Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Lectures on the Coupling Method
Lectures on the Coupling Method
Lectures on the Coupling Method
Ebook418 pages4 hours

Lectures on the Coupling Method

Rating: 0 out of 5 stars

()

Read preview

About this ebook

An important tool in probability theory and its applications, the coupling method is primarily used in estimates of total variation distances. The method also works well in establishing inequalities, and it has proven highly successful in the study of Markov and renewal process asymptotics. This text represents a detailed, comprehensive examination of the method and its broad variety of applications. Readers progress from simple to advanced topics, with end-of-discussion notes that reinforce the preceding material. Topics include renewal theory, Markov chains, Poisson approximation, ergodicity, and Strassen's theorem. A practical and easy-to-use reference, this volume will accommodate the diverse needs of professionals in the fields of statistics, mathematics, and operational research, as well as those of teachers and students.
LanguageEnglish
Release dateAug 15, 2012
ISBN9780486153247
Lectures on the Coupling Method

Related to Lectures on the Coupling Method

Titles in the series (100)

View More

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Lectures on the Coupling Method

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Lectures on the Coupling Method - Torgny Lindvall

    Index

    Introduction

    1. Three examplesbe a Markov chain with a countable state space (equal to ℤ+ = the nonnegative integers), governed by a transition matrix P = (pij), making it aperiodic, irreducible, and positive recurrent. It is a classical result that X approaches stationarity as n → ∞, regardless of the initial distribution λ (that of X0):

    (1.1)

    is the unique stationary distribution, satisfying π = πPis the n, independent of X, governed by P have distribution πby

    where

    We thus make a coupling of X and X’ at T and call that the coupling time. At this instant you should not hesitate to believe that X and X" are equally distributed (this is due to the strong Markov property). This implies that for each j ∈ ℤ+,

    (1.2)

    . Hence

    P(Xn = j)–πj∣ ≤ 2 · P(T > n),

    so (1.1) follows if the coupling is successful (i.e., if T < ∞ a.s.). Actually, this inequality holds without that 2 since

    But (1.2) has more to give; a summation on both sides of that inequality renders

    (now that 2 must be included!) which we may write in a more compact form as

    (1.3)

    or

    ∥λPnπ∥ ≤ 2 ·P(T > n).

    corresponding row vectors (ν0, ν1, . . .), and bounded signed measures on ℤ+, and define

    So if we can prove that the coupling is successful, uniform convergence toward stationarity is established effortlessly.

    A number of questions should now be in your mind. For example:

    Is it easy to prove that the coupling is successful?

    Does (1.3) yield new results?

    Is there an interesting class of Markov chains with general state space to which the coupling method is applicable?

    The common answer is: yes.

    with state space ℤ+. Assume here that the intensities render X nonexplosive and recurrent. With pij(t) short for Pi(Xt = j), where the i indicates that X0 = i and Pt = (pij(t)) for t ≥ 0, the distribution of Xt becomes λPt if that of X0 is λ.

    To investigate the asymptotics of X, introduce the parallel process X’, governed by the same intensities as and independent of X, and with initial distribution μ, say. Now letting

    and

    (1.4)

    we may proceed as in (1.2) to obtain

    (1.5)

    But since the path of a birth and death process is skip-free and hence two independent paths cannot cross without meeting, it is easy to understand that

    where τare the times when X, X’ hit the state 0. Due to the recurrence assumption, we certainly have τa.s., so the coupling is successful and (1.5) yields

    (1.6)

    for any initial distributions λ, μ. Notice that we have not assumed positive recurrence, so a stationary distribution may not exist. If it does, call it π again and put μ = π in (1.6) to obtain

    (1.7)

    While (1.7) is a result on tendency toward stationarity (or equilibrium, as some prefer to say), the statement (1.6) means that a recurrent birth and death process forgets its initial distribution. The term ergodic is used in a number of ways in the literature; in this book we reserve it for results of the type (1.6).

    If X0 = 0, then regardless of μ,

    (1.8)

    again due to the skip-free paths. Since X and X" are equally distributed, we may deduce from (1.8) that

    if g: ℤ+ → ℝ is nondecreasing. In particular,

    (What alternative proof of that do you support?) Hence the coupling method serves us well also for establishing inequalities: You will see many examples of that use below.

    How far-reaching are the consequences of the observation that crossing paths produce a coupling? Very far. In fact, it has made the coupling method an indispensable tool in the study of the Markov processes of birth and death type with state space, for example, {0, 1}S (S countable), that are used to model interacting particle systems.

    For the final example, let Y1, . . . , Yn be independent 0–1 variables with

    P(Yi = 1) = pi

    .When all pi values are small, X is approximately

    , as is well known. To illuminate that, we search for a bound of

    where (i) = exp (–λ) · λi/i! for i . Now suppose that we can find a variable X’ with distribution pλ such that X = X’ with high probability. Then

    (1.9)

    hence

    (1.10)

    so if P(X X’) is small, we get a good Poisson approximation.

    But how is such a variable X’ produced? For one possible answer, let (Yi) be independent for 1 ≤ i n and such that

    (1.11)

    and

    Then Yi is a 0–1 variable with P(Yi = 1) = piis Poisson distributed with parameter pi. has the required distribution, and (1.10) renders

    But

    and we have proved that

    (1.12)

    which is good enough for many applications.

    The use of couplings for approximations such as (1.12) will be only a minor theme of this book. However, there are topics that should not be withheld; among them is a result on Poisson approximation for sums of weakly dependent 0–1 variables, sharp enough to generalize the remarkable Le Cam’s theorem.

    2. An outline. To know a method is to have learned how it works. What we have ahead of us is essentially a collection of applications of a few basic ideas consisting largely of topics of wide common interest with an attempt to maximize diversity. References to specialized studies are given in the Notes at the ends of chapter parts.

    In Chapter I a number of cornerstones are laid. The term coupling is defined, and the coupling inequality, of which (1.3), (1.5), and (1.12) are applications, is settled and elaborated.

    Renewal theory is a main theme of the book. Part 1 of Chapter II is dedicated to a detailed account of coupling of discrete-time renewal processes. The ideas and results here are used repeatedly, with consequences of type (1.3) for Markov chains discussed in the second parts of Chapters II and III, so this is basic. Skill is gained from learning how to couple discrete random walks, shuffle cards, and make Poisson approximations in the remaining parts of Chapter II.

    New challenges face us as soon as we go beyond the discrete models. In Chapter III we meet some of them; the first concerns coupling of continuous-time renewal processes. It turns out that virtually all asymptotic results for such processes may be obtained by suitable couplings, as shown in Part 1. Coupling and related methods have had a strong impact on the asymptotics theory of Markov chains with a general state space. Part 2 is an account of that.

    At this stage it is natural to answer some questions that have emerged. Is there a best possible coupling of two random sequences? Can we generalize our results from renewal to regenerative processes? Are there particular observations to make concerning coupling of Markov processes? Affirmative answers are provided in Parts 3 and 4. In Part 5 of Chapter III we sum up many of our findings about Markov chain ergodicity and coupling in one major theorem.

    Relation (1.8) and its consequences is a simple example of how to use coupling to establish inequalities. That use is the theme of Chapter IV, which has Strassen’s theorem as a basic result. However, many interesting inequalities may be proved by simpler means, in terms of independent, uniformly distributed variables, for example; the last section is a gallery of such inequalities.

    The coupling method is a fine tool in the study of Markov processes which are defined in terms of transition intensities. The first three parts of Chapter V show its use in a range of examples, from birth and death processes, through models for queueing networks and epidemics, to interacting particle systems. The possibilities of embedding in a bivariate Poisson process are demonstrated in Parts 4 and 5; among other things, we find that device useful for urn models and renewal theory.

    Chapter VI is devoted to diffusions. In several cases, coupling of one-dimensional diffusions is easy, due to the path continuity. The multidimensional case is in general difficult, but there are interesting exceptions; it turns out, for example, that Brownian motions may efficiently and easily be coupled.

    We finish with a bit of history of the method in the Appendix, where a tribute is paid to its inventor, Wolfgang Doeblin.

    3. Notes. The reader is supposed to be familiar with probability theory based on measures and Lebesgue integrals learned from, for example, Ash [9], Billingsley [27], Chung [39], Dudley [55], Durrett [58] or Williams [160]. And there are few readers with this background who do not have the slight acquaintance with Markov chains, random walk, Brownian motion, and so on, required, learned either from one of these references or from an undergraduate textbook such as Grimmett and Stirzaker [69]. A few standard results are used without comment; they are easily found in some of the references noted above.

    The enumeration is a standard one: By (3.2) we refer to the second enumerated item in § 3 in the same chapter; (IV 3.2) indicates a reference to another chapter; § IV.3 designates § 3 in Chapter IV The sign = is used not only to mean equal to but also occasionally be defined by and even which we abbreviate to.

    denote the class of measurable mappings from E to Eequals (ℝ, ℛ), the real numbers with the Borel sets, we simply write f in order to say that f is a measurable real-valued function.

    For a class of functions, the prefixes b, c, and i means that f is an increasing (= nondecreasing) and bounded measurable function. The prefix i requires that E be ordered in some way, while c can be used when E has a topology. The prefix b is also used for measures.

    CHAPTER I

    Preliminaries

    1. What is a coupling? .

    (1.1) A coupling of the probability measures P and P’ on a measurable space is a probability measure on such that

    where π(x, x′) = x, π′(x, x′) = x′ for (x, x′) ∈ E² .

    Thus P and P′ .

    Now typical uses of the coupling method are those demonstrated in § Int.1, and we find that our definition (1.1) is not well fit for immediate use. Once again we understand that probability theory is not animated measure theory, but rather, a mathematical discipline in its own right, concerned with random elements.

    to be a quadruple

    (Ω, ℱ, P, X),

    where (Ω, ℱ, P) is the underlying probability space (the sample space) and X , the class of measurable mappings from Ω to E. (1.2) By a coupling of the random elements (Ω, ℱ, P, X) and (Ω′, ℱ′, P′, X′) in we mean a random element in such that

    is a coupling of PX–1 and P′X′–1 in the sense of (1.1).

    as a coupling of the random elements X and X’. When no clarity is lost, we omit superscripts (^ , etc.) on Ω, ℱ, and P.

    For the random elements involved, the superscripts will vary; you will meet X" , , and so on. In the first example of § Int.1 we produced a coupling (X", X’) of the random elements X and X’ ,and the second example was similar in nature. The third, concerning Poisson approximation, is a case where for a comparison of two probability measures P and P’ , a random element (Ω, ℱ, P, (X, Xis constructed. We shall refer to such a construction as a coupling solution of or a coupling approach to a certain problem. Notice that the definitions (1.1) and (1.2) may be extended directly to be valid for any class of probability measures or random elements.

    When should the term coupling be used? There is no general agreement about that; some prefer a usage restricted to topics of the type presented in § Int.1; others find it suitable in any situation where for the comparison of probability measures, a common probability space is constructed. From the latter point of view, embeddings in Brownian motion, for examples, are actually couplings; notice that the definition (1.2) does not reject the use of the term in such connections.

    If you prefer a restrictive usage, you soon find yourself in trouble when trying to state a definition. But there is no urgent need for any.

    2. The coupling inequality. , that norm is defined by

    (2.1)

    such that v(∙ ∩ D) and–v( ∙ ∩ Dc) are positive measures, v+ and v–, say. We have v = v+ –v,

    (2.2)

    and find that

    (2.3)

    . If v has total mass 0, then v+(E) = v–(E), so if P and P’ then (2.2–2.3) renders

    (2.4)

    There are further details about bs and ∥∙∥ in App.2.

    Now suppose that the random element (Ω, ℱ, P, (Z, Z’)) is a coupling in order to estimate ∥PP′∥. Using (2.4), we obtain

    But

    (2.5)

    so

    (2.6)

    This is the (basic) coupling inequality. Before proceeding, notice that we must have {Z = Z′} E , so {Z = Z’} = {(Z, Z′) ∈ ∆} ∈ ℱ.

    , where Xn. If there is a random time T ∈ ℤ+ such that

    (2.7)

    then we call T a coupling time and obtain from (2.6) that

    (2.8)

    .

    , define the shift θn: E∞ → E∞ by

    (2.9)

    for x = (x0, x1, x2,...) ∈ E, where z is a fixed element in E. Notice that Ewith the same coupling times T , and we obtain that (2.8) actually improves itself, to the inequality

    (2.10)

    Consider the space of functions DE[0, ∞), with values in a Polish space E and defined on [0, ∞), which are right-continuous and have left-hand limits at all arguments t. Endowed with the Skorohod topology, that space, which we abbreviate as DE, becomes a Polish space. This is the most general type of path space we shall use; hence we shall not violate the Polish assumption even in our investigations of continuous-time processes. The σ-field on DE ; see also § App.1.

    where X and X’ are DE valued. We call the random time T ∈ ℝ+ a coupling time if

    (2.11)

    The analogs of (2.8) and (2.10) are

    (2.12)

    and

    (2.13)

    where the shift θt is defined in the obvious way:

    (2.14)

    for x DE, where z is a fixed element ∈ E. For (2.12)–(2.13), notice that we must know that θtX and θtX’ , in order to use the argument (2.5) again. But they are; in order that a mapping into DE, h say, shall be measurable, it suffices that h(t) is a measurable mapping into E for each t > 0, as is well known to a reader familiar with the Skorohod topology.

    Let Z and Z′ . From (2.4) we obtain

    (2.15)

    is a coupling of Z and Z′, a

    Enjoying the preview?
    Page 1 of 1