Homogenization of stokes system using bloch waves

  • Received: 01 September 2016 Revised: 01 April 2017
  • Primary: 35P99, 35Q30, 47A75; Secondary: 49J20, 93C20, 93B60

  • In this work, we study the Bloch wave homogenization for the Stokes system with periodic viscosity coefficient. In particular, we obtain the spectral interpretation of the homogenized tensor. The presence of the incompressibility constraint in the model raises new issues linking the homogenized tensor and the Bloch spectral data. The main difficulty is a lack of smoothness for the bottom of the Bloch spectrum, a phenomenon which is not present in the case of the elasticity system. This issue is solved in the present work, completing the homogenization process of the Stokes system via the Bloch wave method.

    Citation: Grégoire Allaire, Tuhin Ghosh, Muthusamy Vanninathan. Homogenization of stokes system using bloch waves[J]. Networks and Heterogeneous Media, 2017, 12(4): 525-550. doi: 10.3934/nhm.2017022

    Related Papers:

    [1] Grégoire Allaire, Tuhin Ghosh, Muthusamy Vanninathan . Homogenization of stokes system using bloch waves. Networks and Heterogeneous Media, 2017, 12(4): 525-550. doi: 10.3934/nhm.2017022
    [2] Vivek Tewary . Combined effects of homogenization and singular perturbations: A bloch wave approach. Networks and Heterogeneous Media, 2021, 16(3): 427-458. doi: 10.3934/nhm.2021012
    [3] Carlos Conca, Luis Friz, Jaime H. Ortega . Direct integral decomposition for periodic function spaces and application to Bloch waves. Networks and Heterogeneous Media, 2008, 3(3): 555-566. doi: 10.3934/nhm.2008.3.555
    [4] 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
    [5] Hiroshi Matano, Ken-Ichi Nakamura, Bendong Lou . Periodic traveling waves in a two-dimensional cylinder with saw-toothed boundary and their homogenization limit. Networks and Heterogeneous Media, 2006, 1(4): 537-568. doi: 10.3934/nhm.2006.1.537
    [6] Patrizia Donato, Florian Gaveau . Homogenization and correctors for the wave equation in non periodic perforated domains. Networks and Heterogeneous Media, 2008, 3(1): 97-124. doi: 10.3934/nhm.2008.3.97
    [7] Hakima Bessaih, Yalchin Efendiev, Florin Maris . Homogenization of the evolution Stokes equation in a perforated domain with a stochastic Fourier boundary condition. Networks and Heterogeneous Media, 2015, 10(2): 343-367. doi: 10.3934/nhm.2015.10.343
    [8] Patrick Henning . Convergence of MsFEM approximations for elliptic, non-periodic homogenization problems. Networks and Heterogeneous Media, 2012, 7(3): 503-524. doi: 10.3934/nhm.2012.7.503
    [9] Xavier Blanc, Claude Le Bris . Improving on computation of homogenized coefficients in the periodic and quasi-periodic settings. Networks and Heterogeneous Media, 2010, 5(1): 1-29. doi: 10.3934/nhm.2010.5.1
    [10] Grigor Nika, Adrian Muntean . Hypertemperature effects in heterogeneous media and thermal flux at small-length scales. Networks and Heterogeneous Media, 2023, 18(3): 1207-1225. doi: 10.3934/nhm.2023052
  • In this work, we study the Bloch wave homogenization for the Stokes system with periodic viscosity coefficient. In particular, we obtain the spectral interpretation of the homogenized tensor. The presence of the incompressibility constraint in the model raises new issues linking the homogenized tensor and the Bloch spectral data. The main difficulty is a lack of smoothness for the bottom of the Bloch spectrum, a phenomenon which is not present in the case of the elasticity system. This issue is solved in the present work, completing the homogenization process of the Stokes system via the Bloch wave method.



    We consider the Stokes system in which the viscosity is a periodically varying function of the space variable with small period ϵ>0. Many physical phenomena (boiling flows, porous media, oil reservoirs, etc.) lead to mixture of fluids with different viscosities. For incompressible slow or creeping flows, such a situation is modeled by the system (1) for a Stokesian fluid with variable viscosity which is further assumed to be a periodic function. From the point of view of application, it is difficult to realize such a periodic distribution of droplets of one fluid in another without deforming the periodic structure, and (1) may seem as too much of an idealized system. Therefore, we also treat another model, which is a variant of the Stokes system and is physically more relevant. Namely, we consider the so-called incompressible elasticity system (10) which corresponds to a mixture of incompressible elastic phases in a composite material (this situation is quite common for rubber or elastomers).

    The first goal of this paper is to study homogenization of the above systems via Bloch Wave Method which is based on the fact that the homogenized operator can be defined using differential properties of the bottom of the so-called Bloch spectrum. The second goal of the paper is to explore this regularity issue which is delicate for the systems under consideration because of the presence of the incompressibility condition. These points are elaborated below.

    Through out the paper, we will follow the usual Einstein summation convention with respect to repeated indices. We introduce now our first model. Assuming that the viscosity (denoted by μ in the sequel) is a periodic function, the goal is to capture the effective viscosity of the mixture. To write down the model we start with a 1-periodic function μ=μ(y)L(Td) or equivalently, a Y-periodic function where Y=]0,1[d which represents the viscosity of the fluids. Here Td is the unit torus in Rd. We assume μ(y)μ0>0 a.e in Td. Denote by μϵ=μϵ(x)=μ(xϵ) the corresponding scaled function which is ϵ-periodic. With f=f(x)L2(Ω)d representing external force, we consider the Stokes system with no-slip boundary condition in a bounded connected nonempty open set ΩRd having Lipschitz boundary, :

    (μϵuϵ)+pϵ=f in Ω,uϵ=0 in Ω,uϵ=0 on Ω.} (1)

    As usual, uϵ and pϵ represent respectively the velocity and pressure fields of the fluid. Well-posedness theory of (1) is classical [15]. We recall some of its elements. To write down the weak formulation, we introduce the spaces

    V={vH10(Ω)d;v=0 in Ω}, (2)

    Here ν denotes unit outward normal to Ω. Multiplying (1) by vV gives the following problem for uϵ which does not involve pϵ: Find uϵV satisfying

    Ωμϵuϵv dx=Ωfv dxvV. (3)

    The classical Lax-Milgram Lemma (essentially, Riesz Representation Theorem due to the symmetry of our bilinear form) ensures existence and uniqueness of a solution uϵV for (3). To get the pressure field one applies de Rham's Theorem in the following form [15]:

    V:={wH1(Ω)d; w,vH1(Ω)d,H10(Ω)d=0,vV}={p;  pL2(Ω)}, (4)

    which implies that the pressure pϵ in (1) belongs to L2(Ω). Since Ω is a connected set, the pressure is defined up to an additive constant. To guarantee the uniqueness of the pressure, we seek p in the space L20(Ω)={fL2(Ω):Ωf dx=0} with L2 norm. Moreover, by using Poincaré inequality and inf-sup inequality [15], one shows that the solution (uϵ,pϵ)(H10(Ω))d×L20(Ω) of (1) are uniformly bounded, namely there exists a constant C, independent of ϵ, such that

    ||uϵ||(H10(Ω))d+||pϵ||L2(Ω)C||f||(L2(Ω))d. (5)

    We are interested here in the homogenization limit of (1), that is the asymptotic limit of the solution (uϵ,pϵ) as ϵ0. This problem is very classical and its solution by means of a combination of two-scale asymptotic expansions and the method of oscillating test functions was provided in various references, including [8,16,20]. We recall their main results and follow the notations of [8,chapter Ⅰ,section 10]. The homogenized tensor (A)klαβ, which represents " effective viscosity", is defined by

    (A)klαβ=Tdμ(y)(χkα+yαek):(χlβ+yβel) dy, (6)

    in which figure the cell test functions {χkα; α,k=1d}H1(Td)d and {Πkα; α,k=1d}L2(Td) solutions of the following problem in the unit torus Td:

    (μ(χkα+yαek))+Πkα=0 in Tdχkα=0 in Td(χkα,Πkα)is Yperiodic. } (7)

    We impose Tdχkα dy=TdΠkα dy=0 to obtain uniqueness of the solutions. It is easy to see that the above homogenized tensor possesses the following ''simple" symmetry, for any indices 1α,β,k,ld,

    (A)klαβ=(A)lkβα, (8)

    which corresponds to the fact that the fourth-order tensor A is a symmetric linear map from the set of all matrices (or second-order tensors) into itself. Since we follow the notations of [8], the simple symmetry (8) seems a bit awkward since it mixes Latin and Greek indices but it is just the usual symmetry for a pair of indices (k,α) and (l,β) in a fourth-order tensor. In other words, (8) holds for a simultaneous permutation of k,l and α,β. It is straight-forward (see [8]) to check that the tensor A is positive-definite so that the following system (9) is well-posed.

    Theorem 1.1. The homogenized limit of the problem (1) is

    xβ((A)klαβukxα)+pxl=fl  in Ω, for l=1,2,...,d,u=0in Ω,u=0 in Ω.} (9)

    More precisely, we have the convergence of solutions:

    (uϵ,pϵ)(u,p) in H10(Ω)d×L20(Ω) weak

    Note that the simple symmetry (8) does not imply that A is symmetric in k,l or in α,β. However, in the homogenized equation (9), since A is constant, only its symmetric version, obtained by symmetrizing in both k,l and α,β, plays a role.

    Let us next consider the second model of incompressible elasticity :

    (μϵE(uϵs))+pϵs=f in Ω,uϵs=0 in Ω,uϵs=0 on Ω.} (10)

    Here the strain rate tensor is given by

    E(v)=12(v+tv) namely Ekl(v)=12(vkxl+vlxk).

    As before, there exists a unique solution (uϵs,pϵs) of the above problem (10) in (H10(Ω))d×L20(Ω) and using Korn's inequality and the inf-sup inequality, the following uniform bound can be proved :

    ||uϵs||(H10(Ω))d+||pϵs||L2(Ω)C||f||(L2(Ω))d, (11)

    where the constant C does not depend on ϵ. Here the homogenized tensor (As)klαβ is given by

    (As)klαβ=Tdμ(y)E(˜χkα+yαek):E(˜χlβ+yβel) dy (12)

    where the cell test functions ˜χkαH1(Td),˜ΠkαL2(Td) are now solutions in the torus Td of

    (μE(˜χkα+yαek))+˜Πkα=0 in Td˜χkα=0 in Td(˜χkα,˜Πkα)is Y periodic } (13)

    We impose Td˜χkα=Td˜Πkα=0. It is known [10] that the above homogenized tensor possesses the following "full" symmetry, for any indices 1α,β,k,ld,

    (As)klαβ=(As)αlkβ=(As)kβαl=(As)lkβα, (14)

    which corresponds to the fact that the fourth-order tensor As is a symmetric linear map from the set of all symmetric matrices into itself (the conditions (14) are the usual symmetry conditions for Hooke's laws in linearized elasticity). The homogenization limit of the problem (10) is again of the form (9) with As replacing A.

    The first goal of this paper is to give an alternate proof of Theorem 1.1 using the Bloch Wave Method instead of two-scale asymptotic expansions and the method of oscillating test functions. The notion of Bloch waves is well-known in physics and mathematics [8,11,19,22]. Bloch waves are eigenfunctions of a family of "shifted" spectral problems in the unit cell Y for the corresponding differential operator. Its link with homogenization theory was first explored in [8,13,17]. The key point is that the homogenized operator can be defined in terms of differential properties of the bottom of the Bloch spectrum. The second goal of this paper is to explore this issue which is especially delicate in the case of Stokes equations. Indeed, it was discovered in [7] that the Bloch spectrum for the Stokes equations is not regular enough at the origin because of the incompressibility constraint. Therefore, its differential properties are all the more intricate to establish. Here we complete the task started in [7] and in particular we prove a conjecture of [7] on the homogenization of the Stokes system (1). Since the treatment of the incompressible elasticity system (10) is almost analogous to that of (1), we focus on (1) and we content ourselves in highlighting the main differences for (10) throughout the sequel.

    The Bloch wave method for scalar equations and systems without differential constraints (like the incompressibility condition) was studied in [12,13,14,21]. In such cases, this approach gives a spectral representation of the homogenized tensor A=(A)klαβ in terms of the lowest energy Bloch waves and their behaviour for small momenta (what we call the bottom of the spectrum). For instance, the homogenized matrix in the scalar case was found to be equal to one -half of the Hessian of the ground energy (or first eigenvalue) at zero momentum. For a system, several bottom eigenvalues play a role and they are merely directionally differentiable by lack of simplicity. In the present case of the Stokes system, the situation is more complicated. The main characteristic of the Stokes system is the presence of the differential constraint expressing incompressibility of the fluid. One of its effects is that the Bloch energy levels are degenerate and the corresponding eigenfunctions are discontinuous at zero momentum. Even though energy levels are continuous at zero momentum, the second order derivatives are not (cf. Theorem 3.1). Thus, we cannot really make sense of the eigenvalue Hessian at zero momentum. Further, it is not clear if the homogenized tensor can be fully recovered from the Bloch spectral data. In fact, this issue is left open in [7]. In the non-self adjoint case treated in [21], only the symmetric part of the homogenized matrix is determined by Bloch spectral data and this is enough to determine the homogenized operator uniquely. Combining all these difficulties, the homogenization of Stokes system using Bloch waves is an interesting issue which is not a direct extension of previous results. Our work, roughly speaking, shows that Bloch spectral data does not determine the homogenized tensor uniquely, but determines the homogenized operator uniquely. This is in sharp contrast with the linear elasticity system treated in [14] in which the homogenized tensor was uniquely determined from Bloch spectral data. We see thus the effect of differential constraints (the incompressibility condition in the case of Stokes equations) on the homogenization process via Bloch wave method. For further discussion on this point, see Section 4. Bloch wave method of homogenization presented in Section 5 consists of localizing (1), taking its Bloch transform and passing to the limit to get the localized version of homogenized system in the Fourier space. Passage to the limit in the Bloch method is straight forward, though arguments are long. We do not run into the classical difficulty of having a product of two weakly convergent sequences. In fact, we use the Taylor approximation of Bloch spectral elements which gives strongly convergent sequences. This is one of the known features of the method. The required homogenized system is obtained by making a passage to the physical space from the Fourier space. Extracting macro constitutive relation and macro balance equation from the localized homogenized equation in the Fourier space turns out to be not very straight forward because of differential constraints.

    Let us end this discussion with two general remarks on Bloch wave method. First one is about the nature of convergence of the homogenization process. It is well-known in the homogenization theory that the convergence in Theorem 1.1 is only weak and not strong. To have strong convergence, we need the so-called correctors [8]. Within Block wave theory, correctors are discussed in [12] for the scalar equation. We do not construct explicitly correctors for Stokes system in this paper, even though all necessary ingredients are presented. Because of the lack of smoothness of the bottom level Bloch spectrum, corrector issue is worth considering separately. The second remark is about non-periodic coefficients. Bloch wave approach to homogenization is well developed only in the case of periodic coefficients. It is known that for some restricted class of locally-periodic/modulated coefficients, new phenomena (like localization) may appear [2,3,4]. We are not aware of a Bloch wave approach for more general coefficients.

    The plan of this paper is as follows. In section 2, we recall from [7] the properties of Bloch waves associated with the Stokes operator. It turns out that the Bloch waves and their energies can be chosen to be directionally regular, upon modifying the spectral cell problem at zero momentum. Bloch transform using eigenfunctions lying at the bottom of the spectrum is also introduced in this section. Its asymptotic behaviour for low momenta is also described. Next, Section 3 is devoted to the computation of directional derivatives of Bloch spectral data. Even though these results are essentially borrowed from [7], some new ones are also included because of their need in the sequel. In particular we derive the so-called propagation relation linking the homogenized tensor A with Bloch spectral data, and the extent to which it determines homogenized tensor is studied in Section 4. Using this information, we prove Theorem 1.1 in Section 5 following the Bloch wave homogenization method.

    Note added in proof. At the end of the introduction we claimed that we were not aware of a Bloch wave approach for non-periodic coefficients. We recently learned about a new work in this direction by A. Benoit and A. Gloria, ''Long-time homogenization and asymptotic ballistic transport of classical waves", which will appear in Annales Scientifiques de l'Ecole Normale Supérieure.

    In this section, we introduce Bloch waves associated to the Stokes operator following the lead of [7]. The Bloch waves are defined by considering the shifted (or translated) eigenvalue problem in the torus Td parametrized by elements in the dual torus which we take to be 2πTd. We denote by y the points of the original torus and by η the points of the dual torus. The spectral Bloch problem amounts to find λ=λ(η)R, ϕ=ϕ(η)(H1(Td))d, with ϕ0 and Π=Π(η)L2(Td), satisfying

    D(η)(μD(η)ϕ)+D(η)Π=λ(η)ϕ in Td,D(η)ϕ=0 in Td,(ϕ,Π) is Y periodic,Y|ϕ|2dy=1.} (15)

    The solutions of (15) ϕ,Π are a priori complex valued, so all functional spaces are complex valued too. Here, we denote

    D(η)=y+iη

    the shifted gradient operator, with i the imaginary root 1. Its action on a vector function ϕ yields a matrix: (D(η)ϕ)kl=ϕlyk+iηkϕl for all k,l=1,,d. The corresponding divergence operation yields a scalar: D(η)ϕ=ϕkyk+iηkϕk. Analogously, if ϕ is a matrix function then its shifted divergence D(η)ϕ is a vector function obtained by acting D(η) on the column vectors of ϕ. At this stage of discussion, spectral problem (15) is stated only formally. Rigourous versions of it with modification will be formulated and used below. For reasons of self-adjointness, its eigenvalues will be real and that is why, we take λ to be real, without any loss of generality. Though we may think of eigen-solution Π as some sort of complex-valued pressure, there is no guarantee that it represents physical pressure field of a real fluid. Let us remark that the system (15) is only a mathematical model which appears as an useful intermediate step in the study of a physical model.

    The main feature of (15) is that the state space keeps varying with η due to the differential constraints defined by the incompressibility of the fluid. That is why, the standard spectral theory for elliptic operators does not apply as such; it has to be modified. This is accomplished in [18]. Secondly, it is easily seen that when η=0, the corresponding eigenvalue λ(0) is equal to zero and its multiplicity is d. In fact, we can take ek,k=1d as eigenvectors (with corresponding eigen-pressure being zero). Because of this degeneracy, spectral elements of (15) are not guaranteed to be smooth at η=0.

    As discussed in introduction, lack of regularity of the Bloch spectrum at η=0 is an issue because the representation of the homogenized tensor in terms of Bloch spectral elements is then not clear. This was not the case for the scalar problem, the lack of regularity of the Bloch spectrum at η=0 does not appear there, in fact the Bloch spectrum is analytic near η=0, see [13].To overcome the difficulty in the present case, the idea is to consider directional regularity as we approach η=0 [14]. Accommodating the directional limit at η=0 requires a modification of the above shifted problem with the addition of a new constraint and corresponding Lagrange multiplier in the equation [7]. Fixing a direction eRd,|e|=1 and taking η=δe, with δ>0, we consider the modified problem: find λ(δ)R, ϕ(.;δ)(H1(Td))d, q(.;δ)L20(Td) where L20(Td)={qL2(Td); Tdq=0} and q0(δ)C satisfying

    D(δe)(μ(y)D(δe)ϕ(y;δ))+D(δe)q(y;δ)+q0(δ)e=λ(δ)ϕ(y;δ) in Td,D(δe)ϕ(y;δ)=0 in Td,eTdϕ(y;δ) dy=0,(ϕ,q) is Y periodic,Td|ϕ(y;δ)|2 dy=1.} (16)

    Note that if δ0 then the relation eTdϕ(.;δ)=0 can be easily obtained from D(δe)ϕ(.;δ)=0 simply by integration. (However, this is not the case if δ=0.) Hence (16) is the same as (15) provided δ0 and η=δe. However for δ=0, (15) is not good because the condition eTdϕ(.;δ)=0 is not included. See [7] on the appearance of this new constraint and the corresponding Lagrange multiplier q0(δ)e.

    It is natural to consider the system (16) with δ small as a perturbation from the following one which corresponds to δ=0. We fix a unit vector ˆηSd1 and we consider the eigenvalue problem: find ν(ˆη)R, w(.,ˆη)(H1(Td))d, q(.;ˆη)L20(Td) and q0(ˆη)C satisfying

    (μw)+q+q0ˆη=ν(ˆη)w in Td,w=0 in Td,ˆηYw dy=0,(w,q) is Y periodic,Y|w|2 dy=1.} (17)

    Existence of eigenvalues and eigenvectors for either (16) or (17) is proved in [7]. Let us recall their result, by specializing to the eigenvalue ν(ˆη)=0 of (17). Note that ν(ˆη)=0 is clearly an eigenvalue of multiplicity (d1) of (17) with corresponding eigenfunctions being constants, namely q0m,ˆη=0, q00,m,ˆη=0 and ϕ0m,ˆη(y) is a constant unit vector of Rd orthogonal to ˆη for m=1,,(d1), say {ϕ01,ˆη,ϕ0d1,ˆη}. Doing perturbation analysis of the above situation, the following result was proved in [7].

    Theorem 2.1. Fix ˆηSd1. Consider the first (d1) eigenvalues of (16). There exists δ0>0 and exactly (d1) analytic functions defined in the real interval |δ|δ0, δ(λm,ˆη(δ), ϕm,ˆη(.;δ), qm,ˆη(.;δ), q0,m,ˆη(δ)), for m=1,,(d1), with values in R×(H1(Td))d×L20(Td)×C, such that

    (ⅰ) λm,ˆη(δ)|δ=0=0, ϕm,ˆη(.;δ)|δ=0=ϕ0m,ˆη, qm,ˆη(.;δ)|δ=0= q0,m,ˆη(δ)|δ=0=0,

    (ⅱ) (λm,ˆη(δ), ϕm,ˆη(.;δ), qm,ˆη(.;δ), q0,m,ˆη(δ) ) satisfies (16).

    (ⅲ) The set {ϕ1,ˆη(.;δ),.ϕ(d1),ˆη(.;δ)} is orthonormal in (L2(Td))d.

    (ⅳ) For each interval IR with ¯I containing exactly the eigenvalue ν(ˆη)=0 of (17) (and no other eigenvalue of (17) then {λ1,ˆη(δ),λ(d1),ˆη(δ)} are the only eigenvalues of (16 (counting multiplicities) lying in the interval I.

    The above theorem says that there are (d1) smooth curves emanating out of the zero eigenvalue as δ varies in an interval (δ0,δ0). We call them Rellich branches. Using them, for m=1,,(d1), we can define the corresponding mth Bloch transform of g(L2(Rd))d via the expression

    Bϵm,ˆηg(ξ)=Rdg(x)¯ϕm,ˆη(xϵ,δ)eixξ dx, (18)

    where δ=δ(ϵ,ξ)=ϵ|ξ| and ˆη=ξ/|ξ|. This is well defined provided ϵ is sufficiently small so that ϵ|ξ|δ0. For other ξ, we define Bϵm,ˆηg(ξ)=0.

    For later purposes we need the Bloch transform for (H1(Rd))d elements also. Let us consider F(g0+dj=1xjgj)(H1(Rd))d, where F,g0,g1,...,gd are valued in Cd and gj((L2(Rd))d for j=0,1,...,d. Then we define Bϵm,ˆηF(ξ) in L2loc(Rdξ) by

    Bϵm,ˆηF(ξ):=Rdg0(x)¯ϕm,ˆη(xϵ;δ)eixξ dx+Rdidj=1ξjgj(x)¯ϕm,ˆη(xϵ;δ)eixξ dxϵ1Rddj=1gj(x)¯ϕm,ˆηyj(xϵ;δ)eixξ dx. (19)

    Definition (19) is independent of the representation used for F(H1(Rd))d in terms of {gj, j=0,...,d} and is consistent with the previous definition (18) whenever F(L2(Rd))d.

    Remark 1. Due to the property (eixξϕϵm,ˆη)=0 in Rd, we see from (19) that

    Bϵm,ˆη(F+ψ)(ξ)=Bϵm,ˆηF(ξ), for all ψL2(Rd). (20)

    In fact, by considering ψ=g0+dj=1xjgj(H1(Rd))d then we can take g0=0 and gj=ψej j=1,,d, in (19) to obtain (20). That is, Bloch transform of gradient field is zero. Therefore the kernel of the Bloch transform Bϵm,ˆη:L2(Rd)dL2(Rd) contains the closed subspace {ψ:ψH1(Rd)} for each m=1,,d1. Roughly speaking since Bloch waves satisfy incompressibility condition the Bloch transform on gradient field vanish. Thus we may anticipate that the pressure effects may not be captured in the Bloch method. This impression is not correct. Indeed, as shown Section 5, by means of localization via a cut-off function, we manage to keep the pressure term.

    Our next result is concerned with the asymptotic behavior of these Bloch transforms as ϵ0. In the physical space such convergence often modeled as two scale convergence (or, multi scale convergence) can be found in [5,1]. Here we will be studying such convergence in the Fourier space. Since ϕm,ˆη(y;0) is a fixed unit vector (=ϕ0m,ˆη) orthogonal to ˆη and independent of y (see Theorem 2.1), we have

    Theorem 2.2. Let gϵ be a sequence in (L2(Rd))d such that its support is contained in a fixed compact set KRd, independent of ϵ. If gϵ converges weakly to g in (L2(Rd))d, then we have

    χϵ1Td(ξ)Bϵm,ˆηgϵ(ξ)ϕ0m,ˆηˆg(ξ),weakly  in  L2loc(Rdξ) for 1md1 (21)

    where ˆg denotes the Fourier transform of g and we recall that ˆη=ξ|ξ|.

    Proof. Let us remark that Bϵm,ˆηgϵ(ξ) is defined for ϵδ0M if |ξ|M. We can write

    Bϵm,ˆηgϵ(ξ)=χϵ1Td(ξ)ϕ0m,ˆηgϵ(ξ)+Kgϵ(x)(¯ϕm,ˆη(xϵ;δ)¯ϕm,ˆη(xϵ;0))eixξ dx.

    By using Cauchy-Schwarz, the second term on the above right hand side can be estimated by the quantity

    CKϕm,ˆη(y;δ)ϕm,ˆη(y;0)(L2(Y))d

    where CK is a constant depending on K but not on ϵ. Recall that δ is a function of (ϵ,ξ), namely δ=ϵ|ξ|. This quantity is easily seen to converge to zero as ϵ0 for each fixed ξ because of the directional continuity of ϕm,ˆη(.,δ)ϕ0m,ˆη in (L2(Td))d as δ0. We merely use the continuity of the mth Rellich branch at δ=0 with values in (L2(Td))d. On the other hand, thanks to our normalization, the integral on K is bounded by a constant independent of (ϵ,ξ). The proof is completed by a simple application of the Dominated Convergence Theorem which guarantees that the second term on the above right hand side converges strongly to 0 in L2loc(Rdξ) as ϵ0.

    Since compactly supported elements are dense in (L2(Rd))d, we have the following:

    Corollary 1. In the setting of Theorem 2.2, if gϵ be a sequence in (L2(Rd))d such that its support is contained in a fixed compact set KRd, independent of ϵ and gϵg in L2(Rd)d then we have the following strong convergence

    χϵ1TdBϵm,ˆηgϵ(ξ)ϕ0m,ˆηˆg,strongly in L2loc(Rdξ) for 1md1. (22)

    We recall the classical orthogonal decomposition [15] :

    L2(Rd)d={ψ:ψH1(Rd)}{ϕL2(Rd)d:ϕ=0}. (23)

    Let us denote

    X={ψ:ψH1(Rd)}, so that, X={ϕL2(Rd)d:ϕ=0}. (24)

    By our choice, {ϕ01,ˆη,,ϕ0d1,ˆη,ˆη} forms an orthonormal basis in Rd, and so we can deduce the following :

    Proposition 1. If gX and ϕ0m,ˆηˆg=0 for all m=1,,d1, then g=0.

    Proof. The proof is immediate, as {ϕ01,ˆη,,ϕ0d1,ˆη} forms an orthogonal basis in Rd1 and ϕ0m,ˆηˆg=0 for all m=1,,d1, so ˆg(ξ)=c(ξ)ξ for some scalar cL2(Rd). Now if c0, it contradicts the hypothesis gX. Thus c=0. Consequently, g=0.

    Corollary 2. In the setting of Theorem 2.2 and Proposition 1 if gϵ be a sequence in XL2(Rd)d such that its support is contained in a fixed compact set KRd, independent of ϵ and gϵg in L2(Rd)d weak, and ϕ0m,ˆηˆg=0 for all m=1,,d1, then g=0.

    Proof. The proof simply follows as X is a closed subspace of L2(Rd)d, so the limit gX and the result follows by applying Proposition 1.

    Theorem 2.2 and its corollaries give some sufficient conditions and specify the sense in which the first Bloch transform tends to Fourier transform, which is a sign of homogenization on the Fourier side. We do not believe that these conditions are sharp and exhaust all possibilities. It would be interesting to explore further in this direction.

    Remark 2. Bloch waves being incompressible are transversal. Longitudinal direction is missing and it has to be added to get the full basis. Naturally, asymptotics of the Bloch transform contains information of the Fourier transform only in transversal directions. It contains no information in the longitudinal direction. Because of this feature, in the homogenization limit also, there is no information in the longitudinal direction. This is however proved to be enough to complete the homogenization process because the limiting velocity field is incompressible. See Section 5.

    In this section, we give the expressions of the derivatives (at δ=0) of the Rellich branches {ϕm,ˆη(y;δ),qm,ˆη(y;δ),q0,m,ˆη(δ),λm,ˆη(δ)} obtained in Theorem 2.2. These results are essentially borrowed from [7] except for the second order derivative of q0,m,ˆη(δ) which is new. We differentiate, with respect to δR, (16) or equivalently the following system, fixing m=1,,d1 and ˆη=ξ|ξ|Sd1,

    D(δˆη)(μ(y)D(δˆη)ϕm,ˆη(y;δ))+D(δˆη)qm,ˆη(y;δ)+q0,m,ˆη(δ)ˆη=λm,ˆη(δ)ϕm,ˆη(y;δ) in Td,D(δˆη)ϕm,ˆη(y;δ)=0 in Td,ˆηTdϕm,ˆη(y;δ) dy=0,(ϕm,ˆη,qm,ˆη) is Y periodic.} (25)

    Zeroth order derivatives : For m=1,,d1 and for a fixed direction ˆηSd1 we have λm,ˆη(0)=0 and a corresponding eigenfunction is such that qm,ˆη(y;0)=0,q0,m,ˆη(0)=0 and ϕm,ˆη(y;0) is a constant unit vector of Rd orthogonal to ˆη. We give a notation for this constant ϕm,ˆη(y;0)=ϕ0m,ˆη. We recall that {ϕ01,ˆη,,ϕ0d1,ˆη,ˆη} is such that they form an orthonormal basis for Rd.

    First order derivatives : Let us differentiate (25) once with respect to δ to obtain (prime denotes derivatives with respect to δ) :

    D(δˆη)(μ(y)D(δˆη)ϕm,ˆη(y;δ))+D(δˆη)qm,ˆη(y;δ)+q0,m,ˆη(δ)ˆηλm,ˆη(δ)ϕm,ˆη(y;δ)=f(δ) in Td,D(δˆη)ϕm,ˆη(y;δ)=g(δ) in Td,ˆηTdϕm,ˆη(y;δ) dy=0,(ϕm,ˆη,qm,ˆη) is Y periodic} (26)

    where,

    f(δ)=λm(δ)ϕm,ˆη(y;δ)iqm,ˆη(y;δ)ˆη+iˆημ(y)D(δˆη)ϕm,ˆη(y;δ)+iD(δˆη)(μ(y)ϕm,ˆη(y;δ)ˆη),g(δ)=iˆηϕm,ˆη(y;δ).

    We put δ=0 in (26) and by integrating over Td, we obtain

    q0,m,ˆη(0)ˆη=λm,ˆη(0)ϕ0m,ˆη.

    Taking scalar product with ˆη, we simply get λm,ˆη(0)=q0,m,ˆη(0)=0 as ˆηϕ0m,ˆη.

    Using the above information in (26), we find that (ϕm,ˆη(y;0),qm,ˆη(y;0)) is a solution of the following cell problem :

    (μ(y)ϕm,ˆη(y;0))+qm,ˆη(y;0)=i(μ(y)ϕ0m,ˆηˆη) in Td,ϕm,ˆη(y;0)=0 in Td,ˆηTdϕm,ˆη(y;0) dy=0,Tdqm,ˆη(y;0) dy=0(ϕm,ˆη(y;0),qm,ˆη(y;0)) is Y periodic.} (27)

    Comparing this with (7), it can be seen that that ϕm,ˆη(y;0) is given by (see [7]) :

    ϕm,ˆη(y;0)=iˆηαχrα(y)(ϕ0m,ˆη)r+ζm,ˆη (28)

    where ζm,ˆηCd is a constant vector (independent of y), orthogonal to ˆη. In other words, the y-dependence of ϕm,ˆη(y;0) is completely determined by the cell test function χrα(y), solution of problem (7).

    In a similar manner, the derivative of the eigenpressure qm,ˆη(y;0) is given by (see [7]):

    qm,ˆη(y;0)=iˆηαΠrα(y)(ϕ0m,ˆη)r, (29)

    That is, the y-dependence of qm,ˆη(y;0) is completely determined by the cell test function Πrα(y), solution of problem (7).

    Second order derivatives : Next we differentiate (26) with respect to δ to obtain :

    D(δˆη)(μ(y)D(δˆη)ϕm,ˆη(y;δ))+D(δˆη)qm,ˆη(y;δ)+q0,m,ˆη(δ)ˆηλm,ˆη(δ)ϕm,ˆη(y;δ)=F(δ) in Td,D(δˆη)ϕm,ˆη(y;δ)=G(δ) in TdˆηTdϕm,ˆη(y;δ) dy=0,(ϕm,ˆη,qm,ˆη) is Y periodic} (30)

    where

    F(δ)=2μ(y)ϕm,ˆη(y;δ)+2iˆημ(y)D(δˆη)ϕm,ˆη(y;δ)+2iD(δˆη)(μ(y)ϕm,ˆη(y;δ)ˆη)2iˆηqm,ˆη(y;δ)+λm,ˆη(δ)ϕm,ˆη(y;δ)+2λm,ˆη(δ)ϕm,ˆη(y;δ),G(δ)=2iˆηϕm,ˆη(y;δ). (31)

    We consider (30) at δ=0 and by integrating over Td, we get

    q0,m,ˆη(0)ˆηk=2Tdμ(y)(ϕ0m,ˆη)k dy2Td[ˆηβμ(y)yχlβ(y)(ϕ0m,ˆη)l]kαˆηα dy+λm,ˆη(0)(ϕ0m,ˆη)k

    or,

    12(q0,m,ˆη(0)ˆηkλm,ˆη(0)(ϕ0m,ˆη)k)=Tdμ(y)[δlkδαβ+(χlβ)kα]dy ˆηαˆηβ(ϕ0m,ˆη)l dy=Tdμ(y)[(yβel):(yαek)+χlβ:(yαek)] dy ˆηαˆηβ(ϕ0m,ˆη)l dy=(A)klαβˆηαˆηβ(ϕ0m,ˆη)l.=[(ϕ0m,ˆη)tM(ˆη,A)]k=[M(ˆη,A)(ϕ0m,ˆη)]k (33)

    where M(ˆη,A) is the symmetric matrix whose entries are given by

    M(ˆη,A)kl=(A)klαβˆηαˆηβ.

    This is nothing but a contraction of the homogenized tensor A. As a simple consequence of (32), we get

    12q0,m,ˆη(0)=M(ˆη,A)ϕ0m,ˆηˆη and 12λm,ˆη(0)=M(ˆη,A)ϕ0m,ˆηϕ0m,ˆη.

    It also follows that M(ˆη,A)ϕ0m,ˆηϕ0m,ˆη for all mm.

    By summarizing the above computations, we have

    Theorem 3.1. For m=1,,d1 and for a fixed direction ˆηSd1 we have

    (ⅰ) λm,ˆη(0)=0 and a corresponding eigenfunction is such that qm,ˆη(y;0)=0,q0,m,ˆη(0)=0 and ϕm,ˆη(y;0)=ϕ0m,ˆη a unit vector orthogonal to ˆη.

    (ⅱ) λm,ˆη(0)=0 and q0,m,ˆη(0)=0.

    (ⅲ) The derivative of the eigenfunction ϕm,ˆη(y;δ) at δ=0 satisfies:

    ϕm,ˆη(y;0)=iˆηαχrα(y)(ϕ0m,ˆη)r+ζm,ˆη

    where ζm,ˆηCd is a constant vector (independent of y), orthogonal to ˆη.

    (ⅳ) The derivative of the eigenfunction qm,ˆη(y;δ) at δ=0 satisfies:

    qm,ˆη(y;0)=iˆηαΠrα(y)(ϕ0m,ˆη)r.

    (v) The second derivative of the eigenvalue λm,ˆη(δ) and q0,m,ˆη(δ) at δ=0 satisfy the relation

    12λm,ˆη(0)ϕ0m,ˆη=12q0,m,ˆη(0)ˆη+M(ˆη,A)ϕ0m,ˆη (34)

    where M(ˆη,A) is the symmetric matrix whose entries are given by

    M(ˆη,A)kl=(A)klαβˆηαˆηβ.

    Remark 3. The above matrix M(ˆη,A) is precisely that which must be positive definite in the Legendre-Hadamard definition of ellipticity. A relation analogous to (34) is called "propagation relation" in [14] in the study of linearized elasticity system and it shows how the homogenized tensor A enters into the Bloch wave analysis. The above relation (34) generalizes the relation (22) in [7].

    Remark 4. In the linearized elasticity system, the propagation relation is an eigenvalue relation. Here, relation (34) can again be seen as an eigenvalue problem, posed in the (d1)-dimensional subspace orthogonal to ˆη. More precisely, 1/2λm,ˆη(0) is an eigenvalue and ϕ0m,ˆη (which is orthogonal to ˆη) is an eigenvector of the restriction of the matrix M(ˆη,A) to the subspace ˆη. In (34) 1/2q0,m,ˆη(0) is the Lagrange multiplier corresponding to the constraint that the eigenvalue problem is posed in the (d1)-dimensional subspace orthogonal to ˆη.

    Case of Symmetrized gradient : We recall the incompressible elasticity system (10) with the symmetrized gradient introduced in Section 1

    (μϵE(uϵs))+pϵs=f in Ω,uϵs=0 in Ω,uϵs=0 on Ω} (35)

    where E(v)=12(v+tv).

    We introduce Bloch waves associated to the Stokes operator defined in (35).

    Find λs=λs(η)R,ϕs=ϕs(η)H1(Td)d, ϕs0 and Πs=Πs(η)L2(Td) satisfying

    D(η)(μE(η)ϕs)+D(η)Πs=λs(η)ϕs in RdD(η)ϕs=0 in Rd(ϕs,Πs) is Y periodicY|ϕs|2 dy=1.} (36)

    As usual D(η)=y+iη is the shifted gradient operator and the shifted strain rate tensor is defined by :

    2E(η)ψ=(+iη)ψ+(+iη)tψ,(2E(η)ψ)kl=(ψkxl+iηlψk)+(ψlxk+iηkψl).

    As earlier, we modify the spectral problem (36) as follows : Find λs(δ)R,ϕs(.;δ)H1(T)d, qs(.;δ)L20(Td) and q0,s(δ)C satisfying

    D(δe)(μ(y)E(δe)ϕs(y;δ))+D(δe)qs(y;δ)+ q0,s(δ)e=λs(δ)ϕs(y;δ) in TdD(δe)ϕs(y;δ)=0 in TdeTdϕs(y;δ) dy=0,(ϕs,qs) is Y periodic,Td|ϕs(y;δ)|2 dy=1.} (37)

    As before, we can compute directional derivatives of the solution of (37) and prove a result completely analogous to Theorem 3.1. In particular, we will have the following propagation relation : For m=1,d1 and for fixed direction ˆηSd1 the second derivative of the eigenvalue λs,m,ˆη(δ) at δ=0 satisfies the relation

    12λs,m,ˆη(0)ϕ0s,m,ˆη=12q0,s,m,ˆη(0)ˆη+M(ˆη,As)ϕ0s,m,ˆη, (38)

    where M(ˆη,As) is the matrix whose entries are given by

    M(ˆη,As)jl=(As)jlαβˆηαˆηβ.

    In the scalar self-adjoint case, it is known that the homogenized matrix is equal to one-half the Hessian of the first Bloch eigenvalue at zero momentum [13]. In the general (non-symmetric) scalar case, treated in [21], it was shown that only the symmetric part of the homogenized matrix is determined by the Bloch spectrum and it is given again by the same one-half of the Hessian of the first Bloch eigenvalue (which exists by virtue of the Krein-Rutman theorem). The fact that only the symmetric part of the homogenized matrix plays a role is not a big surprise since, the homogenized tensor A being constant, the differential operator

    A=dk,l=1Akl2xkxl

    depends only on the symmetric part of A.

    In the case of systems, another phenomenon takes place. For example, the linearized elasticity system (in which there are no differential constraints) was treated in [14] where it was recognized that not only Bloch eigenvalues but also Bloch eigenfunctions at zero momentum are needed to determine the homogenized tensor. More precisely, this connection between Bloch eigenvalues and eigenfunctions, on the one hand, and the homogenized tensor, on the other hand, was expressed via a relation called propagation relation in [14] which uniquely determines the homogenized tensor.

    In the case of Stokes system, a new phenomenon arises because of the presence of a differential constraint (the incompressibility condition). Even though there is an analogue of the propagation relation (see (34) above), it does not determine uniquely the homogenized tensor. In fact the propagation relation (34) is unaltered if we add a multiple of II (where I is the d×d identity matrix) to the homogenized tensor. The homogenized Stokes operator clearly remains the same under such an addition since it corresponds to adding a gradient of the velocity divergence which vanishes because of the incompressibility constraint. The authors in [7] conjectured that the homogenized Stokes tensor is uniquely characterized by the propagation relation up to the addition of a term c(II) (where c is a constant). We prove this assertion in the case of the Stokes system (10) with a symmetrized gradient. For the other Stokes system (1), the homogenized tensor is not uniquely determined by the propagation relation (34). In this section, we investigate this non-uniqueness. Neverheless, we shall prove that for both Stokes systems the homogenized operators (9), and its equivalent for the symmetric gradient case of (10), are uniquely determined.

    Our concern now is the following question: to what extent do the Bloch spectral elements determine the homogenized tensor A via the propagation relation (34)? Since λm,ˆη(0),q0,m,ˆη(0),ϕ0m,ˆη are known from Bloch spectral data, it follows that M(ˆη,A)ϕ0m,ˆη is uniquely determined via the relation (34). But it may happen that different tensors A give rise to the same matrix M(ˆη,A). Three main results are proved in this section and they are stated in the following three propositions.

    Proposition 2. Let A and B be two fourth order tensors possessing the simple symmetry (8). They satisfy the same propagation relation (34), if and only if

    BA=c(II)+N (39)

    where I is the d×d identity matrix and N is a fourth order tensor satisfying, on top of the simple symmetry (8), the following anti-symmetry property

    Njlαβ=Njlβα=Nljαβwhenever(α,β)(j,l) and (β,α)(j,l)Niiii=0.} (40)

    Proof. Let us observe that the addition of c(II) and N, having properties (8) and (40), to A does not alter the propagation relation (34). Indeed, we have,

    M(ˆη,A+c(II)+N)jl=(A)jlαβˆηαˆηβ+cδαjδβlˆηαˆηβ+Njlαβˆηαˆηβ=M(ˆη,A)jl+cˆηjˆηl.

    Since ϕ0m,ˆη is orthogonal to ˆη, we deduce

    M(ˆη,A+c(II)+N)ϕ0m,ˆη=M(ˆη,A)ϕ0m,ˆη.

    Conversely, let us assume that there are two fourth-order tensors A and B, possessing the simple symmetry (8) and such that M(ˆη,A)ϕ0m,ˆη=M(ˆη,B)ϕ0m,ˆη, m=1,...,d1, for all ˆηSd1. We must then deduce (39). For convenience, the proof is divided into five steps.

    Step 1. We begin with showing the matrix M(ˆη,A) is symmetric. By interchanging the dummy indices α and β and using the simple symmetry (8) of the homogenized coefficients, (A)jlαβ=(A)ljβα, we get

    M(ˆη,A)jl=(A)jlαβˆηαˆηβ=(A)jlβαˆηβˆηα=(A)ljαβˆηαˆηβ=M(ˆη,A)lj (41)

    which shows the required symmetry.

    Step 2. For ˜N=BA define M(ˆη)=M(ˆη,˜N)=M(ˆη,B)M(ˆη,A). Since A and B satisfy (34), it follows that M(ˆη)ϕ0m,ˆη=0 for m=1,...,d1. Since the family ϕ0m,ˆη is a basis of the orthogonal space to ˆη, it implies that M(ˆη)=c(ˆη)ˆηˆη for some scalar c(ˆη). Since M(ˆη) depends quadratically on ˆη, it must be that c(ˆη) is independent of ˆη. Thus, for cR, we have M(ˆη)=cˆηˆη, that is, for any ˆηSd1,

    ˜Njlαβˆηαˆηβ=cˆηjˆηl1j,ld. (42)

    Step 3. Under condition (42), we verify that

    ˜Niiii=c    i. (43)
    and˜Njlik+˜Njlki =0 if (i,k)(j,l) and (k,i)(j,l). (44)

    For this purpose, let us take ˆη=ei in (42). We obtain ˜Njlii=cδijδil and so

    ˜Niiii  =  c (45)
    and  ˜Njlii=  0 if ij or il. (46)

    In particular, (43) is proved. Next, choosing ˆη=ei+ek in (42), we get

    ˜Njlii+˜Njlkk+˜Njlik+˜Njlki=c(δji+δjk)(δli+δlk). (47)

    To check (44), there are several cases to consider.

    (ⅰ) (ij and kj). In this case, (44) is a direct consequence of (46) and (47).

    (ⅱ) Similarly, for (kl and il) 44) is a direct consequence of (46) and (47).

    (ⅲ) (ij, k=j). In this case,

    ˜Njljj+˜Njlij+˜Njlji=c(δli+δlj). (48)

    Now together with il we have

    ˜Njljj+˜Njlij+˜Njlji=cδlj. (49)

    Then both j=l or jl cases lead to verify (43) and (44) respectively.

    (ⅳ) Similarly, for (kl and i=l)

    ˜Njiii+˜Njiik+˜Njiki=c(δji+δjk). (50)

    Together with kj we have

    ˜Njiii+˜Njiik+˜Njiki=cδji. (51)

    Then both i=j or ij cases lead to verify (43) and (44) respectively.

    Step 4. Now we consider the two remaining cases not covered in (44).

    (ⅰ) (i,k)=(j,l). Then from (47) we have

    ˜Nikii+˜Nikkk+˜Nikik+˜Nikki=c(1+δik)2.

    For ik it gives using (46)

    ˜Nikik+˜Nikki=c. (52)

    (ⅱ) Similarly, for (k,i)=(j,l), together with ik we have

    ˜Nkiik+˜Nkiki=c (53)

    Step 5. Let us set N=˜Nc(II). Thanks to the properties (43) and (44), we can easily check that N is an anti-symmetric tensor in the sense that it satisfies

    Njlik=Njlki=Nljik.whenever, (i,k)(j,l) and (k,i)(j,l) (54)

    From its very definition N also possesses the symmetry Njlik=Nljki. Thus N has all the properties listed in (40).

    Next we extend Proposition 2 to the Stokes system (10), featuring a symmetric gradient tensor. In this case the propagation relation (34) is replaced by (38) and the homogenized tensor is denoted by As.

    Proposition 3. The propagation relation (38) characterizes uniquely the tensor As, up to the addition of a constant multiple of II. In other words, As and Bs satisfy the same propagation relation (38) if and only if, for some cR,

    BsAs=c(II). (55)

    Proof. The proof continues from the Step 5 of the previous proof of Proposition 2. We defined N=˜Nc(II) satisfying (54) i.e.

    Njlik=Njlki=Nljik.whenever, (i,k)(j,l) and (k,i)(j,l)

    Now as ˜N=BsAs possess with the symmetry of coefficients of linear elasticity, so we have

    Njlik=Niljk=Nljki=Njkilfor all i,j,k,l. (56)

    This symmetry combined with the anti-symmetry established in the previous step implies that N=0. Note that antisymmetry property holds precisely for the interchange of those pairs of indices for which symmetry property does not hold.

    This can be seen as follows: whenever (i,k)(j,l) and (k,i)(j,l)

    Njlik=Nljik=Nijlk=Njilk=Nijkl=Nkjil=Njkil=Njlik (57)
    ThusNjlik=0. (58)

    Similarly, whenever (i,k)=(j,l) or (k,i)=(j,l) together with ik; from (52), (53) we have

    ˜Nikik+˜Nikki = c = ˜Nkiik+˜Nkiki.

    Then using (56) and (46) we clearly have

    Nikik= 0 = Nkiki. (59)

    Therefore (58), (59) imply that N=0 or, ˜N=c(II) and hence BsAs=c(II).

    Remark 5. The conclusion of the above proposition was conjectured in [7] and it is proved here to be true whenever we are working with the system (10) with symmetrized gradient. However, it is not true with the full gradient Stokes system (1) as shown by Proposition 2. However, in both of these cases the propagation relation fixes the homogenized operator(9) uniquely, as is stated in the following proposition.

    Proposition 4. If (39) is satisfied, then A and B give rise to the same homogenized operator (9).

    Proof. We have to check that A and B define the same Stokes differential operator for divergence-free vector fields. Indeed the Fourier symbol of the operator

    u=(uk)1kd  (xβ((AB)klαβukxα))1ld

    is (AB)klαβξαξβ which, by virtue of (42), is equal to cξkξl which is precisely the symbol of the operator uc(u) which vanishes on the space of divergence free functions.

    This section is devoted to a proof of Theorem 1.1, our main homogenization result stated in the first section. It is based on the tools that we have introduced so far. A similar proof is given for the linear elasticity problem in [21]. However, the presence of a pressure and a differential constraint in the Stokes system seriously complexifies the analysis and has a non-trivial effect in the homogenization process. Besides, we also bring some simplifications to the proof given in [21].

    We consider a sequence of solutions (uϵ,pϵ)(H10(Ω))d×L20(Ω) solving the Stokes system (1). It is classical to derive the following bound [8] :

    ||uϵ||(H10(Ω))d+||pϵ||L2(Ω)C||f||(L2(Ω))d, (60)

    where C is independent of ϵ. Then there exist (u,p)(H10(Ω))d×L20(Ω)  and a subsequence (uϵ, pϵ) converging weakly to (u,p) in (H10(Ω))d×L20(Ω). Our aim is to show that (u,p) satisfies the homogenized Stokes system (9). Due to the uniqueness of solutions for the system (9), it follows that the entire sequence (uϵ,pϵ) converges to (u,p) weakly in (H10(Ω))d×L20(Ω).

    There are several steps in the proof. First, we localize the Stokes system (1) by applying a cut-off function technique to the velocity u in order to get the equation (61) in the whole Rd. Next, by taking the Bloch transformation Bϵm,ˆη (1md1) of the equation (61) and passing to the limit, we arrive at the homogenized equation in the Fourier space. Finally, we take the inverse Fourier transform to go back to the physical space which gives our desired result.

    Notation. in the sequel L.H.S. stands for left hand side, and R.H.S. for right hand side.

    Step 1. Localization of the velocity u : Let vD(Ω) be arbitrary. Then vuϵ and pϵ satisfy (for l=1,,d)

    xα(μϵxα)(vuϵl)+pϵxlv=vfl+gϵl+hϵlin Rd, (61)

    where,

    gϵl=2μϵuϵlxαvxαμϵ2vxαxαuϵlandhϵl=μϵxαvxαuϵl. (62)

    Note that, gϵl and hϵl correspond to terms containing zero and first order derivatives of μϵ respectively. In the sequel, we extend uϵ and pϵ by zero outside Ω and such extensions are denoted by the same letters.

    Step 2. Limit of Bϵm,ˆη applied to the L.H.S. of (61) : We consider the following ϵ-scaled spectral problem of (25) as follows : Let ˆη=ξ|ξ|Sd1, δ=ϵ(ξˆη);

    ϕϵm,ˆη(x;δ)=ϕm,ˆη(xϵ;ϵ(ξˆη)), and λϵm,ˆη(δ)=ϵ2λm,ˆη(ϵ(ξˆη))qϵm,ˆη(x;δ)=ϵ1qm,ˆη(xϵ;ϵ(ξˆη)), and qϵ0,m,ˆη(δ)=ϵ2q0,m,ˆη(ϵ(ξˆη)).

    They satisfy the following system because of (25) :

    D(δˆη)(μϵ(x)D(δˆη)ϕϵm,ˆη(x;δ))+D(δˆη)qϵm,ˆη(x;δ)+ qϵ0,m,ˆη(δ)ˆη=λϵm,ˆη(δ)ϕϵm,ˆη(x;δ) in Rd,D(δˆη)ϕϵm,ˆη(x;δ)=0 in Rd,ˆηRdϕϵm,ˆη(x;δ) dx=0,(ϕϵm,ˆη,qϵm,ˆη) is ϵY periodic,ϵTd|ϕϵm,ˆη(x;δ)|2 dx=1.} (63)

    Let us first consider the L.H.S. of (61). For gH1(Rd)d with compact support in Ω, using the definition Bloch transformation (19) and spectral equation (63), we obtain for m=1,d1,

    Bϵm,ˆη(xα(μϵxα)g)(ξ)=eixξϕϵm,ˆη(.;δ),xα(μϵxα)g=g,xα(μϵxα)(eixξϕϵm,ˆη(.;δ))=g,λϵm,ˆη(δ)eixξϕϵm,ˆη(.;ξ)(qϵm,ˆη(.;δ)eixξ)qϵ0,m,ˆη(δ)ˆηeixξ=λϵm,ˆη(δ)Bϵm,ˆηg(ξ)g,(qϵm,ˆη(.;δ)eixξ)g,qϵ0,m,ˆη(δ)ˆηeixξ.

    In the previous equation the duality bracket is between H1comp(Rd)d and H1loc(Rd)d.

    Therefore, Bϵm,ˆη applied to the L.H.S. of (61) (1md1) is equal to

    λϵm,ˆη(δ)Bϵm,ˆη(vuϵ)(ξ)vuϵ,(qϵm,ˆη(.;δ)eixξ)vuϵ,qϵ0,m,ˆη(δ)ˆηeixξ+Bϵm,ˆη(vpϵ)(ξ). (64)

    Below, we treat each term of (64) one by one.

    1st term of (64) : By using the Taylor expansion

    λϵm,ˆη(δ)=ϵ2λm,ˆη(ϵ(ξˆη))=12λm,ˆη(0)(ξˆη)2+O(ϵ(ξˆη)3) (65)

    and then using Theorem 2.2, we get

    χϵ1Td(ξ)λϵm,ˆη(δ)Bϵm,ˆη(vuϵ)(ξ)12λm,ˆη(0)(ξˆη)2ϕ0m,ˆη^(vu)(ξ)in L2loc(Rdξ) strongly, (66)

    where we recall that ϕ0m,ˆη is a constant unit vector of Rd orthogonal to ˆη. Note that λm,ˆη(0) is linked to A via the propagation relation (34). Using this relation, the above limit can be written as

    (ξˆη)2(12q0,m,ˆη(0)ˆη+M(ˆη,A)ϕ0m,ˆη)^(vu)(ξ)=(ξˆη)212q0,m,ˆη(0)ˆηk^(vuk)+(ξˆη)2(A)klαβˆηαˆηβ(ϕ0m,ˆη)l(^vuk)(ξ). (67)

    2nd term of (64) :

    vuϵ,(qϵm,ˆηeixξ)=(vuϵ),eixξqϵm,ˆη=uϵv,eixξqϵm,ˆη (as uϵ=0). (68)

    Using the Taylor expansion of qϵm,ˆη(.;δ) :

    qϵm,ˆη(x;δ)=ϵ1qm,ˆη(xϵ;ϵ(ξˆη))=ϵ1qm,ˆη(xϵ;0)+(ξˆη)qm,ˆη(xϵ;0)+O(ϵ(ξˆη)2), (69)

    (prime denotes the derivative with respect to the second variable), with the properties that (cf. Theorem 2.1)

    qm,ˆη(xϵ;0)=0  and qm,ˆη(xϵ;0)MTd(qm,ˆη(y;0))=0  weakly in L2(Rd);  (as qm,ˆη(y;0)L20(Td)) (70)

    where, MTd(f)=Tdf(y) dy.

    Then by using uϵu strongly in L2(Ω)d from (68) we get

    vuϵ,(qϵm,ˆηeixξ)uv,eixξMTd(qm,ˆη)=0  in L2loc(Rdξ) strongly. (71)

    It is also used that, the error term O(ϵ(ξˆη)2) in the above Taylor expansion tends to 0 in the space L2loc(Rdξ;L2loc(Rd)). Thus the oscillating eigen-pressure qϵm,ˆη does not contribute to the homogenized system.

    3rd term of (64) : We use the Taylor expression of qϵ0,m,ˆη(ξ) with the property q0,m,ˆη(0)=q0,m,ˆη(0)=0 (cf. Theorem 3.1) to have

    qϵ0,m,ˆη(δ)=ϵ2q0,m,ˆη(ϵ(ξˆη))=12q0,m,ˆη(0)+O(ϵ(ξˆη)2). (72)

    So,

    vuϵ,qϵ0,m,ˆη(δ)eixξˆηvu,12q0,m,ˆη(0)(ξˆη)2ˆηeixξ  in L2loc(Rdξ) strongly.=12q0,m,ˆη(0)(ξˆη)2^(vu)ˆη. (73)

    4th term of (64) : Finally, we consider the remaining fourth term in (64), and doing integration by parts we get

    Bϵm,ˆη(vpϵ)(ξ)=vpϵ,eixξϕϵm,ˆη=pϵ,veixξϕϵm,ˆη(as (eixξϕϵm,ˆη)=0). (74)

    We use the Taylor expansion

    ϕϵm,ˆη(x;ξ)=ϕm,ˆη(xϵ;0)+ϵ(ξˆη)ϕm(xϵ;0)+O((ϵ(ξˆη))2)=ϕ0m,ˆη+ϵ(ξˆη)ϕm,ˆη(xϵ;0)+O((ϵ(ξˆη))2)ϕ0m,ˆη in L2loc(Rdξ,(L2(Ω))d) strongly. (75)

    And from (60) as ||pϵ||L2(Ω) is uniformly bounded, so up to a subsequence we have

    pϵp in L2(Ω). (76)

    Thus by passing to the limit in the R.H.S. of (74), we get

    pϵ,veixξϕϵm,ˆηp,veixξϕ0m,ˆη=p,veixξϕ0m,ˆη(as (eixξϕ0m,ˆη)=0). (77)

    Thus

    χϵ1TdBϵm,ˆη(vpϵ)(ξ)ϕ0m,ˆη^(vp)(ξ) in L2loc(Rdξ) strongly. (78)

    This property proved for H1 elements is analogous to Theorem 2.1.

    Summary so far : Combining the previous results, therefore, by taking the Bloch transformation Bϵm,ˆη of the L.H.S. of (61) (1md1) and multiplying by χϵ1Td, we see that it converges to

    (A)klαβˆηαˆηβ(ξˆη)2(ϕ0m,ˆη)l^(vuk)(ξ)+ϕ0m,ˆη^(vp)(ξ)in L2loc(Rdξ) strongly. (79)

    Step 3. Limit of Bϵm,ˆη applied to the R.H.S. of (61) : Applying Bϵm,ˆη to the R.H.S. of (61) (1md1), we obtain

    Bϵm,ˆη(vf)(ξ)+Bϵm,ˆη(gϵ)(ξ)+Bϵm,ˆη(hϵ)(ξ). (80)

    We treat below each of these terms separately. Passing to the limit in the first term is straightforward (cf. Corollary 1) and we obtain

    χϵ1Td(ξ)Bϵm,ˆη(vf)(ξ)ϕ0m,ˆη^(vf) in L2loc(Rdξ) strongly. (81)

    Limit of Bϵm,ˆη(gϵ) : We pose σϵ=μϵuϵ (σϵlα=μϵuϵlxα) which is a bounded matrix in (L2(Ω))d×d and so there exists a weakly convergent subsequence in\break (L2(Ω))d×d. Let σ be its limit as well as its extension by zero outside Ω. Then via Theorem 2.2,

    χϵ1Td(ξ)Bϵm,ˆη(σϵlαvxα)(ξ)^(σlαvxα)(ξ)(ϕ0m,ˆη)lin L2loc(Rdξ) weakly. (82)

    Due to the strong convergence of uϵ in L2(Rd)d, (cf. Corollary 1) we have

    χϵ1Td(ξ)Bϵm,ˆη(μϵΔvuϵ)(ξ)MTd(μ(y))^(Δvu)(ξ)ϕ0m,ˆηin L2loc(Rdξ) weakly. (83)

    Combining the above two convergence results and doing integration by parts, we obtain

    χϵ1Td(ξ)Bϵm,ˆη(gϵ)(ξ)2^(σv)(ξ)ϕ0m,ˆηMTd(μ(y))^(Δvu)(ξ)ϕ0m,ˆη in L2loc(Rdξ) weakly. (84)

    Limit of Bϵm,ˆη(hϵ) : We decompose it into two terms:

    Bϵm,ˆη(hϵ)=Bϵm,ˆη((μϵv)uϵ)(ξ)=(μϵv)uϵ,eixξϕ0m,ˆη(μϵv)uϵ,eixξϵ(ξˆη)ϕm(xϵ;0)+O((ϵ(ξˆη))2).

    We start with the second term. By doing integration by parts, it becomes

    (ξˆη)Rdeixξ(μϵy¯ϕm(xϵ;0)uϵ)v dx+O(ϵ(ξˆη)). (85)

    Thanks to the strong convergence of uϵ in L2(Rd)d, the above quantity converges in L2loc(Rdξ) strongly to

    (ξˆη)Rdeixξ(MTd(μ(y)y¯ϕm(y;0))u)v dx. (86)

    Next, we consider the first term of the R.H.S. of (85). After doing integration by parts, one has

    Rdeixξ[μϵΔv (uϵϕ0m,ˆη)+((μϵuϵ)ϕ0m,ˆη)viμϵ((ϕ0m,ˆηξ)uϵ)v]dx. (87)

    In a manner similar to the above arguments, the limit of (87) would be

    Rdeixξ[MTd(μ(y))Δv (uϕ0m,ˆη)+(σϕ0m,ˆη)v]dxRdeixξ[iMTd(μ(y))((ϕ0m,ˆηξ)u)v] dx. (88)

    Now combining (86) and (88) and using the fact

    ϕm(y;0)iˆηβχlβ(y)(ϕ0m,ˆη)l

    is a constant vector of Cd independent of y, which in turn implies that

    yϕm(y;0)=iˆηβyχlβ(y)(ϕ0m,ˆη)l,

    we see that χϵ1TdBϵm,ˆη(hϵ)(ξ) converges strongly in L2loc(Rdξ) to

    i(ξˆη)[MTd(μ(y)ˆηβyχlβ(y)(ϕ0m,ˆη)l)]kα^(vxαuk)(ξ)+MTd(μ(y))^(Δv uk)(ξ)(ϕ0m,ˆη)k+^(σlβvxβ)(ξ)(ϕ0m,ˆη)liMTd(μ(y))(ϕ0m,ˆη)kξα^(vxαuk)(ξ). (89)

    Step 4. Limit of Bϵm,ˆη applied to (61) : By equating the limiting identities that we have derived in the last two steps, we obtain

    (A)klαβξαξβ^(vuk)(ξ)(ϕ0m,ˆη)l+^(vpxl)(ξ)(ϕ0m,ˆη)l=^(vfl)(ξ)(ϕ0m,ˆη)l2^(σlβvxβ)(ξ)(ϕ0m,ˆη)lMTd(μ(y))^(Δv uk)(ξ)(ϕ0m,ˆη)k i[MTd(μ(y)yχlβ(y))]kα(ϕ0m,ˆη)lξβ^(vxαuk)(ξ)+MTd(μ(y))^(Δv uk)(ξ)(ϕ0m,ˆη)k +^(σlβvxβ)(ξ)(ϕ0m,ˆη)liMTd(μ(y))δαβδlk(ϕ0m,ˆη)lξβ^(vxαuk)(ξ). (90)

    The above equation has to be considered as the localized homogenized equation in the Fourier space. The conclusion of Theorem 1.1 will follow as a consequence of this equation.

    Step 5. Passage from Fourier space (ξ) to physical space (x) : We note that the L.H.S. and the R.H.S. of (90) can be written as L(ξ)ϕ0m,ˆη and R(ξ)ϕ0m,ˆη, respectively, so that we have

    [L(ξ)R(ξ)]ϕ0m,ˆη=0  for m=1,..,(d1).

    Observe that, the quantity [L(ξ)R(ξ)] is independent of m. Varying m=1,..,(d1) and using the fact ξϕ0m,ˆη,ξRd and {ϕ01,ˆηϕ0d1,ˆη} forms a basis of Rd1, we get

    [L(ξ)R(ξ)]=c(ξ)ξ for some scalar c(ξ).

    Therefore, for all test functions w(L2(Rd))d satisfying ξˆw(ξ)=0 (i.e. divw=0 in Rd) we also have

    [L(ξ)R(ξ)]ˆw(ξ)=0.

    Now by using the Plancherel's theorem, we have

    RdF1[L(ξ)R(ξ)](x)ˉw(x) dx=0,w(L2(Rd))d satisfying divw=0 (91)

    where F1 denotes the inverse Fourier transformation.

    We easily compute I(x)=F1[L(ξ)R(ξ)](x) to obtain

    Il(x)=((A)klαβ2(vuk)xβxα+vpxl)(vflσl,βvxβ(A)klαβxβ(vxαuk)) in Rd,

    which simplifies in

    Il=((A)klαβ2ukxβxα+pxlfl)v((A)klαβukxασl,β)vxβ in Rd.

    We pose

    F1l=((A)klαβ2ukxβxα+pxlfl) and F2lβ=F2βl=((A)klαβukxασl,β) (92)

    to write Il in the form

    Il=F1l v+F2lβ vxβ.

    Using (91), it follows from de Rham's theorem that I is a gradient and furthermore this is true whatever be vD(Ω). This imposes restriction on F1,F2. In fact, we show using (91) that F2lβ=qδlβ and F1=q for some scalar qL2(Ω) so that I=vq+qv=(vq).

    Step 5A. To show F2lβ=qδlβ : Let us choose v=v0einxω, where ω is a unit vector in Rd and v0D(Ω) is fixed. Next, we choose w=ψζ,ω(L2(Rd))d where for any two constant perpendicular vectors ζ and ω in Rd, ψζ,ω(L2(Rd))d solves

    divψζ,ω=0 in Rd  with ψζ,ω=ζeinxω in Ω, where ζω. (93)

    The existence of such a function ψζ,ω can be shown as follows. Let R0>0 be such that ¯ΩB(0,R0) and consider the following boundary value problem

    divψζ,ω=0 in B(0,R0)¯Ω,ψζ,ω=0 on B(0,R0),ψζ,ω=ζeinxω on Ω. (94)

    There exists a solution of (94) (see [15,Page No. 24]) since the boundary data satisfies the required compatibility condition (recall that we assume ζω=0)

    Ωζeinxων dσ=Ω(ζω)einxω dx=0.

    Then extending ψζ,ω by 0 outside B(0,R0) and by ζeinxω in Ω, clearly the extended function ψζ,ω solves (93).

    Now using these v and w in (91), we have

    ΩF1lζl v0 dx+ΩF2lβv0xβζl dx+nΩF2lβωβζl v0 dx=0

    and dividing by n and letting n in the above relation, we get

    Ω(F2 ωζ)v0 dx=0. (95)

    As v0D(Ω) is arbitrary, (95) gives F2 ωζ=0 in Ω. As F2 is symmetric, and further using that ω,ζ are arbitrary satisfying ωζ=0, we conclude F2lβ=F2βl=qδlβ for some scalar function qL2(Ω). This means that we have the relation :

    σlβ=qδlβ+(A)klαβukxα. (96)

    Step 5B. To show F1=q : We choose vD(Ω) and w=ψek,0 with ψek,0 as in (93) with ζ=ek and ω=0. Then using these v and w in (93) and using the conclusion from Step 5A, we have

    Ω(F1kv+qvxk)dx=0 for all vD(Ω),

    which implies (F1kqxk)=0 for k=1,..,d or, F1=q.

    Step 5C Using Step 5A and Step 5B in (92), and considering the relation F1F2=0 in Ω, we get the macro balance equation :

    σlβxβ+pxl=fl  in Ω,  l=1,,d. (97)

    Step 5D In this step, we prove that q=0 in Ω by using the divergence-free condition. Indeed, as uϵ=0 in Ω, we have

    σϵll=μϵuϵlxl=0 in Ω.

    Passing to the limit ϵ0, we get

    σll=0 in Ω.

    Using this relation in (96) with β=l, we get

    (A)klαlukxα+qd=0. (98)

    On the other hand, from (6) and (7) we have

    (A)klαl=Tdμ(y)(χkα+yαek):(ylel) dy=Tdμ(y)yl(χkα+yαek)l dy.

    Thus for fixed k,α=1,,d summing over l, since divχkα=0 in Y, we obtain

    (A)klαl=MTd(μ)δkα. (99)

    Using (98) and (99), as divu=0, we deduce

    q=1d(A)klαlukxα=1dMTd(μ)δkαukxα=0.

    Finally, the macro constitutive law follows as a consequence from (96) :

    σlβ=(A)klαβukxα.

    Step 5E. Since q=0, we deduce from Step 5B that F1=0 and from (96) we get the following homogenized Stokes system satisfied by u,p :

    (A)klαβ2ukxαxβ+pxl=fl in Ω  for l=1,..,d.divu=0 in Ωu=0 on Ω. (100)

    This completes the proof of Theorem 1.1.

    This work has been carried out within a project supported by Indo -French Centre for Applied Maths -UMI, IFCAM. G. A. is a member of the DEFI project at INRIA Saclay Ile-de-France. M.V. is in the current address during the preparation of the revised version.

    The authors thank the referees for their useful comments to improve the presentation of this paper.

    [1] Multiscale convergence and reiterated homogenisation. Proc. Roy. Soc. Edinburgh Sect. A (1996) 126: 297-342.
    [2] Homogenization of a spectral problem in neutronic multigroup diffusion. Comput. Methods Appl. Mech. Engrg. (2000) 187: 91-117.
    [3] Homogenization and localization for a 1-D eigenvalue problem in a periodic medium with an interface. Ann. Mat. Pura Appl. (4) (2002) 181: 247-282.
    [4] Homogenization and localization with an interface. Indiana Univ. Math. J. (2003) 52: 1413-1446.
    [5] Homogenization and two-scale convergence. SIAM J. Math. Anal. (1992) 23: 1482-1518.
    [6] G. Allaire, Shape Optimization by the Homogenization Method volume 146 of Applied Mathematical Sciences, Springer-Verlag, New York, 2002. doi: 10.1007/978-1-4684-9286-6
    [7] On Bloch waves for the Stokes equations. Discrete Contin. Dyn. Syst. Ser. B (2007) 7: 1-28 (electronic).
    [8] A. Bensoussan, J. -L. Lions and G. Papanicolaou, Asymptotic Analysis for Periodic Structures, volume 5 of Studies in Mathematics and its Applications, North-Holland Publishing Co. , Amsterdam-New York, 1978.
    [9] Optimal asymptotic estimates for the volume of internal inhomogeneities in terms of multiple boundary measurements. M2AN Math. Model. Numer. Anal. (2003) 37: 227-240.
    [10] Homogenization of the non-stationary {S}tokes equations with periodic viscosity. J. Korean Math. Soc. (2009) 46: 1041-1069.
    [11] C. Conca, J. Planchard and M. Vanninathan, Fluids and Periodic Structures, volume 38 of RAM: Research in Applied Mathematics, John Wiley & Sons, Ltd. , Chichester; Masson, Paris, 1995.
    [12] Bloch approximation in homogenization and applications. SIAM J. Math. Anal. (2002) 33: 1166-1198 (electronic).
    [13] Homogenization of periodic structures via {B}loch decomposition. SIAM J. Appl. Math. (1997) 57: 1639-1659.
    [14] Bloch wave homogenization of scalar elliptic operators. Asymptot. Anal. (2004) 39: 15-44.
    [15] V. Girault and P. -A. Raviart, Finite Element Methods for Navier-Stokes Equations, volume 5 of Springer Series in Computational Mathematics, Springer-Verlag, Berlin, 1986. doi: 10.1007/978-3-642-61623-5
    [16] U. Hornung, Homogenization and Porous Media, Springer, New York, 1997. doi: 10.1007/978-1-4612-1920-0
    [17] An approach for constructing families of homogenized equations for periodic media. SIAM J. Math. Anal. (1991) 22: 16-33.
    [18] Generic simplicity of the eigenvalues of the Stokes system in two space dimensions. Adv. Differential Equations (2001) 6: 987-1023.
    [19] M. Reed and B. Simon, Methods of Modern Mathematical Physics, Academic Press, New York, 1978.
    [20] E. Sanchez-Palencia, Non-Homogeneous Media and Vibration Theory, volume 129 of Springer Lecture Notes in Physics, Springer-Verlag, Berlin, 1980.
    [21] Bloch wave homogenization of linear elasticity system. ESAIM Control Optim. Calc. Var. (2005) 11: 542-573 (electronic).
    [22] Theory of bloch waves. J. Anal. Math. (1978) 33: 146-167.
  • This article has been cited by:

    1. Shu Gu, Jinping Zhuge, Periodic homogenization of Green’s functions for Stokes systems, 2019, 58, 0944-2669, 10.1007/s00526-019-1553-9
    2. Vivek Tewary, Combined effects of homogenization and singular perturbations: A bloch wave approach, 2021, 16, 1556-181X, 427, 10.3934/nhm.2021012
    3. Kirill Cherednichenko, Serena D’Onofrio, Operator-norm homogenisation estimates for the system of Maxwell equations on periodic singular structures, 2022, 61, 0944-2669, 10.1007/s00526-021-02139-7
    4. T. Muthukumar, K. Sankar, Homogenization of the Stokes System in a Domain with an Oscillating Boundary, 2022, 20, 1540-3459, 1361, 10.1137/22M1474345
  • Reader Comments
  • © 2017 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(5203) PDF downloads(132) Cited by(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog