Loading [MathJax]/extensions/TeX/boldsymbol.js

Γ-convergence of quadratic functionals with non uniformly elliptic conductivity matrices

  • Received: 01 June 2021 Revised: 01 July 2021
  • Primary: 35B27, 35B40, 49J45; Secondary: 74Q05

  • We investigate the homogenization through Γ-convergence for the L2(Ω)-weak topology of the conductivity functional with a zero-order term where the matrix-valued conductivity is assumed to be non strongly elliptic. Under proper assumptions, we show that the homogenized matrix A is provided by the classical homogenization formula. We also give algebraic conditions for two and three dimensional 1-periodic rank-one laminates such that the homogenization result holds. For this class of laminates, an explicit expression of A is provided which is a generalization of the classical laminate formula. We construct a two-dimensional counter-example which shows an anomalous asymptotic behaviour of the conductivity functional.

    Citation: Lorenza D'Elia. Γ-convergence of quadratic functionals with non uniformly elliptic conductivity matrices[J]. Networks and Heterogeneous Media, 2022, 17(1): 15-45. doi: 10.3934/nhm.2021022

    Related Papers:

    [1] Lorenza D'Elia . $ \Gamma $-convergence of quadratic functionals with non uniformly elliptic conductivity matrices. Networks and Heterogeneous Media, 2022, 17(1): 15-45. doi: 10.3934/nhm.2021022
    [2] Andrea Braides, Valeria Chiadò Piat . Non convex homogenization problems for singular structures. Networks and Heterogeneous Media, 2008, 3(3): 489-508. doi: 10.3934/nhm.2008.3.489
    [3] 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
    [4] Leonid Berlyand, Volodymyr Rybalko . Homogenized description of multiple Ginzburg-Landau vortices pinned by small holes. Networks and Heterogeneous Media, 2013, 8(1): 115-130. doi: 10.3934/nhm.2013.8.115
    [5] Manuel Friedrich, Bernd Schmidt . On a discrete-to-continuum convergence result for a two dimensional brittle material in the small displacement regime. Networks and Heterogeneous Media, 2015, 10(2): 321-342. doi: 10.3934/nhm.2015.10.321
    [6] Fabio Camilli, Claudio Marchi . On the convergence rate in multiscale homogenization of fully nonlinear elliptic problems. Networks and Heterogeneous Media, 2011, 6(1): 61-75. doi: 10.3934/nhm.2011.6.61
    [7] Brahim Amaziane, Leonid Pankratov, Andrey Piatnitski . Homogenization of variational functionals with nonstandard growth in perforated domains. Networks and Heterogeneous Media, 2010, 5(2): 189-215. doi: 10.3934/nhm.2010.5.189
    [8] Antonio DeSimone, Martin Kružík . Domain patterns and hysteresis in phase-transforming solids: Analysis and numerical simulations of a sharp interface dissipative model via phase-field approximation. Networks and Heterogeneous Media, 2013, 8(2): 481-499. doi: 10.3934/nhm.2013.8.481
    [9] Liselott Flodén, Jens Persson . Homogenization of nonlinear dissipative hyperbolic problems exhibiting arbitrarily many spatial and temporal scales. Networks and Heterogeneous Media, 2016, 11(4): 627-653. doi: 10.3934/nhm.2016012
    [10] Peter Bella, Arianna Giunti . Green's function for elliptic systems: Moment bounds. Networks and Heterogeneous Media, 2018, 13(1): 155-176. doi: 10.3934/nhm.2018007
  • We investigate the homogenization through Γ-convergence for the L2(Ω)-weak topology of the conductivity functional with a zero-order term where the matrix-valued conductivity is assumed to be non strongly elliptic. Under proper assumptions, we show that the homogenized matrix A is provided by the classical homogenization formula. We also give algebraic conditions for two and three dimensional 1-periodic rank-one laminates such that the homogenization result holds. For this class of laminates, an explicit expression of A is provided which is a generalization of the classical laminate formula. We construct a two-dimensional counter-example which shows an anomalous asymptotic behaviour of the conductivity functional.



    In this paper, for a bounded domain Ω of Rd, we study the homogenization through Γ-convergence of the conductivity energy with a zero-order term of the type

    Fε(u):={Ω{A(xε)uu+|u|2}dx,if  uH10(Ω),,if  uL2(Ω)H10(Ω). (1)

    The conductivity A is a Yd-periodic, symmetric and non-negative matrix-valued function in L(Rd)d×d, denoted by Lper(Yd)d×d, which is not strongly elliptic, i.e.

    ess-infyYd(min{A(y)ξξ:ξRd,|ξ|=1})0. (2)

    This condition holds true when the conductivity energy density has missing derivatives. This occurs, for example, when the quadratic form associated to A is given by

    Aξξ:=Aξξforξ=(ξ,ξd)Rd1×R,

    where ALper(Yd)(d1)×(d1) is symmetric and non-negative matrix. It is known (see e.g. [13,Chapters 24 and 25]) that the strongly ellipticity of the matrix A, i.e.

    ess-infyYd(min{A(y)ξξ:ξRd,|ξ|=1})>0, (3)

    combined with the boundedness implies a compactness result of the conductivity functional

    uH10(Ω)ΩA(xε)uudx

    for the L2(Ω)-strong topology. The Γ-limit is given by

    ΩAuudx,

    where the matrix-valued function A is defined by the classical homogenization formula

    Aλλ:=min{YdA(y)(λ+v(y))(λ+v(y))dy:vH1per(Yd)}. (4)

    The Γ-convergence for the Lp(Ω)-strong topology, for p>1, for the class of integral functionals Fε of the form

    Fε(u)=Ωf(xε,Du)dx,foruW1,p(Ω;Rm), (5)

    where f:Ω×Rm×dR is a Borel function, 1-periodic in the first variable satisfying the standard growth conditions of order p, namely c1|M|pf(x,M)c2(|M|p+1) for any xΩ and for any real (m×d)-matrix M, has been widely studied and it is a classical subject (see e.g. [4,Chapter 12] and [13,Chapter 24]). On the contrary, the Γ-convergence of oscillating functionals for the weak topology on bounded sets of Lp(Ω) has been very few analysed. An example of the study of Γ-convergence for the Lp(Ω)-weak topology can be found in the paper [6] where, in the context of double-porosity, the authors compare the Γ-limit for non-linear functionals analogous to 5 computed with respect to different topologies and in particular with respect to Lp(Ω)-weak topology.

    In this paper, we investigate the Γ-convergence for the weak topology on bounded sets (a metrizable topology) of L2(Ω) of the conductivity functional under condition 2. In this case, one has no a priori L2(Ω)-bound on the sequence of gradients, which implies a loss of coerciveness of the investigated energy. To overcome this difficulty, we add a quadratic zeroth-order term of the form u2L2(Ω), so that we immediately obtain the coerciveness in the weak topology of L2(Ω) of Fε, namely, for uH10(Ω),

    Fε(u)Ω|u|2dx.

    This estimate guarantees that Γ-limit for the weak topology on bounded sets of L2(Ω) is characterized by conditions (i) and (ii) of the Definition 1.1 below (see [13,Proposition 8.10]), as well as, thanks to a compactness result (see [13,Corollary 8.12]), Fε Γ-converges for the weak topology of L2(Ω), up to subsequences, to some functional. We will show that, under the following assumptions:

    (H1) any two-scale limit u0(x,y) of a sequence uε of functions in L2(Ω) with bounded energy Fε(uε) does not depend on y (see [1,Theorem 1.2]);

    (H2) the space V defined by

    V:={YdA1/2(y)Φ(y)dy:ΦL2per(Yd;Rd)withdiv(A1/2(y)Φ(y))=0inD(Rd)}

    agrees with the space Rd,

    the Γ-limit is given by

    F0(u):={Ω{Auu+|u|2}dx,if  uH10(Ω),,if  uL2(Ω)H10(Ω), (6)

    where the homogenized matrix A is given through the expected homogenization formula

    Aλλ:=inf{YdA(y)(λ+v(y))(λ+v(y))dy:vH1per(Yd)}. (7)

    We need to make assumption (H1) since for any sequence uε with bounded energy, i.e. supε>0Fε(uε)<, the sequence uε in L2(Ω;Rd) is not bounded due to the lack of ellipticity of the matrix-valued conductivity A(y). Assumption (H2) turns out to be equivalent to the positive definiteness of the homogenized matrix (see Proposition 1).

    In the 2D isotropic elasticity setting of [11], the authors make use of similar conditions as (H1) and (H2) in the proof of the main results (see [11,Theorems 3.3 and 3.4]). They investigate the limit in the sense of Γ-convergence for the L2(Ω)-weak topology of the elasticity functional with a zeroth-order term in the case of two-phase isotropic laminate materials where the phase 1 is very strongly elliptic, while the phase 2 is only strongly elliptic. The strong ellipticity of the effective tensor is preserved through a homogenization process expect in the case when the volume fraction of each phase is 1/2, as first evidenced by Gutiérrez [14]. Indeed, Gutiérrez has provided two and three dimensional examples of 1-periodic rank-one laminates such that the homogenized tensor induced by a homogenization process, labelled 1-convergence, is not strongly elliptic. These examples have been revisited by means of a homogenization process using Γ-convergence in the two-dimensional case of [10] and in the three-dimensional case of [12].

    In the present scalar case, we enlighten assumptions (H1) and (H2) which are the key ingredients to obtain the general Γ-convergence result Theorem 2.1. Using Nguetseng-Allaire [1,16] two-scale convergence, we prove that for any dimension d2, the Γ-limit F0 6 for the weak topology of L2(Ω) actually agrees with the one obtained for the L2(Ω)-strong topology under uniformly ellipticity 3, replacing the minimum in 4 by the infimum in 7. Assumption (H2) implies the coerciveness of the functional F0 showing that its domain is H10(Ω) and that the homogenized matrix A is positive definite. More precisely, the positive definiteness of A turns out to be equivalent to assumption (H2) (see Proposition 1). We also provide two and three dimensional 1-periodic rank-one laminates which satisfy assumptions (H1) and (H2) (see Proposition 2 for the two-dimensional case and Proposition 3 for the three-dimensional case). Thanks to Theorem 2.1, the corresponding homogenized matrix A is positive definite. For this class of laminates, an alternative and independent proof of positive definiteness of A is performed using an explicit expression of A (see Proposition 5). This expression generalizes the classical laminate formula for non-degenerate phases (see [17] and also [2,Lemma 1.3.32], [8]) to the case of two-phase rank-one laminates with degenerate and anisotropic phases.

    The lack of assumption (H1) may induce a degenerate asymptotic behaviour of the functional Fε 1. We provide a two-dimensional rank-one laminate with two degenerate phases for which the functional Fε does Γ-converge for the L2(Ω)-weak topology to a functional F which differs from the one given by 6 (see Proposition 4). In this example, any two-scale limit u0(x,y) of a sequence with bounded energy Fε(uε), depends on the variable y. Moreover, we give two quite different expressions of the Γ-limit F which seem to be original up to the best of our knowledge. The energy density of the first expression is written with Fourier transform of the target function. The second expression appears as a non-local functional due to the presence of a convolution term. However, we do not know if the Γ-limit F is a Dirichlet form in the sense of Beurling-Deny [3], since the Markovian property is not stable by the L2(Ω)-weak topology (see Remark 2).

    The paper is organized as follows. In Section 2, we prove a general Γ-convergence result (see Theorem 2.1) for the functional Fε 1 with any non-uniformly elliptic matrix-valued function A, under assumptions (H1) and (H2). In Section 3 we illustrate the general result of Section 2 by periodic two-phase rank-one laminates with two (possibly) degenerate and anisotropic phases in dimension two and three. We provide algebraic conditions so that assumptions (H1) and (H2) are satisfied (see Propositions 2 and 3). In Section 4 we exhibit a two-dimensional counter-example where assumption (H1) fails, which leads us to a degenerate Γ-limit F involving a convolution term (see Proposition 4). Finally, in the Appendix we give an explicit formula for the homogenized matrix A for any two-phase rank-one laminates with (possibly) degenerate phases. We also provide an alternative proof of the positive definiteness of A using an explicit expression of A for the class of two-phase rank-one laminates introduced in Section 3 (see Proposition 5).

    Notation.

    ● For i=1,,d, ei denotes the i-th vector of the canonical basis in Rd;

    Id denotes the unit matrix of Rd×d;

    H1per(Yd;Rn) (resp. L2per(Yd;Rn), Cper(Yd;Rn)) is the space of those functions in H1loc(Rd;Rn) (resp. L2loc(Rd;Rn), Cloc(Rd;Rn)) that are Yd-periodic;

    ● Throughout, the variable x will refer to running point in a bounded open domain ΩRd, while the variable y will refer to a running point in Yd (or k+Yd, kZd);

    ● We write

    uεu0

    with uεL2(Ω) and u0L2(Ω×Yd) if uε two-scale converges to u0 in the sense of Nguetseng-Allaire (see [1,16])

    F1 and F2 denote the Fourier transform defined on L1(R) and L2(R) respectively. For fL1(R)L2(R), the Fourier transform F1 of f is defined by

    F1(f)(λ):=Re2πiλxf(x)dx.

    Definition 1.1. Let X be a reflexive and separable Banach space endowed with the weak topology σ(X,X), and let Fε:XR be a ε-indexed sequence of functionals. The sequence Fε Γ-converges to the functional F0:XR for the weak topology of X, and we write FεΓ(X)wF0, if for any uX,

    i) uεu, F0(u)lim infε0Fε(uε),

    ii) ¯uεu such that limε0Fε(¯uε)=F0(u).

    Such a sequence ¯uε is called a recovery sequence.

    Recall that the weak topology of L2(Ω) is metrizable on bounded sets, i.e. there exists a metric d on L2(Ω) such that on every norm bounded subset B of L2(Ω) the weak topology coincides with the topology induced on B by the metric d (see e.g. [13,Proposition 8.7]).

    In this section, we will prove the main result of this paper. As previously announced, up to a subsequence, the sequence of functionals , given by 1 with non-uniformly elliptic matrix-valued conductivity , -converges for the weak topology on bounded sets of to some functional. Our aim is to show that -limit is exactly when .

    Theorem 2.1. Let be functionals given by 1 with a -periodic, symmetric, non-negative matrix-valued function in satisfying 2. Assume the following assumptions

    any two-scale limit of a sequence of functions in with bounded energy does not depend on ;

    the space defined by

    (8)

    agrees with the space .

    Then, -converges for the weak topology of to , i.e.

    where is defined by 6 and is given by 7.

    Proof. We split the proof into two steps which are an adaptation of [11,Theorem 3.3] using the sole assumptions (H1) and (H2) in the general setting of conductivity.

    Step 1 - - inequality.

    Consider a sequence converging weakly in to . We want to prove that

    (9)

    If the lower limit is then 9 is trivial. Up to a subsequence, still indexed by , we may assume that is a limit and we can also assume henceforth that, for some ,

    (10)

    As is bounded in , there exists a subsequence, still indexed by , which two-scale converges to a function (see e.g. [1,Theorem 1.2]). In other words,

    (11)

    Assumption (H1) ensures that

    (12)

    where, according to the link between two-scale and weak -convergences (see [1,Proposition 1.6]), is the weak limit of , i.e.

    Since all the components of the matrix are bounded and is non-negative as a quadratic form, in view of 10, for another subsequence (not relabeled), we have

    and also

    (13)

    In particular

    (14)

    Consider such that

    (15)

    or equivalently,

    Take also . Since and in view of 15, an integration by parts yields

    By using [1,Lemma 5.7], is an admissible test function for the two-scale convergence. Then, we can pass to the two-scale limit in the previous expression with the help of the convergences 11 and 13 along with 12, and we obtain

    (16)

    We prove that the target function is in . Setting

    (17)

    and varying in , the equality 16 reads as

    Since the integral in the left-hand side is bounded by a constant times , the right-hand side is a linear and continuous map in . By the Riesz representation theorem, there exists such that, for any ,

    which implies that

    (18)

    In view of assumption (H2), is an arbitrary vector in so that we infer from 18 that

    (19)

    This combined with equality 16 leads us to

    (20)

    By density, the last equality holds if the test functions are replaced by the set of such that

    or equivalently,

    The -orthogonal to that set is the -closure of

    Thus, the equality 20 yields

    for some in the closure of , i.e. there exists a sequence such that

    Due to the lower semi-continuity property of two-scale convergence (see [1,Proposition 1.6]), we get

    Then, by the weak -lower semi-continuity of , we have

    Recalling the definition 7, we immediately conclude that

    provided that .

    It remains to prove that the target function is actually in , giving a complete characterization of -limit. To this end, take a Lebesgue point for and for , the exterior normal to at point . Thanks to 19, we know that , hence, after an integration by parts of the right-hand side of 16, we obtain, for ,

    (21)

    where is given by 17. Varying in , the first two integrals in 21 are equal and bounded by a constant times . It follows that, for any ,

    which leads to -a.e. on . Since is a Lebesgue point, we have

    (22)

    In view of assumption (H2) and the arbitrariness of , we can choose such that so that from 22 we get . Hence,

    This concludes the proof of the - inequality.

    Step 2 - - inequality.

    We use the same arguments of [12,Theorem 2.4] which can easily extend to the conductivity setting. We just give an idea of the proof, which is based on a perturbation argument. For , let be the perturbed matrix of defined by

    where is the unit matrix of . Since the matrix is non-negative, turns out to be positive definite, hence, the functional , defined by 1 with in place of , -converges to the functional given by

    for the strong topology of (see e.g. [13,Corollary 24.5]). Thanks to the compactness result of -convergence (see e.g. [4,Proposition 1.42]), there exists a subsequence such that -converges for the -strong topology to some functional . Let and let be a recovery sequence for which converges to for the -weak topology on bounded sets. Since and since belongs to some bounded set of , from [13,Propositions 6.7 and 8.10] we deduce that

    It follows that converges to as . Then, the -limit of is independent on the subsequence . Repeating the same arguments, any subsequence of has a further subsequence which -converges for the strong topology of to . Thanks to the Urysohn property (see e.g. [4,Proposition 1.44]), the whole sequence -converges to the functional for the strong topology of . On the other hand, in light of the definition 7 of , we get that converges to as , i.e.

    (23)

    Thanks to the Lebesgue dominated convergence theorem and in view of 23, we get that is exactly given by 6. Therefore, -converges to for the -strong topology.

    Now, let us show that -converges to for the weak topology of . Recall that the -weak topology is metrizable on the closed ball of . Fix and let be any metric inducing the -weak topology on the ball centered on and of radius . Let and let be a recovery sequence for for the -strong topology. Since the topology induced by the metric on is weaker than the -strong topology, is also a recovery sequence for for the -weak topology on . Hence,

    which proves the - inequality in . Finally, since any sequence converging weakly in belongs to some ball , as well as its limit, it follows that the - inequality holds true for for -weak topology, which concludes the proof.

    The next proposition provides a characterization of Assumption in terms of homogenized matrix .

    Proposition 1. Assumption is equivalent to the positive definiteness of , or equivalently,

    Proof. Consider . Define

    Recall that if and only if there exists such that (see e.g. [13,Lemma 25.2]). Since is non-negative and symmetric, from 7 it follows that

    Then, there exists a sequence of functions in such that

    which implies that

    (24)

    Now, take such that is a divergence free field in . Recall that, since , we have that , for some . This implies that

    (25)

    where the last equality is obtained by integrating by parts the second integral combined with the fact that is a divergence free field in . In view of convergence 24, the integral on the left-hand side of 25 converges to . Hence, passing to the limit as in 25 yields

    for any such that is a divergence free field in . Therefore which implies that

    Conversely, by 23 we already know that

    where is the homogenized matrix associated with . Since is strongly elliptic, the homogenized matrix is given by

    (26)

    Let be the minimizer of problem 26. Therefore, there exists a constant such that

    which implies that the sequence is bounded in . Then, up to extract a subsequence, we can assume that converges weakly to some in .

    Now, we show that converges strongly to in . Since is a symmetric matrix, there exists an orthogonal matrix-valued function in such that

    where is a diagonal non-negative matrix-valued function in and denotes the transpose of . It follows that , for a.e. . Hence,

    which implies that converges strongly to in .

    Now, passing to the limit as in

    we have

    This along with implies that is a test function for the set given by 8. From 26 it follows that

    Hence, taking into account the strong convergence of in and the weak convergence of in , we have

    which implies that since is a suitable test function for the set . Therefore, for ,

    so that, since is a non-negative matrix, we deduce that . In other words,

    which concludes the proof.

    In this section we provide a geometric setting for which assumptions (H1) and (H2) are fulfilled. We focus on a -periodic rank-one laminates of direction with two phases in , . Specifically, we assume the existence of two anisotropic phases and of given by

    where denotes the volume fraction of the phase . Let and be the associated subsets of , i.e. the open periodic sets

    Let and be unbounded connected components of and in given by

    and we denote by the interface .

    The anisotropic phases are described by two constant, symmetric and non-negative matrices and of which are possibly not positive definite. Hence, the conductivity matrix-valued function , given by

    (27)

    where is the -periodic characteristic function of the phase , is not strongly elliptic, i.e. 2 is satisfied.

    We are interested in two-phase mixtures in with one degenerate phase. We specialize to the case where the non-negative and symmetric matrices and of are such that

    (28)

    for some . The next proposition establishes the algebraic conditions which provide assumptions (H1) and (H2) of Theorem 2.1.

    Proposition 2. Let and be the matrices defined by 28. Assume that and the vectors and are linearly independent in . Then, assumptions and are satisfied. In particular, the homogenized matrix , given by 7, associated to the matrix defined by 27 and 28 is positive definite.

    From Theorem 2.1, we easily deduce that the energy defined by 1 with given by 27 and 28 -converges to the functional given by 6 with conductivity matrix defined by 7. In the present case, the homogenized matrix has an explicit expression given in Proposition 5 in the Appendix.

    Proof. Firstly, let us prove assumption (H1). We adapt the proof of Step 1 of [11,Theorem 3.3] to two-dimensional laminates. In our context, the algebra involved is different due to the scalar setting.

    Denote by the restriction of the two-scale limit in phase or for . In view of 14, for any with compact support in , or due to periodicity in , we deduce that

    so that

    (29)

    Similarly, taking with compact support in , or equivalently in , as test function and repeating the same arguments, we obtain

    (30)

    Due to 29, in phase we have

    where is perpendicular to . Hence, reads as

    (31)

    for some function . On the other hand, since the matrix is positive definite, in phase the relation 30 implies that

    (32)

    for some function . Now, consider a constant vector-valued function defined on such that

    (33)

    Note that condition 33 is necessary for to be an admissible test function for two-scale convergence. In view of 14 and 32, for any , we obtain

    Take now and use the periodized function

    as new test function. Then, we obtain

    (34)

    Recall that , where is such that . This combined with the linear independence of the vectors and implies that the linear map

    is one-to-one. Hence, for any , there exists a unique such that

    (35)

    In view of the arbitrariness of in 35, we can choose such that

    (36)

    Since in the distributional sense and , we deduce that is constant along the direction . Using Fubini's theorem, we may integrate along straight lines parallel to the vector where integration by parts is allowed. Therefore, performing an integration by parts in 34 combined with 36, it follows that for any ,

    where we have set . We conclude that has a trace on for a.e. satisfying

    (37)

    Recall that . Fix . Taking into account 31 and 32, the equality 37 reads as

    Since , it follows that only depends on so that agrees with . Finally, we conclude that is independent of and hence (H1) is satisfied.

    We prove assumption (H2). The proof is a variant of the Step 2 of [11,Theorem 3.4]. For arbitrary , let be a vector-valued function given by

    (38)

    Such a vector field does exist, since is in the range of and thus the right-hand side of 38 belongs pointwise to the range of , or equivalently to the range of . Moreover, the difference of two constant phases in 38 is orthogonal to the laminate direction , so that is a laminate divergence free periodic field in . Its average value is given by

    Hence, due to and the arbitrariness of , the set of the vectors spans , which yields assumption (H2).

    From Proposition 1, it immediately follows that the homogenized matrix is positive definite. For the reader's convenience, the proof of explicit formula of is postponed to Proposition 5 in the Appendix.

    We are going to deal with three-dimensional laminates where both phases are degenerate. We assume that the symmetric and non-negative matrices and of have rank two, hence, there exist such that

    (39)

    The following proposition gives the algebraic conditions so that assumptions required by Theorem 2.1 are satisfied.

    Proposition 3. Let and be the vectors in defined by 39. Assume that the vectors as well as are linearly independent in . Then, assumptions and are satisfied. In particular, the homogenized matrix given by 7 and associated to the conductivity matrix given by 27 and 39 is positive definite.

    Invoking again Theorem 2.1, the energy defined by 1 with given by 27 and 39, -converges for the weak topology of to where the effective conductivity is given by 7. As in two-dimensional laminate materials, has an explicit expression (see Proposition 5 in the Appendix).

    Proof. We first show assumption (H1). The proof is an adaptation of the first step of [11,Theorem 3.3]. Same arguments as in the proof of Proposition 2 show that

    (40)

    In view of 39 and 40, in phase , reads as

    (41)

    for some function and . Now, consider a constant vector-valued function on such that the transmission condition 33 holds. In view of 14, for any , we obtain

    (42)

    Take . Putting the periodized function

    as test function in 42, we get

    (43)

    Since the vectors and are independent in , the linear map

    is surjective. In particular, for any , there exists such that

    (44)

    In view of the arbitrariness of in 44, we can choose such that 36 is satisfied. Due to 40 and 39, we deduce that is constant along the plane perpendicular to , for . This implies that, thanks to Fubini's theorem, we may integrate along the plane where an integration by part may be performed. Hence, an integration by parts in 43 combined with 36, yields for any ,

    which implies that

    (45)

    Fix and recall that . In view of 41, the relation 45 reads as

    (46)

    with for . Due to the independence of in , the linear map defined by

    is a change of variables so that 46 becomes

    This implies that and depend only on and thus and agree with some function . Finally, we conclude that is independent of and hence (H1) is satisfied.

    It remains to prove assumption (H2). To this end, let be the subset of defined by

    (47)

    For , let be the vector-valued function defined by

    (48)

    The existence of such a vector field is guaranteed by the conditions , for , which imply that belongs to the range of and hence the right-hand side of 48 belongs pointwise to the range of , or equivalently to the range of . Moreover, since the difference of the phases and is orthogonal to the laminate direction , is a laminate divergence free periodic field in . Its average value is given by

    Note that is a linear subspace of whose dimension is three. Indeed, let be the linear map defined by

    If we identify the pair with the vector , with , for , the associated matrix of is given by

    with , . In view of the linear independence of , the rank of is three, which implies that the dimension of kernel is also three. Since the kernel agrees with , we conclude that the dimension of is three.

    Now, let be the linear map defined by

    Let us show that is invertible. To this end, consider . From the definition of the map , consists of all vectors of the form

    (49)

    In view of the definition of given by 47, the vector 49 satisfies the conditions

    This combined with the linear independence of implies that

    Hence, which implies along with the fact that the dimension of is three that is invertible. This proves that all the vectors of can be attained through the map so that assumption (H2) is satisfied.

    Thanks to Proposition 1, the homogenized matrix turns out to be positive definite. The proof of the explicit expression of is given in Proposition 5 in the Appendix.

    In this section we are going to construct a counter-example of two-dimensional laminates with two degenerate phases, where the lack of assumption (H1) provides an anomalous asymptotic behaviour of the functional 1.

    Let and let be the laminate direction. We assume that the non-negative and symmetric matrices and of are given by

    for some positive constant . The presence of is essential to have oscillation in the conductivity matrix . In the present case, the matrix-valued conductivity is given by

    (50)

    with

    (51)

    Thus, the energy , defined by 1 with given by 50 and 51 becomes

    (52)

    We denote by the convolution with respect to the variable , i.e. for and

    Throughout this section, denotes the positive constant given by

    (53)

    where is the volume fraction of the phase in . The following result proves the -convergence of for the weak topology of and provides two alternative expressions of the -limit, one of that seems nonlocal due to presence of convolution term (see Remark 2 below).

    Proposition 4. Let be the functional defined by 52. Then, -converges for the weak topology of to the functional defined by

    where denotes the Fourier transform on of parameter with respect to the variable of the function extended by zero outside and

    (54)

    The -limit can be also expressed as

    (55)

    where is given by 53 and is a real-valued function in defined by means of its Fourier transform on

    (56)

    where and are given by

    (57)

    Moreover, any two-scale limit of a sequence with bounded energy depends on the variable .

    Remark 1. From 57, we can deduce that

    for any , so that the Fourier transform of is well-defined.

    Proof. We divide the proof into three steps.

    Step 1 - - inequality.

    Consider a sequence converging weakly in to . Our aim is to prove that

    (58)

    If the lower limit is then 58 is trivial. Up to a subsequence, still indexed by , we may assume that is a limit and we may assume henceforth that, for some ,

    (59)

    It follows that the sequence is bounded in and according to [1,Theorem 1.2], a subsequence, still indexed by , of that sequence two-scale converges to some . In other words,

    (60)

    In view of 51, we know that so that, thanks to 59, for another subsequence (not relabeled) we have

    (61)

    In particular,

    (62)

    Take . By integration by parts, we obtain

    Passing to the limit in both terms with the help of 60 and 62 leads to

    which implies that

    (63)

    Due to the link between two-scale and weak -convergences (see [1,Proposition 1.6]), we have

    (64)

    Now consider such that

    (65)

    Since , an integration by parts leads us to

    In view of the convergences 60 and 61 together with 63, we can pass to the two-scale limit in the previous expression and we obtain

    (66)

    Varying , the left-hand side of 66 is bounded by a constant times so that the right-hand side is a linear and continuous form in . By the Riesz representation theorem, there exists such that, for any ,

    which yields

    (67)

    Then, an integration by parts with respect to of the right-hand side of 66 yields, for any satisfying 65,

    Since for any the first two integrals are equal and bounded by a constant times , we conclude that, for any satisfying 65,

    which implies that

    This combined with 67 yields

    Finally, an integration by parts with respect to of the right-hand side of 66 implies that, for any satisfying 65,

    Since the orthogonal of divergence-free functions is the gradients, from the previous equality we deduce that there exists such that

    (68)

    Now, we show that

    (69)

    To this end, set

    Since , there exists a sequence of functions in such that

    (70)

    hence, by periodicity, we also have

    (71)

    for some positive constant . On the other hand, since given by 68 is in , there exists a sequence of functions in such that

    (72)

    From the inequality

    we get

    (73)

    In view of 71, the first integral on the right-hand side of 73 can be estimated as

    Hence, passing to the limit as in 73 with the help of 61 leads to

    Thanks to 70, we take the limit as in the previous inequality and we obtain

    so that in view of 72, passing to the limit as leads to

    This combined with 68 proves 69.

    By 63, we already know that does not depend on . In view of the periodicity of with respect to , an application of Jensen's inequality leads us to

    This combined with 69 implies that

    (74)

    Now, we extend the functions in by zero with respect to outside so that functions in can be regarded as functions in . Due to the weak -lower semi-continuity of along with 74, we have

    (75)

    We minimize the right-hand side with respect to satisfying 64 where the weak limit of in is fixed. The minimizer, still denoted by , satisfies the Euler equation

    for any such that . Then, there exists independent of such that in distributions sense with respect to the variable ,

    (76)

    for a.e. . Taking the Fourier transform on of parameter with respect to the variables , the equation 76 becomes

    (77)

    Note that 77 proves in particular that the two-scale limit does depend on the variable , since its Fourier transform with respect to the variable depends on through the function .

    In light of the definition 54 of and due to 64, integrating 77 with respect to yields

    (78)

    By using Plancherel's identity with respect to the variable in the right-hand side of 75 and in view of 77 and 78, we obtain

    which proves the - inequality.

    Step 2- - inequality.

    For the proof of the - inequality, we need the following lemma whose proof will be given later.

    Lemma 4.1. Let . For fixed and , let be the distribution (parameterized by ) defined by

    (79)

    where is extended by zero outside . Let be the unique solution to problem

    (80)

    with given by 51. Then is in and is in .

    Let . Thanks to Lemma 4.1, there exists a unique solution

    (81)

    to the problem 80. Taking the Fourier transform on of parameter with respect to of the equation in 80 and taking into account 79, we get

    (82)

    where and are extended by zero outside . Integrating 82 over , we obtain

    (83)

    Let be the sequence in defined by

    Recall that rapidly oscillating -periodic function weakly converges in to the mean value of over . This combined with 83 implies that weakly converges in to . In other words,

    Due to 81, we can apply [1,Lemma 5.5] so that and are admissible test functions for the two-scale convergence. Hence,

    (84)

    where the function is extended by zero outside . In view of the definition 54 of and due to 82, the Plancherel identity with respect to the variable and the Fubini theorem yield

    This together with 84 implies that, for ,

    which proves the - inequality on .

    Now, we extend the previous result to any . To this end, we use a density argument (see e.g. [5,Remark 2.8]). Recall that the weak topology of is metrizable on the closed balls of . Fix and denote any metric inducing the -weak topology on the ball centered on and of radius . Then, can be regarded as a subspace of endowed with the metric . On the other hand, is a Hilbert space endowed with the norm

    The associated metric on induces a topology which is not weaker than that induced by , i.e.

    (85)

    Recall that is a dense subspace of for the metric and that the - inequality holds on for the -weak topology, i.e. for any ,

    (86)

    A direct computation of , given by 54, shows that

    which implies that

    (87)

    where and are given by 57. Hence, there exists a positive constant such that

    (88)

    This combined with the Plancherel identity yields

    (89)

    where is extended by zero outside . Since is a non-negative quadratic form, from 89 we conclude that is continuous with respect to the metric .

    Now, take . By density, there exists a sequence in such that

    (90)

    In particular, due to 85, we also have that as . In view of the weakly lower semi-continuity of - and the continuity of , we deduce from 86 that

    which proves the - inequality in . Since for any the sequence of functions in satisfying 90 belongs to some ball of , as well as its limit, the - property holds true for the sequence on , which concludes the proof of - inequality.

    Step 3 - Alternative expression of -limit.

    The proof of the equality between the two expressions of the -limit relies on the following lemma whose proof will be given later.

    Lemma 4.2. Let and . Then, and

    (91)

    By applying Plancherel's identity with respect to , for any extended by zero with respect to the variable outside , we get

    (92)

    Recall that the Fourier transform of , given by 56, is real. From 92, an application of Lemma 4.2 leads us to

    (93)

    On the other hand, by applying Plancherel's identity with respect to , we obtain

    In view of the expansion of given by 87, the previous equality combined with 92 and 93 implies that, for extended by zero with respect to outside ,

    which concludes the proof.

    Proof of Lemma 4.1. In view of 87, the equality 79 becomes

    (94)

    Since

    the right-hand side of 94 belongs to with respect to , which implies that

    Applying the Plancherel identity, we obtain that with respect to . Since is extended by zero outside , is also equal to zero outside so that

    (95)

    We show that is a continuous function with respect to . Recall that the continuity of is equivalent to

    Thanks to Plancherel's identity, we infer from 79 that

    In view of 88 and thanks to the Plancherel identity, we obtain

    By the Lebesgue dominated convergence theorem and since , from the previous inequality we conclude that the map is continuous. Hence,

    (96)

    To conclude the proof, it remains to show the regularity of . Note that 80 is a Sturm-Liouville problem with constant coefficient with respect to , since and play the role of parameters. By 95, we already know that , so that thanks to a classical regularity result (see e.g. [7] pp. 223-224), the problem 80 admits a unique solution in . Since is embedding into , we have

    On the other hand, the solution to the Sturm-Liouville problem 80 is explicitly given by

    (97)

    where is defined by 79 and 96 and the kernel is given by

    This combined with 96 and 97 proves that

    which concludes the proof.

    We prove now the Lemma 4.2 that we used in Step of Proposition 4.

    Proof of Lemma 4.2. By the convolution property of the Fourier transform on , we have

    (98)

    where denotes the conjugate Fourier transform for . On the other hand, since and due to Riemann-Lebesgue's lemma, we deduce that . This combined with implies that

    Since on , from 98 we deduce that

    which yields 91. This concludes the proof.

    Remark 2. Thanks to the Beurling-Deny theory of Dirichlet forms [3], Mosco [15, Theorem 4.1.2] has proved that the -limit of a family of Markovian form for the -strong topology is a Dirichlet form which can be split into a sum of three forms: a strongly local form , a local form and nonlocal one. More precisely, for with , we have

    (99)

    where is called the diffusion part of , is a positive Radon measure on , called the killing measure, and is a positive Radon measure on , called the jumping measure. Recall that a Dirichlet form is a closed form which satisfies the Markovian property, i.e. for any contraction , such that

    we have . A -limit form obtained with the -weak topology does not a priori satisfy the Markovian property, since the -weak convergence does not commute with all contractions . An example of a sequence of Markovian forms whose -limit for the -weak topology does not satisfy the Markovian property is provided in [9, Theorem 3.1]. Hence, the representation formula 99 does not hold in general when the -strong topology is replaced by the -weak topology.

    In the present context, we do not know if the -limit 55 is a Dirichlet form since the presence of the convolution term makes difficult to prove the Markovian property.

    We are going to give an explicit expression of the homogenized matrix defined by 7, which extends the rank-one laminate formula in the case of a rank-one laminates with degenerate phases. We will recover directly from this expression the positive definiteness of for the class of rank-one laminates introduced in Section 3. Indeed, by virtue of Theorem 2.1 the positive definiteness of also follows from assumption (H2) which is established in Proposition 2 and Proposition 3.

    Set

    (100)

    with being the volume fraction of phase .

    Proposition 5. Let and be two symmetric and non-negative matrices of , . If given by 100 is positive, the homogenized matrix is given by

    (101)

    If , the homogenized matrix is the arithmetic average of the matrices and , i.e.

    (102)

    Furthermore, if one of the following conditions is satisfied:

    i) in two dimensions, and the matrices and are given by 28 with ,

    ii) in three dimensions, , the matrices and are given by 39 and the vectors are independent in ,

    then is positive definite.

    Remark 3. The condition agrees with the -convergence results of Propositions 2 and 3. In the two-dimensional framework, the degenerate case does not agree with Propositions 2. Indeed, implies that in contradiction to positive definiteness of . Similar in the three-dimensional setting, where the independence of is not compatible with . Indeed, implies that , for , which contradicts the fact that and have rank two.

    Proof. Assume that . In view of the convergence 23, we already know that

    (103)

    where, for , is the homogenized matrix associated to conductivity matrix given by

    with . Since and are non-negative matrices, is positive definite and thus the homogenized matrix is given by the lamination formula (see [17] and also [2,Lemma 1.3.32])

    (104)

    If , we easily infer from the convergence 103 combined with the lamination formula 104 the expression 101 for .

    We prove that for any . From the Cauchy-Schwarz inequality, we deduce that

    (105)

    This combined with the definition 101 of implies that, for any ,

    (106)

    In view of definition 100 of , we have that

    Plugging this equality in 106, we deduce that

    (107)

    which proves that is a non-negative definite matrix.

    Now, assume . Since and are non-negative matrices, the condition implies or equivalently . Hence,

    which implies that the lamination formula 104 becomes

    This combined with the convergence 103 yields to the expression 102 for .

    To conclude the proof, it remains to prove the positive definiteness of under the above conditions i) and ii).

    Case (i). , and given by 28.

    Assume . Then, the inequality 107 is an equality, which yields in turn equalities in 105. In particular, we have

    (108)

    Recall that the Cauchy-Schwarz inequality is an equality if and only if one of vectors is a scalar multiple of the other. This combined with 108 leads to for some , so that, since is positive definite or equivalently , we have

    (109)

    From the definition 101 of and due to the assumption , we get

    (110)

    Recall that . This combined with 109 and 110 implies that , which proves that is positive definite.

    Case (ii). , and given by 39.

    Assume that . As in Case (i), we have equalities in 105. In other words,

    (111)
    (112)

    Let be the non-negative polynomials of degree defined by

    In view of 111, the discriminant of is zero, so that there exists such that

    (113)

    Recall that . Since is non-negative matrix, we deduce from 113 that belongs to , so that

    (114)

    Similarly, recalling that and using 112, we have

    (115)

    Since the vectors are independent in , 114 and 115 imply that

    In light of definition 101 of , we have

    which implies that , since . This establishes that is positive definite and concludes the proof.

    Note that when and the assumption is essential to obtain that is positive definite. Otherwise, the homogenized matrix is just non-negative definite as shown by the following counter-example. Let and be symmetric and non-negative matrices of defined by

    Then, it is easy to check that and .

    This problem was pointed out to me by Marc Briane during my stay at Institut National des Sciences Appliquées de Rennes. I thank him for the countless fruitful discussions. My thanks also extend to Valeria Chiadò Piat for useful remarks. The author is also a member of the INdAM-GNAMPA project "Analisi variazionale di modelli non-locali nelle scienze applicate".



    [1] Homogenization and two-scale convergence. SIAM J. Math. Anal. (1992) 23: 1482-1518.
    [2]

    G. Allaire, Shape Optimization by the Homogenization Method, Springer-Verlag, New York, 2002.

    [3] Espaces de dirichlet. Acta Math. (1958) 99: 203-224.
    [4] (2002) -Convergence for Beginners. Oxford: Oxford University Press.
    [5]

    A. Braides, A handbook of -convergence, Handbook of Differential Equations: Stationary Partial Differential Equations Vol. 3, Elsevier, (2006), 101–213.

    [6] A variational approach to double-porosity problems. Asymptot. Anal. (2004) 39: 281-308.
    [7]

    H. Brezis, Functional Analysis, Sobolev Spaces and Partial Differential Equations, Universitext series, Springer, New York, 2010.

    [8] Correctors for the homogenization of a laminate. Adv. Math. Sci. Appl. (1994) 4: 357-379.
    [9] Non-Markovian quadratic forms obtained by homogenization. Boll. Uni. Mate. Ital. Sez. B Artic. Ric. Mat. (2003) 6: 323-337.
    [10] Loss of ellipticity through homogenization in linear elasticity. Math. Mod. Met. Appl. Sci. (2015) 25: 905-928.
    [11] A two-dimensional labile aether through homogenization. Commun. Math. Phys. (2019) 367: 599-628.
    [12]

    M. Briane and A. J. Pallares Martín, Homogenization of weakly coercive integral functionals in three-dimensional linear elasticity, J. Éc. Polytech. Math., 4 (2017), 483–514.

    [13]

    G. Dal Maso, An Introduction to -Convergence, Volume 8 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser, Boston, 1993.

    [14] Laminations in linearized elasticity: The isotropic non-very strongly elliptic case. Q. J. Mech. Appl. Math (2004) 57: 571-582.
    [15] Composite media and asymptotic Dirichlet forms. J. Funct. Anal. (1994) 123: 368-421.
    [16] A general convergence result for a functional related to the theory of homogenization. SIAM J. Math. Anal. (1989) 20: 608-623.
    [17] Estimations fines de coefficients homogénéisés. Ennio De Giorgi Colloquium, Ed. P. Krée, Pitman Research Notes in Mathematics (1985) 125: 168-187.
  • 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(1664) PDF downloads(301) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog