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

Asymptotic behaviour for a time-inhomogeneous Kolmogorov type diffusion

Mihai Gradinaru and Emeline Luirard Univ Rennes, CNRS, IRMAR - UMR 6625, F-35000 Rennes, France
{Mihai.Gradinaru,Emeline.Luirard}@univ-rennes1.fr

Abstract:  We study a kinetic stochastic model with a non-linear time-inhomogeneous drag force and a Brownian-type random force. More precisely, the Kolmogorov type diffusion (V,X)(V,X) is considered: here XX is the position of the particle and VV is its velocity and is solution of a stochastic differential equation driven by a one-dimensional Brownian motion, with the drift of the form tβF(v)t^{-\beta}F(v). The function FF satisfies some homogeneity condition and β\beta is positive. The behaviour of the process (V,X)(V,X) in large time is proved by using stochastic analysis tools.
Keywords:  kinetic stochastic equation; time-inhomogeneous diffusions; explosion times; scaling transformations; asymptotic distributions; ergodicity.
MSC2010 Subject Classification: Primary 60J60; Secondary 60H10; 60J65; 60F17.

1 Introduction

In several domains as fluids dynamics, statistical mechanics, biology, a number of models are based on the Fokker-Planck and Langevin equations driven by Brownian motion or could be non-linear or driven by other random noises. For example, in [CCM10] the persistent turning walker model was introduced, inspired by the modelling of fish motion. An associated two-component Kolmogorov type diffusion solves a kinetic Fokker-Planck equation based on an Ornstein-Uhlenbeck Gaussian process and the authors studied the large time behaviour of this model by using appropriate tools from stochastic analysis. One of the natural questions is the behaviour in large time of the solution to the corresponding stochastic differential equation (SDE). Although the tools of partial differential equations allow us to ask of this kind of questions, since these models are probabilistic, tools based on stochastic processes could be more natural to use.

In the last decade the asymptotic study of solutions of non-linear Langevin’s type was the subject of an important number of papers, see [CNP19], [EG15], [FT21]. For instance, in [FT21] the following system is studied

Vt=v0+Btρ20tF(Vs)ds and Xt=x0+0tVsds.V_{t}=v_{0}+B_{t}-\frac{\rho}{2}\int_{0}^{t}F(V_{s})\mathop{}\!\mathrm{d}s\quad\mbox{ and }\quad X_{t}=x_{0}+\int_{0}^{t}V_{s}\mathop{}\!\mathrm{d}s.

In other words one considers a particle moving such that its velocity is a diffusion with an invariant measure behaving like (1+|v|2)ρ/2(1+|v|^{2})^{-\rho/2}, as |v|+|v|\to+\infty. The authors prove that for large time, after a suitable rescaling, the position process behaves as a Brownian motion or other stable processes, following the values of ρ\rho. Results have been extended to additive functional of VV in [Bét21]. It should be noticed that these cited papers use the standard tools associated with time-homogeneous equations: invariant measure, scale function and speed measure. Several of these tools will not be available when the drag force is depending explicitly on time. In [GO13], a non-linear SDE driven by a Brownian motion but having time-inhomogeneous drift coefficient was studied and its large time behaviour was described. Moreover, sharp rates of convergence are proved for the 1-dimensional marginal of the solution. In the present paper, we consider the velocity process as satisfying the same kind of SDE.

Let us describe our framework: consider a one-dimensional time-inhomogeneous stochastic kinetic model driven by a Brownian motion. We denote by (Xt)t0(X_{t})_{t\geq 0} the one-dimensional process describing the position of a particle at time tt having the velocity VtV_{t}. The velocity process (Vt)t0(V_{t})_{t\geq 0} is supposed to follow a Brownian dynamic in a potential U(t,v)U(t,v), varying in time:

dVt=dBt12vU(t,Vt)dt and Xt=X0+0tVsds.\mathop{}\!\mathrm{d}V_{t}=\mathop{}\!\mathrm{d}B_{t}-\dfrac{1}{2}\partial_{v}U(t,V_{t})\mathop{}\!\mathrm{d}t\quad\mbox{ and }\quad X_{t}=X_{0}+\int_{0}^{t}V_{s}\mathop{}\!\mathrm{d}s.

This system can be viewed as a perturbation of the classical two-component Kolmogorov diffusion

dVt=dBt and Xt=X0+0tVsds.\mathop{}\!\mathrm{d}V_{t}=\mathop{}\!\mathrm{d}B_{t}\quad\mbox{ and }\quad X_{t}=X_{0}+\int_{0}^{t}V_{s}\mathop{}\!\mathrm{d}s.

In the present paper the potential is supposed to grow slowly to infinity, and it will be supposed to be of the form tβ0vF(u)dut^{-\beta}\int_{0}^{v}F(u)\mathop{}\!\mathrm{d}u, with β>0\beta>0 and FF satisfying some homogeneity condition. It describes a one dimensional particle evolving in a force field FtβFt^{-\beta} and undergoing many small random shocks. A natural question is to understand the behaviour of the process (V,X)(V,X) in large time. More precisely we look for the limit in distribution of v(ε)(Vt/ε,εXt/ε)tv(\varepsilon)(V_{t/\varepsilon},\varepsilon X_{t/\varepsilon})_{t}, as ε0\varepsilon\to 0, where v(ε)v(\varepsilon) is some rate of convergence. Our results are proved on the product of path spaces and consequently contain those of [GO13].
If F=0F=0, it is not difficult to see that the rescaled position process (ε1/2Vt/ε,ε3/2Xt/ε)t(\varepsilon^{\nicefrac{{1}}{{2}}}V_{t/\varepsilon},\varepsilon^{\nicefrac{{3}}{{2}}}X_{t/\varepsilon})_{t} converges in distribution towards the Kolmogorov diffusion (Bt,0tBsds)t(B_{t},\int_{0}^{t}B_{s}\mathop{}\!\mathrm{d}s)_{t}. We prove that this kinetic behaviour still holds for sufficiently "small at infinity" potential. The strategy to tackle this problem is based on estimates of moments of the velocity process. The main result can then be extended for the case when the potential is equally weighted in some sense as the random noise. The potential either offsets the random noise (critical regime) or swings with it (sub-critical regime).
As suggested at the beginning of the introduction, other random noises can be considered. In [GL21], the case of a Lévy random noise is analysed. The case of a stochastic system in a harmonic potential is the purpose of a future work (see [Lui22]).


The organisation of our paper is as follows: in the next section we introduce notations, and we state our main results. Results about existence and non-explosion of solutions are stated in Section 3. Estimates of the moments of the velocity process are given in Section 4 while the proofs of our main results are presented in Section 5.

2 Notations and main results

Let (Bt)t0(B_{t})_{t\geq 0} be a standard Brownian motion, β\beta a real number and FF a continuous function which is supposed to satisfy either

for some γ,v,λ>0,F(λv)=λγF(v),\mbox{for some }\gamma\in\mathbb{R},\ \forall v\in\mathbb{R},\ \lambda>0,\ F(\lambda v)=\lambda^{\gamma}F(v), (H1γH_{1}^{\gamma})

or

|F|G where G is a positive function satisfying (H1γ).\left\lvert F\right\rvert\leq G\mbox{ where }G\mbox{ is a positive function satisfying \eqref{hyp1_levy}.} (H2γH_{2}^{\gamma})

Each assumption implies that there exist a positive constant KK such that, for all vv\in\mathbb{R}, |F(v)|K|v|γ\left\lvert F(v)\right\rvert\leq K\left\lvert v\right\rvert^{\gamma}. Obviously (H2γH_{2}^{\gamma}) is a generalisation of (H1γH_{1}^{\gamma}). In the following, sgn\operatorname{sgn} is the sign function with convention sgn(0)=0\operatorname{sgn}(0)=0. As an example of function satisfying (H1γH_{1}^{\gamma}) one can keep in mind F:vsgn(v)|v|γF:v\mapsto\operatorname{sgn}(v)\left\lvert v\right\rvert^{\gamma} (see also [GO13]), and as an example of function satisfying (H2γH_{2}^{\gamma}) (with γ=0\gamma=0) F:vv/(1+v2)F:v\mapsto\nicefrac{{v}}{{(1+v^{2})}} (see also [FT21]).

Remark 2.1.

If a function π\pi satisfies (H1γH_{1}^{\gamma}), then for all xx\in\mathbb{R}, π(x)=π(sgn(x))|x|γ\pi(x)=\pi(\operatorname{sgn}(x))\left\lvert x\right\rvert^{\gamma}.

We consider the following one-dimensional stochastic kinetic model, for tt0>0t\geq t_{0}>0,

dVt=dBttβF(Vt)dt,Vt0=v0>0, and dXt=Vtdt,Xt0=x0.\mathop{}\!\mathrm{d}V_{t}=\mathop{}\!\mathrm{d}B_{t}-t^{-\beta}F(V_{t})\mathop{}\!\mathrm{d}t,\ V_{t_{0}}=v_{0}>0,\quad\mbox{ and }\quad\mathop{}\!\mathrm{d}X_{t}=V_{t}\mathop{}\!\mathrm{d}t,\ X_{t_{0}}=x_{0}\in\mathbb{R}. (SKE)

Most of the convergences take place in the space of continuous functions 𝒞((0,+),){\mathcal{C}}((0,+\infty),\mathbb{R}) endowed by the uniform topology

du:f,g𝒞((0,+),)n=1+12nmin(1,sup[1n,n]|fg|).\displaystyle\mathop{}\!\mathrm{d}_{u}:f,g\in\mathcal{C}((0,+\infty),\mathbb{R})\mapsto\sum_{n=1}^{+\infty}\dfrac{1}{2^{n}}\min\Big{(}1,\sup_{[\frac{1}{n},n]}\left\lvert f-g\right\rvert\Big{)}.

For a family ((Zt(ε))t>0)ε>0((Z_{t}^{(\varepsilon)})_{t>0})_{\varepsilon>0} of continuous processes, we write

(Zt(ε))t>0(Zt)t>0,(Z_{t}^{(\varepsilon)})_{t>0}\quad\Longrightarrow\quad(Z_{t})_{t>0},

if (Zt(ε))t>0(Z_{t}^{(\varepsilon)})_{t>0} converges in distribution to (Zt)t>0(Z_{t})_{t>0} in 𝒞((0,+),){\mathcal{C}}((0,+\infty),\mathbb{R}), as ε0\varepsilon\to 0.
We write

(Zt(ε))t>0f.d.d.(Zt)t>0,(Z_{t}^{(\varepsilon)})_{t>0}\quad\stackrel{{\scriptstyle f.d.d.}}{{\Longrightarrow}}\quad(Z_{t})_{t>0},

if for all finite subsets S(0,+)S\subset(0,+\infty), the vector (Zt(ε))tS(Z_{t}^{(\varepsilon)})_{t\in S} converges in distribution to (Zt)tS(Z_{t})_{t\in S} in S\mathbb{R}^{S}, as ε0\varepsilon\to 0.
Let us state our main results. Set q:=βγ+1q:=\dfrac{\beta}{\gamma+1}.

Theorem 2.2.

Consider γ0\gamma\geq 0, and q>12q>\frac{1}{2}. Assume that either (H1γH_{1}^{\gamma}) or (H2γH_{2}^{\gamma}) is satisfied. Let (Vt,Xt)tt0(V_{t},X_{t})_{t\geq t_{0}} be the solution to (SKE) and (t)t0(\mathcal{B}_{t})_{t\geq 0} be a standard Brownian motion. Furthermore, if γ1\gamma\geq 1, we suppose that for all vv\in\mathbb{R}, vF(v)0vF(v)\geq 0.
Then, as ε0\varepsilon\to 0,

(εVt/ε,ε3/2Xt/ε)tεt0(t,0tsds)t0.\left(\sqrt{\varepsilon}V_{t/\varepsilon},\varepsilon^{\nicefrac{{3}}{{2}}}X_{t/\varepsilon}\right)_{t\geq\varepsilon t_{0}}\stackrel{{\scriptstyle\text{}}}{{\Longrightarrow}}\left(\mathcal{B}_{t},\int_{0}^{t}\mathcal{B}_{s}\mathop{}\!\mathrm{d}s\right)_{t\geq 0}.
Theorem 2.3.

Consider γ0\gamma\geq 0 and q=12q=\frac{1}{2}. Assume that (H1γH_{1}^{\gamma}) is satisfied. Let (Vt,Xt)tt0(V_{t},X_{t})_{t\geq t_{0}} be the solution to (SKE). If γ1\gamma\geq 1, we suppose furthermore that for all vv\in\mathbb{R}, vF(v)0vF(v)\geq 0.
Call H~\widetilde{H} the eternal ergodic process, solution to the homogeneous SDE

dHs=dWsHs2dsF(Hs)ds,\mathop{}\!\mathrm{d}H_{s}=\mathop{}\!\mathrm{d}W_{s}-\dfrac{H_{s}}{2}\mathop{}\!\mathrm{d}s-F\big{(}H_{s}\big{)}\mathop{}\!\mathrm{d}s,

such that the law of HH_{-\infty} is the invariant measure, where (Wt)t0(W_{t})_{t\geq 0} is again a standard Brownian motion. Setting ΛF,t1,,td\Lambda_{F,t_{1},\cdots,t_{d}} for the f.d.d. of H~\widetilde{H}, we call (𝒱t)t0(\mathcal{V}_{t})_{t\geq 0} the process whose finite dimensional distribution (f.d.d.) are TΛF,log(t1),,log(td)T*\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})}, the pushforward measure of ΛF,log(t1),,log(td)\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})} by the linear map T(u1,,ud):=(t1u1,,tdud)T(u_{1},\cdots,u_{d}):=(\sqrt{t_{1}}u_{1},\cdots,\sqrt{t_{d}}u_{d}), that is (𝒱t)t0=(tH~log(t))t0(\mathcal{V}_{t})_{t\geq 0}=(\sqrt{t}\widetilde{H}_{\log(t)})_{t\geq 0}.
Then, as ε0\varepsilon\to 0,

(εVt/ε,ε3/2Xt/ε)tεt0(𝒱t,0t𝒱sds)t0.\left(\sqrt{\varepsilon}V_{t/\varepsilon},\varepsilon^{\nicefrac{{3}}{{2}}}X_{t/\varepsilon}\right)_{t\geq\varepsilon t_{0}}\stackrel{{\scriptstyle\text{}}}{{\Longrightarrow}}\left(\mathcal{V}_{t},\int_{0}^{t}\mathcal{V}_{s}\mathop{}\!\mathrm{d}s\right)_{t\geq 0}.
Remark 2.4.

The one-dimensional distribution of (𝒱t)t0(\mathcal{V}_{t})_{t\geq 0} has already been explicitly computed (see Theorem 4.1 in [GO13]).

Theorem 2.5.

Consider γ1\gamma\geq 1 and q<12q<\frac{1}{2}. Assume that F:vρsgn(v)|v|γF:v\mapsto\rho\operatorname{sgn}(v)\left\lvert v\right\rvert^{\gamma} with ρ>0\rho>0. Let (Vt,Xt)tt0(V_{t},X_{t})_{t\geq t_{0}} be the solution to (SKE). Call H^\widehat{H} the ergodic process, solution to the homogeneous SDE

dHs=dWsF(Hs)ds,\mathop{}\!\mathrm{d}H_{s}=\mathop{}\!\mathrm{d}W_{s}-F\left(H_{s}\right)\mathop{}\!\mathrm{d}s,

where (Wt)t0(W_{t})_{t\geq 0} is a standard Brownian motion. Call ΠF\Pi_{F} its invariant measure. We call (𝒱t)t0(\mathscr{V}_{t})_{t\geq 0} the process whose f.d.d. are T(dΠF)T*\left(\otimes^{d}\Pi_{F}\right), the pushforward measure of dΠF\otimes^{d}\Pi_{F} by the linear map T(u1,,ud):=(t1qu1,,tdqud)T(u_{1},\cdots,u_{d}):=({t_{1}}^{q}u_{1},\cdots,{t_{d}}^{q}u_{d}).
Then, as ε0\varepsilon\to 0,

(εqVt/ε)tεt0f.d.d.(𝒱t)t0.\left(\varepsilon^{q}V_{t/\varepsilon}\right)_{t\geq\varepsilon t_{0}}\stackrel{{\scriptstyle\text{f.d.d.}}}{{\Longrightarrow}}\left(\mathscr{V}_{t}\right)_{t\geq 0}.

Moreover, in the linear case (i.e. γ=1\gamma=1) and if β>12\beta>-\frac{1}{2}, we define (𝒳t)t0(\mathscr{X}_{t})_{t\geq 0} the centered Gaussian process with covariance function K(s,t):=(ρ2(1+2β))1(st)1+2βK(s,t):=(\rho^{2}(1+2\beta))^{-1}(s\wedge t)^{1+2\beta}.
Then, as ε0\varepsilon\to 0,

(εβ+12Xt/ε)tεt0f.d.d.(𝒳t)t0.\left(\varepsilon^{\beta+\frac{1}{2}}X_{t/\varepsilon}\right)_{t\geq\varepsilon t_{0}}\stackrel{{\scriptstyle\text{f.d.d.}}}{{\Longrightarrow}}\left(\mathscr{X}_{t}\right)_{t\geq 0}. (1)
Remark 2.6.

If β=0\beta=0, one can prove using the martingale method, that (εXt/ε)t0(\sqrt{\varepsilon}X_{t/\varepsilon})_{t\geq 0} converges towards a Brownian motion. Assume, by way of contradiction, that the process (εqVt/ε)tεt0(\varepsilon^{q}V_{t/\varepsilon})_{t\geq\varepsilon t_{0}} would converge (i.e. were tight), then by the continuous mapping theorem, the process (εXt/ε)t0(\varepsilon X_{t/\varepsilon})_{t\geq 0} should converge. This is a contradiction with (1). Here is why we deal only with finite-dimensional convergence for the velocity process.

3 Changed-of-time processes

In the following, we suppose that γ>1\gamma>-1 and set Ω=𝒞¯([t0,+))\Omega=\overline{\mathcal{C}}([t_{0},+\infty)) the set of continuous functions, that equal ++\infty after their (possibly infinite) explosion time. Following the idea used in [GO13], we first perform a change of time in (SKE) in order to produce at least one time-homogeneous coefficient in the transformed equation. For every 𝒞2{\mathcal{C}}^{2}-diffeomorphism φ:[0,t1)[t0,+)\varphi:[0,t_{1})\to[t_{0},+\infty), let introduce the scaling transformation Φφ\Phi_{\varphi} defined, for ωΩ\omega\in\Omega, by

Φφ(ω)(s):=ω(φ(s))φ(s), with s[0,t1).\Phi_{\varphi}(\omega)(s):=\dfrac{\omega(\varphi(s))}{\sqrt{\varphi^{\prime}(s)}}\text{, with }s\in[0,t_{1}).

The result containing the change of time transformation can be found in [GO13], Proposition 2.1, p. 187.
Let VV be solution to the equation (SKE). Thanks to Lévy’s characterization theorem of the Brownian motion, (Wt)t0:=(\medint0tdBφ(s)φ(s))t0\left(W_{t}\right)_{t\geq 0}:=\left(\medint\int_{0}^{t}\dfrac{\mathop{}\!\mathrm{d}B_{\varphi(s)}}{\sqrt{\varphi^{\prime}(s)}}\right)_{t\geq 0} is a standard Brownian motion. Then, by a change of variable t=φ(s)t=\varphi(s), one gets

Vφ(t)Vφ(0)=0tφ(s)dWs0tF(Vφ(s))φ(s)βφ(s)ds.V_{\varphi(t)}-V_{\varphi(0)}=\int_{0}^{t}\sqrt{\varphi^{\prime}(s)}\mathop{}\!\mathrm{d}W_{s}-\int_{0}^{t}\dfrac{F(V_{\varphi(s)})}{\varphi(s)^{\beta}}\varphi^{\prime}(s)\mathop{}\!\mathrm{d}s.

The integration by parts formula yields

d(Vφ(s)φ(s))=dWsφ(s)φ(s)βF(Vφ(s))dsφ′′(s)2φ(s)Vφ(s)φ(s)ds.\mathop{}\!\mathrm{d}\left(\dfrac{V_{\varphi(s)}}{\sqrt{\varphi^{\prime}(s)}}\right)=\mathop{}\!\mathrm{d}W_{s}-\dfrac{\sqrt{\varphi^{\prime}(s)}}{\varphi(s)^{\beta}}F(V_{\varphi(s)})\mathop{}\!\mathrm{d}s-\dfrac{\varphi^{\prime\prime}(s)}{2\varphi^{\prime}(s)}\dfrac{V_{\varphi(s)}}{\sqrt{\varphi^{\prime}(s)}}\mathop{}\!\mathrm{d}s.

As a consequence, we can state the following result in our context.

Proposition 3.1.

If VV is a solution to the equation (SKE), then V(φ):=Φφ(V)V^{(\varphi)}:=\Phi_{\varphi}(V) is a solution to

dVs(φ)=dWsφ(s)φ(s)βF(φ(s)Vs(φ))dsφ′′(s)φ(s)Vs(φ)2ds,V0(φ)=Vφ(0)φ(0),\mathop{}\!\mathrm{d}V^{(\varphi)}_{s}=\mathop{}\!\mathrm{d}W_{s}-\dfrac{\sqrt{\varphi^{\prime}(s)}}{\varphi(s)^{\beta}}F(\sqrt{\varphi^{\prime}(s)}V_{s}^{(\varphi)})\mathop{}\!\mathrm{d}s-\dfrac{\varphi^{\prime\prime}(s)}{\varphi^{\prime}(s)}\dfrac{V_{s}^{(\varphi)}}{2}\mathop{}\!\mathrm{d}s,\ V_{0}^{(\varphi)}=\dfrac{V_{\varphi(0)}}{\sqrt{\varphi^{\prime}(0)}}, (2)

where Wt:=\medint0tdBφ(s)φ(s)W_{t}:=\medint\int_{0}^{t}\dfrac{\mathop{}\!\mathrm{d}B_{\varphi(s)}}{\sqrt{\varphi^{\prime}(s)}}.

If V(φ)V^{(\varphi)} is a solution to (2), then Φφ1(V(φ))\Phi_{\varphi}^{-1}(V^{(\varphi)}) is a solution to the equation (SKE), where BtBt0:=\medintt0t(φφ1)(s)dWφ1(s)B_{t}-B_{t_{0}}:=\medint\int_{t_{0}}^{t}\sqrt{(\varphi^{\prime}\circ\varphi^{-1})(s)}\mathop{}\!\mathrm{d}W_{\varphi^{-1}(s)}.

Furthermore, uniqueness in law, pathwise uniqueness or strong existence hold for the equation (SKE) if and only if they hold for the equation (2).

In the following, we will use two particular changes of time, depending on which term of (2) should become time-homogeneous.

  • The exponential change of time: denoting φe:tt0et\varphi_{e}:t\mapsto t_{0}e^{t}, the exponential scaling transformation is defined by Φe(ω):s+ωt0est0es/2\Phi_{e}(\omega):s\in\mathbb{R}^{+}\mapsto\dfrac{\omega_{t_{0}e^{s}}}{\sqrt{t_{0}}e^{\nicefrac{{s}}{{2}}}}, for ωΩ\omega\in\Omega. Thanks to 3.1, the process V(e):=Φe(V)V^{(e)}:=\Phi_{e}(V) satisfies the equation

    dVs(e)=dWsVs(e)2dst01/2βe(1/2β)sF(t0es/2Vs(e))ds,\mathop{}\!\mathrm{d}V_{s}^{(e)}=\mathop{}\!\mathrm{d}W_{s}-\dfrac{V_{s}^{(e)}}{2}\mathop{}\!\mathrm{d}s-t_{0}^{\nicefrac{{1}}{{2}}-\beta}e^{(\nicefrac{{1}}{{2}}-\beta)s}F\big{(}\sqrt{t_{0}}e^{\nicefrac{{s}}{{2}}}V_{s}^{(e)}\big{)}\mathop{}\!\mathrm{d}s,

    where (Wt)t0(W_{t})_{t\geq 0} is a standard Brownian motion.

  • The power change of time: for q=βγ+112q=\frac{\beta}{\gamma+1}\neq\frac{1}{2}, consider φq𝒞2([0,t1))\varphi_{q}\in\mathcal{C}^{2}([0,t_{1})) the solution to the Cauchy problem

    φq=φq2q,φq(0)=t0.\varphi_{q}^{\prime}=\varphi_{q}^{2q},\ \varphi_{q}(0)=t_{0}.

    Clearly, φq(t)=(t012q+(12q)t)1/(12q)\varphi_{q}(t)=\big{(}t_{0}^{1-2q}+(1-2q)t\big{)}^{\nicefrac{{1}}{{(1-2q)}}}, when 2q12q\neq 1, and φq=φe\varphi_{q}=\varphi_{e}, when 2q=12q=1.

    The time t1t_{1} satisfies t1=+t_{1}=+\infty, when 2q12q\leq 1, and t1=t012q(2q1)1t_{1}=t_{0}^{1-2q}(2q-1)^{-1}, when 2q>12q>1. The power scaling transformation is defined by Φq(ω):s+ω(φq(s))φq(s)q\Phi_{q}(\omega):s\in\mathbb{R}^{+}\mapsto\dfrac{\omega(\varphi_{q}(s))}{\varphi_{q}(s)^{q}}. The process V(q):=V(φq)V^{(q)}:=V^{(\varphi_{q})} satisfies the equation

    dVs(q)=dWsφqγq(s)F(φq(s)Vs(q))dsqφq2q1(s)Vs(q)ds,\mathop{}\!\mathrm{d}V_{s}^{(q)}=\mathop{}\!\mathrm{d}W_{s}-\varphi_{q}^{-\gamma q}(s)F\Big{(}\sqrt{\varphi^{\prime}_{q}(s)}V_{s}^{(q)}\Big{)}\mathop{}\!\mathrm{d}s-q\varphi_{q}^{2q-1}(s)V_{s}^{(q)}\mathop{}\!\mathrm{d}s, (3)

    where (Wt)t0(W_{t})_{t\geq 0} is a standard Brownian motion.

Adapting the proof of Propositions 3.2, 3.6 and 3.7 p. 188, in [GO13], one can prove the following proposition.

Proposition 3.2.

For γ0\gamma\geq 0, there exists a pathwise unique strong solution to (SKE), defined up to the explosion time τ\tau_{\infty} of VV.

  • When γ1\gamma\leq 1 or for all vv\in\mathbb{R}, vF(v)0vF(v)\geq 0, then τ\tau_{\infty} is a.s. infinite.

  • When 2q>12q>1, then (τ=+)>0\mathbb{P}(\tau_{\infty}=+\infty)>0.

  • Under (H1γH_{1}^{\gamma}), if γ>1\gamma>1 and (F(1),F(1))((0,+))×[0,+))(×(,0))(F(-1),F(1))\in((0,+\infty))\times[0,+\infty))\cup(\mathbb{R}\times(-\infty,0)), then (τ=+)<1\mathbb{P}(\tau_{\infty}=+\infty)<1.

Remark 3.3.

Assume that (H1γH_{1}^{\gamma}) is satisfied. In the linear case (γ=1\gamma=1), the drift and the diffusion terms are Lipschitz and satisfy locally linear growth condition. The existence and non-explosion of VV follow from Theorem 2.9, p. 289, in [KS98].

For more details, we refer to [Lui22].

4 Moment estimates of the velocity process

In this section, we give estimates for the moment of the velocity process. It will be useful to control some stochastic terms appearing later.

Proposition 4.1.

Assume that γ0\gamma\geq 0 and β\beta\in\mathbb{R}. The inequality

tt0,𝔼[|Vt|κ]Cγ,κ,β,t0tκ2\forall t\geq t_{0},\ \mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq C_{\gamma,\kappa,\beta,t_{0}}t^{\frac{\kappa}{2}}

holds for

  • κ[0,1]\kappa\in[0,1], when γ<1\gamma<1 and βγ+12\beta\geq\frac{\gamma+1}{2},

  • κ0\kappa\geq 0, when for all vv\in\mathbb{R}, vF(v)0vF(v)\geq 0.

If κ[0,1]\kappa\in[0,1], γ<1\gamma<1 and β<γ+12\beta<\frac{\gamma+1}{2}, then

tt0,𝔼[|Vt|κ]Cγ,κ,β,t0tκ1β1γ.\forall t\geq t_{0},\ \mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq C_{\gamma,\kappa,\beta,t_{0}}t^{\kappa\frac{1-\beta}{1-\gamma}}.
Remark 4.2.

When 1<γ<0-1<\gamma<0, it can be proved that for all tt0t\geq t_{0}, 𝔼[|Vt|]Cγ,β,t0t\mathbb{E}\left[\left\lvert V_{t}\right\rvert\right]\leq C_{\gamma,\beta,t_{0}}\sqrt{t}, without hypothesis of the positivity of the function vvF(v)v\mapsto vF(v).

Proof.
  1. Step 1.

    Assume that γ1\gamma\geq 1 and that for all vv\in\mathbb{R}, vF(v)0vF(v)\geq 0.
    Define, for all n0n\geq 0, the stopping times Tn:=inf{tt0,|Vt|n}T_{n}:=\inf\{t\geq t_{0},\ \left\lvert V_{t}\right\rvert\geq n\}. By Itô’s formula, for all tt0t\geq t_{0}, we have

    VtTn2\displaystyle V_{t\wedge T_{n}}^{2} =v02+t0tTn2VsdBst0tTn2sβVsF(Vs)ds+(tTnt0)\displaystyle=v_{0}^{2}+\int_{t_{0}}^{t\wedge T_{n}}2V_{s}\mathop{}\!\mathrm{d}B_{s}-\int_{t_{0}}^{t\wedge T_{n}}2s^{-\beta}V_{s}F(V_{s})\mathop{}\!\mathrm{d}s+({t\wedge T_{n}}-t_{0})
    =v02+t0t𝟙sTn2VsdBst0tTn2sβVsF(Vs)ds+(tTnt0)\displaystyle=v_{0}^{2}+\int_{t_{0}}^{t}\mathbb{1}_{s\leq T_{n}}2V_{s}\mathop{}\!\mathrm{d}B_{s}-\int_{t_{0}}^{t\wedge T_{n}}2s^{-\beta}V_{s}F(V_{s})\mathop{}\!\mathrm{d}s+(t\wedge T_{n}-t_{0})
    v02+t0t𝟙sTn2VsdBs+(tt0).\displaystyle\leq v_{0}^{2}+\int_{t_{0}}^{t}\mathbb{1}_{s\leq T_{n}}2V_{s}\mathop{}\!\mathrm{d}B_{s}+({t}-t_{0}).

    Since t0t4𝟙sTnVs2ds4n2(tt0)<+\int_{t_{0}}^{t}4\mathbb{1}_{s\leq T_{n}}V_{s}^{2}\mathop{}\!\mathrm{d}s\leq 4n^{2}(t-t_{0})<+\infty, taking expectation yields

    𝔼[VtTn2]v02+(tt0)Ct0t.\mathbb{E}\left[V_{t\wedge T_{n}}^{2}\right]\leq v_{0}^{2}+(t-t_{0})\leq C_{t_{0}}t.

    Set κ[0,2]\kappa\in[0,2], we obtain by Jensen’s inequality that

    𝔼[|Vt|κ]𝔼[|Vt|2]κ2(lim infn+𝔼[VtTn2])κ2Cκ,t0tκ2.\mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq\mathbb{E}\left[\left\lvert V_{t}\right\rvert^{2}\right]^{\frac{\kappa}{2}}\leq\left(\liminf_{n\to+\infty}\mathbb{E}\left[V_{t\wedge T_{n}}^{2}\right]\right)^{\frac{\kappa}{2}}\leq C_{\kappa,t_{0}}t^{\frac{\kappa}{2}}. (4)

    When κ>2\kappa>2, the function v|v|κv\mapsto\left\lvert v\right\rvert^{\kappa} is 𝒞2\mathcal{C}^{2}, so by Itô’s formula, we can write for all tt0t\geq t_{0},

    |VtTn|κ=|v0|κ+t0tTnκsgn(Vs)|Vs|κ1dBst0tTnκsβ|Vs|κ1sgn(Vs)F(Vs)ds+t0tTnκ(κ1)2|Vs|κ2ds.\left\lvert V_{t\wedge T_{n}}\right\rvert^{\kappa}=\left\lvert v_{0}\right\rvert^{\kappa}+\int_{t_{0}}^{t\wedge T_{n}}\kappa\operatorname{sgn}(V_{s})\left\lvert V_{s}\right\rvert^{\kappa-1}\mathop{}\!\mathrm{d}B_{s}-\int_{t_{0}}^{t\wedge T_{n}}\kappa s^{-\beta}\left\lvert V_{s}\right\rvert^{\kappa-1}\operatorname{sgn}(V_{s})F(V_{s})\mathop{}\!\mathrm{d}s\\ +\int_{t_{0}}^{t\wedge T_{n}}\dfrac{\kappa(\kappa-1)}{2}\left\lvert V_{s}\right\rvert^{\kappa-2}\mathop{}\!\mathrm{d}s.

    In addition, using the hypothesis on the sign of FF, we have

    |VtTn|κ|v0|κ+t0t𝟙sTnκsgn(Vs)|Vs|κ1dBs+t0tTnκ(κ1)2|Vs|κ2ds.\left\lvert V_{t\wedge T_{n}}\right\rvert^{\kappa}\leq\left\lvert v_{0}\right\rvert^{\kappa}+\int_{t_{0}}^{t}\mathbb{1}_{s\leq T_{n}}\kappa\operatorname{sgn}(V_{s})\left\lvert V_{s}\right\rvert^{\kappa-1}\mathop{}\!\mathrm{d}B_{s}+\int_{t_{0}}^{t\wedge T_{n}}\dfrac{\kappa(\kappa-1)}{2}\left\lvert V_{s}\right\rvert^{\kappa-2}\mathop{}\!\mathrm{d}s. (5)

    We observe that t0tκ2Vs2κ2𝟙sTndsκ2n2κ2(tt0)<+\int_{t_{0}}^{t}\kappa^{2}V_{s}^{2\kappa-2}\mathbb{1}_{s\leq T_{n}}\mathop{}\!\mathrm{d}s\leq\kappa^{2}n^{2\kappa-2}(t-t_{0})<+\infty. Taking expectation in (5), we obtain

    𝔼[|Vt|κ]lim infn+𝔼[|VtTn|κ]|v0|κ+t0tκ(κ1)2𝔼[|Vs|κ2]ds.\mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq\liminf_{n\to+\infty}\mathbb{E}\left[\left\lvert V_{t\wedge T_{n}}\right\rvert^{\kappa}\right]\leq\left\lvert v_{0}\right\rvert^{\kappa}+\int_{t_{0}}^{t}\dfrac{\kappa(\kappa-1)}{2}\mathbb{E}\left[\left\lvert V_{s}\right\rvert^{\kappa-2}\right]\mathop{}\!\mathrm{d}s.

    When 0κ220\leq\kappa-2\leq 2, we can upper bound 𝔼[|Vs|κ2]\mathbb{E}\left[\left\lvert V_{s}\right\rvert^{\kappa-2}\right] by injecting (4) and get

    𝔼[|Vt|κ]|v0|κ+t0tκ(κ1)2Cκ,t0sκ22dsCκ,t0sκ2.\mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq\left\lvert v_{0}\right\rvert^{\kappa}+\int_{t_{0}}^{t}\dfrac{\kappa(\kappa-1)}{2}C_{\kappa,t_{0}}s^{\frac{\kappa-2}{2}}\mathop{}\!\mathrm{d}s\leq C_{\kappa,t_{0}}s^{\frac{\kappa}{2}}.

    The same method is then applied inductively to prove the inequality for all κ>2\kappa>2.

  2. Step 2.

    Assume now that γ[0,1[\gamma\in[0,1[. Fix κ[0,1]\kappa\in[0,1]. Then Jensen’s inequality yields, for all tt0t\geq t_{0}, 𝔼[|Vt|κ]𝔼[|Vt|]κ\mathbb{E}\left[\left\lvert V_{t}\right\rvert^{\kappa}\right]\leq\mathbb{E}\left[\left\lvert V_{t}\right\rvert\right]^{\kappa}, hence it suffices to verify the inequality only for κ=1\kappa=1.
    Define, for all n0n\geq 0, the stopping times Tn:=inf{tt0,|Vt|n}T_{n}:=\inf\{t\geq t_{0},\ \left\lvert V_{t}\right\rvert\geq n\} and let us recall that under both hypotheses (H1γH_{1}^{\gamma}) or (H2γH_{2}^{\gamma}), there exists a positive constant KK, such that |F(v)|K|v|γ\left\lvert F\left(v\right)\right\rvert\leq K\left\lvert v\right\rvert^{\gamma}. We can write, for tt0t\geq t_{0} and n0n\geq 0,

    |VtTn|\displaystyle\left\lvert V_{t\wedge T_{n}}\right\rvert |v0Bt0|+|BtTn|+t0tTnsβ|F(VsTn)|ds\displaystyle\leq\left\lvert v_{0}-B_{t_{0}}\right\rvert+\left\lvert B_{t\wedge T_{n}}\right\rvert+\int_{t_{0}}^{t\wedge T_{n}}s^{-\beta}\left\lvert F(V_{s\wedge T_{n}})\right\rvert\mathop{}\!\mathrm{d}s
    |v0Bt0|+|BtTn|+t0tTnsβK|VsTn|γds.\displaystyle\leq\left\lvert v_{0}-B_{t_{0}}\right\rvert+\left\lvert B_{t\wedge T_{n}}\right\rvert+\int_{t_{0}}^{t\wedge T_{n}}s^{-\beta}K\left\lvert V_{s\wedge T_{n}}\right\rvert^{\gamma}\mathop{}\!\mathrm{d}s.

    By noting that γ[0,1[\gamma\in[0,1[ and (Bt2t)t0(B^{2}_{t}-t)_{t\geq 0} is a martingale, taking expectation we get

    𝔼[|VtTn|]\displaystyle\mathbb{E}\left[\left\lvert V_{t\wedge T_{n}}\right\rvert\right] 𝔼[|v0Bt0|]+𝔼[|BtTn|]+t0tsβK𝔼[|VsTn|γ]ds\displaystyle\leq\mathbb{E}\left[\left\lvert v_{0}-B_{t_{0}}\right\rvert\right]+\mathbb{E}\left[\left\lvert B_{t\wedge T_{n}}\right\rvert\right]+\int_{t_{0}}^{t}s^{-\beta}K\mathbb{E}\left[\left\lvert V_{s\wedge T_{n}}\right\rvert^{\gamma}\right]\mathop{}\!\mathrm{d}s
    𝔼[|v0Bt0|]+𝔼[BtTn2]+t0tsβK𝔼[|VsTn|]γds\displaystyle\leq\mathbb{E}\left[\left\lvert v_{0}-B_{t_{0}}\right\rvert\right]+\sqrt{\mathbb{E}\left[B_{t\wedge T_{n}}^{2}\right]}+\int_{t_{0}}^{t}s^{-\beta}K\mathbb{E}\left[\left\lvert V_{s\wedge T_{n}}\right\rvert\right]^{\gamma}\mathop{}\!\mathrm{d}s
    𝔼[|v0Bt0|]+𝔼[tTn]+t0tsβK𝔼[|VsTn|]γds\displaystyle\leq\mathbb{E}\left[\left\lvert v_{0}-B_{t_{0}}\right\rvert\right]+\sqrt{\mathbb{E}\left[{t\wedge T_{n}}\right]}+\int_{t_{0}}^{t}s^{-\beta}K\mathbb{E}\left[\left\lvert V_{s\wedge T_{n}}\right\rvert\right]^{\gamma}\mathop{}\!\mathrm{d}s
    Ct0t+t0tsβK𝔼[|VsTn|]γds.\displaystyle\leq C_{t_{0}}\sqrt{t}+\int_{t_{0}}^{t}s^{-\beta}K\mathbb{E}\left[\left\lvert V_{s\wedge T_{n}}\right\rvert\right]^{\gamma}\mathop{}\!\mathrm{d}s.

    The function gn:t𝔼[|VtTn|]g_{n}:t\mapsto\mathbb{E}\left[\left\lvert V_{t\wedge T_{n}}\right\rvert\right] is bounded by nn. Applying a Gronwall-type lemma, stated below (4.3) and Fatou’s lemma, for β1\beta\neq 1 and for all tt0t\geq t_{0}, we end up with

    𝔼[|Vt|]lim infn+𝔼[|VtTn|]\displaystyle\ \mathbb{E}\left[\left\lvert V_{t}\right\rvert\right]\leq\liminf_{n\to+\infty}\mathbb{E}\left[\left\lvert V_{t\wedge T_{n}}\right\rvert\right] Cγ[Ct0t+(1γ1βK(t1βt01β))11γ]\displaystyle\leq C_{\gamma}\left[C_{t_{0}}\sqrt{t}+\left(\dfrac{1-\gamma}{1-\beta}K(t^{1-\beta}-t_{0}^{1-\beta})\right)^{\frac{1}{1-\gamma}}\right]
    Cγ,β,t0{t if βγ+12,t1β1γelse.\displaystyle\leq C_{\gamma,\beta,t_{0}}\begin{cases}\sqrt{t}&\mbox{ if }\beta\geq\frac{\gamma+1}{2},\\ t^{\frac{1-\beta}{1-\gamma}}&\mbox{else.}\end{cases}

    The case β=1\beta=1 can be treated similarly.

Lemma 4.3 (Gronwall-type lemma).

Fix r[0,1)r\in[0,1) and t0t_{0}\in\mathbb{R}. Assume that gg is a non-negative real-valued function, bb is a positive function and aa is a differentiable real-valued function. Moreover, suppose that the function bgrbg^{r} is continuous. If

tt0,g(t)a(t)+t0tb(s)g(s)rds,\forall t\geq t_{0},\ g(t)\leq a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s, (6)

then,

tt0,g(t)211r[a(t)+((1r)t0tb(s)ds)11r].\forall t\geq t_{0},\ g(t)\leq 2^{\frac{1}{1-r}}\left[a(t)+\left((1-r)\int_{t_{0}}^{t}b(s)\mathop{}\!\mathrm{d}s\right)^{\frac{1}{1-r}}\right].
Proof.

For tt0t\geq t_{0}, since r0r\geq 0,

g(t)r(a(t)+t0tb(s)g(s)rds)r,g(t)^{r}\leq\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{r}\,,

then, multiplying by b(t)>0b(t)>0,

b(t)g(t)rb(t)(a(t)+t0tb(s)g(s)rds)r.b(t)g(t)^{r}\leq b(t)\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{r}\,.

Now, let us make appear the derivative of HH

a(t)+b(t)g(t)ra(t)+b(t)(a(t)+t0tb(s)g(s)rds)r,a^{\prime}(t)+b(t)g(t)^{r}\leq a^{\prime}(t)+b(t)\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{r},

that is

a(t)+b(t)g(t)r(a(t)+t0tb(s)g(s)rds)rb(t)+a(t)(a(t)+t0tb(s)g(s)rds)rb(t)+a(t)a(t)r.\dfrac{a^{\prime}(t)+b(t)g(t)^{r}}{\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{r}}\leq b(t)+\dfrac{a^{\prime}(t)}{\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{r}}\leq b(t)+\dfrac{a^{\prime}(t)}{a(t)^{r}}.

Integrating, since r1r\neq 1, we obtain

(1r)1[(a(t)+t0tb(s)g(s)rds)1ra(t0)1r](1r)1[a(t)1ra(t0)1r]+t0tb(s)ds(1-r)^{-1}\left[\left(a(t)+\int_{t_{0}}^{t}b(s)g(s)^{r}\mathop{}\!\mathrm{d}s\right)^{1-r}-a(t_{0})^{1-r}\right]\leq(1-r)^{-1}\left[a(t)^{1-r}-a(t_{0})^{1-r}\right]+\int_{t_{0}}^{t}b(s)\mathop{}\!\mathrm{d}s

or equivalently, setting HH for the right-hand side of (6) and using that r<1r<1, we get

H(t)1ra(t)1r+(1r)t0tb(s)ds.H(t)^{1-r}\leq a(t)^{1-r}+(1-r)\int_{t_{0}}^{t}b(s)\mathop{}\!\mathrm{d}s.

Since 11r>0\frac{1}{1-r}>0 and using (6)

g(t)(a(t)1r+(1r)t0tb(s)ds)11rCr[a(t)+((1r)t0tb(s)ds)11r].g(t)\leq\left(a(t)^{1-r}+(1-r)\int_{t_{0}}^{t}b(s)\mathop{}\!\mathrm{d}s\right)^{\frac{1}{1-r}}\leq C_{r}\left[a(t)+\left((1-r)\int_{t_{0}}^{t}b(s)\mathop{}\!\mathrm{d}s\right)^{\frac{1}{1-r}}\right].

This concludes the proof of the lemma. ∎

Remark 4.4.

Call H(t)H(t) the right-hand side of (6). If gg is not continuous, note that the function HH is continuous and satisfies (6) (since bb is positive and gHg\leq H). Therefore, one can apply the lemma to HH and then use the inequality gHg\leq H.

5 Proof of the asymptotic behaviour of the solution

This section is devoted to the proofs of our main results.

5.1 Asymptotic behaviour in the super-critical regime under both assumptions

In this section, we assume that γ0\gamma\geq 0 and q>12q>\frac{1}{2}.

Proof of 2.2.

We split the proof into three steps.

  1. Step 1.

    We note that it is enough to prove that the process

    (Vt(ε))t0:=(εVt/ε)t0(V_{t}^{(\varepsilon)})_{t\geq 0}:=(\sqrt{\varepsilon}V_{t/\varepsilon})_{t\geq 0}

    converges in distribution to a Brownian motion in the space of continuous functions 𝒞([0,+)){\mathcal{C}}([0,+\infty)) endowed by the uniform topology. In order to see V(ε)V^{(\varepsilon)} as a process of 𝒞([0,+))\mathcal{C}([0,+\infty)), let us state for all s[0,εt0]s\in[0,\varepsilon t_{0}], Vs(ε):=Vεt0(ε)=εv0V_{s}^{(\varepsilon)}:=V_{\varepsilon t_{0}}^{(\varepsilon)}=\sqrt{\varepsilon}v_{0}.
    For every ε(0,1]\varepsilon\in(0,1] and tεt0t\geq\varepsilon t_{0}, we can write

    ε3/2Xt/ε=ε3/2x0+εt0tVs(ε)ds.\varepsilon^{\nicefrac{{3}}{{2}}}X_{t/\varepsilon}=\varepsilon^{\nicefrac{{3}}{{2}}}x_{0}+\int_{\varepsilon t_{0}}^{t}V_{s}^{(\varepsilon)}\mathop{}\!\mathrm{d}s.

    Clearly, the theorem will be proved once we show that gε(V(ε)):=(V(ε),εt0Vs(ε)ds)g_{\varepsilon}(V^{(\varepsilon)}_{\bullet}):=(V^{(\varepsilon)}_{\bullet},\int_{\varepsilon t_{0}}^{\bullet}V_{s}^{(\varepsilon)}\mathop{}\!\mathrm{d}s) converges weakly in 𝒞([0,+)){\mathcal{C}}([0,+\infty)) endowed by the uniform topology. Here the mapping gε:v(vt,εt0tvsds)t0g_{\varepsilon}:v\mapsto\left(v_{t},\int_{\varepsilon t_{0}}^{t}v_{s}\mathop{}\!\mathrm{d}s\right)_{t\geq 0} is defined and valued on 𝒞((0,+)){\mathcal{C}}((0,+\infty)). This mapping is converging, as ε0\varepsilon\to 0, to the continuous mapping g:v(vt,0tvsds)t0g:v\mapsto\left(v_{t},\int_{0}^{t}v_{s}\mathop{}\!\mathrm{d}s\right)_{t\geq 0}.
    We have, for every ε(0,1]\varepsilon\in(0,1] and tεt0t\geq\varepsilon t_{0},

    Vt(ε)=εVt/ε=\displaystyle V_{t}^{(\varepsilon)}=\sqrt{\varepsilon}V_{t/\varepsilon}= ε(v0Bt0)+εBt/εεt0t/εF(Vs)sβds\displaystyle\sqrt{\varepsilon}(v_{0}-B_{t_{0}})+\sqrt{\varepsilon}B_{t/\varepsilon}-\sqrt{\varepsilon}\int_{t_{0}}^{t/\varepsilon}F(V_{s})s^{-\beta}\mathop{}\!\mathrm{d}s
    =\displaystyle= ε(v0Bt0)+Bt(ε)εβ1/2εt0tF(Vu/ε)uβdu.\displaystyle\sqrt{\varepsilon}(v_{0}-B_{t_{0}})+B_{t}^{(\varepsilon)}-\varepsilon^{\beta-1/2}\int_{\varepsilon t_{0}}^{t}F(V_{u/\varepsilon})u^{-\beta}\mathop{}\!\mathrm{d}u.

    By self-similarity, B(ε):=(εBt/ε)t0B^{(\varepsilon)}:=(\sqrt{\varepsilon}B_{t/\varepsilon})_{t\geq 0} has the same distribution as a standard Brownian motion.
    Assume that the convergence of the rescaled velocity process is proved in the strong way, that is

    Tt0,supεt0tT|Vt(ε)Bt(ε)|0, as ε0.\forall T\geq t_{0},\ \sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert V_{t}^{(\varepsilon)}-B_{t}^{(\varepsilon)}\right\rvert\stackrel{{\scriptstyle\mathbb{P}}}{{\longrightarrow}}0,\ \mbox{ as }\varepsilon\to 0. (7)

    Then it suffices to prove that gε(B(ε))g()g_{\varepsilon}(B^{(\varepsilon)})\stackrel{{\scriptstyle\text{}}}{{\Longrightarrow}}g(\mathcal{B}) and du(gε(V(ε)),gε(B(ε)))0\mathop{}\!\mathrm{d}_{u}\left(g_{\varepsilon}(V^{(\varepsilon)}),g_{\varepsilon}(B^{(\varepsilon)})\right)\stackrel{{\scriptstyle\mathbb{P}}}{{\longrightarrow}}0, as ε0\varepsilon\to 0 (see Theorem 3.1, p. 27, in [Bil99]).
    On the one hand, the process B(ε)B^{(\varepsilon)} being a Brownian motion and ||2\left\lvert\cdot\right\rvert_{\mathbb{R}^{2}} denoting a norm on 2\mathbb{R}^{2}, the first convergence follows from

    Tt0,supεt0tT|gε(t)g(t)|20, as ε0.\forall T\geq t_{0},\ \sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert g_{\varepsilon}(\mathcal{B}_{t})-g(\mathcal{B}_{t})\right\rvert_{\mathbb{R}^{2}}\stackrel{{\scriptstyle\mathbb{P}}}{{\longrightarrow}}0,\ \mbox{ as }\varepsilon\to 0. (8)

    Let us prove (8). For every εt0tT\varepsilon t_{0}\leq t\leq T, we get

    |gε(t)g(t)|2\displaystyle\left\lvert g_{\varepsilon}(\mathcal{B}_{t})-g(\mathcal{B}_{t})\right\rvert_{\mathbb{R}^{2}} =|0εt0sds|\displaystyle=\left\lvert\int_{0}^{\varepsilon t_{0}}\mathcal{B}_{s}\mathop{}\!\mathrm{d}s\right\rvert
    0εt0|s|ds.\displaystyle\leq\int_{0}^{\varepsilon t_{0}}\left\lvert\mathcal{B}_{s}\right\rvert\mathop{}\!\mathrm{d}s.

    Hence,

    𝔼[supεt0tT|gε(t)g(t)|2]0εt0𝔼|s|dsC0εt0sdsε00.\mathbb{E}\left[\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert g_{\varepsilon}(\mathcal{B}_{t})-g(\mathcal{B}_{t})\right\rvert_{\mathbb{R}^{2}}\right]\leq\int_{0}^{\varepsilon t_{0}}\mathbb{E}\left\lvert\mathcal{B}_{s}\right\rvert\mathop{}\!\mathrm{d}s\leq C\int_{0}^{\varepsilon t_{0}}\sqrt{s}\mathop{}\!\mathrm{d}s\underset{\varepsilon\to 0}{\longrightarrow}0.

    On the other hand, we prove that

    Tt0,supεt0tT|gε(Vt(ε))gε(Bt(ε))|20, as ε0.\forall T\geq t_{0},\ \sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert g_{\varepsilon}(V^{(\varepsilon)}_{t})-g_{\varepsilon}(B_{t}^{(\varepsilon)})\right\rvert_{\mathbb{R}^{2}}\stackrel{{\scriptstyle\mathbb{P}}}{{\longrightarrow}}0,\ \mbox{ as }\varepsilon\to 0. (9)

    For every εt0tT\varepsilon t_{0}\leq t\leq T, using (7)

    |gε(Vt(ε))gε(Bt(ε))|2\displaystyle\left\lvert g_{\varepsilon}(V^{(\varepsilon)}_{t})-g_{\varepsilon}(B_{t}^{(\varepsilon)})\right\rvert_{\mathbb{R}^{2}} =|Vt(ε)Bt(ε)|+|εt0tVs(ε)Bs(ε)ds|\displaystyle=\left\lvert V^{(\varepsilon)}_{t}-B_{t}^{(\varepsilon)}\right\rvert+\left\lvert\int_{\varepsilon t_{0}}^{t}V_{s}^{(\varepsilon)}-B_{s}^{(\varepsilon)}\mathop{}\!\mathrm{d}s\right\rvert
    (1+Tεt0)supεt0tT|Vt(ε)Bt(ε)|0.\displaystyle\leq(1+T-\varepsilon t_{0})\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert V^{(\varepsilon)}_{t}-B_{t}^{(\varepsilon)}\right\rvert\stackrel{{\scriptstyle\mathbb{P}}}{{\longrightarrow}}0.
  2. Step 2.

    Let us prove now (7). Recall that under both hypothesis (H1γH_{1}^{\gamma}) and (H2γH_{2}^{\gamma}), there exists a positive constant KK, such that (ε)γ|F(Vu(ε)ε)|K|Vu(ε)|γ(\sqrt{\varepsilon})^{\gamma}\left\lvert F\left(\dfrac{V_{u}^{(\varepsilon)}}{\sqrt{\varepsilon}}\right)\right\rvert\leq K\left\lvert V_{u}^{(\varepsilon)}\right\rvert^{\gamma}. Modifying the factor in front of the integral part, we get

    Vt(ε)=ε(v0Bt0)+εBt/εεβ(γ+1)/2εt0t(ε)γF(Vu(ε)ε)uβdu.V_{t}^{(\varepsilon)}=\sqrt{\varepsilon}(v_{0}-B_{t_{0}})+\sqrt{\varepsilon}B_{t/\varepsilon}-\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\int_{\varepsilon t_{0}}^{t}(\sqrt{\varepsilon})^{\gamma}F\left(\dfrac{V_{u}^{(\varepsilon)}}{\sqrt{\varepsilon}}\right)u^{-\beta}\mathop{}\!\mathrm{d}u.

    It follows that, for all t0Tt_{0}\leq T,

    supεt0tT|Vt(ε)Bt(ε)|\displaystyle\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert V_{t}^{(\varepsilon)}-B_{t}^{(\varepsilon)}\right\rvert\leq ε|v0Bt0|+εβ(γ+1)/2supεt0tT|εt0t(ε)γF(Vu(ε)ε)uβdu|\displaystyle\sqrt{\varepsilon}\left\lvert v_{0}-B_{t_{0}}\right\rvert+\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert\int_{\varepsilon t_{0}}^{t}(\sqrt{\varepsilon})^{\gamma}F\left(\dfrac{V_{u}^{(\varepsilon)}}{\sqrt{\varepsilon}}\right)u^{-\beta}\mathop{}\!\mathrm{d}u\right\rvert
    \displaystyle\leq ε|v0Bt0|+εβ(γ+1)/2εt0TK|Vu(ε)|γuβdu.\displaystyle\sqrt{\varepsilon}\left\lvert v_{0}-B_{t_{0}}\right\rvert+\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\int_{\varepsilon t_{0}}^{T}K\left\lvert V_{u}^{(\varepsilon)}\right\rvert^{\gamma}u^{-\beta}\mathop{}\!\mathrm{d}u.

    Taking the expectation and using moment estimates (4.1), we obtain, when βγ2+1\beta\neq\frac{\gamma}{2}+1 and since β>γ+12\beta>\frac{\gamma+1}{2},

    εβ(γ+1)/2𝔼[εt0TK|Vu(ε)|γuβdu]\displaystyle\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\mathbb{E}\left[\int_{\varepsilon t_{0}}^{T}K\left\lvert V_{u}^{(\varepsilon)}\right\rvert^{\gamma}u^{-\beta}\mathop{}\!\mathrm{d}u\right] =εβ(γ+1)/2εt0TK𝔼[|Vu(ε)|γ]uβdu\displaystyle=\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\int_{\varepsilon t_{0}}^{T}K\mathbb{E}\left[\left\lvert V_{u}^{(\varepsilon)}\right\rvert^{\gamma}\right]u^{-\beta}\mathop{}\!\mathrm{d}u
    εβ(γ+1)/2εt0TKCγ,β,t0uγ2βdu\displaystyle\leq\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}\int_{\varepsilon t_{0}}^{T}KC_{\gamma,\beta,t_{0}}u^{\frac{\gamma}{2}-\beta}\mathop{}\!\mathrm{d}u
    C(εβ(γ+1)/2Tγ2β+1t0γ2β+1ε)ε00.\displaystyle\leq C\left(\varepsilon^{\beta-\nicefrac{{(\gamma+1)}}{{2}}}T^{\frac{\gamma}{2}-\beta+1}-t_{0}^{\frac{\gamma}{2}-\beta+1}\sqrt{\varepsilon}\right)\underset{\varepsilon\to 0}{\longrightarrow}0.

    Hence, setting r=min(12,β(γ+1)/2)>0r=\min(\frac{1}{2},\beta-\nicefrac{{(\gamma+1)}}{{2}})>0

    𝔼[supεt0tT|Vt(ε)Bt(ε)|]=O(εr).\mathbb{E}\left[\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert V_{t}^{(\varepsilon)}-B_{t}^{(\varepsilon)}\right\rvert\right]=O(\varepsilon^{r}).

    The case β=γ2+1\beta=\frac{\gamma}{2}+1 can be treated similarly to get

    𝔼[supεt0tT|Vt(ε)Bt(ε)|]=O(εln(ε)).\mathbb{E}\left[\sup_{\varepsilon t_{0}\leq t\leq T}\left\lvert V_{t}^{(\varepsilon)}-B_{t}^{(\varepsilon)}\right\rvert\right]=O(\sqrt{\varepsilon}\ln(\varepsilon)).

    This concludes the proof.

Remark 5.1.

One can observe that the only moment in this proof, when we need the condition "γ<1\gamma<1 or for all v,vF(v)v\in\mathbb{R},\ vF(v)" is when we are proving the moment estimates.

5.2 Asymptotic behaviour in the critical regime under (H1γH_{1}^{\gamma})

Assume in this section that β=γ+12\beta=\frac{\gamma+1}{2} and (H1γH_{1}^{\gamma}) is satisfied.

Proof of 2.3.
  1. Step 1.

    As in the first step of the previous section, it suffices to prove the convergence of the rescaled velocity process (εVt/ε)t(\sqrt{\varepsilon}V_{t/\varepsilon})_{t}. Keeping same notations, we prove that gε(V(ε))g_{\varepsilon}(V^{(\varepsilon)}) converges in distribution in 𝒞([0,+))\mathcal{C}([0,+\infty)) to g(𝒱)g(\mathcal{V}). In order to see V(ε)V^{(\varepsilon)} as a process of 𝒞([0,+))\mathcal{C}([0,+\infty)), let us set for all s[0,εt0]s\in[0,\varepsilon t_{0}], Vs(ε):=Vεt0(ε)=εv0V_{s}^{(\varepsilon)}:=V_{\varepsilon t_{0}}^{(\varepsilon)}=\sqrt{\varepsilon}v_{0}. Call PεP_{\varepsilon}, PP the distribution of V(ε)V^{(\varepsilon)}, 𝒱\mathcal{V} respectively. Then, using Pormanteau theorem (see Theorem 2.1 p.16 in [Bil99]), it suffices to prove that for all function h:𝒞([0,+))×𝒞([0,+))h:{\mathcal{C}}([0,+\infty))\times{\mathcal{C}}([0,+\infty))\to\mathbb{R} bounded and uniformly continuous,

    𝒞([0,+))2h(gε(ω))dPε(dω)ε0𝒞([0,+))2h(g(ω))dP(dω).\int_{\mathcal{C}([0,+\infty))^{2}}h(g_{\varepsilon}(\omega))\mathop{}\!\mathrm{d}P_{\varepsilon}(\mathop{}\!\mathrm{d}\omega)\underset{\varepsilon\to 0}{\longrightarrow}\int_{\mathcal{C}([0,+\infty))^{2}}h(g(\omega))\mathop{}\!\mathrm{d}P(\mathop{}\!\mathrm{d}\omega).

    Take a bounded and uniformly continuous function hh. By assumption, one knows that PεPP_{\varepsilon}\Longrightarrow P, hence, by Problem 4.12 p. 64, in [KS98], it suffices to prove that the uniformly bounded sequence (hgε)(h\circ g_{\varepsilon}) of continuous functions on 𝒞([0,+))\mathcal{C}([0,+\infty)) converges uniformly on compact subsets of 𝒞([0,+))\mathcal{C}([0,+\infty)) to the continuous function hgh\circ g. Let KK be a compact set of 𝒞([0,+))\mathcal{C}([0,+\infty)). Then, for all ωK\omega\in K, max[0,εt0]|ω|\max_{[0,\varepsilon t_{0}]}\left\lvert\omega\right\rvert is uniformly bounded by a constant, called MM.
    Fix η>0\eta>0. By the uniform continuity of hh, there exists δ>0\delta>0 such that for all ωK\omega\in K,

    du(gε(ω),g(ω))δ|hgε(ω),hg(ω)|η.\mathop{}\!\mathrm{d}_{u}(g_{\varepsilon}(\omega),g(\omega))\leq\delta\Rightarrow\left\lvert h\circ g_{\varepsilon}(\omega),h\circ g(\omega)\right\rvert\leq\eta.

    However, there exists ε1>0\varepsilon_{1}>0 small enough, such that for all εε1\varepsilon\leq\varepsilon_{1}, for all ωK\omega\in K,

    du(gε(ω),g(ω))C|0εt0ω(s)ds|Cεt0Mδ.\mathop{}\!\mathrm{d}_{u}(g_{\varepsilon}(\omega),g(\omega))\leq C\left\lvert\int_{0}^{\varepsilon t_{0}}\omega(s)\mathop{}\!\mathrm{d}s\right\rvert\leq C\varepsilon t_{0}M\leq\delta.
  2. Step 2.

    We first prove the f.d.d. convergence. The exponential scaling process V(e)V^{(e)} satisfies the time-homogeneous equation

    dVs(e)=dWsVs(e)2dsF(Vs(e))ds,\mathop{}\!\mathrm{d}V_{s}^{(e)}=\mathop{}\!\mathrm{d}W_{s}-\dfrac{V_{s}^{(e)}}{2}\mathop{}\!\mathrm{d}s-F\big{(}V_{s}^{(e)}\big{)}\mathop{}\!\mathrm{d}s, (10)

    where (Wt)t0(W_{t})_{t\geq 0} is a standard Brownian motion.
    Using the bijection induced by the exponential change of time (3.1), we get

    (Vt0ett0et/2)t0=(Ht)t0,\left(\dfrac{V_{t_{0}e^{t}}}{\sqrt{t_{0}}e^{t/2}}\right)_{t\geq 0}=(H_{t})_{t\geq 0},

    as solutions of the same SDE, starting at the same point. This can also be written as

    (Vtt)tt0=(Hlog(t/t0))tt0.\left(\dfrac{V_{t}}{\sqrt{t}}\right)_{t\geq t_{0}}=(H_{\log(t/t_{0})})_{t\geq t_{0}}.

    So, we have, for all ε>0\varepsilon>0, and (t1,,td)[εt0,+)d(t_{1},\cdots,t_{d})\in[\varepsilon t_{0},+\infty)^{d},

    (Vε1t1ε1t1,,Vε1tdε1td)=(Hlog(t1)+log((εt0)1),,Hlog(td)+log((εt0)1)).\left(\dfrac{V_{\varepsilon^{-1}t_{1}}}{\sqrt{\varepsilon^{-1}t_{1}}},\cdots,\dfrac{V_{\varepsilon^{-1}t_{d}}}{\sqrt{\varepsilon^{-1}t_{d}}}\right)=\left(H_{\log(t_{1})+\log((\varepsilon t_{0})^{-1})},\cdots,H_{\log(t_{d})+\log((\varepsilon t_{0})^{-1})}\right). (11)

    As in [GO13], the scale function and the speed measure of HH are respectively

    𝚙(x):=0xexp(y22+2γ+1sgn(y)F(sgn(y))|y|γ+1)dy{\tt p}(x):=\int_{0}^{x}\exp\left(\dfrac{y^{2}}{2}+\dfrac{2}{\gamma+1}\operatorname{sgn}(y)F(\operatorname{sgn}(y))\left\lvert y\right\rvert^{\gamma+1}\right)\mathop{}\!\mathrm{d}y

    and

    νF(dx):=exp(x222γ+1sgn(x)F(sgn(x))|x|γ+1)dx.\nu_{F}(\mathop{}\!\mathrm{d}x):=\exp\left(-\dfrac{x^{2}}{2}-\dfrac{2}{\gamma+1}\operatorname{sgn}(x)F(\operatorname{sgn}(x))\left\lvert x\right\rvert^{\gamma+1}\right)\mathop{}\!\mathrm{d}x.

    By the ergodic theorem (Theorem 23.15 p. 465 in [Kal02]), HH is ΛF\Lambda_{F}-ergodic, where ΛF\Lambda_{F} is the probability measure associated to νF\nu_{F}. Call H~\widetilde{H} the solution of the time homogeneous equation (10) such that the initial condition H~\widetilde{H}_{-\infty} has the distribution ΛF\Lambda_{F}.
    For t1,,td[εt0,+)dt_{1},\cdots,t_{d}\in[\varepsilon t_{0},+\infty)^{d}, let ΛF,t1,,td:=(H~t1,,H~td)\Lambda_{F,t_{1},\cdots,t_{d}}:=\mathcal{L}(\widetilde{H}_{t_{1}},\cdots,\widetilde{H}_{t_{d}}) be the distribution of (H~t1,,H~td)(\widetilde{H}_{t_{1}},\cdots,\widetilde{H}_{t_{d}}). Then, for all s0s\geq 0, ΛF,t1,,td=ΛF,t1+s,,td+s\Lambda_{F,t_{1},\cdots,t_{d}}=\Lambda_{F,t_{1}+s,\cdots,t_{d}+s}. Indeed, thanks to the invariance property of ΛF\Lambda_{F}, (H~t)t(\widetilde{H}_{t})_{t\in\mathbb{R}} and (H~t+s)t(\widetilde{H}_{t+s})_{t\in\mathbb{R}} satisfy the same SDE, starting at the same distribution. As a consequence, for all ε>0\varepsilon>0,

    (H~log(t1)+log((εt0)1),,H~log(td)+log((εt0)1))=ΛF,log(t1),,log(td).\mathcal{L}\left(\widetilde{H}_{\log(t_{1})+\log((\varepsilon t_{0})^{-1})},\cdots,\widetilde{H}_{\log(t_{d})+\log((\varepsilon t_{0})^{-1})}\right)=\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})}. (12)

    Moreover, by exponential ergodicity, for every ψ:d\psi:\mathbb{R}^{d}\to\mathbb{R} continuous and bounded function, we can prove that

    |𝔼[ψ(Hlog(t1/(t0ε)),,Hlog(td/(t0ε)))]𝔼[ψ(H~log(t1/(t0ε)),,H~log(td/(t0ε)))]|ε00.\left\lvert\mathbb{E}\left[\psi\left(H_{\log(t_{1}/(t_{0}\varepsilon))},\cdots,H_{\log(t_{d}/(t_{0}\varepsilon))}\right)\right]-\mathbb{E}\left[\psi\left(\widetilde{H}_{\log(t_{1}/(t_{0}\varepsilon))},\cdots,\widetilde{H}_{\log(t_{d}/(t_{0}\varepsilon))}\right)\right]\right\rvert\underset{\varepsilon\to 0}{\longrightarrow}0. (13)

    We postpone the proof of this convergence in Step 3.
    To conclude this step, gather (11), (12) and (13) to get

    (Vε1t1ε1t1,,Vε1tdε1td)ε0ΛF,log(t1),,log(td).\left(\dfrac{V_{\varepsilon^{-1}t_{1}}}{\sqrt{\varepsilon^{-1}t_{1}}},\cdots,\dfrac{V_{\varepsilon^{-1}t_{d}}}{\sqrt{\varepsilon^{-1}t_{d}}}\right)\underset{\varepsilon\to 0}{\Longrightarrow}\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})}.

    This can be written as

    (εVt1/ε,,εVtd/ε)ε0TΛF,log(t1),,log(td),\left(\sqrt{\varepsilon}V_{t_{1}/\varepsilon},\cdots,\sqrt{\varepsilon}V_{t_{d}/\varepsilon}\right)\underset{\varepsilon\to 0}{\Longrightarrow}T*\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})},

    where TΛF,log(t1),,log(td)T*\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})} is the pushforward of the measure ΛF,log(t1),,log(td)\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})} by the linear map T(u1,,ud):=(t1u1,,tdud)T(u_{1},\cdots,u_{d}):=(\sqrt{t_{1}}u_{1},\cdots,\sqrt{t_{d}}u_{d}).

  3. Step 3.

    Let us now prove (13). Pick εt0st\varepsilon t_{0}\leq s\leq t. Set h0=v0t01h_{0}=v_{0}\sqrt{t_{0}}^{-1}. Actually we prove a more general result, which will also be useful in the last regime. The convergence (13) will be a direct consequence of this lemma.

    Lemma 5.2.

    Let HH be an exponential ergodic process with invariant measure ν\nu, solution to a SDE driven by a Brownian motion. Pick a continuous function ϕ:[t0,+)\phi:[t_{0},+\infty)\to\mathbb{R} satisfying lims+ϕ(s)=+\lim_{s\to+\infty}\phi(s)=+\infty.
    Then, for all integer d1d\geq 1, every continuous and bounded function ψ:d\psi:\mathbb{R}^{d}\to\mathbb{R}, all h0h_{0}\in\mathbb{R} and all t1,,td[εt0,+)dt_{1},\cdots,t_{d}\in[\varepsilon t_{0},+\infty)^{d},

    |𝔼[ψ(Hϕ(ε1t1),,Hϕ(ε1td))|H0=h0]𝔼[ψ(Hϕ(ε1t1),,Hϕ(ε1td))|H0ν]|ε00.\left\lvert\mathbb{E}\left[\psi\left(H_{\phi(\varepsilon^{-1}t_{1})},\cdots,H_{\phi(\varepsilon^{-1}t_{d})}\right)\Big{|}H_{0}=h_{0}\right]-\mathbb{E}\left[\psi\left(H_{\phi(\varepsilon^{-1}t_{1})},\cdots,H_{\phi(\varepsilon^{-1}t_{d})}\right)\Big{|}H_{0}\sim\nu\right]\right\rvert\underset{\varepsilon\to 0}{\longrightarrow}0.
    Proof.

    For the sake of clarity, let us give a proof for d=2d=2. The general case d2d\geq 2 is similar.
    Let ψ:2\psi:\mathbb{R}^{2}\to\mathbb{R} be a continuous and bounded function.

    We set με:=(Hϕ(ε1s)|H0=h0)\mu_{\varepsilon}:=\mathcal{L}\left(H_{\phi(\varepsilon^{-1}s)}\Big{|}H_{0}=h_{0}\right). We use the generalized Markov property of solution to SDE driven by Brownian motion process (see Theorem 21.11 p. 421 in [Kal02]). This leads to

    𝔼[ψ(Hϕ(ε1s),Hϕ(ε1t))|H0=h0]=𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0με]\mathbb{E}\left[\psi\left(H_{\phi(\varepsilon^{-1}s)},H_{\phi(\varepsilon^{-1}t)}\right)\Big{|}H_{0}=h_{0}\right]=\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\mu_{\varepsilon}\right]

    and, since ΛF\Lambda_{F} is invariant,

    𝔼[ψ(Hϕ(ε1s),Hϕ(ε1t))|H0ν]=𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0ν].\mathbb{E}\left[\psi\left(H_{\phi(\varepsilon^{-1}s)},H_{\phi(\varepsilon^{-1}t)}\right)\Big{|}H_{0}\sim\nu\right]=\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\nu\right].

    Then, we are reduced to prove

    |𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0με]𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0ν]|ε00.\left\lvert\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\mu_{\varepsilon}\right]-\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\nu\right]\right\rvert\underset{\varepsilon\to 0}{\longrightarrow}0.

    Hence, setting p(t,x,dy):=x(Htdy)p(t,x,\mathop{}\!\mathrm{d}y):=\mathbb{P}_{x}(H_{t}\in\mathop{}\!\mathrm{d}y) and .TV\|.\|_{TV} for the total variation norm, we get

    |𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0με]𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0ν]||𝔼[ψ(H0,Hϕ(ε1t)ϕ(ε1s))|H0=y](με(dy)ν(dy))|ψ|p(ϕ(ε1s),h0,dy)ν(dy)|ψp(ϕ(ε1s),h0,)νTV.\left\lvert\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\mu_{\varepsilon}\right]-\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}\sim\nu\right]\right\rvert\\ \leq\left\lvert\int_{\mathbb{R}}\mathbb{E}\left[\psi\left(H_{0},H_{\phi(\varepsilon^{-1}t)-\phi(\varepsilon^{-1}s)}\right)\Big{|}H_{0}=y\right]\left(\mu_{\varepsilon}(\mathop{}\!\mathrm{d}y)-\nu(\mathop{}\!\mathrm{d}y)\right)\right\rvert\\ \leq\left\|\psi\right\|_{\infty}\int_{\mathbb{R}}\left\lvert p\left(\phi(\varepsilon^{-1}s),h_{0},\mathop{}\!\mathrm{d}y\right)-\nu(\mathop{}\!\mathrm{d}y)\right\rvert\\ \leq\left\|\psi\right\|_{\infty}\|p\left(\phi(\varepsilon^{-1}s),h_{0},\cdot\right)-\nu\|_{TV}.

    We let ε0\varepsilon\to 0, using the exponential ergodicity of HH. ∎

  4. Step 4.

    Let us prove now the tightness of the family of laws of continuous process (V(ε))tεt0=(εVt/ε)tεt0\left(V^{(\varepsilon)}\right)_{t\geq\varepsilon t_{0}}=\left(\sqrt{\varepsilon}V_{t/\varepsilon}\right)_{t\geq\varepsilon t_{0}} on every compact interval [m,M][m,M], 0<mM0<m\leq M. We prove the Kolmogorov criterion stated in Problem 4.11 p. 64, in [KS98].
    Take ε0\varepsilon_{0} small enough such that for all εε0\varepsilon\leq\varepsilon_{0}, εt0m\varepsilon t_{0}\leq m. Fix mstMm\leq s\leq t\leq M and α>2\alpha>2. Recalling that B(ε)B^{(\varepsilon)} is a Brownian motion, using Jensen’s inequality, moment estimates (4.1) and the relation β=γ+12\beta=\frac{\gamma+1}{2}, we can write

    𝔼[|Vt(ε)Vs(ε)|α]\displaystyle\mathbb{E}\left[\left\lvert V_{t}^{(\varepsilon)}-V_{s}^{(\varepsilon)}\right\rvert^{\alpha}\right] Cα𝔼[|Bt(ε)Bsε|α]+Cα𝔼[|εs/εt/εF(Vu)uβdu|α]\displaystyle\leq C_{\alpha}\mathbb{E}\left[\left\lvert B_{t}^{(\varepsilon)}-B_{s}^{\varepsilon}\right\rvert^{\alpha}\right]+C_{\alpha}\mathbb{E}\left[\left\lvert\sqrt{\varepsilon}\int_{s/\varepsilon}^{t/\varepsilon}F(V_{u})u^{-\beta}\mathop{}\!\mathrm{d}u\right\rvert^{\alpha}\right]
    Cα𝔼[|BtBs|α]+Cαε1α2(ts)α1𝔼[s/εt/ε|F(Vu)|αuβαdu]\displaystyle\leq C_{\alpha}\mathbb{E}\left[\left\lvert B_{t}-B_{s}\right\rvert^{\alpha}\right]+C_{\alpha}\varepsilon^{1-\frac{\alpha}{2}}(t-s)^{\alpha-1}\mathbb{E}\left[\int_{s/\varepsilon}^{t/\varepsilon}\left\lvert F(V_{u})\right\rvert^{\alpha}u^{-\beta\alpha}\mathop{}\!\mathrm{d}u\right]
    Cα𝔼[|Bts|α]+Cαε1α2(ts)α1s/εt/εuγα2βαdu\displaystyle\leq C_{\alpha}\mathbb{E}\left[\left\lvert B_{t-s}\right\rvert^{\alpha}\right]+C_{\alpha}\varepsilon^{1-\frac{\alpha}{2}}(t-s)^{\alpha-1}\int_{s/\varepsilon}^{t/\varepsilon}u^{\frac{\gamma\alpha}{2}-\beta\alpha}\mathop{}\!\mathrm{d}u
    Cα(ts)α2+Cαε1α2(ts)α1s/εt/εuα2du\displaystyle\leq C_{\alpha}(t-s)^{\frac{\alpha}{2}}+C_{\alpha}\varepsilon^{1-\frac{\alpha}{2}}(t-s)^{\alpha-1}\int_{s/\varepsilon}^{t/\varepsilon}u^{-\frac{\alpha}{2}}\mathop{}\!\mathrm{d}u
    Cα(ts)α2+Cα(ts)α1(t1α2s1α2)\displaystyle\leq C_{\alpha}(t-s)^{\frac{\alpha}{2}}+C_{\alpha}(t-s)^{\alpha-1}(t^{1-\frac{\alpha}{2}}-s^{1-\frac{\alpha}{2}})
    Cα(ts)α2+Cα,m,M(ts)α1\displaystyle\leq C_{\alpha}(t-s)^{\frac{\alpha}{2}}+C_{\alpha,m,M}(t-s)^{\alpha-1}
    Cα,m,M(ts)α2.\displaystyle\leq C_{\alpha,m,M}(t-s)^{\frac{\alpha}{2}}.

    Since α>2\alpha>2, then α2>1\frac{\alpha}{2}>1 and the upper bound does not depend on ε\varepsilon. Furthermore, by moment estimates (4.1),

    supεε0𝔼[|Vm(ε)|]m<+.\sup_{\varepsilon\leq\varepsilon_{0}}\mathbb{E}\left[\left\lvert V_{m}^{(\varepsilon)}\right\rvert\right]\leq\sqrt{m}<+\infty.
  5. Conclusion.

    The previous steps yields weak convergence on every compact set (Theorem 13.1 p. 139, in [Bil99]). The conclusion follows from Theorem 16.7 p. 174, in [Bil99], since all processes considered are continuous.

Example 5.1.

We will see that the limiting process 𝒱\mathcal{V} is more explicit in the linear case (γ=1\gamma=1). Choose F(1)=1F(1)=1, F(1)=1F(-1)=-1, the process H~\widetilde{H} solution of (10) is in fact an Ornstein Uhlenbeck process with invariant measure ΛF(dx):=e3x22dx\Lambda_{F}(dx):=e^{-\frac{3x^{2}}{2}}\mathop{}\!\mathrm{d}x. It is a centered Gaussian process, hence for all s1,,sds_{1},\cdots,s_{d}, its f.d.d. ΛF,s1,,sd\Lambda_{F,s_{1},\cdots,s_{d}} are Gaussian. As a consequence, knowing the covariance function KK is enough to provide the law of the process. Since H~\widetilde{H} is a stationary Ornstein-Uhlenbeck process, one has K:s,t13e32|ts|K:s,t\mapsto\frac{1}{3}e^{-\frac{3}{2}\left\lvert t-s\right\rvert}. Hence, the limiting process 𝒱\mathcal{V} having f.d.d TΛF,log(t1),,log(td)T*\Lambda_{F,\log(t_{1}),\cdots,\log(t_{d})} is a centered Gaussian process with covariance function s,t13(st)2sts,t\mapsto\frac{1}{3}\frac{(s\wedge t)^{2}}{s\vee t}.

5.3 Asymptotic behaviour in the subcritical regime under (H1γH_{1}^{\gamma})

Assume in this section that β<γ+12\beta<\frac{\gamma+1}{2} and F:vρsgn(v)|v|γF:v\mapsto\rho\operatorname{sgn}(v)\left\lvert v\right\rvert^{\gamma} with γ1\gamma\geq 1. For simplicity, we shall write φ\varphi instead of φq\varphi_{q}.

Proof of 2.5.
  1. Step 1.

    We first prove the f.d.d. convergence of the velocity process (Vt(ε))tεt0:=(εqVt/ε)tεt0(V_{t}^{(\varepsilon)})_{t\geq\varepsilon t_{0}}:=(\varepsilon^{q}V_{t/\varepsilon})_{t\geq\varepsilon t_{0}}. Again we give a proof only for d=2d=2, since the general case d2d\geq 2 is similar.
    The power scaling process V(q)V^{(q)}, solution to (3) satisfies

    dVs(q)=dWsF(Vs(q))dsqφ2q1(s)Vs(q)ds.\mathop{}\!\mathrm{d}V_{s}^{(q)}=\mathop{}\!\mathrm{d}W_{s}-F\Big{(}V_{s}^{(q)}\Big{)}\mathop{}\!\mathrm{d}s-q\varphi^{2q-1}(s)V_{s}^{(q)}\mathop{}\!\mathrm{d}s.

    We call HH the ergodic process solution to the SDE

    dHs=dWsF(Hs)ds, with H0=h0:=v0t0q.\mathop{}\!\mathrm{d}H_{s}=\mathop{}\!\mathrm{d}W_{s}-F\Big{(}H_{s}\Big{)}\mathop{}\!\mathrm{d}s,\mbox{ with }H_{0}=h_{0}:=v_{0}t_{0}^{-q}. (14)

    We denote by ΠF(dx):=e2ργ+1|x|γ+1dx\Pi_{F}(\mathop{}\!\mathrm{d}x):=e^{-\frac{2\rho}{\gamma+1}\left\lvert x\right\rvert^{\gamma+1}}\mathop{}\!\mathrm{d}x its invariant measure. Using the bijection induced by the power change of time (3.1), as solutions of the same SDE starting at the same point, we have, for all ε>0\varepsilon>0, and s,t[εt0,+)2s,t\in[\varepsilon t_{0},+\infty)^{2},

    (εqVε1ssq,εqVε1ttq)=(Vφ1(ε1s)(q),Vφ1(ε1t)(q)).\left(\varepsilon^{q}\dfrac{V_{\varepsilon^{-1}s}}{s^{q}},\varepsilon^{q}\dfrac{V_{\varepsilon^{-1}t}}{t^{q}}\right)=\left(V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}s)},V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}t)}\right).

    Using Theorem 3.1 p. 27, in [Bil99], it suffices to prove that for all s,t[εt0,+)2s,t\in[\varepsilon t_{0},+\infty)^{2},

    • |(Hφ1(ε1s),Hφ1(ε1t))(Vφ1(ε1s)(q),Vφ1(ε1t)(q))|2ε00\left\lvert\left(H_{\varphi^{-1}(\varepsilon^{-1}s)},H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)-\left(V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}s)},V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\right\rvert_{\mathbb{R}^{2}}\underset{\varepsilon\to 0}{\longrightarrow}0.

    • (Hφ1(ε1s),Hφ1(ε1t))ε0ΠFΠF\left(H_{\varphi^{-1}(\varepsilon^{-1}s)},H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\underset{\varepsilon\to 0}{\Longrightarrow}\Pi_{F}\otimes\Pi_{F}.

  2. Step 2.

    We prove that for all tεt0t\geq\varepsilon t_{0}, 𝔼[(Hφ1(ε1t)Vφ1(ε1t)(q))2]ε00\mathbb{E}\left[\left(H_{\varphi^{-1}(\varepsilon^{-1}t)}-V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}t)}\right)^{2}\right]\underset{\varepsilon\to 0}{\longrightarrow}0.
    Pick tεt0t\geq\varepsilon t_{0}. For simplicity of notation, we write Ht(φ,ε):=Hφ1(ε1t)H^{(\varphi,\varepsilon)}_{t}:=H_{\varphi^{-1}(\varepsilon^{-1}t)} and Vt(φ,ε):=Vφ1(ε1t)(q)V^{(\varphi,\varepsilon)}_{t}:=V^{(q)}_{\varphi^{-1}(\varepsilon^{-1}t)}. We have

    d(Ht(φ,ε)Vt(φ,ε))=ε2q1(F(Ht(φ,ε))F(Vt(φ,ε)))t2qdt+qt1Vt(φ,ε)dt.\mathop{}\!\mathrm{d}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)=-\varepsilon^{2q-1}\left(F(H^{(\varphi,\varepsilon)}_{t})-F(V^{(\varphi,\varepsilon)}_{t})\right)t^{-2q}\mathop{}\!\mathrm{d}t+qt^{-1}V^{(\varphi,\varepsilon)}_{t}\mathop{}\!\mathrm{d}t.

    Pick δ>0\delta>0. By straightforward differentiation, we can write

    d(Ht(φ,ε)Vt(φ,ε))22ε2q1t2q(F(Ht(φ,ε))F(Vt(φ,ε)))(Ht(φ,ε)Vt(φ,ε))𝟙|Ht(φ,ε)Vt(φ,ε)|>δdt+2t1qVt(φ,ε)(Ht(φ,ε)Vt(φ,ε))dt.\mathop{}\!\mathrm{d}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)^{2}\leq-\dfrac{2\varepsilon^{2q-1}}{t^{2q}}\left(F(H^{(\varphi,\varepsilon)}_{t})-F(V^{(\varphi,\varepsilon)}_{t})\right)\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)\mathbb{1}_{\left\lvert H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right\rvert>\delta}\mathop{}\!\mathrm{d}t\\ +2t^{-1}qV^{(\varphi,\varepsilon)}_{t}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)\mathop{}\!\mathrm{d}t.

    Since γ1\gamma\geq 1, the function F1F^{-1} is 1γ\frac{1}{\gamma}-Hölder, therefore there exists Cγ>0C_{\gamma}>0 such that,

    d(Ht(φ,ε)Vt(φ,ε))22ε2q1t2qCγδγ1(Ht(φ,ε)Vt(φ,ε))2𝟙|Ht(φ,ε)Vt(φ,ε)|>δdt+2t1qVt(φ,ε)(Ht(φ,ε)Vt(φ,ε))dt.\mathop{}\!\mathrm{d}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)^{2}\leq-\dfrac{2\varepsilon^{2q-1}}{t^{2q}}C_{\gamma}\delta^{\gamma-1}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)^{2}\mathbb{1}_{\left\lvert H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right\rvert>\delta}\mathop{}\!\mathrm{d}t\\ +2t^{-1}qV^{(\varphi,\varepsilon)}_{t}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)\mathop{}\!\mathrm{d}t. (15)

    We set gε(t)=𝔼[(Ht(φ,ε)Vt(φ,ε))2]g_{\varepsilon}(t)=\mathbb{E}\left[\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)^{2}\right] and g~ε(t)=𝔼[(Ht(φ,ε)Vt(φ,ε))2𝟙|Ht(φ,ε)Vt(φ,ε)|>δ]\widetilde{g}_{\varepsilon}(t)=\mathbb{E}\left[\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)^{2}\mathbb{1}_{\left\lvert H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right\rvert>\delta}\right]. Taking expectation in (15), we get

    g~ε(t)2ε2q1t2qCγδγ1g~ε(t)+bε(t), with g~ε(εt0)=0\widetilde{g}_{\varepsilon}^{\prime}(t)\leq-\dfrac{2\varepsilon^{2q-1}}{t^{2q}}C_{\gamma}\delta^{\gamma-1}\widetilde{g}_{\varepsilon}(t)+b_{\varepsilon}(t),\mbox{ with }\widetilde{g}_{\varepsilon}(\varepsilon t_{0})=0 (16)

    where

    bε(t):=2t1q𝔼[Vt(φ,ε)(Ht(φ,ε)Vt(φ,ε))].b_{\varepsilon}(t):=2t^{-1}q\mathbb{E}\left[V^{(\varphi,\varepsilon)}_{t}\left(H^{(\varphi,\varepsilon)}_{t}-V^{(\varphi,\varepsilon)}_{t}\right)\right].

    Using Cauchy-Schwarz inequality and moment estimates (4.1), we have

    |bε(t)|2t1|q|𝔼[(Vt(φ,ε))2]gε(t)C2t1|q|ε2q1t12qgε(t).\left\lvert b_{\varepsilon}(t)\right\rvert\leq 2t^{-1}\left\lvert q\right\rvert\sqrt{\mathbb{E}\left[\left(V^{(\varphi,\varepsilon)}_{t}\right)^{2}\right]}\sqrt{g_{\varepsilon}(t)}\leq C2t^{-1}\left\lvert q\right\rvert\sqrt{\varepsilon^{2q-1}t^{1-2q}g_{\varepsilon}(t)}.

    Set h(t):=112qCγδγ1t12qh(t):=\dfrac{1}{1-2q}C_{\gamma}\delta^{\gamma-1}t^{1-2q}.We use the comparison theorem for ordinary differential equation on (16) to get

    g~ε(t)εt0tbε(s)exp(2ε2q1(h(t)h(s)))ds.\widetilde{g}_{\varepsilon}(t)\leq\int_{\varepsilon t_{0}}^{t}b_{\varepsilon}(s)\exp(-2\varepsilon^{2q-1}\left(h(t)-h(s)\right))\mathop{}\!\mathrm{d}s.

    As a consequence, we deduce that

    gε(t)\displaystyle g_{\varepsilon}(t) δ2+g~ε(t)\displaystyle\leq\delta^{2}+\widetilde{g}_{\varepsilon}(t)
    δ2+exp(2ε2q1h(t))Cεt0t2s1ε2q1s12qgε(s)exp(2ε2q1h(s))exp(ε2q1h(s))ds.\displaystyle\leq\delta^{2}+\exp(-2\varepsilon^{2q-1}h(t))C\int_{\varepsilon t_{0}}^{t}2s^{-1}\sqrt{\varepsilon^{2q-1}s^{1-2q}}\sqrt{g_{\varepsilon}(s)\exp(2\varepsilon^{2q-1}h(s))}\exp(\varepsilon^{2q-1}h(s))\mathop{}\!\mathrm{d}s.

    Applying a Gronwall-type lemma (4.3) to the function gεexp(2ε2q1h)g_{\varepsilon}\exp(2\varepsilon^{2q-1}h), we obtain

    gε(t)\displaystyle g_{\varepsilon}(t) Cδ2+C(εt0ts1ε2q1s12qexp(ε2q1(h(t)h(s)))ds)2.\displaystyle\leq C\delta^{2}+C\left(\int_{\varepsilon t_{0}}^{t}s^{-1}\sqrt{\varepsilon^{2q-1}s^{1-2q}}\exp(-\varepsilon^{2q-1}(h(t)-h(s)))\mathop{}\!\mathrm{d}s\right)^{2}.

    We conclude, using the dominated convergence theorem, since 12q>01-2q>0, that for all δ>0\delta>0

    0lim supε0gε(t)δ2.0\leq\limsup_{\varepsilon\to 0}g_{\varepsilon}(t)\leq\delta^{2}. (17)

    To prove the domination hypothesis, notice that by optimization of the function xxexp(Ax)x\mapsto\sqrt{x}\exp(-Ax),

    𝟙εt0sts1ε2q1s12qexp(ε2q1(h(t)h(s)))𝟙0sts12q1h(t)h(s).\mathbb{1}_{\varepsilon t_{0}\leq s\leq t}s^{-1}\sqrt{\varepsilon^{2q-1}s^{1-2q}}\exp(-\varepsilon^{2q-1}(h(t)-h(s)))\leq\mathbb{1}_{0\leq s\leq t}s^{-\frac{1}{2}-q}\dfrac{1}{\sqrt{h(t)-h(s)}}.

    This function is integrable, since 12q>01-2q>0.
    We let δ0\delta\to 0 in (17) to conclude that for all t>0t>0, limε0gε(t)=0\lim_{\varepsilon\to 0}g_{\varepsilon}(t)=0.

  3. Step 3.

    Pick s,t[εt0,+)2s,t\in[\varepsilon t_{0},+\infty)^{2}. We prove that the solution HH to (14) satisfies

    (Hφ1(ε1s),Hφ1(ε1t))ε0ΠFΠF.\left(H_{\varphi^{-1}(\varepsilon^{-1}s)},H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\underset{\varepsilon\to 0}{\Longrightarrow}\Pi_{F}\otimes\Pi_{F}. (18)

    Observe that

    φ1(ε1t)φ1(ε1s)=t12qs12qε12qε00.\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)=\dfrac{t^{1-2q}-s^{1-2q}}{\varepsilon^{1-2q}}\underset{\varepsilon\to 0}{\longrightarrow}0. (19)

    By 5.2, for every continuous and bounded function φ\varphi, we can write

    |𝔼[ψ(Hφ1(ε1s),Hφ1(ε1t))|H0=h0]𝔼[ψ(Hφ1(ε1s),Hφ1(ε1t))|H0ΠF]|ε00.\left\lvert\mathbb{E}\left[\psi\left(H_{\varphi^{-1}(\varepsilon^{-1}s)},H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\Big{|}H_{0}=h_{0}\right]-\mathbb{E}\left[\psi\left(H_{\varphi^{-1}(\varepsilon^{-1}s)},H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\Big{|}H_{0}\sim\Pi_{F}\right]\right\rvert\underset{\varepsilon\to 0}{\longrightarrow}0.

    Hence, it suffices to prove that for every bounded continuous functions f,g:f,g:\mathbb{R}\to\mathbb{R}, the following convergence holds

    limε0𝔼[f(Hφ1(ε1s))g(Hφ1(ε1t))|H0ΠF]=ΠF(f)ΠF(g).\lim_{\varepsilon\to 0}\mathbb{E}\left[f\left(H_{\varphi^{-1}(\varepsilon^{-1}s)}\right)g\left(H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\Big{|}H_{0}\sim\Pi_{F}\right]=\Pi_{F}(f)\Pi_{F}(g).

    The following reasoning is inspired from the proof of Lemma 3.2 p. 7-8 in [CCM10]. Since H0H_{0} is starting from the invariant measure, up to considering fΠF(f)f-\Pi_{F}(f) and gΠF(g)g-\Pi_{F}(g), we can assume that ff and gg have zero ΠF\Pi_{F}-mean. We call (Pt)t0(P_{t})_{t\geq 0} the semigroup of HH, then we get, by invariance property of ΠF\Pi_{F},

    𝔼[f(Hφ1(ε1s))g(Hφ1(ε1t))|H0ΠF]\displaystyle\mathbb{E}\left[f\left(H_{\varphi^{-1}(\varepsilon^{-1}s)}\right)g\left(H_{\varphi^{-1}(\varepsilon^{-1}t)}\right)\Big{|}H_{0}\sim\Pi_{F}\right] =Pφ1(ε1s)(fPφ1(ε1t)φ1(ε1s)g)dΠF\displaystyle=\int P_{\varphi^{-1}(\varepsilon^{-1}s)}\left(fP_{\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)}g\right)\mathop{}\!\mathrm{d}\Pi_{F}
    =fPφ1(ε1t)φ1(ε1s)gdΠF.\displaystyle=\int fP_{\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)}g\mathop{}\!\mathrm{d}\Pi_{F}.

    Note that U:v|v|1+γ1+γU:v\mapsto\frac{\left\lvert v\right\rvert^{1+\gamma}}{1+\gamma} is a convex function, thus a λ\lambda-Poincaré inequality holds for the process HH (see [Bob99] p. 1904). This implies the exponential decay of the variance (see Theorem 4.2.5 p. 183, in [BGL14]), i.e. there exists a constant C>0C>0 such that, since ΠF\Pi_{F} is a probability measure,

    |fPφ1(ε1t)φ1(ε1s)gdΠF|\displaystyle\left\lvert\int fP_{\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)}g\mathop{}\!\mathrm{d}\Pi_{F}\right\rvert fPφ1(ε1t)φ1(ε1s)g2\displaystyle\leq\left\|fP_{\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)}g\right\|_{2}
    fPφ1(ε1t)φ1(ε1s)g2\displaystyle\leq\left\|f\right\|_{\infty}\left\|P_{\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)}g\right\|_{2}
    Cfgeλ(φ1(ε1t)φ1(ε1s)).\displaystyle\leq C\left\|f\right\|_{\infty}\left\|g\right\|_{\infty}e^{-\lambda\left(\varphi^{-1}(\varepsilon^{-1}t)-\varphi^{-1}(\varepsilon^{-1}s)\right)}.

    We deduce (18) from (19).

  4. Step 4.

    We prove the f.d.d. convergence of the position process (Xt(ε))tεt0:=(εβ+12Xt/ε)tεt0(X_{t}^{(\varepsilon)})_{t\geq\varepsilon t_{0}}:=(\varepsilon^{\beta+\frac{1}{2}}X_{t/\varepsilon})_{t\geq\varepsilon t_{0}}. Take γ=1\gamma=1 and β(12,1)\beta\in(-\frac{1}{2},1). Pick tεt0t\geq\varepsilon t_{0}. By Itô’s formula applied to tβVtt^{\beta}V_{t}, we get

    ρXt(ε)=εβ+12(t0βv0+x0)ε1β2tβVt(ε)+εβ+12t0t/εsβdBs+εβ+12t0t/εβsβ1Vsds.\rho X_{t}^{(\varepsilon)}=\varepsilon^{\beta+\frac{1}{2}}(t_{0}^{\beta}v_{0}+x_{0})-\varepsilon^{\frac{1-\beta}{2}}t^{\beta}V_{t}^{(\varepsilon)}+\varepsilon^{\beta+\frac{1}{2}}\int_{t_{0}}^{t/\varepsilon}s^{\beta}\mathop{}\!\mathrm{d}B_{s}+\varepsilon^{\beta+\frac{1}{2}}\int_{t_{0}}^{t/\varepsilon}\beta s^{\beta-1}V_{s}\mathop{}\!\mathrm{d}s.

    Since β>12\beta>-\frac{1}{2}, the first term converges to 0 in probability as ε0\varepsilon\to 0. Moreover, by Itô’s formula, for all tt0t\geq t_{0},

    ddt𝔼[Vt2]=2ρsβ𝔼[Vs2]+1.\dfrac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\mathbb{E}\left[V^{2}_{t}\right]=-2\rho s^{-\beta}\mathbb{E}\left[V^{2}_{s}\right]+1.

    Hence, by comparison theorem for ordinary differential equation,

    𝔼[Vt2]exp(2ρt1β1β)(v02+t0texp(2ρs1β1β)ds).\mathbb{E}\left[V^{2}_{t}\right]\leq\exp(-2\rho\dfrac{t^{1-\beta}}{1-\beta})\left(v_{0}^{2}+\int_{t_{0}}^{t}\exp(2\rho\dfrac{s^{1-\beta}}{1-\beta})\mathop{}\!\mathrm{d}s\right).

    Using an integration by parts, we deduce that there exists a positive constant CC such that, for all tt0t\geq t_{0},

    𝔼[Vt2]Ctβ.\mathbb{E}\left[V^{2}_{t}\right]\leq Ct^{\beta}.

    As a consequence, we obtain

    𝔼[|ε1β2tβVt(ε)+εβ+12t0t/εβsβ1Vsds|]\displaystyle\mathbb{E}\left[\left\lvert-\varepsilon^{\frac{1-\beta}{2}}t^{\beta}V_{t}^{(\varepsilon)}+\varepsilon^{\beta+\frac{1}{2}}\int_{t_{0}}^{t/\varepsilon}\beta s^{\beta-1}V_{s}\mathop{}\!\mathrm{d}s\right\rvert\right] ε1β2tβ𝔼[|Vt(ε)|]+εβ+12t0t/εβsβ1𝔼[|Vs|]ds\displaystyle\leq\varepsilon^{\frac{1-\beta}{2}}t^{\beta}\mathbb{E}\left[\left\lvert V_{t}^{(\varepsilon)}\right\rvert\right]+\varepsilon^{\beta+\frac{1}{2}}\int_{t_{0}}^{t/\varepsilon}\beta s^{\beta-1}\mathbb{E}\left[\left\lvert V_{s}\right\rvert\right]\mathop{}\!\mathrm{d}s
    Cε12t3β2+Cε1β2t3β2Cεβ+12t03β2ε00.\displaystyle\leq C\varepsilon^{\frac{1}{2}}t^{\frac{3\beta}{2}}+C\varepsilon^{\frac{1-\beta}{2}}t^{\frac{3\beta}{2}}-C\varepsilon^{\beta+\frac{1}{2}}t_{0}^{\frac{3\beta}{2}}\underset{\varepsilon\to 0}{\longrightarrow}0.

    It remains to study the centered Gaussian process Mt(ε):=εβ+12t0t/εsβdBsM^{(\varepsilon)}_{t}:=\varepsilon^{\beta+\frac{1}{2}}\int_{t_{0}}^{t/\varepsilon}s^{\beta}\mathop{}\!\mathrm{d}B_{s}. By Itô’s isometry and since β>12\beta>-\frac{1}{2}, for all εt0st\varepsilon t_{0}\leq s\leq t, we can write

    Cov(Ms(ε),Mt(ε))=ε2β+1t0s/εu2βdsε0s1+2β1+2β.\textrm{Cov}(M^{(\varepsilon)}_{s},M^{(\varepsilon)}_{t})=\varepsilon^{2\beta+1}\int_{t_{0}}^{s/\varepsilon}u^{2\beta}\mathop{}\!\mathrm{d}s\underset{\varepsilon\to 0}{\sim}\dfrac{s^{1+2\beta}}{1+2\beta}.

    Since the convergence of centered Gaussian processes is characterized by the convergence of their covariance functions, the conclusion follows from Theorem 3.1, p. 27, in [Bil99].

Acknowledgements

The authors would like to thank Jürgen Angst for valuable exchanges and suggestions about this work and Thomas Cavallazzi for his careful reading of the manuscript. We would also like to thank the anonymous referees for her/his careful reading of the manuscript and useful advices.

References

  • [Bét21] Loïc Béthencourt. Stable limit theorems for additive functionals of one-dimensional diffusion processes. arXiv:2104.06027 [math], April 2021.
  • [BGL14] Dominique Bakry, Ivan Gentil, and Michel Ledoux. Analysis and Geometry of Markov Diffusion Operators, volume 348 of Grundlehren Der Mathematischen Wissenschaften. Springer International Publishing, Cham, 2014.
  • [Bil99] Patrick Billingsley. Convergence of Probability Measures. Wiley Series in Probability and Statistics. Probability and Statistics Section. Wiley, 2nd ed edition, 1999.
  • [Bob99] S. G. Bobkov. Isoperimetric and Analytic Inequalities for Log-Concave Probability Measures. The Annals of Probability, 27(4):1903–1921, 1999.
  • [CCM10] Patrick Cattiaux, Djalil Chafai, and Sébastien Motsch. Asymptotic analysis and diffusion limit of the Persistent Turning Walker Model. Asymptotic Analysis, 67(1–2):17–31, 2010.
  • [CNP19] Patrick Cattiaux, Elissar Nasreddine, and Marjolaine Puel. Diffusion limit for kinetic Fokker-Planck equation with heavy tails equilibria: The critical case. Kinetic & Related Models, 12(4):727, 2019.
  • [EG15] Richard Eon and Mihai Gradinaru. Gaussian asymptotics for a non-linear langevin type equation driven by an α\alpha-stable lévy noise. Electron. J. Probab., 20, 2015.
  • [FT21] Nicolas Fournier and Camille Tardif. One dimensional critical Kinetic Fokker-Planck equations, Bessel and stable processes. Communications in Mathematical Physics, 381, January 2021.
  • [GL21] Mihai Gradinaru and Emeline Luirard. Asymptotic behavior for a time-inhomogeneous stochastic differential equation driven by an α\alpha-stable Lévy process. arXiv:2112.07287 [math], December 2021.
  • [GO13] Mihai Gradinaru and Yoann Offret. Existence and asymptotic behaviour of some time-inhomogeneous diffusions. Ann. Inst. H. Poincaré Probab. Statist., 49(1):182–207, February 2013.
  • [Kal02] Olav Kallenberg. Foundations of Modern Probability. Probability and Its Applications. Springer New York, New York, NY, 2002.
  • [KS98] Ioannis Karatzas and Steven Shreve. Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1998.
  • [Lui22] Emeline Luirard. Loi Limite de Modèles Cinétiques Inhomogènes En Temps. PhD thesis, 2022.