We establish the existence of solutions for a class of stochastic reaction-diffusion systems with cross-diffusion terms modeling interspecific competition between two populations. More precisely, we prove the existence of weak martingale solutions employing appropriate Faedo-Galerkin approximations and the stochastic compactness method. The nonnegativity of solutions is proved by a stochastic adaptation of the well-known Stampacchia approach.
Citation: Mostafa Bendahmane, Kenneth H. Karlsen. Martingale solutions of stochastic nonlocal cross-diffusion systems[J]. Networks and Heterogeneous Media, 2022, 17(5): 719-752. doi: 10.3934/nhm.2022024
[1] | José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023 |
[2] | Elisabeth Logak, Isabelle Passat . An epidemic model with nonlocal diffusion on networks. Networks and Heterogeneous Media, 2016, 11(4): 693-719. doi: 10.3934/nhm.2016014 |
[3] | Hakima Bessaih, Yalchin Efendiev, Florin Maris . Homogenization of the evolution Stokes equation in a perforated domain with a stochastic Fourier boundary condition. Networks and Heterogeneous Media, 2015, 10(2): 343-367. doi: 10.3934/nhm.2015.10.343 |
[4] | Simone Göttlich, Stephan Martin, Thorsten Sickenberger . Time-continuous production networks with random breakdowns. Networks and Heterogeneous Media, 2011, 6(4): 695-714. doi: 10.3934/nhm.2011.6.695 |
[5] | Xia Li, Chuntian Wang, Hao Li, Andrea L. Bertozzi . A martingale formulation for stochastic compartmental susceptible-infected-recovered (SIR) models to analyze finite size effects in COVID-19 case studies. Networks and Heterogeneous Media, 2022, 17(3): 311-331. doi: 10.3934/nhm.2022009 |
[6] | Mostafa Bendahmane . Analysis of a reaction-diffusion system modeling predator-prey with prey-taxis. Networks and Heterogeneous Media, 2008, 3(4): 863-879. doi: 10.3934/nhm.2008.3.863 |
[7] | Piotr Gwiazda, Karolina Kropielnicka, Anna Marciniak-Czochra . The Escalator Boxcar Train method for a system of age-structured equations. Networks and Heterogeneous Media, 2016, 11(1): 123-143. doi: 10.3934/nhm.2016.11.123 |
[8] | Arnaud Ducrot, Michel Langlais, Pierre Magal . Multiple travelling waves for an $SI$-epidemic model. Networks and Heterogeneous Media, 2013, 8(1): 171-190. doi: 10.3934/nhm.2013.8.171 |
[9] | Pierre Degond, Sophie Hecht, Nicolas Vauchelet . Incompressible limit of a continuum model of tissue growth for two cell populations. Networks and Heterogeneous Media, 2020, 15(1): 57-85. doi: 10.3934/nhm.2020003 |
[10] | Panpan Xu, Yongbin Ge, Lin Zhang . High-order finite difference approximation of the Keller-Segel model with additional self- and cross-diffusion terms and a logistic source. Networks and Heterogeneous Media, 2023, 18(4): 1471-1492. doi: 10.3934/nhm.2023065 |
We establish the existence of solutions for a class of stochastic reaction-diffusion systems with cross-diffusion terms modeling interspecific competition between two populations. More precisely, we prove the existence of weak martingale solutions employing appropriate Faedo-Galerkin approximations and the stochastic compactness method. The nonnegativity of solutions is proved by a stochastic adaptation of the well-known Stampacchia approach.
This work is devoted to the mathematical analysis of a stochastic reaction-diffusion system with cross-diffusion modeling the interaction between two populations. Cross-diffusion expresses that the population flux of a given subpopulation is affected by the presence of other subpopulations. The (deterministic) dynamics of interacting species with cross-diffusion were investigated by many authors, including Levin [25], Levin and Segel [24], Okubo and Levin [31], Mimura and Murray [27], Mimura and Kawasaki [26], Mimura and Yamaguti [28], Galiano et al. [17,18], Bendahmane et al. [1,6] (see also [2,3,5,7]) to name a few. We consider a spatially distributed population wherein
U(t)=∫Ωu(t,x)dx,V(t)=∫Ωv(t,x)dx, |
whereas the total population is
P(t)=∫Ω(u+v)(t,x)dx=∫Ωp(t,x)dx, |
where
A prototype of stochastic reaction-diffusion systems with nonlocal diffusion and cross-diffusion terms is
du−∇⋅(Du(∫Ωu(t,x)dx)∇u+A11(u,v)∇u+A12(u,v)∇v)dt=F(u,v)dt+σu(u)dWu(t),dv−∇⋅(Dv(∫Ωv(t,x)dx)∇v+A21(u,v)∇u+A22(u,v)∇v)dt=G(u,v)dt+σv(v)dWv(t), | (1) |
which is posed in the time-space cylinder
u(0,x)=u0(x)≥0,v(0,x)=v0(x)≥0,x∈Ω, | (2) |
and zero-flux boundary conditions on
(Du(∫Ωu(t,x)dx)∇u+A11(u,v)∇u+A12(u,v)∇v)⋅ν=0,(Dv(∫Ωv(t,x)dx)∇v+A21(u,v)∇u+A22(u,v)∇v)⋅ν=0. | (3) |
In the system (1),
In (1),
F(u,v):=−θ(u,v)−μuG(u,v):=θ(u,v)−γv−μv | (4) |
are the reaction terms. In the dispersal of an epidemic disease, the constants
θ(u,v)=αuvu+v,u,v≥0. | (5) |
For later use, note that
0≤θ(u,v)≤αmin(u,v),u,v≥0. | (6) |
The diffusion rates (given by
Dw(I)≥Cm,|Dw(I1)−Dw(I2)|≤CM|I1−I2|,∀I,I1,I2∈R. | (7) |
In (1),
A(u,v)(∇u∇v)=(A11(u,v)∇u+A12(u,v)∇vA21(u,v)∇u+A22(u,v)∇v). |
We assume that the matrix
∀u,v≥0,A12(0,v)=0,A21(u,0)=0,∀u,v≥0,∀w:=(w1w2)∈R2d,(A(u,v)w,w)≥1C|A(u,v)||w|2,∀u1,u2,v1,v2≥0,|A(u1,v1)−A(u2,v2)|≤C(|u1−u2|+|v1−v2|), | (8) |
where
A11(u,v)≥0,A22(u,v)≥0,∀u,v≥0. |
A typical example of a cross-diffusion matrix is
A(u,v)=(a11u+a12va13ua21va22u+a23v), |
where the coefficients
Remark 1. For the upcoming analysis we need to extend the definitions of
if u,v≥0, then Aij(u,v)≥0, otherwise Aij(u,v)=0(i≠j) and Aii(u,v)≥0,F(u,v)={−θ(u,v)−μu,if u,v≥0,−μu,if u≥0 and v<0,0,if u<0 and v≥0,G(u,v)={θ(u,v)−γv−μv,if u,v≥0,0,if u≥0 and v<0,−γv−μv,if u<0 and v≥0. |
Our analysis is restricted to positive cross-diffusion matrices
Historically, cross-diffusion models are deterministic, meaning that the input data determine the solution at each moment in time. In deterministic models, non-predictable environmental factors are not considered, although it is well-known that a combination of random perturbations and nonlinearities can strongly influence solutions. Multiple factors may influence the population's growth in the environment, such as food, water, temperature, etc., each element easily being thought of as stochastic. It is natural to employ noise to model these environmental fluctuations by adding a stochastic forcing term to the deterministic system, resulting in (1).
Let us now put the mathematical contributions of this paper into perspective. First, note that the standard theory for parabolic systems does not apply naturally to the cross-diffusion model because of the strong coupling in the highest derivatives. As a result, no traditional maximum principle applies. A stochastic forcing term further complicates the maximum principle approach. The existence result for (1) is based on martingale solutions and the introduction of suitable approximate (Faedo-Galerkin) solutions. We derive a series of system-specific a priori estimates in
In [14], the authors prove the existence of solutions for a related stochastic cross-diffusion system (with
For the existence of martingale solutions for other classes of SPDEs, we refer [8,9,11,12,15,16,20,21,22,34,35], to mention a few inspirational examples.
The paper is organized as follows: In Section 2, we present the stochastic framework and state the noise coefficients' hypotheses. Section 3 supplies the definition of a weak martingale solution and declares the main result. We construct approximate solutions by the Faedo-Galerkin method in Section 4. Uniform estimates for these approximations are established in Sections 5 and 6. Section 7 proves the tightness of the probability laws generated by the Faedo-Galerkin approximations. The tightness and Skorokhod's representation theorem is used to show that a weakly convergent sequence of the probability laws has a limit that can be represented as the law of an almost surely convergent sequence of random variables defined on a common probability space. The limit of this sequence is proved to be a weak martingale solution of the stochastic reaction-diffusion system in Section 8, while its nonnegativity is deferred to Section 9.
Throughout this paper, we will frequently use the letters
This section recalls basic concepts and results from stochastic analysis (see e.g. [11,33] for more details). We consider a complete probability space
In passing, note that the letter
Given a separable Banach space
‖X‖Lp(D;B)=‖X‖Lp(D,F,P;B):=(E[‖X‖pB])1p(p<∞),‖X‖Lp(D;B)=‖X‖L∞(D,F,P;B):=supω∈D‖X(ω)‖B. |
We use the abbreviation a.s. (or almost surely) for "
We refer to
S=(D,F,{Ft}t∈[0,T],P,{Wk}∞k=1) | (9) |
as a (Brownian) stochastic basis, where
A stochastic process
Consider a Hilbert space
For a given cylindrical Wiener process
∫t0σwdWw=∞∑k=1∫t0σw,kdWw,k,σw,k:=σwψk, | (10) |
for any
σ∈L2(D,F,P;L2(0,T;L2(U,L2(Ω)))). |
Throughout the paper, we assume several conditions on the noise coefficients
σw(z)ψk=σw,k(z(⋅)),k≥1, |
for some real-valued functions
∑k≥1|σw,k(z)|2≤Cσ(1+|z|2),∀z∈R,∑k≥1|σw,k(z1)−σw,k(z2)|2≤Cσ|z1−z2|2,∀z1,z2∈R, | (11) |
for a constant
‖σw(z)‖2L2(U,L2(Ω))≤Cσ(1+‖z‖2L2(Ω)),z∈L2(Ω),‖σw(z1)−σw(z2)‖2L2(U,L2(Ω))≤Cσ‖z1−z2‖2L2(Ω),z1,z2∈L2(Ω). | (12) |
Under these conditions (12), the stochastic integral (10) is an
E[supt∈[0,T]‖∫t0σwdWw‖pL2(Ω)]≤CE[(∫T0‖σw‖2L2(U,L2(Ω))dt)p2], | (13) |
where
Lemma 2.1 (convergence of stochastic integrals). For each
Wnn↑∞⟶Win C([0,T];U0), in probabilityGnn↑∞⟶Gin L2(0,T;L2(U;L2(Ω))), in probability. |
Then
∫t0GndWnn↑∞⟶∫t0GdWin L2(0,T;L2(Ω)), in probability. |
Let
Any random variable
We will utilize the following notion of solution for the stochastic cross-diffusion system.
Definition 3.1 (weak martingale solution). Let
1.
2.
3. The elements
L2(D,F,P;L2(0,T;H1(Ω)))⋂L2(D,F,P;L∞(0,T;L2(Ω))), |
and satisfy
√|Aij(u,v)|∇u∈L2(D,F,P;L2(0,T;L2(Ω))),i,j=1,2. |
Finally,
4. The laws of
5. The following equations hold
∫Ωu(t)φudx−∫Ωu0φudx+∫t0∫Ω(Du(∫Ωu(t,x)dx)∇u+A11(u,v)∇u+A12(u,v)∇v)⋅∇φudxds=∫t0∫ΩF(u,v)φudxds+∫t0∫Ωσu(u)φudxdWu(s),∫Ωv(t)φvdx−∫Ωv0φvdx+∫t0∫Ω(Dv(∫Ωv(t,x)dx)∇v+A21(u,v)∇u+A22(u,v)∇v)⋅∇φvdxds=∫t0∫ΩG(u,v)φvdxds+∫t0∫Ωσv(v)φvdxdWv(s), | (14) |
for all
Remark 2. In Definition 3.1, we use the standard Sobolev spaces
H1(Ω)=W1,2(Ω),and for p∈(1,∞),W1,p(Ω)={u∈Lp(Ω):∇u∈Lp(Ω;Rd)}, |
along with the corresponding dual spaces
‖L‖(W1,p(Ω))⋆=sup{|⟨L,ϕ⟩|:ϕ∈W1,p(Ω),‖ϕ‖W1,p(Ω)≤1}, |
where
Recall that
⟨L,ϕ⟩=∫Ωf0ϕ+d∑i=1fi∂xiϕdx,∀ϕ∈W1,p(Ω), |
and
Remark 3. 1. Given the regularity conditions imposed in Definition 3.1, one can show that the deterministic and the stochastic integrals in (14) are all well-defined. Regarding the stochastic terms
2. For martingale solutions, one prescribes the initial data in terms of probability measures
3. Part (3) of Definition 3.1 implies that
Remark 4. A significant difficulty for the analysis of (1) is the strong coupling in the highest derivatives. However, since these terms are zero on the boundary, cf. (3), the nonlinear boundary conditions will "disappear" in the weak martingale formulation.
Our main result is
Theorem 3.2 (existence). Suppose conditions (4), (5), (6), (8), (7), and (11) hold, and thatthe initial data
∫L2(Ω)‖w‖q0L2(Ω)dμw0(w)<∞,for some q0>3,w:=u,v. | (15) |
Then the stochastic cross-diffusionsystem (1), with initial-boundary data (2) and (3), possesses a weak martingale solution in the sense of Definition 3.1. Moreover, assuming
The proof of Theorem 3.2 is organized into several sections. First, in Section 4, we construct the Faedo-Galerkin solutions. Energy-type estimates are derived in Section 5. Convergence of the approximate solutions (along a subsequence) to a limit follows from these estimates, a temporal translation estimate, cf. 6, and the tightness of the probability laws generated by the Faedo-Galerkin solutions, cf. Section 7. In Section 8, we show that the limit is a weak martingale solution. Finally, we prove the nonnegativity of the constructed martingale solution, cf. Section 9.
In this section, we define precisely the Faedo-Galerkin equations and prove that there exists a solution to these equations. We begin by fixing a stochastic basis
Let us make precise the basis functions
L20:={u∈L2(Ω):¯u:=1|Ω|∫Ωudx=0},H2N:={u∈H2(Ω):∂u∂ν=0on ∂Ω},(H1)⋆0:={u∈(H1(Ω))⋆:¯u:=1|Ω|⟨u,1⟩(H1)⋆,H1=0}. |
The embeddings
The Neumann-Laplace operator
⟨−ΔNu,v⟩(H1)⋆,H1=∫Ω∇u⋅∇vdx,u,v∈H1(Ω). |
The Neumann-Laplace operator is positive and self-adjoint. By the Lax-Milgram theorem and the Poincaré inequality, the inverse operator
Πn:L2(Ω)→Xn=Span{e1,…,en},Πnu:=n∑l=1(u,el)el. | (16) |
Then
Denoting the corresponding eigenvalues by
−Δel=λlelin Ω,∂el∂ν=0on ∂Ω, | (17) |
for each
ΔΠnu=n∑l=1(u,el)Δel=n∑l=1(u,−λel)el=n∑l=1(u,Δel)el=n∑l=1(Δu,el)el=ΠnΔu. |
As a result,
‖Πnu‖H2N≤C‖u‖H2N, |
for a constant
From the weak form of (17) with test function
∫Ω∇el⋅∇emdx=λl∫Ωelemdx=λlδlm,∀l,m∈N; |
thus
0=∫Ω∇u⋅∇˜eldx+∫Ωu˜eldx=(1+λl)12∫Ωueldx, |
so that
Let us note that the restriction of
˜Πnu=n∑l=1(u,˜el)H1(Ω)˜el=n∑l=1(1+λl)12(u,el)˜el=n∑l=1(u,el)el=Πnu. |
Consequently,
Πnun↑∞⟶uin H1(Ω),‖Πnu‖H1(Ω)≤‖u‖H1(Ω). |
Finally, we will continue to use the symbol
Πn:X⋆→Span{e1,…,en},Πnu:=n∑l=1⟨u,el⟩X⋆,Xel, |
where
(Πnu,v)=⟨u,Πnv⟩X⋆,X,u∈X⋆,v∈X, |
as
We can now define our Faedo-Galerkin approximations
un,vn:[0,T]→Xn,un(t)=n∑l=1cnl(t)el,vn(t)=n∑l=1dnl(t)el, | (18) |
where the coefficients
(dun,el)+Du(∫Ωun(t,x)dx)(∇un,∇el)dt+(A11(un,vn)∇un+A12(un,vn)∇vn,∇el)dt=(F(un,vn),el)dt+n∑k=1(σnu,k(un),el)dWu,k(t),(dvn,el)+Dv(∫Ωvn(t,x)dx)(∇vn,∇el)dt+(A21(un,vn)∇un+A22(un,vn)∇vn,∇el)dt=(G(un,vn),el)dt+n∑k=1(σnv,k(vn),el)dWv,k(t), | (19) |
and, with reference to the initial data,
un(0)=un0:=n∑l=1cnl(0)el,cnl(0):=(un0,el)L2(Ω),vn(0)=vn0:=n∑l=1dnl(0)el,dnl(0):=(v0,el)L2(Ω). | (20) |
In (19) we have used the following approximations of the noise coefficients:
σnw,k(wn):=n∑l=1σw,k,l(wn)el,whereσw,k,l(wn):=(σw,k(wn),el)L2(Ω),w=u,v. | (21) |
Using the Faedo-Galerkin equations (19), the regularity
un(t)−un0−∫t0Πn[∇⋅(Du(∫Ωun(t,x)dx)∇un)]ds−∫t0Πn[∇⋅(A11(un,vn)∇un+A12(un,vn)∇vn)]ds=∫t0Πn[F(un,vn)]ds+∫t0σnu(un)dWnu(s)in L2(Ω),vn(t)−vn0−∫t0Πn[∇⋅(Dv(∫Ωvn(t,x)dx)∇vn)]ds−∫t0Πn[∇⋅(A21(un,vn)∇un+A22(un,vn)∇vn)]ds=∫t0Πn[G(un,vn)]ds+∫t0σnv(vn)dWnv(s)in L2(Ω), | (22) |
with initial data
Remark 5. Our construction of approximate solutions makes use of Neumann boundary conditions, which are encoded in the space
The existence of pathwise solutions to the finite-dimensional problem (19), (20) is guaranteed by the next lemma.
Lemma 4.1. For each
Proof. We look for a stochastic process
dCn=M(Cn)dt+Γ(Cn)dWn, | (23) |
where
Au(Cn)=−Πn∇⋅(Du(∫Ωun(t,x)dx)∇un)−Πn∇⋅(A11(un,vn)∇un+A12(un,vn)∇vn)+ΠnF(un,vn),Av(Cn)=−Πn∇⋅(Dv(∫Ωvn(t,x)dx)∇vn)−Πn∇⋅(A21(un,vn)∇un+A22(un,vn)∇vn)+ΠnG(un,vn). |
Moreover,
To prove the existence and uniqueness of a pathwise solution to (23), we will use [33,Theorem 3.1.1] (see also Theorem 5.1.3 in [33]), which asks that
(ⅰ) — local weak monotonicity. For all
2(M(C1)−M(C2),C1−C2)+‖Γ(C1)−Γ(C2)‖2L2(Ω)≤K(r)‖C1−C2‖2L2(Ω), | (24) |
for a constant
(ⅱ) — weak coercivity. For all
2(M(C),C)+‖Γ(C)‖2L2(Ω)≤K(1+‖C‖2L2(Ω)), | (25) |
for some constant
The weak coercivity condition (25) is easily verified using the assumption (8) and the global Lipschitz continuity of
Let us verify the weak monotonicity condition (24) in some detail. Fix a real number
(M(C1)−M(C2),C1−C2)+‖Γ(C1)−Γ(C2)‖2L2(Ω)=6∑i=0Ii, | (26) |
where
I1=−∑w=u,vDw(∫Ωw1dx)(∇¯w,∇¯w),I2=−∑w=u,v(Dw(∫Ωw1dx)−Dw(∫Ωw2dx))(∇w2,∇¯w),I3=−(A(u1,v1)(∇¯u∇¯v),(∇¯u∇¯v)),I4=−((A(u1,v1)−A(u2,v2))(∇u2∇v2),(∇¯u∇¯v)),I5=(F(u1,v1)−F(u2,v2),¯u),I6=(G(u1,v1)−G(u2,v2),¯v). |
Recall that the basis functions
|I2|≲∑w=u,v‖w1−w2‖L1(Ω)‖∇w2‖L2(Ω)‖∇¯w‖L2(Ω), |
and so
|I4|≲∑w=u,v‖w1−w2‖L2(Ω)∑w=u,v‖∇w2‖L4(Ω)∑wn=u,v‖∇¯w‖L4(Ω)≲∑w=u,v‖w1−w2‖L2(Ω)∑w=u,v‖∇wn2‖H2N∑w=u,v‖∇¯w‖H2N, |
and so
|I5|+|I6|≲∑w=u,v‖w1−w2‖L2(Ω)∑w=u,v‖wn‖L2(Ω), |
so that
Referring to (26), this implies
We start with a series of basic energy-type estimates.
Lemma 5.1. Let
E[‖un(t)‖2L2(Ω)]+E[‖vn(t)‖2L2(Ω)]≤C,∀t∈[0,T]; | (27) |
E[∫T0∫Ω|∇un|2dxdt]+E[∫T0∫Ω|∇vn|2dxdt]≤C; | (28) |
E[∫T0∫Ω|Aij(un,vn)|(|∇un|2+|∇vn|2)dxdt]≤C,i,j=1,2; | (29) |
E[supt∈[0,T]‖un(t)‖2L2(Ω)]+E[supt∈[0,T]‖vn(t)‖2L2(Ω)]≤C. | (30) |
Proof. By Itô's formula,
12∑w=u,v‖wn(t)‖2L2(Ω)+∑w=u,v∫t0Dw(∫Ωwn(t,x)dx)∫Ω|∇wn|2dxds+∫t0(A11(un,vn)∇un+A12(un,vn)∇vn,∇un)L2(Ω)ds+∫t0(A21(un,vn)∇un+A22(un,vn)∇vn,∇vn)L2(Ω)ds=12∑w=u,v‖wn(0)‖2L2(Ω)+∫t0(F(un,vn),un)L2(Ω)ds+∫t0(G(un,vn),vn)L2(Ω)ds+∑w=u,vn∑k=1∫t0∫Ωwnσnw,k(wn)dxdWw,k+12∑w=u,vn∑k=1∫t0∫Ω(σnw,k(wn))2dxds≤12∑w=u,v‖wn(0)‖2L2(Ω)+C∫t0(1+‖un(t)‖2L2(Ω)+‖vn(t)‖2L2(Ω))ds+∑w=u,vn∑k=1∫t0∫Ωwnσnw,k(wn)dxdWw,k(s), | (31) |
where have put to good use (4), (5), and also (6). By the fundamental assumption (8), the sum of the
∑w=u,v‖wn(t)‖2L2(Ω)+∑w=u,vCm∫t0∫Ω|∇wn|2dxds+∑w=u,v∫t0∫Ω|A(un,vn)||∇wn|2dxds≤∑w=u,v‖wn(0)‖2L2(Ω)+C∫t0(1+∑w=u,v‖wn(t)‖2L2(Ω))ds+∑w=u,vn∑k=1∫t0∫Ωwnσnw,k(wn)dxdWw,k(s). | (32) |
where we have also used (7). Applying
To prove the final estimate (30), we take
∑w=u,vE[supt∈[0,T]‖wn(t)‖2L2(Ω)]≤C(1+∑w=u,vIw), | (33) |
where
|Iw|≤CE[(∫T0n∑k=1|∫Ωwnσnw,k(wn)dx|2dt)12]≤CE[(∫T0(∫Ω|wn|2dx)(n∑k=1∫Ω|σnw,k(wn)|2dx)dt)12]≤CE[(supt∈[0,T]∫Ω|wn|2dx)12(∫T0n∑k=1∫Ω|σnw,k(wn)|2dxdt)12]≤αE[supt∈[0,T]∫Ω|wn|2dx]+C(α)E[∫T0n∑k=1∫Ω|σnw,k(wn)|2dxdt]≤αE[supt∈[0,T]‖wn(t)‖2L2(Ω)]+C, | (34) |
for any number
Later we will need to convert a.s. convergence into
Corollary 1. Let
E[sup0≤t≤T‖wn(t)‖qL2(Ω)]≤C,E[‖∇wn‖qL2((0,T)×Ω)]≤C,w=u,v, | (35) |
and
E[|∫T0∫Ω|Aij(un,vn)|(|∇un|2+|∇vn|2)dxdt|q2]≤C,i,j=1,2. | (36) |
Proof. Starting off from (31), the following estimate holds for any
∑w=u,vsup0≤τ≤t‖wn(τ)‖2L2(Ω)≤∑w=u,v‖wn(0)‖2L2(Ω)+C∑w=u,v∫t0‖wn(s)‖2L2(Ω)ds+C∑w=u,vsup0≤τ≤t|n∑k=1∫τ0∫Ωwnσnw,k(wn)dxdWw,k(s)|, |
for some constant
∑w=u,vE[sup0≤τ≤t‖un(τ)‖qL2(Ω)]≤C∑w=u,vE[‖wn(0)‖qL2(Ω)]+C(1+t)q2+C∑w=u,v∫t0‖wn(s)‖qL2(Ω)ds+∑w=u,vIw, | (37) |
where
Iw≤CE[(∫t0n∑k=1|∫Ωwnσnw,k(wn)dx|2ds)q4]≤CE[(∫t0(∫Ω|wn|2dx)(n∑k=1∫Ω|σnw,k(wn)|2dx)ds)q4]≤CE[(supτ∈[0,t]∫Ω|wn|2dx)q4(∫t0n∑k=1∫Ω|σnk,w(wn)|2dxds)q4]≤αE[(supτ∈[0,t]∫Ω|wn|2dx)q2]+C(α)E[(∫t0n∑k=1∫Ω|σnk,u(wn)|2dxds)q2]≤αE[supτ∈[0,t]‖wn(τ)‖qL2(Ω)]+CE[∫t0‖wn(s)‖qL2(Ω)ds]+C, | (38) |
for any number
∑w=u,vE[sup0≤τ≤t‖wn(τ)‖qL2(Ω)]≤C∑w=u,vE[‖wn(0)‖qL2(Ω)]+C∑w=u,v∫t0E[‖wn(s)‖qL2(Ω)ds]+C, |
for some constant
Finally, we use (32), the first part of (38), and (35) to conclude that there is a constant
∑w=u,vE[|∫t0∫Ω|∇wn|2dxds|q2]≤C,w=u,v, |
and the second part of (35) follows. Similarly, we derive (36).
Given Lemma 5.1, it is easy to see that
Lemma 6.1. Extend the Faedo-Galerkin functions
E[sup|τ|∈(0,δ)‖wn(t+τ)−wn(t)‖(H2N)⋆]≤Cδ1/4,∀t∈[0,T], | (39) |
for any sufficiently small
Proof. In what follows, we write
I(t,τ):=‖un(t+τ,⋅)−un(t,⋅)‖(H2N)⋆=sup{|⟨un(t+τ,⋅)−un(t,⋅),ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}=sup{∫Ω(un(t+τ,x)−un(t,x))ϕ(x)dx:ϕ∈H2N,‖ϕ‖H2N≤1}, |
for
By (18),
I(t,τ):=‖un(t+τ,⋅)−un(t,⋅)‖(H2N)⋆≤4∑i=1Ii(t,τ), |
where
I1(t,τ)=‖∫t+τtΠn[∇⋅(Du(∫Ωun(t,x)dx)∇un)]ds‖(H2N)⋆,I2(t,τ)=‖∫t+τtΠn[∇⋅(A11(un,vn)∇un+A12(un,vn)∇vn)]ds‖(H2N)⋆,I3(t,τ)=‖∫t+τtΠn[F(un,vn)]ds‖(H2N)⋆,I4(t,τ)=‖n∑k=1∫t+τtσnu,k(un)dWu,k(s)‖(H2N)⋆. |
‖∫t+τtLn2,uds‖(H2N)⋆=sup{|⟨∫t+τtLn2,uds,ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}=sup{|∫t+τt∫ΩLn2,uϕdxds|:ϕ∈H2N,‖ϕ‖H2N≤1}=sup{|∫t+τt∫ΩA11(un,vn)∇un⋅∇Πnϕdxds|:ϕ∈H2N,‖ϕ‖H2N≤1} |
by bounding the term
I:=|∫t+τt∫ΩA11(un,vn)∇un⋅∇Πnϕdxds|. |
By the generalised Hölder inequality,
I≤τ1/4‖√|A11(un,vn)|‖L4((0,T)×Ω)×‖√|A11(un,vn)||∇un|‖L2((0,T)×Ω)‖∇Πnϕ‖L4(Ω). |
Now we use that
‖∇Πnϕ‖L4(Ω)≤‖Πnϕ‖W1,4(Ω)≲‖Πnϕ‖H2N. |
As
I≲τ1/4(‖√|A11(un,vn)|‖2L4((0,T)×Ω)+‖√|A11(un,vn)||∇un|‖2L2((0,T)×Ω))‖ϕ‖H2N, |
Note that
‖√|A11(un,vn)|‖2L4((0,T)×Ω)=‖A11(un,vn)‖L2((0,T)×Ω)≲‖1+un+vn‖L2((0,T)×Ω)≲T,Ω1+‖un‖L∞(0,T;L2(Ω))+‖vn‖L∞(0,T;L2(Ω)). |
Consequently, after taking the expectation and using (29) and (30),
E[I]≲T,Ωτ1/4‖ϕ‖H2N. |
Summarising,
E[supτ∈(0,δ)sup{|⟨∫t+τtLn2,uds,ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}]≲δ14, |
i.e.,
E[supτ∈(0,δ)‖∫t+τtLnu,2ds‖(H2N)⋆]≲δ14. |
A similar estimate holds for
E[sup0≤τ≤δI2(t,τ)]≲δ1/4,uniformly in t∈[0,T]. |
|Du(∫Ωun(t,x)dx)|2≲1+(∫Ω|un(t,x)|dx)2≲Ω1+‖un‖2L∞(0,T;L2(Ω)). |
Using this, we bound
|⟨∫t+τtLn1ds,ϕ⟩|=|∫t+τt∫ΩDu(∫Ωun(t,x)dx)∇un⋅∇Πnϕdxds| |
by a constant times
τ1/2(∫T0∫Ω(1+‖un‖2L∞(0,T;L2(Ω)))|∇un|2dxds)12‖∇Πnϕ‖L2(Ω)≲τ1/2(1+‖un‖L∞(0,T;L2(Ω)))‖∇un‖L2((0,T)×Ω)‖Πnϕ‖H1(Ω). |
Recalling that the sequence
‖Πnϕ‖H1(Ω)≤‖ϕ‖H1(Ω≤‖ϕ‖H2N. |
Taking the expectation and using Young's inequality,
E[(1+‖un‖L∞(0,T;L2(Ω)))‖∇un‖L2(Ω)]≲1+E[‖un‖2L∞(0,T;L2(Ω))]+E[‖∇un‖2L∞(0,T;L2(Ω))](28),(30)≲1, |
and thus we conclude that
E[supτ∈(0,δ)sup{|⟨∫t+τtLn1ds,ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}]≲δ1/2, |
i.e.,
E[supτ∈(0,δ)I1(t,τ)]≲δ1/2,uniformly in t∈[0,T]. |
|⟨∫t+τtLn3ds,ϕ⟩|=|∫t+τt∫ΩF(un,vn)Πnϕdxds| |
by a constant times
τ1/2‖1+un+vn‖L2((0,T)×Ω)‖Πnϕ‖L2(Ω)≲τ1/2(1+‖un‖2L2((0,T)×Ω)+‖un‖2L2((0,T)×Ω))‖ϕ‖H2N, |
where we have used Young's inequality and that the sequence
E[supτ∈(0,δ)sup{|⟨∫t+τtLn3ds,ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}]≲δ1/2, |
i.e.,
E[supτ∈(0,δ)I3(t,τ)]≲δ1/2,uniformly in t∈[0,T]. |
|⟨∫t+τtLn3ds,ϕ⟩|=|∫Ωn∑k=1∫t+τtσnu,k(un)dWu,k(s)ϕdx| |
by a constant times
‖n∑k=1∫t+τtσnu,k(un)dWu,k(s)‖L2(Ω)‖ϕ‖L2(Ω) |
where
E[supτ∈(0,δ)‖n∑k=1∫t+τtσnu,k(un)dWu,k(s)‖L2(Ω)]≲E[n∑k=1∫t+δt∫Ω(σnu,k(un))2dxds]12(11)≲Ωδ1/2(1+E[‖un‖L∞(0,T;L2(Ω))]), |
where
E[supτ∈(0,δ)sup{|⟨∫t+τtLn4ds,ϕ⟩|:ϕ∈H2N,‖ϕ‖H2N≤1}]≲δ1/2, |
i.e.,
E[supτ∈(0,δ)I4(t,τ)]≲δ1/2,uniformly in t∈[0,T]. |
Summarising our estimates of
In this section we establish the tightness of the probability measures (laws) generated by the Faedo-Galerkin solutions
We choose the following phase space for the probability laws of the Faedo-Galerkin approximations:
H:=Hu×Hv×HWu×HWv×Hu0×Hv0, |
where
Hu,Hv=L2(0,T;L2(Ω))⋂C(0,T;(H1(Ω))⋆) |
and (
HWu,HWv=C([0,T];U0),Hu0=Hv0=L2(Ω). |
As
Ψn:(D,F,P)→(H,B(H)),Ψn(ω)=(un(ω),vn(ω),Wnu(ω),Wnv(ω),un0(ω),vn0(ω)). |
We define a probability measure
Ln(A)=(P∘Ψ−1)(A)=P(Ψ−1n(A)),A∈B(H). | (40) |
Denote by
Ln=Lun×Lvn×LWnu×LWnv×Lun0×Lvn0. |
Remark 6. As a cartesian product of topological spaces,
Given sequences
Zrm,νm:={z∈L∞(0,T;L2(Ω))∩L2(0,T;H1(Ω)):supm≥11νmsupτ∈(0,rm)‖z(⋅+τ)−z‖L∞(0,T−τ;(H2N)⋆)<∞}. |
It is easy to see that
‖z‖Zrm,νm:=‖z‖L∞(0,T;L2(Ω))+‖w‖L2(0,T;H1(Ω))+supm≥11νmsup0≤τ≤rm‖z(⋅+τ)−z‖L∞(0,T−τ;(H2N)⋆). |
In view of [36], we have
Zrm,νm⊂⊂L2(0,T;L2(Ω))∩C([0,T];(H1(Ω))⋆), |
where
Now we verify that the laws
Lemma 7.1. The sequence
Proof. For each
C1,δ⊂L2(0,T;L2(Ω))⋂C(0,T;(H1(Ω))⋆),andC2,δ⊂C([0,T];U0),C3,δ⊂L2(Ω), |
such that
To this end, pick the sequences
∞∑m=1r1/4mνm<∞, | (41) |
and take
C1,δ:={z∈Zrm,νm:‖z‖Zrm,νm≤R1,δ}, |
where
P({ω∈D:wn(ω)∉C1,δ})≤P({ω∈D:‖wn(ω)‖L∞(0,T;L2(Ω))>R1,δ})+P({ω∈D:‖wn(ω)‖L2(0,T;H1(Ω))>R1,δ})+P({ω∈D:supτ∈(0,rm)‖wn(⋅+τ)−wn‖L∞(0,T−τ;(H2N)⋆)>R1,δνm})=:P1,1+P1,2+P1,3(for any m≥1). |
Repeated applications of the Chebyshev inequality supply
P1,1≤1R1,δE[‖wn(ω)‖L∞(0,T;L2(Ω))]≤CR1,δ,P1,2≤1R1,δE[‖wn(ω)‖L2(0,T;H1(Ω))]≤CR1,δ,P1,3≤∞∑m=11R1,δνmE[sup0≤τ≤rm‖wn(⋅+τ)−wn‖L∞(0,T−τ;(H2N)⋆)]≤CR1,δ∞∑m=1r1/4mνm(41)≤CR1,δ, |
where we have used (28), (30), and (39). From this, we can choose
Lwn(Cc1,δ)=P({ω∈D:wn(ω)∉C1,δ})≤δ6,w=u,v. |
Regarding the finite-dimensional approximations of the Wiener processes, we know that the finite series
LWnw(Cc2,δ)=P({ω∈D:Wnw(ω)∉C2,δ})≤δ6,w=u,v. |
Similarly, the initial data approximations
Lwn0(C3,δ)=P({ω∈D:wn0(ω)∉C3,δ})≤δ6,w=u,v. |
Summarising,
As the probability measures
˜Ψn=(˜un,˜vn,˜Wnu,˜Wnv,˜un0,˜vn0),˜Ψ=(˜u,˜v,˜Wu,˜Wv,˜u0,˜v0), | (42) |
with respective joint laws
˜un→˜u,˜vn→˜vin L2(0,T;L2(Ω)),˜un→˜u,˜vn→˜vin C([0,T];(H1(Ω))⋆),˜Wnu→˜Wu,˜Wnv→˜Wvin C([0,T];U0),˜un0→˜u0,˜vn0→˜v0in L2(Ω). | (43) |
By equality of the laws, the estimates in Lemma 5.1 and Corollary 1 continue to hold for the new random variables
Lemma 7.2. Let
˜E[‖˜wn‖qL∞(0,T;L2(Ω)]≤C,˜E[‖∇˜wn‖qL2((0,T)×Ω)]≤C,w=u,v, | (44) |
for any
Proof. We prove the first estimate in (44), as the other ones can be proved in the same way. Let
Recalling (42), consider the associated stochastic basis
˜Sn=(˜D,˜F,{˜Fnt}t∈[0,T],˜P,˜Wnu,˜Wnv), | (45) |
where
˜Ftn=σ(σ(˜Ψn|[0,t])⋃{N∈˜F:˜P(N)=0}). |
The filtration
A cylindrical Wiener process is fully determined by its law. By equality of the laws and Lévy's martingale characterization of a Wiener process, see [11,Theorem 4.6], we conclude that
Hence, there exist sequences
˜Wnw=∑k≥1˜Wnw,kψk,for w=u,v, | (46) |
recalling that
In what follows, we will use the following
˜W(n)w=n∑k=1˜Wnw,kψk,w=u,v, |
which converges to
Arguing as in [8], using (22) and equality of the laws, the following equations hold
˜un(t)−∫t0Πn[∇⋅(Du(∫Ω˜un(t,x)dx)∇˜un)]ds−∫t0Πn[∇⋅(A11(˜un,˜vn)∇˜un+A12(˜un,˜vn)∇˜vn)]ds=˜un0+∫t0Πn[F(˜un,˜vn)]ds+∫t0σnu(˜un)d˜W(n)u(s)in L2(Ω),˜vn(t)−∫t0Πn[∇⋅(Dv(∫Ω˜vn(t,x)dx)∇˜vn)]ds−∫t0Πn[∇⋅(A21(˜un,˜vn)∇˜un+A22(˜u,˜v)∇˜vn)]ds=˜vn0+∫t0Πn[G(˜un,˜vn)]ds+∫t0σnv(˜vn)d˜W(n)v(s)in L2(Ω), | (47) |
for any
In(ω,t,x):=un(t)−un0−∫t0Πn[∇⋅(Du(∫Ωun(t,x)dx)∇un)]ds−∫t0Πn[∇⋅(A11(un,vn)∇un+A12(un,vn)∇vn)]ds−∫t0Πn[F(un,vn)]ds−∫t0σnu(un)dWnu(s). |
Replacing
In(ω)=‖In(ω,⋅,⋅)‖2L2(0,T;L2(Ω)),˜In(ω)=‖˜In(ω,⋅,⋅)‖2L2(0,T;L2(Ω)). |
By(22),
˜E[˜Iδn1+˜Iδn]=∫XL(Ψ)dLn(Ψ)=E[Iδn1+Iδn]. |
Sending
The next estimate was not stated in Lemma 7.2, but it can be derived from the "tilde" equations in (47), following the proofs of (29) and (36). For any
E[|∫T0∫Ω|Aij(˜un,˜vn)|(|∇˜un|2+|∇˜vn|2)dxdt|q2]≤C,i,j=1,2, | (48) |
where the constant
A stochastic basis is needed for the limit of the Skorokhod representations, i.e., for the variables
˜S=(˜D,˜F,{˜Ft}t∈[0,T],˜P,˜Wu,˜Wv), | (49) |
where
Given the
Lemma 8.1 (convergence). The limits
˜u,˜v∈L2(˜D,˜F,˜P;L2(0,T;H1(Ω)))⋂L2(˜D,˜F,˜P;L∞(0,T;L2(Ω)))⋂L2(˜D,˜F,˜P;C([0,T];(H1(Ω))⋆)), |
and
Let
(i)˜un→˜u,˜vn→˜vin L2(˜D,˜F,˜P;L2(0,T;L2(Ω))),(ii)˜un⇀˜u,˜vn⇀˜vin L2(˜D,˜F,˜P;L2(0,T;H1(Ω))),(iii)˜un⋆⇀˜u,˜vn⋆⇀˜vin L2(˜D,˜F,˜P;L∞(0,T;L2(Ω))),(iv)˜un→˜u,˜vn→˜vin L2(˜D,˜F,˜P;C([0,T];(H1(Ω))⋆)),(v)√|Ai1(˜un,˜vn)|∇˜un⇀√|Ai1(˜u,˜v)|∇˜uin L2(˜D,˜F,˜P;L2(0,T;L2(Ω))),i=1,2,(vi)√|Ai2(˜un,˜vn)|∇˜vn⇀√|Ai2(˜u,˜v)|∇˜vin L2(˜D,˜F,˜P;L2(0,T;L2(Ω))),i=1,2,(vii)˜Wnu→˜Wu,˜Wnv→˜Wvin L2(˜D,˜F,˜P;C([0,T];U0)),(viii)˜un0→˜ui,0,˜vn0→˜v0in L2(˜D,˜F,˜P;L2(Ω)). | (50) |
Proof. The strong convergences (ⅰ) follow from (43), the moment estimate (44) with
Part (ⅳ) is a consequence of (43) and Vitali's convergence theorem, given the moment bounds (with some
˜E‖w‖qC([0,T];(H1(Ω))⋆)≲˜E[‖w‖q−1/2L∞(0,T;L2(Ω))‖w‖1/2C([0,T];(H1(Ω))⋆)]≲1, |
for
Let us verify part (v). Set
Our final step is to pass to the limit in the Faedo-Galerkin equations (47).
Lemma 8.2 (limit equations). The limits
∫Ω˜u(t)φudx−∫Ω˜u0φudx+∫t0∫Ω(Du(∫Ω˜u(t,x)dx)∇˜u+A11(˜u,˜v)∇˜u+A12(˜u,˜v)∇˜v)⋅∇φudxds=∫t0∫ΩF(˜u,˜v)φudxds+∫t0∫Ωσu(˜u)φudxd˜Wu(s), | (51) |
∫Ω˜v(t)φvdx−∫Ω˜v0φvdx+∫t0∫Ω(Dv(∫Ω˜v(t,x)dx)∇˜v+A21(˜u,˜v)∇u+A22(˜u,˜v)∇v)⋅∇φvdxds=∫t0∫ΩG(˜u,˜v)φvdxds+∫t0∫Ωσv(˜v)φvdxd˜Wv(s), | (52) |
for all
Proof. We will focus on (51). The second equation (52) can be treated similarly. First, recall that the space
Fix
‖Iu‖2L2(˜D×(0,T))=˜E∫T0(Iu(ω,t))2dt=0, |
which would imply that
E[∫T01Z(ω,t)Iu(ω,t)]dt=0, | (53) |
for a measurable set
The Faedo-Galerkin equations (47) holds in
∫Ω˜un(t)φudx+∫t0∫ΩDu(∫Ω˜un(t,x)dx)∇˜un⋅∇Πnφudxds+∫t0∫Ω(A11(˜un,˜vn)∇˜un+A12(˜un,˜vn)∇˜vn)⋅∇Πnφudxds=∫Ω˜un0φudx+∫t0∫ΩF(˜un,˜vn)Πnφudxds+∫t0∫Ωσnu(˜un)Πnφudxd˜W(n)u(s). | (54) |
We multiply (54) with
By part (viii) of (50), we obtain
In what follows, we will make repeated use of the following simple fact: If
1Z(ω,t)φu(x)∈L∞(˜D×(0,T)×Ω)=:L∞ω,t,x,1Z(ω,t)∇φu(x)∈L2(˜D×(0,T)×Ω)=:L2ω,t,x. |
The weak convergence in
˜E[∫T01Z(ω,t)(∫t0∫ΩDu(∫Ω˜un(t,x)dx)∇˜un⋅∇Πnφudxds)dt]n↑∞⟶˜E[∫T01Z(ω,t)(∫t0∫ΩDu(∫Ω˜u(t,x)dx)∇˜u⋅∇φudxds)dt], |
where we have used that
Regarding the cross-diffusion terms, set (for
an:=√|A1i(˜un,˜vn)|,bn:=∇Πnφu,cn:=anbn,a:=√|A1i(˜u,˜v)|,b:=∇φu,c:=ab,dn=√|A1i(˜un,˜vn)|∇˜wn,d=√|A1i(˜u,˜v)|∇˜w |
and write
A1i(˜un,˜vn)∇˜wn⋅∇Πnφu=cn⋅dn. |
Recalling that
‖b−bn‖L4(Ω)≤‖φu−Πnφu‖W1,4(Ω)≲‖φu−Πnφu‖H2Nn↑∞⟶0, |
where we have used that
a2n=|A1i(˜un,˜vn)|→|A1i(˜u,˜v)|=a2in L2ω,t,x. |
Thus, by the Brezis-Lieb lemma,
‖a4n‖4L4ω,t,xn↑∞⟶‖a4‖4L4ω,t,x. |
Passing to a subsequence if necessary, we may as well assume that
As a result of
E[∫T01Z(ω,t)(∫t0∫ΩA1i(˜un,˜vn)∇˜wn⋅∇Πnφudxds)dt]n↑∞⟶E[∫T01Z(ω,t)(∫t0∫ΩA1i(˜u,˜v)∇˜w⋅∇φudxds)dt]. |
Using that
E[∫T01Z(ω,t)(∫t0∫ΩF(˜un,˜vn)Πnφudxds)dt]n↑∞⟶E[∫T01Z(ω,t)(∫t0∫ΩF(˜u,˜v)φudxds)dt]. |
For the stochastic integral, we will use Lemma 2.1 to prove that
∫t0σnu(˜un)d˜W(n)u(s)n↑∞⟶∫t0σu(˜u)d˜Wu(s)in L2(0,T;L2(Ω)), | (55) |
in probability (with respect to
σnu(˜un)→σu(˜u)in L2(0,T;L2(U;L2(Ω))),˜P−almost surely. | (56) |
Clearly,
∫T0‖σu(˜u)−σnu(˜un)‖2L2(U;L2(Ω))dt≤∫T0‖σu(˜u)−σu(˜un)‖2L2(U;L2(Ω))dt+∫T0‖σu(˜u)−σnu(˜u)‖2L2(U;L2(Ω))dt=:I1+I2. | (57) |
By (12) and (43), we easily obtain
I1n↑∞⟶0,˜P−almost surely. | (58) |
For the
I2=∫T0∑k≥1‖σu,k(˜u)−σnu,k(˜u)‖2L2(Ω)dt=∫T0∑k≥1‖σu,k(˜u)−n∑l=1σu,k,l(˜u)el‖2L2(Ω)dt=∫T0∑k≥1‖σu,k(˜u)−Πn(σu,k(˜u))‖2L2(Ω)dt=:∫T0Σn(t)dt, |
where
The integrand can be dominated by an
0≤Σn(t)≤4∑k≥1‖σu,k(˜u(t))‖2L2(Ω)=4‖σu(˜u(t))‖2L2(U;L2(Ω))(12)≤C(1+‖˜u(t)‖2L2(Ω))∈L1t, |
recalling that
This calculation also shows that
and
Πn(∑k≥1σu,k(˜u))n↑∞⟶∑k≥1σu,k(˜u)in L2(Ω), |
for a.e.
Σn(t)n↑∞⟶0,a.e. on [0,T] (and a.s), |
an application of Lebesgue's dominated convergence theorem supplies
I2n↑∞⟶0,˜P−almost surely. | (59) |
Combining (57), (58) and (59), we arrive at (56). By Lemma 2.1, this implies (55).
Passing to a subsequence (not relabeled), we may replace "in probability" by "
˜E[‖∫t0σnu(˜un)d˜W(n)u‖qL2((0,T);L2(Ω))]=˜E[(∫T0‖n∑k=1∫t0σnu,k(˜vn)d˜Wnu,k‖2L2(Ω)dt)q2]≤ˉCT˜E[supt∈[0,T]‖n∑k=1∫t0σnu,k(˜un)d˜Wnu,k‖qL2(Ω)]≤CT˜E[(∫T0n∑k=1‖σnu,k(˜un)‖2L2(Ω)dt)q2]≤Cσ,T. |
Hence, by Vitali's convergence theorem, (55) implies
∫t0σnu(˜un)d˜W(n)u(s)→∫t0σu(˜u)d˜Wu(s)in L2(˜D,˜F,˜P;L2(0,T;L2(Ω))). |
Using this and the fact that
˜E[∫T01Z(ω,t)(∫t0∫Ωσnu(˜un)Πnφudxd˜Wnu(s))dt]=˜E[∫T0∫Ω(∫t0σnu(˜un)d˜W(n)u(s))(1Z(ω,t)Πnφu(x))dxdt]n↑∞⟶˜E[∫T01Z(ω,t)(∫t0∫Ωσu(˜u)φudxd˜Wu(s))dt]. |
This concludes the proof of (53), which implies that the desired (51) holds.
Remark 7. We have proved that the Skorokhod representations (42), (43) satisfy the weak formulation (51), (52) for a.e.
This section proves that the martingale solution
The nonnegativity result is contained in
Lemma 9.1. The solution
Proof. In this proof we drop the tildes on the relevant functions, writing for example
Sε(w)={w2−ε26if w<−ε,−w42ε2−4w33εif −ε≤w<0,0if w≥0. |
Observe that
S′ε(w)={2ww<−ε,−2w3ε2−4w2εw∈[−ε,0),0w≥0S″ε(w)={2w<−ε,−6w2ε2−8wεw∈[−ε,0),0w≥0. |
It is easy to see that
∫ΩSε(un(t))dx−∫ΩSε(un(0))dx=−∫t0∫ΩS″ε(un(s))Du(∫Ωun(s,x)dx)|∇un|2dxds−∫t0∫ΩS″ε(un(s))(A11(un,vn)∇un+A12(˜un,˜vn)∇vn)⋅∇undxds+∫t0∫ΩS′ε(un(s))F(un,vn)dxds+n∑k=1∫t0∫ΩS′ε(un(s))σnu,k(un)dxdWnu,k+12n∑k=1∫t0∫ΩS″ε(un(s))(σnu,k(un))2dxds=:5∑i=1Ii. | (60) |
It is easy to see that
S″ε(w)=0for w≥0,andS″ε(w)≥0for w∈R,A11(w,⋅)≥0andA12(w,⋅)=0,for w≤0. | (61) |
As a result,
I2:=−∫t0∫ΩS″ε(un(t))×(A11(un,vn)∇un+A12(˜un,˜vn)∇vn)⋅∇undxds=−∬{un(t,x)≥0}S″ε(un(t))×(A11(un,vn)∇un+A12(un,vn)∇vn)⋅∇undxds−∬{un(t,x)<0}S″ε(un(t))×(A11(un,vn)∇un+A12(un,vn)∇vn)⋅∇undxds=−∬{un(t,x)<0}S″ε(un(t))×(A11(un,vn)∇un+A12(un,vn)∇vn)⋅∇undxds(61)≤0. |
Similarly, from the definition of the function
Keeping in mind the convergences in (50) (see also [10,Section 3.2]), we send
(62) |
Sending
(63) |
for a.e.
1. | M Bendahmane, K H Karlsen, F Mroué, Stochastic electromechanical bidomain model * , 2024, 37, 0951-7715, 075023, 10.1088/1361-6544/ad5132 |