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

Strong convergence of multiscale truncated Euler-Maruyama method for super-linear slow-fast stochastic differential equations

Yuanping Cui School of Mathematics and Statistics, Northeast Normal University, Changchun, Jilin, 130024, China.    Xiaoyue Li Corresponding author (Email: lixy209nenu.edu.cn). School of Mathematics and Statistics, Northeast Normal University, Changchun, Jilin, 130024, China. Research of this author was supported by the National Natural Science Foundation of China (11971096), the National Key R&\&D Program of China (2020YFA0714102), the Natural Science Foundation of Jilin Province (YDZJ202101ZYTS154), the Education Department of Jilin Province (JJKH20211272KJ), and the Fundamental Research Funds for the Central Universities.
Abstract

This work focuses on solving super-linear stochastic differential equations (SDEs) involving different time scales numerically. Taking advantages of being explicit and easily implementable, a multiscale truncated Euler-Maruyama scheme is proposed for slow-fast SDEs with local Lipschitz coefficients. By virtue of the averaging principle, the strong convergence of its numerical solutions to the exact ones in ppth moment is obtained. Furthermore, under mild conditions on the coefficients, the corresponding strong error estimate is also provided. Finally, two examples and some numerical simulations are given to verify the theoretical results.

Keywords. Slow-fast stochastic differential equations; Super-linearity; Explicit multiscale scheme; ppth moment; Strong convergence.

1 Introduction

Stochastic modelling plays an essential role in many branches of science and industry. Especially, super-linear stochastic differential equations (SDEs) are usually used to describe real-world systems in various applications, for examples, the stochastic Lotka-Volterra model in biology for the population growth [33], the elasticity of volatility model arising in finance for the asset price [26] and the stochastic Ginzburg-Landau equation stemming from statistical physics in the study of phase transitions [24]. In many fields, various factors change at different rates: some vary rapidly whereas others evolve slowly. As a result the separation of fast and slow time scales arises in chemistry, fluid dynamics, biology, physics, finance and other fields [5, 14, 17, 25]. Stochastic systems with this characteristic are studied extensively [10, 35, 36, 46] and are often modeled by the following slow-fast SDEs (SFSDEs),

{dxε(t)=b(xε(t),yε(t))dt+σ(xε(t))dW1(t),dyε(t)=1εf(xε(t),yε(t))dt+1εg(xε(t),yε(t))dW2(t),\begin{cases}\mathrm{d}x^{\varepsilon}(t)=b(x^{\varepsilon}(t),y^{\varepsilon}(t))\mathrm{d}t+\sigma(x^{\varepsilon}(t))\mathrm{d}W^{1}(t),\\ \mathrm{d}y^{\varepsilon}(t)=\displaystyle\frac{1}{\varepsilon}f(x^{\varepsilon}(t),y^{\varepsilon}(t))\mathrm{d}t+\frac{1}{\sqrt{\varepsilon}}g(x^{\varepsilon}(t),y^{\varepsilon}(t))\mathrm{d}W^{2}(t),\end{cases} (1.1)

with initial value (xε(0),yε(0))=(x0,y0)n1×n2(x^{\varepsilon}(0),y^{\varepsilon}(0))=(x_{0},y_{0})\in\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}. Here, coefficients

b:n1×n2n1,σ:n1n1×d1,\displaystyle b:\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}\rightarrow\mathbb{R}^{n_{1}},~{}~{}~{}~{}\sigma:\mathbb{R}^{n_{1}}\rightarrow\mathbb{R}^{n_{1}\times d_{1}},\
f:n1×n2n2,g:n1×n2n2×d2\displaystyle f:\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}\rightarrow\mathbb{R}^{n_{2}},~{}~{}~{}~{}g:\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}\rightarrow\mathbb{R}^{n_{2}\times d_{2}}\

are continuous, and {W1(t)}t0\{W^{1}(t)\}_{t\geq 0} and {W2(t)}t0\{W^{2}(t)\}_{t\geq 0} represent mutually independent d1d_{1}-dimensional and d2d_{2}-dimensional Brownian motions, respectively. The parameter ε>0\varepsilon>0 represents the ratio of nature time scales between xε(t)x^{\varepsilon}(t) and yε(t)y^{\varepsilon}(t). Especially, as ε1\varepsilon\ll 1, xε(t)x^{\varepsilon}(t) and yε(t)y^{\varepsilon}(t) are called the slow component and fast component, respectively.

In various applications the time evolution of the slow component xε(t)x^{\varepsilon}(t) is under the spotlight. Due to the existence of super-linear coefficients and multiple time scales as well as the coupling of fast and slow components, it is almost impossible to anticipate the dynamics of slow component directly by solving the full system. Therefore, numerical methods or approximation techniques become the efficient tools. However, on account of the wide separation in time scale, standard computational schemes may be no longer applicable. Hence, our main aim is to construct the appropriate multiscale numerical scheme to approximate the slow component.

On analytical grounds, the averaging principle is essential to describe the asymptotic behavior of the slow component as ε0\varepsilon\rightarrow 0. Precisely, assume that the frozen equation is described by

dyx(s)=f(x,yx(s))ds+g(x,yx(s))dW2(s)\displaystyle\mathrm{d}y^{x}(s)=f(x,y^{x}(s))\mathrm{d}s+g(x,y^{x}(s))\mathrm{d}W^{2}(s) (1.2)

with initial value yx(0)=y0y^{x}(0)=y_{0}, where xx is regarded as a parameter. Let yx,y0(s)y^{x,y_{0}}(s) denote the solution of the frozen equation (1.2) with initial value y0y_{0}. If the transition semigroup of yx,y0(t)y^{x,y_{0}}(t) has a unique invariant probability measure and the following integral

b¯(x)=n2b(x,y)μx(dy).\displaystyle\bar{b}(x)=\int_{\mathbb{R}^{n_{2}}}b(x,y)\mu^{x}(\mathrm{d}y). (1.3)

exists. The averaging principle reveals that under suitable assumptions on the coefficients, the slow component xε(t)x^{\varepsilon}(t) converges to x¯(t)\bar{x}(t), which is the solution of

{dx¯(t)=b¯(x¯(t))dt+σ(x¯(t))dW1(t),x¯(0)=x0.\begin{cases}\mathrm{d}\bar{x}(t)=\bar{b}(\bar{x}(t))\mathrm{d}t+\sigma(\bar{x}(t))\mathrm{d}W^{1}(t),\\ \bar{x}(0)=x_{0}.\end{cases} (1.4)

Much of the study of the averaging principle can be traced back to the original work of Khasminskii [23] for a class of diffusion processes, in which an averaging principle was showed in weak sense. Subsequently, fruitful results on averaging principle are developed. The convergence between the solution of the averaged system and the exact slow component in probability was yielded in [43, 44]. The strong convergence in mean square was obtained in [16], and was further developed under an ergodicity-type assumption in [15]. Furthermore, order 1/41/4 convergence rate was obtained in [13], order 1/21/2 strong convergence rate and order 11 in weak sense were yielded for the degenerate SFSDEs with the deterministic slow component in [8]. The order 1/21/2 strong convergence rate was also yielded for SFSDEs whose coefficients depend on small parameter ε\varepsilon in [30]. Recently, the averaging principle for SFSDEs with the super-linear growth coefficients was obtained in [32]. Furthermore, the strong averaging principles have been developed for various kinds of stochastic slow-fast systems, such as jump-diffusion processes [11, 47], stochastic partial differential equations [2, 9, 4], McKean-Vlasov SDEs [37], and so on.

The averaged equation derived from the averaging principle provides a substantial simplification of the original system. If the coefficient b¯()\bar{b}(\cdot) is explicitly available, the averaged equation (1.4) can be immediately used to simulate the evolution of xε(t)x^{\varepsilon}(t). However, it is often impossible or impractical to do this because the dynamics of the frozen equation (1.2) are too complicated to acquire the invariant measure μx\mu^{x} analytically. The heterogeneous multiscale method (HMM) [6, 7] was proposed to solve the averaged equation. Within the framework of the HMM, the basic idea is to solve the averaged equation (1.4) by a suitable macro solver, in which the averaged coefficient b¯\bar{b} is estimated on the fly by performing a series of constrained microscopic simulations.

Precisely, in 2003 Vanden-Eijnden [41] proposed numerical schemes for Multiscale dynamical systems with stochastic effects. To fully justify the scheme, in 2005, E et al. [8] provided a thorough analysis of convergence and efficiency of the scheme for degenerate SFSDEs of which the slow dynamics is deterministic. In 2006, Givon et al. [13] developed the Projective integration schemes for SFSDEs with the slow diffusion coefficient only depending on the slow component. And in 2008, Givon et al. [12] extended the projective integration schemes to jump-diffusion systems. Furthermore, in 2010, Liu [29] generalized the analysis of E et al. for the fully coupled SFSDEs, whose slow diffusion coefficient depends on both slow and fast components. Bréhier [2, 3] developed the HMM scheme for the slow-fast parabolic stochastic partial differential equations, which promotes the theory in [8] to the case of the infinite dimension.

Even though the averaged equation is not available explicitly, the selection of macro solver relies on its structure. Provided that the coefficients of averaged equation are global Lipschitz continuous, the Euler-Maruyama (EM) scheme is very popular used as the macro solver owing to the simple algebraic structure and the cheap computational cost [41, 13, 8, 29, 12]. However, super-linear growth coefficients are common phenomena in multiscale systems, which may derive a super-linear averaged equation by averaging principle. Hutzenthaler et al. [21] pointed out that the EM approximation errors for a large class of super-linear SDEs diverge to infinity in ppth moment for any p1p\geq 1. Hence, the classical EM scheme is no longer suitable as macro solver for SFSDEs with super-linear coefficients. Besides, chosen as the macro solver for the sup-linear averaged equation, implicit scheme tends to make the algorithm and implementation more involved [8]. As a consequence, there is an urgent need to construct an explicit multiscale numerical scheme for super-linear SFSDEs.

Fortunately, great achievements have been made in the research of explicit numerical methods for super-linear SDEs, for examples, the tamed EM scheme [20, 22, 38, 39], the tamed Milstein scheme [45], the stopped EM scheme [31], the truncated EM scheme [27, 28, 34] and therein. So far the ability of these modified EM methods to approximate the solutions of super-linear diffusion systems have been shown extensively. Inspired by the above works, using modified EM method we are devoted to constructing an explicit multiscale numerical method suitable for super-linear SFSDEs.

In fact, HMM relies heavily on the structure of the averaged equation for the slow variable. Thus we have to overcome two major obstacles: the unexplicit form and super-linear structure of b¯\bar{b} . Inspired by the idea from [34], we design a truncation device to modify the super-linear coefficient bb of original system in advance, so as to achieve the modification of b¯\bar{b}. This modification can avoid possible large excursion due to the super-linearity of b¯\bar{b}. Then fitting into the framework of HMM, we construct an explicit multiscale numerical scheme involving three subroutines as follows.

  • 1.

    The truncated EM (TEM) scheme is selected as the macro solver to evolute the macro dynamics x¯(t)xε(t)\bar{x}(t)\approx x^{\varepsilon}(t) in which the modified averaged coefficient is required to be estimated at each macro time step.

  • 2.

    An appropriate numerical scheme is chosen as the micro solver to solve the frozen equation to produce the data used for approximating the modified coefficient.

  • 3.

    An estimator is established to obtain the desired approximation of the modified averaged coefficient.

Following this line, we construct an easily implementable explicit multiscale numerical scheme for a class of super-linear SFSDEs and obtain its strong convergence.

The rest of this paper is organized as follows. Section 2 gives some notations, hypotheses and preliminaries. Section 3 proposes an explicit multiscale numerical method. Section 4 provides some important pre-estimates. Section 5 yields the strong convergence of MTEM scheme. Section 6 focuses on the error analysis of the explicit MTEM scheme and presents an important example. Section 7 shows two numerical examples and carries out some numerical experiments to verify our theoretical results. Section 8 concludes this paper.

2 Preliminary

Throughout this paper, we use the following notations. Let (Ω,,)(\Omega,\cal{F},\mathbb{P}) be a complete probability space with a natural filtration {t}t0\{\mathcal{F}_{t}\}_{t\geq 0} satisfying the usual conditions (i.e. it is right continuous and increasing while 0\mathcal{F}_{0} contains all \mathbb{P}-null sets), and 𝔼\mathbb{E} be the expectation corresponding to \mathbb{P}. Let |||\cdot| denote the Euclidean norm in n\mathbb{R}^{n} and the trace norm in n×d.\mathbb{R}^{n\times d}. If AA is a vector or matrix, we denote its transpose by ATA^{T}. For a set 𝔻\mathbb{D}, let I𝔻(x)=1I_{\mathbb{D}}(x)=1 if x𝔻x\in\mathbb{D} and 0 otherwise. We set inf=\inf\emptyset=\infty, where \emptyset is empty set. Moreover, for any a,ba,b\in\mathbb{R}, we define ab=max{a,b}a\vee b=\max\{a,b\} and ab=min{a,b}a\wedge b=\min\{a,b\}. We use CC and CpC_{p} to denote the generic positive constants, which may take different values at different appearances, where CpC_{p} is used to emphasize that the constant depends on the parameter pp. In addition, C,CpC,C_{p} are independent of constants Δ1\Delta_{1}, Δ2\Delta_{2}, nn and MM that occur in the next section. CRC_{R} usually denotes some positive function increasing with respect to RR.

Let 𝒫p(n2)\mathcal{P}_{p}(\mathbb{R}^{n_{2}}) be the set of all probability measures on n2\mathbb{R}^{n_{2}} denoted by 𝒫(n2)\mathcal{P}(\mathbb{R}^{n_{2}}), with finite pp-th moment, i.e.,

𝒫p(n2):={μ𝒫(n2):n2|y|pμ(dy)<},\displaystyle\mathcal{P}_{p}(\mathbb{R}^{n_{2}}):=\Big{\{}\mu\in\mathcal{P}(\mathbb{R}^{n_{2}}):\int_{\mathbb{R}^{n_{2}}}|y|^{p}\mu(\mathrm{d}y)<\infty\Big{\}},

which is a Polish space under the Wasserstein distance

𝕎p(μ1,μ2)=infπ𝒞(μ1,μ2)(n2×n2|y1y2|pπ(dy1,dy2))1p,\displaystyle\mathbb{W}_{p}(\mu_{1},\mu_{2})=\inf_{\pi\in\mathcal{C}(\mu_{1},\mu_{2})}\Big{(}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|^{p}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{)}^{\frac{1}{p}},

where 𝒞(μ1,μ2)\mathcal{C}(\mu_{1},\mu_{2}) stands for the set of all probability measures on n2×n2\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}} with marginals μ1\mu_{1} and μ2\mu_{2}, respectively.

To state the main results, we impose some hypotheses on the coefficients bb, σ\sigma of slow system and ff and gg of fast system.

  • (S1)

    There exists a constant θ11\theta_{1}\geq 1 such that for any R>0R>0, x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} with |x1||x2|R|x_{1}|\vee|x_{2}|\leq R,

    |b(x1,y)b(x2,y)|+|σ(x1)σ(x2)|LR|x1x2|(1+|y|θ1),\displaystyle|b(x_{1},y)-b(x_{2},y)|+|\sigma(x_{1})-\sigma(x_{2})|\leq L_{R}|x_{1}-x_{2}|(1+|y|^{\theta_{1}}),\

    here LRL_{R} is a positive constant dependent on RR.

  • (S2 )

    There exist constants θ21\theta_{2}\geq 1 and K1>0K_{1}>0 such that for any xn1x\in\mathbb{R}^{n_{1}} and y1,y2n2y_{1},y_{2}\in\mathbb{R}^{n_{2}}.

    |b(x,y1)b(x,y2)|K1|y1y2|(1+|x|θ2+|y1|θ2+|y2|θ2).\displaystyle|b(x,y_{1})-b(x,y_{2})|\leq K_{1}|y_{1}-y_{2}|\big{(}1+|x|^{\theta_{2}}+|y_{1}|^{\theta_{2}}+|y_{2}|^{\theta_{2}}\big{)}.
  • (S3)

    There exists a constant K2>0K_{2}>0 such that for any xn1x\in\mathbb{R}^{n_{1}},

    |σ(x)|K2(1+|x|).\displaystyle|\sigma(x)|\leq K_{2}(1+|x|).
  • (S4)

    There exist constants θ3,θ41\theta_{3},\theta_{4}\geq 1 and K3>0K_{3}>0 such that for any xn1,yn2x\in\mathbb{R}^{n_{1}},y\in\mathbb{R}^{n_{2}},

    |b(x,y)|K3(1+|x|θ3+|y|θ4).\displaystyle|b(x,y)|\leq K_{3}(1+|x|^{\theta_{3}}+|y|^{\theta_{4}}).
  • (S5)

    There exist constants K4>0K_{4}>0 and λ>0\lambda>0 such that for any xn1,yn2x\in\mathbb{R}^{n_{1}},y\in\mathbb{R}^{n_{2}},

    xTb(x,y)K4(1+|x|2)+λ|y|2.\displaystyle x^{T}b(x,y)\leq K_{4}(1+|x|^{2})+\lambda|y|^{2}.\
  • (F1)

    The functions ff and gg are globally Lipschitz continuous, namely, for any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} and y1,y2n2y_{1},y_{2}\in\mathbb{R}^{n_{2}}, there exists a positive constant LL such that

    |f(x1,y1)f(x2,y2)||g(x1,y1)g(x2,y2)|L(|x1x2|+|y1y2|).\displaystyle|f(x_{1},y_{1})-f(x_{2},y_{2})|\vee|g(x_{1},y_{1})-g(x_{2},y_{2})|\leq L(|x_{1}-x_{2}|+|y_{1}-y_{2}|).
  • (F2)

    There exists a constant β>0\beta>0 such that for any xn1x\in\mathbb{R}^{n_{1}} and y1,y2n2y_{1},y_{2}\in\mathbb{R}^{n_{2}},

    2(y1y2)T(f(x,y1)f(x,y2))+|g(x,y1)g(x,y2)|2β|y1y2|2.\displaystyle 2(y_{1}-y_{2})^{T}\big{(}f(x,y_{1})-f(x,y_{2})\big{)}+|g(x,y_{1})-g(x,y_{2})|^{2}\leq-\beta|y_{1}-y_{2}|^{2}.
  • (𝐅𝟑)({\bf F3})

    For some fixed k2k\geq 2, there exist constants αk>0\alpha_{k}>0 and Lk>0L_{k}>0 such that for any xn1x\in\mathbb{R}^{n_{1}}, yn2y\in\mathbb{R}^{n_{2}},

    yTf(x,y)+k12|g(x,y)|2αk|y|2+Lk(1+|x|2).\displaystyle y^{T}f(x,y)+\frac{k-1}{2}|g(x,y)|^{2}\leq-\alpha_{k}|y|^{2}+L_{k}(1+|x|^{2}).
Remark 2.1.

According to [32, Theorem 2.2], system (1.1) admits a unique global solution (xε(t),yε(t))(x^{\varepsilon}(t),y^{\varepsilon}(t)) under (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}). Obviously, (𝐅𝟏)({\bf{F1}}) guarantees that the frozen equation (1.2) has a unique global solution yx,y0(s)y^{x,y_{0}}(s), which is a time homogeneous Markov process.

Lemma 2.1.

If (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold, then for any fixed xn1x\in\mathbb{R}^{n_{1}}, the transition semigroup {sx}s0\{\mathbb{P}^{x}_{s}\}_{s\geq 0} has a unique invariant probability measure μx𝒫k(n2)\mu^{x}\in\mathcal{P}_{k}(\mathbb{R}^{n_{2}}), which satisfies that

n2|y|kμx(dy)C(1+|x|k).\displaystyle\int_{\mathbb{R}^{n_{2}}}|y|^{k}\mu^{x}(\mathrm{d}y)\leq C(1+|x|^{k}). (2.1)

Furthermore, for any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}},

𝕎2(μx1,μx2)C|x1x2|.\displaystyle\mathbb{W}_{2}(\mu^{x_{1}},\mu^{x_{2}})\leq C|x_{1}-x_{2}|. (2.2)
Proof.

For any fixed xn1x\in\mathbb{R}^{n_{1}} and yn2y\in\mathbb{R}^{n_{2}}, under (𝐅𝟑)({\bf F3}) it follows from [32, Lemma 3.6] that

sups0𝔼|yx,y(s)|kC(1+|x|k)<,\displaystyle\sup_{s\geq 0}\mathbb{E}|y^{x,y}(s)|^{k}\leq C(1+|x|^{k})<\infty,

which implies that δysx𝒫k(n2)𝒫2(n2)\delta_{y}\mathbb{P}^{x}_{s}\in\mathcal{P}_{k}(\mathbb{R}^{n_{2}})\subset\mathcal{P}_{2}(\mathbb{R}^{n_{2}}). It is well known the inequality that for any μ1,μ2𝒫2(n2)\mu_{1},\mu_{2}\in\mathcal{P}_{2}(\mathbb{R}^{n_{2}})

𝕎2(μ1sx,μ2sx)\displaystyle\mathbb{W}_{2}(\mu_{1}\mathbb{P}^{x}_{s},\mu_{2}\mathbb{P}^{x}_{s}) n2×n2𝕎2(δy1sx,δy2sx)π(dy1,dy2)\displaystyle\leq\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}\mathbb{W}_{2}(\delta_{y_{1}}\mathbb{P}^{x}_{s},\delta_{y_{2}}\mathbb{P}^{x}_{s})\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})
n2×n2(𝔼|yx,y1(s)yx,y2(s)|2)12π(dy1,dy2),\displaystyle\leq\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}\big{(}\mathbb{E}|y^{x,y_{1}}(s)-y^{x,y_{2}}(s)|^{2}\big{)}^{\frac{1}{2}}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2}),

here π𝒞(μ1,μ2)\pi\in\mathcal{C}(\mu_{1},\mu_{2}). Then under (F2), by virtue of [32, Lemma 3.7] we derive that

𝕎2(μ1sx,μ2sx)\displaystyle\mathbb{W}_{2}(\mu_{1}\mathbb{P}^{x}_{s},\mu_{2}\mathbb{P}^{x}_{s}) Ceβs2n2×n2|y1y2|π(dy1,dy2)\displaystyle\leq Ce^{-\frac{\beta s}{2}}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\
Ceβs2(n2×n2|y1y2|2π(dy1,dy2))12.\displaystyle\leq Ce^{-\frac{\beta s}{2}}\Big{(}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|^{2}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}}.\

Then due to the arbitrariness of πC(μ1,μ2)\pi\in C(\mu_{1},\mu_{2}), we have

𝕎2(μ1sx,μ2sx)Ceβs2𝕎2(μ1,μ2),\displaystyle\mathbb{W}_{2}(\mu_{1}\mathbb{P}^{x}_{s},\mu_{2}\mathbb{P}^{x}_{s})\leq Ce^{-\frac{\beta s}{2}}\mathbb{W}_{2}(\mu_{1},\mu_{2}),

which yields the uniqueness of invariant measure if it exists. Next we shall prove the existence of invariant probability measure. In fact, it is sufficient to prove that for any fixed xn1x\in\mathbb{R}^{n_{1}} and yn2y\in\mathbb{R}^{n_{2}}, {δysx}s0\{\delta_{y}\mathbb{P}^{x}_{s}\}_{s\geq 0} is a 𝕎2\mathbb{W}_{2}-cauchy sequence due to the completeness of 𝒫2(n2)\mathcal{P}_{2}(\mathbb{R}^{n_{2}}) space. Using the Kolmogorov-Chapman equation and [32, Lemma 3.7], one derives that for any s,t>0s,t>0,

𝕎2(δysx,δys+tx)\displaystyle\mathbb{W}_{2}(\delta_{y}\mathbb{P}^{x}_{s},\delta_{y}\mathbb{P}^{x}_{s+t}) =𝕎2(δysx,δytxsx)Ceβs2𝕎2(δy,δytx)\displaystyle=\mathbb{W}_{2}(\delta_{y}\mathbb{P}^{x}_{s},\delta_{y}\mathbb{P}^{x}_{t}\mathbb{P}^{x}_{s})\leq Ce^{-\frac{\beta s}{2}}\mathbb{W}_{2}(\delta_{y},\delta_{y}\mathbb{P}^{x}_{t})\
Ceβs2(|y|2+𝔼|yx,y(t)|2)12Ceβs2(1+|x|+|y|),\displaystyle\leq Ce^{-\frac{\beta s}{2}}\big{(}|y|^{2}+\mathbb{E}|y^{x,y}(t)|^{2}\big{)}^{\frac{1}{2}}\leq Ce^{-\frac{\beta s}{2}}(1+|x|+|y|),

which implies that as tt\rightarrow\infty, {δysx}s0\{\delta_{y}\mathbb{P}^{x}_{s}\}_{s\geq 0} is a 𝕎2\mathbb{W}_{2}-cauchy sequence whose limit is denoted by μx\mu^{x}. Furthermore, in view of the continuity of 𝕎2\mathbb{W}_{2}-distance(see, [42, Corollary 6.1]) we derive for any t>0t>0

𝕎2(μxtx,μx)=lims𝕎2(δys+tx,δysx)=0,\displaystyle\mathbb{W}_{2}(\mu^{x}\mathbb{P}^{x}_{t},\mu^{x})=\lim_{s\rightarrow\infty}\mathbb{W}_{2}(\delta_{y}\mathbb{P}^{x}_{s+t},\delta_{y}\mathbb{P}^{x}_{s})=0,

which implies that μx\mu^{x} is indeed an invariant probability measure of yx(s)y^{x}(s). On the other hand, it follows from [32, Proposition 3.8] that

n2|y|kμx(dy)C(1+|x|k).\displaystyle\int_{\mathbb{R}^{n_{2}}}|y|^{k}\mu^{x}(\mathrm{d}y)\leq C(1+|x|^{k}).

In addition, using the continuity of 𝕎2\mathbb{W}_{2} again yields that

𝕎22(μx1,μx2)\displaystyle\mathbb{W}^{2}_{2}(\mu^{x_{1}},\mu^{x_{2}}) =lims𝕎22(δysx1,δysx2)\displaystyle=\lim_{s\rightarrow\infty}\mathbb{W}^{2}_{2}(\delta_{y}\mathbb{P}^{x_{1}}_{s},\delta_{y}\mathbb{P}^{x_{2}}_{s})\
lims𝔼|yx1,y(s)yx2,y(s)|2C|x1x2|2,\displaystyle\leq\lim_{s\rightarrow\infty}\mathbb{E}|y^{x_{1},y}(s)-y^{x_{2},y}(s)|^{2}\leq C|x_{1}-x_{2}|^{2},

where the last step follows from the [32, Lemma 3.10]. The proof is complete. ∎

An averaged equation (1.4) derived from the averaging principle, which provides a substantial simplification of original system (1.1). Then the numerical scheme for system (1.1) can be proposed by solving the averaged equation (1.4) numerically. For this purpose, we cite some known results on the averaging principle firstly.

Lemma 2.2 ([32, Theorem 2.3]).

If (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k>4θ12(θ2+1)2θ32θ4k>4\theta_{1}\vee 2(\theta_{2}+1)\vee 2\theta_{3}\vee 2\theta_{4}, then for any 0<p<k0<p<k and T>0T>0,

limε0𝔼(supt[0,T]|xε(t)x¯(t)|p)=0,\displaystyle\lim\limits_{\varepsilon\rightarrow 0}\mathbb{E}\Big{(}\sup\limits_{t\in[0,T]}|x^{\varepsilon}(t)-\bar{x}(t)|^{p}\Big{)}=0,

where xε(t)x^{\varepsilon}(t) and x¯(t)\bar{x}(t) are the solutions of (1.1) and (1.4), respectively.

Lemma 2.3 ([32, Lemma 3.11]).

If (𝐒𝟏)({\bf S1})-(𝐒𝟑)({\bf S3}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with kθ12θ2θ4k\geq\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, then the averaged equation (1.4) has a unique global solution x¯(t)\bar{x}(t) satisfying

𝔼(sup0tT|x¯(t)|p)CT,p>0,T>0.\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq t\leq T}|\bar{x}(t)|^{p}\Big{)}\leq C_{T},~{}~{}~{}~{}\forall~{}p>0,~{}T>0.
Remark 2.2.

For any constant R>|x0|R>|x_{0}|, define the stopping time

τR=inf{t0:|x¯(t)|R}.\tau_{R}=\inf\{t\geq 0:|\bar{x}(t)|\geq R\}.

Then it follows from Lemma 2.3 that

Rp(τRT)𝔼|x¯(TτR)|p𝔼(sup0tT|x¯(t)|p)CT,\displaystyle R^{p}\mathbb{P}(\tau_{R}\leq T)\leq\mathbb{E}|\bar{x}(T\wedge\tau_{R})|^{p}\leq\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)|^{p}\Big{)}\leq C_{T},

which implies that

(τRT)CTRp.\mathbb{P}(\tau_{R}\leq T)\leq\frac{C_{T}}{R^{p}}.

3 The construction of explicit multiscale scheme

With the help of the strong averaging principle, this section is devoted to constructing an easily implementable multiscale numerical scheme for the slow component of original SFSDE (1.1). One notices from (𝐒𝟒)({\bf S4}) that for any yn2y\in\mathbb{R}^{n_{2}},

|b(x,y)|\displaystyle|b(x,y)| K3(1+|x|)(1+|x|θ31)+K3|y|θ4.\displaystyle\leq K_{3}(1+|x|)(1+|x|^{\theta_{3}-1})+K_{3}|y|^{\theta_{4}}.\

Then for any u1u\geq 1 and xn1x\in\mathbb{R}^{n_{1}} with |x|u|x|\leq u

|b(x,y)|K3sup|x|uφ(u)(1+|x|)+K3|y|θ4,\displaystyle|b(x,y)|\leq K_{3}\sup_{|x|\leq u}\varphi(u)(1+|x|)+K_{3}|y|^{\theta_{4}}, (3.1)

where φ(u)=1+u(θ3θ41)\varphi(u)=1+u^{(\theta_{3}\vee\theta_{4}-1)}, and θ3,θ4\theta_{3},\theta_{4} are given in (𝐒𝟒)({\bf S4}). Then for any step size Δ1(0,1]\Delta_{1}\in(0,1], define

x=(|x|φ1(KΔ112))x|x|,xn1,\displaystyle x^{*}=\Big{(}|x|\wedge\varphi^{-1}\big{(}K\Delta_{1}^{-\frac{1}{2}}\big{)}\Big{)}\frac{x}{|x|},~{}~{}~{}~{}~{}~{}x\in\mathbb{R}^{n_{1}},

where x/|x|=𝟎n1x/|x|={\bf 0}\in\mathbb{R}^{n_{1}} if x=𝟎x={\bf 0}, and KK is a constant satisfying K1+|x0|(θ3θ41)K\geq~{}1+|x_{0}|^{(\theta_{3}\vee\theta_{4}-1)}. Clearly, for any xn1x\in\mathbb{R}^{n_{1}},

|b(x,y)|CΔ112(1+|x|)+K3|y|θ4.\displaystyle|b(x^{*},y)|\leq C\Delta_{1}^{-\frac{1}{2}}(1+|x^{*}|)+K_{3}|y|^{\theta_{4}}. (3.2)

Moreover, under (S4), (F1)-(F3) with kθ4k\geq\theta_{4} by the definition (1.3) we derive from the above inequality and (2.1) that

|b¯(x)|\displaystyle|\bar{b}(x^{*})| =|n2b(x,y)μx(dy)|n2|b(x,y)|μx(dy)\displaystyle=\Big{|}\int_{\mathbb{R}^{n_{2}}}b(x^{*},y)\mu^{x^{*}}(\mathrm{d}y)\Big{|}\leq\int_{\mathbb{R}^{n_{2}}}|b(x^{*},y)|\mu^{x^{*}}(\mathrm{d}y)\
CΔ112(1+|x|)+K3n2|y|θ4μx(dy)\displaystyle\leq C\Delta_{1}^{-\frac{1}{2}}(1+|x^{*}|)+K_{3}\int_{\mathbb{R}^{n_{2}}}|y|^{\theta_{4}}\mu^{x^{*}}(\mathrm{d}y)\
CΔ112(1+|x|)+C(1+|x|θ41)(1+|x|)\displaystyle\leq C\Delta_{1}^{-\frac{1}{2}}(1+|x^{*}|)+C(1+|x^{*}|^{\theta_{4}-1})(1+|x^{*}|)\
CΔ112(1+|x|),xn1,\displaystyle\leq C\Delta_{1}^{-\frac{1}{2}}(1+|x^{*}|),~{}~{}~{}~{}x\in\mathbb{R}^{n_{1}}, (3.3)

where μx\mu^{x^{*}} is the unique invariant probability measure of the frozen equation (1.2) with the fixed parameter xx^{*}, and the last step used the increasing of φ\varphi.

Because the analytical form of b¯(x)\bar{b}(x^{*}) is unobtainable, using the ergodicity of the frozen equation (1.2), we approximate b¯(x)\bar{b}(x^{*}) by the time average of b(x,)b(x^{*},\cdot) with respect to the numerical solution of the frozen equation (1.2) with fixed parameter xx^{*}. For convenience, for an integer M>0M>0, we introduce an average function

BM(x,h)=1Mm=1Mb(x,hm),xn1,\displaystyle B_{M}(x,h)=\frac{1}{M}\sum_{m=1}^{M}b(x,h_{m}),~{}~{}~{}\forall x\in\mathbb{R}^{n_{1}},~{} (3.4)

where h={hm}m=1h=\{h_{m}\}_{m=1}^{\infty} is a n2\mathbb{R}^{n_{2}}-valued sequence. Within the framework of HMM, we design an easily implementable multiscale numerical scheme involving a macro solver and a micro solver as well as an estimator. For clarity, we illustrate it as follows. Let Δ1\Delta_{1} and Δ2\Delta_{2} denote macro time step size and micro time step size, respectively.

  • (1)

    Macro solver: For the known Xn\!X_{n}, since the drift coefficient b¯\bar{b} of the averaged equation may be sup-linear, the truncated EM scheme is selected as macro solver to make a macro step and get Xn+1X_{n+1}. Then we have

    Xn+1=Xn+BnΔ1+σ(Xn)ΔWn1,X_{n+1}=X_{n}+B_{n}\Delta_{1}+\sigma(X_{n})\Delta W^{1}_{n},

    where BnB_{n} ia an approximation of b¯(Xn)\bar{b}(X^{*}_{n}) that we obtain in third step, and n1(u):=u/Δ1n_{1}(u):=\lfloor u/\Delta_{1}\rfloor for any u0u\geq 0 with t/δ\lfloor t/\delta\rfloor the integer part of t/δt/\delta, and ΔWn1=W1((n+1)Δ1)W1(nΔ1)\Delta W^{1}_{n}=W^{1}((n+1)\Delta_{1})-W^{1}(n\Delta_{1}).

  • (2)

    Micro solver: To obtain BnB_{n} at each macro time step, for the known Xnn1X_{n}\in\mathbb{R}^{n_{1}}, use the EM method to solve the frozen equation (1.2) with parameter x=Xnx=X^{*}_{n} fixed. Therefore, the micro solver is given by

    {Y0Xn=y0,Ym+1Xn=YmXn+f(Xn,YmXn)Δ2+g(Xn,YmXn)ΔWn,m2,m=0,1,,\!\!\!\begin{cases}Y^{X^{*}_{n}}_{0}=y_{0},\\ Y^{X^{*}_{n}}_{m+1}=Y^{X^{*}_{n}}_{m}\!+f(X^{*}_{n},Y^{X^{*}_{n}}_{m})\Delta_{2}\!+g(X^{*}_{n},Y^{X^{*}_{n}}_{m})\Delta W^{2}_{n,m},~{}~{}m=0,1,\cdot\cdot\cdot,\end{cases}

    where {Wn2()}n0\{W^{2}_{n}(\cdot)\}_{n\geq 0} is a mutually independent Brownian motion sequence and also independent of W1(t)W^{1}(t), and ΔWn,m2=Wn2((m+1)Δ2)Wn2(mΔ2)\Delta W^{2}_{n,m}=W^{2}_{n}((m+1)\Delta_{2})-W^{2}_{n}(m\Delta_{2}).

  • (3)

    Estimator: For the known XnX_{n} and YXn:={YmXn}m1Y^{X^{*}_{n}}:=\{Y^{X_{n}^{*}}_{m}\}_{m\geq 1}, let

    Bn=BM(Xn,YXn)B_{n}=B_{M}(X^{*}_{n},Y^{X^{*}_{n}})

    as an approximation of b¯(Xn)\bar{b}(X^{*}_{n}), where BM(,)B_{M}(\cdot,\cdot) is defined by (3.4) and MM denotes the number of micro time steps used for this approximation.

Overall, for any given Δ1,Δ2(0,1]\Delta_{1},\Delta_{2}\in(0,1] and integer M1M\geq 1 define the multiscale TEM scheme (MTEM) as follows: for any n0n\geq 0,

X0=x0,Xn=(|Xn|φ1(KΔ112))Xn|Xn|,Y0Xn=y0,\displaystyle X_{0}=x_{0},~{}X^{*}_{n}=\Big{(}|X_{n}|\wedge\varphi^{-1}\big{(}K\Delta_{1}^{-\frac{1}{2}}\big{)}\Big{)}\frac{X_{n}}{|X_{n}|},~{}Y^{X_{n}^{*}}_{0}=y_{0}, (3.5a)
Ym+1Xn=YmXn+f(Xn,YmXn)Δ2+g(Xn,YmXn)ΔWn,m2,\displaystyle Y^{X^{*}_{n}}_{m+1}=Y^{X^{*}_{n}}_{m}+f(X^{*}_{n},Y^{X^{*}_{n}}_{m})\Delta_{2}+g(X^{*}_{n},Y^{X^{*}_{n}}_{m})\Delta W^{2}_{n,m}, (3.5b)
m=0,1,,M1,\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}m=0,1,\cdot\cdot\cdot,M-1,\
Xn+1=Xn+BM(Xn,YXn)Δ1+σ(Xn)ΔWn1.\displaystyle X_{n+1}=X_{n}+B_{M}(X^{*}_{n},Y^{X^{*}_{n}})\Delta_{1}+\sigma(X_{n})\Delta W^{1}_{n}. (3.5c)

By this scheme we define the continuous approximation processes

X(t)\displaystyle X(t) =Xn,t[nΔ1,(n+1)Δ1),\displaystyle=X_{n},~{}~{}~{}~{}~{}~{}~{}t\in[n\Delta_{1},(n+1)\Delta_{1}), (3.6)
X¯(t)\displaystyle\bar{X}(t) =x0+0tBM(X(s),YX(s))ds+0tσ(X(s))dW1(s).\displaystyle=x_{0}+\int_{0}^{t}B_{M}(X^{*}(s),Y^{X^{*}(s)})\mathrm{d}s+\int_{0}^{t}\sigma(X(s))\mathrm{d}W^{1}(s). (3.7)

Note that X¯(nΔ1)=X(nΔ1)=Xn\bar{X}(n\Delta_{1})={X}(n\Delta_{1})=X_{n}, that is, X¯(t)\bar{X}(t) and X(t)X(t) coincide with the discrete solution at the grid points, respectively.

4 Some important pre-estimates

In order to better study the strong convergence of the MTEM scheme, we need to study some important properties of the averaged coefficient b¯(x)\bar{b}(x) and its estimator BM(x,Ynx)B_{M}(x,Y^{x}_{n}) (defined in later) in advance. In this section, we mainly provide some pre-estimates for b¯(x)\bar{b}(x) and BM(x,Ynx)B_{M}(x,Y^{x}_{n}).

By virtue of Lemma 2.1, we show that the drift term b¯\bar{b} of the averaged equation (1.4) inherits the local Lipschitz continuity. ∎

Lemma 4.1.

Under (𝐒𝟏)({\bf S1}), (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with kθ12θ2θ4k\geq\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, for any R>0R>0 and x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} with |x1||x2|R|x_{1}|\vee|x_{2}|\leq R, there exists a constant L¯R\bar{L}_{R} such that

|b¯(x1)b¯(x2)|L¯R|x1x2|.\displaystyle|\bar{b}(x_{1})-\bar{b}(x_{2})|\leq\bar{L}_{R}|x_{1}-x_{2}|.
Proof.

For any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}}, according to (1.3) we have

|b¯(x1)b¯(x2)|\displaystyle|\bar{b}(x_{1})-\bar{b}(x_{2})| =|n1×n2(b(x1,y1)b(x2,y2))π(dy1,dy2)|\displaystyle=\Big{|}\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}(b(x_{1},y_{1})-b(x_{2},y_{2}))\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{|}\
n1×n2|b(x1,y1)b(x2,y2)|π(dy1,dy2)\displaystyle\leq\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}\big{|}b(x_{1},y_{1})-b(x_{2},y_{2})|\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\
n1×n2|b(x1,y1)b(x2,y1)|π(dy1,dy2)\displaystyle\leq\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}\big{|}b(x_{1},y_{1})-b(x_{2},y_{1})|\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\
+n1×n2|b(x2,y1)b(x2,y2)|π(dy1,dy2),\displaystyle~{}~{}~{}+\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}\big{|}b(x_{2},y_{1})-b(x_{2},y_{2})|\pi(\mathrm{d}y_{1},\mathrm{d}y_{2}),

here πC(μx1,μx2)\pi\in C(\mu^{x_{1}},\mu^{x_{2}}) is arbitrary. Then for any R>0R>0 and x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} with |x1||x2|R|x_{1}|\vee|x_{2}|\leq R, by the Hölder inequality it follows from (𝐒𝟏)({\bf S1}) and (𝐒𝟐)({\bf S2}) that

|b¯(x1)b¯(x2)|\displaystyle|\bar{b}(x_{1})-\bar{b}(x_{2})| LR|x1x2|n2(1+|y1|θ1)μx1(dy1)\displaystyle\leq L_{R}|x_{1}-x_{2}|\int_{\mathbb{R}^{n_{2}}}(1+|y_{1}|^{\theta_{1}})\mu^{x_{1}}(\mathrm{d}y_{1})\
+K1n1×n2|y1y2|(1+|x2|θ2+|y1|θ2+|y2|θ2)π(dy1,dy2)\displaystyle~{}~{}~{}+K_{1}\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|\big{(}1+|x_{2}|^{\theta_{2}}+|y_{1}|^{\theta_{2}}+|y_{2}|^{\theta_{2}}\big{)}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\
LR|x1x2|n2(1+|y1|θ1)μx1(dy1)+K1(n1×n2|y1y2|2π(dy1,dy2))12\displaystyle\leq L_{R}|x_{1}-x_{2}|\int_{\mathbb{R}^{n_{2}}}(1+|y_{1}|^{\theta_{1}})\mu^{x_{1}}(\mathrm{d}y_{1})+K_{1}\Big{(}\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|^{2}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}}\
×(n1×n2(1+|x2|2θ2+|y1|2θ2+|y2|2θ2)π(dy1,dy2))12.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}\times\Big{(}\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}\big{(}1+|x_{2}|^{2\theta_{2}}+|y_{1}|^{2\theta_{2}}+|y_{2}|^{2\theta_{2}}\big{)}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}}.

Then due to the arbitrariness of πC(μx1,μx2)\pi\in C(\mu^{x_{1}},\mu^{x_{2}}), under (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, applying Lemma 2.1 yields that for any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} with |x1||x2|R|x_{1}|\vee|x_{2}|\leq R,

|b¯(x1)b¯(x2)|\displaystyle|\bar{b}(x_{1})-\bar{b}(x_{2})| CR|x1x2|(1+|x1|θ1)+C𝕎2(μx1,μx2)(1+|x1|θ2+|x2|θ2)\displaystyle\leq C_{R}|x_{1}-x_{2}|(1+|x_{1}|^{\theta_{1}})+C\mathbb{W}_{2}(\mu^{x_{1}},\mu^{x_{2}})(1+|x_{1}|^{\theta_{2}}+|x_{2}|^{\theta_{2}})\
CR|x1x2|+CR𝕎2(μx1,μx2)CR|x1x2|,\displaystyle\leq C_{R}|x_{1}-x_{2}|+C_{R}\mathbb{W}_{2}(\mu^{x_{1}},\mu^{x_{2}})\leq C_{R}|x_{1}-x_{2}|,

which implies the desired result. ∎

Next we reveal that the modified coefficient b¯(x)\bar{b}(x^{*}) preserves the Khasminskii-like condition for all Δ1(0,1]\Delta_{1}\in(0,1], which is used to obtain the moment bound of the auxiliary process Z¯(t)\bar{Z}(t).

Lemma 4.2.

If (𝐒𝟒)({\bf S4}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with kθ4k\geq\theta_{4} then for any xn1x\in\mathbb{R}^{n_{1}}, Δ1(0,1]\Delta_{1}\in(0,1],

xTb¯(x)C(1+|x|2),xn1.\displaystyle x^{T}\bar{b}(x^{*})\leq C(1+|x|^{2}),~{}~{}~{}~{}x\in\mathbb{R}^{n_{1}}.
Proof.

For xn1x\in\mathbb{R}^{n_{1}} with |x|φ1(KΔ11/2)|x|\leq\varphi^{-1}(K\Delta_{1}^{-1/2}), x=xx=x^{*}. Using (𝐒𝟓)({\bf S5}) implies that

xTb¯(x)\displaystyle x^{T}\bar{b}(x^{*}) =xTn2b(x,y)μx(dy)K4(1+|x|2)+λn2|y|2μx(dy).\displaystyle=x^{T}\int_{\mathbb{R}^{n_{2}}}b(x,y)\mu^{x}(\mathrm{d}y)\ \leq K_{4}(1+|x|^{2})+\lambda\int_{\mathbb{R}^{n_{2}}}|y|^{2}\mu^{x}(\mathrm{d}y).

Owing to (F1)-(F3), (2.1) holds. Then inserting (2.1) into the above inequality and using the Hölder inequality, we yield that

xTb¯(x)C(1+|x|2).\displaystyle x^{T}\bar{b}(x^{*})\leq C(1+|x|^{2}). (4.1)

On the other hand, for any xn1x\in\mathbb{R}^{n_{1}} with |x|>φ1(KΔ11/2)|x|>\varphi^{-1}\big{(}K\Delta_{1}^{-1/2}\big{)}, it follows from the definition of xx^{*} that x=(|x|/φ1(KΔ11/2))xx=\Big{(}{|x|}/{\varphi^{-1}(K\Delta_{1}^{-{1}/{2}})}\Big{)}x^{*}. This together with (4.1) implies that

xTb¯(x)\displaystyle x^{T}\bar{b}(x^{*}) =|x|φ1(KΔ112)(x)Tb¯(x)|x|φ1(KΔ112)C(1+|x|2)\displaystyle=\frac{|x|}{\varphi^{-1}(K\Delta_{1}^{-\frac{1}{2}})}(x^{*})^{T}\bar{b}(x^{*})\leq\frac{|x|}{\varphi^{-1}(K\Delta_{1}^{-\frac{1}{2}})}C(1+|x^{*}|^{2})\
C|x|((φ1(K))1+|x|),\displaystyle\leq C|x|\Big{(}\big{(}\varphi^{-1}(K)\big{)}^{-1}+|x|\Big{)},

where the last inequality used the increasing of φ1\varphi^{-1}. Thus the desired assertion follows. ∎

For any fixed xn1x\in\mathbb{R}^{n_{1}} and integer n0n\geq 0 define an auxiliary process ynx(s)y^{x}_{n}(s) described by

dynx(s)=f(x,ynx(s))ds+g(x,ynx(s))dWn2(s)\displaystyle\mathrm{d}y^{x}_{n}(s)=f(x,y^{x}_{n}(s))\mathrm{d}s+g(x,y^{x}_{n}(s))\mathrm{d}W^{2}_{n}(s) (4.2)

on s0s\geq 0 with initial value ynx(0)=y0y^{x}_{n}(0)=y_{0}. The weak uniqueness of solution of the frozen equation (1.2) implies that for any s0s\geq 0, the distribution of ynx(s)y^{x}_{n}(s) coincides with that of yx(s)y^{x}(s) for any n0n\geq 0. Consequently, according to Lemma 2.1, μx\mu^{x} is also the unique invariant probability measure of transition semigroup of ynx(s)y^{x}_{n}(s) for any n0n\geq 0. Then use the EM scheme for (4.2)

{Yn,0x=y0,Yn,m+1x=Yn,mx+f(x,Yn,mx)Δ2+g(x,Yn,mx)ΔWn,m2,m=0,1,.\begin{cases}Y^{x}_{n,0}=y_{0},~{}~{}~{}~{}~{}~{}~{}~{}\\ Y^{x}_{n,m+1}=Y^{x}_{n,m}+f(x,Y^{x}_{n,m})\Delta_{2}+g(x,Y^{x}_{n,m})\Delta W^{2}_{n,m},~{}m=0,1,\cdot\cdot\cdot.\\ \end{cases} (4.3)

Furthermore, define

Ynx(t)\displaystyle Y^{x}_{n}(t) =Yn,mx,t[mΔ2,(m+1)Δ2),\displaystyle=Y^{x}_{n,m},~{}~{}~{}~{}~{}~{}~{}~{}t\in[m\Delta_{2},(m+1)\Delta_{2}),
Y¯nx(t)\displaystyle\bar{Y}^{x}_{n}(t) =y0+0tf(x,Ynx(s))ds+0tg(x,Ynx(s))dWn2(s).\displaystyle=y_{0}+\int_{0}^{t}f(x,Y^{x}_{n}(s))\mathrm{d}s+\int_{0}^{t}g(x,Y^{x}_{n}(s))\mathrm{d}W^{2}_{n}(s). (4.4)

Let Ynx:={Yn,mx}m=1Y^{x}_{n}:=\{Y^{x}_{n,m}\}_{m=1}^{\infty} and BM(x,Ynx)B_{M}(x,Y^{x}_{n}) be the estimator of b¯(x)\bar{b}(x). In particular, one observes that YXn=YnXna.s.Y^{X^{*}_{n}}=Y^{X^{*}_{n}}_{n}~{}\mathrm{a.s.} Thus

B(Xn,YXn)=BM(Xn,YnXn)a.s.\displaystyle B(X^{*}_{n},Y^{X^{*}_{n}})=B_{M}(X^{*}_{n},Y^{X^{*}_{n}}_{n})~{}~{}\mathrm{a.s.} (4.5)

In order to study BM(x,Ynx)B_{M}(x,Y^{x}_{n}), we prepare some important properties for YnxY^{x}_{n} beforehand.

Lemma 4.3.

If (𝐅𝟏)({\bf F1}) and (𝐅𝟑)({\bf F3}) hold, then there exists a Δ^2(0,1]\hat{\Delta}_{2}\in(0,1] such that for any xn1x\in\mathbb{R}^{n_{1}}, integer n0n\geq 0 and Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}],

supm0𝔼|Yn,mx|kC(1+|x|k),\displaystyle\sup_{m\geq 0}\mathbb{E}|Y^{x}_{n,m}|^{k}\leq C(1+|x|^{k}),

and

supt0𝔼|Y¯nx(t)Ynx(t)|kC(1+|x|k)Δ2k2.\displaystyle\sup_{t\geq 0}\mathbb{E}|\bar{Y}^{x}_{n}(t)-{Y}^{x}_{n}(t)|^{k}\leq C(1+|x|^{k})\Delta_{2}^{\frac{k}{2}}.
Proof.

For any t>0t>0, using the Ito^\mathrm{It\hat{o}} formula, we derive from (4) that

𝔼(ekαkt8|Y¯nx(t)|k)\displaystyle\mathbb{E}\Big{(}e^{\frac{k\alpha_{k}t}{8}}|\bar{Y}^{x}_{n}(t)|^{k}\Big{)} |y0|k+𝔼0tekαks8|Y¯nx(s)|k2[kαk8|Y¯nx(s)|2\displaystyle\leq|y_{0}|^{k}+\mathbb{E}\int_{0}^{t}e^{\frac{k\alpha_{k}s}{8}}|\bar{Y}^{x}_{n}(s)|^{k-2}\Big{[}\frac{k\alpha_{k}}{8}|\bar{Y}^{x}_{n}(s)|^{2}\
+k(Y¯nx(s))Tf(x,Ynx(s))+k(k1)2|g(x,Ynx(s))|2]ds.\displaystyle~{}~{}+k(\bar{Y}^{x}_{n}(s))^{T}f(x,Y^{x}_{n}(s))+\frac{k(k-1)}{2}|g(x,Y^{x}_{n}(s))|^{2}\Big{]}\mathrm{d}s. (4.6)

Invoking the Young inequality, (F1) and (F3) implies that

k(Y¯nx(s))Tf(x,Ynx(s))+k(k1)2|g(x,Ynx(s))|2\displaystyle k(\bar{Y}^{x}_{n}(s))^{T}f(x,Y^{x}_{n}(s))+\frac{k(k-1)}{2}|g(x,Y^{x}_{n}(s))|^{2}\
\displaystyle\leq k2[2(Y¯nx(s))Tf(x,Y¯nx(s))+(k1)|g(x,Y¯nx(s))|2]+k(Y¯nx(s))T(f(x,Ynx(s))f(x,Y¯nx(s)))\displaystyle\frac{k}{2}\Big{[}2\big{(}\bar{Y}^{x}_{n}(s)\big{)}^{T}f(x,\bar{Y}^{x}_{n}(s))+(k-1)|g(x,\bar{Y}^{x}_{n}(s))|^{2}\Big{]}+k\big{(}\bar{Y}^{x}_{n}(s)\big{)}^{T}\big{(}f(x,Y^{x}_{n}(s))-f(x,\bar{Y}^{x}_{n}(s))\big{)}\
+k(k1)2|g(x,Ynx(s))g(x,Y¯nx(s))|2+k(k1)|g(x,Y¯nx(s))||g(x,Ynx(s))g(x,Y¯nx(s))|\displaystyle~{}~{}~{}+\frac{k(k-1)}{2}|g(x,Y^{x}_{n}(s))-g(x,\bar{Y}^{x}_{n}(s))|^{2}+k(k-1)|g(x,\bar{Y}^{x}_{n}(s))||g(x,Y^{x}_{n}(s))-g(x,\bar{Y}^{x}_{n}(s))|\
\displaystyle\leq C(1+|x|2)kαk2|Y¯nx(s)|2+C|Ynx(s)Y¯nx(s)|2.\displaystyle C(1+|x|^{2})-\frac{k\alpha_{k}}{2}|\bar{Y}^{x}_{n}(s)|^{2}\ +C|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{2}.

Substituting the above inequality into (4) and using the Young inequality, we get

𝔼(ekαkt8|Y¯nx(t)|k)\displaystyle\mathbb{E}\Big{(}e^{\frac{k\alpha_{k}t}{8}}|\bar{Y}^{x}_{n}(t)|^{k}\Big{)} |y0|k+C(1+|x|2)k2ekαkt8kαk80tekαks8𝔼|Y¯nx(s)|kds\displaystyle\leq|y_{0}|^{k}+C(1+|x|^{2})^{\frac{k}{2}}e^{\frac{k\alpha_{k}t}{8}}{-\frac{k\alpha_{k}}{8}}\int_{0}^{t}e^{\frac{k\alpha_{k}s}{8}}\mathbb{E}|\bar{Y}^{x}_{n}(s)|^{k}\mathrm{d}s
+C0tekαks8𝔼|Ynx(s)Y¯nx(s)|kds.\displaystyle~{}~{}~{}+C\int_{0}^{t}e^{\frac{k\alpha_{k}s}{8}}\mathbb{E}|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{k}\mathrm{d}s. (4.7)

Moreover, it follows from (𝐅𝟏)({\bf F1}) and (4) that

𝔼|Ynx(s)Y¯nx(s)|k\displaystyle\mathbb{E}|{Y}^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{k}\leq 2k1Δ2k𝔼|f(x,Ynx(s))|k+2k1Δ2k2𝔼|g(x,Ynx(s))|k\displaystyle 2^{k-1}\Delta_{2}^{k}\mathbb{E}|f(x,Y^{x}_{n}(s))|^{k}+2^{k-1}\Delta_{2}^{\frac{k}{2}}\mathbb{E}|g(x,Y^{x}_{n}(s))|^{k}\
\displaystyle\leq CΔ2k2𝔼(1+|x|k+|Ynx(s)|k)\displaystyle C\Delta_{2}^{\frac{k}{2}}\mathbb{E}\big{(}1+|x|^{k}+|Y^{x}_{n}(s)|^{k}\big{)}\
\displaystyle\leq CΔ2k2(1+|x|k+𝔼|Y¯nx(s)|k+𝔼|Ynx(s)Y¯nx(s)|k).\displaystyle C\Delta_{2}^{\frac{k}{2}}\Big{(}1+|x|^{k}+\mathbb{E}|\bar{Y}^{x}_{n}(s)|^{k}+\mathbb{E}|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{k}\Big{)}.

Choose a constant Δ2(0,1]{\Delta}^{\prime}_{2}\in(0,1] small enough such that C(Δ2)k21/2C({{\Delta}}_{2}^{\prime})^{\frac{k}{2}}\leq 1/2. Then, for any Δ2(0,Δ2]\Delta_{2}\in(0,{\Delta}_{2}^{\prime}],

𝔼|Ynx(s)Y¯nx(s)|kCΔ2k2(1+|x|k+𝔼|Y¯nx(s)|k).\displaystyle\mathbb{E}|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{k}\leq C\Delta_{2}^{\frac{k}{2}}(1+|x|^{k}+\mathbb{E}|\bar{Y}^{x}_{n}(s)|^{k}). (4.8)

Inserting (4.8) into (4) leads to

𝔼(ekαkt8|Y¯nx(t)|k)|y0|k+C(1+|x|k)ekαkt8(kαk8CΔ2k2)0tekαks8𝔼|Y¯nx(s)|kds.\displaystyle\mathbb{E}\Big{(}e^{\frac{k\alpha_{k}t}{8}}|\bar{Y}^{x}_{n}(t)|^{k}\Big{)}\ \leq|y_{0}|^{k}+C(1+|x|^{k})e^{\frac{k\alpha_{k}t}{8}}-\Big{(}\frac{k\alpha_{k}}{8}-C\Delta_{2}^{\frac{k}{2}}\Big{)}\int_{0}^{t}e^{\frac{k\alpha_{k}s}{8}}\mathbb{E}|\bar{Y}^{x}_{n}(s)|^{k}\mathrm{d}s.

Furthermore, choose Δ^2(0,Δ2]\hat{\Delta}_{2}\in(0,{\Delta}_{2}^{\prime}] small enough such that C(Δ^2)k2qαk/8C\big{(}\hat{\Delta}_{2}\big{)}^{\frac{k}{2}}\leq q\alpha_{k}/8. Then, for any Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}], we derive that

𝔼(ekαkt8|Y¯nx(t)|k)\displaystyle\mathbb{E}\Big{(}e^{\frac{k\alpha_{k}t}{8}}|\bar{Y}^{x}_{n}(t)|^{k}\Big{)} |y0|k+C(1+|x|k)ekαkt8.\displaystyle\leq|y_{0}|^{k}+C(1+|x|^{k})e^{\frac{k\alpha_{k}t}{8}}.

Then a direct computation gives that

𝔼|Y¯nx(t)|k|y0|kekαkt8+C(1+|x|k),\displaystyle\mathbb{E}|\bar{Y}^{x}_{n}(t)|^{k}\leq|y_{0}|^{k}e^{-\frac{k\alpha_{k}t}{8}}+C(1+|x|^{k}), (4.9)

which implies that

supt0𝔼|Y¯nx(t)|kC(1+|x|k),\displaystyle\sup_{t\geq 0}\mathbb{E}|\bar{Y}^{x}_{n}(t)|^{k}\leq C(1+|x|^{k}),

which implies the first desired result. Then substituting the above inequality into (4.8) gives the another desired result. The proof is complete. ∎

Lemma 4.4.

If (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold, then for any fixed xn1x\in\mathbb{R}^{n_{1}}, integer n0n\geq 0 and Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}],

supm0𝔼|Yn,mxynx(mΔ2)|2C(1+|x|2)Δ2.\displaystyle\sup_{m\geq 0}\mathbb{E}\big{|}Y^{x}_{n,m}-y^{x}_{n}(m\Delta_{2})\big{|}^{2}\leq C(1+|x|^{2})\Delta_{2}.
Proof.

In view of (4.2) and (4), define v¯nx(t):=Y¯nx(t)ynx(t)\bar{v}^{x}_{n}(t):=\bar{Y}_{n}^{x}(t)-y^{x}_{n}(t) described by

dv¯nx(t)=(f(x,Ynx(t))f(x,ynx(t)))dt+(g(x,Ynx(t))g(x,ynx(t)))dWn2(t).\displaystyle\!\mathrm{d}\bar{v}^{x}_{n}(t)=\Big{(}f(x,Y^{x}_{n}(t))\!-\!f(x,y^{x}_{n}(t))\Big{)}\mathrm{d}t+\Big{(}g(x,Y^{x}_{n}(t))-g(x,y^{x}_{n}(t))\Big{)}\mathrm{d}W^{2}_{n}(t).\

Using the Ito^\mathrm{It\hat{o}} formula we arrive at

𝔼(eβt4|v¯nx(t)|2)\displaystyle\mathbb{E}\Big{(}e^{\frac{\beta t}{4}}|\bar{v}^{x}_{n}(t)|^{2}\Big{)}\leq 𝔼0t[β4eβs4|v¯nx(s)|2+eβs4(2(v¯nx(s))T[f(x,Ynx(s))\displaystyle\mathbb{E}\int_{0}^{t}\bigg{[}\frac{\beta}{4}e^{\frac{\beta s}{4}}\big{|}\bar{v}^{x}_{n}(s)\big{|}^{2}+e^{\frac{\beta s}{4}}\Big{(}2(\bar{v}^{x}_{n}(s))^{T}\big{[}f(x,Y^{x}_{n}(s))\
f(x,ynx(s))]+|g(x,Ynx(s))g(x,ynx(s))|2)]ds.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}-f(x,y^{x}_{n}(s))\big{]}+\big{|}g(x,Y^{x}_{n}(s))-g(x,y^{x}_{n}(s))\big{|}^{2}\Big{)}\Bigg{]}\mathrm{d}s. (4.10)

Invoking (F1), (F2) and the Young inequality yields that

2(v¯nx(s))T[f(x,Ynx(s))f(x,ynx(s))]+|g(x,Ynx(s))g(x,ynx(s))|2\displaystyle 2(\bar{v}^{x}_{n}(s))^{T}\big{[}f(x,Y^{x}_{n}(s))-f(x,y^{x}_{n}(s))\big{]}+\big{|}g(x,Y^{x}_{n}(s))-g(x,y^{x}_{n}(s))\big{|}^{2}\
\displaystyle\leq 2(v¯nx(s))T[f(x,Y¯nx(s))f(x,ynx(s))]+|g(x,Y¯nx(s))g(x,ynx(s))|2\displaystyle 2(\bar{v}^{x}_{n}(s))^{T}\big{[}f(x,\bar{Y}^{x}_{n}(s))-f(x,y^{x}_{n}(s))\big{]}+\big{|}g(x,\bar{Y}^{x}_{n}(s))-g(x,y^{x}_{n}(s))\big{|}^{2}\
+2(v¯nx(s))T[f(x,Ynx(s))f(x,Y¯nx(s))]+|g(x,Ynx(s))g(x,Y¯nx(s))|2\displaystyle~{}~{}~{}+2(\bar{v}^{x}_{n}(s))^{T}\big{[}f(x,Y^{x}_{n}(s))-f(x,\bar{Y}^{x}_{n}(s))\big{]}+\big{|}g(x,Y^{x}_{n}(s))-g(x,\bar{Y}^{x}_{n}(s))\big{|}^{2}\
+2|g(x,Y¯nx(s))g(x,ynx(s))||g(x,Ynx(s))g(x,Y¯nx(s))|\displaystyle~{}~{}~{}+2\big{|}g(x,\bar{Y}^{x}_{n}(s))-g(x,y^{x}_{n}(s))\big{|}\big{|}g(x,Y^{x}_{n}(s))-g(x,\bar{Y}^{x}_{n}(s))\big{|}\
\displaystyle\leq β|v¯nx(s)|2+|v¯nx(s)||Ynx(s)Y¯nx(s)|+C|Ynx(s)Y¯nx(s)|2\displaystyle-\beta|\bar{v}^{x}_{n}(s)|^{2}+|\bar{v}^{x}_{n}(s)||Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|+C|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{2}\
\displaystyle\leq β2|v¯nx(s)|2+C|Ynx(s)Y¯nx(s)|2.\displaystyle-\frac{\beta}{2}|\bar{v}^{x}_{n}(s)|^{2}+C|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{2}.

Then inserting the above inequality into (4) and using (F1) and (F3), we derive from the result of lemma 4.3 that

eβt4𝔼|v¯nx(t)|2\displaystyle e^{\frac{\beta t}{4}}\mathbb{E}|\bar{v}^{x}_{n}(t)|^{2} C0teβs4𝔼|Ynx(s)Y¯nx(s)|2dsC(1+|x|2)Δ2eβt4,\displaystyle\leq C\int_{0}^{t}e^{\frac{\beta s}{4}}\mathbb{E}|Y^{x}_{n}(s)-\bar{Y}^{x}_{n}(s)|^{2}\mathrm{d}s\leq C(1+|x|^{2})\Delta_{2}e^{\frac{\beta t}{4}},

which yields the desired result. ∎

Lemma 4.5.

Under (𝐅𝟏)({\bf F1}) and (𝐅𝟐)({\bf F2}), there exists a constant Δ¯2(0,1]\bar{\Delta}_{2}\in(0,1] such that for any Δ2(0,Δ¯2)\Delta_{2}\in(0,\bar{\Delta}_{2}), y,zn2y,z\in\mathbb{R}^{n_{2}}, xn1x\in\mathbb{R}^{n_{1}}, integers n0n\geq 0 and m0m\geq 0,

𝔼|Yn,mx,yYn,mx,z|2C|yz|2eβmΔ24.\displaystyle\mathbb{E}|Y^{x,y}_{n,m}-Y^{x,z}_{n,m}|^{2}\leq C|y-z|^{2}e^{\frac{-\beta m\Delta_{2}}{4}}.
Proof.

For notation brevity, set vn,mx:=Yn,mx,yYn,mx,zv^{x}_{n,m}:=Y^{x,y}_{n,m}-Y^{x,z}_{n,m},

F(x,Yn,mx,y,Yn,mx,z):=f(x,Yn,mx,y)f(x,Yn,mx,z),F(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m}):=f(x,Y^{x,y}_{n,m})-f(x,Y^{x,z}_{n,m}),
G(x,Yn,mx,y,Yn,mx,z):=g(x,Yn,mx,y)g(x,Yn,mx,z).G(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m}):=g(x,Y^{x,y}_{n,m})-g(x,Y^{x,z}_{n,m}).

In view of (4.3), for any integer m0m\geq 0 we have

vn,m+1x=vn,mx+F(x,Yn,mx,y,Yn,mx,z)Δ2+G(x,Yn,mx,y,Yn,mx,z)ΔWn,m2.\displaystyle v^{x}_{n,m+1}=v^{x}_{n,m}+F(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta_{2}+G(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta W^{2}_{n,m}.

Then we derive that

|vn,m+1x|2\displaystyle|v^{x}_{n,m+1}|^{2} =|vn,mx|2+2(vn,mx)TF(x,Yn,mx,y,Yn,mx,z)Δ2+|G(x,Yn,mx,y,Yn,mx,z)ΔWn,m2|2\displaystyle=|v^{x}_{n,m}|^{2}+2(v^{x}_{n,m})^{T}F(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta_{2}+|G(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta W^{2}_{n,m}|^{2}\
+|F(x,Yn,mx,y,Yn,mx,z)|2Δ22+2(vn,mx)TG(x,Yn,mx,y,Yn,mx,z)ΔWn,m2\displaystyle~{}~{}~{}+|F(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})|^{2}\Delta_{2}^{2}+2(v^{x}_{n,m})^{T}G(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta W^{2}_{n,m}\
+2FT(x,Yn,mx,y,Yn,mx,z)G(x,Yn,mx,y,Yn,mx,z)ΔWn,m2Δ2.\displaystyle~{}~{}~{}+2F^{T}(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})G(x,Y^{x,y}_{n,m},Y^{x,z}_{n,m})\Delta W^{2}_{n,m}\Delta_{2}. (4.11)

For any integers n1n\geq 1, l1l\geq 1, denote by n,m2\mathcal{F}^{2}_{n,m} the σ\sigma-algebra generated by

{Wn2(s):0smΔ2}.\Big{\{}W^{2}_{n}(s):0\leq s\leq m\Delta_{2}\Big{\}}.

The fact that ΔWn,m2\Delta W^{2}_{n,m} is independent of n,m2\mathcal{F}^{2}_{n,m} implies that

𝔼(AΔWn,m2|n,m2)=0,𝔼(|AΔWn,m2|2|n,m2)=CΔ2,An2×d2.\displaystyle\mathbb{E}\big{(}A\Delta W^{2}_{n,m}|\mathcal{F}^{2}_{n,m}\big{)}=0,~{}\mathbb{E}\big{(}|A\Delta W^{2}_{n,m}|^{2}|\mathcal{F}^{2}_{n,m}\big{)}=C\Delta_{2},~{}\forall A\in\mathbb{R}^{n_{2}\times d_{2}}. (4.12)

Then taking expectation on both sides for (4) and using (F1), (F2) and (4.12) imply that

𝔼|vn,m+1x|2=𝔼|vn,mx|2βΔ2𝔼|vn,mx|2+L2Δ22𝔼|vn,mx|2.\displaystyle\mathbb{E}|v^{x}_{n,m+1}|^{2}=\mathbb{E}|v^{x}_{n,m}|^{2}-\beta\Delta_{2}\mathbb{E}|v^{x}_{n,m}|^{2}+L^{2}\Delta_{2}^{2}\mathbb{E}|v^{x}_{n,m}|^{2}. (4.13)

Choosing Δ¯2<Δ^2(β/2L2)(2/β)\bar{\Delta}_{2}<\hat{\Delta}_{2}\wedge(\beta/2L^{2})\wedge(2/\beta), for any Δ2(0,Δ¯2)\Delta_{2}\in(0,\bar{\Delta}_{2}), we obtain that

𝔼|vn,m+1x|2(1βΔ22)𝔼|vn,mx|2.\displaystyle\mathbb{E}|v^{x}_{n,m+1}|^{2}\leq\Big{(}1-\frac{\beta\Delta_{2}}{2}\Big{)}\mathbb{E}|v^{x}_{n,m}|^{2}. (4.14)

Thus, 𝔼|vn,m+1x|2(1βΔ22)m|yz|2|yz|2eβmΔ22\mathbb{E}|v^{x}_{n,m+1}|^{2}\leq(1-\frac{\beta\Delta_{2}}{2})^{m}|y-z|^{2}\leq|y-z|^{2}e^{-\frac{\beta m\Delta_{2}}{2}}, where the last step used the elementary inequality that 1βΔ22eβΔ221-\frac{\beta\Delta_{2}}{2}\leq e^{-\frac{\beta\Delta_{2}}{2}}, which implies the desired result. ∎

Lemma 4.6.

If (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold, then for any fixed xn1x\in\mathbb{R}^{n_{1}}, integer n0n\geq 0 and Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}], YnxY^{x}_{n} determined by (4.3) admits a unique invariant measure denoted by μx,Δ2𝒫k(n2)\mu^{x,\Delta_{2}}\in\mathcal{P}_{k}(\mathbb{R}^{n_{2}}) which satisfies that

n2|y|kμx,Δ2(dy)C(1+|x|k).\displaystyle\int_{\mathbb{R}^{n_{2}}}|y|^{k}\mu^{x,\Delta_{2}}(\mathrm{d}y)\leq C(1+|x|^{k}).
Proof.

Since the numerical solutions Ynx,n=1,,Y^{x}_{n},~{}n=1,\cdots,\infty  are i.i.d and have Markov property, for any Δ2(0,1)]\Delta_{2}\in(0,1)], use mΔ2x,Δ2\mathbb{P}^{x,\Delta_{2}}_{m\Delta_{2}} to denote the same discrete Markov semigroup of YnxY^{x}_{n}. Under (F1)-(F3), with the help of Lemmas 4.3-4.5, the existence and uniqueness of invariant measure of YnxY^{x}_{n} follows by imitating the argument for Lemma 2.1. On the other hand, since {Wn2(t),t0}n0\{W^{2}_{n}(t),t\geq 0\}_{n\geq 0} is a mutually independent Brownian motion sequence defined on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), one observes from (4.3) that {Ynx}n0\{Y^{x}_{n}\}_{n\geq 0} are independent identically distributed. As a result, YnxY^{x}_{n} has the same invariant probability measure denoted by μx,Δ2\mu^{x,\Delta_{2}} for any n0n\geq 0. Furthermore, it follows from (4.9) that

n2(|y|kN)μx,Δ2(dy)\displaystyle\int_{\mathbb{R}^{n_{2}}}(|y|^{k}\wedge N)\mu^{x,\Delta_{2}}(\mathrm{d}y) =n2𝔼(|Yn,mx,y|kN)μx,Δ2(dy)\displaystyle=\int_{\mathbb{R}^{n_{2}}}\mathbb{E}(|Y^{x,y}_{n,m}|^{k}\wedge N)\mu^{x,\Delta_{2}}(\mathrm{d}y)\
n2(𝔼|Yn,mx,y|kN)μx,Δ2(dy)\displaystyle\leq\int_{\mathbb{R}^{n_{2}}}(\mathbb{E}|Y^{x,y}_{n,m}|^{k}\wedge N)\mu^{x,\Delta_{2}}(\mathrm{d}y)\
n2(|y|keqαkmΔ28N)μx,Δ2(dy)+C(1+|x|k),\displaystyle\leq\int_{\mathbb{R}^{n_{2}}}\Big{(}|y|^{k}e^{-\frac{q\alpha_{k}m\Delta_{2}}{8}}\wedge N\Big{)}\mu^{x,\Delta_{2}}(\mathrm{d}y)+C(1+|x|^{k}),

where the identity is due to the invariance of invariant measure μx,Δ2\mu^{x,\Delta_{2}} and the first inequality holds by Jensen’s inequality since xNx,xx\mapsto N\wedge x,x\in\mathbb{R} is a concave function. Then, taking mm\rightarrow\infty and using the dominated convergence theorem, we deduce that

n2(|y|kN)μx,Δ2(dy)C(1+|x|k).\displaystyle\int_{\mathbb{R}^{n_{2}}}(|y|^{k}\wedge N)\mu^{x,\Delta_{2}}(\mathrm{d}y)\leq C(1+|x|^{k}).

Letting NN\rightarrow\infty and applying the monotone convergence theorem, we get

n2|y|kμx,Δ2(dy)C(1+|x|k).\displaystyle\int_{\mathbb{R}^{n_{2}}}|y|^{k}\mu^{x,\Delta_{2}}(\mathrm{d}y)\leq C(1+|x|^{k}).

The proof is complete. ∎

Taking Lemma 4.4 into consideration, we deduce the convergence rate between numerical invariant measure μx,Δ2\mu^{x,\Delta_{2}} and underlying invariant measure μx\mu^{x} in W2W_{2}-distance.

Lemma 4.7.

Under (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}), for any fixed xn1x\in\mathbb{R}^{n_{1}} and Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}],

W2(μx,μx,Δ2)C(1+|x|)Δ212.\displaystyle W_{2}(\mu^{x},\mu^{x,\Delta_{2}})\leq C(1+|x|)\Delta_{2}^{\frac{1}{2}}.
Proof.

From the proofs of Lemmas 2.1 and 4.6, we know that

limmW2(mΔ2xδ0,μx)=0\displaystyle\lim_{m\rightarrow\infty}W_{2}(\mathbb{P}^{x}_{m\Delta_{2}}\delta_{0},\mu^{x})=0

and

limmW2(mΔ2x,Δ2,μx,Δ2)=0.\displaystyle\lim_{m\rightarrow\infty}W_{2}(\mathbb{P}^{x,\Delta_{2}}_{m\Delta_{2}},\mu^{x,\Delta_{2}})=0.

The above inequalities together with Lemma 4.4 derive that

W2(μx,μx,Δ2)\displaystyle W_{2}(\mu^{x},\mu^{x,\Delta_{2}}) limmW2(μx,mΔ2xδ0)+limmW2(mΔ2xδ0,mΔ2x,Δ2δ0)+limmW2(mΔ2x,Δ2δ0,μx,Δ2)\displaystyle\leq\lim_{m\rightarrow\infty}W_{2}(\mu^{x},\mathbb{P}^{x}_{m\Delta_{2}}\delta_{0})+\lim_{m\rightarrow\infty}W_{2}(\mathbb{P}^{x}_{m\Delta_{2}}\delta_{0},\mathbb{P}^{x,\Delta_{2}}_{m\Delta_{2}}\delta_{0})+\lim_{m\rightarrow\infty}W_{2}(\mathbb{P}^{x,\Delta_{2}}_{m\Delta_{2}}\delta_{0},\mu^{x,\Delta_{2}})\
limm(𝔼|ynx(mΔ2)Yn,mx|2)12C(1+|x|)Δ12.\displaystyle\leq\lim_{m\rightarrow\infty}\big{(}\mathbb{E}|y^{x}_{n}(m\Delta_{2})-Y^{x}_{n,m}|^{2}\big{)}^{\frac{1}{2}}\leq C(1+|x|)\Delta^{\frac{1}{2}}.\

Now we in a position to analyze the property of the estimator BM(x,Ynx)B_{M}(x,Y^{x}_{n}).

Lemma 4.8.

If (𝐒𝟓)({\bf S5}), (𝐅𝟏)({\bf F1}) and (𝐅𝟑)({\bf F3}) hold, then for any 2qk2\leq q\leq k, xn1x\in\mathbb{R}^{n_{1}}, Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}], integers n0n\geq 0 and M1M\geq 1

𝔼|xTBM(x,Ynx)|q2C(1+|x|q).\displaystyle\mathbb{E}\big{|}x^{T}B_{M}(x^{*},Y^{x^{*}}_{n})\big{|}^{\frac{q}{2}}\leq C(1+|x|^{q}).
Proof.

For any xn1x\in\mathbb{R}^{n_{1}} with |x|φ1(KΔ11/2)|x|\leq\varphi^{-1}(K\Delta_{1}^{-1/2}), x=xx=x^{*}. Making use of (S5) and the elementary inequality yields that

𝔼|xTBM(x,Ynx)|q2\displaystyle\mathbb{E}\big{|}x^{T}B_{M}(x^{*},Y^{x^{*}}_{n})\big{|}^{\frac{q}{2}} =𝔼|1Mm=1MxTb(x,Yn,mx)|q2𝔼[K4(1+|x|2)+λMm=1M|Yn,mx|2]q2\displaystyle=\mathbb{E}\Big{|}\frac{1}{M}\sum_{m=1}^{M}x^{T}b(x,Y^{x}_{n,m})\Big{|}^{\frac{q}{2}}\leq\mathbb{E}\Big{[}K_{4}(1+|x|^{2})+\frac{\lambda}{M}\sum_{m=1}^{M}|Y^{x}_{n,m}|^{2}\Big{]}^{\frac{q}{2}}\
C(1+|x|q)+CMm=1M𝔼|Yn,mx|q.\displaystyle\leq C(1+|x|^{q})+\frac{C}{M}\sum_{m=1}^{M}\mathbb{E}|Y^{x}_{n,m}|^{q}.

Then by (F1) and (F3), applying Lemma 4.3 and the Hölder inequality implies that for any Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}],

𝔼|xTBM(x,Ynx)|q2\displaystyle\mathbb{E}\big{|}x^{T}B_{M}(x^{*},Y^{x^{*}}_{n})\big{|}^{\frac{q}{2}} C(1+|x|q)+CMm=1M(𝔼|Yn,mx|k)qkC(1+|x|q).\displaystyle\leq C(1+|x|^{q})+\frac{C}{M}\sum_{m=1}^{M}\big{(}\mathbb{E}|Y^{x}_{n,m}|^{k}\big{)}^{\frac{q}{k}}\leq C(1+|x|^{q}). (4.15)

On the other hand, for xn1x\in\mathbb{R}^{n_{1}} with |x|>φ1(KΔ11/2)|x|>\varphi^{-1}(K\Delta_{1}^{-1/2}), x=|x|x/φ1(KΔ11/2)x=|x|x^{*}/\varphi^{-1}(K\Delta_{1}^{-1/2}). One observes that

xTBM(x,Ynx)\displaystyle x^{T}B_{M}(x^{*},Y^{x^{*}}_{n}) =|x|φ1(KΔ112)(x)TBM(x,Ynx).\displaystyle=\frac{|x|}{\varphi^{-1}(K\Delta_{1}^{-\frac{1}{2}})}(x^{*})^{T}B_{M}(x^{*},Y^{x^{*}}_{n}).\

Due to (4.15), we obtain that

𝔼|xTBM(x,Ynx)|q2\displaystyle\mathbb{E}\big{|}x^{T}B_{M}(x^{*},Y^{x^{*}}_{n})\big{|}^{\frac{q}{2}} |x|q2(φ1(KΔ112))q2C(1+|x|q)C|x|q2(1+|x|q2)C(1+|x|q).\displaystyle\leq\frac{|x|^{\frac{q}{2}}}{\big{(}\varphi^{-1}\big{(}K\Delta_{1}^{-\frac{1}{2}}\big{)}\big{)}^{\frac{q}{2}}}C(1+|x^{*}|^{q})\leq C|x|^{\frac{q}{2}}(1+|x^{*}|^{\frac{q}{2}})\leq C(1+|x|^{q}).

Thus the desired assertion follows. ∎

The error of approximating b¯(x)\bar{b}(x) by BM(x,Ynx)B_{M}(x,Y^{x}_{n}) is key to prove the convergence of the MTEM scheme numerical solution. To estimate this error, we introduce an intermediate quantity. Under (𝐒𝟒)({\bf S4}) and (F1)-(F3) with kθ4k\geq\theta_{4}, by virtue of Lemma 4.6, for any fixed xn1x\in\mathbb{R}^{n_{1}} and Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}],

n2|b(x,y)|μx,Δ2\displaystyle\int_{\mathbb{R}^{n_{2}}}|b(x,y)|\mu^{x,\Delta_{2}} K3n2(1+|x|θ3+|y|θ4)μx,Δ2(dy)\displaystyle\leq K_{3}\int_{\mathbb{R}^{n_{2}}}(1+|x|^{\theta_{3}}+|y|^{\theta_{4}})\mu^{x,\Delta_{2}}(\mathrm{d}y)\
C(1+|x|θ3θ4)<,\displaystyle\leq C(1+|x|^{\theta_{3}\vee\theta_{4}})<\infty, (4.16)

which implies that b(x,)b(x,\cdot) is integrable with respect to μx,Δ2\mu^{x,\Delta_{2}}. Then define

b¯Δ2(x)=n2b(x,y)μx,Δ2(dy).\displaystyle\bar{b}^{\Delta_{2}}(x)=\int_{\mathbb{R}^{n_{2}}}b(x,y)\mu^{x,\Delta_{2}}(\mathrm{d}y). (4.17)

An application of the elementary inequality makes the moment estimate of |b¯(x)BM(x,Ynx)||\bar{b}(x)-B_{M}(x,Y^{x}_{n})| can be obtained by analysing the moments of |b¯(x)b¯Δ2(x)||\bar{b}(x)-\bar{b}^{\Delta_{2}}(x)| and |b¯Δ2(x)BM(x,Ynx)||\bar{b}^{\Delta_{2}}(x)-B_{M}(x,Y^{x}_{n})|, respectively.

Lemma 4.9.

Under (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)(\bf{F3}) with k2θ2θ4k\geq 2\theta_{2}\vee\theta_{4}, for any xn1x\in\mathbb{R}^{n_{1}} and Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}],

|b¯(x)b¯Δ2(x)|C(1+|x|θ2+1)Δ2.\displaystyle|\bar{b}(x)-\bar{b}^{\Delta_{2}}(x)|\leq C(1+|x|^{\theta_{2}+1})\Delta_{2}.
Proof.

Under (𝐒𝟒)({\bf S4}), (F1)-(F3) with kθ4k\geq\theta_{4}, in view of (1.3) and (4.17), using (𝐒𝟐)({\bf S2}) and the Hölder inequality yields that

|b¯(x)b¯Δ2(x)|\displaystyle|\bar{b}(x)-\bar{b}^{\Delta_{2}}(x)| =|n2×n2(b(x,y1)b(x,y2))π(dy1,dy2)|\displaystyle=\Big{|}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}\big{(}b(x,y_{1})-b(x,y_{2})\big{)}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{|}\
n2×n2|b(x,y1)b(x,y2)|π(dy1,dy2)\displaystyle\leq\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}\big{|}b(x,y_{1})-b(x,y_{2})\big{|}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\
C(n2×n2|y1y2|2π(dy1,dy2))12\displaystyle\leq C\Big{(}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|^{2}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}}\
×(n2×n2(1+|x|2θ2+|y1|2θ2+|y2|2θ2)π(dy2,dy2))12,\displaystyle~{}~{}~{}\times\Big{(}\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}(1+|x|^{2\theta_{2}}+|y_{1}|^{2\theta_{2}}+|y_{2}|^{2\theta_{2}})\pi(\mathrm{d}y_{2},\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}},\

where π𝒞(μx,μx,Δ2)\pi\in\mathcal{C}(\mu^{x},\mu^{x,\Delta_{2}}) is arbitrary. Thus we derive that

|b¯(x)b¯Δ2(x)|\displaystyle|\bar{b}(x)-\bar{b}^{\Delta_{2}}(x)| CW2(μx,μx,Δ2)×(1+|x|2θ2+n2|y1|2θ2μx(dy1)+n2|y2|2θ2μx,Δ2(dy2))12.\displaystyle\leq CW_{2}(\mu^{x},\mu^{x,\Delta_{2}})\times\Big{(}1+|x|^{2\theta_{2}}+\int_{\mathbb{R}^{n_{2}}}|y_{1}|^{2\theta_{2}}\mu^{x}(\mathrm{d}y_{1})+\int_{\mathbb{R}^{n_{2}}}|y_{2}|^{2\theta_{2}}\mu^{x,\Delta_{2}}(\mathrm{d}y_{2})\Big{)}^{\frac{1}{2}}.

Then due to (F1)-(F3) with k2θ2k\geq 2\theta_{2}, applying Lemmas 2.1, 4.6 and 4.7 implies that

|b¯(x)b¯Δ2(x)|\displaystyle|\bar{b}(x)-\bar{b}^{\Delta_{2}}(x)| C(1+|x|θ2+1)Δ212.\displaystyle\leq C(1+|x|^{\theta_{2}+1})\Delta_{2}^{\frac{1}{2}}.

The proof is complete. ∎

Now we proceed to estimate 𝔼|b¯Δ2(x)BM(x,Ynx)|2\mathbb{E}|\bar{b}^{\Delta_{2}}(x)-B_{M}(x,Y^{x}_{n})|^{2}. Before that, we prepare a crucial lemma.

Lemma 4.10.

Under (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k2θ2θ4k\geq 2\theta_{2}\vee\theta_{4}, for any xn1x\in\mathbb{R}^{n_{1}}, yn2y\in\mathbb{R}^{n_{2}}, Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}] and integers n0n\geq 0, M1M\geq 1,

|𝔼b(x,Yn,mx,y)b¯Δ2(x)|C(1+|x|θ2+1+|y|θ2+1)eβmΔ28.\displaystyle|\mathbb{E}b(x,Y^{x,y}_{n,m})-\bar{b}^{\Delta_{2}}(x)|\leq C(1+|x|^{\theta_{2}+1}+|y|^{\theta_{2}+1})e^{\frac{-\beta m\Delta_{2}}{8}}.
Proof.

Under (S4) and (F1)-(F3) with kθ4k\geq\theta_{4}, according to (4.17) and the invariance of invariant measure μx,Δ2\mu^{x,\Delta_{2}}, we have

b¯Δ2(x)\displaystyle\bar{b}^{\Delta_{2}}(x) =limkn2b(x,z)I{|z|k}μx,Δ2(dz)\displaystyle=\lim_{k\rightarrow\infty}\int_{\mathbb{R}^{n_{2}}}b(x,z)I_{\{|z|\leq k\}}\mu^{x,\Delta_{2}}(\mathrm{d}z)\
limkn2𝔼(b(x,Yn,mx,z)I{|Yn,mx,z|k})μx,Δ2(dz).\displaystyle\leq\lim_{k\rightarrow\infty}\int_{\mathbb{R}^{n_{2}}}\mathbb{E}\Big{(}b(x,Y^{x,z}_{n,m})I_{\{|Y^{x,z}_{n,m}|\leq k\}}\Big{)}\mu^{x,\Delta_{2}}(\mathrm{d}z). (4.18)

Obviously, limkb(x,Yn,mx,z)I{|Yn,mx,z|k}=b(x,Yn,mx,z),a.s.\lim\limits_{k\rightarrow\infty}b(x,Y^{x,z}_{n,m})I_{\{|Y^{x,z}_{n,m}|\leq k\}}=b(x,Y^{x,z}_{n,m}),~{}\mathrm{a.s.} for any zn2z\in\mathbb{R}^{n_{2}}. In addition, by (𝐒𝟒)({\bf S4}) and (F1)-(F3) with kθ4k\geq\theta_{4}, using (4.9) and Lemma 4.6 yields that

n2𝔼|b(x,Yn,mx,z)|μx,Δ2(dz)\displaystyle\int_{\mathbb{R}^{n_{2}}}\mathbb{E}|b(x,Y^{x,z}_{n,m})|\mu^{x,\Delta_{2}}(\mathrm{d}z) C(1+|x|θ3+n2𝔼|Yn,mx,z|θ4μx,Δ2(dz))\displaystyle\leq C\Big{(}1+|x|^{\theta_{3}}+\int_{\mathbb{R}^{n_{2}}}\mathbb{E}|Y^{x,z}_{n,m}|^{\theta_{4}}\mu^{x,\Delta_{2}}(\mathrm{d}z)\Big{)}\
C(1+|x|θ3θ4+n2|z|θ4μx,Δ2(dz))\displaystyle\leq C\Big{(}1+|x|^{\theta_{3}\vee\theta_{4}}+\int_{\mathbb{R}^{n_{2}}}|z|^{\theta_{4}}\mu^{x,\Delta_{2}}(\mathrm{d}z)\Big{)}\
C(1+|x|θ3θ4)<.\displaystyle\leq C(1+|x|^{\theta_{3}\vee\theta_{4}})<\infty.

Then applying the dominated convergence theorem for (4) we derive that

b¯Δ2(x)=n2𝔼b(x,Yn,mx,z)μx,Δ2(dz).\displaystyle\bar{b}^{\Delta_{2}}(x)=\int_{\mathbb{R}^{n_{2}}}\mathbb{E}b(x,Y^{x,z}_{n,m})\mu^{x,\Delta_{2}}(\mathrm{d}z).

As a result, we have

|𝔼b(x,Yn,mx,y)b¯Δ2(x)|\displaystyle|\mathbb{E}b(x,Y^{x,y}_{n,m})-\bar{b}^{\Delta_{2}}(x)| =|𝔼b(x,Yn,mx,y)n2𝔼b(x,Yn,mx,z)μx,Δ2(dz)|\displaystyle=\Big{|}\mathbb{E}b(x,Y^{x,y}_{n,m})-\int_{\mathbb{R}^{n_{2}}}\mathbb{E}b(x,Y^{x,z}_{n,m})\mu^{x,\Delta_{2}}(\mathrm{d}z)\Big{|}\
n2𝔼|b(x,Yn,mx,y)b(x,Yn,mx,z)|μx,Δ2(dz).\displaystyle\leq\int_{\mathbb{R}^{n_{2}}}\mathbb{E}\big{|}b(x,Y^{x,y}_{n,m})-b(x,Y^{x,z}_{n,m})\big{|}\mu^{x,\Delta_{2}}(\mathrm{d}z).\

Further using (𝐒𝟐)({\bf S2}) and the Hölder inequality gives that

|𝔼b(x,Yn,mx,y)b¯Δ2(x)|\displaystyle|\mathbb{E}b(x,Y^{x,y}_{n,m})-\bar{b}^{\Delta_{2}}(x)|\
\displaystyle\leq K1n2𝔼(|Yn,mx,yYn,mx,z|(1+|x|θ2+|Yn,mx,y|θ2+|Yn,mx,z|θ2))μx,Δ2(dz)\displaystyle K_{1}\int_{\mathbb{R}^{n_{2}}}\mathbb{E}\Big{(}|Y^{x,y}_{n,m}-Y^{x,z}_{n,m}|(1+|x|^{\theta_{2}}+|Y^{x,y}_{n,m}|^{\theta_{2}}+|Y^{x,z}_{n,m}|^{\theta_{2}})\Big{)}\mu^{x,\Delta_{2}}(\mathrm{d}z)\
\displaystyle\leq Cn2[(𝔼|Yn,mx,yYn,mx,z|2)12(𝔼(1+|x|2θ2+|Yn,mx,y|2θ2+|Yn,mx,z|2θ2))12]μx,Δ2(dz).\displaystyle C\int_{\mathbb{R}^{n_{2}}}\Big{[}\big{(}\mathbb{E}|Y^{x,y}_{n,m}-Y^{x,z}_{n,m}|^{2}\big{)}^{\frac{1}{2}}\big{(}\mathbb{E}(1+|x|^{2\theta_{2}}+|Y^{x,y}_{n,m}|^{2\theta_{2}}+|Y^{x,z}_{n,m}|^{2\theta_{2}})\big{)}^{\frac{1}{2}}\Big{]}\mu^{x,\Delta_{2}}(\mathrm{d}z).\

Under (F1)-(F3) with k2θ2k\geq 2\theta_{2}, utilizing Lemmas 4.3 and 4.5 we get

|𝔼b(x,Yn,mx,y)b¯Δ2(x)|\displaystyle|\mathbb{E}b(x,Y^{x,y}_{n,m})-\bar{b}^{\Delta_{2}}(x)| CeβmΔ28n2|yz|(1+|x|θ2+|y|θ2+|z|θ2)μx,Δ2(dz)\displaystyle\leq Ce^{\frac{-\beta m\Delta_{2}}{8}}\int_{\mathbb{R}^{n_{2}}}|y-z|(1+|x|^{\theta_{2}}+|y|^{\theta_{2}}+|z|^{\theta_{2}})\mu^{x,\Delta_{2}}(\mathrm{d}z)\
CeβmΔ28n2(1+|x|θ2+1+|y|θ2+1+|z|θ2+1)μx,Δ2(dz)\displaystyle\leq Ce^{\frac{-\beta m\Delta_{2}}{8}}\int_{\mathbb{R}^{n_{2}}}(1+|x|^{\theta_{2}+1}+|y|^{\theta_{2}+1}+|z|^{\theta_{2}+1})\mu^{x,\Delta_{2}}(\mathrm{d}z)\
C(1+|x|θ2+1+|y|θ2+1)eβmΔ28.\displaystyle\leq C(1+|x|^{\theta_{2}+1}+|y|^{\theta_{2}+1})e^{\frac{-\beta m\Delta_{2}}{8}}.

The proof is complete. ∎

Lemma 4.11.

Under (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k2θ22θ4(θ2+θ4+1)k\geq 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), for any xn1x\in\mathbb{R}^{n_{1}}, Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}] and integers n0n\geq 0, M1M\geq 1,

𝔼|BM(x,Ynx)b¯Δ2(x)|2C(1+|x|2θ32θ4(θ2+θ3θ4+1))(1M+1MΔ2).\displaystyle\mathbb{E}|B_{M}(x,Y^{x}_{n})-\bar{b}^{\Delta_{2}}(x)|^{2}\leq C(1+|x|^{2\theta_{3}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{3}\vee\theta_{4}+1)})\Big{(}\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.
Proof.

In light of (3.4), we derive that for any xn1x\in\mathbb{R}^{n_{1}},

𝔼|BM(x,Ynx)b¯Δ2(x)|2\displaystyle\mathbb{E}\Big{|}B_{M}(x,Y^{x}_{n})-\bar{b}^{\Delta_{2}}(x)\Big{|}^{2} =1M2m,l=1M𝔼Um,l=1M2m=1M𝔼Um,m+2M2l=1Mm=l+1M𝔼Um,l,\displaystyle=\frac{1}{M^{2}}\sum_{m,l=1}^{M}\mathbb{E}U_{m,l}=\frac{1}{M^{2}}\sum_{m=1}^{M}\mathbb{E}U_{m,m}+\frac{2}{M^{2}}\sum_{l=1}^{M}\sum_{m=l+1}^{M}\mathbb{E}U_{m,l}, (4.19)

where

Um,l=(b(x,Yn,mx)b¯Δ2(x))(b(x,Yn,lx)b¯Δ2(x)).\displaystyle U_{m,l}=\Big{(}b\big{(}x,Y^{x}_{n,m}\big{)}-\bar{b}^{\Delta_{2}}(x)\Big{)}\Big{(}b\big{(}x,Y^{x}_{n,l}\big{)}\!-\bar{b}^{\Delta_{2}}(x)\Big{)}.

By (S4), (F1) and (F3) with k2θ4k\geq 2\theta_{4}, invoking Lemma 4.3 and the Ho¨lder\mathrm{H\ddot{o}lder} inequality, we obtain that

𝔼|b(x,Yn,mx)|2\displaystyle\mathbb{E}\big{|}b(x,Y^{x}_{n,m})\big{|}^{2} C𝔼(1+|x|2θ3+|Yn,mx|2θ4)\displaystyle\leq C\mathbb{E}\Big{(}1+|x|^{2\theta_{3}}+|Y^{x}_{n,m}|^{2\theta_{4}}\Big{)}\
C(1+|x|2θ3)+C(𝔼|Yn,mx|k)2θ4kC(1+|x|2(θ3θ4)).\displaystyle\leq C(1+|x|^{2\theta_{3}})+C\Big{(}\mathbb{E}|Y^{x}_{n,m}|^{k}\Big{)}^{\frac{2\theta_{4}}{k}}\leq C(1+|x|^{2(\theta_{3}\vee\theta_{4})}).

Then using the elementary inequality along with the above inequality and (4), for any m,l1m,l\geq 1, we yield that for any xn1x\in\mathbb{R}^{n_{1}},

𝔼|Um,l|\displaystyle\mathbb{E}|U_{m,l}| 𝔼|b(x,Yn,mx)|2+𝔼|b(x,Yn,lx)|2+2𝔼|b¯Δ2(x)|2\displaystyle\leq\mathbb{E}\big{|}b(x,Y^{x}_{n,m})\big{|}^{2}+\mathbb{E}\big{|}b(x,Y^{x}_{n,l})\big{|}^{2}+2\mathbb{E}|\bar{b}^{\Delta_{2}}(x)|^{2}\
C(1+|x|2(θ3θ4))<,\displaystyle\leq C(1+|x|^{2(\theta_{3}\vee\theta_{4})})<\infty, (4.20)

which implies that |Um,l||U_{m,l}| is integrable with respect to \mathbb{P}. To compute precisely, let 𝒢n,l2\mathcal{G}^{2}_{n,l} denote the σ\sigma-algebra generated by

{Wn2(s)Wn2(lΔ2),slΔ2}.\Big{\{}W^{2}_{n}(s)-W^{2}_{n}(l\Delta_{2}),s\geq l\Delta_{2}\Big{\}}.

Obviously, n,l2\mathcal{F}^{2}_{n,l} and 𝒢n,l2\mathcal{G}^{2}_{n,l} are mutually independent. Since Yn,lxY^{x}_{n,l} is n,l2\mathcal{F}^{2}_{n,l}-measurable and independent of 𝒢n,l2\mathcal{G}^{2}_{n,l}, using the result of [40, p.221], we derive that for any xn1x\in\mathbb{R}^{n_{1}} and 1l<mM1\leq l<m\leq M,

𝔼Um,l=\displaystyle\mathbb{E}U_{m,l}= 𝔼[(b(x,Yn,lx)b¯Δ2(x))×𝔼((b(x,Yn,mx)b¯Δ2(x))|n,l2)]\displaystyle\mathbb{E}\Big{[}\big{(}b(x,Y^{x}_{n,l})-\bar{b}^{\Delta_{2}}(x)\big{)}\times\mathbb{E}\Big{(}\big{(}b(x,Y^{x}_{n,m})-\bar{b}^{\Delta_{2}}(x)\big{)}\Big{|}\mathcal{F}^{2}_{n,l}\Big{)}\Big{]}
\displaystyle\leq 𝔼[|b(x,Yn,lx)b¯Δ2(x)|×|𝔼b(x,Yn,mlx,z)b¯Δ2(x)|z=Yn,lx].\displaystyle\mathbb{E}\bigg{[}\big{|}b(x,Y^{x}_{n,l})-\bar{b}^{\Delta_{2}}(x)\big{|}\times\Big{|}\mathbb{E}b\big{(}x,Y^{x,z}_{n,m-l}\big{)}-\bar{b}^{\Delta_{2}}(x)\Big{|}_{z=Y^{x}_{n,l}}\bigg{]}. (4.21)

For any xn1x\in\mathbb{R}^{n_{1}} and yn2y\in\mathbb{R}^{n_{2}}, it follows from (𝐒𝟒)({\bf S4}) and (4) that

|b(x,y)b¯Δ2(x)|\displaystyle|b(x,y)-\bar{b}^{\Delta_{2}}(x)| =|b(x,y)|+|b¯Δ2(x)|C(1+|x|θ3θ4+|y|θ4).\displaystyle=|b(x,y)|+|\bar{b}^{\Delta_{2}}(x)|\leq C(1+|x|^{\theta_{3}\vee\theta_{4}}+|y|^{\theta_{4}}). (4.22)

Owing to (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (F1)-(F3) with k2θ2θ4k\geq 2\theta_{2}\vee\theta_{4}, using Lemma 4.10 derives that

|𝔼b(x,Yn,mlx,z)b¯Δ2(x)|Ceβ(ml)Δ28(1+|x|θ2+1+|z|θ2+1).\displaystyle\Big{|}\mathbb{E}b\big{(}x,Y^{x,z}_{n,m-l})-\bar{b}^{\Delta_{2}}(x)\Big{|}\leq Ce^{-\frac{\beta(m-l)\Delta_{2}}{8}}\big{(}1+|x|^{\theta_{2}+1}+|z|^{\theta_{2}+1}\big{)}.

Using (4.22) and substituting the above inequality into (4) lead to that for any xn1x\in\mathbb{R}^{n_{1}},

𝔼Um,l\displaystyle\mathbb{E}U_{m,l} Ceβ(ml)Δ28𝔼[(1+|x|θ3θ4+|Yn,lx|θ4)\displaystyle\leq Ce^{-\frac{\beta(m-l)\Delta_{2}}{8}}\mathbb{E}\Big{[}\big{(}1+|x|^{\theta_{3}\vee\theta_{4}}+|Y^{x}_{n,l}|^{\theta_{4}}\big{)}\
×(1+|x|θ2+1+|Yn,lx|θ2+1)]\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\times\big{(}1+|x|^{\theta_{2}+1}+|Y^{x}_{n,l}|^{\theta_{2}+1}\big{)}\Big{]}\
Ceβ(ml)Δ28𝔼[(1+|x|θ2+θ3θ4+1+(1+|x|θ3θ4)|Yn,lx|θ2+1\displaystyle\leq Ce^{-\frac{\beta(m-l)\Delta_{2}}{8}}\mathbb{E}\Big{[}\big{(}1+|x|^{\theta_{2}+\theta_{3}\vee\theta_{4}+1}+(1+|x|^{\theta_{3}\vee\theta_{4}})|Y^{x}_{n,l}|^{\theta_{2}+1}\
+(1+|x|θ2+1)|Yn,lx|θ4+|Yn,lx|θ2+θ4+1)].\displaystyle~{}~{}~{}+(1+|x|^{\theta_{2}+1})|Y^{x}_{n,l}|^{\theta_{4}}+|Y^{x}_{n,l}|^{\theta_{2}+\theta_{4}+1}\big{)}\Big{]}.

Due to k(θ2+θ4+1)k\geq(\theta_{2}+\theta_{4}+1), using Lemma 4.3 we deduce that for any 1l<mM1\leq l<m\leq M,

𝔼Um,lCeβ(ml)Δ28(1+|x|θ2+θ3θ4+1).\displaystyle\mathbb{E}U_{m,l}\leq Ce^{-\frac{\beta(m-l)\Delta_{2}}{8}}(1+|x|^{\theta_{2}+\theta_{3}\vee\theta_{4}+1}). (4.23)

Hence, inserting (4) and (4.23) into (4.19) yields that

𝔼|BM(x,Ynx)b¯Δ2(x)|2\displaystyle\mathbb{E}\Big{|}B_{M}(x,Y^{x}_{n})-\bar{b}^{\Delta_{2}}(x)\Big{|}^{2} C(1+|x|2(θ3θ4))M+C(1+|x|θ2+θ3θ4+1)M2l=1Mm=l+1Meβ(ml)Δ28\displaystyle\leq\frac{C(1+|x|^{2(\theta_{3}\vee\theta_{4})})}{M}+\frac{C(1+|x|^{\theta_{2}+\theta_{3}\vee\theta_{4}+1})}{M^{2}}\sum_{l=1}^{M}\sum_{m=l+1}^{M}e^{-\frac{\beta(m-l)\Delta_{2}}{8}}
C(1+|x|2(θ3θ4))M+C(1+|x|θ2+θ3θ4+1)M(eβΔ2/81)\displaystyle\leq\frac{C(1+|x|^{2(\theta_{3}\vee\theta_{4})})}{M}+\frac{C(1+|x|^{\theta_{2}+\theta_{3}\vee\theta_{4}+1})}{M(e^{\beta\Delta_{2}/8}-1)}\
C(1+|x|2θ32θ4(θ2+θ3θ4+1))(1M+1MΔ2),\displaystyle\leq C(1+|x|^{2\theta_{3}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{3}\vee\theta_{4}+1)})\Big{(}\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)},

where the last line used the inequality ex1x,x0e^{x}-1\geq x,\forall x\geq 0. The proof is complete. ∎

Combining Lemmas 4.9 and 4.11, we obtain the estimate of 𝔼|BM(x,Ynx)b¯(x)|2\mathbb{E}|B_{M}(x,Y^{x}_{n})-\bar{b}(x)|^{2} directly.

Lemma 4.12.

Under (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k2θ22θ4(θ2+θ4+1)k\geq 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), for any xn1x\in\mathbb{R}^{n_{1}}, Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}] and integers n0n\geq 0, M1M\geq 1,

𝔼|b¯(x)BM(x,Ynx)|2C(1+|x|2(θ2+1)2θ32θ4)(Δ2+1M+1MΔ2).\displaystyle\mathbb{E}\Big{|}\bar{b}(x)-B_{M}(x,Y^{x}_{n})\Big{|}^{2}\leq C(1+|x|^{2(\theta_{2}+1)\vee 2\theta_{3}\vee 2\theta_{4}})\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.

5 Strong Convergence in ppth moment

With the help of averaging principle, this section aims to prove the strong convergence between the slow component xε(t)x^{\varepsilon}(t) of original system (1.1) and the MTEM scheme numerical solution X(t)X(t) in ppth moment.

Lemma 5.1.

If (𝐒𝟑)({\bf S3})-(𝐒𝟓)({\bf S5}), (𝐅𝟏)({\bf F1}) and (𝐅𝟑)({\bf F3}) hold with k2θ4k\geq 2\theta_{4}, then for any 0<pk/θ40<p\leq k/\theta_{4} and T>0T>0 and M1M\geq 1,

supΔ1(0,1],Δ2(0,Δ^2]𝔼(supt[0,T]|X¯(t)|p)CT,\displaystyle\sup_{\Delta_{1}\in(0,1],\Delta_{2}\in(0,\hat{\Delta}_{2}]}\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)}\leq C_{T},

and

𝔼(sup0tT|X¯(t)X(t)|p)CTΔ1p2.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{X}(t)-X(t)|^{p}\Big{)}\leq C_{T}\Delta_{1}^{\frac{p}{2}}.
Proof.

For 2pk/θ42\leq p\leq k/\theta_{4}, using the Ito^\mathrm{It\hat{o}} formula, we deduce from (3.7) that for any t0t\geq 0,

|X¯(t)|p\displaystyle|\bar{X}(t)|^{p}\leq |x0|p+p0t|X¯(s)|p2[X¯T(s)B(X(s),YX(s))\displaystyle|x_{0}|^{p}+p\int_{0}^{t}|\bar{X}(s)|^{p-2}\Big{[}\bar{X}^{T}(s)B\Big{(}X^{*}(s),Y^{X^{*}(s)}\Big{)}\
+p12|σ(X(s))|2]ds+p0t|X¯(s)|p2X¯T(s)σ(X(s))dW1(s),\displaystyle~{}~{}~{}+\frac{p-1}{2}|\sigma(X(s))|^{2}\Big{]}\mathrm{d}s+p\int_{0}^{t}|\bar{X}(s)|^{p-2}\bar{X}^{T}(s)\sigma(X(s))\mathrm{d}W^{1}(s),

where we write BM(,)B_{M}(\cdot,\cdot) as B(,)B(\cdot,\cdot) for short. Utilizing the Burkholder-Davis-Gundy inequality [33, p.40, Theorem7.2], the Young inequality and the Ho¨lder\mathrm{H\ddot{o}lder} inequality implies that for any T>0T>0,

𝔼(supt[0,T]|X¯(t)|p)\displaystyle\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)} |x0|p+p𝔼0T|X¯(t)|p2[|X¯T(t)B(X(t),YX(t))|+p12|σ(X(t))|2]dt\displaystyle\leq|x_{0}|^{p}+p\mathbb{E}\int_{0}^{T}|\bar{X}(t)|^{p-2}\Big{[}\Big{|}\bar{X}^{T}(t)B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}+\frac{p-1}{2}|\sigma(X(t))|^{2}\Big{]}\mathrm{d}t
+42p𝔼(0T|X¯(t)|2p2|σ(X(t))|2dt)12\displaystyle~{}~{}~{}~{}~{}~{}+4\sqrt{2}p\mathbb{E}\Big{(}\int_{0}^{T}|\bar{X}(t)|^{2p-2}|\sigma(X(t))|^{2}\mathrm{d}t\Big{)}^{\frac{1}{2}}
|x0|p+C0T𝔼|X¯(t)|pdt+C0T𝔼|X¯T(t)B(X(t),YX(t))|p2dt\displaystyle\leq|x_{0}|^{p}+C\int_{0}^{T}\mathbb{E}|\bar{X}(t)|^{p}\mathrm{d}t+C\int_{0}^{T}\mathbb{E}\Big{|}\bar{X}^{T}(t)B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\mathrm{d}t
+C0T𝔼|σ(X(t))|pdt+42p𝔼[supt[0,T]|X¯(t)|p1(0T|σ(X(t))|2dt)12]\displaystyle~{}~{}~{}~{}~{}~{}+C\int_{0}^{T}\mathbb{E}|\sigma(X(t))|^{p}\mathrm{d}t+4\sqrt{2}p\mathbb{E}\Big{[}\sup_{t\in[0,T]}|\bar{X}(t)|^{p-1}\Big{(}\int_{0}^{T}\big{|}\sigma(X(t))\big{|}^{2}\mathrm{d}t\Big{)}^{\frac{1}{2}}\Big{]}
|x0|p+C0T𝔼|X¯(t)|pdt+C0T𝔼|XT(t)B(X(t),YX(t))|p2dt\displaystyle\leq|x_{0}|^{p}+C\int_{0}^{T}\mathbb{E}|\bar{X}(t)|^{p}\mathrm{d}t+C\int_{0}^{T}\mathbb{E}\Big{|}X^{T}(t)B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\mathrm{d}t
+C0T𝔼|(X¯(t)X(t))TB(X(t),YX(t))|p2dt\displaystyle~{}~{}~{}~{}~{}~{}+C\int_{0}^{T}\mathbb{E}\Big{|}(\bar{X}(t)-X(t))^{T}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\mathrm{d}t
+C0T𝔼|σ(X(t))|pdt+12𝔼(supt[0,T]|X¯(t)|p).\displaystyle~{}~{}~{}~{}~{}~{}+C\int_{0}^{T}\mathbb{E}|\sigma(X(t))|^{p}\mathrm{d}t+\frac{1}{2}\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)}.

Then it follows from (𝐒𝟑)({\bf S3}) that

𝔼(supt[0,T]|X¯(t)|p)\displaystyle\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)} |x0|p+C0T𝔼(sup0st|X¯(s)|p)dt+C0T𝔼|XT(t)B(X(t),YX(t))|p2dt\displaystyle\leq|x_{0}|^{p}+C\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{X}(s)|^{p}\Big{)}\mathrm{d}t+C\int_{0}^{T}\mathbb{E}\Big{|}X^{T}(t)B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\mathrm{d}t
+C0T𝔼|(X¯(t)X(t))TB(X(t),YX(t))|p2dt.\displaystyle~{}~{}+C\int_{0}^{T}\mathbb{E}\Big{|}(\bar{X}(t)-X(t))^{T}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\mathrm{d}t. (5.1)

For any t0t\geq 0, one observes that YX(t)=YXn1(t)=Yn1(t)Xn1(t)Y^{X^{*}(t)}=Y^{X^{*}_{n_{1}(t)}}=Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t)}. Due to the independence of Yn1(t)xY^{x}_{n_{1}(t)} and Xn1(t)X_{n_{1}(t)}, for any t0t\geq 0 and 2pk/θ42\leq p\leq k/\theta_{4}, under (𝐒𝟓)({\bf S5}), (𝐅𝟏)({\bf F1}) and (𝐅𝟑)({\bf F3}), we obtain from (4.5) and the result of Lemma\mathrm{Lemma} 4.8 that

𝔼|XT(t)B(X(t),YX(t))|p2=𝔼|Xn1(t)TB(Xn1(t),YXn1(t))|p2\displaystyle\mathbb{E}\Big{|}X^{T}(t)B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}=\mathbb{E}\Big{|}X^{T}_{n_{1}(t)}B\Big{(}X^{*}_{n_{1}(t)},Y^{X^{*}_{n_{1}(t)}}\Big{)}\Big{|}^{\frac{p}{2}}\
=\displaystyle= 𝔼[𝔼(|Xn1(t)TB(Xn1(t),Yn1(t)Xn1(t))|p2|Xn1(t))]𝔼(𝔼|xTB(x,Yn1(t)x)|p2|x=Xn1(t))\displaystyle\mathbb{E}\Big{[}\mathbb{E}\Big{(}\big{|}X^{T}_{n_{1}(t)}B(X^{*}_{n_{1}(t)},Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t)})\big{|}^{\frac{p}{2}}\big{|}X_{n_{1}(t)}\Big{)}\Big{]}\ \leq\mathbb{E}\big{(}\mathbb{E}\big{|}x^{T}B(x^{*},Y^{x^{*}}_{n_{1}(t)})\big{|}^{\frac{p}{2}}\big{|}_{x=X_{n_{1}(t)}}\big{)}
\displaystyle\leq C(1+𝔼|Xn1(t)|p)C+C𝔼(sup0st|X¯(s)|p).\displaystyle C(1+\mathbb{E}|X_{n_{1}(t)}|^{p})\leq C+C\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{X}(s)|^{p}\Big{)}. (5.2)

Under (𝐒𝟒)({\bf S4}), we derive from (3.2) and (4.5) that

𝔼|B(X(t),YX(t))|p\displaystyle\mathbb{E}\Big{|}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{p} =𝔼|B(Xn1(t),Yn1(t)Xn1(t))|p1Mm=1M𝔼|b(Xn1(t),Yn1(t),mXn1(t))|p\displaystyle=\mathbb{E}\Big{|}B\Big{(}X^{*}_{n_{1}(t)},Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t)}\Big{)}\Big{|}^{p}\leq\frac{1}{M}\sum_{m=1}^{M}\mathbb{E}\Big{|}b\big{(}X^{*}_{n_{1}(t)},Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t),m}\big{)}\Big{|}^{p}\
1Mm=1M𝔼[(CΔ112(1+|Xn1(t)|)+K3|Yn1(t),mXn1(t)|θ4)]p\displaystyle\leq\frac{1}{M}\sum_{m=1}^{M}\mathbb{E}\Big{[}\Big{(}C\Delta_{1}^{-\frac{1}{2}}(1+|X^{*}_{n_{1}(t)}|)+K_{3}\big{|}Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t),m}\big{|}^{\theta_{4}}\Big{)}\Big{]}^{p}\
CΔ1p2𝔼(1+|Xn1(t)|)p+CMm=1M𝔼|Yn1(t),mXn1(t)|pθ4.\displaystyle\leq C\Delta_{1}^{-\frac{p}{2}}\mathbb{E}\big{(}1+|X^{*}_{n_{1}(t)}|\big{)}^{p}+\frac{C}{M}\sum_{m=1}^{M}\mathbb{E}|Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t),m}|^{p\theta_{4}}. (5.3)

Owing to (𝐅𝟏)({\bf F1}) and (𝐅𝟑)({\bf F3}) with pθ4kp\theta_{4}\leq k, using the Young inequality and Lemma 4.3 yields that

𝔼|Yn1(t),mXn1(t)|pθ4\displaystyle\mathbb{E}|Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t),m}|^{p\theta_{4}} =𝔼(𝔼(|Yn1(t),mXn1(t)|pθ4|Xn1(t)))=𝔼(𝔼|Yn1(t),mx|pθ4|x=Xn1(t))\displaystyle=\mathbb{E}\Big{(}\mathbb{E}\big{(}|Y^{X^{*}_{n_{1}(t)}}_{n_{1}(t),m}|^{p\theta_{4}}|X^{*}_{n_{1}(t)}\big{)}\Big{)}=\mathbb{E}\big{(}\mathbb{E}|Y^{x}_{n_{1}(t),m}|^{p\theta_{4}}|_{x=X^{*}_{n_{1}(t)}}\big{)}\
C(1+𝔼|Xn1(t)|pθ4)CΔ1p2𝔼(1+|Xn1(t)|)p.\displaystyle\leq C(1+\mathbb{E}|X^{*}_{n_{1}(t)}|^{p\theta_{4}})\leq C\Delta_{1}^{-\frac{p}{2}}\mathbb{E}\big{(}1+|X^{*}_{n_{1}(t)}|\big{)}^{p}.

Inserting the above inequality into (5) implies that

𝔼|B(X(t),YX(t))|pCΔ1p2𝔼(1+|Xn1(t)|)pCΔ1p2𝔼(1+|X(t)|)p.\displaystyle\mathbb{E}\Big{|}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{p}\leq C\Delta_{1}^{-\frac{p}{2}}\mathbb{E}\big{(}1+|X^{*}_{n_{1}(t)}|\big{)}^{p}\leq C\Delta_{1}^{-\frac{p}{2}}\mathbb{E}\big{(}1+|X(t)|\big{)}^{p}. (5.4)

This together with (3.7) implies that for any t0t\geq 0,

𝔼|X¯(t)X(t)|p\displaystyle\mathbb{E}|\bar{X}(t)-X(t)|^{p}\leq 2p1(𝔼|n1(t)Δ1tB(X(s),YX(s))ds|p+𝔼|n1(t)Δ1tσ(X(s))dW1(s)|p)\displaystyle 2^{p-1}\Big{(}\mathbb{E}\Big{|}\int_{n_{1}(t)\Delta_{1}}^{t}B\Big{(}X^{*}(s),Y^{X^{*}(s)}\Big{)}\mathrm{d}s\Big{|}^{p}+\mathbb{E}\Big{|}\int_{n_{1}(t)\Delta_{1}}^{t}\sigma(X(s))\mathrm{d}W^{1}(s)\Big{|}^{p}\Big{)}\
\displaystyle\leq C(Δ1p1n1(t)Δ1t𝔼|B(X(s),YX(s))|pds+Δ1p22n1(t)Δ1t𝔼|σ(X(s))|pds)\displaystyle C\Big{(}\Delta_{1}^{p-1}\int_{n_{1}(t)\Delta_{1}}^{t}\mathbb{E}\Big{|}B\Big{(}X^{*}(s),Y^{X^{*}(s)}\Big{)}\Big{|}^{p}\mathrm{d}s+\Delta_{1}^{\frac{p-2}{2}}\int_{n_{1}(t)\Delta_{1}}^{t}\mathbb{E}|\sigma(X(s))|^{p}\mathrm{d}s\Big{)}\
\displaystyle\leq CΔ1p2𝔼(1+|X(t)|)p.\displaystyle C\Delta_{1}^{\frac{p}{2}}\mathbb{E}\big{(}1+|X(t)|\big{)}^{p}. (5.5)

Invoking the Ho¨lder\mathrm{H\ddot{o}lder} inequality, (5.4) and (5) we obtain

𝔼(|X¯(t)X(t)|p2|B(X(t),YX(t))|p2)\displaystyle\mathbb{E}\Big{(}\big{|}\bar{X}(t)-X(t)\big{|}^{\frac{p}{2}}\Big{|}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{\frac{p}{2}}\Big{)}\leq (𝔼|X¯(t)X(t)|p)12(𝔼|B(X(t),YX(t))|p)12\displaystyle\Big{(}\mathbb{E}\big{|}\bar{X}(t)-X(t)\big{|}^{p}\Big{)}^{\frac{1}{2}}\Big{(}\mathbb{E}\Big{|}B\Big{(}X^{*}(t),Y^{X^{*}(t)}\Big{)}\Big{|}^{p}\Big{)}^{\frac{1}{2}}\
\displaystyle\leq C𝔼(1+|X(t)|)pC+C𝔼(sup0st|X¯(s)|p).\displaystyle C\mathbb{E}\big{(}1+|X(t)|\big{)}^{p}\leq C+C\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{X}(s)|^{p}\Big{)}. (5.6)

Inserting (5) and (5) into (5) yields that

𝔼(supt[0,T]|X¯(t)|p)C+C0T𝔼(sup0st|X¯(s)|p)dt.\displaystyle\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)}\leq C+C\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{X}(s)|^{p}\Big{)}\mathrm{d}t.

A direct application of Gronwall’s inequality derives that

𝔼(supt[0,T]|X¯(t)|p)CT.\displaystyle\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)}\leq C_{T}. (5.7)

Then the second assertion holds directly by substituting (5.7) into (5). The case 0<p<20<p<2 follows from that directly by using the Hölder inequality. The proof is complete. ∎

Remark 5.1.

For any R>|x0|R>|x_{0}|, define the stopping time

ρ¯Δ1,R=inf{t0:|X¯(t)|R}.\displaystyle\bar{\rho}_{\Delta_{1},R}=\inf\{t\geq 0:|\bar{X}(t)|\geq R\}. (5.8)

It follows from Lemma 5.1 that for any T>0T>0,

(ρ¯Δ1,RT)CTRp.\displaystyle\mathbb{P}\big{(}\bar{\rho}_{\Delta_{1},R}\leq T\big{)}\leq\frac{C_{T}}{R^{p}}.\

To prove the strong convergence of the MTEM scheme (3.5), we introduce the auxiliary TEM numerical scheme for averaged equation (1.4)

{Z0=x0,Zn+1=Zn+b¯(Zn)Δ1+σ(Zn)ΔWn1,\begin{cases}Z_{0}=x_{0},\\ Z_{n+1}=Z_{n}+\bar{b}(Z^{*}_{n})\Delta_{1}+\sigma(Z_{n})\Delta W^{1}_{n},\end{cases} (5.9)

and the corresponding continuous-time processes

Z(t)=Zn,t[nΔ1,(n+1)Δ1),Z(t)=Z_{n},~{}~{}~{}~{}~{}~{}~{}t\in[n\Delta_{1},(n+1)\Delta_{1}),\\

and

Z¯(t)=x0+0tb¯(Z(s))ds+0tσ(Z(s))dW1(s).\displaystyle\bar{Z}(t)=x_{0}+\int_{0}^{t}\bar{b}(Z^{*}(s))\mathrm{d}s+\int_{0}^{t}\sigma(Z(s))\mathrm{d}W^{1}(s). (5.10)

One observes that Z¯(nΔ1)=Z(nΔ1)=Zn\bar{Z}(n\Delta_{1})={Z}(n\Delta_{1})=Z_{n}. In what follows, we analyze the strong error between x¯(t)\bar{x}(t) and Z¯(t)\bar{Z}(t) and the strong error between Z¯(t)\bar{Z}(t) and X(t)X(t), respectively. To proceed we begin with the pth moment boundedness of Z¯(t)\bar{Z}(t).

Lemma 5.2.

If (𝐒𝟑)({\bf S3})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with kθ4k\geq\theta_{4}, then for any p>0p>0 and T>0T>0,

supΔ1(0,1]𝔼(sup0tT|Z¯(t)|p)CT,\displaystyle\sup_{\Delta_{1}\in(0,1]}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}\leq C_{T},

and

sup0tT𝔼|Z¯(t)Z(t)|2CTΔ1.\displaystyle\sup_{0\leq t\leq T}\mathbb{E}|\bar{Z}(t)-Z(t)|^{2}\leq C_{T}\Delta_{1}.
Proof.

The case that 0<p<20<p<2 follows directly from the case p2p\geq 2 by using Lyapunov’s inequality. Thus we are only going to deal with the case p2p\geq 2. Applying the Ito^\mathrm{It\hat{o}} formula and Burkholder-Davis-Gundy inequality [33, p.40, Theorem7.2], under (𝐒𝟓)({\bf S5}) and (F1)-(F3), we derive from the result of Lemma 4.2 that for p2p\geq 2 and T>0T>0,

𝔼(sup0tT|Z¯(t)|p)\displaystyle\mathbb{E}\big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\big{)} |x0|p+pC2𝔼0T|Z¯(t)|p2(1+|Z(t)|2)dt\displaystyle\leq|x_{0}|^{p}+\frac{pC}{2}\mathbb{E}\int_{0}^{T}\big{|}\bar{Z}(t)\big{|}^{p-2}\big{(}1+|Z(t)|^{2}\big{)}\mathrm{d}t\
+p𝔼0T|Z¯(t)|p2|Z¯(t)Z(t)||b¯(Z(t)|dt\displaystyle~{}~{}~{}+p\mathbb{E}\int_{0}^{T}|\bar{Z}(t)|^{p-2}|\bar{Z}(t)-Z(t)||\bar{b}(Z^{*}(t)|\mathrm{d}t\
+42p𝔼(0T|Z¯(t)|2p2|σ(Z(t))|2dt)12.\displaystyle~{}~{}~{}+4\sqrt{2}p\mathbb{E}\Big{(}\int_{0}^{T}\big{|}\bar{Z}(t)\big{|}^{2p-2}\big{|}\sigma(Z(t))|^{2}\mathrm{d}t\Big{)}^{\frac{1}{2}}.\

Then by the Young inequality we obtain that for any T>0T>0,

𝔼(sup0tT|Z¯(t)|p)\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)} |x0|p+C0T𝔼(sup0st|Z¯(s)|p)dt+C0T𝔼(|Z¯(t)Z(t)|p2|b¯(Z(t))|p2)dt\displaystyle\leq|x_{0}|^{p}+C\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{Z}(s)|^{p}\Big{)}\mathrm{d}t+C\int_{0}^{T}\mathbb{E}\Big{(}|\bar{Z}(t)-Z(t)|^{\frac{p}{2}}|\bar{b}(Z^{*}(t))|^{\frac{p}{2}}\Big{)}\mathrm{d}t\
+12𝔼(sup0tT|Z¯(t)|p)+C𝔼(0t|σ(Z(t))|2dt)p2.\displaystyle~{}~{}~{}+\frac{1}{2}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}+C\mathbb{E}\Big{(}\int_{0}^{t}|\sigma(Z(t))|^{2}\mathrm{d}t\Big{)}^{\frac{p}{2}}. (5.11)

For any t0t\geq 0, due to (S4), (F1)-(F3) with kθ4k\geq\theta_{4}, (3) hold. Then using (3) and (S3) yields that

𝔼|Z¯(t)Z(t)|p=𝔼|Z¯(t)Zn1(t)|p\displaystyle\mathbb{E}|\bar{Z}(t)-Z(t)|^{p}=\mathbb{E}|\bar{Z}(t)-Z_{n_{1}(t)}|^{p}\
\displaystyle\leq 2p1(𝔼|n1(t)Δ1tb¯(Z(s))ds|p+𝔼|n1(t)Δ1tσ(Z(s))dW1(s)|p)\displaystyle 2^{p-1}\Big{(}\mathbb{E}\Big{|}\int_{n_{1}(t)\Delta_{1}}^{t}\bar{b}(Z^{*}(s))\mathrm{d}s\Big{|}^{p}+\mathbb{E}\Big{|}\int_{n_{1}(t)\Delta_{1}}^{t}\sigma(Z(s))\mathrm{d}W^{1}(s)\Big{|}^{p}\Big{)}\
\displaystyle\leq 2p1(Δ1p𝔼|b¯(Zn1(t))|p+Δ1p2𝔼|σ(Zn1(t))|p)\displaystyle 2^{p-1}\Big{(}\Delta_{1}^{p}\mathbb{E}|\bar{b}(Z^{*}_{n_{1}(t)})|^{p}+\Delta_{1}^{\frac{p}{2}}\mathbb{E}|\sigma(Z_{n_{1}(t)})|^{p}\Big{)}\
\displaystyle\leq CΔ1p2(1+𝔼|Zn1(t)|p)CΔ1p2(1+𝔼|Z(t)|p).\displaystyle C\Delta_{1}^{\frac{p}{2}}\big{(}1+\mathbb{E}|Z_{n_{1}(t)}|^{p}\big{)}\leq C\Delta_{1}^{\frac{p}{2}}(1+\mathbb{E}|Z(t)|^{p}). (5.12)

Then utilizing (3) again and the Ho¨lder\mathrm{H\ddot{o}lder} inequality implies that

𝔼(|Z¯(t)Z(t)|p2|b¯(Z(t))|p2)\displaystyle\mathbb{E}\Big{(}|\bar{Z}(t)-Z(t)|^{\frac{p}{2}}|\bar{b}(Z^{*}(t))|^{\frac{p}{2}}\Big{)} (𝔼|Z¯(t)Z(t)|p)12(𝔼|b¯(Z(t))|p)12\displaystyle\leq\big{(}\mathbb{E}|\bar{Z}(t)-Z(t)|^{p}\big{)}^{\frac{1}{2}}\big{(}\mathbb{E}|\bar{b}(Z(t))|^{p}\big{)}^{\frac{1}{2}}\
CΔ1p4(1+𝔼|Z(t)|p)12Δ1p4(1+𝔼|Z(t)|p)12\displaystyle\leq C\Delta_{1}^{\frac{p}{4}}\big{(}1+\mathbb{E}|Z(t)|^{p}\big{)}^{\frac{1}{2}}\Delta_{1}^{-\frac{p}{4}}\big{(}1+\mathbb{E}|Z(t)|^{p}\big{)}^{\frac{1}{2}}\
C+𝔼(sup0st|Z¯(s)|p).\displaystyle\leq C+\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{Z}(s)|^{p}\Big{)}. (5.13)

Applying (𝐒𝟑)({\bf S3}) and the Hölder inequality we get

𝔼(0T|σ(Z(t))|2dt)p2C+C0T𝔼(sup0st|Z¯(s)|p)dt.\displaystyle\mathbb{E}\Big{(}\int_{0}^{T}\big{|}\sigma(Z(t))|^{2}\mathrm{d}t\Big{)}^{\frac{p}{2}}\leq C+C\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{Z}(s)|^{p}\Big{)}\mathrm{d}t. (5.14)

Hence, substituting (5) and (5.14) into (5) yields that

𝔼(sup0tT|Z¯(t)|p)C+C0T𝔼(sup0st|Z¯(s)|p)dt.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}\leq C+C\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq s\leq t}|\bar{Z}(s)|^{p}\Big{)}\mathrm{d}t.\

An application of the Gronwall\mathrm{Gronwall} inequality gives that

𝔼(sup0tT|Z¯(t)|p)CT.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}\leq C_{T}.

Then inserting the above inequality into (5) implies that the another desired assertion holds. The proof is complete. ∎

Remark 5.2.

From Lemma 5.2, for any constant R>|x0|R>|x_{0}|, define a stopping time

ρΔ1,R:=inf{t0:|Z¯(t)|R}.\displaystyle\rho_{\Delta_{1},R}:=\inf\{t\geq 0:|\bar{Z}(t)|\geq R\}. (5.15)

By a similar argument as Remark 2.2, for any T>0T>0, we have

(ρΔ1,RT)CT/Rp.\mathbb{P}(\rho_{\Delta_{1},R}\leq T)\leq C_{T}/{R^{p}}.
Lemma 5.3.

If (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k>θ12θ2θ4k>\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, then for any T>0T>0,

limΔ10𝔼(sup0tT|x¯(t)Z¯(t)|2)=0.\displaystyle\lim_{\Delta_{1}\rightarrow 0}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-\bar{Z}(t)|^{2}\Big{)}=0.
Proof.

Fix any constant R>|x0|R>|x_{0}|. Define the truncated functions

b¯R(x)=b¯((|x|R)x|x|),σR(x)=σ((|x|R)x|x|).\bar{b}_{R}(x)=\bar{b}\Big{(}(|x|\wedge R)\frac{x}{|x|}\Big{)},~{}~{}~{}\sigma_{R}(x)=\sigma\Big{(}(|x|\wedge R)\frac{x}{|x|}\Big{)}.

Consider the SDE

du¯(t)=b¯R(u¯(t))dt+σR(u¯(t))dW1(t)\displaystyle\mathrm{d}\bar{u}(t)=\bar{b}_{R}(\bar{u}(t))\mathrm{d}t+\sigma_{R}(\bar{u}(t))\mathrm{d}W^{1}(t) (5.16)

with initial value u¯(0)=x0\bar{u}(0)=x_{0}. Under (𝐒𝟏)({\bf S1}), (𝐒𝟐)({\bf S2}) and (F1)-(F3) with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, one observes from (𝐒𝟏)({\bf S1}) and the result of Lemma 4.1 that both b¯R(x)\bar{b}_{R}(x) and σR(x)\sigma_{R}(x) are global Lipschitz continuous. Thus equation (5.16) has a unique global solution u¯(t)\bar{u}(t) on t0t\geq 0. Let U¯(t)\bar{U}(t) denote the continuous extension of the EM numerical solution of (5.16). It is well known [18, 24] that

𝔼(supt[0,T]|u¯(t)U¯(t)|2)CTΔ1,T>0.\displaystyle\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{u}(t)-\bar{U}(t)|^{2}\Big{)}\leq C_{T}\Delta_{1},~{}~{}~{}\forall~{}T>0.\

On the other hand, choose a constant Δ¯1(0,1]\bar{\Delta}_{1}\in(0,1] small sufficiently such that
φ1(K(Δ¯1)1/2)R\varphi^{-1}(K(\bar{\Delta}_{1})^{-1/2})\geq R. One observes that for any Δ1(0,Δ¯1]\Delta_{1}\in(0,\bar{\Delta}_{1}]

b¯R(x)=b¯(x)=b¯(x),xn1with|x|R.\displaystyle\bar{b}_{R}(x)=\bar{b}(x)=\bar{b}(x^{*}),~{}~{}~{}~{}\forall~{}x\in\mathbb{R}^{n_{1}}~{}\hbox{with}~{}|x|\leq R.\

Then it is straightforward to see that that for any t0t\geq 0

x¯(tτR)=u¯(tτR),Z¯(tρΔ1,R)=U¯(tρΔ1,R),a.s.,\displaystyle\bar{x}(t\wedge\tau_{R})=\bar{u}(t\wedge\tau_{R}),~{}~{}~{}~{}\bar{Z}(t\wedge\rho_{\Delta_{1},R})=\bar{U}(t\wedge\rho_{\Delta_{1},R}),~{}~{}\mathrm{a.s.},\

where τR\tau_{R} and ρΔ1,R\rho_{\Delta_{1},R} are defined in Remarks 2.2 and 5.2, respectively. Under (S1)-(S5) and (F1)-(F3) with kθ12θ2θ4k\geq\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, by virtue of Lemmas 2.3 and 5.2, the remainder of the proof follows in a similar manner to that of [34, Theorem 3.5]. To avoid duplication we omit the details. ∎

Then we turn to prove the strong convergence of the MTEM numerical solution X(t)X(t) and auxiliary process Z¯(t)\bar{Z}(t). By virtue of Lemma 5.1, we only need to prove strong convergence of X¯(t)\bar{X}(t) and Z¯(t)\bar{Z}(t).

Lemma 5.4.

If (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k>θ12θ22θ4(θ2+θ4+1)k>\theta_{1}\vee 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), for any T>0T>0 and Δ1(0,1]\Delta_{1}\in(0,1],

limΔ20limMΔ2𝔼(sup0tT|X¯(t)Z¯(t)|2)=0.\displaystyle\lim_{\Delta_{2}\rightarrow 0}\lim_{M\Delta_{2}\rightarrow\infty}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{X}(t)-\bar{Z}(t)|^{2}\Big{)}=0.
Proof.

Define e¯(t)=X¯(t)Z¯(t)\bar{e}(t)=\bar{X}(t)-\bar{Z}(t) for any t0t\geq 0 and βΔ1,R=ρ¯Δ1,RρΔ1,R\beta_{\Delta_{1},R}=\bar{\rho}_{\Delta_{1},R}\wedge\rho_{\Delta_{1},R} for any R>0R>0, where ρ¯Δ1,R\bar{\rho}_{\Delta_{1},R} and ρΔ1,R\rho_{\Delta_{1},R} are given by (5.8) and (5.15), respectively. Due to k>2θ4k>2\theta_{4}, let 2<pk/θ42<p\leq k/\theta_{4}. Fix T>0T>0. For any δ>0\delta>0, using the Young\mathrm{Young} inequality yields that

𝔼(sup0tT|e¯(t)|2)=𝔼(sup0tT|e¯(t)|2I{βΔ1,R>T})+𝔼(sup0tT|e¯(t)|2I{βΔ1,RT})\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}\Big{)}=\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}I_{\{\beta_{\Delta_{1},R}>T\}}\Big{)}+\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}I_{\{\beta_{\Delta_{1},R}\leq T\}}\Big{)}\
\displaystyle\leq 𝔼(sup0tT|e¯(t)|2I{βΔ1,R>T})+2δp𝔼(sup0tT|e¯(t)|p)+p2pδ2p2(βΔ1,RT).\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}I_{\{\beta_{\Delta_{1},R}>T\}}\Big{)}+\frac{2\delta}{p}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{p}\Big{)}+\frac{p-2}{p\delta^{\frac{2}{p-2}}}\mathbb{P}(\beta_{\Delta_{1},R}\leq T).

Owing to (S3)-(S5) and (F1)-(F3), it follows from the results of Lemmas 5.1 and 5.2 that

𝔼(sup0tT|e¯(t)|p)2p1𝔼(sup0tT|Z¯(t)|p)+2p1𝔼(sup0tT|X¯(t)|p)CT.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{p}\Big{)}\leq 2^{p-1}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}+2^{p-1}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{X}(t)|^{p}\Big{)}\leq C_{T}.\

Furthermore, both Remarks\mathrm{Remarks} 5.2 and 5.1 imply that

(βΔ1,RT)(ρΔ1,RT)+(ρ¯Δ1,RT)CTRp.\displaystyle\mathbb{P}(\beta_{\Delta_{1},R}\leq T)\leq\mathbb{P}(\rho_{\Delta_{1},R}\leq T)+\mathbb{P}(\bar{\rho}_{\Delta_{1},R}\leq T)\leq\frac{C_{T}}{R^{p}}.

Consequently we have

𝔼(sup0tT|e¯(t)|2)𝔼(sup0tT|e¯(t)|2I{βΔ1,R>T})+CTδp+CT(p2)pδ2p2Rp.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}\Big{)}\leq\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t)|^{2}I_{\{\beta_{\Delta_{1},R}>T\}}\Big{)}+\frac{C_{T}\delta}{p}+\frac{C_{T}(p-2)}{p\delta^{\frac{2}{p-2}}R^{p}}.

Now, for any ϵ>0\epsilon>0, choose δ>0\delta>0 small sufficiently such that CTδ/pϵ/3{C_{T}\delta}/{p}\leq{\epsilon}/{3}. Then for this δ\delta, choose R>0R>0 large enough such that CT(p2)/(pδ2p2Rp)ϵ/3.C_{T}(p-2)/(p\delta^{\frac{2}{p-2}}R^{p})\leq{\epsilon}/{3}. Hence, for the desired assertion it is sufficient to prove

𝔼(sup0tT|e(t)|2I{βΔ1,R>T})ϵ3.\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\mathrm{e}(t)|^{2}I_{\{\beta_{\Delta_{1},R}>T\}}\Big{)}\leq\frac{\epsilon}{3}. (5.17)

From (3.7) and (5.10) we derive that

e¯(tβΔ1,R)=\displaystyle\bar{e}(t\wedge\beta_{\Delta_{1},R})= 0tβΔ1,R(BM(X(s),YX(s))b¯(Z(s)))ds\displaystyle\int_{0}^{t\wedge\beta_{\Delta_{1},R}}\Big{(}B_{M}\Big{(}X^{*}(s),Y^{X^{*}(s)}\Big{)}-\bar{b}(Z^{*}(s))\Big{)}\mathrm{d}s\
+0tβΔ1,R(σ(X(s))σ(Z(s)))dW1(s).\displaystyle~{}~{}~{}+\int_{0}^{t\wedge\beta_{\Delta_{1},R}}\big{(}\sigma(X(s))-\sigma(Z(s))\big{)}\mathrm{d}W^{1}(s).\

Recalling the definition of the stopping time βΔ1,R\beta_{\Delta_{1},R}, it is straightforward to see that

X(s)=X(s),Z(s)=Z(s),s[0,tβΔ1,R].\displaystyle X^{*}(s)=X(s),~{}~{}~{}Z^{*}(s)=Z(s),~{}~{}~{}\forall s\in[0,t\wedge\beta_{\Delta_{1},R}].

Then we have

e¯(tβΔ1,R)\displaystyle\bar{e}(t\wedge\beta_{\Delta_{1},R}) =0tβΔ1,R(BM(X(s),YX(s))b¯(Z(s)))ds\displaystyle=\int_{0}^{t\wedge\beta_{\Delta_{1},R}}\Big{(}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(Z(s))\Big{)}\mathrm{d}s\
+0tβΔ1,R(σ(X(s))σ(Z(s)))dW1(s).\displaystyle~{}~{}~{}+\int_{0}^{t\wedge\beta_{\Delta_{1},R}}\big{(}\sigma(X(s))-\sigma(Z(s))\big{)}\mathrm{d}W^{1}(s).

Using the Ho¨lder\mathrm{H\ddot{o}lder} inequality, the Burkholder-Davis-Gundy inequality [33, p.40, Theorem 7.2] and the elementary inequality, we arrive at

𝔼(sup0tT|e¯(tβΔ1,R)|2)\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t\wedge\beta_{\Delta_{1},R})|^{2}\Big{)}\leq 2T0T𝔼(|BM(X(s),YX(s))b¯(Z(s))|2I{sβΔ1,R})ds\displaystyle 2T\int_{0}^{T}\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(Z(s))\Big{|}^{2}I_{\{s\leq\beta_{\Delta_{1},R}\}}\Big{)}\mathrm{d}s\
+80T𝔼|σ(X(sβΔ1,R))σ(Z(sβΔ1,R))|2ds\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+8\int_{0}^{T}\mathbb{E}|\sigma(X(s\wedge\beta_{\Delta_{1},R}))-\sigma(Z(s\wedge\beta_{\Delta_{1},R}))|^{2}\mathrm{d}s\
\displaystyle\leq 4T0T𝔼(|BM(X(s),YX(s))b¯(X(s))|2I{sβΔ1,R})ds\displaystyle 4T\int_{0}^{T}\!\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(X(s))\Big{|}^{2}I_{\{s\leq\beta_{\Delta_{1},R}\}}\Big{)}\mathrm{d}s\
+4T0T𝔼|b¯(X(sβΔ1,R))b¯(Z(sβΔ1,R))|2ds\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+4T\!\!\int_{0}^{T}\!\mathbb{E}|\bar{b}(X(s\wedge\beta_{\Delta_{1},R}))-\bar{b}(Z(s\wedge\beta_{\Delta_{1},R}))|^{2}\mathrm{d}s\
+80T𝔼|σ(X(sβΔ1,R))σ(Z(sβΔ1,R))|2ds.\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+8\int_{0}^{T}\mathbb{E}|\sigma(X(s\wedge\beta_{\Delta_{1},R}))-\sigma(Z(s\wedge\beta_{\Delta_{1},R}))|^{2}\mathrm{d}s. (5.18)

For any 0sT0\leq s\leq T, one observes that for any ω{ωΩ:sβΔ1,R}\omega\in\{\omega\in\Omega:s\leq\beta_{\Delta_{1},R}\}, |X(s)|R|X(s)|\leq R. Using this fact and (4.5) implies that

𝔼(|BM(X(s),YX(s))b¯(X(s))|2I{sβΔ1,R})\displaystyle\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(X(s))\Big{|}^{2}I_{\{s\leq\beta_{\Delta_{1},R}\}}\Big{)}\
\displaystyle\leq 𝔼(|BM(X(s),YX(s))b¯(X(s))|2I{|X(s)|R})\displaystyle\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(X(s))\Big{|}^{2}I_{\{|X(s)|\leq R\}}\Big{)}\
=\displaystyle= 𝔼(|BM(Xn1(s),Yn1(s)Xn1(s))b¯(Xn1(s))|2I{|Xn1(s)|R})\displaystyle\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X_{n_{1}(s)},Y^{X_{n_{1}(s)}}_{n_{1}(s)}\Big{)}-\bar{b}(X_{n_{1}(s)})\Big{|}^{2}I_{\{|X_{n_{1}(s)}|\leq R\}}\Big{)}\
=\displaystyle= 𝔼(𝔼[(|BM(Xn1(s),Yn1(s)Xn1(s))b¯(Xn1(s))|2I{|Xn1(s)|R})|Xn1(s)])\displaystyle\mathbb{E}\Bigg{(}\mathbb{E}\Big{[}\Big{(}\Big{|}B_{M}\Big{(}X_{n_{1}(s)},Y^{X_{n_{1}(s)}}_{n_{1}(s)}\Big{)}-\bar{b}(X_{n_{1}(s)})\Big{|}^{2}I_{\{|X_{n_{1}(s)}|\leq R\}}\Big{)}\Big{|}X_{n_{1}(s)}\Big{]}\Bigg{)}\
=\displaystyle= 𝔼(𝔼|BM(x,Yn1(s)x)b¯(x)|x=Xn1(s)2I{|Xn1(s)|R}).\displaystyle\mathbb{E}\Big{(}\mathbb{E}\Big{|}B_{M}\Big{(}x,Y^{x}_{n_{1}(s)}\Big{)}-\bar{b}(x)\Big{|}^{2}_{x=X_{n_{1}(s)}}I_{\{|X_{n_{1}(s)}|\leq R\}}\Big{)}.\

By (S2), (S4) and (F1)-(F3) with k2θ22θ4(θ2+θ4+1)k\geq 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), it follows from the result of Lemma 4.12 that

𝔼(|BM(X(s),YX(s))b¯(X(s))|2I{sβΔ1,R})\displaystyle\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X(s),Y^{X(s)}\Big{)}-\bar{b}(X(s))\Big{|}^{2}I_{\{s\leq\beta_{\Delta_{1},R}\}}\Big{)}\
=\displaystyle= C𝔼[(1+|Xn1(s)|2(θ2+1)2θ32θ4)I{|Xn1(s)|R}](Δ2+1M+1MΔ2)\displaystyle C\mathbb{E}\Big{[}\big{(}1+|X_{n_{1}(s)}|^{2(\theta_{2}+1)\vee 2\theta_{3}\vee 2\theta_{4}}\big{)}I_{\{|X_{n_{1}(s)}|\leq R\}}\Big{]}\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}\
\displaystyle\leq CR(Δ2+1M+1MΔ2).\displaystyle C_{R}\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}. (5.19)

Under (𝐒𝟏)({\bf S1}), (𝐒𝟐)({\bf S2}), (𝐒𝟒)({\bf S4}) and (F1)-(F3) with kθ12θ2θ4k\geq\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, applying Lemma 4.1 yields that

𝔼|b¯(X(sβΔ1,R))b¯(Z(sβΔ1,))|2𝔼|σ(X(sβΔ1,R))σ(Z(sβΔ1,R))|2\displaystyle\mathbb{E}\big{|}\bar{b}(X(s\wedge\beta_{\Delta_{1},R}))-\bar{b}(Z(s\wedge\beta_{\Delta_{1},}))\big{|}^{2}\vee\mathbb{E}|\sigma(X(s\wedge\beta_{\Delta_{1},R}))-\sigma(Z(s\wedge\beta_{\Delta_{1},R}))|^{2}\
\displaystyle\leq (L¯R2LR2)𝔼|e¯(sβΔ1,R)|2(L¯R2LR2)𝔼(sup0rs|e¯(sβΔ1,R)|2).\displaystyle(\bar{L}_{R}^{2}\vee L^{2}_{R})\mathbb{E}|\bar{e}(s\wedge\beta_{\Delta_{1},R})|^{2}\leq(\bar{L}_{R}^{2}\vee L^{2}_{R})\mathbb{E}\Big{(}\sup_{0\leq r\leq s}|\bar{e}(s\wedge\beta_{\Delta_{1},R})|^{2}\Big{)}. (5.20)

Inserting (5) and (5) into (5) we derive that

𝔼(sup0tT|e¯(tβΔ1,R)|2)\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t\wedge\beta_{\Delta_{1},R})|^{2}\Big{)} 4T2CR(Δ2+1M+1MΔ2)\displaystyle\leq 4T^{2}C_{R}\Big{(}\Delta_{2}\!+\!\frac{1}{M}\!+\!\frac{1}{M\Delta_{2}}\Big{)}\
+(8+4T)(L¯R2LR2)0T𝔼(sup0rs|e¯(sβΔ1,R)|2)ds.\displaystyle~{}~{}~{}+(8+4T)(\bar{L}_{R}^{2}\vee L^{2}_{R})\int_{0}^{T}\mathbb{E}\Big{(}\sup_{0\leq r\leq s}|\bar{e}(s\wedge\beta_{\Delta_{1},R})|^{2}\Big{)}\mathrm{d}s.

An application of the Gronwall\mathrm{Gronwall} inequality implies that

𝔼(sup0tT|e¯(tβΔ1,R)|2)CR(Δ2+1M+1MΔ2).\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t\wedge\beta_{\Delta_{1},R})|^{2}\Big{)}\leq C_{R}\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.

For the given RR, choose Δ2(0,Δ^2]\Delta_{2}\in(0,\hat{\Delta}_{2}] small sufficiently such that CRΔ2ϵ/9C_{R}\Delta_{2}\leq\epsilon/9. For the fixed Δ2\Delta_{2}, choose MM large sufficiently such that CR/(MΔ2)ϵ/9{C_{R}}/(M\Delta_{2})\leq\epsilon/9. Therefore, we have

𝔼(sup0tT|e¯(tβΔ1,R)|2)ϵ9+2CRMΔ2ϵ3,\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{e}(t\wedge\beta_{\Delta_{1},R})|^{2}\Big{)}\leq\frac{\epsilon}{9}+\frac{2C_{R}}{M\Delta_{2}}\leq\frac{\epsilon}{3},

which implies that the required assertion (5.17) holds. The proof is complete. ∎

Obviously, combing the second result of Lemma 5.1, Lemmas 5.3 and 5.4 derives the strong convergence between x¯(t)\bar{x}(t) and X(t)X(t).

Lemma 5.5.

Under (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k>θ12θ22θ4(θ2+θ4+1)k>\theta_{1}\vee 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), for any T>0T>0,

limΔ1,Δ20limMΔ2𝔼(sup0tT|X(t)x¯(t)|2)=0.\displaystyle\lim_{\Delta_{1},\Delta_{2}\rightarrow 0}\lim_{M\Delta_{2}\rightarrow\infty}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|X(t)-\bar{x}(t)|^{2}\Big{)}=0.
Theorem 5.1.

If (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k>θ12θ22θ4(θ2+θ4+1)k>\theta_{1}\vee 2\theta_{2}\vee 2\theta_{4}\vee(\theta_{2}+\theta_{4}+1), then for any 0<p<k/θ40<p<k/\theta_{4} and T>0T>0

limΔ1,Δ20limMΔ2𝔼(sup0tT|x¯(t)X(t)|p)=0.\displaystyle\lim_{\Delta_{1},\Delta_{2}\rightarrow 0}\lim_{M\Delta_{2}\rightarrow\infty}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-{X}(t)|^{p}\Big{)}=0. (5.21)
Proof.

For any T>0T>0, combining Lemmas 5.3 and 5.5 implies that the desired assertion holds for p=2p=2. Obviously, (5.21) holds for 0<p<20<p<2 due to the Ho¨lder\mathrm{H\ddot{o}lder} inequality. Next, we consider the case 2<p<k/θ42<p<k/\theta_{4}. Choose a constant q¯\bar{q} such that p<q¯<k/θ4p<\bar{q}<k/\theta_{4}. Utilizing the Ho¨lder\mathrm{H\ddot{o}lder} inequality, Lemmas\mathrm{Lemmas} 2.3 and 5.1 we derive that

𝔼(sup0tT|x¯(t)X(t)|p)\displaystyle\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-{X}(t)|^{p}\Big{)} =𝔼(sup0tT|x¯(t)X(t)|2(q¯p)q¯2|x¯(t)X(t)|p2(q¯p)q¯2)\displaystyle=\mathbb{E}\Big{(}\sup_{0\leq t\leq T}\big{|}\bar{x}(t)-{X}(t)\big{|}^{\frac{2(\bar{q}-p)}{\bar{q}-2}}\big{|}\bar{x}(t)-{X}(t)\big{|}^{p-\frac{2(\bar{q}-p)}{\bar{q}-2}}\Big{)}\
[𝔼(sup0tT|x¯(t)X(t)|2)]q¯pq¯2[𝔼(sup0tT|x¯(t)X(t)|q¯)]p2q¯2\displaystyle\leq\Big{[}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}\big{|}\bar{x}(t)-{X}(t)\big{|}^{2}\Big{)}\Big{]}^{\frac{\bar{q}-p}{\bar{q}-2}}\Big{[}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-{X}(t)|^{\bar{q}}\Big{)}\Big{]}^{\frac{p-2}{\bar{q}-2}}\
CT[𝔼(sup0tT|x¯(t)X(t)|2)]q¯pq¯2.\displaystyle\leq C_{T}\Big{[}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-{X}(t)|^{2}\Big{)}\Big{]}^{\frac{\bar{q}-p}{\bar{q}-2}}.

This, together with the case of p=2p=2, implies the required assertion. The proof is complete. ∎

Theorem 5.2.

If (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k>4θ12(θ2+1)2θ32θ4k>4\theta_{1}\vee 2(\theta_{2}+1)\vee 2\theta_{3}\vee 2\theta_{4}, then for any 0<p<k/θ40<p<k/\theta_{4} and T>0T>0,

limε0limΔ1,Δ20limMΔ2𝔼(sup0tT|xε(t)X(t)|p)=0.\displaystyle\lim_{\varepsilon\rightarrow 0}\lim_{\Delta_{1},\Delta_{2}\rightarrow 0}\lim_{M\Delta_{2}\rightarrow\infty}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|x^{\varepsilon}(t)-X(t)|^{p}\Big{)}=0.\
Proof.

For any 0<p<k/θ40<p<k/\theta_{4}, using the elementary inequality, by virtue of Lemmas 2.2 and Theorem 5.1, yields that

limε0limΔ1,Δ20limMΔ2𝔼(sup0tT|xε(t)X(t)|p)\displaystyle\lim_{\varepsilon\rightarrow 0}\lim_{\Delta_{1},\Delta_{2}\rightarrow 0}\lim_{M\Delta_{2}\rightarrow\infty}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|x^{\varepsilon}(t)-X(t)|^{p}\Big{)}\
\displaystyle\leq 2plimε0𝔼|xε(t)x¯(t)|p+2plimMΔ2limΔ1,Δ20𝔼(sup0tT|x¯(t)X(t)|p)=0.\displaystyle 2^{p}\lim_{\varepsilon\rightarrow 0}\mathbb{E}|x^{\varepsilon}(t)-\bar{x}(t)|^{p}+2^{p}\lim_{M\Delta_{2}\rightarrow\infty}\lim_{\Delta_{1},\Delta_{2}\rightarrow 0}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{x}(t)-{X}(t)|^{p}\Big{)}=0.\

The proof is complete. ∎

6 Strong error estimate

This section focuses on the strong error estimate of the MTEM scheme. To this end, we need somewhat stronger conditions compared with the strong convergence alone. In lieu of (S1) and (S4), we assume

  • (S1’)

    For any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} and yn2y\in\mathbb{R}^{n_{2}}, there exist constants θ11\theta_{1}\geq 1 and K>0K>0 such that

    |b(x1,y)b(x2,y)|+|σ(x1)σ(x2)|K|x1x2|(1+|x1|θ1+|x2|θ1+|y|θ1).\displaystyle|b(x_{1},y)-b(x_{2},y)|+|\sigma(x_{1})-\sigma(x_{2})|\leq K|x_{1}-x_{2}|(1+|x_{1}|^{\theta_{1}}+|x_{2}|^{\theta_{1}}+|y|^{\theta_{1}}).\
  • (S4’)

    For any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}} and y1,y2n2y_{1},y_{2}\in\mathbb{R}^{n_{2}}, there is a constant K5>0K_{5}>0 such that

    2(x1x2)T(b(x1,y1)b(x2,y2))\displaystyle 2(x_{1}-x_{2})^{T}(b(x_{1},y_{1})-b(x_{2},y_{2})) +|σ(x1)σ(x2)|2K5(|x1x2|2+|y1y2|2).\displaystyle+|\sigma(x_{1})-\sigma(x_{2})|^{2}\leq K_{5}\big{(}|x_{1}-x_{2}|^{2}+|y_{1}-y_{2}|^{2}\big{)}.
Remark 6.1.

It follows from (S1’)(\text{\bf S1'}) and (𝐒𝟐)({\bf S2}) that for any (x,y)n1×n2(x,y)\in\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}},

|b(x,y)|\displaystyle|b(x,y)| |b(x,y)b(x,0)|+|b(x,0)b(0,0)|+|b(0,0)|\displaystyle\leq|b(x,y)-b(x,0)|+|b(x,0)-b(0,0)|+|b(0,0)|\
K1|y|(1+|x|θ2+|y|θ2)+K|x|(1+|x|θ1)+|b(0,0)|\displaystyle\leq K_{1}|y|(1+|x|^{\theta_{2}}+|y|^{\theta_{2}})+K|x|\big{(}1+|x|^{\theta_{1}}\big{)}+|b(0,0)|\
C(1+|x|(θ1θ2)+1+|y|θ2+1),\displaystyle\leq C(1+|x|^{(\theta_{1}\vee\theta_{2})+1}+|y|^{\theta_{2}+1}),\

namely, combining (S1’)(\text{\bf S1'}) and (𝐒𝟐)({\bf S2}) can lead to (𝐒𝟒)({\bf S4}) with θ3=θ1θ2+1\theta_{3}=\theta_{1}\vee\theta_{2}+1 and θ4=θ2+1\theta_{4}=\theta_{2}+1.

Remark 6.2.

According to Remark 6.1, we can choose φ(u)=1+uθ1θ2\varphi(u)=1+u^{\theta_{1}\vee\theta_{2}} such that for any u1u\geq 1 and xn1x\in\mathbb{R}^{n_{1}} with |x|u|x|\leq u,

|b(x,y)|Csup|x|uφ(u)(1+|x|)+|y|θ2+1,u1.\displaystyle|b(x,y)|\leq C\sup_{|x|\leq u}\varphi(u)(1+|x|)+|y|^{\theta_{2}+1},~{}~{}~{}\forall~{}u\geq 1.

Using the similar techniques to that of Lemma 4.1, we derive that the averaged coefficient b¯\bar{b} keeps the property of polynomial growth. To avoid duplication we omit the proof.

Lemma 6.1.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, then for any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}}, there is a constant L¯>0\bar{L}>0 such that

|b¯(x1)b¯(x2)|L¯|x1x2|(1+|x1|θ1θ2+|x2|θ1θ2).\displaystyle\big{|}\bar{b}(x_{1})-\bar{b}(x_{2})\big{|}\leq\bar{L}|x_{1}-x_{2}|(1+|x_{1}|^{\theta_{1}\vee\theta_{2}}+|x_{2}|^{\theta_{1}\vee\theta_{2}}).
Lemma 6.2.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (S4’)(\text{\bf S4'}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)(\bf{F3}) hold with kθ2+1k\geq\theta_{2}+1, then for any x1,x2n1x_{1},x_{2}\in\mathbb{R}^{n_{1}},

2(x1x2)T(b¯(x1)b¯(x2))+|σ(x1)σ(x2)|2C|x1x2|2.\displaystyle 2(x_{1}-x_{2})^{T}\big{(}\bar{b}(x_{1})-\bar{b}(x_{2})\big{)}+|\sigma(x_{1})-\sigma(x_{2})|^{2}\leq C|x_{1}-x_{2}|^{2}.
Proof.

Due to (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}) and (F1)-(F3) with kθ2+1k\geq\theta_{2}+1, it follows from the definition of b¯(x)\bar{b}(x) and (S4’)(\text{\bf S4'}) that

2(x1x2)T(b¯(x1)b¯(x2))+|σ(x1)σ(x2)|2\displaystyle 2(x_{1}-x_{2})^{T}(\bar{b}(x_{1})-\bar{b}(x_{2}))+|\sigma(x_{1})-\sigma(x_{2})|^{2}\
=\displaystyle= n2×n2[2(x1x2)T(b(x1,y1)b(x2,y2))+|σ(x1)σ(x2)|2]π(dy1×dy2)\displaystyle\int_{\mathbb{R}^{n_{2}}\times\mathbb{R}^{n_{2}}}\Big{[}2(x_{1}-x_{2})^{T}\big{(}b(x_{1},y_{1})-b(x_{2},y_{2})\big{)}+|\sigma(x_{1})-\sigma(x_{2})|^{2}\Big{]}\pi(\mathrm{d}y_{1}\times\mathrm{d}y_{2})\
\displaystyle\leq K5|x1x2|2+K5n1×n2|y1y2|2π(dy1,dy2),\displaystyle K_{5}|x_{1}-x_{2}|^{2}+K_{5}\int_{\mathbb{R}^{n_{1}}\times\mathbb{R}^{n_{2}}}|y_{1}-y_{2}|^{2}\pi(\mathrm{d}y_{1},\mathrm{d}y_{2}),

here π𝒞(μx1,μx2)\pi\in\mathcal{C}(\mu^{x_{1}},\mu^{x_{2}}) is arbitrary. Then owing to the arbitrariness of π𝒞(μx1,μx2)\pi\in\mathcal{C}(\mu^{x_{1}},\mu^{x_{2}}),

2(x1x2)T(b¯(x1)b¯(x2))+|σ(x1)σ(x2)|2K5|x1x2|2+K5𝕎22(μx1,μx2).\displaystyle 2(x_{1}-x_{2})^{T}(\bar{b}(x_{1})-\bar{b}(x_{2}))+|\sigma(x_{1})-\sigma(x_{2})|^{2}\leq K_{5}|x_{1}-x_{2}|^{2}+K_{5}\mathbb{W}^{2}_{2}(\mu^{x_{1}},\mu^{x_{2}}).

Under (F1)-(F3), we deduce from (2.2) that

2(x1x2)T(b¯(x1)b¯(x2))+|σ(x1)σ(x2)|2C|x1x2|2.\displaystyle 2(x_{1}-x_{2})^{T}(\bar{b}(x_{1})-\bar{b}(x_{2}))+|\sigma(x_{1})-\sigma(x_{2})|^{2}\leq C|x_{1}-x_{2}|^{2}.

The proof is complete. ∎

According to Remark 6.1 and Lemma 4.12, we derive the moment estimate of |BM(x,Yx,n)b¯(x)||B_{M}(x,Y^{x,n})-\bar{b}(x)| as follows.

Lemma 6.3.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k2(θ2+1)k\geq 2(\theta_{2}+1) hold, then for any xn1x\in\mathbb{R}^{n_{1}}, Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}], and integers n0n\geq 0, M1M\geq 1,

𝔼|BM(x,Yx,n)b¯(x)|2C(1+|x|2(θ2+1))(Δ2+1M+1MΔ2).\displaystyle\mathbb{E}\Big{|}B_{M}(x,Y^{x,n})-\bar{b}(x)\Big{|}^{2}\leq C\big{(}1+|x|^{2(\theta_{2}+1)}\big{)}\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.

By the same proof techniques as the strong convergence of the MTEM scheme in Section 5, we give the error estimates of x¯(T)\bar{x}(T) and Z¯(T)\bar{Z}(T), as well as Z¯(T)\bar{Z}(T) and X(T)X(T), respectively.

Lemma 6.4.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (𝐒𝟑)({\bf S3}), (S4’)(\text{\bf S4'}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, then for any T>0T>0 and Δ1(0,1]\Delta_{1}\in(0,1],

𝔼|x¯(T)Z¯(T)|2CTΔ1.\displaystyle\mathbb{E}|\bar{x}(T)-\bar{Z}(T)|^{2}\leq C_{T}\Delta_{1}.
Proof.

Let e(t)=x¯(t)Z¯(t)e(t)=\bar{x}(t)-\bar{Z}(t) for any t0t\geq 0. Define the stopping time

θΔ1=ρΔ1,φ1(KΔ11/2)τφ1(KΔ11/2).\theta_{\Delta_{1}}=\rho_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\wedge\tau_{\varphi^{-1}(K\Delta_{1}^{-1/2})}.

Choosing p2(θ1θ2+1)p\geq 2(\theta_{1}\vee\theta_{2}+1) and then using the Young inequality for p>2p>2, we derive that for any T>0T>0,

𝔼|e(T)|2\displaystyle\mathbb{E}|e(T)|^{2} =𝔼(|e(T)|2I{θΔ1>T})+𝔼(|e(T)|2I{θΔ1T})\displaystyle=\mathbb{E}\big{(}|e(T)|^{2}I_{\{\theta_{\Delta_{1}}>T\}}\big{)}+\mathbb{E}\big{(}|e(T)|^{2}I_{\{\theta_{\Delta_{1}}\leq T\}}\big{)}\
𝔼(|e(T)|2I{θΔ1>T})+2Δ1𝔼|e(T)|pp+(p2)(θΔ1T)pΔ12p2.\displaystyle\leq\mathbb{E}\big{(}|e(T)|^{2}I_{\{\theta_{\Delta_{1}}>T\}}\big{)}+\frac{2\Delta_{1}\mathbb{E}|e(T)|^{p}}{p}+\frac{(p-2)\mathbb{P}(\theta_{\Delta_{1}}\leq T)}{p\Delta_{1}^{\frac{2}{p-2}}}. (6.1)

Under (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (𝐒𝟑)({\bf S3}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, it follows from the results of Lemmas 2.3 and 5.2 that

𝔼|e(T)|pC(𝔼|x¯(T)|p+𝔼|Z¯(T)|p)CT.\displaystyle\mathbb{E}|e(T)|^{p}\leq C\big{(}\mathbb{E}|\bar{x}(T)|^{p}+\mathbb{E}|\bar{Z}(T)|^{p}\big{)}\leq C_{T}.

Furthermore, by Remarks 2.2 and 5.2 we deduce that

(θΔ1T)\displaystyle\mathbb{P}(\theta_{\Delta_{1}}\leq T) (τφ1(KΔ11/2)T)+(ρΔ1,φ1(KΔ11/2)T)\displaystyle\leq\mathbb{P}(\tau_{\varphi^{-1}(K\Delta_{1}^{-1/2})}\leq T)+\mathbb{P}(\rho_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\leq T)\
CT(φ1(KΔ11/2))p.\displaystyle\leq\frac{C_{T}}{\big{(}\varphi^{-1}(K\Delta_{1}^{-1/2})\big{)}^{p}}.

Then inserting the above two inequalities into (6) and using p2(θ1θ2+1)p\geq 2(\theta_{1}\vee\theta_{2}+1) yield that

𝔼|e(T)|pCTΔ1+𝔼|e(TθΔ1)|2.\displaystyle\mathbb{E}|e(T)|^{p}\leq C_{T}\Delta_{1}+\mathbb{E}|e(T\wedge\theta_{\Delta_{1}})|^{2}.

Thus for the desired result it is sufficient to prove

𝔼|e(TθΔ1)|2CTΔ1.\displaystyle\mathbb{E}|e(T\wedge\theta_{\Delta_{1}})|^{2}\leq C_{T}\Delta_{1}.

Recalling the definition of the stopping time θΔ1\theta_{\Delta_{1}}, one observes that Z(t)=Z(t),0tTθΔ1Z^{*}(t)=Z(t),~{}0\leq t\leq T\wedge\theta_{\Delta_{1}}. Thus using the Ito^\mathrm{It\hat{o}} formula for (1.4) and (5.10) yields that

𝔼|e(TθΔ1)|2=\displaystyle\mathbb{E}|e(T\wedge\theta_{\Delta_{1}})|^{2}= 𝔼0TθΔ1[2eT(t)(b¯(x¯(t))b¯(Z(t)))+|σ(x¯(t))σ(Z(t))|2]dt\displaystyle\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}\Big{[}2e^{T}(t)\big{(}\bar{b}(\bar{x}(t))-\bar{b}(Z(t))\big{)}+|\sigma(\bar{x}(t))-\sigma(Z(t))|^{2}\Big{]}\mathrm{d}t\
\displaystyle\leq 𝔼0TθΔ1[2eT(t)(b¯(x¯(t))b¯(Z¯(t)))+|σ(x¯(t))σ(Z¯(t))|2]dt\displaystyle\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}\Big{[}2e^{T}(t)\big{(}\bar{b}(\bar{x}(t))-\bar{b}(\bar{Z}(t))\big{)}+|\sigma(\bar{x}(t))-\sigma(\bar{Z}(t))|^{2}\Big{]}\mathrm{d}t\
+2𝔼0TθΔ1eT(t)(b¯(Z¯(t))b¯(Z(t)))dt+𝔼0TθΔ1|σ(Z¯(t))σ(Z(t))|2dt\displaystyle~{}~{}~{}+2\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}e^{T}(t)\big{(}\bar{b}(\bar{Z}(t))-\bar{b}(Z(t))\big{)}\mathrm{d}t+\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}|\sigma(\bar{Z}(t))-\sigma(Z(t))|^{2}\mathrm{d}t\
+2𝔼0TθΔ1|σ(x¯(t))σ(Z¯(t))||σ(Z¯(t))σ(Z(t))|dt.\displaystyle~{}~{}~{}+2\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}|\sigma(\bar{x}(t))-\sigma(\bar{Z}(t))||\sigma(\bar{Z}(t))-\sigma(Z(t))|\mathrm{d}t.

Under (S4’) and (F1)-(F3) with kθ2+1k\geq\theta_{2}+1, utilizing the Lemma 6.2 and the Young inequality we derive that

𝔼|e(TθΔ1)|2C𝔼0TθΔ1|e(t)|2dt+J1+J2,\displaystyle\mathbb{E}|e(T\wedge\theta_{\Delta_{1}})|^{2}\leq C\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}|e(t)|^{2}\mathrm{d}t+J_{1}+J_{2}, (6.2)

where

J1=C𝔼0TθΔ1(|b¯(Z¯(t))b¯(Z(t))|2+|σ(Z¯(t))σ(Z(t))|2)dt,\displaystyle J_{1}=C\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}\Big{(}|\bar{b}(\bar{Z}(t))-\bar{b}(Z(t))|^{2}+|\sigma(\bar{Z}(t))-\sigma(Z(t))|^{2}\Big{)}\mathrm{d}t,\
J2=C𝔼0TθΔ1|σ(x¯(t))σ(Z¯(t))||σ(Z¯(t))σ(Z(t))|dt.\displaystyle J_{2}=C\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}|\sigma(\bar{x}(t))-\sigma(\bar{Z}(t))||\sigma(\bar{Z}(t))-\sigma(Z(t))|\mathrm{d}t.

Due to (S1’), (S2), (S3), (S5) and (F1)-(F3) with kθ12θ2θ4k\geq\theta_{1}\vee 2\theta_{2}\vee\theta_{4}, it follows from the results of Lemmas 5.2 and 6.1 that

J1\displaystyle J_{1} C0T𝔼[|Z¯(t)Z(t)|2(1+|Z¯(t)|2(θ1θ2)+|Z(t)|2(θ1θ2))]dt\displaystyle\leq C\int_{0}^{T}\mathbb{E}\big{[}|\bar{Z}(t)-Z(t)|^{2}(1+|\bar{Z}(t)|^{2(\theta_{1}\vee\theta_{2})}+|Z(t)|^{2(\theta_{1}\vee\theta_{2})})\big{]}\mathrm{d}t\
C0T(𝔼|Z¯(t)Z(t)|4)12[𝔼(1+|Z¯(t)|4(θ1θ2)+|Z(t)|4(θ1θ2))]12dtCTΔ1.\displaystyle\leq C\int_{0}^{T}\Big{(}\mathbb{E}|\bar{Z}(t)-Z(t)|^{4}\Big{)}^{\frac{1}{2}}\Big{[}\mathbb{E}\big{(}1+|\bar{Z}(t)|^{4(\theta_{1}\vee\theta_{2})}+|Z(t)|^{4(\theta_{1}\vee\theta_{2})}\big{)}\Big{]}^{\frac{1}{2}}\mathrm{d}t\leq C_{T}\Delta_{1}. (6.3)

In addition, using the Young inequality and the Hölder inequality yields that

J2\displaystyle J_{2} C𝔼0TθΔ1|e(t)||Z¯(t)Z(t)|(1+|x¯(t)|2θ1+|Z¯(t)|2θ1+|Z(t)|2θ1)dt\displaystyle\leq C\mathbb{E}\int_{0}^{T\wedge\theta_{\Delta_{1}}}|e(t)||\bar{Z}(t)-Z(t)|(1+|\bar{x}(t)|^{2\theta_{1}}+|\bar{Z}(t)|^{2\theta_{1}}+|Z(t)|^{2\theta_{1}})\mathrm{d}t\
C0T(𝔼|Z¯(t)Z(t)|4)12[𝔼(1+|x¯(t)|8θ1+|Z¯(t)|8θ1+|Z(t)|8θ1)]12dt\displaystyle\leq C\int_{0}^{T}\big{(}\mathbb{E}|\bar{Z}(t)-Z(t)|^{4}\big{)}^{\frac{1}{2}}\Big{[}\mathbb{E}\big{(}1+|\bar{x}(t)|^{8\theta_{1}}+|\bar{Z}(t)|^{8\theta_{1}}+|Z(t)|^{8\theta_{1}}\big{)}\Big{]}^{\frac{1}{2}}\mathrm{d}t\
+C0T𝔼|e(tθΔ1)|2dt.\displaystyle~{}~{}~{}+C\int_{0}^{T}\mathbb{E}|e(t\wedge\theta_{\Delta_{1}})|^{2}\mathrm{d}t.

Similarly to (6), applying Lemmas 2.3 and 5.2 we show that

J2\displaystyle J_{2} CTΔ1+CT0T𝔼|e(tθΔ1)|2dt.\displaystyle\leq C_{T}\Delta_{1}+C_{T}\int_{0}^{T}\mathbb{E}|e(t\wedge\theta_{\Delta_{1}})|^{2}\mathrm{d}t. (6.4)

Inserting (6) and (6.4) into (6.2) and then using Gronwall’s inequality derive that

𝔼|e(TθΔ1)|2CTΔ1,\displaystyle\mathbb{E}|e(T\wedge\theta_{\Delta_{1}})|^{2}\leq C_{T}\Delta_{1},\

which implies the desired result. The proof is complete. ∎

Lemma 6.5.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (𝐒𝟑)({\bf S3}), (S4’)(\text{\bf S4'}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) with k[2(2θ1+1)2(θ1θ2+1)](θ2+1)k\geq[2(2\theta_{1}+1)\vee 2(\theta_{1}\vee\theta_{2}+1)](\theta_{2}+1) hold, then for any T>0T>0, Δ1(0,1]\Delta_{1}\in(0,1], Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}] and M1M\geq 1,

𝔼|X¯(T)Z¯(T)|2CT(Δ1+Δ2+1M+1MΔ2).\displaystyle\mathbb{E}|\bar{X}(T)-\bar{Z}(T)|^{2}\leq C_{T}\Big{(}\Delta_{1}+\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.
Proof.

Define the stopping time

θ¯Δ1=ρ¯Δ1,φ1(KΔ11/2)ρΔ1,φ1(KΔ11/2),\bar{\theta}_{\Delta_{1}}=\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\wedge\rho_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})},

where ρ¯Δ1,φ1(KΔ11/2)\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})} and ρΔ1,φ1(KΔ11/2)\rho_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})} are given by (5.8) and (5.15). Due to k[2(2θ1+1)2(θ1θ2+1)](θ2+1)k\geq[2(2\theta_{1}+1)\vee 2(\theta_{1}\vee\theta_{2}+1)](\theta_{2}+1), we can choose a constant pp such that

2<2(θ1θ2+1)2(2θ1+1)pk/(θ2+1).2<2(\theta_{1}\vee\theta_{2}+1)\vee 2(2\theta_{1}+1)\leq p\leq k/(\theta_{2}+1).

By (S1’), (S2), (S3), (S5) and (F1)-(F3), using Lemmas 5.1 and 5.2 as well as the Hölder inequality yields that

supΔ1(0,1]𝔼(sup0tT|Z¯(t)|p)supΔ1(0,1],Δ2(0,Δ^2]𝔼(supt[0,T]|X¯(t)|p)CT.\displaystyle\sup_{\Delta_{1}\in(0,1]}\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|\bar{Z}(t)|^{p}\Big{)}\vee\sup_{\Delta_{1}\in(0,1],\Delta_{2}\in(0,\hat{\Delta}_{2}]}\mathbb{E}\Big{(}\sup_{t\in[0,T]}|\bar{X}(t)|^{p}\Big{)}\leq C_{T}. (6.5)

Then applying the Young\mathrm{Young} inequality, for any δ>0\delta>0 we obtain that

𝔼|e¯(T)|2=𝔼(|e¯(T)|2I{θ¯Δ1>T})+𝔼(|e¯(T)|2I{θ¯Δ1T})\displaystyle\mathbb{E}|\bar{e}(T)|^{2}=\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}>T\}}\big{)}+\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}\leq T\}}\big{)}\
\displaystyle\leq 𝔼(|e¯(T)|2I{θ¯Δ1>T})+2Δ1p𝔼|e¯(T)|p+p2pΔ12p2(θ¯Δ1T).\displaystyle\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}>T\}}\big{)}+\frac{2\Delta_{1}}{p}\mathbb{E}|\bar{e}(T)|^{p}+\frac{p-2}{p\Delta_{1}^{\frac{2}{p-2}}}\mathbb{P}(\bar{\theta}_{\Delta_{1}}\leq T). (6.6)

It follows from (6.5) that

𝔼|e¯(T)|p2p1𝔼|Z¯(T)|p+2p1𝔼|X¯(t)|pCT.\displaystyle\mathbb{E}|\bar{e}(T)|^{p}\leq 2^{p-1}\mathbb{E}|\bar{Z}(T)|^{p}+2^{p-1}\mathbb{E}|\bar{X}(t)|^{p}\leq C_{T}.\

Furthermore, by the Markov inequality and (6.5) we derive that

(ρ¯Δ1,φ1(KΔ11/2)T)\displaystyle\mathbb{P}\Big{(}\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\leq T\Big{)} (|X¯(Tρ¯Δ1,φ1(KΔ11/2))|φ1(KΔ11/2))\displaystyle\leq\mathbb{P}\Big{(}|\bar{X}(T\wedge\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})})|\geq\varphi^{-1}(K\Delta_{1}^{-1/2})\Big{)}\
𝔼|X¯(Tρ¯Δ1,φ1(KΔ11/2))|p(φ1(KΔ11/2))pCT(φ1(KΔ11/2))p.\displaystyle\leq\frac{\mathbb{E}\Big{|}\bar{X}(T\wedge\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})})\Big{|}^{p}}{\big{(}\varphi^{-1}(K\Delta_{1}^{-1/2})\big{)}^{p}}\leq\frac{C_{T}}{\big{(}\varphi^{-1}(K\Delta_{1}^{-1/2})\big{)}^{p}}.

Then combining the above inequality and Remark 5.2 gives that

(θ¯Δ1T)(ρΔ1,φ1(KΔ11/2)T)+(ρ¯Δ1,φ1(KΔ11/2)T)CT(φ1(KΔ112))p.\displaystyle\mathbb{P}(\bar{\theta}_{\Delta_{1}}\leq T)\leq\mathbb{P}(\rho_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\leq T)+\mathbb{P}(\bar{\rho}_{\Delta_{1},\varphi^{-1}(K\Delta_{1}^{-1/2})}\leq T)\leq\frac{C_{T}}{(\varphi^{-1}(K\Delta_{1}^{-\frac{1}{2}}))^{p}}.

Due to p2(θ1θ2+1)p\geq 2(\theta_{1}\vee\theta_{2}+1), inserting the above inequality into (6) shows that

𝔼|e¯(T)|2\displaystyle\mathbb{E}|\bar{e}(T)|^{2} 𝔼(|e¯(T)|2I{θ¯Δ1>T})+CTΔ1p+CTpΔ2p2(φ1(KΔ112))p\displaystyle\leq\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}>T\}}\big{)}+\frac{C_{T}\Delta_{1}}{p}+\frac{C_{T}}{p\Delta^{\frac{2}{p-2}}(\varphi^{-1}(K\Delta_{1}^{-\frac{1}{2}}))^{p}}\
𝔼(|e¯(T)|2I{θ¯Δ1>T})+CTΔ1.\displaystyle\leq\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}>T\}}\big{)}+C_{T}\Delta_{1}.

Hence for the desired result it remains to prove that

𝔼(|e¯(T)|2I{θ¯Δ1>T})CTΔ1.\displaystyle\mathbb{E}\big{(}|\bar{e}(T)|^{2}I_{\{\bar{\theta}_{\Delta_{1}}>T\}}\big{)}\leq C_{T}\Delta_{1}.

Obviously, X(t)=X(t)X^{*}(t)=X(t) and Z(t)=Z(t)Z^{*}(t)=Z(t) for any 0tTθ¯Δ10\leq t\leq T\wedge\bar{\theta}_{\Delta_{1}}. Using the Ito^\mathrm{It\hat{o}} formula for (3.7) and (5.10) and the Young inequality, under (S4’) and (F1)-(F3), by Lemma 6.2 we arrive at that for any T>0T>0,

𝔼|e¯(Tθ¯Δ1)|2\displaystyle\mathbb{E}|\bar{e}(T\wedge\bar{\theta}_{\Delta_{1}})|^{2} =𝔼0Tθ¯Δ1[2e¯T(t)(BM(X(t),YX(t))b¯(Z(t)))+|σ(X(t))σ(Z(t))|2]dt\displaystyle=\mathbb{E}\int_{0}^{T\wedge\bar{\theta}_{\Delta_{1}}}\Big{[}2\bar{e}^{T}(t)\Big{(}B_{M}\Big{(}X(t),Y^{X(t)}\Big{)}-\bar{b}(Z(t))\Big{)}+|\sigma(X(t))-\sigma(Z(t))|^{2}\Big{]}\mathrm{d}t\
𝔼0Tθ¯Δ1[2e¯T(t)(b¯(X¯(t))b¯(Z¯(t)))+|σ(X¯(t))σ(Z¯(t))|2]dt\displaystyle\leq\mathbb{E}\int_{0}^{T\wedge\bar{\theta}_{\Delta_{1}}}\Big{[}2\bar{e}^{T}(t)\Big{(}\bar{b}(\bar{X}(t))-\bar{b}(\bar{Z}(t))\Big{)}+|\sigma(\bar{X}(t))-\sigma(\bar{Z}(t))|^{2}\Big{]}\mathrm{d}t\
+0TC𝔼|e¯(tθ¯Δ1)|2dt+I1+I2+I3+I4\displaystyle~{}~{}~{}+\int_{0}^{T}C\mathbb{E}|\bar{e}(t\wedge\bar{\theta}_{\Delta_{1}})|^{2}\mathrm{d}t+I_{1}+I_{2}+I_{3}+I_{4}\
0TC𝔼|e¯(tθ¯Δ1)|2dt+I1+I2+I3+I4,\displaystyle\leq\int_{0}^{T}C\mathbb{E}|\bar{e}(t\wedge\bar{\theta}_{\Delta_{1}})|^{2}\mathrm{d}t+I_{1}+I_{2}+I_{3}+I_{4}, (6.7)

where

I1=0T𝔼|BM(X(t),YX(t))b¯(X(t))|2dt,\displaystyle I_{1}=\int_{0}^{T}\mathbb{E}\Big{|}B_{M}\Big{(}X(t),Y^{X(t)}\Big{)}-\bar{b}(X(t))\Big{|}^{2}\mathrm{d}t,\
I2=C0T𝔼(|b¯(X(t))b¯(X¯(t))|2+|σ(X(t))σ(X¯(t))|2)dt,\displaystyle I_{2}=C\int_{0}^{T}\mathbb{E}\big{(}|\bar{b}(X(t))-\bar{b}(\bar{X}(t))|^{2}+|\sigma(X(t))-\sigma(\bar{X}(t))|^{2}\big{)}\mathrm{d}t,\
I3=C0T𝔼(|b¯(Z¯(t))b¯(Z(t))|2+|σ(Z¯(t))σ(Z(t))|2)dt,\displaystyle I_{3}=C\int_{0}^{T}\mathbb{E}\big{(}|\bar{b}(\bar{Z}(t))-\bar{b}(Z(t))|^{2}+|\sigma(\bar{Z}(t))-\sigma(Z(t))|^{2}\big{)}\mathrm{d}t,\
I4=C𝔼0Tθ¯Δ12|σ(X¯(t))σ(Z¯(t))|(|σ(X(t))σ(X¯(t))|+|σ(Z¯(t))σ(Z(t))|)dt.\displaystyle I_{4}=C\mathbb{E}\int_{0}^{T\wedge\bar{\theta}_{\Delta_{1}}}2|\sigma(\bar{X}(t))-\sigma(\bar{Z}(t))|\big{(}|\sigma(X(t))-\sigma(\bar{X}(t))|+|\sigma(\bar{Z}(t))-\sigma(Z(t))|\big{)}\mathrm{d}t.

In addition, owing to (S1’), (S2) and (F1)-(F3) with k2(θ2+1)k\geq 2(\theta_{2}+1), applying (4.5) and Lemma 6.3 implies that for any 0tT0\leq t\leq T,

𝔼|BM(X(t),YX(t))b¯(X(t))|2\displaystyle\mathbb{E}\Big{|}B_{M}\Big{(}X(t),Y^{X(t)}\Big{)}-\bar{b}(X(t))\Big{|}^{2}\ =𝔼|BM(Xn1(t),YXn1(t))b¯(Xn1(t))|2\displaystyle=\mathbb{E}\Big{|}B_{M}\Big{(}X_{n_{1}(t)},Y^{X_{n_{1}(t)}}\Big{)}-\bar{b}(X_{n_{1}(t)})\Big{|}^{2}
=𝔼[𝔼(|BM(Xn1(t),Yn1(t)Xn1(t))b¯(Xn1(t))|2|Xn1(t))]\displaystyle=\mathbb{E}\Big{[}\mathbb{E}\Big{(}\Big{|}B_{M}\Big{(}X_{n_{1}(t)},Y^{X_{n_{1}(t)}}_{n_{1}(t)}\Big{)}-\bar{b}(X_{n_{1}(t)})\Big{|}^{2}\Big{|}X_{n_{1}(t)}\Big{)}\Big{]}\
=𝔼(𝔼|BM(x,Yn1(t)x)b¯(x)|2|x=Xn1(t))\displaystyle=\mathbb{E}\Big{(}\mathbb{E}\Big{|}B_{M}\Big{(}x,Y^{x}_{n_{1}(t)}\Big{)}-\bar{b}(x)\Big{|}^{2}\Big{|}_{x=X_{n_{1}(t)}}\Big{)}\
C(Δ2+1M+1MΔ2)(1+𝔼|Xn1(t)|2(θ2+1))).\displaystyle\leq C\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}\big{(}1+\mathbb{E}|X_{n_{1}(t)}|^{2(\theta_{2}+1)}\big{)}\big{)}.

Furthermore, due to p>2(θ2+1)p>2(\theta_{2}+1), utilizing (6.5) and the Hölder inequality we deduce that

I1\displaystyle I_{1} C(Δ2+1M+1MΔ2)0T(1+𝔼|Xn1(t)|2(θ2+1))dt\displaystyle\leq C\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}\int_{0}^{T}\big{(}1+\mathbb{E}|X_{n_{1}(t)}|^{2(\theta_{2}+1)}\big{)}\mathrm{d}t\
C(Δ2+1M+1MΔ2)0T(1+(𝔼|Xn1(t)|p)2(θ2+1)p)dt\displaystyle\leq C\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}\int_{0}^{T}\big{(}1+\big{(}\mathbb{E}|X_{n_{1}(t)}|^{p}\big{)}^{\frac{2(\theta_{2}+1)}{p}}\big{)}\mathrm{d}t\
CT(Δ2+1M+1MΔ2).\displaystyle\leq C_{T}\Big{(}\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}. (6.8)

Under (S1’), (S2) and (F1)-(F3) with kθ12θ2k\geq\theta_{1}\vee 2\theta_{2}, by Lemma 6.1 and the Hölder inequality we derive that

I2+I3\displaystyle I_{2}+I_{3} C0T𝔼(|X(t)X¯(t)|2(1+|X(t)|2(θ1θ2)+|X¯(t)|2(θ1θ2)))dt\displaystyle\leq C\int_{0}^{T}\mathbb{E}\Big{(}|X(t)-\bar{X}(t)|^{2}\big{(}1+|X(t)|^{2(\theta_{1}\vee\theta_{2})}+|\bar{X}(t)|^{2(\theta_{1}\vee\theta_{2})}\big{)}\Big{)}\mathrm{d}t\
+C0T𝔼(|Z(t)Z¯(t)|2(1+|Z(t)|2(θ1θ2)+|Z¯(t)|2(θ1θ2)))dt\displaystyle~{}~{}~{}+C\int_{0}^{T}\mathbb{E}\Big{(}|Z(t)-\bar{Z}(t)|^{2}\big{(}1+|Z(t)|^{2(\theta_{1}\vee\theta_{2})}+|\bar{Z}(t)|^{2(\theta_{1}\vee\theta_{2})}\big{)}\Big{)}\mathrm{d}t\
C0T(𝔼|X(t)X¯(t)|p)2p(𝔼(1+|X(t)|2p(θ1θ2)p2+|X¯(t)|2p(θ1θ2)p2))p2pdt\displaystyle\leq C\int_{0}^{T}\Big{(}\mathbb{E}|X(t)-\bar{X}(t)|^{p}\Big{)}^{\frac{2}{p}}\Big{(}\mathbb{E}\big{(}1+|X(t)|^{\frac{2p(\theta_{1}\vee\theta_{2})}{p-2}}+|\bar{X}(t)|^{\frac{2p(\theta_{1}\vee\theta_{2})}{p-2}}\big{)}\Big{)}^{\frac{p-2}{p}}\mathrm{d}t\
+C0T(𝔼|Z(t)Z¯(t)|p)2p(𝔼(1+|Z(t)|2p(θ1θ2)p2+|Z¯(t)|2p(θ1θ2)p2))p2pdt.\displaystyle~{}~{}~{}+C\int_{0}^{T}\Big{(}\mathbb{E}|Z(t)-\bar{Z}(t)|^{p}\Big{)}^{\frac{2}{p}}\Big{(}\mathbb{E}\big{(}1+|Z(t)|^{\frac{2p(\theta_{1}\vee\theta_{2})}{p-2}}+|\bar{Z}(t)|^{\frac{2p(\theta_{1}\vee\theta_{2})}{p-2}}\big{)}\Big{)}^{\frac{p-2}{p}}\mathrm{d}t.\

Thanks to 2(θ1θ2+1)pk/(θ2+1)2(\theta_{1}\vee\theta_{2}+1)\leq p\leq k/(\theta_{2}+1), we have 2p(θ1θ2)/(p2)pk/(θ2+1)2p(\theta_{1}\vee\theta_{2})/(p-2)\leq p\leq k/(\theta_{2}+1). Then applying Lemmas 5.1 and 5.2 and the Hölder inequality yields that

I2+I3\displaystyle I_{2}+I_{3} C0T(𝔼|X(t)X¯(t)|p)2p(𝔼(1+|X(t)|p+|X¯(t)|p))p2pdt\displaystyle\leq C\int_{0}^{T}\Big{(}\mathbb{E}|X(t)-\bar{X}(t)|^{p}\Big{)}^{\frac{2}{p}}\Big{(}\mathbb{E}(1+|X(t)|^{p}+|\bar{X}(t)|^{p})\Big{)}^{\frac{p-2}{p}}\mathrm{d}t\
+C0T(𝔼|Z(t)Z¯(t)|p)2p(𝔼(1+|Z(t)|p+|Z¯(t)|p))p2pdtCTΔ1.\displaystyle~{}~{}~{}+C\int_{0}^{T}\Big{(}\mathbb{E}|Z(t)-\bar{Z}(t)|^{p}\Big{)}^{\frac{2}{p}}\Big{(}\mathbb{E}(1+|Z(t)|^{p}+|\bar{Z}(t)|^{p})\Big{)}^{\frac{p-2}{p}}\mathrm{d}t\leq C_{T}\Delta_{1}. (6.9)

In view of (S1’), together with using the Young inequality and the Hölder inequality, we also obtain that

I4\displaystyle I_{4} C𝔼0Tθ¯Δ1|e¯(t)|(|X(t)X¯(t)|+|Z(t)Z¯(t)|)\displaystyle\leq C\mathbb{E}\int_{0}^{T\wedge\bar{\theta}_{\Delta_{1}}}|\bar{e}(t)|\big{(}|X(t)-\bar{X}(t)|+|Z(t)-\bar{Z}(t)|\big{)}\
×(1+|X(t)|2θ1+|X¯(t)|2θ1+|Z¯(t)|2θ1+|Z(t)|2θ1)dt\displaystyle~{}~{}~{}~{}~{}~{}\times\big{(}1+|X(t)|^{2\theta_{1}}+|\bar{X}(t)|^{2\theta_{1}}+|\bar{Z}(t)|^{2\theta_{1}}+|Z(t)|^{2\theta_{1}}\big{)}\mathrm{d}t
C0T𝔼|e¯(tθ¯Δ1)|2dt+C0T[𝔼(|X(t)X¯(t)|p+|Z(t)Z¯(t)|p)]2p\displaystyle\leq C\int_{0}^{T}\mathbb{E}|\bar{e}(t\wedge\bar{\theta}_{\Delta_{1}})|^{2}\mathrm{d}t+C\int_{0}^{T}\Big{[}\mathbb{E}\big{(}|X(t)-\bar{X}(t)|^{p}+|Z(t)-\bar{Z}(t)|^{p}\big{)}\Big{]}^{\frac{2}{p}}\
×[𝔼(1+|X(t)|4pθ1p2+|X¯(t)|4pθ1p2+|Z(t)|4pθ1p2+|Z¯(t)|4pθ1p2)]p2pdt.\displaystyle~{}~{}~{}\times\Big{[}\mathbb{E}\big{(}1+|X(t)|^{\frac{4p\theta_{1}}{p-2}}+|\bar{X}(t)|^{\frac{4p\theta_{1}}{p-2}}+|Z(t)|^{\frac{4p\theta_{1}}{p-2}}+|\bar{Z}(t)|^{\frac{4p\theta_{1}}{p-2}}\big{)}\Big{]}^{\frac{p-2}{p}}\mathrm{d}t.

Similarly, owing to 2(2θ1+1)pk/(θ2+1)2(2\theta_{1}+1)\leq p\leq k/(\theta_{2}+1), 4pθ1/(p2)pk/(θ2+1)4p\theta_{1}/(p-2)\leq p\leq k/(\theta_{2}+1). By means of Lemmas 5.1 and 5.2 and using the Hölder inequality we deduce that

I4\displaystyle I_{4} C0T𝔼|e¯(tθ¯Δ1)|2dt+C0T[𝔼(|X(t)X¯(t)|p+|Z(t)Z¯(t)|p)]2p\displaystyle\leq C\int_{0}^{T}\mathbb{E}|\bar{e}(t\wedge\bar{\theta}_{\Delta_{1}})|^{2}\mathrm{d}t+C\int_{0}^{T}\Big{[}\mathbb{E}\big{(}|X(t)-\bar{X}(t)|^{p}+|Z(t)-\bar{Z}(t)|^{p}\big{)}\Big{]}^{\frac{2}{p}}\
×[𝔼(1+|X(t)|p+|X¯(t)|p+|Z(t)|p+|Z¯(t)|p)]p2pdt\displaystyle~{}~{}~{}\times\Big{[}\mathbb{E}\big{(}1+|X(t)|^{p}+|\bar{X}(t)|^{p}+|Z(t)|^{p}+|\bar{Z}(t)|^{p}\big{)}\Big{]}^{\frac{p-2}{p}}\mathrm{d}t\
C0T𝔼|e¯(tθ¯Δ1)|2dt+CTΔ1.\displaystyle\leq C\int_{0}^{T}\mathbb{E}|\bar{e}(t\wedge\bar{\theta}_{\Delta_{1}})|^{2}\mathrm{d}t+C_{T}\Delta_{1}. (6.10)

Then inserting (6)-(6) into (6) implies that

𝔼|e¯Δ1(TβΔ1)|C0T𝔼|e¯(tβΔ1)|2dt+CT(Δ1+Δ2+1M+1MΔ2).\displaystyle\mathbb{E}|\bar{e}_{\Delta_{1}}(T\wedge\beta_{\Delta_{1}})|\leq C\int_{0}^{T}\mathbb{E}|\bar{e}(t\wedge\beta_{\Delta_{1}})|^{2}\mathrm{d}t+C_{T}\Big{(}\Delta_{1}+\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.

Using the Gronwall inequality shows that

𝔼|e¯Δ1(TβΔ1)|CT(Δ1+Δ2+1M+1MΔ2),\displaystyle\mathbb{E}|\bar{e}_{\Delta_{1}}(T\wedge\beta_{\Delta_{1}})|\leq C_{T}\Big{(}\Delta_{1}+\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)},

which implies the desired result. The proof is complete. ∎

Combining Lemmas 5.1 and 6.4 as well as 6.5, the theorem on error estimate of the MTEM scheme can be obtained immediately as follows.

Theorem 6.1.

If (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (𝐒𝟑)({\bf S3}), (S4’)(\text{\bf S4'}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with k[2(2θ1+1)2(θ1θ2+1)](θ2+1)k\geq[2(2\theta_{1}+1)\vee 2(\theta_{1}\vee\theta_{2}+1)](\theta_{2}+1), then for any T>0T>0, Δ1(0,Δ¯1]\Delta_{1}\in(0,\bar{\Delta}_{1}], Δ2(0,Δ¯2]\Delta_{2}\in(0,\bar{\Delta}_{2}] and M1M\geq 1,

𝔼|x¯(T)X(T)|2CT(Δ1+Δ2+1M+1MΔ2).\displaystyle\mathbb{E}|\bar{x}(T)-{X}(T)|^{2}\leq C_{T}\Big{(}\Delta_{1}+\Delta_{2}+\frac{1}{M}+\frac{1}{M\Delta_{2}}\Big{)}.
Remark 6.3.

Theorem 6.1 provides the strong error estimate between the exact solution of the averaged equation (1.4) and the MTEM numerical solution is provided. Suppose that the strong convergence rate of averaging principle is further obtained, the strong error estimate between the slow component of original system and the MTEM numerical solution is also acquired.

7 Numerical examples

This section gives two examples and carries out some numerical experiments by the MTEM scheme to verify the theoretical results.

Example 7.1.

Consider the following SFSDE

{dxε(t)=((xε(t))3yε(t))dt+xε(t)dW1(t),dyε(t)=1ε(xε(t)yε(t))dt+1εdW2(t)\begin{cases}\mathrm{d}x^{\varepsilon}(t)=\big{(}-(x^{\varepsilon}(t))^{3}-y^{\varepsilon}(t)\big{)}\mathrm{d}t+x^{\varepsilon}(t)\mathrm{d}W^{1}(t),\\ \mathrm{d}y^{\varepsilon}(t)=\displaystyle\frac{1}{\varepsilon}\big{(}x^{\varepsilon}(t)-y^{\varepsilon}(t)\big{)}\mathrm{d}t+\frac{1}{\sqrt{\varepsilon}}\mathrm{d}W^{2}(t)\end{cases} (7.1)

with the initial value (x0,y0)=(1,1)(x_{0},y_{0})=(1,1), where W1(t)W^{1}(t) and W2(t)W^{2}(t) are mutually independent 11-dimensional Brownian motions, respectively. Obviously,

b(x,y)=x3y,σ(x)=x,f(x,y)=xy,g(x,y)=1.b(x,y)=-x^{3}-y,~{}~{}~{}\sigma(x)=x,~{}~{}~{}f(x,y)=x-y,~{}~{}~{}g(x,y)=1.

It can be verified that (S1’)(\text{\bf S1'}), (𝐒𝟐)({\bf S2}), (𝐒𝟑)({\bf S3}), (S4’)(\text{\bf S4'}), (𝐒𝟓)({\bf S5}) and (𝐅𝟏){\bf(F1)}-(𝐅𝟑){\bf(F3)} hold with θ1=2,θ2=1\theta_{1}=2,\theta_{2}=1 and any k2k\geq 2. The corresponding frozen equation is described by

dyx(s)=(xyx(s))ds+dW2(s),\displaystyle\mathrm{d}y^{x}(s)=(x-y^{x}(s))\mathrm{d}s+\mathrm{d}W^{2}(s), (7.2)

with initial value y0=1y_{0}=1. By solving the Fokker-Planck equation, the invariant probability density of (7.2) is μx(dy)=e(yx)2πdy.\mu^{x}(dy)=\frac{e^{-(y-x)^{2}}}{\sqrt{\pi}}\mathrm{d}y. Then the averaged equation is described by

dx¯(t)=(x¯3(t)x¯(t))dt+x¯(t)dW1(t)\mathrm{d}\bar{x}(t)=\big{(}-\bar{x}^{3}(t)-\bar{x}(t)\big{)}\mathrm{d}t+\bar{x}(t)\mathrm{d}W^{1}(t) (7.3)

with x¯(0)=x0\bar{x}(0)=x_{0}. Its exact solution has a closed form (see, e.g., [21, 24])

x¯(t)=x0exp(32t+W1(t))1+2x020texp(3s+2W1(s))ds.\displaystyle\bar{x}(t)=\frac{x_{0}\exp(-\frac{3}{2}t+W^{1}(t))}{\sqrt{1+2x_{0}^{2}\int_{0}^{t}\exp(-3s+2W^{1}(s))\mathrm{d}s}}.\

According to Remark 6.2, we can choose φ(u)=1+u2,u1\varphi(u)=1+u^{2},~{}\forall~{}u\geq 1. For the fixed Δ1,Δ2(0,1]\Delta_{1},\Delta_{2}\in(0,1] and integer M1M\geq 1, define the MTEM scheme for (7.1): for any n0n\geq 0,

{X0=x0,Xn=(|Xn|(2Δ1121)12)Xn|Xn|,Y0Xn=y0,Ym+1Xn=YmXn+(XnYmXn)Δ2+ΔWn,m2,m=0,1,,M1,BM(Xn,YXn)=1Mm=1M((Xn)3+YmXn),Xn+1=Xn+BM(Xn,YXn)Δ1+XnΔWn1.\begin{cases}X_{0}=x_{0},X^{*}_{n}=\Big{(}|X_{n}|\wedge\big{(}2\Delta_{1}^{-\frac{1}{2}}-1\big{)}^{\frac{1}{2}}\Big{)}\frac{X_{n}}{|X_{n}|},~{}Y^{X_{n}^{*}}_{0}=y_{0},\\ Y^{X^{*}_{n}}_{m+1}=Y^{X^{*}_{n}}_{m}+(X^{*}_{n}-Y^{X^{*}_{n}}_{m})\Delta_{2}+\Delta W^{2}_{n,m},~{}~{}~{}m=0,1,\dots,M-1,\\ B_{M}(X^{*}_{n},Y^{X^{*}_{n}})=-\displaystyle{\frac{1}{M}}\sum_{m=1}^{M}\Big{(}\big{(}X^{*}_{n}\big{)}^{3}+Y^{X^{*}_{n}}_{m}\Big{)},\\ X_{n+1}=X_{n}+B_{M}(X^{*}_{n},Y^{X^{*}_{n}})\Delta_{1}+X_{n}\Delta W^{1}_{n}.\end{cases}\vspace{-1mm}\!\!\! (7.4)

Owing to Theorem 2.2, one notices that xε(t)x^{\varepsilon}(t) converges to x¯(t)\bar{x}(t) as ε0\varepsilon\rightarrow 0. Next we pay attention to the strong convergence between x¯(t)\bar{x}(t) and the numerical solution X(t)X(t) by the MTEM scheme (7.4) as Δ1,Δ20\Delta_{1},\Delta_{2}\rightarrow 0 and MΔ2M\Delta_{2}\rightarrow\infty revealed by Theorem 6.1. To verify this result, we carry out some numerical experiments by the MTEM scheme. Provided that we want to bound the error by 𝒪(2q)(q>0)\mathcal{O}(2^{-q})(q>0), the optimal parameters are derived by Theorem 6.1 as follows:

Δ1=𝒪(2q),Δ2=𝒪(2q),M=𝒪(22q).\displaystyle\Delta_{1}=\mathcal{O}(2^{-q}),~{}~{}\Delta_{2}=\mathcal{O}(2^{-q}),~{}~{}M=\mathcal{O}(2^{2q}).

In the numerical calculations, using 500500 sample points we compute the sample mean square of the error (SMSE)

𝔼|x¯(t)X(t)|21500j=1500|x¯(j)(nΔ1)Xn(j)|2,\displaystyle\mathbb{E}|\bar{x}(t)-X(t)|^{2}\approx\frac{1}{500}\sum_{j=1}^{500}|\bar{x}^{(j)}(n\Delta_{1})-X^{(j)}_{n}|^{2}, (7.5)

where x¯(j)(nΔ1)\bar{x}^{(j)}(n\Delta_{1}) and Xn(j)X^{(j)}_{n} are sequences of independent copies of x¯(nΔ1)\bar{x}(n\Delta_{1}) and XnX_{n}, respectively. Note that for the fixed nn and jj, x¯(j)(nΔ1)\bar{x}^{(j)}(n\Delta_{1}) and Xn(j)X^{(j)}_{n} are generated by a same Brownian motion. Then we carry out numerical experiments by implementing (7.4) using MATLAB. In Figure 1, the blue solid line depicts the SMSE for q=2,3,4,5,6,7q=2,3,4,5,6,7 with 500500 sample points. The red dotted line plots the reference line with the slope -1. In addition, we plot 10 groups of sample paths of x¯(t)\bar{x}(t) and X(t)X(t) for t[0,5]t\in[0,5] with (Δ1,Δ2,M)=(28,26,212)(\Delta_{1},\Delta_{2},M)=(2^{-8},2^{-6},2^{12}). The Figure 2 only depicts four groups of them.

Refer to caption
Figure 1: The SMSE for q=2,3,4,5,6,7q=2,3,4,5,6,7 with 500500 sample points. The red dashed line is the reference with slope -1.
Refer to caption
Figure 2: Four pairs of sample paths of x¯(t)\bar{x}(t) and X(t)X(t) for t[0,5]t\in[0,5] with (Δ1,Δ2,M)=(28,26,212)(\Delta_{1},\Delta_{2},M)=(2^{-8},2^{-6},2^{12}).
Remark 7.1.

A scalar nonlinear SFSDE is addressed in [29], which is described by

{dxε(t)=[(yε(t))3+sin(2πt)]dt+yε(t)dW1(t),dyε(t)=1ε(yε(t)xε(t))dt+1εdW2(t).\begin{cases}\mathrm{d}x^{\varepsilon}(t)=\big{[}-(y^{\varepsilon}(t))^{3}+\sin(2\pi t)\big{]}\mathrm{d}t+y^{\varepsilon}(t)\mathrm{d}W^{1}(t),\\ \mathrm{d}y^{\varepsilon}(t)=-\displaystyle{\frac{1}{\varepsilon}}(y^{\varepsilon}(t)-x^{\varepsilon}(t))\mathrm{d}t+\frac{1}{\sqrt{\varepsilon}}\mathrm{d}W^{2}(t).\end{cases} (7.6)

Its averaged equation is

dx¯(t)=(x¯3(t)32x¯(t))dt+x¯2(t)+12dW1(t).\displaystyle\mathrm{d}\bar{x}(t)=\Big{(}-\bar{x}^{3}(t)-\frac{3}{2}\bar{x}(t)\Big{)}\mathrm{d}t+\sqrt{\bar{x}^{2}(t)+\frac{1}{2}}\mathrm{d}W^{1}(t). (7.7)

In [29], using EM scheme as macro-solver to obtain the macro dynamics numerically for nonlinear SFSDE (7.6), which seems that the EM scheme works perfectly for SFSDEs. However, we want to state two facts:

  • (1)

    In 2011, Hutzenthaler et al. [21] had pointed out that the EM method is not applicable any more for a large class of super-linear SDEs, such as the Ginzburg-Landau equation (7.3) we listed in Example 7.1. As a consequence, it is not appropriate to use the EM scheme as macro solver if the corresponding averaged equation is super-linear. The explicit MTEM scheme defined by (3.5) is suitable for a wide class of SFSDEs.

  • (2)

    In [30], an example is provided to illustrate that the strong averaging principle does not hold when the slow diffusion coefficient of original system depends on fast component. Thus we only consider the case that the slow diffusion coefficient is independent of the fast component.

Example 7.2.

Consider the following SFSDE

{dxε(t)=[xε(t)xε(t)(yε(t))2+yε(t)]dt+xε(t)dW1(t),dyε(t)=1ε(xε(t)4yε(t))dt+1ε(xε(t)+yε(t))dW2(t)\begin{cases}\mathrm{d}x^{\varepsilon}(t)=\big{[}x^{\varepsilon}(t)-x^{\varepsilon}(t)(y^{\varepsilon}(t))^{2}+y^{\varepsilon}(t)\big{]}\mathrm{d}t+x^{\varepsilon}(t)\mathrm{d}W^{1}(t),\\ dy^{\varepsilon}(t)=\displaystyle{\frac{1}{\varepsilon}}(x^{\varepsilon}(t)-4y^{\varepsilon}(t))\mathrm{d}t+\frac{1}{\sqrt{\varepsilon}}(x^{\varepsilon}(t)+y^{\varepsilon}(t))dW^{2}(t)\end{cases} (7.8)

with the initial value (x0,y0)=(1,1)(x_{0},y_{0})=(1,1). Assume that

b(x,y)=xx2y+y,σ(x)=x,f(x,y)=x4y,g(x,y)=x+y.\displaystyle b(x,y)=x-x^{2}y+y,~{}~{}\sigma(x)=x,~{}~{}f(x,y)=x-4y,~{}~{}g(x,y)=x+y. (7.9)

It can be verified that (𝐒𝟏)({\bf S1})-(𝐒𝟓)({\bf S5}) and (𝐅𝟏)({\bf F1})-(𝐅𝟑)({\bf F3}) hold with θ1=θ2=2,θ3=θ4=2\theta_{1}=\theta_{2}=2,\theta_{3}=\theta_{4}=2 and 6<k<96<k<9. Then using lemma 2.2 yields that the strong convergence between xε(t)x^{\varepsilon}(t) and the averaged equation x¯(t)\bar{x}(t) in ppth (0<p<k)(0<p<k) moment. Although the averaged equation provides a substantial simplification for SFSDE (7.8), the closed form of the averaged equation is unavailable. Then classical numerical approximation techniques can’t be used directly. This is where MTEM scheme defined by (3.5) comes in.

First, by (3.1) we take φ(u)=1+u2,u1\varphi(u)=1+u^{2},u\geq 1. Then for any Δ1(0,1],Δ2(0,1]\Delta_{1}\in(0,1],\Delta_{2}\in(0,1] and integer M>0M>0, define the MTEM scheme for (7.8): for any n0n\geq 0,

{X0=1,Xn=(|Xn|(2Δ1121)12)Xn|Xn|,Y0Xn=1,Ym+1Xn=YmXn+(Xn4YmXn)Δ2+YmXnΔWn,m2,m=0,1,,M1,BM(Xn,YXn)=Xn+1Mm=1M(Xn(YmXn)2+YmXn),Xn+1=Xn+BM(Xn,YXn)Δ1+XnΔWn1.\begin{cases}X_{0}=1,X^{*}_{n}=\Big{(}|X_{n}|\wedge\big{(}2\Delta_{1}^{-\frac{1}{2}}-1\big{)}^{\frac{1}{2}}\Big{)}\frac{X_{n}}{|X_{n}|},Y^{X^{*}_{n}}_{0}=1,\ \\ Y^{X^{*}_{n}}_{m+1}=Y^{X^{*}_{n}}_{m}+(X^{*}_{n}-4Y^{X^{*}_{n}}_{m})\Delta_{2}+Y^{X^{*}_{n}}_{m}\Delta W^{2}_{n,m},~{}~{}~{}m=0,1,\dots,M-1,\\ B_{M}(X^{*}_{n},Y^{X^{*}_{n}})=X^{*}_{n}+\displaystyle{\frac{1}{M}}\sum_{m=1}^{M}\big{(}-X^{*}_{n}\big{(}Y^{X^{*}_{n}}_{m}\big{)}^{2}+Y^{X^{*}_{n}}_{m}\big{)},\\ X_{n+1}=X_{n}+B_{M}(X^{*}_{n},Y^{X^{*}_{n}})\Delta_{1}+X_{n}\Delta W^{1}_{n}.\end{cases}\vspace{-1mm}\!\!\!

Therefore, by Theorem 5.2, using this scheme we can approximate the slow component of SFSDE (7.8) in the ppth (0<p<k/2)(0<p<k/2) moment. In order to test the efficiency of the scheme, we carry out numerical experiments by implementing (7.2) using MATLAB. Let (Δ1,Δ2,M)=(210,28,216)(\Delta_{1},\Delta_{2},M)=(2^{-10},2^{-8},2^{16}). The Figure 3 depicts the five sample paths of |X(t)||X(t)| and sample mean value of 100 sample points in different time interval [0,T][0,T], where T=5T=5 (left), T=10T=10 (middle) and T=20T=20(right), respectively.

Refer to caption
Figure 3: Five sample paths and sample mean value of |X(t)||X(t)| for 100 sample points in different time intervals

8 Concluding remarks

This paper developed the explicit numerical scheme for a class of super-linear SFSDEs in which the slow drift coefficient grows polynomially. An explicit multiscale numerical scheme called MTEM is proposed by a truncation device. The strong convergence and error estimate of the numerical solutions are provided under weak restrictions. However, more delicate questions, such as the super-linearity of fast component, haven’t been solved yet. They will be considered in our future work.

References

  • [1] J. Bao, J. Shao, C. Yuan, Approximation of invariant measures for regime-switching diffusions, Potential Anal., 44 (2016), pp. 707-727.
  • [2] C.-E. Bréhier, Analysis of an HMM time-discretization scheme for a system of stochastic PDEs, SIAM J. Numer. Anal., 51 (2013), pp. 1185–1210.
  • [3] C.-E. Bréhier, Orders of convergence in the averaging principle for SPDEs: the case of a stochastically forced slow component, Stochastic Process. Appl., 130 (2020), pp. 3325–3368.
  • [4] 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), pp. 2843-2884.
  • [5] W. E, Principles of Multiscale Modeling, Cambridge University Press, Cambridge, 2011.
  • [6] W. E, B. Engquist, The heterogeneous multiscale methods, Commun. Math. Sci., 1 (2003), pp. 87–132.
  • [7] W. E, B. Engquist, X. Li, W. Ren, E. Vanden-Eijnden, Heterogeneous multiscale methods: a review, Commun. Comput. Phys., 2 (2007), pp. 367-450.
  • [8] W. E, D. Liu, E. Vanden-Eijnden, Analysis of multiscale methods for stochastic differential equations, Comm. Pure Appl. Math., 58 (2005), pp. 1544–1585.
  • [9] H. Fu, J. Liu, Strong convergence in stochastic averaging principle for two time-scales stochastic partial differential equations, J. Math. Anal. Appl., 384 (2011), pp. 70-86.
  • [10] H. Gao, Y. Shi, Averaging principle for a stochastic coupled fast-slow atmosphere-ocean model, J. Differential Equations, 298 (2021), pp. 248-297.
  • [11] D. Givon, Strong convergence rate for two-time-scale jump-diffusion stochastic differential systems, Multiscale Model. Simul., 6 (2007), pp. 577–594.
  • [12] D. Givon, I. G. Kevrekidis, Multiscale integration schemes for jump-diffusion systems, Multiscale Model. Simul., 7 (2008), pp. 495–516.
  • [13] D. Givon, I. G. Kevrekidis, R. Kupferman, Strong convergence of projective integration schemes for singularly perturbed stochastic differential systems, Commun. Math. Sci., 4 (2006), pp. 707–729.
  • [14] J. Glimm, D. H. Sharp, Multiscale science: A challenge for the twenty-first century, Advances in Mechanics, 28 (1998), pp. 545–551.
  • [15] J. Golec, Stochastic averaging principle for systems with pathwise uniqueness, Stochastic Anal. Appl., 13 (1995), pp. 307–322.
  • [16] J. Golec, G. Ladde, Averaging principle and systems of singularly perturbed stochastic differential equations, J. Math. Phys., 31 (1990), pp. 1116–1123.
  • [17] E. Harvey, V. Kirk, M. Wechselberger, J. Sneyd, Multiple timescales, mixed mode oscillations and canards in models of intracellular calcium dynamics, J. Nonlinear Sci., 21 (2011), pp. 639–683.
  • [18] D. J. Higham, X. Mao, A. M. Stuart, Strong convergence of Euler-type methods for nonlinear stochastic differential equations, SIAM J. Numer. Anal., 40 (2002), pp. 1041–1063.
  • [19] W. Hong, S. Li, W. Liu, Strong convergence rates in averaging principle for slow-fast McKean-Vlasov SPDEs, J. Differential Equations, 316 (2022), pp. 94-135.
  • [20] M. Hutzenthaler, A. Jentzen, Numerical approximations of stochastic differential equations with non-globally Lipschitz continuous coefficients, Mem. Amer. Math. Soc., 236 (2015), 99 pp.
  • [21] M. Hutzenthaler, A. Jentzen, P. E. Kloeden, Strong and weak divergence in finite time of Euler’s method for stochastic differential equations with non-globally Lipschitz continuous coefficients, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 467 (2011), pp. 1563–1576.
  • [22] M. Hutzenthaler, A. Jentzen, P. E. Kloeden, Strong convergence of an explicit numerical method for SDEs with nonglobally Lipschitz continuous coefficients, Ann. Appl. Probab., 22 (2012), pp. 1611–1641.
  • [23] R. Z. Khassminskii, On the principle of averaging the Itô’s stochastic differential equations, Kybernetika (Prague), 4 (1968), pp. 260–279.
  • [24] P. E. Kloeden, E. Platen, Numerical Solution of Stochastic Differential Equations, Springer, Berlin, 1992.
  • [25] C. Kuehn, Multiple Time Scale Dynamics, Springer, Cham, 2015.
  • [26] A. L. Lewis, Option Valuation under Stochastic Volatility, Finance Press, Newport Beach, CA, 2000.
  • [27] X. Li, X. Mao, H. Yang, Strong convergence and asymptotic stability of explicit numerical schemes for nonlinear stochastic differential equations, Math. Comp., 90 (2021), pp. 2827–2872.
  • [28] X. Li, X. Mao, G. Yin, Explicit numerical approximations for stochastic differential equations in finite and infinite horizons: truncation methods, convergence in ppth moment and stability, IMA J. Numer. Anal., 39 (2019), pp. 847–892.
  • [29] D. Liu, Analysis of multiscale methods for stochastic dynamical systems with multiple time scales, Multiscale Model. Simul., 8 (2010), pp. 944–964.
  • [30] D. Liu, Strong convergence of principle of averaging for multiscale stochastic dynamical systems, Commun. Math. Sci., 8 (2010), pp. 999–1020.
  • [31] W. Liu, X. Mao, Strong convergence of the stopped Euler-Maruyama method for nonlinear stochastic differential equations, Appl. Math. Comput., 223 (2013), pp. 389–400.
  • [32] W. Liu, M. Ro¨ckner\mathrm{R\ddot{o}ckner}, X. Sun, Y. Xie, Averaging principle for slow-fast stochastic differential equations with time dependent locally Lipschitz coefficients, J. Differential Equations, 268 (2020), pp. 2910–2948.
  • [33] X. Mao, Stochastic Differential Equations and Applications, second ed., Horwood Publishing Limited, Chichester, 2008.
  • [34] X. Mao, The truncated Euler-Maruyama method for stochastic differential equations, J. Comput. Appl. Math., 290 (2015), pp. 370–384.
  • [35] E. Papageorgiou, R. Sircar, Multiscale intensity models for single name credit derivatives, Appl. Math. Finance, 15 (2008), pp. 73–105.
  • [36] G. A. Pavliotis, A. M. Stuart, Multiscale Methods: Averaging and Homogenization, Springer, 2007.
  • [37] M. Röckner, X. Sun, Y. Xie, Strong convergence order for slow-fast McKean-Vlasov stochastic differential equations, Ann. Inst. Henri Poincaré Probab. Stat., 57 (2021), pp. 547-576.
  • [38] S. Sabanis, A note on tamed Euler approximations, Electron. Commun. Probab., 18 (2013), 10 pp.
  • [39] S. Sabanis, Euler approximations with varying coefficients: the case of superlinearly growing diffusion coefficients, Ann. Appl. Probab., 26 (2016), pp. 2083–2105.
  • [40] A. N. Shiryaev, Probability, second ed., Springer-Verlag, New York, 1996.
  • [41] E. Vanden-Eijnden, Numerical techniques for multi-scale dynamical systems with stochastic effects, Commun. Math. Sci., 1 (2003), pp. 385–391.
  • [42] C. Villani, Optimal transport, old and new, Springer, 2009.
  • [43] A. Y. Veretennikov, On the averaging principle for systems of stochastic differential equations, Math. USSR Sb., 69 (1991), pp. 271–284.
  • [44] A. Y. Veretennikov, On large deviations in the averaging principle for SDEs with “a full dependence”, Ann. Probab., 27 (1999), pp. 284-296.
  • [45] X. Wang, S. Gan, The tamed Milstein method for commutative stochastic differential equations with non-globally Lipschitz continuous coefficients, J. Difference Equ. Appl., 19 (2013), pp. 466–490.
  • [46] F. Wu, T. Tian, J. B. Rawlings, G. Yin, Approximate method for stochastic chemical kinetics with two-time scales by chemical langevin equations, J. Chem. Phys., 144 (2016): 174112.
  • [47] Y. Xu, J. Duan, W. Xu, An averaging principle for stochastic dynamical systems with Le´vy\mathrm{L\acute{e}vy} noise, Phys. D, 240 (2011), pp. 1395–1401.