c 2006 Society for Industrial and Applied Mathematics 

SIAM J. CONTROL OPTIM. Vol. 45, No. 5, pp. 1633–1656

SIMULATION-BASED UNIFORM VALUE FUNCTION ESTIMATES OF MARKOV DECISION PROCESSES∗ RAHUL JAIN† AND PRAVIN P. VARAIYA† Abstract. The value function of a Markov decision process (MDP) assigns to each policy its expected discounted reward. This expected reward can be estimated as the empirical average of the reward over many independent simulation runs. We derive bounds on the number of runs needed for the uniform convergence of the empirical average to the expected reward for a class of policies, in terms of the Vapnik-Chervonenkis or P-dimension of the policy class. Further, we show through a counterexample that whether we get uniform convergence or not for an MDP depends on the simulation method used. Uniform convergence results are also obtained for the average-reward case, for partially observed Markov decision processes, and can be easily extended to Markov games. The results can be viewed as a contribution to empirical process theory and as an extension of the probably approximately correct (PAC) learning theory for partially observable MDPs and Markov games. Key words. Markov decision processes, Markov games, empirical process theory, PAC learning, value function estimation, uniform rate of convergence AMS subject classifications. 90C40, 62C12, 62M05 DOI. 10.1137/040619508

1. Introduction. We address the following question: Given a Markov decision process (MDP) with an unknown policy from a given set of policies, how can we estimate its value function from computer simulations? The question is motivated by the system identification problem in stochastic dynamical systems, such as pursuit-evasion games [38], and the estimation problem in econometric analysis [32]. The question is intimately related to empirical process theory (EPT) [35, 44], which studies the uniform behavior of a class G of measurable functions in the law of large numbers [34] (as well as the central limit theorem [15]) regime. In particular, EPT studies the conditions for which   n 1 Pr sup | (1.1) g(Xi ) − EP [g(X)]| >  → 0 g∈G n i=1 and the rate of convergence. Convergence results in EPT typically use concentration of measure inequalities such as those of Chernoff [11] and Hoeffding [19]. The rate of convergence of an empirical average to the expected value depends on the exponent in the upper bound of such inequalities. Thus, there has been an effort to improve the exponent [23]. Talagrand [40] introduced new concentration of measure inequalities for product probability spaces that are significantly tighter than the Hoeffding–Chernoff type inequalities. The setting of general product probability spaces [27, 41] instead of just independent, identically distributed (i.i.d.) product probability spaces [11, ∗ Received

by the editors November 23, 2004; accepted for publication (in revised form) May 24, 2006; published electronically November 14, 2006. This research supported by National Science Foundation Award ECS-0424445. http://www.siam.org/journals/sicon/45-5/61950.html † EECS Department, University of California, Berkeley, CA 94720 ([email protected], [email protected]). 1633

1634

RAHUL JAIN AND PRAVIN P. VARAIYA

19] has greatly expanded the applications [28, 39]. Moreover, many applications involve dependent processes. The Hoeffding inequality has been extended to various dependent cases [8, 26, 43], while an extension of Talagrand’s inequality for Markov chains and certain mixing processes was provided by [37]. The goal of this paper is to extend the reach of this rich and rapidly developing theory in a new direction. We provide the beginnings of an empirical process theory for MDPs. This essentially involves considering empirical averages of iterates of functions; i.e., if f is a map from R to itself, then we consider g = f t for some fixed integer t, where f t denotes f ◦· · ·◦f , with iteration being done t times. This case is not subsumed in the existing results in the empirical process theory discussed above (see also [13]). Interestingly, we discover that the method used to obtain the sample trajectories of the MDPs from computer simulation affects the rate of convergence. Thus, such a theory fills an important void in the EPT [35, 44] and the stochastic control literatures [36]. It also underlines the importance of choosing a suitable computer simulation. We now make the question above more precise and explain the contribution of this paper. Consider an MDP with a set of policies Π. The value function assigns to each π ∈ Π its expected discounted reward V (π). We estimate V from independent samples of the discounted reward by the empirical mean, Vˆ (π). We obtain the number of samples n(, δ) (or sample complexity) needed so that the probability   ˆ Pr sup |V (π) − V (π)| >  < δ. (1.2) π∈Π

Our approach is broadly inspired by [17, 45, 46] and influenced by [21]. Thus, we would like to reduce the problem in (1.2) to understanding the geometry of Π in terms of its covering number. (If the covering number is finite, it is the minimal number of elements of a set needed to approximate any element in the set Π with a given accuracy.) We first relate the covering numbers of the space of stationary stochastic policies and the space of Markov chains that they induce. We relate these to the space of simulation functions that simulate the Markov chains when the set of transition probabilities of the latter is convex. These results together yield the rate of convergence of the empirical estimate to the expected value for the discounted-reward MDPs. What makes the problem nontrivial is that obtaining an empirical discounted reward from simulation involves an iteration of simulation functions. The geometry of the space of iterated simulation functions is much more complex than that of the original space. One of the key contributions of this paper is the observation that how we simulate an MDP matters for obtaining uniform estimates. We show through an example (see Example 5.4) that uniform estimates of an MDP may converge under one simulation model but fail to do so under another. This is a new (and surprising) observation in the Markov decision theory as well as in the EPT literature. We then consider the average-reward case. This appears to be the first attempt at nonparametric uniform value estimation for the average-reward case when simulation is done with just one sample path. Ergodicity and weak mixing are exploited to obtain uniform rates of convergence of estimates to expected values. We extend the results to dynamic Markov games and to the case when the Markov decision process is partially observable and policies are nonstationary and have memory. The problem of uniform convergence of the empirical average to the value function for discounted MDPs was studied in [20, 30] in a machine learning context. While [20] considered only finite state and action spaces, [30] obtains the conditions for uniform

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1635

convergence in terms of the simulation model rather than the geometric characteristics (such as covering numbers or the P-dimension) of the simulation function space as opposed to that of the more natural policy space. Large-deviations results for finite state and action spaces for the empirical state-action frequencies and general reward functions were obtained in [4, 24]. A different approach more akin to importance sampling is explored in [33]. While the problem of uniform estimation of the value function for discounted and average-reward partially observed MDPs is of interest in itself, it is also connected with the system identification problem [10, 48]. Also interesting and important for many applications are computationally tractable methods (such as through simulation) for approximating the optimal policy [31]. The simulation-based estimates, such as those proposed in this paper, have been used in a gradient-based method for finding Nash equilibrium policies in a pursuit-evasion game problem [38], though the theoretical understanding is far from complete. Other simulation-based methods for finding approximations to the optimal policy include [25], a likelihood-ratio type gradient estimation method for finite state spaces, and [6], which imposes certain differentiability and regularity assumptions on the derivatives of the policies with respect to the parameters. The rest of the paper is organized as follows. Section 2 relates the work to the probably approximately correct (PAC) learning model and the system identification problem. Section 3 is a warm-up section presenting preliminaries and relating the covering numbers of the space of transition probabilities of the induced Markov chains and the policy space. Section 4 presents the estimation methodology using the “simple” simulation model and discusses its combinatorial complexity. Section 5 obtains uniform sample complexity results for estimation of values of discounted-reward MDPs. Section 6 considers average-reward MDPs. Section 7 provides the extension to partially observed MDPs with general policies. Some proofs are relegated to the appendices to maintain a smooth exposition. 2. Relation to PAC learning and system identification. System identification is studied in a general function learning setting in the PAC learning model [17, 42]. A fundamental relationship between system identification and empirical process theory was established by Vapnik and Chervonenkis in [47]. Consider a bounded real-valued measurable function f ∈ F over a set X with a probability measure P . We equip F with a pseudometric such as the one defined in (2.1) below. Unless necessary, we ignore all measurability issues throughout the paper. These have been discussed at length in [15, 34]. The goal is to estimate or “learn” f from independent samples S = {(x1 , f (x1 )), . . . , (xn , f (xn ))}. Say that F is PAC-learnable if there is an algorithm that maps S to hn,f ∈ F such that for any  > 0, the probability that the empirical error  err(f, hn,f ) := |f (x) − hn,f (x)|P (dx) (2.1) is greater than  goes to zero as n → ∞. (Note that hn,f is a function of S.) In other words, for n large enough the probability that the error is larger than  is smaller than some given δ > 0. The class of functions F has the uniform convergence of empirical means (UCEM) property if   n 1     sup  f (Xi ) − EP [f (X)] → 0  f ∈F  n i=1

1636

RAHUL JAIN AND PRAVIN P. VARAIYA

in probability. It is known that a class of bounded real-valued functions with the UCEM property is not only PAC-learnable but PUAC (probably uniformly approximately correct)-learnable [47], i.e.,   lim Pn

n→∞

sup err(f, hn,f ) > 

f ∈F

= 0.

Thus if the mean value of each function in a family can be determined with small error and high probability, the function itself can be “identified” with small error and high probability [45, 46, 47, 49]. One such (minimum empirical risk) algorithm was discovered in [7] when the function class F satisfies a certain (finite covering number) condition. The PAC learning model has been generalized to the case when the inputs are Markovian [1, 16], but it has not been extended to MDPs and games. We provide that extension in this paper. 3. Preliminaries. Consider an MDP M with countable state space X and action space A, transition probability function Pa (x, x ), initial state distribution λ, and a measurable reward function r(x) with values in [0, R]. The value function for a policy π is the expected discounted reward ∞

 V (π) = E γ t r(xt ) , t=1

where 0 < γ < 1 is a discount factor and xt is the state at time t under policy π. For the average-reward case the value function is

T 1 V (π) = E lim inf r(xt ) . T →∞ T t=1 Let Π0 denote the space of all stationary stochastic policies {π(x, a) : a ∈ A, x ∈ X, a π(x, a) = 1} and let Π ⊆ Π0 be the subset of policies of interest. The MDP M under a fixed stationary policy π induces a Markov chain with transition probability function Pπ (x, x ) = a Pa (x, x )π(x, a). The initial distribution on the Markov chains is λ, and we identify Pπ with the Markov chain. Denote P := {Pπ : π ∈ Π}. We seek conditions on the policy space Π such that a simulation-based estimate Vˆ (π) converges to the value function V (π) in probability uniformly over all policies in Π. For this, as we will see in section 5, it is essential to understand the geometry of the space P, and hence of Π. We do this by relating the covering numbers of Π with that of P, which are then related to a space of (simulation) functions F that we define in section 4. Let X be an arbitrary set and let λ be a probability measure on X. Given a set F of real-valued functions on X, ρ a metric on R, let dρ(λ) be the pseudometric on F with respect to measure λ,  dρ(λ) (f, g) = ρ(f (x), g(x))λ(dx). A subset G ⊆ F is an -net for F if ∀f ∈ F, ∃g ∈ G with dρ(λ) (f, g) < . The size of the minimal -net is the -covering number, denoted N (, F, dρ(λ) ). The -capacity of F under the ρ metric is C(, F, ρ) = supλ N (, F, dρ(λ) ). Essentially, the -net can be

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1637

seen as a subset of functions that can -approximate any function in F. The covering number is a measure of the richness of the function class. The richer it is, the more approximating functions we will need for a given measure of approximation . The capacity makes it independent of the underlying measure λ on X. (See [21] for an elegant treatment of covering numbers.) Let σ be a probability measure on A. We now define the following L1 -pseudometric on Π:   dL1 (σ×λ) (π, π  ) := σ(a) λ(x)|π(x, a) − π  (x, a)|, a∈A

x∈X

and the total variation pseudometric on P,         λ(x)(P (x, y) − P (x, y)) . dT V (λ) (P, P ) :=    y∈X x∈X

Note that covering numbers of function spaces can be defined for pseudometrics, and a metric structure is not necessary (see [17, 21, 49]). Bounds on covering numbers are obtained in terms of various combinatorial dimensions.1 Thus, we first relate the covering number of P with a combinatorial dimension of Π. Recall some measures of combinatorial dimension. Let F be a set of binary-valued functions from X to {0, 1}. Say that F shatters {x1 , . . . , xn } if the set {(f (x1 ), . . . , f (xn )), f ∈ F} has cardinality 2n . The largest such n is the VC-dim(F). Intuitively, this means that the function class F can distinguish between a set of n points from the set X. Let F be a set of real-valued functions from X to [0, 1]. Say that F P-shatters {x1 , . . . , xn } if there exists a witness vector c = (c1 , . . . , cn ) such that the set {(η(f (x1 )− c1 ), . . . , η(f (xn )−cn )), f ∈ F} has cardinality 2n ; η(·) is the sign function. The largest such n is the P-dim(F). This is a generalization of VC-dim, and for {0, 1}-valued functions, the two definitions are equivalent. Other combinatorial dimensions such as the fat-shattering dimension introduced in [2] yield both an upper and lower bound on the covering numbers but in this paper we will use the P-dim. Results using fat-shattering dimension can be established similarly. Given a policy space Π, let P denote the set of transition probabilities of the Markov chains it induces. We relate the covering numbers of the two spaces under the pseudometrics defined above. Lemma 3.1. Suppose P-dim(Π) = d. Suppose there is a probability measure λ on X and a probability measure σ on A such that π(x, a)/σ(a) ≤ K ∀x ∈ X, a ∈ A, π ∈ Π. Then, for 0 <  < e/4 log2 e,

d

2eK Π N (, P, dT V (λ) ) ≤ N , , dL1 (σ×λ) ≤ 2 Φ , σ  where Φ(x) = x log x. The proof is in the appendix. We now give an example illustrating the intuition behind the concept of P-dim. It is similar in spirit to well-known results about finite-dimensional linear spaces. It 1 These are different from the algebraic dimension. Examples of combinatorial dimensions are VCdimension and P-dimension. A class of real-valued functions may have infinite algebraic dimension but finite P-dimension. See [3, 49] for more details.

1638

RAHUL JAIN AND PRAVIN P. VARAIYA

shows that for an MDP with finite state and action spaces, the set of all stationary stochastic policies has a finite P-dimension equal to the number of free parameters needed to represent the set of policies being considered. Example 3.2. Let X = {1, . . . , N } and A = {1, . . . , M }. Then, P-dim (Π0 ) = N (M − 1) for M > 1. Proof. Consider the set of all stochastic policies:    π(x, a) = 1 ∀x ∈ X . Π0 = π : X × A → [0, 1] | a∈A

Let S = {(x1 , a1 ), . . . , (xN , a1 ), . . . , (x1 , aM −1 ), . . . , (xN , aM −1 )}. We can find c11 , . . . , cN M −1 such that the N (M − 1)-dimensional vector ( η(π(x1 , a1 ) − c11 ), . . . , η(π(x1 , aM −1 ) − c1M −1 ), . . . .. .. . . . . . , η(π(xN , a1 ) − cN 1 ), . . . , η(π(xN , aM −1 ) − cN M −1 ) ) yields all possible binary vectors as π runs over Π; η(·) is the sign function. Consider the first row. Note that the probabilities there together with π(x1 , aM ) sum to 1. Choose all c1j to be 1/M . Then, we can get all possible binary vectors in the first row. Since the subsequent rows are independent, we can do the same for all of them. Thus, we can get all possible binary vectors of length N (M − 1). So Π0 shatters S. However, if we add another point, say (x1 , aM ), to S, the first row will sum to 1. In this case we cannot get all the 2M possible binary vectors. Thus, the P-dimension of Π0 is N (M − 1). 4. The simulation model. We estimate the value V (π) of policy π ∈ Π from independent samples of the discounted rewards. The samples are generated by a simulation “engine” h. This is a deterministic function to which we feed a “noise” sequence ω = (ω1 , ω2 , . . .) (with ωi being i.i.d. from uniform distribution over Ω = [0, 1]) and an initial state x0 (drawn from distribution λ). The engine h generates a sample trajectory with the same distribution as the Markov chain corresponding to π. The function h : X × A × Ω → X gives the next state x given the current state x, action taken a, and noise ωi . Several such sample trajectories are generated using i.i.d. noise sequences and initial states. Each sample trajectory yields an empirical total discounted reward. The estimate of V (π), Vˆ (π) is the average of the empirical total discounted reward for the various sample trajectories. Because simulation cannot be performed indefinitely, we stop the simulation at some time T , after which the contribution to the total discounted reward falls below /2 for required estimation error bound . T is the /2-horizon time. Many simulation functions are possible. We will work with the following simple simulation model. For the rest of this paper, we consider the state space to be X = N. Definition 4.1 (simple simulation model). The simple simulation model h for a given MDP is given by h(x, a, ω) = inf{y ∈ X : ω ∈ [Fa,x (y − 1), Fa,x (y))}, in which Fa,x (y) := y ≤y Pa (x, y  ) is the cummulative distribution function (c.d.f.) corresponding to the transition probability function Pa (x, y). Similarly, with a slight abuse of notation, we define the simple simulation model h for the Markov chain P as h(x, P, ω) = inf{y ∈ X : ω ∈ [FP,x (y − 1), FP,x (y))},

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1639

where FP,x (y) := y ≤y P (x, y  ) is the c.d.f. corresponding to the transition probability function P (x, y). This is the simplest method of simulation. For example, to simulate a probability distribution on a discrete state space, we partition the unit interval so that the first subinterval has length equal to the mass on the first state, the second subinterval has length equal to the mass on the second state, and so on. Perhaps surprisingly, there are other simulation functions h that generate the same Markov chain, but which have a much larger complexity than h. The sample trajectory {xt } for policy π is obtained by xt+1 = fPπ (xt , ωt+1 ) = h(xt , Pπ , ωt+1 ), in which Pπ is the transition probability function of the Markov chain induced by π and ωt+1 ∈ Ω is noise. The initial state x0 is drawn according to the given initial state distribution λ. The function fPπ : X × Ω → X is called the simulation function for the Markov chain transition probability function Pπ . (The reader may note that the above definition is to ease understanding in this section. In the next section, we will redefine the domain and range of the simulation functions.) As before, P = {Pπ : π ∈ Π}. We denote by F = {fP : P ∈ P} the set of all simulation functions induced by P. To every P ∈ P, there corresponds a function f ∈ F. Observe that f ∈ F simulates P ∈ P given by P (x, y) = μ0 {ω : f (x, ω) = y}, where μ0 is the Lebesgue measure on Ω. Unless specified otherwise, F will denote the set of simulation functions for the class P under the simple simulation model. In the previous section, we related the covering numbers of policy space Π and P. However, as we shall see in the next section, the convergence properties of our estimate of the value function really depend on the covering number of F. Thus, we now show that the complexity of the space F is the same as that of P if P is convex. The result is in the same spirit as Theorem 13.9 in [12] for finite-dimensional linear vector spaces. However, the setting here is different. We provide an independent proof. Lemma 4.2. Suppose P is convex (being generated by a convex space of policies) with P-dimension d. Let F be the corresponding space of simple simulation functions induced by P. Then, P-dim (F) = d. Moreover, the algebraic dimension of P is also d. Proof. There is a one-to-one map between the space of simple simulation functions F and the space of c.d.f.’s F˜ corresponding to P. (F˜ = {F˜ : F˜ (x, y) =  ˜ y  ≤y P (x, y ), P ∈ P}.) F and F have the same P-dimension because, for any ˜ F˜ (x, y) < ω. Thus, F ∈ F, F (x, ω) > y if and only if for the corresponding F˜ ∈ F, F˜ shatters {(x1 , y1 ), . . . , (xd , yd )} with witness vector (ω1 , . . . , ωd ) if and only if F shatters {(x1 , ω1 ), . . . , (xd , ωd )} with witness vector (y1 , . . . , yd ). So in the following discussion we treat them as the same space F. Because P has P-dimension d, there exists S = {(x1 , y1 ), . . . , (xd , yd )} that is shattered by P with some witness vector c = (c1 , . . . , cd ). Consider the projection of the set P on the S coordinates: P|S = {(P (x1 , y1 ), . . . , P (xd , yd )) : P ∈ P}. The definition of shattering implies that there is a d-dimensional hypercube contained in P|S with center c. Also note that P|S is convex and its algebraic dimension is d. To argue that the algebraic dimension of P cannot be d + 1, suppose that it is. Then it would contain d + 1 coordinates such that the projection of P along those coordinates contains a hypercube of dimension d + 1. Thus, P would shatter d + 1 points with the

1640

RAHUL JAIN AND PRAVIN P. VARAIYA

center of the hypercube being a witness vector. But that contradicts the assumption that the P-dimension of P is d. Thus for convex spaces, the algebraic dimension and P-dimension are equal. Next, F is obtained from P by an invertible linear transformation; hence its algebraic dimension is also d. Thus, it has d coordinates S such that the projected space F|S has algebraic dimension d. Moreover, it contains a hypercube of dimension d. Hence, its P-dimension is at least d. Since the argument is reversible starting from space F to space P, it implies P-dim (P) = P-dim (F). It may be noted that convexity is essential to the above argument. From several examples it appears that the result is not true without convexity. But we are unable to offer a concrete counterexample. 5. Discounted-reward MDPs. We now consider uniform value function estimation from simulation for discounted-reward MDPs. For the rest of the paper, we redefine F to be a set of measurable functions from Y := X × Ω∞ onto itself which simulate P, the transition probabilities induced by Π under the simple simulation model. However, each function depends only on the first component of the sequence ω = (ω1 , ω2 , . . .). Thus the results and discussion of the previous section hold. Let θ be the left-shift operator on Ω∞ , θ(ω1 , ω2 , . . .) = (ω2 , ω3 , . . .). For a policy π, our simulation system is (xt+1 , θω) = fPπ (xt , ω), in which xt+1 is the next state starting from xt and the simulator also outputs the shifted noise sequence θω. This definition of the simulation function is introduced to facilitate the iteration of simulation functions. Denote F := {fP : X × Ω∞ → X × Ω∞ , P ∈ P} and F 2 := {f ◦ f : Y → Y, f ∈ F} and F t its generalization to t iterations. Similarly, we redefine the reward function as r : X × Ω∞ → [0, R]. The estimation procedure is the following. Obtain n initial states x10 , . . . , xn0 drawn as i.i.d. according to λ, and n noise sequences ω 1 , . . . , ω n ∈ Ω∞ drawn according to μ. Denote the samples by S = {(x10 , ω 1 ), . . . , (xn0 , ω n )}. Under the simple simulation model, fP (x, ω) := (h(x, P, ω1 ), θω) and as before, F := {fP : P ∈ P}. For a given initial state and noise sequence, the simulation function yields a reward sequence, the reward at time t given by Rt (x0 , ω) := r ◦ fP ◦ · · · ◦ fP (x0 , ω), with fP composed t times. The empirical total discounted reward for a given state sequence then is ∞ t of V (π) from n simulations, each conducted for /2t=0 γ Rt (x0 , ω). Our estimate n T horizon time T , is Vˆn (π) := n1 i=1 [ t=0 γ t Rtπ (xi0 , ω i )]. We first present a key technical result which relates the covering number of the iterated functions F t under the ρ pseudometric with the covering number for F under the L1 pseudometric, for which bounds are known in terms of the P-dim of F. Let μ be any probability measure on Ω∞ and λ the initial distribution on X. Denote the product measure on Y by P = λ × μ, and on Y n by Pn . Define two pseudometrics on F,  λ(x)μ{ω : f (x, ω) = g(x, ω)} ρP (f, g) = x

and dL1 (P) (f, g) :=

 x

 λ(x)

|f (x, ω) − g(x, ω)|dμ(ω).

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1641

Here, we take |f (x, ω)−g(x, ω)| to denote |x −x |+||θω−θω||, where f (x, ω) = (x , θω) and g(x, ω) = (x , θω), and || · || is the l1 -norm on Ω∞ . Recall that x , x ∈ N. Then, |f (x, ω) − g(x, ω)| = |x − x |, the usual l1 distance. Lemma 5.1. Let λ be the initial distribution on X and let λf be the (one-step) distribution given by λf (y) = x λ(x)μ{ω : f (x, ω) = (y, θω)} for f ∈ F. Suppose that   λf (y) K := max sup (5.1) , 1 < ∞. f ∈F ,y∈X λ(y) Then, ρP (f t , g t ) ≤ K t dL1 (P) (f, g) and N (, F t , ρP ) ≤ N (/K t , F, dL1 (P) ). The proof is in the appendix. The condition of the lemma essentially means that under distribution λ the change in the probability mass on any state under any policy after one transition is bounded. It should be noted that for simulation, we can choose the initial state distribution and it should be such that λ(y) > 0 ∀ y. Further, if λ(y) = 0, the Markov chains are such that we must have λf (y) = 0 as well, i.e., λf λ. A particular case where this is satisfied is a set of positive recurrent Markov chains, say with the same invariant distribution π. If we choose λ = π, then λf = π and the condition is trivially satisfied. We now show that the estimate converges to the expected discounted reward uniformly over all policies in Π, and we also obtain the uniform rate of convergence. Theorem 5.2. Let (X, Γ, λ) be a measurable state space. Let A be the action space and r the [0, R]-valued reward function. Let Π ⊆ Π0 be the space of stationary stochastic policies, P the space of Markov chain transition probabilities induced by Π, and F the space of simulation functions of P under the simple simulation model h. Suppose that P-dim(F) ≤ d and the initial state distribution λ is such that K := λf (x) max{supf ∈F ,x∈X λ(x) , 1} is finite. Let Vˆn (π) be the estimate of V (π) obtained by averaging the reward from n samples. Then, given any , δ > 0,   Pn sup |Vˆn (π) − V (π)| >  < δ π∈Π

for (5.2)

32R2 n≥ α2





4 32eR log + 2d log + T log K . δ α

T is the /2-horizon time and α = /2(T + 1). Proof. Fix a policy π. Let P be the induced Markov chain transition probability function simulated by the simple simulation function fP . Let Rt (x0 , ω) := r ◦fP ◦· · ·◦ fP (x0 , ω), with fP composed t times, be the reward at time t, and denote Rt := {Rt : X × Ω∞ → [0, R], P ∈ P}. Let V T (π) be the expected discounted reward truncated n T up to T steps and Vˆn (π) = n1 i=1 [ t=0 γ t Rtπ (xi0 , ω i )] its estimate from n finite time

1642

RAHUL JAIN AND PRAVIN P. VARAIYA

simulations. Then, |V (π) − Vˆn (π)| ≤ |V (π) − V T (π)| + |V T (π) − Vˆn (π)|,  ≤ |V T (π) − VˆnT (π)| + , 2   T   n    1 π i i π  ≤ [Rt (x0 , ω ) − E(Rt )] + .  n  2 t=0 i=1 Here, the expectation is with respect to the product measure Pπt × λ × μ. We show that with high probability, each term in the sum over t is bounded by α = /2(T + 1). Note that   |r(f t (x, ω)) − r(g t (x, ω))|dμ(ω)dλ(x) ≤ R · λ(x)μ{ω : f t (x, ω) = g t (x, ω)}, x

which as in Lemma 5.1 implies that dL1 (P) (r ◦ f t , r ◦ g t ) ≤ R · ρP (f t , g t ) ≤ R · K T dL1 (P) (f, g). Applying Theorem 3 from [17] with the “α” in the statement of that theorem set equal to /4R and ν = 2R, and using Lemma 5.1 and the inequality above, we get     n 1     n i i P sup  Rt (x0 , ω ) − E(Rt ) > α  Rt ∈Rt  n i=1

   α 2 (− nα 2 ) 32R , Rt |S , dl1 e ,1 ≤ 2E min 2N 16

 α nα2 , Rt , dL1 (P) exp − ≤ 4 sup N 16 32R2 P

α  2 nα ≤ 4C , Rt , dL1 exp − 16 32R2

d

32eRK T nα2 32eRK T log . exp − ≤4 α α 32R2 This implies that the estimation error is bounded by α, with probability at least δ, if the number of samples is

4 32R2 32eR log + 2d log n≥ + T log K . α2 δ α Remarks. 1. Theorem 5.2 implies that supπ∈Π |Vˆn (π) − V (π)| converges to zero in probability; hence the policy space Π is PAC-learnable. As in [30], the theorem assumes that the P-dimension of the F space is finite. Combined with Lemma 4.2, this gives the following corollary. Corollary 5.3. Under assumption (5.1) and if P is convex with P-dim(P) = d, result (5.2) of Theorem 5.2 holds. 2. Our sample complexity is of the same order, in terms of δ, , T, R, and d, as the results of [30], but the two results are not directly comparable due to the different assumptions made. In fact, the major challenge in obtaining the uniform rate of convergence is relating the covering numbers and P-dimensions of the policy space Π

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1643

and P with the space F. This is what we accomplished in this paper and what is missing in [30]. Also, unlike in [30], we do not require the simulation functions to be Lipschitz continuous. For a discrete state space, this is not a realistic assumption, as the following examples show. (a) Consider the Markov chain on N such that the only transitions are from state 1 to state 2 with probability 1/2, to state 4 with probability 1/4, . . . , to state k+l−1 −i 2k with probability 1/2k , etc. Let ωlk = 2 and kl = 2−(k+l) and define i=1 k k k k k k k k ω ˆ = (ω1 − 1 /2, ω2 − 2 /2, . . .) and ω ˇ = (ω1 + 1 /2, ω2k + k2 /2, . . .). Thus, with noise k ˆ = 2k , while with noise sequence sequence ω ˆ , the transition from state 1 is to state x k k+1 ˇ=2 ω ˇ , the transition from state 1 is to state x . Define the metric ρ on X × Ω∞ , k k ˆ ), (x2 , ω ˇ )) = |x1 − x2 | + |ˆ ρ((x1 , ω ω1 − ω ˇ 1 |. Then, it can be verified that ρ(f (1, ω ˆ k ), f (1, ω ˇ k )) = |ˆ x−x ˇ| + k2 ≥ 2k ∀k, whereas Lipschitz continuity would require that it be less that C2−(k+1) for some positive constant C and every k. Thus, f is not Lipschitz continuous on Ω∞ . (b) Consider the following Markov chain: state space X = N again endowed with the same metric ρ as in the above example. Transitions are deterministic: transition from an even state n is to state 2n, and from an odd state n + 1 is to 3n. Then, ρ(f (n + 1, ω), f (n, ω)) = n and so is not Lipschitz continuous on X. These examples demonstrate that Lipschitz continuity of the simulation functions on a discrete state space is not the right assumption to make. 3. Markov games. The generalization of the results of this section to discountedreward Markov games is relatively straightforward. It is of considerable interest for many applications [38, 32]. Consider two players playing a Markov game with action spaces A1 and A2 , state space X, and transition function Pa1 ,a2 (x, y), a1 ∈ A1 , a2 ∈ A2 . We consider only stationary policy spaces Π1 and Π2 . The two reward functions r1 and r2 depend only on the state and have values in [0, R]. Denote the discountedreward functions by V1 (π1 , π2 ) and V2 (π1 , π2 ) with discount factor 0 < γ < 1. Denote the set of Markov chains induced by the policies in Π1 ×Π2 by P = {Pπ1 ,π2 : (π1 , π2 ) ∈ Π1 × Π2 )} with Pπ1 ,π2 (x, y) = a1 ,a2 Pa1 ,a2 (x, y)π1 (x, a1 )π2 (x, a2 ). Of interest is a uniform sample complexity bound such that the error in estimating both V1 (π1 , π2 ) and V2 (π1 , π2 ) is within  with probability at least 1 − δ. This is now easily obtained by using Theorem 5.2 and bounding the maximum of the estimation error in V1 and V2 . Such estimates may be used to compute the Nash equilibria [38]. A bound on the -covering number of P can be easily obtained by constructing /2-covers for Π1 and Π2 under the L1 metric, and then the  cover for Π1 × Π2 is obtained by taking the product of the two sets. The rest of the argument is the same as in Lemma 3.1. 4. We now show through two examples that the choice of simulation model to use for generating the sample trajectories affects whether we get uniform convergence or not. Example 5.4. (i) Consider the three-state Markov chain in Figure 5.1 with the following transition matrix: ⎡ ⎤ 0 p 1−p 0 ⎦ P =⎣ 0 1 0 0 1 and with state 0 always the initial state. Let P = {P as defined above with 0 ≤ p ≤ 1} denote the set of all such Markov chains as p varies. It is easy to check that P is

1644

RAHUL JAIN AND PRAVIN P. VARAIYA State 0

p

1-p

1

1 State -1

State 1

Fig. 5.1. A three-state Markov chain with initial state 0.

convex and has P-dimension 1. Let the reward in state 0 be 0, in state −1 be −1, and in state 1 be 1. This defines a Markov reward process, and for a given p we get the value Vp = 1 − 2p. Let F be the set of simple simulation functions that simulate the Markov chains in P. By Lemma 4.2, P-dim(F) is 1 as well. Now, consider the following set of simulations functions. Fix p ∈ [0, 1]. Let x1 , x2 , . . . be rational numbers in [0, 1] such that (5.3)

0 ≤ x1 ≤ x1 + p/2 ≤ x2 ≤ x2 + p/4 ≤ · · · ≤ 1.

Generate a uniform [0, 1] random number ω. If it falls in [x1 , x1 +p/2]∪[x2 , x2 +p/4]∪ · · · , then the next state is −1; otherwise it is 1. This simulation function simulates a Markov chain with parameter p. Let F˜ denote the set of all such simulation functions as p varies in [0, 1] and for every possible sequence of rational numbers satisfying (5.3). ˜ is not finite. Now, consider any ω1 , ω2 , . . . , ωn ∈ It is easy to check that P-dim(F) [0, 1]. Then, there exists a p and x1 , x2 , . . . such that ωk ∈ [xk , xk + p/2k ], which means there is a simulation function which always goes to state −1 for this sequence of random numbers. Since n is arbitrary, this implies Vˆp = −1 = Vp for some p. Thus, in this case, even though the second set of simulation functions F˜ simulates the same set of Markov chains P, we do not get uniform convergence. In the example above, there are many simulation functions in F that simulate the same Markov chain in P. We now modify the example so that there is only one simulation function for each Markov chain. (ii) Fix a positive integer m. Consider the [0, 1] interval divided into m2 subin 2 tervals of length 1/m2 . Pick any m of these subintervals, which can be done in m m ways. Arrange these choices in any order. Define 2 1 k m 1 ≤ , + m2  k = 1, . . . , . pm,k = m+1 m m m(m + 1) m

Scale the length of each of the m subintervals of the kth such choice to pm,k /m ≤ 1/m2 , keeping the beginning point of the subintervals the same. Now, for every pm,k  2 as defined above with k = 1, . . . , m m , and m = 1, 2, . . . , define a simulation function fm,k such that for the Markov chain with parameter pm,k , if the current state is zero, and the uniform [0, 1] random number falls in any of the subintervals, then the next state is −1; otherwise it is 1. All other Markov chains are simulated according to the simple simulation model. Thus, we get a new set of simulation functions F¯ with each Markov chain being simulated by only one simulation function. It is easy to check ¯ is not finite. Further, given any ω1 , ω2 , . . . , ωn ∈ [0, 1], we can always that P-dim(F)

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1645

pick an m large enough such that there is a km so that all of the ωi lie outside the m subintervals, each of length pm,km /m. Thus, the simulation goes to state −1 for each of the ωi . Thus, Vˆpm,km = Vpm,km and we do not get uniform convergence. These examples demonstrate that uniform convergence for MDPs depends not only on the geometry of the policy space but also on how the MDP is simulated to obtain the sample trajectories. The simple simulation model is in some sense the best model since it has the same complexity under convexity as the set of Markov chains it simulates (Lemma 4.2). However, these examples do not satisfy the conditions provided in Theorem 5.2. Note that those are sufficient conditions only. An alternative simulation method for finite-dimensional convex spaces. An important special case is when the policy space is the convex hull of a finite number of policies, i.e., when all policies are (random) mixtures of finitely many policies. While the previous simulation method would still work, we present an alternative simulation method that exploits the convex structure. The simulation method is the following. Suppose there are two policies πi for i = 0, 1, each inducing a Markov chain Pi with simulation function fi . Consider the mixture πw = wπ0 + (1 − w)π1 , so Pw = wP0 + (1 − w)P1 . For t steps, we have Pwt = wt P0t + wt−1 (1 − w)P0t−1 P1 + wt−1 (1 − w)P0t−2 P1 P0 + · · · + (1 − w)t P1t . Note that P0 and P1 need not commute. To obtain Vˆn (π), we first determine the rewards at time t for t = 0, . . . , T . To estimate the rewards, first draw the initial states x10 , . . . , xn0 from λ, and the noise sequences ω 1 , . . . , ω n from μ. Then carry out 2t simulations, one for each term in the sum of the equation above. For example, if it · · · i1 is the binary representation of k, then the contribution to reward from the kth term is determined by r ◦ fit ◦ · · · ◦ fi1 (xi0 , ω i ). The estimate of the contribution to ˆ k , is the mean over the n initial state and noise sequence reward from the kth term, R t pairs. The estimate of the reward at time t is ˆ t = wt R ˆ 1 + · · · + (1 − w)t R ˆ t0 + wt−1 (1 − w)R ˆ t2t −1 R t T ˆ t . This can be generalized to a and the value function estimate Vˆn (π) is t=0 γ t R policy space which is a convex hull of any finite number of policies. Theorem 5.5. Let Π = conv{π0 , . . . , πd−1 }, P = conv{P0 , . . . , Pd−1 } be the space of Markov chains induced by Π, and F be the space of simulation functions of P under the simple simulation model. Let Vˆn (π) be the estimate of V (π) obtained from n samples. Then, given any , δ > 0   n ˆ P sup |Vn (π) − V (π)| >  < δ π∈Π 2





R for n ≥ 2α log 1δ + T log 2d , where T is the /2-horizon time and α = /2(T + 1). 2 d−1 Proof. Consider any π ∈ Π andt the corresponding P ∈ P. Let P = i=0 ai Pi with ai ≥ 0 and i ai = 1. Thus, P can be written as

(5.4)

t

P =

t d −1

wkt Pit · · · Pi1 ,

k=0

where it · · · i1 is a d-ary representation of k and the nonnegative weights wkt are such dt −1 that k=0 wkt = 1 and can be determined easily.

1646

RAHUL JAIN AND PRAVIN P. VARAIYA

To obtain Vˆn (π), we need to simulate P for T steps as before. The t-step reward is determined by the state at time t, whose distribution is given by νt = P t λ. To estimate the reward, first draw the initial states x10 , . . . , xn0 from λ, and the noise sequence ω 1 , . . . , ω n from μ. We carry out dt simulations, one for each term in the sum of (5.4). Recall that it · · · i1 is a d-ary representation of k. Thus, to determine the contribution to the empirical reward at time t due to the kth term in (5.4), the state at time t is determined by r ◦ fit ◦ · · · ◦ fi1 (xi0 , ω i ). Thus, an estimate of the expected reward at time t is  n  t d −1 1 t t i i ˆ wk (5.5) r ◦ fit ◦ · · · ◦ fi1 (x0 , ω ) , Rt (w ) = n i=1 k=0

T ˆ t (wt (π)), where a consistent estimator of EP t λ [r(x)]. Vˆn (π) is now given by t=0 γ t R t w (π) are the weights determined in 5.4 by policy π at time t. Denote Wt := {wt (π) : π ∈ Π}. Note that E

(5.6)

P tλ

[r(x)] =

t d −1

  wkt EPit ···Pi1 λ [r(x)] ,

k=0

ˆ k ) is an estimator for the where the quantity in the brackets of (5.5) (denote it by R t quantity in the brackets of (5.6). Thus,     n n t ˆ ˆ t P sup |Vn (π) − V (π)| >  ≤ P sup |Rt (w ) − EP λ [r(x)]| > α π∈Π

wt ∈Wt

ˆ k − EP ···P λ [r(x)]| > α} ≤ dT max Pn {|R t it i1 k

2nα2 , ≤ 2dT exp − 2 R where α = /2(T + 1) and the last inequality follows from Hoeffding’s inequality. From the above, the sample complexity can be obtained as in the proof of Theorem 5.2. Remarks. This alternative simulation model has the same order of sample complexity as in the earlier case, but it has greater computational complexity since, for each chain, dT simulations need to be carried out. However, if several Markov chains need to be simulated, as when performing an optimal policy search, the simulation is carried out only once for all the chains because the estimates for various mixtures are ˆk . obtained by appropriately weighing the dt estimates, R t Also, to obtain the estimates for t ≤ T , one need not repeat the simulation for each t. Instead, the T -step simulation suffices to yield the t-step estimate by simply ignoring the simulation beyond t + 1. Thus, only dT simulations are needed to obtain an estimate for a reward for any step t ≤ T and any P ∈ P. 6. Average-reward MDPs. Some MDP problems use the average-reward criterion. However, there are no published results on simulation-based uniform estimates of value functions of average-reward MDPs. We present such a result in this section. Unlike the discounted-reward case, where we simulated several different sample paths, starting with different initial states, here the estimate is obtained from only one sample path. This is possible only when the policies we consider are such that

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1647

the Markov chains are stationary, ergodic, and weakly mixing. (The conditions under which a policy induces a stationary, ergodic Markov chain can be found in Chapter 5 of [9].) A related problem is addressed in [22], which uses Csiszer’s concentration inequality to bound reward functionals for Doeblin chains. However, the bound is not uniform and distribution dependent. Let λπ denote the invariant measure of the Markov chain {Xk }∞ k=0 with transition probability function Pπ , and Λ be the set of all such invariant measures. Let P denote the probability measure for the process. We assume that there is a Markov chain P0 with invariant measure (steady state distribution) λ0 such that λπ λ0 , i.e., λπ is absolutely continuous with respect to λ0 , meaning that λπ (A) = 0 if λ0 (A) = 0 for any measurable set A [18]. We call such a chain a reference Markov chain. Let π (x) := λπ (x)/λ0 (x) be the Radon–Nikodym derivative and assume that it is uniformly bounded, π (x) ≤ K ∀ x and π. Let H be the set of all such Radon– Nikodym derivatives. Our simulation methodology is to generate a sequence {xk }nk=1 according to P0 . We then multiply r(xt ) by π (xt ) to obtain the tth-step reward for the Markov chain induced by policy π. The estimate of the value function is then obtained by taking an empirical average of the rewards, i.e., 1 r˜π (xt ), Vˆn (π) = n t=1 n

in which r˜π (x) := rπ (x)π (x). Let R = {˜ rπ : π ∈ Π}. Furthermore, in some problems, such as when the state space is multidimensional or complex, it may not be possible to integrate the reward function with respect to the stationary measure. In such cases, Monte Carlo type methods such as importance sampling are useful for estimating integrals. The method proposed here falls into such a category, and we present a uniform rate of convergence result for it. This approach is useful when it is difficult to sample from the stationary distributions of the Markov chains but easy to compute the derivative π (x). We first state some definitions and results needed in the proof below. Let ∞ ∞ {Xn }∞ n=−∞ be a process on the measurable space (X−∞ , S−∞ ). Consider a stationary,  ergodic process with measure P. Let P be the one-dimensional marginal, and ¯ = ∞ P , the product measure under which the process is i.i.d. P −∞ Let P0−∞ and P∞ 1 be the semi-infinite marginals of P. Define the β-mixing coefficients [14] as 0 ∞ β(k) := sup{|P(A) − P0−∞ × P∞ 1 (A)| : A ∈ σ(X−∞ , Xk )}.

If β(k) → 0 as k → ∞, the process is said to be β-mixing (or weakly mixing). From the definition of the β-mixing coefficients we get the following. ¯ Fact 1 (see [29, 50]). If A ∈ σ(X0 , Xk , . . . , X(m−1)k ), |P(A) − P(A)| ≤ mβ(k). We assume that the Markov chain P0 is β-mixing with mixing coefficients β0 . We also need the following generalization of Hoeffding’s bound. Fact 2 (McDiarmid–Azuma inequality [5, 27]). Let X1 , . . . , Xn be i.i.d. drawn according to P and g : X n → R. Suppose g has bounded differences, i.e., |g(x1 , . . . , xi , . . . , xn ) − g(x1 , . . . , xi , . . . , xn )| ≤ ci . Then, ∀τ > 0,



2τ 2 P {g(X ) − Eg(X ) ≥ τ } ≤ exp − n n

n

n

2 i=1 ci

.

1648

RAHUL JAIN AND PRAVIN P. VARAIYA

We now show that the estimation procedure enunciated above for the averagereward case produces estimates Vˆ (π) that converge uniformly over all policies to the true value V (π). Moreover, we can obtain the rate of convergence, whose explicit form depends on the specific problem. Theorem 6.1. Suppose the Markov chains induced by π ∈ Π are stationary and ergodic. Assume there exists a Markov chain P0 with invariant measure λ0 and mixing coefficient β0 such that λπ λ0 and the π are bounded by a constant K with P-dim(H) ≤ d. Denote by Vˆn (π) the estimate of V (π) from n samples. Then, given any , δ > 0,   P0 sup |Vˆn (π) − V (π)| >  < δ, π∈Π 2

1 1/2 for n large enough such that γ(m)+Rmβ0 (k)+τ (n) ≤ , where τ (n) := ( 2R n log δ ) and 

d

 32eRK 32eRK mα2 log γ(m) := inf α + 8R , exp − α>0 α α 32R2

with n = mn kn such that mn → ∞ and kn → ∞ as n → ∞. Proof. The idea of the proof is to use the β-mixing property and reduce the problem to one with i.i.d. samples. We then use techniques similar to those used in the proof of Theorem 5.2. The problem of iteration of simulation functions does not occur in this case, which makes the proof easier. Let x1 , x2 , . . . , xn be the state sequence generated according to P0 which can be done using the simple simulation model. Note that E0 [r(x)π (x)] = Eλπ [r(x)], in which Eλπ and E0 denote expectation taken with respect to the stationary measures n ˆ n [˜ rπ ; xn ] := n1 t=1 r˜π (xt ) and observe that λπ and λ0 , respectively. Denote E  

 (1) k−1  1 m−1    1   ˆ n [˜ E0 sup |E rπ ; xn ] − E0 [˜ rπ (x)]| ≤ E0 sup  r˜π (xlk+j ) − E0 [˜ rπ (x)]   k j=0 π∈Π π∈Π m l=0      1 m−1  (2)    ≤ α + RP0 sup  r˜π (xlk ) − E0 [˜ rπ (x)] > α  π∈Π  m l=0

(3)

≤ α + Rmβ0 (k)      1 m−1     ¯ +RP0 sup  r˜π (xlk ) − E0 [˜ rπ (x)] > α ,  π∈Π  m l=0

¯ 0 is the i.i.d. product measure in which E0 is expectation with respect to P0 , and P corresponding to P0 . Inequality (1) follows by triangle inequality, (2) by stationarity and the fact that the reward function is bounded by R, and (3) by definition of the β-mixing coefficients. Claim 1. Suppose π (x) ≤ K for all x and π and that P-dim(H) ≤ d. Then, ∀α > 0

d 2eRK 2eRK log C(α, R, dL1 ) ≤ 2 . α α

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1649

Proof. Let λ0 be a probability measure on X. Observe that for r˜1 , r˜2 ∈ R,  r1 , r˜2 ) = |˜ r1 (x) − r˜2 (x)|λ0 (x) dL1 (λ0 ) (˜ x

≤R·



|1 (x) − 2 (x)|λ0 (x)

x

= R · dL1 (λ0 ) (1 , 2 ). As argued for similar results earlier in the paper, this implies the desired conclusion. From Theorem 5.7 in [49], we then get    

 1 m−1   α mα2    ¯ , R, dL1 exp − P0 sup  r˜π (xlk ) − E0 [˜ rπ (x)] > α ≤ 4C  16 32R2 π∈Π  m l=0

d

32eRK 32eRK mα2 log ≤8 . exp − α α 32R2 Substituting above, we get   ˆ n [˜ E0 sup |E rπ ; xn ] − E0 [˜ rπ (x)]| ≤ γ(m) + Rmβ0 (k). π∈Π

Now, defining g(xn ) as the argument of E0 above and using the McDiarmid– Azuma inequality with ci = R/n, we obtain that   n ˆ P0 sup |En [˜ rπ ; x ] − E0 [˜ rπ (x)]| ≥ γ(m) + Rmβ0 (k) + τ (n) < δ, π∈Π

where δ = exp(−2nτ 2 (n)/R2 ), and hence we get the desired result. Note that by assumption of the mixing property, β0 (kn ) → 0, and for fixed  and δ, γ(mn ), τ (n) → 0 as n → ∞. The sample complexity result is implicit here since given functions γ, β0 , and τ , we can determine n, mn , and kn such that n = mn kn , mn → ∞, kn → ∞, and γ(mn ) + Rmn β0 (kn ) + τ ≤  for given  and δ. The existence of mn and kn sequences such that mn β(kn ) → 0 is guaranteed by Lemma 3.1 in [49]. This implies δ → 0 as n → ∞, and thus the policy space Π is PAC-learnable under the hypothesis of the theorem. One of the assumptions we have made is that for policy π, the Radon–Nikodym derivative π is bounded by K. This essentially means that all the Markov chains are close to the reference Markov chain in the sense that the probability mass on any state does not differ by more than a multiplicative factor of K from that of P0 . The assumption that H has finite P-dimension is less natural but essential to the argument. Let us now show the existence of reference Markov chains through an example. Example 6.2. Consider any Markov chain with invariant measure λ0 . Consider any parametrized class of functions π : X → [0, 1], π ∈ Π such that x λ0 (x)π (x) = 1, and P-dim({π : π ∈ Π}) = d. Denote λπ (x) = λ0 (x)π (x) and consider a set of Markov chains P with invariant measures λπ . Then, clearly, λ0 is an invariant measure of a reference Markov chain for the set P of Markov chains. 7. Partially observable MDPs with general policies. We now consider partially observed discounted-reward MDPs with general policies (nonstationary with

1650

RAHUL JAIN AND PRAVIN P. VARAIYA

memory). The setup is as before, except that the policy depends on observations y ∈ Y, governed by the (conditional) probability ν(y|x) of observing y ∈ Y when the state is x ∈ X. Let ht denote the history (y0 , a1 , y1 , . . . , at , yt ) of observations and actions up to time t. The results of section 5 extend when the policies are nonstationary; however, there are many subtleties regarding the domain and range of simulation functions, and measures, and some details are different. Let Ht = {ht = (y0 , a1 , y1 , . . . , at , yt ) : as ∈ A, ys ∈ Y, 0 ≤ s ≤ t}. Let Π be the set of policies π = (π1 , π2 , . . .), with πt : Ht × A → [0, 1] a probability measure on A conditioned on ht ∈ Ht . Let Πt denote the set of all policies πt at time t with π ∈ Π. This gives rise to a conditional state transition function Pt (x, x ; ht ), the probability of transition from state x to x given history ht up to time t. Under π,  Pa (x, x )πt (ht , a). Pt (x, x ; ht ) = a

Let Pt denote the set of all Pπt induced by the policies πt with π ∈ Π. Then, defining the usual dT V (λ) metric on Pt and the usual L1 metric on Πt , we get the next result. Lemma 7.1. Suppose Πt and Pt are as defined above with P-dim(Πt ) = d. Assume λ and ρ are probability measures on X and Ht , respectively, and σ a probability measure on A such that πt (ht , a)/σ(a) ≤ K ∀ht ∈ Ht , a ∈ A, πt ∈ Πt . Then, for 0 <  < e/4 log2 e,

d 2eK 2eK N (, Pt , dT V (λ×ρ) ) ≤ N (, Πt /σ, dL1 (σ×ρ) ) ≤ . log   The proof can be found in the appendix. Let Ft be the set of simulation functions of Pt under the simple simulation model. Thus, ft ∈ Ft for t ≥ 2 is defined on ft : X × Ht−1 × Ω∞ → X × Ht × Ω∞ , while f1 ∈ F1 shall be defined on f1 : X × Ω∞ → X × H1 × Ω∞ . This is because at time t = 1, there is no history, and the state transition depends only on the initial state and the noise. For t > 1, the state transition depends on the history as well. Further, the function definitions have to be such that the composition ft ◦ ft−1 ◦ · · · ◦ f1 is well defined. It is straightforward to verify that Lemma 4.2 extends to Lemma 7.2. Lemma 7.2. Suppose P is convex (being generated by a convex space of policies Π). Let P-dim(Pt ) = d. Let Ft be the corresponding space of simple simulation functions induced by Pt . Then, P-dim(Ft ) = d. By F t we shall denote the set of functions f t = ft ◦ · · · ◦ f1 , where fs ∈ Fs and they arise from a common policy π. Note that f t : X × Ω∞ → Zt × Ω∞ , where Zt = X × Ht . We shall consider the following pseudometric on Ft with respect to a measure λt on Zt−1 for t ≥ 2 and measure σ on Ω∞ ,  λt (z)σ{ω : ft (z, ω) = gt (z, ω)}. ρt (ft , gt ) := z∈Zt−1

We shall take ρ1 as the pseudometric on F t with respect to the product measure λ×σ. Let  λf t (z) := λ(x)σ{ω : f t (x, ω) = (z, θt ω)} z∈Zt

be a probability measure on Zt . We now state the extension of the technical lemma needed for the main theorem of this section.

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1651

Lemma 7.3. Let λ be a probability measure on X and λf t be the probability measure on Zt as defined above. Suppose that P-dim (Ft ) ≤ d and there exists probability λf t (z) measures λt on Zt such that K := max{supt supf t ∈F t ,z∈Zt λt+1 (z) , 1} < ∞. Then, for 1 ≤ t ≤ T ,

dt     2eKt   2eKt , Ft , ρt · · · N , F1 , ρ1 ≤ log N (, F t , ρ1 ) ≤ N . Kt Kt   The proof can be found in the appendix. We now obtain our sample complexity result. Theorem 7.4. Let (X, Γ, λ) be the measurable state space, A the action space, Y the observation space, Pa (x, x ) the state transition function, and ν(y|x) the conditional probability measure that determines the observations. Let r(x) be the realvalued reward function bounded in [0, R]. Let Π be the set of stochastic policies (nonstationary and with memory in general), Pt be the set of state transition functions induced by Πt , and Ft the set of simulation functions of Pt under the simple simulation model. Suppose that P-dim (Pt ) ≤ d. Let λ and σ be probability measures on X and A, respectively, and λt+1 a probability measure on Zt such that λf t (z) t K := max{supt supf t ∈F t ,z∈Zt λt+1 (z) , 1} is finite, where λf is as defined above. Let Vˆn (π) be the estimate of V (π) obtained from n samples. Then, given any , δ > 0, and with probability at least 1 − δ, sup |Vˆn (π) − V (π)| <  π∈Π

 2   log 4δ + 2dT (log 32eR for n ≥ 32R α2 α + log KT ) , where T is the 2 -horizon time and  α = 2 (T + 1). Remarks. A special case is when the policies are stationary and memoryless, i.e., πt = π1 ∀ t, and π1 : Y×A → [0, 1] depends only on the current observation. Let Π1 be Markov chain the set of all π1 . Then, each policy π ∈ Π induces a time-homogeneous with probability transition function given by Pπ (x, x ) = a,y Pa (x, x )π1 (y, a)ν(y|x). Let P denote the set of all probability transition functions induced by Π. In general if Π is convex, P is convex. We will denote the set of simulation functions under the simple simulation model for P ∈ P by F. Suppose that P-dim (P) = d; then by Lemma 4.2, P-dim (P) = P-dim (F) = d. This implies that the sample complexity result of Theorem 5.2 for discounted-reward MDPs holds for the case when the state is partially observable and the policies are stationary and memoryless. Thus, for uniform convergence and estimation, partial observability of the state does not impose any extra cost in terms of sample complexity. Note that in the case of general polices, the sample complexity is O(T log T ) times more. 8. Conclusions. The paper considers simulation-based value function estimation methods for Markov decision processes (MDPs). Uniform sample complexity results are presented for the discounted-reward case. The combinatorial complexity of the space of simulation functions under the proposed simple simulation model is shown to be the same as that of the underlying space of induced Markov chains when the latter is convex. Using ergodicity and weak mixing leads to similar uniform sample complexity result for the average-reward case, when a reference Markov chain exists. Extensions of the results are obtained when the MDP is partially observable with general policies. Remarkably, the sample complexity results have the same order for both completely and partially observed MDPs when stationary and memoryless

1652

RAHUL JAIN AND PRAVIN P. VARAIYA

policies are used. Sample results for discounted-reward Markov games can be deduced easily as well. The results can be seen as an extension of the theory of PAC (probably approximately correct) learning for partially observable Markov decision processes (POMDPs) and games. PAC theory is related to the system identification problem. One of the key contributions of this paper is the observation that how we simulate an MDP matters for obtaining uniform estimates. This is a new (and surprising) observation. Thus, the results of this paper can also be seen as the first steps towards developing an empirical process theory for MDPs. Such a theory would go a long way in establishing a theoretical foundation for computer simulation of complex engineering systems. We have used Hoeffding’s inequality in obtaining the rate of convergence for discounted-reward MDPs and used the McDiarmid–Azuma inequality for the averagereward MDPs, though more sophisticated and tighter inequalities of Talagrand [41] (see also Samson [37]) can be used as well. This would yield better results and is part of future work. Appendix A. Proof of Lemma 3.1. Proof. We first relate the dT V (λ) pseudometric on P with the dL1 (λ×σ) pseudometric on Π. Pick any π, π  ∈ Π and denote P = Pπ and P  = Pπ . Then,          λ(x) Pa (x, y)(π(x, a) − π (x, a)) dT V (λ) (P, P ) =    y x a  ≤ λ(x)Pa (x, y)|π(x, a) − π  (x, a)| a





x

σ(a)

a

y



λ(x)|π(x, a) − π  (x, a)|/σ(a)

x

= dL1 (σ×λ) (π/σ, π  /σ). The second inequality above follows from the triangle inequality and the fact that the order of the sums over y, x, and a can be changed by Fubini’s theorem [18] and noting that y Pa (x, y) = 1. Thus, if Π /σ ⊆ Π/σ is an -net for Π/σ, {Pπ , π ∈ Π } is an -net for P. Further, the spaces Π and Π/σ have the same P-dimension, as can be easily verified. The bound on the covering number is then given by a standard result (see Theorem 4.3 of [49]). Appendix B. Proof of Lemma 5.1. Proof. Consider any f, g ∈ F, and x ∈ X . Then, μ{ω : f t (x, ω) = g t (x, ω)} = μ{ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω)} + μ{ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω)} = μ{∪y (ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω) = (y, θt−1 ω))} + μ{∪y (ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω), f t−1 (x, ω) = (y, θt−1 ω))} ≤ μ{∪y (ω : f (y, θt−1 ω) = g(y, θt−1 ω), f t−1 (x, ω) = g t−1 (x, ω) = (y, θt−1 ω))} + μ{∪y (ω : f t−1 (x, ω) = g t−1 (x, ω), f t−1 (x, ω) = (y, θt−1 ω))}  ≤ μ{ω : f (y, θt−1 ω) = g(y, θt−1 ω)|f t−1 (x, ω) = (y, θt−1 ω)}μ{ω : f t−1 (x, ω) = y} y

+ μ{ω : f t−1 (x, ω) = g t−1 (x, ω)}.

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1653

It is easy to argue that λf t−1 (y) ≤ K t−1 λ(y), where λf t−1 (y) = x λ(x)μ{ω : f t−1 (x, ω) = (y, θt−1 ω)}. Thus, multiplying both sides of the above sequence of inequalities and summing over x, and observing that μ{ω : f (y, θt−1 ω) = g(y, θt−1 ω)} = μ{ω  : f (y, ω  ) = g(y, ω  )}, we get that the first part of the right side is   λf t−1 (y)μ{ω  : f (y, ω  ) = g(y, ω  )} ≤ K t−1 · λ(y)μ{ω  : f (y, ω  ) = g(y, ω  )}. y

y

This implies that ρP (f t , g t ) ≤ K t−1 ρP (f, g) + ρP (f t−1 , g t−1 ) ≤ (K t−1 + K t−2 + · · · + 1)ρP (f, g) ≤ K t ρP (f, g), where the second inequality is obtained by induction. Now,    λ(x)μ{ω : f (x, ω) = g(x, ω)} ≤ λ(x) |f (x, ω) − g(x, ω)|dμ(ω), x

x

and thus ρP (f t , g t ) ≤ K t · dL1 (f, g), which proves the required assertion. Appendix C. Proof of Lemma 7.1. Proof. Pick any πt , πt ∈ Πt , and denote P = Pπt and P  = Pπt . Then,             dT V (λ×ρ) (P, P ) = λ(x) Pa (x, x )(πt (ht , a) − πt (ht , a)) ρ(ht )   x x ∈X a∈A,ht ∈Ht    ≤ ρ(ht ) λ(x) Pa (x, x )|πt (ht , a) − πt (ht , a)| ht

x

a

x

     πt (y, a) πt (y, a)    ρ(ht ) σ(a)  ≤ −  σ(a) σ(a) a ht

πt πt = dL1 (σ×ρ) , . σ σ  The second inequality above follows by changing the order of the sums over a and x ,  noting that x Pa (x, x ) = 1 and denoting x λ(x) = 1. The rest of the argument is the same as in Lemma 3.1.

Appendix D. Proof of Lemma 7.3. The proof of Lemma 7.3 is similar but the details are somewhat more involved. Proof. Consider any f t , g t ∈ F t , and x ∈ X . Then, μ{ω : f t (x, ω) = g t (x, ω)} = μ{ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω)} + μ{ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω)} = μ{∪z∈Zt−1 (ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω) = (z, θt−1 ω))} + μ{∪z∈Zt−1 (ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω)

= g t−1 (x, ω), f t−1 (x, ω) = (z, θt−1 ω))}

1654

RAHUL JAIN AND PRAVIN P. VARAIYA

≤ μ{∪z∈Zt−1 (ω : f t (x, ω) = g t (x, ω), f t−1 (x, ω) = g t−1 (x, ω) = (z, θt−1 ω))} + μ{∪z∈Zt−1 (ω : f t−1 (x, ω) = g t−1 (x, ω), f t−1 (x, ω) = (z, θt−1 ω))}  ≤ μ{ω : ft (z, θt−1 ω) = gt (z, θt−1 ω)|f t−1 (x, ω) z∈Zt−1

= (z, θt−1 ω)}μ{ω : f t−1 (x, ω) = (z, θt−1 ω)} + μ{ω : f t−1 (x, ω) = g t−1 (x, ω)}. Multiplying both sides of the above sequence of inequalities and summing over x, and observing again that μ{ω : ft (z, θt−1 ω) = gt (z, θt−1 ω)} = μ{ω  : ft (z, ω  ) = gt (z, ω  )}, we get that the first part of the right side is   λf t−1 (z)μ{ω  : ft (z, ω  ) = gt (z, ω  )} ≤ K· λt (z)μ{ω  : ft (z, ω  ) = gt (z, ω  )}. z∈Zt−1

z∈Zt−1

This by induction implies ρ1 (f t , g t ) ≤ K(ρt (ft , gt ) + · · · + ρ1 (f1 , g1 )), which implies the first inequality. For the second inequality, note that the ρ pseudometric and the L1 pseudometric are related; thus,    λt (z)μ{ω : ft (z, ω) = gt (z, ω)} ≤ λt (z) |ft (z, ω) − gt (z, ω)|dμ(ω), z

z

which relates their covering numbers. Also, the covering number under the L1 pseudometric can be bounded in terms of the P-dim of appropriate spaces. Acknowledgments. We thank Peter Bartlett, Tunc Simsek, and Antonis Dimakis for many helpful discussions. The counterexample 5.4(ii) was jointly developed with Antonis Dimakis. We also thank the anonymous referees for their comments which helped to greatly improve the quality of the paper. REFERENCES [1] D. Aldous and U. Vazirani, A Markovian extension of Valiant’s learning model, Inform. Comput., 117 (1995), pp. 181–186. [2] N. Alon, S. Ben-David, N. Cesa-Bianchi, and D. Haussler, Scale-sensitive dimension, uniform convergence and learnability, J. ACM, 44 (1997), pp. 615–631. [3] M. Anthony and P. Bartlett, Neural Network Learning: Theoretical Foundations, Cambridge University Press, Cambridge, UK, 1999. [4] E. Altman and O. Zeitouni, Rate of convergence of empirical measures and costs in controlled Markov chains and transient optimality, Math. Oper. Res., 19 (1994), pp. 955–974. [5] K. Azuma, Weighted sums of certain dependent random variables, Tˆ ohoku Math. J. (2), 19 (1967), pp. 357–367. [6] J. Baxtar and P. L. Bartlett, Infinite-horizon policy-gradient estimation, J. Artificial Intelligence Res., 15 (2001), pp. 319–350. [7] G. M. Benedek and A. Itai, Learning with respect to fixed distributions, Theoret. Comput. Science, 86 (1991), pp. 377–390. [8] V. Bentkus, On Hoeffding’s inequalities, Ann. Probab., 32 (2004), pp. 1650–1673. [9] V. Borkar, Topics in Controlled of Markov Chains, Pitman Res. Notes Math. 240, Longman Scientific Technical, Harlow, John Wiley, New York, 1991.

UNIFORM ESTIMATION FOR MARKOV DECISION PROCESSES

1655

[10] M. C. Campi and P. R. Kumar, Learning dynamical systems in a stationary environment, Systems Control Lett., 34 (1998), pp. 125–132. [11] H. Chernoff, A measure of asymptotic efficiency of tests of a hypothesis based on the sum of observations, Ann. Math. Statist., 23 (1952), pp. 493–507. [12] L. Devroye, L. Gyorfi, and G. Lugosi, A Probabilistic Theory of Pattern Recognition, Springer-Verlag, New York, 1996. [13] P. Diaconis and D. Freedman, Iterated random functions, SIAM Rev., 41 (1999), pp. 45–76. [14] P. Doukhan, Mixing: Properties and Examples, Lecture Notes in Statist., 85, Springer-Verlag, Berlin, 1994. [15] R. Dudley, Uniform Central Limit Theorems, Cambridge University Press, Cambridge, UK, 1999. [16] D. Gamarnik, Extensions of the PAC framework to finite and countable Markov chains, IEEE Trans. Inform. Theory, 49 (2003), pp. 338–345. [17] D. Haussler, Decision theoretic generalizations of the PAC model for neural nets and other learning applications, Inform. Comput., 100 (1992), pp. 78–150. [18] P. R. Halmos, Measure Theory, Springer-Verlag, Berlin, 1974. [19] W. Hoeffding, Probability inequalities for sums of bounded random variables, J. Amer. Statist. Assoc., 58 (1963), pp. 13–30. [20] M. Kearns, Y. Mansour, and A. Y. Ng, Approximate planning in large POMDPs via reusable trajectories, in Advances in Neural Information Processing Systems 12, MIT Press, 2000. [21] A. N. Kolmogorov and V. M. Tihomirov, -entropy and -capacity of sets in functional space, Amer. Math. Soc. Transl. Ser. 2, 17 (1961), pp. 277–364. [22] I. Kontoyiannis, L. A. L. Montano, and S. P. Meyn, Relative entropy and exponential deviation bounds for general Markov chains, in Proceedings of the International Symposium on Information Theory, IEEE, 2005, pp. 1563–1567. [23] M. Ledoux, The Concentration of Measure Phenomenon, Math. Surveys Monogr. 89, AMS, Providence, RI, 2001. [24] S. Mannor and J. N. Tsitsiklis, On the empirical state-action frequencies in Markov decision processes under general policies, Math. Oper. Res., 30 (2005), pp. 545–561. [25] P. Marbach and J. N. Tsitsiklis, Approximate gradient methods in policy-space optimization of Markov reward processes, Discrete Event Dyn. Sys., 13 (2003), pp. 111–148. [26] K. Marton, Measure concentration for Euclidean distance in the case of dependent random variables, Ann. Probab., 32 (2004), pp. 2526–2544. [27] C. McDiarmid, On the method of bounded differences, in Surveys in Combinatorics, Cambridge University Press, Cambridge, UK, 1989, pp. 148–188. [28] P. Massart, Some applications of concentration inequalities to statistics, Ann. Fac. Sci. Toulouse Math. (6), 9 (2000), pp. 245–303. [29] A. Nobel and A. Dembo, A note on uniform laws of averages for dependent processes, Statist. Probab. Lett., 17 (1993), pp. 169–172. [30] A. Y. Ng and M. I. Jordan, Pegasus: A policy search method for large MDPs and POMDPs, in Proceedings of the 16th Annual Conference on Uncertainty in Artificial Intelligence, Kaufman, San Franciso, CA, 2000, pp. 405–414. [31] C. H. Papadimitriou and J. N. Tsitsiklis, The complexity of Markov decision processes, Math. Oper. Res., 12 (1987), pp. 441–450. [32] M. Pesendorfer and P. S. Dengler, Identification and Estimation of Dynamic Games, Working Paper 9726, National Bureau of Economic Research, Cambridge, MA, 2003. [33] L. Peshkin and S. Mukherjee, Bounds on sample size for policy evaluation in Markov environments, in Computational Learning Theory, Lecture Notes in Comput. Sci. 2111, Springer, Berlin, 2001, pp. 616–629. [34] D. Pollard, Convergence of Stochastic Processes, Springer-Verlag, Berlin, 1984. [35] D. Pollard, Empirical Process Theory and Applications, Institute of Mathematical Statistics, Hayward, CA, 1990. [36] M. Puterman, Markov Decision Processes: Discrete Stochastic Dynamic Programming, John Wiley and Sons, New York, 1994. [37] P-M. Samson, Concentration of measure inequalities for Markov chains for Φ-mixing processes, Ann. Probab., 28 (2000), pp. 416–461. [38] D. H. Shim, H. J. Kim, and S. Sastry, Decentralized nonlinear model predictive control of multiple flying robots in dynamic environments, in Proceedings of the 42nd Conference IEEE Conference on Decision and Control, IEEE, Los Alamitos, 2003, pp. 3621–3626. [39] J. M. Steele, Probability Theory and Combinatorial Optimization, SIAM, Philadelphia, PA, 1997. [40] M. Talagrand, Concentration of measure and isoperimetric inequalities in product spaces, ´ Inst. Haute Etudes, Sci. Publ. Math., 81 (1995), pp. 73–205.

1656

RAHUL JAIN AND PRAVIN P. VARAIYA

[41] M. Talagrand, A new look at independence, Ann. Probab., 24 (1996), pp. 1–34. [42] L. Valiant, A theory of the learnable, Commun. ACM, 27 (1984), pp. 1134–1142. [43] S. Van de Geer, On Hoeffding’s inequality for dependent random variables, in Empirical Process Techniques for Dependent Data, H. Dehling, T. Mikosch, and M. Sorensen, eds., Birkh¨ auser Boston, Boston, MA, 2002, pp. 161–170. [44] A. W. van der Vaart and J. A. Wellner, Weak Convergence and Empirical Processes, Springer-Verlag, Berlin, 1996. [45] V. N. Vapnik and A. Ya. Chervonenkis, On the uniform convergence of relative frequencies to their probabilities, Theory Probab. Appl., 16 (1971), pp. 264–280. [46] V. N. Vapnik and A. Ya. Chervonenkis, Necessary and sufficient conditions for the uniform convergence of means to their expectations, Theory Probab. Appl., 26 (1981), pp. 532–553. [47] V. Vapnik and A. Chervonenkis, The necessary and sufficient conditions for consistency in the empirical risk minimization method, Pattern Recognition and Image Analysis, 1 (1991), pp. 283–305. [48] M. Vidyasagar and R. L. Karandikar, System identification: A learning theory approach, in Proceedings of the 40th IEEE Conference on Decision and Control, IEEE, Los Alamitos, CA, 2001, pp. 2001–2006 [49] M. Vidyasagar, Learning and Generalization: With Applications to Neural Networks, 2nd ed., Springer-Verlag London, LTD., London, 2003. [50] B. Yu, Rates of convergence of empirical processes for mixing sequences, Ann. Probab., 22 (1994), pp. 94–116.

Simulation-based uniform value function estimates of ...

for discounted MDPs was studied in [20, 30] in a machine learning context. ...... Scale the length of each of the m subintervals of the kth such choice to pm,k/m ≤.

254KB Sizes 4 Downloads 119 Views

Recommend Documents

2.4. Average Value of a Function (Mean Value Theorem) 2.4.1 ...
f(x)dx . 2.4.2. The Mean Value Theorem for Integrals. If f is con- tinuous on [a, b], then there exists a number c in [a, b] such that f(c) = fave = 1 b − a. ∫ b a f(x)dx ,.

Uniform value in dynamic programming - CiteSeerX
that for each m ≥ 0, one can find n(m) ≥ 1 satisfying vm,n(m)(z) ≤ v−(z) − ε. .... Using the previous construction, we find that for z and z in Z, and all m ≥ 0 and n ...

Uniform value in dynamic programming - CiteSeerX
Uniform value, dynamic programming, Markov decision processes, limit value, Black- ..... of plays giving high payoffs for any (large enough) length of the game.

Uniform value in Dynamic Programming
We define, for every m and n, the value vm,n as the supremum payoff the decision maker can achieve when his payoff is defined as the average reward.

Uniform value in dynamic programming
the supremum distance, is a precompact metric space, then the uniform value v ex- .... but then his payoff only is the minimum of his next n average rewards (as if ...

Selberg's zeta function and Dolgopyat's estimates for ...
(see for example [1, 24, 13, 14, 18] ) on the asymptotic distribution of closed orbits for various flows of hyperbolic nature and counting problems on higher rank symmetric spaces. Some recent applications of Dolgopyat's techniques to number theory a

Supplement to “A Critical Value Function Approach, with an ...
Jun 6, 2016 - transformations on the data g ◦ (y, x)=(y + Dg, x) , where g ... Fig S2: Critical Value Function (intercept and time trend). (a) ρ = 0.95. (b) ρ = −0.95.

Supplement to “A Critical Value Function Approach, with an ...
Jun 6, 2016 - this graph with Figure 1, it is evident that both bootstrap methods and subsampling have null rejection probabilities farther away from the 10% ...

A Critical Value Function Approach, with an Application ...
Jun 6, 2016 - Application to Persistent Time-Series. Marcelo J. ...... model. If the model has one lag, Hansen [9] suggests a “grid” bootstrap which delivers. 15 ...

Uniform Distribution - Khadi Uniform Clarification.pdf
Uniform Distribution - Khadi Uniform Clarification.pdf. Uniform Distribution - Khadi Uniform Clarification.pdf. Open. Extract. Open with. Sign In. Details. Comments. General Info. Type. Dimensions. Size. Duration. Location. Modified. Created. Opened

POINTWISE AND UNIFORM CONVERGENCE OF SEQUENCES OF ...
Sanjay Gupta, Assistant Professor, Post Graduate Department of .... POINTWISE AND UNIFORM CONVERGENCE OF SEQUENCES OF FUNCTIONS.pdf.

Nonparametric Estimates of the Labor-Supply Effects of ...
Unfortunately, responses to the data collection instrument (which depended on .... program designers to gauge how much the program will actually cost once.

Structure and function of mucosal immun function ...
beneath the epithelium directly in contact with M cells. Dendritic cells in ... genitourinary tract, and the breast during lactation (common mucosal immune system).

Paired-Uniform Scoring - University of Pittsburgh
where one of the two bags is selected with equal probability. Two-thirds of the ... To determine your payment the computer will randomly draw two numbers. ... the previous month's unemployment rate is lower than a year ago. .... tractable, and capabl

Constitution of uniform Committee.PDF
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Constitution of ...Missing:

Uniform Crl. Rules Of Practice.pdf
criminal justice and to usher in a certain amount of. uniformity, and acceptance of best practices prevailing. over various parts of India, this Court may consider.

Uniform Package.pdf
Page 1 of 1. ORDER FORM. NAME. PHONE. EMAIL. S M L XL 2XL 3XL. Package Total: $255.00. TRACK & FIELD. TRACK & FIELD. TRACK & FIELD. UA Women's Singlet & Shorts UA Men's Singlet & Shorts. S M L XL 2XL 3XL S M L XL 2XL 3XL. S M L XL 2XL S M L XL 2XL S

ENTROPY ESTIMATES FOR A FAMILY OF EXPANDING ... - CiteSeerX
where Js denotes the Bessel function of the first kind and order s. We recall ..... by evaluation of the Lyapunov exponent on a grid of stepsize 10−3 and use of the.

Age-of-acquisition and subjective frequency estimates ...
Because collecting such data is time consum- ing, it is good practice to make these ratings ... Psychonomic Society's Archive of Norms, Stimuli, and Data, www.psychonomic.org/archive. Behavior Research Methods. 2008, 40 (4) ... 2004; Perry, Ziegler,

Uniform Committee.PDF
Loading… Whoops! There was a problem loading more pages. Whoops! There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Uniform Committee.PDF. Uniform Committee.

uniform-guidance.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps. ... of the apps below to open or edit this item. uniform-guidance.pdf.

Making sense of heat tolerance estimates in ectotherms ...
The confounding effects of stochasticity, resource depletion (or fatigue) and short-term accli- matory responses are ... zation that patterns emerging from alternative protocols .... concern during thermal tolerance assays than food or energy.

Asymptotic behavior of RA-estimates in autoregressive ...
article (e.g. in Word or Tex form) to their personal website or institutional ..... The least square estimator of /0 based on ˜XM, with domain ̂M ⊂ is defined as the ...