• 検索結果がありません。

We denote by the characteristic size of the heterogeneity in the medium

N/A
N/A
Protected

Academic year: 2022

シェア "We denote by the characteristic size of the heterogeneity in the medium"

Copied!
27
0
0

読み込み中.... (全文を見る)

全文

(1)

ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu ftp ejde.math.txstate.edu (login: ftp)

HOMOGENIZED MODEL FOR FLOW IN PARTIALLY FRACTURED MEDIA

CATHERINE CHOQUET

Abstract. We derive rigorously a homogenized model for the displacement of one compressible miscible fluid by another in a partially fractured porous reservoir. We denote by the characteristic size of the heterogeneity in the medium. A function α characterizes the cracking degree of the rock. Our starting point is an adapted microscopic model which is scaled by appropriate powers of . We then study its limit as 0. Because of the partially fractured character of the medium, the equation expressing the conservation of total mass in the flow is of degenerate parabolic type. The homogenization process for this equation is thus nonstandard. To overcome this difficulty, we adapt two-scale convergence techniques, convexity arguments and classical compactness tools. The homogenized model contains both single porosity and double porosity characteristics.

1. Introduction and main result

We consider the displacement of a two-component mixture through a highly contrasted porous medium, with fractures and matrix blocks. Assuming that the matrix blocks are disconnected, one usually models this type of setting using the concept of double porosity introduced by Barenblatt et al [5]. The fractured part is responsible for the macro-scale transport and the matrix part can store a concentra- tion longer than is to be expected in a single porous material. The less permeable part of the rock thus contributes as global sink or source terms for the transported solutes in the fracture (see for instance [4, 9]). By the way, the matrix of cells may also be connected so that some flow occurs directly within the cell matrix. We consider here such a partially fissured medium. Most commercial simulators have the added feature of including matrix-matrix connections. But there are very few theoretical derivations of a model for this phenomenon. The uncertainties relating to the size of the physical structure and the fluid content of the reservoir make understanding fluid flow through homogenization a pragmatic approach.

The present paper is an extension of the works [14, 13, 11]. In these latter ref- erences, the degree of interconnection between matrix and fractured part of the

2000Mathematics Subject Classification. 76S05, 35K55, 35B27, 76M50.

Key words and phrases. Miscible compressible displacement; porous medium;

partially fractured reservoir; double porosity; homogenization;

two-scale limit of a degenerate parabolic equation.

c

2009 Texas State University - San Marcos.

Submitted November 20, 2008. Published January 2, 2009.

1

(2)

medium, was characterized by a constant averaged parameter. In the present pa- per, a function αdescribes the interconnection. This function may be zero in the totally fractured part Ω0 of the domain Ω, and some technical challenges are en- countered there. The required estimates for the homogenization process involve the degenerating functionαand are technically more challenging than in [11]. The final homogenized equation are different in Ω0and in Ω\Ω0. From a more physical viewpoint, this contribution aims to give a model more adapted to the local geom- etry of a natural domain. The interconnection functionα is considered as a first order property of the medium, comparable, for instance, with the fracture intensity function.

We begin by recalling the equations describing the transport of two miscible species in a slightly compressible flow through a homogeneous porous medium, see [6, 20, 15] for details. The unknowns of the problem are the pressure p and the concentrationcof one of the two species of the mixture. Denoting byφthe porosity of the rock and byk its permeability, the mass conservation principles during the displacement are expressed by the equations

φ∂tp+ div(v) =qs, v=− k

µ(c)∇p, (1.1)

φ∂tc+v· ∇c−div(D(v)∇c) =qs(ˆc−c). (1.2) The average velocity of the flow v is given by the Darcy law in (1.1). We neglect the gravitational terms. The viscosity µis a nonlinear function depending on the concentration. For instance, in the Koval model [17], µis defined forc∈(0,1) by µ(c) =µ(0)(1 + (M1/4−1)c)−4, the constant M =µ(0)/µ(1) being the mobility ratio. Analogous to Fick’s law the dispersive flux is considered proportional to the concentration gradient and the dispersion tensor is

D(v) =φf DmId+Dp(v)

f DmId+|v|(αlE(v) +αt(Id− E(v)))

, (1.3) whereE(v)ij =vivj/|v|2, αl andαtare the longitudinal and transverse dispersion constants and Dm is the molecular diffusion. For the usual rates of flow, these real numbers are such thatαl ≥αt≥Dm >0. The terms containing qs are the injection and production terms.

We now aim to study a similar flow in a partially fractured porous medium. We thus consider a domain Ω⊂R3with a periodic structure, controlled by a parameter > 0 which represents the size of each block of the matrix. The C1 boundary of Ω is Γ and ν is the corresponding exterior normal. The standard period (= 1) is a cellY consisting of a matrix block Ym of externalC1 boundary ∂Ym and of a fracture domainYf. We assume that|Y|= 1. The-reservoir consists of copiesY covering Ω. The two subdomains of Ω are defined by

f = Ω∩

ξ∈A(Yf+ξ) , Ωm= Ω∩

ξ∈A(Ym+ξ) ,

whereAis an appropriate infinite lattice. The fracture-matrix interface is denoted by Γf m=∂Ωf∩∂Ωm∩Ω andνf mis the corresponding unit normal pointing out Ωf. See [14] and [1] for some illustrations of admissible structures. To homogenize the reservoir, we shall let tend to zero the sizeof the cells.

Following [14], we assume that the flow is made of two parts. The first component accounts for the global diffusion in the fracture system. The second one corresponds to high frequency spatial variations which lead to local storage in the matrix. A

(3)

functionα∈ C1(Ω) characterizes the interconnection between fractures and matrix.

It is assumed such that

0≤α(x)<1, α(x) +β(x) = 1, α(x) = 0 if and only ifx∈Ω0,

where Ω0is a bounded open subset of Ω. It may be given by experimental data on samples of porous media and by stochastic reconstruction (see [7] and the references therein). The functionαdescribing theinterconnection intensityis obviously linked with the commonly used concepts of fracture intensity and fracture-size distribution (see for instance [18]). Note that in [14, 13] and [11], the cracking degree was characterized by a constantα∈(0,1).

We thus adapt System (1.1)-(1.2) to such a decomposition of the flow. We also scale the equations for the rapidly varying part by appropriate powers of to conserve the flow between the matrix and the fractures as → 0 (cf [4, 14]).

The complete derivation of the microscopic model is justified in [11]. Denoting by J = (0, T),T >0, the time interval of interest, we consider:

φftf1+vf· ∇f1−div(D(vf)∇f1) =qs( ˆf1−f1) in Ωf×J, (1.4) φftpf+ div(vf) =qs, vf =− kf

µ(f1)∇pf in Ωf×J, (1.5) φtC1+V· ∇C1−div(D(V)∇C1) =qs( ˆC1−C1) in Ωm×J, (1.6) φtc1+V· ∇c1−div(D(V)∇c1) =qs(ˆc1−c1) in Ωm×J, (1.7) φtc2+V· ∇c2−div(D(V)∇c2) =qs(ˆc2−c2) in Ωm×J, (1.8) φtp+ div(V) =qs, V=Vs+Vh, in Ωm×J, (1.9) Vs=−α(c1+c2) k

µ(m1)∇p, Vh=−(1−α(c1+c2)) k

µ(m1)∇p, (1.10) wherem1=αc1+βC1. The flow in the fractures is described by (1.4)-(1.5). The matrix behavior is described by (1.6)-(1.10). In particular, (1.6) governs the slowly varying component while (1.7)-(1.8) governs the high frequency varying ones. We note that the former system becomes of double degenerate type as→0. Indeed, in the subset Ω0, the parabolic character of (1.6)-(1.9) is only ensured by the term 2. Moreover Eq. (1.9) is also of degenerate parabolic type in Ω\Ω0 since we can solely state that (c1+c2)(x, t)≥0 in Ω×J.

The model is completed by the following boundary and initial conditions. We begin by the transmission relations across the interface Γf m×J.

βD(vf)∇f1·νf m=D(V)∇C1·νf m, (1.11) αD(vf)∇f1·νf m=D(V)∇c1·νf m, (1.12) αD(vf)∇(1−f1)·νf m=−αD(vf)∇f1·νf m=D(V)∇c2·νf m, (1.13) f1=αc1+βC1, α(c1+c2) =α, (1.14) vf·νf m=V·νf m, pf =p. (1.15) We add a zero flux condition out of the full domain Ω

D(vf)∇f1·ν = 0 on∂Ωf∩Γ, (1.16) D(V)∇C1·ν=D(V)∇c1·ν=D(V)∇c2·ν= 0 on∂Ωm∩Γ, (1.17) vf·ν = 0 on∂Ωf∩Γ, V·ν= 0 on∂Ωm∩Γ, (1.18)

(4)

and the following initial conditions in Ω

(f1(x,0), C1(x,0), c1(x,0), c2(x,0)) = (f1o(x), C1o(x), co1(x), co2(x)), (1.19) pf(x,0) =χf(x)po(x), p(x,0) =χm(x)po(x). (1.20) Let us now enumerate the assumptions. The source term qs is a nonnegative function ofLq(Ω×J),q >2, and

αˆc1+βCˆ1= ˆf1, 0≤fˆ1≤1, ˆc1+ ˆc2= 1.

As we assume a periodic structure in the reservoir, the porosities (φf(x), φ(x)) = (φf(x), φ(x)) and the permeabilities (kf(x), k(x)) = (kf(x), k(x)) of the fracture and of the matrix are periodic of period (Yf, Ym). These quantities are assumed to be smooth and bounded, but globally they are discontinuous across Γf m. We assume moreover

0< φ≤φf(x), φ(x)≤φ−1 , k|ξ|2≤kf(x)ξ·ξ, k(x)ξ·ξ≤k−1 |ξ|2, k>0, a.e. in Ω, for allξ∈R3. The viscosityµ∈W1,∞(Ω×(0,1)) is such that

0< µ≤µ(x, c)≤µ+ ∀c∈(0,1), µ(x, c) =µ∈R+ in Ω0.

For sake of simplicity we have assumed that the viscosity is constant in Ω0. We then can pass to the limit in Ω0without introducing a dilation operator (see [11] Section 4 for the details). The tensor Dis already defined in (1.3). The tensor D has a similar structure but its diffusive part (α+β2)DmIdcontains the proportions of slowly and rapidly varying flows in the matrix. The main property of these tensors is

D(vf)ξ·ξ≥φ(Dmt|vf|)|ξ|2, ∀ξ∈R3,

D(V)ξ·ξ≥φ(Dm(α+β2) +αt|Vs+2Vh|)|ξ|2, ∀ξ∈R3. (1.21) We assume thatpobelongs toH1(Ω), and that (f1o, C1o, co1, co2)∈(L(Ω))4satisfies 0≤f1o(x)≤1 a.e. in Ω, (1.22) γ≤co1(x)≤γ+, (γ, γ+)∈R2, a.e. in Ω, (1.23) αco1(x) +βC1o(x) =χmf1o(x), 0≤co1(x) +co2(x)≤1 a.e. in Ωm. (1.24) The main result of the paper is the following.

Theorem 1.1. As the scaling parameter tends to zero, the microscopic model (1.4)-(1.20) converges to the following macroscopic model. The homogenized pres- sure problem is

φfYfΩ\Ω0φYm

tpf−div KHα µ(f1)∇pf

=qs−χ0 Z

Ym

φ ∂tp0dy inΩ×J, φ(y)∂tp0+ div

y (V0) =qs, V0=−k(y)

µ(f1) inΩ0×Ym×J, pf(x, t) =p0(x, y, t) ify∈Γf m, (x, t)∈Ω×J,

KHα∇pf·ν= 0 on∂Ω×J, pf(x,0) =p0(x, y,0) =po(x) in Ω×Ym. The homogenized concentrations problem is inΩ×J:

φfYftf1Ω\Ω01

βφYmtC101 β

Z

Ym

φ(y)∂tC10dy

(5)

− KYH

f

µ(f1)∇pf· ∇f1−χΩ\Ω0 KYH

m

βµ(f1)∇pf· ∇C1−χ0 β

Z

Ym

k(y)

µ ∇yp0· ∇yC10dy

−div(DfH(∇pf)∇f1)−χΩ\Ω01

βdiv(DmH(∇pf)∇C1)

=qs|Yf|( ˆf1−f1) +1 βqs

1−χΩ\Ω0|Ym|C1−χ0

Z

Ym

C10(·, y,·)dy ,

φYmχΩ\Ω0tc10

Z

Ym

φ(y)∂tc01dy−α

βφYmχΩ\Ω0tC1

−α βχ0

Z

Ym

φ(y)∂tC10dy−χΩ\Ω0KYH

m

µ(f1)∇pf· ∇c1−α β∇C1

−χ0 Z

Ym

k(y)

µ ∇yp0· ∇yc01−α β∇yC10

dy

−χΩ\Ω0div(DHm(∇pf)∇c1) +χΩ\Ω0α

βdiv(DmH(∇pf)∇C1)

=qs ˆ

c1−χΩ\Ω0|Ym|c1−χ0 Z

Ym

c01(·, y,·)dy

−α βqs

1−χΩ\Ω0|Ym|C1−χ0

Z

Ym

C10(·, y,·)dy , and inΩ0×Ym×J

φ(y)∂tC10−k(y)

µ ∇yp0· ∇yC10−div

y D(k(y)

µ ∇yp0)∇yC10

=qs( ˆC1−C10), φ(y)∂tc01−k(y)

µ ∇yp0· ∇yc01−div

y D(k(y)

µ ∇yp0)∇yc01

=qs(ˆc1−c01), completed by

DHf (∇pf)∇f1−1

βDHm(∇pf)∇C1

·ν

Γ×J= 0, DHm(∇pf)∇c1−α

βDHm(∇pf)∇C1

·ν

Γ×J = 0, f1

t=0=f1o, c1

t=0=c01

t=0=co1, C1

t=0=C10

t=0=C1o, f1=αc1+βC1 a.e. inΩ×J, f1=βC10 a.e. inΩ0×Γf m×J.

The homogenized quantities KHα, DHf and DHm are defined in (3.3), (3.17) and (3.18) below.

The homogenization process then leads to a macroscopic model containing both single porosity and double porosity characteristics. We show that the double poros- ity part of the model almost disappears as soon as a direct flow occurs in the matrix (see the equations in Ω\Ω0). It emphasizes in particular the role of the dispersion tensor which models all the velocities heterogeneity at the microscopic level. It is characteristic of a miscible flow (see [3] and the references therein). The result is thus quite different of the one obtained in [14, 13]. Nevertheless, even in Ω\Ω0, the model captures the interactions between the matrix and the fractured part.

Indeed, the homogenized permeability and diffusion tensors strongly depend on the transmission functionα. One could compare this effects with some models where

(6)

the permeability is concentration dependent: propagation in clays (see [16] and the references therein) or blood flow in micro vessels (see [22] and the references therein) for instance. And in the subset Ω0 where no direct transmission occurs, the model is of double porosity type.

This paper is organized as follows. Section 2 is devoted to the analysis of the microscopic model. We derive uniform estimates for the solutions. Convergence results are stated using two-scale convergence techniques, convexity arguments and classical compactness tools. In Section 3, we pass to the limit→0 and we get the homogenized model described in Theorem 1.1.

2. Analysis of the microscopic model

The existence of weak solutions for the problem (1.4)-(1.20) is proved in [11].

The proof is of course inspired by the statement of the existence of solutions for Problem (1.1)-(1.2) in a homogeneous porous medium (see [10]). But the decom- position of the flow in the matrix part of the domain induces additional difficulties.

Appropriate concentrations spaces for the problem are introduced following [13]:

His the Hilbert space H=L2(Ωf)×L2(Ωm)×L2(Ωm) with the inner product [uf, um, Um],[ψf, ψmm]

H

= Z

f

uf(x)ψf(x)dx+ Z

m

um(x)ψm(x)dx+ Z

m

Um(x) Ψm(x)dx, andV is the Banach space

V=H

(uf, um, Um)∈H1(Ωf)×H1(Ωm)×H1(Ωm);

γfuf =αγm um+βγmUm on Γf m endowed with the norm

k(uf, um, Um)kV=kχfufkL2(Ω)+kχmumkL2(Ω)+kχmUmkL2(Ω)

+kχf∇ufk(L2(Ω))3+kχm∇umk(L2(Ω))3+kχm∇Umk(L2(Ω))3, where γj :H1(Ωj)→L2(∂Ωj) is the usual trace map and χj is the characteristic function associated with Ωj,j=f, m. We also introduce the Banach spaceVc

Vc=L2(Ωm)×L2(Ωm)∩

(u1, u2)∈H1(Ωm)×H1(Ωm);

α=αγm (u1+u2) on Γf m endowed with the norm

k(u1, u2)kVc =kχmu1kL2(Ω)+kχmu2kL2(Ω)+kχm∇u1k(L2(Ω))3+kχm∇u2k(L2(Ω))3. We note that for any fixed >0, the problem is of parabolic type. Then, adapting the proof of [10] to the present piecewise structure, one can state the following existence result (see [11] for a detailed proof).

Theorem 2.1. Let 0 < < 1. There exists a solution (pf, p, f1, c1, C1, c2) of Problem (1.4)-(1.20)in the following sense.

(i) The pressure part (pf, p) belongs to L2(J;H1(Ωf))×L2(J;H1(Ωm)) and is a weak solution of (1.5), (1.9)-(1.10), (1.15), (1.18) and (1.20). Indeed, for any

(7)

function ψ∈ C1(J;H1(Ω)) such thatψ|t=T= 0,

− Z

Ω×J

fφfpfmφp)∂tψ +

Z

Ω×J

χf kf

µ(f1)∇pfm(α(c1+c2)(1−2) +2) k µ(m1)∇p

· ∇ψ

=− Z

fφfmφ)poψ(x,0) + Z

Ω×J

qsψ.

(2.1)

(ii) The concentration part (f1, c1, C1, c2) is such that (f1, c1, C1) ∈ L2(J;V)∩ H1(J; (V)0) and (c1, c2) ∈ L2(J;V c)∩H1(J; (Vc)0). It satisfies for any test functions(df, d1, D1)∈L2(J;V)andd2∈L2(J;H1(Ωm))the following relations.

Z

f×J

φftf1df+ Z

m×J

φtc1d1+ Z

m×J

φtC1D1+ Z

f×J

(vf· ∇f1)df +

Z

m×J

V·(d1∇c1+D1∇C1) + Z

f×J

D(vf)∇f1· ∇df +

Z

m×J

D(V)∇c1· ∇d1+ Z

m×J

D(V)∇C1· ∇D1

= Z

f×J

qs( ˆf1−f1)df+ Z

m×J

qs(ˆc1−c1)d1+ Z

m×J

qs( ˆC1−C1)D1, (2.2) and

Z

m×J

φtc2d2+ Z

m×J

(V· ∇c2)d2+ Z

m×J

D(V)∇c2· ∇d2

− Z

∂Ωm×J

(D(V)∇c2·νmm d2

= Z

m×J

qs(1−c2)d2.

(2.3)

Furthermore, the following maximum principles hold:

0≤f1(x, t)≤fˆ1 a.e. inΩf ×J, (2.4) 0≤m1(x, t)≤fˆ1, 0≤c1(x, t) +c2(x, t)≤1 a.e. inΩm×J, (2.5) γ≤c1(x, t)≤γ+ a.e. inΩm×J. (2.6) We now state some uniform estimates for the solutions of the microscopic sys- tem. We begin by stating the following properties of the pressure solutions of the problem (1.5), (1.9)–(1.10), (1.15), (1.18), (1.20). One of the main difficulties of the homogenization problem appears in the following lemma. Indeed, letting to 0, Equation (1.9) is of degenerate parabolic type because one can only ensure that c1+c2≥0. It is a main difference with our former work in [11].

Lemma 2.2. The pressure satisfies the following uniform estimates kpfkL(J;Lq(Ωf))+kpfkL2(J;H1(Ωf)) ≤C,

kvfk(L2(J;L2(Ωf)))2≤C, kpkL(J;Lq(Ωm))≤C,

(8)

1/2(c1+c2)1/2∇pk(L2(J;L2(Ωm)))3+k∇pk(L2(J;L2(Ωm)))3≤C, kVsk(L2(J;L2(Ωm)))3 ≤C,

kVhk(L2(J;L2(Ωm)))3 ≤C.

Furthermore the time derivative (χfφftpfmφtp) is uniformly bounded in L2(J; (H1(Ω))0).

Proof. The estimates are derived from integration by parts. We multiply (1.5) by pf and integrate over Ωf×J. We multiply (1.9) byp and integrate over Ωm×J. Summing up the resulting relations, we obtain

1 2 Z

f

φf|pf|2dx+1 2 Z

m

φ|p|2dx+ Z

f×J

kf

µ(f1)∇pf· ∇pfdxdt +

Z

m×J

α(c1+c2)(1−2) +2 k

µ(m1)∇p· ∇pdxdt

= 1 2 Z

fφf(x) +χmφ(x))|po(x)|2dx+ Z

Ω×J

qsfpfmp)dxdt.

Applying the Cauchy-Schwarz and Young inequalities with the properties ofφf, kf,k andµin the latter relation, we get

φ 2

Z

f

|pf|2dx+φ 2

Z

m

|p|2dx+k µ+

Z

f×J

|∇pf|2dxdt +k

µ+ Z

m×J

α(c1+c2)|∇p|2+2(1−α(c1+c2))|∇p|2 dxdt

≤C kpokL2(Ω),kqskL2(Ω×J) +

Z

f×J

|pf|2dxdt+ Z

m×J

|p|2dxdt.

Using the Gronwall lemma, we prove the desired estimates, but inL2instead ofLq. The result on the time derivatives then follows straightforward from (1.5), (1.9)- (1.10). It remains to show that the pressure is uniformly bounded inL(J;Lq(Ω)).

Letη >0. We multiply Eq. (1.5) (respectively (1.9)) by qpf(pf2+η)q/2−1 (resp.

qp(p2+η)q/2−1) and we integrate by parts over Ωf (resp. Ωm). We obtain d

dt Z

χf(pf2+η)q/2m(p2+η)q/2 +

Z

f

kf

µ(f1)(pf2+η)q/2−1∇pf· ∇pf +

Z

f

kf

µ(f1)qpf2(pf2+η)q/2−2∇pf· ∇pf +

Z

m

(α(c1+c2)(1−2) +2) k

µ(m1)(p2+η)q/2−1∇p· ∇p +

Z

m

(α(c1+c2)(1−2) +2) k

µ(m1)qp2(p2+η)q/2−2∇p· ∇p

= Z

qsq χfpf(pf2+η)q/2−1mp(p2+η)q/2−1 .

(9)

The four last terms of the left hand side of the latter relation are nonnegative. The right hand side term is estimated as follows using the H¨older inequality.

Z

qsq χfpf(pf2+η)q/2−1mp(p2+η)q/2−1

≤C Z

|qs| χf(pf2+η)(q−1)/2m(p2+η)(q−1)/2

≤CZ

χf(pf2+η)q/2m(p2+η)q/2(q−1)/qZ

|qs|q1/q

≤CZ

χf(pf2+η)q/2m(p2+η)q/2(q−1)/q .

We conclude with the Gronwall lemma that χf(pf2 +η)1/2m(p2+η)1/2 is uniformly bounded inL(J;Lq(Ω)). It follows thatχfpfmp is also uniformly

bounded inL(J;Lq(Ω)).

We now establish the following results concerning the concentrations functions (f1, C1, c1, c2).

Lemma 2.3. (i) The functions (f1, C1, c1, c2) are uniformly bounded in the spaceL(J;L2(Ωf))×(L(J;L2(Ωf)))3 and are such that

0≤f1(x, t)≤fˆ1≤1 almost everywhere in Ωf×J, 0≤αc1(x, t) +βC1(x, t)≤fˆ1≤1 almost everywhere in Ωm×J

0≤c1(x, t) +c2(x, t)≤1 almost everywhere inΩm×J, γ≤c1(x, t)≤γ+ almost everywhere in Ωm×J;

(ii) the sequence ((D1/2mt1/2|vf|1/2)∇f1)is uniformly bounded in (L2(Ωf× J))3;

(iii) fori= 1,2, the diffusive termsα1/2(1 + (c1+c2)1/2|∇p|1/2)∇ci and(1 +

|∇p|1/2)∇ciare uniformly bounded in(L2(Ωm×J))3. The same estimates hold forC1.

Proof. The maximum principles of(i)are a direct consequence of the construction of the solution (f1, C1, c1, c2) in Theorem 2.1. We write the variational formulation (2.2) with the test function (df, d1, D1) = (f1, c1, C1). We get

1 2

Z

f

φf|f1|2dx+1 2 Z

m

φ(|c1|2+|C1|2)dx+ Z

f×J

D(vf)∇f1· ∇f1dx dt +

Z

m×J

(D(V)∇c1· ∇c1+D(V)∇C1· ∇C1)dx dt +

Z

f×J

(vf· ∇f1)f1dx dt+ Z

m×J

V·(c1∇c1+C1∇C1)dx dt +

Z

Ω×J

qsf|f1|2m(|c1|2+|C1|2))dx dt (2.7)

= Z

Ω×J

qs1f1dx dt+ Z

m×J

qs(ˆc1c1+ ˆC1C1)dx dt +1

2 Z

φf|f1o|2(|co1|2+|C1o|2) dx.

(10)

The convective terms in (2.7) are estimated as follows using the Cauchy-Schwarz and Young inequalities. In the fractured part, we write

Z

f×J

(vf· ∇f1)f1dx ≤

Z

f×J

αt

2 |vf| |∇f1|2dx+Ckf1k2 Z

f

|vf|dx, where 0≤f1(x, t)≤1 a.e. in Ωf×J andvf is uniformly bounded in (L1(Ωf×J))3 thanks to Lemma 2.2. In the matrix part, we get firstly

Z

m×J

V·(c1∇c1+C1∇C1) ≤

Z

m×J

αt

2 |Vs+2Vh|(|∇c1|2+|∇C1|2) +C

Z

f

(|Vs|+|Vh|) (|c1|2+|C1|2)dx.

The second term of the right-hand side of the latter relation is treated as follows using Lemma 2.2.

Z

f

|Vs+Vh|(|c1|2+|C1|2)

≤ k+ µ

Z

f

α(c1+c2)(1−) +

|∇p| |c1|2+|C1|2

≤CZ

f

α2(c1+c2)2+2(1−α(c1+c2)2

|∇p|21/2

× kc1k2L(Ωm)+kC1k2L(Ωm)

≤ C

δ kc1k2L(Ωm)+kC1k2L(Ωm)

δ Z

f

(α+2)Dm |∇c1|2+|∇C1|2 , for any δ >0. The last term in the left-hand side of (2.7) is nonnegative. Using the latter estimates, the Cauchy-Schwarz and Young inequalities for the right-hand side source terms and the basic properties (1.21) of the tensorsDandD, it follows from (2.7) that

φ 2

Z

f|f1|2m(|c1|2+|C1|2))dx+φ Z

f×J

(Dmt

2 |vf|)|∇f1|2dxdt +φ

Z

m×J

((α+2)(1−δ)Dmt

2 |Vs+2Vh|) |∇c1|2+|∇C1|2 dxdt

≤C δ +C

Z

f×J

|f1|2dxdt.

We choose 0< δ <1. We use the Gronwall lemma to infer from the latter relation that p

α+β2∇c1 and |α(c1+c2) +3(1−α(c1+c2))∇p|1/2∇c1 are uniformly bounded in (L2(Ω×J))3. The estimates for f1, c1 andC1 follow. Once we know the estimate for c1, we obtain similar ones forc2 by multiplying (1.7) by c1, (1.8) byc2, integrating over Ωm and summing up the results to kill the terms on Γf m.

Our claim is proved.

We now have sufficient estimates to state the first convergence result. The proof of the homogenization process will be carried out by using the two-scale convergence introduced by G.Nguetseng in [19] and developed by Allaire in [2]. The basic definition and properties of this concept follow.

参照

関連したドキュメント

Recently, Velin [44, 45], employing the fibering method, proved the existence of multiple positive solutions for a class of (p, q)-gradient elliptic systems including systems

By applying the Schauder fixed point theorem, we show existence of the solutions to the suitable approximate problem and then obtain the solutions of the considered periodic

Theorem 4.8 shows that the addition of the nonlocal term to local diffusion pro- duces similar early pattern results when compared to the pure local case considered in [33].. Lemma

A monotone iteration scheme for traveling waves based on ordered upper and lower solutions is derived for a class of nonlocal dispersal system with delay.. Such system can be used

Xiang; The regularity criterion of the weak solution to the 3D viscous Boussinesq equations in Besov spaces, Math.. Zheng; Regularity criteria of the 3D Boussinesq equations in

We consider the Cauchy problem for nonstationary 1D flow of a compressible viscous and heat-conducting micropolar fluid, assuming that it is in the thermodynamical sense perfect

“Breuil-M´ezard conjecture and modularity lifting for potentially semistable deformations after

Then it follows immediately from a suitable version of “Hensel’s Lemma” [cf., e.g., the argument of [4], Lemma 2.1] that S may be obtained, as the notation suggests, as the m A