• Aucun résultat trouvé

The effective permeability of fractured porous media subject to the Beavers-Joseph contact law

N/A
N/A
Protected

Academic year: 2021

Partager "The effective permeability of fractured porous media subject to the Beavers-Joseph contact law"

Copied!
18
0
0

Texte intégral

(1)

HAL Id: hal-00904846

https://hal-univ-rennes1.archives-ouvertes.fr/hal-00904846

Submitted on 15 Nov 2013

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

The effective permeability of fractured porous media subject to the Beavers-Joseph contact law

Isabelle Gruais, Dan Polisevski, Florentina-Alina Stanescu

To cite this version:

Isabelle Gruais, Dan Polisevski, Florentina-Alina Stanescu. The effective permeability of fractured porous media subject to the Beavers-Joseph contact law. Asymptotic Analysis, IOS Press, 2014, 90 (3-4), pp.267-280. �10.3233/ASY-141248�. �hal-00904846�

(2)

The effective permeability of fractured porous media subject to the Beavers-Joseph contact law

Isabelle Gruaisa, Dan Poliˇsevskib, Florentina-Alina Stanescub

aUniversit´e de Rennes 1, I.R.M.A.R, Campus de Beaulieu, 35042 Rennes Cedex (France)

bI.M.A.R., P.O. Box 1-764, Bucharest (Romania).

Abstract

We are interested in the asymptotic behaviour of a fluid flow contained in a microscopic periodic distribution of fissures perturbating a porous medium where the Darcy law is valid, when the coupling between both systems is modeled by the Beavers-Joseph interface condition. As the small period of the distribution tends to zero, the interface condition is preserved on a microscopic scale under the additional assumption that the permeability co- efficients behave like the squared period of the distribution which is also the squared size of the fissures. Moreover, the resulting pressure is purely macroscopic unlike the velocity field which also depends on the microscopic variable.

Keywords: Fractured porous media, Stokes flow, Beavers-Joseph interface, Homogenization, Two-scale convergence.

2000 MSC: 35B27, 76M50, 76S05, 76T99 1. Introduction

The perturbation of a porous medium by fissures containing a Stokes flow give rise to nontrivial phenomena that cannot be modeled by the laws of fluid mechanics. Given a microscopic periodically distributed collection of thin fissures, we are interested in the asymptotic behaviour of a medium where the Stokes equations of a fluid constrained in the fissures are coupled with the Darcy equations of the surrounding porous medium through a contact law of Beavers-Joseph type [6] known as the Saffman’s variant [25]. The pioneering work by Beavers and Joseph yields the existence of a slip velocity at the interface between the fluid part and the fluid-saturated porous solid part. Proper rescaling of the fissures shows that the Beavers-Joseph condition

(3)

influences the asymptotic behaviour of the system as long as the permeability coefficients obey one of two admissible alternatives. As one of them was studied in [10], we concentrate on the second alternative, namely the case where permeability is of the same order of magnitude as the squared period of the distribution. Our aim is is to show that although this influence is very small in comparison with the previous case it still plays a nongligeable part that could not be simply deduced from the first one.

As we use arguments of the homogenization theory, we briefly recall basic facts about the homogenization of fluids.

An argument favoring the use of periodic homogenization in fluid flows is that it was already used to justify Darcy’s law [28]. Therefore, the present framework may be seen as a further development of this theory, although it does not deal with the formal method of asymptotic expansions [14], [15], [26] but rather extends the first rigorous proof based on the construction of a pressure extension due to [28] and followed by contextual variants [1], [16], [19]. with the restriction that, unlike previous works [1], [12], [28] relying on specific constructions, the velocity and pressure of the fluid have natural bounded extensions in the porous medium. We refer to [12] and references therein for developments about the physics and mathematics of this subject.

However unrealistic, the periodicity assumption allows to concentrate ideas on the actual process. Homogenization of phenomena in fractured media were studied later in [22], [1] and [23] that is, when an assumption of nonconect- edness could be dropped. Several models of fluid flows through fractured porous media (see [4], [5], [27], [9] and [24]) have been obtained by means of asymptotic methodswhere a homogeneous porous medium is altered by a possibly periodic distribution of microscopic fissures. Then, ε-periodicity allows to use procedures of the homogenization theory.

We consider an incompressible viscous fluid flow in a fractured porous media represented by a periodically structured domain consisting of two in- terwoven regions, separated by an interface. The first region represents the system of fissures which form the fracture, which is connected and where the flow is governed by the Stokes system. The second region, which is also connected, stands for the system of porous blocks, which have a certain per- meability and where the flow is governed by Darcy’s law. These two flows are coupled on the interface by the Saffman’s variant [25] of the Beavers-Joseph condition [6], [17] which was confirmed by [13] as the limit of a homogeniza- tion process. Besides the continuity of the normal component of the velocity, it imposes the proportionality of the tangential velocity with the tangen-

(4)

tial component of the viscous stress on the fluid-side of the interface. We prove here the existence and uniqueness of the solution of this model in our ε-periodic structure.

The paper is organized as follows. The problem of the flow is introduced in Section 2. More precisely, considering a microscopic periodic distribution of fissures we are interested in the behaviour of the equivalent material when the size of the fissures is of the same order of magnitude as the vanishing period of the distribution. The porous medium is described by a Darcy model in the solid part of the system and is coupled with Stokes equations in the fissures through a law of Beavers-Joseph type, see [6]. An adequate scaling shows that two cases arise depending on whether the permeability coefficients is of unity order or behave like the squared period of the distribution. Letting aside the first alternative which was studied in [10], we concentrate on the second one which displays new a priori estimates. The homogenization process is initiated in Section 3 thanks to a priori estimates and compacity arguments of the two-scale convergence theory (see [3], [18] and [21]) to identify convergent subsequences.

2. The flow through the ε-periodic structure

Let Ω be an open connected bounded set in RN(N ≥2), locally located on one side of the boundary ∂Ω, a Lipschitz manifold composed of a finite number of connected components.

LetYf be a Lipschitz open connected subset of the unit cubeY =]0,1[N, such that the intersections of ∂Yf with ∂Y are reproduced identically on the opposite faces of the cube and 0 ∈/ Yf. The outward normal on ∂Yf is denoted by ν. RepeatingY by periodicity, we assume that the reunion of all the ¯Yf parts, denoted by RNf , is a connected domain inRN with a boundary of class C2. Defining Ys=Y \Yf, we assume also that the reunion of all the Y¯s parts is a connected domain in RN.

For anyε∈]0,1[ we denote

Zε ={k ∈ZN, εk+εY ⊆Ω} (2.1) Iε ={k∈Zε, εk±εei+εY ⊆Ω,∀i∈1, N} (2.2) where ei are the unit vectors of the canonical basis in RN.

Finally, we define the system of fissures by Ωεf = int ∪k∈Iε(εk+εY¯f)

(2.3)

(5)

and the porous matrix of our structure by Ωεs = Ω\Ω¯εf. The interface between the porous blocks and the fluid is denoted by Γε=∂Ωεf. Its normal is:

νε(x) = νx ε

, x∈Γε (2.4)

where ν has been periodically extended to RN.

Let us remark that Ωεs and Ωεf are connected and that the fracture ratio of this structure is given by

m=|Yf| ∈]0,1[, as |Ωεf|

|Ω| →m when ε→0. (2.5) To the previous structure we associate a model of fluid flow through a fractured porous medium by assuming that there is a filtration flow in Ωεs obeying the Darcy’s law and that there is a viscous flow in Ωεf governed by the Stokes system. These two flows are coupled by a Saffman’s variant [25]

of the Beavers-Joseph condition [6], [17]. This system is completed by an impermeability condition on ∂Ω:

divvεs = 0 in Ωεs (2.6)

µεvεs=Kε(gε− ∇pεs) in Ωεs, (2.7)

divvεf = 0 in Ωεf, (2.8)

σεij =−pεfδij + 2µεeij(vεf) in Ωεf (2.9)

− ∂

∂xj

σεij =giε in Ωεf (2.10) vεs·νε =vεf ·νε on Γε, (2.11)

−pεsνiε−σijενjεεµεβε−1(vεfi −(vεf ·νεiε) on Γε, (2.12) vεs·n= 0 on ∂Ω, n the outward normal on∂Ω, (2.13) where vεs, vεf and pεs, pεf stand for the corresponding velocities and pres- sures, µε > 0 is the viscosity of the fluid, αε ∈ L(Ω) is the positive non- dimensional Beavers-Joseph number, gε ∈ L2(Ω)N is the exterior force and e(v) denotes the symmetric tensor of the velocity gradient defined by

eij(v) = 1 2

∂vi

∂xj + ∂vj

∂xi

.

(6)

Finally, the positive tensor of permeability is defined by:

Kε(x) = βε2Kx ε

, (2.14)

where K ∈ L(Y)N×N and βε > 0 stands for the magnitude of (T rKε)1/2 with respect to ε→0.

As usual, we use the notations:

H0(div,Ω) ={v ∈H(div,Ω), v ·ν= 0 on ∂Ω} (2.15) L20(Ω) ={p∈L2(Ω),

Z

p= 0} (2.16)

V0(div,Ω) ={v ∈H0(div,Ω), divv = 0 in Ω} (2.17) Next, we define

Hε={v ∈H0(div,Ω), v ∈H1(Ωεf)N}, (2.18) the Hilbert space endowed with the scalar product

(u, v)Hε= Z

εs

uv+ Z

εs

divudivv+ε2 Z

εf

e(u)e(v) +ε Z

Γε

εu−(γνεu)νεεv (2.19) whereγεandγνε denote respectively the trace and the normal trace operators on Γε with respect to Ωεf. Its corresponding subspace of incompressible velocities is

Vε ={v ∈V0(div,Ω), v ∈H1(Ωεf)N} (2.20) A useful property of the present structure is the existence of a bounded extension operator similar to that introduced in [7], [8] and [2] in the case of isolated fractures.

Theorem 2.1. There exists an extension operator Pε : H1(Ωεf) → H01(Ω) such that

Pεu=u in Ωεf (2.21)

|e(Pεu)|L2(Ω) ≤C|e(u)|L2(Ωεf), ∀u∈H1(Ωεf) (2.22) where C is independent of ε.

A straightforward consequence, via the corresponding Korn inequality, is

(7)

Lemma 2.2. There exists some constant C >0, independent ofε, such that

|uε|L2(Ωεf)+ε|∇u|L2(Ωεf) ≤C|u|Hε, ∀u∈Hε. (2.23) Proof. We prove that:

|ey(u)|2Yf +|γu−(γνu)ν|2Γ (2.24) is a norm on H1(Yf). Indeed, let

|ey(uk)|2Y

f +|γuk−(γνuk)ν|2Γ→0 (2.25) and

|uk|2Yf +|∇yuk|2Yf = 1. (2.26) From (2.26), we deduce that, at least for some subsequence and for some u∈H1(Yf):

uk * u in H1(Yf). (2.27)

Moreover, (2.25) implies that

ey(u) = 0 and γu−(γνu)ν = 0.

It follows that u is a rigid displacement with vanishing the tangential trace on our Γ, that is u≡ 0. The compactness of the inclusion L2(Yf)⊂H1(Yf) leads to the strong convergence

uk →u in L2(Yf).

Applying (2.26) again and Korn’s inequality, we infer that (2.27) is a strong convergence, which leads to a contradiction between (2.26) and u = 0. Fi- nally, (2.23) is obtained by rescaling the fact that (2.24) is a norm onH1(Yf).

Denoting

Aε = (Kε)−1, (2.28)

and using the positivity of Kε, we can assume without loss of genenerality that

∃a0 >0 such that Aεij(·)ξiξj ≥a0|ξ|2, ∀ξ ∈RN, a.e. in Ω. (2.29) Rescaling the velocity by

uε =

uεs in Ωεs uεf in Ωεf

= µε βε2

vεs in Ωεs vεf in Ωεf

(2.30)

(8)

then, for any u, v ∈Hε and q ∈L20(Ω), we define aε(u, v) =

Z

εs

Aεuv+βε2 Z

εf

e(u)e(v) +βε

Z

Γε

αεεu−(γνεu)νεεv (2.31)

bε(q, v) =− Z

q divv. (2.32)

We see that if the pair (uε, pε) is a smooth solution of the problem (2.6)–

(2.13), then it is also a solution of the following problem: To find (uε, pε)∈ Hε×L20(Ω) such that

aε(uε, v) +bε(pε, v) = Z

gεv, ∀v ∈Hε (2.33) bε(q, uε) = 0, ∀q∈L20(Ω) (2.34) Theorem 2.3. There exists a unique pair (uε, pε)∈Hε×L20(Ω) solution of (2.33)–(2.34).

Proof. AsH01(Ω) is obviously included inHε, the following inf-sup condition is easily satisfied by bε:

∃C1ε >0 such that inf

q∈L20(Ω)

sup

v∈Hε

bε(q, v)

|v|Hε|q|L2

0(Ω)

≥C1ε.

The coercivity condition (2.29) implies that

∃C2ε >0 such that aε(v, v)≥C2ε|v|2Hε, ∀v ∈Hε, that is the Vε-ellipticity of aε. As we also have

Vε ={v ∈Hε, bε(q, v) = 0, ∀q∈L20(Ω)}, (2.35) the proof is completed by Corollary 4.1, Ch. 1 of [11].

In the rest of the paper we shall study the asymptotic behaviour (when ε→0) of (uε, pε), the unique solution of (2.33)–(2.34).

As Aε defined by (2.28) is of ε0-order, we assume that

∃A∈Lper(Y)N2 such that Aε(x) =Ax ε

, x∈Ω (2.36)

∃g ∈L2(Ω)N such that gε→g strongly in L2(Ω). (2.37)

(9)

Because |Γε| is of ε−1-order, we expect that macroscopic effects of the Beavers-Joseph condition will appear only whenαεβεis ofε1-order (see [20]).

Therefore, we shall work under the hypothesis:

∃α∈Cper1 (Y) andα0 >0 such thatε−1βεαε(x) =αx ε

≥α0, x∈Ω. (2.38) Thus, for the study of the asymptotic behaviour it remains only the order of βε to be taken into account.

In the sequel we shall study the case when βε = O(ε). Without loss of generality we can consider from now on that

∃β >0 such that βε22β, ∀ε >0. (2.39) 3. The homogenization process

From now on, for any function ϕ defined on Ω ×Y we shall use the notations

ϕh =ϕ|Ω×Yh, ϕ˜h = 1

|Yh| Z

Yh

ϕ(·, y)dy, h∈ {s, f}, (3.1)

˜ ϕ=

Z

Y

ϕ(·, y)dy, that is ϕ˜= (1−m) ˜ϕs+mϕ˜f. (3.2) Also, for any sequence (ϕε)ε, bounded in L2(Ω×Y), we denote

ϕε* ϕ2

iff ϕε is two-scale convergent to ϕ∈L2(Ω×Y) in the sense of [3].

Noticing thatuε ∈Vε and settingv =uε in (2.33) we get

|uε|2Hε ≤C|uε|L2(Ω). (3.3) Applying (2.23) we find that

{uε}εis bounded in Vε and in V0(div,Ω), (3.4)

|uε|L2(Ωεf)+ε|∇uε|L2(Ωεf) ≤C, C being independent of ε. (3.5) It follows that ∃u∈L2(Ω×Y)N such that, on some subsequence

uε* u2 (3.6)

(10)

uε * Z

Y

u(·, y)dy∈V0(div,Ω) weakly in L2(Ω)N (3.7) Denoting χεf(x) = χfx

ε

and χεs(x) = χsx ε

, where χf and χs are the characteristic functions of Yf and Ys inY, we see that (χεsuε)ε, (χεfuε)ε and

χεf∂uε

∂xi

ε

are bounded in (L2(Ω))N, ∀i∈ {1,2,· · ·, N}.

Let us denote:

per1 (Yf) = {ϕ∈Hloc1 (RNf ), ϕ is Y-periodic, Z

Yf

ϕ= 0}. (3.8)

Lemma 3.1. uf ∈L2(Ω,( ˜Hper1 (Yf))N) and satisfies:

εχεf∇uεi * χ2 fyufi. (3.9) Proof. Using the compacity result of [21], it follows that∃ηi ∈L2(Ω×Y)N such that, on some subsequence we have

εχεf∇uεi * η2 i. (3.10) Next, we easily remark that ηi|Ω×Ys = 0.

Let us introduce

V0per(div, Yf) = {ϕ∈Hloc(div,RNf ), divyϕ= 0 in RNf ,

ϕ·ν = 0 on Γ, ϕ isY −periodic}. (3.11) Let ψ ∈L2(Ω, V0per(div, Yf)); denotingψε =ψ x,xε

,we have

ψε ∈H1(Ωεf)N and ψενε= 0 on Γε (3.12) Z

εf

ε∇uεi(x)ψε(x)dx=− Z

εf

εuεi(x)divxψ

x,x ε

dx. (3.13)

Using the two-scale convergences (3.6)-(3.10) we find Z

Ω×Yf

ηi(x, y)ψ(x, y)dxdy = 0, ψ ∈L2(Ω, V0per(div, Yf)). (3.14) As the orthogonal space of V0per(div, Yf) in L2(Ω×Y) is:

∇H˜per1 (Yf) ={∇q, q ∈H˜per1 (Yf)} (3.15)

(11)

(see Th.2.7, Ch.I[11]), it follows that there exists w∈L2(Ω; ˜Hper1 (Yf)N) such that

ηi(x, y) = χf(y)∇ywi(x, y), ∀i∈ {1, ..., N}. (3.16) Further we prove that

wi(x, y) = ufi(x, y)−u˜fi(x), ∀i∈ {1, ..., N}, which complets the proof.

Proposition 3.2. There holds:

divyuf = 0, in Ω×Yf (3.17) divyus= 0, in Ω×Ys (3.18) ufn =usn, on Ω×Γ (3.19) (1−m)div˜us+mdiv˜uf = 0, in Ω. (3.20) Proof. First, let a∈ {s, f}and let ζ ∈ D(Ω), θ ∈ D(Ω×Ya). Define

θε=θ x,x

ε

. There holds

0 = Z

εs

divuε(ζ+εθε) + Z

εf

divuε(ζ+εθε) =

=− Z

εs

uε(∇ζ+ε(∇θ)ε+∇yθε)− Z

εf

uε(∇ζ+ε(∇θ)ε+∇yθε)

→ − Z

Ω×Yf

uf(∇ζ+∇yθ)− Z

Ω×Ys

us(∇ζ+∇yθ) =

= Z

Ω×Yf

divyufθ+ Z

Ω×Ys

divyusθ+ Z

(mdiv˜uf + (1−m)div˜us)ζ from which we deduce (3.17), (3.18) and (3.20).

Second, let ζ ∈ D(Ω), θ ∈ D(Ω;Cper(Y)) and define θε

x,x ε

. There holds

0 = Z

divuε(ζ+εθε) =− Z

uε(∇ζ+ε(∇θ)ε+∇yθε) + Z

Γε

[uεn](ζ+εθε) =

(12)

=− Z

uε(∇ζ+ε(∇θ)ε+∇yθε)→ − Z

Ω×Yf

uf(∇ζ+∇yθ)−

Z

Ω×Ys

us(∇ζ+∇yθ) =

= Z

Ω×Yf

divyufθ+ Z

Ω×Ys

divyusθ+|Y| Z

(div˜uf + div˜us)ζ− Z

Ω×Γ

(ufn−usn

=− Z

Ω×Γ

(ufn−usn)θ which yields (3.19) and achieves the proof.

The same arguments as in [10] yield:

Proposition 3.3. There exists a constant C > 0independent of ε such that

|pε|L2(Ω)+|∇pε|L2(Ωεs) ≤C. (3.21) Lemma 3.4. There exists p ∈ L20(Ω×Y) with ps = ˜ps ∈ H1(Ω) and pf =

˜

pf ∈L2(Ω) such that, up to some subsequence, we have:

pε * p.2 (3.22)

Moreover: ps =pf =p and thus p∈H˜1(Ω).

Proof. The same arguments as in [10] yield ps= ˜ps∈L2(Ω). Let Qε: H1(Ωεs)→H1(Ω)

denote the continuous extension operator introduced in [7]. From (3.21), we deduce that, at least for some subsequence, ∃z ∈H1(Ω) such that

Qεpε* z in H1(Ω).

Noticing that χεsQεpε = χεspε in Ω, we deduce that that z = ˜ps in H1(Ω), which yields

ps = ˜ps ∈H1(Ω). (3.23)

Using (2.9)–(2.10) and noticing that Murat-Tartar inequality together with Korn’s inequality yield, for some constant C > 0 independent of ε:

|v|L2(Ωεf) ≤C|e(v)|L2(Ωεf), ∀v ∈H01(Ωεf) we find that there exists a constant C > 0 such that

|∇pε|H−1(Ωεf) ≤Cε.

(13)

Theorem 3.2 of [23] yields the existence of pf ∈ L2(Ω) such that, on some subsequence:

χεfpε * χ2 fpf. (3.24) To conclude, let ϕ∈ D(Ω), ψ ∈ Cper(Y)N such that

Z

Γ

ψ·ν 6= 0

and set

viε(x) = εϕ(x)ψix ε

, ∀i∈ {1,· · ·, N}.

Usingvε in the variational formulation (2.33)–(2.34) and passing to the limit as ε→0 yields

Z

(pf(x)−ps(x))ϕ(x)dx Z

Γ

ψ·ν

= 0, ∀ϕ∈ D(Ω) that is, ps=pf. We conclude thanks to (3.23).

4. The homogenized problem Consider the Hilbert space:

X ={u∈L2(Ω×Y), uf ∈L2(Ω,H˜per1 (Yf)), (˜us,u˜f)∈H0(div,Ω)2, divyu= 0}

endowed with the scalar product:

(u, v)X = Z

Ω×Ys

u v+

Z

div˜udiv˜v+ Z

Ω×Yf

ey(u)ey(v)+

Z

Ω×Γ

α(y)(γuγv−γννv)

and set:

X0 ={u∈X, div˜u= 0}, M =L20(Ω).

We can present our first homogenization result:

Proposition 4.1. The limit problem reads: Find (v, q)∈X×M such that a(v, ϕ) +b(q, ϕ) =

Z

gϕ,˜ ∀ϕ∈X (4.1)

b(π, v) = 0, ∀π ∈M (4.2)

(14)

where we set a(v, ϕ) =

Z

Ω×Ys

Avϕ+β Z

Ω×Yf

ey(v) :ey(ϕ) + Z

Ω×Γ

α(y)(γv−(γνv)ν)γϕ (4.3)

b(π, v) = − Z

πdiv˜v, (4.4)

Proof. Letϕ∈X∩ C(Ω×Y) and set ϕε(x) = ϕ

x,x ε

.

Then ϕε ∈ Vε. Passing to the limit in (2.33) with ϕε as the test function, we obtain (4.1). Notice that the integral on Ω×Γ results from the same arguments as in [10].

Let π∈M and set likewise:

πε(x) = π x,x

ε

.

Then, πε ∈L20(Ω) and we may pass to the limit in (2.34) withπε as the test function to obtain (4.2).

Proposition 4.2. The problem (4.1)–(4.2) is well-posed.

Proof. Let π∈M. Using Theorem 4.2 of [10], let v ∈H01(Ω)N ⊂X be the unique solution of π= divv such that |v|H1

0(Ω) ≤C|π|L2(Ω). Noticing that

|v|2X ≤C|v|2H1(Ω) ≤C|π|2L2(Ω)

we deduce that:

sup

ϕ∈X

b(π, ϕ)

|ϕ|X|π| ≥ b(π, v)

|v|X|π| = |π|2

|v|X|π| ≥ 1 C.

The coercivity of a(·,·) follows from the coercivity of the matrix A and that X may be equipped with the norm (2.24). We conclude as in Theorem 4.2 of [10] using Corollary 4.1, Ch. 1 of [11].

Proposition 4.3. The problem (4.1)–(4.2) equivalently reads: Find u∈X0 such that

a(u, ϕ) = Z

(g− ∇p) ˜ϕ, ∀ϕ∈X0.

(15)

Consider the cell problem: for every i∈ {1,· · ·, N}, let wi ∈W, where:

W ={w∈L2(Y), uf ∈H˜per1 (Yf), divyw= 0 in Y}, such that

Z

Ys

Awiψ+β Z

Yf

ey(wi) :ey(ψ) + Z

Γ

α(γwi−(γνwi)ν)ψ = Z

Y

ψi, ∀ψ ∈W.

Then, using the convention of repeated indices, the solution of the homoge- nized problem reads:

u(x, y) = wi(y)

gi(x)− ∂p

∂xi

. (4.5)

We set

Kji = Z

Y

wij. (4.6)

Proposition 4.4. The effective permeability tensorK is symmetric and pos- itively defined.

Proof. By definition of wi, the matrix K is symmetric. Introducing wξ = PN

i=1ξiwi for every ξ∈RN, we see that K is also positively definite. There- fore, the positivity is given by the lowest, necessarily positive, eigenvalue of K.

From (4.5) we get the Darcy’s law

˜

u=K(g− ∇p), (4.7)

where ˜u ∈H0(div,Ω) and p∈H˜1(Ω) is the unique solution of the following boundary value problem:

div(K∇p) = div(Kg) in Ω, (4.8)

K∇p·n=Kg·n on ∂Ω, (4.9)

where obviously div(Kg)∈H−1(Ω).

Proposition 4.5. The problem (4.8)–(4.9) is well-posed.

(16)

Proof. AsK is symmetric and positively defined, (4.8)–(4.9) is a clasical non homogeneous Neumann problem. The compatibility condition is obviously satisfied. Then there exists a unique solution in ˜H1(Ω).

Acknowledgements. This work has been accomplished during the visit of Florentina-Alina Stanescu at the I.R.M.A.R’s Department of Mechanics (University of Rennes 1); her contribution to this paper is supported by the Sectorial Operational Programme Human Resources Development (SOP HRD), financed from the European Social Fund and by the Romanian Gov- ernment under the contract number SOP HRD/107/1.5/S/82514.

References

[1] G. Allaire, Homogenization of the Stokes flow in a connected porous medium, Asymptotic Analysis 2 (1989) 203-222.

[2] G. Allaire, Homogenization of the Navier-Stokes Equations with a Slip Boundary Condition, Communications in Pure and Applied Mathematics 44 (1991) 605-641.

[3] G. Allaire, Homogenization and two-scale convergence, S.I.A.M. J. Math.

Anal. 23 (1992) 1482-151.

[4] G.I. Barenblatt, Y.P. Zheltov, I.N. Kochina, On basic conceptions of the theory of homogeneous fluids seepage in fractured rocks (in Russian), Prikl.Mat. i Mekh. 24 (1960) 852-864.

[5] G.I. Barenblatt, V.M. Entov, V.M. Ryzhik, Theory of Fluid Flows Through Natural Rocks, Kluwer Acad. Pub., Dordrecht, 1990.

[6] G.S. Beavers, D.D. Joseph, Boundary conditions at a naturally permeable wall, J. Fluid Mech. 30 (1967) 197–207.

[7] D. Cioranescu, J. Saint-Jean-Paulin, Homogenization in open sets with holes, J. Math. Anal. Appl. 71 (2)(1979) 590–607

[8] C. Conca, On the application of the homogenization theory to a class of problems arising in fluid mechanics, J. Math. Pures et Appl. 64(1985) 31–75

(17)

[9] H.I. Ene, D. Poliˇsevski, Model of diffusion in partially fissured media, Z.A.M.P. 53 (2002) 1052–1059.

[10] I. Gruais, D. Poliˇsevski, Fluid flows through fractured porous media along Beavers-Joseph interfaces Preprint I.R.M.A.R. 2013.

[11] V. Girault, P.-A. Raviart Finite Element Methods for Navier-Stokes Equations, Springer-Verlag, Berlin, 1986.

[12] U. Hornung, Homogenization and porous media,Interdiscip. Appl.

Math.6, Springer,New York(1997).

[13] W. J¨ager, A. Mikeli´c, Modeling effective interface laws for transport phenomena between an unconfined fluid and a porous medium using ho- mogenization, Transp. Porous Med. 78 (2009) 489–508.

[14] J.B. Keller, Darcy’s law for flow in porous media and the two-space method, Nonlinear partial differential equations in engineering and ap- plied science (Proc. Conf., Univ. Rhode Island, Kingston, R.I., 1979), Lecture Notes in Pure and Appl. Math. 54 429–443, Dekker,New York(1980)

[15] J.-L Lions, Some methods in the mathematical analysis of systems and their control, Kexue Chubanshe (Science Press), Beijing(1981)

[16] R. Lipton and M. Avellaneda, Darcy’s law for slow viscous flow past a stationary array of bubbles, Proc. Roy. Soc. Edinburgh Sect. A 114 (1-2)(1990) 71–79.

[17] I.P. Jones, Low Reynolds number flow past a porous spherical shell, Proc Camb. Phil. Soc. 73 (1973) 231–238.

[18] D. Lukkassen, G. Nguetseng, P. Wall, Two-scale convergence. Int. J.

Pure Appl. Math. 2 (2002) 35–86.

[19] A. Mikeli´c, Effets inertiels pour un ´ecoulement stationnaire visqueux incompressible dans un milieu poreux, C. R. Acad. Sci. Paris S´er. I Math.320 (10) (1995) 1289–1294.

[20] M. Neuss-Radu, Some extensions of two-scale convergence. C.R. Acad.

Sci. Paris, t. 322, Serie I (1996) 899-904.

(18)

[21] G. Nguetseng, A general convergence result for a functional related to the theory of homogenization, S.I.A.M. J.Math.Anal. 20 (1989) 608-623.

[22] D. Poliˇsevski, On the homogenization of fluid flows through periodic media, Rend. Sem. Mat. Univers. Politecn. Torino 45 (2) (1987) 129–139.

[23] D. Poliˇsevski, Basic homogenization results for a biconnected -periodic structure, Appl. Anal. 82 (4) (2003) 301–309.

[24] D. Poliˇsevski, The Regularized Diffusion in Partially Fractured Porous Media, in: L. Dragos (Ed.) Current Topics in Continuum Mechanics, Volume 2, Ed. Academiei, Bucharest, 2003.

[25] P.G. Saffman, On the boundary condition at the interface of a porous medium. Stud. Appl. Math. 50 (1971) 93–101.

[26] E. Sanchez-Palencia, Non-Homogeneous Media and Vibration Theory.

Lecture Notes in Physics, Vol. 127, Springer-Verlag, Berlin, 1980.

[27] R.E. Showalter, N.J. Walkington, Micro-structure models of diffusion in fissured media, J. Math. Anal. Appl. 155 (1991) 1-20.

[28] L. Tartar, Incompressible fluid flow in a porous medium. Convergence of the homogenization process. Appendix of [26].

Références

Documents relatifs

High order homogenization of the Stokes system in a periodic porous medium..

Finally, note that the idea of plugging an approximation of the solution into the relative entropy method should certainly apply to other degenerate particle systems and allow to

By using different methods, we show how to derive the correct mechanical route to the pressure of a fluid adsorbed in a random porous medium.. Some discussions are also made

Based on the obtained results a simple equation is presented for the equivalent permeability of a randomly cracked porous material as a function of the matrix permeability and the

As mentioned earlier, the great advantage of the effective stress concept is the simplification of the analysis by using one single variable σ ′ instead of two independent

Experimentally, it has long been known that a non linear relationship between the flow rate and the pressure drop imposed on the porous medium is obtained for non-Newtonian fluids

First, the 3D micro structures of the porous media are constructed with a parametric program which allow to control the pore sizes, pore orientations,

On the Understanding of the Macroscopic Inertial Effects in Porous Media: Investigation of the Microscopic Flow Structure.. 9th International Conference on Porous Media &