Loading [MathJax]/jax/element/mml/optable/MathOperators.js

Homogenization of stiff inclusions through network approximation

  • Received: 01 June 2021 Revised: 01 January 2022 Published: 18 February 2022
  • 35R60, 94C05, 78M40

  • We investigate the homogenization of inclusions of infinite conductivity, randomly stationary distributed inside a homogeneous conducting medium. A now classical result by Zhikov shows that, under a logarithmic moment bound on the minimal distance between the inclusions, an effective model with finite homogeneous conductivity exists. Relying on ideas from network approximation, we provide a relaxed criterion ensuring homogenization. Several examples not covered by the previous theory are discussed.

    Citation: David Gérard-Varet, Alexandre Girodroux-Lavigne. Homogenization of stiff inclusions through network approximation[J]. Networks and Heterogeneous Media, 2022, 17(2): 163-202. doi: 10.3934/nhm.2022002

    Related Papers:

    [1] Mohamed Belhadj, Eric Cancès, Jean-Frédéric Gerbeau, Andro Mikelić . Homogenization approach to filtration through a fibrous medium. Networks and Heterogeneous Media, 2007, 2(3): 529-550. doi: 10.3934/nhm.2007.2.529
    [2] Xiangdong Du, Martin Ostoja-Starzewski . On the scaling from statistical to representative volume element in thermoelasticity of random materials. Networks and Heterogeneous Media, 2006, 1(2): 259-274. doi: 10.3934/nhm.2006.1.259
    [3] Marc Briane, David Manceau . Duality results in the homogenization of two-dimensional high-contrast conductivities. Networks and Heterogeneous Media, 2008, 3(3): 509-522. doi: 10.3934/nhm.2008.3.509
    [4] Mario Ohlberger, Ben Schweizer, Maik Urban, Barbara Verfürth . Mathematical analysis of transmission properties of electromagnetic meta-materials. Networks and Heterogeneous Media, 2020, 15(1): 29-56. doi: 10.3934/nhm.2020002
    [5] Mohamed Camar-Eddine, Laurent Pater . Homogenization of high-contrast and non symmetric conductivities for non periodic columnar structures. Networks and Heterogeneous Media, 2013, 8(4): 913-941. doi: 10.3934/nhm.2013.8.913
    [6] Paolo Luzzini, Paolo Musolino . Perturbation analysis of the effective conductivity of a periodic composite. Networks and Heterogeneous Media, 2020, 15(4): 581-603. doi: 10.3934/nhm.2020015
    [7] Natalia O. Babych, Ilia V. Kamotski, Valery P. Smyshlyaev . Homogenization of spectral problems in bounded domains with doubly high contrasts. Networks and Heterogeneous Media, 2008, 3(3): 413-436. doi: 10.3934/nhm.2008.3.413
    [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] Giovanni Scilla . Motion of discrete interfaces in low-contrast periodic media. Networks and Heterogeneous Media, 2014, 9(1): 169-189. doi: 10.3934/nhm.2014.9.169
    [10] 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
  • We investigate the homogenization of inclusions of infinite conductivity, randomly stationary distributed inside a homogeneous conducting medium. A now classical result by Zhikov shows that, under a logarithmic moment bound on the minimal distance between the inclusions, an effective model with finite homogeneous conductivity exists. Relying on ideas from network approximation, we provide a relaxed criterion ensuring homogenization. Several examples not covered by the previous theory are discussed.



    The classical theory of homogenization deals with conductivity matrices Aε=A(xε) having uniform lower and upper bounds: αIdAβId, with 0<α<β<+. Degenerate cases, for which A is allowed to vanish or to take infinite values, still lead to open questions. A problem typical of the first situation is as follows. Given a union F=III of closed connected domains I (the inclusions), either periodic or random stationary, one considers the following system:

    {Δuε=fin UFε,uε|U=0,νuε|Fε=0. (1)

    Here, U is a smooth bounded domain of R3, and Fε=εI where the union is restricted to inclusions satisfying εIU. The Neumann condition corresponds to zero conductivity inside the inclusions, while conductivity is normalized to 1 outside. The problem is to determine a homogenized limit model. Namely, one tries to understand under which conditions the solution uε of 1 converges as ε0 to the solution u0 of

    {div (A0u0)=(1λ)fin U,u0|U=0, (2)

    for some non-degenerate effective conductivity matrix A0. The constant λ refers to the density of the inclusions, and is the weak limit of 1Fε. This homogenization problem has become standard, and is discussed extensively in the classical book [27]. In brief, homogenization holds under some mild but uniform regularity requirement on the geometry of the inclusions, and in a general random stationary ergodic setting, as long as R3F is connected. One strategy to show this result is through an adaptation of the classical div-curl approach. With regards to this strategy, a nice feature is that the first equation and the Neumann condition are equivalent to the single equation

    div (1Fεuε)=1Fεf in U

    meaning that the divergence of the flux 1Fεuε is relatively compact in H1(U). On the contrary, one difficulty is that uε is defined only on UFε and there is no canonical extension as a gradient of an H1 function over the whole domain U. Still, construction of such a potential field extension inside the inclusions is possible under mild requirements. Let us stress that such extension is a local process, in the sense that extension in one inclusion can be considered independently from the others.

    As regards the opposite case of inclusions with infinite conductivity, the theory of homogenization is less complete. The analogue of 1 reads:

    {Δuε=fin UFε,uε=0in Fε,uε|U=0,Iενuε=0,IεCC(Fε). (3)

    Here, we denote Iε:=εI, that belongs to the set CC(Fε) of connected components of Fε. On such inclusions, the condition uε=0 ensures that the potential is constant, which corresponds to infinite conductivity. The integral condition Iενuε=0 corresponds to the fact that the total flux through the boundary of the inclusion is zero (a continuous analogue of Kirchoff's nodal rule). Let us note that system 3 has an important extension to modelling of suspensions in fluid mechanics, in which the Laplace operator is replaced by the Stokes operator, the electric field being replaced by the Newtonian stress tensor of the fluid. The analogue question of the effective viscosity of passive suspensions has attracted a lot of attention recently, cf. [25,2,30,26,20,21,22,19,16,15,17] among many.

    Again, the point is to show convergence to the effective system 2. However, contrary to the case of soft inclusions 1, the case of stiff inclusions (in the terminology of [27]) requires more than the connectedness of R3F. Some condition on the distance between particles is needed. The main difference between 1 and 3, responsible for extra assumptions in the latter case, can be seen when trying to adapt the div-curl approach. The difficulty is reversed. While the potential field uε has a natural extension as a potential field over U (just extend uε by its constant value inside each inclusion), the divergence of the flux 1Fεuε is no longer controlled in H1(U). The point is then to find a nice extension inside the inclusions of a field (namely uε) with given divergence (namely f). This constraint on the divergence makes such extension process non-local, and implies global assumptions on the configuration of the inclusions.

    Up to our knowledge, homogenization was so far only established by Zhikov for a random spherical structure F=iNBi, under the assumption that almost surely,

    lim supN+1N3Bi(N,N)3μi<+, where μi:=|ln(d(Bi,FBi))|. (4)

    See Theorem 2.6 below, or [27,chapter 8] for more. Obviously, in the case where there is a minimal distance δ>0 between the spheres, such condition is satisfied. As will be discussed later in the paper, the logarithmic term μi comes from the fact that given any smooth Dirichlet data φi on Bi, there is an H1 function ϕ with ϕ=φi on Bi and ϕ=0 on all other Bj's, such that it's energy R3|ϕ|2 explodes at most like μi. This gives a glimpse on how such assumption may help to build the extension operator alluded to above.

    In the opposite very dense setting where for all i, |μi|1, homogenization is not always possible. As a consequence, one can find configurations F and boundary data problems of the form

    {Δuε=0in UFε,uε=0in Fε,uε|U=φ,Iενuε=0,IεCC(Fε) (5)

    for which

    lim infε0U|uε|2=+.

    Such negative results for dense settings rely notably on the so-called network approximation method, as described in the monograph [7]. The idea behind this method is that when the spheres get close to one another, the analysis of systems of type 5 can be simplified: the asymptotic behaviour of the system (notably of its energy), can be deduced from the properties of an underlying weighted graph where:

    ● nodes of the graph are the spheres of infinite conductivity

    ● edges are pairs of spheres close to one another, in the sense that they belong to adjacent Voronoi cells

    ● each edge e={Bi,Bj} has weight μe:=|lnd(Bi,Bj)|.

    For instance, the energy of the system can be approximated as the parameter δ(ε):=ε1maxIεFεd(Iε,FεIε) goes to zero when ε goes to zero by a reduced discrete energy associated to the graph. Such energy may then be analyzed by tools of graph theory, and its divergence as ε goes to zero established. We refer to [13,6,8,9], and again to the monograph [7] and its bibliography. See also [4,5,10] in the context of fluid mechanics and suspensions.

    The goal of this paper is to investigate some intermediate situations, in which the condition on the distance between the inclusions 4 is not necessarily satisfied, but homogenization is still possible. To identify situations of this kind, we will rely on two notions:

    ● the notion of multigraph of inclusions, reminiscent of the network approximation just mentioned

    ● the notion of short of inclusions, reminiscent of the study of electrical circuits.

    Thanks to these notions, we will formulate two assumptions H1-H2. The weaker one, H1, allows to define the homogenized matrix A0, while the stronger one, H2, allows for homogenization. This set of two assumptions is implied by a logarithmic moment bound on the minimal distance, but is much more general. For instance, our homogenization theorems apply to the case of inclusions or clusters of inclusions of arbitrairily large size, with a mere moment bound on their diameter. Further examples, like inclusions with anisotropic structure, will be discussed.

    The outline of the paper is as follows. In Section 2, after a brief reminder on stationary closed sets, we describe the class of inclusions under consideration. For such class, we define multigraphs and shorts of inclusions. We conclude the section by the statement of our homogenization results. In Section 3, we show certain extension properties for potential and solenoidal vector fields, crucial to the proof of homogenization. This proof is then given in Section 4. Eventually, Section 5 provides complements on our main hypothesis H1-H2.

    We follow here the definition in the lecture notes [11,chapter 13]. A random closed set is a random variable F=F(ω)R3 from a probability space to the set

    CL={FR3,Fclosed}

    equipped with the borelian σ-algebra B(CL). We remind that the topology on CL is the topology induced by the sets CLK={F,FK}, K describing the compact subsets of R3. Note that by considering the law PF of F on CL, we can always assume that the probability space is (CL,B(CL),PF) and that F(ω)=ω. This is the canonical representation of the random closed set. We now introduce the shift τx:CLCL, τx(ω)=ωx, xR3. We say that the random closed set F is R3-stationary, resp. Z3 stationary, if PFτ1x=PF for all xR3, resp. xZ3. We say that it is ergodic if under the assumption that τx(A)=A for all xR3, resp. τx(A)=A for all xZ3, one has PF(A){0,1}.

    Note that there is a trick to turn a Z3-stationary ergodic random closed set into an R3-stationary ergodic random closed set. Namely, given F a Z3-stationary random closed set, one defines

    ˜F:(CL×(0,1)3,PFLeb)CL,(ω,x)ωx.

    Then, one considers on CL the measure image of PFLeb by ˜F, and denoting P˜F such a measure, one can check that P˜Fτ1x=P˜F for all xR3. Indeed, given a borelian A of CL, by Fubini's theorem,

    P˜Fτ1x(A)=PF×Leb({(ω,y),ωyA+x})=(0,1)3CL1A+x(ωy)dPF(ω)dy=(0,1)3PF(τxy(A))dy.

    By Z3-stationarity of PF, the integrand is Z3-periodic, and the result follows by the change of variable y=y+x. Moreover, ˜F is ergodic if F is. Indeed, if τy(A)=A fordiv all yR3, we find (take x=0 in the previous identity):

    P˜F(A)=(0,1)3PF(τy(A))dy=PF(A){0,1}.

    Note that a property that is almost sure with respect to P˜F will hold for realizations of the form ωx, that is F(ω)x, for PF-almost every ω and almost every x. As the choice x=0 characterizing F made at the beginning is irrelevant, this gives results about the original random closed set F.

    More generally, let (Fi)iI a family of random closed sets defined on the same probability space, and P its joint law on CLI. We define the shift τx:CLICLI, τx((ωi)iI)=(x+ωi)iI, xR3. We say that (Fi)iI is R3-stationary, resp. Z3-stationary, if Pτ1x=P for all xR3, resp. xZ3. We say that it is ergodic if under the assumption that τx(A)=A for all xR3, resp. τx(A)=A for all xZ3, one has P(A){0,1}. Of course, stationarity of the family is stronger than the stationarity of each of its elements, and is the right notion as soon as one deals with events involving intersections, unions of several random closed sets Fi.

    A convenient unified description, that we adopt from now on, is the following. We consider a probability space (Ω,A,P), equipped with a family of maps τx:ΩΩ, xR3, which satisfies:

    i) (x,ω)τx(ω) measurable

    ii) x,y,τx+y=τxτy (shift)

    iii) x,P=Pτ1x (measure preserving)

    iv) (x,τx(A)=A)P(A){0,1} (ergodicity).

    Under this description, an (ergodic) stationary closed set is then a r. v. F:ΩCL s.t.

    F(τx(ω))=F(ω)x,ωΩ,xR3.

    Introducing the subset F of Ω defined by

    F={ωΩ,0F(ω)} (6)

    one can notice that

    F(ω)={x,τx(ω)F}.

    This is the point of view taken in [27].

    Our homogenization results apply to a class of ergodic stationary closed sets F=F(ω) that we now describe. Let F be a closed set and CC(F) its family of connected components, so that F=ICC(F)I. Each of this connected component is of course closed. We introduce the following geometric conditions:

    (G1) Regularity of the inclusions : there exists d>0, such that all ICC(F) is the closure of a C2 bounded domain satisfying an interior and exterior ball condition with uniform radius d.

    (G2) Geometry of the gaps : there exists δ,a>0, such that for all ICC(F), the set

    I{d(x,FI)δ}

    has a finite number of connected components Iα, α=1,,NI, with

    supINI<+, and with at most one couple (J,β)(I,α) such that d(Iα,Jβ)2δ. Moreover, for an appropriate local system of cylindrical coordinates (r,θ,z):

    Iα{zd(Iα,Iβ)/2+ar2},Iβ{zd(Iα,Iβ)/2ar2}, (7)

    that is Iα and Iβ are separated by paraboloids with uniform curvature.

    Remark 1. By 7, there is a unique couple of points (xI,α,xJ,β)Iα×Jβ such that

    |xI,αxJ,β|=d(Iα,Jβ).

    See Figure 1.

    Figure 1. 

    Geometry of the inclusions with a close-up on a gap

    .

    Definition 2.1. (Admissible set of inclusions)

    We say that an ergodic stationary closed set F=F(ω) is an admissible set of inclusions if it satisfies (G1)-(G2) almost surely, with d,δ,a and supINI bounded by a deterministic constant.

    Remark 2. At this stage, we do not assume anything on the diameter of the inclusions.

    We now associate to a closed set satisfying (G1)-(G2) an unoriented multigraph that we call multigraph of inclusions. Roughly, the nodes of the multigraph are the connected components of the closed set, and we link pieces of these connected components that are δ-close. See Figure 2 for an illustration. In a more formal way:

    Figure 2. 

    Spherical set up. The graph obtained with the whites lines is isomorphic to the multigraph of inclusions

    .

    Definition 2.2. (Multigraph of inclusions)

    Let F a closed set satisfying (G1)-(G2). For δ the constant in (G2), the δ-multigraph of inclusions associated to F, called multigraph of inclusions for brevity, is the unoriented multigraph Gr(F)=(CC(F),Ed(F)) with set of nodes CC(F) and set of edges Ed(F) made of elements of the form

    e=[xI,α,xJ,β], with IJCC(F),1αNI,1βNJ,d(Iα,Jβ)δ.

    We say that an edge e as above connects I to J, and denote it IeJ. If there exists eEd(F) such that IeJ, we simply note IJ. Note that several edges can connect the same pair of nodes, hence the multigraph structure. This corresponds to multiple gaps for a given pair of inclusions. To each of the edges, we associate a weight μe, by the formula

    μe:=|ln|e||=|lnd(Iα,Jβ)| for e=[xI,α,xJ,β],|e|=|xJ,βxI,α|. (8)

    As explained in the introduction, there is a strong analogy between our multigraph of inclusions and the network approximation of [7].

    For later use, we further define

    Definition 2.3. (Cluster of inclusions)

    Let F satisfying (G1)-(G2). A cluster of F is a union of all the inclusions that are the nodes of a connected component of the δ-multigraph Gr(F) (not to be confused with a connected component of F itself, which corresponds to a single inclusion/node).

    Remark 3. For F a set of inclusions, and yR3, we shall denote:

    Iy,FCC(F) the inclusion containing y, with the convention that Iy,F= if yF.

    Cy,FCC(F) the cluster containing y, with the convention that Cy,F= if yF.

    In the case F=F(ω) is stationary, one has clearly

    Iy,F(ω)=I0,F(τy(ω)),Cy,F(ω)=C0,F(τy(ω)).

    This implies that (y,ω)diam(Iy,F(ω)), (y,ω)diam(Cy,F(ω)) and (y,ω)(Cy,F(ω)) are stationary random fields. We will apply the ergodic theorem several times to these fields.

    A last notion we need to explain before stating our main results is the notion of short of inclusions. It is directly inspired from the study of electrical networks and their associated multigraphs:

    Definition 2.4. (Short of a multigraph)

    Let G=(V,E) a multigraph. Let I,JV. The short of G at {I,J} is the multigraph obtained by identifying nodes I and J (and suppressing all edges joining I and J). More generally, given a set of pairs of nodes S, the short of G at S is the multigraph obtained from G by identifying all pairs in S (and suppressing all edges joining these nodes):

    We say that G is a short of G if there exists a set S of pairs of nodes such that G is the short of G at S.

    For electrical circuits, it is well-known that nodes with the same potential can be shorted, without changing the values of the currents through the remaining edges. This is a useful fact, as one can short the multigraph of an electrical circuit to simplify calculations. In other words, shorting two nodes is the same as imposing the same potential on each of the node, which can be interpreted as having zero resistance between those nodes. We refer to [12,chapter 2] for more.

    At the level of the inclusions, an analogue of a short between two nodes consists in bridging the gap between two inclusions, so as to obtain one single connected component out of the two, with a single potential. We introduce the following

    Definition 2.5. (Short of inclusions)

    Let F a closed set satisfying (G1)-(G2) and Gr(F) its multigraph of inclusions, see Definition 2.2. We say that a closed set FF is a short of F if

    F satisfies (G1)-(G2) (with some constants d,δ,a,NI instead of d,δ,a,NI)

    Ed(F)Ed(F)

    ● there exists η>0, such that for all eEd(F)Ed(F), one has

    {d(x,e)<η}FFeEd(F)Ed(F){d(x,e)<2η}.

    Furthermore, if F is an admissible set of inclusions, see Definition 2.1, F is an admissible short of F if it is itself an admissible set, and a short of F almost surely for some deterministic η.

    As explained above, a short of F is obtained by filling some gaps of F. See Figure 3. If F is a short of F, Gr(F) is isomorphic to a short of Gr(F).

    Figure 3. 

    Two inclusions configuration, shorted on the right

    .

    We now turn to the homogenization problem described roughly in the introduction. For the rest of this section, we assume that F=F(ω)=ICC(F(ω))I(ω) is an admissible set of inclusions, see Definition 2.1. Given a parameter ε>0, δ0>0 and a bounded domain U, we denote

    Iε:=εIICC(F),Fε:=IεU,d(Iε,U)>εδ0Iε. (9)

    Let fL6/5(U), and uε=uε(ω)H10(U) the solution of 3. Existence and uniqueness of uε is standard, and comes with the estimate

    uεL2(U)CfL6/5(U).

    In particular, uε has a subsequence that weakly converges in H10(U). The point is to identify the limit u0, and to determine under which conditions it satisfies a system like 2, where λ=E1F is the average density of the set of inclusions.

    As mentioned in the introduction, Zhikov has tackled the homogenization problem, for a collection of random spheres of unit radius, under a logarithmic moment bound on the minimal distance between the spheres. The result extends easily to admissible sets of inclusions, under a uniform bound on their diameter. It can further be expressed in terms of multigraphs of inclusions. Namely, for all N>0, let

    QN:=(N,N)3,FN:=ICC(F),IQNI. (10)

    Then the homogenization theorem of Zhikov reads

    Theorem 2.6. (Zhikov, [27]) Let F an admissible set of inclusions satisfying almost surely: supICC(F)diam(I)<+, and such that

    lim supN+1|QN|eEd(FN)μe<+. (11)

    Then, almost surely, the whole sequence uε converges weakly in H10(U) to the solution u0 of a system of type 2.

    Remark 4. As pointed out to us by A. Gloria, it can be seen from the ergodic theorem that: supICC(F)diam(I)<+   a.s.  supICC(F)diam(I)D a.s for some deterministic D.

    As usual in stochastic homogenization, the effective conductivity matrix A0 can be expressed in terms of a variational problem in probability. Following [27,chapter 8], we first introduce the effective resistance matrix B0 : it is the symmetric matrix defined by

    ξR3,B0ξξ:=infzEQ1F|ξ+z|2,

    where the infimum is taken over the set of vector fields z=z(y,ω)L2(Ω,L2loc(R3)) satisfying

    i) z solenoidal, that is divyz=0

    ii) z stationary, that is z(y+y,ω)=z(y,τy(ω))

    iii) Ez=0.

    Then, part of the proof of Theorem 2.6 is to show that B0 is invertible, and that the effective conductivity matrix is given by

    A0=(B0)1. (12)

    Our goal is to relax the assumption in the previous theorem, notably to identify configurations for which no condition on the minimal distance is needed. Our criterion for homogenization will be expressed again through multigraphs of inclusions. A key role will be played by the following discrete energy functional: for F a closed set satisfying (G1)-(G2), and two families {uI}, indexed by ICC(F), and {bIJe} indexed by triplets with I,JCC(F), and eEd(F) s.t. IeJ, we introduce the energy functional

    E(F,{uI},{bIJe})=I,JCC(F)eEd(F),IeJμe|bIJebJIe+uIuJ|2+ICC(F)|I||uI|2. (13)

    We start with the definition of the homogenized matrix. We remind the notation I0,F from Remark 3. We denote xS the center of mass of a subset S of R3.

    Proposition 1. (Existence of the homogenized matrix)

    Let F an admissible set of inclusions. If Ediam(I0,F)2<+ and if almost surely

    lim supN+inf{uI}1|QN|E(FN,{uI},{bIJe=ξxI})<+ (H1)

    where E is given in 13, and FN in 10, then the matrix A0 introduced in 12 is well-defined.

    Compared to the assumptions in Theorem 2.6, those of Proposition 1 are better in two regards.

    a) The uniform bound on the diameter of the inclusions (see Remark 4) is replaced by a moment bound. This possibility of considering inclusions of arbitrary large size, interesting in its own, will turn very useful when combined to our homogenization result involving shorts, cf. Theorem 2.7.

    b) As mentioned in the introduction, H1 is weaker than 11. Indeed, for inclusions satisfying supICC(F)diam(I)<+, the latter clearly implies

    lim supN+1|QN|E(FN,{0},{ξxI})<+.

    In other words, Zhikov's condition 11 corresponds to the simplest choice uI=0 for every inclusions I, while H1 corresponds to an optimization over all possible constants uI. This is not just a technical improvement: as an interesting application, we will prove in Section 5 that H1 is verified by sets F satisfying the moment bound

    Ediam(C0,F)2<+

    where C0,F is the cluster of F containing 0, cf. Remark 3. Further discussion of H1 will be provided in Section 5, notably its relation to the graph Laplacian or to the subadditive ergodic theorem.

    We now turn to the homogenization of 3. It is an interesting open problem to be able to perform homogenization under the mere assumption H1. We must here strengthen it a little.

    Theorem 2.7. (Homogenization of stiff inclusions)

    Let F an admissible set of inclusions. We assume that there exists an admissible short F of F and some s(3,6) (s(3,) in the case F=F) such that, almost surely:

    lim supN+1|QN|sup{bIJe}inf{uI}E(FN,{uI},{bIJe})/(bIJe)2s<+ (H2)

    where

    (bIJe)s=(1|QN|I,JCC(FN)eEd(FN),IeJ|bIJe|s)1s.

    Then, there exists p=p(s) such that under the additional condition:

    Ediam(I0,F)p<+, (14)

    the solution uε of 3 converges a.s. weakly in H10(U) to the solution u0 of 2, with A0 defined by 12.

    Here are a few remarks, to be complemented in Section 5.

    i) Of course, in previous statements, it is enough that all assumptions involving the δ-multigraph of F hold for some δ>0. In practice, one should take δ small, to have a reduced number of edges in the multigraph.

    ii) In practice, the following reformulation of H2 will be used:

    One can find almost surely an M=M(ω)>0 satisfying: for all N>0, for any family {bIJe} indexed by I,J,eCC(FN)2×Ed(FN) with IeJ, there exists a family {uI}ICC(FN) such that

    E(FN,{uI},{bIJe})M|QN|(1|QN|I,JCC(FN)eEd(FN),IeJ|bIJe|s)2/s.

    iii) Let F an admissible set of inclusions, F an admissible short of F. If

    Ediam(I0,F)s<+

    then H2 implies H1. See Lemma 5.5. In particular, by Proposition 1, the matrix A0 is well-defined.

    iv) H2 is implied by the following logarithmic moment bound, see Lemma 5.4:

    lim supN+1|QN|eEd(FN)μke<+,k=(s2)=ss2. (15)

    In the case where F=F, s can be taken arbitrarily large in Theorem 2.7, so that this sufficient condition is almost the standard one (one can take any k>1). For a general short F, we are limited to s<6 for technical reasons, hence to k>32.

    iv) A corollary of Theorem 2.7, to be established in Section 5 and illustrated in Figure 4, is the following:

    Figure 4. 

    Cycle-free configuration set up

    .

    Corollary 1. Let F an admissible set of inclusions with a.s. supICC(F)diam(I)<+. Assume that Gr(F) is cycle-free, and that

    E(C0,F)p<+,for somep>2,

    where C0,F is the cluster of F containing 0, cf. Remark 3. Then, H2 is satisfied with F=F and s=2pp2, so that homogenization holds.

    v) Another important corollary of Theorem 2.7 is:

    Corollary 2. Let F an admissible set of inclusions. Let s(3,6) and p=p(s) as in Theorem 2.7. If

    Ediam(C0,F)p<+

    then homogenization holds.

    Indeed, if we simply bridge all pairs of nodes I,J with IJ, we obtain in this way an admissible short F of F with Ediam(I0,F)p<+, and with Ed(F)=, hence trivially satisfying H2 and 14. See Figure 5 for an illustration.

    Figure 5. 

    On the left, all clusters are far away from the others. On the right, groups of inclusions joined by a grey line form a short F that verifies Ed(F)=

    .
    Figure 6. 

    Separation of the domain P=P1P2 with the boundary Σ

    .
    Figure 7. 

    Geometry of the gap

    .

    Let us point out that the sharper bound

    Eexp(diam(C0,F)γ)<+, for one γ>1

    has been shown to hold for several examples of admissible sets of inclusions, satisfying strong mixing assumptions, below the critical percolation threshold. We refer to the recent article [18] for details, where, moreover, the homogenization of both Laplace and Stokes equation is obtained in this special case by a different method.

    vi) The local blow-up of the energy between two close inclusions I and J linked by an edge e depends on the geometry of the narrow gap between the two inclusions. The assumption (G2) ensures that the gap is well separated by paraboloids of uniform curvature a, which leads to a blow-up of the energy of order |ln|e||. More precisely, one can show (see Lemma A.1), that the explosion is of order |ln|e||a, which means that flatter boundaries lead to a greater explosion. One could change the expression of the weight μe in 8 and track the dependence on the curvature a. Beyond that, let us remark that the analysis of the local energy explosion as a function of gap geometry was extensively studied in [24] and could be used to relax our geometrical assumption (G2) to include more degenerate curvatures.

    vii) In this work, the dimension n=3 case was considered as it is the most pertinent for applications. A similar analysis could be performed in other dimensions. In dimension n>3, a straightforward adaptation of Lemma A.1 shows that there is no blow-up of the energy when inclusions are close. In that case, a simple deterministic bound on the inclusions diameter is enough to perform the whole homogenization process. In dimension n=2, the explosion between two close inclusions in Lemma A.1 is of order (ν)1. Replacing the weight μe=|ln|e|| by (|e|)1 in 8, the Proposition 1 still holds and the homogenization theorem 2.7 remains correct under the same assumption H2 but with exponent s>4 instead of s>3.

    In this whole section, F is an admissible set of inclusions, U is a bounded domain, and Fε is defined as in 9. We will show how assumptions H1 or H2 allow to construct suitable extensions of H1 fields given in the inclusions, that is in Fε, resp. divergence-free L2 fields given outside the inclusions, that is in UFε. By suitable, we mean that the extension operator will be bounded uniformly in ε. These extension results will be central to the homogenization process.

    The two main results of this paragraph are

    Proposition 2. Let ξR3. Assume H1. If Ediam(I0,F)2<+, one can find almost surely C>0 independent of ε and a field ϕεH10(U) with

    ϕε=ξinFε,ε1ϕεL2(U)+ϕεL2(U)C|ξ|.

    Proposition 3. Assume H2 with F=F and s>3. For any ˜s>s, there exists p=p(˜s) such that if

    Ediam(I0,F)p<+,

    one can find almost surely C>0 independent of ε satisfying: for all φεW1,˜s(Fε), there exists a field ϕεH10(U) with

    ϕε=φεinFε,ε1ϕεL2(U)+ϕεL2(U)CφεL˜s(Fε).

    We will focus on the proof of the latter proposition, as the former requires only minor modifications.

    Proof. In all the proof, the realization ω is fixed and does not show up in the notations. We will use the following Poincaré-Wirtinger inequality: there exists r>0 such that for all t[1,+), one can find C>0 satisfying:

    ICC(F),uW1,t(I),u(u)ILt(I)Cdiam(I)ruLt(I), (16)

    with (u)I=1|I|Iu. A main point here is that the constant is bounded by a power of the diameter. This inequality is known to be true with r=1 for convex domains, and more generally for star-shaped domains, see [23,chapter 12]. We indicate in Appendix B how to show that this inequality holds with r=12. Obviously, in the special case where u(x)=ξx, relevant to Proposition 2, the exponent r=1 is enough. Let

    N:=ε1,˜FN:=NFε=IFNU,d(I,NU)>δ0I

    the set of all the connected components of F included in NU and δ0 far from the boundary. By a scaling argument, it is enough to prove that there exists a constant C>0 independent of N such that for any φW1,˜s(˜FN), one can find ϕNH10(NU) with

    ϕN=φ in ˜FN,ϕNH1(NU)CN3˜s62˜sφL˜s(˜FN). (17)

    Let φW1,˜s(˜FN) and denote φI:=φ|I for any ICC(˜FN). The proof is split into two main steps:

    1.Given an arbitrary family (uI)ICC(˜FN), and any e=[xI,α,xJ,β]Ed(˜FN), we build a local extension ϕe, defined in a neighborhood Ve of e that contains Iα and Jβ, such that

    ϕe|I=φIIφI+uI,ϕe|J=φJJφJ+uJ (18)

    and

    ϕe2H1(Ve)Cμe|φI(xI,α)IφIφJ(xJ,β)+JφJ+uIuJ|2+C(|uI|2+|uJ|2)+C(diam(I)2rφI2L2(I)+diam(J)2rφJ2L2(J))+C(diam(I)2rφI2Ls(I)+diam(J)2rφJ2Ls(J)) (19)

    2.We show that for a proper choice of the family (uI) and with the help of the previous local extensions, there exists a global extension ϕN satisfying 17.

    Step 1. Let e=[xI,α,xJ,β]Ed(˜FN). Clearly, to show the existence of ϕe satisfying 18-19, one can restrict to the case where

    IφI=0I, so that, for all  t2,φIW1,t(I)Cdiam(I)rφILt(I)

    thanks to 16. Let now Pα and Pβ two paraboloids that enclose Iα and Jβ. By assumption (G2), for appropriate local cylindrical coordinates (r,θ,z) centered at xI,α+xJ,β2 and of axis along e, we can write

    Pα={z|e|2+ar2},Pβ={z|e|2ar2}.

    By Lemma A.1, given d>0 there exists a function weH1(PαPβFe(d)) where in the same system of local coordinates:

    Fe(d):={r2d2,ar2+|e|2z|e|2ar2}

    and where

    0we1,we|Pα=1,we|Pβ=0,Fe(d)|we|2dxCμe,Fe(d)|we|2|xxI,α|2γdxCγγ>0.

    We take d>0 large enough so that IαJβ lies in the interior of PαPβFe(d).

    Now, using the Stein extension operator from W1,t(I) to W1,t(R3), see [31,Chapter 6], one can find an extension ˜φI of φI on R3 such that for all t[2,˜s],

    ˜φIW1,t(R3)cφIW1,t(I)cdiam(I)rφILt(I) ICC(F)

    We remark that the analysis in [31,Chapter 6] provides constants c and c that are independent of the size of the inclusions. This is however not a crucial point here, as we could handle constants diverging polynomially in the diameter of the inclusions. Eventually, we define

    ϕe:=we(˜φI+uI)+(1we)(˜φJ+uJ)

    that we consider as a function of W1,s(Ve), with

    Ve:=PαPβFe(d){|z|R} (20)

    where R is large enough so that IαJβ lies in the interior of Ve. The intersection with {|z|R} is just here to make Ve bounded. Clearly, 18 is satisfied. We then compute

    Fe(d)|ϕe|2dxFe(d)|we(˜φI+uI˜φJuJ)+we˜φI+(1we)˜φJ|2dx.

    Hence,

    Fe(d)|ϕe|2dxC(Fe(d)|we|2|˜φI+uI˜φJuJ|2dx+Fe(d)|˜φI|2+|˜φJ|2dx)C(Fe(d)|we|2|φI(xI,α)+uIφJ(xJ,β)uJ|2dx+Fe(d)|we|2|˜φIφI(xI,α)+φJ(xJ,β)˜φJ|2dx+diam(I)2rφI2L2(I)+diam(J)2rφJ2L2(J)).

    Thanks to Morrey's inequality ˙W1,s(R3)C1,γ(R3), γ=13s(0,1), one can write

    |˜φI(x)φI(xI,α)|C|xxI,α|γ˜φILs(R3)C|xxI,α|γdiam(I)rφILs(I)

    for any x in the gap Fe(d). Thus, we have the following

    Fe(d)|we|2|˜φIφI(xI,α)+φJ(xJ,β)˜φJ|2dxCdiam(I)2rφI2Ls(I)Fe(d)|we|2|xxI,α|2γdx+Cdiam(J)2rφJ2Ls(J)Fe(d)|we|2|xxJ,β|2γdxC(diam(I)2rφI2Ls(I)+diam(J)2rφJ2Ls(J)).

    Finally, combining the previous inequalities entails

    Fe(d)|ϕe|2dxC(μe|φI(xI,α)+uIφJ(xJ,β)uJ|2+diam(I)2rφI2L2(I)+diam(J)2rφJ2L2(J)+diam(I)2rφI2Ls(I)+diam(J)2rφJ2Ls(J)).

    It is even simpler to show that

    VeFe(d)|ϕe|2dx+Ve|ϕe|2dxC(|uI|2+|uJ|2+diam(I)2rφI2L2(I)+diam(J)2rφJ2L2(J))

    which concludes the derivation of 19, and the first step.

    Step 2. We now explain how to construct a global extension ϕNH10(NU) with

    ϕN|I=φIIφI+uI,ICC(˜FN). (21)

    For all e=[xI,α,xJ,β]Ed(˜FN), we first introduce a function χeCc(R3), with values in [0,1], satisfying

    χe=1 in a neighborhood of IαJβ

    χe=0 in a neighborhood of R3Ve, where Ve was introduced in 20

    χe=0 in a δ0/2 neighborhood of NU

    ● the supports of χe and χe are disjoint for all eeEd(˜FN)

    |χe|C for some constant C that is uniform in e and N.

    Existence of such functions is easily deduced from our geometric assumptions (G1)-(G2). We now set

    ϕN,1:=eEd(˜FN)χeϕe.

    By our choice of functions χe and by property 18, one has ϕN,1H10(NU), and

    ϕN,1|Iα=φIIφI+uI,ICC(˜FN),α=1,,NI. (22)

    Moreover, by estimate 19,

    ϕN,12H1(NU)Ce=[xI,α,xJ,β]Ed(˜FN)(μe|φI(xI,α)IφIφJ(xJ,β)+JφJ+uIuJ|2+|uI|2+|uJ|2diam(I)2rφI2L2(I)+diam(J)2rφJ2L2(J)+diam(I)2rφI2Ls(I)+diam(J)2rφJ2Ls(J))

    It remains to construct some ϕN,2 satisfying

    ϕN,2|I=ψI,ψI:=φIIφI+uIϕN,1,ICC(˜FN), (23)

    in order for ϕN:=ϕN,1+ϕN,2 to satisfy 21. By 22, ψI is zero on Iα for all ICC(˜FN), for all 1αNI. Thanks to this property and (G1)-(G2), one can find a constant ν>0 independent of I or N such that for all I, there exists ˜ψIH1(R3) satisfying

    ˜ψI|I=ψI,˜ψIH1(R3)CψIH1(I)

    and for all (J,β) with e=[xI,α,xJ,β]Ed(˜FN), for the same local coordinates around the edge e as seen before

    ˜ψI=0on Jβ,ν:=Jβ{|z|ν}.

    Now, for each ICC(˜FN), we introduce ν>0 and a smooth function χI which is 1 in a ν-neighborhood of I and 0 outside a 2ν-neighborhood of I. Thanks to our geometric assumptions, by taking ν small enough (but independent of I and N), we can ensure that for all JI connected by an edge eEd(˜FN),

    Supp(χJ)I1αNIIα,ν.

    We finally set

    ϕN,2=ICC(˜FN)χI˜ψI.

    The keypoint in the definition of ϕN,2 is that for a given I, and for any JI, the term χJ˜ψJ is zero on I: indeed, for all xI, either χJ(x)=0, or χJ(x)0 which implies that xαIα,ν so that ˜ψJ(x)=0. Hence, 23 is satisfied, as expected. Moreover, it is easily seen that

    ϕN,22H1(NU)C(ϕN,12H1(NU)+ICC(˜FN)(diam(I)2rφI2rL2(I)+|I||uI|2))

    so that eventually

    ϕN2H1(NU)Ce=[xI,α,xJ,β]Ed(˜FN)(μe|φI(xI,α)IφIφJ(xJ,β)+JφJ+uIuJ|2+CICC(˜FN|I||uI|2+ICC(˜FN)diam(I)2r(φI2L2(I)+φI2Ls(I)).

    The final step of the proof is to show that for a proper choice of the family (uI)ICC(˜FN), ϕN satisfies the bound 17. This is done using assumption H2. Namely, we denote bIJe:=φI(xI,α)IφI, where e=[xI,α,xJ,β]Ed(˜FN). Remembering Definition 13, one has clearly

    ϕN2H1(NU)CE(˜FN,{uI},{bIJe})+CICC(˜FN)diam(I)2r(φI2L2(I)+φI2Ls(I))CE(˜FN,{uI},{bIJe})+CICC(˜FN)diam(I)2r(|I|˜s2˜s+|I|2(˜ss)s˜s)φI2L˜s(I)CE(˜FN,{uI},{bIJe})+C(ICC(˜FN)diam(I)2r˜s˜s2|I|)˜s2˜sφ2L˜s(˜FN).

    Now, taking N=2NsupxU|x|, one has NUQN and |QN|=C|QN| for a constant C independent of N. Furthermore, we can write FN=˜FNG, where the union is disjoint and Gr(FN) is deduced from Gr(˜FN) by the addition of nodes and edges. Using the property (35) proved in Section 5, we have

    E(˜FN,{uI},{bIJe})E(FN,{ˉuI},{ˉbIJe})

    for any extensions {¯uI},{¯bIJe} of {uI},{bIJe}. We make the choice ˉbIJe=0 if eEd(FN)Ed(˜FN). Now, using property H2 (in the form mentioned in Remark i) after Theorem 2.7), there exists, almost surely, a family {ˉuI}ICC(FN) such that

    E(FN,{ˉuI},{¯bIJe})M|QN|(1|QN|I,JCC(FN)eEd(FN),IeJ|ˉbIJe|s)2/sM|QN|(1|QN|I,JCC(˜FN)eEd(˜FN),IeJ|bIJe|s)2/s.

    Using one last time the Morrey injection yields

    |bIJe|s=|φI(xI,α)φI(xI)|s|xIxI,α|s3˜φIsLs(R3)Cdiam(I)s3+rsφIsLs(I).

    Setting uI=¯uI for ICC(˜FN), we get, back to ϕN:

    ϕN2H1(NU)M|QN|(1|QN|I,JCC(˜FN)diam(I)s3+rsφIsLs(I))2s+C(ICC(˜FN)diam(I)2r˜s˜s2|I|)˜s2˜sφ2L˜s(˜FN)CN3˜s6˜s(1|QN|ICC(˜FN)diam(I)p|I|)˜s2˜sφ2L˜s(˜FN)

    where p=max((s3+rs)˜s˜ss,2r˜s˜s2). As

    1|QN|ICC(˜FN)diam(I)p|I|1|QN|QNdiam(Iy,F)pdy

    we find by the ergodic theorem that

    lim supN+1|QN|ICC(˜FN)diam(I)p|I|Ediam(I0,F)p<+

    which concludes the proof.

    Our proof of homogenization, based on the div-curl lemma, will require proper extensions of solenoidal vector fields, or of fields with given divergence, inside the inclusions. This is the purpose of

    Proposition 4. Assume that H2 holds with F=F and s>3. LetfεL6/5(U), pεL2(UFε) such that divpε=fεinUFε, satisfying the following compatibility conditions:

    Iεpεν=Iεfε,IεCC(Fε).

    Then, there exists a field PεL2(U) satisfyingPε|UFε=pε, divPε=fε in U. Moreover, given any t<s=ss1, there exists p=p(t) such that under the additional hypothesis Ediam(I0,F)p<+, one can choose Pε satisfying the uniform estimate

    PεLt(U)Ct(pεL2(UFε)+fεL6/5(U)).

    Proof. Let t<s. We introduce ˜s such that s<˜s<t. We also introduce the solution w of

    Δw=fε on U,w|U=0.

    It satisfies the estimate

    wLq(U)CqfεW1,q(U)CqfεL3q/(3+q)(U)q(1,2].

    Denoting sε=pεw, it remains to find Sε satisfying Sε|UFε=sε, div Sε=0 in U, and

    SεLt(U)CtsεL2(UFε).

    Then Pε:=Sε+w will meet all requirements. The idea is to search for Sε in the form of a gradient in the inclusions. Strictly speaking, for a fixed realization, we introduce a field vε (depending on ω) defined on Fε that verifies in each inclusion Iε the Neumann problem

    {Δvε=0in ˚Iε,νvε=sενin Iε.

    This Neumann problem is well-posed thanks to the compatibility condition

    Iεsεν=IεpενIενw=Iεfε˚IεΔw=0,IεCC(Fε).

    We then define the random field Sε by Sε=sε in UFε and Sε=vε in Fε. It is divergence-free on U thanks to the continuity of its normal component through each Iε. To establish the uniform estimate on Sε in Lt(U), we proceed by duality. Let ΦLt(U), with t the conjugate of t. It admits the following Helmholtz decomposition in each inclusion:

    Φ|Iε:=P˚IεΦ+φε,IεCC(Fε)

    where for any open set O, PO is the Leray projector, continuous over L˜s(O). More precisely, we claim that

    P˚IεΦL˜s(Iε)+φεL˜s(Iε)C˜sdiam(I)RΦL˜s(Iε), for some R>0.

    Indeed, by scaling, it is enough to show this inequality for ε=1. To show that the operator norm of P˚I (or equivalently IdP˚I) is bounded by a power of diam(I), one writes (IdP˚I)f=uf, where

    Δuf=div f on ˚I,νu|I=fν.

    One must then look carefully at the proof of the inequality ufL˜s(˚I)CIfL˜s(˚I), and track the dependence of CI with respect to I. The derivation of this inequality follows the usual scheme: by local charts and straightening of the boundary, one can use that the inequality holds in R3 and R3+. A tedious verification shows that the constant in the inequality involves the constant d in (G1), the number of charts and the constant in the Poincaré inequality. Under our regularity assumptions, it is controlled by diam(I)R for large enough R. We skip the details for brevity.

    We now introduce the function ϕεH10(U) associated to φε in Proposition 3. In particular, ϕε and φε coincide on each Iε up to a constant function. We compute

    FεvεΦ=IεCC(Fε)IεvεφεIεvεdiv P˚IεΦ+IεvεP˚IεΦν=0=IεCC(Fε)Isενφε using the equation on  vε=IεCC(Fε)Isενϕε using that   Isεν=0=UFεsεϕεsεL2(UFε)ϕεL2(UFε)CsεL2(UFε)φεL˜s(Fε)

    where the last inequality comes from Proposition 3. Now,

    φε˜sL˜s(Fε)=IεCC(Fε)φε˜sL˜s(Iε)CIεCC(Fε)diam(I)R˜sΦ˜sL˜s(Iε)C(IεCC(Fε)|Iε|diam(I)p)t˜stΦ˜sLt(Fε),p=R˜stt˜s.

    Again,

    IεCC(Fε)|Iε|diam(I)pCUdiam(Ix/ε,F)pε0C|U|Ediam(Ip0,F)<+

    so that we end up with

    USεΦ=UFεsεΦ+FεvεΦCsεL2(UFε)(ΦL2(UFε)+ΦLt(Fε))CsεL2(UFε)ΦLt(U).

    This concludes the proof.

    Here, again, F is an admissible set of inclusions.

    The goal of this section is to define properly the matrix A0, describing the effective viscosity of the conducting medium. We follow here the approach developed in [27,chapter 8]. We first introduce the so-called resistance matrix, that is the symmetric matrix defined by:

    ξR3,B0ξξ:=infzEQ1F|ξ+z|2,

    where the infimum is taken over vector fields z=z(y,ω)L2(Ω,L2loc(R3)) that are solenoidal, stationary and mean-free. An equivalent formulation of the variational problem is

    ξR3,B0ξξ:=infZV2sol(Ω)ΩF|ξ+Z|2, (24)

    where:

    F is the subset of Ω defined in 6, so that F(ω)={xR3,τx(ω)F}

    V2sol(Ω)={ZL2(Ω),EZ=0,yZ(τy(ω))solenoidal vector field}.

    We remind that introducing the other subspace of vector fields

    L2pot(Ω)={UL2(Ω),yU(τy(ω))potential vector field}

    one has the orthogonal decomposition L2(Ω)=V2sol(Ω)L2pot(Ω).

    Still following [27,chapter 8], if we now denote

     X : the closure in  L2(ΩF)   of the space   {Z|ΩF,ZV2sol(Ω)} (25)

    then there exists a unique minimizer ZX attaining the infimum, and it satisfies the Euler-Lagrange equation:

    ΩF(ξ+Z)Z=0,ZV2sol(Ω).

    In particular, 1F(ξ+Z)L2pot(Ω), and B0ξ=E1F(ξ+Z).

    The last step of proof of Proposition 1 is showing that the matrix B0 above is invertible. Therefore, we use Lemma 8.7 of [27], which provides a sufficient condition:

    Lemma 4.1. [27,Lemma 8.7] Assume that for any ξR3, and for any ω in a subset of positive measure, there exists a sequence of potential vector fields vεL2(U) satisfying

    vε|Fε(ω)=0,vεξweakly in L^2(U) ,lim supε0vεL2C|ξ| for some C > 0 .

    Then, B0 is positive definite.

    The keypoint is that under H1, the assumptions of the lemma are granted by Proposition 2: one can take vε=ξϕε=(ξxϕε), with ϕε as in Proposition 2. This concludes the proof of Proposition 1.

    We prove in this section part of Theorem 2.7. Namely, we focus on the case where H2 is satisfied with F=F, for some s>3. The reason for treating this special case separately is that it is much easier : indeed, the arguments of [27,chapter 8] rely on the existence of proper extensions of solenoidal vector fields, or of fields with given divergence, inside the inclusions. As such extensions are granted by Proposition 4, they adapt straightforwardly. The proof of the general case, given in the next section, will be more involved (and due to technical difficulty limited to s<6).

    First, by Remark iii) after Theorem 2.7, H2 implies H1. Hence, we can apply Proposition 1, so that A0=(B0)1 is well-defined. Let fL6/5(U), and uεH10(U) the solution of 3 (with implicit dependence on ω), where domain Fε is defined in 9. From a simple energy estimate, uε is bounded in H10(U) uniformly in ω and ε. Hence, almost surely, uε has a subsequence that converges weakly to some u0. The goal is to show that u0 satisfies 2. By uniqueness of this accumulation point, this will mean that the whole sequence converges to u0. From now on, for the sake of brevity, we denote uε the converging subsequence.

    Let now ξR3, and Z the minimizer of problem 24. As ZX, cf. 25, there exists a sequence ZνV2sol(Ω) and ZνZL2(ΩF)0 as ν0. Let ˉZ, resp. ˉZν, the extension of Z, resp. of Zν|ΩF, by ξ on F. We remind that ξ+ˉZL2pot(Ω), with E(ξ+ˉZ)=B0ξ. We finally set

    ˉz(y,ω)=ˉZ(τy(ω)),ˉzν(y,ω)=ˉZν(τy(ω)),zν(y,ω)=Zν(τy(ω)).

    Let pε=uε. By Proposition 4, for any t<s, assuming 14 for large enough p, one can extend pε into a field Pε such that

    Pε|UFε=pε,div Pε=f1UFε in U,PεLt(U)Ct.

    The last bound implies weak convergence of (a subsequence of) Pε towards some P0 in Lt(U). By the ergodic theorem, f1UFε converges weakly to f(1λ), with λ=E1F, in L6/5(U). Hence, div P0=(1λ)f in U. Let now φCc(U), and u0 the weak limit of (a subsequence of) uε in H10(U). The point is to show that, as ε0:

    Uφ(x)uε(x)(ξ+ˉz(x/ε))dxUφ(x)u0(x)ξdx (26)

    as well as

    Uφ(x)uε(x)(ξ+ˉz(x/ε))dxUφ(x)P0B0ξdx. (27)

    Identifying the limits, it follows that B0P0=u0, so that P0=A0u0 and as div P0=(1λf), we recover system 2.

    The proof of 26-27 is an adaptation of the one in [27], so that we indicate only the main elements and the changes that are needed. As regards 26, we write

    Uφ(x)uε(x)(ξ+ˉz(x/ε))dx=Uφ(x)uε(x)(ˉz(x/ε)ˉzν(x/ε))dx+Uφ(x)uε(x)(ˉzν(x/ε)zν(x/ε))dx+Uφ(x)uε(x)(ξ+zν(x/ε))dx. (28)

    The first term at the r.h.s. satisfies

    |Uφ(x)uε(x)(ˉz(x/ε)ˉzν(x/ε))dx|φuεL2(U)ˉz(x/ε)ˉzν(x/ε)L2(U)

    so that, by the uniform L2 bound on uε and the ergodic theorem:

    |lim supεUφ(x)uε(x)(ˉz(x/ε)ˉzν(x/ε))dx|CˉZˉZνL2(Ω)=CZZνL2(ΩF)

    and finally

    lim supνlim supε|Uφ(x)uε(x)(ˉz(x/ε)zν(x/ε))dx|=0.

    For the second term at the r.h.s. of 28, we notice that uε(ˉzν(/ε)zν(/ε)) is zero in Fε, because uε is zero there, and in UεF, because ˉzν=zν there. However, it does not a priori vanish in (εF)UFε. This corresponds to inclusions Iε in εF that intersect Uδ0ε, where

    Uη:={xU,d(x,U)η},η>0.

    A crucial point is that, under the moment condition Ediam(I0,F)3<+, by a direct adaptation of the proof of Lemma 4.2 and 30 below, one has almost surely,

    sup{diam(Iε),IεCC(εF),IεUδ0ε}=o(1) as ε0.

    Hence, for any η>0, for ε small enough, one has (εFU)FεUη, so that

    |Uφ(x)uε(x)(ˉzν(x/ε)zν(x/ε))dx|=|(εFU)Fεφ(x)uε(x)(ξ+zν(x/ε))dx|φLuεL2(Uη)ξ+zν(/ε)L2(Uη)Cξ+zν(/ε)L2(Uη)

    and by the ergodic theorem

    lim supε|Uφ(x)uε(x)(ˉzν(x/ε)zν(x/ε))dx|C1UηL2ξ+ZνL2(Ω)Cη1/2.

    As η is arbitrary, it follows that

    lim supε|Uφ(x)uε(x)(ˉzν(x/ε)zν(x/ε))dx|=0.

    Finally, as regards the third term at the r.h.s. of 28, by the div-curl lemma and the ergodic theorem, for any given ν,

    limεUφ(x)uε(x)(ξ+zν(x/ε))dx=Uφ(x)u0(x)(ξ+EZν)dx=Uφ(x)u0(x)ξdx

    where the last equality comes from the property EZν=0. Combining all previous relations yields 26.

    As regards 27, we want again to rely on the div-curl lemma but switching the potential and solenoidal vector fields. Therefore, we write

    Uφ(x)uε(x)(ξ+ˉz(x/ε))dx=Uφ(x)Pε(x)wεdx

    taking into account that (ξ+ˉz(x/ε)) is a potential vector field, hence can be written wε. Moreover,

    wεE(ξ+ˉZ)=B0ξweakly in  L2(U), almost surely.

    If (Pε)ε>0 was bounded in L2(U), one could conclude directly by the div-curl lemma. As it is only bounded in Lt(U) for t<s, one must use an approximation of wε by the truncation

    wε,l(x)=wε(x)if |wε(x)|l,wε,l(x)=lif wε(x)l,wε,l(x)=lif wε(x)l.

    We refer to [27,chapter 8,page 286] for implementation of this argument.

    We tackle the proof of Theorem 2.7 in the general case where F is an admissible short of F.

    First, we introduce the sequence of admissible shorts (Fκ)κ(0,1], defined by the following properties : for all κ(0,1), F is a short of Fκ and Fκ is a short of F, with

    Ed(Fκ)=Ed(F){eEd(F)Ed(F),|e|κ}.

    In other words, Fκ is deduced from F by removing bridges corresponding to gaps of size larger than κ. Obviously, almost surely, for every closed ball B, FκB=FB for κκB small enough.

    Lemma 4.2. If F satisfies H2 and the moment bound 14 for p=3, then Fκ satisfies H2 for all κ>0.

    Proof. We will first show that,

     almost surely, κQ(0,1],supICC(Fκ),IQNdiam(I)=o(N). (29)

    Indeed, let κQ(0,1]. Clearly, diam(I0,Fκ)diam(I0,F), so Ediam(I0,Fκ)3<+. Let η>0, and consider the event

    AN={ω,there exists  ICC(Fκ(ω)),IQN,IQcN(1+η)}.

    We recall that all inclusions satisfy an inner sphere condition with uniform deterministic radius. Hence, there exists a (deterministic) set of points x1,,xKN of QN+12 with KNCN2 for a deterministic constant C and such that any ICC(Fκ) with IQN, IQcN(1+η) contains at least an xi. It follows that

    P(AN)KNi=1P(diam(Ixi,Fκ)ηN)CN2P(diam(I0,Fκ)ηN).

    The moment bound implies that P(AN)<+, and it follows from Borel-Cantelli Lemma that P(lim supAN)=0. In other words, for all η>0, for ω in a set of full measure, there exists N such that

    supICC(Fκ),IQNdiam(I)ηN.

    By taking a countable subset of η (and as κ describes the countable subset Q(0,1]), one can find a set of full measure independent of κ and η, which proves 29. Let us remark that for κ large enough, namely for κδ with δ the constant in (G1) associated to F, one has Fκ=F, so that 29 implies

     almost surely, supICC(F),IQNdiam(I)=o(N). (30)

    We now turn to the proof of the lemma. Let N1, and {bIJe} a family indexed by I,JFκN, eEd(FκN), IeJ. By 30, for N large enough each ICC(FκN) is included in a connected component of F2N. We define a family {bIJe} indexed by I,JF2N, eEd(F2N), IeJ in the following way:

    ● if I or J does not contain any element of CC(FκN), bIJe=0

    ● if I and J contain elements of CC(FNκ), but eEd(FκN), bIJe=0

    ● if I and J contain elements of CC(FNκ), and eEd(FκN), bIJe=bIJe, where I,J are the unique elements in CC(FκN) such that IeJ.

    We introduce the family {uI} indexed by CC(F2N) such that

    E(F2N,{uI},{bIJe})=inf{tI}E(F2N,{tI},{bIJe}).

    We then define a family {uI} indexed by ICC(FκN), as follows:

    uI=uI for   I the single c.c. of  F2N containing   I.

    With this choice, we have

    I,JCC(F2N)eEd(F2N),IeJ|bIJe|sI,JCC(FκN)eEd(FκN),IeJ|bIJe|s

    and

    ICC(FκN)|I||uI|2CICC(F2N)|I||uI|2

    and

    E(FκN,{uI},{bIJe})CE(F2N,{uI},{bIJe})+ICC(F2N)I,JCC(FκN),I,JI.eEd(FκN),IeJ|bIJebJIe|2μe.

    Now, by definition of Fκ, connected components I,J of Fκ that are included in a single connected component I of F are at distance at least κ, so that μe|lnκ|. Hence, the last term is bounded by

    C|lnκ|I,JCC(FκN)eEd(FκN),IeJ|bIJe|2C|lnκ||QN|(1|QN|I,JCC(FκN)eEd(FκN),IeJ|bIJe|s)2s.

    The result follows easily from assumption H2 for F applied with 2N and previous inequalities.

    We have now all ingredients to perform the proof of our main Theorem 2.7, for a general admissible short F. First, by Remark iii) after Theorem 2.7, A0=(B0)1 is well-defined. As in section 4.2, given fL6/5(U), one has (uε)ε bounded in H10(U), and the goal is to show that any weak accumulation point u0 satisfies 2.

    Similarly, we introduce ξR3, Z the minimizer of problem 24, ˉZ its extension by ξ in F, and ˉz(y,ω)=ˉZ(τy(ω)). Let pε=uε. Let Fκ,ε defined as in 9, replacing F by Fκ and δ0 by δ0/2 :

    Iκ,ε:=εIκIκCC(Fκ),Fκ,ε=Iκ,εU,d(Iκ,ε,U)δ02εIκ,ε.

    We would like to extend the function pε|UFκ,ε into some Pκ,ε satisfying

    div Pκ,ε=f1UFεin U,

    relying on Proposition 4 and the fact that Fκ satisfies H2. Note that for all Iκ,ε in CC(Fκ,ε), one has the compatibility condition

    Iκ,ενuε=IεCC(Fε),IεIκ,εIενuε+(Iκ,εF)νuε=(Iκ,εFε)νuε=Iκ,εFεΔuε=Iκ,εFεf=Iκ,εf1UFε.

    But there is a little technicality here, due to the fact that UFκ,ε is not necessarily included in UFε so that a priori div pεf on UFκ,ε. This is due to the connected components I of F contained in connected components Iκ of Fκ, such that IεFε and Iκ,ε. Still, one can easily replace such connected components by smaller connected closed sets with , , and such that Proposition 4 applies to instead of . Roughly speaking, one has just to erase in the "beads" that intersect . For brevity, we leave to the reader to verify that no complication occurs replacing by , and keep the former notation.

    Eventually, by applying Proposition 4, we obtain a field with

    Moreover, for all , where in the exponent in H2, if the moment condition 14 is satisfied for large enough, one has

    By diagonal extraction, there exists a subsequence in common to all , and in such that ignoring the subsequence in the notation:

    Let . Proceeding exactly as in Section 4.2 for the proof of 26, we find

    (31)

    The novel difficulty lies in the adaptation of the proof of 27. We shall prove that

    (32)

    Comparing 31 and 32, we get

    which shows that converges in the sense of distributions to . But we also have

    so that sending to zero,

    and finally, sending to zero, we get 2.

    It remains to show 31. We take into account that , in , and write:

    The first integral can be treated as in Section 4.2, resulting in

    The second integral is bounded by

    where we have used the uniform bound on in . From the ergodic theorem, we infer that

    The integral at the right-hand side converges to zero as : it follows from the dominated convergence theorem and the pointwise convergence to zero of , because

    Hence,

    (33)

    We still have to control

    where is the center of mass of . We recall that is a potential field that converges weakly in to as . By a proper choice of the additive constant in , we can always assume that converges weakly in to . Now, we write

    Note that for the last equality, we have used that for all , and that is a constant in each . Finally,

    resulting in

    By using the uniform bound on , the uniform bound on and the ergodic theorem, we end up with

    as seen above. It remains to treat

    using the ergodic theorem to bound the factor . The last difficulty is to bound , because we only have so far a control in , for any . Still, under a large moment bound on , cf. 14, we will now show that

    (34)

    Indeed, following the proof of Proposition 4, we see that inside each inclusion , one has , where solves

    so that in particular

    while is the solution, mean-free over , of

    A crucial point is that is a union of bridges, that are at some uniform distance from all other inclusions. Denoting , such bridges, and introducing for all , the neighborhood of , we claim that

    Indeed, by a scaling argument, it is enough to consider the case . The first inequality follows then from a standard elliptic regularity result, while the second one follows from the usual bound

    applied in the domain . Here, we rely on the fact that such domains are far away from other inclusions, so that the constant can be taken uniform in .

    Summing over all 's and all inclusions , we end up with

    As , one has . Hence, for close enough to , , with

    Again, the power of is deduced from a scaling argument, while the factor comes from the Poincaré inequality 16 applied to the mean-free function . We get eventually

    Using Hölder inequality, we find for any such that :

    where . The first factor is bounded thanks to the ergodic theorem and 14, while the second one is bounded thanks to the uniform bound for in . Inequality 34 follows.

    Back to , we deduce that,

    which goes to zero taking close enough to , by the condition . This concludes the proof of the theorem.

    We start here an extended discussion of the assumptions H1-H2. We remind the definition of :

    It follows from this definition that for closed sets where the union is disjoint (so that is deduced from by the addition of nodes and/or edges), one has

    (35)

    for any extensions of , meaning that

    Indeed, the sum in the right-hand side of 35 has more (positive) terms than the one at the left-hand side.

    H1 for a short implies H1. An important property of the discrete energy above concerns closed sets with a short of , cf. Definition 2.5. Given , indexed by , a family associated to , one can associate a family indexed by as follows:

    (36)

    We remind that for any set , is the center of mass of . Note that by definition of a short, any connected component of contains at least one connected component of . We then claim that for finite,

    (37)

    Indeed, introducing , resp. , we write

    where . As when are included in the same inclusion of , the first term can be bounded by

    For the third inequality, we have used the fact that , by our definition of the family , as well as assumption (G2): the number of gaps between two inclusions and is finite, so that

    Moreover, , by our definition of the family ). Eventually,

    which yields 37. We are now ready to show

    Lemma 5.1. Let a short of satisfying

    as well as H1. Then, itself satisfies H1.

    Remark 5. In the case where is an admissible set of inclusions, and is an admissible short of , see Definition 2.5, the first two conditions above are consequences of the relation

    see the proof of 30.

    Proof. One must realize once again that for , and such that , one does not have necessarily . Indeed, it may happen that is contained in , while crosses . Still, by the second assumption of the lemma, , for large enough. It follows that can be seen as the short of a closed set with , and with a disjoint union.

    Let now , indexed by , satisfying

    We associate to the family , indexed by , as in 36 (replacing by and by ). By 37, we have

    Furthermore, by using 35, we have

    so that combining everything we find

    As satisfies the first assumption of the lemma and H1, dividing by and sending to infinity, we see that satisfies H1.

    Clusters with a moment bound on the diameter. We prove here

    Lemma 5.2. Let an admissible set of inclusions satisfying the moment bound

    where is the cluster of containing , cf. Remark 3. Then, satisfies H1.

    Proof. We want to consider the discrete energy for a suitable choice of 's. For each cluster of , and for each , , we set , where as before is the center of mass of . Then,

    Sending to infinity, we end up with

    which shows that satisfies H1 and concludes the proof of the lemma.

    Link with the graph Laplacian. Another interesting result starts with the following observation. Assumption H1 is verified by if almost surely, there exists such that for any , any ,

    Writing the Euler equations of the minimization problem at the left-hand side leads to the following linear system

    where we remind that (and is therefore zero if and are not linked by an edge). This linear system can be written into the matrix form :

    where , and is a symmetric matrix of size defined by

    This kind of matrix arises in the graph literature as the weighted Laplacian matrix for the pondered unoriented graph , see the first section of [3]. It can be seen as a discrete version of a continuous problem of the form

    The energy of this problem is a superadditive quantity over sets and one can expect our discrete minimization problem to verify a similar property. We state

    Lemma 5.3. Let a bounded set of and denote

    The quantity is superadditive over sets, which means that for any decomposition where are pairwise disjoint, one has almost surely

    Proof. It is enough to prove the results for a simple decomposition , with a boundary between and . This leads to the following decomposition of the nodes of the graph

    where is the set of all connected of components of that intersect the boundary without being included in or . The following figure explains the decomposition. In white is the graph and in black is the graph . In dotted lines is what remains from the graph of .

    Let the solution that minimizes , that we split, up to some permutations, into a vector of the form . We compute

    which ends the argument.

    We can then use the superadditive ergodic theorem (cf. [1,14]), which yields a sufficient condition for H1 to hold :

    Proposition 5. Assume that

    then the admissible set verifies assumption H1.

    Logarithmic moment bound.

    Lemma 5.4. Let a random closed set satisfying (G1)-(G2). Let and . If

    (38)

    then verifies H2 with exponent .

    Proof. Let , and a family indexed by in and such as . We want to control the quantity by

    for a suitable choice of . With our logarithmic bound, we may take for all . We find

    Using Hölder inequality with and , we have

    which together with the bound 38 gives the expected result.

    H2 implies H1.

    Lemma 5.5. Let an admissible set of inclusions, an admissible short of , . If

    and if H2 is satisfied by , then H1 is satisfied by .

    Proof. By Lemma 5.1 and Remark 5, it is enough to show that satisfies H1. This can be seen by setting

    Clearly, from assumptions (G1)-(G2), one has

    and . It follows that

    Thanks to the ergodic theorem, we end up with

    Cycle-free graphs. The previous lemma comes from a trivial choice of the family . We will show that if the multigraph of inclusions is cycle-free, there is a better choice, that enables to relax the logarithmic moment bound, and prove Corollary 1.

    Proof of Corollary 1. Let a family indexed by the triplet , , with . As is cycle-free, there is a single edge linking the nodes and , so that we can note instead of for brevity. Given an arbitrary reference inclusion in each cluster of , we then define a family , , as follows. For all such inclusion, there is a unique integer and a unique branch connecting to (with in the case ). We define

    Note that in particular, . Doing this for all cluster , we get a family . We then compute

    Using that , and the bound

    we find

    We get, for any , denoting the conjugate exponent of :

    Thanks to the cycle-free hypothesis, we get that , so much that

    Finally, we notice that , so that

    This concludes the proof.

    The object of this appendix is the following

    Lemma A.1. Let and be two paraboloids be defined in cylindrical coordinates as

    Furthermore we note the gap of width between the two paraboloids defined by

    (see the picture below).Then there exists such that

    where are constants independent of .

    Proof. For , we set

    that we extend by on and on . Clearly, is smooth, with values in . As regards the bound on the gradient, it is easily verified that all derivatives are bounded uniformly in in except for , for which one can compute:

    by direct computation of the integral. The other bounds can be computed similarly. Note that functions of the type of are sometimes referred as Keller functions, cf [28].

    We explain here how to obtain inequality (16) with for inclusions satisfying (G1). A crucial point, beside the regularity of the boundary, is that the inclusions do not shrink, thanks to the interior ball condition with uniform radius .

    Our starting point is the following statement: given an open set Lipschitz-diffeomorphic to the unit ball, with diffeomorphism satisfying

    one has

    where depends only on . This can be seen by a direct adaptation of the proof of the Poincaré-Wirtinger inequality for convex domains given in [29,chapter 3,page 5]: just take in this proof, instead of . It follows easily that for all , with ,

    (39)

    and then

    (40)

    where just depends on (besides ).

    To prove (16), we first introduce a covering , satisfying the following properties:

    i) for each , is an open set Lipschitz-diffeomorphic to the unit ball, with diffeomorphisms s.t. for some independent of (and ).

    ii) The cardinal of the covering is bounded by .

    There are of course many possible choices to satisfy such properties. As regards i), our assumption (G1) ensures that we can cover a vicinity of the boundary by such type of open sets, while the remaining part of can be covered directly by balls. As regards ii), if the covering satisfies i) and is uniformly locally finite, meaning:

    then the cardinal is comparable to , hence bounded by .

    Then, given this covering, we write

    where we used property ii) and the fact that , with the constant in (G1). Now, for any fixed couple , we take a sequence such that

    is a ball for all , whose radius is bounded by some uniform in and .

    , for some uniform in and .

    ● The cardinal is bounded by for some constant uniform in .

    We finally write: for all ,

    where we used (39)-(40) for the second inequality.

    The authors acknowledge the support of the SingFlows project, grant ANR-18-CE40-0027 of the French National Research Agency (ANR). David Gérard-Varet acknowledges the support of the Institut Universitaire de France. Alexandre Girodroux-Lavigne acknowledges the support of the DIM Math Innov de la Région Ile-de-France.



    [1] Ergodic theorems for superadditive processes. J. Reine Angew. Math. (1981) 323: 53-67.
    [2]

    H. Ammari, P. Garapon, H. Kang and H. Lee, Effective viscosity properties of dilute suspensions of arbitrarily shaped particles, Asymptot. Anal., 80 (2012) 189–211.

    [3]

    G. Berkolaiko and P. Kuchment, Introduction to Quantum Graphs, American Mathematical Soc., 2013.

    [4] Network approximation for effective viscosity of concentrated suspensions with complex geometry. SIAM J. Math. Anal. (2005) 36: 1580-1628.
    [5] Fictitious fluid approach and anomalous blow-up of the dissipation rate in a two-dimensional model of concentrated suspensions. Arch. Ration. Mech. Anal. (2009) 193: 585-622.
    [6] Network approximation in the limit of small interparticle distance of the effective properties of a high-contrast random dispersed composite. Arch. Ration. Mech. Anal. (2001) 159: 179-227.
    [7]

    L. Berlyand, A. G Kolpakov and A. Novikov, Introduction to the Network Approximation Method for Materials Modeling, Cambridge University Press, 148, 2013.

    [8] Increase and decrease of the effective conductivity of two phase composites due to polydispersity. J. Stat. Phys. (2005) 118: 481-509.
    [9] Error of the network approximation for densely packed composites with irregular geometry. SIAM Journal on Mathematical Analysis (2002) 34: 385-408.
    [10] Strong and weak blow-up of the viscous dissipation rates for concentrated suspensions. Journal of Fluid Mechanics (2007) 578: 1-34.
    [11]

    B. Blaszczyszyn, Lecture notes on random geometric models. random graphs, point processes and stochastic geometry, 2017.

    [12]

    B. Bollobás, Modern Graph Theory, Springer-Verlag, New York, 1998.

    [13] Network approximation for transport properties of high contrast materials. SIAM J. Appl. Math. (1998) 58: 501-539.
    [14] Nonlinear stochastic homogenization. Ann. Mat. Pura Appl. (1986) 144: 347-389.
    [15]

    M. Duerinckx, Effective viscosity of random suspensions without uniform separation, arXiv: 2008.13188.

    [16] Corrector equations in fluid mechanics: Effective viscosity of colloidal suspensions. Arch. Ration. Mech. Anal. (2021) 239: 1025-1060.
    [17]

    M. Duerinckx and A. Gloria, On Einstein's effective viscosity formula, arXiv: 2008.03837.

    [18]

    M. Duerinckx and A. Gloria, Continuum percolation in stochastic homogenization and the effective viscosity problem, arXiv: 2108.09654.

    [19]

    D. Gérard-Varet, Derivation of the Batchelor-Green formula for random suspensions, Journal de Mathématiques Pures et Appliquées, 152 (2021), 211–250.

    [20] Analysis of the viscosity of dilute suspensions beyond Einstein's formula. Arch. Ration. Mech. Anal. (2020) 238: 1349-1411.
    [21] Mild assumptions for the derivation of Einstein's effective viscosity formula. Communications in Partial Differential Equations (2021) 46: 611-629.
    [22]

    D. Gérard-Varet and A. Mecherbet, On the correction to Einstein's formula for the effective viscosity, arXiv: 2004.05601.

    [23]

    L. Giovanni, A First Course in Sobolev Spaces, American Mathematical Society, Providence, RI, 2017.

    [24] The Effective conductivity of densely packed high contrast composites with inclusions of optimal shape. Continuum Models and Discrete Systems (2004) 158: 63-74.
    [25] A proof of Einstein's effective viscosity for a dilute suspension of spheres. SIAM Journal on Mathematical Analysis (2012) 44: 2120-2145.
    [26] Effective viscosity of a polydispersed suspension. Journal de Mathématiques Pures et Appliquées. Neuvième Série (2020) 138: 413-447.
    [27]

    V. V. Jikov, S. M. Kozlov and O. A. Oleinik, Homogenization of Differential Operators, Springer-Verlag, Berlin Heidelberg, 1994.

    [28] A theorem on the conductivity of a composite medium. Journal of Mathematical Physics (1964) 5: 548-549.
    [29]

    S. Mischler, An introduction to evolution PDEs, (2017).

    [30] A local version of Einstein's formula for the effective viscosity of suspensions. SIAM J. Math. Anal. (2020) 52: 2561-2591.
    [31]

    E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, 1994.

  • This article has been cited by:

    1. Mitia Duerinckx, Antoine Gloria, Continuum Percolation in Stochastic Homogenization and the Effective Viscosity Problem, 2023, 247, 0003-9527, 10.1007/s00205-023-01857-w
    2. Azdine Nait-Ali, Sami Ben Elhaj Salah, Variational modeling adapted to the medium with gradient properties, 2024, 352, 1873-7234, 159, 10.5802/crmeca.254
    3. Thuyen Dang, Yuliya Gorb, Silvia Jiménez Bolaños, Homogenization of High-Contrast Dielectric Elastomer Composites, 2025, 23, 1540-3459, 668, 10.1137/24M1631420
  • Reader Comments
  • © 2022 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(1176) PDF downloads(163) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog