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

Rate of convergence in the Smoluchowski-Kramers approximation for mean-field stochastic differential equations

Ta Cong Son University of Science, Vietnam National University, Hanoi [email protected] Dung Quang Le École Polytechnique, France [email protected]  and  Manh Hong Duong University of Birmingham, UK. [email protected].
Abstract.

In this paper we study a second-order mean-field stochastic differential systems describing the movement of a particle under the influence of a time-dependent force, a friction, a mean-field interaction and a space and time-dependent stochastic noise. Using techniques from Malliavin calculus, we establish explicit rates of convergence in the zero-mass limit (Smoluchowski-Kramers approximation) in the LpL^{p}-distances and in the total variation distance for the position process, the velocity process and a re-scaled velocity process to their corresponding limiting processes.

Key words and phrases:
Smoluchowski-Kramers approximation, Stochastic differential by mean-field, Total variation distance, Malliavin calculus
2010 Mathematics Subject Classification:
60G22, 60H07, 91G30

1. Introduction

In this paper, we are interested in the following second-order mean-field stochastic differential equations

{dXtα=Ytαdt,1αdYtα=[κYtαg(t,Xtα)γ(Ytα𝔼(Ytα))]dt+σ(t,Xtα)dWt,X0α=x0,Y0α=y0.\begin{cases}dX^{\alpha}_{t}=Y^{\alpha}_{t}\,dt,\\ \frac{1}{\alpha}dY^{\alpha}_{t}=[-\kappa Y^{\alpha}_{t}-g(t,X^{\alpha}_{t})-\gamma(Y^{\alpha}_{t}-\mathbb{E}(Y^{\alpha}_{t}))]\,dt+\sigma(t,X^{\alpha}_{t})\,dW_{t},\\ X^{\alpha}_{0}=x_{0},Y^{\alpha}_{0}=y_{0}.\end{cases} (1.1)

Here α,γ\alpha,\gamma and κ\kappa are positive constants, g(t,x):[0,T]g(t,x):[0,T]\rightarrow\mathbb{R} is a given function, x0,y0x_{0},y_{0}\in\mathbb{R} are given points in the real line, and (Wt)t0(W_{t})_{t\geq 0} is the standard one-dimensional Wiener process. The notation 𝔼\mathbb{E} denotes the expectation with respect to the probability measure of the underlying probability space in which the Wiener process is defined.

System (1.1) describes the movement of a particle at position (displacement) XtαX_{t}^{\alpha}\in\mathbb{R} and with velocity YtαY_{t}^{\alpha}\in\mathbb{R}, at time tt, under the influence of four different forces: an external, possibly time-dependent and non-potential, force g(t,Xtα)-g(t,X^{\alpha}_{t}); a friction κYtα-\kappa Y^{\alpha}_{t}; a (McKean-Vlasov type) mean-field interaction force γ(Ytα𝔼(Ytα))-\gamma(Y^{\alpha}_{t}-\mathbb{E}(Y^{\alpha}_{t})) (noting that here the mean-field term is acting on the velocity rather than the position) and a stochastic noise σ(t,Xtα)W˙t\sigma(t,X^{\alpha}_{t})\dot{W}_{t}. Physically, α\alpha is the inverse of the mass, κ\kappa is the friction coefficient and γ\gamma is the strength of the interaction. We use the superscript α\alpha in (1.1) to emphasize the dependence on α\alpha since in the subsequent analysis we are concerned with the asymptotic behaviour of (1.1) as α\alpha tends to ++\infty.

Under Assumptions 1.1 (see below) of this paper, system (1.1) can also be obtained as the mean-field (hydrodynamic) limit of the following interacting particle system as NN tends to ++\infty

{dXtα,i=Ytα,idt,dYtα,i=[ακYtα,iαg(t,Xtα,i)αγNj=1N(Ytα,iYtα,j)]dt+ασ(t,Xtα,i)dWti,X0α=x0,Y0α=y0,\begin{cases}dX^{\alpha,i}_{t}=Y^{\alpha,i}_{t}\,dt,\\ dY^{\alpha,i}_{t}=[-\alpha\kappa Y^{\alpha,i}_{t}-\alpha g(t,X^{\alpha,i}_{t})-\frac{\alpha\gamma}{N}\sum_{j=1}^{N}(Y^{\alpha,i}_{t}-Y^{\alpha,j}_{t})]\,dt+\alpha\sigma(t,X^{\alpha,i}_{t})\,dW^{i}_{t},\\ X^{\alpha}_{0}=x_{0},Y^{\alpha}_{0}=y_{0},\end{cases} (1.2)

where {Wi}i=1N\{W^{i}\}_{i=1}^{N} are independent one-dimensional Wiener processes. In fact, under Assumptions 1.1 the above interacting system satisfies the property of propagation of chaos, that is as NN tends to infinity, it behaves more and more like a system of independent particles, in which each particle evolves according to (1.1) where the interaction term in (1.2) is replaced by the expectation one. For a detailed account on the propagation of chaos phenomenon, we refer the reader to classical papers [Kac56, Szn91] and more recent papers [BGM10, Duo15, JW17] and references therein for degenerate diffusion systems like (1.1). The interacting particle system (1.2) and its mean-field limit (1.1) and more broadly systems of these types have been used extensively in biology, chemistry and statistical physics for the modelling of molecular dynamics, chemical reactions, flockings, social interactions, just to name a few, see for instance, the monographs [RF96, Pav14].

In this paper, we are interested in the zero-mass limit (as also known as the Smoluchowski-Kramers approximation) of (1.1), that is its asymptotic behaviour as α\alpha tends to ++\infty. By employing techniques from Malliavin calculus, we obtain explicitly rate of convergences, in LpL^{p}-distances and in total variation distances, for both the position and velocity processes.

1.1. Main results

Before stating our main results, we make the following assumptions.

Assumption 1.1.
  • (A)(A)

    The coefficients g,σ:[0,T]×g,\sigma:[0,T]\times\mathbb{R}\longrightarrow\mathbb{R} have linear growth, i.e. there exists K>0K>0 such that

    |g(t,x)|+|σ(t,x)|K(1+|x|)x,t[0,T].|g(t,x)|+|\sigma(t,x)|\leq K(1+|x|)\quad\forall x\in\mathbb{R},t\in[0,T].
  • (B)(B)

    The coefficients g,σ:[0,T]×g,\sigma:[0,T]\times\mathbb{R}\longrightarrow\mathbb{R} are Lipschitz, i.e. there exists L>0L>0 such that

    |g(t,x)g(t,y)|+|σ(t,x)σ(t,y)|L|xy|x,y,t[0,T].|g(t,x)-g(t,y)|+|\sigma(t,x)-\sigma(t,y)|\leq L|x-y|\quad\forall x,y\in\mathbb{R},t\in[0,T].
Assumption 1.2.

g(t,x),σ(t,x)g(t,x),\sigma(t,x) are twice differentiable in xx and the derivatives are bounded by some constant M>0\mathrm{M}>0.

Let F,GF,G be random variables, we denote by dTV(F,G)d_{TV}(F,G) the total variation distance between the laws of FF and GG, that is,

dTV(F,G)\displaystyle d_{TV}(F,G) =supA()|P(FA)P(GA)|\displaystyle=\sup\limits_{A\in\mathcal{B}(\mathbb{R})}|P(F\in A)-P(G\in A)|
=12sup{|ϕ(F)ϕ(G)|:ϕ:which is bounded by 1}.\displaystyle=\dfrac{1}{2}\sup\{|\phi(F)-\phi(G)|:\phi:\mathbb{R}\to\mathbb{R}\ \mbox{which is bounded by}\ 1\}.

Consider the following first-order stochastic differential equation, which will be the limiting system for the displacement process

(κ+γ)dXt=[g(t,Xt)γκ𝔼[g(t,Xt)]]dt+σ(t,Xt)dWt,X0=x0.(\kappa+\gamma)dX_{t}=-\Big{[}g(t,X_{t})-\dfrac{\gamma}{\kappa}\mathbb{E}[g(t,X_{t})]\Big{]}\,dt+\sigma(t,X_{t})dW_{t},\quad X_{0}=x_{0}\in\mathbb{R}. (1.3)

Our first main result provides an explicit rates of convergence for the displacement process.

Theorem 1.1 (Quantitative rates of convergence of the displacement process).

Under Assumptions 1.1 and 1.2, systems (1.1) and (1.3) have unique solutions and the following statements hold.

  1. (1)

    (rate of convergence in LpL^{p}-distances) For all p2p\geq 2, α1\alpha\geq 1 and t[0,T]t\in[0,T],

    𝔼[sup0st|XsαXs|p]C[(λ(t,α(κ+γ)))p2+(λ(t,ακ))p],\mathbb{E}\Big{[}\sup_{0\leq s\leq t}|X^{\alpha}_{s}-X_{s}|^{p}\Big{]}\leq C\left[\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{\frac{p}{2}}+\left(\lambda(t,\alpha\kappa)\right)^{p}\right],

    where λ(t,a)=(1/a)[1exp(at)]\lambda(t,a)=(1/a)[1-\exp(-at)] for t,a>0t,a>0 and CC is a positive constant depending on {x0,y0,κ,γ,K,L,p,T}\{x_{0},y_{0},\kappa,\gamma,K,L,p,T\} but not on α\alpha and tt.

  2. (2)

    (rate of convergence in the total variation distance). We further assume that |σ(t,x)|σ0>0|\sigma(t,x)|\geq\sigma_{0}>0 for all (t,x)[0,T]×.(t,x)\in[0,T]\times\mathbb{R}. Then, for each α1\alpha\geq 1 and t(0,T],t\in(0,T],

    dTV(Xtα,Xt)Ct1(λ(t,α(κ+γ))+(λ(t,ακ))2),d_{TV}(X^{\alpha}_{t},X_{t})\leq C\sqrt{t^{-1}(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})},

    where CC is a constant depending only on {x0,y0,σ0,κ,γ,K,L,M,T}\{x_{0},y_{0},\sigma_{0},\kappa,\gamma,K,L,\,M,T\} but not on α\alpha and tt. As a corollary, if |σ(t,x)|σ0>0|\sigma(t,x)|\geq\sigma_{0}>0 for all (t,x)[0,T]×(t,x)\in[0,T]\times\mathbb{R} then we have.

    dTV(Xtα,Xt)Ct1/2α1/2.d_{TV}(X^{\alpha}_{t},X_{t})\leq Ct^{-1/2}\alpha^{-1/2}.

Theorem 1.1 combines Theorem 3.1 (for the LpL^{p}-distances) and Theorem 3.2 (for the total variation distance) in Section 3.1.

We are also interested in the asymptotic behavior, when α\alpha\to\infty, of the velocity process YtαY^{\alpha}_{t} of (1.1) and of a re-scaled velocity process, Y~tα\tilde{Y}^{\alpha}_{t}, which is defined by

Y~tα:=1αYt/αα.\tilde{Y}^{\alpha}_{t}:=\frac{1}{\sqrt{\alpha}}Y^{\alpha}_{t/\alpha}.

The re-scaled process Y~tα\tilde{Y}^{\alpha}_{t} satisfies the following stochastic differential equation

{Y~tα=y0α(κ+γ)0tY~sα𝑑s1α0tg(sα,Xsαα)𝑑sγ0t𝔼(Y~sα)𝑑s+0tσ(sα,Xsαα)𝑑W~sX0α=x0,\begin{cases}\tilde{Y}^{\alpha}_{t}=\dfrac{y_{0}}{\sqrt{\alpha}}-(\kappa+\gamma)\int_{0}^{t}\tilde{Y}^{\alpha}_{s}ds-\dfrac{1}{\sqrt{\alpha}}\int_{0}^{t}g(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})ds-\gamma\int_{0}^{t}\mathbb{E}(\tilde{Y}^{\alpha}_{s})ds+\int_{0}^{t}\sigma(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})d\tilde{W}_{s}\\ X^{\alpha}_{0}=x_{0},\end{cases} (1.4)

where W~t:=αWt/α\tilde{W}_{t}:=\sqrt{\alpha}W_{t/\alpha} is a rescaled Brownian process.

Now we consider the following stochastic differential equation, which will be the limiting process of the rescaled velocity process

{dY~t=(κ+γ)dY~t+σ(0,x0)dW~t,Y~(0)=0.\begin{cases}d\tilde{Y}_{t}=-(\kappa+\gamma)d\tilde{Y}_{t}+\sigma(0,x_{0})d\tilde{W}_{t},\\ \tilde{Y}(0)=0.\end{cases} (1.5)

We now describe our result for the rescaled velocity process first since for this process we also work with a general setting where both gg and σ\sigma can depend on both spatial and temporal variables. We only assume additionally the following condition.

Assumption 1.3.
|σ(t,x)σ(s,y)|L(|ts|+|xy|)x,y,t,s[0,T].|\sigma(t,x)-\sigma(s,y)|\leq L(|t-s|+|x-y|)\quad\forall x,y\in\mathbb{R},t,s\in[0,T].

In the next theorem, we provide explicit rates of convergence, both in LpL^{p}-distances and in the total variation distance, for the rescaled velocity process.

Theorem 1.2 (Quantitative rates of convergence for the rescaled velocity processes).

Under Assumptions 1.1 and 1.3 the following hold.

  1. (1)

    (rate of convergence in LpL^{p}-distance for the rescaled velocity process) For all p2p\geq 2 and α1\alpha\geq 1,

    𝔼[sup0tT|Y~tαY~t|p]Cαp/2,\mathbb{E}\left[\sup_{0\leq t\leq T}|\tilde{Y}^{\alpha}_{t}-\tilde{Y}_{t}|^{p}\right]\leq\dfrac{C}{\alpha^{p/2}},

    where CC is a positive constant depending on pp and other parameters but not on α\alpha.

  2. (2)

    (rate of convergence in the total variation distance for the rescaled velocity process) Assume that Assumptions 1.1 and 1.3 hold and |σ(0,x0)|>0|\sigma(0,x_{0})|>0 for all (t,x)[0,T]×.(t,x)\in[0,T]\times\mathbb{R}. Then, for each α1\alpha\geq 1 and t(0,T],t\in(0,T],

    dTV(Y~tα,Y~t)C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}(\tilde{Y}^{\alpha}_{t},\tilde{Y}_{t})\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

    where CC is a constant depending only on {x0,y0,κ,γ,K,L,p,T,σ(0,x0)}\{x_{0},y_{0},\kappa,\gamma,K,L,p,T,\sigma(0,x_{0})\}.

Theorem 1.2 summarizes Theorem 3.3 (for the LpL^{p}-distances) and Theorem 3.4 (for the total variation distance) in Section 3.2.

When g(t,x)=g(x)g(t,x)=g(x) and σ(t,x)=δ\sigma(t,x)=\delta, [Nar94, Theorem 2.3] shows that the velocity process YtαY^{\alpha}_{t} converges to the normal distribution as α\alpha\to\infty. The third aim of this paper is to generalize this result to a more general setting where gg depends on both xx and tt while σ\sigma depends only on tt, i.e. σ(t,x)=σ(t)\sigma(t,x)=\sigma(t), obtaining rates of convergence in the total variation distance. The following theorem is the content of Theorem 3.5 in Section 3.2.

Theorem 1.3 (Quantitative rates of convergence for the velocity processes).

Under Assumptions 1.1 the following hold. Assume additionally that σ(t)\sigma(t) is continuously differentiable on [0,T][0,T] and that σ(t)0\sigma(t)\not=0 for each t(0,T]t\in(0,T]. Let NN be a normal random variable with mean 0 and variance σ2(t)2(κ+γ)\dfrac{\sigma^{2}(t)}{2(\kappa+\gamma)}, Then, for each α1\alpha\geq 1 and t(0,T]t\in(0,T]

dTV(Ytαα,N)C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},N\right)\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where C>0C>0 is a constant not depending on α\alpha and tt.

Theorem 1.3 is Theorem 3.5 in Section 3.2. We emphasize that in the main theorems, to obtain the existence and uniqueness as well as the rate of convergence in LpL^{p}-distances we only use Assumptions 1.1. Assumptions 1.2 and 1.3 are needed to employ techniques from Malliavin calculus, in particular to derive estimates for the Malliavin derivatives.

Corollary 1.1 (Rate of convergence in Wasserstein distance for the laws of the displacement and velocity processes).

Let μ\mu and ν\nu be two probability measures with finite second moments, then the pp-Wasserstein distance, Wp(μ,ν)W_{p}(\mu,\nu), between them can be defined by

Wp(μ,ν)=(inf{𝔼[|XY|p]:Xμ,Yν})1/p.W_{p}(\mu,\nu)=\Big{(}\inf\{\mathbb{E}\big{[}|X-Y|^{p}\big{]}:X\sim\mu,Y\sim\nu\}\Big{)}^{1/p}.

Using this formulation, as a direct consequence of our main results, we also obtain explicit rates of convergence in pp-Wasserstein distances for the laws of the displacement and the rescaled velocity processes to the corresponding limiting ones

supt[0,T]Wpp(law(Xtα),law(Xt))Cαp/2,\displaystyle\sup_{t\in[0,T]}W_{p}^{p}(\mathrm{law}(X^{\alpha}_{t}),\mathrm{law}(X_{t}))\leq\frac{C}{\alpha^{p/2}},
supt[0,T]Wpp(law(Y~tα),law(Y~t))Cαp/2.\displaystyle\sup_{t\in[0,T]}W_{p}^{p}(\mathrm{law}(\tilde{Y}^{\alpha}_{t}),\mathrm{law}(\tilde{Y}_{t}))\leq\frac{C}{\alpha^{p/2}}.

1.2. Comparison with existing literature and future work

The zero-mass limit of second order differential equations has been studied intensively in the literature. In the seminal work [Kra40], Kramers formally discusses this problem, in the context of applications to chemical reactions, for the classical underdamped Langevin dynamics, which corresponds to (1.1) with g=Vg=-\nabla V (a gradient potential force), γ=0\gamma=0 (no interaction force) and a constant diffusion coefficient. Due to this seminal work, this limit has become known in the literature as the Smoluchowski-Kramers approximation. Nelson rigorously shows that, under suitable rescaling, the solution to the Langevin equation converges almost surely to the solution of (3.1) with ψ=0\psi=0  [Nel67]. Since then various generalizations and related results have been proved using different approaches such as stochastic methods, asymptotic expansions and variational techniques, see for instance [Nar91b, Nar91a, Nar94, Fre04, CF06, HVW12, DLPS17, DLP+18, NN20]. The most relevant papers to the present one include [Nar91b, Nar91a, Nar94, DLPS17, NN20]. The main novelty of the present paper lies in the fact that we consider interacting (mean-field) systems allowing time-dependent external forces and diffusion coefficients, and providing explicit rates of convergence in both LpL^{p}-distances and total variation distances for both displacement and velocity processes. Existing papers lack at least one of these features. More specifically,

Papers that consider mean-field (interaction) systems. The papers [Nar91b, Nar91a, Nar94, DLPS17] consider second order mean-field stochastic differential equations establishing the zero-mass limit, but they require much more stringent conditions that g(t,x)=g(x)g(t,x)=g(x) (time-independent force) and σ(t,x)=δ\sigma(t,x)=\delta (constant diffusivity). On top of that, they do not provide a rate of convergence. Furthermore, our approach using Malliavin calculus is also different: Narita’s papers use direct arguments while [DLPS17] employs variational methods based on Gamma-convergence and large deviation principle.

Papers that provide a rate of convergence. The papers [NN20, DLP+18] provide a rate of convergence but only consider non-interacting systems (also using different measurements). Like our paper, [NN20] also utilizes techniques from Malliavin calculus, but [DLP+18] uses a completely different variational method. The recent paper [CT22], which studies the kinetic Vlasov-Fokker-Planck equation, is particularly interesting since it considers both interacting systems and provides a rate of convergence, but this paper is different to ours in a couple of aspects. First, the interaction force is acting on the position instead of the velocity; second, it works on the Fokker Planck equations and obtains a rate of convergence in Wasserstein distance while we work on the stochastic differential equations and obtain error quantifications in both LpL^{p}-distances and total variation distances; third, as mentioned, we use Malliavin calculus while [CT22] applied variational techniques like in [DLPS17, DLP+18]. We also mention the paper [Ta20], which provides similar rate of convergence to ours but it consider non mean-field stochastic differential equations driven by fractional Brownian motions.

Future work. The Lipschitz boundedness and differentiability Assumptions 1.1-1.2-1.3 are standard, but rather restricted since they do not cover some physically interesting interacting singular, such as Coloumb or Newton, forces. It would be interesting and challenging to extend our work to non-Lipschtizian and singular coefficients. Initial attempts in this direction for related models exist in the literature, see [Bre09] for non-Lipschitzian coefficients and recent papers [XY22, CT22] for singular forces. Another interesting problem for future work is to study the Kramers-Smoluchowski approximation for the NN-particle system (1.2) obtaining a rate of convergence that is independent of NN.

1.3. Overview of the proofs

To prove the main theorems for the general setting, with time-dependent coefficients, and obtain LpL^{p}-distances and total variations distances for the position and velocity processes, several technical improvements have been carried out.

On existence and uniqueness. Under Assumptions 1.1, the existence and uniqueness, as well as the boundedness of the moments, of the second-order system (1.1) and the limiting first-order one (1.3) are standard results following Hölder’s and the Burkholder-Davis-Gundy inequalities.

On rate of convergence in LpL^{p}-distances. Combining the mentioned inequalities and known estimates from [Nar91b] we can directly estimate 𝔼[sup0st|XsαXs|p]\mathbb{E}\Big{[}\sup_{0\leq s\leq t}|X^{\alpha}_{s}-X_{s}|^{p}\Big{]} and 𝔼[sup0tT|Y~tαY~t|p]\mathbb{E}\left[\sup_{0\leq t\leq T}|\tilde{Y}^{\alpha}_{t}-\tilde{Y}_{t}|^{p}\right] and obtain the rate of convergences in LpL^{p}-distances, proving parts (1) of both theorems.

On rate of convergence in total-variation distances. The Malliavin differentiablity of the processes is followed from similar arguments as in [Nua06]. Obtaining the rate of convergence in total variation distances is the most technically challenging. Lemma 2.1, which provides an upper bound estimate for the total variation between two random variables in terms of their Malliavin derivatives, is the key in our analysis. This lemma enables us to obtain the desired rates of convergence by estimating the corresponding quantities appearing in the right-hand side of Lemma 2.1.

1.4. Organization of the paper

The rest of of the paper is organized as follows. In Section 2, we give an overview of some elements of Malliavin calculus and mean-field stochastic differential equations. The proofs of the main theorems are given in Section 3.

2. Preliminaries

In this section, we provide some basic and directly relevant knowledge on the Malliavin calculus and mean-field stochastic differential equations.

2.1. Malliavin calculus

Let us recall some elements of stochastic calculus of variations (for more details see [Nua06]). We suppose that (Wt)t[0,T](W_{t})_{t\in[0,T]} is defined on a complete probability space (Ω,,𝔽,P)(\Omega,\mathcal{F},\mathbb{F},P), where 𝔽=(t)t[0,T]\mathbb{F}=(\mathcal{F}_{t})_{t\in[0,T]} is a natural filtration generated by the Brownian motion W.W. For hL2[0,T]:=,h\in L^{2}[0,T]:=\mathcal{H}, we denote by W(h)W(h) the Wiener integral

W(h)=0Th(t)𝑑Wt.W(h)=\int\limits_{0}^{T}h(t)dW_{t}.

Let 𝒮\mathcal{S} denote the dense subset of L2(Ω,,P):=L2(Ω)L^{2}(\Omega,\mathcal{F},P):=L^{2}(\Omega) consisting of smooth random variables of the form

F=f(W(h1),,W(hn)),F=f(W(h_{1}),...,W(h_{n})), (2.1)

where n,fCb(n),h1,,hn.n\in\mathbb{N},f\in C_{b}^{\infty}(\mathbb{R}^{n}),h_{1},...,h_{n}\in\mathcal{H}. If FF has the form (2.1), we define its Malliavin derivative as the process DF:={DtF,t[0,T]}DF:=\{D_{t}F,t\in[0,T]\} given by

DtF=k=1nfxk(W(h1),,W(hn))hk(t).D_{t}F=\sum\limits_{k=1}^{n}\frac{\partial f}{\partial x_{k}}(W(h_{1}),...,W(h_{n}))h_{k}(t).

More generally, for each k1k\geq 1 we can define the iterated derivative operator on a cylindrical random variable by setting

Dt1,,tkkF=Dt1DtkF.D^{k}_{t_{1},...,t_{k}}F=D_{t_{1}}...D_{t_{k}}F.

For any p,k1,p,k\geq 1, we shall denote by 𝔻k,p\mathbb{D}^{k,p} the closure of 𝒮\mathcal{S} with respect to the norm

Fk,pp:=𝔼[|F|p]+𝔼[0T|Dt11F|p𝑑t1]++𝔼[0T0T|Dt1,,tkkF|p𝑑t1𝑑tk].\|F\|^{p}_{k,p}:=\mathbb{E}\big{[}|F|^{p}\big{]}+\mathbb{E}\bigg{[}\int_{0}^{T}|D^{1}_{t_{1}}F|^{p}dt_{1}\bigg{]}+...+\mathbb{E}\bigg{[}\int_{0}^{T}...\int_{0}^{T}|D^{k}_{t_{1},...,t_{k}}F|^{p}dt_{1}...dt_{k}\bigg{]}.

A random variable FF is said to be Malliavin differentiable if it belongs to 𝔻1,2.\mathbb{D}^{1,2}.

An important operator in the Malliavin’s calculus theory is the divergence operator δ\delta, which is the adjoint of the derivative operator DD. The domain of δ\delta is the set of all functions uL2(Ω,)u\in L^{2}(\Omega,\mathcal{H}) such that

𝔼[|DF,u|]C(u)FL2(Ω),\mathbb{E}\big{[}|\langle DF,u\rangle_{\mathcal{H}}|\big{]}\leq C(u)\|F\|_{L^{2}(\Omega)},

where C(u)C(u) is some positive constant depending on uu. In particular, if uDom(δ)u\in\mathrm{Dom}(\delta), then δ(u)\delta(u) is characterized by the following duality relationship

𝔼[DF,u]=𝔼[Fδ(u)].\mathbb{E}\big{[}\langle DF,u\rangle_{\mathcal{H}}\big{]}=\mathbb{E}[F\delta(u)].

The following lemma provides an upper bound on the total variation distance between two random variables in terms of their Malliavin derivatives. This lemma will play an important role in the analysis of the present paper.

Lemma 2.1.

Let F1𝔻2,2F_{1}\in\mathbb{D}^{2,2} be such that DF1>0a.s.\|DF_{1}\|_{\mathcal{H}}>0\,\,a.s. Then, for any random variable F2𝔻1,2F_{2}\in\mathbb{D}^{1,2} we have

dTV(F1,F2)F1F21,2[3(𝔼D2F14)1/4(𝔼DF18)1/4+2(𝔼DF12)1/2],\displaystyle d_{TV}(F_{1},F_{2})\leq\|F_{1}-F_{2}\|_{1,2}\left[3\left(\mathbb{E}\|D^{2}F_{1}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|DF_{1}\|_{\mathcal{H}}^{-8}\right)^{1/4}+2\left(\mathbb{E}\|DF_{1}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right], (2.2)

provided that the expectations exist.

Proof.

From [Ta20, Lemma 2.1] we have

dTV(F1,F2)F1F21,2[(𝔼δ(DF1DF12)2)1/2+(𝔼DF12)1/2].\displaystyle d_{TV}(F_{1},F_{2})\leq\|F_{1}-F_{2}\|_{1,2}\left[\left(\mathbb{E}\delta\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)^{2}\right)^{1/2}+\left(\mathbb{E}\|DF_{1}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right]. (2.3)

Now using [Nua06, Proposition 1.3.1], we get

𝔼δ(DF1DF12)2\displaystyle\mathbb{E}\delta\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)^{2} 𝔼DF1DF122+𝔼D(DF1DF12)2\displaystyle\leq\mathbb{E}\left\|\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right\|^{2}_{\mathcal{H}}+\mathbb{E}\left\|D\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)\right\|_{\mathcal{H}\bigotimes\mathcal{H}}^{2}
=𝔼DF12+𝔼D(DF1DF12)2.\displaystyle=\mathbb{E}\left\|DF_{1}\right\|^{-2}_{\mathcal{H}}+\mathbb{E}\left\|D\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)\right\|_{\mathcal{H}\bigotimes\mathcal{H}}^{2}. (2.4)

Moreover, observing that

D(DF1DF12)=D2F1DF122D2F1,DF1DF1DF14,D\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)=\dfrac{D^{2}F_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}-2\dfrac{\langle D^{2}F_{1},DF_{1}\bigotimes DF_{1}\rangle_{\mathcal{H}\bigotimes\mathcal{H}}}{\|DF_{1}\|_{\mathcal{H}}^{4}},

which implies that

D(DF1DF12)3D2F1DF12.\left\|D\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)\right\|_{\mathcal{H}\bigotimes\mathcal{H}}\leq\dfrac{3\|D^{2}F_{1}\|_{\mathcal{H}\bigotimes\mathcal{H}}}{\|DF_{1}\|_{\mathcal{H}}^{2}}. (2.5)

Substituting the inequality (2.5) into (2.1) and using Hölder’s inequality, one can derive that

𝔼δ(DF1DF12)2\displaystyle\mathbb{E}\delta\left(\dfrac{DF_{1}}{\|DF_{1}\|_{\mathcal{H}}^{2}}\right)^{2} 𝔼DF12+9𝔼(D2F12DF14)\displaystyle\leq\mathbb{E}\left\|DF_{1}\right\|^{-2}_{\mathcal{H}}+9\mathbb{E}\left(\dfrac{\|D^{2}F_{1}\|^{2}_{\mathcal{H}\bigotimes\mathcal{H}}}{\|DF_{1}\|_{\mathcal{H}}^{4}}\right)
𝔼DF12+9(𝔼D2F14)1/2(𝔼DF18)1/2.\displaystyle\leq\mathbb{E}\left\|DF_{1}\right\|^{-2}_{\mathcal{H}}+9\left(\mathbb{E}\|D^{2}F_{1}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/2}\left(\mathbb{E}\|DF_{1}\|_{\mathcal{H}}^{-8}\right)^{1/2}.

Finally, substituting the above estimate back into (2.3) and using the fundamental inequality (a+b)1/2a1/2+b1/2(a+b)^{1/2}\leq a^{1/2}+b^{1/2} for all a,b0a,b\geq 0, we obtain (2.2), which completes the proof of this lemma. ∎

2.2. Mean-field stochastic differential equations

Let (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) be a probability space with an increasing family {t;t0}\{\mathcal{F}_{t};t\geq 0\} of sub-σ\sigma-algebras of \mathcal{F} and let {Wt;t0}\{W_{t};t\geq 0\} be a one-dimensional Brownian motion process adapted to t\mathcal{F}_{t}.

The following lemma provides equivalent formulations of (1.1) and (1.3) as stochastic integral equations.

Lemma 2.2.

Equations (1.1) and (1.3) are, respectively, equivalent to the following equations

Xtα\displaystyle X^{\alpha}_{t} =x01κ+γ0tg(s,Xsα)𝑑sγκ+γ0tGα(s)𝑑s+1κ+γ0tσ(s,Xsα)𝑑Ws\displaystyle=x_{0}-\dfrac{1}{\kappa+\gamma}\int_{0}^{t}g(s,X^{\alpha}_{s})\ ds-\dfrac{\gamma}{\kappa+\gamma}\int_{0}^{t}G^{\alpha}(s)\ ds+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\sigma(s,X^{\alpha}_{s})dW_{s}
+I0α(t)+I1α(t)I2α(t)I3α(t)I4α(t),\displaystyle\qquad+I_{0}^{\alpha}(t)+I_{1}^{\alpha}(t)-I_{2}^{\alpha}(t)-I_{3}^{\alpha}(t)-I_{4}^{\alpha}(t), (2.6)
Xt\displaystyle X_{t} =x01κ+γ0tg(s,Xs)𝑑sγκ+γ0tG(s)𝑑s+1κ+γ0tσ(s,Xs)𝑑Ws,\displaystyle=x_{0}-\dfrac{1}{\kappa+\gamma}\int_{0}^{t}g(s,X_{s})\ ds-\dfrac{\gamma}{\kappa+\gamma}\int_{0}^{t}G(s)\ ds+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\sigma(s,X_{s})dW_{s}, (2.7)

where Gα(t)=𝔼[g(t,Xtα)]/κG^{\alpha}(t)=\mathbb{E}[g(t,X^{\alpha}_{t})]/\kappa, G(t)=𝔼[g(t,Xt)]/κG(t)=\mathbb{E}[g(t,X_{t})]/\kappa, and

I0α(t)=y0λ(t;α(κ+γ)),\displaystyle I_{0}^{\alpha}(t)=y_{0}\lambda(t;\alpha(\kappa+\gamma)),
I1α(t)=1κ+γ0texp[α(κ+γ)(ut)]g(u,Xuα)𝑑u,\displaystyle I_{1}^{\alpha}(t)=\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(u-t)]}g(u,X^{\alpha}_{u})du,
I2α(t)=γκ+γ0texp[α(κ+γ)(ut)]nα(u)𝑑u,\displaystyle I_{2}^{\alpha}(t)=\dfrac{\gamma}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(u-t)]}n^{\alpha}(u)\ du,
I3α(t)=1κ+γ0texp[α(κ+γ)(ut)]σ(u,Xuα)𝑑Wu,\displaystyle I_{3}^{\alpha}(t)=\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(u-t)]}\sigma(u,X^{\alpha}_{u})dW_{u},
I4α(t)=γκ+γ(y0λ(t;ακ)+0texp[ακ(ut)]Gα(u)𝑑u),\displaystyle I_{4}^{\alpha}(t)=\dfrac{\gamma}{\kappa+\gamma}\left(y_{0}\lambda(t;\alpha\kappa)+\int_{0}^{t}\exp{[\alpha\kappa(u-t)]}G^{\alpha}(u)du\right),

where λ(t;a)=(1/a)(1exp[at]),nα(t)=E[Ytα]\lambda(t;a)=(1/a)(1-\exp{[-at]}),n^{\alpha}(t)=E[Y^{\alpha}_{t}].

Proof.

Firstly, we can rewrite the second equation of (1.1) as follows

dYtα=[ακYtααg(t,Xtα)αγ(Ytαnα(t))]dt+ασ(t,Xtα)dWtwith Y0α=y0.\displaystyle dY^{\alpha}_{t}=[-\alpha\kappa Y^{\alpha}_{t}-\alpha g(t,X^{\alpha}_{t})-\alpha\gamma(Y^{\alpha}_{t}-n^{\alpha}(t))]dt+\alpha\sigma(t,X^{\alpha}_{t})dW_{t}\ \mbox{with }\ Y^{\alpha}_{0}=y_{0}.

Using Itô formula, we have the following expression

d(exp[α(κ+γ)t]Ytα)\displaystyle d(\exp{[\alpha(\kappa+\gamma)t]}Y^{\alpha}_{t}) =αexp[α(κ+γ)t]g(t,Xtα)dt+αγexp[α(κ+γ)t]nα(t)dt\displaystyle=-\alpha\exp{[\alpha(\kappa+\gamma)t]}g(t,X_{t}^{\alpha})dt+\alpha\gamma\exp{[\alpha(\kappa+\gamma)t]}n^{\alpha}(t)dt
+αexp[α(κ+γ)t]σ(t,Xtα)dWt,\displaystyle\qquad+\alpha\exp{[\alpha(\kappa+\gamma)t]}\sigma(t,X_{t}^{\alpha})dW_{t},

which implies

Ytα\displaystyle Y^{\alpha}_{t} =y0exp[α(κ+γ)t]α0texp[α(κ+γ)(st)]g(s,Xsα)𝑑s\displaystyle=y_{0}\exp{[-\alpha(\kappa+\gamma)t]}-\alpha\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}g(s,X^{\alpha}_{s})ds
+αγ0texp[α(κ+γ)(st)]nα(s)𝑑s+α0texp[α(κ+γ)(st)]σ(s,Xsα)𝑑Ws.\displaystyle\qquad+\alpha\gamma\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}n^{\alpha}(s)\ ds+\alpha\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}\sigma(s,X^{\alpha}_{s})dW_{s}. (2.8)

Secondly, substituting this equation into the first equation of (1.1) we get

Xtα\displaystyle X^{\alpha}_{t} =x0+y00texp[α(κ+γ)s]𝑑sα0t0sexp[α(κ+γ)(us)]g(u,Xuα)𝑑u𝑑s\displaystyle=x_{0}+y_{0}\int_{0}^{t}\exp{[-\alpha(\kappa+\gamma)s]}ds-\alpha\int_{0}^{t}\int_{0}^{s}\exp{[\alpha(\kappa+\gamma)(u-s)]}g(u,X^{\alpha}_{u})duds
+αγ0t0sexp[α(κ+γ)(us)]nα(u)𝑑u𝑑s+α0t0sexp[α(κ+γ)(us)]σ(u,Xuα)𝑑Wu𝑑s.\displaystyle+\alpha\gamma\int_{0}^{t}\int_{0}^{s}\exp{[\alpha(\kappa+\gamma)(u-s)]}n^{\alpha}(u)\ duds+\alpha\int_{0}^{t}\int_{0}^{s}\exp{[\alpha(\kappa+\gamma)(u-s)]}\sigma(u,X^{\alpha}_{u})dW_{u}ds.

Now, we use integration by parts for the non-stochastic integral and Ito’s product rule for the stochastic integral to get

Xtα\displaystyle X^{\alpha}_{t} =x01κ+γ0tg(u,Xuα)𝑑u+γκ+γ0tnα(u)𝑑u+1κ+γ0tσ(s,Xα)𝑑W(s)\displaystyle=x_{0}-\dfrac{1}{\kappa+\gamma}\int_{0}^{t}g(u,X^{\alpha}_{u})\ du+\dfrac{\gamma}{\kappa+\gamma}\int_{0}^{t}n^{\alpha}(u)\ du+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\sigma(s,X^{\alpha})dW(s)
+I0α(t)+I1α(t)I2α(t)I3α(t),\displaystyle\ +I_{0}^{\alpha}(t)+I_{1}^{\alpha}(t)-I_{2}^{\alpha}(t)-I_{3}^{\alpha}(t), (2.9)

where the terms Iiα(t)I_{i}^{\alpha}(t) (i=0,1,2,3i=0,1,2,3) are defined in the statement of the lemma. On the other hand, from the second equation of (1.1) we have

ddtnα(t)=ακnα(t)ακGα(t)withnα(0)=y0.\dfrac{d}{dt}n^{\alpha}(t)=-\alpha\kappa n^{\alpha}(t)-\alpha\kappa G^{\alpha}(t)\ \ \mbox{with}\ n^{\alpha}(0)=y_{0}.

This implies that

nα(t)=y0exp[ακt]ακ0texp[ακ(ut)]Gα(u)𝑑u.\displaystyle n^{\alpha}(t)=y_{0}\exp{[-\alpha\kappa t]}-\alpha\kappa\int_{0}^{t}\exp{[\alpha\kappa(u-t)]}G^{\alpha}(u)du. (2.10)

Integrating this equation over the interval [0,t][0,t] and changing the order of integration in the double integral, we get

0tnα(s)𝑑s=y0λ(t,ακ)0tGα(s)𝑑s+0texp[ακ(st)]Gα(s)𝑑s.\displaystyle\int_{0}^{t}n^{\alpha}(s)\,ds=y_{0}\lambda(t,\alpha\kappa)-\int_{0}^{t}G^{\alpha}(s)ds+\int_{0}^{t}\exp[\alpha\kappa(s-t)]G^{\alpha}(s)ds. (2.11)

Substituting (2.11) back into (2.9) we obtain (2.6).

The proof for Equation (2.7) is similar. ∎

The existence and uniqueness of solutions to (2.6) and (2.7) under Assumptions 1.1 is stated in the [McK67].

3. Proof of the main results

In this section, we present the proofs of the main theorems 1.1, 1.2 and 1.3. We start with the displacement process (Theorems 3.1 and 3.2 give Theorem 1.1) in Section 3.1. Then in Section 3.2 we deal with the rescaled velocity process and the velocity process (Theorems 3.3 and 3.4 give Theorem 1.2 and Theorem 1.3 is Theorem 3.5).

3.1. Approximation of the displacement process

In this section, we give explicit bounds on LpL^{p}-distances and the total variation distance between the solution XtαX^{\alpha}_{t} of (1.1) and the solution XtX_{t} of (1.3). We will repeatedly use the following fundamental inequalities.

  1. (i)

    Minkowski’s inequality: for p1p\geq 1 and nn real numbers a1,,ana_{1},\ldots,a_{n}, we have

    |i=1nai|pnp1i=1n|ai|p.\Big{|}\sum_{i=1}^{n}a_{i}\Big{|}^{p}\leq n^{p-1}\sum_{i=1}^{n}|a_{i}|^{p}. (3.1)
  2. (ii)

    Hölder’s inequality: for p1p\geq 1, t>0t>0 and measurable functions ff we have

    (0t|f(s)|𝑑s)ptp10t|f(s)|p𝑑s.\Big{(}\int_{0}^{t}|f(s)|\,ds\Big{)}^{p}\leq t^{p-1}\int_{0}^{t}|f(s)|^{p}\,ds. (3.2)
  3. (iii)

    The Burkholder-Davis-Gundy (BDG) inequality for Brownian stochastic integrals, see for instance [SP12, Section 17.7]: for 0<p<0<p<\infty and fL2([0,t],Ω)f\in L^{2}([0,t],\Omega) we have

    𝔼[sups[0,t]|0sfr𝑑Wr|p]Cp𝔼[(0t|fs|2𝑑s)p/2],\mathbb{E}\left[\sup_{s\in[0,t]}\Big{|}\int_{0}^{s}f_{r}\,dW_{r}\Big{|}^{p}\right]\leq C_{p}\mathbb{E}\left[\Big{(}\int_{0}^{t}|f_{s}|^{2}\,ds\Big{)}^{p/2}\right], (3.3)

    where CpC_{p} is a positive constant depending only on pp.

Applying the BDG inequality (3.3) to solutions of (2.6) and (2.7) we obtain

𝔼[sups[0,t]|0sσ(r,Xrα)𝑑Wr|p]Cp𝔼[(0t|σ(s,Xsα)|2𝑑s)p/2],\displaystyle\mathbb{E}\left[\sup_{s\in[0,t]}\Big{|}\int_{0}^{s}\sigma(r,X^{\alpha}_{r})\,dW_{r}\Big{|}^{p}\right]\leq C_{p}\mathbb{E}\left[\Big{(}\int_{0}^{t}|\sigma(s,X^{\alpha}_{s})|^{2}\,ds\Big{)}^{p/2}\right], (3.4)
𝔼[sups[0,t]|0s(σ(r,Xrα)σ(r,Xr))𝑑Wr|p]Cp𝔼[(0t|σ(s,Xsα)σ(s,Xs)|2𝑑s)p/2].\displaystyle\mathbb{E}\left[\sup_{s\in[0,t]}\Big{|}\int_{0}^{s}(\sigma(r,X^{\alpha}_{r})-\sigma(r,X_{r}))\,dW_{r}\Big{|}^{p}\right]\leq C_{p}\mathbb{E}\left[\Big{(}\int_{0}^{t}|\sigma(s,X^{\alpha}_{s})-\sigma(s,X_{s})|^{2}\,ds\Big{)}^{p/2}\right]. (3.5)

The next lemma provides important estimates on the moments of the displacement process {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\}, which will be helpful to prove the main results of this section. Hereafter, we denote by CC a generic constant which may vary at each appearance.

Lemma 3.1.

Let {Xα(t),t[0,T]}\{X^{\alpha}(t),t\in[0,T]\} be the solution of (2.6) under Assumptions 1.1. Then, for all p2p\geq 2,

supα>0𝔼[sup0tT|Xα(t)|p]C,\sup_{\alpha>0}\mathbb{E}\Big{[}\sup_{0\leq t\leq T}|X^{\alpha}(t)|^{p}\Big{]}\leq C, (3.6)

and for all  0tT0\leq t\leq T,

supα>0𝔼[|Xα(t)x0|p]Ctp/2,\sup_{\alpha>0}\mathbb{E}\Big{[}|X^{\alpha}(t)-x_{0}|^{p}\Big{]}\leq Ct^{p/2}, (3.7)

where CC is a positive constant depending only on {x0,y0,κ,γ,K,p,T}\{x_{0},y_{0},\kappa,\gamma,K,p,T\}.

Proof.

We first prove (3.6). We shall divide the proof into two steps.

Step 1: We evaluate the upper bound of the moments of each Iiα(t),i=1,2,3,4I_{i}^{\alpha}(t),i=1,2,3,4.

From definition of I1α(t)I_{1}^{\alpha}(t) and Assumptions 1.1 we have

sup0st|I1α(s)|\displaystyle\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)| Kκ+γsup0st0sexp[α(κ+γ)(us)](1+|Xuα|)𝑑u\displaystyle\leq\dfrac{K}{\kappa+\gamma}\sup\limits_{0\leq s\leq t}\int_{0}^{s}\exp{[\alpha(\kappa+\gamma)(u-s)]}(1+|X^{\alpha}_{u}|)du
Ktκ+γ+Kκ+γ0tsup0us|Xuα|ds.\displaystyle\leq\dfrac{Kt}{\kappa+\gamma}+\dfrac{K}{\kappa+\gamma}\int_{0}^{t}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|ds.

Now Minkowski’s inequality (3.1) with n=2n=2 and Hölder’s inequality (3.2) yield

𝔼(sup0st|I1α(s)|p)2p1Kptp(κ+γ)p+2p1Kptp1(κ+γ)p0t𝔼(sup0us|Xuα|p)𝑑s\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}\leq\dfrac{2^{p-1}K^{p}t^{p}}{(\kappa+\gamma)^{p}}+\dfrac{2^{p-1}K^{p}t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}\Big{(}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|^{p}\Big{)}ds (3.8)

For I2α(t),I_{2}^{\alpha}(t), by substituting (2.10) into I2α(t)I_{2}^{\alpha}(t) and changing the order of integration in the double integral, one can derive that

I2α(t)\displaystyle I_{2}^{\alpha}(t) =y0γλ(t,αγ)κ+γexp[α(κ+γ)t]1κ+γ0texp[αγ(st)]E[g(s,Xsα)]𝑑s\displaystyle=\dfrac{y_{0}\gamma\lambda(t,\alpha\gamma)}{\kappa+\gamma}\exp[-\alpha(\kappa+\gamma)t]-\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha\gamma(s-t)]}E[g(s,X_{s}^{\alpha})]ds
+1κ+γ0texp[α(κ+γ)(st)]E[g(s,Xsα)]𝑑s.\displaystyle\qquad\qquad\qquad\qquad+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}E[g(s,X_{s}^{\alpha})]ds. (3.9)

Using Assumptions 1.1, Minkowski’s inequality (3.1) with n=2n=2 and Hölder’s inequality (3.2) with noting that λ(t,a)t\lambda(t,a)\leq t for all t0,a0t\geq 0,a\geq 0, we get

sup0st|I2α(s)|p\displaystyle\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p} (y0γtκ+γ+2Kκ+γ0t(1+𝔼|Xsα|)𝑑s)p\displaystyle\leq\left(\dfrac{y_{0}\gamma t}{\kappa+\gamma}+\dfrac{2K}{\kappa+\gamma}\int_{0}^{t}(1+\mathbb{E}|X_{s}^{\alpha}|)ds\right)^{p}
2p1(y0γ+2K)ptp(κ+γ)p+22p1Kptp1(κ+γ)p0t𝔼(sup0us|Xuα|p)𝑑s.\displaystyle\leq\dfrac{2^{p-1}\left(y_{0}\gamma+2K\right)^{p}t^{p}}{(\kappa+\gamma)^{p}}+\dfrac{2^{2p-1}K^{p}t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}(\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|^{p})ds. (3.10)

For I3αI_{3}^{\alpha}, using the BDG inequality (3.4), Hölder’s inequality (3.2) and Assumptions 1.1, we get

𝔼(sup0st|I3α(s)|p)\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)} Cp(κ+γ)p𝔼(0texp[2α(κ+γ)(st)]|σ(s,Xsα)|2𝑑s)p/2\displaystyle\leq\dfrac{C_{p}}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{0}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}|\sigma(s,X^{\alpha}_{s})|^{2}ds\right)^{p/2}
2p1Kptp22Cp(κ+γ)p0texp[pα(κ+γ)(st)](1+𝔼|Xsα|p)𝑑s\displaystyle\leq\dfrac{2^{p-1}K^{p}t^{\frac{p-2}{2}}C_{p}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\exp{[p\alpha(\kappa+\gamma)(s-t)]}(1+\mathbb{E}|X^{\alpha}_{s}|^{p})ds
2p1Kptp2Cp(κ+γ)p+2p1Kptp22Cp(κ+γ)p0t𝔼(sup0us|Xuα|p)𝑑s.\displaystyle\leq\dfrac{2^{p-1}K^{p}t^{\frac{p}{2}}C_{p}}{(\kappa+\gamma)^{p}}+\dfrac{2^{p-1}K^{p}t^{\frac{p-2}{2}}C_{p}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}\Big{(}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|^{p}\Big{)}ds. (3.11)

Next, from Assumptions 1.1, Minkowski’s inequality (3.1) with n=2n=2 and Hölder’s inequality (3.2) with noting that λ(t,a)t\lambda(t,a)\leq t for all t0,a0t\geq 0,a\geq 0, we get

sup0st|I4α(s)|p\displaystyle\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p} γp(κ+γ)p[y0t+Kκ0t(1+𝔼(sup0us|Xuα|)ds)𝑑s]p\displaystyle\leq\dfrac{\gamma^{p}}{(\kappa+\gamma)^{p}}\left[y_{0}t+\dfrac{K}{\kappa}\int_{0}^{t}\left(1+\mathbb{E}\Big{(}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|\Big{)}ds\right)ds\right]^{p}
2p1γp(κ+γ)p[(y0+Kκ)ptp+Kptp1κp0t𝔼(sup0us|Xuα|p)𝑑s].\displaystyle\leq\dfrac{2^{p-1}\gamma^{p}}{(\kappa+\gamma)^{p}}\left[\left(y_{0}+\dfrac{K}{\kappa}\right)^{p}t^{p}+\dfrac{K^{p}t^{p-1}}{\kappa^{p}}\int_{0}^{t}\mathbb{E}\Big{(}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|^{p}\Big{)}ds\right]. (3.12)

Step 2: We estimate the integrand in the integrals in the right hand side of the above expressions. From (2.6), applying Minkowski’s inequality with n=9n=9, Hölder’s inequality (3.2), the BDG inequality (3.4) as well as Assumptions 1.1, we obtain

𝔼(sup0st|Xsα|p)\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|X^{\alpha}_{s}|^{p}\Big{)} 9p1[x0p+tp1(κ+γ)p0t𝔼|g(s,Xsα)|pds+γptp1(κ+γ)p0t|Gα(s)|pds\displaystyle\leq 9^{p-1}\Big{[}x_{0}^{p}+\dfrac{t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}|g(s,X^{\alpha}_{s})|^{p}ds+\dfrac{\gamma^{p}t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}|G^{\alpha}(s)|^{p}\ ds
+Cp(κ+γ)p𝔼(0t|σ(s,Xsα)|2𝑑s)p/2+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad+\dfrac{C_{p}}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{0}^{t}|\sigma(s,X^{\alpha}_{s})|^{2}ds\right)^{p/2}+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p]\displaystyle\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Big{]}
9p1[x0p+2p1Kp(tp1(κ+γ)p+γptp1κp(κ+γ)p)0t(1+𝔼|Xsα|p)ds\displaystyle\leq 9^{p-1}\Big{[}x_{0}^{p}+2^{p-1}K^{p}\left(\dfrac{t^{p-1}}{(\kappa+\gamma)^{p}}+\dfrac{\gamma^{p}t^{p-1}}{\kappa^{p}(\kappa+\gamma)^{p}}\right)\int_{0}^{t}(1+\mathbb{E}|X^{\alpha}_{s}|^{p})ds
+2p1Kptp22Cp(κ+γ)p0t(1+𝔼|Xsα|p)𝑑s+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad+\dfrac{2^{p-1}K^{p}t^{\frac{p-2}{2}}C_{p}}{(\kappa+\gamma)^{p}}\int_{0}^{t}(1+\mathbb{E}|X_{s}^{\alpha}|^{p})ds+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p].\displaystyle\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Big{]}.

From this, together with (3.8), (3.1), (3.1) and (3.1), we deduce that

𝔼(sup0st|Xsα|p)C+C0t𝔼(sup0us|Xuα|p)𝑑s,\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|X^{\alpha}_{s}|^{p}\Big{)}\leq C+C\int_{0}^{t}\mathbb{E}\Big{(}\sup\limits_{0\leq u\leq s}|X^{\alpha}_{u}|^{p}\Big{)}ds, (3.13)

where CC is a positive constant depending on {x0,y0,κ,γ,K,p,T}\{x_{0},y_{0},\kappa,\gamma,K,p,T\}. From (3.13), by applying Gronwall’s lemma, we have

supα>0𝔼[sup0tT|Xα(t)|p]C,\sup_{\alpha>0}\mathbb{E}\Big{[}\sup_{0\leq t\leq T}|X^{\alpha}(t)|^{p}\Big{]}\leq C,

which completes the proof of (3.6).

Next we prove (3.7). From expression (2.6), applying Minkowski’s inequality (3.1) with n=8n=8, Hölder’s inequality (3.2), the BDG inequality (3.4) and Assumptions 1.1 again, we get

𝔼|Xsαx0|p\displaystyle\mathbb{E}|X^{\alpha}_{s}-x_{0}|^{p} 8p1[tp1(κ+γ)p0t𝔼|g(s,Xsα)|pds+γptp1(κ+γ)p0t|Gα(s)|pds\displaystyle\leq 8^{p-1}\Big{[}\dfrac{t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}|g(s,X^{\alpha}_{s})|^{p}ds+\dfrac{\gamma^{p}t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}|G^{\alpha}(s)|^{p}\ ds
+Cp(κ+γ)p𝔼(0t|σ(s,Xsα)|2𝑑s)p/2+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad+\dfrac{C_{p}}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{0}^{t}|\sigma(s,X^{\alpha}_{s})|^{2}ds\right)^{p/2}+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p]\displaystyle\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Big{]}
8p1[2p1Kp(tp1(κ+γ)p+γptp1κp(κ+γ)p)0t(1+𝔼|Xsα|p)ds\displaystyle\leq 8^{p-1}\Big{[}2^{p-1}K^{p}\left(\dfrac{t^{p-1}}{(\kappa+\gamma)^{p}}+\dfrac{\gamma^{p}t^{p-1}}{\kappa^{p}(\kappa+\gamma)^{p}}\right)\int_{0}^{t}(1+\mathbb{E}|X^{\alpha}_{s}|^{p})ds
+2p1Kptp22Cp(κ+γ)p0t(1+𝔼|Xsα|p)𝑑s+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad+\dfrac{2^{p-1}K^{p}t^{\frac{p-2}{2}}C_{p}}{(\kappa+\gamma)^{p}}\int_{0}^{t}(1+\mathbb{E}|X_{s}^{\alpha}|^{p})ds+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p].\displaystyle\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Big{]}.

This, together with (3.8), (3.1), (3.1), (3.1) and (3.6), we get (3.7). ∎

In the following theorem, we obtain a rate of convergence in LpL^{p}-distances in the Smoluchowski-Kramers approximation for the displacement process.

Theorem 3.1.

Let {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} and {Xt,t[0,T]}\{X_{t},t\in[0,T]\} be respectively the solution of (2.6) and of (2.7) under Assumptions 1.1. Then, for all p2p\geq 2, α1\alpha\geq 1 and t[0,T]t\in[0,T],

𝔼[sup0st|XsαXs|p]C[(λ(t,α(κ+γ)))p2+(λ(t,ακ))p],\mathbb{E}\Big{[}\sup_{0\leq s\leq t}|X^{\alpha}_{s}-X_{s}|^{p}\Big{]}\leq C\left[\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{\frac{p}{2}}+\left(\lambda(t,\alpha\kappa)\right)^{p}\right],

where CC is a positive constant depending on {x0,y0,κ,γ,K,L,p,T}\{x_{0},y_{0},\kappa,\gamma,K,L,p,T\} but not on α\alpha and tt.

Proof.

From (2.6) and (2.7), we have

XtαXt\displaystyle X_{t}^{\alpha}-X_{t} =1κ+γ0t(g(s,Xsα)g(s,Xs))𝑑sγκ(κ+γ)0t(𝔼[g(t,Xtα)]𝔼[g(t,Xt)])𝑑s\displaystyle=\dfrac{1}{\kappa+\gamma}\int_{0}^{t}(g(s,X^{\alpha}_{s})-g(s,X_{s}))\ ds-\dfrac{\gamma}{\kappa(\kappa+\gamma)}\int_{0}^{t}(\mathbb{E}[g(t,X^{\alpha}_{t})]-\mathbb{E}[g(t,X_{t})])ds
+1κ+γ0t(σ(s,Xsα)σ(s,Xs))𝑑W(s)+I0α(t)+I1α(t)I2α(t)I3α(t)I4α(t).\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}(\sigma(s,X^{\alpha}_{s})-\sigma(s,X_{s}))dW(s)+I_{0}^{\alpha}(t)+I_{1}^{\alpha}(t)-I_{2}^{\alpha}(t)-I_{3}^{\alpha}(t)-I_{4}^{\alpha}(t).

Similar to the proof of the previous lemma, by applying Minkowski’s inequality (3.1) with n=8n=8, Hölder’s inequality (3.2), the BDG inequality (3.5) and Assumptions 1.1, we get

𝔼[sup0st|XsαXs|p]\displaystyle\mathbb{E}\Big{[}\sup\limits_{0\leq s\leq t}|X_{s}^{\alpha}-X_{s}|^{p}\Big{]}
8p1[tp1(κ+γ)p0tE|g(s,Xsα)g(s,Xs)|pds+γptp1κp(κ+γ)p0t|𝔼[g(t,Xtα)]𝔼[g(t,Xt)]|pds\displaystyle\qquad\leq 8^{p-1}\Bigg{[}\dfrac{t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}E|g(s,X^{\alpha}_{s})-g(s,X_{s})|^{p}ds+\dfrac{\gamma^{p}t^{p-1}}{\kappa^{p}(\kappa+\gamma)^{p}}\int_{0}^{t}\Big{|}\mathbb{E}[g(t,X^{\alpha}_{t})]-\mathbb{E}[g(t,X_{t})]\Big{|}^{p}\ ds
+Cp(κ+γ)p𝔼(0t|σ(s,Xsα)σ(s,Xs)|2𝑑s)p/2+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad\qquad+\dfrac{C_{p}}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{0}^{t}|\sigma(s,X^{\alpha}_{s})-\sigma(s,X_{s})|^{2}ds\right)^{p/2}+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p]\displaystyle\qquad\qquad\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Bigg{]}
8p1[Lptp1(κ+γ)p0t𝔼|XsαXs|pds+Lpγptp1κp(κ+γ)p0t𝔼|XsαXs|pds\displaystyle\qquad\leq 8^{p-1}\bigg{[}\dfrac{L^{p}t^{p-1}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}|X^{\alpha}_{s}-X_{s}|^{p}ds+\dfrac{L^{p}\gamma^{p}t^{p-1}}{\kappa^{p}(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}|X^{\alpha}_{s}-X_{s}|^{p}ds
+CpLptp2(κ+γ)p0t𝔼|XsαXs|p𝑑s+sup0st|I0α(s)|p+𝔼(sup0st|I1α(s)|p)\displaystyle\qquad\qquad+\dfrac{C_{p}L^{p}t^{\frac{p}{2}}}{(\kappa+\gamma)^{p}}\int_{0}^{t}\mathbb{E}|X^{\alpha}_{s}-X_{s}|^{p}ds+\sup\limits_{0\leq s\leq t}|I_{0}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}
+sup0st|I2α(s)|p+𝔼(sup0st|I3α(s)|p)+sup0st|I4α(s)|p].\displaystyle\qquad\qquad\qquad+\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}+\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}+\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\Bigg{]}.

Next we estimate the terms 𝔼(sup0st|Iiα(s)|p)\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{i}^{\alpha}(s)|^{p}\Big{)}, i=1,2,3,4i=1,2,3,4. We start with 𝔼(sup0st|I1α(s)|p)\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)}. From definition of I1α(t)I_{1}^{\alpha}(t) and Assumptions 1.1 and Minkowski’s inequality (3.1) with n=2n=2 we obtain that

|I1α(t)|p\displaystyle|I_{1}^{\alpha}(t)|^{p} 2pKp(κ+γ)p(0texp[α(κ+γ)(st)](1+|Xsα|)𝑑s)p\displaystyle\leq\dfrac{2^{p}K^{p}}{(\kappa+\gamma)^{p}}\left(\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}(1+|X^{\alpha}_{s}|)ds\right)^{p}
2pKp(κ+γ)p(1+sup0tT|Xtα|p)(λ(t;(κ+γ)α))p.\displaystyle\leq\dfrac{2^{p}K^{p}}{(\kappa+\gamma)^{p}}\left(1+\sup_{0\leq t\leq T}|X^{\alpha}_{t}|^{p}\right)(\lambda(t;(\kappa+\gamma)\alpha))^{p}.

This, together with the fact that the function tλ(t,a)t\mapsto\lambda(t,a) is increasing and Lemma 3.1, implies

𝔼(sup0st|I1α(s)|p)\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{1}^{\alpha}(s)|^{p}\Big{)} 2pKp(κ+γ)p[1+𝔼(sup0tT|Xtα|p)](λ(t;(κ+γ)α))p\displaystyle\leq\dfrac{2^{p}K^{p}}{(\kappa+\gamma)^{p}}\left[1+\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|X^{\alpha}_{t}|^{p}\Big{)}\right](\lambda(t;(\kappa+\gamma)\alpha))^{p}
C(λ(t;(κ+γ)α))p,\displaystyle\leq C(\lambda(t;(\kappa+\gamma)\alpha))^{p}, (3.14)

where CC is a positive constant depending on {x0,y0,κ,γ,K,p,T}\{x_{0},y_{0},\kappa,\gamma,K,p,T\}. Next, using (3.1) and Lemma 3.1, we get

|I2α(t)|y0λ(t;α(κ+γ))+Kκ+γ0texp[ακ(st)][1+𝔼(|Xsα|)]𝑑s+Kκ+γ0texp[α(κ+γ)(st)][1+𝔼(|Xsα|)]𝑑sC[λ(t;α(κ+γ)+0t(exp[ακ(st)]+exp[α(κ+γ)(st)])ds]C[λ(t;α(κ+γ)+λ(t,ακ)],\begin{split}|I_{2}^{\alpha}(t)|&\leq y_{0}\lambda(t;\alpha(\kappa+\gamma))+\dfrac{K}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha\kappa(s-t)]}[1+\mathbb{E}(|X^{\alpha}_{s}|)]\ ds\\ &\qquad+\dfrac{K}{\kappa+\gamma}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}[1+\mathbb{E}(|X^{\alpha}_{s}|)]\ ds\\ &\leq C\left[\lambda(t;\alpha(\kappa+\gamma)+\int_{0}^{t}\left(\exp{[\alpha\kappa(s-t)]}+\exp{[\alpha(\kappa+\gamma)(s-t)]}\right)ds\right]\\ &\leq C\left[\lambda(t;\alpha(\kappa+\gamma)+\lambda(t,\alpha\kappa)\right],\end{split}

where CC is constant depending on {x0,y0,κ,γ,K,p}\{x_{0},y_{0},\kappa,\gamma,K,p\}. Thus,

sup0st|I2α(s)|pC[(λ(t,α(κ+γ)))p+(λ(t,ακ))p].\sup\limits_{0\leq s\leq t}|I_{2}^{\alpha}(s)|^{p}\leq C\left[\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{p}+\left(\lambda(t,\alpha\kappa)\right)^{p}\right]. (3.15)

Applying the BDG inequality (3.3), Hölder’s inequality (3.2) and Lemma 3.1, one can derive that

𝔼(sup0st|I3α(s)|p)Cp(κ+γ)pE(0texp[2α(κ+γ)(st)]|σ(s,Xsα)|2𝑑s)p/2\displaystyle\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}\leq\dfrac{C_{p}}{(\kappa+\gamma)^{p}}E\left(\int_{0}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}|\sigma(s,X^{\alpha}_{s})|^{2}ds\right)^{p/2}
Cp(κ+γ)p(0texp[pp1α(κ+γ)(st)]𝑑s)p/210texp[p2α(κ+γ)(st)]𝔼|σ(s,Xsα)|p𝑑s\displaystyle\quad\leq\dfrac{C_{p}}{(\kappa+\gamma)^{p}}\left(\int_{0}^{t}\exp{\left[\frac{p}{p-1}\alpha(\kappa+\gamma)(s-t)\right]}ds\right)^{p/2-1}\int_{0}^{t}\exp{\left[\frac{p}{2}\alpha(\kappa+\gamma)(s-t)\right]}\mathbb{E}|\sigma(s,X^{\alpha}_{s})|^{p}\,ds
2p1Cp(κ+γ)p(0texp[pp1α(κ+γ)(st)]𝑑s)p/210texp[p2α(κ+γ)(st)](1+𝔼|Xsα|p)𝑑s\displaystyle\quad\leq\dfrac{2^{p-1}C_{p}}{(\kappa+\gamma)^{p}}\left(\int_{0}^{t}\exp{\left[\frac{p}{p-1}\alpha(\kappa+\gamma)(s-t)\right]}ds\right)^{p/2-1}\int_{0}^{t}\exp{\left[\frac{p}{2}\alpha(\kappa+\gamma)(s-t)\right]}(1+\mathbb{E}|X^{\alpha}_{s}|^{p})ds
C(λ(t,pp1α(κ+γ)))p/21λ(t,p2α(κ+γ)),\displaystyle\quad\leq C\left(\lambda(t,\frac{p}{p-1}\alpha(\kappa+\gamma))\right)^{p/2-1}\lambda(t,\frac{p}{2}\alpha(\kappa+\gamma)),

where CC is constant depending on {x0,y0,κ,γ,K,p,T}\{x_{0},y_{0},\kappa,\gamma,K,p,T\}. On the other hand, for all t>0t>0 and a>0a>0, we have

λ(t,a)a=(1+at)eat1a2<0.\frac{\partial\lambda(t,a)}{\partial a}=\dfrac{(1+at)e^{-at}-1}{a^{2}}<0.

Thus the function aλ(t,a)a\mapsto\lambda(t,a) is decreasing. Hence we get

𝔼(sup0st|I3α(s)|p)C(λ(t,α(κ+γ)))p/2.\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{3}^{\alpha}(s)|^{p}\Big{)}\leq C\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{p/2}.

Now we consider sup0st|I4α(s)|p\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}. Using Lemma 3.1 we can derive that

sup0st|I4α(s)|γκ+γ[y0λ(t;ακ)+Kκ0texp[α(κ+γ)(ut)][1+𝔼(|Xuα|)]𝑑u].\begin{split}\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|\leq\dfrac{\gamma}{\kappa+\gamma}\left[y_{0}\lambda(t;\alpha\kappa)+\dfrac{K}{\kappa}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(u-t)]}\left[1+\mathbb{E}(|X^{\alpha}_{u}|)\right]\,du\right].\end{split}

Thus,

sup0st|I4α(s)|pC[(λ(t,α(κ+γ)))p+(λ(t,ακ))p],\sup\limits_{0\leq s\leq t}|I_{4}^{\alpha}(s)|^{p}\leq C\left[\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{p}+\left(\lambda(t,\alpha\kappa)\right)^{p}\right],

where CC is constant depending on {x0,y0,κ,γ,K,L,p}\{x_{0},y_{0},\kappa,\gamma,K,L,p\}. From the above estimates, together with the fact that λ(t,α(κ+γ))tT,\lambda(t,\alpha(\kappa+\gamma))\leq t\leq T, one sees that

𝔼[sup0st|XsαXs|p]C[t2p10t𝔼|XsαXs|p𝑑s+i=04𝔼(sup0st|Iiα(s)|p)]C[0t𝔼[sup0us|XuαXu|p]𝑑s+(λ(t,α(κ+γ)))p2+(λ(t,ακ))p],\begin{split}\mathbb{E}\Big{[}\sup\limits_{0\leq s\leq t}|X_{s}^{\alpha}-X_{s}|^{p}\Big{]}&\leq C\left[t^{2p-1}\int_{0}^{t}\mathbb{E}|X^{\alpha}_{s}-X_{s}|^{p}ds+\sum_{i=0}^{4}\mathbb{E}\Big{(}\sup\limits_{0\leq s\leq t}|I_{i}^{\alpha}(s)|^{p}\Big{)}\right]\\ &\leq C\left[\int_{0}^{t}\mathbb{E}\Big{[}\sup\limits_{0\leq u\leq s}|X_{u}^{\alpha}-X_{u}|^{p}\Big{]}ds+\left(\lambda(t,\alpha(\kappa+\gamma))\right)^{\frac{p}{2}}+\left(\lambda(t,\alpha\kappa)\right)^{p}\right],\\ \end{split}

where CC is constant depending only on {x0,y0,κ,γ,K,L,p}\{x_{0},y_{0},\kappa,\gamma,K,L,p\}. Using Growwall’s inequality, we obtain the claimed estimate and complete the proof. ∎

In the following lemma, we show Malliavin differentiability of XtαX^{\alpha}_{t} and XtX_{t}.

Lemma 3.2.

Under Assumptions 1.1, the solutions {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} and {Xt,t[0,T]}\{X_{t},t\in[0,T]\} of (2.6) and (2.7) respectively are Malliavin differentiable random variables. Moreover, the derivatives DrXtα,DrXtD_{r}X^{\alpha}_{t},D_{r}X_{t} satisfy DrXtα=DrXt=0D_{r}X^{\alpha}_{t}=D_{r}X_{t}=0 for rtr\geq t and for 0r<tT,0\leq r<t\leq T,

DrXtα\displaystyle D_{r}X^{\alpha}_{t} =σ(r,Xrα)κ+γ(exp[α(κ+γ)(rt)]1)1κ+γrtg¯α(s)DrXsα𝑑s\displaystyle=\dfrac{\sigma(r,X_{r}^{\alpha})}{\kappa+\gamma}(\exp{[\alpha(\kappa+\gamma)(r-t)]}-1)-\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\bar{g}^{\alpha}(s)D_{r}X_{s}^{\alpha}ds
+1κ+γrtσ¯α(s)DrXsα𝑑Ws+1κ+γrtexp[α(κ+γ)(st)]g¯α(s)DrXsα𝑑s\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\bar{\sigma}^{\alpha}(s)D_{r}X_{s}^{\alpha}\,dW_{s}+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}\bar{g}^{\alpha}(s)D_{r}X_{s}^{\alpha}ds
+1κ+γrtexp[α(κ+γ)(st)]σ¯α(s)DrXsα𝑑Ws,\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}\bar{\sigma}^{\alpha}(s)D_{r}X_{s}^{\alpha}\,dW_{s}, (3.16)
DrXt\displaystyle D_{r}X_{t} =σ(r,Xr)κ+γ1κ+γrtg¯(s)DrXs𝑑s+1κ+γrtσ¯(s)DrXs𝑑Ws,\displaystyle=\dfrac{\sigma(r,X_{r})}{\kappa+\gamma}-\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\bar{g}(s)D_{r}X_{s}ds+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\bar{\sigma}(s)D_{r}X_{s}\,dW_{s}, (3.17)

where g¯(s),g¯α(s),σ¯(s),σ¯α(s)\bar{g}(s),\bar{g}^{\alpha}(s),\bar{\sigma}(s),\bar{\sigma}^{\alpha}(s) are adapted stochastic processes and are bounded by LL.

Proof.

From the second equation of (1.1) we have

Ytα=y0α(κ+γ)0tYsα𝑑sα0tg(s,Xsα)𝑑sαγ0t𝔼(Ysα)𝑑s+α0tσ(s,Xsα)𝑑Ws.Y^{\alpha}_{t}=y_{0}-\alpha(\kappa+\gamma)\int_{0}^{t}Y^{\alpha}_{s}ds-\alpha\int_{0}^{t}g(s,X^{\alpha}_{s})ds-\alpha\gamma\int_{0}^{t}\mathbb{E}(Y^{\alpha}_{s})\,ds+\alpha\int_{0}^{t}\sigma(s,X^{\alpha}_{s})\,dW_{s}.

Using Minkowski’s inequality (3.1) with n=5n=5, Hölder’s inequality (3.2) with p=2p=2, the BDG inequality (3.3), Assumptions 1.1 and Lemma 3.1 we get

𝔼(sup0st|Ysα|2)\displaystyle\mathbb{E}(\sup\limits_{0\leq s\leq t}|Y^{\alpha}_{s}|^{2}) 5(y02+tα2(κ+γ)20t𝔼|Ysα|2ds+tα20tg2(s,Xsα)dstα2γ20t𝔼|Ysα|2ds\displaystyle\leq 5\bigg{(}y^{2}_{0}+t\alpha^{2}(\kappa+\gamma)^{2}\int_{0}^{t}\mathbb{E}|Y^{\alpha}_{s}|^{2}ds+t\alpha^{2}\int_{0}^{t}g^{2}(s,X^{\alpha}_{s})\,ds-t\alpha^{2}\gamma^{2}\int_{0}^{t}\mathbb{E}|Y^{\alpha}_{s}|^{2}\,ds
+α20tσ2(s,Xsα)ds)\displaystyle\qquad\qquad+\alpha^{2}\int_{0}^{t}\sigma^{2}(s,X^{\alpha}_{s})\,ds\bigg{)}
C(1+0t𝔼(sup0us|Yuα|2)𝑑s+0t(1+𝔼|Xsα|2)𝑑s)\displaystyle\leq C\left(1+\int_{0}^{t}\mathbb{E}(\sup\limits_{0\leq u\leq s}|Y^{\alpha}_{u}|^{2})\,ds+\int_{0}^{t}(1+\mathbb{E}|X^{\alpha}_{s}|^{2})\,ds\right)
C(1+0t𝔼(sup0us|Yuα|2)𝑑s).\displaystyle\leq C\left(1+\int_{0}^{t}\mathbb{E}(\sup\limits_{0\leq u\leq s}|Y^{\alpha}_{u}|^{2})\,ds\right).

Applying Gronwall’s lemma we obtain

𝔼[sup0tT|Ytα|2]C.\mathbb{E}\Big{[}\sup_{0\leq t\leq T}|Y^{\alpha}_{t}|^{2}\Big{]}\leq C.

This, together with Lemma 3.1 we deduce that 𝔼[sup0tT|Ytα|]\mathbb{E}[\sup\limits_{0\leq t\leq T}|Y^{\alpha}_{t}|] and 𝔼[sup0tT|Xtα|]\mathbb{E}[\sup\limits_{0\leq t\leq T}|X^{\alpha}_{t}|] are bounded. Then, by Assumptions 1.1 and the dominated convergence theorem, the integrals I0α(t),I2α(t)I_{0}^{\alpha}(t),I_{2}^{\alpha}(t) and I4α(t)I_{4}^{\alpha}(t) are continuous functions.

Let us define

f(t):=x0γκ+γ0tGα(u)𝑑u+I0α(t)I2α(t)I4α(t).f(t):=x_{0}-\dfrac{\gamma}{\kappa+\gamma}\int_{0}^{t}G^{\alpha}(u)\ du+I_{0}^{\alpha}(t)-I_{2}^{\alpha}(t)-I_{4}^{\alpha}(t).

Then f(t)f(t) is continuous function in [0,T][0,T] and Equation (2.6) becomes

Xtα=f(t)+1κ+γ0t(exp[α(κ+γ)(st)]1)g(s,Xsα)𝑑s+1κ+γ0t(exp[α(κ+γ)(st)]1)σ(s,Xs)𝑑Ws.\begin{split}X^{\alpha}_{t}&=f(t)+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\left(\exp{[\alpha(\kappa+\gamma)(s-t)]}-1\right)g(s,X^{\alpha}_{s})\ ds\\ &\qquad+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}(\exp{[\alpha(\kappa+\gamma)(s-t)]}-1)\sigma(s,X_{s})\ dW_{s}.\end{split} (3.18)

Now, we consider the Picard approximation sequence {Xtα,n,t[0,T]}n0\{X^{\alpha,n}_{t},t\in[0,T]\}_{n\geq 0} given by

{Xtα,0=f(t),Xtα,n+1=f(t)+1κ+γ0t(exp[α(κ+γ)(st)]1)g(s,Xsα,n)𝑑s+δκ+γ0t(exp[α(κ+γ)(st)]1)σ(s,Xsα,n)𝑑Ws,t[0,T],n0.\begin{cases}X^{\alpha,0}_{t}=f(t),\\ \begin{split}X^{\alpha,n+1}_{t}&=f(t)+\dfrac{1}{\kappa+\gamma}\int_{0}^{t}\left(\exp{[\alpha(\kappa+\gamma)(s-t)]}-1\right)g(s,X^{\alpha,n}_{s})\ ds\\ &\quad+\dfrac{\delta}{\kappa+\gamma}\int_{0}^{t}(\exp{[\alpha(\kappa+\gamma)(s-t)]}-1)\sigma(s,X_{s}^{\alpha,n})\ dW_{s},\,t\in[0,T],\,n\geq 0.\end{split}\end{cases}

From this, using the same method as in the proof of [Nua06, Theorem 2.2.1], we conclude that the solution {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} of (3.18) (thus, of (2.6)) is Malliavin’s differentable. Obviously, the solution {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} is 𝔽\mathbb{F}-adapted. Hence, we always have DθXtα=0D_{\theta}X^{\alpha}_{t}=0 for θ>t.\theta>t. For θt,\theta\leq t, from [Nua06, Proposition 1.2.4] and Lipschitz property of gg and σ\sigma, there exist adapted processes g¯α(s)\bar{g}^{\alpha}(s), σ¯α(s)\bar{\sigma}^{\alpha}(s) uniformly bounded by LL such that Dθg(s,Xsα)=g¯α(s)DθXsαD_{\theta}g(s,X^{\alpha}_{s})=\bar{g}^{\alpha}(s)D_{\theta}X^{\alpha}_{s} and Dθσ(s,Xsα)=σ¯α(s)DθXsαD_{\theta}\sigma(s,X^{\alpha}_{s})=\bar{\sigma}^{\alpha}(s)D_{\theta}X^{\alpha}_{s}. Then we obtain (3.2) by applying the operator DD to the equation (3.18).

The proof for the solution XtX_{t} of (2.7) is similar. ∎

Remark 3.1.

If gg and σ\sigma are continuously differentiable, then g¯(s)=g2(s,Xs)\bar{g}(s)=g^{\prime}_{2}(s,X_{s}), g¯α(s)=g2(s,Xsα),σ¯(s)=σ2(s,Xs)\bar{g}^{\alpha}(s)=g^{\prime}_{2}(s,X_{s}^{\alpha}),\bar{\sigma}(s)=\sigma^{\prime}_{2}(s,X_{s}) and σ¯α(s)=σ2(s,Xsα)\bar{\sigma}^{\alpha}(s)=\sigma^{\prime}_{2}(s,X_{s}^{\alpha}). Here, for a function h(t,x),h(t,x), we use the convention h2(t,x)=h(t,x)x.h^{\prime}_{2}(t,x)=\frac{\partial h(t,x)}{\partial x}.

In the next lemma, we show that the moments of the Malliavin’s derivative of solutions of (2.7) are bounded.

Lemma 3.3.

Let {Xt,t[0,T]}\{X_{t},t\in[0,T]\} be the solution of (2.7) with Assumptions 1.1. Then, for all p2p\geq 2, we have

sup0t,rT𝔼(|DrXt|p)<.\sup_{0\leq t,r\leq T}\mathbb{E}(|D_{r}X_{t}|^{p})<\infty.
Proof.

Using Minkowski’s inequality (3.1) with n=3n=3, Hölder’s inequality (3.2), the BDG inequality (3.3), Assumptions 1.1 and Lemma 3.1, and noting that |g¯(s)|L,|σ¯(s)|L|\bar{g}(s)|\leq L,|\bar{\sigma}(s)|\leq L, it follows from (3.17) that

𝔼[|DrXt|p]\displaystyle\mathbb{E}[|D_{r}X_{t}|^{p}]
3p1(𝔼[|σ(r,Xr)|p](κ+γ)p1(κ+γ)p𝔼(rtg¯(s)DrXs𝑑s)p+1(κ+γ)p𝔼(rt|σ¯(s)DrXs|2𝑑s)p/2)\displaystyle\quad\leq 3^{p-1}\left(\dfrac{\mathbb{E}[|\sigma(r,X_{r})|^{p}]}{(\kappa+\gamma)^{p}}-\dfrac{1}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{r}^{t}\bar{g}(s)D_{r}X_{s}\,ds\right)^{p}+\dfrac{1}{(\kappa+\gamma)^{p}}\mathbb{E}\left(\int_{r}^{t}|\bar{\sigma}(s)D_{r}X_{s}|^{2}\,ds\right)^{p/2}\right)
3p1(Kp(1+𝔼[|Xt|p])(κ+γ)pLp(tr)p1(κ+γ)prt𝔼[|DrXs|p]𝑑s+Lp(tr)p/21(κ+γ)prt𝔼[|DrXs|p]𝑑s)\displaystyle\quad\leq 3^{p-1}\left(\dfrac{K^{p}(1+\mathbb{E}[|X_{t}|^{p}])}{(\kappa+\gamma)^{p}}-\dfrac{L^{p}(t-r)^{p-1}}{(\kappa+\gamma)^{p}}\int_{r}^{t}\mathbb{E}[|D_{r}X_{s}|^{p}]\,ds+\dfrac{L^{p}(t-r)^{p/2-1}}{(\kappa+\gamma)^{p}}\int_{r}^{t}\mathbb{E}[|D_{r}X_{s}|^{p}]ds\right)
C[1+rt|DrXα(s)|p𝑑s],\displaystyle\quad\leq C\left[1+\int_{r}^{t}|D_{r}X^{\alpha}(s)|^{p}\,ds\right],

where CC is a positive constant depending only on {κ,γ,K,L,p,T}\{\kappa,\gamma,K,L,p,T\}.

Taking the expectation and using Gronwall’s inequality, we obtain the claimed estimate. ∎

The following lemma provides an upper bound for the difference between the derivatives of the solutions of (2.6) and (2.7).

Lemma 3.4.

Let {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} and {Xt,t[0,T]}\{X_{t},t\in[0,T]\} be respectively the solution of (2.6) and of (2.7) under Assumptions 1.1 and 1.2. Then, for all α1\alpha\geq 1,

𝔼[DXtαDXt2]C(λ(t,α(κ+γ))+(λ(t,ακ))2),\mathbb{E}\left[\|DX^{\alpha}_{t}-DX_{t}\|_{\mathcal{H}}^{2}\right]\leq C(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}),

where CC is a positive constant depending only on {x0,y0,κ,γ,K,L,M,T}\{x_{0},y_{0},\kappa,\gamma,K,L,M,T\}.

Proof.

Under Assumptions 1.2, gg and σ\sigma are twice differentiable, thus (see Remark 3.1) g¯(s)=g2(s,Xs)\bar{g}(s)=g^{\prime}_{2}(s,X_{s}), g¯α(s)=g2(s,Xsα),σ¯(s)=σ2(s,Xs)\bar{g}^{\alpha}(s)=g^{\prime}_{2}(s,X_{s}^{\alpha}),\bar{\sigma}(s)=\sigma^{\prime}_{2}(s,X_{s}) and σ¯α(s)=σ2(s,Xsα)\bar{\sigma}^{\alpha}(s)=\sigma^{\prime}_{2}(s,X_{s}^{\alpha}). Furthermore,

|g2(s,Xsα)g2(s,Xs)|M|XsαXs|,|σ2(s,Xsα)σ2(s,Xs|M|XsαXs|.|g^{\prime}_{2}(s,X_{s}^{\alpha})-g^{\prime}_{2}(s,X_{s})|\leq M|X_{s}^{\alpha}-X_{s}|,\quad|\sigma^{\prime}_{2}(s,X_{s}^{\alpha})-\sigma^{\prime}_{2}(s,X_{s}|\leq M|X_{s}^{\alpha}-X_{s}|. (3.19)

From (3.2) and (3.17) we have

DrXtαDrXt\displaystyle D_{r}X^{\alpha}_{t}-D_{r}X_{t} =(σ(r,Xrα)κ+γ(exp[α(κ+γ)(rt)]1)σ(r,Xr)κ+γ)\displaystyle=\left(\dfrac{\sigma(r,X_{r}^{\alpha})}{\kappa+\gamma}(\exp{[\alpha(\kappa+\gamma)(r-t)]}-1)-\dfrac{\sigma(r,X_{r})}{\kappa+\gamma}\right)
1κ+γrt(g2(s,Xsα)DrXsαg2(s,Xs)DrXs)𝑑s\displaystyle\qquad-\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\left(g^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X_{s}^{\alpha}-g^{\prime}_{2}(s,X_{s})D_{r}X_{s}\right)\,ds
+1κ+γrt(σ2(s,Xsα)DrXsασ2(s,Xs)DrXs)𝑑Ws\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\left(\sigma^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X_{s}^{\alpha}-\sigma^{\prime}_{2}(s,X_{s})D_{r}X_{s}\right)\,dW_{s}
+1κ+γrtexp[α(κ+γ)(st)]g2(s,Xsα)DrXsα𝑑s\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}g^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X^{\alpha}_{s}\,ds
+1κ+γrtexp[α(κ+γ)(st)]σ2(s,Xs)DrXsα𝑑Ws.\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}\sigma^{\prime}_{2}(s,X_{s})D_{r}X^{\alpha}_{s}\,dW_{s}. (3.20)

Now, we shall estimate each term in the right hand side of (3.1). First, using Assumptions 1.1, Lemma 3.1 and Theorem 3.1 for p=2p=2, we can derive that

𝔼(σ(r,Xrα)κ+γ(exp[α(κ+γ)(rt)]1)σ(r,Xr)κ+γ)2\displaystyle\mathbb{E}\left(\dfrac{\sigma(r,X_{r}^{\alpha})}{\kappa+\gamma}(\exp{[\alpha(\kappa+\gamma)(r-t)]}-1)-\dfrac{\sigma(r,X_{r})}{\kappa+\gamma}\right)^{2}
2[𝔼[|σ(r,Xrα)σ(r,Xr)|2](κ+γ)2+𝔼[|σ(r,Xrα)|2](κ+γ)2exp[2α(κ+γ)(rt)]]\displaystyle\qquad\leq 2\left[\dfrac{\mathbb{E}\big{[}|\sigma(r,X_{r}^{\alpha})-\sigma(r,X_{r})|^{2}\big{]}}{(\kappa+\gamma)^{2}}+\dfrac{\mathbb{E}\big{[}|\sigma(r,X_{r}^{\alpha})|^{2}\big{]}}{(\kappa+\gamma)^{2}}\exp{[2\alpha(\kappa+\gamma)(r-t)]}\right]
2[L2𝔼|XrαXr|2(κ+γ)2+2K2(1+𝔼[|Xrα|2])(κ+γ)2exp[2α(κ+γ)(rt)]]\displaystyle\qquad\leq 2\left[\dfrac{L^{2}\mathbb{E}|X_{r}^{\alpha}-X_{r}|^{2}}{(\kappa+\gamma)^{2}}+\dfrac{2K^{2}(1+\mathbb{E}[|X_{r}^{\alpha}|^{2}])}{(\kappa+\gamma)^{2}}\exp{[2\alpha(\kappa+\gamma)(r-t)]}\right]
C[λ(t,α(κ+γ))+(λ(t,ακ))2+exp[2α(κ+γ)(rt)]],\displaystyle\qquad\leq C\left[\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\exp{[2\alpha(\kappa+\gamma)(r-t)]}\right],

where CC is constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}. From Hölder’s inequality, Assumptions 1.1, Lemma 3.1, Lemma 3.3 and Theorem 3.1 for p=4p=4, together with (3.19) we get

𝔼(rt(g2(s,Xsα)DrXsαg2(s,Xs)DrXs)𝑑s)2\displaystyle\mathbb{E}\left(\int_{r}^{t}\left(g^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X_{s}^{\alpha}-g^{\prime}_{2}(s,X_{s})D_{r}X_{s}\right)\,ds\right)^{2}
2𝔼(rt(g2(s,Xsα)g2(s,Xs))DrXs𝑑s)2+2𝔼(rtg2(s,Xsα)(DrXsαDrXs)𝑑s)2\displaystyle\qquad\leq 2\mathbb{E}\left(\int_{r}^{t}\left(g^{\prime}_{2}(s,X_{s}^{\alpha})-g^{\prime}_{2}(s,X_{s})\right)D_{r}X_{s}\,ds\right)^{2}+2\mathbb{E}\left(\int_{r}^{t}g^{\prime}_{2}(s,X_{s}^{\alpha})\left(D_{r}X_{s}^{\alpha}-D_{r}X_{s}\right)\,ds\right)^{2}
2(tr)[rt𝔼|(g2(s,Xsα)g2(s,Xs))DrXs|2ds\displaystyle\qquad\leq 2(t-r)\Bigg{[}\int_{r}^{t}\mathbb{E}|\left(g^{\prime}_{2}(s,X_{s}^{\alpha})-g^{\prime}_{2}(s,X_{s})\right)D_{r}X_{s}|^{2}\,ds
+rt𝔼|g2(s,Xsα)(DrXsαDrXs)|2ds]\displaystyle\qquad\qquad\qquad+\int_{r}^{t}\mathbb{E}\left|g^{\prime}_{2}(s,X_{s}^{\alpha})\left(D_{r}X_{s}^{\alpha}-D_{r}X_{s}\right)\right|^{2}\,ds\Bigg{]}
2M2(tr)rt(𝔼[|XsαXs|4])1/2(𝔼|DrXs|4)1/2𝑑s\displaystyle\qquad\leq 2\mathrm{M}^{2}(t-r)\int_{r}^{t}\left(\mathbb{E}\left[|X_{s}^{\alpha}-X_{s}|^{4}\right]\right)^{1/2}\left(\mathbb{E}|D_{r}X_{s}|^{4}\right)^{1/2}\,ds
+2L2(tr)rt𝔼[|DrXsαDrXs|2]𝑑s\displaystyle\qquad\qquad\qquad+2L^{2}(t-r)\int_{r}^{t}\mathbb{E}\left[|D_{r}X_{s}^{\alpha}-D_{r}X_{s}|^{2}\right]\,ds
C(λ(t,α(κ+γ))+(λ(t,ακ))2+rt𝔼[|DrXsαDrXs|2]𝑑s).\displaystyle\qquad\leq C\left(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\int_{r}^{t}\mathbb{E}\left[|D_{r}X_{s}^{\alpha}-D_{r}X_{s}|^{2}\right]\,ds\right).

By Itô’s isometry formula, Hölder’s inequality, Assumptions 1.1, Lemma 3.1, Lemma 3.3 and Theorem 3.1 for p=4p=4, together with (3.19), we have

𝔼(rt(σ2(s,Xsα)DrXsασ2(s,Xs)DrXs)𝑑Ws)2\displaystyle\mathbb{E}\left(\int_{r}^{t}\left(\sigma^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X_{s}^{\alpha}-\sigma^{\prime}_{2}(s,X_{s})D_{r}X_{s}\right)\,dW_{s}\right)^{2}
2𝔼(rt(σ2(s,Xsα)σ2(s,Xs))DrXs𝑑Ws)2+2𝔼(rtσ2(s,Xsα)(DrXsαDrXs)𝑑Ws)2\displaystyle\qquad\leq 2\mathbb{E}\left(\int_{r}^{t}\left(\sigma^{\prime}_{2}(s,X_{s}^{\alpha})-\sigma^{\prime}_{2}(s,X_{s})\right)D_{r}X_{s}\,dW_{s}\right)^{2}+2\mathbb{E}\left(\int_{r}^{t}\sigma^{\prime}_{2}(s,X_{s}^{\alpha})\left(D_{r}X_{s}^{\alpha}-D_{r}X_{s}\right)\,dW_{s}\right)^{2}
2rt𝔼|(σ2(s,Xsα)σ2(s,Xs))DrXs|2𝑑s+2rt𝔼|σ2(s,Xsα)(DrXsαDrXs)|2𝑑s\displaystyle\qquad\leq 2\int_{r}^{t}\mathbb{E}|\left(\sigma^{\prime}_{2}(s,X_{s}^{\alpha})-\sigma^{\prime}_{2}(s,X_{s})\right)D_{r}X_{s}|^{2}\,ds+2\int_{r}^{t}\mathbb{E}|\sigma^{\prime}_{2}(s,X_{s}^{\alpha})\left(D_{r}X_{s}^{\alpha}-D_{r}X_{s}\right)|^{2}\,ds
2M2rt(𝔼|XsαXs|4)1/2(𝔼|DrXs|4)1/2𝑑s+2L2rt𝔼|DrXsαDrXs|2𝑑s\displaystyle\qquad\leq 2\mathrm{M}^{2}\int_{r}^{t}\left(\mathbb{E}|X_{s}^{\alpha}-X_{s}|^{4}\right)^{1/2}\left(\mathbb{E}|D_{r}X_{s}|^{4}\right)^{1/2}ds+2L^{2}\int_{r}^{t}\mathbb{E}|D_{r}X_{s}^{\alpha}-D_{r}X_{s}|^{2}ds
C[λ(t,α(κ+γ))+(λ(t,ακ))2+rt𝔼[|DrXsαDrXs|2]𝑑s].\displaystyle\qquad\leq C\left[\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\int_{r}^{t}\mathbb{E}\big{[}|D_{r}X_{s}^{\alpha}-D_{r}X_{s}|^{2}\big{]}ds\right].

Next, using Hölder’s inequality, Assumptions 1.1 and Lemma 3.3 one sees that

𝔼(rtexp[α(κ+γ)(st)]g2(s,Xsα)DrXsα𝑑s)2\displaystyle\mathbb{E}\left(\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}g^{\prime}_{2}(s,X_{s}^{\alpha})D_{r}X^{\alpha}_{s}ds\right)^{2}
(tr)L2rtexp[2α(κ+γ)(st)]𝔼[|DrXsα|2]𝑑s\displaystyle\qquad\leq(t-r)L^{2}\int_{r}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}\mathbb{E}\big{[}|D_{r}X^{\alpha}_{s}|^{2}\big{]}ds
Crtexp[2α(κ+γ)(st)]𝑑s\displaystyle\qquad\leq C\int_{r}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}ds
Cλ(t,α(κ+γ)).\displaystyle\qquad\leq C\lambda(t,\alpha(\kappa+\gamma)).

By the same estimate for the last term in the right hand side of (3.1), we can obtain

𝔼(rtexp[α(κ+γ)(st)]σ2(s,Xs)DrXsα𝑑Ws)2\displaystyle\mathbb{E}\left(\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}\sigma^{\prime}_{2}(s,X_{s})D_{r}X^{\alpha}_{s}dW_{s}\right)^{2}
L2rtexp[2α(κ+γ)(st)]E[|DrXsα|2]𝑑s\displaystyle\qquad\leq L^{2}\int_{r}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}E\big{[}|D_{r}X^{\alpha}_{s}|^{2}\big{]}\,ds
Crtexp[2α(κ+γ)(st)]𝑑s\displaystyle\qquad\leq C\int_{r}^{t}\exp{[2\alpha(\kappa+\gamma)(s-t)]}ds
Cλ(t,α(κ+γ)).\displaystyle\qquad\leq C\lambda(t,\alpha(\kappa+\gamma)).

From the above estimates, together with the fact that the function aλ(t,a)a\mapsto\lambda(t,a) is decreasing one can derive that

0t𝔼[|DrXtαDrXt|2]𝑑rC0t(λ(t,α(κ+γ))+(λ(t,ακ))2+exp[2α(κ+γ)(rt)]+rt𝔼[|DrXuαDrXu|2]𝑑u)𝑑rC[λ(t,α(κ+γ))+(λ(t,ακ))2+λ(t,2α(κ+γ))+0t𝑑rrt𝔼[|DrXuDrXuα|2]𝑑u]C[λ(t,α(κ+γ))+(λ(t,ακ))2+0t𝑑u0u𝔼[|DrXuDrXuα|2]𝑑r],\begin{split}&\int_{0}^{t}\mathbb{E}\left[|D_{r}X^{\alpha}_{t}-D_{r}X_{t}|^{2}\right]\ dr\\ &\leq C\,\int_{0}^{t}\left(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\exp{[2\alpha(\kappa+\gamma)(r-t)]}+\int_{r}^{t}\mathbb{E}\left[|D_{r}X^{\alpha}_{u}-D_{r}X_{u}|^{2}\right]\ du\right)\,dr\\ &\leq C\,\left[\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\lambda(t,2\alpha(\kappa+\gamma))+\int_{0}^{t}dr\int_{r}^{t}\mathbb{E}\left[|D_{r}X_{u}-D_{r}X^{\alpha}_{u}|^{2}\right]\,du\right]\\ &\leq C\,\left[\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\int_{0}^{t}du\int_{0}^{u}\mathbb{E}\left[|D_{r}X_{u}-D_{r}X^{\alpha}_{u}|^{2}\right]\,dr\right],\end{split}

where CC is constant depending only on {x0,y0,κ,γ,K,L,M,T}\{x_{0},y_{0},\kappa,\gamma,K,L,\mathrm{M},T\}.

Let ϕ(t):=0t𝔼[|DrXtDrXtα|2]𝑑r\phi(t):=\displaystyle\int_{0}^{t}\mathbb{E}\left[|D_{r}X_{t}-D_{r}X^{\alpha}_{t}|^{2}\right]dr, then we have

ϕ(t)C[λ(t,α(κ+γ))+(λ(t,ακ))2+0tϕ(u)𝑑u].\phi(t)\leq C\left[\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2}+\displaystyle\int_{0}^{t}\phi(u)\ du\right].

Thus, applying Gronwall’s inequality, we get

ϕ(t)C(λ(t,α(κ+γ))+(λ(t,ακ))2)exp(Ct),\phi(t)\leq C(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})\exp{(Ct)},

where CC is constant depending only on {x0,y0,κ,γ,K,L,M,T}\{x_{0},y_{0},\kappa,\gamma,K,L,\mathrm{M},T\}. This completes the proof of the lemma. ∎

Now, we give explicit bounds on the total variation distance between the solution Xα(t)X^{\alpha}(t) of (2.6) and the solution XtX_{t} of (2.7).

Theorem 3.2.

Let {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} and {Xt,t[0,T]}\{X_{t},t\in[0,T]\} be, respectively, the solution of (2.6)and of (2.7) with Assumptions 1.1 and 1.2. We further assume that |σ(t,x)|σ0>0|\sigma(t,x)|\geq\sigma_{0}>0 for all (t,x)[0,T]×.(t,x)\in[0,T]\times\mathbb{R}. Then, for each α1\alpha\geq 1 and t(0,T],t\in(0,T],

dTV(Xtα,Xt)Ct1(λ(t,α(κ+γ))+(λ(t,ακ))2),d_{TV}(X^{\alpha}_{t},X_{t})\leq C\sqrt{t^{-1}(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})},

where CC is a constant depending only on {x0,y0,σ0,κ,γ,K,L,M,T}\{x_{0},y_{0},\sigma_{0},\kappa,\gamma,K,L,\,M,T\}.

Proof.

Lemma 2.1 gives us

dTV(Xtα,Xt)XtαXt1,2[3(𝔼D2Xt4)1/4(𝔼DXt8)1/4+2(𝔼DXt2)1/2].d_{TV}(X^{\alpha}_{t},X_{t})\leq\|X^{\alpha}_{t}-X_{t}\|_{1,2}\left[3\left(\mathbb{E}\|D^{2}X_{t}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|DX_{t}\|_{\mathcal{H}}^{-8}\right)^{1/4}+2\left(\mathbb{E}\|DX_{t}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right].

Thanks to Theorem 3.1 and Lemma 3.4, we obtain

dTV(Xtα,Xt)\displaystyle d_{TV}(X^{\alpha}_{t},X_{t}) C(λ(t,α(κ+γ))+(λ(t,ακ))2)\displaystyle\leq C\sqrt{(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})}
×[3(𝔼D2Xt4)1/4(𝔼DXt8)1/4+2(𝔼DXt2)1/2],\displaystyle\times\left[3\left(\mathbb{E}\|D^{2}X_{t}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|DX_{t}\|_{\mathcal{H}}^{-8}\right)^{1/4}+2\left(\mathbb{E}\|DX_{t}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right], (3.21)

where CC is a constant depending only on {x0,y0,κ,γ,K,L,M,T}\{x_{0},y_{0},\kappa,\gamma,K,L,\mathrm{M},T\}. Now, from (3.17), one sees that

DθXt=σ(θ,Xθ)κ+γexp(θt(g2(s,Xs)κ+γ12(κ+γ)2σ2(s,Xs)2)𝑑s+1κ+γθtσ2(s,Xs)𝑑Bs),D_{\theta}X_{t}=\dfrac{\sigma(\theta,X_{\theta})}{\kappa+\gamma}\exp\bigg{(}\int_{\theta}^{t}\bigg{(}\dfrac{g_{2}^{\prime}(s,X_{s})}{\kappa+\gamma}-\frac{1}{2(\kappa+\gamma)^{2}}\sigma_{2}^{\prime}(s,X_{s})^{2}\bigg{)}ds+\dfrac{1}{\kappa+\gamma}\int_{\theta}^{t}\sigma_{2}^{\prime}(s,X_{s})dB_{s}\bigg{)},

which implies that

|DθXt|2σ02(κ+γ)2e(2Lκ+γ+L2(κ+γ)2)Texp(2κ+γθtσ2(s,Xs)𝑑Bs),0θtT.|D_{\theta}X_{t}|^{2}\geq\dfrac{\sigma_{0}^{2}}{(\kappa+\gamma)^{2}}e^{-(\frac{2L}{\kappa+\gamma}+\frac{L^{2}}{(\kappa+\gamma)^{2}})T}\exp\bigg{(}\dfrac{2}{\kappa+\gamma}\int_{\theta}^{t}\sigma_{2}^{\prime}(s,X_{s})dB_{s}\bigg{)},\quad 0\leq\theta\leq t\leq T.

Define the stochastic process Mt:=0tσ2(s,Xs)𝑑Bs,0tT,M_{t}:=\int_{0}^{t}\sigma_{2}^{\prime}(s,X_{s})dB_{s},0\leq t\leq T, one derives that

DXt20t|DθXt|2𝑑θ\displaystyle\|DX_{t}\|^{2}_{\mathcal{H}}\geq\int\limits_{0}^{t}|D_{\theta}X_{t}|^{2}d\theta σ02(κ+γ)2e(2Lκ+γ+L2(κ+γ)2)Te2Mtκ+γ0te2Mθκ+γ𝑑θ\displaystyle\geq\dfrac{\sigma_{0}^{2}}{(\kappa+\gamma)^{2}}e^{-(\frac{2L}{\kappa+\gamma}+\frac{L^{2}}{(\kappa+\gamma)^{2}})T}e^{\frac{2M_{t}}{\kappa+\gamma}}\int_{0}^{t}e^{\frac{2M_{\theta}}{\kappa+\gamma}}d\theta
tσ02(κ+γ)2e(2Lκ+γ+L2(κ+γ)2)Te4κ+γmin0tTMt,\displaystyle\geq t\dfrac{\sigma_{0}^{2}}{(\kappa+\gamma)^{2}}e^{-(\frac{2L}{\kappa+\gamma}+\frac{L^{2}}{(\kappa+\gamma)^{2}})T}e^{\frac{4}{\kappa+\gamma}\min\limits_{0\leq t\leq T}M_{t}},

for 0tT0\leq t\leq T. We observe that MtM_{t} is a martingale with bounded quadratic variation. Indeed, Mt=0tσ2(s,Xs)2𝑑sL2T.\langle M\rangle_{t}=\int_{0}^{t}\sigma_{2}^{\prime}(s,X_{s})^{2}ds\leq L^{2}T. So, by Dubins and Schwarz’s theorem (see, e.g. Theorem 3.4.6 in [KSSS91]) there exists a Wiener process (W^t)t0(\widehat{W}_{t})_{t\geq 0} such that Mt=W^Mt.M_{t}=\widehat{W}_{\langle M\rangle_{t}}. Then, we arrive at the following

DXt2tσ02(κ+γ)2e(2Lκ+γ+L2(κ+γ)2)Te4κ+γmin0tL2TW^t,  0tT.\displaystyle\|DX_{t}\|^{2}_{\mathcal{H}}\geq t\dfrac{\sigma_{0}^{2}}{(\kappa+\gamma)^{2}}e^{-(\frac{2L}{\kappa+\gamma}+\frac{L^{2}}{(\kappa+\gamma)^{2}})T}e^{\frac{4}{\kappa+\gamma}\min\limits_{0\leq t\leq L^{2}T}\widehat{W}_{t}},\,\,0\leq t\leq T.

This implies that, for each r2r\geq 2 and 0<tT,0<t\leq T,

𝔼(DXtr)tr/2(κ+γ)rσ0re(rLκ+γ+rL22(κ+γ)2)T𝔼[e2rκ+γmax0tL2T(W^t)].\displaystyle\mathbb{E}\left(\|DX_{t}\|_{\mathcal{H}}^{-r}\right)\leq t^{-r/2}\dfrac{(\kappa+\gamma)^{r}}{\sigma_{0}^{r}}e^{(\frac{rL}{\kappa+\gamma}+\frac{rL^{2}}{2(\kappa+\gamma)^{2}})T}\mathbb{E}\left[e^{\frac{2r}{\kappa+\gamma}\max\limits_{0\leq t\leq L^{2}T}(-\widehat{W}_{t})}\right].

On the other hand, by Fernique’s theorem, we always have

𝔼[e2rκ+γmax0tL2T(W^t)]<.\mathbb{E}\left[e^{\frac{2r}{\kappa+\gamma}\max\limits_{0\leq t\leq L^{2}T}(-\widehat{W}_{t})}\right]<\infty.

Hence

𝔼(DXtr)Ctr/2.\mathbb{E}\left(\|DX_{t}\|_{\mathcal{H}}^{-r}\right)\leq Ct^{-r/2}. (3.22)

We observe that, from (3.2), for γ,θt,\gamma,\theta\leq t, under Assumptions 1.2,

DγDθXt\displaystyle D_{\gamma}D_{\theta}X_{t} =σ(θ,Xθ)DγXθ+σ(γ,Xθ)DθXγ\displaystyle=\sigma^{\prime}(\theta,X_{\theta})D_{\gamma}X_{\theta}+\sigma^{\prime}(\gamma,X_{\theta})D_{\theta}X_{\gamma}
+θγt(g′′(s,Xs)DθXsDγXs+g(s,Xs)DγDθXs)𝑑s\displaystyle\qquad+\int_{\theta\vee\gamma}^{t}\left(g^{\prime\prime}(s,X_{s})D_{\theta}X_{s}D_{\gamma}X_{s}+g^{\prime}(s,X_{s})D_{\gamma}D_{\theta}X_{s}\right)ds
+θγt(σ′′(s,Xs)DθXsDγXs+σ(s,Xs)DγDθXs)𝑑Ws,\displaystyle\qquad+\int_{\theta\vee\gamma}^{t}\left(\sigma^{\prime\prime}(s,X_{s})D_{\theta}X_{s}D_{\gamma}X_{s}+\sigma^{\prime}(s,X_{s})D_{\gamma}D_{\theta}X_{s}\right)dW_{s},

Now, using Minkowski’s inequality (3.1) with n=4n=4, Hölder’s inequality (3.2), the BDG inequality (3.3), Assumptions 1.1 and 1.2, we can deduce

𝔼|DγDθXt|4\displaystyle\mathbb{E}|D_{\gamma}D_{\theta}X_{t}|^{4} 64[L4𝔼|DγXθ|4+L4𝔼|DθXγ|4\displaystyle\leq 64\Bigg{[}L^{4}\mathbb{E}|D_{\gamma}X_{\theta}|^{4}+L^{4}\mathbb{E}|D_{\theta}X_{\gamma}|^{4}
+8(t3+C4t)θγt(M4(𝔼|DθXs|2)2(𝔼|DγXs|2)2+L4𝔼|DγDθXs|4)ds].\displaystyle\qquad+8(t^{3}+C_{4}t)\int_{\theta\vee\gamma}^{t}\left(M^{4}(\mathbb{E}|D_{\theta}X_{s}|^{2})^{2}(\mathbb{E}|D_{\gamma}X_{s}|^{2})^{2}+L^{4}\mathbb{E}|D_{\gamma}D_{\theta}X_{s}|^{4}\right)ds\Bigg{]}.

This, with together Lemma 3.3, gives us

𝔼[|DγDθXt|4]C+Cθγt𝔼[|DγDθXs|4]𝑑s,\displaystyle\mathbb{E}\left[|D_{\gamma}D_{\theta}X_{t}|^{4}\right]\leq C+C\int_{\theta\vee\gamma}^{t}\mathbb{E}\big{[}|D_{\gamma}D_{\theta}X_{s}|^{4}\big{]}\,ds,

where CC is a positive constant. By Gronwall’s inequality, we can verify that

𝔼[|DγDθXt|4]CeC(tθγ)C 0θ,γtT.\mathbb{E}\left[|D_{\gamma}D_{\theta}X_{t}|^{4}\right]\leq Ce^{C(t-\theta\vee\gamma)}\leq C\ \ \forall\ 0\leq\theta,\gamma\leq t\leq T.

Therefore,

𝔼D2Xt4\displaystyle\mathbb{E}\|D^{2}X_{t}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}} t20t0t𝔼[|DγDθXt|4]𝑑θ𝑑γt20t0tC𝑑θ𝑑γ=Ct4,\displaystyle\leq t^{2}\int_{0}^{t}\int_{0}^{t}\mathbb{E}\big{[}|D_{\gamma}D_{\theta}X_{t}|^{4}\big{]}\,d\theta\,d\gamma\leq t^{2}\int_{0}^{t}\int_{0}^{t}Cd\theta d\gamma=Ct^{4}, (3.23)

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}.

Combining (3.1), (3.22) and (3.23), we can conclude that

dTV(Xtα,Xt)\displaystyle d_{TV}(X^{\alpha}_{t},X_{t})\leq C(λ(t,α(κ+γ))+(λ(t,ακ))2)×[C+Ct1/2]\displaystyle C\sqrt{(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})}\times\left[C+Ct^{-1/2}\right]
Ct1(λ(t,α(κ+γ))+(λ(t,ακ))2),\displaystyle\leq C\sqrt{t^{-1}(\lambda(t,\alpha(\kappa+\gamma))+\left(\lambda(t,\alpha\kappa)\right)^{2})},

where CC is a constant depending only on {x0,y0,σ0,κ,γ,K,L,M,T}\{x_{0},y_{0},\sigma_{0},\kappa,\gamma,K,L,\,M,T\}. This completes the proof. ∎

From Theorem 3.2, together with the fact that for all t>0t>0 and a>0a>0, λ(t,a)<1a\lambda(t,a)<\dfrac{1}{a}, then we get the following Corollary.

Corollary 3.1.

Let {Xtα,t[0,T]}\{X^{\alpha}_{t},t\in[0,T]\} and {Xt,t[0,T]}\{X_{t},t\in[0,T]\} be, respectively, the solution of (2.6) and of(2.7) with Assumptions 1.1 and 1.2. We further assume that |σ(t,x)|σ0>0|\sigma(t,x)|\geq\sigma_{0}>0 for all (t,x)[0,T]×.(t,x)\in[0,T]\times\mathbb{R}. Then, for each α1\alpha\geq 1 and t(0,T],t\in(0,T],

dTV(Xtα,Xt)Ct1/2α1/2,d_{TV}(X^{\alpha}_{t},X_{t})\leq Ct^{-1/2}\alpha^{-1/2},

where CC is a constant depending only on {x0,y0,σ0,κ,γ,K,L,M,T}\{x_{0},y_{0},\sigma_{0},\kappa,\gamma,K,L,\,M,T\}.

3.2. Approximation the velocity and rescaled velocity processes

In this section, we establish rates of convergence in LpL^{p}-distances and in the total variation distance for the velocity and rescaled velocity processes. We will discuss the re-scaled velocity process first since in this case, our results are applicable to more general settings where both external forces and diffusion coefficients can be dependent on both xx and tt, i.e. g=g(t,x)g=g(t,x) and σ=σ(t,x)\sigma=\sigma(t,x).

3.2.1. The re-scaled velocity process

From the second equation of (1.1) we can see that the process YtααY^{\alpha}_{\frac{t}{\alpha}} satisfies

{Ytαα=y0(κ+γ)0tYsαα𝑑s0tg(sα,Xsαα)𝑑sγ0tE(Ysαα)𝑑s+α0tσ(sα,Xsαα)𝑑WsαX0α=x0.\begin{cases}Y^{\alpha}_{\frac{t}{\alpha}}=y_{0}-(\kappa+\gamma)\int_{0}^{t}Y^{\alpha}_{\frac{s}{\alpha}}ds-\int_{0}^{t}g(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})ds-\gamma\int_{0}^{t}E(Y^{\alpha}_{\frac{s}{\alpha}})ds+\alpha\int_{0}^{t}\sigma(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})dW_{\frac{s}{\alpha}}\\ X^{\alpha}_{0}=x_{0}.\end{cases} (3.24)

We recall the definition of the re-scaled velocity process introduced in the Introduction

Y~tα=1αYtαα.\tilde{Y}^{\alpha}_{t}=\dfrac{1}{\sqrt{\alpha}}Y^{\alpha}_{\frac{t}{\alpha}}.

Then Y~tα\tilde{Y}^{\alpha}_{t} satisfies (1.4), that is

{Y~tα=y0α(κ+γ)0tY~sα𝑑s1α0tg(sα,Xsαα)𝑑sγ0tE(Y~sα)𝑑s+α0tσ(sα,Xsαα)𝑑WsαXα(0)=x0.\begin{cases}\tilde{Y}^{\alpha}_{t}=\dfrac{y_{0}}{\sqrt{\alpha}}-(\kappa+\gamma)\int_{0}^{t}\tilde{Y}^{\alpha}_{s}ds-\dfrac{1}{\sqrt{\alpha}}\int_{0}^{t}g(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})ds-\gamma\int_{0}^{t}E(\tilde{Y}^{\alpha}_{s})ds+\sqrt{\alpha}\int_{0}^{t}\sigma(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})dW_{\frac{s}{\alpha}}\\ X^{\alpha}(0)=x_{0}.\end{cases} (3.25)

Now, we put W~t=αWt/α\tilde{W}_{t}=\sqrt{\alpha}W_{t/\alpha}, then (W~t)t0(\tilde{W}_{t})_{t\geq 0} is a Brownian motion process and (3.25) can be rewritten in the form

{Y~tα=y0α(κ+γ)0tY~sα𝑑s1α0tg(sα,Xsαα)𝑑sγ0tE(Y~sα)𝑑s+0tσ(sα,Xsαα)𝑑W~sX0α=x0.\begin{cases}\tilde{Y}^{\alpha}_{t}=\dfrac{y_{0}}{\sqrt{\alpha}}-(\kappa+\gamma)\int_{0}^{t}\tilde{Y}^{\alpha}_{s}ds-\dfrac{1}{\sqrt{\alpha}}\int_{0}^{t}g(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})ds-\gamma\int_{0}^{t}E(\tilde{Y}^{\alpha}_{s})ds+\int_{0}^{t}\sigma(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})d\tilde{W}_{s}\\ X^{\alpha}_{0}=x_{0}.\end{cases} (3.26)

Our goal in this section is to study the rate of convergence in LpL^{p}-distance and in the total variation distance between Y~tα\tilde{Y}^{\alpha}_{t} and Y~t\tilde{Y}_{t}. Here, Y~t\tilde{Y}_{t} is the solution of Ornstein-Uhlembeck process (1.5), which is

{dY~t=(κ+γ)dY~t+σ(0,x0)dW~t,Y~(0)=0.\begin{cases}d\tilde{Y}_{t}=-(\kappa+\gamma)d\tilde{Y}_{t}+\sigma(0,x_{0})d\tilde{W}_{t},\\ \tilde{Y}(0)=0.\end{cases} (3.27)

First, we obtain the rate of convergence in LpL^{p}-distances between Y~tα\tilde{Y}^{\alpha}_{t} and Y~t\tilde{Y}_{t} in the following lemma.

Theorem 3.3.

Let {Y~tα,t[0,T]}\{\tilde{Y}^{\alpha}_{t},t\in[0,T]\} and {Y~t,t[0,T]}\{\tilde{Y}_{t},t\in[0,T]\} be, respectively, the solution of (3.26) and of (3.27) with Assumptions 1.1 and 1.3. Then, for all p2p\geq 2 and α1\alpha\geq 1,

𝔼[sup0tT|Y~tαY~t|p]Cαp/2,\mathbb{E}\left[\sup_{0\leq t\leq T}|\tilde{Y}^{\alpha}_{t}-\tilde{Y}_{t}|^{p}\right]\leq\dfrac{C}{\alpha^{p/2}},

where CC is a positive constant depending on pp but not on α\alpha.

Proof.

From (3.26) and (3.27), together with the fact that 𝔼[Y~t]=0\mathbb{E}[\tilde{Y}_{t}]=0 for all t[0,T]t\in[0,T], we get

Y~tαY~t\displaystyle\tilde{Y}^{\alpha}_{t}-\tilde{Y}_{t} =y0α(κ+γ)0t(Y~sαY~s)𝑑s1α0tg(sα,Xsαα)𝑑s\displaystyle=\dfrac{y_{0}}{\sqrt{\alpha}}-(\kappa+\gamma)\int_{0}^{t}(\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s})ds-\dfrac{1}{\sqrt{\alpha}}\int_{0}^{t}g(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})ds
γ0tE(Y~sαY~s)𝑑s+0t(σ(sα,Xsαα)σ(0,x0))𝑑W~s.\displaystyle\qquad-\gamma\int_{0}^{t}E(\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s})ds+\int_{0}^{t}(\sigma(\frac{s}{\alpha},X^{\alpha}_{\frac{s}{\alpha}})-\sigma(0,x_{0}))d\tilde{W}_{s}.

Using Minkowski’s inequality (3.1) with n=5n=5, Hölder’s inequality, the BDG inequality and Assumptions 1.1 and 1.3, one can derive that

𝔼[sup0st|Y~sαY~s|p]\displaystyle\mathbb{E}\left[\sup_{0\leq s\leq t}|\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s}|^{p}\right] 5p1[|y0|pαp/2+tp1(κ+γ)p0t𝔼[|Y~sαY~s|p]ds\displaystyle\leq 5^{p-1}\Big{[}\dfrac{|y_{0}|^{p}}{\alpha^{p/2}}+t^{p-1}(\kappa+\gamma)^{p}\int_{0}^{t}\mathbb{E}\big{[}|\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s}|^{p}\big{]}\,ds
+Kp(2t)p1αp/20t(1+𝔼[|Xsαα|p])𝑑s+tp1γp0t𝔼[|Y~sαY~s|p]𝑑s\displaystyle\qquad\quad+\dfrac{K^{p}(2t)^{p-1}}{\alpha^{p/2}}\int_{0}^{t}(1+\mathbb{E}\big{[}|X^{\alpha}_{\frac{s}{\alpha}}|^{p}\big{]})\,ds+t^{p-1}\gamma^{p}\int_{0}^{t}\mathbb{E}\big{[}|\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s}|^{p}\big{]}\,ds
+2p1Cptp/210t(spαp+𝔼[|Xsααx0|p)]ds].\displaystyle\qquad\quad+2^{p-1}C_{p}t^{p/2-1}\int_{0}^{t}(\dfrac{s^{p}}{\alpha^{p}}+\mathbb{E}\big{[}|X^{\alpha}_{\frac{s}{\alpha}}-x_{0}|^{p})\big{]}\,ds\Big{]}.

By Lemma 3.1, with noting that 0sαstT0\leq\frac{s}{\alpha}\leq s\leq t\leq T, we have

𝔼[sup0st|Y~sαY~s|p]\displaystyle\mathbb{E}\left[\sup_{0\leq s\leq t}|\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s}|^{p}\right] Cαp/2+C0t𝔼[|Y~sαY~s|p]𝑑s+C0t(spαp+sp/2αp/2)𝑑s\displaystyle\leq\dfrac{C}{\alpha^{p/2}}+C\int_{0}^{t}\mathbb{E}\left[|\tilde{Y}^{\alpha}_{s}-\tilde{Y}_{s}|^{p}\right]\,ds+C\int_{0}^{t}(\dfrac{s^{p}}{\alpha^{p}}+\dfrac{s^{p/2}}{\alpha^{p/2}})ds
Cαp/2+C0t𝔼[sup0us|Y~uαY~u|p]𝑑s,\displaystyle\leq\dfrac{C}{\alpha^{p/2}}+C\int_{0}^{t}\mathbb{E}\left[\sup_{0\leq u\leq s}|\tilde{Y}^{\alpha}_{u}-\tilde{Y}_{u}|^{p}\right]\,ds,

where CC is constant depending only on {x0,y0,κ,γ,K,L,p}\{x_{0},y_{0},\kappa,\gamma,K,L,p\}. Using Growwall’s inequality, we obtain the claimed inequality and complete the proof. ∎

From (3.25) and (3.27), under Assumptions 1.1 the Malliavin differentiability of the solutions Y~tα\tilde{Y}^{\alpha}_{t} and Y~t\tilde{Y}_{t} can be proved by using the same method as in the proof of Lemma 3.2. Moreover, the Malliavin derivatives DθY~tαD_{\theta}\tilde{Y}^{\alpha}_{t} and DθY~tD_{\theta}\tilde{Y}_{t} satisfy DrY~tα=DrY~t=0D_{r}\tilde{Y}^{\alpha}_{t}=D_{r}\tilde{Y}_{t}=0 for rt/αr\geq t/\alpha and 0αr<tT,0\leq\alpha r<t\leq T,

DrY~tα\displaystyle D_{r}\tilde{Y}^{\alpha}_{t} =ασ(r,Xrα)(κ+γ)αrtDrY~sα𝑑s1ααrtg¯(sα)DrXsαα𝑑s+ααrtσ¯(sα)DrXsαα𝑑Wsα,\displaystyle=\sqrt{\alpha}\sigma(r,X^{\alpha}_{r})-(\kappa+\gamma)\int_{\alpha r}^{t}D_{r}\tilde{Y}^{\alpha}_{s}ds-\dfrac{1}{\sqrt{\alpha}}\int_{\alpha r}^{t}\bar{g}(\frac{s}{\alpha})D_{r}X^{\alpha}_{\frac{s}{\alpha}}ds+\sqrt{\alpha}\int_{\alpha r}^{t}\bar{\sigma}(\frac{s}{\alpha})D_{r}X^{\alpha}_{\frac{s}{\alpha}}dW_{\frac{s}{\alpha}},
DrY~t\displaystyle D_{r}\tilde{Y}_{t} =ασ(0,x0)(κ+γ)αrtDrY~s𝑑s,\displaystyle=\sqrt{\alpha}\sigma(0,x_{0})-(\kappa+\gamma)\int_{\alpha r}^{t}D_{r}\tilde{Y}_{s}ds, (3.28)

where g¯(s),g¯α(s),σ¯(s),σ¯α(s)\bar{g}(s),\bar{g}^{\alpha}(s),\bar{\sigma}(s),\bar{\sigma}^{\alpha}(s) are adapted stochastic processes and bounded by L.L. Furthermore, if gg and σ\sigma are continuously differentiable, then g¯(s)=g2(s,Xs)\bar{g}(s)=g^{\prime}_{2}(s,X_{s}), g¯α(s)=g2(s,Xsα),σ¯(s)=σ2(s,Xs)\bar{g}^{\alpha}(s)=g^{\prime}_{2}(s,X_{s}^{\alpha}),\bar{\sigma}(s)=\sigma^{\prime}_{2}(s,X_{s}) and σ¯α(s)=σ2(s,Xsα).\bar{\sigma}^{\alpha}(s)=\sigma^{\prime}_{2}(s,X_{s}^{\alpha}).

Lemma 3.5.

Let {Y~tα,t[0,T]}\{\tilde{Y}^{\alpha}_{t},t\in[0,T]\} and {Y~t,t[0,T]}\{\tilde{Y}_{t},t\in[0,T]\} be defined as above. Assume that Assumptions 1.1 and 1.3 hold. Then, for all α1\alpha\geq 1,

𝔼[DY~tαDY~t2]Cα1,\mathbb{E}\left[\|D\tilde{Y}^{\alpha}_{t}-D\tilde{Y}_{t}\|_{\mathcal{H}}^{2}\right]\leq C\alpha^{-1},

where CC is constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}.

Proof.

It follows from (3.2.1) that, for 0αr<t,0\leq\alpha r<t,

DY~tαDY~t\displaystyle D\tilde{Y}^{\alpha}_{t}-D\tilde{Y}_{t} =α(σ(r,Xrα)σ(0,x0))(κ+γ)αrt[DY~sαDY~s]𝑑s\displaystyle=\sqrt{\alpha}\left(\sigma(r,X^{\alpha}_{r})-\sigma(0,x_{0})\right)-(\kappa+\gamma)\int_{\alpha r}^{t}\left[D\tilde{Y}^{\alpha}_{s}-D\tilde{Y}_{s}\right]ds
1ααrtg¯(sα)DrXsαα𝑑s+ααrtσ¯(sα)DrXsαα𝑑Wsα.\displaystyle\qquad-\dfrac{1}{\sqrt{\alpha}}\int_{\alpha r}^{t}\bar{g}(\frac{s}{\alpha})D_{r}X^{\alpha}_{\frac{s}{\alpha}}ds+\sqrt{\alpha}\int_{\alpha r}^{t}\bar{\sigma}(\frac{s}{\alpha})D_{r}X^{\alpha}_{\frac{s}{\alpha}}dW_{\frac{s}{\alpha}}.

Using Cauchy-Schwarz inequality, the Itô isometry formula, Assumptions 1.1 and 1.3, Lemma 3.1, with noting that 0sαstT0\leq\frac{s}{\alpha}\leq s\leq t\leq T, we get

𝔼[|DrY~tαDrY~t|2]\displaystyle\mathbb{E}\left[|D_{r}\tilde{Y}^{\alpha}_{t}-D_{r}\tilde{Y}_{t}|^{2}\right] 4(2α(r2+𝔼[|Xrαx0|2])+(κ+γ)2(tαr)αrt𝔼[|DrY~sαDrY~s|2]ds\displaystyle\leq 4\Big{(}2\alpha\left(r^{2}+\mathbb{E}\left[|X^{\alpha}_{r}-x_{0}|^{2}\right]\right)+(\kappa+\gamma)^{2}(t-\alpha r)\int_{\alpha r}^{t}\mathbb{E}\Big{[}\big{|}D_{r}\tilde{Y}^{\alpha}_{s}-D_{r}\tilde{Y}_{s}\big{|}^{2}\Big{]}\,ds
+L2Tααrt𝔼[|DrXsαα|2]ds+L2ααrt𝔼[|DrXsαα|2]dsα)\displaystyle\qquad+\dfrac{L^{2}T}{\alpha}\int_{\alpha r}^{t}\mathbb{E}\big{[}|D_{r}X^{\alpha}_{\frac{s}{\alpha}}|^{2}\big{]}\,ds+L^{2}\alpha\int_{\alpha r}^{t}\mathbb{E}\big{[}|D_{r}X^{\alpha}_{\frac{s}{\alpha}}|^{2}\big{]}\frac{ds}{\alpha}\Big{)}
C(α(r2+r)+αrt𝔼[|DrY~sαDrY~s|2]𝑑s+(tαr)α+(tαr)),\displaystyle\leq C\Big{(}\alpha\left(r^{2}+r\right)+\int_{\alpha r}^{t}\mathbb{E}\Big{[}\big{|}D_{r}\tilde{Y}^{\alpha}_{s}-D_{r}\tilde{Y}_{s}\big{|}^{2}\Big{]}\,ds+\dfrac{(t-\alpha r)}{\alpha}+(t-\alpha r)\Big{)},

where CC is constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}. From this, we have

0t/α𝔼[|DrY~tαDrY~t|2]𝑑r\displaystyle\int_{0}^{t/\alpha}\mathbb{E}\big{[}|D_{r}\tilde{Y}^{\alpha}_{t}-D_{r}\tilde{Y}_{t}|^{2}\big{]}\,dr
C(α(t33α3+t22α2)+0t/α(αrt𝔼[|DrY~sαDrY~s|2]𝑑s)𝑑r+t22α2+t22α)\displaystyle\qquad\leq C\Big{(}\alpha\left(\frac{t^{3}}{3\alpha^{3}}+\frac{t^{2}}{2\alpha^{2}}\right)+\int_{0}^{t/\alpha}\left(\int_{\alpha r}^{t}\mathbb{E}\Big{[}\big{|}D_{r}\tilde{Y}^{\alpha}_{s}-D_{r}\tilde{Y}_{s}\big{|}^{2}\Big{]}\,ds\right)dr+\dfrac{t^{2}}{2\alpha^{2}}+\dfrac{t^{2}}{2\alpha}\Big{)}
Cα+C0t(0s/αE|DrY~sαDrY~s|2𝑑r)𝑑s.\displaystyle\qquad\leq\dfrac{C}{\alpha}+C\int_{0}^{t}\left(\int_{0}^{s/\alpha}E\left|D_{r}\tilde{Y}^{\alpha}_{s}-D_{r}\tilde{Y}_{s}\right|^{2}dr\right)ds.

Denote ϕ(t)=0t/αE|DrY~tαDrY~t|2𝑑r\phi(t)=\int_{0}^{t/\alpha}E|D_{r}\tilde{Y}^{\alpha}_{t}-D_{r}\tilde{Y}_{t}|^{2}dr, using Gronwall’s inequality, one sees easily that

𝔼[DY~tαDY~t2]=ϕ(t)CαeCtCα1,\mathbb{E}\Big{[}\|D\tilde{Y}^{\alpha}_{t}-D\tilde{Y}_{t}\|_{\mathcal{H}}^{2}\Big{]}=\phi(t)\leq\dfrac{C}{\alpha}e^{Ct}\leq C\alpha^{-1},

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}. This completes our proof. ∎

Bringing the above lemmas together, we can get the following result.

Theorem 3.4.

Let {Y~tα,t[0,T]}\{\tilde{Y}^{\alpha}_{t},t\in[0,T]\} and {Y~t,t[0,T]}\{\tilde{Y}_{t},t\in[0,T]\} be as before. Assume that Assumptions 1.1 and 1.3 hold and |σ(0,x0)|>0|\sigma(0,x_{0})|>0 for all (t,x)[0,T]×.(t,x)\in[0,T]\times\mathbb{R}. Then, for each t(0,T],t\in(0,T],

dTV(Y~tα,Y~t)C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}(\tilde{Y}^{\alpha}_{t},\tilde{Y}_{t})\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T,σ(0,x0)}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma(0,x_{0})\}.

Proof.

Using Lemma 2.1 we have

dTV(Y~tα,Y~t)Y~tαY~t1,2[3(𝔼D2Y~t4)1/4(𝔼DY~t8)1/4+2(𝔼DY~t2)1/2].d_{TV}(\tilde{Y}^{\alpha}_{t},\tilde{Y}_{t})\leq\|\tilde{Y}^{\alpha}_{t}-\tilde{Y}_{t}\|_{1,2}\left[3\left(\mathbb{E}\|D^{2}\tilde{Y}_{t}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|D\tilde{Y}_{t}\|_{\mathcal{H}}^{-8}\right)^{1/4}+2\left(\mathbb{E}\|D\tilde{Y}_{t}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right].

Thanks to Theorem 3.3 and Lemma 3.5, we obtain

dTV(Y~tα,Y~t)Cα1[3(𝔼D2Y~t4)1/4(𝔼DY~t8)1/4+2(𝔼DY~t2)1/2].d_{TV}(\tilde{Y}^{\alpha}_{t},\tilde{Y}_{t})\leq C\alpha^{-1}\left[3\left(\mathbb{E}\|D^{2}\tilde{Y}_{t}\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|D\tilde{Y}_{t}\|_{\mathcal{H}}^{-8}\right)^{1/4}+2\left(\mathbb{E}\|D\tilde{Y}_{t}\|_{\mathcal{H}}^{-2}\right)^{1/2}\right]. (3.29)

Moreover, we have DrY~t=0D_{r}\tilde{Y}_{t}=0 for rt/αr\geq t/\alpha and DrY~t=ασ(0,x0)(κ+γ)αrtDrY~s𝑑s,D_{r}\tilde{Y}_{t}=\sqrt{\alpha}\sigma(0,x_{0})-(\kappa+\gamma)\int_{\alpha r}^{t}D_{r}\tilde{Y}_{s}ds, for 0αr<t.0\leq\alpha r<t. Solving this ODE directly yields

DrY~t={0if rαtασ(0,x0)exp[(κ+γ)(rαt)]if rα<t.D_{r}\tilde{Y}_{t}=\begin{cases}0\quad&\text{if }r\alpha\geq t\\ \sqrt{\alpha}\sigma(0,x_{0})\exp{[(\kappa+\gamma)(r\alpha-t)]}\quad&\text{if }r\alpha<t.\end{cases} (3.30)

Thus, one can easily show that

DY~t2=ασ2(0,x0)0t/αexp[2(κ+γ)(rαt)]𝑑r=σ2(0,x0)λ(t,2(κ+γ)).\|D\tilde{Y}_{t}\|^{2}_{\mathcal{H}}=\alpha\sigma^{2}(0,x_{0})\int_{0}^{t/\alpha}\exp{[2(\kappa+\gamma)(r\alpha-t)]}dr=\sigma^{2}(0,x_{0})\lambda(t,2(\kappa+\gamma)).

This implies, for all p2p\geq 2,

𝔼[DY~tp]=σp(0,x0)(λ(t,2(κ+γ)))p/2.\mathbb{E}\big{[}\|D\tilde{Y}_{t}\|^{-p}_{\mathcal{H}}\big{]}=\sigma^{-p}(0,x_{0})\left(\lambda(t,2(\kappa+\gamma))\right)^{-p/2}. (3.31)

Applying the Malliavin derivative to (3.30), we have

DθDrY~t=0.D_{\theta}D_{r}\tilde{Y}_{t}=0. (3.32)

Combining (3.29), (3.31) and (3.32), we obtain

dTV(Y~tα,Y~t)C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}(\tilde{Y}^{\alpha}_{t},\tilde{Y}_{t})\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T,σ(0,x0)}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma(0,x_{0})\}. This completes our proof. ∎

3.2.2. The velocity process

As mentioned in the introduction, when g(t,x)=g(x)g(t,x)=g(x) and σ(t,x)=δ\sigma(t,x)=\delta, [Nar94, Theorem 2.3] shows that the velocity process YtαY^{\alpha}_{t} converges to the normal distribution as α\alpha\to\infty. In the rest of this section, we generalize this result to a much more general setting where gg depends on both xx and tt while σ\sigma depends only on tt, i.e. σ(t,x)=σ(t)\sigma(t,x)=\sigma(t).

From (2.2) we get

Ytα=y0exp[α(κ+γ)t]α(κ+γ)I1α(t)+α(κ+γ)I2α(t)+α(κ+γ)I3α(t).Y^{\alpha}_{t}=y_{0}\exp{[-\alpha(\kappa+\gamma)t]}-\alpha(\kappa+\gamma)I_{1}^{\alpha}(t)+\alpha(\kappa+\gamma)I_{2}^{\alpha}(t)+\alpha(\kappa+\gamma)I_{3}^{\alpha}(t). (3.33)

Since XtαX^{\alpha}_{t} is Malliavin differentiable, YtαY^{\alpha}_{t} is also Malliavin differentiable satisfying DrYtα=0D_{r}Y^{\alpha}_{t}=0 for rtr\geq t, and for 0r<tT0\leq r<t\leq T

DrYtα=ασ(r)exp[α(κ+γ)(rt)]αrtexp[α(κ+γ)(ts)]g(s,Xsα)DrXsα𝑑s.D_{r}Y^{\alpha}_{t}=\alpha\sigma(r)\exp{[\alpha(\kappa+\gamma)(r-t)]}-\alpha\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(t-s)]}g^{\prime}(s,X^{\alpha}_{s})D_{r}X^{\alpha}_{s}\,ds. (3.34)

Define

Wα(t)=α(κ+γ)I3α(t)=α0texp[α(κ+γ)(ut)]σ(u)𝑑Wu.W^{\alpha}(t)=\sqrt{\alpha}(\kappa+\gamma)I_{3}^{\alpha}(t)=\sqrt{\alpha}\int_{0}^{t}\exp{[\alpha(\kappa+\gamma)(u-t)]}\sigma(u)dW_{u}.

Then Wα(t)W^{\alpha}(t) is also Malliavin differentiable and DrWtα=0D_{r}W^{\alpha}_{t}=0 for rtr\geq t and for 0r<tT0\leq r<t\leq T

DrWα(t)=ασ(r)exp[α(κ+γ)(rt)].D_{r}W^{\alpha}(t)=\sqrt{\alpha}\sigma(r)\exp{[\alpha(\kappa+\gamma)(r-t)]}. (3.35)
Lemma 3.6.

Let YtαY^{\alpha}_{t} be the solution of (3.24) with Assumptions 1.1. Then, for all p2p\geq 2 and t(0,T],t\in(0,T],

𝔼[|YtααWα(t)|p]Cαp/2,\mathbb{E}\left[\Big{|}\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}}-W^{\alpha}(t)\Big{|}^{p}\right]\leq C\alpha^{-p/2},

where CC is constant depending only on {y0,κ,γ,K,L,p}\{y_{0},\kappa,\gamma,K,L,p\}.

Proof.

From (3.33), we get

YtααWα(t)=y0αexp[α(κ+γ)t]α(κ+γ)I1α(t)+α(κ+γ)I2α(t).\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}}-W^{\alpha}(t)=\dfrac{y_{0}}{\sqrt{\alpha}}\exp{[-\alpha(\kappa+\gamma)t]}-\sqrt{\alpha}(\kappa+\gamma)I_{1}^{\alpha}(t)+\sqrt{\alpha}(\kappa+\gamma)I_{2}^{\alpha}(t).

Then, by (3.1), (3.15) and Lemma 3.3, we have the following estimation

𝔼[|YtααWα(t)|p]3p1[|y0|pαp/2+αp/2(κ+γ)p𝔼(|I1α(t)|p)+αp/2(κ+γ)p|I2α(t)|p]C[1αp/2+αp/2(2+𝔼(sup0tT|Xtα|p))(λ(t;(κ+γ)α))p+αp/2(λ(t,ακ))p]Cαp/2[1+𝔼(sup0tT|Xtα|p)]Cαp/2,\begin{split}\mathbb{E}\left[\Big{|}\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}}-W^{\alpha}(t)\Big{|}^{p}\right]&\leq 3^{p-1}\left[\dfrac{|y_{0}|^{p}}{\alpha^{p/2}}+\alpha^{p/2}(\kappa+\gamma)^{p}\mathbb{E}\big{(}|I_{1}^{\alpha}(t)|^{p}\big{)}+\alpha^{p/2}(\kappa+\gamma)^{p}|I_{2}^{\alpha}(t)|^{p}\right]\\ &\leq C\left[\dfrac{1}{\alpha^{p/2}}+\alpha^{p/2}\left(2+\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|X^{\alpha}_{t}|^{p}\Big{)}\right)(\lambda(t;(\kappa+\gamma)\alpha))^{p}+\alpha^{p/2}\left(\lambda(t,\alpha\kappa)\right)^{p}\right]\\ &\leq\dfrac{C}{\alpha^{p/2}}\left[1+\mathbb{E}\Big{(}\sup_{0\leq t\leq T}|X^{\alpha}_{t}|^{p}\Big{)}\right]\\ &\leq\dfrac{C}{\alpha^{p/2}},\end{split}

where CC is a constant depending only on {y0,κ,γ,K,L}\{y_{0},\kappa,\gamma,K,L\}. This completes the proof of the lemma. ∎

Lemma 3.7.

Let YtαY^{\alpha}_{t} be the solution of (3.24) with Assumptions 1.1. Assume that σ(t)\sigma(t) is continuous on [0,T][0,T]. Then, for all α1\alpha\geq 1,

𝔼0t|DrYtααDrWα(t)|p𝑑rCσpαp/2,\mathbb{E}\int_{0}^{t}\left|\dfrac{D_{r}Y^{\alpha}_{t}}{\sqrt{\alpha}}-D_{r}W^{\alpha}(t)\right|^{p}dr\leq C\|\sigma\|^{p}_{\infty}\alpha^{-p/2},

where σ=supt[0,T]|σ(t)|\|\sigma\|_{\infty}=\sup\limits_{t\in[0,T]}|\sigma(t)| and CC is constant depending only on {κ,γ,K,L,p,T}\{\kappa,\gamma,K,L,p,T\}.

Proof.

From (3.34) and (3.35) we have

𝔼0t|DrYtααDrWα(t)|p𝑑r\displaystyle\mathbb{E}\int_{0}^{t}\left|\dfrac{D_{r}Y^{\alpha}_{t}}{\sqrt{\alpha}}-D_{r}W^{\alpha}(t)\right|^{p}dr =𝔼0t|αrtexp[α(κ+γ)(ts)]g(s,Xsα)DrXsα𝑑s|p𝑑r\displaystyle=\mathbb{E}\int_{0}^{t}\left|\sqrt{\alpha}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(t-s)]}g^{\prime}(s,X^{\alpha}_{s})D_{r}X^{\alpha}_{s}ds\right|^{p}\,dr
Cαp/20t𝔼|rtexp[α(κ+γ)(ts)]|DrXsα|ds|p𝑑r.\displaystyle\leq C\alpha^{p/2}\int_{0}^{t}\mathbb{E}\left|\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(t-s)]}|D_{r}X^{\alpha}_{s}|ds\right|^{p}dr. (3.36)

On the other hand, from (3.2), together with Assumptions 1.1 and the fact that σ¯α(s)=0,\bar{\sigma}^{\alpha}(s)=0, we have

|DrXtα|\displaystyle|D_{r}X^{\alpha}_{t}| |σ(r)|κ+γ(1exp[α(κ+γ)(rt)])+1κ+γrt|g¯α(s)||DrXsα|𝑑s\displaystyle\leq\dfrac{|\sigma(r)|}{\kappa+\gamma}(1-\exp{[\alpha(\kappa+\gamma)(r-t)]})+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}|\bar{g}^{\alpha}(s)||D_{r}X_{s}^{\alpha}|ds
+1κ+γrtexp[α(κ+γ)(st)]|b¯α(s)||DrXsα|𝑑s\displaystyle\qquad+\dfrac{1}{\kappa+\gamma}\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(s-t)]}|\bar{b}^{\alpha}(s)||D_{r}X^{\alpha}_{s}|ds
σκ+γ+2Lκ+γrt|DrXsα|𝑑s.\displaystyle\leq\dfrac{\|\sigma\|_{\infty}}{\kappa+\gamma}+\dfrac{2L}{\kappa+\gamma}\int_{r}^{t}|D_{r}X_{s}^{\alpha}|\,ds.

Thus, by Gronwall’s inequality one sees that

|DrXtα|σκ+γe2L(tr)κ+γCσ,|D_{r}X^{\alpha}_{t}|\leq\dfrac{\|\sigma\|_{\infty}}{\kappa+\gamma}e^{\frac{2L(t-r)}{\kappa+\gamma}}\leq C\|\sigma\|_{\infty},

where CC is constant depending only on {κ,γ,K,L,T}\{\kappa,\gamma,K,L,T\}. Substituting this into (3.2.2) yields

𝔼0t|DrYtααDrWα(t)|p𝑑r\displaystyle\mathbb{E}\int_{0}^{t}\left|\dfrac{D_{r}Y^{\alpha}_{t}}{\sqrt{\alpha}}-D_{r}W^{\alpha}(t)\right|^{p}dr Cσpαp/20t|rtexp[α(κ+γ)(ts)]𝑑s|p𝑑r\displaystyle\leq C\|\sigma\|^{p}_{\infty}\alpha^{p/2}\int_{0}^{t}\left|\int_{r}^{t}\exp{[\alpha(\kappa+\gamma)(t-s)]}ds\right|^{p}dr
Cσpαp/20tλp(tr;α(κ+γ))𝑑r\displaystyle\leq C\|\sigma\|^{p}_{\infty}\alpha^{p/2}\int_{0}^{t}\lambda^{p}(t-r;\alpha(\kappa+\gamma))dr
Cσpαp/2,\displaystyle\leq C\|\sigma\|^{p}_{\infty}\alpha^{-p/2},

which is the desired conclusion. ∎

Now we are ready to prove the rate of convergence in total variation distance for the velocity process YtαY^{\alpha}_{t} as α\alpha\to\infty.

Theorem 3.5.

Let YtαY^{\alpha}_{t} be the solution of (3.24) with Assumptions 1.1. Assume that σ(t)\sigma(t) is a continuously differentiable function on [0,T][0,T] and that σ(t)0\sigma(t)\not=0 for each t(0,T]t\in(0,T]. Then, for each α1\alpha\geq 1 and t(0,T]t\in(0,T]

dTV(Ytαα,N)C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},N\right)\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where NN is a normal random variable with mean 0 and variance σ2(t)2(κ+γ)\dfrac{\sigma^{2}(t)}{2(\kappa+\gamma)}, and CC is a constant depending only on {x0,y0,κ,γ,K,L,T,σ}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma\}.

Proof.

Using Lemma 2.1, we have:

dTV(Ytαα,Wα(t))YtααWα(t)1,2\displaystyle d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},W^{\alpha}(t)\right)\leq\left\|\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}}-W^{\alpha}(t)\right\|_{1,2} [3(𝔼D2Wα(t)4)1/4(𝔼DWα(t)8)1/4\displaystyle\Bigg{[}3\left(\mathbb{E}\|D^{2}W^{\alpha}(t)\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|DW^{\alpha}(t)\|_{\mathcal{H}}^{-8}\right)^{1/4}
+2(𝔼DWα(t)2)1/2].\displaystyle\qquad\qquad+2\left(\mathbb{E}\|DW^{\alpha}(t)\|_{\mathcal{H}}^{-2}\right)^{1/2}\Bigg{]}.

By Lemmas 3.6 and 3.7, one can derive that

dTV(Ytαα,Wα(t))Cα1/2\displaystyle d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},W^{\alpha}(t)\right)\leq C\alpha^{-1/2} [3(𝔼D2Wα(t)4)1/4(𝔼DWα(t)8)1/4\displaystyle\Bigg{[}3\left(\mathbb{E}\|D^{2}W^{\alpha}(t)\|^{4}_{\mathcal{H}\bigotimes\mathcal{H}}\right)^{1/4}\left(\mathbb{E}\|DW^{\alpha}(t)\|_{\mathcal{H}}^{-8}\right)^{1/4}
+2(𝔼DWα(t)2)1/2],\displaystyle\ \ \qquad+2\left(\mathbb{E}\|DW^{\alpha}(t)\|_{\mathcal{H}}^{-2}\right)^{1/2}\Bigg{]}, (3.37)

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T}\{x_{0},y_{0},\kappa,\gamma,K,L,T\}.

Now we calculate the derivatives of Wα(t)W^{\alpha}(t). One can easily show that

DrWα(t)=ασ(r)exp[α(κ+γ)(rt)]andDθDrWα(t)=0.D_{r}W^{\alpha}(t)=\sqrt{\alpha}\sigma(r)\exp{[\alpha(\kappa+\gamma)(r-t)]}\ \mbox{and}\ \ D_{\theta}D_{r}W^{\alpha}(t)=0. (3.38)

Therefore,

DWα(t)2\displaystyle\|DW^{\alpha}(t)\|_{\mathcal{H}}^{2} =0tσ2(r)αexp[2α(κ+γ)(rt)]𝑑r\displaystyle=\int_{0}^{t}\sigma^{2}(r)\alpha\exp{[2\alpha(\kappa+\gamma)(r-t)]}dr
αλ(t;2α(κ+γ))mint[0,T]|σ(t)|2\displaystyle\geq\alpha\lambda(t;2\alpha(\kappa+\gamma))\min\limits_{t\in[0,T]}|\sigma(t)|^{2}
λ(t;2(κ+γ))mint[0,T]|σ(t)|2.\displaystyle\geq\lambda(t;2(\kappa+\gamma))\min\limits_{t\in[0,T]}|\sigma(t)|^{2}.

Thus, for all p2p\geq 2,

𝔼[DY~tp]1λp/2(t;2(κ+γ))mint[0,T]|σ(t)|p.\mathbb{E}\big{[}\|D\tilde{Y}_{t}\|^{-p}_{\mathcal{H}}\big{]}\leq\dfrac{1}{\lambda^{p/2}(t;2(\kappa+\gamma))\min\limits_{t\in[0,T]}|\sigma(t)|^{p}}. (3.39)

From (3.2.2), (3.38), (3.39), we obtain

dTV(Ytαα,Wα(t))Cα1/2(2(λ(t;2(κ+γ)))1/2mint[0,T]|σ(t)|),d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},W^{\alpha}(t)\right)\leq C\alpha^{-1/2}\left(\dfrac{2}{(\lambda(t;2(\kappa+\gamma)))^{1/2}\min\limits_{t\in[0,T]}|\sigma(t)|}\right),

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T,σ}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma\}. Thus,

dTV(Ytαα,Wα(t))C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},W^{\alpha}(t)\right)\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where CC is a constant depending only on {x0,y0,κ,γ,K,L,T,σ}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma\}.

Note that by Itô’s isometry and using integration by parts for the non-stochastic integral, we have

𝔼[[Wα(t)]2]\displaystyle\mathbb{E}\big{[}\left[W^{\alpha}(t)\right]^{2}\big{]} =0tασ2(r)exp[2α(κ+γ)(rt)]𝑑r\displaystyle=\int_{0}^{t}\alpha\sigma^{2}(r)\exp{[2\alpha(\kappa+\gamma)(r-t)]}dr
=σ2(t)2(κ+γ)σ2(0)exp[2α(κ+γ)t]2(κ+γ)0tσ(r)σ(r)exp[2α(κ+γ)(rt)]𝑑r.\displaystyle=\dfrac{\sigma^{2}(t)}{2(\kappa+\gamma)}-\dfrac{\sigma^{2}(0)\exp{[-2\alpha(\kappa+\gamma)t]}}{2(\kappa+\gamma)}-\int_{0}^{t}\sigma(r)\sigma^{\prime}(r)\exp{[2\alpha(\kappa+\gamma)(r-t)]}dr.

Thus, we can deduce that Wα(t)W^{\alpha}(t) is random variable with normal distribution with mean 0 and variance

σ2(t)2(κ+γ)σ2(0)exp[2α(κ+γ)t]2(κ+γ)0tσ(r)σ(r)exp[2α(κ+γ)(rt)]𝑑r.\dfrac{\sigma^{2}(t)}{2(\kappa+\gamma)}-\dfrac{\sigma^{2}(0)\exp{[-2\alpha(\kappa+\gamma)t]}}{2(\kappa+\gamma)}-\int_{0}^{t}\sigma(r)\sigma^{\prime}(r)\exp{[2\alpha(\kappa+\gamma)(r-t)]}dr.

Now, applying Lemma 4.9, [Kla07], we derive that

dTV(Wα(t),N)\displaystyle d_{TV}(W^{\alpha}(t),N) C(σ2(0)exp[2α(κ+γ)t]2(κ+γ)+0t|σ(r)σ(r)|exp[2α(κ+γ)(rt)]𝑑r)\displaystyle\leq C\left(\dfrac{\sigma^{2}(0)\exp{[-2\alpha(\kappa+\gamma)t]}}{2(\kappa+\gamma)}+\int_{0}^{t}|\sigma(r)\sigma^{\prime}(r)|\exp{[2\alpha(\kappa+\gamma)(r-t)]}dr\right)
C(σ2(0)exp[2α(κ+γ)t]2(κ+γ)+σσλ(t,2α(κ+γ)))\displaystyle\leq C\left(\dfrac{\sigma^{2}(0)\exp{[-2\alpha(\kappa+\gamma)t]}}{2(\kappa+\gamma)}+\|\sigma\|_{\infty}\|\sigma^{\prime}\|_{\infty}\lambda(t,2\alpha(\kappa+\gamma))\right)
Cσσα1,\displaystyle\leq C\|\sigma\|_{\infty}\|\sigma^{\prime}\|_{\infty}\alpha^{-1},

where σ=supt[0,T]|σ(t)|\|\sigma\|_{\infty}=\sup\limits_{t\in[0,T]}|\sigma(t)|, σ=supt[0,T]|σ(t)|\|\sigma^{\prime}\|_{\infty}=\sup\limits_{t\in[0,T]}|\sigma^{\prime}(t)| and CC is an universal constant. Thus,

dTV(Ytαα,N)dTV(Wα(t),N)+dTV(Ytαα,Wα(t))C(λ(t,2(κ+γ)))1/2α1/2,d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},N\right)\leq d_{TV}(W^{\alpha}(t),N)+d_{TV}\left(\dfrac{Y^{\alpha}_{t}}{\sqrt{\alpha}},W^{\alpha}(t)\right)\leq C\left(\lambda(t,2(\kappa+\gamma))\right)^{-1/2}\alpha^{-1/2},

where CC is a constant depending on {x0,y0,κ,γ,K,L,T,σ}\{x_{0},y_{0},\kappa,\gamma,K,L,T,\sigma\}. This completes the proof. ∎

Acknowledgment

Research of MHD was supported by EPSRC Grants EP/W008041/1 and EP/V038516/1.

References

  • [BGM10] Bolley, F., Guillin, A., and Malrieu, F. Trend to equilibrium and particle approximation for a weakly selfconsistent vlasov-fokker-planck equation. ESAIM: M2AN, 44(5):867–884, 2010.
  • [Bre09] Norbert Breimhorst. Smoluchowski-Kramers Approximation for Stochastic Differential Equations with non-Lipschitzian coefficients. PhD thesis, 2009.
  • [CF06] S. Cerrai and M. Freidlin. On the Smoluchowski-Kramers approximation for a system with an infinite number of degrees of freedom. Probab. Theory Related Fields, 135(3):363–394, 2006.
  • [CT22] Y.-P. Choi and O. Tse. Quantified overdamped limit for kinetic vlasov–fokker–planck equations with singular interaction forces. Journal of Differential Equations, 330:150–207, 2022.
  • [DLP+18] M H Duong, A Lamacz, M A Peletier, A Schlichting, and U Sharma. Quantification of coarse-graining error in langevin and overdamped langevin dynamics. Nonlinearity, 31(10):4517–4566, aug 2018.
  • [DLPS17] M. H. Duong, A. Lamacz, M. A. Peletier, and U. Sharma. Variational approach to coarse-graining of generalized gradient flows. Calculus of Variations and Partial Differential Equations, 56(4):100, Jun 2017.
  • [Duo15] Manh Hong Duong. Long time behaviour and particle approximation of a generalised vlasov dynamic. Nonlinear Analysis: Theory, Methods & Applications, 127:1–16, 2015.
  • [Fre04] M. Freidlin. Some remarks on the Smoluchowski-Kramers approximation. Journal of Statistical Physics, 117(3-4):617–634, 2004.
  • [HVW12] S. Hottovy, G. Volpe, and J. Wehr. Noise-induced drift in stochastic differential equations with arbitrary friction and diffusion in the Smoluchowski-Kramers limit. J. Stat. Phys., 146(4):762–773, 2012.
  • [JW17] P.-E. Jabin and Z. Wang. Mean Field Limit for Stochastic Particle Systems, pages 379–402. Springer International Publishing, Cham, 2017.
  • [Kac56] M. Kac. Foundations of kinetic theory. Proceedings of the Third Berkeley Symposium on Mathematical Statistics and Probability, 1954–1955, vol. III, University of California Press, Berkeley, Los Angeles, 1956.
  • [Kla07] B. Klartag. A central limit theorem for convex sets. Inventiones mathematicae, 168(1):91–131, Apr 2007.
  • [Kra40] H.A. Kramers. Brownian motion in a field of force and the diffusion model of chemical reactions. Physica, 7(4):284 – 304, 1940.
  • [KSSS91] I. Karatzas, I.K.S. Shreve, S. Shreve, and S.E. Shreve. Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics (113) (Book 113). Springer New York, 1991.
  • [McK67] H. P. McKean. Propagation of chaos for a class of non-linear parabolic equations, pages 41–57. Stochastic Differential Equations (Lecture Series in Differential Equations, Session 7, Catholic Univ. 1967.
  • [Nar91a] Kiyomasa Narita. Asymptotic behavior of velocity process in the smoluchowski–kramers approximation for stochastic differential equations. Advances in Applied Probability, 23(2):317–326, 1991.
  • [Nar91b] Kiyomasa Narita. The smoluchowski–kramers approximation for the stochastic liénard equation by mean-field. Advances in Applied Probability, 23(2):303–316, 1991.
  • [Nar94] K. Narita. Asymptotic behavior of fluctuation and deviation from limit system in the smoluchowski-kramers approximation for sde. Yokohama Mathematical Journal, 42:41–76, 1994.
  • [Nel67] Edward Nelson. Dynamical Theories of Brownian Motion, volume 17. Princeton University Press Princeton, 1967.
  • [NN20] V. T. Nguyen and T. D. Nguyen. A berry–esseen bound in the smoluchowski–kramers approximation. Journal of Statistical Physics, 179(4):871–884, May 2020.
  • [Nua06] D. Nualart. The Malliavin Calculus and Related Topics. Probability and Its Applications. Springer Berlin Heidelberg, 2006.
  • [Pav14] G.A. Pavliotis. Stochastic Processes and Applications: Diffusion Processes, the Fokker-Planck and Langevin Equations. Texts in Applied Mathematics. Springer New York, 2014.
  • [RF96] H. Risken and T. Frank. The Fokker-Planck Equation: Methods of Solution and Applications. Springer Series in Synergetics. Springer Berlin Heidelberg, 1996.
  • [SP12] R. L. Schilling and L. Partzsch. Brownian Motion: An Introduction to Stochastic Processes. Walter de Gruyter, 2012.
  • [Szn91] A.-S. Sznitman. Topics in propagation of chaos. In Paul-Louis Hennequin, editor, Ecole d’Eté de Probabilités de Saint-Flour XIX — 1989, volume 1464 of Lecture Notes in Mathematics, pages 165–251. Springer Berlin Heidelberg, 1991.
  • [Ta20] C. S. Ta. The rate of convergence for the smoluchowski-kramers approximation for stochastic differential equations with fbm. Journal of Statistical Physics, 181(5):1730–1745, Dec 2020.
  • [XY22] Longjie Xie and Li Yang. The smoluchowski–kramers limits of stochastic differential equations with irregular coefficients. Stochastic Processes and their Applications, 150:91–115, 2022.