Publication Cover
Applicable Analysis
An International Journal
Volume 101, 2022 - Issue 14
1,274
Views
5
CrossRef citations to date
0
Altmetric
Research Article

On pressure-driven Hele–Shaw flow of power-law fluids

ORCID Icon, ORCID Icon & ORCID Icon
Pages 5107-5137 | Received 25 Jun 2020, Accepted 14 Jan 2021, Published online: 03 Feb 2021

Abstract

We analyze the asymptotic behavior of a non-Newtonian Stokes system, posed in a Hele–Shaw cell, i.e. a thin three-dimensional domain which is confined between two curved surfaces and contains a cylindrical obstacle. The fluid is assumed to be of power-law type defined by the exponent 1< p<. By letting the thickness of the domain tend to zero we obtain a generalized form of the Poiseuille law, i.e. the limit velocity is a nonlinear function of the limit pressure gradient. The flow is assumed to be driven by an external pressure which is applied as a normal stress along the lateral part of the boundary. On the remaining part of the boundary we impose a no-slip condition. The two-dimensional limit problem for the pressure is a generalized form of the p-Laplace equation, 1/p+1/p=1, with a coefficient called ‘flow factor’, which depends on the geometry as well as the power-law exponent. The boundary conditions are preserved in the limit as a Dirichlet condition for the pressure on the lateral boundary and as a Neumann condition for the pressure on the solid obstacle.

2010 Mathematics Subject Classifications:

1. Introduction

In this paper we consider the stationary flow of an incompressible non-Newtonian fluid in a thin domain Ωϵ in R3, where ε is a positive small parameter related to the thickness of the domain. More precisely, we consider a Hele–Shaw cell, i.e. the domain is confined between two curved surfaces in close proximity and contains a cylindrical obstacle. The fluid is modeled as a power-law fluid, i.e. the viscosity depends on the rate of strain via a power-law. The boundary Ωϵ consists of two disjoint part ΓDϵ (Dirichlet boundary) and ΓNϵ (Neumann boundary). The Dirichlet part corresponds to the upper and lower surfaces which are separated by a non-uniform thickness, as well as the lateral surface of the obstacle. The Neumann part, which corresponds to the lateral boundary of the cell, is regarded as a penetrable inlet/outlet zone where an external pressure gradient is applied as a surface force along ΓNϵ. Moreover, the flow is assumed to be governed by the Stokes equation, therefore inertial effects are neglected. It is also assumed that the flow is isothermal and that gravity can be neglected.

The main aim of the present work is to derive a lower-dimensional model for the flow by studying the asymptotic behavior of the system described above, as ϵ0. This dimensional reduction will be achieved by using an adaptation of a technique called two-scale convergence for thin domains which was developed by Marušić and Marušić-Paloka [Citation1]. The present results generalize [Citation2] where the Newtonian case was considered without any obstacle in the geometry. Moreover, we generalize the classical Poiseuille law, see e.g. [Citation3, p. 222], for Newtonian flow. In particular the Newtonian Hele–Shaw flow is suitable to visualize streamlines around an obstacle [Citation4], as well as, connecting the velocity to the gradient of the pressure by linear dependence. The main result of the present paper is a nonlinear Poiseuille law, in which the limit velocity and the limit pressure gradient follows a power-law [Citation5, Sec. 7.3.1]. A novelty is that we consider pressure-driven flow that can be found in many realistic applications, see e.g. [Citation6–9].

Non-Newtonian flow in Hele–Shaw like domains appears in many industrial applications, e.g. polymer processing, transportation of oil in pipelines, hydrology, as well as in food processing. In particular, we focus on some works where analysis and mathematical modeling of Poiseuille flow of power-law fluids are considered. Aronsson and Janfalk [Citation10] derived the equations of motion and proved some exact solutions and a representation theorem. A more rigorous study was performed by Mikelić and Tapiéro [Citation11] who derived a nonlinear Poiseuille law as a limit case of the Navier–Stokes equations when the thickness of the cell tends to zero. However, to avoid technical difficulties the authors chose to employ a no-slip condition on the whole boundary. The typical situation in a Hele–Shaw cell is that some part of the boundary should be penetrable, thus allowing fluid particles to enter and leave the domain. At such boundaries it is natural to prescribe the mass flux or impose a boundary condition for the pressure, i.e. by prescribing the momentum flux. Note also that, most examples in Hele–Shaw's original paper showed domains where a solid obstacle was placed in the interior of the domain, which is another characteristic of Hele–Shaw flow. Thus, the present work complements [Citation11] in the sense that a mixed boundary condition is employed, both surfaces may be curved and the domain may contain an interior obstacle. Related models that take thermal effects into account can be found in [Citation12,Citation13]. Non-Newtonian Hele–shaw flow of other type than power-law relation has been studied in [Citation14,Citation15].

Poiseuille's law also plays an important role in lubrication theory, formulated in the Newtonian case by Reynolds [Citation16], see also [Citation17, Ch. 22]. Recall that the fundamental problem in lubrication theory is to describe fluid flow in the gap between adjacent surfaces which are in relative motion to each other and in general there are no obstacles in the domain. Bayada and Chambat [Citation18], gave the first rigorous mathematical derivation of Reynolds equation, considering the Stokes problem in a thin domain and assuming that the velocity field satisfies a Dirichlet condition on the whole boundary. The authors proved that the limit pressure satisfies the classical Reynolds equation with a Neumann condition. In contrast, our limit equation has a Dirichlet condition for the pressure. The connection between Hele–Shaw theory and lubrication theory has been further explored in [Citation19]. A lubrication problem with power-law fluids was considered in [Citation20], under the assumption that the lower surface is flat and moving whereas the upper surface is rough and stationary. Flow of power-law fluids through a thin porous medium, i.e. periodic array of vertical cylinders confined between two parallel plates with no-slip boundary condition on the whole boundary was studied in [Citation21].

The paper is organized as follows: In Section 2 we give a precise description of the Hele–shaw cell and present some notation and preliminary results. In Section 3 we set the problem and formulate our main result. To prove the main result we need several a priori estimates and some results related to two-scale convergence for thin domains. These results are proved in Section 4– 5. Finally, the main result is proved in Section 56. In addition, several technical results are proved in Appendices 1–4.

2. Preliminaries, notation and some technical results

2.1. Euclidean structure

Let Rm×n denote the set of real m×n matrices X=(xij)i,j=1m,n equipped with the Euclidean scalar product X:Y=tr(XTY)=j=1ni=1mxijyij,|X|=(X:X)1/2,where XT=(xji)i,j=1n,m in Rm×n denotes the transpose of X, the symmetrical and skew-symmetric parts of X are defined respectively e(X)=12(X+XT),ω(X)=12(XXT).The identity element in Rn×n is denoted as I=(δij)i,j=1n,n. We identify Rn with Rn×1 (column vectors) and denote the scalar product in Rn as xy=xTy=i=1nxiyi,|x|=(xx)1/2.

2.2. Geometry of the domain

Let ω denote a perforated Lipschitz domain in R2, say ω=(L,L)×(L,L)D,where L>0 and D is the closed disc D={(x1,x2):x12+x22R2}of radius 0<R<L. The boundary of ω is decomposed into two disjoint parts γD=DγN=ωγD.Thus, both γD and γN have positive measure (arc length). Note that the main result of this paper is valid for more general domains ω that share the essential properties of the definition given above.

The (unscaled) fluid domain is defined as Ω={(x1,x2,x3)R3:(x1,x2)ω, h(x1,x2)<x3<h+(x1,x2)},where h+,h are Lipschitz functions defined on the closure of ω, i.e. h+,hC0,1(ω¯) and satisfy the conditions (1) 12h(x1,x2)14and14h+(x1,x2)12h(x)=defh+(x1,x2)h(x1,x2)andm(x1,x2)=defh+(x1,x2)+h(x1,x2)2,(1) where h(x1,x2) with 12h(x1,x2)1 is the thickness of the fluid domain and m(x1,x2) is the center surface of the fluid domain.

A point in the domain Ω will be denoted as (x,y), where x=(x1,x2) and y=x3. Similarly, the components of a vector function u:ΩR3 are denoted as u=(u,u3), where u=(u1,u2).

The boundary of Ω is divided into two disjoint parts ΓD=Ω{(x,y):y=h+(x) or y=h(x) or xD}ΓN=ΩΓD,where ΓN is referred to as the ‘lateral’ surface.

The thin domain Ωϵ is defined by anisotropic scaling of the unscaled domain Ω of unit thickness, using the parameter 0<ϵ1, i.e. Ωϵ={(x,x3)R3:xω,ϵh(x)<x3<ϵh+(x)}.In other words Ωϵ is a Hele–Shaw cell of thickness ϵh(x). Also, the boundary of Ωϵ is divided into two disjoint parts ΓDϵ=Ωϵ{(x,x3):x3=ϵh+(x) or x3=ϵh(x) or xD}ΓNϵ=ΩϵΓDϵ.

2.3. Function spaces

We shall work mainly with the following function spaces Wp(Ωϵ)={vW1,p(Ωϵ;R3):v=0 on ΓDϵ}Vp(Ωϵ)={vWp(Ωϵ):divv=0 in Ωϵ,v=0 on ΓDϵ},where 1<p< with conjugate exponent p=p/(p1) and the divergence operator div:Wp(Ωϵ)Lp(Ωϵ) is defined by (2) divv=i=13vixi.(2) Vp(Ωϵ)Wp(Ωϵ) are closed subspace of W1,p(Ωϵ;R3). Since ΓDϵ has positive surface measure we can equip Wp(Ωϵ) with the norm vWp(Ωϵ)=e(v)Lp(Ωϵ)=Ωϵe(v)pdx1/pwhich is equivalent to the usual W1,p-norm by Korn's inequality [Citation22]. Moreover, Wp(Ωϵ)/Vp(Ωϵ) denotes the quotient space of Wp(Ωϵ) by Vp(Ωϵ) defined as the set of all equivalence classes {u}=u+Vp(Ωϵ),equipped with the norm {v}Wp(Ωϵ)/Vp(Ωϵ)=infuVp(Ωϵ)vuWp(Ωϵ).For more details concerning the characterization of this space, see e.g. [Citation23,Citation24].

We introduce the operator a:R3×3R3×3 defined by (3) a(X)=2μ0|e(X)|p2e(X)(XR3×3),(3) where e(X) is the symmetrical part of X. Then a has the following properties

  1. a is continuous (1<p<), with a(0)=0;

  2. a is (p1)-homogeneous, i.e. (4) a(λX)=λp1a(X),(λ>0);(4)

  3. a is monotone, i.e. (5) (a(X)a(Y)):(XY)0,(5) holds for all X,YR3×3.

For more details, see e.g. [Citation25].

In the following subsections, we give some variants of Korn's inequality, Bogovski and de Rham's operators for our thin domain.

2.4. Korn's inequality

These estimates show how the constants depend on the parameter in the thin domain Ωϵ.

Theorem 2.1

Korn's inequality

There exist constants K1 and K2 depending only on Ω and ΓD such that (6) ||v||Lp(Ωϵ)ϵK1||e(v)||Lp(Ωϵ)(6) (7) ||v||Lp(Ωϵ)K2||e(v)||Lp(Ωϵ),(7) for all vWp(Ωϵ). These constants are such that sup0<ϵ1ϵ1K1ϵK1,sup0<ϵ1K2ϵK2,where K1ϵ and K2ϵ denote the best constants in the inequalities (Equation6) and (Equation7).

Proof.

See Appendix 1.

2.5. The Bogovski operator

The divergence operator is onto, see [Citation24, Theorem 5.4] where the case p = 2 was considered. The generalization to 1<p< is straight-forward since the vector field vˆ in [Citation24, Lemma 5.1] is of class C0,1. Related results can also be found in [Citation26–28]. However, the divergence operator is not one-to-one since Null div=Vp(Ωϵ) is always nontrivial. By collapsing the nullspace to zero we can construct an invertible operator A:Wp(Ωϵ)/Vp(Ωϵ)Lp(Ωϵ)defined by A{u}=divu, for all uWp(Ωϵ), which is one-to-one, onto and therefore A is an isomorphism. From Jensen's inequality it follows that ||A{u}||Lp(Ωϵ)=||divu||Lp(Ωϵ)3p||u||Lp(Ωϵ),for all u{u}, hence A is continuous. The Open Mapping theorem asserts that the inverse B:Lp(Ωϵ)Wp(Ωϵ)/Vp(Ωϵ)of A is continuous. The operator B is called the Bogovski operator.

2.6. De Rham's operator

Since B is a continuous isomorphism, we can identify Wp(Ωϵ)/Vp(Ωϵ)Vp(Ωϵ),where Vp(Ωϵ) is the annihilator of V(Ωϵ) defined as Vp(Ωϵ)={FWp(Ωϵ):F,φWp(Ωϵ),Wp(Ωϵ)=0,  φVp(Ωϵ)},and by the Riesz Representation theorem, we have (Lp(Ωϵ))Lp(Ωϵ). The dual operator B:Vp(Ωϵ)Lp(Ωϵ),of B is usually called De Rham's operator or the ‘pressure operator’. Note that B is also continuous with ||B||=||B||. To emphasize that B depends on the domain Ωϵ, we write B=BΩϵ. Thus, to obtain a Lp-bound for the pressure we need to investigate how the operator norm ||BΩϵ||=||BΩϵ||=defsuphLp(Ωϵ)h=1infvWp(Ωϵ)divv=hΩϵ|e(v)|pdx1/pdepends on ε.

Theorem 2.2

Let BΩϵ:Lp(Ωϵ)Wp(Ωϵ)/Vp(Ωϵ),denote the Bogovski operator. Then there exist positive constants C1 and C2 depending only on Ω and ΓD such that (8) C1inf0<ϵ1ϵBΩϵsup0<ϵ1ϵBΩϵC2.(8)

Proof.

See Appendix 2.

Remark 2.3

We would like to point out that the lower bound in the Theorem 2.2 for the norm of the Bogovski operator is not needed to prove the main result, it is included only to show that the norm BΩϵ as ϵ0 and that an upper bound of the form BΩϵϵ1C2 is the best possible estimate that can be obtained.

Theorem 2.4

De Rham's Theorem

Let FWp(Ωϵ) be such that F,φWp(Ωϵ),Wp(Ωϵ)=0,for all φVp(Ωϵ).Then, there exists a unique qLp(Ωϵ), such that F,φWp(Ωϵ),Wp(Ωϵ)=Ωϵqdivφdx,for all φWp(Ωϵ).

Proof.

Since FWp(Ωϵ) satisfying F,φ=0 for all φVp(Ωϵ) this implies that F  Vp(Ωϵ), therefore, we can choose q in Lp(Ωϵ) such that q=BF. Furthermore, by duality we have BF,ϕLp(Ωϵ),Lp(Ωϵ)=F,BϕWp(Ωϵ),Wp(Ωϵ).Then, we define {φ}=BϕWp(Ωϵ)/Vp(Ωϵ), which implies that ϕ=A{φ}=divφ, consequently, q,divφLp(Ωϵ),Lp(Ωϵ)=F,φWp(Ωϵ),Wp(Ωϵ),i.e. F,φWp(Ωϵ),Wp(Ωϵ)=Ωϵqdivφdx,as desired.

3. Setting the problem and the main result

3.1. Power-law fluid

Let us suppose that uϵ is the velocity field, qϵ is the pressure and uϵ=(uiϵ/xj)i,j=13,3 is the gradient of velocity. The rate of strain tensor e=(eij)i,j=13,3 is defined by eij=12(uiϵxj+ujϵxi).The stress tensor σϵ=(σijϵ)i,j=13,3 is defined by the following non-Newtonian constitutive law (9) σijϵ=qϵδij+2μ(e)eij.(9) When the function μ (dynamic viscosity) is constant, we say that the fluid is Newtonian. In the present paper we study the flow of a power-law fluid, where the viscosity μ:[0,)R is a nonlinear function. More precisely, we assume that (10) μ(t)=μ0tp2(t>0),(10) which defines the rheology of the fluid, where 1<p< and μ0>0 are constant parameters called power-law index and the consistency respectively. For more details see [Citation29, pp. 16–22] or [Citation30, p. 55]. The parameter p is dimensionless while μ0 has units which depend on the value of p. When p = 2 we recover the case of a Newtonian fluid studied in [Citation2].

3.2. Boundary value problem

Let us consider stationary pressure-driven isothermal fluid flow in the Hele–Shaw cell Ωϵ. We assume that inertial and body forces can be neglected. Moreover, we assume that the fluid is incompressible and obeys the non-Newtonian constitutive relation (Equation9). We model the flow by the following boundary value problem in Ωϵ (11) divσϵ=0in Ωϵdivuϵ=0in Ωϵuϵ=0on ΓDϵσϵnˆ=pbnˆon ΓNϵ,(11) where uϵ and qϵ are the unknowns, nˆ is the outward unit normal of Ωϵ and pb, the external pressure, is a given function. We assume that pb is defined on the whole domain, more precisely pbW1,p(ω), i.e. pb depends only on the variable x.

3.3. Existence and uniqueness

In order to prove existence and uniqueness of solutions, we consider the problem (Equation11) in the general form (12) divσϵ=fin Ωϵdivuϵ=0in Ωϵuϵ=0on ΓDϵσϵnˆ=gon ΓNϵ,(12) where fLp(Ωϵ;R3) and gLp(ΓNϵ;R3) are given vector functions, which are interpreted as body force and surface force respectively.

We use the standard methods of the calculus of variations to interpret the problem (Equation12) as a problem of minimization.

Lemma 3.1

Let J be the functional defined as J(v)=Ωϵ2μ0pe(v)pfvdxΓNϵgvdS.Then, there exists a unique minimizer uϵ in Vp(Ωϵ), i.e. J(uϵ)J(v) for all v in the admissible class Vp(Ωϵ).

Proof.

See Appendix 3.

3.4. Weak formulation

Take a test function vWp(Ωϵ) and apply the Divergence theorem to σϵnˆv, we have ΩϵσϵnˆvdS=Ωϵi,j=1nσijϵnjvidS=i=1nΩϵj=1nxj(σijϵvi)dx=i,j=1nΩϵviσijϵxj+σijϵvixjdx=Ωϵvdivσϵ+σϵ:vdx.Since v = 0 on ΓDϵ and σϵnˆ=g, from the definition of the stress tensor σϵ we see that, ΓNϵgvdS+Ωϵfvdx=Ωϵ(qϵI+2μ(|e(uϵ)|)e(uϵ)):vdx=Ωϵqϵdivvdx+Ωϵ2μ0e(uϵ)p2e(uϵ):vdx.Thus, the weak formulation of problem (Equation12) is (13) Ωϵqϵdivvdx=Ωϵ(2μ0e(uϵ)p2e(uϵ):vfv)dxΓNϵgvdS,(13) for all v in Wp(Ωϵ).

Theorem 3.2

Given fLp(Ωϵ) and gLp(ΓNϵ), there exists a unique weak solution (uϵ,qϵ)Vp(Ωϵ)×Lp(Ωϵ) of the problem (Equation12) satisfying the weak formulation (Equation13).

Proof.

By Lemma 3.1, there exists a unique minimizer uϵVp(Ωϵ) which satisfies the Euler–Lagrange equation, (14) Ωϵ2μ0e(uϵ)p2e(uϵ):vdx=Ωϵfvdx+ΓNϵgvdS,(14) for all vVp(Ωϵ). Define FϵWp(Ωϵ) by Fϵ,v=defΩϵ(2μ0e(uϵ)p2e(uϵ):vfv)dxΓNϵgvdS.In this notation (Equation14) becomes Fϵ,v=0for all vVp(Ωϵ).From De Rham's theorem we deduce the existence and uniqueness of a pressure function qϵ  Lp(Ωϵ), such that Fϵ,v=Ωϵqϵdivvdx,for all vWp(Ωϵ).This is the weak formulation (Equation13) and therefore (uϵ,qϵ)Vp(Ωϵ)×Lp(Ωϵ) is the unique weak solution of (Equation12).

3.5. Main result

Taking f = 0 and g=pbnˆ in the weak formulation (Equation13) we have that the weak solution uϵVp(Ωϵ) of (Equation11) satisfies (15) Ωϵqϵdivvdx=Ωϵ2μ0e(uϵ)p2e(uϵ):vdx+ΩϵpbvnˆdS,(15) for all vWp(Ωϵ). Applying the Divergence theorem to the surface integral and using that v = 0 on ΓDϵ, yields (16) Ωϵπϵdivvdx=Ωϵ(2μ0e(uϵ)p2e(uϵ):v+pbv)dx,(16) for all vWp(Ωϵ), where πϵ denotes the normalized pressure defined by (17) πϵ=qϵpb.(17) Recall that, by assumption, pb depends only of xω. This implies that the third component in pb is zero, i.e. pb=(pb/x1,pb/x2,0).

To give a precise asymptotic description of the system (Equation11), we need the definition of two-scale convergence for thin domains introduced by Marušić and Marušić-Paloka in [Citation1].

Definition 3.3

We say that a sequence uϵ, where ϵ>0, in Lp(Ωϵ) two-scale converges to u in Lp(Ω) provided that (18) limϵ01ΩϵΩϵuϵ(x)vx,ϵ1x3dx=1ΩΩu(x,y)v(x,y)dxdy,(18) for all v in Lp(Ω) and we write uϵ2u. Moreover, we say that uϵ two-scale converges strongly to u if (19) limϵ01ΩϵΩϵuϵ(x)u(x,ϵ1x3)pdx=0(19) and we write uϵ2u (strongly).

To characterize the limit velocity, we introduce the so called permeability function and the flow factor for a thin domain, both depending on the rheology of the fluid and the geometry of the domain.

Definition 3.4

The solution ψ of the boundary value problem (20) Δp,yψ=1in Ωψ=0on ΓD±,(20) where ΓD±=(x,y)ΓD:y=h±(x),is called the permeability function of Ω. That is, for each xω, ψ(x,) is the solution of Δp,yψ(x,)=1in h(x),h+(x)ψ(x,)=0on y=h±(x),where Δp,y()=y(|y()|p2y()) is the p-laplacian in the variable y.

Definition 3.5

The flow factor of Ω is defined as (21) ϱ(x)=h(x)h+(x)ψ(x,y)dy,xω,(21) where ψ is the permeability function, i.e. the solution of boundary value problem (Equation20).

Lemma 3.6

For our geometry the permeability function ψ and the flow factor ϱ for Ω, are given by (22) ψ(x,y)=1ph(x)2p|m(x)y|p,(22) for all y[h(x),h+(x)] and (23) ϱ(x)=2pp+1hp+1(x),(23) for all xω where h(x) and m(x) are defined in (Equation1).

Proof.

The proof of this lemma will be given in Section 6.

Remark 3.7

The smoothness of ψ depends on the smoothness of Ω. From (Equation22) we see that ψ is of class C0,1(Ω¯), since h is a Lipschitz function on ω¯. From (Equation23) and the fact that h is bounded from below by the positive constant by (Equation1), we deduce also that the flow factor ϱ and 1/ϱ belong to W1,(ω).

Before we formulate the main result, let us introduce the following definitions that will be used throughout this paper: Let ϕ be a scalar function and u=(u1,u2,u3) be a vector function, then we define x, y, divx and divy as xϕ=ϕx1,ϕx2,0,yϕ=0,0,ϕyand divxu=u1x1+u2x2,divyu=u3y.

Theorem 3.8

Main result

For each 0<ϵ1 the boundary value problem (Equation11) has a unique solution (uϵ,qϵ)Vp(Ωϵ)×Lp(Ωϵ) such that (24) 1|Ωϵ|1/pϵpuϵLp(Ωϵ)+ϵ1puϵLp(Ωϵ)C1pbLp(Ω)p11|Ωϵ|1/pqϵLp(Ωϵ)+||ϵ1a(uϵ)||Lp(Ωϵ)C2pbLp(Ω),(24) where the constants C1 and C2 depends only on Ω and ΓD. Let ψ be the permeability function defined by (Equation22). Then, the following two-scale convergence holds, ϵpuϵ2uandqϵ2q,where uLp(Ω;R3) is the limit velocity defined by u(x,y)=ψ(x,y)1μ0xq(x)p21μ0xq(x),and qW1,p(ω) is the unique solution of the boundary value problem (25) divx(ϱ|xq|p2xq)=0in ωq=pbon γNϱ|xq|p2xqnˆ=0on γD,(25) where ϱ is the flow factor defined by (Equation23) and nˆ is the outward unit normal to the surface of the obstacle.

We observe that the Dirichlet condition on ΓDϵ for the velocity field in the original problem becomes a Neumann condition on γD for the limit pressure and the Neumann condition on ΓNϵ for the stress tensor in the original problem becomes a Dirichlet condition on γN for the limit pressure.

Furthermore, by contrast, if one imposes a non-homogeneous Dirichlet condition for the velocity field on ΓNϵ one ends up with a Neumann condition on γN (this is shown in [Citation18, equation (5.7)]). A major difference between the two boundary conditions is that the limit pressure is uniquely determined by the Dirichlet condition, whereas it is only determined up to an additive constant in the case of a Neumann condition. Moreover, under the Neumann condition the obstacle becomes a streamline of the y-averaged flow (see Lemma 5.4 (iv)). The limit velocity u is uniquely determined in both formulations and the equation qy=0in Ω,tells us that pressure variation in the thickness of Ω can be neglected. In the Newtonian case, p = 2, the flow factor is given by ϱ(x)=h3(x)/12 and therefore, the pressure equation (Equation25) becomes the classical coefficient in the Reynolds equation formulated in [Citation16]. This shows that our results can be extended to the context of lubrication theory, see e.g. [Citation19]. In addition, for future work, it would be interesting to allow the upper and lower surfaces to be in contact, i.e. h+(x)=h(x) in some points x of ω, thus creating more obstacles in the domain. Then, the problem (Equation12) becomes a singular Hele–Shaw flow problem and therefore, the limit problem (Equation25) will be degenerate. For some results in the Newtonian case and h+=h on γD see [Citation1, Sec. 3.3.].

4. Estimates

To prove the main result (Theorem 3.8) we need uniform a priori estimates for the velocity, the monotone operator and the pressure with respect to the parameter ε.

Theorem 4.1

Velocity estimates

For the velocity field uϵ the following estimates hold (26) ϵpuϵLp(Ωϵ)K1K12μ0pbLp(Ωϵ)p1ϵ1puϵLp(Ωϵ)K2K12μ0pbLp(Ωϵ)p1,(26) where K1 and K2 are the constants in Theorem 2.1.

Proof.

Taking v=uϵ in (Equation16) and using the fact that divuϵ=0 we have, (27) Ωϵ2μ0|e(uϵ)|pdx=Ωϵpbuϵdx.(27) Applying the Hölder inequality in (Equation27) and the Korn inequality (Equation6) we find Ωϵ2μ0|e(uϵ)|pdxuϵLp(Ωϵ)pbLp(Ωϵ)ϵK1e(uϵ)Lp(Ωϵ)pbLp(Ωϵ).Thus, (28) e(uϵ)Lp(Ωϵ)ϵK12μ0pbLp(Ωϵ)p1.(28) This together with the first and second inequalities in the Korn inequality again in the left hand side, we obtain uϵLp(Ωϵ)ϵK1ϵK12μ0pbLp(Ωϵ)p1uϵLp(Ωϵ)K2ϵK12μ0pbLp(Ωϵ)p1.

Theorem 4.2

Monotone operator estimate

The monotone operator (Equation3) satisfies the following estimate (29) ||ϵ1a(uϵ)||Lp(Ωϵ)K1||pb||Lp(Ωϵ),(29) where K1 is the constants in Theorem 2.1.

Proof.

Follows from the estimate obtained in the proof of Theorem 4.1.

Theorem 4.3

Pressure estimate

The normalized pressure πϵ satisfies the estimate (30) πϵLp(Ωϵ)C2K1(K2+1)pbLp(Ωϵ),(30) where K1,K2 and C2 are the constants defined in Theorems 2.1 and 2.2 respectively.

Proof.

Let FϵWp(Ωϵ) be defined by Fϵ,v=Ωϵ(2μ0e(uϵ)p2e(uϵ):v+pbv)dx.From (Equation16) we see that Fϵ belongs to (Vp(Ω)), and by the Hölder inequality and the Korn inequalities (Equation6) and (Equation7) we obtain, |Fϵ,v|Ωϵ(2μ0|e(uϵ)|p1|v|+|pb||v|)dx2μ0K2e(uϵ)Lp(Ωϵ)p1e(v)Lp(Ωϵ)+ϵK1pbLp(Ωϵ)e(v)Lp(Ωϵ).This together with inequality (Equation28) it follows |Fϵ,v|ϵK1(K2+1)pbLp(Ωϵ)vWp(Ωϵ),whence, we deduce that FϵWp(Ωϵ)ϵK1(K2+1)pbLp(Ωϵ).Let πϵ in Lp(Ωϵ) be defined by the pressure operator, i.e. πϵ=BΩϵFϵΩϵπϵdivvdx=Fϵ,v vWp(Ωϵ).This means that the unknown pressure function in (Equation11) can be recovered via (Equation17), i.e. qϵ=πϵ+pb. Note that qϵ depends only on the boundary values of pb on ΓNϵ. From (Equation8) in Theorem 2.2 we infer, πϵLp(Ωϵ)BΩϵFϵWp(Ωϵ)ϵK1(K2+1)BΩϵpbLp(Ωϵ)C2K1(K2+1)pbLp(Ωϵ).

5. Two-scale convergence results

Here we prove some compactness results for two-scale convergence

5.1. Compactness results

Since pbW1,p(ω) we have 1|Ωϵ|1/ppbLp(Ωϵ)=1|Ω|1/ppbLp(Ω).So, raising both sides to power p1 and using the Theorems 4.1, 4.2 and 4.3 yield ϵpuϵLp(Ωϵ)+ϵ1puϵLp(Ωϵ)K12μ0p1K1+K2pbLp(Ωϵ)p1and πϵLp(Ωϵ)+||ϵ1a(uϵ)||Lp(Ωϵ)K1[C2(K2+1)+1]pbLp(Ωϵ).Thus, we can choose some constants C and C~, which are independent of ε as C=K12μ0p1(K1+K2)1|Ω|1/ppbLp(Ω)p1andC~=K1[C2(K2+1)+1]1|Ω|1/ppbLp(Ω)such that (31) 1|Ωϵ|1/pϵpuϵLp(Ωϵ)+ϵ1puϵLp(Ωϵ)C,1|Ωϵ|1/pπϵLp(Ωϵ)+||ϵ1a(uϵ)||Lp(Ωϵ)C~.(31)

Lemma 5.1

For any sequence (uϵ,πϵ) with 0<ϵ1 in Wp(Ωϵ)×Lp(Ωϵ) satisfying the bounds (Equation31), there exist a uLp(Ω;R3) with u/yLp(Ω,R3×1), and a π0Lp(Ω) such that, up to a subsequence

  1. ϵpuϵ2u=(u,u3),

  2. ϵ1puϵ2D(u)=0uy0u3y,

  3. ϵ1pe(uϵ)2e(D(u))=120uyuyT2u3y,

  4. πϵ2π0.

Proof.

For the sake of completeness we provide the proof which follows the same ideas as in [Citation1, Theorem 1]. Since (Equation31) holds, there exist a subsequence which is still denoted by uϵ and uLp(Ω;R3) such that limϵ01ΩϵΩϵϵpuϵ(x)vx,ϵ1x3dx=1ΩΩu(x,y)v(x,y)dxdy,for any vLp(Ω;R3). Similarly, there exists dLp(Ω;R3×3) such that, up to a subsequence limϵ01|Ωϵ|Ωϵϵ1puϵ(x):Φ(x,ϵ1x3)dx=1|Ω|Ωd(x,y):Φ(x,y)dxdy,for any ΦLp(Ω;R3×3).

Since uϵ vanishes on ΓDϵ we have (32) 0=Ωϵuϵ(x)Φ(x,ϵ1x3)nˆdS=Ωϵdiv (ΦTuϵ)dx=Ωϵuϵ(x):Φ(x,ϵ1x3)+uϵ(x)div xΦ+ϵ1div yΦ(x,ϵ1x3)dx,(32) for all ΦW1,p(Ω;R3×3) such that Φnˆ=0 on ΓN. Multiplying (Equation32) by ϵ1p/|Ωϵ| and passing to the limit as ϵ0 we see that (33) Ωd(x,y):Φ(x,y)+u(x,y)div yΦ(x,y)dxdy=0.(33) Let us characterize the subsequence limit d, by considering it block by block, i.e. d=d11d12d21d22=2×22×11×21×1.Similarly the blocks of the matrix Φ are denoted as Φ=Φ11Φ12Φ21Φ22=2×22×11×21×1.Choosing Φ12=0 and Φ22=0 in (Equation33) gives Ω(d11:Φ11+d21:Φ21)dxdy=0for all ΦCc(Ω;R3×3).We conclude that d11=0 and d21=0. Next, choosing Φ11=0 and Φ21=0 in (Equation33) gives Ωd12Φ12+d22Φ22+uΦ12y+u3Φ22ydxdy=0for all ΦCc(Ω;R3×3).By definition of weak derivative we deduce that d12=u/y and d22=u3/y. Thus, d=0uy0u3y=defD(u).Hence the trace of u on ΓD is well defined by [Citation1, Lemma 4 (i)]. Applying the Divergence theorem to (Equation33) we obtain 0=ΓDu(x,y)Φ(x,y)nˆdS,for all ΦW1,p(Ω;R3×3) such that Φnˆ=0 on ΓN. We conclude that u satisfies the Dirichlet condition u = 0 on ΓD. The convergence of the symmetrical part of gradient u follows by linearity. The two-scale convergence for the normalized pressure follows directly from the definition.

5.2. Two-scale convergence and the monotone operator

We introduce the monotone operator (Equation3) in the weak formulation (Equation16), i.e. (34) Ωϵ(πϵdivv+a(uϵ):v+pbv)dx=0,for all vWp(Ω).(34)

Lemma 5.2

Given vWp(Ω), let vϵWp(Ωϵ) be defined as vϵ(x)=ϵpv(x,ϵ1x3). Then, we have the strong two-scale convergences

  1. ϵ1pvϵ2D(v),

  2. a(ϵ1pvϵ)2a(D(v)).

In particular, vϵ is an admissible test function in the two-scale sense.

Proof.

The gradient of vϵWp(Ωϵ) is defined as vϵ(x)=ϵpxvϵp1vyϵpxv3ϵp1v3y(x,ϵ1x3),for all vWp(Ωϵ). Therefore, the weak two-scale convergence (i) follows from Lemma 5.1 (ii). The weak two-scale convergence (ii) is deduced from the homogeneity property of a defined in (Equation4). Thus, it is enough to prove (35) limϵ01Ωϵ1pa(ϵ1pvϵ)Lp(Ωϵ)=1Ω1pa(D(v))Lp(Ω).(35) Then, by the Change of Variable theorem, and noting that |Ωϵ|=ϵ|Ω|, we obtain 1|Ωϵ|1/pΩϵ|a(vϵ)|pdx1/p=1|Ω|1/pΩaϵpxvϵp1vyϵpxv3ϵp1v3ypdxdy1/p.Multiplying by ϵ1=ϵ(p1)(p1) and homogeneity property, we get 1|Ωϵ|1/pΩϵ|a(ϵ1pvϵ)|pdx1/p=1|Ω|1/pΩaϵxvvyϵxv3v3ypdxdy1/p.Passing to the limit as ϵ0, using the Dominated Convergence theorem and the continuity of a, we obtain (Equation35).

Lemma 5.3

The sequence a(uϵ), (0<ϵ1) in Lp(Ωϵ) satisfies the bound (Equation31). There exists ζLp(Ω;R3×3) such that up to a subsequence ϵ1a(uϵ)2ζ.Moreover, (36) limϵ0inf1|Ωϵ|Ωϵϵ1a(uϵ):ϵ1puϵdx1|Ω|Ωζ:D(u)dxdy(36) and if (Equation36) holds as an equality, then (37) Ωζ:D(v)dxdy=Ωa(D(u)):D(v)dxdy,(37) for all test functions v in Wp(Ω).

Proof.

In view of the monotonicity condition (Equation5), we have (38) Ωϵ(a(uϵ)a(vϵ)):(uϵvϵ)dx0,(38) where vϵ is a test function of the same form as in Lemma 5.2. This inequality yields Ωϵa(uϵ):uϵdxΩϵ(a(uϵ):vϵ+a(vϵ):(uϵvϵ))dx.Multiplying this inequality by the factor ϵp|Ωϵ|=ϵ1|Ωϵ|1/pϵ1p|Ωϵ|1/p, we find (39) 1|Ωϵ|Ωϵϵ1a(uϵ):ϵ1puϵdx1|Ωϵ|Ωϵϵ1a(uϵ):ϵ1pvϵdx+1|Ωϵ|Ωϵϵ1a(vϵ):ϵ1p(uϵvϵ)dx.(39) Let us analyze the integrals on the right hand side in (Equation39) in terms of limits as ϵ0. The limit of the first integral exists by hypothesis and is given by (40) limϵ01|Ωϵ|Ωϵϵ1a(uϵ):ϵ1pvϵdx =1|Ω|Ωζ:D(v)dxdy.(40) The second integral is convergent by Lemma 5.1 (ii) and Lemma 5.2, as the integrand is a product of strongly two-scale convergent and two-scale convergent sequences and [Citation1, Lemma 2]. (41) limϵ01|Ωϵ|Ωϵϵ1a(vϵ):ϵ1p(uϵvϵ)dx=1|Ω|Ωa(D(v)):D(uv)dxdy.(41) Thus, we can estimate the inferior limit of the integral on the left hand side in (Equation39), by means of limits (Equation40) and (Equation41), as ϵ0, obtaining (42) limϵ0inf1|Ωϵ|Ωϵϵ1a(uϵ):ϵ1puϵdx1|Ω|Ω(ζ:D(v)+a(D(v)):D(uv))dxdy,(42) and the inequality (Equation36) is established by taking u = v.

Assuming that equality holds in (Equation36), we see from (Equation42) that 1|Ω|Ωζ:D(u)dxdy1|Ω|Ω(ζ:D(v)+a(D(v)):D(uv))dxdy.After some rearranging, it follows that (43) Ω(ζD(v)):(D(u)D(v))dxdy0.(43) Fix wWp(Ω) and set v=defutw with (t>0) in (Equation43), we obtain tΩ(ζa(D(u)tD(w))):D(w)dxdy0.Dividing by t, using the continuity of a, and letting t0 give Ω(ζa(D(u))):D(w)dxdy0.Finally, replacing w by w, we deduce that Ωζ:D(w)dxdy=Ωa(D(u)):D(w)dxdy,for all wWp(Ω).

5.3. Conservation of volume

In this subsection we investigate the conservation of the incompressibility of the flow when we let ϵ0. To this end it is convenient to introduce the following space of functions.

Definition 5.4

Let Wyp(Ω) be defined as the closure of Wp(Ω) in the norm vWyp(Ω)=defvLp(Ω)+vyLp(Ω).

Lemma 5.5

Let uϵ, (0<ϵ1) be a sequence as in Lemma 5.1 with the additional condition that div uϵ=0 in Ωϵ. Then any subsequential two-scale limit u=(u,u3) satisfies (44) u3=0in Ω(44) and (45) divxh(x)h+(x)u(x,y)dy=0in ωh(x)h+(x)u(x,y)dynˆ=0on γD.(45)

Proof.

As in [Citation1, Proposition 4], the divergence free condition for uϵ can be stated as ΩϵuϵvnˆdS=Ωϵuϵvdx,for all scalar test functions vW1,p(Ωϵ). It follows that (46) 0=Ωϵ(uϵ)(x)xv(x,ϵ1x3)+ϵ1u3ϵ(x)vy(x,ϵ1x3)dx,(46) for all vW1,p(Ω) such that v = 0 on ΓN. Multiplying (Equation46) by ϵ1p/|Ωϵ| and passing to the limit, as ϵ0 yield 1|Ω|Ωu3(x,y)vy(x,y)dxdy=0,hence u3/y=0. From this and the no-slip boundary conditions on ΓD± we deduce (Equation44).

By choosing v in (Equation46) which does not depend on y, i.e. v=v(x), such that v = 0 on γN and passing to the limit, as ϵ0, we obtain 0=1|Ω|Ωu(x,y)xv(x)dxdy=1|Ω|ωh(x)h+(x)u(x,y)dyxv(x)dx,for all vW1,p(ω) such that v = 0 on γN. This implies that the line integral on ω depends only on γD, whence by the Divergence theorem, we have ωdivxh(x)h+(x)u(x,y)dyv(x)dx=γDh(x)h+(x)u(x,y)dyv(x)nˆdS.Since uLp(Ω;R2) we see that divxh(x)h+(x)u(x,y)dyLp(ω).It follows from [Citation22, Proposition 2] that the trace in the dual sense is the restriction of (h(x)h+(x)u(x,y)dy)nˆ on γD, which is well defined as, (47) h(x)h+(x)u(x,y)dynˆW1/p,p(γD;R2).(47) Thus, (Equation45) is proved.

Lemma 5.6

All functions vWyp(Ω) satisfy

  1. v/yLp(Ω;R3×1),

  2. v = 0 on ΓD,

  3. divxh(x)h+(x)v(x,y)dyLp(ω),

  4. h(x)h+(x)v(x,y)dynˆ=0 on γD.

Proof.

The proofs of (i)(ii) follow by the same arguments used in the proof of Lemma 5.1 and (iii)–(iv) can be deduced by following the ideas used in the proof of Lemma 5.5.

5.4. Momentum equation

We end this Section, by proving the main result of this paper. Indeed, we will derive a lower-dimensional model for the flow from the Stokes Equation (Equation11).

Lemma 5.7

The thin film equation

Let u=(u,u3) and π0 be as in Lemma 5.1 and let ζ denote a subsequential limit of ϵ1a(uϵ) as in Lemma 5.3. Then the first component u of u and π0 satisfy (48) x(π0+pb)+μ0Δp,yu=0in Ωπ0y=0in Ωu=0on ΓDπ0μ0uyp2uy0π0nˆ=0on ΓN.(48)

Proof.

By choosing test functions in (Equation34), of the form vϵ(x)=(v(x,ϵ1x3),ϵv3(x,ϵ1x3)),we have 1|Ωϵ|Ωϵxvϵ1vyxv3v3yπϵ(x)divxv+v3y(x,ϵ1x3)+ϵ1a(uϵ)(x):ϵxvϵ1vyxv3v3y(x,ϵ1x3)+xpb(x)vϵ(x)dx=0.Passing to the limit as ϵ0, we obtain (49) 1|Ω|Ωπ0divxv+v3y+ζ:0vy00+xpbvdxdy=0,(49) for all (v,v3)Wp(Ω). Choosing v=0, we see that 0=Ωπ0v3ydxdyfor all v3Wp(Ω),or in other words, π0/y=0 in Ω. This shows that π0 does not depend of variable y, i.e. π0(x,y)=π0(x).

Furthermore, multiplying (Equation34) with ϵp/|Ωϵ|, choosing vϵ=uϵ and using that divuϵ=0 in Ωϵ, we can assert that (50) limϵ01|Ωϵ|Ωϵϵ1a(uϵ):ϵ1puϵdx=limϵ01|Ωϵ|Ωϵpbϵpuϵdx=1|Ω|Ωxpbudxdy.(50) Using π0(x,y)=π0(x), we can rewrite the first term in (Equation49) as Ωπ0divxv+v3ydxdy=ωπ0(x)divxh(x)h+(x)vdydxwe deduce (51) ωπ0divxh(x)h+(x)vdydx+Ωζ:0vy00+xpbvdxdy=0,(51) which holds for all v=(v,v3)Wyp(Ω) (see Definition 5.4), since Wp(Ω) is dense in Wyp(Ω). Then taking v = u in (Equation51) and using Lemma 5.5, i.e. (Equation44) and (Equation45), we have (52) Ωζ:D(u)dxdy=Ωxpbudxdy.(52) From (Equation50) and (Equation52) we deduce that equality holds in (Equation36). Thus (Equation37) implies that (Equation49) can be written as (53) ωπ0divxh(x)h+(x)vdydx+Ωa(D(u)):0vy00+xpbvdxdy=0.(53) From the definition of the operator a, i.e. equality (Equation3) and (Equation44), we obtain a(D(u))=μ02p20uyuyT0p20uyuyT0.Inserting this into (Equation53) yields (54) ωπ0divxh(x)h+(x)vdy+h(x)h+(x)μ0uyp2uyvy+xpbvdydx=0,(54) for all (v,0)Wyp(Ω). This is the weak formulation of the boundary value problem (Equation48).

6. Regularity of pressure and pressure equation

In this section, we will prove some properties of solutions (u,q) to the lower-dimensional problem. Indeed, we will prove a regularity result for the pressure q=π0+pb and a nonlinear relation between the velocity u and the gradient of q. However, to be able to do this we first present the proof of Lemma 3.6.

Proof

Proof of Lemma 3.6

We note that the boundary value problem (Equation20) is formally an ordinary differential equation in variable y, with the parameter xω, i.e. yψy(x,y)p2ψy(x,y)=1.By formal integration, we see that there exist a scalar function C(x) such that ψy(x,y)p2ψy(x,y)=C(x)y.From the observation, ψy(x,y)p1=|C(x)y|ψy(x,y)=|C(x)y|p1  ψyp2=|C(x)y|2p,it follows that ψy(x,y)=|C(x)y|p2(C(x)y).Whence by integration once again the last equality becomes (55) ψ(x,y)=h(x)y|C(x)s|p2(C(x)s)ds=h(x)ydds1p|C(x)s|pds=1p(|C(x)h(x)|p|C(x)y|p).(55) From the boundary condition ψ(x,h(x))=0 and ψ(x,h+(x))=0, we see that C(x)=m(x),where m(x) is defined in (Equation1) and therefore we get (Equation22) from (Equation55). To obtain the flow factor ϱ(x), we integrate the permeability function on (h(x),h+(x)), i.e. ϱ(x)=h(x)h+(x)ψ(x,y)dy=2pp|h(x)|p+11ph(x)h+(x)|m(x)y|pdy.Observe that the last integral can be written as h(x)h+(x)|m(x)y|pdy=h(x)m(x)|m(x)y|p1(m(x)y)dy+m(x)h+(x)|m(x)y|p1(ym(x))dy=2p+1h+(x)h(x)2p+1.

Lemma 6.1

Let ψW1,(Ω) be the permeability function defined by (Equation22) and (u,π0)Wyp(Ω)×Lp(ω) be any solution of the boundary value problem (Equation48). Then π0 belongs to W1,p(ω), with π0=0 on γN.

Proof.

To show the W1,p-regularity for π0, in view of the Lemma 5.4, we take test functions v(x,y) of the form v(x,y)=ψ(x,y)Θ(x),v3(x,y)=0where Θ is any vector field in W1,p(ω;R2) such that Θnˆ=0 on γD, in (Equation54), we obtain ωπ0divx(ϱΘ)dx=ωμ0h(x)h+(x)ψyuyp2uyΘdy+ϱxpbΘdx,where ϱ is the flow factor defined in (Equation23). Since π0Lp(ω), we can define a linear functional with xπ0W1,p(ω) determined by xπ0,ϱ(x)Θ(x)=defωπ0divx(ϱΘ)dx,i.e. xπ0,ϱ(x)Θ(x)=Ωμ0ψyuyp2uyΘdxdy+ωϱxpbΘdx.Thus, (56) |xπ0,ϱ(x)Θ(x)|CΘLp(ω),(56) where C is the a positive constant given by C=defμ0ψyL(ω)uyLp(Ω)p1+ϱL(ω)xpbLp(ω).Furthermore, since the functions ϱ and ϱ1=1/ϱ belong to W1,(ω), (see Remark 3.7) we can replace Θ with ϱ1Θ in (Equation56) which gives |xπ0,Θ|Cϱ1ΘLp(ω)Cϱ1L(ω)ΘLp(ω)C~ΘLp(ω),for all ΘW1,p(ω;R2) such that Θnˆ=0 on γD, and therefore, in view of Lemma A.5 in Appendix 4, we conclude that π0 belongs to W1,p(ω) with π0=0 on γN.

Remark 6.2

Note that q=π0+pbW1,p(ω) with q=pb on γN.

Lemma 6.3

Let (u,π0)Wyp(Ω)×W1,p(ω) be any solution of (Equation48). Then u is uniquely determined by π0, more precisely (57) u(x,y)=ψ(x,y)1μ0xq(x)p21μ0xq(x),(57) where q=π0+pb and ψ must satisfy the scalar ordinary differential Equation (Equation20).

Proof.

In view of Lemma 6.1, the problem (Equation48) can be written as an ordinary differential equation in the variable y, i.e. (58) Δp,yu(x,)=ξin h(x),h+(x)u(x,)=0on y=h±(x),(58) where xω is a parameter and ξ=1μ0xq(x) can be regarded as a constant vector in R2.

We want to show that (Equation57) is a solution of the form u(x,y)=ψ(x,y)ξp2ξ(ξR2).Indeed, yuy(x,y)p2uy(x,y)=ξ.As in [Citation11, Propostion 3.4], by formal integration, we see that there exist a vector function C(x)Lp(ω;R2) such that uy(x,y)p2uy(x,y)=C(x)yξ.Thus, uy(x,y)=|C(x)yξ|p2(C(x)yξ)whence by integration once again and taking into account the no-slip condition at y=h(x) we obtain the velocity field (59) u(x,y)=h(x)y|C(x)zξ|p2(C(x)zξ)dz.(59) Since u(x,y)=0 on y=h+(x), as well, we find that g(C)=defh(x)h+(x)|C(x)yξ|p2(C(x)yξ)dy=0.Observe that g(C) is the gradient with respect to C of a the integral functional J(C)=defωF(x,C)dx,with convex Lagrangian F(x,C)=def1ph(x)h+(x)|C(x)yξ|pdy.Consequently using analogue of Lemma 3.1 we see that C is uniquely determined and the unique solution to the Euler equation is C(x)=m(x)ξ,where m(x) is defined as in (Equation1) and so u is uniquely determined, and finally (Equation57) is deduced from (Equation59), using (Equation55).

To summarize this section, from Lemma 5.5, we deduce that q satisfies the boundary value problem (Equation25), hence q in W1,p(ω) is uniquely determined by pb. Thereafter it follows from Lemma 6.3 that u in Wyp(Ω) is also unique. Since all subsequential limits u and q are the same, we conclude that the whole sequence of solutions (ϵpuϵ,qϵ) to the problem (Equation11) two-scale converges to (u,q).

Acknowledgements

We thank the anonymous referees for many valuable comments and suggestions which improved the final version of this paper.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Marušić S, Marušić-Paloka E. Two-scale convergence for thin domains and its applications to some lower-dimensional models in fluid mechanics. Asymptot Anal. 2000;23(1):23–57.
  • Fabricius J, Miroshnikova E, Tsandzana A, et al. Pressure-driven flow in thin domains. Asymptot Anal. 2020;116(1):1–26.
  • Batchelor GK. An introduction to fluid dynamics. Cambridge: Cambridge University Press; 2000.
  • Hele-Shaw HS. The flow of water. Nature. 1898;58:34–36.
  • Huilgol RR. Fluid mechanics of viscoplasticity. Berlin: Springer; 2015.
  • Davis AMJ, Mai T-Z. Steady pressure-driven non-Newtonian flow in a partially filled pipe. J Non-Newton Fluid Mech. 1991;41(1–2):81–100.
  • Jang HK, Hong SO, Lee SB, et al. Viscosity measurement of non-Newtonian fluids in pressure-driven flows of general geometries based on energy dissipation rate. J Non-Newton Fluid Mech. 2019;274(104204):00–00.
  • Marušić-Paloka E. Mathematical modeling of junctions in fluid mechanics via two-scale convergence. J Math Anal Appl. 2019;480(1):25.
  • Pereira GG. Effect of variable slip boundary conditions on flows of pressure-driven non-Newtonian fluids. J Non-Newton Fluid Mech. 2009;157(3):197–206.
  • Aronsson G, Janfalk U. On Hele-Shaw flow of power-law fluids. European J Appl Math. 1992;3(4):343–366.
  • Mikelić A, Tapiéro R. Mathematical derivation of the power law describing polymer flow through a thin slab. RAIRO-Modél Math Anal Numér.. 1995;29(1):3–21.
  • Gilbert RP, Fang M. An obstacle problem in non-isothermal and non-Newtonian Hele-Shaw flows. Commun Appl Anal. 2004;8(4):459–489.
  • Nassehi V. Generalized Hele-Shaw models for non-Newtonian, non-isothermal flow in thin curved layers. IMA J Math Appl Bus Ind. 1996;7(1):71–88.
  • Kondic L, Palffy-Muhoray P, Shelley M. Models of non-Newtonian Hele-Shaw flow. Phys Rev E. 1996;54(5):4536–4539.
  • Paşa G. Some non-Newtonian effects in Hele-Shaw displacements. Rev Roumaine Math Pured Appl. 2016;61(4):293–304.
  • Reynolds O. On the theory of lubrication and its application to Mr. Beauchamp Tower's experiments. Phil Trans Roy Soc London. 1886;177:157–234.
  • Panton RL. Incompressible flow. 4th ed. Hoboken: Wiley; 2013.
  • Bayada G, Chambat M. The transition between the Stokes equations and the Reynolds equation: a mathematical proof. Appl Math Optim. 1986;14(1):73–93.
  • Bayada G, Boukrouche M, El-A. Talibi M. The transient lubrication problem as a generalized Hele-Shaw type problem. Z Anal Anwend. 1995;14(1):59–87.
  • Anguiano M, Suárez-Grau FJ. Nonlinear Reynolds equations for non-Newtonian thin-film fluid flows over a rough boundary. IMA J Appl Math. 2019;84(1):63–95.
  • Anguiano M, Suárez-Grau FJ. Homogenization of an incompressible non-Newtonian flow through a thin porous medium. Z Angew Math Phys. 2017;68(2):25.
  • Geymonat G, Suquet P. Functional spaces for Norton-Hoff materials. Math Meth Appl Sci. 1986;8(2):206–222.
  • Boyer F, Fabrie P. Mathematical tools for the study of the incompressible Navier-Stokes equation and related models. New York: Springer; 2013.
  • Fabricius J. Stokes flow with kinematic and dynamic boundary conditions. Quart Appl Math. 2019;77(3):525–544.
  • Zeidler E. Nonlinear functional analysis and its applications II/B. New York: Springer; 1990.
  • Bogovski ME. Solution of the first boundary value problem for the equation of continuity of an incompressible medium. Dokl Akad Nauk SSSR. 1979;248(5):1037–1040.
  • Nečas J. Direct methods in the theory of elliptic equations. New York: Springer; 2012.
  • Tartar L. Topics in nonlinear analysis. Orsay: Publications Mathématiques d'Orsay, Université de Paris-Sud; 1978.
  • Barnes HA, Hutton JF, Walters K. An introduction to rheology. Amsterdam: Elsevier; 1989.
  • Astarita G, Marrucci G. Principles of non-Newtonian fluid mechanics. Maidenhead: McGraw-Hill; 1974.
  • Grisvard P. Elliptic problems in nonsmooth domains. Boston (MA): Pitman; 1985.
  • Lindqvist P. On the equation div (|∇u|p−2∇u)+λ|u|p−2u=0. Proc Amer Math Soc. 1990;109(1):157–164.
  • Brezis H. Analyse fonctionnelle - Théorie et applications. Paris: Masson; 1983.

Appendices

Appendix 1. Proof of Theorem 2.1

In this appendix, we generalize the Korn inequalities proved in [Citation2], where the case p = 2 was considered. Let □ be the cube in R3 defined by =(1/2,1/2)3 and let ϵ be the cube defined by ϵ=(ϵ/2,ϵ/2)3. Moreover we define the planes Γ±={(x1,x2,x3)R3:x3=±1/2}and Γ±ϵ={(x1,x2,x3)R3:x3=±ϵ/2}.

Lemma A.1

Let K1 and K2 denote the best constants in the inequalities (A1) ||v||Lp()K1||e(v)||Lp()(A1) (A2) ||v||Lp()K2||e(v)||Lp()(A2) for all vW1,p(,R3) such that v = 0 on Γ+Γ. Then (A3) ||v||Lp(ϵ)ϵK1||e(v)||Lp(ϵ)(A3) (A4) ||v||Lp(ϵ)K2||e(v)||Lp(ϵ)(A4) for all vW1,p(ϵ,R3) such that v = 0 on Γ+ϵΓϵ.

Proof.

Suppose vWp(ϵ) and define vˆ in Wp() as the function which satisfies v(x)=vˆ(x/ϵ),(x).Making the change of variable xˆ=x/ϵ and applying (EquationA1) to vˆ we obtain ϵv(x)pdx=|vˆ(xˆ)|pϵ3dxˆK1pe(vˆ)(xˆ)pϵ3dxˆ=K1pϵe(ϵv)(x)pdx,which implies (EquationA3). Similarly, but using (EquationA2) instead of (EquationA1), we obtain (EquationA4).

To estimate the constant in the Korn inequality under anisotropic scaling of the domain, we use an extension operator in combination with a covering argument. In view of assumptions in (Equation1) we can deduce that Ωϵω×(ϵ/2,ϵ/2),is also thin domain, so that we can consider the unscaled domain Ω defined as Ω=ω×(1/2,1/2).Indeed, due to the Dirichlet condition on ΓD±, we can extend any vWp(Ω) by zero to ω×(1/2,1/2), which is confined between the hyperplane Γ±.

Finally, we can assume a sufficiently large domain Ω~ of Ω given by Ω~=(L1,L+1)2×(1/2,1/2),such that ΩΩ~.

A.1. Extension operator

By consideration above, we prove the result

Lemma A.2

There exists a bounded linear operator E:Wp(Ω)Wp(Ω~),such that:

  1. Ev = v, for all v in Wp(Ω);

  2. for some positive constant Cω which depends only on ω, it holds e((Ev))Lp(Ω~)Cωe(v)Lp(Ω),for all v in Wp(Ω).

Proof.

Since Ω is a bounded open subset of R3 with Lipschitz boundary, there exists a continuous linear extension operator E from Wp(Ω) into Wp(Ω~), such that Ev(xy)=v(x,y),for a.e (x,y)Ω.See for instance, [Citation31, p. 25. Theorem 1.4.3.4], whence (i) follows. To show (ii), first we note that e(v)Lp(Ω)vLp(Ω).Applying this inequality to Ev yields e((Ev))Lp(Ω~)(Ev)Lp(Ω~)EvW1,p(Ω~).By the continuity of the extension operator, there exists a constant C1>0, which depends only on Ω, such that EvW1,p(Ω~)C1vW1,p(Ω).From the Korn inequality for W1,p(Ω) (see [Citation24, Theorem 4.8]) there exists a positive constant C2 depends only in Ω, vW1,p(Ω)C2e(v)Lp(Ω).Hence, e((Ev))Lp(Ω~)Cωe(v)Lp(Ω),where Cω=C1C2, and thus, we conclude the proof of the lemma.

A.2. Covering argument

For zR2, let Q(z,ϵ) denote the square in R2 of length ε centered in z, i.e. Q(z,ϵ)={yR2: |yizi|<ϵ/2,i=1,2}.For each 0<ϵ1, there exist a finite collection of points {zi} in ϵZ2 such that ωiQ(zi,ϵ)¯(L1,L+1)2.Define the cube iϵ=Q(zi,ϵ)×(ϵ/2,ϵ/2).Then Ωϵiiϵ¯Ω~,for all 0<ϵ1. Since each cube iϵ is a translation of the cube ϵ=(ϵ/2,ϵ/2)3 we can use the inequality (EquationA3) locally in Lemma A.1.

Proof

Proof of Theorem 2.1

Suppose vWp(Ωϵ). To use the extension operator in Lemma A.2, let v¯ in Wp(Ω) be defined by extension v¯=v,in Ωϵ0,in ΩΩϵ.From Lemma A.1(i), the covering argument and inequality (EquationA3) of Lemma A.2 it follows that Ωϵ|v|pdx=ΩϵEv¯pdxi=1iϵEv¯pdxi=1ϵpk1piϵe((Ev¯))pdxϵpk1pΩ~e((Ev¯))pdx,or vLp(Ωϵ)ϵK1e((Ev¯))Lp(Ω~).This together with Lemma A.2(ii) yields vLp(Ωϵ)ϵK1Cωe((v¯))Lp(Ω)=ϵK1Cωe((v))Lp(Ωϵ).This proves the first inequality in Theorem 2.1. The second inequality is proved in a similar way.

Appendix 2. Proof of Theorem 2.2

We will need a simple scaling  argument to prove the upper bound of norm of the Bogovski operator. To obtain a lower bound we consider an eigenvalue problem.

Simple scaling argument

For any 0<ϵ1, let Sϵ denote the linear scaling operator defined by Sϵf(x)=f(x,ϵ1x3)(f isascalarfunction)Sϵv(x)=(v(x,ϵ1x3),ϵv3(x,ϵ1x3))(v is a vector function),where x=(x,x3)R2×R and similarly for v=(v,v3). We note that Sϵ is invertible with (Sϵ)1=S1/ϵ. Using a simple change of variables we show that Sϵ maps Lp(Ω) onto Lp(Ωϵ) with 1|Ωϵ|1/pSϵfLp(Ωϵ)p=1|Ω|1/pfLp(Ω)p.Moreover, Sϵ maps Wp(Ω) onto Wp(Ωϵ) such that div(Sϵv)=Sϵ(divv)in Ωϵ.In particular, we define Sϵ acting on {v}Wp(Ω)/Vp(Ω), as Sϵ{v}=def{Sϵv}.It is readily checked that Sϵ{v} is well defined in Wp(Ω)/Vp(Ω).

Lemma A.3

Let B:Lp(Ω)Wp(Ω)/Vp(Ω),Bϵ:Lp(Ωϵ)Wp(Ωϵ)/Vp(Ωϵ)denote the Bogovski operators for Ω and Ωϵ respectively. Then (A5)  SϵB(Sϵ)1f=Bϵf fLp(Ωϵ)(A5) (A6) (Sϵ)1BϵSϵf=Bf  fLp(Ω).(A6)

Proof.

The statements (EquationA5) and (EquationA6) are equivalent, so it suffices to prove the first one. Given fLp(Ωϵ), suppose that u in Wp(Ωϵ) satisfies divu=fin Ωϵ{u}=Bϵf.Then (Sϵ)1uWp(Ω) satisfies div((Sϵ)1u)=(Sϵ)1(divu)=(Sϵ)1f  (Sϵ)1{u}=B(Sϵ)1f.Inverting, the last relation we see that Bϵf={u}=SϵB(Sϵ)1f.

Lemma A.4

For all v in Wp(Ω) and all 0<ϵ1 we have 1|Ωϵ|1/pSϵ{v}Wp(Ωϵ)/Vp(Ωϵ)ϵ1K2|Ω|1/p{v}Wp(Ω)/Vp(Ω),where K2 is the constant in Theorem 2.1.

Proof.

From (Sϵv)= xvϵ1vyϵxv3v3y(x,ϵ1x3),we obtain |(Sϵv)(x)|pϵp|v(x,ϵ1x3)|p.Integrating this inequality over Ωϵ gives Ωϵ|(Sϵv)(x)|pdxϵpΩϵ|v(x,ϵ1x3)|pdxdx3=ϵp|Ωϵ||Ω|Ω|v(x,y)|pdxdy.By the Korn inequality (Equation7) and some rearrangement, we take infimum on both sides, the lemma is proved.

Proof

Proof of Theorem 2.2

Assume fLp(Ωϵ). From Lemma A.3 and Lemma A.4 we deduce, 1|Ωϵ|1/pBϵfWp(Ωϵ)/Vp(Ωϵ)=1|Ωϵ|1/pSϵB(Sϵ)1fWp(Ωϵ)/Vp(Ωϵ)ϵ1K2|Ω|1/pB(Sϵ)1fWp(Ω)/Vp(Ω)ϵ1BK2|Ω|1/p(Sϵ)1fWp(Ω)/Vp(Ω)=ϵ1BK2|Ωϵ|1/pfLp(Ωϵ),which implies the upper bound in (Equation8) BΩϵK2ϵ1BΩ.To prove the lower bound, we consider the eigenfunction ψ in W01,p(ω) corresponding to the eigenvalue problem Δpψ=λ|ψ|p2ψin ωψ=0on ω.where Δp=div(|ψ|p2ψ), and λ is the smallest eigenvalue of Δp on ω. It is well know that λ can be characterized as the minimum of the Rayleigh quotient λ=infψψLp(ω)ψLp(ω),see [Citation32] for more details. Now suppose u in Wp(Ωϵ) satisfies divu=ψin Ωϵ.Since u = 0 on ΓDϵ and ψ=0 on ΓNϵ we have (A7) Ωϵ|ψ|pdx=Ωϵ(|ψ|p2ψ)divudx=Ωϵ(|ψ|p2ψ)udxΩϵ|(|ψ|p2ψ)|pdx1/pΩϵ|u|pdx1/p.(A7) Using Ωϵ|(|ψ|p2ψ)|pdx=λΩϵ|ψ|pdx,and applying the Korn inequality (Equation6) in the last integral on right hand side of (EquationA7) we deduce that Ωϵ|ψ|pdx1/pK1ϵλ1/pΩϵ|e(u)|pdx1/p,for all uWp(Ωϵ) such that divu=ψ. Consequently, taking the infimum on the last integral we obtain ψLp(Ωϵ)K1ϵλ1/p{u}Wp(Ωϵ)/Vp(Ωϵ).Estimating the right hand side with the norm of the Bogovski operator, we see that {u}Wp(Ωϵ)/Vp(Ωϵ)=BΩϵψWp(Ωϵ)/Vp(Ωϵ)BΩϵψLp(Ωϵ),whence, we deduce, 1K1λ1/pϵ1BΩϵ.This proves the lower bound in (Equation8) with C1=1K1λ1/p.

Appendix 3. Proof of Lemma 3.1

Proof

Proof of Lemma 3.1

For fixed ϵ>0, consider the minimization problem J(uϵ)=minvVp(Ωe)J(v)J(v)=Ωϵ2μ0pe(v)pfvdxΓNϵgvdS

(1) Let us begin to show that J(v) is weakly coercive on W1,p(Ωϵ). By the Young inequality with xyrpp|x|p+rpp|y|p,for appropriate r>0, the Trace theorem (see [Citation23, p. 153, Sec. 2.5]) and the Korn inequality (Equation6), we obtain J(v)Ωϵ2μ0rpC+1ϵK1|e(v)|ppdxrppΩϵfpdx+ΓNϵgpdx,where C is a constant from the Trace inequality in the Trace theorem. We choose rp=ϵK1μ0/(C+1), to obtain that J(v)μ0p||v||Vp(Ωϵ)pKp||f||Lp(Ωϵ)p+||g||Lp(ΓNϵ)p.where K is a positive constant which depends on Ωϵ. Hence, J(v) if ||v||Vp(Ωϵ), this proves the weak coercivity of J.

(2) Choose any minimizing sequence {uk}k=1 such that limkJ(uk)=infvVp(Ωϵ)J(v).Then the weak coercivity condition implies that {uk}k=1 is bounded in W1,p(Ωϵ), 1<p<, and since, W1,p(Ωϵ) is reflexive there exist a subsequence {ukj}j=1 and a function uϵ in W1,p(Ωϵ) such that ukjuϵweakly in W1,p(Ωϵ).Since Vp(Ωϵ) is a closed subspace of W1,p(Ωϵ), then Vp(Ωϵ) is weakly closed i.e. ukjuϵweakly in Vp(Ωϵ),and so uϵVp(Ωϵ).

(3) The weak lower semi continuity of J implies that J(uϵ)limjinfJ(ukj)=infvVp(Ωϵ)J(v),therefore uϵ is a minimizer of J over Vp(Ωϵ).

(4) We turn to deriving the corresponding Euler–Lagrange equation. Let uϵ in Vp(Ωϵ) be a minimizer of J and pick vVp(Ωϵ). Then uϵ+tv belongs the admissible class Vp(Ωϵ) for all tR. Let j(t)=defJ(uϵ+tv), then (A8) j(t)j(0)t=ΩϵLt(x)fvdxΓNϵgvdS,(A8) where, Lt(x)=def2μ0pe(uϵ+tv)pe(uϵ)pt.Note that, Lt(x)2μ0e(uϵ)p2e(uϵ):e(v)as t0.Consequently, by Dominated Convergence theorem, the limit (EquationA8) exist as t0, and thus, j(0)=Ωϵ(2μ0e(uϵ)p2e(uϵ):e(v)fv)dxΓNϵgvdS.Since j(0)j(t) for all t it must hold that j(0)=0, hence the minimizer uϵ satisfies the weak form of the Euler–Lagrange equation Ωϵ2μ0e(uϵ)p2e(uϵ):e(v)dx=Ωϵfvdx+ΓNϵgvdS.( 5) Finally, let us prove the uniqueness of the minimizer uϵ. Let u1ϵ and u2ϵ be two different minimizers, then by Korn's inequality (Equation6) we have e(u1ϵ)e(u2ϵ) on a set of positive measure. Since J is strictly convex it follows that J12u1ϵ+12u2ϵ<12J(u1ϵ)+12J(u2ϵ)=J(u1ϵ),which contradicts that u1ϵ is a minimizer, hence u1ϵ=u2ϵ.

Appendix 4. Characterization of the normalized limit pressure π0

Following [Citation33, p. 153. Proposition IX.3], we present a characterization of the subspace of W1,p(ω) to which the normalized limit pressure π0 belongs.

Lemma A.5

The following statements are equivalent

  1. π0Lp(ω) and there exists a constant C~ such that |xπ0,Θ|C~ΘLp(ω),for all ΘW1,p(ω;R2) such that Θnˆ=0 on γD;

  2. π0W1,p(ω) with π0=0 on γN.

Proof.

If π0W1,p(ω) with π0=0 on γN, then by the Divergence theorem we have that ωπ0ΘnˆdS=ω(xπ0Θ+π0divΘ)dx,for all ΘW1,p(ω;R2) such that Θnˆ=0 on γD. And therefore ωπ0divΘdx=ωxπ0Θdx.Hence, |xπ0,Θ|C~ΘLp(ω),where C~=xπ0Lp(ω). By density [Citation22, Lemma 2] this holds for all ΘLp(ω;R2) such that divΘLp(ω) and Θnˆ=0 on γD.

Conversely, assuming that π0Lp(ω), we can define the gradient of π0 as the linear functional xπ0W1,p(ω;R2), given by xπ0,Θ=ωπ0divΘdx,defined on the subspace {ΘW1,p(ω;R2): Θnˆ=0 on γD}Lp(ω;R2). By assumption |xπ0,Θ|C~ΘLp(ω),so, xπ0 is bounded. According to the Hahn–Banach theorem the xπ0 can be extended to a bounded linear functional on Lp(ω;R2). Thus by the Riesz representation theorem, there exists gLp(ω,R2) such that xπ0,Θ=ωgΘdx,for all ΘLp(ω;R2).Therefore, ωπ0divΘdx=ωgΘdx,for all ΘW1,p(ω;R2) such that Θnˆ=0 on γD. This implies that g=π0Lp(ω) and therefore π0W1,p(ω). Now by Divergence theorem we have γNπ0ΘnˆdS=ω(xπ0Θ+π0divΘ)dx=0,for all ΘW1,p(ω;R2) such that Θnˆ=0 on γD. We conclude that π0W1,p(ω) with π0=0 on γN.