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

The Vlasov-Poisson and Vlasov-Poisson-Fokker-Planck systems in stochastic electromagnetic fields: local well-posedness

Jacob Bedrossian Department of Mathematics, University of Maryland, College Park, MD 20742, USA [email protected]. Both J.B. and S.P. were supported by National Science Foundation Award DMS-2108633.    Stavros Papathanasiou Department of Mathematics, University of Maryland, College Park, MD 20742, USA [email protected]
Abstract

In this paper, we construct unique, local-in-time strong solutions to the Vlasov-Poisson (VP) and Vlasov-Poisson-Fokker-Planck (VPFP) systems subjected to external, spatially regular, white-in-time electromagnetic fields in 𝕋d×d\mathbb{T}^{d}\times\mathbb{R}^{d}. Initial conditions are taken HσH^{\sigma} with σ>d/2+1\sigma>d/2+1 (in addition to polynomial velocity weights). We additionally show that solutions to the VPFP are instantly Cx,vC^{\infty}_{x,v} due to hypoelliptic regularization if the external force fields are smooth. The external forcing arises in the kinetic equation as a stochastic transport in velocity, which means, together with the anisotropy between xx and vv in the nonlinearity, that the local theory is a little more complicated than comparable fluid mechanics equations subjected to either additive stochastic forcing or stochastic transport. Although stochastic electromagnetic fields are often discussed in the plasma physics literature, to our knowledge, this is the first mathematical study of strong solutions to nonlinear stochastic kinetic equations.

1 Introduction

In this paper we prove the local-in-time existence and uniqueness of (probabilistically strong) solutions of the Vlasov and Vlasov-Fokker-Planck equations for a distribution of charged particles subjected to a stochastic external electric field

{df+vxfdt+Evfdt+vfdWt=νv(vf+fv)dtρ=fdvE=x(Δx)1(ρ1)ρ(t,x)=df(t,x,v)dv,\displaystyle\begin{cases}\mathrm{d}f+v\cdot\nabla_{x}f\mathrm{d}t+E\cdot\nabla_{v}f\mathrm{d}t+\nabla_{v}f\odot\mathrm{d}W_{t}=\nu\nabla_{v}\cdot(\nabla_{v}f+fv)\mathrm{d}t\\ \rho=\int f\mathrm{d}v\\ E=\nabla_{x}(-\Delta_{x})^{-1}(\rho-1)\\ \rho(t,x)=\int_{\mathbb{R}^{d}}f(t,x,v)\mathrm{d}v,\end{cases} (1.1)

where ν0\nu\geq 0 is the collision frequency – we treat both the case ν>0\nu>0 and ν=0\nu=0 (i.e. the Vlasov–Poisson equations). Below we denote the Fokker-Planck operator

f:=Δvf+v(vf),\mathcal{L}f:=\Delta_{v}f+\nabla_{v}\cdot(vf),

which is a commonly used simplification for collisions of charged particles against a background (see e.g. [boyd2003physics]).

Here, we consider the problem in the periodic box (x,v)𝕋d×d(x,v)\in\mathbb{T}^{d}\times\mathbb{R}^{d}, although the case (x,v)d×d(x,v)\in\mathbb{R}^{d}\times\mathbb{R}^{d} could be approached with similar arguments. The process WtW_{t} is a white-in-time, colored-in-space, vector-valued Gaussian process which plays the role of an external fluctuating electric field which we describe in more detail below. Our analysis works for general 1d31\leq d\leq 3 and also applies to external magnetic fields. For simplicity, we will take initial conditions in the velocity-weighted Sobolev space HmσH^{\sigma}_{m} defined by the norm:

fHmσ2:=|α|+|β|σ𝕋d×d|xαvβf(x,v)|2(1+|v|2)m/2dvdx.\left\|f\right\|_{H^{\sigma}_{m}}^{2}:=\sum_{|\alpha|+|\beta|\leq\sigma}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f(x,v)|^{2}(1+|v|^{2})^{m/2}\mathrm{d}v\mathrm{d}x.

Stochastic and randomly fluctuating electromagnetic fields are a classical topic in the plasma physics literature where they are used as a model for studying various dynamics in “turbulent”-like situations in both confined fusion and astrophysical applications. Much of the work is on studying the motion of charged particles (i.e. Lagrangian trajectories or passive scalars) subjected to stochastic electromagnetic fields of various kinds; see e.g. [hall1967diffusion, hasselmann1968scattering, jaekel1992fokker, hall1969particle, balescu1994langevin, krommes1983plasma, wingen2006influence, wang1995diffusive, vanden1996statistical] and the references therein for a tiny fraction of the existing work on the topic. Another line of work regards subjecting gyrokinetic equations or other macroscopic models to randomly fluctuating external force fields of this type for the purpose of studying plasma turbulence; see e.g. [tenbarge2014oscillating, navarro2016structure, told2015multiscale] and the references therein. The purpose of this work is to begin laying down some rigorous mathematical theory for studying nonlinear kinetic theory models of plasmas in these kinds of settings.

It is sometimes useful to make a concrete representation of WtW_{t} and for simplicity we will show how to do this in d=3d=3; the extension to other dimensions is straightforward and is omitted. To make this concrete representation of WtW_{t}, we define a real Fourier basis on L2(𝕋3;3)L^{2}(\mathbb{T}^{3};\mathbb{R}^{3}) by defining for each k=(,i)𝕂:=0d×{1,2,3}k=(\ell,i)\in\mathbb{K}:=\mathbb{Z}^{d}_{0}\times\{1,2,3\}

ek(x)={cdγisin(x),+dcdγicos(x),d,e_{k}(x)=\begin{cases}c_{d}\gamma_{\ell}^{i}\sin(\ell\cdot x),\quad&\ell\in\mathbb{Z}^{d}_{+}\\ c_{d}\gamma_{\ell}^{i}\cos(\ell\cdot x),\quad&\ell\in\mathbb{Z}^{d}_{-},\end{cases}

where 0d:=d{0,,0}\mathbb{Z}^{d}_{0}:=\mathbb{Z}^{d}\setminus\left\{0,\ldots,0\right\}, +d={0d:(d)>0}{0d:(1)>0,(d)=0}\mathbb{Z}_{+}^{d}=\{\ell\in\mathbb{Z}^{d}_{0}:\ell^{(d)}>0\}\cup\{\ell\in\mathbb{Z}^{d}_{0}\,:\,\ell^{(1)}>0,\ell^{(d)}=0\} and d=+d\mathbb{Z}_{-}^{d}=-\mathbb{Z}_{+}^{d}, and for each 0d\ell\in\mathbb{Z}_{0}^{d}, {γi}i=13\{\gamma_{\ell}^{i}\}_{i=1}^{3} is a set of three orthonormal vectors with {γ1,γ2}\left\{\gamma_{\ell}^{1},\gamma_{\ell}^{2}\right\} spanning the plane perpendicular to 3\ell\in\mathbb{R}^{3} with the property that γi=γi\gamma_{-\ell}^{i}=-\gamma_{\ell}^{i} and γ3\gamma_{\ell}^{3} parallel to \ell. The constant cd=2(2π)d/2c_{d}=\sqrt{2}(2\pi)^{-d/2} is a normalization factor so that ek(x)e_{k}(x) are a complete orthonormal basis on L2L^{2}. With this, we define our external electric field as

Wt(x)=k𝕂σkek(x)Wt(k),\displaystyle W_{t}(x)=\sum_{k\in\mathbb{K}}\sigma_{k}e_{k}(x)W_{t}^{(k)},

with {Wt(k)}k𝕂\left\{W_{t}^{(k)}\right\}_{k\in\mathbb{K}} is a family of independent standard Wiener processes defined on a given stochastic basis (Ω,,t,𝐏)(\Omega,\mathcal{F},\mathcal{F}_{t},\mathbf{P}). The σk\sigma_{k} are coloring coefficients satisfying at least k𝕂|σk|2<\sum_{k\in\mathbb{K}}\left|\sigma_{k}\right|^{2}<\infty, however, more stringent regularity requirements will be assumed below (here we make the natural definition |k|=||\left|k\right|=\left|\ell\right| for k=(,i)𝕂k=(\ell,i)\in\mathbb{K}). We can also treat the case of fluctuating magnetic fields; see Remark 1.4 below.

Local well-posedness of strong solutions for the deterministic problem is classical; see e.g. [horst1981classical, horst1987global] for the Vlasov equations and [neunzert1984vlasov, victory1990classical] for the Vlasov-Fokker-Planck equation. Global existence for the deterministic problems was proved in [pfaffelmoser1992global] (see also [schaeffer1991global, batt1991global]) for the Vlasov equations and [bouchut1993existence] for the Vlasov-Fokker-Planck equations; we will consider global existence for (1.1) in a follow up work. Notice that in Itô form, the SPDE becomes

df+vxfdt+Evfdt+vfdWt=νfdt+12k(σkekv)2fdt,\mathrm{d}f+v\cdot\nabla_{x}f\mathrm{d}t+E\cdot\nabla_{v}f\mathrm{d}t+\nabla_{v}f\cdot\mathrm{d}W_{t}=\nu\mathcal{L}f\mathrm{d}t+\frac{1}{2}\sum_{k}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f\mathrm{d}t, (1.2)

so it is clear that stochastic transport cannot be treated perturbatively with respect to the deterministic evolution, as the Stratonovich-Itô correction term is of second order. However, this correction term is subelliptic, and so stochastic transport enjoys a special structure that makes it possible to develop a strong well-posedness theory, and in fact, it is sometimes possible to produce a better well-posedness theory for stochastic transport than for deterministic transport [flandoli2010well]. Due to this special structure and the many physical applications, there have been a great number of works studying stochastic transport equations recently; see for example [fedrizzi2011pathwise, fedrizzi2013noise, beck2019stochastic, mohammed2015sobolev, champagnat2018strong] and also [fedrizzi2017regularity, de2018invariant] in the kinetic case.

There have been many works on fluid equations subjected to multiplicative or transport-type stochastic forcing. For the Navier-Stokes equations see for example [brzezniak1992stochastic, mikulevicius2004stochastic, brzezniak2013existence, capinski1993navier, mikulevicius2005global]. The 2D Euler equations in vorticity form subjected to transport noise was studied in, for example [brzezniak2016existence, crisan2019well], where strong solutions with bounded vorticity were constructed (see also [crisan2019solution]). The aforementioned papers [crisan2019well, crisan2019solution] belong to the so-called theory of \sayStochastic Advection by Lie Transport (SALT), see the foundational paper [holm2015variational], as well as [crisan2020local, alonso2020well]. The work [brzezniak2020well] studies the 3D primitive equations with transport noise.

The works [debussche2011local, debussche2012global] provide a fairly general framework to study a wide class of dissipative fluid equations forced with multiplicative noise, such as the Navier-Stokes equations or the primitive equations. For the 2D Euler equations with various types of general multiplicative noise, see [GV14] and the references therein.

In comparison to stochastic fluid dynamics, the work on nonlinear, stochastically forced kinetic equations is significantly thinner. The paper [punshon2018boltzmann] constructs global-in-time renormalized martingale (probabilistically weak) solutions of the Boltzmann equations with external stochastic forcing similar to that used in (1.1). In work with a clear relationship with our own, [delarue2014noise] constructs global solutions of interacting point charges (i.e. Vlasov–Poisson with solutions given by a finite number of Dirac masses) subjected to stochastic external electric fields; see also [coghi2016propagation].

In this paper we continue the study of stochastic kinetic theory by proving local existence and pathwise uniqueness of strong solutions. Let us recall some standard notions for probabilistically strong solutions of SPDEs that may experience finite-time blow up (we follow the presentation used in [debussche2012global, GV14]), which are nothing more than the natural stochastic analogues of the deterministic notions of local-in-time existence, uniqueness, and maximally-extended solutions.

Definition 1.1.

A local pathwise solution of (1.1) is a pair (f,τ)(f,\tau) with τ\tau an almost-surely strictly positive stopping time and ff an adapted stochastic process satisfying the regularity

f(τ)C([0,);Hmσ)\displaystyle f(\cdot\wedge\tau)\in C([0,\infty);H^{\sigma}_{m})

and for t0t\geq 0 satisfies,

f(tτ)f(0)+0tτ(vxf(s)+E(s)vfνf(s))𝑑τ=0tτvf(s)dWs.\displaystyle f(t\wedge\tau)-f(0)+\int_{0}^{t\wedge\tau}\left(v\cdot\nabla_{x}f(s)+E(s)\cdot\nabla_{v}f-\nu\mathcal{L}f(s)\right)d\tau=\int_{0}^{t\wedge\tau}\nabla_{v}f(s)\circ\mathrm{d}W_{s}.

Moreover, we say such pathwise solutions are unique if for any pair (f1,τ1)(f_{1},\tau_{1}), (f2,τ2)(f_{2},\tau_{2}) we have

(f1(t)f2(t)=00t<τ1τ2|f1(0)=f2(0))=1.\displaystyle\mathbb{P}\left(f_{1}(t)-f_{2}(t)=0\quad\forall 0\leq t<\tau_{1}\wedge\tau_{2}|f_{1}(0)=f_{2}(0)\right)=1.

In this case, f1f_{1} and f2f_{2} are called indistinguishable.

The following definition of maximal pathwise solution provides a continuation criterion for strong solutions. For this we will use Hd/2+d/2+1+H^{d/2+1+}_{d/2+}; sharper continuation criteria will be considered in future work. That is, we show that local HmσH^{\sigma}_{m} solutions can be uniquely extended provided some Hm0s0H^{s_{0}}_{m_{0}} norm remains finite for s0>1+d/2s_{0}>1+d/2 and m0>dm_{0}>d fixed and arbitrary.

Definition 1.2.

Fix s0>d/2+1s_{0}>d/2+1 an integer and m0>dm_{0}>d. We call a maximal pathwise solution a triple of a solution ff, an increasing sequence of almost-surely positive stopping times {τn}n0\left\{\tau_{n}\right\}_{n\geq 0}, and a limiting stopping time ξ\xi such that each pair (f,τn)(f,\tau_{n}) is a local pathwise solution, limnτn=ξ\lim_{n\to\infty}\tau_{n}=\xi, and

sup0tτnf(t)Hm0s0n on the set {ξ<}.\displaystyle\sup_{0\leq t\leq\tau_{n}}\left\|f(t)\right\|_{H^{s_{0}}_{m_{0}}}\geq n\quad\textup{ on the set }\left\{\xi<\infty\right\}.

In this paper, we prove the following local existence and uniqueness theorem. Global existence of these strong solutions will be considered in a future work.

Theorem 1.3.

Let 1d31\leq d\leq 3. Let σ>s0\sigma>s_{0} and m>m0m>m_{0} be fixed integers and assume that

k𝕂|k|2σ|σk|2<\sum_{k\in\mathbb{K}}\left|k\right|^{2\sigma^{\prime}}\left|\sigma_{k}\right|^{2}<\infty (1.3)

for some σ>σ+4\sigma^{\prime}>\sigma+4 (integer). Suppose that the initial condition f0f_{0} is an 0\mathcal{F}_{0}-measurable random variable such that f0Hmσf_{0}\in H^{\sigma}_{m} almost-surely. Then, there exists a unique, maximal pathwise solution to (1.1) for any ν0\nu\geq 0.

Remark 1.4.

Our proof also applies when there is a stochastic magnetic field. We may similarly treat the case of independent electric and magnetic fields as the following, for example (denoting c>0c>0 the speed of light),

Wt(x)\displaystyle W_{t}(x) k𝕂σk(E)ek(x)Wtk;E\displaystyle\mapsto\sum_{k\in\mathbb{K}}\sigma_{k}^{(E)}e_{k}(x)W_{t}^{k;E}
+1cv×03σ(B;1)e(,1)(x)Wt(,1);B+1cv×03σ(B;2)e(,2)(x)Wt(,2);B,\displaystyle\quad+\frac{1}{c}v\times\sum_{\ell\in\mathbb{Z}^{3}_{0}}\sigma_{\ell}^{(B;1)}e_{(\ell,1)}(x)W_{t}^{(\ell,1);B}+\frac{1}{c}v\times\sum_{\ell\in\mathbb{Z}^{3}_{0}}\sigma_{\ell}^{(B;2)}e_{(\ell,2)}(x)W_{t}^{(\ell,2);B},

with

k=(,j)𝕂|k|2σ(|σk(E)|2+|σ(B;1)|2+|σ(B;2)|2)<,\displaystyle\sum_{k=(\ell,j)\in\mathbb{K}}\left|k\right|^{2\sigma^{\prime}}\left(\left|\sigma_{k}^{(E)}\right|^{2}+\left|\sigma_{\ell}^{(B;1)}\right|^{2}+\left|\sigma_{\ell}^{(B;2)}\right|^{2}\right)<\infty, (1.4)

or when electromagnetic fields are correlated, for example one could use forcing of the following potentially natural form

Wt\displaystyle W_{t} 03σ(1)(e(,1)(x)+v×e(,2)(x)c)Wt(,1)+03σ(2)(e(,2)(x)+v×e(,1)(x)c)Wt(,2).\displaystyle\mapsto\sum_{\ell\in\mathbb{Z}^{3}_{0}}\sigma_{\ell}^{(1)}\left(e_{(\ell,1)}(x)+\frac{v\times e_{(\ell,2)}(x)}{c}\right)W_{t}^{(\ell,1)}+\sum_{\ell\in\mathbb{Z}^{3}_{0}}\sigma_{\ell}^{(2)}\left(e_{(\ell,2)}(x)+\frac{v\times e_{(\ell,1)}(x)}{c}\right)W_{t}^{(\ell,2)}.

Sufficiently regular-in-space deterministic external electromagnetic fields or random fields that are smoother in time than white noise (for example, Ornstein-Uhlenbeck processes and variations thereof as in the Langevin antenna forcing used in the plasma physics literature [tenbarge2014oscillating]) can also be easily included in the analysis without any significant changes. For simplicity of presentation, we will mainly focus on the case of external electric fields and simply make comments about what changes when considering an additional magnetic field.

Remark 1.5.

The methods of this paper can also deal with with more general mean-field interactions, replacing the self-consistent electric field with:

E=xK(ρ1),\displaystyle E=\nabla_{x}K\ast(\rho-1),

for any kernel KK such that xj+1ELppxjρLp\left\|\nabla_{x}^{j+1}E\right\|_{L^{p}}\lesssim_{p}\left\|\left\langle\nabla_{x}\right\rangle^{j}\rho\right\|_{L^{p}} for all 1<p<1<p<\infty.

Remark 1.6.

It should be straightforward to extend to d4d\geq 4. It should also be possible to treat non-integer σ\sigma, s0s_{0}, and σ\sigma^{\prime}, however, this would require more delicate (anisotropic) commutator estimates.

Remark 1.7.

We believe our methods could be extended to the Landau collision operators for initial data f0f_{0} sufficiently close to a global Maxwellian to prove local-in-time existence and uniqueness of strong solutions to e.g. the Vlasov–Poisson–Landau equations with stochastic external electromagnetic fields. This extension may be considered in future work.

Remark 1.8.

In light of the classical deterministic theory of bounded solutions of the Vlasov equations (see e.g. [lions1991propagation, loeper2006uniqueness]), it is natural to expect an analogue of [brzezniak2016existence] in kinetic theory. Similarly, we expect local (and global) existence and uniqueness of the Vlasov-Fokker-Planck equations using only e.g. f0Lm2f_{0}\in L^{2}_{m}. These extensions may be considered in future work.

Finally, in Section 5 we present a proof of the following hypoelliptic regularization result. This is proved using a time-dependent hypocoercivity norm in the spirit of [dric2009hypocoercivity].

Theorem 1.9.

Let ff be a maximal pathwise solution to (1.1) as in Theorem 1.3. Suppose that for all NN there holds

|σk|N|k|N.\displaystyle\left|\sigma_{k}\right|\lesssim_{N}\left|k\right|^{-N}.

Then if ν>0\nu>0, then f(t)Cx,vf(t)\in C^{\infty}_{x,v} for all t(0,ξ)t\in(0,\xi).

2 Outline

Let us outline the general idea of how to prove Theorem 1.3 and then provide the details in the main body of the text. See Section 5 for how to prove Theorem 1.9.

As in e.g. [debussche2012global, GV14], we first construct solutions to (1.1) with smoother initial data HmσH^{\sigma^{\prime}}_{m^{\prime}} with σ>σ+4\sigma^{\prime}>\sigma+4 and m>m+3m^{\prime}>m+3 (both integers) with trajectories in Lt,locHmσCt,locHmσL^{\infty}_{t,loc}H^{\sigma^{\prime}}_{m^{\prime}}\cap C_{t,loc}H^{\sigma}_{m}. This procedure is done in Section 3. Then we regularize the initial condition and pass to the limit to obtain solutions with initial data in HmσH^{\sigma}_{m} that take values in Ct,locHmσC_{t,loc}H^{\sigma}_{m}. In addition to obtaining solutions with lower regularity, what is more important for many purposes, is that this constructs solutions which take values continuously in the highest regularity available. This latter procedure is done in Section 4.

To construct maximal pathwise solutions to (1.1) we first introduce a standard trick for regularizing the nonlinearity in a way which allows to close necessary probabilistic moment estimates. Consider a smooth nonnegative and nonincreasing function θ:[0,)\theta:[0,\infty)\to\mathbb{R} such that:

θ(x)={1 if x1,0 if x2,\theta(x)=\begin{cases}1&\text{ if }x\leq 1,\\ 0&\text{ if }x\geq 2,\end{cases} (2.1)

and define:

θR(x)θ(xR).\theta_{R}(x)\equiv\theta\left(\frac{x}{R}\right). (2.2)

Then we define the regularized SPDE

df+vxfdt+θR(fHm0s0)Evfdt+vfdWt=νfdt,\mathrm{d}f+v\cdot\nabla_{x}f\mathrm{d}t+\theta_{R}(\left\|f\right\|_{H^{s_{0}}_{m_{0}}})E\cdot\nabla_{v}f\mathrm{d}t+\nabla_{v}f\odot\mathrm{d}W_{t}=\nu\mathcal{L}f\mathrm{d}t, (2.3)

We show in Section 3 that this SPDE admits global-in-time, unique, pathwise solutions (i.e. ξ=\xi=\infty with probability 11 in the definition of maximal solutions) starting from HmσH^{\sigma^{\prime}}_{m^{\prime}} initial conditions. Specifically, we prove the following.

Lemma 2.1.

Let f0f_{0} be a 0\mathcal{F}_{0}-measurable random variable such that p2\forall p\geq 2,

𝐄f0Hmσp<.\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}^{p}<\infty.

Then, there exists an fC([0,);Hm3σ4)Lt,loc([0,);Hm1σ1)f\in C([0,\infty);H_{m^{\prime}-3}^{\sigma^{\prime}-4})\cap L_{t,\text{loc}}^{\infty}([0,\infty);H_{m^{\prime}-1}^{\sigma^{\prime}-1}) 𝐏\mathbf{P}–a.s. which is a solution to (2.3) in the sense that

f(t)=f0+0t(vxf(s)θR(fHm0s0)E(s)vf(s)+νf(s))ds0tvf(s)dWs.𝐏–a.s.,f(t)=f_{0}+\int_{0}^{t}\left(-v\cdot\nabla_{x}f(s)-\theta_{R}(\left\|f\right\|_{H^{s_{0}}_{m_{0}}})E(s)\cdot\nabla_{v}f(s)+\nu\mathcal{L}f(s)\right)\mathrm{d}s-\int_{0}^{t}\nabla_{v}f(s)\circ\mathrm{d}W_{s}.\quad\mathbf{P}\text{--a.s.}, (2.4)

where the equality holds in C([0,);Hm3σ4)C([0,\infty);H^{\sigma^{\prime}-4}_{m^{\prime}-3}). Moreover, if f~\tilde{f} is any other solution in the above sense, then f=f~f=\tilde{f} almost surely in the sense that

(f(t)f~(t)=00t<|f(0)=f~(0))=1.\displaystyle\mathbb{P}\left(f(t)-\tilde{f}(t)=0\quad\forall 0\leq t<\infty|f(0)=\tilde{f}(0)\right)=1.

It is clear that solutions of (2.3) are also solutions to (1.1) for as long as fHm0s0<R\left\|f\right\|_{H_{m_{0}}^{s_{0}}}<R, and so by considering the increasing sequence of R=nR=n and defining the stopping times

τn=inf{t0:f(t)Hm0s0>n},\displaystyle\tau_{n}=\inf\left\{t\geq 0:\left\|f(t)\right\|_{H^{s_{0}}_{m_{0}}}>n\right\},

we may use (2.3) to construct local pathwise solutions to (1.1). A standard cutting procedure (described below) also shows how to remove the moment requirement on the initial data.

Lemma 2.2.

Let f0f_{0} be an 0\mathcal{F}_{0}-measurable random variable with f0Hmσf_{0}\in H^{\sigma^{\prime}}_{m^{\prime}} almost surely. Then, Lemma 2.1 implies the pathwise existence and uniqueness of a maximal solution (f,τ)(f,\tau) to (1.1) with initial data f0f_{0} with trajectories ff satisfying

f(τ)Lt,loc(0,;Hm1σ1)Ct,loc([0,);Hm3σ4).\displaystyle f(\cdot\wedge\tau)\in L^{\infty}_{t,loc}(0,\infty;H^{\sigma^{\prime}-1}_{m^{\prime}-1})\cap C_{t,loc}([0,\infty);H^{\sigma^{\prime}-4}_{m^{\prime}-3}).
Proof.

First consider the case that f0Hmσ<M\left\|f_{0}\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}<M almost-surely. Then, we choose R=M+1R=M+1 in (2.3), and define the stopping time:

τ=inf{t0:f(t)Hmσ>R},\tau=\inf\{t\geq 0:\,\left\|f(t)\right\|_{H_{m}^{\sigma}}>R\},

where ff is the solution to (2.3) with initial data f0f_{0}, guaranteed to exist and be unique from Lemma 2.1. Note that up to time τ\tau, the process ff also solves (1.1), since for t<τt<\tau we have f(t)Hm0s0f(t)HmσR\|f(t)\|_{H_{m_{0}}^{s_{0}}}\leq\|f(t)\|_{H_{m}^{\sigma}}\leq R and therefore θR(f(t)Hm0s0)=1.\theta_{R}(\|f(t)\|_{H_{m_{0}}^{s_{0}}})=1. Clearly τ>0\tau>0 almost surely since Hm2σ2HmσH_{m^{\prime}-2}^{\sigma^{\prime}-2}\subset H_{m}^{\sigma} and ff takes values continuously in Hm2σ2H_{m^{\prime}-2}^{\sigma^{\prime}-2}. The pair (f,τ)(f,\tau) is thus a local solution of (1.1) within the higher regularity framework of this lemma, which is unique by Lemma 3.13 below. Now we will extend ff to a maximal solution.

Let 𝒯\mathcal{T} be the collection of all stopping times corresponding to a local solution and define ξ=sup𝒯\xi=\sup\mathcal{T}. Define also:

τn:=inf{t0:f(t)Hm0s0>n}.\tau_{n}:=\inf\{t\geq 0:\,\left\|f(t)\right\|_{H_{m_{0}}^{s_{0}}}>n\}.

Fix T>0T>0 finite but arbitrary and assume 𝐏(ξ=τnT)>0\mathbf{P}(\xi=\tau_{n}\wedge T)>0 for some nn. This implies that

suptξf(t)Hm0s0n on the set {ξ=τnT},\displaystyle\sup_{t\leq\xi}\left\|f(t)\right\|_{H_{m_{0}}^{s_{0}}}\leq n\textup{ on the set }\{\xi=\tau_{n}\wedge T\},

and thus ff can be continued to a solution of (2.3) with R=n+1R=n+1 and thus of (1.1) up to a stopping time past ξ\xi - contradicting ξ\xi’s maximality. Since TT was arbitrary, we either have ξ=\xi=\infty, or τn<ξ<\tau_{n}<\xi<\infty for all nn. In the latter case, we also get supt<ξf(t)Hm0s0n\sup_{t<\xi}\left\|f(t)\right\|_{H_{m_{0}}^{s_{0}}}\geq n for all nn and thus supt<ξf(t)Hm0s0=.\sup_{t<\xi}\left\|f(t)\right\|_{H_{m_{0}}^{s_{0}}}=\infty.

Now we drop the almost-sure uniform boundedness requirement. If f0Hmσ<\left\|f_{0}\right\|_{H_{m^{\prime}}^{\sigma^{\prime}}}<\infty almost surely, we decompose f0=k=0f0,k,f_{0}=\sum_{k=0}^{\infty}f_{0,k}, where f0,k:=𝟙{kf0Hmσ<k+1}f0f_{0,k}:=\mathbbm{1}_{\{k\leq\left\|f_{0}\right\|_{H_{m^{\prime}}^{\sigma^{\prime}}}<k+1\}}f_{0}. Now each f0,kf_{0,k} generates a maximal solution (fk,τk)(f_{k},\tau_{k}) where τk\tau_{k} is the corresponding maximal existence time, and we define the \saytotal maximal solution (in high regularity) of (1.1) as (f¯,τ¯)(\bar{f},\bar{\tau}) with:

f¯=k=0𝟙{kf0Hmσ<k+1}(ω)fk,\displaystyle\bar{f}=\sum_{k=0}^{\infty}\mathbbm{1}_{\{k\leq\left\|f_{0}\right\|_{H_{m^{\prime}}^{\sigma^{\prime}}}<k+1\}}(\omega)f_{k}, (2.5)
τ¯=k=0𝟙{kf0Hmσ<k+1}(ω)τk.\displaystyle\bar{\tau}=\sum_{k=0}^{\infty}\mathbbm{1}_{\{k\leq\left\|f_{0}\right\|_{H_{m^{\prime}}^{\sigma^{\prime}}}<k+1\}}(\omega)\tau_{k}. (2.6)

Solutions to (2.3) are constructed using a two-step procedure. First, we regularize the nonlinearity again and use an iteration procedure to construct a solution to the regularized SPDE and then second, we pass to the limit in the additional regularization parameter. Let φCc(B(0,2))\varphi\in C^{\infty}_{c}(B(0,2)) with nφdx=1\int_{\mathbb{R}^{n}}\varphi\mathrm{d}x=1 and define φϵ=ϵdφ(ϵ1)\varphi_{\epsilon}=\epsilon^{-d}\varphi(\epsilon^{-1}\cdot). Specifically, we seek a solution to the following regularized SPDE (here the convolution in xx has been periodized),

df~+vxf~dt+vf~dWt+θR(f~Hm0s0)(φϵE~)vf~=νf~dt\displaystyle\mathrm{d}\tilde{f}+v\cdot\nabla_{x}\tilde{f}\mathrm{d}t+\nabla_{v}\tilde{f}\circ\mathrm{d}W_{t}+\theta_{R}(\left\|\tilde{f}\right\|_{H_{m_{0}}^{s_{0}}})(\varphi_{\epsilon}\ast\tilde{E})\cdot\nabla_{v}\tilde{f}=\nu\mathcal{L}\tilde{f}\mathrm{d}t (2.7)
E~=x(Δx)1(df~(t,,v)dv1).\displaystyle\tilde{E}=\nabla_{x}(-\Delta_{x})^{-1}\left(\int_{\mathbb{R}^{d}}\tilde{f}(t,\cdot,v)\mathrm{d}v-1\right).

This is done by an iteration method, specifically the following

fj(0)=f0\displaystyle f^{j}(0)=f_{0} (2.8)
df0+vxf0dt+vf0dWt=νf0dt\displaystyle\mathrm{d}f^{0}+v\cdot\nabla_{x}f^{0}\mathrm{d}t+\nabla_{v}f^{0}\odot\mathrm{d}W_{t}=\nu\mathcal{L}f^{0}\mathrm{d}t (2.9)
dfj+1+vxfj+1dt+θR(fjHm0s0)(φϵE[fj])vfj+1dt+vfj+1dWt=νfj+1dt,\displaystyle\mathrm{d}f^{j+1}+v\cdot\nabla_{x}f^{j+1}\mathrm{d}t+\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}})(\varphi_{\epsilon}\ast E[f^{j}])\cdot\nabla_{v}f^{j+1}\mathrm{d}t+\nabla_{v}f^{j+1}\odot\mathrm{d}W_{t}=\nu\mathcal{L}f^{j+1}\mathrm{d}t, (2.10)

where we denote

E[f]:=x(Δx)1(df(t,,v)dv1).\displaystyle E[f]:=\nabla_{x}(-\Delta_{x})^{-1}\left(\int_{\mathbb{R}^{d}}f(t,\cdot,v)\mathrm{d}v-1\right).

The solutions to (2.9) - (2.10) are constructed by the method of characteristics. Indeed, (2.10) is the forward Kolmogorov equation associated to the SDE

{dXt=VtdtdVt=νVtdt+θR(fjHm0s0)(φϵE[fj])(t,Xt)dt+2νdW~t+kσkek(Xt)dWt(k),\begin{cases}\mathrm{d}X_{t}=V_{t}\mathrm{d}t\\ \mathrm{d}V_{t}=-\nu V_{t}\mathrm{d}t+\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}})(\varphi_{\epsilon}\ast E[f^{j}])(t,X_{t})\mathrm{d}t+\sqrt{2\nu}\mathrm{d}\tilde{W}_{t}+\sum_{k}\sigma_{k}e_{k}(X_{t})\circ\mathrm{d}W_{t}^{(k)},\end{cases} (2.11)

where (W~t)(\tilde{W}_{t}) is a dd-dimensional Brownian motion defined in a new stochastic basis (Ω,,𝐏)(\Omega^{\prime},\mathcal{F}^{\prime},\mathbf{P}^{\prime}) (independent of the original basis). That is, (2.11) are the stochastic characteristics corresponding to (2.10), which generates a global stochastic flow of volume-preserving diffeomorphisms ϕt\phi_{t} on 𝕋d×d\mathbb{T}^{d}\times\mathbb{R}^{d}, defined on the product space

(Ω×Ω,,𝐏×𝐏),(\Omega\times\Omega^{\prime},\mathcal{F}\otimes\mathcal{F}^{\prime},\mathbf{P}\times\mathbf{P}^{\prime}),

which map HmσH^{\sigma^{\prime}}_{m^{\prime}} back to itself for all finite times almost-surely. The multiplicative (linear!) SPDE (2.10) is then solved by a \saypartial Feynman-Kac formula with respect to 𝐏:\mathbf{P}^{\prime}:

fj+1=𝐄𝐏f0ϕt1.f^{j+1}=\mathbf{E}_{\mathbf{P}^{\prime}}f_{0}\circ\phi_{t}^{-1}. (2.12)

See [kunita1997stochastic] for more details.

Remark 2.3.

Note that this type of regularization procedure has the added benefit of retaining non-negativity of ff as well as the preservation of the Casimir conservation laws, e.g. if ν=0\nu=0 then fjLp=f0Lp\left\|f^{j}\right\|_{L^{p}}=\left\|f_{0}\right\|_{L^{p}} and for ν>0\nu>0 one at least has fjLpedνtf0Lp\left\|f^{j}\right\|_{L^{p}}\leq e^{d\nu t}\left\|f_{0}\right\|_{L^{p}}. However, these properties do not play an important role here.

Next, we need uniform a priori estimates to enable passing jj\to\infty. These are obtained via Eulerian energy methods and come out as p<\forall p<\infty, α(0,1/2)\alpha\in(0,1/2), and T<T<\infty,

supj1𝐄fjL(0,T;Hmσ)pp,T,ϵ1\displaystyle\sup_{j\geq 1}\mathbf{E}\left\|f^{j}\right\|_{L^{\infty}(0,T;H^{\sigma^{\prime}}_{m^{\prime}})}^{p}\lesssim_{p,T,\epsilon}1
supj1𝐄fjWα,p(0,T;Hm1σ2)pp,T,ϵ,α1.\displaystyle\sup_{j\geq 1}\mathbf{E}\left\|f^{j}\right\|_{W^{\alpha,p}(0,T;H^{\sigma^{\prime}-2}_{m^{\prime}-1})}^{p}\lesssim_{p,T,\epsilon,\alpha}1.

See Lemma 3.1 for the proof of these estimates. Several previous works, for example [debussche2012global, GV14, brzezniak2020well] have used compactness to pass to similar limits, extract martingale solutions (i.e. probabilistically weak) using the Skorokhod embedding theorem, and then subsequently upgrade these solutions using a Gyöngy-Krylov lemma [gyongy1996existence] argument and pathwise uniqueness. However, this technique seems not to apply in a clear manner to the Lagrangian iteration (2.10). Instead, we prove directly that there is a stopping time ξ\xi which is almost-surely greater than 11 such that {fj}j0\left\{f^{j}\right\}_{j\geq 0} forms a Cauchy sequence in L2(Ω;Ct([0,ξ);Hm0s0))L^{2}(\Omega;C_{t}([0,\xi);H^{s_{0}}_{m_{0}})), at which point it is not hard to pass to the limit, iterate in tt, and construct global solutions to (2.7) in the desired regularity classes. This is proved in Lemma 3.6, where, in analogy with a classical Picard iteration, we show that fj+1fjf^{j+1}-f^{j} is nearly comparable in size to the jj-th term of a power series in powers of t\sqrt{t} of the solution. This procedure finally yields

Lemma 2.4.

Let f0f_{0} be an 0\mathcal{F}_{0}-measurable random variable such that p2\forall p\geq 2,

𝐄f0Hmσp<.\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}^{p}<\infty.

Then, there exists an fC([0,);Hm2σ3)Lt,loc([0,);Hmσ)f\in C([0,\infty);H_{m^{\prime}-2}^{\sigma^{\prime}-3})\cap L_{t,\text{loc}}^{\infty}([0,\infty);H_{m^{\prime}}^{\sigma^{\prime}}) 𝐏\mathbf{P}–a.s. which is a solution to (2.7) in the sense that , 𝐏\mathbf{P}–a.s.:

f(t)=f0+0t(vxf(s)θR(fHm0s0)φϵE(s)vf(s)+νf(s))ds0tvf(s)dWs,f(t)=f_{0}+\int_{0}^{t}\left(-v\cdot\nabla_{x}f(s)-\theta_{R}(\left\|f\right\|_{H^{s_{0}}_{m_{0}}})\varphi_{\epsilon}*E(s)\cdot\nabla_{v}f(s)+\nu\mathcal{L}f(s)\right)\mathrm{d}s-\int_{0}^{t}\nabla_{v}f(s)\circ\mathrm{d}W_{s}, (2.13)

where the equality holds in C([0,);Hm2σ3)C([0,\infty);H^{\sigma^{\prime}-3}_{m^{\prime}-2}). Moreover, if f~\tilde{f} is any other solution in the above sense, then f=f~f=\tilde{f} almost surely in the sense that

(f(t)f~(t)=00t<|f(0)=f~(0))=1.\displaystyle\mathbb{P}\left(f(t)-\tilde{f}(t)=0\quad\forall 0\leq t<\infty|f(0)=\tilde{f}(0)\right)=1.

The next step in the proof of Lemma 2.1 is to remove the superfluous mollifier φϵ\varphi_{\epsilon}, which begins with obtaining ϵ\epsilon-independent estimates (now indexing solutions to (2.7) by ϵ\epsilon),

supϵ(0,1)𝐄fϵL(0,T;Hmσ)pp,T,R1\displaystyle\sup_{\epsilon\in(0,1)}\mathbf{E}\left\|f_{\epsilon}\right\|_{L^{\infty}(0,T;H^{\sigma^{\prime}}_{m^{\prime}})}^{p}\lesssim_{p,T,R}1
supϵ(0,1)𝐄fϵWα,p(0,T;Hm2σ2)pp,T,R,α1.\displaystyle\sup_{\epsilon\in(0,1)}\mathbf{E}\left\|f_{\epsilon}\right\|_{W^{\alpha,p}(0,T;H^{\sigma^{\prime}-2}_{m^{\prime}-2})}^{p}\lesssim_{p,T,R,\alpha}1.

See Lemma 3.10 for the proof of these estimates. These estimates can be considered the probabilistic analogue of the common deterministic method of sharpening a continuation criterion a posteriori, specifically, the thrust of the estimates is to show that the Hm0s0H^{s_{0}}_{m_{0}} norm controls all HmσH^{\sigma^{\prime}}_{m^{\prime}} norms for m>m0m^{\prime}>m_{0} and σ>m0\sigma^{\prime}>m_{0}. At this step, it does not seem straightforward to prove that {fϵ}ϵ(0,1)\left\{f_{\epsilon}\right\}_{\epsilon\in(0,1)} is Cauchy, and so we follow the martingale approach. Specifically, we use these uniform bounds to apply the Skorokhod embedding theorem to produce probabilistically weak solutions to (2.3) (see Proposition 3.11 below). These solutions are subsequently upgraded to probabilistically strong solutions by proving pathwise uniqueness (Lemma 3.13) and an application of the Gyöngy-Krylov lemma (from [gyongy1996existence]; see Lemma 3.12 below). This general procedure is rather standard at this point; see for example [debussche2012global, GV14, brzezniak2020well]. This step completes the proof of Lemma 2.1.

The final step in the proof of Theorem 1.3 is to pass to a suitable limit in order to construct solutions in CtHmσC_{t}H^{\sigma}_{m} from HmσH^{\sigma}_{m} initial data, which is done in Section 4. We perform a regularization procedure on the initial data by defining a sequence of initial conditions

f0;n=nf0θn(v)n2dη(n)x,vf0,\displaystyle f_{0;n}=\mathcal{R}^{n}f_{0}\equiv\theta_{n}(v)n^{2d}\eta\left(\frac{\cdot}{n}\right)\ast_{x,v}f_{0},

where ηCc(2d)\eta\in C^{\infty}_{c}(\mathbb{R}^{2d}) and satisfies η0\eta\geq 0 and 2dηdxdv=1\int_{\mathbb{R}^{2d}}\eta\mathrm{d}x\mathrm{d}v=1. Note these have been both mollified and cut-off in velocity (to improve both regularity and localization). For all f0HmσL+1f_{0}\in H_{m}^{\sigma}\cap L^{1}_{+}, we hence have f0;nHmσL+1f_{0;n}\in H^{\sigma^{\prime}}_{m^{\prime}}\cap L^{1}_{+} for all σ,m\sigma^{\prime},m^{\prime}. Subsequently, there are unique local pathwise solutions to (1.1) (fn,τn)(f_{n},\tau_{n}) with

fn(τn)Lloc(0,;Hmσ)C([0,);Hmσ).\displaystyle f_{n}(\cdot\wedge\tau_{n})\in L^{\infty}_{loc}(0,\infty;H^{\sigma^{\prime}}_{m^{\prime}})\cap C([0,\infty);H^{\sigma}_{m}).

By obtaining suitable uniform-in-nn upper bounds on the CtHmσC_{t}H^{\sigma}_{m} norm, we may pass to the limit nn\to\infty and hence extract local pathwise solutions to the original problem in CtHmσC_{t}H^{\sigma}_{m}; see Lemmas 4.2 and 4.3 for details.

Notation and conventions

For technical reasons, it is sometimes necessary (particularly when passing to the limit in the proofs of Lemmas 3.11 and 2.1) to view the fluctuating field as coming from a cylindrical Wiener process. Specifically, let 𝔘\mathfrak{U} be a separable Hilbert space, with an orthonormal basis (gk)k𝕂.(g_{k})_{k\in\mathbb{K}}. We can formally define a cylindrical Wiener process 𝒲t\mathcal{W}_{t} on 𝔘\mathfrak{U} by the formula

𝒲t:=k𝕂gkWt(k).\mathcal{W}_{t}:=\sum_{k\in\mathbb{K}}g_{k}W_{t}^{(k)}.

Since this sum is divergent on 𝔘,\mathfrak{U}, one frequently employs the larger Hilbert space:

𝔘0:={kαkgk:kk2αk2<},\displaystyle\mathfrak{U}_{0}:=\left\{\sum_{k}\alpha_{k}g_{k}:\sum_{k}k^{-2}\alpha_{k}^{2}<\infty\right\},
kαkgk𝔘02=kk2αk2,\displaystyle\left\|\sum_{k}\alpha_{k}g_{k}\right\|_{\mathfrak{U}_{0}}^{2}=\sum_{k}k^{-2}\alpha_{k}^{2},

where it can be shown that the formal sum for 𝒲t\mathcal{W}_{t} converges and defines a process whose paths are almost surely in C([0,T);𝔘0)C([0,T);\mathfrak{U}_{0}). Moreover, the embedding 𝔘𝔘0\mathfrak{U}\subset\mathfrak{U}_{0} is Hilbert–Schmidt. For any separable Hilbert space XX, we denote the space of all Hilbert–Schmidt operators from 𝔘\mathfrak{U} to XX by L2(𝔘;X)L_{2}(\mathfrak{U};X); the definition of this space is:

L2(𝔘;X):={TL(𝔘;X):kTgkX2<},\displaystyle L_{2}(\mathfrak{U};X):=\left\{T\in L(\mathfrak{U};X):\sum_{k}\|Tg_{k}\|_{X}^{2}<\infty\right\},
TL2(𝔘;X)2=kTgkX2.\displaystyle\left\|T\right\|_{L_{2}(\mathfrak{U};X)}^{2}=\sum_{k}\|Tg_{k}\|_{X}^{2}.

For more details on cylindrical Wiener processes and the relevant functional analytic setting, see [da2014stochastic].

At various points, we use the notation fLpf\in L^{p-} to signify that ff is in any LqL^{q} space for q<pq<p.

We often employ the common notation:

A(f)p1,p2,B(f)A(f)\lesssim_{p_{1},p_{2},\dots}B(f)

which means that there exists a constant C>0C>0 depending only on the parameters p1,p2,p_{1},p_{2},\dots but not on the argument ff, such that A(f)CB(f)A(f)\leq CB(f) for all relevant ff. We omit the parameters if they are unimportant or clear from the context.

For the velocity-weighted L2L^{2} norms and inner products, we set:

f,gm:=𝕋d×df(x,v)g(x,v)vmdvdx\displaystyle\left\langle f,g\right\rangle_{m}:=\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}f(x,v)g(x,v)\left<v\right>^{m}\mathrm{d}v\mathrm{d}x
fLm22:=f,fm.\displaystyle\|f\|_{L_{m}^{2}}^{2}:=\left\langle f,f\right\rangle_{m}.

Finally, at various points we use the mixed weighted norms:

fLv,n2Lxp:=(d[𝕋d|f(x,v)|pdx]2/pvndv)1/2,\displaystyle\|f\|_{L_{v,n}^{2}L_{x}^{p}}:=\left(\int_{\mathbb{R}^{d}}\left[\int_{\mathbb{T}^{d}}|f(x,v)|^{p}\mathrm{d}x\right]^{2/p}\left<v\right>^{n}\mathrm{d}v\right)^{1/2},
fLv,n2Hxs:=(|α|sd𝕋d|xαf(x,v)|2vndxdv)1/2.\displaystyle\|f\|_{L_{v,n}^{2}H_{x}^{s}}:=\left(\sum_{|\alpha|\leq s}\int_{\mathbb{R}^{d}}\int_{\mathbb{T}^{d}}|\partial_{x}^{\alpha}f(x,v)|^{2}\left<v\right>^{n}\mathrm{d}x\mathrm{d}v\right)^{1/2}.

3 Very smooth solutions and pathwise uniqueness

3.1 Proof of Lemma 2.4

As discussed in Section 2. a key step in proving Lemma 2.4 is constructing a convergent sequence of approximate solutions derived from a Lagrangian iteration scheme for (2.7). In particular, consider a sequence fjf^{j} defined inductively as:

{df0+vxf0dt+vf0dWt=νf0dt,dfj+1+vxfj+1dt+vfj+1dWt=νfj+1dtθR(fjHm0s0)(φϵEj)vfj+1dt,fj(0)=f0.\begin{cases}\mathrm{d}f^{0}+v\cdot\nabla_{x}f^{0}\mathrm{d}t+\nabla_{v}f^{0}\circ\mathrm{d}W_{t}=\nu\mathcal{L}f^{0}\mathrm{d}t,\\ \mathrm{d}f^{j+1}+v\cdot\nabla_{x}f^{j+1}\mathrm{d}t+\nabla_{v}f^{j+1}\circ\mathrm{d}W_{t}=\nu\mathcal{L}f^{j+1}\mathrm{d}t-\theta_{R}(\|f^{j}\|_{H_{m_{0}}^{s_{0}}})(\varphi_{\epsilon}\ast E^{j})\cdot\nabla_{v}f^{j+1}\mathrm{d}t,\\ f^{j}(0)=f_{0}.\end{cases} (3.1)

As discussed in Section 2, for a given fjfC([0,);Hm2σ3)Lt,loc([0,),Hmσ)f^{j}\in f\in C([0,\infty);H_{m^{\prime}-2}^{\sigma^{\prime}-3})\cap L_{t,loc}^{\infty}([0,\infty),H_{m^{\prime}}^{\sigma^{\prime}}), the solution fj+1f^{j+1} is constructed via the method of stochastic characteristics.

First, we provide jj-independent estimates in order to pass to the limit j,j\to\infty, for which we need appropriate compactness estimates for the iterates fjf^{j} defined above. We remark that studying the stochastic flow of diffeomorphisms could show that fjfC([0,);Hm2σ3)Lt,loc([0,),Hmσ)f^{j}\in f\in C([0,\infty);H_{m^{\prime}-2}^{\sigma^{\prime}-3})\cap L_{t,loc}^{\infty}([0,\infty),H_{m^{\prime}}^{\sigma^{\prime}}), however, providing jj-independent (and eventually ϵ\epsilon-independent) bounds seems to be significantly more complicated than an Eulerian energy method approach, which is hence the approach we take. The main ingredient is provided by the following lemma:

Lemma 3.1.

Let (fj)j1(f^{j})_{j\geq 1} be a sequence of global solutions to the iterative scheme (3.1) with 𝐄f0Hmσp<Mp<\mathbf{E}\left\|f_{0}\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}^{p}<M_{p}<\infty for all p2.p\geq 2. For α(0,12),p2,\alpha\in(0,\frac{1}{2}),p\geq 2, we have the uniform estimates:

supj1𝐄suptTfj(t)HmσpCT,R,ϵ,M\sup_{j\geq 1}\mathbf{E}\sup_{t\leq T}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\leq C_{T,R,\epsilon,M} (3.2)

and

supj1𝐄fjWα,p([0,T];Hm1σ2)pCT,R,ϵ,M.\sup_{j\geq 1}\mathbf{E}\|f^{j}\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq C_{T,R,\epsilon,M}. (3.3)

Before we begin, let us begin by recalling a few standard estimates. The first shows how to estimate the density in terms of the distribution function using sufficiently many velocity moments.

Lemma 3.2.

For any m>dm>d there exists a constant Cm,dC_{m,d} such that

f(x,v)dvLx2CfLm2\left\|\int f(x,v)\mathrm{d}v\right\|_{L_{x}^{2}}\leq C\|f\|_{L_{m}^{2}} (3.4)

Next, we recall the following Gagliardo-Nirenberg-Sobolev estimate: for all integers 0iσ0\leq i\leq\sigma and functions fHσf\in H^{\sigma} (in 𝕋d\mathbb{T}^{d} or n\mathbb{R}^{n}) there holds

|α|=iαfL2σiCfL1iσ(|α|=σαfL2)iσ.\displaystyle\sum_{\left|\alpha\right|=i}\left\|\partial^{\alpha}f\right\|_{L^{\frac{2\sigma}{i}}}\leq C\left\|f\right\|_{L^{\infty}}^{1-\frac{i}{\sigma}}\left(\sum_{\left|\alpha\right|=\sigma}\left\|\partial^{\alpha}f\right\|_{L^{2}}\right)^{\frac{i}{\sigma}}. (3.5)

The next estimate recalls how to adapt Sobolev space product rules to the anisotropic nonlinearity. We give a proof for the readers’ convenience.

Lemma 3.3.

Let gHn0s1,fHnsg\in H_{n_{0}}^{s-1},f\in H_{n}^{s} for some n0>dn_{0}>d, s>d2+1s>\frac{d}{2}+1, and n0n\geq 0 arbitrary. Denote Eg:=xΔx1(gdv1).E^{g}:=\nabla_{x}\Delta_{x}^{-1}\left(\int g\mathrm{d}v-1\right). Then, for a constant CC that does not depend on gg or ff:

EgvfHnsCgHn0s1vfHns\|E^{g}\cdot\nabla_{v}f\|_{H_{n}^{s}}\leq C\|g\|_{H_{n_{0}}^{s-1}}\|\nabla_{v}f\|_{H_{n}^{s}} (3.6)

and

|α|+|β|sxαvβ(Egvf),xαvβfnC(EgW1,+gHn0s1)fHns2.\sum_{|\alpha|+|\beta|\leq s}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{g}\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\leq C(\|E^{g}\|_{W^{1,\infty}}+\|g\|_{H_{n_{0}}^{s-1}})\|f\|_{H_{n}^{s}}^{2}. (3.7)
Proof of Lemma 3.3.

In what follows, η(0,12)\eta\in(0,\frac{1}{2}) will be fixed, which implies Hxd2+ηLx.H_{x}^{\frac{d}{2}+\eta}\subset L_{x}^{\infty}. Denote the multivariate binomial coefficients by

(αα):=j=1dαj!j=1d(αj)!j=1d(αjαj)!.\begin{pmatrix}\alpha\\ \alpha^{\prime}\end{pmatrix}:=\frac{\prod_{j=1}^{d}\alpha_{j}!}{\prod_{j=1}^{d}(\alpha_{j}^{\prime})!\prod_{j=1}^{d}(\alpha_{j}-\alpha_{j}^{\prime})!}.

We begin the proof of (3.6) by using the product rule and the triangle inequality:

EgvfHns|α|+|β|sα1α(αα1)xαα1Egvxα1vβfLn2.\|E^{g}\cdot\nabla_{v}f\|_{H_{n}^{s}}\leq\sum_{\begin{subarray}{c}|\alpha|+|\beta|\leq s\\ \alpha_{1}\leq\alpha\end{subarray}}\begin{pmatrix}\alpha\\ \alpha_{1}\end{pmatrix}\left\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\right\|_{L_{n}^{2}}. (3.8)

Now, we split into four separate cases:

Case 1 α1=α\alpha_{1}=\alpha:

In this case, we use Hölder’s inequality, the embedding Hxd/2+ηLxH^{d/2+\eta}_{x}\subset L^{\infty}_{x}, and Lemma 3.2

EgvxαvβfLn2\displaystyle\|E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{n}^{2}}\leq EgLvxαvβfLn2\displaystyle\|E^{g}\|_{L^{\infty}}\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{n}^{2}}
\displaystyle\leq CgHn0d/21+ηvfHns.\displaystyle C\|g\|_{H_{n_{0}}^{d/2-1+\eta}}\|\nabla_{v}f\|_{H_{n}^{s}}. (3.9)

In the following cases, α1<α.\alpha_{1}<\alpha.

Case 2 |α1|+|β|=|α|+|β|1|\alpha_{1}|+|\beta|=|\alpha|+|\beta|-1 or |α|+|β|=1|\alpha|+|\beta|=1:

Here, we have |αα1|=1|\alpha-\alpha_{1}|=1, and hence similarly to the previous case

xαα1Egvxα1vβfLn2\displaystyle\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\|_{L_{n}^{2}}\leq xEgLvfHns\displaystyle\|\nabla_{x}E^{g}\|_{L^{\infty}}\|\nabla_{v}f\|_{H_{n}^{s}}
\displaystyle\leq CgHn0d2+ηvfHns.\displaystyle C\|g\|_{H_{n_{0}}^{\frac{d}{2}+\eta}}\|\nabla_{v}f\|_{H^{s}_{n}}. (3.10)
Case 3 |αα1|=|α|+|β|2|\alpha-\alpha_{1}|=|\alpha|+|\beta|\geq 2:

Here, we necessarily have α1=β=0,\alpha_{1}=\beta=0, and hence (using the Sobolev embedding now on vf\nabla_{v}f),

xαEgvfLn2\displaystyle\|\partial_{x}^{\alpha}E^{g}\cdot\nabla_{v}f\|_{L_{n}^{2}}\leq xαEgL2vfLxLv,m2\displaystyle\|\partial_{x}^{\alpha}E^{g}\|_{L^{2}}\left\|\left\|\nabla_{v}f\right\|_{L_{x}^{\infty}}\right\|_{L_{v,m^{\prime}}^{2}}
\displaystyle\leq CgHn0s1vfHnd/2+η.\displaystyle C\|g\|_{H_{n_{0}}^{s-1}}\|\nabla_{v}f\|_{H_{n}^{d/2+\eta}}. (3.11)
Case 4 |α1|+|β||α|+|β|2|\alpha_{1}|+|\beta|\leq|\alpha|+|\beta|-2:

Here |αα1|2,|\alpha-\alpha_{1}|\geq 2, so:

xαα1Egvxα1vβfLn2\displaystyle\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\|_{L_{n}^{2}}\leq xαα1EgLx2vxα1vβfLxLv,n2\displaystyle\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\|_{L^{2}_{x}}\left\|\left\|\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\right\|_{L_{x}^{\infty}}\right\|_{L_{v,n}^{2}}
\displaystyle\leq CgHn0|αα1|1vxα1vβfHnd/2+η\displaystyle C\|g\|_{H_{n_{0}}^{|\alpha-\alpha_{1}|-1}}\|\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\|_{H_{n}^{d/2+\eta}}
\displaystyle\leq CgHn0s1vfHns.\displaystyle C\|g\|_{H_{n_{0}}^{s-1}}\|\nabla_{v}f\|_{H_{n}^{s}}. (3.12)

Summing over the various cases, (3.6) follows.

The proof of (3.7) is similar but just slightly more subtle after using the cancellation that occurs when all of the derivatives land on vf\nabla_{v}f. First, distribute the derivatives with Leibniz’s rule.

|α|+|β|sxαvβ(Egvf),xαvβfn|α|+|β|sα1α(αα1)|xαα1Egvxα1vβf,xαvβfn|.\displaystyle\sum_{|\alpha|+|\beta|\leq s}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{g}\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\leq\sum_{|\alpha|+|\beta|\leq s}\sum_{\alpha_{1}\leq\alpha}\begin{pmatrix}\alpha\\ \alpha_{1}\end{pmatrix}\left|\left<\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\right|. (3.13)

We distinguish the following cases:

Case 1 α1=α\alpha_{1}=\alpha:

by integrating by parts the v\nabla_{v} onto the weight, we have

|Egvxαvβf,xαvβfn|=\displaystyle\left|\left<E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\right|= 12𝕋d×dEgv|xαvβf|2(vm)dvdx\displaystyle\frac{1}{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}E^{g}\cdot\nabla_{v}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f|^{2}(\left<v\right>^{m})\mathrm{d}v\mathrm{d}x
\displaystyle\leq CEgLxαvβfLn22.\displaystyle C\|E^{g}\|_{L^{\infty}}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{n}^{2}}^{2}. (3.14)
Case 2 |αα1|=1|\alpha-\alpha_{1}|=1:

Cauchy-Schwarz gives

|xαα1Egvxα1vβf,xαvβfn|\displaystyle\left|\left<\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\right|\leq xαα1EgLfHns2\displaystyle\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\|_{L^{\infty}}\|f\|_{H_{n}^{s}}^{2}
\displaystyle\leq CxEgLfHns2.\displaystyle C\|\nabla_{x}E^{g}\|_{L^{\infty}}\|f\|_{H_{n}^{s}}^{2}. (3.15)
Case 3 |αα1|2|\alpha-\alpha_{1}|\geq 2:

We have:

|xαα1Egvxα1vβf,xαvβfn|\displaystyle\left|\left<\partial_{x}^{\alpha-\alpha_{1}}E^{g}\cdot\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{n}\right|
\displaystyle\leq xαα1EgL2s1|αα1|1vxα1vβfLv,n2Lx2s1|α1|+|β|xαvβfLn2\displaystyle\|\partial_{x}^{\alpha-\alpha_{1}}E^{g}\|_{L^{2\frac{s-1}{|\alpha-\alpha_{1}|-1}}}\left\|\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f\right\|_{L_{v,n}^{2}L_{x}^{2\frac{s-1}{|\alpha_{1}|+|\beta|}}}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{n}^{2}}
\displaystyle\leq CxEgL|α1|+|β|s1xEgHxs|αα1|1s1fHns2.\displaystyle C\|\nabla_{x}E^{g}\|_{L^{\infty}}^{\frac{|\alpha_{1}|+|\beta|}{s-1}}\|\nabla_{x}E^{g}\|_{H_{x}^{s}}^{\frac{|\alpha-\alpha_{1}|-1}{s-1}}\|f\|_{H_{n}^{s}}^{2}. (3.16)

In the above we used Hölder’s inequality, Gagliardo-Nirenberg interpolation (3.5) on xEg\nabla_{x}E^{g} and Sobolev embedding on the vxα1vβf\nabla_{v}\partial_{x}^{\alpha_{1}}\partial_{v}^{\beta}f term, where we note that the order of integrability p:=2s1|α1|+|β|p:=2\frac{s-1}{|\alpha_{1}|+|\beta|} corresponds to the embedding of Hxσ~H_{x}^{\tilde{\sigma}} in LxpL_{x}^{p} for σ~\tilde{\sigma} given by:

σ~=d2|αα1|1s1\tilde{\sigma}=\frac{d}{2}\frac{|\alpha-\alpha_{1}|-1}{s-1} (3.17)

which satisfies σ~<|αα1|1\tilde{\sigma}<|\alpha-\alpha_{1}|-1 exactly if s>d2+1s>\frac{d}{2}+1.

Summing over the above cases, we obtain (3.7), which completes the proof of Lemma 3.3. ∎

Next, we prove Lemma 3.1.

Proof of Lemma 3.1.

We begin with an estimate for fHmσ2\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}, which we then use to derive an estimate for fHmσp\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}, for p>2.p>2. Applying Itô’s formula to xαvβfjLm22\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\|_{L_{m^{\prime}}^{2}}^{2}, we have:

dxαvβfjLm22=\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\|_{L_{m^{\prime}}^{2}}^{2}= 2xαvβ(vxfj),xαvβfjmdt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t
+2Δvxαvβfj,xαvβfjmdt\displaystyle+2\left<\Delta_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t
+2xαvβ(divv(fjv)),xαvβfjmdt\displaystyle+2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\operatorname{\mathrm{div}}_{v}(f^{j}v)),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t
2θRxαvβ(φϵEj1vfj),xαvβfjmdt\displaystyle-2\left<\theta_{R}\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*E^{j-1}\cdot\nabla_{v}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t
2xαvβ(vfjdWt),xαvβfjm\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f^{j}\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
+kxαvβ[(σkekv)2fj],xαvβfjmdt\displaystyle+\sum_{k}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}[(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f^{j}],\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t
+kxαvβ(σkekvfj)Lm22dt\displaystyle+\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f^{j})\|_{L_{m^{\prime}}^{2}}^{2}\mathrm{d}t
=\displaystyle= :𝒯α,β(fj)+𝒟α,β(fj)+α,β(fj)+𝒩α,β(fj)+α,β(fj)+𝒞α,β(fj),\displaystyle:\mathcal{T}_{\alpha,\beta}(f^{j})+\mathcal{D}_{\alpha,\beta}(f^{j})+\mathcal{F}_{\alpha,\beta}(f^{j})+\mathcal{N}_{\alpha,\beta}(f^{j})+\mathcal{M}_{\alpha,\beta}(f^{j})+\mathcal{C}_{\alpha,\beta}(f^{j}), (3.18)

where

𝒞α,β(fj)=kxαvβ[(σkekv)2fj],xαvβfjmdt+kxαvβ(σkekvfj)Lm22dt.\displaystyle\mathcal{C}_{\alpha,\beta}(f^{j})=\sum_{k}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}[(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f^{j}],\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t+\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f^{j})\|_{L_{m^{\prime}}^{2}}^{2}\mathrm{d}t.

Here, the 𝒯,𝒟,,𝒩,,𝒞\mathcal{T},\mathcal{D},\mathcal{F},\mathcal{N},\mathcal{M},\mathcal{C} terms abbreviate transport, dissipation, friction, nonlinear electric field, martingale, and correction contributions, respectively. We begin by observing that by integration by parts,

𝒯α,β(fj)+α,β(fj)CfjHm|α|+|β|2dt.\mathcal{T}_{\alpha,\beta}(f^{j})+\mathcal{F}_{\alpha,\beta}(f^{j})\leq C\|f^{j}\|_{H_{m^{\prime}}^{|\alpha|+|\beta|}}^{2}\mathrm{d}t. (3.19)

Similarly, for the dissipative term, integrating by parts gives:

𝒟α,β(fj)=\displaystyle\mathcal{D}_{\alpha,\beta}(f^{j})= 2vxαvβfjLm22dt+𝕋d×d|xαvβfj|2Δvvmdvdxdt\displaystyle-2\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\|_{L_{m^{\prime}}^{2}}^{2}\mathrm{d}t+\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}|^{2}\Delta_{v}\left<v\right>^{m}\mathrm{d}v\mathrm{d}x\mathrm{d}t
\displaystyle\leq 2vxαvβfjLm22dt+CfHm|α|+|β|2dt\displaystyle-2\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\|_{L_{m^{\prime}}^{2}}^{2}\mathrm{d}t+C\|f\|_{H_{m^{\prime}}^{|\alpha|+|\beta|}}^{2}\mathrm{d}t (3.20)

Next, turn to the Itô correction terms, which need to be treated carefully in order to not lose derivatives. Distributing derivatives gives

𝒞α,β(fj)=\displaystyle\mathcal{C}_{\alpha,\beta}(f^{j})= k(xαvβ(σkekv)fjLm22+xαvβ(σkekv)2fj,xαvβfjm)dt\displaystyle\sum_{k}\left(\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})f^{j}\|_{L_{m^{\prime}}^{2}}^{2}+\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\right)\mathrm{d}t
=\displaystyle= kα<α(αα)σk2xαα(ekek):v2xαvβfj,xαvβfjmdt\displaystyle\sum_{k}\sum_{\alpha^{\prime}<\alpha}\begin{pmatrix}\alpha\\ \alpha^{\prime}\end{pmatrix}\sigma_{k}^{2}\left<\partial_{x}^{\alpha-\alpha^{\prime}}(e_{k}\otimes e_{k}):\nabla_{v}^{2}\partial_{x}^{\alpha^{\prime}}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t (3.21)
+kσk2(ekek):v2xαvβfj,xαvβfjmdt\displaystyle+\sum_{k}\sigma_{k}^{2}\left<(e_{k}\otimes e_{k}):\nabla_{v}^{2}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\mathrm{d}t (3.22)
+kxαvβ(σkekv)fjLm22dt.\displaystyle+\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})f^{j}\|_{L_{m^{\prime}}^{2}}^{2}\mathrm{d}t. (3.23)

Now, (3.22) provides a term of highest order that cancels the Itô correction xαvβ(σkekv)fjLm22\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})f^{j}\|_{L_{m^{\prime}}^{2}}^{2}, and terms of lower order that can either be readily controlled by fjHmσ2\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2} or cancel out with a corresponding term in (3.21). Integrating by parts in (3.22) we have,

(3.22)=σk2(ekek):v2xαvβfj,xαvβfjm\displaystyle\eqref{c:2:it}=\sigma_{k}^{2}\left<(e_{k}\otimes e_{k}):\nabla_{v}^{2}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
=\displaystyle= σkekvxαvβfj,σkekvxαvβfjm\displaystyle-\left<\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j},\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
𝕋d×dxαvβfj(σkekvxαvβfj)σkekv(vm)dvdx\displaystyle-\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}(\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j})\sigma_{k}e_{k}\cdot\nabla_{v}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x
=\displaystyle= xαvβ(σkekvfj),σkekvxαvβfjm[σkekv,xαvβ]fj,σkekvxαvβfjm\displaystyle-\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f^{j}),\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}-\left<[\sigma_{k}e_{k}\cdot\nabla_{v},\partial_{x}^{\alpha}\partial_{v}^{\beta}]f^{j},\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
+12𝕋d×d|xαvβfj|2(σkekv)2(vm)dvdx\displaystyle+\frac{1}{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}|^{2}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x
=\displaystyle= σkekv(xαvβfj)Lm22\displaystyle-\|\sigma_{k}e_{k}\cdot\nabla_{v}(\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j})\|_{L_{m^{\prime}}^{2}}^{2}
xαvβ(σkekvfj),[σkekv,xαvβ]fjm[σkekv,xαvβ]fj,σkekvxαvβfjm\displaystyle-\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f^{j}),[\sigma_{k}e_{k}\cdot\nabla_{v},\partial_{x}^{\alpha}\partial_{v}^{\beta}]f^{j}\right>_{m^{\prime}}-\left<[\sigma_{k}e_{k}\cdot\nabla_{v},\partial_{x}^{\alpha}\partial_{v}^{\beta}]f^{j},\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
+12𝕋d×d|xαvβfj|2(σkekv)2(vm)dvdx\displaystyle+\frac{1}{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}|^{2}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x
=\displaystyle= σkekv(xαvβfj)Lm22\displaystyle-\|\sigma_{k}e_{k}\cdot\nabla_{v}(\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j})\|_{L_{m^{\prime}}^{2}}^{2} (3.24)
2[σkekv,xαvβ]fj,σkekvxαvβfjm\displaystyle-2\left<[\sigma_{k}e_{k}\cdot\nabla_{v},\partial_{x}^{\alpha}\partial_{v}^{\beta}]f^{j},\sigma_{k}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}} (3.25)
[xαvβ,σkekv]fj,[σkekv,xαvβ]fjm\displaystyle-\left<[\partial_{x}^{\alpha}\partial_{v}^{\beta},\sigma_{k}e_{k}\cdot\nabla_{v}]f^{j},[\sigma_{k}e_{k}\cdot\nabla_{v},\partial_{x}^{\alpha}\partial_{v}^{\beta}]f^{j}\right>_{m^{\prime}} (3.26)
+12𝕋d×d|xαvβfj|2(σkekv)2(vm)dvdx.\displaystyle+\frac{1}{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}|^{2}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x. (3.27)

Thus, we observe that (3.24) cancels the Itô correction (3.23), (3.27) is bounded above by CxαvβfjLm22C\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\|_{L_{m^{\prime}}^{2}}^{2}, and (3.26) only contains derivatives of order lower than |α|+|β||\alpha|+|\beta| and is thus bounded above by CfjHmσ2C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}. Next, we turn to (3.25). This term contains (A) terms from the commutator where the total number of derivatives on fjf^{j} is strictly less than |α|+|β||\alpha|+|\beta|, which can be treated by integration by parts of the σkekv\sigma_{k}e_{k}\cdot\nabla_{v} and are thus bounded above by CfjHmσ2C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}; and (B) a highest order term which we deal with as follows:

2α<α|α|=1σk2xαekvxαα1vβfj,ekvxαvβfjm\displaystyle 2\sum_{\begin{subarray}{c}\alpha^{\prime}<\alpha\\ |\alpha^{\prime}|=1\end{subarray}}\sigma_{k}^{2}\left<\partial_{x}^{\alpha^{\prime}}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha-\alpha_{1}}\partial_{v}^{\beta}f^{j},e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}
=\displaystyle= 2α<α|α|=1σk2(xαekv)(ekv)xααvβfj,xαvβfjm\displaystyle-2\sum_{\begin{subarray}{c}\alpha^{\prime}<\alpha\\ |\alpha^{\prime}|=1\end{subarray}}\sigma_{k}^{2}\left<(\partial_{x}^{\alpha^{\prime}}e_{k}\cdot\nabla_{v})(e_{k}\cdot\nabla_{v})\partial_{x}^{\alpha-\alpha^{\prime}}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m}
α<α|α|=1σk2𝕋d×dxαvβfjxαekvxααvβfjekv(vm)dvdx\displaystyle-\sum_{\begin{subarray}{c}\alpha^{\prime}<\alpha\\ |\alpha|=1\end{subarray}}\sigma_{k}^{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\partial_{x}^{\alpha^{\prime}}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha-\alpha^{\prime}}\partial_{v}^{\beta}f^{j}e_{k}\cdot\nabla_{v}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x
=\displaystyle= α<α|α|=1σk2xα(ekek):v2xααvβfj,xαvβfjm\displaystyle-\sum_{\begin{subarray}{c}\alpha^{\prime}<\alpha\\ |\alpha^{\prime}|=1\end{subarray}}\sigma_{k}^{2}\left<\partial_{x}^{\alpha^{\prime}}(e_{k}\otimes e_{k}):\nabla_{v}^{2}\partial_{x}^{\alpha-\alpha^{\prime}}\partial_{v}^{\beta}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}} (3.28)
α<α|α|=1σk2𝕋d×dxαvβfjxαekvxααvβfjekv(vm)dvdx.\displaystyle-\sum_{\begin{subarray}{c}\alpha^{\prime}<\alpha\\ |\alpha|=1\end{subarray}}\sigma_{k}^{2}\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\partial_{x}^{\alpha^{\prime}}e_{k}\cdot\nabla_{v}\partial_{x}^{\alpha-\alpha^{\prime}}\partial_{v}^{\beta}f^{j}e_{k}\cdot\nabla_{v}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x. (3.29)

Notice that to obtain the prefactor 11 in (3.28) we used the symmetry of the tensor ekeke_{k}\otimes e_{k}. Now, (3.29) can be directly bounded by CfjHmσ2C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}, while (3.28) cancels out the highest order term in (3.21). Therefore, we finally conclude using (3.21)–(3.29) that we have

𝒞α,β(fj)CfjHmσ2dt.\mathcal{C}_{\alpha,\beta}(f^{j})\leq C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t. (3.30)

Next we treat the contribution of the electric field term. It follows from (3.9)-(3.12) in the proof of (3.7) that:

𝒩α,β(fj)\displaystyle\mathcal{N}_{\alpha,\beta}(f^{j})\leq CθR(fj1Hm0s0)(φϵEj1W1,+φϵEj1Hxσ1)fjHmσ2dt\displaystyle C\theta_{R}(\|f^{j-1}\|_{H_{m_{0}}^{s_{0}}})(\|\varphi_{\epsilon}*E^{j-1}\|_{W^{1,\infty}}+\|\varphi_{\epsilon}*E^{j-1}\|_{H_{x}^{\sigma^{\prime}-1}})\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t
ϵ,R\displaystyle\lesssim_{\epsilon,R} fjHmσ2dt.\displaystyle\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t. (3.31)

Finally, the martingale contribution is given by

α,β(fj)=\displaystyle\mathcal{M}_{\alpha,\beta}(f^{j})= 𝕋d×d|xαvβfj|2dWtv(vm)dvdx2[xαvβ,dWt]vfj,xαvβfjm.\displaystyle\iint_{\mathbb{T}^{d}\times\mathbb{R}^{d}}|\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}|^{2}\mathrm{d}W_{t}\cdot\nabla_{v}(\left<v\right>^{m^{\prime}})\mathrm{d}v\mathrm{d}x-2\left<[\partial_{x}^{\alpha}\partial_{v}^{\beta},\mathrm{d}W_{t}]\nabla_{v}f^{j},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}. (3.32)

We sum (3.19)-(3.32) over |α|+|β|σ|\alpha|+|\beta|\leq\sigma^{\prime} and obtain

dfjHmσ2\displaystyle\mathrm{d}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\leq CfjHmσ2dt2νvfjHmσ2dt+0|α|+|β|σα,β(fj),\displaystyle C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t-2\nu\|\nabla_{v}f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t+\sum_{0\leq|\alpha|+|\beta|\leq\sigma}\mathcal{M}_{\alpha,\beta}(f^{j}), (3.33)

so integrating in time and using the Burkhölder–Davis–Gundy inequality (see e.g. [da1996ergodicity]) (hereinafter abbreviated as BDG) we obtain:

𝐄suptTfj(t)Hmσ2\displaystyle\mathbf{E}\sup_{t\leq T}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\leq 𝐄f0Hmσ2+C0T𝐄fj(t)Hmσ2dt\displaystyle\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}+C\int_{0}^{T}\mathbf{E}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t
+C𝐄(0Tfj(t)Hmσ4dt)12\displaystyle+C\mathbf{E}\left(\int_{0}^{T}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{4}\mathrm{d}t\right)^{\frac{1}{2}}
\displaystyle\leq f0Hmσ2+C0T𝐄fj(t)Hmσ2dt\displaystyle\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}+C\int_{0}^{T}\mathbf{E}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\mathrm{d}t
+12𝐄suptTfj(t)Hmσ2,\displaystyle+\frac{1}{2}\mathbf{E}\sup_{t\leq T}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}, (3.34)

where the second line followed from Hölder’s inequality. After rearranging and applying Grönwall’s inequality we obtain the uniform-in-jj estimate:

𝐄suptTfj(t)Hmσ2C𝐄f0Hmσ2,\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\leq C\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}, (3.35)

where the constant CC depends on ϵ,R,T,m,σ\epsilon,R,T,m^{\prime},\sigma^{\prime}, but not f0f_{0} or jj. Thus we have obtained (3.2) for p=2p=2.

Now, we use Itô’s formula again, this time for fjHmσp\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}, with p>2p>2:

dfjHmσp=\displaystyle\mathrm{d}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}= p2fjHmσp2dfHmσ2\displaystyle\frac{p}{2}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p-2}\mathrm{d}\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}
+p2p22fjHmσp4k|α|+|β|σ|xαvβ(σkekvfj),xαvβfjm|2dt.\displaystyle+\frac{p}{2}\frac{p-2}{2}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p-4}\sum_{k}\sum_{|\alpha|+|\beta|\leq\sigma^{\prime}}\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}\right|^{2}\mathrm{d}t.

The latter term is treated by a straightforward commutator estimate, and together with the above estimates on dfjHmσ2\mathrm{d}\left\|f^{j}\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}^{2}, we obtain

dfjHmσp\displaystyle\mathrm{d}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\leq CfjHmσpdtpfjHmσp2|α|+|β|σxαvβ(vfjdWt),xαvβfjm.\displaystyle C\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\mathrm{d}t-p\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p-2}\sum_{|\alpha|+|\beta|\leq\sigma^{\prime}}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f^{j}\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{j}\right>_{m^{\prime}}.

After integrating in time, using the BDG inequality, and applying Hölder’s inequality, we have

𝐄suptTfjHmσp\displaystyle\mathbf{E}\sup_{t\leq T}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\leq 𝐄f0jHmσp+C0T𝐄fjHmσpds\displaystyle\mathbf{E}\|f_{0}^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}+C\int_{0}^{T}\mathbf{E}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\mathrm{d}s
+C𝐄(0TfjHmσ2pds)12\displaystyle+C\mathbf{E}\left(\int_{0}^{T}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2p}\mathrm{d}s\right)^{\frac{1}{2}}
\displaystyle\leq 𝐄f0Hmσp+C0T𝐄fjHmσpds+12𝐄suptTfjHmσp.\displaystyle\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}+C\int_{0}^{T}\mathbf{E}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\mathrm{d}s+\frac{1}{2}\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}. (3.36)

By rearranging and using Grönwall’s lemma, we obtain (3.2).

We now turn to the proof of (3.3). We have:

𝐄fjWα,p([0,T];Hm1σ2)p\displaystyle\mathbf{E}\left\|f^{j}\right\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq Cf0Hm1σ2p\displaystyle C\|f_{0}\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-2}}^{p}
+C𝐄0t(vxfj+Δvfj+divv(fjv))dsW1,p([0,T];Hm1σ2)p\displaystyle+C\mathbf{E}\left\|\int_{0}^{t}\left(-v\cdot\nabla_{x}f^{j}+\Delta_{v}f^{j}+\operatorname{\mathrm{div}}_{v}(f^{j}v)\right)\mathrm{d}s\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}
+C𝐄0tθRφϵEj1vfjW1,p([0,T];Hm1σ2)p\displaystyle+C\mathbf{E}\left\|\int_{0}^{t}\theta_{R}\varphi_{\epsilon}*E^{j-1}\cdot\nabla_{v}f^{j}\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}
+C𝐄120tk(σkekv)2fjdsW1,p([0,T];Hm1σ2)p\displaystyle+C\mathbf{E}\left\|\frac{1}{2}\int_{0}^{t}\sum_{k}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f^{j}\mathrm{d}s\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}
+C𝐄0tvfjdWtWα,p([0,T];Hm1σ2)p.\displaystyle+C\mathbf{E}\left\|\int_{0}^{t}\nabla_{v}f^{j}\cdot\mathrm{d}W_{t}\right\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}.

The terms that are regular in time are estimated in a straightforward manner using the available regularity:

𝐄0t(vxfj+divv(fjv))dsW1,p([0,T];Hm1σ2)pC𝐄suptTfj(t)Hmσ1p\displaystyle\mathbf{E}\left\|\int_{0}^{t}\left(v\cdot\nabla_{x}f^{j}+\operatorname{\mathrm{div}}_{v}(f^{j}v)\right)\mathrm{d}s\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq C\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t^{\prime})\|_{H_{m^{\prime}}^{\sigma^{\prime}-1}}^{p}
𝐄0tΔvfjdsW1,p([0,T];Hm1σ2)pC𝐄suptTfj(t)Hm1σp\displaystyle\mathbf{E}\left\|\int_{0}^{t}\Delta_{v}f^{j}\mathrm{d}s\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq C\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t^{\prime})\|_{H_{m^{\prime}-1}^{\sigma^{\prime}}}^{p}
𝐄0tk(σkekv)2fjW1,p([0,T];Hm1σ2)pC𝐄suptTfj(t)Hm1σp\displaystyle\mathbf{E}\left\|\int_{0}^{t}\sum_{k}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f^{j}\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq C\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t^{\prime})\|_{H_{m^{\prime}-1}^{\sigma^{\prime}}}^{p}
𝐄0tθRφϵEj1vfjdsW1,p([0,T];Hm1σ2)pCR𝐄suptTfj(t)Hm1σ1p.\displaystyle\mathbf{E}\left\|\int_{0}^{t}\theta_{R}\varphi_{\epsilon}*E^{j-1}\cdot\nabla_{v}f^{j}\mathrm{d}s\right\|_{W^{1,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p}\leq C_{R}\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t^{\prime})\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{p}.

The time-regularity is only limited by the stochastic integral, which is estimated by a variant of the BDG inequality adapted to fractional regularity estimates in time (see e.g. [Lemma 2.1; [flandoli1995martingale]] for a proof), namely

𝐄0tvfjdWtWα,p([0,T];Hm1σ2)p\displaystyle\mathbf{E}\left\|\int_{0}^{t}\nabla_{v}f^{j}\cdot\mathrm{d}W_{t}\right\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-1}^{\sigma^{\prime}-2})}^{p} C𝐄0Tvfj(s)Hm1σ2p𝑑s\displaystyle\leq C\mathbf{E}\int_{0}^{T}\left\|\nabla_{v}f^{j}(s)\right\|_{H^{\sigma^{\prime}-2}_{m^{\prime}-1}}^{p}ds
C𝐄suptTfj(t)Hm1σ1p.\displaystyle\leq C\mathbf{E}\sup_{t^{\prime}\leq T}\|f^{j}(t^{\prime})\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{p}.

Therefore, using that W1,p([0,T];Hm2σ2)Wα,p([0,T];Hm2σ2)W^{1,p}([0,T];H_{m^{\prime}-2}^{\sigma^{\prime}-2})\subset W^{\alpha,p}([0,T];H_{m^{\prime}-2}^{\sigma^{\prime}-2}) continuously and (3.2), we obtain:

𝐄fjWα,p([0,T];Hm2σ2)pCR,T𝐄f0Hmσp\mathbf{E}\|f^{j}\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-2}^{\sigma^{\prime}-2})}^{p}\leq C_{R,T}\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p} (3.37)

uniformly in jj, which implies (3.3), completing the proof of Lemma 3.1. ∎

Remark 3.4.

By examining the proof above, one can see that one can also treat magnetic fields, due to the special structure of the Lorentz force v×B(x)v\times B(x), which ensures both v(v×B)=0\nabla_{v}\cdot(v\times B)=0 and, despite the power of vv, the estimates do not lose any moments in vv as v×Bv\times B is orthogonal to vv (nor does the vv dependence create any issues controlling higher regularity).

We continue the proof of Lemma (2.4). The approximation procedure mixes fjf^{j} and fj+1f^{j+1} in a way that makes it difficult to apply the usual method of using tightness of the laws in pathspace and applying the Skorohod embedding theorem to construct probabilistically weak solutions which are subsequently upgraded to strong solutions (see e.g. [debussche2011local, debussche2012global, GV14, brzezniak2020well]). Instead we will prove that {fj}j=1\left\{f^{j}\right\}_{j=1}^{\infty} is Cauchy in a suitable topology. For this we first need the following consequence of Lemma 3.1 and the Borel-Cantelli lemma.

Lemma 3.5.

For all δ>0\delta>0, \exists a 1\mathcal{F}_{1}-measurable, almost-surely finite, random constant C0C_{0} such that for all j0j\geq 0 there holds

sups<1fj(s)Hmσ<C0(ω,δ)jδ.\displaystyle\sup_{s<1}\left\|f^{j}(s)\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}<C_{0}(\omega,\delta)\left\langle j\right\rangle^{\delta}.

Moreover, α,n\forall\alpha,n there holds,

𝐏(C0>n)δ,αnα.\displaystyle\mathbf{P}(C_{0}>n)\lesssim_{\delta,\alpha}n^{-\alpha}.
Proof.

Recall the uniform in jj bound (3.2) for the iterates for T=1T=1:

supj1𝐄sups1fj(s)HmσpCp,R,ϵ,M<.\sup_{j\geq 1}\mathbf{E}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\leq C_{p,R,\epsilon,M}<\infty.

This estimate implies:

𝐄j=0sups1fj(s)HmσpjδpCp,R,ϵ,M,\mathbf{E}\sum_{j=0}^{\infty}\frac{\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}}{\left\langle j\right\rangle^{\delta p}}\leq C_{p,R,\epsilon,M},

for p>1δp>\frac{1}{\delta}. Denote by AjA_{j} the sets:

Aj:={ωΩ:sups1fj(s)Hmσ>jδ},A_{j}:=\{\omega\in\Omega:\,\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}>\left\langle j\right\rangle^{\delta}\},

and note that by Chebyshev’s inequality:

j=0𝐏(Aj)j=0𝐄sups1fj(s)Hmσpjδp<.\sum_{j=0}^{\infty}\mathbf{P}(A_{j})\leq\sum_{j=0}^{\infty}\frac{\mathbf{E}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}}{\left\langle j\right\rangle^{\delta p}}<\infty. (3.38)

It then follows by the Borel–Cantelli lemma that

𝐏(lim supjAj)=0,\mathbf{P}(\limsup_{j\to\infty}A_{j})=0,

implying that 𝐏\mathbf{P}-a.s., sups1fj(s)Hmσ>jδ\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}>\left\langle j\right\rangle^{\delta} at most for a finite number of jj’s. Denote the largest such jj by j0(ω)j_{0}(\omega). We then see that there is a random constant C0(ω,δ)C_{0}(\omega,\delta) such that

supj0(jδsups1fj(s)Hmσ)<C0(ω,δ)\sup_{j\geq 0}\left(\left\langle j\right\rangle^{-\delta}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}\right)<C_{0}(\omega,\delta)

𝐏\mathbf{P}–almost surely. In particular, we can take:

C0(ω,δ):=inf{n:supjj0(ω)(jδsups1fj(s)Hmσ)<n}.C_{0}(\omega,\delta):=\inf\left\{n\in\mathbb{N}:\,\sup_{j\leq j_{0}(\omega)}\left(\left\langle j\right\rangle^{-\delta}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}\right)<n\right\}.

To bound the probability that C0C_{0} is large, we observe:

𝐏(C0>n)\displaystyle\mathbf{P}(C_{0}>n)\leq 𝐏(supj0(jδsups1fj(s)Hmσ>n))\displaystyle\mathbf{P}\left(\sup_{j\geq 0}\left(\left\langle j\right\rangle^{-\delta}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}>n\right)\right)
\displaystyle\leq j=0jδp𝐄sups1fj(s)Hmσpnp\displaystyle\sum_{j=0}^{\infty}\left\langle j\right\rangle^{-\delta p}\mathbf{E}\sup_{s\leq 1}\|f^{j}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}n^{-p}
\displaystyle\lesssim np.\displaystyle n^{-p}.

This completes the proof of the lemma. ∎

The next lemma is the crucial convergence estimate.

Lemma 3.6.

There exists an increasing sequence of stopping times τn\tau_{n} such that {fj}j=1\left\{f^{j}\right\}_{j=1}^{\infty} is Cauchy in Lω2C([0,τn];Hm0s0)L^{2}_{\omega}C([0,\tau_{n}];H^{s_{0}}_{m_{0}}) and the stopping time

limnτn=ξ,\displaystyle\lim_{n\to\infty}\tau_{n}=\xi,

is almost-surely greater than 11.

Proof.

Define the increasing sequence of stopping times

τn=inf{t:j:fj(t)Hm0s0+1>njδ}.\displaystyle\tau_{n}=\inf\left\{t:\exists j:\left\|f^{j}(t)\right\|_{H^{s_{0}+1}_{m_{0}}}>n\left\langle j\right\rangle^{\delta}\right\}.

Note that by Lemma 3.5 there holds

𝐏(τn1)\displaystyle\mathbf{P}(\tau_{n}\geq 1) =𝐏(supj0jδsupt<1fj(t)Hm0s0+1<n)\displaystyle=\mathbf{P}\left(\sup_{j\geq 0}\left\langle j\right\rangle^{-\delta}\sup_{t<1}\left\|f^{j}(t)\right\|_{H^{s_{0}+1}_{m_{0}}}<n\right)
𝐏(C0<n)\displaystyle\geq\mathbf{P}(C_{0}<n)
1𝐏(C0>n)\displaystyle\geq 1-\mathbf{P}(C_{0}>n)
1𝒪(nα).\displaystyle\geq 1-\mathcal{O}(n^{-\alpha}).

Therefore, limn𝐏(τn>1)=1\lim_{n\to\infty}\mathbf{P}(\tau_{n}>1)=1 and so if we define

ξ=limnτn,\displaystyle\xi=\lim_{n\to\infty}\tau_{n},

then ξ\xi is almost-surely greater than or equal to 11.

Let δ(0,1/6)\delta\in(0,1/6) be fixed arbitrary. We will show by induction that K0>0\exists K_{0}>0 (deterministic constant depending on δ\delta) such that for all j1j\geq 1, there holds

𝐄sups<tτnfjfj1Hm0s02\displaystyle\mathbf{E}\sup_{s<t\wedge\tau_{n}}\left\|f^{j}-f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}^{2} (K0n4t)jj4δjj!.\displaystyle\leq\frac{(K_{0}n^{4}t)^{j}j^{4\delta j}}{j!}. (3.39)

First consider the case j=1j=1. The calculation of dxαvβ(f1f0)Lm022\mathrm{d}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{1}-f^{0})\right\|^{2}_{L^{2}_{m_{0}}} is the same in Lemma 3.1 except for the nonlinear terms. That is, for |α|+|β|s0|\alpha|+|\beta|\leq s_{0} we have for some constant C>0C>0

dxαvβ(f1f0)Lm022\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{1}-f^{0})\|_{L_{m_{0}}^{2}}^{2} Cf1f0Hm0s02dt\displaystyle\leq C\left\|f^{1}-f^{0}\right\|_{H^{s_{0}}_{m_{0}}}^{2}\mathrm{d}t
2θR(f0Hm0s0)xαvβ(φϵE0vf1),xαvβ(f1f0)m0dt\displaystyle\quad-2\left<\theta_{R}(\left\|f^{0}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{0}\cdot\nabla_{v}f^{1}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{1}-f^{0})\right>_{m_{0}}\mathrm{d}t
2xαvβ(v(f1f0)dWt),xαvβ(f1f0)m0.\displaystyle\quad-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f^{1}-f^{0})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{1}-f^{0})\right>_{m_{0}}.

For the nonlinear term we note that by (3.7) we have, recalling the definition of τn\tau_{n}

|θR(f0Hm0s0)xαvβ(φϵE0vf1),xαvβ(f1f0)m0|\displaystyle\left|\left<\theta_{R}(\left\|f^{0}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{0}\cdot\nabla_{v}f^{1}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{1}-f^{0})\right>_{m_{0}}\right| ϵ,Rnf1f0Hm0s0.\displaystyle\lesssim_{\epsilon,R}n\left\|f^{1}-f^{0}\right\|_{H^{s_{0}}_{m_{0}}}.

Integrating in time and using the BDG inequality as above, we obtain (note that f1f^{1} and f0f^{0} have the same initial data),

𝐄sups<tτnf1(s)f0(s)Hm0s02\displaystyle\mathbf{E}\sup_{s<t\wedge\tau_{n}}\|f^{1}(s)-f^{0}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}\leq C0tτn𝐄f1(s)f0(s)Hm0s02ds\displaystyle C\int_{0}^{t\wedge\tau_{n}}\mathbf{E}\|f^{1}(s)-f^{0}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}\mathrm{d}s
+Cn2t+C𝐄(0tτnf1(s)f0(s)Hm0s04ds)12\displaystyle\quad+Cn^{2}t+C\mathbf{E}\left(\int_{0}^{t\wedge\tau_{n}}\|f^{1}(s)-f^{0}(s)\|_{H_{m_{0}}^{s_{0}}}^{4}\mathrm{d}s\right)^{\frac{1}{2}}
\displaystyle\leq Cn2t+C0tτn𝐄f1(s)f0(s)Hm0s02ds\displaystyle Cn^{2}t+C\int_{0}^{t\wedge\tau_{n}}\mathbf{E}\|f^{1}(s)-f^{0}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}\mathrm{d}s
+12𝐄supstτnf1(s)f0(s)Hm0s02,\displaystyle+\frac{1}{2}\mathbf{E}\sup_{s\leq t\wedge\tau_{n}}\|f^{1}(s)-f^{0}(s)\|_{H_{m_{0}}^{s_{0}}}^{2},

Therefore, Grönwall’s inequality verifies (3.39) for j=1j=1 and some large K0K_{0}.

Next consider the inductive step. Hence, suppose that (3.39) holds for jj and we wish to verify that it holds for j+1j+1. As above, for some constant C>0C>0

dxαvβ(fj+1fj)Lm022\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\|_{L_{m_{0}}^{2}}^{2} Cfj+1fjHm0s02dt\displaystyle\leq C\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}^{2}\mathrm{d}t
2θR(fjHm0s0)xαvβ(φϵEjvfj+1),xαvβ(fj+1fj)m0dt\displaystyle\quad-2\left<\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{j}\cdot\nabla_{v}f^{j+1}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}\mathrm{d}t
+2θR(fj1Hm0s0)xαvβ(φϵEj1vfj),xαvβ(fj+1fj)m0dt\displaystyle\quad+2\left<\theta_{R}(\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{j-1}\cdot\nabla_{v}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}\mathrm{d}t
2xαvβ(v(fj+1fj)dWt),xαvβ(fj+1fj)m0.\displaystyle\quad-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f^{j+1}-f^{j})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}.

The nonlinearity separates into several natural terms, namely

()\displaystyle(\ast) =2θR(fjHm0s0)xαvβ(φϵEj(vfj+1vfj)),xαvβ(fj+1fj)m0dt\displaystyle=-2\left<\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{j}\cdot(\nabla_{v}f^{j+1}-\nabla_{v}f^{j})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}\mathrm{d}t
2(θR(fjHm0s0)θR(fj1Hm0s0))xαvβ(φϵEjvfj),xαvβ(fj+1fj)m0dt\displaystyle\quad-2\left<\left(\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}})-\theta_{R}(\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})\right)\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}\ast E^{j}\cdot\nabla_{v}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}\mathrm{d}t
2θR(fj1Hm0s0)xαvβ((φϵEjφϵEj1)vfj),xαvβ(fj+1fj)m0dt\displaystyle\quad-2\left<\theta_{R}(\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}((\varphi_{\epsilon}\ast E^{j}-\varphi_{\epsilon}\ast E^{j-1})\cdot\nabla_{v}f^{j}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{j+1}-f^{j})\right>_{m_{0}}\mathrm{d}t
=𝒩1+𝒩2+𝒩3.\displaystyle=\mathcal{N}_{1}+\mathcal{N}_{2}+\mathcal{N}_{3}.

The term 𝒩1\mathcal{N}_{1} is treated via (3.7) in the same manner as in Lemma 3.1, giving

𝒩1R,ϵfj+1fjHm0s02.\displaystyle\mathcal{N}_{1}\lesssim_{R,\epsilon}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}^{2}.

The terms 𝒩2,𝒩3\mathcal{N}_{2},\mathcal{N}_{3} however are different. The term 𝒩3\mathcal{N}_{3} is estimated via the following for t<τnt<\tau_{n}:

𝒩3\displaystyle\mathcal{N}_{3} RfjHm0s0+1fjfj1Hm0s0fj+1fjHm0s0\displaystyle\lesssim_{R}\left\|f^{j}\right\|_{H^{s_{0}+1}_{m_{0}}}\left\|f^{j}-f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}
njδfjfj1Hm0s0fj+1fjHm0s0.\displaystyle\lesssim nj^{\delta}\left\|f^{j}-f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}.

The term 𝒩2\mathcal{N}_{2} requires a control on the difference θR(fjHm0s0)θR(fj1Hm0s0)\theta_{R}(\left\|f^{j}\right\|_{H_{m_{0}}^{s_{0}}})-\theta_{R}(\left\|f^{j-1}\right\|_{H_{m_{0}}^{s_{0}}}):

θR(fjHm0s0)\displaystyle\theta_{R}(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}}) θR(fj1Hm0s0)\displaystyle-\theta_{R}(\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})
=01θR(zfjHm0s0+(1z)fj1Hm0s0)(fjHm0s0fj1Hm0s0)dz.\displaystyle=\int_{0}^{1}\theta_{R}^{\prime}(z\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}}+(1-z)\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})(\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}}-\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}})\mathrm{d}z.

Therefore, for t<τnt<\tau_{n}

𝒩2\displaystyle\mathcal{N}_{2} |fjHm0s0fj1Hm0s0|fjLm02fjHm0s0+1fj+1fjHm0s0\displaystyle\lesssim\left|\left\|f^{j}\right\|_{H^{s_{0}}_{m_{0}}}-\left\|f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}\right|\left\|f^{j}\right\|_{L^{2}_{m_{0}}}\left\|f^{j}\right\|_{H^{s_{0}+1}_{m_{0}}}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}
fjfj1Hm0s0fjLm02fjHm0s0+1fj+1fjHm0s0\displaystyle\lesssim\left\|f^{j}-f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}\left\|f^{j}\right\|_{L^{2}_{m_{0}}}\left\|f^{j}\right\|_{H^{s_{0}+1}_{m_{0}}}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}
n2j2δfjfj1Hm0s0fj+1fjHm0s0.\displaystyle\lesssim n^{2}j^{2\delta}\left\|f^{j}-f^{j-1}\right\|_{H^{s_{0}}_{m_{0}}}\left\|f^{j+1}-f^{j}\right\|_{H^{s_{0}}_{m_{0}}}.

Integrating in time and using the BDG inequality as above, we obtain (noting that fj+1f^{j+1} and fjf^{j} have the same initial data) for t<τnt<\tau_{n}:

𝐄sups<tfj+1(s)fj(s)Hm0s02\displaystyle\mathbf{E}\sup_{s<t}\|f^{j+1}(s)-f^{j}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}\leq C0t𝐄fj+1(s)fj(s)Hm0s02ds\displaystyle C\int_{0}^{t}\mathbf{E}\|f^{j+1}(s)-f^{j}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}\mathrm{d}s
+C0n4j4δ𝐄0tfj(s)fj1(s)Hm0s02ds\displaystyle\quad+C_{0}n^{4}j^{4\delta}\mathbf{E}\int_{0}^{t}\left\|f^{j}(s)-f^{j-1}(s)\right\|_{H^{s_{0}}_{m_{0}}}^{2}\mathrm{d}s
+12𝐄supstfj+1(s)fj(s)Hm0s02\displaystyle\quad+\frac{1}{2}\mathbf{E}\sup_{s\leq t}\|f^{j+1}(s)-f^{j}(s)\|_{H_{m_{0}}^{s_{0}}}^{2}

By the inductive hypothesis

Cn4j4δ𝐄0tfj(s)fj1(s)Hm0s02ds\displaystyle Cn^{4}j^{4\delta}\mathbf{E}\int_{0}^{t}\left\|f^{j}(s)-f^{j-1}(s)\right\|_{H^{s_{0}}_{m_{0}}}^{2}\mathrm{d}s C0n4j4δ0t(C0n4s)jj4δjj!ds\displaystyle\leq C_{0}n^{4}j^{4\delta}\int_{0}^{t}\frac{(C_{0}n^{4}s)^{j}j^{4\delta j}}{j!}\mathrm{d}s
(C0n4t)j+1j4δ(j+1)(j+1)!,\displaystyle\leq\frac{(C_{0}n^{4}t)^{j+1}j^{4\delta(j+1)}}{(j+1)!},

and so we have verfied (3.39).

Finally, we show that (3.39) implies that {fj}\left\{f^{j}\right\} is Cauchy in Lω2Lt2([0,τn];Hm0s0)L_{\omega}^{2}L_{t}^{2}([0,\tau_{n}];H_{m_{0}}^{s_{0}}). Indeed, let k<k<\ell and

𝐄sups<tτnffkHm0s02\displaystyle\mathbf{E}\sup_{s<t\wedge\tau_{n}}\left\|f^{\ell}-f^{k}\right\|_{H^{s_{0}}_{m_{0}}}^{2} j=k(K0n4t)jj4δjj!.\displaystyle\leq\sum_{j=k}^{\ell}\frac{(K_{0}n^{4}t)^{j}j^{4\delta j}}{j!}. (3.40)

Hence, if we choose k>(2C0n4t)1/δk>(2C_{0}n^{4}t)^{1/\delta}, then

𝐄sups<tτnffkHm0s02\displaystyle\mathbf{E}\sup_{s<t\wedge\tau_{n}}\left\|f^{\ell}-f^{k}\right\|_{H^{s_{0}}_{m_{0}}}^{2} j=k12δjkjδj4δjj!j=k12δjj5δjj!.\displaystyle\leq\sum_{j=k}^{\ell}\frac{1}{2^{-\delta j}}\frac{k^{j\delta}j^{4\delta j}}{j!}\leq\sum_{j=k}^{\ell}\frac{1}{2^{-\delta j}}\frac{j^{5\delta j}}{j!}.

By Stirling’s formula we have the following uniformly in jj (using 5δ<15\delta<1),

j5δjj!δ1,\displaystyle\frac{j^{5\delta j}}{j!}\lesssim_{\delta}1,

therefore

𝐄sups<tτnffkHm0s02\displaystyle\mathbf{E}\sup_{s<t\wedge\tau_{n}}\left\|f^{\ell}-f^{k}\right\|_{H^{s_{0}}_{m_{0}}}^{2} j=k12δj12δk.\displaystyle\lesssim\sum_{j=k}^{\ell}\frac{1}{2^{-\delta j}}\lesssim\frac{1}{2^{-\delta k}}.

We conclude that the sequence is Cauchy as claimed in the lemma. ∎

Lemma 3.7.

For each n,n, the iterates {fj}j=1\left\{f^{j}\right\}_{j=1}^{\infty} converge uniformly in Hm0s0H_{m_{0}}^{s_{0}} on compact subintervals of [0,τn][0,\tau_{n}] to a strong pathwise solution of the SPDE (2.7) on the set {τn>0}Ω\left\{\tau_{n}>0\right\}\subset\Omega.

Proof.

Consider only ω{τn>0}Ω\omega\in\left\{\tau_{n}>0\right\}\subset\Omega. Let ff be the limiting process of the fjf^{j} in Lω2C([0,τn];Hm0s0)L_{\omega}^{2}C([0,\tau_{n}];H_{m_{0}}^{s_{0}}) - whose existence is guaranteed by Lemma 3.6. We will show that each term in (3.1) converges to the corresponding term in (2.7). The convergence of the linear terms is straightforward:

𝐄suptTτn0tvx(fj+1f)dsHm01s012\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}v\cdot\nabla_{x}(f^{j+1}-f)\mathrm{d}s\right\|_{H_{m_{0}-1}^{s_{0}-1}}^{2} m0T2𝐄suptTτnfj+1fHm0s020,\displaystyle\lesssim_{m_{0}}T^{2}\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}\to 0, (3.41)
𝐄suptTτn0tΔv(fj+1f)dsHm0s022\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\Delta_{v}(f^{j+1}-f)\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-2}}^{2} T2𝐄suptTτnfj+1fHm0s020,\displaystyle\lesssim T^{2}\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}\to 0, (3.42)
𝐄suptTτn0tdivv(fj+1vfv)dsHm01s012\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\operatorname{\mathrm{div}}_{v}(f^{j+1}v-fv)\mathrm{d}s\right\|_{H_{m_{0}-1}^{s_{0}-1}}^{2} m0T2𝐄suptTτnfj+1fHm0s020,\displaystyle\lesssim_{m_{0}}T^{2}\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}\to 0, (3.43)
𝐄suptTτn0tk(σkekv)2(fj+1f)dsHm0s022\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\sum_{k}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(f^{j+1}-f)\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-2}}^{2} T2𝐄suptTτnfj+1fHm0s020.\displaystyle\lesssim T^{2}\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}\to 0. (3.44)

For the nonlinear electric field terms, we have:

𝐄suptTτn0t(θRjφϵEjvfjθRφϵEvf)dsHm0s01𝒩1+𝒩2+𝒩3,\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}(\theta_{R}^{j}\varphi_{\epsilon}*E^{j}\cdot\nabla_{v}f^{j}-\theta_{R}\varphi_{\epsilon}*E\cdot\nabla_{v}f)\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-1}}\lesssim\mathcal{N}_{1}+\mathcal{N}_{2}+\mathcal{N}_{3},

where:

𝒩1:=𝐄suptTτn0t(θRjθR)φϵEjvfjdsHm0s01,\displaystyle\mathcal{N}_{1}:=\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}(\theta_{R}^{j}-\theta_{R})\varphi_{\epsilon}*E^{j}\cdot\nabla_{v}f^{j}\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-1}},
𝒩2:=𝐄suptTτn0tθRφφϵ(EjE)vfjdsHm0s01,\displaystyle\mathcal{N}_{2}:=\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\theta_{R}\varphi\varphi_{\epsilon}*(E^{j}-E)\cdot\nabla_{v}f^{j}\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-1}},
𝒩3:=𝐄suptTτn0tθRφϵEv(fjf)dsHm0s01.\displaystyle\mathcal{N}_{3}:=\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\theta_{R}\varphi_{\epsilon}*E\cdot\nabla_{v}(f^{j}-f)\mathrm{d}s\right\|_{H_{m_{0}}^{s_{0}-1}}.

These terms are estimated as follows:

𝒩1R\displaystyle\mathcal{N}_{1}\lesssim_{R} 𝐄suptTτn0t|fjHm0s0fHm0s0|fjHm0s02ds\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\int_{0}^{t}\left|\|f^{j}\|_{H_{m_{0}}^{s_{0}}}-\|f\|_{H_{m_{0}}^{s_{0}}}\right|\|f^{j}\|_{H_{m_{0}}^{s_{0}}}^{2}\mathrm{d}s
\displaystyle\lesssim 𝐄suptTτn(fjfL2([0,t];Hm0s0)fjL4([0,T];Hm0s0)2)\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left(\|f^{j}-f\|_{L^{2}([0,t];H_{m_{0}}^{s_{0}})}\|f^{j}\|_{L^{4}([0,T];H_{m_{0}}^{s_{0}})}^{2}\right)
\displaystyle\to 0,\displaystyle 0,
𝒩2𝐄suptTτn(fjfL2([0,t];Hm0s0)fjL2([0,t];Hm0s0))0,\displaystyle\mathcal{N}_{2}\lesssim\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left(\left\|f^{j}-f\right\|_{L^{2}([0,t];H_{m_{0}}^{s_{0}})}\|f^{j}\|_{L^{2}([0,t];H_{m_{0}}^{s_{0}})}\right)\to 0,
𝒩3𝐄suptTτn(fL2([0,t];Hm0s0)fjfHm0s0)0.\displaystyle\mathcal{N}_{3}\lesssim\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left(\|f\|_{L^{2}([0,t];H_{m_{0}}^{s_{0}})}\|f^{j}-f\|_{H_{m_{0}}^{s_{0}}}\right)\to 0.

Lastly, for the martingale terms we use the BDG inequality:

𝐄suptTτn0tkσkekv(fj+1f)dWskHm0s012\displaystyle\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\left\|\int_{0}^{t}\sum_{k}\sigma_{k}e_{k}\cdot\nabla_{v}(f^{j+1}-f)\cdot\mathrm{d}W_{s}^{k}\right\|_{H_{m_{0}}^{s_{0}-1}}^{2}\lesssim 𝐄0Tfj+1fHm0s02ds\displaystyle\mathbf{E}\int_{0}^{T}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}\mathrm{d}s
\displaystyle\lesssim T2𝐄suptTτnfj+1fHm0s02\displaystyle T^{2}\mathbf{E}\sup_{t\leq T\wedge\tau_{n}}\|f^{j+1}-f\|_{H_{m_{0}}^{s_{0}}}^{2}
\displaystyle\to 0.\displaystyle 0. (3.45)

Combining the above, we see that ff is a solution of (2.7). ∎

Corollary 3.8.

There exists a global, strong pathwise solution of the SPDE (2.7) such that p[2,\forall p\in[2,\infty, fLωpCt,locHm2σ3LωpLt,locHmσf\in L_{\omega}^{p}C_{t,loc}H_{m^{\prime}-2}^{\sigma^{\prime}-3}\cap L_{\omega}^{p}L_{t,loc}^{\infty}H_{m^{\prime}}^{\sigma^{\prime}}.

Proof.

By sending nn\to\infty and using that τn\tau_{n} is a non-decreasing sequence such that limn(τn>1)=1\lim_{n\to\infty}\mathbb{P}(\tau_{n}>1)=1 we see that almost-surely, {fj}j=1\left\{f^{j}\right\}_{j=1}^{\infty} converges uniformly in Hm0s0H_{m_{0}}^{s_{0}} on compact subintervals of [0,1)[0,1) to a limiting function fCt([0,1);Hm0s0)f\in C_{t}([0,1);H_{m_{0}}^{s_{0}}). By Sobolev interpolation, and uniform boundedness in Lt,locHmσL^{\infty}_{t,\mathrm{loc}}H^{\sigma^{\prime}}_{m^{\prime}}, we obtain similar uniform convergence in Hm′′s′′H_{m^{\prime\prime}}^{s^{\prime\prime}} for all 0s′′<s0\leq s^{\prime\prime}<s^{\prime} and m′′<mm^{\prime\prime}<m^{\prime}. At the same time, the uniform bounds on {fj}\left\{f^{j}\right\} from Lemma 3.1 imply that p[2,)\forall p\in[2,\infty), fLωpCtHm2σ3LωpLt,locHmσf\in L_{\omega}^{p}C_{t}H_{m^{\prime}-2}^{\sigma^{\prime}-3}\cap L_{\omega}^{p}L_{t,loc}^{\infty}H_{m^{\prime}}^{\sigma^{\prime}} by the lower semicontinuity of weak convergence. By Lemma 3.7, the limiting function ff is also a solution of (2.7). Now, we simply iterate the construction starting at t=1/2,3/2,t=1/2,3/2,... to obtain the existence of a global solution satisfying the desired bounds. ∎

The following lemma proves uniqueness of solutions to (2.7), thus completing the proof of Lemma 2.4.

Lemma 3.9.

Let f,ff,f^{\prime} be two global pathwise solutions to (2.7) on the same stochastic basis with f(0)=f(0)=f0f(0)=f^{\prime}(0)=f_{0} for some 0\mathcal{F}_{0}-measurable f0f_{0} with 𝐄f0Hmσp<\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}<\infty for some p>2p>2 and such that for all ϵ>0\epsilon>0, f,fLωpϵCtHm2σ3LωpLt,locHmσ.f,f^{\prime}\in L_{\omega}^{p-\epsilon}C_{t}H_{m^{\prime}-2}^{\sigma^{\prime}-3}\cap L_{\omega}^{p-}L_{t,loc}^{\infty}H_{m^{\prime}}^{\sigma^{\prime}}. Then f,ff,f^{\prime} are indistinguishable, that is:

𝐏(f(t)=f(t) for all 0t)=1.\mathbf{P}\left(f(t)=f^{\prime}(t)\,\text{ for all }0\leq t\right)=1. (3.46)
Proof.

This is proved by an energy estimate on ffHm1σ12.\|f-f^{\prime}\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}. Similarly to the proof of Lemma 3.1, for |α|+|β|σ1|\alpha|+|\beta|\leq\sigma^{\prime}-1 we have:

dxαvβ(ff)Lm122\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\|_{L_{m^{\prime}-1}^{2}}^{2}\leq CffHm1σ12dt\displaystyle C\|f-f\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}\mathrm{d}t
2xαvβ(θRφϵEvfθRφϵEvf),xαvβ(ff)m1dt\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}\varphi_{\epsilon}*E\cdot\nabla_{v}f-\theta_{R}^{\prime}\varphi_{\epsilon}*E^{\prime}\cdot\nabla_{v}f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1}\mathrm{d}t
20txαvβ(v(ff)dWt),xαvβ(ff)m1.\displaystyle-2\int_{0}^{t}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f-f^{\prime})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1}. (3.47)

We split the electric field contributions as:

xαvβ(θRφϵEvfθRφϵEvf),xαvβ(ff)m1=𝒩1+𝒩2+𝒩3,\displaystyle\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}\varphi_{\epsilon}*E\cdot\nabla_{v}f-\theta_{R}^{\prime}\varphi_{\epsilon}*E^{\prime}\cdot\nabla_{v}f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1}=\mathcal{N}_{1}+\mathcal{N}_{2}+\mathcal{N}_{3},

where:

𝒩1:=(θRθR)xαvβ(φϵEvf),xαvβ(ff)m1,\displaystyle\mathcal{N}_{1}:=\left\langle(\theta_{R}-\theta_{R}^{\prime})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1},
𝒩2:=θRxαvβ(φϵ(EE)vf),xαvβ(ff)m1,\displaystyle\mathcal{N}_{2}:=\left\langle\theta_{R}^{\prime}\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*(E-E^{\prime})\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1},
𝒩3:=θRxαvβ(φϵEv(ff)),xαvβ(ff)m1.\displaystyle\mathcal{N}_{3}:=\left\langle\theta_{R}^{\prime}\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*E^{\prime}\cdot\nabla_{v}(f-f^{\prime})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right\rangle_{m^{\prime}-1}.

These are estimated as follows:

|𝒩1|RfHmσ2ffHm1σ12,|\mathcal{N}_{1}|\lesssim_{R}\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{2}\|f-f^{\prime}\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}, (3.48)
|𝒩2|fHmσffHm1σ12,|\mathcal{N}_{2}|\lesssim\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}\|f-f\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}, (3.49)
|𝒩3|RffHm1σ12,|\mathcal{N}_{3}|\lesssim_{R}\|f-f^{\prime}\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}, (3.50)

where in (3.48) we used the mean value theorem for θR\theta_{R} and (3.6) , in (3.49) we used (3.6), and in (3.50) we used (3.7) - in addition to Lemma 3.2 for each electric field.

Now, fix K>0K>0. Since f,fLωpLt,locHmσf,f^{\prime}\in L_{\omega}^{p-}L_{t,loc}^{\infty}H_{m^{\prime}}^{\sigma^{\prime}}, the stopping time:

ξK=inf{t0:supstf(s)Hmσ+supstf(s)Hmσ>K}ττ\xi_{K}=\inf\{t\geq 0:\sup_{s\leq t}\|f(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}+\sup_{s\leq t}\|f^{\prime}(s)\|_{H_{m^{\prime}}^{\sigma^{\prime}}}>K\}\wedge\tau\wedge\tau^{\prime}

is almost surely finite. Even though it is not clear that ξK\xi_{K} is almost surely positive in general, for almost every ωΩ\omega\in\Omega there exists K>0K>0 such that ξK>0,\xi_{K}>0, and in addition ξKττ\xi_{K}\to\tau\wedge\tau^{\prime} 𝐏\mathbf{P}–a.s. as KK\to\infty. With this in mind, we fix T>0T>0 and use (3.48)-(3.50) and the BDG inequality in (3.47), to obtain:

𝐄supstξKf(s)f(s)Hm1σ12\displaystyle\mathbf{E}\sup_{s\leq t\wedge\xi_{K}}\|f(s)-f^{\prime}(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}\lesssim 𝐄0tsupssξKf(s)f(s)Hm1σ12ds\displaystyle\mathbf{E}\int_{0}^{t}\sup_{s\leq s^{\prime}\wedge\xi_{K}}\|f(s)-f^{\prime}(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}\mathrm{d}s^{\prime}
+𝐄(0tsupssξKf(s)f(s)Hm1σ14ds)12\displaystyle+\mathbf{E}\left(\int_{0}^{t}\sup_{s\leq s^{\prime}\wedge\xi_{K}}\|f(s)-f^{\prime}(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{4}\mathrm{d}s^{\prime}\right)^{\frac{1}{2}}
\displaystyle\leq C𝐄0tsupssξKf(s)f(s)Hm1σ12ds\displaystyle C\mathbf{E}\int_{0}^{t}\sup_{s\leq s^{\prime}\wedge\xi_{K}}\|f(s)-f^{\prime}(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}\mathrm{d}s^{\prime}
+12𝐄supstξKf(s)Hm1σ12,\displaystyle+\frac{1}{2}\mathbf{E}\sup_{s\leq t\wedge\xi_{K}}\|f(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}, (3.51)

for all tTt\leq T, whereby the usual rearrangement and Grönwall’s lemma give:

𝐄supsTξKf(s)f(s)Hm1σ12=0.\mathbf{E}\sup_{s\leq T\wedge\xi_{K}}\|f(s)-f^{\prime}(s)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}=0.

Taking KK\to\infty and then TT\to\infty, the conclusion follows. ∎

3.2 Proof of Lemma 2.1

Next, we want to pass to the limit ϵ0\epsilon\to 0, for which we need uniform-in-ϵ\epsilon estimates similar to those of Lemma 3.1, but this time for a family {fϵ}ϵ>0\left\{f_{\epsilon}\right\}_{\epsilon>0} of solutions to (2.7). Note that since fϵLt,locHmσCtHm2σ3,f_{\epsilon}\in L_{t,loc}^{\infty}H_{m^{\prime}}^{\sigma^{\prime}}\cap C_{t}H_{m^{\prime}-2}^{\sigma^{\prime}-3}, the highest norm in which we know fϵf_{\epsilon} is continuous is CtHm1σ1C_{t}H_{m^{\prime}-1}^{\sigma^{\prime}-1} - and thus we use this as the base for our estimates.

Lemma 3.10.

Let ff be a solution of (2.7). For α(0,12),\alpha\in(0,\frac{1}{2}), p2,p\geq 2, we have the uniform in ϵ\epsilon estimates:

𝐄suptTf(t)Hm1σ1pp,T,R,f01\mathbf{E}\sup_{t\leq T}\|f(t)\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{p}\lesssim_{p,T,R,f_{0}}1 (3.52)

and

𝐄fWα,p([0,T];Hm2σ3)pp,T,R,f01.\mathbf{E}\|f\|_{W^{\alpha,p}([0,T];H_{m^{\prime}-2}^{\sigma^{\prime}-3})}^{p}\lesssim_{p,T,R,f_{0}}1. (3.53)
Proof.

The proof proceeds by induction in the number of derivatives111see for instance [luk2016strichartz] for similar inductive energy estimates for the relativistic Vlasov–Maxwell system.. The inductive hypothesis is that for s>d/2s>d/2 derivatives on a solution ff of (2.7), we have:

𝐄suptTfHm1spp,R,T,f01.\mathbf{E}\sup_{t^{\prime}\leq T}\|f\|_{H_{m^{\prime}-1}^{s}}^{p}\lesssim_{p,R,T,f_{0}}1. (3.54)

We show that this implies the same estimate for s+1s+1. Begin by using Itô’s formula on xαvβfLm122\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{m^{\prime}-1}^{2}}^{2} for |α|+|β|=s+1,|\alpha|+|\beta|=s+1, where similarly to (3.18) we obtain:

dxαvβfLm122=\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{m^{\prime}-1}^{2}}^{2}= 2xαvβ(vxf),xαvβfm1dt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\mathrm{d}t
+2Δvxαvβf,xαvβfm1dt\displaystyle+2\left<\Delta_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\mathrm{d}t
+2xαvβ(divv(fv)),xαvβfm1dt\displaystyle+2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\operatorname{\mathrm{div}}_{v}(fv)),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\mathrm{d}t
2θR(fHm0s0)xαvβ(φϵEvf),xαvβfm1dt\displaystyle-2\left<\theta_{R}(\|f\|_{H_{m_{0}}^{s_{0}}})\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\mathrm{d}t
2xαvβ(vfdWt),xαvβfm1\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}
+kxαvβ[(σkekv)2f],xαvβfm1dt\displaystyle+\sum_{k}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}[(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f],\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\mathrm{d}t
+kxαvβ(σkekvf)Lm122\displaystyle+\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\|_{L_{m^{\prime}-1}^{2}}^{2}
=\displaystyle= 𝒯α,β(f)+𝒟α,β(f)+α,β(f)+𝒩α,β(f)+α,β(f)+𝒞α,β(f).\displaystyle\mathcal{T}_{\alpha,\beta}(f)+\mathcal{D}_{\alpha,\beta}(f)+\mathcal{F}_{\alpha,\beta}(f)+\mathcal{N}_{\alpha,\beta}(f)+\mathcal{M}_{\alpha,\beta}(f)+\mathcal{C}_{\alpha,\beta}(f). (3.55)

The linear terms are treated as in the proof of Lemma 3.1, and the only term that requires new attention is 𝒩α,β(f).\mathcal{N}_{\alpha,\beta}(f). By the classical Gagliardo-Nirenberg inequality (see e.g. [Proposition A.3 [tao2006nonlinear]] we have:

|xαvβ(φϵEvf),xαvβfm1|\displaystyle\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\varphi_{\epsilon}*E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}-1}\right|
\displaystyle\leq CφϵEW1,fHm1s+12dt\displaystyle C\|\varphi_{\epsilon}*E\|_{W^{1,\infty}}\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\mathrm{d}t
+γ<α|αγ|2xαγφϵELx2s|αγ|1vvβxγfLv,m2Lx2s|β|+|γ|+1xαvβfLm12dt\displaystyle+\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\|\partial_{x}^{\alpha-\gamma}\varphi_{\epsilon}*E\|_{L_{x}^{2\frac{s}{|\alpha-\gamma|-1}}}\|\nabla_{v}\partial_{v}^{\beta}\partial_{x}^{\gamma}f\|_{L_{v,m}^{2}L_{x}^{2\frac{s}{|\beta|+|\gamma|+1}}}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\|_{L_{m^{\prime}-1}^{2}}\mathrm{d}t
\displaystyle\leq CφϵEW1,fHm1s+12dt\displaystyle C\|\varphi_{\epsilon}*E\|_{W^{1,\infty}}\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\mathrm{d}t
+Cγ<α|αγ|2xφϵELx|β|+|γ|+1sφϵEHxs+1|αγ|1svvβfLm121ηfLm121+ηdt,\displaystyle+C\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\|\nabla_{x}\varphi_{\epsilon}*E\|_{L_{x}^{\infty}}^{\frac{|\beta|+|\gamma|+1}{s}}\|\varphi_{\epsilon}*E\|_{H_{x}^{s+1}}^{\frac{|\alpha-\gamma|-1}{s}}\|\nabla_{v}\partial_{v}^{\beta}f\|_{L_{m^{\prime}-1}^{2}}^{1-\eta}\|f\|_{L_{m^{\prime}-1}^{2}}^{1+\eta}\mathrm{d}t, (3.56)

where for fixed γ,\gamma, the interpolation parameter η\eta is given by:

η=\displaystyle\eta= |γ||α|1+d|α|1(12|β|+|γ|+12s)\displaystyle\frac{|\gamma|}{|\alpha|-1}+\frac{d}{|\alpha|-1}\left(\frac{1}{2}-\frac{|\beta|+|\gamma|+1}{2s}\right)
=\displaystyle= |γ||α|1+d|α|1|αγ|12s\displaystyle\frac{|\gamma|}{|\alpha|-1}+\frac{d}{|\alpha|-1}\cdot\frac{|\alpha-\gamma|-1}{2s} (3.57)

and thus η<1\eta<1 provided s>d2s>\frac{d}{2}. By Young’s inequality and (3.56) it follows that

𝒩α,βC(1+fHm1sp)dt+CRfHm1s+12dt.\mathcal{N}_{\alpha,\beta}\leq C\left(1+\|f\|_{H_{m^{\prime}-1}^{s}}^{p}\right)\mathrm{d}t+CR\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\mathrm{d}t. (3.58)

Plugging this back into (3.55), and using the same procedure as in the proof of Lemma 3.1, we obtain:

dfHm1s+12\displaystyle\mathrm{d}\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\leq C(1+R)fHm1s+12dt+|α|+|β|=s+1α,β(f)+0,0(f)\displaystyle C(1+R)\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\mathrm{d}t+\sum_{|\alpha|+|\beta|=s+1}\mathcal{M}_{\alpha,\beta}(f)+\mathcal{M}_{0,0}(f)
+C(1+fHm1sp)dt.\displaystyle+C(1+\|f\|_{H_{m^{\prime}-1}^{s}}^{p})\mathrm{d}t. (3.59)

We again integrate in time and apply the BDG inequality as in the proof of (3.2) for p=2p=2, where the only difference is the term fHm1sp,\|f\|_{H_{m^{\prime}-1}^{s}}^{p}, which is now controlled by the inductive hypothesis, and we get:

𝐄suptTfHm1s+12R,p,T,f01.\mathbf{E}\sup_{t\leq T}\|f\|_{H_{m^{\prime}-1}^{s+1}}^{2}\lesssim_{R,p,T,f_{0}}1. (3.60)

With this in hand, we can directly transfer the proof of (3.2) for p>2p>2 and obtain (3.52). Then the same argument as the proof of (3.3) (i.e. using the variant of BDG from [flandoli1995martingale]*Lemma 2.1) gives (3.53).

The last thing that remains is to demonstrate the inductive base of the preceding scheme. Here this is done by first estimating the Hm2H_{m^{\prime}}^{2} norm of ff. This is sufficient to start the inductive scheme above in 1d31\leq d\leq 3 as 2>d/22>d/2. As the linear terms are always controlled in the same way, we only focus on the electric field contributions. As always, we have:

|xαvβ(Evf),xαvβfm|CEW1,fHm22+γ<α|αγ|2xαγEvvβxγf,xαvβfm,\displaystyle\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}}\right|\leq C\|E\|_{W^{1,\infty}}\|f\|_{H_{m^{\prime}}^{2}}^{2}+\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left<\partial_{x}^{\alpha-\gamma}E\cdot\nabla_{v}\partial_{v}^{\beta}\partial_{x}^{\gamma}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right>_{m^{\prime}}, (3.61)

but since only two derivatives are acting on ff at this point, the terms in the summation are only present when |α|=2|\alpha|=2 and |β|=|γ|=0.|\beta|=|\gamma|=0. Let q=4q=4 in d=1,2d=1,2, and for d3d\geq 3 let qq be arbitrary such that 2<q<2dd22<q<\frac{2d}{d-2}. Then by Hölder’s inequality and Sobolev embeddings we have

|xαEvf,xαfm|\displaystyle\left|\left<\partial_{x}^{\alpha}E\cdot\nabla_{v}f,\partial_{x}^{\alpha}f\right>_{m^{\prime}}\right|\leq x2ELxqq1vfLv,m2LxqfHm2\displaystyle\|\nabla_{x}^{2}E\|_{L_{x}^{\frac{q}{q-1}}}\|\nabla_{v}f\|_{L_{v,m}^{2}L_{x}^{q}}\|f\|_{H_{m^{\prime}}^{2}}
\displaystyle\lesssim x2EHxd(q2)2qfHm22\displaystyle\|\nabla_{x}^{2}E\|_{H_{x}^{\frac{d(q-2)}{2q}}}\|f\|_{H_{m^{\prime}}^{2}}^{2}
\displaystyle\lesssim fHm0s0fHm22,\displaystyle\|f\|_{H_{m_{0}}^{s_{0}}}\|f\|_{H_{m^{\prime}}^{2}}^{2}, (3.62)

where we have used that s0>d2+1>d(q2)2qs_{0}>\frac{d}{2}+1>\frac{d(q-2)}{2q} and the embedding Hx1LxqH_{x}^{1}\subset L_{x}^{q} which holds for all d1d\geq 1 due to our choice of qq. From this point on the procedure is the same as in the inductive step. We plug (3.62) into (3.55) for α,β\alpha,\beta with |α|+|β|=2|\alpha|+|\beta|=2, sum over all such α,β\alpha,\beta as well as the case when α=β=0,\alpha=\beta=0, integrate in time, apply the BDG inequality and Grönwall’s lemma and obtain:

𝐄suptTf(t)Hm22T,R,f01.\mathbf{E}\sup_{t\leq T}\|f(t)\|_{H_{m^{\prime}}^{2}}^{2}\lesssim_{T,R,f_{0}}1. (3.63)

Then applying the same argument as in the proof of (3.2) for p>2,p>2, we also obtain for p>2p>2:

𝐄suptTf(t)Hm2pp,T,R,f01.\mathbf{E}\sup_{t\leq T}\|f(t)\|_{H_{m^{\prime}}^{2}}^{p}\lesssim_{p,T,R,f_{0}}1. (3.64)

This provides the inductive base and therefore the proof of the lemma is complete for 1d31\leq d\leq 3. ∎

For solutions to (2.7), it is unclear how to prove {fϵ}ϵ>0\left\{f^{\epsilon}\right\}_{\epsilon>0} forms a Cauchy sequence as ϵ0\epsilon\to 0. Instead, we employ a standard procedure based on the Skorokhod embedding theorem (see e.g. [ikeda2014stochastic]) to produce probabilistically weak (called martingale) solutions on a new stochastic basis, and then upgrade them to probabilistically strong using the Gyöngy–Krylov lemma from [gyongy1996existence] (see Lemma 3.12 below). We let (ϵn)n=1(\epsilon_{n})_{n=1}^{\infty} be a decreasing sequence of positive numbers with ϵn0\epsilon_{n}\to 0 as nn\to\infty and define the corresponding sequence fn:=fϵnf_{n}:=f^{\epsilon_{n}} of solutions to (2.7), which we have shown satisfy the uniform bounds (3.52) and (3.53). For α(0,12)\alpha\in(0,\frac{1}{2}) and p>2p>2 such that αp>1\alpha p>1, we define the pathspace

𝐗:=Wlocα,p([0,);Hm2σ3)Lloc([0,);Hm1σ1).\mathbf{X}:=W_{loc}^{\alpha,p}([0,\infty);H_{m^{\prime}-2}^{\sigma^{\prime}-3})\cap L_{loc}^{\infty}([0,\infty);H_{m^{\prime}-1}^{\sigma^{\prime}-1}). (3.65)

Recall that since αp>1\alpha p>1 and Hm2σ3Hm3σ4H_{m^{\prime}-2}^{\sigma^{\prime}-3}\subset H_{m^{\prime}-3}^{\sigma^{\prime}-4} compactly, from [flandoli1995martingale]*Theorem 2.2, we have:

Wlocα,p([0,);Hm2σ3)CtHm3σ4.W_{loc}^{\alpha,p}([0,\infty);H_{m^{\prime}-2}^{\sigma^{\prime}-3})\subset C_{t}H_{m^{\prime}-3}^{\sigma^{\prime}-4}.

By the uniform estimates (3.52) and (3.53), the laws νn:=(fn)\nu^{n}:=\mathcal{L}(f_{n}) are bounded in probability in 𝐗\mathbf{X}, and thus they are tight in the smaller pathspace

𝐗c:=C([0,);Hm3σ4)Lloc([0,);Hm1σ1).\mathbf{X}_{c}:=C([0,\infty);H_{m^{\prime}-3}^{\sigma^{\prime}-4})\cap L_{loc}^{\infty}([0,\infty);H_{m^{\prime}-1}^{\sigma^{\prime}-1}). (3.66)

Note that the tightness in Lloc([0,);Hm1σ1)L_{loc}^{\infty}([0,\infty);H_{m^{\prime}-1}^{\sigma^{\prime}-1}) is in the weak-\star topology. We now use this to obtain a martingale solution to (2.3) in high regularity.

Proposition 3.11.

Let μ0\mu_{0} be a probability measure on HmσH_{m^{\prime}}^{\sigma^{\prime}} so that HmσfHmσpdμ0(f)\int_{H_{m^{\prime}}^{\sigma^{\prime}}}\|f\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}\mathrm{d}\mu_{0}(f) for some p>2p>2. Then there exists a stochastic basis S~=(Ω~,~,{~t},P~)\tilde{S}=(\tilde{\Omega},\tilde{\mathcal{F}},\{\tilde{\mathcal{F}}_{t}\},\tilde{P}) and a predictable process

f~LωpCtHm3σ4LωpLt,locHm1σ1\tilde{f}\in L_{\omega}^{p-}C_{t}H_{m^{\prime}-3}^{\sigma^{\prime}-4}\cap L_{\omega}^{p-}L_{t,loc}^{\infty}H_{m^{\prime}-1}^{\sigma^{\prime}-1} (3.67)

such that (f~(0))=μ0\mathcal{L}(\tilde{f}(0))=\mu_{0} and f~\tilde{f} solves (2.3) in the sense that, there is a sequence of i.i.d Brownian motions {W~t(k)}\left\{\widetilde{W}_{t}^{(k)}\right\} such that the following equality holds in C([0,);Hm2σ3)C([0,\infty);H^{\sigma^{\prime}-3}_{m^{\prime}-2})

f~(t)=f0+0t(vxf~(s)θR(fHm0s0)E~(s)vf~(s)+νf~(s))ds0tvf~(s)dW~s𝐏–a.s.,\displaystyle\tilde{f}(t)=f_{0}+\int_{0}^{t}\left(-v\cdot\nabla_{x}\tilde{f}(s)-\theta_{R}(\left\|f\right\|_{H^{s_{0}}_{m_{0}}})\tilde{E}(s)\cdot\nabla_{v}\tilde{f}(s)+\nu\mathcal{L}\tilde{f}(s)\right)\mathrm{d}s-\int_{0}^{t}\nabla_{v}\tilde{f}(s)\circ\mathrm{d}\widetilde{W}_{s}\,\mathbf{P}\text{--a.s.},

with

E~=x(Δx)1(df~(t,,v)dv1).\displaystyle\tilde{E}=\nabla_{x}(-\Delta_{x})^{-1}\left(\int_{\mathbb{R}^{d}}\tilde{f}(t,\cdot,v)\mathrm{d}v-1\right).
Proof.

Let μn=(fn,𝒲)\mu^{n}=\mathcal{L}(f_{n},\mathcal{W}) in 𝐗c×C([0,);𝔘0).\mathbf{X}_{c}\times C([0,\infty);\mathfrak{U}_{0}). The sequence (μn)n=1(\mu^{n})_{n=1}^{\infty} is tight by the uniform estimates (3.52) and (3.53) combined with the fact that its projection onto C([0,);𝔘0)C([0,\infty);\mathfrak{U}_{0}) is the same for each nn. By Prokhorov’s theorem, (μn)n=1(\mu^{n})_{n=1}^{\infty} has a weakly convergent subsequence - reindexed to μn\mu^{n}. By Skorokhod’s embedding theorem, there exists a new probability space (Ω~,~,𝐏~)(\tilde{\Omega},\tilde{\mathcal{F}},\tilde{\mathbf{P}}) and on it random elements (f~n,𝒲~n)(\tilde{f}_{n},\tilde{\mathcal{W}}_{n}) with laws μn\mu^{n} which converge 𝐏~\tilde{\mathbf{P}}–a.s. to some limit (f~,𝒲~)(\tilde{f},\tilde{\mathcal{W}}) in the product topology of 𝐗c×C([0,);𝔘0)\mathbf{X}_{c}\times C([0,\infty);\mathfrak{U}_{0}). Then by a variation of the mollification technique employed in the proof of [bensoussan1995stochastic]*Equation 4.17 222See also [brzezniak2020well]*Proposition 3.2, (iii) for an application to the primitive equations where the noise is present as a stochastic transport term. the random elements (f~n,𝒲~n)(\tilde{f}_{n},\tilde{\mathcal{W}}_{n}) satisfy (2.7) just like (fn,𝒲)(f_{n},\mathcal{W}), but in the new probability space (Ω~,~,𝐏~)(\tilde{\Omega},\tilde{\mathcal{F}},\tilde{\mathbf{P}}):

f~n(t)f~n(0)+0t(vxf~n+E~nvf~n12k(σkekv)2f~nνf~n)ds+0tvf~ndW~tn=0,\tilde{f}_{n}(t)-\tilde{f}_{n}(0)+\int_{0}^{t}\left(v\cdot\nabla_{x}\tilde{f}_{n}+\tilde{E}_{n}\cdot\nabla_{v}\tilde{f}_{n}-\frac{1}{2}\sum_{k}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}\tilde{f}_{n}-\nu\mathcal{L}\tilde{f}_{n}\right)\mathrm{d}s+\int_{0}^{t}\nabla_{v}\tilde{f}_{n}\cdot\mathrm{d}\tilde{W}_{t}^{n}=0,

where we have denoted by W~tn\tilde{W}_{t}^{n} the external electric field corresponding to 𝒲~n\tilde{\mathcal{W}}_{n}; that is to say, if:

𝒲~n=kgkW~kn,\tilde{\mathcal{W}}_{n}=\sum_{k}g_{k}\tilde{W}_{k}^{n},

then W~n\tilde{W}^{n} is simply given by:

W~n=kσkekW~kn.\tilde{W}^{n}=\sum_{k}\sigma_{k}e_{k}\tilde{W}_{k}^{n}.

The passage to the limit nn\to\infty in the SPDEs (2.7) satisfied by (f~n,𝒲~n)(\tilde{f}_{n},\tilde{\mathcal{W}}_{n}) to obtain that the limit (f~,𝒲~)(\tilde{f},\tilde{\mathcal{W}}) solves (2.3) can be carried out by combining the convergences f~nf~\tilde{f}_{n}\to\tilde{f} and 𝒲n~𝒲~\tilde{\mathcal{W}_{n}}\to\tilde{\mathcal{W}} with [debussche2011local]*Lemma 2.1, so we omit the proof for technicalities. We simply note that the presence of the transport noise does not cause any additional difficulties in our setting. ∎

As our goal is to construct solutions in 𝒮\mathcal{S}, we need to \sayupgrade the martingale solutions of the preceding lemma to probabilistically strong solutions. With the above in mind, we now state the Gyöngy–Krylov lemma from [gyongy1996existence], which will allow us to combine the tightness of (fn)n=1(f_{n})_{n=1}^{\infty} with the pathwise uniqueness of the limit (Lemma 3.13 below) to show that in fact (2.3) has (unique) global solutions on the original stochastic basis 𝒮\mathcal{S}.

Lemma 3.12 (Gyöngy–Krylov).

Let (Yn)n=1(Y_{n})_{n=1}^{\infty} be a sequence of XX-valued random variables, where XX is a complete separable metric space. Then (Yn)n(Y_{n})_{n} converges in probability if and only if for every two subsequences Ynk,YlkY_{n_{k}},Y_{l_{k}} the joint sequence (Ynk,Ylk)(Y_{n_{k}},Y_{l_{k}}) has a subsequence (Ynk,Ylk)(Y_{n_{k^{\prime}}},Y_{l_{k^{\prime}}}) whose laws converge weakly to a probability measure ν\nu supported on the diagonal of X×XX\times X:

ν({(x,y)X×X:x=y})=1.\nu\left(\left\{(x,y)\in X\times X:\,x=y\right\}\right)=1. (3.68)

With this lemma at hand, we now set to prove pathwise uniqueness of solutions to (2.3), which is the content of the following:

Lemma 3.13.

Let f,ff,f^{\prime} be global solutions to (2.3) on the same stochastic basis with f(0)=f(0)=f0f(0)=f^{\prime}(0)=f_{0} almost surely, where 𝐄f0Hmσp<\mathbf{E}\|f_{0}\|_{H_{m^{\prime}}^{\sigma^{\prime}}}^{p}<\infty for some p>2p>2 and such that f,fLωpCtHm3σ4LωpLt,locHm1σ1f,f^{\prime}\in L_{\omega}^{p-}C_{t}H_{m^{\prime}-3}^{\sigma^{\prime}-4}\cap L_{\omega}^{p-}L_{t,loc}^{\infty}H_{m^{\prime}-1}^{\sigma^{\prime}-1}. Then f,ff,f^{\prime} are indistinguishable, that is:

𝐏(f(t)=f(t) for all t0)=1.\mathbf{P}\left(f(t)=f^{\prime}(t)\text{ for all }t\geq 0\right)=1. (3.69)
Proof.

First of all, notice that since f,fLt,locHm1σ1,f,f^{\prime}\in L_{t,loc}^{\infty}H_{m^{\prime}-1}^{\sigma^{\prime}-1}, for K>0K>0 the stopping times

ξK:=inf{t>0:fHm1σ12+fHm1σ12K}\xi_{K}:=\inf\{t>0:\|f\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}+\|f^{\prime}\|_{H_{m^{\prime}-1}^{\sigma^{\prime}-1}}^{2}\geq K\} (3.70)

are well defined and satisfy ξK\xi_{K}\to\infty as KK\to\infty, 𝐏\mathbf{P}–almost surely. We now perform an energy estimate on Hm0s0H_{m_{0}}^{s_{0}}. We use Itô’s formula on the quantity xαvβ(ff)Lm022,\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\|_{L_{m_{0}}^{2}}^{2}, for |α|+|β|=s0|\alpha|+|\beta|=s_{0}:

dxαvβ(ff)Lm22=\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\|_{L_{m^{\prime}}^{2}}^{2}= 2xαvβ(vx(ff)),xαvβ(ff)m0dt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}(f-f^{\prime})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\mathrm{d}t
+2Δvxαvβ(ff),xαvβ(ff)m0dt\displaystyle+2\left<\Delta_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\mathrm{d}t
+2xαvβdivv((ff)v),(ff)m0dt\displaystyle+2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}\operatorname{\mathrm{div}}_{v}((f-f^{\prime})v),(f-f^{\prime})\right>_{m_{0}}\mathrm{d}t
2xαvβ(θREfvfθREfvf),xαvβ(ff)m0dt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}E^{f}\cdot\nabla_{v}f-\theta_{R}^{\prime}E^{f^{\prime}}\cdot\nabla_{v}f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\mathrm{d}t
2xαvβ(v(ff)dWt),xαvβ(ff)m0\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f-f^{\prime})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}
+kxαvβ((σkekv)2(ff)),xαvβ(ff)m0dt\displaystyle+\sum_{k}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}((\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(f-f^{\prime})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\mathrm{d}t
kxαvβ(σkekv(ff))Lm022dt.\displaystyle-\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}(f-f^{\prime}))\|_{L_{m_{0}}^{2}}^{2}\mathrm{d}t. (3.71)

Clearly, all terms except those involving the electric fields can be estimated as in the proof of (3.2), so we only examine the electric field term:

|xαvβ(θREfvfθREfvf),xαvβ(ff)m0|\displaystyle\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}E^{f}\cdot\nabla_{v}f-\theta_{R}^{\prime}E^{f^{\prime}}\cdot\nabla_{v}f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\right|
\displaystyle\leq |(θRθR)xαvβ(Efvf),xαvβ(ff)m0|\displaystyle\left|(\theta_{R}-\theta_{R}^{\prime})\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{f}\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\right|
+|θRxαvβ((EfEf)vf),xαvβ(ff)m0|\displaystyle+\left|\theta_{R^{\prime}}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}((E^{f}-E^{f^{\prime}})\cdot\nabla_{v}f^{\prime}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\right|
+|θRxαvβ(Efv(ff)),xαvβ(ff)m0|\displaystyle+\left|\theta_{R^{\prime}}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{f^{\prime}}\cdot\nabla_{v}(f-f^{\prime})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f-f^{\prime})\right>_{m_{0}}\right|
\displaystyle\leq CR|fHm0s0fHm0s0|EfHxs0ffHm0s0\displaystyle C_{R}\left|\|f\|_{H_{m_{0}}^{s_{0}}}-\|f^{\prime}\|_{H_{m_{0}}^{s_{0}}}\right|\cdot\|E^{f}\|_{H_{x}^{s_{0}}}\|f-f^{\prime}\|_{H_{m_{0}}^{s_{0}}}
+CEfEfHs0vfHm0s0ffHm0s0\displaystyle+C\|E^{f}-E^{f^{\prime}}\|_{H^{s_{0}}}\|\nabla_{v}f\|_{H_{m_{0}}^{s_{0}}}\|f-f^{\prime}\|_{H_{m_{0}}^{s_{0}}}
+CRffHm0s02\displaystyle+CR\|f-f\|_{H_{m_{0}}^{s_{0}}}^{2}
R,K\displaystyle\lesssim_{R,K} ffHm0s02\displaystyle\|f-f^{\prime}\|_{H_{m_{0}}^{s_{0}}}^{2} (3.72)

for tξKt\leq\xi_{K}, where we have used the mean value theorem on θR\theta_{R} and Lemma 3.3. Thus, arguing similarly to the proof of Lemma 3.9 (i.e. by the BDG and Grönwall inequalities), we have:

𝐄supttξKTffHm0s02CR,K,Tf(0)f(0)Hm0s02=0.\mathbf{E}\sup_{t^{\prime}\leq t\wedge\xi_{K}\wedge T}\|f-f^{\prime}\|_{H_{m_{0}}^{s_{0}}}^{2}\leq C_{R,K,T}\|f(0)-f^{\prime}(0)\|_{H_{m_{0}}^{s_{0}}}^{2}=0. (3.73)

Since ξK\xi_{K}\to\infty as K,K\to\infty, the monotone convergence theorem implies that for all T>0T>0 we have:

𝐄suptTf(t)f(t)Hm0s02=0,\mathbf{E}\sup_{t\leq T}\|f(t)-f^{\prime}(t)\|_{H_{m_{0}}^{s_{0}}}^{2}=0, (3.74)

which implies (3.69) since TT is arbitrary. ∎

We now have everything we need to (subsequentially) pass to the limit ϵ0\epsilon\to 0 in the original stochastic basis.

Proof of Lemma 2.1.

We define the joint laws μn,l=(fn,fl,𝒲),\mu^{n,l}=\mathcal{L}(f_{n},f_{l},\mathcal{W}), Similarly to the discussion in Lemma 3.11, for any sequence μnk,lk\mu^{n_{k},l_{k}} with nk,lkn_{k},l_{k}\to\infty as k,k\to\infty, by Prokhorov’s theorem the estimates (3.52) and (3.53) (and the fact that {𝒲}\{\mathcal{W}\} is a singleton) provide a weakly convergent subsequence of probability measures in 𝐗c×𝐗c×C([0,);𝔘0)\mathbf{X}_{c}\times\mathbf{X}_{c}\times C([0,\infty);\mathfrak{U}_{0}), which we still denote (after relabelling) by μnk,lk\mu^{n_{k},l_{k}}, and we denote its limit by μ\mu. By the Skorokhod embedding theorem, we can construct a new stochastic basis again denoted by 𝒮~=(Ω~,F~,𝐏~)\tilde{\mathcal{S}}=(\tilde{\Omega},\tilde{F},\tilde{\mathbf{P}}) and on it a sequence of random elements (f~nk,f~lk,𝒲~k)(\tilde{f}_{n_{k}},\tilde{f}_{l_{k}},\tilde{\mathcal{W}}^{k}) and (f~,f~~,𝒲~)(\tilde{f},\tilde{\tilde{f}},\tilde{\mathcal{W}}) such that (f~nk,f~lk,𝒲~k)=μnk,lk\mathcal{L}(\tilde{f}_{n_{k}},\tilde{f}_{l_{k}},\tilde{\mathcal{W}}^{k})=\mu^{n_{k},l_{k}}, (f~,f~~,𝒲~)\mathcal{L}(\tilde{f},\tilde{\tilde{f}},\tilde{\mathcal{W}}) and:

(f~nk,f~lk,𝒲k~)(f~,f~~,𝒲~) in 𝐗c×𝐗c×C([0,);𝔘0),𝐏~–a.s..\displaystyle(\tilde{f}_{n_{k}},\tilde{f}_{l_{k}},\tilde{\mathcal{W}^{k}})\to(\tilde{f},\tilde{\tilde{f}},\tilde{\mathcal{W}})\quad\text{ in }\mathbf{X}_{c}\times\mathbf{X}_{c}\times C([0,\infty);\mathfrak{U}_{0}),\quad\tilde{\mathbf{P}}\text{--a.s..}

As in Lemma 3.11, (f~nk,𝒲~k)(\tilde{f}_{n_{k}},\tilde{\mathcal{W}}^{k}) and (f~lk,𝒲~k)(\tilde{f}_{l_{k}},\tilde{\mathcal{W}}^{k}) satisfy the SPDE (2.7) in the new stochastic basis (by the method of [bensoussan1995stochastic]*Section 4.3.4), so we can pass to the limit kk\to\infty in all the terms of (2.7) for (f~nk,𝒲~k)(\tilde{f}_{n_{k}},\tilde{\mathcal{W}}^{k}) and (f~lk,𝒲k~)(\tilde{f}_{l_{k}},\tilde{\mathcal{W}^{k}}) (using [debussche2011local]*Lemma 2.1 for the stochastic integrals) to show that (f~,𝒲~)(\tilde{f},\tilde{\mathcal{W}}) (f~~,𝒲~)(\tilde{\tilde{f}},\tilde{\mathcal{W}}) are solutions to (2.3) on the new stochastic basis. Since 𝐏(fnk(0)=flk(0))=1\mathbf{P}(f_{n_{k}}(0)=f_{l_{k}}(0))=1, we also have 𝐏~(f~nk(0)=f~lk(0))=1\tilde{\mathbf{P}}(\tilde{f}_{n_{k}}(0)=\tilde{f}_{l_{k}}(0))=1, and thus in the limit kk\to\infty we obtain 𝐏~(f~(0)=f~~(0))=1\tilde{\mathbf{P}}(\tilde{f}(0)=\tilde{\tilde{f}}(0))=1. Therefore, by Lemma 3.13, f~\tilde{f} and f~~\tilde{\tilde{f}} are indistinguishable. This means that the measure μ¯\bar{\mu}, defined as the projection of μ\mu onto the first two components 𝐗c×𝐗c\mathbf{X}_{c}\times\mathbf{X}_{c}, is in fact supported on the diagonal of 𝐗c×𝐗c\mathbf{X}_{c}\times\mathbf{X}_{c}. Thus, by Lemma 3.12, a subsequence of (fn)(f_{n}) converges in probability in the original stochastic basis 𝒮\mathcal{S} in the topology of 𝐗c\mathbf{X}_{c}^{\prime} to a limiting process ff which solves (2.3). This concludes the proof of Lemma 2.1. ∎

4 Proof of main theorem

In this section, with the results of Section 3 at hand, we prove the main result of the paper, Theorem 1.3. At first, we consider initial data f0f_{0} satisfying f0HmσM<\|f_{0}\|_{H_{m}^{\sigma}}\leq M<\infty almost surely, for a fixed deterministic MM. This assumption can be removed at the end by a cutting argument similar to that of Lemma 2.2. We treat the initial data f0f_{0} with a sequence of regularization and velocity cutoff operators n,\mathcal{R}^{n}, obtaining a sequence of regularized data

f0n:=nf0,f_{0}^{n}:=\mathcal{R}^{n}f_{0},

defined as

nf:=θn(v)n2dη(n)x,vf,\mathcal{R}^{n}f:=\theta_{n}(v)n^{2d}\eta\left(\frac{\cdot}{n}\right)\ast_{x,v}f,

where ηCc(2d)\eta\in C_{c}^{\infty}(\mathbb{R}^{2d}) satisfies η(x,v)0\eta(x,v)\geq 0 and η(x,v)dvdx=1\iint\eta(x,v)\mathrm{d}v\mathrm{d}x=1. We note the following properties of this regularization. The proofs are standard and are omitted for brevity.

Lemma 4.1.

Let ss0s^{\prime}\geq s\geq 0 be integers and mm0m^{\prime}\geq m\geq 0 be arbitrary. Then,

  • (i)

    The regularization operators n\mathcal{R}^{n} are uniformly bounded on HmσH^{\sigma}_{m}

    supn1nfHmσm,σfHmσ.\displaystyle\sup_{n\geq 1}\left\|\mathcal{R}^{n}f\right\|_{H^{\sigma}_{m}}\lesssim_{m,\sigma}\left\|f\right\|_{H^{\sigma}_{m}}.
  • (ii)

    The regularization operators satisfy: for fHmσf\in H^{\sigma}_{m}

    nfHmσnσσn(mm)/2fHmσ.\displaystyle\left\|\mathcal{R}^{n}f\right\|_{H^{\sigma^{\prime}}_{m^{\prime}}}\lesssim n^{\sigma^{\prime}-\sigma}n^{(m^{\prime}-m)/2}\left\|f\right\|_{H^{\sigma}_{m}}.
  • (iii)

    The regularization operators converge in the following senses: for fHmσf\in H^{\sigma}_{m} there holds

    limnnffHmσ=0\displaystyle\lim_{n\to\infty}\left\|\mathcal{R}^{n}f-f\right\|_{H^{\sigma}_{m}}=0 (4.1)
    limnnnffHmσ1=0.\displaystyle\lim_{n\to\infty}n\left\|\mathcal{R}^{n}f-f\right\|_{H^{\sigma-1}_{m}}=0. (4.2)

In the previous section, we showed that each of the f0nf_{0}^{n} generates a maximal solution fnf^{n} of (1.1) in C([0,τn);Hmσ)Lloc([0,τn);Hm1σ1)C([0,\tau_{n});H_{m}^{\sigma})\cap L_{\text{loc}}^{\infty}([0,\tau_{n});H_{m^{\prime}-1}^{\sigma^{\prime}-1}), where τn\tau_{n} is the maximal time of existence of fnf^{n}. We now show that the sequence (fn)n=1(f^{n})_{n=1}^{\infty} of approximate solutions has a strongly convergent subsequence.

We start by defining the stopping times:

τnT:=inf{t0:fn(t)Hmσ>f0nHmσ+2}T,\displaystyle\tau_{n}^{T}:=\inf\left\{t\geq 0:\,\|f^{n}(t)\|_{H_{m}^{\sigma}}>\|f_{0}^{n}\|_{H_{m}^{\sigma}}+2\right\}\wedge T, (4.3)
τn,lT=τnTτkT\displaystyle\tau_{n,l}^{T}=\tau_{n}^{T}\wedge\tau_{k}^{T} (4.4)

The following is similar to [mikulevicius2004stochastic]*Lemma 37 or [GV14]*Lemma 7.1:

Lemma 4.2.

Let τn\tau_{n} be a sequence of stopping times and suppose that a sequence of predictable processes fnC([0,τn];Hmσ)f^{n}\in C([0,\tau_{n}];H_{m}^{\sigma}) satisfy:

limnsupln𝐄suptτn,lTfnflHmσ2=0,\lim_{n\to\infty}\sup_{l\geq n}\mathbf{E}\sup_{t^{\prime}\leq\tau_{n,l}^{T}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}=0, (4.5)
limϵ0supn1𝐏[suptτnTϵfnHmσ>f0nHmσ+1]=0.\lim_{\epsilon\to 0}\sup_{n\geq 1}\mathbf{P}\left[\sup_{t\leq\tau_{n}^{T}\wedge\epsilon}\|f^{n}\|_{H_{m}^{\sigma}}>\|f_{0}^{n}\|_{H_{m}^{\sigma}}+1\right]=0. (4.6)

Then, there exists a stopping time τ\tau with 𝐏(0<τT)=1,\mathbf{P}(0<\tau\leq T)=1, a predictable process fC([0,τ];Hmσ),f\in C([0,\tau];H_{m}^{\sigma}), and a subsequence (fnj)j=1(f^{n_{j}})_{j=1}^{\infty} of (fn)n=1(f^{n})_{n=1}^{\infty} such that

suptτfnj(t)f(t)Hmσ0 as j, 𝐏–a.s.\sup_{t\leq\tau}\|f^{n_{j}}(t)-f(t)\|_{H_{m}^{\sigma}}\to 0\text{ as }j\to\infty,\text{ }\mathbf{P}\text{--a.s.} (4.7)

and

suptτf(t)Hmσ2+supnf0nHmσ, 𝐏–a.s.\sup_{t\leq\tau}\|f(t)\|_{H_{m}^{\sigma}}\leq 2+\sup_{n}\|f_{0}^{n}\|_{H_{m}^{\sigma}},\text{ }\mathbf{P}\text{--a.s.} (4.8)

We now verify that the regularized solutions {fn}n1\left\{f^{n}\right\}_{n\geq 1} satisfy the conditions of Lemma 4.2.

Lemma 4.3.

The solutions (fn)n=1(f^{n})_{n=1}^{\infty} generated by the regularized data (f0n)n=1(f_{0}^{n})_{n=1}^{\infty} satisfy (4.5) and (4.6).

Proof.

We begin with proving that the Cauchy property (4.5) holds for the sequence fnf^{n} of solutions to (2.3) with initial data f0nf_{0}^{n}. This is done via an energy estimate with some similarities with the uniqueness and convergence proofs in Section 3 with one significant difference. For tτn,lTt\leq\tau_{n,l}^{T}, we have:

dxαvβ(fnfl)Lm22=\displaystyle\mathrm{d}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\|_{L_{m}^{2}}^{2}= 2xαvβ(vx(fnfl)),xαvβ(fnfl)mdt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}(f^{n}-f^{l})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\mathrm{d}t
+2Δv(xαvβ(fnfl)),xαvβ(fnfl)mdt\displaystyle+2\left<\Delta_{v}(\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\mathrm{d}t
+2xαvβdivv((fnfl)v),xαvβ(fnfl)mdt\displaystyle+2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}\operatorname{\mathrm{div}}_{v}((f^{n}-f^{l})v),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\mathrm{d}t
2xαvβ(EnvfnElvfl),xαvβ(fnfl)mdt\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{n}\cdot\nabla_{v}f^{n}-E^{l}\cdot\nabla_{v}f^{l}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\mathrm{d}t
2xαvβ(v(fnfl)dWt),xαvβ(fnfl)m\displaystyle-2\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}
+kxαvβ(σkekv)2(fnfl),xαvβ(fnfl)mdt\displaystyle+\sum_{k}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}(f^{n}-f^{l}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\mathrm{d}t
+kxαvβ(σkekv(fnfl))Lm22dt.\displaystyle+\sum_{k}\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}(f^{n}-f^{l}))\|_{L_{m}^{2}}^{2}\mathrm{d}t. (4.9)

We will control the above for |α|+|β|=σ|\alpha|+|\beta|=\sigma. Of course, the linear terms are treated in the same way as in the estimates of Section 3. We now explain how the electric field terms are to be estimated. For tτn,lt\leq\tau_{n,l}, we have:

|xαvβ(EnvfnElvfl),xαvβ(fnfl)m|\displaystyle\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(E^{n}\cdot\nabla_{v}f^{n}-E^{l}\cdot\nabla_{v}f^{l}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\right|
\displaystyle\leq |xαvβ((EnEl)vfn),xαvβ(fnfl)m|+|xαvβ((Elv(fnfl)),xαvβ(fnfl)m|\displaystyle\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}((E^{n}-E^{l})\cdot\nabla_{v}f^{n}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\right|+\left|\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}((E^{l}\cdot\nabla_{v}(f^{n}-f^{l})),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}\right|
\displaystyle\leq CfnflHmσ1fnHmσ+1fnflHmσ+CflHmσ1fnflHmσ2\displaystyle C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}\|f^{n}\|_{H_{m}^{\sigma+1}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}+C\|f^{l}\|_{H_{m}^{\sigma-1}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}
\displaystyle\leq CfnflHmσ2+fnflHmσ12fnHmσ+12,\displaystyle C\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}+\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}, (4.10)

where we have used Lemma 3.3. Combining our estimates from the previous section with (4.10) and the fact that we are taking tτj,kTt\leq\tau_{j,k}^{T}, we have:

d(fnflHmσ2)\displaystyle\mathrm{d}\left(\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}\right)\leq CfnflHmσ2dt+CfnflHmσ12fnHmσ+12dt\displaystyle C\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}\mathrm{d}t+C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t
2|α|+|β|=σxαvβ(v(fnfl)dWt),xαvβ(fnfl)m\displaystyle-2\sum_{|\alpha|+|\beta|=\sigma}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}
2v(fnfl)dWt,fnflm.\displaystyle-2\left<\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t},f^{n}-f^{l}\right>_{m}. (4.11)

In what follows we denote

σ(fnfl):=\displaystyle\mathcal{M}_{\sigma}(f^{n}-f^{l}):= 2|α|+|β|=σxαvβ(v(fnfl)dWt),xαvβ(fnfl)m\displaystyle-2\sum_{|\alpha|+|\beta|=\sigma}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}
2v(fnfl)dWt,fnflm.\displaystyle\quad-2\left<\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t},f^{n}-f^{l}\right>_{m}. (4.12)

The estimate (4.11) would close similarly to before (i.e. by using BDG and Grönwall’s inequalities), save for the fact that we do not a priori know that the term fnflHmσ12fnHmσ+12\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2} is in Lω1Lt1L_{\omega}^{1}L_{t}^{1}, so we now estimate it separately333This loss of probabilistic moments was addressed by the cutoff in the approximation scheme of Section 3. (compare to [GV14]*Lemma 7.2). The stochastic product rule gives:

d(fnflHmσ12fnHmσ+12)=\displaystyle\mathrm{d}(\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2})= fnHmσ+12dfnflHmσ1+fnflHmσ12dfnHmσ+12\displaystyle\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}+\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\mathrm{d}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}
+(dfnflHmσ12)(dfnHmσ+12).\displaystyle+(\mathrm{d}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2})(\mathrm{d}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}). (4.13)

The correction term in (4.13) is:

(dfnflHmσ12)(dfnHmσ+12)=kBk,σ1(fnfl)Bk,σ+1(fn)(\mathrm{d}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2})(\mathrm{d}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2})=\sum_{k}B_{k,\sigma-1}(f^{n}-f^{l})B_{k,\sigma+1}(f^{n}) (4.14)

where:

Bk,s(h):=σkekv(h),hm+|α|+|β|=sxαvβ(σkekvh),xαvβhm.B_{k,s}(h):=\left<\sigma_{k}e_{k}\cdot\nabla_{v}(h),h\right>_{m}+\sum_{|\alpha|+|\beta|=s}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}h),\partial_{x}^{\alpha}\partial_{v}^{\beta}h\right>_{m}. (4.15)

Note that Bk,s(h)CσkekWs,hHms2,B_{k,s}(h)\leq C\sigma_{k}\|e^{k}\|_{W^{s,\infty}}\|h\|_{H_{m}^{s}}^{2}, so:

d(fnflHmσ12)d(fnHmσ12)CfnflHmσ12fnHmσ+12dt.\mathrm{d}(\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2})\mathrm{d}(\|f^{n}\|_{H_{m}^{\sigma-1}}^{2})\leq C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t. (4.16)

For the main terms of (4.13) we have similar estimates as before. For the difference fnflf^{n}-f^{l} and for tτn,lt\leq\tau_{n,l} we have (recalling the definition (4.12)):

dfnflHmσ12\displaystyle\mathrm{d}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}
CfnflHmσ12dt+CfnflHmσ12fnHmσ2dt\displaystyle\leq C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\mathrm{d}t+C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma}}^{2}\mathrm{d}t
2|α|+|β|=σ1xαvβ[v(fnfl)dWt],xαvβ(fnfl)m\displaystyle-2\sum_{|\alpha|+|\beta|=\sigma-1}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}\left[\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t}\right],\partial_{x}^{\alpha}\partial_{v}^{\beta}(f^{n}-f^{l})\right>_{m}
2v(fnfl)dWt,fnflm\displaystyle-2\left<\nabla_{v}(f^{n}-f^{l})\cdot\mathrm{d}W_{t},f^{n}-f^{l}\right>_{m}
\displaystyle\leq CfnflHmσ12dt+σ1(fnfl),\displaystyle C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\mathrm{d}t+\mathcal{M}_{\sigma-1}(f^{n}-f^{l}), (4.17)

where we used (4.10) for σ1\sigma-1 instead of σ\sigma derivatives and the definition of the stopping time τnT\tau_{n}^{T}. Similarly, for the norm of fnf^{n} and for tτn,lt\leq\tau_{n,l} we have:

dfnHmσ+12\displaystyle\mathrm{d}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\leq CfnHmσ+12dt\displaystyle C\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t
+CfnHmσfnHmσ+12dt\displaystyle+C\|f^{n}\|_{H_{m}^{\sigma}}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t
2|α|+|β|=σ+1xαvβ(vfndWt),xαvβfnm\displaystyle-2\sum_{|\alpha|+|\beta|=\sigma+1}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f^{n}\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{n}\right>_{m}
2vfndWt,fnm\displaystyle-2\left<\nabla_{v}f^{n}\cdot\mathrm{d}W_{t},f^{n}\right>_{m}
\displaystyle\leq CfnHmσ+12dt+σ+1(fn),\displaystyle C\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t+\mathcal{M}_{\sigma+1}(f^{n}), (4.18)

where we again used the definition of the stopping time τnT\tau_{n}^{T}. Now, plugging (4.16), (4.17) (4.18) into (4.13), we obtain:

d(fnflHmσ12fnHmσ+12)CfnflHmσ12fnHmσ+12dt\displaystyle\mathrm{d}(\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2})\leq C\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}t
+σ+1(fn)fnflHmσ12+σ1(fnfl)fnHmσ+12,\displaystyle+\mathcal{M}_{\sigma+1}(f^{n})\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}+\mathcal{M}_{\sigma-1}(f^{n}-f^{l})\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}, (4.19)

Integrating (4.19) in time and using the BDG inequality, we obtain:

𝐄supttτn,l(fnflHmσ12fnHmσ+12)\displaystyle\mathbf{E}\sup_{t^{\prime}\leq t\wedge\tau_{n,l}}\left(\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\right)\leq 𝐄(f0nf0lHmσ12f0nHmσ+12)\displaystyle\mathbf{E}\left(\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f_{0}^{n}\|_{H_{m}^{\sigma+1}}^{2}\right)
+C𝐄0tsuptsτn,lfnflHmσ12fnHmσ+12ds\displaystyle+C\mathbf{E}\int_{0}^{t}\sup_{t^{\prime}\leq s\wedge\tau_{n,l}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\mathrm{d}s
+C𝐄(0tfnflHmσ14fnHmσ+14ds)1/2,\displaystyle+C\mathbf{E}\left(\int_{0}^{t}\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{4}\|f^{n}\|_{H_{m}^{\sigma+1}}^{4}\mathrm{d}s\right)^{1/2},

so after rearranging and using Grönwall’s inequality as done previously in e.g. the proof of Lemma 3.1, we get:

𝐄supttτn,l(fnflHmσ12fnHmσ+12)C𝐄(f0nf0lHmσ12f0nHmσ+12).\mathbf{E}\sup_{t^{\prime}\leq t\wedge\tau_{n,l}}\left(\|f^{n}-f^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f^{n}\|_{H_{m}^{\sigma+1}}^{2}\right)\leq C\mathbf{E}\left(\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f_{0}^{n}\|_{H_{m}^{\sigma+1}}^{2}\right). (4.20)

Now returning to (4.11), integrating in time, using the BDG inequality, plugging in (4.20), we obtain:

𝐄supttτn,lfnflHmσ2\displaystyle\mathbf{E}\sup_{t^{\prime}\leq t\wedge\tau_{n,l}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}\leq 𝐄f0nf0lHmσ2\displaystyle\mathbf{E}\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma}}^{2}
+C0t𝐄suptsτn,lfnflHmσ2ds\displaystyle+C\int_{0}^{t}\mathbf{E}\sup_{t^{\prime}\leq s\wedge\tau_{n,l}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}\mathrm{d}s
+C𝐄(0tτn,lfnflHmσ4ds)12\displaystyle+C\mathbf{E}\left(\int_{0}^{t\wedge\tau_{n,l}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{4}\mathrm{d}s\right)^{\frac{1}{2}}
+C𝐄(f0nf0lHmσ12f0nHmσ+12).\displaystyle+C\mathbf{E}\left(\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f_{0}^{n}\|_{H_{m}^{\sigma+1}}^{2}\right).

Therefore we have

𝐄suptτn,lfnflHmσ2\displaystyle\mathbf{E}\sup_{t^{\prime}\leq\tau_{n,l}}\|f^{n}-f^{l}\|_{H_{m}^{\sigma}}^{2}\leq C𝐄f0nf0lHmσ2\displaystyle C\mathbf{E}\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma}}^{2}
+C𝐄(f0nf0lHmσ12f0nHmσ+12).\displaystyle+C\mathbf{E}\left(\|f_{0}^{n}-f_{0}^{l}\|_{H_{m}^{\sigma-1}}^{2}\|f_{0}^{n}\|_{H_{m}^{\sigma+1}}^{2}\right). (4.21)

Then (4.5) follows from (4.21) and Lemma 4.1 (in particular, note (4.2)).

Next, we move to the proof of (4.6). By Itô’s formula, we have:

dfnHmσ2\displaystyle\mathrm{d}\|f^{n}\|_{H_{m}^{\sigma}}^{2}\leq CfnHmσ2dt+EnHσ2dt\displaystyle C\|f^{n}\|_{H_{m}^{\sigma}}^{2}\mathrm{d}t+\|E^{n}\|_{H^{\sigma}}^{2}\mathrm{d}t
2|α|+|β|=σxαvβ(vfndWt),xαvβfnm\displaystyle-2\sum_{|\alpha|+|\beta|=\sigma}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f^{n}\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{n}\right>_{m}
2vfndWt,fnm.\displaystyle-2\left<\nabla_{v}f^{n}\cdot\mathrm{d}W_{t},f^{n}\right>_{m}. (4.22)

Let us denote

σ(fn):=2|α|+|β|=σxαvβ(vfn)dWt,xαvβfnm+2vfndWt,fnm\mathcal{M}_{\sigma}(f^{n}):=2\sum_{|\alpha|+|\beta|=\sigma}\left<\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f^{n})\cdot\mathrm{d}W_{t},\partial_{x}^{\alpha}\partial_{v}^{\beta}f^{n}\right>_{m}+2\left<\nabla_{v}f^{n}\cdot\mathrm{d}W_{t},f^{n}\right>_{m}

so that for tτnTϵt\leq\tau_{n}^{T}\wedge\epsilon, after integrating in time, (4.22) gives:

fn(t)Hmσ2f0nHmσ2+C0tfn(s)Hmσ2ds+0tn(s).\displaystyle\|f^{n}(t)\|_{H_{m}^{\sigma}}^{2}\leq\|f_{0}^{n}\|_{H_{m}^{\sigma}}^{2}+C\int_{0}^{t}\|f^{n}(s)\|_{H_{m}^{\sigma}}^{2}\mathrm{d}s+\int_{0}^{t}\mathcal{M}^{n}(s). (4.23)

Therefore, by Chebyshev’s inequality for the usual deterministic integral and Doob’s ienquality for the martingale, we have

𝐏(suptτnTϵfnHmσ2>f0nHmσ2+1)\displaystyle\mathbf{P}(\sup_{t^{\prime}\leq\tau_{n}^{T}\wedge\epsilon}\|f^{n}\|_{H_{m}^{\sigma}}^{2}>\|f_{0}^{n}\|_{H_{m}^{\sigma}}^{2}+1)\leq 𝐏(C0τnTϵfn(s)Hmσ2ds>12)\displaystyle\mathbf{P}\left(C\int_{0}^{\tau_{n}^{T}\wedge\epsilon}\|f^{n}(s)\|_{H_{m}^{\sigma}}^{2}\mathrm{d}s>\frac{1}{2}\right)
+\displaystyle+ 𝐏(suptτnTϵ|0tn(s)|>12)\displaystyle\mathbf{P}\left(\sup_{t^{\prime}\leq\tau_{n}^{T}\wedge\epsilon}\left|\int_{0}^{t^{\prime}}\mathcal{M}^{n}(s)\right|>\frac{1}{2}\right)
\displaystyle\leq C𝐄0τnTϵfn(s)Hmσ2ds\displaystyle C\mathbf{E}\int_{0}^{\tau_{n}^{T}\wedge\epsilon}\|f^{n}(s)\|_{H_{m}^{\sigma}}^{2}\mathrm{d}s
+C𝐄0τnTϵfn(s)Hmσ4ds\displaystyle+C\mathbf{E}\int_{0}^{\tau_{n}^{T}\wedge\epsilon}\|f^{n}(s)\|_{H_{m}^{\sigma}}^{4}\mathrm{d}s
\displaystyle\leq Cϵ𝐄f0nHmσ2\displaystyle C\epsilon\mathbf{E}\|f_{0}^{n}\|_{H_{m}^{\sigma}}^{2}
\displaystyle\leq CMϵ;\displaystyle CM\epsilon; (4.24)

note we also used that tτnTt\leq\tau_{n}^{T} implies fn(t)HmσC\|f^{n}(t)\|_{H_{m}^{\sigma}}\leq C for a constant C>0C>0 that depends on the size of the initial data f0f_{0} uniformly in nn, since f0nHmσCf0Hmσ\|f_{0}^{n}\|_{H_{m}^{\sigma}}\leq C\|f_{0}\|_{H_{m}^{\sigma}} independently of nn. Taking ϵ0\epsilon\to 0, we obtain (4.6). ∎

Combining Lemmas 4.2 and 4.3, we obtain the existence of a local strong solutions to (1.1) when f0HmσM<\|f_{0}\|_{H_{m}^{\sigma}}\leq M<\infty almost surely. A splitting of the general random initial condition similar to the one in Lemma 2.2 can now provide a local solution whenever f0f_{0} is 0\mathcal{F}_{0}-measurable with f0Hmσ<\|f_{0}\|_{H_{m}^{\sigma}}<\infty 𝐏\mathbf{P}–a.s.. Specifically, since f0=M=0𝟙Mf0Hmσ<M+1f0,f_{0}=\sum_{M=0}^{\infty}\mathbbm{1}_{M\leq\|f_{0}\|_{H_{m}^{\sigma}}<M+1}f_{0}, each component f0,M:=𝟙Mf0Hmσ<M+1f0f_{0,M}:=\mathbbm{1}_{M\leq\|f_{0}\|_{H_{m}^{\sigma}}<M+1}f_{0} generates a local strong solution (fM,τM)(f_{M},\tau_{M}) to (1.1) and we re-construct the full ff and τ\tau using

f=M=0𝟙Mf0Hmσ<M+1fM,f=\sum_{M=0}^{\infty}\mathbbm{1}_{M\leq\|f_{0}\|_{H_{m}^{\sigma}}<M+1}f_{M}, (4.25)

and

τ=M=0𝟙Mf0Hmσ<M+1τM.\tau=\sum_{M=0}^{\infty}\mathbbm{1}_{M\leq\|f_{0}\|_{H_{m}^{\sigma}}<M+1}\tau_{M}. (4.26)

This completes the proof of Theorem 1.3.

5 Hypoelliptic regularization for Vlasov-Poisson-Fokker-Planck

Theorem 1.9 follows by a priori regularization estimates of (2.3), specifically, it suffices to prove that solutions to (2.3) are almost-surely Cx,vC^{\infty}_{x,v} for t>0t>0.

We first prove that if f0Hmσf_{0}\in H^{\sigma}_{m}, then the solution lies in f(t)Hmσ+1f(t)\in H^{\sigma+1}_{m} for t>0t>0 (with size depending only on the HσH^{\sigma} norm of the initial condition). As mentioned in Section 1, this hypoelliptic regularization is proved using a time-weighted variation of the classical hypocoercive energy functional for the kinetic Fokker–Planck equation (see [dric2009hypocoercivity]). For the linear case, a related hypoelliptic regularization estimate can be found in [de2018invariant]. Taking the standard energy from [dric2009hypocoercivity] and scaling derivatives with the powers of tt expected from known hypoelliptic regularization estimates (alternatively, one can deduce them from scaling arguments; see e.g. [bouchut1993existence]) we have

1[t,f]=f(t)Lm22+atvf(t)Lm22+bt2vf(t),xf(t)m+ct3xf(t)Lm22.\displaystyle\mathcal{E}_{1}[t,f]=\left\|f(t)\right\|_{L_{m}^{2}}^{2}+at\left\|\nabla_{v}f(t)\right\|_{L_{m}^{2}}^{2}+bt^{2}\left\langle\nabla_{v}f(t),\nabla_{x}f(t)\right\rangle_{m}+ct^{3}\left\|\nabla_{x}f(t)\right\|_{L_{m}^{2}}^{2}.

For HmσH^{\sigma}_{m} estimates we hence define

σ[t,f]:=0qσ1,m[t,xσqvqf].\mathcal{E}_{\sigma}[t,f]:=\sum_{0\leq q\leq\sigma}\mathcal{E}_{1,m}[t,\nabla_{x}^{\sigma-q}\nabla_{v}^{q}f]. (5.1)

The constants are chosen (indepedent of σ\sigma) such that 0<c<b<a0<c<b<a and b2<acb^{2}<\sqrt{ac} so that

σfHmσ+tvfHmσ2+t3xfHmσ2.\displaystyle\mathcal{E}_{\sigma}\approx\left\|f\right\|_{H^{\sigma}_{m}}+t\left\|\nabla_{v}f\right\|_{H^{\sigma}_{m}}^{2}+t^{3}\left\|\nabla_{x}f\right\|_{H^{\sigma}_{m}}^{2}.

The parameters a,b,ca,b,c are chosen more specifically to satisfy: for some sufficiently small ε1\varepsilon\ll 1 we require

{1aεbε2cε3aε1b,bεac.\begin{cases}1\geq\frac{a}{\varepsilon}\geq\frac{b}{\varepsilon^{2}}\geq\frac{c}{\varepsilon^{3}}\\ a\leq\varepsilon\sqrt{1\cdot b},\quad b\leq\varepsilon\sqrt{a\cdot c}.\end{cases} (5.2)

We recall the proof that such a,b,ca,b,c exist in Lemma 5.2 below. Note that these conditions imply bt2εat+εct3bt^{2}\leq\varepsilon at+\varepsilon ct^{3}, a fact we use repeatedly below.

Hence, an estimate on σ\mathcal{E}_{\sigma} in terms of f0Hmσ\left\|f_{0}\right\|_{H^{\sigma}_{m}} implies the desired regularization estimates (along with some more quantitative information that we will not directly use here). The main result of this Section is the following.

Proposition 5.1.

Let f0f_{0} be a 0\mathcal{F}_{0}-measurable initial data and suppose that for all p<p<\infty, Mp>0\exists M_{p}>0 such that for some σ>d2+1\sigma>\frac{d}{2}+1 we have:

𝐄f0Hmσp<Mp.\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H^{\sigma}_{m}}^{p}<M_{p}. (5.3)

Let R<R<\infty, let ff be the unique pathwise solution of (2.3).

Then T>0\exists T>0 depending only on σ\sigma such that there holds for all p<p<\infty,

𝐄(sup0<t<Tσ[t,f(t)])pC(R,p,M2,M3,).\displaystyle\mathbf{E}\left(\sup_{0<t<T}\mathcal{E}_{\sigma}[t,f(t)]\right)^{p}\leq C(R,p,M_{2},M_{3},\ldots).

Therefore, almost surely f(t)Hmσ+1f(t)\in H^{\sigma+1}_{m} for all 0<t<T0<t<T.

Proposition 5.1 implies a corresponding instantaneous regularization for the maximal pathwise solution of (1.1). Once the above proposition is proved, one may simply iterate it, observing that for all δ>0\delta>0, f(δ)f(\delta) is an δ\mathcal{F}_{\delta}-measurable random variable with

𝐄f(δ)Hmσ+1p<.\displaystyle\mathbf{E}\left\|f(\delta)\right\|_{H^{\sigma+1}_{m}}^{p}<\infty.

Therefore, we may apply Proposition 5.1 to the initial data f(δ)f(\delta) with σσ+1\sigma\mapsto\sigma+1. Finally, similar to the proof of Lemma 2.2, a simple cutting procedure can be applied to remove the moment constraint on the initial condition. Hence, to prove Theorem 1.9, it suffices to prove Proposition 5.1.

Proof.

For notational simplicity, we will take ν=1\nu=1 but the same arguments (up to a suitable rescaling of the coefficients a,b,ca,b,c) apply for any ν>0\nu>0. Define the dissipation rate:

𝔻σ(t,f(t))\displaystyle\mathbb{D}_{\sigma}(t,f(t)) :=|α|+|β|σ(vxαvβf(t)Lm22+atv2xαvβf(t)Lm22\displaystyle:=\sum_{\left|\alpha\right|+\left|\beta\right|\leq\sigma}\Bigl{(}\left\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}+at\left\|\nabla_{v}^{2}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}
+b2t2xxαvβf(t)Lm22+ct3vxxαvβf(t)Lm22),\displaystyle\quad+\frac{b}{2}t^{2}\left\|\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}+ct^{3}\left\|\nabla_{v}\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}\Bigr{)}, (5.4)

which we show arises from dσ\mathrm{d}\mathcal{E}_{\sigma}. Note that this is almost the same as the contribution from dσ\mathrm{d}\mathcal{E}_{\sigma} that arises when the time derivative lands on the powers of tt in front of the higher-order terms. In order to reduce some of the notation in the ensuing calculation, we use B(h,g)B(h,g) to denote an Lm2L^{2}_{m}-bounded bilinear form, the exact form of which is irrelevant, i.e, a form which is linear in both arguments and such that for any h,gLm2h,g\in L^{2}_{m}

B(h,g)Lm2hLm2gLm2.\displaystyle\left\|B(h,g)\right\|_{L^{2}_{m}}\lesssim\left\|h\right\|_{L^{2}_{m}}\left\|g\right\|_{L^{2}_{m}}.

The main step of the proof is to calculate the following

dσ\displaystyle\mathrm{d}\mathcal{E}_{\sigma} =|α|+|β|σdxαvβf(t)Lm22+d(atvxαvβf(t)Lm22)\displaystyle=\sum_{\left|\alpha\right|+\left|\beta\right|\leq\sigma}\mathrm{d}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}+\mathrm{d}\left(at\left\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}\right)
+|α|+|β|σd(bt2vxαvβf(t),xxαvβf(t)m)+d(ct3xxαvβf(t)Lm22).\displaystyle\quad+\sum_{\left|\alpha\right|+\left|\beta\right|\leq\sigma}\mathrm{d}\left(bt^{2}\left\langle\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t),\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\rangle_{m}\right)+\mathrm{d}\left(ct^{3}\left\|\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f(t)\right\|_{L_{m}^{2}}^{2}\right).

As in the proof of the various bounds in Sections 34, we have:

dxαvβfLm22=\displaystyle\mathrm{d}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{2}= 2xαvβ(vxf),xαvβfmdt\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
+2Δvxαvβf,xαvβfmdt\displaystyle+2\left\langle\Delta_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
+2xαvβdivv(fv),xαvβfmdt\displaystyle+2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}\operatorname{\mathrm{div}}_{v}(fv),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
2xαvβ(θR(fHm0s0)Evf),xαvβfmdt\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}(\left\|f\right\|_{H_{m_{0}}^{s_{0}}})E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
2xαvβ(vfdWt),xαvβfm\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
+kxαvβ((σkekv)2f),xαvβfmdt\displaystyle+\sum_{k}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}((\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
kxαvβ(σkekvf)Lm22dt.\displaystyle-\sum_{k}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\|_{L_{m}^{2}}^{2}\mathrm{d}t. (5.5)

This formula, and its straightforward variations, are then used to expand most of the terms of dσ\mathrm{d}\mathcal{E}_{\sigma}, with the exception of the cross-terms (i.e. those multiplied by bb). For the cross-terms we instead have

dxαvβ+ejf,xα+ejvβfm=\displaystyle\mathrm{d}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}=
(xαvβ+ej(vxf),xα+ejvβfm+xαvβ+ejf,xα+ejvβ(vxf)m)dt\displaystyle-\left(\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(v\cdot\nabla_{x}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(v\cdot\nabla_{x}f)\right\rangle_{m}\right)\mathrm{d}t
(θRxαvβ+ej(Evf),xα+ejvβfm+xαvβ+ejf,θRxα+ejvβ(Evf)m)dt\displaystyle-\left(\left\langle\theta_{R}\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\theta_{R}\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(E\cdot\nabla_{v}f)\right\rangle_{m}\right)\mathrm{d}t
+(xαvβ+ej(Δvf),xα+ejvβfm+xαvβ+ejf,xα+ejvβ(Δvf)m)dt\displaystyle+\left(\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\Delta_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\Delta_{v}f)\right\rangle_{m}\right)\mathrm{d}t
+(xαvβ+ej(divv(fv)),xα+ejvβfm+xαvβ+ejf,xα+ejvβ(divv(fv))m)dt\displaystyle+\left(\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\operatorname{\mathrm{div}}_{v}(fv)),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\operatorname{\mathrm{div}}_{v}(fv))\right\rangle_{m}\right)\mathrm{d}t
xαvβ+ej(vfdWt),xα+ejvβfmxαvβ+ejf,xα+ejvβ(vfdWt)m\displaystyle-\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}-\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t})\right\rangle_{m}
+12kxαvβ+ej(σkekv)2f,xα+ejvβfmdt\displaystyle+\frac{1}{2}\sum_{k}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t
+12kxαvβ+ejf,xα+ejvβ(σkekv)2fmdt\displaystyle+\frac{1}{2}\sum_{k}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f\right\rangle_{m}\mathrm{d}t
+xαvβ+ej(σkekvf),xα+ejvβ(σkekvf)mdt\displaystyle+\sum\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\sigma_{k}e_{k}\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\rangle_{m}\mathrm{d}t
:=\displaystyle:= 𝒯c,α,β,j(f)+𝒩c,α,β,j(f)+𝒟c,α,β,j(f)+c,α,β,j(f)+c,α,β,j(f)+𝒞c,α,β,j(f),\displaystyle\mathcal{T}_{c,\alpha,\beta,j}(f)+\mathcal{N}_{c,\alpha,\beta,j}(f)+\mathcal{D}_{c,\alpha,\beta,j}(f)+\mathcal{F}_{c,\alpha,\beta,j}(f)+\mathcal{M}_{c,\alpha,\beta,j}(f)+\mathcal{C}_{c,\alpha,\beta,j}(f), (5.6)

where we abbreviated θR:=θR(fHm0s0)\theta_{R}:=\theta_{R}(\left\|f\right\|_{H_{m_{0}}^{s_{0}}}) and 𝒯c,α,β,j,𝒩c,α,β,j,𝒟c,α,β,j,c,α,β,j,c,α,β,j,𝒞c,α,β,j\mathcal{T}_{c,\alpha,\beta,j},\mathcal{N}_{c,\alpha,\beta,j},\mathcal{D}_{c,\alpha,\beta,j},\mathcal{F}_{c,\alpha,\beta,j},\mathcal{M}_{c,\alpha,\beta,j},\mathcal{C}_{c,\alpha,\beta,j} indicate transport, nonlinear (electric field), dissipation, friction, martingale, and (Itô) correction contributions (which incorporate the last three terms) to the cross-terms, respectively.

Linear, deterministic contributions:

First, we collect the contributions of the linear terms, namely those that arise from the vxv\cdot\nabla_{x} free transport and the Fokker–Planck operator. The main effect of these terms is to introduce the dissipation 𝔻σ\mathbb{D}_{\sigma}. The calculation is standard (see [dric2009hypocoercivity]) and so we omit most of the details. We define the total contribution of the linear terms of the SPDE for ff to d\mathrm{d}\mathcal{E} by:

Lin(α,β)\displaystyle Lin(\alpha,\beta) :=avxαvβfm2+2btvxαvβf,xxαvβfm+3ct2xxαvβfLm22\displaystyle:=a\left\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\|_{m}^{2}+2bt\left\langle\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f,\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}+3ct^{2}\left\|\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\|_{L^{2}_{m}}^{2}
2xαvβ(vxf),xαvβfm+2Δvxαvβf,xαvβfm\displaystyle\quad-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}+2\left\langle\Delta_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
+2xαvβdivv(fv),xαvβfm\displaystyle\quad+2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}\operatorname{\mathrm{div}}_{v}(fv),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
2atvxαvβ(vxf),vxαvβfm+2atvxαvβΔvf,vxαvβfm\displaystyle\quad-2at\left\langle\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}+2at\left\langle\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}\Delta_{v}f,\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
+2atxαvβvdivv(fv),vxαvβfm\displaystyle\quad+2at\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}\nabla_{v}\operatorname{\mathrm{div}}_{v}(fv),\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
bt2(j𝒯c,α,β,j+𝒟c,α,β,j+c,α,β,j)\displaystyle\quad-bt^{2}\left(\sum_{j}\mathcal{T}_{c,\alpha,\beta,j}+\mathcal{D}_{c,\alpha,\beta,j}+\mathcal{F}_{c,\alpha,\beta,j}\right)
2ct3xxαvβ(vxf),xxαvβfm\displaystyle\quad-2ct^{3}\left\langle\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
+2ct3Δxα+ejvβf,xα+ejvβfm+2ct3xα+ejvβ(divv(fv)),xα+ejvβf\displaystyle\quad+2ct^{3}\left\langle\Delta\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+2ct^{3}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\operatorname{\mathrm{div}}_{v}(fv)),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle (5.7)

For the rest of this proof, we denote p:=|α|,q:=|β|p:=|\alpha|,q:=|\beta|. By integration by parts we may write

𝒯c,α,β,j(f)=\displaystyle\mathcal{T}_{c,\alpha,\beta,j}(f)= xα+ejvβ(vxf),xαvβ+ejfmdtxα+ejvβf,xαvβ+ej(vxf)mdt\displaystyle-\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(v\cdot\nabla_{x}f),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}\mathrm{d}t-\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(v\cdot\nabla_{x}f)\right\rangle_{m}\mathrm{d}t
=\displaystyle= xα+ejvβfLm22dt\displaystyle-\left\|\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{2}\mathrm{d}t
vx(xα+ejvβf),xαvβ+ejfmdtxα+ejvβf,vx(xαvβ+ejf)m\displaystyle-\left\langle v\cdot\nabla_{x}(\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}\mathrm{d}t-\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,v\cdot\nabla_{x}(\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f)\right\rangle_{m}
+β<β|β|=1(ββ)(xα+ej+βvββf,xαvβ+ejfm+xα+ejvβf,xα+βvββ+ejfm)dt\displaystyle+\sum_{\begin{subarray}{c}\beta^{\prime}<\beta\\ |\beta^{\prime}|=1\end{subarray}}\begin{pmatrix}\beta\\ \beta^{\prime}\end{pmatrix}\left(\left\langle\partial_{x}^{\alpha+e_{j}+\beta^{\prime}}\partial_{v}^{\beta-\beta^{\prime}}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha+\beta^{\prime}}\partial_{v}^{\beta-\beta^{\prime}+e_{j}}f\right\rangle_{m}\right)\mathrm{d}t
=\displaystyle= {xα+ejfLm22dt, if β=0xα+ejvβfLm22dt+q=q1qB(xp+1vqf,xp+1vqf)dt if |β|>0,\displaystyle\begin{cases}-\left\|\partial_{x}^{\alpha+e_{j}}f\right\|_{L_{m}^{2}}^{2}\mathrm{d}t,&\text{ if }\beta=0\\ -\left\|\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{2}\mathrm{d}t+\sum_{q^{\prime}=q-1}^{q}B(\nabla_{x}^{p+1}\nabla_{v}^{q^{\prime}}f,\nabla_{x}^{p+1}\nabla_{v}^{q}f)\mathrm{d}t&\text{ if }|\beta|>0,\end{cases} (5.8)

where recall from above that B(,)B(\cdot,\cdot) denotes an Lm2L^{2}_{m} bounded bilinear form, the exact form of which is not relevant. The dissipation term is more easily treated, yielding

𝒟c,α,β(f)=\displaystyle\mathcal{D}_{c,\alpha,\beta}(f)= xαvβ+ejΔvf,xα+ejvβfmdt+xαvβ+ejf,xα+ejvβΔvfmdt\displaystyle\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}\Delta_{v}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}\Delta_{v}f\right\rangle_{m}\mathrm{d}t
=\displaystyle= 2xαvβ+ejvf,xα+ejvβvfmdt\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}\nabla_{v}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}\nabla_{v}f\right\rangle_{m}\mathrm{d}t
+B(xpvq+2f,xp+1vqf)dt+B(xpvq+1f,xp+1vq+1f)dt.\displaystyle+B(\nabla_{x}^{p}\nabla_{v}^{q+2}f,\nabla_{x}^{p+1}\nabla_{v}^{q}f)\mathrm{d}t+B(\nabla_{x}^{p}\nabla_{v}^{q+1}f,\nabla_{x}^{p+1}\nabla_{v}^{q+1}f)\mathrm{d}t. (5.9)

The friction term can be re-arranged as follows

c,α,β(f)=\displaystyle\mathcal{F}_{c,\alpha,\beta}(f)= xαvβ+ej(divv(fv)),xα+ejvβfmdt+xαvβ+ejf,xα+ejvβ(divv(fv))mdt\displaystyle\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\operatorname{\mathrm{div}}_{v}(fv)),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\mathrm{d}t+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\operatorname{\mathrm{div}}_{v}(fv))\right\rangle_{m}\mathrm{d}t
=\displaystyle= divv(xα+ejvβfv),xαvβ+ejfmdt+xα+ejvβf,divv(xαvβ+ejfv)mdt\displaystyle\left\langle\operatorname{\mathrm{div}}_{v}(\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}fv),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}\mathrm{d}t+\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\operatorname{\mathrm{div}}_{v}(\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}fv)\right\rangle_{m}\mathrm{d}t
+β<β|β|=1xα+ejvβ+βf,xαvβ+ejfm+β<β|β|=1xα+ejvβf,xαvβ+β+ejfmdt\displaystyle+\sum_{\begin{subarray}{c}\beta^{\prime}<\beta\\ |\beta^{\prime}|=1\end{subarray}}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta+\beta^{\prime}}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}+\sum_{\begin{subarray}{c}\beta^{\prime}<\beta\\ |\beta^{\prime}|=1\end{subarray}}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+\beta^{\prime}+e_{j}}f\right\rangle_{m}\mathrm{d}t
=\displaystyle= 2divv(xα+ejvβfxαvβ+ejfv),1mdtvv(xα+ejvβfxαvβ+ejf),1mdt\displaystyle 2\left\langle\operatorname{\mathrm{div}}_{v}(\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}fv),1\right\rangle_{m}\mathrm{d}t-\left\langle v\cdot\nabla_{v}(\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f),1\right\rangle_{m}\mathrm{d}t
+\displaystyle+ β<β|β|=1xα+ejvβ+βf,xαvβ+ejfm+β<β|β|=1xα+ejvβf,xαvβ+β+ejfmdt.\displaystyle\sum_{\begin{subarray}{c}\beta^{\prime}<\beta\\ |\beta^{\prime}|=1\end{subarray}}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta+\beta^{\prime}}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}+\sum_{\begin{subarray}{c}\beta^{\prime}<\beta\\ |\beta^{\prime}|=1\end{subarray}}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+\beta^{\prime}+e_{j}}f\right\rangle_{m}\mathrm{d}t. (5.10)

The fundamental structure of the hypocoercive norm \mathcal{E} is that the 𝒯\mathcal{T} term gives rise to the x\nabla_{x} dissipation term that would otherwise be missing from the dissipation of a kinetic equation. That is, from (5.8), we obtain:

bt2𝒯c,α,β,j(f)+bt2xα+ejvβfLm22dt\displaystyle bt^{2}\mathcal{T}_{c,\alpha,\beta,j}(f)+bt^{2}\left\|\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{2}\mathrm{d}t\leq bt2B(v(xp+1vq1f),v(xp+1vq1f))dt\displaystyle bt^{2}B\left(\nabla_{v}(\nabla_{x}^{p+1}\nabla_{v}^{q-1}f),\nabla_{v}(\nabla_{x}^{p+1}\nabla_{v}^{q-1}f)\right)\mathrm{d}t
+bt2B(xp+1vq1f,v(xp+1vq1f))dt\displaystyle+bt^{2}B\left(\nabla_{x}^{p+1}\nabla_{v}^{q-1}f,\nabla_{v}(\nabla_{x}^{p+1}\nabla_{v}^{q-1}f)\right)\mathrm{d}t
\displaystyle\lesssim bat𝔻σdt+bt2σ\displaystyle\frac{b}{a}t\mathbb{D}_{\sigma}\mathrm{d}t+bt^{2}\mathcal{E}_{\sigma}
\displaystyle\lesssim εt𝔻σdt+bt2σdt,\displaystyle\varepsilon t\mathbb{D}_{\sigma}\mathrm{d}t+bt^{2}\mathcal{E}_{\sigma}\mathrm{d}t, (5.11)

and similarly, from (5.9) and (5.10):

bt2𝒟c,α,β,j+bt2c,α,β,j\displaystyle bt^{2}\mathcal{D}_{c,\alpha,\beta,j}+bt^{2}\mathcal{F}_{c,\alpha,\beta,j}\lesssim (ba+bac)t𝔻σdt+σdt\displaystyle\left(\frac{b}{a}+\frac{ba}{c}\right)t\mathbb{D}_{\sigma}\mathrm{d}t+\mathcal{E}_{\sigma}\mathrm{d}t
\displaystyle\lesssim t𝔻σdt+σdt.\displaystyle t\mathbb{D}_{\sigma}\mathrm{d}t+\mathcal{E}_{\sigma}\mathrm{d}t. (5.12)

Putting together the negative definite terms that arise from 𝒯c,α,β,j\mathcal{T}_{c,\alpha,\beta,j} and those in (5.7), we obtain for short tt and for some C>0C>0:

Lin(α,β)(2Ct)𝔻σ+Cσ.\displaystyle Lin(\alpha,\beta)\leq-\left(2-Ct\right)\mathbb{D}_{\sigma}+C\mathcal{E}_{\sigma}.

In fact this is somewhat sub-optimal, as the second term on the right-hand side above can be taken in weaker norms. However, such refinements will be irrelevant here as we are only interested in short time regularization.

Nonlinear contributions:

Next, we collect the contributions of the nonlinear electric field. Namely,

NL(α,β)\displaystyle NL(\alpha,\beta) :=2xαvβ(θR(fHm0s0)Evf),xαvβfm\displaystyle:=-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\theta_{R}(\left\|f\right\|_{H_{m_{0}}^{s_{0}}})E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
2atxαvβv(θR(fHm0s0)Evf),xαvβvfm\displaystyle\quad-2at\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}\nabla_{v}(\theta_{R}(\left\|f\right\|_{H_{m_{0}}^{s_{0}}})E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}\nabla_{v}f\right\rangle_{m}
bt2(θRxαvβ+ej(Evf),xα+ejvβfm+xαvβ+ejf,θRxα+ejvβ(Evf)m)\displaystyle\quad-bt^{2}\left(\left\langle\theta_{R}\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\theta_{R}\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(E\cdot\nabla_{v}f)\right\rangle_{m}\right)
2ct3xαvβx(θR(fHm0s0)Evf),xαvβxfm.\displaystyle\quad-2ct^{3}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}\nabla_{x}(\theta_{R}(\left\|f\right\|_{H_{m_{0}}^{s_{0}}})E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}\nabla_{x}f\right\rangle_{m}.

We first analyze the electric field’s contribution to (5.5), similarly to the various nonlinear estimates of Section 3:

|xαvβ(Evf),xαvβfm|\displaystyle\left|\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\right|
γ<α|αγ|2|xαγEvvβxγf,xαvβfm|dt+EW1,fHmσ2\displaystyle\lesssim\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left|\left\langle\partial_{x}^{\alpha-\gamma}E\cdot\nabla_{v}\partial_{v}^{\beta}\partial_{x}^{\gamma}f,\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\right|\mathrm{d}t+\left\|E\right\|_{W^{1,\infty}}\left\|f\right\|_{H_{m}^{\sigma}}^{2}
γ<α|αγ|2x|αγ|ELx2σ1|αγ|1vvβxγfLv,m2Lx2σ1|β|+|γ|xαvβfLm2\displaystyle\leq\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left\|\nabla_{x}^{|\alpha-\gamma|}E\right\|_{L_{x}^{2\frac{\sigma-1}{|\alpha-\gamma|-1}}}\left\|\nabla_{v}\partial_{v}^{\beta}\partial_{x}^{\gamma}f\right\|_{L_{v,m}^{2}L_{x}^{2\frac{\sigma-1}{|\beta|+|\gamma|}}}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}
+EW1,fHmσ2\displaystyle+\left\|E\right\|_{W^{1,\infty}}\left\|f\right\|_{H_{m}^{\sigma}}^{2}
\displaystyle\lesssim γ<α|αγ|2EW1,1θEHσθvvβfLm21ηfHmσ1+η+EW1,fHmσ2,\displaystyle\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left\|E\right\|_{W^{1,\infty}}^{1-\theta}\left\|E\right\|_{H^{\sigma}}^{\theta}\left\|\nabla_{v}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{1-\eta}\left\|f\right\|_{H_{m}^{\sigma}}^{1+\eta}+\left\|E\right\|_{W^{1,\infty}}\left\|f\right\|_{H_{m}^{\sigma}}^{2}, (5.13)

where for each α,β,γ,\alpha,\beta,\gamma, the interpolation index η\eta is given as:

η=\displaystyle\eta= |γ||α|1+d|α|1(12|β|+|γ|2(σ1))\displaystyle\frac{|\gamma|}{|\alpha|-1}+\frac{d}{|\alpha|-1}\left(\frac{1}{2}-\frac{|\beta|+|\gamma|}{2(\sigma-1)}\right)
=\displaystyle= |γ||α|1+d|α|1|αγ|12(σ1),\displaystyle\frac{|\gamma|}{|\alpha|-1}+\frac{d}{|\alpha|-1}\cdot\frac{|\alpha-\gamma|-1}{2(\sigma-1)}, (5.14)

Note that in the above, η<1\eta<1 since d2(σ1)<1\frac{d}{2(\sigma-1)}<1. Therefore,

EW1,1θEHσθvvβfLm21ηfHmσ1+η\displaystyle\left\|E\right\|_{W^{1,\infty}}^{1-\theta}\left\|E\right\|_{H^{\sigma}}^{\theta}\left\|\nabla_{v}\partial_{v}^{\beta}f\right\|_{L_{m}^{2}}^{1-\eta}\left\|f\right\|_{H_{m}^{\sigma}}^{1+\eta} fHmσ2+(EW1,1θEHσθ)21ηfHmσ12\displaystyle\lesssim\left\|f\right\|_{H^{\sigma}_{m}}^{2}+\left(\left\|E\right\|_{W^{1,\infty}}^{1-\theta}\left\|E\right\|_{H^{\sigma}}^{\theta}\right)^{\frac{2}{1-\eta}}\left\|f\right\|_{H_{m}^{\sigma-1}}^{2}
fHmσ2+fHmσ142η1η.\displaystyle\lesssim\left\|f\right\|_{H^{\sigma}_{m}}^{2}+\left\|f\right\|_{H_{m}^{\sigma-1}}^{\frac{4-2\eta}{1-\eta}}. (5.15)

Note that in this term, it was not necessary to make use of the dissipation, as the first term in the final inequality above is controlled by at most HmσH^{\sigma}_{m} and the second term, which is derived from the factor EW1,η1θEHσθvvβfLm21η\left\|E\right\|_{W^{1,\eta}}^{1-\theta}\left\|E\right\|_{H^{\sigma}}^{\theta}\|\nabla_{v}\partial_{v}^{\beta}f\|_{L_{m}^{2}}^{1-\eta}, contains at most σ1\sigma-1 derivatives (since |β|σ2\left|\beta\right|\leq\sigma-2 whenever |αγ|2|\alpha-\gamma|\geq 2).

In a similar manner, the corresponding \saysecond term in the electric field contributions to the xα+ejvβf\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f and xαvβ+ejf\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f terms of the energy will contain at most σ\sigma derivatives. This means that all in all, for short tt we can bound the electric field contributions to (5.5), as well as those to the higher order terms in the definition of σ\mathcal{E}_{\sigma}, from (5.13):

|xαvβ(Evf),xαvβfm|fHmσ1p+σ\displaystyle\left|\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}\right|\lesssim\left\|f\right\|_{H_{m}^{\sigma-1}}^{p}+\mathcal{E}_{\sigma} (5.16)
at|xαvβ+ej(Evf),xαvβ+ejfm|fHmσp+σ\displaystyle at\left|\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}\right|\lesssim\left\|f\right\|_{H_{m}^{\sigma}}^{p}+\mathcal{E}_{\sigma} (5.17)
ct3|xα+ejvβ(Evf),xα+ejvβfm|fHmσp+σ,\displaystyle ct^{3}\left|\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\right|\lesssim\left\|f\right\|_{H_{m}^{\sigma}}^{p}+\mathcal{E}_{\sigma}, (5.18)

for some p>2p>2 fixed. Note that the high power of fHmσ\left\|f\right\|_{H^{\sigma}_{m}} is a priori controlled (in Lω1Lt,locL_{\omega}^{1}L_{t,loc}^{\infty}) by the finite pp-th moment assumptions (5.3).

We now move to estimating the electric field contribution to the cross term. First, we integrate by parts for convenience, in order to \saysymmetrize 𝒩c,α,β,j\mathcal{N}_{c,\alpha,\beta,j} up to a lower order term:

𝒩c,α,β,j=\displaystyle\mathcal{N}_{c,\alpha,\beta,j}= xα+ejvβ(Evf),xαvβ+ejfm+xα+ejvβf,xαvβ+ej(Evf)m\displaystyle\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f)\right\rangle_{m}
=\displaystyle= xαvβ(Evf),xα+ejvβ+ejfm+xα+ejvβf,xαvβ+ej(Evf)m\displaystyle-\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}+\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f)\right\rangle_{m}
=\displaystyle= 2xα+ejvβf,xαvβ+ej(Evf)m\displaystyle\quad 2\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f,\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f)\right\rangle_{m}
+xαvβ(Evf)xα+ejvβfvj(vm)dvdx.\displaystyle+\iint\partial_{x}^{\alpha}\partial_{v}^{\beta}(E\cdot\nabla_{v}f)\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\partial_{v^{j}}(\left<v\right>^{m})\mathrm{d}v\mathrm{d}x. (5.19)

We analyze the first term in the final equality above:

|xαvβ+ej(Evf)xα+ejvβfm|\displaystyle\left|\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f)\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\right|
\displaystyle\lesssim γα|xαγEvvβ+ejxγf,xα+ejvβfm|\displaystyle\sum_{\gamma\leq\alpha}\left|\left\langle\partial_{x}^{\alpha-\gamma}E\cdot\nabla_{v}\partial_{v}^{\beta+e_{j}}\partial_{x}^{\gamma}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\right|
ELxvq+2xpfLm2vqxp+1fLm2\displaystyle\leq\left\|E\right\|_{L_{x}^{\infty}}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p}f\right\|_{L_{m}^{2}}\left\|\nabla_{v}^{q}\nabla_{x}^{p+1}f\right\|_{L_{m}^{2}}
+xELxvq+2xp1fLm2xp+1vqfLm2\displaystyle+\left\|\nabla_{x}E\right\|_{L_{x}^{\infty}}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p-1}f\right\|_{L_{m}^{2}}\left\|\nabla_{x}^{p+1}\nabla_{v}^{q}f\right\|_{L_{m}^{2}}
+γ<α|αγ|2x|αγ|ELx2σ|αγ|1vq+2x|γ|fLv,m2Lx2σ|β|+|γ|vqxp+1fLm2.\displaystyle+\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left\|\nabla_{x}^{|\alpha-\gamma|}E\right\|_{L_{x}^{2\frac{\sigma}{|\alpha-\gamma|-1}}}\left\|\nabla_{v}^{q+2}\nabla_{x}^{|\gamma|}f\right\|_{L_{v,m}^{2}L_{x}^{2\frac{\sigma}{|\beta|+|\gamma|}}}\left\|\nabla_{v}^{q}\nabla_{x}^{p+1}f\right\|_{L_{m}^{2}}. (5.20)

To estimate each term in the summation above, we again interpolate as in the proof of Lemma 3.10;

x|αγ|ELx2σ|αγ|1\displaystyle\left\|\nabla_{x}^{|\alpha-\gamma|}E\right\|_{L_{x}^{2\frac{\sigma}{|\alpha-\gamma|-1}}}\lesssim xELx1|αγ|1σxσ+1ELx2|αγ|1σ,\displaystyle\left\|\nabla_{x}E\right\|_{L_{x}^{\infty}}^{1-\frac{|\alpha-\gamma|-1}{\sigma}}\left\|\nabla_{x}^{\sigma+1}E\right\|_{L_{x}^{2}}^{\frac{|\alpha-\gamma|-1}{\sigma}}, (5.21)
vq+2x|γ|fLv,m2Lx2σ|β|+|γ|\displaystyle\left\|\nabla_{v}^{q+2}\nabla_{x}^{|\gamma|}f\right\|_{L_{v,m}^{2}L_{x}^{2\frac{\sigma}{|\beta|+|\gamma|}}}\lesssim vq+2fLm21ηvq+2xpfLm2η,\displaystyle\left\|\nabla_{v}^{q+2}f\right\|_{L_{m}^{2}}^{1-\eta}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p}f\right\|_{L_{m}^{2}}^{\eta}, (5.22)

where

η=\displaystyle\eta= |γ||α|+d|α|(12|β|+|γ|2σ)\displaystyle\frac{|\gamma|}{|\alpha|}+\frac{d}{|\alpha|}\left(\frac{1}{2}-\frac{|\beta|+|\gamma|}{2\sigma}\right)
=\displaystyle= |γ||α|+d|α||αγ|2σ,\displaystyle\frac{|\gamma|}{|\alpha|}+\frac{d}{|\alpha|}\frac{|\alpha-\gamma|}{2\sigma},

which again satisfies η<1\eta<1 since σ>d/2.\sigma>d/2. With these exponents, plugging (5.21)-(5.22) into (5.20) we get:

|xαvβ+ej(Evf),xα+ejvβfm|\displaystyle\left|\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(E\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}\right|
\displaystyle\lesssim ELxvq+2xpfLm2vqxp+1fLm2+xELxvq+2xp1fLm2xp+1vqfLm2\displaystyle\left\|E\right\|_{L_{x}^{\infty}}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p}f\right\|_{L_{m}^{2}}\left\|\nabla_{v}^{q}\nabla_{x}^{p+1}f\right\|_{L_{m}^{2}}+\left\|\nabla_{x}E\right\|_{L_{x}^{\infty}}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p-1}f\right\|_{L_{m}^{2}}\left\|\nabla_{x}^{p+1}\nabla_{v}^{q}f\right\|_{L_{m}^{2}}
+γ<α|αγ|2xELx|β|+|γ|+1σxσ+1ELx2|αγ|1σvq+2fLm21ηvq+2xpfLm2ηvqxp+1fLm2.\displaystyle+\sum_{\begin{subarray}{c}\gamma<\alpha\\ |\alpha-\gamma|\geq 2\end{subarray}}\left\|\nabla_{x}E\right\|_{L_{x}^{\infty}}^{\frac{|\beta|+|\gamma|+1}{\sigma}}\left\|\nabla_{x}^{\sigma+1}E\right\|_{L_{x}^{2}}^{\frac{|\alpha-\gamma|-1}{\sigma}}\left\|\nabla_{v}^{q+2}f\right\|_{L_{m}^{2}}^{1-\eta}\left\|\nabla_{v}^{q+2}\nabla_{x}^{p}f\right\|_{L_{m}^{2}}^{\eta}\left\|\nabla_{v}^{q}\nabla_{x}^{p+1}f\right\|_{L_{m}^{2}}. (5.23)

The lower order term in (5.19) produces a less significant contribution - as it contains a smaller number of derivatives - and hence we omit the treatment for brevity. Therefore, since |αγ|2|\alpha-\gamma|\geq 2 implies q+2σq+2\leq\sigma, we obtain from (5.23) that C(ε)>0\exists C(\varepsilon)>0 such that

bt2𝒩c,α,β,jε10𝔻σdt+C(ε)σdt+C(ε)f(t)Hmσpdt.\displaystyle bt^{2}\mathcal{N}_{c,\alpha,\beta,j}\leq\frac{\varepsilon}{10}\mathbb{D}_{\sigma}\mathrm{d}t+C(\varepsilon)\mathcal{E}_{\sigma}\mathrm{d}t+C(\varepsilon)\left\|f(t)\right\|_{H_{m}^{\sigma}}^{p}\mathrm{d}t. (5.24)

This completes the required estimates on the electric field.

Itô corrections:

Next, let us analyze the contributions of the corrections to the cross term. This equals:

𝒞c,α,β,j(f)=\displaystyle\mathcal{C}_{c,\alpha,\beta,j}(f)= 12k(xαvβ+ej(σkekv)2f,xα+ejvβfm\displaystyle\frac{1}{2}\sum_{k}\Bigl{(}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}
+xαvβ+ejf,xα+ejvβ(σkekv)2fm)dt\displaystyle+\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f\right\rangle_{m}\Bigr{)}\mathrm{d}t
+kxαvβ+ej(σkekvf),xα+ejvβ(σkekvf)mdt\displaystyle+\sum_{k}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\sigma_{k}e_{k}\cdot\nabla_{v}f),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\rangle_{m}\mathrm{d}t
=\displaystyle= pp(B(xpvq+2f,xp+1vq+1f)+B(xp+1vq+2f,xpvq+1f))dt\displaystyle\sum_{p^{\prime}\leq p}\left(B(\nabla_{x}^{p^{\prime}}\nabla_{v}^{q+2}f,\nabla_{x}^{p+1}\nabla_{v}^{q+1}f)+B(\nabla_{x}^{p+1}\nabla_{v}^{q+2}f,\nabla_{x}^{p^{\prime}}\nabla_{v}^{q+1}f)\right)\mathrm{d}t
+pp(B(xpvq+2f,xp+1vqf)+B(xp+1vq+1f,xpvq+1f))dt\displaystyle+\sum_{p^{\prime}\leq p}\left(B(\nabla_{x}^{p^{\prime}}\nabla_{v}^{q+2}f,\nabla_{x}^{p+1}\nabla_{v}^{q}f)+B(\nabla_{x}^{p+1}\nabla_{v}^{q+1}f,\nabla_{x}^{p^{\prime}}\nabla_{v}^{q+1}f)\right)\mathrm{d}t
+ppp′′p+1B(xpvq+2f,xp′′vq+1f)dt.\displaystyle+\sum_{p^{\prime}\leq p}\sum_{p^{\prime\prime}\leq p+1}B(\nabla_{x}^{p^{\prime}}\nabla_{v}^{q+2}f,\nabla_{x}^{p^{\prime\prime}}\nabla_{v}^{q+1}f)\mathrm{d}t. (5.25)

As mentioned above, here BB denotes a bilinear form which is bounded on Lm2×Lm2L_{m}^{2}\times L_{m}^{2}, the exact form of which is irrelevant. It follows that

bt2𝒞c,α,β,j(ε+t)𝔻σ+σ.\displaystyle bt^{2}\mathcal{C}_{c,\alpha,\beta,j}\lesssim(\varepsilon+t)\mathbb{D}_{\sigma}+\mathcal{E}_{\sigma}.

As in (5.5), from similar calculations to those in the proof of (3.2) in Lemma 3.1 for p=2p=2, we have

kxαvβ((σkekv)2f),xαvβfmkxαvβ(σkekvf)Lm22\displaystyle\sum_{k}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}((\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}-\sum_{k}\left\|\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\|_{L_{m}^{2}}^{2} σ\displaystyle\lesssim\mathcal{E}_{\sigma}
atk{vxαvβ((σkekv)2f),vxαvβfmkvxαvβ(σkekvf)Lm22}\displaystyle at\sum_{k}\left\{\left\langle\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}((\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f),\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}-\sum_{k}\left\|\nabla_{v}\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\|_{L_{m}^{2}}^{2}\right\} σ\displaystyle\lesssim\mathcal{E}_{\sigma}
ct3k{xxαvβ((σkekv)2f),xxαvβfmkxxαvβ(σkekvf)Lm22}\displaystyle ct^{3}\sum_{k}\left\{\left\langle\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}((\sigma_{k}e_{k}\cdot\nabla_{v})^{2}f),\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}-\sum_{k}\left\|\nabla_{x}\partial_{x}^{\alpha}\partial_{v}^{\beta}(\sigma_{k}e_{k}\cdot\nabla_{v}f)\right\|_{L_{m}^{2}}^{2}\right\} σ.\displaystyle\lesssim\mathcal{E}_{\sigma}.

This completes the necessary estimates on the Itô correction terms.

Final estimate:

For tt sufficiently small, combining the estimates on the linear terms of (5.5) from the above arguments with (5.11), (5.12), (5.24) and (5.16)-(5.18), we ultimately obtain

dσCσdt(2Cε)𝔻σdt+σ+CfHmσpdt,\displaystyle\mathrm{d}\mathcal{E}_{\sigma}\leq C\mathcal{E}_{\sigma}\mathrm{d}t-(2-C\varepsilon)\mathbb{D}_{\sigma}\mathrm{d}t+\mathcal{M}_{\sigma}+C\left\|f\right\|_{H_{m}^{\sigma}}^{p}\mathrm{d}t, (5.26)

where σ\mathcal{M}_{\sigma} denotes all of the martingale terms:

σ=\displaystyle\mathcal{M}_{\sigma}= 2xαvβ(vfdWt),xαvβfm\displaystyle-2\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta}f\right\rangle_{m}
2atj=1dxαvβ+ej(vfdWt),xαvβ+ejfm\displaystyle-2at\sum_{j=1}^{d}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f\right\rangle_{m}
bt2j=1dxαvβ+ej(vfdWt),xα+ejvβfm\displaystyle-bt^{2}\sum_{j=1}^{d}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}
bt2j=1dxαvβ+ejf,xα+ejvβ(vfdWt)m\displaystyle-bt^{2}\sum_{j=1}^{d}\left\langle\partial_{x}^{\alpha}\partial_{v}^{\beta+e_{j}}f,\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t})\right\rangle_{m}
2ct3j=1dxα+ejvβ(vfdWt),xα+ejvβfm.\displaystyle-2ct^{3}\sum_{j=1}^{d}\left\langle\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}(\nabla_{v}f\cdot\mathrm{d}W_{t}),\partial_{x}^{\alpha+e_{j}}\partial_{v}^{\beta}f\right\rangle_{m}. (5.27)

Therefore, integrating in time and using the BDG inequality, we obtain

𝐄suptτσ(t)+(2Cε)𝐄0τ𝔻σ(s)ds\displaystyle\mathbf{E}\sup_{t^{\prime}\leq\tau}\mathcal{E}_{\sigma}(t^{\prime})+(2-C\varepsilon)\mathbf{E}\int_{0}^{\tau}\mathbb{D}_{\sigma}(s)\mathrm{d}s\leq 𝐄f0Hmσ2+C𝐄0τσ(s)ds+C𝐄0τf(s)Hmσpds\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H_{m}^{\sigma}}^{2}+C\mathbf{E}\int_{0}^{\tau}\mathcal{E}_{\sigma}(s)\mathrm{d}s+C\mathbf{E}\int_{0}^{\tau}\left\|f(s)\right\|_{H_{m}^{\sigma}}^{p}\mathrm{d}s
+C𝐄(0τ(σ(s)+εσ(s)𝔻σ(s))2ds)12\displaystyle+C\mathbf{E}\left(\int_{0}^{\tau}(\mathcal{E}_{\sigma}(s)+\varepsilon\sqrt{\mathcal{E}_{\sigma}(s)\mathbb{D}_{\sigma}(s)})^{2}\mathrm{d}s\right)^{\frac{1}{2}}
\displaystyle\leq 𝐄f0Hmσ2+C𝐄0τσ(s)ds+C𝐄0τf(s)Hmσpds\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H_{m}^{\sigma}}^{2}+C\mathbf{E}\int_{0}^{\tau}\mathcal{E}_{\sigma}(s)\mathrm{d}s+C\mathbf{E}\int_{0}^{\tau}\left\|f(s)\right\|_{H_{m}^{\sigma}}^{p}\mathrm{d}s
+C𝐄suptτ12(t)(0τ(σ(s)12+ε𝔻σ(s)12)2ds)12\displaystyle+C\mathbf{E}\sup_{t^{\prime}\leq\tau}\mathcal{E}^{\frac{1}{2}}(t^{\prime})\left(\int_{0}^{\tau}(\mathcal{E}_{\sigma}(s)^{\frac{1}{2}}+\varepsilon\mathbb{D}_{\sigma}(s)^{\frac{1}{2}})^{2}\mathrm{d}s\right)^{\frac{1}{2}}
\displaystyle\leq 𝐄f0Hmσ2+C𝐄0τσ(s)ds+C𝐄0τf(s)Hmσpds\displaystyle\mathbf{E}\left\|f_{0}\right\|_{H_{m}^{\sigma}}^{2}+C\mathbf{E}\int_{0}^{\tau}\mathcal{E}_{\sigma}(s)\mathrm{d}s+C\mathbf{E}\int_{0}^{\tau}\left\|f(s)\right\|_{H_{m}^{\sigma}}^{p}\mathrm{d}s
+12𝐄suptτσ(t)+Cε2𝐄0τ𝔻σ(s)ds.\displaystyle+\frac{1}{2}\mathbf{E}\sup_{t^{\prime}\leq\tau}\mathcal{E}_{\sigma}(t^{\prime})+C\varepsilon^{2}\mathbf{E}\int_{0}^{\tau}\mathbb{D}_{\sigma}(s)\mathrm{d}s. (5.28)

Rearranging terms and applying Grönwall’s inequality, we obtain

𝐄suptτσ(t)1.\mathbf{E}\sup_{t^{\prime}\leq\tau}\mathcal{E}_{\sigma}(t^{\prime})\lesssim 1. (5.29)

As is standard (and as in the proofs of the p>2p>2 estimates (3.2) and (3.52) in Section 3), a straightforward variation of the above argument extends to prove

𝐄(suptτσ(t))p1,\mathbf{E}\left(\sup_{t^{\prime}\leq\tau}\mathcal{E}_{\sigma}(t^{\prime})\right)^{p}\lesssim 1, (5.30)

completing the desired estimates. ∎ For the reader’s convenience, we include the elementary proof that such constants a,b,ca,b,c as prescribed in (5.2) do indeed exist. A more general situation is treated in [villani2002review]*Lemma A.16.

Lemma 5.2.

Let ε>0\varepsilon>0. Then there exist constants a,b,ca,b,c such that (5.2) holds:

{1aεbε2cε3aε1b,bεac\begin{cases}1\geq\frac{a}{\varepsilon}\geq\frac{b}{\varepsilon^{2}}\geq\frac{c}{\varepsilon^{3}}\\ a\leq\varepsilon\sqrt{1\cdot b},\quad b\leq\varepsilon\sqrt{a\cdot c}\end{cases}
Proof.

Let ϑ>0\vartheta>0, and pick m1,m2,m3>0m_{1},m_{2},m_{3}>0 such that

m1=1,m2(1,2),m3(m2,2m21)m_{1}=1,\quad m_{2}\in(1,2),\quad m_{3}\in(m_{2},2m_{2}-1)

and set:

a=ϑm1=ϑ,b=ϑm2,c=ϑm3.a=\vartheta^{m_{1}}=\vartheta,\quad b=\vartheta^{m_{2}},\quad c=\vartheta^{m_{3}}.

The proof of the lemma is concluded by picking ϑ\vartheta sufficiently small. ∎

References