Processing math: 36%
Research article Special Issues

Well-posedness of a nonlinear stochastic model for a chemical reaction in porous media and applications

  • Received: 11 June 2024 Revised: 01 August 2024 Accepted: 07 August 2024 Published: 26 August 2024
  • MSC : 60H15, 60H30, 60H35, 76V05

  • In this paper, we considered a stochastic model of chemical reactive flows acting through porous media under the influence of nonlinear external random fluctuations, where the interchanges of chemical flow across the skeleton's surface are represented by a nonlinear function. We studied the existence and uniqueness of strong probabilistic solutions for the model under consideration. We also show the positivity for the concentration of the solute in the fluid face as well as the concentration of reactants on the surface of the skeleton under extra reasonable assumptions on the data. Initially, we approximated the solution of the nonlinear stochastic diffusion equation using Galerkin's approximation, and obtained important bound estimates along with probabilistic compactness results. Thereafter, we passed the limit and obtained a weak probabilistic solution. This was followed by the path-wise uniqueness of the solution, which leads to the existence and uniqueness of strong probabilistic solutions as a result of Yamada-Watanabe's theorem. Finally, we discuss some important numerical applications such as Langmuir and Freundlich kinetics using the extended stochastic non-conforming finite element method to illustrate the efficiency of this approach and compare it to the deterministic approach in both cases. Let us mention that well-posedness, positivity, and numerical simulations have not been considered so far for such a nonlinear stochastic model.

    Citation: Mhamed Eddahbi, Mogtaba Mohammed, Hammou El-Otmany. Well-posedness of a nonlinear stochastic model for a chemical reaction in porous media and applications[J]. AIMS Mathematics, 2024, 9(9): 24860-24886. doi: 10.3934/math.20241211

    Related Papers:

    [1] Xiaoming Wang, Muhammad W. Yasin, Nauman Ahmed, Muhammad Rafiq, Muhammad Abbas . Numerical approximations of stochastic Gray-Scott model with two novel schemes. AIMS Mathematics, 2023, 8(3): 5124-5147. doi: 10.3934/math.2023257
    [2] Yongchang Wei, Zongbin Yin . Long-time dynamics of a stochastic multimolecule oscillatory reaction model with Poisson jumps. AIMS Mathematics, 2022, 7(2): 2956-2972. doi: 10.3934/math.2022163
    [3] Khalil Ur Rehman, Nosheen Fatima, Wasfi Shatanawi, Nabeela Kousar . Mathematical solutions for coupled nonlinear equations based on bioconvection in MHD Casson nanofluid flow. AIMS Mathematics, 2025, 10(1): 598-633. doi: 10.3934/math.2025027
    [4] Wedad Albalawi, Muhammad Imran Liaqat, Fahim Ud Din, Kottakkaran Sooppy Nisar, Abdel-Haleem Abdel-Aty . Well-posedness and Ulam-Hyers stability results of solutions to pantograph fractional stochastic differential equations in the sense of conformable derivatives. AIMS Mathematics, 2024, 9(5): 12375-12398. doi: 10.3934/math.2024605
    [5] Muhammad Aslam, Muhammad Farman, Hijaz Ahmad, Tuan Nguyen Gia, Aqeel Ahmad, Sameh Askar . Fractal fractional derivative on chemistry kinetics hires problem. AIMS Mathematics, 2022, 7(1): 1155-1184. doi: 10.3934/math.2022068
    [6] W. Abbas, Ahmed M. Megahed, Osama M. Morsy, M. A. Ibrahim, Ahmed A. M. Said . Dissipative Williamson fluid flow with double diffusive Cattaneo-Christov model due to a slippery stretching sheet embedded in a porous medium. AIMS Mathematics, 2022, 7(12): 20781-20796. doi: 10.3934/math.20221139
    [7] Nadeem Abbas, Wasfi Shatanawi, Fady Hasan, Taqi A. M. Shatnawi . Numerical analysis of Darcy resistant Sutterby nanofluid flow with effect of radiation and chemical reaction over stretching cylinder: induced magnetic field. AIMS Mathematics, 2023, 8(5): 11202-11220. doi: 10.3934/math.2023567
    [8] Dan-Dan Dai, Wei Zhang, Yu-Lan Wang . Numerical simulation of the space fractional (3+1)-dimensional Gray-Scott models with the Riesz fractional derivative. AIMS Mathematics, 2022, 7(6): 10234-10244. doi: 10.3934/math.2022569
    [9] Nafissa T. Trouba, Huiying Xu, Mohamed E. M. Alngar, Reham M. A. Shohib, Haitham A. Mahmoud, Xinzhong Zhu . Soliton solutions and stability analysis of the stochastic nonlinear reaction-diffusion equation with multiplicative white noise in soliton dynamics and optical physics. AIMS Mathematics, 2025, 10(1): 1859-1881. doi: 10.3934/math.2025086
    [10] Xiaodong Wang, Kai Wang, Zhidong Teng . Global dynamics and density function in a class of stochastic SVI epidemic models with Lévy jumps and nonlinear incidence. AIMS Mathematics, 2023, 8(2): 2829-2855. doi: 10.3934/math.2023148
  • In this paper, we considered a stochastic model of chemical reactive flows acting through porous media under the influence of nonlinear external random fluctuations, where the interchanges of chemical flow across the skeleton's surface are represented by a nonlinear function. We studied the existence and uniqueness of strong probabilistic solutions for the model under consideration. We also show the positivity for the concentration of the solute in the fluid face as well as the concentration of reactants on the surface of the skeleton under extra reasonable assumptions on the data. Initially, we approximated the solution of the nonlinear stochastic diffusion equation using Galerkin's approximation, and obtained important bound estimates along with probabilistic compactness results. Thereafter, we passed the limit and obtained a weak probabilistic solution. This was followed by the path-wise uniqueness of the solution, which leads to the existence and uniqueness of strong probabilistic solutions as a result of Yamada-Watanabe's theorem. Finally, we discuss some important numerical applications such as Langmuir and Freundlich kinetics using the extended stochastic non-conforming finite element method to illustrate the efficiency of this approach and compare it to the deterministic approach in both cases. Let us mention that well-posedness, positivity, and numerical simulations have not been considered so far for such a nonlinear stochastic model.



    Solute transfer in porous substances is a topic that chemists, geologists, and environmentalists are all interested in. Convection, diffusion, and potential chemical reactions that the solutes may go through are the phenomena that impact solute transport. Most of the mathematical modeling literature for solute transportation in porous medium encompass deterministic diffusion and conviction equations, see for example [2,10,11,15] and the references therein. The earliest rigorous mathematical investigation for this type of model dates back to [15], when the authors investigated a deterministic linear model for convection, adsorption, diffusion, and chemical reaction in perforated materials. In [11], the authors proposed expansions to nonlinear models, including linear diffusion, and chemical reaction equations with nonlinear chemical flow across the skeleton's surface. Related models dealing with Homogenization of reactive flows in porous media was discussed [2,10]. Nonlinear chemical reactive fluxes have been widely explored in well-posedness of elliptic linear and nonlinear PDEs (among others, we refer to [6,7,13]). However, stochastic differential equations account for the inherent randomness impacting the solute's movement through a porous medium. In [4], Bessaih et al. explored the asymptotic analysis and well-posedness of stochastic systems in diffusion-convection flows. Their system integrates a deterministic diffusion equation with a stochastic linear convection equation. Additionally, in [17], the author examined the asymptotic behavior of a nonlinear stochastic model featuring nonlinear random forces for chemical reactive fluxes in porous media.

    We aim to establish the existence and uniqueness of strong probabilistic solutions, demonstrate positivity, and conduct numerical simulations for the nonlinear stochastic version of the model in [11] (see also [17]). Specifically, we examine a stochastic model of chemical reactive flows subject to nonlinear external random fluctuations within porous media. A porous medium is a material made of holes or pores, and the skeletal part of a porous medium is called a matrix or a frame. We consider a situation where chemical substances are dissolved in a fluid and transported by diffusion through a porous medium. These chemicals adsorb onto the surface of the skeleton and react with substances bound to the surface of the skeleton. Adsorption and reaction mechanisms continuously impact solute concentrations in fluids and skeletal surfaces. In addition, we suppose that exogenous nonlinear random forces influence the solute concentration in the fluid. A porous medium is described mathematically as a perforated domain where the perforations represent the obstacles in the porous medium. The perforated domain consists of both the fluid phase (pores) and the solid phase (skeleton). Diffusion occurs in the fluid phase, while reactions occur on the solid phase's surface. The model includes a stochastic diffusion system in the fluid phase, a stochastic response system on the skeleton's surface, and a boundary condition that connects the two. It is important to note that well-posedness, positivity, and numerical simulations have not yet been addressed for this type of nonlinear stochastic model. However, some results related to the analysis of nonlinear stochastic evolution equations and nonlinear parabolic stochastic differential equations can be found in [1,16,25] and the references therein. This study represents a significant advancement in the modeling of solute transport in porous media by integrating stochastic elements and addressing nonlinear chemical reactions. Its contributions are valuable across multiple disciplines, offering both theoretical and practical benefits. However, the complexity and computational demands of the models, along with the challenges in parameter estimation and potential limitations in generalization, should be considered when applying these findings to real-world scenarios.

    Let DRn be an open and bounded set, Sδ is small hole (smooth enough) in Rn of diameter δ, where δ>0 is small number that allows to go to zero, and B(Sδ) be is the set of all periodically distributed holes Sδ in Rn such that B(Sδ)D=. Define the perforated domain as Dδ=DB(Sδ), where we denote by Γδ the boundary (RnB(Sδ)), therefore, Dδ=DΓδ. An illustration of the domain D and holes are presented in Figure 1 below.

    Figure 1.  Sketch of domain D and its related notations.

    The solute concentration in the fluid is modeled on the fluid phase Dδ by the following nonlinear stochastic diffusion equation with initial and boundary conditions, where the stochastic term captures external random fluctuations:

    {duδ=κΔuδdt+fδ1(t,uδ,uδ)dt+gδ1(t,uδ)dW1, in Ω×Dδ×(0,T),uδ=0 on Ω ×D×(0,T),uδ(0) = uδ0, in Dδ,κuδν=δfδ2=δγ(h(t,uδ)vδ), in Ω×Γδ×(0,T), (1.1)

    where T(0,), uδ represents the solute concentration in the fluid, uδ0 represents the starting solute concentration in the fluid phase, ν is the exterior unit normal to Dδ, κ>0 is the diffusion coefficient, fδ1 is a nonlinear function representing an external source of energy that depends on the solute concentration in the fluid phase and its diffusion, fδ2 is a nonlinear function representing interchanges of chemical flow across the surface of the skeleton Γδ. As in the deterministic case, see [11]. The solute concentration on the surface of the skeleton is represented by vδ, γ>0 is the adsorption factor and gδ1W1 is an external nonlinear random fluctuation affecting the concentration of the solute in the fluid phase with gδ1 representing the intensity of the noise and W1 an m-dimensional Brownian motion defined on a complete (Ω,F,P,(Ft)0tT) probability space.

    On the surface of the skeleton, the concentration of the solute vδ is modeled using the following stochastic reaction equation:

    {dvδ=(fδ2avδ)dt+gδ2dW2, in Ω×Γδ×(0,T),vδ(0)=vδ0, in Γδ, (1.2)

    where a>0 is the reaction factor, vδ0 is the initial concentration of reactants on the surface of the skeleton, and gδ2dW2 is an external linear random fluctuation affecting the concentration of the solute on the surface of the skeleton with gδ2 representing the intensity of the noise and W2 defined on the same probability space as above. Let us note that the boundary condition κuδν=δγ(h(t,uδ)vδ), in Ω×Γδ×(0,T) is meant to couple the nonlinear stochastic diffusion system (1.1) in the fluid phase Dδ and the reaction stochastic system (1.2) in the skeleton.

    The rest of this paper is organized as follows: In Section 2, we introduce several assumptions to define a solution for the coupled stochastic partial differential equations. In Section 3, we establish Theorem 2.1 and derive a mild-weak probabilistic solution for the system (1.1)-(1.2). This is followed by a path-wise uniqueness result in Section 4, which establishes a strong probabilistic solution. In Section 5, we show that, given additional data assumptions, the solute concentration in the fluid phase and the reactant concentration on the skeleton's surface remain positive. We conclude this work with Section 6, where we present numerical results. In this section, we illustrate smooth Langmuir and Freundlich kinetics, employing extended stochastic non-conforming finite element methods for discretizing the coupled SPDE. Additionally, we numerically, analyze the convergence order of this method.

    The general probabilistic functional spaces, which we will frequently employ in this paper, will now be introduced. Let B be a Banach space, and p,q[1,]. The space Lp(0,T;B) is defined to be the set of all measurable functions

    u:t[0,T]u(t)B

    such that ϕ(t)B is p integrable in the set [0,T]. The space Lp,q(Ω;(0,T);B)) is the set of all of functions v:(ω,t)Ω×[0,T]v(ω,t,)B, where v(ω,t,x) is Ft-measurable in ω for all t. This space can be equipped with the following norm:

    ||v||Lp,q(Ω;(0,T);B))=(E||v||pLq(0,T;B))1/p.

    For more general properties of these functional spaces, we refer to [9,18,20,21]. In the following, we provide the necessary assumptions on the data of our problem.

    Assumptions:

    (A1) The initial condition (uδ0,vδ0) are uniformly bounded in L2(Dδ)×L2(Γδ).

    (A2) fδ1(t,uδ,uδ)=Fδ1(t,uδ)uδ such that

    (ⅰ) Fδ1=(Fδ1,1,Fδ1,2,,Fδ1,n), where Fδ1,i(,v)L2,2(Ω;(0,T);L2(Dδ)) for all vL2,2(Ω;(0,T);L2(Dδ)) and Fδ1,i satisfies Lipschitz condition with respect to the second argument in the L2-norm;

    (ⅱ) (fδ1(t,uδ,uδ),ϕ)L2(Dδ)CuδL2(Dδ)ϕL2(Dδ).

    (A3) gδ1=(gδ1,1,gδ1,2,,gδ1,m) such that gδ1,i(,v) satisfies Lipschitz condition with respect to the second argument in the L2-norm and the following linear growth:

    gδ1(t,ϕ)L2(Dδ)C(1+ϕL2(Dδ)),for allϕL2(Dδ).

    (A4) The interchanges of chemical flows across the boundary are given by the nonlinear function h:Ω×(0,T)×RR satisfying

    (ⅰ) h(,0)=0 and h is monotonously non-decreasing function,

    (ⅱ) h is continuously differentiable with respect to the second argument,

    (ⅲ) |hz(,z)|C(1+|z|p1) for all zR, C>0, and

    {p[1,),for n=2,p[1,nn2],for n>2.

    (A5) gδ1L2(0,T;L2(Dδ)).

    Remark 2.1. Note that, from Assumption (A4(i)) and (A4(ii)), one can deduces the following:

    zh(z)0,

    |h(z)|C(1+|z|p), for all zR.

    We now, introduce the notion of a weak probabilistic solution to our model. The notion of a weak probabilistic solution to our model is now introduced.

    Definition 2.1. The system (Ω,F,P,uδ,vδ,W1,W2) is said to be a weak probabilistic solution to the model (1.1)-(1.2), if:

    (a) W1 and W2 are Ft-Wiener processes (m-dimensional), where Ft is a filtration defined on (Ω,F,P).

    (b) The mappings (ω,t)uδ(ω,t) and (ω,t)vδ(ω,t) are progressively measurable.

    (c) (uδ,vδ) satisfy, for σ(dx) a surface measure on the boundary Γδ

    Dδuδ(t)ψdxDδuδ(0)ψdx+κt0Dδuδψdxds+δγt0Γδh(uδ)ψσ(dx)ds=δγt0Γδvδψσ(dx)ds+t0Dδfδ1(uδ,uδ)ψdxds+t0Dδgδ1(uδ)ψdxdW1(s), (2.1)

    for all 0tT and ψVδ where Vδ={ϕH1(Dδ) | ϕ=0 on Dδ}, and

    vδ(t,x)=v0(x)e(a+γ)t+t0e(a+γ)(ts)gδ2(s)dW2(s)+γt0e(a+γ)(ts)h(uδ(s,x))ds. (2.2)

    Definition 2.2. Let (Ω,F,P,uδ1,vδ1,W1,W2) and (Ω,F,P,uδ2,vδ2,W1,W2) be two probabilistic weak solutions of system (1.1)-(1.2) with the same initial condition (uδ0,vδ0), we say that (1.1)-(1.2) admits path-wise uniqueness if

    (uδ1,vδ1)(t)=(uδ2,vδ2)(t),forall0tT,Pa.s.

    Then, we have the following result.

    Theorem 2.1. Suppose that the Assumptions (A1)–(A4) hold true. Then for fixed δ and for all positive real number T, there exists a weak probabilistic solution (Ω,F,P,uδ,vδ,W1,W2) of system (1.1)-(1.2) by means of Definition 2.1.

    Theorem 2.2. Let (A1)–(A4) hold, then the system (1.1)-(1.2) has the path-wise uniqueness property in the sense of Definition 2.2.

    In this section, we establish the existence of a weak probabilistic solution employing the Galerkin approximation method, Prokhorov tightness property, and Skorokhod representation theorem.

    Using the Galerkin approximation method, we demonstrate the existence of solutions. Let (δk)k=1[VδL(Dδ)] be sequence of linearly independent vectors and consider Hn=span{δk, k=1,2,,n} to be dense subset of Vδ and define the projection Pn:L2(Dδ)Hn. We now let

    φδ(t,x)=v0(x)e(a+γ)t+t0e(a+γ)(ts)gδ2(s)dW2(s), (3.1)

    where φδ in L2(Ω;C([0,T];L2(Γδ))), we seek an approximate vector solution to our problem in finite dimensions as

    uδn(t,ω,x) = ni=1cin(t,ω)δi(x)andρδn(t,ω,x) = ni=1din(t,ω)δi(x),

    we demand that uδn and ρδn meet the following system:

    Dδduδnδidx+κDδuδnδidxdt+δγΓδh(uδn)δiσ(dx)dt=δΓδ(ρδn+φδ)δiσ(dx)dt+Dδfδ1(uδn,uδn)δidxdt+Dδgδ1(uδn)δidxdW1(t),uδn(0)=Pnuδ0, (3.2)

    and

    ρδnt+(a+γ)ρδn = γh(uδn),ρδn(0 = 0, (3.3)

    where

    ρδn = vδnφδ. (3.4)

    The corresponding system of stochastic ODEs satisfied by the Fourier coefficients cin and din has a unique solution. This gives the existence of uδn and ρδn that solve the system (3.2)-(3.3).

    In this subsection, we derive crucial a priori estimates that will be of great importance when studying the asymptotic behaviour of the solutions of the considered model. For the proof of the following lemma, we refer to [10].

    For notational simplification we shall use uδn(t) to denote uδn(t,) and ρδn(t) to denote ρδn(t,).

    Lemma 3.1. For Dδ and Sδ as constructed above, we have

    w2L2(Γδ)α(1δw2L2(Dδ)+δw2L2(Dδ)),

    where α>0, for all wW1,2δ, where W1,2δ={ϕW1,2(Dδ) | ϕ=0 on Dδ}.

    Theorem 3.1. The constructed approximation uδn satisfies the following estimates:

    Esupt[0,T]uδn(t)2L2(Dδ)+ET0uδn(t)2L2(Dδ)dtC, (3.5)

    for some positive constant C.

    Proof. We first note that the solution of (3.3) is given by

    ρδn(t)=γt0e(a+γ)(ts)h(uδn(s))ds, (3.6)

    from which one obtains

    ET0ρδn(t)2L2(Γδ)dtγCET0h(uδn(t))2L2(Γδ)dt. (3.7)

    We apply Itô's formula to the integral equation (3.2) using the function ϕ(uδn(t))=|uδn(t)|2L2(Dδ), to get

    duδn(t)2L2(Dδ)+2κuδn(t)2L2(Dδ)dt+2δγ(h(uδn(t)),uδn(t))L2(Γδ)dt=2δγ((ρδn+φδ),uδn(t))L2(Γδ)dt+2(fδ1(uδn(t),uδn(t)),uδn(t))L2(Dδ)dt+2(gδ1(uδn(t))dW1,uδn(t))L2(Dδ)+gδ1(uδn(t))2L2(Dδ)dt. (3.8)

    Using Cauchy-Schwartz's inequality flowed by Young's inequality and the assumptions on the nonlinear term h(uδn), we obtain

    duδn(t)2L2(Dδ)+2κuδn(t)2L2(Dδ)dtδγ{uδn(t)2L2(Γδ)+ρδn+φδ2L2(Γδ)}dt+2(fδ1(uδn(t),uδn(t)),uδn(t))L2(Dδ)dt+2(gδ1(uδn(t))dW1,uδn(t))L2(Dδ)+gδ12L2(Dδ)dt. (3.9)

    Integrating over the interval (0,t), taking the supremum and the expectation to both sides of (3.9), we get

    Esupt[0,T]uδn(t)2L2(Dδ)+2κET0uδn(t)2L2(Dδ)dtPnuδ(0)2L2(Dδ)+CET0{uδn(t)2L2(Γδ)+ρδn+φδ2L2(Γδ)}dt+2ET0|(fδ1(uδn,uδn(t)),uδn(t))L2(Dδ)|dt+ET0gδ1(uδn(t))2L2(Dδ)dt+2Esupt[0,T]|T0(gδ1(uδn(t))dW1,uδn(t))L2(Dδ)|. (3.10)

    Using the bound (3.7), Remark 2.1, Assumptions (A1) and (A5), and Lemma 3.1, we obtain

    Esupt[0,T]uδn(t)2L2(Dδ)+2κET0uδn(t)2L2(Dδ)dtC1+ET0{C2uδn(t)2L2(Dδ)+C3δ2uδ2L2(Dδ)}dt+2ET0|(fδ1(uδn,uδn(t)),uδn(t))L2(Dδ)|dt+ET0gδ1(uδn(t))2L2(Dδ)dt+2Esupt[0,T]|T0(gδ1(uδn(t))dW1,uδn(t))L2(Dδ)|. (3.11)

    However, we have from (A2(ⅱ)) and Young's inequality

    2ET0|(fδ1(uδn,uδn(t)),uδn(t))L2(Dδ)|dt2CET0uδn(t)L2(Dδ)uδn(t)[L2(Dδ)]nλEsupt[0,T]uδn(t)2L2(Dδ)+C(λ)ET0uδn(t)2[L2(Dδ)]ndt, (3.12)

    for any λ>0. Also from (A3), one sees that

    ET0gδ1(uδn)2L2(Dδ)dtTC+CET0uδn(t)2L2(Dδ)dt. (3.13)

    Regarding the stochastic term in (3.11) we use Burkholder-Davis-Gundy's and Young inequalities as follows:

    2Esupt[0,T]|t0(gδ1(τ,uδn(τ))dW1(τ),uδn(τ))L2(Dδ)|2C1E[T0(gδ1(t,uδn(t)),uδn(t))2L2(Dδ)dt]122CE[T0gδ1(t,uδn(t))2L2(Dδ)uδn(t)2L2(Dδ)dt]122CEsupt[0,T]uδn(t)L2(Dδ)[T0gδ1(t,uδn(t)),uδn(t)2L2(Dδ)dt]121ηEsupt[0,T]uδn(t)2L2(Dδ)+CηET0gδ1(t,uδn(t)),uδn(t)2L2(Dδ)dt, (3.14)

    for any η>0. From (3.11)–(3.14) and Gronwall's inequality we have

    Esupt[0,T]uδn(t)2L2(Dδ)+ET0uδn(t)2L2(Dδ)dtC.

    We can achieve only weak limits for sub-sequences of the Galerkin approximate solution of our problem using the above estimates. Strong convergence requires some compactness, but because we are dealing with stochastic processes, bounds on the time derivative of the solution cannot be obtained. We have the following bounds for the finite difference of uδn.

    Theorem 3.2. Let uδn be the approximate solution of problem (1.1). Then under Assumptions (A1)–(A5) there exists a positive constant C, independent of δ, such that for all ρ1 and all s with t+s[0,T], we have

    Esup|s|ρT0uδn(t+s)uδn(t)2W1,2(Dδ)dtCρ. (3.15)

    Proof. For any s>0, the solution of problem (Pδ1) satisfies

    uδn(t+s)uδn(t)=κt+stΔuδn(τ)dτ+t+stfδ1(τ,uδn(τ),uδn(τ))dτ+t+stgδ1(τ,uδn(τ))dW1(τ),

    from this we obtain

    uδn(t+s)uδn(t)W1,2(Dδ)κt+stΔuδn(τ)dτW1,2(Dδ)+t+stfδ1(τ,uδn(τ),uδn(τ))dτW1,2(Dδ)+t+stgδ1(τ,uδn(τ))dW1(τ)W1,2(Dδ). (3.16)

    Let us estimates the terms on the right hand side of (3.16) as follows:

    κt+stΔuδn(τ)dτW1,2(Dδ)=supψW1,2δ,ψ=1|t+stκΔuδn(τ)dτ,ψ|supψW1,2δ,ψ=1Dδt+stκΔuδn(τ)dψdτdxsupψW1,2δ,ψ=1Dδt+stκuδn(τ)ψdτdx+supψW1,2δ,ψ=1δΓδt+stfδ2ψdτσ(dx)κsupψW1,2δ,ψ=1[t+stuδn(τ)L2(Dδ)ψL2(Dδ)dτ+δt+stfδ2L2(Γδ)ψL2(Γδ)dτ]Ct+stuδn(τ)L2(Dδ)+δ[h(uδn(τ))L2(Γδ)+vδn(τ)L2(Γδ)]dτ.

    Next, from the Cauchy-Schwartz inequality, we have

    ET0(t+stuδn(τ)L2(Dδ)dτ)2dtsCET0uδn(t)2L2(Dδ)dt. (3.17)

    Using Lemma 3.1, Eqs (3.4) and (3.7), we obtain

    EδT0(t+st[h(uδn(τ))L2(Γδ)+vδn(τ)L2(Γδ)]dτ)2dtCsET0[C1+C2uδn(τ)2L2(Dδ)+C3δ2uδn(τ)2L2(Dδ)]dt. (3.18)

    We also have from the embedding of L2(Dδ) into W1,2(Dδ), Assumption (A2(ⅱ)) and Cauchy inequality that

    t+stfδ1(τ,uδn(τ),uδn(τ))dτW1,2(Dδ)t+stfδ1(τ,uδn,uδn(τ))dτL2(Dδ)Ct+stuδn(τ)LDδdτCs12(t+stuδn(τ)LDδdτ)12. (3.19)

    Once again, we use the embedding of L2(Dδ) into W1,2(Dδ), followed by Itô's isometry, Cauchy-Schwartz inequality, and Fubini's theorem that

    Esup|s|ρT0t+stgδ1(τ,uδ1(τ))dW1(τ)2W1,2(Dδ)dtEsup|s|ρT0t+stgδ1(τ,uδ1(τ))dW1(τ)2L2(Dδ)dtET0sup|s|ρt+stgδ1(τ,uδ1(τ))dW1(τ)2L2(Dδ)dtT0(DδEsup|s|ρ(t+stgδ1(τ,uδ1(τ))dW1(τ))2dx)dtT0(Et+ρtgδ1(τ,uδ1(τ))2L2(Dδ)dτ)dtCT0(Et+ρt(1+uδ1(τ))2L2(Dδ))dτ)dtρCT. (3.20)

    From estimates (3.5), the assumptions on fδ1 and gδ1 (3.16)–(3.20), the proof is complete.

    As stated earlier, the goal is to take n goes to infinity in (uδn, ρδn) in order to arrive at (uδ, ρδ) a solution of the system (2.1)-(2.1). The estimates (3.5) aid in the extraction of convergent sub-sequences from (uδ, ρδ). However, the presence of the nonlinear function h in (2.1) necessitates strong compactness for uδn. Theorem 3.2, together with the estimates (3.5), we make sense of the set

    Zδ,n={ϕδnL,2(Ω;(0,T);L2(Dδ))L2,2(Ω;(0,T);Vδ)Esup|s|ρTs0ϕδn(t+s)ϕδn(t)2W1,2(Dδ)dtCρ.}

    Following Bensoussan [3] (and [19,24]), it is easily seen that Zδ,n is compactly embedded in L2,2(Ω;(0,T);L2(Dδ)). Having this, we define the set

    I=L2,2(Ω;(0,T);L2(Dδ))×C((0,T);Rm),

    and the mapping

    Φn: ΩI,ω(ϕδn,Wn).

    On the Borel set B(I) of the set I we define a probability measure as pullback of the mapping Φn as follows:

    μn(B)=P(Φ1n(A)),for alln and allBB(I).

    Again, following Bensoussan [3] (see also Sango [24] and Mohammed and Sango [19]), we show that the family {μn} is tight. From Prokhorov's compactness results [5], we obtain a subsequence {μnν} of {μn} and a probability measure μ such that μnν converging weakly to μ. As a consequence, by Skorokhod's compactness theorem [5], we are able to construct a new probability space (Ω1;F1;P1), set of random variables {(uδnν,Wnν)B(I):nνN} and (uδ,W) such that the probability law of (uδnν,Wnν) is μnν and the law of (uδ,W) is μ. In addition to that we have

    (uδnν,Wnν)(uδ,W)inB(I)  with probability  1. (3.21)

    In the following lemma, we transform our problem to a new setting in which the sequence of solutions has nicer properties and, more precisely, has a strong convergence in probability.

    Lemma 3.2. Define on the probability space(Ω1;F1;P1) the filtration F1t=σ({uδ(s),Ws}s[0,t]). Then

    W is an (F1t)t0 Brownian motion;

    (uδnν,Wnν) satisfies with probability 1,

    Dδuδnν(t)δidx+κt0Dδuδnν(s)δidxds+δγt0Γδh(uδnν(s))δiσ(dx)ds=DδPnνuδ0δidx+δt0Γδ(ρδnν+φδ)δiσ(dx)ds+t0Dδfδ1(uδnν(s),uδnν(s))δidxds+t0Dδgδ1(uδnν(s))δidxdWnν(s), (3.22)

    and

    ρδnνt+(a+γ)ρδnν=γh(uδnν),ρδnν(0)=0, (3.23)

    where

    ρδnν=vδnνφδ. (3.24)

    Before we proceed to the passage to the limit, we should note that nonlinearities pose a significant challenge in this regard. As for the nonlinear operator h that represents the interchanges of chemical flows across the boundary, Bituin and Donato [6] have examined a number of its properties in the elliptic framework. In the following lemma, we give the evaluation version of its major properties (see [6,Theorem 3.2,p. 179]).

    Lemma 3.3. Let h be as in (A4) and s and its conjugate s, i.e., (1s+1s=1), are such that

    {s[1,)andsps,forn=2,s=2(n1)n2ands=22n,forn>2.

    Then, the application

    vL2(0,T;W1,2(Dδ))h(t,v)L2(0,T;Ls(Γδ))

    is weakly continuous, i.e., for all bounded sequence {vn} in L2(0,T;W1,2(Dδ)), one has up to a sub-sequence

    h(t,vn)h(t,v)weaklyinL2(0,T;Ls(Γδ)). (3.25)

    Note that, estimates (3.5) still true for the variational form (3.22), which gives the P1–a.s. weak and weak convergence

    uδnνuδ,weaklyinL,2(Ω;(0,T);L2(Dδ)), (3.26)
    uδnνuδ,weaklyinL2,2(Ω;(0,T);Vδ)). (3.27)

    The limits (3.26) and (3.27) allow us to pass to the limit in the first two terms on the left-hand side of (3.22). In order to pass to the limit in the third term on the left-hand side of (3.22), we use the weak convergence (3.25) and obtain

    limnνδγt0Γδh(uδnν(s))δiσ(dx)ds = δγt0Γδh(uδ(s))δiσ(dx)dsP1–a.s. (3.28)

    Regarding the initial condition, it is known that the projection Pnνuδ0 converges strongly to uδ0 in the L2-norm. We also have

    ρδnν(t)=γt0e(a+γ)(ts)h(uδnν(s))ds, (3.29)

    this together with (3.25) give

    ρδnν(t)ρδ(t),weaklyinC([0,T];Ls(Γδ))P1–a.s., (3.30)

    this gives the corresponding limit of the second term on the right-hand side of (3.22). It also gives

    limnνvδnν=ρδ+φδ = vδ.

    For the third term on the right-hand side of (3.22), we have

    t0Dδfδ1(uδnν(s),uδnν(s))δidxds=t0DδFδ1(uδnν(s))uδnν(s)δidxds=t0Dδ(Fδ1(uδnν(s))Fδ1(uδ))uδnν(s)δidxds+t0DδFδ1(uδ(s))uδnν(s)δidxds. (3.31)

    The first term on the right-hand side of (3.31), will converge to zero due to the assumption (A2(ⅰ)) and the strong convergence (3.21). Then, we pass to the limit on the second term using the weak limit (3.27), we obtain

    limnνt0Dδfδ1(uδnν(s),uδnν(s))δidxds=t0Dδfδ1(uδ(s),uδ(s))δidxds P1–a.s. (3.32)

    Now, let us show that

    limnνt0Dδgδ1(uδnν(s))δidxdWnν(s)=t0Dδgδ1(uδ(s))δidxdW(s)P1–a.s., (3.33)

    for this, we write

    t0Dδgδ1(uδnν(s))δidxdWnν(s)=t0Dδgδ1(uδnν(s))δidxd(Wnν(s)W(s))+t0Dδgδ1(uδnν(s))δidxdW(s). (3.34)

    Because the variation (Wnν(s)W(s)) is unbounded, the passage to the limit on the first term of the right-hand side of (3.34) requires special attention, so for one to benefit from the strong P1–a.s. convergence (3.21). We utilize the concept of regularization

    gδ,η1(t,uδnν(t))=1ηt0ψ(tsη)gδ1(s,uδnν(s))ds,

    where the function ψ is the standard mollifier. By definition, we have

    gδ,η1(t,uδnν(t))gδ1(t,uδnν(t))inL2,2(Ω;(0,T);L2(Dδ))asη0, (3.35)

    also, gδ,η1 is bounded by gδ1 and differentiable in time. We write

    t0Dδgδ1(s,uδnν(s))δidxd(Wnν(s)W(s))=t0Dδ[gδ1(s,uδnν(s))gδ,η1(s,uδnν(s))]δidxd(Wnν(s)W(s))+t0Dδgδ,η1(s,uδnν(s))δidxd(Wnν(s)W(s)). (3.36)

    We deal with the right-hand-side as follows: For the first term we use Burkholder-Davis-Gundy's inequality and the strong convergence (3.35),

    Esup0st|t0Dδ(gδ1(s,uδnν(s))gδ,η1(s,uδnν(s)))δidxd(Wnν(s)W(s))|CE[t0Dδ(gδ1(s,uδnν(s))gδ,η1(s,uδnν(s)))2(δi)2dxds]12CE[t0gδ1(s,uδnν(s))gδ,η1(s,uδnν(s))2L2(Dδ)δi2L2(Dδ)ds]12Cgδ1(s,uδnν(s))gδ,η1(s,uδnν(s))L2,2(Ω;(0,T);L2(Dδ)). (3.37)

    Clearly, the last term converges to zero P–a.s as η tends towards 0.

    Integration by parts, assumptions on the nonlinear function (gδ1(t,uδnν(t)), and the strong convergence (3.21) lead to

    t0Dδgδ,η1(s,uδnν(s))δidxd(Wnν(s)W(s))=(Wnν(t)W(t))Dδgδ,η1(s,uδnν(s))δidxt0(Wnν(s)W(s))(Dδtgδ,η1(s,uδnν(s))δidx)ds. (3.38)

    The right hand side of (3.38) tends towards 0 P–a.s. as nν tends to .

    Using (3.37) and (3.38) into (3.34), we arrive at

    limnνt0Dδgδ1(s,uδ(s))δidxdWnν(s)=t0Dδgδ1(s,uδ(s))δidxdW(s). (3.39)

    However,

    gδ1(t,uδ(t))gδ1(t,uδ(t))inL2,2(Ω;(0,T);L2(Dδ))asnν. (3.40)

    We use Rozovskii's convergence theorem for stochastic integrals [23,Theorem 4, p. 63] to assert the following:

    limnνt0Dδgδ1(s,uδnν(s))δidxdW(s)=t0Dδgδ1(s,uδ(s))δidxdW(s). (3.41)

    Combining the above convergence, we obtain

    Dδuδ(t)δidx+κt0Dδuδ(s)δidxds+δγt0Γδh(uδ(s))δiσ(dx)ds=Dδuδ0δidx+δt0Γδ(ρδ+φδ)δiσ(dx)ds+t0Dδfδ1(s,uδ(s),uδ(s))δidxds+t0Dδgδ1(s,uδ(s))δidxdW(s), (3.42)

    and

    ρδt+(a+γ)ρδ=γh(uδ),ρδ(0)=0, (3.43)

    where

    ρδ=vδφδ. (3.44)

    This concludes the demonstration that (Ω,P,F,uδ,vδ) is a weak probabilistic solution to the system (1.1)-(1.2).

    In this part, we prove the path-wise uniqueness of the solution of problem (1.1). We assume that uδ1 and uδ2 solve the system (3.42)–(3.44) and let wδ=uδ1uδ2. Applying Itô's formula to the function ϕ(wδ(t))=wδ(t)2L2(Dδ), we get

    dwδ(t)2L2(Dδ)+2κwδ(t)2L2(Dδ)dt+2δγ(h(uδ1(t))h(uδ2(t)),uδ1(t)uδ2(t))L2(Γδ)dt=2δγ(ρδ1ρδ1,uδ1(t)uδ2(t))L2(Γδ)dt+2(fδ1(uδ1(t),uδ1(t))fδ1(uδ2(t),uδ2(t)),uδ1(t)uδ2(t))L2(Dδ)dt+2(gδ1(uδ1(t))gδ1(uδ2(t)),uδ1(t)uδ2(t))L2(Dδ)dW(t)+gδ1(uδ1(t))gδ1(uδ1(t))2L2(Dδ)dt, (4.1)

    where

    ρδ1(t)=γt0e(a+γ)(ts)h(uδ1(s))ds, (4.2)

    and

    ρδ2(t)=γt0e(a+γ)(ts)h(uδ2(s))ds. (4.3)

    Integrating (4.1) over (0,t), taking the expectation, the supremum, and using the Assumptions (A4(ⅰ)), we obtain

    Esups[0,t]wδ(s)2L2(Dδ)+2κEt0wδ(s)2L2(Dδ)ds2δγEt0|(ρδ1(s)ρδ2(s),uδ1(s)uδ2(s))L2(Γδ)|ds+2Et0|(fδ1(uδ1(s),uδ1(s))fδ1(uδ2(s),uδ2(s)),uδ1(s)uδ2(s))L2(Dδ)|ds+2Esups[0,t]|t0(gδ1(uδ1(s))gδ1(uδ2(s)),uδ1(s)uδ2(s))L2(Dδ)dW(s)|+Et0gδ1(uδ1(s))gδ1(uδ1(s))2L2(Dδ)dt. (4.4)

    Now, we shall estimate each of the terms on the right side of (4.4). We start by

    Et0|(ρδ1(s)ρδ2(s),uδ1(s)uδ2(s))L2(Γδ)|dsEt0ρδ1(s)ρδ1(s)L2(Γδ)wδ(s)L2(Γδ)ds. (4.5)

    We use Assumption (A4) to derive the following:

    ρδ1(t)ρδ1(t)L2(Γδ)γt0e(a+γ)(tτ)h(uδ1)(τ)h(uδ2)(τ)L2(Γδ)dτγsup0sth(uδ1)(s)h(uδ2)(s)L2(Γδ)t0e(a+γ)(tτ)dτγa+γsup0sth(uδ1)(s)h(uδ2)(s)L2(Γδ)γLa+γsup0stuδ1(s)uδ2(s)L2(Γδ). (4.6)

    From (4.5), (4.6) and Lemma 3.1, we arrive at

    \begin{eqnarray} 2\delta\gamma\mathbb{E}\int_{0}^{t}|(\rho^{\delta}_{1}(s)-\rho^{\delta}_{1}(s), w^{\delta}(s))_{L^{2}(\Gamma^\delta)}|ds &\leq &\frac{2\delta\gamma^2 L}{a+\gamma} \mathbb{E}\int_{0}^{t} \sup\limits_{0\leq s \leq \tau}\|w^{\delta}(s)\|_{L^{2}(\Gamma^\delta)}^{2}d\tau\\ &\leq & \beta \alpha \mathbb{E}\int_{0}^{t} \sup\limits_{0\leq s \leq \tau} \Big\{\|w^{\delta}(s)\|_{L^{2}(D^\delta)}^{2}+\delta^2\|\nabla w^{\delta}(s)\|_{L^{2}(D^\delta)}^{2}\Big\}d\tau, \end{eqnarray} (4.7)

    where \beta = \frac{2\gamma^2 L}{a+\gamma} . Second, we use the estimate in Theorem 3.1, Cauchy-Schwartz and Poincaré inequalities and Assumption (A2), we have

    \begin{eqnarray} && \mathbb{E}\int_{0}^{t}|(f^{\delta}_{1}(u^{\delta}_{1}(s), \nabla u^{\delta}_{1}(s))-f^{\delta}_{1}(u^{\delta}_{2}(s), \nabla u^{\delta}_{2}(s)), u^{\delta}_{1}(s)-u^{\delta}_{2}(s))_{L^{2}(D^{\delta})}|ds\\ & = &\mathbb{E}\int_{0}^{t}|\left(\left(F^{\delta}_{1}(u^{\delta}_{1}(s))\cdot\nabla u^{\delta}_{1}(s)-F^{\delta}_{1}(u^{\delta}_{2}(s))\cdot \nabla u^{\delta}_{2}\right)(s), w^{\delta}(s)\right)_{L^{2}(D^{\delta})}|ds\\ & = & \mathbb{E}\int_{0}^{t}|(F^{\delta}_{1}(u^{\delta}_{1}(s))\cdot\nabla u^{\delta}_{1}(s)-F^{\delta}_{1}(u^{\delta}_{1}(s))\cdot \nabla u^{\delta}_{2}(s)\\ & & \qquad +F^{\delta}_{1}(u^{\delta}_{1}(s)) \cdot \nabla u^{\delta}_{2}(s)-F^{\delta}_{1}(u^{\delta}_{2}(s)) \cdot \nabla u^{\delta}_{2}(s), w^{\delta}(s))_{L^{2}(D^{\delta})}|ds\\ & = &\mathbb{E}\int_{0}^{t}|(F^{\delta}_{1}(u^{\delta}_{1}(s)) \cdot\nabla w^{\delta}(s) +(F^{\delta}_{1}(u^{\delta}_{1}(s))-F^{\delta}_{1}(u^{\delta}_{2}(s)))\cdot \nabla u^{\delta}_{2}(s), w^{\delta}(s))_{L^{2}(D^{\delta})}|ds\\ & \leq & C_1\mathbb{E}\int_{0}^{t}\{\|\nabla w^{\delta}(s)\|_{L^{2}(D^{\delta})}\|w^{\delta}(s)\|_{L^{2}(D^{\delta})} +\| w^{\delta}(s)\|_{L^{2}(D^{\delta})}^{2}\}ds\\ & \leq & C_1\mathbb{E}\int_{0}^{t}\{\lambda\|\nabla w^{\delta}(s)\|_{L^{2}(D^{\delta})}^{2} +(1+C_\lambda)\| w^{\delta}(s)\|_{L^{2}(D^{\delta})}^{2}\}ds, \end{eqnarray} (4.8)

    where \lambda>0 is small enough. Regarding the last two terms on the right-hand side of (4.4), we use Burkholder-Davis-Gundy's, Cauchy-Schwartz, and Young's inequalities and Assumption (A3) to get

    \begin{eqnarray} && 2\mathbb{E}\sup\limits_{s\in [0, T]} \left|\int_{0}^{t}(g^{\delta}_{1}(u^{\delta}_{1})(s))-g^{\delta}_{1}(u^{\delta}_{2}(s)), u^{\delta}_{1}(s)-u^{\delta}_{2}(s))_{L^{2}(D^\delta)}dW(s)\right|\\ &&+\mathbb{E}\int_{0}^{t}\|g^{\delta}_{1}(u^{\delta}_{1}(s))-g^{\delta}_{1}(u^{\delta}_{1}(s))\|^{2}_{L^{2}(D^\delta)}ds \\ &\leq & 2C\mathbb{E} \left(\int_{0}^{T}(g^{\delta}_{1}(u^{\delta}_{1}(s))-g^{\delta}_{1}(u^{\delta}_{2}(s)), u^{\delta}_{1}-u^{\delta}_{2}(s))_{L^{2}(D^\delta)}^{2}ds\right)^{\frac{1}{2}}\\ && +\mathbb{E}\int_{0}^{T}\|g^{\delta}_{1}(u^{\delta}_{1}(s))-g^{\delta}_{1}(u^{\delta}_{1}(s))\|^{2}_{L^{2}(D^\delta)}ds\\ &\leq & C \mathbb{E}\int_{0}^{T}\|w^{\delta}(s)\|^{2}_{L^{2}(D^\delta)}ds. \end{eqnarray} (4.9)

    Using estimates (4.7)–(4.9) into (4.4), we obtain

    \begin{eqnarray} && \mathbb{E}\sup\limits_{s\in [0, T]}\|w^{\delta}(s)\|^{2}_{L^{2}(D^\delta)}+2\left(\kappa-\frac{\beta \alpha \delta^2}{2}-\frac{C_1 }{2}\lambda\right)\mathbb{E} \int_{0}^{T}\|\nabla w^{\delta}(s)\|^{2}_{L^{2}(D^\delta)}ds\\ &\leq & C\mathbb{E}\int_{0}^{T} \sup\limits_{0\leq s\leq \tau }\|w^{\delta}(s)\|_{L^{2}(D^\delta)}^{2}d \tau, \end{eqnarray} (4.10)

    choosing \lambda = \frac{1}{C_1}(\kappa-\frac{\beta \alpha \delta^2}{2}) and setting \Lambda = \kappa-\frac{\beta \alpha \delta^2}{2} , we obtain

    \begin{eqnarray} \mathbb{E}\sup\limits_{s\in [0, T]}\|w^{\delta}(s)\|^{2}_{L^{2}(D^\delta)}+\Lambda\mathbb{E} \int_{0}^{T}\|\nabla w^{\delta}(s)\|^{2}_{L^{2}(D^\delta)}ds &\leq & C\mathbb{E}\int_{0}^{T} \sup\limits_{0\leq s\leq \tau }\|w^{\delta}(s)\|_{L^{2}(D^\delta)}^{2}d \tau. \end{eqnarray} (4.11)

    It is worth noting that, because \delta represents the hole size, we may always choose it so that \kappa>\frac{\beta \alpha \delta^2}{2} . We come to the conclusion that for each t \in [0, T] , we have w^{\delta}(t) = 0 , \mathbb{P} –a.s. i.e., u^{\delta}_{1}(t) = u^{\delta}_{2}(t) , \mathbb{P} –a.s., thanks to the Gronwall's inequality. We also have

    \begin{eqnarray} \|v^{\delta}_{2}(t)-v^{\delta}_{1}(t)\|_{L^{2}(\Gamma^\delta)} = \|\rho^{\delta}_{2}(t)-\rho^{\delta}_{1}(t)\|_{L^{2}(\Gamma^\delta)}. \end{eqnarray} (4.12)

    This and (4.6) give the uniqueness of the solution of (1.2). As previously stated, in light of the Yamada-Watanabe theorem [22], we obtain the existence and uniqueness of a strong probabilistic solution to our model based on the obtained path-wise uniqueness and probabilistic weak solution of the previous section.

    The goal of this section is to prove the positivity of the solutions of our system, which can be rewritten as follows:

    \begin{equation} \label{system} \left\{ \begin{array}{rcll} du^\delta & = & \kappa\Delta u^\delta dt + f^\delta_1(t, u^{\delta}, \nabla u^{\delta} )dt+g_{1}^{\delta}(t, u^{\delta})dW_{1}, & \text{ in }\Omega \times D^\delta \times (0, T), \\ dv^{\delta} & = & (f^\delta_{2}-av^\delta )dt +g^{\delta}_{2}dW_{2}, & \text{ in }\Omega \times \Gamma^\delta \times (0, T), \\ u^{\delta} & = & 0, & \text{ in } \Omega\ \times \partial D \times (0, T), \\ u^\delta(0)& = & u^\delta_{0}, & \text{ in }D^\delta , \notag \\ -\kappa \frac{\partial u^\delta}{\partial \nu} & = &\delta f^\delta_{2} = \delta\gamma(h(t, u^{\delta})-v^\delta), & \text{ in }\Omega \times \Gamma^\delta \times (0, T), \\ v^{\delta}(0)& = & v^\delta_{0}, & \text{ in }\Gamma^\delta. \end{array} \right. \end{equation}

    More precisely, we shall prove under some additional conditions that u^{\delta}(t, x) \geq 0 \mathbb{P} –a.s. and v^{\delta}(t, x) \geq 0 \mathbb{P} –a.s. To this end, we apply comparison principles for SPDEs and SDEs. Since the solution u^{\delta} depends only on v^{\delta} via Neumann's boundary condition, it is enough to check the positivity for the separated equations

    \begin{equation} \left\{ \begin{array}{rcll} du^\delta & = & \kappa\Delta u^\delta dt + f^\delta_1(t, u^{\delta}, \nabla u^{\delta} )dt+g_{1}^{\delta}(t, u^{\delta})dW_{1}, & \text{ in }\Omega \times D^\delta \times (0, T), \\ u^\delta(0)& = & u^\delta_{0}, & \text{ in }D^\delta, \end{array} \right. \end{equation} (5.1)

    and

    \begin{equation} \left\{ \begin{array}{rcll} dv^{\delta} & = & (\gamma(h(t, u^{\delta}) - v^\delta )-av^\delta )dt +g^{\delta}_{2}dW_{2}, & \text{ in }\Omega \times \Gamma^\delta \times (0, T), \\ v^{\delta}(0)& = & v^\delta_{0}, & \text{ in }\Gamma^\delta. \end{array} \right. \end{equation} (5.2)

    Theorem 5.1. (i) If u^\delta_{0}\geq 0 and f^\delta_1(t, u, \nabla u)\geq 0 for all t and u , then u^\delta \geq 0 \mathbb{P} –a.s.

    (ii) If v^\delta_{0}\geq 0 and \gamma(h(t, u)-v) \geq av for all t and u, v , then v^\delta \geq 0 \mathbb{P} –a.s.

    Proof. For (ⅰ), according to the result of Denis et al. [12, Theorem 5, p. 518], it follows that if the initial condition u^\delta_{0} and the drift coefficient f^\delta_1(t, u, v) are non-negative, then the solution of the SPDE (5.1) is also non-negative.

    Similarly, for (ⅱ), we apply the comparison principle for SDEs and refer to [14] for more details on related results. Hence, if the initial condition v^\delta_{0} and the drift coefficient \gamma(h(t, u)-v) - av are non-negative, then the solution of the SDE (5.2) is also non-negative.

    This section deals with numerical experiments to illustrate and support the above theoretical results. To perform these numerical experiments, we consider the coupled stochastic partial differential system (1.1)-(1.2) with Neumann boundary conditions and initial conditions on the perforated domain D = (0, 1)\times (0, 1) with D^\delta = D \setminus \bigcup_{i = 1}^4 \Gamma^\delta_i and \Gamma^\delta_i = \{(x_i, y_i) \in \mathbb{R}^2; (x - x_i)^2 + (y - y_i)^2 \leq 0.042\}, where x_i, y_i \in \{1/8, 1/4, 3/8, 1/2, 5/8, 6/8\}, see Figure 2.

    \begin{eqnarray} \left\{ \begin{array}{rlll} \frac{\partial u^\delta }{\partial t}&\! = \! & \left(\kappa_x\frac{\partial^2u^\delta}{\partial x^2}+\kappa_y\frac{\partial^2u^\delta}{\partial y^2}\right) + \frac{\kappa\cdot\nabla u^\delta}{1+e^{-u^\delta}}+g_1(u^\delta)\frac{\partial W_1(t) }{\partial t}, &\text{ in }\Omega \times D^{\delta} \times (0, T), \\ \frac{\partial v^\delta }{\partial t}+av^\delta &\! = \! & \gamma(\phi(t, u^\delta)- v^\delta ) +g_2(t, u^\delta)\frac{\partial W_{2}(t)}{\partial t}, &\text{ in }\Omega \times \Gamma^{\delta} \times (0, T), \\ u^\delta(t, x, y) &\! = \! & 0, &\text{ on } \partial D \times (0, T), \\ \quad u^\delta (0, x, y) &\! = \! & u^\delta _{0}, &\text{ in }D^{\delta} , \notag \\ -\kappa \frac{\partial u }{\partial n} &\! = \! & \delta \gamma (\phi(t, u^\delta)- v^\delta ), &\text{ in }\Omega \times \Gamma ^{\delta} \times (0, T), \\ v^\delta(0, x, y) &\! = \! & v^\delta _{0}, &\text{ in }\Gamma^{\delta}. \end{array} \right. \end{eqnarray} (6.1)
    Figure 2.  Mesh of perforated media D used in the simulations.

    In order to discretize the preceding system in time, we partition the interval I = (0, T) into M subintervals I_n = (t_{n-1}, t_n], where t_0 = 0 < t_1 < \ldots < t_N = T. We set the time length of each step as \Delta t : = t_n-t_{n-1}. The main idea behind variational time discretization is that we approximate the solution v defined in I via a piecewise polynomial function with unstructured mesh. For the stochastic terms g^\delta_{1}(t, u^\delta) \frac{\partial W_1}{\partial t} and g^\delta_{2}(t, u^\delta) \frac{\partial W_2}{\partial t}, we use the Euler-Maruyama method. All the numerical experiments are developed in FEniCSx Project and performed with Python on an Intel Xeon (4-core) with 3.60 GHz and 16 GB RAM. The DOLFINx version utilized was 0.5.0. The analysis for nonconforming function spaces was carried out with both the direct and Krylov solvers. This transient analysis solves the variational problems T = 2 times, resulting in a substantially larger analysis to run than the steady-state analysis. As a result, the execution time of this program is significantly longer than that of the steady-state program. The discrete problem with Neumann and Boundary conditions is addressed in a For loop from t = 0 to t = T with the time step \Delta t. Thus, the discrete problem is solved at each time step, and the result is saved in an XDMF file within the loop.

    To fix the ideas, we choose in all experiments \gamma = 0.35, a = 0.25, \vartheta = 0.5, \kappa = 2, \delta = 7 and the diffusion coefficient \kappa = 0.035 (but any other combination of coefficients would serve our purposes), and the initial conditions are given by (u^{\delta}_{0}, v^{\delta}_{0}) = (x^{\frac{-1}{4}}, x^{\frac{-1}{4}})\in L^{2}(D^\delta)\times L^{2}(\Gamma^{\delta}). The nonlinear function \phi describing the interchanges of chemical flow across the surface of the skeleton will be discussed in the two following examples.

    In this test-case, we consider the case where the reaction of the chemical flow on the skeleton surface is modeled by Langmuir kinetics, i.e., we let

    \begin{eqnarray} \phi(u) & = &\vartheta \frac{\kappa u}{1+\kappa u}, \quad \kappa, \ \vartheta > 0, \end{eqnarray} (6.2)

    where \phi(u) represents the adsorption isotherm in the fluid phase, \vartheta is the adsorption capacity, and \kappa is the isotherm Langmuir constant (see [6,7,8,10,11]), and we let u\neq \frac{-1}{\kappa}, \text{for all} (t, x)\in (0, T)\times D^{\delta}. With this setting, if u^{\delta}\geq 0 in D^{\delta}, it is clear that the Assumption (A4) is met by Langmuir kinetics and as a result, we have that (u^{\delta}, v^{\delta}) satisfies

    \begin{eqnarray} && \int_{D^{\delta}} u^{\delta}(t)\psi dx - \int_{D^{\delta}} u^{\delta}(0)\psi dx +\kappa\int_{0}^{t}\int_{D^{\delta}}\nabla u^{\delta}\nabla \psi dx ds +\delta \gamma \kappa \vartheta\int_{0}^{t}\int_{\Gamma^{\delta}} \frac{ u^{\delta}\psi}{1+\kappa u^{\delta}} \sigma(dx) ds \\ & = & \delta \gamma \int_{0}^{t}\int_{\Gamma^{\delta}} v^{\delta} \psi \sigma(dx) ds +\int_{0}^{t}\int_{D^{\delta}}f_{1}^{\delta}(u^\delta, \nabla u^{\delta}) \psi dx ds + \int_{0}^{t}\int_{D^{\delta}}g_{1}^{\delta}(u^\delta) \psi dx dW_1(s), \end{eqnarray} (6.3)

    for all 0\leq t \leq T and \psi \in \mathcal{V}_\delta and

    \begin{eqnarray} v^{\delta}(t, x) & = & v^{\delta}_{0}(x) e^{-(a+\gamma)t}+\gamma \kappa \vartheta \int_{0}^{t} e^{-(a+\gamma)(t-s)}\frac{ u^{\delta}(s, x)}{1+\kappa u^{\delta}(s, x)}ds + \int_{0}^{t}e^{-(a+\gamma)(t-s)} g_{2}^{\delta}(s)dW_2(s), \end{eqnarray} (6.4)

    and Theorems 2.1 and 2.2 hold true.

    The deterministic solute concentration represents the average behavior of the system under fixed parameters, providing a reliable baseline for comparison. In contrast, the stochastic solute concentration accounts for the inherent randomness and variability present in the system, capturing a more comprehensive view of the possible fluctuations in concentration levels due to uncertainties in the parameters or external influences. This dual analysis allows us to better understand the solute behavior under different scenarios and is critical for validating the effectiveness of the numerical methods employed.

    In the Figure 3 below, the solution was computed using a non-conforming finite element method with a spatial mesh size of h and a time step size of \Delta t. This approach enables the analysis of both deterministic and stochastic solute concentrations within the considered model. We can observe that all oscillations are strongly distributed and extend far from the skeleton's surface due to the use of Langmuir kinetics functions and the associated parameters. This explains the absence of significant disturbances at the skeleton level.

    Figure 3.  Solution computed with nonconforming finite element for spatial mesh h = 2^{-6} and time mesh \Delta t = 10^{-6}. The deterministic solute concentration (left), and stochastic solute concentration (right) along the axes X and Y.

    To analyze the quality of our numerical approximations, we look at an estimate of a strong mean-square error in L^2 -norm, for spatial mesh-size h and temporal mesh-size \Delta t, by computing the following formula:

    \begin{equation} \|u-u_h\|_{2} = \left(\mathbf{E}\big[\|u-u_h\|^2_{L^2(D))}\big]\right)^{1/2}, \end{equation} (6.5)

    where u_h represents a finite element approximation of u within the finite element space V_h, and this approximation can be expressed in terms of the finite element basis \{\psi^i_n\}_{i = 1}^N of V_h and for each time t_n :

    \begin{equation} u_h(t) = \sum\limits_{i = 1}^N m_i \psi^i_n(t), \quad m_i\in\mathbf{R}. \end{equation} (6.6)

    Since an exact solution for stochastic partial differential equation u cannot easily be found, we must instead use a numerical solution for partial differential equation within g_1 = g_2 = 0. This approximation is chosen to be the finite element solution of PDE in a deterministic case, denoted by u^{deter}_h, solved in refined uniform mesh (see Figure 2) with a total of N = 2^{11}+1 vertices and mesh-size h_{ref} = 2^{-11}. We now estimate the strong error for u_h^{stoch} on the mesh sizes h = 2^{-k}, \, k\in\{1, 2, 3, \ldots, 10\} with the refined solutions solved on a grid with mesh-size h_{ref} and time mesh \Delta t \in \{10^{-2}; 5.10^{-2}; 2, 5.10^{-3}; 1, 25.10^{-3}; 6, 25.10^{-4} \}.

    In Table 1, we present the computed errors for the solute concentration in the fluid in energy and L^2 norms and the solute concentration on the surface of the skeleton in L^2 norm, on a sequence of uniformly refined meshes for space, as well as the convergence order computed from the errors on two successively refined meshes (with N and 4N elements). Similarly, in Table 2 we present the computed errors for accuracy time. The convergent rate in both time and space is assessed by means of mean-square approximation errors at the endpoint T = 0.2, which are induced by both temporal and spatial discretizations. The anticipated amount of error is approximated by taking the mean of 500 samples for every efficiency test.

    Table 1.  Errors and convergence orders for SPDE respect to spatial discretization N.
    N \|u^{stoch}_{N}-u^{deter}_{N}\|_{2} order \|v^{stoch}_{N}-v^{deter}_{N}\|_{2} order
    64 0.99345676 - 0.089876544 -
    256 0.266874324 1.896296676 0.049976436 0.846696627
    1 024 0.067711235 1.978693384 0.025293729 0.982468273
    4 096 0.016987634 1.994910291 0.009999097 1.338910092
    16 384 0.004213546 2.011378151 0.004719324 1.083217588

     | Show Table
    DownLoad: CSV
    Table 2.  Errors and convergence orders for SPDE respect to time discretization ∆t.
    \Delta \|u^{stoch}_{N}-u^{deter}_{N}\|_{2} order \|v^{stoch}_{N}-v^{deter}_{N}\|_{2} order
    1\times 10^{-2} 0.008251298 - 0.006546875 -
    5\times 10^{-3} 0.005597649 0.559800051 0.004553468 0.523840815
    2.5\times 10^{-3} 0.003847823 0.540778563 0.003245677 0.488446184
    1.25\times 10^{-3} 0.002636498 0.545419753 0.002284771 0.506469862
    6.25\times 10^{-4} 0.001854372 0.507691996 0.001585238 0.527350117

     | Show Table
    DownLoad: CSV

    Utilizing the theoretical findings, we numerically obtain the optimal convergence rates concerning spatial discretization and validate the theoretical results for the spatial mesh size

    \begin{eqnarray} \left\{ \begin{array}{rcl} \|u^{stoch}_{N}-u^{deter}_{N}\|_{2}& = &\mathcal{O}(N^{-1}) = \mathcal{O}(h^2), \\ \|v^{stoch}_{N}-v^{deter}_{N}\|_{2}& = &\mathcal{O}(N^{-1/2}) = \mathcal{O}(h), \end{array} \right. \end{eqnarray} (6.7)

    and for time-mesh \Delta t,

    \begin{eqnarray} \left\{ \begin{array}{rcl} \|u^{stoch}_{\Delta t}-u^{deter}_{\delta t}\|_{2} & = &\mathcal{O}(\Delta t^{-1/2}), \\ \|v^{stoch}_{\Delta t}-v^{deter}_{\Delta t}\|_{2}& = &\mathcal{O}(\Delta t^{-1/2}). \end{array} \right. \end{eqnarray} (6.8)

    Another well-known example is Freundlich kinetics, where we allow the chemical reaction to proceed on the skeleton surface with the formula

    \begin{eqnarray*} \phi(t, u) = |u|^{\lambda-1}u, \quad 0 < \lambda < 1, \end{eqnarray*}

    where \lambda is called the reaction order. It is seen that this function is not smooth, although \phi(t, 0) = 0, \phi(t, u) is defined for all values u(t, x) \in \mathbb{R} and continuous and |\phi(t, u)|\leq C(1+|u|)) for some constant C>0. One of this function's key properties, as stated in [11], is the existence of a convex and lower semi-continuous function \Phi :(0, T)\times\mathbb{R}\rightarrow \mathbb{R}, such that

    \begin{eqnarray*} \Phi(t, u)& = &\int_{0}^{u}\phi(t, s)ds. \end{eqnarray*}

    To study the existence and uniqueness in this case, one could approximate the system (1.1)-(1.2), using Yosida regularization, so for \eta >0, we have

    \begin{eqnarray} && \int_{D^{\delta}} u^{\delta, \eta}\psi dx - \int_{D^{\delta}} u^{\delta, \eta}_{0}\psi dx +\kappa\int_{0}^{t}\int_{D^{\delta}}\nabla u^{\delta, \eta}\nabla \psi dx ds +\delta \gamma \kappa \vartheta\int_{0}^{t}\int_{\Gamma^{\delta}} \phi^{\eta}(s, u^{\delta, \eta} )\psi \sigma(dx) ds \\ & = & \delta \gamma \int_{0}^{t}\int_{\Gamma^{\delta}} v^{\delta, \eta} \psi \sigma(dx)ds +\int_{0}^{t}\int_{D^{\delta}}f_{1}^{\delta}(u^{\delta, \eta}, \nabla u^{\delta, \eta}) \psi dx ds + \int_{0}^{t}\int_{D^{\delta}}g_{1}^{\delta}(u^{\delta, \eta}) \psi dx dW_1(s), \end{eqnarray} (6.9)

    for all 0\leq t \leq T and \psi \in \mathcal{V}_\delta and

    \begin{eqnarray} v^{\delta, \eta}(t, x)& = & v^{\delta, \eta}_{0}(x) e^{-(a+\gamma)t}+\gamma \kappa \vartheta \int_{0}^{t} \phi^{\eta}(s, u^{\delta, \eta}(s, x) )ds + \int_{0}^{t}e^{-(a+\gamma)(t-s)} g_{2}^{\delta, \eta}(s)dW_2(s), \end{eqnarray} (6.10)

    where

    \begin{eqnarray*} \phi^{\eta}(t, u^{\delta, \eta})& = &\frac{I-(I+\eta\partial_{x} \Phi(t, u^{\delta, \eta} ))^{-1}}{\eta}. \end{eqnarray*}

    Now, the function \phi^{\eta}(t, u^{\delta, \eta}) can be smoothed out with the standard mollifier, and the existence and uniqueness results for the system (1.1)-(1.2) can be easily obtained.

    Similarly to the previous numerical example, in the Figure 4 below, the solution was computed using a non-conforming finite element method with a spatial mesh size of h and a time step size of \Delta t. This approach enables the analysis of both deterministic and stochastic solute concentrations within the considered model. Unlike the previous example, all oscillations are sparsely distributed and have a small size, but they extend widely around the skeleton's surface due to the shape of the Freundlich kinetics function and the associated parameter \lambda \in (0, 1). This explains the presence of significant disturbances at the skeleton level.

    Figure 4.  Solution computed with non-conforming finite element for spatial mesh h = 2^{-6} and time mesh \Delta t = 10^{-6}. The deterministic solute concentration (left), and stochastic solute concentration (right) along the axes X and Y.

    To illustrate the structure of the solute concentration u in dependence on the mesh size of the perforated domain, we include samples next to the convergence plots by using the previous formula (6.5). In this case, we consider the numerical solute concentration obtained by the finite element solution of PDE as an exact solution, denoted by u^{deter}_h which is solved in refined uniform mesh (see Figure 2) with a total of N = 2^{11}+1 vertices and spatial mesh h = 2^{-11} and time mesh \Delta t = 10^{-6}. Similarly to the previous test case, we estimate the strong error for u_h^{deter} on the mesh sizes h = 2^{-k}, \, k\in\{1, 2, 3, \cdots, 10\} with the refined solutions solved on a grid with mesh-size h_{ref} and time mesh \Delta t \in \{10^{-2}; 5.10^{-2}; 2, 5.10^{-3}; 1, 25.10^{-3}; 6, 25.10^{-4} \}.

    In Tables 3 and 4, we present the computed errors for the solute concentration in the fluid in energy and L^2 -norms and the solute concentration on the surface of the skeleton in L^2-norm on a sequence of uniformly refined meshes for space and time, as well as the convergence order computed from the errors on two successively refined meshes. The convergent rate in both time and space is assessed by means of mean-square approximation errors at the endpoint T = 0.2, produced by both temporal and spatial discretizations, and \lambda = 1/2. The anticipated amount of error is approximated by finding the mean of 500 samples for every efficiency test.

    Table 3.  Errors and convergence orders for SPDE respect to spatial discretization N.
    N \|u^{stoch}_{N}-u^{deter}_{N}\|_{2} order \|v^{stoch}_{N}-v^{deter}_{N}\|_{2} order
    64 0.763456768 - 0.057684321 -
    256 0.194523452 1.972602365 0.029578974 0.963607207
    1 024 0.048876453 1.992732602 0.014562340 1.022329832
    4 096 0.009976938 2.292470577 0.006845354 1.089045079
    16 384 0.002254494 2.145793759 0.003189323 1.101875000

     | Show Table
    DownLoad: CSV
    Table 4.  Errors and convergence orders for SPDE respect to time discretization ∆t.
    \Delta \|u^{stoch}_{N}-u^{deter}_{N}\|_{2} order \|v^{stoch}_{N}-v^{deter}_{N}\|_{2} order
    1\times 10^{-2} 0.005251298 - 0.004001222 -
    5\times 10^{-3} 0.004451682 0.548333579 0.002735681 0.548540665
    2.5\times 10^{-3} 0.003781534 0.527838102 0.001926523 0.505900602
    1.25\times 10^{-3} 0.003216459 0.533737177 0.001377659 0.483780575
    6.25\times 10^{-4} 0.002764550 0.531446991 0.000982452 0.487760006

     | Show Table
    DownLoad: CSV

    By analyzing the above results, we numerically retrieve the following optimal convergence orders with respect to spatial and temporal discretization cited in (6.7) and (6.8).

    In the present work, we have systematically developed a framework to solve coupled stochastic partial differential equations by introducing key assumptions in Section 2. We established a mild-weak probabilistic solution through Theorem 2.1 in Section 3, and further demonstrated path-wise uniqueness to confirm a strong probabilistic solution in Section 4. Our analysis in Section 5 confirmed the non-negativity of solute and reactant concentrations under specific data conditions. Finally, Section 6 showcased numerical results, demonstrating the efficacy of extended stochastic non-conforming finite element methods for discretizing the coupled SPDE and provided a numerical analysis of the convergence order of these methods, with examples illustrating smooth Langmuir and Freundlich kinetics.

    Future research could extend this work by addressing more complex and higher-dimensional systems of coupled stochastic partial differential equations, investigating diverse boundary conditions and noise types, and refining numerical methods for better accuracy and efficiency. Additionally, applying these models to real-world scenarios such as environmental, biological, and financial systems, and integrating machine learning for parameter estimation and predictive modeling, could enhance the practical impact and interdisciplinary relevance of the research.

    M. Eddahbi: Supervision; M. Mohammed: Writing-original draft; M. Eddahbi and M. Mohammed: Conceptualization; M. Eddahbi, M. Mohammed and H. El-Otmany: Formal analysis, Methodology, Validation, Writing-review & editing; M. Eddahbi and H. El-Otmany: Visualization. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The first named author extends his appreciation to the Researchers Supporting Project, Grant number (RSPD2024R1075), King Saud University, Riyadh, Saudi Arabia.

    The corresponding author extends the appreciation to the Deanship of Postgraduate Studies and Scientific Research at Majmaah University for funding this research work through the project number (R-2024-1232).

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.



    [1] Z. I. Ali, M. Sango, Probabilistic weak solutions for nonlinear stochastic evolution problems involving pseudomonotone operators, Ukr. Math. J., 74 (2022), 997–1020. https://doi.org/10.1007/s11253-022-02117-y doi: 10.1007/s11253-022-02117-y
    [2] G. Allaire, H. Hutridurga, Homogenization of reactive flows in porous media and competition between bulk and surface diffusion, IMA J. Appl. Math., 77 (2012), 788–815. https://doi.org/10.1093/imamat/hxs049 doi: 10.1093/imamat/hxs049
    [3] A. Bensoussan, Some existence results for stochastic partial differential equations, Pitman Res. Notes Math. Ser., 268 (1992), 37–53.
    [4] H. Bessaih, Y. Efendiev, R. F. Maris, Stochastic homogenization of a convection-diffusion equation, SIAM J. Math. Anal., 53 (2021), 2718–2745. https://doi.org/10.1137/19M130277 doi: 10.1137/19M130277
    [5] P. Billingsley, Convergence of probability measures, John Wiley & Sons, 2013.
    [6] B. Cabarrubias, P. Donato, Existence and uniqueness for a quasilinear elliptic problem with nonlinear Robin conditions, Carpathian J. Math., 27 (2011), 173–184.
    [7] I. Chourabi, P. Donato, Bounded solutions for a quasilinear singular problem with nonlinear Robin boundary conditions, Differ. Integral Equ., 26 (2013), 975–1008. https://doi.org/10.57262/die/1372858558 doi: 10.57262/die/1372858558
    [8] I. Chourabi, P. Donato, Homogenization of elliptic problems with quadratic growth and nonhomogenous Robin conditions in perforated domains, Chin. Ann. Math. Ser. B, 37 (2016), 833–852. https://doi.org/10.1007/s11401-016-1008-y doi: 10.1007/s11401-016-1008-y
    [9] P. L. Chow, Stochastic partial differential equations, CRC Press, 2014.
    [10] C. Conca, J. I. Díaz, A. Linan, C. Timofte, Homogenization in chemical reactive flows, Electron. J. Differ. Equ., 2004 (2004), 1–22.
    [11] C. Conca, J. I. Díaz, C. Timofte, Effective chemical processes in porous media, Math. Models Methods Appl. Sci., 13 (2003), 1437–1462. https://doi.org/10.1142/S0218202503002982 doi: 10.1142/S0218202503002982
    [12] L. Denis, A. Matoussi, L. Stoica, Maximum principle and comparison theorem for quasi-linear stochastic PDE's, Electron. Commun. Probab., 14 (2009), 500–530. https://doi.org/10.1214/EJP.v14-629 doi: 10.1214/EJP.v14-629
    [13] P. Donato, S. Monsurrò, F. Raimondi, Existence and uniqueness results for a class of singular elliptic problems in perforated domains, Ric. Mat., 66 (2017), 333–360. https://doi.org/10.1007/s11587-016-0303-y doi: 10.1007/s11587-016-0303-y
    [14] G. Ferreyra, P. Sundar, Comparison of solutions of stochastic equations and applications, Stoch. Anal. Appl., 18 (2000), 211–229. https://doi.org/10.1080/07362990008809665 doi: 10.1080/07362990008809665
    [15] U. Hornung, W. Jäer, Diffusion, convection, adsorption, and reaction of chemicals in porous media, J. Differ. Equ., 92 (1991), 199–225. https://doi.org/10.1016/0022-0396(91)90047-D doi: 10.1016/0022-0396(91)90047-D
    [16] M. Mohammed, Well-posedness for nonlinear parabolic stochastic differential equations with nonlinear Robin conditions, Symmetry, 14 (2022), 1–19. https://doi.org/10.3390/sym14081722 doi: 10.3390/sym14081722
    [17] M. Mohammed, Homogenization of a nonlinear stochastic model with nonlinear random forces for chemical reactive flows in porous media, Discrete Contin. Dyn. Syst. Ser. B, 28 (2023), 4598–4624. https://doi.org/10.3934/dcdsb.2023032 doi: 10.3934/dcdsb.2023032
    [18] M. Mohammed, N. Ahmed, Homogenization and correctors of Robin problem for linear stochastic equations in periodically perforated domains, Asymptot. Anal., 120 (2020), 123–149. https://doi.org/10.3233/ASY-191582 doi: 10.3233/ASY-191582
    [19] M. Mohammed, M. Sango, Homogenization of linear hyperbolic stochastic partial differential equation with rapidly oscillating coefficients: the two scale convergence method, Asymptot. Anal., 91 (2015), 341–371. https://doi.org/10.3233/ASY-141269 doi: 10.3233/ASY-141269
    [20] M. Mohammed, M. Sango, Homogenization of Neumann problem for hyperbolic stochastic partial differential equations in perforated domains, Asymptot. Anal., 97 (2016), 301–327. https://doi.org/10.3233/ASY-151355 doi: 10.3233/ASY-151355
    [21] M. Mohammed, M. Sango, A Tartar approach to periodic homogenization of linear hyperbolic stochastic partial differential equation, Int. J. Modern Phys. B, 30 (2016), 1640020. https://doi.org/10.1142/S0217979216400208 doi: 10.1142/S0217979216400208
    [22] M. Ondreját, Uniqueness for stochastic evolution equations in Banach spaces, Dissertationes Math., 426 (2004), 1–63.
    [23] B. L. Rozovsky, S. V. Lototsky, Stochastic evolution systems: linear theory and applications to non-linear filtering, Cham: Springer, 2018. https://doi.org/10.1007/978-3-319-94893-5
    [24] M. Sango, Splitting-up scheme for nonlinear stochastic hyperbolic equations, Forum Math., 25 (2013), 931–965. https://doi.org/10.1515/form.2011.138 doi: 10.1515/form.2011.138
    [25] M. Sango, Stochastic Navier-Stokes variational inequalities with unilateral boundary conditions: probabilistic weak solvability, Ukr. Math. J., 75 (2023), 600–620. https://doi.org/10.1007/s11253-023-02219-1 doi: 10.1007/s11253-023-02219-1
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(868) PDF downloads(56) Cited by(0)

Figures and Tables

Figures(4)  /  Tables(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog