Abstract

We investigate the stochastic 3D Navier-Stokes-𝛼 model which arises in the modelling of turbulent flows of fluids. Our model contains nonlinear forcing terms which do not satisfy the Lipschitz conditions. The adequate notion of solutions is that of probabilistic weak solution. We establish the existence of a such of solution. We also discuss the uniqueness.

1. Introduction

In this paper, we are interested in the study of probabilistic weak solutions of the 3D Navier-Stokes-𝛼(NS𝛼) model (also known as the Lagrangien averaged Navier-Stokes-alpha model or the viscous Camassa-Holm equations) with homogeneous Dirichlet boundary conditions in a bounded domains in the case in which random perturbations appear. To be more precise, let 𝐷 be a connected and bounded open subset of 𝑅3 with 𝐶2 boundary 𝜕𝐷 and a final time 𝑇>0. We denote by 𝐴 the Stokes operator and consider the system

𝜕𝑡(𝑢𝛼Δ𝑢)+𝜈(𝐴𝑢𝛼Δ(𝐴𝑢))+(𝑢)(𝑢𝛼Δ𝑢)𝛼𝑢Δ𝑢+𝑝=𝐹(𝑡,𝑢)+𝐺(𝑡,𝑢)𝑑𝑊,𝑑𝑡in𝐷×(0,𝑇),𝑢=0,in𝐷×(0,𝑇),𝑢=0,𝐴𝑢=0,on𝜕D×(0,T),𝑢(0)=𝑢0,in𝐷,(1.1) where 𝑢=(𝑢1,𝑢2,𝑢3) and 𝑝 are unknown random fields on 𝐷×(0,𝑇), representing, respectively, the large-scale velocity and the pressure, in each point of 𝐷×(0,𝑇). The constant 𝜈>0 and 𝛼>0 are given, and represent, respectively, the kinematic viscosity of the fluid, and the square of the spatial scale at which fluid motion is filtered. The terms 𝐹(𝑡,𝑢) and 𝐺(𝑡,𝑢)(𝑑𝑊/𝑑𝑡) are external forces depending on 𝑢, where 𝑊 is an 𝑅𝑚-valued standard Wiener process. Finally, 𝑢0 is a given nonrandom velocity field.

The deterministic version of (1.1), that is, when 𝐺=0 has been the object of intense investigations over the last years [15] and the initial motivation was to find a closure model for the 3D turbulence-averaged Reynolds model. A key interest in the model is the fact that it serves as a good approximation to the 3D Navier-Stokes equations. One of the main reasons justifying its use is the high computational cost that the Navier-Stokes model requires. Many important results have been obtained in the deterministic case. More precisely, the global well posedness of weak solutions for the NS-𝛼 model on bounded domains has been established in [6, 7] amongst others, and the asymptotic behavior can be found in [6]. Similar results have been proved by Foias et al. [8] in the case of periodic boundary conditions.

However, in order to consider a more realistic model our problem, it is sensible to introduce some kind of noise in the equations. This may reflect, some environmental effects on the phenomena, some external random forces, and so forth. To the best of our knowledge, the existence and uniqueness of solutions of the stochastic version (1.1) which we consider in this paper has only been analyzed in [9] (see also [10]) in the case of Lipschitz assumptions on 𝐹 and 𝐺. The case of non-Lipschitz assumptions on the coefficients 𝐹 and 𝐺 is the main concern of the present paper. This question has been opened till now. The general motivation for studying weak rather than strong solutions of stochastic equations is that existence of weak solutions can be carried through under weaker regularity on the coefficients. This was pointed out, for instance, in [11].

In this paper, we will establish the existence of probabilistic weak solutions for the problem (1.1) under appropriate conditions on the data. Under the strong assumptions on 𝐹 and 𝐺, we prove the uniqueness of weak solutions. The method used for the proof of our existence results is different from the method in [9]. To prove the existence, we use the Galerkin approximation method employing special bases, combined with some famous theorems of probabilistic nature due to Prokhorov [12] and Skorokhod [13].

The paper is organized as follows. In Section 2, we establish some properties of nonlinear term appearing in our equations. The rigorous statement of our problem as well as the main results are included in Section 3 and we show how our problem can be reformulated as an abstract stochastic model. Section 4 is devoted to the proof of our main results.

2. Properties of the Nonlinear Terms in (1.1)

Following [9], we establish some properties of the nonlinear term (𝑢)(𝑢𝛼Δ𝑢)𝛼𝑢Δ𝑢 appearing in (1.1).

We denote by (,) and ||, respectively, the scalar product and associated norm in (𝐿2(𝐷))3, and by (𝑢,𝑣), the scalar product in ((𝐿2(𝐷))3)3 of the gradients of 𝑢 and 𝑣. We consider the scalar product in (𝐻10(𝐷))3 defined by

𝐻((𝑢,𝑣))=(𝑢,𝑣)+𝛼(𝑢,𝑣),𝑢,𝑣10(𝐷)3,(2.1) where its associated norm is, in fact, equivalent to the usual gradient norm. We denote by 𝐻 the closure in (𝐿2(𝐷))3 of the set

𝒱=𝑣(𝒟(𝐷))3𝑣=0in𝐷,(2.2) and by 𝑉 the closure of 𝒱 in (𝐻10(𝐷))3. Then 𝐻 is a Hilbert space equipped with the inner product of (𝐿2(𝐷))3, and 𝑉 is a Hilbert subspace of (𝐻10(𝐷))3.

Denote by 𝐴 the Stokes operator, with domain 𝐷(𝐴)=(𝐻2(𝐷))3𝑉, defined by

𝐴𝑤=𝒫(Δ𝑤),𝑤𝐷(𝐴),(2.3) where 𝒫 is the projection operator from (𝐿2(𝐷))3 onto 𝐻. Recall that as 𝜕𝐷 is 𝐶2,|𝐴𝑤| defines in 𝐷(𝐴) a norm which is equivalent to the (𝐻2(𝐷))3 norm, that is, there exists a constant 𝑐1(𝐷), depending only on 𝐷, such that

𝑤(𝐻2(𝐷))3𝑐1||||(𝐷)𝐴𝑤,𝑤𝐷(𝐴),(2.4) and so 𝐷(𝐴) is a Hilbert space with respect to the scalar product

(𝑣,𝑤)𝐷(𝐴)=(𝐴𝑣,𝐴𝑤).(2.5) For 𝑢𝐷(𝐴) and 𝑣(𝐿2(𝐷))3, we define (𝑢)𝑣 as the element of (𝐻1(𝐷))3 given by

(𝑢)𝑣,𝑤1=3𝑖,𝑗=1𝜕𝑖𝑣𝑗,𝑢𝑖𝑤𝑗1𝐻,𝑤10(𝐷)3,(2.6) where by 𝑢,𝑣1, we denote either the duality product between (𝐻1(𝐷))3 and (𝐻10(𝐷))3 or between 𝐻1(𝐷) and 𝐻10(𝐷).

Observe that (2.6) is meaningful, since 𝐻2(𝐷)𝐿(𝐷) and 𝐻10(𝐷)𝐿6(𝐷) with continuous injections. This implies that 𝑢𝑖𝑤𝑗𝐻10(𝐷), and there exists a constant 𝑐2(𝐷)>0, depending only on 𝐷, such that

||(𝑢)𝑣,𝑤1||𝑐2||||𝐿(𝐷)𝐴𝑢|𝑣|𝑤,(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐻10(𝐷)3.(2.7) Observe also that if 𝑣(𝐻1(𝐷))3, then the definition above coincides with the definition of (𝑢)𝑣 as the vector function whose components are 3𝑖=1𝑢𝑖𝜕𝑖𝑣𝑗, for 𝑗=1,2,3. However, as it not known whether the solutions of the stochastic problem (1.1) have the same regularity as the deterministic case (we only can ensure 𝐻2 instead of 𝐻3), the present extension is necessary.

Now, if 𝑢𝐷(𝐴), then 𝑢(𝐻1(𝐷))3×3(𝐿6(𝐷))3×3, and consequently, for 𝑣(𝐿2(𝐷))3, we have that 𝑢𝑣(𝐿3/2(𝐷))3(𝐻1(𝐷))3, with

𝑢𝑣,𝑤1=3𝑖,𝑗=1𝐷𝜕𝑗𝑢𝑖𝑣𝑖𝑤𝑗𝐻𝑑𝑥,𝑤10(𝐷)3.(2.8) It follows that there exists a constant 𝑐3(𝐷), depending only on 𝐷, such that

||𝑢𝑣,𝑤1||𝑐3||||𝐿(𝐷)𝐴𝑢|𝑣|𝑤,(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐻10(𝐷)3.(2.9) We have the following results.

Proposition 2.1. For all (𝑢,𝑤)𝐷(𝐴)×𝐷(𝐴) and for all 𝑣(𝐿2(𝐷))3, it follows that (𝑢)𝑣,𝑤1=𝑤𝑣,𝑢1.(2.10)

Proof. If (𝑢,𝑤)𝐷(𝐴)×𝐷(𝐴), then for each 𝑖,𝑗=1,2,3, we have 𝑢𝑖𝑤𝑗𝐻10(𝐷) and consequently 𝜕𝑖𝑣𝑗,𝑢𝑖𝑤𝑗1=𝐷𝑣𝑗𝜕𝑖𝑢𝑖𝑤𝑗𝑑𝑥=𝐷𝑣𝑗𝑤𝑗𝜕𝑖𝑢𝑖𝑑𝑥𝐷𝑣𝑗𝑢𝑖𝜕𝑖𝑤𝑗𝑑𝑥,(2.11) using 𝑢=0, we have (2.10).

Consider now the bilinear form defined by

𝑏(𝑢,𝑣,𝑤)=(𝑢)𝑣,𝑤1+𝑢𝑣,𝑤1,𝐿(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐻10(𝐷)3.(2.12)

Proposition 2.2. The bilinear form 𝑏 satisfies 𝑏(𝑢,𝑣,𝑤)=𝑏𝐿(𝑤,𝑣,𝑢),(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐷(𝐴),(2.13) and consequently, 𝑏𝐿(𝑢,𝑣,𝑢)=0,(𝑢,𝑣)𝐷(𝐴)×2(𝐷)3.(2.14) Moreover, there exists a constant 𝑐(𝐷)>0, depending only on 𝐷, such that ||𝑏||||||𝐿(𝑢,𝑣,𝑤)𝑐(𝐷)𝐴𝑢|𝑣|𝑤,(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐻10(𝐷)3,||𝑏||||||𝐿(𝑢,𝑣,𝑤)𝑐(𝐷)𝑢|𝑣|𝐴𝑤,(𝑢,𝑣,𝑤)𝐷(𝐴)×2(𝐷)3×𝐷(𝐴).(2.15) Thus, in particular, 𝑏 is continuous on 𝐷(𝐴)×(𝐿2(𝐷))3×(𝐻10(𝐷))3.

Proof. The proof is straightforward consequences of (2.7), (2.9), and (2.10).

3. Statement of the Problem and the Main Results

We now introduce some probabilistic evolutions spaces.

Let (Ω,𝐹,{𝐹𝑡}0𝑡𝑇,𝑃) be a filtered probability space and let 𝑋 be a Banach space. For 𝑟,𝑞1, we denote by

𝐿𝑝(Ω,𝐹,𝑃;𝐿𝑟(0,𝑇;𝑋))(3.1) the space of functions 𝑢=𝑢(𝑥,𝑡,𝜔) with values in 𝑋 defined on [0,𝑇]×Ω and such that

(1)𝑢 is measurable with respect to (𝑡,𝜔) and for almost all 𝑡, 𝑢 is 𝐹𝑡 measurable, (2)𝑢𝐿𝑝(Ω,𝐹,𝑃;𝐿𝑟(0,𝑇;𝑋))=𝐸𝑇0𝑢𝑟𝑋𝑑𝑡𝑝/𝑟1/𝑟<,(3.2) where 𝐸 denotes the mathematical expectation with respect to the probability measure 𝑃.

The space 𝐿𝑝(Ω,𝐹,𝑃;𝐿𝑟(0,𝑇;𝑋)) so defined is a Banach space.

When 𝑟=, the norm in 𝐿𝑝(Ω,𝐹,𝑃;𝐿(0,𝑇;𝑋)) is given by

𝑢𝐿𝑝(Ω,𝐹,𝑃;𝐿(0,𝑇;𝑋))=𝐸supess0𝑡𝑇𝑢𝑝𝑋1/𝑝.(3.3)

We make precise our assumptions on (1.1).

We start with the nonlinear function 𝐹 and 𝐺. We assume that

𝐻𝐹(0,𝑇)×𝑉1(𝐷)3,ameasurable.e.𝑡,𝑢𝐹(𝑡,𝑢)continuousfrom𝑉to𝐻1(𝐷)3𝐹(𝑡,𝑢)𝐻1(𝐷)3𝑐(1+𝑢),𝐺(0,𝑇)×𝑉(𝐿2(𝐷))3𝑚,ameasurable.e.𝑡,𝑢𝐺(𝑡,𝑢)continuousfrom𝑉to𝐿2(𝐷)3𝑚||||𝐺(𝑡,𝑢)((𝐿2(𝐷))3)𝑚𝑐(1+𝑢).(3.4) We will define the concept of weak solution of the problem (1.1), namely, the following.

Definition 3.1. A weak solution of (1.1) means a system (Ω,,{𝑡}0𝑡𝑇,𝒫,𝑊,𝑢) such that (1)(Ω,,𝒫) is a probability space, ({𝑡},0𝑡𝑇) is a filtration, (2)𝑊 is an 𝑚-dimensional {𝑡}  standard Wiener process, (3)𝑢(𝑡) is 𝑡 adapted for all 𝑡[0,𝑇]𝑢𝐿𝑝Ω,,𝒫;𝐿2(0,𝑇,𝐷(𝐴))𝐿𝑝(Ω,,𝒫;𝐿(0,𝑇,𝑉)),1𝑝<,(3.5)(4)for almost all 𝑡(0,𝑇), the following equation holds 𝒫-a.s.((𝑢(𝑡),Φ))+𝜈𝑡0(𝑢(𝑠)+𝛼𝐴𝑢(𝑠),𝐴Φ)𝑑𝑠+𝑡0𝑏(=𝑢𝑢(𝑠),𝑢(𝑠)𝛼Δ𝑢(𝑠),Φ)𝑑𝑠0+,Φ𝑡0𝐹(𝑠,𝑢(𝑠)),Φ1𝑑𝑠+𝑡0𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠),Φ(3.6) for all Φ𝐷(𝐴).

Our two major results are as follows.

Theorem 3.2 (Existence). Assume (3.4) and 𝑢0𝑉. Then there exists a weak solution (Ω,,{𝑡}0𝑡𝑇,𝒫,𝑊,𝑢) of (1.1) in the sense of Definition 3.1.
Moreover 𝑢𝐿𝑝(Ω,,𝒫;𝐶([0,𝑇];𝑉)), and there exists a unique ̃𝑝𝐿2(Ω,𝑡,𝒫;𝐻1(0,𝑡;𝐻1(𝐷)), for all 𝑡[0,𝑇], such that 𝒫-a.s.
𝜕𝑡(𝑢𝛼Δ𝑢)+𝜈(𝐴𝑢𝛼Δ(𝐴𝑢))+(𝑢)(𝑢𝛼Δ𝑢)𝛼𝑢Δ𝑢+̃𝑝=𝐹(𝑡,𝑢)+𝐺(𝑡,𝑢)𝑑𝑊,𝑑𝑡in𝒟((0,𝑇)×𝐷)3,𝐷̃𝑝𝑑𝑥=0,in𝒟(0,𝑇),(3.7) where 𝐺(𝑡,𝑢)(𝑑𝑊/𝑑𝑡) denotes the time derivative of 𝑡0𝐺(𝑠,𝑢(𝑠))𝑑𝑊𝑠, that is, by definition 𝐺(𝑡,𝑢)𝑑𝑊𝑑𝑡=𝜕𝑡.0𝐺(𝑠,𝑢(𝑠))𝑑𝑊𝑠,in𝒟𝐿0,𝑇;2(𝐷)3,𝒫-a.s.(3.8)

Corollary 3.3 (Uniqueness). Assume that 𝐹 and 𝐺 are Lipschitz with respect to the second variable 𝑢0𝑉. Then there exists a unique weak solution of problem (1.1) in the sense of Definition 3.1.
Moreover, two strong solutions on the same Brownian stochastic basis coincide a.s.

3.1. Formulation of Problem (1.1) as an Abstract Problem

We will rewrite our model as an abstract problem.

We identify 𝑉 with its topological dual 𝑉 and we have the Gelfand triple 𝐷(𝐴)𝑉𝐷(𝐴).

We denote by , the duality product between 𝐷(𝐴) and 𝐷(𝐴). We define

𝐴𝑢,𝑣=𝜈(𝐴𝑢,𝑣)+𝜈𝛼(𝐴𝑢,𝐴𝑣),𝑢,𝑣𝐷(𝐴).(3.9) It is clear that for all 𝑣𝐷(𝐴),

||||2𝐴𝑢,𝑣=2𝜈(𝐴𝑣,𝑣)+2𝜈𝛼(𝐴𝑣,𝐴𝑣)2𝜈𝛼𝐴𝑣2,(3.10) and, if we denote by 𝜆𝑘 and 𝑤𝑘,𝑘1, the eigenvalues, and their corresponding eigenvalues associated to 𝐴, then

𝐴𝑤𝑘,𝑣=𝜈𝜆𝑘𝑤𝑘.,𝑣(3.11) Thus, taking

𝛼=2𝜈𝛼,(3.12) we have

(a)𝐴 is a linear continuous operator 𝐴(𝐷(𝐴),𝐷(𝐴)), such that (a𝐴1)isself-adjoint(a2)thereexists𝛼>0,suchthat2𝐴𝑣,𝑣𝛼𝑣2𝐷(𝐴),𝑣𝐷(𝐴).(3.13) On the other hand, denote 𝐵(𝑢,𝑣),𝑤=𝑏(𝑢,𝑣𝛼Δ𝑣,𝑤),(𝑢,𝑣,𝑤)𝐷(𝐴)×𝐷(𝐴)×𝐷(𝐴),𝐹(𝑡,𝑢),𝑤=𝐹(𝑡,𝑢),𝑤1,(𝑢,𝑤)𝑉×𝑉.(3.14) Thus it is straightforward to check that if we take 𝑐1=(1+𝛼)𝑐1(𝐷)𝑐(𝐷),(3.15) then we obtain that (b)𝐵𝐷(𝐴)×𝐷(𝐴)𝐷(𝐴) is a bilinear mapping such that (b1𝐵)(𝑢,𝑣),𝑢=0,𝑢,𝑣𝐷(𝐴),(3.16)(b2)𝐵(𝑢,𝑣)𝐷(𝐴)𝑐1𝑢𝑣𝐷(𝐴),𝑢,𝑣𝐷(𝐴)×𝐷(𝐴),(3.17)(b||||3)𝐵(𝑢,𝑣),𝑤𝑐1𝑢𝐷(𝐴)𝑣𝐷(𝐴)𝑤,𝑢,𝑣,𝑤𝐷(𝐴).(3.18)(c)𝐹(0,𝑇)×𝑉𝑉, measurable such that (c1)a.e𝐹.𝑡,𝑢(𝑡,𝑢)continuousfrom𝑉to𝑉(c2)𝐹(𝑡,𝑢)𝑐(1+𝑢).(3.19) Now, let 𝐼 denote the identity operator in 𝐻, and define 𝐺(𝑡,𝑢) as 𝐺(𝑡,𝑢)=(𝐼+𝛼𝐴)1𝒫𝐺(𝑡,𝑢),𝑢𝑉.(3.20)𝐼+𝛼𝐴 is bijective from 𝐷(𝐴) onto 𝐻, and (𝐼+𝛼𝐴)1𝑓,𝑤=(𝑓,𝑤),𝑓𝐻,𝑤𝑉.(3.21) Thus, for each𝑓𝐻,(𝐼+𝛼𝐴)1𝑓2||𝑓||=(𝑓,𝑢)|𝑢|,(3.22) where 𝑢=(𝐼+𝛼𝐴)1𝑓, that is, (𝑢,𝑤𝑘)+𝛼(𝐴𝑢,𝑤𝑘)=(𝑓,𝑤𝑘), for all 𝑘1, so (1+𝛼𝜆𝑘)(𝑢,𝑤𝑘)=(𝑓,𝑤𝑘), which implies 𝑢,𝑤𝑘=11+𝛼𝜆𝑘𝑓,𝑤𝑘11+𝛼𝜆1𝑓,𝑤𝑘,|𝑢|2=𝑘=1𝑢,𝑤𝑘211+𝛼𝜆12𝑘=1𝑓,𝑤𝑘2=11+𝛼𝜆12||𝑓||2.(3.23) Therefore, (𝐼+𝛼𝐴)1𝑓211+𝛼𝜆1||𝑓||2,(3.24) and, consequently, taking 𝑐̃𝑐=1+𝛼𝜆1,(3.25) we obtain that (d)𝐺(0,𝑇)×𝑉𝑉𝑚, measurable such that (d1)a.e.𝑡,𝑢𝐺(𝑡,𝑢)continuousfrom𝑉to𝑉𝑚(d2)𝐺(𝑡,𝑢)𝑉𝑚̃𝑐(1+𝑢),(3.26)

where 𝑉𝑚𝑖𝑠𝑡𝑒𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑜𝑓𝑚𝑐𝑜𝑝𝑖𝑒𝑠𝑜𝑓𝑉. Next, for each 𝑗1, and all (𝑡,𝑢,Φ)(0,𝑇)×𝑉×𝐷(𝐴), we have

=,(𝐺(𝑡,𝑢),Φ)=(𝐼+𝛼𝐴)𝐺(𝑡,𝑢),Φ𝐺(𝑡,𝑢),Φ(3.27) and, for all 𝑢𝐿2(Ω,,𝒫;𝐿(0,𝑇;𝑉)), (𝑡,Φ)(0,𝑇)×𝐷(𝐴), it follows that

𝑡0=𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠),Φ𝑑𝑗=1𝑡0𝐺𝑗(𝑠,𝑢(𝑠)),Φ𝑑𝑊𝑗=(𝑠)𝑑𝑗=1𝑡0𝐺𝑗(𝑠,𝑢(𝑠),Φ)𝑑𝑊𝑗=(𝑠)𝑡0.𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠),Φ(3.28) Consequently, in this abstract framework, a weak solution (Ω,,{𝑡}0𝑡𝑇,𝒫,𝑊,𝑢) of (1.1) is equivalently as follows.

Definition 3.4. It holds that(1)(Ω,,𝒫) is a probability space, ({𝑡},0𝑡𝑇) is a filtration, (2)𝑊 is a 𝑚-dimensional {𝑡} standard Wiener process, (3)𝑢(𝑡) is 𝑡 adapted for all 𝑡[0,𝑇]𝑢𝐿𝑝Ω,,𝒫;𝐿2(0,𝑇,𝐷(𝐴))𝐿𝑝(Ω,,𝒫;𝐿(0,𝑇,𝑉)),1𝑝<,(3.29)(4)for almost all 𝑡(0,𝑇), the following equation holds 𝒫-a.s.𝑢(𝑡)+𝑡0𝐴𝑢(𝑠)𝑑𝑠+𝑡0𝐵(𝑢(𝑠),𝑢(𝑠))𝑑𝑠=𝑢0+𝑡0𝐹(𝑠,𝑢(𝑠))𝑑𝑠+𝑡0𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠),(3.30) as an equality in 𝐷(𝐴).

Remark 3.5. However, (3.30) implies that 𝑢𝒞(0,𝑇;𝐷(𝐴)), then 𝑢 is weakly continuous in 𝑉 [14, page 263] and the initial condition is meaningful.

4. Proofs of the Main Results

4.1. Proof of Theorem 3.2

We make use of the Galerkin approximation combined with the method of compactness.

We will split the proof into six steps.

4.1.1. Step  1. Construction of an Approximating Sequence

As the injection 𝐷(𝐴)𝑉 is compact, consider an orthonormal basis {𝑒𝑗}𝑗=1,2,in𝐷(𝐴) which is orthogonal in 𝑉 such that 𝑒𝑗 are eigenfunctions of the spectral problem

𝑒𝑗,𝑣𝐷(𝐴)=𝜆𝑗𝑒𝑗,𝑣,𝑣𝐷(𝐴),(4.1) where (.,.)𝐷(𝐴) denotes the scalar product in 𝐷(𝐴). For each 𝑁, let 𝑉𝑁 be the span of {𝑒1,,𝑒𝑁}.

Consider the probabilistic system

Ω,𝐹,𝐹𝑡0𝑡𝑇,𝑃,𝑊.(4.2) We denote by 𝐸 the mathematical expectation with respect to (Ω,𝐹,𝑃).

We look for a sequence of functions 𝑢𝑁(𝑡) in 𝑉𝑁, that is,

𝑢𝑁(𝑡)=𝑁𝑗=1𝑐𝑁𝑗𝑡,𝜔𝑒𝑗(𝑥),(4.3) solutions of the following stochastic ordinary differential equations in 𝑉𝑁

𝑑𝑢𝑁,𝑒𝑗+𝐴𝑢𝑁(𝑡),𝑒𝑗𝐵𝑢+𝑁(𝑡),𝑢𝑁(𝑡),𝑒𝑗=𝐹𝑑𝑡𝑡,𝑢𝑁(𝑡),𝑒𝑗𝐺𝑑𝑡+𝑡,𝑢𝑁(𝑡),𝑒𝑗𝑑𝑢𝑊,𝑗=1,2,,𝑁𝑁(0)=𝑢𝑁0,(4.4) where 𝑢𝑁0𝑉𝑁 and is chosen with the requirements that

𝑢𝑁0𝑢0in𝑉as𝑁.(4.5) There exists a a maximal solution to (4.4), that is, a stopping time 𝑇𝑁𝑇 such that (4.4) holds for 𝑡<𝑇𝑁 [11]. Solvability over (0,𝑇) will follow from a priori estimates for 𝑢𝑁 that we derive in the following section.

We have the following Fourier expansion:

𝑢𝑁(𝑡)=𝑁𝑗=1𝑢𝑁(𝑡),𝑒𝑗𝐷(𝐴)𝑒𝑗=𝑁𝑗=1𝜆𝑗𝑢𝑁(𝑡),𝑒𝑗𝑒𝑗,𝑢𝑁(𝑡)2=𝑁𝑗=1𝜆𝑗𝑢𝑁(𝑡),𝑒𝑗2.(4.6)

4.1.2. Step  2. A Priori Estimates

Throughout 𝐶 and 𝐶𝑖(𝑖=1,) denotes a positive constant independent of 𝑁.

We have the following Lemma.

Lemma 4.1. It holds that 𝑢𝑁 satisfies the following a priori estimates: 𝐸sup0𝑡𝑇𝑢𝑁(𝑠)2+2𝛼𝐸𝑇0𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠𝐶1,(4.7) where 𝐶1 is a constant independent of 𝑁.

Proof. By Ito’s formula, we obtain from (3.16) and (4.4) that 𝑑𝑢𝑁(𝑡)2+2𝐴𝑢𝑁(𝑡),𝑢𝑁2𝐹(𝑡)𝑑𝑡=𝑡,𝑢𝑁(𝑡),𝑢𝑁+(𝑡)𝑁𝑗=1𝜆𝑗𝐺𝑡,𝑢𝑁(𝑡),𝑒𝑗2𝐺𝑑𝑡+2𝑡,𝑢𝑁(𝑡),𝑢𝑁𝑑(𝑡)𝑊.(4.8) Integrating (4.8) with respect to 𝑡, and using (3.13) and (3.19), we have 𝑢𝑁(𝑡)2+𝛼𝑡0𝑢𝑁(𝑠)2𝐷(𝐴)𝑢𝑑𝑠𝑁02+𝐶+𝐶𝑡0𝑢𝑁(𝑠)2𝑑𝑠+2𝑡0𝐺𝑠,𝑢𝑁(𝑠),𝑢𝑁𝑑(𝑠)𝑊(𝑠).(4.9) Let us estimate the stochastic integral in this inequality. By Burkholder-Davis Gundy’s inequality [15], we have 𝐸sup0𝑠𝑡||||𝑠0𝐺𝑠,𝑢𝑁(𝑠),𝑢𝑁𝑑(𝑠)𝑊||||(𝑠)𝐶𝐸𝑡0𝐺𝑠,𝑢𝑁(𝑠),𝑢𝑁(𝑠)2𝑑𝑠1/2𝜖𝐸sup0𝑠𝑡𝑢𝑁(𝑠)2+𝐶𝜖𝑡0𝑢1+𝑁(𝑠)2𝑑𝑠,(4.10) here we have used 𝐻̈𝑜𝑙𝑑𝑒𝑟𝑠 and Young’s inequalities; 𝜖 is an arbitrary positive number.
Using (4.10) and (4.9) together with appropriate choice of 𝜖, we obtain
𝐸sup0𝑠𝑡𝑢𝑁(𝑠)2+2𝛼𝐸𝑡0𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠𝐶+𝐶𝐸𝑡0𝑢𝑁(𝑠)2𝑑𝑠.(4.11) By Gronwall’s lemma, we obtain the sought estimate (4.7).

The following result is related to the higher integrability of 𝑢𝑁.

Lemma 4.2. It holds that 𝐸sup0𝑠𝑇𝑢𝑁(𝑠)𝑝𝐶𝑝1𝑝<.(4.12)

Proof. By Ito’s formula, it follows from (4.4) that for 𝑝4, we have 𝑑𝑢𝑁(𝑡)𝑝/2=𝑝2𝑢𝑁(𝑡)𝑝/22𝐴𝑢𝑁(𝑡),𝑢𝑁𝐵𝑢(𝑡)2𝑁(𝑡),𝑢𝑁(𝑡),𝑢𝑁𝐹(𝑡)+2𝑡,𝑢𝑁(𝑡),𝑢𝑁+1(𝑡)2𝑁𝑖=1𝜆𝑖𝐺𝑡,𝑢𝑁(𝑡),𝑒𝑖2+𝑝44𝐺𝑢𝑁(𝑡),𝑢𝑁(𝑡)2𝑢𝑁(𝑡)2+𝑝𝑑𝑡2𝑢𝑁(𝑡)𝑝/22𝐺𝑡,𝑢𝑁(𝑡),𝑢𝑁𝑑(𝑡)𝑊.(4.13) Using the assumptions (3.16), (3.19), (3.26), it follows that sup0𝑠𝑡𝑢𝑁(𝑠)𝑝/2𝑢𝑁0𝑝/2+𝐶𝑡0𝑢1+𝑁(𝑠)𝑝/2+𝑝𝑑𝑠2sup0𝑠𝑡||||𝑠0𝑢𝑁(𝑠)𝑝/22𝐺𝑠,𝑢𝑁(𝑠),𝑢𝑁𝑑(𝑠)𝑊||||.(4.14) Squaring the both sides of this inequality and passing to mathematical expectation, we deduce from the Martingale inequality, that is, 𝐸sup0𝑠𝑡𝑢𝑁(𝑠)𝑝𝑢𝐶𝑁0𝑝+𝑇+𝐸𝑡0𝑢𝑁(𝑠)𝑝.𝑑𝑠(4.15) From Gronwall’s inequality, we deduce that 𝐸sup0𝑠𝑡𝑢𝑁(𝑠)𝑝𝐶𝑝(4.16) for all 1𝑝<.

We also have the following lemma.

Lemma 4.3. It holds that 𝑢𝑁 satisfies 𝐸𝑇0𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠𝑝𝐶𝑝(4.17) for all 1𝑝<.

Proof. Using (4.9), we have 𝛼𝑝𝑡0𝑢𝑁(𝑠)2𝐷(𝐴)𝑝𝑢𝐶𝑁02𝑝+𝐶+𝐶𝑡0𝑢𝑁(𝑠)2𝑑𝑠𝑝||||+𝐶𝑡0((𝐺(𝑠,𝑢𝑁(𝑠)),𝑢𝑁(𝑠)))𝑑𝑊||||𝑝.(4.18) Taking the mathematical expectation and use the Burkholder-Gundy’s inequality, the proof of the lemma follows from Lemma 4.2.

Lemma 4.4. It holds that 𝐸sup||𝜃||0𝛿1𝑇0𝑢𝑁(𝑡+𝜃)𝑢𝑁(𝑡)2𝐷(𝐴)𝑑𝑡𝐶𝛿.(4.19)

Proof. We note that the functions {𝜆𝑗𝑒𝑗}𝑗=1,2, form an orthonormal basis in the dual 𝐷(𝐴) of 𝐷(𝐴). Let 𝑃𝑁 be the orthogonal projection of 𝐷(𝐴) onto the span {𝜆1𝑒1,,𝜆𝑁𝑒𝑁}, that is, 𝑃𝑁=𝑁𝑗=1𝜆𝑗,𝑒𝑗𝑒𝑗.(4.20)
Thus (4.4) can be rewritten in an integral form as the equality between random variables with values in 𝐷(𝐴) as
𝑢𝑁(𝑡)+𝑡0𝑃𝑁𝐴𝑢𝑁(𝐵𝑢𝑠)+𝑁(𝑠),𝑢𝑁(𝐹𝑠)𝑠,𝑢𝑁(𝑠)𝑑𝑠=𝑢𝑁0+𝑡0𝑃𝑁𝐺𝑠,𝑢𝑁𝑑(𝑠)𝑊.(4.21) For any positive 𝜃, we have 𝑢𝑁(𝑡+𝜃)𝑢𝑁(𝑡)𝐷(𝐴)𝑡𝑡+𝜃𝐴𝑢𝑁𝐵𝑢(𝑠)+𝑁(𝑠),𝑢𝑁𝐹(𝑠)𝑠,𝑢𝑁(𝑠)𝑑𝑠𝐷(𝐴)+𝑡𝑡+𝜃𝐺𝑠,𝑢𝑁𝑑(𝑠)𝑊𝐷(𝐴).(4.22) Taking the square and use the properties of 𝐵𝐴, and 𝐹, we have 𝑢𝑁(𝑡+𝜃)𝑢𝑁(𝑡)2𝐷(𝐴)𝐶𝜃2+𝐶𝑡𝑡+𝜃𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠2+𝐶sup0𝑡𝑇𝑢𝑁(𝑠)2𝑡𝑡+𝜃𝑢𝑁(𝑠)𝐷(𝐴)𝑑𝑠2+𝐶𝜃2sup0𝑠𝑇𝑢𝑁(𝑠)2+𝑡𝑡+𝜃𝐺𝑠,𝑢𝑁𝑑(𝑠)𝑊2.(4.23) For fixed 𝛿, taking the supremun over 𝜃𝛿, integrating with respect to 𝑡, and taking the mathematical expectation, we have 𝐸sup0𝜃𝛿𝑇0𝑢𝑁(𝑡+𝜃)𝑢𝑁(𝑡)2𝐷(𝐴)𝑑𝑡𝐶𝛿2+𝐶𝐸𝑇0𝑡𝑡+𝛿𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠2𝑑𝑡+𝐶𝐸sup0𝑠𝑇𝑢𝑁(𝑠)2𝑇0𝑡𝑡+𝛿𝑢𝑁(𝑠)𝐷(𝐴)𝑑𝑠2𝑑𝑡+𝐶𝛿2𝐸sup0𝑠𝑇𝑢𝑁(𝑠)2+𝐸𝑇0sup0𝜃𝛿𝑡𝑡+𝜃𝐺𝑠,𝑢𝑁𝑑(𝑠)𝑊2𝑑𝑡.(4.24) We estimate the integrals in this inequality.
We have by 𝐻̈𝑜𝑙𝑑𝑒𝑟𝑠 inequality
𝐼1=𝐸sup0𝑠𝑇𝑢𝑁(𝑠)2𝑇0𝑡𝑡+𝛿𝑢𝑁(𝑠)𝐷(𝐴)𝑑𝑠2𝑑𝑡𝛿2𝐸sup0𝑠𝑇𝑢𝑁(𝑠)2𝑇0𝑢𝑁(𝑠)2𝐷(𝐴)𝑑𝑠.(4.25) Using the 𝐻̈𝑜𝑙𝑑𝑒𝑟𝑠 inequality and the estimates of Lemmas 4.2 and 4.3, we have 𝐼1𝐶𝛿2.(4.26) By Martingale’s inequality, we have 𝐼2=𝐸𝑇0sup0𝜃𝛿𝑡𝑡+𝜃𝐺𝑠,𝑢𝑁𝑑(𝑠)𝑊2𝑑𝑡𝐸𝑇0𝑡𝑡+𝛿𝐺(𝑠,𝑢𝑁(𝑠))2𝑑𝑠𝑑𝑡.(4.27) Using the assumptions on 𝐺 and the estimate of Lemma 4.2, we have 𝐼2𝐶𝛿.(4.28) Collecting the results and making a similar reasoning with 𝜃<0, we obtain from (4.24) that 𝐸sup||𝜃||0𝛿𝑇0𝑢𝑁(𝑡+𝜃)𝑢𝑁(𝑡)2𝐷(𝐴)𝐶𝛿(4.29)

The following lemma is from [16], and it is a compactness results which represents a variation of the compactness theorems in [17, Chapter I, Section 5]. It will be useful for us to prove the tightness property of Galerkin solution.

Proposition 4.5. For any sequences of positives reals number 𝜈𝑚,𝜇𝑚 which tend to 0 as 𝑚, the injection of 𝑌𝜇𝑛,𝜈𝑛=𝑦𝐿2(0,𝑇;𝐷(𝐴))𝐿(0,𝑇;𝑉)sup𝑚1𝜈𝑚sup||𝜃||𝜇𝑚𝑇0𝑦(𝑡+𝜃)𝑦(𝑡)2𝐷(𝐴)1/2<(4.30) in 𝐿2(0,𝑇;𝑉) is compact.

Furthermore 𝑌𝜇𝑛,𝜈𝑛 is a Banach space with the norm

𝑦𝑌𝜇𝑛𝑛,𝜈=sup0𝑡𝑇𝑦(𝑡)+𝑇0𝑦(𝑡)2𝐷(𝐴)𝑑𝑡1/2+sup𝑛1𝜈𝑛sup||𝜃||𝜇𝑛𝑇0𝑦(𝑡+𝜃)𝑦(𝑡)2𝐷(𝐴)𝑑𝑡1/2.(4.31) Alongside with 𝑌𝜇𝑛,𝜈𝑛, we also consider the space 𝑋𝑝,𝜇𝑛,𝜈𝑛(1𝑝<) of random variables 𝑦 such that

𝐸sup0𝑡𝑇𝑦(𝑡)𝑝<;𝐸𝑇0𝑦(𝑡)2𝐷(𝐴)𝑑𝑡𝑝/2<;𝐸sup𝑛1𝜈𝑛sup||𝜃||𝜇𝑛𝑇0𝑦(𝑡+𝜃)𝑦(𝑡)2𝐷(𝐴)𝑑𝑡<.(4.32) Endowed with the norm

𝑦𝑋𝑛𝑛𝑝,𝜈,𝜇=𝐸sup0𝑡𝑇𝑦(𝑡)𝑝1/𝑝+𝐸𝑇0𝑦(𝑡)2𝐷(𝐴)𝑑𝑡𝑝/2𝑝/2+𝐸sup𝑛1𝜈𝑛sup||𝜃||𝜇𝑛𝑇0𝑦(𝑡+𝜃)𝑦(𝑡)2𝐷(𝐴)𝑑𝑡1/2,(4.33)𝑋𝑝,𝜇𝑛,𝜈𝑛 is a Banach space. The priori estimates of the preceding lemmas enable us to claim that for any 1𝑝< and for 𝜇𝑛,𝜈𝑛 such that the series 𝑛=1(𝜇𝑛/𝜈𝑛) converges, the sequence of Galerkin solutions {𝑢𝑁𝑁𝑁} is bounded in 𝑋𝑝,𝜇𝑛,𝜈𝑛.

4.1.3. Step  3. Tightness Property of Galerkin Solutions

Now, we consider the set

𝑆=𝐶(0,𝑇;𝑅𝑚)×𝐿2(0,𝑇;𝑉),(4.34) and 𝐵(𝑆) the𝜎-algebra of the Borel sets of 𝑆.

For each 𝑁, let 𝜙 be the map

𝜙Ω𝑆𝜔𝑊𝜔,,𝑢𝑁.𝜔,(4.35) For each 𝑁, we introduce a probability measure Π𝑁 on (𝑆,𝐵(𝑆)) by

Π𝑁(𝐴)=𝑃𝜙1(𝐴)(4.36) for all 𝐴𝐵(𝑆). The main result of this subsection is the following.

Theorem 4.6. The family of probability measures {Π𝑁;𝑁𝑁} is tight.

Proof. For 𝜀>0, we should find the compact subsets Σ𝜀𝐶(0,𝑇;𝑅𝑚),𝑌𝜀𝐿2(0,𝑇;𝑉),(4.37) such that 𝑃𝜔𝑊𝜔,Σ𝜀𝜀2,(4.38)𝑃𝜔𝑢𝑁𝜔,𝑌𝜀𝜀2.(4.39) The quest for Σ𝜀 is made by taking account of some fact about the Wiener process such as the formula 𝐸||𝑊𝑡2𝑊𝑡1||2𝑗𝑡=(2𝑗1)!2𝑡1𝑗,𝑗=1,2,.(4.40) For a constant 𝐿𝜀 depending on 𝜀 to be chosen later and 𝑛𝑁, we consider the set Σ𝜀=𝑊()𝐶(0,𝑇;𝑅𝑚)sup𝑡1,𝑡2||𝑡[0,𝑇],2𝑡1||1/𝑛6𝑛||𝑊𝑡2𝑡𝑊1||𝐿𝜀(4.41) Making use of Markov’s inequality: 1𝑃(𝜔𝜉(𝜔)𝛼)𝛼𝑘𝐸||||𝜉(𝜔)𝑘(4.42) for a random variable 𝜉 on (Ω,𝐹,𝑃) and positives variables 𝛼 and 𝑘, we get 𝑃𝜔𝑊𝜔,Σ𝜀𝑃𝑛𝜔sup𝑡1,𝑡2[]||𝑡0,𝑇2𝑡1||<1/𝑛6||𝑊𝑡2𝑊𝑡1||>𝐿𝜀𝑛𝑛𝑛=161𝑖=0𝑛𝐿𝜀4𝐸sup𝑖𝑇/𝑛6𝑡(𝑖+1)𝑇/𝑛6||𝑊(𝑡)𝑊𝑖𝑇𝑛6||4𝑐𝑛=1𝑛𝐿𝜀4𝑇𝑛62𝑛6=𝑐𝐿4𝜀𝑛=11𝑛2,(4.43) we choose 𝐿4𝜀=2𝐶𝜀1𝑛=11𝑛2(4.44) to get (4.38).
Next we choose 𝑌𝜀 as a ball of radius 𝑀𝜀 in 𝑌𝜇𝑛,𝜈𝑛 centered at zero and with 𝜇𝑛,𝜈𝑛, independent of 𝜀, converging to zero, and such that 𝑛(𝜇𝑛/𝜈𝑛) converges.
From Proposition 4.5, 𝑌𝜀 is a compact subset of 𝐿2(0,𝑇;𝑉).
We have further
𝑃𝜔𝑢𝑁(𝜔,)𝑌𝜀𝑃𝑢𝜔𝑁𝑌𝜇𝑛𝑛,𝜈>𝑀𝜀1𝑀𝜀𝐸𝑢𝑁𝑌𝜇𝑛𝑛,𝜈𝑐𝑀𝜀,(4.45) choosing 𝑀𝜀=2𝑐𝜀1, we get (4.39).
From (4.38) and (4.39), we have
𝑃𝜔𝑊(𝜔,)Σ𝜀;𝑢𝑁(𝜔,)𝑌𝜀1𝜀,(4.46) this proves that Π𝑁Σ𝜀×𝑌𝜀1𝜀,𝑁.(4.47)

4.1.4. Step  4. Applications of Prokhorov and Skorokhod Results

From the tightness property of {Π𝑁} and Prokhorov’s theorem [12], we have that there exist a subsequence {Π𝑁𝑗} and a measure Π such that Π𝑁𝑗Π weakly.

By Skorokhod’s theorem [13], there exist a probability space (Ω,,𝑃) and random variables (𝑊𝑁𝑗,𝑢𝑁𝑗),(𝑊,𝑢) on (Ω,,𝑃) with values in 𝑆 such that

thelawof𝑊𝑁𝑗,𝑢𝑁𝑗isΠ𝑁𝑗,(4.48)thelawof(𝑊,𝑢)isΠ,(4.49)𝑊𝑁𝑗,𝑢𝑁𝑗(𝑊,𝑢)in`𝑆,𝑃-a.s.(4.50) Hence, {𝑊𝑁𝑗} is a sequence of an 𝑚-dimensional standard Wiener process.

Let 𝑡=𝜎{𝑊(𝑠),𝑢(𝑠),0𝑠𝑡}.

Arguing as in [16], we prove that 𝑊(𝑡) is an 𝑚-dimensional 𝑡 standard Wiener process and the pair (𝑊𝑁𝑗,𝑢𝑁𝑗) satisfies the equation

𝑢𝑁𝑗(𝑡)+𝜈𝑡0𝑃𝑁𝑗𝐴𝑢𝑁𝑗(𝑠)𝑑𝑠+𝑡0𝑃𝑁𝑗𝐵𝑢𝑁𝑗(𝑠),𝑢𝑁𝑗(=𝑠)𝑑𝑠𝑡0𝑃𝑁𝑗𝐹𝑠,𝑢𝑁𝑗(𝑠)𝑑𝑠+𝑡0𝑃𝑁𝑗𝐺𝑠,𝑢𝑁𝑗(𝑠)𝑑𝑊𝑁𝑗+𝑢𝑁𝑗0.(4.51)

4.1.5. Step  5. Passage to the Limit

From (4.51), it follows that 𝑢𝑁𝑗 satisfies the results of the Lemmas 4.2, 4.3, and 4.4. Therefore, we have for 𝑝1 the a priori estimates

𝐸sup0𝑡𝑇𝑢𝑁𝑗(𝑡)𝑝𝐸𝐶;𝑇0𝑢𝑁𝑗(𝑡)2𝐷(𝐴)𝑑𝑡𝑝𝐶;𝐸sup0𝜃𝛿𝑇0𝑢𝑁𝑗(𝑡+𝜃)𝑢𝑁𝑗2𝐷(𝐴)𝑑𝑡𝐶(𝛼)𝛿(4.52) thus modulo the extraction of a subsequence denoted again by 𝑢𝑁𝑗, we have

𝑢𝑁𝑗𝑢weaklyin𝐿𝑝(Ω,,𝑃;𝐿𝑢(0,𝑇;𝑉));𝑁𝑗𝑢weaklyin𝐿𝑝Ω,,𝑃;𝐿2;(0,𝑇;𝐷(𝐴))𝐸sup0𝑡𝑇𝑢(𝑡)𝑝𝐶;𝐸𝑇0𝑢(𝑡)2𝐷(𝐴)𝑑𝑡𝑝𝐶;𝐸sup0𝜃𝛿𝑇0𝑢(𝑡+𝜃)𝑢(𝑡)2𝐷(𝐴)𝑑𝑡𝐶𝛿.(4.53) By (4.50) and the first estimate in (4.52) and Vitali’s theorem, we have

𝑢𝑁𝑗𝑢stronglyin𝐿2Ω,,𝑃;𝐿2,(0,𝑇,𝑉)(4.54) and thus modulo the extraction of a subsequence and for almost every (𝜔,𝑡) with respect to the measure 𝑑𝑃𝑑𝑡:

𝑢𝑁𝑗𝑢in𝑉.(4.55) This convergence together with the condition on 𝐹, the first estimate in (4.52) and Vitali’s theorem, give

𝐹,𝑢𝑁𝑗𝐹()(,𝑢())stronglyin𝐿2Ω,,𝑃;𝐿2,(0,𝑇,𝑉)𝑡0𝐹𝑠,𝑢𝑁𝑗(𝑠)𝑑𝑠𝑡0𝐹(𝑠,𝑢(𝑠))𝑑𝑠stronglyin𝐿2Ω,,𝑃;𝐿2.(0,𝑇,𝑉)(4.56) As

𝑢𝑁𝑗𝑢weaklyin𝐿2Ω,,𝑃;𝐿2,(0,𝑇;𝐷(𝐴))(4.57) then

𝑡0𝐴𝑢𝑁𝑗(𝑠)𝑑𝑠𝑡0𝐴𝑢(𝑠)𝑑𝑠weaklyin𝐿2Ω,,𝑃;𝐿20,𝑇;𝐷(𝐴).(4.58) We also have

𝑡0𝐵𝑢𝑁𝑗(𝑠),𝑢𝑁𝑗(𝑠)𝑑𝑠𝑡0𝐵(𝑢(𝑠),𝑢(𝑠))𝑑𝑠weaklyin𝐿2Ω,,𝑃;𝐿20,𝑇;𝐷(𝐴).(4.59) In fact, since 𝐿(Ω×(0,𝑇),𝑑𝑃×𝑑𝑡;𝐷(𝐴)) is dense in 𝐿2(Ω,,𝑃;𝐿2(0,𝑇;𝐷(𝐴))), and 𝐵(𝑢𝑁𝑗(𝑠),𝑢𝑁𝑗(𝑠)) is bounded in 𝐿2(Ω,,𝑃;𝐿2(0,𝑇;𝐷(𝐴))) it suffices to prove that forall𝜑𝐿(Ω×(0,𝑇),𝑑𝑃×𝑑𝑡;𝐷(𝐴)),

𝐸𝑇0𝐵𝑢𝑁𝑗(𝑠),𝑢𝑁𝑗(𝑠),𝜑(𝑠)𝐷(𝐴)𝑑𝑠𝐸𝑇0𝐵(𝑢(𝑠),𝑢(𝑠)),𝜑(𝑠)𝐷(𝐴)𝑑𝑠.(4.60) Indeed, we have

𝐸𝑇0𝐵𝑢𝑁𝑗(𝑠),𝑢𝑁𝑗(𝑠)𝐵(𝑢(𝑠),𝑢(𝑠)),𝜑(𝑠)𝐷(𝐴)𝑑𝑠=𝐸𝑇0𝐵(𝑢𝑁𝑗(𝑠)𝑢(𝑠),𝑢𝑁𝑗(𝑠)),𝜑(𝑠)𝐷(𝐴)𝑑𝑠+𝐸𝑇0𝐵(𝑢(𝑠),𝑢𝑁𝑗(𝑠)𝑢(𝑠)),𝜑(𝑠)𝐷(𝐴)𝑑𝑠=𝐼1𝑗+𝐼2𝑗,𝐼1𝑗=𝐸𝑇0𝐵𝑢𝑁𝑗(𝑠)𝑢(𝑠),𝑢𝑁𝑗(𝑠),𝜑(𝑠)𝐷(𝐴)𝑑𝑠(4.61) By the property (3.17) of 𝐵, we have

𝐼1𝑗𝐶𝐸𝑇0𝑢𝑁𝑗𝑢(𝑠)𝑢(𝑠)𝑁𝑗(𝑠)𝐷(𝐴)||||𝐴𝜑(𝑠)𝑑𝑠,(4.62) applying Cauchy-Schwarz inequality

𝐼1𝑗𝐶𝜑𝐸𝑇0𝑢𝑁𝑗(𝑠)𝑢(𝑠)2𝑑𝑠1/2𝐸𝑇0𝑢𝑁𝑗(𝑠)2𝐷(𝐴)𝑑𝑠1/2.(4.63) Since

𝑢𝑁𝑗𝑢stronglyin𝐿2Ω,,𝑃;𝐿2,(0,𝑇;𝑉)(4.64) and 𝑢𝑁𝑗 is bounded in 𝐿2(Ω,,𝑃;𝐿2(0,𝑇;𝐷(𝐴))), we conclude that

𝐼1𝑗0as𝐼𝑗.2𝑗=𝐸𝑇0𝐵𝑢(𝑠),𝑢𝑁𝑗𝑢(𝑠),𝜑(𝑠)𝐷(𝐴)𝑑𝑠.(4.65) Again thanks to the property (3.18) of 𝐵, as

𝑢𝑁𝑗𝑢weaklyin𝐿2Ω,,𝑃;𝐿2,(0,𝑇;𝐷(𝐴))(4.66) we obtain 𝐼2𝑗0 as 𝑗 since any strongly continuous linear operator is weakly continuous. We are now left with the proof of

𝑡0𝐺𝑠,𝑢𝑁𝑗(𝑠)𝑑𝑊𝑁𝑗(𝑠)𝑡0𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠)weakly𝐿2Ω,,𝑃;𝐿0,𝑇;𝐷(𝐴),(4.67) which can be prove with the same argument like in [16].

Collecting all the convergence results, we deduce that

𝑢(𝑡)+𝜈𝑡0𝐴𝑢(𝑠)𝑑𝑠+𝑡0=𝐵(𝑢(𝑠),𝑢(𝑠))𝑑𝑠𝑡0𝐹(𝑠,𝑢(𝑠))𝑑𝑠+𝑡0𝐺(𝑠,𝑢(𝑠))𝑑𝑊(𝑠)+𝑢0,𝑃-a.s.(4.68) as the equality in 𝐷(𝐴).

We have 𝐵(𝑢,𝑢)𝐿2(Ω,,𝑃;𝐿(0,𝑇;𝐷(𝐴))), 𝐴𝑢𝐹(𝑡,𝑢)𝐿2(Ω,,𝑃;𝐿(0,𝑇;𝐷(𝐴))), 𝐺(𝑡,𝑢)𝐿2(Ω,,𝑃;𝐿(0,𝑇;𝑉𝑚)).

Thus, from the classical results in [18] (see also [19]), we deduce from (4.68) that 𝑢 is 𝑃-a.s. continuous with values in 𝑉.

4.1.6. Step  6. Existence of the Pressure

For the existence of the pressure, we use a generalization of the Rham’s theorem processes [20, Theorem 4.1, Remark 4.3]. From (3.6), we have for all 𝑣𝒱,

𝜕𝑡(𝑢𝛼Δ𝑢)𝜈(𝐴𝑢𝛼Δ(𝐴𝑢))(𝑢)(𝑢𝛼Δ𝑢)+𝛼𝑢Δ𝑢+𝐹(,𝑢)+𝐺(,𝑢)𝑑𝑊𝑑𝑡,𝑣𝒟(𝐷)3×(𝒟(𝐷))3=0.(4.69) We denote

=𝜕𝑡(𝑢𝛼Δ𝑢)𝜈(𝐴𝑢𝛼Δ(𝐴𝑢))(𝑢)(𝑢𝛼Δ𝑢)+𝛼𝑢Δ𝑢+𝐹(,𝑢)+𝐺(,𝑢)𝑑𝑊.𝑑𝑡(4.70) We will prove that the regularity on 𝑢, implies that

𝐿2Ω,𝑡,𝑃;𝐻1𝐻0,𝑡;2(𝐷)3.(4.71) By (2.7) and (2.9), we have as 𝑢𝐿4(Ω,,𝑃;𝐿2(0,𝑇;𝐷(𝐴))),

(𝑢(𝑢𝛼Δ𝑢))+𝑢Δ𝑢𝐿2Ω,𝑡,𝑃;𝐿1𝐻0,𝑡;1(𝐷)3,𝐴𝑢𝛼Δ(𝐴𝑢)𝐿4Ω,𝑡,𝑃;𝐿2𝐻0,𝑡;2(𝐷)3.(4.72) We also have

𝑢𝛼Δ𝑢𝐿4Ω,𝑡,𝑃;𝐿2𝐿0,𝑡;2(𝐷)3,𝜕𝑡(𝑢𝛼Δ𝑢)𝐿4Ω,𝑡,𝑃;𝐻1𝐿0,𝑡;2(𝐷)3[].,𝑡0,𝑇(4.73) Again, as 𝑢𝐿4(Ω,,𝑃;𝐶([0,𝑇];𝑉)), then its follows that

𝐹(𝑡,𝑢)𝐿4Ω,𝑡,𝑃;𝐿2𝐻0,𝑡;1(𝐷)3,𝐺(𝑡,𝑢)𝑑𝑊𝑑𝑡𝐿4Ω,𝑡,𝑃;𝑊1,𝐿0,𝑡;2(𝐷)3,(4.74) for all 𝑡[0,𝑇].

Then 𝐿2(Ω,𝑡,𝑃;𝐻1(0,𝑡;(𝐻2(𝐷))3), and

,𝑣𝒟(𝐷)3×(𝒟(𝐷))3=0,𝑣𝒱.(4.75) Therefore, by a generalization of the Rham theorem processes [20], there exists a unique ̃𝑝𝐿2(Ω,𝑡,𝑃;𝐻1(0,𝑡;(𝐻1(𝐷))3)such that𝑃-a.s.

̃𝑝=,𝐷̃𝑝𝑑𝑥=0,thatis,(3.7).(4.76) Theorem 3.2 is proved.

4.2. Proof of Corollary 3.3

Proof. We will prove the pathwise uniqueness which implies uniqueness of weak solutions. Let 𝐿𝐹 and 𝐿𝐺 be two real such that 𝐹(𝑡,𝑢)𝐹(𝑡,𝑣)(𝐻1(𝐷))3𝐿𝐹𝑢𝑣,𝐺(𝑡,𝑢)𝐺(𝑡,𝑣)((𝐿2(𝐷))3)𝑚𝐿𝐺𝑢𝑣.(4.77) Then 𝐹 and 𝐺 are defined, respectively, by (3.14) and (3.20) satisfying 𝐹(𝑡,𝑢)𝐹(𝑡,𝑣)𝑉𝐹𝐿𝑢𝑣,𝐺(𝑡,𝑢)𝐺(𝑡,𝑢)𝑉𝑚𝐺𝐿𝑢𝑣.(4.78) Let 𝑢1 and 𝑢2 two weak solutions of problem (1.1) defined on the same probability space together with the same Wiener process and starting from the same initial value 𝑢0.
We denote 𝑢=𝑢1𝑢2. Take 𝜇>0 to be defined later and 𝜌(𝑡)=exp(𝜇𝑡0𝑢2(𝑠)2𝐷(𝐴)𝑑𝑠), 0𝑡𝑇.
Applying Ito’s formula to the real process 𝜌(𝑡)𝑢(𝑡)2, we obtain from (3.13), (3.18), (3.19), and (3.26) that 𝜌(𝑡)𝑢(𝑡)2+𝛼𝑡0𝜌(𝑠)𝑢(𝑠)2𝐷(𝐴)𝑑𝑠𝐿2𝐺𝑡0𝜌(𝑠)𝑢(𝑠)2𝑑𝑠+2̃𝑐𝑡0𝑢𝜌(𝑠)2(𝑠)𝐷(𝐴)𝑢(𝑠)𝐷(𝐴)𝐹𝑢(𝑠)𝑑𝑠+2𝐿𝑡0𝜌(𝑠)𝑢(𝑠)𝐷(𝐴)𝑢(𝑠)𝑑𝑠+2𝑡0𝐺𝜌(𝑠)𝑠,𝑢1(𝐺𝑠)𝑠,𝑢2(,𝑠)𝑢(𝑠)𝑑𝑊(𝑠)𝜇𝑡0𝑡0𝑢𝜌(𝑠)2(𝑠)2𝐷(𝐴)𝑢(𝑠)2𝑑𝑠,(4.79) for all 𝑡[0,𝑇].
By young’s inequality, we have
𝑢2̃𝑐2(𝑠)𝐷(𝐴)𝑢(𝑠)𝐷(𝐴)𝑢(𝑠)𝛼2𝑢(𝑠)2𝐷(𝐴)+2̃𝑐2𝑢𝛼2(𝑠)2𝐷(𝐴)𝑢(𝑠)2𝐹2𝐿𝑢(𝑠)𝐷(𝐴)𝑢(𝑠)𝛼2𝑢(𝑠)2𝐷(𝐴)+2𝐿2𝐹𝛼𝑢(𝑠)2.(4.80) If we take 𝜇=2(̃𝑐2/𝛼), we obtain from (4.79) that 𝜌(𝑡)𝑢(𝑡)2𝐿2𝐺+2𝐿2𝐹𝛼𝑡0𝜌(𝑠)𝑢(𝑠)2𝑑𝑠+2𝑡0𝐺𝜌(𝑠)𝑠,𝑢1𝐺(𝑠)𝑠,𝑢2,(𝑠)𝑢(𝑠)𝑑𝑊.(4.81) As 0<𝜌(𝑡)1, the expectation of the stochastic integral in (4.81) vanishes and 𝐸𝜌(𝑡)𝑢(𝑡)2𝐿2𝐺+2𝐿2𝐹𝐸𝛼𝑡0𝜌(𝑠)𝑢(𝑠)2𝑑𝑠.(4.82) The Gronwall lemma implies that 𝑢(𝑡)=0, 𝑃-a.s. for all 𝑡[0,𝑇]. Also, the corollary is proved.

Remark 4.7. Using the famous Yamada-Watanabe theorem [11], Corollary 3.3 implies the existence of a unique strong solution of (1.1).

Acknowledgment

The research of the authors is supported by the University of Pretoria and the National Research Foundation South Africa.