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

Corrector results for a class of elliptic problems with nonlinear Robin conditions and L1 data

  • Received: 05 December 2022 Revised: 22 March 2023 Accepted: 03 April 2023 Published: 25 April 2023
  • In this paper, we consider a class of elliptic problems in a periodically perforated domain with L1 data and nonlinear Robin conditions on the boundary of the holes. Using the framework of renormalized solutions, which is well adapted to this situation, we show a convergence result for the truncated energy in the quasilinear case. When the operator is linear, we also prove a corrector result. Since we cannot expect to have solutions belonging to H1, the main difficulty is to express the corrector result through the truncations of the solutions, together with the fact that the definition of a renormalized solution contains test functions which are nonlinear functions of the solution itself.

    Citation: Patrizia Donato, Olivier Guibé, Alip Oropeza. Corrector results for a class of elliptic problems with nonlinear Robin conditions and L1 data[J]. Networks and Heterogeneous Media, 2023, 18(3): 1236-1259. doi: 10.3934/nhm.2023054

    Related Papers:

    [1] 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
    [2] 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
    [3] Mogtaba Mohammed, Mamadou Sango . Homogenization of nonlinear hyperbolic stochastic partial differential equations with nonlinear damping and forcing. Networks and Heterogeneous Media, 2019, 14(2): 341-369. doi: 10.3934/nhm.2019014
    [4] Junlong Chen, Yanbin Tang . Homogenization of nonlinear nonlocal diffusion equation with periodic and stationary structure. Networks and Heterogeneous Media, 2023, 18(3): 1118-1177. doi: 10.3934/nhm.2023049
    [5] Sara Monsurrò, Carmen Perugia . Homogenization and exact controllability for problems with imperfect interface. Networks and Heterogeneous Media, 2019, 14(2): 411-444. doi: 10.3934/nhm.2019017
    [6] Shin-Ichiro Ei, Toshio Ishimoto . Effect of boundary conditions on the dynamics of a pulse solution for reaction-diffusion systems. Networks and Heterogeneous Media, 2013, 8(1): 191-209. doi: 10.3934/nhm.2013.8.191
    [7] Renata Bunoiu, Claudia Timofte . Homogenization of a thermal problem with flux jump. Networks and Heterogeneous Media, 2016, 11(4): 545-562. doi: 10.3934/nhm.2016009
    [8] Martin Heida, Benedikt Jahnel, Anh Duc Vu . Regularized homogenization on irregularly perforated domains. Networks and Heterogeneous Media, 2025, 20(1): 165-212. doi: 10.3934/nhm.2025010
    [9] Filipa Caetano, Martin J. Gander, Laurence Halpern, Jérémie Szeftel . Schwarz waveform relaxation algorithms for semilinear reaction-diffusion equations. Networks and Heterogeneous Media, 2010, 5(3): 487-505. doi: 10.3934/nhm.2010.5.487
    [10] Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo . On boundary optimal control problem for an arterial system: First-order optimality conditions. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027
  • In this paper, we consider a class of elliptic problems in a periodically perforated domain with L1 data and nonlinear Robin conditions on the boundary of the holes. Using the framework of renormalized solutions, which is well adapted to this situation, we show a convergence result for the truncated energy in the quasilinear case. When the operator is linear, we also prove a corrector result. Since we cannot expect to have solutions belonging to H1, the main difficulty is to express the corrector result through the truncations of the solutions, together with the fact that the definition of a renormalized solution contains test functions which are nonlinear functions of the solution itself.



    The aim of this work is to prove convergence of energies and corrector results for an elliptic problem with nonlinear Robin conditions and L1 data in a periodically perforated domain. This study completes the homogenization results given by the authors in [11], where the convergence of the solutions to a limit problem explicitly described, including its unfolded version, is proved.

    More precisely, we consider the following elliptic problem in a periodically perforated domain Ωε:

    {div(A(xε,uε)uε)=f in Ωε,uε=0 on Γε0,A(xε,uε)uεn+εγτ(xε)h(uε)=εg(xε) on Γε1, (1.1)

    where fL1(Ω), gL1(T), A(,t) is a coercive Y-periodic matrix which is bounded where t is bounded, h is a monotone continuous function verifying a sign condition and τ is a positive function in L(T).

    Here, as in [11] (see also [4] and the references therein), the perforated domain is obtained by removing from a fixed domain Ω a set of ε-periodic holes of size ε. Its boundary consists of two parts on which we prescribe two different boundary conditions. Roughly speaking, on the boundary of the holes which are completely contained in Ω, we prescribe the nonlinear Robin condition, and on the remaining part of the boundary is a homogeneous Dirichlet condition. We refer the reader to Section 2 for a rigorous definition of the domain.

    Heterogeneous media are widely studied since they have many interesting applications in sciences, in industry and, more recently, even in biology and environmental sciences. Let us recall that the mathematical homogenization theory (see, for instance, [3,8]) allows to describe the microscopical behavior of a problem with periodic oscillations in the coefficients and/or in the domain. Theory provides a limit homogenized problem described through a problem posed in the periodicity reference cell. It represents a good approximation of the initial problem, and it is easier to compute since it does not present oscillations anymore.

    It is well known already, in the classical case where fL2, that the gradient of the solution uε converges weakly (never strongly) in L2 to that of the solution u0 of the homogenized problem. This is the reason why one looks for corrector results improving the weak convergence. To do that, one replaces u0 by Cεu0, where Cε is the corrector matrix field, described via the cell problem. Hence, one proves that uεCεu0 strongly converges to zero in L1. As a first step of the proof, one has to prove the convergence of the energy of the problem to the one of the homogenized problem.

    In this paper, we prove similar results in the more delicate case where f is only in L1. For physical motivations and references of related works, we refer the reader to [11].

    As usual, the presence of the L1 data requires a specific framework: we use here that of renormalized solutions (see [9] and the references therein and Definition 2.1 below). Since, in this case, the solution does not belong to H1, the notion of a renormalized solution consists first of imposing the regularity of the truncations of the solution, and second, of making use of test functions of the type S(u)φ where SC1(R) has a compact support and φLH1. The test functions depend on the solution and vanish for large values of the solution. To counterbalance the lack of information where |u| is large, a decay of the truncated energy is imposed.

    The existence and the uniqueness of a renormalized solution of this problem have been proved in [12]. Successively, in [11], the authors, using the periodic unfolding method introduced in [5] (see for a complete presentation the book [7]), studied the homogenization of problem (1.1), proving that the renormalized solution converges to the renormalized solution of a homogenized problem posed in the whole domain.

    We prove first the convergence of the energies for problem (1.1) (see Theorem 3.1 and Theorem 3.3), and then a corrector result (Theorem 4.1) for the corresponding linear equation, where the matrix field does not depend on the solution that is A(y,t)=A(y) (see Remark 4.2). As far as we know, the results presented here are new, even in the case of a fixed domain (where there are no holes, so that Ωε=Ω). With respect to the classical situation with L2 data, since the solutions are not in H1, we cannot expect to have for the renormalized solutions a convergence result for uεCεu0, and we can only describe the convergences in terms of the truncated solutions and of the truncated limit function (at a fixed level). Our corrector result states the following convergence:

    limε0Tk(uε)CεTk(u0)L1(Ωε)=0, (1.2)

    where u0 is the solution of the homogenized problem, Tk is the truncation at level k and Cε is the corrector matrix of the classical linear case in perforated domains (see [10]). This is not surprising, since, in the homogenization results proved in [11] for the case fL1, all of the convergences concern the truncations Tk(uε). In fact, the use of the truncation is standard in the literature of renormalized solutions.

    The proofs are quite technical, since one cannot merely replace the solutions by their truncations and follow the usual arguments because the definition of a renormalized solution (see (2.21) in Definition 2.1) contains test functions which are nonlinear functions of the solution itself. This is the main difficulty all along the proofs. In addition, since the truncation function is not differentiable, we need to approach it by using suitable and more regular functions.

    In Section 2, we introduce the problem and we recall some results on the periodic unfolding method, as well as the homogenization results from [11]. In Section 3, we prove the convergence of both (unfolded and not) types of truncated energy to those of the homogenized problem. Section 4 and 5 are devoted to the statement of the corrector result, and to the related proofs.

    In this paper, we study some corrector results for an elliptic problem with nonlinear Robin conditions and L1 data, in a periodically perforated domain Ωε.

    In Subsection 2.1, we define the perforated domain and set the problem, together with its variational formulation. In Subsection 2.2, we recall the definition of the periodic unfolding operator and the homogenization results obtained in [11].

    Let us introduce the geometrical framework used in [11] (see also [4]). In what follows, Ω is a connected open bounded subset of RN (N2) with a Lipschitz-continuous boundary and b=(b1,,bN) as a given basis of RN.

    We define the reference periodicity cell Y by

    Y={RN:=Ni=1libi , (l1,,lN)(0,1)N},

    and denote by {ε}ε>0 a positive sequence converging to zero. We set

    G={ξRN:ξ=Ni=1kibi , (k1,,kN)ZN}.

    As is usual in the periodic unfolding method, (see for instance [6], and the exhaustive book [7]), we construct the interior of the largest union of cells ε(ξ+¯Y) contained in Ω, as well as its complement, that is,

    ˆΩε=interior{ξΞεε(ξ+¯Y)Ω},Λε=ΩˆΩε,where Ξεε={ξG:ε(ξ+Y)Ω}. (2.1)

    We now denote by T the reference hole, which is a compact subset of Y, and by Y=YT the perforated reference cell. We suppose that the boundary T is Lipschitz-continuous with a finite number of connected components.

    Then, the holes and the perforated domain Ωε (see Figure 1) are defined by

    Tε=ξGε(ξ+T),Ωε=ΩTε, (2.2)
    Figure 1.  The reference cell Y and the perforated domain Ωε.

    respectively, while the perforated sets corresponding to (2.1) are

    ˆΩε=ˆΩεTεandΛε=ΩεˆΩε. (2.3)

    Finally, we decompose the boundary of the perforated domain Ωε as

    Ωε=Γε0Γε1,whereΓε1=ˆΩεTε and Γε0=ΩεΓε1. (2.4)

    In the sequel, we denote by

    ˜v, the extension by zero outside B of a function v defined on any set B,

    θ=|Y||Y|, the proportion of the material,

    χA, the characteristic function of a measurable set A,

    MT(v)=1|T|Tv(y)dσy, the mean value over T of a function vL1(T).

    Let us recall that, as ε0,

    χΩεθweakly in L(Ω). (2.5)

    We are concerned with the following problem:

    {div(Aε(x,uε)uε)=f in Ωε,uε=0 on Γε0,Aε(x,uε)uεn+εγτε(x)h(uε)=gε on Γε1, (2.6)

    where γ1 and n is the unit exterior normal to Ωε.

    We suppose that the following assumptions hold true:

    ● The functions f, g, h and τ are such that

     1. fL1(Ω) (2.7)
     2. h:RR is an increasing continuous function, with h(0)=0 (2.8)
     3. τ is a positive Y-periodic function in L(T) with  (2.9)
    τε(x)=τ(xε).
    4. Either (2.10)

    (i) gε(x)=εg(xε), with gL1(T) Y-periodic with MT(g)0

    or

    (ii) gε0.

    ● Let A:(y,t)Y×RA(y,t)RN2 be a real matrix field such that the matrix field A(,t)={aij(,t)}i,j=1...N is Y-periodic for every t.

    We suppose that A is a Carathéodory function, i.e., for almost every yY, the map tA(y,t) is continuous, and for every tR, the map yA(y,t) is measurable.

    For some constant α>0, we suppose further that the matrix A satisfies the following:

     1. A(y,t)ξξα|ξ|2, for a.e. yY,tR,ξRN (2.11)
     2. k>0,A(y,t)L(Y×(k,k))N×N (2.12)

    3. The matrix field A(y,t) is locally Lipschitz-continuous with respect to the second variable, that is, for every r>0, there exists a positive constant Mr such that

    |A(y,s)A(y,t)|<Mr|st|,s,t[r,r],yY, (2.13)

    and we set

    Aε(x,t)=A(xε,t) for every (x,t)Ω×R. (2.14)

    In order to define a renormalized solution of problem (2.6), let us introduce the space

    Vε={vH1(Ωε):v=0onΓε0}, (2.15)

    equipped with the norm

    vVε=vL2(Ωε)for all vVε. (2.16)

    Observe that (2.16) defines a norm since a Poincaré inequality holds in Vε, namely,

    uL2(Ωε)CuL2(Ωε)uVε, (2.17)

    where the constant C is independent of ε. Also, the Sobolev continuous and compact embedding theorems on Vε hold with constants independent of ε.

    We recall now the definition of the truncation, which plays a crucial role in our work. For any k>0, the truncation function Tk:RR at height ±k is given by

    Tk(t)=min(k,max(t,k)) (2.18)

    for all tR (see Figure 2).

    Figure 2.  The function Tk.

    Let us now present the definition of a renormalized solution to our problem, introduced in [12].

    Definition 2.1. We say that uε is a renormalized solution of (2.6) if

    Tk(uε)Vε for any k>0, (2.19)
    limn+1n{xΩε:|uε|<n}Aε(x,uε)uεuεdx=0, (2.20)

    and for any ψC1(R) (or equivalently for any ψW1,(R)) with compact support, uε satisfies

    Ωεψ(uε)Aε(x,uε)uεvdx+Ωεψ(uε)Aε(x,uε)uεuεvdx+Γε1εγτε(x)ψ(uε)h(uε)vdσx=Ωεfψ(uε)vdx+Γε1gεψ(uε)vdσx (2.21)

    for all vVεL(Ωε).

    Remark 2.2.

    1. Proposition 2.3 in [12] (see also [2]) guarantees that the gradient and the trace along the boundaries of any function verifying (2.19) and (2.20) are well defined almost everywhere in Ωε and Γε1, respectively. This shows that every term in (2.21) is well defined.

    2. Observe that, for every k>0, we have

    vTk(v)=Tk(v)Tk(v) (2.22)

    for any function v such that Tk(v)Vε for all k>0.

    3. It has been proved in [12] that, under assumptions (2.7)–(2.12), there exists a renormalized solution to (2.6) in the sense of Definition 2.1. Moreover, assumption (2.13) provides the uniqueness of a solution.

    In this subsection, we recall the homogenization results proved in [11] by using the periodic unfolding method, and we state them in the particular case where assumption (2.13) holds. This condition is needed in the following sections, since it provides the uniqueness of the solution to the problem we consider here.

    Let us start by recalling the definitions of the unfolding operator and the boundary unfolding operator. For a detailed and extensive presentation of the method, see [6,4,7]. For the properties used in this paper, we refer the reader to [11,Section 3].

    For a.e. zRN, we denote by [z]Y=Ni=1libi, liZ for i=1,,n, the unique integer combination such that z[z]YY and set {z}Y=z[z]YY.

    Thus, for a positive ε, we can write

    x=ε({xε}Y+[xε]Y) for a.e. xRN.

    Definition 2.3. Suppose φ is a Lebesgue-measurable function. The unfolding operator Tε is defined as

    Tε(φ)(x,y)={φ(ε[xε]Y+εy) for a.e. (x,y)ˆΩε×Y,0 for a.e. (x,y)Λε×Y. (2.23)

    Definition 2.4. Suppose that φ is a Lebesgue-measurable function on ˆΩεTε. The boundary unfolding operator Tbε is defined as

    Tbε(φ)(x,y)={φ(ε[xε]Y+εy) for a.e. (x,y)ˆΩεT,0 for a.e. (x,y)Λε×T. (2.24)

    Remark 2.5. For a given a continuous function r(x), with r(0)=0, one has

    Tε(r(uε))=r(Tε(uε)) (2.25)

    in Ω×Y.

    Nevertheless, for any Lebesgue measurable function φ, we can write

    Tε(r(uε))Tε(φ)=r(Tε(uε))Tε(φ) (2.26)

    even if r(0)0. This is due to the fact that, if (x,y)Λε×Y, equality is still obtained since Tε(φ)(x,y)=0. Further, this implies that (2.26) holds for all (x,y)Ω×Y.

    Similar properties hold for Tbε.

    Let us state now the homogenization results proved in [11].

    Theorem 2.6 ([11]). Let uε be the renormalized solution of (2.6) under assumptions (2.7)–(2.14), with γ1. Set J(γ)

    J(γ)={|T| if γ=1,0 if γ>1. (2.27)

    Then, as ε tends to zero, there exists u0:ΩR, measurable and finite almost everywhere, and for every kN, ^ukL2(Ω,H1 per (Y)) with MY(^uk)=0 satisfying

    {(i).Tε(uε)u0 a.e. in Ω×Y,(ii).Tbε(uε)u0 a.e. in Ω×T, (2.28)

    and

    {(i).Tε(Tk(uε))Tk(u0) strongly in L2(Ω,H1(Y)),(ii).Tε(Tk(uε))Tk(u0)+y^uk weakly in L2(Ω×Y),(iii).Tk(~uε)=~Tk(uε)θTk(u0) weakly in L2(Ω),(iv).Tk(uε)Tk(u0)L2(Ωε)0. (2.29)

    Further, there exists a unique measurable function ˆu:Ω×YR such that, for every function RW1,(R) with compact support such that suppR[n,n] for some nN, we have

    R(u0)^uk=R(u0)ˆu a.e. in Ω×Y,kn. (2.30)

    Moreover, if S,S1 are functions in C1(R) with compact supports, then the pair (u0,ˆu) is the unique solution of the limit problem

    {Ω×YA(y,u0)(u0+yˆu)((S(u0)η0)+S1(u0)yΨ(x,y))dxdy+ J(γ)MT(τ)Ωh(u0)S(u0)η0dx= |Y|ΩfS(u0)η0dx+|T|MT(g)ΩS(u0)η0dx, for every  η0H10(Ω)L(Ω) and for every  ΨL2(Ω,H1 per (Y)). (2.31)

    We also have the convergence

    limk1kΩ×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy=0. (2.32)

    As a consequence, we prove the following result, used in the sequel:

    Corollary 2.7. Under the assumptions of Theorem 2.6, for any bounded continuous function H:RR such that H(0)=0, we have

    H(uε)H(u0)L2(Ωε)0. (2.33)

    Consequently, using (2.5),

    ~H(uε)θH(u0) in L(Ω) -weak* . (2.34)

    Proof. Using the properties of H, from Remark 2.5, convergence (2.28)(i) and the dominated convergence Lebesgue theorem, we get

    Tε(H(uε))=H(Tε(uε))wH(u0)strongly in L2(Ω×Y).

    Applying Corollary 1.19 of [4], from the boundedness of H we derive convergence (2.33). Convergence (2.34) is then straightforward.

    We also recall the next theorem, which identifies ˆu in terms of the limit function u0.

    Theorem 2.8 ([11]). Under the same assumptions and notations of Theorem 2.6, the function ˆu can be expressed as

    ˆu(y,x)=Nj=1^χej(y,u0(x))u0xj(x), (2.35)

    where (ej)Nj=1 is the canonical basis of RN and ^χej(,t) is the solution of

    {div(A(,t)y^χλ(,t))=div(A(,t)λ) in Y,A(,t)(λy^χλ(,t))n=0 on T,^χλ(,t)Y -periodic ,MY(^χλ(,t))=0 (2.36)

    for every tR and λRN.

    The next result shows that u0 is a renormalized solution to a homogenized elliptic problem corresponding to the homogenized matrix A0.

    Theorem 2.9 ([11]). Let u0 be the function given in Theorem 2.6. Then, u0 is the renormalized solution of the problem

    {div(A0(u0)u0)+J(γ)|Y|MT(τ)h(u0)=θf+|T||Y|MT(g) in Ω,u0=0 on Ω, (2.37)

    that is, u0 satisfies

    Tk(u0)H10(Ω) for any k>0, (2.38)
    limn+1n{xΩ:|u0|<n}A0(u0)u0u0dx=0, (2.39)

    and for every SC1(R) with compact support, u0 satisfies

    ΩS(u0)A0(u0)u0η0dx+ΩS0(u0)A0(u0)u0u0η0dx+J(γ)|Y|MT(τ)ΩS(u0)h(u0)η0dx=θΩfS(u0)η0dx+|T||Y|MT(g)ΩS(u0)η0dx (2.40)

    for all η0H10(Ω)L(Ω).

    The homogenized matrix A0(t) is defined, for every fixed tR, as

    A0(t)λ=1|Y|YA(y,t)y^wλ(y,t)dyλRN, (2.41)

    in which

    ^wλ(y,t)=λy^χλ(y,t), (2.42)

    and where the function ^χλ(,t) is the solution of the problem (2.36).

    Consequently, in view of Theorem 2.8,

    A0(u0)u0=1|Y|YA(y,u0)(u0+yˆu)dy,a.e. in Ω. (2.43)

    The result below, proved in [11] (Proposition 6.1), plays an important role in the proof of the corrector results to our problem.

    Proposition 2.10. Under the assumptions of Theorem 2.6, for every kN and ε>0, we have

    limk+lim supε01k{|uε|<k}Aε(x,uε)uεuεdx=0. (2.44)

    The first result of this section states the convergence of the truncated energies associated with our problem. This convergence is important in itself, and it is essential in the proof of our corrector results.

    To this aim, for nN, we define the function ψn (see Figure 3) by

    ψn(x)={xn+2,2nxn1,nxnxn+2,nx2n0,|x|2n, (3.1)
    Figure 3.  The function ψn.

    which is Lipschitz-continuous and has a compact support given by suppψn=[2n,2n].

    Further, ψn satisfies

    0ψn1,|ψn(s)|1n for |s|2n, a.e. in R. (3.2)

    Theorem 3.1. Under assumptions (2.7)–(2.14), let uε be the renormalized solution to (2.6). Let also GW1,(R) be a nondecreasing function such that G has a compact support and G(0)=0.

    Then,

    limε0ΩεAε(x,uε)uεG(uε)dxΩA0(u0)u0G(u0)dx (3.3)

    as ε tends to zero, where u0 and A0 are given by Theorem 2.6.

    In particular, for every fixed kN,

    ΩεAε(x,uε)Tk(uε)Tk(uε)dx=ΩεAε(x,uε)uεTk(uε)dxΩA0(u0)Tk(u0)u0dx=ΩA0(u0)Tk(u0)Tk(u0) (3.4)

    as ε tends to zero.

    Proof. Let GW1,(R) be a nondecreasing function such that, for some kN, supp G[k,k]. Since, for nk,

    ΩεAε(x,uε)uεG(uε)dx=Ωεψn(uε)Aε(x,uε)uεG(uε)dx; (3.5)

    it suffices to prove that

    limn+limε0Ωεψn(uε)Aε(x,uε)uεG(uε)dx=ΩA0(u0)u0G(u0)dx, (3.6)

    where ψn is defined by (3.1). Using ψ=ψn and v=G(uε) in (2.21), we have

    Ωεψn(uε)Aε(x,uε)uεG(uε)dx=Ωεfψn(uε)G(uε)dx+Γε1gεψn(uε)G(uε)dσxΓε1εγτε(x)ψn(uε)h(uε)G(uε)dσxΩεψn(uε)Aε(x,uε)uεuεG(uε)dx. (3.7)

    Let us first prove that, for any nN,

    limε0(Ωεfψn(uε)G(uε)dx+Γε1gεψn(uε)G(uε)dσxΓε1εγτε(x)ψn(uε)h(uε)G(uε)dσx)=θΩfψn(u0)G(u0)dx+|T||Y|MT(g)Ωψn(u0)G(u0)dxJ(γ)|Y|MT(τ)Ωψn(u0)h(u0)G(u0)dx, (3.8)

    where J(γ) is given by (2.27).

    Corollary 2.7 applied to ψnG, and the first convergence in (2.28), give

    limε0Ωεfψn(uε)G(uε)dx=θΩfψn(u0)G(u0)dx. (3.9)

    From the properties of the boundary unfolding operator (see [4]) and Remark 2.5, we have

    εΓε1τε(x)ψn(uε)h(uε)G(uε)dσx=1|Y|Ω×Tτ(y)Tbε(ψn(uε))Tbε(h(uε))Tbε(G(uε))dxdσy=1|Y|Ω×Tτ(y)ψn(Tbε(uε))h(Tbε(uε))G(Tbε(uε))dxdσy. (3.10)

    Using convergence (ii) of (2.28), we obtain

    ψn(Tbε(uε))ψn(u0)a.e. in Ω×T, (3.11)
    G(Tbε(uε))G(u0)a.e. in Ω×T. (3.12)

    Also, from the assumptions on h and Remark 2.5, we have

    Tbε(h(uε))=h(Tbε(uε))h(u0)a.e. in Ω×T. (3.13)

    Thus, combining the convergences above, equality (3.10) gives

    limε0εΓε1τε(x)ψn(uε)h(uε)G(uε)dσx=1|Y|Ω×Tτ(y)ψn(u0)h(u0)G(u0)dxdσy=1|Y|(Tτ(y)dσy)(Ωψn(u0)h(u0)G(u0)dx)=|T||Y|MT(τ)Ωψn(u0)h(u0)φdx, (3.14)

    since u0 is independent of y. When γ>1, we deduce from (3.14) that

    limε0 εγΓε1τε(x)ψn(uε)h(uε)G(uε)dσx=0. (3.15)

    Concerning the second integral on the left-hand side of (3.8), for the case MT(g)0, we again use (2.10), Remark 2.5 and properties of the boundary unfolding operator to write

    εΓε1g(xε)ψn(uε)G(uε)dσx=1|Y|Ω×Tg(y)Tbε(ψn(uε))Tbε(G(uε))dxdσy=1|Y|Ω×Tg(y)ψn(Tbε(uε))G(uε)(Tbε)dxdσy.

    Arguing as above, we obtain

    limε0εΓε1g(xε)ψn(uε)G(uε)dσx=|T||Y|MT(g)Ωψn(u0)φ(x)dx, (3.16)

    which completes the proof of (3.8).

    Now, using the properties of ψn and setting mG=maxR|G|, we obtain

    |Ωεψn(uε)Aε(x,uε)uεuεG(uε)dx|=|{|uε|<2n}ψn(uε)Aε(x,uε)uεuεG(uε)dx|mGn{|uε|<2n}Aε(x,uε)uεuεdx.

    Then, from Proposition 2.10, we deduce that

    lim supε0|Ωεψn(uε)Aε(x,uε)(uε)(uε)G(uε)dx|=ω1(n), (3.17)

    where ω1(n) goes to zero as n. On the other hand, taking η0=G(u0) and S=ψn as test functions in (2.40) gives

    Ωψn(u0)A0(u0)u0G(u0)dx=θΩfψn(u0)G(u0)dx+|T||Y|MT(g)Ωψn(u0)G(u0)dxJ(γ)|Y|MT(τ)Ωψn(u0)h(u0)G(u0)dxΩψn(u0)A0(u0)u0u0G(u0)dx. (3.18)

    This, combined with (3.7), and using (3.8) and (3.17), yields

    limε0Ωεψn(uε)Aε(x,uε)uεG(uε)dx=Ωψn(u0)A0(u0)u0G(u0)dx+ω1(n).

    Now, since ψn1 as n, by the Lebesgue dominated convergence theorem,

    limnΩψn(u0)A0(u0)u0G(u0)dx=ΩA0(u0)u0G(u0)dx. (3.19)

    Therefore, passing to the limit as n+ in (3.19), we get (3.6), which, in view of (3.5), proves (3.3).

    Proposition 3.2. Under the assumptions of Theorem 3.1, for every kN, we have

    ΩA0(u0)Tk(u0)Tk(u0)dx=1|Y|Ω×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy (3.20)

    and

    Ω×YA(y,uε)Tε(Tk(uε))Tε(Tk(uε))dxdy1|Y|Ω×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy (3.21)

    as ε tends to zero.

    Proof. We prove first the following inequality for the two energies:

    ΩA0(u0)Tk(u0)Tk(u0)dx=1|Y|Ω×Yχ{|u0|<k}A(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy1|Y|Ω×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy. (3.22)

    Observe that

    Tk(u0)=u0χ{|u0|<k}a.e in Ω. (3.23)

    Then, using (2.30) and equality (2.43) from Theorem 2.8, we can write

    1|Y|Ω×Yχ{|u0|<k}A(y,u0)(Tk(u0)+y^uk)Tk(u0)dxdy=1|Y|Ω×YA(y,u0)(u0+yˆu)Tk(u0)dxdy=ΩA0(u0)u0Tk(u0)dx=ΩA0(u0)Tk(u0)Tk(u0)dx.

    Hence, to prove the equality in (3.22), it suffices to show that

    1|Y|Ω×Yχ{|u0|<k}A(y,u0)(Tk(u0)+y^uk)y^ukdxdy=1|Y|Ω×Yχ{|u0|<k}A(y,u0)(Tk(u0)+yˆu)y^ukdxdy=0, (3.24)

    where, again, we used (2.30) in the first equality.

    To do that, for any δ>0, let S1δC1(R) be a bounded sequence function with compact support contained in [k,k], and such that

    0S1δ(r)1,andlimδ0S1δ(r)χ{|r|<k}for every rR. (3.25)

    Then, choosing in (2.31) η0=0, Ψ=^uk and S1=S1δ, we obtain

    Ω×YA(y,u0)(Tk(u0)+yˆu)S1δ(u0)y^ukdxdy=Ω×YA(y,u0)(u0+yˆu)S1δ(u0)y^ukdxdy=0.

    Passing to the limit as δ0, from (3.25), we deduce (3.24), which concludes the proof of (3.22).

    Let us prove now convergence (3.21). From (2.11)–(2.12), by the lower semi-continuity of the limit and using convergence (2.29)(ii), the properties of the unfolding operator and convergence (3.4) from Theorem 3.1, we have

    Ω×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdylim infε0Ω×YA(y,uε)Tε(Tk(uε))Tε(Tk(uε))dxdylim supε0Ω×YA(y,u0)Tε(Tk(uε))Tε(Tk(uε))dxdy=lim supε0|Y|ˆΩεAε(x,uε)Tk(uε)Tk(uε)dx=|Y|ΩA0(u0)Tk(u0)Tk(u0)dxΩ×YA(y,u0)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy,

    where we also used (3.22). This implies the equality of all terms above and proves both equality (3.20) and convergence (3.21).

    The following result shows that convergence (ii) of (2.29) is actually strong.

    Theorem 3.3. Under the assumptions of Theorem 3.1, for all kN,

    Tε(Tk(uε))Tk(u0)+y^uk strongly in L2(Ω×Y). (3.26)

    Proof. By the ellipticity of A, we have

    αTε(Tk(uε))(Tk(u0)+y^uk)2L2(Ω×Y)Jε, (3.27)

    where

    Jε=Ω×YA(y,uε)(Tε(Tk(uε))(Tk(u0)+y^uk))(Tε(Tk(uε))(Tk(u0)+y^uk))dxdy. (3.28)

    It suffices to show that

    Jε0 as ε0. (3.29)

    Now, from (3.28), we can write Jε as

    Jε=Jε,1Jε,2Jε,3+Jε,4, (3.30)

    where

    Jε,1=Ω×YA(y,uε)Tε(Tk(uε))Tε(Tk(uε))dxdy,Jε,2=Ω×YA(y,uε)Tε(Tk(uε))(Tk(u0)+y^uk)dxdy,Jε,3=Ω×YA(y,uε)(Tk(u0)+y^uk)Tε(Tk(uε))dxdy,Jε,4=Ω×YA(y,uε)(Tk(u0)+y^uk)(Tk(u0)+y^uk)dxdy.

    By Proposition 3.2, we get

    Jε,1Jε,4. (3.31)

    From the convergences of (2.29) in Theorem 2.6, we also have

    Jε,2Jε,4, (3.32)
    Jε,3Jε,4. (3.33)

    Hence, using (3.31)–(3.33), from (3.30), we get (3.29).

    In this section, we suppose that the equation is linear, that is, A(y,t)=A(y). Then, (2.12) reads as A(L(Y))N×N, and we set

    β=AL(Y). (4.1)

    Moreover, the homogenized matrix A0 is constant and is given by

    A0λ=1|Y|YA(y)y^wλ(y)dyλRN. (4.2)

    We recall the well-known inequality (see, for instance, [8,Prop. 8.3])

    A0L(Y)β2α. (4.3)

    In this case, the functions ^χλ and ^wλ defined in (2.36) and (2.42), respectively, are the classical functions used in the linear homogenization. That is, for any λRN, the function ^χλ is the unique solution of

    {div(A^χλ)=div(Aλ)in Y,A(λ^χλ)n=0on T,^χλY-periodic,MY(^χλ)=0, (4.4)

    and ^wλ is defined in Y by

    ^wλ(y)=λy^χλ(y). (4.5)

    Moreover, setting

    ^wελ(x)=εˆwλ(xε) (4.6)

    in Ωε, one has

    Aε~wελA0λweakly in (L2(Ω))n, (4.7)

    and

    ΩεAεwελvdx =0,vVε. (4.8)

    For any ε, the corrector matrix for perforated domain Cε=(Cεij)1i,jN, introduced in [10], is defined by

    {Cε(x)=C(xε) a.e. inΩε,Cij(y)=^wjyi(y)i,j=1,,Na.e. onY, (4.9)

    where ^wj=^wej and {ej}Nj=1 is the canonical basis of RN.

    We are now ready to present our main corrector result.

    Theorem 4.1. Let uε be the renormalized solution to (2.6) under the assumptions of Theorem 2.6. Then, for any fixed kN, we have

    limε0Tk(uε)CεTk(u0)L1(Ωε)=0. (4.10)

    The proof of Theorem 4.1 is given at the end of this section and makes use of the following results whose proof is given in next section. It makes use of a somehow "regularized truncation" function Hδk (see (4.12)–(4.13)).

    Remark 4.2. Let us observe that, in the quasilinear case, the corrector would have the form Cε(x,t)=C(xε,t), and then, in the corrector result, one should replace t by a function of x. To do that, it is necessary to have at least the measurability of C with respect to t, and the proof requires that C (hence, each derivative of wj) is Lipschitz-continuous in t. This is not the case under our assumptions on A, since this regularity is essentially true only under very strong and global regularity assumptions on A, as proved in [1]. This is not adapted to homogenization, and this is why, in this section, we suppose that A is independent of t.

    Theorem 4.3. Let uε be the renormalized solution to (2.6) under the assumptions of Theorem 2.6. Let kN be fixed, and let H be a nondecreasing function in C2(R) such that H(0)=0, and such that H has a compact support included in [k,k]. Then, for any Φ=(Φ1,Φ2,,ΦN)(D(Ω))N,

    lim supε0H(uε)CεΦL2(Ωε)βαH(u0)ΦL2(Ω),

    where α and β are given by (2.11) and (4.1), respectively.

    Proof of Theorem 4.1. For a fixed k, for all δ>0, there exists Φδ(D(Ω))N such that

    ΦδTk(u0)L2(Ω)δ. (4.11)

    Further, for any δ>0, let HδkC2(R) be a smooth approximation of Tk verifying (see Figure 4)

    Hδk(r)={r, if |r|k2δ,kδ, if rk,k+δ, if rk (4.12)
    Figure 4.  The function Hδk.

    with

    0(Hδk)2 in R. (4.13)

    Observe that, by construction, Hδk satisfies the assumptions of Theorem 4.3; in particular, its support is contained in [k,k].

    In view of the regularity of Tk(u0), and by construction of the function Hδk, we have

    Hδk(u0)Tk(u0)L2(Ω)=ω2(δ), (4.14)

    with limδ0ω2(δ)=0. Let us prove that

    lim supε0Hδk(uε)Tk(uε)L2(Ωε)βαω2(δ). (4.15)

    By the definitions of Hδk and Tk, and using the ellipticity condition (2.11), we have

    Hδk(uε)Tk(uε)2L2(Ωε)=ΩεuεGδk(uε)dx=Ωε(Gδk)(uε)|uε|2dx1αΩε(Gδk)Aεuεuεdx=1αΩεAεuεGδk(uε)dx, (4.16)

    where

    Gδk=r0((Hδk)χ{|s|k})2ds.

    Since the function G=Gδk satisfies the assumptions of Theorem 3.1, from (2.12) and (4.16), and by using (4.3), we obtain

    lim supε0Hδk(uε)Tk(uε)2L2(Ωε)1αΩA0u0G(u0)dxβ2α2Hδk(u0)Tk(u0)2L2(Ω),

    which proves (4.15).

    Hence, from Theorem 4.3, (4.11) and (4.15), we have

    0lim infε0Tk(uε)CεTk(u0)L1(Ωε)lim supε0Tk(uε)CεTk(u0)L1(Ωε)lim supε0Tk(uε)Hδk(uε)L1(Ωε)+lim supε0Hδk(uε)CεΦδL1(Ωε)+lim supε0CεΦδCεTk(u0)L1(Ωε)c1lim supε0Tk(uε)Hδk(uε)L2(Ωε)+c1lim supε0Hδk(uε)CεΦδL2(Ωε)+CεL2(Ω)Tk(u0)ΦδL2(Ω)c1βαω2(δ)+c1βαHδk(u0)ΦδL2(Ω)+c2Tk(u0)ΦδL2(Ω)cω3(δ), (4.17)

    where, for a fixed k, limδ0ω3(δ)=0. This proves (4.10).

    For any ε>0, let uε be a renormalized solution of (2.6). Let kN be fixed, and let H be a nondecreasing function in C2(R) such that H(0)=0 and H has a compact support included in [k,k]; it follows that H(uε)=H(Tk(uε)) belongs to L(Ωε)H10(Ωε).

    In the whole proof of Theorem 4.3, to shorten the notations, we set

    uεk=Tk(uε) and u0k=Tk(u0).

    We consider the quantity αH(uε)CεΦ2L2(Ωε), where Φ=(Φ1,Φ2,,ΦN)(D(Ω))N.

    Since Aε is uniformly coercive,

    αH(uε)CεΦ2L2(Ωε)ΩεAε(H(uε)CεΦ)(H(uε)CεΦ)dx=ΩεAεH(uε)H(uε)dxΩεAεH(uε)(CεΦ)dxΩεAε(CεΦ)H(uε)dx+ΩεAε(CεΦ)(CεΦ)dxI1εI2εI3ε+I4ε. (5.1)

    We will pass to the limit in (5.1) in each term as ε0.

    Let us point out that the difficulties in our situation concern the first three terms studied below in Step 1, Step 2 and Step 3, respectively. In particular, Step 2 requires the most delicate arguments due to the fact that we are dealing with renormalized solutions. Passing to the limit in the last term is standard.

    Step 1. Limit of I1ε

    By defining G(r)=r0H(s)2ds and recalling that the support of H is included in [k,k], we can write

    I1ε=ΩεAεH(uε)H(uε)dx=ΩεAεuεuε(H(uε))2dx=ΩεAεuεG(uε)dx=ΩεAεuεG(uε)dx.

    Since G is a nondecreasing element of W1,(R) such that G(0)=0, Theorem 3.1 leads to

    limε0I1ε=ΩA0u0G(u0)dx=ΩA0H(u0)H(u0)dx. (5.2)

    Step 2. Limit of I2ε

    For the second integral I2ε on the right-hand side of (5.1), by the definition (4.9) of Cε, we have

    I2ε=ΩεAεH(uε)(CεΦ)dx=Ni=1ΩεAεH(uε)(Φi^wεi)dx=Ni=1(ΩεAεH(uε)(Φi^wεi)dxΩεAεH(uε)Φi^wεidx). (5.3)

    Since H belongs to C1(R) and has a compact support, we have H(uε)=H(uε)uε almost everywhere in Ωε. On the other hand, the function ^wεi given by (4.6) belongs to L(Ωε)H1(Ωε), so that Φi^wεiL(Ωε)H10(Ωε). Then, choosing Φi^wεi as a test function and ψ=H in (2.21), we get, for 1iN,

    ΩεH(uε)Aε(x)uε(Φi^wεi)dx=ΩεfH(uε)Φi^wεidx+Γε1gεH(uε)Φi^wεidσxΓε1εγτε(x)H(uε)h(uε)Φi^wεidσxΩεH(uε)Aε(x)uεuεΦi^wεidx. (5.4)

    Thus, to study the behavior of I2ε, it remains to determine the limit of ΩεAεH(uε)Φi^wεidx and the limit of the right hand-side of (5.4) as ε goes to zero.

    By the properties of the unfolding operator, Remark 2.5, the convergences in (2.29) and definition (4.6), we compute

    limε0ΩεAεH(uε)Φi^wεidx=limε0ΩεH(uε)AεTk(uε)Φi^wεidx=limε01|Y|Ω×YH(Tε(uε))Tε(Aε)Tε(Tk(uε))Tε(Φi)Tε(^wεi)dxdy=limε01|Y|Ω×YH(Tε(uε))A(y)Tε(Tk(uε))Tε(Φi)εˆwi(y)dxdy=limε01|Y|Ω×YH(Tε(uε))A(y)Tε(Tk(uε))Tε(Φi)(xiεˆχi(y))dxdy=1|Y|Ω×YH(u0)A(y)(Tk(u0)+y^uk)Φixidxdy. (5.5)

    We now study the behavior of the terms on the right-hand side of (5.4) when ε goes to zero. In view of (4.6) and the properties of H, and applying Corollary 2.7, we have

    limε0ΩεfH(uε)Φi^wεidx=|Y||Y|ΩfH(u0)Φixidx. (5.6)

    The boundary unfolding operator properties and Remark 2.5 give

    Γε1gεH(uε)Φi^wεidσx=εΓε1g(xε)H(uε)Φi^wεidσx=1|Y|Ω×Tg(y)Tbε(H(uε))Tbε(Φi)Tbε(^wεi)dxdσy=1|Y|Ω×Tg(y)H(Tbε(uε))Tbε(Φi)(xiεˆξ(y))dxdσy.

    Next, due to (3.11) and the properties of the boundary unfolding operator, we obtain

    limε0Γε1gεH(uε)Φi(x)^wεidσx=1|Y|Ω×Tg(y)H(u0)Φixidxdσy.

    Since H(u0), Φi and xxi are independent of y, we get, at last,

    limε0Γε1gεH(uε)Φi(x)^wεidσx=|T||Y|MT(g)ΩH(u0)Φi(x)xidx. (5.7)

    Using similar arguments, we get

    limε0Γε1εγτε(x)H(uε)h(uε)Φi^wεidσx=J(γ)|Y|MT(τ)ΩH(u0)h(u0)Φixidx, (5.8)

    where J(γ) is given by (2.27).

    We now turn to the last term of the right-hand side of (5.4). Again, the unfolding operator and Remark 2.5 allow us to write

    ΩεH(uε)Aε(x)uεuεΦi^wεidx=1|Y|Ω×YH(Tε(uε))Tε(Aε)Tε(Tk(uε))Tε(Tk(uε))Tε(Φi)Tε(^wεi)dxdy=1|Y|Ω×YH(Tε(uε))A(y)Tε(Tk(uε))Tε(Tk(uε))Tε(Φi)(xiεˆχi(y))dxdy. (5.9)

    From Theorem 3.3, we have

    Tε(Tk(uε))Tk(u0)+y^uk strongly in L2(Ω×Y) as ε0,

    and since H is a continuous and bounded function, convergence (2.28) implies that

    H(Tε(uε))H(u0) in L(Ω×Y) weak star as ε0.

    By the properties of the unfolding operator, the function Tε(Φi) goes to Φ in L(Ω×Y) weak star as ε0. It follows that

    limε0ΩεH(uε)Aε(x)uεuεΦi^wεidx=1|Y|Ω×YH(u0)A(y)(Tk(u0)+y^uk)(Tk(u0)+y^uk)Φ(x)xidxdy. (5.10)

    Gathering (5.3), (5.4), (5.5), (5.6), (5.7), (5.8) and (5.10), we obtain

    limε0I2ε=Ni=1(|Y||Y|ΩfH(u0)Φixidx+|T||Y|MT(g)ΩH(u0)Φi(x)xidx1|Y|J(γ)MT(τ)ΩH(u0)h(u0)Φixidx1|Y|Ω×YH(u0)A(y)(Tk(u0)+y^uk)(Tk(u0)+y^uk)Φ(x)xidxdy1|Y|Ω×YH(u0)A(y)(Tk(u0)+y^uk)Φixidxdy). (5.11)

    For 1iN, using η0=Φixi and Ψ=^ukΦixi as test functions in (2.31) with S=H and S1=H, and recalling that supp(H)[k,k], we get

    Ω×YH(u0)A(y)(Tk(u0)+y^uk)(Tk(u0)+y^uk)Φ(x)xidxdy+Ω×YH(u0)A(y)(Tk(u0)+y^uk)(Φixi)dxdy+J(γ)MT(τ)ΩH(u0)h(u0)Φixidx=|Y|ΩfH(u0)Φixidx+|T|MT(g)ΩH(u0)Φi(x)xidx.

    Because (Φixi)=Φixi+Φiei, using (2.43) written for Tk(u0), we obtain

    limε0I2ε=Ni=1(1|Y|Ω×YH(u0)A(y)(Tk(u0)+y^uk)eiΦidxdy)=Ni=1(ΩH(u0)A0Tk(u0)eiΦidx)=ΩH(u0)A0Tk(u0)Φdx=ΩA0H(u0)Φdx. (5.12)

    Step 3. Limit of I3ε and I4ε

    From (4.8), we have

    I3ε=ni=1ΩεAεwελΦiH(uε)dx=ni=1ΩεAεwελH(uε)Φidx. (5.13)

    Using convergence (2.34) given in Corollary 2.7, and (4.7), we can pass to the limit in (5.13). Since, here, A0 is constant, we obtain

    limε0I3ε=ΩA0H(u0)Φdx=ΩA0ΦH(u0)dx. (5.14)

    On the other hand, it has been proved in [10] that

    limε0I4ε=ΩA0ΦΦdx, (5.15)

    which ends this step.

    Step 4. Conclusion

    Collecting (5.2), (5.12), (5.14) and (5.15), we have

    limε0(I1εI2εI3ε+I4ε)=ΩA0H(u0)H(u0)dxΩA0H(u0)ΦdxΩA0ΦH(u0)dx+ΩA0ΦΦdx=ΩA0(H(u0)Φ)(H(u0)Φ)dx=ΩA0(H(u0)Φ)(H(u0)Φ)dx. (5.16)

    From (5.1), the ellipticity of the matrix A (see (2.11)) and (4.3), we obtain

    αlim supε0H(uε)CεΦ2L2(Ω)lim supε0(I1εI2εI3ε+I4ε)β2αΩ|H(u0)Φ|2dx, (5.17)

    which concludes the proof.

    The authors declare that there is no conflict of interest.



    [1] M. Artola, G. Duvaut, Un résultat d'homogénéisation pour une classe de problèmes de diffusion non linéaires stationnaires, Ann. Fac. Sci. Toulouse Math., 4 (1982), 1–28. https://doi.org/10.5802/afst.572 doi: 10.5802/afst.572
    [2] P. Bénilan, L. Boccardo, T. Gallouët, R. Gariepy, M. Pierre, J. L. Vazquez, An L1-theory of existence and uniqueness of solutions of nonlinear elliptic equations, Ann. Sc. Norm. Super. Pisa Cl. Sci., 22 (1995), 241–273. https://eudml.org/doc/84205
    [3] G. Papanicolau, A. Bensoussan, J. L. Lions, Studies in Mathematics and its Applications, Asymptotic Analysis for Periodic Structures. Amsterdam: Elsevier, 1978.
    [4] D. Cioranescu, A. Damlamian, P. Donato, G. Griso, R. Zaki, The periodic unfolding method in domains with holes, SIAM J. Math. Anal., 44 (2012), 718–760. https://doi.org/10.1137/100817942 doi: 10.1137/100817942
    [5] D. Cioranescu, A. Damlamian, G. Griso, Periodic unfolding and homogenization, C. R. Math., 335 (2022), 99–104. https://doi.org/10.1016/S1631-073X(02)02429-9 doi: 10.1016/S1631-073X(02)02429-9
    [6] D. Cioranescu, A. Damlamian, G. Griso, The periodic unfolding method in homogenization, SIAM J. Math. Anal., 40 (2008), 1585–1620. https//doi.org/10.1137/080713148 doi: 10.1137/080713148
    [7] D. Cioranescu, A. Damlamian, G. Griso, The Periodic Unfolding Method, Theory and Applications to Partial Differential Problems, Singapore: Springer, 2018. https//doi.org/10.1007/978-981-13-3032-2
    [8] D. Cioranescu, P. Donato, An Introduction to Homogenization, Oxford: Oxford University Press, 1999.
    [9] G. Dal Maso, F. Murat, L. Orsina, A. Prignet, Renormalized solutions of elliptic equations with general measure data, Ann. Sc. Norm. Super. Pisa, Cl. Sci., IV. Ser., 28 (1999), 741–808. https://eudml.org/doc/84396
    [10] P. Donato, A. Gaudiello, L. Sgambati, Homogenization of bounded solutions of elliptic equations with quadratic growth in periodically perforated domains, Asymptotic Anal., 16 (1998), 223–243.
    [11] P. Donato, O. Guibé, A. Oropeza, Homogenization of quasilinear elliptic problems with nonlinear Robin conditions and L1 data, J. Math. Pures Appl., 120 (2018), 91–129. https//doi.org/10.1016/j.matpur.2017.10.002 doi: 10.1016/j.matpur.2017.10.002
    [12] O. Guibé, A. Oropeza, Renormalized solutions of elliptic equations with Robin boundary conditions, Acta Math. Sci., 37 (2017), 889–910. https//doi.org/10.1016/S0252-9602(17)30046-2 doi: 10.1016/S0252-9602(17)30046-2
  • This article has been cited by:

    1. Patrizia Donato, Rheadel Fulgencio, Olivier Guibé, Convergence of the Energy and Correctors for Some Elliptic Problems in a Two-Component Domain With Weak Data, 2025, 22, 1660-5446, 10.1007/s00009-025-02858-7
  • Reader Comments
  • © 2023 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(1824) PDF downloads(79) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog