• Aucun résultat trouvé

On stochastic modified 3D Navier-Stokes equations with anisotropic viscosity

N/A
N/A
Protected

Academic year: 2021

Partager "On stochastic modified 3D Navier-Stokes equations with anisotropic viscosity"

Copied!
39
0
0

Texte intégral

(1)On stochastic modified 3D Navier-Stokes equations with anisotropic viscosity Hakima Bessaih, Annie Millet. To cite this version: Hakima Bessaih, Annie Millet. On stochastic modified 3D Navier-Stokes equations with anisotropic viscosity. Journal of Mathematical Analysis and Applications, Elsevier, 2018, 462 (1), pp.915-956. �10.1016/j.jmaa.2017.12.053�. �hal-01502048v2�. HAL Id: hal-01502048 https://hal.archives-ouvertes.fr/hal-01502048v2 Submitted on 20 Nov 2017. HAL is a multi-disciplinary open access archive for the deposit and dissemination of scientific research documents, whether they are published 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..

(2) ON STOCHASTIC MODIFIED 3D NAVIER-STOKES EQUATIONS WITH ANISOTROPIC VISCOSITY HAKIMA BESSAIH AND ANNIE MILLET Abstract. Navier-Stokes equations in the whole space R3 subject to an anisotropic viscosity and a random perturbation of multiplicative type is described. By adding a term of Brinkman-Forchheimer type to the model, existence and uniqueness of global weak solutions in the PDE sense are proved. These are strong solutions in the probability sense. The Brinkman-Forchheirmer term provides some extra regularity in the space L2α+2 (R3 ), with α > 1. As a consequence, the nonlinear term has better properties which allow to prove uniqueness. The proof of existence is performed through a control method. A Large Deviations Principle is given and proven at the end of the paper.. 1. Introduction The Navier-Stokes equations describe the time evolution of the velocity u of an incompressible fluid in a bounded or unbounded domain of Rn , n = 2, 3 and are described by: ∂t u − ν ∆u + u · ∇u + ∇p = 0, divu = 0,. u|t=0 = u0 ,. where ν > 0 is the viscosity of the fluid and p denotes the pressure. If existence and uniqueness is known to hold in dimension 2, the case of dimension 3 is still only partially solved. Indeed, there exists a solution in some homogeneous Sobolev space H˙ 1/2 either on a small time interval or on an arbitrary time interval if the norm of the initial condition is small enough. The difficulty in dimension 3 comes from the nonlinear term (u · ∇)u that requires more regularity. However, this regularity is not satisfied by the energy estimates while it is in dimension 2. In particular, the lack of this regularity is essentially the reason the uniqueness cannot be proved for weak solutions. Many regularizations have been introduced to overcome this difficulty. Here, we will discuss only two of them; a regularization by a rotating term u × e3 and a regularization by a Brickman-Forchheimer

(3)

(4)

(5)

(6) term u u. Of course these two different regularizations give rise to different models. One is related to some rotating flows while the other is related to some porous media models. We refer to [19] and the references therein, where the following system has been investigated (in an even more general formulation)

(7)

(8) 2α ∂t u − ν ∆u + u · ∇u + ∇p + a

(9) u

(10) u = f, divu = 0,. u|t=0 = u0 ,. 2000 Mathematics Subject Classification. Primary 60H15, 60F10; Secondary 76D06, 76M35. Key words and phrases. Navier-Sokes equations, anisotropic viscosity, Brinkman-Forchheimer model, nonlinear convectivity, stochastic PDEs, large deviations. Hakima Bessaih was partially supported by NSF grant DMS 1418838. 1.

(11) 2. H. BESSAIH AND A. MILLET. where a > 0 and α > 0 and f is an external force. Under some assumptions on the coefficient α, the authors in [19] prove the existence and uniqueness of global strong solutions. A slightly different regularization has been investigated by Kalantarov and Zelik in [17]; more precisely they studied some versions of the following model: ∂t u − ν ∆u + u · ∇u + g(u) + ∇p = f, div u = 0,. u|t=0 = u0 ,. where g ∈ C 2 (R3 , R3 ) satisfies the following properties: . g 0 (u)v · v ≥ (−K + κ|u|r−1 )|v|2 , ∀u, v ∈ R3 , |g 0 (u)| ≤ C(1 + |u|r−1 ), ∀u ∈ R3 ,. (1.1). where K, C, κ are some positive constants, r ∈ [1, ∞) and u · v stands for the inner product in R3 . When the forcing is of random type, that is f = σ(t, u)dW (t), M. R¨ockner, T. Zhang and X. Zhang tackled a stochastic version of a modification of the previous model (1.1), that they called the tamed stochastic Navier-Stokes equations, in several papers such as [23], and [24]. Let us mention that in both the deterministic and the stochastic versions of (1.1), the solutions are investigated when the regularity of initial condition is at least H 1 and the viscosity acts in all three directions. In this paper, we are interested in the 3D Navier-Stokes equations with anisotropic viscosity that is acting only in the horizontal directions. These models have some applications in atmospheric dynamics where some informations are missing. The relevance of the anisotropic viscosity is explained through the Ekeman law (see e.g. [22] or the introduction of [10]). The aim of this paper is to study an anisotropic Navier-Stokes equation in dimension 3 that is subject to some multiplicative random forcing. More precisely, we consider the following model of a modified 3D anisotropic Navier-Stokes system on a fixed time interval [0, T ] which can be written formally as follows:

(12)

(13) 2α ˙ for (t, x) ∈ [0, T ] × R3 , (1.2) ∂t u − ν ∆h u + u · ∇u + a

(14) u

(15) u + ∇p = σ(t, u) W ∇ · u=0. for (t, x) ∈ [0, T ] × R3 ,. with the initial condition u0 independent of the driving noise W . Here the viscosity ν and the coefficient a of the nonlinear convective term are strictly positive, α > 1, ∂t denotes the time partial derivative, ∆h := ∂12 + ∂22 and ∂i denotes the partial derivative in the direction xi , i = 1, 2, 3. Thus the viscosity is only smoothing in the horizontal directions. ˙ is a As usual the fluid is incompressible, p denotes the pressure; the forcing term σ(t, u) W multiplicative noise driven by an infinite dimensional Brownian motion W which is white in time with spatial correlation. The convective term a|u|2α u is of Brinkman-Forchheimer type and has a regularizing effect which can balance on one hand the vertical partial derivative of the bilinear term to prove existence, and on the other hand provide some control to obtain uniqueness. Note that the space L2α+2 (R3 ) appears naturally in the analysis of (1.2); it is equal to L4 (R3 ) if α = 1. Furthermore, the homogeneous critical Sobolev space H˙ 1/2 for the Navier-Stokes equation is included in L4 . Hence it is natural to impose α > 1. The deterministic counterpart of (1.2), that is equation (1.2) with σ = 0, has been studied by H. Bessaih, S. Trabelsi and H. Zorgati in [5]. The authors have proved that if the ˜ 0,1 , for any T > 0 there exists a unique solution in L∞ (0, T ; H ˜ 0,1 )∩ initial condition u0 ∈ H.

(16) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 3. ˜ 1,1 ) which belongs to C([0, T ], L2 ), for some anisotropic Sobolev spaces which L2 (0, T ; H will be defined in the next section (see (2.1)). We generalize this result by allowing the system to be subject to some random external force whose intensity may depend on the solution u and on its horizontal gradient ∇h u. Note that since no smoothing is provided by a viscosity in the vertical direction, in the anisotropic case, one requires that the initial condition u0 is square integrable as well as its vertical partial derivative. In the deterministic setting (that is σ = 0), replacing the Brinkman-Forchheimer term a|u|2α u by the rotating term 1 u × e3 , J.Y. Chemin, B. Desjardin, I. Gallagher and E. Grenier [9] have studied an anisotropic modified Navier Stokes equation on R3 with a vertical viscosity νv ≥ 0, which is allowed to vanish. Using some homogeneous anisotropic spaces, they have proved that if u0 ∈ H 0,s with s > 23 , there exists 0 depending only on ν and u0 such that for  ∈ (0, 0 ], 1 ∂t u − ν ∆h u + u · ∇u + u × e3 + ∇p = 0, for (t, x) ∈ [0, T ] × R3 ,  ∇ · u = 0 for (t, x) ∈ [0, T ] × R3 , u|t=0 = u0 has a unique global solution in L∞ (0, T ; H 0,s ) ∩ L2 (0, T ; H 1,s ). The dispersive BrickmanForchheimer term is ”larger” than the rotating term used in [9] but the regularity required on the initial condition is weaker and we allow a stochastic forcing term. The paper is organized as follows. In section 2 we describe the functional setting of our anisotropic model and prove some technical properties of the deterministic terms. Several results were already proved in [5] and we sketch the arguments for the sake of completeness. We also describe the random forcing term and the growth and Lipschitz ˜ 0,1 ) assumptions on the diffusion coefficient σ. In section 3 we prove that if u0 ∈ L4 (Ω, H is independent of W and σ satisfies some general assumptions (in particular cases σ may contain some ”small multiple” of the horizontal gradient ∇h u), equation (1.2) has a unique ˜ 0,1 )) ∩ L2 (Ω; L2 (0, T : H ˜ 1,1 )) ∩ L2α+2 (Ω × (0, T ) × R3 ), which solution in L4 (Ω; L∞ (0, T ; H is almost surely continuous from [0, T ] to H, where H denotes the set of square integrable divergence free functions. Examples of such coefficients σ are provided. Since we are working on the whole space R3 , and not on a bounded domain, the martingale approach used in [4], which depends on tightness properties, does not seem appropriate. We use instead the control method introduced in [20] for the 2D Navier-Stokes equation; see also [26], [15], [13] and [24], where this method has been used for the stochastic 2D NavierStokes equations, stochastic 2D general hydrodynamical B´enard models and the stochastic 3D tamed Navier-Stokes equations. In section 4, under stronger assumptions on σ (which may no longer depend on the horizontal gradient ∇h u), we also prove a large deviations 2 ˜ 1,0 result in C([0, √ T ]; H) ∩ L (0, T ; H ) when the noise intensity is multiplied by a small parameter  converging to 0. The proof uses the weak-convergence approach introduced by A. Budhiraja , P. Dupuis and R.S. Ellis in [16] and [6]; see also the references [26], [15], [13] and [24] where this approach, based on the equivalence of the Large Deviations and Laplace principles, is used for various stochastic 2D Hydrodynamical models and the stochastic 3D tamed Navier-Stokes equation. For the sake of completeness, some technical well-posedness result for a stochastic controlled equation and estimates which only depend on the norm stochastic control, whose proofs are similar to that of the original equation in section 3, are given in the appendix. The proof of the weak convergence and compactness arguments, which have also been used in some papers on Large Deviations Principles of stochastic hydrodynamical models, are also described in the appendix.

(17) 4. H. BESSAIH AND A. MILLET. 2. The functional setting 2.1. Some notations. Let us describe some further notations and the functional framework we will use throughout the paper. Given a vector x = (x1 , x2 , x3 ) let xh := (x1 , x2 ) denote the horizontal variable, which does not play the same role as the vertical variable x3 . Due to the anisotropic feature of the model, we use anisotropic Sobolev spaces defined 0 0 as follows: given s, s0 ∈ R let H s,s denote the set of tempered distributions ψ ∈ S (R3 ) such that Z

(18)

(19) 2s 

(20)

(21) 2s0 

(22)

(23)

(24) Fψ(ξ)

(25) 2 dξ < ∞, 1 +

(26) (ξ1 , ξ2 )

(27) 1 +

(28) ξ3

(29) kψk2s,s0 := (2.1) R3. 0. where F denotes the Fourier transform. The set H s,s endowed with the norm k · ks,s0 is a Hilbert space. 3 Set divh u = ∂1 u1 + ∂2 u2 . Note that for u ∈ H 1,0 ∩ H 0,1 ∇ ·u=0. implies. divh u = −∂3 u3 .. (2.2). For exponents p, q ∈ [1, ∞) let k · kp denote the Lp (R3 ) norm while Lph (Lqv ) denotes the space Lp (Rx1 Rx2 , Lq (Rx3 )) endowed with the norm nZ Z

(30) p o1

(31)

(32) φ(xh , x3 )

(33) q dx3 q dxh p . kφkLp (Lqv ) := h. R2. R. Lqv (Lph ). = Lq (Rx3 ; Lp (Rx1 Rx2 )) is defined in a similar way and endowed with the

(34) q 1 R R

(35)

(36) φ(xh , x3 )

(37) p dxh p dx3 q . Note that in the above definitions norm kφkLqv (Lp ) := 2 R R h we may assume that p or q is ∞ changing the norm accordingly. Let V be the space of infinitely differentiable vector fields u on R3 with compact support and satisfying ∇ · u = 0. Let us denote by H the closure of V in L2 (R3 ; R3 ), that is . H = u ∈ L2 (R3 ; R3 ) : ∇ · u = 0 in R3 . The space. The space H is a separable Hilbert space with the inner product inherited from L2 , denoted in the sequel by (., .) with corresponding norm | . |L2 . To ease notations, when no confusion arises let Lp (resp. Lqv (Lph )) also denote the set of triples of functions u = (u1 , u2 , u3 ) such that each component uj belongs to Lp (resp. to Lqv (Lph )), j = 1, 2, 3, that is u ∈ Lp (R3 ; R3 ) (resp. u ∈ Lqv (Lph )(R3 ; R3 )). For non negative indices s, s0 we set  ˜ s,s0 := H s,s0 3 ∩ H and again k · ks,s0 for the corresponding norm. H ˜ 0,1 , that is for u, v ∈ H ˜ 0,1 : We denote by (·, ·)0,1 the scalar product in the Hilbert space H 3 Z 3 Z X X  u , v 0,1 = uj (x) vj (x) dx + ∂3 uj (x) ∂3 vj (x) dx. j=1. R3. j=1. R3. As defined previously, we set ∆h := ∂12 + ∂22 ; integration by parts implies that given u ∈ (H 2,0 )3 we have 3 Z X  ∆h u , u = − |∇h uj |2L2 dx, where ∇h uj = (∂1 uj , ∂2 uj , uj ). j=1. R3. To we write ∇h u to denote the triple of functions (∇h uj , j = 1, 2, 3) so that. ease notation,. ˜ 1,0 . ∆h u , u = −|∇h u|2L2 for u ∈ H.

(38) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 5. ˜ 0,1 and projecting equation Note that as usual, starting with an initial condition u0 ∈ H (1.2) on the space of divergence-free fields, we get rid of the pressure and rewrite the evolution equation as follows:

(39)

(40) 2α ˙ for (t, x) ∈ [0, T ] × R3 , (2.3) ∂t u − ν Ah u + B(u, u) + a

(41) u

(42) u = σ(t, u) W where Ah u = Pdiv ∆h u,. |u|α u = Pdiv (|u|α u),.  B(u, v) = Pdiv u · ∇v ,. (2.4). and Pdiv denotes the projection on divergence free functions. For u ∈ H 1 such that ∇ · u = 0, set B(u) := B(u, u). 2.2. Some properties of the non linear terms. In this section, we describe some properties of the non linear terms B(u) = u · ∇u and |u|2α u in equation (2.3). They will be crucial to obtain apriori estimates and prove global well posedness. First, for u, v, w in the classical (non isotropic) Sobolev space H 1 such that ∇ · u = ∇ · v = ∇ · w = 0, the classical antisymmetry property is satisfied:. B(u, v) , w = − B(u, w) , v , and B(u, v) , v = 0. (2.5) We will prove that under proper assumptions on the initial condition u0 and on the stochastic forcing term, the solution u to the SPDE (2.3) belongs a.s. to the set X defined by ˜ 0,1 ) ∩ L2 (0, T ; H ˜ 1,1 ) ∩ L2(α+1) ((0, T ) × R3 ; R3 ) X := L∞ (0, T ; H (2.6) and endowed with the norm 3 h Z X kukX := ess sup kuj (t)k0,1 + j=1. 0. t∈[0,T ]. T. 1 i 2 kuj (t, .)k21,1 dt + kuj kL2(α+1) ([0,T ]×R3 ) .. For random processes, we set ΩT := Ω × (0, T ) endowed with the product measure dP ⊗ ds on F ⊗ B(0, T ), and   ˜ 1,1 ) ∩ L2(α+1) (ΩT × R3 ; R3 ). ˜ 0,1 ) ∩ L4 Ω; (L2 (0, T ; H (2.7) X := L4 Ω; L∞ (0, T ; H First, let us prove some integral upper estimates of the bilinear term. ˜ 1,0 ) and v ∈ L∞ (0, T ; H) ∩ L2 (0, T ; H ˜ 1,1 ). Lemma 2.1. Let u ∈ L∞ (0, T ; H) ∩ L2 (0, T ; H Then Z T Z T 1 Z T 1

(43).

(44) 2 2 2

(45) B(u(t)), v(t)

(46) dt ≤ C kv(t)k1,1 dt |∇h u(t)|2L2 dt , ess sup |u(t)|L2 0. 0. 0. t∈[0,T ]. (2.8)

(47).

(48)

(49)

(50)

(51)

(52)

(53) B(u(t)) − B(v(t)), (u − v)(t)

(54) ≤ C kvk1,1

(55) ∇h (u(t) − v(t))

(56) 2

(57) (u − v)(t)

(58) 2 , L L Z T Z T 1

(59).

(60) 2

(61) B(u(t)) − B(v(t)), (u − v)(t)

(62) dt ≤ C kv(t)k21,1 dt 0. (2.9). 0. × ess sup |(u − v)(t)|L2 t∈[0,T ]. Z 0. T. 1  2 |∇h (u − v)(t) |2L2 dt . (2.10). ˜ 1,0 and v ∈ H ˜ 1,1 . Proof. Let us prove some upper estimates of B(ϕ, ψ), v for ϕ, ψ ∈ H Since ∇ · ϕ = ∇ · ψ = ∇ · v = 0, using notations similar to that in [5] and part of the.

(63) 6. H. BESSAIH AND A. MILLET. arguments in this reference used to prove the uniqueness of the solution, the antisymmetry (2.5) of B yields. . − B(ϕ, ψ) , v = B(ϕ, v) , ψ = J1 + J2 , (2.11) where 2 X 3 Z 3 Z X X J1 := ϕk (x) ∂k vl (x) ψl (x) dx, J2 := ϕ3 (x) ∂3 vl (x) ψl (x) dx. R3. k=1 l=1. l=1. R3. The Fubini theorem and H¨ older’s inequality applied to the Lebesgue integral with respect to dxh imply that for almost every t ∈ [0, T ]: 2 X 3 Z X |J1 | ≤ |∂k vl (., x3 )|L2 kϕk (., x3 )kL4 kψl (., x3 )kL4 dx3 ≤. h. k=1 l=1 R 2 X 3  X. Z. sup |∂k vl (., x3 )|L2. h. x3. k=1 l=1. h. h. kϕk (., x3 )kL4 kψl (., x3 )kL4 dx3 . h. R. h. The Gagliardo-Nirenberg inequality implies that for almost every x3 ∈ R we have for φ = ϕk (., x3 ) and φ = ψl (., x3 ): 1. 1. kφkL4 ≤ C |∇h φ|L2 2 |φ|L2 2 . h. h. (2.12). h. On the other hand, for almost every x3 ∈ R the Cauchy-Schwarz inequality for the Lebesgue measure on R3 implies for k = 1, 2 and l = 1, 2, 3: Z x3 Z Z x3 d |∂k vl (., z)|2L2 dz = 2 ∂k vl (xh , z) ∂z ∂k vl (xh , z)dxh dz |∂k vl (., x3 )|2L2 = h h −∞ R2 −∞ dz ≤ C|∇h v|L2 |∂3 ∇h v|L2 ≤ Ckvk21,1 . Therefore, the H¨ older inequality with respect to the Lebesgue measure dx3 implies that Z 1  Z 1 4 4 |J1 | ≤C kvk1,1 |∇h ϕ(., x3 )|2L2 dx3 |∇h ψ(., x3 )|2L2 dx3 h h R R Z 1  Z 1 4 4 × |ϕ(., x3 )|2L2 dx3 |ψ(., x3 )|2L2 dx3 h. R. h. R 1 2. 1 2. 1 2. 1 2. ≤C kvk1,1 |∇h ϕ|L2 |∇h ψ|L2 |ϕ|L2 |ψ|L2 .. (2.13). Using once more the Fubini theorem and H¨older’s inequality with respect to dxh we deduce that 3 Z X |J2 | ≤ k∂3 vl (., x3 )kL4 |ϕ3 (., x3 )|L2 kψl (., x3 )kL4 dx3 ≤. h. l=1 R 3  X. sup |ϕ3 (., x3 )|L2. l=1. h. x3. h. Z R. h. k∂3 vl (., x3 )kL4 kψl (., x3 )kL4 dx3 . h. h.  Furthermore, since ∇ · ϕ = 0, we deduce that ∂3 ϕ3 (xh , x3 ) = −divϕh (xh , x3 ) := − ∂1 ϕ1 (xh , x3 )+ ∂2 ϕ2 (xh , x3 ) . Therefore, the Cauchy-Schwarz inequality with respect to the Lebesgue measure on R3 yields for almost every t ∈ [0, T ] and x3 ∈ R: Z x3 Z 2 |ϕ3 (., x3 )|L2 = 2 ϕ3 (xh , z) ∂z ϕ3 (x3 , z) dxh dz h. −∞. R2.

(64) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. Z. x3. = −2 −∞. 7. Z R2. ϕ3 (xh , z) divϕh (xh , z) dxh dz ≤ 2|∇h ϕ|L2 |ϕ|L2 .. Plugging the above upper estimate, using again the Gagliardo-Nirenberg inequality (2.12) for φ = ∂3 vl (., x3 ) and φ = ψl (., x3 ), using the H¨older inequality with respect to the Lebesgue measure dxh we obtain: Z 3 X 1 1 1 1 2 2 |J2 | ≤ C |∇h ϕ|L2 |ϕ|L2 |∇h ∂3 vl (., x3 )|L2 2 |∂3 vl (., x3 )|L2 2 h. R. l=1. h. 1 2. 1 2. × |∇h ψl (t, ., x3 )|L2 |ψl (t, ., x3 )|L2 dx3 1 2. h 1 2 L2. 1 2. h. 1 2. ≤ C |∇h ϕ|L2 |ϕ|L2 |∇h ∂3 v| 1. 1. 1. |∂3 v|L2 |∇h ψ|L2 2 |ψ|L2 2. 1. 1. 1. ≤ C kvk1,1 |∇h ϕ|L2 2 |∇h ψ|L2 2 |ϕ|L2 2 |ψ|L2 2 .. (2.14). The upper estimates (2.11), (2.13) and (2.14) imply the existence of a positive constant C such that 1 1 1 1

(65).

(66)

(67) B(ϕ, ψ), v

(68) ≤ C kvk1,1 |∇h ϕ| 2 2 |∇h ψ| 2 2 |ϕ| 2 2 |ψ| 2 2 . (2.15) L L L L ˜ 1,0 ) and v ∈ L∞ (0, T ; H)∩L2 (0, T ; H ˜ 1,1 ). Since for almost Let u ∈ L∞ (0, T ; H)∩L2 (0, T ; H 0,1 1,1 ˜ ˜ every t ∈ [0, T ] we have u(t, .) ∈ H and v(t, .) ∈ H , using (2.15) for ϕ = ψ = u(t) and H¨older’s inequality with respect to the Lebesgue measure on [0, T ], we obtain Z T Z T 1

(69).

(70) 2 2 2

(71) B(u(t)), v(t)

(72) dt ≤ C kvk 2 |∇ u(t, .)| |u(t, .)| dt ˜ 1,1 ) h L2 L2 L (0,T ;H 0. 0. ≤ C kvkL2 (0,T ;H˜ 1,1 ) ess sup |u(t)|L2 t∈[0,T ]. Z 0. T. |∇h u(t)|2L2 dt. 1 2. .. (2.16). This concludes the proof of (2.8). Expanding B(u(t)) − B(v(t)) and using the antisymmetry property (2.5) we deduce that. . . B(u(t, .)) − B(v(t, .)) , (u − v)(t, .) = B (u − v)(t, .), v(t, .) , (u − v)(t, .) . Using once more the antisymmetry and the upper estimate (2.15) with ϕ = ψ = (u−v)(t), we conclude the proof of (2.9). Integrating (2.9) on [0, T ] and using the Cauchy Schwarz inequality, we deduce (2.10).  Using H¨ older’s inequality with respect to the expected value in the upper estimates of Lemma 2.1, we deduce the following analog for stochastic processes.    ˜ 1,0 ) and v ∈ L4 Ω; L∞ (0, T ; H) ∩ Lemma 2.2. Let u ∈ L4 Ω; L∞ (0, T ; H) ∩L4 Ω; L2 (0, T ; H  ˜ 1,1 ) . Then L4 Ω; L2 (0, T ; H Z T

(73).

(74)

(75) B(u(t)), v(t)

(76) dt ≤ C kvk 4 E ˜ 1,1 )) L (Ω;L2 (0,T ;H 0.  hZ × kukL4 (Ω;L∞ (0,T ;H)) E 0. Z E 0. T

(77). T. i2  1 4 |∇h u(t)|2L2 dt ..

(78)

(79) B(u(t)) − B(v(t)), (u − v)(t)

(80) dt ≤ C kvk 4 ˜ 1,1 )) L (Ω;L2 (0,T ;H. (2.17).

(81) 8. H. BESSAIH AND A. MILLET.  hZ × ku − vkL4 (Ω;L∞ (0,T ;H)) E. T. 0. i2  1 4 |∇h (u − v)(t)|2L2 dt .. (2.18). The following lemma proves upper estimates for the third partial derivatives of the bilinear term; it is essentially contained in [5]. This results shows the crucial role of the other non linear term |u|2α u of (2.3) in the control of the partial derivative ∂3 of the bilinear term. Lemma 2.3. There exists a positive constant C such that for any α ∈ (1, ∞) there exists Cα > 0, 0 , 1 > 0, s ∈ [0, T ] and u ∈ X:

(82). h

(83) 2

(84)

(85) 1

(86)

(87)

(88) |u(s)|α ∂3 u(s)

(89) L2 ∂ B(u(s)), ∂ u(s) ≤ C 0 |∇h ∂3 u(s)|2L2 +

(90) 3

(91) 3 40 i 1 − + Cα 0 −1 1 α−1 |∂3 u(s)|2L2 . (2.19) Proof. We briefly sketch the proof in order. to be self contained.  Since divh ∂3 u(s) = ∂3 divh u(s), the antisymmetry (2.5) yields B u(s), ∂3 u(s) , ∂3 u(s) = 0; hence for s ∈ [0, T ]: 3 Z X. ∂3 B(u(s)), ∂3 u(s) = ∂3 uk (s, x)∂k ul (s, x)∂3 ul (s, x)dx := J¯1 (s) + J¯2 (s), 3 k,l=1 R. where integration by parts with respect to ∂k , k = 1, 2 yields 2 X 3 Z X ¯ ∂k ∂3 uk (s, x) ul (s, x) ∂3 ul (s, x) dx J1 (s) = − −. 3 k=1 l=1 R 2 X 3 Z X. k=1 l=1. J¯2 (s) =. 3 Z X l=1. R3. ∂3 uk (s, x) ul (s, x) ∂k ∂3 ul (s, x) dx,. R3. 3 X 2 ∂3 u3 (s, x) ∂3 ul (s, x) dx = − l=1. Z. 2 divh uh (s, x) ∂3 ul (s, x) dx;. R3. the last identity comes from the fact that ∇ · u(s) = 0. Since α > 1, the H¨older and Young inequalities imply that for functions f, g, h : R3 → R, 0 > 0 and then 1 > 0, we have for some Cα > 0:

(92) Z

(93) . 1 1

(94)

(95) f (x)g(x)h(x)dx

(96) ≤ |f | |g| α L2α |g|1− α α−1 2α |h|L2

(97) R3. L. 1 − α−1 1

(98)

(99) α

(100)

(101) 2  (2.20) ≤ 0 |h|2L2 + |f | g L2 + Cα −1 |g|2L2 . 0 1 40 Using this inequality for f = ul (s), g = ∂3 ul (s) and h = ∂k ∂3 uk (s) (resp. g = ∂3 uk (s), h = ∂k ∂3 ul (s)) we deduce the existence of C > 0 such that for any α > 1, 0 , 1 > 0 and some constant Cα > 0: h

(102) i 1

(103) 2 − α−1 1

(104)

(105) 2 |J¯1 (s)| ≤ C 0

(106) ∇h ∂3 u(s)|2L2 + |u(s)|α ∂3 u(s)

(107) L2 + Cα −1  |∂ u(s)| 2 3 0 1 L . 40 P P R Integration by parts implies that J¯2 (s) = 2 2k=1 3l=1 R3 uk (s, x)∂k ∂3 ul (s, x)∂3 ul (s, x)dx. Using (2.20) with f = uk (s), g = ∂3 ul (s) and h = ∂k ∂3 ul (s), we deduce the existence of C > 0 such that for any α > 1, 0 , 1 > 0 and Cα > 0: h

(108) i 1

(109) 2 − α−1 1

(110)

(111) 2 |J¯2 (s)| ≤ C 0

(112) ∇h ∂3 u(s)|2L2 + |u(s)|α ∂3 u(s)

(113) L2 + Cα −1  |∂ u(s)| 2 3 1 0 L . 40.

(114) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 9. The upper estimates of J¯1 (s) and J¯2 (s) conclude the proof.. . For any regular enough function ϕ : R3 → R3 , let F (ϕ) be the function defined by F (ϕ) = ν∆h ϕ − B(ϕ) − a |ϕ|2α ϕ.. (2.21). The following lemma proves that for u ∈ X (resp. u ∈ X ), F (u) belongs to the dual space ˜ 1,1 ) ∩ L2(α+1) ((0, T ) × R3 ) (resp. to the dual space of L4 (Ω; L2 (0, T ; H ˜ 1,1 )) ∩ of L2 (0, T ; H L2(α+1) (ΩT × R3 )). ˜ 1,1 ) ∩ L2(α+1) ((0, T ) × R3 ; R3 ); then Lemma 2.4. (i) Let u ∈ X and v ∈ L2 (0, T ; H Z T h

(115).

(116)

(117) F (u(t, .)), v(t, .)

(118) dt ≤ C kvk 2 ˜ 1,0 ) kukL2 (0,T ;H ˜ 1,0 ) + kvkL2(α+1) ([0,T ]×R3 ) L (0,T ;H 0. ×. kuk2α+1 L2(α+1) ((0,T )×R3 ). + kvkL2 (0,T ;H˜ 1,1 ) sup |u(t)|L2 t∈[0,T ]. Z 0. T. 1 i. |∇h u(t)|2L2 dt. 2. . (2.22). ˜ 1,1 )) ∩ L2(α+1) (ΩT × R3 ). Then (ii) Let u ∈ X and v ∈ L4 (Ω; L2 (0, T ; H Z T h

(119).

(120)

(121) F (u(t, .)), v(t, .)

(122) dt ≤ C kvk 2 E ˜ 1,0 ) kukL2 (ΩT ;H ˜ 1,0 ) + kvkL2(α+1) (ΩT ×R3 ) L (ΩT ;H 0 i × kuk2α+1 + kvk 4 (Ω;L2 (0,T ;H 1,1 )) kukL4 (Ω;L∞ (0,T ;H)) kukL4 (Ω;L2 (0,T ;H 1,0 )) . ˜ ˜ 2(α+1) 3 L L (ΩT ×R ) (2.23) Proof. (i) Integration by parts and the Cauchy-Schwarz inequality with respect to dt ⊗ dx yield Z TZ

(123)

(124)

(125)

(126) Z T

(127)

(128)

(129)

(130) h∆h u(t, .), v(t, .)idt

(131) =

(132) − ν ∇h u(t, x) ∇h v(t, x) dxdt

(133)

(134) ν 0. 0. R3. ≤ ν kukL2 (0,T ;H˜ 1,0 ) kvkL2 (0,T ;H˜ 1,0 ) . Note that 2α+2 and. 2α+2 2α+1. (2.24). are conjugate H¨older exponents. Since u ∈ L2(α+1) ((0, T )×R3 ), 2(α+1). the function |u|2α u belongs to L 2α+1 ((0, T ) × R3 ) and

(135) Z T Z.

(136) |u(t, x)|2α u(t, x) v(t, x) dx dt ≤ |u|2α u 2(α+1)

(137) 0. R3. L. 2α+1. ((0,T )×R3 ). kvkL2(α+1) ((0,T )×R3 ). ≤ kuk2α+1 kvkL2(α+1) ((0,T )×R3 ) . L2(α+1) ((0,T )×R3 ). (2.25). The inequalities (2.24), (2.8) and (2.25) conclude the proof of (2.22). ˜ 1,1 )) ∩ L2(α+1) (ΩT × R3 ). Then a.s. we may (ii) Let u ∈ X and v ∈ L4 (Ω; L2 (0, T ; H apply part (i) to u(t)(ω) and v(t)(ω). The Cauchy Schwarz and H¨older inequalities with respect to the expectation conclude the proof.  To prove uniqueness lemma which provides. of the solution, we will need the following. an upper estimate of F (u(t, .)) − F (v(t, .)), u(t, .) − v(t, .) for u, v ∈ X and t ∈ [0, T ]. Lemma 2.5. There exists a positive constant κ depending on α, and for any η ∈ (0, ν) a ˜ 1,1 ∩ L2(α+1) (R3 ); positive constant Cη such that for u, v ∈ H. F (u) − F (v), u − v ≤ −η |∇h (u − v)|2L2

(138) 2

(139) α + Cη kvk21,1 |u − v|2L2 − a κ

(140) |u| + |v| (u − v)

(141) L2 . (2.26).

(142) 10. H. BESSAIH AND A. MILLET. Proof. Integration by parts implies that. ν ∆h (u − v) , u − v = −ν |∇h (u − v)|2L2 .. (2.27). It is well-known (see [2]; see also [19] where it is used) that there exists a constant κ depending on α such that 2α κ|u(x) − v(x)|2 |u(x)| + |v(x)|   ≤ |u(x)|2α u(x) − |v(x)|2α v(x) · u(x) − v(x) , which clearly implies: Z a R3.   |u(x)|2α u(x) − |v(x)|2α v(x) . u(x) − v(x) dx

(143) 2

(144) α ≥ a κ

(145) |u| + |v| (u − v)

(146) L2 .. (2.28). Using Young’s inequality in (2.9) we deduce that for any η ∈ (0, ν) there exists Cη > 0 such that

(147)

(148)

(149) hB(u) − B(v) , u − vi

(150) ≤ (ν − η)|∇h (u − v)|2 2 + Cη kvk21,1 |(u − v)|2 2 . L. L. This upper estimate, (2.27) and (2.28) conclude the proof of (2.26).. . 2.3. The stochastic perturbation. We will consider an external random force in equation (2.3) driven by a Wiener process W and whose intensity may depend on the solution u. More precisely, let (ek , k ≥ 1) be an orthonormal basis of H whose elements belong ˜ 0,1 . For integers k, l ≥ 1 with k 6= l, we to H 2 := W 2,2 (R3 ; R3 ) and are orthogonal in H deduce that (∂32 ek , el ) = −(∂3 ek , ∂3 el ) = −(ek , el )0,1 − (ek , el ) = 0. Therefore, ∂32 ek is a constant multiple of ek . Let Hn = span (e1 , · · · , en ) and let Pn (resp. ˜ 0,1 ) to Hn . We deduce that for P˜n ) denote the orthogonal projection from H (resp. H 0,1 ˜ ˜ ˜ 0,1 : u ∈ H we have Pn u = Pn u. Indeed, for v ∈ Hn , we have ∂32 v ∈ Hn and for any u ∈ H (Pn u, v) = (u, v),. and (∂3 Pn u, ∂3 v) = −(Pn u, ∂32 v) = −(u, ∂32 v) = (∂3 u, ∂3 v).. ˜ 0,1 , we have (Pn u, v)0,1 = (u, v)0,1 for any v ∈ Hn ; this proves that Pn Hence given u ∈ H ˜ 0,1 . and P˜n coincide on H ˜ 0,1 -valued Wiener process with covariance operator Q on a Let (W (t), t ≥ 0) be a H ˜ 0,1 to itself filtered probability space (Ω, F, (Ft ), P); that is Q is a positive operator from H whichP is trace class, and hence compact. Let (qk , k ≥ 1) be the set of eigenvalues of Q with k≥1 qk < ∞, and let (ψk , k ≥ 1) denote the corresponding eigenfunctions (that is Qψk = qk ψk ). The process W is Gaussian, has independent time increments, and for ˜ 0,1 , s, t ≥ 0, f, g ∈ H     E (W (s), f )0,1 = 0 and E (W (s), f )0,1 (W (t), g)0,1 = s ∧ t) (Qf, g)0,1 . We also have the following representation ˜ 0,1 ) with Wn (t) = W (t) = lim Wn (t) in L (Ω; H 2. n→∞. n X. 1/2. qk βk (t)ψk ,. (2.29). k=1. where βk are standard (scalar) mutually independent Wiener processes and ψk are the above eigenfunctions of Q. For details concerning this Wiener process we refer to [14]..

(151) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 11. 1. ˜ 0,1 ; then H0 is a Hilbert space with the scalar product Let H0 = Q 2 H 1. 1. (φ, ψ)0 = (Q− 2 φ, Q− 2 ψ)0,1 , ∀φ, ψ ∈ H0 , p ˜ 0,1 is Hilberttogether with the induced norm | · |0 = (·, ·)0 . The embedding i : H0 → H Schmidt and hence compact; moreover, i i∗ = Q. ˜ 0,1 ) ) be the space of linear operators Let L ≡ L(2) (H0 , H) (resp. Le ≡ L(2) (H0 , H ˜ 0,1 ) such that SQ 12 is a Hilbert-Schmidt operator from S : H0 7→ H (resp. S : H0 7→ H ˜ 0,1 to H (resp. from H ˜ 0,1 to itself). Clearly, Le ⊂ L. Set H |S|2L = traceH ([SQ1/2 ][SQ1/2 ]∗ ) =. ∞ X. kSQ1/2 φk k2L2 ,. k=1 ∞ X. |S|2Le = traceH˜ 0,1 ([SQ1/2 ][SQ1/2 ]∗ ) =. |SQ1/2 φk |20,1. (2.30) (2.31). k=1. ˜ 0,1 . Let (·, ·)L and (·, ·) ˜ denote the associated scalar for any orthonormal basis {φk } in H L products. ˜ 1,1 → Le which we put The noise intensity of the stochastic perturbation σ : [0, T ] × H in (2.3) satisfies the following classical growth and Lipschitz conditions (i) and (ii). Note that due to the anisotropic feature of our model, we have to impose growth conditions both for the | · |L and | · |L˜ norms.  ˜ 1,1 ; L) e is a linear operator such Condition (C): The diffusion coefficient σ ∈ C [0, T ]× H that: ˜ i such that for (i) Growth condition There exist non negative constants Ki and K ˜ 1,1 : every t ∈ [0, T ] and u ∈ H |σ(t, u)|2L ≤ K0 + K1 |u|2L2 + K2 |∇h u|2L2 ,  ˜0 + K ˜ 1 kuk2 + K ˜ 2 |∇h u|2 2 + |∂3 ∇h u|2 2 . |σ(t, u)|2Le ≤ K 0,1 L L. (2.32) (2.33). (ii) Lipschitz condition There exists constants L1 and L2 such that: |σ(t, u) − σ(t, v)|2L ≤ L1 |u − v|2L2 + L2 |∇h (u − v)|2L2 ,. ˜ 1,1 . t ∈ [0, T ] and u, v ∈ H. Definition 2.6. An (Ft )-predictable stochastic process u(t, ω) is called a weak solution in C([0, T ]; H) ∩ X for the stochastic equation (2.3) on [0, T ] with initial condition u0 if u ∈ C([0, T ]; H) ∩ X a.s., where X is defined in (2.6), and u satisfies Z th. . i (u(t), v) − (u0 , v) + − ν u(s), ∆h v − B(u(s), v) , u(s) ds 0 Z t Z tZ

(152)

(153) 2α 

(154)

(155) +a u(s, x) u(s, x)v(x)dxds = σ(s, u(s))dW (s), v , a.s., 0. R3. for every test function v ∈ L2(α+1) ([0, T ] × R3 ). 0. H 2 (R3 ). and all t ∈ [0, T ]. All terms are well defined since. u∈ for almost every s ∈ [0, T ]; this implies |u(s)|2α u(s) ∈ L which is the dual space of L2(α+1) (R3 ).. 2(α+1) 2α+1. (R3 ). Furthermore the Gagliardo-Nirenberg inequality implies Dom(−∆) ⊂ Lp (R3 ) for any p ∈ [2, ∞). Note that this solution is a strong one in the probabilistic meaning, that is the trajectories of u are written in terms of stochastic integrals with respect to the given Brownian motion W ..

(156) 12. H. BESSAIH AND A. MILLET. 3. Existence and uniqueness of global solutions The aim of this section is to prove that equation (2.3) has a unique solution in X defined in (2.7). We at first prove local well posedeness of a Galerkin approximation of u and apriori estimates. 3.1. Galerkin approximation and apriori estimates. Let (en , n ≥ 1) be the orthonormal basis of H defined in section 2.3 (that is made of functions in H 2 which are ˜ 0,1 ). Recall that for every integer n ≥ 1 we set Hn := span(e1 , · · · , en ) also orthogonal in H ˜ 0,1 coincides with the and that the orthogonal projection Pn from H to Hn restricted to H 0,1 ˜ orthogonal projection from H to Hn . P √ Let Πn denote the projection in H0 on Q1/2 (Hn ). Let Wn (t) = nj=1 qj ψj βj (t) = Πn W (t) be defined by (2.29). Fix n ≥ 1 and consider the following stochastic ordinary differential equation on the n-dimensional space Hn defined by un (0) = Pn u0 , and for t ∈ [0, T ] and v ∈ Hn :. d(un (t), v) = F (un (t)), v dt + (Pn σ(t, un (t)) Πn dW (t), v). (3.1) Then for k = 1, · · · , n we have for t ∈ [0, T ]: n X 1.  d(un (t), ek ) = F (un (t)), ek dt + qj2 Pn σ(t, un (t))ψj , ek dβj (t). j=1. Note that for v ∈ Hn the map u ∈ Hn 7→ hF (u) , vi is locally Lipschitz. Indeed, H 2 ⊂ L2α+2 and there exists some constant C(n) such that kvkH 2 ≤ C(n)|v|L2 for v ∈ Hn . Let ϕ, ψ, v ∈ Hn ; integration by parts implies that |h∆h ϕ − ∆h ψ, vi| ≤ kϕ − ψk1,0 kvk1,0 ≤ C(n)2 |ϕ − ψ|L2 |v|L2 . In the polynomial nonlinear term, the H¨older and Gagliardo-Nirenberg inequalities imply:

(157) Z

(158) 

(159)

(160) |ϕ(x)|2α ϕ(x)−|ψ(x)|2α ψ(x) v(x)dx

(161)

(162) R3  2α kϕ − ψkL2α+2 kvkL2α+2 + kψk ≤ C(α) kϕk2α 2α+2 2α+2 L L  2(α+1) 2α 2α ≤ C(α) C(n) |ϕ|L2 + |ψ|L2 |ϕ − ψ|L2 |v|L2 . Finally, using (2.15) and integration by parts we deduce:

(163)

(164) |hB(ϕ) − B(ψ), vi| =

(165) − hB(ϕ − ψ, v) , ϕi − hB(ψ, v) , ϕ − ψi

(166)  ≤ C kϕ − ψk1,0 kϕk1,0 + kψk1,0 kvk1,1  ≤ CC(n)3 |ϕ − ψ|L2 |ϕ|L2 + |ψ|L2 |v|L2 .   √ Condition (C) implies that the map u ∈ Hn 7→ qj σ(t, u)ψj , ek : 1 ≤ j, k ≤ n satisfies the classical global linear growth and Lipschitz conditions from Hn to n × n matrices uniformly in t ∈ [0, T ]; indeed, the growth and Lipschitz conditions (2.32) and (C)(ii) imply: p p p

(167)

(168) 

(169)

(170)

(171) σ(t, u)√qj ψj , ek

(172)

(173) σ(t, u)√qj ψj

(174) |ek |L2 ≤ K0 + K1 |u|L2 + K2 |∇h u|L2 H  ≤ C(n) 1 + |u|L2 , p

(175) 

(176) p

(177) [σ(t, u) − σ(t, v)]√qj ψj , ek

(178) ≤ L1 |u − v|L + L2 |∇h (u − v)|2 2 ≤ CC(n)|u − v|L . 2 2 L Hence by a well-known result about existence and uniqueness of solutions  difPto stochastic ferential equations (see e.g. [18]), there exists a maximal solution un = nk=1 (un , ek ek ∈.

(179) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 13. Hn to (3.1), i.e., a stopping time τn∗ ≤ T such that (3.1) holds for t < τn∗ and as t ↑ τn∗ < T , |un (t)|L2 → ∞. The following proposition shows that τn∗ = T a.s., that is provides the (global) existence and uniqueness of the finite dimensional approximations un . It also gives apriori estimates of un which do not depend on n; this will be crucial to prove well posedeness of (2.3). Proposition 3.1. Let u0 be a F0 measurable random variable such that Eku0 k40,1 < ∞, ˜ 2 < 2ν . Then (3.1) has a unique global solution T > 0 and σ satisfy condition (C) with K 21 ∗ (i.e., τn = T a.s.) with a modification un ∈ C([0, T ], Hn ). Furthermore, there exists a constant C > 0 such that: i 2 Z T h Z T  2(α+1) 2 4 kun (s)kL2(α+1) ds ≤ C Eku0 k40,1 + 1 . kun (s)k1,1 ds + sup E sup kun (t)k0,1 + n. 0. 0. t∈[0,T ]. (3.2) Proof. Let un (t) be the maximal solution to (3.1) described above. For every N > 0, set τN = inf{t : kun (t)k0,1 ≥ N } ∧ τn∗ . Itˆo’s formula applied to k .k0,1 and the antisymmetry relation (2.5) of the bilinear term yield that for t ∈ [0, T ]: Z t∧τN Z t∧τN 2 2 2 |∇h ∂3 un (s)|2L2 ds (3.3) |∇h un (s)|L2 ds − 2ν kun (t ∧ τN )k0,1 = kPn u0 k0,1 − 2ν 0. 0. t∧τN. Z − 2a 0. kun (s)k2α+2 ds − 2a(2α + 1) L2α+2. Z 0. t∧τN Z. |un (s, x)|2α |∂3 un (s, x)|2 ds +. R3. 3 X. Tj (t),. j=1. where Z. t∧τN. T1 (t) = −2 h∂3 B(un (s)), ∂3 un (s)i ds, 0 Z t∧τN  σ(s, un (s))dWn (s), un (s) 0,1 , T2 (t) = 2 0 Z t∧τN

(180)

(181)

(182) Pn σ(s, un (s))Πn

(183) 2e ds. T3 (t) = L 0. The growth condition (2.33) implies that Z t∧τN   ˜0 + K ˜ 1 kun (s)k20,1 + K ˜ 2 |∇h un (s)|2 2 + |∂3 ∇h un (s)|2 2 ds, T3 (t) ≤ K L L 0. while (2.19) in Lemma 2.3 yields the existence of positive constants C, Cα , 0 and 1 such that Z t∧τN h Z t∧τN

(184)

(185) 1 2

(186) |un (s)|α ∂3 un (s)

(187) 2 2 ds |T1 (t)| ≤ 2C 0 |∇h ∂3 un (s)|L2 ds + L 40 0 0 Z i t∧τN 1 − α−1 + Cα −1 |∂3 un (s)|2L2 ds . 0 1 0. Finally, the Burkholder-Davies-Gundy and Young inequalities as well as (2.33) imply that for β ∈ (0, 1):

(188) Z s∧τN  nZ t∧τN

(189) o1

(190) 

(191)

(192) 

(193)

(194) Pn σ(r, un (r))Πn

(195) 2ekun (r)k20,1 dr 2 E sup

(196) 2 σ(r, un (r))dWn (r), un (r) 0,1

(197) ≤ 6E L s≤t. 0. 0.

(198) 14. H. BESSAIH AND A. MILLET.  ≤β E +. sup s≤inf t∧τN Z t∧τN . 9 E β. kun (s)k20,1. .  ˜0 + K ˜ 1 kun (s)k20,1 + K ˜ 2 |∇h un (s)|2 2 + |∂3 ∇h un (s)|2 2 ds. K L L. 0.  ˜ 2 < and  ∈ (0, 2ν − 10K ˜ 2 > , ˜ 2 ), we may choose β ∈ (0, 1) such that 2ν − 9 + 1 K If K β  C then 0 > 0 such that 2C0 < 2 , and finally 1 > 0 such that 2a(2α + 1) − 21 0 > . For this choice of constants, the inequality kPn u0 k0,1 ≤ ku0 k0,1 and the above upper estimates yield (neglecting some non negative terms in the left hand side of (3.3)):    ˜0 9 + 1 (1 − β)E sup kun (s ∧ τN )k20,1 ≤ Eku0 k20,1 + T K β s∈[0,t] Z h i t  ˜ 1 9 + 1 + 2CCα E ||un (s ∧ τN )k2 ds. (3.4) + K 0,1 1/(α−1) β 0 0 1  Gronwall’s lemma implies that E sups∈[0,T ] kun (s ∧ τN )k20,1 ≤ C for some constant C which does not depend on n and N . Note that kφk21,1 = kφk20,1 + |∇h φ|2L2 + |∂3 ∇h φ|2L2 . We use (3.4) and the upper estimates of Ti (t) for i = 1, 2, 3 for the same choice of constants β, 0 and 1 ; this yields Z τN     2 E sup kun (s ∧ τN )k0,1 + E kun (s)k21,1 + kun (s)k2α+2 ds ≤ C 1 + Eku0 k20,1 (3.5) L2α+2 ν 5. 0. s∈[0,T ]. ˜ i , i = 0, 1, 2, β, 0 and 1 but independent of for some positive constant C depending on K n and N . Apply once more the Itˆ o formula to the square of k . k20,1 . This yields Z t∧τN   4 4 kun (s)k20,1 |∇h un (s)|2L2 + |∂3 ∇h un (s)|2L2 ds kun (t ∧ τN )k0,1 = kPn u0 k0,1 − 4ν 0 Z t∧τN − 4a kun (s)k20,1 kun (s)k2α+2 ds L2α+2 0. Z. t∧τN. − 4a(2α + 1) 0. 4 X

(199) 2

(200) T˜j (t), kun (s)k20,1

(201) |un (s)|α ∂3 un (s)

(202) L2 ds +. (3.6). j=1. where we let Z t∧τN ˜ T1 (t) = − 4 h∂3 B(un (s)), ∂3 un (s)i kun (s)k20,1 ds, 0 Z t∧τN  ˜ T2 (t) = 4 Pn σ(s, un (s))dWn (s), un (s) 0,1 kun (s)k20,1 , 0 Z t∧τN

(203)

(204)

(205) Pn σ(s, un (s))Πn

(206) 2e kun (s)k20,1 ds, T˜3 (t) = 2 L 0 Z t∧τN

(207)

(208) 

(209) Πn σ(s, un (s))Pn ∗ un (s)

(210) 2 kun (s)k20,1 ds. T˜4 (t) = 4 0 0. The growth condition (2.33) implies that Z t∧τN   ˜ ˜ ˜ 0 +K ˜ 1 kun (s)k20,1 +K ˜ 2 |∇h un (s)|2 2 +|∂3 ∇h un (s)|2 2 kun (s)k20,1 ds, T3 (t)+T4 (t) ≤ 6 K L L 0.

(211) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 15. while (2.19) implies Z t∧τN  1

(212) 2 − α−1 1

(213)

(214) 2  |un (s)|α ∂3 un (s)

(215) L2 + Cα −1 |T˜1 (t)| ≤ 4C |∂ u (s)| 0 |∇h ∂3 un (s)|2L2 + 2 3 n 0 1 L 40 0 × kun (s)k20,1 ds. The Burkholder-Davies-Gundy inequality, the growth condition (2.33) and Young’s inequality imply that for β ∈ (0, 1):   n Z t∧τN

(216) o1

(217)

(218) σ(r, un (r))

(219) 2e kun (r)k60,1 dr 2 E sup T˜2 (s) ≤ 12E s≤t. L. 0.   ≤ βE sup kun (s)k40,1 +. s≤t∧τN Z t∧τN . 36 E β. 0.  ˜0 + K ˜ 1 kun (s)k2 + K ˜ 2 |∇h un (s)|2 2 + |∂3 ∇h un (s)|2 2 kun (s)k2 ds. K 0,1 0,1 L L. ˜ 2 < 2ν we may choose β ∈ (0, 1) and  > 0 such that  < 4ν − 6 1 + 6/β)K ˜ 2 , then If K 21 4C 0 > 0 such that 4C0 < 2 , and finally 1 > 0 such that 401 +  < 4a(2α + 1). For this choice of constants, neglecting some non positive integrals in the right hand side of (3.6), we deduce:    Z t∧τN   4 (1 − β)E sup kun (s ∧ τN )k0,1 + E kun (s)k20,1 |∇h un (s)|2L2 + |∂3 ∇h un (s)|2L2 ds 2 0 s∈[0,t] Z t Z t h 36  ˜ 36 i ˜ ≤ Eku0 k40,1 + 6 + K1 E kun (s ∧ τN )k40,1 ds + 6 + K0 E kun (s ∧ τN )k20,1 ds. β β 0 0  This inequality, (3.5) and Gronwall’s lemma yield supn E sups∈[0,T ] kun (s ∧ τN )k40,1 < ∞. We deduce the existence of a constant C, which does not depend on n and N , such that: Z τN    4 kun (s)k21,1 kun (s)k20,1 ds ≤ C 1 + Eku0 k40,1 . (3.7) E sup kun (s ∧ τN )k0,1 + E 0. s∈[0,T ]. We now prove that (3.2) holds. As N → ∞, the sequence of stopping times τN increases to τn∗ , and on the set {τn∗ < T } we have sups∈[0,τN ] kun (s)k0,1 → +∞. Hence (3.5) proves that P (τn∗ < T ) = 0 and that for almost every ω, for N (ω) large enough, τN (ω) (ω) = T . The monotone convergence theorem used in (3.5) and (3.7), we deduce the following upper estimates for some constant which does not depend on n: Z T     2 ds ≤ C 1 + Eku0 k20,1 , (3.8) E sup kun (s)k0,1 + E kun (s)k21,1 + kun (s)k2α+2 L2α+2 0. s∈[0,T ]. Z   4 E sup kun (s)k0,1 + E. T.  kun (s)k21,1 kun (s)k20,1 ds ≤ C 1 + Eku0 k40,1 .. (3.9). 0. s∈[0,T ]. To complete the proof and check (3.2), we finally prove that

(220) 2  

(221) Z T 

(222)

(223) kun (s)k21,1 ds

(224) ≤ C 1 + Eku0 k40,1 . sup E

(225) n. (3.10). 0. ˜ 2 < 2ν, The identity (3.3) and the upper estimates of T1 (t) and T3 (t) imply that for K ˜ 2 and 1 small enough we have for every t ∈ [0, T ] and : 2C0 < K Z t∧τN  2 ˜ kun (t ∧ τN )k0,1 + (2ν − K2 ) |∇h un (s)|2L2 + |∂3 ∇h un (s)|2L2 ds 0.

(226) 16. H. BESSAIH AND A. MILLET. ≤ ku0 k20,1 + sup |T2 (s)| + J(t),. (3.11). s≤t. where for some positive constant C: Z t∧τN h ˜ 1 kun (s)k20,1 + K ˜0 + J(t) = K 0. Z t∧τN i 2 |∂ u (s)| ds ≤ C kun (s)k20,1 ds. L2 1/(α−1) 3 n 0 0 1 2CCα. ˜ 2 < 2ν, using the Doob and Cauchy Schwarz inequalities as well as (2.33), we Hence for K deduce: Z τN hn  o2 i 2 ˜ E sup kun (s ∧ τN )k0,1 + (2ν − K2 ) |∇h un (s)|2L2 + |∂3 ∇h un (s)|2L2 ds s≤T. 0.   ≤ 3E(J(T ) ) + 3E sup T22 (s) + 3E(ku0 k40,1 ) s≤T Z τN Z τN 4 kun (s)k20,1 |σ(un (s))Πn |2L˜ds + 3E(ku0 k40,1 ) kun (s)k0,1 ds + 3CE ≤ 3CT E 0 Z τ0N   2 ˜ 2 kun (s)k kun (s)k2 + (K ˜ 1 + T )kun (s)k4 + K ˜ 0 kun (s)k2 ds 2K ≤ 3C E 2. 0,1. 1,1. 0,1. 0,1. 0. + 3E(ku0 k40,1 ). Let N → ∞ in this equation. Since τN (ω) (ω) = T for N (ω) large enough, the above inequality where τN is replaced by T (which is deduced by means of the monotone convergence theorem) coupled with (3.8) and (3.9) yield (3.10). This completes the proof.  3.2. Well posedeness of equation (2.3). The aim of this section is to prove that if the ˜ 0,1 ), equation (2.3) has a unique (weak) solution in the space initial condition u0 ∈ L4 (Ω; H X which belongs a.s. to C([0, T ]; H), where X has been defined in (2.7). ˜ 2 < 2ν and u0 be independent of Theorem 3.2. Let σ satisfy condition (C) with K 21 (W (t), t ≥ 0) such that E(ku0 k40,1 ) < ∞. Then there exists a weak solution u ∈ X to (2.3) with initial condition u0 . This solution belongs to C([0, T ]; H) a.s. Furthermore, there exists a constant C > 0 such that this solution satisfies the following upper estimate:  Z T 2 Z T Z   4 2 E sup ku(t)k0,1 + ku(t)k1,1 dt + |u(t, x)|2(α+1) dxdt ≤ C 1 + Eku0 k40,1 . 0≤t≤T. 0. 0. R3. (3.12) If L2 < 2ν, then (2.3) has a pathwise unique weak solution in X which belongs a.s. to C([0, T ]; H). Proof. The proof is decomposed in several steps. Recall that L is defined by (2.31) and that σ satisfies (2.32). Step 1: Weak convergence of the solution The inequalities (3.2) and (2.23)  imply the existence of a subsequence of (un , n ≥ 1) (resp. of Pn σ(., un ) ◦ Πn , n ≥ 1 and of  F (un ), n ≥ 1 ), still denoted by the same notation, of processes u ∈ X (resp. S˜ ∈    ˜ 1,1 ) ∩ L2(α+1) (ΩT × R3 ) ∗ ), and finally of a random L2 (ΩT ; L) and F˜ ∈ L4 Ω; L2 (0, T ; H ˜ 0,1 ), for which the following properties hold: variable u ˜(T ) ∈ L2 (Ω; H  ˜ 1,1 ) ∩ L2(α+1) (ΩT × R3 ), (i) un → u weakly in L4 Ω; L2 (0, T ; H  ˜ 0,1 ) , (ii) un is weak star converging to u in L4 Ω; L∞ 0, T ; H ˜ 0,1 ), (iii) un (T ) → u ˜(T ) weakly in L2 (Ω; H.

(227) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 17.    ˜ 1,1 ) ∩ L2(α+1) (ΩT × R3 ) ∗ , (iv) F (un ) → F˜ weakly in L4 Ω; L2 (0, T ; H (v) Pn σ(., un (.))Πn → S˜ weakly in L2 (ΩT ; L). Indeed, (i) and (ii) are straightforward consequences of Proposition 3.1, of (3.2), and of RT uniqueness of the limit of E 0 (un (t), v(t))dt for appropriate v. The upper estimate (2.23) proves (iv). The definition of Pn , Πn , the growth condition (2.32) and (3.2) imply: Z T Z T   |Pn σ(s, un (s))Πn |2L ds ≤ sup E K0 + K1 |un (s)|2L2 + K2 |∇h un (s)|2L2 ds < ∞. sup E n. n. 0. 0. This proves (v). Finally, (3.7) and the equality τN = T a.s. imply that supn Ekun (T )k40,1 < ∞, which proves (iii). Furthermore, properties (i) and (ii) and (3.2) imply that i 2 Z T h Z T 2 ku(s)k2α+2 ds ≤ C(1 + Eku0 k40,1 ), ku(s)k1,1 ds + E 2α+2 L 0 0   E sup ku(s)k40,1 ≤ C(1 + Eku0 k40,1 ). s∈[0,T ]. Step 2: An equation for the weak limits The approach is that used in [21]. We prove that u ˜(T ) = u(T ) a.s. and that for t ∈ [0, T ]: Z t Z t ˜ u(t) = u0 + F˜ (s)ds + S(s)dW (s). (3.13) 0. 0. For δ > 0, let f ∈ H 1 (−δ, T +δ) be such that kf k∞ = 1, f (0) = 1 and for any integer j ≥ 1 set gj (t) = f (t)ej , where {ej }j≥1 is the previous orthonormal basis of H made of elements ˜ 0,1 , such that for every n ≥ 1, Hn = span (e1 , · · · , en ). of H 2 which are also orthogonal in H The Itˆ o formula implies that for any j ≥ 1, and for 0 ≤ t ≤ T : 3   X i un (T ) , gj (T ) = un (0) , gj (0) + In,j ,. (3.14). i=1. where 1 In,j. Z. T. =. (un (s), ej ) f (s)ds, 0. 3 In,j. 0. Z =. 2 In,j. Z. T. hF (un (s)), gj (s)ids,. = 0. T.  Pn σ(s, un (s))Πn dW (s), gj (s) .. 0. We study the convergence of all terms in (3.14). Since f 0 ∈ L2 (0, T ) and α > 1, for every  ˜ 0,1 ) ⊂ Z ∈ L2(α+1) (Ω) ⊂ L4 (Ω), the map (t, ω) 7→ ej Z(ω) f 0 (t) belongs to L4 Ω; L2 (0, T ; H 4 ˜ 0,1 )). Hence, the weak-star convergence (ii) above implies that as n → ∞, L 3 (Ω; L1 (0, T ; H   R 1 → T u(s), e f 0 (s)ds weakly in L2(α+1) (Ω) . In,j j 0 Furthermore, the Gagliardo-Nirenberg inequality implies that H 1 (−δ, T +δ) ⊂ L2(α+1) (0, T ) ˜ 1,1 ); and H 2 ⊂ L2(α+1) (R3 ). Hence (t, ω) 7→ gj (t)Z(ω) ∈ L2(α+1) (ΩT ×R3 )∩L4 (Ω; L2 (0, T ; H  R 2 → T hF ˜ (s), gj (s)ids weakly in L2(α+1) (Ω) . therefore, (iv) implies that as n → ∞, In,j 0.

(228) 18. H. BESSAIH AND A. MILLET. 3 , as in [26] (see also [13]), let P denote the class of To prove the convergence of In,j T 2 predictable processes in L (ΩT , L) with the inner product Z T Z T  (G, J)PT = E G(s), J(s) L ds = E traceH (G(s)QJ(s)∗ ) ds. 0. 0.  RT The map T : PT → defined by T (G)(t) = 0 G(s)dW (s), gj (s) is linear and continuous because of the Itˆ o isometry. Furthermore, (v) shows that for every G ∈ PT ,  ˜ as n → ∞, Pn σ(., un (.))Πn , G P → (S(.), G)PT weakly in L2 (Ω). Hence, as n → ∞, T   RT RT P σ(s, u (s)) Π dW (s) , g (s) converges to 0 S˜s dW (s) , gj (s) . n n n j 0 Finally, as n → ∞, Pn u0 = un (0) → u0 in H. By (iii), (un (T ), gj (T )) converges to (˜ u(T ), gj (T )) weakly in L2 (Ω). Therefore, as n → ∞, (3.14) leads to: L2 (Ω). (˜ u(T ), ej ) f (T ) =. Z. . T. . Z. 0. u(s), ej f (s)ds +. u0 , ej + Z. T. hF˜ (s), gj (s)ids. 0. 0.  ˜ S(s)dW (s), gj (s) a.s.. +. T. (3.15). 0. Choosing f in an appropriate way, we next prove a similar identity for any fixed t ∈ [0, T ]. For δ > 0, k > 1δ , t ∈ [0,T ], let fk ∈ H 1 (−δ, T + δ) be such that kfk k∞ = 1, fk = 1 on (−δ, t − k1 ) and fk = 0 on t, T + δ . Then fk → 1(−δ,t) in L2 , and fk0 → −δt in the sense of distributions. Hence as k → ∞, (3.15) written with f := fk yields Z t Z t   ˜ ˜ 0 = u0 − u(t), ej + hF (s), ej ids + S(s)dW (s), ej 0. 0. for almost all (ω, t) ∈ ΩT . Here, the weak continuity (after some modification) of u(t) in H for almost all ω ∈ Ω is deduced by using Lemma 1.4 in Chapter III in Temam [27]. Indeed, it is easy to see that (3.15) provides weak continuity with values in H −1 . Using the fact that the solution is also a.s L∞ (0, T ; H), Lemma 1.4 from [27] provides that the solution is a.s. in Cw ([0, T ]; H). RT 2 ds < ∞; hence for 0 ≤ t ≤ T and almost every ˜ Note that j is arbitrary and E 0 |S(s)| L Rt ω, we deduce (3.13). Moreover 0 F˜ (s)ds ∈ H a.s. Let f = 1(−δ,T +δ) ; using again (3.15) we obtain Z T Z T ˜ u ˜(T ) = u0 + F˜ (s)ds + S(s)dW (s). 0. 0. This equation and (3.13) yield that u ˜(T ) = u(T ) a.s. Step 3: Identification of the limits In (3.13) we still have to prove that dP ⊗ ds a.e. on ΩT , we have: ˜ S(s) = σ(s, u(s)) and F˜ (s) = F (u(s)) . To establish these relations we use the same idea as in [20] (see also [26]). More precisely, we introduce a discounting factor which enables us to cancel out terms where both elements in the scalar products depend on t. Let v ∈ X , where X has been defined in (2.7). Since σ satisfies the Lipschitz condition (C)(ii) with a constant L2 < 2ν, we may choose η ∈ (0, ν) such that L2 < 2η. For this choice of η, let Cη > 0 be defined by (2.26) and for every t ∈ [0, T ], set Z th i r(t) = Cη kv(s)k21,1 + L1 ds. (3.16) 0.

(229) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 19. Then almost surely, 0 ≤ r(t) < ∞ for all t ∈ [0, T ]. Moreover, we also have that   r ∈ L1 Ω; L∞ (0, T ) , e−r ∈ L∞ (ΩT ), r0 ∈ L1 (ΩT ), r0 e−r ∈ L2 Ω; L1 (0, T ) .. (3.17). The weak convergence in (iii) and the property Pn u0 → u0 in H imply that h i   E |u(T )|2L2 e−r(T ) − E|u0 |2L2 ≤ lim inf E |un (T )|2L2 e−r(T ) − E|Pn u0 |2L2 .. (3.18). n. We now apply Itˆ o’s formula to |φ(t)|2L2 e−r(t) for φ = u and φ = un . This gives the relation Z T Z T .  −r(s) 2 2 −r(T ) 2 r0 (s)e−r(s) |φ(s)|2L2 ds, e d |φ(s)|L2 − E E |φ(T )|L2 e − E|φ(0)|L2 = E 0. 0. which can be justified due to (3.17) and the property ∈ L1 (Ω, L∞ ((0, T )) for both choices of φ. Using (3.13), (3.1) and letting u = v + (u − v) after simplification, from (3.18) we obtain Z T

(230) 2    

(231) 2 ˜ e−r(s) − r0 (s)

(232) u(s) − v(s)

(233) L2 + 2 u(s) − v(s) , v(s) } + 2hF˜ (s), u(s)i + |S(s)| E L ds |φ|2. 0. ≤ lim inf Xn ,. (3.19). n. where Z. T.

(234) 2  

(235)  e−r(s) − r0 (s)

(236) un (s) − v(s)

(237) L2 + 2 un (s) − v(s) , v(s) 0  + 2hF (un (s)), un (s)i + |Pn σ(s, un (s))Πn |2L ds. P We write Xn = Yn + 3i=1 Zni , where Yn need not converge but is non positive, while the sequences Zni , i = 1, 2, 3 converge as n → ∞. The upper estimate in (2.26) and the Lipschitz condition (C)(ii) imply that for s ∈ [0, T ] and L2 < 2η < 2ν:

(238)

(239) 2 2hF (un (s)) − F (v(s)) , un (s) − v(s)i +

(240) Pn σ(s, un (s))Πn − Pn σ(s, v(s))Πn

(241) L

(242) 

(243) 2 ≤ −2η

(244) ∇h un (s) − v(s)

(245) L2 + Cη kv(t)k21,1 |un (s) − v(s)|2L2 + |σ(s, un (s)) − σ(s, v(s))|2L  ≤ −(2η − L2 )|∇h (un (s) − v(s)|2L2 + Cη kv(s)k21,1 + L1 |un (s) − v(s)|2L2 . Xn = E. Hence the definition of r in (3.16) implies that Z T  Yn := E e−r(s) − r0 (s)|un (s) − v(s)|2L2 + 2hF (un (s)) − F (v(s)), un (s) − v(s)i 0

(246)  

(247) 2 i +

(248) Pn σ(s, un (s)) − σ(s, v(s)) Πn

(249) L ds ≤ 0. (3.20) P Furthermore, Xn = Yn + 3j=1 Znj , where Z T h  Zn1 = E e−r(s) − 2r0 (s) un (s)) − v(s), v(s) + 2hF (un (s)), v(s)i + 2hF (v(s)), un (s)i 0  i − 2hF (v(s)), v(s)i + 2 Pn σ(s, un (s))Πn , σ(s, v(s)) L ds, Z T  2 Zn = 2 E e−r(s) Pn σ(s, un (s))Πn , Pn σ(s, v(s))Πn − σ(s, v(s)) L ds, 0. Zn3. Z =−E 0. T.

(250)

(251) 2 e−r(s)

(252) Pn σ(s, v(s))Πn

(253) L ds..

(254) 20. H. BESSAIH AND A. MILLET. ˜ We next study the convergence of Znj , j = 1, 2, 3, and first prove that S(s) = σ(s, u(s)) a.e.  0 −r 2 1 0,1 ˜ on ΩT . The definition of X and (3.17) imply that r e v ∈ L Ω; L (0, T ; H ) . Hence the weak star convergence (ii) implies that as n → ∞: Z T Z T   e−r(s) r0 (s) u(s) − v(s) , v(s) ds. e−r(s) r0 (s) un (s) − v(s) , v(s) ds → E E 0. 0. ∗ Since (2.23) implies that F (v) ∈ ∩ L2(α+1) (ΩT × R3 ) , the weak RT RT convergence (i) implies that E 0 e−r(s) hF (v(s)), un (s)ids → E 0 e−r(s) hF (v(s)), u(s)ids. ˜ 1,1 )) ∩ L2(α+1) (ΩT × R3 ), the weak convergence (iv) implies that Since v ∈ L4 (Ω; L2 (0, T ; H R T −r(s) RT E 0 e hF (un (s)), v(s)ids → E 0 e−r(s) hF˜ (s), v(s)ids. Finally, the weak convergence (v) implies that as n → ∞: Z T Z T   ˜ , σ(s, v(s)) ds. E e−r(s) Pn σ(s, un (s))Πn , σ(s, v(s)) L ds → E e−r(s) S(s) L ˜ 1,1 )) L4 (Ω; L2 (0, T ; H. 0. 0. Hence as n → ∞, Z T h  1 Zn → E e−r(s) − 2r0 (s) u(s) − v(s), v(s) + 2hF˜ (s), v(s)i + 2hF (v(s)), u(s)i 0  i ˜ , σ(s, v(s)) ds. (3.21) − 2hF (v(s)), v(s)i + 2 S(s) L P For almost every (ω, t) ∈ ΩT and any orthonormal basis ψj of H0 , j≥n+1 qj |σ(s, v(s))ψj |2L2 converges to 0 as n → ∞. This sequence is dominated by |σ(s, v(s))|2L which belongs to L1 (P ) by means of the growth condition (2.32) and the definition of X . Furthermore, the inequality |Pn σ(s, un (s)) ◦ Πn |L ≤ |σ(s, un (s))|L , the growth condition (2.32), (3.7) and the Cauchy-Schwarz inequality yield  Z T 1  Z T 1 X 2 2 Zn2 ≤ E e−r(s) |Pn σ(s, un (s)) ◦ Πn |2L ds E e−r(s) qj |σ(s, v(s))ψj |2L2 ds . 0. 0. j≥n+1. In the above right hand side, the first factor remains bounded, while as n → ∞ the second one converges to 0 by the dominated convergence theorem. This yields Zn2 → 0 as n → ∞.. (3.22). Finally, the definition of Pn , Πn and the growth condition (2.32) imply that for a.e. (ω, s) ∈ ΩT ,

(255) X

(256) X

(257) 2 2

(258) q |P σ(s, v(s))Π ψ | − |σ(s, v(s))| qj |σ(s, v(s))ψj |2L2

(259) j n n j L2 L

(260) ≤ 2 j≥1. j≥n+1. + 2|(Pn − Id)σ(s, v(s))|2L → 0 as n → ∞. Furthermore, the growth conditon (2.32) implies that for every n:

(261) X

(262)

(263)

(264) qj |Pn σ(s, v(s))Πn ψj |2L2 − |σ(s, v(s))|2L

(265) ≤ 2|σ(s, v(s))|2L

(266) j≥1.   ≤ 2 K0 + K1 |v(s)|2L2 + K2 |∇h v(s)|2L2 ∈ L1 (ΩT ). Hence the dominated convergence theorem implies that Z T 3 Zn → −E e−r(s) |σ(s, v(s))|2L ds. 0. (3.23).

(267) STOCHASTIC ANISOTROPIC 3D NAVIER-STOKES. 21. Using the inequalities (3.19)–(3.23) we obtain: Z T h E e−r(s) − r0 (s)|u(s) − v(s)|2L2 + 2hF˜ (s) − F (v(s)) , u(s) − v(s)i 0 i ˜ − σ(s, v(s))|2L ds ≤ 0. + |S(s). (3.24). ˜ Let v = u ∈ X ; then we deduce that for almost every (ω, s) ∈ ΩT we have S(s) = σ(s, u(s)). ˜ Using another choice of v, we next trove that F (s) = F (u(s)) a.e. on ΩT . Let λ ∈ R and v˜ ∈ X and set vλ = u + λ˜ v ∈ X . Then if rλ is defined in terms on vλ using (3.16), the inequality (3.24) yields Z T Z T 2 −rλ (s) 0 2 e−rλ (s) hF˜ (s) − F (u(s)) , v˜(s)ids λ E e rλ (s)|˜ v (s)|L2 ds + 2λE 0. 0. Z +2λE. T. e−rλ (s) hF (u(s)) − F (vλ (s)) , v˜(s)ids ≤ 0.. (3.25). 0. The upper estimate (2.26) and H¨ older’s inequality imply that for η ∈ (0, ν) and λ ∈ (0, 1],

(268)

(269)

(270)

(271)

(272) hF (vλ (s)) − F (u(s)) , v˜(s)i

(273) = 1

(274) hF (vλ (s)) − F (u(s)) , vλ (s) − u(s)i

(275) ≤ |λ|φ(t), |λ| where by H¨ older’s inequality we have  φ(t) =ηk˜ v (t)k21,1 + 2Cη ku(s)k21,1 + k˜ v k21,1 |˜ v (s)|2L2  + Caκ ku(s)k2α + k˜ v (s)k2α k˜ v (s)k2L2(α+1) . L2(α+1) L2(α+1) Using once more H¨ older’s inequality, we deduce that Z T

(276) 2  1 i

(277) 2  1 

(278) Z T hZ T h

(279) Z T

(280)

(281) 2

(282)

(283) 2 2 2 +

(284) k˜ v (s)k21,1 ds

(285) E φ(t)dt ≤ CE k˜ v (s)k1,1 ds +

(286) ku(s)k1,1 ds

(287) 0. ×. . 0. sup |˜ v (s)|4L2. 1. 2. s∈[0,T ]. 0. 0. h. i 2α + C kuk2α + k˜ v (s)k k˜ v k2L2(α+1) (ΩT ×R3 ) < ∞. 2(α+1) 2(α+1) 3 3 L (ΩT ×R ) L (ΩT ×R ). Since rλ (s) ≥ 0, the dominated convergence theorem implies that Z T E e−rλ (s) hF (u(s)) − F (vλ (s)), v˜(s)ids → 0 as λ → 0. 0.  ˜ 1,1 )) ∩ L2(α+1) (ΩT × R3 ) ∗ , using Furthermore, since F˜ (s) − F (u(s)) ∈ L4 (Ω; L2 (0, T ; H once more the dominated convergence theorem we deduce that as λ → 0: Z T Z T −rλ (s) ˜ E e hF (s) − F (u(s)) , v˜(s)ids → E e−r0 (s) hF˜ (s) − F (u(s)) , v˜(s)ids. 0. 0. Dividing (3.25) by λ and letting λ → and λ → 0− , we deduce that for every v˜ ∈ X , Z T E e−r0 (s) hF˜ (s) − F (u(s)) , v˜(s)ids = 0. 0+. 0. This implies that F˜ (s) = F (u(s)) a.e. on ΩT . Step 4: Continuity of the solution We next prove that u ∈ C([0, T ]; H) a.s. The proof, based on some regularization of the solution, is similar to that in [13]; however, the functional setting is different which requires some changes. Set A = Pdiv ∆; then e−δA ˜ 1,1 )∗ ⊂ H −2 to H for any δ > 0. Furthermore, the Gagliardo Nirenberg inequality maps (H implies that H 2 (R3 ) ⊂ L2(α+1) (R3 ), so that the semi-group e−δA also maps L. 2(α+1) 2α+1. (R3 ) =.

Références

Documents relatifs

In Sections 4 and 5, we study the mixed finite volume approximation respectively for Stokes and Navier-Stokes equations: we show that this method leads to systems (linear in the case

The main results of this paper can be summarised as follows: first, suitable con- trollability properties of a non-linear PDE imply the uniqueness and exponential mixing for

The very first work concerning NBC was done by Solonnikov and ˇ Sˇ cadilov [10] for α = 0 where the authors considered stationary Stokes system with Dirichlet condition on some part

For d = 2, we explicit a core, identify the abstract generator with the differential operator L on this core and prove existence and uniqueness for the corresponding

Odasso, Markov solutions for the 3D stochastic Navier Stokes equations with state dependent noise, Journal of Evolution Equations, Birkhuser Basel, vol.. Sinai, Gibbsian dynamics

In this paper, we deal with the existence of insensitizing controls for the Navier–Stokes equations in a bounded domain with Dirichlet boundary conditions. We prove that there

Nirenberg, Partial regularity of suitable weak solutions of the Navier–Stokes equations, Comm..

— We proved in Section 2 of [7] a local existence and uniqueness theorem (Theorem 2.2) for solutions of the Navier- Stokes equations with initial data in a space 7-^1 ^2,&lt;5s ^