Research article

Averaging principle on infinite intervals for stochastic ordinary differential equations with Lévy noise

  • Received: 16 January 2021 Accepted: 03 March 2021 Published: 12 March 2021
  • MSC : 34C29, 60H10, 60G51, 37B20, 34C27

  • In this paper, we establish an averaging principle on the infinite time intervals for semilinear stochastic ordinary differential equations with Lévy noise. In particular, under suitable conditions we prove that if the coefficients are Poisson stable (including periodic, quasi-periodic, almost periodic, almost automorphic etc), then there exists a unique L2-bounded solution of the original equation, which inherits the recurrence property of the coefficients, and the recurrent solution uniformly converges to the stationary solution of the averaged equation on the whole real axis in distribution sense.

    Citation: Xin Liu, Yan Wang. Averaging principle on infinite intervals for stochastic ordinary differential equations with Lévy noise[J]. AIMS Mathematics, 2021, 6(5): 5316-5350. doi: 10.3934/math.2021314

    Related Papers:

    [1] Yifei Wang, Haibo Gu, Ruya An . Averaging principle for space-fractional stochastic partial differential equations driven by Lévy white noise and fractional Brownian motion. AIMS Mathematics, 2025, 10(4): 9013-9033. doi: 10.3934/math.2025414
    [2] Xin Liu, Yongqi Hou . Almost automorphic solutions for mean-field SDEs driven by Lévy noise. AIMS Mathematics, 2025, 10(5): 11159-11183. doi: 10.3934/math.2025506
    [3] Michal Fečkan, Július Pačuta, Michal Pospíśil, Pavol Vidlička . Averaging methods for piecewise-smooth ordinary differential equations. AIMS Mathematics, 2019, 4(5): 1466-1487. doi: 10.3934/math.2019.5.1466
    [4] Jinghuai Liu, Litao Zhang . Square-mean asymptotically almost periodic solutions of second order nonautonomous stochastic evolution equations. AIMS Mathematics, 2021, 6(5): 5040-5052. doi: 10.3934/math.2021298
    [5] Meijiao Wang, Qiuhong Shi, Maoning Tang, Qingxin Meng . Stochastic differential equations in infinite dimensional Hilbert space and its optimal control problem with Lévy processes. AIMS Mathematics, 2022, 7(2): 2427-2455. doi: 10.3934/math.2022137
    [6] Huanhuan Zhang, Jia Mu . Periodic problem for non-instantaneous impulsive partial differential equations. AIMS Mathematics, 2022, 7(3): 3345-3359. doi: 10.3934/math.2022186
    [7] Hamood Ur Rehman, Aziz Ullah Awan, Sayed M. Eldin, Ifrah Iqbal . Study of optical stochastic solitons of Biswas-Arshed equation with multiplicative noise. AIMS Mathematics, 2023, 8(9): 21606-21621. doi: 10.3934/math.20231101
    [8] Yazid Alhojilan, Islam Samir . Investigating stochastic solutions for fourth order dispersive NLSE with quantic nonlinearity. AIMS Mathematics, 2023, 8(7): 15201-15213. doi: 10.3934/math.2023776
    [9] Shabir Ahmad, Saud Fahad Aldosary, Meraj Ali Khan . Stochastic solitons of a short-wave intermediate dispersive variable (SIdV) equation. AIMS Mathematics, 2024, 9(5): 10717-10733. doi: 10.3934/math.2024523
    [10] Boubaker Smii . Representation of the solution of a nonlinear molecular beam epitaxy equation. AIMS Mathematics, 2024, 9(12): 36012-36030. doi: 10.3934/math.20241708
  • In this paper, we establish an averaging principle on the infinite time intervals for semilinear stochastic ordinary differential equations with Lévy noise. In particular, under suitable conditions we prove that if the coefficients are Poisson stable (including periodic, quasi-periodic, almost periodic, almost automorphic etc), then there exists a unique L2-bounded solution of the original equation, which inherits the recurrence property of the coefficients, and the recurrent solution uniformly converges to the stationary solution of the averaged equation on the whole real axis in distribution sense.



    The core idea of the averaging method is to simplify the original system so as to get an effective one which can in some sense reflect the dynamics of the original one. It was first introduced in perturbation theory by Clairaut, Laplace and Lagrange. Some rigorous results on averaging principles can date back to Krylov and Bogolyubov's work [1], now called Krylov-Bogolyubov-Mitropolsky method [2,3,4]. Driven by applications the averaging principle has been developed in mechanics, mathematics, control and other areas. There are lots of works on averaging for deterministic systems which we will not mention here. Stochastic averaging principle is proposed by Stratonovich [5] for nonlinear oscillation problems with random noise. For averaging principles of stochastic differential equations among others [6,7,8,9,10,11,12,13,14].

    In previous studies for stochastic averaging, researchers usually consider Gaussian noise. Although it is an ideal noise, we do agree on another fact: Gaussian noise cannot describe discontinuous situations and cannot simulate large fluctuations. It is evident that random noises in practise are more likely to be non-Gaussian. Lévy processes are essentially stochastic processes with stationary and independent increments, and they are viewed as an important and useful class of non-Gaussian processes since they are the simplest examples of random motions whose sample paths are right-continuous and have a number of (at most countable) random jump discontinuities occurring at random times, on each finite time interval. The method of averaging has been applied to stochastic differential equations with Lévy noise, but in general to the initial problem on a finite interval, such as [15,16]. As for the averaging method on infinite intervals to deterministic equations, the book of Burd [17] provides a detail description.

    Motivated by the work of Cheban and Liu [18], we investigate the averaging principle on infinite intervals for linear and semilinear stochastic differential equations based on Lévy noise with Poisson stable (including stationary, periodic, quasi-periodic, almost periodic, almost automorphic, Birkhoff recurrent, almost recurrent in the sense of Bebutov, Levitan almost periodic, pseudo-periodic, pseudo-recurrent and Poisson stable) coefficients. Under some suitable conditions, the original equation has a unique bounded solution with the same recurrent property as the coefficients, see [19,20,21] for details. Besides we show that this recurrent solution uniformly converges to the unique stationary solution of the averaged equation on the whole real axis in distribution when the time scale goes to zero.

    The paper is organized as follows. Section 2 begins with definitions of Poisson stable functions, Lévy processes and their basic properties. We simply review Lévy-Itô decomposition, B. A. Shcherbakov's comparable method by character of recurrence, and the existence of recurrent solutions for stochastic differential equations. In the third and fourth sections, we respectively investigate the averaging principles for the following equations

    dY(t)=ε(A(t)Y(t)+f(t))dt+εg(t)dW(t)+ε|x|U<1F(t,x)˜N(dt,dx)+ε|x|U1G(t,x)N(dt,dx)

    and

    dY(t)=ε(A(t)Y(t)+f(t,Y(t)))dt+εg(t,Y(t))dW(t)+ε|x|U<1F(t,Y(t),x)˜N(dt,dx)+ε|x|U1G(t,Y(t),x)N(dt,dx),

    with ε a small positive parameter, operator A non-stationary but bounded, coefficients f,g,F,G Poisson stable in time, tR an infinite time interval.

    Let (Y,ρ) be a complete metric space. Denote by C(R,Y) the space of all continuous functions φ:RY equipped with the metric

    d(φ,ψ):=n=112ndn(φ,ψ)1+dn(φ,ψ)

    for any φ,ψC(R,Y), where dn(φ,ψ):=sup|t|nρ(φ(t),ψ(t)). Note that the metric d generates the compact-open topology on C(R,Y) and the space (C(R,Y),d) is a complete metric space [22,23,24,25].

    Remark 2.1. Let φ,φnC(R,Y) (nN). Then the following statements are equivalent:

    (i) d(φn,φ)0 as n;

    (ii) for each l>0, limnmax|t|lρ(φn(t),φ(t))=0;

    (iii) there exists a sequence {ln}+ such that limnmax|t|lnρ(φn(t),φ(t))=0.

    For given φC(R,Y), we use φh to denote the h-translation of φ, where φh(t):=φ(h+t) for tR. The hull of φ, denoted by H(φ), is the set of all the limits of φhn in C(R,Y), i.e. H(φ):={ψC(R,Y):ψ=limnφhn for some sequence {hn}R}.

    Remark 2.2. The mapping π:R×C(R,Y)C(R,Y) defined by π(h,φ)=φh is a dynamical system, i.e. π(0,φ)=φ, π(h1+h2,φ)=π(h2,π(h1,φ)) and the mapping π is continuous (see [22,26]). In particular, the mapping π restricted to R×H(φ) is a dynamical system.

    Now we recall the types of Poisson stable functions to be studied in this paper and the relations among them; see [22,23,24,25] for further details.

    Definition 2.3. A function φC(R,Y) is called stationary (respectively, τ-periodic) if φ(t)=φ(0) (respectively, φ(t+τ)=φ(t)) for all tR.

    Definition 2.4. A function φC(R,Y) is called quasi-periodic with the spectrum of frequencies ν1,ν2,,νm if the following conditions are fulfilled:

    (i) the numbers ν1,ν2,,νm are rationally independent;

    (ii) there exists a continuous function Φ:RmY such that Φ(t1+2π,t2+2π,,tm+2π)=Φ(t1,t2,,tm) for all (t1,t2,,tm)Rm;

    (iii) φ(t)=Φ(ν1t,ν2t,,νmt) for all tR.

    Definition 2.5. A function φC(R,Y) is said to be almost periodic if for each ε>0, the set T(φ,ε):={τR:suptRρ(φ(t+τ),φ(t))<ε} is relatively dense on R, i.e. there exists l=l(ε)>0 such that T(φ,ε)[a,a+l] for any aR. The set T(φ,ε) is called the set of ε-almost period of the function φ.

    Definition 2.6. A function φC(R,Y) is said to be pseudo-periodic in the positive (respectively, negative) direction if for each ε>0 and l>0 there exists an ε-almost period τ>l (respectively, τ<l) of the function φ. The function φ is called pseudo-periodic if it is pseudo-periodic in both directions.

    Remark 2.7. A function φC(R,Y) is pseudo-periodic in the positive (respectively, negative) direction if and only if there is a sequence {tn}+ (respectively, {tn}) such that φtn converges to φ uniformly with respect to (w.r.t.) tR as n.

    Definition 2.8. A function φC(R,Y) is called almost automorphic if and only if for any sequence {tn}R there are a subsequence {tn} and some function ψ:RY such that

    φ(t+tn)ψ(t)  and  ψ(ttn)φ(t)

    uniformly in t on every compact subset from R.

    Definition 2.9. (i) A number τR is said to be ε-shift for φC(R,Y) if d(φτ,φ)<ε; a function φC(R,Y) is called almost recurrent (in the sense of Bebutov) if for every ε>0 the set {τ:d(φτ,φ)<ε} is relatively dense.

    (ii) A function φC(R,Y) is called Lagrange stable if {φh:hR} is a relatively compact subset of C(R,Y).

    (iii) A function φC(R,Y) is called Birkhoff recurrent if it is almost recurrent and Lagrange stable.

    In what follows, we denote by (X,γ) a complete metric space.

    Definition 2.10. A function φC(R,Y) is called Levitan almost periodic if there exists an almost periodic function ψC(R,X) such that for any ε>0 there exists δ=δ(ε)>0 such that d(φτ,φ)<ε for all τT(ψ,δ).

    Remark 2.11. ([27,ChIV])

    (i) Every almost periodic function is Levitan almost periodic but the inverse statement is not true in general.

    (ii) A function φC(R,Y) is said to be almost automorphic if it is Levitan almost periodic and Lagrange stable.

    Definition 2.12. A function φC(R,Y) is called pseudo-recurrent if for any ε>0 and lR there exists a constant Ll such that for any τ0R there exists a number τ[l,L] satisfying

    sup|t|ε1ρ(φ(t+τ0+τ),φ(t+τ0))ε.

    Definition 2.13. A function φC(R,Y) is called Poisson stable in the positive (respectively, negative) direction if for every ε>0 and l>0 there exists τ>l (respectively, τ<l) such that d(φτ,φ)<ε. The function φ is called Poisson stable provided it is Poisson stable in both directions.

    Remark 2.14. ([23,24,25,28])

    (i) Every Birkhoff recurrent function is pseudo-recurrent, but the inverse is not always true.

    (ii) If the function φC(R,Y) is pseudo-recurrent, then every function ψH(φ) is pseudo-recurrent.

    (iii) If the function φC(R,Y) is Lagrange stable and every function ψH(φ) is Poisson stable, then φ is pseudo-recurrent.

    Finally, we remark that a Lagrange stable function is not Poisson stable in general, but all other types of functions introduced above are Poisson stable.

    Let φC(R,Y). Denote by Nφ (respectively, Mφ) the family of all sequences {tn}R such that φtnφ (respectively, {φtn} converges) in C(R,Y) as n. We denote by Nuφ (respectively, Muφ) the family of sequences {tn}Nφ (respectively, {tn}Mφ) such that φtn converges to φ (respectively, {φtn} converges) uniformly w.r.t. tR as n.

    Definition 2.15. A function φC(R,Y) is said to be comparable (by character of recurrence) with ψC(R,X) if NψNφ; φ is said to be strongly comparable (by character of recurrence) with ψ if MψMφ.

    Theorem 2.16. ([23,29,Chapter II]) Let φC(R,Y), ψC(R,X). Then the following statements hold.

    (i) MψMφ implies NψNφ, and hence strong comparability implies comparability.

    (ii) Let φ be comparable with ψ. If the function ψ is stationary (respectively, τ-periodic, Levitan almost periodic, almost recurrent, Poisson stable), then so is φ.

    (iii) Let φ be strongly comparable with ψ. If the function ψ is quasi-periodic with the spectrum of frequencies ν1, ν2, , νm (respectively, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable), then so is φ.

    (iv) Let φ be strongly comparable with ψ and ψ be Lagrange stable. If ψ is pseudo-periodic (respectively, pseudo-recurrent), then so is φ.

    Lemma 2.17. ([19]) Let φC(R,Y), ψC(R,X). The following statements hold:

    (i) If MuψMuφ, then NuψNuφ.

    (ii) If MuψMuφ and the function ψ is almost periodic, then so is φ.

    (iii) If NuψNuφ and the function ψ is pseudo periodic, then so is φ.

    Denote by BUC(R×Y,X) the space of all functions f:R×YX which are continuous in t uniformly w.r.t. y on every bounded subset QY and bounded on every bounded subset from R×Y. We endow the function space with the metric

    d(f,g):=n=112ndn(f,g)1+dn(f,g), (2.1)

    where

    dn(f,g):=sup|t|n, yQnγ(f(t,y),g(t,y))

    with QnY being bounded, QnQn+1 and Y=n1Qn. Note that (BUC(R×Y,X),d) is a complete metric space and d(fn,f)0 if and only if fn(t,y)f(t,y) uniformly w.r.t. (t,y) on every bounded subset from R×Y. For given fBUC(R×Y,X) and τR, denote the translation of f by fτ, i.e. fτ(t,y):=f(t+τ,y) for (t,y)R×Y, and the hull of f by H(f):=¯{fτ:τR} with the closure being taken under the metric d given by (2.1). Note that the mapping π:R×BUC(R×Y,X)BUC(R×Y,X) defined by π(τ,f):=fτ is a dynamical system, i.e. π(0,f)=f, π(τ1+τ2,f)=π(τ2,π(τ1,f)) and the mapping π is continuous. See [22,26] or [§ 1.1] for details.

    We use BC(Y,X) to denote the space of all continuous functions f:YX which are bounded on every bounded subset QY and equip the space with the metric

    d(f,g):=n=112ndn(f,g)1+dn(f,g),

    where dn(f,g):=supyQnγ(f(y),g(y)) with Qn the same as (2.1). It is immediate to check that (BC(Y,X),d) is a complete metric space.

    Remark 2.18. Let FBUC(R×Y,X) and f:RBC(Y,X) be a mapping defined by f(t):=F(t,). Note that for any FBUC(R×Y,X), we have MF=Mf and MuF=Muf. Here MF is the set of all sequences {tn} such that {Ftn} converges in the space BUC(R×Y,X); MuF is the set of all sequences {tn} such that {Ftn} converges in the space BUC(R×Y,X) uniformly w.r.t. tR.

    Throughout the paper, we fix a complete probability space (Ω,F,P), two real separable Hilbert spaces (H,||) and (U,||U), and a real separable Banach space (B,||B). Denote by L(H) (respectively, L(B)) the Hilbert (respectively, Banach) space of all bounded linear operators on H (respectively, B) endowed with operator norm (respectively, B). Denote by L(U,H) the Banach space of all bounded linear operators from U to H with the norm L(U,H).

    We now review the definition of Lévy processes and the important Lévy-Itô decomposition theorem (see [30,31]). In this paper, the Lévy processes we consider are U-valued.

    Definition 2.19. A U-valued stochastic process L=(L(t),t0) is called Lévy process if it has the following three properties:

    (i) L(0)=0 almost surely.

    (ii) L has stationary and independent increments, i.e. the law of L(t+h)L(t) does not depend on t and for all 0t0<t1<t2<...<tn<, random variables L(t1)L(t0), L(t2)L(t1), ..., L(tn)L(tn1) are independent.

    (iii) L is stochastically continuous, i.e. for all ϵ>0 and for all s>0

    limtsP(|L(t)L(s)|U>ϵ)=0.

    Since a Lévy process L is càdlàg, the associated jump process ΔL=(ΔL(t),t0) is given by ΔL(t)=L(t)L(t). Let B(U{0}) be the Borel field of U{0} and BB(U{0}). Define the random counting measure

    N(t,B)(ω):={0st:ΔL(s)(ω)B}=0stχB(ΔL(s)(ω)),

    where χB is the indicator function of B. We call ν():=E(N(1,)) the intensity measure of L. We say that BB(U{0}) is bounded below if 0¯B, the closure of B. Note that if B is bounded below, then N(t,B)< holds almost surely for all t0 and (N(t,B),t0) is a Poisson process with intensity ν(B). N is called Poisson random measure. For each t0 and B bounded below, define the compensated Poisson random measure by

    ˜N(t,B)=N(t,B)tν(B).

    Proposition 2.20 (Lévy-Itô decomposition). If L is a Lévy process in U, then there exists aU, a U-valued Q-Wiener process W and an independent Poisson random measure N on R+×(U{0}) with intensity measure ν such that for any t0,

    L(t)=at+W(t)+|x|U<1x˜N(t,dx)+|x|U1xN(t,dx). (2.2)

    Here the intensity measure ν satisfies

    U(|x|2U1)ν(dx)< (2.3)

    and ˜N is the compensated Poisson random measure of N.

    As for Q-Wiener processes and the stochastic integral based on them, the monograph [32] provides a thorough description. Assume that L1 and L2 are two independent, identically distributed Lévy processes with decompositions as in Proposition 2.19 and let

    L(t)={L1(t), for t0,L2(t), for t<0.

    Then L is a two-sided Lévy process. In this paper, we consider two-sided Lévy process L which is defined on the filtered probability space (Ω,F,P,(Ft)tR) and suppose that the covariance operator Q of W is of trace class, i.e. TrQ<.

    Remark 2.21. It follows from (2.3) that |x|U1ν(dx)<. For convenience, we set hereafter

    b:=|x|U1ν(dx).

    Remark 2.22. Note that the stochastic process ˜L=(˜L(t),tR) given by ˜L(t):=L(t+s)L(s) for some sR is also a two-sided Lévy process which shares the same law as L. In particular, when sR+, the similar conclusion holds for one-sided Lévy processes.

    Consider the linear homogeneous equation

    ˙y=A(t)y (2.4)

    on the space B, where AC(R,L(B)). Denote by TA(t,τ) the evolution (solving) operator of Eq (2.4), where TA(t,τ):=UA(t)U1A(τ) with UA(t) the Cauchy operator of Eq (2.4) (see [33]).

    Definition 2.23. Eq (2.4) is said to be uniformly asymptotically stable if there are positive constants K and ω such that

    TA(t,τ)BKeω(tτ)  for any tτ (t,τR). (2.5)

    Lemma 24. ([34,Chapter III]) Suppose that Eq (2.4) is uniformly asymptotically stable such that Eq (2.5) holds. Then for any tτ (t,τR) and ˜AH(A)

    T˜A(t,τ)BKeω(tτ),

    where H(A) denotes the closure in the space C(R,L(B)) of all translations {Ah:hR} with Ah(t):=A(t+h) for tR.

    We now consider the stochastic differential equation driven by Lévy noise

    dY(t)=(A(t)Y(t)+f(t,Y(t)))dt+g(t,Y(t))dL(t), (2.6)

    where A:RL(H), f:R×HH, g:R×HL(U,H); L is a U-valued Lévy process. By Lévy-Itô decomposition (2.2), Eq (2.6) reads

    dY(t)=(A(t)Y(t)+f(t,Y(t)))dt+g(t,Y(t))dW(t)+|x|U<1F(t,Y(t),x)˜N(dt,dx)+|x|U1G(t,Y(t),x)N(dt,dx), (2.7)

    where F and G are H-valued. It allows us to study large jumps with considerable probability. We set Ft:=σ{L(u):ut}.

    Definition 2.25. An Ft-adapted process {Y(t)}tR is called a mild solution of Eq (2.7) if it satisfies the corresponding stochastic integral equation

    Y(t)=TA(t,r)Y(r)+trTA(t,s)f(s,Y(s))ds+trTA(t,s)g(s,Y(s))dW(s)+tr|x|U<1TA(t,s)F(s,Y(s),x)˜N(ds,dx)+tr|x|U1TA(t,s)G(s,Y(s),x)N(ds,dx), (2.8)

    for all tr and each rR.

    Let P(B) be the space of all Borel probability measures on the space B endowed with the β metric:

    β(μ,ν):=sup{|fdμfdν|B:fBL1},for μ,νP(B).

    Here f varies in the space of bounded Lipschitz continuous real-valued functions on the space B with the norm

    fBL=Lip(f)+f,

    where

    Lip(f)=supxy|f(x)f(y)|B|xy|B,f=supxB|f(x)|B.

    A sequence {μn}P(B) is said to weakly converge to μ if fdμnfdμ for all fCb(B), where Cb(B) is the space of all bounded continuous real-valued functions on the space B. As we know that (P(B),β) is a separable complete metric space and that a sequence {μn} weakly converges to μ if and only if β(μn,μ)0 as n. See [35,§11.3] for β metric and related properties.

    Definition 2.26. A sequence of random variables {Yn} is said to converge in distribution to the random variable Y if the corresponding laws {μn} of {Yn} weakly converge to the law μ of Y, i.e. β(μn,μ)0.

    Definition 2.27. Let {φ(t)}tR be a mild solution of Eq (2.7). Then φ is called compatible (respectively, strongly compatible) in distribution if N(A,f,g,F,G)˜Nφ (respectively, M(A,f,g,F,G)˜Mφ), where ˜Nφ (respectively, ˜Mφ) means the set of all sequences {tn}R such that the sequence {φ(+tn)} converges to φ() (respectively, {φ(+tn)} converges) in distribution uniformly on any compact interval.

    Define for p2

    Lp(P;H):=Lp(Ω,F,P;H)={Y:ΩH|E|Y|p=Ω|Y|pdP<}

    with the norm

    YLp(P;H):=(Ω|Y|pdP)1p.

    Then (Lp(P;H),Lp(P;H)) is a Banach space. Set

    L2(P;L(U,H)):=L2(Ω,F,P;L(U,H))={Y:ΩL(U,H)|EY2L(U,H)=ΩY2L(U,H)dP<}

    and define a norm by

    YL2(P;L(U,H)):=(ΩY2L(U,H)dP)12.

    Note that (L2(P;L(U,H)),L2(P;L(U,H))) is a Banach space.

    Remark 2.28. If the operator QL(U), the space of bounded linear operators from U to U, is nonnegative, symmetric and TrQ<, then LQ12L2(U,H) for all LL(U,H), where the space L2(U,H) is a separable Hilbert space that consists of all Hilbert-Schmidt operators from U to H with inner product A,BL2(U,H):=kNAek,Bek and {ek}kN an orthonormal basis of U. For gL2(P;L(U,H)), we have gQ12L2(P;L2(U,H)) and denote gQ12L2(P;L2(U,H))=(EgQ122L2(U,H))12.

    Define

    L2(Pν;H):=L2(Ω×U,PU,Pν;H)={Y:Ω×UH|Ω×U|Y|2dPν=UE|Y|2ν(dx)<},

    where PU is the product σ-algebra on Ω×U and Pν=Pν. For YL2(Pν;H), let

    YL2(Pν;H):=(UE|Y|2ν(dx))12.

    Then L2(Pν;H) is a Hilbert space equipped with the norm L2(Pν;H).

    Denote by Cb(R,B) the Banach space of all continuous and bounded mappings φ:RB equipped with the norm φ:=sup{|φ(t)|B:tR}. Note that if fCb(R,B) and ˜fH(f), the hull of f, then ˜ff.

    Theorem 2.29. Consider the linear stochastic differential equation

    dY(t)=(A(t)Y(t)+f(t))dt+g(t)dW(t)+|x|U<1F(t,x)˜N(dt,dx)+|x|U1G(t,x)N(dt,dx). (2.9)

    Assume that ACb(R,L(H)); Eq (2.4) is uniformly asymptotically stable such that (2.5) holds; fCb(R,L2(P;H)), gCb(R,L2(P;L(U,H))), F,GCb(R,L2(Pν;H)). Suppose that W and N are the same as in Section 2.3. Then Eq (2.9) has a unique mild solution φCb(R,L2(P;H)) which satisfies

    φ(t)=tTA(t,τ)f(τ)dτ+tTA(t,τ)g(τ)dW(τ)+t|x|U<1TA(t,τ)F(τ,x)˜N(dτ,dx)+t|x|U1TA(t,τ)G(τ,x)N(dτ,dx),

    and this unique L2-bounded solution is strongly compatible in distribution (i.e. M(A,f,g,F,G)˜Mφε). Furthermore, Mu(A,f,g,F,G)˜Muφ, where ˜Muφ is the set of all sequences {tn} such that the sequence {φ(t+tn)} converges in distribution uniformly in tR.

    Proof. The proof is analogous to Theorem 3.3 in [21].

    Define

    L2(ν;H):=L2(U,BU,ν;H)={Y:UH|U|Y|2ν(dx)<},

    with BU being the Borel σ-algebra on U. For YL2(ν;H), let

    YL2(ν;H):=(U|Y|2ν(dx))12.

    Denote by C(R×Y,X) the space of all continuous functions f:R×YX, recalling that (Y,ρ) and (X,γ) are complete metric spaces.

    Consider Eq (2.7). Like in [21], assume that fC(R×H,H), gC(R×H,L(U,H)), F,GC(R×H,L2(ν;H)) and f,g,F,G satisfy the following conditions:

    (E1) There exists a number M0 such that for all tR

    |f(t,0)|M, g(t,0)Q12L2(U,H)M,

    |x|U<1|F(t,0,x)|2ν(dx)M2, |x|U1|G(t,0,x)|2ν(dx)M2.

    (E1') There exists a number M0 such that for some constant p>2 and all tR

    |f(t,0)|M, g(t,0)Q12L2(U,H)M,

    |x|U<1|F(t,0,x)|pν(dx)Mp, |x|U1|G(t,0,x)|pν(dx)Mp.

    (E2) There exists a number L0 such that for all tR and y1,y2H

    |f(t,y1)f(t,y2)|L|y1y2|, (g(t,y1)g(t,y2))Q12L2(U,H)L|y1y2|,

    |x|U<1|F(t,y1,x)F(t,y2,x)|2ν(dx)L2|y1y2|2,

    |x|U1|G(t,y1,x)G(t,y2,x)|2ν(dx)L2|y1y2|2.

    (E2 ') There exists a number L0 such that for some constant p>2 and tR, y1,y2H

    |f(t,y1)f(t,y2)|L|y1y2|, (g(t,y1)g(t,y2))Q12L2(U,H)L|y1y2|,

    |x|U<1|F(t,y1,x)F(t,y2,x)|pν(dx)Lp|y1y2|p,

    |x|U1|G(t,y1,x)G(t,y2,x)|pν(dx)Lp|y1y2|p.

    (E3) f,g,F,G are continuous in t uniformly w.r.t. Y on each bounded subset QH.

    Remark 2.30. (i) If f,g,F,G satisfy (E1)(E3), then fBUC(R×H,H), gBUC(R×H,L(U,H)), F,GBUC(R×H,L2(ν;H)) and H(f,g,F,G)BUC(R×H,H)×BUC(R×H,L(U,H))×BUC(R×H,L2(ν;H))×BUC(R×H,L2(ν;H)).

    (ii) If f,g,F,G satisfy (E1),(E2) (or (E1), (E2)) with the constants M and L, then every quadruplet (˜f,˜g,˜F,˜G)H(f,g,F,G):=¯{(fτ,gτ,Fτ,Gτ):τR}, the hull of (f,g,F,G), also possesses the same properties with the same constants.

    Theorem 2.31. Consider Eq (2.7). Suppose that AC(R,L(H)); Eq (2.4) is uniformly asymptotically stable such that (2.5) holds; fC(R×H,H), gC(R×H,L(U,H)), F,GC(R×H,L2(ν;H)). Suppose that W and N are the same as in Section 2.3; f,g,F,G satisfy the conditions (E1) and (E2). Then the following statements hold:

    (i) If L<ω2K1+2ω+2b, then Eq (2.7) has a unique solution ξC(R,B[0,r]) defined by equality

    ξ(t)=tTA(t,τ)f(τ,ξ(τ))dτ+tTA(t,τ)g(τ,ξ(τ))dW(τ)+t|x|U<1TA(t,τ)F(τ,ξ(τ),x)˜N(dτ,dx)+t|x|U1TA(t,τ)G(τ,ξ(τ),x)N(dτ,dx),

    where B[0,r]:={YL2(P;H):YL2(P;H)r} and r=2KM1+2ω+2bω2KL1+2ω+2b.

    (ii) If L<{ω2K2+8ω+4bω2K1+10ω+2b} and f,g,F,G satisfy (E1),(E2) and (E3) additionally, then we have Mu(A,f,g,F,G)˜Muξ and the solution ξ is strongly compatible in distribution (i.e. M(A,f,g,F,G)˜Mξ).

    Proof. The proof is similar to Theorem 4.6 in [21].

    Corollary 2.32. Consider Eq (2.7). Assume that the conditions of Theorem 2.31 hold.

    (i) If A,f,g,F,G are jointly stationary (respectively, τ-periodic, quasi-periodic with the spectrum of frequencies ν1,ν2,,νm, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable, Levitan almost periodic, almost recurrent, Poisson stable) in tR uniformly w.r.t. YH on every bounded subset, then so is the unique bounded solution ξ of Eq (2.7) in distribution.

    (ii) If A,f,g,F,G are jointly pseudo-periodic (respectively, pseudo-recurrent) and jointly Lagrange stable in tR uniformly w.r.t. YH on every bounded subset, then the unique bounded solution ξ of Eq (2.7) is pseudo-periodic (respectively, pseudo-recurrent) in distribution.

    Proof. This statement follows from Theorems 2.16, 2.31 and Remark 2.18.

    Let ε0 be some fixed positive number. We now study an averaging principle for the following equation

    dY(t)=ε(A(t)Y(t)+f(t))dt+εg(t)dW(t)+ε|x|U<1F(t,x)˜N(dt,dx)+ε|x|U1G(t,x)N(dt,dx), (3.1)

    where AC(R,L(H)), fC(R,L2(P;H)), gC(R,L2(P;L(U,H))), F,GC(R,L2(Pν;H)), and ε(0,ε0] is a small parameter. Here W and N are the Lévy-Itô decomposition components of the two-sided Lévy process L as in Section 2.3.

    Denote by Ψ the family of all decreasing, positive bounded functions ψ:R+R+ with limt+ψ(t)=0. Let AL(H). Denote by σ(A) the spectrum of A. We respectively impose the following conditions on A,f,g,F,G:

    (A1) AC(R,L(H)) and there exists ˉAL(H) such that

    limT+1Tt+TtA(s)ds=ˉA

    uniformly w.r.t. tR;

    (A2) fC(R,L2(P;H)) and there exist functions ω1Ψ, ˉfL2(P;H) such that

    1Tt+Tt[f(s)ˉf]dsL2(P;H)ω1(T) (3.2)

    for any T>0 and tR;

    (A3) gC(R,L2(P;L(U,H))) and there exist functions ω2Ψ, ˉgL2(P;L(U,H)) such that

    1Tt+TtE(g(s)ˉg)Q122L2(U,H)dsω2(T) (3.3)

    for any T>0 and tR;

    (A4) FC(R,L2(Pν;H)) and there exist functions ω3Ψ, ˉFL2(Pν;H) such that

    1Tt+Tt|x|U<1E|F(s,x)ˉF(x)|2ν(dx)dsω3(T) (3.4)

    for any T>0 and tR;

    (A5) GC(R,L2(Pν;H)) and there exist functions ω4Ψ, ˉGL2(Pν;H) such that

    1Tt+Tt|x|U1E|G(s,x)ˉG(x)|2ν(dx)dsω4(T) (3.5)

    for any T>0 and tR.

    Theorem 3.1. ([36,Chapter IV]) Suppose that ACb(R,L(B)) and

    limT+1Tt+TtA(s)ds=ˉA

    uniformly w.r.t. tR and the operator ˉA is Hurwitz, i.e. Re λ<0 for any λσ(ˉA).

    Then the following statements hold:

    (i) There exists a positive constant αε0 such that the equation

    dy(t)=Aε(t)y(t)dt,

    where Aε(t):=A(tε) for any tR, is uniformly asymptotically stable for any ε(0,α].

    (ii) There exists γ0>0 such that

    limε0sup(tτ;t,τR)eγ0(tτ)TAε(t,τ)TˉA(t,τ)B=0. (3.6)

    Remark 3.2. (i) Under the conditions of Theorem 3.1 there are positive constants α, K and ω such that

    TAε(t,τ)B,TˉA(t,τ)BKeω(tτ) (3.7)

    for any tτ and ε(0,α].

    (ii) By Eq (3.6) there exists a function K:(0,α)R+ such that K(ε)0 as ε0 and

    TAε(t,τ)TˉA(t,τ)BK(ε)eγ0(tτ)

    for any tτ (t,τR).

    Lemma 3.3. ([18]) Let fεC(R,B) for ε(0,α] be functions satisfying the following conditions:

    (i) there exists a positive constant A such that |fε(t)|BA for any tR and ε(0,α];

    (ii) for any l>0

    limε0sup|s|l, tR|t+stfε(σ)dσ|B=0.

    Then for any ω>0 we have

    limε0sup tR|teω(tτ)fε(τ)dτ|B=0.

    Remark 3.4. ([18]) If the function fCb(R,B) and ˉfB are such that

    limL+1L|t+Lt[f(s)ˉf]ds|=0

    uniformly w.r.t. tR, then the function fε(σ):=f(σε)ˉf satisfies the conditions of Lemma 3.3. Similarly, if the function g (respectively, F, G) in (A3) (respectively, (A4), (A5)) is L2-bounded, then the function fε(σ):=E(g(σε)ˉg)Q122L2(U,H) (respectively, fε(σ):=|x|U<1E|F(σε,x)ˉF(x)|2ν(dx), fε(σ):=|x|U1E|G(σε,x)ˉG(x)|2ν(dx)) satisfies the conditions of Lemma 3.3.

    Theorem 3.5. Consider Eq (3.1). Suppose that ACb(R,L(H)), fCb(R,L2(P;H)), gCb(R,L2(P;L(U,H))), F,GCb(R,L2(Pν;H)) and conditions (A1)-(A5) are satisfied. Suppose further that ˉA in (A1) is Hurwitz such that (3.6) and (3.7) hold. Then for any ε(0,α], Eq (3.1) has a unique solution φεCb(R,L2(P;H)) and it is strongly compatible in distribution (i.e. M(A,f,g,F,G)˜Mφε) and Mu(A,f,g,F,G)˜Muφε. Besides we have

    limε0suptRβ(L(φε(tε)),L(ˉϕ(t)))=0,

    where L(X) denotes the law of random variable X and ˉϕ is the unique stationary solution of

    dY(t)=(ˉAY(t)+ˉf)dt+ˉgdW(t)+|x|U<1ˉF(x)˜N(dt,dx)+|x|U1ˉG(x)N(dt,dx). (3.8)

    Proof. Consider the following equations

    dYε(t)=(Aε(t)Yε(t)+fε(t))dt+gε(t)dWε(t)+|x|U<1Fε(t,x)˜Nε(dt,dx)+|x|U1Gε(t,x)Nε(dt,dx) (3.9)

    and

    dYε(t)=(Aε(t)Yε(t)+fε(t))dt+gε(t)dW(t)+|x|U<1Fε(t,x)˜N(dt,dx)+|x|U1Gε(t,x)N(dt,dx), (3.10)

    where Aε(t):=A(tε), fε(t):=f(tε), gε(t):=g(tε), Fε(t,x):=F(tε,x) and Gε(t,x):=G(tε,x) for tR and ε(0,ε0]. Here Wε(t):=εW(tε), Nε(t,x):=εN(tε,x) and ˜Nε(t,x):=ε˜N(tε,x). By Theorem 2.29, Eq (3.9) has a unique bounded solution ψεCb(R,L2(P;H)) defined by equality

    ψε(t)=tTAε(t,τ)fε(τ)dτ+tTAε(t,τ)gε(τ)dWε(τ)+t|x|U<1TAε(t,τ)Fε(τ,x)˜Nε(dτ,dx)+t|x|U1TAε(t,τ)Gε(τ,x)Nε(dτ,dx)

    and the solution is strongly compatible in distribution (i.e. M(Aε,fε,gε,Fε,Gε)˜Mψε) and Mu(Aε,fε,gε,Fε,Gε)˜Muψε. Let φε(t):=ψε(εt), tR. Then φεCb(R,L2(P;H)) is the unique bounded solution of Eq (3.1). If {tn}M(A,f,g,F,G) (respectively, {tn}Mu(A,f,g,F,G)), then by Theorem 2.29 {εtn}M(Aε,fε,gε,Fε,Gε)˜Mψε (respectively, {εtn}Mu(Aε,fε,gε,Fε,Gε)˜Muψε). Further we have {tn}˜Mφε (respectively, {tn}˜Muφε).

    We now prove the solution φε of the original Eq (3.1) converges to the stationary solution ˉϕ of the averaged Eq (3.8) uniformly in tR in distribution sense. By Theorem 2.29, Eq (3.10) has a unique bounded solution ϕεCb(R,L2(P;H)) defined by equality

    ϕε(t)=tTAε(t,τ)fε(τ)dτ+tTAε(t,τ)gε(τ)dW(τ)+t|x|U<1TAε(t,τ)Fε(τ,x)˜N(dτ,dx)+t|x|U1TAε(t,τ)Gε(τ,x)N(dτ,dx). (3.11)

    By Theorem 2.29, Eq (3.8) has a unique bounded and stationary solution ˉϕ, which is given by the formula

    ˉϕ(t)=tTˉA(t,τ)ˉfdτ+tTˉA(t,τ)ˉgdW(τ)+t|x|U<1TˉA(t,τ)ˉF(x)˜N(dτ,dx)+t|x|U1TˉA(t,τ)ˉG(x)N(dτ,dx), (3.12)

    where TˉA(t,τ)=exp{ˉA(tτ)} for t,τR. From (3.11) and (3.12), by the basic inequality (ni=1ai)2n(ni=1a2i),nN, we have

    E|ϕε(t)ˉϕ(t)|2=E|tTAε(t,τ)fε(τ)dτ+tTAε(t,τ)gε(τ)dW(τ)+t|x|U<1TAε(t,τ)Fε(τ,x)˜N(dτ,dx)+t|x|U1TAε(t,τ)Gε(τ,x)N(dτ,dx)tTˉA(t,τ)ˉfdτtTˉA(t,τ)ˉgdW(τ)t|x|U<1TˉA(t,τ)ˉF(x)˜N(dτ,dx)t|x|U1TˉA(t,τ)ˉG(x)N(dτ,dx)|24E|t[TAε(t,τ)fε(τ)TˉA(t,τ)ˉf]dτ|2+4E|t[TAε(t,τ)gε(τ)TˉA(t,τ)ˉg]dW(τ)|2+4E|t|x|U<1[TAε(t,τ)Fε(τ,x)TˉA(t,τ)ˉF(x)]˜N(dτ,dx)|2+4E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]N(dτ,dx)|2=:I1(t,ε)+I2(t,ε)+I3(t,ε)+I4(t,ε). (3.13)

    Note that

    I1(t,ε)=4E|t[TAε(t,τ)fε(τ)TˉA(t,τ)ˉf]dτ|28E|tTAε(t,τ)(fε(τ)ˉf)dτ|2+8E|t[TAε(t,τ)ˉfTˉA(t,τ)ˉf]dτ|2=:8I11(t,ε)+8I12(t,ε). (3.14)

    Since

    TA(t,τ)τ=TA(t,τ)A(τ),

    we have

    TAε(t,t+s)sKAeωs

    for any tR and s<0.

    Similar to the proof of [18,Theorem 3.9], by making the change of variable s:=τt, integrating by parts and letting l be an arbitrary positive number we have

    tTAε(t,τ)(fε(τ)ˉf)dτL2(P;H)=0TAε(t,t+s)(fε(t+s)ˉf)dsL2(P;H)=0TAε(t,t+s)dds(t+st[fε(σ)ˉf]dσ)dsL2(P;H)=0TAε(t,t+s)s(t+st[fε(σ)ˉf]dσ)dsL2(P;H)=lTAε(t,t+s)s(t+st[fε(σ)ˉf]dσ)dsL2(P;H)+0lTAε(t,t+s)s(t+st[fε(σ)ˉf]dσ)dsL2(P;H)KA(2feωl(lω+1ω2)+1ω(1eωl)sup|s|l,tRt+st[fε(σ)ˉf]dσL2(P;H)). (3.15)

    By Remark 3.4, passing to limit in (3.15) as ε0 we obtain

    lim supε0suptR0TAε(t,t+s)s(t+st[fε(σ)ˉf]dσ)dsL2(P;H)2KAfeωl(lω+1ω2).

    Since l is an arbitrary positive number, by letting l we get

    limε0suptRI11(t,ε)=0.

    Note that by Cauchy-Schwarz inequality, we have

    I12(t,ε):=E|t[TAε(t,τ)TˉA(t,τ)]ˉfdτ|2(ˉfL2(P;H)K(ε)γ0)20

    as ε0. Consequently,

    limε0suptRI1(t,ε)=0.

    As for I2(t,ε), by Itô's isometry property and Remark 3.2-(ii), we have

    I2(t,ε)=4E|t[TAε(t,τ)gε(τ)TˉA(t,τ)ˉg]dW(τ)|2=4Et(TAε(t,τ)gε(τ)TˉA(t,τ)ˉg)Q122L2(U,H)dτ8EtTAε(t,τ)(gε(τ)ˉg)Q122L2(U,H)dτ+8Et(TAε(t,τ)TˉA(t,τ))ˉgQ122L2(U,H)dτ8K2te2ω(tτ)E(gε(τ)ˉg)Q122L2(U,H)dτ+4(K(ε))2γ0ˉgQ122L2(P;L2(U,H)). (3.16)

    Since by Lemma 3.3 the integral

    te2ω(tτ)E(gε(τ)ˉg)Q122L2(U,H)dτ

    goes to 0 as ε0 uniformly w.r.t. tR, taking to the limit in (3.16) we obtain

    limε0suptRI2(t,ε)=0.

    Note that

    I3(t,ε)=4E|t|x|U<1[TAε(t,τ)Fε(τ,x)TˉA(t,τ)ˉF(x)]˜N(dτ,dx)|2=4Et|x|U<1|TAε(t,τ)Fε(τ,x)TˉA(t,τ)ˉF(x)|2ν(dx)dτ8Et|x|U<1|TAε(t,τ)(Fε(τ,x)ˉF(x))|2ν(dx)dτ+8Et|x|U<1|(TAε(t,τ)TˉA(t,τ))ˉF(x)|2ν(dx)dτ8K2t|x|U<1e2ω(tτ)E|Fε(τ,x)ˉF(x)|2ν(dx)dτ+4(K(ε))2γ0ˉF2L2(Pν;H). (3.17)

    According to Lemma 3.3 we have

    limε0suptRt|x|U<1e2ω(tτ)E|Fε(τ,x)ˉF(x)|2ν(dx)dτ=0. (3.18)

    Passing to the limit in (3.17) and considering (3.18), we get

    limε0suptRI3(t,ε)=0.

    By properties of integrals for Poisson random measures, we obtain

    I4(t,ε)=4E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]N(dτ,dx)|2=4E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]˜N(dτ,dx)+t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]ν(dx)dτ|28E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]˜N(dτ,dx)|2+8E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]ν(dx)dτ|2=:I41(t,ε)+I42(t,ε). (3.19)

    Using the similar arguments as (3.17), we get

    I41(t,ε)16K2t|x|U1e2ω(tτ)E|Gε(τ,x)ˉG(x)|2ν(dx)dτ+8(K(ε))2γ0ˉG2L2(Pν;H). (3.20)

    By Cauchy-Schwarz inequality we have

    I42(t,ε)=8E|t|x|U1[TAε(t,τ)Gε(τ,x)TˉA(t,τ)ˉG(x)]ν(dx)dτ|216E|t|x|U1TAε(t,τ)(Gε(τ,x)ˉG(x))ν(dx)dτ|2+16E|t|x|U1(TAε(t,τ)TˉA(t,τ))ˉG(x)ν(dx)dτ|216t|x|U1K2eω(tτ)ν(dx)dτt|x|U1eω(tτ)E|Gε(τ,x)ˉG(x)|2ν(dx)dτ+16t|x|U1(K(ε))2eγ0(tτ)ν(dx)dτt|x|U1eγ0(tτ)E|ˉG(x)|2ν(dx)dτ16K2bωt|x|U1eω(tτ)E|Gε(τ,x)ˉG(x)|2ν(dx)dτ+16(K(ε))2bγ20ˉG2L2(Pν;H). (3.21)

    According to (3.19)-(3.21), we obtain

    I4(t,ε)16K2(1+bω)t|x|U1eω(tτ)E|Gε(τ,x)ˉG(x)|2ν(dx)dτ+8(K(ε))2(1γ0+2bγ20)ˉG2L2(Pν;H). (3.22)

    By Lemma 3.3 we have

    limε0suptRt|x|U1eω(tτ)E|Gε(τ,x)ˉG(x)|2ν(dx)dτ=0,

    so taking limits in (3.22) we obtain

    limε0suptRI4(t,ε)=0.

    Consequently we have

    limε0suptRE|ϕε(t)ˉϕ(t)|2=0. (3.23)

    Since the L2 convergence implies convergence in probability, it follows from (3.23) that

    limε0suptRβ(L(ϕε(t)),L(ˉϕ(t)))=0.

    On the other hand taking into consideration that L(W)=L(Wε) and L(N)=L(Nε) with compensated Poisson random measure ˜N we have L(ψε(t))=L(ϕε(t)) for any tR, and by the relationship between φε and ψε (i.e. φε(t):=ψε(εt), tR) we get

    limε0suptRβ(L(φε(tε)),L(ˉϕ(t)))=0.

    The proof is complete.

    Corollary 3.6. Under the assumptions of Theorem 3.5, it follows from Theorems 2.16 and 3.5 that

    (i) if A,f,g,F,G are jointly stationary (respectively, τ-periodic, quasi-periodic with the spectrum of frequencies ν1,,νk, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable, Levitan almost periodic, almost recurrent, Poisson stable), then Eq (3.1) has a unique solution φεCb(R,L2(P;H)) which is stationary (respectively, τ-periodic, quasi-periodic with the spectrum of frequencies ν1,,νk, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable, Levitan almost periodic, almost recurrent, Poisson stable) in distribution;

    (ii) if A,f,g,F,G are Lagrange stable and jointly pseudo-periodic (respectively, pseudo-recurrent), then Eq (3.1) has a unique solution φεCb(R,L2(P;H)) which is pseudo-periodic (respectively, pseudo-recurrent) in distribution;

    (iii)

    limε0suptRβ(L(φε(tε),L(ˉϕ(t)))=0.

    Let ε0 be some fixed positive number. Consider the stochastic differential equation driven by Lévy noise of type

    dY(t)=ε(A(t)Y(t)+f(t,Y(t)))dt+εg(t,Y(t))dW(t)+ε|x|U<1F(t,Y(t),x)˜N(dt,dx)+ε|x|U1G(t,Y(t),x)N(dt,dx), (4.1)

    where AC(R,L(H)), fC(R×H,H), gC(R×H,L(U,H)), F,GC(R×H,L2(ν;H)); ε(0,ε0] is a small parameter; W as well as N is the Lévy-Itô decomposition components of the two-sided Lévy process L with assumptions stated in Section 2.3. Assume that conditions (E1)-(E2) in Section 2.4 are satisfied. We also impose the following additional conditions on A,f,g,F,G:

    (H1) there exists ˉAL(H) such that

    limT+1Tt+TtA(s)ds=ˉA

    uniformly w.r.t. tR;

    (H2) there exist functions ω1Ψ and ˉfC(H,H) such that

    1T|t+Tt[f(s,y)ˉf(y)]ds|ω1(T)(1+|y|)

    for any T>0, yH and tR;

    (H3) there exist functions ω2Ψ and ˉgC(H,L(U,H)) such that

    1Tt+Tt(g(s,y)ˉg(y))Q122L2(U,H)dsω2(T)(1+|y|2)

    for any T>0, yH and tR;

    (H4) there exist functions ω3Ψ and ˉFC(H,L2(ν;H)) such that

    1Tt+Tt|x|U<1|F(s,y,x)ˉF(y,x)|2ν(dx)dsω3(T)(1+|y|2)

    for any T>0, yH and tR;

    (H5) there exist functions ω4Ψ and ˉGC(H,L2(ν;H)) such that

    1Tt+Tt|x|U1|G(s,y,x)ˉG(y,x)|2ν(dx)dsω4(T)(1+|y|2)

    for any T>0, yH and tR.

    Remark 4.1. Under the conditions (E1), (E2) and (H2)-(H5) the functions ˉf,ˉg,ˉF,ˉG also possess the properties (E1), (E2) with the same constants M and L.

    Lemma 4.2. Suppose that fC(R×H,H), gC(R×H,L(U,H)), F,GC(R×H,L2(ν;H)) and the conditions (E1) and (E2) hold. If φ is an L2-bounded solution of the equation

    dY(t)=f(t,Y(t))dt+g(t,Y(t))dW(t)+|x|U<1F(t,Y(t),x)˜N(dt,dx)+|x|U1G(t,Y(t),x)N(dt,dx),

    then there exists a constant C>0 depending only on M,L,φ,b, such that

    E|φ(t+h)φ(t)|2Ch2+Ch

    and

    Esuptst+h|φ(s)|2Ch2+Ch+C

    for any tR and h>0.

    Proof. Note that

    φ(t+h)=φ(t)+t+htf(τ,φ(τ))dτ+t+htg(τ,φ(τ))dW(τ)+t+ht|x|U<1F(τ,φ(τ),x)˜N(dτ,dx)+t+ht|x|U1G(τ,φ(τ),x)N(dτ,dx).

    Since f satisfies (E1) and (E2), we have for any τR

    E|f(τ,φ(τ))|22E|f(τ,φ(τ))f(τ,0)|2+2E|f(τ,0)|22L2φ2+2M2.

    Using the same arguments as above we have the same estimates for functions g,F,G. By Cauchy-Schwartz inequality, Itô's isometry property and properties of integrals for Poisson random measures we have

    E|φ(t+h)φ(t)|24E|t+htf(τ,φ(τ))dτ|2+4E|t+htg(τ,φ(τ))dW(τ)|2+4E|t+ht|x|U<1F(τ,φ(τ),x)˜N(dτ,dx)|2+4E|t+ht|x|U1G(τ,φ(τ),x)N(dτ,dx)|24ht+htE|f(τ,φ(τ))|2dτ+4t+htEg(τ,φ(τ))Q122L2(U,H)dτ+4t+ht|x|U<1E|F(τ,φ(τ),x)|2ν(dx)dτ+8E|t+ht|x|U1G(τ,φ(τ),x)˜N(dτ,dx)|2+8E|t+ht|x|U1G(τ,φ(τ),x)ν(dx)dτ|28ht+ht(M2+L2φ2)dτ+8t+ht(M2+L2φ2)dτ+8t+ht(M2+L2φ2)dτ+16t+ht(M2+L2φ2)dτ+8t+ht|x|U1E|G(τ,φ(τ),x)|2ν(dx)dτt+ht|x|U11ν(dx)dτ8ht+ht(M2+L2φ2)dτ+32t+ht(M2+L2φ2)dτ+16bht+ht(M2+L2φ2)dτCh2+Ch.

    Note that the BDG inequality for stochastic integrals with ˜N is very different from the BDG inequality with Brownian stochastic integrals. Here we need the following Kunita's first inequality ([31,Theorem 2.11])

    Esups[t0,t]|st0ZF(τ,x)˜N(dτ,dx)|pdp{E(tt0Z|F(τ,x)|2m(dx)dτ)p2+Ett0Z|F(τ,x)|pm(dx)dτ}, (4.2)

    where (Z,Z,m) is a measurable space, F:Ω×[t0,t]×ZH is a predictable process, p2, and dp is continuous w.r.t. p.

    Employing the BDG inequality (see [32,Theorem 4.36]), Kunita's first inequality, Cauchy-Schwartz inequality, Itô's isometry property and properties of integrals for Poisson random measures, we have

    Esuptst+h|φ(s)|25E|φ(t)|2+5Esuptst+h|stf(τ,φ(τ))dτ|2+5Esuptst+h|stg(τ,φ(τ))dW(τ)|2+5Esuptst+h|st|x|U<1F(τ,φ(τ),x)˜N(dτ,dx)|2+10Esuptst+h|st|x|U1G(τ,φ(τ),x)˜N(dτ,dx)|2+10Esuptst+h|st|x|U1G(τ,φ(τ),x)ν(dx)dτ|25φ2+5Esuptst+h|st(M+L|φ(τ)|)dτ|2+5CEt+htg(τ,φ(τ))Q122L2(U,H)dτ+5CEt+ht|x|U<1|F(τ,φ(τ),x)|2ν(dx)dτ+10CEt+ht|x|U1|G(τ,φ(τ),x)|2ν(dx)dτ+10Esuptst+hst|x|U1|G(τ,φ(τ),x)|2ν(dx)dτt+ht|x|U11ν(dx)dτ5φ2+5ht+ht2(M2+L2φ2)dτ+5Ct+ht2(M2+L2φ2)dτ+5Ct+ht2(M2+L2φ2)dτ+10Ct+ht2(M2+L2φ2)dτ+10bht+ht2(M2+L2φ2)dτCh2+Ch+C,

    with C denoting some positive constants and each C maybe different.

    Theorem 4.3. Consider Eq (4.1). Suppose that the operator A is bounded; the functions A,f,g,F,G satisfy the conditions (E1), (E2), (H1)-(H5), and the operator ˉA in (H1) is Hurwitz, i.e. Re λ<0 for any λσ(ˉA). Then there exists a positive constant ε1α such that for any ε(0,ε1]

    (i) if L<ω2K1+2ω+2b, Eq (4.1) has a unique solution φεCb(R,L2(P;H)) and φεr, where

    r:=2KM1+2ω+2bω2KL1+2ω+2b;

    (ii) if L<{ω2K2+8ω+4bω2K1+10ω+2b} and f,g,F,G satisfy (E1), (E2) and (E3) additionally, then we have Mu(A,f,g,F,G)˜Muφε and the solution φε is strongly compatible in distribution (i.e. M(A,f,g,F,G)˜Mφε);

    (iii) if L<ω2K3(1+2ω+2b), we have

    limε0suptRβ(L(φε(tε)),L(ˉϕ(t)))=0,

    where ˉϕ is the unique stationary solution of the averaged equation

    dY(t)=(ˉAY(t)+ˉf(Y(t)))dt+ˉg(Y(t))dW(t)+|x|U<1ˉF(Y(t),x)˜N(dt,dx)+|x|U1ˉG(Y(t),x)N(dt,dx).

    Proof. The first and second statements follow from Theorem 2.31.

    We now consider the following equations

    dY(t)=(Aε(t)Y(t)+fε(t,Y(t)))dt+gε(t,Y(t))dW(t)+|x|U<1Fε(t,Y(t),x)˜N(dt,dx)+|x|U1Gε(t,Y(t),x)N(dt,dx), (4.3)

    and

    dY(t)=(Aε(t)Y(t)+fε(t,Y(t)))dt+gε(t,Y(t))dWε(t)+|x|U<1Fε(t,Y(t),x)˜Nε(dt,dx)+|x|U1Gε(t,Y(t),x)Nε(dt,dx),

    where Aε(t):=A(tε), fε(t,y):=f(tε,y), gε(t,y):=g(tε,y), Fε(t,y,x):=F(tε,y,x) and Gε(t,y,x):=G(tε,y,x) for tR, yH and ε(0,ε0]. Here as before Wε(t):=εW(tε), ˜Nε(t,x):=ε˜N(tε,x) and Nε(t,x):=εN(tε,x). Since (fε,gε,Fε,Gε) satisfy conditions (E1) and (E2) with the same constants as (f,g,F,G), according to Theorem 2.31 Eq (4.3) has a unique solution ϕε from Cb(R,L2(P;H)) with ϕεC(R,B[0,r]), where

    r:=2KM1+2ω+2bω2KL1+2ω+2b.

    Note that

    E|ϕε(t)ˉϕ(t)|2=E|tTAε(t,τ)fε(τ,ϕε(τ))dτ+tTAε(t,τ)gε(τ,ϕε(τ))dW(τ)+t|x|U<1TAε(t,τ)Fε(τ,ϕε(τ),x)˜N(dτ,dx)+t|x|U1TAε(t,τ)Gε(τ,ϕε(τ),x)N(dτ,dx)tTˉA(t,τ)ˉf(ˉϕ(τ))dτtTˉA(t,τ)ˉg(ˉϕ(τ))dW(τ)t|x|U<1TˉA(t,τ)ˉF(ˉϕ(τ),x)˜N(dτ,dx)t|x|U1TˉA(t,τ)ˉG(ˉϕ(τ),x)N(dτ,dx)|24E|t(TAε(t,τ)fε(τ,ϕε(τ))TˉA(t,τ)ˉf(ˉϕ(τ)))dτ|2+4E|t(TAε(t,τ)gε(τ,ϕε(τ))TˉA(t,τ)ˉg(ˉϕ(τ)))dW(τ)|2+4E|t|x|U<1(TAε(t,τ)Fε(τ,ϕε(τ),x)TˉA(t,τ)ˉF(ˉϕ(τ),x))˜N(dτ,dx)|2+4E|t|x|U1(TAε(t,τ)Gε(τ,ϕε(τ),x)TˉA(t,τ)ˉG(ˉϕ(τ),x))N(dτ,dx)|2=:4(I1(t,ε)+I2(t,ε)+I3(t,ε)+I4(t,ε)). (4.4)

    Similar to the proof of [18,Theorem 4.3] with minor modifications, we have

    I1(t,ε)+I2(t,ε)3K2L2(12ω+1ω2)suptRE|ϕε(t)ˉϕ(t)|2+A(ε)+B(ε), (4.5)

    for any tR and ε(0,α). Here A,B:(0,α)R+ are functions such that A(ε),B(ε)0 as ε0.

    Next we estimate I3(t,ε) and I4(t,ε). By properties of integrals for Poisson random measures, we have

    I3(t,ε):=E|t|x|U<1(TAε(t,τ)Fε(τ,ϕε(τ),x)TˉA(t,τ)ˉF(ˉϕ(τ),x))˜N(dτ,dx)|23E|t|x|U<1TAε(t,τ)(Fε(τ,ϕε(τ),x)Fε(τ,ˉϕ(τ),x))˜N(dτ,dx)|2+3E|t|x|U<1(TAε(t,τ)TˉA(t,τ))Fε(τ,ˉϕ(τ),x)˜N(dτ,dx)|2+3E|t|x|U<1TˉA(t,τ)(Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x))˜N(dτ,dx)|23K2L2te2ω(tτ)E|ϕε(τ)ˉϕ(τ)|2dτ+6(K(ε))2te2γ0(tτ)(M2+L2ˉϕ2)dτ+3K2t|x|U<1e2ω(tτ)E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ3K2L22ωsuptRE|ϕε(t)ˉϕ(t)|2+3(K(ε))2γ0(M2+L2ˉϕ2)+3K2t|x|U<1e2ω(tτ)E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ. (4.6)

    To prove

    limε0suptRt|x|U<1e2ω(tτ)E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ=0,

    by Lemma 3.3 we only need to illustrate

    limε0sup|s|l,tR|t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ|=0.

    Divide [0,l] into intervals of size δ, where δ>0 is a fixed constant depending only on ε. Denote an adapted process ˆϕ such that ˆϕ(σ)=ˉϕ(t+kδ) for any σ[t+kδ,t+(k+1)δ). We may assume s>0 without loss of generality. Then by Lemma 4.2 we have

    t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ=t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)Fε(τ,ˆϕ(τ),x)+Fε(τ,ˆϕ(τ),x)ˉF(ˆϕ(τ),x)+ˉF(ˆϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ3t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)Fε(τ,ˆϕ(τ),x)|2ν(dx)dτ+3t+st|x|U<1E|Fε(τ,ˆϕ(τ),x)ˉF(ˆϕ(τ),x)|2ν(dx)dτ+3t+st|x|U<1E|ˉF(ˆϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ6L2l(Cδ2+Cδ)+3t+st|x|U<1E|Fε(τ,ˆϕ(τ),x)ˉF(ˆϕ(τ),x)|2ν(dx)dτ=:6L2l(Cδ2+Cδ)+3J1.

    Denote s(δ):=[|s|δ]. For J1, we have

    J1:=Et+st|x|U<1|Fε(τ,ˆϕ(τ),x)ˉF(ˆϕ(τ),x)|2ν(dx)dτEs(δ)1k=0t+(k+1)δt+kδ|x|U<1|Fε(τ,ˉϕ((t+kδ)),x)ˉF(ˉϕ((t+kδ)),x)|2ν(dx)dτ+Et+st+s(δ)δ|x|U<1|Fε(τ,ˉϕ((t+s(δ)δ)),x)ˉF(ˉϕ((t+s(δ)δ)),x)|2ν(dx)dτ=:J11+J21.

    Then

    J11:=Es(δ)1k=0t+(k+1)δt+kδ|x|U<1|Fε(τ,ˉϕ((t+kδ)),x)ˉF(ˉϕ((t+kδ)),x)|2ν(dx)dτ[lδ]max0ks(δ)1Et+(k+1)δt+kδ|x|U<1|Fε(τ,ˉϕ((t+kδ)),x)ˉF(ˉϕ((t+kδ)),x)|2ν(dx)dτ=[lδ]max0ks(δ)1Et+(k+1)δεt+kδε|x|U<1|F(τ,ˉϕ((t+kδ)),x)ˉF(ˉϕ((t+kδ)),x)|2εν(dx)dτlω3(δε)(1+ˉϕ2)

    and

    J21:=Et+st+s(δ)δ|x|U<1|Fε(τ,ˉϕ((t+s(δ)δ)),x)ˉF(ˉϕ((t+s(δ)δ)),x)|2ν(dx)dτ8(M2+L2ˉϕ2)δ.

    Hence

    sup|s|l,tR|t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ|6L2l(Cδ2+Cδ)+3lω3(δε)(1+ˉϕ2)+24(M2+L2ˉϕ2)δ. (4.7)

    Taking δ=ε and letting ε0 in (4.7), we have

    limε0sup|s|l,tR|t+st|x|U<1E|Fε(τ,ˉϕ(τ),x)ˉF(ˉϕ(τ),x)|2ν(dx)dτ|=0. (4.8)

    From (4.6) and (4.8) it follows that

    I3(t,ε)3K2L22ωsuptRE|ϕε(t)ˉϕ(t)|2+C(ε), (4.9)

    where C:(0,α)R+ is a function so that C(ε)0 as ε0.

    Note that

    I4(t,ε):=E|t|x|U1(TAε(t,τ)Gε(τ,ϕε(τ),x)TˉA(t,τ)ˉG(ˉϕ(τ),x))N(dτ,dx)|22E|t|x|U1(TAε(t,τ)Gε(τ,ϕε(τ),x)TˉA(t,τ)ˉG(ˉϕ(τ),x))˜N(dτ,dx)|2+2E|t|x|U1(TAε(t,τ)Gε(τ,ϕε(τ),x)TˉA(t,τ)ˉG(ˉϕ(τ),x))ν(dx)dτ|2=:2(I41(t,ε)+I42(t,ε)). (4.10)

    The similar arguments as I3(t,ε) yield that

    I41(t,ε)3K2L22ωsuptRE|ϕε(t)ˉϕ(t)|2+D1(ε), (4.11)

    where D1(ε) is some constant and D1(ε)0 as ε0. For I42(t,ε), by Cauchy-Schwartz inequality we have

    I42(t,ε):=E|t|x|U1(TAε(t,τ)Gε(τ,ϕε(τ),x)TˉA(t,τ)ˉG(ˉϕ(τ),x))ν(dx)dτ|23E|t|x|U1(TAε(t,τ)(Gε(τ,ϕε(τ),x)Gε(τ,ˉϕ(τ),x))ν(dx)dτ|2+3E|t|x|U1(TAε(t,τ)TˉA(t,τ))Gε(τ,ˉϕ(τ),x)ν(dx)dτ|2+3E|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|23K2L2bω2suptRE|ϕε(t)ˉϕ(t)|2+6(K(ε))2bγ20(M2+L2ˉϕ2)+3E|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|2. (4.12)

    We now show that

    limε0suptRE|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|2=0.

    To this end, making the change of variable s=τt and integrating by parts, we obtain for any l0

    E|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|2=E|0|x|U1TˉA(t,t+s)[Gε(t+s,ˉϕ((t+s),x)ˉG(ˉϕ((t+s),x)]ν(dx)ds|2=E|0TˉA(t,t+s)dds(t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ)ds|22E|0TˉA(t,t+s)s(t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ)ds|24E|lTˉA(t,t+s)s(t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ)ds|2+4E|0lTˉA(t,t+s)s(t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ)ds|24E(l|t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|KˉAeωsds)2+4Esupls0|t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|2|0lKˉAeωsds|2=:J2+J3. (4.13)

    For J2, by using Cauchy-Schwartz inequality we have

    J2:=4E(l|t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|KˉAeωsds)24K2ˉA2leωsdslE|t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|2eωsds4K2ˉA2ωeωll(t+st|x|U1E|Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)|2ν(dx)dσ)(t+st|x|U11ν(dx)dσ)eωsds4K2ˉA2ωeωll(t+st8(M2+L2ˉϕ2)dσ)bseωsds32K2ˉA2bω(M2+L2ˉϕ2)(l2ω+2lω2+2ω3)e2ωl. (4.14)

    Denote an adapted process ˜ϕ such that ˜ϕ(σ)=ˉϕ(tkδ) for any σ(t(k+1)δ,tkδ]. By Lemma 4.2 and Cauchy-Schwartz inequality, we obtain

    Esupls0|t+st|x|U1[Gε(σ,ˉϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|2=Esupls0|t+st|x|U1[Gε(σ,ˉϕ(σ),x)Gε(σ,˜ϕ(σ),x)+Gε(σ,˜ϕ(σ),x)ˉG(˜ϕ(σ),x)+ˉG(˜ϕ(σ),x)ˉG(ˉϕ(σ),x)]ν(dx)dσ|26Esupls0|t+stL|ˉϕ(σ)˜ϕ(σ)|dσ|2+3Esupls0|t+st|x|U1[Gε(σ,˜ϕ(σ),x)ˉG(˜ϕ(σ),x)]ν(dx)dσ|26L2l2(Cδ2+Cδ)+3Esupls0|t+st|x|U1[Gε(σ,˜ϕ(σ),x)ˉG(˜ϕ(σ),x)]ν(dx)dσ|2=:6L2l2(Cδ2+Cδ)+i1. (4.15)

    For i1, recalling that s(δ):=[|s|δ] by Lemma 4.2 we have

    i1:=3Esupls0|t+st|x|U1[Gε(τ,˜ϕ(τ),x)ˉG(˜ϕ(τ),x)]ν(dx)dτ|2=3Esupls0|s(δ)1k=0t(k+1)δtkδ|x|U1[Gε(τ,ˉϕ((tkδ)),x)ˉG(ˉϕ((tkδ)),x)]ν(dx)dτ+t+sts(δ)δ|x|U1[Gε(τ,ˉϕ((ts(δ)δ)),x)ˉG(ˉϕ((ts(δ)δ)),x)]ν(dx)dτ|26[lδ]Esupls0s(δ)1k=0|t(k+1)δtkδ|x|U1[Gε(τ,ˉϕ((tkδ)),x)ˉG(ˉϕ((tkδ)),x)]ν(dx)dτ|2+6Esupls0|t+sts(δ)δ|x|U1[Gε(τ,ˉϕ((ts(δ)δ)),x)ˉG(ˉϕ((ts(δ)δ)),x)]ν(dx)dτ|26l2δ2Esupls0max0ks(δ)1|t(k+1)δεtkδε|x|U1[G(τ,ˉϕ((tkδ)),x)ˉG(ˉϕ((tkδ)),x)]εν(dx)dτ|2+6Esupls0bδts(δ)δt+s|x|U1|Gε(τ,ˉϕ((ts(δ)δ)),x)ˉG(ˉϕ((ts(δ)δ)),x)|2ν(dx)dτ6l2δ2Esupls0max0ks(δ)1t(k+1)δεtkδε|x|U1|G(τ,ˉϕ((tkδ)),x)ˉG(ˉϕ((tkδ)),x)|2εν(dx)dτt(k+1)δεtkδε|x|U1εν(dx)dτ+6bδEsupls0ts(δ)δt+s8(M2+L2|ˉϕ(ts(δ)δ)|2)dτ6l2bEsupls0max0ks(δ)1ω4(δε)(1+|ˉϕ(t+kδ)|2)+6bδttl8(M2+L2Esuptlτt|ˉϕ(τ)|2)dτ6l2b(Cl2+Cl+C+1)ω4(δε)+48lbδ(M2+L2(Cl2+Cl+C)) (4.16)

    Therefore, by (4.15) and (4.16) we get

    J3[6L2l2(Cδ2+Cδ)+6l2b(Cl2+Cl+C+1)ω4(δε)+48lbδ(M2+L2(Cl2+Cl+C))]4K2ˉA2ω2(1eωl)2. (4.17)

    Hence, according to (4.13), (4.14) and (4.17) we have

    E|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|232K2ˉA2bω(M2+L2ˉϕ2)(l2ω+2lω2+2ω3)e2ωl+4K2ˉA2ω2(1eωl)2[6L2l2(Cδ2+Cδ)+6l2b(Cl2+Cl+C+1)ω4(δε)+48lbδ(M2+L2(Cl2+Cl+C))]. (4.18)

    Taking δ=ε and letting ε0 in (4.11), we have

    limε0suptRE|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|232K2ˉA2bω(M2+L2ˉϕ2)(l2ω+2lω2+2ω3)e2ωl.

    Since l is arbitrary, by letting l0 we have

    limε0suptRE|t|x|U1TˉA(t,τ)[Gε(τ,ˉϕ(τ),x)ˉG(ˉϕ(τ),x)]ν(dx)dτ|2=0. (4.19)

    From (4.12) and (4.19) it follows that

    I42(t,ε)3K2L2bω2suptRE|ϕε(t)ˉϕ(t)|2+D2(ε), (4.20)

    where D2(ε) is some positive constant such that D2(ε)0 as ε0. So by (4.10), (4.11) and (4.20) we have

    I4(t,ε)(3K2L2ω+6K2L2bω2)suptRE|ϕε(t)ˉϕ(t)|2+D1(ε)+D2(ε). (4.21)

    Combing (4.4), (4.5), (4.9) and (4.21), we have

    (112K2L2(1ω2+2ω+2bω2))suptRE|ϕε(t)ˉϕ(t)|24(A(ε)+B(ε)+C(ε)+D1(ε)+D2(ε)).

    By the assumption L<ω2K3(1+2ω+2b) the coefficient is positive, so

    limε0suptRE|ϕε(t)ˉϕ(t)|2=0.

    Since L2-convergence implies convergence in distribution,

    limε0suptRβ(L(ϕε(t)),L(ˉϕ(t)))=0.

    We note that L(φε(tε))=L(ϕε(t)). Then we have

    limε0suptRβ(L(φε(tε)),L(ˉϕ(t)))=0.

    The proof is complete.

    Corollary 4.4. Assume that the conditions of Theorem 4.3 hold.

    (i) If A,f,g,F,G are jointly stationary (respectively, τ-periodic, quasi-periodic with the spectrum of frequencies ν1,,νk, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable, Levitan almost periodic, almost recurrent, Poisson stable), then Eq (4.1) has a unique solution φεCb(R,L2(P;H)) which is stationary (respectively, τ-periodic, quasi-periodic with the spectrum of frequencies ν1,,νk, almost periodic, almost automorphic, Birkhoff recurrent, Lagrange stable, Levitan almost periodic, almost recurrent, Poisson stable) in distribution.

    (ii) If A,f,g,F,G are Lagrange stable and jointly pseudo-periodic (respectively, pseudo-recurrent), then Eq (4.1) has a unique solution φεCb(R,L2(P;H)) which is pseudo-periodic (respectively, pseudo-recurrent) in distribution.

    (iii)

    limε0suptRβ(L(φε(tε),L(ˉϕ(t)))=0.

    Proof. This statement follows from Theorems 2.16, 4.3 and Remark 2.18.

    In this section, we illustrate our theoretical results by two examples.

    Example 5.1. Consider the following stochastic ordinary differential equation driven by a two-sided Lévy noise:

    dy=ε(3y+14ysin2t)dt+14εdW+ε|x|<115y˜N(dt,dx)+ε|x|118y(sint+cos3t)N(dt,dx), (5.1)

    where ε is a small positive parameter; W is a one-dimensional two-sided Brownian motion and N is a Poisson random measure in R, which is independent of W. Let

    ˉf(y)=limT+1TT014ysin2tdt=18y,
    ˉG(y,x)=limT+1TT018y(sint+cos3t)dt=0,

    and define the corresponding averaged stochastic differential equation

    dˉy=(3ˉy+18ˉy)dt+14dW+|x|<115ˉy˜N(dt,dx). (5.2)

    It is clear that A generates a dissipative semigroup on R with K=1, ω=3. Since the functions A,f,g,F,G are respectively stationary, periodic, stationary, stationary and quasi-periodic in t, uniformly with respect to y on any bounded subset of R, it follows that functions A,f,g,F,G are jointly quasi-periodic. Conditions (E1) and (E1) always hold for any constant M, and the Lipschitz constants of functions f,g,F,G in Conditions (E2) and (E2) can be chosen as 14 if

    |x|<1(15)pν(dx)(14)pand|x|1(14)pν(dx)(14)p

    for p=2 and some constant p>2, i.e.

    ν(1,1)<2516andb1.

    As for Condition (E3), it naturally holds for stochastic ordinary differential equation. Conditions (H1)-(H5) are also satisfied for functions A,f,g,F,G.

    Since f,g,F,G satisfy (E1) and (E2), by Theorem 4.2-(i) Eq (5.1) has a unique L2-bounded solution provided ν(1,1)<2516, b1. The restrictions in Theorem 4.2-(ii) and (iii) respectively become

    14<{322+24+4b321+30+2b}and14<323(1+6+2b),

    i.e. b<52. According to Corollary 4.3, the unique L2-bounded solution of Eq (5.1) is quasi-periodic in distribution and it uniformly converges to the unique stationary solution of the averaged Eq (5.2) on R in distribution sense.

    Example 5.2. Consider the following equations:

    dyi=ε[aiyi+fi(t,y)]dt+εgi(t,y)dW+ε|x|<1Fi(t,y,x)˜N(dt,dx)+ε|x|1Gi(t,y,x)N(dt,dx),(i=1,2,...,n,...) (5.3)

    where 0<ωaiκ for i=1,2,..., and κ and ω are constants; fi,giC(R×l2,l2), Fi,GiC(R×l2,L2(ν;l2)), l2:={y:y=(y1,y2,...,yn,...),i=1y2i<}; W is a one-dimensional two-sided Brownian motion and N is a Poisson random measure independent of W. (5.3) can also be written in the following form:

    dY=ε[AY+f(t,Y)]dt+εg(t,Y)dW+ε|x|<1F(t,Y,x)˜N(dt,dx)+ε|x|1G(t,Y,x)N(dt,dx), (5.4)

    where Y=(y1yn), A=(a1an), f(t,Y)=(f1(t,Y)fn(t,Y)), g(t,Y)=(g1(t,Y)gn(t,Y)),

    F(t,Y,x)=(F1(t,Y,x)Fn(t,Y,x)), G(t,Y,x)=(G1(t,Y,x)Gn(t,Y,x)).

    According to (H1)-(H5), we define the corresponding averaged equations:

    dˉyi=[aiˉyi+ˉfi(ˉy)]dt+ˉgi(ˉy)dW+|x|<1ˉFi(ˉy,x)˜N(dt,dx)+|x|1ˉGi(ˉy,x)N(dt,dx),(i=1,2,...). (5.5)

    Eq (5.5) can also be written in the following form:

    dˉY=[AˉY+ˉf(ˉY)]dt+ˉg(ˉY)dW+|x|<1ˉF(ˉY,x)˜N(dt,dx)+|x|1ˉG(ˉY,x)N(dt,dx), (5.6)

    where ˉY=(ˉy1ˉyn), A=(a1an), ˉf(ˉY)=(ˉf1(ˉY)ˉfn(ˉY)), ˉg(ˉY)=(ˉg1(ˉY)ˉgn(ˉY)),

    ˉF(ˉY,x)=(ˉF1(ˉY,x)ˉFn(ˉY,x)), ˉG(ˉY,x)=(ˉG1(ˉY,x)ˉGn(ˉY,x)).

    Assume that the Lipschitz condition, growth condition and all the conditions of Theorem 4.3 are satisfied for fi,gi,Fi,Gi, (i = 1, 2, ...). Thus Theorem 4.3 and Corollary 4.4 hold. If every fi,gi,Fi,Gi, (i = 1, 2, ...) are periodic or quasi-periodic in t uniformly with respect to y on any bounded subset of l2, then fi,gi,Fi,Gi are jointly almost periodic. Basis on Theorem 4.3 and Corollary 4.4, Eq (5.4) has a unique almost periodic solution and it converges to the stationary solution of the averaged Eq (5.6) in distribution sense on the whole real axis.

    Now we give a concrete example with infinite dimension:

    dyi=ε[4yi+14(yi1+yi+yi+1)cos2it]dt+ε[15(yi1+yi+yi+1)sin2it]dW+ε|x|<1[15yi1sinit+15yisin3it+15yi+1cos5it]˜N(dt,dx),(i=1,2,...) (5.7)

    and the averaged equations

    dˉyi=4ˉyidt+[110(ˉyi1+ˉyi+ˉyi+1)]dW,(i=1,2,...). (5.8)

    It is immediate to check that the drift, diffusion, and jump coefficients of Eqs (5.7) and (5.8) satisfy all the conditions of Theorem 4.3 and Corollary 4.4. So the almost periodic solution of Eq (5.7) converges to the stationary solution of Eq (5.8) on R in distribution.

    In this paper we study an averaging principle on infinite time intervals for semilinear stochastic ordinary differential equations with Lévy noise. We establish the existence of Poisson stable (including periodic, quasi-periodic, almost periodic, almost automorphic etc) solutions of the equations by a unified framework. We show that the Poisson stable solution of the original equation converges to the stationary solution of the averaged equation uniformly on the whole real axis in distribution sense, as the time scale goes to zero. Moreover, we illustrate our theoretical results with some examples.

    In this paper we only consider stochastic ordinary differential equations case. The case of the averaging on R for stochastic partial differential equations is more complicated, but it is both interesting and important. It deserves a separate paper.

    This work is partially supported by NSFC Grant 11871132, and Xinghai Jieqing and DUT19TD14 funds from Dalian University of Technology. We sincerely thank Professor Zhenxin Liu for helpful discussions. We are grateful to the anonymous referees for their careful reading of our paper and valuable suggestions, which lead to significant improvement of the paper.

    All authors declare no conflicts of interest in this paper.



    [1] N. N. Bogolyubov, N. M. Krylov, La theorie generale de la mesure dans son application a l'etude de systemes dynamiques de la mecanique non-lineaire, Ann. Math., 38 (1937), 65-113.
    [2] N. M. Krylov, N. N. Bogolyubov, Introduction to Non-Linear Mechanics, Princeton, N. J.: Princeton University Press, 1943.
    [3] N. N. Bogolyubov, On Some Statistical Methods in Mathematical Physics, Kiev: Akademiya Nauk Ukrainskoi SSR, 1945.
    [4] N. N. Bogolyubov, Y. A. Mitropolsky, Asymptotic Methods in the Theory of Non-Linear Oscillations, New York: Gordon and Breach Science Publishers, 1961.
    [5] R. L. Stratonovich, Topics in the Theory of Random Noise, New York: Gordon and Breach, 1963.
    [6] S. Cerrai, M. Freidlin, Averaging principle for a class of stochastic reaction-diffusion equations, Probab. Theory Relat. Fields, 144 (2009), 137-177. doi: 10.1007/s00440-008-0144-z
    [7] S. Cerrai, A. Lunardi, Averaging principle for nonautonomous slow-fast systems of stochastic reaction-diffusion equations: The almost periodic case, SIAM J. Math. Anal., 49 (2017), 2843-2884. doi: 10.1137/16M1063307
    [8] J. Duan, W. Wang, Effective Dynamics of Stochastic Partial Differential Equations, Amsterdam: Elsevier, 2014.
    [9] M. Freidlin, A. Wentzell, Random Perturbations of Dynamical Systems, Heidelberg: Springer, 2012.
    [10] R. Z. Khasminskii, On the principle of averaging the Itô's stochastic differential equations, Kybernetika, 4 (1968), 260-279.
    [11] A. V. Skorokhod, Asymptotic Methods in the Theory of Stochastic Differential Equations, Providence, R. I.: American Mathematical Society, 1989.
    [12] A. Yu. Veretennikov, On large deviations in the averaging principle for SDEs with a "full dependence", Ann. Probab., 27 (1999), 284-296. doi: 10.1214/aop/1022677263
    [13] I. Vrkoc, Weak averaging of stochastic evolution equations, Math. Bohem., 120 (1995), 91-111. doi: 10.21136/MB.1995.125891
    [14] W. Wang, A. J. Roberts, Average and deviation for slow-fast stochastic partial differential equations, J. Differ. Equations, 253 (2012), 1265-1286. doi: 10.1016/j.jde.2012.05.011
    [15] Y. Xu, J. Duan, W. Xu, An averaging principle for stochastic dynamical systems with Lévy noise, Phys. D, 240 (2011), 1395-1401. doi: 10.1016/j.physd.2011.06.001
    [16] W. Mao, S. You, X. Wu, X. Mao, On the averaging principle for stochastic delay differential equations with jumps, Adv. Differ. Equations, 70 (2015), 1-19.
    [17] V. Burd, Method of Averaging for Differential Equations on an Infinite Interval: Theory and applications, New York: Chapman and Hall, 2007.
    [18] D. Cheban, Z. Liu, Averaging principle on infinite intervals for stochastic ordinary differential equations, Electron. Res. Arch., 2021. Available from: http://www.aimsciences.org/article/doi/10.3934/era.2021014.
    [19] D. Cheban, Z. Liu, Periodic, quasi-periodic, almost periodic, almost automorphic, Birkhoff recurrent and Poisson stable solutions for stochastic differential equations, J. Differ. Equations, 269 (2020), 3652-3685.
    [20] M. Cheng, Z. Liu, Periodic, almost periodic and almost automorphic solutions for SPDEs with monotone coefficients, 2021. Available from: https://arXiv.org/abs/1911.02169.
    [21] X. Liu, Z. Liu, Poisson stable solutions for stochastic differential equations with Lévy noise, 2021. Available from: https://arXiv.org/abs/2002.00395.
    [22] G. R. Sell, Topological Dynamics and Ordinary Differential Equations, London: Van Nostrand Reinhold Co., 1971.
    [23] B. A. Shcherbakov, Topologic Dynamics and Poisson Stability of Solutions of Differential Equations, Chişinǎu: Ştiinţa, 1972.
    [24] B. A. Shcherbakov, Poisson Stability of Motions of Dynamical Systems and Solutions of Differential Equations, Chişinǎu: Ştiinţa, 1985.
    [25] K. S. Sibirsky, Introduction to Topological Dynamics, Leiden: Noordhoff International Publishing, 1975.
    [26] D. Cheban, Global Attractors of Nonautonomous Dynamical and Control Systems, 2Eds., Hackensack, NJ: World Scientific Publishing Co. Pte. Ltd., 2015.
    [27] B. M. Levitan, V. V. Zhikov, Almost Periodic Functions and Differential Equations, Moscow: Moscow State University Press, 1978.
    [28] B. A. Shcherbakov, A certain class of Poisson stable solutions of differential equations, Differencial'nye Uravnenija, 4 (1968), 238-243.
    [29] B. A. Shcherbakov, The comparability of the motions of dynamical systems with regard to the nature of their recurrence, Differentcial'nye Uravnenija, 11 (1975), 1246-1255.
    [30] D. Applebaum, Lévy Process and Stochastic Calculus, 2Eds., Cambridge: Cambridge University Press, 2009.
    [31] H. Kunita, Stochastic differential equations based on Lévy processes and stochastic flows of diffeomorphisms, In: M. M. Rao, Real and Stochastic Analysis, Boston: Birkhäuser, (2004), 305-373.
    [32] G. Da Prato, J. Zabczyk, Stochastic Equations in Infinite Dimensions, 2Eds., Cambridge: Cambridge University Press, 2014.
    [33] Ju. L. Daleckii, M. G. Krein, Stability of Solutions of Differential Equations in Banach Space, Providence, R. I.: American Mathematical Society, 1974.
    [34] D. Cheban, Asymptotically Almost Periodic Solutions of Differential Equations, New York: Hindawi Publishing Corporation, 2009.
    [35] R. M. Dudley, Real Analysis and Probability, 2Eds., Cambridge: Cambridge University Press, 2002.
    [36] M. A. Krasnoselskii, V. Burd, Yu. S. Kolesov, Nonlinear Almost Periodic Oscillations, Moscow: Nauka, 1970.
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2676) PDF downloads(164) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog