Loading [MathJax]/jax/output/SVG/jax.js

Homogenization of Bingham flow in thin porous media

  • Received: 01 March 2019
  • Primary: 76A05, 76A20; Secondary: 76M50, 35B27

  • By using dimension reduction and homogenization techniques, we study the steady flow of an incompresible viscoplastic Bingham fluid in a thin porous medium. A main feature of our study is the dependence of the yield stress of the Bingham fluid on the small parameters describing the geometry of the thin porous medium under consideration. Three different problems are obtained in the limit when the small parameter ε tends to zero, following the ratio between the height ε of the porous medium and the relative dimension aε of its periodically distributed pores. We conclude with the interpretation of these limit problems, which all preserve the nonlinear character of the flow.

    Citation: María Anguiano, Renata Bunoiu. Homogenization of Bingham flow in thin porous media[J]. Networks and Heterogeneous Media, 2020, 15(1): 87-110. doi: 10.3934/nhm.2020004

    Related Papers:

    [1] María Anguiano, Renata Bunoiu . Homogenization of Bingham flow in thin porous media. Networks and Heterogeneous Media, 2020, 15(1): 87-110. doi: 10.3934/nhm.2020004
    [2] María Anguiano, Francisco Javier Suárez-Grau . Newtonian fluid flow in a thin porous medium with non-homogeneous slip boundary conditions. Networks and Heterogeneous Media, 2019, 14(2): 289-316. doi: 10.3934/nhm.2019012
    [3] Leda Bucciantini, Angiolo Farina, Antonio Fasano . Flows in porous media with erosion of the solid matrix. Networks and Heterogeneous Media, 2010, 5(1): 63-95. doi: 10.3934/nhm.2010.5.63
    [4] Andro Mikelić, Giovanna Guidoboni, Sunčica Čanić . Fluid-structure interaction in a pre-stressed tube with thick elastic walls I: the stationary Stokes problem. Networks and Heterogeneous Media, 2007, 2(3): 397-423. doi: 10.3934/nhm.2007.2.397
    [5] Alexei Heintz, Andrey Piatnitski . Osmosis for non-electrolyte solvents in permeable periodic porous media. Networks and Heterogeneous Media, 2016, 11(3): 471-499. doi: 10.3934/nhm.2016005
    [6] Tasnim Fatima, Ekeoma Ijioma, Toshiyuki Ogawa, Adrian Muntean . Homogenization and dimension reduction of filtration combustion in heterogeneous thin layers. Networks and Heterogeneous Media, 2014, 9(4): 709-737. doi: 10.3934/nhm.2014.9.709
    [7] Michiel Bertsch, Carlo Nitsch . Groundwater flow in a fissurised porous stratum. Networks and Heterogeneous Media, 2010, 5(4): 765-782. doi: 10.3934/nhm.2010.5.765
    [8] Catherine Choquet, Ali Sili . Homogenization of a model of displacement with unbounded viscosity. Networks and Heterogeneous Media, 2009, 4(4): 649-666. doi: 10.3934/nhm.2009.4.649
    [9] Tom Freudenberg, Michael Eden . Homogenization and simulation of heat transfer through a thin grain layer. Networks and Heterogeneous Media, 2024, 19(2): 569-596. doi: 10.3934/nhm.2024025
    [10] Edoardo Mainini . On the signed porous medium flow. Networks and Heterogeneous Media, 2012, 7(3): 525-541. doi: 10.3934/nhm.2012.7.525
  • By using dimension reduction and homogenization techniques, we study the steady flow of an incompresible viscoplastic Bingham fluid in a thin porous medium. A main feature of our study is the dependence of the yield stress of the Bingham fluid on the small parameters describing the geometry of the thin porous medium under consideration. Three different problems are obtained in the limit when the small parameter ε tends to zero, following the ratio between the height ε of the porous medium and the relative dimension aε of its periodically distributed pores. We conclude with the interpretation of these limit problems, which all preserve the nonlinear character of the flow.



    In this paper we study the asymptotic behavior of the flow of a viscoplastic Bingham fluid in a thin porous medium which contains an array of bodies modelized as vertical cylindrical obstacles (the pores). We refer the reader to the very recent paper [8] and the references therein for the application of our study to problems issued from the real life applications. As a first example one can mention the flow of the volcanic lava through dense forests (see [30]). Another important application is the flow of fresh concrete spreading through networks of steel bars (see [33]).

    The model of thin porous medium of thickness much smaller than the distance between the pores was introduced in [34], where a stationary incompressible Navier-Stokes flow was studied. Recently, the model of thin porous medium under consideration in this paper was introduced in [20], where the flow of an incompressible viscous fluid described by the stationary Navier-Stokes equations was studied by the multiscale expansion method, which is a formal but powerful tool to analyse homogenization problems. These results were rigorously proved in [5] using an adaptation introduced in [4] of the periodic unfolding method from [16] and [17]. This adaptation consists of a combination of the unfolding method with a rescaling in the height variable, in order to work with a domain of fixed height, and to use monotonicity arguments to pass to the limit. In [4], in particular, the flow of an incompressible stationary Stokes system with a nonlinear viscosity, being a power law, was studied. For non-stationary incompressible viscous flow in a thin porous medium see [1], where a non-stationary Stokes system is considered, and [2], where a non-stationary non-newtonian Stokes system, where the viscosity obeyed the power law, is studied. For the periodic unfolding method applied to the study of problems stated in other type of thin periodic domains we refer for instance to [23] for crane type structures and to [24], [25] for thin layers with thin beams structures, where elasticity problems are considered. In [32], the homogenization of elasticity problems in thin periodic domains of planar grids type is studied. For problems involving arrays of bodies in high-contrast materials we refer the reader to [6], Chapter 2.

    If Π is a three-dimensional domain with smooth boundary Π and f=(f1,f2,f3) are external given forces defined on Π, then the velocity u=(u1,u2,u3) of a fluid and its pressure p satisfy the equations of motion

    3j=1xi(σ(p,u))ij=fiinΠ,1i3, (1)

    completed with fluid's incompressibility condition divu=3i=1xiui=0 in Π, and the no-slip boundary condition u=0 on the boundary Π. What distinguishes different fluids is the expression of the stress tensor σ. Newtonian fluids are the most encountered ones in real life and as typical examples one can mention the water and the air. For a newtonian fluid, the entries of the stress tensor σ(p,u) are given by

    (σ(p,u))ij=pδij+2μ(D(u))ij,1i,j3 (2)

    where δij is the Kronecker symbol, the real positive μ is the viscosity of the fluid and the entries of the strain tensor are (D(u))ij=(xjui+xiuj)/2. If f belongs to (L2(Π))3 and the space V is defined by V={v(H10(Π))3|divv=0}, then u and p satisfying (1) with (2) are such that (see for instance [22]):

    (Stokes) There is a unique uV and a unique (up to an additive real constant) pL2(Π) such that (if <,> is the dual pairing between (H1(Π))3 and (H10(Π))3)

    a(u,v)=l(v)<p,v>,v(H10(Π))3, (3)

    with a(u,v)=2μΠD(u):D(v)dx and l(v)=Πfvdx.

    A fluid whose stress is not defined by relation (2) is called a non-newtonian fluid. There are several classes of non-newtonian fluids, as the power law, Carreau, Cross, Bingham fluids. It is on the study of the last type of fluid that we are interested in this paper. We refer to [18] for a review on non-newtonian fluids. For a Bingham fluid, the nonlinear stress tensor is defined by (see [19])

    (σ(p,u))ij=pδij+2μ(D(u))ij+2g(D(u))ij|D(u)|, (4)

    where |D(u)|2=D(u):D(u)0 and the positive number g represents the yield stress of the fluid. If g=0, then (4) becomes (2). Viscoplastic Bingham fluids are quite often encountered in real life. As examples one can mention volcanic lava, fresh concrete, the drilling mud, oils, clays and some paintings. For ug and pg satisfying (1) with (4), according to [19], one has the following result:

    (Bingham) There is a unique ugV and a (non-unique) pgL2(Π)/R such that

    a(ug,vug)+j(v)j(ug)l(vug)<pg,vug>,v(H10(Π))3. (5)

    Here a,l,<,> are as before and

    j(v)=2gΠ|D(v)|dx,v(H10(Π))3.

    If the yield stress of the Bingham fluid is of the form g(ε), with ε]0,1[ and such that g(ε) tends to zero when ε tends to zero, then, according to [[19], Chapter 6, Théorème 5.1.], the following result holds

    When ε tends to zero, one has for the solution uε of problem (5) corresponding to g(ε) the following convergence

    uεuweaklyinV,

    where u is the solution of problem (3).

    This means that, in a fixed domain, the nonlinear character of the Bingham flow is lost in the limit when the yield stress tends to zero, as it is expected. A natural question that arises is the following: If the yield stress g(ε) is as before and, moreover, the domain Π itself depends on the small parameter ε, what happens when ε tends to zero? The answer is that, in the limit, the nonlinear character of the flow may be preserved. For instance, if Πε is a classical rigid porous medium, it was proven in [29] with the asymptotic expansion method that, in a range of parameters, the nonlinear character of the Bingham flow is preserved in the homogenized problem, which is a nonlinear Darcy equation. The convergence corresponding to the above mentioned result was proven in [10] with the two-scale convergence method and then recovered in [12] with the periodic unfolding method. The case of a doubly periodic rigid porous medium was studied in [11], where a more involved nonlinear Darcy equation is derived. Another class of domains for which the nonlinear character of the flow may be preserved in the limit is those of thin domains. The case of a domain Πε which is thin in one direction was addressed in [14] and [15]. We refer to [13] for the asymptotic analysis of a Bingham fluid in a thin T-like shaped domain. In all these cases, a lower-dimensional Bingham-like law was exhibited in the limit. This law was already encountered in engineering (see [31]), but no rigurous mathematical justification was previously known. A first mathematical result combining both periodic and thin domains for the Bingham flow was announced in [3]. For the shallow flow of a viscoplastic fluid we refer the reader to [21], [9], [26], [27] and [28]. For a homogenized non-newtonian viscoelastic model we refer to [7].

    The paper is organized as follows. In Section 2. we state the problem: we define in (6) the thin porous medium Ωε (see also Figure 1), of height ε and relative dimension aε of its periodically distributed pores. In Ωε we consider the flow of a viscoplastic Bingham fluid with velocity uε and pressure pε verifying the nonlinear variational inequality (9). In Section 3. we give some a priori estimates for the velocity and for the pressure obtained after the change of variables (10) and verifying (12), and then for the velocity and for the pressure defined in (21). In Section 4. by passing to the limit ε0, we prove the main convergence results of our paper, stated in Theorems 4.2, 4.4 and 4.6, respectively. Up to our knowledge, problems (37), (59) and (81) are new in the mathematical literature. We conclude in Section 5. with the interpretation of these limit problems, which all three preserve the nonlinear character of the flow; both effects of a nonlinear Darcy equation and a lower dimensional Bingham-like law appear. The paper ends with a list of References.

    Figure 1.  View of the domain Ωε.

    A periodic porous medium is defined by a domain ω and an associated microstructure, or periodic cell Y=[1/2,1/2]2, which is made of two complementary parts: the fluid part Yf, and the solid part Ys (YfYs=Y and YfYs=). More precisely, we assume that ω is a smooth, bounded, connected set in R2, and that Ys is an open connected subset of Y with a smooth boundary Ys, such that ¯Ys is strictly included in Y.

    The microscale of a porous medium is a small positive number aε. The domain ω is covered by a regular mesh of size aε: for kZ2, each cell Yk,aε=aεk+aεY is divided in a fluid part Yfk,aε and a solid part Ysk,aε, i.e. is similar to the unit cell Y rescaled to size aε. We define Y=Y×(0,1)R3, which is divided in a fluid part Yf and a solid part Ys, and consequently Yk,aε=Yk,aε×(0,1)R3, which is also divided in a fluid part Yfk,aε and a solid part Ysk,aε.

    We denote by τ(¯Ysk,aε) the set of all translated images of ¯Ysk,aε. The set τ(¯Ysk,aε) represents the solids in R2. The fluid part of the bottom ωεR2 of the porous medium is defined by ωε=ωkKε¯Ysk,aε, where Kε={kZ2:Yk,aεω}. The whole fluid part ΩεR3 in the thin porous medium is defined by

    Ωε={(x1,x2,x3)ωε×R:0<x3<ε}. (6)

    We make the assumption that the solids τ(¯Ysk,aε) do not intersect the boundary ω. We define Yεsk,aε=Ysk,aε×(0,ε). Denote by Sε the set of the solids contained in Ωε. Then, Sε is a finite union of solids, i.e. Sε=kKε¯Yεsk,aε.

    We define ˜Ωε=ωε×(0,1), Ω=ω×(0,1), and Qε=ω×(0,ε). We observe that ˜Ωε=ΩkKε¯Ysk,aε, and we define Tε=kKε¯Ysk,aε as the set of the solids contained in ˜Ωε.

    We denote by : the full contraction of two matrices; for A=(ai,j)1i,j3 and B=(bi,j)1i,j3, we have A:B=3i,j=1aijbij.

    In order to apply the unfolding method, we will need the following notation. For kZ2, we define κ:R2Z2 by

    κ(x)=kxYk,1. (7)

    Remark that κ is well defined up to a set of zero measure in R2 (the set kZ2Yk,1). Moreover, for every aε>0, we have

    κ(xaε)=kxYk,aε.

    We denote by C a generic positive constant which can change from line to line.

    The points xR3 will be decomposed as x=(x,x3) with x=(x1,x2)R2, x3R. We also use the notation x to denote a generic vector of R2.

    In Ωε we consider the stationary flow of an incompressible Bingham fluid. As already seen in the Introduction, following Duvaut and Lions [19], the problem is formulated in terms of a variational inequality.

    For a vectorial function v=(v,v3), we define

    (D(v))i,j=12(xjvi+xivj),1i,j3,|D(v)|2=D(v):D(v).

    We introduce the following spaces

    V(Ωε)={v(H10(Ωε))3|divv=0 in Ωε},
    H(Ωε)={v(L2(Ωε))3|divv=0 in Ωε,vn=0 on Ωε}.

    For u,v(H10(Ωε))3, we introduce

    a(u,v)=2μΩεD(u):D(v)dx,j(v)=2g(ε)Ωε|D(v)|dx,(u,v)Ωε=Ωεuvdx,

    where the yield stress g(ε) will be made precise in Section 3.1. Let f(L2(Ω))3 be given such that f=(f,0). Let fε(L2(Ωε))3 be defined by

    fε(x)=f(x,x3/ε), a.e. xΩε.

    The model of the flow is described by the following variational inequality:

    Find uεV(Ωε) such that

    a(uε,vuε)+j(v)j(uε)(fε,vuε)Ωε,vV(Ωε). (8)

    From Duvaut and Lions [19], we know that there exists a unique uεV(Ωε) solution of problem (8). Moreover, from Bourgeat and Mikelić [10], we know that if pε is the pressure of the fluid in Ωε, then problem (8) is equivalent to the following one: Find uεV(Ωε) and pεL20(Ωε) such that

    a(uε,vuε)+j(v)j(uε)(fε,vuε)Ωε+(pε,div(vuε))Ωε,v(H10(Ωε))3. (9)

    Problem (9) admits a unique solution uεV(Ωε) and a (non) unique solution pεL20(Ωε), where L20(Ωε) denotes the space of functions belonging to L2(Ωε) and of mean value zero.

    Our aim is to study the asymptotic behavior of uε and pε when ε tends to zero. For this purpose, we first use the dilatation of the domain Ωε in the variable x3, namely

    y3=x3ε, (10)

    in order to have the functions defined in an open set with fixed height, denoted ˜Ωε.

    Namely, we define ˜uε(H10(˜Ωε))3, ˜pεL20(˜Ωε) by

    ˜uε(x,y3)=uε(x,εy3),˜pε(x,y3)=pε(x,εy3) a.e. (x,y3)˜Ωε.

    Let us introduce some notation which will be useful in the following. For a vectorial function v=(v,v3) and a scalar function w, we will denote Dx[v]=12(Dxv+Dtxv) and y3[v]=12(y3v+ty3v), where we denote y3=(0,0,y3)t. Moreover, associated to the change of variables (10), we introduce the operators: Dε, Dε, divε and ε, defined by

    (Dεv)i,j=xjvi for i=1,2,3, j=1,2,(Dεv)i,3=1εy3vi for i=1,2,3,
    Dε[v]=12(Dεv+Dtεv),|Dε[v]|2=Dε[v]:Dε[v],
    divεv=divxv+1εy3v3,εw=(xw,1εy3w)t.

    We introduce the following spaces

    V(˜Ωε)={˜v(H10(˜Ωε))3|divε˜v=0 in ˜Ωε},
    H(˜Ωε)={˜v(L2(˜Ωε))3|divε˜v=0 in ˜Ωε,˜vn=0 on ˜Ωε}.

    For ˜u,˜vV(˜Ωε), we introduce

    aε(˜u,˜v)=2μ˜ΩεDε[˜u]:Dε[˜v]dxdy3,jε(˜v)=2g(ε)˜Ωε|Dε[˜v]|dxdy3,

    and

    (˜u,˜v)˜Ωε=˜Ωε˜u˜vdxdy3.

    Using the transformation (10), the variational inequality (8) can be rewritten as:

    Find ˜uεV(˜Ωε) such that

    aε(˜uε,˜v˜uε)+jε(˜v)jε(˜uε)(f,˜v˜uε)˜Ωε,˜vV(˜Ωε), (11)

    and (9) can be rewritten as:

    Find ˜uεV(˜Ωε) and ˜pεL20(˜Ωε) such that

    aε(˜uε,˜v˜uε)+jε(˜v)jε(˜uε)(f,˜v˜uε)˜Ωε+(˜pε,divε(˜v˜uε))˜Ωε,˜v(H10(˜Ωε))3. (12)

    Our goal now is to describe the asymptotic behavior of this new sequence (˜uε, ˜pε).

    We start by obtaining some a priori estimates for ˜uε.

    Lemma 3.1. There exists a constant C independent of ε, such that if ˜uε(H10(˜Ωε))3 is the solution of problem (11), one has

    i) if aεε, with aε/ελ, 0<λ<+, or aεε, then

    ˜uε(L2(˜Ωε))3Cμa2ε,Dε[˜uε](L2(˜Ωε))3×3Cμaε,Dε˜uε(L2(˜Ωε))3×3Cμaε, (13)

    ii) if aεε, then

    ˜uε(L2(˜Ωε))3Cμε2,Dε[˜uε](L2(˜Ωε))3×3Cμε,Dε˜uε(L2(˜Ωε))3×3Cμε. (14)

    Proof. Setting successively ˜v=2˜uε and ˜v=0 in (11), we have

    2μ˜ΩεDε[˜uε]:Dε[˜uε]dxdy3+2g(ε)˜Ωε|Dε[˜uε]|dxdy3=˜Ωεf˜uεdxdy3. (15)

    Using Cauchy-Schwarz's inequality and the assumption of f, we obtain that

    ˜Ωεf˜uεdxdy3C˜uε(L2(˜Ωε))3,

    and taking into account that ˜Ωε|Dε[˜uε]|dxdy30, by (15), we have

    Dε[˜uε]2(L2(˜Ωε))3×3Cμ˜uε(L2(˜Ωε))3.

    For the cases aεε or aεε, taking into account Remark 4.3(ⅰ) in [4], we obtain the second estimate in (13), and, consequently, from classical Korn's inequality we obtain the last estimate in (13). Now, from the second estimate in (13) and Remark 4.3(ⅰ) in [4], we deduce the first estimate in (13). For the case aεε, proceeding similarly with Remark 4.3(ⅱ) in [4], we obtain the desired result.

    We extend the velocity ˜uε by zero to the Ω˜Ωε and denote the extension by the same symbol. Obviously, estimates (13)-(14) remain valid and the extension is divergence free too.

    We study in the sequel the following cases for the value of the yield stress g(ε):

    ⅰ) if aεε, with aε/ελ, 0<λ<+, or aεε, then g(ε)=gaε,

    ⅱ) if aεε, then g(ε)=gε,

    where g is a positive number. These choices are the most challenging ones and they answer to the question adressed in the paper, namely they all preserve in the limit the nonlinear character of the flow.

    In order to extend the pressure to the whole domain Ω, the mapping Rε (defined in Lemma 4.5 in [4] as Rε2) allows us to extend the pressure pε to Qε by introducing Fε in (H1(Qε))3:

    Fε,wQε=pε,RεwΩε, for any w(H10(Qε))3. (16)

    Setting succesively v=uε+Rεw and v=uεRεw in (9) we get the inequality

    |Fε,wQε||a(uε,Rεw)|+|(fε,Rεw)Ωε|+j(Rεw). (17)

    Moreover, if divw=0 then Fε,wQε=0, and the DeRham Theorem gives the existence of Pε in L20(Qε) with Fε=Pε.

    Using the change of variables (10), we get for any ˜w(H10(Ω))3 where ˜w(x,y3)=w(x,εy3),

    ε˜Pε,˜wΩ=Ω˜Pεdivε˜wdxdy3=ε1QεPεdivwdx=ε1Pε,wQε.

    Then, using the identification (16) of Fε and the inequality (17),

    |ε˜Pε,˜wΩ|ε1(|a(uε,Rεw)|+|(fε,Rεw)Ωε|+j(Rεw)).

    and applying the change of variables (10),

    |ε˜Pε,˜wΩ||aε(˜uε,˜Rε˜w)|+|(f,˜Rε˜w)˜Ωε|+jε(˜Rε˜w), (18)

    where ˜Rε˜w=Rεw for any ˜w(H10(Ω))3.

    Now, we estimate the right-hand side of (18) using the estimates given in Lemma 4.6 in [4].

    Lemma 3.2. There exists a constant C independent of ε, such that the extension ˜PεL20(Ω) of the pressure ˜pε satisfies

    ˜PεL20(Ω)C. (19)

    Proof. Let us estimate ε˜Pε in the cases aεε or aεε. We estimate the right-hand side of (18). Using Cauchy-Schwarz's inequality and from the second estimate in (13) we have

    |aε(˜uε,˜Rε˜w)|2μDε[˜uε](L2(˜Ωε))3×3Dε˜Rε˜w(L2(˜Ωε))3×3CaεDε˜Rε˜w(L2(˜Ωε))3×3.

    Using the assumption made on the function f, we obtain

    |(f,˜Rε˜w)˜Ωε|C˜Rε˜w(L2(˜Ωε))3,

    and by Cauchy-Schwarz's inequality and taking into account that |˜Ωε||Ω|, we obtain

    jε(˜Rε˜w)CaεDε˜Rε˜w(L2(˜Ωε))3×3.

    Then, from (18), we deduce

    |ε˜Pε,˜wΩ|CaεDε˜Rε˜w(L2(˜Ωε))3×3+C˜Rε˜w(L2(˜Ωε))3.

    Taking into account the third point in Lemma 4.6 in [4], we have

    |ε˜Pε,˜wΩ|Caε(1aε˜w(L2(Ω))3+Dε˜w(L2(Ω))3×3)+C(˜w(L2(Ω))3+aεDε˜w(L2(Ω))3×3).

    If aεε we take into account that aε1, and if aεε we take into account that aε/ε1 and aε1, and we see that there exists a positive constant C such that

    |ε˜Pε,˜wΩ|C˜w(H10(Ω))3,˜w(H10(Ω))3,

    and consequently

    ε˜Pε(H1(Ω))3C.

    It follows that (see for instance Girault and Raviart [22], Chapter I, Corollary 2.1) there exists a representative of ˜PεL20(Ω) such that

    ˜PεL20(Ω)C˜Pε(H1(Ω))3Cε˜Pε(H1(Ω))3C.

    Finally, let us estimate ε˜Pε in the case aεε. Similarly to the previous case, we estimate the right side of (18) by using Cauchy-Schwarz's inequality and from the second estimate in (14), and we have

    |ε˜Pε,˜wΩ|CεDε˜Rε˜w(L2(˜Ωε))3×3+C˜Rε˜w(L2(˜Ωε))3.

    Taking into account the proof in Lemma 4.5 in [4], the change of variables (10) and that aεε, we can deduce

    |ε˜Pε,˜wΩ|Cε(1ε˜w(L2(Ω))3+1εDx˜w(L2(Ω))3×2+1εy3˜w(L2(Ω))3)+C(˜w(L2(Ω))3+aεDx˜w(L2(Ω))3×2+y3˜w(L2(Ω))3),

    and using that aε1, we see that there exists a positive constant C such that

    |ε˜Pε,˜wΩ|C˜w(H10(Ω))3,˜w(H10(Ω))3,

    and reasing as the previous case, we have the estimate (19).

    According to these extensions, problem (12) can be written as:

    2μΩDε[˜uε]:Dε[˜v˜uε]dxdy3+2g(ε)Ω|Dε[˜v]|dxdy32g(ε)Ω|Dε[˜uε]|dxdy3Ωf(˜v˜uε)dxdy3+Ω˜Pεdivε(˜v˜uε)dxdy3, (20)

    for every ˜v that is the extension by zero to the whole Ω of a function in (H10(˜Ωε))3.

    The change of variable (10) does not provide the information we need about the behavior of ˜uε in the microstructure associated to ˜Ωε. To solve this difficulty, we use an adaptation introduced in [4] of the unfolding method from [16] and [17].

    Let us recall this adaptation of the unfolding method in which we divide the domain Ω in cubes of lateral length aε and vertical length 1. For this purpose, given (˜uε,˜Pε)(H10(Ω))3×L20(Ω), we define (ˆuε,ˆPε) by

    ˆuε(x,y)=˜uε(aεκ(xaε)+aεy,y3),ˆPε(x,y)=˜Pε(aεκ(xaε)+aεy,y3), (21)

    a.e. (x,y)ω×Y, where the function κ is defined in (7).

    Remark 1. For kKε, the restriction of (ˆuε,ˆPε) to Yk,aε×Y does not depend on x, whereas as a function of y it is obtained from (˜uε,˜Pε) by using the change of variables y=xaεkaε, which transforms Yk,aε into Y.

    We are now in position to obtain estimates for the sequences (ˆuε,ˆPε), as in the proof of Lemma 4.9 in [4].

    Lemma 3.3. There exists a constant C independent of ε, such that the couple (ˆuε,ˆPε) defined by (21) satisfies

    i) if aεε, with aε/ελ, 0<λ<+, or aεε,

    ˆuε(L2(ω×Y))3Ca2ε,Dy[ˆuε](L2(ω×Y))3×2Ca2ε,y3[ˆuε](L2(ω×Y))3Cεaε,

    ii) if aεε,

    ˆuε(L2(ω×Y))3Cε2,Dy[ˆuε](L2(ω×Y))3×2Caεε,y3[ˆuε](L2(ω×Y))3Cε2,

    and, moreover, in every cases,

    ˆPεL20(ω×Y)C.

    When ε tends to zero, we obtain for problem (20) different behaviors, depending on the magnitude of aε with respect to ε. We will analyze them in the next sections.

    First, we obtain some compactness results about the behavior of the sequences (˜uε,˜Pε) and (ˆuε,ˆPε) satisfying the a priori estimates given in Lemmas 3.1-ⅰ) and 3.3-ⅰ), respectively.

    Lemma 4.1 (Critical case). For a subsequence of ε still denote by ε, there exist ˜uH1(0,1;L2(ω)3), where ˜u3=0 and ˜u=0 on y3={0,1}, ˆuL2(ω;H1(Y)3) ("" denotes Y-periodicity), with ˆu=0 on ω×Ys and ˆu=0 on y3={0,1} such that Yˆu(x,y)dy=10˜u(x,y3)dy3 with Yˆu3dy=0, and ˆPL20(ω×Y), independent of y, such that

    ˜uεa2ε(˜u,0)inH1(0,1;L2(ω)3), (22)
    ˆuεa2εˆuinL2(ω;H1(Y)3),ˆPεˆPinL20(ω×Y), (23)
    divx(10˜u(x,y3)dy3)=0inω,(10˜u(x,y3)dy3)n=0onω, (24)
    divλˆu=0inω×Y,divx(Yˆu(x,y)dy)=0inω, (25)
    (Yˆu(x,y)dy)n=0onω, (26)

    where divλ=divy+λy3.

    Proof. We refer the reader to Lemmas 5.2, 5.3 and 5.4 in [4] for the proof of (22)-(26). Here, we prove that ˆP does not depend on the microscopic variable y. To do this, we choose as test function ˜v(x,y)D(ω;C(Y)3) with ˜v(x,y)=0ω×Ys (thus, ˜v(x,x/aε,y3)(H10(˜Ωε))3). Setting aε˜v(x,x/aε,y3) in (20) (we recall that g(ε)=gaε)) and using that divε˜uε=0, we have

    2μaεΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μΩ|Dε[˜uε]|2dxdy3+2ga2εΩ|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32gaεΩ|Dε[˜uε]|dxdy3aεΩf˜vdxdy3Ωf˜uεdxdy3+aεΩ˜Pεdivx˜vdxdy3+Ω˜Pεdivy˜vdxdy3+aεεΩ˜Pεy3˜v3dxdy3. (27)

    By the change of variables given in Remark 1 and by Lemma 3.3, we get for the first term in relation (27)

    ΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy3=ω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy+Oε, (28)

    and for the second term in relation (27)

    Ω|Dε[˜uε]|2dxdy3=ω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|2dxdy=Oε. (29)

    Moreover, applying the change of variables given in Remark 1 to the fourth term in relation (27), we have

    Ω|Dε[˜uε]|dxdy3=ω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy. (30)

    Therefore, applying the change of variables given in Remark 1 to relation (27), we obtain

    2μaεω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy+2ga2εω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gaεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεaεω×Yf˜vdxdyω×Yfˆuεdxdy+aεω×YˆPεdivx˜vdxdy+ω×YˆPεdivy˜vdxdy+aεεω×YˆPεy3˜v3dxdy+Oε. (31)

    According with (23), the first term in relation (31) can be written by the following way

    2μaεω×Y(1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]):(Dy[˜v]+aεεy3[˜v])dxdy0, as ε0. (32)

    In order to pass to the limit in the first nonlinear term, we have

    2gaεω×Y|aεDx[˜v]+Dy[˜v]+aεεy3[˜v]|dxdy0, as ε0. (33)

    Now, in order to pass the limit in the second nonlinear term, we are taking into account that

    aεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy=a2εω×Y|1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]|dxdy,

    and using (23) and the fact that the function E(φ)=|φ| is proper convex continuous, we can deduce that

    lim infε02gaεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy0. (34)

    Moreover, using (23) the two first terms in the right hand side of (31) can be written by

    aεω×Yf˜vdxdya2εω×Yfˆuεa2εdxdy0, as ε0. (35)

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (23), passing to the limit when ε tends to zero, we have

    ω×YˆPdivλ˜vdxdy. (36)

    Therefore, taking into account (32)-(36), when we pass to the limit in (31) when ε tends to zero, we have 0ω×YˆPdivλ˜vdxdy. Now, if we choose as test function aε˜v(x,x/aε,y3) in (20) and we argue similarly, we obtain ω×YˆPdivλ˜vdxdy0. Thus, we can deduce that ω×YˆPdivλ˜vdxdy=0, which shows that ˆP does not depend on y.

    Theorem 4.2 (Critical case). If aεε, with aε/ελ, 0<λ<+, then (ˆuε/a2ε,ˆPε) converges to (ˆu,ˆP) in L2(ω;H1(Y)3)×L20(ω×Y), which satisfies the following variational inequality

    2μω×YDλ[ˆu]:(Dλ[˜v]Dλ[ˆu])dxdy+2gω×Y|Dλ[˜v]|dxdy2gω×Y|Dλ[ˆu]|dxdyω×Yf(˜vˆu)dxdyω×YxˆP(˜vˆu)dxdy, (37)

    where Dλ[]=Dy[]+λy3[] and for every ˜vL2(ω;H1(Y)3) such that

    ˜v(x,y)=0inω×Ys,divλ˜v=0inω×Y,(Y˜v(x,y)dy)n=0onω.

    Proof. We choose a test function ˜v(x,y)D(ω;C(Y)3) with ˜v(x,y)=0ω×Ys (thus, we have that ˜v(x,x/aε,y3)(H10(˜Ωε))3). We first multiply (20) by a2ε and we use that divε˜uε=0. Then, we take as test function a2ε˜vε=a2ε(˜v(x,x/aε,y3),λε/aεv3(x,x/aε,y3)), with ˜v(x,y)=0 in ω×Ys and satisfying the incompressibility conditions (25)-(26), that is, divλ˜v=0 in ω×Y and (Y˜v(x,y)dy)n=0 on ω, and we have

    2μΩDε[˜uε]:(Dx[˜vε]+1aεDy[˜vε]+1εy3[˜vε])dxdy32μ1a2εΩ|Dε[˜uε]|2dxdy3+2gaεΩ|Dx[˜vε]+1aεDy[˜vε]+1εy3[˜vε]|dxdy32g1aεΩ|Dε[˜uε]|dxdy3Ωf˜vdxdy31a2εΩf˜uεdxdy3+Ω˜Pεdivx˜vdxdy3+1aεΩ˜Pεdivy˜vdxdy3+λaεΩ˜Pεy3˜v3dxdy3. (38)

    By the change of variables given in Remark 1 and by Lemma 3.3, we have (28) for the first term in relation (38), and for the second term in relation (38) we obtain

    Ω|Dε[˜uε]|2dxdy3=ω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|2dxdy. (39)

    Moreover, applying the change of variables given in Remark 1 to the fourth term in relation (38), we have (30). Therefore, applying the change of variables given in Remark 1 to relation (38), we obtain

    2μω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜vε]+1εy3[˜vε])dxdy2μ1a2εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|2dxdy+2gaεω×Y|Dx[˜vε]+1aεDy[˜vε]+1εy3[˜vε]|dxdy2g1aεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεω×Yf˜vdxdy1a2εω×Yfˆuεdxdy+ω×YˆPεdivx˜vdxdy+1aεω×YˆPεdivy˜vdxdy+λaεω×YˆPεy3˜v3dxdy+Oε. (40)

    According with (23), the first term in relation (40) can be written

    2μω×Y(1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]):(Dy[˜vε]+aεεy3[˜vε])dxdy,

    and, taking into account that λε/aε1, this term tends to the following limit

    2μω×Y(Dy[ˆu]+λy3[ˆu]):(Dy[˜v]+λy3[˜v])dxdy. (41)

    The second term in relation (40) writes

    2μω×Y(1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]):(1a2εDy[ˆuε]+aεε1a2εy3[ˆuε])dxdy,

    and, taking into account that the function B(φ)=|φ| is proper convex continuous and λε/aε1, we get that the lim infε0 of this second is greater or equal than

    2μω×Y(Dy[ˆu]+λy3[ˆu]):(Dy[ˆu]+λy3[ˆu])dxdy. (42)

    In order to pass to the limit in the first nonlinear term, we have

    |aεω×Y|Dx[˜vε]+1aεDy[˜vε]+1εy3[˜vε]|dxdyω×Y|Dy[˜v]+λy3[˜v]|dxdy|ω×Y|aεDx[˜vε]+Dy[˜vε]+aεεy3[˜vε]Dy[˜v]λy3[˜v]|dxdyω×Y|aεDx[˜vε]|dxdy+ω×Y|Dy[˜vε]Dy[˜v]|dxdy+ω×Y|aεεy3[˜vε]λy3[˜v]|dxdy0, as ε0,

    and we can deduce that the first nonlinear term tends to the following limit

    2gω×Y|Dy[˜v]+λy3[˜v]|dxdy. (43)

    Now, in order to pass the limit in the second nonlinear term, we are taking into account that

    1aεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy=ω×Y|1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]|dxdy,

    and using (23) and the fact that the function E(φ)=|φ| is proper convex continuous, we can deduce that

    lim infε02g1aεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy2gω×Y|Dy[ˆu]+λy3[ˆu]|dxdy. (44)

    Moreover, using (23) the two first terms in the right hand side of (40) tend to the following limit

    ω×Yf(˜vˆu)dxdy. (45)

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (23) the first term of the pressure tends to the following limit ω×YˆPdivx˜vdxdy, and using (25) and taking into account that ˆP does not depend on y, we have

    ω×YˆPdivx˜vdxdy=ω×YˆPdivx˜vdxdyωˆP(divxYˆudy)dx=ω×YxˆP(˜vˆu)dxdy. (46)

    Finally, using that divλ˜v=0, we have

    1aεω×YˆPεdivy˜vdxdy+λaεω×YˆPεy3˜v3dxdy=0. (47)

    Therefore, taking into account (41)-(47), we have (37).

    We obtain some compactness results about the behavior of the sequences (˜uε,˜Pε) and (ˆuε,ˆPε) satisfying the a priori estimates given in Lemmas 3.1-ⅰ) and 3.3-ⅰ), respectively.

    Lemma 4.3 (Subcritical case). For a subsequence of ε still denoted by ε, there exist ˜u(L2(Ω))3, where ˜u3=0 and ˜u=0 on y3={0,1}, ˆuL2(Ω;H1(Y)3) ("" denotes Y-periodicity), with ˆu=0 in ω×Ys and ˆu=0 on y3={0,1} such that Yˆu(x,y)dy=10˜u(x,y3)dy3 with Yˆu3dy=0 and ˆu3 independent of y3, and ˆPL20(ω×Y), independent of y, such that

    ˜uεa2ε(˜u,0)in(L2(Ω))3, (48)
    ˆuεa2εˆuinL2(Ω;H1(Y)3),ˆPεˆPinL20(ω×Y), (49)
    divx(10˜u(x,y3)dy3)=0inω,(10˜u(x,y3)dy3)n=0onω, (50)
    divyˆu=0inω×Y,divx(Yˆu(x,y)dy)=0inω, (51)
    (Yˆu(x,y)dy)n=0onω. (52)

    Proof. See Lemmas 5.2, 5.3 and 5.4 in [4] for the proof of (48)-(52). In order to prove that ˆP does not depend on y we argue as in the proof of Lemma 4.1 using that aεε, and we obtain ω×YˆPdivy˜vdxdy=0, which shows that ˆP does not depend on y. Now, in order to prove that ˆP does not depend on y3, setting ε˜v=ε(0,˜v3(x,x/aε,y3)) in (20) (we recall that g(ε)=gaε)) and using that divε˜uε=0, we have

    2μεΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μΩ|Dε[˜uε]|2dxdy3+2gaεεΩ|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32gaεΩ|Dε[˜uε]|dxdy3Ωf˜uεdxdy3+Ω˜Pεy3˜v3dxdy3. (53)

    Applying the change of variables given in Remark 1 to relation (53) and taking into account (28)-(30), we obtain

    2μεω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy+2gaεεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gaεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεω×Yfˆuεdxdy+ω×YˆPεy3˜v3dxdy+Oε. (54)

    According with (49) and using that aεε, the first term in relation (54) can be written by the following way

    2μεω×Y(1a2εDy[ˆuε]+aεε1a2εy3[ˆuε]):(Dy[˜v]+aεεy3[˜v])dxdy0, as ε0. (55)

    In order to pass to the limit in the first nonlinear term, we have

    2gεω×Y|aεDx[˜v]+Dy[˜v]+aεεy3[˜v]|dxdy0, as ε0. (56)

    In order to pass to the limit in the second nonlinear term, we proceed as in Lemma 4.1. Moreover, using (49) the first term in the right hand side of (54) can be written by

    a2εω×Yfˆuεa2εdxdy0, as ε0. (57)

    We consider now the term which involves the pressure. Taking into account the convergence of the pressure (49), passing to the limit when ε tends to zero, we have

    ω×YˆPy3˜v3dxdy. (58)

    Therefore, taking into account (34) and (55)-(58), when we pass to the limit in (54) when ε tends to zero, we have 0ω×YˆPy3˜v3dxdy. Now, if we choose as test function ε˜v=ε(0,˜v3(x,x/aε,y3)) in (20) and we argue similarly, we can deduce that ˆP does not depend on y3, so ˆP does not depend on y.

    Theorem 4.4 (Subcritical case). If aεε, then (ˆuε/a2ε,ˆPε) converges to (ˆu,ˆP) in L2(Ω;H1(Y)3)×L20(ω×Y), which satisfies the following variational inequality

    2μω×YDy[ˆu]:(Dy[˜v]Dy[ˆu])dxdy+2gω×Y|Dy[˜v]|dxdy2gω×Y|Dy[ˆu]|dxdyω×Yf(˜vˆu)dxdyω×YxˆP(˜vˆu)dxdy, (59)

    for every ˜vL2(Ω;H1(Y)3) such that

    ˜v(x,y)=0inω×Ys,divy˜v=0inω×Y,(Y˜v(x,y)dy)n=0onω.

    Proof. We choose a test function ˜v(x,y)D(ω;C(Y)3) with ˜v(x,y)=0ω×Ys (thus, we have that ˜v(x,x/aε,y3)(H10(˜Ωε))3). We first multiply (20) by a2ε and we use that divε˜uε=0. Then, we take a test function a2ε˜v(x,x/aε,y3), with ˜v3 independent of y3 and with ˜v(x,y)=0 in ω×Ys and satisfying the incompressibility conditions (51)-(52), that is, divy˜v=0 in ω×Y and (Y˜v(x,y)dy)n=0 on ω, and we have

    2μΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μ1a2εΩ|Dε[˜uε]|2dxdy3+2gaεΩ|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32g1aεΩ|Dε[˜uε]|dxdy3Ωf˜vdxdy31a2εΩf˜uεdxdy3+Ω˜Pεdivx˜vdxdy3+1aεΩ˜Pεdivy˜vdxdy3. (60)

    Applying the change of variables given in Remark 1 to relation (60) and taking into account (28), (30) and (39), we obtain

    2μω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy2μ1a2εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|2dxdy+2gaεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2g1aεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεω×Yf˜vdxdy1a2εω×Yfˆuεdxdy+ω×YˆPεdivx˜vdxdy+1aεω×YˆPεdivy˜vdxdy+Oε. (61)

    In the left-hand side, we only give the details of convergence for the first nonlinear term, the most challenging one.

    |2gaεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gω×Y|Dy[˜v]|dxdy|2gω×Y|aεDx[˜v]+Dy[˜v]+aεεy3[˜v]Dy[˜v]|dxdy2gω×Y|aεDx[˜v]|dxdy+2gω×Y|aεεy3[˜v]|dxdy0, as ε0.

    Using (49) the two first terms in the right hand side of (61) tend to the following limit

    ω×Yf(˜vˆu)dxdy.

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (49) the first term of the pressure tends to the following limit ω×YˆPdivx˜vdxdy, and using (51) and taking into account that ˆP does not depend on y, we have (46). Finally, using that divy˜v=0, we have

    1aεω×YˆPεdivy˜vdxdy=0. (62)

    It is straightforward to obtain that ˆu3=0 and therefore we get (59).

    We obtain some compactness results about the behavior of the sequences (˜uε,˜Pε) and (ˆuε,ˆPε) satisfying the a priori estimates given in Lemmas 3.1-ⅱ) and 3.3-ⅱ), respectively.

    Lemma 4.5 (Supercritical case). For a subsequence of ε still denote by ε, there exist ˜uH1(0,1;L2(ω)3), where ˜u3=0 and ˜u=0 on y3={0,1}, ˆuH1(0,1;L2(ω×Y)3) (" " denotes Y-periodicity), with ˆu=0 in ω×Ys, ˆu=0 on y3={0,1} such that Yˆu(x,y)dy=10˜u(x,y3)dy3 with Yˆu3dy=0 and ˆu3 independent of y3, and ˆPL20(ω×Y), independent of y, such that

    ˜uεε2(˜u,0)inH1(0,1;L2(ω)3), (63)
    ˆuεε2ˆuinH1(0,1;L2(ω×Y)3),ˆPεˆPinL20(ω×Y), (64)
    divx(10˜u(x,y3)dy3)=0inω,(10˜u(x,y3)dy3)n=0onω, (65)
    divyˆu=0inω×Y,divx(Yˆu(x,y)dy)=0inω, (66)
    (Yˆu(x,y)dy)n=0onω. (67)

    Proof. See Lemmas 5.2, 5.3 and 5.4 in [4] for the proof of (63)-(67). Here, we prove that ˆP does not depend on the microscopic variable y. To do this, we choose as test function ˜v(x,y)D(ω;C(Y)3) with ˜v(x,y)=0ω×Ys (thus, ˜v(x,x/aε,y3)(H10(˜Ωε))3). In order to prove that ˆP does not depend on y3, we set ε˜v(x,x/aε,y3) in (20) (we recall that g(ε)=gε))and using that divε˜uε=0, we have

    2μεΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μΩ|Dε[˜uε]|2dxdy3+2gε2Ω|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32gεΩ|Dε[˜uε]|dxdy3εΩf˜vdxdy3Ωf˜uεdxdy3+εΩ˜Pεdivx˜vdxdy3+εaεΩ˜Pεdivy˜vdxdy3+Ω˜Pεy3˜v3dxdy3. (68)

    Applying the change of variables given in Remark 1 to relation (68) and taking into account (28)-(30), we obtain

    2μεω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy+2gε2ω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεεω×Yf˜vdxdyω×Yfˆuεdxdy+εω×YˆPεdivx˜vdxdy+εaεω×YˆPεdivy˜vdxdy+ω×YˆPεy3˜v3dxdy+Oε. (69)

    According with (64) and using that aεε, one has for the first term in relation (69)

    2μεω×Y(εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]):(εaεDy[˜v]+y3[˜v])dxdy0, as ε0. (70)

    We pass to the limit in the first nonlinear term and we have

    2gεω×Y|εDx[˜v]+εaεDy[˜v]+y3[˜v]|dxdy0, as ε0. (71)

    In order to pass the limit in the second nonlinear term, we taking into account that

    εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy=ε2ω×Y|εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]|dxdy,

    and using (64), with aεε, and the fact that the function E(φ)=|φ| is proper convex continuous, we can deduce that

    lim infε02gεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy0. (72)

    Moreover, using (64) the two first terms in the right hand side of (69) can be written by

    εω×Yf˜vdxdyε2ω×Yfˆuεε2dxdy0, as ε0. (73)

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (64) and aεε, passing to the limit when ε tends to zero, we have

    ω×YˆPy3˜v3dxdy. (74)

    Therefore, taking into account (70)-(74), when we pass to the limit in (69) when ε tends to zero, we have 0ω×YˆPy3˜v3dxdy. Now, if we choose as test function ε˜v(x,x/aε,y3) in (20) and we argue similarly, we can deduce that ˆP does not depend on y3. Now, in order to prove that ˆP does not depend on y, we set aε˜v=aε(˜v(x,x/aε,y3),0) in (20) and using that divε˜uε=0, we have

    2μaεΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μΩ|Dε[˜uε]|2dxdy3+2gεaεΩ|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32gεΩ|Dε[˜uε]|dxdy3aεΩf˜vdxdy3Ωf˜uεdxdy3+aεΩ˜Pεdivx˜vdxdy3+Ω˜Pεdivy˜vdxdy3. (75)

    Applying the change of variables given in Remark 1 to relation (75) and taking into account (28)-(30), we obtain

    2μaεω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy+2gεaεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gεω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεaεω×Yf˜vdxdyω×Yfˆuεdxdy+aεω×YˆPεdivx˜vdxdy+ω×YˆPεdivy˜vdxdy. (76)

    According with (64) and using that aεε, the first term in relation (76) can be written by the following way

    2μaεω×Y(εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]):(εaεDy[˜v]+y3[˜v])dxdy0, as ε0. (77)

    In order to pass to the limit in the first nonlinear term, we have

    2gaεω×Y|εDx[˜v]+εaεDy[˜v]+y3[˜v]|dxdy0, as ε0. (78)

    Moreover, using (64) the two first terms in the right hand side of (76) can be written by

    aεω×Yf˜vdxdyε2ω×Yfˆuεε2dxdy0, as ε0. (79)

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (64), passing to the limit when ε tends to zero, we have

    ω×YˆPdivy˜vdxdy. (80)

    Therefore, taking into account (72) and (77)-(80), when we pass to the limit in (76) when ε tends to zero, we have 0ω×YˆPdivy˜vdxdy. Now, if we choose as test function aε˜v=aε(˜v(x,x/aε,y3),0) in (20) and we argue similarly, we can deduce that ˆP does not depend on y, so ˆP does not depend on y.

    Theorem 4.6 (Supercritical case). If aεε, then (ˆuε/ε2,ˆPε) converges to (ˆu,ˆP) in H1(0,1;L2(ω×Y)3)×L20(ω×Y), which satisfies the following variational equality

    2μω×Yy3[ˆu]:(y3[˜v]y3[ˆu])dxdy+2gω×Y|y3[˜v]|dxdy2gω×Y|y3[ˆu]|dxdyω×Yf(˜vˆu)dxdyω×YxˆP(˜vˆu)dxdy, (81)

    for every ˜vH1(0,1;L2(ω×Y)3) such that

    ˜v(x,y)=0inω×Ys,divy˜v=0inω×Y,(Y˜v(x,y)dy)n=0onω.

    Proof. We choose a test function ˜v(x,y)D(ω;C(Y)3) with ˜v(x,y)=0ω×Ys (thus, ˜v(x,x/aε,y3)(H10(˜Ωε))3). We first multiply (20) by ε2 and we use that divε˜uε=0. Then, we take a test function ε2˜v(x,x/aε,y3), with ˜v3 independent of y3 and with ˜v(x,y)=0 in ω×Ys and satisfying the incompressibility conditions (66)-(67), that is, divy˜v=0 in ω×Y and (Y˜v(x,y)dy)n=0 on ω, and we have

    2μΩDε[˜uε]:(Dx[˜v]+1aεDy[˜v]+1εy3[˜v])dxdy32μ1ε2Ω|Dε[˜uε]|2dxdy3+2gεΩ|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy32g1εΩ|Dε[˜uε]|dxdy3Ωf˜vdxdy31ε2Ωf˜uεdxdy3+Ω˜Pεdivx˜vdxdy3+1aεΩ˜Pεdivy˜vdxdy3. (82)

    Applying the change of variables given in Remark 1 to relation (82), arguing as in the critical case, we obtain

    2μω×Y(1aεDy[ˆuε]+1εy3[ˆuε]):(1aεDy[˜v]+1εy3[˜v])dxdy2μ1ε2ω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|2dxdy+2gεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2g1εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy+Oεω×Yf˜vdxdy1ε2ω×Yfˆuεdxdy+ω×YˆPεdivx˜vdxdy+1aεω×YˆPεdivy˜vdxdy+Oε. (83)

    According with (64), the first term in relation (83) can be written by the following way

    2μω×Y(εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]):(εaεDy[˜v]+y3[˜v])dxdy,

    and, taking into account that aεε, this term tends to the following limit

    2μω×Yy3[ˆu]:y3[˜v]dxdy. (84)

    The second term in relation (83) writes

    2μω×Y(εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]):(εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε])dxdy,

    and, taking into account that the function B(φ)=|φ| is proper convex continuous and aεε, we get that the lim infε0 of this second is greater or equal than

    2μω×Yy3[ˆu]:y3[ˆu]dxdy. (85)

    In order to pass to the limit in the first nonlinear term, using that aεε, we have

    |2gεω×Y|Dx[˜v]+1aεDy[˜v]+1εy3[˜v]|dxdy2gω×Y|y3[˜v]|dxdy|2gω×Y|εDx[˜v]+εaεDy[˜v]+y3[˜v]y3[˜v]|dxdy2gω×Y|εDx[˜v]|dxdy+2gεaεω×Y|Dy[˜v]|dxdy0, as ε0.

    Now, in order to pass the limit in the second nonlinear term, taking into account that

    1εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy=ω×Y|εaε1ε2Dy[ˆuε]+1ε2y3[ˆuε]|dxdy,

    and using (64) and the fact that the function E(φ)=|φ| is proper convex continuous and aεε, we can deduce that

    lim infε02g1εω×Y|1aεDy[ˆuε]+1εy3[ˆuε]|dxdy2gω×Y|y3[ˆu]|dxdy. (86)

    Moreover, using (64) the two first terms in the right hand side of (83) tend to the following limit

    ω×Yf(˜vˆu)dxdy. (87)

    We consider now the terms which involve the pressure. Taking into account the convergence of the pressure (64) the first term of the pressure tends to the following limit ω×YˆPdivx˜vdxdy, and using (66) and taking into account that ˆP does not depend on y, we have (46). Finally using that divy˜v=0, we have (62). Therefore, taking into account (46), (62) and (84)-(87), we get (81).

    By using dimension reduction and homogenization techniques, we studied the limiting behavior of the velocity and of the pressure for a nonlinear viscoplastic Bingham flow with small yield stress, in a thin porous medium of small height ε and for which the relative dimension of the pores is aε. Three cases are studied following the value of λ=limε0aε/ε and, at the limit, they all preserve the nonlinear character of the flow. More precisely, according to [29], each of the limit problems (37), (59) and (81), is written as a nonlinear Darcy equation:

    {~U(x)=Kλ(f(x)xˆP(x))inω,divx~U(x)=0inω,~U(x)n=0onω. (88)

    The velocity of filtration ˜U(x)=(~U(x),˜U3(x)) is defined by

    ˜U(x)=Yˆu(x,y)dy=10(Yˆu(x,y,y3)dy)dy3=10˜u(x,y3)dy3.

    We remark that in all three cases, the vertical component ˜U3 of the velocity of filtration equals zero and this result is in accordance with the previous mathematical studies of the flow in this thin porous medium, for newtonian fluids (Stokes and Navier-Stokes equations) and for power law fluids (see [20], [1], [2], [4], [5]). Moreover, despite the fact that the limit pressure is not unique, the velocity of filtration is uniquely determined (see Section 4.3 in [29]). In (88), the function Kλ:R2R2 is nonlinear and its expression can not be made explicit for the Bingham flow (see [29]). Nevertheless, in each case, for a given ξR2, one has Kλ(ξ)=Yχξλ(y)dy, with χξλ solution of a local problem stated in the cell Y. If 0<λ<+, the local problem is a 3-D Bingham problem. If λ=0, the local problem is a 2-D Bingham problem (defined for each y3]0,1[), while if λ=+ the 1-D local problem (defined for each yY) corresponds to a lower-dimensional Bingham-like law (see [15]).

    We end with the remark that if in the initial problem (9) we take g=0, then the problem under study becomes the Stokes problem. We refer to [4] (case p=2) for the asymptotic analysis of the Stokes problem. If we set g=0 in the limit problems (37), (59) and (81), they become exactly the ones in [4], Theorem 6.1 (case p=2), corresponding to the Stokes case.

    We are grateful to the anonymous referee whose suggestions and questions were very valuable.



    [1] Darcy's laws for non-stationary viscous fluid flow in a thin porous medium. Math. Meth. Appl. Sci. (2017) 40: 2878-2895.
    [2] On the non-stationary non-Newtonian flow through a thin porous medium. ZAMM-Z. Angew. Math. Mech. (2017) 97: 895-915.
    [3] M. Anguiano and R. Bunoiu, On the flow of a viscoplastic fluid in a thin periodic domain,, in Integral Methods in Science and Engineering (eds. C. Constanda and P. Harris), Birkhäuser, Cham, (2019), 15–24. doi: 10.1007/978-3-030-16077-7_2
    [4] M. Anguiano and F. J. Suárez-Grau, Homogenization of an incompressible non-newtonian flow through a thin porous medium,, Z. Angew. Math. Phys., 68 (2017), Art. 45, 25 pp. doi: 10.1007/s00033-017-0790-z
    [5] M. Anguiano and F. J. Suárez-Grau, The transition between the Navier-Stokes equations to the Darcy equation in a thin porous medium, Mediterr. J. Math., 15 (2018), Art. 45, 21 pp. doi: 10.1007/s00009-018-1086-z
    [6] (2013) Introduction to the Network Approximation Method for Materials Modeling.Cambridge University Press.
    [7] Homogenized non-newtonian viscoelastic rheology of a suspension of interacting particles in a viscous newtonian fluid,. SIAM Journal of Applied Mathematics (2004) 64: 1002-1034.
    [8] Laminar shallow viscoplastic fluid flowing through an array of vertical obstacles. J. Non-Newtonian Fluid Mech. (2018) 257: 59-70.
    [9] D. Bresch, E. D. Fernandez-Nieto, I. Ionescu and P. Vigneaux, Augmented lagrangian method and compressible visco-plastic flows: Applications to shallow dense avalanches, in New Directions in Mathematical Fluid Mechanics. Advances in Mathematical Fluid Mechanics (eds. A.V. Fursikov, G.P. Galdi and V.V. Pukhnachev), Birkhäuser Basel, (2010), 57–89.
    [10] A note on homogenization of Bingham flow through a porous medium. J. Math. Pures Appl. (1993) 72: 405-414.
    [11] Bingham flow in porous media with obstacles of different size. Math. Meth. Appl. Sci. (2017) 40: 4514-4528.
    [12] R. Bunoiu, G. Cardone and C. Perugia, Unfolding method for the homogenization of Bingham flow, in Modelling and Simulation in Fluid Dynamics in Porous Media. Springer Proceedings in Mathematics & Statistics (eds. J. Ferreira, S. Barbeiro, G. Pena and M. Wheeler), Vol 28, Springer, New York, NY, (2013), 109–123. doi: 10.1007/978-1-4614-5055-9_7
    [13] Asymptotic analysis of a Bingham fluid in a thin T-like shaped structure. J. Math. Pures Appl. (2019) 123: 148-166.
    [14] Fluide de Bingham dans une couche mince. Annals of University of Craiova, Math. Comp. Sci. Ser. (2003) 30: 71-77.
    [15] Asymptotic behaviour of a Bingham fluid in thin layers. Journal of Mathematical Analysis and Applications (2004) 293: 405-418.
    [16] The periodic unfolding method in homogenization. SIAM J. Math. Anal. (2008) 40: 1585-1620.
    [17] D. Cioranescu, A. Damlamian and G. Griso, The Periodic Unfolding Method: Theory and Applications to Partial Differential Problems, Series in Contemporary Mathematics, 3, Springer, 2018. doi: 10.1007/978-981-13-3032-2
    [18] D. Cioranescu, V. Girault and K. R. Rajagopal, Mechanics and Mathematics of Fluids of the Differential Type, Advances in Mechanics and Mathematics, 35, Springer, 2016. doi: 10.1007/978-3-319-39330-8
    [19] G. Duvaut and J. L. Lions, Les Inéquations en Mécanique et en Physique, Travaux et Recherches Mathématiques, 21, Dunod, Paris, 1972.
    [20] Darcy's law for flow in a periodic thin porous medium confined between two parallel plates. Transp. Porous Med. (2016) 115: 473-493.
    [21] Shallow water equations for power law and Bingham fluids. Science China Mathematics (2012) 55: 277-283.
    [22] V. Girault and P. A. Raviart, Finite Element Methods for Navier-Stokes Equations. Theory and Algorithms, Springer Series in Computational Mathematics, 5, Springer-Verlag, 1986. doi: 10.1007/978-3-642-61623-5
    [23] Asymptotic behavior of a crane. C.R.Acad.Sci. Paris, Ser. Ⅰ (2004) 338: 261-266.
    [24] Junctions between two plates and a family of beams. Math. Meth. Appl. Sci. (2018) 41: 58-79.
    [25] Asymptotic analysis for domains separated by a thin layer made of periodic vertical beams. Journal of Elasticity (2017) 128: 291-331.
    [26] Onset and dynamic shallow flow of a viscoplastic fluid on a plane slope. J. Non-Newtonian Fluid Mech. (2010) 165: 1328-1341.
    [27] Augmented lagrangian for shallow viscoplastic flow with topography. Journal of Computational Physics (2013) 242: 544-560.
    [28] Viscoplastic shallow flow equations with topography. J. Non-Newtonian Fluid Mech. (2013) 193: 116-128.
    [29] Écoulement d'un fluide viscoplastique de Bingham dans un milieu poreux. J. Math. Pures Appl. (1981) 60: 341-360.
    [30] P. Lipman, J. Lockwood, R. Okamura, D. Swanson and K. Yamashita, Ground Deformation Associated with the 1975 Magnitude-7.2 Earthquake and Resulting Changes in Activity of Kilauea Volcano, Hawaii, Professional Paper, 1276, Technical report, US Goverment Printing Office, 1985. doi: 10.3133/pp1276
    [31] Approximate equations for the slow spreading of a thin sheet of Bingham plastic fluid. Physics of Fluids A: Fluid Dynamics (1990) 2: 30-36.
    [32] Asymptotic analysis in elasticity problems on thin periodic structures. Networks and Heterogeneous Media (2009) 4: 577-604.
    [33] Flow of fresh concrete through steel bars: A porous medium analogy. Cement and Concrete Research (2011) 41: 496-503.
    [34] Homogenization of a stationary Navier-Stokes flow in porous medium with thin film. Acta Mathematica Scientia (2008) 28: 963-974.
  • This article has been cited by:

    1. Renata Bunoiu, Antonio Gaudiello, On the Bingham Flow in a Thin Y-Like Shaped Structure, 2022, 24, 1422-6928, 10.1007/s00021-021-00657-0
    2. María Anguiano, Matthieu Bonnivard, Francisco J Suárez-Grau, Carreau law for non-newtonian fluid flow through a thin porous media, 2022, 75, 0033-5614, 1, 10.1093/qjmam/hbac004
    3. Giuseppe Cardone, Carmen Perugia, Manuel Villanueva Pesqueira, Asymptotic Behavior of a Bingham Flow in Thin Domains with Rough Boundary, 2021, 93, 0378-620X, 10.1007/s00020-021-02643-7
    4. Francisco J. Suárez-Grau, Mathematical modeling of micropolar fluid flows through a thin porous medium, 2021, 126, 0022-0833, 10.1007/s10665-020-10075-2
    5. María Anguiano, Francisco J. Suárez-Grau, Lower-Dimensional Nonlinear Brinkman’s Law for Non-Newtonian Flows in a Thin Porous Medium, 2021, 18, 1660-5446, 10.1007/s00009-021-01814-5
    6. Francisco J. Suárez‐Grau, Theoretical derivation of Darcy's law for fluid flow in thin porous media, 2022, 295, 0025-584X, 607, 10.1002/mana.202000184
    7. Renata Bunoiu, Claudia Timofte, Upscaling of a double porosity problem with jumps in thin porous media, 2022, 101, 0003-6811, 3497, 10.1080/00036811.2020.1854232
    8. Reine Gladys Noucheun, Jean Louis Woukeng, Homogenization of Smoluchowski Equations in Thin Heterogeneous Porous Domains, 2023, 11, 2227-7390, 3796, 10.3390/math11173796
    9. Renata Bunoiu, Karim Karim, Claudia Timofte, T-coercivity for the asymptotic analysis of scalar problems with sign-changing coefficients in thin periodic domains, 2021, 2021, 1072-6691, 59, 10.58997/ejde.2021.59
    10. María Anguiano, Francisco J. Suárez-Grau, Sharp Pressure Estimates for the Navier–Stokes System in Thin Porous Media, 2023, 46, 0126-6705, 10.1007/s40840-023-01514-1
    11. María Anguiano, Francisco Javier Suárez-Grau, Mathematical derivation of a Reynolds equation for magneto-micropolar fluid flows through a thin domain, 2024, 75, 0044-2275, 10.1007/s00033-023-02169-5
    12. María Anguiano, Matthieu Bonnivard, Francisco J. Suárez‐Grau, Effective models for generalized Newtonian fluids through a thin porous medium following the Carreau law, 2025, 105, 0044-2267, 10.1002/zamm.202300920
  • Reader Comments
  • © 2020 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(3672) PDF downloads(261) Cited by(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog