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

Large deviations principle for stationary solutions of stochastic differential equations with multiplicative noise

Peipei Gao [email protected] Yong Liu [email protected] Yue Sun [email protected] Zuohuan Zheng [email protected] Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100190, China School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China LMAM, School of Mathematical Sciences, Peking University, Beijing 100871, China College of Mathematics and Statistics, Hainan Normal University, Haikou, Hainan 571158, China
Abstract

We study the large deviations principle (LDP) for stationary solutions of a class of stochastic differential equations (SDE) in infinite time intervals by the weak convergence approach, and then establish the LDP for the invariant measures of the SDE by the contraction principle. We further point out the equivalence of the rate function of the LDP for invariant measures induced by the LDP for stationary solutions and the rate function defined by quasi-potential. This fact gives another view of the quasi-potential introduced by Freidlin and Wentzell.

keywords:
Large deviations principle; Stationary solutions; Invariant measures; Quasi-potential

1 Introduction

Real world phenomena are often effected by random perturbations or under the influences of noise, random dynamical systems arise to model those problems. The stationary solution, random periodic solution, random quasi-periodic solution (see, for example [22, 19, 27, 12, 10, 13, 11]) are natural extensions correspond to the fixed point, periodic solution, quasi-periodic solution from deterministic dynamical systems to random dynamical systems, which are the fundamental concepts that describe the long time behavior of random dynamical systems. In particular, there are many phenomena with small random perturbations in nature, and small perturbations essentially influence the long time behaviour of the system in general. Following Freidlin and Wentzell’s perspective in [14], we use the LDP to characterize the long time asymptotic behavior of dynamical systems as the small random perturbation converges to zero. Therefore, it is interesting to investigate the LDP for stationary solutions, random periodic solutions and random quasi-periodic solutions in infinite time intervals with small noises.

The LDP for solutions in infinite time intervals is also an interesting problem in mathematics. The LDP for Wiener process in infinite time intervals is shown by Deuschel and Stroock (cf. Schilder’s theorem in Section 1.3 of [8]), the proof depends on the properties of the Gaussian measure. Although we can get the LDP for stationary solutions of Eq. (1.1) with additive noise by using the contraction principle, it is not feasible for multiplicative noise. The exponential tightness for stationary solutions of Eq. (1.1) with multiplicative noise in the infinite intervals plays an important role in the proof of LDP if we follow the approach of Deuschle and Stroock [8], but it is hard to obtain the exponential tightness for the lack of certain accurate estimates, such as the Fernique theorem for Gaussian measure (cf. Theorem 1.3.24 in [8]). Although, the proof of exponential tightness for non-Gaussian measures in the continuous function space is an interesting problem, unfortunately we have not found a suitable way to solve it by now.

Therefore, we use another well-known method, the weak convergence approach to prove the family of stationary solutions of Eq. (1.1) satisfies the LDP as follows. Noticing that the large deviations principle is equivalent to the Laplace principle (LP) in Polish space, it is sufficient to prove the LP by the Boué-Dupuis formula (see, for example, [3, 5, 9, 26]). This method is often referred as the weak convergence approach. Using the weak convergence approach and the Boué-Dupuis formula, the Freidlin-Wentzell type LDP for the family of solutions of SDE or SPDE driven by Wiener process in finite intervals has been re-proved (see, for example, [14, 25]). However, considering the LDP in infinite intervals, we need the Boué-Dupuis formula in infinite intervals. Since the compactness of space and the integrability of bounded functions on infinite intervals are different from those on finite intervals, we have to extend the Boué-Dupuis formula from finite intervals to infinite intervals. As far as we know, the Boué-Dupuis formula in infinite intervals has been used directly without proof in [23] by Barashkov and Gubinelli, we have to present the proof of Boué-Dupuis formula and the weak convergence approach in infinite intervals for the completeness of the present paper in Appendix A. In fact, the most difficult problem in our proof is the well-posedness with respect to the skeleton Eq. (3.1) in infinite time intervals, as we need to prove the uniqueness for the backward infinite horizon integral Eq. (5.1) for the skeleton Eq. (3.1) and construct the solution of integral equation (5.1). Besides, verifying the two conditions (cf. Condition 2.1 in Section 2.2 below) of the weak convergence approach in infinite time intervals is also quite technical.

The main purpose of this paper is to establish the LDP of stationary solutions for a class of stochastic differential equations in infinite time intervals. Moreover, since the one dimensional distribution of the stationary solution always generates an invariant measure of the SDE, here we give another illuminating view of the LDP for invariant measures by the contraction principle (cf. Section 4.2.1 in [7]). In fact, the invariant measure is unable to provide the properties for the trajectories of the dynamical systems, and there exist many systems with different dynamical behaviors but with the same invariant measure (cf. Example 4 below). By comparison, the stationary solution show more accurate information about dynamical systems, and the LDP for them will give more explicit characterization for the long time asymptotic behavior of random dynamical systems with small perturbations.

As far as we know, Freidlin and Wentzell [14], Cerrai and Röckner [6], and Brzeźniak and Cerrai [4] studied the LDP for invariant measures by taking the quasi-potential as the rate function. Moreover, it is not difficult to prove that the rate function (4.3) defined in Theorem 4.2 is equivalent to the rate function defined by quasi-potential, and the rate function (4.3) gives an natural explanation of quasi-potential. Roughly speaking, Freidlin and Wentzell [14], Cerrai and Röckner [6], and Brzeźniak and Cerrai [4] studied the LDP for invariant measures by using the result of the LDP of solutions in finite intervals and let the length of time intervals tends to infinity. It is different to our method that we directly consider the asymptotic behavior of stationary solutions for random dynamical systems in infinite intervals as perturbation converges to zero, and by using the contraction principle to get the LDP for invariant measures. Besides, we believe that our methods to show the LDP for stationary solutions is also available to explore the LDP for random periodic solutions and random quasi-periodic solutions. We give two examples (cf. Example 2 and 3 below) to show the random periodic solutions of Eq. (3.27) (see, for example, Feng, Liu and Zhao [10]) and random quasi-periodic solutions (see, for example, Hopf [15]) of equation (3.31) satisfy the LDP, and we will research the LDP for random periodic solutions and random quasi-periodic solutions for others equation in future papers.

In this paper, we consider the stochastic differential equation as follows

dXε=AXεdt+F(Xε)dt+εB(Xε)dWt.\mathrm{d}X_{\varepsilon}=AX_{\varepsilon}\mathrm{d}t+F(X_{\varepsilon})\mathrm{d}t+\sqrt{\varepsilon}B(X_{\varepsilon})\mathrm{d}W_{t}. (1.1)

All definitions of the symbols involved can be found later in Section 2.1.

Liu and Zhao give the representation of the stationary solution for Burgers equation with large viscosity in [19]. Under our Hypothesis 2.1, similar to the proof of [21] and [19], we have proved that there exists ε0>0\varepsilon_{0}>0 such that for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the stochastic equation (1.1) exists a unique stationary solution Xε(t,ω)X^{*}_{\varepsilon}(t,\omega) satisfies the following equation in HH for any tt\in\mathbb{R}

Xε(t)=tSA(ts)F(Xε)ds+εtSA(ts)B(Xε)dWs,X_{\varepsilon}^{*}(t)=\int_{-\infty}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}\right)\mathrm{d}W_{s}, (1.2)

where HH be a separable Hilbert space, which is defined in Section 2.1.

The first major conclusion as follows.

Theorem 1.

(cf. Theorem 3.2 below) If the operators AA, FF and BB satisfy Hypothesis 2.1, then the family of stationary solutions {Xε:ε>0}\{X^{*}_{\varepsilon}:\varepsilon>0\} satisfies the LDP in C(;H)C(\mathbb{R};H) with good rate function II given in (3.23).

We could get the following result for the family of invariant measures of Eq. (1.1) satisfies the LDP by using the contraction principle .

Theorem 2.

(cf. Theorem 4.2 below) If the operators AA, FF and BB satisfy Hypothesis 2.1, then the family of invariant measures {νε}ε>0\left\{\nu_{\varepsilon}\right\}_{\varepsilon>0} for Eq. (1.1) satisfies the LDP in HH, with good rate function given in (4.3).

Noticing that under Hypothesis 2.1, we actually only deal with the case where Eq. (1.1) has only one asymptotically stable fixed point for ε=0\varepsilon=0. And even the stochastic Burgers equation and 22-dimensional Navier-Stokes on tours are all satisfies Hypothesis 2.1 (cf. Example 1, 5 below), but there are also many types of stochastic partial differential equations that do not satisfy Hypothesis 2.1, such as the stochastic wave equations. Therefore, there are many more complex and interesting cases that we might follow, such as the stochastic wave equations, random dynamical systems with random periodic solutions, random quasi-periodic solutions. In particular, the cases that the deterministic dynamical systems for ε=0\varepsilon=0 have multiple equilibrium points (see, for example, [20]) is also an attractive problem. Furthermore, the existence of invariant measure does not imply the existence of stationary solution, unless we consider in an extended probability space (see, for example, Arnold [1]). The asymptotic behavior for the stationary solutions in the extended probability space is also an interesting problem.

This paper is structured as follows. In Section 2, we present the basic notations and some definitions for the stationary solution, LDP and quasi-potential. We give Hypothesis 2.1 which the operators AA, FF and BB satisfy throughout the paper and the definition for stationary solution in Section 2.1, the preliminary knowledge for LDP in Section 2.2 and the definition of quasi-potential in Section 2.3.

In Section 3, we prove the family of stationary solutions of Eq. (1.1) satisfies LDP. We use the weak convergence approach to prove the LDP for stationary solution of Eq. (1.1), and verify that the Burger’s equation satisfies the Hypothesis 2.1 in Example 1. Moreover, we state two examples to show the LDP for random periodic solutions and random quasi-periodic solutions in Example 2 and Example 3.

In Section 4, we prove the LDP for invariant measures of Eq. (1.1) in Theorem 4.2 and prove the rate function defined in Theorem 4.2 is equivalent to the rate function (2.10) defined by quasi-potential in Lemma 4.1. Moreover, we give two examples to show that the rate function (4.3) defined in Theorem 4.2 of the invariant measures is consistent with quasi-potential (see, for example, [4, 6, 14]), and the LDP of the stationary solutions can provide more dynamic information.

In Section 5, we consider the well-posedness of the skeleton equation in Section 5.1, and prove there exists a ε0\varepsilon_{0} such that for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), Eq. (1.1) exists a unique stationary solution in Section 5.2.

In Appendix A, we prove the Boué-Dupuis formula and the weak convergence approach in infinite intervals.

2 Preliminaries

2.1 Notations

Let VV be a reflexive Banach space, and VV^{*} be the dual space of VV. Let HH be a separable Hilbert space with inner product ,H\langle\cdot,\cdot\rangle_{H}, HH and its dual space HH^{*} are consistent by the Riesz isomorphism. Let VHV\subset H continuously and densely, then (V,H,V)\left(V,H,V^{*}\right) is called a Gelfand triple.

Let L(H)L(H) be the space of all bounded linear operators from HH to HH with the operator norm ||L|\cdot|_{L}. Let L2(H)L_{2}(H) be the space of all Hilbert-Schmidt operators from HH to HH with Hilbert-Schmidt norm |T|L2:=kTakH2|T|_{L_{2}}:=\sum_{k\in\mathbb{N}}\left\|Ta_{k}\right\|_{H}^{2}, where {ak}k\left\{a_{k}\right\}_{k\in\mathbb{N}} is an orthogonal basis of HH. Let QQ be a trace class operator, LQ(H)L_{Q}(H) is denoted by the space of linear operators TT such that TQ1/2TQ^{1/2} is a Hilbert-Schmidt operator from HH to HH, with the norm |T|LQ:=|TQ1/2|L2|T|_{L_{Q}}:=|TQ^{1/2}|_{L_{2}}.

Let H0=Q1/2HH_{0}=Q^{1/2}H. Then H0H_{0} is a Hilbert space with the inner product

u,vH0=Q1/2u,Q1/2vH,u,vH0.\langle u,v\rangle_{H_{0}}=\langle Q^{-1/2}u,Q^{-1/2}v\rangle_{H},\quad\forall u,v\in H_{0}.

Clearly, the imbedding of H0H_{0} in HH is Hilbert-Schmidt for QQ is a trace class operator.

Let {W(t)}t\{W(t)\}_{t\in\mathbb{R}} be a QQ-Wiener process on HH with respect to a probability space (Ω,0,)(\Omega,\mathcal{F}^{0},\mathbb{P}), where 0\mathcal{F}^{0} is Borel σ\sigma-field of Ω\Omega, \mathbb{P} is the Wiener measure on Ω\Omega. Let \mathcal{F} be the \mathbb{P}-completion of 0\mathcal{F}^{0}. Define

stσ{WuWv:su,vt}𝒩,tstst,st,\mathcal{F}_{s}^{t}\equiv\sigma\left\{W_{u}-W_{v}:s\leq u,v\leq t\right\}\vee\mathcal{N},\quad\mathcal{F}_{-\infty}^{t}\equiv\bigvee_{s\leq t}\mathcal{F}_{s}^{t},\quad\forall s\leq t\in\mathbb{R},

where 𝒩\mathcal{N} are the null sets of \mathcal{F} (see, for example, [19]).

We consider the stochastic differential equation

dX=AXdt+F(X)dt+εB(X)dWt,\mathrm{d}X=AX\mathrm{d}t+F(X)\mathrm{d}t+\sqrt{\varepsilon}B(X)\mathrm{d}W_{t}, (2.1)

where A:VVA:V\rightarrow V^{*}, F:VHF:V\rightarrow H, B:VL(H)B:V\rightarrow L(H) be progressively measurable and satisfies the following Hypothesis.

Hypothesis 2.1.

(i). Assume there exist constants λ,C0>0\lambda,C_{0}>0 such that

AuAv+F(u)F(v),uvVVλuvV2+C0uvH2uV2,{}_{V^{*}}\langle A\textbf{u}-A\textbf{v}+F(\textbf{u})-F(\textbf{v}),\textbf{u}-\textbf{v}\rangle_{V}\leq-\lambda\left\|\textbf{u}-\textbf{v}\right\|_{V}^{2}+C_{0}\left\|\textbf{u}-\textbf{v}\right\|_{H}^{2}\left\|\textbf{u}\right\|_{V}^{2},

and

AX+F(u),uVVλuV2,u,vV.{}_{V^{*}}\langle AX+F(\textbf{u}),\textbf{u}\rangle_{V}\leq-\lambda\left\|\textbf{u}\right\|_{V}^{2},\quad\forall\textbf{u},\textbf{v}\in V.

(ii). Let SA(t)S_{A}(t) be the associated semigroup on H corresponding to AA, for any N+N\in\mathbb{N}^{+}, t0,t[N,N]t_{0},t\in[-N,N] and t0<tt_{0}<t, there exists constant C>0C>0 and 1<α<0-1<\alpha<0 such that the semi-group satisfies

t0tSA(s)(F(u)F(v))dsH2C(N)t0tsαuvH2ds.\Big{\|}\int_{t_{0}}^{t}S_{A}(s)(F(\textbf{u})-F(\textbf{v}))\mathrm{d}s\Big{\|}_{H}^{2}\leq C(N)\int_{t_{0}}^{t}s^{\alpha}\|\textbf{u}-\textbf{v}\|_{H}^{2}\mathrm{d}s.

(iii). The function BB will not be zero operator and there exist positive constants β0,D0\beta_{0},D_{0} such that

|B(u)B(v)|Lβ0uvH,u,vV,|B(\textbf{u})-B(\textbf{v})|_{L}\leq\beta_{0}\left\|\textbf{u}-\textbf{v}\right\|_{H},\quad\forall\textbf{u},\textbf{v}\in V,

and

|B(u)|LD0,uV.|B(\textbf{u})|_{L}\leq D_{0},\quad\forall\textbf{u}\in V.
Remark 2.1.

(i). Since VV imbedes to HH, there exists a constant C1C_{1} such that C1uH2uV2,uV.C_{1}\|\textbf{u}\|_{H}^{2}\leq\|\textbf{u}\|_{V}^{2},\forall\textbf{u}\in V.
(ii). Let β=β0trQ\beta=\beta_{0}trQ and D=D0trQD=D_{0}trQ. It follows from QQ is a trace class operator and combining Hypothesis 2.1 (iii), we could obtain that

|B(u)B(v)|LQ=|(B(u)B(v))Q1/2|L2|(B(u)B(v))|L|Q1/2|L2βuvH,\displaystyle|B(\textbf{u})-B(\textbf{v})|_{L_{Q}}=|(B(\textbf{u})-B(\textbf{v}))Q^{1/2}|_{L_{2}}\leq|(B(\textbf{u})-B(\textbf{v}))|_{L}|Q^{1/2}|_{L_{2}}\leq\beta\left\|\textbf{u}-\textbf{v}\right\|_{H},

and

|B(u)|LQ|B(u)Q1/2|L2|B(u)|L|Q1/2|L2D,u,vV.|B(\textbf{u})|_{L_{Q}}\leq|B(\textbf{u})Q^{1/2}|_{L_{2}}\leq|B(\textbf{u})|_{L}|Q^{1/2}|_{L_{2}}\leq D,\quad\forall\textbf{u},\textbf{v}\in V.

(iii). Let u(t):=SA(t)u0\textbf{u}(t):=S_{A}(t)\textbf{u}_{0} be the solution of the equation u˙=Au\dot{\textbf{u}}=A\textbf{u} with initial value u0\textbf{u}_{0}, where u˙\dot{\textbf{u}} be the deririaive of u with respect to time tt, then we have the estimates

12duH2dt=Vu,AuVλuV2,\frac{1}{2}\frac{\mathrm{d}\|\textbf{u}\|_{H}^{2}}{\mathrm{d}t}=_{V}\langle\textbf{u},A\textbf{u}\rangle_{V^{\ast}}\leq-\lambda\|\textbf{u}\|_{V}^{2},

then 12u(t)H2+λ0tuV2ds12u0H2\frac{1}{2}\|\textbf{u}(t)\|_{H}^{2}+\lambda\int_{0}^{t}\|\textbf{u}\|_{V}^{2}\mathrm{d}s\leq\frac{1}{2}\|\textbf{u}_{0}\|_{H}^{2}, thus

SA(t)u0H2u0H2,0SA(s)u0H2ds1C10SA(s)u0V2ds12λC1u0H2.\|S_{A}(t)\textbf{u}_{0}\|_{H}^{2}\leq\|\textbf{u}_{0}\|_{H}^{2},\quad\int_{0}^{\infty}\|S_{A}(s)\textbf{u}_{0}\|_{H}^{2}\mathrm{d}s\leq\frac{1}{C_{1}}\int_{0}^{\infty}\|S_{A}(s)\textbf{u}_{0}\|_{V}^{2}\mathrm{d}s\leq\frac{1}{2\lambda C_{1}}\|\textbf{u}_{0}\|_{H}^{2}.

For any u0H,u0H=1\textbf{u}_{0}\in H,\|\textbf{u}_{0}\|_{H}=1, it implies that

|SA(t)|L1,limt|SA(t)|L=0.|S_{A}(t)|_{L}\leq 1,\quad\lim_{t\rightarrow\infty}|S_{A}(t)|_{L}=0. (2.2)
Definition 2.2.

(Mild solution). For any ε,T>0\varepsilon,T>0, ξH\xi\in H, an HH-valued predictable process Xε(t,ω;ξ)X_{\varepsilon}(t,\omega;\xi), t[0,T]t\in[0,T], is called a mild solution of Eq. (2.1) with initial value ξ\xi if

Xε(t,ω;ξ)=\displaystyle X_{\varepsilon}(t,\omega;\xi)= SA(t)ξ+0tSA(ts)F(Xε(s,ω;ξ))ds+ε0tSA(ts)B(Xε(s,ω;ξ))dWs(ω)-a.s.\displaystyle S_{A}(t)\xi+\int_{0}^{t}S_{A}(t-s)F(X_{\varepsilon}(s,\omega;\xi))\mathrm{d}s+\sqrt{\varepsilon}\int_{0}^{t}S_{A}(t-s)B(X_{\varepsilon}(s,\omega;\xi))\mathrm{d}W_{s}(\omega)\quad\mathbb{P}\text{-a.s. }

for each t[0,T]t\in[0,T]. In particular, the appearing integrals are well defined.

Theorem 2.3.

For any ε,T>0\varepsilon,T>0, ξH\xi\in H, there exists a unique mild solution XεC([0,T];H),a.s.X_{\varepsilon}\in C([0,T];H),a.s. of Eq. (2.1) in the sense of definition 2.2.

Proof.

It is obvious that the Hypothesis 2.1 satisfies the Hypothesis H1H1 - H4H4 in Chapter 4 of [18], which guarantees that the variational solution of Eq. (2.1) exists and is unique. Combining the Appendix FF in [24], we get the conclusion. ∎

To define the stationary solution of random dynamical system, let (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) be a complete probability space, θ:×ΩΩ\theta:\mathbb{R}\times\Omega\rightarrow\Omega be a group of \mathbb{P}-preserving ergodic transformations on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}).

Definition 2.4.

(cf. [1]) A crude cocycle (U,θ)(U,\theta) on HH is a ((+)\left(\mathcal{B}\left(\mathbb{R}^{+}\right)\otimes\right. (H),(H))\mathcal{B}(H)\otimes\mathcal{F},\mathcal{B}(H))-measurable random field U:+×H×ΩHU:\mathbb{R}^{+}\times H\times\Omega\rightarrow H with the following properties:
(i). U(t1+t2,,ω)=U(t2,U(t1,,ω),θ(t1,ω))U\left(t_{1}+t_{2},\cdot,\omega\right)=U\left(t_{2},U\left(t_{1},\cdot,\omega\right),\theta\left(t_{1},\omega\right)\right) for fixed t1t_{1}, and all t2t_{2}, \mathbb{P}-a.s. (where the exceptional set 𝒩t1\mathcal{N}_{t_{1}} can depend on t1t_{1}).
(ii). U(0,ξ,ω)=ξU(0,\xi,\omega)=\xi for all ξH\xi\in H, ωΩ\omega\in\Omega.

Definition 2.5.

An \mathcal{F} measurable random variable v:ΩH\textbf{v}:\Omega\rightarrow H is said to be a stationary solution for the crude cocycle (U,θ)(U,\theta) if it satisfies for any rr

U(,Y(r,ω),θ(r,ω))=Y(+r,ω)=Y(,θ(r,ω)),a.s.U(\cdot,Y(r,\omega),\theta(r,\omega))=Y(\cdot+r,\omega)=Y(\cdot,\theta(r,\omega)),\quad a.s. (2.3)
Remark 2.2.

We denote by θ:×ΩΩ\theta:\mathbb{R}\times\Omega\rightarrow\Omega be the standard \mathbb{P}-preserving ergodic Wiener shift on Ω\Omega, θ(t,ω(s)):=ω(t+s)ω(t)\theta(t,\omega(s)):=\omega(t+s)-\omega(t), t,st,s\in\mathbb{R}. Noticing that the stationary solution defined in Definition 2.5 means that for any rr, Eq. (2.3) holds almost surely, which is slightly different from the definition in [19, 22] needs Eq. (2.3) holds for all rr for all ωΩ\omega\in\Omega with respect to the perfect cocycle (cf. [1, 19, 22]). We prove the unique solution of the pullback integral Eq. (5.10) be the stationary solution (in the sense of Definition 2.5), and the distribution of the solution for (5.10) generates an invariant measure. Our results about the LDP are independent of the perfect cocycle, since the LDP describes the properties of the distribution of stochastic processes, but not the pathwise properties. Even the well-posedness of the pullback infinite horizon integral equation (5.10) is enough to show the LDP and depicts the long time asymptotic behavior of random dynamical systems with small perturbations.,

Liu and Zhao [19] have proved the stationary solution for Burgers equation with additive noise by using the proof of Mattingly [21]. However, we consider Eq. (2.1) with multiplicative noise, we prove the following theorem for the completeness of the present paper in Section 5.2.

Theorem 2.6.

There exists ε0>0\varepsilon_{0}>0, for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), Eq. (2.1) exists a unique stationary solution (in the sense of Definition 2.5) Xε(t,ω)(t)X^{*}_{\varepsilon}(t,\omega)(t\in\mathbb{R}) be a ((),(H))(\mathcal{B}(\mathbb{R})\otimes\mathcal{F},\mathcal{B}(H))-measurable, (t)t(\mathcal{F}_{-\infty}^{t})_{t\in\mathbb{R}}-adapted process. Moreover, Xε(t,ω)X_{\varepsilon}^{*}(t,\omega) satisfies the following equation in HH for any tt\in\mathbb{R}

Xε(t)=tSA(ts)F(Xε)ds+εtSA(ts)B(Xε)dWs,X_{\varepsilon}^{*}(t)=\int_{-\infty}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}\right)\mathrm{d}W_{s}, (2.4)

and

supt𝔼Xε(t)H2<.\sup_{t\in\mathbb{R}}\mathbb{E}\left\|X_{\varepsilon}^{*}(t)\right\|_{H}^{2}<\infty. (2.5)

2.2 Preliminary knowledge for LDP

We recall some standard definitions and results of the large deviations theory (cf. [3, 9, 7]). Let EE be a Polish space, {Xε}\{X^{*}_{\varepsilon}\} be a family of EE-valued random variables defined on a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}). The large deviation theory concerns the exponential decay of the probability measures of rare events. The rate of such exponential decay is expressed by the rate function.

Definition 2.7.

(Rate function). A function II mapping EE to [0,+)[0,+\infty) is called a rate function if II is lower semi-continuous. A rate function II is called a good rate function if for each M<M<\infty, the level set {xE:I(x)M}\{x\in E:I(x)\leq M\} is compact.

Definition 2.8.

(Large deviation principle). The sequence {Xε}\{X^{*}_{\varepsilon}\} is said to satisfies the LDP with rate function II if for each Borel subset AA of EE,

infxAI(x)lim infε0εlog(XεA)lim supε0εlog(XεA)infxA¯I(x),-\inf_{x\in A^{\circ}}I(x)\leq\liminf_{\varepsilon\longrightarrow 0}\varepsilon\log\mathbb{P}(X^{*}_{\varepsilon}\in A)\leq\limsup_{\varepsilon\longrightarrow 0}\varepsilon\log\mathbb{P}(X^{*}_{\varepsilon}\in A)\leq-\inf_{x\in\bar{A}}I(x),

where AA^{\circ} and A¯\bar{A} are the interior and the closure of AA in EE, respectively.

Since we can not prove the exponential tightness for stationary solutions of Eq. (2.1) with multiplicative noise in infinite intervals, which extremely depends on the properties of the Gaussian measure. We choose the weak convergence method to study the LDP of stationary solution for Eq. (2.1). Some fundamental concepts and results about the weak convergence method are stated as follows.

Definition 2.9.

(Laplace principle). The sequence {Xε}\{X_{\varepsilon}\} is said to satisfy the Laplace principle with a rate function II if for each bounded continuous real-valued function hh defined on EE, we have

limε0εlog𝔼{sup[1εh(Xε)]}=infxE{h(x)+I(x)}.\lim_{\varepsilon\rightarrow 0}\varepsilon\log\mathbb{E}\Big{\{}\sup\Big{[}-\frac{1}{\varepsilon}h(X_{\varepsilon})\Big{]}\Big{\}}=-\inf_{x\in E}\{h(x)+I(x)\}. (2.6)

We consider the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) satisfies LDP in the Polish space E:=C(;H)E:=C(\mathbb{R};H) with the norm

fC(;H):=f𝒞=k=12k(sups[k,k]f(s)H1),fC(;H).\|f\|_{C(\mathbb{R};H)}:=\|f\|_{\mathcal{C}}=\sum_{k=1}^{\infty}2^{-k}\left(\sup_{s\in[-k,k]}\left\|f(s)\right\|_{H}\wedge 1\right),\quad f\in C(\mathbb{R};H). (2.7)

According to Theorems 1.2.1 and 1.2.3 in Dupuis and Ellis [9], if EE is a Polish space and II is a good rate function, then the LDP and Laplace principle are equivalent, which is the basis for the weak convergence approach. Moreover, we need the Boué-Dupuis formula in infinite intervals to prove the weak convergence approach in infinite intervals. Although the Boué-Dupuis formula in infinite intervals has been used in [23] by Barashkov and Gubinelli, for the completeness of the article, we still give the proof of Boué-Dupuis formula and the weak convergence approach in infinite intervals in Appendix A, and illustrate the results in the following.
Let

𝒜=\displaystyle\mathcal{A}= {v:v is theH0-valuedt-predictable process and +v(s)H02ds<a.s.},\displaystyle\left\{v:\quad v\text{~{}is the}~{}H_{0}\text{-valued}~{}\mathcal{F}_{t}\text{-predictable process and~{}}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s<\infty~{}a.s.\right\},

and

SM:\displaystyle S_{M}: ={vL2(;H0):v(s)H02dsM}.\displaystyle=\left\{v\in L^{2}(\mathbb{R};H_{0}):\int_{-\infty}^{\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\leq M\right\}.

It is similar to prove the set SMS_{M} endowed with the weak topology is a Polish space (cf. [17], Theorem III.1’). In this paper, except for special instructions, the topology of SMS_{M} is always weak topology. We also define

𝒜M:={v𝒜,v(ω)SM,-a.s.}.\mathcal{A}_{M}:=\Big{\{}v\in\mathcal{A},\quad v(\omega)\in S_{M},\quad\mathbb{P}\text{-}a.s.\Big{\}}.

We provide the sufficient condition for the Laplace principle (equivalently, the LDP), which is similar to Assumption 4.3 in [5], in the follwoing.

Condition 2.1.

There exists a measurable map 𝒢0:C(;H0)E\ \mathcal{G}^{0}:C(\mathbb{R};H_{0})\longrightarrow E such that the following hold:
(i). Let {vε:ε>0}𝒜M\{v^{\varepsilon}:\varepsilon>0\}\subset\mathcal{A}_{M} for some M<M<\infty. If vεv^{\varepsilon} converges to vv in distribution as SMS_{M}-valued random elements, then 𝒢ε(W()+1εvεds)\mathcal{G}^{\varepsilon}(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}^{\cdot}v^{\varepsilon}\mathrm{d}s) converges to 𝒢0(v(s)ds)\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s) in distribution as ε0\varepsilon\rightarrow 0.
(ii). For every M<M<\infty, the set KM={𝒢0(v(s)ds):vSM}K_{M}=\{\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s):v\in S_{M}\} is a compact subset of EE.

Similar to the proof of Budhiraja and Dupuis in [3], we prove the following result in Appendix A.4.

Theorem 2.10.

If Xε=𝒢ε(W())X_{\varepsilon}=\mathcal{G}^{\varepsilon}(W(\cdot)) satisfies Condition 2.1, then the family {Xε:ε>0}\{X_{\varepsilon}:\varepsilon>0\} satisfies the Laplace principle in EE with good rate function

I(f)=inf{vL2(;H0):f=𝒢0(v(s)ds)}{12+v(s)H02ds}.I(f)=\inf_{\left\{v\in L^{2}(\mathbb{R};H_{0}):f=\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\right\}. (2.8)

where the infimum over an empty set is taken as ++\infty.

In order to verify Condition 2.1 for the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) in space C(;H)C(\mathbb{R};H), we need to consider 𝒢0\mathcal{G}^{0}, which is defined as the unique solution of the skeleton equation (3.1). Furthermore, we need to prove the well-posedness with respect to Eq. (3.1) in infinite time intervals.

2.3 Definition of the quasi-potential

It is well known that [14] and [6] have proved the family of invariant measures for stochastic equations satisfies LDP with the quasi-potential as rate function. We give the definition of quasi-potential below, and compare the rate function of the LDP for invariant measures induced by the LDP for stationary solutions and the rate function defined by quasi-potential in Section 4.

For any t1<t2+-\infty\leq t_{1}<t_{2}\leq+\infty and vL2(t1,t2;H0)v\in L^{2}(t_{1},t_{2};H_{0}), we denote by ut1x(v)u^{x}_{t_{1}}(v) any solution belonging to C(t1,t2;H)C(t_{1},t_{2};H) of the control equation

dudt=Au+F(u)+B(u)v,u(t1)=xH.\frac{\mathrm{d}u}{\mathrm{d}t}=Au+F(u)+B(u)v,\quad u(t_{1})=x\in H. (2.9)

And we define the action functionals by

St0,t1(u):=12inf{t0t1v(t)H02dt;u=u(v)},S_{t_{0},t_{1}}(u):=\frac{1}{2}\inf\left\{\int_{t_{0}}^{t_{1}}\left\|v(t)\right\|_{H_{0}}^{2}\mathrm{d}t;\quad u=u(v)\right\},

where u(v)u(v) is the solution of Eq. (2.9) in the intervals [t1,t2][t_{1},t_{2}] corresponding to the control vv, and inf=+.\inf\varnothing=+\infty.

Moreover, we denote

ST:=ST,0,ST:=S0,T, for every T>0.S_{-T}:=S_{-T,0},\quad S_{T}:=S_{0,T},\quad\text{ for every }T>0.

In particular, when t0=t_{0}=-\infty and t1=0t_{1}=0, we set

S(u):=12inf{0v(t)H02dt;u=u(v)}.S_{-\infty}(u):=\frac{1}{2}\inf\left\{\int_{-\infty}^{0}\left\|v(t)\right\|_{H_{0}}^{2}\mathrm{d}t;\quad u=u(v)\right\}.

Similar to [14] and [6], we define the quasi-potential VV associated with Eq. (2.1), by setting

V(x)\displaystyle V(x) :=inf{ST(u):T>0,uC([0,T];H),u(0)=0,u(T)=x}\displaystyle:=\inf\left\{S_{T}(u):T>0,u\in C([0,T];H),u(0)=0,u(T)=x\right\}
=inf{ST(u):T>0,uC([T,0];H),u(T)=0,u(0)=x},xH.\displaystyle=\inf\left\{S_{-T}(u):T>0,u\in C([-T,0];H),u(-T)=0,u(0)=x\right\},\quad x\in H.

It follows from the Proposition 5.4 in [6], which is also mentioned in Chapter 4 of [14], quasi-potential VV has a good characterization as follows.

V(x)=inf{S(u):uC((,0];H),u(0)=x,limtu(t)H=0},V(x)=\inf\Big{\{}S_{-\infty}(u):u\in C((-\infty,0];H),u(0)=x,\lim_{t\rightarrow-\infty}\|u(t)\|_{H}=0\Big{\}}, (2.10)

for all xHx\in H, V(x)<.V(x)<\infty.

We will prove that the rate function (4.3) defined in Theorem 4.2 is equivalent to the rate function (2.10) defined by quasi-potential for invariant measures.

3 Large deviation principle for stationary solution

3.1 The proof of LDP for Large deviation principle for stationary solution

We verify that the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) satisfies Condition 2.1, which is a sufficient condition to XεX^{*}_{\varepsilon} satisfies LDP in this section.

To define 𝒢0\mathcal{G}^{0} in Condition 2.1, we consider the following control equation

dX(t)dt=AX(t)+F(X(t))+B(X)v.\frac{\mathrm{d}X(t)}{\mathrm{d}t}=AX(t)+F(X(t))+B(X)v. (3.1)

We illustrate the following result for the skeleton equation (3.1), and Theorem 3.1 will be proved in Section 5.1.

Theorem 3.1.

For some finite M>0M>0 and vSMv\in S_{M}, under Hypothesis 2.1, there exists a unique solution XvX^{*}_{v} of the backward infinite horizon integral equation

Xv(t,ω)=tSA(tr)F(Xv(r,ω))dr+tSA(tr)B(Xv(r,ω))v(r)dr,X^{*}_{v}(t,\omega)=\int_{-\infty}^{t}S_{A}(t-r)F(X^{*}_{v}(r,\omega))\mathrm{d}r+\int_{-\infty}^{t}S_{A}(t-r)B(X^{*}_{v}(r,\omega))v(r)\mathrm{d}r, (3.2)

and satisfies

suptXv(t)H2<.\sup_{t\in\mathbb{R}}\left\|X^{*}_{v}(t)\right\|_{H}^{2}<\infty. (3.3)

Therefore, we could define the measurable map 𝒢0:C(;H0)C(;H)\mathcal{G}^{0}:C(\mathbb{R};H_{0})\longrightarrow C(\mathbb{R};H) by

𝒢0(v(s)ds):=Xv(),\mathcal{G}^{0}\Big{(}\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\Big{)}:=X_{v}^{*}(\cdot),

where XvX_{v}^{*} be the unique solution of Eq. (3.2) with control term vv and satisfies (3.3).

We consider the LDP for the family of stationary solutions of Eq. (2.1) in space C(;H)C(\mathbb{R};H), which norm is defined by (2.7). According to Theorem 2.6, we have proved that there exists ε0>0\varepsilon_{0}>0 such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), there exists a unique stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) and satisfies the integral equation

Xε(t,ω)=tSA(tr)F(Xε(r,ω))dr+εtSA(tr)B(Xε(r,ω))dW(r).X^{*}_{\varepsilon}(t,\omega)=\int_{-\infty}^{t}S_{A}(t-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-r)B\left(X^{*}_{\varepsilon}(r,\omega)\right)\mathrm{d}W(r). (3.4)

Then we could define 𝒢ε:C(;H0)C(;H)\mathcal{G}^{\varepsilon}:C(\mathbb{R};H_{0})\rightarrow C(\mathbb{R};H) by 𝒢ε(W()):=Xε()\mathcal{G}^{\varepsilon}(W(\cdot)):=X_{\varepsilon}^{*}(\cdot). We define Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} by

𝒢ε(W()+1εvε(s)ds),\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}^{\cdot}v_{\varepsilon}(s)\mathrm{d}s\right),

with the help of the Girsanov transform, Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} be the unique stationary solution of the control equation

dXε,vε=AXε,vε(t,x)dt+F(Xε,vε(t,x))dt+εB(Xε,vε)dW(t)+B(Xε,vε)vεdt.\mathrm{d}X_{\varepsilon,v_{\varepsilon}}^{*}=AX_{\varepsilon,v_{\varepsilon}}^{*}(t,x)\mathrm{d}t+F(X_{\varepsilon,v_{\varepsilon}}^{*}(t,x))\mathrm{d}t+\sqrt{\varepsilon}B(X_{\varepsilon,v_{\varepsilon}}^{*})\mathrm{d}W(t)+B(X_{\varepsilon,v_{\varepsilon}}^{*})v_{\varepsilon}\mathrm{d}t. (3.5)

Before we prove the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) satisfies Condition 2.1, we first give the following Lemma, which will be used in the proof of Lemma 3.2.

Lemma 3.1.

If {vε:ε>0}𝒜M\{v^{\varepsilon}:\varepsilon>0\}\subset\mathcal{A}_{M}, for some M<M<\infty, then for every fixed k+k\in\mathbb{N}^{+}, there exists ε0,η>0\varepsilon_{0},\eta>0 and a constant CC, which depend only on M,ε0,k,DM,\varepsilon_{0},k,D such that the solution Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} of Eq. (3.5) satisfies

𝔼supt(,k)e2ηtXε,vεH2C,ε(0,ε0),\mathbb{E}\sup_{t\in(-\infty,k)}e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\leq C,\quad\forall\varepsilon\in(0,\varepsilon_{0}),

where DD is defined by Remark 2.1 (ii).

Proof.

By using the Itô formula, Hypothesis 2.1 (i) , Remark 2.1 (ii) and Young inequality, we get

e2ηtXε,vεH2\displaystyle e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}
=\displaystyle= 2ηte2ηsXε,vεH2ds+2te2ηsAXε,vε+F(Xε,vε),Xε,vε(s)Vds\displaystyle 2\eta\int_{-\infty}^{t}e^{2\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\mathrm{d}s+2\int_{-\infty}^{t}e^{2\eta s}{{}_{V^{*}}}\left\langle AX_{\varepsilon,v_{\varepsilon}}^{*}+F(X_{\varepsilon,v_{\varepsilon}}^{*}),X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right\rangle\mathrm{d}s
+2te2ηsB(Xε,vε(s))vε(s),Xε,vε(s)Hds\displaystyle+2\int_{-\infty}^{t}e^{2\eta s}\left\langle B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)v_{\varepsilon}(s),X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right\rangle_{H}\mathrm{d}s
+2εte2ηsXε,vε(s),B(Xε,vε(s))dW(s)H+εte2ηs|B(Xε,vε(s))|LQ2ds\displaystyle+2\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\langle X_{\varepsilon,v_{\varepsilon}}^{*}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\rangle_{H}+\varepsilon\int_{-\infty}^{t}e^{2\eta s}\left|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\right|_{L_{Q}}^{2}\mathrm{~{}\mathrm{d}}s
\displaystyle\leq 2ηte2ηsXε,vεH2ds2λte2ηsXε,vεV2ds+δte2ηsXε,vεH2ds\displaystyle 2\eta\int_{-\infty}^{t}e^{2\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\mathrm{d}s-2\lambda\int_{-\infty}^{t}e^{2\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{V}^{2}\mathrm{d}s+\delta\int_{-\infty}^{t}e^{2\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\mathrm{d}s
+D2δte2ηsvε(s)H02ds+2εte2ηsXε,vε(s),B(Xε,vε(s))dW(s)H+εte2ηsD2ds.\displaystyle+\frac{D^{2}}{\delta}\int_{-\infty}^{t}e^{2\eta s}\|v_{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+2\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle X_{\varepsilon,v_{\varepsilon}}^{*}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}+\varepsilon\int_{-\infty}^{t}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s.

After choosing η,δ\eta,\delta small enough, it follows from Remark 2.1 (i) that

e2ηtXε,vεH2\displaystyle e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\leq D2δte2ηsvε(s)H02ds+εte2ηsD2ds\displaystyle\frac{D^{2}}{\delta}\int_{-\infty}^{t}e^{2\eta s}\|v_{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+\varepsilon\int_{-\infty}^{t}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s
+2εte2ηsXε,vε(s),B(Xε,vε(s))dW(s)H,\displaystyle+2\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle X_{\varepsilon,v_{\varepsilon}}^{*}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H},

thus for any k+k\in\mathbb{N}^{+}, we can obtain that

𝔼supt(,k)e2ηtXε,vεH2\displaystyle\mathbb{E}\sup_{t\in(-\infty,k)}e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\leq D2e2ηkδ𝔼+vε(s)H02ds+εke2ηsD2ds\displaystyle\frac{D^{2}e^{2\eta k}}{\delta}\mathbb{E}\int_{-\infty}^{+\infty}\|v_{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+\varepsilon\int_{-\infty}^{k}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s
+2ε𝔼supt(,k)te2ηsXε,vε(s),B(Xε,vε(s))dW(s)H.\displaystyle+2\sqrt{\varepsilon}\mathbb{E}\sup_{t\in(-\infty,k)}\int_{-\infty}^{t}e^{2\eta s}\left\langle X_{\varepsilon,v_{\varepsilon}}^{*}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}.

It follows from the Burkholder-Davis-Gundy inequality (cf. Proposition 3.26 in [16]), Young inequality and Remark 2.1 (ii) that

2ε𝔼{supt(,k]te2ηsXε,vε(s),B(Xε,vε(s))dW(s)H}12\displaystyle 2\sqrt{\varepsilon}\mathbb{E}\left\{\sup_{t\in(-\infty,k]}\int_{-\infty}^{t}e^{2\eta s}\left\langle X_{\varepsilon,v_{\varepsilon}}^{*}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}\right\}^{\frac{1}{2}}
\displaystyle\leq 2ε𝔼{ke4ηsXε,vεH2|B(Xε,vε)|LQ2ds}12\displaystyle 2\sqrt{\varepsilon}\mathbb{E}\left\{\int_{-\infty}^{k}e^{4\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}\right)|_{L_{Q}}^{2}\mathrm{d}s\right\}^{\frac{1}{2}}
\displaystyle\leq 2ε𝔼{supt(,k]eηtXε,vεH{ke2ηs|B(Xε,vε)|LQ2ds}12}\displaystyle 2\sqrt{\varepsilon}\mathbb{E}\left\{\sup_{t\in(-\infty,k]}e^{\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}\left\{\int_{-\infty}^{k}e^{2\eta s}|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}\right)|_{L_{Q}}^{2}\mathrm{d}s\right\}^{\frac{1}{2}}\right\}
\displaystyle\leq ε𝔼{supt(,k]e2ηtXε,vεH2+D2ke2ηsds},\displaystyle\sqrt{\varepsilon}\mathbb{E}\left\{\sup_{t\in(-\infty,k]}e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}+D^{2}\int_{-\infty}^{k}e^{2\eta s}\mathrm{d}s\right\},

thus for every ε<14\varepsilon<\frac{1}{4}, we could have

𝔼supt(,k)e2ηtXε,vεH2\displaystyle\mathbb{E}\sup_{t\in(-\infty,k)}e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\leq 2D2e2ηkδ𝔼+vε(s)H02ds+2εke2ηsD2ds\displaystyle\frac{2D^{2}e^{2\eta k}}{\delta}\mathbb{E}\int_{-\infty}^{+\infty}\|v_{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+2\varepsilon\int_{-\infty}^{k}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s (3.6)
+2εD2ke2ηsds.\displaystyle+2\sqrt{\varepsilon}D^{2}\int_{-\infty}^{k}e^{2\eta s}\mathrm{d}s.

For every ε>0\varepsilon>0, it follows from (3.6) and vε𝒜Mv_{\varepsilon}\in\mathcal{A}_{M}, there exists ε0(0,14)\varepsilon_{0}\in(0,\frac{1}{4}) such that

𝔼supt(,k)e2ηtXε,vεH2C(M,ε0,D,k),ε(0,ε0).\mathbb{E}\sup_{t\in(-\infty,k)}e^{2\eta t}\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H}^{2}\leq C(M,\varepsilon_{0},D,k),\quad\forall\varepsilon\in(0,\varepsilon_{0}).

The following lemmas will verify the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) satisfies Condition 2.1 in space C(;H)C(\mathbb{R};H), which is a sufficient condition to XεX^{*}_{\varepsilon} satisfies the LDP in space C(;H)C(\mathbb{R};H).

Lemma 3.2.

Let {vε:ε>0}𝒜M\{v^{\varepsilon}:\varepsilon>0\}\subset\mathcal{A}_{M}, for some M<M<\infty. If vεv^{\varepsilon} converges to vv as SMS_{M}-valued random element in distribution, then Xε,vεXvX_{\varepsilon,v_{\varepsilon}}^{*}\rightarrow X_{v}^{*} in distribution as ε0\varepsilon\rightarrow 0.

Proof.

Since Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} converges to XvX_{v}^{*} in probability could deduce Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} converges to XvX^{*}_{v} in distribution, it is sufficient to prove Xε,vεX_{\varepsilon,v_{\varepsilon}}^{*} converges to XvX^{*}_{v} in probability. For convenience, let wε:=Xε,vεXvw_{\varepsilon}:=X_{\varepsilon,v_{\varepsilon}}^{*}-X_{v}^{*}. Since for any δ>0\delta>0, there exists N>0N>0 such that k=N+112k<δ2\sum_{k=N+1}^{\infty}\frac{1}{2^{k}}<\frac{\delta}{2}, then it follows from the definition of the norm (2.7) in space C(;H)C(\mathbb{R};H),

(wε𝒞>δ)\displaystyle\mathbb{P}(\|w_{\varepsilon}\|_{\mathcal{C}}>\delta)\leq (k=1Nsupt[k,k]wε(t)H12k>δ2)\displaystyle\mathbb{P}\left(\sum_{k=1}^{N}\frac{\sup_{t\in[-k,k]}\|w_{\varepsilon}(t)\|_{H}\wedge 1}{2^{k}}>\frac{\delta}{2}\right)
\displaystyle\leq k=1N(supt[k,k]wε(t)H12k>δ2N).\displaystyle\sum_{k=1}^{N}\mathbb{P}\left(\frac{\sup_{t\in[-k,k]}\|w_{\varepsilon}(t)\|_{H}\wedge 1}{2^{k}}>\frac{\delta}{2N}\right).

Thus it is sufficient to prove that supt[k,k]wε(t)H0\sup_{t\in[k,k]}\|w_{\varepsilon}(t)\|_{H}\rightarrow 0 as ε0\varepsilon\rightarrow 0 in probability for every k+k\in\mathbb{N}^{+}.

By using the Itô formula, Hypothesis 2.1 (i) and Remark 2.1 (ii), we can get that

12e2ηtwεH2=\displaystyle\frac{1}{2}e^{2\eta t}\|w_{\varepsilon}\|_{H}^{2}= ηte2ηswεH2ds+te2ηsAwε+F(Xε,vε)F(Xv),wε(s)Vds\displaystyle\eta\int_{-\infty}^{t}e^{2\eta s}\|w_{\varepsilon}\|_{H}^{2}\mathrm{d}s+\int_{-\infty}^{t}e^{2\eta s}{{}_{V^{*}}}\left\langle Aw_{\varepsilon}+F(X_{\varepsilon,v_{\varepsilon}}^{*})-F(X_{v}^{*}),w_{\varepsilon}(s)\right\rangle\mathrm{d}s
+te2ηsB(Xε,vε(s))vε(s)B(Xv(s))v(s),wε(s)Hds\displaystyle+\int_{-\infty}^{t}e^{2\eta s}\left\langle B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)v_{\varepsilon}(s)-B\left(X_{v}^{*}(s)\right)v(s),w_{\varepsilon}(s)\right\rangle_{H}\mathrm{d}s
+εte2ηswε(s),B(Xε,vε(s))dW(s)H+ε2te2ηs|B(Xε,vε(s))Q1/2|L22ds\displaystyle+\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}+\frac{\varepsilon}{2}\int_{-\infty}^{t}e^{2\eta s}\left|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)Q^{1/2}\right|_{L_{2}}^{2}\mathrm{~{}\mathrm{d}}s
\displaystyle\leq ηte2ηswεH2dsλte2ηswε(s)V2+C0te2ηswε(s)H2Xv(s)V2ds\displaystyle\eta\int_{-\infty}^{t}e^{2\eta s}\|w_{\varepsilon}\|_{H}^{2}\mathrm{d}s-\lambda\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{V}^{2}+C_{0}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{H}^{2}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{d}s
+te2ηs|B(Xε,vε(s))B(Xv(s))Q1/2|L2vε(s)H0wε(s)Hds\displaystyle+\int_{-\infty}^{t}e^{2\eta s}\left|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)-B\left(X_{v}^{*}(s)\right)Q^{1/2}\right|_{L_{2}}\left\|v_{\varepsilon}(s)\right\|_{H_{0}}\left\|w_{\varepsilon}(s)\right\|_{H}\mathrm{d}s
+te2ηsB(Xv(s))(vε(s)v(s))Hwε(s)Hds\displaystyle+\int_{-\infty}^{t}e^{2\eta s}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}\left\|w_{\varepsilon}(s)\right\|_{H}\mathrm{d}s
+εte2ηswε(s),B(Xε,vε(s))dW(s)H+ε2te2ηsD2ds.\displaystyle+\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}+\frac{\varepsilon}{2}\int_{-\infty}^{t}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s.

It follows from the Young inequality, Remark 2.1 (i) and (ii), after choosing η\eta small enough, we can obtain that

12e2ηtwεH2+3λ4te2ηswε(s)V2ds\displaystyle\frac{1}{2}e^{2\eta t}\|w_{\varepsilon}\|_{H}^{2}+\frac{3\lambda}{4}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s
\displaystyle\leq C0te2ηswε(s)H2Xv(s)V2ds+βte2ηswε(s)Hvε(s)H0wε(s)Hds\displaystyle C_{0}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{H}^{2}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s+\beta\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{H}\left\|v_{\varepsilon}(s)\right\|_{H_{0}}\left\|w_{\varepsilon}(s)\right\|_{H}\mathrm{d}s
+te2ηsB(Xv(s))(vε(s)v(s))Hwε(s)Hds\displaystyle+\int_{-\infty}^{t}e^{2\eta s}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}\left\|w_{\varepsilon}(s)\right\|_{H}\mathrm{d}s
+εte2ηswε(s),B(Xε,vε(s))dW(s)H+ε2te2ηsD2ds\displaystyle+\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}+\frac{\varepsilon}{2}\int_{-\infty}^{t}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s
\displaystyle\leq C0te2ηswε(s)H2Xv(s)V2ds+λ4te2ηswε(s)V2ds+2β2λC12te2ηs|vε(s)|H02wεH2ds\displaystyle C_{0}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{H}^{2}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s+\frac{\lambda}{4}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s+\frac{2\beta^{2}}{\lambda C_{1}^{2}}\int_{-\infty}^{t}e^{2\eta s}\left|v_{\varepsilon}(s)\right|_{H_{0}}^{2}\|w_{\varepsilon}\|_{H}^{2}\mathrm{d}s
+λ4te2ηswε(s)V2ds+2λC12te2ηsB(Xv(s))(vε(s)v(s))H2ds\displaystyle+\frac{\lambda}{4}\int_{-\infty}^{t}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s+\frac{2}{\lambda C_{1}^{2}}\int_{-\infty}^{t}e^{2\eta s}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s
+εte2ηswε(s),B(Xε,vε(s))dW(s)H+ε2te2ηsD2ds.\displaystyle+\sqrt{\varepsilon}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}+\frac{\varepsilon}{2}\int_{-\infty}^{t}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s.

Define

τN,εk:=kinf{\displaystyle\tau_{N,\varepsilon}^{k}:=k\wedge\inf\Big{\{} t:tXv(s)V2ds>Norsups(,t]Xv(s)H2>N\displaystyle t:\int_{-\infty}^{t}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{d}s>N\quad\text{\rm or}\quad\sup_{s\in(-\infty,t]}\|X_{v}^{*}(s)\|_{H}^{2}>N
orsups(,t]e2ηsXε,vε(s)H2>N},\displaystyle\text{\rm or}\quad\sup_{s\in(-\infty,t]}e^{2\eta s}\|X_{\varepsilon,v_{\varepsilon}}^{*}(s)\|_{H}^{2}>N\Big{\}},

τN,εk\tau_{N,\varepsilon}^{k} is a continuous random process for XvH\|X_{v}^{*}\|_{H} and Xε,vεH\|X_{\varepsilon,v_{\varepsilon}}^{*}\|_{H} are continuous (cf. Theorem 5.16 or Theorem 2.2 in [19]).

Then for any fixed k+k\in\mathbb{N}^{+}, T[k,k]T\in[-k,k], there exists a constant CC, which only depend on λ,C0,β,D,C1\lambda,C_{0},\beta,D,C_{1} such that

supt(,TτN,εk]e2ηtwε(t)H2+TτN,εke2ηswε(s)V2ds\displaystyle\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}e^{2\eta t}\|w_{\varepsilon}(t)\|_{H}^{2}+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s
\displaystyle\leq C{TτN,εke2ηswε(s)H2Xv(s)V2ds+TτN,εke2ηsvεH02wε(s)H2ds\displaystyle C\left\{\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}\left\|w_{\varepsilon}(s)\right\|_{H}^{2}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}\|v_{\varepsilon}\|_{H_{0}}^{2}\left\|w_{\varepsilon}(s)\right\|_{H}^{2}\mathrm{d}s\right.
+TτN,εke2ηsB(Xv(s))(vε(s)v(s))H2ds+εTτN,εke2ηsD2ds\displaystyle+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s+\varepsilon\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s
+εsupt(,TτN,εk]te2ηswε(s),B(Xε,vε(s))dW(s)H}.\displaystyle\left.+\sqrt{\varepsilon}\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}\right\}.

Let Mt:=sups(,t]e2ηswε(s)H2M_{t}:=\sup_{s\in(-\infty,t]}e^{2\eta s}\|w_{\varepsilon}(s)\|_{H}^{2}, then we have

MTτN,εk\displaystyle M_{T\wedge\tau_{N,\varepsilon}^{k}}\leq C{TτN,εkMsXv(s)V2ds+TτN,εkMs|vε(s)|H02ds\displaystyle C\left\{\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}M_{s}\left\|X_{v}^{*}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}M_{s}\left|v_{\varepsilon}(s)\right|_{H_{0}}^{2}\mathrm{d}s\right.
+TτN,εkB(Xv(s))(vε(s)v(s))H2ds+εTτN,εke2ηsD2ds\displaystyle+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}\left\|B\left(X_{v}^{*}(s)\right)(v_{\varepsilon}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s+\varepsilon\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s
+εsupt(,TτN,εk]te2ηswε(s),B(Xε,vε(s))dW(s)H}.\displaystyle\left.+\sqrt{\varepsilon}\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}\right\}.

By using the Burkholder-Davis-Gundy inequality, Young inequality and Remark 2.1 (ii), there exists a constant CC such that

𝔼{supt(,TτN,εk]te2ηswε(s),B(Xε,vε(s))dW(s)H}12\displaystyle\mathbb{E}\left\{\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}\right\}^{\frac{1}{2}}
\displaystyle\leq C𝔼{TτN,εke4ηswεH2|B(Xε,vε)|LQ2ds}12\displaystyle C\mathbb{E}\left\{\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{4\eta s}\|w_{\varepsilon}\|_{H}^{2}|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}\right)|_{L_{Q}}^{2}\mathrm{d}s\right\}^{\frac{1}{2}}
\displaystyle\leq C𝔼{supt(,TτN,εk]eηtwεH{TτN,εke2ηs|B(Xε,vε)|LQ2ds}12}\displaystyle C\mathbb{E}\left\{\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}e^{\eta t}\|w_{\varepsilon}\|_{H}\left\{\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}|B\left(X_{\varepsilon,v_{\varepsilon}}^{*}\right)|_{L_{Q}}^{2}\mathrm{d}s\right\}^{\frac{1}{2}}\right\}
\displaystyle\leq C2𝔼{supt(,TτN,εk]e2ηtwεH2+D2TτN,εke2ηsds}<.\displaystyle\frac{C}{2}\mathbb{E}\left\{\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}e^{2\eta t}\|w_{\varepsilon}\|_{H}^{2}+D^{2}\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}\mathrm{d}s\right\}<\infty.

Then using the Gronwall inequality, we could obtain that

MTτN,εk\displaystyle M_{T\wedge\tau_{N,\varepsilon}^{k}} CeC(N+M){εsupt(,TτN,εk]te2ηswε(s),B(Xε,vε(s))dW(s)H\displaystyle\leq Ce^{C(N+M)}\left\{\sqrt{\varepsilon}\sup_{t\in(-\infty,T\wedge\tau_{N,\varepsilon}^{k}]}\int_{-\infty}^{t}e^{2\eta s}\left\langle w_{\varepsilon}(s),B\left(X_{\varepsilon,v_{\varepsilon}}^{*}(s)\right)\mathrm{d}W(s)\right\rangle_{H}\right. (3.8)
+εTτN,εke2ηsD2ds+TτN,εkB(Xv(s))(vε(s)v(s))H2ds}.\displaystyle\left.+\varepsilon\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}e^{2\eta s}D^{2}\mathrm{~{}\mathrm{d}}s+\int_{-\infty}^{T\wedge\tau_{N,\varepsilon}^{k}}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s\right\}.

It follows from B()Q1/2B(\cdot)Q^{1/2} is a Hilbert-Schmidt operator on HH, Remark 2.1 (ii), and vεv_{\varepsilon} converges to vv as SMS_{M}-value random variable in distribution,

limε0𝔼+B(Xv(s))(vε(s)v(s))H2dslimε0D2𝔼+(vε(s)v(s))H02ds=0.\lim_{\varepsilon\rightarrow 0}\mathbb{E}\int_{-\infty}^{+\infty}\left\|B(X_{v}^{*}(s))(v_{\varepsilon}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s\leq\lim_{\varepsilon\rightarrow 0}D^{2}\mathbb{E}\int_{-\infty}^{+\infty}\left\|(v_{\varepsilon}(s)-v(s))\right\|_{H_{0}}^{2}\mathrm{d}s=0. (3.9)

It follows from Lemma 5.2, Lemma 3.1, and the Chebyshev inequality, for any fixed kk, it is easy to obtain that there exists ε0\varepsilon_{0}, which is defined by Lemma 3.1, and a constant CC such that

{τN,εk=k}1CN,ε(0,ε0),\mathbb{P}\{\tau_{N,\varepsilon}^{k}=k\}\geq 1-\frac{C}{N},\quad\forall\varepsilon\in(0,\varepsilon_{0}),

it implies that limNτN,εk=k,a.e.ε(0,ε0)\lim\limits_{N\rightarrow\infty}\tau_{N,\varepsilon}^{k}=k,~{}a.e.~{}\forall\varepsilon\in(0,\varepsilon_{0}). Thus combining inequalitys (3.1), (3.8) and (3.9), let NN\rightarrow\infty and ε0\varepsilon\rightarrow 0, we get that

supt(,k]e2ηtwε(t)H20\sup_{t\in(-\infty,k]}e^{2\eta t}\|w_{\varepsilon}(t)\|_{H}^{2}\rightarrow 0 (3.10)

in probability. Moreover, for any fixed kk, it follows from

eηksupt[k,k]wε(t)H2supt[k,k]e2ηtwε(t)H2supt(,k]e2ηtwε(t)H2,e^{-\eta k}\sup_{t\in[-k,k]}\|w_{\varepsilon}(t)\|_{H}^{2}\leq\sup_{t\in[-k,k]}e^{2\eta t}\|w_{\varepsilon}(t)\|_{H}^{2}\leq\sup_{t\in(-\infty,k]}e^{2\eta t}\|w_{\varepsilon}(t)\|_{H}^{2},

and (3.10) to obtain that supt[k,k]wε(t)H20\sup_{t\in[-k,k]}\|w_{\varepsilon}(t)\|_{H}^{2}\rightarrow 0 in probability as ε0\varepsilon\rightarrow 0. ∎

Lemma 3.3.

For any fixed finite positive number MM, the set KM={𝒢0(v(s)ds);vSM}K_{M}=\Big{\{}\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s);v\in S_{M}\Big{\}} is a compact subset of C((,0];H)C((-\infty,0];H).

Proof.

Let {𝒢0(vn(s)ds)}n=1\big{\{}\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v_{n}(s)\mathrm{d}s)\big{\}}_{n=1}^{\infty} be a sequence in KMK_{M}, for convenience let Xn:=𝒢0(vn(s)ds)X^{n}:=\mathcal{G}^{0}\big{(}\int_{-\infty}^{\cdot}v_{n}(s)\mathrm{d}s\big{)} where XnX^{n} corresponds to the solution of Eq. (3.1) with vnSMv_{n}\in S_{M} in place of vv. By weak compactness of SMS_{M}, there exists a subsequence of {vn}\{v_{n}\} which converges to a limit vv weakly in L2(;H0)L^{2}(\mathbb{R};H_{0}). The subsequence is indexed by nn for ease of notation. Let X:=𝒢0(v(s)ds)X:=\mathcal{G}^{0}\big{(}\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\big{)} be the solution of Eq. (3.1) responding to vv and wn=XnXw^{n}=X^{n}-X. Then we obtain

dwndt=Awn+F(Xn)F(X)+B(Xn)vnB(X)v,\frac{\mathrm{d}w^{n}}{\mathrm{d}t}=Aw_{n}+F(X^{n})-F(X)+B(X^{n})v_{n}-B(X)v,

it follows from Hypothesis 2.1 (i) and Remark 2.1 (i), (ii) that

12wn(t)H2=\displaystyle\frac{1}{2}\left\|w^{n}(t)\right\|_{H}^{2}= tAwn+F(Xn)F(X),wnVV\displaystyle\int_{-\infty}^{t}{{}_{V^{*}}}\langle Aw_{n}+F(X^{n})-F(X),w^{n}\rangle_{V}
+t{(B(Xn(s))B(X(s))vn(s),wn(s)H\displaystyle+\int_{-\infty}^{t}\Big{\{}\langle\left(B\left(X^{n}(s)\right)-B(X(s)\right)v_{n}(s),w^{n}(s)\rangle_{H}
+B(X(s))(vn(s)v(s)),wn(s)H}ds\displaystyle+\langle B(X(s))\left(v_{n}(s)-v(s)\right),w^{n}(s)\rangle_{H}\Big{\}}\mathrm{d}s
\displaystyle\leq λtwn(s)V2ds+C0twn(s)H2X(s)V2ds\displaystyle-\lambda\int_{-\infty}^{t}\left\|w^{n}(s)\right\|_{V}^{2}\mathrm{d}s+C_{0}\int_{-\infty}^{t}\left\|w^{n}(s)\right\|_{H}^{2}\|X(s)\|_{V}^{2}\mathrm{d}s
+βC1twn(s)Vwn(s)Hvn(s)H0ds\displaystyle+\frac{\beta}{C_{1}}\int_{-\infty}^{t}\left\|w^{n}(s)\right\|_{V}\left\|w^{n}(s)\right\|_{H}\left\|v_{n}(s)\right\|_{H_{0}}\mathrm{~{}\mathrm{d}}s
+tB(X(s))(vn(s)v(s))Hwn(s)Hds.\displaystyle+\int_{-\infty}^{t}\left\|B(X(s))(v_{n}(s)-v(s))\right\|_{H}\left\|w^{n}(s)\right\|_{H}\mathrm{d}s.

For any k+k\in\mathbb{N}^{+} and T[k,k]T\in[-k,k], by using the Young inequality and Remark 2.1 (i), we have

12supt(,T]wn(t)H2+λ2Twn(s)V2ds\displaystyle\frac{1}{2}\sup_{t\in(-\infty,T]}\left\|w^{n}(t)\right\|_{H}^{2}+\frac{\lambda}{2}\int_{-\infty}^{T}\left\|w^{n}(s)\right\|_{V}^{2}\mathrm{~{}\mathrm{d}}s\leq CTwn(s)H2(X(s)V2+vn(s)H02)ds\displaystyle C\int_{-\infty}^{T}\left\|w^{n}(s)\right\|_{H}^{2}\left(\|X(s)\|_{V}^{2}+\left\|v_{n}(s)\right\|_{H_{0}}^{2}\right)\mathrm{~{}\mathrm{d}}s
++B(X(s))(vn(s)v(s))H2ds,\displaystyle+\int_{-\infty}^{+\infty}\left\|B(X(s))(v_{n}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s,

where CC is a constant only depend on λ,C0,C1,β\lambda,C_{0},C_{1},\beta.

It follows from the Gronwall inequality that

12supt(,k]wn(t)H2\displaystyle\frac{1}{2}\sup_{t\in(-\infty,k]}\left\|w^{n}(t)\right\|_{H}^{2} (3.11)
\displaystyle\leq C+B(X(s))(vn(s)v(s))H2dsexp{k(X(s)V2+vn(s)H02)ds}.\displaystyle C\int_{-\infty}^{+\infty}\left\|B(X(s))(v_{n}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s\cdot\exp\left\{\int_{-\infty}^{k}\left(\|X(s)\|_{V}^{2}+\left\|v_{n}(s)\right\|_{H_{0}}^{2}\right)\mathrm{~{}\mathrm{d}}s\right\}.

It follows from vnSMv_{n}\in S_{M}, and Lemma 5.2 that

exp{k(X(s)V2+vn(s)H02)ds}<.\exp\left\{\int_{-\infty}^{k}\left(\|X(s)\|_{V}^{2}+\left\|v_{n}(s)\right\|_{H_{0}}^{2}\right)\mathrm{~{}\mathrm{d}}s\right\}<\infty. (3.12)

Since vn,vSMv_{n},v\in S_{M}, there exits simple function sequences v~nk\tilde{v}_{n_{k}} and v~k\tilde{v}_{k} strong convergence to vnv_{n} and vv respectively, we can choose a subsequence still record it as nkn_{k}, such that

limnlimk+v~nk(s)vn(s)H02ds=0,\lim_{n\rightarrow\infty}\lim_{k\rightarrow\infty}\int_{-\infty}^{+\infty}\|\tilde{v}_{n_{k}}(s)-v_{n}(s)\|_{H_{0}}^{2}\mathrm{d}s=0, (3.13)

since vnvv_{n}-v weak converges to 0 in SMS_{M}, v~nkv~k\tilde{v}_{n_{k}}-\tilde{v}_{k} also weak converges to 0. By using Remark 2.1 (ii) and Hypothesis 2.1 (iii), we can obtain that

+B(X(s))(vn(s)v(s))H2ds\displaystyle\int_{-\infty}^{+\infty}\left\|B(X(s))(v_{n}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s
\displaystyle\leq +B(X(s))(vn(s)v~nk(s))H2ds++B(X(s))(v~nk(s)v~k(s))H2ds\displaystyle\int_{-\infty}^{+\infty}\left\|B(X(s))(v_{n}(s)-\tilde{v}_{n_{k}}(s))\right\|_{H}^{2}\mathrm{d}s+\int_{-\infty}^{+\infty}\left\|B(X(s))(\tilde{v}_{n_{k}}(s)-\tilde{v}_{k}(s))\right\|_{H}^{2}\mathrm{d}s (3.14)
++B(X(s))(v~k(s)v(s))H2ds\displaystyle+\int_{-\infty}^{+\infty}\left\|B(X(s))(\tilde{v}_{k}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s
\displaystyle\leq D2+vn(s)v~nk(s)H02ds+D2+v~k(s)v(s)H02ds\displaystyle D^{2}\int_{-\infty}^{+\infty}\left\|v_{n}(s)-\tilde{v}_{n_{k}}(s)\right\|_{H_{0}}^{2}\mathrm{d}s+D^{2}\int_{-\infty}^{+\infty}\left\|\tilde{v}_{k}(s)-v(s)\right\|_{H_{0}}^{2}\mathrm{d}s
+\displaystyle+ D02+Q1/2Q1/2(v~nk(s)v~k(s))H2ds.\displaystyle D_{0}^{2}\int_{-\infty}^{+\infty}\left\|Q^{1/2}Q^{-1/2}(\tilde{v}_{n_{k}}(s)-\tilde{v}_{k}(s))\right\|_{H}^{2}\mathrm{d}s.

It follows from Q1/2Q^{1/2} is a Hilbert-Schmidt operator on HH and v~nkv~k\tilde{v}_{n_{k}}-\tilde{v}_{k} weak converges to 0 in SMS_{M} that there exists a subsequence, still set nkn_{k}, such that

limk+Q1/2Q1/2(v~nk(s)v~k(s))H2ds=0.\lim_{k\rightarrow\infty}\int_{-\infty}^{+\infty}\left\|Q^{1/2}Q^{-1/2}(\tilde{v}_{n_{k}}(s)-\tilde{v}_{k}(s))\right\|_{H}^{2}\mathrm{d}s=0. (3.15)

It follows from (3.13), (3.1), (3.15) and v~k\tilde{v}_{k} strong converges to vv in SMS_{M} that

limn+B(X(s))(vn(s)v(s))H2ds=0.\lim_{n\rightarrow\infty}\int_{-\infty}^{+\infty}\left\|B(X(s))(v_{n}(s)-v(s))\right\|_{H}^{2}\mathrm{d}s=0. (3.16)

Combining (3.11), (3.12) and (3.16), we could get that for any k+k\in\mathbb{N}^{+},

limnsupt(,k]wn(t)H2=0.\lim_{n\rightarrow\infty}\sup_{t\in(-\infty,k]}\left\|w^{n}(t)\right\|_{H}^{2}=0.

By using the dominated convergence theorem, we obtain that

limnXnX𝒞\displaystyle\lim_{n\rightarrow\infty}\left\|X^{n}-X\right\|_{\mathcal{C}} =limnk=1supt[k,k]XnXH12klimnk=1supt(,k]XnXH12k\displaystyle=\lim_{n\rightarrow\infty}\sum_{k=1}^{\infty}\frac{\sup_{t\in[-k,k]}\|X^{n}-X\|_{H}\wedge 1}{2^{k}}\leq\lim_{n\rightarrow\infty}\sum_{k=1}^{\infty}\frac{\sup_{t\in(-\infty,k]}\|X^{n}-X\|_{H}\wedge 1}{2^{k}}
=k=1limnsupt(,k]XnXH12k=0,\displaystyle=\sum_{k=1}^{\infty}\frac{\lim_{n\rightarrow\infty}\sup_{t\in(-\infty,k]}\|X^{n}-X\|_{H}\wedge 1}{2^{k}}=0,

it implies that KMK_{M} is compact. ∎

Lemma 3.2 and 3.3 have proved that the stationary solution XεX^{*}_{\varepsilon} of Eq. (2.1) satisfies Condition 2.1 in Polish Space C(;H)C(\mathbb{R};H). By using Theorem 2.10 and the equivalence of LDP and LP in Polish space, the following theorem holds.

Theorem 3.2.

Under the Hypothesis 2.1, the family {Xε:ε>0}\{X^{*}_{\varepsilon}:\varepsilon>0\} satisfies the LDP in C(;H)C(\mathbb{R};H) with rate function

I(f)=inf{vL2(;H0):f=𝒢0(v(s)ds)}{12+v(s)H02ds},I(f)=\inf_{\left\{v\in L^{2}(\mathbb{R};H_{0}):f=\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\right\}, (3.17)

where the infimum over an empty set is taken as ++\infty.

3.2 Some Examples

Next, we will give an example that satisfies the Hypothesis 2.1.

Example 1.

We consider the stochastic Burgers equation with multiplicative noise as follows,

{du(t,x)=Δu(t,x)dt+12x[u(t,x)]2dt+εB(u)dW(t),ts,x(0,1),u(s,x)=us(x),u(t,1)=u(t,0)=0.\left\{\begin{array}[]{l}\mathrm{d}\textbf{u}(t,x)=\Delta\textbf{u}(t,x)\mathrm{d}t+\frac{1}{2}\partial_{x}[\textbf{u}(t,x)]^{2}\mathrm{d}t+\sqrt{\varepsilon}B(\textbf{u})\mathrm{d}W(t),t\geqslant s,x\in(0,1),\\ \textbf{u}(s,x)=\textbf{u}_{s}(x),\\ \textbf{u}(t,1)=\textbf{u}(t,0)=0.\end{array}\right. (3.18)

To correspond to the notation of Hypothesis 2.1, let Au+F(u):=Δu(t,x)+12x[u(t,x)]2A\textbf{u}+F(\textbf{u}):=\Delta\textbf{u}(t,x)+\frac{1}{2}\partial_{x}[\textbf{u}(t,x)]^{2}.
Let

H:={f:[0,1]1:f(0)=f(1)=0,01f2(x)dx<},\displaystyle H:=\left\{f:[0,1]\rightarrow\mathbb{R}^{1}:f(0)=f(1)=0,\quad\int_{0}^{1}f^{2}(x)\mathrm{d}x<\infty\right\},
V:={fH:01(xf(x))2dx<}.\displaystyle V:=\left\{f\in H:\int_{0}^{1}(\partial_{x}f(x))^{2}\mathrm{d}x<\infty\right\}.

We will verify AuA\textbf{u} and F(u)F(\textbf{u}) satisfies Hypothesis 2.1 (i) and (ii) below.
By using integration by parts, we can obtain that

01u2(t,x)x(u(t,x))dx\displaystyle\int_{0}^{1}\textbf{u}^{2}(t,x)\partial_{x}(\textbf{u}(t,x))\mathrm{d}x =0,\displaystyle=0,
01u(t,x)w(t,x)x(w(t,x))dx\displaystyle\int_{0}^{1}\textbf{u}(t,x)\textbf{w}(t,x)\partial_{x}(\textbf{w}(t,x))\mathrm{d}x =1201w2(t,x)xu(t,x)dx,\displaystyle=-\frac{1}{2}\int_{0}^{1}\textbf{w}^{2}(t,x)\partial_{x}\textbf{u}(t,x)\mathrm{d}x,

for any u,wV\textbf{u},\textbf{w}\in V, then

F(u(t,x))F(w(t,x)),uwVV=\displaystyle{}_{V^{*}}\langle F(\textbf{u}(t,x))-F(\textbf{w}(t,x)),\textbf{u}-\textbf{w}\rangle_{V}= uxuwxw,uwH\displaystyle\langle\textbf{u}\partial_{x}\textbf{u}-\textbf{w}\partial_{x}\textbf{w},\textbf{u}-\textbf{w}\rangle_{H} (3.19)
=\displaystyle= ux(uw)+(uw)xu(uw)x(uw),uwH\displaystyle\langle\textbf{u}\partial_{x}(\textbf{u}-\textbf{w})+(\textbf{u}-\textbf{w})\partial_{x}\textbf{u}-(\textbf{u}-\textbf{w})\partial_{x}(\textbf{u}-\textbf{w}),\textbf{u}-\textbf{w}\rangle_{H}
=\displaystyle= 1201w2(t,x)xu(t,x)dx.\displaystyle\frac{1}{2}\int_{0}^{1}\textbf{w}^{2}(t,x)\partial_{x}\textbf{u}(t,x)\mathrm{d}x.

According to Lemma A.1 in [19], we could have the following property.

|01v(x)xw(x)u(x)dx|2γ2vHvVwV2uHuV,u,v,wV,\left|\int_{0}^{1}\textbf{v}(x)\partial_{x}\textbf{w}(x)\textbf{u}(x)\mathrm{d}x\right|^{2}\leq\gamma^{2}\|\textbf{v}\|_{H}\|\textbf{v}\|_{V}\|\textbf{w}\|_{V}^{2}\|\textbf{u}\|_{H}\|\textbf{u}\|_{V},\quad\forall\textbf{u},\textbf{v},\textbf{w}\in V, (3.20)

where γ\gamma is the minimal constant such that Sobolev’s inequality,

maxx[0,1]|u(x)|γuV,uV\max_{x\in[0,1]}|\textbf{u}(x)|\leq\gamma\|\textbf{u}\|_{V},\quad\forall u\in V

holds.

For linear operator AA, we have known that

VAu,uV=VΔu,uVuV2,uV._{V^{*}}\langle A\textbf{u},\textbf{u}\rangle_{V}=_{V^{*}}\langle\Delta\textbf{u},\textbf{u}\rangle_{V}\leq-\|\textbf{u}\|_{V}^{2},\quad\forall\textbf{u}\in V. (3.21)

It follows from (3.19), (3.20), (3.21) and the Young inequality that

AuAw+F(v)F(w),vwVV=\displaystyle{{}_{V^{*}}}\langle A\textbf{u}-A\textbf{w}+F(\textbf{v})-F(\textbf{w}),\textbf{v}-\textbf{w}\rangle_{V}= Δ(uw),uwVV+1201w2(t,x)xu(t,x)dx\displaystyle{{}_{V^{*}}}\langle\Delta(\textbf{u}-\textbf{w}),\textbf{u}-\textbf{w}\rangle_{V}+\frac{1}{2}\int_{0}^{1}\textbf{w}^{2}(t,x)\partial_{x}\textbf{u}(t,x)\mathrm{d}x
\displaystyle\leq uwV2+γ2uwHuwVuV\displaystyle-\|\textbf{u}-\textbf{w}\|_{V}^{2}+\frac{\gamma}{2}\|\textbf{u}-\textbf{w}\|_{H}\|\textbf{u}-\textbf{w}\|_{V}\|\textbf{u}\|_{V}
\displaystyle\leq 12uwV2+γ28uwH2uV2,\displaystyle-\frac{1}{2}\|\textbf{u}-\textbf{w}\|_{V}^{2}+\frac{\gamma^{2}}{8}\|\textbf{u}-\textbf{w}\|_{H}^{2}\|\textbf{u}\|_{V}^{2},

it implies that Eq. (3.18) satisfies Hypothesis 2.1 (i).

Let SAS_{A} be the heat semigroup, through the similar calculation to Lemma 3.3 in [19], for any N+N\in\mathbb{N}^{+}, t0<t[N,N]t_{0}<t\in[-N,N], we can get

t0tSA(s)(F(v)F(w))dsH2C(N)t0ts34vwH2ds,\left\|\int_{t_{0}}^{t}S_{A}(s)(F(\textbf{v})-F(\textbf{w}))\mathrm{d}s\right\|_{H}^{2}\leq C(N)\int_{t_{0}}^{t}s^{-\frac{3}{4}}\|\textbf{v}-\textbf{w}\|_{H}^{2}\mathrm{d}s,

it deduces that Eq. (3.18) satisfies Hypothesis 2.1 (ii).

Therefore, assume that the function BB satisfies Hypothesis 2.1 (iii). It follows from Lemma 2.6 that there exists ε0>0\varepsilon_{0}>0 such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the stationary solution uε\textbf{u}^{*}_{\varepsilon} of Burgers equation exists and is unique, which satisfies the following equation in HH for any tt\in\mathbb{R},

uε(t,ω)=12tSA(tr)x[uε(r,ω)]2dr+εtSA(tr)B(uε(r,ω))dW(r),\textbf{u}^{*}_{\varepsilon}(t,\omega)=\frac{1}{2}\int_{-\infty}^{t}S_{A}(t-r)\partial_{x}[\textbf{u}^{*}_{\varepsilon}(r,\omega)]^{2}\mathrm{d}r+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-r)B(\textbf{u}^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r),

and

supt𝔼uε(t)H2<.\sup_{t\in\mathbb{R}}\mathbb{E}\left\|\textbf{u}^{*}_{\varepsilon}(t)\right\|_{H}^{2}<\infty.

The skeleton equation of Burgers equation is

{du(t,x)dt=Δu(t,x)+12x[u(t,x)]2+B(u)v(t),ts,x(0,1),u(s,x)=us(x),u(t,1)=u(t,0)=0.\left\{\begin{array}[]{l}\frac{\mathrm{d}\textbf{u}(t,x)}{\mathrm{d}t}=\Delta\textbf{u}(t,x)+\frac{1}{2}\partial_{x}[\textbf{u}(t,x)]^{2}+B(\textbf{u})v(t),\quad t\geqslant s,x\in(0,1),\\ \textbf{u}(s,x)=\textbf{u}_{s}(x),\\ \textbf{u}(t,1)=\textbf{u}(t,0)=0.\end{array}\right. (3.22)

It follows from Lemma 5.5 and Lemma 5.4, for any vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), there exits a unique solution uv\textbf{u}^{*}_{v} of Eq. (3.22), which satisfies the following equation in HH for any tt\in\mathbb{R},

uv(t,ω)=12tSA(tr)x[uv(r,ω)]2dr+tSA(tr)B(uv(r,ω))v(r)dr.\textbf{u}^{*}_{v}(t,\omega)=\frac{1}{2}\int_{-\infty}^{t}S_{A}(t-r)\partial_{x}[\textbf{u}^{*}_{v}(r,\omega)]^{2}\mathrm{d}r+\int_{-\infty}^{t}S_{A}(t-r)B(\textbf{u}^{*}_{v}(r,\omega))v(r)\mathrm{d}r.

We define 𝒢0:C(;H0)C(;H)\mathcal{G}^{0}:C(\mathbb{R};H_{0})\longrightarrow C(\mathbb{R};H) by

𝒢0(v(s)ds):=uv().\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right):=\textbf{u}^{*}_{v}(\cdot).

By using Theorem 3.2, the family {uε:ε>0}\{\textbf{u}^{*}_{\varepsilon}:\varepsilon>0\} satisfies the LDP in C(;H)C(\mathbb{R};H) with rate function

I(f)=inf{vL2(;H0):f=𝒢0(v(s)ds)}{12+v(s)H02ds},I(f)=\inf_{\left\{v\in L^{2}(\mathbb{R};H_{0}):f=\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\right\}, (3.23)

where the infimum over an empty set is taken as ++\infty.

The following two examples show that we can also study the LDP for random periodic solutions and random quasi-periodic solutions. We give the definition of random periodic solution (see, for example, [13]).

Definition 3.3.

A random periodic solution of period TT of the random dynamical system Φ:+×\Phi:\mathbb{R}^{+}\times Ω×𝕏𝕏\Omega\times\mathbb{X}\rightarrow\mathbb{X} is an \mathcal{F}-measurable map Y:×Ω𝕏Y:\mathbb{R}\times\Omega\rightarrow\mathbb{X} such that for almost all ωΩ\omega\in\Omega,

Φ(t,θ(s,ω))Y(s,ω)=Y(t+s,ω),Y(s+T,ω)=Y(s,θ(T,ω)),\Phi(t,\theta(s,\omega))Y(s,\omega)=Y(t+s,\omega),Y(s+T,\omega)=Y(s,\theta(T,\omega)), (3.24)

for any t+,st\in\mathbb{R}^{+},s\in\mathbb{R}. It is called a random periodic solution with the minimal period TT if T>0T>0 is the smallest number such that (3.24) holds.

The following lemma is the result of the LDP for nn-dimensional Brownian motion.
For convenience, let EnE_{n} be the space C(;n)C(\mathbb{R};\mathbb{R}^{n}) with norm

Bn:=suptBtn1+|t|,\|B\|_{\mathcal{E}_{n}}:=\sup_{t\in\mathbb{R}}\frac{\|B_{t}\|_{\mathbb{R}^{n}}}{1+|t|}, (3.25)

where xn2:=i=1n|xi|2,x=(x1,x2,,xn)n\|x\|_{\mathbb{R}^{n}}^{2}:=\sum_{i=1}^{n}|x_{i}|^{2},x=(x_{1},x_{2},\cdots,x_{n})\in\mathbb{R}^{n}.

Lemma 3.4.

(cf. Schilder’s theorem in section 1.3 of [8]) nn-dimensional Brownian motion BB satisfies the LDP in space E1E_{1} with rate function

J(ϕ)=12+ϕ˙(s)n2ds,ϕC(;n),J(\phi)=\frac{1}{2}\int_{-\infty}^{+\infty}\|\dot{\phi}(s)\|_{\mathbb{R}^{n}}^{2}\mathrm{d}s,\quad\phi\in C(\mathbb{R};\mathbb{R}^{n}), (3.26)

where ϕ˙\dot{\phi} be the derivative of ϕ\phi with respect to time tt.

Example 2.

Consider the one-dimension equation shown by Feng, Liu and Zhao in [10],

dx=5xdt+(sin(x)+0.3sin(2πt))dt+εdBt,\mathrm{d}x=-5x\mathrm{d}t+(\sin(x)+0.3\sin(2\pi t))\mathrm{d}t+\sqrt{\varepsilon}\mathrm{d}B_{t}, (3.27)

where BtB_{t} is one-dimension Brownian motion.
According to [10], for every ε>0\varepsilon>0, the random periodic solution is

x~ε(t)=te5(ts)(sin(x~ε)+0.3sin(2πt))ds+εte5(ts)dBs.\tilde{x}_{\varepsilon}(t)=\int_{-\infty}^{t}e^{-5(t-s)}(\sin(\tilde{x}_{\varepsilon})+0.3\sin(2\pi t))\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}e^{-5(t-s)}\mathrm{d}B_{s}.

Let G1G_{1} be the space C(;)C(\mathbb{R};\mathbb{R}) with another norm

g𝒞1=k=12k(sups[k,k]g(s)1),gC(;).\|g\|_{\mathcal{C}_{1}}=\sum_{k=1}^{\infty}2^{-k}\left(\sup_{s\in[-k,k]}\left\|g(s)\right\|_{\mathbb{R}}\wedge 1\right),\quad g\in C(\mathbb{R};\mathbb{R}). (3.28)

Let x~:=x~ε=1\tilde{x}:=\tilde{x}_{\varepsilon=1}, and f:E1G1f:E_{1}\rightarrow G_{1} defined by

f(ω)(t):=te5(ts)(sin(x~)+0.3sin(2πt))ds+te5(ts)dω(s).f(\omega)(t):=\int_{-\infty}^{t}e^{-5(t-s)}(\sin(\tilde{x})+0.3\sin(2\pi t))\mathrm{d}s+\int_{-\infty}^{t}e^{-5(t-s)}\mathrm{d}\omega(s).

For any ω1,ω2E1\omega_{1},\omega_{2}\in E_{1}, then after simple calculation, we obtain that

|f(ω1)(t)f(ω2)(t)|\displaystyle|f(\omega_{1})(t)-f(\omega_{2})(t)|
\displaystyle\leq {|te5(tr)[sin(x~(r,ω1))sin(x~(r,ω2))]dr|\displaystyle\left\{\left|\int_{-\infty}^{t}e^{-5(t-r)}[\sin(\tilde{x}(r,\omega_{1}))-\sin(\tilde{x}(r,\omega_{2}))]\mathrm{d}r\right|\right.
+|te5(tr)dω1(r)te5(tr)dω2(r)|}\displaystyle+\left.\left|\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{1}(r)-\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{2}(r)\right|\right\}
\displaystyle\leq {|te5(tr){[x~(r,ω1)x~(r,ω2)]}dr|\displaystyle\left\{\left|\int_{-\infty}^{t}e^{-5(t-r)}\left\{[\tilde{x}(r,\omega_{1})-\tilde{x}(r,\omega_{2})]\right\}\mathrm{d}r\right|\right.
+|te5(tr)dω1(r)te5(tr)dω2(r)|},\displaystyle+\left.\left|\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{1}(r)-\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{2}(r)\right|\right\},

by using the Grownall inequality, there exists a constant CC such that

|F(ω1)(t)F(ω2)(t)|C|te5(tr)dω1(r)te5(tr)dω2(r)|.|F(\omega_{1})(t)-F(\omega_{2})(t)|\leq C\left|\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{1}(r)-\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{2}(r)\right|. (3.29)

It follows from the integration by parts that

|te5(tr)dω1(r)te5(tr)dω2(r)|\displaystyle\left|\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{1}(r)-\int_{-\infty}^{t}e^{-5(t-r)}\mathrm{d}\omega_{2}(r)\right|
=\displaystyle= |(ω1(t)ω2(t))5te5(tr)ω1(r)dr+5te5(tr)ω2(r)dr|\displaystyle\left|(\omega_{1}(t)-\omega_{2}(t))-5\int_{-\infty}^{t}e^{-5(t-r)}\omega_{1}(r)\mathrm{d}r+5\int_{-\infty}^{t}e^{-5(t-r)}\omega_{2}(r)\mathrm{d}r\right|
\displaystyle\leq |ω1(t)ω2(t)|+5te5(tr)|ω1(r)ω2(r)|dr,\displaystyle|\omega_{1}(t)-\omega_{2}(t)|+5\int_{-\infty}^{t}e^{5(t-r)}|\omega_{1}(r)-\omega_{2}(r)|\mathrm{d}r,

for any N+N\in\mathbb{N}^{+}, it implies that there exists a constant only depends on NN such that

supt[N,N]|te5(tr)𝑑ω1(r)te5(tr)𝑑ω2(r)|2C(N)ω1ω21.\sup_{t\in[-N,N]}\left|\int_{-\infty}^{t}e^{-5(t-r)}d\omega_{1}(r)-\int_{-\infty}^{t}e^{-5(t-r)}d\omega_{2}(r)\right|^{2}\leq C(N)\|\omega_{1}-\omega_{2}\|_{\mathcal{E}_{1}}. (3.30)

For any η>0\eta>0, we choose NN such that k=N+2kη2\sum_{k=N}^{+\infty}2^{-k}\leq\frac{\eta}{2}, it follows from (3.29) and (3.30) that

f(ω1)f(ω2)𝒞1\displaystyle\|f(\omega_{1})-f(\omega_{2})\|_{\mathcal{C}_{1}}\leq k=1N2k(sups[k,k]|f(ω1)(s)f(ω2)(s)|1)+k=N+2k\displaystyle\sum_{k=1}^{N}2^{-k}\left(\sup_{s\in[-k,k]}|f(\omega_{1})(s)-f(\omega_{2})(s)|\wedge 1\right)+\sum_{k=N}^{+\infty}2^{-k}
\displaystyle\leq sups[N,N]|f(ω1)(s)f(ω2)(s)|1+η2\displaystyle\sup_{s\in[-N,N]}|f(\omega_{1})(s)-f(\omega_{2})(s)|\wedge 1+\frac{\eta}{2}
\displaystyle\leq C(N)ω1ω2𝒞11+η2,\displaystyle C(N)\|\omega_{1}-\omega_{2}\|_{\mathcal{C}_{1}}\wedge 1+\frac{\eta}{2},

then there exists δη2C(N)\delta\leq\frac{\eta}{2C(N)} such that for any ω1ω21δ\|\omega_{1}-\omega_{2}\|_{\mathcal{E}_{1}}\leq\delta, it deduces that f(ω1)f(ω2)𝒞1η\|f(\omega_{1})-f(\omega_{2})\|_{\mathcal{C}_{1}}\leq\eta. It imply that ff is continuous, by using the contraction principle, we obtain that the random periodic solutions of Eq. (3.27) satisfies the LDP in space G1G_{1} with rate function

I(u)=12+|u˙(t)+5u(t)sin(u(t))0.3sin(2πt)|2𝑑t,I(u)=\frac{1}{2}\int_{-\infty}^{+\infty}\left|\dot{u}(t)+5u(t)-\sin(u(t))-0.3\sin(2\pi t)\right|^{2}dt,

where u˙\dot{u} be the derivative of uu with respect to time tt.

The following graphs with respect to the numerical of random periodic solutions of Eq. (3.27) with ε=0,0.01,1\varepsilon=0,0.01,1.
[Uncaptioned image] [Uncaptioned image] [Uncaptioned image]

The following example is to show that we can also study the LDP for random quasi-periodic solutions. We give the definition of random quasi-periodic solution (see, for example, [11]).

Definition 3.4.

Let F(ω,z)F({\omega},z) be a measurable map from Ω×Td\Omega\times T^{d} to HH, and ϕ(t,ω)\phi(t,\omega) be a random dynamical systems from HH to HH. Let α=(α1,,αd)\alpha=\left(\alpha_{1},\cdots,\alpha_{d}\right) be a d-dimension vector, which is rationally independent. Then, we say ϕ(t,ω)\phi(t,\omega) has a random quasi periodic solution F(ω,αt)F({\omega},\alpha\cdot t), if they satisfy
(i). (shift invariant of orbit) ϕ(t,ω)F(ω,z)=F(θ(t,ω),z)\phi(t,\omega)F(\omega,z)=F(\theta({t},\omega),z).
(ii). (quasi periodic property) ϕ(t,ω)F(ω,z)=F(θ(t,ω),αt+z)a.s.\phi(t,\omega)F(\omega,z)=F(\theta({t},\omega),\alpha\cdot t+z)\quad a.s.

Example 3.

We consider a random Hopf’s bifurcation model for turbulence. The original deterministic model given by Hopf in [15] is

ρ(x,t)t=I(ρ)+L(ρ)+μ2ρ(x,t)x2,\frac{\partial\rho(x,t)}{\partial t}=I(\rho)+L(\rho)+\mu\frac{\partial^{2}\rho(x,t)}{\partial x^{2}}, (3.31)

where ρ=u+iv\rho=u+\mathrm{i}v, and

L(ρ)=12π02πρ(x+y)F¯(y)dy,\displaystyle L(\rho)=\frac{1}{2\pi}\int_{0}^{2\pi}\rho(x+y)\bar{F}(y)\mathrm{d}y,
I(ρ)=14π202π02πρ(y)ρ(y)ρ¯(y+yx)dydy.\displaystyle I(\rho)=-\frac{1}{4\pi^{2}}\int_{0}^{2\pi}\int_{0}^{2\pi}\rho(y)\rho\left(y^{\prime}\right)\bar{\rho}\left(y+y^{\prime}-x\right)\mathrm{d}y\mathrm{d}y^{\prime}.

We denote u+iv¯=uiv\overline{u+\mathrm{i}v}=u-\mathrm{i}v. Here, FF is regarded as the external force acting on the “velocity field” ρ\rho. By the Fourier transform, we know that

ρ(x,t)=n=ρn(t)en(x), and F(x)=n=Fnen(x),\rho(x,t)=\sum_{n=-\infty}^{\infty}\rho_{n}(t){\rm e}_{n}(x),\text{ and }F(x)=\sum_{n=-\infty}^{\infty}F_{n}e_{n}(x),

where {en(x)=einx,xS1,n}\{e_{n}(x)=e^{{\rm i}nx},x\in S^{1},n\in\mathbb{Z}\} is the orthonormal basis of L2(S1)L^{2}(S^{1}), and L2(S1)L^{2}(S^{1}) with norm

hL2(S1):=n=|hn|,h=n=hnen(x)L2(S1).\|h\|_{L^{2}(S^{1})}:=\sum_{n=-\infty}^{\infty}|h_{n}|,\quad h=\sum_{n=-\infty}^{\infty}h_{n}{\rm e}_{n}(x)\in L^{2}(S^{1}).

So, we get

dρn(t)dt=ρn2ρn¯+(Fn¯μn2)ρn,Fn=anibn.\frac{\mathrm{d}\rho_{n}(t)}{dt}=-\rho_{n}^{2}\overline{\rho_{n}}+\left(\overline{F_{n}}-\mu n^{2}\right)\rho_{n},\quad F_{n}=a_{n}-\mathrm{i}b_{n}.

Thus, we consider the following equation

dρ(t)dt=ρ2ρ¯+(F¯ν)ρ,F=aib,\frac{\mathrm{d}\rho(t)}{dt}=-\rho^{2}\bar{\rho}+(\bar{F}-\nu)\rho,\quad F=a-\mathrm{i}b,

let ρ=reiθ\rho=re^{i\theta}, it is equivalent to

{dr(t)dt=(aνr2)r,dθ(t)dt=b,\left\{\begin{array}[]{l}\frac{\mathrm{d}r(t)}{\mathrm{d}t}=\left(a-\nu-r^{2}\right)r,\\ \frac{\mathrm{d}\theta(t)}{\mathrm{d}t}=b,\end{array}\right. (3.32)

where

ν=μn2,a=an=ReFn,b=bn=ImFn,r=rn,θ=θn,ρn=rneiθn.\displaystyle\nu=\mu n^{2},\quad a=a_{n}=\operatorname{Re}F_{n},\quad b=b_{n}=-\operatorname{Im}F_{n},\quad r=r_{n},\quad\theta=\theta_{n},\quad\rho_{n}=r_{n}e^{{\rm i}\theta_{n}}.

Now, we consider that ρ\rho is perturbed by random external force FF. For simplicity and illustrating our ideas, we assume that F=n(an+ibn)en+WF=\sum_{n}(a_{n}+{\rm i}b_{n})\mathrm{e}_{n}+W, where Wt(x)=ncnBn(t)en(x)W_{t}(x)=\sum_{n}c_{n}B_{n}(t)\mathrm{e}_{n}(x) and {Bn}\{B_{n}\} is a sequence independent Brownian motion, then Hopf’s model (3.31) or (3.33) with random external force is equivalent to the following stochastic equation

dr(t)dt=(aν+12c2r2)r+crdB(t),\displaystyle\frac{\mathrm{d}r(t)}{\mathrm{d}t}=\left(a-\nu+\frac{1}{2}c^{2}-r^{2}\right)r+cr\mathrm{d}B(t), (3.33)
dθ(t)=bdt,\displaystyle\mathrm{d}\theta(t)=b\mathrm{d}t, (3.34)

where

ν=μn2,a=an,b=bn,c=cn,r=rn,θ=θn,ρn=rneiθn,B(t)=Bn(t).\displaystyle\nu=\mu n^{2},\quad a=a_{n},\quad b=b_{n},\quad c=c_{n},\quad r=r_{n},\quad\theta=\theta_{n},\quad\rho_{n}=r_{n}e^{{\rm i}\theta_{n}},\quad B(t)=B_{n}(t).

For some more complex random Hopf’s models, please see [2].

The Eq. (3.33) has the non-trivial stationary solution

r(ω)=(0e2(aν)s+2cB(s)(ω)ds)12, if a>ν,r^{*}(\omega)=\left(\int_{-\infty}^{0}e^{2(a-\nu)s+2cB(s)(\omega)}\mathrm{d}s\right)^{-\frac{1}{2}},\quad\text{ if }a>\nu,

and unique trivial stationary solution

r(ω)=0, if aν.r^{*}(\omega)=0,\quad\text{ if }a\leq\nu.

Moreover, assume Fn=0,n0F_{n}=0,n\leq 0, and ann20(n>0)\frac{a_{n}}{n^{2}}\downarrow 0(n>0), b1,b2,,bn,b_{1},b_{2},\cdots,b_{n},\cdots is rationally independent. Let μn=ann2(n>0)\mu_{n}=\frac{a_{n}}{n^{2}}(n>0). Set θk=bkt+θk0\theta_{k}=b_{k}t+\theta^{0}_{k}, if μm>μμm+1\mu_{m}>\mu\geq\mu_{m+1} for some fixed m+m\in\mathbb{N}^{+}, then the solution ρ(x,t)\rho(x,t) of Eq. (3.31) converges to a random quasi-periodic solution ρ(θj,1jm;μ,ω)=j=1mrj(t,ω)ei(θj0+bjt)ej(x)\rho\left(\theta_{j},1\leq j\leq m;\mu,\omega\right)=\sum_{j=1}^{m}r_{j}^{*}(t,\omega)e^{{\rm i}(\theta^{0}_{j}+b_{j}t)}{\rm e}_{j}(x) with the angle variables θ\theta form a manifold of the type of a mm-d torus 𝕋m\mathbb{T}^{m}. Specially, we can choose cj=ε,1jmc_{j}=\sqrt{\varepsilon},1\leq j\leq m, cm+1=cm+2==0c_{m+1}=c_{m+2}=\cdots=0, ajj2μ=3212c2,1jma_{j}-j^{2}\mu=\frac{3}{2}-\frac{1}{2}c^{2},1\leq j\leq m, let

rε,j(t,ω)=1[2te(3ε)s+2εBj(s)(ω)ds]12,r_{\varepsilon,j}^{*}(t,\omega)=\frac{1}{\left[2\int_{-\infty}^{t}e^{(3-\varepsilon)s+2\sqrt{\varepsilon}B_{j}(s)(\omega)}\mathrm{d}s\right]^{\frac{1}{2}}},

where Bj(s)(ω)=ωj(s)B_{j}(s)(\omega)=\omega_{j}(s), ω=(ω1,ω1,,ωm)\omega=(\omega_{1},\omega_{1},\cdots,\omega_{m}), j=1,2,,mj=1,2,\cdots,m. Then for every ε>0\varepsilon>0, the solution ρε(x,t)\rho_{\varepsilon}(x,t) of Eq. (3.31) converges to a random quasi-periodic solution ρε:Ω×𝕋mL2(S1)\rho_{\varepsilon}:\Omega\times\mathbb{T}^{m}\rightarrow L^{2}{(S^{1})}

ρε(θj,1jm;μ,ω):=j=1mrε,j(t,ω)ei(θj0+bjt)ej(x),\rho_{\varepsilon}\left(\theta_{j},1\leq j\leq m;\mu,\omega\right):=\sum_{j=1}^{m}r_{\varepsilon,j}^{*}(t,\omega)e^{{\rm i}(\theta^{0}_{j}+b_{j}t)}{\rm e}_{j}(x),

where ρε(θj,1jm;μ,ω)L2(S1)=j=1m|rε,j(t,ω)|\|\rho_{\varepsilon}\left(\theta_{j},1\leq j\leq m;\mu,\omega\right)\|_{L^{2}{(S^{1})}}=\sum_{j=1}^{m}|r_{\varepsilon,j}^{*}(t,\omega)|.

Let G2G_{2} be the space C(;L2(S1))C(\mathbb{R};L^{2}(S^{1})) with another norm

h𝒞2=k=12k(sups[k,k]h(s)L2(S1)1),hC(;L2(S1)).\|h\|_{\mathcal{C}_{2}}=\sum_{k=1}^{\infty}2^{-k}\left(\sup_{s\in[-k,k]}\left\|h(s)\right\|_{L^{2}(S^{1})}\wedge 1\right),\quad h\in C(\mathbb{R};L^{2}(S^{1})). (3.35)

Let ri:=rε=1,i,i=1,2,,mr^{*}_{i}:=r^{*}_{\varepsilon=1,i},i=1,2,...,m, and g:EmG2g:E_{m}\rightarrow G_{2} defined by

g(ω)(t):=j=1mrj(t,ω)ei(θj0+bjt)ej(x),g(\omega)(t):=\sum_{j=1}^{m}r_{j}^{*}(t,\omega)e^{{\rm i}(\theta^{0}_{j}+b_{j}t)}{\rm e}_{j}(x),

where ω=(ω1,ω2,ωm)Em\omega=(\omega_{1},\omega_{2}\cdots,\omega_{m})\in E_{m}. We prove the function gg is continuous in the following.

For any ω1=(ω11,ω21,,ωm1),ω2=(ω12,ω22,,ωm2)Em\omega^{1}=(\omega_{1}^{1},\omega_{2}^{1},\cdots,\omega_{m}^{1}),\omega^{2}=(\omega_{1}^{2},\omega_{2}^{2},\cdots,\omega_{m}^{2})\in E_{m}, then after simple calculation, we obtain that

g(ω1)(t)g(ω2)(t)L2(S1)=j=1m|rj(t,ωj1)rj(t,ωj2)|.\displaystyle\|g(\omega^{1})(t)-g(\omega^{2})(t)\|_{L^{2}{(S^{1})}}=\sum_{j=1}^{m}\left|r_{j}^{*}(t,\omega_{j}^{1})-r_{j}^{*}(t,\omega_{j}^{2})\right|. (3.36)

For any η>0\eta>0, we choose NN big enough such that k=N+2kη2\sum_{k=N}^{+\infty}2^{-k}\leq\frac{\eta}{2}, it follows from Eqs. (3.35) and (3.36) that

g(ω1)g(ω2)𝒞2\displaystyle\|g(\omega_{1})-g(\omega_{2})\|_{\mathcal{C}_{2}}\leq k=1N2k(sups[k,k]g(ω1)(s)g(ω2)(s)L2(S1)1)+k=N+2k\displaystyle\sum_{k=1}^{N}2^{-k}\left(\sup_{s\in[-k,k]}\|g(\omega_{1})(s)-g(\omega_{2})(s)\|_{L^{2}{(S^{1})}}\wedge 1\right)+\sum_{k=N}^{+\infty}2^{-k}
\displaystyle\leq sups[N,N]g(ω1)(s)g(ω2)(s)L2(S1)1+k=N+2k\displaystyle\sup_{s\in[-N,N]}\|g(\omega_{1})(s)-g(\omega_{2})(s)\|_{L^{2}{(S^{1})}}\wedge 1+\sum_{k=N}^{+\infty}2^{-k}
\displaystyle\leq j=1msups[N,N](|rj(t,ωj1)rj(t,ωj2)|)1+η2.\displaystyle\sum_{j=1}^{m}\sup_{s\in[-N,N]}\left(\left|r_{j}^{*}(t,\omega_{j}^{1})-r_{j}^{*}(t,\omega_{j}^{2})\right|\right)\wedge 1+\frac{\eta}{2}.

Define (+)m(\mathbb{R}^{+})^{m} be the product space of +\mathbb{R}^{+} times mm, let f:Em(+)mf:E_{m}\rightarrow(\mathbb{R}^{+})^{m} be

f(ω)(t)=\displaystyle f(\omega)(t)= (f1(ω)(t),f2(ω)(t),,fm(ω)(t))\displaystyle(f_{1}(\omega)(t),f_{2}(\omega)(t),\cdots,f_{m}(\omega)(t))
:=\displaystyle:= (te2s+2B1(s)(ω)ds,te2s+2B2(s)(ω)ds,,te2s+2Bm(s)(ω)ds),\displaystyle\left(\int_{-\infty}^{t}e^{2s+2B_{1}(s)(\omega)}\mathrm{d}s,\int_{-\infty}^{t}e^{2s+2B_{2}(s)(\omega)}\mathrm{d}s,\cdots,\int_{-\infty}^{t}e^{2s+2B_{m}(s)(\omega)}\mathrm{d}s\right),

it is sufficient to prove that f|[N,N]f|_{[-N,N]} is a continuous function from EmE_{m} to space C([N,N];(+)m)C([-N,N];(\mathbb{R}^{+})^{m}) with norm fC([N,N];+×+):=j=1msupt[N,N]|fj(t)|\|f\|_{C([-N,N];\mathbb{R}^{+}\times\mathbb{R}^{+})}:=\sum_{j=1}^{m}\sup_{t\in[-N,N]}|f_{j}(t)|, for any N+N\in\mathbb{N}^{+}.

For any N+N\in\mathbb{N}^{+}, δ>0\delta>0, ω1Em\omega^{1}\in E_{m} satisfies ω1mK\|\omega^{1}\|_{\mathcal{E}_{m}}\leq K and ω1ω2mδ\|\omega^{1}-\omega^{2}\|_{\mathcal{E}_{m}}\leq\delta, we have

|fi(ω1)(t)fi(ω2)(t)|=|te2s+2ωi1(s)dste2s+2ωi2(s)ds|\displaystyle|f_{i}(\omega^{1})(t)-f_{i}(\omega^{2})(t)|=\left|\int_{-\infty}^{t}e^{2s+2\omega_{i}^{1}(s)}\mathrm{d}s-\int_{-\infty}^{t}e^{2s+2\omega_{i}^{2}(s)}\mathrm{d}s\right| (3.37)
=\displaystyle= |t0e2s+2ωi1(s)dst0e2s+2ωi2(s)ds|+|0t0e2s+2ωi1(s)ds0t0e2s+2ωi2(s)ds|\displaystyle\left|\int_{-\infty}^{t\wedge 0}e^{2s+2\omega_{i}^{1}(s)}\mathrm{d}s-\int_{-\infty}^{t\wedge 0}e^{2s+2\omega_{i}^{2}(s)}\mathrm{d}s\right|+\left|\int_{0}^{t\vee 0}e^{2s+2\omega_{i}^{1}(s)}\mathrm{d}s-\int_{0}^{t\vee 0}e^{2s+2\omega_{i}^{2}(s)}\mathrm{d}s\right|
:=\displaystyle:= Ii1+Ii2,i=1,2,,m.\displaystyle I_{i}^{1}+I_{i}^{2},\quad i=1,2,\cdots,m.

We first estimate Ii1,i=1,2,,mI_{i}^{1},i=1,2,\cdots,m, choose δ\delta small enough such that eK1+2δeK1+δ<1e^{K-1+2\delta}-e^{K-1+\delta}<1 then

Ii1=\displaystyle I_{i}^{1}= |t0e2s[exp{ωi1(s)(1s)}exp{ωi2(s)(1s)}]2(1s)ds|\displaystyle\left|\int_{-\infty}^{t\wedge 0}e^{2s}\left[\exp\left\{\frac{\omega_{i}^{1}(s)}{(1-s)}\right\}-\exp\left\{\frac{\omega_{i}^{2}(s)}{(1-s)}\right\}\right]^{2(1-s)}\mathrm{d}s\right| (3.38)
\displaystyle\leq |t0e2s[exp{ωi2(s)(1s)+δ}exp{ωi2(s)(1s)}]2(1s)ds|\displaystyle\left|\int_{-\infty}^{t\wedge 0}e^{2s}\left[\exp\left\{\frac{\omega_{i}^{2}(s)}{(1-s)}+\delta\right\}-\exp\left\{\frac{\omega_{i}^{2}(s)}{(1-s)}\right\}\right]^{2(1-s)}\mathrm{d}s\right|
\displaystyle\leq |t0e2s[eK+2δeK+δ]2(1s)ds|\displaystyle\left|\int_{-\infty}^{t\wedge 0}e^{2s}\left[e^{K+2\delta}-e^{K+\delta}\right]^{2(1-s)}\mathrm{d}s\right|
=\displaystyle= [eK+2δeK+δ]2|t0[eK1+2δeK1+δ]2sds|\displaystyle\left[e^{K+2\delta}-e^{K+\delta}\right]^{2}\left|\int_{-\infty}^{t\wedge 0}\left[e^{K-1+2\delta}-e^{K-1+\delta}\right]^{-2s}\mathrm{d}s\right|
=\displaystyle= 12(eK+2δeK+δ)2(eK1+2δeK1+δ)2(t0)[ln(eK1+2δeK1+δ)]1\displaystyle\frac{1}{2}\left(e^{K+2\delta}-e^{K+\delta}\right)^{2}\left(e^{K-1+2\delta}-e^{K-1+\delta}\right)^{-2(t\wedge 0)}\left[\ln\left(e^{K-1+2\delta}-e^{K-1+\delta}\right)\right]^{-1}
\displaystyle\leq 12(eK+2δeK+δ)2[ln(eK1+2δeK1+δ)]1.\displaystyle\frac{1}{2}\left(e^{K+2\delta}-e^{K+\delta}\right)^{2}\left[\ln\left(e^{K-1+2\delta}-e^{K-1+\delta}\right)\right]^{-1}.

Similar estimate for Ii2,i=1,2,,mI_{i}^{2},i=1,2,\cdots,m, choose δ\delta small enough such that eK+1+2δeK+1+δ<1e^{K+1+2\delta}-e^{K+1+\delta}<1, we get that

Ii2=\displaystyle I_{i}^{2}= |0t0e2s[exp{ωi1(s)(1+s)}exp{ωi2(s)(1+s)}]2(1+s)ds|\displaystyle\left|\int_{0}^{t\vee 0}e^{2s}\left[\exp\left\{\frac{\omega_{i}^{1}(s)}{(1+s)}\right\}-\exp\left\{\frac{\omega_{i}^{2}(s)}{(1+s)}\right\}\right]^{2(1+s)}\mathrm{d}s\right| (3.39)
\displaystyle\leq |0t0e2s[exp{ωi2(s)(1+s)+δ}exp{ωi2(s)(1+s)}]2(1+s)ds|\displaystyle\left|\int_{0}^{t\vee 0}e^{2s}\left[\exp\left\{\frac{\omega_{i}^{2}(s)}{(1+s)}+\delta\right\}-\exp\left\{\frac{\omega_{i}^{2}(s)}{(1+s)}\right\}\right]^{2(1+s)}\mathrm{d}s\right|
\displaystyle\leq |0t0e2s[eK+2δeK+δ]2(1+s)ds|\displaystyle\left|\int_{0}^{t\vee 0}e^{2s}\left[e^{K+2\delta}-e^{K+\delta}\right]^{2(1+s)}\mathrm{d}s\right|
=\displaystyle= [eK+2δeK+δ]2|0t0[eK+1+2δeK+1+δ]2sds|\displaystyle\left[e^{K+2\delta}-e^{K+\delta}\right]^{2}\left|\int_{0}^{t\vee 0}\left[e^{K+1+2\delta}-e^{K+1+\delta}\right]^{2s}\mathrm{d}s\right|
=\displaystyle= 12(eK+2δeK+δ)2[(eK+1+2δeK+1+δ)2(t0)1][ln(eK+1+2δeK+1+δ)]1\displaystyle\frac{1}{2}\left(e^{K+2\delta}-e^{K+\delta}\right)^{2}\left[\left(e^{K+1+2\delta}-e^{K+1+\delta}\right)^{2(t\wedge 0)}-1\right]\left[\ln\left(e^{K+1+2\delta}-e^{K+1+\delta}\right)\right]^{-1}
\displaystyle\leq 12(eK+2δeK+δ)2[ln(eK+1+2δeK+1+δ)]1.\displaystyle\frac{1}{2}\left(e^{K+2\delta}-e^{K+\delta}\right)^{2}\left[\ln\left(e^{K+1+2\delta}-e^{K+1+\delta}\right)\right]^{-1}.

It follows from (3.37), (3.38) and (3.39) that f|[N,N]f|_{[-N,N]} is a continuous function from EmE_{m} to space C([N,N];(+)m)C([-N,N];(\mathbb{R}^{+})^{m}), thus gg is a continuous function from EmE_{m} to G2G_{2}.

By using the contraction principle, we obtain that the random quasi-periodic solutions of Eq. (3.31) satisfies the LDP in space G2G_{2} with rate function

I(ρ)=inf{12+ϕ˙(s)m2ds:ϕEm,ρ=g(ϕ)}.\displaystyle I(\rho)=\inf\left\{\frac{1}{2}\int_{-\infty}^{+\infty}\|\dot{\phi}(s)\|_{\mathbb{R}^{m}}^{2}\mathrm{d}s:\quad\phi\in E_{m},\quad\rho=g(\phi)\right\}.

4 Invariant measure, rate function and quasi-potential

4.1 LDP for invariant measure

In this section, we will use the contraction principle to prove that the LDP for the family of stationary solutions {Xε}ε>0\{X^{*}_{\varepsilon}\}_{\varepsilon>0} deduce the LDP for the family of invariant measures {νε}ε>0\{\nu_{\varepsilon}\}_{\varepsilon>0} for (2.1).

It follows from Lemma 2.6 that there exists ε0>0\varepsilon_{0}>0 such that for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), (2.1) exists a unique stationary solution XεX^{*}_{\varepsilon}. Let νε\nu_{\varepsilon} be the distribution of the stationary solution Xε(,ω)X^{*}_{\varepsilon}(\cdot,\omega) at time 0, i.e.

νε(A):={ω;Xε(0,ω)A},A(H),\nu_{\varepsilon}(A):=\mathbb{P}\left\{\omega;X^{*}_{\varepsilon}(0,\omega)\in A\right\},\quad\forall A\in\mathcal{B}(H),

then νε\nu_{\varepsilon} is an invariant measure of Eq. (2.1) in HH. Moreover, Lemma 5.11 easily implies the invariant measure of Eq. (2.1) is unique in HH, for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}). Thus νε\nu_{\varepsilon} is the unique invariant measure of Eq. (2.1) in HH.

For any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), Xε(0,ω)X^{*}_{\varepsilon}(0,\omega) is an 0\mathcal{F}_{-\infty}^{0}-measurable random variable, therefore, in this section, we only consider the LDP of stationary solutions for Eq. (2.1) in space C((,0];H)C((-\infty,0];H) with the norm

C((,0];H)=k=12k(sups[k,0]H1).\|\cdot\|_{C((-\infty,0];H)}=\sum_{k=1}^{\infty}2^{-k}\left(\sup_{s\in[-k,0]}\left\|\cdot\right\|_{H}\wedge 1\right). (4.1)

Similar to the proof in Section 3, we could have the following theorem.

Theorem 4.1.

Under the Hypothesis 2.1, the family {Xε:ε>0}\{X^{*}_{\varepsilon}:\varepsilon>0\} satisfies the LDP in C((,0];H)C((-\infty,0];H) with good rate function

I~(f)=inf{vL2((,0];H0):f=𝒢0(v(s)ds)}{120v(s)H02ds},\tilde{I}(f)=\inf_{\left\{v\in L^{2}((-\infty,0];H_{0}):f=\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{0}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\right\}, (4.2)

where the infimum over an empty set is taken as ++\infty.

We could get the following Theorem by using the contraction principle (cf. Theorem 4.2.1 in [7]).

Theorem 4.2.

Under the Hypothesis 2.1, the family of invariant measures {νε}ε>0\left\{\nu_{\varepsilon}\right\}_{\varepsilon>0} for Eq. (2.1) satisfies the LDP in HH, with good rate function II^{\prime}

I(x)=inf{120v(s)H02ds:u=𝒢0(v(s)ds),u(0)=x}.I^{\prime}(x)=\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s:u=\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right),u(0)=x\right\}. (4.3)
Proof.

It follows from Theorem 4.1 that the stationary solution family {Xε(,ω)}ε>0\left\{X^{*}_{\varepsilon}(\cdot,\omega)\right\}_{\varepsilon>0} for Eq. (2.1) satisfies the LDP in C((,0];H)C((-\infty,0];H) with rate function I~\tilde{I}.

Let G:C((,0];H)HG:C((-\infty,0];H)\longrightarrow H by G(f)=f(0)G(f)=f(0), it is obvious that GG is continuous, then it follows from the contraction principle that the invariant measure family {νε}ε>0\left\{\nu_{\varepsilon}\right\}_{\varepsilon>0} of Eq. (2.1) satisfies the LDP in HH, with rate function

I(x)=\displaystyle I^{\prime}(x)= inf{x=G(u)}{I~(u)}=inf{120v(s)H02ds:u=𝒢0(v(s)ds),u(0)=x}.\displaystyle\inf_{\{x=G(u)\}}\{\tilde{I}(u)\}=\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s:u=\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right),u(0)=x\right\}.

4.2 Rate function and quasi-potential

It is well known that [14] and [6] have proved the family of invariant measures for stochastic equations satisfies LDP with the quasi-potential as rate function. The definition of quasi-potential is given in Section 2.3, we will prove the rate function II^{\prime} defined in (4.3) and quasi-potential are equivalent below.

Lemma 4.1.

The rate function (4.3) defined in Theorem 4.2 are equivalent to the rate function (2.10) defined by quasi-potential of the LDP for invariant measures of Eq. (2.1).

Proof.

It follows from the definition of 𝒢0\mathcal{G}^{0} and (2.10) that

I(x)=\displaystyle I^{\prime}(x)= inf{120v(s)H02ds:u=𝒢0(v(s)ds),u(0)=x}\displaystyle\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s:u=\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right),u(0)=x\right\} (4.4)
=\displaystyle= inf{120v(t)H02dt;uC((,0];H),u=u(v),u(0)=x,limtu(t)H=0}\displaystyle\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\left\|v(t)\right\|_{H_{0}}^{2}\mathrm{d}t;u\in C((-\infty,0];H),u=u(v),u(0)=x,\lim_{t\rightarrow\infty}\|u(t)\|_{H}=0\right\}
=\displaystyle= inf{S(u);uC((,0];H),u(0)=x,limtu(t)H=0}\displaystyle\inf\left\{S_{-\infty}(u);u\in C((-\infty,0];H),u(0)=x,\lim_{t\rightarrow-\infty}\|u(t)\|_{H}=0\right\}
=\displaystyle= V(x),xH,V(x)<.\displaystyle V(x),\quad\forall x\in H,\ V(x)<\infty.

Remark 4.1.

Theorem 4.2 states that the LDP for the family of stationary solutions can deduce the LDP for the family of invariant measures of Eq. (2.1). And for the rate function II^{\prime} consistent with quasi-potential VV, II^{\prime} defined in Theorem 4.2 gives another explain of quasi-potential.

The next example illustrate that there exists two systems with different dynamical behaviors but with the same invariant measure. It implies that the LDP for the stationary solutions gives more dynamical information than the LDP of invariant measures.

Example 4.

Define two 2×22\times 2 metrics A1A_{1} and A2A_{2} by

A1=[λ00λ],A2=[λββλ].A_{1}=\begin{bmatrix}-\lambda&0\\ 0&-\lambda\end{bmatrix},\quad A_{2}=\begin{bmatrix}-\lambda&-\beta\\ \beta&-\lambda\end{bmatrix}.

We consider two stochastic equations

dX=A1Xdt+εdBt,\mathrm{d}X=A_{1}X\mathrm{d}t+\sqrt{\varepsilon}\mathrm{d}B_{t}, (4.5)

and

dX=A2Xdt+εdBt.\mathrm{d}X=A_{2}X\mathrm{d}t+\sqrt{\varepsilon}\mathrm{d}B_{t}. (4.6)

It follows from Theorem 4.2 that the family of invariant measures of Eqs. (4.5) and (4.6) satisfies LDP with good rate function as follows

Vi(X)=inf{120Y˙(t)AiY(t)22dt:Y(0)=X,limtY(t)2=0},i=1,2,V_{i}(X)=\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\left\|\dot{Y}(t)-A_{i}Y(t)\right\|_{\mathbb{R}^{2}}^{2}\mathrm{d}t:Y(0)=X,\lim_{t\rightarrow-\infty}\|Y(t)\|_{\mathbb{R}^{2}}=0\right\},\quad i=1,2,

where X=(x1,x2)X=(x_{1},x_{2}) with norm X22:=x12+x22\|X\|_{\mathbb{R}^{2}}^{2}:=x_{1}^{2}+x_{2}^{2}, and Y˙\dot{Y} be the derivative of YY with respect to time tt.
By using Theorem 3.1 in [14], we can get that

Vi(X)=λ(x12+x22),i=1,2.V_{i}(X)=\lambda(x_{1}^{2}+x_{2}^{2}),\quad i=1,2.

It implies that the rate functions for invariant measures of two Eqs. (4.5) and (4.6) are equivalent. Furthermore, let Zε:=2eλ(x12+x22)εdXZ_{\varepsilon}:=\int_{\mathbb{R}^{2}}e^{-\frac{\lambda(x_{1}^{2}+x_{2}^{2})}{\varepsilon}}\mathrm{d}X, the two Eqs. (4.5) and (4.6) have the same invariant measure νε()=1Zεeλ(x12+x22)εdX\nu_{\varepsilon}(\cdot)=\frac{1}{Z_{\varepsilon}}\int_{\cdot}e^{-\frac{\lambda(x_{1}^{2}+x_{2}^{2})}{\varepsilon}}\mathrm{d}X, for every ε\varepsilon. However, we have known that the determinate equations dX=A1Xdt\mathrm{d}X=A_{1}X\mathrm{d}t and dX=A2Xdt\mathrm{d}X=A_{2}X\mathrm{d}t with respect to Eqs. (4.5) and (4.6) have different asymptotic behavior.

Define the stationary solutions of Eqs. (4.5) and (4.6) by

Xi,ε()=εeAi(s)dBs,i=1,2.X^{*}_{i,\varepsilon}(\cdot)=\sqrt{\varepsilon}\int_{-\infty}^{\cdot}e^{A_{i}(\cdot-s)}\mathrm{d}B_{s},\quad i=1,2.

It follows from Theorem 2.10 that the family of the stationary solutions {Xi,ε}ε>0\{X^{*}_{i,\varepsilon}\}_{\varepsilon>0} of Eqs. (4.5) and (4.6) satisfies LDP in space C(;2)C(\mathbb{R};\mathbb{R}^{2}) with different rate function

Ii(Y)=12+Y˙(t)AiY(t)22dt,YC(;2),i=1,2.I_{i}(Y)=\frac{1}{2}\int_{-\infty}^{+\infty}\left\|\dot{Y}(t)-A_{i}Y(t)\right\|_{\mathbb{R}^{2}}^{2}\mathrm{d}t,Y\in C(\mathbb{R};\mathbb{R}^{2}),\quad i=1,2.

It implies that it is meaningful to research the LDP for stationary solutions, which gives more dynamical information. The numerical simulation in the following graphs will give a more intuitive explanation. We choose λ=0.3,β=2\lambda=0.3,\beta=2, the following three graphs are the numerical approximate of solutions for Eq. (4.5) correspond to the cases ε=0,ε=0.01,ε=1\varepsilon=0,\varepsilon=0.01,\varepsilon=1.
[Uncaptioned image] [Uncaptioned image] [Uncaptioned image]
The following three graphs are the numerical approximate of solutions for Eq. (4.5) correspond to the cases ε=0,ε=0.01,ε=1\varepsilon=0,\varepsilon=0.01,\varepsilon=1.
[Uncaptioned image] [Uncaptioned image] [Uncaptioned image]

Brzeźniak and Cerrai have researched the LDP for the invariant measures of the 22-dimensional stochastic Navier-Stokes equations on a torus in [4]. We will consider the LDP for the solutions of the pullback integral equations of the 22-dimensional stochastic Navier-Stokes equations in the following.

Example 5.

For convenience, we write 22-dimensional stochastic Navier-Stokes equations perturbed by a small additive noise in a functional form shown by Brzeźniak and Cerrai in [4], as

du(t)+Au(t)dt+B(u(t),u(t))dt=εdwt,u(0)=u0,\mathrm{d}\textbf{u}(t)+\mathrm{A}\textbf{u}(t)\mathrm{d}t+\mathrm{B}(\textbf{u}(t),\textbf{u}(t))\mathrm{d}t=\sqrt{\varepsilon}\mathrm{d}w_{t},\quad\textbf{u}(0)=\textbf{u}_{0}, (4.7)

for 0<ε<<10<\varepsilon<<1 on a two-dimensional torus 𝕋2\mathbb{T}^{2}.

Let us recall that AA is the Stokes operator, roughly speaking, equal to the Laplace operator composed with the Leray-Helmholtz projection PP, the convection B(u,u)\mathrm{B}(\textbf{u},\textbf{u}) is equal to P(uu)P(\textbf{u}\nabla\textbf{u}), and w(t)w(t) is a QQ Wiener process.
The space H\mathrm{H} is defined by

H={uL2(𝕋2),𝕋2u(x)dx=0}.\mathrm{H}=\left\{\textbf{u}\in L^{2}\left(\mathbb{T}^{2}\right),\quad\int_{\mathbb{T}^{2}}\textbf{u}(x)\mathrm{d}x=0\right\}.

We also define the space V\mathrm{V} by setting

V={uH:DjuL2(𝕋2,2)},\mathrm{V}=\left\{\textbf{u}\in\mathrm{H}:D_{j}\textbf{u}\in L^{2}\left(\mathbb{T}^{2},\mathbb{R}^{2}\right)\right\},

where Dj,j=1,2D_{j},j=1,2, are the 1st order weak derivatives of the torus.

The proof is similar to the proof of Mattingly in [21] and Appendix 5.2, that there exists ε0>0\varepsilon_{0}>0 such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), the solution of the following pullback integral equation exists and is unique.

uε(r)=rSA(ts)B(uε(s),uε(s))ds+εrSA(ts)dws.\textbf{u}_{\varepsilon}^{*}(r)=\int_{-\infty}^{r}S_{A}(t-s)B\left(\textbf{u}_{\varepsilon}^{*}(s),\textbf{u}_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{r}S_{A}(t-s)\mathrm{d}w_{s}. (4.8)

Similar to Example 1, it follows from Lemma 5.5 and Lemma 5.4, for any vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), that there exits a unique solution uv\textbf{u}^{*}_{v} of skeleton Eq. (4.9), which satisfies the following equation in HH for any tt\in\mathbb{R},

uv(t,ω)=tSA(tr)B(uv(s),uv(s))dr+tSA(tr)v(r)dr.\textbf{u}^{*}_{v}(t,\omega)=\int_{-\infty}^{t}S_{A}(t-r)B\left(\textbf{u}^{*}_{v}(s),\textbf{u}^{*}_{v}(s)\right)\mathrm{d}r+\int_{-\infty}^{t}S_{A}(t-r)v(r)\mathrm{d}r. (4.9)

We define 𝒢0:C(;H0)C(;H)\mathcal{G}^{0}:C(\mathbb{R};H_{0})\longrightarrow C(\mathbb{R};H) by

𝒢0(v(s)ds):=uv().\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right):=\textbf{u}^{*}_{v}(\cdot).

By using Theorem 3.2, the family {uε:ε>0}\{\textbf{u}^{*}_{\varepsilon}:\varepsilon>0\} satisfies the LDP in C(;H)C(\mathbb{R};H) with rate function

I(f)=inf{vL2(;H0):f=𝒢0(v(s)ds)}{12+v(s)H02ds},I(f)=\inf_{\left\{v\in L^{2}(\mathbb{R};H_{0}):f=\mathcal{G}^{0}(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s\right\},

where the infimum over an empty set is taken as ++\infty.

For any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), although we have not proved the solutions of stochastic Navier-Stokes Eq. (4.7) form a CkC^{k} perfect cocycle, the distribution of the solution for Eq. (4.8) is also the unique invariant measure of Eq. (4.7). Moreover, it follows from Theorem 4.2 that the family of invariant measures of Eq. (4.7) satisfies LDP with rate function

I(x)=inf{120v(s)H02ds:u=𝒢0(v(s)ds),u(0)=x}.\displaystyle I^{\prime}(x)=\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s:\textbf{u}=\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right),\textbf{u}(0)=x\right\}. (4.10)

It follows from the definition of 𝒢0\mathcal{G}^{0} that uv()\textbf{u}^{*}_{v}(\cdot) is a very weak solution (cf. Definition 3.4 in [4]) of Eq. (4.7). And similar to the Definition 3.6 of [4], we define (u)\mathcal{H}(\textbf{u}) by

[(u)](t):=u˙v(t)+Auv(t)+B(uv(t),uv(t)),t,[\mathcal{H}(\textbf{u})](t):=\dot{\textbf{u}}^{*}_{v}(t)+\mathrm{A}\textbf{u}^{*}_{v}(t)+\mathrm{B}(\textbf{u}^{*}_{v}(t),\textbf{u}^{*}_{v}(t)),\quad t\in\mathbb{R}, (4.11)

where u˙v(t)\dot{\textbf{u}}^{*}_{v}(t) be the derivative of uv\textbf{u}^{*}_{v} with respect to time tt.
And

S(u):=120(u)(s)H02ds.S_{-\infty}(\textbf{u}):=\frac{1}{2}\int_{-\infty}^{0}\|\mathcal{H}(\textbf{u})(s)\|_{H_{0}}^{2}\mathrm{d}s. (4.12)

Combining Eqs. (4.10), (4.11) and (4.12), we get that

I(x)=inf{120(u)(s)H02ds:u(0)=x}=inf{S(u):u(0)=x},I^{\prime}(x)=\inf\left\{\frac{1}{2}\int_{-\infty}^{0}\|\mathcal{H}(\textbf{u})(s)\|_{H_{0}}^{2}\mathrm{d}s:\textbf{u}(0)=x\right\}=\inf\Big{\{}S_{-\infty}(\textbf{u}):\textbf{u}(0)=x\Big{\}},

which is equal to the rate function defined by quasi-potential in (4.13) of [4].

5 Proofs of the well-posedness of skeleton equation in infinite intervals and stationary solution

This section is divided into two subsections, in the first subsection we prove Theorem 3.1 about the well-posedness of the skeleton Eq. (3.1). In the second subsection we prove Theorem 2.6 about the existence and uniqueness of stationary solution XεX^{*}_{\varepsilon} for Eq. (2.1).

5.1 The well-posedness of the skeleton equation

Similar to Theorem 4.4 of Sritharan and Sundar [25], we could have the following lemma.

Lemma 5.1.

For any T>0T>0, X(0)HX(0)\in H and vSMv\in S_{M}, for some M<M<\infty, under Hypothesis 2.1, there exists a unique mild solution XX of Eq. (3.1) with initial value X(0)X(0) such that

X(t)=SA(t)X(0)+0tSA(tr)F(X(r))dr+0tSA(tr)B(X(r))v(r)dr,t[0,T]X(t)=S_{A}(t)X(0)+\int_{0}^{t}S_{A}(t-r)F(X(r))\mathrm{d}r+\int_{0}^{t}S_{A}(t-r)B(X(r))v(r)\mathrm{d}r,\quad\forall t\in[0,T]

in space C([0,T];H)C([0,T];H).

Before proving Theorem 3.1, we illustrate some lemmas below.

Lemma 5.2.

(Priori estimate of skeleton equation) For any t0t_{0}\in\mathbb{R}, X(t0)HX(t_{0})\in H, vSMv\in S_{M}, for some M<M<\infty, under the Hypothesis 2.1, there exists a constant CC, which only depend on X(t0)X(t_{0}), MM, λ\lambda, C1C_{1}, DD such that the unique mild solution XX of Eq. (3.1) with initial value X(0)X(0) at time t0t_{0} satisfies

X(t)H2+λt0tXV2dsC,tt0,\|X(t)\|_{H}^{2}+\lambda\int_{t_{0}}^{t}\|X\|_{V}^{2}\mathrm{d}s\leq C,\quad\forall t\geq t_{0},

where λ\lambda is defined in Hypothesis 2.1 (i), C1C_{1} and DD are defined in Remark 2.1 (i) and (ii) respectively.

Proof.

It follows from the Hypothesis 2.1 (i), Remark 2.1 (ii) and Young inequality,

dXH2dt=2X,dXdtH=2VAX+F(X),XV+2B(X)v,XH\displaystyle\frac{\mathrm{d}\|X\|_{H}^{2}}{\mathrm{d}t}=2\Big{\langle}X,\frac{\mathrm{d}X}{\mathrm{d}t}\Big{\rangle}_{H}=2_{V^{*}}\langle AX+F(X),X\rangle_{V}+2\langle B(X)v,X\rangle_{H}
\displaystyle\leq 2λXV2+2|B(X)|LQvH0XH2λXV2+2DvH0XH\displaystyle-2\lambda\|X\|_{V}^{2}+2|B(X)|_{L_{Q}}\|v\|_{H_{0}}\|X\|_{H}\leq-2\lambda\|X\|_{V}^{2}+2D\|v\|_{H_{0}}\|X\|_{H}
\displaystyle\leq λXV2+(δλC1)XH2+D2δvH02,\displaystyle-\lambda\|X\|_{V}^{2}+(\delta-\lambda C_{1})\|X\|_{H}^{2}+\frac{D^{2}}{\delta}\|v\|_{H_{0}}^{2},

by taking integral and since vSMv\in S_{M}, it implies that

X(t)H2+λt0tXV2ds\displaystyle\|X(t)\|_{H}^{2}+\lambda\int_{t_{0}}^{t}\|X\|_{V}^{2}\mathrm{d}s\leq X(t0)H2+(δλC1)t0tXH2ds+D2δ+vH02ds\displaystyle\|X(t_{0})\|_{H}^{2}+(\delta-\lambda C_{1})\int_{t_{0}}^{t}\|X\|_{H}^{2}\mathrm{d}s+\frac{D^{2}}{\delta}\int_{-\infty}^{+\infty}\|v\|_{H_{0}}^{2}\mathrm{d}s
\displaystyle\leq X(t0)H2+D2Mδ+(δλC1)t0tXH2ds.\displaystyle\|X(t_{0})\|_{H}^{2}+\frac{D^{2}M}{\delta}+(\delta-\lambda C_{1})\int_{t_{0}}^{t}\|X\|_{H}^{2}\mathrm{d}s.

By using the Gronwall inequality, we then obtain

X(t)H2+λt0tXV2ds\displaystyle\|X(t)\|_{H}^{2}+\lambda\int_{t_{0}}^{t}\|X\|_{V}^{2}\mathrm{d}s\leq (X(t0)H2+D2Mδ)exp{(δλC1)(tt0)}.\displaystyle\Big{(}\left\|X(t_{0})\right\|_{H}^{2}+\frac{D^{2}M}{\delta}\Big{)}\exp\{(\delta-\lambda C_{1})(t-t_{0})\}.

After choosing δ\delta small enough such that δ<λC1\delta<\lambda C_{1}, then

X(t)H2+t0tXV2dsX(t0)H2+D2Mδ.\|X(t)\|_{H}^{2}+\int_{t_{0}}^{t}\|X\|_{V}^{2}\mathrm{d}s\leq\left\|X(t_{0})\right\|_{H}^{2}+\frac{D^{2}M}{\delta}.

The following lemmas are aim to study the asymptotic stability of dynamical systems, which is the basis of the definition 𝒢0\mathcal{G}^{0}. For any M>0M>0, t0t_{0}\in\mathbb{R}, tt0t\geq t_{0}, X0,X^0HX_{0},\hat{X}_{0}\in H and vSMv\in S_{M}. Let X(t,t0;ω)X0X(t,t_{0};\omega)X_{0} and X^(t,t0;ω)X^0\hat{X}(t,t_{0};\omega)\hat{X}_{0} denote the solutions of Eq. (3.1) starting at different initial value X0X_{0} and X^0\hat{X}_{0} at time t0t_{0}, respectively. For convenience, let ρ(t,t0;X0,X^0)=X(t,t0;ω)X0X^(t,t0;ω)X^0\rho(t,t_{0};X_{0},\hat{X}_{0})=X(t,t_{0};\omega)X_{0}-\hat{X}(t,t_{0};\omega)\hat{X}_{0}.

Lemma 5.3.

Let Γ(l,t0;X)=λC1ηC0(1lt0t0+lX(s)V2ds)\Gamma(l,t_{0};X)=\lambda C_{1}-\eta-C_{0}(\frac{1}{l}\int_{t_{0}}^{t_{0}+l}\left\|X(s)\right\|_{V}^{2}\mathrm{d}s), then

ρ(t,t0;X0,X^0)H2e2Γ(tt0,t0;X)(tt0)(X0X^0H2+2D2Mη),\left\|\rho(t,t_{0};X_{0},\hat{X}_{0})\right\|_{H}^{2}\leq e^{-2\Gamma(t-t_{0},t_{0};X)(t-t_{0})}\Big{(}\Big{\|}X_{0}-\hat{X}_{0}\Big{\|}_{H}^{2}+\frac{2D^{2}M}{\eta}\Big{)},

where λ\lambda and C0C_{0} are defined in Hypothesis 2.1 (i), C1C_{1} and DD are defined in Remark 2.1 (i) and (ii) respectively.

Proof.

It follows from the Hypothesis 2.1 (i), Remark 2.1 (i), (ii) and Young inequality,

12ddtρ(t)H2\displaystyle\frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}\left\|\rho(t)\right\|_{H}^{2} =Vρ,AXAX^+F(X)F(X^)V+B(X)vB(X^)v,ρH\displaystyle=_{V}\langle\rho,AX-A\hat{X}+F(X)-F(\hat{X})\rangle_{V^{*}}+\langle B(X)v-B(\hat{X})v,\rho\rangle_{H}
λρV2+C0ρH2XV2+2DvH0ρH\displaystyle\leq-\lambda\left\|\rho\right\|_{V}^{2}+C_{0}\left\|\rho\right\|_{H}^{2}\left\|X\right\|_{V}^{2}+2D\|v\|_{H_{0}}\|\rho\|_{H}
λC1ρH2+C0ρH2XV2+ηρH2+D2ηvH02\displaystyle\leq-\lambda C_{1}\left\|\rho\right\|_{H}^{2}+C_{0}\left\|\rho\right\|_{H}^{2}\left\|X\right\|_{V}^{2}+\eta\|\rho\|_{H}^{2}+\frac{D^{2}}{\eta}\|v\|_{H_{0}}^{2}
[λC1C0XV2η]ρH2+D2ηvH02,\displaystyle\leq-\Big{[}\lambda C_{1}-C_{0}\left\|X\right\|_{V}^{2}-\eta\Big{]}\left\|\rho\right\|_{H}^{2}+\frac{D^{2}}{\eta}\|v\|_{H_{0}}^{2},

after taking integral, it follows that

ρ(t)H2ρ(0)H2+2D2η+vH02ds2t0t[λC1C0XV2η]ρH2ds,\left\|\rho(t)\right\|_{H}^{2}\leq\left\|\rho(0)\right\|_{H}^{2}+\frac{2D^{2}}{\eta}\int_{-\infty}^{+\infty}\|v\|_{H_{0}}^{2}\mathrm{d}s-2\int_{t_{0}}^{t}\Big{[}\lambda C_{1}-C_{0}\left\|X\right\|_{V}^{2}-\eta\Big{]}\left\|\rho\right\|_{H}^{2}\mathrm{d}s,

by the Gronwall inequality, we can get the desired result. ∎

Corollary 5.1.

For any λ0(0,λC1)\lambda_{0}\in(0,\lambda C_{1}), there exists a constant N0N_{0} such that for any tt0>N0t-t_{0}>N_{0},

X(t,t0;X0)X^(t,t0;X^0)H2=ρ(t,t0;X0,X^0)H2e2λ0(tt0)(X0X^0H2+2D2Mη),\left\|X(t,t_{0};X_{0})-\hat{X}(t,t_{0};\hat{X}_{0})\right\|_{H}^{2}=\left\|\rho(t,t_{0};X_{0},\hat{X}_{0})\right\|_{H}^{2}\leq e^{-2\lambda_{0}(t-t_{0})}\Big{(}\left\|X_{0}-\hat{X}_{0}\right\|_{H}^{2}+\frac{2D^{2}M}{\eta}\Big{)},

which imply that

limt+X(t,t0;X0)X^(t,t0;X^0)H=0.\lim_{t\rightarrow+\infty}\left\|X(t,t_{0};X_{0})-\hat{X}(t,t_{0};\hat{X}_{0})\right\|_{H}=0.
Proof.

Combining Lemma 5.2 and Lemma 5.3, we could choose η\eta small enough and N0N_{0} big enough such that Γ(tt0,t0;X)>λ0,tt0>N0\Gamma(t-t_{0},t_{0};X)>\lambda_{0},\forall t-t_{0}>N_{0}, it implies the conclusion. ∎

We will give the following two lemmas to prove there exist a unique solution of the backward infinite horizon integral equation for Eq. (3.1). And then we define 𝒢0\mathcal{G}^{0} as the unique solution of the backward infinite horizon integral equation for Eq. (3.1).

Lemma 5.4.

For any vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), assume that for any N+,Y(t,ω)|t[N,N]C([N,N];H).N\in\mathbb{Z}^{+},Y(t,\omega)|_{t\in[-N,N]}\in C([-N,N];H). Moreover, if Y(t,ω)Y(t,\omega) satisfies the following equation in HH for any tt\in\mathbb{R},

Y(t,ω)=tSA(tr)F(Y(r,ω))dr+tSA(tr)B(Y(r,ω))v(r)dr,Y(t,\omega)=\int_{-\infty}^{t}S_{A}(t-r)F(Y(r,\omega))\mathrm{d}r+\int_{-\infty}^{t}S_{A}(t-r)B(Y(r,\omega))v(r)\mathrm{d}r, (5.1)

and

suptY(t)H2<,\sup_{t\in\mathbb{R}}\left\|Y(t)\right\|_{H}^{2}<\infty, (5.2)

then Y(t)Y(t) is unique.

Proof.

For any t<tt^{\prime}<t, it follows from Eq. (5.1) that

Y(t,ω)=SA(tt)Y(t,ω)+ttSA(tr)F(Y(r,ω))dr+ttSA(tr)B(Y(r,ω))v(r)dr.Y(t,\omega)=S_{A}(t-t^{\prime})Y(t^{\prime},\omega)+\int_{t^{\prime}}^{t}S_{A}(t-r)F(Y(r,\omega))\mathrm{d}r+\int_{t^{\prime}}^{t}S_{A}(t-r)B(Y(r,\omega))v(r)\mathrm{d}r.

Therefore, for any t>tt>t^{\prime}\in\mathbb{R}, Y(,ω)Y(\cdot,\omega) is the mild solution of Eq. (3.1) with initial value Y(t,ω)Y(t^{\prime},\omega). We will show the uniqueness of Eq. (5.1). Assume Y(,ω)Y(\cdot,\omega) and Z(,ω)Z(\cdot,\omega) are two solutions of Eq. (5.1), then for any n+,n<tn\in\mathbb{Z}^{+},-n<t,

Y(t,ω)=SA(t+n)Y(n,ω)+ntSA(tr)F(Y(r,ω))dr+ntSA(tr)B(Y(r,ω))v(r)dr,\displaystyle Y(t,\omega)=S_{A}(t+n)Y(-n,\omega)+\int_{-n}^{t}S_{A}(t-r)F(Y(r,\omega))\mathrm{d}r+\int_{-n}^{t}S_{A}(t-r)B(Y(r,\omega))v(r)\mathrm{d}r,
Z(t,ω)=SA(t+n)Z(n,ω)+ntSA(tr)F(Z(r,ω))dr+ntSA(tr)B(Z(r,ω))v(r)dr.\displaystyle Z(t,\omega)=S_{A}(t+n)Z(-n,\omega)+\int_{-n}^{t}S_{A}(t-r)F(Z(r,\omega))\mathrm{d}r+\int_{-n}^{t}S_{A}(t-r)B(Z(r,\omega))v(r)\mathrm{d}r.

It implies that Y(,ω)Y(\cdot,\omega) and Z(,ω)Z(\cdot,\omega) are mild solutions of Eq. (3.1) with initial value Y(n,ω)Y(-n,\omega) and Z(n,ω)Z(-n,\omega) at time n-n, then by using Corollary 5.1 and assumption suptY(t)H2<\sup_{t\in\mathbb{R}}\left\|Y(t)\right\|_{H}^{2}<\infty and suptZ(t)H2\sup_{t\in\mathbb{R}}\left\|Z(t)\right\|_{H}^{2} <<\infty, let nn\rightarrow\infty, we can have

Y(l,ω)Z(l,ω)H20,l>0.\left\|Y(l,\omega)-Z(l,\omega)\right\|_{H}^{2}\rightarrow 0,\quad\forall l>0.

The uniqueness have been proved. ∎

Finally, we will construct the solution of (5.1). For any n+n\in\mathbb{Z}^{+} and fixed vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), for convenience we set Xv(t,s;Xs)X_{v}(t,s;X_{s}) be the mild solution of (3.1) with initial value XsX_{s} at initial time ss, let

Xvn(t)={Xv(t,n;0),t>n,0,tn.X_{v}^{n}(t)=\begin{cases}X_{v}(t,-n;0),&t>-n,\\ 0,&t\leq-n.\\ \end{cases}
Lemma 5.5.

Assume that vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), for any N+N\in\mathbb{Z}^{+}, Xvn()Xv()X_{v}^{n}(\cdot)\rightarrow X_{v}^{*}(\cdot) in C([N,N];H)C([-N,N];H) as nn\rightarrow\infty. Moreover, XvX_{v}^{*} satisfies the backward infinite horizon integral Eq. (5.1) and (5.2).

Proof.

It follows from Corollary 5.1 that XvnX_{v}^{n} is a Cauchy sequence in C([N,N];H)C([-N,N];H). Since the space C([N,N];H)C([-N,N];H) is complete, there exists XvX_{v}^{*} such that limnXvn=Xv\lim\limits_{n\rightarrow\infty}X_{v}^{n}=X_{v}^{*} in C([N,N];H)C([-N,N];H). For NN is arbitrary, Xv()X_{v}^{*}(\cdot) is defined for all time, and from Lemma 5.2, we have supnsuptXvn(t)H2<\sup_{n}\sup_{t\in\mathbb{R}}\left\|X_{v}^{n}(t)\right\|_{H}^{2}<\infty, this implies that

suptXv(t)H2<.\sup_{t\in\mathbb{R}}\left\|X_{v}^{*}(t)\right\|_{H}^{2}<\infty. (5.3)

Finally we will show XvX_{v}^{*} satisfies Eq. (5.1).

Step 1: For any tt\in\mathbb{R} and t0<tt_{0}<t, we will show that XvX_{v}^{*} satisfies

Xv(t)=SA(tt0)Xv(t0)+t0tSA(tr)F(Xv(r))2dr+t0tSA(tr)B(Xv(r))v(r)dr.\displaystyle X_{v}^{*}(t)=S_{A}(t-t_{0})X_{v}^{*}(t_{0})+\int_{t_{0}}^{t}S_{A}(t-r)F(X_{v}^{*}(r))^{2}\mathrm{d}r+\int_{t_{0}}^{t}S_{A}(t-r)B(X_{v}^{*}(r))v(r)\mathrm{d}r.

For any t0<tt_{0}<t\in\mathbb{R}, we can find N+N\in\mathbb{N}^{+} such that t0,t[N,N]t_{0},t\in[-N,N]. Fixed NN, it follows from Hypothesis 2.1 (ii), we obtain that

t0tSA(tr)F(Xv(r))drt0tSA(tr)F(Xvn(r))drH2\displaystyle\left\|\int_{t_{0}}^{t}S_{A}(t-r)F(X_{v}^{*}(r))\mathrm{d}r-\int_{t_{0}}^{t}S_{A}(t-r)F(X_{v}^{n}(r))\mathrm{d}r\right\|^{2}_{H}
\displaystyle\leq C(N)t0t(tr)αXv(r)Xvn(r)H2dr\displaystyle C(N)\int_{t_{0}}^{t}(t-r)^{\alpha}\left\|X_{v}^{*}(r)-X_{v}^{n}(r)\right\|_{H}^{2}\mathrm{d}r
\displaystyle\leq C(N)supr[N,N]Xv(r)Xvn(r)H2t0t(tr)αdr,\displaystyle C(N)\sup_{r\in[-N,N]}\left\|X_{v}^{*}(r)-X_{v}^{n}(r)\right\|_{H}^{2}\int_{t_{0}}^{t}(t-r)^{\alpha}\mathrm{d}r,

Since Xvn()Xv()X_{v}^{n}(\cdot)\rightarrow X_{v}^{*}(\cdot) in C([N,N];H)C([-N,N];H), it deduces that

t0tSA(tr)F(Xv(r))drt0tSA(tr)F(Xvn(r))drH20.\left\|\int_{t_{0}}^{t}S_{A}(t-r)F(X_{v}^{*}(r))\mathrm{d}r-\int_{t_{0}}^{t}S_{A}(t-r)F(X_{v}^{n}(r))\mathrm{d}r\right\|^{2}_{H}\rightarrow 0.

It follows from Remark 2.1 (ii) and (iii),

t0tSA(tr)B(Xv(r))v(r)drt0tSA(tr)B(Xvn(r))v(r)drH2\displaystyle\left\|\int_{t_{0}}^{t}S_{A}(t-r)B(X_{v}^{*}(r))v(r)\mathrm{d}r-\int_{t_{0}}^{t}S_{A}(t-r)B(X_{v}^{n}(r))v(r)\mathrm{d}r\right\|_{H}^{2}
\displaystyle\leq Ct0tB(Xv(r))v(r)B(Xvn(r))v(r)H2ds\displaystyle C\int_{t_{0}}^{t}\left\|B\left(X_{v}^{*}(r)\right)v(r)-B(X_{v}^{n}(r))v(r)\right\|_{H}^{2}\mathrm{d}s
\displaystyle\leq Cβ2t0tXv(r)Xvn(r)H2vH02ds\displaystyle C\beta^{2}\int_{t_{0}}^{t}\|X_{v}^{*}(r)-X_{v}^{n}(r)\|_{H}^{2}\|v\|_{H_{0}}^{2}\mathrm{d}s
\displaystyle\leq Cβ2supt[N,N]Xv(r)Xvn(r)H2t0tvH02ds.\displaystyle C\beta^{2}\sup_{t\in[-N,N]}\|X_{v}^{*}(r)-X_{v}^{n}(r)\|_{H}^{2}\int_{t_{0}}^{t}\|v\|_{H_{0}}^{2}\mathrm{d}s.

Since Xvn()Xv()X_{v}^{n}(\cdot)\rightarrow X_{v}^{*}(\cdot) in C([N,N];H)C([-N,N];H), for any vL2(;H0)v\in L^{2}(\mathbb{R};H_{0}), it implies that

limnt0tSA(tr)B(Xvn)v(r)dr=t0tSA(tr)B(Xv(r))v(r)dr,inC([N,N];H).\lim_{n\rightarrow\infty}\int_{t_{0}}^{t}S_{A}(t-r)B\left(X_{v}^{n}\right)v(r)\mathrm{d}r=\int_{t_{0}}^{t}S_{A}(t-r)B\left(X_{v}^{*}(r)\right)v(r)\mathrm{d}r,\quad\text{in}~{}C([-N,N];H).

At the same time it follows form Remark 2.1 (iii) that Xvn(t)X_{v}^{n}(t), SA(tt0)Xvn(t0)S_{A}(t-t_{0})X_{v}^{n}(t_{0}) converge strongly to XvX_{v}^{*} and SA(tt0)Xv(t0)S_{A}(t-t_{0})X_{v}^{*}(t_{0}) in HH respectively, hence Eq. (5.1) holds.

Step 2: We next prove that XvX_{v}^{*} satisfies Eq. (5.1). From Eq. (5.1), it is easy to know that for any 0<m<n0<m<n,

nmSA(r)[F(Xv(r))+B(Xv(r))v(r)]dr=SA(n)Xv(n)+SA(m)Xv(m),\int_{-n}^{-m}S_{A}(-r)[F(X_{v}^{*}(r))+B(X_{v}^{*}(r))v(r)]\mathrm{d}r=-S_{A}(n)X_{v}^{*}(-n)+S_{A}(m)X_{v}^{*}(-m),

moreover,

SA(n)Xv(n)H|SA(n)|LXv(n)H,SA(m)Xv(m)H|SA(m)|LXv(m)H2.\|S_{A}(n)X_{v}^{*}(-n)\|_{H}\leq|S_{A}(n)|_{L}\left\|X_{v}^{*}(-n)\right\|_{H},\quad\|S_{A}(m)X_{v}^{*}(-m)\|_{H}\leq|S_{A}(m)|_{L}\left\|X_{v}^{*}(-m)\right\|_{H}^{2}.

Thus combining (5.3) and Remark 2.1 (iii), we obtain that SA(n)Xv(n)H\|S_{A}(n)X_{v}^{*}(-n)\|_{H} and SA(m)Xv(m)H\|S_{A}(m)X_{v}^{*}(-m)\|_{H} converge to 0 as m,nm,n\rightarrow\infty. Therefore,

ntSA(r)[F(Xv(r))+B(Xv(r))v(r)]dr\int_{-n}^{t}S_{A}(-r)[F(X_{v}^{*}(r))+B(X_{v}^{*}(r))v(r)]\mathrm{d}r

is a Cauchy sequence in HH with respect to nn for any tt\in\mathbb{R}. Let nn\rightarrow\infty, we can obtain that

ntSA(r)[F(Xv(r))+B(Xv(r))v(r)]drtSA(r)[F(Xv(r))+B(Xv(r))v(r)]dr.\displaystyle\int_{-n}^{t}S_{A}(-r)[F(X_{v}^{*}(r))+B(X_{v}^{*}(r))v(r)]\mathrm{d}r\rightarrow\int_{-\infty}^{t}S_{A}(-r)[F(X_{v}^{*}(r))+B(X_{v}^{*}(r))v(r)]\mathrm{d}r.

Moreover, by using Remark 2.1 (iii), SA(t+n)Xv(n)H2|SA(t+n)|LXv(n)H20,asn.\left\|S_{A}(t+n)X_{v}^{*}(-n)\right\|_{H}^{2}\leq|S_{A}(t+n)|_{L}\left\|X_{v}^{*}(-n)\right\|_{H}^{2}\rightarrow 0,~{}\text{as}~{}n\rightarrow\infty. Thus it follows from Eq. (5.1) that XvX_{v}^{*} satisfies Eq. (5.1). ∎

The proof of Theorem 3.1 in the following.

Proof.

Combining Lemma 5.5 and Lemma 5.4, we have known that the solution of the backward infinite horizon integral Eq. (3.2) for skeleton Eq. (3.1) exists and is unique, and satisfies (3.3). ∎

5.2 Stationary solution

Similar to the proofs of Mattingly [21] and Liu and Zhao [19], we prove Theorem 2.6 in this subsection. This subsection under Hypothesis 2.1, and we first give the energy estimate for Eq. (2.1) in the following lemma.

Lemma 5.6.

(Energy Estimate) For convenience, we fixed t0t_{0} and denote Xε(t)=Xε(t,t0;X0)X_{\varepsilon}(t)=X_{\varepsilon}(t,t_{0};X_{0}) be the solution of Eq. (2.1) with initial value X0X_{0} at time t0t_{0}, then we have

𝔼Xε(t)H2p\displaystyle\mathbb{E}\left\|X_{\varepsilon}(t)\right\|_{H}^{2p}\leq 𝔼X0H2p2λp𝔼t0tXε(s)H2(p1)Xε(s)V2ds\displaystyle\mathbb{E}\left\|X_{0}\right\|_{H}^{2p}-2\lambda p\mathbb{E}\int_{t_{0}}^{t}\left\|X_{\varepsilon}(s)\right\|_{H}^{2(p-1)}\left\|X_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s
+𝔼t0t2εp(p1)Xε(s)H2(p2)(Xε(s)H2|B(Xε(s))|LQ2)ds\displaystyle+\mathbb{E}\int_{t_{0}}^{t}2\varepsilon p(p-1)\left\|X_{\varepsilon}(s)\right\|_{H}^{2(p-2)}\left(\|X_{\varepsilon}(s)\|_{H}^{2}|B(X_{\varepsilon}(s))|_{L_{Q}}^{2}\right)\mathrm{d}s
+𝔼t0tεpXε(s)H2(p1)|B(Xε(s))|LQ2ds,ε>0.\displaystyle+\mathbb{E}\int_{t_{0}}^{t}\varepsilon p\left\|X_{\varepsilon}(s)\right\|_{H}^{2(p-1)}|B(X_{\varepsilon}(s))|_{L_{Q}}^{2}\mathrm{d}s,\quad\forall\varepsilon>0.
Proof.

For every ε>0\varepsilon>0, it follows from the Itô formula that

dXε(t)H2=2AXε,XεVVdt+2F(Xε),XεHdt+2εXε,B(Xε)dWH+ε|B(Xε)|LQ2dt.\mathrm{d}\|X_{\varepsilon}(t)\|_{H}^{2}=2{{}_{V^{*}}}\langle AX_{\varepsilon},X_{\varepsilon}\rangle_{V}\mathrm{d}t+2\langle F(X_{\varepsilon}),X_{\varepsilon}\rangle_{H}\mathrm{d}t+2\sqrt{\varepsilon}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}+\varepsilon|B(X_{\varepsilon})|_{L_{Q}}^{2}\mathrm{d}t.

For p1p\geq 1, it follows from the Itô formula and Hypothesis 2.1 (i) that

dXε(t)H2p\displaystyle\mathrm{d}\left\|X_{\varepsilon}(t)\right\|_{H}^{2p} =\displaystyle= pXε(t)H2(p1)dXε(t)H2+12p(p1)Xε(t)H2(p2)dXε(t)H2t\displaystyle p\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-1)}\mathrm{d}\left\|X_{\varepsilon}(t)\right\|_{H}^{2}+\frac{1}{2}p(p-1)\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-2)}{\left\langle\mathrm{d}\left\|X_{\varepsilon}(t)\right\|_{H}^{2}\right\rangle}_{t}
\displaystyle\leq pXε(t)H2(p1)[2λXεV+2εXε,B(Xε)dWH]\displaystyle p\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-1)}\left[-2\lambda\|X_{\varepsilon}\|_{V}+2\sqrt{\varepsilon}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}\right]
+2εp(p1)Xε(t)H2(p2)(XεH2|B(Xε)|LQ2)dt\displaystyle+2\varepsilon p(p-1)\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-2)}\left(\|X_{\varepsilon}\|_{H}^{2}|B(X_{\varepsilon})|_{L_{Q}}^{2}\right)\mathrm{d}t
+εpXε(t)H2(p1)|B(Xε)|LQ2dt.\displaystyle+\varepsilon p\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-1)}|B(X_{\varepsilon})|_{L_{Q}}^{2}\mathrm{d}t.

For convenience set Mε(t):=Mε(t,t0;X0)=2εpt0tXεH2(p1)Xε,B(Xε)dWHM_{\varepsilon}(t):=M_{\varepsilon}(t,t_{0};X_{0})=2\sqrt{\varepsilon}p\int_{t_{0}}^{t}\left\|X_{\varepsilon}\right\|_{H}^{2(p-1)}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}, then for Mε(t)M_{\varepsilon}(t) is a local martingale there exists a sequence of stopping time {Tn}\left\{T_{n}\right\}, with {Tn}+\left\{T_{n}\right\}\rightarrow+\infty as n+n\rightarrow+\infty, such that Mε(tTn)M_{\varepsilon}({t\wedge T_{n}}) is a martingale, so the Optional Stopping Time Theorem implies 𝔼Mε(tTn)=0\mathbb{E}M_{\varepsilon}({t\wedge T_{n}})=0, we denote fn(t)f_{n}(t) by

fn(t)\displaystyle f_{n}(t) =\displaystyle= X0H2p+t0t2εp(p1)XεH2(p2)(XεH2|B(Xε)|LQ2)ds\displaystyle\left\|X_{0}\right\|_{H}^{2p}+\int_{t_{0}}^{t}2\varepsilon p(p-1)\left\|X_{\varepsilon}\right\|_{H}^{2(p-2)}\left(\|X_{\varepsilon}\|_{H}^{2}|B(X_{\varepsilon})|_{L_{Q}}^{2}\right)\mathrm{d}s
+t0tpXε(t)H2(p1)|B(Xε)|LQ2ds+Mε(tTn).\displaystyle+\int_{t_{0}}^{t}p\left\|X_{\varepsilon}(t)\right\|_{H}^{2(p-1)}|B(X_{\varepsilon})|_{L_{Q}}^{2}\mathrm{d}s+M_{\varepsilon}({t\wedge T_{n}}).

When tTnt\leq T_{n}, fn(t)Xε(t)H2p+λt0t2pXε(s)H2(p1)Xε(s)V2ds>0f_{n}(t)\geq\left\|X_{\varepsilon}(t)\right\|_{H}^{2p}+\lambda\int_{t_{0}}^{t}2p\left\|X_{\varepsilon}(s)\right\|_{H}^{2(p-1)}\left\|X_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s>0. And when t>Tnt>T_{n}, fn(t)f(Tn)=Tnt2εp(p1)XεH2(p2)(XεH2|B(Xε)|LQ2)ds+TntpXεH2(p1)|B(Xε)|LQ2ds>0f_{n}(t)-f(T_{n})=\int_{T_{n}}^{t}2\varepsilon p(p-1)\left\|X_{\varepsilon}\right\|_{H}^{2(p-2)}\left(\|X_{\varepsilon}\|_{H}^{2}|B(X_{\varepsilon})|_{L_{Q}}^{2}\right)\mathrm{d}s+\int_{T_{n}}^{t}p\left\|X_{\varepsilon}\right\|_{H}^{2(p-1)}|B(X_{\varepsilon})|_{L_{Q}}^{2}\mathrm{d}s>0. So, fn(t)f_{n}(t) is non-negative for all tt. Thus we can use Fatou’s Lemma to get

𝔼Xε(t)H2p+𝔼t0t2λpXε(s)H2(p1)Xε(s)V2ds=𝔼limn+fnlimn+𝔼fn,\displaystyle\mathbb{E}\left\|X_{\varepsilon}(t)\right\|_{H}^{2p}+\mathbb{E}\int_{t_{0}}^{t}2\lambda p\left\|X_{\varepsilon}(s)\right\|_{H}^{2(p-1)}\left\|X_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s=\mathbb{E}\lim_{n\rightarrow+\infty}f_{n}\leq\lim_{n\rightarrow+\infty}\mathbb{E}f_{n},

which imply the conclusion. ∎

Set 𝔼p,ε(t):=𝔼Xε(t)H2p\mathbb{E}_{p,{\varepsilon}}(t):=\mathbb{E}\left\|X_{\varepsilon}(t)\right\|_{H}^{2p}, we prove the following lemma by using Lemma 5.6.

Lemma 5.7.

Assume the initial condition satisfies 𝔼X(t0)H2p<\mathbb{E}\left\|X(t_{0})\right\|_{H}^{2p}<\infty for some fixed p1p\geq 1, then for any ε0>0\varepsilon_{0}>0, there exists a constant CC such that

𝔼Xε(t,t0;X(t0))H2pC,tt0,ε(0,ε0),\mathbb{E}\left\|X_{\varepsilon}(t,t_{0};X(t_{0}))\right\|_{H}^{2p}\leq C,\quad\forall t\geq t_{0},\ \varepsilon\in(0,\varepsilon_{0}),

where C depend on 𝔼X(t0)H2j,1jp,j\mathbb{E}\left\|X(t_{0})\right\|_{H}^{2j},1\leq j\leq p,~{}j\in\mathbb{Z} and p,ε0,λ,C1,Dp,\varepsilon_{0},\lambda,C_{1},D, where λ\lambda is defined in Hypothesis 2.1 (i), C1C_{1} and DD are defined in Remark 2.1 (i) and (ii) respectively.

Proof.

For any tt0t\geq t_{0}, it follows from the Energy Estimate Lemma 5.6 and Remark 2.1 (i),(ii) that

d𝔼p,ε(t)2pλC1𝔼p,ε(t)+(2εp(p1)D2+εpD2)𝔼p1,ε(t),\mathrm{d}\mathbb{E}_{p,{\varepsilon}}(t)\leq-2p\lambda C_{1}\mathbb{E}_{p,{\varepsilon}}(t)+\left(2\varepsilon p(p-1)D^{2}+\varepsilon pD^{2}\right)\mathbb{E}_{p-1,\varepsilon}(t),

and

d(e2pλC1t𝔼p,ε(t))=\displaystyle\mathrm{d}\left(e^{2p\lambda C_{1}t}\mathbb{E}_{p,{\varepsilon}}(t)\right)= 2pλC1e2pλC1t𝔼p,ε(t)dt+e2pλC1td𝔼p,ε(t).\displaystyle 2p\lambda C_{1}e^{2p\lambda C_{1}t}\mathbb{E}_{p,{\varepsilon}}(t)\mathrm{d}t+e^{2p\lambda C_{1}t}\mathrm{d}\mathbb{E}_{p,{\varepsilon}}(t).

Thus

𝔼p,ε(t)e2pλC1(tt0)𝔼p,ε(t0)+(2εp(p1)D2+εpD2)t0te2pλC1(ts)𝔼p1,ε(s)ds,\mathbb{E}_{p,{\varepsilon}}(t)\leq e^{-2p\lambda C_{1}(t-t_{0})}\mathbb{E}_{p,{\varepsilon}}(t_{0})+\left(2\varepsilon p(p-1)D^{2}+\varepsilon pD^{2}\right)\int_{t_{0}}^{t}e^{-2p\lambda C_{1}(t-s)}\mathbb{E}_{p-1,\varepsilon}(s)\mathrm{d}s,

then 𝔼1,ε(t)𝔼1,ε(t0)+εD2t0te2λC1(ts)ds𝔼1(t0)+εD22λC1\mathbb{E}_{1,\varepsilon}(t)\leq\mathbb{E}_{1,\varepsilon}(t_{0})+\varepsilon D^{2}\int_{t_{0}}^{t}e^{-2\lambda C_{1}(t-s)}\mathrm{d}s\leq\mathbb{E}_{1}(t_{0})+\frac{\varepsilon D^{2}}{2\lambda C_{1}}, set 𝔼1max(t0)=𝔼1(t0)+ε0D22λC1\mathbb{E}_{1}^{\max}(t_{0})=\mathbb{E}_{1}(t_{0})+\frac{\varepsilon_{0}D^{2}}{2\lambda C_{1}}, then by induction we get that

𝔼p,ε(t)𝔼p,ε(t0)+Cp𝔼p1max(t0)=𝔼p,εmax(t0),\mathbb{E}_{p,{\varepsilon}}(t)\leq\mathbb{E}_{p,{\varepsilon}}(t_{0})+C_{p}^{{}^{\prime}}\mathbb{E}_{p-1}^{\max}(t_{0})=\mathbb{E}_{p,{\varepsilon}}^{\max}(t_{0}),

where CpC_{p}^{{}^{\prime}} is constant only depends on p,ε0,λ,C1,Dp,\varepsilon_{0},\lambda,C_{1},D, which imply that 𝔼p,εmax\mathbb{E}_{p,{\varepsilon}}^{\max} is just the linear combination of the moments of order less that or equal to pp of 𝔼p,εmax(t0)\mathbb{E}_{p,{\varepsilon}}^{\max}(t_{0}). ∎

Let {Xn}\{X_{n}\} be a sequence of real random variables. Let g:++g:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+}. Define the random variable Tbound({Xn},g)T_{bound}\left(\{X_{n}\},g\right) to be the smallest positive integer such that for almost every ω\omega,

m>Tbound({Xn},g)(ω)|Xm(ω)|g(m),ωΩ.m>T_{bound}\left(\{X_{n}\},g\right)(\omega)\Rightarrow|X_{m}(\omega)|\leq g(m),~{}\quad\forall\omega\in\Omega.

Mattingly has proved the following Bounding Lemma in [21].

Lemma 5.8.

(Bounding Lemma) Assume that

(|Xn|εnδ)𝔼|Xn|pnpδεpCnpδrεp,\mathbb{P}(|X_{n}|\geq\varepsilon n^{\delta})\leq\frac{\mathbb{E}|X_{n}|^{p}}{n^{p\delta}\varepsilon^{p}}\leq\frac{C}{n^{p\delta-r}\varepsilon^{p}},

for some ε,δ,p,C>0\varepsilon,\delta,p,C>0 and r0r\geq 0, then
(i). if pδ>1+rp\delta>1+r then Tbound({Xn},εnδ)<+T_{bound}\left(\{X_{n}\},\varepsilon n^{\delta}\right)<+\infty a.s.ω\omega.
(ii). 𝔼[Tbound({Xn},εnδ)]q\mathbb{E}\left[T_{bound}\left(\{X_{n}\},\varepsilon n^{\delta}\right)\right]^{q} is finite for q(0,pδ(1+r))q\in(0,p\delta-(1+r)).

Let Xε(t,t0;ω,X0)X_{\varepsilon}(t,t_{0};\omega,X_{0}) and X~ε(t,t0;ω,X~0)\tilde{X}_{\varepsilon}(t,t_{0};\omega,\tilde{X}_{0}) be the solutions of Eq. (2.1) starting from different initial value X0,X~0X_{0},\tilde{X}_{0} respectively. Define ρε(t,t0;X0,X~0)=Xε(t,t0;ω,X0)X~ε(t,t0;ω,X~0)\rho_{\varepsilon}(t,t_{0};X_{0},\tilde{X}_{0})=X_{\varepsilon}(t,t_{0};\omega,X_{0})-\tilde{X}_{\varepsilon}(t,t_{0};\omega,\tilde{X}_{0}), we will consider the asymptotically stable property of Eq. (2.1) in the following lemmas.

Lemma 5.9.

Let Γ(l,t0;Xε)=2λC1εβ22C0(1lt0t0+lXε(s)V2ds)\Gamma(l,t_{0};X_{\varepsilon})=2\lambda C_{1}-\varepsilon\beta^{2}-2C_{0}\Big{(}\frac{1}{l}\int_{t_{0}}^{t_{0}+l}\left\|X_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s\Big{)}, for any ε0>0\varepsilon_{0}>0, there exists δ>1\delta>1 such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) there exists a almost finite random variable τε\tau_{\varepsilon} satisfies

ρε(t0+n,t0;X0,X~0)H2enΓ(n,t0;Xε)(X0X~0H2+2εnδ),\left\|\rho_{\varepsilon}(t_{0}+n,t_{0};X_{0},\tilde{X}_{0})\right\|_{H}^{2}\leq e^{-n\Gamma(n,t_{0};X_{\varepsilon})}\Big{(}\left\|X_{0}-\tilde{X}_{0}\right\|_{H}^{2}+2\sqrt{\varepsilon}n^{\delta}\Big{)},

for every n>τεn>\tau_{\varepsilon}.

Proof.

It follows from the Itô formula and Hypothesis 2.1 (i) and Remark 2.1 (i),(ii) that

dρε(t)H2=\displaystyle\mathrm{d}\left\|\rho_{\varepsilon}(t)\right\|_{H}^{2}= 2AXεAX~ε+F(Xε)F(X~ε),ρεVV\displaystyle 2{{}_{V^{*}}}\langle AX_{\varepsilon}-A\tilde{X}_{\varepsilon}+F(X_{\varepsilon})-F(\tilde{X}_{\varepsilon}),\rho_{\varepsilon}\rangle_{V} (5.5)
+2ερε,(B(Xε)B(X~ε))dWtH+ε|B(Xε)B(X~ε)|LQ2\displaystyle+2\sqrt{\varepsilon}\langle\rho_{\varepsilon},(B(X_{\varepsilon})-B(\tilde{X}_{\varepsilon}))\mathrm{d}W_{t}\rangle_{H}+\varepsilon|B(X_{\varepsilon})-B(\tilde{X}_{\varepsilon})|_{L_{Q}}^{2}
\displaystyle\leq 2λρεV2+2C0ρεH2XεV2+εβ2ρεH2+2ερε,(B(Xε)B(X~ε))dWtH\displaystyle-2\lambda\|\rho_{\varepsilon}\|_{V}^{2}+2C_{0}\left\|\rho_{\varepsilon}\right\|_{H}^{2}\left\|X_{\varepsilon}\right\|_{V}^{2}+\varepsilon\beta^{2}\|\rho_{\varepsilon}\|_{H}^{2}+2\sqrt{\varepsilon}\langle\rho_{\varepsilon},(B(X_{\varepsilon})-B(\tilde{X}_{\varepsilon}))\mathrm{d}W_{t}\rangle_{H}
\displaystyle\leq (2λC12C0XεV2εβ2)ρεH2+2ερε,(B(Xε)B(X~ε))dWtH.\displaystyle-\Big{(}2\lambda C_{1}-2C_{0}\|X_{\varepsilon}\|_{V}^{2}-\varepsilon\beta^{2}\Big{)}\|\rho_{\varepsilon}\|_{H}^{2}+2\sqrt{\varepsilon}\langle\rho_{\varepsilon},(B(X_{\varepsilon})-B(\tilde{X}_{\varepsilon}))\mathrm{d}W_{t}\rangle_{H}.

Set Mε(l,t0;ρ):=t0t0+lρε,[B(Xε)B(X~ε)]dWtHM_{\varepsilon}(l,t_{0};\rho):=\int_{t_{0}}^{t_{0}+l}\Big{\langle}\rho_{\varepsilon},\big{[}B(X_{\varepsilon})-B(\tilde{X}_{\varepsilon})\big{]}\mathrm{d}W_{t}\Big{\rangle}_{H}, it follows from the Burkholder-Davis-Gundy inequality, Remark 2.1 (ii) and Lemma 5.7, for any ε0>0\varepsilon_{0}>0 there exists a constant CC such that

𝔼(supt0slMε(s,t0;ρ))2C𝔼(M(,t0;ρ)l)CD2t0t0+l𝔼ρε(s)H2dsCD2l𝔼1,εmax,\displaystyle\mathbb{E}\left(\sup_{t_{0}\leq s\leq l}M_{\varepsilon}(s,t_{0};\rho)\right)^{2}\leq C\mathbb{E}\left(\left\langle M(\cdot,t_{0};\rho)\right\rangle_{l}\right)\leq CD^{2}\int_{t_{0}}^{t_{0}+l}\mathbb{E}\left\|\rho_{\varepsilon}(s)\right\|_{H}^{2}\mathrm{d}s\leq CD^{2}l\mathbb{E}_{1,\varepsilon}^{\max},

for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}).

Let Mεn=supt0snMε(s,t0;ρ)M^{n}_{\varepsilon}=\sup_{t_{0}\leq s\leq n}M_{\varepsilon}(s,t_{0};\rho), by the Chebyshev’s inequality, we have

(|Mεn|nδ)𝔼|Mεn|2n2δCD2𝔼1,εmaxn2δ1,\mathbb{P}(|M^{n}_{\varepsilon}|\geq n^{\delta})\leq\frac{\mathbb{E}|M^{n}_{\varepsilon}|^{2}}{n^{2\delta}}\leq\frac{CD^{2}\mathbb{E}_{1,\varepsilon}^{\max}}{n^{2\delta-1}},

for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and δ>1\delta>1, it follows from Bounding Lemma 5.8, τε:=Tbound({Mεn},nδ)\tau_{\varepsilon}:=T_{bound}(\left\{M^{n}_{\varepsilon}\right\},n^{\delta}) is almost finite. Combining (5.5), when l>τεl>\tau_{\varepsilon}

ρε(t0+l)H2ρε(t0)H2+2εlδt0t0+l(2λC12C0XεV2εβ2)ρεH2dt.\|\rho_{\varepsilon}(t_{0}+l)\|_{H}^{2}\leq\|\rho_{\varepsilon}(t_{0})\|_{H}^{2}+2\sqrt{\varepsilon}l^{\delta}-\int_{t_{0}}^{t_{0}+l}\Big{(}\frac{2\lambda}{C_{1}}-2C_{0}\|X_{\varepsilon}\|_{V}^{2}-\varepsilon\beta^{2}\Big{)}\left\|\rho_{\varepsilon}\right\|_{H}^{2}\mathrm{d}t.

By using the Gronwall inequality, we get the desired result. ∎

It follows from the Itô formula and Hypothesis 2.1 that

2λlt0t0+lXε(s,t0;X0)V2ds\displaystyle\frac{2\lambda}{l}\int_{t_{0}}^{t_{0}+l}\left\|X_{\varepsilon}(s,t_{0};X_{0})\right\|_{V}^{2}\mathrm{d}s \displaystyle\leq X0H2Xε(t0+l,t0;X0)H2l+εlt0t0+l|B(Xε)|LQ2ds\displaystyle\frac{\left\|X_{0}\right\|_{H}^{2}-\left\|X_{\varepsilon}(t_{0}+l,t_{0};X_{0})\right\|_{H}^{2}}{l}+\frac{\varepsilon}{l}\int_{t_{0}}^{t_{0}+l}|B(X_{\varepsilon})|_{L_{Q}}^{2}\mathrm{d}s
+2εlt0t0+lXε,B(Xε)dWH\displaystyle+\frac{2\sqrt{\varepsilon}}{l}\int_{t_{0}}^{t_{0}+l}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}
\displaystyle\leq εD2+1l(X0H2+2εt0t0+lXε,B(Xε)dWH).\displaystyle\varepsilon D^{2}+\frac{1}{l}\Big{(}\left\|X_{0}\right\|_{H}^{2}+2\sqrt{\varepsilon}\int_{t_{0}}^{t_{0}+l}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}\Big{)}.

Set M~ε(l,t0;Xε)=t0t0+lXε,B(Xε)dWH\tilde{M}_{\varepsilon}(l,t_{0};X_{\varepsilon})=\int_{t_{0}}^{t_{0}+l}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}, for control Γ(l,t0;Xε)\Gamma(l,t_{0};X_{\varepsilon}) we need to control M~ε(l,t0;Xε)\tilde{M}_{\varepsilon}(l,t_{0};X_{\varepsilon}). By using the Burkholder-Davis-Gundy inequality and Lemma 5.7, for any ε0>0\varepsilon_{0}>0 there exists a constant CC such that

𝔼(supt0slM~ε(s,t0;Xε))2p\displaystyle\mathbb{E}\left(\sup_{t_{0}\leq s\leq l}\tilde{M}_{\varepsilon}(s,t_{0};X_{\varepsilon})\right)^{2p} 𝔼(M~ε(l,t0;Xε)lp)\displaystyle\leq\mathbb{E}\left(\left\langle\tilde{M}_{\varepsilon}(l,t_{0};X_{\varepsilon})\right\rangle_{l}^{p}\right)
CD2plp1t0t0+l𝔼Xε(s)L22pds\displaystyle\leq CD^{2p}l^{p-1}\int_{t_{0}}^{t_{0}+l}\mathbb{E}\left\|X_{\varepsilon}(s)\right\|_{L^{2}}^{2p}\mathrm{d}s
CD2plp𝔼p,εmax,ε(0,ε0),\displaystyle\leq CD^{2p}l^{p}\mathbb{E}_{p,{\varepsilon}}^{\max},\quad\forall\varepsilon\in(0,\varepsilon_{0}),

the second inequality comes from Hölder’s inequality.

Let M~εn=supt0snM~ε(s,t0;Xε)\tilde{M}^{n}_{\varepsilon}=\sup_{t_{0}\leq s\leq n}\tilde{M}_{\varepsilon}(s,t_{0};X_{\varepsilon}). By the Chebyshev’s inequality, we have

(|M~εn|ηnδ)𝔼|M~εn|2pn2pδη2pCD2p𝔼p,εmaxn2pδpη2p.\mathbb{P}(|\tilde{M}^{n}_{\varepsilon}|\geq\eta n^{\delta})\leq\frac{\mathbb{E}|\tilde{M}^{n}_{\varepsilon}|^{2p}}{n^{2p\delta}\eta^{2p}}\leq\frac{CD^{2p}\mathbb{E}_{p,{\varepsilon}}^{\max}}{n^{2p\delta-p}\eta^{2p}}. (5.7)

By using Bounding Lemma 5.8, we can get the following Lemma.

Lemma 5.10.

Let X0X_{0} be a random variable, measurable with respect to t0\mathcal{F}_{t_{0}}, such that 𝔼X0H2p\mathbb{E}\left\|X_{0}\right\|_{H}^{2p} is finite, then
(i). For any fixed ε0>0\varepsilon_{0}>0, if p>1p>1, then for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), ηε,ηε>0\eta_{\varepsilon},\eta^{\prime}_{\varepsilon}>0, Tbound({M~εn},ηεn)T_{bound}\left(\left\{\tilde{M}^{n}_{\varepsilon}\right\},\eta_{\varepsilon}n\right) and Tbound({X0H2},ηεn)T_{bound}\left(\left\{\left\|X_{0}\right\|_{H}^{2}\right\},\eta^{\prime}_{\varepsilon}n\right) is finite almost surely.
(ii). 𝔼Tbound({M~εn},ηεn)q\mathbb{E}T_{bound}\left(\left\{\tilde{M}^{n}_{\varepsilon}\right\},\eta_{\varepsilon}n\right)^{q} and 𝔼Tbound({X0H2},ηεn)q\mathbb{E}T_{bound}\left(\left\{\left\|X_{0}\right\|_{H}^{2}\right\},\eta^{\prime}_{\varepsilon}n\right)^{q} is finite for q(0,p1)q\in(0,p-1).

Proof.

Noticing that (5.7) and

(X0H2ηεn)𝔼X0H2pnp(ηε)pCnp(ηε)p,\mathbb{P}(\left\|X_{0}\right\|_{H}^{2}\geq\eta^{\prime}_{\varepsilon}n)\leq\frac{\mathbb{E}\left\|X_{0}\right\|_{H}^{2p}}{n^{p}(\eta^{\prime}_{\varepsilon})^{p}}\leq\frac{C}{n^{p}(\eta^{\prime}_{\varepsilon})^{p}},

Bounding Lemma 5.8 imply the conclusion. ∎

Fixed ε0>0\varepsilon_{0}>0, for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), combining the definition of Γ(l,t0;Xε)\Gamma(l,t_{0};X_{\varepsilon}), (5.2) and Lemma 5.10, when

l>max{Tbound({M~εn},εn),Tbound({X0H2},εn)},l>\max\left\{T_{bound}\left(\left\{\tilde{M}^{n}_{\varepsilon}\right\},\sqrt{\varepsilon}n\right),\quad T_{bound}\left(\left\{\left\|X_{0}\right\|_{H}^{2}\right\},{\varepsilon}n\right)\right\},

we have

Γ(l,t0;Xε)=\displaystyle\Gamma(l,t_{0};X_{\varepsilon})= 2λC1εβ22C0(1lt0t0+lXε(s)V2ds)\displaystyle 2\lambda C_{1}-\varepsilon\beta^{2}-2C_{0}\Big{(}\frac{1}{l}\int_{t_{0}}^{t_{0}+l}\left\|X_{\varepsilon}(s)\right\|_{V}^{2}\mathrm{d}s\Big{)}
\displaystyle\geq 2λC1εβ2C0λ{εD2+1l(X0H2+2εt0t0+lXε,B(Xε)dWH)}\displaystyle 2\lambda C_{1}-\varepsilon\beta^{2}-\frac{C_{0}}{\lambda}\left\{\varepsilon D^{2}+\frac{1}{l}\left(\left\|X_{0}\right\|_{H}^{2}+2\sqrt{\varepsilon}\int_{t_{0}}^{t_{0}+l}\langle X_{\varepsilon},B(X_{\varepsilon})\mathrm{d}W\rangle_{H}\right)\right\}
\displaystyle\geq 2λC1εβ2C0λ(εD2+3ε).\displaystyle 2\lambda C_{1}-\varepsilon\beta^{2}-\frac{C_{0}}{\lambda}(\varepsilon D^{2}+3\varepsilon). (5.8)
Lemma 5.11.

For any λ0(0,2λC1)\lambda_{0}\in(0,2\lambda C_{1}) and t0t_{0}\in\mathbb{R}. Let X0,X~0HX_{0},\tilde{X}_{0}\in H be initial condition, measurable with respect to t0\mathcal{F}_{-\infty}^{t_{0}}, such that 𝔼X0H2p<\mathbb{E}\left\|X_{0}\right\|_{H}^{2p}<\infty, 𝔼X~0H2p<\mathbb{E}\|\tilde{X}_{0}\|_{H}^{2p}<\infty for some p>1p>1. Let Xε(t,t0;X0)X_{\varepsilon}(t,t_{0};X_{0}) and X~ε(t,t0;X~0)\tilde{X}_{\varepsilon}(t,t_{0};\tilde{X}_{0}) be the solutions of Eq. (2.1) starting from different initial value X0,X~0X_{0},\tilde{X}_{0} respectively. Then there exists ε0,δ\varepsilon_{0},\delta and a sequence almost finite random variable {lε}ε(0,ε0)\{l_{\varepsilon}\}_{\varepsilon\in(0,\varepsilon_{0})} such that

Xε(t,t0;X0)X~ε(t,t0;X~0)H2(X0X~0H2+ε(tt0)δ)eλ0(tt0),\left\|X_{\varepsilon}(t,t_{0};X_{0})-\tilde{X}_{\varepsilon}(t,t_{0};\tilde{X}_{0})\right\|_{H}^{2}\leq\left(\left\|X_{0}-\tilde{X}_{0}\right\|_{H}^{2}+\sqrt{\varepsilon}(t-t_{0})^{\delta}\right)e^{-\lambda_{0}(t-t_{0})},

for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), t>t0+lεt>t_{0}+l_{\varepsilon}.

Proof.

We can chose ε0\varepsilon_{0} small enough such that

2λC1εβ2C0λ(εD2+3ε)>λ0,2\lambda C_{1}-\varepsilon\beta^{2}-\frac{C_{0}}{\lambda}(\varepsilon D^{2}+3\varepsilon)>\lambda_{0}, (5.9)

for every ε<ε0\varepsilon<\varepsilon_{0}. We define lεl_{\varepsilon} by

lε=max{τε,Tbound({M~εn},εn),Tbound({X0H2},εn)},l_{\varepsilon}=\max\left\{\tau_{\varepsilon},\quad T_{bound}\left(\left\{\tilde{M}^{n}_{\varepsilon}\right\},\sqrt{\varepsilon}n\right),\quad T_{bound}\left(\left\{\left\|X_{0}\right\|_{H}^{2}\right\},{\varepsilon}n\right)\right\},

where τε\tau_{\varepsilon} is defined in Lemma 5.9, the conclusion follows from Lemma 5.9, (5.2) and (5.9). ∎

Lemma 5.12.

(cf. Theorem 2 in [21]) Fix λ0(0,2λC1)\lambda_{0}\in(0,2\lambda C_{1}) and tt\in\mathbb{R}. Let {X0(n)}\left\{X_{0}(n)\right\} be a sequence of random variable with n+n\in\mathbb{Z}^{+}. Assume that X0(n)X_{0}(n) is measurable with respect to tn\mathcal{F}^{t-n}_{-\infty} and that 𝔼X0(n)H2p\mathbb{E}\left\|X_{0}(n)\right\|_{H}^{2p} is uniformly bounded in nn for some p>2p>2. Then there exists ε0,δ\varepsilon_{0},\delta such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) the following hold:
(i). With probability one, there exists a random time nε>0\vec{n}_{\varepsilon}>0 such that for every s>0s>0 and all nn\in\mathbb{Z} with n>nεn>\vec{n}_{\varepsilon}, we have

supX~0AnXε(t+s,tn;X0(n))Xε(t+s,tn;X~0)H2(n+ε(s+n)δ)eλ0(s+n).\sup_{\tilde{X}_{0}\in A_{n}}\left\|X_{\varepsilon}(t+s,t-n;X_{0}(n))-X_{\varepsilon}(t+s,t-n;\tilde{X}_{0})\right\|_{H}^{2}\leq\left(n+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)}.

Here AnA_{n} is the set {X~0:X~0H2n4}.\left\{\tilde{X}_{0}:\left\|\tilde{X}_{0}\right\|_{H}^{2}\leq\frac{n}{4}\right\}. In addition 𝔼(nεq)<\mathbb{E}(\vec{n}_{\varepsilon}^{q})<\infty for any q(0,p2)q\in(0,p-2).
(ii). Let {X¯0(n)}\left\{\bar{X}_{0}(n)\right\} be a second sequence of random variable with n+n\in\mathbb{Z}^{+} measurable with respect to tn\mathcal{F}^{t-n}_{-\infty} and that 𝔼X¯0(n)H2p\mathbb{E}\left\|\bar{X}_{0}(n)\right\|_{H}^{2p} is uniformly bounded in nn for some p>2p>2. Then with probability one, there exists a random time nε(ε,δ,t,ω)>0\vec{n}^{\prime}_{\varepsilon}(\varepsilon,\delta,t,\omega)>0 such that for every s>0s>0 and all nn\in\mathbb{Z} with n>nεn>\vec{n}^{\prime}_{\varepsilon}, we have

Xε(t+s,tn;X0(n))Xε(t+s,tn;X¯0(n))H2(n+ε(s+n)δ)eλ0(s+n).\left\|X_{\varepsilon}(t+s,t-n;X_{0}(n))-X_{\varepsilon}(t+s,t-n;\bar{X}_{0}(n))\right\|_{H}^{2}\leq\left(n+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)}.

And 𝔼[(nε)q]<\mathbb{E}\left[(\vec{n}^{\prime}_{\varepsilon})^{q}\right]<\infty for any q(0,p2)q\in(0,p-2).

Proof.

Only need to prove (i), it follows from Lemma 5.11, there exists ε0,δ\varepsilon_{0},\delta and a sequence almost finite random variable {lε}ε(0,ε0)\{l_{\varepsilon}\}_{\varepsilon\in(0,\varepsilon_{0})} such that

Xε(t+s,tn;X0(n))Xε(t+s,tn;X~0)H2(X0(n)X~0H2+ε(s+n)δ)eλ0(s+n),\left\|X_{\varepsilon}(t+s,t-n;X_{0}(n))-X_{\varepsilon}(t+s,t-n;\tilde{X}_{0})\right\|_{H}^{2}\leq\left(\left\|X_{0}(n)-\tilde{X}_{0}\right\|_{H}^{2}+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)},

for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), t>t0+lεt>t_{0}+l_{\varepsilon}.

Set TX0=Tbound({X0(n)H2},n4)T_{X_{0}}=T_{bound}\left(\left\{\left\|X_{0}(n)\right\|_{H}^{2}\right\},\frac{n}{4}\right), then using Bounding Lemma 5.8 we have that TX0T_{X_{0}} is almost finite and 𝔼(TX0)q,q(0,p2)\mathbb{E}(T_{X_{0}})^{q},q\in(0,p-2) is finite a.s. Set nε=max{TX0,lε}{\vec{n}}_{\varepsilon}=max\left\{T_{X_{0}},l_{\varepsilon}\right\}, then for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and n>nεn>\vec{n}_{\varepsilon}

supX~0AnXε(t+s,tn;X0(n))Xε(t+s,tn;X~0)H2\displaystyle\sup_{\tilde{X}_{0}\in A_{n}}\left\|X_{\varepsilon}(t+s,t-n;X_{0}(n))-X_{\varepsilon}(t+s,t-n;\tilde{X}_{0})\right\|_{H}^{2}
\displaystyle\leq (X0(n)X~0H2+ε(s+n)δ)eλ0(s+n)\displaystyle\left(\left\|X_{0}(n)-\tilde{X}_{0}\right\|_{H}^{2}+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)}
\displaystyle\leq (2X0(n)H2+2X~0H2+ε(s+n)δ)eλ0(s+n)\displaystyle\left(2\left\|X_{0}(n)\right\|_{H}^{2}+2\left\|\tilde{X}_{0}\right\|_{H}^{2}+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)}
\displaystyle\leq (n+ε(s+n)δ)eλ0(s+n).\displaystyle\left(n+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)}.

Lemma 5.13.

(cf. Corollary 1 in [21]) Fix t1t_{1}\in\mathbb{Z}, λ0(0,2λC1)\lambda_{0}\in(0,2\lambda C_{1}) and u0=0u_{0}=0. There exist ε0,δ\varepsilon_{0},\delta such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), with probability one, there is a positive random variable nεn^{*}_{\varepsilon} such that for all l0l\geq 0 and all n1,n2n_{1},n_{2}\in\mathbb{Z}, if n1,n2<nt1nεn_{1},n_{2}<n\leq t_{1}-n^{*}_{\varepsilon}, we have

Xε(t1+l,n1;X0)Xε(t1+l,n2;X0)H2(n+ε(t1+ln)δ)eλ0(t1+ln).\left\|X_{\varepsilon}(t_{1}+l,n_{1};X_{0})-X_{\varepsilon}(t_{1}+l,n_{2};X_{0})\right\|_{H}^{2}\leq\left(n+\sqrt{\varepsilon}(t_{1}+l-n)^{\delta}\right)e^{-\lambda_{0}(t_{1}+l-n)}.
Proof.

Assume n1<n2n_{1}<n_{2}, then according to Lemma 5.7 and Lemma 5.12 (ii) that

Xε(t1+l,n1;X0)Xε(t1+l,n2;X0)H2\displaystyle\left\|X_{\varepsilon}(t_{1}+l,n_{1};X_{0})-X_{\varepsilon}(t_{1}+l,n_{2};X_{0})\right\|_{H}^{2}
=\displaystyle= Xε(t1+l,n;Xε(n,n1;X0))Xε(t1+l,n;Xε(n,n2;X0))H2\displaystyle\left\|X_{\varepsilon}(t_{1}+l,n;X_{\varepsilon}(n,n_{1};X_{0}))-X_{\varepsilon}(t_{1}+l,n;X_{\varepsilon}(n,n_{2};X_{0}))\right\|_{H}^{2}
\displaystyle\leq (n+ε(t1+ln)δ)eλ0(t1+ln).\displaystyle\left(n+\sqrt{\varepsilon}(t_{1}+l-n)^{\delta}\right)e^{-\lambda_{0}(t_{1}+l-n)}.

We first consider the uniqueness of the solution of the infinite horizon stochastic integral Eq. (5.10), which is a crucial technical condition to obtain the stationarity solution XεX_{\varepsilon}^{*} of Eq. (2.1).

Lemma 5.14.

For every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), where ε0\varepsilon_{0} is defined in Lemma 5.11. Assume that Xε(t,ω)(t)X^{*}_{\varepsilon}(t,\omega)(t\in\mathbb{R}) is a ((),(H))\big{(}\mathcal{B}(\mathbb{R})\otimes\mathcal{F},\mathcal{B}(H)\big{)}-measurable, (t)t(\mathcal{F}_{-\infty}^{t})_{t\in\mathbb{R}}-adapted process, and for any N+,Xε(t,ω)|t[N,N]N\in\mathbb{Z}^{+},X^{*}_{\varepsilon}(t,\omega)|_{t\in[-N,N]} C([N,N];H)\in C([-N,N];H). Moreover, if Xε(t,ω)X_{\varepsilon}^{*}(t,\omega) satisfies the following equation in HH for any tt\in\mathbb{R}

Xε(t)=tSA(ts)F(Xε)ds+εtSA(ts)B(Xε)dWs.X_{\varepsilon}^{*}(t)=\int_{-\infty}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}\right)\mathrm{d}W_{s}. (5.10)

and

supt𝔼Xε(t)H2<,\sup_{t\in\mathbb{R}}\mathbb{E}\left\|X_{\varepsilon}^{*}(t)\right\|_{H}^{2}<\infty, (5.11)

then Xε(t)X_{\varepsilon}^{*}(t) is unique.

Proof.

For any n+n\in\mathbb{N}^{+} and n<t-n<t,

Xε(t)=\displaystyle X_{\varepsilon}^{*}(t)= tSA(ts)F(Xε)ds+εtSA(ts)B(Xε(s))dWs\displaystyle\int_{-\infty}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s}
=\displaystyle= nSA(ts)F(Xε(s))ds+εnSA(ts)B(Xε(s))dWs\displaystyle\int_{-\infty}^{-n}S_{A}(t-s)F\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{-n}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s}
+ntSA(ts)F(Xε(s))ds+εntSA(ts)B(Xε(s))dWs\displaystyle+\int_{-n}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-n}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s}
=\displaystyle= Xε(n)+ntSA(ts)F(Xε(s))ds+εntSA(ts)B(Xε(s))dWs.\displaystyle X_{\varepsilon}^{*}(-n)+\int_{-n}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-n}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s}.

Assume that YεY_{\varepsilon}^{*} is another solution of Eq. (5.10), then

Yε(t)=Yε(n)+ntSA(ts)F(Yε(s))ds+εntSA(ts)B(Yε(s))dWs,Y_{\varepsilon}^{*}(t)=Y_{\varepsilon}^{*}(-n)+\int_{-n}^{t}S_{A}(t-s)F\left(Y_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-n}^{t}S_{A}(t-s)B\left(Y_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s},

which imply that XεX_{\varepsilon}^{*} and YεY_{\varepsilon}^{*} are solutions of Eq. (2.1) with initial value Xε(n)X_{\varepsilon}^{*}(-n) and Yε(n)Y_{\varepsilon}^{*}(-n) at time n-n. It follows from Lemma 5.12 (ii) that there exists ε0,δ,λ0>0\varepsilon_{0},\delta,\lambda_{0}>0 and a sequence almost surely finite stopping time {nε}ε(0,ε0)\{{\vec{n}}_{\varepsilon}\}_{\varepsilon\in(0,\varepsilon_{0})} such that for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), s>0,ts>0,t\in\mathbb{R} when n>nεn>{\vec{n}}_{\varepsilon},

Xε(t+s,tn;Xε(tn))Yε(t+s,tn;Yε(tn))H2\displaystyle\|X_{\varepsilon}^{*}(t+s,t-n;X_{\varepsilon}^{*}(t-n))-Y_{\varepsilon}^{*}(t+s,t-n;Y_{\varepsilon}^{*}(t-n))\|_{H}^{2}
(n+ε(s+n)δ)eλ0(s+n),\displaystyle\leq\left(n+\sqrt{\varepsilon}(s+n)^{\delta}\right)e^{-\lambda_{0}(s+n)},

under the condition (5.11), let nn\rightarrow\infty, we can get that for any η>0\eta>0

Xε(t+s)Yε(t+s)H2η,\|X_{\varepsilon}^{*}(t+s)-Y_{\varepsilon}^{*}(t+s)\|_{H}^{2}\leq\eta,

it implies Xε()=Yε(),a.s.X_{\varepsilon}^{*}(\cdot)=Y_{\varepsilon}^{*}(\cdot),\quad a.s.

The following Lemma illustrate that if Eq. (5.10) has unique solution XεX_{\varepsilon}^{*}, then XεX_{\varepsilon}^{*} is also the stationary solution of Eq. (2.1).

Lemma 5.15.

If XεX_{\varepsilon}^{*} satisfies all conditions and (5.10), (5.11) in Lemma 5.14, then for any ll\in\mathbb{R}

Xε(+l,ω)=Xε(,θ(l,ω)),a.s.X_{\varepsilon}^{*}(\cdot+l,\omega)=X_{\varepsilon}^{*}(\cdot,\theta(l,\omega)),\quad a.s.
Proof.

Since θ\theta is \mathbb{P}-preserving on the probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), for any ll\in\mathbb{R}, we have, almost surly,

Xε(t,θ(l,ω))\displaystyle X_{\varepsilon}^{*}(t,\theta(l,\omega)) =\displaystyle= tSA(ts)F(Xε(s,θ(l,ω)))ds+εtSA(ts)B(Xε(s,θ(l,ω)))dWs(θ(l,ω))\displaystyle\int_{-\infty}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}\left(s,\theta(l,\omega)\right)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s,\theta(l,\omega))\right)\mathrm{d}W_{s}(\theta(l,\omega))
=\displaystyle= t+lSA(t+ls)F(Xε(sl,θ(l,ω)))ds\displaystyle\int_{-\infty}^{t+l}S_{A}(t+l-s)F\left(X_{\varepsilon}^{*}(s-l,\theta(l,\omega))\right)\mathrm{d}s
+εt+lSA(t+ls)B(Xε(sl,θ(l,ω)))dWs(ω),\displaystyle+\sqrt{\varepsilon}\int_{-\infty}^{t+l}S_{A}(t+l-s)B\left(X_{\varepsilon}^{*}(s-l,\theta(l,\omega))\right)\mathrm{d}W_{s}(\omega),

set X~ε(t,ω)=Xε(tl,θ(l,ω))\tilde{X}_{\varepsilon}^{*}(t,\omega)=X_{\varepsilon}^{*}(t-l,\theta(l,\omega)), then

X~ε(t+l,ω)=t+lSA(t+ls)F(X~ε(s,ω))ds+εt+lSA(t+ls)B(X~ε(s,ω))dWs(ω),\tilde{X}_{\varepsilon}^{*}(t+l,\omega)=\int_{-\infty}^{t+l}S_{A}(t+l-s)F\left(\tilde{X}_{\varepsilon}^{*}(s,\omega)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-\infty}^{t+l}S_{A}(t+l-s)B\left(\tilde{X}_{\varepsilon}^{*}(s,\omega)\right)\mathrm{d}W_{s}(\omega),

it follows from the uniqueness of Eq. (5.10) that for any ll\in\mathbb{R}

Xε(+l,ω)=Xε(,θ(l,ω)),a.s.X_{\varepsilon}^{*}(\cdot+l,\omega)=X_{\varepsilon}^{*}(\cdot,\theta(l,\omega)),\quad a.s.

Next, we will construct the solution of Eq. (5.10). For n+n\in\mathbb{Z}^{+}, ε>0\varepsilon>0, we define {Xεn}n=1\left\{X^{n}_{\varepsilon}\right\}_{n=1}^{\infty} by

Xεn(t,ω)={Xε(t,n;ω,0),tn,0,t<n.X^{n}_{\varepsilon}(t,\omega)=\begin{cases}X_{\varepsilon}(t,-n;\omega,0),&t\geq-n,\\ 0,&t<-n.\\ \end{cases}
Lemma 5.16.

There exists ε0>0\varepsilon_{0}>0 such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), N+N\in\mathbb{Z}^{+}, Xεn(,ω)Xε(,ω)X^{n}_{\varepsilon}(\cdot,\omega)\rightarrow X_{\varepsilon}^{*}(\cdot,\omega) in C([N,N];H)C([-N,N];H) as nn\rightarrow\infty. Moreover, XεX_{\varepsilon}^{*} satisfies the backward infinite horizon stochastic integral Eq. (5.10) and (5.11).

Proof.

It follows from Lemma 5.13 that there exists ε0>0\varepsilon_{0}>0, such that for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), N+N\in\mathbb{Z}^{+}, XεnX^{n}_{\varepsilon} is a Cauchy sequence in C([N,N];H)C([-N,N];H). And since the space C([N,N];H)C([-N,N];H) is complete, there exists a XεX_{\varepsilon}^{*} such that limnXεn=Xε\lim_{n\rightarrow\infty}X^{n}_{\varepsilon}=X_{\varepsilon}^{*} in C([N,N];H)C([-N,N];H). Since NN is arbitrary, Xε(t,ω)X_{\varepsilon}^{*}(t,\omega) is defined for all time, and from Lemma 5.7, we have supnsupt𝔼Xεn(t)H2<\sup_{n}\sup_{t\in\mathbb{R}}\mathbb{E}\left\|X^{n}_{\varepsilon}(t)\right\|_{H}^{2}<\infty. This implies that

supt𝔼Xε(t)H2<.\sup_{t\in\mathbb{R}}\mathbb{E}\left\|X_{\varepsilon}^{*}(t)\right\|_{H}^{2}<\infty. (5.12)

We will through two steps to prove XεX_{\varepsilon}^{*} satisfies Eq. (5.10).

Step 1. Firstly, we will prove that for any tt0t\geq t_{0}, XX^{*} satisfies

Xε(t)=SA(tt0)Xε(t0)+t0tSA(ts)F(Xε(s))ds+εt0tSA(ts)B(Xε(s))dWs.X_{\varepsilon}^{*}(t)=S_{A}(t-t_{0})X_{\varepsilon}^{*}(t_{0})+\int_{t_{0}}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s)\right)\mathrm{d}W_{s}. (5.13)

For any fixed t0t_{0}, there exists N+N\in\mathbb{N}^{+} such that t0Nt_{0}\geq-N,

Xεn(t,ω)=\displaystyle X^{n}_{\varepsilon}(t,\omega)= nt0SA(ts)F(Xε(s,n;ω,0))ds+εnt0SA(ts)B(Xε(s,n;ω,0))dWs\displaystyle\ \int_{-n}^{t_{0}}S_{A}(t-s)F\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{-n}^{t_{0}}S_{A}(t-s)B\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}W_{s}
+t0tSA(ts)F(Xε(s,n;ω,0))ds+εt0tSA(ts)B(Xε(s,n;ω,0))dWs\displaystyle+\int_{t_{0}}^{t}S_{A}(t-s)F\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}s+\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)B\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}W_{s}
=\displaystyle= Stt0Xε(t0,n;ω,0)+t0tSA(ts)F(Xε(s,n;ω,0))ds\displaystyle S_{t-t_{0}}X_{\varepsilon}(t_{0},-n;\omega,0)+\int_{t_{0}}^{t}S_{A}(t-s)F\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}s
+εt0tSA(ts)B(Xε(s,n;ω,0))dWs.\displaystyle+\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)B\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}W_{s}.

We have proved that Xεn(,ω)Xε(,ω)X^{n}_{\varepsilon}(\cdot,\omega)\rightarrow X_{\varepsilon}^{*}(\cdot,\omega) as nn\rightarrow\infty in C([N,N];H)C([-N,N];H) for any ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), then according to Hypothesis 2.1 (ii), for any t[N,N]t\in[-N,N],

limnt0tSA(ts)F(Xε(s,n;ω,0))dst0tSA(ts)F(Xε(s,ω))dsH2\displaystyle\lim_{n\rightarrow\infty}\left\|\int_{t_{0}}^{t}S_{A}(t-s)F\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}s-\int_{t_{0}}^{t}S_{A}(t-s)F\left(X_{\varepsilon}^{*}(s,\omega)\right)\mathrm{d}s\right\|_{H}^{2}
\displaystyle\lesssim limnt0t(ts)α(Xε(s,n;ω,0)Xε(s,ω))H2ds\displaystyle\lim_{n\rightarrow\infty}\int_{t_{0}}^{t}(t-s)^{\alpha}\|(X_{\varepsilon}(s,-n;\omega,0)-X_{\varepsilon}^{*}(s,\omega))\|_{H}^{2}\mathrm{d}s
\displaystyle\lesssim limnsups[N,N](Xε(s,n;ω,0)Xε(s,ω))H2t0t(ts)αds=0,\displaystyle\lim_{n\rightarrow\infty}\sup_{s\in[-N,N]}\|(X_{\varepsilon}(s,-n;\omega,0)-X_{\varepsilon}^{*}(s,\omega))\|_{H}^{2}\int_{t_{0}}^{t}(t-s)^{\alpha}\mathrm{d}s=0,

where the sign """\lesssim" means that the left side is less than or equal to the right side of a constant multiple.

It follows from the Burkholder-Davis-Gundy inequality and Remark 2.1 (ii) and (iii) that

limn𝔼{supt[N,N]εt0tSA(ts)[B(Xε(s,n;ω,0))B(Xε(s,ω))]dWs}2\displaystyle\lim_{n\rightarrow\infty}\mathbb{E}\left\{\sup_{t\in[-N,N]}\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)\left[B\left(X_{\varepsilon}(s,-n;\omega,0)\right)-B\left(X_{\varepsilon}^{*}(s,\omega)\right)\right]\mathrm{d}W_{s}\right\}^{2}
\displaystyle\lesssim εlimnt0N𝔼|SA(ts)[B(Xε(s,n;ω,0))B(Xε(s,ω))]|LQ2ds\displaystyle\varepsilon\lim_{n\rightarrow\infty}\int_{t_{0}}^{N}\mathbb{E}|S_{A}(t-s)\left[B\left(X_{\varepsilon}(s,-n;\omega,0)\right)-B\left(X_{\varepsilon}^{*}(s,\omega)\right)\right]|_{L_{Q}}^{2}\mathrm{d}s
\displaystyle\lesssim εlimnt0N𝔼|SA(ts)|L2|B(Xε(s,n;ω,0))B(Xε(s,ω)))|LQ2ds\displaystyle\varepsilon\lim_{n\rightarrow\infty}\int_{t_{0}}^{N}\mathbb{E}|S_{A}(t-s)|_{L}^{2}|B\left(X_{\varepsilon}(s,-n;\omega,0)\right)-B\left(X_{\varepsilon}^{*}(s,\omega))\right)|_{L_{Q}}^{2}\mathrm{d}s
\displaystyle\lesssim εβ2limnt0N𝔼Xε(s,n;ω,0)Xε(s,ω)H2ds=0.\displaystyle\varepsilon\beta^{2}\lim_{n\rightarrow\infty}\int_{t_{0}}^{N}\mathbb{E}\left\|X_{\varepsilon}(s,-n;\omega,0)-X_{\varepsilon}^{*}(s,\omega)\right\|_{H}^{2}\mathrm{d}s=0.

Thus there exists a subsequence still set as {Xεn}n=1\{X^{n}_{\varepsilon}\}_{n=1}^{\infty} such that

limnεt0tSA(ts)B(Xε(s,n;ω,0))dWs=εt0tSA(ts)B(Xε(s,ω))dWs,\lim_{n\rightarrow\infty}\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)B\left(X_{\varepsilon}(s,-n;\omega,0)\right)\mathrm{d}W_{s}=\sqrt{\varepsilon}\int_{t_{0}}^{t}S_{A}(t-s)B\left(X_{\varepsilon}^{*}(s,\omega)\right)\mathrm{d}W_{s},

in C([N,N];H)C([-N,N];H) a.s.a.s. At the same time, Xεn(t),SA(tt0)Xεn(t0)X^{n}_{\varepsilon}(t),S_{A}(t-t_{0})X^{n}_{\varepsilon}(t_{0}) converge strongly to XεX^{*}_{\varepsilon} and SA(tt0)Xε(t0)S_{A}(t-t_{0})X^{*}_{\varepsilon}(t_{0}) in HH respectively, hence let NN\rightarrow\infty, Eq. (5.13) holds for any tt0t\geq t_{0}.

Step 2. Finally, we will prove that XεX^{*}_{\varepsilon} satisfies Eq. (5.10).
It follows from Eq. (5.13), for any 0<m<n0<m<n,

nmSA(r)F(Xε(r,ω))dr\displaystyle\int_{-n}^{-m}S_{A}(-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r
=\displaystyle= εnmSA(r)B(Xε(r,ω))dW(r)T(n)Xε(n)+T(m)Xε(m).\displaystyle-\sqrt{\varepsilon}\int_{-n}^{-m}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r)-T(n)X^{*}_{\varepsilon}(-n)+T(m)X^{*}_{\varepsilon}(-m).

Thus

𝔼nmSA(r)F(Xε(r,ω))drH2\displaystyle\mathbb{E}\left\|\int_{-n}^{-m}S_{A}(-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r\right\|_{H}^{2}
\displaystyle\leq 2(𝔼εnmSA(r)B(Xε(r,ω))dW(r)H2+𝔼SA(n)Xε(n)H2+𝔼SA(m)Xε(m)H2)\displaystyle 2\left(\mathbb{E}\left\|\sqrt{\varepsilon}\int_{-n}^{-m}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r)\right\|_{H}^{2}+\mathbb{E}\left\|S_{A}(n)X^{*}_{\varepsilon}(-n)\right\|_{H}^{2}+\mathbb{E}\left\|S_{A}(m)X^{*}_{\varepsilon}(-m)\right\|_{H}^{2}\right)
=\displaystyle= 2(I+II+III).\displaystyle 2(I+II+III).

It follows from the Itô equality and Remark 2.1 (ii), (iii) that

I=\displaystyle I= 𝔼εnmSA(r)B(Xε(r,ω))dW(r)H2=ε𝔼nm|SA(r)B(Xε(r,ω))Q12|L22dr\displaystyle\mathbb{E}\left\|\sqrt{\varepsilon}\int_{-n}^{-m}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r)\right\|_{H}^{2}=\varepsilon\mathbb{E}\int_{-n}^{-m}|S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))Q^{\frac{1}{2}}|_{L_{2}}^{2}\mathrm{d}r
\displaystyle\leq ε𝔼nm|SA(r)|L2|B(Xε(r,ω))|LQ2dr\displaystyle\varepsilon\mathbb{E}\int_{-n}^{-m}|S_{A}(-r)|_{L}^{2}|B(X^{*}_{\varepsilon}(r,\omega))|_{L_{Q}}^{2}\mathrm{d}r
\displaystyle\leq εD2nm|SA(r)|L2dr,\displaystyle\varepsilon D^{2}\int_{-n}^{-m}|S_{A}(-r)|_{L}^{2}\mathrm{d}r,

is a Cauchy sequence, we obtain that I0I\rightarrow 0 as m,nm,n\rightarrow\infty.

Moreover, because II|SA(n)|L2𝔼Xε(n)H2II\leq|S_{A}(n)|_{L}^{2}\mathbb{E}\left\|X^{*}_{\varepsilon}(-n)\right\|_{H}^{2} and III|SA(n)|L2𝔼Xε(m)H2III\leq|S_{A}(n)|_{L}^{2}\mathbb{E}\left\|X^{*}_{\varepsilon}(-m)\right\|_{H}^{2}. Then form Remark (2.1) (iii) we know IIII and IIIIII converge to 0 as m,nm,n\rightarrow\infty. So

ntSA(r)F(Xε(r,ω))drandntSA(r)B(Xε(r,ω))dW(r)\int_{-n}^{t}S_{A}(-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r\quad\text{and}\quad\int_{-n}^{t}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r)

are Cauchy sequences in L2(Ω,H)L^{2}(\Omega,H) with respect to nn for any tt\in\mathbb{R}, we get that

ntSA(r)F(Xε(r,ω))dr\displaystyle\int_{-n}^{t}S_{A}(-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r \displaystyle\longrightarrow tSA(r)F(Xε(r,ω))dr,\displaystyle\int_{-\infty}^{t}S_{A}(-r)F(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}r,
ntSA(r)B(Xε(r,ω))dW(r)\displaystyle\int_{-n}^{t}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r) \displaystyle\longrightarrow tSA(r)B(Xε(r,ω))dW(r),asninH.\displaystyle\int_{-\infty}^{t}S_{A}(-r)B(X^{*}_{\varepsilon}(r,\omega))\mathrm{d}W(r),\quad\text{as}~{}n\rightarrow\infty~{}\text{in}~{}H.

Furthermore, 𝔼SA(t+n)Xε(n)H2|SA(t+n)|L2𝔼Xε(n)H20,asn.\mathbb{E}\left\|S_{A}(t+n)X^{*}_{\varepsilon}(-n)\right\|_{H}^{2}\leq|S_{A}(t+n)|_{L}^{2}\mathbb{E}\left\|X^{*}_{\varepsilon}(-n)\right\|_{H}^{2}\rightarrow 0,~{}as~{}n\rightarrow\infty. Thus combining Eq. (5.13) we can get that XεX^{*}_{\varepsilon} satisfies Eq. (5.10). ∎

The proof of Theorem 2.6 in the following.

Proof.

Combining Lemma 5.15 and Lemma 5.16, XεX_{\varepsilon}^{*} be the stationary solution of Eq. (2.1) (in the sense of Definition 2.5). Moreover, Xε(t,ω)X_{\varepsilon}^{*}(t,\omega) satisfies Eq. (5.10) in HH for any tt\in\mathbb{R} and (5.11). ∎

Lemma 5.17.

The solution XεX_{\varepsilon}^{*} of the Eq. (5.10) is the unique stationary solution of Eq. (2.1).

Proof.

Let X~ε(,ω)\tilde{X}_{\varepsilon}^{*}(\cdot,\omega) be another stationary solution of Eq. (2.1). Denote X~ε(ω)=X~ε(0,ω)\tilde{X}_{\varepsilon}^{*}(\omega)=\tilde{X}_{\varepsilon}^{*}(0,\omega) and Xε(ω)=Xε(0,ω)X_{\varepsilon}^{*}(\omega)=X_{\varepsilon}^{*}(0,\omega). It follows from the cocycle property and Lemma 5.11 that there exists ε0,δ,λ0>0\varepsilon_{0},\delta,\lambda_{0}>0 and a sequence almost surely finite stopping time {lε}ε(0,ε0)\{l_{\varepsilon}\}_{\varepsilon\in(0,\varepsilon_{0})} such that for every ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), s>0,ts>0,t\in\mathbb{R} when t>lεt>l_{\varepsilon},

X~ε(θ(t;ω))Xε(θ(t,ω))H2\displaystyle\left\|\tilde{X}_{\varepsilon}^{*}(\theta(t;\omega))-X_{\varepsilon}^{*}(\theta(t,\omega))\right\|_{H}^{2} =X~ε(t,0;ω,X~ε(ω))Xε(t,0;ω,Xε(ω))H2\displaystyle=\left\|\tilde{X}_{\varepsilon}(t,0;\omega,\tilde{X}_{\varepsilon}^{*}(\omega))-X_{\varepsilon}\left(t,0;\omega,X_{\varepsilon}^{*}(\omega)\right)\right\|_{H}^{2} (5.14)
X~ε(ω)Xε(ω)+εtδH2eλ0t.\displaystyle\leq\left\|\tilde{X}_{\varepsilon}^{*}(\omega)-X_{\varepsilon}^{*}(\omega)+\sqrt{\varepsilon}t^{\delta}\right\|_{H}^{2}e^{-\lambda_{0}t}.

As θ(t,ω)\theta(t,\omega) is a \mathbb{P}-preserving ergodic Wiener shift, we have, for any η>0,t>0\eta>0,t>0,

(X~ε(θ(t,ω))Xε(θ(t,ω))H2>η)=(X~ε(ω)Xε(ω)H2>η).\mathbb{P}\left(\left\|\tilde{X}_{\varepsilon}^{*}(\theta(t,\omega))-X_{\varepsilon}^{*}(\theta(t,\omega))\right\|_{H}^{2}>\eta\right)=\mathbb{P}\left(\left\|\tilde{X}_{\varepsilon}^{*}(\omega)-X_{\varepsilon}^{*}(\omega)\right\|_{H}^{2}>\eta\right).

In fact, by the inequality (5.14), we know that

{ω:X~ε(θ(t,ω))Xε(θ(t,ω))H2>η}{ω:X~ε(ω)Xε(ω)+εtδH2eλ0t>η},\left\{\omega:\left\|\tilde{X}_{\varepsilon}^{*}(\theta(t,\omega))-X_{\varepsilon}^{*}(\theta(t,\omega))\right\|_{H}^{2}>\eta\right\}\subset\left\{\omega:\left\|\tilde{X}_{\varepsilon}^{*}(\omega)-X_{\varepsilon}^{*}(\omega)+\sqrt{\varepsilon}t^{\delta}\right\|_{H}^{2}e^{-\lambda_{0}t}>\eta\right\},

and

(X~ε(θ(t,ω))Xε(θ(t,ω))H2>η)(X~ε(ω)Xε(ω)+εtδH2eλ0t>η).\mathbb{P}\left(\left\|\tilde{X}_{\varepsilon}^{*}(\theta(t,\omega))-X_{\varepsilon}^{*}(\theta(t,\omega))\right\|_{H}^{2}>\eta\right)\leq\mathbb{P}\left(\left\|\tilde{X}_{\varepsilon}^{*}(\omega)-X_{\varepsilon}^{*}(\omega)+\sqrt{\varepsilon}t^{\delta}\right\|_{H}^{2}e^{-\lambda_{0}t}>\eta\right).

However,

limt(X~ε(ω)Xε(ω)+εtδH2eλ0t>η)=0.\lim_{t\rightarrow\infty}\mathbb{P}\left(\left\|\tilde{X}_{\varepsilon}^{*}(\omega)-X_{\varepsilon}^{*}(\omega)+\sqrt{\varepsilon}t^{\delta}\right\|_{H}^{2}e^{-\lambda_{0}t}>\eta\right)=0.

This implies that (X~ε(ω)=Xε(ω))=1\mathbb{P}\left(\tilde{X}_{\varepsilon}^{*}(\omega)=X_{\varepsilon}^{*}(\omega)\right)=1. ∎

Appendix A: Weak convergence method in infinite intervals

The Boué-Dupuis formula in infinite intervals has been used directly without proof in [23] by Barashkov and Gubinelli. Although we believe that the experts in field of LDP and Gaussian measures are familiar with the Boué-Dupuis formula in infinite intervals, we still present the proof of Boué-Dupuis formula and the weak convergence approach in infinite intervals for the convenience of readers.

The main difference between the proof in infinite intervals and finite intervals appears in the lower bounded proof of Theorem A.5, since bounded functions are integrable in finite intervals, but not in infinite intervals. However, the integrable function of infinite intervals can be approximated by simple function, and then the proof for infinite intervals can be transformed into finite intervals. Other routine proofs are shown for the completeness of the present paper.

A.3 Boué-Dupuis formula in infinite interval

Lemma A.18.

(cf. Problem 3.19 in [16]) The following conditions are equivalent for a continuous martingale {Xt,t;0t<}\{X_{t},\mathcal{F}_{t};0\leq t<\infty\}.
(i). It is a uniformly integrable family of random variables.
(ii). It converges in L1L^{1}, as tt\longrightarrow\infty.
(iii). It converges \mathbb{P}-a.s. (as tt\longrightarrow\infty) to an integrable random variable XX_{\infty}, such that

{Xt,t;0t<}\{X_{t},\mathcal{F}_{t};0\leq t<\infty\}

is a martingale.
(iv). There exists an integrable random variable YY, such that Xt=𝔼(Y|)X_{t}=\mathbb{E}(Y|\mathcal{F}) \mathbb{P}-a.s., for every t0t\geq 0.

Let (𝕎,,μ)(\mathbb{W},\mathbb{H},\mu) be an abstract Wiener space. Namely, (𝕎,𝕎)(\mathbb{W},\|\cdot\|_{\mathbb{W}}) is a separable Banach space, (,)(\mathbb{H},\|\cdot\|_{\mathbb{H}}) is a separable Hilbert space densely and continuously embedded in 𝕎\mathbb{W}, and μ\mu is the Gaussian measure over 𝕎\mathbb{W}. If we identify the dual space \mathbb{H}^{\ast} with itself, then 𝕎\mathbb{W}^{\ast} may be viewed as a dense linear subspace of \mathbb{H} so that l(w)=l,wl(w)=\langle l,w\rangle_{\mathbb{H}} whenever l𝕎l\in\mathbb{W}^{\ast} and ww\in\mathbb{H} , where ,\langle\cdot,\cdot\rangle_{\mathbb{H}} denotes the inner product in \mathbb{H}.

We now recall some notations from [26] about the filtration in abstract Wiener space. In what follows, we fix a continuous and strictly monotonic resolution π={πt,t[0,1]}\pi=\{\pi_{t},t\in[0,1]\} of the identity in \mathbb{H},
(i). For each t[0,1]t\in[0,1], πt\pi_{t} is an orthogonal projection.
(ii). π0=0,π1=I\pi_{0}=0,\pi_{1}=I.
(iii). For 0s<t10\leq s<t\leq 1, πsπt\pi_{s}\mathbb{H}\subsetneq\pi_{t}\mathbb{H}.
(iv). For any hh\in\mathbb{H} and t[0,1]t\in[0,1], limstπsh=πth\lim_{s\rightarrow t}\pi_{s}h=\pi_{t}h.
For any hh\in\mathbb{H}, there exists a sequence hn𝕎h_{n}\in\mathbb{W}^{\ast} such that limnhnh=0\lim\limits_{n\rightarrow\infty}\|h_{n}-h\|_{\mathbb{H}}=0. Thus, there exists a δ(h)L2(𝕎,(𝕎),μ)\delta(h)\in L^{2}(\mathbb{W},\mathcal{B}(\mathbb{W}),\mu) such that

limn𝔼|hn()δ(h)|2=0.\lim_{n\rightarrow\infty}\mathbb{E}|h_{n}(\cdot)-\delta(h)|^{2}=0.

The δ(h)(w)\delta(h)(w) is also written as h,w\langle h,w\rangle, called the Skorohod integral of hh.

After taking f(x)=x1xf(x)=\frac{x}{1-x}, we can obtain a continuous and strictly monotonic resolution π={πt,0t}\pi=\{\pi_{t},0\leq t\leq\infty\} of the identity in \mathbb{H}, i.e.
(i). For each t[0,]t\in[0,\infty], πt\pi_{t} is an orthogonal projection.
(ii). π0=0,π=I\pi_{0}=0,\pi_{\infty}=I.
(iii). For 0s<t0\leq s<t\leq\infty, πsπt\pi_{s}\mathbb{H}\subsetneq\pi_{t}\mathbb{H}.
(iv). For any hh\in\mathbb{H} and t[0,]t\in[0,\infty], limstπsh=πth\lim_{s\rightarrow t}\pi_{s}h=\pi_{t}h.

If we take another transform, we can also get the filtration on \mathbb{R}.

Definition A.1.

The continuous filtation on (𝕎,μ)(\mathbb{W},\mu) is defined by

t:={δ(πth),h}𝒩,\mathcal{F}_{t}:=\mathcal{B}\{\delta(\pi_{t}h),h\in\mathbb{H}\}\vee\mathcal{N},

where 𝒩\mathcal{N} is the collection of all the null sets in 𝕎\mathbb{W} with respect to μ\mu. We write \mathcal{F}_{\infty} as \mathcal{F}, and remark that (𝕎)\mathcal{B}(\mathbb{W})\subset\mathcal{F}.

Below, we shall consider the filtered probability space (𝕎,,(t)0t,μ)(\mathbb{W},\mathcal{F},(\mathcal{F}_{t})_{0\leq t\leq\infty},\mu). If there is no special declaration, the expectation 𝔼\mathbb{E} and the term ``a.s"``a.s" are always taken with respect to the Wiener measure μ\mu.

Definition A.2.

For every 0t0\leq t\leq\infty, let 𝒞t\mathcal{C}_{t} be the collection of all cylindrical function with the form

F(w)=g(πth1,w,,πthn,w);gCb(n),h1,,hn𝕎.F(w)=g(\langle\pi_{t}h_{1},w\rangle,...,\langle\pi_{t}h_{n},w\rangle);\quad g\in C_{b}^{\infty}(\mathbb{R}^{n}),\quad h_{1},...,h_{n}\in\mathbb{W}^{\ast}. (A.1)

In particular, the elements in 𝒞t\mathcal{C}_{t} are measurable with respect to t\mathcal{F}_{t}, We write 𝒞1\mathcal{C}_{1} as 𝒞\mathcal{C}.

We have the following simple approximation result.

Lemma A.19.

For a fixed 0t0\leq t\leq\infty, lat FF be an t\mathcal{F}_{t} measurable and bounded function on 𝕎\mathbb{W} with bound NN. There exists a sequence Fk𝒞tF_{k}\in\mathcal{C}_{t} such that

Fk:=supw𝕎|Fk(w)|NandFkF,a.s.\|F_{k}\|_{\infty}:=\sup_{w\in\mathbb{W}}|F_{k}(w)|\leq N\quad and\quad F_{k}\rightarrow F,\quad a.s.

In particular, 𝒞t\mathcal{C}_{t} is dense in L2(𝕎,t,μ)L^{2}(\mathbb{W},\mathcal{F}_{t},\mu).

Definition A.3.

An \mathbb{H} valued random variable vv is called adapted to t\mathcal{F}_{t} if for every 0t0\leq t\leq\infty and h,πth,wth\in\mathbb{H},\langle\pi_{t}h,w\rangle_{\mathbb{H}}\in\mathcal{F}_{t}. All the adapted \mathbb{H} valued random variables in L2(𝕎,,μ;)L^{2}(\mathbb{W},\mathcal{F},\mu;\mathbb{H}) is denoted by a\mathcal{H}^{a}. The set of all bounded elements in a\mathcal{H}^{a} is denoted by ba\mathcal{H}_{b}^{a}, i.e.

ba:={va:v(w)Na.s.wforsomeN>0}.\mathcal{H}_{b}^{a}:=\{v\in\mathcal{H}^{a}:\|v(w)\|_{\mathbb{H}}\leq N\ a.s.-w\ for\ some\ N>0\}.

A vav\in\mathcal{H}^{a} is called simple if it has the following form

v(w)=i=0n1ξi(w)(πti+1πti)hi,ξ𝒞ti,hi,v(w)=\sum_{i=0}^{n-1}\xi_{i}(w)(\pi_{t_{i+1}}-\pi_{t_{i}})h_{i},\quad\xi\in\mathcal{C}_{t_{i}},\quad h_{i}\in\mathbb{H},

where 0=t0<t1<<tn<0=t_{0}<t_{1}<...<t_{n}<\infty. The set of all simple elements in a\mathcal{H}^{a} is denoted by 𝒮a\mathcal{S}^{a}. We write 𝒮ba:=𝒮aba\mathcal{S}_{b}^{a}:=\mathcal{S}^{a}\cap\mathcal{H}_{b}^{a}.

Proposition A.1.

a\mathcal{H}^{a} is a closed subspace of L2(𝕎,,μ;)L^{2}(\mathbb{W},\mathcal{F},\mu;\mathbb{H}), and 𝒮ba\mathcal{S}_{b}^{a} is dense in a\mathcal{H}^{a}.

Basing on this Proposition, for any vav\in\mathcal{H}^{a}, we can define Itô’s integral δ(v)\delta(v) such that

𝔼|δ(v)|2=𝔼v2.\mathbb{E}|\delta(v)|^{2}=\mathbb{E}\|v\|_{\mathbb{H}}^{2}.

On the other hand, for any L2\|\cdot\|_{L^{2}} norm finite and real Borel measurable function ff on [0,)[0,\infty) and hh\in\mathbb{H}, we may define the following integral with respect to the vector valued measure

0f(s)dπsh\int_{0}^{\infty}f(s)\mathrm{d}\pi_{s}h

such that

0f(s)dπsh2=0|f(s)|2dπsh,h.\Big{\|}\int_{0}^{\infty}f(s)\mathrm{d}\pi_{s}h\Big{\|}_{\mathbb{H}}^{2}=\int_{0}^{\infty}|f(s)|^{2}\mathrm{d}\langle\pi_{s}h,h\rangle_{\mathbb{H}}.

It is standard to prove the following result.

Lemma A.20.

Let ff be a left-continuous t\mathcal{F}_{t} adapted process the L2(+)\|\cdot\|_{L^{2}(\mathbb{R}^{+})} and L\|\cdot\|_{L^{\infty}} bounded by NN. Then for any hh\in\mathbb{H}, there exists a sequence vkh𝒮bav_{k}^{h}\in\mathcal{S}_{b}^{a} such that

vkh(w)Nh,a.s.\|v_{k}^{h}(w)\|_{\mathbb{H}}\leq N\cdot\|h\|_{\mathbb{H}},\quad a.s.

and

limk𝔼vkh0f(s)dπsh2=0.\lim_{k\rightarrow\infty}\mathbb{E}\|v_{k}^{h}-\int_{0}^{\infty}f(s)\mathrm{d}\pi_{s}h\|_{\mathbb{H}}^{2}=0.
Proof.

First of all, we define for every nn\in\mathbb{N}

fn(s):=j=0n2n1f(j2n)I[j2n,(j+1)2n)(s).f_{n}(s):=\sum_{j=0}^{n2^{n}-1}f(j2^{-n})I_{[j2^{-n},(j+1)2^{-n})}(s).

Then, by the dominated convergence theorem we have

limn𝔼0(fn(s)f(s))dπsh2=limn𝔼0|fn(s)f(s))|2dπsh,h=0.\lim_{n\rightarrow\infty}\mathbb{E}\Big{\|}\int_{0}^{\infty}(f_{n}(s)-f(s))\mathrm{d}\pi_{s}h\Big{\|}_{\mathbb{H}}^{2}=\lim_{n\rightarrow\infty}\mathbb{E}\int_{0}^{\infty}|f_{n}(s)-f(s))|^{2}\mathrm{d}\langle\pi_{s}h,h\rangle_{\mathbb{H}}=0.

For each nn\in\mathbb{N} and j=0,,n2n1j=0,...,n2^{-n}-1, by Lemma A.19 one can find Fn,kj𝒞j2nF_{n,k}^{j}\in\mathcal{C}_{j2^{-n}} such that

Fn,kjN,Fn,kjf(j2n),ask.\|F_{n,k}^{j}\|_{\infty}\leq N,\quad F_{n,k}^{j}\rightarrow f(j2^{-n}),\ as\ k\rightarrow\infty.

Finally, we define

vn,kh(w):=j=0n2n1Fn,kj(w)(π(j+1)2nπj2n)h.v_{n,k}^{h}(w):=\sum_{j=0}^{n2^{-n}-1}F_{n,k}^{j}(w)\cdot(\pi_{(j+1)2^{-n}}-\pi_{j2^{-n}})h.

By the diagonalization method, we may find the desired sequence vkhv_{k}^{h}. In fact, the condition L<N\|\cdot\|_{L^{\infty}}<N can be ignored. ∎

Proposition A.2.

Let 0<cFC0<c\leq F\leq C be a Borel measurable function on 𝕎\mathbb{W}. Then there exists a vav\in\mathcal{H}^{a} such that

𝔼(F|t)=𝔽exp{δ(πtv)12πtv2},0t.\mathbb{E}(F|\mathcal{F}_{t})=\mathbb{F}\cdot\exp\Big{\{}\delta(\pi_{t}v)-\frac{1}{2}\|\pi_{t}v\|_{\mathbb{H}}^{2}\Big{\}},\quad 0\leq t\leq\infty.
Proof.

Set Mt:=𝔼(F|t)M_{t}:=\mathbb{E}(F|\mathcal{F}_{t}). Then {Mt,t}\{M_{t},\mathcal{F}_{t}\} is a uniformly martingale bounded from above by CC and from below by cc. By Lemma A.18 and the representation formula of martingales, there is a uau\in\mathcal{H}^{a} such that

Mt=𝔽+δ(πtu).M_{t}=\mathbb{F}+\delta(\pi_{t}u).

Now define

v:=0dπtuMt,mt:=δ(πtv).v:=\int_{0}^{\infty}\frac{\mathrm{d}\pi_{t}u}{M_{t}},\ m_{t}:=\delta(\pi_{t}v).

Then, clearly vav\in\mathcal{H}^{a} and {mt,t}\{m_{t},\mathcal{F}_{t}\} is a martingale with square variation process tπtvt\rightarrow\|\pi_{t}v\|_{\mathbb{H}}. Thus, we have

Mt=𝔽+0tMs𝑑ms.M_{t}=\mathbb{F}+\int_{0}^{t}M_{s}dm_{s}.

The desired formula follows. ∎

We also need the following Clark-Ocone formula.

Proposition A.3.

For any F𝒞F\in\mathcal{C} with the form (A.1), it then holds

𝔼(F|t)=𝔼F+δ(πtv),0t,\mathbb{E}(F|\mathcal{F}_{t})=\mathbb{E}F+\delta(\pi_{t}v),\quad\forall 0\leq t\leq\infty,

where

v:=i=1n0𝔼[(ig)(h1,,,hn,)|t]dπthiba.v:=\sum_{i=1}^{n}\int_{0}^{\infty}\mathbb{E}\big{[}(\partial_{i}g)(\langle h_{1},\cdot\rangle,...,\langle h_{n},\cdot\rangle)|\mathcal{F}_{t}\big{]}\mathrm{d}\pi_{t}h_{i}\in\mathcal{H}_{b}^{a}.

Let ca\mathcal{H}_{c}^{a} be the set of all vav\in\mathcal{H}^{a} satisfying

𝔼[exp{δ(v)12v2}]=1.\mathbb{E}\big{[}\exp\{\delta(v)-\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\}\big{]}=1.

For vcav\in\mathcal{H}_{c}^{a}, we define

Tv(w):=wv(w)T_{v}(w):=w-v(w)

and

dμv=exp{δ(v)12v2}dμ.\mathrm{d}\mu_{v}=\exp\big{\{}\delta(v)-\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{\}}\mathrm{d}\mu. (A.2)

Then by the Girsanov theorem, we have for any A(𝕎)A\in\mathcal{B}(\mathbb{W})

μv(w:TvwA)=μ(A).\mu_{v}(w:T_{v}w\in A)=\mu(A).
Lemma A.21.

For vcav\in\mathcal{H}_{c}^{a}, let μv\mu_{v} be defined by (A.2). Then

R(μv||μ)=12𝔼μvv2.R(\mu_{v}||\mu)=\frac{1}{2}\mathbb{E}^{\mu_{v}}\|v\|_{\mathbb{H}}^{2}.
Theorem A.4.

Let FF be any bounded Borel measurable function on 𝕎\mathbb{W}. Then

log𝔼(eF)=infvca𝔼μv(F+12v2),-\log\mathbb{E}(e^{-F})=\inf_{v\in\mathcal{H}_{c}^{a}}\mathbb{E}^{\mu_{v}}\big{(}F+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)},

where μv\mu_{v} is defined by (A.2). Moreover, the infimum is unique attained at some v0cav_{0}\in\mathcal{H}_{c}^{a}.

Proposition A.4.

Let FF be any bounded Borel measurable function on 𝕎\mathbb{W}. For any v𝒮bav\in\mathcal{S}_{b}^{a}, there are two v~,v^𝒮ba\tilde{v},\hat{v}\in\mathcal{S}_{b}^{a} such that

𝔼μv~(F+12v~2)\displaystyle\mathbb{E}^{\mu_{\tilde{v}}}\big{(}F+\frac{1}{2}\|\tilde{v}\|_{\mathbb{H}}^{2}\big{)} =𝔼(F(+v)+12v2).\displaystyle=\mathbb{E}\big{(}F(\cdot+v)+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)}.
𝔼μv(F+12v2)\displaystyle\mathbb{E}^{\mu_{v}}\big{(}F+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)} =𝔼(F(+v^)+12v^2).\displaystyle=\mathbb{E}\big{(}F(\cdot+\hat{v})+\frac{1}{2}\|\hat{v}\|_{\mathbb{H}}^{2}\big{)}.

Moreover

R(μ(+v)||μ)=12𝔼v2,R(\mathcal{L}_{\mu}(\cdot+v)||\mu)=\frac{1}{2}\mathbb{E}\|v\|_{\mathbb{H}}^{2}, (A.3)

where μ(+v)\mathcal{L}_{\mu}(\cdot+v) denotes the law of ww+v(w)w\mapsto w+v(w) in (𝕎,)(\mathbb{W},\mathcal{F}) under μ\mu.

Reader interested in proof can refer to [26].

Theorem A.5.

Let FF be a bounded Borel measurable function on 𝕎\mathbb{W}. Then we have

log𝔼(eF)\displaystyle-\log\mathbb{E}(e^{-F}) =infva𝔼(F(+v)+12v2)\displaystyle=\inf_{v\in\mathcal{H}^{a}}\mathbb{E}\big{(}F(\cdot+v)+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)}
=infv𝒮ba𝔼(F(+v)+12v2).\displaystyle=\inf_{v\in\mathcal{S}_{b}^{a}}\mathbb{E}\big{(}F(\cdot+v)+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)}. (A.4)
Proof.

(Upper bound)Let vav\in\mathcal{H}^{a}. By Proposition A.1 we may choose a sequence of vn𝒮bav_{n}\in\mathcal{S}_{b}^{a} such that

limn𝔼vnv2.\lim_{n\rightarrow\infty}\mathbb{E}\|v_{n}-v\|_{\mathbb{H}}^{2}.

So, (w,vn)(w,v_{n}) converges in distribution to (w,v)(w,v) in 𝕎×\mathbb{W}\times\mathbb{H}, and μ(+vn)\mathcal{L}_{\mu}(\cdot+v_{n}) converges weakly to μ(+v)\mathcal{L}_{\mu}(\cdot+v). Noting that by Eq. (A.3)

supnR(μ(+vn)||μ)=12supn𝔼vn2<,\sup_{n}R(\mathcal{L}_{\mu}(\cdot+v_{n})||\mu)=\frac{1}{2}\sup_{n}\mathbb{E}\|v_{n}\|_{\mathbb{H}}^{2}<\infty,

and we have ([26] Lemma2.1(ii))

limn𝔼(F(+vn))=𝔼(F(+v)).\lim_{n\rightarrow\infty}\mathbb{E}(F(\cdot+v_{n}))=\mathbb{E}(F(\cdot+v)).

Therefore,by Theorem A.4, we get the upper bound

log𝔼(eF)limn𝔼(F(+vn)+12vn2)=𝔼(F(+v)+12v2).-\log\mathbb{E}(e^{-F})\leq\lim_{n\rightarrow\infty}\mathbb{E}\big{(}F(\cdot+v_{n})+\frac{1}{2}\|v_{n}\|_{\mathbb{H}}^{2}\big{)}=\mathbb{E}\big{(}F(\cdot+v)+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)}.

(Lower bound) We divided the proof into two steps.

Step 1: We first assume that F𝒞F\in\mathcal{C} with the form

F(w):=g(h1,w,,hn,w),gCb(n),h1,,hn𝕎.F(w):=g\big{(}\langle h_{1},w\rangle,...,\langle h_{n},w\rangle\big{)},\quad g\in C_{b}^{\infty}(\mathbb{R}^{n}),\quad h_{1},...,h_{n}\in\mathbb{W}^{\ast}.

By Proposition A.3, there is a vav\in\mathcal{H}^{a} such that for all 0t0\leq t\leq\infty

𝔼(eF|t)=𝔼(eF)exp{δ(πtv)12πtv2}.\mathbb{E}(e^{-F}|\mathcal{F}_{t})=\mathbb{E}(e^{-F})\cdot\exp\{\delta(\pi_{t}v)-\frac{1}{2}\|\pi_{t}v\|_{\mathbb{H}}^{2}\}. (A.5)

By Proposition A.3, we in fact find from the proof of Proposition A.2 that

v:=i=1n01𝔼(eF|t)𝔼[eF(ig)(h1,,,hn,)|t]dπthi.v:=\sum_{i=1}^{n}\int_{0}^{\infty}\frac{1}{\mathbb{E}(e^{-F}|\mathcal{F}_{t})}\cdot\mathbb{E}\big{[}e^{-F}\cdot(\partial_{i}g)(\langle h_{1},\cdot\rangle,...,\langle h_{n},\cdot\rangle)|\mathcal{F}_{t}\big{]}\mathrm{d}\pi_{t}h_{i}.

Thus, vbav\in\mathcal{H}_{b}^{a}. By Lemma A.20 there exists a sequence vn𝒮bav_{n}\in\mathcal{S}_{b}^{a} satisfying for some CF>0C_{F}>0

vnCFa.s.\|v_{n}\|_{\mathbb{H}}\leq C_{F}\quad a.s.

such that

limkn𝔼vnv2=0.\lim_{{k_{n}}\rightarrow\infty}\mathbb{E}\|v_{n}-v\|_{\mathbb{H}}^{2}=0.

By extracting a subsequence if necessary, we may further assume that

vnv20a.s,δ(vn)δ(v)a.s.\|v_{n}-v\|_{\mathbb{H}}^{2}\rightarrow 0\quad a.s,\quad\delta(v_{n})\rightarrow\delta(v)\ a.s.

we have by the dominated convergence theorem

𝔼μn(F+12vn2)𝔼μ(F+12v2)asn.\mathbb{E}^{\mu_{n}}\big{(}F+\frac{1}{2}\|v_{n}\|_{\mathbb{H}}^{2}\big{)}\rightarrow\mathbb{E}^{\mu}\big{(}F+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}\big{)}\quad as\ n\rightarrow\infty.

Here we have used the uniform integrability of {eδ(vn),n}\{e^{\delta(v_{n})},n\in\mathbb{N}\}. Moreover, by Eq. (A.5) and Lemma A.21

log𝔼(eF)=𝔼μv(F)+R(μv||μ)=𝔼μv(F+12v2).-\log\mathbb{E}(e^{-F})=\mathbb{E}^{\mu_{v}}(F)+R(\mu_{v}||\mu)=\mathbb{E}^{\mu_{v}}(F+\frac{1}{2}\|v\|_{\mathbb{H}}^{2}).

Step 2: Let FF be a bounded measurable function on (𝕎,)(\mathbb{W},\mathcal{F}). By Lemma A.19 we can choose a sequence Fn𝒞F_{n}\in\mathcal{C} such that FnF<\|F_{n}\|_{\infty}\leq\|F\|_{\infty}<\infty, and limnFn=Fμ\lim\limits_{n\rightarrow\infty}F_{n}=F\ \mu-a.s.. For any ϵ>0\epsilon>0 and FnF_{n}, by Step 1 there exists a vn𝒮bav_{n}\in\mathcal{S}_{b}^{a} such that

log𝔼(eFn)𝔼(Fn(+vn)+12vn2)ϵ.-\log\mathbb{E}(e^{-F_{n}})\geq\mathbb{E}\big{(}F_{n}(\cdot+v_{n})+\frac{1}{2}\|v_{n}\|_{\mathbb{H}}^{2}\big{)}-\epsilon. (A.6)

In view (A.3)\eqref{7.4} and (A.6)\eqref{7.6}, we have

supnR(μ(+vn)||μ)12supn𝔼vn22F+ϵ.\sup_{n}R(\mathcal{L}_{\mu}(\cdot+v_{n})||\mu)\leq\frac{1}{2}\sup_{n}\mathbb{E}\|v_{n}\|_{\mathbb{H}}^{2}\leq 2\|F\|_{\infty}+\epsilon.

So there is a subsequence nkn_{k} such that 𝔼vnk2\mathbb{E}\|v_{n_{k}}\|_{\mathbb{H}}^{2} convergence. We have

limk𝔼|Fnk(+vnk)F(+vnk)|=0.\lim_{k\rightarrow\infty}\mathbb{E}|F_{n_{k}}(\cdot+v_{n_{k}})-F(\cdot+v_{n_{k}})|=0.

Dominated convergence theorem gives that for sufficiently large kk,

log𝔼(eF)𝔼(F(+vnk)+12vnk2)2ϵ.-\log\mathbb{E}(e^{-F})\geq\mathbb{E}\big{(}F(\cdot+v_{n_{k}})+\frac{1}{2}\|v_{n_{k}}\|_{\mathbb{H}}^{2}\big{)}-2\epsilon.

Since vnk𝒮bav_{n_{k}}\in\mathcal{S}_{b}^{a}, we thus complete the proof of the lower bound. ∎

A.4 The proof of Laplace principle for infinite intervals

Since the large deviation principe is equivalent to Laplace principle in Polish space, we only need to prove the Laplace principle for infinite intervals, i.e. Theorem 2.10.

Proof.

In order to prove the Laplace principle, we must show that Eq. (2.6) holds for all real valued, bounded and continuous functions hh on space EE.

(Lower bound) Define

𝕎\displaystyle\mathbb{W} =C(;H),={h=hds;hH02ds<},\displaystyle=C(\mathbb{R};H),\quad\mathbb{H}=\big{\{}h=\int_{-\infty}^{\cdot}h^{\prime}\mathrm{d}s;\quad\int_{-\infty}^{\infty}\|h^{\prime}\|_{H_{0}}^{2}\mathrm{d}s<\infty\big{\}},
δ(v)\displaystyle\delta(v) =tB(v(t))dW(t),πθh=πθhhds=𝟙(,θ)hds,h.\displaystyle=\int_{-\infty}^{t}B(v(t))\mathrm{d}W(t),\quad\pi_{\theta}h=\pi_{\theta}h\int_{-\infty}^{\cdot}h^{\prime}\mathrm{d}s=\int_{-\infty}^{\cdot}\mathbbm{1}_{(-\infty,\theta)}h^{\prime}\mathrm{d}s,\quad\forall h\in\mathbb{H}.

It follows from the variational presentation (A.5) that

\displaystyle- log𝔼{exp[1εh(Xε())]}\displaystyle\log\mathbb{E}\left\{\exp\left[-\frac{1}{\varepsilon}h\left(X^{\varepsilon}(\cdot)\right)\right]\right\}
=infv𝒜𝔼(ε2+v(s)H02ds+h𝒢ε(W()+v(s)ds)).\displaystyle=\inf_{v\in\mathcal{A}}\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{\cdot}v(s)\mathrm{d}s\right)\right).

Then for every δ>0\delta>0, there exists v~ε𝒜\tilde{v}^{\varepsilon}\in\mathcal{A} such that

infv𝒜𝔼(ε2+v(s)H02ds+h𝒢ε(W()+v~(s)ds))\displaystyle\inf_{v\in\mathcal{A}}\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{+\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{\cdot}\tilde{v}(s)\mathrm{d}s\right)\right) (A.7)
𝔼(ε2+vε(s)H02ds+h𝒢ε(W()+.vε(s)ds))δ.\displaystyle\geq\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{+\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{.}v^{\varepsilon}(s)\mathrm{d}s\right)\right)-\delta.

We will prove that

lim infε0𝔼\displaystyle\liminf_{\varepsilon\longrightarrow 0}\mathbb{E} (ε2+vε(s)H02ds+h𝒢ε(W()+.vε(s)ds))\displaystyle\left(\frac{\varepsilon}{2}\int_{-\infty}^{+\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{.}v^{\varepsilon}(s)\mathrm{d}s\right)\right)
infxE{I(x)+h(x)}.\displaystyle\geq\inf_{x\in E}\{I(x)+h(x)\}.

We claim that we can assume without loss of generality that for all ε>0\varepsilon>0 and a.s.

ε+vε(s)H02dsN\varepsilon\int_{-\infty}^{+\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s\leq N (A.8)

for some finite number NN.

To see this, observe that if M:=hM:=\|h\|_{\infty} then supε>0𝔼(ε2+vε(s)H02ds)2M+δ<\sup_{\varepsilon>0}\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{+\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s\right)\leq 2M+\delta<\infty. Now define random variable

τNε:=inf{t:ε2tvε(s)H02dsN}.\tau_{N}^{\varepsilon}:=\inf\left\{t\in\mathbb{R}:\frac{\varepsilon}{2}\int_{-\infty}^{t}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s\geq N\right\}\wedge\infty.

The processes vε,N(s):=vε(s)𝟙(,τNε](s)v^{\varepsilon,N}(s):=v^{\varepsilon}(s)\mathbbm{1}_{(-\infty,\tau_{N}^{\varepsilon}]}(s) are in 𝒜\mathcal{A}, 𝟙\mathbbm{1} being as before the indicator function, and furthermore

{vεvε,N}{ε2vε(s)H02dsN}2M+δN.\mathbb{P}\left\{v^{\varepsilon}\neq v^{\varepsilon,N}\right\}\leq\mathbb{P}\left\{\frac{\varepsilon}{2}\int_{-\infty}^{\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s\geq N\right\}\leq\frac{2M+\delta}{N}.

This observation implies that the right side of inequality (A.7) is at most

𝔼(ε2vε,N(s)H02ds+h𝒢ε(W()+.vε,N(s)ds))2M(2M+δ)Nδ.\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{\infty}\|v^{\varepsilon,N}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{.}v^{\varepsilon,N}(s)\mathrm{d}s\right)\right)-\frac{2M(2M+\delta)}{N}-\delta.

Hence it suffices to prove with vε(s)v^{\varepsilon}(s) replaced by vε,N(s)v^{\varepsilon,N}(s). This proves the claim.

Henceforth we will assume that (A.8) holds. Pick a subsequence along which vε~:=εvε\tilde{v^{\varepsilon}}:=\sqrt{\varepsilon}v^{\varepsilon} converges in distribution to v~\tilde{v} as SNS_{N} -valued random elements.

We now have from Condition 2.1 (i) that

lim infε0𝔼(ε2vε(s)H02ds+h𝒢ε(W()+.vε(s)ds))\displaystyle\liminf_{\varepsilon\rightarrow 0}\mathbb{E}\left(\frac{\varepsilon}{2}\int_{-\infty}^{\infty}\|v^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\int_{-\infty}^{.}v^{\varepsilon}(s)\mathrm{d}s\right)\right)
=\displaystyle= lim infε0𝔼(12v~ε(s)H02ds+h𝒢ε(W()+1ε.v~ε(s)ds))\displaystyle\liminf_{\varepsilon\rightarrow 0}\mathbb{E}\left(\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}^{\varepsilon}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}^{.}\tilde{v}^{\varepsilon}(s)\mathrm{d}s\right)\right)
\displaystyle\geq 𝔼(12v~(s)H02ds+h(𝒢0(v~(s)ds)))\displaystyle\mathbb{E}\left(\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\left(\mathcal{G}^{0}\left(\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right)\right)\right)
\displaystyle\geq inf{(x,v)E×L2(;H0):x=𝒢0(.v~(s)ds)}{12v~(s)H02ds+h(x)}\displaystyle\inf_{\left\{(x,v)\in E\times L^{2}\left(\mathbb{R};H_{0}\right):x=\mathcal{G}^{0}\left(\int_{-\infty}^{.}\tilde{v}(s)\mathrm{d}s\right)\right\}}\left\{\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}(s)\|_{H_{0}}^{2}\mathrm{d}s+h(x)\right\}
\displaystyle\geq infxE{I(x)+h(x)}.\displaystyle\inf_{x\in E}\{I(x)+h(x)\}.

This completes the proof of the lower bound.

(Upper bound) Since hh is bounded infxE{I(x)+h(x)}<\inf_{x\in E}\{I(x)+h(x)\}<\infty. Let δ>0\delta>0 be arbitrary, and let x~E\tilde{x}\in E be such that

I(x~)+h(x~)infxE{I(x)+h(x)}+δ/2.I\left(\tilde{x}\right)+h\left(\tilde{x}\right)\leq\inf_{x\in E}\{I(x)+h(x)\}+\delta/2.

Choose v~L2(;H0)\tilde{v}\in L^{2}(\mathbb{R};H_{0}) such that

12v~(t)H02dtI(x~)+δ/2\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}(t)\|_{H_{0}}^{2}\mathrm{d}t\leq I\left(\tilde{x}\right)+\delta/2

and

x~=𝒢0(v~(s)ds).\tilde{x}=\mathcal{G}^{0}\left(\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right).

By using the variational presentation, for bounded and continuous functions hh, we could have

lim supε0εlog𝔼(exp{h(Xε(t,,ω)x0)/ε})\displaystyle\limsup_{\varepsilon\rightarrow 0}-\varepsilon\log\mathbb{E}\left(\exp\left\{-h\left(X^{\varepsilon}(t,-\infty,\omega)x_{0}\right)/\varepsilon\right\}\right)
=\displaystyle= lim supε0infv𝒜𝔼(12v(s)H02ds+h𝒢ε(W()+1εv(s)ds))\displaystyle\limsup_{\varepsilon\rightarrow 0}\inf_{v\in\mathcal{A}}\mathbb{E}\left(\frac{1}{2}\int_{-\infty}^{\infty}\|v(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}v(s)\mathrm{d}s\right)\right)
\displaystyle\leq lim supε0𝔼(12v~(s)H02ds+h𝒢ε(W()+1εv~(s)ds))\displaystyle\limsup_{\varepsilon\rightarrow 0}\mathbb{E}\left(\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}(s)\|_{H_{0}}^{2}\mathrm{d}s+h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right)\right)
=\displaystyle= 12v~(s)H02ds+lim supε0𝔼(h𝒢ε(W()+1εv~(s)ds))\displaystyle\frac{1}{2}\int_{-\infty}^{\infty}\|\tilde{v}(s)\|_{H_{0}}^{2}\mathrm{d}s+\limsup_{\varepsilon\rightarrow 0}\mathbb{E}\left(h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right)\right)
\displaystyle\leq I(x~)+δ/2+lim supε0𝔼(h𝒢ε(W()+1εv~(s)ds)).\displaystyle I\left(\tilde{x}\right)+\delta/2+\limsup_{\varepsilon\rightarrow 0}\mathbb{E}\left(h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right)\right).

Now from Condition 2.1 (i), as ε0\varepsilon\rightarrow 0

𝔼(h𝒢ε(W()+1εv~(s)ds)),\mathbb{E}\left(h\circ\mathcal{G}^{\varepsilon}\left(W(\cdot)+\frac{1}{\sqrt{\varepsilon}}\int_{-\infty}\tilde{v}(s)\mathrm{d}s\right)\right),

converges to h(𝒢0(v~(s)ds))=h(x~)h\left(\mathcal{G}^{0}\left(\int_{-\infty}^{\cdot}\tilde{v}(s)\mathrm{d}s\right)\right)=h\left(\tilde{x}\right). Thus

lim supε0εlog𝔼(exp{h(Xε(t,,ω)x0)/ε})infxE{I(x)+h(x)}+δ.\limsup_{\varepsilon\rightarrow 0}-\varepsilon\log\mathbb{E}\left(\exp\left\{-h\left(X^{\varepsilon}(t,-\infty,\omega)x_{0}\right)/\varepsilon\right\}\right)\leq\inf_{x\in E}\{I(x)+h(x)\}+\delta.

Since δ\delta is arbitrary, the proof is complete. The Condition 2.1 (ii) guaranteed that II is a good rate function. ∎

Acknowledgments

We are very grateful to Professor Z. Dong for his help and suggestions, and also to Dr. W.L. Zhang for participating in the discussion. Y. Liu appreciates Professor H.Z. Zhao and Professor C.R. Feng for their discussion about random stationary solutions, random periodic solutions, etc. Y. Liu thanks to Mr. C. Bai for their discussion about the random Hopf model.

Y. Liu is supported by CNNSF (No. 11731009, No.11926327) and Center for Statistical Science, PKU. Z.H. Zheng is supported by CNNSF (No.12031020, No.12090014), the Key Lab. of Random Complex Structures and Data Sciences, CAS and National Center for Mathematics and Interdisplinary Sciences, CAS.

References

References

  • [1] L. Arnold, Random dynamical systems, Springer Monographs in Mathematics, Springer-Verlag, Berlin, 1998.
  • [2] C. Bai, Some problems of ergodic properties of Hopf Model with perturbation, Master’s thesis, Peking University, 2019.
  • [3] M. Boué, P. Dupuis, A variational representation for certain functionals of Brownian motion, Ann. Probab. 26 (4) (1998) 1641-1659.
  • [4] Z. Brzezniak, S. Cerrai, Large deviations principle for the invariant measures of the 2D stochastic Navier-Stokes equations on a torus, J. Funct. Anal. 273 (6) (2017) 1891-1930.
  • [5] A. Budhiraja, P. Dupuis, A variational representation for positive functionals of infinite dimensional Brownian motion, Probab. Math. Statist. 20 (1) (2000) 39-61.
  • [6] S. Cerrai, M. Röckner, Large deviations for invariant measures of stochastic reaction-diffusion systems with multiplicative noise and non-Lipschitz reaction term, Ann. Inst. H. Poincaré Probab. Statist. 41 (1) (2005) 69-105.
  • [7] A. Dembo, O. Zeitouni, Large deviations techniques and applications, Vol. 38 of Stochastic Modelling and Applied Probability, Springer-Verlag, Berlin, 2010, corrected reprint of the second (1998) edition.
  • [8] J.D. Deuschel, D.W. Stroock, Large deviation, Springer-Verlag, Universitext, New York, 1984.
  • [9] P. Dupuis, R.S. Ellis, A weak convergence approach to the theory of large deviations, Wiley Series in Probability and Statistics: Probability and Statistics, John Wiley & Sons, Inc., New York, 1997, a Wiley-Interscience Publication.
  • [10] C.R. Feng, Y. Liu, H.Z. Zhao, Numerical approximation of random periodic solutions of stochastic differential equations, Z. Angew. Math. Phys. 68 (5) (2017) 68-199.
  • [11] C.R. Feng, B.Y. Qu, H.Z. Zhao, Random quasi-periodic paths and quasi-periodic measures of stochastic differential equations, J. Differential Equations 286 (2021) 119-163.
  • [12] C.R. Feng, H.Z. Zhao, Random periodic solutions of SPDEs via integral equations and Wiener-Sobolev compact embedding, J. Funct. Anal. 262 (10) (2012) 4377-4422.
  • [13] C.R. Feng, H.Z. Zhao, B. Zhou, Pathwise random periodic solutions of stochastic differential equations, J. Differential Equations 251 (1) (2011) 119-149.
  • [14] M.I. Freidlin, A.D. Wentzell, Random perturbations of dynamical systems, 3rd Edition, Vol. 260 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Springer, Heidelberg, 2012, translated from the 1979 Russian original by Joseph Szücs.
  • [15] E. Hopf, Repeated branching through loss of stability, an example. Proceedings of the conference on differential equations (dedicated to A. Weinstein), University of Maryland Book Store, College Park, Md., 1956, 49056. Also in Selected works of Eberhard Hopf with commentaries. Edited by Cathleen S. Morawetz, James B. Serrin and Yakov G. Sinai. American Mathematical Society, Providence, RI, 2002.
  • [16] I. Karatzas, S.E. Shreve, Brownian motion and stochastic calculus, 2nd Edition, Vol. 113 of Graduate Texts in Mathematics, Springer-Verlag, New York, 1991.
  • [17] A.N. Kolmogorov, S.V. Fomin, Elements of the theory of functions and functional analysis. Vol. 1. Metric and normed spaces, Graylock Press, Rochester, N.Y., 1957, translated from the first Russian edition by Leo F. Boron.
  • [18] W. Liu, M. Röckner, Stochastic partial differential equations: an introduction, Universitext, Springer, Cham, 2015.
  • [19] Y. Liu, H. Z. Zhao, Representation of pathwise stationary solutions of stochastic burgers’ equations, Stoch. Dynam. 9 (04) (2009) 613-634.
  • [20] D. Martirosyan, Large deviations for stationary measures of stochastic nonlinear wave equations with smooth white noise, Comm. Pure Appl. Math. 70 (9) (2017) 1754-1797.
  • [21] J. C. Mattingly, Ergodicity of 2d navier stokes equations withrandom forcing and large viscosity, Commun. Math. Phys. 206 (2) (1999) 273-288.
  • [22] S.-E.A. Mohammed, T.S. Zhang, H.Z. Zhao, The stable manifold theorem for semilinear stochastic evolution equations and stochastic partial differential equations, Mem. Amer. Math. Soc. 196 (917) (2008) 1-105.
  • [23] B. Nikolay, G. Massimiliano, A variational method for Φ34\Phi^{4}_{3} , Duke Math. J. 169 (17) (2020) 3339-3415.
  • [24] C. Prévôt, M. Röckner, A concise course on stochastic partial differential equations, Vol. 1905 of Lecture Notes in Mathematics, Springer, Berlin, 2007.
  • [25] S. Sritharan, P. Sundar, Large deviations for the two-dimensional Navier-Stokes equations with multiplicative noise, Stochastic Process. Appl. 116 (11) (2006) 1636-1659.
  • [26] X.C. Zhang, A variational representation for random functionals on abstract wiener spaces, J. Math. Kyoto. U 49 (3) (2009) 475-490.
  • [27] H.Z. Zhao, Z.H. Zheng, Random periodic solutions of random dynamical systems, J. Differential Equations 246 (5) (2009) 2020-2038.