The Vlasov-Poisson and Vlasov-Poisson-Fokker-Planck systems in stochastic electromagnetic fields: local well-posedness
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 . Initial conditions are taken with (in addition to polynomial velocity weights). We additionally show that solutions to the VPFP are instantly 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 and 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
(1.1) |
where is the collision frequency – we treat both the case and (i.e. the Vlasov–Poisson equations). Below we denote the Fokker-Planck operator
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 , although the case could be approached with similar arguments. The process 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 and also applies to external magnetic fields. For simplicity, we will take initial conditions in the velocity-weighted Sobolev space defined by the norm:
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 and for simplicity we will show how to do this in ; the extension to other dimensions is straightforward and is omitted. To make this concrete representation of , we define a real Fourier basis on by defining for each
where , and , and for each , is a set of three orthonormal vectors with spanning the plane perpendicular to with the property that and parallel to . The constant is a normalization factor so that are a complete orthonormal basis on . With this, we define our external electric field as
with is a family of independent standard Wiener processes defined on a given stochastic basis . The are coloring coefficients satisfying at least , however, more stringent regularity requirements will be assumed below (here we make the natural definition for ). 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
(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 with an almost-surely strictly positive stopping time and an adapted stochastic process satisfying the regularity
and for satisfies,
Moreover, we say such pathwise solutions are unique if for any pair , we have
In this case, and are called indistinguishable.
The following definition of maximal pathwise solution provides a continuation criterion for strong solutions. For this we will use ; sharper continuation criteria will be considered in future work. That is, we show that local solutions can be uniquely extended provided some norm remains finite for and fixed and arbitrary.
Definition 1.2.
Fix an integer and . We call a maximal pathwise solution a triple of a solution , an increasing sequence of almost-surely positive stopping times , and a limiting stopping time such that each pair is a local pathwise solution, , and
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 . Let and be fixed integers and assume that
(1.3) |
for some (integer). Suppose that the initial condition is an -measurable random variable such that almost-surely. Then, there exists a unique, maximal pathwise solution to (1.1) for any .
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 the speed of light),
with
(1.4) |
or when electromagnetic fields are correlated, for example one could use forcing of the following potentially natural form
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:
for any kernel such that for all .
Remark 1.6.
It should be straightforward to extend to . It should also be possible to treat non-integer , , and , 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 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. . 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].
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 with and (both integers) with trajectories in . 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 that take values in . 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 such that:
(2.1) |
and define:
(2.2) |
Then we define the regularized SPDE
(2.3) |
We show in Section 3 that this SPDE admits global-in-time, unique, pathwise solutions (i.e. with probability in the definition of maximal solutions) starting from initial conditions. Specifically, we prove the following.
Lemma 2.1.
Let be a -measurable random variable such that ,
Then, there exists an –a.s. which is a solution to (2.3) in the sense that
(2.4) |
where the equality holds in . Moreover, if is any other solution in the above sense, then almost surely in the sense that
It is clear that solutions of (2.3) are also solutions to (1.1) for as long as , and so by considering the increasing sequence of and defining the stopping times
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.
Proof.
First consider the case that almost-surely. Then, we choose in (2.3), and define the stopping time:
where is the solution to (2.3) with initial data , guaranteed to exist and be unique from Lemma 2.1. Note that up to time , the process also solves (1.1), since for we have and therefore Clearly almost surely since and takes values continuously in . The pair 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 to a maximal solution.
Let be the collection of all stopping times corresponding to a local solution and define . Define also:
Fix finite but arbitrary and assume for some . This implies that
and thus can be continued to a solution of (2.3) with and thus of (1.1) up to a stopping time past - contradicting ’s maximality. Since was arbitrary, we either have , or for all . In the latter case, we also get for all and thus
Now we drop the almost-sure uniform boundedness requirement. If almost surely, we decompose where . Now each generates a maximal solution where is the corresponding maximal existence time, and we define the \saytotal maximal solution (in high regularity) of (1.1) as with:
(2.5) | |||
(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 with and define . Specifically, we seek a solution to the following regularized SPDE (here the convolution in has been periodized),
(2.7) | |||
This is done by an iteration method, specifically the following
(2.8) | |||
(2.9) | |||
(2.10) |
where we denote
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
(2.11) |
where is a -dimensional Brownian motion defined in a new stochastic basis (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 on , defined on the product space
which map 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
(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 as well as the preservation of the Casimir conservation laws, e.g. if then and for one at least has . However, these properties do not play an important role here.
Next, we need uniform a priori estimates to enable passing . These are obtained via Eulerian energy methods and come out as , , and ,
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 which is almost-surely greater than such that forms a Cauchy sequence in , at which point it is not hard to pass to the limit, iterate in , 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 is nearly comparable in size to the -th term of a power series in powers of of the solution. This procedure finally yields
Lemma 2.4.
Let be an -measurable random variable such that ,
Then, there exists an –a.s. which is a solution to (2.7) in the sense that , –a.s.:
(2.13) |
where the equality holds in . Moreover, if is any other solution in the above sense, then almost surely in the sense that
The next step in the proof of Lemma 2.1 is to remove the superfluous mollifier , which begins with obtaining -independent estimates (now indexing solutions to (2.7) by ),
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 norm controls all norms for and . At this step, it does not seem straightforward to prove that 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 from initial data, which is done in Section 4. We perform a regularization procedure on the initial data by defining a sequence of initial conditions
where and satisfies and . Note these have been both mollified and cut-off in velocity (to improve both regularity and localization). For all , we hence have for all . Subsequently, there are unique local pathwise solutions to (1.1) with
By obtaining suitable uniform-in- upper bounds on the norm, we may pass to the limit and hence extract local pathwise solutions to the original problem in ; 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 be a separable Hilbert space, with an orthonormal basis We can formally define a cylindrical Wiener process on by the formula
Since this sum is divergent on one frequently employs the larger Hilbert space:
where it can be shown that the formal sum for converges and defines a process whose paths are almost surely in . Moreover, the embedding is Hilbert–Schmidt. For any separable Hilbert space , we denote the space of all Hilbert–Schmidt operators from to by ; the definition of this space is:
For more details on cylindrical Wiener processes and the relevant functional analytic setting, see [da2014stochastic].
At various points, we use the notation to signify that is in any space for .
We often employ the common notation:
which means that there exists a constant depending only on the parameters but not on the argument , such that for all relevant . We omit the parameters if they are unimportant or clear from the context.
For the velocity-weighted norms and inner products, we set:
Finally, at various points we use the mixed weighted norms:
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 defined inductively as:
(3.1) |
As discussed in Section 2, for a given , the solution is constructed via the method of stochastic characteristics.
First, we provide -independent estimates in order to pass to the limit for which we need appropriate compactness estimates for the iterates defined above. We remark that studying the stochastic flow of diffeomorphisms could show that , however, providing -independent (and eventually -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 be a sequence of global solutions to the iterative scheme (3.1) with for all For we have the uniform estimates:
(3.2) |
and
(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 there exists a constant such that
(3.4) |
Next, we recall the following Gagliardo-Nirenberg-Sobolev estimate: for all integers and functions (in or ) there holds
(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 for some , , and arbitrary. Denote Then, for a constant that does not depend on or :
(3.6) |
and
(3.7) |
Proof of Lemma 3.3.
In what follows, will be fixed, which implies Denote the multivariate binomial coefficients by
We begin the proof of (3.6) by using the product rule and the triangle inequality:
(3.8) |
Now, we split into four separate cases:
- Case 1 :
-
In this case, we use Hölder’s inequality, the embedding , and Lemma 3.2
(3.9) In the following cases,
- Case 2 or :
-
Here, we have , and hence similarly to the previous case
(3.10) - Case 3 :
-
Here, we necessarily have and hence (using the Sobolev embedding now on ),
(3.11) - Case 4 :
-
Here so:
(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 . First, distribute the derivatives with Leibniz’s rule.
(3.13) |
We distinguish the following cases:
- Case 1 :
-
by integrating by parts the onto the weight, we have
(3.14) - Case 2 :
-
Cauchy-Schwarz gives
(3.15) - Case 3 :
-
We have:
(3.16) In the above we used Hölder’s inequality, Gagliardo-Nirenberg interpolation (3.5) on and Sobolev embedding on the term, where we note that the order of integrability corresponds to the embedding of in for given by:
(3.17) which satisfies exactly if .
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 , which we then use to derive an estimate for , for Applying Itô’s formula to , we have:
(3.18) |
where
Here, the terms abbreviate transport, dissipation, friction, nonlinear electric field, martingale, and correction contributions, respectively. We begin by observing that by integration by parts,
(3.19) |
Similarly, for the dissipative term, integrating by parts gives:
(3.20) |
Next, turn to the Itô correction terms, which need to be treated carefully in order to not lose derivatives. Distributing derivatives gives
(3.21) | ||||
(3.22) | ||||
(3.23) |
Now, (3.22) provides a term of highest order that cancels the Itô correction , and terms of lower order that can either be readily controlled by or cancel out with a corresponding term in (3.21). Integrating by parts in (3.22) we have,
(3.24) | ||||
(3.25) | ||||
(3.26) | ||||
(3.27) |
Thus, we observe that (3.24) cancels the Itô correction (3.23), (3.27) is bounded above by , and (3.26) only contains derivatives of order lower than and is thus bounded above by . Next, we turn to (3.25). This term contains (A) terms from the commutator where the total number of derivatives on is strictly less than , which can be treated by integration by parts of the and are thus bounded above by ; and (B) a highest order term which we deal with as follows:
(3.28) | ||||
(3.29) |
Notice that to obtain the prefactor in (3.28) we used the symmetry of the tensor . Now, (3.29) can be directly bounded by , while (3.28) cancels out the highest order term in (3.21). Therefore, we finally conclude using (3.21)–(3.29) that we have
(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:
(3.31) |
Finally, the martingale contribution is given by
(3.32) |
We sum (3.19)-(3.32) over and obtain
(3.33) |
so integrating in time and using the Burkhölder–Davis–Gundy inequality (see e.g. [da1996ergodicity]) (hereinafter abbreviated as BDG) we obtain:
(3.34) |
where the second line followed from Hölder’s inequality. After rearranging and applying Grönwall’s inequality we obtain the uniform-in- estimate:
(3.35) |
where the constant depends on , but not or . Thus we have obtained (3.2) for .
Now, we use Itô’s formula again, this time for , with :
The latter term is treated by a straightforward commutator estimate, and together with the above estimates on , we obtain
After integrating in time, using the BDG inequality, and applying Hölder’s inequality, we have
(3.36) |
By rearranging and using Grönwall’s lemma, we obtain (3.2).
We now turn to the proof of (3.3). We have:
The terms that are regular in time are estimated in a straightforward manner using the available regularity:
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
Therefore, using that continuously and (3.2), we obtain:
(3.37) |
uniformly in , 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 , which ensures both and, despite the power of , the estimates do not lose any moments in as is orthogonal to (nor does the dependence create any issues controlling higher regularity).
We continue the proof of Lemma (2.4). The approximation procedure mixes and 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 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 , a -measurable, almost-surely finite, random constant such that for all there holds
Moreover, there holds,
Proof.
Recall the uniform in bound (3.2) for the iterates for :
This estimate implies:
for . Denote by the sets:
and note that by Chebyshev’s inequality:
(3.38) |
It then follows by the Borel–Cantelli lemma that
implying that -a.s., at most for a finite number of ’s. Denote the largest such by . We then see that there is a random constant such that
–almost surely. In particular, we can take:
To bound the probability that is large, we observe:
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 such that is Cauchy in and the stopping time
is almost-surely greater than .
Proof.
Define the increasing sequence of stopping times
Note that by Lemma 3.5 there holds
Therefore, and so if we define
then is almost-surely greater than or equal to .
Let be fixed arbitrary. We will show by induction that (deterministic constant depending on ) such that for all , there holds
(3.39) |
First consider the case . The calculation of is the same in Lemma 3.1 except for the nonlinear terms. That is, for we have for some constant
For the nonlinear term we note that by (3.7) we have, recalling the definition of
Integrating in time and using the BDG inequality as above, we obtain (note that and have the same initial data),
Therefore, Grönwall’s inequality verifies (3.39) for and some large .
Next consider the inductive step. Hence, suppose that (3.39) holds for and we wish to verify that it holds for . As above, for some constant
The nonlinearity separates into several natural terms, namely
The term is treated via (3.7) in the same manner as in Lemma 3.1, giving
The terms however are different. The term is estimated via the following for :
The term requires a control on the difference :
Therefore, for
Integrating in time and using the BDG inequality as above, we obtain (noting that and have the same initial data) for :
By the inductive hypothesis
and so we have verfied (3.39).
Finally, we show that (3.39) implies that is Cauchy in . Indeed, let and
(3.40) |
Hence, if we choose , then
By Stirling’s formula we have the following uniformly in (using ),
therefore
We conclude that the sequence is Cauchy as claimed in the lemma. ∎
Lemma 3.7.
For each the iterates converge uniformly in on compact subintervals of to a strong pathwise solution of the SPDE (2.7) on the set .
Proof.
Consider only . Let be the limiting process of the in - 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:
(3.41) | ||||
(3.42) | ||||
(3.43) | ||||
(3.44) |
For the nonlinear electric field terms, we have:
where:
These terms are estimated as follows:
Lastly, for the martingale terms we use the BDG inequality:
(3.45) |
Combining the above, we see that is a solution of (2.7). ∎
Corollary 3.8.
There exists a global, strong pathwise solution of the SPDE (2.7) such that , .
Proof.
By sending and using that is a non-decreasing sequence such that we see that almost-surely, converges uniformly in on compact subintervals of to a limiting function . By Sobolev interpolation, and uniform boundedness in , we obtain similar uniform convergence in for all and . At the same time, the uniform bounds on from Lemma 3.1 imply that , by the lower semicontinuity of weak convergence. By Lemma 3.7, the limiting function is also a solution of (2.7). Now, we simply iterate the construction starting at 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 be two global pathwise solutions to (2.7) on the same stochastic basis with for some -measurable with for some and such that for all , Then are indistinguishable, that is:
(3.46) |
Proof.
This is proved by an energy estimate on Similarly to the proof of Lemma 3.1, for we have:
(3.47) |
We split the electric field contributions as:
where:
These are estimated as follows:
(3.48) |
(3.49) |
(3.50) |
where in (3.48) we used the mean value theorem for 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 . Since , the stopping time:
is almost surely finite. Even though it is not clear that is almost surely positive in general, for almost every there exists such that and in addition –a.s. as . With this in mind, we fix and use (3.48)-(3.50) and the BDG inequality in (3.47), to obtain:
(3.51) |
for all , whereby the usual rearrangement and Grönwall’s lemma give:
Taking and then , the conclusion follows. ∎
3.2 Proof of Lemma 2.1
Next, we want to pass to the limit , for which we need uniform-in- estimates similar to those of Lemma 3.1, but this time for a family of solutions to (2.7). Note that since the highest norm in which we know is continuous is - and thus we use this as the base for our estimates.
Lemma 3.10.
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 derivatives on a solution of (2.7), we have:
(3.54) |
We show that this implies the same estimate for . Begin by using Itô’s formula on for where similarly to (3.18) we obtain:
(3.55) |
The linear terms are treated as in the proof of Lemma 3.1, and the only term that requires new attention is By the classical Gagliardo-Nirenberg inequality (see e.g. [Proposition A.3 [tao2006nonlinear]] we have:
(3.56) |
where for fixed the interpolation parameter is given by:
(3.57) |
and thus provided . By Young’s inequality and (3.56) it follows that
(3.58) |
Plugging this back into (3.55), and using the same procedure as in the proof of Lemma 3.1, we obtain:
(3.59) |
We again integrate in time and apply the BDG inequality as in the proof of (3.2) for , where the only difference is the term which is now controlled by the inductive hypothesis, and we get:
(3.60) |
With this in hand, we can directly transfer the proof of (3.2) for 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 norm of . This is sufficient to start the inductive scheme above in as . As the linear terms are always controlled in the same way, we only focus on the electric field contributions. As always, we have:
(3.61) |
but since only two derivatives are acting on at this point, the terms in the summation are only present when and Let in , and for let be arbitrary such that . Then by Hölder’s inequality and Sobolev embeddings we have
(3.62) |
where we have used that and the embedding which holds for all due to our choice of . From this point on the procedure is the same as in the inductive step. We plug (3.62) into (3.55) for with , sum over all such as well as the case when integrate in time, apply the BDG inequality and Grönwall’s lemma and obtain:
(3.63) |
Then applying the same argument as in the proof of (3.2) for we also obtain for :
(3.64) |
This provides the inductive base and therefore the proof of the lemma is complete for . ∎
For solutions to (2.7), it is unclear how to prove forms a Cauchy sequence as . 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 be a decreasing sequence of positive numbers with as and define the corresponding sequence of solutions to (2.7), which we have shown satisfy the uniform bounds (3.52) and (3.53). For and such that , we define the pathspace
(3.65) |
Recall that since and compactly, from [flandoli1995martingale]*Theorem 2.2, we have:
By the uniform estimates (3.52) and (3.53), the laws are bounded in probability in , and thus they are tight in the smaller pathspace
(3.66) |
Note that the tightness in is in the weak- topology. We now use this to obtain a martingale solution to (2.3) in high regularity.
Proposition 3.11.
Let be a probability measure on so that for some . Then there exists a stochastic basis and a predictable process
(3.67) |
such that and solves (2.3) in the sense that, there is a sequence of i.i.d Brownian motions such that the following equality holds in
with
Proof.
Let in The sequence is tight by the uniform estimates (3.52) and (3.53) combined with the fact that its projection onto is the same for each . By Prokhorov’s theorem, has a weakly convergent subsequence - reindexed to . By Skorokhod’s embedding theorem, there exists a new probability space and on it random elements with laws which converge –a.s. to some limit in the product topology of . 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 satisfy (2.7) just like , but in the new probability space :
where we have denoted by the external electric field corresponding to ; that is to say, if:
then is simply given by:
The passage to the limit in the SPDEs (2.7) satisfied by to obtain that the limit solves (2.3) can be carried out by combining the convergences and 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 , 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 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 .
Lemma 3.12 (Gyöngy–Krylov).
Let be a sequence of -valued random variables, where is a complete separable metric space. Then converges in probability if and only if for every two subsequences the joint sequence has a subsequence whose laws converge weakly to a probability measure supported on the diagonal of :
(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 be global solutions to (2.3) on the same stochastic basis with almost surely, where for some and such that . Then are indistinguishable, that is:
(3.69) |
Proof.
First of all, notice that since for the stopping times
(3.70) |
are well defined and satisfy as , –almost surely. We now perform an energy estimate on . We use Itô’s formula on the quantity for :
(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:
(3.72) |
for , where we have used the mean value theorem on and Lemma 3.3. Thus, arguing similarly to the proof of Lemma 3.9 (i.e. by the BDG and Grönwall inequalities), we have:
(3.73) |
Since as the monotone convergence theorem implies that for all we have:
(3.74) |
which implies (3.69) since is arbitrary. ∎
We now have everything we need to (subsequentially) pass to the limit in the original stochastic basis.
Proof of Lemma 2.1.
We define the joint laws Similarly to the discussion in Lemma 3.11, for any sequence with as by Prokhorov’s theorem the estimates (3.52) and (3.53) (and the fact that is a singleton) provide a weakly convergent subsequence of probability measures in , which we still denote (after relabelling) by , and we denote its limit by . By the Skorokhod embedding theorem, we can construct a new stochastic basis again denoted by and on it a sequence of random elements and such that , and:
As in Lemma 3.11, and 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 in all the terms of (2.7) for and (using [debussche2011local]*Lemma 2.1 for the stochastic integrals) to show that are solutions to (2.3) on the new stochastic basis. Since , we also have , and thus in the limit we obtain . Therefore, by Lemma 3.13, and are indistinguishable. This means that the measure , defined as the projection of onto the first two components , is in fact supported on the diagonal of . Thus, by Lemma 3.12, a subsequence of converges in probability in the original stochastic basis in the topology of to a limiting process 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 satisfying almost surely, for a fixed deterministic . This assumption can be removed at the end by a cutting argument similar to that of Lemma 2.2. We treat the initial data with a sequence of regularization and velocity cutoff operators obtaining a sequence of regularized data
defined as
where satisfies and . We note the following properties of this regularization. The proofs are standard and are omitted for brevity.
Lemma 4.1.
Let be integers and be arbitrary. Then,
-
(i)
The regularization operators are uniformly bounded on
-
(ii)
The regularization operators satisfy: for
-
(iii)
The regularization operators converge in the following senses: for there holds
(4.1) (4.2)
In the previous section, we showed that each of the generates a maximal solution of (1.1) in , where is the maximal time of existence of . We now show that the sequence of approximate solutions has a strongly convergent subsequence.
We start by defining the stopping times:
(4.3) | |||
(4.4) |
The following is similar to [mikulevicius2004stochastic]*Lemma 37 or [GV14]*Lemma 7.1:
Lemma 4.2.
Let be a sequence of stopping times and suppose that a sequence of predictable processes satisfy:
(4.5) |
(4.6) |
Then, there exists a stopping time with a predictable process and a subsequence of such that
(4.7) |
and
(4.8) |
We now verify that the regularized solutions satisfy the conditions of Lemma 4.2.
Proof.
We begin with proving that the Cauchy property (4.5) holds for the sequence of solutions to (2.3) with initial data . This is done via an energy estimate with some similarities with the uniqueness and convergence proofs in Section 3 with one significant difference. For , we have:
(4.9) |
We will control the above for . 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 , we have:
(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 , we have:
(4.11) |
In what follows we denote
(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 is in , 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:
(4.13) |
The correction term in (4.13) is:
(4.14) |
where:
(4.15) |
Note that so:
(4.16) |
For the main terms of (4.13) we have similar estimates as before. For the difference and for we have (recalling the definition (4.12)):
(4.17) |
where we used (4.10) for instead of derivatives and the definition of the stopping time . Similarly, for the norm of and for we have:
(4.18) |
where we again used the definition of the stopping time . Now, plugging (4.16), (4.17) (4.18) into (4.13), we obtain:
(4.19) |
Integrating (4.19) in time and using the BDG inequality, we obtain:
so after rearranging and using Grönwall’s inequality as done previously in e.g. the proof of Lemma 3.1, we get:
(4.20) |
Now returning to (4.11), integrating in time, using the BDG inequality, plugging in (4.20), we obtain:
Therefore we have
(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:
(4.22) |
Let us denote
so that for , after integrating in time, (4.22) gives:
(4.23) |
Therefore, by Chebyshev’s inequality for the usual deterministic integral and Doob’s ienquality for the martingale, we have
(4.24) |
note we also used that implies for a constant that depends on the size of the initial data uniformly in , since independently of . Taking , we obtain (4.6). ∎
Combining Lemmas 4.2 and 4.3, we obtain the existence of a local strong solutions to (1.1) when 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 is -measurable with –a.s.. Specifically, since each component generates a local strong solution to (1.1) and we re-construct the full and using
(4.25) |
and
(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 for .
We first prove that if , then the solution lies in for (with size depending only on the 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 expected from known hypoelliptic regularization estimates (alternatively, one can deduce them from scaling arguments; see e.g. [bouchut1993existence]) we have
For estimates we hence define
(5.1) |
The constants are chosen (indepedent of ) such that and so that
The parameters are chosen more specifically to satisfy: for some sufficiently small we require
(5.2) |
We recall the proof that such exist in Lemma 5.2 below. Note that these conditions imply , a fact we use repeatedly below.
Hence, an estimate on in terms of 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 be a -measurable initial data and suppose that for all , such that for some we have:
(5.3) |
Let , let be the unique pathwise solution of (2.3).
Then depending only on such that there holds for all ,
Therefore, almost surely for all .
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 , is an -measurable random variable with
Therefore, we may apply Proposition 5.1 to the initial data with . 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 but the same arguments (up to a suitable rescaling of the coefficients ) apply for any . Define the dissipation rate:
(5.4) |
which we show arises from . Note that this is almost the same as the contribution from that arises when the time derivative lands on the powers of in front of the higher-order terms. In order to reduce some of the notation in the ensuing calculation, we use to denote an -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
The main step of the proof is to calculate the following
As in the proof of the various bounds in Sections 3–4, we have:
(5.5) |
This formula, and its straightforward variations, are then used to expand most of the terms of , with the exception of the cross-terms (i.e. those multiplied by ). For the cross-terms we instead have
(5.6) |
where we abbreviated and 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 free transport and the Fokker–Planck operator. The main effect of these terms is to introduce the dissipation . 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 to by:
(5.7) |
For the rest of this proof, we denote . By integration by parts we may write
(5.8) |
where recall from above that denotes an bounded bilinear form, the exact form of which is not relevant. The dissipation term is more easily treated, yielding
(5.9) |
The friction term can be re-arranged as follows
(5.10) |
The fundamental structure of the hypocoercive norm is that the term gives rise to the dissipation term that would otherwise be missing from the dissipation of a kinetic equation. That is, from (5.8), we obtain:
(5.11) |
and similarly, from (5.9) and (5.10):
(5.12) |
Putting together the negative definite terms that arise from and those in (5.7), we obtain for short and for some :
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,
We first analyze the electric field’s contribution to (5.5), similarly to the various nonlinear estimates of Section 3:
(5.13) |
where for each the interpolation index is given as:
(5.14) |
Note that in the above, since . Therefore,
(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 and the second term, which is derived from the factor , contains at most derivatives (since whenever ).
In a similar manner, the corresponding \saysecond term in the electric field contributions to the and terms of the energy will contain at most derivatives. This means that all in all, for short we can bound the electric field contributions to (5.5), as well as those to the higher order terms in the definition of , from (5.13):
(5.16) | |||
(5.17) | |||
(5.18) |
for some fixed. Note that the high power of is a priori controlled (in ) by the finite -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 up to a lower order term:
(5.19) |
We analyze the first term in the final equality above:
(5.20) |
To estimate each term in the summation above, we again interpolate as in the proof of Lemma 3.10;
(5.21) |
(5.22) |
where
which again satisfies since With these exponents, plugging (5.21)-(5.22) into (5.20) we get:
(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 implies , we obtain from (5.23) that such that
(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:
(5.25) |
As mentioned above, here denotes a bilinear form which is bounded on , the exact form of which is irrelevant. It follows that
As in (5.5), from similar calculations to those in the proof of (3.2) in Lemma 3.1 for , we have
This completes the necessary estimates on the Itô correction terms.
Final estimate:
For 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
(5.26) |
where denotes all of the martingale terms:
(5.27) |
Therefore, integrating in time and using the BDG inequality, we obtain
(5.28) |
Rearranging terms and applying Grönwall’s inequality, we obtain
(5.29) |
As is standard (and as in the proofs of the estimates (3.2) and (3.52) in Section 3), a straightforward variation of the above argument extends to prove
(5.30) |
completing the desired estimates. ∎ For the reader’s convenience, we include the elementary proof that such constants as prescribed in (5.2) do indeed exist. A more general situation is treated in [villani2002review]*Lemma A.16.
Lemma 5.2.
Let . Then there exist constants such that (5.2) holds:
Proof.
Let , and pick such that
and set:
The proof of the lemma is concluded by picking sufficiently small. ∎