First Hitting Time Analysis of the Independence Metropolis Sampler

Similar documents
First Hitting Time Analysis of the Independence Metropolis Sampler

Lect4: Exact Sampling Techniques and MCMC Convergence Analysis

Minicourse on: Markov Chain Monte Carlo: Simulation Techniques in Statistics

Introduction to Machine Learning CMU-10701

The Particle Filter. PD Dr. Rudolph Triebel Computer Vision Group. Machine Learning for Computer Vision

Computer Vision Group Prof. Daniel Cremers. 11. Sampling Methods: Markov Chain Monte Carlo

6 Markov Chain Monte Carlo (MCMC)

Computer Vision Group Prof. Daniel Cremers. 11. Sampling Methods

MARKOV CHAIN MONTE CARLO

Computer Vision Group Prof. Daniel Cremers. 14. Sampling Methods

A Geometric Interpretation of the Metropolis Hastings Algorithm

A quick introduction to Markov chains and Markov chain Monte Carlo (revised version)

Convex Optimization CMU-10725

Stochastic optimization Markov Chain Monte Carlo

Markov Chain Monte Carlo The Metropolis-Hastings Algorithm

STA 294: Stochastic Processes & Bayesian Nonparametrics

Markov Chains CK eqns Classes Hitting times Rec./trans. Strong Markov Stat. distr. Reversibility * Markov Chains

The simple slice sampler is a specialised type of MCMC auxiliary variable method (Swendsen and Wang, 1987; Edwards and Sokal, 1988; Besag and Green, 1

Computer Vision Group Prof. Daniel Cremers. 10a. Markov Chain Monte Carlo

Monte Carlo Methods. Leon Gu CSD, CMU

STA 4273H: Statistical Machine Learning

Introduction to Machine Learning CMU-10701

Markov Chain Monte Carlo

P i [B k ] = lim. n=1 p(n) ii <. n=1. V i :=

1 Probabilities. 1.1 Basics 1 PROBABILITIES

Quantitative Non-Geometric Convergence Bounds for Independence Samplers

Ch5. Markov Chain Monte Carlo

University of Toronto Department of Statistics

Linear Algebra March 16, 2019

G METHOD IN ACTION: FROM EXACT SAMPLING TO APPROXIMATE ONE

6.842 Randomness and Computation March 3, Lecture 8

Lecture 8: The Metropolis-Hastings Algorithm

Computational statistics

MCMC notes by Mark Holder

Sampling Methods (11/30/04)

Winter 2019 Math 106 Topics in Applied Mathematics. Lecture 9: Markov Chain Monte Carlo

Markov Chain Monte Carlo (MCMC)

Lecture 20 : Markov Chains

Minimal basis for connected Markov chain over 3 3 K contingency tables with fixed two-dimensional marginals. Satoshi AOKI and Akimichi TAKEMURA

INTRODUCTION TO MARKOV CHAIN MONTE CARLO

On the Optimal Scaling of the Modified Metropolis-Hastings algorithm

1 Probabilities. 1.1 Basics 1 PROBABILITIES

Markov chain Monte Carlo Lecture 9

Markov Chains and Stochastic Sampling

Model reversibility of a two dimensional reflecting random walk and its application to queueing network

CS242: Probabilistic Graphical Models Lecture 7B: Markov Chain Monte Carlo & Gibbs Sampling

Doubly Indexed Infinite Series

SAMPLING ALGORITHMS. In general. Inference in Bayesian models

Robert Collins CSE586, PSU Intro to Sampling Methods

ELEC633: Graphical Models

SMSTC (2007/08) Probability.

Markov and Gibbs Random Fields

RENEWAL THEORY STEVEN P. LALLEY UNIVERSITY OF CHICAGO. X i

Lecture 11: Introduction to Markov Chains. Copyright G. Caire (Sample Lectures) 321

16 : Approximate Inference: Markov Chain Monte Carlo

AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES

CS 246 Review of Linear Algebra 01/17/19

Boolean Inner-Product Spaces and Boolean Matrices

Session 3A: Markov chain Monte Carlo (MCMC)

16 : Markov Chain Monte Carlo (MCMC)

MCMC Methods: Gibbs and Metropolis

Detailed Proofs of Lemmas, Theorems, and Corollaries

ON THE REGULARITY OF SAMPLE PATHS OF SUB-ELLIPTIC DIFFUSIONS ON MANIFOLDS

Markov Chains, Stochastic Processes, and Matrix Decompositions

Fundamentals of Engineering Analysis (650163)

Quantifying Uncertainty

ON SUM OF SQUARES DECOMPOSITION FOR A BIQUADRATIC MATRIX FUNCTION

17 : Markov Chain Monte Carlo

Markov chain Monte Carlo

On Reparametrization and the Gibbs Sampler

Discrete Applied Mathematics

Brief introduction to Markov Chain Monte Carlo

Markov Chain Monte Carlo

Stochastic Simulation

Lecture 7 and 8: Markov Chain Monte Carlo

A regeneration proof of the central limit theorem for uniformly ergodic Markov chains

Simulated Annealing for Constrained Global Optimization

Reinforcement Learning

1 Multiply Eq. E i by λ 0: (λe i ) (E i ) 2 Multiply Eq. E j by λ and add to Eq. E i : (E i + λe j ) (E i )

Chapter 2: Markov Chains and Queues in Discrete Time

April 20th, Advanced Topics in Machine Learning California Institute of Technology. Markov Chain Monte Carlo for Machine Learning

Exponential families also behave nicely under conditioning. Specifically, suppose we write η = (η 1, η 2 ) R k R p k so that

ON CONVERGENCE RATES OF GIBBS SAMPLERS FOR UNIFORM DISTRIBUTIONS

Review. DS GA 1002 Statistical and Mathematical Models. Carlos Fernandez-Granda

Distributed Randomized Algorithms for the PageRank Computation Hideaki Ishii, Member, IEEE, and Roberto Tempo, Fellow, IEEE

CPSC 540: Machine Learning

Asymptotics and Simulation of Heavy-Tailed Processes

ST 740: Markov Chain Monte Carlo

The Distribution of Mixing Times in Markov Chains

MCMC: Markov Chain Monte Carlo

Advances and Applications in Perfect Sampling

Sampling Contingency Tables

F denotes cumulative density. denotes probability density function; (.)

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Spanning and Independence Properties of Finite Frames

LECTURE 15 Markov chain Monte Carlo

Gärtner-Ellis Theorem and applications.

STOCHASTIC PROCESSES Basic notions

Math 471 (Numerical methods) Chapter 3 (second half). System of equations

A misère-play -operator

Transcription:

(To appear in the Journal for Theoretical Probability) First Hitting Time Analysis of the Independence Metropolis Sampler Romeo Maciuca and Song-Chun Zhu In this paper, we study a special case of the Metropolis algorithm, the Independence Metropolis Sampler (IMS), in the finite state space case. The IMS is often used in designing components of more complex Markov Chain Monte Carlo algorithms. We present new results related to the first hitting time of individual states for the IMS. These results are expressed mostly in terms of the eigenvalues of the transition kernel. We derive a simple form formula for the mean first hitting time and we show tight lower and upper bounds on the mean first hitting time with the upper bound being the product of two factors: a local factor corresponding to the target state and a global factor, common to all the states, which is expressed in terms of the total variation distance between the target and the proposal probabilities. We also briefly discuss properties of the distribution of the first hitting time for the IMS and analyze its variance. We conclude by showing how some non-independence Metropolis-Hastings algorithms can perform better than the IMS and deriving general lower and upper bounds for the mean first hitting times of a Metropolis- Hastings algorithm. Keywords: Eigenanalysis, Expectation, First Hitting Time, Independence Metropolis Sampler, Metropolis-Hastings. The authors are with the Department of Statistics, 825 Math. Science Bldg, Box 95554, University of California, Los Angeles, CA 90095. emails: {rmaciuca, sczhu}@stat.ucla.edu

Hitting time analysis for IMS 2 Introduction In this paper, we study a special case of the celebrated Metropolis algorithm the Independence Metropolis Sampler (IMS), for finite state spaces. The IMS is often used in designing components of more complex Markov Chain Monte Carlo algorithms. Using an acceptance-rejection mechanism described in section 3, the IMS simulates a Markov chain with target probability p = (p, p 2,..., p n ), by drawing samples from a more tractable probability q = (q, q 2,..., q n ). In the last two decades a considerable number of papers have been devoted to studying properties of the IMS. Without trying to be comprehensive, we shall briefly review some of the results that were of interest to us. For finite state spaces, Diaconis and Hanlon [3] and Liu [9] proved various upper bounds for the total variation distance between updated and target distributions for the IMS. They showed that the convergence rate of the Markov chain is upper bounded by a quantity that depends on the second largest eigenvalue: λ slem = min{ q i }. i p i A complete eigenanalysis of the IMS kernel was performed by Liu (996). He also compared the IMS with other two well known sampling techniques, rejection sampling and importance sampling. By making use of Liu s results, Smith and Tierney [2] obtained exact m-step transition probabilities for IMS, for both discrete and continuous state spaces. In the continuous case, if denoting by r = inf x {q(x) p(x) }, Mengersen and Tweedie [0] showed that if r is strictly less than, the chain is uniformly ergodic, while if r is equal to, the convergence is not even geometric anymore. Similar results were obtained by Smith and Tierney. These results show that the convergence rate of the Markov chain for the IMS is subject to a worst-case scenario. For the finite case, the state corresponding to the least probability ratio q i /p i is determining the rate of convergence, that is just one state from a potentially huge state space decides the rate of convergence of the Markov chain. A similar situation occurs in continuous spaces. To illustrate it, let us consider the following simple example. Example: Let q and p be two Gaussians having equal variances and the means slightly shifted. Then q, as proposal distribution, will approximate the target p very well. However,

Hitting time analysis for IMS 3 it is easy to see that inf x {q(x)/p(x)} = 0 and therefore the IMS will not have a geometric rate of convergence. This dismal behavior motivated our interest for studying the mean first hitting time as a measure of speed for Markov chains. This is particularly appropriate when dealing with stochastic search algorithms, when the focus could be on finding individual states rather than on the global convergence of the chain. For instance, in computer vision problems, one is often searching for the most probable interpretation of a scene and, to this end, various Metropolis-Hastings type algorithms can be employed. See Tu and Zhu [3] for examples and discussions. In such a context, knowledge of the behavior of the first hitting time of some states, like the modes of the posterior distribution of a scene given the input images, is of interest. The IMS was thoroughly studied, but most of the analysis focused on its convergence properties. Here, we analyze its first hitting time and derive formulas for its expectation and variance. These formulas are expressed mostly in terms of the eigenvalues of the transition kernel. We first review some general formulas for first hitting times. Then, we derive a formula for the mean f.h.t for ergodic kernels in terms of its eigen-elements and show that when the starting distribution of the chain is equal to one of the rows of the transition kernel, the mean f.h.t will have a particularly simple form. Using this result together with the eigen-analysis of the IMS kernel (briefly reviewed in section 3), we prove the main result, which gives an analytical formula for the mean f.h.t of individual states, as well as bounds. We show that, if in running an IMS chain the starting distribution is the same as the proposal distribution q, then after ordering the states according to their probability ratio, and if denoting by λ i the i th eigenvalue of the transition kernel, we have: i) E[τ(i)] = p i ( λ i ) ii) min{q i, p i } E[τ(i)] min{q i, p i }, p q T V where τ(i) stands for the f.h.t of i, and p q T V denotes the total variation distance between the proposal and target distributions. The result can be extended from individual sets to some subsets of state space, as we shall see in section 3. We then illustrate these results by a simple example.

Hitting time analysis for IMS 4 We conclude the section by proving that when starting from j i, the mean f.h.t of i are decreasing, with the smallest being equal to the mean f.h.t of i when starting from q: If q /p q 2 /p 2... q n /p n then : E [τ(i)] E 2 [τ(i)]... E i [τ(i)] E i+ [τ(i)] =... = E n [τ(i)] = E[τ(i)], i. Next, in section 3.5 and section 3.6 we focus on the tail distribution and the variance of the f.h.t for the IMS and we determine: An exponential upper bound on the tail: P (τ(i) > m) exp{ m(p i w )}, m > 0. If Z denotes the fundamental matrix associated with the IMS kernel then: V ar[τ(i)] = 2Z ii( λ i ) 3p i ( λ i ) + 2p i p 2 i ( λ, i. i) 2 Various bounds on the variance are also presented. Further, in section 4 we show how a special class of Metropolis-Hastings algorithms can outperform the IMS in terms of mean first hitting times. We prove that if Q is a stochastic proposal matrix satisfying Q ji /p i, Q ij /p j, i, j i, and R is the corresponding Metropolis-Hastings kernel then, for any initial distribution q, and as a corollary, E Q q [τ(i)] + q i p i, E Q q [τ(i)] min{q i, p i } EIMS q [τ(i)] i, where we denoted by Eq IMS [τ(i)] the mean f.h.t of the IMS kernel associated to q and p. We conclude by presenting lower and upper bounds on the hitting times for general Metropolis-Hastings kernels.

Hitting time analysis for IMS 5 2 General f.h.t for finite spaces Consider an ergodic Markov chain {X m } m on the finite space Ω = {, 2,..., n}. Let K be the transition kernel, p its unique stationary probability, and q the starting distribution. For each state i Ω, the first hitting time is defined below. Definition 2. The first hitting time for a state i is the number of steps for reaching i for the first time in the Markov chain sequence, τ(i) = min{m : X m = i}. E[τ(i)] is the mean first hitting time of i for the Markov chain governed by K. For any i, let us denote by K i deleting the i th the (n ) (n ) matrix obtained from K by column and row, that is, K i (k, j) = K(k, j), k i, j i. Also let q i = (q,..., q i, q i+,..., q n ). Then, it is immediate that P (τ(i) > m) = q i K m i, where := (,,..., ). This leads to the following formula for the expectation: E q [τ(i)] = + q i (I K i ), (2.) where I denotes the identity matrix. The existence of the inverse of I K i is implied by the sub-stochasticity of K i and the irreducibility of K (Bremaud [2]). More generally, the mean f.h.t of a subset A of Ω is given by E q [τ(a)] = + q A (I K A ), A Ω. (2.2) A different route is to consider the first hitting times if starting from a fixed j i. Here, we should define a different stopping time, by not counting starting from j as an initial step, but for simplicity, we will use the same notation and refer to E j [τ(i)] as the mean f.h.t of i when starting from state j. Then, for all j i, one has E j [τ(i)] = (Z ii Z ji )/p i. Z denotes the fundamental matrix, which, we recall, is defined to be Z = (I K + P ), where P denotes the matrix having all rows equal to p. When starting from q instead from a fixed state j, one has: E q [τ(i)] = + q j E j [τ(i)] = + q j (Z ii Z ji ). (2.3) p i j i j i For the rest of the paper we shall drop the subscript q whenever this will not create any notation confusion.

Hitting time analysis for IMS 6 The variance of the f.h.t can also be derived from the fundamental matrix Z. It is known that the second moment of τ(i), when starting from j, is determined by: E j [τ(i)] 2 = 2 p i (Z 2 ii Z 2 ji) p i (Z ii Z ji ) + 2 p 2 i Z ii (Z ii Z ji ), j i, (2.4) where the first term refers to the matrix Z 2. Hence, it is immediate that the second moment of the f.h.t when starting from q is just: E[τ(i) 2 ] = + 2 q j (Zii 2 Z p ji) 2 q j (Z ii Z ji ) + 2Z ii i p i p 2 i j j q j (Z ii Z ji ), (2.5) which readily leads to a formula for the variance. For more on the properties of the fundamental matrix Z and its connections to hitting times refer to Kemeny and Snell [6]. Next, we will show how knowing the eigen-structure of the transition matrix allows the direct computation of the mean f.h.t. Let {λ j } 0 j n be the eigenvalues of K and let v k = {v kl } 0 l n, and u k = {u kl } 0 l n be their corresponding right and left eigenvectors, such that U V = I, where U = {u k } k, V = {v k } k. As K is a stochastic matrix with stationary probability p, we have λ 0 = and we can fix v 0 = and u 0 = p respectively. Moreover, all the eigenvalues have real values and λ j <, j > 0. Proposition 2.. Using the same notations as before, for any ergodic kernel K and any initial distribution q, the mean first hitting time of i Ω is E[τ(i)] = + n p i k= λ k u ki (v ki l j q l v kl ). In particular, if q is chosen to be row j th of K for arbitrary j Ω, then Proof: Using (2.3), E[τ(i)] = n u ki (v ki v kj ) + δ ij. p i λ k p i k= E[τ(i)] = + q j (Z ii Z ji ). (2.6) p i Let us recall that Z and K share the same system of eigenvectors, while the eigenvalues of Z are β 0 =, β j = /( λ j ), j n. Therefore, we can apply the spectral j i decomposition theorem (see Bremaud [2]), to get: n Z li = β k v kl u ki = v 0l u 0i + k=0 n k= n v kl u ki = p i + λ k k= λ k v kl u ki, l, i.

Hitting time analysis for IMS 7 Therefore, From (2.7) and (2.6), we get Z ii Z ji = n k= E[τ(i)] = + q j (Z ii Z ji ) = + p i p i j i which, by changing the summation order, turns into E[τ(i)] = + n p i k= λ k (v ki v kj )u ki. (2.7) j i n q j k= λ k (v ki v kj )u ki, u ki q j (v ki v kj ). (2.8) λ k Noting that j i q j(v ki v kj ) can be rewritten as v ki l q lv kl, the first part of the proof is completed. For the second part, assume that q = K j. This implies that l q lv kl = l K jlv kl = (Kv k ) j. But as v k is a right eigenvector for λ k, we get l q lv kl = λ k v kj and by plugging this into the general formula just proved, E[τ(i)] = + n p i Or k= We have to consider two cases: j i u ki (v ki λ k v kj ) = + n λ k p i k= k= λ k u ki (v ki v kj + ( λ k )v kj ). E[τ(i)] = + n u ki (v ki v kj ) + n u ki v kj. (2.9) p i λ k p i i) j = i. In this case, n k= u kiv kj = n k=0 u kiv ki p i = p i since n k=0 u kiv ki =. Therefore, from (2.9) it follows that E[τ(i)] = /p i, the first sum cancelling for j = i. ii) j i. Then, again, n k= u kiv kj = n k=0 u kiv kj p i = δ ij p i = p i. Now, using (2.9) E[τ(i)] = + n p i k= u ki (v ki v kj ) = n λ k p i k= 3 Hitting time analysis for the IMS k= λ k u ki (v ki v kj ). Here, we shall capitalize on the previous result to prove our main theorem. But first, let us set the stage by briefly introducing the IMS.

Hitting time analysis for IMS 8 3. The Independence Metropolis Sampler The IMS is a Metropolis-Hastings type algorithm with the proposal independent of the current state of the chain. It has also been called Metropolized Independent Sampling (Liu [9]). The goal is to simulate a Markov chain {X m } m 0 taking values in Ω and having stationary distribution p (the target probability). To do this, at each step a new state j Ω is sampled from the proposal probability q = (q, q 2,..., q n ) according to j q j, which is then accepted with probability α(i, j) = min{, q i p i p j q j }. Therefore, the transition from X m to X m+ is decided by the transition kernel having the form q j α(i, j), j i, K(i, j) = k i K(i, k), j = i. The initial state could be either fixed or generated from a distribution whose natural choice in this case is q. In section 3.3, we show why it is more efficient to generate the initial state from q instead of choosing it deterministically. It is easy to show that p is the invariant (stationary) distribution of the chain. In other words, p K = p. Since from q > 0 it follows that K is ergodic, then p is also the equilibrium distribution of the chain. Therefore, the marginal distribution of the chain at step m, for m large enough, is approximately p. However, instead of trying to sample from the target distribution p, one may be interested in searching for a state i with maximum probability: i = arg max i Ω p i. Here is where the mean f.h.t can come into play. E[τ(i)] is a good measure for the speed of search in general. As a special case we may want to know E[τ(i )] for the optimal state. As it shall become clear later, a key quantity to the analysis is the probability ratio w i = q i /p i. It measures how much knowledge the heuristic q i has about p i, or in other words how informed is q about p for state i. Therefore we define the following concepts. Definition 3. A state i is said to be over-informed if q i > p i and under-informed if q i < p i. There are three special states defined below.

Hitting time analysis for IMS 9 Definition 3.2 A state i is exactly-informed if q i = p i. A state i is most-informed (or least-informed) if it has the highest (or lowest) ratio w i : i max = arg max i Ω { w i }, i min = arg min i Ω { w i }. Liu [9] noticed that the transition kernel can be written in a simpler form by reordering the states increasingly according to their informedness. Since for i j, K ij = q j min{, w i /w j }, if w w 2... w n it follows that w i p j i < j, K ij = k<i q k w i k>i p k i = j, q j = w j p j i > j. Without loss of generality, we shall assume for the rest of the paper that the states are indexed such that w w 2... w n, to allow for this more tractable form of the transition kernel. Proposition 2. can be used to compute mean first hitting times whenever an eigenanalysis for the transition kernel is available. In practice, this situation is quite rare though. However, such an eigen-analysis is available for the IMS. We review these results below and then proceed with our results. 3.2 The eigenstructure of the IMS A first result concerns the eigenvalues and right eigenvectors of the IMS kernel. Theorem 3. (J. Liu, 996) Let T k = i k q i and S k = i k p i. Then, the eigenvalues of the transition matrix K are λ k = T k w k S k, k n, λ 0 =, and they are decreasing as λ 0 > λ λ 2 λ n 0. Moreover, the right eigenvector corresponding to λ k, k > 0 is v k = (0,, 0, S k+, p k,, p k ), with the first k entries being 0 and v 0 = (,,..., ). Remark: It is easy to see now that the eigenvalues of K are incorporated in the diagonal terms of K through the equality K ii = λ i + q i, which will be often used later on. Smith and Tierney [2] computed the exact k-step transition probabilities for the IMS. One of their results reveals in fact the very structure of the left eigenvectors. Suppose δ k is the unit vector with in the k th position ( k n) and 0 everywhere else. They showed that:

Hitting time analysis for IMS 0 Proposition 3.2 (Smith and Tierney, 996) For k n, δ k = p k v 0 + k v k p k S k j= v j S j S j+ while for k = n, n v j δ n = p n v 0 p n. S j S j+ j= As a corollary, the left eigenvectors of K are given by: Corollary 3.3 u 0 = p, u k = (0, 0,..., 0,, p k+,..., ) T, k n, S k S k S k+ S k S k+ where for k > 0 the first k entries are 0. p n 3.3 Main Result We are now able to compute the mean f.h.t for the IMS and provide bounds for it, by making use of the eigen-structure of the IMS kernel as well as of Proposition 2.. Theorem 3.4 Assume a Markov chain starting from q is simulated according to the IMS transition kernel having proposal q and target probability p. Then, using previous notations: i) E[τ(i)] =, i Ω, p i ( λ i ) ii) min{q i, p i } E[τ(i)] min{q i, p i }, p q T V where we define λ n to be equal to zero and p q T V denotes the total variation distance between p and q. Equality is attained for the three special states from Definition 3.2. Proof: i) Let us first note that we are in the situation from the second part of Proposition 2.. That is, after reordering the states according to their probability ratios, our initial distribution q is equal to the n th row of K as it can easily be seen. Then, from Proposition 2., one has: E[τ(i)] = n u ki (v ki v kn ) + δ in. (3.) p i λ k p i k=

Hitting time analysis for IMS From Theorem 3., v ki = v kn, k < i, while from Corollary 3.3, u ki = 0 for k > i. Hence u ki (v ki v kn ) = 0, k i. If i = n, then E[τ(i)] = δ in /p i = /[p n ( λ n )], while for i < n one has E[τ(i)] = u ii(v ii v in ) p i ( λ i ). Using the eigen-analysis for the IMS, we can write u ii (v ii v in ) = (S i+ ( p i ))/S i =, so the expectation becomes E[τ(i)] = /[p i ( λ i )], and the proof of i) is completed. ii) By using i) it is obvious that E[τ(i)] /p i since 0 λ i <. Therefore, we only need to show that λ i w i which would imply that E[τ(i)] /q i. Noting that λ i = q i + q i+ +... + q n (p i + p i+ +... + p n )w i, we need to prove that w i = q i p i q + q 2 +... + q i p + p 2 +... + p i. This is follows quickly since for any j < i, w j w i q j p j w i. To prove the upper bound, let us first get a more tractable form for p q T V. We partition the state space into two sets: under-informed and over-informed with the exactlyinformed states in either set: Ω = Ω under Ω over. As the states are sorted, let k n be their dividing point Ω under = {i k : q i p i }, Ω over = {i > k : q i > p i }, where Ω over can be the empty set if q = p. By definition, p q T V = 2 i Ω (p i q i ) = 0, we have i p i q i. Since p q T V = p i q i = (p i q i ) + (q i p i ) 2 2 2 i Ω i Ω under i Ω over = (q i p i ) = T k+ S k+, (3.2) i Ω over where we define T n+ = S n+ = 0. We prove the upper bound for the under-informed and over-informed states respectively. Case I. upper bound for under-informed states i k. For under-informed states, q i = min{p i, q i }. As λ i = T i w i S i, it follows that: p i ( λ i ) = p i ( T i ) + q i S i = p i ( T i+ ) p i q i + q i S i+ + q i p i = p i ( T i+ ) + q i S i+. Therefore, p i ( λ i ) q i ( T i+ +S i+ ). By using (3.2), we get min{p i, q i }( p q T V ) = q i ( T k+ +S k+ ). Thus, we only need to show that S i+ S k+ T i+ T k+. By definition,

Hitting time analysis for IMS 2 this is equivalent to p i+ + p i+2 +... + p k q i+ + q i+2 +... q k, which is obviously true because states i +,..., k are under-informed. Equality is attained for p j = q j, j [i, k], which is at the exactly-informed states. Case II. upper bound for over-informed states i > k. As min{p i, q i } = p i, it suffices to show that p i ( λ i ) p i ( T k+ + S k+ ), or λ i T k+ S k+. Because λ i λ k+, it suffices to prove that λ k+ T k+ S k+, or T k+ w k+ S k+ T k+ S k+, which is trivial since w k+ for over-informed states. Equality in this case is obtained if λ i = λ i =... = λ k+ and w k+ = which is equivalent to w k+ = w k+2 =... = w i =. Theorem 3.4 can be extended by considering the first hitting time of some particular sets. The following corollary holds true. Corollary 3.5 Let A Ω of the form A = {i+, i+2,..., i+k}, with w w 2... w n. Denote p A := p i+ + p i+2 +... + p i+k, q A := q i+ + q i+2 +... + q i+k, w A := q A /p A and λ A := (q i+ +... + q n ) (p i+ +... + p n )w A. Then i) and ii) from Theorem 3.4 hold ad-literam with i replaced by A. Proof: We will only prove part i) since the proof of ii) is analogous to the one in Theorem 3.4. Let A = {i +, i + 2,..., i + k}. We notice that w w 2... w i w A w i+k+... w n. Therefore, if we consider A to be a singleton, the problem of computing the mean f.h.t of A reduces to computing the mean f.h.t of the singleton A in the reduced space Ω A := {, 2,..., i, {A}, i + k +,..., n}. The new proposal (respectively target) probability would be q restricted to the space Ω A by putting mass q A on the state {A} (similarly for p). It is easy to check that K A = K {A}, where the last matrix is obtained if we consider A to be a singleton (it is essential that the ordering of the states according to the probability ratios is the same in Ω as in Ω A ). Now, we can apply (2.2) to obtain E Ω [τ(a)] = + q A (I K A ) = + q {A} (I K {A} ), and by using Theorem 3.4 for Ω A, E Ω [τ(a)] = E ΩA [τ({a})] = /[p A ( λ A )]. We used the subscripts Ω or Ω A to indicate which space we are working on. In the introduction part we hinted at showing why generating the initial state from q is preferable to starting from a fixed state j i. The following result attempts to clarify this issue.

Hitting time analysis for IMS 3 Proposition 3.6 Assuming that w w 2... w n, the following inequalities hold true: E [τ(i)] E 2 [τ(i)]... E i [τ(i)] E i+ [τ(i)] =... = E n [τ(i)] = E[τ(i)], i Ω. Proof: We saw that E j [τ(i)] = n u ki (v ki v kj ), j i. p i λ k k= i) j > i. Then one has u ki (v ki v kj ) = u ki (v ki v kn ), k > 0 and therefore, E j [τ(i)] = n u ki (v ki v kn ) = E[τ(i)]. p i λ k k= ii) j < i. Let us compute the difference E j [τ(i)] E j+ [τ(i)] for arbitrary j. E j [τ(i)] E j+ [τ(i)] = n p i k= k= (v ki v kj )u ki n λ k p i k= λ k (v ki v k(j+) )u ki = = n (v k(j+) v kj )u ki. (3.3) p i λ k If j < i then for k < j we have v k(j+) = 0 = v kj, while for j + < k < i, v k(j+) = p k = v kj, so in both cases the difference is zero, which cancels the corresponding terms in (3.3). The terms for k > i cancel too, because u ki = 0. The only remaining terms are those for k = j, j +. Therefore, E j [τ(i)] E j+ [τ(i)] = [ ] (v j(j+) v jj )u ji + (v (j+)(j+) v (j+)j )u (j+)i. p i λ j λ j+ We note that (v j(j+) v jj )u ji = ( p j S j+ )( p i /(S j S j+ )) = p i /S j+. Similarly, (v (j+)(j+) v (j+)j )u (j+)i = S j+2 ( p i /(S j+ S j+2 )) = p i /S j+. Hence, E j [τ(i)] E j+ [τ(i)] = p i ( λ j λ j+ ) pi S j+ = S j+ ( λ j ) 0. λ j+ Equality case is obtained if w j = w j+ which implies λ j = λ j+. Therefore, if states j and j + have the same informedness, it would make no difference from which one of them the sampler would start. The only thing left to prove is that E i [τ(i)] E[τ(i)]. To do this, we note that one can write (3.3) with i in the place of j and i + instead of j +. This gives E i [τ(i)] E i+ [τ(i)] = p i i k= λ k (v k(i+) v k(i ) )u ki.

Hitting time analysis for IMS 4 As before, all the terms cancel except for k = i, i and analogously, E i [τ(i)] E i+ [τ(i)] = ( ) 0. S i λ i λ i As E i+ [τ(i)] = E j [τ(i)] = E[τ(i)], j > i, the proof of Proposition 3.6 is completed. 3.4 Example We can illustrate the main results in Theorem 3.4 through a simple example. We consider a space with n = 000 states. Let p and q be mixtures of two discretized Gaussians with tails truncated and then normalized to one. They are plotted as solid (p), dashed (q) curves in Fig.a. Fig.b plots the logarithm of the expected first hitting-time ln E[τ(i)]. The lower and upper bounds from Theorem 3.4 are plotted in logarithm scale as dashed curves which almost coincide with the hitting-time plot. For better resolution we focused on a portion of the plot around the mode, the three curves becoming more distinguishable in Fig.c. We can see that the mode x = 333 has p(x ) 0.02 and it is hit in E[τ x ] 62 times on average for q. This is much smaller than n/2 = 500 which would be the average time for exhaustive search. In comparison, for an uninformed (i.e uniform) proposal the result is E[τ x ] = 000. Thus, it becomes visible how a good proposal q can influence the speed of such a stochastic sampler. p (solid) vs. q (dashed) 70 logarithm of the expected hit time: lne 7.5 logarithm of the expected hit time:zoom in 0.02 60 7 0.0 50 6.5 0.008 6 40 0.006 5.5 30 5 0.004 20 4.5 0.002 0 4 0 00 200 300 400 500 600 700 800 900 000 0 0 00 200 300 400 500 600 700 800 900 000 3.5 320 330 340 350 360 370 380 (a) p (solid) vs q (dashed) (b) ln E[τ(i)] and bounds (c) zoom in for (b) Figure : Mean f.h.t and bounds

Hitting time analysis for IMS 5 3.5 Tail Distribution It is known (Abadi []) that the distribution of first hitting times is generally well approximated by an exponential distribution. This can be illustrated on a small example. Consider a state space with N = 0 states, with p and q being discretized mixtures of Gaussians as before. Fig. 2a plots the tail distributions of the f.h.t for all the states of the space. It is apparent that their shapes generally resemble exponential tails. In Fig. 2b, we plotted both the tail distribution of the f.h.t (in solid) and the corresponding exponential distribution (dashed) for an arbitrary state (i = 3). Even though we are not able to quantify the approximation error in general, we can give an exponential upper bound on the tail distribution of the f.h.t. The bound is shown in Fig. 2c. 0.9 0.9 Tail distrib of the f.h.t and the corresponding exponential distrib for i=3,n=0. f.h.t exponential 0.9 Tail distrib of the f.h.t for i=3 with upper bound f.h.t upper bound 0.8 0.8 0.8 0.7 0.7 0.7 0.6 0.6 0.6 0.5 0.5 0.5 0.4 0.4 0.4 0.3 0.3 0.3 0.2 0.2 0.2 0. 0. 0. 0 0 50 00 50 200 250 300 350 400 450 500 0 0 0 20 30 40 50 60 70 80 90 00 0 0 50 00 50 200 250 (a) Tail distributions (b) F.h.t and the corresponding exponential distribution Figure 2: Tails, exponential approximation and the upper bound (c) Upper bound for the tail Proposition 3.7 For all i Ω, P (τ(i) > m) ( q i )( p i w ) m exp{ m(p i w )}, m > 0. Proof: For all j i we can write K ji = p i min{w i, w j }. This shows that K ji p i w, j i. Or, equivalently, K ji p i w, j i. By writing this set of inequalities in matrix form, one gets K l i ( p i w ) l, l. K i ( p i w ). Now, we can iterate this inequality and therefore, We recall that P (τ(i) > m) = q i K m i. Hence, by taking l = m we shall obtain P (τ(i) > m) ( p i w ) m q i, or finally, P (τ(i) > m) ( q i )( p i w ) m.

Hitting time analysis for IMS 6 For the second of the inequalities we note that w i w, so q i p i w, which readily gives P (τ(i) > m) ( p i w ) m. But ( p i w ) m exp{ m(p i w )}, since exp{ x} x, x, and the proof is completed. Remarks: ) We note that the last of the inequalities in Proposition 3.7 holds also for the exponential distribution µ(i), having mean equal to E[τ(i)]. That is, P (µ(i) > m) exp{ m(p i w )}, m > 0. To see why, note that λ i λ = w, so E[τ(i)] = /[p i ( λ i )] /(p i w ) or /E[τ(i)] p i w. Then, obviously, P (µ(i) > m) = exp{ m/e[τ(i)]} exp{ m(p i w )}. 2) We can always use Markov s inequality to upper bound the tail distribution with respect to the expectation of the f.h.t of some i. That is, for any k positive integer, one has: P (τ(i) > ke[τ(i)]) k. If we were to avoid the use of eigenvalues, whose values might be too difficult to compute, we could rely on the upper bound from Theorem 3.4 which combined with the above gives: P (τ(i) k min{p i, q i }( p q ) ) k. 3.6 Variance In this section, we derive a formula for the variance of the f.h.t for the IMS. We show that: Theorem 3.8 If Z is the fundamental matrix associated to the IMS kernel and using the same notations as before, the variance of the first hitting time of i is given by: V ar[τ(i)] = 2Z ii( λ i ) 3p i ( λ i ) + 2p i p 2 i ( λ i) 2, i Ω. Proof: Already knowing the expectation of the f.h.t reduces the problem of computing the variance to finding E(τ(i) 2 ). This is given by (2.5): E[τ(i) 2 ] = + 2 q j (Zii 2 Z p ji) 2 q j (Z ii Z ji ) + 2Z ii i p i p 2 i We can rewrite the above as: j j q j (Z ii Z ji ). j E[τ(i) 2 ] = + 2 p i (Z 2 ii j q j Z 2 ji) p i (Z ii j q j Z ji ) + 2Z ii (Z p 2 ii i j q j Z ji ). (3.4)

Hitting time analysis for IMS 7 Let us note that j q jz ji = j K njz ji = (KZ) ni. Also, recall that KZ = Z + P I. (3.5) Thus, j q jz ji = Z ni + p i δ ni. Similarly, j q jz 2 ji = j K njz 2 ji = (KZ 2 ) ni. From (3.5) it also follows that KZ 2 = Z 2 + P Z, hence j q jz 2 ji = Z 2 ni + p i Z ni. By transforming (3.4) we get E[τ(i) 2 ] = + 2 (Zii 2 Zni 2 p i + Z ni ) (Z ii Z ni p i + δ ni )+ 2Z ii (Z p i p i p 2 ii Z ni p i + δ ni ), i which further reduces to E[τ(i) 2 ] = 2 (Zii 2 Z p ni) 2 3 (Z ii Z ni ) + 2Z ii (Z i p i p 2 ii Z ni ) + δ ni ( 2Z ii ). (3.6) i p i p i For i = n (3.6) becomes E[τ(n) 2 ] = (2Z nn p n )/p 2 n, so V ar[τ(n)] = E[τ(n) 2 ] (E[τ(n)]) 2 = (2Z nn p n )/p 2 n /p 2 n or finally, which is what I wanted since λ n = 0. If i < n, let us rewrite (3.6), for clarity: V ar[τ(n)] = 2Z nn p n, p 2 n E[τ(i) 2 ] = 2 (Zii 2 Z p ni) 2 3 (Z ii Z ni ) + 2Z ii (Z i p i p 2 ii Z ni ). (3.7) i We use the spectral decomposition theorem for Z 2 and obtain Therefore, we have which, just as before, leads to Z 2 li = p i + Z 2 ii Z 2 ni = n k= n k= ( λ k ) 2 v klu ki, l, i. ( λ k ) 2 u ki(v ki v kn ), Z 2 ii Z 2 ni = u ii(v ii v in ) ( λ i ) 2 = ( λ i ) 2. (3.8) It is also noted that Z ii Z ni = p i E n [τ(i)]. At the same time, from Proposition 3.6, E n [τ(i)] = E[τ(i)] = /[p i ( λ i )]. Therefore, Z ii Z ni = λ i. (3.9)

Hitting time analysis for IMS 8 Now using (3.8) and (3.9) in (3.7) we obtain Or E[τ(i) 2 ] = 2 p i ( λ i ) 3 2 p i ( λ i ) + 2Z ii p 2 i ( λ i). E[τ(i) 2 ] = 2Z ii( λ i ) 3p i ( λ i ) + 2p i p 2 i ( λ i) 2. Since V ar[τ(i)] = E[τ(i) 2 ] E[τ(i)] 2 and E[τ(i)] = /(p i ( λ i )), it is immediate that 3.6. Bounds for the variance V ar[τ(i)] = 2Z ii( λ i ) 3p i ( λ i ) + 2p i p 2 i ( λ i) 2. Two corollaries of Theorem 3.8 will offer bounds on the variance of the f.h.t. By bounding the term Z ii in Theorem 3.8, we obtain Corollary 3.9, which gives bounds for the variance mainly in terms of the expectation of the f.h.t: Corollary 3.9 Let us denote E i := E[τ(i)], for any i Ω.Then, E i (E i ) E i [( + 2q i )E i 3] V ar[τ(i)] E i [ 2( + q i) w p i E i 3], with equality if w i = w i =... = w. Proof: For the proof we first need to prove the following lemma. Lemma 3.0 + q i p i λ i Z ii + q i p i w, with equality if and only if w i = w i =... = w. Proof of the lemma: We use again the identity Z ii = p i + n k= λ k v ki u ki. As /( λ k ) = + λ k /( λ k ), we can rewrite Z ii as n Z ii = p i + v ki u ki + j= n k= λ k n v ki u ki = + λ k k= λ k λ k v ki u ki = + i k= λ k λ k v ki u ki.

Hitting time analysis for IMS 9 Note that /( λ i ) /( λ j ) /( λ ), j i, which gives + λ i i k= λ k v ki u ki Z ii + λ i λ k v ki u ki. Since i k= λ kv ki u ki = n k= λ kv ki u ki = K ii p i = q i + λ i p i, it follows that + (q i + λ i p i )/( λ i ) Z ii + (q i + λ i p i )/( λ ), or further, ( + q i p i )/( λ i ) Z ii ( + q i p i + λ i λ )/( λ ). The lemma is proved if, for the right hand side term, we use λ = w and λ i λ. Clearly, equality on both sides is obtained if and only if w i = w i =... = w. Going back to the proof of Corollary 3.9, we note that, starting from the left side, the first inequality is trivial since E i /q i. Also, proving that E i [( + 2q i )E i 3] V ar[τ(i)] is just a matter of applying Lemma 3.0 and regrouping the terms. For the upper bound we notice that w = λ λ i, which gives p i p i ( λ i )/w. This combined with the upper bound for Z ii will show that Z ii ( λ i ) + p i ( + q i p i )( λ i ) w + p i( λ i ) w = ( + q i)( λ i ) w. Since from Theorem 3.8, V ar[τ(i)] = [2Z ii ( λ i ) 3p i ( λ i ) + 2p i ]/[p 2 i ( λ i ) 2 ], it follows that V ar[τ(i)] [2( + q i )( λ i )/w 3p i ( λ i ) ]/[p 2 i ( λ i ) 2 ], which easily turns into the pursued upper bound since, from Theorem 3.4, E i = /[p i ( λ i )]. The equality case shows up if λ i = λ i =... = λ, which is equivalent to w i = w i =... = w. The bounds given by Corollary 3.9 can be further simplified, but weakened at the same time, if one uses the known lower bound for E i on the left and maximizes the upper bound with respect to E i. Thus, one gets: Corollary 3. If M i := / min{q i, p i }, for any i Ω, then M i (M i ) M i [M i ( + 2q i ) 3] V ar[τ(i)] k= ( + qi 3 ) 2. w p i 2 Proof: Obviously, for the lower bounds we apply inequality E i M i to the previous corollary. To prove the upper bound, we refer again to Corollary 3.9 and for simplicity, let us denote 2 ( + q i) w p i 3 := a. Then, Corollary 3.9 gives V ar[τ(i)] E i (a E i ). Consequently, a E i > 0, and since the maximum value of function f(x) := x(a x) on (0, a) is obtained for x = a/2, we conclude that f(e i ) a 2 /4, which is the upper bound.

Hitting time analysis for IMS 20 4 IMS vs. a special class of Metropolis-Hastings kernels We have seen that for the IMS the mean f.h.t is always bounded below by /p i, for all proposal probabilities q. We shall prove that for more general Metropolis kernels, the mean f.h.t can be lower than /p i, and thus show formally, what was otherwise clear intuitively, that, because of its independence from the current state, the IMS kernel can be inferior to other samplers in terms of speed of hitting a certain state. Firstly, we recall that a Metropolis-Hastings kernel R, induced by a proposal stochastic matrix Q can be written as R ij = Q ij min{, Q ji p j /(Q ij p i )}, for any i j (Hastings [5]). Theorem 4. Let Q be a stochastic proposal matrix satisfying the condition Q ji p i, j i. Then, for any initial distribution q, the Metropolis-Hastings Markov chain that uses Q as a proposal matrix has the property that E Q q [τ(i)] + q i p i, i Ω, with equality for Q equal to the stationary matrix. Proof: Let R be the Metropolis-Hastings kernel associated to the proposal Q and the target probability p. Let i Ω. As Q ji p i and Q ij p j, it follows that R ji = min{q ji, Q ij p i /p j } p i, j i. This implies that R ji p i or R j n p i. As the previous inequality holds true for all j i, we get that R i ( p i ) or equivalently (I R i ) p i. The inverse of I R i exists and it is equal to m Rm i and therefore (I R i ) 0. This said, we can multiply the inequality (I R i ) p i by q i (I R i ) and get q i (I R i ) ( q i )/p i or finally, E Q q [τ(i)] + ( q i )/p i, where we have used formula (2.) for the mean f.h.t when starting from q. We have equality if R ji = p i, j i, which is fulfilled if Q equals the stationary matrix. Naturally, there are also other Q s that accomplish equality, the condition being that either Q ji = p i or Q ij = p j, j i. Combining Theorem 3.4 and Theorem 4., one gets:

Hitting time analysis for IMS 2 Corollary 4.2 For any initial distribution q and Q satisfying the assumption in Theorem 4., Eq Q [τ(i)] max{, } Eq IMS [τ(i)], p i q i where we denoted by E IMS q [τ(i)] the mean f.h.t of the IMS kernel associated to q and p. Proof: The proof is immediate since, obviously, + ( q i )/p i max{/p i, /q i }, with equality if and only if q i = p i or, in other words, if i is an exactly-informed state for q. The above corollary thus gives a simple way to construct Metropolis-Hastings samplers that would perform better than a corresponding IMS sampler in terms of first hitting times. It is worth noting that there are known examples of samplers that satisfy the condition in Theorem 4.. Such a sampler is the Metropolized Gibbs Sampler (Liu [8]) or simply MGS. A recent application of this sampler is described in Tu and Zhu [3]. For the MGS, the proposal matrix Q is defined as: Q ij satisfies the above mentioned condition. = p j /( p i ), i j which Interestingly, the MGS can also be viewed as a particular case of the IMS. To see this, let us remark that after metropolizing Q through the usual acceptance-rejection mechanism, one gets the transition kernel having elements: p j if i < j, pi R ij = k i R ki if i = j, p j if i > j. pj Without loss of generality, we assume that p p 2... p n. Now, if we denote by q i := p i /( p i ), i < n and q n := i<n q i, we note that R has the same form as the IMS transition matrix corresponding to p and q for w w 2... w n. Therefore, if using as initial distribution the newly defined q, all the previous results pertaining to the IMS apply also to the MGS. Remark: The MGS is a modified Gibbs sampler, the main difference being that it will never propose the current state. Thus, it travels through the state space in a more efficient manner. However, a M-H acceptance probability needs to be introduced to maintain the correct invariant distribution. Rejections could still cause the sampler to stay in the same state. Nevertheless, Liu [8] showed that the MGS is more efficient than the ordinary Gibbs

Hitting time analysis for IMS 22 sampler in the sense that the asymptotic variance of the estimators based on the Markov chain samples is smaller for the MGS than for the Gibbs sampler. Thus, the expected gain in efficiency would justify metropolizing the stationary probability p. A recent review of various types of efficiency definitions for MCMC samplers as well as theoretical results linking these types of efficiency notions can be found in Mira []. Our approach is to consider the mean first hitting time as an indicator of efficiency when the focus is on searching for a few states through a finite state space. 5 General bounds for expected f.h.ts for Metropolis- Hastings kernels It turns out that one can get lower and upper bounds on the expected f.h.t for any Metropolis-Hastings kernel by a reasoning similar to the one in Theorem 4. as shown with the result below. Theorem 5. Let p and Q be the target probability and the proposal matrix respectively for a Metropolis-Hasting sampler. Let M = max i,j Q ij /p j and m = min i,j Q ij /p j. We assume m > 0. Then for any initial distribution q, the expected f.h.ts are bounded by p i + q i M p ieq Q [τ(i)] p i + q i m, i. Equality is attained if Q ij = p j, i, j. Proof: The proof is similar to the one for Theorem 4. so it will only be sketched. Firstly one shows that mp i K ji Mp i which then leads to (mp i ) (I K i ) (Mp i ), which in turn, by an argument analogous to the one in Theorem 4., gives ( q i )/Mp i q i (I K i ) ( q i )/mp i. Now using the corresponding identity for expected f.h.t, Eq Q [τ(i)] = + q i (I K i ), one immediately gets the result stated in Theorem 5.. For some particular choices of q, the bounds can be made more concrete. Two particular choices seem both intuitive and convenient: i) q i = j α j Q ji, i, ( j α j = ) ii) q i = max j Q ji k max j Q jk, i.

Hitting time analysis for IMS 23 It is immediate to check that both i) and ii) are valid probability distributions. The first one is just a linear combination of the elements on the ith column, which in the particular case when the proposal does not depend on the current state (the IMS), reduces to the proposal distribution for the IMS. The second distribution described above would use the maximum value of the proposal matrix on each column as an initial step, after normalization. Using them as initial distributions for the Markov chain and applying Theorem 5. one can derive the corollary below: Corollary 5.2 Within the setup from Theorem 5., the following hold: ) If the initial distribution is given by i) then /M p i Eq Q [τ(i)] /m, i. 2) If the initial distribution is given by ii), then max{/n, /M} p i Eq Q [τ(i)] < + /m, i. Proof: To prove ), let us first notice that from the way M and m where defined it follows that mp i q i Mp i. Therefore p i + q i M p i + Mp i M = M and analogously p i + q i m which proves ) by means of Theorem 5.. p i + mp i m = m For 2), we first need to show that mp i /n < q i Mp i, i. We note that = k Q k k max Q jk < j k = n Therefore, max j Q ji n < q i = max j Q ji k max max j Q jk Hence mp i /n < q i Mp i, i. Now, as for 2), one will get M p ieq Q [τ(i)] < p i + p im/n m j Q ji < + m The only thing left to prove is that p i E Q q [τ(i)] /n. In order to show this, we employ the second basic identity for the kernel K, which is pk = p. If we denote by u the n dimensional vector obtained from row i of K after deleting component K ii, we can write p i K i + p i u = p i or p i (I K i ) = p i u. We note that K ij Q ij max i Q ij < nq j, j.

Hitting time analysis for IMS 24 Therefore, u < nq i so p i (I K i ) < (np i )q i. As before, it follows that p i < (np i )q i (I K i ) or p i < (np i )(Eq Q [τ(i)] ). Hence p i Eq Q [τ(i)] > p i + ( p i )/n, implying that p i Eq Q [τ(i)] > /n which concludes the proof of the corollary. Naturally, because of their generality the bounds developed in this section are quite weak in general, as m and M can take very extreme values in practice, rendering the bounds useless for such cases. 6 Conclusion We were able to perform a detailed first hitting analysis of one special type of Metropolis- Hastings sampler, the Independence Metropolis Sampler. More practical general nonindependence Metropolis-Hastings samplers seem to be too complex to allow for a detailed analysis. In the spirit of this paper, such an analysis could only be done if the eigenstructure of the kernel matrix would be available. This is typically not the case for most of the practical applications. However, even when the eigenstructure is unknown, insights into the behavior of mean first hitting time are possible, as seen in Theorem 4.. Also, we have derived lower and upper bounds for the expected first hitting time for general Metropolis- Hastings algorithms. We are hoping that future work will allow obtaining better bounds in the case of the tail distribution for the IMS and further for more general cases. This would make our results more useful and amenable for using them in practice. Acknowledgment This work was supported by an NSF grant (IIS-0244763) and a Sloan fellowship. References [] Abadi, M. and Galves, A. (200). Inequalities for the occurrence times of rare events in mixing processes. The state of the art, Markov Proc. Relat. Fields 7, 97-2.

Hitting time analysis for IMS 25 [2] Bremaud, P. (999). Markov Chains: Gibbs Fields, Monte Carlo Simulation and Queues, Springer, New York. [3] Diaconis, P. and Hanlon, P. (992). Eigen Analysis for Some Examples of the Metropolis Algorithm, Contemporary Mathematics 38, 99-7. [4] Diaconis, P. and Saloff-Coste, L. (998). What Do We Know about the Metropolis Algorithm?, Journal of Computer and System Sciences 57, 20-36. [5] Hastings, W.K. (970). Monte Carlo Sampling Methods using Markov Chains and Their Applications, Biometrika 57, 97-09. [6] Kemeny, J. G. and Snell, J. L. (976). Finite Markov Chains, Springer Verlag. [7] Liu, J.S. (200). Monte Carlo Strategies in Scientific Computing, Springer Verlag. [8] Liu, J.S. (996). Metropolized Gibbs sampler: an improvement. Technical report, Dept. Statistics, Stanford Univ. [9] Liu, J.S. (996), Metropolized Independence Sampling with Comparisons to Rejection Sampling and Importance Sampling, Statistics and Computing 6, 3-9. [0] Mengersen, K.L. and Tweedie, R.L. (994). Rates of Convergence of the Hastings and Metropolis Algorithms, Annals of Statistics 24, 0-2. [] Mira, A. (2002). Ordering and improving the performance of Monte Carlo Markov Chains. Statistical Science 6, 340-350. [2] Smith, R.L. and Tierney, L. (996). Exact Transition Probabilities for Metropolized Independence Sampling, Technical Report, Dept. of Statistics, Univ. of North Carolina. [3] Tu, Z.W. and Zhu, S.C. (2003). Parsing Images into Regions, Curves, and Curve Groups, submitted.