b

DiscoverSearch
About
My stuff
Target-Based Temporal Difference Learning
2019·arXiv
Abstract
Abstract

The use of target networks has been a popular and key component of recent deep Q-learning algorithms for reinforcement learning, yet little is known from the theory side. In this work, we introduce a new family of target-based temporal difference (TD) learning algorithms and provide theoretical analysis on their convergences. In contrast to the standard TD-learning, target-based TD algorithms maintain two separate learning parameters–the target variable and online variable. Particularly, we introduce three members in the family, called the averaging TD, double TD, and periodic TD, where the target variable is updated through an averaging, symmetric, or periodic fashion, mirroring those techniques used in deep Q-learning practice.

We establish asymptotic convergence analyses for both averaging TD and double TD and a finite sample analysis for periodic TD. In addition, we also provide some simulation results showing potentially superior convergence of these target-based TD algorithms compared to the standard TD-learning. While this work focuses on linear function approximation and policy evaluation setting, we consider this as a meaningful step towards the theoretical understanding of deep Q-learning variants with target networks.

Deep Q-learning [Mnih et al., 2015] has recently captured significant attentions in the reinforcement learning (RL) community for outperforming human in several challenging tasks. Besides the effective use of deep neural networks as function approximators, the success of deep Q-learning is also indispensable to the utilization of a separate target network for calculating target values at each iteration. In practice, using target networks is proven to substantially improve the performance of Q-learning algorithms, and is gradually adopted as a standard technique in modern implementations of Q-learning.

To be more specific, the update of Q-learning with target network can be viewed as follows:

image

where  yt = r(st, at) + γ maxa Q(st+1, a; θ′t), θt is the online variable, and  θ′t is the target variable. Here the state-action value function  Q(s, a; θ) is parameterized by  θ. The update of the online variable  θtresembles the stochastic gradient descent step. The term  r(st, at) stands for the intermediate reward of taking action  at in state st, and ytstands for the target value under the target variable,  θ′t. When the target variable is set to be the same as the online variable at each iteration, this reduces to the standard Q-learning algorithm [Watkins & Dayan, 1992], and is known to be unstable with nonlinear function approximations. Several choices of target networks are proposed in the literature to overcome such instability: (i) periodic update, i.e., the target variable is copied from the online variable every  τ >0 steps, as used for deep Q-learning [Gu et al., 2016, Mnih et al., 2015, 2016, Wang et al., 2016]; (ii) symmetric update, i.e., the target variable is updated symetrically as the online variable; this is first introduced in double Q-learning [Hasselt, 2010, Van Hasselt et al., 2016]; and (iii) Polyak averaging update, i.e., the target variable takes weighted average over the past values of the online variable; this is used in deep deterministic policy gradient [Heess et al., 2015, Lillicrap et al., 2015] as an example. In the following, we simply refer these as target-based Q-learning algorithms.

While the integration of Q-learning with target networks turns out to be successful in practice, its theoretical convergence analysis remains largely an open yet challenging question. As an intermediate step towards the answer, in this work, we first study target-based temporal difference (TD) learning algorithms and establish their convergence analysis. TD algorithms [Sutton, 1988, Sutton et al., 2009a,b] are designed to evaluate a given policy and are the fundamental building blocks of many RL algorithms. Comprehensive surveys and comparisons among TD-based policy evaluation algorithms can be found in Dann et al. [2014]. Motivated by the target-based Q-learning algorithms [Mnih et al., 2015, Wang et al., 2016], we introduce a target variable into the TD framework and develop a family of target-based TD algorithms with different updating rules for the target variable. In particular, we propose three members in the family, the averaging TD, double TD, and periodic TD, where the target variable is updated through an averaging, symmetric or periodic fashion, respectively. Meanwhile, similar to the standard TD-learning, the online variable takes stochastic gradient steps of the Bellman residual loss function while freezing the target variable. As the target variable changes slowly compared to the online variable, target-based TD algorithms are prone to improve the stability of learning especially if large neural networks are used, although this work will focus mainly on TD with linear function approximators.

Theoretically, we prove the asymptotic convergence of both averaging TD and double TD. We also provide a finite sample analysis for the periodic TD algorithm. Practically, we also run some simulations showing superior convergence of the proposed target-based TD algorithms compared to the standard TD-learning. In particular, our empirical case studies demonstrate that the target TD-learning algorithms outperforms the standard TD-learning in the long run with smaller errors and lower variances, despite their slower convergence at the very beginning. Moreover, our analysis reveals an important connection between the TD-learning and the target-based TD-learning. We consider the work as a meaningful step towards the theoretical understanding of deep Q-learning with general nonlinear function approximation.

Related work. The first target-based reinforcement learning was proposed in [Mnih et al., 2015] for policy optimization problems with nonlinear function approximation, where only empirical results were given. To our best knowledge, target-based reinforcement learning for policy evaluation has not been specifically studied before. A somewhat related family of algorithms are the gradient TD (GTD) learning algorithms [Dai et al., 2017, Mahadevan et al., 2014, Sutton et al., 2009a,b], which minimize the projected Bellman residual through the primal-dual algorithms. The GTD algorithms share some similarities with the proposed target-based TD-learning algorithms in that they also maintain two separate variables – the primal and dual variables, to minimize the objective. Apart from this connection, the GTD algorithms are fundamentally different from the averaging TD and double TD algorithms that we propose. The proposed periodic TD algorithm can be viewed as approximately solving least squares problems across cycles, making it closely related to two families of algorithms, the least-square TD (LSTD) learning algorithms [Bertsekas, 1995, Bradtke & Barto, 1996] and the least squares policy evaluation (LSPE) [Bertsekas & Yu, 2009, Yu & Bertsekas, 2009]. But they also distinct from each other in terms of the subproblems and subroutines used in the algorithms. Particularly, the periodic TD executes stochastic gradient descent steps while LSTD uses the least-square parameter estimation method to minimize the projected Bellman residual. On the other hand, LSPE directly solves the subproblems without successive projected Bellman operator iterations. Moreover, the proposed periodic TD algorithm enjoys a simple finite-sample analysis based on existing results on stochastic approximation.

In this section, we briefly review the basics of the TD-learning algorithm with linear function approximation. We first list a few notations that will be used throughout the paper.

Notation The following notation is adopted: for a convex closed set  S, ΠS(x) is the projection of x onto the set  S, i.e., ΠS(x) := argminy∈S∥x − y∥2; diam(S) := supx∈S,y∈S ∥x − y∥2 is thediameter of the set  S; ∥x∥D :=√xT Dxfor any positive-definite  D; λmin(A) and λmax(A) denotesthe minimum and maximum eigenvalues of a symmetric matrix A, respectively.

2.1 Markov Decision Process (MDP)

In general, a (discounted) Markov decision process is characterized by the tuple  M := (S, A, P, r, γ),where S is a finite state space, A is a finite action space,  P(s, a, s′) := P[s′|s, a] represents the (unknown) state transition probability from state  s to s′ given action  a, r : S × A → [0, σ] is auniformly bounded stochastic reward, and  γ ∈ (0,1) is the discount factor. If action a is selected with the current state s, then the state transits to  s′ with probability  P(s, a, s′) and incurs a random reward  r(s, a) ∈ [0, σ] with expectation R(s, a). A stochastic policy is a distribution  π ∈ ∆|S|×|A|representing the probability  π(s, a) = P[a|s], P π denotes the transition matrix whose (s, s′) entryis  P[s′|s] = �a∈A P(s, a, s′)π(s, a), and d ∈ ∆|S| denotes the stationary distribution of the state s ∈ Sunder policy  π, i.e., d = dP π. The following assumption is standard in the literature.

Assumption 1. We assume that  d(s) > 0 for all s ∈ S.

We also define  rπ(s) and Rπ(s) as the stochastic reward and its expectation given the policy  πand the current state s, i.e.

image

The infinite-horizon discounted value function given policy  π is

image

where  s ∈ S, Estands for the expectation taken with respect to the state-action-reward trajectories.

2.2 Linear Function Approximation

Given pre-selected basis (or feature) functions  φ1, . . . , φn : S → R, Φ ∈ R|S|×n is defined as

image

Here  n ≪ |S|is a positive integer and  φ(s) is a feature vector. It is standard to assume that the columns of Φ do not have any redundancy up to linear combinations. We make the following assumption.

Assumption 2. Φhas full column rank.

2.3 Reinforcement Learning (RL) Problem

In this paper, the goal of RL with the linear function approximation is to find the weight vector θ ∈ Rn such that Jθ := Φθapproximates the true value function  Jπ. This is typically done by minimizing the mean-square Bellman error loss function [Sutton et al., 2009a]

image

where D is defined as a diagonal matrix with diagonal entries equal to a stationary state distribution d under the policy  π. Note that due to Assumption 1,  D ≻0. In typical RL setting, the model is unknown, while only samples of the state-action-reward are observed. Therefore, the problem can only be solved in stochastic way using the observations. In order to formally analyze the sample complexity, we consider the following assumption on the samples.

Assumption 3. There exists a Sampling Oracle (SO) that takes input (s, a) and generates a new state  s′ with probabilities  P(s, a, s′)and a stochastic reward  r(s, a) ∈ [0, σ].

This oracle model allows us to draw i.i.d. samples (s, a, r, s′) from s ∼ d(·), a ∼ π(s, ·), s′ ∼P(s, a, ·). While such an i.i.d. assumption may not necessarily hold in practice, it is commonly adopted for complexity analysis of RL algorithms in the literature [Bhandari et al., 2018, Dalal et al., 2018, Sutton et al., 2009a,b]. It’s worth mentioning that several recent works also provide complexity analysis when only assuming Markovian noise or exponentially  β-mixing properties of the samples [Antos et al., 2008, Bhandari et al., 2018, Dai et al., 2018, Srikant & Ying., 2019]. For sake of simplicity, this paper only focuses on the i.i.d. sampling case.

A naive idea for solving 1 is to apply the stochastic gradient descent steps,  θk+1 = θk−αk ˜∇θl(θk),where  αk >0 is a step-size and ˜∇θl(θk) is a stochastic estimator of the true gradient of  l at θ = θk,

image

This approach is called the residual method [Baird, 1995]. Its main drawback is the double sampling issue [Bertsekas & Tsitsiklis, 1996, Lemma 6.10, pp. 364]: to obtain an unbiased stochastic estimation of  ∇θl(θk), we need two independent samples given any pair (s, a) ∈ S × A. This ispossible under Assumption 3, but hardly implementable in most real applications.

2.4 Standard TD-Learning

In the standard TD-learning [Sutton, 1988], the gradient term  Es′[γ∇θJθk(s′)] in the last line (∇θl(θk)) is omitted [Bertsekas & Tsitsiklis, 1996, pp. 369]. The resulting update rule is

image

While the algorithm avoids the double sampling problem and is simple to implement, a key issue here is that the stochastic gradient  η(θk) does not correspond to the true gradient of the loss function l(θ) or any other objective functions, making the theoretical analysis rather subtle. Asymptotic convergence of the TD-learning was given in the original paper [Sutton, 1988] in tabular case and in Tsitsiklis & Van Roy [1997] with linear function approximation. Finite-time convergence analysis was recently established in Bhandari et al. [2018], Dalal et al. [2018], Srikant & Ying. [2019].

Remark. The TD-learning can also be interpreted as minimizing the modified loss function at each iteration

image

where  θstands for an online variable and  θ′ stands for a target variable. At each iteration step k, it sets the target variable to the value of current online variable and performs a stochastic gradient step

image

A full algorithm is described in Algorithm 1.

image

Inspired by the the recent target-based deep Q-learning algorithms [Mnih et al., 2015], we consider several alternative updating rules for the target variable that are less aggressive and more general. This then leads to the so-called target-based TD-learning. One of the potential benefits is that by slowing down the update for the target variable, we can reduce the correlation of the target value, or the variance in the gradient estimation, which would then improve the stability of the algorithm. To this end, we introduce three variants of target-based TD: averaging TD, double TD, and periodic TD, each of which corresponds to a different strategy of the target update. In the following sections, we discuss these algorithms in details and provide their convergence analysis.

We start by integrating TD-learning with the Polyak averaging strategy for target variable update. This is motivated by the recent deep Q-learning [Mnih et al., 2015] and DDPG [Lillicrap et al., 2015]. It’s worth pointing out that such a strategy has been commonly used in the deep Q-learning framework, but the convergence analysis remains absent to our best knowledge. Here we first study this for the TD-learning. The basic idea is to minimize the modified loss,  l(θ; θ′), with respect to  θwhile freezing  θ′, and then enforce  θ′ → θ(target tracking). Roughly speaking, the tracking step, θ′ → θ, is executed with the update

image

where  δ >0 is the parameter used to adjust the update speed of the target variable and ˜∇θl(θ; θ′k)is a stochastic estimation of  ∇θl(θ; θ′k). A full algorithm is summarized in Algorithm 2, which is called averaging TD (A-TD).

Compared to the standard TD-learning in Algorithm 1, the only difference comes from the target variable update in the last line of Algorithm 2. In particular, if we set  αk = 1/δand replace θk with θk+1in the second update, then it reduces to the TD-learning.

image

Next, we prove its convergence under certain assumptions. The convergence proof is based on the ODE (ordinary differential equation) approach [Bhatnagar et al., 2012], which is standard technique used in the RL literature [Sutton et al., 2009b]. In the approach, a stochastic recursive algorithm is converted to the corresponding ODE, and the stability of the ODE is used to prove the convergence. The ODE associated with A-TD is as follows:

image

We arrive at the following convergence result,.

Theorem 1. Assume that with a fixed policy  π, the Markov chain is ergodic and the step-sizes satisfy

image

Then,  θ′k → θ∗ and θk → θ∗ as k → ∞with probability one, where

image

Remark 1. Note that θ∗ in (5)is not identical to the optimal solution of the original problem in (1). Instead, it is the solution of the projected Bellman equation defined as

image

where F is the projected Bellman operator defined by

image

Π is the projection onto the range space of Φ, denoted by  R(Φ): Π(x) := arg minx′∈R(Φ) ∥x − x′∥2D.The projection can be performed by the matrix multiplication: we write Π(x) := Πx, where Π :=Φ(ΦT DΦ)−1ΦT D.

Theorem 1 implies that both the target and online variables of the A-TD converge to  θ∗ whichsolves the projected Bellman equation. The proof of Theorem 1 is provided in Appendix A based on the stochastic approximation approach, where we apply the Borkar and Meyn theorem [Bhatnagar et al., 2012, Appendix D]. Alternatively, the multi-time scale stochastic approximation [Bhatnagar et al., 2012, pp. 23] can be used with slightly different step-size rules. Due to the introduction of target variable updates, deriving a finite-sample analysis for the modified TD-learning is far from straightforward [Bhandari et al., 2018, Dalal et al., 2018]. We will leave this for future investigation.

In this section, we introduce a natural extension of the A-TD, which has a more symmetric form. The algorithm mirrors the double Q-learning [Van Hasselt et al., 2016], but with a notable difference. Here, both the online variable and target variable are updated in the same fashion by switching roles. To enforce  θ′ → θ, we also add a correction term  δ(θ − θ′) to the gradient update. The algorithm is summarized in Algorithm 3, and referred to as the double TD-learning (D-TD).

image

We provide the convergence of the D-TD with linear function approximation below. The proof is similar to the proof of Theorem 1, and is contained in Appendix B. Noting that asymptotic convergence for double Q-learning has been established in Hasselt [2010] for tabular case, but no result is yet known when linear function approximation is used.

Theorem 2. Assume that with a fixed policy  π, the Markov chain is ergodic and the step-sizes satisfy (4). Then,  θk → θ∗ and θ′k → θ∗ as k → ∞with probability one.

If D-TD uses identical initial values for the target and online variables, then the two updates remain identical, i.e.,  θk = θ′k for k ≥0. In this case, D-TD is equivalent to the TD-learning with a variant of the step-size rule. In practice, this problem can also be resolved if we use different samples for each update, and the convergence result will still apply to this variation of D-TD.

Compared to the corresponding form of the double Q-learning [Hasselt, 2010], D-TD has two modifications. First, we introduce an additional term,  δ(θ′k − θk) or δ(θk − θ′k), linking the target and online parameter to enforce a smooth update of the target parameter. This covers double Q-learning as a special case by setting  δ= 0. Moreover, the D-TD updates both target and online parameters in parallel instead of randomly. This approach makes more efficient use of the samples in a slight sacrifice of the computation cost. The convergence of the randomized version is proved with slight modification of the corresponding proof (see Appendix C for details).

In this section, we propose another version of the target-based TD-learning algorithm, which more resembles that used in the deep Q-learning [Mnih et al., 2015]. It corresponds to the periodic update form of the target variable, which differs from previous sections. Roughly speaking, the target variable is only periodically updated as follows:

image

where ˜∇θl(θ; θk−(k mod L)) is a stochastic estimator of the gradient  ∇θl(θ; θk−(k mod L)). The standard TD-learning is recovered by setting L = 1.

Alternatively, one can interpret every L iterations of the update as contributing to minimizing the modified loss function

image

while freezing the target variable. In other words, the above subproblem is approximately solved at each iteration through L steps of stochastic gradient descent. We formally present the algorithmic idea in a more general way as depicted in Algorithm 4 and call it the periodic TD algorithm (P-TD).

For the P-TD, given a fixed target variable  θ′k, the subroutine,  SGD(θk, θ′k, Lk), runs stochastic gradient descent steps  Lktimes in order to approximately solve the subproblem arg minθ∈Rn l(θ; θ′k),for which an unbiased stochastic gradient estimator is obtained by using observations. Upon solving the subproblem after  Lksteps, the next target variable is replaced with the next online variable. This makes it similar to the original deep Q-learning [Mnih et al., 2015] as it is periodic if  Lkis set to a constant. Moreover, P-TD is also closely related to the TD-learning Algorithm 1. In particular, if  Lk= 0 for all  k = 0, 1, . . . , T −1, then P-TD corresponds to the standard TD.

Based on the standard results in Bottou et al. [2018, Theorem 4.7], the SGD subroutine converges to the optimal solution,  θ∗k+1 := arg minθ∈Rn l(θ; θ′k). But as we only apply a finite number  Lksteps of SGD, the subroutine will return an approximate solution with a certain error bound  εk inexpectation, i.e.,  E[∥θk+1 − θ∗k+1∥22|θk] ≤ εk+1.

image

In the following, we establish a finite-time convergence analysis of P-TD. We first present a result in terms of the expected error of the solution and its bounds with high probability.

Theorem 3. Consider Algorithm 4. We have

image

Moreover,

image

The second result implies that P-TD achieves an  ϵ-optimal solution with an arbitrarily high probability by approaching  T → ∞and controlling the error bounds  εk. In particular, if  εk = ε forall  k ≥ 0, then

image

One can see that the error is essentially decomposed into two terms, one from the approximation errors induced from SGD procedures and one from the contraction property of solving the subproblems, which can also be viewed as solving the projected Bellman equations. Full details of the proof can be found in Appendix D.

To further analyze the approximation error from the SGD procedure, existing convergence results in Bottou et al. [2018, Theorem 4.7] can be applied with slight modifications.

Proposition 1. Suppose that the SGD method in Algorithm 4 is run with a stepsize sequence such that, for all  t ≥ 0,

image

for some  β > 1/λmin(ΦT DΦ) and κ > 0 such that

image

Then, for all 0  ≤ t ≤ Lk − 1, the expected optimality gap satisfies

image

where

image

and

image

Proposition 1 ensures that the subroutine iterate,  θk, converges to the solution of the subproblem at the rate of  O(1/Lk). Combining Proposition 1 with Theorem 3, the overall sample complexity is derived in the following proposition. We defer the proofs to Appendix E and Appendix F.

Proposition 2 (Sample Complexity). The ϵ-optimal solution,  E[∥θT − θ∗∥D] ≤ ϵ, is obtained by Algorithm 4 with SO calls at most  ρ1(ρ2ϵ−2 + 4χ2) ln(ϵ−1), where

image

and  χ1 and χ2are defined in Proposition 1.

As a result, the overall sample complexity of P-TD is bounded by  O((1/ϵ2) ln(1/ϵ)). As mentioned earlier, non-asymptotic analysis for even the standard TD algorithm is only recently developed in a few work [Bhandari et al., 2018, Dalal et al., 2018, Srikant & Ying., 2019]. Our sample complexity result on P-TD, which is a target-based TD algorithm, matches with that developed in Bhandari et al. [2018] with similar decaying step-size sequence, up to a log factor. Yet, our analysis is much simpler and builds directly upon existing results on stochastic gradient descent. Moreover, from the computational perspective, although P-TD runs in two loops, it is as the efficient as standard TD.

P-TD also shares some similarity with the least squares temporal difference (LSTD, Bradtke & Barto [1996]) and its stochastic approximation variant (fLSTD-SA, Prashanth et al. [2014]). LSTD is a batch algorithm that directly estimates the optimal solution as described in (5) through samples, which can also be viewed as exactly computing the solution to a least squares subproblem. fLSTD-SA alleviates the computation burden by applying the stochastic gradient descent (the same as TD update) to solve the subproblems. The key difference between fLSTD-SA and P-TD lies in that the objective for P-TD is adjusted by the target variables across cycles. Lastly, P-TD is also closely related to and can be viewed as a special case of the least-squares fitted Q-iteration [Antos et al., 2008]. Both of them solves a similar least squares problems using target values. However, for P-TD, we are able to directly apply the stochastic gradient descent to address the subproblems to near-optimality.

In this section, we provide some preliminary numerical simulation results showing the efficiency of the proposed target-based TD algorithms. We stress that the main goal of this paper is to introduce the family of target-based TD algorithms with linear function approximation and provide theoretical convergence analysis for target TD algorithms, as an intermediate step towards the understanding of target-based Q-learning algorithms. Hence, our numerical experiments simply focus on testing the convergence, sensitivity in terms of the tuning parameters of these target-based algorithms, as well as effects of using target variables as opposed to the standard TD-learning.

6.1 Convergence of A-TD and D-TD

In this example, we consider an MDP with  γ = 0.9, |S| = 10,

image

and  rπ(s) ∼ U[0,20], where U[0, 20] denotes the uniform distribution in [0, 20] and rπ(s) stands for the reward given policy  πand the current state s. The action space and policy are not explicitly defined here. For the linear function approximation, we consider the feature vector with the radial basis function [Geramifard et al., 2013] (n = 2)

image

image

Figure 1: (a) Blue line: error evolution of the standard TD-learning with the step-size  αk =1000/(k + 10000); Red line: error evolution of A-TD with the step-size  αk = 1000/(k+ 10000) and δ = 0.9. The shaded areas depict empirical variances obtained with several realizations. (a) Error over the interval [0, 3000]; (b) Error over the interval [2000, 3000].

image

Figure 2: Blue line: error evolution of the standard TD-learning with the step-size  αk = 1000/(k+10000); Red line: error evolution of D-TD with the step-size  αk = 1000/(k+ 10000) and  δ = 0.9.The shaded areas depict empirical variances obtained with several realizations. (a) Error over the interval [0, 3000]; (b) Error over the interval [2000, 3000].

Simulation results are given in Figure 1, which illustrate error evolution of the standard TDlearning (blue line) with the step-size,  αk = 1000/(k+ 10000) and the proposed A-TD (red line) with the  αk = 1000/(k+ 10000) and  δ = 0.9.The design parameters of both approaches are set to demonstrate reasonably the best performance with trial and errors. Additional simulation results in Appendix G provide comparisons for several different parameters. Figure 1(b) provides the results in the same plot over the interval [2000, 3000]. The results suggest that although A-TD with  δ = 0.9 initially shows slower convergence, it eventually converges faster than the standard

image

Figure 3: Blue line: error evolution of the standard TD-learning with the step-size  αk = 10000/(k+10000). Red line: error of P-TD with the step-size  βt = (10000· (0.997)k)/(10000+ t) and Lk = 40.The shaded areas depict empirical variances obtained with several realizations. (a) Error over the interval [0, 30000]; (b) Error over the interval [29000, 30000].

TD with lower variances after certain iterations. With the same setting, comparative results of D-TD are given in Figure 2.

6.2 Convergence of P-TD

In this section, we provide empirical comparative analysis of P-TD and the standard TD-learning. The convergence results of both approaches are quite sensitive to the design parameters to be determined, such as the step-size rules and total number of iterations of the subproblem. We consider the same example as above but with an alternative linear function approximation with the feature vector consisting of the radial basis function

image

From our own experiences, applying the same step-size rule,  βt, for every  k ∈ {0, 1, . . . , T − 1}yields unstable fluctuations of the error in some cases. For details, the reader is referred to Ap- pendix G, which provides comparisons with different design parameters. The results motivate us to apply an adaptive step-size rules for the subproblem of P-TD so that smaller and smaller step-sizes are applied as the outer-loop steps increases. In particular, we employ the adaptive step-size rule,  βk,t = (10000 · (0.997)k)/(10000 + t) with Lk= 40 for P-TD, and the corresponding simulation results are given in Figure 3, where P-TD outperforms the standard TD with the step-size, αk = 10000/(k+ 10000), best tuned for comparison. Figure 3(b) provides the results in Figure 3 in the interval [29000, 30000], which clearly demonstrates that the error of P-TD is smaller with lower variances.

In this paper, we propose a new family of target-based TD-learning algorithms, including the averaging TD, double TD, and periodic TD, and provide theoretical analysis on their convergences. The proposed TD algorithms are largely inspired by the recent success of deep Q-learning using target networks and mirror several of the practical strategies used for updating target network in the literature. Simulation results show that integrating target variables into TD-learning can also help stabilize the convergence by reducing variance of and correlations with the target. Our convergence analysis provides some theoretical understanding of target-based TD algorithms. We hope this would also shed some light on the theoretical analysis for target-based Q-learning algorithms and non-linear RL frameworks.

Possible future topics include (1) developing finite-time convergence analysis for A-TD and DTD; (2) extending the analysis of the target-based TD-learning to the Q-learning case w/o function approximation; and (3) generalizing the target-based framework to other variations of TD-learning and Q-learning algorithms.

Antos, A., Szepesv´ari, C., and Munos, R. Learning near-optimal policies with Bellman-residual minimization based fitted policy iteration and a single sample path. Machine Learning, 71(1): 89–129, Apr 2008.

Antsaklis, P. J. and Michel, A. N. A linear systems primer. 2007.

Baird, L. Residual algorithms: reinforcement learning with function approximation. In Machine Learning Proceedings, pp. 30–37. 1995.

Bertsekas, D. P. Dynamic programming and optimal control. Athena Scientific Belmont, MA, 1995.

Bertsekas, D. P. and Tsitsiklis, J. N. Neuro-dynamic programming. Athena Scientific Belmont, MA, 1996.

Bertsekas, D. P. and Yu, H. Projected equation methods for approximate solution of large linear systems. Journal of Computational and Applied Mathematics, 227(1):27–50, 2009.

Bhandari, J., Russo, D., and Singal, R. A finite time analysis of temporal difference learning with linear function approximation. arXiv preprint arXiv:1806.02450, 2018.

Bhatnagar, S., Prasad, H. L., and Prashanth, L. A. Stochastic recursive algorithms for optimization: simultaneous perturbation methods, volume 434. Springer, 2012.

Bottou, L., Curtis, F. E., and Nocedal, J. Optimization methods for large-scale machine learning. Siam Review, 60(2):223–311, 2018.

Boyd, S. and Vandenberghe, L. Convex optimization. Cambridge University Press, 2004.

Bradtke, S. J. and Barto, A. G. Linear least-squares algorithms for temporal difference learning. Machine Learning, 22(1):33–57, Mar 1996.

Bubeck, S. et al. Convex optimization: Algorithms and complexity.  Foundations and Trends R⃝ inMachine Learning, 8(3-4):231–357, 2015.

Chen, C.-T. Linear System Theory and Design. Oxford University Press, Inc., 1995.

Dai, B., He, N., Pan, Y., Boots, B., and Song, L. Learning from conditional distributions via dual embeddings. In Artificial Intelligence and Statistics, pp. 1458–1467, 2017.

Dai, B., Shaw, A., Li, L., Xiao, L., He, N., Liu, Z., Chen, J., and Song, L. SBEED: Convergent reinforcement learning with nonlinear function approximation. In Dy, J. and Krause, A. (eds.), Proceedings of the 35th International Conference on Machine Learning, volume 80 of Proceedings of Machine Learning Research, pp. 1125–1134. PMLR, 10–15 Jul 2018.

Dalal, G., Sz¨or´enyi, B., Thoppe, G., and Mannor, S. Finite sample analyses for TD(0) with function approximation. In Thirty-Second AAAI Conference on Artificial Intelligence, 2018.

Dann, C., Neumann, G., and Peters, J. Policy evaluation with temporal differences: A survey and comparison. Journal of Machine Learning Research, 15(1):809–883, 2014.

Geramifard, A., Walsh, T. J., Tellex, S., Chowdhary, G., Roy, N., How, J. P., et al. A tutorial on linear function approximators for dynamic programming and reinforcement learning. Foundations and Trends R⃝ in Machine Learning, 6(4):375–451, 2013.

Gu, S., Lillicrap, T., Sutskever, I., and Levine, S. Continuous deep q-learning with model-based acceleration. In International Conference on Machine Learning, pp. 2829–2838, 2016.

Hasselt, H. V. Double Q-learning. In Advances in Neural Information Processing Systems, pp. 2613–2621, 2010.

Heess, N., Hunt, J. J., Lillicrap, T. P., and Silver, D. Memory-based control with recurrent neural networks. arXiv preprint arXiv:1512.04455, 2015.

Lillicrap, T. P., Hunt, J. J., Pritzel, A., Heess, N., Erez, T., Tassa, Y., Silver, D., and Wierstra, D. Continuous control with deep reinforcement learning. arXiv preprint arXiv:1509.02971, 2015.

Mahadevan, S., Liu, B., Thomas, P., Dabney, W., Giguere, S., Jacek, N., Gemp, I., and Liu, J. Proximal reinforcement learning: A new theory of sequential decision making in primal-dual spaces. arXiv preprint arXiv:1405.6757, 2014.

Mnih, V., Kavukcuoglu, K., Silver, D., Rusu, A. A., Veness, J., Bellemare, M. G., Graves, A., Ried- miller, M., Fidjeland, A. K., Ostrovski, G., et al. Human-level control through deep reinforcement learning. Nature, 518(7540):529, 2015.

Mnih, V., Badia, A. P., Mirza, M., Graves, A., Lillicrap, T., Harley, T., Silver, D., and Kavukcuoglu, K. Asynchronous methods for deep reinforcement learning. In International conference on machine learning, pp. 1928–1937, 2016.

Prashanth, L. A., Korda, N., and Munos, R. Fast lstd using stochastic approximation: Finite time analysis and application to traffic control. In Calders, T., Esposito, F., H¨ullermeier, E., and Meo, R. (eds.), Machine Learning and Knowledge Discovery in Databases, pp. 66–81. Springer Berlin Heidelberg, 2014.

Srikant, R. and Ying., L. Finite-time error bounds for linear stochastic approximation and TD learning. arXiv preprint arXiv:1902.00923, 2019.

Sutton, R. S. Learning to predict by the methods of temporal differences. Machine learning, 3(1): 9–44, 1988.

Sutton, R. S., Maei, H. R., Precup, D., Bhatnagar, S., Silver, D., Szepesv´ari, C., and Wiewiora, E. Fast gradient-descent methods for temporal-difference learning with linear function approximation. In Proceedings of the 26th Annual International Conference on Machine Learning, pp. 993–1000, 2009a.

Sutton, R. S., Maei, H. R., and Szepesv´ari, C. A convergent O(n) temporal-difference algorithm for off-policy learning with linear function approximation. In Advances in neural information processing systems, pp. 1609–1616, 2009b.

Tsitsiklis, J. N. and Van Roy, B. An analysis of temporal-difference learning with function approx- imation. IEEE Transactions on Automatic Control, 42(5):674–690, 1997.

Van Hasselt, H., Guez, A., and Silver, D. Deep reinforcement learning with double Q-learning. In AAAI, volume 2, pp. 5. Phoenix, AZ, 2016.

Wang, Z., Schaul, T., Hessel, M., Hasselt, H., Lanctot, M., and Freitas, N. Dueling network architectures for deep reinforcement learning. In International Conference on Machine Learning, pp. 1995–2003, 2016.

Watkins, C. J. C. H. and Dayan, P. Q-learning. Machine learning, 8(3-4):279–292, 1992.

Yu, H. and Bertsekas, D. P. Convergence results for some temporal difference methods based on least squares. IEEE Transactions on Automatic Control, 54(7):1515–1531, 2009.

A Proof of Theorem 1

The proof is based on the analysis of the general stochastic recursion

image

where h is a mapping  h : Rn → Rn.If only the asymptotic convergence is our concern, the ODE (ordinary differential equation) approach [Bhatnagar et al., 2012] is a convenient tool. Before starting the main proof, we review essential knowledge of the linear system theory [Chen, 1995].

Definition 1 (Chen [1995, Definition 5.1]). The ODE, ˙x(t) = Ax(t), t ≥ 0, where A ∈ Rn×n andx(t) ∈ Rn, is asymptotically stable if for every finite initial state  x(0) = x0, x(t) → 0 as t → ∞.

Definition 2 (Hurwitz matrix). A complex square matrix  A ∈ Cn×n is Hurwitz if all eigenvalues of A have strictly negative real parts.

Lemma 1 (Chen [1995, Theorem 5.4]). The ODE, ˙x(t) = Ax(t), t ≥ 0, is asymptotically stable if and only if A is Hurwitz.

Lemma 2 (Lyapunov theorem [Chen, 1995, Theorem 5.5]). A complex square matrix  A ∈ Cn×n isHurwitz if and only if there exists a positive definite matrix  M = MH ≻ 0 such that AHM +MA ≺0, where  AH is the complex conjugate transpose of A.

Lemma 3 (Schur complement [Boyd & Vandenberghe, 2004, pp. 651]). For any complex block

image

Convergence of many RL algorithms rely on the ODE approaches [Bhatnagar et al., 2012]. One of the most popular approach is based on the Borkar and Meyn theorem [Bhatnagar et al., 2012, Appendix D]. Basic technical assumptions are given below.

Assumption 4.

1. The mapping  h : Rn → Rn is globally Lipschitz continuous and there exists a function  h∞ :Rn → Rn such that

image

2. The origin in  Rn is an asymptotically stable equilibrium for the ODE ˙θ(t) = h∞(θ(t)).

3. There exists a unique globally asymptotically stable equilibrium  θe ∈ Rn for the ODE ˙θ(t) =h(θ(t)), i.e., θ(t) → θe as t → ∞.

4. The sequence  {εk, Gk, k ≥ 1} with Gk = σ(θi, εi, i ≤ k)is a Martingale difference sequence. In addition, there exists a constant  C0 < ∞such that for any initial  θ0 ∈ Rn, we haveE[∥εk+1∥2|Gk] ≤ C0(1 + ∥θk∥2), ∀k ≥ 0.

5. The step-sizes satisfy (4).

Lemma 4 (Borkar and Meyn theorem). Suppose that Assumption 4 holds. For any initial  θ0 ∈ Rn,supk≥0 ∥θk∥ < ∞with probability one. In addition,  θk → θe as k → ∞with probability one.

Based on the technical results, we are in position to prove Theorem 1.

Proof of Theorem 1: The ODE (3) can be expressed as the linear system with an affine term

image

where

image

Therefore, the mapping  h : Rn → Rn, defined by  h(¯θ) = A¯θ + b, is globally Lipschitz continuous. Moreover, we have

image

Therefore, the first condition in Assumption 4 holds. To meet the second condition of Assump- tion 4, by Lemma 1, it suffices to prove that A is Hurwitz. The reason is explained below. Suppose that A is Hurwitz. If A is Hurwitz, it is invertible, and there exists a unique equilibrium ¯θe ∈ Rnfor the ODE ˙¯θ = A¯θ + bsuch that 0 =  A¯θe + b, i.e., ¯θe = −A−1b. Due to the constant term b, it is not clear if such equilibrium point, ¯θe, is globally asymptotically stable. From [Antsaklis & Michel, 2007, pp. 143], by letting  x = ¯θ − ¯θe, the ODE can be transformed to ˙x = Ax, where the origin is the globally asymptotically stable equilibrium point since A is Hurwitz. Therefore, ¯θe is globally asymptotically stable equilibrium point of ˙¯θ = A¯θ + b, and the third condition of Assumption 4 is satisfied. Therefore, it remains to prove that A is Hurwitz. We first provide a simple analysis and prove that there exists a  δ∗ >0 such that for all  δ ≥ δ∗, Ais Hurwitz. To this end, we use the property of the similarity transformation [Antsaklis & Michel, 2007, pp. 88], i.e., A is Hurwitz if and only if  BAB−1 is Hurwitz for any invertible matrix  B. Letting B =� I 0

image

To prove that  BAB−1 is Hurwitz, we use Lemma 2 with M = I and check the sufficient condition

image

=�ΦT D(−I + γP π)Φ + ΦT (−I + γP π)T DΦ −ΦT (−I + γP π)T DΦ + γΦT DP πΦ−ΦT D(−I + γP π)Φ + γΦT (P π)T DΦ −2δI − γΦT DP πΦ − γΦT (P π)T DΦ�

image

To check the above matrix inequality, note that ΦT D(γP π − I)Φ is negative definite [Bertsekas & Tsitsiklis, 1996, Lemma 6.6, pp. 300]. By using the Schur complement Lemma 3, (7) holds if and only if

image

The above inequality holds for a sufficiently large  δ, i.e., there exists  δ∗ >0 such that the above inequality holds for all  δ > δ∗. Therefore,  BAB−1 and Aare Hurwitz for all  δ > δ∗. A natural question is whether or not  δ∗ = 0. We prove that this is indeed the case. The proof requires rather more involved analysis.

Claim: A is Hurwitz for all  δ > 0.

image

where �xy �∈ C2n is an eigenvector and  λ ∈ Cis an eigenvalue of A. Equivalently, it is written by

image

For any  δ >0, the complex number in the above equation

image

where  λ∗ is the complex conjugate of  λ ∈ C and | · |is the absolute value of a complex number (·), has the absolute value less than or equal to 1, i.e.,  |s| = δ|λ+δ| <1. Now, we prove that the complex matrix, ΦT D(−I + sγP π)Φ, is Hurwitz for any  s ∈ C such that |s| ≤1. For any real vector  v ∈ R|S|, we have

image

image

where the first inequality is due to the Cauchy-Schwarz inequality, the second inequality is due to Tsitsiklis & Van Roy [1997, Lemma 1], and the final inequality follows from the fact that  |s| ≤ 1implies  −2 ≤ s + s∗ ≤2. The last result ensures  vT (sγDP π + s∗γ(P π)T D)v ⪯ γ2vT Dv for anyv ∈ R|S|, and equivalently,

image

Multiplying both sides of the above inequality by Φ from the right and its transpose from the left, one gets

image

By Lemma 2 with M = I, we conclude that the complex matrix, ΦT D(−I + sγP π)Φ, isHurwitz for any  s ∈ C such that |s| ≤1. Based on this observation, we return to (11) and conclude that  −ΦT DΦ + γ δδ+λΦT DP πΦ is Hurwitz for any  λ ∈ C. By the definition of a Hurwitz matrix in Definition 2 and the eigenvalue, we conclude that the real part of  λshould be always strictly negative. Therefore, A is Hurwitz for any  δ >0. This completes the proof. ■.

Next, we prove the remaining parts. Since  εk+1can be expressed as an affine map of ¯θk =[¯θk, θ′k]T, it can be easily proved that the fourth condition of Assumption 4 is satisfied. In particular, if we define  mk := �ki=0 εi, then mkis Martingale, and  εkis a Martingale difference sequence. Therefore, the fourth condition is met.

Finally, by Lemma 4, ¯θkconverges to ¯θe such that

image

B Proof of Theorem 2

The ODE corresponding to Algorithm 3 can be expressed as the linear system with an affine term

image

where

image

The proof follows the same lines as the proof of Theorem 1. Therefore, we only prove that A is Hurwitz here. In particular, A can be represented by  A = B + CT BC, where

image

Pre- and post-multiplying the left-hand side of the about inequality by the full rank matrix �I CT �and its transpose, respectively, yields

image

By Lemma 2 with M = I, this implies that A is Hurwitz. This completes the proof.

C Randomized version of D-TD

We consider a randomized version of D-TD in Algorithm 5, which updates either the target or online parameters randomly.

image

We have the convergence result similar to Theorem 2.

Theorem 4. Consider Algorithm 5 and assume that with a fixed policy  π, the Markov chain is ergodic and the step-sizes satisfy (4). Then,  θk → θ∗ and θ′k → θ∗ as k → ∞with probability one.

Proof: The proof is a slight modification of the proof of Theorem 2. The ODE corresponding to Algorithm 5 can be expressed as the linear system with an affine term

image

ΛA is Hurwitz. From the proof of Theorem 2, we know  AT + A ≺0, which is equivalent to (AT Λ)Λ−1 + Λ−1(ΛA) ≺ 0. By Lemma 2 with M = Λ−1, this implies that ΛA is Hurwitz. This completes the proof. ■

D Proof of Theorem 3

Before presenting the proof, we first introduce a deterministic version of P-TD summarized in Al- gorithm 6 in order to make smooth steps forward. For a fixed  θ′k (target variable), the subroutine, GradientDecent, runs gradient descent steps  Lktimes in order to approximately solve the subproblem, arg minθ∈Rn l(θ; θ′k). By the standard results in Bubeck et al. [2015, Theorem 10.3], the gradient descent iterations converge to the optimal solution  θ∗k+1 := arg minθ∈Rn l(θ; θ′k) linearly, the finite iterates reache an approximate solution within a certain error bound  εk. Upon solving the subproblem, the next target variable is replaced with the next online variable.

The overall convergence relies on the fact that approximately solving the subproblem can be interpreted as approximately solving a projected Bellman equation defined below.

Definition 3 (Projected Bellman equation). The projected Bellman equation is defined as

image

where F is the projected Bellman operator defined by

image

Π is the projection onto the range space of Φ, denoted by  R(Φ): Π(x) := arg minx′∈R(Φ) ∥x − x′∥2D.The projection can be performed by the matrix multiplication: we write Π(x) := Πx, where Π :=Φ(ΦT DΦ)−1ΦT D.

By direct calculations, we can conclude that the solution of the projected Bellman equation is not identical to the solution of the value function evaluation problem in (1), while it only approximates the solution of (1). The solution of the projected Bellman equation is denoted by  θ∗, i.e.,

image

Therefore, Algorithm 6 executes an approximate dynamic programming procedure. Based on these observations, the convergence of Algorithm 6 is given below.

Proposition 3. Consider Algorithm 6. We have

image

image

To prove Proposition 3, we first summarize some essential technical lemmas. The first lemma states that the operator F is a contraction.

Lemma 5. The operator  F is a γ-contraction with respect to  ∥ · ∥D, i.e.,

image

Proof: We have

image

where the first inequality is due to the non-expansive mapping property of the projection, and the second inequality is due to Tsitsiklis & Van Roy [1997, Lemma 1]. This completes the proof. ■

Lemma 6. θ∗k+1 inAlgorithm 6 satisfies Φθ∗k+1 = F(Φθ′k).

Proof: The result follows by solving the optimality condition  ∇θl(θ; θ′k) = −ΦT D(Rπ +γP πΦθ′k − Φθ) = 0. In particular, it implies

image

Multiplying both sides by (ΦT DΦ)−1 from the left, we have

image

Again, we multiply both sides by Φ from the left to obtain

image

where Π := Φ(ΦT DΦ)−1ΦT D. This completes the proof. ■

Lemma 7. l(θ; θ′k) := 12∥Rπ + γP πΦθ′k − Φθ∥2D is µ-strongly convex with  µ := λmin(ΦT DΦ).

Proof: Noting that

image

and that ΦT DΦ − λmin(ΦT DΦ)I ⪰0, we conclude that  l(θ; θ′k) − 12∥θ∥22λmin(ΦT DΦ) is convex. Therefore, by the definition of the strongly convex function, the desired conclusion holds. ■

image

where the second equality is due to Lemma 6 and the last inequality is due to Lemma 5. Combining the last inequality over  k = 0, 1, . . . , T −1, one gets the desired result. The last result is obtained by using the Markov inequality. ■

Note that the second term in the inequality of Proposition 3 vanishes as  T → ∞. The firstterms depend on the error incurred at each iteration. In particular, if  εk = ε for all k ≥ 0, then

image

Therefore, we have

image

The remaining error term can vanish if  ε →0, and it can be done by increasing  Lk → ∞.

Finally, the proof of Theorem 3 follows similar lines to the proof of Proposition 3 except for the expectation.

image

where the third inequality is due to  E[�∥θk+1 − θ∗k+1∥22] ≤�E[∥θk+1 − θ∗k+1∥22] ≤ √εk+1. There-fore, we have

image

Combining the last inequality over  k = 0, 1, . . . , T −1, the desired result is obtained. ■

E Proof of Proposition 1

The convergence results in Bottou et al. [2018, Theorem 4.7] can be applied to the procedure SGD of Algorithm 4. We first summarize the results in Bottou et al. [2018]. Consider the optimization problem

image

where  F : Rn → R, let g(θt) be an unbiased i.i.d. stochastic estimation of  ∇θF(θ) at θ = θt, andconsider the stochastic gradient descent method in Algorithm 7.

image

With appropriate assumptions, its convergence can be proved. We first list the assumptions.

Assumption 5. The objective function, F, and SGD Algorithm 7 satisfy the following conditions:

1. F is continuously differentiable and  ∇θFis Lipschitz continuous with Lipschitz constant L < 0, i.e.,  ∥∇F(θ) − ∇F(θ′)∥2 ≤ L∥θ − θ′∥2 for all θ, θ′ ∈ Rn.

2. F is c-strongly convex.

3. The sequence of iterates  {θt}∞t=0 is contained in an open set over which F is bounded below by a scalar  Finf.

4. There exist scalars  µG ≥ µ > 0such that, for all  t ≥ 0,

image

Under Assumption 5, the convergence of iterates of Algorithm 7 in expectation can be proved.

Lemma 8. Under Assumption 5 (with  Finf = F(θ∗)), suppose that the SGD method in Algorithm 7 is run with a stepsize sequence such that, for all  t ≥ 0,

image

for some  β > 1/(cµ) and κ > 0 such that β0 ≤ µ/(L(M + µ2G)). Then, for all  t ≥ 0, the expected optimality gap satisfies

image

where

image

To apply Lemma 8 to SGD of Algorithm 4, we will prove that all the conditions in Assumption 5 are satisfied with  F(θ) = l(θ; θ′k) := 12∥Rπ + γP πΦθ′k − Φθ∥2D. The strong convexity is established in Lemma 7. In the following lemmas, we prove the Lipschitz continuity of the gradient and the remaining conditions in Assumption 5.

Lemma 9 (Lipschitz continuous gradient). F satisfies

image

with L =�λmax(ΦT DΦΦT DΦ).

Proof: Noting that  ∇F(θ) = ΦT D(Rπ + P πΦθk − Φθ), we have

∥∇F(θ) − ∇F(θ′)∥2 =∥ΦT D(Rπ + P πΦθk − Φθ) − ΦT D(Rπ + P πΦθk − Φθ′)∥2=∥ΦT DΦ(θ − θ′)∥2

image

which proves the desired result. ■

Lemma 10. For SGD of Algorithm 4, we have

image

Proof: By the definition of  g(θk,t) in SGD ofAlgorithm 4, we have

image

proving the first equation. For the second result, we have

image

proving the second result. ■

The first result in Lemma 10 implies that  g(θk,t) is an unbiased stochastic estimation of  ∇F(θk,t).The second result in Lemma 10 means that the second moment of the stochastic gradient estimation is bounded by a quantity which is dependent on  ∥θk∥22. Based on Lemma 10, we bound the variance of the gradient in the next lemma. Before proceeding, we introduce an inequality which will be frequently used.

Lemma 11. For any a, b ∈ Rn, we have

image

where  ε > 0is any real number.

Proof: We obtain the first upper bound by

image

for any  ε >0, where the last inequality is due to the Young’s inequality,  |aT b| ≤ ε∥a∥22/2+ε−1∥b∥22/2.Similarly, the lower bound can be obtained by

image

This completes the proof. ■

Lemma 12 (Bounded variance). The variance of the gradient is bounded as follows:

image

where

image

Proof: Using the definition of  V[g(θt)|θk,t, θk] inAssumption 5 and the bound on  E[∥g(θk,t)∥22|θk,t, θk]in Lemma 10, we have

image

for any K > 0. A main issue in (13) is the presence of the term depending on  θk,t. We will obtain a bound on the first two terms which does not depend on  θk,t. To this end, a lower bound on ∥∇F(θk,t)∥22 is obtained as follows:

image

for any  ε >0 such that 1  − ε−1 >0, where the first inequality is due to Lemma 11. Combining the last inequality with (13) yields

image

Note that by appropriately choosing  ε > 0 and K >0, the term related to  ∥θk,t∥22 canbe removed. In particular, we can choose  ε > 0 and K >0 such that 3∥Φ∥42 − (1 + K)(1 −ε−1)λmin(ΦT DΦΦT DΦ) = 0. A solution is

image

and

image

for any  δ >0. Setting  δ= 1 and substituting these expressions for  ε and K in (14) result in

image

image

where

image

Applying Lemma 11 again for  ∥ΦT DRπ + ΦT DP πΦθk∥22 in (15) yields

image

where

image

which is the desired conclusion. ■

We are now ready to prove Proposition 1.

Proof of Proposition 1: The first statement of Proposition 1 is proven in Lemma 12. To prove the remaining conditions of Proposition 1, we note that all the results of this section prove that Assumption 5 is satisfied with  µ = µG = 1, c = λmin(ΦT DΦ), L =�λmax(ΦT DΦΦT DΦ), M =ξ1 + ξ2∥θk∥22, MV = ξ3, and Finf = minθ F(θ), where the positive real numbers  ξ1, ξ2, and ξ3are given in Lemma 12. Then, by using Lemma 8, it can be proved that if the SGD method in Algorithm 4 is run with a stepsize sequence such that, for all  t ≥ 0,

image

for some  β > 1/λmin(ΦT DΦ) and κ >0 such that

image

then, for all 0  ≤ t ≤ Lk −1, the expected optimality gap satisfies

image

with

image

Note that  θ∗k+1 is the solution that minimizes F, which is different from  θ∗. By the definition of the strong convexity, we have

image

image

For later analysis, we will further polish the upper bound. Using  F(θ∗k+1) ≥0 and the triangle inequality, we have

image

where  θ∗ is the solution of the projected Bellman equation, Φθ∗ = F(Φθ∗), that we want to find, and it should not be confused with  θ∗k+1, which is the solution that minimizes F.

Next,  F(θk) is bounded as

image

where the first inequality is due to the inequality (a + b)2 ≤ 2a2 + 2b2. We combine this result with (18) to obtain

image

(κ + 1)λmax((P πΦ − Φ)T D(P πΦ − Φ))∥θk − θ∗∥22 + (γ + 1)∥Rπ + P πΦθ∗ − Φθ∗∥2D�≤β2�λmax(ΦT DΦΦT DΦ)�ξ1 + ξ2∥θ∗∥22 + ξ2∥θk − θ∗∥22�2(βλmin(ΦT DΦ) − 1)

+ (κ + 1)λmax((P πΦ − Φ)T D(P πΦ − Φ))∥θk − θ∗∥22 + (κ + 1)∥Rπ + P πΦθ∗ − Φθ∗∥2D=χ1 + χ2∥θk − θ∗∥22,

where the second inequality is due to the inequality max{a, b} ≤ a + b and

image

F Proof of Proposition 2

To prove the sample complexity, we will make use of Proposition 1. A tricky part is due to the fact that the constant factor of the convergence rate depends on  ∥θk − θ∗∥22. We will prove that ∥θk − θ∗∥22 is bounded by a constant in expectation, which plays a key role in the proof.

Lemma 13. Suppose that Algorithm 4 is run with  εi = ε for all k ≥ i ≥ 1. Then,

image

where

image

Proof: We follow the procedure similar to that of Theorem 3. The main difference relies on the fact that we need a bound on the squared norm. First, we obtain the chain of inequalities

image

where 0  ≤ i ≤ k−1, the first equality is due to Lemma 11, the second equality follows from Lemma 6, and the last inequality is due to Lemma 5. Since  γ ∈ [0,1), there exists  δ >0 such that (1+δ)γ2 < 1,which is equivalent to  δ < 1−γ2γ2. We simply choose  δ = 1−γ22γ2, yielding

image

Taking the total expectation on both sides and using the hypothesis,  E[∥θi+1 − θ∗i+1∥22] ≤ ε forall 0  ≤ i ≤ k − 1, yield

image

By the induction argument in i, we have

image

image

where the second inequality is obtained by letting  i → ∞and the last inequality is due to�γ2+12 �i<1. Since the first term on the right hand side is nonnegative, the last inequality holds for i = 0. By using  E[∥Φθk − Φθ∗∥2D] ≥ λmin(ΦT DΦ)E[∥θk − θ∗∥22] and arranging terms, we arrive at the conclusion. ■

Lemma 13 states that if the subproblems are solved such that  E[∥θi − θ∗i ∥22] ≤ ε for k ≥ i ≥ 1,then  E[∥θi − θ∗∥22] is bounded by a constant depending on  ε for all k ≥ i ≥0. Using this property, we introduce another version of Proposition 1 which drops the dependency of  ∥θk − θ∗∥22.

Proposition 4. Suppose that the SGD method in Algorithm 4 is run with a stepsize sequence such that, for all  t ≥ 0,

image

for some  β > 1/λmin(ΦT DΦ) and κ > 0 such that

image

Moreover, suppose that Algorithm 4 is run with  εi = ε for all k − 1 ≥ i ≥ 1. Then, for all 0  ≤ t ≤ Lk − 1, the expected optimality gap satisfies

image

Proof: The proof is completed by taking the total expectation on both sides of (6) in Propo- sition 1 and using the bound in Lemma 13. ■

From Proposition 4, we concludes that with the number of subproblem iterations such that

image

each subproblem achieves  ε-optimality in expectation. Based on this observation, we will now prove the sample complexity. By Theorem 3,  E[∥θT+1 − θ∗∥D] ≤ ϵ holds if

image

Again, it holds if

image

and

image

for any real numbers a, b > 0 such that a + b = 1. The condition (21) holds if

image

The condition (20) holds if

image

Combined with Proposition 1 and Lemma 13, a sufficient condition of (23) is

image

Using the upper bound in (23) and arranging terms, we have that the condition (23) (and hance (20)) holds if the number of iteration,  Lk, for the subproblem at iteration k is lower bounded by

image

Combining (22) and (24), we conclude that (19) holds with SO calls at most

image

To simplify the expression, a > 0 and b > 0 are set to be  a =�maxs∈S d(s) and b = 1 −�maxs∈S d(s), respectively. Plugging the explicit expressions for  ω1, ω2 in Lemma 13and further simplifications lead to the desired conclusion.

G Additional Simulations for Section 6

We consider the same MDP as in Section 6 with a linear function approximation using the feature vector

image

Figure 4(a) depicts the error evolution of the standard TD-learning with different step-sizes,  αk =α/(k + 10000), α = 1000,4000, by which one concludes that the step-size  αk = 1000/(k + 10000)provides reasonable performance. Figure 4(b) illustrates the error evolution of A-TD with step-size αk = 1000/(k+ 10000) and  δ = 0.1, 0.2, 0.5, 0.7, 0.9. From Figure 4(b), we can observe that the smaller the  δ, the slower the convergence rate.

Next, we consider the same MDP as in Section 6 with a linear function approximation using the feature vector

image

image

Figure 4: (a) Errors of the standard TD-learning with different step-sizes,  αk = α/(k+10000), α =1000, 4000. (b) Errors of A-TD with step-size  αk = 1000/(k+ 10000) and  δ = 0.1, 0.2, 0.5, 0.7, 0.9.The shaded areas depict empirical variances obtained with several realizations.

Simulation results (error evolution) for the standard TD are given in Figure 5(a) with different step-sizes  αk = α/(10000 + k) and α = 1000, 2000, . . . , 10000.Moreover, simulation results of P-TD are given in Figure 5(b) with  Lk = 5, 10, 20, 40, 80, 160,320, where the following different step-sizes are used:  βt = 4000/(10000 + t) in Figure 5(b), βt = 6000/(10000 + t) in Figure 5(c),βt = 8000/(10000 + t) in Figure 5(d).

Figure 6 illustrates the error plots of P-TD for step-sizes,  βt = β/(10000+t), β = 1000, 2000, . . . , 8000and different  Lk = 10 (Figure 6(a)), Lk = 20 (Figure 6(b)), and 40 (Figure 6(c)).

From Figure 4 and Figure 6, one observes that the error evolution for  β= 8000 has large fluctuations.

image

Figure 5: (a) Error evolution of the standard TD-learning with different step-sizes,  αk = α/(k +10000),  α = 1000, 2000, . . . ,10000. Error evolution of P-TD with  Lk = 5, 10, 20, 40, 80, 160, 320 andthe step-sizes, (b)  βt = 4000/(10000 + t), (c) βt = 6000/(10000 + t), (d) βt = 8000/(10000 + t).The shaded areas depict empirical variances obtained with several realizations.

image

Figure 6: Error evolution of P-TD with different step-sizes,  βt = β/(10000 + t), β =1000, 2000, . . . , 8000. Each subplot uses different  Lk: (a) Lk = 10; (b) Lk = 20; (c) Lk = 40.The shaded areas depict empirical variances obtained with several realizations.


Designed for Accessibility and to further Open Science