• Aucun résultat trouvé

A Bhatnagar-Gross-Krook Approximation to Stochastic Scalar Conservation Laws

N/A
N/A
Protected

Academic year: 2021

Partager "A Bhatnagar-Gross-Krook Approximation to Stochastic Scalar Conservation Laws"

Copied!
37
0
0

Texte intégral

(1)

HAL Id: hal-00826681

https://hal.archives-ouvertes.fr/hal-00826681

Submitted on 28 May 2013

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Scalar Conservation Laws

Martina Hofmanova

To cite this version:

Martina Hofmanova. A Bhatnagar-Gross-Krook Approximation to Stochastic Scalar Conservation

Laws. Annales de l’Institut Henri Poincaré (B) Probabilités et Statistiques, Institut Henri Poincaré

(IHP), 2015, 51 (4), pp.1500-1528. �10.1214/14-AIHP610�. �hal-00826681�

(2)

to Stochastic Scalar Conservation Laws

Martina Hofmanov´ a

Abstract. We study a BGK-like approximation to hyperbolic conserva- tion laws forced by a multiplicative noise. First, we make use of the stochastic characteristics method and establish the existence of a solu- tion for any fixed parameterε. In the next step, we investigate the limit asεtends to 0 and show the convergence to the kinetic solution of the limit problem.

Mathematics Subject Classification (2010).60H15, 35R60, 35L65.

Keywords. stochastic conservation laws, kinetic solution, BGK model, hydrodynamic limit, stochastic characteristics method.

1. Introduction

In the present paper, we consider a scalar conservation law with stochastic forcing

du+ div A(u)

dt=Φ(u) dW, t∈(0, T), x∈TN, u(0) =u0

(1.1) and study its approximation in the sense of Bhatnagar-Gross-Krook (a BGK- like approximation for short). In particular, we aim to describe the conser- vation law (1.1) as the hydrodynamic limit of the stochastic BGK model, as the microscopic scaleεgoes to 0.

The literature devoted to the deterministic counterpart, i.e. correspond- ing to the situationΦ= 0, is quite extensive (see [1], [11], [15], [16], [17], [18], [19], [20]). In that case, the BGK model is given as follows

t+a(ξ)· ∇

fεuε−fε

ε , t >0, x∈TN, ξ∈R, (1.2)

This research was supported in part by the GA ˇCR Grant no. P201/10/0752 and the GA UK Grant no. 556712.

(3)

whereχuε, the so-called equilibrium function, is defined by χuε(ξ) =10<ξ<uε−1uε<ξ<0,

andais the derivative ofA. The differential operator∇is with respect to the space variablex. The additional real-valued variableξ is called velocity; the solutionfεis then a microscopic density of particles at (t, x) with velocityξ.

The local density of particles is defined by uε(t, x) =

Z

R

fε(t, x, ξ) dξ.

The collisions of particles are given by the nonlinear kernel on the right hand side of (1.2). The idea is that, as ε→0, the solutions fε of (1.2) converge to χu where uis the unique kinetic or entropy solution of the deterministic scalar conservation law.

The addition of the stochastic term to the basic governing equation is rather natural for both practical and theoretical applications. Such a term can be used for instance to account for numerical and empirical uncertainties and therefore stochastic conservation laws has been recently of growing inter- est, see [2], [5], [7], [10], [12], [22], [23], [24]. The first complete well-posedness result for multi-dimensional scalar conservation laws driven by a general mul- tiplicative noise was obtained by Debussche and Vovelle [5] for the case of kinetic solutions. In the present paper, we extend this result and show that the kinetic solution is the macroscopic limit of stochastic BGK approxima- tions. As the latter are much simpler equations that can be solved explicitly, this analysis can be used for developing innovative numerical schemes for hyperbolic conservation laws.

The BGK model in the stochastic case reads dFε+a(ξ)· ∇Fεdt=1uε−Fε

ε dt−∂ξFεΦdW −1

2∂ξ G2(−∂ξFε) dt, Fε(0) =F0ε,

(1.3) where the functionFεcorresponds tofε+10>ξ, the local densityuεis given as above, and the function G2 will be defined in (2.1). Note, that setting Φ = 0 in (1.3) yields an equation which is equivalent to the deterministic BGK model (1.2). Our purpose here is twofold. First, we make use of the stochastic characteristics method as developed by Kunita in [14] to study a certain auxiliary problem. With this in hand, we fixεand prove the existence of a unique weak solution to the stochastic BGK model (1.3). Second, we establish a series of estimates uniform inεwhich together with the results of Debussche and Vovelle [5] justify the limit argument, asε→0, and give the convergence of the weak solutions of (1.3) to the kinetic solution of (1.1).

Let us make some comments on the deterministic BGK model (1.2).

Even though the general concept of the proof is analogous, we point out that the techniques required by the stochastic case are significantly different. In particular, the characteristic system for the deterministic BGK model consists

(4)

of independent equations dxi(t)

dt =ai(ξ), i= 1, . . . , N,

and the ξ-coordinate of the characteristic curve is constant. Accordingly, it is much easier to control the behavior offεfor largeξ. Namely, if the initial dataf0εare compactly supported (inξ), the same remains valid also for the solution itself and also the convergence proof simplifies. On the contrary, in the stochastic case, theξ-coordinate of the characteristic curve is governed by an SDE and therefore this property is, in general, lost. Similar issues has to be dealt with in order to obtain all the necessary uniform estimates. To overcome this difficulty, it was needed to develop a suitable method to control the decay at infinity in connection with the remaining variablesω, t, x. (cf.

Proposition5.3).

Using this approach we are able to prove the convergence of the BGK model under a slightly weaker hypothesis on the initial datumu0than usually assumed in the deterministic case: it is not supposed to be bounded, we only assumeu0∈Lp(Ω×TN) for allp∈[1,∞). Note, that under this condition, the initial data for the deterministic BGK model, for instance f0ε = χu0, are not compactly supported and so the usual methods are not applicable. In the deterministic case, however, the boundedness assumption is fairly natural since also the solutionuto the conservation law remains bounded. Obviously, this is not true for the stochastic case as it is impossible to get any Lω estimates due to the active white noise term.

There is another difficulty coming from the complex structure of the characteristic system for the stochastic BGK model (1.3). Namely, the finite speed of propagation that is an easy consequence of boundedness of the solu- tionuof the conservation law in the deterministic case (see for instance [19]) is no longer valid and therefore some growth assumptions on the transport coefficient aare in place. The hypothesis of bounded derivatives is natural for the stochastic characteristics method as it implies the existence of global stochastic flows. Even though this already includes one important example of Burgers’ equation it is of essential interest to handle also more general co- efficients having polynomial growth. This was achieved by a suitable cut-off procedure which also guarantees all the necessary estimates.

The exposition is organized as follows. In Section 2, we introduce the basic setting and state the main result, Theorem 2.1. In order to make the paper more self-contained, Section3provides a brief overview of two concepts which are the keystones of our proof of existence and convergence of the BGK model. On the one hand, it is the notion of kinetic solution to stochastic hyperbolic conservation laws, on the other hand, the method of stochastic characteristics for first-order linear SPDEs. Section 4 is mainly devoted to the existence proof for stochastic BGK model, however, in the Subsection4.2 we establish some important estimates useful in Section5. This final section contains technical details of the passage to the limit and completes the proof of Theorem2.1.

(5)

2. Setting and the main result

We now give the precise assumptions on each of the terms appearing in the above equations (1.1) and (1.3). We work on a finite-time interval [0, T], T >0,and consider periodic boundary conditions:x∈TN where TN is the N-dimensional torus. The flux function

A= (A1, . . . , AN) :R−→RN

is supposed to be of classC4,η, for someη >0, with a polynomial growth of its first derivative, denoted bya= (a1, . . . , aN).

Regarding the stochastic term, let (Ω,F,(Ft)t≥0,P) be a stochastic basis with a complete, right-continuous filtration. The initial datum may be random in general, i.e. F0-measurable, and we assumeu0∈Lp(Ω;Lp(TN)) for allp∈[1,∞). As we intend to apply the stochastic characteristics method developed by Kunita [14], we restrict ourselves to finite-dimensional noise.

However, our results extend to infinite-dimensional setting once the corre- sponding properties of stochastic flows are established. Let U be a finite- dimensional Hilbert space and (ek)dk=1its orthonormal basis. The processW is ad-dimensional (Ft)-Wiener process:W(t) =Pd

k=1βk(t)ek with (βk)dk=1 being mutually independent real-valued standard Wiener processes relative to (Ft)t≥0. The diffusion coefficientΦis then defined as

Φ(z) :U−→L2(TN) h7−→

d

X

k=1

gk(·, z(·))hek, hi, z∈L2(TN),

where the functions g1, . . . , gd : TN ×R → R are of class C4,η, for some η >0, with linear growth and bounded derivatives of all orders. Under these assumptions, the following estimate holds true

G2(x, ξ) =

d

X

k=1

|gk(x, ξ)|2≤C 1 +|ξ|2

, x∈TN, ξ∈R. (2.1) However, in order to get all the necessary estimates (cf. Corollary4.11, Re- mark4.12), we restrict ourselves to two special cases: either

gk(x,0) = 0, x∈TN, k= 1, . . . , d, (2.2) hence (2.1) rewrites as

G2(x, ξ)≤C|ξ|2, x∈TN, ξ∈R, or we strengthen (2.1) in the following way

G2(x, ξ)≤C, x∈TN, ξ∈R. (2.3) Note, that the latter is satisfied for instance in the case of additive noise.

In this setting, we can assume without loss of generality that the σ- algebra F is countably generated and (Ft)t≥0 is the completed filtration generated by the Wiener process and the initial condition. Let us denote by P the predictable σ-algebra on Ω×[0, T] associated to (Ft)t≥0 and byPs

(6)

the predictableσ-algebra on Ω×[s, T] associated to (Ft)t≥s. For notational simplicity, we writeLPs(Ω×[s, T]×TN ×R) to denote1

L Ω×[s, T]×TN ×R,Ps⊗ B(TN)⊗ B(R),dP⊗dt⊗dx⊗dξ . Concerning the initial data for the BGK model (1.3), one possibility is to consider simplyF0ε=1u0, however, one can also take some suitable approximations of1u0. Namely, let{uε0;ε∈(0,1)}be a set of approximate F0-measurable initial data, which is bounded in Lp(Ω;Lp(TN)) for all p∈ [1,∞), and assume in addition thatuε0→u0inL1(Ω;L1(TN)). Thus, setting F0ε=1uε0,f0εuε0 yields the convergencef0ε→f0u0 inL1(Ω×TN× R).

Let us close this section by stating the main result to be proved precisely.

Theorem 2.1 (Hydrodynamic limit of the stochastic BGK model). Let the above assumptions hold true. Then, for anyε >0, there existsFε∈LP(Ω×

[0, T]×TN×R)which is a unique weak solution to the stochastic BGK model (1.3)with initial conditionF0ε=1uε0. Furthermore, iffε=Fε−10>ξthen (fε)converges inLp(Ω×[0, T]×TN×R), for allp∈[1,∞), to the equilibrium functionχu, whereuis the unique kinetic solution to the stochastic hyperbolic conservation law(1.1). Besides, the local densities(uε)converge to the kinetic solutionuin Lp(Ω×[0, T]×TN), for allp∈[1,∞).

Throughout the paper, we use the letterC to denote a generic positive constant, which can depend on different quantities but ε and may change from one line to another. We also employ a shortened notation for various Lp-type norms, e.g. we writek · kLp

ω,x,ξ for the norm inLp(Ω×TN ×R) and similarly for other spaces.

3. Preliminary results

As we are going to apply the well-posedness theory for kinetic solutions of hyperbolic scalar conservation laws (1.1) as well as the theory of stochas- tic flows generated by stochastic differential equations, we provide a brief overview of these two concepts.

3.1. Kinetic formulation for scalar conservation laws

The main reference for this subsection is the paper of Debussche and Vovelle [5]. For further reading about the kinetic approach used in different settings, we refer the reader to [3], [9], [15], [16], or [20]. In the paper [5], the notion of kinetic and generalized kinetic solution to (1.1) was introduced and the existence, uniqueness and continuous dependence on initial data were proved.

In the following, we present the main ideas and results while skipping all the technicalities.

1B(TN) andB(R), respectively, denotes the Borelσ-algebra onTN andR, respectively.

(7)

Letube a smooth solution to (1.1). It follows from the Itˆo formula that ualso satisfies the kinetic formulation of (1.1)

tF+a(ξ)· ∇F =δu=ξΦ(u) ˙W+∂ξ

m−1

2G2δu=ξ

, (3.1)

where F =1u>ξ andm is an unknown kinetic measure, i.e. a random non- negative bounded Borel measure on [0, T]×TN×Rthat vanishes for largeξ in the following sense: ifBcR={ξ∈R;|ξ| ≥R} then

R→∞lim

Em TN ×[0, T]×BcR

= 0.

Hence we arrive at the notion of kinetic solution: letu∈Lp(Ω×[0, T],P,dP⊗ dt;Lp(TN)),∀p∈[1,∞). It is said to be a kinetic solution to (1.1) provided F = 1u>ξ is a solution, in the sense of distributions over [0, T]×TN ×R, to the kinetic formulation (3.1) for some kinetic measurem. Replacing the indicator function by a general kinetic functionF we obtain the definition of a generalized kinetic solution. It corresponds to the situation where one does not know the exact value of u(t, x) but only its law given by a probability measure νt,x. More precisely, let F(t), t ∈ [0, T], be a kinetic function on Ω×TN ×R and νt,x(ξ) = −∂ξF(t, x, ξ). Then F is a generalized kinetic solution to (1.1) provided: F(0) = 1u0 and for any test function ϕ ∈ Cc([0, T)×TN ×R),

Z T

0

F(t), ∂tϕ(t) dt+

F(0), ϕ(0) +

Z T

0

F(t), a(ξ)· ∇ϕ(t) dt

=−

d

X

k=1

Z T

0

Z

TN

Z

R

gk(x, ξ)ϕ(t, x, ξ) dνt,x(ξ) dxdβk(t)

−1 2

Z T

0

Z

TN

Z

R

G2(x, ξ)∂ξϕ(t, x, ξ)dνt,x(ξ) dxdt+m(∂ξϕ)

(3.2)

holds true P-a.s.. The assumptions considered in [5] are the following: the flux functionA is of classC1with a polynomial growth of its derivative; the processW is a (generally infinite-dimensional) cylindrical Wiener process, i.e.

W(t) =P

k≥1βk(t)ek with (βk)k≥1 being mutually independent real-valued standard Wiener processes and (ek)k≥1 a complete orthonormal system in a separable Hilbert spaceU; the mapping Φ(z) : U → L2(TN) is defined for each z ∈ L2(TN) by Φ(z)ek = gk(·, z(·)) where gk ∈ C(TN ×R) and the following conditions

X

k≥1

|gk(x, ξ)|2≤C 1 +|ξ|2 , X

k≥1

|gk(x, ξ)−gk(y, ζ)|2≤C |x−y|2+|ξ−ζ|h(|ξ−ζ|) ,

are fulfilled for everyx, y ∈TN, ξ, ζ∈R, withhbeing a continuous nonde- creasing function onR+ satisfying, for someα >0,

h(δ)≤Cδα, δ <1.

(8)

Under these hypotheses, the well-posedness result [5, Theorem 11, Theorem 19] states: For anyu0 ∈Lp(Ω×TN) for allp∈[1,∞) there exists a unique kinetic solution to (1.1). Besides, any generalized kinetic solutionFis actually a kinetic solution, i.e. there exists a processusuch thatF =1u>ξ. Moreover, ifu1, u2are kinetic solutions with initial datau1,0andu2,0, respectively, then for allt∈[0, T]

Eku1(t)−u2(t)kL1x≤Eku1,0−u2,0kL1x. 3.2. Stochastic flows and stochastic characteristics method

The results mentioned in this subsection are due to Kunita and can be found in [13] and [14]. To begin with, we introduce some notation. We denote by Cbl,δ(Rd) the space of all l-times continuously differentiable functions with bounded derivatives up to orderl (the function itself is only required to be of linear growth) andδ-H¨older continuousl-th derivatives.

Let Bt = (B1t, . . . , Btm) be an m-dimensional Wiener process and let bk :Rd→Rd, k= 0, . . . , m.We study the following system of Stratonovich’s stochastic differential equations

t=b0t) dt+

m

X

k=1

bkt)◦dBtk. (3.3) Under the hypothesis that b1, . . . , bm ∈ Cbl+1,δ(Rd) and b0 ∈ Cbl,δ(Rd) for some l ≥ 1 and δ > 0, and for any given y ∈ Rd, s ∈ [0, T], the problem (3.3) possesses a unique solution starting fromyat times. Let us denote this solution by φs,t(y). It enjoys several important properties. Namely, it is a continuousCl,ε-semimartingale for anyε < δand defines a forward Brownian stochastic flow ofCl-diffeomorphisms, i.e. there exists a null setN of Ω such that for anyω∈Nc, the family of continuous maps{φs,t(ω); 0≤s≤t≤T} satisfies

(i) φs,t(ω) =φr,t(ω)◦φs,r(ω) for all 0≤s≤r≤t≤T, (ii) φs,s(ω) = Id for all 0≤s≤T,

(iii) φs,t(ω) : Rd → Rd is l-times differentiable with respect to y, for all 0≤s≤t≤T, and the derivatives are continuous in (s, t, y),

(iv) φs,t(ω) :Rd→Rd is aCl-diffeomorphism for all 0≤s≤t≤T, (v) φti,ti+1, i = 0, . . . , n−1, are independent random variables for any

0≤t0≤ · · · ≤tn≤T.

Therefore, for each 0 ≤ s ≤ t ≤ T, the mapping φs,t(ω) has the inverse ρs,t(ω) =φs,t(ω)−1which satisfies

(vi) ρs,t(ω) : Rd → Rd is l-times differentiable with respect to y, for all 0≤s≤t≤T, and the derivatives are continuous in (s, t, y),

(vii) ρs,t(ω) =ρs,r(ω)◦ρr,t(ω) for all 0≤s≤r≤t≤T,

and consequentlyρs,tis a stochastic flow ofCl-diffeomorphisms for the back- ward direction. Indeed, the following holds true: For any 0 ≤ s ≤ t ≤ T, the processρs,t(y) satisfies the backward Stratonovich stochastic differential

(9)

equation with the terminal conditiony ρs,t(y) =y−

Z t

s

b0 ρr,t(y) dr−

m

X

k=1

Z t

s

bk ρr,t(y)

◦dBˆ rk,

where the last term is a backward Stratonovich integral defined by Kunita [14] using the time-reversing method. To be more precise, the Brownian mo- tionB is regarded as a backward martingale with respect to its natural two parametric filtration

σ Br1−Br2;s≤r1, r2≤t

, 0≤s≤t≤T,

the integral is then defined similarly to the forward case and both stochastic flowsφs,tas well asρs,tare adapted to this filtration. Furthermore, we have a growth control for both forward and backward stochastic flow. Fix arbitrary δ∈(0,1), then the following convergences hold uniformly ins, t,P-a.s.,

|y|→∞lim

s,t(y)|

(1 +|y|)1+δ = 0, lim

|y|→∞

s,t(y)|

(1 +|y|)1+δ = 0,

|y|→∞lim

(1 +|y|)δ

1 +|φs,t(y)| = 0, lim

|y|→∞

(1 +|y|)δ 1 +|ρs,t(y)| = 0.

In the remainder of this subsection we will discuss the stochastic charac- teristics method where the theory of stochastic flows plays an important role.

We restrict our attention to a first-order linear stochastic partial differential equation of the form

dv=b0(y)· ∇yvdt+

m

X

k=1

bk(y)· ∇yv◦dBtk, v(0) =v0,

(3.4)

with coefficientsbk:Rd →Rd,k= 0, . . . , m.The associated stochastic char- acteristic system is defined by a system of Stratonovich stochastic differential equations

t=b0t) dt+

m

X

k=1

bkt)◦dBtk, (3.5) A solution of (3.5) starting atyis the so-called stochastic characteristic curve of (3.4) and will be denoted byφt(y). Assume thatb1, . . . , bm∈Cbl+1,δ(Rd) and b0 ∈ Cbl,δ(Rd) for some l ≥3 and δ > 0. If the initial function v0 lies in Cl,δ(Rd), then the problem (3.4) has a unique strong solution which is a continuousCl,ε-semimartingale for someε >0 and is represented by

v(t, y) =v0 φ−1t (y)

, t∈[0, T], (3.6)

where the inverse mapping φ−1t is well defined according to the previous paragraph. It satisfies (3.4) in the following sense

v(t, y) =v0(y) +b0(y)· Z t

0

yv(r, y) dr+

m

X

k=1

bk(y)· Z t

0

yv(r, y)◦dBrk.

(10)

Moreover, if the initial condition v0 is rapidly decreasing then so does the solution itself and

E sup

t∈[0,T]

Z

Rd

|v(t, y)|(1 +|y|)ndy p

<∞, ∀n∈N0, p∈[1,∞).

The choice of the Stratonovich integral is more natural here and is given by application of the Itˆo-Wentzell-type formula in the proof of the explicit representation of the solution (3.6). Indeed, in this case it is close to the classical differential rule formula for composite functions (cf. [13, Theorem I.8.1, Theorem I.8.3]).

4. Solution to the stochastic BGK model

This section is devoted to the existence proof for the stochastic BGK model (1.3). Let us start with the definition of its solution.

Definition 4.1. Let ε > 0. Then Fε ∈ LP(Ω×[0, T]×TN ×R) satisfying Fε−10>ξ∈L1(Ω×[0, T]×TN×R) is called a weak solution to the stochastic BGK model (1.3) with initial conditionF0εprovided the following holds true for a.e.t∈[0, T],P-a.s.,

Fε(t), ϕ

= F0ε, ϕ

+ Z t

0

Fε(s), a· ∇ϕ ds +1

ε Z t

0

1uε(t)>ξ−Fε(t), ϕ(t) dt+

d

X

k=1

Z t

0

Fε(s), ∂ξ(gkϕ) dβk(s) +1

2 Z t

0

Fε(s), ∂ξ(G2ξϕ) ds.

Remark 4.2. In particular, for any ϕ ∈ Cc(TN ×R), there exists a rep- resentative of hFε(t), ϕi ∈L(Ω×[0, T]) which is a continuous stochastic process.

In order to solve the stochastic BGK model (1.3), we intend to employ the stochastic characterics method introduced in the previous section. Hence we need to reformulate the problem in Stratonovich form. It will be seen from the following lemma (see Corollary4.4) that on the level of above defined weak solutions the problem (1.3) is equivalent to

dFε+a(ξ)· ∇Fεdt= 1uε−Fε

ε dt−∂ξFεΦ◦dW+1

4∂ξFεξG2dt, Fε(0) =F0ε.

Lemma 4.3. If X be a C1(TN ×R)-valued continuous (Ft)-semimartingale whose martingale part is given by −Rt

0ξXΦdW, then

− Z t

0

ξXΦdW+1 2

Z t

0

ξ G2ξX dt=−

Z t

0

ξXΦ◦dW+1 4

Z t

0

ξX∂ξG2dt.

(4.1)

(11)

Moreover, the same is valid in the sense of distributions as well: let X be a D(TN×R)-valued continuous(Ft)-semimartingale whose martingale part is given by−Rt

0ξXΦdW, i.e. hX(t), ϕi is a continuous (Ft)-semimartingale with martingale part −Rt

0h∂ξXΦ, ϕidW for any ϕ ∈ Cc(TN ×R). Then (4.1)holds true inD(TN×R).

Proof. We will only prove the second part of the statement as the first one is straightforward and follows similar arguments. Let us recall the relation between Itˆo and Stratonovich integrals (see [13] or [14]). Let Y be a contin- uous local semimartingale and Ψ be a continuous semimartingale. Then the Stratonovich integral is well defined and satisfies

Z t

0

Ψ◦dY = Z t

0

Ψ dY +1

2hhΨ, Yiit,

wherehh·,·iitdenotes the cross-variation process. Therefore, we need to calcu- late the cross variation of−∂ξXgk and the Wiener processβk,k= 1, . . . , d.

Towards this end, we take a test function ϕ∈Cc(TN ×R) and derive the martingale part ofh∂ξXgk, ϕi(in the following, we emphasize only the cor- responding martingale parts).

hX, ϕi=· · · − Z t

0

ξXgk, ϕ dβk(s), hX, gkϕi=· · · −

Z t

0

ξXgk, gkϕ dβk(s), h∂ξX, gkϕi=· · ·+

Z t

0

ξXgk, ∂ξ(gkϕ) dβk(s), where

ξXgk, ∂ξ(gkϕ)

=−

ξ(∂ξXgk), gkϕ

=−

ξ2Xg2k, ϕ

−1 2

ξX∂ξgk2, ϕ

=−

ξ(g2kξX), ϕ +1

2

ξX∂ξgk2, ϕ . Consequently

h−∂ξXgk, ϕi, βk

t= Z t

0

ξ(gk2ξX), ϕ ds−1

2 Z t

0

ξX∂ξg2k, ϕ ds

and the claim follows by summing up overk.

Corollary 4.4. Let ε > 0. If Fε ∈ LP(Ω×[0, T]×TN ×R) is such that Fε−10>ξ ∈ L1(Ω×[0, T]×TN ×R) then it is a weak solution to (1.3) if and only if, for any ϕ ∈ Cc(TN ×R), there exists a representative of hFε(t), ∂ξ(gkϕ)i ∈L(Ω×[0, T])which is a continuous(Ft)-semimartingale

(12)

and the following holds true for a.e. t∈[0, T],P-a.s., Fε(t), ϕ

= F0ε, ϕ

+ Z t

0

Fε(s), a· ∇ϕ ds +1

ε Z t

0

1uε(t)>ξ−Fε(t), ϕ(t) dt+

d

X

k=1

Z t

0

Fε(s), ∂ξ(gkϕ)

◦dβk(s)

−1 4

Z t

0

Fε(s), ∂ξ(ϕ∂ξG2) ds.

As the first step in order to show the existence of a solution to the stochastic BGK model, we shall study the following auxiliary problem:

dX+a(ξ)· ∇Xdt=−∂ξXΦ◦dW+1

4∂ξX∂ξG2dt, X(s) =X0.

(4.2) It will be shown in Corollary4.10that this problem possesses a unique weak solution providedX0∈L(Ω×TN×R). Let

S={S(t, s); 0≤s≤t≤T}

be its solution operator, i.e. for any 0 ≤s ≤t ≤T we define S(t, s)X0 to be the solution to (4.2). Then we have the following existence result for the stochastic BGK model.

Theorem 4.5. For any ε > 0, there exists a unique weak solution of the stochastic BGK model (1.3)and is represented by

Fε(t) = etεS(t,0)F0ε+1 ε

Z t

0

etεsS(t, s)1uε(s)>ξds. (4.3) The proof of Theorem 4.5 will be divided into several steps. First, we have to concentrate on the problem (4.2).

4.1. Application of the stochastic characteristics method

In this subsection, we prove the existence of a unique solution to (4.2) and study the behavior of the solution operatorS. The equation (4.2) is a first- order linear stochastic partial differential equation of the form (3.4), however, the coefficienta, as well as∂ξG2in the case of (2.2), is not supposed to have bounded derivatives. For this purpose we introduce the following truncated problem: let (kR) be a smooth truncation on R, i.e. let kR(ξ) = k(R−1ξ), wherekis a smooth function with compact support satisfying 0≤k≤1 and

k(ξ) =

(1, if |ξ| ≤ 12, 0, if |ξ| ≥1,

and define gkR(x, ξ) = gk(x, ξ)kR(ξ), k = 1, . . . , d, and aR(ξ) = a(ξ)kR(ξ).

CoefficientsΦRandGR,2, respectively, can be defined similarly asΦandG2,

(13)

respectively, usinggkR instead ofgk.2 Then

dX+aR(ξ)· ∇Xdt=−∂ξR◦dW +1

4∂ξX∂ξGR,2dt, X(s) =X0

(4.4) can be solved by the method of stochastic characteristics. Indeed, the sto- chastic characteristic system associated with (4.4) is defined by the following system of Stratonovich’s stochastic differential equations

0t =−1

4∂ξGR,2t) dt+

d

X

k=1

gkRt)◦dβk(t), dϕit=aRi0t) dt, i= 1, . . . , N,

(4.5)

where the processesϕ0t andϕit, i= 1, . . . , N,respectively, describe the evolu- tion of theξ-coordinate andxi-coordinate, i= 1, . . . , N,respectively, of the characteristic curve.

Let us denote by ϕRs,t(x, ξ) the solution of (4.5) starting from (x, ξ) at times. ThenϕRdefines a stochastic flow ofC3-diffeomorphisms and we de- note byψRthe corresponding inverse flow. It is the solution to the backward problem

t0=1

4∂ξGR,2t) ˆdt−

d

X

k=1

gRkt)◦dβˆ k(t), dψit=−aRit0) ˆdt, i= 1, . . . , N.

(4.6)

Remark 4.6. Note, that unlike the deterministic BGK model (i.e. gk = 0, k= 1, . . . , d), the stochastic case is not time homogeneous:ϕRs,t6=ϕR0,t−s. Proposition 4.7. LetR >0. IfX0∈C3,η(TN×R)almost surely,3there exists a unique strong solution to (4.4)which is a continuous C3,ϑ-semimartingale for someϑ >0, i.e. it satisfies (4.4)in the following sense

X(t, x, ξ;s) =X0(x, ξ)−aR(ξ)· Z t

s

∇X(r, x, ξ;s) dr

d

X

k=1

gRk(x, ξ) Z t

s

ξX(r, x, ξ;s)◦dβk(r) +1

4∂ξGR,2(x, ξ) Z t

s

ξX(r, x, ξ;s) dr, Moreover, it is represented by

X(t, x, ξ;s) =X0 ψs,tR(x, ξ) .

2For notational simplicity we writeGR,2 as an abbreviation for GR2

and similarlygR,k 2 instead of gRk2

.

3η >0 is the H¨older exponent from Section2.

(14)

Proof. The above representation formula corresponds to (3.6). It can be shown in a straightforward manner using the Itˆo-Wentzell formula (see [14,

Theorem 6.1.9]).

It is obvious, that the domain of definition of the solution operator to (4.4), hereafter denoted bySR, can be extended to more general functions which do not necessarily fulfil the assumptions of Proposition 4.7. In this case, we define consistently

SR(t, s)X0=X0 ψs,tR(x, ξ)

, 0≤s≤t≤T.

Since diffeomorphisms preserve sets of measure zero the above is well defined also ifX0is only defined almost everywhere. The resulting process cannot be a strong solution to (4.4), however, as it will be seen in Corollary4.9 it can still satisfy (4.4) in a weak sense. In the following proposition we establish basic properties of the operatorSR.

Proposition 4.8. Let R >0. Let SR ={SR(t, s),0 ≤s≤t≤T} be defined as above. Then

(i) SR is a family of bounded linear operators onL1(Ω×TN ×R)having unit operator norm, i.e. for anyX0∈L1(Ω×TN ×R),0≤s≤t≤T,

SR(t, s)X0

L1ω,x,ξ≤ kX0kL1

ω,x,ξ, (4.7)

(ii) SR verifies the semigroup law

SR(t, s) =SR(t, r)◦ SR(r, s), 0≤s≤r≤t≤T,

SR(s, s) = Id, 0≤s≤T.

Proof. Fix arbitrary 0 ≤ s ≤ t ≤ T. The linearity of SR(t, s) follows eas- ily from its definition. In order to prove (4.7), we will proceed in several steps. First, we make an additional assumption upon the initial condition X0, namely,

X0∈L1(Ω×TN×R)∩L(Ω×TN ×R). (4.8) Let us now consider a suitable smooth approximation of X0. In particular, let (hδ) be an approximation to the identity onTN ×R, and (kδ) a smooth truncation onR, i.e. definekδ(ξ) =k(δξ), wherekwas defined at the begin- ning of this subsection. Then the regularizationX0δ, defined in the following way

X0δ(ω) = X0(ω)∗hδ kδ,

is bounded, pathwise smooth and compactly supported and X0δ−→X0 in L1(Ω×TN×R);

X0δ L1

ω,x,ξ ≤ kX0kL1

ω,x,ξ. (4.9) Furthermore, also all the partial derivatives∂ξX0δ, ∂xiX0δ, i= 1, . . . , N, are bounded, pathwise smooth and compactly supported.

(15)

Next, the processXδ=SR(t, s)X0δ is the unique strong solution to (4.4) or equivalently

dX+aR(ξ)· ∇Xdt=−∂ξRdW+1

2∂ξ GR,2ξX dt, X(s) =X0δ

(4.10) which follows by a similar approach as in Lemma4.3. For anyx∈TN, ξ∈R, the above stochastic integral is a well defined martingale with zero expected value. Indeed, for eachk= 1, . . . , d, we have4

E Z T

s

ξXδgRk(x, ξ)

2dr=CE Z T

s

x,ξX0δ ψs,rR (x, ξ)

·∂ξψRs,r(x, ξ)

2dr

≤CE Z T

s

ξψs,rR (x, ξ)

2dr <∞

sincegkR is bounded and the process∂ξψs,rR (x, ξ) solves a backward bilinear stochastic differential equation with bounded coefficients (see [14, Theorem 4.6.5]) and therefore possesses moments of any order which are bounded in 0 ≤ s ≤ r ≤ T, x ∈ TN, ξ ∈ R. Nevertheless, we point out the same is not generally true without the assumption (4.8). In this case, the stochastic integral can happen to be a local martingale only, which would significantly complicate the subsequent steps.

We intend to integrate the equation (4.10) with respect to the variables ω, x, ξ and expect the stochastic integral to vanish. Towards this end, it is needed to verify the interchange of integrals with respect to x, ξ and the stochastic one. We make use of the stochastic Fubini theorem [4, Theorem 4.18]. In order to verify its assumptions, the following quantity

Z

TN

Z

R

E

Z T

s

ξXδgkR(x, ξ)

2dr 12

dξdx

= Z

TN

Z

R

|gkR(x, ξ)|

E

Z T

s

x,ξX0δ ψs,rR (x, ξ)

·∂ξψRs,r(x, ξ)

2dr 12

dξdx should be finite. Recall thatgkR, k= 1, . . . , d,are bounded and the moments of ∂ξψs,rR (x, ξ) are finite and bounded in s, r, x, ξ. Thus, since ∇x,ξX0δ is bounded and pathwise compactly supported it is sufficient to show that so does ∇x,ξX0δ ψRs,r(x, ξ)

. However, this fact follows immediately from the growth control on the stochastic flow ψR. Indeed, all the assertions of [14, Section 4.5], in particular Exercise 4.5.9 and 4.5.10, can be modified in order to obtain corresponding results for the componentψs,rR,0 only. Hence, for any η∈(0,1), we have uniformly ins, r, x, P-a.s.,

|ξ|→∞lim

s,rR,0(x, ξ)|

(1 +|ξ|)1+η = 0, lim

|ξ|→∞

(1 +|ξ|)η

1 +|ψs,rR,0(x, ξ)| = 0.

4Byx,ξwe denote the gradient with respect to the variablesx, ξ.

(16)

Consequently, it yields: for any fixed L > 0, there existsl > 0 such that if

|ξ|> l then it holds uniformly ins, r, x,P-a.s., (1 +|ξ|)η ≤L(1 +|ψs,rR,0(x, ξ)|).

The support ofX0δ as well as∇x,ξX0δ in the variableξis included in [−1δ,1δ].

Therefore, if in addition (1 +|ξ|)η > L(1 + 1δ) then |ψR,0s,r(x, ξ)| > 1δ for all s, r, x, P-a.s., and accordingly∇x,ξX0δ ψRs,r(x, ξ)

= 0 for all s, r, x, P-a.s..

As a consequence, the stochastic Fubini theorem can be applied.

Therefore, integrating the equation (4.10) with respect toω, x, ξ yields E

Z

TN

Z

R

Xδ(t, x, ξ) dξdx+E Z t

s

Z

R

aR(ξ)· Z

TN

∇Xδ(r, x, ξ) dxdξdr

=E Z

TN

Z

R

X0δdξdx+1 2E

Z t

s

Z

TN

Z

R

ξ GR,2(x, ξ)∂ξXδ(r, x, ξ)

dξdxdr where the second term on the left hand side vanishes due to periodic boundary conditions and the second one on the right hand side due to the compact support ofGR,2 in ξ. Hence we obtain

E Z

TN

Z

R

SR(t, s)X0δdξdx=E Z

TN

Z

R

X0δdξdx

where the integrals on both sides are finite. Note, that ifX0δ is nonnegative (nonpositive) then also SR(t, s)X0δ stays nonnegative (nonpositive). There- fore,

SR(t, s)X0δ+

=SR(t, s)(X0δ)+, SR(t, s)X0δ

=SR(t, s)(X0δ), and by splitting the initial data into positive and negative part we obtain that (4.7) is satisfied with equality in this case.

In addition to (4.9), also the convergence SR(t, s)X0δ → SR(t, s)X0

holds true inL1(Ω×TN ×R). Indeed, let us fixδ1, δ2∈(0,1). Then (4.7) is also fulfilled byX0δ1−X0δ2 hence the set{SR(t, s)X0δ; δ∈(0,1)}is Cauchy in L1(Ω×TN×R) and the limit is necessarilySR(t, s)X0since diffeomorphisms preserve sets of zero measure. Finally, application of the Fatou lemma gives (4.7) forX0.

As the next step, we avoid the hypothesis (4.8). LetX0∈L1(Ω×TN×R) and consider the following approximations

X0n=X01|X0|≤n, n∈N. Then clearly

X0n−→X0 in L1(Ω×TN ×R), X0n

L1ω,x,ξ≤ kX0kL1

ω,x,ξ

andX0n∈L(Ω×TN ×R) hence the estimate (4.7) is valid for allX0n. As above, it is possible to show thatSR(t, s)X0n→ SR(t, s)X0inL1(Ω×TN×R) and by the lower semicontinuity of the norm we obtain the claim.

Finally, item(ii)can be shown by the flow property ofψ:

SR(t, r)◦ SR(r, s)X0=X0 ψRs,r ψr,tR(x, ξ)

=X0 ψs,tR(x, ξ)

=SR(t, s)X0.

(17)

Corollary 4.9. Let R > 0. For any Fs⊗ B(TN)⊗ B(R)-measurable initial datum X0 ∈ L(Ω×TN ×R) there exists a unique X ∈ LPs Ω×[s, T]× TN ×R

that is a weak solution to (4.10), i.e. the following holds true for anyφ∈Cc(TN ×R), a.e. t∈[s, T],P-a.s.,

X(t), φ

= X0, φ

+ Z t

s

X(r), aR· ∇φ dr +

d

X

k=1

Z t

s

X(r), ∂ξ(gRkφ)

k(r) +1 2

Z t

s

X(r), ∂ξ(GR,2ξφ) dr.

(4.11)

Furthermore, it is represented byX =SR(t, s)X0.

Proof. Let us start with the proof of uniqueness. Due to linearity, it is enough to prove that anyL-weak solution to (4.10) starting from the originX0= 0 vanishes identically. LetX be such a solution. First, let (hτ) be a symmetric approximation to the identity onTN×Rand test (4.10) byφ(x, ξ) =hτ(y− x, ζ−ξ). (Here, we employ the parameter τ in order to distinguish from the regularization defined in Proposition4.8, which will also be used in this proof.) ThenXτ(t) :=X(t)∗hτ, for a.e.t∈[s, T], satisfies

Xτ(t, y, ζ) =− Z t

s

aR· ∇X(r)τ

(y, ζ) dr−

d

X

k=1

Z t

s

ξX(r)gRkτ

(y, ζ) dβk(r) +1

2 Z t

s

ξ GR,2ξX(r)τ

(y, ζ) dr

hence is smooth in (y, ζ) and can be extended to become continuous on [s, T].

Now, we will argue as in [8, Theorem 20] and make use of the stochastic flow ϕR. From the Itˆo-Wentzell formula for the Itˆo integral [14, Theorem 3.3.1]

(18)

we deduce Xτ t, ϕRs,t(˜y,ζ)˜

=− Z t

s

aR· ∇X(r)τ

ϕRs,r(˜y,ζ)˜ dr

d

X

k=1

Z t

s

ξX(r)gkRτ

ϕRs,r(˜y,ζ)˜ dβk(r) +1

2 Z t

s

ξ GR,2ξX(r)τ

ϕRs,r(˜y,ζ)˜ dr +

Z t

s

∇Xτ r, ϕRs,r(˜y,ζ)˜

·aR ϕR,0s,r(˜y,ζ)˜ dr +

d

X

k=1

Z t

s

ξXτ r, ϕRs,r(˜y,ζ)˜

gkR ϕRs,r(˜y,ζ)˜ dβk(r) +1

2 Z t

s

2ξXτ r, ϕRs,r(˜y,ζ)˜

GR,2 ϕRs,r(˜y,ζ)˜ dr

d

X

k=1

Z t

s

ξ

ξX(r)gkRτ

ϕRs,r(˜y,ζ)˜

gkR ϕRs,r(˜y,ζ)˜ dr

= J1+ J2+ J3+ J4+ J5+ J6+ J7.

As the next step, we intend to show that J1+ J4→0,J2+ J5→0,and J3+ J6+ J7→0 inD(TN×R),P-a.s., asτ→0. Remark, that unlike [8], working with the Stratonovich form of (4.10) would not bring any simplifications here. To be more precise, the Stratonovich version of the Itˆo-Wentzell formula (see [14, Theorem 3.3.2]) is close to the classical differential rule formula for composite functions hence any correction terms (as J6,J7in the Itˆo version) are not necessary; however, due to the dependence onx, ξ of the coefficients gkR, the corresponding Stratonovich integrals would not cancel and therefore in order to guarantee their convergence to zero, one would need to control the correction terms J6,J7anyway.

Let us proceed with the proof of the above sketched convergence. To- wards this end, we employ repeatedly the arguments of the commutation lemma of DiPerna and Lions [6, Lemma II.1]. In particular, in the case of J1+ J4 we obtain for a.e.r∈[s, t],P-a.s., that

aR· ∇Xτ(r)−

aR· ∇X(r)τ

−→0 in D(TN ×R). (4.12) Indeed, since

aR(ξ)· ∇Xτ(r, x, ξ)−

aR· ∇X(r)τ

(x, ξ)

= Z

TN

Z

R

X(r, y, ζ)

aR(ξ)−aR(ζ)

· ∇hτ(x−y, ξ−ζ)dζdy

(19)

and τ|∇hτ|(·) ≤ Ch(·), we obtain the following bound by standard esti- mates on convolutions : for anyφ∈Cc(TN ×R)

DaR· ∇Xτ(r)−

aR· ∇X(r)τ , φE

≤C aR

W1,(R)kX(r)kLp(Kφ)kφkLq(TN×R),

whereKφ⊂TN ×Ris a suitable compact set andp, q∈[1,∞] are arbitrary conjugate exponents. As a consequence, it is sufficient to considerX(r) con- tinuous in (x, ξ) as the general case can be concluded by a density argument.

We have Z

TN

Z

R

X(r, y, ζ)

aR(ξ)−aR(ζ)

· ∇hτ(x−y, ξ−ζ)dζdy

= Z

TN

Z

R

Z 1

0

X(r, y, ζ)DaR ζ+σ(ξ−ζ)

(ξ−ζ)· ∇hτ(x−y, ξ−ζ)dσdζdy

= Z

TN

Z

R

Z 1

0

X r, x−τy, ξ˜ −τζ˜

DaR ξ−(1−σ)τζ˜ζ˜· ∇h(˜y,ζ)dσd˜˜ ζd˜y

−→X(r, x, ξ)DaR(ξ)· Z

TN

Z

R

ζ˜∇h(˜y,ζ)d˜˜ ζd˜y= 0

hence (4.12) follows by the dominated convergence theorem. Moreover, we deduce also that for a.e.r∈[s, t],P-a.s.,

aR ϕR,0s,r

· ∇Xτ r, ϕRs,r

aR· ∇X(r)τ

ϕRs,r

−→0 in D(TN ×R).

(4.13) It can be seen by using the change of variables formula: let JψRs,rdenote the Jacobian of the inverse flowψs,rR , then

DaR ϕR,0s,r

· ∇Xτ r, ϕRs,r

aR· ∇X(r)τ

ϕRs,r , φE

=

DaR· ∇Xτ(r)−

aR· ∇X(r)τ

, φ ψRs,rRs,r

E

≤C aR

W1,(R)kX(r)kLp(K) φ ψRs,r

Rs,r Lq(K)

≤C aR

W1,∞(R)ess sup

s≤r≤T

kX(r)kLp(K)kφkL(K) sup

s≤r≤T

Rs,r

Lq(K)<∞, which holds for a suitably chosen compact set K ⊂ TN ×R as φ(ψs,rR ) is compactly supported inTN ×Rand any conjugate exponents p, q∈[1,∞].

The estimate of sups≤r≤TkJψs,rR kLq(K) is an immediate consequence of the fact that for almost every ω ∈ Ω the mapping (r, x, ξ) 7→ DψRs,r(ω, x, ξ) is continuous due to the properties of stochastic flows (see Subsection3.2,(vi)) and therefore (r, x, ξ)7→Jψs,rR (ω, x, ξ) is bounded on the given compact set [s, T]×K. Having this bound in hand, we infer (4.13) by using density again.

Accordingly, the almost sure convergence J1+ J4→0 inD(TN×R) follows by the dominated convergence theorem.

In order to pass to the limit in the case of J2+ J5, we employ the same approach as above so we will only write the main points of the proof. We

Références

Documents relatifs

• Arnaud Debussche et Julien Vovelle [ DV10b ] ou [ DV14 ] pour leur article de référence, base de réflexion pour mon mémoire de M2 et origine de cette thèse « Scalar

exists a unique L ∞ (0, T ; BV (Ω)) entropy solution of the onservation law satisfying (pointwise on Σ ) the BLN boundary ondition; and that this solution is the limit of the

Tadmor, A kinetic formulation of multidimensional scalar conser- vation laws and related equations, Journal of the American Mathematical Society 7 (1994), no. Miyakawa, A

After a short introduction to scalar conservation laws and stochastic differential equations, we give the proof of the existence of an invariant measure for the stoschastic

We show that the Cauchy Problem for a randomly forced, periodic multi-dimensional scalar first-order conservation law with additive or multiplicative noise is well-posed: it admits

Our result is applied to the convergence of the finite volume method in the companion paper (Dotti and Vovelle in Convergence of the finite volume method for scalar conservation

http://dmsweb.daa.asn.au/files/Info%20for%20Professionals/Texture_Mod_Poster.pd f ;Irlande http://iaslt.ie/docs/public/Irish%20consistency%20descriptors%20poster.pdf ) Après avoir

Main outcomes were the level of consciousness observed at time of ICU discharge and the Glasgow Outcome Scale-Extended (GOS-E) obtained 3 months following the hemorrhage via