This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Asymptotic and Finite Sample Analysis of
Nonexpansive Stochastic Approximations with Markovian Noise

Ethan Blaser    Shangtong Zhang
Abstract

Stochastic approximation is an important class of algorithms, and a large body of previous analysis focuses on stochastic approximations driven by contractive operators, which is not applicable in some important reinforcement learning settings. This work instead investigates stochastic approximations with merely nonexpansive operators. In particular, we study nonexpansive stochastic approximations with Markovian noise, providing both asymptotic and finite sample analysis. Key to our analysis are a few novel bounds of noise terms resulting from the Poisson equation. As an application, we prove, for the first time, that the classical tabular average reward temporal difference learning converges to a sample path dependent fixed point.

Machine Learning, ICML

1 Introduction

Stochastic approximation (SA) algorithms (Robbins & Monro, 1951; Kushner & Yin, 2003; Borkar, 2009) form the foundation of many iterative optimization and learning methods by updating a vector incrementally and stochastically. Prominent examples include stochastic gradient descent (SGD) (Kiefer & Wolfowitz, 1952) and temporal difference (TD) learning (Sutton, 1988). These algorithms generate a sequence of iterates {xn}\quantity{x_{n}} starting from an initial point x0dx_{0}\in\mathbb{R}^{d} through the recursive update:

xn+1xn+αn+1(H(xn,Yn+1)xn)x_{n+1}\doteq x_{n}+\alpha_{n+1}\quantity(H\quantity(x_{n},Y_{n+1})-x_{n}) (SA)

where {αn}\left\{\alpha_{n}\right\} is a sequence of deterministic learning rates, {Yn}\{Y_{n}\} is a sequence of random noise in a space 𝒴\mathcal{Y}, and a function H:d×𝒴dH:\mathbb{R}^{d}\times\mathcal{Y}\rightarrow\mathbb{R}^{d} maps the current iterate xnx_{n} and noise Yn+1Y_{n+1} to the actual incremental update. We use hh to denote the expected update, i.e., h(x)𝔼[H(x,y)]h(x)\doteq\mathbb{E}\quantity[H(x,y)], where the expectation will be formally defined shortly.

Despite the foundational role of SA in analyzing reinforcement learning (RL) (Sutton & Barto, 2018) algorithms, most of the existing literature assumes that the expected mapping hh is a contraction, ensuring the stability and convergence of the iterates {xn}\quantity{x_{n}} under mild conditions. Table 1 highlights the relative scarcity of results concerning nonexpansive mappings. However, in many problems in RL, particularly those involving average reward formulations (Tsitsiklis & Roy, 1999; Puterman, 2014; Wan et al., 2021b, a; He et al., 2022), hh is only guaranteed to be non-expansive, not contractive.

Table 1: Overview of stochastic approximation methods, with a focus on those that consider non-expansive mappings. “Non-expansive hh” refers to works where the expected mapping is non-expansive, as opposed to strictly a contraction. “Markovian {Yn}\quantity{Y_{n}}” indicates cases where the noise term {Yn}\quantity{Y_{n}} is Markovian. “Asymptotic” refers to works that prove almost sure convergence, which is not necessarily weaker than non-asymptotic convergence results. Note that we present only a representative subset of results for SA with contractive mappings due to an abundance of literature in the area. For a more comprehensive treatment, see (Benveniste et al., 1990; Kushner & Yin, 2003; Borkar, 2009).
Nonexpansive hh Markovian {Yn}\{Y_{n}\} Asymptotic Non-Asymptotic
Krasnosel’skii (1955)
Ishikawa (1976)
Reich (1979)
Benveniste et al. (1990)
Liu (1995)
Szepesvári (1997)
Abounadi et al. (2002)
Tadić (2002)
Kushner & Yin (2003)
Koval & Schwabe (2003)
Tadic (2004)
Kim & Xu (2007)
Borkar (2009)
Cominetti et al. (2014)
Bravo et al. (2019)
Chen et al. (2021)
Borkar et al. (2021)
Karandikar & Vidyasagar (2024)
Bravo & Cominetti (2024)
Qian et al. (2024)
Liu et al. (2025)
Ours

One tool for analyzing (SA) with nonexpansive hh which has recently gained renewed attention, is Krasnoselskii-Mann iterations. In their simplest deterministic form, these iterations are given by:

xt+1=xt+αt+1(Txtxt),\displaystyle x_{t+1}=x_{t}+\alpha_{t+1}(Tx_{t}-x_{t}), (KM)

where T:ddT:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} is some nonexpansive mapping. Under some other restrictive conditions, Krasnosel’skii (1955) first proves the convergence of (KM) to a fixed point of TT and this result is further generalized by Edelstein (1966); Ishikawa (1976); Reich (1979); Liu (1995). More recently, Cominetti et al. (2014) use a novel fox-and-hare model to connect KM iterations with Bernoulli random variables, providing a sharper convergence rate for xkTxk0\norm{x_{k}-Tx_{k}}\to 0.

In practice, algorithms often deviate from (KM) due to noise, leading to the study of inexact KM iterations (IKM) with deterministic noise (Kim & Xu, 2007; Bravo et al., 2019):

xt+1=xt+αt+1(Txtxt+et+1),\displaystyle x_{t+1}=x_{t}+\alpha_{t+1}(Tx_{t}-x_{t}+e_{t+1}), (IKM)

where {et}\quantity{e_{t}} is a sequence of deterministic noise. Bravo et al. (2019) extend Cominetti et al. (2014) and establish the convergence of (IKM), under some mild conditions on {et}\quantity{e_{t}}.

However, deterministic noise is still not desirable in many problems. To this end, a stochastic version of (IKM) is studied, which considers the iterates

xt+1=xt+αt+1(Txtxt+Mt+1),\displaystyle x_{t+1}=x_{t}+\alpha_{t+1}(Tx_{t}-x_{t}+M_{t+1}), (SKM)

where {Mt}\quantity{M_{t}} is a Martingale difference sequence. Under mild conditions, Bravo & Cominetti (2024) proves the almost sure convergence of (SKM) to a fixed point of TT. If we write (SA) as

xn+1=xn+αn+1(h(xn)xn+H(xn,Yn+1)h(xn)),\displaystyle x_{n+1}=x_{n}+\alpha_{n+1}\quantity(h(x_{n})-x_{n}+H\quantity(x_{n},Y_{n+1})-h(x_{n})), (1)

we observe that the convergence result from Bravo & Cominetti (2024) implies the almost sure convergence of (SA) when {Yn}\quantity{Y_{n}} is i.i.d., since this makes {H(xn,Yn+1)h(xn)}\quantity{H(x_{n},Y_{n+1})-h(x_{n})} a Martingale difference sequence.

Bravo & Cominetti (2024) is the first to introduce this SKM based method in RL, by using it to prove the almost sure convergence and non-asymptotic convergence rate of a synchronous version of RVI QQ-learning (Abounadi et al., 2001). However, the assumption that {Yn}\quantity{Y_{n}} is i.i.d only holds for some synchronous RL algorithms. In most practical settings where the RL algorithm is asynchronous, the noise {Yn}\quantity{Y_{n}} is Markovian, meaning {H(xn,Yn+1)h(xn)}\quantity{H(x_{n},Y_{n+1})-h(x_{n})} is not a Martingale difference sequence and the results of Bravo & Cominetti (2024) do not apply.

Contribution

Our primary contribution is to close the aforementioned gap by extending the results of Bravo & Cominetti (2024) to the Markovian noise setting. Namely, this work allows {Yn}\quantity{Y_{n}} to be a Markov chain, and HH to be a 11-Lipschitz continuous noisy estimate of a non-expansive operator hh, providing both the first proof of almost sure convergence, and also the first non-asymptotic convergence rate in this setting (Table 1).

  • Theorem 2.6 proves that the sequence {xn}\quantity{x_{n}} generated by (SA) with Markovian {Yn}\quantity{Y_{n}} and nonexpansive hh, converges almost surely to some random point x𝒳x^{*}\in\mathcal{X}_{*}, where 𝒳\mathcal{X}_{*} is the set of fixed points of hh. Importantly, xx_{*} may depend on the entire sample path.

  • Theorem 3.1 provides the convergence rate of the expected residuals 𝔼[xnh(xn)]\mathbb{E}\quantity[\norm{x_{n}-h(x_{n})}].

  • Theorem 4.2 utilizes our SKM results to provide the first proof of almost sure convergence of tabular average reward temporal difference learning (TD) to a (possibly sample path dependent) fixed point.

By extending Bravo & Cominetti (2024) to Markovian noise, we are the first to use the SKM method to analyze asynchronous RL algorithms.

The key idea of our approach is to use Poisson’s equation to decompose the error {H(xn,Yn+1)h(xn)}\quantity{H(x_{n},Y_{n+1})-h(x_{n})} into boundable error terms (Benveniste et al., 1990). While the use of Poisson’s equation for handling Markovian noise is well-established, our method departs from prior techniques for bounding these error terms in almost sure convergence analyses. Specifically, Benveniste et al. (1990) and Konda & Tsitsiklis (1999) use stopping times, while Borkar et al. (2021) employ a Lyapunov function and use the scaled iterates technique. In contrast, we leverage a 1-Lipschitz continuity assumption on HH to directly control the growth of error terms.

Notations

In this paper, all vectors are column. We use \norm{\cdot} to denote a generic operator norm and use ee to denote an all-one vector. We use 2\norm{\cdot}_{2} and \norm{\cdot}_{\infty} to denote 2\ell_{2} norm and infinity norm respectively. We use 𝒪()\mathcal{O}(\cdot) to hide deterministic constants for simplifying presentation, while the letter ζ\zeta is reserved for sample-path dependent constants.

2 Almost Sure Convergence of Stochastic Krasnoselskii-Mann Iterations with Markovian and Additive Noise

To extend the analysis of (SKM) in Bravo et al. (2019); Bravo & Cominetti (2024) to SKM with Markovian and additive noise, we consider the following iterates

xn+1=xn+αn+1(H(xn,Yn+1)xn+ϵn+1(1)).x_{n+1}=x_{n}+\alpha_{n+1}\left(H(x_{n},Y_{n+1})-x_{n}+{\epsilon}^{{\left(1\right)}}_{n+1}\right). (SKM with Markovian and Additive Noise)

Here, {xn}\quantity{x_{n}} are stochastic vectors evolving in d\mathbb{R}^{d}, {Yn}\quantity{Y_{n}} is a Markov chain evolving in a finite state space 𝒴\mathcal{Y}, H:d×𝒴dH:\mathbb{R}^{d}\times\mathcal{Y}\to\mathbb{R}^{d} defines the update, {ϵn+1(1)}\quantity{{\epsilon}^{{\left(1\right)}}_{n+1}} is a sequence of stochastic noise evolving in d\mathbb{R}^{d}, and {αn}\quantity{\alpha_{n}} is a sequence of deterministic learning rates. Although the primary contribution of this work is to allow {Yn}\quantity{Y_{n}} to be Markovian, we also include the deterministic noise term ϵn(1){\epsilon}^{{\left(1\right)}}_{n} in (SKM with Markovian and Additive Noise), as it will later be instrumental in proving the almost sure convergence of average reward TD in Section 4.

We make the following assumptions.

Assumption 2.1 (Ergodicity).

The Markov chain {Yn}\quantity{Y_{n}} is irreducible and aperiodic.

The Markov chain {Yn}\quantity{Y_{n}} thus adopts a unique invariant distribution, denoted dμd_{\mu}. We use PP to denote the transition matrix of {Yn}\quantity{Y_{n}}.

Assumption 2.2 (1-Lipschitz).

The function HH is 1-Lipschitz continuous in its first argument w.r.t. some operator norm \norm{\cdot} and uniformly in its second argument, i.e., for any x,x,yx,x^{\prime},y, it holds that

H(x,y)H(x,y)xx.\displaystyle\norm{H(x,y)-H(x^{\prime},y)}\leq\norm{x-x^{\prime}}. (2)

This assumption has two important implications. First, it implies that H(x,y)H(x,y) can grow at most linearly. Indeed, let x=0x^{\prime}=0, we get H(x,y)H(0,y)+x\norm{H(x,y)}\leq\norm{H(0,y)}+\norm{x}. Define CHmaxyH(0,y)C_{H}\doteq\max_{y}\norm{H(0,y)}, we get

H(x,y)CH+x.\displaystyle\norm{H(x,y)}\leq C_{H}+\norm{x}. (3)

Second, define the function h:ddh:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} as the expectation of HH over the stationary distribution dμd_{\mu}:

h(x)𝔼ydμ[H(x,y)].\displaystyle h(x)\doteq\mathbb{E}_{y\sim d_{\mu}}[H(x,y)]. (4)

We then have that hh is non-expansive. Namely,

h(x)h(x)\displaystyle\textstyle\norm{h(x)-h(x^{\prime})} ydμ(y)H(x,y)H(x,y)\displaystyle\leq\sum_{y}d_{\mu}(y)\norm{H(x,y)-H(x^{\prime},y)} (5)
xx.\displaystyle\leq\norm{x-x^{\prime}}. (6)

This hh is exactly the non-expansive operator in the SKM literature. We, of course, need to assume that the problem is solvable.

Assumption 2.3 (Fixed Points).

The non-expansive operator hh adopts at least one fixed point.

We use 𝒳\mathcal{X}_{*}\neq\emptyset to denote the set of fixed points of hh.

Assumption 2.4 (Learning Rate).

The learning rate {αn}\quantity{\alpha_{n}} has the form

αn=1(n+1)b,α0=0,\displaystyle\textstyle\alpha_{n}=\frac{1}{(n+1)^{b}},\alpha_{0}=0, (7)

where b(45,1]b\in(\frac{4}{5},1].

The primary motivation for requiring b(45,1]b\in(\frac{4}{5},1] is that our learning rates αn\alpha_{n} need to decrease quickly enough for certain key terms in the proof to be finite. The specific need for b>45b>\frac{4}{5} can be seen in the proof of (79) in Lemma B.1.

Next, using this definition of the learning rates, we will define two useful shorthands,

αk,n\displaystyle\alpha_{k,n} αkj=k+1n(1αj),αn,nαn,\displaystyle\doteq\alpha_{k}\prod_{j=k+1}^{n}{\left(1-\alpha_{j}\right)},\,\alpha_{n,n}\doteq\alpha_{n}, (8)
τn\displaystyle\tau_{n} k=1nαk(1αk).\displaystyle\doteq\sum_{k=1}^{n}\alpha_{k}{\left(1-\alpha_{k}\right)}. (9)

We now impose assumptions on the additive noise.

Assumption 2.5 (Additive Noise).
k=1αkϵk(1)<\displaystyle\textstyle\sum_{k=1}^{\infty}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}<  a.s.,\displaystyle\infty\mbox{\quad a.s.,\quad} (10)
𝔼[ϵn(1)2]=\displaystyle\textstyle\mathbb{E}\left[\norm{{\epsilon}^{{\left(1\right)}}_{n}}^{2}\right]= 𝒪(1n).\displaystyle\textstyle\mathcal{O}\quantity(\frac{1}{n}). (11)

The first part of Assumption 2.5 can be interpreted as a requirement that the total amount of additive noise remains finite, akin to the assumption on ete_{t} in (IKM) in Bravo et al. (2019). Additionally, we impose a condition on the second moment of this noise, requiring it to converge at the rate 𝒪(1n)\mathcal{O}\quantity(\frac{1}{n}). While these assumptions on ϵn(1){{\epsilon}^{{\left(1\right)}}_{n}} may seem restrictive, it should be noted that even if ϵn(1){\epsilon}^{{\left(1\right)}}_{n} were absent, our work would still extend the results of (Bravo & Cominetti, 2024) to cases involving Markovian noise, as the Markovian noise component is already incorporated in YnY_{n}, which represents a significant result. For most RL applications involving algorithms which have only one set of weights, the additional noise ϵk(1){\epsilon}^{{\left(1\right)}}_{k} will simply be 0. We are now ready to present the main convergence result.

Theorem 2.6.

Let Assumptions 2.1 - 2.5 hold. Then the iterates {xn}\quantity{x_{n}} generated by 
(SKM with Markovian and Additive Noise) satisfy

limnxn=x a.s.,\displaystyle\lim_{n\to\infty}x_{n}=x_{*}\mbox{\quad a.s.,\quad} (12)

where x𝒳x_{*}\in\mathcal{X}_{*} is a possibly sample-path dependent fixed point. Or more precisely speaking, let ω\omega denote a sample path (w0,Y0,Y1,)(w_{0},Y_{0},Y_{1},\dots) and write xn(ω)x_{n}(\omega) to emphasize the dependence of xnx_{n} on ω\omega. Then there exists a set Ω\Omega of sample paths with Pr(Ω)=1\Pr(\Omega)=1 such that for any ωΩ\omega\in\Omega, the limit limnxn(ω)\lim_{n\to\infty}x_{n}(\omega) exists, denoted as x(ω)x_{*}(\omega), and satisfies x(ω)𝒳x_{*}(\omega)\in\mathcal{X}_{*}.

Proof.

We start with a decomposition of the error H(x,Yn+1)h(x)H(x,Y_{n+1})-h(x) using Poisson’s equation akin to Métivier & Priouret (1987); Benveniste et al. (1990). Namely, thanks to the finiteness of 𝒴\mathcal{Y}, it is well known (see, e.g., Theorem 17.4.2 of Meyn & Tweedie (2012) or Theorem 8.2.6 of Puterman (2014)) that there exists a function ν(x,y):d×𝒴d\nu(x,y):\mathbb{R}^{d}\times\mathcal{Y}\to\mathbb{R}^{d} such that

H(x,y)h(x)=ν(x,y)(Pν)(x,y).\displaystyle H(x,y)-h(x)=\nu(x,y)-(P\nu)(x,y). (13)

Here, we use PνP\nu to denote the function (x,y)yP(y,y)ν(x,y)(x,y)\mapsto\sum_{y^{\prime}}P(y,y^{\prime})\nu(x,y^{\prime}). The error can then be decomposed as

H(x,Yn+1)h(x)=Mn+1+ϵn+1(2)+ϵn+1(3),\displaystyle H(x,Y_{n+1})-h(x)=M_{n+1}+{\epsilon}^{{\left(2\right)}}_{n+1}+{\epsilon}^{{\left(3\right)}}_{n+1}, (14)

where

Mn+1\displaystyle M_{n+1} ν(xn,Yn+2)(Pν)(xn,Yn+1),\displaystyle\doteq\nu(x_{n},Y_{n+2})-(P\nu)(x_{n},Y_{n+1}), (15)
ϵn+1(2)\displaystyle{\epsilon}^{{\left(2\right)}}_{n+1} ν(xn,Yn+1)ν(xn+1,Yn+2),\displaystyle\doteq\nu{\left(x_{n},Y_{n+1}\right)}-\nu{\left(x_{n+1},Y_{n+2}\right)}, (16)
ϵn+1(3)\displaystyle{\epsilon}^{{\left(3\right)}}_{n+1} ν(xn+1,Yn+2)ν(xn,Yn+2).\displaystyle\doteq\nu{\left(x_{n+1},Y_{n+2}\right)}-\nu{\left(x_{n},Y_{n+2}\right)}. (17)

Here {Mn+1}\quantity{M_{n+1}} is a Martingale difference sequence. We then use

ξn+1\displaystyle\xi_{n+1} ϵn+1(1)+ϵn+1(2)+ϵn+1(3),\displaystyle\doteq{\epsilon}^{{\left(1\right)}}_{n+1}+{\epsilon}^{{\left(2\right)}}_{n+1}+{\epsilon}^{{\left(3\right)}}_{n+1}, (18)

to denote all the non-Martingale noise, yielding

xn+1\displaystyle x_{n+1} =(1αn+1)xn+αn+1(h(xn)+Mn+1+ξn+1).\displaystyle={\left(1-\alpha_{n+1}\right)}x_{n}+\alpha_{n+1}{\left(h{\left(x_{n}\right)}+M_{n+1}+\xi_{n+1}\right)}. (19)

We now define an auxiliary sequence {Un}\quantity{U_{n}} to capture how the noise evolves

U0\displaystyle U_{0}\doteq  0,\displaystyle\,0, (20)
Un+1\displaystyle U_{n+1}\doteq (1αn+1)Un+αn+1(Mn+1+ξn+1).\displaystyle\,{\left(1-\alpha_{n+1}\right)}U_{n}+\alpha_{n+1}{\left(M_{n+1}+\xi_{n+1}\right)}. (21)

If we are able to prove that the total noise is well controlled in the following sense

k=1αkUk1\displaystyle\sum_{k=1}^{\infty}\alpha_{k}\norm{U_{k-1}} < a.s.,\displaystyle<\infty\mbox{\quad a.s.,\quad} (22)
limnUn\displaystyle\lim_{n\rightarrow\infty}\norm{U_{n}} =0 a.s.,\displaystyle=0\mbox{\quad a.s.,\quad} (23)

then a result from Bravo & Cominetti (2024) concerning the convergence of (IKM) can be applied on each sample path to complete the almost sure convergence proof. The rest of the proof is dedicated to the verification of those two conditions.

Telescoping (21) yields

Un=\displaystyle U_{n}= k=1nαk,nMkM¯n+k=1nαk,nϵk(1)ϵ¯n(1)+\displaystyle\underbrace{\sum_{k=1}^{n}\alpha_{k,n}M_{k}}_{{\overline{M}}_{n}}+\underbrace{\sum_{k=1}^{n}\alpha_{k,n}{\epsilon}^{{\left(1\right)}}_{k}}_{{\overline{\epsilon}}^{{\left(1\right)}}_{n}}+ (24)
k=1nαk,nϵk(2)ϵ¯n(2)+k=1nαk,nϵk(3)ϵ¯n(3).\displaystyle\quad\underbrace{\sum_{k=1}^{n}\alpha_{k,n}{\epsilon}^{{\left(2\right)}}_{k}}_{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}+\underbrace{\sum_{k=1}^{n}\alpha_{k,n}{\epsilon}^{{\left(3\right)}}_{k}}_{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}. (25)

Then, we can upper-bound (22) as

k=1nαkUk1\displaystyle\sum_{k=1}^{n}\alpha_{k}\norm{U_{k-1}} k=1nαkM¯k1M¯¯n+k=1nαkϵ¯k1(1)ϵ¯¯n(1)\displaystyle\leq\underbrace{\sum_{k=1}^{n}\alpha_{k}\norm{{\overline{M}}_{k-1}}}_{{\overline{\overline{M}}}_{n}}+\underbrace{\sum_{k=1}^{n}\alpha_{k}\norm{{\overline{\epsilon}}^{{\left(1\right)}}_{k-1}}}_{{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}} (26)
+k=1nαkϵ¯k1(2)ϵ¯¯n(2)+k=1nαkϵ¯k1(3)ϵ¯¯n(3).\displaystyle+\underbrace{\sum_{k=1}^{n}\alpha_{k}\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{k-1}}}_{{\overline{\overline{\epsilon}}}^{{\left(2\right)}}_{n}}+\underbrace{\sum_{k=1}^{n}\alpha_{k}\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{k-1}}}_{{\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n}}. (27)

Lemmas B.8B.9, and B.10 respectively prove that M¯¯n,ϵ¯¯n(1),{\overline{\overline{M}}}_{n},{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}, and ϵ¯¯n(3){\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n} in (27) are bounded almost surely. We bound the remaining term ϵ¯¯n(2){\overline{\overline{\epsilon}}}^{{\left(2\right)}}_{n} needed to verify (22) here as an example of the novelty in bounding these terms. Starting with the definition of ϵ¯n(2){\overline{\epsilon}}^{{\left(2\right)}}_{n} from (25), we have,

ϵ¯n(2)\displaystyle{\overline{\epsilon}}^{{\left(2\right)}}_{n} =k=1nαk,nϵk(2)\displaystyle=\sum_{k=1}^{n}\alpha_{k,n}{\epsilon}^{{\left(2\right)}}_{k} (28)
=k=1nαk,n(ν(xk,Yk+1)ν(xk1,Yk)),\displaystyle=-\sum_{k=1}^{n}\alpha_{k,n}\quantity(\nu\quantity(x_{k},Y_{k+1})-\nu\quantity(x_{k-1},Y_{k})), (29)
=k=1nαk,nν(xk,Yk+1)αk1,nν(xk1,Yk)\displaystyle=-\sum_{k=1}^{n}\alpha_{k,n}\nu\quantity(x_{k},Y_{k+1})-\alpha_{k-1,n}\nu{\left(x_{k-1},Y_{k}\right)} (30)
+αk1,nν(xk1,Yk)αk,nν(xk1,Yk),\displaystyle\quad\quad+\alpha_{k-1,n}\nu{\left(x_{k-1},Y_{k}\right)}-\alpha_{k,n}\nu{\left(x_{k-1},Y_{k}\right)}, (31)
=αn,nν(xn,Yn+1)\displaystyle=-\alpha_{n,n}\nu{\left(x_{n},Y_{n+1}\right)} (32)
k=1n(αk1,nαk,n)ν(xk1,Yk).\displaystyle\quad\quad-\sum_{k=1}^{n}\quantity(\alpha_{k-1,n}-\alpha_{k,n})\nu\quantity(x_{k-1},Y_{k}). (33)

where the last inequality holds because α00\alpha_{0}\doteq 0. Additionally, since αn,n=αn\alpha_{n,n}=\alpha_{n}, taking the norm gives

ϵ¯n(2)\displaystyle\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}} (34)
αnν(xn,Yn+1)+k=1n|αk1,nαk,n|ν(xk1,Yk),\displaystyle\leq\alpha_{n}\norm{\nu{\left(x_{n},Y_{n+1}\right)}}+\sum_{k=1}^{n}\absolutevalue{\alpha_{k-1,n}-\alpha_{k,n}}\norm{\nu{\left(x_{k-1},Y_{k}\right)}}, (35)
ζB.5(αnτn+k=1n|αk1,nαk,n|τk1),\displaystyle\leq\zeta_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\sum_{k=1}^{n}\left|\alpha_{k-1,n}-\alpha_{k,n}\right|\tau_{k-1}\right)}, (36)
2ζB.5αnτn,\displaystyle\leq 2\zeta_{\ref{lem:v_norm}}\alpha_{n}\tau_{n}, (37)

where the second inequality holds by Lemma B.5, and the last inequality holds because α00\alpha_{0}\doteq 0, and that αi,n\alpha_{i,n} and τi\tau_{i} are monotonically increasing (Lemma A.2).

Then, from the definition of ϵ¯¯n(2){\overline{\overline{\epsilon}}}^{{\left(2\right)}}_{n} in (22), we have

ϵ¯¯n(2)=k=1nαkϵ¯k1(2)2ζB.5k=1nαk2τk.\displaystyle{\overline{\overline{\epsilon}}}^{{\left(2\right)}}_{n}=\sum_{k=1}^{n}\alpha_{k}\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{k-1}}\leq 2\zeta_{\ref{lem:v_norm}}\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}. (38)

where the inequality holds because α00\alpha_{0}\doteq 0 and αk\alpha_{k} is decreasing. Then, by Lemma B.1, we have supnk=1nαk2τk<\sup_{n}\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}<\infty, which when combined with the monotone convergence theorem, proves that limnϵ¯¯n(2)<\lim_{n\rightarrow\infty}{\overline{\overline{\epsilon}}}^{{\left(2\right)}}_{n}<\infty, verifying (22).

We now verify (23). This time, rewrite UnU_{n} as

Un\displaystyle U_{n} =k=1nαkUk1+αk(Mk+ϵk(1)+ϵk(2)+ϵk(3)).\displaystyle=-\sum_{k=1}^{n}\alpha_{k}U_{k-1}+\alpha_{k}{\left(M_{k}+{\epsilon}^{{\left(1\right)}}_{k}+{\epsilon}^{{\left(2\right)}}_{k}+{\epsilon}^{{\left(3\right)}}_{k}\right)}. (39)

Lemma B.11, Assumption 2.5, and Lemmas B.12B.13 prove that supnk=1nαkMk<\sup_{n}\norm{\sum_{k=1}^{n}\alpha_{k}M_{k}}<\infty and supnk=1nαkϵk(j)<\sup_{n}\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(j\right)}}_{k}}<\infty for j{1,2,3}j\in\quantity{1,2,3} respectively.

Together with (25), this means that supnUn<\sup_{n}\norm{U_{n}}<\infty. In other words, we have established the stability of (21). Then, it can be shown (Lemma B.14), using an extension of Theorem 2.1 of Borkar (2009) (Lemma D.7), that {Un}\quantity{U_{n}} converges to the globally asymptotically stable equilibrium of the ODE dU(t)dt=U(t)\derivative{U(t)}{t}=-U(t), which is 0. This verifies (23). Lemma B.15 then invokes a result from Bravo & Cominetti (2024) and completes the proof. ∎

Remark 2.7.

We want to highlight that the technical novelty of our work comes from two sources. The first is that while the use of Poisson’s equation for handling Markovian noise is well-established, including the noise representation in (14), previous works with such error decomposition (e.g., Benveniste et al. (1990); Konda & Tsitsiklis (1999); Borkar et al. (2021)) usually only need to bound terms like kαkϵk(1){\sum_{k}\alpha_{k}{\epsilon}^{{\left(1\right)}}_{k}}. In contrast, our setup requires bounding additional terms such as ϵ¯n(1)=kαk,nϵk(1){\overline{\epsilon}}^{{\left(1\right)}}_{n}=\sum_{k}\alpha_{k,n}{\epsilon}^{{\left(1\right)}}_{k} and ϵ¯¯n(1)=iαiϵ¯k1(1){\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}=\sum_{i}\alpha_{i}\norm{{\overline{\epsilon}}^{{\left(1\right)}}_{k-1}} which appear novel and more challenging. Second, our work extends Theorem 2.1 of Borkar (2009) by relaxing an assumption on the convergence of the deterministic noise term. Instead of requiring the noise to converge to 0, we only require more mild condition on the asymptotic rate of change of this noise term. We believe this extension, detailed in Appendix D, has independent utility beyond this work.

3 Convergence Rate

The previous analysis not only guarantees the almost sure convergence of the iterates, but can also be used to obtain estimates of the expected fixed-point residuals.

Theorem 3.1.

Consider the iteration (SKM with Markovian and Additive Noise) and let Assumptions 2.1 - 2.5 hold. There there exists a constant C3.1C_{\ref{thm:conv_rate}} such that

𝔼[xnh(xn)]C3.1τn={𝒪(1/n1b)if45<b<1,𝒪(1/logn)ifb=1.\displaystyle\mathbb{E}\left[\norm{x_{n}-h{\left(x_{n}\right)}}\right]\leq\frac{C_{\ref{thm:conv_rate}}}{\sqrt{\tau_{n}}}=\begin{cases}\mathcal{O}{\left(1/\sqrt{n^{1-b}}\right)}&\text{if}\ \frac{4}{5}<b<1,\\ \mathcal{O}{\left(1/\sqrt{\log n}\right)}&\text{if}\ b=1.\end{cases} (40)
Proof.

Considering the sequence znxnUnz_{n}\doteq x_{n}-U_{n} we have,

xnh(xn)\displaystyle\norm{x_{n}-h\quantity(x_{n})} znh(zn)+2znxn,\displaystyle\leq\norm{z_{n}-h\quantity(z_{n})}+2\norm{z_{n}-x_{n}}, (41)
=znh(zn)+2Un.\displaystyle=\norm{z_{n}-h\quantity(z_{n})}+2\norm{U_{n}}. (42)

where the inequality holds due to the non-expansivity of hh as proven in (6). Then, our proof of Theorem 2.6 guarantees the conditions under which the znz_{n}’s are bounded. Specifically, we proved in Lemma B.15 that if n=1αkUk1<\sum_{n=1}^{\infty}\alpha_{k}\norm{U_{k-1}}<\infty (22) and Un0\norm{U_{n}}\rightarrow 0 (23) almost surely, then with ek=Uk1e_{k}=U_{k-1}, Lemma A.1 can be invoked to bound znh(zn)\norm{z_{n}-h(z_{n})}. This yields,

xnh(xn)\displaystyle\norm{x_{n}-h{\left(x_{n}\right)}} (43)
ζA.1σ(τn)+k=2n2αkσ(τnτk)Uk1+4Un.\displaystyle\leq\zeta_{\ref{lem:bravo_2.1}}\sigma\quantity(\tau_{n})+\sum_{k=2}^{n}2\alpha_{k}\sigma{\left(\tau_{n}-\tau_{k}\right)}\norm{U_{k-1}}+4\norm{U_{n}}. (44)

for ζA.1=2dist(x0,𝒳)+k=2αkUk1\zeta_{\ref{lem:bravo_2.1}}=2dist(x_{0},\mathcal{X}_{*})+\sum_{k=2}^{\infty}\alpha_{k}\norm{U_{k-1}}. However, ζA.1\zeta_{\ref{lem:bravo_2.1}} is a sample-path dependent constant whose order is unknown, and the random sequence Un\norm{U_{n}} may occasionally become very large. Therefore, we compute the non-asymptotic error bound of the expected residuals 𝔼[xnh(xn)]\mathbb{E}\left[\norm{x_{n}-h(x_{n})}\right], which gives,

𝔼[xnh(xn)]𝔼[ζA.1]σ(τn)R1\displaystyle\mathbb{E}\quantity[\norm{x_{n}-h{\left(x_{n}\right)}}]\leq\underbrace{\mathbb{E}\quantity[\zeta_{\ref{lem:bravo_2.1}}]\sigma\quantity(\tau_{n})}_{R_{1}} (45)
+k=2n2αkσ(τnτk)𝔼[Uk1]R2+4𝔼[Un]R3.\displaystyle\quad+\underbrace{\sum_{k=2}^{n}2\alpha_{k}\sigma\quantity(\tau_{n}-\tau_{k})\mathbb{E}\quantity[\norm{U_{k-1}}]}_{R_{2}}+\underbrace{4\mathbb{E}\quantity[\norm{U_{n}}]}_{R_{3}}. (46)

Recalling that σ(y)min{1,1/πy}\sigma(y)\doteq\min\quantity{1,1/\sqrt{\pi y}}, we can see that if there exists a deterministic constant C3.1C_{\ref{thm:conv_rate}} such that 𝔼[ζA.1]C3.1\mathbb{E}\quantity[\zeta_{\ref{lem:bravo_2.1}}]\leq C_{\ref{thm:conv_rate}}, we obtain that R1=𝒪(1/τn)R_{1}=\mathcal{O}\quantity(1/\sqrt{\tau_{n}}). Therefore, in order to prove the Theorem, it is sufficient to find such a constant C3.1C_{\ref{thm:conv_rate}} such that 𝔼[ζA.1]C3.1\mathbb{E}\quantity[\zeta_{\ref{lem:bravo_2.1}}]\leq C_{\ref{thm:conv_rate}}, and prove that R2R_{2}, and R3R_{3} are also 𝒪(1/τn)\mathcal{O}\quantity(1/\sqrt{\tau_{n}}).

We proceed by first upper-bounding 𝔼[Un]\mathbb{E}\quantity[\norm{U_{n}}]. Taking the expectation of (25), we have,

𝔼[Un]\displaystyle\mathbb{E}\quantity[\norm{U_{n}}] (47)
\displaystyle\leq 𝔼[M¯n]+𝔼[ϵ¯n(1)]+𝔼[ϵ¯n(2)]+𝔼[ϵ¯n(3)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{M}}_{n}}]+\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(1\right)}}_{n}}]+\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}]+\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}] (48)
\displaystyle\leq CC.1τnαn+1+i=1nαi,n𝔼[ϵi(1)]+CC.2αnτn\displaystyle C_{\ref{lem: M rate}}\tau_{n}\sqrt{\alpha_{n+1}}+\sum_{i=1}^{n}\alpha_{i,n}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}]+C_{\ref{lem: e2 rate}}\alpha_{n}\tau_{n} (49)
+CC.3αni=1nαiτi(Corollaries C.1C.2C.3)\displaystyle\quad+C_{\ref{lem: e3 rate}}\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i}\quad\text{(Corollaries \ref{lem: M rate}, \ref{lem: e2 rate}, \ref{lem: e3 rate})} (50)
\displaystyle\doteq ωn\displaystyle\omega_{n} (51)

It can be shown (Lemma C.4) that ωn=𝒪(τnαn+1)\omega_{n}=\mathcal{O}(\tau_{n}\sqrt{\alpha_{n+1}}). Then, to prove 𝔼[ζA.1]C3.1\mathbb{E}\quantity[\zeta_{\ref{lem:bravo_2.1}}]\leq C_{\ref{thm:conv_rate}}, since

k=2αk𝔼[Uk1]k=2αkωk1=𝒪(k=2αk3/2τk1),\displaystyle\sum_{k=2}^{\infty}\alpha_{k}\mathbb{E}\quantity[\norm{U_{k-1}}]\leq\sum_{k=2}^{\infty}\alpha_{k}\omega_{k-1}=\mathcal{O}\quantity(\sum_{k=2}^{\infty}\alpha_{k}^{3/2}\tau_{k-1}), (52)

which converges almost surely by Lemma B.1, there exists a C3.1C_{\ref{thm:conv_rate}} such that 𝔼[ζA.1]=2dist(x0,𝒳)+k=2αk𝔼[Uk1]C3.1\mathbb{E}\quantity[\zeta_{\ref{lem:bravo_2.1}}]=2dist(x_{0},\mathcal{X}_{*})+\sum_{k=2}^{\infty}\alpha_{k}\mathbb{E}\quantity[\norm{U_{k-1}}]\leq C_{\ref{thm:conv_rate}} almost surely.

Additionally, our ωn\omega_{n} is of the same order as the analogous νn\nu_{n} in Theorem 2.10 of Bravo & Cominetti (2024). Therefore, we can invoke Lemma C.5, which is a combination of Theorems 2.11 and 3.1 from Bravo & Cominetti (2024), which proves that R2=𝒪(1/τn)R_{2}=\mathcal{O}\quantity(1/\sqrt{\tau_{n}}). Finally, by (51), we directly have that R3=𝒪(τnαn+1)R_{3}=\mathcal{O}(\tau_{n}\sqrt{\alpha_{n+1}}) which is dominated by R2R_{2} and R1R_{1}. ∎

4 Application in Average Reward Temporal Difference Learning

In this section, we provide the first proof of almost sure convergence to a fixed point for average reward TD in its simplest tabular form. Remarkably, this convergence result has remained unproven for over 25 years despite the algorithm’s fundamental importance and simplicity.

4.1 Reinforcement Learning Background

In reinforcement learning (RL), we consider a Markov Decision Process (MDP; Bellman (1957); Puterman (2014)) with a finite state space 𝒮\mathcal{S}, a finite action space 𝒜\mathcal{A}, a reward function r:𝒮×𝒜r:\mathcal{S}\times\mathcal{A}\to\mathbb{R}, a transition function p:𝒮×𝒮×𝒜[0,1]p:\mathcal{S}\times\mathcal{S}\times\mathcal{A}\to[0,1], an initial distribution p0:𝒮[0,1]p_{0}:\mathcal{S}\to[0,1]. At time step 0, an initial state S0S_{0} is sampled from p0p_{0}. At time tt, given the state StS_{t}, the agent samples an action Atπ(|St)A_{t}\sim\pi(\cdot|S_{t}), where π:𝒜×𝒮[0,1]\pi:\mathcal{A}\times\mathcal{S}\to[0,1] is the policy being followed by the agent. A reward Rt+1r(St,At)R_{t+1}\doteq r(S_{t},A_{t}) is then emitted and the agent proceeds to a successor state St+1p(|St,At)S_{t+1}\sim p(\cdot|S_{t},A_{t}). In the rest of the paper, we will assume the Markov chain {St}\quantity{S_{t}} induced by the policy π\pi is irreducible and thus adopts a unique stationary distribution dμd_{\mu}. The average reward (a.k.a. gain, Puterman (2014)) is defined as

J¯πlimT1Tt=1T𝔼[Rt].\textstyle\bar{J}_{\pi}\doteq\lim_{T\rightarrow\infty}\frac{1}{T}\sum_{t=1}^{T}\mathbb{E}\left[R_{t}\right]. (53)

Correspondingly, the differential value function (a.k.a. bias, Puterman (2014)) is defined as

vπ(s)limT1Tτ=1T𝔼[i=1τ(Rt+iJ¯π)St=s].\displaystyle\textstyle v_{\pi}(s)\doteq\lim_{T\to\infty}\frac{1}{T}\sum_{\tau=1}^{T}\mathbb{E}\left[\sum_{i=1}^{\tau}(R_{t+i}-\bar{J}_{\pi})\mid S_{t}=s\right]. (54)

The corresponding Bellman equation (a.k.a. Poisson’s equation) is then

v=rπJ¯πe+Pπv,\displaystyle v=r_{\pi}-\bar{J}_{\pi}e+P_{\pi}v, (55)

where v|𝒮|v\in\mathbb{R}^{|\mathcal{S}|} is the free variable, rπ|𝒮|r_{\pi}\in\mathbb{R}^{|\mathcal{S}|} is the reward vector induced by the policy π\pi, i.e., rπ(s)aπ(a|s)r(s,a)r_{\pi}(s)\doteq\sum_{a}\pi(a|s)r(s,a), and Pπ|𝒮|×|𝒮|P_{\pi}\in\mathbb{R}^{{|\mathcal{S}|}\times{|\mathcal{S}|}} is the transition matrix induced by the policy π\pi, i.e., Pπ(s,s)π(a|s)p(s|s,a)P_{\pi}(s,s^{\prime})\doteq\pi(a|s)p(s^{\prime}|s,a). It is known (Puterman, 2014) that all solutions to (55) form a set

𝒱{vπ+cec}.\displaystyle\mathcal{V}_{*}\doteq\quantity{v_{\pi}+ce\mid c\in\mathbb{R}}. (56)

The policy evaluation problem in average reward MDPs is to estimate vπv_{\pi}, perhaps up to a constant offset cece.

4.2 Average Reward Temporal Difference Learning

Temporal Difference learning (TD; Sutton (1988)) is a foundational algorithm in RL (Sutton & Barto, 2018). Inspired by its success in the discounted setting, Tsitsiklis & Roy (1999) proposed using the update rule (Average Reward TD) to estimate vπv_{\pi} (up to a constant offset) for average reward MDPs. The updates are given by:

Jt+1\displaystyle J_{t+1} =Jt+βt+1(Rt+1Jt),\displaystyle=J_{t}+\beta_{t+1}(R_{t+1}-J_{t}), (Average Reward TD)
vt+1(St)\displaystyle v_{t+1}(S_{t}) =vt(St)+αt+1(Rt+1Jt+vt(St+1)vt(St)),\displaystyle=v_{t}(S_{t})+\alpha_{t+1}\big{(}R_{t+1}-J_{t}+v_{t}(S_{t+1})-v_{t}(S_{t})\big{)}, (57)

where {S0,R1,S1,}\{S_{0},R_{1},S_{1},\dots\} is a trajectory of states and rewards from an MDP under a fixed policy in a finite state space 𝒮\mathcal{S}, JtJ_{t}\in\mathbb{R} is the scalar estimate of the average reward J¯π\bar{J}_{\pi}, vt|𝒮|v_{t}\in\mathbb{R}^{|\mathcal{S}|} is the tabular value estimate, and {αt,βt}\{\alpha_{t},\beta_{t}\} are learning rates.

To utilize Theorem 2.6 to prove the almost sure convergence of  (Average Reward TD), we first rewrite it in a compact form to match that of (SKM with Markovian and Additive Noise). Define the augmented Markov chain Yt+1(St,At,St+1)Y_{t+1}\doteq(S_{t},A_{t},S_{t+1}). It is easy to see that {Yt}\quantity{Y_{t}} evolves in the finite space 𝒴{(s,a,s)π(a|s)>0,p(s|s,a)>0}\mathcal{Y}\doteq\quantity{(s,a,s^{\prime})\mid\pi(a|s)>0,p(s^{\prime}|s,a)>0}. We then define a function H:|𝒮|×𝒴|𝒮|H:\mathbb{R}^{|\mathcal{S}|}\times\mathcal{Y}\to\mathbb{R}^{|\mathcal{S}|} by defining the ss-th element of H(v,(s0,a0,s1))H(v,(s_{0},a_{0},s_{1})) as

H(v,(s0,a0,s1))[s]\displaystyle H(v,(s_{0},a_{0},s_{1}))[s]\doteq (58)
𝕀{s=s0}(r(s0,a0)J¯π+v(s1)v(s0))+v(s).\displaystyle\quad\mathbb{I}\quantity{s=s_{0}}(r(s_{0},a_{0})-\bar{J}_{\pi}+v(s_{1})-v(s_{0}))+v(s). (59)

Then, the update to {vt}\quantity{v_{t}} in (Average Reward TD) can then be expressed as

vt+1=vt+αt+1(H(vt,Yt+1)vt+ϵt+1).\displaystyle v_{t+1}=v_{t}+\alpha_{t+1}\quantity(H(v_{t},Y_{t+1})-v_{t}+\epsilon_{t+1}). (60)

Here, ϵt+1|𝒮|\epsilon_{t+1}\in\mathbb{R}^{|\mathcal{S}|} is the random noise vector defined as ϵt+1(s)𝕀{s=St}(JtJ¯π)\epsilon_{t+1}(s)\doteq\mathbb{I}\quantity{s=S_{t}}(J_{t}-\bar{J}_{\pi}). This ϵt+1\epsilon_{t+1} is the current estimate error of the average reward estimator JtJ_{t}. Intuitively, the indicator 𝕀{s=St}\mathbb{I}\quantity{s=S_{t}} reflects the asynchronous nature of (Average Reward TD). For each tt, only the StS_{t}-indexed element in vtv_{t} is updated.

We are now ready to prove the convergence of (Average Reward TD). Throughout the rest of the section, we utilize the following assumption.

Assumption 4.1 (Ergodicity).

Both 𝒮\mathcal{S} and 𝒜\mathcal{A} are finite. The Markov chain {St}\quantity{S_{t}} induced by the policy π\pi is aperiodic and irreducible.

Theorem 4.2.

Let Assumption 4.1 hold. Consider the learning rates in the form of αt=1(t+1)b,βt=1t\alpha_{t}=\frac{1}{(t+1)^{b}},\beta_{t}=\frac{1}{t} with b(45,1]b\in(\frac{4}{5},1]. Then the iterates {vt}\quantity{v_{t}} generated by (Average Reward TD) satisfy

limtvt=v a.s.,\displaystyle\lim_{t\to\infty}{v_{t}}=v_{*}\mbox{\quad a.s.,\quad} (61)

where v𝒱v_{*}\in\mathcal{V}_{*} is a possibly sample-path dependent fixed point.

Proof.

We proceed via verifying assumptions of Theorem 2.6. In particular, we consider the compact form (60).

Under Assumption 4.1, it is obvious that {Yt}\quantity{Y_{t}} is irreducible and aperiodic and adopts a unique stationary distribution.

To verify Assumption 2.2, we demonstrate that HH is 11-Lipschitz in vv w.r.t \norm{\cdot}_{\infty}. For notation simplicity, let y=(s0,a0,s1)y=(s_{0},a_{0},s_{1}). We have,

H(v,y)[s]H(v,y)[s]=\displaystyle H(v,y)[s]-H(v^{\prime},y)[s]= (62)
𝕀{s=s0}(v(s1)v(s1)v(s0)+v(s0))+v(s)v(s).\displaystyle\,\mathbb{I}\quantity{s=s_{0}}(v(s_{1})-v^{\prime}(s_{1})-v(s_{0})+v^{\prime}(s_{0}))+v(s)-v^{\prime}(s). (63)

Separating cases based on ss, if ss0s\neq s_{0}, we have

|H(v,y)[s]H(v,y)[s]|=|v(s)v(s)|vv.\absolutevalue{H(v,y)[s]-H(v^{\prime},y)[s]}=\absolutevalue{v(s)-v^{\prime}(s)}\leq\norm{v-v^{\prime}}_{\infty}. (64)

For the case when s=s0s=s_{0}, we have

|H(v,y)[s]H(v,y)[s]|=|v(s1)v(s1)|vv.\absolutevalue{H(v,y)[s]-H(v^{\prime},y)[s]}=\absolutevalue{v(s_{1})-v^{\prime}(s_{1})}\leq\norm{v-v^{\prime}}_{\infty}. (65)

Therefore

H(v,y)H(v,y)\displaystyle\norm{H(v,y)-H(v,y)}_{\infty} =maxs𝒮|H(v,y)[s]H(v,y)[s]|\displaystyle=\max_{s\in\mathcal{S}}\absolutevalue{H(v,y)[s]-H(v^{\prime},y)[s]} (66)
vv.\displaystyle\leq\norm{v-v^{\prime}}_{\infty}. (67)

It is well known that the set of solutions to Poisson’s equation 𝒱\mathcal{V}_{*} defined in (56) is non-empty (Puterman, 2014), verifying Assumption 2.3. Assumption 2.4 is directly met by the definition of αt\alpha_{t}.

To verify Assumption 2.5, we first notice that for (Average Reward TD), we have ϵt(1)=|J¯πJt|\norm{{\epsilon}^{{\left(1\right)}}_{t}}_{\infty}=\absolutevalue{\bar{J}_{\pi}-J_{t}}. It is well-known from the ergodic theorem that JtJ_{t} converges to J¯π\bar{J}_{\pi} almost surely. To verify Assumption 2.5, however, requires both an almost sure convergence rate and an L2L^{2} convergence rate. To this end, we rewrite the update of {Jt}\quantity{J_{t}} as

Jt+1=Jt+βt+1(Rt+1+γJtϕ(St+1)Jtϕ(St))ϕ(St),\displaystyle J_{t+1}=J_{t}+\beta_{t+1}\left(R_{t+1}+\gamma J_{t}\phi(S_{t+1})-J_{t}\phi(S_{t})\right)\phi(S_{t}), (68)

where we define γ0\gamma\doteq 0 and ϕ(s)1s\phi(s)\doteq 1\,\forall s. It is now clear that the update of {Jt}\quantity{J_{t}} is a special case of linear TD in the discounted setting (Sutton, 1988). Given our choice of βt=1t\beta_{t}=\frac{1}{t}, the general result about the almost sure convergence rate of linear TD (Theorem 1 of Tadić (2002)) ensures that

|JtJ¯π|ζ4.2lnlntt a.s.,\displaystyle\absolutevalue{J_{t}-\bar{J}_{\pi}}\leq\frac{\zeta_{\ref{thm:avg_rew_td}}\sqrt{\ln\ln t}}{\sqrt{t}}\mbox{\quad a.s.,\quad} (69)

where ζ4.2\zeta_{\ref{thm:avg_rew_td}} is a sample-path dependent constant. This immediately verifies (10). We do note that this almost sure convergence rate can also be obtained via a law of the iterated logarithm for Markov chains (Theorem 17.0.1 of Meyn & Tweedie (2012)). The general result about the L2L^{2} convergence rate of linear TD (Theorem 11 of Srikant & Ying (2019)) ensures that

𝔼[|JtJ¯π|2]=𝒪(1t).\displaystyle\mathbb{E}\quantity[\absolutevalue{J_{t}-\bar{J}_{\pi}}^{2}]=\mathcal{O}\quantity(\frac{1}{t}). (70)

This immediately verifies (11) and completes the proof. ∎

5 Related Work

ODE and Lyapunov Methods for Asymptotic Convergence

A large body of research has employed ODE-based methods to establish almost sure convergence of SA algorithms (Benveniste et al., 1990; Kushner & Yin, 2003; Borkar, 2009). These methods typically begin by proving stability of the iterates {xn}\quantity{x_{n}} (i.e. supnxn<\sup_{n}\norm{x_{n}}<\infty). Abounadi et al. (2002) uses this ODE-method to study the convergence of (SKM), but they require the noise sequence {Mn}\quantity{M_{n}} to be uniformly bounded, and that the set of fixed points of the nonexpansive map TT be a singleton to prove the stability of the iterates.

The ODE@\infty technique (Borkar & Meyn, 2000; Borkar et al., 2021; Meyn, 2024; Liu et al., 2025) is a powerful stability technique in RL. If the so called “ODE@\infty is globally asymptotically stable, existing results such as Meyn (2022); Borkar et al. (2021); Liu et al. (2025) can be used to establish the desired stability of {xt}\quantity{x_{t}}. However, if we consider a generic non-expansive operator hh which may admit multiple fixed points or induce oscillatory behavior, we cannot guarantee the global asymptotic stability of the ODE@\infty without additional assumptions. This limits the ODE method’s utility in analyzing (SKM with Markovian and Additive Noise).

In addition to the ODE method, there are other works that use Lyapunov methods such as (Bertsekas & Tsitsiklis, 1996; Konda & Tsitsiklis, 1999; Srikant & Ying, 2019; Borkar et al., 2021; Chen et al., 2021; Zhang et al., 2022, 2023) to provide asymptotic and nonasymptotic results of various RL algorithms. Both the ODE and Lyapunov based methods are distinct from the fox-and-hare based approach for (IKM) introduced by (Cominetti et al., 2014) that our work is built upon.

Average Reward TD

The (Average Reward TD) algorithm introduced by Tsitsiklis & Roy (1999) is the most fundamental policy evaluation algorithm in average reward settings.

In addition to the tabular setting we study here, (Average Reward TD) has also been extended to linear function approximation (Tsitsiklis & Roy, 1999; Konda & Tsitsiklis, 1999; Wu et al., 2020; Zhang et al., 2021). Instead of using a look-up table v|𝒮|v\in\mathbb{R}^{|\mathcal{S}|} to store the value estimate, linear function approximation approximates v(s)v(s) with ϕ(s)w\phi(s)^{\top}w. Let Φ|𝒮|×K\Phi\in\mathbb{R}^{{|\mathcal{S}|}\times K} be the feature matrix, whose ss-th row is the ϕ(s)\phi(s)^{\top}, and ww is the learnable weights. Linear function approximation reduces to the tabular method when Φ=I\Phi=I. While Tsitsiklis & Roy (1999) proves the almost sure convergence under assumptions such as linear independence of columns in Φ\Phi and Φwce\Phi w\neq ce for any cc\in\mathbb{R}, these conditions fail to hold in the most straightforward tabular case (where Ie=eIe=e). However, under a non-trivial construction of Φ\Phi, it can be shown that the results from Tsitsiklis & Roy (1999) can be used to prove the almost sure convergence of (Average Reward TD) to a set in the tabular case.

Zhang et al. (2021) establishes the L2L^{2} convergence of (Average Reward TD), and also provides a convergence rate. However, it is well known that L2L^{2} convergence and almost sure convergence do not imply each other. Our work improves upon both of these works by proving that the iterates converge to a fixed point almost surely.

Finally, the (Average Reward TD) algorithm has inspired the design of many other TD algorithms for average reward MDPs, for both policy evaluation and control, including Konda & Tsitsiklis (1999); Yang et al. (2016); Wan et al. (2021a); Zhang & Ross (2021); Wan et al. (2021b); He et al. (2022); Saxena et al. (2023). We envision that our work will shed light on the almost sure convergence of those follow-up algorithms.

6 Conclusion

In this work, we provide the first proof of almost sure convergence as well as non-asymptotic finite sample analysis of stochastic approximations under nonexpansive maps with Markovian noise. As an application, we provide the first proof of almost sure convergence of (Average Reward TD) to a potentially sample-path dependent fixed point. This result highlights the underappreciated strength of SKM iterations, a tool whose potential is often overlooked in the RL community. Addressing several follow-up questions could open the door to proving the convergence of many other RL algorithms. Do SKM iterations converge in LpL^{p}? Do they follow a central limit theorem or a law of the iterated logarithm? Can they be extended to two-timescale settings? And can we develop a finite sample analysis for them? Resolving these questions could pave the way for significant advancements across RL theory. We leave them for future investigation.

Acknowledgements

This work is supported in part by the US National Science Foundation (NSF) under grants III-2128019 and SLES-2331904. EB acknowledges support from the NSF Graduate Research Fellowship (NSF-GRFP) under award 1842490. This work was also supported in part by the Coastal Virginia Center for Cyber Innovation (COVA CCI) and the Commonwealth Cyber Initiative (CCI), an investment in the advancement of cyber research and development, innovation, and workforce development. For more information about CCI, visit www.covacci.org and www.cyberinitiative.org.

Impact Statement

This paper presents work whose goal is to advance the field of reinforcement learning. There are many potential societal consequences of our work, none of which we feel must be specifically highlighted here.

References

  • Abounadi et al. (2001) Abounadi, J., Bertsekas, D., and Borkar, V. S. Learning algorithms for markov decision processes with average cost. SIAM Journal on Control and Optimization, 2001.
  • Abounadi et al. (2002) Abounadi, J., Bertsekas, D. P., and Borkar, V. Stochastic approximation for nonexpansive maps: Application to q-learning algorithms. SIAM Journal on Control and Optimization, 41(1):1–22, 2002.
  • Bellman (1957) Bellman, R. A markovian decision process. Journal of mathematics and mechanics, pp.  679–684, 1957.
  • Benveniste et al. (1990) Benveniste, A., Métivier, M., and Priouret, P. Adaptive Algorithms and Stochastic Approximations. Springer, 1990.
  • Bertsekas & Tsitsiklis (1996) Bertsekas, D. P. and Tsitsiklis, J. N. Neuro-Dynamic Programming. Athena Scientific Belmont, MA, 1996.
  • Borkar et al. (2021) Borkar, V., Chen, S., Devraj, A., Kontoyiannis, I., and Meyn, S. The ode method for asymptotic statistics in stochastic approximation and reinforcement learning. arXiv preprint arXiv:2110.14427, 2021.
  • Borkar (2009) Borkar, V. S. Stochastic approximation: a dynamical systems viewpoint. Springer, 2009.
  • Borkar & Meyn (2000) Borkar, V. S. and Meyn, S. P. The ode method for convergence of stochastic approximation and reinforcement learning. SIAM Journal on Control and Optimization, 2000.
  • Bravo & Cominetti (2024) Bravo, M. and Cominetti, R. Stochastic fixed-point iterations for nonexpansive maps: Convergence and error bounds. SIAM Journal on Control and Optimization, 62(1):191–219, 2024.
  • Bravo et al. (2019) Bravo, M., Cominetti, R., and Pavez-Signé, M. Rates of convergence for inexact krasnosel’skii–mann iterations in banach spaces. Mathematical Programming, 175:241–262, 2019.
  • Chen et al. (2021) Chen, Z., Maguluri, S. T., Shakkottai, S., and Shanmugam, K. A lyapunov theory for finite-sample guarantees of asynchronous q-learning and td-learning variants. arXiv preprint arXiv:2102.01567, 2021.
  • Cominetti et al. (2014) Cominetti, R., Soto, J. A., and Vaisman, J. On the rate of convergence of krasnosel’skii-mann iterations and their connection with sums of bernoullis. Israel Journal of Mathematics, 199:757–772, 2014.
  • Edelstein (1966) Edelstein, M. A remark on a theorem of m. a. krasnoselski. American Mathematical Monthly, 1966.
  • Folland (1999) Folland, G. B. Real analysis: modern techniques and their applications, volume 40. John Wiley & Sons, 1999.
  • He et al. (2022) He, J., Wan, Y., and Mahmood, A. R. The emphatic approach to average-reward policy evaluation. In Deep Reinforcement Learning Workshop NeurIPS 2022, 2022.
  • Ishikawa (1976) Ishikawa, S. Fixed points and iteration of a nonexpansive mapping in a banach space. Proceedings of the American Mathematical Society, 59(1):65–71, 1976.
  • Karandikar & Vidyasagar (2024) Karandikar, R. L. and Vidyasagar, M. Convergence rates for stochastic approximation: Biased noise with unbounded variance, and applications. Journal of Optimization Theory and Applications, pp.  1–39, 2024.
  • Kiefer & Wolfowitz (1952) Kiefer, J. and Wolfowitz, J. Stochastic estimation of the maximum of a regression function. Annals of Mathematical Statistics, 1952.
  • Kim & Xu (2007) Kim, T.-H. and Xu, H.-K. Robustness of mann’s algorithm for nonexpansive mappings. Journal of Mathematical Analysis and Applications, 327(2):1105–1115, 2007.
  • Konda & Tsitsiklis (1999) Konda, V. R. and Tsitsiklis, J. N. Actor-critic algorithms. In Advances in Neural Information Processing Systems, 1999.
  • Koval & Schwabe (2003) Koval, V. and Schwabe, R. A law of the iterated logarithm for stochastic approximation procedures in d-dimensional euclidean space. Stochastic processes and their applications, 105(2):299–313, 2003.
  • Krasnosel’skii (1955) Krasnosel’skii, M. A. Two remarks on the method of successive approximations. Uspekhi matematicheskikh nauk, 10(1):123–127, 1955.
  • Kushner & Yin (2003) Kushner, H. and Yin, G. G. Stochastic approximation and recursive algorithms and applications. Springer Science & Business Media, 2003.
  • Liu (1995) Liu, L.-S. Ishikawa and mann iterative process with errors for nonlinear strongly accretive mappings in banach spaces. Journal of Mathematical Analysis and Applications, 194(1):114–125, 1995.
  • Liu et al. (2025) Liu, S., Chen, S., and Zhang, S. The ODE method for stochastic approximation and reinforcement learning with markovian noise. Journal of Machine Learning Research, 2025.
  • Métivier & Priouret (1987) Métivier, M. and Priouret, P. Théorèmes de convergence presque sure pour une classe d’algorithmes stochastiques à pas décroissant. Probability Theory and related fields, 74:403–428, 1987.
  • Meyn (2022) Meyn, S. Control systems and reinforcement learning. Cambridge University Press, 2022.
  • Meyn (2024) Meyn, S. The projected bellman equation in reinforcement learning. IEEE Transactions on Automatic Control, 2024.
  • Meyn & Tweedie (2012) Meyn, S. P. and Tweedie, R. L. Markov chains and stochastic stability. Springer Science & Business Media, 2012.
  • Puterman (2014) Puterman, M. L. Markov decision processes: discrete stochastic dynamic programming. John Wiley & Sons, 2014.
  • Qian et al. (2024) Qian, X., Xie, Z., Liu, X., and Zhang, S. Almost sure convergence rates and concentration of stochastic approximation and reinforcement learning with markovian noise. arXiv preprint arXiv:2411.13711, 2024.
  • Reich (1979) Reich, S. Weak convergence theorems for nonexpansive mappings in banach spaces. J. Math. Anal. Appl, 67(2):274–276, 1979.
  • Robbins & Monro (1951) Robbins, H. and Monro, S. A stochastic approximation method. The Annals of Mathematical Statistics, 1951.
  • Saxena et al. (2023) Saxena, N., Khastagir, S., Kolathaya, S., and Bhatnagar, S. Off-policy average reward actor-critic with deterministic policy search. In International Conference on Machine Learning, pp.  30130–30203. PMLR, 2023.
  • Srikant & Ying (2019) Srikant, R. and Ying, L. Finite-time error bounds for linear stochastic approximation andtd learning. In Proceedings of the Conference on Learning Theory, 2019.
  • Sutton (1988) Sutton, R. S. Learning to predict by the methods of temporal differences. Machine Learning, 1988.
  • Sutton & Barto (2018) Sutton, R. S. and Barto, A. G. Reinforcement Learning: An Introduction (2nd Edition). MIT press, 2018.
  • Szepesvári (1997) Szepesvári, C. The asymptotic convergence-rate of q-learning. Advances in neural information processing systems, 10, 1997.
  • Tadić (2002) Tadić, V. B. On the almost sure rate of convergence of temporal-difference learning algorithms. IFAC Proceedings Volumes, 35(1):455–460, 2002.
  • Tadic (2004) Tadic, V. B. On the almost sure rate of convergence of linear stochastic approximation algorithms. IEEE Transactions on Information Theory, 50(2):401–409, 2004.
  • Tsitsiklis & Roy (1999) Tsitsiklis, J. N. and Roy, B. V. Average cost temporal-difference learning. Automatica, 1999.
  • Wan et al. (2021a) Wan, Y., Naik, A., and Sutton, R. Average-reward learning and planning with options. Advances in Neural Information Processing Systems, 34:22758–22769, 2021a.
  • Wan et al. (2021b) Wan, Y., Naik, A., and Sutton, R. S. Learning and planning in average-reward markov decision processes. In Proceedings of the International Conference on Machine Learning, 2021b.
  • Wu et al. (2020) Wu, Y., Zhang, W., Xu, P., and Gu, Q. A finite-time analysis of two time-scale actor-critic methods. In Advances in Neural Information Processing Systems, 2020.
  • Yang et al. (2016) Yang, S., Gao, Y., An, B., Wang, H., and Chen, X. Efficient average reward reinforcement learning using constant shifting values. In Proceedings of the AAAI Conference on Artificial Intelligence, volume 30, 2016.
  • Zhang et al. (2021) Zhang, S., Zhang, Z., and Maguluri, S. T. Finite sample analysis of average-reward td learning and qq-learning. Advances in Neural Information Processing Systems, 2021.
  • Zhang et al. (2022) Zhang, S., des Combes, R. T., and Laroche, R. Global optimality and finite sample analysis of softmax off-policy actor critic under state distribution mismatch. Journal of Machine Learning Research, 2022.
  • Zhang et al. (2023) Zhang, S., Des Combes, R. T., and Laroche, R. On the convergence of sarsa with linear function approximation. In International Conference on Machine Learning, 2023.
  • Zhang & Ross (2021) Zhang, Y. and Ross, K. W. On-policy deep reinforcement learning for the average-reward criterion. In International Conference on Machine Learning, pp.  12535–12545. PMLR, 2021.

Appendix A Mathematical Background

Lemma A.1 (Theorem 2.1 from Bravo & Cominetti (2024)).

Let {zn}\quantity{z_{n}} be a sequence generated by (IKM). Let Fix(T)\text{Fix}(T) denote the set of fixed points of TT (assumed to be nonempty). Additionally, let τn\tau_{n} be defined according to (9) and the real function σ:(0,)(0,)\sigma:(0,\infty)\rightarrow(0,\infty) as

σ(y)=min{1,1/πy}.\displaystyle\sigma(y)=\min\quantity{{1,1/\sqrt{\pi y}}}. (71)

If ζA.10\zeta_{\ref{lem:bravo_2.1}}\geq 0 is such that Tznx0ζA.1\norm{Tz_{n}-x_{0}}\leq\zeta_{\ref{lem:bravo_2.1}} for all n1n\geq 1, then

znTznζA.1σ(τn)+k=1n2αkekσ(τnτk)+2en+1.\norm{z_{n}-Tz_{n}}\leq\zeta_{\ref{lem:bravo_2.1}}\sigma{\left(\tau_{n}\right)}+\sum_{k=1}^{n}2\alpha_{k}\norm{e_{k}}\sigma{\left(\tau_{n}-\tau_{k}\right)}+2\norm{e_{n+1}}. (72)

Moreover, if τn\tau_{n}\rightarrow\infty and en0\norm{e_{n}}\rightarrow 0 with Sn=1αnen<S\doteq\sum_{n=1}^{\infty}\alpha_{n}\norm{e_{n}}<\infty, then (72) holds with ζA.1=2infxFix(T)x0x+S\zeta_{\ref{lem:bravo_2.1}}=2\inf_{x\in\text{Fix}(T)}\norm{x_{0}-x}+S, and we have znTzn0\norm{z_{n}-Tz_{n}}\rightarrow 0 as well as znxz_{n}\rightarrow x_{*} for some fixed point xFix(T)x_{*}\in\text{Fix}(T)

Lemma A.2 (Monotonicity of αk,n\alpha_{k,n} from Lemma B.1 in Bravo & Cominetti (2024)).

For αn=1(n+1)b\alpha_{n}=\frac{1}{{\left(n+1\right)}^{b}} with 0<b10<b\leq 1 and αi,n\alpha_{i,n} in (8), we have αk,nαk+1,n\alpha_{k,n}\leq\alpha_{k+1,n} for k1k\geq 1 so that αk+1,nαn,n=αn\alpha_{k+1,n}\leq\alpha_{n,n}=\alpha_{n}.

Lemma A.3 (Lemma B.2 from (Bravo & Cominetti, 2024)).

For αn=1(n+1)b\alpha_{n}=\frac{1}{{\left(n+1\right)}^{b}} with 0<b10<b\leq 1 and αi,n\alpha_{i,n} in (8), we have k=1nαk,n2αn+1\sum_{k=1}^{n}\alpha_{k,n}^{2}\leq\alpha_{n+1} for all n1n\geq 1.

Lemma A.4 (Monotone Convergence Theorem from Folland (1999)).

Given a measure space (X,M,μ)\quantity(X,M,\mu), define L+L^{+} as the space of all measurable functions from XX to [0,][0,\infty]. Then, if {fn}\quantity{f_{n}} is a sequence in L+L^{+} such that fjfj+1f_{j}\leq f_{j+1} for all j, and f=limnfnf=\lim_{n\rightarrow\infty}f_{n}, then f𝑑μ=limnfn𝑑μ\int fd\mu=\lim_{n\rightarrow\infty}\int f_{n}d\mu.

Appendix B Additional Lemmas from Section 2

In this section, we present and prove the lemmas referenced in Section 2 as part of the proof of Theorem 2.6. Additionally, we establish several auxiliary lemmas necessary for these proofs.

We begin by proving several convergence results related to the learning rates.

Lemma B.1 (Learning Rates).

With τn\tau_{n} defined in (9) we have,

τn={𝒪(n1b)if45<b<1,𝒪(logn)ifb=1.\displaystyle\tau_{n}=\begin{cases}\mathcal{O}\quantity(n^{1-b})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\log n)&\text{if}\quad b=1.\end{cases} (73)

This further implies,

supnk=1nαk2τk\displaystyle\sup_{n}\,\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k} <,\displaystyle<\infty, (74)
supnk=1nαk2τk2\displaystyle\sup_{n}\,\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}^{2} <,\displaystyle<\infty, (75)
supnk=1nαk3/2τk1\displaystyle\sup_{n}\,\sum_{k=1}^{n}\alpha_{k}^{3/2}\tau_{k-1} <,\displaystyle<\infty, (76)
supnk=0n1|αkαk+1|τk\displaystyle\sup_{n}\sum_{k=0}^{n-1}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k} <,\displaystyle<\infty, (77)
supnk=1nαk2j=1i1αjτj\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}^{2}\sum_{j=1}^{i-1}\alpha_{j}\tau_{j} <,\displaystyle<\infty, (78)
supnk=1nαkj=1k1αj,k12τj12\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}\sqrt{\sum_{j=1}^{k-1}\alpha_{j,k-1}^{2}\tau_{j-1}^{2}} <,\displaystyle<\infty, (79)

Since this Lemma is comprised of several short proofs regarding the deterministic learning rates defined in Assumption 2.4, we will decompose each result into subsections. Recall that αn1(n+1)b\alpha_{n}\doteq\frac{1}{\quantity(n+1)^{b}} where 45<b1\frac{4}{5}<b\leq 1.

(73):

Proof.

From the definition of τn\tau_{n} in (9), we have

τn\displaystyle\tau_{n} k=1nαk(1αk)k=1nαk=k=1n1(k+1)b.\displaystyle\doteq\sum_{k=1}^{n}\alpha_{k}\quantity(1-\alpha_{k})\leq\sum_{k=1}^{n}\alpha_{k}=\sum_{k=1}^{n}\frac{1}{(k+1)^{b}}. (81)

Case 1: b=1b=1. It is easy to see τn=𝒪(logn)\tau_{n}=\mathcal{O}\quantity(\log n).

Case 2: When b<1b<1, we can approximate the sum with an integral, with

k=1n1(k+1)b1n1kb𝑑k=n1b11b\displaystyle\sum_{k=1}^{n}\frac{1}{\quantity(k+1)^{b}}\leq\int_{1}^{n}\frac{1}{k^{b}}\,dk=\frac{n^{1-b}-1}{1-b} (82)

Therefore we have τn=𝒪(n1b)\tau_{n}=\mathcal{O}\quantity(n^{1-b}) when b<1b<1. ∎

In analyzing the subsequent equations, we will use the fact that τn=𝒪(logn)\tau_{n}=\mathcal{O}\quantity(\log n) when b=1b=1 and τn=𝒪(n1b)\tau_{n}=\mathcal{O}\quantity(n^{1-b}) when 45<b<1\frac{4}{5}<b<1. Additionally, we have αn=(1nb)\alpha_{n}=\quantity(\frac{1}{n^{b}}).

(74):

Proof.

We have an order-wise approximation of the sum

k=1nαk2τk={𝒪(k=1n1k3b1)if45<b<1,𝒪(k=1nlog(k)k2)ifb=1..\displaystyle\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}=\begin{dcases}\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{3b-1}})&\text{if}\ \frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\sum_{k=1}^{n}\frac{\log(k)}{k^{2}})&\text{if}\ b=1.\end{dcases}. (83)

In both cases of b=1b=1 and 45<b<1\frac{4}{5}<b<1, the series clearly converge as nn\rightarrow\infty. ∎

(76):

Proof.

We have an order-wise approximation of the sum

k=1nαk3/2τk={𝒪(k=1n1k52b1)if45<b<1,𝒪(k=1nlog(k)k3/2)ifb=1..\displaystyle\sum_{k=1}^{n}\alpha_{k}^{3/2}\tau_{k}=\begin{dcases}\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{\frac{5}{2}b-1}})&\text{if}\ \frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\sum_{k=1}^{n}\frac{\log(k)}{k^{3/2}})&\text{if}\ b=1.\end{dcases}. (84)

In both cases of b=1b=1 and 45<b<1\frac{4}{5}<b<1, the series clearly converge as nn\rightarrow\infty. ∎

(75):

Proof.

We can give an order-wise approximation of the sum

k=1nαk2τk2={𝒪(k=1n1k4b2)if45<b<1,𝒪(k=1nlog2(k)k2)ifb=1..\displaystyle\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}^{2}=\begin{dcases}\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{4b-2}})&\text{if}\ \frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\sum_{k=1}^{n}\frac{\log^{2}(k)}{k^{2}})&\text{if}\ b=1.\end{dcases}. (85)

In both cases of b=1b=1 and 45<b<1\frac{4}{5}<b<1, the series clearly converge as nn\rightarrow\infty. ∎

(77):

Proof.

Since αn\alpha_{n} is strictly decreasing, we have |αkαk+1|=αkαk+1\absolutevalue{\alpha_{k}-\alpha_{k+1}}=\alpha_{k}-\alpha_{k+1}.

Case 1: For the case where b=1b=1, it is trivial to see that,

k=1n|αkαk+1|τk=𝒪(k=1nlog(k)k2+k).\displaystyle\sum_{k=1}^{n}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k}=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{\log(k)}{k^{2}+k}). (86)

This series clearly converges.

Case 2: For the case where 45<b<1\frac{4}{5}<b<1, we have

αnαn+1\displaystyle\alpha_{n}-\alpha_{n+1} =𝒪(1nb1(n+1)b),\displaystyle=\mathcal{O}\quantity(\frac{1}{n^{b}}-\frac{1}{(n+1)^{b}}), (87)
=𝒪((n+1)bnbnb(n+1)b).\displaystyle=\mathcal{O}\quantity(\frac{(n+1)^{b}-n^{b}}{n^{b}(n+1)^{b}}). (88)

To analyze the behavior of this term for large nn we first consider the binomial expansion of (n+1)b(n+1)^{b},

(n+1)b\displaystyle(n+1)^{b} =nb(1+1n)b=nb(1+b1n+b(b1)21n2+)\displaystyle=n^{b}\quantity(1+\frac{1}{n})^{b}=n^{b}(1+b\frac{1}{n}+\frac{b(b-1)}{2}\frac{1}{n^{2}}+\dots) (89)

Subtracting nbn^{b} from (n+1)b(n+1)^{b}:

(n+1)bnb=nb(1+b1n+b(b1)21n2+)nb=𝒪(bnb1).\displaystyle(n+1)^{b}-n^{b}=n^{b}(1+b\frac{1}{n}+\frac{b(b-1)}{2}\frac{1}{n^{2}}+\dots)-n^{b}=\mathcal{O}\quantity(bn^{b-1}). (90)

The leading order of the denominator of (88) is clearly n2bn^{2b}, which gives

αnαn+1=𝒪(bnb1n2b)=𝒪(bnb+1).\displaystyle\alpha_{n}-\alpha_{n+1}=\mathcal{O}\quantity(\frac{bn^{b-1}}{n^{2b}})=\mathcal{O}\quantity(\frac{b}{n^{b+1}}). (91)

Therefore with τn=𝒪(n1b)\tau_{n}=\mathcal{O}\quantity(n^{1-b}),

k=1n|αkαk+1|τk=𝒪(bk=1n1k2b)\displaystyle\sum_{k=1}^{n}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k}=\mathcal{O}\quantity(b\sum_{k=1}^{n}\frac{1}{k^{2b}}) (92)

which clearly converges as nn\rightarrow\infty for 45<b<1\frac{4}{5}<b<1. ∎

(78):

Proof.

Case 1: In the proof for (73) we prove that k=1nαk=𝒪(logn)\sum_{k=1}^{n}\alpha_{k}=\mathcal{O}\quantity(\log n) when b=1b=1. Then since τk\tau_{k} is increasing, we have

k=1nαk2j=1k1αjτjk=1nαk2τkj=1k1αj=𝒪(k=1nlog2kk2),\sum_{k=1}^{n}\alpha_{k}^{2}\sum_{j=1}^{k-1}\alpha_{j}\tau_{j}\leq\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}\sum_{j=1}^{k-1}\alpha_{j}=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{\log^{2}k}{k^{2}}), (93)

which clearly converges as nn\rightarrow\infty.

Case 2: For the case when b(45,1)b\in(\frac{4}{5},1), we first consider the inner sum of (78),

j=1k1αjτj=𝒪(j=1k11j2b1),\displaystyle\sum_{j=1}^{k-1}\alpha_{j}\tau_{j}=\mathcal{O}\quantity(\sum_{j=1}^{k-1}\frac{1}{j^{2b-1}}), (94)

which we can approximate by an integral,

1k1x2b1𝑑x=𝒪(k22b).\displaystyle\int_{1}^{k}\frac{1}{x^{2b-1}}\ dx=\mathcal{O}\quantity(k^{2-2b}). (95)

Therefore,

k=1nαk2j=1k1αjτj=𝒪(k=1nk22bk2b)=𝒪(k=1n1k4b2),\displaystyle\sum_{k=1}^{n}\alpha_{k}^{2}\sum_{j=1}^{k-1}\alpha_{j}\tau_{j}=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{k^{2-2b}}{k^{2b}})=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{4b-2}}), (96)

which converges for 45<b1\frac{4}{5}<b\leq 1 as nn\rightarrow\infty. ∎

(79):

Proof.

Case 1: For b=1b=1, because we have αj,i<αj+1,i\alpha_{j,i}<\alpha_{j+1,i} and αi,i=αi\alpha_{i,i}=\alpha_{i} from Lemma A.2, we have the order-wise approximation,

i=1nαij=1i1αj,i12τj12\displaystyle\sum_{i=1}^{n}\alpha_{i}\sqrt{\sum_{j=1}^{i-1}\alpha_{j,i-1}^{2}\tau_{j-1}^{2}} i=1nαiαi12τi12j=1i11,\displaystyle\leq\sum_{i=1}^{n}\alpha_{i}\sqrt{\alpha_{i-1}^{2}\tau_{i-1}^{2}\sum_{j=1}^{i-1}1}, (τi\tau_{i} is increasing) (97)
=i=1nαiαi1τi1i1.\displaystyle=\sum_{i=1}^{n}\alpha_{i}\alpha_{i-1}\tau_{i-1}\sqrt{i-1}. (98)
=𝒪(i=1nlog(i1)i(i1))\displaystyle=\mathcal{O}\quantity(\sum_{i=1}^{n}\frac{\log(i-1)}{i\sqrt{(i-1)}}) (99)
=𝒪(i=1nlog(i1)i3/2),\displaystyle=\mathcal{O}\quantity(\sum_{i=1}^{n}\frac{\log(i-1)}{i^{3/2}}), (100)

which clearly converges.

Case 2: For the case when b(45,1)b\in(\frac{4}{5},1), we have,

i=1nαij=1i1αj,i12τj12\displaystyle\sum_{i=1}^{n}\alpha_{i}\sqrt{\sum_{j=1}^{i-1}\alpha_{j,i-1}^{2}\tau_{j-1}^{2}} i=1nαiτi1j=1i1αj,i12,\displaystyle\leq\sum_{i=1}^{n}\alpha_{i}\tau_{i-1}\sqrt{\sum_{j=1}^{i-1}\alpha_{j,i-1}^{2}}, (τi\tau_{i} is increasing) (101)
=i=1nαiτi1αi.\displaystyle=\sum_{i=1}^{n}\alpha_{i}\tau_{i-1}\sqrt{\alpha_{i}}. (Lemma A.3) (102)
=𝒪(i=1ni1bibib)\displaystyle=\mathcal{O}\quantity(\sum_{i=1}^{n}\frac{i^{1-b}}{i^{b}\sqrt{i^{b}}}) (103)
=𝒪(i=1n1i5b/21),\displaystyle=\mathcal{O}\quantity(\sum_{i=1}^{n}\frac{1}{i^{5b/2-1}}), (104)

which converges for 45<b<1\frac{4}{5}<b<1. ∎

Then, under Assumption 2.5, we prove additional results about the convergence of the first and second moments of the additive noise {ϵn(1)}\quantity{{\epsilon}^{{\left(1\right)}}_{n}}.

Lemma B.2.

Let Assumptions 2.4 and 2.5 hold. Then, we have

𝔼[ϵn(1)]\displaystyle\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{n}}] =𝒪(1n),\displaystyle=\mathcal{O}\quantity(\frac{1}{\sqrt{n}}), (105)
supnk=1nαk𝔼[ϵk(1)]\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}] <,\displaystyle<\infty, (106)
supnk=1nαk𝔼[ϵk(1)2]\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2}] <,\displaystyle<\infty, (107)
supnk=1nαk2𝔼[ϵk(1)2]\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}^{2}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2}] <,\displaystyle<\infty, (108)
supnk=1nαkj=1k1αj,k1𝔼[ϵj(1)]\displaystyle\sup_{n}\sum_{k=1}^{n}\alpha_{k}\sum_{j=1}^{k-1}\alpha_{j,k-1}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{j}}] <.\displaystyle<\infty. (109)
Proof.

Recall that by Assumption 2.5 we have 𝔼[ϵn(1)2]=𝒪(1n)\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{n}}^{2}]=\mathcal{O}\quantity(\frac{1}{n}). Also recall that αk=𝒪(1nb)\alpha_{k}=\mathcal{O}\quantity(\frac{1}{n^{b}}) with 45<b1\frac{4}{5}<b\leq 1. Then, we can prove the following equations:

(105):

By Jensen’s inequality, we have

𝔼[ϵn(1)]𝔼[ϵn(1)2]=𝒪(1n).\displaystyle\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{n}}]\leq\sqrt{\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{n}}^{2}]}=\mathcal{O}\quantity(\frac{1}{\sqrt{n}}). (110)

(106):

k=1nαk𝔼[ϵk(1)]=𝒪(k=1n1kb+12)\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}]=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{b+\frac{1}{2}}}) (111)

which clearly converges as nn\rightarrow\infty for 45<b1\frac{4}{5}<b\leq 1.

(107):

k=1nαk𝔼[ϵk(1)2]=𝒪(k=1n1kb+1)\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2}]=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{b+1}}) (112)

which clearly converges as nn\rightarrow\infty for 45<b1\frac{4}{5}<b\leq 1.

(108):

k=1nαk2𝔼[ϵk(1)2]=𝒪(k=1n1k2b+1)\sum_{k=1}^{n}\alpha_{k}^{2}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2}]=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{2b+1}}) (113)

which clearly converges as nn\rightarrow\infty for 45<b1\frac{4}{5}<b\leq 1.

(109):

k=1nαkj=1k1αj,k1𝔼[ϵj(1)]\displaystyle\sum_{k=1}^{n}\alpha_{k}\sum_{j=1}^{k-1}\alpha_{j,k-1}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{j}}] k=1nαk2j=1k1𝔼[ϵj(1)],\displaystyle\leq\sum_{k=1}^{n}\alpha_{k}^{2}\sum_{j=1}^{k-1}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{j}}], (Lemma A.2) (114)
=𝒪(k=1n1k2bj=1k11j).\displaystyle=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{2b}}\sum_{j=1}^{k-1}\frac{1}{\sqrt{j}}). (Lemma B.2) (115)

It can be easily verified with an integral approximation that j=1k11j=𝒪(k)\sum_{j=1}^{k-1}\frac{1}{\sqrt{j}}=\mathcal{O}(\sqrt{k}). This further implies

k=1nαkj=1k1αj,k1𝔼[ϵj(1)]\displaystyle\sum_{k=1}^{n}\alpha_{k}\sum_{j=1}^{k-1}\alpha_{j,k-1}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{j}}] =𝒪(k=1n1k2b12),\displaystyle=\mathcal{O}\quantity(\sum_{k=1}^{n}\frac{1}{k^{2b-\frac{1}{2}}}), (116)

which converges as nn\rightarrow\infty for 45<b1\frac{4}{5}<b\leq 1. ∎

Next, in Lemma B.3, we upper-bound the iterates {xn}\quantity{x_{n}}.

Lemma B.3.

For each {xn}\quantity{x_{n}}, we have

xnx0+CHk=1nαk+k=1nαkϵk(1)CB.3τn+k=1nαkϵk(1),\|x_{n}\|\leq\norm{x_{0}}+C_{H}\sum_{k=1}^{n}\alpha_{k}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}\leq C_{\ref{lem:xn_norm}}\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}, (117)

where CB.3C_{\ref{lem:xn_norm}} is a deterministic constant.

Proof.

Applying \|\cdot\| to both sides of (SKM with Markovian and Additive Noise) gives,

xn+1\displaystyle\|x_{n+1}\| =(1αn+1)xn+αn+1(H(xn,Yn+1)+ϵn+1(1)),\displaystyle=\norm{(1-\alpha_{n+1})x_{n}+\alpha_{n+1}\quantity(H{\left(x_{n},Y_{n+1}\right)}+{\epsilon}^{{\left(1\right)}}_{n+1})}, (118)
(1αn+1)xn+αn+1H(xn,Yn+1)+αn+1ϵn+1(1),\displaystyle\leq(1-\alpha_{n+1})\|x_{n}\|+\alpha_{n+1}\norm{H(x_{n},Y_{n+1})}+\alpha_{n+1}\norm{{\epsilon}^{{\left(1\right)}}_{n+1}}, (119)
(1αn+1)xn+αn+1(CH+xn)+αn+1ϵn+1(1),\displaystyle\leq(1-\alpha_{n+1})\norm{x_{n}}+\alpha_{n+1}{\left(C_{H}+\norm{x_{n}}\right)}+\alpha_{n+1}\norm{{\epsilon}^{{\left(1\right)}}_{n+1}}, (By (3)) (120)
=xn+αn+1CH+αn+1ϵn+1(1).\displaystyle=\|x_{n}\|+\alpha_{n+1}C_{H}+\alpha_{n+1}\norm{{\epsilon}^{{\left(1\right)}}_{n+1}}. (121)

A simple induction shows that almost surely,

xnx0+CHk=1nαk+k=1nαkϵk(1).\displaystyle\norm{x_{n}}\leq\norm{x_{0}}+C_{H}\sum_{k=1}^{n}\alpha_{k}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}. (122)

Since {αn}\quantity{\alpha_{n}} is monotonically decreasing, we have

xn\displaystyle\norm{x_{n}} x0+CH(1α1)k=1nαk(1αk)+k=1nαkϵk(1),\displaystyle\leq\norm{x_{0}}+\frac{C_{H}}{{\left(1-\alpha_{1}\right)}}\sum_{k=1}^{n}\alpha_{k}(1-\alpha_{k})+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}, (123)
=x0+CH(1α1)τn+k=1nαkϵk(1),\displaystyle=\norm{x_{0}}+\frac{C_{H}}{{\left(1-\alpha_{1}\right)}}\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}, (124)
max{x0,CH(1α1)}(1+τn)+k=1nαkϵk(1).\displaystyle\leq\max\left\{\norm{x_{0}},\frac{C_{H}}{{\left(1-\alpha_{1}\right)}}\right\}{\left(1+\tau_{n}\right)}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}. (125)

Therefore, since τn\tau_{n} is monotonically increasing, there exists some constant we denote as CB.3C_{\ref{lem:xn_norm}} such that

xn\displaystyle\norm{x_{n}} CB.3τn+k=1nαkϵk(1).\displaystyle\leq C_{\ref{lem:xn_norm}}\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}. (126)

Lemma B.4.

With ν(x,y)\nu(x,y) as defined in (13), we have

ν(x,y)ν(x,y)CB.4xx,\norm{\nu{\left(x,y\right)}-\nu{\left(x^{\prime},y\right)}}\leq C_{\ref{lem:v Lipschitz}}\norm{x-x^{\prime}}, (127)

which further implies

ν(x,y)CB.4(CB.4+x),\norm{\nu{\left(x,y\right)}}\leq C_{\ref{lem:v Lipschitz}}\quantity(C_{\ref{lem:v Lipschitz}}^{\prime}+\norm{x}), (128)

where CB.4,CB.4C_{\ref{lem:v Lipschitz}},C_{\ref{lem:v Lipschitz}}^{\prime} are deterministic constants.

Proof.

Since we work with a finite 𝒴\mathcal{Y}, we will use functions and matrices interchangeably. For example, given a function f:𝒴df:\mathcal{Y}\to\mathbb{R}^{d}, we also use ff to denote a matrix in (|𝒴|×d)\mathbb{R}^{\quantity({|\mathcal{Y}|}\times d)} whose yy-th row is f(y)f(y)^{\top}. Similarly, a matrix in (|𝒴|×d)\mathbb{R}^{\quantity({|\mathcal{Y}|}\times d)} also corresponds to a function 𝒴d\mathcal{Y}\to\mathbb{R}^{d}.

Let νx|𝒴|×d\nu_{x}\in\mathbb{R}^{{|\mathcal{Y}|}\times d} denote the function yν(x,y)y\mapsto\nu(x,y) and let Hx|𝒴|×dH_{x}\in\mathbb{R}^{{|\mathcal{Y}|}\times d} denote the function yH(x,y)y\mapsto H(x,y). Theorem 8.2.6 of Puterman (2014) then ensures that

νx=H𝒴Hx,\displaystyle\nu_{x}=H_{\mathcal{Y}}H_{x}, (129)

where H𝒴|𝒴|×|𝒴|H_{\mathcal{Y}}\in\mathbb{R}^{{|\mathcal{Y}|}\times{|\mathcal{Y}|}} is the fundamental matrix of the Markov chain depending only on the chain’s transition matrix PP. The exact expression of H𝒴H_{\mathcal{Y}} is inconsequential and we refer the reader to Puterman (2014) for details. Then we have for any i=1,,di=1,\dots,d,

νx[y,i]=yH𝒴[y,y]Hx[y,i].\displaystyle\nu_{x}[y,i]=\sum_{y^{\prime}}H_{\mathcal{Y}}[y,y^{\prime}]H_{x}[y^{\prime},i]. (130)

This implies that

|νx[y,i]νx[y,i]|\displaystyle\absolutevalue{\nu_{x}[y,i]-\nu_{x^{\prime}}[y,i]}\leq yH𝒴[y,y]|Hx[y,i]Hx[y,i]|\displaystyle\sum_{y^{\prime}}H_{\mathcal{Y}}[y,y^{\prime}]\absolutevalue{H_{x}[y^{\prime},i]-H_{x^{\prime}}[y^{\prime},i]} (131)
\displaystyle\leq yH𝒴[y,y]H(x,y)H(x,y)\displaystyle\sum_{y^{\prime}}H_{\mathcal{Y}}[y,y^{\prime}]\norm{H(x,y)-H(x^{\prime},y^{\prime})}_{\infty} (132)
\displaystyle\leq yH𝒴[y,y]xx\displaystyle\sum_{y^{\prime}}H_{\mathcal{Y}}[y,y^{\prime}]\norm{x-x^{\prime}}_{\infty} (Assumption 2.2) (133)
\displaystyle\leq H𝒴xx,\displaystyle\norm{H_{\mathcal{Y}}}_{\infty}\norm{x-x^{\prime}}_{\infty}, (134)

yielding

ν(x,y)ν(x,y)H𝒴xx.\displaystyle\norm{\nu(x,y)-\nu(x^{\prime},y)}_{\infty}\leq\norm{H_{\mathcal{Y}}}_{\infty}\norm{x-x^{\prime}}_{\infty}. (135)

The equivalence between norms in finite dimensional space ensures that there exists some CB.4C_{\ref{lem:v Lipschitz}} such that (127) holds. Letting x=0x^{\prime}=0 then yields

ν(x,y)CB.4(ν(0,y)+x).\displaystyle\norm{\nu(x,y)}\leq C_{\ref{lem:v Lipschitz}}\quantity(\norm{\nu(0,y)}+\norm{x}). (136)

Define CB.4maxyν(0,y)C_{\ref{lem:v Lipschitz}}^{\prime}\doteq\max_{y}\norm{\nu(0,y)}, we get

ν(x,y)CB.4(CB.4+x).\displaystyle\norm{\nu(x,y)}\leq C_{\ref{lem:v Lipschitz}}\quantity(C_{\ref{lem:v Lipschitz}}^{\prime}+\norm{x}). (137)

Lemma B.5.

We have for any y𝒴y\in\mathcal{Y},

ν(xn,y)ζB.5τn,\norm{\nu{\left(x_{n},y\right)}}\leq\zeta_{\ref{lem:v_norm}}\tau_{n}, (138)

where ζ\zeta is a possibly sample-path dependent constant. Additionally, we have

𝔼[ν(xn,y)]\displaystyle\mathbb{E}\quantity[\norm{\nu{\left(x_{n},y\right)}}] CB.5τn,\displaystyle\leq C_{\ref{lem:v_norm}}\tau_{n}, (139)

where CB.5C_{\ref{lem:v_norm}} is a deterministic constant.

Proof.

Having proven that ν(x,y)\nu{\left(x,y\right)} is Lipschitz continuous in xx in Lemma B.4, we have

ν(xn,y)\displaystyle\norm{\nu{\left(x_{n},y\right)}} CB.4(CB.4+xn),\displaystyle\leq C_{\ref{lem:v Lipschitz}}(C_{\ref{lem:v Lipschitz}}^{\prime}+\norm{x_{n}}), (Lemma B.4) (140)
CB.4(CB.4+CB.3τn+k=1nαkϵk(1)).\displaystyle\leq C_{\ref{lem:v Lipschitz}}\quantity(C_{\ref{lem:v Lipschitz}}^{\prime}+C_{\ref{lem:xn_norm}}\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}). (Lemma B.3) (141)
=𝒪(τn+k=1nαkϵk(1)).\displaystyle=\mathcal{O}\quantity(\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}). (142)

Since (10) in Assumption 2.5 assures us that k=1αkϵk(1)\sum_{k=1}^{\infty}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}} is finite almost surely while τn\tau_{n} is monotonically increasing, then there exists some possibly sample-path dependent constant ζB.5\zeta_{\ref{lem:v_norm}} such that

ν(xn,y)ζB.5τn.\norm{\nu{\left(x_{n},y\right)}}\leq\zeta_{\ref{lem:v_norm}}\tau_{n}. (143)

We can also prove a deterministic bound on the expectation of ν(xn,Yn+1)\norm{\nu{\left(x_{n},Y_{n+1}\right)}},

𝔼[ν(xn,y)]\displaystyle\mathbb{E}\quantity[\norm{\nu{\left(x_{n},y\right)}}] =𝒪(𝔼[τn+k=1nαkϵk(1)]),\displaystyle=\mathcal{O}\quantity(\mathbb{E}\quantity[\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}]), (144)
=𝒪(τn+k=1nαk𝔼[ϵk(1)]).\displaystyle=\mathcal{O}\quantity(\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}]). (145)

By Lemma B.2, its easy to see that k=1nαk𝔼[ϵk(1)]<\sum_{k=1}^{n}\alpha_{k}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}]<\infty. Therefore, there exists some deterministic constant CB.5C_{\ref{lem:v_norm}} such that

𝔼[ν(xn,y)]CB.5τn.\mathbb{E}\quantity[\norm{\nu{\left(x_{n},y\right)}}]\leq C_{\ref{lem:v_norm}}\tau_{n}. (146)

Although the two statements in Lemma B.5 appear similar, their difference is crucial. Assumption 2.5 and (10) only ensure the existence of a sample-path dependent constant ζB.5\zeta_{\ref{lem:v_norm}} but its form is unknown, preventing its use for expectations or explicit bounds. In contrast, using (11) from Assumption 2.5, we derive a universal constant CB.5C_{\ref{lem:v_norm}}.

Lemma B.6.

For each {Mn}\quantity{M_{n}}, defined in (15), we have

Mn+1ζB.6τn,\norm{M_{n+1}}\leq\zeta_{\ref{lem:M_norm_bound}}\tau_{n}, (147)

where ζB.6\zeta_{\ref{lem:M_norm_bound}} is a the sample-path dependent constant.

Proof.

Applying \norm{\cdot} to (15) gives

Mn+1\displaystyle\norm{M_{n+1}} =ν(xn,Yn+2)Pν(xn,Yn+1),\displaystyle=\norm{\nu{\left(x_{n},Y_{n+2}\right)}-P\nu{\left(x_{n},Y_{n+1}\right)}}, (148)
Pν(xn,Yn+1)+ν(xn,Yn+2),\displaystyle\leq\norm{P\nu{\left(x_{n},Y_{n+1}\right)}}+\norm{\nu{\left(x_{n},Y_{n+2}\right)}}, (149)
=y𝒴P(Yn+1,y)ν(xn,y)+ν(xn,Yn+2),\displaystyle=\norm{\sum_{y^{\prime}\in\mathcal{Y}}P(Y_{n+1},y^{\prime})\nu(x_{n},y^{\prime})}+\norm{\nu{\left(x_{n},Y_{n+2}\right)}}, (150)
y𝒴P(Yn+1,y)ν(xn,y)+ν(xn,Yn+2),\displaystyle\leq\sum_{y^{\prime}\in\mathcal{Y}}\norm{P(Y_{n+1},y^{\prime})\nu(x_{n},y^{\prime})}+\norm{\nu{\left(x_{n},Y_{n+2}\right)}}, (151)
=(maxy𝒴ν(xn,y))y𝒴|P(Yn+1,y)|+ν(xn,Yn+2),\displaystyle=\quantity(\max_{y\in\mathcal{Y}}\norm{\nu(x_{n},y)})\sum_{y^{\prime}\in\mathcal{Y}}\absolutevalue{P(Y_{n+1},y^{\prime})}+\norm{\nu(x_{n},Y_{n+2})}, (152)
2maxy𝒴ν(xn,y)\displaystyle\leq 2\max_{y\in\mathcal{Y}}\norm{\nu(x_{n},y)} (153)

Under Assumption 2.5, we can apply the sample-path dependent bound from Lemma B.5,

Mn+1\displaystyle\norm{M_{n+1}} 2ζB.5τn,\displaystyle\leq 2\zeta_{\ref{lem:v_norm}}\tau_{n}, (Lemma B.5) (154)
=ζB.6τn,\displaystyle=\zeta_{\ref{lem:M_norm_bound}}\tau_{n}, (155)

with ζB.62ζB.5\zeta_{\ref{lem:M_norm_bound}}\doteq 2\zeta_{\ref{lem:v_norm}}. ∎

Lemma B.7.

For each {Mn}\quantity{M_{n}}, defined in (15), we have

𝔼[Mn+12n+1]CB.7(1+xn2),\mathbb{E}\quantity[\norm{M_{n+1}}^{2}\mid\mathcal{F}_{n+1}]\leq C_{\ref{lem: M second moment}}^{\prime}(1+\norm{x_{n}}^{2}), (156)

and

𝔼[Mn+122]CB.72τn2,\mathbb{E}\quantity[\norm{M_{n+1}}_{2}^{2}]\leq C_{\ref{lem: M second moment}}^{2}\tau_{n}^{2}, (157)

where CB.7C_{\ref{lem: M second moment}}^{\prime} and CB.7C_{\ref{lem: M second moment}} are deterministic constants and

n+1σ(x0,Y1,,Yn+1)\displaystyle\mathcal{F}_{n+1}\doteq\sigma(x_{0},Y_{1},\dots,Y_{n+1}) (158)

is the σ\sigma-algebra until time n+1n+1.

Proof.

First, to prove (156), we have

𝔼[Mn+12n+1]\displaystyle\mathbb{E}\quantity[\norm{M_{n+1}}^{2}\mid\mathcal{F}_{n+1}] 4maxy𝒴ν(xn,y)2=𝒪(1+xn2),\displaystyle\leq 4\max_{y\in\mathcal{Y}}\norm{\nu(x_{n},y)}^{2}=\mathcal{O}\quantity(1+\norm{x_{n}}^{2}), (159)

where the first inequality results form (153) in Lemma B.6 and the second inequality results from Lemma B.4.

Then, to prove (157), from Lemma B.3 we then have,

𝔼[ν(xn,y)2]\displaystyle\mathbb{E}\quantity[\norm{\nu\quantity(x_{n},y)}^{2}] 𝔼[1+(CB.3τn+k=1nαkϵk(1))2]=𝒪(τn2+𝔼[(k=1nαkϵk(1))2]).\displaystyle\leq\mathbb{E}\quantity[1+\quantity(C_{\ref{lem:xn_norm}}\tau_{n}+\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}})^{2}]=\mathcal{O}\quantity(\tau_{n}^{2}+\mathbb{E}\quantity[\quantity(\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}})^{2}]). (160)

Recall that by Assumption 2.5, 𝔼[ϵk(1)2]=𝒪(1k)\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2}]=\mathcal{O}\quantity(\frac{1}{k}). Examining the right-most term we then have,

𝔼[(k=1nαkϵk(1))2]\displaystyle\mathbb{E}\quantity[\quantity(\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}})^{2}] 𝔼[(k=1nαk)(k=1nαkϵk(1)2)],\displaystyle\leq\mathbb{E}\quantity[\quantity(\sum_{k=1}^{n}\alpha_{k})\quantity(\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}^{2})], (Cauchy-Schwarz) (161)
=𝒪(k=1nαk),\displaystyle=\mathcal{O}\quantity(\sum_{k=1}^{n}\alpha_{k}), (By (107) in Lemma B.2) (162)
=𝒪(11α1k=1nαk(1α1)),\displaystyle=\mathcal{O}\quantity(\frac{1}{1-\alpha_{1}}\sum_{k=1}^{n}\alpha_{k}(1-\alpha_{1})), (163)
=𝒪(k=1nαk(1αk));\displaystyle=\mathcal{O}\quantity(\sum_{k=1}^{n}\alpha_{k}(1-\alpha_{k})); (164)
=𝒪(τn).\displaystyle=\mathcal{O}\quantity(\tau_{n}). (165)

We then have

𝔼[ν(xn,y)2]=𝒪(τn2).\displaystyle\mathbb{E}\quantity[\norm{\nu\quantity(x_{n},y)}^{2}]=\mathcal{O}(\tau_{n}^{2}). (166)

Because our bound on 𝔼[ν(xn,y)2]\mathbb{E}\quantity[\norm{\nu\quantity(x_{n},y)}^{2}] is independent of yy, we have

𝔼[Mn+12]\displaystyle\mathbb{E}\quantity[\norm{M_{n+1}}^{2}] =𝒪(𝔼[ν(xn,y)2])=𝒪(τn2).\displaystyle=\mathcal{O}\quantity(\mathbb{E}\quantity[\norm{\nu(x_{n},y)}^{2}])=\mathcal{O}(\tau_{n}^{2}). (By (166)) (167)

Due to the equivalence of norms in finite-dimensional spaces, there exists a deterministic constant CB.7C_{\ref{lem: M second moment}} such that (157) holds. ∎

Now, we are ready to present four additional lemmas which we will use to bound the four noise terms in (27).

Lemma B.8.

With {M¯¯n}\quantity{{\overline{\overline{M}}}_{n}} defined in (27),

limnM¯¯n<, a.s. \lim_{n\rightarrow\infty}{\overline{\overline{M}}}_{n}<\infty,\mbox{\quad a.s.\quad} (168)
Proof.

We first observe that the sequence {M¯¯n}\quantity{{\overline{\overline{M}}}_{n}} defined in (27) is positive and monotonically increasing. Therefore by the monotone convergence theorem, it converges almost surely to a (possibly infinite) limit which we denote as,

M¯¯limnM¯¯n a.s. {\overline{\overline{M}}}_{\infty}\doteq\lim_{n\rightarrow\infty}{\overline{\overline{M}}}_{n}\mbox{\quad a.s.\quad} (169)

Then, we will utilize a generalization of Lebesgue’s monotone convergence theorem (Lemma A.4) to prove that the limit M¯¯{\overline{\overline{M}}}_{\infty} is finite almost surely. From Lemma A.4, we see that

𝔼[M¯¯]=limn𝔼[M¯¯n].\displaystyle\mathbb{E}\quantity[{\overline{\overline{M}}}_{\infty}]=\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}]. (170)

Therefore, to prove that M¯¯{\overline{\overline{M}}}_{\infty} is almost surely finite, it is sufficient to prove that limn𝔼[M¯¯n]<\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}]<\infty. To this end, we proceed by bounding the expectation of {M¯¯n}\quantity{{\overline{\overline{M}}}_{n}}, by first starting with {M¯n}\quantity{{\overline{M}}_{n}} from (25). We have,

𝔼[M¯n]\displaystyle\mathbb{E}\left[\norm{{\overline{M}}_{n}}\right] =𝔼[i=1nαi,nMi],\displaystyle=\mathbb{E}\left[\norm{\sum_{i=1}^{n}\alpha_{i,n}M_{i}}\right], (171)
=𝒪(𝔼[i=1nαi,nMi22]),\displaystyle=\mathcal{O}\quantity(\sqrt{\mathbb{E}\left[\norm{\sum_{i=1}^{n}\alpha_{i,n}M_{i}}_{2}^{2}\right]}), (Jensen’s Ineq.) (172)
=𝒪(i=1nαi,n2𝔼[Mi22]),\displaystyle=\mathcal{O}\quantity(\sqrt{\sum_{i=1}^{n}\alpha_{i,n}^{2}\mathbb{E}\left[\norm{M_{i}}_{2}^{2}\right]}), (MiM_{i} is a Martingale Difference Series) (173)
=𝒪(i=1nαi,n2τi2),\displaystyle=\mathcal{O}\quantity(\sqrt{\sum_{i=1}^{n}\alpha_{i,n}^{2}\tau_{i}^{2}}), (Lemma B.7) (174)

Then using the definition of {M¯¯n}\quantity{{\overline{\overline{M}}}_{n}} from (27), we have

𝔼[M¯¯n]\displaystyle\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}] =i=1nαi𝔼[M¯i1]=𝒪(i=1nαij=1i1αj,i12τj12).\displaystyle=\sum_{i=1}^{n}\alpha_{i}\mathbb{E}\left[\norm{{\overline{M}}_{i-1}}\right]=\mathcal{O}\quantity(\sum_{i=1}^{n}\alpha_{i}\sqrt{\sum_{j=1}^{i-1}\alpha_{j,i-1}^{2}\tau_{j-1}^{2}}). (175)

Then, by (79) in Lemma B.1, we have

supn𝔼[M¯¯n]<,\sup_{n}\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}]<\infty, (176)

and since {𝔼[M¯¯n]}\quantity{\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}]} is also monotonically increasing, we have

limn𝔼[M¯¯n]<,\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{M}}}_{n}]<\infty, (177)

which implies that M¯¯<{\overline{\overline{M}}}_{\infty}<\infty almost surely. ∎

Lemma B.9.

With {ϵ¯¯n(1)}\quantity{{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}} defined in (27),

limnϵ¯¯n(1)<, a.s. \lim_{n\rightarrow\infty}{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}<\infty,\ \mbox{\quad a.s.\quad} (178)
Proof.

We first observe that the sequence {ϵ¯¯n(1)}\quantity{{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}} defined in (27) is positive and monotonically increasing. Therefore by the monotone convergence theorem, it converges almost surely to a (possibly infinite) limit which we denote as,

ϵ¯¯(1)limnϵ¯¯n(1) a.s. {\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{\infty}\doteq\lim_{n\rightarrow\infty}{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}\mbox{\quad a.s.\quad} (179)

Then, we utilize a generalization of Lebesgue’s monotone convergence theorem (Lemma A.4) to prove that the limit ϵ¯¯(1){\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{\infty} is finite almost surely. By Lemma A.4, we have

𝔼[ϵ¯¯(1)]=limn𝔼[ϵ¯¯n(1)].\displaystyle\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{\infty}]=\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}]. (180)

Therefore, to prove that ϵ¯¯(1){\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{\infty} is almost surely finite, it is sufficient to prove that limn𝔼[ϵ¯¯n(1)]<\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}]<\infty. To this end, we proceed by bounding the expectation of {ϵ¯¯n(1)}\quantity{{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}},

𝔼[ϵ¯¯n(1)]=i=1nαi𝔼[ϵ¯i1(1)]i=1nαij=1i1αj,i1𝔼[ϵj(1)].\displaystyle\mathbb{E}\left[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}\right]=\sum_{i=1}^{n}\alpha_{i}\mathbb{E}\left[\norm{{\overline{\epsilon}}^{{\left(1\right)}}_{i-1}}\right]\leq\sum_{i=1}^{n}\alpha_{i}\sum_{j=1}^{i-1}\alpha_{j,i-1}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{j}}]. (181)

Then, by (109) in Lemma B.2, we have,

supn𝔼[ϵ¯¯n(1)]<,\sup_{n}\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}]<\infty, (182)

and since {𝔼[ϵ¯¯n(1)]}\quantity{\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}]} is also monotonically increasing, we have

limn𝔼[ϵ¯¯n(1)]<.\lim_{n\rightarrow\infty}\mathbb{E}\quantity[{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{n}]<\infty. (183)

which implies that ϵ¯¯(1)<{\overline{\overline{\epsilon}}}^{{\left(1\right)}}_{\infty}<\infty almost surely.

Lemma B.10.

With {ϵ¯¯n(3)}\quantity{{\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n}} defined in (27), we have

limnϵ¯¯n(3)<, a.s. \lim_{n\rightarrow\infty}\ {\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n}<\infty,\mbox{\quad a.s.\quad} (184)
Proof.

Beginning with the definition of ϵ¯n(3){\overline{\epsilon}}^{{\left(3\right)}}_{n} in (25), we have

ϵ¯n(3)\displaystyle\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}} =i=1nαi,n(ν(xi,Yi+1)ν(xi1,Yi+1)),\displaystyle=\norm{\sum_{i=1}^{n}\alpha_{i,n}{\left(\nu{\left(x_{i},Y_{i+1}\right)}-\nu{\left(x_{i-1},Y_{i+1}\right)}\right)}}, (185)
i=1nαi,nν(xi,Yi+1)ν(xi1,Yi+1),\displaystyle\leq\sum_{i=1}^{n}\alpha_{i,n}\norm{\nu{\left(x_{i},Y_{i+1}\right)}-\nu{\left(x_{i-1},Y_{i+1}\right)}}, (186)
CB.4i=1nαi,nxixi1,\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{i=1}^{n}\alpha_{i,n}\norm{x_{i}-x_{i-1}}, (Lemma B.4) (187)
CB.4i=1nαi,nαi(H(xi1,Yi)+xi1+ϵi(1)),\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\quantity(\norm{H{\left(x_{i-1},Y_{i}\right)}}+\norm{x_{i-1}}+\norm{{\epsilon}^{{\left(1\right)}}_{i}}), (By (SKM with Markovian and Additive Noise)) (188)
CB.4i=1nαi,nαi(2xi1+CH+ϵi(1)),\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\quantity(2\norm{x_{i-1}}+C_{H}+\norm{{\epsilon}^{{\left(1\right)}}_{i}}), (By (3)) (189)
CB.4i=1nαi,nαi(2CB.3τi1+2k=1i1αkϵk(1)+CH+ϵi(1)),\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\quantity(2C_{\ref{lem:xn_norm}}\tau_{i-1}+2\sum_{k=1}^{i-1}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}+C_{H}+\norm{{\epsilon}^{{\left(1\right)}}_{i}}), (Lemma B.3) (190)

Because Assumption 2.5 assures us that k=1αkϵk(1)\sum_{k=1}^{\infty}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}} is almost surely finite, then there exists some sample-path dependent constant we denote as ζB.10\zeta_{\ref{lem:sup_e3}} where,

ϵ¯n(3)\displaystyle\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}} ζB.10i=1nαi,nαi(τi1+ϵi(1)),\displaystyle\leq\zeta_{\ref{lem:sup_e3}}\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\quantity(\tau_{i-1}+\norm{{\epsilon}^{{\left(1\right)}}_{i}}), (Assumption 2.5) (191)
ζB.10(i=1nαi,nαiτi+i=1nαi,nαiϵi(1)),\displaystyle\leq\zeta_{\ref{lem:sup_e3}}\quantity(\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\tau_{i}+\sum_{i=1}^{n}\alpha_{i,n}\alpha_{i}\norm{{\epsilon}^{{\left(1\right)}}_{i}}), (τi\tau_{i} is increasing) (192)
ζB.10αn(i=1nαiτi+i=1nαiϵi(1)).\displaystyle\leq\zeta_{\ref{lem:sup_e3}}\alpha_{n}\quantity(\sum_{i=1}^{n}\alpha_{i}\tau_{i}+\sum_{i=1}^{n}\alpha_{i}\norm{{\epsilon}^{{\left(1\right)}}_{i}}). (Lemma A.2).\displaystyle\text{(Lemma \ref{lem:bravo b1})}. (193)

Again, from Assumption 2.5 we can conclude that there exists some other sample-path dependent constant we denote as ζB.10\zeta_{\ref{lem:sup_e3}}^{\prime} where

ϵ¯n(3)ζB.10αni=1nαiτi.\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}\leq\zeta_{\ref{lem:sup_e3}}^{\prime}\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i}. (194)

Therefore, from the definition of ϵ¯¯n(3){\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n} in (22)

ϵ¯¯n(3)ζB.10i=1nαi2j=1i1αjτj.{\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n}\leq\zeta_{\ref{lem:sup_e3}}^{\prime}\sum_{i=1}^{n}\alpha_{i}^{2}\sum_{j=1}^{i-1}\alpha_{j}\tau_{j}. (195)

So, by (78) in Lemma B.1

supnϵ¯¯n(3)supnζB.10i=1nαi2j=1i1αjτj< a.s. \sup_{n}\ {\overline{\overline{\epsilon}}}^{{\left(3\right)}}_{n}\leq\sup_{n}\ \zeta_{\ref{lem:sup_e3}}^{\prime}\sum_{i=1}^{n}\alpha_{i}^{2}\sum_{j=1}^{i-1}\alpha_{j}\tau_{j}<\infty\mbox{\quad a.s.\quad} (196)

Then, the monotone convergence theorem proves the lemma. ∎

To prove (23) holds almost surely, we introduce four lemmas which we will subsequently use to prove an extension of Theorem 2 from (Borkar, 2009) in Section D.

Lemma B.11.

We have

supnk=1nαkMk< a.s. \sup_{n}\norm{\sum_{k=1}^{n}\alpha_{k}M_{k}}<\infty\mbox{\quad a.s.\quad} (197)
Proof.

Recall that MkM_{k} is a Martingale difference series. Then, the Martingale sequence

{k=1nαkMk}\quantity{\sum_{k=1}^{n}\alpha_{k}M_{k}} (198)

is bounded in L2L^{2} with,

𝔼[k=1nαkMk2]\displaystyle\mathbb{E}\left[\norm{\sum_{k=1}^{n}\alpha_{k}M_{k}}_{2}\right] 𝔼[k=1nαkMk22],\displaystyle\leq\sqrt{\mathbb{E}\left[\norm{\sum_{k=1}^{n}\alpha_{k}M_{k}}_{2}^{2}\right]}, (Jensen’s Ineq.) (199)
=k=1nαk2𝔼[Mk22],\displaystyle=\sqrt{\sum_{k=1}^{n}\alpha_{k}^{2}\mathbb{E}\left[\norm{M_{k}}_{2}^{2}\right]}, (MiM_{i} is a Martingale Difference Series) (200)
CB.7k=1nαk2τk2.\displaystyle\leq C_{\ref{lem: M second moment}}\sqrt{\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}^{2}}. (Lemma B.7) (201)

Lemma B.1 then gives

supnCB.7k=1nαk2τk2\displaystyle\sup_{n}\ C_{\ref{lem: M second moment}}\sqrt{\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}^{2}} <\displaystyle<\infty (202)

Doob’s martingale convergence theorem implies that {k=1nαkMk}\quantity{\sum_{k=1}^{n}\alpha_{k}M_{k}} converges to an almost surely finite random variable, which proves the lemma. ∎

Lemma B.12.

We have,

supnk=1nαkϵk(2)< a.s.\displaystyle\sup_{n}\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}}<\infty\mbox{\quad a.s.\quad} (203)
Proof.

Utilizing the definition of ϵk(2){\epsilon}^{{\left(2\right)}}_{k} in (16), we have

k=1nαkϵk(2)\displaystyle\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k} =k=1nαk(ν(xk,Yk+1)ν(xk1,Yk)),\displaystyle=-\sum_{k=1}^{n}\alpha_{k}{\left(\nu{\left(x_{k},Y_{k+1}\right)}-\nu{\left(x_{k-1},Y_{k}\right)}\right)}, (204)
=k=1nαkν(xk,Yk+1)αk1ν(xk1,Yk)+αk1ν(xk1,Yk)αkν(xk1,Yk),\displaystyle=-\sum_{k=1}^{n}\alpha_{k}\nu{\left(x_{k},Y_{k+1}\right)}-\alpha_{k-1}\nu{\left(x_{k-1},Y_{k}\right)}+\alpha_{k-1}\nu{\left(x_{k-1},Y_{k}\right)}-\alpha_{k}\nu{\left(x_{k-1},Y_{k}\right)}, (205)
=αnν(xn,Yn+1)k=1n(αk1αk)ν(xk1,Yk).\displaystyle=-\alpha_{n}\nu{\left(x_{n},Y_{n+1}\right)}-\sum_{k=1}^{n}{\left(\alpha_{k-1}-\alpha_{k}\right)}\nu{\left(x_{k-1},Y_{k}\right)}. (α0\alpha_{0} = 0) (206)

The triangle inequality gives

k=1nαkϵk(2)\displaystyle\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}} αnν(xn,Yn+1)+k=1n|αk1αk|ν(xk1,Yk),\displaystyle\leq\alpha_{n}\norm{\nu{\left(x_{n},Y_{n+1}\right)}}+\sum_{k=1}^{n}\left|\alpha_{k-1}-\alpha_{k}\right|\norm{\nu{\left(x_{k-1},Y_{k}\right)}}, (207)
ζB.5(αnτn+k=1n|αk1αk|τk1),\displaystyle\leq\zeta_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\sum_{k=1}^{n}\left|\alpha_{k-1}-\alpha_{k}\right|\tau_{k-1}\right)}, (Lemma B.5) (208)
=ζB.5(αnτn+α1τ1+k=1n1|αkαk+1|τk)\displaystyle=\zeta_{\ref{lem:v_norm}}\quantity(\alpha_{n}\tau_{n}+\alpha_{1}\tau_{1}+\sum_{k=1}^{n-1}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k}) (α00).\displaystyle\text{($\alpha_{0}\doteq 0$)}. (209)

Its easy to see that limnαnτn=0\lim_{n\rightarrow\infty}\alpha_{n}\tau_{n}=0, and α1τ1\alpha_{1}\tau_{1} is simply a deterministic and finite constant. Therefore, by Lemma B.1 we have

supnk=1n|αkαk+1|τk\displaystyle\sup_{n}\sum_{k=1}^{n}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k} < a.s.\displaystyle<\infty\mbox{\quad a.s.\quad} (210)

which proves the lemma.

Lemma B.13.

We have,

supnk=1nαkϵk(3)< a.s. \sup_{n}\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}}<\infty\mbox{\quad a.s.\quad} (211)
Proof.

Utilizing the definition of ϵk(3){\epsilon}^{{\left(3\right)}}_{k} in (17), we have

k=1nαkϵk(3)\displaystyle\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}} =k=1nαk(ν(xk,Yk+1)ν(xk1,Yk+1)),\displaystyle=\norm{\sum_{k=1}^{n}\alpha_{k}\quantity(\nu\quantity(x_{k},Y_{k+1})-\nu\quantity(x_{k-1},Y_{k+1}))}, (212)
k=1nαkν(xk,Yk+1)ν(xk1,Yk+1),\displaystyle\leq\sum_{k=1}^{n}\alpha_{k}\norm{\nu\quantity(x_{k},Y_{k+1})-\nu\quantity(x_{k-1},Y_{k+1})}, (213)
CB.4k=1nαkxkxk1,\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k}\norm{x_{k}-x_{k-1}}, (Lemma B.4) (214)
CB.4k=1nαk2(H(xk1,Yk)+xk1+ϵk(1)),\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k}^{2}\quantity(\norm{H{\left(x_{k-1},Y_{k}\right)}}+\norm{x_{k-1}}+\norm{{\epsilon}^{{\left(1\right)}}_{k}}), (215)
(By (SKM with Markovian and Additive Noise)) (216)
CB.4k=1nαk2(2xk1+CH+ϵk(1)),\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k}^{2}\quantity(2\norm{x_{k-1}}+C_{H}+\norm{{\epsilon}^{{\left(1\right)}}_{k}}), (By (3)) (217)
CB.4k=1nαk2(2CB.3τk1+2i=1k1αiϵi(1)+CH+ϵk(1)).\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k}^{2}\quantity(2C_{\ref{lem:xn_norm}}\tau_{k-1}+2\sum_{i=1}^{k-1}\alpha_{i}\norm{{\epsilon}^{{\left(1\right)}}_{i}}+C_{H}+\norm{{\epsilon}^{{\left(1\right)}}_{k}}). (Lemma B.3) (218)

Because Assumption 2.5 assures us that k=1αkϵk(1)\sum_{k=1}^{\infty}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}} is finite, then there exists some sample-path dependent constant we denote as ζB.13\zeta_{\ref{lem:ak_e3}} where,

k=1nαkϵk(3)\displaystyle\norm{\sum_{k=1}^{n}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}} ζB.13k=1nαk2(τk1+ϵk(1)),\displaystyle\leq\zeta_{\ref{lem:ak_e3}}\sum_{k=1}^{n}\alpha_{k}^{2}\quantity(\tau_{k-1}+\norm{{\epsilon}^{{\left(1\right)}}_{k}}), (Assumption 2.5) (219)
ζB.13(k=1nαk2τk+k=1nαk2ϵk(1)),\displaystyle\leq\zeta_{\ref{lem:ak_e3}}\quantity(\sum_{k=1}^{n}\alpha_{k}^{2}\tau_{k}+\sum_{k=1}^{n}\alpha_{k}^{2}\norm{{\epsilon}^{{\left(1\right)}}_{k}}), (τk\tau_{k} is increasing) (220)

Lemma B.1 and Assumption 2.5 then prove the lemma. ∎

Lemma B.14.

Let UnU_{n} be the iterates defined in (21). Then if supnUn<\sup_{n}\norm{U_{n}}<\infty, we have Un0U_{n}\rightarrow 0 almost surely.

Proof.

We use a stochastic approximation argument to show that Un0U_{n}\rightarrow 0. The almost sure convergence of Un0U_{n}\rightarrow 0 is given by a generalization of Theorem 2.1 of (Borkar, 2009), which we present as Theorem D.6 in Appendix D for completeness.

We now verify the assumptions of Theorem D.6. Beginning with the definition of ξk\xi_{k} in (18), we have

limnsupjnk=njαkξk\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}\xi_{k}} =limnsupjnk=njαk(ϵk(1)+ϵk(2)+ϵk(3)),\displaystyle=\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\left({\epsilon}^{{\left(1\right)}}_{k}+{\epsilon}^{{\left(2\right)}}_{k}+{\epsilon}^{{\left(3\right)}}_{k}\right)}}, (221)
limnsupjnk=njαkϵk(1)S1+limnsupjnk=njαkϵk(2)S2+limnsupjnk=njαkϵk(3)S3.\displaystyle\leq\underbrace{\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(1\right)}}_{k}}}_{S_{1}}+\underbrace{\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}}}_{S_{2}}+\underbrace{\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}}}_{S_{3}}. (222)

We now bound the three terms in the RHS.

For S1S_{1}, we have

limnsupjnk=njαkϵk(1)limnsupjnk=njαkϵk(1)limnk=nαkϵk(1)=0,\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(1\right)}}_{k}}\leq\lim_{n\to\infty}\sup_{j\geq n}\sum_{k=n}^{j}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}\leq\lim_{n\to\infty}\sum_{k=n}^{\infty}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}}=0, (223)

where we have used the fact that the series k=1nαkϵk(1)\sum_{k=1}^{n}\alpha_{k}\norm{{\epsilon}^{{\left(1\right)}}_{k}} converges by Assumption 2.5 almost surely.

For S2S_{2}, from (206) in Lemma B.12, we have

k=njαkϵk(2)\displaystyle\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k} =k=1jαkϵk(2)k=1n1αkϵk(2),\displaystyle=\sum_{k=1}^{j}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}-\sum_{k=1}^{n-1}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}, (224)
=αn1ν(xn,Yn)αjν(xj,Yj+1)k=nj(αk1αk)ν(xk1,Yk).\displaystyle=\alpha_{n-1}\nu(x_{n},Y_{n})-\alpha_{j}\nu(x_{j},Y_{j+1})-\sum_{k=n}^{j}(\alpha_{k-1}-\alpha_{k})\nu(x_{k-1},Y_{k}). (225)

Taking the norm and applying the triangle inequality, we have

limnsupjnk=njαkϵk(2)\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}} limnsupjn(αn1ν(xn,Yn)+αjν(xj,Yj+1)\displaystyle\leq\lim_{n\to\infty}\sup_{j\geq n}\bigg{(}\alpha_{n-1}\norm{\nu(x_{n},Y_{n})}+\alpha_{j}\norm{\nu(x_{j},Y_{j+1})} (226)
+k=nj(αk1αk)ν(xk1,Yk)),\displaystyle\quad+\sum_{k=n}^{j}\norm{(\alpha_{k-1}-\alpha_{k})\nu(x_{k-1},Y_{k})}\bigg{)}, (227)
limnsupjnζB.5(αn1τn1+αjτj+k=n|αk1αk|τk1),\displaystyle\leq\lim_{n\to\infty}\sup_{j\geq n}\zeta_{\ref{lem:v_norm}}\quantity(\alpha_{n-1}\tau_{n-1}+\alpha_{j}\tau_{j}+\sum_{k=n}^{\infty}\absolutevalue{\alpha_{k-1}-\alpha_{k}}\tau_{k-1}), (Lemma B.5) (228)

where the last inequality holds because k=nj|αk1αk|τk1\sum_{k=n}^{j}\absolutevalue{\alpha_{k-1}-\alpha_{k}}\tau_{k-1} is monotonically increasing. Note that

αnτn={𝒪(n12b)if45<b<1,𝒪(lognn)ifb=1.\displaystyle\alpha_{n}\tau_{n}=\begin{cases}\mathcal{O}\quantity(n^{1-2b})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\frac{\log n}{n})&\text{if}\quad b=1.\end{cases} (229)

Since we have jnj\geq n, then

limnsupjnk=njαkϵk(2)\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(2\right)}}_{k}} limnζB.5(2αn1τn1+k=n|αk1αk|τk1)=0\displaystyle\leq\lim_{n\to\infty}\zeta_{\ref{lem:v_norm}}\quantity(2\alpha_{n-1}\tau_{n-1}+\sum_{k=n}^{\infty}\absolutevalue{\alpha_{k-1}-\alpha_{k}}\tau_{k-1})=0 (230)

where we used the fact that (77) in Lemma B.1 and the monotone convergence theorem prove that the series k=1n|αkαk+1|τk\sum_{k=1}^{n}\absolutevalue{\alpha_{k}-\alpha_{k+1}}\tau_{k} converges almost surely.

For S3S_{3}, following the steps in Lemma B.13 (which we omit to avoid repetition), we have,

limnsupjnk=njαkϵk(3)\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}} limnsupjnζB.13(k=njαk2τk+k=njαk2ϵk(1)).\displaystyle\leq\lim_{n\to\infty}\sup_{j\geq n}\zeta_{\ref{lem:ak_e3}}\quantity(\sum_{k=n}^{j}\alpha_{k}^{2}\tau_{k}+\sum_{k=n}^{j}\alpha_{k}^{2}\norm{{\epsilon}^{{\left(1\right)}}_{k}}). (231)

which further implies that

limnsupjnk=njαkϵk(3)\displaystyle\lim_{n\to\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}{\epsilon}^{{\left(3\right)}}_{k}} limnζB.13(k=nαk2τk+k=nαk2ϵk(1))=0,\displaystyle\leq\lim_{n\to\infty}\zeta_{\ref{lem:ak_e3}}\quantity(\sum_{k=n}^{\infty}\alpha_{k}^{2}\tau_{k}+\sum_{k=n}^{\infty}\alpha_{k}^{2}\norm{{\epsilon}^{{\left(1\right)}}_{k}})=0, (232)

where we use the fact that, by (74) in Lemma B.1, Assumption 2.5, and the monotone convergence theorem, both series on the RHS series converge almost surely. Therefore we have proven that,

limnsupjnk=njαkξk=0 a.s. \lim_{n\rightarrow\infty}\sup_{j\geq n}\norm{\sum_{k=n}^{j}\alpha_{k}\xi_{k}}=0\mbox{\quad a.s.\quad} (233)

thereby verifying Assumption D.1.

Assumption D.2 is satisfied by (6) which is the result of Assumption 2.2. Assumption D.3 is clearly met by the definition of the deterministic learning rates in Assumption 2.4. Demonstrating Assumption D.4 holds, Lemma B.7 demonstrates {Mn}\quantity{M_{n}} is square-integrable martingale difference series.

Therefore, by Theorem D.6, the iterates {Un}\quantity{U_{n}} converge almost surely to a possibly sample-path dependent compact connected internally chain transitive set of the following ODE:

dU(t)dt=U(t).\displaystyle\derivative{U(t)}{t}=-U(t). (234)

Since the origin is the unique globally asymptotically stable equilibrium point of (234), we have that Un0U_{n}\rightarrow 0 almost surely. ∎

Lemma B.15.

With {xn}\quantity{x_{n}} defined in (18) and {Un}\quantity{U_{n}} defined in (21), if k=1αkUk1\sum_{k=1}^{\infty}\alpha_{k}\norm{U_{k-1}} and limnUn=0\lim_{n\rightarrow\infty}U_{n}=0, then limnxn=x\lim_{n\rightarrow\infty}x_{n}=x_{*} where x𝒳x_{*}\in\mathcal{X}_{*} is a possibly sample-path dependent fixed point.

Proof.

Following the approach of Bravo & Cominetti (2024), we utilize the estimate for inexact Krasnoselskii-Mann iterations of the form (IKM) presented in Lemma A.1 to prove the convergence of (SKM with Markovian and Additive Noise). Using the definition of {Un}\left\{U_{n}\right\} in (21), we then let z0=x0z_{0}=x_{0} and define znxnUnz_{n}\doteq x_{n}-U_{n}, which gives

zn+1\displaystyle z_{n+1} =(1αn+1)xn+αn+1(h(xn)+Mn+1+ξn+1)\displaystyle={\left(1-\alpha_{n+1}\right)}x_{n}+\alpha_{n+1}{\left(h(x_{n})+M_{n+1}+\xi_{n+1}\right)} (235)
((1αn+1)Un+αn+1(Mn+1+ξn+1))\displaystyle\quad\quad-{\left({\left(1-\alpha_{n+1}\right)}U_{n}+\alpha_{n+1}{\left(M_{n+1}+\xi_{n+1}\right)}\right)} (236)
=(1αn+1)zn+αn+1h(xn)\displaystyle=\quantity(1-\alpha_{n+1})z_{n}+\alpha_{n+1}h(x_{n}) (237)
=zn+αn+1(h(zn)zn+en+1)\displaystyle=z_{n}+\alpha_{n+1}{\left(h(z_{n})-z_{n}+e_{n+1}\right)} (238)

which matches the form of (IKM) with en=h(xn1)h(zn1)e_{n}=h{\left(x_{n-1}\right)}-h{\left(z_{n-1}\right)}. Due to the non-expansivity of hh from (6), we have

en+1=h(xn)h(zn)xnzn=Un\norm{e_{n+1}}=\norm{h\quantity(x_{n})-h\quantity(z_{n})}\leq\norm{x_{n}-z_{n}}=\norm{U_{n}} (239)

The convergence of xnx_{n} then follows directly from Lemma A.1 which gives limnzn=x\lim_{n\rightarrow\infty}z_{n}=x_{*} for some x𝒳x_{*}\in\mathcal{X}_{*}, and therefore limnxn=limnzn+Un=x\lim_{n\rightarrow\infty}x_{n}=\lim_{n\rightarrow\infty}z_{n}+U_{n}=x_{*}. We note that here ene_{n} is stochastic while the (IKM) result in Lemma A.1 considers deterministic noise. This means we apply Lemma A.1 for each sample path. ∎

Appendix C Additional Lemmas from Section 3

Corollary C.1.

We have

𝔼[M¯n]CC.1τnαn+1\mathbb{E}\quantity[\norm{{\overline{M}}_{n}}]\leq C_{\ref{lem: M rate}}\tau_{n}\sqrt{\alpha_{n+1}} (240)

where CC.1C_{\ref{lem: M rate}} is a deterministic constant.

Proof.

Starting from (174) from Lemma B.8 to avoid redundancy, we directly have

𝔼[M¯n]\displaystyle\mathbb{E}\left[\norm{{\overline{M}}_{n}}\right] =𝒪(i=1nαi,n2τi2).\displaystyle=\mathcal{O}\quantity(\sqrt{\sum_{i=1}^{n}\alpha_{i,n}^{2}\tau_{i}^{2}}). (241)

Additionally, by Lemma A.3, we have i=1nαi,n2τi2τnαn+1\sqrt{\sum_{i=1}^{n}\alpha_{i,n}^{2}\tau_{i}^{2}}\leq\tau_{n}\sqrt{\alpha_{n+1}}. Therefore, there exists a deterministic constant such that the corollary holds. ∎

Corollary C.2.

We have

𝔼[ϵ¯n(2)]CC.2αnτn\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}]\leq C_{\ref{lem: e2 rate}}\alpha_{n}\tau_{n} (242)

where CC.2C_{\ref{lem: e2 rate}} is a deterministic constant.

Proof.

Starting from (35) to avoid repetition, we have,

ϵ¯n(2)\displaystyle\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}} αnν(xn,Yn+1)+i=1n|αi1,nαi,n|ν(xi1,Yi).\displaystyle\leq\alpha_{n}\norm{\nu{\left(x_{n},Y_{n+1}\right)}}+\sum_{i=1}^{n}\left|\alpha_{i-1,n}-\alpha_{i,n}\right|\norm{\nu{\left(x_{i-1},Y_{i}\right)}}. (243)

Now we can take the expectation and apply the sample-path independent bound from Lemma B.5 with,

𝔼[ϵ¯n(2)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}] CB.5(αnτn+i=1n|αi1,nαi,n|τi1)\displaystyle\leq C_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\sum_{i=1}^{n}\left|\alpha_{i-1,n}-\alpha_{i,n}\right|\tau_{i-1}\right)} (Lemma B.5) (244)
=CB.5(αnτn+k=0n1|αk,nαk+1,n|τk)\displaystyle=C_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\sum_{k=0}^{n-1}\left|\alpha_{k,n}-\alpha_{k+1,n}\right|\tau_{k}\right)} (245)

Lemma B.1 and τk\tau_{k} being monotonically increasing for k1k\geq 1 yields,

𝔼[ϵ¯n(2)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}] CB.5(αnτn+α1,nτ0+τnk=1n1(αk+1,nαk,n)),\displaystyle\leq C_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\alpha_{1,n}\tau_{0}+\tau_{n}\sum_{k=1}^{n-1}{\left(\alpha_{k+1,n}-\alpha_{k,n}\right)}\right)}, (246)
=CB.5(αnτn+α1,n+τn(αn,nα1,n)),\displaystyle=C_{\ref{lem:v_norm}}{\left(\alpha_{n}\tau_{n}+\alpha_{1,n}+\tau_{n}{\left(\alpha_{n,n}-\alpha_{1,n}\right)}\right)}, (τ01\tau_{0}\doteq 1) (247)
=𝒪(αnτn).\displaystyle=\mathcal{O}\quantity(\alpha_{n}\tau_{n}). (Lemma A.2) (248)

Therefore, there exists a deterministic constant we denote as CC.2C_{\ref{lem: e2 rate}} such that

𝔼[ϵ¯n(2)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(2\right)}}_{n}}] CC.2αnτn.\displaystyle\leq C_{\ref{lem: e2 rate}}\alpha_{n}\tau_{n}. (249)

Corollary C.3.

We have

𝔼[ϵ¯n(3)]CC.3αni=1nαiτi.\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}]\leq C_{\ref{lem: e3 rate}}\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i}. (250)
Proof.

Starting with (190) from Lemma B.10 to avoid redundancy, we have

ϵ¯n(3)\displaystyle\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}} CB.4k=1nαk,nαk(2CB.3τk1+2i=1k1αiϵi(1)+CH+ϵk(1)).\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k,n}\alpha_{k}\quantity(2C_{\ref{lem:xn_norm}}\tau_{k-1}+2\sum_{i=1}^{k-1}\alpha_{i}\norm{{\epsilon}^{{\left(1\right)}}_{i}}+C_{H}+\norm{{\epsilon}^{{\left(1\right)}}_{k}}). (251)

Taking the expectation gives,

𝔼[ϵ¯n(3)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}] CB.4k=1nαk,nαk(2CB.3τk1+2i=1k1αi𝔼[ϵi(1)]+CH+𝔼[ϵk(1)]).\displaystyle\leq C_{\ref{lem:v Lipschitz}}\sum_{k=1}^{n}\alpha_{k,n}\alpha_{k}\quantity(2C_{\ref{lem:xn_norm}}\tau_{k-1}+2\sum_{i=1}^{k-1}\alpha_{i}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}]+C_{H}+\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}]). (252)

Recall that τk\tau_{k} is monotonically increasing. Additionally, by Lemma B.2, i=1k1αi𝔼[ϵi(1)]\sum_{i=1}^{k-1}\alpha_{i}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}] converges and limk𝔼[ϵk(1)]=0\lim_{k\rightarrow\infty}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{k}}]=0. Therefore, there exists a deterministic constant CC.3C_{\ref{lem: e3 rate}} such that

𝔼[ϵ¯n(3)]\displaystyle\mathbb{E}\quantity[\norm{{\overline{\epsilon}}^{{\left(3\right)}}_{n}}] CC.3k=1nαk,nαkτk1,\displaystyle\leq C_{\ref{lem: e3 rate}}\sum_{k=1}^{n}\alpha_{k,n}\alpha_{k}\tau_{k-1}, (253)
CC.3αni=1nαiτi\displaystyle\leq C_{\ref{lem: e3 rate}}\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i} (Lemma A.2).\displaystyle\text{(Lemma \ref{lem:bravo b1})}. (254)

Lemma C.4.

For ωn\omega_{n} defined in (51), we have

ωn=𝒪(τnαn+1)\displaystyle\omega_{n}=\mathcal{O}(\tau_{n}\sqrt{\alpha_{n+1}}) (255)
Proof.

From (51), we have

ωnCB.7τnαn+1K1+i=1nαi,n𝔼[ϵi(1)]K2+CC.2αnτnK3+CC.3αni=1nαiτiK4\displaystyle\omega_{n}\doteq\underbrace{C_{\ref{lem: M second moment}}\tau_{n}\sqrt{\alpha_{n+1}}}_{K_{1}}+\underbrace{\sum_{i=1}^{n}\alpha_{i,n}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}]}_{K_{2}}+\underbrace{C_{\ref{lem: e2 rate}}\alpha_{n}\tau_{n}}_{K_{3}}+\underbrace{C_{\ref{lem: e3 rate}}\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i}}_{K_{4}} (256)

To prove the Lemma, we will examine each of the four terms and prove they are 𝒪(τnαn+1)\mathcal{O}(\tau_{n}\sqrt{\alpha_{n+1}}). For K1K_{1}, this is trivial. For K2K_{2}, we first recall from Lemma B.1 that αn=𝒪(1nb)\alpha_{n}=\mathcal{O}(\frac{1}{n^{b}}) and

τn={𝒪(n1b)if45<b<1,𝒪(logn)ifb=1.\displaystyle\tau_{n}=\begin{cases}\mathcal{O}\quantity(n^{1-b})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\log n)&\text{if}\quad b=1.\end{cases} (257)

Then we have,

τnαn+1={𝒪(1n32b1)if45<b<1,𝒪(lognn)ifb=1.\displaystyle\tau_{n}\sqrt{\alpha_{n+1}}=\begin{cases}\mathcal{O}\quantity(\frac{1}{n^{\frac{3}{2}b-1}})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\frac{\log n}{\sqrt{n}})&\text{if}\quad b=1.\end{cases} (258)

Then by Lemma B.2 we have

i=1nαi,n𝔼[ϵi(1)]\displaystyle\sum_{i=1}^{n}\alpha_{i,n}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}] αni=1n𝔼[ϵi(1)],\displaystyle\leq\alpha_{n}\sum_{i=1}^{n}\mathbb{E}\quantity[\norm{{\epsilon}^{{\left(1\right)}}_{i}}], (Lemma A.2) (259)
=𝒪(αni=1n1i),\displaystyle=\mathcal{O}\quantity(\alpha_{n}\sum_{i=1}^{n}\frac{1}{\sqrt{i}}), (260)
=𝒪(αnn)\displaystyle=\mathcal{O}\quantity(\alpha_{n}\sqrt{n}) (261)
=𝒪(1nbn),\displaystyle=\mathcal{O}\quantity(\frac{1}{n^{b}}\sqrt{n}), (262)
=𝒪(1nb1/2)\displaystyle=\mathcal{O}\quantity(\frac{1}{n^{b-1/2}}) (263)

Because we have 32b1b12\frac{3}{2}b-1\leq b-\frac{1}{2} for b(45,1]b\in(\frac{4}{5},1], we can see from (258), that K2K_{2} is dominated by K1K_{1}.

For K3K_{3}, by Lemma B.1 we have,

αnτn={𝒪(1n2b1)if45<b<1,𝒪(lognn)ifb=1.\displaystyle\alpha_{n}\tau_{n}=\begin{cases}\mathcal{O}\quantity(\frac{1}{n^{2b-1}})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\frac{\log n}{n})&\text{if}\quad b=1.\end{cases} (264)

It is clear from (258), K3K_{3} is dominated by K1K_{1}.

For K4K_{4}, for the case when b=1b=1, we have

αni=1nαiτi\displaystyle\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i} αnτni=1nαi\displaystyle\leq\alpha_{n}\tau_{n}\sum_{i=1}^{n}\alpha_{i} (τn\tau_{n} increasing) (265)
=𝒪(lognni=1n1i),\displaystyle=\mathcal{O}\quantity(\frac{\log n}{n}\sum_{i=1}^{n}\frac{1}{i}), (266)
=𝒪(log2nn).\displaystyle=\mathcal{O}\quantity(\frac{\log^{2}n}{n}). (267)

For the case when 45<b<1\frac{4}{5}<b<1, we have

αni=1nαiτi\displaystyle\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i} =𝒪(1nbi=1n1i2b1)\displaystyle=\mathcal{O}\quantity(\frac{1}{n^{b}}\sum_{i=1}^{n}\frac{1}{i^{2b-1}}) (268)

which we can approximate by an integral,

1n1x2b1𝑑x=𝒪(n22b).\displaystyle\int_{1}^{n}\frac{1}{x^{2b-1}}\ dx=\mathcal{O}\quantity(n^{2-2b}). (269)

Therefore,

αni=1nαiτi\displaystyle\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i} =𝒪(n23b)\displaystyle=\mathcal{O}\quantity(n^{2-3b}) (270)

Combining our results from the two cases, we have for K4K_{4}

αni=1nαiτi={𝒪(1n3b2)if45<b<1,𝒪(log2nn)ifb=1.\displaystyle\alpha_{n}\sum_{i=1}^{n}\alpha_{i}\tau_{i}=\begin{cases}\mathcal{O}\quantity(\frac{1}{n^{3b-2}})&\text{if}\quad\frac{4}{5}<b<1,\\ \mathcal{O}\quantity(\frac{\log^{2}n}{n})&\text{if}\quad b=1.\end{cases} (271)

Comparing with K1K_{1} in (258), since we have 32b1<3b2\frac{3}{2}b-1<3b-2 for b(45,1)b\in(\frac{4}{5},1), we can see that K4K_{4} is dominated by K1K_{1}, thereby proving the lemma.

Lemma C.5.

We have,

k=2n2αkσ(τnτk)𝔼[Uk1]=𝒪(1/τn).\displaystyle\sum_{k=2}^{n}2\alpha_{k}\sigma\quantity(\tau_{n}-\tau_{k})\mathbb{E}\quantity[\norm{U_{k-1}}]=\mathcal{O}(1/\sqrt{\tau_{n}}). (272)
Proof.

The proof of this Lemma is a straightforward combination of the existing results of Theorems 2.11 and 3.1 from (Bravo & Cominetti, 2024). First, from (51), we have

k=2n2αkσ(τnτk)𝔼[Uk1]k=2n2αkσ(τnτk)ωk1.\displaystyle\sum_{k=2}^{n}2\alpha_{k}\sigma\quantity(\tau_{n}-\tau_{k})\mathbb{E}\quantity[\norm{U_{k-1}}]\leq\sum_{k=2}^{n}2\alpha_{k}\sigma\quantity(\tau_{n}-\tau_{k})\omega_{k-1}. (273)

In the proof of Theorem 2.11 of (Bravo & Cominetti, 2024), they prove that if there exists a decreasing convex function f:(0,)(0,)f:(0,\infty)\rightarrow(0,\infty) of class C2C^{2}, and a constant γ1\gamma\geq 1, such that for k2k\geq 2,

{wk1(1αk)f(τk),αk(1αk)γαk+1(1αk+1),\displaystyle\begin{cases}w_{k-1}\leq(1-\alpha_{k})f(\tau_{k}),\\ \alpha_{k}(1-\alpha_{k})\leq\gamma\alpha_{k+1}(1-\alpha_{k+1}),\end{cases} (274)

then,

k=2n2αkσ(τnτk)ωk12γπτ1τnf(x)τnx𝑑x+2αnwn1.\displaystyle\sum_{k=2}^{n}2\alpha_{k}\sigma\quantity(\tau_{n}-\tau_{k})\omega_{k-1}\leq\frac{2\gamma}{\sqrt{\pi}}\int_{\tau_{1}}^{\tau_{n}}\frac{f(x)}{\sqrt{\tau_{n}-x}}dx+2\alpha_{n}w_{n-1}. (275)

Using the fact that ωn=𝒪(τnαn+1)\omega_{n}=\mathcal{O}(\tau_{n}\sqrt{\alpha_{n+1}}), which aligns with the analogous νn\nu_{n} from Bravo & Cominetti (2024), and adopting their definition of τn\tau_{n}, we avoid redundant derivations here.

Theorem 3.1 in Bravo & Cominetti (2024) establishes that for the step size schedule specified in Assumption 2.4, there exist constants γ1\gamma\geq 1 and a function f(x)f(x) satisfying (274). Specifically, they show with

f(x)={κx(1+x)b/2(1b)if b<1,κxex/2if b=1,\displaystyle f(x)=\begin{cases}\kappa x(1+x)^{-b/2(1-b)}&\text{if }b<1,\\ \kappa xe^{-x/2}&\text{if }b=1,\end{cases} (276)

for some constant κ\kappa and γ=3227\gamma=\frac{32}{27}, (274) is satisfied. Moreover, they demonstrate that the resulting convolution integral in (275) evaluates to 𝒪(1/τn)\mathcal{O}(1/\sqrt{\tau_{n}}).

Combining these results, the right-hand side of (275) simplifies to 𝒪(1/τn)\mathcal{O}(1/\sqrt{\tau_{n}}), which completes the proof. For detailed steps, we refer the reader to Bravo & Cominetti (2024) to avoid repetition. ∎

Appendix D Extension of Theorem 2.1 of Borkar (2009)

In this section, we present a simple extension of Theorem 2 from (Borkar, 2009) for completeness. Readers familiar with stochastic approximation theory should find this extension fairly straightforward. Originally, Chapter 2 of (Borkar, 2009) considers stochastic approximations of the form,

yn+1=yn+αn(h(yn)+Mn+1+ξn+1)\displaystyle y_{n+1}=y_{n}+\alpha_{n}{\left(h(y_{n})+M_{n+1}+\xi_{n+1}\right)} (277)

where it is assumed that ξn0\xi_{n}\rightarrow 0 almost surely. However, our work requires that we remove the assumption that ξn0\xi_{n}\rightarrow 0, and replace it with a more mild condition on the asymptotic rate of change of ξn\xi_{n}, akin to Kushner & Yin (2003).

Assumption D.1.

For any T>0T>0,

limnsupnjm(n,T)i=njαiξi=0 a.s.\displaystyle\lim_{n\to\infty}\sup_{n\leq j\leq m(n,T)}\norm{\sum_{i=n}^{j}\alpha_{i}\xi_{i}}=0\mbox{\quad a.s.\quad} (278)

where m(n,T)min{k|i=nkα(i)T}m(n,T)\doteq\min\quantity{k|\sum_{i=n}^{k}\alpha(i)\geq T}.

The next four assumptions are the same as the remaining assumptions in Chapter 2 of (Borkar, 2009).

Assumption D.2.

The map hh is Lipschitz: h(x)h(y)Lxy\norm{h{\left(x\right)}-h{\left(y\right)}}\leq L\norm{x-y} for some 0<L<0<L<\infty.

Assumption D.3.

The step sizes {αn}\quantity{\alpha_{n}} are positive scalars satisfying

nαn=,nαn2<\displaystyle\sum_{n}\alpha_{n}=\infty,\sum_{n}\alpha_{n}^{2}<\infty (279)
Assumption D.4.

{Mn}\quantity{M_{n}} is a martingale difference sequence w.r.t the increasing family of σ\sigma-algebras

nσ(ym,Mm,mn)=σ(y0,M1,,Mn),n0.\displaystyle\mathcal{F}_{n}\doteq\sigma\quantity(y_{m},M_{m},m\leq n)=\sigma\quantity(y_{0},M_{1},\ldots,M_{n}),\,n\geq 0. (280)

That is,

𝔼[Mn+1|n]=0 a.s. ,n0.\displaystyle\mathbb{E}\left[M_{n+1}|\mathcal{F}_{n}\right]=0\mbox{\quad a.s.\quad},n\geq 0. (281)

Furthermore, {Mn}\quantity{M_{n}} are square-integrable with

𝔼[Mn+12|n]K(1+xn2) a.s. ,n0,\displaystyle\mathbb{E}\left[\norm{M_{n+1}}^{2}|\mathcal{F}_{n}\right]\leq K{\left(1+\norm{x_{n}}^{2}\right)}\ \mbox{\quad a.s.\quad},\ n\geq 0, (282)

for some constant K>0K>0

Assumption D.5.

The iterates of (277) remain bounded almost surely, i.e.,

supnyn<\displaystyle\sup_{n}\norm{y_{n}}<\infty (283)
Theorem D.6 (Extension of Theorem 2.1 from (Borkar, 2009)).

Let Assumptions D.1, D.2, D.3, D.4, D.5 hold. Almost surely, the sequence {yn}\quantity{y_{n}} generated by (277) converges to a (possibly sample-path dependent) compact connected internally chain transitive set of the ODE

dy(t)dt=h(y(t)).\displaystyle\derivative{y(t)}{t}=h(y(t)). (284)
Proof.

We now demonstrate that even with the relaxed assumption on ξn\xi_{n}, we can still achieve the same almost sure convergence of the iterates achieved by (Borkar, 2009). Following Chapter 2 of (Borkar, 2009), we construct a continuous interpolated trajectory y¯(t),t0\bar{y}{\left(t\right)},t\geq 0, and show that it asymptotically approaches the solution set of (284) almost surely. Define time instants t(0)=0,t(n)=m=0n1αm,n1t{\left(0\right)}=0,t{\left(n\right)}=\sum_{m=0}^{n-1}\alpha_{m},n\geq 1. By assumption D.3, t(n)t{\left(n\right)}\uparrow\infty. Let In[t(n),t(n+1)],n0I_{n}\doteq\left[t{\left(n\right)},t{\left(n+1\right)}\right],n\geq 0. Define a continuous, piece-wise linear y¯(t),t0\bar{y}{\left(t\right)},t\geq 0 by y¯(t(n))=yn,n0\bar{y}{{\left(t{\left(n\right)}\right)}}=y_{n},\ n\geq 0, with linear interpolation on each interval InI_{n}:

y¯(t)=yn+(yn+1yn)tt(n)t(n+1)t(n),tIn\bar{y}{\left(t\right)}=y_{n}+{\left(y_{n+1}-y_{n}\right)}\frac{t-t{\left(n\right)}}{t{\left(n+1\right)}-t{\left(n\right)}},t\in I_{n} (285)

It is worth noting that supt0y¯(t)=supnyn<\sup_{t\geq 0}\norm{\bar{y}{\left(t\right)}}=\sup_{n}\norm{y_{n}}<\infty almost surely by Assumption D.5. Let ys(t),ts,y^{s}{\left(t\right)},t\geq s, denote the unique solution to (284)\eqref{eq:Un_ode} ‘starting at s’:

dys(t)dt\displaystyle\derivative{y^{s}(t)}{t} =h(ys(t)),ts,\displaystyle=h{\left(y^{s}{\left(t\right)}\right)},t\geq s, (286)

with ys(s)=y¯(s),sy^{s}{\left(s\right)}=\bar{y}{\left(s\right)},s\in\mathbb{R}. Similarly, let ys(t),ts,y_{s}{\left(t\right)},t\geq s, denote the unique solution to (284)\eqref{eq:Un_ode} ‘ending at s’:

dys(t)dt\displaystyle\derivative{y_{s}(t)}{t} =h(ys(t)),ts,\displaystyle=h{\left(y_{s}{\left(t\right)}\right)},t\leq s, (287)

with ys(s)=y¯(s),sy_{s}{\left(s\right)}=\bar{y}{\left(s\right)},s\in\mathbb{R}. Define also

ζn\displaystyle\zeta_{n} =m=0n1αm(Mm+1+ξm+1),n1\displaystyle=\sum_{m=0}^{n-1}\alpha_{m}{\left(M_{m+1}+\xi_{m+1}\right)},\ n\geq 1 (288)

After invoking Lemma D.7, the analysis and proof presented for Theorem 2 in (Borkar, 2009) applies directly, yielding our desired extended result. ∎

Lemma D.7 (Extension of Theorem 1 from (Borkar, 2009)).

Let D.1 - D.5 hold. We have for any T>0T>0,

lims\displaystyle\lim_{s\rightarrow\infty} supt[s,s+T]y¯(t)ys(t)=0, a.s.\displaystyle\sup_{t\in\left[s,s+T\right]}\norm{\bar{y}{\left(t\right)}-y^{s}{\left(t\right)}}=0,\mbox{\quad a.s.\quad} (289)
lims\displaystyle\lim_{s\rightarrow\infty} supt[s,s+T]y¯(t)ys(t)=0, a.s.\displaystyle\sup_{t\in\left[s,s+T\right]}\norm{\bar{y}{\left(t\right)}-y_{s}{\left(t\right)}}=0,\mbox{\quad a.s.\quad} (290)
Proof.

Let t(n+m)t{\left(n+m\right)} be in [t(n),t(n)+T]\left[t(n),t(n)+T\right]. Let [t]max{t(k):t(k)t}\left[t\right]\doteq\max\quantity{t(k):t(k)\leq t}. Then,

y¯(t(n+m))=y¯(t(n))+k=0m1αn+kh(y¯(t(n+k)))+δn,n+m\displaystyle\bar{y}{\left(t{\left(n+m\right)}\right)}=\bar{y}{\left(t{\left(n\right)}\right)}+\sum_{k=0}^{m-1}\alpha_{n+k}h{\left(\bar{y}{\left(t{\left(n+k\right)}\right)}\right)}+\delta_{n,n+m} (2.1.6 in (Borkar, 2009)) (291)

where δn,n+mζn+mζn\delta_{n,n+m}\doteq\zeta_{n+m}-\zeta_{n}. Borkar (2009) then compares this with

yt(n)(t(n+m))=y¯(t(n))+k=0m1αn+kh(yt(n)(t(n+k)))\displaystyle y^{t(n)}{\left(t{\left(n+m\right)}\right)}=\bar{y}{\left(t{\left(n\right)}\right)}+\sum_{k=0}^{m-1}\alpha_{n+k}h{\left(y^{t(n)}{\left(t{\left(n+k\right)}\right)}\right)} (292)
+t(n)t(n+m)(h(yt(n)(z))h(yt(n)([z])))𝑑z.\displaystyle\ +\int_{t(n)}^{t(n+m)}{\left(h{\left(y^{t(n)}{\left(z\right)}\right)}-h{\left(y^{t(n)}{\left(\left[z\right]\right)}\right)}\right)}dz. (2.1.7 in (Borkar, 2009)) (293)

Next, Borkar (2009) bounds the integral on the right-hand side by proving

t(n)t(n+m)(h(yt(n)(t))h(yt(n)([t])))𝑑tCTLk=0αn+k2n0, a.s.\displaystyle\norm{\int_{t(n)}^{t(n+m)}{\left(h{\left(y^{t(n)}{\left(t\right)}\right)}-h{\left(y^{t(n)}{\left(\left[t\right]\right)}\right)}\right)}dt}\leq C_{T}L\sum_{k=0}^{\infty}\alpha_{n+k}^{2}\xrightarrow{n\uparrow\infty}0,\ \mbox{\quad a.s.\quad} (2.1.8 in (Borkar, 2009)) (294)

where CTh(0)+L(C0+h(0)T)eLT<C_{T}\doteq\norm{h{\left(0\right)}}+L{\left(C_{0}+\norm{h{\left(0\right)}}T\right)}e^{LT}<\infty almost surely and C0supnyn<C_{0}\doteq\sup_{n}\norm{y_{n}}<\infty a.s. by Assumption D.5.

Then, we can subtract (2.1.7) from (2.1.6) and take norms, yielding

y¯(t(n+m))yt(n)(t(n+m))\displaystyle\norm{\bar{y}{\left(t{\left(n+m\right)}\right)}-y^{t(n)}{\left(t{\left(n+m\right)}\right)}} Li=0m1αn+iy¯(t(n+i))yt(n)(t(n+i))\displaystyle\leq L\sum_{i=0}^{m-1}\alpha_{n+i}\norm{\bar{y}{\left(t{\left(n+i\right)}\right)}-y^{t(n)}{\left(t{\left(n+i\right)}\right)}} (295)
+CTLk0αn+k2+sup0km(n,T)δn,n+k.\displaystyle\quad+C_{T}L\sum_{k\geq 0}\alpha_{n+k}^{2}+\sup_{0\leq k\leq m(n,T)}\norm{\delta_{n,n+k}}. (296)

The key difference between (296) and the analogous equation in Borkar (2009) Chapter 2, is that we replace the supk0\sup_{k\geq 0} with a sup0km(n,T)\sup_{0\leq k\leq m(n,T)}. The reason we can make this change is that we defined t(n+m)t(n+m) to be in the range [t(n),t(n)+T]\left[t(n),t(n)+T\right]. Recall that we also defined m(n,T)min{k|i=nkα(i)T}m(n,T)\doteq\min\quantity{k|\sum_{i=n}^{k}\alpha(i)\geq T} in Assumption D.1, so we therefore know that mm(n,T)m\leq m(n,T) in (291). Borkar (2009) unnecessarily relaxes this for notation simplicity, but a similar argument can be found in (Kushner & Yin, 2003).

Also, we have,

δn,n+k\displaystyle\norm{\delta_{n,n+k}} =ζn+kζn,\displaystyle=\norm{\zeta_{n+k}-\zeta_{n}}, (297)
=i=nkαi(Mi+1+ξi+1),\displaystyle=\norm{\sum_{i=n}^{k}\alpha_{i}{\left(M_{i+1}+\xi_{i+1}\right)}}, (by (288)) (298)
i=nkαiMi+1+i=nkαiξi+1.\displaystyle\leq\norm{\sum_{i=n}^{k}\alpha_{i}M_{i+1}}+\norm{\sum_{i=n}^{k}\alpha_{i}\xi_{i+1}}. (299)

Borkar (2009) proves that (i=0n1αiMi+1,n),n1{\left(\sum_{i=0}^{n-1}\alpha_{i}M_{i+1},\mathcal{F}_{n}\right)},\ n\geq 1 is a zero mean, square-integrable martingale. By D.3, D.4, D.5,

n0𝔼[i=0nαiMi+1i=0n1αiMi+1|n]=n0𝔼[Mn+12|n]<.\displaystyle\sum_{n\geq 0}\mathbb{E}\left[\norm{\sum_{i=0}^{n}\alpha_{i}M_{i+1}-\sum_{i=0}^{n-1}\alpha_{i}M_{i+1}}\,\bigg{|}\,\mathcal{F}_{n}\right]=\sum_{n\geq 0}\mathbb{E}\left[\norm{M_{n+1}}^{2}\,|\,\mathcal{F}_{n}\right]<\infty. (300)

Therefore, the martingale convergence theorem gives the almost sure convergence of (i=nkαiMi+1,n){\left(\sum_{i=n}^{k}\alpha_{i}M_{i+1},\mathcal{F}_{n}\right)} as nn\rightarrow\infty. Combining this with assumption D.1 yields,

limnsup0km(n,T)δn,n+k=0 a.s.\displaystyle\lim_{n\rightarrow\infty}\sup_{0\leq k\leq m(n,T)}\norm{\delta_{n,n+k}}=0\mbox{\quad a.s.\quad} (301)

Using the definition of KT,nCTLk0αn+k2+sup0km(n,T)δn,n+kK_{T,n}\doteq C_{T}L\sum_{k\geq 0}\alpha_{n+k}^{2}+\sup_{0\leq k\leq m(n,T)}\norm{\delta_{n,n+k}} given by (Borkar, 2009), we have proven that our slightly relaxed assumption still yields KT,n0K_{T,n}\rightarrow 0 almost surely as nn\rightarrow\infty. The rest of the argument for the proof of the theorem in Borkar (2009) holds without any additional modification. ∎