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

Some Large Deviations Asymptotics in Small Noise Filtering Problems

Anugu Sumith Reddy Anugu Sumith Reddy, International Centre for Theoretical Sciences-Tata Institute of Fundamental Research, Bangalore, India [email protected] Amarjit Budhiraja Amarjit Budhiraja, Department of Statistics and Operations Research, 304 Hanes Hall, University of North Carolina, Chapel Hill, NC 27599 [email protected]  and  Amit Apte Amit Apte, International Centre for Theoretical Sciences-Tata Institute of Fundamental Research, Bangalore, India [email protected]
Abstract.

We consider nonlinear filters for diffusion processes when the observation and signal noises are small and of the same order. As the noise intensities approach zero, the nonlinear filter can be approximated by a certain variational problem that is closely related to Mortensen’s optimization problem(1968). This approximation result can be made precise through a certain Laplace asymptotic formula. In this work we study probabilities of deviations of true filtering estimates from that obtained by solving the variational problem. Our main result gives a large deviation principle for Laplace functionals whose typical asymptotic behavior is described by Mortensen-type variational problems. Proofs rely on stochastic control representations for positive functionals of Brownian motions and Laplace asymptotics of the Kallianpur-Striebel formula.

Key words and phrases:
Laplace asymptotics, large deviation principle, nonlinear filtering, small observation and signal noise, minimum energy estimate, 44DVAR
2010 Mathematics Subject Classification:
60F10, 60G35, 93E11

1. Introduction

In this work we study certain large deviation asymptotics for nonlinear filtering problems with small signal and observation noise. As the noise in the signal and observation processes vanish, the filtering problem can formally be replaced by a variational problem and one may approximate the filtering estimates (namely suitable conditional probabilities or expectations) by solutions of certain deterministic optimization problems. However due to randomness there will be occasional large deviations of the true nonlinear filter estimates from the variational problem solutions. The main goal of this work is to investigate the probabilities of such deviations by establishing a suitable large deviation principle. Large deviations and related asymptotic problems in the context of small noise nonlinear filtering have been investigated, under different settings, in many works [15, 13, 2, 16, 21, 3, 24, 18, 19, 11, 22, 1]. We summarize the main results of these works and their relation to the asymptotic questions considered in the current work at the end of this section.

In order to describe our results precisely, we begin by introducing the filtering model that we study. We consider a signal process XεX^{\varepsilon} given as the solution of the dd-dimensional stochastic differential equation (SDE)

(1.1) dXε(t)=b(Xε(t))dt+εσ(Xε(t))dW(t),Xε(0)=x0, 0tT,dX^{\varepsilon}(t)=b(X^{\varepsilon}(t))dt+\varepsilon\sigma(X^{\varepsilon}(t))dW(t),\;X^{\varepsilon}(0)=x_{0},\;0\leq t\leq T,

and an mm-dimensional observation process YεY^{\varepsilon} governed by the equation

(1.2) Yε(t)=0th(Xε(s))𝑑s+εB(t), 0tTY^{\varepsilon}(t)=\int_{0}^{t}h(X^{\varepsilon}(s))ds+\varepsilon B(t),\;0\leq t\leq T

on some probability space (Ω¯,¯,¯)(\bar{\mathnormal{\Omega}},\bar{{\mathcal{F}}},\bar{{\mathbb{P}}}). Here ε(0,)\varepsilon\in(0,\infty) is a small parameter, T(0,)T\in(0,\infty) is some given finite time horizon, WW and BB are mutually independent standard Brownian motions in k{\mathbb{R}}^{k} and m{\mathbb{R}}^{m} respectively, x0dx_{0}\in{\mathbb{R}}^{d} is known deterministic initial condition of the signal, and the functions b,σb,\sigma and hh are required to satisfy the following condition.

Assumption 1.

The following hold.

  1. (a)

    The functions b,σ,hb,\sigma,h from dd{\mathbb{R}}^{d}\to{\mathbb{R}}^{d}, dd×k{\mathbb{R}}^{d}\to{\mathbb{R}}^{d\times k}, dm{\mathbb{R}}^{d}\to{\mathbb{R}}^{m} are Lipschitz: For some clip(0,)c_{\mbox{\tiny{lip}}}\in(0,\infty)

    b(x)b(y)+σ(x)σ(y)+h(x)h(y)clipxy for all x,yd.\|b(x)-b(y)\|+\|\sigma(x)-\sigma(y)\|+\|h(x)-h(y)\|\leq c_{\mbox{\tiny{lip}}}\|x-y\|\mbox{ for all }x,y\in{\mathbb{R}}^{d}.
  2. (b)

    The function σ\sigma is bounded: For some cσ(0,)c_{\sigma}\in(0,\infty)

    supxdσ(x)cσ.\sup_{x\in{\mathbb{R}}^{d}}\|\sigma(x)\|\leq c_{\sigma}.
  3. (c)

    The function hh is twice continuously differentiable with bounded first and second derivatives.

Note that under Assumption 1 there is a unique pathwise solution of (1.1) and the solution is a stochastic process with sample paths in 𝒞d{\mathcal{C}}_{d} (the space of continuous functions from [0,T][0,T] to d{\mathbb{R}}^{d} equipped with the uniform metric).

The filtering problem is concerned with the computation of the conditional expectations of the form

(1.3) 𝔼¯[ϕ(Xε)𝒴Tε]\bar{\mathbb{E}}\left[\phi(X^{\varepsilon})\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]

where 𝒴Tεσ{Yε(s):0sT}{\mathcal{Y}}^{\varepsilon}_{T}\doteq\sigma\{Y^{\varepsilon}(s):0\leq s\leq T\} and ϕ:𝒞d\phi:{\mathcal{C}}_{d}\to{\mathbb{R}} is a suitable map. The stochastic process with values in the space of probability measures on 𝒞d{\mathcal{C}}_{d}, given by

¯[Xε𝒴Tε]\bar{\mathbb{P}}\left[X^{\varepsilon}\in\cdot\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]

is usually referred to as the nonlinear filter.

In this work we are interested in the study of the asymptotic behavior of the nonlinear filter as ε0\varepsilon\to 0. Denote by ξ𝒞d\xi^{*}\in{\mathcal{C}}_{d} the unique solution of

(1.4) dξ(t)=b(ξ(t))dt,ξ(0)=x0.d\xi^{*}(t)=b(\xi^{*}(t))dt,\;\;\xi^{*}(0)=x_{0}.

It can be shown that, under additional conditions (see discussion in Section 2), that, as ε0\varepsilon\to 0,

(1.5) ¯[Xε𝒴Tε]δξ, in probability, under ¯,\bar{\mathbb{P}}\left[X^{\varepsilon}\in\cdot\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]\to\delta_{\xi^{*}},\mbox{ in probability, under }\bar{\mathbb{P}},

weakly. In particular for Borel subsets AA of 𝒞d{\mathcal{C}}_{d} whose closure does not contain ξ\xi^{*} one will have ¯[XεA𝒴Tε]0\bar{\mathbb{P}}\left[X^{\varepsilon}\in A\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]\to 0 in probability as ε0\varepsilon\to 0. It is of interest to study the rate of decay of conditional probabilities of such non-typical state trajectories. As a special case of the results of the current paper (see Corollary 4.2) it will follow that for every real continuous and bounded function ϕ\phi on 𝒞d{\mathcal{C}}_{d}, denoting

(1.6) Uε[ϕ]𝔼¯[e1ε2ϕ(Xε)𝒴Tε],U^{\varepsilon}[\phi]\doteq\bar{\mathbb{E}}\left[e^{-\frac{1}{\varepsilon^{2}}\phi(X^{\varepsilon})}\mid{\mathcal{Y}}^{\varepsilon}_{T}\right],
(1.7) ε2logUε[ϕ]¯infη𝒞d[ϕ(η)+120Th(η(s))h(ξ(s))2+J(η)],-\varepsilon^{2}\log U^{\varepsilon}[\phi]\stackrel{{\scriptstyle\bar{\mathbb{P}}}}{{\longrightarrow}}\inf_{\eta\in{\mathcal{C}}_{d}}\left[\phi(\eta)+\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}+J(\eta)\right],

where ¯\stackrel{{\scriptstyle\bar{\mathbb{P}}}}{{\longrightarrow}} denotes convergence in probability under ¯\bar{\mathbb{P}}, and JJ is the rate function on 𝒞d{\mathcal{C}}_{d} associated with the large deviation principle for {Xε}ε>0\{X^{\varepsilon}\}_{\varepsilon>0} (see Section 2). From this convergence it follows using standard arguments (see e.g. [6, Theorem 1.8]), that, for all Borel subsets AA of 𝒞d{\mathcal{C}}_{d})

(1.8) lim¯ε0¯ε2log¯[XεA𝒴Tε]\displaystyle\underline{\lim}^{\bar{\mathbb{P}}}_{\varepsilon\to 0}\varepsilon^{2}\log\bar{\mathbb{P}}\left[X^{\varepsilon}\in A\mid{\mathcal{Y}}^{\varepsilon}_{T}\right] infηAo[120Th(η(s))h(ξ(s))2+J(η)]\displaystyle\geq-\inf_{\eta\in A^{o}}\left[\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}+J(\eta)\right]
lim¯ε0¯ε2log¯[XεA𝒴Tε]\displaystyle\overline{\lim}^{\bar{\mathbb{P}}}_{\varepsilon\to 0}\varepsilon^{2}\log\bar{\mathbb{P}}\left[X^{\varepsilon}\in A\mid{\mathcal{Y}}^{\varepsilon}_{T}\right] infηA¯[120Th(η(s))h(ξ(s))2+J(η)],\displaystyle\leq-\inf_{\eta\in\bar{A}}\left[\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}+J(\eta)\right],

where for real random variables ZεZ^{\varepsilon} and a constant α\alpha\in{\mathbb{R}} we say lim¯ε0¯Zεα\overline{\lim}^{\bar{\mathbb{P}}}_{\varepsilon\to 0}Z^{\varepsilon}\leq\alpha [ resp. lim¯ε0¯Zεα\underline{\lim}^{\bar{\mathbb{P}}}_{\varepsilon\to 0}Z^{\varepsilon}\geq\alpha] if (Zεα)+(Z^{\varepsilon}-\alpha)^{+} [resp. (αZε)+(\alpha-Z^{\varepsilon})^{+}] converges to 0 in ¯\bar{\mathbb{P}}-probability, and for a set AA, AoA^{o} and A¯\bar{A} denote its interior and closure respectively..

Thus the convergence in (1.7) gives information on rate of decays of conditional probabilities of non-typical state trajectories. Formally, denoting the infimum in the above display as S(ξ,A)S(\xi^{*},A), we can write approximations for conditional probabilities:

(1.9) ¯[XεA𝒴Tε]exp{1ε2S(ξ,A)}.\bar{\mathbb{P}}\left[X^{\varepsilon}\in A\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]\approx\exp\left\{-\frac{1}{\varepsilon^{2}}S(\xi^{*},A)\right\}.

However, due to stochastic fluctuations, one may find that for some ‘rogue’ observation trajectories the conditional probabilities on the left side of (1.9) are quite different from the deterministic approximation on the right side of (1.9). In order to quantify the probabilities of observing such rogue observation trajectories that cause deviations from the bounds in (1.8), a natural approach is to study a large deviation principle for {\mathbb{R}} valued random variables {ε2logUε[ϕ]}\{-\varepsilon^{2}\log U^{\varepsilon}[\phi]\} whose typical (law of large numbers) behavior is described by the right side of (1.7). Establishing such a large deviation principle is the goal of this work. Such a result gives information on decay rates of probabilities of the form

¯{|ε2log¯[XεA𝒴Tε]\displaystyle\bar{\mathbb{P}}\Big{\{}\Big{|}\varepsilon^{2}\log\bar{\mathbb{P}}\left[X^{\varepsilon}\in A\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]
+infηA[120Th(η(s))h(ξ(s))2+J(η)]|>δ}\displaystyle\quad\quad\quad+\inf_{\eta\in A}\left[\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}+J(\eta)\right]\Big{|}>\delta\Big{\}}

for suitable sets A(𝒞d)A\in{\mathcal{B}}({\mathcal{C}}_{d}) and δ>0\delta>0. Our main result is Theorem 2.1 which gives a large deviation principle for {ε2logUε[ϕ]}\{-\varepsilon^{2}\log U^{\varepsilon}[\phi]\}, for every continuous and bounded function ϕ\phi on 𝒞d{\mathcal{C}}_{d} with a rate funcion defined by the variational formula in (2.16)-(2.17).

Notation. The following notation and definitions will be used. For pp\in{\mathbb{N}} the Euclidean norm in p{\mathbb{R}}^{p} will be denoted as .\|.\| and the corresponding inner product will be written as ,\langle\cdot,\cdot\rangle. The space of finite positive measures (resp. probability measures) on a Polish space SS will be denoted by (S){\mathcal{M}}(S) (resp. 𝒫(S){\mathcal{P}}(S)). The space of bounded measurable (resp. continuous and bounded) functions from SS\to{\mathbb{R}} will be denoted by BM(S)\mbox{BM}(S) and Cb(S)C_{b}(S) respectively. For ϕBM(S)\phi\in\mbox{BM}(S), ϕsupxS|ϕ(x)|\|\phi\|_{\infty}\doteq\sup_{x\in S}|\phi(x)|. For ϕBM(S)\phi\in\mbox{BM}(S) and μ(S)\mu\in{\mathcal{M}}(S), μ[ϕ]ϕ𝑑μ\mu[\phi]\doteq\int\phi\,d\mu. Borel σ\sigma-field on a Polish space SS will be denoted as (S){\mathcal{B}}(S). For pp\in{\mathbb{N}} and T(0,)T\in(0,\infty), 𝒞p,T{\mathcal{C}}_{p,T} will denote the space of continuous functions from [0,T][0,T] to p{\mathbb{R}}^{p} which is equipped with the supremum norm, defined as f,Tsup0tTf(t)\|f\|_{*,T}\doteq\sup_{0\leq t\leq T}\|f(t)\|, f𝒞p,Tf\in{\mathcal{C}}_{p,T}. Since T(0,)T\in(0,\infty) will be fixed in most of this work, frequently the subscript TT in 𝒞p,T{\mathcal{C}}_{p,T} and f,T\|f\|_{*,T} will be dropped. We denote by p2L2([0,T]:p){\mathcal{L}}^{2}_{p}\doteq L^{2}([0,T]:{\mathbb{R}}^{p}) the Hilbert space of square-integrable functions from [0,T][0,T] to p{\mathbb{R}}^{p}. By convention, the infimum over an empty set will be taken to be \infty. For random variables XnX_{n}, XX with values in some Polish space SS, convergence in distribution of XnX_{n} to XX will be denoted as XnXX_{n}\Rightarrow X. A function II from a Polish space SS to [0,][0,\infty] is called a rate function if it has compact sub-level sets, namely the set {xS:I(x)m}\{x\in S:I(x)\leq m\} is a compact set of SS for every m(0,)m\in(0,\infty). Given a function a:(0,)(0,)a:(0,\infty)\to(0,\infty) such that a(ε)a(\varepsilon)\to\infty as ε0\varepsilon\to 0, and a rate function II on a Polish space SS, a collection {Uε}ε>0\{U^{\varepsilon}\}_{\varepsilon>0} of SS valued random variables is said to satisfy a large deviation principle (LDP) with rate function II and speed a(ε)a(\varepsilon) if for every ϕCb(S)\phi\in C_{b}(S)

limε0a(ε)1log𝔼ea(ε)ϕ(Uε)=infxS[I(x)+ϕ(x)].\lim_{\varepsilon\to 0}-a(\varepsilon)^{-1}\log{\mathbb{E}}e^{-a(\varepsilon)\phi(U^{\varepsilon})}=\inf_{x\in S}[I(x)+\phi(x)].

Relation with existing body of work. Denote by 𝒞m1{\mathcal{C}}^{1}_{m} the collection of absolutely continuous functions y𝒞my\in{\mathcal{C}}_{m} that satisfy 0Ty˙(s)2𝑑s<\int_{0}^{T}\|\dot{y}(s)\|^{2}ds<\infty. For y𝒞m1y\in{\mathcal{C}}^{1}_{m} define Iy:𝒞d[0,]I_{y}:{\mathcal{C}}_{d}\to[0,\infty] as

(1.10) Iy(η)=120Th(η(s))y˙(s)2𝑑s+J(η)\displaystyle I_{y}(\eta)=\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-\dot{y}(s)\|^{2}ds+J(\eta)

where JJ is the rate function of {Xε}\{X^{\varepsilon}\} defined in (2.9). The functional IyI_{y} was introduced in Mortensen[20] as the objective function in an optimization problem whose minima describes the most probable trajectory given the data in a nonlinear filtering problem in an appropriate asymptotic sense. This functional is also used in implementing the popular 44DVAR data assimilation algorithm (cf. [7, Section 3.2], [12, Chapter 16]). Connection of the optimization problem associated with the objective function in (1.10) with the asymptotics of classical small noise filtering problem has been studied by several authors [15, 14, 16]. We now describe this connection.

In Section 2 we will introduce a continuous map Λ^Tε:𝒞m𝒫(𝒞d)\hat{\Lambda}^{\varepsilon}_{T}:{\mathcal{C}}_{m}\to{\mathcal{P}}({\mathcal{C}}_{d}) such that Λ^Tε(Yε)=¯(Xε𝒴Tε)\hat{\Lambda}^{\varepsilon}_{T}(Y^{\varepsilon})=\bar{\mathbb{P}}(X^{\varepsilon}\in\cdot\mid{\mathcal{Y}}^{\varepsilon}_{T}) a.s. In [15], Hijab established, under conditions (that include boundedness and smoothness of various coefficients functions), a large deviation principle for the collection of probability measures (on 𝒞d{\mathcal{C}}_{d}) {Λ^Tε(y)}ε>0\{\hat{\Lambda}^{\varepsilon}_{T}(y)\}_{\varepsilon>0} (with speed ε2\varepsilon^{-2}), for a fixed yy in 𝒞m1{\mathcal{C}}^{1}_{m}, with rate function I^y:𝒞d[0,]\hat{I}_{y}:{\mathcal{C}}_{d}\to[0,\infty] given by

(1.11) I^y(η)=Iy(η)infη^𝒞d{Iy(η^)}.\hat{I}_{y}(\eta)=I_{y}(\eta)-\inf_{\hat{\eta}\in{\mathcal{C}}_{d}}\{I_{y}(\hat{\eta})\}.

In a related direction, Hijab’s Ph.D. dissertation [14], studied asymptotics of the unnormalized conditional density and established, under conditions, an asymptotic formula of the form

qε(x,t)=exp{1ε2(W(x,t)+o(1))},q^{\varepsilon}(x,t)=\exp\left\{-\frac{1}{\varepsilon^{2}}(W(x,t)+o(1))\right\},

where qε(x,t)q^{\varepsilon}(x,t) denotes the solution of the Zakai equation associated with the nonlinear filter (cf. [17]). The deterministic function W(x,t)W(x,t) coincides with Mortensen’s (deterministic) minimum energy estimate [20] which is given as solution of a certain minimization problem related to the objective function Iy(η)I_{y}(\eta). Results of Hijab were extended to random initial conditions in [16], once again assuming boundedness and smoothness of coefficients. In related work, the problem of constructing observers for dynamical systems as limits of stochastic nonlinear filters is studied in [2]. Heunis[13] studies a somewhat different asymptotic problem for small noise nonlinear filters. Specifically, it is shown in [13], that for every ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}), w𝒞mw\in{\mathcal{C}}_{m}, and for any η𝒞d\eta\in{\mathcal{C}}_{d} for which the map defined in (2.13) has a unique minimizer (at say η\eta^{*}),

Λ^Tε(0h(η(s))𝑑s+εw)[ϕ]ϕ(η), as ε0.\hat{\Lambda}^{\varepsilon}_{T}\left(\int_{0}^{\cdot}h(\eta(s))ds+\varepsilon w\right)[\phi]\to\phi(\eta^{*}),\;\mbox{ as }\varepsilon\to 0.

This result and its connection to our work are further discussed in Section 2. In particular the statement in (1.5) follows readily on using similar ideas as in [13]. The work of Pardoux and Zeitouni[21] considers a one dimensional nonlinear filtering problem where the observation noise is small while the signal noise is O(1)O(1) (specifically, the term εσ(Xε(t))\varepsilon\sigma(X^{\varepsilon}(t)) in (1.1) is replaced by 11). In this case the conditional distribution of X(T)X(T) given 𝒴Tε{\mathcal{Y}}^{\varepsilon}_{T} converges a.s. to a Dirac measure δX(T)\delta_{X(T)} as ε0\varepsilon\to 0. The paper [21] proves a quenched LDP for this conditional distribution (regarded as a collection of probability measures on 𝒞d{\mathcal{C}}_{d} parametrized by X(T)(ω)X(T)(\omega)) in 𝒞d{\mathcal{C}}_{d}. In a somewhat different direction, in a sequence of papers [24, 19, 18], the authors have studied asymptotics of the filtering problem under a small signal to noise ratio limit, under various types of model settings. In this case the nonlinear filter converges to the unconditional law of the signal and the authors establish large deviation principles characterizing probabilities of deviation of the filter from the above deterministic law. An analogous result in a correlated signal-observation noise case was studied in [3]. Finally, yet another type of large deviation problem in the context of nonlinear filtering (with correlated signal-observation noise) when the observation noise is O(1)O(1) and the signal noise and drift are suitably small has been considered in a series of papers [11, 22, 1].

The closest connections of the current work are with [15] and [13]. Specifically, the asymptotic statements in (1.7) and (1.8) which follow as a special case of our results (see Corollary 4.2) is analogous to results in [15], except that instead of a fixed observation path we consider the actual observation process YεY^{\varepsilon} (also we make substantially weaker assumption on coefficients than [15]). However our main interest is in a large deviation principle for the convergence to the deterministic limit in (1.7) , thus roughly speaking we are interested in quantifying the probability of deviations from the convergence statement in [15] (when a fixed observation path is replaced with the observation process YεY^{\varepsilon}). This large deviation result, given in Theorem 2.1, is the main contribution of our work.

Proof idea. The proof of Theorem 2.1 is based on a variational representation for functionals of Brownian motions obtained in [4] (see also [5]) using which the proof of the large deviation principle reduces to proving a key weak convergence result given in Lemma 4.1. Proof of Lemma 4.1 is the technical heart of this work. Important use is made of some key estimates obtained in [13] (see in particular Proposition 5.3). One of the key steps is to argue that terms of order ε1\varepsilon^{-1} can be ignored in the exponent when studying Laplace asymptotics for the quantity on the left side of (3.6). This relies on several careful large deviation exponential estimates which are developed in Section 5. Once Lemma 4.1 is available the proof of the large deviation principle in Theorem 2.1 follows readily using the now well developed weak convergence approach for the study of large deviation problems (cf. [6]).

Organization. It will be convenient to formulate the filtering problem on canonical path spaces and also to represent the nonlinear filter through a map given on the path space of the observation process. This formulation and our main result (Theorem 2.1) are given in Section 2. The key idea in the proof of the LDP is a variational representation from [4]. A somewhat simplified version of this representation (cf. [6]) that is used in this work is presented in Section 3. Section 4 presents a key lemma (Lemma 4.1) that is needed for implementing the weak convergence method for proving the large deviation result in Theorem 2.1. Section 5 is devoted to the proof of Lemma 4.1. Using this lemma, the proof of Theorem 2.1 is completed in Section 6.

2. Setting and Main Result

Recall that XεX^{\varepsilon} has sample paths in 𝒞d{\mathcal{C}}_{d}. Similarly, the processes Yε,W,BY^{\varepsilon},W,B have sample paths in 𝒞m,𝒞k,𝒞m{\mathcal{C}}_{m},{\mathcal{C}}_{k},{\mathcal{C}}_{m} respectively. It will be convenient to formulate the filtering problem on suitable path spaces. Denote, for pp\in{\mathbb{N}}, the standard Wiener measure on (𝒞p,(𝒞p))({\mathcal{C}}_{p},{\mathcal{B}}({\mathcal{C}}_{p})) as 𝒲p{\mathcal{W}}_{p} and the Wiener measure with variance parameter ε2\varepsilon^{2} as 𝒲pε{\mathcal{W}}_{p}^{\varepsilon}. Denote the canonical coordinate process on (𝒞k,(𝒞k))({\mathcal{C}}_{k},{\mathcal{B}}({\mathcal{C}}_{k})) as {γ(t):0tT}\{\gamma(t):0\leq t\leq T\} and consider the SDE on the probability space (𝒞k,(𝒞k),𝒲k)({\mathcal{C}}_{k},{\mathcal{B}}({\mathcal{C}}_{k}),{\mathcal{W}}_{k}),

dxε(t)=b(xε(t))dt+εσ(xε(t))dγ(t),xε(0)=x0, 0tT.dx^{\varepsilon}(t)=b(x^{\varepsilon}(t))dt+\varepsilon\sigma(x^{\varepsilon}(t))d\gamma(t),\;x^{\varepsilon}(0)=x_{0},\;0\leq t\leq T.

From Assumption 1, the above SDE has a unique strong solution with sample paths in 𝒞d{\mathcal{C}}_{d}.

Consider the map 𝒞kΩx𝒞d×𝒞k{\mathcal{C}}_{k}\to\mathnormal{\Omega}_{x}\doteq{\mathcal{C}}_{d}\times{\mathcal{C}}_{k} defined as ω(xε(ω),γ(ω))\omega\mapsto(x^{\varepsilon}(\omega),\gamma(\omega)) and let

με𝒲k(xε,γ)1.\mu^{\varepsilon}\doteq{\mathcal{W}}_{k}\circ(x^{\varepsilon},\gamma)^{-1}.

Next, let Ωy𝒞m\mathnormal{\Omega}_{y}\doteq{\mathcal{C}}_{m} and consider the probability space

(Ω,,ε)(Ωx×Ωy,(Ωx)(Ωy),με𝒲mε).(\mathnormal{\Omega},{\mathcal{F}},{\mathbb{Q}}^{\varepsilon})\doteq(\mathnormal{\Omega}_{x}\times\mathnormal{\Omega}_{y},{\mathcal{B}}(\mathnormal{\Omega}_{x})\otimes{\mathcal{B}}(\mathnormal{\Omega}_{y}),\mu^{\varepsilon}\otimes{\mathcal{W}}^{\varepsilon}_{m}).

Abusing notation, denote the coordinate maps on the above probability space as ξ,γ,ζ\xi,\gamma,\zeta, namely

ξ(ω)=ω1,γ(ω)=ω2,ζ(ω)=ω3 for ω=((ω1,ω2),ω3)Ωx×Ωy.\xi(\omega)=\omega_{1},\;\;\gamma(\omega)=\omega_{2},\;\;\zeta(\omega)=\omega_{3}\mbox{ for }\omega=((\omega_{1},\omega_{2}),\omega_{3})\in\mathnormal{\Omega}_{x}\times\mathnormal{\Omega}_{y}.

We will frequently write ξ(ω)(s)\xi(\omega)(s) as ξ(s)\xi(s) for (ω,s)Ω×[0,T](\omega,s)\in\mathnormal{\Omega}\times[0,T]. Similar notational shorthand will be followed for other coordinate maps.

Note that, under ε{\mathbb{Q}}^{\varepsilon}, ξ(0)=x0\xi(0)=x_{0}, γ\gamma and ε1ζ\varepsilon^{-1}\zeta are independent standard Brownian motions in k{\mathbb{R}}^{k} and m{\mathbb{R}}^{m} respectively and

(2.1) ξ(t)=x0+0tb(ξ(s))𝑑s+ε0tσ(ξ(s))𝑑γ(s), 0tT.\xi(t)=x_{0}+\int_{0}^{t}b(\xi(s))ds+\varepsilon\int_{0}^{t}\sigma(\xi(s))d\gamma(s),\;0\leq t\leq T.

Define, for ε{\mathbb{Q}}^{\varepsilon} a.e. ω=((ω1,ω2),ω3)\omega=((\omega_{1},\omega_{2}),\omega_{3}), for t[0,T]t\in[0,T],

Ltε(ω)exp{1ε20th(ξ(s)),dζ(s)12ε20th(ξ(s))2𝑑s}.L^{\varepsilon}_{t}(\omega)\doteq\exp\left\{\frac{1}{\varepsilon^{2}}\int_{0}^{t}\langle h(\xi(s)),d\zeta(s)\rangle-\frac{1}{2\varepsilon^{2}}\int_{0}^{t}\|h(\xi(s))\|^{2}ds\right\}.

Note that, since under ε{\mathbb{Q}}^{\varepsilon}, ε1ζ\varepsilon^{-1}\zeta is a standard Brownian martingale with respect to the filtration t0σ{γ(s),ξ(s),ζ(s):0st}{\mathcal{F}}_{t}^{0}\doteq\sigma\{\gamma(s),\xi(s),\zeta(s):0\leq s\leq t\}, the first integral in the exponent is well-defined as an Itô integral. From the independence of ξ\xi and ζ\zeta under ε{\mathbb{Q}}^{\varepsilon} and Assumption 1 it follows that LtεL^{\varepsilon}_{t} is a {t0}\{{\mathcal{F}}^{0}_{t}\}-martingale under ε{\mathbb{Q}}^{\varepsilon}. Define a probability measure ε{\mathbb{P}}^{\varepsilon} on (Ω,)(\mathnormal{\Omega},{\mathcal{F}}) as

dεdε(ω)LTε(ω),ε a.e. ω.\frac{d{\mathbb{P}}^{\varepsilon}}{d{\mathbb{Q}}^{\varepsilon}}(\omega)\doteq L^{\varepsilon}_{T}(\omega),\;{\mathbb{Q}}^{\varepsilon}\mbox{ a.e. }\omega.

Note that, by Girsanov’s theorem, under ε{\mathbb{P}}^{\varepsilon}

(2.2) β(t)1εζ(t)1ε0th(ξ(s))𝑑s, 0tT\beta(t)\doteq\frac{1}{\varepsilon}\zeta(t)-\frac{1}{\varepsilon}\int_{0}^{t}h(\xi(s))ds,\;0\leq t\leq T

is a standard mm-dimensional Brownian motion which is independent of (ξ,γ)(\xi,\gamma). Rewriting the above equation as

ζ(t)=0th(ξ(s))𝑑s+εβ(t), 0tT,\zeta(t)=\int_{0}^{t}h(\xi(s))ds+\varepsilon\beta(t),\;0\leq t\leq T,

we see that

(2.3) ¯(Xε,Yε)1=ε(ξ,ζ)1.\bar{\mathbb{P}}\circ(X^{\varepsilon},Y^{\varepsilon})^{-1}={\mathbb{P}}^{\varepsilon}\circ(\xi,\zeta)^{-1}.

Next, for ε>0\varepsilon>0, define ΓTε:𝒞m(𝒞d)\Gamma^{\varepsilon}_{T}:{\mathcal{C}}_{m}\to{\mathcal{M}}({\mathcal{C}}_{d}) as

(2.4) ΓTε(ω3)[A]Ωx1A(ω1)LTε((ω1,ω2),ω3)𝑑με(ω1,ω2),ω3𝒞m,A(𝒞d).\Gamma^{\varepsilon}_{T}(\omega_{3})[A]\doteq\int_{\mathnormal{\Omega}_{x}}1_{A}(\omega_{1})L^{\varepsilon}_{T}((\omega_{1},\omega_{2}),\omega_{3})d\mu^{\varepsilon}(\omega_{1},\omega_{2}),\;\omega_{3}\in{\mathcal{C}}_{m},\;A\in{\mathcal{B}}({\mathcal{C}}_{d}).

The maps are well defined ε{\mathbb{P}}^{\varepsilon}-a.s. and using results of [8, 9, 10], one can obtain versions of these maps (denoted as Γ^Tε\hat{\Gamma}^{\varepsilon}_{T}) which are continuous on 𝒞m{\mathcal{C}}_{m}. Also, define ΛTε:𝒞m𝒫(𝒞d)\Lambda^{\varepsilon}_{T}:{\mathcal{C}}_{m}\to{\mathcal{P}}({\mathcal{C}}_{d}) as

(2.5) ΛTε(ω)[A]ΓTε(ω)[A]ΓTε(ω)[𝒞d],ε-a.e. ω𝒞m,A(𝒞d).\Lambda^{\varepsilon}_{T}(\omega)[A]\doteq\frac{\Gamma^{\varepsilon}_{T}(\omega)[A]}{\Gamma^{\varepsilon}_{T}(\omega)[{\mathcal{C}}_{d}]},\;{\mathbb{P}}^{\varepsilon}\mbox{-a.e. }\omega\in{\mathcal{C}}_{m},\;A\in{\mathcal{B}}({\mathcal{C}}_{d}).

Once again, for each ε>0\varepsilon>0, this map is well defined ε{\mathbb{P}}^{\varepsilon}-a.s. and a continuous version of the map exists (which we denote as Λ^Tε\hat{\Lambda}^{\varepsilon}_{T}) from [8, 9, 10] . Write, for fBM(𝒞d)f\in\mbox{BM}({\mathcal{C}}_{d})

ΓTε(f,ω)𝒞df(ω~)ΓTε(ω)[dω~],ΛTε(f,ω)𝒞df(ω~)ΛTε(ω)[dω~],ε-a.e. ω𝒞m.\Gamma^{\varepsilon}_{T}(f,\omega)\doteq\int_{{\mathcal{C}}_{d}}f(\tilde{\omega})\Gamma^{\varepsilon}_{T}(\omega)[d\tilde{\omega}],\;\Lambda^{\varepsilon}_{T}(f,\omega)\doteq\int_{{\mathcal{C}}_{d}}f(\tilde{\omega})\Lambda^{\varepsilon}_{T}(\omega)[d\tilde{\omega}],\;{\mathbb{P}}^{\varepsilon}\mbox{-a.e. }\omega\in{\mathcal{C}}_{m}.

Then with (Xε,Yε)(X^{\varepsilon},Y^{\varepsilon}) as in (1.1)-(1.2), for ϕBM(𝒞d)\phi\in\mbox{BM}({\mathcal{C}}_{d})

(2.6) 𝔼¯[ϕ(Xε)𝒴Tε]=ΛTε(ϕ,Yε) a.s. ¯.\bar{\mathbb{E}}\left[\phi(X^{\varepsilon})\mid{\mathcal{Y}}^{\varepsilon}_{T}\right]=\Lambda^{\varepsilon}_{T}(\phi,Y^{\varepsilon})\mbox{ a.s. }\bar{\mathbb{P}}.

Also,

(2.7) 𝔼ε[ϕ(ξ)σ{ζ(s):0sT}]=ΛTε(ϕ,ζ) a.s. ε,{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon}}\left[\phi(\xi)\mid\sigma\{\zeta(s):0\leq s\leq T\}\right]=\Lambda^{\varepsilon}_{T}(\phi,\zeta)\mbox{ a.s. }{\mathbb{P}}^{\varepsilon},

where 𝔼ε{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon}} denotes the expectation under the probability measure ε{\mathbb{P}}^{\varepsilon}, and

(2.8) ¯(Xε,Yε,W,B,ΛTε(ϕ,Yε))1=ε(ξ,ζ,γ,β,ΛTε(ϕ,ζ))1.\bar{\mathbb{P}}\circ(X^{\varepsilon},Y^{\varepsilon},W,B,\Lambda^{\varepsilon}_{T}(\phi,Y^{\varepsilon}))^{-1}={\mathbb{P}}^{\varepsilon}\circ(\xi,\zeta,\gamma,\beta,\Lambda^{\varepsilon}_{T}(\phi,\zeta))^{-1}.

Let for ξ0𝒞d\xi_{0}\in{\mathcal{C}}_{d},

(2.9) J(ξ0)infφ𝒰(ξ0)[120Tφ(t)2𝑑t]J(\xi_{0})\doteq\inf_{\varphi\in{\mathcal{U}}(\xi_{0})}\left[\frac{1}{2}\int_{0}^{T}\|\varphi(t)\|^{2}dt\right]

where 𝒰(ξ0){\mathcal{U}}(\xi_{0}) is the collection of all φ\varphi in k2{\mathcal{L}}^{2}_{k} such that

(2.10) ξ0(t)=x0+0tb(ξ0(s))𝑑s+0tσ(ξ0(s))φ(s)𝑑s,t[0,T].\xi_{0}(t)=x_{0}+\int_{0}^{t}b(\xi_{0}(s))ds+\int_{0}^{t}\sigma(\xi_{0}(s))\varphi(s)ds,\;t\in[0,T].

Note that, by Assumption 1, for every φk2\varphi\in{\mathcal{L}}^{2}_{k} there is a unique solution of (2.10). By classical results of Freidlin and Wentzell (see e.g. [6, Theorem 10.6]) the collection {Xε}\{X^{\varepsilon}\} of 𝒞d{\mathcal{C}}_{d} valued random variables satisfies a LDP with rate function JJ and speed ε2\varepsilon^{-2}, namely, for all FCb(𝒞d)F\in C_{b}({\mathcal{C}}_{d})

(2.11) limε0ε2logΩxexp{1ε2F(ξ^)}𝑑με=infξ0𝒞d[F(ξ0)+J(ξ0)],\lim_{\varepsilon\to 0}-\varepsilon^{2}\log\int_{\mathnormal{\Omega}_{x}}\exp\left\{-\frac{1}{\varepsilon^{2}}F(\hat{\xi})\right\}d\mu^{\varepsilon}=\inf_{\xi_{0}\in{\mathcal{C}}_{d}}\left[F(\xi_{0})+J(\xi_{0})\right],

where we denote the first coordinate process on Ωx\mathnormal{\Omega}_{x} by ξ^\hat{\xi}, i.e. ξ^(ω)=ω1\hat{\xi}(\omega)=\omega_{1} for ω=(ω1,ω2)Ωx=𝒞d×𝒞k\omega=(\omega_{1},\omega_{2})\in\mathnormal{\Omega}_{x}={\mathcal{C}}_{d}\times{\mathcal{C}}_{k}. In [13] it is shown that for every w𝒞mw\in{\mathcal{C}}_{m}, and a given η𝒞d\eta\in{\mathcal{C}}_{d}, the probability measure

(2.12) Λ^Tε(0h(η(s))𝑑s+εw())δη\hat{\Lambda}_{T}^{\varepsilon}\left(\int_{0}^{\cdot}h(\eta(s))ds+\varepsilon w(\cdot)\right)\to\delta_{\eta^{*}}

weakly, if the map

(2.13) η~J(η~)+120Th(η(s))h(η~(s))2𝑑s\tilde{\eta}\mapsto J(\tilde{\eta})+\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|^{2}ds

attains its infimum over 𝒞d{\mathcal{C}}_{d} uniquely at η\eta^{*}, where recall that Λ^Tε\hat{\Lambda}^{\varepsilon}_{T} is the continuous version of ΛTε\Lambda^{\varepsilon}_{T}. We remark that [13] assumes in addition to (1) that hh and bb are bounded, but an examination of the proof shows (see calculations in Section 5) that these conditons can be replaced by linear growth conditions that are implied by Assumption 1 .

Recall the function ξ𝒞d\xi^{*}\in{\mathcal{C}}_{d} from (1.4). Then using similar ideas as in [13], under Assumption 1, and assuming in addition that either σσ\sigma\sigma^{\dagger} is positive definite or hh is a one-to-one function, it follows that

(2.14) ΛTεδξ, in probability, under ε,\Lambda_{T}^{\varepsilon}\to\delta_{\xi^{*}},\mbox{ in probability, under }{\mathbb{P}}^{\varepsilon},

weakly, as ε0\varepsilon\to 0. This is a consequence of the fact that when η=ξ\eta=\xi^{*} the map in (2.13) achieves its minimum (which is 0) uniquely at ξ\xi^{*}.

As a consequence of the results of the current paper (see Corollary 4.2) one can show the Laplace asymptotic formula in (1.7). Recall from the discussion in the Introduction that the convergence in (1.7) gives information on asymptotics of conditional probabilities of non-typical state trajectories. In order to quantify the decay rate of probabilities of observing rare observation trajectories that cause deviations from the deterministic variational quantity in (1.7), we will establish a large deviation principle for {ε2logUε[ϕ]}\{-\varepsilon^{2}\log U^{\varepsilon}[\phi]\} defined in (1.6).

We now present the rate function associated with this LDP.

Define the map H:𝒞d×𝒞d×m2+H:{\mathcal{C}}_{d}\times{\mathcal{C}}_{d}\times{\mathcal{L}}^{2}_{m}\to{\mathbb{R}}_{+} as

(2.15) H(η,η~,ψ)120Th(η(s))h(η~(s))ψ(s)2𝑑s.H(\eta,\tilde{\eta},\psi)\doteq\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))-\psi(s)\|^{2}ds.

Also, for φk2\varphi\in{\mathcal{L}}^{2}_{k}, let ξ0φ\xi_{0}^{\varphi} be given as the unique solution of (2.10).

We now introduce the rate function that will govern the large deviation asymptotics of ε2logUε[ϕ]-\varepsilon^{2}\log U^{\varepsilon}[\phi].

Fix ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}) and define Iϕ:[0,]I^{\phi}:{\mathbb{R}}\to[0,\infty] as

(2.16) Iϕ(z)=inf(φ,ψ)𝒮(z)[120Tφ(t)2𝑑t+120Tψ(t)2𝑑t]I^{\phi}(z)=\inf_{(\varphi,\psi)\in{\mathcal{S}}(z)}\left[\frac{1}{2}\int_{0}^{T}\|\varphi(t)\|^{2}dt+\frac{1}{2}\int_{0}^{T}\|\psi(t)\|^{2}dt\right]

where 𝒮(z){\mathcal{S}}(z) is the collection of all (φ,ψ)(\varphi,\psi) in k2×m2{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m} such that

(2.17) infη𝒞d[H(η,ξ0φ,ψ)+ϕ(η)+J(η)]infη𝒞d[H(η,ξ0φ,ψ)+J(η)]=z.\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+J(\eta)\right]=z.

The following is the main result of the work.

Theorem 2.1.

Suppose that Assumption 1 is satisfied. Then for every ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}), the collection {ε2logUε[ϕ]}\{-\varepsilon^{2}\log U^{\varepsilon}[\phi]\} satisfies a large deviation principle on {\mathbb{R}} with rate function IϕI^{\phi} and speed ε2\varepsilon^{-2}.

3. A Variational Representation

Fix ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}). Recall the functional Uε[ϕ]U^{\varepsilon}[\phi] from (1.6). From (2.6), note that one can write Uε[ϕ]U^{\varepsilon}[\phi] as

Uε[ϕ]=ΛTε(exp{ε2ϕ()},Yε)U^{\varepsilon}[\phi]=\Lambda^{\varepsilon}_{T}\left(\exp\{-\varepsilon^{-2}\phi(\cdot)\},Y^{\varepsilon}\right)

whose distribution under ¯\bar{\mathbb{P}} is same as the distribution of ΛTε(exp{ε2ϕ()},ζ)\Lambda^{\varepsilon}_{T}\left(\exp\{-\varepsilon^{-2}\phi(\cdot)\},\zeta\right) under ε{\mathbb{P}}^{\varepsilon}. Let

Vε[ϕ]=ε2logΛTε(exp{ε2ϕ()},ζ).V^{\varepsilon}[\phi]=-\varepsilon^{2}\log\Lambda^{\varepsilon}_{T}\left(\exp\{-\varepsilon^{-2}\phi(\cdot)\},\zeta\right).

Using this equality of laws and the equivalence between Large deviation principles and Laplace principles (see e.g. [6, Theorems 1.5 and 1.8]), in order to prove Theorem 2.1 it suffices to show that, IϕI^{\phi} has compact sub-level sets, i.e.,

(3.1)  for every m+,{z:Iϕ(z)m} is compact,\mbox{ for every }m\in{\mathbb{R}}_{+},\{z\in{\mathbb{R}}:I^{\phi}(z)\leq m\}\mbox{ is compact,}

and for every GCb()G\in C_{b}({\mathbb{R}})

(3.2) limε0ε2log𝔼ε[exp{ε2G(Vε[ϕ])}]=infz{G(z)+Iϕ(z)}.\lim_{\varepsilon\to 0}-\varepsilon^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon}}\left[\exp\left\{-\varepsilon^{-2}G(V^{\varepsilon}[\phi])\right\}\right]=\inf_{z\in{\mathbb{R}}}\{G(z)+I^{\phi}(z)\}.

The proof of the identity in (3.2) will use a variational representation for nonnegative functionals of Brownian motions given by Boué and Dupuis[4]. We now use this representation to give a variational formula for the left side of the above equation. Let t{\mathcal{F}}_{t} denote the ε{\mathbb{P}}^{\varepsilon}-completion of t0{\mathcal{F}}^{0}_{t} and denote by 𝒜k{\mathcal{A}}^{k} [resp. 𝒜m{\mathcal{A}}^{m}] the collection of all {t}\{{\mathcal{F}}_{t}\}-progressively measurable k{\mathbb{R}}^{k} [resp. m{\mathbb{R}}^{m}] valued processes gg such that for some M=M(g)(0,)M=M(g)\in(0,\infty)

0Tg(s)2𝑑sMa.s.\int_{0}^{T}\|g(s)\|^{2}ds\leq M\;\;\mbox{a.s.}

For (u,v)𝒜k×𝒜m(u,v)\in{\mathcal{A}}^{k}\times{\mathcal{A}}^{m}, let ξu\xi^{u} be given as the unique solution of the SDE on (Ω,,{t},ε)(\mathnormal{\Omega},{\mathcal{F}},\{{\mathcal{F}}_{t}\},{\mathbb{P}}^{\varepsilon}):

(3.3) ξu(t)=x0+0tb(ξu(s))𝑑s+ε0tσ(ξu(s))𝑑γ(s)+0tσ(ξu(s))u(s)𝑑s, 0tT.\xi^{u}(t)=x_{0}+\int_{0}^{t}b(\xi^{u}(s))ds+\varepsilon\int_{0}^{t}\sigma(\xi^{u}(s))d\gamma(s)+\int_{0}^{t}\sigma(\xi^{u}(s))u(s)ds,\;0\leq t\leq T.

Also define

(3.4) ζu,v(t)=0th(ξu(s))𝑑s+εβ(t)+0tv(s)𝑑s, 0tT.\zeta^{u,v}(t)=\int_{0}^{t}h(\xi^{u}(s))ds+\varepsilon\beta(t)+\int_{0}^{t}v(s)ds,\;0\leq t\leq T.

Occasionally, to emphasize the dependence of above processes on ε\varepsilon we will write (ξu,ζu,v)(\xi^{u},\zeta^{u,v}) as (ξε,u,ζε,u,v)(\xi^{\varepsilon,u},\zeta^{\varepsilon,u,v}).

Now let

(3.5) V¯ε,u,v[ϕ]ε2logΛTε(exp{ε2ϕ()},ζε,u,v).\bar{V}^{\varepsilon,u,v}[\phi]\doteq-\varepsilon^{2}\log\Lambda^{\varepsilon}_{T}\left(\exp\{-\varepsilon^{-2}\phi(\cdot)\},\zeta^{\varepsilon,u,v}\right).

When clear from context we will drop (u,v,ϕ)(u,v,\phi) from the notation in V¯ε,u,v[ϕ]\bar{V}^{\varepsilon,u,v}[\phi] and simply write V¯ε\bar{V}^{\varepsilon}. Then it follows from [4] (cf. [6, Theorems 3.17]) that

(3.6) ε2log𝔼ε[exp{ε2G(Vε[ϕ])}]\displaystyle-\varepsilon^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon}}\left[\exp\left\{-\varepsilon^{-2}G(V^{\varepsilon}[\phi])\right\}\right]
=inf(u,v)𝒜k×𝒜m𝔼ε[G(V¯ε,u,v[ϕ])+120T(u(s)2+v(s)2)𝑑s].\displaystyle=\inf_{(u,v)\in{\mathcal{A}}^{k}\times{\mathcal{A}}^{m}}{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon}}\left[G(\bar{V}^{\varepsilon,u,v}[\phi])+\frac{1}{2}\int_{0}^{T}(\|u(s)\|^{2}+\|v(s)\|^{2})ds\right].

4. A Key Lemma

For M(0,)M\in(0,\infty), let

SM{(φ,ψ)k2×m2:0T(φ(s)2+ψ(s)2)𝑑sM}.S_{M}\doteq\{(\varphi,\psi)\in{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m}:\int_{0}^{T}(\|\varphi(s)\|^{2}+\|\psi(s)\|^{2})ds\leq M\}.

We equip, SMS_{M} with the weak topology under which (φn,ψn)(φ,ψ)(\varphi_{n},\psi_{n})\to(\varphi,\psi) as nn\to\infty if and only if for all (f,g)k2×m2(f,g)\in{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m}

0T[φn(s),f(s)+ψn(s),g(s)]𝑑s0T[φ(s),f(s)+ψ(s),g(s)]𝑑s\int_{0}^{T}[\langle\varphi_{n}(s),f(s)\rangle+\langle\psi_{n}(s),g(s)\rangle]ds\to\int_{0}^{T}[\langle\varphi(s),f(s)\rangle+\langle\psi(s),g(s)\rangle]ds

as nn\to\infty. This topology can be metrized so that SMS_{M} is a compact metric space.

Recall ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}) in the statement of Theorem 2.1. For (φ,ψ)k2×m2(\varphi,\psi)\in{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m} define

(4.1) V0φ,ψ[ϕ]infη𝒞d[H(η,ξ0φ,ψ)+ϕ(η)+J(η)]infη𝒞d[H(η,ξ0φ,ψ)+J(η)].V_{0}^{\varphi,\psi}[\phi]\doteq\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+J(\eta)\right].

Note that with this notation 𝒮(z){\mathcal{S}}(z) (introduced below (2.16)) is the collection of all (φ,ψ)(\varphi,\psi) in k2×m2{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m} such that V0φ,ψ[ϕ]=zV_{0}^{\varphi,\psi}[\phi]=z.

The following lemma will be the key to the proof of Theorem 2.1.

Lemma 4.1.

Fix M(0,)M\in(0,\infty). Let {(un,vn)}\{(u_{n},v_{n})\} be a sequence of SMS_{M} valued random variables such that (un,vn)𝒜k×𝒜m(u_{n},v_{n})\in{\mathcal{A}}^{k}\times{\mathcal{A}}^{m} for every nn. Suppose that (un,vn)(u_{n},v_{n}) converges in distribution to (u,v)(u,v). Suppose εn\varepsilon_{n} is a sequence of positive reals converging to 0 as nn\to\infty. Then V¯εn,un,vn[ϕ]V0u,v[ϕ]\bar{V}^{\varepsilon_{n},u_{n},v_{n}}[\phi]\to V_{0}^{u,v}[\phi], in distribution, as nn\to\infty.

As an immediate corollary of the lemma we have the following.

Corollary 4.2.

As ε0\varepsilon\to 0,

ε2logUε[ϕ]εinfη𝒞d[ϕ(η)+120Th(η(s))h(ξ(s))2𝑑s+J(η)].-\varepsilon^{2}\log U^{\varepsilon}[\phi]\stackrel{{\scriptstyle{\mathbb{P}}^{\varepsilon}}}{{\longrightarrow}}\inf_{\eta\in{\mathcal{C}}_{d}}\left[\phi(\eta)+\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}ds+J(\eta)\right].
Proof.

The proof follows on observing that, V¯ε,0,0[ϕ]=Vε[ϕ]\bar{V}^{\varepsilon,0,0}[\phi]=V^{\varepsilon}[\phi] which has the same distribution as ε2logUε[ϕ]-\varepsilon^{2}\log U^{\varepsilon}[\phi], for η𝒞d\eta\in{\mathcal{C}}_{d},

H(η,ξ00,0)=120Th(η(s))h(ξ(s))2𝑑sH(\eta,\xi_{0}^{0},0)=\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}ds

and that

infη𝒞d[H(η,ξ00,0)+J(η)]=infη𝒞d[120Th(η(s))h(ξ(s))2𝑑s+J(η)]=0.\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{0},0)+J(\eta)\right]=\inf_{\eta\in{\mathcal{C}}_{d}}\left[\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi^{*}(s))\|^{2}ds+J(\eta)\right]=0.

5. Proof of Lemma 4.1.

Let (u,v)𝒜k×𝒜m(u,v)\in{\mathcal{A}}_{k}\times{\mathcal{A}}_{m}. Define canonical coordinate processes on Ωx\mathnormal{\Omega}_{x} as ξ^(ω~)=ω~1\hat{\xi}(\tilde{\omega})=\tilde{\omega}_{1} and γ^(ω~)=ω~2\hat{\gamma}(\tilde{\omega})=\tilde{\omega}_{2}, ω~=(ω~1,ω~2)𝒞d×𝒞k\tilde{\omega}=(\tilde{\omega}_{1},\tilde{\omega}_{2})\in{\mathcal{C}}_{d}\times{\mathcal{C}}_{k}. Note that

exp[ε2V¯ε,u,v[ϕ]]=ΓTε(exp{ε2ϕ()},ζu,v)ΓTε(1,ζε,u,v)\exp\left[-\varepsilon^{-2}\bar{V}^{\varepsilon,u,v}[\phi]\right]=\frac{\Gamma^{\varepsilon}_{T}\left(\exp\{-\varepsilon^{-2}\phi(\cdot)\},\zeta^{u,v}\right)}{\Gamma^{\varepsilon}_{T}\left(1,\zeta^{\varepsilon,u,v}\right)}

and for fCb(𝒞d)f\in C_{b}({\mathcal{C}}_{d}), ε{\mathbb{P}}^{\varepsilon} a.s.,

ΓTε(f,ζε,u,v)\displaystyle\Gamma^{\varepsilon}_{T}\left(f,\zeta^{\varepsilon,u,v}\right) =Ωxf(ξ^(ω~))e1ε20th(ξ^(ω~)(s)),dζu,v(s)12ε20th(ξ^(ω~(s)))2𝑑sμε(dω~).\displaystyle=\int_{\Omega_{x}}f(\hat{\xi}(\tilde{\omega}))e^{\frac{1}{\varepsilon^{2}}\int_{0}^{t}\langle h(\hat{\xi}(\tilde{\omega})(s)),d\zeta^{u,v}(s)\rangle-\frac{1}{2\varepsilon^{2}}\int_{0}^{t}\|h(\hat{\xi}(\tilde{\omega}(s)))\|^{2}ds}\mu^{\varepsilon}(d\tilde{\omega}).

Suppressing ω~\tilde{\omega} in notation, we have

1ε20th(ξ^(s)),dζu,v(s)12ε20th(ξ^(s))2𝑑s\displaystyle\frac{1}{\varepsilon^{2}}\int_{0}^{t}\langle h(\hat{\xi}(s)),d\zeta^{u,v}(s)\rangle-\frac{1}{2\varepsilon^{2}}\int_{0}^{t}\|h(\hat{\xi}(s))\|^{2}ds
=1ε0Th(ξ^(s)),dβ(s)+1ε20Th(ξ^(s))v(s)𝑑s\displaystyle=\frac{1}{\varepsilon}\int_{0}^{T}\langle h(\hat{\xi}(s)),d\beta(s)\rangle+\frac{1}{\varepsilon^{2}}\int_{0}^{T}h(\hat{\xi}(s))\cdot v(s)ds
12ε20Th(ξ^(s))h(ξu(s))2𝑑s+12ε20Th(ξu(s))2𝑑s\displaystyle\quad-\frac{1}{2\varepsilon^{2}}\int_{0}^{T}\|h(\hat{\xi}(s))-h(\xi^{u}(s))\|^{2}ds+\frac{1}{2\varepsilon^{2}}\int_{0}^{T}\|h(\xi^{u}(s))\|^{2}ds
=1ε0Th(ξ^(s)),dβ(s)1ε2H(ξ^,ξu,v)\displaystyle=\frac{1}{\varepsilon}\int_{0}^{T}\langle h(\hat{\xi}(s)),d\beta(s)\rangle-\frac{1}{\varepsilon^{2}}H(\hat{\xi},\xi^{u},v)
+12ε20T(h(ξu(s))2+v(s)2)𝑑s+1ε20Th(ξu(s))v(s)𝑑s.\displaystyle\quad+\frac{1}{2\varepsilon^{2}}\int_{0}^{T}(\|h(\xi^{u}(s))\|^{2}+\|v(s)\|^{2})ds+\frac{1}{\varepsilon^{2}}\int_{0}^{T}h(\xi^{u}(s))\cdot v(s)ds.

Thus, letting

(5.1) F(ω~,β)0Th(ξ^(ω~)(s)),dβ(s)F(\tilde{\omega},\beta)\doteq\int_{0}^{T}\langle h(\hat{\xi}(\tilde{\omega})(s)),d\beta(s)\rangle

we can write

(5.2) eε2V¯ε,u,v[ϕ]=Ωxe1εF(ω~,β)1ε2(ϕ(ξ^(ω~))+H(ξ^(ω~),ξu,v))με(dω~)Ωxe1εF(ω~,β)1ε2H(ξ^(ω~),ξu,v)με(dω~).e^{-\varepsilon^{-2}\bar{V}^{\varepsilon,u,v}[\phi]}=\frac{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon}F(\tilde{\omega},\beta)-\frac{1}{\varepsilon^{2}}(\phi(\hat{\xi}(\tilde{\omega}))+H(\hat{\xi}(\tilde{\omega}),\xi^{u},v))}\mu^{\varepsilon}(d\tilde{\omega})}{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon}F(\tilde{\omega},\beta)-\frac{1}{\varepsilon^{2}}H(\hat{\xi}(\tilde{\omega}),\xi^{u},v)}\mu^{\varepsilon}(d\tilde{\omega})}.

Let now εn,un,vn,u,v\varepsilon_{n},u_{n},v_{n},u,v be as in the statement of Lemma 4.1. Using Assumption 1 it is immediate that

(5.3) (un,vn,ξεn,un,ζεn,un,vn,β)(u,v,ξ0u,ζ0u,v,β)(u_{n},v_{n},\xi^{\varepsilon_{n},u_{n}},\zeta^{\varepsilon_{n},u_{n},v_{n}},\beta)\Rightarrow(u,v,\xi^{u}_{0},\zeta_{0}^{u,v},\beta)

in SM×𝒞d×𝒞m×𝒞mS_{M}\times{\mathcal{C}}_{d}\times{\mathcal{C}}_{m}\times{\mathcal{C}}_{m}, where

ζ0u,v(t)=0th(ξ0u(s))𝑑s+0tv(s)𝑑s,t[0,T].\zeta_{0}^{u,v}(t)=\int_{0}^{t}h(\xi_{0}^{u}(s))ds+\int_{0}^{t}v(s)ds,\;t\in[0,T].

By appealing to Skorohod representation theorem we can obtain, on some probability space (Ω,,)(\mathnormal{\Omega}^{*},{\mathcal{F}}^{*},{\mathbb{P}}^{*}), random variables (u~n,v~n,ξ~n,ζ~n,β~n)(\tilde{u}_{n},\tilde{v}_{n},\tilde{\xi}^{n},\tilde{\zeta}^{n},\tilde{\beta}^{n}) with same law as the random vector on the left side of (5.3) and (u~,v~,ξ~0,ζ~0,β~)(\tilde{u},\tilde{v},\tilde{\xi}_{0},\tilde{\zeta}_{0},\tilde{\beta}) with same law as the vector on the right side of (5.3), such that

(5.4) (u~n,v~n,ξ~n,ζ~n,β~n)(u~,v~,ξ~0,ζ~0,β~), a.s.(\tilde{u}_{n},\tilde{v}_{n},\tilde{\xi}^{n},\tilde{\zeta}^{n},\tilde{\beta}^{n})\to(\tilde{u},\tilde{v},\tilde{\xi}_{0},\tilde{\zeta}_{0},\tilde{\beta}),\;{\mathbb{P}}^{*}-\mbox{ a.s.}

Henceforth, to simplify notation we will drop the ~\tilde{\cdot} from the notation in the above vectors and denote the corresponding process V¯εn,un,vn[ϕ]\bar{V}^{\varepsilon_{n},u_{n},v_{n}}[\phi] as V¯n[ϕ]\bar{V}^{n}[\phi]. Then, from (5.2), and the distributional equality noted above, it follows that

(5.5) eεn2V¯n[ϕ]=Ωxe1εnF(ω~,βn)1εn2(ϕ(ξ^(ω~))+H(ξ^(ω~),ξn,vn))μεn(dω~)Ωxe1εnF(ω~,βn)1εn2H(ξ^(ω~),ξn,vn)μεn(dω~)=Ωxe1εnF(ω~,βn)1εn2(ϕ(ξ^(ω~))+H(ξ^(ω~),ξn,v)0Th(ξ^(ω~)(s))(vn(s)v(s))𝑑s)μεn(dω~)Ωxe1εnF(ω~,βn)1εn2(H(ξ^(ω~),ξn,v)0Th(ξ^(ω~)(s))(vn(s)v(s))𝑑s)μεn(dω~).e^{-\varepsilon_{n}^{-2}\bar{V}^{n}[\phi]}=\frac{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}(\phi(\hat{\xi}(\tilde{\omega}))+H(\hat{\xi}(\tilde{\omega}),\xi^{n},v^{n}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})}{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}H(\hat{\xi}(\tilde{\omega}),\xi^{n},v^{n})}\mu^{\varepsilon_{n}}(d\tilde{\omega})}\\ =\frac{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}\left(\phi(\hat{\xi}(\tilde{\omega}))+H(\hat{\xi}(\tilde{\omega}),\xi^{n},v)-\int_{0}^{T}h(\hat{\xi}(\tilde{\omega})(s))\cdot(v^{n}(s)-v(s))ds\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega})}{\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}\left(H(\hat{\xi}(\tilde{\omega}),\xi^{n},v)-\int_{0}^{T}h(\hat{\xi}(\tilde{\omega})(s))\cdot(v^{n}(s)-v(s))ds\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega})}.

In order to prove Lemma 4.1 it now suffices to show that, for all ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}), as nn\to\infty,

(5.6) Υ¯1n[ϕ]εn2log[Ωxe1εnF(ω~,βn)1εn2(ϕ(ξ^(ω~))+H(ξ^(ω~),ξn,v)0Th(ξ^(ω~)(s))(vn(s)v(s))𝑑s)μεn(dω~)]infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)] a.s. .\bar{\Upsilon}_{1}^{n}[\phi]\doteq-\varepsilon_{n}^{-2}\log\left[\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}\left(\phi(\hat{\xi}(\tilde{\omega}))+H(\hat{\xi}(\tilde{\omega}),\xi^{n},v)-\int_{0}^{T}h(\hat{\xi}(\tilde{\omega})(s))\cdot(v^{n}(s)-v(s))ds\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega})\right]\\ \to\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right]\mbox{ a.s. }{\mathbb{P}}^{*}.

Define Δ1n:𝒞d×Ω\Delta^{n}_{1}:{\mathcal{C}}_{d}\times\mathnormal{\Omega}^{*}\to{\mathbb{R}} as

(5.7) Δ1n(η)\displaystyle\Delta^{n}_{1}(\eta) =H(η,ξ0,v)H(η,ξn,v)+0Th(η(s))(vn(s)v(s))𝑑s\displaystyle=H(\eta,\xi_{0},v)-H(\eta,\xi^{n},v)+\int_{0}^{T}h(\eta(s))\cdot(v^{n}(s)-v(s))ds
=120T(2(h(η(s))v(s))(h(ξn(s))h(ξ0(s)))+h(ξ0(s))2h(ξn(s))2)𝑑s\displaystyle=\frac{1}{2}\int_{0}^{T}\left(2\left(h(\eta(s))-v(s)\right)\cdot\left(h(\xi^{n}(s))-h(\xi_{0}(s))\right)+\|h(\xi_{0}(s))\|^{2}-\|h(\xi^{n}(s))\|^{2}\right)ds
+0Th(η(s))(vn(s)v(s))𝑑s.\displaystyle\quad+\int_{0}^{T}h(\eta(s))\cdot\left(v^{n}(s)-v(s)\right)ds.

Then from the continuity of hh and the a.s. convergence in (5.4), we see that for every η𝒞d\eta\in{\mathcal{C}}_{d}

(5.8)  as n,Δ1n(η)0, a.s. .\mbox{ as }n\to\infty,\;\;\Delta_{1}^{n}(\eta)\to 0,\mbox{ a.s. }{\mathbb{P}}^{*}.

Furthermore, with Δn(ω~,ω)Δ1n(ξ^(ω~),ω)\Delta^{n}(\tilde{\omega},\omega^{*})\doteq\Delta_{1}^{n}(\hat{\xi}(\tilde{\omega}),\omega^{*}),

(5.9) Υ¯1n[ϕ]=εn2log[Ωxe1εnF(ω~,βn)1εn2(ϕ(ξ^(ω~))+H(ξ^(ω~),ξ0,v)Δn)μεn(dω~)].\bar{\Upsilon}_{1}^{n}[\phi]=-\varepsilon_{n}^{2}\log\left[\int_{\Omega_{x}}e^{\frac{1}{\varepsilon_{n}}F(\tilde{\omega},\beta^{n})-\frac{1}{\varepsilon_{n}^{2}}\left(\phi(\hat{\xi}(\tilde{\omega}))+H(\hat{\xi}(\tilde{\omega}),\xi_{0},v)-\Delta^{n}\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega})\right].

In order to prove (5.6) we will show

(5.10) lim supnΥ¯1n[ϕ]infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)] a.s. \limsup_{n\to\infty}\bar{\Upsilon}_{1}^{n}[\phi]\leq\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right]\mbox{ a.s. }{\mathbb{P}}^{*}

and

(5.11) lim infnΥ¯1n[ϕ]infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)] a.s. .\liminf_{n\to\infty}\bar{\Upsilon}_{1}^{n}[\phi]\geq\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right]\mbox{ a.s. }{\mathbb{P}}^{*}.

The fact that FF can be neglected in the asymptotic formula follows along the lines of [13], however since, unlike [13], we do not assume that hh is bounded and our functional of interest is somewhat different from the one considered in [13], we provide the details.

5.1. Proof of (5.11)

We begin with the following lemmas.

Lemma 5.1.

For any C(0,)C\in(0,\infty),

lim supε0ε2log𝒞xexp(Cε2ξ^(ω~))με(dω~)<.\displaystyle\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{{\mathcal{C}}_{x}}\exp\left(C\varepsilon^{-2}\|\hat{\xi}(\tilde{\omega})\|_{*}\right)\mu^{\varepsilon}(d\tilde{\omega})<\infty.
Proof.

Note that for t[0,T]t\in[0,T]

ξ^(t)=x0+0tb(ξ^(s))𝑑s+ε0tσ(ξ^(s))𝑑γ^(s).\hat{\xi}(t)=x_{0}+\int_{0}^{t}b(\hat{\xi}(s))ds+\varepsilon\int_{0}^{t}\sigma(\hat{\xi}(s))d\hat{\gamma}(s).

Let M(t)0tσ(ξ^(s))𝑑γ^(s)M(t)\doteq\int_{0}^{t}\sigma(\hat{\xi}(s))d\hat{\gamma}(s). Then by an application of Gronwall’s lemma, it suffices to show that

lim supε0ε2logEμεexp(Cε1M)<\limsup_{\varepsilon\to 0}\varepsilon^{2}\log E_{\mu^{\varepsilon}}\exp\left(C\varepsilon^{-1}\|M\|_{*}\right)<\infty

where EμεE_{\mu^{\varepsilon}} is the expectation under the probability measure με\mu^{\varepsilon}. Since σ\sigma is bounded and under με\mu^{\varepsilon}, γ^\hat{\gamma} is a Brownian motion, there is C1(0,)C_{1}\in(0,\infty) such that Eμεexp(Cε1M)C1exp(C1ε2)E_{\mu^{\varepsilon}}\exp\left(C\varepsilon^{-1}\|M\|_{*}\right)\leq C_{1}\exp\left(C_{1}\varepsilon^{-2}\right) for every ε>0\varepsilon>0. The result follows. ∎

Lemma 5.2.

Let for ε>0\varepsilon>0, ¯ε\bar{\mathcal{R}}^{\varepsilon} and A¯ε\bar{A}^{\varepsilon} be measurable maps from 𝒞d{\mathcal{C}}_{d} to {\mathbb{R}} such that

(5.12) supε>0¯ε(η)cR(1+η),supε>0|A¯ε(η)|cA(1+η) for all η𝒞d.\sup_{\varepsilon>0}\bar{\mathcal{R}}^{\varepsilon}(\eta)\leq c_{R}(1+\|\eta\|_{*}),\;\;\sup_{\varepsilon>0}|\bar{A}^{\varepsilon}(\eta)|\leq c_{A}(1+\|\eta\|_{*})\mbox{ for all }\eta\in{\mathcal{C}}_{d}.

Then

(5.13) lim supε0ε2logΩxeε1A¯ε(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~)lim supε0ε2logΩxeε2¯ε(ξ^(ω~))με(dω~)\displaystyle\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega})\leq\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega})

and for every c0(0,)c_{0}\in(0,\infty)

(5.14) lim supMlim supε0ε2logΩxeε1A¯ε(ξ^(ω~))+ε2c0(1+ξ^(ω~))1{A¯ε(ξ^(ω~)M}με(dω~)=.\displaystyle\limsup_{M\to\infty}\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}c_{0}(1+\|\hat{\xi}(\tilde{\omega})\|_{*})}1_{\{\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega})\geq M\}}\mu^{\varepsilon}(d\tilde{\omega})=-\infty.
Proof.

For M(0,)M\in(0,\infty), let AMεA¯εMA_{M}^{\varepsilon}\doteq\bar{A}^{\varepsilon}\wedge M. Then

Ωxeε1A¯ε(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~)\displaystyle\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega}) Ωxeε1AMε(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~)\displaystyle\leq\int_{\Omega_{x}}e^{\varepsilon^{-1}A_{M}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega})
+Ωxeε1A¯ε(ξ^(ω~))+ε2¯ε(ξ^(ω~))1{A¯ε(ξ^(ω~)M}με(dω~).\displaystyle\quad+\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}1_{\{\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega})\geq M\}}\mu^{\varepsilon}(d\tilde{\omega}).

Thus

lim supε0ε2logΩxeε1A¯ε(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~)max{lim supε0ε2logΩxeε1AMε(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~),lim supε0ε2logΩxeε1A¯ε(ξ^(ω~))+ε2¯ε(ξ^(ω~))1{A¯ε(ξ^(ω~)M}με(dω~)}.\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega})\\ \leq\max\Big{\{}\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}A_{M}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega}),\\ \limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}1_{\{\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega})\geq M\}}\mu^{\varepsilon}(d\tilde{\omega})\Big{\}}.

Since

lim supε0ε2logΩxeε1AM(ξ^(ω~))+ε2¯ε(ξ^(ω~))με(dω~)=lim supε0ε2logΩxeε2¯ε(ξ^(ω~))με(dω~),\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}A_{M}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega})=\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-2}\bar{\mathcal{R}}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))}\mu^{\varepsilon}(d\tilde{\omega}),

in order to prove the lemma it suffices to show (5.14) for every c0(0,)c_{0}\in(0,\infty). Fix ε(0,1)\varepsilon\in(0,1). Using the fact that, on the set {A¯ε(ξ^(ω~))M}\{\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))\geq M\},

ε1A¯ε(ξ^(ω~))ε2(A¯ε(ξ^(ω~))M)+ε1M,\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))\leq\varepsilon^{-2}(\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))-M)+\varepsilon^{-1}M,

and the bound in (5.12), we see that

lim supε0ε2logΩxeε1A¯ε(ξ^(ω~))+ε2c0(1+ξ^(ω~))1{A¯ε(ξ^(ω~)M}με(dω~)\displaystyle\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-1}\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega}))+\varepsilon^{-2}c_{0}(1+\|\hat{\xi}(\tilde{\omega})\|_{*})}1_{\{\bar{A}^{\varepsilon}(\hat{\xi}(\tilde{\omega})\geq M\}}\mu^{\varepsilon}(d\tilde{\omega})
M+lim supε0ε2logΩxeε2(cA+c0)(1+ξ^(ω~))με(dω~).\displaystyle\leq-M+\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\int_{\Omega_{x}}e^{\varepsilon^{-2}(c_{A}+c_{0})(1+\|\hat{\xi}(\tilde{\omega})\|_{*})}\mu^{\varepsilon}(d\tilde{\omega}).

The inequality in (5.14) now follows on applying Lemma 5.1. ∎

Note that, by Itô’s formula,

F(ω~,βn)\displaystyle F(\tilde{\omega},\beta^{n}) =0Th(ξ^(ω~)(s)),dβn(s)\displaystyle=\int_{0}^{T}\langle h(\hat{\xi}(\tilde{\omega})(s)),d\beta^{n}(s)\rangle
=h(ξ^(ω~)(T)),βn(T)l=1m0Tβln(s)hl(ξ^(ω~)(s)),dξ^(ω~)(s)\displaystyle=\langle h(\hat{\xi}(\tilde{\omega})(T)),\beta^{n}(T)\rangle-\sum_{l=1}^{m}\int_{0}^{T}\beta^{n}_{l}(s)\langle\nabla h_{l}(\hat{\xi}(\tilde{\omega})(s)),d\hat{\xi}(\tilde{\omega})(s)\rangle
ε22i,j=1kl=1m0Tβln(s)(σσ)ij(ξ^(ω~)(s))2hlxixj(ξ^(ω~)(s))𝑑s\displaystyle-\frac{\varepsilon^{2}}{2}\sum_{i,j=1}^{k}\sum_{l=1}^{m}\int_{0}^{T}\beta^{n}_{l}(s)(\sigma\sigma^{\dagger})_{ij}(\hat{\xi}(\tilde{\omega})(s))\frac{\partial^{2}h_{l}}{\partial x_{i}\partial x_{j}}(\hat{\xi}(\tilde{\omega})(s))ds
=h(ξ^(ω~)(T)),βn(T)l=1m0Tβln(s)hl(ξ^(ω~)(s)),b(ξ^(ω~)(s))𝑑s\displaystyle=\langle h(\hat{\xi}(\tilde{\omega})(T)),\beta^{n}(T)\rangle-\sum_{l=1}^{m}\int_{0}^{T}\beta^{n}_{l}(s)\langle\nabla h_{l}(\hat{\xi}(\tilde{\omega})(s)),b(\hat{\xi}(\tilde{\omega})(s))\rangle ds
ε22i,j=1kl=1m0Tβln(s)(σσ)ij(ξ^(ω~)(s))2hlxixj(ξ^(ω~)(s))𝑑s\displaystyle-\frac{\varepsilon^{2}}{2}\sum_{i,j=1}^{k}\sum_{l=1}^{m}\int_{0}^{T}\beta^{n}_{l}(s)(\sigma\sigma^{\dagger})_{ij}(\hat{\xi}(\tilde{\omega})(s))\frac{\partial^{2}h_{l}}{\partial x_{i}\partial x_{j}}(\hat{\xi}(\tilde{\omega})(s))ds
l=1m0Tβln(s)hl(ξ^(ω~)(s)),(dξ^(ω~)(s)b(ξ^(ω~)(s))ds)\displaystyle-\sum_{l=1}^{m}\int_{0}^{T}\beta^{n}_{l}(s)\left\langle\nabla h_{l}(\hat{\xi}(\tilde{\omega})(s)),\left(d\hat{\xi}(\tilde{\omega})(s)-b(\hat{\xi}(\tilde{\omega})(s))ds\right)\right\rangle
=AT(ξ^(ω~),βn)+KT(ξ^(ω~),βn),\displaystyle=A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n})+K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}),

where, ε{\mathbb{P}}^{\varepsilon} a.s.,

KT(ξ,β)l=1m0Tβl(s)hl(ξ(s)),(dξ(s)b(ξ(s))ds)K_{T}(\xi,\beta)\doteq-\sum_{l=1}^{m}\int_{0}^{T}\beta_{l}(s)\left\langle\nabla h_{l}(\xi(s)),\left(d\xi(s)-b(\xi(s))ds\right)\right\rangle

and AT(ξ,β)=0Th(ξ(s)),dβ(s)KT(ξ,β)A_{T}(\xi,\beta)=\int_{0}^{T}\langle h(\xi(s)),d\beta(s)\rangle-K_{T}(\xi,\beta).

The following result is taken from Heunis[13](cf. page 940 therein).

Proposition 5.3 (Heunis[13]).

The maps KTK_{T} and ATA_{T} are measurable and continuous, respectively, from 𝒞d×𝒞m{\mathcal{C}}_{d}\times{\mathcal{C}}_{m} to {\mathbb{R}}, and there are c1,c2(0,)c_{1},c_{2}\in(0,\infty) such that for all x>0x>0, n1n\geq 1,

μεn(ω~:|KT(ξ^(ω~),βn)|>x)2exp{c1x2εn2(1+βn2)}, a.s. \displaystyle\mu^{\varepsilon_{n}}(\tilde{\omega}:|K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n})|>x)\leq 2\exp\left\{-c_{1}\frac{x^{2}}{\varepsilon_{n}^{2}(1+\|\beta^{n}\|_{*}^{2})}\right\},\;\text{ a.s. }{\mathbb{P}}^{*}

and

(5.15) |AT(ξ^(ω~),βn)|c2(1+ξ^(ω~)+βn) a.s. μεn.|A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n})|\leq c_{2}(1+\|\hat{\xi}(\tilde{\omega})\|_{*}+\|\beta^{n}\|_{*})\mbox{ a.s. }\mu^{\varepsilon_{n}}\otimes{\mathbb{P}}^{*}.

Define

(5.16) Gn(ω~,ω)ϕ(ξ^(ω~))H(ξ^(ω~),ξ0(ω),v(ω))+Δn(ω~,ω),(ω~,ω)Ωx×Ω.G^{n}(\tilde{\omega},\omega^{*})\doteq-\phi(\hat{\xi}(\tilde{\omega}))-H(\hat{\xi}(\tilde{\omega}),\xi_{0}(\omega^{*}),v(\omega^{*}))+\Delta^{n}(\tilde{\omega},\omega^{*}),\;(\tilde{\omega},\omega^{*})\in\Omega_{x}\times\mathnormal{\Omega}^{*}.
Proposition 5.4.

For any δ(0,)\delta\in(0,\infty), and {\mathbb{P}}^{*} a.e. ω\omega^{*}

lim supnεn2log{|εnKT(ξ^(ω~),βn(ω))|>δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)=,\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{|\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))|>\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty,
(5.17) lim supnεn2log{εnKT(ξ^(ω~),βn(ω))<δ}eεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))μεn(dω~)=.\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.
Proof.

Note that on the set {εnKT(ξ^(ω~),βn(ω))<δ}\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}

εn2Gn(ω~,ω)+εn1F(ω~,βn(ω))εn2(Gn(ω~,ω)δ)+εn1AT(ξ^(ω~),βn(ω)).\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))\leq\varepsilon_{n}^{-2}(G^{n}(\tilde{\omega},\omega^{*})-\delta)+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*})).

Also note that, using the linear growth of hh, one can find a measurable map θ:Ω+\theta:\mathnormal{\Omega}^{*}\to{\mathbb{R}}_{+} such that

(5.18) Gn(ω~,ω)θ(ω)(1+ξ^(ω~)), for all ω~Ωx, a.e. ω.G^{n}(\tilde{\omega},\omega^{*})\leq\theta(\omega^{*})(1+\|\hat{\xi}(\tilde{\omega})\|_{*}),\mbox{ for all }\tilde{\omega}\in\Omega_{x},\;{\mathbb{P}}^{*}\mbox{ a.e. }\omega^{*}.

Using these observations, we have

{εnKT(ξ^(ω~),βn(ω))<δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
(5.19) eεn2(θ(ω)δ){εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)+εn1AT(ξ^(ω~),βn(ω))μεn(dω~).\displaystyle\leq e^{\varepsilon_{n}^{-2}(\theta(\omega^{*})-\delta)}\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

Next, for every M(0,)M\in(0,\infty)

{εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)+εn1AT(ξ^(ω~),βn(ω))μεn(dω~)\displaystyle\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
{εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)+εn1Mμεn(dω~)\displaystyle\leq\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}+\varepsilon_{n}^{-1}M}\mu^{\varepsilon_{n}}(d\tilde{\omega})
(5.20) +{εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)+εn1AT(ξ^(ω~),βn(ω))1{AT(ξ^(ω~),βn(ω))M}μεn(dω~).\displaystyle\quad+\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}1_{\{A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\geq M\}}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

We now consider the two terms in the above display separately. For the first term, from Cauchy-Schwarz inequality,

{εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)μεn(dω~)\displaystyle\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}}\mu^{\varepsilon_{n}}(d\tilde{\omega})
[Ωxe2εn2θ(ω)ξ^(ω~)μεn(dω~)]1/2[μεn{εnKT(ξ^(ω~),βn(ω))<δ}]1/2\displaystyle\leq\left[\int_{\Omega_{x}}e^{2\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}}\mu^{\varepsilon_{n}}(d\tilde{\omega})\right]^{1/2}\left[\mu^{\varepsilon_{n}}\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}\right]^{1/2}

and therefore

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))<δ}eεn2θ(ω)ξ^(ω~)μεn(dω~)\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}}\mu^{\varepsilon_{n}}(d\tilde{\omega})
lim supnεn22logΩxe2εn2θ(ω)ξ^(ω~)μεn(dω~)+lim supnεn22logμεn{KT(ξ^(ω~),βn(ω))<δεn}\displaystyle\leq\limsup_{n\to\infty}\frac{\varepsilon_{n}^{2}}{2}\log\int_{\Omega_{x}}e^{2\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}}\mu^{\varepsilon_{n}}(d\tilde{\omega})+\limsup_{n\to\infty}\frac{\varepsilon_{n}^{2}}{2}\log\mu^{\varepsilon_{n}}\{K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<\frac{-\delta}{\varepsilon_{n}}\}
lim supnεn22logΩxe2εn2θ(ω)ξ^(ω~)μεn(dω~)c1δ22εn2(1+βn2)\displaystyle\leq\limsup_{n\to\infty}\frac{\varepsilon_{n}^{2}}{2}\log\int_{\Omega_{x}}e^{2\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}}\mu^{\varepsilon_{n}}(d\tilde{\omega})-c_{1}\frac{\delta^{2}}{2\varepsilon_{n}^{2}(1+\|\beta^{n}\|_{*}^{2})}
=\displaystyle=-\infty

where in the next to last line we have used Proposition 5.3 and in the last line we have appealed to Lemma 5.1 and the fact that supnβn<\sup_{n}\|\beta^{n}\|_{*}<\infty {\mathbb{P}}^{*} a.s.

For the second term on the right side in (5.20), we have from Lemma 5.2 (see (5.14)) and (5.15) that

lim supMlim supnεn2logΩxeεn2θ(ω)ξ^(ω~)+εn1AT(ξ^(ω~),βn(ω))1{AT(ξ^(ω~),βn(ω)M}με(dω~)=.\displaystyle\limsup_{M\to\infty}\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}\theta(\omega^{*})\|\hat{\xi}(\tilde{\omega})\|_{*}+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}1_{\{A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*})\geq M\}}\mu^{\varepsilon}(d\tilde{\omega})=-\infty.

Using the last two displays in (5.20) and combining with (5.19) we have (5.17) and

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))<δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)=.\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.

Next, from [13, Proposition 4.74.7], it follows that

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))>δ}e2εn1KT(ξ^(ω~),βn(ω))μεn(dω~)=.\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))>\delta\}}e^{2\varepsilon_{n}^{-1}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.

Now using Cauchy-Schwarz inequality and arguing as before, we see that

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))>δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)=.\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))>\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.

We omit the details.

The following proposition shows that the term involving FF in the definition of Υ¯1n[ϕ]\bar{\Upsilon}_{1}^{n}[\phi] can be ignored in proving the bound in (5.11).

Proposition 5.5.

For {\mathbb{P}}^{*} a.e. ω\omega^{*},

lim supnεn2logΩxeεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)lim supnεn2logΩxeεn2Gn(ω~,ω)μεn(dω~).\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})\leq\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega}).
Proof.

Fix δ(0,)\delta\in(0,\infty) and write

Ωxeεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
={εnKT(ξ^(ω~),βn(ω))>δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle=\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))>\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
+{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~).\displaystyle\quad+\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\leq\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

From Proposition 5.4,

(5.21) lim supnεn2log{εnKT(ξ^(ω~),βn(ω))>δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)=.\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))>\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.

Next note that

{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\leq\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+δεn2+εn1AT(ξ^(ω~),βn(ω))μεn(dω~).\displaystyle\leq\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\leq\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\delta\varepsilon_{n}^{-2}+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

Now recalling (5.15) and (5.18) and applying the first inequality in Lemma 5.2 (i.e. (5.13)), we get

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\leq\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
δ+lim supnεn2logΩxeεn2Gn(ω~,ω)μεn(dω~).\displaystyle\leq\delta+\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

Since δ>0\delta>0 is arbitrary, the result follows on combining the above with (5.21). ∎

The proof of the following lemma follows along the lines of Varadhan’s lemma (cf. [23, Theorem 2.6], [6, Theorem 1.18 ]). We provide details for reader’s convenience.

Lemma 5.6.

Let {Zε}ε>0\{Z^{\varepsilon}\}_{\varepsilon>0} be a collection of random variables with values in a Polish space (𝒳,d(,))({\mathcal{X}},d(\cdot,\cdot)) that satisfies a LDP with rate function JJ and speed ε2\varepsilon^{-2}. Let ϕ:𝒳\phi:{\mathcal{X}}\rightarrow\mathbb{R} be a continuous function bounded from above, namely supx𝒳ϕ(x)<\sup_{x\in{\mathcal{X}}}\phi(x)<\infty, and let {ψε}ε>0\{\psi^{\varepsilon}\}_{\varepsilon>0} be a collection of real measurable maps on 𝒳{\mathcal{X}} such that supε>0supx𝒳|ψε(x)|<\sup_{\varepsilon>0}\sup_{x\in{\mathcal{X}}}|\psi^{\varepsilon}(x)|<\infty. Further suppose that

for every δ>0 and x𝒳,\displaystyle\text{ for every $\delta>0$ and $x\in{\mathcal{X}}$}, there exist ε0(x)δ1(x)(0,) such that |ψε(y)|<δ,\displaystyle\text{ there exist $\varepsilon_{0}(x)$, $\delta_{1}(x)\in(0,\infty)$ such that }|\psi^{\varepsilon}(y)|<\delta,\;
for all d(x,y)<δ1(x) and all 0<ε<ε0(x).\displaystyle\mbox{ for all }d(x,y)<\delta_{1}(x)\text{ and all }0<\varepsilon<\varepsilon_{0}(x).

Then

limε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})]=supx𝒳[ϕ(x)J(x)].\displaystyle\lim_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)]=\sup_{x\in{\mathcal{X}}}\left[\phi(x)-J(x)\right].
Proof.

Define Rsupx𝒳(ϕ(x)+supε>0|ψε(x)|)R\doteq\sup_{x\in{\mathcal{X}}}(\phi(x)+\sup_{\varepsilon>0}|\psi^{\varepsilon}(x)|), Ssupx𝒳(ϕ(x)J(x))S\doteq\sup_{x\in{\mathcal{X}}}(\phi(x)-J(x)) and K{x𝒳:J(x)|S|+R}K\doteq\{x\in{\mathcal{X}}:J(x)\leq|S|+R\}. Since JJ is a rate function, KK is a compact subset of 𝒳{\mathcal{X}}.

Fix δ(0,)\delta\in(0,\infty). From the hypothesis of the lemma, for each x𝒳x\in{\mathcal{X}}, there exist δ1(x),ε0(x)(0,)\delta_{1}(x),\varepsilon_{0}(x)\in(0,\infty) such that |ψε(y)|<δ|\psi^{\varepsilon}(y)|<\delta for every yB(x,δ1(x))y\in B(x,\delta_{1}(x)) and ε(0,ε0(x))\varepsilon\in(0,\varepsilon_{0}(x)), where B(z,γ){x𝒳:d(x,z)<γ}B(z,\gamma)\doteq\{x\in{\mathcal{X}}:d(x,z)<\gamma\} is an open ball of radius γ\gamma in 𝒳{\mathcal{X}}. Also, from the continuity of ϕ\phi, for every x𝒳x\in{\mathcal{X}} there exists δ2(x)(0,)\delta_{2}(x)\in(0,\infty) such that

|ϕ(x)ϕ(y)|<δ,yB(x,δ2(x)).\displaystyle|\phi(x)-\phi(y)|<\delta,\;\forall y\in B(x,\delta_{2}(x)).

Next, from the lower semi-continuity of JJ, for every x𝒳x\in{\mathcal{X}}, there exists δ3(x)(0,)\delta_{3}(x)\in(0,\infty) such that

J(x)infyB(x,δ3(x))¯J(y)+δ.\displaystyle J(x)\leq\inf_{y\in\overline{B(x,\delta_{3}(x))}}J(y)+\delta.

Let δ¯(x)min{δ1(x),δ2(x),δ3(x)}\bar{\delta}(x)\doteq\min\{\delta_{1}(x),\delta_{2}(x),\delta_{3}(x)\}. Now define an open cover xKU(x)\cup_{x\in K}U(x) of KK using the following open sets:

U(x)B(x,δ¯(x)),xK.\displaystyle U(x)\doteq B(x,\bar{\delta}(x)),\;x\in K.

Note that for any, xKx\in K, yU(x)y\in U(x) and ε<ε0(x)\varepsilon<\varepsilon_{0}(x), we have

(5.22) |ψε(y)|<δ, |ϕ(x)ϕ(y)|<δ and J(x)infzU(x)¯J(z)+δ.\displaystyle|\psi^{\varepsilon}(y)|<\delta,\text{ }|\phi(x)-\phi(y)|<\delta\text{ and }J(x)\leq\inf_{z\in\overline{U(x)}}J(z)+\delta.

Since KK is compact, there exists NN\in{\mathbb{N}} and {xi}i=1NK\{x_{i}\}_{i=1}^{N}\subset K such that {UiU(xi)}i=1N\{U_{i}\doteq U(x_{i})\}_{i=1}^{N}cover KK. For i=1,,Ni=1,\ldots,N, we can find 0<ε(xi)ε0(xi)0<\varepsilon(x_{i})\leq\varepsilon_{0}(x_{i}) such that with ε¯0mini=1,,Nε(xi)\bar{\varepsilon}_{0}\doteq\min_{i=1,\ldots,N}\varepsilon(x_{i}), for every ε<ε¯0\varepsilon<\bar{\varepsilon}_{0},

(5.23) [ZεUi¯]exp[ε2(bi+δ)],[ZεF]exp[ε2(infxFJ(x)+δ)]\displaystyle\mathbb{P}[Z^{\varepsilon}\in\overline{U_{i}}]\leq\exp\left[\varepsilon^{-2}(-b_{i}+\delta)\right],\;\;\;\mathbb{P}[Z^{\varepsilon}\in F]\leq\exp\left[\varepsilon^{-2}(-\inf_{x\in F}J(x)+\delta)\right]

where, F(i=1NUi)cF\doteq\left(\cup_{i=1}^{N}{U_{i}}\right)^{c} and biinfxUi¯J(x)b_{i}\doteq\inf_{x\in\overline{U_{i}}}J(x).

Next note that

𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})]\displaystyle\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)] =𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1i=1NUi(Zε)]\displaystyle=\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{\cup_{i=1}^{N}{U_{i}}}(Z^{\varepsilon})]
+𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1F(Zε)]\displaystyle\quad+\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{F}(Z^{\varepsilon})]
i=1N𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1Ui(Zε)]\displaystyle\leq\sum_{i=1}^{N}\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{{U}_{i}}(Z^{\varepsilon})]
(5.24) +𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1F(Zε)].\displaystyle\quad+\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{F}(Z^{\varepsilon})].

Defining aiinfxUi¯ϕ(x)a_{i}\doteq\inf_{x\in\overline{U_{i}}}\phi(x), we have |aiϕ(x)|<2δ|a_{i}-\phi(x)|<2\delta, for xUix\in{U_{i}}. Thus, using (5.22) and (5.23)

lim supε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1Ui(Zε)]\displaystyle\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{{U}_{i}}(Z^{\varepsilon})]
(5.25) (aibi+4δ)ϕ(xi)J(xi)+5δsupx𝒳[ϕ(x)J(x)]+5δ.\displaystyle\leq(a_{i}-b_{i}+4\delta)\leq\phi(x_{i})-J(x_{i})+5\delta\leq\sup_{x\in{\mathcal{X}}}\left[\phi(x)-J(x)\right]+5\delta.

Also

lim supε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1F]RinfxFJ(x)+δ\displaystyle\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{F}]\leq R-\inf_{x\in F}J(x)+\delta
(5.26) |S|+δsupx𝒳[ϕ(x)J(x)]+δ,\displaystyle\quad\leq-|S|+\delta\leq\sup_{x\in{\mathcal{X}}}\left[\phi(x)-J(x)\right]+\delta,

where the second inequality is a consequence of the observation that FKcF\subset K^{c}. Since δ>0\delta>0 is arbitrary, using (5.25) and (5.26) in (5.24) we now see that

(5.27) lim supε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})]supx𝒳[ϕ(x)J(x)].\limsup_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)]\leq\sup_{x\in{\mathcal{X}}}\left[\phi(x)-J(x)\right].

For the lower bound, choose x0x_{0} such that ϕ(x0)J(x0)Sδ\phi(x_{0})-J(x_{0})\geq S-\delta. Let δ(x0),ε(x0)(0,)\delta(x_{0}),\varepsilon(x_{0})\in(0,\infty) be such that for all xUB(x0,δ(x0))x\in U\doteq B(x_{0},\delta(x_{0})), |ϕ(x)ϕ(x0)|<δ|\phi(x)-\phi(x_{0})|<\delta and |ψε(x)|<δ|\psi^{\varepsilon}(x)|<\delta, for ε<ε(x0)\varepsilon<\varepsilon(x_{0}). Then

lim infε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})]\displaystyle\liminf_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}\left[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)\right]
lim infε0ε2log𝔼[exp(ε2{ϕ(Zε)+ψε(Zε)})1U(Zε)]\displaystyle\geq\liminf_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{E}\left[\exp\left(\varepsilon^{-2}\left\{\phi(Z^{\varepsilon})+\psi^{\varepsilon}(Z^{\varepsilon})\right\}\right)1_{U}(Z^{\varepsilon})\right]
ϕ(x0)2δ+lim infε0ε2log[ZεU]\displaystyle\geq\phi(x_{0})-2\delta+\liminf_{\varepsilon\to 0}\varepsilon^{2}\log\mathbb{P}\left[Z^{\varepsilon}\in U\right]
ϕ(x0)2δinfxUJ(x)ϕ(x0)2δJ(x0)supx𝒳[ϕ(x)J(x)]3δ.\displaystyle\geq\phi(x_{0})-2\delta-\inf_{x\in U}J(x)\geq\phi(x_{0})-2\delta-J(x_{0})\geq\sup_{x\in{\mathcal{X}}}\left[\phi(x)-J(x)\right]-3\delta.

Sending δ0\delta\to 0 we have the lower bound and combining it with the upper bound in (5.27), we have the result. ∎

Recall the definition of Δ1n\Delta^{n}_{1} from (5.7). The following lemma will allow us to apply Lemma 5.6.

Lemma 5.7.

For {\mathbb{P}}^{*} a.e. ω\omega^{*} and every δ(0,)\delta\in(0,\infty) and η𝒞d\eta\in{\mathcal{C}}_{d} there exist n0n_{0}\in{\mathbb{N}} and δ1(0,)\delta_{1}\in(0,\infty) such that

|Δ1n(η~)|<δ whenever η~𝒞d,ηη~δ1 and nn0.|\Delta^{n}_{1}(\tilde{\eta})|<\delta\mbox{ whenever }\tilde{\eta}\in{\mathcal{C}}_{d},\;\|\eta-\tilde{\eta}\|_{*}\leq\delta_{1}\mbox{ and }n\geq n_{0}.
Proof.

Consider ω\omega^{*} in the set of full measure on which the convergence in (5.4) (and thus in (5.8)) holds. From (5.8), for any fixed δ(0,)\delta\in(0,\infty) and η𝒞d\eta\in{\mathcal{C}}_{d}, we can find n0n_{0}\in{\mathbb{N}} such that for all nn0n\geq n_{0}

(5.28) |Δ1n(η,ω)|δ2.|\Delta^{n}_{1}(\eta,\omega^{*})|\leq\frac{\delta}{2}.

Also, from continuity of hh, we can find a δ1(0,)\delta_{1}\in(0,\infty) such that for all η~𝒞d\tilde{\eta}\in{\mathcal{C}}_{d} with ηη~δ1\|\eta-\tilde{\eta}\|_{*}\leq\delta_{1}

supn0Th(η(s))h(η~(s))(h(ξn(s))+h(ξ0(s)))𝑑sδ4\sup_{n\in{\mathbb{N}}}\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|(\|h(\xi^{n}(s))\|+\|h(\xi_{0}(s))\|)ds\leq\frac{\delta}{4}

and

supn0Th(η(s))h(η~(s))(vn(s)+v(s))𝑑sδ4.\sup_{n\in{\mathbb{N}}}\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|(\|v^{n}(s)\|+\|v(s)\|)ds\leq\frac{\delta}{4}.

Thus for all nn0n\geq n_{0} and η~𝒞d\tilde{\eta}\in{\mathcal{C}}_{d} with ηη~δ1\|\eta-\tilde{\eta}\|_{*}\leq\delta_{1}

|Δ1n(η~)|\displaystyle|\Delta^{n}_{1}(\tilde{\eta})| |Δ1n(η~)Δ1n(η)|+|Δ1n(η)|\displaystyle\leq|\Delta^{n}_{1}(\tilde{\eta})-\Delta^{n}_{1}(\eta)|+|\Delta^{n}_{1}(\eta)|
0Th(η(s))h(η~(s))(h(ξn(s))+h(ξ0(s)))𝑑s\displaystyle\leq\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|(\|h(\xi^{n}(s))\|+\|h(\xi_{0}(s))\|)ds
+0Th(η(s))h(η~(s))(vn(s)+v(s))𝑑s+δ2δ.\displaystyle\quad+\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|(\|v^{n}(s)\|+\|v(s)\|)ds+\frac{\delta}{2}\leq\delta.

We now complete the proof of (5.11).

Completing the proof of (5.11). Note that, from Proposition 5.5, {\mathbb{P}}^{*} a.s.,

lim supnΥ¯1n[ϕ]\displaystyle\limsup_{n\to\infty}-\bar{\Upsilon}_{1}^{n}[\phi] =lim supnεn2logΩxeεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle=\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
(5.29) lim supnεn2logΩxeεn2Gn(ω~,ω)μεn(dω~).\displaystyle\leq\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

For Q(0,)Q\in(0,\infty), let Δn,Q(ΔnQ)(Q)\Delta^{n,Q}\doteq(\Delta^{n}\wedge Q)\vee(-Q). Then

Ωxeεn2Gn(ω~,ω)μεn(dω~)\displaystyle\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega}) Ωxeεn2Gn(ω~,ω)1{|Δn|Q}μεn(dω~)\displaystyle\leq\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}1_{\{|\Delta^{n}|\geq Q\}}\mu^{\varepsilon_{n}}(d\tilde{\omega})
(5.30) +Ωxeεn2(ϕ(ξ^(ω~))H(ξ^(ω~),ξ0(ω),v(ω))+Δn,Q(ω,ω))μεn(dω~).\displaystyle\quad+\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}\left(-\phi(\hat{\xi}(\tilde{\omega}))-H(\hat{\xi}(\tilde{\omega}),\xi_{0}(\omega^{*}),v(\omega^{*}))+\Delta^{n,Q}(\omega,\omega^{*})\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

Note that ϕ\phi is a continuous and bounded map on 𝒞d{\mathcal{C}}_{d}, ηH(η,ξ0,v)\eta\mapsto H(\eta,\xi_{0},v) is a continuous, nonnegative map on 𝒞d{\mathcal{C}}_{d} and ηΔ1n(η,ω)Q(Q)\eta\mapsto\Delta^{n}_{1}(\eta,\omega^{*})\wedge Q\vee(-Q) is a map uniformly bounded in nn which satisfies the properties in Lemma 5.7. Thus applying Lemma 5.6 and the large deviations result from (2.11), we have

(5.31) lim supnεn2logΩxeεn2(ϕ(ξ^(ω~))H(ξ^(ω~),ξ0(ω),v(ω))+Δn,Q(ω,ω))μεn(dω~)\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}\left(-\phi(\hat{\xi}(\tilde{\omega}))-H(\hat{\xi}(\tilde{\omega}),\xi_{0}(\omega^{*}),v(\omega^{*}))+\Delta^{n,Q}(\omega,\omega^{*})\right)}\mu^{\varepsilon_{n}}(d\tilde{\omega})
infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)].\displaystyle\leq-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right].

Next, using the linear growth property of hh

supn|Δ1n(η)|cΔ(ω)(1+η), a.s. \sup_{n}|\Delta^{n}_{1}(\eta)|\leq c_{\Delta}(\omega^{*})(1+\|\eta\|_{*}),\;{\mathbb{P}}^{*}\mbox{ a.s. }

for some measurable map cΔ:Ω+c_{\Delta}:\mathnormal{\Omega}^{*}\to{\mathbb{R}}_{+}. Thus, using the boundedness of ϕ\phi and the nonnegativity of HH, we have

lim supQlim supnεn2logΩxeεn2Gn(ω~,ω)1{|Δn|Q}μεn(dω~)\displaystyle\limsup_{Q\to\infty}\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})}1_{\{|\Delta^{n}|\geq Q\}}\mu^{\varepsilon_{n}}(d\tilde{\omega})
lim supQlim supnεn2logΩxeεn2(cΔ+ϕ)(1+ξ^(ω~))1{cΔ(1+ξ^(ω~))Q}μεn(dω~)=\displaystyle\leq\limsup_{Q\to\infty}\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}(c_{\Delta}+\|\phi\|_{\infty})(1+\|\hat{\xi}(\tilde{\omega})\|_{*})}1_{\{c_{\Delta}(1+\|\hat{\xi}(\tilde{\omega})\|_{*})\geq Q\}}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty

where the last equality follows from Lemma 5.2 (see (5.14)). Using the last bound together with (5.31) in (5.30) and (5.29) we now have the inequality in (5.11). ∎

5.2. Proof of (5.10)

Recall the convergence from (5.4). We begin with the following lemma.

Lemma 5.8.

For {\mathbb{P}}^{*} a.e. ω\omega^{*}

lim infnεn2logΩxeεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))με(dω~)\displaystyle\liminf_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon}(d\tilde{\omega}) infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)].\displaystyle\geq-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right].
Proof.

Fix η0𝒞d\eta_{0}\in{\mathcal{C}}_{d} and δ(0,)\delta\in(0,\infty). From continuity of ϕ\phi on 𝒞d{\mathcal{C}}_{d}, of ATA_{T} on 𝒞d×𝒞m{\mathcal{C}}_{d}\times{\mathcal{C}}_{m}, and of ηH(η,ξ0(ω),v(ω))\eta\mapsto H(\eta,\xi_{0}(\omega^{*}),v(\omega^{*})) (for {\mathbb{P}}^{*} a.e. ω\omega^{*}) on 𝒞d{\mathcal{C}}_{d}, a.s. convergence of βn\beta^{n} to β\beta, and Lemma 5.7, we can find, for {\mathbb{P}}^{*} a.e. ω\omega^{*}, a neighbourhood GG of η0\eta_{0} and n1n_{1}\in{\mathbb{N}} such that

infη~GAT(η~,βn(ω))AT(η0,βn(ω))δ, for all nn1,\displaystyle\inf_{\tilde{\eta}\in G}A_{T}(\tilde{\eta},\beta^{n}(\omega^{*}))\geq A_{T}(\eta_{0},\beta^{n}(\omega^{*}))-\delta,\text{ for all }n\geq n_{1},
infη~G[ϕ(η~)H(η~,ξ0(ω),v(ω)][ϕ(η0)H(η0,ξ0(ω),v(ω)]δ,\displaystyle\quad\inf_{\tilde{\eta}\in G}[-\phi(\tilde{\eta})-H(\tilde{\eta},\xi_{0}(\omega^{*}),v(\omega^{*})]\geq[-\phi(\eta_{0})-H(\eta_{0},\xi_{0}(\omega^{*}),v(\omega^{*})]-\delta,
supη~G|Δ1n(η~)|<δ for all nn1.\displaystyle\quad\sup_{\tilde{\eta}\in G}|\Delta^{n}_{1}(\tilde{\eta})|<\delta\mbox{ for all }n\geq n_{1}.

Observe that

Ωxeεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))με(dω~)\displaystyle\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon}(d\tilde{\omega})
Ωxeεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))1{ξ^(ω~)G}με(dω~)\displaystyle\geq\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}1_{\{\hat{\xi}(\tilde{\omega})\in G\}}\mu^{\varepsilon}(d\tilde{\omega})
eεn2[ϕ(η0)H(η0,ξ0(ω),v(ω)2δ]+εn1(AT(η0,βn(ω))δ)με(G),\displaystyle\geq e^{\varepsilon_{n}^{-2}[-\phi(\eta_{0})-H(\eta_{0},\xi_{0}(\omega^{*}),v(\omega^{*})-2\delta]+\varepsilon_{n}^{-1}(A_{T}(\eta_{0},\beta^{n}(\omega^{*}))-\delta)}\mu^{\varepsilon}(G),

Noting that supn|AT(η0,βn(ω))|<\sup_{n}|A_{T}(\eta_{0},\beta^{n}(\omega^{*}))|<\infty {\mathbb{P}}^{*} a.s. and applying the large deviation result from (2.11), we now have

lim infnεn2logΩxeεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))με(dω~)\displaystyle\liminf_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon}(d\tilde{\omega})
[ϕ(η0)H(η0,ξ0(ω),v(ω)2δ]infη~GJ(η~)\displaystyle\geq[-\phi(\eta_{0})-H(\eta_{0},\xi_{0}(\omega^{*}),v(\omega^{*})-2\delta]-\inf_{\tilde{\eta}\in G}J(\tilde{\eta})
ϕ(η0)H(η0,ξ0(ω),v(ω)J(η0)2δ.\displaystyle\geq-\phi(\eta_{0})-H(\eta_{0},\xi_{0}(\omega^{*}),v(\omega^{*})-J(\eta_{0})-2\delta.

Since δ(0,)\delta\in(0,\infty) and η0𝒞d\eta_{0}\in{\mathcal{C}}_{d} are arbitrary, the result follows. ∎

We now complete the proof of (5.10).

Completing the proof of (5.10). Fix δ(0,)\delta\in(0,\infty). Then

Ωxeεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+εn1F(ω~,βn(ω))μεn(dω~)\displaystyle\geq\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\geq-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}F(\tilde{\omega},\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
{εnKT(ξ^(ω~),βn(ω))δ}eεn2Gn(ω~,ω)+εn1(δεn1+AT(ξ^(ω~),βn(ω))μεn(dω~)\displaystyle\geq\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))\geq-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}(-\delta\varepsilon_{n}^{-1}+A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})
=Ωxeεn2(Gn(ω~,ω)δ)+εn1AT(ξ^(ω~),βn(ω)μεn(dω~)\displaystyle=\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}(G^{n}(\tilde{\omega},\omega^{*})-\delta)+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega})
{εnKT(ξ^(ω~),βn(ω))<δ}eεn2(Gn(ω~,ω)δ)+εn1AT(ξ^(ω~),βn(ω)μεn(dω~).\displaystyle\quad-\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}(G^{n}(\tilde{\omega},\omega^{*})-\delta)+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega}).

From Proposition 5.4 (see (5.17))

lim supnεn2log{εnKT(ξ^(ω~),βn(ω))<δ}eεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω))μεn(dω~)=.\displaystyle\limsup_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\{\varepsilon_{n}K_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))<-\delta\}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*}))}\mu^{\varepsilon_{n}}(d\tilde{\omega})=-\infty.

Thus to prove (5.10) it suffice to show that, {\mathbb{P}}^{*} a.s.,

(5.32) lim infnεn2logΩxeεn2Gn(ω~,ω)+εn1AT(ξ^(ω~),βn(ω)μεn(dω~)infη𝒞d[H(η,ξ0,v)+ϕ(η)+J(η)].\liminf_{n\to\infty}\varepsilon_{n}^{2}\log\int_{\Omega_{x}}e^{\varepsilon_{n}^{-2}G^{n}(\tilde{\omega},\omega^{*})+\varepsilon_{n}^{-1}A_{T}(\hat{\xi}(\tilde{\omega}),\beta^{n}(\omega^{*})}\mu^{\varepsilon_{n}}(d\tilde{\omega})\geq-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0},v)+\phi(\eta)+J(\eta)\right].

However the above is an immediate consequence of Lemma 5.8. This completes the proof of (5.10). ∎

Finally we complete the proof of Lemma 4.1.

Completing the proof of Lemma 4.1.

As noted above (5.6), in order to prove Lemma 4.1 it suffices to show (5.6) for every ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}). Also, for this it is enough to show (5.11) and (5.10). The inequality in (5.11) was shown in Section 5.1 and the proof of the inequality in (5.10) was provided in Section 5.2. Combining these we have Lemma 4.1. ∎

6. Proof of Theorem 2.1.

In order to prove the theorem it suffices to show (3.1) and (3.2). Proof of (3.2) is given in Section 6.1 while the proof of (3.1) is provided in Section 6.2.

6.1. Proof of (3.2).

Let {εn}n\{\varepsilon_{n}\}_{n\in{\mathbb{N}}} be a sequence of positive reals such that εn0\varepsilon_{n}\to 0 as nn\to\infty. To show (3.2) it suffices to show that for every GCb()G\in C_{b}({\mathbb{R}})

(6.1) lim infnεn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]infz{G(z)+Iϕ(z)},\liminf_{n\to\infty}-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]\geq\inf_{z\in{\mathbb{R}}}\{G(z)+I^{\phi}(z)\},
(6.2) lim supnεn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]infz{G(z)+Iϕ(z)}.\limsup_{n\to\infty}-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]\leq\inf_{z\in{\mathbb{R}}}\{G(z)+I^{\phi}(z)\}.

We begin with the proof of (6.1). Fix δ(0,1)\delta\in(0,1) and using (3.6) choose (u~n,v~n)𝒜k×𝒜m(\tilde{u}_{n},\tilde{v}_{n})\in{\mathcal{A}}^{k}\times{\mathcal{A}}^{m} such that

(6.3) εn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]\displaystyle-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]
𝔼εn[G(V¯εn,u~n,v~n[ϕ])+120T(u~n(s)2+v~n(s)2)𝑑s]δ.\displaystyle\geq{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[G(\bar{V}^{\varepsilon_{n},\tilde{u}_{n},\tilde{v}_{n}}[\phi])+\frac{1}{2}\int_{0}^{T}(\|\tilde{u}_{n}(s)\|^{2}+\|\tilde{v}_{n}(s)\|^{2})ds\right]-\delta.

Note that

(6.4) supn𝔼εn[0T(u~n(s)2+v~n(s)2)𝑑s]2(2G+1)cG.\sup_{n\in{\mathbb{N}}}{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\int_{0}^{T}(\|\tilde{u}_{n}(s)\|^{2}+\|\tilde{v}_{n}(s)\|^{2})ds\right]\leq 2(2\|G\|_{\infty}+1)\doteq c_{G}.

We now use a standard localization argument (cf. [6, Theorem 3.17]). For M(0,)M\in(0,\infty) let

τMninf{t0:0t(u~n(s)2+v~n(s)2)𝑑sM}\tau^{n}_{M}\doteq\inf\{t\geq 0:\int_{0}^{t}(\|\tilde{u}_{n}(s)\|^{2}+\|\tilde{v}_{n}(s)\|^{2})ds\geq M\}

and define

u~n,M(s)u~n(s)1[0,τMn](s),v~n,M(s)v~n(s)1[0,τMn](s),s[0,T].\tilde{u}_{n,M}(s)\doteq\tilde{u}_{n}(s)1_{[0,\tau^{n}_{M}]}(s),\;\;\tilde{v}_{n,M}(s)\doteq\tilde{v}_{n}(s)1_{[0,\tau^{n}_{M}]}(s),\;s\in[0,T].

Denoting the expectation on the right side of (6.3) by RnR^{n} and denoting the corresponding expectation, with (u~n,v~n)(\tilde{u}_{n},\tilde{v}_{n}) replaced by (u~n,M,v~n,M)(\tilde{u}_{n,M},\tilde{v}_{n,M}), as Rn,MR^{n,M} we see that

RnRn,M\displaystyle R^{n}-R^{n,M} Gεn(τMnT)\displaystyle\geq-\|G\|_{\infty}{\mathbb{P}}^{\varepsilon_{n}}(\tau^{n}_{M}\leq T)
=Gεn(0T(u~n(s)2+v~n(s)2)𝑑sM)GcGM,\displaystyle=-\|G\|_{\infty}{\mathbb{P}}^{\varepsilon_{n}}(\int_{0}^{T}(\|\tilde{u}_{n}(s)\|^{2}+\|\tilde{v}_{n}(s)\|^{2})ds\geq M)\geq-\|G\|_{\infty}\frac{c_{G}}{M},

where the last inequality uses (6.4). Now choose MM such that GcGMδ\|G\|_{\infty}\frac{c_{G}}{M}\leq\delta and denote u~n,M=un\tilde{u}_{n,M}=u_{n}, v~n,M=vn\tilde{v}_{n,M}=v_{n}. Then

(6.5) εn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]\displaystyle-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]
𝔼εn[G(V¯εn,un,vn[ϕ])+120T(un(s)2+vn(s)2)𝑑s]2δ.\displaystyle\geq{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[G(\bar{V}^{\varepsilon_{n},u_{n},v_{n}}[\phi])+\frac{1}{2}\int_{0}^{T}(\|u_{n}(s)\|^{2}+\|v_{n}(s)\|^{2})ds\right]-2\delta.

Note that {(un,vn)}\{(u_{n},v_{n})\} is a sequence of SMS_{M} valued random variable and since SMS_{M} is weakly compact, every subsequence of {(un,vn)}\{(u_{n},v_{n})\} has a weakly convergent subsubsequence. It suffices to show (6.1) along such a subsubsequence which we denote once more as {n}\{n\}. Denoting the limit as (u,v)(u,v), given on some probability space (Ω0,0,0)(\mathnormal{\Omega}^{0},{\mathcal{F}}^{0},{\mathbb{P}}^{0}), we have from Lemma 4.1 that, as nn\to\infty, V¯εn,un,vn[ϕ]V0u,v[ϕ]\bar{V}^{\varepsilon_{n},u_{n},v_{n}}[\phi]\to V_{0}^{u,v}[\phi], in distribution. Using the fact that GCb()G\in C_{b}({\mathbb{R}}) and Fatou’s lemma, we now have

lim infn𝔼εn[G(V¯εn,un,vn[ϕ])+120T(un(s)2+vn(s)2)𝑑s]\displaystyle\liminf_{n\to\infty}{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[G(\bar{V}^{\varepsilon_{n},u_{n},v_{n}}[\phi])+\frac{1}{2}\int_{0}^{T}(\|u_{n}(s)\|^{2}+\|v_{n}(s)\|^{2})ds\right]
𝔼0[G(V0u,v[ϕ])+120T(u(s)2+v(s)2)𝑑s]\displaystyle\geq{\mathbb{E}}_{{\mathbb{P}}^{0}}\left[G(V^{u,v}_{0}[\phi])+\frac{1}{2}\int_{0}^{T}(\|u(s)\|^{2}+\|v(s)\|^{2})ds\right]
𝔼0[G(V0u,v[ϕ])+Iϕ(V0u,v[ϕ])]infz[G(z)+Iϕ(z)],\displaystyle\geq{\mathbb{E}}_{{\mathbb{P}}^{0}}\left[G(V^{u,v}_{0}[\phi])+I^{\phi}(V^{u,v}_{0}[\phi])\right]\geq\inf_{z\in{\mathbb{R}}}[G(z)+I^{\phi}(z)],

where the second inequality uses the fact that, by definition (u,v)𝒮(V0u,v[ϕ])(u,v)\in{\mathcal{S}}(V^{u,v}_{0}[\phi]) a.s. Combining the above display with (6.5) and recalling that δ>0\delta>0 is arbitrary, we have (6.1).

We now give the proof of (6.2). Fix δ(0,1)\delta\in(0,1) and let zz^{*}\in{\mathbb{R}} be such that

(6.6) G(z)+Iϕ(z)infz[G(z)+Iϕ(z)]+δ.G(z^{*})+I^{\phi}(z^{*})\leq\inf_{z\in{\mathbb{R}}}[G(z)+I^{\phi}(z)]+\delta.

Now choose (φ,ψ)𝒮(z)(\varphi,\psi)\in{\mathcal{S}}(z^{*}) such that

(6.7) 120Tφ(t)2𝑑t+120Tψ(t)2𝑑tIϕ(z)+δ.\frac{1}{2}\int_{0}^{T}\|\varphi(t)\|^{2}dt+\frac{1}{2}\int_{0}^{T}\|\psi(t)\|^{2}dt\leq I^{\phi}(z^{*})+\delta.

Since (φ,ψ)𝒜k×𝒜m(\varphi,\psi)\in{\mathcal{A}}_{k}\times{\mathcal{A}}_{m} (as they are non-random and square-integrable), we have from (3.6) that, for every nn\in{\mathbb{N}},

(6.8) εn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]\displaystyle-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]
𝔼εn[G(V¯εn,φ,ψ[ϕ])+120T(φ(s)2+ψ(s)2)𝑑s].\displaystyle\leq{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[G(\bar{V}^{\varepsilon_{n},\varphi,\psi}[\phi])+\frac{1}{2}\int_{0}^{T}(\|\varphi(s)\|^{2}+\|\psi(s)\|^{2})ds\right].

Also, from Lemma 4.1, as nn\to\infty, V¯εn,φ,ψ[ϕ]V0ϕ,ψ[ϕ]\bar{V}^{\varepsilon_{n},\varphi,\psi}[\phi]\to V_{0}^{\phi,\psi}[\phi], in distribution. Since (φ,ψ)𝒮(z)(\varphi,\psi)\in{\mathcal{S}}(z^{*}), (2.17) holds with zz replaced with zz^{*} and so V0ϕ,ψ[ϕ]=zV_{0}^{\phi,\psi}[\phi]=z^{*}. Thus sending nn\to\infty in (6.8), we have

lim supnεn2log𝔼εn[exp{εn2G(Vεn[ϕ])}]\displaystyle\limsup_{n\to\infty}-\varepsilon_{n}^{2}\log{\mathbb{E}}_{{\mathbb{P}}^{\varepsilon_{n}}}\left[\exp\left\{-\varepsilon_{n}^{-2}G(V^{\varepsilon_{n}}[\phi])\right\}\right]
G(z)+120T(φ(s)2+ψ(s)2)𝑑sG(z)+Iϕ(z)+δinfz[G(z)+Iϕ(z)]+2δ,\displaystyle\leq G(z^{*})+\frac{1}{2}\int_{0}^{T}(\|\varphi(s)\|^{2}+\|\psi(s)\|^{2})ds\leq G(z^{*})+I^{\phi}(z^{*})+\delta\leq\inf_{z\in{\mathbb{R}}}[G(z)+I^{\phi}(z)]+2\delta,

where the second inequality uses (6.7) while the third uses (6.6). Since δ>0\delta>0 is arbitrary, we have (6.2), and, together with (6.1), completes the proof of (3.2). ∎

6.2. Proof of (3.1).

Fix ϕCb(𝒞d)\phi\in C_{b}({\mathcal{C}}_{d}) and M(0,)M\in(0,\infty). Consider the set {z:Iϕ(z)M}EM\{z\in{\mathbb{R}}:I^{\phi}(z)\leq M\}\doteq E_{M} and let {zn}n\{z_{n}\}_{n\in{\mathbb{N}}} be a sequence in this set. Since for each nn\in{\mathbb{N}}, Iϕ(zn)MI^{\phi}(z_{n})\leq M, we can find (φn,ψn)𝒮(zn)k2×m2(\varphi_{n},\psi_{n})\in{\mathcal{S}}(z_{n})\subset{\mathcal{L}}^{2}_{k}\times{\mathcal{L}}^{2}_{m} such that

(6.9) 120T(φn(s)2+ψn(s)2)𝑑sM+1n.\frac{1}{2}\int_{0}^{T}(\|\varphi_{n}(s)\|^{2}+\|\psi_{n}(s)\|^{2})ds\leq M+\frac{1}{n}.

Since (φn,ψn)𝒮(zn)(\varphi_{n},\psi_{n})\in{\mathcal{S}}(z_{n}),

(6.10) zn=V0φn,ψn[ϕ]=infη𝒞d[H(η,ξ0φn,ψn)+ϕ(η)+J(η)]infη𝒞d[H(η,ξ0φn,ψn)+J(η)].z_{n}=V_{0}^{\varphi_{n},\psi_{n}}[\phi]=\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+J(\eta)\right].

Note that, we can write

H(η,ξ0φn,ψn)\displaystyle H(\eta,\xi_{0}^{\varphi_{n}},\psi_{n}) =120Th(η(s))h(ξ0φn(s))ψn(s)2𝑑s\displaystyle=\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))-\psi_{n}(s)\|^{2}ds
=120Th(η(s))h(ξ0φn(s))2𝑑s0T[h(η(s))h(ξ0φn(s))]ψn(s)𝑑s\displaystyle=\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}ds-\int_{0}^{T}[h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))]\cdot\psi_{n}(s)ds
+120Tψn(s)2𝑑s\displaystyle\quad+\frac{1}{2}\int_{0}^{T}\|\psi_{n}(s)\|^{2}ds
=H~(η,ξ0φn,ψn)+120Tψn(s)2𝑑s,\displaystyle=\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+\frac{1}{2}\int_{0}^{T}\|\psi_{n}(s)\|^{2}ds,

where for η,η~𝒞d\eta,\tilde{\eta}\in{\mathcal{C}}_{d} and ψm2\psi\in{\mathcal{L}}^{2}_{m}

H~(η,η~,ψ)120Th(η(s))h(η~(s))2𝑑s0T[h(η(s))h(η~(s))]ψ(s)𝑑s.\tilde{H}(\eta,\tilde{\eta},\psi)\doteq\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\tilde{\eta}(s))\|^{2}ds-\int_{0}^{T}[h(\eta(s))-h(\tilde{\eta}(s))]\cdot\psi(s)ds.

From (6.10) and relation between HH and H~\tilde{H} it follows that

(6.11) zninfη𝒞d[H~(η,ξ0φn,ψn)+ϕ(η)+J(η)]infη𝒞d[H~(η,ξ0φn,ψn)+J(η)].z_{n}\doteq\inf_{\eta\in{\mathcal{C}}_{d}}\left[\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+J(\eta)\right].

Also, from (6.9) it follows that {(φn,ψn)}nS2(M+1)\{(\varphi_{n},\psi_{n})\}_{n\in{\mathbb{N}}}\subset S_{2(M+1)}. Since S2(M+1)S_{2(M+1)} is compact, we can find a subsequence along which (φn,ψn)(\varphi_{n},\psi_{n}) converges to some (φ,ψ)S2(M+1)(\varphi,\psi)\in S_{2(M+1)}. In fact, from (6.9) and lower semicontinuity it follows that (φ,ψ)S2M(\varphi,\psi)\in S_{2M}. Define

(6.12) zV0φ,ψ[ϕ]\displaystyle z^{*}\doteq V_{0}^{\varphi,\psi}[\phi] =infη𝒞d[H(η,ξ0φ,ψ)+ϕ(η)+J(η)]infη𝒞d[H(η,ξ0φ,ψ)+J(η)]\displaystyle=\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[H(\eta,\xi_{0}^{\varphi},\psi)+J(\eta)\right]
=infη𝒞d[H~(η,ξ0φ,ψ)+ϕ(η)+J(η)]infη𝒞d[H~(η,ξ0φ,ψ)+J(η)].\displaystyle=\inf_{\eta\in{\mathcal{C}}_{d}}\left[\tilde{H}(\eta,\xi_{0}^{\varphi},\psi)+\phi(\eta)+J(\eta)\right]-\inf_{\eta\in{\mathcal{C}}_{d}}\left[\tilde{H}(\eta,\xi_{0}^{\varphi},\psi)+J(\eta)\right].

In order to complete the proof of (3.1) it suffices to show that

(6.13)  as n,znz.\mbox{ as }n\to\infty,\;\;z_{n}\to z^{*}.

We first argue that in the infimum appearing in (the second line of )(6.12) and (6.11), {η𝒞d}\{\eta\in{\mathcal{C}}_{d}\} can be replaced by {ηK}\{\eta\in K\} for some fixed compact set KK. To see this, note that, with ξ\xi^{*} as in (1.4),

infη𝒞d[H~(η,ξ0φn,ψn)+ϕ(η)+J(η)]\displaystyle\inf_{\eta\in{\mathcal{C}}_{d}}\left[\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+\phi(\eta)+J(\eta)\right] H~(ξ,ξ0φn,ψn)+ϕ+J(ξ).\displaystyle\leq\tilde{H}(\xi^{*},\xi_{0}^{\varphi_{n}},\psi_{n})+\|\phi\|_{\infty}+J(\xi^{*}).

Also, note that J(ξ)=0J(\xi^{*})=0 and

H~(ξ,ξ0φn,ψn)\displaystyle\tilde{H}(\xi^{*},\xi_{0}^{\varphi_{n}},\psi_{n}) =120Th(ξ(s))h(ξ0φn(s))2𝑑s0T[h(ξ(s))h(ξ0φn(s))]ψn(s)𝑑s\displaystyle=\frac{1}{2}\int_{0}^{T}\|h(\xi^{*}(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}ds-\int_{0}^{T}[h(\xi^{*}(s))-h(\xi_{0}^{\varphi_{n}}(s))]\cdot\psi_{n}(s)ds
0Th(ξ(s))h(ξ0φn(s))2+120Tψn(s)2𝑑s\displaystyle\leq\int_{0}^{T}\|h(\xi^{*}(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}+\frac{1}{2}\int_{0}^{T}\|\psi_{n}(s)\|^{2}ds
2Th(ξ())2+20Th(ξ0φn(s))2𝑑s+120Tψn(s)2𝑑s\displaystyle\leq 2T\|h(\xi^{*}(\cdot))\|_{*}^{2}+2\int_{0}^{T}\|h(\xi_{0}^{\varphi_{n}}(s))\|^{2}ds+\frac{1}{2}\int_{0}^{T}\|\psi_{n}(s)\|^{2}ds
2Th(ξ())2+κ1(M+1)κ2,\displaystyle\leq 2T\|h(\xi^{*}(\cdot))\|_{*}^{2}+\kappa_{1}(M+1)\doteq\kappa_{2},

where κ1(0,)\kappa_{1}\in(0,\infty) depends only on x0,Tx_{0},T and the linear growth coefficients of h,b,σh,b,\sigma. Thus, taking κ3κ2+ϕ+1\kappa_{3}\doteq\kappa_{2}+\|\phi\|_{\infty}+1, we see that the first infimum in (6.11) can be replaced by infimum over the set

K0n{η𝒞d:H~(η,ξ0φn,ψn)+ϕ(η)+J(η)κ3}.K_{0}^{n}\doteq\{\eta\in{\mathcal{C}}_{d}:\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})+\phi(\eta)+J(\eta)\leq\kappa_{3}\}.

Using the relation ab14a2b2a\cdot b\geq-\frac{1}{4}\|a\|^{2}-\|b\|^{2},

H~(η,ξ0φn,ψn)\displaystyle\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n}) 120Th(η(s))h(ξ0φn(s))2\displaystyle\geq\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}
140Th(η(s))h(ξ0φn(s))20Tψn(s)2𝑑s2M.\displaystyle\quad-\frac{1}{4}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}-\int_{0}^{T}\|\psi_{n}(s)\|^{2}ds\geq-2M.

Thus, with κ4κ3+ϕ+1+2M\kappa_{4}\doteq\kappa_{3}+\|\phi\|_{\infty}+1+2M, K0nK_{0}^{n} is contained in the compact set

K{η𝒞d:J(η)κ4}.K\doteq\{\eta\in{\mathcal{C}}_{d}:J(\eta)\leq\kappa_{4}\}.

Thus the first infimum in (6.11) can be replaced by infimum over the set KK. Similarly, the second infimum in (6.11) and both infima in (second line of) (6.12) can be replaced by infima over the same compact set KK. Note that if Bn,BB_{n},B are maps from KK\to{\mathbb{R}} such that BnBB_{n}\to B uniformly on compacts, then

infηK[Bn(η)+J(η)]infηK[B(η)+J(η)].\inf_{\eta\in K}[B_{n}(\eta)+J(\eta)]\to\inf_{\eta\in K}[B(\eta)+J(\eta)].

Thus, to complete the proof of (6.13) it suffices to show that,

(6.14)  as n,H~(η,ξ0φn,ψn)H~(η,ξ0φ,ψ), uniformly for ηK.\mbox{ as }n\to\infty,\;\;\tilde{H}(\eta,\xi_{0}^{\varphi_{n}},\psi_{n})\to\tilde{H}(\eta,\xi_{0}^{\varphi},\psi),\;\mbox{ uniformly for }\eta\in K.

For this note that from Assumption 1 and the convergence of φnφ\varphi_{n}\to\varphi it follows that, ξ0φnξ0φ\xi_{0}^{\varphi_{n}}\to\xi_{0}^{\varphi} in 𝒞d{\mathcal{C}}_{d} as nn\to\infty. Also, since KK is compact, supηKη<\sup_{\eta\in K}\|\eta\|_{*}<\infty. Combining these observations with the continuity and linear growth of hh we have that, as nn\to\infty,

(6.15) 120Th(η(s))h(ξ0φn(s))2𝑑s120Th(η(s))h(ξ0φ(s))2𝑑s\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi_{n}}(s))\|^{2}ds\to\frac{1}{2}\int_{0}^{T}\|h(\eta(s))-h(\xi_{0}^{\varphi}(s))\|^{2}ds

uniformly for ηK\eta\in K. Also, writing

0Th(ξ0φn(s))ψn(s)𝑑s=0T[h(ξ0φn(s))h(ξ0φ(s))]ψn(s)𝑑s+0Th(ξ0φ(s))ψn(s)𝑑s\int_{0}^{T}h(\xi_{0}^{\varphi_{n}}(s))\cdot\psi_{n}(s)ds=\int_{0}^{T}[h(\xi_{0}^{\varphi_{n}}(s))-h(\xi_{0}^{\varphi}(s))]\cdot\psi_{n}(s)ds+\int_{0}^{T}h(\xi_{0}^{\varphi}(s))\cdot\psi_{n}(s)ds

and using the convergence (ξ0φn,ψn)(ξ0φ,ψ)(\xi_{0}^{\varphi_{n}},\psi_{n})\to(\xi_{0}^{\varphi},\psi), the bound in (6.9), and the Lipschitz property of hh, we have that, as nn\to\infty

(6.16) 0Th(ξ0φn(s))ψn(s)𝑑s0Th(ξ0φ(s))ψ(s)𝑑s.\int_{0}^{T}h(\xi_{0}^{\varphi_{n}}(s))\cdot\psi_{n}(s)ds\to\int_{0}^{T}h(\xi_{0}^{\varphi}(s))\cdot\psi(s)ds.

Finally we claim that, as nn\to\infty,

(6.17) 0Th(η(s))ψn(s)𝑑s0Th(η(s))ψ(s)𝑑s,\int_{0}^{T}h(\eta(s))\cdot\psi^{n}(s)ds\to\int_{0}^{T}h(\eta(s))\cdot\psi(s)ds,

uniformly for ηK\eta\in K. Indeed, to show the claim, it suffices to show that if ηnη\eta^{n}\to\eta in KK then

(6.18) 0Th(ηn(s))ψn(s)𝑑s0Th(η(s))ψ(s)𝑑s.\int_{0}^{T}h(\eta^{n}(s))\cdot\psi^{n}(s)ds\to\int_{0}^{T}h(\eta(s))\cdot\psi(s)ds.

Write the right hand side as

0Th(ηn(s))ψn(s)𝑑s=0T(h(ηn(s))h(η(s)))ψn(s)𝑑s+0Th(η(s))ψn(s)𝑑s.\int_{0}^{T}h(\eta^{n}(s))\cdot\psi^{n}(s)ds=\int_{0}^{T}(h(\eta^{n}(s))-h(\eta(s)))\cdot\psi^{n}(s)ds+\int_{0}^{T}h(\eta(s))\cdot\psi^{n}(s)ds.

The convergence in (6.18) is now immediate from the above display on using, the Lipschitz property of hh, the bound in (6.9), and the convergence of (ηn,ψn)(\eta^{n},\psi^{n}) to (η,ψ)(\eta,\psi), which proves the claim. Combining the convergence properties in (6.15), (6.16), and (6.17), we now have the statement in (6.14), which, as noted previously, proves (3.1). ∎

Acknowledgement: The research of AB was supported in part by the NSF (DMS-1814894 and DMS-1853968). Part of this research was carried out when AB was visiting the International Centre for Theoretical Sciences-TIFR and Imperial College, London, and he will like to thank his hosts Amit Apte and Dan Crisan, respectively, at these institutes, for their generous hospitality. AA and ASR acknowledge the support of the Department of Atomic Energy, Government of India, under projects no.12-R&D-TFR-5.10-1100, and no.RTI4001.

References

  • [1] Gérard Ben Arous and Fabienne Castell. Flow decomposition and large deviations. Journal of functional Analysis, 140(1):23–67, 1996.
  • [2] J.S. Baras, A. Bensoussan, and M.R. James. Dynamic observers as asymptotic limits of recursive filters: Special cases. SIAM Journal on Applied Mathematics, 48(5):1147–1158, 1988.
  • [3] Abdelkarem Berkaoui, Boualem Djehiche, and Youssef Ouknine. Sur les grandes déviations en théorie de filtrage non linéaire. Studia Mathematica, 148:5–21, 2001.
  • [4] Michelle Boué and Paul Dupuis. A variational representation for certain functionals of Brownian motion. The Annals of Probability, 26(4):1641–1659, 1998.
  • [5] Amarjit Budhiraja and Paul Dupuis. A variational representation for positive functionals of infinite dimensional Brownian motion. Probability and Mathematical Statistics-Wroclaw University, 20(1):39–61, 2000.
  • [6] Amarjit Budhiraja and Paul Dupuis. Analysis and Approximation of Rare events: Representations and Weak Convergence Methods, volume 94 of Probability Theory and Stochastic Modeling. Springer, 2019.
  • [7] Alberto Carrassi, Marc Bocquet, Laurent Bertino, and Geir Evensen. Data assimilation in the geosciences: An overview of methods, issues, and perspectives. Wiley Interdisciplinary Reviews: Climate Change, 9(5):e535, 2018.
  • [8] J.M.C. Clark. The design of robust approximations to the stochastic differential equations of nonlinear filtering. Communication systems and random process theory, 25:721–734, 1978.
  • [9] Mark H.A. Davis. On a multiplicative functional transformation arising in nonlinear filtering theory. Zeitschrift für Wahrscheinlichkeitstheorie und verwandte Gebiete, 54(2):125–139, 1980.
  • [10] Mark H.A. Davis. A pathwise solution of the equations of nonlinear filtering. Theory of Probability & Its Applications, 27(1):167–175, 1982.
  • [11] H. Doss. Un nouveau principe de grandes déviations en théorie du filtrage non linéaire. Annales de l’IHP Probabilités et Statistiques, 27(3):407–423, 1991.
  • [12] Steven J Fletcher. Data assimilation for the geosciences: From theory to application. Elsevier, 2017.
  • [13] A. J. Heunis. Non-linear filtering of rare events with large signal-to-noise ratio. Journal of Applied Probability, 24(4):929–948, 1987.
  • [14] Omar Hijab. Minimum energy estimation. PhD thesis, University of California, Berkeley, 1980.
  • [15] Omar Hijab. Asymptotic Bayesian estimation of a first order equation with small diffusion. The Annals of Probability, pages 890–902, 1984.
  • [16] M. R. James and J. S. Baras. Nonlinear filtering and large deviations: A PDE-control theoretic approach. Stochastics, 23(3):391–412, 1988.
  • [17] Gopinath Kallianpur. Stochastic Filtering Theory, volume 13. Springer Science & Business Media, 2013.
  • [18] Vasileios Maroulas, Xiaoyang Pan, and Jie Xiong. Large deviations for the optimal filter of nonlinear dynamical systems driven by Lévy noise. Stochastic Processes and their Applications, 130(1):203–231, 2020.
  • [19] Vasileios Maroulas and Jie Xiong. Large deviations for optimal filtering with fractional brownian motion. Stochastic Processes and their Applications, 123(6):2340–2352, 2013.
  • [20] R. Mortensen. Maximum-likelihood recursive nonlinear filtering. Journal of Optimization Theory and Applications, 2:386–394, 1968.
  • [21] Étienne Pardoux and Ofer Zeitouni. Quenched large deviations for one dimensional nonlinear filtering. SIAM journal on control and optimization, 43(4):1272–1297, 2004.
  • [22] Toussaint Joseph Rabeherimanana. Petites perturbations de systemes dynamiques et algebres de Lie Nilpotentes. PhD thesis, 1992.
  • [23] Daniel W Stroock. An Introduction to the Theory of Large Deviations. Springer Science & Business Media, 2012.
  • [24] Jie Xiong. Large deviation principle for optimal filtering. Applied Mathematics & Optimization, 47(2), 2003.