + Lp (s, s ). Since e and s (resp. e∗ and s∗ ) are independent, we have, passing to the minimal metrics, that 1 ℓp (w, w∗ ) ≤ ℓp (e, e∗ ) + ℓp (s, s∗ ). 2

(7.6.64)

By (7.6.61) and since ej (j ∈ IN) are constants, we obtain the bound ∗

ℓp (e, e ) =

∞ 

−j

2

Lp (ej , e∗j )

=

j=1

∞ 

2−j ℓp (ej , e∗j ).

(7.6.65)

j=1

To estimate ℓp (s, s∗ ) in (7.6.64) we use the (ℓp , ε)-independence characteristic defined in (7.6.62): ℓp (s, s∗ ) ≤ IND(s) + IND(s∗ ) + ℓp (s, s∗ ),

(7.6.66) d

d

where s (resp. s∗ ) has independent components and sj = sj (resp. s∗j = s∗j ). We now invoke the “regularity” property of the Kantorovich metric: 2∞ 3 ∞ ∞      (n) (n) (n) (n) ℓp X , Y ≤ ℓp X , Y (7.6.67) n=1

n=1

n=1

for sequences {X (n) }n≥1 ⊂ Xp (IR∞ ), {Y (n) }n≥1 ⊂ Xp (IR∞ ) of independent components. Let E j be a sequence with components all equal to zero except for the jth component, which equals 1. Then by (7.6.67), ⎞ ⎛ ∞ ∞   (7.6.68) sE j , s∗ E j ⎠ ℓp (s, s∗ ) = ℓp ⎝ j=1



∞  j=1



j=1

j



ℓp sE , s E

j



=

∞  j=1

2−j ℓp (sj , s∗j ).

78

7. Relaxed or Additional Constraints

Combining (7.6.64), (7.6.65), (7.6.66), and (7.6.68) proves the lemma.

2

The estimate (7.6.63) suggests that the approximating model should be chosen with s∗ having independent components. If this is the case, then IND(s∗ ) = 0, and the first problem is to estimate IND(s). Lemma 7.6.33 (a) Suppose that the only information known about the “real” service times are the moments ESjq1 = βj ,

Esqj 2 = γj ,

j ∈ IN,

(7.6.69)

and that the support of Fsj is [0, ∞). Then IND(s) ≤

∞ 

2−j Δj ,

(7.6.70)

j=1

where

Δj :=

⎧  pq ⎪ ⎨ 2β 1/q1 , if 0 < p ≤ q1 , 1 ≤ q1 < q2 , j ⎪ ⎩ +∞,

if 0 < q1 < q2 < p and

1/q βj 1

=

1/q γj 2 ,

(7.6.71)

and q = min(1, 1/p). (b) Suppose the support Fsj is the compact interval [cj , dj ], and βj = Esj . Then (7.6.70) holds with Δj Tj



1/p     j = −2(dj − cj )p , D j δj + T j = , p ≥ 1, where D     d j − βj dj − βj  = 2 1− , and δj = max 0, 1 − 2 . dj − cj dj − cj

(7.6.72)

Proof: Assertion (a) follows from Corollary 2 of Kuznezova-Sholpo and Rachev (1989) and (b) from Theorem 7.6.5. 2

Lemma 7.6.34 Suppose that for every j ∈ IN the first two moments of ej are known: mj := Eej ,

(2)

mj

and let aj ≤ ej ≤ bj a.s.

:= Ee2j ,

σj2 := Var ej ,

(7.6.73)

7.6 Moment Problems of Stochastic Processes and Rounding Problems

79

(i) If p ≥ 2 and −∞ < aj < bj < ∞ and if e∗j is chosen to be the midpoint of [aj , bj ], then ℓp (ej , e∗j ) ≤

*#

bj − aj 2

$p−2 

+  1/p (2) . (7.6.74) mj − mj (aj + bj ) + e∗2 j

(ii) Suppose 0 < p ≤ 2 and either −∞ = aj , +∞ = bj , or −∞ < aj < bj < ∞ and σj ≤ min[mj − aj , bj − mj ].

(7.6.75)

Then the “optimal” d∗j for the approximating model is given by e∗j = mj , and in this case ℓp (ej , e∗j ) ≤ σjpq ,

q = min(1, 1/p).

(7.6.76)

Proof: This follows from Theorems 7.6.10, 7.6.11, and 7.6.12 after some obvious arguments. The estimates (7.6.74) and (7.6.76) are sharp. 2

Lemma 7.6.35 (a) If 0 < p ≤ q1 , 1 ≤ q1 < q2 , then

& ' (2) ∗(2) q1 q2 ∗q1 ∗q2 ∗ ∗ sup ℓp (sj , sj ); nj = Esj , nj = Esj , nj = Esj , nj = Esj pq  ∗1/q1 1/q1 = nj + nj , q = min(1, 1/p). (7.6.77)

(b) Suppose p ≥ 1, cj ≤ sj ≤ dj , c∗j ≤ s∗j ≤ d∗j a.s., and nj = Esj , n∗j = Es∗j . Then ℓp (sj , s∗j )





 j δj + Tj D

1/p

,

(7.6.78)

 j = Dj (cj , dj , c∗ , d∗ ), δj = δj (cj , dj , c∗ , d∗ , nj , n∗ ), where D j j j j j ∗ ∗ ∗  Tj = Tj (cj , dj , cj , dj , nj , nj ) are given by (7.6.48)–(7.6.51).

Proof: Assertion (a) follows from Corollary 2 of Kuznezova-Sholpo and Rachev (1989) and (b) from Theorem 7.6.5. 2 Lemmas 7.6.32–7.6.35 lead us to the main result. Theorem 7.6.36 Let the approximating queueing model be of type D|G|1|∞. Assume that the sequences e and s of the “real” queueing model are independent. Then the Kantorovich metric between the sequences of waiting

80

7. Relaxed or Additional Constraints

times of the “approximating” and “real” models is bounded as follows: ℓp (w, w∗ )

(7.6.79) ∗

≤ 2 IND(s) + 2 IND(s ) +

∞ 

2−j+1 (ℓp (ej , e∗j ) + ℓp (sj , s∗j )).

j=1

Each term in the right-hand side of (7.6.79) can be estimated as follows: (a) An appropriate choice for the approximating sequence of service times will be IND(s∗ ) = 0. (b) If (7.6.69) holds, a bound for IND(s) is given by (7.6.70). (c) If the means and variances of the ej ’s are known, then ℓp (ej , e∗j ) can be estimated from above by (7.6.74). (d) The last term in (7.6.78), ℓp (sj , s∗j ), can be estimated by (7.6.77) (resp. (7.6.78)), provided that the corresponding moment conditions hold. In the next theorem we shall omit the restriction that e and s are independent, but we shall assume that the approximating model is of completely deterministic type D|D|1|∞. Theorem 7.6.37 If the approximation queueing model is of type D|D|1|∞, then ∗

ℓp (w, w ) ≤

∞ 

2−j+1 (ℓp (ej , e∗j ) + ℓp (sj , s∗j )).

(7.6.80)

j=1

If the first moments of ej and sj are fixed, then ℓp (ej , e∗j ) and ℓp (sj , s∗j ) can be estimated as in Lemma 7.6.34. The proof is similar to that of Theorem 7.6.36.

7.6.5

Rounding Random Numbers with Fixed Moments

In this part we shall discuss the interplay between Moment problems 3 and 4 (on pages 53, 54) and the problem of rounding of random proportions. Given a vector X = (X1 , . . . , Xn ) of r.v.s consider the sum X1 +· · ·+Xn . If the Xi ’s are uniformly distributed on the simplex {(si ) ≥ 0; s1 +· · ·+sn = 1}, then they can be treated as proportions, and clearly Sn := X1 + · · · + Xn = 1. If Sn∗ is the sum of conventional roundings [X1 ]1/2 + · · · + [Xn ]1/2 , then Mosteller, Youtz, and Zahn (1967), Diaconis and Freedman (1979), and Balinski and Rachev (1993) have estimated the probability that Sn =

7.6 Moment Problems of Stochastic Processes and Rounding Problems

81

Sn∗ . Here we shall examine the closeness between Sn and Sn∗ in the case of i.i.d. observations Xi where only one or two moments are known. Suppose {Xi }i∈IN are nonnegative i.i.d. r.v.s with known moments EX1 = d1 ,

EX1r = dr .

(7.6.81)

The c-rounding [·]c (c ∈ [0, 1]) (see Section 7.6.2) gives us the sequence of i.i.d. roundings {[Xi ]c }i∈IN . Let Vi := [Xi ]c − Xi be the rounding error, and n Sn,c = i=1 Vi is the total rounding error. Then the normalized rounding error n−1 Sn,c converges by the LLN to E[X1 ]c − d1 . Our objective here is to find sharp bounds for the distribution function of n−1 Sn,c subject to (7.6.81). In other words, for a suitably chosen metric μ in the distribution functions space, the problem is to determine the “radius” of the set of probabilistic laws, i.e.,   Dn = Dn (μ) := sup μ n−1 Sn,c , E[X]c − d1 ; (7.6.82)

EX = d1 , EX r = dr . In (7.6.82), X has the same distribution as the Xi ’s, and thus E[X]c − d1 = d

EV , where V = V1 . Clearly, there is a great variety of metrics μ from which one can choose in (7.6.82). We shall consider two metrics, one especially designed for the problem, one the ideal metric θs (s > 1) and the other the L´evy metric L. Note that Theorems 7.6.18–7.6.21 provide us with sharp bounds for E[X]c − d1 , in the case of the conventional rounding c = 12 .(6) In fact, with [X] = [X]1/2 , sup{E[X] − d1 ; EX = d1 , EX r = dr } inf{E[X] − d1 ; EX = d1 , EX r = dr }

= U − d1 ,

= L − d1 ,

(7.6.83) (7.6.84)

where the exact values of U and L are given in Theorems 7.6.18–7.6.21. Next we can chose μ in the definition of Dn = Dn (μ) to be the L´evy metric L(X, Y ) =

inf{ε > 0; FX (x − ε) − ε ≤ FY (x) ≤ FX (x + ε) + ε for all x ∈ IR},

and thus for the distribution function Fn of n−1 Sn,c we obtain the following bounds: 0 ≤ Fn (x) ≤ Dn

0 1 − Dn

(6) The

≤ Fn (x) ≤ 1 ≤ Fn (x) ≤ 1

for 0 ≤ x ≤ L − Dn ,

for L − Dn ≤ x ≤ U − Dn , for U + Dn ≤ x.

general case of c ≤ 1 was treated in Anastassiou and Rachev (1992).

(7.6.85) (7.6.86) (7.6.87)

82

7. Relaxed or Additional Constraints

From Theorems 7.6.18–7.6.21 it follows that the above bounds are sharp. Our next step is to find a good estimate for Dn = Dn (L). To this end we first estimate Dn (θs ) for θs (X, Y ) = sup |Ef (X) − Ef (Y )|.

(7.6.88)

Here, the supremum is taken over all bounded functions f on IR with q1 . The next integrable second derivative |f  |q ≤ 1, 1 < s < 2, q = 2−s lemma shows that the θs -radius Dn (θs ) = O(n1−s ) for all 1 < s < 2. We use the notation ∨ := max. Lemma 7.6.38 For 1 < s < 2, Dn (θs ) ≤  cn1−s ,

 c :=

2 (c ∨ (1 − c))s . s

Proof: For any X and Y with equal means, θs (X, Y )

≤ :=

1 κs (X, Y ) s d d =  X|  s−1 − Y |Y |s−1 ; X inf{E|X| X, Y = Y }.

Therefore, from the ideality of θs (7) Dn (θs )

1 ≤ n1−s θs (V, EV ) ≤ n1−s κs (V, EV ) s   1 = n1−s E V |V |s−1 − (EV )|EV |s−1  s 1 ≤ n1−s . s

The latter follows since |V | = |X − [x]c | ∈ (0, c ∨ (1 − c)).

2 

− s−1 1+s

In the next theorem we bound Dn = Dn (L) in (7.6.82) as O n



.

Theorem 7.6.39 For any 1 < s < 2, 0 < c < 1, 1

1−s

Dn (L) ≤ (4 c) 1+s n 1+s , where the constant  c is defined as in Lemma 7.6.38.



(7) θ

s is an ideal |c |s θs (Xi , Yi ), for i i

  metric of order s > 0; that is, θs ( i ci Xi , i ci Yi ) all independent Xi , Yi and constants ci ∈ IR.



7.6 Moment Problems of Stochastic Processes and Rounding Problems

83

Proof: The following claim was proved by Grigorevski and Shiganov (1976) for the case s = 2; i.e., in (7.6.88) the functions f have a.e. f  and |f  | ≤ 1 a.e.; see also Maejima and Rachev (1987) and Rachev and R¨ uschendorf (1992). Claim: For any 1 < s < 2, θ(X, Y ) ≥

1 1+s L (X, Y ). 4

Proof of the Claim: Let L(X, Y ) > ε. Then there exists x0 such that either FX (x0 ) > FY (x0 +ε)+ε or FY (x0 ) > FY (x0 +ε)+ε. Say the first inequality takes place. Define ⎧ 1 for x ≤ x0 ; ⎪ ⎪  2 ⎪ ⎪ 2(x − x0 ) ε ⎪ ⎪ for x0 < x ≤ x0 + ; ⎨ 1− ε 2  2 f0 (x) := ⎪ 2(x0 + ε + x) ε ⎪ ⎪ for x ≤ x < x0 + ε; −1 + + ⎪ 0 ⎪ ε 2 ⎪ ⎩ 1 for x ≥ x0 + ε. Observe that |f0 (x)| ≤ 1, f  (x) exists a.e., and

⎤1/q ⎡ x +ε 0 |f0 (x)|q dx⎦ = 8ε−s =: c(ε) ||f0 ||q = ⎣ x0

1 . Recalling the definition of θs , we have 2−s  # $ # $   f (X) (Y ) f 0 0  θs (X, Y ) ≥ E −E c(ε) c(ε)     1   = (f (x) + 1) d [F (x) − F (x)] 0 X Y  c(ε)   x  0 ∞ 1  = (f0 (x) + 1) dFX (x) + (f0 (x) + 1) dFX (x) c(ε) 

for q =

−∞

≥ ≥

x0

  ∞  (f0 (x) + 1) dFY (x) − (f0 (x) + 1) dFY (x) −  x0 +ε −∞ ⎡ x ⎤ 0 ∞ 1 ⎣ (f0 (x) + 1) dFX (x) − (f0 (x) + 1) dFY (x)⎦ c(ε) −∞

x0 +ε

2 [FX (x0 ) − FY (x0 + ε)] c(ε)

x0 +ε

84

7. Relaxed or Additional Constraints

≥ =

2ε c(ε) 1 1+s ε . 4

Letting ε → L(X, Y ) completes the proof of the claim.

Now the desired estimate follows from Lemma 7.6.38 and the claim.

6:26 pm, 3/14/05

2

8 Application of Kantorovich-Type Metrics to Various Probabilistic-Type Limit Theorems

We have discussed already in detail the Kantorovich metric as the solution of mass transportation and mass transshipment problems with a metric cost function; cf. Section 2.5 and Chapter 4. In Chapter 7 we studied generalized transshipment problems, leading to extensions of the Kantorovich metric to encompass a variety of ideal probability metrics. This chapter is devoted to applications of these metrics to the rate of convergence problem in the central limit theorem (CLT) and different summability methods for random vectors. We also discuss applications to the asymptotics of various rounding rules.

8.1 Rate of Convergence in the CLT with Respect to the Kantorovich Metric In this section, we investigate bounds for the rate of convergence in the CLT with respect to the Kantorovich metric for random variables with values in separable Banach spaces. In the first part, the rate in stable limit theorems for sums of i.i.d. random variables is considered. The method of proof is an extension of the Bergstr¨ om convolution method. All assumptions regarding the domain of attraction are given in a metric form. In the second part an extension is given to the martingale case. The proof is based on smoothing properties of suitable conditonal versions of the Kantorovich metric. Smoothing inequalities for the Kantorovich metric will be established, and

86

8. Probabilistic-Type Limit Theorems

the Bergstr¨om convolution method (cf. Zolotarev (1977, 1979, 1983, 1986), Senatov (1980), Sazonov (1981), Rachev and Yukich (1989, 1991), Rachev (1991c)) will be extended to the case of stable limit theorems and at the same time to the Kantorovich metric. All assumptions concerning the domain of attraction and the order of convergence are described in terms of finiteness conditions for certain convolution-type metrics. As a consequence of the results for the Kantorovich metric, one obtains rate of convergence results in stable limit theorems for martingales with respect to the Prohorov metric.(1) We start with the rate of convergence in the i.i.d. case. Consider a separable Banach space (U,  · ) and the space X (U ) of U -valued r.v.s defined on a rich enough probability space. The r.v. ϑ ∈ X (U ) is said to be α-stable (0 < α ≤ 2) if n

−1/α

n 

d

ϑi = ϑ,

(8.1.1)

i=1

where the ϑi ’s are i.i.d. copies of ϑ. We are interested in the rate of convergence of the normalized sum Zn = n

−1/α

n 

Xi

(8.1.2)

i=1

of i.i.d. r.v.s to ϑ with respect to the Kantorovich metric: ℓ1 (X, Y ) :=

sup {|E(f (X) − f (Y ))|; f : U → IR bounded, (8.1.3) |f (x) − f (y)| ≤ x − y} .

Recall from Chapters 2 and 4 that ℓ1 -convergence is equivalent to convergence in distribution and convergence of the moments E ·  (existence assumed); moreover, for U = IR, ℓ1 (X, Y ) = |FX (x) − FY (x)| dx. The Prohorov metric π(X, Y ) :=

inf{ε > 0; P (X ∈ A) ≤ P (Y ∈ Aε ) + ε,

(8.1.4) for all Borel sets A in U }(Aε := {x; |x − A|} < ε)

and the Kantorovich metric satisfy the well-known inequality π 2 ≤ ℓ1 , (1) Some

(8.1.5)

results in the literature are formulated more generally but use bounds involving moments of order ≥ 2 and, therefore, are restricted to the Gaussian case. For some recent literature we refer to Bolthausen (1982), H¨ aussler (1988), Bentkus et al. (1990), and Rackauskas (1990). Our method involves various extensions of an idea in Gudynas (1985) on suitably conditioned versions of probability metrics. The results in this section are based on Rachev and R¨ uschendorf (1994a).

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

87

which is in fact an immediate consequence of the Strassen and Kantorovich theorems. In particular, ℓ1 -convergence rates imply convergence rates for π. Theorem 8.1.1 For any 0 < α < 1, ℓ1 (Zn , ϑ) ≤ n1−1/α ℓ1 (X1 , ϑ).

(8.1.6)

Proof: The result follows from (8.1.1)–(8.1.3) and the contraction properties of ℓ1 ; in fact, if ϑi are i.i.d. copies of ϑ, then   n      −1/α E  Zn − n ϑi  ≤ n1−1/α E|X1 − ϑ1 |. (8.1.7)   i=1

Next, we take in both sides of (8.1.7) the infimum over all joint distributions P X1 ,ϑ1 with fixed marginals P X1 and P ϑ . The result is (8.1.6) as desired. 2 Note that (8.1.6) is a general property for every ideal metric of order 1; see for example Zolotarev (1979). Recall that a probability metric μ is said to be ideal of order r if μ(X + Z, Y + Z) ≤ μ(X, Y )

(8.1.8)

for all r.v.s X, Y, Z such that Z is independent of (X, Y ) and μ(cX, cY ) = |c|r μ(X, Y )

for all c ∈ IR,

(8.1.9)

see Sections 6.3 and 6.4. Consider next the rate of convergence in ℓ1 (Zn , ϑ) → 0

(8.1.10)

for 1 < α ≤ 2. Define the following ideal (smoothing) Kantorovich metric of order r > 1: ℓr (X, Y ) = sup hr−1 ℓ1 (X + hϑ, Y + hϑ),

r > 1,

(8.1.11)

h>0

and σ r (X, Y ) = sup hr σ(X + hϑ, Y + hϑ),

r > 0.

(8.1.12)

h>0

Here ϑ in (8.1.11) and (8.1.12) is assumed to be independent of X and Y , and σ is the total variation metric: σ(X, Y ) = sup{|E(f (X) − f (Y ))|; f : U → [0, 1] continuous} = 2 sup |P (X ∈ A) − P (Y ∈ A)|. A∈B(U )

(8.1.13)

88

8. Probabilistic-Type Limit Theorems

Note that ℓr and σ r are ideal metrics of order r. Throughout this section ℓr stands for the smoothed ℓ1 -metric of order r. The notion ℓp has been used in previous sections for the minimal Lp -metric. So, we have increased the level of “ideality” for ℓ1 and σ (recall that ℓ1 is an ideal metric of order 1, while σ is ideal of order 0) by appropriate smoothing; see (8.1.11) and (8.1.12). The next theorem provides an estimate of the convergence rate in (8.1.10). In what follows C stands for an absolute constant that can be different in different places. Set ℓ1 = ℓ1 (X1 , ϑ), ℓr = ℓr (X1 , ϑ), σ = σ(X1 , ϑ), σ r = σ r (X1 , ϑ). We always assume r > 0. The results in this section are due to Rachev and R¨ uschendorf (1994). Theorem 8.1.2 Suppose that (a)

Eϑ < ∞;

(b)

ℓ1 + ℓr + σ1 + σ r < ∞.

Then 

ℓ1 (Zn , ϑ) ≤ C n

1−r/α

ℓr + τr n

−1/α



,

(8.1.14)

where   . τr = max ℓ1 , σ1 , σ 1/(r−α) r

(8.1.15)

Remark 8.1.3 Zolotarev (1986, §5.4) provides a similar bound for ℓ1 (Zn , ϑ) in the normal univariate case. Zolotarev’s bound contains ζr metrics in the right-hand side of (8.1.14), which can be easily estimated from above in the normal case. In the stable case, however, we need more refined bounds. The problem of finiteness of σ r was discussed in Rachev and Yukich (1989) (see also Section 8.3); for the finiteness of ℓr see the next corollary. Further in this section the sum of any random variables X + Y d d  + Y , where X  and Y are independent and X = means X X, Y = Y . ϑ, and ϑi are defined as in (8.1.1) and satisfy (a). Proof: The proof is similar to that of Theorem 8.1.16, further in this section, which we shall give in detail. Here we give only a short sketch of the proof. It uses the following two properties of the metrics ℓ1 , ℓr , σ r ; see Zolotarev (1986, §5.4). Smoothing Property 1. For any X, Y ∈ X (U ),

ℓ1 (X, Y ) ≤ ℓ1 (X + εϑ, Y + εϑ) + 2εEϑ.

(8.1.16)

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

89

Smoothing Property 2. For any X, Y, Z, W independent, ℓ1 (X + Z, Y + Z) ≤ ℓ1 (Z, W )σ(X, Y ) + ℓ1 (X + W, Y + W ). (8.1.17) Next, let m = [ n2 ]; then by (8.1.16), ℓ1 (Zn , ϑ1 ) ≤ ℓ1 (Zn + εϑ, ϑ1 + εϑ) + Cε   ϑ1 + X 1 + · · · + X n + εϑ ≤ ℓ1 Zn + εϑ, n1/α  m  ϑ1 + · · · + ϑj + Xj+1 + · · · + Xn · ℓ1 + εϑ, 1/α n j=1

 ϑ1 + · · · + ϑj+1 + Xj+2 + · · · + Xn + εϑ n1/α   ϑ1 + · · · + εm+1 + Xm+2 + · · · + Xn + ℓ1 + εϑ, ϑ1 + εϑ n1/α m  = I0 + Ij + Im+1 . j=1

By (8.1.17),    X2 + · · · + Xn ϑ2 + · · · + ϑn −1/α −1/α , σ n X1 + εϑ, n ϑ1 + εϑ I0 ≤ ℓ1 n1/α n1/α   X1 + ϑ2 + · · · + ϑ n ϑ1 + · · · + ϑn + ℓ1 + εϑ, + εϑ . n1/α n1/α 

Similar upper bounds are obtained for Ij , 1 ≤ j ≤ m + 1. Some of the terms obtained in this way can be estimated using the ideality properties of the For example, a term of the following form, Δ = (m +   Xmetrics. 1 +ϑ2 +···+ϑn , ϑ , can be estimated by 1)ℓ1 n1/α Δ =

(m + 1)ℓ1

≤ (m + 1) ≤ (m + 1)





2

n−1/α X1 +

n n−1 n n−1

 r−1 α  r−1 α



n−1 n

1/α

ϑ, n−1/α ϑ1 +

ℓr (n−1/α X1 , n−1/α ϑ) n−r/α ℓr ≤ Cn1−r/α ℓr ,

where in the first inequality we use the obvious relation ℓr (X, Y ) ≥ hr−1 ℓ1 (X + hϑ, Y + hϑ).



n−1 n

1/α 3 ϑ

90

8. Probabilistic-Type Limit Theorems





X1 +···+Xj ϑ1 +···+ϑj we , j 1/α j 1/α Bj ≤ C(ℓr j 1−r/α + τr j −1/α ).

For terms of the form Bj := ℓ1

argument to get the bound see the proof of Theorem 8.1.16.

use an induction For more details 2

Corollary 8.1.4 Suppose that U = IRK and that ϑ has a Fr´echet differentiable density pϑ and let  C(ϑ) = sup |pϑ (y)(z)| dz < ∞. (8.1.18) z ≤1

Suppose that Eϑ < ∞ and ℓ1 + ℓr < ∞. Then ℓ1 (Zn , ϑ) ≤ C(n1−r/α ℓr + τr∗ n−1/α ), where τ1∗ = max(ℓ1 , ℓr

1/(r−α)

(8.1.19)

).

For an integer r, ℓr can be estimated from above by the ζr -metric (see  (r) Zolotarev (1983, p. 294)): ℓr ≤ Cζr if sup z ≤1 |pϑ (y)(z)| dz is finite. We shall discuss the finiteness of ℓr in Section 8.5 in more detail. Proof: Claim 1. For any X and Y ∈ X (IRk ) and δ > 0, σ(X + δϑ, Y + δϑ) ≤ C(r)δ −r ℓr (X, Y ),

(8.1.20)

with C(r) = 2(2−3)/α C(ϑ). To prove the claim we first use the obvious bound σ(X + δϑ, Y + δϑ) ≤ δ −r σ r (X, Y ).

(8.1.21)

Next, we show that for any δ > 0, σ(X + δϑ, Y + δϑ) ≤ δ −1 C(ϑ)ℓ1 (X, Y ).

(8.1.22)

Indeed, by the ideality of σ and ℓ1 it is enough to show (8.1.22) for δ = 1. Then     σ(X + ϑ, Y + ϑ) ≤ sup  f (x)(PX ( dx) − PY ( dx)) , |f |≤1



f (x + y)pϑ (y) dy. Since |f | ≤ 1,  f  (x) = sup |f  (x)(z)| ≤ sup |pϑ (y)(z)| dy =: C(ϑ),

where f (x) =

z ≤1

z ≤1

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

91

and thus |f (x) − f (y)| ≤ C(ϑ)x − y, which obviously implies (8.1.22). To show (8.1.19) we use (8.1.21), (8.1.22), and the following bound: σ r (X, Y )

=

sup hr σ(X + hϑ, Y + hϑ) h>0

≤ sup hr ℓ1 (X + 2−1/α hϑ, Y + 2−1/α hϑ) h>0

21/α C(ϑ) h

= C(ϑ)2r/α ℓr (X, Y ). This completes the proof of the claim as well as that of (8.1.19).

2

Remark 8.1.5 (Rate of convergence in the CLT for random elements with LePage representation) Consider a symmetric α-stable U -valued random variable ϑ with LePage representation d

ϑ =

∞ 

−1/α

Γj

ηj Yj ,

(8.1.23)

j=1

where (i) Yj are i.i.d. with EY1 r < ∞; (ii) ηj are i.i.d. symmetric real-valued random variables with η1 α = (E|η1 |α )1/α < ∞; (iii) (Γj ) is a sequence of successive times of jump of a standard Poisson process; (iv) we assume that the three sequences are independent; see Ledoux and Talagrand (1991, Sect. 5.1) and Samorodnitsky and Taqqu (1994). Suppose X has a similar representation d

X =

∞ 

−1/α ∗ ∗ ηj Yj ,

Γj

(8.1.24)

j=1

where (Yj∗ ) and (ηj∗ ) are chosen as in (i) and (ii) with the only difference that they are not identically distributed. Write Zn , the normalized sum of i.i.d. copies Xi as in (8.1.2). Then Theorem 8.1.2 yields the following rate of convergence of Zn to ϑ in the ℓ1 -metric. Corollary 8.1.6 Let 1 ∨ α < r < 2, and E||Y1 ||r + supj≥1 E||Yj∗ ||r + E||η1 ||r + supj≥1 E||ηj∗ ||r < ∞. Then ℓ1 (Zn , ϑ) ≤ C(n1−r/α ℓ∗r + τr∗ n−1/α ),

(8.1.25)

92

8. Probabilistic-Type Limit Theorems

& where ℓ∗r := supj≥1 (ℓr (Yj∗ , Y1 ) + ℓr (ηj∗ Y1 , η1 Y1 )) and τr∗ = max ℓ∗1 , σ1∗ , ' ∗1/(r−α) σr with σr∗ := supj≥1 (σ r (Yj∗ , Y1 ) + σ r (ηj∗ Y1 , η1 Y1 )). Proof: In view of (8.1.14), (8.1.15) we need only show the finiteness of σ r and ℓr . For σ r = σ r (X, ϑ) we use the ideality of order r and the asymptotics −r/α EΓj ∼ j −r/α (j → ∞) to obtain σ r (X, ϑ)

=



−r/α

EΓj

ℓr (ηj Yj , ηj∗ Yj∗ )

j≥1



≤ ⎝





−r/α ⎠

EΓj

j≥1

sup{E|ηj∗ |r σ r (Yj∗ , Yj ) + σ r (ηj∗ Yj , ηj Yj )} j≥1

≤ C sup(σ r (Yj∗ , Y1 ) + σ r (ηj∗ Y1 , η1 Y1 )). j≥1

The same type estimate is valid for ℓr .

2

Since in the LePage representations Yj , Yj∗ can have any high enough moment, examples with finite ℓ∗r and τr∗ can be readily constructed. Take, d

for example, U to be a Hilbert space with basis (hm )m≥1 , and set Yj∗ =  ∗ d  ∗ ζm hm , where (ζm )m≥1 , (ζj,m )m≥1 are sequences of ζj,m hm , Y1 =

m≥1

m≥1

independent random variables. Then, by the ideality of σ r ,   ∗ ∗ σ r (Yj∗ , Y1 ) ≤ σ r (ζj,m , ζm ) ≤ C κr (ζj,m , ζm ), m≥1

(8.1.26)

m≥1

where κr is the rth pseudomoment,  ∗ κr (ζ , ζ) = r |x|r−1 |Fζ ∗ (x) − Fζ (x)|dx, see Zolotarev (1983). Similarly,  ∗ ℓr (Yj∗ , Y1 ) ≤ C κr (ζj,m , ζm ).

(8.1.27)

m≥1

The same example is valid if we relax the independence assumption to “independence in finite blocks,” requiring only that (ζ1+ℓ , . . . , ζL+ℓ ), ℓ = 0, L, 2L, . . ., are independent. Remark 8.1.7 (Finite-dimensional approximation) An alternative use of the explicit upper bounds of the smoothing metrics in the finite-dimensional case is to combine Theorem 8.1.2 with an approximation step by the finitedimensional case. To be concrete, let X, Y be C(S) valued processes, (S, d)

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

93

a totally bounded metric space. For ε > 0 let Vε denote a finite covering ε-net and let Pε X = (Xt )t∈Vε be the corresponding finite-dimensional approximation of X = (Xt )t∈S . If E sup |Xs − Xt | ≤ a(ε)

(8.1.28)

d(s,t)≤ε

and E sup d(s,t)≤ε

|Ys − Yt | ≤ b(ε),

then ℓ1 (X, Y ) ≤ ℓ1 (Pε X, Pε Y ) + a(ε) + b(ε).

(8.1.29)

So we can combine fluctuation inequalities (8.1.28) with the finite-dimensional bounds derived in (8.1.14) for the normalized sum Zn in order to choose an optimal rate of approximation ε = ε(n) → 0. A general and simple useful tool to derive fluctuation inequalities as in (8.1.28) is Pollard’s lemma, which applied to (8.1.28) yields % 4 Nε max E sup |Xs − Xt | , (8.1.30) E sup |Xs − Xt | ≤ d(s,t)≤ε

1≤i≤Nε

d(s,ti ) α,

ℓ1 (Zn , ϑ) ≤ Br,α τ r n1−r/α , where Br,α ≥ 8(r−1)/α + 2(2r/α + 3r/α ).

(8.1.34)

94

8. Probabilistic-Type Limit Theorems

The proof uses the following analogue of (8.1.17): For any independent X, Y, Z, W , ℓ1 (X + Z, Y + Z) ≤ ℓ1 (X, Y )σ(Z, W ) + ℓ1 (X + W, Y + W ). (8.1.35) The proof is similar to that of the smoothing inequality in Zolotarev (1986, §5.4) (see also Rachev (1991c, Theorem 15.2.2)) and thus is omitted. The theorem is of interest for 1 ≤ α ≤ 2, as for 0 < α < r < 1 we get from (8.1.6), ℓ1 (Zn , ϑ) ≤ n1−1/α τ¯r .

Our next objective is the extension of Theorem 8.1.2 to the martingale case. Let (Ω, A, P ) be a rich enough probability space, (Fi ) an increasing sequence of sub σ-algebras of A, and let (Xi , Fi ) be an adapted martingale difference sequence with values in a separable Banach space (U,  · ); that is, E(Xi |Fi−1 ) = 0 a.s., i ∈ IN. For a given probability metric μ and a sub σ-algebra G ⊂ A define the G-dependence metric μ(·G) by μ(X, Y G) = sup μ(X + V, Y + V ),

(8.1.36)

V ∈G

where V ∈ G denotes that V is a G-measurable random variable. This notion generalizes an idea due to Gudynas (1985). Lemma 8.1.10 If μ is homogeneous of order r, that is, μ(cX, cY ) ≤ |c|r μ(X, Y ),

(8.1.37)

then μ(·G) also is homogeneous of order r. We shall use the following metrics: ℓr (·G), σ r (·G), where ℓr , σ r are respectively the smoothed Kantorovich metric and the total variation metric (cf. (8.1.11), (8.1.12)). Lemma 8.1.11 Let the regular conditional distributions PX|G , PY |G exist. Then ℓr (X, Y G) ≤ Eℓr (PX|G , PY |G )

(8.1.38)

σ r (X, Y G) ≤ Eσ r (PX|G , PY |G ).

(8.1.39)

and

Proof: Let ϑ be independent of X, Y , and G. Then ℓr (X, Y G) = =

sup ℓr (X + V, Y + V )  sup sup sup hr−1 E(f (X + V + hϑ)

V ∈G

f L ≤1 h>0 V ∈G

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

95

 −f (Y + V + hϑ))

 ≤ E sup sup sup hr−1 E(f (X + V + hϑ)|G)

 −E(f (Y + V + hϑ)|G)

f L ≤1 h>0 V ∈G

 = E sup sup sup hr−1 E(fV (X + hϑ)|G)

 −E(fV (Y + hϑ)|G),

f L ≤1 h>0 V ∈G

where fV (·) = f (· + V ) is the translation by V , which is again a Lipschitz (y)| is the Lipschitz norm. We function, and where f L = supx =y |f (x)−f x−y arrive at   ℓr (X, Y G) ≤ E sup sup hr−1 E(f (X + hϑ)|G) − E(f (Y + hϑ)|G) f L ≤1 h>0

= Eℓr (PX|G PY |G ).

The proof for the metric σ r is similar.

2

As a consequence we obtain the following regularity property of ℓr and σr . Lemma 8.1.12 Let (Xi , Fi ) be a stochastic sequence and (Gi ) a decreasing sequence of sub σ-algebras such that Yj are Gi -measurable for j ≥ i. Suppose that the following condition holds: (c)

Xi and Gi+1 are conditionally independent given Fi−1 , and Yi and Gi+1 are conditionally independent given Fi−1 .

Then, for ci ∈ IR, ℓr

2 n 

ci Xi ,

i=1

n 

ci Yi

i=1

3



n 



n 

i=1

|ci |r Eℓr (PXi |Fi−1 , PYi |Fi−1 )

(8.1.40)

and σr

2 n  i=1

ci Xi ,

n  i=1

ci Yi

3

i=1

|ci |r Eσ r (PXi |Fi−1 , PYi |Fi−1 ),

assuming that the conditional distributions exist.

(8.1.41)

96

8. Probabilistic-Type Limit Theorems

Proof: By Lemma 8.1.10, 2 n 3 n   ℓr ci Xi , ci Yi i=1



n 



n 

= ≤

i=1

i=1

ℓr (c1 X1 + · · · + ci Xi + ci+1 Yi+1 + · · · + cn Yn , c1 X1 + · · · + ci−1 Xi−1 + ci Yi + · · · + cn Yn ) sup hr−1

i=1 h>0 n  i=1 n  i=1

sup V ∈Fi−1 ∨Gi+1

ℓ1 (ci Xi + V + hϑ, ci Yi + V + hϑ)

ℓr (ci Xi , ci Yi Fi−1 ∨ Gi+1 ) |ci |r ℓr (X1 , Y1 Fi−1 ∨ Gi+1 ),

where Fi−1 ∨ Gi+1 is the σ-algebra generated by Fi−1 and Gi+1 . From Lemma 8.1.11 and the conditional independence assumption, ℓr (Xi , Yi Fi−1 ∨ Gi+1 ) ≤ Eℓr (PXi |Fi−1 ∨Gi+1 , PYi |Fi−1 ∨Gi+1 ) = Eℓr (PXi |Fi−1 , PYi |Fi−1 ).

As for the metric σ r , the proof is similar.

2

Remark 8.1.13 If Yi are independent of Fi−1 , EYi = 0, then ℓr (PXi |Fi−1 , PYi ) ≤ Cr ζr (PXi |Fi−1 , PYi )

(8.1.42)

≤ Cr κr (PXi |Fi−1 , PYi ),

where ζr is the Zolotarev metric and κr is the pseudo-difference moment (cf. (8.1.26) and Rachev (1991, p. 377)). In the α-stable case 1 < α < 2 and r = 2, the finiteness of κr implies that E(Xi |Fi−1 ) = EYi = 0,

(8.1.43)

which is fulfilled in the martingale case. In the normal case α = 2 and r = 3, the finiteness of ζr implies in the Euclidean case that the conditional covariance Cov (Xi |Fi−1 ) = Cov (Yi )

(8.1.44)

is almost surely constant. This and related conditions have been assumed in several papers on the martingale convergence theorem (cf. Basu (1976), Dvoretzky (1970), Bolthausen (1982), Butzer et al. (1983), H¨ aussler (1988), Rackauskas (1990)).

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

97

Lemma 8.1.14 ℓr (X, Y ) ≥ (ℓ1 (X, Y ))r



r−1 r

r−1

1 2Eϑ. r

(8.1.45)

Proof: By the triangle inequality and from the definition of ℓr , ℓr (X, Y ) ≥ ℓ1 (X + εϑ, Y + εϑ)εr−1

≥ (ℓ1 (X, Y ) − 2εEϑ)εr−1 := ϕ(ε).

Maximizing ϕ(ε) with respect to ε, we obtain (8.1.45).

2

In the next step we extend the smoothing inequality (8.1.17). Lemma 8.1.15 Suppose that X, Z, Y, W are random variables with values in U such that (X, Z) is independent of (Y, W ) and Y, W are independent. Then ℓ1 (X + Z, Y + Z) ≤ ℓ1 (Z, W )σ(X, Y ) + ℓ1 (X + W, Y + W )  + X), + ℓ1 (Z + X, Z

(8.1.46)

σ(X + Z, Y + Z) ≤ σ(Z, W )σ(X, Y ) + σ(X + W, Y + W )  + X). + σ(Z + X, Z

(8.1.47)

d =  is independent of X, and Z and Z where Z

Proof: By the triangle inequality, ℓ1 (X + Z, Y + Z) =

sup |E [(f (X + Z) − f (X + W ))

f L ≤1

− (f (Y + Z) − f (Y + W ))]|

+ sup |E(f (X + W ) − f (Y + W ))|. f L ≤1

Furthermore, |E [f (X + Z) − f (X + W ) − (f (Y + Z) − f (Y + W ))]|   =  (E(f (Z + x)|X = x) − Ef (W + x)) dPX (x)    − (Ef (Z + x) − Ef (W + x)) dPY (x)     ≤  (E(f (Z + x)|X = x) − Ef (Z + x)) dPX (x)   +  Ef (Z + x)( dPX (x) − dPY (x))

98

8. Probabilistic-Type Limit Theorems





  Ef (W + x)( dPX (x) − dPY (x))

 + X) + ℓ1 (Z, W )σ(X, Y ). ≤ ℓ1 (Z + X, Z

The proof of (8.1.47) is similar.

2

The last term in (8.1.46) is a measure of dependence of Z, X, which disappears if Z, X are independent. Making use of the smoothing properties, we next extend Theorem 8.1.2 to the martingale case. Let (Xi , Fi ) be a martingale difference sequence, n 1/α Zn = n j=1 Xj , and as in (8.1.4) let ϑ, ϑi be independent, α-stable distributed r.v.s. For r > α we define ℓr = sup ℓr (Xj , ϑj ), τr = sup Eℓr (PXj |Fj−1 , Pϑj ), ℓr = ℓr ∨ τr , (8.1.48) j

j

τr = sup Eℓr (PXj |Fj−1 , PXj ), τ5r = sup Eℓr (PXj |G5j+1 , PXj ), j

j

where G5j+1 = σ(Xj+1 , Xj+2 , . . .), and σ r = supj σ r (Xj , ϑj ), the conditional distributions, are assumed to exist. Theorem 8.1.16 Suppose that Eϑ < ∞. Then ℓ1 (Zn , ϑ) ≤ C(n1−r/α ℓr + n−1/α tr ),   1 1 r−α r−α where tr = max ℓ1 , σ1 , σ r , τ5r , τ1 .

(8.1.49)

Proof: Applying (8.1.16) we shall estimate ℓ1 (Zn + εϑ, ϑ1 + εϑ). Set m = [ n2 ]. Then 2 3 n  ℓ1 Zn + εϑ, n−1/α ϑi + εϑ ≤

i=1



−1/α

ℓ1 Zn + εϑ, n +

m  j=1



(ϑ1 + X2 + · · · + Xn ) + εϑ



ℓ1 n−1/α (ϑ1 + · · · + ϑj + Xj+1 + · · · + Xn ) + εϑ, −1/α

n ⎛

(ϑ1 + · · · + ϑj+1 + Xj+2 + · · · + Xn ) + εϑ

+ ℓ1 ⎝n−1/α (ϑ1 + · · · + ϑm+1 + Xm+2 + · · · + Xn ) + εϑ,



8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

n−1/α

n  j=1

=: I0 +

m 

99



ϑj + εϑ⎠

Ij + Im+1 .

j=1

From the generalized smoothing inequality (8.1.46),      X1 ϑ1 X2 + · · · + Xn X2 + · · · + Xn I0 = ℓ1 + + εϑ , + + εϑ n1/α n1/α n1/α n1/α     X 2 + · · · + X n ϑ2 + · · · + ϑ n −1/α −1/α ≤ ℓ1 , σ n X + εϑ, n ϑ + εϑ 1 1 n1/α n1/α   X 1 + ϑ2 + · · · + ϑ n ϑ1 + · · · + ϑ n + εϑ, + εϑ + ℓ1 n1/α n1/α 2 3 n X1 + · · · + X n X1 + · · · + Xn−1 + X + ℓ1 + εϑ, + εϑ n1/α n1/α =: Δ1 + Δ2 + Δ3 ,

d n = n is independent of X1 , . . . , Xn−1 , ϑ. Similarly, where X Xn and X

m  j=1

Ij



m  j=1

ℓ1





Xj+2 + · · · + Xn ϑj+2 + · · · + ϑn , n1/α n1/α



 ϑ1 + · · · + ϑj + Xj+1 ϑ1 + · · · + ϑj+1 ·σ + εϑ, + εϑ n1/α n1/α 2 m  ϑ1 + · · · + ϑj + Xj+1 + ϑj+2 + · · · + ϑn + ℓ1 + εϑ, 1/α n j=1 3 n ϑ j=1 j + εϑ n1/α  m  ϑ1 + · · · + ϑj + Xj+1 + · · · + Xn + ℓ1 + εϑ, 1/α n j=1

j+1 + Xj+2 + · · · + Xn ϑ1 + · · · + ϑj + X + εϑ n1/α

3

=: Δ4 + Δ5 + Δ6 .

We first estimate Δ5 . By the ideality of ℓr , 2  1/α m  n−1 Xj+1 Δ5 = ϑ1 + εϑ, + ℓ1 1/α n n j=1

(8.1.50)

100

8. Probabilistic-Type Limit Theorems

3  1/α ϑj+1 n−1 + ϑ + εϑ n n1/α 2   1/α 1/α 3 m  n−1 n−1 Xj+1 ϑj+1 ≤ + ℓ1 ϑ, 1/α + ϑ 1/α n n n n j=1 ≤ Cn1−r/α ℓr .

Similarly, by Lemma 8.1.12, Δ7

=: Im+1 (8.1.51) 2 1/α m+1 Xm+1 + · · · + Xn , ϑ+ ≤ ℓ1 n n1/α 3  1/α m+1 ϑm+1 + · · · + ϑn ϑ+ n n1/α  (1−r)/α n    m+1 −r/α ≤ n Eℓr PXi |Fi−1 , Pϑi n j=m+1 ≤

Cn1−r/α τr ,

and in the same way as for Δ5 , we obtain Δ2 ≤ Cn1−r/α ℓr .

(8.1.52)

The remaining terms are dealt with by induction. Assume next that for j < n,     X1 + · · · + X j ϑ 1 + · · · + ϑ j r j 1−r/α + tr j −1/α , ≤ B (8.1.53) ℓ ℓ1 j 1/α j 1/α and let

ε = A max



1/(r−α) 1/(r−α) σ1 , σ 1/(r−α) , τ5r , ℓr r



n−1/α

(8.1.54)

with a constant A ≥ 0 that we shall fix later in the proof. Then Δ1

≤ BC(n1−r/α ℓr + n−1/α tr )ε−1 n−1/α σ1 (X1 , ϑ) 1 ≤ BC(n1−r/α ℓr + n−1/α tr ). A

In the same way,   1−r/α −1/α Δ4 ≤ CB ℓr (n − m − 2) + tr (n − m − 2) −r/α   ∞   j Xj+1 ϑj+1 α +ε , σr · 1/α n1/α n n j=1

(8.1.55)

(8.1.56)

8.1 Rate of Convergence in the CLT with Respect to Kantorovich Metric

  ≤ CB ℓr n1−r/α + tr n−1/α ·

∞  j=1



σ r (Xj+1 , ϑj+1 )

j+

Aα σ

r

(Xj+1 , ϑj+1 )

α/(r−α)

≤ CB(ℓr n1−r/α + tr n−1/α )/Ar−α ,

101

r/α

using that εα ≥ Aα σ r α/(r−α) n−1 . To estimate Δ3 we apply the G-dependence metric ℓ1 (·Fn−1 ): 3 2  Xn Xn (8.1.57) + εϑ, 1/α + εϑFn−1 Δ3 ≤ ℓ1 1/α n n 2 3 n Xn X ≤ ℓ1 , 1/α Fn−1 ≤ n−1/α Eℓ1 (PXn |Fn−1 , PXn ) 1/α n n ≤ n−1/α τ1 .

Finally, we estimate Δ6 as follows: 2 1/α  m  j Xj+1 Xj+2 + · · · + Xn α Δ6 = +ε ℓ1 + + ϑ, 1/α 1/α n n n j=1 1/α 3  j+1 j X Xj+2 + · · · + Xn ϑ + + + εα 1/α 1/α n n n (1−r)/α m   j ≤ + εα n j=1 3 2  Xj+1 Xj+2 + · · · + Xn Xj+1 Xj+2 + . . . + Xn + , 1/α + · ℓr 1/α 1/α n n n n1/α ≤

m  (j + nεα )(1−r)/α

n(1−r)/α

j=1

≤ τr n

−1/α

≤ Cn−1/α

m  j=1

j+1 G5j+2 ) n−r/α ℓr (Xj+1 , X

(j + nAα τ5rα/(r−α) n−1 )(1−r)/α

1

τ5r1/(r−α) . r−1−α A

Gathering all the inequalities, we obtain  B  1−r/α  −1/α n ℓ1 (Zn , ϑ) ≤ C1 tr + C2 n1−r/α ℓr + C3 n−1/α τ1 ℓr + n A  B  + C4 r−α n1−r/α ℓr + n−1/α tr + C5 n1−r/α ℓr A

102

8. Probabilistic-Type Limit Theorems

+ C6

1 τ5r1/(r−α) n−1/α r−1−α A

+ 2Eϑn

−1/α



max σ1 , σ r

+ C7 n1−r/α τr

1/(r−α)

, ℓr

1/(r−α)

, τ5r1/(r−α)



.

4 ≤ 21 and then choose B Choose A large enough such that CA1 + ACr−α large enough such that C2 + C3 + C5 + C6 A1+α−r + C7 + 2Eϑ ≤ B2 . Thus we obtain (8.1.49). 2

8.2 Application to Stable Limit Theorems Zolotarev (1976) introduced the ζr -metric as an extension of the Kantorovich metric. For any pair of random vectors X, Y on IRk and r = m + a it is defined by ζr (X, Y ) = sup{|E(f (X) − f (Y ))|; f ∈ Fr },

(8.2.1)

where Fr is the class of all functions f : IRk → IR such that f (m) (x) − f (m) (y) ≤ x − ya1 , 0 < a ≤ 1; f (m) denotes the mth Fr´echet derivative of f supplied with the usual supremum norm of multilinear functionals (cf. Zolotarev (1986, Section 6.3) and Rachev (1991, p. 264)), and x − y1 denotes the L1 -norm in IRk . Indeed, ζ1 is merely the Kantorovich metric; see (8.1.3). ζr is ideal of order r and therefore is suitable for analyzing the rate of convergence in various central limit theorems. (The definition of an ideal metric was given in (8.1.8) and (8.1.9).) A disadvantage of ζr is that only for integers r can ζr be estimated by difference pseudomoments from above, while for r ∈ / IN the known upper estimates involve absolute moments br = max(EXr , EY r ) or absolute pseudomoments of order r and therefore are not suitable for approximation by stable distributions of order α < 2. In IR an alternative ideal metric of order r that does not have this drawback of ζr was found by Maejima and Rachev (1987) and applied to prove convergence to self-similar processes; see also Rachev (1991c, Section 17.1). In this section we introduce a new ideal metric ϑs,p (with respect to summation of independent random vectors in IRk ), which generalizes the construction in Maejima and Rachev (1987). This ideal metric has the following properties. It is ideal of order r = s − 1 + p1 . It can be estimated from above by a Zolotarev-type metric and, what is more important, by a pseudo difference moment, which allows applications to stable distributions. Finally, it can be bound from below by the L´evy metric, and thus ϑs,p describes weak convergence of distributions. The degree of ideality of this

8.2 Application to Stable Limit Theorems

103

metric does not depend on the dimension. This is an important property, which is not satisfied by some obvious generalizations of one-dimensional ideal metrics of order greater than 1, see Sections 6.1, 6.3. We shall establish relations between ϑ1,p and ϑs,p and prove various smoothing inequalities. In the second part of this section we give an application to the rate of convergence in stable limit theorems. The upper bounds in the limit theorem are formulated in metric terms. We establish some new results ensuring the finiteness of these bounds and apply these results to show that random vectors in a neighborhood of the LePage decomposition of a stable law satisfy the central limit theorem with rate. Further applications are to the convergence of summability methods of i.i.d. random vectors and to the approximation by compound Poisson distributions. All these applications are based on the thorough analysis of the metric properties of ideal metrics having a structure close to that of the Kantorovich metric. The results in this section are due to Rachev and R¨ uschendorf (1992). We start with the construction of the ϑs,p -metric. Let X, Y ∈ X (IRk ), the class of k-dimensional random vectors, and define for s ∈ IN, 1 ≤ p ≤ ∞, ϑs,p (X, Y ) = sup{|E(f (X) − f (Y ))|; f ∈ Gs,p }.

(8.2.2)

Here Gs,p is the class of functions f : IRk → IR, such that for 1 ≤ i1 ≤ · · · ≤ is ≤ k and for 1 ≤ j ≤ k, x = (x1 , . . . , xj−1 , xj+1 , . . . , xk ) ∈ IRk−1, and q with p1 + 1q = 1,

Dis1 ,...,is f q,j (x) ⎧ ⎛ ⎞1/q ⎪  ⎪ ⎪ ⎪ q ⎨ ⎝ |Ds ⎠ i1 ,...,is f (x1 , . . . , xj , . . . , xk )| dxj = IR ⎪ ⎪ ⎪ ⎪ ess sup |Dis1 ,...,is f (x1 , . . . , xj , . . . , xk )| ⎩ ≤ 1

xj ∈IR

(8.2.3) if q < ∞, if q = ∞

a.s. with respect to the Lebesgue measure.

Lemma 8.2.1 For any 1 ≤ p ≤ ∞ and s ∈ IN, the metric ϑs,p is an ideal metric of order r = s − 1 + p1 . Proof: If f ∈ Gs,p and z ∈ IRk , then f (· + z) ∈ Gs,p and hence ϑs,p (X + Z, Y + Z) ≤ ϑs,p (X, Y ) for any Z independent of X and Y . Further, when q < ∞, for any c ∈ IR, x ∈ IRk−1 , 1 ≤ j ≤ k, fc (x) := f (cx), ⎡ ⎤1/q  Dis1 ,...,is fc q,j (x) = ⎣ |Dis1 ,...,is fc (x1 , . . . , xk )|q dxj ⎦ IR

(8.2.4)

104

8. Probabilistic-Type Limit Theorems

⎡ ⎤1/q  = |c|s−1/q ⎣ |Dis1 ,...,is f (y1 , . . . , yk )|q dyj ⎦ = |c|

r

IR s Di1 ,...,is

f q,j ,

which yields the ideality of ϑs,p of order r. The case q = ∞ can be handled similarly. 2

Remark 8.2.2 Note that the direct generalization of the Maejima–Rachev (1987) construction leads to   1/q  ϑs,p (X, Y ) = sup |E(f (X) − f (Y ))|; f (s) (x)q dx ≤ 1 , (8.2.5)

which is an ideal metric of order s − kq = s − k(1 − p1 ). This unpleasant dependence on the dimensionality is avoided in the definition of ϑs,p by the restriction to one-dimensional integration in (8.2.3). We next show that ϑs,p is estimated from above by the following modification ζ r of the ζr -metric: ζ r (X, Y ) = sup{|E(f (X) − f (Y ))|; f ∈ F r },

(8.2.6)

where F r is the class of functions f : IRk → IR with f

(m)

(x) − f

(m)

(y) ≤ dα (x, y) :=

k  i=1

|xi − yi |a ,

(8.2.7)

m = 0, 1, . . . , a ∈ (0, 1], and r = m + a. In fact, (8.2.7) is equivalent to |f (x) − f (y)| ≤ dα (x, y), ∀x, y ∈ IRk ,

if m = 0,

(8.2.8)

and to sup 1≤i1 ≤···≤im ≤k

|Dim1 ,...,im f (x) − Dim1 ,...,im f (y)| ≤ da (x, y), if m ≥ 1.

Since x − ya1 ≤ da (x, y), we have ζr ≤ ζ r . Lemma 8.2.3 (a)

(8.2.9) For any integer r,

ϑr,1 = ζ r = ζr .

(8.2.10)

8.2 Application to Stable Limit Theorems

105

For any r > 0, if ϑs,p (X, Y ) < ∞, then

(b)

ϑs,p (X, Y ) ≤ ζ r ≤ where r = s − 1 + 

v r (X, Y ) :=

1 p

(8.2.11)

= m + a, and 

da (x, 0)

1≤i1 ,···,im ≤k

IRk

Proof: (a)

Γ(1 + a) v r (X, Y ), Γ(1 + r)

|xi1 · · · xim | |PX − PY |( dx).

(8.2.12)

It is enough to check that Gr,1 = F r . If f ∈ Gr,1 , then

sup 1≤i1 ,...,ir−1 ≤k

f (y)| f (x) − Dir−1 |Dir−1 1 ,...,ir−1 1 ,...,ir−1

 ⎧ y  ⎨  1  r  Di ,...,i ,1 f (t, x2 , . . . , xk ) dt ≤ sup 1 r−1   1≤i1 ,...,ir−1 ≤k ⎩  x1 y   2    r +  Di1 ,...,ir−1 ,2 f (y1 , t, x3 , . . . , xk ) dt   x2 y ⎫  k ⎬   r + · · · +  Di1 ,...,ir−1 ,k f (y1 , . . . , yk−1 , t) dt  ⎭ xk

≤ x − y1 =: d1 (x, y); i.e., f ∈ F r .

Conversely, for f ∈ F r we have from (8.2.8), sup 1≤i1 ,...,ir ≤k

=

|Dir1 ,...,ir f (x)|

sup

lim

1≤i1 ,...,ir ≤k yr →xr

(f (x1 , . . . , xk ) − f (x1 , . . . , yr , . . . , xk ))| |Dir−1 1 ,...,ir−1 |xr − yr |

≤ 1 a.s.; i.e., f ∈ Gr,1 . (b) If f ∈ Gs,p and 1 < p ≤ ∞, then similarly to the proof in (a), sup 1≤i1 ,...,is−1 ≤k



f (y)| f (x) − Dis−1 |Dis−1 1 ,...,is−1 1 ,...,is−1

sup

(8.2.13)

yi

k  

1≤i1 ,...,is−1 ≤k i=1 xi

|Dis1 ,...,is−1 ,i f (y1 , . . . , t, xi+1 , . . . , xk )| dt

106

8. Probabilistic-Type Limit Theorems



k  i=1

Dis1 ,...,is f q,i (y1 , . . . , yi−1 , xi+1 , . . . , xk )|xi − yi |1/p

≤ da (x, y). For the second part of (b) first note that ϑs,p (X, Y ) < ∞ implies that for any 1 ≤ i1 ≤ · · · ≤ ij ≤ k, j ≤ s − 1, E(Xi1 · · · Xij − Yi1 · · · Yij ) = 0.

(8.2.14)

This follows, by taking fc (x) = c xi1 · · · xij , j ≤ s − 1, and the obvious inequality ϑs,p (X, Y ) ≥ supc>0 |E fc (X) − E fc (Y )|. Following the argument in the first part of this proof, (8.2.14) is also a consequence of the condition ζ r (X, Y ) < ∞. We obtain from the Taylor expansion and applying (8.2.14) with m = s − 1 that |E(f (X) − f (Y ))|   1   (1 − t)m−1  =  (m − 1)! 1≤i1 ,...,im ≤k 0

 · E Dim1 ,...,im f (tX)Xi1 · · · Xim









1

1≤i1 ,...,im ≤k 0





1 ⎝ (m − 1)! ·

=

(1 − t)m−1 ⎝ (m − 1)!

1≤i1 ,...,im ≤k 0

·⎝ ≤



1



0

|Dim1 ,...,im f (tx)xi1 · · · xim

IRk



− Dim1 ,...,im f (t0)xi1 · · · xim | |PX − PY |( dx)⎠ dt

(1 − t)m−1 (m − 1)!

IRk

1



    m − Di1 ...im f (tY )Yi1 · · · Yim dt 



da (tx, t0)|xi1 · · · xim | |PX − PY |(dx)⎠ dt ⎞

(1 − t)m−1 ta dt⎠



1≤i1 ,...,im ≤k

Γ(1 + α) v r (X, Y ). Γ(1 + r)



da (x, 0)|xi1 · · · xim | |PX − PY |( dx)

IRk

2

8.2 Application to Stable Limit Theorems

xr1



k i=1

107

a m r  k k |xi | = i=1 |xi | i=1 |xi |

Remark 8.2.4 (a) Since =    k a ≤ 1≤i1 ,...,im ≤k |xi1 . . . xim | , and on the other hand kxr1 ≥ i=1 |xi | k  a |x . . . x | i i 1 m 1≤i1 ,...,im ≤k i=1 |xi | , we have  vr (X, Y ) := xr1 |PX − PY |( dx) ≤ v r (X, Y ) ≤ k vr (X, Y ). (8.2.15) IRk

(b) By the same arguments as in (a) we have ζr ≤ ζ r ≤ k ζ r .

(8.2.16)

In particular, by (8.2.11), ϑs,p is also estimated from above by the Zolotarev metric ζr (up to a constant). The following theorem gives an estimate of ϑs,p in terms of certain pseudomoments, which allows one to apply ϑs,p to stable distributions. For random vectors X, Y with densities uX , uY , define αs,p (X, Y ) =

k 



i=1 1≤i1 ,...,is ≤k

·



IRk−1

⎡ 1   s−1  ⎣ |yi1 . . . yis | t−s−k (1 − t) (uX − uY )  (s − 1)! 0

IR

⎤1/p

 yk   dt |yi |p dyi ⎦ ,..., t t

y

1

dyi . . . dyi−1 dyi+1 . . . dyk .

If k = 1, then after some transformations we obtain αs,p (X, Y )

= =

Fs,X − Fs,Y p (8.2.17) ⎛  x ⎞1/p   s−1  (x − t) ⎝  d(FX − FY )(t)|p dx ⎠ ,  (s − 1)!  −∞

Fs,X (x) :=

1 E(x − X)s−1 + (s − 1)!

(see Maejima and Rachev (1987), Rachev and R¨ uschendorf (1990)). Indeed, αs,p is an ideal metric of order r. Representation (8.2.17) shows that αs,p depends only on the difference pseudomoments. A similar representation holds also for k ≥ 1. Theorem 8.2.5 αs,p is the upper bound for ϑs,p ; i.e., ϑs,p ≤ αs,p .

(8.2.18)

108

8. Probabilistic-Type Limit Theorems

Proof: By (8.2.14) and the Taylor expansion, E(f (X) − f (Y ))   = IRk

1

1≤i1 ≤···≤is ≤k 0

(1 − t)s−1 (s − 1)!

  · Dis1 ,...,is f (tx1 , . . . , txk )xi1 · · · xis dt d(FX − FY )(x) ⎡ 1    (1 − t)s−1 ⎣ = Dis1 ,...,is f (y1 , . . . , yk )yi1 . . . yis (s − 1)! 1≤i1 ,...,is ≤k 0 IRk ⎤ y · (uX − uY ) t−s−k dy ⎦ dt. t

This implies, by making use of H¨older’s inequality, that |E(f (X) − f (Y ))| ≤

k  s−1 



i=1 j=0 1≤i1 ,...,ij ≤k,ij =i



IRk−1

|yi1 . . . yij |



IR

 1      s−1   (1 − t) y −s−k ·  t (uX − uY ) dt (s − 1)! t   0 ⎤



⎣Di1 ,...,ij ,i,...,i f (y)

· |yi |s−j dyi ⎦ dy1 . . . dyi−1 dyi+1 . . . dyk



k  s−1 



i=1 j=0 1≤i1 ,...,ij ≤k,ij =i



|yi1 . . . yij |Di1 ,...,ij ,i,...,i f q,i (y)

IRk−1

( 1 ( ( (   s−1 ( ( y (1 − t) s−j −s−k ( (u ·( | dt − u ) |y t i X Y ( ( (s − 1)! t ( ( 0

p

dy1 . . . dyi−1 dyi+1 . . . dyk ,

which is equivalent to the representation in (8.2.17).

2

For random vectors in IRk we define the L´evy distance by L(X, Y ) =

inf{ε > 0; P (X ∈ Bx ) ≤ P (Y ∈ Bx (ε)) + ε, P (Y ∈ Bx ) ≤ P (X ∈ Bx (ε)) + ε, ∀x ∈ IRk },

(8.2.19)

8.2 Application to Stable Limit Theorems

109

where Bx := {y ∈ IRk ; yi ≤ xi , 1 ≤ i ≤ k} and Bx (ε) := {y ∈ IRk ; y − Bx  ≤ ε},

y =

2 k 

yi2

i=1

31/2 ;

note that P (X ∈ Bx ) = FX (x). L metrizes the topology of weak convergence. If X has a bounded density uX , then ̺(X, Y ) := sup |FX (x) − FY (x)| 2 3 k  ≤ 1+ sup uXi (x) L(X, Y ), i=1

x

(8.2.20) X = (X1 , . . . , Xk ).

Next, we establish that ϑs,p convergence implies weak convergence by providing a lower bound of ϑs,p in terms of L. Theorem 8.2.6 Let s = 1, 2, . . . , p ∈ [1, ∞], r = s − 1 + p1 . Then ϑs,p (X, Y ) ≥ a(s, k)Lr+1 (X, Y ),

(8.2.21)

where a(s, k) :=

Vr 2s+k s!

,

Vr :=



(1 − x2 )r+1 dx.

(8.2.22)

{ x ≤1}

Proof: Let L(X, Y ) > ε. Then without loss of generality we can assume that for some z = (z1 , ..., zk ), P (X ∈ Bz ) − P (Y ∈ Bz (ε)) > ε.

(8.2.23)

We define gr (x) = (1 − x2 )r+1 + ,

x ∈ IRk (a+ = max(0, a)),

(8.2.24)

and “normalize” gr by g r (x) :=

gr (x) . Vr

(8.2.25)

Finally, we define the smoothed version of the indicator of Bz :  &  ε ' ε g r (y) dy I x − y ∈ Bz uε (x) = 2 2 IRk

 k  &   ε '  2 2 = gr I y ∈ Bz (x − y) d y. ε 2 ε IRk

(8.2.26)

110

8. Probabilistic-Type Limit Theorems

Since 0 ≤ uε (x) ≤



IRk

g r (y) dy ≤ 1, we have

|fε | ≤ 1, where fε (x) := 2 uε (x) − 1; and furthermore, ⎧ ⎨ 1 if x ∈ B , z fε (x) = ⎩ 0 if x ∈ / Bz (ε).

(8.2.27)

(8.2.28)

In fact, we have for x ∈ Bz ,  &  ε ' ε uε (x) = g r (y) dy I x − y ∈ Bz 2 2 IRk  & ' ε ε = I x − y ∈ Bz , y ≤ 1 g r (y) dy 2 2 IRk  = g r (y) dy = 1. IRk

Similarly, for x ∈ / Bz (ε), uε (x) = 0.

In the next step we establish bounds on the derivatives of fε . To that purpose let Ls (f ) =

sup ess sup Dis1 ,...,is f q,i (x).

(8.2.29)

2ε−r . a(s, k)

(8.2.30)

1≤i≤k x∈IRk−1

Then Ls (fε ) ≤

To show (8.2.30), observe that Dis1 ,...,is fε (x) (8.2.31)   k+s  &   ε ' 2 2 (x − y) d y = 2 I y ∈ Bz Dis1 ,...,is g s ε 2 ε IRk

 s  &  ε ' 2 = 2 Dis1 ,...,is g s (y) dy. I x − εy ∈ Bz ε 2 IRk

By Minkowski’s inequality, we get the following bound for the norm of the above quantity: Dis1 ,...,is fε q,1 (x)

(8.2.32)

8.2 Application to Stable Limit Theorems



= ⎣



IR

111

⎤1/q

|Dis1 ,...,is fε (x1 , x2 , . . . , xn )|q dx1 ⎦

⎧  s ⎨  2 = 2 ε ⎩

 s 2 = 2 ε

  & '     I x − εy ∈ Bz ε , y ≤ 1  2  IR IRk ⎫1/q ⎬ q s ·I{Bz (ε)\Bz }Di1 ,...,is g s (y) dy| dx1 ⎭

⎫1/q ⎬ ε s dx1 dy I{x − εy ∈ Bz |Di1 ,...,is g s (y)| ⎭ ⎩ 2



⎧ ⎨

IR

{ y ≤1}∩(Bz (ε)\Bz )

 s 2 ≤ 2 ε



|Dis1 ,...,is g s (y)|

{ y ≤1}∩(Bz (ε)\Bz )

·

⎧ ⎨ ⎩

I{z1 ≤ x1 − εy1 ≤ z1 + ε} dx1

IR

⎫1/q ⎬ ⎭

dy.

In fact, the inequality is valid a.e. with respect to Lebesgue measure λ\k−1 . The last integrals are estimated as follows: ⎧ ⎫1/q ⎬ ⎨ = ε1/q , (8.2.33) I{z1 ≤ x1 − εy1 ≤ z1 + ε} dx1 ⎩ ⎭ IR

and



|Dis1 ,...,is g s (y)| dy

(8.2.34)

{ y ≤1}∩(Bz (ε)\Bz )

=

=

1 Vr 1 Vr



{ y ≤1}



  3s−1 2   n    s−1 2yis  dy yi2 I{Bz (ε)\Bz } Di1 ,...,is−1 s 1 −   i=1 + I{Bz (ε)\Bz }s!2s |yi1 · · · yis | dy

{ y ≤1}



1 s!2s Vr

1 1 · · · |yi1 · · · yis | dy ≤

−1

−1

1 1 s!2s · 2k−s = s!2k . Vr Vr

Similarly, we can argue for any index 1 ≤ i ≤ k, and thus (8.2.30) follows from (8.2.33) and (8.2.34).

112

8. Probabilistic-Type Limit Theorems

From the inequality in (8.2.30) we obtain that given fε (x) , Ls (fε )

f ∗ (x) :=

ϑs,p (X, Y ) ≥ E(f ∗ (X) − f ∗ (Y )) εr ≥ a(s, k)E(fε (X) − fε (Y )). 2

(8.2.35)

Applying (8.2.27), (8.2.28) we arrive at the following decomposition: 

E(fε (X) − fε (Y )) =

(fε (x) + 1)(PX − PY )( dx)

IRk



⎜ = ⎝



+



+



Bz Bz (ε)\Bz IRk \Bz (ε)

=: I1 + I2 + I3 ,



⎟ ⎠(fε (x) + 1)(PX − PY )( dx)

where I1

=



(fε (x) + 1)(PX − PY )( dx) = 2(PX − PY )(Bz );

Bz

I2 I3

≥ −2 PY (Bz (ε)\Bz ); = 0.

Thus by (8.2.23), I1 + I2 + I3 ≥ 2(PX (Bz ) − PY (Bz )) − 2(PY (Bz (ε)) − PY (Bz )) ≥ 2ε. From (8.2.35) we finally obtain ϑs,p (X, Y ) ≥ εr+1 a(s, k). With ε → L(X, Y ), this implies (8.2.21). 2

Remark 8.2.7 (a)

Let us use the polar transformation

x1 x2 x3 .. .

= ̺ cos ϑ1 = ̺ cos ϑ1 = ̺ cos ϑ1

xk

= ̺ sin ϑ1 ,

· · · cos ϑk−2 cos ϑk−1 , · · · cos ϑk−2 sin ϑk−1 , · · · sin ϑk−2 ,

(8.2.36)

8.2 Application to Stable Limit Theorems

113

where ̺ > 0, 0 ≤ ϑ1 ≤ 2π, 0 ≤ ϑj ≤ π, 2 ≤ j ≤ k − 1, and ∂(x1 , . . . , xk ) ∂(̺, ϑ1 , . . . , ϑk−1 )

Dk (ϑ)

̺k−1 Dk (ϑ), ⎛ sin ϑ1 ⎜ sin ϑ1 ⎜ ⎜ := det ⎜ sin ϑ1 ⎜ .. ⎝ . =

cos ϑ1

Then we have Vr

= Dk

1

(1 − ̺2 )r+1 ̺k−1 d̺

1

(1 − ̺2 )r+1 (̺2 )

0

= Dk

0

k−2 2

· · · sin ϑk−1 · · · sin ϑk−2 · · · cos ϑk−2 0

···

(8.2.37) ⎞

sin ϑk−1 cos ϑk−1 0 0

⎟ ⎟ ⎟ ⎟. ⎟ ⎠

(8.2.38)

d̺2

1 Γ(r + 2)Γ( k2 ) = Dk , 2 Γ(r + 1 + k2 ) where Dk :=



Dk (ϑ) dϑ.

/ IN) in terms of (b) Note that lower bound of ϑs,p (r = s − 1 + p1 ∈ the Prohorov metric exists. This follows from an example in Maejima and Rachev (1987) in the case k = 1. We next investigate smoothing inequalities, which play an important role in the proof of Berry–Ess´een-type theorems. They are also of interest for the study of intrinsic properties of probability metrics. Lemma 8.2.8 (a) Let Z be independent of X, Y, ε > 0, r = s − 1 + p1 . Then ϑs,p (X, Y ) ≤ ϑs,p (X + εZ, Y + εZ) + 2

Γ(1 + p1 ) Γ(1 + r)

εr kEZr1 . (8.2.39)

(b) If Z, W are independent of X, Y , then ϑs,p (X + Z, Y + Z) ≤ ϑs,p (X, Y )σ(W, Z) + ϑs,p (X + W, Y + W ); (8.2.40) and moreover, ϑs,p (X + Z, Y + Z) ≤ ϑs,p (W, Z)σ(X, Y ) + ϑs,p (X + W, Y + W ), (8.2.41)

114

8. Probabilistic-Type Limit Theorems

where σ is the total variation distance. Proof: (a) By the regularity of ϑs,p (cf. 8.2.1) we have ϑs,p (X, Y ) ≤ ϑs,p (X + εZ, Y + εZ) + 2ϑs,p (0, εZ) ≤ ϑs,p (X + εZ, Y + εZ) + 2εr ϑs,p (0, Z). By (8.2.11) and (8.2.15), Γ(1 + α) ϑs,p (0, Z) ≤ Γ(1 + r)



1≤i1 ,...,is−1 ≤k

E

k  i=1

|Zi |α |Zi1 · · · Zis−1 |.

(b) For any f ∈ Gs,p we have |E(f (X + Z) − f (Y + Z))| (8.2.42) ≤ |E(f (X + Z) − f (X + W )) − E(f (Y + Z) − f (Y + W ))| + |E(f (X + W ) − f (Y + W ))|. If f ∈ Gs,p , then the translates fx (z) := f (x + z) are also in Gs,p , and therefore, the first term is estimated by conditioning on X (respectively Y ):       E(f (x + Z) − f (x + W )) dPX (x) − E(f (x + Z) − f (x + W )) dPY (x)       =  E(f (x + Z) − f (x + W )) d(PX − PY )(x) (8.2.43) ≤ ϑs,p (Z, W )σ(X, Y ).

Indeed, the inequalities (8.2.42), (8.2.43) imply ϑs,p (X + Z, Y + Z) ≤ ϑs,p (Z, W )σ(X, Y ) + ϑs,p (X + W, Y + W ). The other case is derived similarly.

2

Lemma 8.2.9 If Z is independent of X, Y and PZ has a density pZ having integrable (s − 1)-fold derivatives  Cs,Z := p (x)| dx < ∞, (8.2.44) sup |Dis−1 1 ,...,is−1 Z 1≤i1 ,...,is−1 ≤k IRk

then ϑ1,p (X + Z, Y + Z) ≤ Cs,Z ϑs,p (X, Y ).

(8.2.45)

8.2 Application to Stable Limit Theorems

Proof: For any f ∈ G1,p , E(f (X + Z) − f (Y + Z)) =



f (x) d(FX+Z − FY +Z )(x)

IRk

=



IRk

=



⎛ ⎝



115

(8.2.46) ⎞

IRk

f (x)pZ (x − z) d(FX − FY )(z)⎠ dx

f ∗ (z) d(FX − FY )(z),

IRk

where f ∗ (z) =



IRk



f (x)pZ (x − z) dx. From the Taylor expansion, ∗

f (z) = f (0) +

s−1 



j=1 1≤i1 ,...,ij ≤k



+

1

1≤i1 ,...,is ≤k 0

Dij1 ,...,ij f ∗ (0)zi1 · · · zij

(8.2.47)

(1 − t)s−1 s Di1 ,...,is f ∗ (tz)zi1 · · · zis dt. (s − 1)!

Since f ∈ G1,p , i.e., ⎛ ⎝



IR

⎞1/q

|Di1 f (x1 , . . . , xi , . . . , xn )|q dxi ⎠

≤ 1

a.s.,

(8.2.48)

we have the following bound for the qth norm of f ∗ -derivatives: ⎛ ⎝



IR

⎞1/q

|Dis1 ,...,is f ∗ (z1 , . . . , zn )|q dzi ⎠

 q ⎞1/q       = ⎝ Dis1 ,...,is f (x)pZ (x − z) dx dzi ⎠   IR IRk q ⎞1/q ⎛         dzi ⎠ p (x − z) dx = ⎝ Di11 f (x)Dis−1 Z  2 ,...,is   k IR IR q ⎞1/q ⎛         dzi ⎠ = ⎝ Di11 p (x) dx f (x + z)Dis−1 Z  2 ,...,is   k ⎛

IR

IR

(8.2.49)

116

8. Probabilistic-Type Limit Theorems

 q ⎞1/q        1  dzi ⎠ = ⎝  p (x)) dx Di1 f (x + z) (Dis−1 Z ,...,i  2 s   k IR IR ⎫1/q ⎧  ⎨ ⎬  1  (Di f (x + z))(Ds−1 pZ (x))q dzi dx ≤ i2 ,...,is 1 ⎭ ⎩ ⎛

IRk

=



IRk





IR

⎧ ⎫1/q  ⎨ ⎬  s−1   1 q D    D p (x) f (x + z) dz dx i i1 i2 ,...,is Z ⎩ ⎭ IR

pZ (x)| dx = Cs,Z |Dis−1 2 ,...,is

by (8.2.48)).

IRk

Summarizing the results in (8.2.46), (8.2.47), and (8.2.48), we derive the desired inequality (8.2.45). 2 As a consequence of Lemmas 8.2.8, 8.2.9 we next obtain an estimate between ϑ1,p and ϑs,p . Theorem 8.2.10 For every s = 1, 2, . . . , p ∈ [1, ∞), r := s − 1 + p1 , and random vectors X, Y on IRk we have 1 pr

ϑ1,p (X, Y ) ≤ A(s, p, k)ϑs,p (X, Y ),

(8.2.50)

where 1

A(s, p, k) := a pr b

s−1 r

1

(p(s − 1)) pr

r s−1

(8.2.51)

and 1 a := √ s



2s π

s/2 ,

2 8

b := 2k k

2 π

31/p .

(8.2.52) d

Proof: Recall the inequality (8.2.39). Then for any ε > 0 and any Z = N (0, 1) independent of X, Y , we have 1/p

ϑ1,p (X, Y ) ≤ ϑ1,p (X + εZ, Y + εZ) + 2ε1/p k EZ1 ;

(8.2.53)

and furthermore, 1/p

EZ1

≤ (EZ1 )1/p =

2 k  i=1

31/p

E|Zi |

2 8 31/p 2 . = k π

(8.2.54)

8.2 Application to Stable Limit Theorems

117

Now we apply Lemma 8.2.9 to get ϑ1,p (X + εZ, Y + εZ) ≤ Cs,εZ ϑs,p (X, Y ).

(8.2.55)

Next, assuming that Z is standard normal with independent components, we bound the constant in (8.2.55) as follows:  x sup pZ ( )|ε−s−n+1 dx |Dis−1 (8.2.56) Cs,εZ = 1 ,...,is−1 ε i1 ,...,is−1 IRk

= ε1−s Cs,Z = ε1−s Cs,s−1/2 (Z1 +···+Zs ) . Here, Z1 , . . . , Zs are i.i.d. copies of Z, and thus (see Zolotarev (1977, 1983, 1986)) Cs,εZ

1

= ε1−s s− 2 (1−s) Cs,Z1 +···+Zs s−1 2

≤ ε1−s s

(8.2.57)

(C1,Z )s ,

where C1,Z

=

sup



1≤i≤k IRk

=

sup i

 

IRk

=

sup i



IR

|Di1 pZ (x)| dx 1 √ 2π

n

|xi |e



x2 i 2

9



e

x2 j 2

dx

j =i

x2 1 i √ |xi |e− 2 dxi = 2π

8

2 . π

Therefore, from (3.15), (3.17), (3.19) we obtain 1 ϑ1,p (X, Y ) ≤ ε1−s √ s



2s π

2 8 31/p s/2 2 ϑs,p (X, Y ) + 2ε1/p k k . (8.2.58) π

Define ϕ(x) := ax1−s + bx1/p , 1 a := √ s



s/2 2s ϑs,p (X, Y ), π

2 8

b := 2k k

Minimizing ϕ with respect to x yields (8.2.50).

2 π

31/p

. 2

As a consequence of the smoothing properties we shall establish a Berry– Ess´een-type result, that will provide the right order estimate in the stable central limit theorem in terms of the metric ϑ1,1 . Let (Xi ) be an i.i.d.

118

8. Probabilistic-Type Limit Theorems

sequence of random vectors in IRk ; let (Θ, Θi ) be an i.i.d. sequence of d

symmetric α-stable distributed random vectors, i.e. n−1/α (Θ1 +· · ·+Θn ) = Θ; and define ϑr := ϑr (X1 , Θ) := sup hr−1 ϑ1,1 (X1 + hΘ, Θ1 + hΘ),

(8.2.59)

h>0

σr := σr (X1 , Θ) := sup hr σ(X1 + hΘ, Θ1 + hΘ), (8.2.60) h>0   1 ϑ := ϑ1,1 (X1 , Θ), σ := σ(X1 , Θ), τr := max ϑ, σ, σr(r−α) . (8.2.61) Theorem 8.2.11Suppose that 1 < α ≤ 2, α < r, and ϑ+ϑr +σ+σr < ∞. n Let Zn = n−1/α i=1 Xi . Then for some absolute constant C = C(k) depending only on the dimension k,   r 1− α −1/α ϑ1,1 (Zn , Θ) ≤ C n ϑr + τ r n . (8.2.62)

Proof: We shall use the notation ϑ(X, Y ) = ϑ1,1 (X, Y ) during this proof. Note that by (8.2.10), ϑ(X, Y ) = ζ 1 (X, Y ) = ζ1 (X, Y ). From the smoothing inequality (8.2.39) we obtain the following bound: For any ε > 0 with Θi , Θ independent and identically distributed, ϑ(Zn , Θ1 ) ≤ ϑ(Zn + εΘ, Θ1 + εΘ) + Cε,

(8.2.63)

where C := 2 k EΘ1 . Our proof will be based on the Bergstr¨om convolution method (cf. Rachev (1991, Chapter 18) and the references therein). We start by making use of the triangle inequality: ϑ(Zn + εΘ, Θ1 + εΘ)   Θ 1 + X 2 + · · · + Xn ≤ ϑ Zn + εΘ, + εΘ n1/α  m  Θ1 + · · · + Θj + Xj+1 + · · · + Xn + εΘ, + ϑ 1/α n j=1

(8.2.64)

 Θ1 + · · · + Θj+1 + Xj+2 + · · · + Xn + εΘ n1/α   Θ1 + · · · + Θm+1 + Xm+2 + · · · + Xn +ϑ + εΘ, Θ1 + εΘ n1/α m  =: Io + Ij + Im+1 , m = [n/2]. j=1

Applying the smoothing property (8.2.41), we obtain   X2 + · · · + Xn Θ2 + · · · + Θn Io ≤ ϑ , n1/α n1/α

(8.2.65)

8.2 Application to Stable Limit Theorems

119

  · σ n−1/α X1 + εΘ, n−1/α Θ1 + εΘ   X 1 + Θ2 + · · · + Θn Θ 1 + · · · + Θn + εΘ, + εΘ . +ϑ n1/α n1/α Similarly, for 1 ≤ j ≤ m, we have Ij



 Xj+2 + · · · + Xn Θj+2 + · · · + Θn ≤ ϑ , (8.2.66) n1/α n1/α   Θ1 + · · · + Θj + Xj+1 Θ1 + · · · + Θj+1 ·σ + εΘ, + εΘ n1/α n1/α   Θ1 + · · · + Θj + Xj+1 + Θj+2 + · · · + Θn +ϑ + εΘ, Θ1 + εΘ . n1/α

Summarizing the above inequalities, we get the bound ϑ(Zn , Θ1 ) ≤

5 

Δj ,

(8.2.67)

j=1

where Δ1 := Δ2 :=

Δ3 := Δ4 := Δ5 :=



   X2 + · · · + Xn Θ2 + · · · + Θn X1 Θ1 ϑ , + εΘ, 1/α + εΘ , σ n1/α n1/α n1/α n   m  Xj+2 + · · · + Xn Θj+2 + · · · + Θn , ϑ 1/α n n1/α j=1   Θ1 + · · · + Θj + Xj+1 Θ1 + · · · + Θj+1 + εΘ, + εΘ , ·σ n1/α n1/α   X 1 + Θ2 + · · · + Θ n , Θ , (m + 1)ϑ n1/α   Θ1 + · · · + Θm+1 + Xm+2 + · · · + Xn ϑ + εΘ, Θ1 + εΘ , n1/α Cε = 2k EΘ1 ε.

To estimate the terms Δi , we introduce smoothed versions of the metrics ϑ, σ defined as follows: For r > 1, ϑr (X, Y )

:=

σr (X, Y )

:=

sup hr−1 ϑ(X + hΘ, Y + hΘ),

(8.2.68)

h>0

sup hr σ(X + hΘ, Y + hΘ)

(8.2.69)

h>0

(cf. also (8.2.59), (8.2.60)). It is easy to see that both ϑr , σr are ideal metrics of order r.

120

8. Probabilistic-Type Limit Theorems

We first estimate Δ3 and Δ4 using the ideality of ϑr . In the rest of the proof, c stands for a general constant, which may be different at different places: 2   1/α 1/α 3 n−1 n−1 Θ, n−1/α Θ1 + Θ Δ3 = (m + 1)ϑ n−1/α X1 + n n   r−1   α n −1/α −1/α ϑr n X1 , n Θ (8.2.70) ≤ (m + 1) n−1 r

≤ c n1− α ϑr .

Similarly, Δ4

≤ ϑr



Xm+2 + · · · + Xn Θm+2 + · · · + Θn , n1/α n1/α r

≤ c n1− α ϑr .



n m+1

 r−1 α

(8.2.71)

Define 

1 r−α

ε := A max σ1 , σr



n−1/α ,

A > 0.

(8.2.72)

The proof continues by induction on n. The induction hypothesis states that for all j < n,   r X1 + · · · + Xj Θ1 + · · · + Θj ϑ , ≤ B(ϑr j 1− α + τr j −1/α ). (8.2.73) 1/α 1/α j j (For n = 1, . . . , n0 , n0 fixed, (8.2.62) follows from τr ≥ ϑ and the ideality of ϑ1,1 .) Then, for Δ1 = Δ1 (n), we obtain r

Δ1 ≤ Bc(n1− α ϑr + n−1/α τr )σ(n−1/α X1 + εΘ, n−1/α Θ1 + εΘ) (8.2.74) r ≤ Bc(n1− α ϑr + n−1/α τr )ε−1 n−1/α σ1   r 1 Bc n1− α ϑr + n−1/α τr . ≤ A Similarly, we estimate Δ2 : Δ2

r

≤ cB(ϑr (n − m − 2)1− α + τr (n − m − 2)−1/α ) −r/α ∞ j   −1/α −1/α α · X1 , n Θ σr n +ε n i=1 

≤ cB ϑr n 

r 1− α

r

−1/α

+ τr n

≤ cB ϑr n1− α + τr n−1/α

∞ 



j=1

σr (j + εα n)r/α

1

Ar−α

.

(8.2.75)

8.2 Application to Stable Limit Theorems

From (8.2.70)–(8.2.75) we infer     r 1 1 1− α −1/α B ϑr n + + τr n ϑ(Zn , Θ) ≤ C1 A Ar−α   r + C2 ϑr n1− α + Aτr n−1/α . 1 Choosing A big enough so that C1 ( A1 + Ar−α )< B/2 > C2 (1 + A), we complete the proof.

1 2

121

(8.2.76)

and then B such that 2

Remark 8.2.12 (a) We note that the conditions concerning the domain of attraction of Θ are given solely in terms of the metrics appearing in the upper bounds. (b) Since Θ has a density pΘ with integrable derivatives, we can get, similarly to the proof of Lemma 8.2.9, that σ(X + δΘ, Y + δΘ) ≤ Cδ −1 ϑ1,1 (X, Y ),

(8.2.77)

where C = C(Θ). This implies that for any 0 < ε < 1, σr (X, Y )

sup hr σ(X + hΘ, Y + hΘ) h>0  r = sup h σ X + h(1 − ε)1/α Θ1 + hε1/α Θ2 , h>0  1/α 1/α Y + h(1 − ε) Θ1 + hε Θ2   −1/α 1/α r Θ1 , Y + h(1 − ε) Θ2 ≤ sup h ϑ1,1 X + h(1 − ε) =

h>0

· C(Θ)/(h(1 − ε)1/α )

= C(Θ)(1 − ε)(1−r)/α ε−1/α ϑr (X, Θ). The minimum in the right-hand side is attained for ε = 1/r, implying that σr (X, Y )

≤ C(Θ)rr/α (r − 1)(1−r)/α ϑr (X, Θ)

(8.2.78)

≤ C(Θ)2r/α ϑr (X, Θ).

Similarly to the proof of Theorem 8.2.10, relations (8.2.77), (8.2.78) allow us to replace (8.2.62) by a bound involving only ϑ, ϑr :   r 1− α 1/(r−α) −1/α ϑr + max{ϑ, ϑr }n ϑ1,1 (Zn , Θ) ≤ C n . (8.2.79)

For r ∈ IN, ϑr , the r-smoothed ϑ metric, can be estimated from above by the ζr metric: ϑr ≤ cr ζr ,

(8.2.80)

122

8. Probabilistic-Type Limit Theorems

 (r) where cr depends on sup z ≤1 |pΘ (y)(z)| dy (cf. Zolotarev (1983, p. 294, property 6)). Also, for r ∈ IN, ζr is estimated from above by the rth difference pseudomoment kr , defined by & kr (X, Y ) = sup |E(f (X) − f (Y ))|; f bounded, f : IRk → IR, (8.2.81) ( (' r−1 r−1 ( ( |f (x) − f (y)| ≤ xx − y y and

r

kr (X, Y ) ≤ 2 vr (X, Y ) = 2



r

xr |PX − PY |( dx),

(8.2.82)

where vr is the absolute pseudomoment of order r. From (8.2.79)– (8.2.82) we obtain easy-to-check criteria for finiteness of the upper bounds. In particular, in the normal case α = 2, we take r = 3, and so the finiteness of the third moments of Xi implies the Berry–Ess´een result. In the case 1 < α < 2 we use the boundness of kr in (8.2.81). (c) In the case k = 1, α = 2 (normal case, dimension one) the result of Theorem 8.2.11 is due to Zolotarev (1987), based on the proof of Senatov (1980). From part (b) of the above remark, it follows that we can replace the terms ϑ, σ, ϑr , σr in the upper bound in (8.2.62) by k1 and kr . Since kr is topologically weaker than vr , it is of interest to obtain alternative bounds for kr . To this end let us recall the minimal ℓr -metric: for r > 0, & ' d r (1/r)∧1  d    ; X = X, Y = Y . (8.2.83) ℓr (X, Y ) := inf (EX − Y  ) Then

kr (X, Y ) =

& ' d d r−1 r−1       inf |EXX − Y Y  |; X = X, Y = Y (8.2.84)

= ℓ1 (XXr−1 , Y Y |r−1 ) & ' d d r−1 r−1 = inf EU − V ; U = XX , V = Y Y 

(cf. Rachev and R¨ uschendorf (1990)). If X and Y have densities fX and fY , respectively, then (cf. Rachev (1991, pp. 249–252)) we use that k1 = ℓ1 to get the bound kr (X, Y )



αr (X, Y ) (8.2.85)   1           −k−1 x x − fY Y r−1 dt dx. := fX X r−1 x1  t t t   k IR

0

8.2 Application to Stable Limit Theorems

123

For some examples with an equality in (8.2.85), see Rachev (1991, p. 252). The densities of XXr−1 , Y Y r−1 are obtainable from the transformation formula. In particular, this gives explicit bounds in the case r = 1, where the expression in (8.2.85) simplifies. The following upper bound for kr , r > 1, will turn out to be useful. Lemma 8.2.13 If EXr < ∞, EY r < ∞, r > 1, then kr (X, Y ) ≤ cℓr (X, Y ),

(8.2.86)

where the constant c depends on the rth moments of X, Y . Proof: Let r :=

r r−1

d

d

and U = X, V = Y . Then

( ( E (U U r−1− V V r−1 ( ≤

  EU r−1 U −V  + EV  U r−1− V r−1 

=: I1 + I2 .

For I1 , I2 we readily get the bounds I1 ≤ (EU r )1/r (EU − V r )1/r and ⎧ 1/r 1/r ⎪ ⎪ (EV r ) (EU − V r ) , 1 < r ≤ 2, ⎪ ⎨ I2 ≤ (r − 1)(EU − V r )1/r ⎪   ⎪ r−2 1 ⎪ r r−1 r 1/r r r−1 ⎩ (EU  ) · (EV  ) + (EV  ) , r > 2.

So I1 + I2 ≤ c(EU − V r )1/r , and passing to the corresponding minimal 2 metrics, we get kr (X, Y ) ≤ cℓr (X, Y ), as required. In some examples one can determine kr explicitly. Suppose that for some radial transformation ⎧ ⎨ α( x ) x if x = 0, x k k φ : IR → IR , φ(x) := ⎩ 0 if x = 0, d

with α monotonically nondecreasing, we have Y = φ(X). Examples of this relation include spherically invariant distributions and spherically equivalent distributions, as for example, the uniform distribution on a p-ball in IRk and the product of Weibull distributions (cf. Section 3.2). By (8.2.84), it is easy to see that the pair (X, φ(X)) is an optimal coupling with respect to kr , and so we obtain ( ( (8.2.87) kr (X, Y ) = E (XXr−1 − φ(X)φ(X)r−1 ( = E |Xr − α(X)r | .

124

8. Probabilistic-Type Limit Theorems

(A related explicit formula was derived for the ℓr distance in Section 3.2.) Note that α is determined by the equation F Y (y) = P (α(X) ≤ y) = −1 F X (α−1 (y)), which in the case of F X continuous leads to α(t) = F Y ◦ F X (t). We illustrate the above resutls, invoking again the stable limit theorem 8.2.11. Let Θ be a k-dimensional α-stable random vector with spectral  α measure m such that x dm(x) < ∞; i.e., IRk

⎫ ⎧ ⎬ ⎨ 1 α |t, s| dm(s) . E exp{it, Θ} = exp − ⎭ ⎩ 2

(8.2.88)

IRk

We apply the LePage representation for symmetric α-stable laws. Let 1. (Yj ) be an i.i.d. sequence of random vectors with distribution m/|m| and let Yj := |m|1/α Yj ;

2. ( ηj ) be i.i.d. symmetric random variables with  η1 α = (E| η1 |α )1/α < ∞ and let ηj := ηj/ ηj α ;

3. (Γj ) be the sequence of successive times of jump of a standard Poisson process and assume that the three sequences are independent. Then d

Θ = cα

∞ 

−1/α

Γj

ηj Y j ,

(8.2.89)

j=1

where the constant cα is determined by the tail behavior of the law of Θ. Without loss of generality we set cα = 1 (cf. Ledoux and Talagrand (1991, Section 5.1) and Samorodnitsky and Taqqu (1994)). Suppose that the distribution of X has a similar representation in distribution d

X =

∞ 

−1/α ∗ ∗ ηj Yj ,

(8.2.90)

Γj

j=1

where (Yj∗ ), (ηj∗ ) are independent but not necessarily identically distributed. Recall the bound (8.2.80) to see that what we need is an estimate for ζr (X, Θ) from above. Proposition 8.2.14 Let r > max{1, α}. Suppose that Yj , Yj∗ , ηj , ηj∗ have finite rth moments with supj E|ηj∗ |r < ∞. Then ⎛ ⎞ ∞ ζr (X, Θ) ≤ C sup ⎝ℓr (Yj∗ , Yj ) + |x|r−1 |Fηj∗ (x) − Fηj (x)| dx⎠. (8.2.91) j≥1

−∞

8.2 Application to Stable Limit Theorems

125

Proof: By the ideality of ζr , ζr (X, Θ) ≤

∞ 

−r/α

E(Γj

)ζr (ηj∗ Yj∗ , ηj Yj ).

j=1

Since r > α, and for j > r/α, −r/α

EΓj

the series Sr =

=

Γ(j − r/α) ∼ j −r/α , Γ(j)

∞

j=1

−r/α

E(Γj

) converges. Furthermore,

ζr (ηj∗ Yj∗ , ηj Yj ) ≤ ζr (ηj∗ Yj∗ , ηj∗ Yj ) + ζr (ηj∗ Yj , ηj Yj ) ≤

(E|ηj∗ |r )ζr (Yj∗ , Yj )

≤ C



ζr (Yj∗ , Yj )

+

+

k 

(E|Yj,i |r )ζr (ηj∗ , ηj )

i=1  ∗ ζr (ηj , ηj ) ,

where Yj,i is the ith compoment of Yj . Since 1 1 ζr (ηj∗ , ηj ) ≤ kr (ηj∗ , ηj ) = r! (r − 1)!

∞ |x|r−1 |Fηj∗ (x) − Fηj (x)| dx,

−∞

we obtain (8.2.91) after applying the inequality ζr (Yj∗ , Yj ) ≤ and Lemma 8.2.13.

1 ∗ r! κr (Yj , Yj )

2

Under the additional assumption sup EYj∗ r < ∞

(8.2.92)

j

we obtain (by the obvious bound ℓr (Yj∗ , Yj ) ≤ (EYj∗ r )1/r + (EYj r )1/r ) the finiteness of the upper bound in the limit theorem in (8.2.62). In this way we establish a stable limit theorem (with an estimate of the rate of convergence) for random vectors in the ℓr -neighborhood of a stable symmetric law in the sense of the LePage representation. For r = 1 we use the estimate in (8.2.85). For r > 1 (and in particular r = 2) explicit expressions for ℓr are known in several cases (cf. Section 3.2), for example for the distance ℓr (X, Y ) between normal distributed random vectors X and Y , between uniform distributions on balls and multivariate normal, or Weibull, distributions, and between spherically equivalent distributions.

126

8. Probabilistic-Type Limit Theorems

8.3 Application to Summability Methods and Compound Poisson Approximation In this section we apply the ϑs,p -metric (see (8.2.2)) to obtain rate of convergence results in stable limit theorems for multivariate summability methods, thus extending some results of Maejima (1985) in the real case. We also study the approximation of sums of independent random variables by compound Poisson distributions. Let (Xn )n≥0 be an i.i.d. sequence of random vectors in IRk and consider the weighted sums T (λ) :=

∞ 

cj (λ)Xj ,

j=0

cj (λ) ≥ 0,

(8.3.1)

where for λ > 0 or λ ∈ IN, (cj (λ)), j ≥ 0, is a summability method Some classical summability methods are ⎧ ⎨ 1 , 0 ≤ j ≤ n, n+1 (8.3.2) “C´esaro method” cj (λ) = ⎩ 0, otherwise; “Borel method”

“Euler method”

“Abel method”

λj −λ e , λ > 0, j ∈ IN0 ; j!   n j λ (1 − λ)n−j , cj (λ) = j 0 ≤ j ≤ n, 0 < λ < 1;

cj (λ) =

(8.3.3)

(8.3.4)

cj (λ) = (1 − e−1/λ )e−j/λ , 0 ≤ j < ∞;

(8.3.5)

0 ≤ j < ∞,

(8.3.6)

“random walk method” cj (n) = P (Sn = j),

where Sn is a random walk on the integers IN0 . For a review and discussion of these methods in the univariate case we refer to Maejima (1985). Let Θ(α) denote a random vector with symmetric stable distribution of index α, 0 < α ≤ 2. Recall (see Samorodnitsky and Taqqu (1994)) that for 0 < α < 2, Θ(α) is symmetric α-stable in IRk if and only if there exists a (unique) symmetric finite measure Γ on the unit sphere Sk such that ϕΘ(α) (t)



E exp it, Θ(α)  ⎧ ⎫ ⎨  ⎬ α = exp − |(t, s)| Γ( ds) , ⎩ ⎭

=

Sk

(8.3.7) t ∈ IRk .

8.3 Summability Methods, Compound Poisson Approximation

127

Define then ⎛

dα (λ) = ⎝

∞  j=0

⎞1/α

cj (λ)α ⎠ .

(8.3.8)

Theorem 8.3.1 Let 0 < α < r = s − 1 + p1 . Then ϑs,p



1 T (λ), Θ(α) dα (λ)



≤ R(λ)ϑs,p ,

(8.3.9)

where ⎛ ⎞r ∞  R(λ) = ⎝ cj (λ)/dα (λ)⎠,

ϑs,p := ϑs,p (X0 , Θ(α) ).

(8.3.10)

j=0

Proof: Let (Θj ) be an i.i.d. sequence with the same distribution as Θ(α) . Let us show first that ∞

1  d cj (λ)Θj = Θ(α) . dα (λ) i=0

(8.3.11)

Consider the characteristic function of the right-hand side quantity in (8.3.11): i d

Ee

1 α (λ)

∞

j=0 cj (λ)Θj ,t

=

∞ 9

j=0

i

Ee

:

cj (λ) Θ ,t dα (λ) j

;

(8.3.12)

⎫ ⎧ =α ⎬ ⎨  < c (λ)  j = exp −  t, s  Γ( ds) ⎭ ⎩ dα (λ) j=0 Sk ⎧ ⎫ α  ∞  ⎨  ⎬ cj (λ) = exp − |t, s|α Γ( ds) ⎩ ⎭ dα (λ) j=0 ∞ 9

Sk

= ϕΘ(α) (t).

By Lemma 8.2.3, ϑs,p is ideal of order r = s − 1 + 

1 p

> α. Therefore,

 1 T (λ), Θ(α) ϑs,p dα (λ) ⎛ ⎞ ∞ ∞   1 1 = ϑs,p ⎝ cj (λ)Xj , cj (λ)Θj ⎠ dα (λ) j=0 dα (λ) j=0

(8.3.13)

128

8. Probabilistic-Type Limit Theorems −r

≤ dα (λ)

∞ 

cj (λ)r ϑs,p (X0 , Θ(α) )

j=0

= R(λ)ϑs,p (X0 , Θ(α) ). 2 Note that various upper bounds for ϑs,p were established in Section 8.2. In particular, if r ∈ IN, or if X0 has a density, we have obtained upper bounds in terms of difference pseudomoments. Maejima (1985) showed that R(λ) ≤ cλ−(r−α)/α

(8.3.14)

for the C´esaro and Abel methods, and R(λ) ≤ cλ−(r−α)/2α

(8.3.15)

for the random walk method (which includes the Euler method and the Borel method as particular cases). In the Gaussian case, for r = 3 the metric ϑs,p in (8.3.9) is finite, provided that (i) if Cov (X0 ) = Ik , the unity matrix, and (ii) the components have finite third moments. Furthermore, the corresponding rate of convergence is λ−1/2 for the C´esaro and Abel methods and λ−1/4 for the random walk method. We complete this section with an application of the ideality properties of our metrics to the approximation of the distribution of sums of nonidentically distributed random vectors by a compound Poisson law. Let X1 , . . . , Xn be independent random vectors in IRk with distributions P1 , . . . , Pn of the form Pi = (1 − pi )δ0 + pi Qi ,

0 ≤ pi ≤ 1, 1 ≤ i ≤ n.

(8.3.16)

Here, δ0 stands for the one point distribution at zero. We can write Xi in the form Xi = Ci Di ,

1 ≤ i ≤ n,

(8.3.17)

where Ci has distribution Qi , Di is B(1, pi )-distributed, and Ci , Di are independent. We shall consider the approximation of S ind :=

n 

Xi

(8.3.18)

i=1

by a multivariate compound Poisson distribution P(μ, Q). P(μ, Q) is defined as the distribution of S coll :=

N  i=1

Zi ,

(8.3.19)

8.3 Summability Methods, Compound Poisson Approximation

129

where (Zi ) is an i.i.d. sequence; P Zi = Q, N is Poisson distributed with parameter μ, P N = P(μ); and N , (Zi ) are independent. The notation S ind , S coll is taken from risk theory. Recall that in the risk-theory framework in the “individual model” pi is the probability of a claim Ci with distribution Qi , corresponding to k different types of claims. S coll denotes the approximation of S ind by the “collective model”; we refer to the books of Gerber (1981) and Hipp and Michel (1990) for these and related notions. The usual choice of Q, μ in risk theory is n 

μ=μ  :=

 := Q=Q

pi ,

i=1

n  pi i=1

μ

Qi .

(8.3.20)

This leads to the following representation of S coll : S coll =

n 

Sicoll ,

(8.3.21)

i=1

where Sicoll ∼ P(pi , Qi ) (X ∼ Q denoting that X has distribution Q)  and and {Sicoll } independent. Note that with this choice μ = μ , Q = Q, moreover, E S ind = E S coll ,

(8.3.22)

if the expectations exist. If Σi = Cov (Ci ), αi = (αi,1 , . . . , αi,k ) = ECi , then 







Cov S ind

=

n 

pi Σ i +

i=1

n 

pi qi αiT αi ,

(8.3.23)

pi αiT αi .

(8.3.24)

i=1

while Cov S coll

=

n 

pi Σi +

n  i=1

i=1

As a consequence we obtain the following majorization result: 

Cov S ind



0, ∗

μ 5(z)t = μ 5(tB z)eib(t) ,

for all z ∈ IRd .

(8.4.1)

Here μ 5 is the characteristic function of μ, B ∗ is the adjoint operator of B,  ∞ and tA = exp{(ln t)A} = k=0 (k!)−1 (ln t)k Ak . The distribution μ is called strictly operator-stable if we can choose b(t) ≡ 0. In this section, we always assume that μ is a full strictly operator-stable distribution on IRd . Sharpe (1969) showed that if 1 is not in the spectrum of B, then the operatorstable law can be centered so as to become strictly operator-stable. Thus, the assumption of strict operator-stability is not so restrictive. The invertible linear operator B in (8.4.1) is called an exponent of μ. When μ is operator-stable with an exponent B, μ may satisfy (8.4.1) for other B’s; i.e., the exponent of μ is not necessarily unique. Further, we fix the value of the exponent B and denote by θ the random vector in IRd having the full strictly operator-stable distribution μ with this fixed B. It is known that every eigenvalue of B has its real part not less than 12 (see Sharpe (1969)). Recall that for a given sequence X1 , X2 , . . . of i.i.d. random vectors in d IR for which the normalized sum converges to θ, namely, n

−B

n  i=1

w

Xi → θ,

(8.4.2)

we say that {Xi } belongs to the domain of normal attraction of μ. As in the previous sections, we will be interested in the rate of convergence of (8.4.2). Remark 8.4.1 Some of the results in this section can be extended to Banach space–valued random variables. Also, our arguments can be used for similar rate of convergence problems of the max-operator-stable limit theorem. We start with some notation. Let  · 0 be the usual Euclidean norm of IR and let S(μ) be the symmetry group associated with μ, that is, the group of all invertible linear operators A on IRd such that for some a ∈ IRd , μ 5(z) = μ 5(A∗ z)eia . Since by assumption μ is full, S(μ) is compact, and thus there exists a Haar probability H on S(μ). We introduce the following norm  · , which depends on the particular operator-stable law μ but not on the choice of exponent: d

x =

 1

S(μ) 0

gtB x0

dt dH(g) t

(8.4.3)

8.4 Operator-Stable Limit Theorems

133

(see Hudson et al. (1986) and Hahn et al. (1989)). It has the following properties: (i)  ·  does not depend on the choice of the exponent B. (ii) The map t → tB x is strictly increasing on (0, ∞) for x = 0. Define the norm of the linear operator A on IRd in the usual way by A = sup x =1 Ax. Then property (ii) implies (iii) The map t → tB  is strictly increasing on (0, ∞); i.e., t → t−B  = (t−1 )B  is strictly decreasing on (0, ∞). Further, we will need estimates of the growth rate of R(t) = tB x. Meerschaert (1989) showed that for every x, the function R0 (t) = tB x0 varies regularly with index between λB and ΛB , where λB and ΛB are the minimum and the maximum of the real parts of the eigenvalues of B, respectively. Clearly, for every norm  ·  on IRd , the function R(t) = tB x will be of the same order as the regularly varying function R0 (t). In particular, for any η > 0, there exists t0 > 0 such that for any t > t0 , tλB −η x < tB x < tΛB +η x,

(8.4.4)

t−ΛB −η x < t−B x < t−λB +η x.

(8.4.5)

and

Let X (IRd ) be the class of all random vectors in IRd , and ̺ the Kolmogorov metric in X (IRd ), ̺(X, Y ) :=

sup |P (X ≤ x) − P (Y ≤ x)|.

(8.4.6)

x∈IRd

Here, and throughout this section, x ≤ y or x < y, x, y ∈ IRd , means component-wise inequality. Also, all the probability metrics μ that we shall use are in fact metrics in the space of probability laws: We write μ(X, Y ) instead of μ(PX , PY ) only for the sake of simplicity, where PX , PY stand for the probability distributions of X, Y , respectively. Next, we define a uniform metric depending on the exponent B, ̺∗ (X, Y ) := sup ̺(tB X, tB Y ).

(8.4.7)

t>0

This metric plays a crucial role in our approach to the rate of convergence problem (8.4.2). Let Var be the total variation distance in X (IRd ),

134

8. Probabilistic-Type Limit Theorems

Var(X, Y ) := 2 =



sup A∈B(IRd )

|P (X ∈ A) − P (Y ∈ A)|

(8.4.8)

|PX − PY |( dx)

IRd

=

sup{|Ef (X) − Ef (Y )|; f : IRd → IR, continuous,

|f (x)| ≤ 1 for all x ∈ IRd }.

Remark 8.4.2 It is not difficult to check that ̺∗ is topologically “between” ̺ and Var; that is, top

top

̺ ≺ ̺∗ ≺ Var . top

Here we use the standard notation μ ≺ ν, meaning that ν-convergence implies μ-convergence, but the inverse is not generally valid. Remark 8.4.3 Our aim is to present a general approach to the rate of convergence problems associated with (8.4.2) that is designed to work for different metrics in terms of which we want to obtain estimates of the rate of convergence. We start with uniform-type metrics (̺, ̺∗ , Var), and then we will proceed with Kantorovich-type minimal distances. For r > 0, define a convolution-type metric associated to Var: μr (X, Y ) := sup tB −r Var (tB X + θ, tB Y + θ).

(8.4.9)

t>0

Here and in what follows, the notation X1 + X2 means the sum of two independent random vectors X1 and X2 . We shall first list our results and then prove them, extending the general method we have outlined in Section 8.1. Theorem 8.4.4 Let θ be a full strictly operator-stable random vector in B IRd , and B an exponent of θ. Let r > Λ (≥ λ1B ) and take p such that λ2 1 λB

0, 2

̺ n−B

n 

Xi , θ

i=1

3

≤ ̺∗

2

n−B



≤ K nn

n 

Xi , θ

i=1 −B r

3

 μr + n−B τr



for all n ≥ 1.

Remark 8.4.5 In our theorem, we do not explicitly assume that {Xi } belongs to the domain of normal attraction of θ. However, since λB > 12 and rλB > 1, nn−B r μr + n−B τr → 0

as

n → ∞,

because of (8.4.5). Consequently, conditions (8.4.10) and (8.4.11) are sufficient for {Xi } to be in the domain of normal attraction of θ. As to the decreasing rate of n−B , by (8.4.5), for every η > 0, there exists n0 such that n−B  ≤ n−λB +η for every n ≥ n0 . However, we also see that for any η > 0, n−B  ≤ M n−λB +η for all n ≥ 1, where M = supt≥1 t−B+(λB −η)I  (< ∞). Note, however, that rate of convergence theorems typically describe only a relatively small subset of that domain of attraction. Letting B =

1 αI

, 0 < α ≤ 2, we have the following:

Corollary 8.4.6 Let θ be a strictly α-stable random vector with index 0 < α ≤ 2 . Let α < p < r and {Xi } be a sequence of i.i.d. random vectors in IRd satisfying τr < ∞. Then for some absolute constant K = K(d, α, p) > 0, ∗

̺

2

−1/α

n

n  i=1

Xi , θ

3



≤K n

r 1− α

μr + n

1 −α

τr



for all n ≥ 1. (8.4.12)

Resnick and Greenwood (1979) studied the limit theorem for (α1 , α2 )stable laws, which corresponds to the operator-stable limit theorem with exponent ⎞ ⎛ 1/α1 0 ⎠. B=⎝ 0 1/α2

Theorem 8.4.4 provides a bound for the rate of convergence in this particular case.

Corollary 8.4.7 Let θ = (θ(1) , θ(2) ) be a strictly (α1 , α2 )-stable bivariate α2 vector, 0 < α1 ≤ α2 ≤ 2. Let r > α21 (≥ α2 ) and take p such that α2
λB , ( 5 ( , (8.4.13) ( ( ( (r ( (r 1 ( −B (r 2 , c = (2B ( + (3B ( , a = bc

and

( 1 ( ( r I−B (r M = sup (x ( (< ∞).

(8.4.14)

x≥1

d Theorem 8.4.8 Let {Xi }∞ i=1 be a sequence of i.i.d. random vectors in IR satisfying

νr = νr (X1 , θ) := max{Var(X1 , θ), μr } ≤

a . M

(8.4.15)

Then 2

Var n−B

n 

Xi , θ

i=1

3

( (r ≤ cn (n−B ( νr ≤

( ( ( 1 ( (2−B (r n (n−B (r bM

(8.4.16) for all n ≥ 1.

It would be interesting to have a version of this theorem without condition (8.4.15). Our next theorem concerns the rate of convergence of a third type of uniform metric χ that lies “between” ̺ and Var, top

top

̺ ≺ χ ≺ Var, namely, the uniform distance between characteristic functions:   is,X , (8.4.17) χ(X, Y ) := sup |φX (s) − φY (s)|, φX (s) := E e s∈IRd

where ·, · is the inner product in IRd . The corresponding “tB -uniform” (recall the definition of ̺∗ (8.4.7)) and “smoothed” versions of χ are defined by χ∗ (X, Y ) := sup χ(tB X, tB Y ) t>0

(8.4.18)

8.4 Operator-Stable Limit Theorems

137

and ( (−r   χr (X, Y ) := sup (tB ( χ∗ tB X + θ, tB Y + θ .

(8.4.19)

t>0

d Theorem 8.4.9 Let {Xi }∞ i=1 be a sequence of i.i.d. random vectors in IR satisfying

νr∗ = νr∗ (X1 , θ) := max{χ∗ (X1 , θ), χr (X1 , θ)} ≤ Then for all n ≥ 1, 2 3 n  ( (r χ∗ n−B Xi , θ ≤ cn (n−B ( νr∗ ≤ i=1

a . M

( ( ( 1 ( (2−B (r n (n−B (r . bM

Let us now denote the density of the random vector X by pX (x) (when it exists) and define the fourth type of uniform metric d(X, Y ) := ess sup |pX (x) − pY (x)|.

(8.4.20)

x∈IRd

This is “topologically” the strongest: top

top

top

̺ ≺ χ ≺ Var ≺ d. Let K(d, B) :=

max

     max  xB ij 

2≤x≤3 1≤i,j≤d

(8.4.21)

(< ∞),

where Aij is the (i, j) component of the matrix A, and put C(d, B) = d! K(d, B)d . Let dr be the smoothed version of d, ( (−r   dr (X, Y ) := sup (tB ( d tB X + θ, tB Y + θ .

(8.4.22)

(8.4.23)

t>0

Applying Theorem 8.4.8, we obtain the following rate of convergence bound in the local central limit theorem for operator-stable random vectors. Theorem 8.4.10 Suppose X1 has a density. Let     ( B (r 6 ( B (r A = max C(d, B) (2 ( + (3 ( , 1 5 and

( (r D ≥ C(d, B) (3B ( .

138

8. Probabilistic-Type Limit Theorems

If Tr = Tr (X1 , θ) := max{d(X1 , θ), dr (X1 , θ)} < ∞ and νr ≤ min



a 1 , M M cD



,

then 2

d n−B

n  i=1

Xi , θ

3

≤ Ann−B r Tr

for all n ≥ 1.

Remark 8.4.11 Operator-stable random vectors have bounded densities (Hudson (1980)). The rest of our rate of convergence results are concerned with the minimal Lp -metrics, 0 ≤ p ≤ ∞, and in particular with ℓ1 , the Kantorovich metric, 5 1 ; see Section 8.1. Recall that the total variation distance Var is ℓ1 = L in fact the minimal L0 -metric. Recall the definition of the Lp -compound metric: For any X, Y ∈ X (IRd ), (8.4.24) Lp (X, Y ) := {E[X − Y p ]}min(1,1/p) , 0 < p < ∞, L0 (X, Y ) := E[I[X = Y ]] = P (X = Y ), (8.4.25) L∞ (X, Y ) := ess sup X −Y  = inf{ε > 0; P (X −Y  > ε) = 0}, (8.4.26) where I[A] is the indicator function of a set A. As always in this book, we assume that all random vectors X ∈ X (IRd ) are defined on a nonatomic probability space (Ω, A, P ); in this way the space of all joint laws PX,Y coincides with the space of all probability measures on IR2d . The Lp -minimal metric for 0 ≤ p ≤ ∞ was defined in (8.2.23): 5 p (X, Y ) = L 5 p (PX , PY ) := inf Lp (X,  Y ), (8.4.27) ℓp (X, Y ) = L

 d where the infimum is taken over all PX,  Y  with fixed marginals X ∼ X, d Y ∼ Y .

Remark 8.4.12 For every p ∈ [0, ∞] fixed, we shall be interested in the   5 p n−B n Xi , θ → 0. As a consequence, we rate of convergence of L i=1 shall derive the rate of convergence results in terms of the Prohorov metric π(X, Y ) = inf{ε; P (X ∈ A) ≤ P (Y ∈ Aε ) + ε for all A ∈ B(IRd )}, (8.4.28) where Aε := {x; x − A ≤ ε}.

8.4 Operator-Stable Limit Theorems

139

5 0 = 1 Var, and so Theorem 8.4.8 gives the Case p = 0. For p = 0, L 2 desired bound for the rate of convergence. Case 0< p ≤ 1. Suppose first that B, the exponent of θ, satisfies   ΛB 1 < p ≤ 1. (8.4.29) ≤ λB λ2B Then by (8.4.5), nn−B p → 0 as n → ∞. Theorem 8.4.13 Suppose 0 < p ≤ 1 and (8.4.29) holds. Let X, X1 , X2 , . . . be a sequence of i.i.d. random vectors satisfying

Then

5 p (X, θ) < ∞. 5 p := L L 5p L

2

n−B

n 

Xi , θ

i=1

3

(8.4.30)

( (p 5p , ≤ n (n−B ( L

and furthermore, 3 2 n 1  ( ( p 1 5 pp+1 . Xi , θ ≤ n p+1 (n−B ( p+1 L π n−B i=1

In the case where (8.4.29) is not satisfied, we shall prove a result similar to that in Theorem 8.4.10. Define the convolution-type metric associated 5 p : For r > 0, with L 5 p (tB X + θ, tB Y + θ). 5 p,r (X, Y ) := sup tB −r L L

(8.4.31)

t>0

Theorem 8.4.14 Let 0 < p ≤ 1 and X, X1 , X2 , . . . be a sequence of i.i.d. random vectors in IRd . Let r > λ1b , 

and

   ( −B (p ( B (r 6 ( B (r ( ( ( ( ( ( 2 3 A = max 2 + ,1 5 ( (p ( (r D ≥ (2−B ( (3B ( .

Suppose

5 p (X, θ), L 5 p,r (X, θ)} < ∞ Rp,r = Rp,r (X, θ) := max{L

140

8. Probabilistic-Type Limit Theorems

and νr ≤ min



1 a , M M cD



,

where a, c, and M are defined in (8.4.13) and (8.4.14). Then, for all n ≥ 1,  2

π n−B

n 

Xi , θ

i=1

3p+1

5p ≤ L

2

n−B

n 

Xi , θ

i=1

3

≤ Ann−B r Rp,r .

a , M1cD In the next theorem we shall relax the assumption νr ≤ min M at the cost of losing a little of the order of convergence nn−B r . The next result has a form resembling Theorem 8.4.4. Theorem 8.4.15 Let 0 < p ≤ 1. Let r > 1 λB

0, 2 3 n  (   ( (r ( 5 p,r + (n−B (p Qp,r , 5 p n−B Xi , θ ≤ K n (n−B ( L L i=1

for all n ≥ 1.

Case 1< p ≤2. For 1 < p ≤ 2, we use a completely different method in the rate of convergence problem, which relies on the minimality property 5 p and was introduced in Rachev and R¨ uschendorf (1992). of L

B Theorem 8.4.16 Suppose 1 < p ≤ 2 and p > Λ (≥ λ2B X2 , . . . be a sequence of i.i.d. random vectors satisfying

5p = L 5 p (X, θ) < ∞ L

and

E[X − θ] = 0.

Then there exists Cr > 0 such that for all n ≥ 1, 3 2 n  ( 1 ( 5 p n−B 5p , L Xi , θ ≤ Cp n p (n−B ( L i=1

1 λB ).

Let X, X1 ,

8.4 Operator-Stable Limit Theorems

141

and moreover, the right-hand side vanishes as n → ∞. Furthermore, 2 3 n p p  ( ( p 1 5 pp+1 . π n−B Xi , θ ≤ Cpp+1 n p+1 (n−B ( p+1 L i=1

Corollary 8.4.17 Let θ = (θ(1) , θ(2) ) be a strictly (α1 , α2 )-stable bivariα2 ate vector, 0 < α1 ≤ α2 ≤ 2. Let 2 ≥ p > α12 (≥ α2 ). Let {Xi = (2) (1) 5p = (Xi , Xi )}i≥1 be a sequence of i.i.d. random vectors satisfying L 5 p (X1 , θ) < ∞, and if p > 1, we additionally assume that E[X1 − θ] = 0. L Then for all n ≥ 1, 3 2 n n   (1) (2) Xi , n−1/α2 Xi ), θ π (n−1/α1 i=1

5p ≤ L ≤

i=1

22

n−1/α1

n 

(1)

Xi , n−1/α2

i=1

⎧   1 p ⎪ ⎨ n1− α1 L 5 p p+1   p ⎪ ⎩ Cp n p1 − α11 L 5 p p+1

n  i=1

(2)

Xi

3

, θ

for

0 < p ≤ 1,

for

1 < p ≤ 2.

3 max(1,p) p+1

5 p can be Remark 8.4.18 Our approach based on the use of “ideality” of L >n extended to bound the distance between the maxima MX (n) := n−B k=1 k   >n >n k −B and M X θ (n) := n θ i=1 i i=1 i . (Here k=1 stands for the k=1 componentwise maximum, and {θi } are i.i.d. copies of θ.) Also, we can   ?n k ?n k −B compare?mX (n) := n−B k=1 X θ (n) := n with m i i θ i=1 k=1 i=1 n (where k=1 stands for the componentwise minimum) and aX (n) := n−B ( ( ( >n ( (k Xi ( with a (n) := n−B >n (k θi (. θ i=1 k=1 i=1 k=1 Theorem 8.4.19 Let 1 < p ≤ 2 and θ be a full strictly operator-stable random vector with exponent B such that n2 n−B p → 0

as

n → ∞.

Let X, X1 , X2 , . . . be i.i.d. random vectors in IRd with E[X − θ] = 0 and such that Lp (X, θ) < ∞. Then there exists Cp,d > 0 such that for every n ≥ 1, & ' 5 p (M (n), M (n)) , L 5 p (m (n), m (n)) , L 5 p (a (n), a (n)) max L X X X θ θ θ ( ( 5 p (X, θ). ≤ Cp,d n2/p (n−B ( L

142

8. Probabilistic-Type Limit Theorems

Remark 8.4.20 Note that aX (n) and aθ (n) are positive random variables, and therefore ⎞1/p ⎛ 1   p 5 p (a (n), a (n)) = ⎝ F −1 (t) − F −1 (t) dt⎠ , L X θ a (n) a (n) X

θ

0

−1 where FX is the generalized inverse of FX ; cf. Theorem 3.1.2. Also, from the above bound we can get the rate of convergence for π by making use of p

the bound π ≤ Lpp+1 . Let us compare Theorem 8.4.16 with a similar result on the rate of convergence in terms of the Zolotarev metric ζr , r > 0; see (8.2.1). Theorem 8.4.21 Let X, X1 , X2 , . . . be i.i.d. random vectors in IRd , and r a positive constant satisfying the conditions ζr := ζr (X, θ) < ∞

and

nn−B r → 0 as n → ∞.

Then for every n ≥ 1, 2 3 n  ζr n−B Xi , θ ≤ nn−B r ζr , i=1

and for some Cr > 0, 3 2 n  1 1 ( r ( 1 π n−B Xi , θ ≤ Crr+1 n r+1 (n−B ( r+1 ζrr+1 . i=1

It is known that if r is an integer, then ζr on the right-hand sides of the above bounds can be estimated by the rth difference pseudomoment κr from above (see Zolotarev (1993)). Namely, if all mixed moments of order less than or equal to r − 1 for X and Y agree, then ζr (X, Y ) ≤

1 κr (X, Y ), r!

r ∈ IN,

(8.4.32)

where κr is rth difference pseudomoment & κr (X, Y ) = sup |E[f (X) − f (Y )]| ; f : IRd → IR, (8.4.33) ( ( ' ( r−1 r−1 ( d |f (x) − f (y)| ≤ (x x − y y ( for all x, y ∈ IR .

For arbitrary r > 0, ζr is bounded from above by the absolute pseudomoment, namely, if all mixed moments of X and Y of order less than or equal to m (r = m + α, m ∈ IN, α ∈ (1, 2]) agree, then ζr ≤

Γ(1 + α) ξr , Γ(1 + r)

(8.4.34)

8.4 Operator-Stable Limit Theorems

where ξr is the rth absolute pseudomoment  xr |PX − PY |( dx). ξr (X, Y ) :=

143

(8.4.35)

IRd

Let us now compare the rate of convergence in Theorem 8.4.21 with that in Theorem 8.4.16 for r = p ∈ (1, 2]. Recall that (8.3.32) is true only for r ∈ IN, and the known estimates for ζr from above by κr (r being noninteger) involve E[Xr ] and E[Y r ]. However, for any p ≥ 1, 5 pp (X, Y ) ≤ 2p κp (X, Y ) ≤ 2p ξp (X, Y ). L

(8.4.36)

5 p (X, θ) < ∞ in Theorem 8.4.16 is preferable Therefore, the restriction L to ζr (X, θ) < ∞ in Theorem 8.4.21. On the other hand, the estimate for n −B ζr (n i=1 Xi , θ) holds for any r > 0 and provides us with the exact order of convergence (as n → ∞) under the assumption ζr (X1 , θ) < ∞. Case 2< p ≤ ∞. Theorem 8.4.22 T8.4.21 Let θ be a full strictly operator-stable random vector that does not have a Gaussian component, or equivalently, whose exponent B satisfies n1/2 n−B  → 0

as

n → ∞.

5 p (X, θ) < ∞. Then Let X, X1 , X2 , . . . be i.i.d. random vectors such that L for some C(d, p) > 0, 2 3 n  5 p n−B 5 p (X, θ). L Xi , θ ≤ C(d, p)n1/2 n−B L i=1

Before starting with the proof of our theorems, we introduce a notion of ideality for a probability metric, designed for problem (8.4.2).

Definition 8.4.23 A metric ζ : X (IRd ) × X (IRd ) → [0, ∞) is called operator-ideal of order r ≥ 0 if (i) (homogeneity) ζ(aB X, aB Y ) ≤ aB r ζ(X, Y ) for any a > 0, and (ii) (regularity) ζ(X + Z, Y + Z) ≤ ζ(X, Y ) for any Z independent of X and Y . We next show a few lemmas needed for the proof of our main results.

144

8. Probabilistic-Type Limit Theorems

Lemma 8.4.24 ̺, κr , and ℓ are regular (that is, (ii) holds); ̺∗ , Var, and χ∗ are operator-ideal of order r = 0; and ̺ ≤ ̺∗ ≤ 21 Var, χ ≤ χ∗ ≤ Var. This follows directly from the definitions of the metrics. 5 p,r , and ζr are operator-ideal of order r > 0. Lemma 8.4.25 μr , χr , dr , L 5 p is operator-ideal of order p ∧ 1. L

Proof: We first show the operator-ideality of μr . For any a > 0, ( (−r μr (aB X, aB Y ) = sup (tB ( Var((ta)B X + θ, (ta)B Y + θ) t>0

(  (−r 3 2   B B ( t B( t t ( ( = sup ( a X + θ, a Y +θ ( Var ( a a t>0 ( a ( (−r ≤ aB r sup (tB ( Var(tB X + θ, tB Y + θ), t>0

since ( B ( ( B (−1 (t ( (a (

( ( ( ( (−1 ( = (tB a−B aB ( (aB ( ≤ (tB a−B ( = aB r μr (X, Y ),

which shows the homogeneity of μr of order r > 0. We also have μr (X + Z, Y + Z) = sup tB −r Var(tB (X + Z) + θ, tB (Y + Z) + θ) t>0

≤ sup tB −r Var(tB X + θ, tB Y + θ) t>0

= μr (X, Y ), since tB Z is independent of tB X and θ, and Var is regular. This demon5 p,r strates the regularity of μr . One can check the ideality of χr , dr , and L in a similar fashion.

We next show the operator-ideality of ζr . The regularity of ζr is known. As for the homogeneity, we have & ζr (aB X, aB Y ) = sup |E[f (aB X) − f (aB Y )]|; ' (r−1) (r−1) f (x) − f (y) ≤ x − y . (8.4.37) Let fa (x) := f (aB x). Then

  (r−1) fa(r−1) (x)(h)(r−1) = f (r−1) aB x aB h ,

8.4 Operator-Stable Limit Theorems

145

implying that ( ( ( B (r−1 (  B ( ( (r−1) ( ( ( (r−1)  B  (r−1) (r−1) ( ( (x) − fa (y)( ≤ a a y (. a x −f (fa (f Then the side condition in (8.4.37), ( ( ( (r−1) ( (r−1) (x) − f (y)( ≤ x − y, (f

results in ( ( ( (r−1 ( B ( ( ( ( (r−1) ( (r−1) (a x − aB y ( ≤ (aB (r x − y. (x) − fa (y)( ≤ (aB ( (fa

Consequently, by (8.4.37),   ζr aB X, aB Y & ' ≤ sup |E[fa (X) − fa (Y )]| ; fa(r−1) (x) − fa(r−1) (y) ≤ aB r x − y = aB r ζr (X, Y ),

which shows the regularity of ζr .

2

Lemma 8.4.26 If Z, W ∈ X (IRd ) are independent of X, Y ∈ X (IRd ), then ̺∗ (X + Z, Y + Z) ≤ ̺∗ (Z, W ) Var(X, Y ) + ̺∗ (X + W, Y + W ) and ̺∗ (X + Z, Y + Z) ≤ ̺∗ (X, Y ) Var(Z, W ) + ̺∗ (X + W, Y + W ). Lemma 8.4.27 If Z, W ∈ X (IRd ) are independent of X, Y ∈ X (IRd ), then Var(X + Z, Y + Z) ≤ Var(Z, W ) Var(X, Y ) + Var(X + W, Y + W ) and χ∗ (X + Z, Y + Z) ≤ χ∗ (Z, W )χ∗ (X, Y ) + χ∗ (X + W, Y + W ). Lemma 8.4.28 If Z, W ∈ X (IRd ) are independent of X, Y ∈ X (IRd ), then d(X + Z, Y + Z) ≤ d(Z, W ) Var(X, Y ) + d(X + W, Y + W ), d(X + Z, Y + Z) ≤ d(X, Y ) Var(Z, W ) + d(X + W, Y + W ), 5p . and for 0 < p ≤ 1 both inequalities hold with d replaced by L

146

8. Probabilistic-Type Limit Theorems

Proof: The proofs are very similar to those in Lemma 8.1.15; cf. also Lemma 2 in Senatov (1980) or Lemmas 14.3.3 and 14.3.6 in Rachev (1991). 5p , 0 < We shall demonstrate only the proof of the smoothing inequality for L 5p : p ≤ 1. We use the dual representation for L 5 p (X + Z, Y + Z) L = sup |E[f (X + Z) − f (Y + Z)]| f ∈ Lipb (p)

(recall that Lipb (p) consists of all bounded continuous functions on IRd satisfying |f (x) − f (y)| ≤ x − yp for all x, y ∈ IRd )     = sup  PZ ( dz)(E[f (X + z)] − E[f (Y + z)]) f ∈ Lipb (p)





    sup  (PZ − PW )( dz) (E[f (X + z)] − E[f (Y + z)]) f ∈ Lipb (p)     + sup  PW ( dz) (E[f (X + z)] − E[f (Y + z)])



f ∈ Lipb (p)

|PZ − PW |( dz)

sup

f ∈ Lipb (p)

|(E[f (X + z)] − E[f (Y + z)])|

+ Lp (X + W, Y + W ) =

Var(Z, W )Lp (X, Y ) + Lp (X + W, Y + W ),

as desired.

2

5 p , 0 < p ≤ 1) Let θ Lemma 8.4.29 (Smoothing inequalities for ̺∗ and L and θ1 be independent random vectors in IRd having the same full strictly operator-stable distribution with exponent B. Then for any X ∈ X (IRd ) independent of θ1 and δ > 0, ̺∗ (X, θ) ≤ C1 ̺∗ (X + δ B θ1 , θ + δ B θ1 ) + C2 δ B , and for 0 < p ≤ 1, if E[θp ] < ∞, 5 p (X, θ) ≤ Lp (X + δ B θ1 , θ + δ B θ1 ) + 2δ B p E[θp ]. L

Here and in what follows, the Ci ’s are absolute constants depending only on d and B, unless stated otherwise explicitly. B   = tB Proof: Fix ε ∈ (0, 1) and choose X ε X, θ = tε θ for some tε > 0 such  + ε. We first show the inequality  θ) that ̺∗ (X, θ) ≤ ̺(X,

 ≤ C1 ̺(X  θ)  + δ B θ1 , θ + δ B θ1 ) + C2 δ B  ̺(X,

(8.4.38)

8.4 Operator-Stable Limit Theorems

147

d   = ̺(cX,  for any c > 0, we can shrink X,   θ)  cθ)  θ, for θ1 ∼ θ. Since ̺(X, and θ1 without any loss of generality. So we assume

 < 1) > 2 . P (θ 3

(8.4.39)

For brevity we shall delete the “  ” from now on. Let θ(i) ∈ IR, i = 1, . . . , d, be the ith component of the operator-stable random vector θ ∈ IRd . Then for each i = 1, . . . , d, θ(i) has a bounded density; that is, for some M < ∞,    d  P (θ(i) ≤ x) =: sup |pθ(i) (x)| ≤ M for all i. sup  x∈IR x∈IR dx

(See Hudson (1976).) The idea of the following proof is taken from Lemma 12.1 in Bhattacharya and Rao (1976). First consider the case ̺ := ̺(X, θ)

=

sup |P (X ≤ x) − P (θ ≤ x)|

(8.4.40)

x∈IRd

= − inf (P (X ≤ x) − P (θ ≤ x)) . x∈IRd

Given η ∈ (0, ̺), there exists x0 ∈ IRd such that P (X ≤ x0 ) − P (θ ≤ x0 ) < −̺ + η.

(8.4.41)

We then have I := P (X + δ B θ1 ≤ x0 − δ B e) − P (θ + δ B θ1 ≤ x0 − δ B e)    = P (X + z ≤ x0 − δ B e)−P (θ + z ≤ x0 − δ B e) P (δ B θ1 ∈ dz) IRd

=



E

+



,

Ec

where E := {z ∈ IRd − δ B e < z < δ B e} and e = (1, 1, . . . , 1)t ∈ IRd . Then estimating both integrals in the representation for I, we get    I ≤ P (X ≤ x0 ) − P (θ ≤ x0 − z − δ B e) P (δ B θ1 ∈ dz) E

+ ̺P (δ B θ1 ∈ E c ).

(8.4.42)

148

8. Probabilistic-Type Limit Theorems

To estimate the last term observe that 2 , 3

β := P (δ B θ1 ∈ E) ≥ P (δ B  θ1  < δ B ) >

(8.4.43)

by (8.4.39). On the other hand, denoting the distribution function of θ by F (x), x = (x(1) , x(2) , . . . , x(d) )t ∈ IRd , and ε = (ε(1) , ε(2) , . . . , ε(d) )t ∈ IRd , we have |P (θ ≤ x + ε) − P (θ ≤ x)| d    ≤ F (x(1) , . . . , x(i−1) , x(i) + ε(i) , . . . , x(d) + ε(d) ) i=1

(1)

−F (x



d  i=1

(i)

(i+1)

,...,x ,x



(i+1)

(d)

,...,x



(d)

P (θ(i) ∈ Ii ).

  )

Here Ii := (x(i) , x(i) + ε(i) ] or := (x(i) + ε(i) x(i) ] depending on the sign of ε(i) . Therefore, d 

P (θ

(i)

i=1

∈ Ii ) ≤

d  i=1

|ε(i) | sup |pθ(i) (x)| ≤ M ε1 , x∈IR

where  · 1 is the L1 -norm. Hence, −P (θ ≤ x + ε) ≤ −P (θ ≤ x) + M ε1 .

(8.4.44)

Thus we have, by (8.4.41), (8.4.42), and (8.4.44) with ε = −z − δ B e, that    I ≤ P (X ≤ x0 ) − P (θ ≤ x0 ) + M (z1 + dδ B ) P (δ B θ1 ∈ dz) E





E

+ ̺P (δ B θ1 ∈ E c )   −̺ + η + M (z1 + dδ B ) P (δ B θ1 ∈ dz) + ̺P (δ B θ1 ∈ E c ).

Since z1 ≤ dδ B  on E, it follows that I

≤ (−̺ + η + 2M dδ B )P (δ B θ1 ∈ E) + ̺P (δ B θ1 ∈ E c ) ≤ ̺(1 − 2β) + η + 2M dδ B .

Consequently, (2β − 1)̺ ≤ ̺(X + δ B θ1 , θ + δ B θ1 ) + 2M dδ B  + η.

8.4 Operator-Stable Limit Theorems

149

Since η can be taken arbitrarily small, we have  1  ̺(X + δ B θ1 , θ + δ B θ1 ) + 2M dδ B  . 2β − 1   Since β > 23 by (8.4.43), ̺ ≤ 3 ̺(X + δ B θ1 , θ + δ B θ1 ) + 2M dδ B  . This proves the inequality ̺(X, θ) ≤ C1 ̺(X + δ B θ1 , θ + δ B θ1 ) + C2 δ B  with C1 = 3 and C2 = 6M d, provided that (8.4.40) holds. ̺≤

If, on the other hand, ̺ = supx∈IRd (P (X ≤ x) − P (θ ≤ x)), then given η ∈ (0, ̺), there exists x0 such that P (X ≤ x0 ) − P (θ ≤ x0 ) > ̺ − η. Then we similarly have P (X + δ B θ1 ≤ x0 + δ B e) − P (θ + δ B θ1 ≤ x0 + δ B e)    = P (X + z ≤ x0 + δ B e) − P (θ + z ≤ x0 + δ B e) P (δ B θ1 ∈ dz) IRd

=



E





E





E

+



Ec

  P (X ≤ x0 ) − P (θ ≤ x0 − z + δ B e) P (δ B θ1 ∈ dz) − ̺P (δ B θ1 ∈ E c )

  P (X ≤ x0 ) − P (θ ≤ x0 ) − M (z1 + dδ B ) P (δ B θ1 ∈ dz) − ̺P (θB θ1 ∈ E c )

≥ ̺(2β − 1) − η − 2M dδ B . B B B Hence (2β  − 1)̺B ≤ ̺(XB+ δ θ1 , θ + Bδ θ1 ) + 2M dδ  + η, so that ̺ ≤ 3 ̺(X + δ θ1 , θ + δ θ1 ) + 2M dδ  . This completes the proof of the inequality (8.4.28), and reintroducing the symbol “ ”, we write

 ≤ C1 ̺(X  + δ B θ1 , θ + δ B θ1 ) + C2 δ B .  θ) ̺(X,

  and θ, Therefore, by the definition of ̺∗ , X, ̺∗ (X, θ)

 +ε  θ) ≤ ̺(X,  + δ B θ1 , θ + δ B θ1 ) + C2 δ B  + ε ≤ C1 ̺(X

B B B B ≤ C1 ̺(tB ε (X + δ θ1 ), tε (θ + δ θ1 )) + C2 δ  + ε ≤ C1 ̺∗ (X + δ B θ1 , θ + δ B θ1 ) + C2 δ B  + ε,

150

8. Probabilistic-Type Limit Theorems

which yields the first smoothing inequality of Lemma 8.4.29. 5 p (0 < p ≤ 1). By Now let us show an inequality of a similar type for L 5p , the triangle inequality and the regularity of L

5 p (X, θ) ≤ L 5 p (X, X + δ B θ1 ) + L 5 p (X + δ B θ1 , θ+ δ B θ1 ) + L 5 p (θ, θ+ δ B θ1 ) L 5 p (X + δ B θ1 , θ + δ B θ1 ) + 2L 5 p (0, δ B θ1 ). ≤ L

5 p as a minimal metric with respect to the Lp -metric, From the definition of L it follows that 5 p (0, δ B θ) = E[δ B θp ] ≤ δ B p E[θp ], L

which completes the proof of Lemma 8.4.29.

2

The proof of the next two lemmas is obvious. Lemma 8.4.30 For any a > 0 and r > 0, Var(aB X + θ, aB Y + θ) ≤ aB r μr (X, Y ), χ∗ (aB X + θ, aB Y + θ) ≤ aB r χr (X, Y ), d(aB X + θ, aB Y + θ) ≤ aB r dr (X, Y ),

and for 0 < p ≤ 1, 5 p (aB X + θ, aB Y + θ) ≤ aB r L 5 p,r (X, Y ), L

where X, Y ∈ X (IRd ) are independent of θ.

Lemma 8.4.31 Let Aut(IRd ) be the set of all invertible linear operators (automorphisms). Then, for any A ∈ Aut(IRd ), d(AX, AY ) = |JA−1 |d(X, Y ), where JA is the Jacobian of the matrix A. Lemma 8.4.32 For x ∈ [2, 3], |JxB | ≤ C(d, B), where C(d, B) is defined in (8.4.22). Proof: Note that |JxB | = | det xB |.

8.4 Operator-Stable Limit Theorems

151

If A is d × d matrix, then | det A| ≤ d! | max Aij |d , 1≤i,j≤d

which proves the lemma.

2

Lemma 8.4.33 Let θ be a full strictly operator-stable random vector in IRd with exponent B. Then for any two independent copies θ1 and θ2 of θ and for any t, s > 0, d

tB θ1 + sB θ2 ∼ (t + s)B θ. Proof: By (8.4.1) with b(t) ≡ 0, 

iz,tB θ1 +sB θ2 

E e





iz,tB θ1 





iz,sB θ2 



E e = E e  ∗   ∗  = μ 5 tB z μ 5 sB z = μ 5(z)t μ 5(z)s   t+s B∗ = μ 5(z) = μ 5 (t + s) z   iz,(t+s)B θ = E e .

2

The following lemmas are proved in Sections 2.5 and 2.6. Further, Cb (IRd ) stands for the set of all bounded continuous functions on IRd . 5p ) L 5 p admits the following repreLemma 8.4.34 (Duality theorems for L sentation: (i) For p = 0, 5 0 (X, Y ) L

= =

sup{|E[f (X) − f (Y )]|; f ∈ Cb (IRd ) such that |f (x) − f (y)| ≤ 1 for all x, y ∈ IRd } 1 Var(X, Y ). 2

(ii) For 0 < p ≤ 1, 5 p (X, Y ) L

=

sup{|E[f (X) − f (Y )]|; f ∈ Cb (IRd )

such that |f (x) − f (y)| ≤ x − yp for all x, y ∈ IRd }.

152

8. Probabilistic-Type Limit Theorems

(iii) For 1 < p < ∞, 5 p (X, Y ) L

=

sup{E[f (X)] − E[g(Y )]; f, g ∈ Cb (IRd )

such that |f (x) − g(y)| ≤ x − yp for all x, y ∈ IRd }.

(iv) For p = ∞, 5 ∞ (X, Y ) L

inf{ε > 0; P (X ∈ A) ≤ P (Y ∈ Aε ) for all A ∈ B(IRd )}.

=

5 p -convergence) (i) For any 0 ≤ p ≤ ∞, L 5 p -convergence Lemma 8.4.35 (L implies weak convergence. Moreover, if π is the Prohorov metric, then 1

and

5 pp+1 π ≤ L

for all 0 ≤ p ≤ 1

p

5 pp+1 π ≤ L

for all 1 ≤ p ≤ ∞.

(ii) Let 0 < p < ∞ and E[Xn p ] + E[Xp ] < ∞. Then 5 p (Xn , X) → 0 L

if and only if w

Xn → X

E[Xn p ] → E[Xp ].

and

5 p in X (IR)) For d = 1, 1 ≤ Lemma 8.4.36 (Explicit representations for L p ≤ ∞, ⎛

5 p (X, Y ) = ⎝ L

5 ∞ (X, Y ) L

=

1 0

⎞1/p

−1 |FX (t) − FY−1 (t)|p dt⎠

,

1 ≤ p < ∞,

−1 sup |FX (t) − FY−1 (t)|.

0≤t≤1

5 p ) For 1 ≤ p < ∞, Lemma 8.4.37 (Upper bounds for L 5 p ≤ κp ≤ ξp , 2−p+1 L

where κp (resp. ξp ) is the pth difference (resp. absolute) pseudomoment.

8.5 Proofs of the Rate of Convergence Results

153

8.5 Application to Operator-Stable Limit Theorems: Proofs of the Rate of Convergence Results In this section we give the proofs of the rate of convergence results stated in Section 8.4. Proof of Theorem 8.4.4: All probability metrics for the random vectors in this section are defined by their marginal distributions and are consequently independent of their joint distributions. So, without loss of generality, we assume that {Xi } and θ are independent of each other.

Let {θi } be independent copies of θ and assume that {θi } are independent of {Xi } and θ. Then by the definition of θ (or by (8.4.1) with b(t) ≡ 0), −B

n

n  i=1

d

θi ∼ θ

for any n = 1, 2, . . . .

(8.5.1)

Now, by Lemma 8.4.29 and (8.5.1), for any δ > 0, 2 3 n  Xi , θ ̺∗ n−B i=1

≤ C1 ̺ ∗ = C1 ̺∗

2

n−B

2

n−B

n 

i=1 n 

Xi + δ B θ1 , θ + δ B θ1 Xi + δ B θ, n−B

i=1

≤ ̺∗ +

n−B

n  i=1

̺

j=1

+ ̺∗

θi + δ B θ

3

i=1

2

m 

+ C2 δ B 

i=1

Furthermore, by the triangle inequality, 2 3 n n   Xi + δ B θ, n−B θi + δ B θ ̺∗ n−B i=1

n 

3

(8.5.2)

2



Xi + δ B θ, n−B θ1 + n−B ⎛

∗⎝ −B⎝

n

n−B

j  i=1

2m+1  i=1

θi +

n 

i=j+1



n 

Xi + δ B θ

i=2

+ C2 δ B .

3

Xi ⎠ + δ B θ,

⎛ ⎞ ⎞ j+1 n   n−B ⎝ θi + Xi ⎠ + δ B θ⎠ i=1

θi +

n 

i=m+2

Xi

3

i=j+2

+ δ B θ, n−B

n  i=1

3

θi + δ B θ ,

154

8. Probabilistic-Type Limit Theorems

where m = [ n2 ], n ≥ 5. By Lemma 8.4.26, the above is ≤

̺∗

2

n−B

+ ̺∗

+

2

n 

Xi , n−B

i=2

n−B (X1 +

m  j=1



n 

i=2 n 

⎣̺∗ ⎝n−B

θi ) + δ B θ, n−B

n 

× Var n−B j 

+ ̺∗ ⎝n−B ( + ̺∗ 4 

=:

n−B

2 j 

n 

i=j+2

θi + δ B θ



i=i

n 

i=1

3

θi ⎠

θi + Xj+1

θi + Xj+1 +

2m+1 

n  i=1

Xi , n−B

i=j+2

2

2

  Var n−B X1 + δ B θ, n−B θ1 + δ B θ

i=2





θi

3

3

+ δ B θ, n−B

i=j+2

θi +

i=1

n 

Xi

i=m+2

θi + δ B θ

i=1

θi ) + δ B θ, n−B

3

j+1 

n  i=1

+ δ B θ, n−B

n 

⎞⎤

θi + δ B θ⎠⎦

θi + δ B θ

i=1

3

3

Δk .

k=1

Here, the summands Δk are defined as follows: 2 3 n n     Δ1 = ̺∗ n−B Xi , n−B θi Var n−B X1 + δ B θ, n−B θ1 + δ B θ , Δ2

=

m  j=1



i=2

̺∗ ⎝n−B 2

× Var n−B Δ3

= (m + 1)̺∗

Δ4

2

= ̺∗

n−B

2

i=2

n 

Xi , n−B

i=j+2

2 j 

2m+1  i=1

i=j+2

θi + Xj+1

i=1

n−B

n 

2

X1 +

θi +

n 

3

θi

i=2

n 

Xi

i=m+2

θi ⎠

+ δ B θ, n−B

3

3



i=1

k=1

3

θi + δ B θ , 3

+ δ B θ, θ1 + δ B θ ,

+ δ B θ, n−B

Next, by (8.5.2), 2 3 n 4   −B ∗ ̺ n Δk + Δ5 , Xi , θ ≤ C1 i=1

j+1 

n  i=1

3

θi + δ B θ .

(8.5.3)

8.5 Proofs of the Rate of Convergence Results

155

with Δ5 = C2 δ B . We shall estimate each Δk separately. (I) Estimate for Δ3 . We have 2

Δ3 = (m + 1)̺

−B



n

X1 + n

−B

B

(n − 1) (n − 1)

−B

n 

θi + δ B θ,

i=2

n−B θ1 + n−B (n − 1)B (n − 1)−B 

= (m + 1)̺∗ n−B X1 + n−B (n − 1)B θ2 + δ B θ, −B

n

θ1 + n

−B

B

B

(n − 1) θ2 + δ θ

n  i=2

θi + δ B θ

3



[by (8.5.1)]   ≤ (m + 1)̺∗ n−B X1 + n−B (n − 1)B θ2 , n−B θ1 + n−B (n − 1)B θ2 [by the regularity of ̺∗ ]   1 (m + 1) Var n−B X1 + n−B (n − 1)B θ, n−B θ1 + n−B (n − 1)B θ ≤ 2 1 [since ̺∗ ≤ Var] 2   1 ≤ (m + 1) Var (n − 1)−B X1 + θ, (n − 1)−B θ1 + θ 2

[by the homogeneity of Var]. Thus, we have Δ3 ≤

(r 1 ( n ((n − 1)−B ( μr (X1 , θ1 ), 2

(8.5.4)

invoking Lemma 8.4.30. Now, using the fact that t → tB  is strictly increasing on (0, ∞), we have ( ( ( n − 1 −B ( ( ( ( ( ( −B ( ((n − 1)−B ( ≤ ( (8.5.5) ( ( n ( ( ( n (  ( ( 1 −B ( ( ( ( ( ( ( ( ( −B ( ≤ ( ( n ( = (2B ( (n−B ( . ( 2 ( Thus it follows from (8.5.4) and (8.5.5) that ( (r Δ3 ≤ C3 n (n−B ( μr ,

with C3 = 12 2B r .

(8.5.6)

156

8. Probabilistic-Type Limit Theorems

(II) Estimate for Δ4 . Similarly, we have 2 Δ4



= ̺

−B

n

B

−B

(m + 1) (m + 1)

m+1  i=1

n−B (m + 1)B (m + 1)−B ≤ ̺∗

θi + n

−B

n 

n−B (m + 1)B θ1 + n−B

m+1 

θi + n−B

n 

θi

3

[by (8.5.1) and the regularity of ̺ ] 2 n  1 −B B −B Xi , ≤ Var n (m + 1) θ1 + n 2 i=m+2 n−B (m + 1)B θ1 + n−B ≤ ≤

2

n 

i=m+2 n 

θi + δ B θ

3

Xi ,

i=m+2 ∗

n 

i=m+2

i=m+2

n−B (m + 1)B θ1 + n−B

Xi + δ B θ,

i=m+2

i=1

2

n 

θi

3 n 

1 Xi + θ, (m + 1)−B θi + θ Var (m + 1)−B 2 i=m+2 i=m+2 2 n 3 n   1 −B r Xi , θi (m + 1)  μr 2 i=m+2 i=m+2

3

[by Lemma 8.4.30] ( 1( ((m + 1)−B (r (n − m − 1)μr (X1 , θ1 ) ≤ 2

by the triangle the repeated use of the regularity of μr . ( and ( inequality (  ( ( B( ( n −B ( 1 −B ( (2 (, we have ( ≤ ( = Finally, by ( ( ( 2 ( ( m+1 ( (r Δ4 ≤ C3 n (n−B ( μr .

(8.5.7)

We prove the theorem by induction. For n = 1, the theorem is valid because ̺∗ (X1 , θ) ≤ τr (X1 , θ). For n = 2, 3, and 4, the estimates are similar to those for n ≥ 5, the case we are going to prove. However, the absolute constants in the bounds for n = 2, 3, and 4 have smaller values.

8.5 Proofs of the Rate of Convergence Results

157

In the following we assume n ≥ 5. Assume that for all j < n, 2 3 j    ̺∗ j −B Xi , θ ≤ K jj −B r μr + j −B τr . i=1

Since we have already estimated Δ3 and Δ4 independently of the induction hypothesis, we shall estimate only Δ1 , Δ2 , and Δ5 . To this end, take δ > 0 such that nδ = ε−1 for some ε > 0, where ε will be suitably chosen later. (III) A bound for Δ1 . By the definition of ̺∗ , 2 3 n n   ̺∗ n−B Xi , n−B θi i=2

=

2

sup ̺ tB n−B t>0

2

i=2 n 

Xi , tB n−B

i=2

n 

θi

i=2 n 

3

n−1 B ) (n − 1)−B n t>0 i=2 2 3 n n   ≤ sup ̺ uB (n − 1)−B Xi , uB (n − 1)−B θi

=

sup ̺ tB (

n−1 B ) (n − 1)−B n

u>0

= ̺∗

2

(n − 1)−B



Xi , tB (

i=2

n  i=2

Xi , (n − 1)−B

n  i=2

θi

3

n  i=2

θi

3

i=2

≤ K (n − 1)(n − 1)−B r μr + (n − 1)−B τr



by the induction hypothesis. Furthermore,     Var n−B X1 + δ B θ, n−B θ1 + δ B θ ≤ Var (nδ)−B X1 + θ, (nδ)−B θ1 + θ ≤ (nδ)−B μ1 (X1 , θ1 )

by the homogeneity of Var of order 0 and Lemma 8.4.30. Thus, we have  ( ( ( ( ( ( −B (r −B ( ( ( Δ1 ≤ K (n − 1) (n − 1) τr ((nδ)−B ( μ1 (8.5.8) μr + (n − 1)  ( ( (r ( ( ( ≤ KC4 n (n−B ( μr + (n−B ( τr ((nδ)−B ( μ1  ( ( −B (  ( B ( ( −B (r ( μr + (n ( τr (ε ( τr , ≤ KC4 n n

with C4 = max{2B r , 2B }. (IV) A bound for Δ2 .

158

8. Probabilistic-Type Limit Theorems

We assume n ≥ 5. Then, as in the case for Δ3 , we have, for j ≤ m = [ n2 ], ⎛ ⎞ n n   ̺∗ ⎝n−B Xi , n−B θi ⎠ i=j+2

i=j+2



≤ ̺∗ ⎝(n − j − 1)−B 

n 

Xi , (n − j − 1)−B

i=j+2

n 

i=j+2



θi ⎠

 ≤ K (n − j − 1)(n − j − 1)−B r μr + (n − j − 1)−B τr .

Also, we have 2

j j+1   B −B −B θi + δ B θ θi + Xj+1 ) + δ θ, n Var n ( i=1

i=1

2

≤ Var Xj+1 +

j 

θi + (nδ)B θ, θj+1 +

i=1 j B θ1

j 

3

θi + (nδ)B θ

i=1 B j θ1

3

 = Var Xj+1 + + (nδ) θ, θj+1 + + (nδ)B θ   = Var X1 + (j + nδ)B θ, θ1 + (j + nδ)B θ (by Lemma 8.4.33)   ≤ Var (j + nδ)−B X1 + θ, (j + nδ)−B θ1 + θ (by the homogeneity of Var of order 0,) ≤ (j + nδ)−B r μr (X1 , θ) by Lemma 8.4.30. Thus we have 

Δ2



m  j=1

B



K (n − j − 1)(n − j − 1)−B r μr + (n − j − 1)

−B



(8.5.9)

τr (j + nδ)−B r μr .

Now, for 1 ≤ j ≤ m = [ n2 ], n ≥ 5, we have

n−j−1 n



n−m−1 n



(n − j − 1)−B  ≤ 4B  n−B ,

1 4.

Hence

(8.5.10)

and so by (8.5.9) and (8.5.10), 

Δ2 ≤ KC5 nn

−B r

 μr + n

−B

τr

∞  j=1

(j + nδ)−B r μr ,

(8.5.11)

where C5 = max{4B r , 4B }. Furthermore, ∞  j=1

(j + nδ)−B r



∞ 0

(x + nδ)−B r dx

(8.5.12)

8.5 Proofs of the Rate of Convergence Results

=

∞

y −B r dy ≤ (nδ)(nδ)−B r



∞ 1

159

z −B r dz.

Recall that r > ΛB /λ2B ≥ 1/λB . Take q such that 1/λB < q < r. Then, by 1 1 (8.4.5), x q I−B  → 0 as x → ∞, and hence M1 := supx≥1 x q I−B  < ∞. Thus for z ≥ 1, z −B r ≤ M1r z −r/q , and hence ∞ 1

∞ z −B r dz ≤ M1 z −r/q dz =: C6 < ∞.

(8.5.13)

1

It also follows from the assumption for p (p > 1/λB ) that for z ≥ 1, 1 z p ≤ (M2 z −B −1 ≤ M2 z B  since z B  z −B  ≥ 1, where M2 = ( ( ( 1 supx≥1 (x p I−B (. Thus if nδ ≥ 1, then nδ ≤ M2p (nδ)B p .

(8.5.14)

Finally, we have by (8.5.11)–(8.5.14) and (8.4.5) that if nδ = ε−1 (where ε > 0 will be taken small), then ( (r   (8.5.15) Δ2 ≤ KC5 nn−B r μr + n−B τr C6 (nδ) ((nδ)−B ( μr ( ( ( (   p r ≤ KC5 C6 nn−B r μr + n−B τr M2p (ε−B ( (εB ( μr  ( ( −B (  ( −B (p ( B (r r−p ( p −B (r ( ≤ KC5 C6 M2 n n μr + (n ( τr (ε ( (ε ( τr . (V) A bound for Δ5 .

We have Δ5 ≤ C2 n−B  ε−B  = C2 n−B τr ε−B τr−1 .

(8.5.16)

Altogether we have from (8.5.3), (8.5.6), (8.5.7), (8.5.8), (8.5.15), and (8.5.16) that 2 3 n  ̺ n−B Xi , θ (8.5.17) i=1

( B (  ( −B (r ( −B (  ( ( ( ( μr + (n ( τr ≤ 2C1 C3 nn  μr + KC1 C4 ε τr n n  ( ( −B (  ( ( ( ( ( p ( −B (p ( B (r r−p −B (r ( ε μr + (n ( τr τr n n + KC1 C5 C6 M2 ε ( ( ( ( + C2 (ε−B ( τr−1 (n−B ( τr &  ( ( ( B (r  ( B( p( B −1 (p ( ( ( τr ε ( ≤ K C1 C4 τr ε + C1 C5 C6 M2 (τr ε ) ( (  ( (  ( −B (r  n (n ( μr + (n−B ( τr . + 2C1 C3 + C2 ((τr εB )−1 ( −B r

160

8. Probabilistic-Type Limit Theorems

We first show that ( (p ( (r lim (t−B ( (tB ( = 0. t→0

(8.5.18)

It follows from and (8.4.5) that for any η > 0 and for some small ( −B ( (8.4.4) −Λ t0 > 0, (t ( ≤ t B −η , t < t0 , and tB  ≤ tλB −η , t < t0 . Thus for t < t0 , ( −B (p ( B (r (t ( (t ( ≤ t−(ΛB +η)p+(λB −η)r ≤ t−ΛB p+λB r−η(p+r) ,

where by the restrictions on r and p, −ΛB p + λB r > 0. Thus, taking η > 0 sufficiently small, we get (8.5.18). Also, of course, limt→0 tB  = 0. Therefore, we can find a sufficiently small ε > 0 such that the matrix τr εB satisfies ( ( B(  (p ( B (r 1 p ( B −1 ( ( ( ( . (8.5.19) C1 C4 τr ε + C1 C5 C6 M2 ( τr ε ( τr ε ( ≤ 2

Then choose K such that

( ( K 2C1 C3 + C2 ((τr εB )−1 ( ≤ . 2

(8.5.20)

Finally, we obtain from (8.5.17), (8.5.19), and (8.5.20) that 2 3 n  (  ( ( −B (  −B −B (r ( ̺ n Xi , θ ≤ K n n μr + (n ( τr . i=1

2

Proof of Theorems 8.4.8 and 8.4.9: By the same reasoning as that mentioned at the beginning of this section, we can assume that {Xi } and θ are independent of each other and that the {θi } are independent copies of θ and are independent of the {Xi } and θ. We prove the theorem by induction. For n = 1, the assertion is trivial. For n = 2, we have     Var 2−B (X1 + X2 ), θ = Var 2−B (X1 + X2 ), 2−B (θ1 + θ2 ) [by (8.5.1)]

≤ Var(X1 + X2 , θ1 + θ2 ) [by the homogeneity of Var of order r = 0] ≤ Var(X1 + X2 , X1 + θ2 ) + Var(X1 + θ2 , θ1 + θ2 ) [by the triangle inequality] ≤ Var(X2 , θ2 ) + Var(X1 , θ1 ) [by the regularity of Var]

( (r = 2 Var (X1 , θ) ≤ 2νr ≤ 2c (2−B ( νr ,

8.5 Proofs of the Rate of Convergence Results

161

since ( (r ( (r  ( ( (r ( (r (r ( −B (r ( ( c 2 = (2B ( + (3B ( (2−B ( ≥ (2B ( (2−B ( ≥ 1.

For n = 3, we similarly have ( (r   Var 3−B (X1 + X2 + X3 ), θ ≤ 3c (3−B ( νr .

Now suppose that for all j < n, 3 2 j  ( (r Var j −B Xi , θ ≤ cj (j −B ( νr .

(8.5.21)

i=1

Then for any j < n, 2 3 j  ( 1( (2−B (r Var j −B Xi , θ ≤ cM νr ≤ ca = b i=1

(8.5.22)

by our assumptions.

For any integer n ≥ 4 and m = [ n2 ], we have 2 3 n  Var n−B Xi , θ i=1

2

= Var n−B 2

≤ Var n−B 2

n 

i=1 n 

Xi , n−B

n 

θi

i=1

Xi , n−B

i=1

2m 

3

+ Var n−B (

θi +

i=1

n 

n 

θi +

i=1

m 

≤ Var n−B 2

n 

Xi ), n−B

i=m+1

i=m+1

+ Var n−B 2

Xi , n−B

2m  i=1

m 

+ Var n−B (

θi +

i=1

[by Lemma 8.4.26]

n 

n 

i=m+1 n 

i=m+1

n 

33

θi

i=1

θi

i=m+1

Xi +

Xi

i=m+1

[by the triangle inequality] 2

(8.5.23)

θi

3

3

3

2

Var n−B

, n−B

Xi ), n−B

i=1

n  i=1

n  i=1

m 

θi

θi

3

3

Xi , n−B

m  i=1

θi

3

162

8. Probabilistic-Type Limit Theorems

=: I1 + I2 + I3 . By the induction hypotheses (8.5.21) and (8.5.22), 2  −B −B 3 n  n n I1 = Var Xi , (n − m)−B θ1 n−m n − m i=m+1 2 3 n  n −B  n −B  × Var Xi , m−B θ1 m m i=m+1 2 3 3 2 m n   ≤ Var (n − m)−B Xi , θ1 Xi , θ1 Var m−B i=1

i=m+1



( ( ( 1( (2−B (r cm (m−B (r νr . b

Note that m = [ n2 ] ≥ 52 n for n ≥ 4. Hence (  (r (  (r ( 2 −B ( ( 2 −B ( ( ( −B (r n n r ( ( ( ( ( −B ( (m ( ≤ ( ≤ ( n ( . ( n ( ( ( 2( 5 2( 5 Thus I1

(  (r −B ( (r ( (r ( 1 ( −B (r 1 ( 2 ( ( ( ( 2 2 ≤ c ( cn (n−B ( νr (8.5.24) ( n (n−B ( νr = ( b 2( 5 5

by the definition of b. To estimate I2 , observe that 2 n m   n −B −B −B ) (n − m) θi , Xi + ( I2 = Var n n − m i=m+1 i=1 3   n m −B   n n−B (n − m)−B θi θi + n − m i=m+1 i=1 3 2 m m   −B −B Xi + θ, (n − m) θi + θ . ≤ Var (n − m) i=1

Then we have I2

( (r ≤ ((n − m)−B ( μr

2m  i=1

i=1

Xi ,

m  i=1

θi

3

[by Lemma 8.4.30]

( (r ≤ ((n − m)−B ( mμr (X1 , θ) [by the triangle inequality and the repeated use of the regularity of μr ] (  (r −B ( (r n−m 1( 1 ( 1 ( ( ≤ ≥ . Hence ( ( n (n−B ( μr since ( 2( 2 n 2

8.5 Proofs of the Rate of Convergence Results

I2 ≤

( ( ( 1( (2B (r n (n−B (r νr . 2

163

(8.5.25)

As to I3 , we have 2 3 n n   Xi + n−B mB θ, n−B θi + n−B mB θ I3 = Var n−B 2

i=m+1

≤ Var m−B

n 

i=m+1

Xi + θ, m−B

i=m+1

n 

θi + θ

i=m+1

3

(8.5.26)

( ( ( −B (r ( −B (r ( n B (r ( ≤ (m ( (n − m)μr (X1 , θ) ≤ (n ( ( ( m ( (n − m)μr ( ( ( 3( (3B (r n (n−B (r νr , ≤ 5

since

n m

≤ 3 for n ≥ 4 and n − m ≤ 35 n.

Altogether, we have from (8.5.23)–(8.5.26), 2 3   n  ( (r 2 3 1 Var n−B Xi , θ ≤ c + 2B r + 3B r n (n−B ( νr 5 2 5 i=1 ≤ cnn−B r νr .

This completes the proof of Theorem 8.4.8. The proof of Theorem 8.4.9 is similar and is therefore omitted. 2 Proof of Theorem 8.4.10: Again, we assume that the {Xi } and θ are independent of each other. Let {θi } be independent copies of θ, and assume that the {θi } are independent of {Xi } and θ. We prove the theorem by induction. For n = 1, d(X1 , θ) ≤ Ad(X1 , θ) ≤ ATr . For n = 2, we have by Lemma 8.4.31, the regularity of d, and the triangle inequality,     d 2−B (X1 + X2 ), θ = d 2−B (X1 + X2 ), 2−B (θ1 + θ2 ) = |J2B | d(X1 + X2 , θ1 + θ2 ) ≤ 2 |J2B | d(X1 , θ) ≤ 2C(d, B)Tr [by Lemma 8.4.32] ≤ 2A2−B r Tr , ( ( (r ( (r (r since A (2−B ( ≥ C(d, B) (2B ( (2−B ( ≥ C(d, B). Similarly, we have for n = 3, (r (   d 3−B (X1 + X2 + X3 ), θ ≤ 3C(d, B)Tr ≤ 3A (3−B ( Tr ,

164

8. Probabilistic-Type Limit Theorems

( (r ( (r ( (r since A (3−B ( ≥ C(d, B) (3B ( (3−B ( ≥ C(d, B).

To prove the theorem by induction, assume for all j < n that 2 3 j  ( (r d j −B Xj , θ ≤ Aj (j −B ( Tr . i=1

For any n ≥ 4 and m = [ n2 ], we have by Lemma 8.4.28, 2

d n−B ≤

n 

Xi , θ

i=1

2

3

n 

d n−B

i=1

2

+ d n−B ≤

2

d n−B 2

Xi , n−B

2m 

2m 

m  i=1 −B

m  i=1

Xi +

+ d n−B

Xi

3

3

i=1

n 

θi

n 

Xi

3

3

,n

33

n 

θi

i=1

θi Var n−B

i=m+1

=: I1 + I2 + I3 .

Xi

, n−B

2

i=m+1

θi +

n 

i=m+1

i=m+1

i=1

2m 

n 

θi +

Xi , n−B 2m 

θi +

i=1

i=1

+d n 2

(8.5.27)

n 

3

Xi , n−B

i=m+1 −B

n 

θi

i=1

, n−B

n  i=1

θi

3

n 

i=m+1

3

θi

3

By Lemma 8.4.31, 3 2 3 2 m n  n −B  n −B   m−B θ Var (n − m)−B Xi , Xi , θ I1 ≤ d m m i=1 i=m+1 2 3 2 3 m n      n  −B Xi , θ Var (n − m)−B Xi , θ . (8.5.28) = J( m )B  d m i=1

i=m+1

By the induction hypothesis and Lemma 8.4.32, 2 3 m    (r (  n  Xi , θ ≤ C(d, B)Am (m−B ( Tr J( m )B  d m−B i=1

(8.5.29)

(  (r ( m −B ( ( n( r ( Tr ≤ C(d, B)A (n−B ( ( ( ( 2 n (r 1 ( (r ( ≤ C(d, B)A (3B ( n (n−B ( Tr . 2

8.5 Proofs of the Rate of Convergence Results a , by Theorem 8.4.8, On the other hand, since νr ≤ M 2 3 n  ( (r Var (n − m)−B Xi , θ ≤ c(n − m) ((n − m)−B ( νr i=m+1

where we have used νr ≤

1 , D

1 M cD .

Therefore, we have, by (8.5.28)–(8.5.30),

(r (r ( 1 ( (r 1 ( 1 I1 ≤ C(d, B)A (3B ( n (n−B ( Tr ≤ An (n−B ( Tr , 2 D 2 ( B (r since D ≥ C(d, B) (3 ( . Similarly, for the estimate of I2 , I2

(8.5.30)

(r ( 1 ( ( I−B = c ((n − m) r ( νr ≤ cM νr ≤

2

165

(8.5.31)

 −B −B 3 m  n n Xi + = d n−B θi + θ, n−B θ n − m n − m i=1 i=1 2 3 m m       −B n Xi + θ, (n − m)−B θi + θ = J( n−m )B  d (n − m) m 



i=1

i=1

[by Lemma 8.4.31]

≤ C(d, B)(n − m)−B r dr

2m  i=1

Xi ,

m  i=1

θi

3

[by Lemma 8.4.32]

( (r ≤ C(d, B) ((n − m)−B ( mdr (X1 , θ) by the triangle inequality and the repeated use of the regularity of dr . Hence (r 1 ( (r ( I2 ≤ C(d, B) (2B ( n (n−B ( Tr . 2

(8.5.32)

Finally, we have 3 2 n n  n −B  n −B   I3 = d θ + n−B θ + n−B Xi , θi m m i=m+1 i=m+1 2 3 n n   ≤ C(d, B)d m−B Xi + θ, m−B θi + θ i=m+1

( (r ≤ C(d, B) (m−B ( (n − m)dr (X1 , θ) (r 3 ( (r ( ≤ C(d, B) (3B ( n (n−B ( Tr . 5

i=m+1

(8.5.33)

166

8. Probabilistic-Type Limit Theorems

Combining the estimates for I1 , I2 , and I3 , we finally obtain from (8.5.27) and (8.5.31)–(8.5.33) that 2 3 n  Xi , θ d n−B ≤ I1 + I 2 + I3 ≤



i=1

 ( B (r 3 ( B (r ( (r 1 1 A + C(d, B) (2 ( + C(d, B) (3 ( n (n−B ( Tr 2 2 5

≤ Ann−B r Tr by the definition of A.

2

Proof of Theorem 8.4.13: We apply the “minimality” property of the 5 Lp -metric : 3 2 2 3 n n n    i , n−B 5 p n−B Xi , θ ≤ inf Lp n−B X L θi , (8.5.34) i=1

i=1

i=1

where the infimum is taken over all independent identically distributed d d d  with fixed marginals X  θ) ∼ i , θi ) ∼ pairs (X (X, X and θ ∼ θ. The righthand side in (8.5.34) is less than or equal to ' &( (p ( (p d d  X  θ); ∼ 5 p (X, θ). inf (n−B ( nLp (X, X, θ ∼ θ = n (n−B ( L The bound for π follows from Lemma 8.4.33.

2

Proof of Theorem 8.4.14: The proof is similar to that of Theorem 8.4.10 and is therefore omitted. 2 Proof of Theorem 8.4.15: The proof resembles that of Theorem 8.4.4 5 p in with the replacement of the smoothing inequality for ̺∗ by that for L Lemma 8.4.27 and hence is omitted. 2 Proof of Theorem 8.4.16: Using the Marcinkiewicz–Zygmund inequality (see for example Kawata (1972, Theorem 13.6.1)), if 1 < p ≤ 2 and {ξi }i≥1 are independent random vectors with E[ξ] = 0, then *( n (p + n ( (  ( ( E ( ≤ Cp E[ξi p ], (8.5.35) ξi ( ( ( i=1

i=1

i − θi ), where {(X i − θi )}i≥1 are i.i.d. for some Cp > 0. Take ξi = n−B (X d d i ∼ X and θi ∼ θ. Thus we get pairs, X (p + *( n n ( ( (p  (   ( −B ( ( −B  (    ≤ Cp E (n (Xi − θi )( (Xi − θi )( E (n ( ( i=1

i=1

8.5 Proofs of the Rate of Convergence Results

167

(  ( −B (p ( (  (p ( ( = Cp n n E (X − θ( .

Passing to the minimal metrics gives the necessary inequality. Finally, note Λ that p > λ2B implies nn−B p → 0 as n → ∞. The bound for the Prohorov B

p

5 pp+1 for p ≥ 1. metric π comes from the inequality π ≤ L

2

5 p to get Proof of Theorem 8.4.19: We use the minimality property of L 'p & 5 p (M (n), M (n)) L X θ 2 k 3 2 k 33p  2 n n @  @  5 p n−B = Xi , n−B θi L k=1

i=1

k=1

i=1

*( 2 k 3 2 k 3(p + n n ( ( @  @  ( ( ≤ E (n−B Xi − n−B θi ( ( ( k=1 i=1 k=1 i=1 2 k 3(p + *( n 2 k 3 n ( (@  @  ( ( −B p ≤ n  E ( θi ( Xi − ( ( k=1 i=1 k=1 i=1 (p + * n ( k k (   @ ( ( (  −B p ≤ Cp,d n  E Xi − θi ( ( ( ( k=1

i=1

i=1

0

  }1/p  · 0 for some Cp,d > 0 and (here we have used  ·  ≤ {Cp,d ( ( > >( > (  xk − yk 0 ≤ xk −yk 0 )

 ≤ Cp,d n−B p

n 

k=1

(p + *( k ( ( ( ( E ( (Xi − θi )( . ( ( i=1

0

Since E[X − θ] = 0 and 1 < p ≤ 2, by (8.5.35) again the above is, for some other constant Cp,d > 0, k n   ( −B (p p ( ( n ≤ E [X − θ ] Cp,d k=1 i=1

= Cp,d

( n(n + 1) ( (n−B (p E [X − θp ] . 2

Passing to the minimal metric gives the necessary bound for 5 p (M (n), M (n)). L X θ

168

8. Probabilistic-Type Limit Theorems

5 p (m (n), m (n)). Further, The same argument leads to the bound for L X θ

& 'p 5 Lp (aX (n), aθ (n)

( k ( p + ( * n ( k n ( (  @ ( ( @ ( −B (p ( ( ( (  ( ( n ≤ E  θi ( Xi ( − ( ( ( ( ( (  k=1 i=1 k=1 i=1 ( (p + ( * n ( k k ( ( (  @ ( ( ( ( ( −B p ≤ n  E θ i ( Xi ( − ( ( ( ( ( ( i=1 i=1 k=1 (p + * n ( k k ( @(   ( ( −B p ≤ n  E Xi − θi ( ( ( ( k=1

≤ Cp

i=1

i=1

( n(n + 1) ( (n−B (p E [X − θp ] 2

as before. Combining our estimates, we complete the proof of the theorem. 2 Proof of Theorem 8.4.21: From the definition of the Zolotarev metric ζr and its ideality of order r, we get the first bound: 2 2 3 3 n n n    ζr n−B Xi , θ ≤ ζr n−B Xi , n−B θi i=1

i=1

( (r ≤ (n−B ( ζr

2n 

Xi ,

i=1

i=1 3 n 

θi

i=1

( (r ≤ n (n−B ( ζr (X1 , θ) .

Applying the universal bound for the Prohorov metric π by ζr , π r+1 ≤ Crr+1 ζr

on X (IRd ) for some Cr > 0 (cf. Zolotarev (1983)), we obtain the final estimate. 2 d i − θi ), where X i ∼ Proof of Theorem 8.4.22: Let ξi = n−B (X X and d i , θi ) are i.i.d. Then by the Rosenthal inequality (see, for θi ∼ θ, and (X example, Araujo and Gin´e (1980, p. 205)), for 2 < p < ∞, ⎧2  *( n (p +1/p 31/p2 n 31/2 ⎫ n ( ( ⎨  ⎬  ( ( p 2 E ( ξi ( ≤ C(d, p) max E[ξi  ] , E[ξi  ] , ( ( ⎩ ⎭ i=1

i=1

for some C(d, p) > 0. Then 2 3 n n   i , n−B X θi Lp n−B i=1

i=1

i=1

8.5 Proofs of the Rate of Convergence Results

≤ C(d, p) max

⎧2 n ⎨  ⎩

i=1

i − θi )p ] E[n−B (X 2 n  i=1

169

31/p ,

31/2 ⎫ ⎬ −B  2  E[n (Xi − θi ) ] ⎭

 1/2  1/p  ( (2 ( −B (p 2 −B p    − θ ]  − θ ] , n(n ( E[X ≤ C(d, p) max n(n ( E[X & ' ( ( ( ( 1/2 ( −B ( 1/p ( −B (     L2 (X, θ) ≤ C(d, p) max n n Lp (X, θ), n n ( (   θ), ≤ C(d, p)n1/2 (n−B ( Lp (X,

since L2 ≤ Lp and n1/p ≤ n1/2 for p > 2. The case p = ∞ is similar.

2

In Theorems 8.4.4, 8.4.8, 8.4.9, 8.4.10, and 8.4.14 we have assumed the 5 p,r . Since χr ≤ μr , the natural finiteness of the metrics μr , χr , dr , and L 5 p,r , which may question is how we can assure the finiteness of μr , dr and L not be easily checked just by a direct use of the definitions. The rest of this 5 p,r , section is devoted to the construction of upper bounds for μr , dr , and L where the metrics used in the upper bounds are more familiar distances in the literature. We shall construct bounds from above for μr , dp , and Lp,r , using the Zolotarev ζr -metric. Define the following probability metrics: For X, Y ∈ X (IRd ), μr (X, Y )

=

dr (X, Y )

=

sup T ∈Aut(IRd )

sup T ∈Aut(IRd )

T −r Var(T X + θ, T Y + θ), T −r d(T X + θ, T Y + θ),

and 5 (X, Y ) = L p,r

sup T ∈Aut(IRd )

T −r Lp (T X + θ, T Y + θ).

5 . In the next two theorems we 5 p,r ≤ L Clearly, μr ≤ μr , dr ≤ dr , L p,r 5 are going to estimate μr , dr , and Lp,r from above by ζr . Let pθ (x) be the density function of the strictly operator-stable random vector θ ∈ IRd . For m ∈ IN let  (m) Cm (θ) := sup |pθ (x)(h)m | dx h =1 IRd

170

8. Probabilistic-Type Limit Theorems

and Dm (θ) :=

(m)

sup sup |pθ

x∈IRd h =1

(x)(h)m |.

Theorem 8.5.1 (i) For m ∈ IN, μm (X, Y ) ≤ Cm (θ)ζm (X, Y ). (ii) If r = m + p, m ∈ IN, 0 < p ≤ 1, then 5 (X, Y ) ≤ C (θ)ζ (X, Y ). L p,r m r

Theorem 8.5.2

dm (X, Y ) ≤ Dm (θ)ζm (X, Y ),

m = 1, 2, . . . .

Proof of Theorem 8.5.1: (i) For any X and Y , Var(X + θ, Y + θ) =

sup A∈B(IRd )

|P (X + θ ∈ A) − P (Y + θ ∈ A)|

(8.5.36)

sup{|E[f (X + θ) − f (Y + θ)]|; f ∈ F, f ∞ ≤ 1} = sup{|E[g(X) − g(Y )]|; f ∈ F, f ∞ ≤ 1},

=

where g(x) := E[f (x + θ)]. Since pθ (x) is differentiable infinitely many times (see Hudson (1980)), g(x) =



f (x + y)pθ (y) dy =



f (z)pθ (z − x) dz

IRd

IRd

has derivatives of every order, and furthermore,       (m) |g (m) (x)(h)m | =  f (z)pθ (z − x)(h)m dz   d  IR       (m) =  f (x + y)pθ (y)(h)m dy  .  d  IR

Since for f with f ∞ ≤ 1, sup sup |g (m) (x)(h)m | ≤ Cm (θ), we have x∈IRd h =1

g (m−1) (x) − g (m−1) (y) ≤ Cm (θ)x − y.

(8.5.37)

8.5 Proofs of the Rate of Convergence Results

171

Hence by (8.5.36) and (8.5.37), Var(X + θ, Y + θ) ( ( & ( ( (m−1) (m−1) ≤ sup |E [g(X) − g(Y )]| ; (g (x) − g (y)(

' ≤ Cm (θ) x − y

and μm (X, Y ) =

sup T ∈Aut(IRd )

T −m Var(T X + θ, T Y + θ)

−m

≤ sup T  T

(8.5.38)

&

sup |E [g(T X) − g(T Y )]| ; ( ( ' ( (m−1) ( (m−1) (x) − g (y)( ≤ Cm (θ) x − y . (g

Let gT (x) := g(T x). Then

gT(m−1) (x)(h)m−1 = g (m−1) (T x)(T h)m−1

for any x, h ∈ IRd ,

implying that ( ( ( ( ( (m−1) ( ( m−1 ( (m−1) (m−1) (m−1) (x) − gT (y)( ≤ T  (T x) − g (T y)( . (gT (g Then the side condition in (8.5.38), ( ( ( (m−1) ( (m−1) (x) − g (y)( ≤ Cm (θ)x − y, (g

results in ( ( ( (m−1) ( (m−1) (y)( ≤ Cm (θ)T m−1 T x − T y (x) − gT (gT ≤ Cm (θ)T m x − y.

Consequently, by (8.5.38), −m

μm (X, Y ) ≤ sup T  T

& sup |E[gT (X) − gT (Y )]|;

(m−1) gT (x)

= Cm (θ)ζm (X, Y ),



(m−1) (y) gT

' ≤ Cm (θ)T  x − y m

as desired. 5 . Let r = m + p, m ∈ IN, (ii) Let us now prove a similar bound for L p,r 0 < p ≤ 1. Then by Lemma 8.4.34 (ii), 5 p (X + θ, Y + θ) = sup {|E [f (X + θ) − f (Y + θ)]| ; f ∈ Lipb (p)} L = sup {|E [g(X) − g(Y )]| ; f ∈ Lipb (p)} ,

172

8. Probabilistic-Type Limit Theorems

where g(x) := E[f (x + θ)]. Since pθ (x) is differentiable infinitely many  times, the function g(x) = f (x + y)pθ (y) dy = f (z)pθ (z − x) dz has IRd

IRd

derivatives of all orders, and for m ∈ IN, r = m + p,

      (m) (m) m m |g (x)(h) | =  f (x + y)pθ (y)(h) dy  .  d  IR

By the requirement for f , ( ( ( (m) ( (m) (g (x) − g (y)( =

   (m) m (m) m sup g (x)(h) − g (y)(h) 

h =1

      (m) m = sup  [f (x + z) − f (y + z)] pθ (z)(h) dz  h =1   IRd    p  (m) m ≤ sup x − y pθ (z)(h)  dz h =1 IRd p

≤ x − y Cm (θ).

& ( ( 5 Therefore Lp (X + θ, Y + θ) ≤ sup |E [g(X) − g(Y )]|; (g (m) (x) − g (m) (y)( ' d p ≤ Cm (θ)x − y for any x, y ∈ IR . 5 (X, Y ): Next consider L p

5 (X, Y ) L p,r

=

sup

T ∈Aut(IRd )

T −r Lp (T X + θ, T Y + θ)

& ≤ sup T  sup |E[g(T X) − g(T Y )]|; T ( ( ' ( (m) ( (m) p (g (x) − g (y)( ≤ Cr (θ)x − y . −r

Let gT (x) := g(T x). Then for all x, h ∈ IRd , gT(m) (x)(h)m = g (m) (T x)(T h)m for any x, h ∈ IRd , implying that ( ( ( ( ( (m) ( ( (m) m ( (m) (m) (gT (x) − gT (y)( = T  (g (T x) − g (T y)( .

Applying g (m) (x) − g (m) (y) ≤ Cm (θ)x − yp , we get that ( ( ( (m) ( (m) (gT (x) − gT (y)( ≤ Cm (θ)T m T x − T yp ≤ Cm (θ)T m+p x − yp ,

8.5 Proofs of the Rate of Convergence Results

173

and m + p = r by assumption. Similarly, we get 5 (X, Y ) ≤ sup T −r sup {|E[g (X) − g (Y )]|; L p,r T T T ( ( ' ( (m) ( (m) r p (gT (x) − gT (y)( ≤ Cm (θ)T  x − y ≤ Cm (θ)ζr (X, Y ).

2

Proof of Theorem 8.5.2: We have dm (X, Y ) =

(8.5.39)

sup T ∈Aut(IRd )

T −m d(T X + θ, T Y + θ)

sup T −m sup |pT X+θ (x) − pT Y +θ (x)| T x∈IRd       = sup T −m sup  pθ (x − y)[P (T X ∈ dy) − P (T Y ∈ dy)] . T x∈IR   d

=

IR

Let

g(y) = pθ (x − y).

(8.5.40)

Then      (m)  (m) m m sup sup sup g (y)(h)  ≤ sup pθ (x − y)(h)  = Dm (θ). y,x,h

y∈IRd x∈IRd h =1

Hence

( ( ( (m−1) ( (m−1) (x) − g (y)( ≤ Dm (θ)x − y. (g

(8.5.41)

From by (8.5.39)–(8.5.41) we have dm (X, Y )

& ≤ sup T −m sup |E[g(T X) − g(T Y )]|; T ( ( ' ( ( (m−1) (m−1) (x) − g (y)( ≤ Dm (θ)x − y . (g

This upper bound is the same as that of μm (X, Y ) in (8.5.38) if Cm (θ) is replaced by Dm (θ). Therefore the proof of Theorem 8.5.1 also implies that dm (X, Y ) ≤ Dm (θ)ζm (X, Y ).

2

The next question is the finiteness of Cm (θ) and Dm (θ). Theorem 8.5.3 We have Dm (θ) < ∞, m = 1, 2, . . . , and if ΛB < 1, then Cm (θ) < ∞, m = 1, 2, . . . .

174

8. Probabilistic-Type Limit Theorems

Proof: We first show the finiteness of Dm (θ). Let μ 5(z), z ∈ IRd , be the characteristic function of θ. As was shown in Hudson (1980), for some c > 0, (8.5.42) |5 μ(z)| ≤ exp{−cz1/ B } for every z with z > 1.  zm |5 μ(z)| dz < ∞, implying the exisHence, for every m = 1, 2, . . . , IRd

(m)

tence of pθ (x), and furthermore the finiteness of Dm (θ). To prove the finiteness of Cm (θ), we assume ΛB < 1, which implies E[θ] < ∞. (See Hudson et al. (1988).) We start with Carlson’s inequality for one-variable functions f . If f5, the Fourier transform of f , is in L2 (IR), and f5 exists and is in L2 (IR), then ⎞ ⎛ ∞ ⎞⎛ ∞ ⎛ ∞ ⎞4    ⎝ |f (x)| dx⎠ ≤ K ⎝ f5(z)2 dz ⎠ ⎝ f5 (z)2 dz ⎠. (8.5.43) −∞

−∞

−∞

A version of this inequality for several variable functions f is, for each h ∈ IRd , ⎛ ⎞4 ⎛ ⎞⎛ ⎞    ⎝ |f (x)| dx⎠ ≤ K ⎝ f5(z)2 dz ⎠ ⎝ (Df5(z)h)2 dz ⎠, (8.5.44) IRd

IRd

IRd

where Df5(z) is the gradient (row) vector of f5(z). The proof of (8.5.44) can be carried out in the same manner as for (8.5.43), so we omit it. Since we are assuming E[θ] < ∞, μ 5(z) is differentiable. Fix h ∈ IRd and apply (8.5.44) to f (x) := p(m) (x)(h)m , x ∈ IRd . θ  Then f5(z) = eiz,x p(m) (x)(h)m dx, and θ

IRd

 Df5(z)h = i x, heiz,x pθ(m) (x)(h)m dx. IRd

Thus



f5(z)2 dz ≤

IRd

sup



h =1 IRd



z2m |5 μ(z)|2 dz < ∞ by (8.5.42), and

IRd

Df5(z)h

2

dz ≤



z2m |D5 μ(z)|2 dz =: I.

IRd

So it remains to show the finiteness of I. We recall that the characteristic function μ 5(z) is given by 

μ 5(z)

=

exp iz, c + z, Az

8.5 Proofs of the Rate of Convergence Results

+



γ( dx)

∞  0

S

175

  B 1 eiz,s x − 1 − iz, sB xIQ (sB x) 2 ds . s

Here z ∈ IRd , c ∈ IRd , A is a nonnegative definite symmetric matrix, S = {x ∈ IRd ; x = 1 and tB x > 1 for all t > 1}, Q = {x ∈ IRd ; x ≤ 1}, γ is a probability measure on S. We write μ 5(z) = eψ(z) . Since ΛB < 1,   sB x M1 := γ( dx) ds < ∞. (8.5.45) s2 S

{ sB x >1}

Note that if γ(S) > 0, the non-Gaussian part exists, and the restriction of B to the support of the measure γ (we shall call it B again) satisfies λB > 12 . (See Hudson and Mason (1981).) Hence we also have   sB x2 M2 := γ( dx) ds < ∞. (8.5.46) s2 S

{ sB x ≤1}

We have, for h ∈ IRd , D5 μ(z)h = μ 5(z)Dψ(z)h.

Let z = (z1 , . . . , zd )t , c = (c1 , . . . , cd )t, A = (aij ), sBx =  B  t s x d , and h = (h1 , . . . , hd )t . Then ∂ ∂zj ψ(z)



 sB x 1 , . . . ,

= icj + 2(Az)j

 ∞  B  1  B   B  iz,sB x ds. − i s x j IQ s x + γ( dx) i s x j e s2 S

0

Thus ∞       ∂  B    i  B  1 − IQ (s x) 2 ds  ∂zj ψ(z) ≤ |cj | + 2Az + γ( dx)  s x j  e s 0 S   ( B ( 1 (s x( ≤ c + 2Az + γ( dx) ds s2 +

 S

S

γ( dx)



{ sB x >1}

 1 ( B (  i  (s x( e − 1 2 ds s

{ sB x ≤1}

≤ c + 2Az + M1 + M2 z,

where M1 and M2 are finite by (8.5.45) and (8.5.46). We thus finally have    d     ∂  ψ(z)h |Dψ(z)h| =  j ∂zj   j=1 

176

8. Probabilistic-Type Limit Theorems

≤ dh (c + 2Az + M1 + M2 z) ≤ C1 + C2 z,

and |D5 μ(z)h| ≤ (C1 + C2 z) |5 μ(z)| . Hence by (8.5.42) we conclude that  z2m |D5 μ(z)h|2 dz < ∞. I = IRd

The proof of Theorem 8.5.3 is now complete.

2

The final question is the finiteness of ζm (X1 , θ). As we have noted in (8.4.32), ζm (X1 , θ) ≤

1 κm (X1 , θ), m!

m = 1, 2, . . . ,

where κm (X1 , θ) is the difference pseudomoment, namely κm (X1 , θ)

=

sup {|E[f (X1 ) − f (θ)]|; |f (x) − f (y)| ≤ dm (x, y) ' d for any x, y ∈ IR ,

( ( ( m−1 m−1 ( − y y where dm (x, y) = (x x (.

It would be difficult to check conditions implying κm (X1 , θ) < ∞. Instead we give an example for laws of X1 with ζm (X1 , θ) < ∞. The idea is to use the series representation for θ (compare Remark 8.1.5). Let {Wj }∞ j=1 be a sequence of i.i.d. random variables taking their values in S = {x ∈ IRd ; x = 1} with a common probability distribution λ on S and Γj = δ1 +· · ·+δj . Here {δj } is a sequence of independent exponentially distributed random variables with E[δ1 ] = 1 that are independent of {Wj }. It is known that the series ∞    −B   Γj Wj − E Γ−B j I[Γj ≥ 1] E[Wj ] j=1

converges almost surely and is distributed as an operator-stable random vector with exponent B. Suppose E[Wj ] = 0 for all j, and set d

θ =

∞  j=1

Γ−B j Wj .

8.5 Proofs of the Rate of Convergence Results

177

Let r > 1/λB and j0 = [rλB ]. Then it is easy to show that if j > j0 , then r E[Γ−B j  ] < ∞.

Consider another sequence of independent random variables {Vj } on S, which are independent of {Γj }, such that d

for j ≤ j0 ,

Vj = Wj

and for j > j0 , {Vj } are arbitrary but not identically distributed random variables on S. Define d

X =

∞ 

Γ−B j Vj ,

j=1

assuming that the series converges almost surely. Theorem 8.5.4 Suppose that all mixed moments of order less than or equal to r − 1 coincide; that is,   E V1α1 · · · Vpαp − W1α1 · · · Wpαp = 0 (8.5.47) p for any αi ≥ 0 with i=1 αi ≤ r − 1. Then ζr (X, θ) < ∞.

Proof: By the operator-ideality of ζr of order r and the triangle inequality, ⎞ ⎛   ⎠ ζr (X, θ) = ζr ⎝ Γ−B V , Γ−B j j j Wj j



∞ 

j

−B ζr (Γ−B j Vj , Γj Wj )

j=1

=

∞ 

−B ζr (Γ−B j Vj , Γj Wj )

j=j0 +1

(by Zolotarev (1983, Property 4 on p. 293)) ≤

∞ 

j=j0 +1

r E[Γ−B j  ]ζr (Vj , Wj )

≤ sup ζr (Vj , Wj ) j

∞ 

j=j0 +1

  r E Γ−B j  .

Since rλB > 1, the final series converges. This is shown as follows. Note that for any ε > 0, there exists C > 0 such that x−B  < Cx−(ΛB +ε) ,

0 < x ≤ 1,

178

8. Probabilistic-Type Limit Theorems

and x−B  < Cx−(λB −ε) ,

x > 1.

Thus we have for j > j0 = [rλB ], −B r −B r r E[Γ−B j  ] = E[Γj  I[Γj ≤ 1]] + E[Γj  I[Γj > 1]] −r(ΛB +ε)

≤ E[Γj

−r(λB −ε)

] + E[Γj

].

On the other hand, if j > p, then E[Γ−p j ] =

Γ(j − p) ∼ j −p . Γ(j)

Hence r −r(Λ+ε) E[Γ−B + j −r(λB +ε) ) ≤ Cj −r(λB +ε) ) j  ] ≤ C(j

for large j, because ΛB ≥ λB . This shows the convergence of the series in question. Also, as we have noted in (8.4.34), under (8.5.47), ζr (Vj , Wj ) ≤

Γ(1 + α) ξr (Vj , Wj ), Γ(1 + r)

where r = m + α, m ∈ IN, α ∈ (1, 2]. In our case, Vj and Wj take their values on the unit sphere, and therefore, ξr (Vj , Wj ) = Var(Vj , Wj ) ≤ 1. This concludes the proof of the theorem.

2

8.6 Ideal Metrics in the Problem of Rounding It is a widely accepted fact that sums of rounded proportions often fail to add to 1. In their pioneering works, Mosteller, Youtz, and Zahn (1967) and Diaconis and Freedman (1979) assessed the probability that a vector of conventionally rounded percentages adds to 100. The conventional rule picks the midpoint of each interval as the threshold for rounding. However, the goal of rounding to maximize the probability that the sum of roundings is the rounding of the sum may well not be a “significant” question: Instead, it seems that the goal of rounding to obtain a distribution as much like the original one as possible is more fundamental.

8.6 Ideal Metrics in the Problem of Rounding

179

Suppose, for example, that q1 , . . . , qs are s independent identically [0, 1]uniformly distributed random numbers and that each is to be rounded to either 0, 12 , or 1. The usual method is to let x = 0 if 0 ≤ q < 14 , x = 12 if 41 ≤ q < 34 , and x = 1 if 34 ≤ q ≤ 1. Then Ex = Eq = 21 , and 1 . On the other hand, if instead of rounding “at” Var x = 81 = Var q = 12 1 3 1 5 1 1 ∗ ∗ 4 and 4 one rounds at 6 and 6 , that is, x = 0 if 0 ≤ q < 6 , x = 2 if 61 ≤ q < 65 , and x∗ = 1 if 56 ≤ q ≤ 1, then Ex∗ = Eq = 12 and 1 . The importance of the deviations between the sums Var x∗ = Var q = 12 of the resulting roundings xS = x1 + · · · + xs and x∗S = x∗1 + · · · + x∗s from qS = q1 + · · · + qs may be seen by comparing the differences △s = sup |P (a < qS < b) − P (a < xS < b)| a0 |E(aX − aY )| = +∞ if E(X − Y ) = 0. Thus, a necessary condition for x∗ = ̺xt (q) to be an optimal stationary rule with respect to θr is the equality of the first moments of q1 and of its (1/t)rounding: Eq1 =

Ex∗1



1 = P (q1 > d(k)/t). t 0

or Eq1

(8.6.19)

This condition is sufficient under the mild assumption that Eq1r < ∞.

(8.6.20)

In fact, (8.6.19) implies θr (qs , x∗s ) ≤ s where κr (X, Y ) = r

1 κr (q1 , r1 ), Γ(r − 1)

∞ 0

xr−1 |FX (x)−FY (x)| dx is the Kantorovich rth pseu-

domoment, and by (8.6.20), κr (q1 , x∗1 )



Eq1r



1 + E q1 + t

r

< ∞.

Summarizing, this yields the following theorem. Theorem 8.6.3 Suppose that the vector q = (q1 , . . . , qs ) consists of i.i.d. random variables, and Eq1r is finite for some r ∈ (1, 2). Then   1 1 θr θr (qs , xs ) = ∞, = ∞ qs , xs s s for any rule of (1/t)-rounding x = ̺t (q) with Eq1 = Ex1 . However, if Eq1 = Ex∗1 for some stationary rule x∗ = ̺∗t (q), then ̺∗t is an optimal rule with respect to μ over the class of all stationary rules. Moreover,   1 1 1 ∗ θr ≤ s1−r qs , xs κr (q1 , x∗1 ) = O(s1−r ). s s Γ(r − 1)

8.6 Ideal Metrics in the Problem of Rounding

185

Corollary 8.6.4 If in Theorem 8.6.3 ̺t is stationary, then it is optimal with respect to θr if and only if tEq1 =

∞ 

P (tq1 > k + C).

(8.6.21)

0

Equation (8.6.21) always has a solution C that is unique for any t > 0 under the condition that Fq1 (x) is strictly increasing. Thus, an optimal stationary rule with respect to θr exists and is unique over the set of stationary rules. Example 8.6.5 Suppose q1 is uniform on the interval (0, 1). Then (8.6.21) becomes  [t−C]  k+C 1 1  1− . = 2 t 0 t If t ∈ IN = {1, 2, . . .}, then this is equivalent to  t−1  1 k+C 1 1− = , 2 t 0 t whose solution is C(t) = 12 , so the Webster rule is optimal. However, if t ∈ N + 21 , then the solution is C(t) = (t − 14 )/(2t + 1) < 12 , so the Webster rule is only asymptotically optimal. The set of stationary rules are clearly a very restrictive class within the class of all divisor rules. Moreover, the rate of convergence of (8.6.15) for θr can be very slow, as can be seen from Theorem 8.6.3, where 1 − r ∈ (−1, 0). Indeed, simple examples show that the order of convergence O(s1−r ) is exact. Given two optimal rounding rules with respect to some μ, x∗ = ̺∗t (q) and x = ̺t (q), ̺∗t is preferred to ̺t if μ((1/s)qs , (1/s)x∗s ) → 0 at a faster rate than μ((1/s)qs , (1/s)xs ) → 0 as s → ∞. And ̺∗t is optimal of order λ > 0 with respect to μ over a class R of rules if for any q it is optimal, and   1 1 ∗ → O(s−λ ) as s → ∞. (8.6.22) qs , xs μ s s Theorem 8.6.3 tells us that there exists an optimal stationary rule ̺∗t of order λ = r − 1 with respect to θr if the rth moment of q1 exists and is finite. Is it possible that an ideal metric μ other than θr would determine an optimal stationary rule different from ̺∗t ? The answer is negative for all “nonpathological” metrics.

186

8. Probabilistic-Type Limit Theorems

Corollary 8.6.6 Suppose μ is an ideal metric of order r > 1 such that the law of large numbers holds with respect to μ; that is,   X 1 + · · · + Xn , EX1 μ → 0 as n → ∞ (8.6.23) n for any nonnegative i.i.d. Xi ’s with finite EX1 . Then there is a unique stationary rule of (1/t)-rounding that is optimal of order r − 1, and it is determined by the solution C to (8.6.21). Proof: Suppose two different stationary rules x∗ = ̺∗t (q) and x = ̺t (q) are optimal of orders λ∗ and λ (where λ∗ and λ may be different). Then     1 ∗ 1 ∗  ∗ 1 ∗ μ(Ex1 , Ex1 ) ≤ μ Ex1 , xs + μ x , qs s s s s     1 1  1  +μ qs , xs + μ x , Ex1 . s s s s However, the right-hand side goes to 0 as s → ∞ by (8.6.16) and (8.6.23), and therefore Ex∗1 = Ex1 . Since the rules are stationary, they are both determined by (8.6.4), and so they are the same. 2 To obtain faster rates of convergence one must consider a wider class of divisor rules than the stationary ones. It is our objective to show that if one extends the analysis to K-stationary rules, then one can find an optimal rule of order λ = K + 1. To do this it is necessary to generalize the definition of θr (since by the previous definition r ∈ (1, 2)), which can be done as follows: θr (X, Y ) = sup{|E(f (X) − f (Y ))|; f ∈ Fr },

(8.6.24)

where r = r0 + 1/p > 0, r0 ≥ 0 an integer, and p ≥ 1. Furthermore, in (8.6.24), Fr is the set of all functions f with (r0 + 1)-derivative f (r0 +1) satisfying f (r0 +1) q = [ |f (r0 +1) |q ]1/q ≤ 1, with (1/p) + (1/q) = 1. As before, it may be checked that   1 1 θr ≤ s1−r θr (q1 , x1 ), qs , xs s s so that (8.6.18) holds, provided that θr (q1 , x1 ) < ∞. In addition, θr (q1 , x1 ) < ∞

implies

E(q1k − xk1 ) = 0

for 0 < k ≤ r0 . (8.6.25)

Conversely, if Eq1r < ∞, then E(q1k − xk1 ) = 0 for 0 < k ≤ r0 implies θr (q1 , x1 ) < ∞. (8.6.26)

8.6 Ideal Metrics in the Problem of Rounding

187

In fact, θr (q1 , x1 ) can be bounded by θr (q1 , x1 ) ≤ Cr K κr (q1 , x1 ) ≤ Cr Eq1r < ∞,

(8.6.27)

where Cr and Cr are constants, and κr is the pseudomoment of order r. It is possible to say more. θr (q1 , x1 ) → 0 means that the distributions of q1 and of x1 “θ-merge”; that is, L(q1 , x1 ) → 0 and E(q1r − xr1 ) → 0, where L is the L´evy metric between the distributions of q1 and x1 (for merging, the L´evy metric, and related concepts see D’Aristotile, Diaconis, and Freedman (1988) and Rachev (1991)). A K-stationary rule x∗ = ̺∗t (q) of order λ = r − 1 is optimal with respect to the metric θr over the class of K-stationary rules if for any q, θr (qs , x∗s ) = min{θr (qs , xs ); x = ̺t (q), ̺t K-stationary}, (8.6.28) ̺t

and furthermore, the rate of θr -merging is θr



1 1 qs , x∗s s s



= O(s1−r )

as s → ∞

(8.6.29)

whenever Eq1r < ∞. In fact, (8.6.25) through (8.6.27) imply that if for a rule x = ̺t (q) the moment conditions E(q1k − xk1 ) = 0 do not hold for some k = 0, . . . , r0 , then θr (q1 , x1 ) = ∞. So (8.6.28) fails for s = 1, and thus ̺t is not optimal. On the other hand, if these moment conditions do hold for all k for some rule x∗ = ̺∗t (q), then by (8.6.27) and the ideality of θr , θr (qs , x∗s ) ≤ sCr Eq < ∞ for any fixed s, so (8.6.28) is fulfilled. This proves Theorem 8.6.7 Suppose Eq1r < ∞ with r = K +1+1/p. Then x∗ = ̺∗t (q) is an optimal K-stationary rule of order λ = r − 1 with respect to the metric θr if and only if the thresholds C0 , . . . , CK−1 , C = CK = CK+1 . . . are chosen such that E(q1j − x∗j 1 )=0

for j = 1, . . . , K + 1.

(8.6.30)

This means that the K +1 thresholds must be chosen such that the following system of equations is satisfied Eq1j

=

∞  j  k

k=1

t

P (k − 1 + Ck−1 < tq1 < k + Ck ),

j = 1, . . . , K + 1,

where Ck = CK for k ≥ K.

(8.6.31)

188

8. Probabilistic-Type Limit Theorems

It has been observed above that when q1 is uniform over the interval (0, 1), t is an integer, and K = 0, then the optimal rule is Webster’s. But if K > 0, there is no stationary rule that meets (8.6.31) even in the uniform case. Example 8.6.8 Suppose q1 is uniform over the interval (0, 1), t is an integer, and K = 1.(4) Then (8.6.31) yields the solution C0 = 13 , C = (3t − 2)/(6t − 6), so the optimal 1-stationary rule is determined by d(0) = 1 3 , d(k) = k + (3t − 2)/(6t − 6) for k = 0. Suppose t = 2. Then q1 is rounded to either 0, 21 , or 1 as follows: x∗1 = 0 if 0 < q1 ≤ 61 , x∗1 = 12 if 61 < q1 ≤ 56 , and x∗1 = 1 if 65 < q1 ≤ 1. The first 1 two moments of q1 and x∗1 agree: Eq1 = 12 = Ex∗1 , Var q1 = 12 = Var x∗1 . (Indeed, the third moments also agree: Eq13 = E(x∗1 )3 = 41 .) The Webster rule, on the other hand, rounds q1 as follows: xw 1 = 0 if 1 1 3 1 3 w w 0 < q1 ≤ 4 , x1 = 2 if 4 < q ≤ 4 , and x1 = 1 if 4 < q1 ≤ 1. The first 1 w moments of q1 and xw 1 agree but not the second: Eq1 = 2 = Ex1 , Var q1 = 1 1 w 12 < 8 = Var x1 . Since the first two moments of q1 and x∗1 are equal for the optimal 1stationary rule, the central limit theorem applies, and for the Kolmogorov metric ̺(qs , x∗s ) = sup{|P (qs ≤ x) − P (x∗s ≤ x)|; x ∈ IR} we have # $ qs − s/2 x∗s − s/2 ∗ ̺(qs , xs ) = ̺ √ , √ ≈ O(s−1/2 ) as s → ∞. 12s 12s In contrast, the Webster rule gives * 8 + # $ w w qs − s/2 xs − s/2 qs − s/2 xs − s/2 8 √ ̺(qs , xw ̺ √ , √ = ̺ , √ s) = 12 12s 12s 12s 8s   √ → ̺ N(0,1) , N(0, 2/3) > 0 as s → ∞,

where N(m,σ) is the normal distribution with mean m and standard deviation σ. In terms of the ideal metric θ3 there is even better evidence for the advantage of the optimal 1-stationary rule over the standard Webster rounding. For   1 1 ∗ ≤ s−2 θ3 (q1 , x∗1 ) ≤ constant s−2 , qs , xs θ3 s s the last inequality is due to the optimality of x∗1 , whereas   1 1 ∗ qs , xs θ3 = +∞ for any s. s s (4) For

(1993).

other examples including the case K = 2, we refer to Balinski and Rachev

8.6 Ideal Metrics in the Problem of Rounding

189

For more results concerning the vector problem with more i.i.d. observations and the problem of rounding tables we refer to Balinski and Rachev (1993).

9 Mass Transportation Problems and Recursive Stochastic Equations

In this chapter we use the regularity properties of metrics and distances defined via mass transportation problems in order to investigate the asymptotic behavior of stochastic algorithms and recursive equations. The recursive structure allows us to apply fixed-point and approximation techniques to the space of probability measures supplied with adapted probability metrics in order to describe the limiting behavior of various algorithms.

9.1 Recursive Algorithms and Contraction of Transformations Several different approaches to the asymptotic analysis of algorithms have been given in the literature. Interesting results have been obtained by the transformation method, the method of branching processes, the method based on stochastic approximations, the martingale method, and others. The analysis of algorithms is an important application of stochastics in computer science and poses challenging questions and problems. It has led to some new developments also in stochastics. Based on the properties of minimal metrics as introduced in Chapter 8, a promising new method for asymptotic analysis has recently been introduced. R¨ osler (1991) gave an asymptotic analysis of the quicksort algorithm based on the minimal ℓp -metric. His proof has been extended in several papers by Rachev and R¨ uschendorf to a general “contraction method” with

192

9. Mass Transportation Problems and Recursive Stochastic Equations

a wide range of possible applications. A series of examples and further developments of the method have been found in some recent work. The contraction method (in its basic form) uses the following sequence of steps: 1. Find the correct normalization of the algorithms. (Typically by studying the first moments or tails.) 2. Determine the recursion for the normalized algorithm. 3. Determine the limiting form of the normalized algorithms. The limiting equation typically is defined via a transformation T on the set of probability measures. 4. Choose an ideal metric μ such that T has good contraction properties with respect to μ. This ideal metric has to reflect the structure of the algorithm. It also has to have good bounds in terms of interpretable other metrics and has to allow one to estimate bounds (in terms of moments usually). As a consequence one obtains 5. The conjectured limiting distribution is the unique fixed point of T . Finally, one should ensure that the recursion is stable enough so that the contraction in the limit can be pulled over to establish contraction properties of the recursion itself for n → ∞. This is the technically most involved step in the analysis. 6. Establish convergence of the algorithms to the fixed point. Applications of this method have been given to several sorting algorithms, to the communication resolution interval (CRI) algorithm, to generalized branching-type algorithms, to bootstrap estimators, iterated function systems, learning algorithms, and others. For several examples modifications of this method have been considered. There are examples where the contraction factors converge to one. In several cases there is a trivial limiting recursion that gives no clue to a possible limit distribution. Also, logarithmic normalizations and convergence rates have to be handled by special considerations. We begin with a discussion of contraction properties of transformations T on the set of probability distributions on a basis space U. Stochastic algorithms are typically directly described by the iterates of a transformation T on the set of all probability distributions on the basic space U , or else they are asymptotically closely related to the iterations (and thereby to the fixed points) of a transformation T that describes the limiting equation. They can, for example, differ from iterations by a stochastic sequence converging to zero. Examples for recursive algorithms

9.1 Recursive Algorithms and Contraction of Transformations

193

that are asymptotically related to iterations of transformations T are studied in Section 9.2. Consider a contraction transformation T : M 1 (U ) → M 1 (U ), where M 1 (U ) is the set of probability measures on U supplied with probability metrics as in Chapter 8. Applying the fixed-point theorems for complete, separable metric spaces, we can infer the convergence of the iterates (T n F ), F ∈ M 1 (U ), to a fixed point of T . Some of the following examples serve to describe the influence of the choice of the metrics, while others indicate the range of applicability to different fields. (a) Consider at first a transformation of the form d

TF =

N 

ai (τ )Yi + C(τ ).

(9.1.1)

i=1

Here, for F ∈ M 1 (U ), (Yi )1≤i≤N is an i.i.d. sequence with distribution F . Furthermore, C(τ ), ai (τ ), τ are real random variables independent of (Yi ), N and finally, T F is the law of i=1 ai (τ )Yi + C(τ ). Consider Zolotarev’s ideal metric ζr of order r > 0: For r = m + α, m ∈ IN, and 0 < α ≤ 1, ζr (X, Y ) :=

(9.1.2) ' & (m) (m) α sup |E(f (X) − f (Y ))|; |f (x) − f (y)| ≤ x − y .

Here f (m) (x) denotes the Fr´echet derivative of order m, and  ·  is a norm on U . Next, suppose that F, G ∈ M 1 (U ), where (U,  · ) is a Banach space. Proposition 9.1.1 ζr (T F, T G) ≤

2N 

E|ai (τ )|r

i=1

3

ζr (F, G).

(9.1.3)

Proof: The proof of (9.1.3) uses the ideality properties of ζr ; that is, (i) ζr (X + Z, Y + Z) ≤ ζr (X, Y )

for Z independent of X, Y,

(9.1.4)

and (ii) ζr (cX, cY ) = |c|r ζr (X, Y ), for all c ∈ IR. d

d

(9.1.5)

Let Yi = F , Zi = G be independent r.v.s. Then, with r = m + α,  &      ai (τ )Yi + C(τ ) −Ef ai (τ )Zi + C(τ )  ; ζr (T F, T G) = sup Ef ' |f (m) (x) − f (m) (y)| ≤ |x − y|α

194

9. Mass Transportation Problems and Recursive Stochastic Equations

≤ ≤ ≤



ζr



ζr 





ai (t)Yi + C(t), ai (t)Yi ,





 ai (t)Zi + C(t) dP τ (t)

 ai (t)Zi ) dP τ (t)

|ai (t)|r dP τ (t)ζr (Yi , Zi ) =



E|ai (τ )|r ζr (F, G),

which proves (9.1.3).

2

For the property to hold it suffices to require that 

E|ai (τ )|r < 1

and

ζr (F, G) < ∞.

(9.1.6)

In some cases, the last condition can be established by making use of the inequality

ζr ≤

Γ(1 + α) vr , Γ(1 + r)

(9.1.7)

where

vr (X, Y ) :=



xr d|PX − PY |(x)

is the absolute pseudomoment of order r. For random vectors X and Y , and m ≤ r < m + 1, (9.1.7) requires that all moments X and Y of order ≤ m coincide. Recall that the minimal Lp -metrics ℓp are ideal of order r = min(1, p).

Proposition 9.1.2 For F, G ∈ M 1 (U ),

ℓp (T F, T G) ≤

2N  i=1

E|ai (τ )|r

3

ℓp (F, G).

(9.1.8)

9.1 Recursive Algorithms and Contraction of Transformations d

195

d

Proof: Let Yi = F, Zi = G, 1 ≤ i ≤ N , be independent pairs of random variables with laws F, G, and Lp (Yi , Zi ) = ℓp (F, G). Then    ai (τ )Zi + C(τ ) ℓp (T F, T G) = ℓp ai (τ )Yi + C(τ ), ( (  ( ( ai (τ )Yi − ≤ ( ai (τ )Zi ( p



=

N  i=1

E|ai (τ )|r Yi − Zi p

2N 

E|ai (τ )|r

i=1

3

ℓp (F, G). 2

So, the contraction property will hold if N 

E|ai (τ )|r < 1

i=1

and ℓp (F, G) < ∞.

(9.1.9)

Under additional assumptions we may improve (9.1.9). Let U be a Hilbert space and let F, G have identical first moments. Then the following is a refinement of Proposition 9.1.2. Proposition 9.1.3 2 31/2 N  ℓ2 (T F, T G) ≤ E|ai (τ )|2 ℓ2 (F, G).

(9.1.10)

i=1

Proof: With Yi , Zi as in the proof of (9.1.8), we have ( (2 ( ( 2 ℓ2 (T F, T G) ≤ ( ai (τ )(Yi − Zi )( 2

=

N  i=1

=

E|ai (τ )|2 Yi − Zi 22



2

E|ai (τ )|



ℓ22 (F, G). 2

If the Banach space U is of type p, 1 ≤ p ≤ 2, and F, G have identical d d first moments (more precisely, E(Y − Z) = 0 for Y = F, Z = G), then for 1 ≤ p ≤ 2, ℓp (T F, T G) ≤ Bp1/p

2N  i=1

E|ai (τ )|p

31/p

ℓp (F, G).

(9.1.11)

196

9. Mass Transportation Problems and Recursive Stochastic Equations

Here Bp is the constant arising in the Woyczinski inequality (cf. Rachev and R¨ uschendorf (1992a)). For U = Lp (μ) (Lp (μ) is the space of all r.v.s X with  finite |X|p dμ), one can choose the constants B1 = 1, Bp = 18p3/2 /(p − 1)1/2 for 1 < p ≤ 2. The proof of (9.1.11) is similar to that of (9.1.10), but in (9.1.11) we use the Woyczinski inequality instead of the Hilbert space structure. If the underlying space is Euclidean, we can derive similar contraction properties with respect to other metrics defined in Chapter 8. Example 9.1.4 Let N = 2, let τ be uniformly distributed on (0,1), and a1 (τ ) = τ, a2 (τ ) = 1 − τ . Then the contraction factor α with respect to the ℓp -metric is given in the following list: ℓ1 -metric : α ℓ2 -metric : α ζ2 -metric : α ζ3 -metric : α

= Eτ + E(1 − τ ) = 1,

i.e., “no contraction”; (9.1.12) % = (Eτ 2 + E(1 − τ )2 )1/2 = 2/3; = Eτ 2 + E(1 − τ )2 = 2/3;

= 1/2.

Clearly, if for a probability metric μ on M 1 (U ) the contraction factor is α < 1, μ(T F, T G) ≤ αμ(F, G), then μ(T n+1 F, T n F ) ≤ αn μ(T F, F );

(9.1.13)

i.e., one obtains an exponential convergence rate to a fixed point. In the d example above we consider the recursion Xn+1 = τn Xn + (1 − τn )X n + d

d

C(τn ), τn = τ, where X n = Xn , and τn , Xn , X n are independent. The d

corresponding fixed point equation is X = τ X + (1 − τ )X + C(τ ). So under the condition of equal first moments, the convergence rate is (2/3)n for the “ideal” metric ζ2 , in comparison to (2/3)n/2 for the ℓ2 -metric. √ √ If a1 (τ ) = τ , a2 (τ ) = 1 − τ , then with respect to the ℓ2 -metric the contraction factor is α = Eτ + E(1 − τ ) = 1; i.e., there is no contraction. The same “no contraction” property is valid for ζ2 . For ζ3 the contraction factor is α = Eτ 3/2 + E(1 − τ )3/2 = 54 < 1; so the contraction property holds if ζ3 (F, G) < ∞. (b) We next consider the transformation d

TF =

max {ai (τ )Yi }.

1≤i≤N

(9.1.14)

For U = IRk , k = 1, 2, . . . , ∞, F ∈ M 1 (U ), let (Yi )1≤i≤N , τ be as in (a). Let ai (τ ) ≥ 0 and consider d

TF =

max {ai (τ )Yi }.

1≤i≤N

(9.1.15)

9.1 Recursive Algorithms and Contraction of Transformations

197

We shall study the contraction property of T by making use of the weighted uniform metric ̺r : ̺r (X, Y )

=

sup M (x)r |FX (x) − FY (x)|,

(9.1.16)

x∈IRk

where M (x) := mini≤k |xi |. In the next proposition we use the fact that ̺r is an ideal metric of order r with respect to the maxima of i.i.d. r.v.s. Proposition 9.1.5 ̺r (T F, T G) ≤



r

E(ai (τ ))



̺r (F, G).

(9.1.17)

d

Proof: Let (Zi ) be i.i.d. and Z1 = G. Then using the max-ideality of ̺r , we have   ̺r (T F, T G) = ̺r max{ai (τ )Yi }, max{ai (τ )Zi } i≤N i≤N     = sup |x|r  (Fmax{ai (t)Yi } (x) − Fmax{ai (t)Zi } (x)) dP τ (t) x  ≤ ̺r (max{ai (t)Yi }, max{ai (t)Zi }) dP τ (t) 2 3   ≤ ai (t)r dP τ (t)̺r (Yi , Zi ) = Eai (τ )r ̺r (Y1 , Z1 ). i

i

2 For more general maxima we can again use the ℓp -metrics. Let U = L (μ), 1 ≤ λ < ∞. For F ∈ M 1 (U ) and ai (τ ) ≥ 0 consider λ

d

TF =

max ai (τ )Yi + C(τ ).

i≤i≤N

(9.1.18)

d

Here (Yi ) are i.i.d., Y1 = F, τ is independent of (Yi ), and C(τ ) has values in U . For any p, λ we have 2N 3  Eai (τ )r ℓp (F, G), r = min(1, p). (9.1.19) ℓp (T F, T G) ≤ i=1

For 1 ≤ p ≤ λ < ∞ we have the following improvement: Proposition 9.1.6 If 1 ≤ p ≤ λ < ∞, then ℓp (T F, T G) ≤

N  i=1

(Eai (τ )p )1/p ℓp (F, G).

(9.1.20)

198

9. Mass Transportation Problems and Recursive Stochastic Equations d

d

Proof: Let Yi = F , Zi = G satisfy Yi − Zi pλ,μ = ℓp (F, G)p , where  Xλ,μ = ( |X(t)|λ dμ(t))1/λ . Then ℓp (T F, T G) ≤ (E max ai (τ )Yi − max ai (τ )Zi pλ,μ )1/p 31/p 2  p/λ dP τ (t) | max ai (t)Yi (s)−max ai (t)Zi (s)|λ dμ(s) = E

⎞1/p ⎛ 3p/λ 2    ai (t)λ |Yi (s) − Zi (s)|λ dμ(s) dP τ (t)⎠ ≤ ⎝ E i



2  i

E

#

31/p $p/λ ai (t)λ |Yi (s) − Zi (s)|λ dμ(s) dP τ (t)

(since p/λ ≤ 1) =

2  i

Eai (τ )EYi − Zi pλ,μ

31/p

=

2 

Eai (τ )p

i

31/p

ℓp (F, G). 2

(c) Bootstrap Estimators

 For a separable Banach space U and F ∈ M 1 (U ), let μ(F ) = x dF (x), n d μn (F ) = n1 i=1 Xi , where (Xi ) are i.i.d., Xi = F , and Fn is the empirical measure of X1 , . . . , Xn . For p > 0 denote by Γp the class of distributions with finite pth moment. From the strong law of large numbers, for any F ∈ Γp we then obtain (cf. Chapter 8) ℓp (Fn , F ) → 0 a.s. and Eℓp (Fn , F ) → 0.

(9.1.21) d

∗ Let now X1∗ , . . . , Xm be a bootstrap sample; i.e., the (Xi∗ ) are i.i.d., X1∗ = ∗ is the empirical distribution of Fn (conditionally on X1 , . . . , Xn ), and Fn,m ∗ ∗ ∗ X1 , . . . , Xm , m = m(n). The condition ℓp (Fn,m , F ) → 0 a.s. (conditionally on X) is equivalent to the joint convergence m

1  f (Xi∗ ) → m i=1 m

1  p ∗ d (Xi , a) m i=1







f dF a.s.,

f ∈ Cb (U ),

dp (x, y) dF (x) a.s.

9.1 Recursive Algorithms and Contraction of Transformations

199

(cf. Chapter 8), representing a special form of the SLLN for real-valued r.v.s. In the case (U, d) = (IRr ,  · ), p > 1, we are able to obtain a rate of convergence for the approximation. Let γ = kr/[(k − r)(k − 2)],  bootstrap γ k > r, k > 2, and x F ( dx) < ∞. Then 1

Eℓpp (Fn , F ) ≤ C(r, k, p)n−(1− p )/k

(9.1.22)

(cf. Rachev (1984d, pp. 667–668)), and thus 1

∗ ∗ , F ) ≤ 2p EEX1 ,...,Xn ℓpp (Fn,m , Fn ) + 2p C(r, k, p)n−(1− p )/k Eℓpp (Fn,m   1 1 −(1− p )/k −(1− p )/k ∗ . (9.1.23) ≤ C (r, k, p) m +n

If, however, the (Xi ) are in the domain of an α-stable distribution, then it is more natural to choose the bootstrap estimator from a distribution Fn 1 , . . . , X n be a bootstrap which has a tail behavior similar to F . Let then X sample with adapted tail behavior such that ℓp (Fn , F ) → 0.

Consider Vn =

1 n1/α

n

(9.1.24)

d i ) =   X Tn (Fn ) (1 ≤ α ≤ 2) as a i=1 (Xi − EF n n d 1 Vn = n1/α i=1 (Xi − EF Xi ) = Tn (F ). Then

boot-

strap estimator of for a Banach space of type p, 1 ≤ p ≤ 2 and 1 ≤ α ≤ p, it follows from (9.1.11) that 1 1 1 − E X  1 , X1 − E F X 1 ) ℓp (Vn , Vn ) ≤ Bp n p − α ℓp (X 1 1 1 − EF X1 |) → 0. ≤ Bp n p − α (ℓp (Fn , F ) + |E X

(9.1.25)

Since ℓp (Vn , Y(α) ) → 0, where Y(α) an α-stable r.v., then in the case ℓp (X1 , Y(α) ) < ∞, it follows from the bound in (9.1.25) that ℓp (Vn , Y(α) ) → 0.

(9.1.26)

Moreover, the rate of convergence is of order o(n1/p−1/α ). In the case p = 2 and F ∈ Γ2 , the condition (9.1.24) is satisfied for Fn = Fn∗ ; for Euclidean spaces this case has been considered by Bickel and Freedman (1981). Their investigation of more general functionals on the set of empirical measures can also be extended to the setting we described. (d) Transformation by Markov Kernels, Image Encoding Let (U, d) be a separable metric space and let wi : U → U, 1 ≤ i ≤ N , be mappings satisfying d(wi x, wi y) ≤ si d(x, y).

(9.1.27)

200

9. Mass Transportation Problems and Recursive Stochastic Equations

Given a probability distribution (pi )1≤i≤N define the Markov kernel K(x, ·) =

N 

pi εwi (x) ,

i=1

N ≤ ∞.

(9.1.28)

The implied transformation on M 1 (U ) is denoted by  T F = KF, where KF (A) = K(x, A)F ( dx).

(9.1.29)

Let now Lip(U ) be the set of Lipschitz  functions, |f (x) − f (y)| ≤ d(x, y) for all x, y ∈ U . Then for Kf (x) = pi f ◦ wi (x), we have  |Kf (x) − Kf (y)| ≤ pi |f ◦ wi (x) − f ◦ wi (y)| (9.1.30)     pi si d(x, y). ≤ pi d(wi (x), wi (y)) ≤ Let us look at the contraction properties for the mapping T with respect to the Kantorovich metric d

d

μL (F, G) = sup{|E(f (X) − f (Y ))|; X = F, Y = G, f ∈ Lip(U )}. (9.1.31) We have then μL (T F, T G) = sup{|E(f (K(X, ·) − f (K(Y, ·)))|; f ∈ Lip(U )} (9.1.32) = sup{|E(Kf (X) − Kf (Y ))|; f ∈ Lip(U )}   pi si sup{|E(g(X) − g(Y ))|; g ∈ Lip(U )} ≤   pi si μL (F, G). =

 If pi si < 1, then T is a contractive mapping. By the Kantorovich–Rubinstein theorem μL coincides with the minimal L1 -metric, and therefore,   pi si ℓ1 (F, G). ℓ1 (T F, T G) ≤ (9.1.33) Moreover, for any p > 0, we can extend this result as follows. Proposition 9.1.7 ℓp (T F, T G) ≤ d



pi spi d

1/p∧1

ℓp (F, G).

Proof: Suppose X = F, Y = G satisfy ⎧ ⎨ (E d(X, Y )p )1/p , if p ≥ 1, ℓp (F, G) = ⎩ E d(X, Y )p , if p < 1.

(9.1.34)

9.1 Recursive Algorithms and Contraction of Transformations

201

Take I to be a random variable with values in {1, 2, . . . , N } and distribution (pi ) that is independent of X, Y . Then for 1 ≤ p, ℓpp (T F, T G)

≤ E d(wI(X), wI(Y ))p =

N  i=1

=

N 

E(d(wi (X), wi (Y ))p I(I = i)) pi E d(wi (X), wi (Y ))p

i=1



(9.1.35)

2N 

pi spi

i=1

3

E d(X, Y ) =



pi spi



ℓpp (F, G).

The proof for the case 0 < p < 1 is similar.

2

Remark 9.1.8 Another proof of (9.1.32) can be given via the dual representation of ℓp . Indeed, for 0 < p < ∞, p = 1 ∨ p, ′ ℓpp (F, G)

=

sup



f dF +



g dG; f, g bounded continuous,  p f (x) + g(y) ≤ d (x, y), ∀x, y ∈ U .

(9.1.36)

Therefore, ′

ℓpp (T F, T G) = d

(9.1.37)

sup{E f (K(X, ·)) + E g(K(Y, ·)); f (x) + g(y) ≤ dp (x, y)}, d

where X = F, Y = G. Since Kf (x) + Kg(y) ≤ we obtain



pi dp (wi (x), wi (y)) ≤



 pi spi dp (x, y),



ℓpp (T F, T G) = sup{E Kf (X) + E Kg(Y )), f (x) + g(y) ≤ dp (x, y)}   ′ pi spi ℓpp (F, G). ≤ (9.1.38) Remark 9.1.9 Hutchinson (1981) was the first to prove convergence with respect to the metric μL in the case si ≤ 1. Barnsley and Elton (1988) used the above Markov chain to “construct images” by so-called iterated

202

9. Mass Transportation Problems and Recursive Stochastic Equations

function systems (IFS). They established the existence of a unique attractive invariant measure μ under the assumption that N 9

i=1

d(wi (x), wi (y))pi ≤ r d(x, y),

r < 1.

(9.1.39)

The above inequality is indeed implied by the condition (9.1.27) with N 9

spi i < 1.

(9.1.40)

i=1

In the case of affine maps on IRk we can improve the arguments in the following way (see Proposition 9.1.10 below, cf. also Burton and R¨ osler (1995)). Define d

T F = AX + b,

(9.1.41)

where A is a random matrix, b a random vector, (A, b) independent of X, d

and X = F . Consider the operator norm of the expected product EAT A, EAT A := sup

x∈IRk x=0

(EAT A)x . x

(9.1.42)

Then EAT A = sup x =0

EAx, Ax = EA, A, x2

(9.1.43)

where the right-hand side is the L2 -norm of EA, A. Proposition 9.1.10 Assume that  b 2 < ∞. Then ℓ2 (T μ, T ν) ≤

A

EAT A ℓ2 (μ, ν)

for any μ, ν ∈ M 1 (IRk ) with finite second moments.

(9.1.44)

9.1 Recursive Algorithms and Contraction of Transformations

203

Proof: Let Y, Z be random vectors with distributions μ, ν and ℓ2 (μ, ν) = (EY − Z2 )1/2 , where Y, Z are independent of (A, b). Then ℓ2 (T μ, T ν) ≤  AY − AZ 2 % EA(Y − Z), A(Y − Z) ≤ A = EY − Z, E(AT A)(Y − Z) A % ≤ EAT A EY − Z, Y − Z A = EAT A ℓ2 (μ, ν).

(9.1.45)

2

Notice that the estimate from above defined in (9.1.45) is an improvement (in the case p = 2) over the general estimate  AX p ≤  A · X p ≤  A p ·  X p .

(9.1.46)

In fact, the above general bound requires the stronger condition  A p < 1 to yield the contraction property. (e) Environmental Processes Let (Yi , Zi ) be a sequence of i.i.d. pairs of r.v.s with values in U × IR, where U is a separable Banach space. Define a sequence of r.v.s (Sn ) by Sn+1 = (Yn + Sn )Zn ,

S0 ≥ 0.

(9.1.47)

This kind of process has found several applications in environmental modeling and has been studied intensively. If we write τn = (Yn , Zn ), and a(τn ) = Zn , C(τn ) = Yn Zn , then Sn+1 = a(τn )Sn + C(τn ),

(9.1.48)

so we have a special case of (9.1.1). Under the condition that E|a(τ )|r < 1, d

the operator T S = a(τ )S + C(τ ) is contractive. Therefore, (Sn ) converges (with respect to some ideal metric of order r such as ζr , for example) to a fixed point, i.e., a solution of d

S = (Y + S)Z.

(9.1.49)

Numerous properties of the solutions of the above equation have been studied in the literature; see, for example, Rachev and Samorodnitsky (1995) and Rachev and R¨ uschendorf (1995).

204

9. Mass Transportation Problems and Recursive Stochastic Equations

9.2 Convergence of Recursive Algorithms In this section we apply the contraction properties established in Section 9.1 to study limits for recursive algorithms. We shall use the “method of probability metrics.” The main idea of this method is to transform the recursive equations in such a way that with respect to a suitable metric we can derive contraction properties in the limit; i.e., we consider decompoBn such that (Yn ) has contraction properties and W Bn sitions Xn = Yn + W converges to zero. This idea will be demonstrated in various examples. n The approach is natural from the following point of view. If Sn = i=1 Yi is a sum of independent (centered) random variables and Xn = n−1/α Sn is the normalized sum, then Xn satisfies the following simple recursion: Xn+1 =



n n+1

−1/α

Xn + (n + 1)−1/α Yn+1 .

(9.2.1)

Thus the central limit theorem can be considered as the limit theorem of this simple (stochastic) recursion. The form of the recursion corresponding to the strong law of large numbers is even simpler.

9.2.1

Learning Algorithm

Let Y1 , Y2 , . . . be an i.i.d. sequence of r.v.s with values in a separable Banach space with first moment μ. Define the following recursive sequence: Let X1 be arbitrary with finite first moment, and let Xn+1 =

n 1 Xn + Yn+1 . n+1 n+1

(9.2.2)

Xn can be viewed as an easy recursive algorithm designed to “learn” about the unknown theoretical mean μ given the sample (Y1 , . . . , Yn ). Proposition 9.2.1

ζr (Xn , μ) → 0

if

ζr (X1 , μ) < ∞.

Proof: Let ℓn = EXn . Claim 1: ℓn → μ.

For the proof of Claim 1 note that from (9.2.2), we obtain ℓn+1

= = =

1 n ℓn + μ n+1 n+1 n−1 2 ℓn−1 + μ n+1 n+1 1 n ℓ1 + μ, n+1 n+1

(9.2.3)

9.2 Convergence of Recursive Algorithms

205

where the last step follows from the inductive argument. This implies Claim 1. Define next Z n = Xn − ℓ n ,

Wn = Yn − μ.

(9.2.4)

Then, Zn+1 + ℓn+1 Zn+1

1 n (Zn + ℓn ) + (Wn+1 + μ), (by (9.2.2)) n+1 n+1   n 1 1 n = Zn + Wn+1 − ℓn+1 − ℓn −μ n+1 n+1 n+1 n+1 1 n + Wn+1 (by (9.2.3)). (9.2.5) = Zn n+1 n+1 =

Now let μr be an ideal metric of order r, 1 < r < 2, and bn = μr (Zn , 0) (for example we can choose μr = ζr ). Claim 2. μr (Zn , 0) → 0 if a = μr (W1 , 0) < ∞. For the proof of this claim note that bn+1

= ≤

= =



 1 n + Wn+1 ,0 μr (Zn+1 , 0) = μr Zn n+1 n+1     1 n μr Z n , 0 + μr Wn+1 , 0 n+1 n+1 (since Zn is independent of Wn+1 )   r r 1 n μr (Zn , 0) + μr (Wn+1 , 0) n+1 n+1  r r  1 n bn + a. n+1 n+1

Therefore, bn+1

 r $  r 1 1 bn−1 + a + ≤ n n+1 a   r r n−1 1 = bn−1 + 2 a n+1 n+1  r r  1 1 b1 + n a. (9.2.6) ≤ n+1 n+1 

n n+1

r #

n−1 n

r

Since 1 < r, it follows that bn → 0. In particular, for μr = ζr , we obtain from Claim 1 that

206

9. Mass Transportation Problems and Recursive Stochastic Equations

ζr (Xn , μ) → 0

if ζr (X1 , μ) < ∞.

(9.2.7) 2

For the case of Euclidean spaces the condition ζr (Y1 , μ) < ∞ is satisfied if Y1 has a finite absolute rth moment, r > 1. Therefore, under the assumption of a finite rth moment we obtain convergence of Xn to μ. The sequence (Xn ) provides a simple example of a “learning algorithm” (for μ). Its convergence to μ in the real case can also be obtained as an application of the Robbins–Siegmund lemma (cf. Robbins and Siegmund (1971)) under the stronger assumption of a finite second moment. In this simple example we can, of course, directly prove the convergence of Xn to μ under the assumption of a finite first moment. The arguments above illustrate the general idea behind the method of probability metrics and show that in this simple case the method of probability metrics works with weaker assumptions than the method of stochastic approximation based on the Robbins–Siegmund lemma. Some further simple examples of the Robbins–Monroe-type recursion Xn+1 = fn (Xn , Yn+1 ) can be treated similarly. Note that our method only needs a metric ideal of order r > 1 such that μr (Xn − ℓn , 0) → 0 implies that Xn − ℓn → 0 in distribution. The ℓp metric will not work in this example, since its degree of ideality is only r = min(1, p).

9.2.2

Branching-Type Recursion

Consider the following recursive sequence (Ln ): L0 ≡ 1,

d

Ln =

K 

(i)

Xi Ln−1 + Y.

(9.2.8)

i=1

(i)

Here Ln−1 are i.i.d. copies of Ln−1 , (Xi ) is a real random sequence, K is a random number in IN0 , and Y is a random “immigration” such that d (i) K, {(Xi ), Y }, (Ln−1 ) are independent. As usual, = denotes equality in distribution. (9.2.8) induces a transformation T on M 1 , the set of probability distributions on (IR1 , B 1 ). This is achieved by letting T (μ) be the distriK bution of i=1 Xi Zi + Y , where the (Zi ) are i.i.d. μ-distributed r.v.s, and moreover, (Zi ), {(Xi ), Y }, K are independent.

Some special cases of those transformation and recursion have been studied intensively in the literature. If Xi ≡ 1, then (9.2.8) describes a Galton– Watson process with immigration Y with the number of descendants of a parent described by K. The recursion (9.2.8) can be viewed as a branching process with random multiplicative weights. The special case where K is constant, Y = 0, and (Xi ) are i.i.d. and nonnegative was introduced by Mandelbrot (1974) in his analysis of the Yaglom–Kolmogorov

9.2 Convergence of Recursive Algorithms

207

turbulence model. This case has been also studied by Kahane and Peyri`ere (1976) and Guivarch (1990), who considered the question of nontrivial fixed points of T , the existence of moments of the fixed points, and the convergence of (Ln ). For Xi ≡ K −1/α , the solutions of the fixed-point equation d K Z = i=1 K −1/α Zi are Paretian stable distributions (if Zi ≥ 0). For that reason the solutions are called semistable in Guivarch (1990). In this section we will be mainly interested in the case of multipliers Xi and solutions Zi with moments of order ≥ 2. While the analysis of Kahane and Peyri`ere (1976) is based on an associated martingale, Guivarch (1990) uses a more elementary martingale property together with a conjugation relation and moment-type estimates for the Lp -distance, 0 < p < 1. Motivated by some problems in infinite particle systems, Holley and Liggett (1981) and Durrett and Liggett (1983) considered a smoothing transformation with (Xi ) that are not not necessarily independent and assume that Xi ≥ 0, K constant, and Y = 0. In Durrett and Liggett (1983) a complete analysis of the case is given. In particular, a necessary and sufficient condition for the existence and characterization of (all) fixed points as well as a general sufficient condition for convergence was derived, as well as a generalization of the result of Kahane and Peyri`ere on the existence of moments. The method of Durrett and Liggett is based on an associated branching random walk. The use of contraction properties of minimal Lp -metrics in this section allows us to obtain quantitative approximation results for the recursion (9.2.8). Under moment assumptions used in this section, the recursion converges to the limiting distribution exponentially fast. This is demonstrated by simulations for several examples. Also, it is possible to remove the assumption of nonnegativity, to deal with a random number K, and to add immigration Y . This allows us to include applications to branching processes as well as to study the development of the total mass in the construction of multifractal measures (cf., for example, Arbeiter (1991)). For details we refer to Cramer and R¨ uschendorf (1996b). (a) Branching-Type Recursion with Multiplicative Weights In this section we shall study the recursion (9.2.8) allowing for dependent multipliers Xi but setting the immigration Y ≡ 0. In other words, we consider the recursion L0 ≡ 1, 

(i)



d

Ln =

K 

(i)

Xi Ln−1 ,

(9.2.9)

i=1

are i.i.d. copies of Ln−1 , (Xi ) is a square integrable real   (i) random sequence, K is a random number in IN0 , and K, (Xi ), Ln−1 are independent r.v.s. where

Ln−1

208

9. Mass Transportation Problems and Recursive Stochastic Equations

To determine the correct normalization of (L n ) we first   consider the first K moments of (Ln ). Set ℓn := ELn , c := E i=1 Xi , vn := Var(Ln ),     K K 2 a := E i=1 Xi . i=1 Xi , and b := Var

Proposition 9.2.2 Let ℓ0 = 1, ℓn = cn . Suppose that b > 0, c = 0, a = c2 . Then vn = bc

1 − ( ca2 )n , 1 − ca2

2n−2

n ≥ 1, v0 = 0.

(9.2.10)

If a = c2 = 0, then vn = nban−1 . Proof: Using the independence assumption in (9.2.9) and the conditional expectations, we obtain ℓn

3 33 2K 2 2K   (i) (i) = E EXi Ln−1 = E E Xi Ln−1 |K = E

2K 

i=1

EXi

i=1

3

i=1

ℓn−1 = c ℓn−1 ;

i.e., ℓn = cn . Similarly, vn

2

= EL2n − (ELn ) ⎡ ⎛2 32  ⎞⎤ K    (i) = E⎣ E⎝ Xi Ln−1 K ⎠⎦ − c2 ℓ2n−1  i=1 ⎤ ⎡ K       2 (j) (i) (i) = E⎣ E Xi Ln−1 + E Xi Xj Ln−1 Ln−1 ⎦ − c2 ℓ2n−1 ⎡

i=1

i =j

= E ⎣EL2n−1

K 

EXi2 + ℓ2n−1

i=1

 i =j



E(Xi Xj )⎦ − c2 ℓ2n−1

⎡ ⎤ K     = E⎣ EXi2 Var Ln−1 + ℓ2n−1 + ℓ2n−1 E(Xi Xj )⎦ − c2 ℓ2n−1 i=1

= E

2K  i=1

Xi2

3

vn−1 + Var

2K 

= a vn−1 + b c2(n−1) = b

i=1

n−1  k=0

Xi

3

i =j

ℓ2n−1

ak c2(n−1−k)

9.2 Convergence of Recursive Algorithms

=

⎧ ⎨ b c2n−2 ⎩

1−( ca2 )n 1− ca2

= bc2n

1−( ca2 )n c2 −a

, if

nban−1 , if

209

a = c2 = 0, a = c2 .

2

In the case b = 0, we have vn = 0 for all n. Therefore, we consider only the case b > 0. √ From (9.2.10) we obtain that for a < c2 , vn is of the same order as ℓn . This makes it possible to use a simple normalization by ℓn . Define for c = 0,  n := Ln /cn . L

(9.2.11)

 n = 1, and Var(L n ) → Then E L recursion n L

b c2 −a .

K

1   (i) = Xi Ln−1 , c i=1 d

 n satisfies the modified Moreover, L (9.2.12)

(i)

n−1  (i) := Ln−1 . Define D2 to be the set of distributions on (IR1 , B 1 ) where L n−1 c with finite second moments and first moment equal to one. Next, define the mapping T : D2 → D2 by

T (G) = L

2

K 1

c

Xi Zi

i=1

3

,

(9.2.13)

where the (Zi ) are i.i.d. random variables with distribution G, and such that (Xi ), (Zi ), K are independent r.v.s. Let ℓ2 denote the minimal L2 metric on D2 : & '  d d 2 1/2 ℓ2 (μ, ν) = inf E(V − W ) ; V = μ, W = ν (9.2.14) ⎞1/2 ⎛ 1   −1 2 = ⎝ F (u) − G−1 (u) du⎠ . 0

Here F, G are the distribution functions of μ, ν respectively. If a < c2 , then T is a contraction with respect to ℓ2 . Proposition 9.2.3 Assume that a < c2 . Then for F, G ∈ D2 , ℓ2 (T (F ), T (G)) ≤

8

a ℓ2 (F, G). c2

(9.2.15)

210

9. Mass Transportation Problems and Recursive Stochastic Equations d

d

Proof: Let the r.v.s U (i) = F, V (i) = G, i ∈ IN, be choosen on (Ω, A, P ) in such a way that ||U (i) −V (i) ||2 = ℓ2 (F, G); for all i and K, (Xi ), (U (1) , V (1) ), (U (2) , V (2) ), . . . are all assumed to be independent. Then ( K (2 K ( (1   1 ( (i) (i) ( 2 Xi U − Xi V ( ℓ2 (T (F ), T (G)) ≤ ( (c ( c i=1 i=1 2 ⎛ ⎡2 3 2  ⎤⎞ K K   1 ⎝ ⎣   ⎦⎠ (i) (i) E E Xi U − Xi V = K 2  c i=1 i=1 *K 2      1 (i) (i) 2 K = E E Xi U − V c2 i=1 +

 i =j

= =













E Xi U (i) − V (i) Xj U (j) − V (j) |K ⎦

2K 



1 E EXi2 E U (i) − V (i) c2 i=1 a 2 ℓ (F, G). c2 2

2

3

2 As a consequence of Proposition 9.2.3 it follows that T has exactly one fixed point in D2 with variance equal to b/(c2 − a). The fixed-point equad

tion is given in terms of the independent random variables Z, Zi ∈ D2 , Zi = Z, (Zi ) as follows: K

1 Z = Xi Zi . c i=1 d

(9.2.16)

As a corollary we obtain Theorem 9.2.4 If a = E  n , Z) ≤ ℓ2 (L



 a n/2 c2

K i=1



 Xi2 < c2 , then



b . c2 − a

(9.2.17)

 n converges in distribution to Z. In particular, L Proposition 9.2.5 If K is constant and E 2 ≤ k ≤ h, then E|Z|h < ∞.





K i=1

|Xj |

k



< ck for all

9.2 Convergence of Recursive Algorithms

211

 n can be equivalently represented by Yn in the following form: Proof: L Y0 = 1,

Yn

1 = n c



n 9

Xj1 ,...,jk ,

(j1 ,...,jn )∈{1,...K}n k=1 d

where (Xj1 ,...,jk−1 ,1 , ..., Xj1 ,...,jk−1 ,K ) = (X1 , ..., XK ) (cf. Guivarch (1990)). Moreover, (Yn ) is a martingale, and therefore |Yn |k is a submartingale. K (j) Representing the Yn in the recursive way Yn = 1c j=1 Xj Yn−1 , where (j)

Yn−1 are independent copies of Yn−1 , we have ⎞ ⎛ K  |Xj |k ⎠ E|Yn−1 |k ck E|Yn |k ≤ ⎝E j=1



+

k1 + · · · + kK = k ki ≤ k − 1



 9 K K 9 k kj E |Xj | E|Yn−1 |kj . k1 , . . . , kK j=1 j=1

We can infer from Theorem 9.2.4 that E|Yn |k is uniformly bounded for k ≤ 2. By induction over k ≤ h, we see that the lower-order terms in the above equation are uniformly bounded, say by a constant C. Since E|Yn |k ≥ E|Yn−1 |k , we obtain * 3+ 2K  ≤ C. |Xj |k E|Yn |k ck − E i=1

Therefore, the assumptions of this proposition ensure that E|Yn |k is uniformly bounded for all k ≤ h. The submartingale convergence theorem now yields the existence of an integrable almost sure limit of |Yn |h . Since d n =  n is absolutely h-integrable. Yn , the weak limit Z of L 2 L

We can also obtain a “stability” result for the stationary equation (9.2.16). This will be achieved in terms of the ℓp metrics defined as in (9.2.14) with 2 replaced by p. Suppose we want to approximate the solution S of the equation d

S =

K 

Xi Si

i=1

by the solution of the “approximate” equation S



d

=

K  i=1

Xi∗ Si∗ .

(9.2.18)

212

9. Mass Transportation Problems and Recursive Stochastic Equations

Here we assume without loss of generality that c = 1 and consider the case of independent sequences (Xi ), (Xi∗ ) so that the pairs (Xi ), (Si ) and (Xi∗ ), (Si∗ ) are independent, and K is constant. Proposition 9.2.6 If K is constant, < 1, then

K

∗ i=1 ℓp (Xi , Xi ) < ε, and

ε||S ∗ ||p ℓp (S, S ) ≤ . K 1 − i=1 ||Xi ||p ∗

K

i=1

||Xi ||p

(9.2.19)

Proof: From the definition of S, S ∗ , 3 2K K   Xi Si , Xi∗ Si∗ ℓp (S, S ∗ ) = ℓp i=1

≤ ≤ ≤

K 

i=1

ℓp (Xi Si , Xi∗ Si∗ )

i=1

K 

(ℓp (Xi Si , Xi Si∗ ) + ℓp (Xi Si∗ , Xi∗ Si∗ )

i=1

2K  i=1

||Xi ||p

3

ℓp (S, S ∗ ) + ||S ∗ ||p · ε.

This implies that ℓp (S, S ∗ ) ≤

ε||S ∗ ||p . K 1 − i=1 ||Xi ||p

2

A similar idea for establishing robustness of equations can be found in Rachev (1991, Chapter 19.3). For the case of a random K we replace Proposition 9.2.6 by the following one. 

Proposition 9.2.7 If E   K 2 a=E < 1, then i=1 Xi ℓ2 (S, S ∗ ) ≤

K i=1

(Xi −

ε √ ||S ∗ ||2 . 1− a

2 Xi∗ )



≤ ε2 , EXi = EXi∗ , and

(9.2.20)

Proof: By the triangle inequality and the independence assumption and the assumption EXi = EXi∗ , 2K 3 K   Xi Si , Xi∗ Si∗ ℓ2 (S, S ∗ ) = ℓ2 i=1

i=1

9.2 Convergence of Recursive Algorithms

≤ ℓ2

2K 

Xi∗ Si∗ ,

i=1

K 

Xi Si∗

i=1

3

+ ℓ2

2K 

Xi Si∗ ,

i=1

K  i=1

X i Si

213

3

2 K 331/2 2 2K 31/2   ≤ E ℓ2 (S, S ∗ ) + ||S ∗ ||2 E (Xi − Xi∗ )2 Xi2

=



i=1

i=1

a ℓ2 (S, S ∗ ) + ε||S ∗ ||2 .

Therefore, ε||S ∗ ||2 √ . ℓ2 (S, S ) ≤ 1− a ∗

2 Remark 9.2.8 In the case of constant K and nonnegative Xi , Durrett and Liggett (1983) proved that the stationary solution Z of (9.2.16) has moments of order β if and only if 3 2 K  1 < 0. (9.2.21) EXiβ v(β) = log β c i=1 For β = 2, (9.2.4) is equivalent to the condition a < c2 used in Proposition 9.2.3. In this sense this condition is sharp when using ℓ2 -distances. Guivarch (1990) has shown how to relax the second-moment assumption. Remark 9.2.9 For the normalized recursion (9.2.12) with (Xi ) i.i.d. r.v.s, K being a constant (we assume without loss of generality that c = 1), we can use the form  0 = 1, L

n = L

(j1 ,...,jn



)∈{1,...,K}n

n 9

Xj1 ,...,jk ,

(9.2.22)

k=1

 n is a sum where (Xj1 ,...,jk ) are independent and distributed as X1 ; i.e., L over product weights in the complete K-ary tree; cf. the proof of Proposition 9.2.5. For nonnegative multipliers Xi we also can consider functionals of the type Mn = max Pn

n 9

Xj1 ,...,jk ,

(9.2.23)

k=1

where the maximum is taken over all paths of length n. Taking logarithms, − ln Mn = − max Pn

n 

k=1

ln (Xj1 ,...,jk ) = min Pn

n 

k=1

(− ln (Xj1 ,...,jk )) ,

214

9. Mass Transportation Problems and Recursive Stochastic Equations

and applying Kingman’s subadditive ergodic theorem yields that for some constant β, 1 log Mn → β a.s. n

(9.2.24)

This shows that in some sense the max product weight is not larger in order of magnitude than the average product weight. In some cases, the constant d β is explicitly known, for example, for Xi = U [0, 1], β ≈ −0.23196 (cf. Mahmoud (1992, p. 165)). Remark 9.2.10 In some cases explicit solutions of (9.2.16) are known. (1) If K is constant and d

d 1 c Xi =

a , a − a ) is beta distributed, then β( K K

Z = Γ(a, β) is gamma distributed (cf. Guivarch (1990)). d d 1 K (2) Suppose that Z1 = K i=1 Xi Zi , (Yi ) are i.i.d. r.v.s, X = X1 , and d d d 1 K Y1 = X1 Z1 holds. Then Y1 = K i=1 Yi X. Conversely, if Y1 = K d 1 K 1 i=1 Yi X1 and (Xi ) are i.i.d. r.v.s, then Zi = K j=1 Yj . The K  d 1 K sequence (Zi ) solves the equation Z1 = K i=1 Xi Zi (cf. Durrett and Liggett (1983)). d K 1/ϑ (3) Suppose (Zi ) solves Z1 = i=1 Xi Zi , Xi ≥ 0. Then Yi = Zi Wi , where 0 < ϑ ≤ 2 and Wi are stable r.v.s with index ϑ, satisfy K 

1/ϑ

Xi

d

Yi = Y1 .

(9.2.25)

i=1

To prove (9.2.25), observe that K 

1/ϑ

Xi

1/ϑ

Zi

d

Wi =

i=1

2K  i=1

Xi Zi

31/ϑ

d

1/ϑ

W1 = Z1

W1 = Y1 .

This interesting transformation property is used in Guivarch (1990) to reduce the case with moments of Xi of higher order to the case of moments of lower order. K d d (4) If i=1 Xi2 ≡ c2 = 0, then the normally distributed r.v.s Z = Zi = N (0, σ 2 ) satisfy (9.2.16).

(5) If Z solves (9.2.16) and Z is an independent copy of Z, then Z ∗ := Z − Z solves Z



K

1 ∗ ∗ X Z . = c i=1 i i d

9.2 Convergence of Recursive Algorithms

215

Here Xi∗ = τi Xi , and the τi are arbitrary random signs. In particular, d

if K = 2, and the r.v.s Xi∗ = U [−1, 1] are independent, then (9.2.16) d

is solved by Z ∗ := Z − Z, where Z = Γ(2, 21 , 0). n , Remark 9.2.11 The following simulations (Figures 9.1 and 9.2) of L d

d

d

0.4 0.2 0.0

0.0

0.2

0.4

0.6

0.6

0.8

0.8

1.0

1.0

with K = 2, X1 , X2 independent r.v.s, X1 = X2 = U [0, 1], Xi = β(2, 2), show good approximation of the empirical d.f. by the theoretical gamma distribution.

0

1

2

3

4

0

d

FIGURE 9.1. Empirical d.f., X1 = U [0, 1], n = 10, theoretical Gamma Γ(2, 12 , 0)

1

2

3

4

d

FIGURE 9.2. Empirical d.f., X1 = β(2, 2), n = 8, theoretical Gamma Γ(4, 14 , 0)

d

d

0.8

1.0

Remark  1 9  9.2.12 In the case K = 2, X1 , X2 independent r.v.s, X1 = X2 = U − 8 , 8 , no explicit solution of (9.2.16) is known. Nevertheless, the fol n converges very fast to the lowing simulation (Figure 9.3) shows that L  10 and L  12 fixed point of (9.2.16). The empirical distribution functions of L can hardly be distinguished. Therefore, they may be regarded as the limiting  6 is already distribution function. The empirical distribution function of L very close to that limit (cf. Figure 9.3).

0.0

0.2

0.4

0.6

 FIGURE 9.3. Empirical d.f. # of Ln$ for d n = 6, 10, and 12, X1 = U − 81 , 98 -4

-2

0

2

4

6

216

9. Mass Transportation Problems and Recursive Stochastic Equations

Remark 9.2.13 (Branching processes) Equation (9.2.9) includes the Galton–Watson process as special case. A Galton–Watson process is defined by the recursion Z0 = 1,

Zn 

Zn+1 =

Xkn ,

(9.2.26)

k=1 d

d

where Xkn = X are i.i.d. r.v.s, n ∈ IN0 . Define K = X and Xi ≡ 1. Then d

Ln = Zn

(9.2.27)

for all n. This equality can be checked by induction on n. In fact, take first d Z0 = L0 = 1. If Zk = Lk for k ≤ n, then Zn+1

d

=

Ln 

d

Xkn =

k=l

d

=

K  i=1

d

=

K 

⎛ ⎝

Ln−1



k=1

K 

i

i=1 k=1+ i−1 L(j) j=1 n−1

⎞(i)

Xkn ⎠ d

(j)

L

n−1

j=1

d

=

K  i=1



Xkn (i)

Zn−1

⎞(i)

⎜  n⎟ Xk ⎠ ⎝

=

k=1

K 

Zn(i)

i=1

Ln(i) = Ln+1 .

i=1

The assumption a < c2 is equivalent to the condition EX > 1. From (9.2.27) we can derive explicit stationary distributions and even the extinction probabilities in some cases. If, for example, X is geometrically distributed, P (X = k) = p(1 − p)k , k ∈ IN0 , then c = EX = 1−p p > 1 if p < 1−p 1 √ Zn 2 and Var(X) = p2 . The normalized Galton–Watson process Var(Zn ) converges to a (unique) solution of the fixed-point equation 4 X  1 EX(EX − 1) d Z = Zi , EZ = . (9.2.28) EX i=1 Var(X)

p The extinction probability is easily seen to be 1−p . For the normalized continuous part an equation identical to (9.2.28) (but with different variances) is also valid. It is well known that this equation is solved by the geometric stable distribution of order 1, i.e., the exponential distribution. This finally implies √  1 − 2p p 1 − 2p d Z = , (9.2.29) δ0 + exp 1−p 1−p 1−p √ since EZ = 1 − 2p, EZ 2 = 2(1 − p).

9.2 Convergence of Recursive Algorithms

217

(b) A Random Immigration Term In this section we admit an additional immigration term; i.e., we consider the recursion d

Ln =

K 

(i)

Xi Ln−1 + Y,

(9.2.30)

i=1

(2)

(1)

where {(Xi ), Y }, K, Ln−1 , Ln−2 , . . . are independent r.v.s. The analysis of (9.2.30) is essentially simplified if we assume ℓ0 := EL0 , v0 := Var(L0 ), K Var(ℓ0 i=1 Xi + Y ) EY (if c = 1), v0 = , where a < 1.(9.2.31) ℓ0 = 1−c 1−a If c = 1, then EY = 0 and ℓ0 is arbitrary. Lemma 9.2.14 Under the assumption (9.2.31), ℓn = ELn = ℓ0 , vn = Var(Ln ) = v0 , Proof: From (9.2.31), ℓn = c ℓn−1 + EY = vn

EY 1−c

for all n ∈ IN. (9.2.32)

= ℓn−1 ,

Var(Ln ) = EL2n − ℓ2n ⎡  K  2      (j) (i) (i) 2 ⎣ E Xi Xj Ln−1 Ln−1 |K = E E Xi Ln−1 |K + =

i=1

+ E(Y 2 |K) + 2E

2K  i=1

i =j

(i)

Y Xi Ln−1 |K

⎞ ⎛  EXi Xj ⎠ ℓ20 = a(vn−1 + ℓ20 ) + ⎝ i =j

+ EY 2 + ℓ0 2E 2

2K 

= avn−1 + Var ℓ0

i=1

K  i=1

Y Xi

3

Xi + Y

3+

− ℓ20

− ℓ20 3

= vn−1 . 2

Condition (9.2.31) is fulfilled for a two-point distribution of L0 . Indeed, it allows us to use the method in the proof of section 9.2.2(a). A change of the initial condition leads to the necessity to change the method of proof and leads to a great variety of different cases to be considered. We therefore restrict ourselves to (9.2.31) in this section.

218

9. Mass Transportation Problems and Recursive Stochastic Equations

As in section (a), we introduce the operator T : M (ℓ0 , v0 ) → M (ℓ0 , v0 ),

T (G) = L

2K 

Xi Vi + Y

i=1

3

.

(9.2.33)

Here M (ℓ0 , v0 ) is the set of distributions with mean ℓ0 and variance v0 , d

(Vi ) are i.i.d. r.v.s, and the random quantities V1 = G, (Vi ), {(Xi ), Y }, K are independent. Similarly as in Proposition 9.2.3, we obtain the contractive inequality ℓ2 (T (F ), T (G)) ≤



a ℓ2 (F, G).

(9.2.34)

This implies the convergence of Ln to a unique fixed point for the mapping √ T in M (ℓ0 , v0 ) with respect to the ℓ2 -metric. The contraction factor is a. A sharper result (i.e., a smaller contraction factor) is obtained by the use of the Zolotarev metric ζr instead of ℓ2 . Recall the definition for ζr (cf. (9.1.2)): ζr (F, G) = sup{|E(f (X) − f (Y ))|; |f (m) (x) − f (m) (y)| ≤ |x − y|α }(9.2.35) for r = m + α, m ∈ IN0 , 0 < α ≤ 1. Proposition 9.2.15 ζr (T (F ), T (G)) ≤ E

2K  i=1

|Xi |r

3

ζr (F, G).

(9.2.36)

Proof: Recall that ζr is an ideal metric of order r with respect to summation; i.e., ζr (X + Z, Y + Z) ≤ ζr (X, Y ) for Z independent of X, Y , and moreover, ζr (cX, cY ) = |c|r ζr (X, Y ). Then, for (Zi ), (Wi ) being i.i.d. r.v.s distributed according to F, G, we have ζr (T F, T G)

=

2K  3 2K 3       sup Ef Xi Zi + Y − Ef Xi Wi + Y  ;   i=1 i=1  |f (m) (X) − f (m) (Y )| ≤ |x − y|2





ζr

2K  i=1

xi Zi + y,

K  i=1

xi Wi + y

3

dP (X,Y,K) (x, y, k)

9.2 Convergence of Recursive Algorithms



  k i=1

= E

219

|xi |r ζr (Zi , Wi ) dP (X,Y,K) (x, y, k)

2K  i=1

|Xi |r

3

ζr (F, G).

2

Note that for the recursion defined by T the first two moments are matched. Therefore, we can apply (ζr ) with r ≤ 3 and obtain as a corollary the following theorem. EY or c = 1. SupTheorem 9.2.16 Suppose either c = 1 and ℓ0 = 1−c K Var(ℓ0 i=1 Xi +Y ) pose also that EY = 0 and v0 = for a < 1. Then for 1−a 0 < r ≤ 3, the inequality

ar := E

K  i=1

implies

|Xi |r < 1 anr ζr (L0 , L1 ) < ∞, 1 − ar

ζr (Ln , Z) ≤

where Z is a fixed point of T in M (ℓ0 , v0 ). In particular, Ln converges in distribution to Z. Therefore, in the case with immigration we also obtain an exponential rate of convergence. As a consequence, after a few iterations, the limiting distribution is already well approximated. d

1 Consider the following example: L0 = 10 δ−5 + 52 δ0 + 12 δ2 , K = 2, X1 , X2   d d d 5 25 independent, X1 = X2 = U − 21 , 1 , and Y = 17 32 δ−1 + 64 δ0 + 64 δ2 .

0.8

1.0

In this situation the assumptions of Theorem 9.2.16 are fulfilled. The fast convergence is confirmed by the closeness of the empirical distribution functions of L6 and L8 in the simulation described in Figure 9.4.

0.0

0.2

0.4

0.6

FIGURE 9.4. Empirical distribution functions for L6 and L8 ; the difference between the two curves is hardly visible

-4

-2

0

2

4

6

220

9. Mass Transportation Problems and Recursive Stochastic Equations

9.2.3

Limiting Distribution of the Collision Resolution Interval

In this section we apply the method of probability metrics to investigate the contraction properties of stochastic algorithms arising in random-access communication protocols. The results are due to Feldman, Rachev, and R¨ uschendorf (1994); see also Rachev and R¨ uschendorf (1995). The Capetanakis–Tsybakov–Mikhailov (CTM) protocol is one of the most elegant solutions to the classical multiple-access problem, in which a large population of users share a single communication channel. Throughput of this protocol is close to the throughput of the slotted Aloha protocol. The CTM protocol, unlike the classical “slotted Aloha,” is inherently stable. The “tree splitting protocols,” of which the CTM protocol is an example, pose some interesting mathematical problems and have been the subject of intensive study. We briefly review the definition of the CTM protocol; see Bertsekas and Gallager (1987). Time is divided into slots of equal duration. During each slot, one of the following events occurs: 1. The slot is wasted because no one transmits. 2. Exactly one user transmits a message, in which case the message is successfully received. 3. The slot is wasted because two or more users transmit, interfering with each other. This is called a collision. At the end of each slot, every user knows which of these three events occurred (this is sometimes called “trinary feedback”). When a collision occurs, all users involved (those that transmitted during the slot) divide themselves into two groups on a random basis. Each user performs the equivalent of an independent coin toss in order to make its decision; p is the probability that a user selects the first group. Users in the first group retransmit their messages during the slot following the one in which the collision occurred; users in the second group defer their retransmissions until all users in the first group have successfully transmitted their messages. If one of these groups contains more than one user, another collision will occur, in which case this group divides in the same way. Collisions are resolved on a last-come first-served (LCFS) basis; i.e., the most recent collision is resolved before any prior collisions. We assume that new messages are generated according to a Poisson process with aggregate rate λ. Users who have transmitted a message that collided do not generate any new messages until their messages have been transmitted; however, since only a finite number of users are involved in

9.2 Convergence of Recursive Algorithms

221

any collision, the rate λ remains constant when the total user population is infinite. Denote by Ln the number of slots required for resolution of a collision between n users. Ln includes the slot in which the initial collision occurred, plus the times for the two groups of users to transmit their messages. It is easily seen that the following stochastic recursion holds: d  n−I +Y , Ln = 1 + LIn +X + L n

n ≥ 2,

(9.2.37) d

with initial conditions L0 = L1 = 1. Here In = B(n, p) is the number of users who retransmit immediately, X is the number of new arrivals in the collision slot, and Y is the number of new arrivals during the slot in d  which the deferred retransmissions occur. Moreover, Ln = L n , and the  n )n≥0 are assumed to be mutually random quantities X, Y, (Ln )n≥0 , (L independent. For real systems, the total number of users sharing a multipleaccess channel might be as large as 103 or 104 , but the number n of users involved in any collision would be a small fraction of this. Fayolle et al. (1985) showed that limn→∞ ELn /n exists if log p/ log(1 − p) is irrational;

(9.2.38)

otherwise, ELn /n oscillates around a certain value. In a subsequent paper, Fayolle et al. (1986) proved the linearity of the variance of Ln under (9.2.38) and the finiteness of all moments of Ln . Confirming a conjecture of Massey (1981), Regnier and Jacquet (1989) proved that the variance of d

Ln is not linear for In = B(n, p), p = 1/2, and X = Y = 0. In Jacquet and Regnier (1988) and Regnier and Jacquet (1989) the asymptotic normality of the standardized sequence {Ln } (for X = Y = 0 or both Poisson) was established. In this section we examine the asymptotic normality of the law of Ln without the specific assumptions on the distribution type of In , X, and Y , provided that the variance of Ln is asymptotically linear. In the second part of the section we numerically investigate the influence of nonlinearity d in the case In = B(n, p), X = Y = 0, and p = 21 . It turns out that E Ln /n and (Var Ln )/n increase monotonically with n until n reaches a large value (n = 39, 488). After that, the linearity breaks down, in agreement with the theoretical results. We consider the simple normalization √ when ℓn = E Ln . (9.2.39) Yn = (Ln − ℓn )/ n, The main theoretical result indicates that normality holds if the variance behaves linearly and the numbers of retransmissions are not concentrated too much in the extremes. In this sense the result can be considered as a stability result for the asymptotic distribution.

222

9. Mass Transportation Problems and Recursive Stochastic Equations

This idea of stability is confirmed by simulations for some cases of immigration in part (b) of this section. In the numerical study we detect the theoretically predicted instability but only for extremely large n and with a practically negligible order of magnitude. Our simulation study confirms the stability in the standard model concerning dependence on p. Moreover, a simulation study of the empirical d.f. of Yn confirms the normality for 102 ≤ n ≤ 104 . The “instability” of E Ln /n and Var Ln /N and hence the fluctuation of the limit distribution of Yn arises for very large values of n, n ≫ 104 . The order of magnitude of the instability is seen from our numerical results and simulation study to be extremely small (but existent; in accordance with the theoretical results) and can be neglected from the practical point of view. This has the valuable consequence that in practical applications one can use just simple linear normalizations as in (9.2.39) and the normal approximation also for n “moderately” large, 102 ≤ n ≤ 104 . (a) Asymptotic Normality of the Law of Ln In this section the asymptotic normality of Yn is shown under the following assumptions: For some r ∈ (2, 3], (a) E X r/2 + E Y r/2 < ∞ and

In Lr −→ p ∈ (0, 1); n

(b) σn2 = (Var Ln )/n → σ 2 ; (c) supn E|Yn |r < ∞ for some r ∈ (2, 3]. Conditions (b), (c) amount to the correctness of the normalization in (9.2.39). Condition (a) implies that the subgroups are not allowed to be extremely large or small. Note that the number of retransmitting users In is not necessarily binomial in our assumptions. This allows us, for example, to consider departures from independence in the protocol. Regnier and d Jacquet (1989) showed that (a), (b), and (c) hold for In = B(n, p), (9.2.38), d

d

and X = Y = 0. More generally, one can allow X = Y = Pois(λ). Theorem 9.2.17 Under (a), (b), and (c), the distribution of Yn is asymptotically N (0, σ 2 ). Proof: From the definitions of Ln , Yn , and (9.2.37), (9.2.39), Yn

d

=



1/2 In + X YIn +X n  1/2 n − In + Y + Yn−In +Y + Cn (In , X, Y ). n

(9.2.40)

9.2 Convergence of Recursive Algorithms

223

Here Yn is an independent copy of Yn , and

Cn (k, m, m)  := n−1/2 (1 + ℓk+m + ℓn−k+m  − ℓn ).

Define a sequence of normal N (0, σn2 )-distributed independent r.v.s Zn that are independent of (In ), X, and Y , and let Zn∗



=

In + X n

1/2 1/2  n − In + Y n−I +Y + Cn (In , X, Y ), Z ZIn +X + n n

n is an independent version of Zn . Z ∗ is an accompanying sequence where Z n to Yn . Let μr be one of the following ideal metrics of order r > 0: & ' (s) μ(1) (X, Y ) = sup |E(f (X) − f (Y ))|; f  ≤ 1 q r with

μ(2) r (X, Y ) = μ(3) r (X, Y

) =

r = s + 1/ p, s ∈ IN, p ∈ [1, ∞], sup |t|−r |E eit X − E eit Y |;

t∈IR

sup |h|r

h∈IR

1 1 + = 1, p q

sup |P (X + hN ∈ A) − P (Y + hN ∈ A)|,

A∈B(IR)

where N is a standard normal r.v. independent of X and Y . Claim 1. (μr -closeness of Zn∗ and Yn ) Set an = μr (Zn , Yn ) and suppose a := supn an < ∞. Then sup μr(i) (Zn∗ , Yn ) ≤ a[pr/2 + (1 − p)r/2 ].

(9.2.41)

n

(i)

For μr = μr (i = 1, 2, 3), μr (Zn∗ , Yn )  ≤ P (In = k, X = m, Y = m)  k,m,m 

28

8

n−k+m    Zn−k+m  + cn (k, m, m), n 3 8 8 k+m n−k+m   Yk+m + Yn−k+m   + cn (k, m, m) n n * r/2  r/2 +  k+m n−k+m  ≤ P (In = k, X = m, Y = m)a  + n n k,m,m  * r/2  r/2 + n − In + Y In + X . + = aE n n μr

k+m Zk+m + n

224

9. Mass Transportation Problems and Recursive Stochastic Equations

Using assumption (a), the right-hand side of the above inequality converges to a[pr/2 + (1 − p)r/2 ]. Claim 2. (Condition (9.2.41) holds) a ≤ C sup(E|Yn |r + E|Zn |r ) < ∞.

(9.2.42)

n

(Throughout this section, C stands for an absolute constant that can have different values in different places.) For the proof, note that for i = 1, 2, or 3, μr(i) (X, Y ) ≤ C(E|X|r + E|Y |r ) < ∞, provided that E(X j − Y j ) = 0 for j = 1, 2 (see, for example, Rachev (1991, Chapters 14, 15)). Thus (9.2.42) holds. Claim 3. (Asymptotic normality of Zn∗ ) For n → ∞, bn = μr (Zn , Zn∗ ) → 0. (1)

We consider the case μr = μr only. Let κr be the rth pseudomoment,  κr (X, Y ) = r |x|r−1 |FX (x) − FY (x)| dx. IR

Then, since the mean and variance of Zn matched those of Zn∗ (μr (Zn∗ , Yn ) < ∞ implies E((Zn∗ )j − Ynj ) = 0, j = 1, 2), it follows that bn ≤ C κr (Zn , Zn∗ ). Recall that (Zn )n≥1 is independent of (In )n≥1 and X, Y . Let N0 denote a standard normal r.v. independent of (In ) and X, Y . Consequently, 8 8 I + X n − In + Y  n ZIn +X + Zn∗ = Zn−In +Y + Cn (In , X, Y ) n n  1/2 In + X 2 n − In + Y 2 d = N0 + Cn (In , X, Y ) σIn +X + σn−In +Y n n =: ηn N0 + Cn (In , X, Y ). From assumptions (a), (b) we get the convergence of ηn in probability: P

ηn −→ (p σ 2 + (1 − p)σ 2 )1/2 = σ. d

Since Zn∗ = ηn N0 + Cn (In , X, Y ) has the same mean and variance as Zn = σn N0 , then σn2

= E(ηn N0 + Cn (In , X, Y ))2 = E ηn2 + E(Cn (In , X, Y ))2 .

9.2 Convergence of Recursive Algorithms L2

225

P

As ηn −→ σ, we conclude that Cn (In , X, Y ) −→ 0. This implies that bn = μr (Zn , Zn∗ ) → 0, as desired in Claim 3.

With an = μr (Zn , Yn ) ≤ μr (Zn∗ , Yn )+bn and a = lim an we finally obtain from claims 1–3 the following result: Claim 4. a = 0. To prove the claim, choose n0 = n0 (ε) (ε > 0) such that ak ≤ a + ε for k > n0 . Then for n ≥ n0 , as in the proof of Claim 1, we have

an

≤ μr (Zn∗ , Yn ) + bn 3 2n −1 n 0   P (In = k) ≤ + k=n−n0

k=0

×E +

*

k+X n

r/2

n−n 0 −1 

P (In = k)

×E

k+X n

k=n0

*

r/2

+



(ak+X + an−k+Y )

sup 0≤k≤n0 −1, n−n0 ≤k ≤n

n−k+Y n

(a + ε) +



r/2 +

n−k+Y n

r/2

+

(a + ε) + bn .

Recall Claim 2, a = supn an < ∞, and thus as n → ∞, 3 2n −1 n 0   P (In = k)2a E(X r/2 + Y r/2 ) + a ≤ lim sup n

=

k=0

k=n−n0

+ (a + ε)(pr/2 + (1 − p)r/2 ) + lim sup bn

0 + (a + ε)(pr/2 + (1 − p)r/2 ) + 0.

Since r > 2, we have pr/2 + (1 − p)r/2 < 1, which implies that a = 0, and thus the proof of the theorem is complete, since μr -convergence implies weak convergence. 2

Remark 9.2.18 Theorem 9.2.17 shows a remarkable stability of the central limit theorem for Ln . It says that the central limit theorem can be expected if the variance behaves approximately linearly and that it is even true under protocols that are not based on a binomial number of retransmitting users. In concrete examples it is not easy to obtain the asymptotic behavior of the first moments. Our method of proof separates this problem and establishes a general structural stability property concerning the asymptotic distribution. This should be of some interest for the application of the algorithm, too.

226

9. Mass Transportation Problems and Recursive Stochastic Equations

This stability is not clear or expected from the methods that established the central limit theorem up to now in some very special cases. (b) Numerical Results In the first part of this section we study numerically the extent of nonlind earity of ELn , Var Ln in the special case of (9.2.37) where X = Y = 0, In = B(n, p), log p/ log(1 − p) rational.

Initial investigation of the behavior of the mean ℓn of Ln at p = 0.5 failed to show the predicted instability of ℓn /n. The normalized value ℓn /n seemed to converge rapidly, reaching a value of about 2885 for n = 2400, and showing no variation out to 7 decimal places with further increase in n. The increments ℓn /n − ℓn−1 /(n − 1) were observed always to be positive, another indication of convergence. At n = 38, 488, a negative increment appears, and subsequently, values of the increment oscillate in a sinusoidal fashion, with a peak magnitude of about 1 × 10−10 . The behavior of the increments is shown graphically in Figure 9.5 on a logarithmic scale.

FIGURE 9.5. Increments of ℓn /n, p = 0.5, n = number of users who initially collide

Based on recursions for the first moments, the numerical √ results for evaluℓn−1 ℓn ation of ℓn /n, Δ(ℓn /n) := n − n−1 , Varn := Var(Ln / n), and Δ(Varn ) := Varn − Varn−1 are shown in Table 9.1.

On the other hand, a change in the initial conditions disturbs the value of ℓn /n and Varn (see Table 9.2).

Table 9.1 confirms the stability of ℓnn ≈ 2.88, Varn ≈ 3.38 for moderate n ∈ (102 , 104 ) and p = 0.5. Slight perturbation of p around 0.5 does not change the overall stability of ℓnn for practically relevant n; see Figure 9.6. Summarizing the numerical findings, it appears that for reasonably large n ≥ 100 and p = 0.5 the nonlinearity of ℓn /n and Varn is not observed

9.2 Convergence of Recursive Algorithms

n 2 3 4 5 10 100 500 1000 5000 10000

n 2 3 4 5 10 100 500 1000

227

TABLE 9.1. numerical results p = 0.5, L0 = L1 = 1 ℓn /n Δ(ℓn /n) Varn Δ(Varn ) 2.5000D+00 1.5000D+00 4.0000D+00 4.000D+00 2.5556D+00 5.5556D−02 3.2593D+00 −7.4074D−01 2.6310D+00 7.5397D−02 3.3832D+00 1.2396D−01 2.6838D+00 5.2857D−02 3.3875D+00 4.2812D−03 2.7853D+00 1.0985D−02 3.3832D+00 1.1672D−04 2.8754D+00 1.0113D−04 3.3834D+00 9.1046D−07 2.8834D+00 4.0528D−06 3.3834D+00 −8.5624D−08 2.8844D+00 1.0224D−06 3.3834D+00 −4.1963D−08 2.8852D+00 3.4639D−08 3.3834D+00 2.1844D−08 2.8853D+00 7.3428D−09 3.3835D+00 −4.1539D−07

ℓn /n 1.0000E+00 1.1111E+00 1.1905E+00 1.2419E+00 1.3427E+00 1.4327E+00 1.4407E+00 1.4417E+00

TABLE 9.2. p = 0.5, L0 = L1 Δ(ℓn /n) Varn 1.000E+00 1.000E+00 1.1111E−01 8.1481E−01 7.9365E−02 8.4580E−01 5.1429E−02 8.4688E−01 1.1048E−02 8.4579E−01 1.0107E−04 8.4586E−01 4.0304E−06 8.4586E−01 1.0117E−06 8.4586E−01

=0 Δ(Varn ) 1.000E+00 −1.8519E−01 3.0990E−02 1.0703E−03 2.9179E−05 2.2762E−07 −2.1503E−08 −1.2471E−08

FIGURE 9.6. |Δ(ℓn /n)|, p = 0.499

in a practically relevant magnitude. Also, in this range of values of n the behavior of ℓn and Varn is stable with respect to p. The following simulations (Figures 9.7, 9.8, and 9.9) show a good agreement with the normal approximation for n ≥ 100. For n = 20 or n = 30 the normal fit is no longer good. Further simulation results indicate stability with respect to the value of p.

228

9. Mass Transportation Problems and Recursive Stochastic Equations

FIGURE 9.7. Simulation curve for Yn = (Ln − ℓn )/σn for n = 1000, p = 0.5, L0 = L1 = 1, based on 936,725 trials, and the fitted normal curve with mean zero and variance 3.3834 as given in Table 9.1

FIGURE 9.8. Normal fit to empirical d.f. with n = 50, p = 0.5

FIGURE 9.9. Normal fit with σ 2 = 3.3874 to the simulated Yn ’s; n = 1000, p = 0.49, L0 = L1 = 1 based on 697,675 trials

In the final simulations (Figures 9.10, 9.11) we consider the case with nonzero immigrations X, Y in a symmetric and a nonsymmetric case with masses in 0,1,2. These examples confirm the general robustness idea that asymptotic normality is approximatively valid if the variances behave approximately lin-

9.2 Convergence of Recursive Algorithms

FIGURE 9.10. Normal fit for n = 40/100 and X ∼ (nonsymmetric case)

3 δ 4 0

229

+ 18 δ1 + 81 δ2 , Y ∼ δ0 ,

3 δ1 + FIGURE 9.11. Normal fit for n = 50/10000 and X ∼ Y ∼ δ0 16 metric case)

1 δ 16 2

(sym-

early (which is observed in these examples empirically).

9.2.4

Quicksort

The quicksort algorithm, which was introduced by C.A.A. Hoare in 1961– 1962, represents a standard sorting procedure in computer systems. From a list of n arbitrary (but different) real numbers it selects an element x randomly. Then the remaining numbers are divided into two groups, the group of numbers smaller and that of numbers larger than x. The same procedure is applied to each of these groups if they contain more than one element. The algorithm ends with a sorted list of the original numbers.

230

9. Mass Transportation Problems and Recursive Stochastic Equations

If Ln denotes the number of comparisons in the quicksort algorithm on its way to ordering n elements x1 , . . . , xn , then Ln satisfies the following recursive equation: d

Ln = n − 1 + LIn + Ln−In ,

L0 = L1 = 0,

L2 = 1. (9.2.43)

Here In , n − In are the sizes of the subgroups, and they are assumed to be uniformly distributed on {0, . . . , n−1}. The expectation ℓn = ELn satisfies then the resursion ℓn = n − 1 +

n−1 

P (In = i)(ℓi + ℓn−i ),

i=0

and therefore, n+1 1 ℓn 2 ℓn−1 2(n − 1) + − 4. = + = n+1 n n(n + 1) i n + 1 i=1

This yields ℓn = 2n ln n + n(2γ − 4) + 2 ln n + 2γ + 1 + O(n−1 ln n), (9.2.44) where γ ≈ 0.5772 is the Euler constant. Similarly, for vn = Var(Ln ), we have   2 2 vn = (9.2.45) 7 − π n2 + o(n2 ). 3 The normalized random sequence Yn =

Ln − ℓn n

(9.2.46)

satisfies the recursion d

Yn =

n − In In Y n−In + Cn (In ), YIn + n n

(9.2.47)

where Cn (i) =

n−1 E(Li + Ln−i − Ln ). n

As n → ∞, Inn converges to some random variable τ that is uniformly distributed on [0, 1]. Moreover, Cn (In ) = Cn (n Inn ) can be uniformly approximated as follows: sup |Cn (⌈nx⌉) − C(x)| ≤

x∈(0,1)

6 n ln n + O(n−1 ). n

(9.2.48)

9.2 Convergence of Recursive Algorithms

231

Here C(x) = 2x log x+2(1−x) log(1−x)+1, and ⌈x⌉ is the smallest integer larger than or equal to x (cf. R¨ osler (1991)). As a result we obtain the limiting equation Y

d

= τ Y + (1 − τ )Y + C(τ ).

(9.2.49)

In particular, it yields recursive formulas for the moments of Y . Using as an accompanying sequence Yn := τ Y + (1 − τ )Y + Cn (In ), R¨ osler (1991) established the convergence of Yn to Y for the ℓp -metrics. From Proposition 9.1.3 there exists a unique solution Y (in distribution) of the fixed-point equation (9.2.49). Theorem 9.2.19 Let Y denote the solution of (9.2.49). Then ℓp (Yn , Y ) → 0.

(9.2.50)

0.6

The simulation result described on Figure 9.12 shows that the density of Y is very well approximated by a lognormal distribution (cf. Cramer (1996)). The maximal deviation of the fitted lognormal density and the smoothed empirical density is about 0.004.

0.0

0.2

0.4

FIGURE 9.12. Smoothed empirical density of quicksort. Simultaneously, a lognormal approximation is given

-2

-1

9.2.5

0

1

2

3

Limiting Behavior of Random Maxima

A sample of size n is divided into two parts of random size In and n − In , where In is a random variable. We consider the recursion of “maxima type” d

Ln = cn + LIn ∨ Ln−In ,

(9.2.51)

where Ln , Ln are independent and identically distributed r.v.s, (In ) are independent, and (cn ) is a sequence of real numbers. Given α > 0, let us introduce the normalizations Yn = n−1/α Ln ,

Y n = n−1/α Ln .

(9.2.52)

232

9. Mass Transportation Problems and Recursive Stochastic Equations

By (9.2.51), d

Yn = cn n−1/α +



In n

1/α

YIn ∨



n − In n

1/α

Y n−In .

(9.2.53)

Suppose that n−1/α is the right normalization to obtain the weak converD D gence results, Yn −→ Z, Y n −→ Z, and moreover, let Inn → τ , where τ a random variable independent of Z, Z. Then, in the limit, we obtain the fixed-point equation d

Z = τ 1/α Z ∨ (1 − τ )1/α Z.

(9.2.54)

It is easy to check that, for example, the extreme value distribution 

FZ (x) =

e−x 0,

−α

, x > 0, x≤0

(9.2.55)

satisfies (9.2.53). As a motivation for the study of equation (9.3.46), consider Ln = max{X1 , . . . , Xn }, cn = 0, and assume that (Xi ) are i.i.d. r.v.s of Paretian type (F (x) ∼ 1 − x−α for x → ∞). Then by Gnedenko’s extreme-value D theorem, Yn = n−1/α Ln −→ Z, with FZ as in (9.3.50). Note also that formula (9.3.46) concerns some modifications of this recursion, where the maxima are produced by a (random) scheme determined by In (for examd

ple In = B(n, p) ) and cn corresponding to some weighting of the number of steps in this reduction (for example cn = 1). Furthermore, note that (9.2.51) with cn = 1 also describes the maximum search length of a search algorithm dividing a slot of size n succesively into two parts of size In and n − In , respectively.

Define next ak := ℓr (Yk , Z) for 0 < α < r ≤ 1 or ak := ̺r (Yk , Z), ̺r the weighted Kolmogorov metric (cf. (9.1.16)) for 1 ≤ α < r < α + 1, and consider the following assumptions: lim ak cn n−1/α In n


0 if for any c > 0 and independent X, Y , and Z, μ(cX ∨ Z, cY ∨ Z) ≤ cr μ(X, Z); see Rachev (1991) and Rachev and R¨ uschendorf (1991).

234

9. Mass Transportation Problems and Recursive Stochastic Equations

for any r ≤ 1 that is a singleconsequence of the Monge–Kantorovich theorem; recall that ℓ1 (X, Y ) = |FX (x) − FY (x)| dx. Claim 1. bn → 0.

To show the claim, we apply (9.2.57) to obtain 3 2  1/α  1/α I n − I n n Z n−In , Zn ZI n ∨ ≤ ℓ1 n−1/α cn + b1/r n n n 2 3   1/α 1/α n  n−k k Zk ∨ P (In = k)ℓ1 n−1/α cn + Z n−k , Zn ≤ n n k=1 3 2  1/α n  n−k k + Z, Z . P (In = k)ℓ1 cn n−1/α + = n n k=1

In the above bound we have used that the extreme value distributions Zn satisfy the max-stability property:  1/α   1/α 1/α n−k k k n−k d + Z n−k = Zk ∨ Z = Z. n n n n 1/r

Therefore, bn

≤ cn n−1/α → 0, proving the claim.

Applying the triangle inequality and (9.2.58), (9.2.59), we have an ≤ ℓr (Yn , Zn∗ ) + ℓr (Zn∗ , Zn ),

(9.2.60)

and therefore, a ≤ aE(τ r/α + (1 − τ )r/α ) + 0,

with a := lim an ,

(9.2.61)

implying that a = 0. Next, we shall make use of the weighted Kolmogorov metric ̺r (cf. (9.1.16)). It is easy to check that for ε ≥ 1 and X, Y ≥ 0, ̺r (X + a, Y + ε) ≤ (εa)r + εr ̺r (X, Y ).

(9.2.62)

Define then ak = ̺r (Yk , Z), bk = ̺r (Zk∗ , Zk ). By (9.2.62) (with ε = 1), 2   1/α 1/α In n − In ∗ r −r/α + ̺r Z n−In , ZIn ∨ ̺r (Yn , Zn ) ≤ cn n n n 3  1/α  1/α In n − In YIn ∨ Y n−In n n   r/α r/α n − In In r −r/α ≤ cn n aIn + E an−In . +E n n

9.2 Convergence of Recursive Algorithms

235

This implies that lim ̺r (Yn , Zn∗ ) ≤ E(τ r/α + (1 − τ )r/α )a for r > α.

(9.2.63)

If α < r ≤ α + 1, then ̺r (Zn , Zn∗ )



n 

P (In = k)̺r (cn n−1/α + Zn , Zn )

k=1

= ̺r (cn n−1/α + Z, Z). We next prove that ̺r (a + Z, Z) → 0 for a → 0. Let χ(x, a) := xr |FZ (x) − FZ+a (x)| = xr |e−x

−α

− e−(x−a)

−α

−α

Then sup0≤x≤a χ(x, a) ≤ ar , supa≤x≤2a χ(x, a) ≤ (2a)r e−a more, sup xr |e−x

−α

2a≤x≤1

−e

−(x−a)α

| =

sup xr α 2a≤x≤1

x

y −α−1 e−y

−α

|.

. Further-

dy

x−a

1 −x−α e (x − a)α+1 2a≤x≤1 r  x αa e−1 ≤ sup α+1−r (x − a) 2a≤x≤1 x − a ≤

sup xr αa

≤ 2r αar−α e−1 and sup xr α 1≤x 0 we have the Poisson approximation P (τ = n + r) ≈ e−λ λn /n!. This, in turn, leads to a Poisson model for the time of the disastrous event. Assuming that the Ni are i.i.d. Poiss(λ) r.v.s, Ti = N1 +· · ·+Ni is viewed as the time of the ith disastrous event. We shall study in the framework of the model (9.2.66) the distributions of the sums T1 

ST1 =

i=0

Yi

i 9

Zj ,

(9.2.69)

j=0

and Tk+1



STk+1 − STk =

Yi

i=Tk +1

i 9

Zj ,

k = 1, 2, . . . .

j=0

Here for the sake of convenience we start the sequence {(Yn , Zn )} at n = 0. Similarly to (9.2.69) we shall be interested in the laws of Mτ

=

and MT1

=

Zj

i=1

i 9

j=1

T1 @

i 9

Zj

τ @

Yi

Yi

i=0

(9.2.70)

j=0

and Tk+1

@

Yi

i 9

Zj ,

k = 1, 2, . . . .

j=0

i=Tk +1

Note that if Sn (or Sn∗ ) converges in distribution, then the limiting (in distribution) random variable S satisfies the equation d

S = (S + Y )Z.

(9.2.71)

d

Here (Y, Z) = (Yn , Zn ), and the random quantities S and (Y, Z) in the right-hand side of (9.2.71) are independent. In many cases the solution of the equation (9.2.71) turns out to be an infinitely divisible random variable. Similarily, the distributional limit M of Mn satisfies d

M = (M ∨ Y )Z.

(9.2.72)

9.2 Convergence of Recursive Algorithms

239

It is interesting to note that the total “wealth” until the disastrous geometrical event Sτ also satisfies a distributional equation: d

Sτ = (Y + δSτ )Z.

(9.2.73)

d

Here, as before, (Y, Z) = (Yn , Zn ), δ is Bernoulli with P (δ = 1) = 1 − p); S, δ, and (Y, Z) in the right-hand side of (9.2.73) are independent.(7) Similarily to (9.2.73), d

Mτ = (Y ∨ δMτ )Z;

(9.2.74)

if Z ≡ 1, Mτ is said to be max-geometric infinitely divisible.(8)

We start with results on the limiting behavior of the recursions defined above. In the next five theorems {(Yn , Zn )}n≥1 is, unless explicitely stated otherwise, a sequence of nonnegative i.i.d. random vectors such that P (Yn > 0) > 0 and P (Zn > 0) = 1. Set Sn and Mn as in (9.2.66) and (9.2.67), and let Xn = Yn

n 9

Zj .

j=1

Set ξn = log Zn , ν = Eξn (when they exist). Lemma 9.2.21 Let {(Yn , Zn )}n≥1 be a sequence of random vectors living on a common probability space such that {Yn }n≥1 is a sequence of nonnegative i.i.d. random variables with P (Yn > 0) > 0, and {Zn }n≥1 is a sequence of positive i.i.d. random variables. Suppose that E log(1 + Zn ) < ∞. (a) If ν > 0, then with probability 1, Xn does not converge to 0, and thus Sn → ∞. The same is true in the case ν = 0, provided that the sequence {Yn , Zn }n≥1 is a sequence of i.i.d. random vectors. Moreover, in both cases Mn → ∞ (unless P (Zn = 1) = 1). (b) If −∞ < ν < 0, the following are equivalent as n → ∞: (b-i) Xn → 0 a.s.. (b-ii) Sn converges to a finite limit S a.s..

(7) See

Rachev and Todorovich (1990) for some examples of distributions of Sτ ; if Z ≡ 1, Sτ is said to be geometrically infinitely divisible; see Klebanov, Maniya, and Melamed (1984). (8) See Rachev and Resnick (1991).

240

9. Mass Transportation Problems and Recursive Stochastic Equations

(b-iii) Mn converges to a finite limit a.s. (b-iv) 0 < E log(1 + Yn ) < ∞. Moreover, (b-iv) implies (b-i)–(b-iii) even if ν = −∞. The proofs of this and the further assertions in this section can be found in Rachev and Samorodnitsky (1995). Remark 9.2.22 Given a sequence of nonnegative i.i.d. random vectors   (Yn , Zn ) = Yn(1) , . . . , Yn(d) , Zn(1) , . . . , Zn(d) ∈ IR2d , (1)

(d)

we consider the vector of “wealths” Sn = (Sn , . . . , Sn ) given by Sn(k)

=

n 

(k) Yi

i=1

i 9

(k)

Zj ,

n = 1, 2, . . . .

(9.2.75)

j=1

Then Lemma 9.2.21 applied componentwise yields convergence. Our next theorem is the CLT for the “total wealth” Sn in (9.2.66). We assume that ξn = log Zn belongs to the domain of attraction of an α-stable r.v. ηα (1 < α ≤ 2); i.e., there exist an > 0 and bn ∈ IR such that an

n 

D

an = n−1/α L(n),

ξi + bn =⇒ ηα ;

(9.2.76)

i=1

where L(n) is a slowly varying function. Theorem 9.2.23 Suppose that E log(1 + Zn ) < ∞. D

(a) If ν > 0 and E log(1 + Yn ) < ∞, then an log Sn + bn =⇒ ηα . D

(b) If ν < 0 and E log(1 + Yn ) < ∞, then an log(S − Sn ) + bn =⇒ ηα . (c) Let ν = 0, and assume (without loss of generality) that bn ≡ 0. Suppose also that the sequences {Yn }n≥1 and {Zn }n≥1 are independent and that P (log Y1 > 1/an ) = o(n−1 ),

n → ∞.

(9.2.77)

Then, as n → ∞, D

an log Sn =⇒

sup L(t),

(9.2.78)

0≤t≤1

d

where L is a Levy stable motion on [0, 1] with L(1) = ηα .

9.2 Convergence of Recursive Algorithms

241

Remark 9.2.24 The results of Theorem 9.2.23 can be extended both to the multivariate setting and to the form of a functional CLT. We give just one example of such an extension, which is obtainable using Theorem 1 of Resnick and Greenwood (1979).   (1) (2) = In the notation of Remark 9.2.22 let d = 2 and set ξn = ξn , ξn   (1) (2) log Zn , log Zn . Assume that there exists an ∈ IR2+ and bn ∈ IR2 such that   D (1) (2) (2) ξ , a ξ a(1) + bn =⇒ ηα . n n n n (i)

Here α = (α1 , α2 ), 1 < αi ≤ 2, i = 1, 2 and an = n−1/αi Li (n), i = 1, 2, (1) where the Li ’s are slowly varying functions. If E log(1 + Zn ) < ∞ and (i) ν (i) := E log Zn > 0 for i = 1, 2, then &  ' D (1) (2) (2) a(1) =⇒ {L(t)}t≥0 ; log S , a log S + b n n [nt] n [nt] t≥0

  the weak convergence is in the space D [0, ∞), IR2 . {L(t) =  (1)  d L (t), L(2) (t) , t ≥ 0} is a L´evy process with L(1) = ηα and such that  ' & d 1/α2 (2) 1/α1 (1) L (1) + β(t), t ≥ 0 L (1), t t {L(t), t ≥ 0} =

for some β(t) ∈ IR2 prescribed by the marginal convergence. Moreover, (i) if (α1 , α2 ) = (2, 2), then L is an IR2 -valued Wiener process;

(ii) if (α1 , α2 ) = (α, 2), 1 < α < 2, then {L(t) = (L(1) (t), L(2) (t)), t ≥ 0}, where L(1) is an α-stable process and L(2) is a Wiener process independent of L(1) ;

 (iii) if 1 < αi < 2, i = 1, 2, then L has L´evy measure ℓ defined by ℓ◦T = ℓ, where   1/α1 1/α2 T X = (sign x1 )|x1 | , (sign x2 )|x2 | . The measure ℓ is determined by

 ∈ IR2 ; |x| > r, θ(x) ∈ H} = r−1 S(H) ℓ{x

for r > 0; H is a Borel subset of [0, 2π], where |x| and θ(x) are the polar coordinates of x ∈ IR2 , and S is a finite Borel measure on [0, 2π].(9) (9) A

more detailed analysis of L can be obtained using further the results of Resnick and Greenwood (1979) and de Haan et al. (1984). An even more general case where an

242

9. Mass Transportation Problems and Recursive Stochastic Equations

Propositions (b) and (c) of Theorem 9.2.23 can be extended in a similar fashion. As far as the maximal “wealth change” (9.2.65) for n years is concerned, we have the following analogue of Theorem 9.2.23.(10) Theorem 9.2.25 Under the assumptions of Theorem 9.2.23 the following hold: D

(a) If ν > 0, then an log Mn + bn =⇒ ηα . D

(b) If ν < 0, then an log(∨j>n Xj ) + bn =⇒ ηα . D

(c) If ν = 0 and (9.2.78) hold, then an log Mn =⇒ sup0≤t≤1 L(t). Next, we examine the geometric random sum Sτ as defined above. We say that ξn = log Zn belongs to the domain of attraction of a geometric α-stable r.v. Gα if there exist functions a = a(p) > 0 and b = b(p) on [0, 1] such that a

τ  i=1

D

(ξi + b) =⇒ Gα

as p → 0.

(9.2.79)

Here a(p) = p1/α L(1/p), where L is slowly varying function.(11) Theorem 9.2.26 Suppose that E log(1 + Zn ) < ∞ and (9.2.79) holds. D

(a) If ν > 0 and E log(1+Yn ) < ∞, then a(log Sτ +τ b) =⇒ Gα as p → 0. (b) If ν < 0 and E log(1 + Yn ) < ∞, then a(log as p → 0.



j≥τ +1

D

Xj + τ b) =⇒ Gα

(c) Let ν = 0 and b ≡ 0. Assume also that the sequences {Yn }n≥1 and {Zn }n≥1 are independent and P (log Y1 > n1/α L(n)−1 ) = o(n−1 ),

n → ∞.

is a (2 × 2) matrix can be treated using the theory of operator stable random vectors; see Meerschaert (1991). (10) Extensions similar to the ones discussed in Remark 9.2.24 are possible here as well; we may use the multivariate extreme value theory as in de Haan and Resnick (1977). (11) The ch.f. of G admits the representation f α Gα (t) = 1/(1 − log φα (t)), where φα is the ch.f. of an α-stable r.v. (Klebanov, Maniya, and Melamed (1984)). Similarly, fξn = 1/(1 − log ψ), where ψ is the ch.f. of a distribution in the domain of attraction of an α-stable r.v. with ch.f. φα (Mittnik and Rachev (1991)). Examples of geometric α-stable distributions are the exponential law (α = 1) and the Laplace law (α = 2).

9.2 Convergence of Recursive Algorithms

243

Then D

a log Sτ =⇒

sup G(t)

0≤t≤1

as p → 0.

(9.2.80)

Here G is a “geometric L´evy stable motion”; i.e., the weak limit in D[0, 1] [τ t] of Gp (t) = a j=1 ξj , 0 ≤ t ≤ 1. Remark 9.2.27 Regarding the existence of the process G as the weak limit of Gp , one can check the following:

(a) The finite-dimensional distributions Gp (t1 ), . . . , Gp (td ) (0 ≤ t1 < · · · < td ≤ 1) converge to “geometric strictly stable distributions” G(t1 ), . . . , G(td ) with ch.f. g(θ) of the form 1/(1 − log ψ(θ)), where ψ(θ) is the ch.f. of a strictly α-stable random vector on IRd . (b) The set of laws of Gp (·) (0 < p < 1) is tight. Remark 9.2.28 Under the assumptions listed in Remark 9.2.24 we also have '  & w (1) (1) (2) (2) =⇒ {L(νt)}t≥0 , an S[τ t] , an S[τ t] + τn tbn t≥0

where τn = τ (1/n) and ν is an exponential random variable with mean 1 independent of the bivariate L´evy process L. An important observation is that the above limit relation remains true if we choose any sequence of D positive integer-valued random variables τn such that τn /n ⇒ τ , where τ is a positive random variable. Choosing therefore different laws for τ , we arrive at different models for the total “asset value process.” We list below some of these models, assuming that L is a zero mean bivariate Wiener process; that is, we are in case (i) discussed in Remark 9.2.24. (a) If τn is a mixture (by the values of a fixed random variable U ) of Poisson random variables with mean nU , then we may take τ = U , and L(τ ·) is a mixture of Wiener processes; see Boness et al. (1979). √ (b) If τ = 1/ Xm , where Xm is a chi-square random variable with m degrees of freedom, then the one-dimensional marginals of L(τ ·) are Student’s t distributed; this model was used in Blattberg and Genodes (1974) to model stock prices. (c) If τ is positive strictly stable with index α/2, 0 < α < 2, then L(τ ·) is an α-stable motion. This subordinated process was used in Mandelbrot and Taylor (1967) to explain the nonnormality of stock price changes.

244

9. Mass Transportation Problems and Recursive Stochastic Equations

(d) If τ is a lognormal random variable, then L(τ ·) is the Clark (1973) alternative to the Mandelbrot and Taylor (1967) subordinated process. Note that in contrast to (c), here L(τ ·) has finite variances. Similarly to Theorems 9.2.25 and 9.2.26, we obtain the following limit theorem for the distribution of the maximal “wealth change.” Theorem 9.2.29 Under the assumptions of Theorem 9.2.26 the following holds: D

(a) If ν > 0, then a(log Mτ + τ b) =⇒ Gα as p → 0. D

(b) If ν < 0, then a(log ∨j≥τ +1 Xj + τ b) =⇒ Gα as p → 0. (c) Suppose that the conditions of Theorem 9.2.26(c) hold. Then as p → 0, D

a log Mτ =⇒

sup G(t).

0≤t≤1

Finally, let us consider the total “wealth” until a Poisson (λ) random moment T = T (λ). Let the sequence {Yn , Zn }n≥0 be as before and independent of T . Suppose also that the ch.f. fξn of ξn = log Zn satisfies lim |u|−α (1 − fξn −a (u)) = μ

u→0

(9.2.81)

for some μ > 0, real a, and 1 < α ≤ 2. Note that a = Eξn , and at least when α = 2, (9.2.81) is equivalent to assuming that the ξn ’s are in the domain of normal attraction of an α-stable distribution (Feller (1971, p. 596)). Theorem 9.2.30 Suppose that E log(1 + Zn ) < ∞ and (9.2.81) holds. Let ST =

T  i=0

Xi =

T  i=0

Yi

i 9

Zj .

j=0

(a) If ν = Eξn > 0 and E log(1 + Yn ) < ∞, then as λ → ∞, D

λ−1/α (log ST − aT ) =⇒ Y(α) , where Y(α) is a symmetric stable r.v. with ch.f. exp(−μ|θ|α ).

9.2 Convergence of Recursive Algorithms

245

(b) If ν < 0 and E log(1 + Yn ) < ∞, then as λ → ∞, ⎛

λ−1/α ⎝log ⎛

λ−1/α ⎝log

∞ 

j=T1 +1

D

Xj − aT ⎠

=⇒ Y(α) ,

Xj − aT1 ⎠

=⇒ Y(α) ,

j=T +1 Tk 





D

where T1 , Tk are as in (9.2.69). (c) Let ν = 0 and suppose that the sequences {Yn }n≥0 and {Zn }n≥0 are independent and P (log Yn > u) = o(u−α ) as u → ∞. Then as λ → ∞, D

λ−1/α log ST =⇒

sup L(t),

0≤t≤1

d

where L(·) is a L´evy stable motion on [0, 1] with L(1) = Y(α) . Analogous theorems can be established for the limit distributions of MT1 Tk+1 and ∨i=T Xi . k +1

The remaining results in this section deal with characterizations of the limit laws of Sn (cf. (9.2.66) and (9.2.64)) and Mn (cf. (9.2.67)), which can arise for any given distribution of Zn ’s in a given parametric family of distributions. We will assume that the sequences {Yn }n≥1 and {Zn }n≥1 are independent. Also, we will concentrate our attention on the distributions of Zn supported by (0.1).(12) Invoking Lemma 9.2.21(b), we conclude that Sn (resp. Mn ) converges to a finite limit S (resp. M ) if (and only if, in the case E log Zn > −∞) 0 ≤ E log(1 + Yn ) < ∞.

(9.2.82)

Given (9.2.82), the limits S and M satisfy the equations (9.2.71) and (9.2.72).(13) We start with a characterization of the class S1 (resp. M1 ) of laws L(S) of S (resp. L(M )) such that for any L(Z) ∈ Z1 there exist Y = Y (Z) (12) The

case Zn ∈ (0, 1) a.s. corresponds to “deteriorating environment,” the case being close to the soil erosion model of Todorovic and Gani (1987). (13) Moreover, the converse is also true. Namely, if S, (Y, Z) (M, (Y, Z) respectively) is a solution of (9.2.71) (or (9.2.72) respectively), then the distribution of S(M ) is equal d

to the limiting distribution in the model (9.2.66) ((9.2.67) respectively) with (Yn , Zn ) = (Y, Z). This is a simple consequence of the uniquenes principle (the so-called Letac principle); see Letac (1986) or Goldie (1991).

246

9. Mass Transportation Problems and Recursive Stochastic Equations

such that (9.2.71) (resp. (9.2.72)) holds. The class(14) Z1 of Z-laws L(Z) consists of distributions on (0, 1) with densities fα (u) = (1 + α)z α ,

0 < z < 1, α ≥ 0.

(9.2.83)

In the sequel, for any 0 < β < 1 and an r.v. Y , define ⎧ ⎨ 0 with probability 1 − β, Yβ := ⎩ Y with probability β.

(9.2.84)

A complete description of the class Z1 is given in the following theorem. d

Theorem 9.2.31 The class S1 of the laws L(S) solving S = (S + Y )Z consists of all nonnegative infinitely divisible r.v.s S with Laplace transform ⎫ ⎧ ∞ ⎬ ⎨  1 (1 − e−θx )MS ( dx) . φS (θ) = exp − ⎭ ⎩ x 0

Moreover, the L´evy measure MS is of the following form: MS ≪ Leb

and

MS ( dx) = H(x) dx,

where H(0) ∈ [0, 1], H is nonincreasing on [0, ∞) and vanishing at ∞. The corresponding Y has 1 − H as its distribution function. Remark 9.2.32 Suppose that S is a solution of (9.2.71) for a given Y with 0 ≤ E log(1 + Y ) < ∞ and Z uniform. Then then S is also a solution of (9.2.71) with Z having density (9.2.83) and Y replaced by Y1/(1+α) . Note that Z1 is a subclass of the class of self-decomposable random variables; see Vervaat (1979). Also, allowing α in (9.2.83) to take values in the whole range (−1, ∞) would have made the class Z1 degenerate (consisting of Z = 0 a.s.). Remark 9.2.32, in particular, has no counterpart for α’s in the range (−1, 0). Our next task is the characterization of the class M1 of laws L(M ) such that for every L(Z) ∈ Z1 there exists Y = Y (Z) such that (9.2.72) holds. Theorem 9.2.33 The class M1 consists of all absolutely continuous laws L(M ) with density fM and d.f. FM satisfying the following conditions: (14) The

class Z1 was considered by Vervaat (1979) (who discussed a wider family, allowing α > −1 in (9.2.83)) and Todorovich and Gani (1987). Some particular examples of laws L(S) ∈ S1 , L(M ) ∈ M1 , were studied by Todorovich and Gani (1987), Todorovich (1987), and Rachev and Todorovich (1990).

9.2 Convergence of Recursive Algorithms

247

(i) fM (x) is nonincreasing on (0, ∞). (ii) x fM (x)/FM (x) is nonincreasing on (0, ∞). Suppose that L(M ) ∈ M1 and let Z (α) have density fZ α (z) = (1 + d

α)z α , 0 < z < 1. Then M = (M ∨ Y )Z (α) is equivalent to F Y (x) =

1 x fM (x) , 1 + α FM (x)

x > 0.

(9.2.85)

By (9.2.85), for any L(M ) ∈ M1 and 0 < α < 1, d

d

M = (M ∨ Yα )Z (α) ⇐⇒ M = (M ∨ Y0 )Z (0) , where Yα is determined by (9.2.84). The last relation is parallel to the corresponding relation in the scheme of summation (cf. Remark 9.2.32). Note also that gamma Γ(p, λ)-distributions with 0 < p ≤ 1 belong to M1 , while those with p > 1 do not. Next, we consider the class S2 (resp. M2 )(15) of laws L(S) (resp. L(M )) such that for every L(Z) ∈ Z2 ≡ {δz , 0 < z < 1} there is a Y = Y (Z) such that (9.2.71) (resp. (9.2.72)) holds. Theorem 9.2.34 The class S2 coincides with the family of all nonnegative infinitely divisible r.v.s with Laplace tranform of the form ⎫ ⎧ ∞ ⎬ ⎨ 1 −tx (1 − e )MS ( dx) , φS (t) = exp −at − ⎭ ⎩ x 0

where a ≥ 0 and the L´evy measure MS ≪ Leb is absolutely continuous, whose Radon-Nikodym derivative is nonincreasing a.s. For any S ∈ S2 and z ∈ (0, 1), the corresponding Y in the equation d

S = (S + Y )z is a nonnegative infinitely divisible r.v. with Laplace transform ⎫ ⎧ ⎬ ⎨ at(1 − z) ∞ 1 −tx − (1 − e )MY ( dx) . φY (t) = exp − ⎭ ⎩ z x 0

(15) It

turns out that the class S2 coincides with the class L of Khinchine (cf. Feller (1971, Sect. 8, Chapter XVII) of nonnegative r.v.s. We shall state here a more explicit description of S2 than that in Feller (1971, Theorem XVII.8). Moreover, S1 ⊂ S2 ; see also Vervaat (1979, Remark 4.9). The class of M2 coincides with the class of the laws of max self-decomposable r.v.s (see Balkema et al. (1990) and the references there). The next theorem, similar to the Mejzler (1956) result, is based on a characterization of the weak limits of the normalized maxima an {max(X1 , X2 , . . . , Xn ) − bn } when the Xi ’s are independent and nonidentically distributed.

248

9. Mass Transportation Problems and Recursive Stochastic Equations

Moreover, MY ≪ Leb, and dMS dMS dMY (x) = (zx) − (x), dλ dλ dλ where λ = Leb. Theorem 9.2.35 The class M2 consists of the laws of positive absolutely continuous r.v.s M such that xfM (x)/FM (x) is a nonincreasing function on (0, ∞). Also, M1 ⊂ M2 .

9.2.7

Random Recursion Arising in Probabilistic Modeling: Rate of Convergence

Throughout this section, (B, || · ||) is the separable Banach space C(T ) of continuous mappings x : T → IR, where T is compact and || · || is the usual supremum norm in C(T ). For any x, y ∈ B we set (x · y)(t) = x(t) · y(t), (x ∨ y)(t) = x(t) ∨ y(t), t ∈ T . Given a nonatomic probability space, let X (B) be the space of all random fields (r.f.s) X of B-valued random variables, and let L(B) be the space of all laws PX . Suppose {(Yn , Zn )}n≥1 is a sequence of i.i.d. pairs of r.f.s, and define Sn =

n 

Xi ,

Xi = Yi

i 9

Zj .

(9.2.86)

j=1

i=1

The r.f. Sn can be interpreted as the “wealth” accumulated in different commodities {At , t ∈ T } for a period of n years. We take T = T (U ), where U is a compact metric space, U = (U, ̺), and T (U ) is the set of all closed subsets (think, for example, of crop-producing areas) t of U endowed with the Hausdorff metric h(t1 , t2 ) = inf{ε > 0; t1 ⊂ tε2 , t2 ⊂ tε1 }. Here tε stands for the ε-neighborhood of t (cf. Hausdorff (1957, Sect. 29)).(16) Similarly, we define the maximal “wealth changes” Mn =

n @

Xi .

(9.2.87)

i=1

Next, we are interested in conditions providing an exponential rate of convergence of Sn and Mn to finite limits S and M , respectively. The rate (16) Then

(T, h) is a compact metric space (cf. Hausdorff (1957, Sect. 29); see also Kuratowski (1966, §21), Kuratowski (1969, §31), and Matheron (1975)).

9.2 Convergence of Recursive Algorithms

249

of convergence of the laws PXn to PX will be expressed, as usual in the Banach space setting, in terms of the Prohorov metric π(X, Y ) := π(PX , PY ) := inf{ε > 0; P (X ∈ A) ≤ P (Y ∈ Aε ) + ε

(9.2.88)

for all Borel subsets A in B},

where Aε is the open ε-neighborhood of A. Further, we shall use also the following metrics and functions in X (B) and L(B):(17) (i) χp -metric in X (B): 1 # $ 1+p χp (X, Y ) := sup tp P (||X − Y || > t) ,

p > 0;

t>0

(ii) χp -minimal metric in L(B): 5p (PX , PY ) χ 5p (X, Y ) = χ

d d  Y ); X,  Y ∈ X (B), X = X, Y = Y }, p > 0; := inf{χp (X,

ωp,N (X)p+1

(iii)

:=

sup tp P (||X|| > t), t>N

ωp (X)

:= ωp,0 (X),

Np (X)

:= {E||X||p }1/(1+p) ,

p > 0.

Note that χ 5p is a metric in L(B), χ 5p ≥ π, and the following convergence criterion holds: if ωp,N (Xn ) + ωp,N (X) → 0

as N → ∞ for any n = 1, 2, . . . ,

(9.2.89)

then D

and

χ 5p (Xn , X) → 0 as n → ∞ if and only if Xn →X as n → ∞ lim lim sup ωp,N (Xn ) = 0.

N →∞ n→∞ (17) (cf.

Pisier and Zinn (1977), de Haan and Rachev (1989), and Rachev (1991c).

250

9. Mass Transportation Problems and Recursive Stochastic Equations

Similarly, if (9.2.21) holds, then P

χp (Xn , X) → 0 as n → ∞ if and only if Xn →X as n → ∞ and lim lim sup ωp,N (Xn ) = 0.

N →∞ n→∞

Theorem 9.2.36 (a) If for some p > 0, Np (Z1 ) < 1 and ωp (Y1 Z 1 ) < ∞,  ∞ then Sn converges in probability to a P -a.e. finite limit S = i=1 Xi . Moreover, χp (Sn , S) ≤ φn (Z1 )ωp (Y1 Z1 ),

(9.2.90)

where φn (Z1 ) := Np (Z1 )n /(1 − Np (Z1 )). (b) Under the above conditions, suppose additionally that {(Yi∗ , Zi )}i≥1 is a sequence of i.i.d. pairs of r.f.s satisfying the “tail” condition ωp (Y1∗ Z1 ) < ∞. Let S ∗ be the limit of Sn∗ , i.e., Sn∗

:=

n  i=1

Yi∗

i 9

j=1

P

Zj → S ∗ .

i ’s are i.i.d. copies of the Z1 independent of the Suppose now that the Z 5p (Y1 Z1 , Y1∗ Z1 ) < ∞. Then as n → ∞,(18) (Yi , Zi )’s and let χ

1 · · · Z n · S) ≤ φn (Z1 )5 π(S ∗ − Sn∗ , Z χp (Y1 Z1 , Y1∗ Z1 ) → 0.

(9.2.91)

Proof: a) First note that χp ≥ κ, where κ is the distance in probability (the Ky Fan metric): κ(X, Y ) = inf{u > 0; P (||X − Y || > u) < u}. P

Indeed, to prove Sn → S it is enough to show that Sn is χp -fundamental. Actually, for k = 1, 2, . . . , χp (Sn+k , Sn ) 2 n+k 3  ≤ Xi = ωp i=n+1

(9.2.92) n+k 

ωp (Xi )

i=n+1

1 ⎧ ⎛ ⎞p+1 ⎫ 1+p ⎪ i−1 n+k ⎨ ⎬  ⎪ 9 EZ1 =z1 ,...,Zi−1 =zi−1 ωp ⎝Yi ≤ zj Zi ⎠ ⎪ ⎪ ⎭ i=n+1 ⎩ j=1

(18) The

CLT.

1 · · · Z n plays the same role as the normalizing scaling in the usual factor Z

9.2 Convergence of Recursive Algorithms

=

n+k 

i=n+1



n+k 





⎣EZ1 =z1 ,...,Zi−1 =zi−1 ⎝

i−1 9

j=1

1 ⎞p ⎤ 1+p

||zj ||⎠ ⎦

251

ωp (Yi Zi )

Np (Z1 )i−1 ωp (Y1 Z1 )

i=n+1

= φn (Z1 )ωp (Y1 Z1 ) → 0 as n → ∞, which indeed implies that Sn is χp -fundamental. The bound (9.2.90) follows by the same arguments we used to show (9.2.92). (b) By definition, χ 5p is the minimal metric with respect to χp , and thus for any joint distribution of Y1 and Y1∗ , ⎛ ⎞ i i  9  9 1 · · · Z n · S) ≤ χp ⎝ Yi∗ Zj , Yi χ 5p (S ∗ − Sn∗ , Z Zj ⎠ . i>n

j=1

i>n

j=1

Now proceed as in (9.2.92) to obtain that the right-hand side is not greater than φn (Z1 )5 χp (Y1 Z1 , Y1∗ Z1 ). We next take the infimum in the last inequality over all joint distribu5p ≥ π to tions of (Y1 , Y1∗ ) with fixed marginals, and use the inequality χ complete the proof of (9.2.92). 2 Theorem 9.2.37 Suppose the Yi ’s and Zi ’s are nonnegative r.v.s. Then P under the assumptions of Theorem 9.2.36a), Mn → M , and moreover, χp (Mn , M ) ≤ φn (Z1 )ωp (Y1 Z1 ) → 0 as n → ∞. If also χ 5p (Y1 , Y1∗ ) < ∞, then under the assumptions of Theorem 9.2.36(b), ⎞ ⎛ i ∞ @ 9 1 · · · Z n · M ⎠ ≤ φn (Z1 )5 π⎝ χp (Y1 Z1 , Y1∗ Z1 ) → 0. Yi∗ Zj , Z i=n+1

j=1

Proof: (a) For any k = 1, 2, . . ., 2 n+k 3  χp (Mn+k , Mn ) ≤ ωp ≤ Xi i=n+1

p

n+k 

ωp (Xi ),

i=n+1

and therefore, Mn → M follows by the same arguments as in the proof of Theorem 9.2.36, and the required bound for χp (Mn , M ) is obtained in the same way as we did in (9.2.92).

252

9. Mass Transportation Problems and Recursive Stochastic Equations

b) With Xi∗ = Yi∗ χ 5p

2

@

i>n

Ci

j=1

1 · · · Z n · M Xi∗ , Z

Zj we have 3

≤ χp ≤ ≤

2

2

@

Xi∗ ,

i>n

sup tp P t>0

 i>n

@

Xi

i>n

2  i>n

3

||Xi∗ − Xi || > t

331/1+p

ωp (Xi∗ − Xi ).

The last inequality follows from the triangle inequality for χp in the space of real-valued random variables. Conditioning as in the proof of Theorem 9.2.36(a), we obtain  i>n

ωp (Xi∗ − Xi ) ≤ φn (Z1 )χp (X1∗ , X1 ).

Passing to the minimal metrics and using again π ≤ χ 5p , we obtain the necessary bound. 2 Suppose N is an integer valued r.v. independent of the Yi ’s and Zi ’s. Then under the conditions of Theorem 9.2.36, π(SN , S) ≤ ψN (Z1 )ωp (Z1 ), where ψN (Z1 ) = (ENp (Z1 )N )/(1 − Np (Z1 )), and moreover, ∗ 1 · · · Z N · S) ≤ ψN (Z1 )5 , Z χp (Y1 , Y1∗ ). π(S ∗ − SN

Similar results on the limiting behavior of of the maximum be obtained as a consequence of Theorem 9.2.36.

>N

i=1

Xi can

Remark 9.2.38 Vervaat (1979) showed the following limiting result for Sn∗ (see (9.2.65)): let un ↑ ∞ as n → ∞ be a sequence of reals, and assume, in addition, that (i) E log+ |B1 | < ∞, E| log |A1 ||2+η < ∞ for some η > 0, μ := E log |A1 | < 0; d

(ii) the solution S ∗ of the equation S ∗ = A1 S ∗ + B1 , S ∗ and (A1 , B1 ) independent, has a density f that is ultimately nonincreasing and such that f (t) = O(t−1 ) as t → ∞;

9.2 Convergence of Recursive Algorithms

253

(iii) there are positive reals b and ε < |μ| and a positive nonincreasing integrable function φ on [1, ∞) such that the function T ← (T+ (x)φ(y))x−1 e(μ+ε)y (where T+ (x) = P (|S ∗ | > x), T (x) = P (S ∗ > x), and T ← is its generalized inverse) is bounded on the set {(x, y); x ≥ b, y ≥ 1}. Then

∞ 

P (S ∗ > un )

n=1



< ∞, = ∞,

implies P (Sn∗

> un i.o.) =



0, 1.

Following our Theorem 9.2.36, let us compare the tail behavior of the distribution of the Sn∗ ’s and their limit S ∗ . We consider again the Banach 5 p be the minimal metric space setting for An , Bn , Sn∗ , and S ∗ . Let ℓp = L with respect to Lp (X, Y ), 0 ≤ p ≤ ∞, L0 (X, Y ) = P (X = Y ), Lp (X, Y ) = {EX − Y p }min(1,1/p) ,

0 < p < ∞,

and L∞ (X, Y ) = ess supX − Y . Then, as in Theorem 9.2.36(a), if for some p ∈ [0, ∞], Np∗ (A1 ) := Lp (Z, 0) < 1 and Np∗ (B1 ) < ∞, then as n → ∞, ℓp (Sn∗ , S ∗ ) ≤

Np (A1 )n Np (B1 ) → 0. 1 − Np (A1 )

In the case of real-valued Sn∗ and S ∗ , the last bound gives us conditions for exponential rate of convergence in the total variation metric and in the ℓp -Kantorovich metrics: ℓ0 (Sn∗ , S ∗ )

=

ℓ1 (Sn∗ , S ∗ ) =

sup |P (Sn∗ ∈ A) − P (S ∗ ∈ A)| → 0;

A Borel ∞

−∞

|FSn∗ (x) − FS ∗ (x)| dx → 0;

254

9. Mass Transportation Problems and Recursive Stochastic Equations

⎞ 1/p ∞ ℓp (Sn∗ , S ∗ ) = ⎝ |FS←n∗ (x) − FS←∗ (x)|p dx⎠ → 0, ⎛

−∞

1 ≤ p < ∞;

and ℓ∞ (Sn∗ , S ∗ ) =

sup |FS←n∗ (x) − FS←∗ (x)| → 0.

0≤x≤1

Here as usual, FSn∗ and FS ∗ are the corresponding distribution functions, and F ← stands for the generalized inverse of F .

9.3 Extensions of the Contraction Method A well-known problem in the theory of probability metrics is the extension of the method of ideal metrics to limit theorems for sums or maxima with “nonregular” normalizations of logarithmic type. Moreover, this problem is quite typical in a wide range of stochastic algorithms, since the logarithmictype normalization is not reflected in the regularity structure of probability metrics, while power normalizations na can be captured easily by ideal metrics of order a. The second difficulty arises when the contraction factors converge to one. In this section we study several examples that show solutions to this problem by the use of a modified version of the contraction method. In Sections 9.3.1 and 9.3.2 we consider the number of inversions for random permutations and the “MAX”-algorithm. In Sections 9.3.3 and 9.3.4 we study successful and unsuccessful searching in binary random trees. Each of these examples needs some special arguments in order to achieve approximation by a limit distribution; so in general, the contraction method cannot be considered an “automatic” method. The advantage of the contraction method is its generality, which allows us, for example, to consider recursions in very general spaces, as well as the fact that it often allows us to obtain quantitative approximations. The examples in this section are due to Cramer and R¨ uschendorf (1996a).

9.3.1

The Number of Inversions of a Random Permutation

Given a permutation σ = (a1 , . . . , an ), the pair (ai , aj ), i < j, is called an inversion if ai > aj . Denote by In the number of inversions in a random permutation of size n. Then the following recursion holds: d

In = In−1 + Xn ,

I1 = 0,

(9.3.1)

9.3 Extensions of the Contraction Method

255

where Xn ∼ U({0, . . . , n − 1}) is uniformly distributed on 0, . . . , n − 1 and the r.v.s In−1 , Xn are independent. This leads to explicit expressions for the moment generating function, the mean, and the variance: Gn (z) = Ez

E In =

1 (1 − z 2 ) · · · (1 − z n ) , · = n! (1 − z)n−1

In

n (n − 1) , 4

Var In =

(n − 1) n (2 n + 5) 72

(9.3.2)

(9.3.3)

(cf. Hofri (1987, pp. 122–124)). For the normalized version In − E In I5n := √ Var In

(9.3.4)

we obtain the following Berry–Ess´een-type result. (Note that we assume that all the occurring random variables are defined on one and the same probability space.) Theorem 9.3.1 For n ≥ 7,   1 5 ̺ In , N (0, 1) ≤ C · n− 2 ,

with C = 2.75 ·

84 6·128

A

(9.3.5)

7 6.

n Proof: Without loss of generality, we assume that In = i=1 Xi , where the Xi are independent, Xi ∼ U({0, . . . , i − 1}). By the Berry–Ess´een theorem (cf. Bhattacharya and Ranga Rao (1976, Th. 12.4)),  5 ̺ In , N (0, 1) ≤ 2.75 

Sn,3 , (Sn,2 )3/2

(9.3.6)

where

Sn,m :=

n 

E|Xk − E Xk |m .

k=1

(9.3.7)

We have, for k ≥ 2, 3

E|Xk − E Xk |

k3 , ≤ 32

Var Xk

k2 − 1 . = 12

3 n by some tedious calculations. This implies that k=1 Var Xk ≥ (n−1) and 36 4 n (n+1) 3 k=1 E|Xk − E Xk | ≤ 128 . Thus, from (9.3.6), we obtain, for n ≥ 7,

256

9. Mass Transportation Problems and Recursive Stochastic Equations

 4 8  3 1 1 n n + 1 6 n− 2 ≤ C n− 2 . ̺ I5n , N (0, 1) ≤ 2.75 128 n − 1 n−1 

2

Recursion (9.3.1) leads to a sum of independent variables and therefore allows the application of the classical tools for the central limit theorem. On the other hand, it is an interesting “test” rate of convergence example for the contraction method, since the contraction factors of the normalized recursion converge to one. Furthermore, the approximation result (in terms of the ζ3 -metric) is of independent interest. It gives the same convergence rate as in Theorem 2.1 uniformly on the set of functions f (In ) with f (3) ∞ ≤ 1, when we study the limiting behavior of In − E I n In := . n3/2

(9.3.8)

Theorem 9.3.2 Let σn2 := Var(In ) and Zn ∼ N (0, σn2 ). Then for some C > 0, and for all n ∈ IN, 1 ζ3 (In , Zn ) ≤ C n− 2 .

(9.3.9)

Proof: First note that In satisfies the modified recursion  3/2 n − 1 d n , In−1 + X In = n

(9.3.10)

Xn n := Xn −E where X . Let the sequence (Zn ) be independent, Zn ∼ n3/2 2 N (0, σn ), and define the accompanying sequence

Zn∗ :=



n−1 n

3/2

Bn . Zn−1 + X

Let Yi ∼ N (0, τi2 ) be independent   3 2 Xi τi2 := σi2 − i−1 ≥ 0. Then σi−1 = Var i i3 d

Zi =



i−1 i

3/2

(9.3.11) of

i , Zi−1 ), (X

Zi−1 + Yi .

where

(9.3.12)

Using the homogeneity of order three of the ideal metric ζ3 , we obtain 2 3  3/2 3/2 n − 1 n − 1 n , n In−1 + X ζ3 (In , Zn ) ≤ ζ3 Zn−1 + X n n + ζ3 (Zn∗ , Zn )  9/2 n−1 ζ3 (In−1 , Zn−1 ) + ζ3 (Zn∗ , Zn ). ≤ n

9.3 Extensions of the Contraction Method

By iteration, using Z1 = I1 = 0, we obtain the “ground estimate” n  9/2  i ζ3 (Zi∗ , Zi ). ζ3 (In , Zn ) ≤ n i=2

257

(9.3.13)

Note that E Zi = E Zi∗ = 0 and E Zi2 = E(Zi∗ )2 . Therefore, by making T (1+ 1 )

α use of the estimate ζr ≤ Γ(1+r) κr for r = m + 2, by (9.3.12) and some calculations (cf. Cramer (1995)) we have 2 3   3/2 3/2 i−1 Bi + i − 1 ζ3 (Zi∗ , Zi ) = ζ3 X Zi−1 , Yi + Zi−1 i i     Γ(2)  B ≤ ζ3 Xi , Yi ≤ κ3 Xi , Yi Γ(4) 0 x2 |FX i (x) − FYi (x)| dx =

−∞



7 1 −5/2 −3/2 i + i . 26 · 32 25

Therefore, by some additional calculations, ζ3 (In , Zn ) ≤



≤ =

n  9/2  i ζ3 (Zi∗ , Zi ) n i=2  n  9/2   i 1 −5/2 7 −3/2 i + 6 2i 5 n 2 2 ·3 i=2    $ # 1 1 7 1 1 n2 + n3 + 6 2 n3 + n4 5 9/2 2 3 2 ·3 4 n   1 7 − 23 √ + O n . · 28 · 32 n

2   32 Note that the contraction factor in this example is of order n−1 only, n and consequently we cannot obtain a uniform bound, implying that we need to estimate more precisely the individual terms. The exponential con√ vergence rate is reduced to the rate n.

9.3.2

The Number of Records

The “MAX”-algorithm determines the maximum element of a random sequence (cf. Hofri (1987, pp. 112–113)). Its complexity is essentially given

258

9. Mass Transportation Problems and Recursive Stochastic Equations

by the number of records in a random permutation. Let Mn denote the number of maxima of a random permutation read from left to right. Then Mn satisfies the recursion d

Mn = Mn−1 + Xn ,

(9.3.14)

1 where  n has a Bernoulli distribution with success probability n , Xn ∼  1X B 1, n , and Xn , Mn−1 are assumed independent. Define M1 = 0. Then d

Mn =

n 

Xi ,

(9.3.15)

i=2

when the (Xi ) are independent. Furthermore, E Mn = Hn − 1, (k)

where Hn = (2)

Hn

n

1 j=1 j k ,

−→ ζ(2) =

n→∞

π2 6

Var Mn = Hn − Hn(2) ,

(9.3.16)

  (1) Hn = Hn = ln n + γ + O n−1 , and

(cf. Hofri (1987)).

Define next the normalized sum Dn := M√n − E Mn . M Var Mn

(9.3.17)

Then as in Section 9.3.1, we obtain the normal approximation, but with a “very slow” logarithmic rate of convergence. Theorem 9.3.3 For all n ∈ IN and some absolute constant C > 0, the following uniform rate of convergence holds:   C D ̺ Mn , N (0, 1) ≤ √ . (9.3.18) ln n

Proof: We invoke the Berry–Ess´een bound (9.3.6), where E Xk = k1 , k3 −3 k2 +4 k−2 3 , and E|X − E X | = . Var Xk = k−1 2 k k k k4 n  n Therefore, k=2 E|Xk − EXk |3 ∼ ln n, and k=2 Var Xk ∼ ln n, leading to (9.3.18). The constant C can be easily explicitly calculated. 2 The normalization of Mn is logarithmic in n. To get a rate of convergence result similar to that in (9.3.18), we shall make use of the ζ3 -metric. AIt turns out that in this example we obtain contraction factors of order ln(n−1) ln n that converge to one. Nevertheless, the method described in the proof of Theorem 9.3.2 can also be applied in this case. To this end, define Bn := Mn√− E Mn . M ln n

(9.3.19)

9.3 Extensions of the Contraction Method

259

Bn and Zn ∼ N (0, σ 2 ), we have Theorem 9.3.4 For σn2 := Var M n   Bn , Zn ) = O √ 1 ζ3 (M . (9.3.20) ln n Bn satisfies the recursion Proof: Indeed, M 8 ln(n − 1) B d n , Bn = Mn−1 + X M ln n

(9.3.21)

n := Xn√−E Xn . Let (Zn ) be independent normally distributed r.v.s, where X ln n Zn ∼ N (0, σn2 ), and let 8 ln(n − 1) n Zn−1 + X Zn∗ := (9.3.22) ln n be the accompanying sequence. Further, let Yn ∼ N (0, τn2 ), and τn2 := σn2 − Then d

Zn =

8

ln(n − 1) 2 n . σn−1 = Var X ln n

ln(n − 1) Zn−1 + Yn , ln n

(9.3.23)

(9.3.24)

and using the same arguments as in Section 9.2, we get Bn , Zn ) ≤ ζ3 (M Bn , Zn∗ ) + ζ3 (Zn∗ , Zn ) ζ3 (M  3/2 ln(n − 1) Bn−1 , Zn−1 ) + ζ3 (Yn , X n ). ≤ ζ3 (M ln n

By iteration, this yields the bound Bn , Zn ) ≤ ζ3 (M



ln 2 ln n

3/2 3/2 n   ln i B2 , Z2 ) + i ). ζ3 (M ζ3 (Yi , X ln n i=3

B2 , Z2 ) < ∞, and since By the moment estimate ζ3 (M i = Var Yi = τ 2 = 1 · i−1 Var X i ln i i2 , we have i ) ≤ ζ3 (Yi , X



 1 3 3  E|Yi | + E|Xi | 6 2 3 √ 1 8 1 1 1 · +√ · √ ; π i i (ln i)3/2 6 i

 3 here we also used the estimate E Xi − 1i  ≤ 1i .

(9.3.25)

(9.3.26)

260

9. Mass Transportation Problems and Recursive Stochastic Equations

From (9.3.25) we finally obtain Bn , Zn ) ≤ ζ3 (M

= ≤ =

9.3.3

 3/2 1 ln 2 · 6 ln n 3 2 √ 3/2 n   8 ln i 1 1 1 1 +√ · √ + · · 3/2 ln n 6 i π i i (ln i) i=3 * √ + n n   1 1 1 8 3/2 √ √ (ln 2) + · + i i=3 i i π 6 (ln n)3/2 i=3   1 3/2 (ln 2) + 2 ln n 6 (ln n)3/2   1 1 −3/2 + O (ln n) ·√ . 3 ln n

2

Unsuccessful Searching in Binary Search Trees

In this and the following section we deal with the analysis of inserting and retrieving randomly ordered data in binary search trees by the contraction method; we refer to Mahmoud (1992) for an introduction to random search tree algorithms. Let Un denote the number of comparisons that are necessary in order to insert a new random element in a random search tree. A search tree is called random if it arises from a random permutation. An element (to be inserted in a tree) is called random if each of the n + 1 free leaves of the 1 of being chosen. tree has probability n+1 Un satisfies the recursion d

Un = Un−1 + Yn ,

U0 = 0,

(9.3.27)

2 where Un−1 , Yn are independent, Yn ∼ B(1, n+1 ). For n = 1, one comparison with the root is necessary. For n ≥ 2, insertion of the (n + 1)th element needs as many comparisons in the n-tree as in the (n − 1)-tree except in the case that one comparison with the nth element is necessary. The probability that no comparison with this element is necessary equals n−1 n+1 .

From (9.3.27) we have E Un = 2 (Hn+1 − 1),

(2)

Var Un = 2 Hn+1 − 4 Hn+1 + 2.

(9.3.28)

9.3 Extensions of the Contraction Method

261

Brown and Shubert (1984) (cf. Mahmoud (1992, p. 76)) proved a central limit theorem for Un making use of the Lyapunov theorem and the method generating functions. Since by (9.3.27), d

Un =

n 

Yi ,

Yi

i=1



2 ∼ B 1, i+1



,

(Yi ) independent, (9.3.29)

this argument can be simplified to yield the following theorem. Theorem 9.3.5 Define n − E Un 5n := U√ . U Var Un

Then for some constant C > 0 and all n, 

5n , N (0, 1) ̺ U



C ≤ √ . ln n Sn,3

(9.3.30)

1 (cf. Mahmoud (1992, p. 77)). There3/2 2 ln n Sn,2 fore, (9.3.30) is a consequence of (9.3.6). 2

Proof: Observe that

∼√

Applying the results of Deheuvels and Pfeifer (1988) we obtain that ln1n 5n by a Poisson distribution. This is the exact order of approximation of U indicates that the logarithmic rate in the Berry–Esseen bound (9.3.30) should give essentially the right order of approximation. The following rate of convergence result, obtained by the contraction method, supports the fact that the logarithmic order is sharp. The contraction method can be applied in the theorem below in much the same way as in Section 9.3.2. We therefore only give a sketch of the proof. For more details we refer to Cramer (1995a). n , and Zn ∼ N (0, σn2 ). n := Un√−E Un , σn2 := Var U Theorem 9.3.6 Define U ln n Then, for some C > 0 and all n ∈ IN, we have ζ3



n , Zn U



C ≤ √ . ln n

n satifies the recursion Proof: U 8 ln (n − 1)  d n = Un−1 + Yn , U ln n

(9.3.31)

Yn − E Y n √ . Yn := ln n

(9.3.32)

262

9. Mass Transportation Problems and Recursive Stochastic Equations

Define then Zn∗ :=

8

and

ln (n − 1) Zn−1 + Yn ln n

τn2 := σn2 −

ln (n − 1) 2 σn−1 = Var Yn . ln n

(9.3.33)

(9.3.34)

Let the normal random variables Wn ∼ N (0, τn2 ) be independent of the sequences (Zn ), (Yn ). Then the sequences 8 ln(n − 1) d (9.3.35) Zn−1 + Wn . Zn = ln n Consequently, as in Section 9.3.2, we have the bound 3/2  3/2   n       ln i ln 2    ζ3 Un , Zn ≤ ζ3 U2 , Z2 + ζ3 Wi , Yi . ln n ln n i=3

(9.3.36)

2 = 0 = E Z2 , Var U 2 = σ 2 = Var Z2 , it follows that Next, since E U 2 

2 , Z2 ζ3 U



1 ≤ 6

      3 3 < ∞. E U2  + E |Z2 |

Furthermore,   3    1   3 ζ3 Wi , Yi ≤ E|Wi | + E Yi  6 * √ 2  3  3 3+ 1 2 2 3 i−1 1 2 2 i−1 √ τi + = + 3/2 6 i+1 i+1 i+1 i+1 π (ln i) * + 4 2 1+ % . ≤ 6 (ln i)3/2 (i + 1) π (i + 1)

Therefore,    ζ3 Un , Zn ≤

≤ as required.

1 1 (ln n)3/2 6 + √

1 (ln n)3/2

1 ln n

 10 8 + √ 81 27 π 2 3 n  4 1 1 1+ % · i+1 3 π(i + 1) i=3



for n ≥ n0 ,

(9.3.37) 2

9.3 Extensions of the Contraction Method

263

Remark 9.3.7 Studying the recursion (9.3.32) we can also obtain rate of convergence under alternative distributional assumptions on Yn (resp. Yn ). For example, if μr is any (r, +)-ideal, simple metric, then (as in (9.3.36)) 



n , Zn ≤ μr U



ln 2 ln n

r/2  r/2  n     ln i   μr U2 , Z2 + μr Wi , Yi . (9.3.38) ln n i=3

This indeed implies that   n , Zn μr U −→ 0,

(9.3.39)

n→∞

provided that the following conditions hold:       (a) μr U2 , Z2 < ∞, μr Wi , Yi < ∞, (b)





μr Wi , Yi = o



1 i ln i



i ≥ 3.

.

(9.3.40)

ε To show (9.3.39) for ε > 0, choose k0 ∈ IN such that μr (Wk , Yk ) ≤ , k ln k for k ≥ k0 . Then lim sup μr n→∞



n , Zn U



≤ lim sup n→∞



ln 2 ln n

r/2

  2 , Z2 μr U

k 0 −1   1 r/2  + lim sup (ln i) μr Wi , Yi r/2 n→∞ (ln n) i=3 n  1 r/2−1 1 ε + lim sup (ln i) r/2 i n→∞ (ln n) i=k 0

n 1  1 ≤ 0 + 0 + lim sup ε ≤ ε. ln n i n→∞ i=k0

In the preceding example of unsuccessful searching, the estimate of the  rate of “merging” of the sequences (Wi ) and (Yi ) in terms of μr (W i , Yi ) is √ of order 1/i(ln i)3/2 , allowing us to reach the convergence rate 1/ ln n.

9.3.4

Successful Searching in Binary Search Trees

Given a random binary search tree as in Section 9.3.3, let Sn denote the number of comparisons to retrieve a randomly chosen element in the tree. Brown and Shubert (1984) derived a formula for P (Sn = k), and Louchard (1987) proved a central limit theorem for Sn using the generating function

264

9. Mass Transportation Problems and Recursive Stochastic Equations

method in Mahmoud (1992, pp. 78–82). We shall next derive a quantitative version of the central limit theorem. Our main tool will be the contraction method and moment formulas based on the following recursion for Sn : d

Sn = 1 + SIn ,

S0 = 0,

S1 = 1.

Here In is independent of (Si ), and P (In = 0) = 1 ≤ j ≤ n − 1.

(9.3.41) 1 n,

P (In = j) =

2j n2 ,

It can be shown that this recursion does not transform itself to a sum of independent random variables as was done in the random search algorithm in Rachev and R¨ uschendorf (1991) (cf. (9.3.59)). Therefore, (9.3.41) does not allow the application of the Berry–Esseen-type or Poisson-type approximation result. In fact, it arises from the recursion n 

P (Sn = k) =

P (Sn = k, j chosen)

(9.3.42)

j=1 n 

n  1  1{i=j} δ1k + 1{ij} P (Si−1 = k − 1) n δ1k  n − i P (Sn−i = k − 1) + 2 n n i=1

=

=

+

n  i−1 i=1

n2

P (Si−1 = k − 1)

n−1

δ1k  2j P (Sj = k − 1) · 2 . + n n j=1

=

An explicit formula for P (Sn = k) is due to Brown and Shubert (1984) (cf. Mahmoud (1992, p. 79)). Making use of the Brown–Shubert result, Mahmoud (1992, p. 80) desired formulas for the first two moments of Sn . The recursion (9.3.41) leads to a direct calculation of those moments, as we shall see in the next proposition. Proposition 9.3.8 (a)

E Sn

(b)

Var Sn

 1 (9.3.43) = 2 1+ Hn − 3. n  # 2 $   Hn 10 1 (2) = 2+ + Hn + 4. (9.3.44) Hn − 4 1 + n n n 

9.3 Extensions of the Contraction Method

265

Proof: = 1 + E (E(SIn |In )) = 1 +

(a) E Sn

= 1+

n−1  k=0

n−1 

P (In = k) E Sk

k=0

2k E Sk . n2

With Qn := n · E Sn , the recursion (9.3.45) leads to Qn = n + n2 n+1 Q1 = 1, which implies Qn+1 = 2n+1 + n+2 n+1 Qn . Iteratively, Qn

= = = = =

k=1

Qk ,

1 + 2 E SIn + E SI2n 1 + 2 (E Sn − 1) +

n−1  j=1

2j E Sj2 . 2 n

With Pn := n · E Sn2 , we obtain Pn = −n + 2 Qn + This yields

n−1

n−1  2k − 1 n + 1 2n − 1 + · n k k+1 k=1 * n + n    2 1 1 (n + 1) − − k+1 k k+1 k=1 k=1 # $ 1 (n + 1) 2 (Hn+1 − 1) − 1 + n+1 2(n + 1) Hn − 3 n.

=

(b) E Sn2

(9.3.45)

n+1 2

Pn+1 −

Pn+1 = 8 Hn −

n 2

Pn =

2 n+1 2

2 n

n−1 j=1

Pj .

+ 2 Qn + Pn . By (a), we now have

10 n − 1 n + 2 Pn , + n+1 n+1

and iterating the above expression, we get Pn =

n   j=1

10 j − 3 8 Hj − j



n+1 . j+1

The relation n  Hj j=1

j

(2)

Hn + Hn2 = 2

leads to an explicit calculation of Pn , which yields (9.3.44).

2

266

9. Mass Transportation Problems and Recursive Stochastic Equations

Our next step is to show that (Sn ) after a logarithmic normalization merges to a sequence of normal r.v.s. Define the following normalized version of (Sn ):

Let

Sn − E Sn √ , Sn := 2 ln n

S0 = S1 = 0.

a(k, n) := 1 − E Sn + E Sk , b(k) := Var Sk , σn2 := Var Sn .

(9.3.46)

(9.3.47)

For our derivation we need the following (so far unchecked): 3 2 3+ * 2  n−1  2k y y − a(k, n)  2 % −Φ (C) lim sup y  Φ %  dy < ∞.(9.3.48)   n2 n→∞ b(n) b(k) k=2

Here, Φ is the standard normal d.f.

Let (Zn ) be independent of (Sn ), and Zn ∼ N (0, σn2 ). Theorem 9.3.9 Suppose that (C) holds. Then there exists a constant K < ∞ such that   K  . (9.3.49) ζ3 Sn , Zn ≤ √ ln n

Proof: Note first that (Sn ) satisfies the recursion 8 ln In  d (9.3.50) Sn = SI + cn (In ), ln n n √ where cn (k) := 1 − E Sn + E Sk / 2 ln n. Define then the accompanying sequence 8 ln In d ZIn + cn (In ). Zn∗ = (9.3.51) ln n Applying the “ideality” properties of the metric ζ3 , we obtain the following recursive bound for ζ3 (Sn , Zn );     ∗   ζ3 Sn , Zn ≤ ζ3 Sn , Zn + ζ3 (Zn∗ , Zn ) (9.3.52) 3 28 8 n−1  ln k  ln k Zk +cn (k) P (In = k) ζ3 Sk +cn (k), ≤ ln n ln n k=0

+ ζ3 (Zn∗ , Zn ) n−1   2 k  ln k 3/2   S , Z ζ ≤ + ζ3 (Zn∗ , Zn ) . k k 3 2 n ln n k=2

9.3 Extensions of the Contraction Method

267

To estimate the (ζ3 )-distance between Zn∗ , and Zn we compute the first two moments of Zn∗ : E Zn∗

n−1 

=

P (In = k) E

k=0

n−1 

=

P (In = k)

k=0

28

3 ln k Zk + cn (k) ln n

1 − E Sn + E Sk √ 2 ln n

= (2 ln n)−1/2 [1 − E Sn + E SIn ] = 0 = E Zn , 2

and similarly, E (Zn∗ ) =

1 2 ln n

Var Sn = Var Sn . Now we obtain

1 1 ≤ κ3 (Zn∗ , Zn ) = 6 2

ζ3 (Zn∗ , Zn )



  x2 FZn∗ (x) − FZn (x) dx.

 √  % Furthermore, FZn (x) = Φ(x/σn ) = Φ x 2 ln n/ b(n) , and FZn∗ (x) =

n−1  k=0

P (Zn∗ ≤ x | In = k) · P (In = k)

3 2 √ √ x 2 ln n − a(k, n) 1 % = + 1[1−E Sn ,∞) (x 2 ln n) n b(k) k=2 √ 2 + 2 1[2−E Sn ,∞) (x 2 ln n). n √ Applying the substitution y = x · 2 ln n, the above implies n−1 

2k Φ n2

ζ3 (Zn∗ , Zn ) ≤

1 [An + Bn + Cn ] , 2 · (2 ln n)3/2

(9.3.53)

where An

:=

1 n

Bn

:=

2 n2

and Cn :=



 2 3   y  2  y 1[1−E Sn ,∞) (y) − Φ %  dy,  b(n)   2 3    y   y 2 1[2−E Sn ,∞) (y) − Φ %  dy,  b(n) 



n−1 3 2 3+ * 2   2 k y y − a(k, n)  2  % % y  − Φ Φ  dy.  n2 b(k) b(n)  k=2

268

9. Mass Transportation Problems and Recursive Stochastic Equations

Invoking the assumption (C), we obtain Cn ≤ MC for all n ∈ IN and a fixed constant MC . For n ≥ n0 we have E Sn ≥ 1 and  2 3     y 1 1  2  y Φ % y 2 1[1−E Sn ,0) (y) dy An ≤ − 1[0,∞) (y)  dy +   n n b(n) √ 3 1 1 1 1 2% ≤ b(n) + · (E Sn − 1)3 −→ 0. · 2√ n→∞ n 3 n 3 π The last bound follows from the follwoing asymptotics: b(n) = Var Sn ∼ 2 ln n, and E Sn ∼ 2 ln n. Therefore, An ≤ MA , and similarly Bn ≤ MB , for all n, and we obtain M , (ln n)3/2

ζ3 (Zn , Zn∗ ) ≤

(9.3.54)

where M is a fixed constant. Next, we need to apply the Euler summation formula (cf. Hofri (1987, p. 19)) to the function f (x) = x ln x, x ≥ 1: n−1  j=1

n m   Bk  (k−1) f (j) = f (x) dx + (n) − f (k−1) (1) + Rm , f k!

(9.3.55)

k=1

1

where (Bk ) are the Bernoulli numbers. In (9.3.55) the term Rm has the form Rm

(−1)m+1 = m!

n Bm ({x}) f (m) (x) dx, 1

{x} = x − ⌊x&,

   m−k where Bm (x) = k≥0 m is the mth Bernoulli polynomial. After k Bk x some calculations, (9.3.55) with m = 2 yields n−1 

j ln j =

j=2

1 2 1 1 n ln n − n2 − n ln n + O(ln n). 2 4 2

(9.3.56)

n−1 ≤ Consider a sufficiently large n0 such that for n ≥ n0 , j=2 j ln j 1 2 1 2 B B 2 n ln n − 4 n . Choose M large enough that  (9.3.49)(with M instead of B, 2 M . So from (9.3.52), K) holds for n < n0 and define K := max M (9.3.54), using inductive arguments and assuming (9.3.49) for all k < n, we obtain the final bound: ζ3



Sn , Zn





n−1  k=2

2k n2



ln k ln n

3/2

K M √ + ln k (ln n)3/2

9.3 Extensions of the Contraction Method

= ≤ ≤

269

n−1  M 2 1 · K k ln k + · (ln n)3/2 n2 (ln n)3/2 k=2 #  $  2K 1 2 1 2 1 + M n ln n − n 4 (ln n)3/2 n2 2 # $ K K K 1 √ K ln n − + = . 2 2 (ln n)3/2 ln n

2

Remark 9.3.10 In the preceding example, a direct proof of the convergence of Sn based on direct application of the method of probability metric seems impossible. We were able to obtain the rate of convergence by induction arguments that use the Euler summation formula in a crucial way. This extension of the contraction technique seems to be potentially useful also for other examples in the theorey of probability metrics.

0.358

0.358

0.360

0.360

0.362

0.362

0.364

0.364

0.366

0.366

Remark 9.3.11 Numerical simulations (for n ≤ 10, 000) indicate that (C) is correct. Let us denote the integral in (9.3.48) for n ∈ IN by f (n). Numerical calculation in the range −25 to 25 (with a Newton–Cote algorithm with precision 10−5 ) leads to the graphs in Figures 9.13 and 9.14 of f (n) against n, respectively against ln(ln(ln n)). These graphs indicate the boundedness of f .

0

2000

4000

6000

8000

FIGURE 9.13. f (n) against n

9.3.5

0.5

10000

0.6

FIGURE 9.14. ln(ln(ln(n)))

0.7

f (n)

0.8

against

A Random Search Algorithm

In this section we consider a random search in a set of n ordered states {1, 2, . . . , n}, starting in the largest state n. Let (Tn ) be an independent sequence of random natural numbers, Tn ≤ n − 1. After one step of the

270

9. Mass Transportation Problems and Recursive Stochastic Equations

search we reach state Tn ≤ n − 1. The search is continued in the smaller set {1, . . . , Tn } in the same way, reaching in the next step the state TTn ≤ Tn − 1. The search ends if state 1 is reached. Let Sn denote the number of steps needed for this random search to reach this final state 1. Then Sn satisfies the recursion d

Sn = 1 + STn ,

S1 = 1.

(9.3.57)

With r.v.s Tn being uniformly distributed on {1, . . . , n − 1}, this model has been used by Ross (1982, p. 118) and Bickel and Freeman (1981) in a search for an estimate of mean number of steps in the simplex method (with n extreme points). For applications to max-search problems we refer to Nevzorov (1988), and Pfeifer (1991). In their setting there are given independent r.v.s X1 , . . . , Xn , and Tn is the largest index k ≤ n − 1 such that Xk > Xn . We add the index 0 to the state space, Tn = 0 meaning that no value larger than Xn occurs. Consider now the r.v.s I1 , . . . , In , where Ik is defined as 1 or 0 as state k is visited by the search process or not. Then d

Sn =

n 

Ik .

(9.3.58)

k=1

Let ai ∈ [0, 1], i ≥ 1, a1 = 1, and consider the special search strategy P (Tn = k) =

2

n−1 9

bm

m=k+1

where bm = 1 − am and Special cases: (a)

Cn−1

3

m=n bm

ak ,

1 ≤ k ≤ n − 1,

(9.3.59)

= 1.

If ak = 1/k, bk = (k − 1)/k, then

αn,k =

2

n−1 9

bm

m=k+1

3

ak =

n−2 1 1 k ··· = , kk+1 n−1 n−1

that is, this special case corresponds to the uniform search on {1, . . . , n−1}; (b)

If ak = 1 − e−αk , bk = e−αk , (α1 = −∞), then αn,k = e−

 n−1

m=k+1

αk

(1 − e−αn ).

9.3 Extensions of the Contraction Method

271

With our choice of the search probabilities in (9.3.59) we can easily see that the random variables d

I1 , . . . , In

are independent, and Ii = B(1, ai ).

(9.3.60)

The above implies that d

Sn =

n 

Ii

(9.3.61)

i=1

is a sum of independent binomial random variables; in particular, ESn =

n 

ai ,

Var(Sn ) =

i=1

n 

ai bi .

(9.3.62)

i=1

In the uniform search case this leads to ESn

  1 = log n + γ + O n

(9.3.63)

and π2 +O Var(Sn ) = log n + γ − 6

  1 , n

where γ = 0.5772 is the Euler constant. n n n 2 Suppose that λn = a , and a / i i=2 i=2 i i=2 ai = rn is small for n → ∞. Consider then for the Kolmogorov distance ̺(X, Y ) = sup |FX (x) − FY (x)|

(9.3.64)

x

between Sn − 1 and a Poisson distributed random variable Zn with mean λn . From the results of Deheuvels and Pfeifer (1988) we have the following asymptotic approximation: 2  n 3  3/2  1 p2i / ̺(Sn − 1, Zn ) = √ rn + O max pi , rn2 ; (9.3.65) 2 2πe 2 that is, P (Sn = k + 1) = e−λn λkn /k! + O(rn ).

(9.3.66)

272

9. Mass Transportation Problems and Recursive Stochastic Equations

Some alternative approximations of Sn in terms of various probability metrics were studied in Rachev and R¨ uschendorf (1990).

9.3.6

Bucket Algorithms

Consider now n i.i.d. r.v.s X1 , . . . , Xn with density f on [0, 1], and let i us divide [0, 1] into m intervals Ai = [ i−1 m , m ], 1 ≤ i ≤ m. Let N = (N1 , . . . , Nm ) be the vector of the  number of r.v.s in the “m-buckets” n A1 , . . . , Am ; in other words, Ni = j=1 1Ai (Xj ). The total number of comparisons needed to sort n random numbers by the bucket algorithm is given by

Cn =

 m   Ni i=1

2

1 (Tn − n), = 2

Tn =

m 

Ni2

(9.3.67)

i=1

(cf. Devroye  (1986)). Since N is multinomial M (m; p1 , . . . , pm )-distributed with pi = f (x) dx, we obtain Ai

m

ECn

n(n − 1)  2 = pj . 2 i=1

(9.3.68)

Therefore, in the case of a uniform distribution pj = 1/m and for m/n → α ∈ (0, 1), we have ECn ≈ n/2α. In the general case we have the following asymptotics for the first two moments of Cn :

ECn

n ≈ 2α

1

f 2 (x) dx

(9.3.69)

0

and 2 2 3   1 2 4 2 3 f 2 (x) dx. (9.3.70) + f (x) dx f (x) dx − Var Cn → 2 n α α  2 ≥ f 2 (x) dx α

9.3 Extensions of the Contraction Method

273

We shall demonstrate the method of probability metrics in order to obtain the asymptotic distribution of Cn in the special case m = 2 and n → ∞. We have 32 2 n 32 2 n   ζi , + n− Tn = ζi

1 Cn = (Tn − n), n

i=1

(9.3.71)

i=1

where the ζi ’s are i.i.d. Bernoulli random variables with success probability p. Define the approximating U-statistic based on a normal sample: 1 (Sn − n), = 2

Dn

Sn

32 2 n 32 2 n   = ηi + n− ηi , (9.3.72) i=1

i=1

where ηi ∼ N (p, pq), q := 1 − p. (A detailed analysis of the distribution of Sn can be found in Seidel (1988).) Next, by use of ◦ we shall ◦ √ √ ◦ denote the normalized quantities ζi = (ζi − p)/ pq, ηi = (ηi − p)/ pq, ◦  ◦  ◦ Tn = ( ζi )2 + (n − ζi )2 , etc. ◦



The next theorem provides estimates of closeness of Cn and Dn in terms of the Kantorovich ℓp -metrics for p = 1 and p = 2.

Theorem 9.3.12 For m = 2 and n → ∞, ℓ2

2

Cn Dn , n3/2 n3/2

ℓ1

2

Cn Dn , n3/2 n3/2





3

= O(n−1/4 ),

(9.3.73)

3

= O(n−1/2 ).

(9.3.74)

and ◦



Proof: To show (9.3.73) note that the normalization n3/2 is of the right order. In fact,

Var

2



Dn n3/2

3

= =

◦ 1 Var(n−3/2 Sn ) 4 2 n n ◦ ◦  ◦3 2 i=1 j=1 ηi ηj − 2n ηi 1 Var 4 n3/2

274

9. Mass Transportation Problems and Recursive Stochastic Equations

≈ constant Var



 ◦ 2n ηi ≈ constant > 0. n3/2

Since ℓ2 is the minimal L2 -metric, it follows that ◦



ℓ2 (Cn , Dn ) ≤ L2



 ◦ ◦ 1 ◦ 1 ◦ 1 = (Tn − n), (Sn − n) L2 (Tn , Sn ). 2 2 2

Thus ℓ2

2





Cn Dn , n3/2 n3/2

3

2

3 ◦ ◦ Tn Sn , n3/2 n3/2 ⎛⎛ ⎞  ◦ 1 ⎝⎝   ◦ ◦ 2 ζi ζj + n2 − 2n ζi ⎠ n−3/2 , L2 2 i j ⎛ ⎞ ⎞  ◦ ◦  ◦ ⎝2 ηi ηj + n2 − 2n ηi ⎠ n−3/2 ⎠ 1 L2 2

≤ =

i



≤ L2 ⎝n−3/2 

 i

+ L2 n−1/2 ◦

j

j



◦ ◦

ζi ζj , n−3/2

 i

j



ηi ηj ⎠ ◦ ◦

 ◦ ◦ −1/2 ζi , n ηi =: I1 + I2 .



Assuming that the pairs ( ζi , ηi ) are independent, we obtain

I1 = n−3/2 L2

2  i

◦2

ζi ,

 i

η2i ◦

3



+ n−3/2 L2 ⎝

 i =j

◦ ◦

ζi ζj ,

 i =j



ηi ηj ⎠ ◦ ◦

(9.3.75)

⎛ ⎛ ⎞2 ⎞1/2  ◦2    ◦ ◦ ⎟ ◦2 ◦ ◦ −1/2 −3/2 ⎜ ⎝ ≤ n L2 ζ1 , η1 + n ζi ζj − ηi ηj ⎠ ⎠ ⎝E i =j

⎛ ⎞1/2  ◦ ◦ 2  ◦2   ◦ ◦ ◦ ⎠ = n−1/2 L2 ζ1 , η21 + n−3/2 ⎝ E ζi ζj − ηi ηj i =j

◦ ◦ 1/2  ◦2   ◦ ◦ ◦2 −3/2 = n L2 ζ1 , η1 + n n(n − 1)L1 ζ1 ζ2 , η1 η2    ◦ ◦  ◦ 2  ◦ 2 ◦ ◦ −1/2 ≤ n Var ζ1 + Var ( η1 ) + n−1/2 Var ζ1 ζ2 + Var ( η1 η2 ) −1/2

≤ c n−1/2 .

9.3 Extensions of the Contraction Method

275

Here and in the sequel c stands for an absolute constant, which may be different at different places. Similarly, I22

 ◦ ζi , n ηi (9.3.76) n        ◦  ◦  −1/2  ◦   −1/2  ◦  −1/2 −1/2 ζi  +E n ηi  ζi , n ηi E n L1 n    ◦  ◦ ◦ ◦ −1/2 −1/2 ζi , n ηi Var ζ1 + Var η1 L1 n   ◦  ◦ −1/2 −1/2 ζi , n ηi . 2 L1 n

L22

= ≤ ≤ =



−1/2





−1/2

Passing to the minimal metric, this yields ◦



−1/2

ℓ2 (Cn , Dn ) ≤ c n

 ◦ ◦ √ A + 2 ℓ1 (n−1/2 ζi , η2 ) .

(9.3.77)

The rate of convergence in the CLT for the ℓ1 = ζ1 -metric has been discussed by Zolotarev (1986, Theorem 5.4.7) (see also Rachev and R¨ uschendorf (1990, Lemma 3.3)). It is given by −1/2

ℓ1 (n



 ζi , η1 ) ≤ 11.5 max ℓ1 ( ζ1 , η1 ), ζ2 ( ζ1 , η1 ) n−1/2 , ◦













(9.3.78) ◦



where ζr is the Zolotarev ideal metric of order r > 0. This implies ℓ2 (Cn , Dn ) ≤ C n−1/4 . We can argue similarly to show (9.3.76). The bound (9.3.75) can be replaced by −1/2

I1 ≤ n

     ◦ ◦     ◦ 2 ◦ 2 ◦ ◦ −1/2 E  ζ1  + E | η1 | + n E  ζ1 ζ2  + E | η1 η2 | (9.3.79)

≤ c n−1/2 .

Invoking (9.3.78), we see that the term I2 is of order n−1/2 .

2

We can extend our results to the cases m = 3, 4. However, the proofs are computationally quite involved. For some general results on the asymptotic distributions of quadratic forms we refer to de Jong (1989).

10 Stochastic Differential Equations and Empirical Measures

10.1 Propagation of Chaos and Contraction of Stochastic Mappings In this section we use contraction properties of stochastic mappings with respect to suitably chosen metrics in order to study some new examples of propagation of chaos. In particular, systems of stochastic differential equations (SDEs) with mean field type interactions and the corresponding nonlinear SDEs of McKean–Vlasov type for the limiting cases will be considered. We shall also study the rate of convergence to the corresponding limit. Assumptions on the smoothness and growth properties of the coefficients of the SDEs are to be reflected in the choice of the probability metric in order to obtain the required contraction properties. This allows us to investigate new types of interactions as well as to consider systems with relaxed Lipschitz assumptions.

10.1.1 Introduction The notion “propagation of chaos” was introduced by Kac in his investigation of the relationship between simple Markov models of interacting particles and nonlinear Boltzmann-type equations; for an introduction to the propagation theory of chaos we refer to Sznitman (1989). A formal def-

278

10. Stochastic Differential Equations and Empirical Measures

inition follows. Let (uN ) be a sequence of symmetric probability measures on E N , E a separable metric space, and let u be a probability on E. Then w (uN ) is called u-chaotic if πk un −→ u(k) . Here πk stands for the k-marginal w distribution, u(k) is the k-fold product, and −→ denotes weak convergence. A basic example for chaotic sequences is McKean’s interacting diffusion (cf. the “laboratory” example in Sznitman (1989, p. 172)). Consider a system of interacting diffusions dXti,N X0i,N

=

dWti

= xi0 ,

N 1  b(Xti,N , Xtj,N )dt, + N j=1

i = 1, . . . , N,

(10.1.1)

where the W i are independent Brownian motions and b satisfies a certain Lipschitz condition. Let uN denote the distribution of (X 1,N , . . . , X N,N ). The nonlinear limiting equation is given by the McKean–Vlasov equation dXt = dBt +



b(Xt , y)ut (dy) dt,

(10.1.2)

when Bt is a Brownian motion and ut is the distribution of Xt . Then uN is u-chaotic, where u is the distribution of X on C(IR+ , IRd ). An alternative example of chaotic behavior of particles, not described by SDEs, are uniform distributions on “p-spheres.” Let uN denote the uniform distribution on the p-sphere of radius N in IRN + , that is, on Sp,N := {x ∈ p N IR+ ; Σxi = N }. Let u denote the probability measure on IR+ with density fp (x) =

p1−1/p −xp /p e , Γ(1/p)

x ≥ 0.

Then for N > k + p, and k and N big enough, πk uN − u(k)  ≤

2(k + p) + 1 , N −k−p

(10.1.3)

where  ·  denotes the total variation distance (cf. Kuelbs (1977), Rachev and R¨ uschendorf (1991)). In particular, we obtain that uN is u-chaotic. This example has its origin in Poincar´e’s theorem on the asymptotic behavior of particle systems. More general examples of this kind have been developed in the statistical physics literature in connection with the “equivalence of ensembles” but typically without a quantitative estimate as in (10.1.3).

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

279

The main goal of this section is the study of propagation of chaos in several modifications of McKean’s example. We shall be concerned with the form of the interaction and the regularity assumptions on the coefficients. To this end we introduce suitable probability metrics, allowing us to derive contraction properties of the stochastic equations defined by the corresponding linear equations. Dobrushin (1979) introduced the use of the Kantorovich metric for the interacting diffusions in the model (10.1.1), (10.1.2). The success of this metric is based on a coupling argument inherent in its definition. This metric has been applied since then in several other papers. For some modifications of the model (10.1.1), (10.1.2) we shall need alternatives to the Kantorovich metric that provide suitable regularity and ideality properties for the equations considered. In particular, we need metrics that are “ideal” of higher order when we relax the Lipschitz conditions in equations (10.1.1), (10.1.2). Our modifications allow us to treat much more complicated forms of interactions than those in the McKean example. In particular, we consider nonlinear interactions via some general energy function, as for example the p-norm of the vector of all pair interactions. We also consider interactions with “outside” particles over the whole past (history) of the process, describing some non-Markovian systems. We demonstrate the flexibility of the approach based on suitable probability metrics to analyze with nonstandard forms of interactions and develop the tools to study complex physical systems.

10.1.2 Equations with p-Norm Interacting Drifts Consider a system of N interacting diffusions with p-norm interacting drifts; that is, the drift is given by the pth norm of the vector of all pair interactions (which can be viewed as the driving force in the system):

dXti,N X0i,N

⎫1/p ⎧ N ⎨1  ⎬  dt bp Xti,N , Xtj,N = dWti + ⎭ ⎩N j=1 = X0i ,

1 ≤ i ≤ N,

(10.1.4)

280

10. Stochastic Differential Equations and Empirical Measures

b ≥ 0, p ≥ 1. ((Wti ), X0i ) are independent, identically distributed for all i.) We shall establish that each X i,N has a natural limit X i , where the (X i ) are independent copies of the solutions of a nonlinear equation ⎧ ⎨ ⎩

dXt

= dBt +

Xt=0

= X,



1/p dt, b(Xt , y) ut ( dy) p

(10.1.5)

d

with B = W 1 a process on CT , and ut = P Xt . In order to obtain the necessary contraction properties of these equations, we consider the L∗p -metric and the corresponding minimal L∗p -metric (ℓ∗p ), defined for processes X, Y (or the corresponding probability measures m1 , m2 ∈ M 1 (CT )). Here and in what follows M 1 (CT ) denotes the class of all probability distributions on CT , L∗p,t (X, Y ) := (E sup |Xs − Ys |p )1/p ,

(10.1.6)

s≤t

and d

d

ℓ∗p,t (m1 , m2 ) := inf{L∗p,t (X, Y ); X = m1 , Y = m2 }.

(10.1.7)

In (10.1.7) we tacitly assumed that the underlying probability space is rich enough to support all possible couplings of m1 , m2 , which is true, for example, in the case of atomless probability spaces. Define, for m0 ∈ M 1 (CT ), Mp (CT , m0 ) := {m1 ∈ M 1 (CT ); ℓ∗p,T (m0 , m1 ) < ∞},

(10.1.8)

and let Xp (CT , m0 ) be the class of processes on CT with distribution m ∈ Mp (CT , m0 ). For m0 = δa (the one-point measure in a ∈ CT ), this is the class of all distributions on CT with finite pth moment of the process norm. For m ∈ Mp (CT , m0 ) consider the linear equation corresponding to (10.1.5)

Xt = Bt +

t 0

⎛ ⎝



CT

⎞1/p

b(Xs , ys )p dm(y)⎠

ds,

(10.1.9)

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

281

where ys is the value of y at time s. Let (Bt ) be a real-valued process on CT = C[0, T ] with finite pth absolute moment (E sups≤T |Bs |p < ∞), and let b ≥ 0 be a Lipschitz function in x; that is, |b(x1 , y) − b(x2 , y)| ≤ c|x1 − x2 |,

for all x1 , x2 and Y in CT . (10.1.10)

As usual, a strong solution of the SDE (10.1.9) means a solution measurable with respect to the augmented filtration of the process (Bt ). In constrast, a weak solution of (10.1.9) is defined on a suitable filtered space of distributions.

Lemma 10.1.1 Assume that (10.1.10) holds, and let 1/p   b(0, ys )p dm(y) ds < ∞. (a) Then equation (10.1.9) has a unique strong solution X. (b) If Φ(m) is the law of X, then Φ(m) ∈ Mp (CT , m0 ); that is, Φ : Mp (CT , m0 ) → Mp (CT , m0 ). Proof: Let X ∈ Xp (CT , m0 ), and define

(SX)t := Bt +

 t 

b(Xs , ys )p dm(y)

0

1/p

ds.

Then, for Y ∈ Xp (CT , m0 ),

|(SX)t − (SY )t | ≤ ≤

t 0

 1/p  1/p     b(Xs , ys )p dm(y) − b(Ys , y)p dm(y) ds    

 t  0

1/p ds |b(Xs , ys ) − b(Ys , ys )| dm(y)

t ≤ c |Xs − Ys | ds. 0

p

282

10. Stochastic Differential Equations and Empirical Measures

This implies sups≤t |(SX)s − (SY )s | ≤ c thermore, L∗p,t (SX, SY

) =



t 0

supu≤s |Xu − Yu | ds, and fur-

p

E sup |(SX)s − (SY )s | s≤t

⎛ ⎛

≤ c ⎝E ⎝

t 0

1/p

⎞p ⎞ 1/p

sup |Xu − Yu | ds⎠ ⎠ u≤s

t ≤ c L∗p,s (X, Y ) ds. 0

Define inductively X 0 := B and X n := SX n−1 . Then the above bound yields L∗p,T (X n , X n−1 ) ≤ cn

Tn ∗ Lp,T (X 1 , X 0 ). n!

For the L∗p,T -distance in the right-hand side we have the following estimate: $1/p T#  p p ∗ 1 0  ds E|Bs | + b(0, ys ) dm(y) Lp,T (X , X ) ≤ c 0



≤ c

T 0

E sup |Bu |

p

u≤s



1/p

p 1/p

≤ c T (E sup |Bs | ) s≤T

ds + c

+c

T



0



T 0

1/p ds b(0, ys ) dm(y) p

1/p ds. b(0, ys ) dm(y) p

From the assumptions on B and b, the above bound implies that L∗p,T (X 1 , X 0 ) < ∞. Consequently, ∞ 

n=1

L∗p,T (X n , X n−1 ) ≤ ecT L∗p,T (X 1 , X 0 ) < ∞,

∞ by the Gronwall lemma. This results in n=1 sups≤T |Xsn −Xsn−1 | < ∞ a.s., and therefore, X n converges to some process X a.s., uniformly on bounded intervals. The limiting process X is a.s. continuous, has finite pth moments

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

283

(i.e., X∗p,t := E sups≤t |Xs |p < ∞), and is a fixed point of the mapping S. So, Φ(m) = P X ∈ Mp (CT , m0 ); this holds because B∗p,T < ∞ and 2 L∗p,T (X, B) < ∞. In addition, suppose that b is Lipschitz in both arguments; that is, for all x1 , x2 , y1 , and y2 in CT , |b(x1 , y1 ) − b(x2 , y2 )| ≤ c[|x1 − y1 | + |x2 − y2 |],

(10.1.11)

and consider the mapping Φ : Mp (CT , m0 ) → Mp (CT , m0 ). Lemma 10.1.2 (Contraction of Φ with respect to the ℓ∗p,t -minimal metric) Under the Lipschitz condition (10.1.11) and the assumptions of Lemma 10.1.1, for t ≤ T and m1 , m2 ∈ Mp (CT , m0 ), the following holds: ℓ∗p,t (Φ(m1 ), Φ(m2 )) ≤ cect

t

ℓ∗p,u (m1 , m2 ) du.

(10.1.12)

0

Proof: Let for i = 1, 2 and t ≤ T , (i)

Xt

:= Bt +

t 0

⎛ ⎝



CT

⎞1/p

b(Xs(i) , ys )p dmi (y)⎠

ds,

and let m ∈ M 1 (m1 , m2 ), the class of probability measures on CT × CT with marginals m1 , m2 . Then sup |Xs(1) − Xs(2) | s≤t ⎡ ⎤1/p  ⎤1/p ⎡ t      ⎣ (2) ⎦ (2) (2) p (1) ⎦ (1) (1) p ⎣ ds  ≤ b(Xs , ys ) dm2 (y ) b(Xs , ys ) dm1 (y ) −      CT 0 CT ⎡ ⎤1/p   t p   ≤ ds ⎣ b(Xs(1) , ys(1) ) − b(Xs(2) , ys(2) ) dm(y (1) , y (2) )⎦ 0

≤ c

t 0

CT ×CT



ds |Xs(1) − Xs(2) | +

#

$1/p  |ys(1) − ys(2) |p dm(y (1) , y (2) ) .

284

10. Stochastic Differential Equations and Empirical Measures

Minimizing the right-hand side with respect to all couplings m, we obtain

sup |Xs(1) s≤t



Xs(2) |

t t (1) (2) ≤ c ds sup |Xu − Xu | + c ds ℓ∗p,s (m1 , m2 ). (10.1.13) 0

u≤s

0

Consequently, by Gronwall’s lemma, t sup |Xs(1) − Xs(2) | ≤ c ect ℓ∗p,s (m1 , m2 ) ds. s≤t

(10.1.14)

0

Finally, passing to the pth norm in the left-hand side of (10.1.14) and then to the corresponding minimal metric ℓ∗p,t proves the lemma. 2

Theorem 10.1.3 Under the Lipschitz condition (10.1.11) and assuming T  that ( b(0, ys )p dm0 (y))1/p ds < ∞, equation (10.1.1) has a unique weak 0

and strong solution in Xp (CT , m0 ).

Proof: From Lemma 10.1.2 we obtain that for m ∈ Mp (CT , m0 ), Tk ∗ (cT = c ecT ) ℓp,T (Φ(m), m) k! Tk ∗ (ℓ (Φ(m), m0 ) + ℓ∗p,T (m, m0 )) < ∞. ≤ ckT k! p,T

ℓ∗p,T (Φk+1 (m), Φk (m)) ≤ ckT

Consequently, (Φk (m)) is a Cauchy sequence in (CT , ℓ∗p,T ) and thus converges to a fixed point of Φ. Let X k+1 , X k denote the couplings of Φk+1 (m), Φk (m). Then, by (10.1.12), we have that L∗p,T (X k+1, X k ) ≤ ckT

Tk ∗ ℓ (Φ(m), m), k! p,T

and therefore, we determine a unique strong solution with finite pth moment. 2

Remark 10.1.4 While the linear equation in Lemma 10.1.1 can be handled with the L1 -metric, in Lemma 10.1.2 we obtain only a contraction with respect to the minimal ℓp -metric ℓ∗p,T (cf. equation (10.1.12)).

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

285

Remark 10.1.5 The result of Theorem 10.1.3 can be extended to the case p = ∞ by applying the metric L∗∞,T (X, Y ) = ess sup sup |Xs − Ys |

(10.1.15)

s≤T

and the corresponding minimal metric d

d

ℓ∗∞,T (m1 , m2 ) = inf{L∗∞,T (X, Y ); X = m1 , Y = m2 }.

(10.1.16)

Then the equation

Xt

t = Bt + ess sup b(Xs , y) ds 0

(10.1.17)

us (dy)

has a unique solution in M∞ (CT , m0 ) if B is a.s. bounded, that is, if ess sups≤T |Bs | < ∞. Remark 10.1.6 Several extensions of equation (10.1.5) can be handled in a similar way, as for example

Xt = Bt +

t 0

b(Xs , y)

p

1/p

us(k) ( dy)

ds,

(10.1.18)

Ek (k) Xs where us = stands for the k-fold product of us and y = i=1 P k (y1 , . . . , yk ) ∈ IR . More generally, b = b(s, x, y) can be dependent upon s and the past of the process y = (yu )u≤s . In this case, us has to be replaced by u(s) := P (Xu )u≤s (the distribution of the past), and we need to assume a functional Lipschitz condition on b. In a similar way one can also investigate the d-dimensional case. Taking as a starting point Theorem 10.1.3, we next investigate the system of interacting equations in (10.1.4). The following theorem asserts that as N goes to infinity, each X i,N has a natural limit X i . Here, the (X i ) are independent copies of the solutions of the nonlinear equation (10.1.5). Theorem 10.1.7 Let b satisfy the Lipschitz condition (10.1.11) and suppose that |b(X 1s , ys )|2p us ( dys ) < ∞, a.s. Then √ sup N E 1/p sup |Xti,N − X it |p < ∞ N

t≤T

for

p ≥ 2,

(10.1.19)

286

10. Stochastic Differential Equations and Empirical Measures

and N p−1 E sup |Xti,N − X it |p = o(1)

for

t≤T

1 ≤ p ≤ 2.

Proof: For notational convenience we drop the superscript N ; then t

Xti − X it =

0

⎛⎧ ⎫1/p ⎞ ⎫1/p ⎧  N ⎬ ⎨ ⎨ ⎬ ⎟ ⎜ 1  p i j p i − b(X s , y) us ( dy) b(Xs , Xs ) ⎠ ds ⎝ ⎩N ⎭ ⎭ ⎩ j=1

⎧⎡⎛ ⎤ ⎞ ⎞ 1/p ⎛ 1/p ⎪ ⎨ 1  ⎢⎝ 1  ⎥ i j p⎠ ⎝ − b(Xs , Xs ) b(X is , Xsj ⎠ ⎦ ds ⎣ = ⎪ N j N j ⎩ 0 ⎡⎛ ⎤ ⎞ ⎞ 1/p ⎛ 1/p 1  p i j ⎠ ⎥ ⎢ 1  + ⎣⎝ b(X is , Xsj )p ⎠ − ⎝ b (X s , X s ) ⎦ N j N j t

⎡⎛ ⎤⎫ ⎞ ⎞ 1/p 1/p ⎛ ⎪  ⎬ ⎥ ⎢⎝ 1  p i j p⎠ i ⎠ ⎝ +⎣ − b(X s , y) us ( dy) b(X s , X s ) ⎦ . ⎪ N j ⎭

Set |X|T := sups≤T |Xs |. Then by the Minkowski inequality and the Lipschitz condition on b, the above equality implies i

   1/p i p i  := E X − X T

i

X − X T,p ⎧ ⎪ T ⎪ N ⎨ 1  i i ds cXs − X s p + c ≤ Xsj − X js p ⎪ N j=1 ⎪ ⎩ 0

⎫  ⎞ ⎛ ⎛ 1/p ⎞1/p ⎛ ⎞ 1/p p  ⎪ ⎪    1  ⎬  ⎜ ⎝ ⎟ p i j p⎠ i ⎝ ⎠ . b(X s , y) us ( dy) + ⎝E  − b(X s , X s )  ⎠  ⎪  N ⎪ j   ⎭

Summing up over i and using the symmetry, we find

1

1

N X − X T,p =

N  i=1

X i − X i T,p

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

≤ 2c

T

ds

⎧ ⎪ N ⎨

287

Xsi − X is p

⎪ ⎩ i=1  ⎫ ⎛ ⎞ ⎞ 1/p p 1/p ⎛   N N  1   ⎪ ⎬  ⎝  p i j p⎠ i . b(X s , X s ) + E − ⎝ b(X s , y) us ( dy)⎠   N  ⎪ j=1 i=1   ⎭ 0

This amounts to

X i − X i T,p ≤ 2c

T 0

⎧ ⎪ ⎨ ds X i − X i s,p ⎪ ⎩

 ⎤⎫ ⎡ ⎛ ⎞ ⎞ 1/p p 1/p ⎛   N  ⎪ ⎬ 1  ⎢ ⎝ 1   ⎥ p i j p⎠ i ⎠ ⎝ b(X s , X s ) + − b(X s , y) us ( dy)  ⎦ , ⎣E   N  ⎪ N i=1 j   ⎭ and consequently, by the Gronwall lemma, ⎡ ⎛ ⎞ 1/p  N N    1 ⎢ ⎝ 1 b(X is , X js )p⎠ X i − X i T,p ≤ 2c e2cT ds ⎣E   N N i=1 j=1  0 ⎛ ⎞1/p p ⎤    ⎥ i p ⎠ ⎝ − b(X s , y) us ( dy)  ⎦   ⎛  ⎞ ⎞ 1/p p 1/p ⎛  T   1   ⎝  2cT 1 j p⎠ 1 p ⎝ ⎠ ds E  − b(X s , X s ) = 2c e b(X s , y) us ( dy)  .  N  j   0 T

By the Taylor expansion and with Yj := b(X is , X js )p (conditionally on X is ) we obtain p  p  1/p   S 1 1−p  SN   N 1/p  , +a −a  ≤ p a E E   N p N 

(10.1.20)

288

10. Stochastic Differential Equations and Empirical Measures

 where SN = (Yj − a), a = EYj > 0. Therefore, from the Marcinkiewicz– Zygmund inequality (cf. Chow and Teicher, (1978, p. 357)), we conclude that √

p   p 1/p   S   N 1/p  SN  1/p  1/p  +a − a  ≤ const. E  √  = O(1). NE    N N

This yields (10.1.19) for p ≥ 2. For 1 ≤ p < 2, the claim follows from N |p = o(1). the moment bounds of Pyke and Root (1968), giving E| NS1/p Therefore, p  1/p   S N p−1  1/p  +a N E − a  = o(1).   N

(10.1.21)

2

We next interpret Theorem 10.1.7 as a chaotic property of the diffusions governed by (10.1.4). Recall that by Proposition 2.2 in Sznitman (1989), a sequence (uN ) of symmetric probability measures on E (N ) is u-chaotic, u ∈ M 1 (E), if for (X1 , . . . , XN ) distributed as uN , N 1  w δXi −→ u. N i=1

For X N :=

1 N

w

N

i=1 δX i,N

X N −→ X,

(10.1.22)

we obtain from Theorem 10.1.7 that (10.1.23)

where X is the solution of equation (10.1.5). Therefore, with m denoting the law of X and mN denoting the law of (X 1,N , . . . , X N,N ) we obtain from (10.1.22) the following corollary.

Corollary 10.1.8 Under the assumptions of Theorems 10.1.3 and 10.1.7, the sequence (mN ) is m-chaotic.

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

289

Remark 10.1.9 For p = ∞ (see (10.1.17)) the propagation of chaos property does not hold. Also, the case 0 < p < 1 does not lead to propagation of chaos, and there does not exist a unique strong solution of

Xt = Bt +

t 

b(Xs , ys )p dm(y) ds.

(10.1.24)

0

Remark 10.1.10 (An example leading to a Burger type equation.) Consider the stochastic system

dXti

⎞1/p N  1 b(Xti , Xtj )p ⎠ dt, = dWti + ⎝ N j=1 ⎛

i = 1, . . . , N,

(10.1.25)

with Lipschitz (in both arguments) interactive term b(·, ·). Then the instantaneous drift term seen by particle i is

Δi

⎞1/p N  1 = ⎝ b(Xti , Xtj )p ⎠ . N j=1 ⎛

Under the assumptions of Theorem 10.1.7, we have ⎡

lim E ⎣

N →∞

1 N

N  i=1

2

Δpi −



⎤ 1/p 32 ⎦ = 0, bp (Xti , y)ut ( dy)

as well as  +2  N   1 lim E Δpi − bp (Xti , y)ut ( dy) = 0. N →∞ N i=1 *

Similarly to the above limit we shall examinethe average behavior nrelations p 1 of the “pseudo drift” N i=1 Zi . Here, Zip := N 1−1 j =i φpN,a (Xti − Xtj ), and φN,a (x − y) = N ad/p φ(N a (x − y)), where φ(·) ≥ 0 is smooth, compactly  supported on IRd , and φ(x) dx = 1. We consider the vector-valued case here. Note that N

Z1p

:=

1  p φN,a (Xt1 − Xtj ) N − 1 j=2

290

10. Stochastic Differential Equations and Empirical Measures N

=

1  ad p a i N φ (N (Xt − Xtj )), N − 1 j=1

and consequently, EZ1p = N ad (Eφp (N a (Xt1 − Xt2 )))  ad p a 1  N Eφ (N (Xt − Xt2 ))  p = φp φ

−→

N →∞

ut 2L2



φp =: ut,p (Xt ).

Consider next

an

*

+2 N 1  p := E (Zi − ut,p (Xt1 )) N i=1 ⎛ ⎞⎤2 ⎡ N 1 ⎝ 1  p φN,a (Xti − Xtj ) − ut,p (Xt1 )⎠⎦ = E⎣ N i=1 N − 1 j =i

=



E⎣

1 N −1

N  j=1

⎤2

φpN,a (Xt1 − Xtj ) − ut,p (Xt1 )⎦ .

Arguing as in Sznitman (1989, p. 196), we find that

aN →

⎧ ⎪ ⎪ 0 ⎪ ⎪ ⎪  ⎨  ⎪ ⎪ ⎪ ⎪ ⎪ ⎩



φ

2p

   p 2 φ dx ut 2L2  2p φ

if 0 < a < d1 , if a = d1 , if a > d1 .

Therefore, only in the case of moderate interaction do we obtain Burger’s equation in the limit.

10.1.3 A Random Number of Particles Let (W i )i∈IN be a sysetm of i.i.d. real-valued processes (as in (10.1.4)) with finite pth moments and let (Nn )n≥1 be an i.i.d. integer-valued sequence of

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

291

r.v.s independent of (W i ). Consider the following system of SDE with a random number of particles and interactions:

dXti,n

⎞1/p Nn  1 b(Xti,n , Xtj,n )p ⎠ dt, i = 1, . . . , Nn . (10.1.26) = dWti + ⎝ n j=1 ⎛

We assume that the following asymptotic stability condition holds: Nn → Y n

a.s. as n → ∞.

(10.1.27)

As in Section 10.1.2, it turns out that X i,N has a natural limit X i that is a solution of the nonlinear SDE dXt = dBt + Y 1/p



1/p b(Xt , y)p ut ( dy) .

(10.1.28)

d

Here B = W 1 , and Y is assumed to be independent of B. For m0 ∈ M 1 (CT ) let Mp (CT , m0 ), L∗p,T , ℓ∗p,T be defined as in Section 10.1.2.

Lemma 10.1.11 Suppose that

t 0

|Bs | ds < ∞ a.s. Then for any m ∈

Mp (CT , m0 ), there exists a unique strong solution of the equation

Xt = Bt + Y 1/p

t 0

Proof: Set (SX)t := Y 1/p

⎛ ⎝

CT

t 0

⎞1/p



⎛ ⎝

b(Xs , ys )p dm(y)⎠ 

CT

ds.

⎞1/p

b(Xs , ys )p dm(y)⎠

(10.1.29)

ds. Then arguing

in a similar fashion as in the proof of Lemma 10.1.1, we obtain the bound

sup |(SX)s − (SY )s | ≤ cY 1/p s≤t

t 0

sup |Xu − Yu | du.

0≤u≤s

292

10. Stochastic Differential Equations and Empirical Measures

Defining inductively X 0 := B, X n := SX n−1 , we have sup |Xsn s≤t



Xsn−1 |

n

≤ c Y

n/p t

n

n!

sup |Xs1 − Xs0 | s≤t

 ⎤ ⎡ t 1/p     t    t  ds⎦ ≤ cn Y (n+1)/p ⎣ Bs ds + |ys |p dm(y)  n!   n

0

0

< ∞.

This indeed implies the existence of a unique strong solution.

2

Given m ∈ Mp (CT , m0 ), let Φ(m) denote the distribution of the solution of (10.1.29). Then we have the following contraction-type property for the mapping Φ.

Lemma 10.1.12 Suppose that Ap := cY 1/p ecY T, m1 , m2 ∈ Mp (CT , m0 ),

1/p

p < ∞. Then for t ≤

t ℓ∗p,t (Φ(m1 ), Φ(m2 )) ≤ Ap ℓ∗p,s (m1 , m2 ) ds.

(10.1.30)

0

Proof: Let X (i) be the solution of the SDE

(i)

Xt

= Bt + Y 1/p

t 0

⎛ ⎝



CT

⎞1/p

b(Xs(i) , ys )p dmi (y)⎠

ds.

Then, as in the proof of Lemma 10.1.2,

sup |Xs(1) − Xs(2) | ≤ cY 1/p s≤t

By the Gronwall lemma, implying that

t 0

t sup |Xu(1) − Xu(2) | + c ds ℓp,s (m1 , m2 ). u≤s

sups≤t |Xs1 − Xs2 |

0

≤ cY

1/p cY 1/p

e

t 0

ℓp,s (m1 , m2 ) ds,

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

293

( t ( 1/p ( ( ℓ∗p,s (m1 , m2 ) ds. ℓ∗p,t (Φ(m1 ), Φ(m2 )) ≤ (cY 1/p ecY ( p

0

2

From Lemmas 10.1.11 and 10.1.12 we conclude that (10.1.28) has a unique solution. The proof is similar to that of Theorem 10.1.3. Theorem 10.1.13 Under the assumptions of Lemmas 10.1.11 and 10.1.12, equation (10.1.28) has a unique solution, provided that B∗p,T

10.1.4

< ∞

and

 

1/p ds < ∞. b(0, ys ) dm0 (y) p

pth Mean Interactions in Time: A Non-Markovian Case

Suppose (Xti,N )i=1,...,N determines a system of N particles and let b(Xsi,N , ·) := (b(Xsi,N , Xsj,N ))1≤i≤N describe the interaction vector. Recall that in Section 10.1.2 we considered equation (10.1.4) with a drift of the form b(Xsi,N , ·)p corresponding to the pth norm of the interaction vector. In this section we shall study SDEs with mean interactions in time. In fact, let     N  1  i,N j,N  (10.1.31) b Xs , Xs  Fi (s) :=  N   j=1 be the average of the interaction vector and consider the equations

Xti,N X0i,N



= Wti + ⎝ =

X0i ,

t 0

⎞1/p

|Fi (s)|p ds⎠

1 ≤ i ≤ N,

;

(10.1.32)

for 1 ≤ p < ∞;

Xti,N

= Wti + ess sup |Fi (s)|;

X0i,N

= X0i , 1 ≤ i ≤ N,

s≤t,λ\

for p = ∞;

(10.1.33)

294

10. Stochastic Differential Equations and Empirical Measures

Xti,N

= Wti +

X0i,N

X0i

t 0

=

|Fi (s)|p ds;

; 1 ≤ i ≤ N,

(10.1.34)

for 0 < p < 1.

In other words, we consider SDEs with a drift resulting from the pth mean in time of the average of the interaction vector. From the definition it is clear that this model describes a system that no longer behaves as a Markovian one, since the instantaneous drift |Fi (t)|p is weighted by the mean interac⎛ t ⎞1/p−1  1 over the whole past of the process. From this tion ⎝ |Fi (s)|p ds⎠ p 0

point of view the propagation of chaos property seems to be not so obvious in this model.

First we consider the case 1 ≤ p < ∞. The nonlinear limiting equation is given by

Xt

⎞1/p p  t    = Bt + ⎝  b(Xs , y)us ( dy) ds⎠ , us = P Xs . (10.1.35) ⎛

0

Here Xt , Bt , b are real-valued, Bt is a process in CT = C[0, T ], and |b(x1 , y) − b(x2 , y)| ≤ c|x1 − x2 | for some c > 0.

(10.1.36)

Define, for m0 ∈ M 1 (ℓT ), Mp (CT , m0 ) := {m1 ∈ M 1 (CT ); ℓ∗p,t (m1 , m0 ) < ∞}.

(10.1.37)

Then, for m ∈ Mp (CT , m0 ), consider the linear equation

Xt

p ⎞1/p   t     = Bt + ⎝  b(Xs , ys ) dm(y) ds⎠ ,   ⎛

0

CT

where ys is the value of y at time s.

(10.1.38)

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

295

Lemma 10.1.14 Assume that the Lipschitz condition (10.1.36) holds, and furthermore,  p  T      b(0, ys )ms ( dy) ds < ∞,     0

CT

where ms is the distribution of Xs at time s under m. Then (a) Equation (10.1.38) has a unique strong solution X. (b) If Φ(m) is the law of X, then Φ(m) ∈ Mp (CT , m0 ); that is, Φ : Mp (CT , m0 ) → Mp (CT , m0 ). Proof: Let X ∈ Xp (CT , m0 ) and define ⎛

⎞1/p p  t    := Bt + ⎝  b(Xs , y)ms ( dy) ds⎠ .

(SX)t

(10.1.39)

0

Then |(SX)t − (SY )t |p ⎛⎛ ⎞1/p p  t    ⎜ = ⎝⎝  b(Xs , y)ms ( dy) ds⎠ 0



−⎝

t 0

⎞1/p ⎞p  p    b(Ys , y)ms ( dy) ds⎠ ⎟ ⎠  

⎛ t ⎞1/p p     ≤ ⎝  (b(Xs , y) − b(Ys , y))ms ( dy) ds⎠ 0

(Minkowski inequality)

t cp |Xs − Ys |p ds (by the Lipschitz condition (10.1.36)). ≤ 0

296

10. Stochastic Differential Equations and Empirical Measures

This implies t ≤ cp sup |Xu − Yu |p ds,

sup |(SX)s − (SY )s |p s≤t

0

(10.1.40)

u≤s

t ∗p p and furthermore, L∗p (SX, SY ) ≤ c Lp,s (X, Y ) ds. Define, inductively, p,t 0

n

1 0 n n−1 X 0 := B, X n := SX n−1 . Then L∗p ) ≤ cpn Tn! L∗p p,t (X , X p,T (X , X ).  By (10.1.36), the integral b(Xs , ys )m( dy) is a Lipschitz function of Xs . CT

Thus

p  t    = E sup  b(Bs , y)ms ( dy) ds t≤T

1 0 L∗p p,T (X , X )

0



≤ E

T

≤ c

T 

0

0

(|b(0, y)| + c|Bs |)ms ( dy) |b(0, y)|ms ( dy)

p

p

ds

T ds + c E |Bs |p ds < ∞, 0

as by the assumptions the integrals in the right-hand side are finite. Therefore, 

L∗p,T (X n , X n−1 )

n≥1

This implies 

n≥1



n≥1





n≥1

c

n



Tn n!

1/p

L∗p,T (X 1 , X 0 ) < ∞.

L∗p,T (X n , X n−1 ) < ∞. Then

L∗1,T (X n , X n−1 ) < ∞.

In consequence, X n converges to some process X a.s. uniformly on bounded intervals. X is a.s. continuous, and E sups≤t |Xs |p < ∞, since E sups≤t |Bs |p < ∞. This yields Φ(m) ∈ Mp (CT , m0 ). 2 In addition, suppose that b is Lipschitz in both arguments; that is, |b(x1 , y1 ) − b(x2 , y2 )| ≤ c[|x1 − x2 | + |y1 − y2 |]

(10.1.41)

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

297

for all x1 , x2 , y1 , and y2 in CT , and consider the map Φ : Mp (XT , m0 ) → Mp (CT , m0 ). Lemma 10.1.15 (Contraction of Φ with respect to ℓ∗p,t ) Suppose that (10.1.41) and the assumption of Lemma 10.1.14 hold. Then for t < T and m1 , m2 ∈ Mp (CT , m0 ), t ℓ∗p,t (Φ(m1 ), Φ(m2 )) ≤ cp ecp t ℓ∗p p,s (m1 , m2 ) ds,

(10.1.42)

0

where cp := c 2p−1 . Proof: For i = 1, 2 and t ≤ T , set (i)

Xt

p ⎞1/p ⎛ t        (i) = Bt ⎝  b Xs , ys dmi (y) ds⎠ ,   0

CT

and let m ∈ M 1 (m1 , m2 ), the class of probabilities on CT × CT with marginals m1 and m2 . Then ⎛  p ⎞ 1/p  t          sup |Xs(1) − Xs(2) |p = ⎝  b Xs(1) , ys(1) dm1 (y (1) ) ds⎠  s≤t   0  CT  p ⎞ ⎛ t 1/p p           (2) (2) (2)   dm2 (y ) ds⎠  − ⎝  b X s , ys     0 CT ⎤ ⎡ p t           ≤ ds ⎣ b Xs(1) , ys(1) − b Xs(2) , ys(2)  dm y (1) , y (2) ⎦ 0



t 0

CT ×CT

#      $p     (1) . ds c Xs − Xs(2)  + ys(1) − ys(2)  dm y (1) , y (2)

Minimizing the right-hand side over all couplings, we get  p  (1) (2)  sup Xs − Xs  s≤t

298

10. Stochastic Differential Equations and Empirical Measures

≤ Hc · IJ 2p−1K =:cp

t 0

t p    2p−1K ds ℓ∗p ds sup Xu(1) − Xu(2)  + cH · IJ 1,s (m1 , m2 ). u≤s

=:cp

0

Consequently, for p ≥ 1, by the Gronwall lemma and ℓ∗1,s ≤ ℓ∗p,s , t p    sup Xs(1) − Xs(2)  ≤ cp ecp t ds ℓ∗p p,s (m1 , m2 ). s 0, √

sup N N



E

sup |Xti,N t≤T

− X it |p

1/p

< ∞

for p ≥ 2

(10.1.45)

and N

(1/p)−1



E

sup |Xti,N t≤T



X it |p

1/p

= o(1)

for 1 ≤ p < 2.

Proof: We drop further the superscript N . Then Xti − X it p ⎞ ⎛ t ⎞ 1/p ⎛ t  p 1/p      N 1       = ⎝  b Xsi , Xsj  ds⎠ − ⎝  b X is , y us ( dy) ds⎠  N j=1  0 0 ⎤ ⎡⎛   p ⎞  ⎞ ⎛ 1/p 1/p p t     t   N N 1   i j   i j  ⎥ ⎢⎝  1   ⎠ ⎝ X s , Xs  ds⎠ ⎦ ds − b X , X b = ⎣ s s  N N    j=1  j=1 0 0 ⎤ ⎡⎛  p ⎞ p ⎞ 1/p ⎛ t  1/p      t   N N 1 1     ⎥ ⎢ b X is , Xsj  ds⎠ − ⎝  b X is , X js  ds⎠ ⎦ + ⎣⎝     N j=1  N j=1 0

0

300

10. Stochastic Differential Equations and Empirical Measures

⎡⎛  p ⎞ ⎞ ⎤ 1/p ⎛ t  p 1/p   t   N  1      ⎥ ⎢ b X is , X js  ds⎠ − ⎝  b X is , y us ( dy) ds⎠ ⎦ . + ⎣⎝    N j=1 0

0

Applying the Minkowski inequality and setting XT := sups≤T |Xs |, we obtain

X i − X i pT,p = EX i − X i pT  p ⎡  t   N        1 ≤ 4p−1 ⎣E ds  b Xsi , Xsj − b X is , Xsj   N j=1  0  p  T   N 1  i j    i j + E ds  b X s , Xs − b X s , X s   N j=1  0 p ⎤    T   N  1  i   i j + E ds  b X s , X s − b X s , y us ( dy) ⎦   N j=1 0 ⎧ ⎤p ⎡ T ⎨ N 1  j ≤ 4p−1 ds cp E|Xsi − X is |p + cp E ⎣ |X − X js |⎦ ⎩ N j=1 s 0 p ⎫   ⎬  T   N  1 . b(X is , X js ) − b(X is , y)us ( dy) + E ds   ⎭ N j=1

0

Summing up over i and using the symmetry, we find that

i

N X −

X i pT,p

=

N  i=1

X i − X i pT,p

⎧ ⎤ ⎡ ⎛ ⎞ 1/p p T ⎪ N N ⎨  ⎥ ⎢ 1 ⎝ j p i p−1 i p p |Xs − X js |p ⎠ ⎦ ds c EXs − X s p + c N E ⎣ ≤ 4 ⎪ N j=1 ⎩ i=1 0 p ⎫   ⎬   N N  1   b(X is , X js ) − b(X is , y)us ( dy) . E  + cp ⎭ N   j=1 i=1

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

301

Therefore, ⎧ ⎨ i p−1 p ≤ 4 c ds X i − X i ps,p + cp X i − X ps,p ⎩ 0  p ⎫   ⎬  N N  1   1  i j i + E b(X s , X s ) − b(X s , y)us ( dy) . N i=1  N j=1  ⎭ T

X i − X i pT,p

Consequently, by the Gronwall lemma, X i − X i pT,p ≤ Cp eCp T

T 0

 p ⎤    N 1   1 i j i E  b(X s , X s ) − b(X s , y)us ( dy) ⎦ ds ⎣ N i=1  N j=1  ⎡

N 

(with Cp = 2 · 4p−1 cp )  p   T  N    1 ≤ Cp eCp T ds E  b(X is , X js ) − b(X is , y)us ( dy)  N j=1  0 #  $p 1 Cp T = Cp e T ·E 0 √ ; N

here we also used the Marcinkiewicz–Zygmund inequality (cf. Chow and Teicher (1978, p. 357)) for p ≥ 2, and the Pyke and Root (1968) inequality for 1 ≤ p < 2. 2

Corollary 10.1.18 Let m denote the law of X satisfying (10.1.43), and let mN be the law of (X 1,N , . . . , X N,N ). Then, under the assumptions of Theorems 10.1.16 and 10.1.17, mN is m-chaotic. We next study the limiting case p = ∞ (cf.(10.1.33)). In contrast to the limiting case in Section 10.1.4 of pth norm interaction, we obtain the propagation of chaos property for pth mean interaction in time under a stronger Lipschitz condition. Consider for m ∈ M 1 (CT ), Xt

      = Bt + ess sup  b(Xs , y)m( dy) . s≤t   CT

(10.1.46)

302

10. Stochastic Differential Equations and Empirical Measures

Here Xt , Bt , and b are real-valued, Bt is a process on CT having finite pth moment, E ess sups≤T |Bs |p < ∞, and b satisfies the Lipschitz condition for all x1 , x2 , and y ∈ CT : |b(x1 , y) − b(x2 , y)| ≤ c|x1 − x2 |,

with 0 < c < 1.

(10.1.47)

We shall use the following Lp -type metric for p ≥ 1:  ∗p,t (X, Y ) := (E ess sup |Xs − Ys |p )1/p L s≤t

Let

in X (CT ).

5  ∗p,t (m1 , m2 ) ℓ∗p,t (m1 , m2 ) = L

(10.1.48)

(10.1.49)

be the corresponding minimal metric. Consider the set of measures on M 2 (CT ): Bp (CT , m0 ) = {m1 ∈ M 1 (CT ); ℓ∗p,T (m1 , m0 ) < ∞}, M

(10.1.50)

and let Xp (CT , m0 ) denote the corresponding class of processes. For m0 ∈ Bp (CT , m0 ), consider the linear equation M 1 (CT ) and m ∈ M Xt

      = Bt + ess sup  b(Xs , ys ) dm(y) . s≤t  

(10.1.51)

CT

Lemma 10.1.19 Assume that the Lipschitz condition (10.1.47) holds, and let       ess sup  b(0, ys )m( dy) < ∞. s≤T   CT

Then

(a) Equation (10.1.51) has a unique solution X. (b) If Φ(m) is the law of X, then Φ(m) ∈ Mp (CT , m0 ); that is, Bp (CT , m0 ) → M Bp (CT , m0 ). Φ:M

10.1 Propagation of Chaos and Contraction of Stochastic Mappings

Proof: Let X ∈ Xp (CT , m0 ) and define

      := Bt + ess sup  b(Xs , ys )m( dy) . s≤t  

(SX)t

CT

Then

|(SX)t − (SY )t |p    p             = ess sup  b(Xs , ys )m( dy) − ess sup  b(Ys , ys )m( dy)  0≤s≤t   0≤s 0.

∞ etx dμ(x) < ∞,

−∞

(10.3.8)

Then there exist positive constants C, A, λ, depending only on μ, and for each natural number s > 0 two s-tuples (x1 , . . . , xs ) and (y1 , . . . , ys ), each consisting of i.i.d. real-valued r.v.s, with L(x1 ) = μ and y1 being standardnormally distributed, such that for each a > 0, P



    k  max  i=1 (xi − yi ) > C ln s + a < Ae−λa .

1≤k≤s

For translating this estimate into an estimate with the distance used in the previous chapters, we need the following lemma. Lemma 10.3.12 Assume that there exist constants C, A, λ > 0 with λC ≥ 1 and for any two natural numbers r, s ≥ 1 an r-tuple (δ1,s , . . . , δr,s ) of i.i.d. positive real-valued r.v.s satisfying P (δ1,s > C ln s + a) < Ae−λa

for all a > 0.

(10.3.9)

Then for each p ∈ [0, ∞), there exists a constant Mp > 0 such that for all natural r, s ≥ 1, p ≤ Mp (1 + ln r + ln s)p . E max δi,s 1≤i≤r

10.3 Stochastic Differential Equations

347

The following well-known result (cf., for example, Shortt (1983), Rachev (1991)) is used in the proof of the following theorem. Lemma 10.3.13 Let S1 , S2 , and S3 be Polish spaces (i.e., topological spaces that are metrizable with a complete separable metric), and let π12 : S1 × S2 × S3 → S1 × S2 ; π23 : S1 × S2 × S3 → S2 × S3 ; π212 : S1 × S2 → S2 , and π223 : S2 × S3 → S2 denote the projections defined by dropping one component. Then for any two measures ν12 ∈ P(S1 × S2 ) and ν23 ∈ P(S2 × S3 ) with ν12 ◦ (π212 )−1 = ν23 ◦ (π223 )−1 , i.e., with identical marginal distributions on S2 , there exists a measure ν123 ∈ P(S1 ×S2 ×S3 ) with ν123 ◦ (π12 )−1 = ν12 and ν123 ◦ (π23 )−1 = ν23 . Now we can prove the estimates for the chance discretization step: Theorem 10.3.14 Let p ∈ [2, ∞) and μ ∈ P(IR) satisfy (10.3.8). Then we can define a q-dimensional standard Wiener process (w(t))t∈[t0 ,T ] and k a set of i.i.d. r.v.s {ξji : j = 1, . . . , q; i = 1, . . . , mk ; k = 0, . . . , n − 1} 0 with distribution L(ξ11 ) = μ on a common probability space such that for the methods (E2), (E3), (M2), and (M3) constructed with them we have (a) If (AS1) and (AS2) hold, then E

E

p

y (t) − z (t) ≤ K E sup  t0 ≤t≤T



1+ln m(h)



m(h)

p

.

(b) If (AS1), (AS2), (AS3), and (AS4) hold, then y M (t) − z M (t)p ≤ K E sup  t0 ≤t≤T



1+ln m(h)



m(h)

p

.

The preceding results yield the following theorem, which gives bounds for the Lp -norm of the differences between the exact solution x of (SE) and the approximate solutions z E and z M defined in (E3) and (M3). Again, as in Theorem 10.3.14, this is a result in the weak sense. Theorem 10.3.15 Let p ∈ [2, ∞) and μ ∈ P(IR) satisfy (10.3.8). Then we can define a q-dimensional standard Wiener process (w(t))t∈[t0 ,T ] and k a set of i.i.d. r.v.s {ξji , j = 1, . . . , q, i = 1, . . . , mk , k = 0, . . . , n − 1} with 0 distribution L(ξ11 ) = μ on a common probability space such that for (SE) and the methods (E3) and (M3) constructed with them we have

348

10. Stochastic Differential Equations and Empirical Measures

(a) If (AS1) and (AS2) hold, then 

E sup x(t) − z E (t)p ≤ K hp/2 + t0 ≤t≤T



1+ln m(h)



m(h)

p 

.

(b) If (AS1), (AS2), (AS3), and (AS4) hold, then M

E sup x(t) − z (t)

p

t0 ≤t≤T

  p  1+ln m(h) p √ ≤ K h + . m(h)

To show that both assertions (a) and (b) follow from the Theorems 10.3.6, 10.3.10, and 10.3.14, it suffices to verify that h m(h)



1 + ln



m(h) h



≤ K



1+ln m(h)



m(h)

2

,

which follows easily from (G1). Since Theorem 10.3.15 provides results in the weak sense, it is appropriate to formulate it as an estimate for the Lp -Wasserstein metric between the distributions of the exact solution and the approximate solutions: Corollary 10.3.16 Let p ∈ [1, ∞) and μ ∈ P(IR) have the properties (10.3.8). Moreover, let w(t) t∈[t ,T ] be a q-dimensional standard Wiener 0

k {ξji

process and : j = 1, . . . , q; i = 1, . . . , mk ; k = 0, . . . , n − 1} a set of 0 ) = μ. Then for (SE) and the methods i.i.d. r.v.s with distribution L(ξ11 (E3) and (M3) constructed with them we have (a) If (AS1) and (AS2) hold, then E



1/2

ℓp (x, z ) ≤ K h

1+ln m(h)

+ √

m(h)



.

(b) If (AS1), (AS2), (AS3), and (AS4) hold, then   √ m(h) . ℓp (x, z M ) ≤ K h + 1+ln m(h)

For p ∈ [2, ∞) the assertions follow directly from Theorem 10.3.15 and applying Lemma 10.3.3 to the right-hand sides. Then the assertions are also true for p ∈ [1, 2), since ℓp1 ≤ ℓp2 for 1 ≤ p1 ≤ p2 < ∞. The estimates in Theorem 10.3.15 and Corollary 10.3.16 give convergence rates

10.3 Stochastic Differential Equations

349

with respect to h for the methods (E3) and (M3) and for any grid sequence in G(m, Λ, α, β). These rates consist of two summands, one depending on h and the other depending on m(h), representing the rates of time and chance discretization, respectively. It is desirable to tune the rates of both summands, i.e., to equal the powers of h in both summands. This means to choose m(h) to be increasing like 1/h for method (E3) and like 1/h2 for the method (M3). Corollary 10.3.17 Let p ∈ [2, ∞) and μ ∈ P(IR) satisfy (10.3.8). Then we can construct solutions in (SE), (E3), and (M3) on a common probability space (as in Theorem 10.3.15) with the following properties. & (a) If (AS1) and (AS2) hold and max sup0