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

On the way to allergy prevention—future perspective or illusory aim?

  • Received: 22 April 2017 Accepted: 24 April 2017 Published: 24 April 2017
  • Citation: Claudia Traidl-Hoffmann. On the way to allergy prevention—future perspective or illusory aim?[J]. AIMS Allergy and Immunology, 2017, 1(1): 15-20. doi: 10.3934/Allergy.2017.1.15

    Related Papers:

    [1] Vivek Tewary . Combined effects of homogenization and singular perturbations: A bloch wave approach. Networks and Heterogeneous Media, 2021, 16(3): 427-458. doi: 10.3934/nhm.2021012
    [2] Grégoire Allaire, Tuhin Ghosh, Muthusamy Vanninathan . Homogenization of stokes system using bloch waves. Networks and Heterogeneous Media, 2017, 12(4): 525-550. doi: 10.3934/nhm.2017022
    [3] Carlos Conca, Luis Friz, Jaime H. Ortega . Direct integral decomposition for periodic function spaces and application to Bloch waves. Networks and Heterogeneous Media, 2008, 3(3): 555-566. doi: 10.3934/nhm.2008.3.555
    [4] Laura Caravenna, Laura V. Spinolo . New interaction estimates for the Baiti-Jenssen system. Networks and Heterogeneous Media, 2016, 11(2): 263-280. doi: 10.3934/nhm.2016.11.263
    [5] Riccardo Bonetto, Hildeberto Jardón Kojakhmetov . Nonlinear diffusion on networks: Perturbations and consensus dynamics. Networks and Heterogeneous Media, 2024, 19(3): 1344-1380. doi: 10.3934/nhm.2024058
    [6] Rémi Goudey . A periodic homogenization problem with defects rare at infinity. Networks and Heterogeneous Media, 2022, 17(4): 547-592. doi: 10.3934/nhm.2022014
    [7] 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
    [8] Hiroshi Matano, Ken-Ichi Nakamura, Bendong Lou . Periodic traveling waves in a two-dimensional cylinder with saw-toothed boundary and their homogenization limit. Networks and Heterogeneous Media, 2006, 1(4): 537-568. doi: 10.3934/nhm.2006.1.537
    [9] Hakima Bessaih, Yalchin Efendiev, Florin Maris . Homogenization of the evolution Stokes equation in a perforated domain with a stochastic Fourier boundary condition. Networks and Heterogeneous Media, 2015, 10(2): 343-367. doi: 10.3934/nhm.2015.10.343
    [10] Antonin Chambolle, Gilles Thouroude . Homogenization of interfacial energies and construction of plane-like minimizers in periodic media through a cell problem. Networks and Heterogeneous Media, 2009, 4(1): 127-152. doi: 10.3934/nhm.2009.4.127


  • The formation of shear bands in elasticity is described by a degenerate operator of elliptic-hyperbolic type [1, 23, 4]. The shear bands that are mathematically obtained in this model are infinitesimally thin. To overcome this non-physical description, it is customary to penalize the elasticity operator by a fourth-order singular perturbation [18]. Subsequently, it was suggested that the penalization should be followed by a homogenization procedure, which results in different regimes depending on the order of penalization as compared to the length-scale of the periodic heterogeneities. This was first carried out in [6, 18]. Similar problems in the framework of Γ-convergence have been studied in [32, 8, 39]. A quantitative analysis of this problem appeared in [27, 26, 29]. The aim of the present work is to revisit this problem by employing the homogenization framework developed by Conca and Vanninathan [13]. Bloch wave method in homogenization is a spectral method of homogenization. It also offers an alternative approximation scheme in the form of Bloch approximation [11, 12] which provides sharp convergence estimates in homogenization under minimal regularity requirements. In this paper, we shall establish new characterizations of homogenized tensor for the singularly perturbed problem in terms of the first Bloch eigenvalue and use Bloch wave method to recover the homogenization result for singularly perturbed periodic homogenization problem.

    We will study the simultaneous homogenization and singular perturbation limits of the following operator

    Aκ,ε:=κ2Δ2A(xε), (1.1)

    where 0<κ,ε1, and A(y)=(ajk)dj,k=1 is a matrix whose entries are real, bounded, measurable functions of yRd. Further, the matrix A satisfies the following hypotheses:

    A1. A is elliptic, i.e., there exists α>0 such that for all ξRd and a.e. yRd, A(y)ξξα|ξ|2.

    A2. A is Y-periodic, i.e., A(y+2πp)=A(y) for all pZd, a.e. yRd. The set Y:=[0,2π)d is called a basic periodicity cell and may also be interpreted as a parametrization of the d-dimensional torus Td.

    A3. The matrix A is symmetric.

    We mention briefly the function spaces that make an appearance in this problem. The solutions of the cell problem associated with homogenization of (1.1), as well as Bloch eigenfunctions, are sought in the space H2(Y), which is the space of all periodic distributions u for which the norm ||u||H2=(nZd(1+|n|2)2|ˆu(n)|2)12 is finite. The space H2(Y) may be identified with H2(Td). The spaces Hs(Y) for sR are similarly defined. For s>0, Hs(Y) forms a subspace of the space of all periodic L2 functions in Rd, denoted by L2(Y) or L2(Td). We shall denote the mean value or average of a periodic function u on the basic periodicity cell Y by MY(u):=1|Y|Yu(y)dy. Averaged integrals such as 1|Y|Yu(y)dy are sometimes denoted as .

    The method of Bloch waves rests on decomposition of a periodic operator in terms of Bloch waves which may be thought of as a periodic analogue of plane waves. As plane waves decompose a linear operator with constant coefficients by means of the Fourier transform, Bloch waves diagonalize a linear operator with periodic coefficients. This decomposition begins with a direct integral decomposition of a periodic operator A in Rd.

    ATdA(η)dη.

    The fiber operator A(η) has compact resolvent for each fixed ηTd. The eigenfunctions viewed as functions of η are called Bloch waves. Finally, the operator A(η) is diagonalized by means of Bloch waves.

    The homogenization limits for a highly oscillating scalar periodic operator are obtained from its first Bloch mode. The rest of the Bloch modes do not contribute to the homogenization limit. This is a consequence of the separation of the first Bloch eigenvalue from the rest of the spectrum. Such an interpretation of homogenization is also called spectral threshold effect [7]. Moreover, the homogenized tensor is obtained from Hessian of the first Bloch eigenvalue. Therefore, the second-order nature of the differential operator is reflected in the quadratic nature of the first Bloch eigenvalue near the bottom of the spectrum. Indeed, homogenization of higher-even-order periodic operators can also be obtained by the Bloch wave method, where the first Bloch eigenvalue behaves like a polynomial of the corresponding order near the bottom of the spectrum [37, 36].

    The homogenization result in Theorem 8.1 exhibits three different regimes depending on the ratio of κ and ε, where κ is to be interpreted as a function of ε satisfying limε0κ=0. Define ρ:=κε. The three different regimes correspond to

    limε0ρ=0,

    0<limε0ρ<, and

    limε0ρ=.

    In the first regime, the subcritical case, that is, limε0ρ=0, homogenization dominates over the singular perturbation. Hence, it may be thought of as taking the homogenization limit followed by the singular perturbation limit. In the critical regime, that is, 0<limε0ρ<, we can see the combined effects of singular perturbation and homogenization. The cell functions are solutions to a fourth order operator. On the other hand, in the supercritical regime, that is, limε0ρ=, singular perturbation dominates the homogenization limit and homogenized coefficients are given by the average of the periodic coefficients.

    The approach is to treat ρ as a fixed number and obtain Bloch wave decomposition for the unscaled operator Aρ=ρ2Δ2A(y). This decomposition is employed to obtain the homogenization limit as ε,κ0. While the qualitative homogenization result is not new, the presentation is original. Some novel features of the proof are characterization of homogenized tensor and cell functions in terms of derivatives of the first Bloch eigenvalue and eigenfunctions. This requires us to prove analyticity of the first Bloch eigenvalue and eigenfunction in a neighbourhood of zero in the dual parameter. Interestingly, the region of analyticity does not depend on the singular perturbation, which plays an important role in the simultaneous passage to 0 of κ and ε. In order to study the stability of the homogenized tensor in the different regimes, we obtain uniform in ρ estimates for Bloch eigenvalues, eigenfunctions and their derivatives of all orders in the dual parameter. We also prove that only the first Bloch mode contributes to homogenization and the higher modes are negligible. The analysis of three separate regimes provides considerable challenges in the Bloch wave method, particularly in obtaining uniform estimates in these regimes.

    While the motivation for the problem (1.1) comes from the theory of elasticity, it is for the sake of simplicity that we only study the scalar operator. However, it must be noted that Bloch wave homogenization of systems carries some unique difficulties, such as the presence of multiplicity at the bottom of the spectrum. Indeed, these challenges have been surmounted by the use of directional analyticity of Bloch eigenvalues in [35, 7, 3]. Further, the assumption of symmetry, while customary in elasticity, is made for a simplified presentation. A Bloch wave analysis of homogenization of non-selfadjoint operators may be found in [19].

    In a forthcoming work, we will obtain quantitative estimates for the combined effects of singular perturbation and homogenization through the notion of Bloch approximation, which was introduced in [11]. Higher order estimates in homogenization have been obtained by these methods, particularly for the dispersive wave equation [14, 15, 2, 22].

    The plan of the paper is as follows: In Section 2, we obtain Bloch waves for the singularly perturbed operator Aρ. In Section 3, we prove that the first Bloch eigenpair are analytic functions of the dual parameter in a neighbourhood of 0. In Section 4, we prove that the neighbourhood of analyticity is independent of ρ. In Section 5, we recall the cell problem for the operator Aρ and the estimates associated with it. In Section 6, we characterize the homogenized tensor in the three regimes by way of Bloch method. In Section 7, we shall define the first Bloch transform and analyze its asymptotic properties. In Section 8, we obtain the qualitative homogenization theorem by means of the first Bloch transform, which is the periodic analogue of Fourier transform. Finally, in Section 9, we quantify the contribution of the higher Bloch transforms towards the homogenization limit.

    In this section, we will prove the existence of Bloch waves for the singular operator given by

    Aρ:=ρ2Δ2A(y). (2.1)

    Recall that ρ was earlier set as κε, however in this section, ρ will be assumed to be a fixed positive number. Bloch waves for (2.1) refers to eigenfunctions of (2.1) satisfying the so-called (ηY)-periodicity condition, that is, we look for functions ψ satisfying the following eigenvalue problem:

    {ρ2Δ2ψA(y)ψ=λψψ(y+2πp)=e2πipηψ(y),pZd,ηRd.

    The above problem is invariant under Zd-shifts of η, hence it suffices to restrict η to Y:=[12,12)d. Now, if we set ψ(y)=eiyηϕ(y) where ϕ is a Y-periodic function, then the above eigenvalue problem is transformed into:

    {Aρ(η)ϕ:=ρ2(+iη)4ϕ(+iη)A(y)(+iη)ϕ=λϕϕ(y+2πp)=ϕ(y),pZd,ηY. (2.2)

    The operator Aρ(η) is often called the shifted operator associated with Aρ where the shift iη appears as a magnetic potential. In order to prove the existence of eigenvalues for (2.2), we shall begin by proving that a zeroth-order perturbation of Aρ(η) is elliptic on H2(Y). This amounts to a Gå rding type inequality for the operator Aρ(η). Combined with Rellich compactness theorem, this will allow us to prove compactness of the inverse in L2(Y). Then, a standard application of the spectral theorem for compact self-adjoint operators will guarantee the existence of eigenvalues for each fixed ρ and η.

    The bilinear form aρ[η](,) defined on H2(Y)×H2(Y) by

    aρ[η](u,v):=YA(+iη)u¯(+iη)vdy+ρ2Y(+iη)2u¯(+iη)2vdy, (2.3)

    is associated to the operator Aρ(η). We shall prove the following Gå rding-type inequality for aρ[η].

    Lemma 2.1. There exists a positive real number C not depending on η but depending on ρ such that for all uH2(Y) and all ηY, we have

    aρ[η](u,u)+C||u||2L2(Y)ρ26||Δu||2L2(Y)+α2||u||H1(Q). (2.4)

    Proof. We have

    aρ[η](u,u)=YA(+iη)u¯(+iη)udyI+ρ2Y(+iη)2u¯(+iη)2udyII. (2.5)

    We shall estimate the two summands separately. For the first summand, observe that

    I=YA(+iη)u¯(+iη)udy=YAu¯udy+2Re{YAiηu¯udy}+YAηuη¯udy, (2.6)

    where Re denotes the real part. Now we shall estimate each term on the RHS above. The first term of I is estimated as follows:

    YAu¯udy=YAu¯udyαY|u|2dy. (2.7)

    For the second term of I, observe that

    |2Re{YAηu¯udy}|2Y|Aηu¯u|dyC1Y|ηu¯u|dyC1||ηu||L2(Y)||u||L2(Y)C1C2||u||2L2(Y)+C1C2||u||L2(Y). (2.8)

    Finally, the third term of I is dominated by L2(Y) norm of u as follows:

    |YAηuη¯udy|C3Y|ηuη¯u|dyC4||u||L2(Y). (2.9)

    Now, we may choose C2 so that C1C2=α2, then

    Iα2||u||L2(Y)+α2||u||L2(Y)(α2+C1C2+C4)||u||2L2(Y). (2.10)

    For the second summand, observe that

    II=ρ2Y|(+iη)2u|2dy=ρ2Y|Δu|2dy+ρ2Y|η|4|u|2dy+4ρ2Y|ηu|2dy+2ρ2iIm{Y|η|2u¯Δudy}+4iρ2Re{Y(ηu)Δ¯udy}+4iρ2Re{Y|η|2u(η¯u)dy}. (2.11)

    We estimate the last three terms as follows.

    ρ2|2iIm{Y|η|2u¯Δudy}|2ρ2Y|η|2|u||Δu|dy2ρ2|η|2||u||L2||Δu||L248ρ2|η|4||u||2L2+ρ248||Δu||2L2. (2.12)
    ρ2|4iRe{Y(η)uΔ¯udy}|4ρ2Y|(η)u||Δu|dy4ρ2||(η)u||L2||Δu||L216ρ23||(η)u||2L2+3ρ24||Δu||2L2. (2.13)
    ρ2|4iRe{Y|η|2u(η)¯udy}|4ρ2Y|(η)u||η|2|u|dy4ρ2|η|2||u||L2||(η)u||L2192ρ2|η|4||u||2L2+ρ248||(η)u||2L2. (2.14)

    The previous threee estimate use Cauchy-Schwarz inequality for the second step and Young's inequality for the third step. Substituting the inequalities (2.12), (2.13) and (2.14) into (2.11), we get

    II11ρ248||Δu||2L2(Y)240ρ2|η|4||u||2L2(Y)65ρ248||(η)u||2L2(Y). (2.15)

    Since |η|12, we get

    II11ρ248||Δu||2L2(Y)15ρ2||u||2L2(Y)65ρ2192||u||2L2(Y).

    Now, notice that

    ||u||2L2=Y|u|2dy=YuΔudy6548||u||2L2+1265||Δu||2L2. (2.16)

    Substituting (2.16) in (2.15), we get

    IIρ26||Δu||2L2(Y)16ρ2||u||2L2(Y). (2.17)

    Combining (2.10) and (2.17), we obtain (2.4) with

    C=(α2+C1C2+C4+16ρ2). (2.18)

    Remark 2.2. In the Gå rding type inequality for the operator Aρ(η), the perturbation CI in zeroth term depends on the parameter ρ. It is possible to avoid the dependence of C on ρ by forgoing the shift iη in the biharmonic term. This simplification was employed in [34].

    Now that we have the coercivity estimate (2.4), we can prove the existence of Bloch eigenvalues and eigenfunctions for the operator Aρ.

    Theorem 2.3. For each ηY and ρ>0, the singularly perturbed Bloch eigenvalue problem (2.2) admits a countable sequence of eigenvalues and corresponding eigenfunctions in the space H2(Q).

    Proof. Lemma 2.1 shows that for every ηY the operator Aρ(η)+CI is elliptic on H2(Y). Hence, for fL2(Y), this shows that Aρ(η)u+Cu=f is solvable and the solution is in H2(Y). As a result, the solution operator Sρ(η) is continuous from L2(Y) to H2(Y). Since the space H2(Y) is compactly embedded in L2(Y), Sρ(η) is a self-adjoint compact operator on L2(Y). Therefore, by an application of the spectral theorem for self-adjoint compact operators, for every ηY we obtain an increasing sequence of eigenvalues of Aρ(η)+CI and the corresponding eigenfunctions form an orthonormal basis of L2(Y). However, note that both the operators Aρ(η) and Aρ(η)+CI have the same eigenfunctions but each eigenvalue of the two operators differ by C. We shall denote the eigenvalues and eigenfunctions of the operator Aρ(η) by η(λρm(η),ϕρm(,η)).

    Remark 2.4. We can prove the existence of Bloch eigenvalues and eigenfunctions for the case ρ=0 by a similar method. This corresponds to the standard Bloch eigenvalue problem considered in [13].

    Now that we have proved the existence of Bloch eigenvalues and eigenfunctions, we can state the Bloch Decomposition Theorem which offers a partial diagonalization of the operator Aρ in terms of its Bloch eigenvalues. This is facilitated by the Bloch transform which is a mapping from L2(Rd) to 2(N;L2(Y)). The proof is similar to the one in [6] and is therefore omitted. Note that the proof relies on a measurable selection of the Bloch eigenfunctions with respect to η. A measurable selection of Bloch eigenfunctions for the Schrödinger operator was first demonstrated in [38]. In contrast, the Bloch eigenvalues are Lipschitz continuous in the dual parameter η. We will prove this fact in Section 4.

    Theorem 2.5. Let ρ>0. Let gL2(Rd). Define the mth Bloch coefficient of g as

    Bρmg(η):=Rdg(y)eiyη¯ϕρm(y;η)dy,mN,ηY. (2.19)

    1. The following inverse formula holds

    g(y)=Ym=1Bρmg(η)ϕρm(y;η)eiyηdη. (2.20)

    2. Parseval's identity

    ||g||2L2(Rd)=m=1Y|Bρmg(η)|2dη. (2.21)

    3. Plancherel formula For f,gL2(Rd), we have

    Rdf(y)¯g(y)dy=m=1YBρmf(η)¯Bρmg(η)dη. (2.22)

    4. Bloch Decomposition in H1(Rd) For an element F=u0(y)+Nj=1uj(y)yj of H1(Rd), the following limit exists in L2(Y):

    BρmF(η)=Rdeiyη{u0(y)¯ϕρm(y;η)+iNj=1ηjuj(y)¯ϕρm(y;η)}dyRdeiyηNj=1uj(y)¯ϕρmyj(y;η)dy. (2.23)

    The definition above is independent of the particular representative of F.

    5. Finally, for gD(Aρ),

    Bρm(Aρg)(η)=λρm(η)Bρmg(η). (2.24)

    The Bloch wave method of homogenization requires differentiability of the Bloch eigenvalues and eigenfunctions in a neighbourhood of η=0. In this section, we will prove the following theorem. As before, ρ>0 will be treated as a fixed number.

    Theorem 3.1. For every ρ>0, there exists δρ>0 and a ball Uρ:=Bδρ(0):={ηY:|η|<δρ} such that

    1. The first Bloch eigenvalue ηλρ(η) of Aρ is analytic for ηUρ.

    2. There is a choice of corresponding eigenfunctions ϕρ1(,η) such that ηUρϕρ1(,η)H2(Y) is analytic.

    For the proof, we will make use of Kato-Rellich theorem which establishes the existence of a sequence of eigenvalues and eigenfunctions associated with a selfadjoint holomorphic family of type (B). The definition of selfadjoint holomorphic family of type (B) and other related notions may be found in Kato [21]. Nevertheless, they are stated below for completeness. We begin with the definition of a holomorphic family of forms of type (a).

    Definition 3.2.

    1. The numerical range of a form a is defined as

    Θ(a)={a(u,u):uD(a),||u||=1}

    where D(a) denotes the domain of the form a. Here, D(a) is a subspace of a Hilbert space H.

    2. The form a is called sectorial if there are numbers cR and θ[0,π/2) such that

    Θ(a)Sc,θ:={λC:|arg(λc)|θ)}.

    3. A sectorial form a is said to be closed if given a sequence unD(a) with unu in H and a(unum)0 as n,m, we have uD(a) and a(unu)0 as n.

    Definition 3.3. [Kato] A family of forms a(z),zD0CM is called a holomorphic family of type (a) if

    1. each a(z) is sectorial and closed with domain DH independent of z and dense in H,

    2. a(z)[u,u] is holomorphic for zD0CM for each uD.

    A family of operators is called a holomorphic family of type (B) if it generates a holomorphic family of forms of type (a).

    In [21, 31], Kato-Rellich theorem is stated only for a single parameter family. In [5], one can find the proof of Kato-Rellich theorem for multiple parameters with the added assumption of simplicity for the eigenvalue at η=0.

    Theorem 3.4. (Kato-Rellich) Let D(˜η) be a self-adjoint holomorphic family of type (B) defined for ˜η in an open set in CM. Further let λ0=0 be an isolated eigenvalue of D(0) that is algebraically simple. Then there exists a neighborhood R0CM containing 0 such that for ˜ηR0, the following holds:

    1. There is exactly one point λ(˜η) of σ(D(˜η)) near λ0=0. Also, λ(˜η) is isolated and algebraically simple. Moreover, λ(˜η) is an analytic function of ˜η.

    2. There is an associated eigenfunction ϕ(˜η) depending analytically on ˜η with values in H.

    The proof of Theorem 3.1 proceeds by complexifying the shifted operator Aρ(η) before verifying the hypothesis of Kato-Rellich Theorem.

    Proof. (Proof of Theorem 3.1)

    (i) Complexification of Aρ(η): The form aρ[η](,) is associated with the operator Aρ(η). We define its complexification as

    t(˜η)=YA(+iστ)u(iσ+τ)¯udy+ρ2Y|(+iσ+τ)2u|2dy

    for ˜ηR where

    R:={˜ηCM:˜η=σ+iτ,σ,τRM,|σ|<1/2,|τ|<1/2}.

    (ii) the form t(˜η) is sectorial: We have

    t(˜η)=YA(+iστ)u(iσ+τ)¯udy+ρ2Y|(+iσ+τ)2u|2dy=YA(+iσ)u(iσ)¯udyYA(τu)¯udy+YAu(τ¯u)dyYAτuτ¯udy+iYAσuτ¯udy+iYAτuσ¯udy+ρ2Y(Δ|σ|2+|τ|2)u(Δ|σ|2+|τ|2)¯udy+2ρ2Y(Δ|σ|2+|τ|2)u(τiσiστ)¯udy+2ρ2Y(iστiστ)u(Δ|σ|2+|τ|2)¯udy+4ρ2Y(iστiστ)u(τiσiστ)¯udy.

    From above, it is easy to write separately the real and imaginary parts of the form t(˜η).

    t(˜η)[u]=YA(+iσ)u(iσ)¯udyYAτuτ¯udy+ρ2Y(Δ|σ|2+|τ|2)u(Δ|σ|2+|τ|2)¯udy4ρ2Y|(τ)u|2dy+4ρ2Y|(σ)u|2dy4ρ2Y|(τσ)u|2dy+8ρ2Re{Yi(τ)u(στ)¯udy}+4ρ2Re{Yi(σ)uΔ¯udy}+4ρ2Re{Yi|σ|2u(σ)¯udy}.
    t(˜η)[u]=YAσuτ¯udy+YAτuσ¯udy+2{YAuτ¯udy}+8ρ2Im{Yi(τσ)u(σ)¯udy}+8ρ2Im{Yi(σ)u(στ)¯udy}+4ρ2Im{YΔu(τ)¯udy}4ρ2Im{YiΔu(στ)¯udy}+4ρ2Im{Y(τ)u|σ|2¯udy}+4ρ2Im{Yi|σ|2u(στ)¯udy}+4ρ2Im{Y|τ|2u(τ)¯udy}4ρ2Im{Yi|τ|2u(στ)¯udy}.

    The following coercivity estimate can be easily found for the real part:

    t(˜η)[u]+C5||u||2L2(Y)α2(||u||2L2(Y)+||u||2L2(Y))+ρ26||Δu||2L2(Y). (3.1)

    Let us define the new form ˜t(˜η) by ˜t(˜η)[u,v]=t(˜η)[u,v]+(C5+C6)(u,v)L2(Y). Then it holds that

    ˜t(˜η)[u]α2(||u||2L2(Y)+||u||2L2(Y))+ρ26||Δu||2L2(Y)+C6||u||2L2(Y).

    Also, the imaginary part of ˜t(˜η) can be estimated as follows:

    ˜t(˜η)[u]C7||u||2L2(Y)+C8||u||2L2(Y)+C9||Δu||2L2(Y)C10=max{2C8α,6C9ρ2}=C6=C10C7C10(C6||u||2L2(Y)+α2||u||2L2(Y)+ρ26||Δu||2L2(Y))C10(˜t(˜η)[u]α2||u||2L2(Y)).

    This shows that ˜t(˜η) is sectorial. However, sectoriality is invariant under translations by scalar multiple of identity operator in L2(Y), therefore the form t(˜η) is also sectorial.

    (iii) The form t(˜η) is closed: Suppose that untu. This means that unu in L2(Q) and t(˜η)[unum]0. As a consequence, t(˜η)[unum]0. By (3.1), ||unum||H2(Y)0, i.e., (un) is Cauchy in H2(Y). Therefore, there exists vH2(Y) such that unv in H2(Y). Due to uniqueness of limit in L2(Y), v=u. Therefore, the form is closed.

    (iv) The form t(˜η) is holomorphic: The holomorphy of t follows as a consequence of t being a quartic polynomial in η.

    (v) 0 is an isolated eigenvalue: Zero is an eigenvalue because constants are eigenfunctions of Aρ(0)=A+ρ2Δ2. As a result, C is an eigenvalue of Aρ(0)+CI. We proved using Lemma 2.1 that Aρ(0)+CI has compact resolvent. Therefore, C1 is an eigenvalue of (Aρ(0)+CI)1 and C1 is isolated. Hence, zero is an isolated point of the spectrum of Aρ(0).

    (vi) 0 is a geometrically simple eigenvalue: Denote by kerAρ(0) the kernel of operator Aρ(0). Let vkerAρ(0), then YAvvdy+ρ2Y|Δv|2=0. Due to the coercivity of the matrix A, we obtain ||v||L2(Y)=0. Hence, v is a constant. This shows that the eigenspace corresponding to eigenvalue 0 is spanned by constants, therefore, it is one-dimensional.

    (vii) 0 is an algebraically simple eigenvalue: Suppose that vH2(Y) such that Aρ(0)2v=0, i.e., Aρ(0)vkerAρ(0). This implies that Aρ(0)v=C for some generic constant C. However, by the compatibility condition for the solvability of this equation, we obtain C=0. Therefore, vkerAρ(0). This shows that the eigenvalue 0 is algebraically simple.

    In Section 3, we have proved that the first Bloch eigenvalue and eigenfunction is analytic in a neighbourhood of η=0. However, a priori, this neighbourhood depends on the parameter ρ. In this section, we will prove that the neighbourhood is, in fact, independent of ρ. This requirement is essential for problems where simultaneous limits with respect to two parameters are studied, such as [28, 16]; as well as for quantitative estimates, such as [10].

    We begin by proving that Bloch eigenvalues are Lipschitz continuous in the dual parameter.

    Lemma 4.1. For all mN and ρ>0, λρm is a Lipschitz continuous function of ηY.

    Proof. The following form is associated with Aρ(η):

    aρ[η](u,u)=YA(+iη)u¯(+iη)udy+ρ2Y(+iη)2u¯(+iη)2udy. (4.1)

    Hence, for η,ηY, we have

    aρ[η]aρ[η]=2Re{YAi(ηη)u¯udy}+YAηuη¯udy+YAηuη¯udy+ρ2Y(|η|2|η|2)|u|2dy+4ρ2Y|ηu|2|ηu|2dy+2ρ2iIm{Y(|η|2|η|2)uΔ¯udy}+4ρ2iRe{Y(ηη)uΔ¯udy}+4ρ2iRe{Y|η|2u(η)¯u|η|2u(η)¯udy}C|ηη|||u||2H1(Y)+C|ηη|ρ2{||Δu||2L2(Y)+||u||2L2(Y)},

    where C and C are generic constants independent of ρ. By the Courant-Fischer minmax characterization of eigenvalues, we obtain

    λρm(η)λρm(η)+C|ηη|μm+Cρ2|ηη|νm, (4.2)

    where μm is the mth eigenvalue of the following spectral problem:

    {Δum+um=μmuminYumisYperiodic,

    and νm is the mth eigenvalue of the following spectral problem:

    {Δ2vm+vm=νmvminYvmisYperiodic.

    By interchanging the role of η and η, we obtain

    |λρm(η)λρm(η)|C(μm+ρ2νm)|ηη|. (4.3)

    Here, C is a generic constant independent of ρ.

    Now, we will prove a spectral gap result, viz. the second Bloch eigenvalue is bounded below.

    Lemma 4.2. For all m2 and for all ηY, we have

    λρm(η)αλN2, (4.4)

    where λN2 is the second Neumann eigenvalue of the operator Δ in Y.

    Proof. Notice that λρm(η)λρ2(η) for all m2 and for all ηY. Next, observe that

    λρ2(η)=infWH2(Y)dim(W)=2maxϕWϕ0YA(eiηyϕ)(eiηy¯ϕ)dy+ρ2Y|Δ(eiηyϕ)|2dyY|ϕ|2dyinfWH2(Y)dim(W)=2maxψWψ0YAψ¯ψdy+ρ2Y|Δψ|2dyY|ψ|2dyinfWH2(Y)dim(W)=2maxψWψ0YAψ¯ψdyY|ψ|2dyinfWH1(Y)dim(W)=2maxψWψ0YAψ¯ψdyY|ψ|2dyαinfWH1(Y)dim(W)=2maxψWψ0Y|ψ|2dyY|ψ|2dy=αλN2.

    The bound obtained in Lemma 4.2 will be useful for the small ρ regime, however, for the large ρ regime, that is, for ρ, we need a different lower bound.

    Lemma 4.3. For all m2 and for all ηY, we have

    λρm(η)Cρ2κ2C, (4.5)

    where κ2 is the 2nd eigenvalue of periodic bilaplacian on Y, where C and C are generic constants independent of ρ and η.

    Proof. Recall the following Gå rding type estimate (2.4) for the form aρ[η] associated with the operator Aρ(η):

    aρ[η](u,u)+C||u||2L2(Y)ρ26||Δu||2L2(Y)+α2||u||H1(Q).

    The inequality in Lemma 4.3 follows readily from above by applying the minmax characterization.

    Remark 4.4. In Lemma 4.2 and Lemma 4.3, we have avoided estimating the second Bloch eigenvalue by using the spectral problem associated with Neumann bilaplacian as it is known to be ill-posed [30]. Moreover, polyharmonic Neumann eigenvalue problems on polygonal domains (such as Y) are less well understood [20, 17]. However, suitable natural boundary conditions associated with the operator Δ2τΔ are obtained in [9].

    We are finally in a position to prove that the neighbourhood of analyticity of the first Bloch eigenvalue does not depend on the parameter ρ.

    Theorem 4.5. There exists a neighbourhood U=Bδ(0) of η=0 , not depending on ρ, such that λρ1(η) is analytic on Bδ(0).

    Proof. It was proved in Theorem 3.1 that the first Bloch eigenvalue is analytic in a neighbourhood of η=0. However, a priori it is not clear whether this neighbourhood is independent of ρ. To prove this, it suffices to prove that the first Bloch eigenvalue is simple in a neighbourhood of η=0 independently of ρ. Observe that

    |λρ1(η)λρ2(η)|λρ2(η)|λρ1(η)λρ1(0)||λρ2(η)λρ2(0)|(4.3)λρ2(η)2(C+ρ2)|η|, (4.6)

    where C is a generic constant independent of ρ and η.

    ● For sufficiently large ρ,

    |λρ1(η)λρ2(η)|λρ2(η)2(C+ρ2)|η|Lemma4.3(Cρ2C)2(C+ρ2)|η|largeρCρ22ρ2|η|>0

    for |η|<C2. Here, C,C,C and C are all generic constants independent of ρ and η.

    ● For remaining values of ρ,

    |λρ1(η)λρ2(η)|λρ2(η)2(C+ρ2)|η|Lemma4.2αλN22(C+ρ2)|η|αλN22C|η|>0

    for |η|<αλN22C.

    Remark 4.6. In the papers [33, 34], an additional artificial parameter is introduced in the Bloch eigenvalue problem to facilitate the homogenization method. Unlike (1.1), these papers employ successive limits of the two parameters instead of simultaneous limits. Therefore, the non-dependence of the neighbourhood of analyticity on the second parameter is not required in [33, 34].

    In this section, we will consider the classical cell problem associated with (1.1) and the estimates for the corrector field. This section will allow us to characterize the homogenized tensor for (1.1) and the corrector field in terms of Bloch eigenvalues and eigenfunctions.

    For 1jd, consider the following cell problem associated with the operator (1.1):

    (5.1)

    By a simple application of Lax-Milgram lemma on H2(Y), we obtain solution to above for every ρ>0 (For ρ=0, Lax-Milgram lemma is applied for H1(Y)). Further, since the equation is also satisfied in the sense of distributions, we conclude that χρjH3(Y) for ρ>0. By using χρj as a test function, we obtain the following bound:

    ρ||Δχρj||L2(Y)+||χρj||H1(Y)C. (5.2)

    If we use Δχρj as a test function, we obtain

    ρ2||3χρj||L2(Y)C. (5.3)

    We also collect below a few estimates which will be required later. Similar estimates have been proved in [26] to which we refer for more details.

    Lemma 5.1. Let ρ1,ρ2>0. Let χρ1,χρ2H2(Y) be solutions to (5.1) for ρ=ρ1 and ρ=ρ2 respectively, then the following estimate holds.

    ||χρ1χρ2||L2(Y)C|1(ρ1/ρ2)2|. (5.4)

    Proof. Define z=χρ1jχρ2j, then z satisfies the following equation

    ρ21Δ2zdivA(y)z=(ρ22ρ21)Δ2χρ2. (5.5)

    Now, the quoted estimate readily follows by taking z as the test function, applying uniform ellipticity of A, (5.3) and an application of Poincaré inequality.

    Lemma 5.2. Let ρ>0. Let χρH2(Y) be the solution to (5.1) and let χ0H1(Y) solve

    (5.6)

    Then, there is q(1,) such that for every ϰ>0 there exists a matrix B with entries in C(Y) such that ||AB||Lq(Y)ϰ and the following estimates hold.

    ||χ0jχBj||L2(Y)Cϰ (5.7)
    ||χρχB||L2(Y)C{ρ||χB||H2(Y)+ϰ}, (5.8)

    where χBH1(Y) solve

    (5.9)

    Proof. Observe that given any q(1,) and ϰ>0, we can find a smooth periodic matrix B with the same ellipticity constant and upper bound as A such that

    ||AB||Lq(Y)ϰ.

    For example, this can be achieved by a standard smoothing by convolution. Now, by regularity theory, χBjH2(Y). Define z=χρjχBj, then z satisfies the following equation

    ρ2Δ2zdivB(y)z=ρ2Δ2χBj+div(AB)χρj. (5.10)

    We test this equation against z to obtain:

    ρ2Y|Δz|2dy+YA(y)zzdyρ2Y|ΔχBj||Δz|dy+Y|AB||χρj||z|dy.

    This leads to

    ρ2||Δz||2L2+α||z||2L2ρ2||ΔχBj||L2||Δz||L2+||z||L2(Y|AB|2|χρj|2dy)1/2.

    By Young's inequality,

    ||z||L2C{ρ||ΔχBj||L2+(Y|AB|2|χρj|2dy)1/2}.

    On the last term, we apply a form of Meyers estimate for the Lp integrability of χρj proved in [26, Page 7, Theorem 2.3]. This fixes the choice of q. This finishes the proof of the estimate (5.8). The proof of (5.7) is similar and simpler and hence omitted.

    For every fixed 0ρ<, the homogenized tensor for the operator Aρ,ε=ρ2Δ2divA(xε) is given by

    Aρ,hom:=MY(A+Aχρ) (5.11)

    Definition 5.3 (Homogenized Tensor for Aκ,ε)

    Ahom:={MY(A+Aχθ)for0<θ<whereρ=κεθ,MY(A+Aχ0)whenρ=κε0,MY(A)whenρ=κε. (5.12)

    In this section, we will give a new characterization of the homogenized tensor (see Definition 5.3), and corrector field (5.1) in terms of the first Bloch eigenvalue and eigenfunction. These characterizations are obtained by differentiating the Bloch spectral problem (2.2) with respect to the dual parameter η. Indeed, this is possible since we proved the analyticity of the first Bloch eigenvalue and eigenfunction with respect to η in a neighbourhood of η=0 in Theorem 3.1. Moreover we will use the properties of the first Bloch eigenvalue to prove the stability of the homogenized tensor with respect to the limits ρ0 and ρ.

    We recall the Bloch eigenvalue problem for the operator Aρ here:

    ρ2(+iη)4ϕρ1(y;η)(+iη)A(y)(+iη)ϕρ1(y;η)=λρ1(η)ϕρ1(y;η). (6.1)

    We know that λρ1(0)=0. For ηY, recall that Aρ(η)=ρ2(+iη)4(+iη)A(y)(+iη). In this section, for notational convenience we will hide the dependence on y.

    We shall normalize the average value of the first Bloch eigenfunction ϕρ1(;η) to be (2π)d/2, that is,

    MY(ϕρ1(,η))=(2π)d/2 (6.2)

    for all η in the neighbourhood of analyticity. We shall use β to denote a multiindex, such as β=(β1,β2,,βd)Nd{0} and |β|:=|β1|+|β2|++|βd|. We will use the shorthand βη to denote βη:=β1ηβ11βdηηdd. For simplicity, we will also use the notation β0u=β1uηβ11βduηηdd|η=0. Differentiating (6.2), we obtain

    MY(β0ϕρ1)=0 (6.3)

    for all |β|>0. Later on, we will see this as the compatibility condition associated with the equation satisfied by β0ϕρ1. Denote by β0Aρ:=βAηβ|η=0. Then, it holds true that

    βηA0forall|β|>4, (6.4)

    since Aρ(η) is a fourth order polynomial in η. Direct calculation shows that

    Aρ(0)=ρ24A(y)ej0Aρ=4iρ2ej2iejAiAejej+ek0Aρ=4ρ2δjk28ρ2yjyk+2ajkej+ek+el0Aρ=8iρ2(δjkyl+δjlyk+δklyj)ej+ek+el+em0Aρ=8ρ2(δjkδlm+δjlδkm+δjmδkl), (6.5)

    where ej denotes the standard Euclidean unit vector with 1 in the jth place and 0 elsewhere. Now, we are in a position to write down the differential equations satisfied by the derivatives of ϕρ1 of all orders. To this end, we recall the Leibniz's formula for the derivatives of product of functions, viz.,

    β(fg)=γNd{0}(βγ)γfβγg, (6.6)

    where (βγ)=(β1γ1)(βdγd) and the sum is always finite since (βjγj)=0 whenever βj<γj.

    Cell problems for β0ϕρ1_ On differentiating (6.1) with respect to η and applying (6.6) we obtain

    Aρ(0)β0ϕρ1+dj=1βjej0Aρβej0ϕρ1+j,k(βej+ek)ej+ek0Aρβejek0ϕρ1
    +j,k,l(βej+ek+el)ej+ek+el0Aρβejekel0ϕρ1+j,k,l,m(βej+ek+el+em)ej+ek+el+em0Aρβejekelem0ϕρ1=γNd{0}(βγ)γ0λρ1βγ0ϕρ1. (6.7)

    Substituting (6.5) in (6.7), we obtain

    (ρ24A(y))β0ϕρ1=dj=1βj(4iρ2ej2+iejA+iAej)βej0ϕρ1j,k(βej+ek)(4ρ2δjk28ρ2yjyk+2ajk)βejek0ϕρ1+j,k,l(βej+ek+el)(8iρ2(δjkyl+δjlyk+δklyj))βejekel0ϕρ1j,k,l,m(βej+ek+el+em)(8ρ2(δjkδlm+δjlδkm+δjmδkl))βejekelem0ϕρ1+γNd(βγ)γ0λρ1βγ0ϕρ1. (6.8)

    Expression for β0λρ1_ It is easy to see that λρ1(0)=0 since 0 is the first eigenvalue of the operator Aρ(0). On the other hand, λρ1 is an even function of η since aρ[η](ϕ,ϕ)=aρ[η](¯ϕ,¯ϕ) and ¯ϕH2(Y) if and only if ϕH2(Y) (see also [15, Page 41, Lemma 4.4]).

    Rearranging (6.8), we get

    β0λρ1ϕρ1(0)=(ρ24A(y))β0ϕρ1+dj=1βj(4iρ2ej2+iejA+iAej)βej0ϕρ1j,k(βej+ek)(4ρ2δjk28ρ2yjyk+2ajk)βejek0ϕρ1+j,k,l(βej+ek+el)(8iρ2(δjkyl+δjlyk+δklyj))βejekel0ϕρ1j,k,l,m(βej+ek+el+em)(8ρ2(δjkδlm+δjlδkm+δjmδkl))βejekelem0ϕρ1+γNdγβ(βγ)γ0λρ1βγ0ϕρ1. (6.9)

    Integrating (6.9) over Y, using (6.2) and (6.3) and the fact that integrals over Y of derivatives of periodic functions vanish due to Green's identity, we obtain the following formula for the derivatives of the first Bloch eigenvalue at η=0:

    β0λρ1={2j,k(βej+ek)MY(ajkβejek0ϕρ1)idj=1βjMY(ejAβej0ϕρ1)when|β|4.2j,k(βej+ek)MY(ajkβejek0ϕρ1)idj=1βjMY(ejAβej0ϕρ1)j,k,l,m8ρ2(δjkδlm+δjlδkm+δjmδkl)ϕρ1(0)when|β|=4. (6.10)

    We specialize to β=el in (6.10) to get λρ1ηl(0)=0 for all l=1,2,,d.

    On the other hand, if we set β=el in (6.8), we obtain

    (A(y)+ρ2Δ2)ϕρ1ηl(0)=A(y)eliϕρ1(0).

    Comparing with (5.1), we conclude that χρl1iϕρ1(0)ϕρ1ηl(0) is a constant.

    We also specialize to β=el+ek in (6.10) to get

    122λρ1ηkηl(0)=1|Y|Y(ekAel+12ekAχρl+12elAχρk)dy. (6.11)

    On comparing (6.11) with (5.11), we obtain the following theorem:

    Theorem 6.1. The first Bloch eigenvalue and eigenfunction satisfy:

    1. λρ1(0)=0.

    2. The eigenvalue λρ1(η) has a critical point at η=0, i.e.,

    λρ1ηl(0)=0,l=1,2,,d. (6.12)

    3. For l=1,2,,d, the derivative of the eigenvector (ϕρ1/ηl)(0) satisfies:

    (ϕρ1/ηl)(y;0)iϕρ1(y;0)χρl(y) is a constant in y where χρl solves the cell problem (5.1).

    4. The Hessian of the first Bloch eigenvalue at η=0 is twice the homogenized matrix Aρ,hom as defined in (5.11), i.e.,

    122λρ1ηkηl(0)=ekAρ,homel. (6.13)

    Now, we will prove the stability of homogenized tensor in the limits ρ0 and ρ.

    Lemma 6.2. Let 0θ< such that ρθ then λρ1(η)λθ1(η) uniformly on compact sets K contained in their common domain of definition.

    Proof. The fact that the first Bloch eigenvalues {λρ1(η)}ρ0 are analytic on a common domain in Y has been proved in Theorem 4.5. Since ρθ where θ is finite, the sequence ρ is bounded, that is, 0<ρ<ρ0 for some 0<ρ0<. We will prove that the family {λρ1(η)}0ρ<ρ0 is locally uniformly bounded, that is, for every compact set K,

    |λρ1(η)||λρ1(η)λρ1(0|(4.3)C(μ1+ν1ρ2)|η|<CforallηK,

    where C is independent of ρ and η. It is an easy consequence of Montel's Theorem [25, Page 9, Prop. 7] that pointwise convergence implies uniform convergence for locally uniformly bounded sequences. We will show that λρ1(η) converges pointwise to λθ1(η). Applying minmax characterization to the form below

    aρ[η](u,u)=YA(+iη)u¯(+iη)udy+ρ2Y(+iη)2u¯(+iη)2udy=aθ[η](u,u)+(ρ2θ2)Y(+iη)2u¯(+iη)2udy

    we obtain the following inequality:

    λρ1(η)λθ1(η)(ρ2θ2)ϑ1(η),

    where ϑ1(η) is the first Bloch eigenvalue of the bilaplacian and λθ1(η) is the first Bloch eigenvalue of the operator div(A)+θ2Δ2 for θ[0,). On the other hand, we also have

    aρ[η](u,u)aθ[η](u,u)forρθ,

    so that an application of minmax characterization yields:

    λρ1(η)λθ1(η)forρθ.

    Thus, we obtain

    0λρ1(η)λθ1(η)(ρ2θ2)ϑ1(η)forρθ.

    As a consequence, for each ηY, λρ1(η)λθ1(η) as ρθ.

    Theorem 6.3. Let 0θ such that ρθ then Aρ,homAhom, with Ahom as in Definition 5.3.

    Proof.

    Case 1. θ[0,): By the characterization in Theorem 6.1, Aρ,hom=122ηλρ1(0). In Lemma 6.2, we have proved that λρ1 conveges uniformly to λθ1 when ρθ for θ[0,). By Theorem of Weierstrass [25, Page 7, Prop. 5], derivatives of all orders of λρ1(η) converge uniformly on compact sets to corresponding derivatives of λθ1(η). In particular, 2ηλρ1 converges uniformly on compact sets to 2λθ1. As a consequence,

    Aρ,hom=122ηλρ1(0)122ηλθ1(0)=Ahomasρθ[0,).

    Case 2. θ=: Observe that

    |Aρ,homMY(A)|=|YA(y)χρ(y)dy|C||χρ||L2(Y)Cρ2,

    where the last inequality follows from Poincaré inequality and (5.3). Therefore, as ρ,

    Aρ,homMY(A)=Ahomasρ.

    Remark 6.4. As a consequence of Lemma 6.2 and the discussion in Case 1 of Theorem 6.3, for any R>0, and 0ρ<R, the first Bloch eigenvalue and all its derivatives are bounded uniformly in their domain of analyticity independent of ρ. We may also conclude from the same that the corresponding (suitably normalized) Bloch eigenfunction and all their derivatives with respect to dual parameter are bounded in H1 independent of ρ. For the first Bloch eigenfunction, the boundedness is a consequence of the following calculation, which is a consequence of (2.4):

    |λρ1(η)|+Cλρ1(η)+C=aρ[η](ϕρ1(η),ϕρ1(η))+Cα2||ϕρ1(η)||2H1(Y).

    Recall that the number C is explicitly given in (2.18). The boundedness in H1(Y) of the derivatives of the first Bloch eigenfunction in the dual parameter is a consequence of its analyticity [25, Page 5, Prop. 3]. This precludes the case ρ which is covered in Theorem 6.5.

    We end this section by finding boundedness estimates for higher order derivatives of the first Bloch eigenvalue and eigenfunction in the dual parameter in the regime ρ.

    Theorem 6.5. For 1ρ<,

    1. ||ej0ϕρ1||H1, \partial^{e_j}_0\lambda^\rho_1 = 0 .

    2. ||\partial_0^{e_j+e_k}\phi^\rho_1||_{H^1}\lesssim\frac{1}{\rho^2} , \left|\partial^{e_j+e_k}_0\lambda^\rho_1-\mathcal{M}_Y(e_j\cdot Ae_k)\right|\lesssim\frac{1}{\rho^2} .

    3. ||\partial_0^{e_j+e_k+e_l}\phi^\rho_1||_{H^1}\lesssim\frac{1}{\rho^2} , \partial^{e_j+e_k+e_l}_0\lambda^\rho_1 = 0 .

    4. ||\partial_0^{\beta}\phi^\rho_1||_{H^1}\lesssim 1 for all |\beta|\geq 4 .

    5. |\partial^{\beta}_0\lambda^\rho_1|\lesssim{\begin{cases} &\rho^2\;\mathit{for}\;|\beta| = 4.\\ &1\;\mathit{for}\;|\beta|>4.\\ \end{cases} } .

    Proof. The estimates are computed in tandem from the equations (6.8) and (6.10). One begins by proving the estimate on the derivative of the first Bloch eigenfunction at \eta = 0 , followed by the estimate on the corresponding derivative of the first Bloch eigenvalue. The solvability of (6.8) in H^2_\sharp(Y) follows from a standard application of Lax-Milgram lemma. On the other hand, one can also read off from (6.8) that the solution \partial^\beta_0\phi^\rho_1\in H^3_\sharp(Y) . As a consequence, the estimate on the derivatives of the first Bloch eigenfunction are obtained by employing the test function \Delta_y\partial^\beta_0\phi^\rho_1 in (6.8) and repeated applications of Poincaré inequality. The computations are standard and therefore, omitted.

    In this section, we will relate the Bloch spectral problem (2.2) to the Bloch spectral problem at the \varepsilon -scale:

    \begin{eqnarray} \begin{cases} \mathcal{A}^{\kappa, \varepsilon}(\eta)\phi^{\kappa, \varepsilon}&:= \kappa^2(\nabla+i\xi)^4\phi^{\kappa, \varepsilon}(x)-(\nabla+i\xi)\cdot A\left(\frac{x}{ \varepsilon}\right)(\nabla+i\xi)\phi^{\kappa, \varepsilon}(x)\\ &\qquad = \lambda^{\kappa, \varepsilon}(\xi)\phi^{\kappa, \varepsilon}(x)\\ \phi^{\kappa, \varepsilon}(x+2\pi p \varepsilon)& = \phi^{\kappa, \varepsilon}(x), p\in\mathbb{Z}^d,\xi\in \frac{Y^{'}}{ \varepsilon}. \end{cases} \end{eqnarray} (7.1)

    Comparing to (2.2), by homothety and \kappa = \rho \varepsilon , we conclude that

    \begin{align} \lambda^{\kappa, \varepsilon}(\xi) = \varepsilon^{-2}\lambda^\rho( \varepsilon\xi){\rm{ and }}\phi^{\kappa, \varepsilon}(x,\xi) = \phi^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right). \end{align} (7.2)

    Now, we can state the Bloch decomposition theorem of L^2(\mathbb{R}^d) at \varepsilon -scale. We shall normalize \phi^\rho_1(y;0) to be (2\pi)^{-d/2} .

    Theorem 7.1. Let \rho >0 . Let g\in L^2(\mathbb{R}^d) . Define the m^{th} Bloch coefficient of g at \varepsilon -scale as

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_mg(\xi):=\int_{\mathbb{R}^d}g(x)e^{-ix\cdot\xi}\overline{\phi_m^{\kappa, \varepsilon}(x;\xi)}\,dx,\; m\in\mathbb{N},\; \xi\in \frac{Y^{'}}{ \varepsilon}. \end{align} (7.3)

    1. The following inverse formula holds

    \begin{align} g(x) = \int_{\frac{Y^{'}}{ \varepsilon}}\sum\limits_{m = 1}^{\infty}\mathcal{B}^{\kappa, \varepsilon}_mg(\xi)\phi_m^{\kappa, \varepsilon}(x;\xi)e^{ix\cdot\xi}\,d\xi. \end{align} (7.4)

    2. Parseval's identity

    \begin{align} ||g||^2_{L^2(\mathbb{R}^d)} = \sum\limits_{m = 1}^{\infty}\int_{\frac{Y^{'}}{ \varepsilon}}|\mathcal{B}^{\kappa, \varepsilon}_mg(\xi)|^2\,d\xi. \end{align} (7.5)

    3. Plancherel formula For f,g\in L^2(\mathbb{R}^d) , we have

    \begin{align} \int_{\mathbb{R}^d}f(x)\overline{g(x)}\,dx = \sum\limits_{m = 1}^{\infty}\int_{\frac{Y^{'}}{ \varepsilon}}\mathcal{B}^{\kappa, \varepsilon}_mf(\xi)\overline{\mathcal{B}^{\kappa, \varepsilon}_mg(\xi)}\,d\xi. \end{align} (7.6)

    4. Bloch Decomposition in H^{-1}(\mathbb{R}^d) For an element F = u_0(x)+\sum_{j = 1}^N\frac{\partial u_j(x)}{\partial x_j} of H^{-1}(\mathbb{R}^d) , the following limit exists in L^2\left(\frac{Y^{'}}{ \varepsilon}\right) :

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_mF(\xi) = \int_{\mathbb{R}^d}e^{-ix\cdot\xi}\left\{u_0(x)\overline{\phi^{\kappa, \varepsilon}_m(x;\xi)}+i\sum\limits_{j = 1}^N\xi_ju_j(x)\overline{\phi^{\kappa, \varepsilon}_m(x;\xi)}\right\}\,dx\\-\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\sum\limits_{j = 1}^Nu_j(x)\frac{\partial\overline{\phi^{\kappa, \varepsilon}_m}}{\partial x_j}(x;\xi)\,dx. \end{align} (7.7)

    The definition above is independent of the particular representative of F .

    5. Finally, for g\in D(\mathcal{A}^{\kappa, \varepsilon}) ,

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_m(\mathcal{A}^{\kappa, \varepsilon}g)(\xi) = \lambda^{\kappa, \varepsilon}_m(\xi)\mathcal{B}^{\kappa, \varepsilon}_mg(\xi). \end{align} (7.8)

    In order to compute the homogenization limit, we need to know the limit of Bloch Transform of a sequence of functions. The following theorem proves that for a sequence of functions convergent in a suitable way, the first Bloch transform converges to the Fourier transform of the limit.

    Theorem 7.2. Let K\subseteq\mathbb{R}^d be a compact set and (g^ \varepsilon) be a sequence of functions in L^2(\mathbb{R}^d) such that g^ \varepsilon = 0 outside K . Suppose that g^ \varepsilon\rightharpoonup g in L^2(\mathbb{R}^d) -weak for some function g\in L^2(\mathbb{R}^d) . Then it holds that

    in L^2_{\;\mathit{loc}\;}(\mathbb{R}^d_\xi) -weak, where \widehat{g} denotes the Fourier transform of g and denotes the characteristic function of the set { \varepsilon^{-1}U} .

    Proof. In Theorem 4.5, the existence of the set U indepedent of \rho was proved. The function \mathcal{B}_1^{\kappa, \varepsilon} g^ \varepsilon is defined for \xi\in \varepsilon^{-1}Y^{'} . However, we shall treat it as a function on \mathbb{R}^d by extending it outside \varepsilon^{-1}U by zero. We can write

    \begin{align*} \mathcal{B}_1^{\kappa, \varepsilon} g^ \varepsilon(\xi)& = \int_{\mathbb{R}^d} g(x)e^{-ix\cdot\xi}\overline{{\phi}_1^{\kappa, \varepsilon}}(x;0)\,dx\\ &\qquad+\int_{\mathbb{R}^d} g(x)e^{-ix\cdot\xi}\left(\overline{\phi_1^\rho}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) -\overline{\phi^\rho}\left(\frac{x}{ \varepsilon};0\right) \right)dx. \end{align*}

    Now, we need to distinguish between the regimes:

    Case 1. \theta\in [0,\infty) : The first term above converges to the Fourier transform of g on account of the normalization of \phi_1(y;0) whereas the second term goes to zero since it is O( \varepsilon\xi) due to the Lipschitz continuity of the first regularized Bloch eigenfunction which follows from (4.3).

    Case 2. \theta = \infty : In this case, again, the first term converges to the Fourier transform of g on account of the normalization of \phi_1(y;0) . However, for the second term, we make use of the analyticity of \phi_1^{\kappa, \varepsilon} in \varepsilon^{-1}U and the estimates in Theorem 6.5 to conclude that the second term is O( \varepsilon\xi) independent of \rho .

    In this section, we will prove the qualitative homogenization result for the singularly perturbed homogenization problem. There are three regimes according to convergence of \rho = \frac{\kappa}{ \varepsilon} , viz., \rho\to 0 , \rho\to\theta\in(0,\infty) and \rho\to\infty .

    Theorem 8.1. Let \Omega be an arbitrary domain in \mathbb{R}^d and f\in L^2(\Omega) . Let u^ \varepsilon\in H^2(\Omega) be such that u^ \varepsilon converges weakly to u^* in H^1(\Omega) , \kappa\Delta u^ \varepsilon is uniformly bounded in L^2(\Omega) , and

    \begin{align} \mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon = f\,\;\mathit{in}\;\,\; \Omega, \end{align} (8.1)

    where \kappa\to 0 as \varepsilon\to 0 and \lim_{ \varepsilon\to 0}\frac{\kappa}{ \varepsilon} = \theta\in [0,\infty] . Let A^{hom} = (a^*_{kl})_{k,l = 1}^d be as defined in Definition 5.3. Then

    1. For all k = 1,2,\ldots,d , we have the following convergence of fluxes:

    \begin{align} A\left(\frac{x}{ \varepsilon}\right)\nabla u^ \varepsilon(x)\rightharpoonup A^{hom}\nabla u^*(x) \;\mathit{in}\; (L^2(\Omega))^d\;\mathit{-weak}\;. \end{align} (8.2)

    2. The limit u^* satisfies the homogenized equation:

    \begin{align} \mathcal{A}^{hom}u^* = -\nabla\cdot A^{hom}\nabla u^* = f\,\;\mathit{in}\;\,\Omega. \end{align} (8.3)

    Remark 8.2. In the spirit of H-convergence [24], we do not impose any boundary condition on the equation. The H-convergence compactness theorem concerns convergence of sequences on which certain differential constraints have been imposed. In homogenization, the weak convergence of solutions is a consequence of uniform bounds on them, which follow from boundary conditions imposed on the equation. In the theorem quoted above, the uniform boundedness on \kappa\Delta u^ \varepsilon would have followed if appropriate boundary conditions were imposed.

    The proof of Theorem 8.1 is divided into the following steps. We begin by localizing the equation (8.1) which is posed on \Omega , so that it is posed on \mathbb{R}^d . We take the first Bloch transform \mathcal{B}^{\kappa, \varepsilon}_1 of this equation and pass to the limit \kappa, \varepsilon\to 0 . The proof relies on the analyticity of the first Bloch eigenvalue and eigenfunction in a neighborhood of 0\in Y^{'} . The limiting equation is an equation in Fourier space. The homogenized equation is obtained by taking the inverse Fourier transform. We will use the notation a_{kl}^ \varepsilon(x) to denote a_{kl}\left(\frac{x}{ \varepsilon}\right) . Further, we will assume the Einstein convention of summing over repeated indices. The proof has been divided into separate cases for the regimes \rho\to\theta\in(0,\infty) , \rho\to\infty , and \rho = 0 .

    Let \psi_0 be a fixed smooth function supported in a compact set K\subset\mathbb{R}^d . Since u^ \varepsilon satisfies \mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon = f , \psi_0 u^ \varepsilon satisfies

    \begin{align} \mathcal{A}^{\kappa, \varepsilon}(\psi_0 u^ \varepsilon)(x) = \psi_0f(x)+g^ \varepsilon(x)+h^{ \varepsilon}(x)+\sum\limits_{m = 1}^4l^{\kappa, \varepsilon}_m(x)\,\;{\rm{ in }}\;\,\mathbb{R}^d, \end{align} (8.4)

    where

    \begin{align} g^ \varepsilon(x)&:=-\frac{\partial \psi_0}{\partial x_k}(x)a^ \varepsilon_{kl}(x)\frac{\partial u^ \varepsilon}{\partial x_l}(x), \end{align} (8.5)
    \begin{align} h^{ \varepsilon}(x)&:=-\frac{\partial}{\partial x_k}\left(\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x)\right), \end{align} (8.6)
    \begin{align} l^{\kappa, \varepsilon}_1(x)&:=\kappa^2\frac{\partial^4\psi^0}{\partial x_k^4}(x)u^ \varepsilon(x) . \end{align} (8.7)
    \begin{align} l^{\kappa, \varepsilon}_2(x)&:= 4\kappa^2\frac{\partial^3\psi^0}{\partial x_k^3}(x)\frac{\partial u^ \varepsilon}{\partial x_k}(x) . \end{align} (8.8)
    \begin{align} l^{\kappa, \varepsilon}_3(x)&:= 2\kappa^2\frac{\partial^2\psi^0}{\partial x_k^2}(x)\frac{\partial^2u^ \varepsilon}{\partial x_k^2}(x) . \end{align} (8.9)
    \begin{align} l^{\kappa, \varepsilon}_4(x)&:= 4\kappa^2\frac{\partial\psi^0}{\partial x_k}(x)\frac{\partial^3u^ \varepsilon}{\partial x_k^3}(x)+ 4\kappa^2\frac{\partial^2\psi^0}{\partial x_k^2}(x)\frac{\partial^2u^ \varepsilon}{\partial x_k^2}(x) = 4\kappa^2\frac{\partial}{\partial x_k}\left(\frac{\partial\psi_0}{\partial x_k}\frac{\partial^2u^ \varepsilon}{\partial x_k^2}\right) . \end{align} (8.10)

    While the sequence g^ \varepsilon is bounded in L^2(\mathbb{R}^d) , the sequence h^{ \varepsilon} is bounded in H^{-1}(\mathbb{R}^d) . Taking the first Bloch transform of both sides of the equation (8.4), we obtain for \xi\in \varepsilon^{-1}U a.e.

    \begin{align} \lambda^{\kappa, \varepsilon}_1(\xi)\mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon)(\xi) = \mathcal{B}^{\kappa , \varepsilon}_1(\psi_0 f)(\xi)+\mathcal{B}^{\kappa, \varepsilon}_1g^ \varepsilon(\xi)+\mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi)+\sum\limits_{m = 1}^4\mathcal{B}^{\kappa, \varepsilon}_1l^{\kappa, \varepsilon}_m(\xi). \end{align} (8.11)

    We shall now pass to the limit \kappa, \varepsilon\to 0 in the equation (8.11).

    We expand the first Bloch eigenvalue about \eta = 0 in \lambda^{\kappa, \varepsilon}_1(\xi)\mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon) to write

    \begin{align*} \left(\frac{1}{2}\frac{\partial^2\lambda^\rho_1}{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t + O( \varepsilon^2) \right) \mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon). \end{align*}

    The higher order derivatives of \lambda^\rho_1(\eta) are bounded uniformly in \rho (see Remark 6.4). Hence, their contribution is O( \varepsilon^2) . Now, we can pass to the limit \kappa, \varepsilon\to 0 in L^2_{\rm{loc}}(\mathbb{R}^d_\xi) -weak by applying Lemma 7.2 to obtain:

    \begin{align} e_s\cdot A^{hom}e_t{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t\widehat{\psi_o u^*}(\xi). \end{align} (8.12)

    An application of Lemma 7.2 yields the convergence of \mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 f) to (\psi_0 f)^{\bf\widehat{}} in L^2_{\rm{loc}}(\mathbb{R}^d_\xi) -weak.

    The sequence g^ \varepsilon as defined in (8.5) is bounded in L^2(\mathbb{R}^d) and hence has a weakly convergent subsequence with limit g^*\in L^2(\mathbb{R}^d) . This sequence is supported in a fixed set K . Also, note that the sequence \sigma_k^ \varepsilon(x):= a^ \varepsilon_{kl}(x)\frac{\partial u^ \varepsilon}{\partial x_l}(x) is bounded in L^2(\Omega) , hence has a weakly convergent subsequence whose limit is denoted by \sigma^*_k for k = 1,2,\ldots,d . Extend \sigma^*_k by zero outside \Omega and continue to denote the extension by \sigma^*_k . Thus, g^* is given by -\frac{\partial \psi_0}{\partial x_k}\sigma^*_k . Therefore, by Lemma 7.2, we obtain the following convergence in L^2_{\rm{loc}}(\mathbb{R}^d_\xi) -weak:

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\mathcal{B}^{\kappa, \varepsilon}_1g^ \varepsilon(\xi)\rightharpoonup-\left(\frac{\partial \psi_0}{\partial x_k}(x)\sigma^*_k(x)\right)^{\bf\widehat{}}(\xi). \end{align} (8.13)

    We have the following weak convergence in L^2_{\rm{loc}}(\mathbb{R}^d_{\xi}) .

    \begin{align} \lim\limits_{ \varepsilon\to 0}\,\chi_{ \varepsilon^{-1}U}(\xi)\mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi) = -i\xi_ka^*_{kl}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi) \end{align} (8.14)

    We shall prove this in the following steps.

    Step 1. By the definition of the Bloch transform (7.7) for elements of H^{-1}(\mathbb{R}^d) , we have

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi) = -i\xi_k\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x)\overline{\phi^\rho_1\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)}\,dx\\+\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x)\frac{\partial\overline{\phi^{\rho}_1}}{\partial x_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\,dx. \end{align} (8.15)

    Step 2. The first term on RHS of (8.15) is the Bloch transform of the expression -i\xi_k\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x) which converges weakly to -i\xi_k\mathcal{M}_Y(a_{kl})\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right) .

    Step 3. Now, we analyze the second term on RHS of (8.15). To this end, we make use of analyticity of first Bloch eigenfunction with respect to the dual parameter \eta near 0 . We have the following power series expansion in H^1_\sharp(Y) for \phi_1^\rho(\eta) about \eta = 0 :

    \begin{align} \phi_1^\rho(y;\eta) = \phi_1^\rho(y;0)+\eta_s\frac{\partial \phi^\rho_1}{\partial \eta_s}(y;0)+\gamma^\rho(y;\eta). \end{align} (8.16)

    We know that \gamma^\rho(y;0) = 0 and (\partial \gamma^\rho/\partial \eta_s)(y;0) = 0 , therefore, \gamma^\rho(\cdot;\eta) = O(|\eta|^2) in L^\infty(U;H^1_\sharp(Y)) . We also have (\partial\gamma^\rho/\partial y_k)(\cdot;\eta) = O(|\eta|^2) in L^\infty(U;L^2_\sharp(Y)) . These orders are uniform in \rho by Remark 6.4. Now,

    \begin{align} \phi_1^{\kappa, \varepsilon}(x;\xi) = \phi_1^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) = \phi_1^\rho\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon\xi_s\frac{\partial \phi^\rho_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+\gamma^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right). \end{align} (8.17)

    Differentiating the last equation with respect to x_k , we obtain

    \begin{align} \frac{\partial}{\partial x_k}\phi_1^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) = \xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^\rho_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon^{-1}\frac{\partial \gamma^\rho}{\partial y_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right). \end{align} (8.18)

    For \xi belonging to the set \{\xi: \varepsilon\xi\in U{\rm{ and }}|\xi|\leq M\} , we have

    \begin{align} \frac{\partial \gamma^\rho}{\partial y_k}(\cdot; \varepsilon\xi) = O(| \varepsilon\xi|^2) = \varepsilon^2O(|\xi|^2)\leq CM^2 \varepsilon^2. \end{align} (8.19)

    As a consequence,

    \begin{align} \varepsilon^{-2}\frac{\partial \gamma^\rho}{\partial y_k}(x/ \varepsilon; \varepsilon\xi)\in L^\infty_{\rm{loc}}(\mathbb{R}^d_\xi;L^2_\sharp( \varepsilon Y)). \end{align} (8.20)

    The second term on the RHS of (8.15) is given by

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\frac{\partial}{\partial x_k}\left(\overline{\phi^{\rho}_1}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\right)\,dx. \end{align} (8.21)

    Substituting (8.18) in (8.21), we obtain

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\biggl[\xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^{\rho}_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon^{-1}\frac{\partial \gamma^\rho}{\partial y_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\biggr]\,dx. \end{align} (8.22)

    In the last expression, the term involving \gamma^\rho goes to zero as \varepsilon\to 0 in view of (8.19), whereas the other term has the following limit as \rho\to\theta\in(0,\infty) :

    \begin{align} \mathcal{M}_Y\left(a_{kl}(y)\frac{\partial \chi^{\theta}_s}{\partial y_k}(y)\right)\xi_s\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)u^*(x)\,dx. \end{align} (8.23)

    To see this, we write the second term as

    \begin{align*} \int_K& e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^{\rho}_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)\,dx\\ & = \int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\xi_s\frac{\partial \chi^{\rho}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)\,dx\\ & = \int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\xi_s\left(\frac{\partial \chi^{\theta}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)+\left[\frac{\partial \chi^{\rho}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)-\frac{\partial \chi^{\theta}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)\right]\right)\,dx. \end{align*}

    The first term in parantheses goes to (8.23) due to strong convergence of u^ \varepsilon in L^2(K) and weak convergence of a_{kl}\frac{\partial \chi^{\theta}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right) , whereas the expression in the square brackets goes to zero due to Lemma 5.1.

    Step 4. By Theorem 6.1 and Remark 6.2, it follows that

    \begin{align} \mathcal{M}_Y\left(a_{kl}(y)\frac{\partial}{\partial y_k}\left(\frac{\partial\phi^{\theta}_1}{\partial\eta_s}(y;0)\right)\right) = -i(2\pi)^{-d/2}\mathcal{M}_Y\left(a_{kl}(y)\frac{\partial \chi^{\theta}_s}{\partial y_k}(y)\right). \end{align} (8.24)

    Therefore, we have the following convergence in L^2_{\rm{loc}}(\mathbb{R}^d_\xi) -weak:

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)&\mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi)\\ &\rightharpoonup-i\xi_s\biggl\{\mathcal{M}_Y(a_{kl})+\mathcal{M}_Y\left(a_{kl}(y)\frac{\partial \chi^{\theta}_s}{\partial y_k}(y)\right)\biggr\}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi)\\ & = -i\xi_s a_{kl}^{*} \left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi) \end{align} (8.25)

    We shall prove that

    \begin{align} \lim\limits_{ \varepsilon\to 0}\mathcal{B}_1^{\kappa, \varepsilon}l_1^{\kappa, \varepsilon} = 0. \end{align} (8.26)

    Observe that

    \begin{align} \mathcal{B}_1^{\kappa, \varepsilon}l^{\kappa, \varepsilon}_1(\xi) = \kappa^2\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial^4\psi^0}{\partial x_k^4}(x)u^ \varepsilon(x)\overline{\phi^{\kappa, \varepsilon}(x,\xi)}\,dx. \end{align} (8.27)

    The integral is the Bloch transform of \frac{\partial^4\psi^0}{\partial x_k^4}(x)u^ \varepsilon(x) which converges to the Fourier transform of \frac{\partial^4\psi^0}{\partial x_k^4}(x)u^*(x) . However, since \kappa\to 0 , the whole expression goes to zero.

    We shall prove that

    \begin{align} \lim\limits_{ \varepsilon\to 0}\mathcal{B}_1^{\kappa, \varepsilon}l_2^{\kappa, \varepsilon} = 0. \end{align} (8.28)

    Observe that

    \begin{align} \mathcal{B}_1^{\kappa, \varepsilon}l^{\kappa, \varepsilon}_2(\xi) = 4\kappa^2\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial^3\psi^0}{\partial x_k^3}(x)\frac{\partial u^ \varepsilon}{\partial x_k}(x)\overline{\phi^{\kappa, \varepsilon}(x,\xi)}\,dx. \end{align} (8.29)

    The integral is the Bloch transform of \frac{\partial^3\psi^0}{\partial x_k^3}(x)\frac{\partial u^ \varepsilon}{\partial x_k}(x) which converges to the Fourier transform of \frac{\partial^3\psi^0}{\partial x_k^3}(x)\frac{\partial u^*}{\partial x_k}(x) . However, since \kappa\to 0 , the whole expression goes to zero.

    We shall prove that

    \begin{align} \lim\limits_{ \varepsilon\to 0}\mathcal{B}_1^{\kappa, \varepsilon}l_3^{\kappa, \varepsilon} = 0. \end{align} (8.30)

    Observe that

    \begin{align} \mathcal{B}_1^{\kappa, \varepsilon}l^{\kappa, \varepsilon}_3(\xi) = 2\kappa\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial^2\psi^0}{\partial x_k^2}(x)\kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k^2}(x)\overline{\phi^{\kappa, \varepsilon}(x,\xi)}\,dx. \end{align} (8.31)

    The integral is the Bloch transform of \frac{\partial^2\psi^0}{\partial x_k^2}(x)\kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k^2}(x) which converges for a subsequence since \kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k^2}(x) is bounded in L^2(\Omega) (and hence converges weakly in L^2(\Omega) for a subsequence). However, since \kappa\to 0 , the whole expression goes to zero.

    We shall prove that

    \begin{align} \lim\limits_{ \varepsilon\to 0}\mathcal{B}_1^{\kappa, \varepsilon}l_4^{\kappa, \varepsilon} = 0. \end{align} (8.32)

    Observe that

    \begin{align*} l^{\kappa, \varepsilon}_4(x) = 4\kappa\frac{\partial}{\partial x_k}\left(\frac{\partial\psi_0}{\partial x_k}\kappa\frac{\partial^2u^ \varepsilon}{\partial x_k^2}\right) \end{align*}

    belongs to H^{-1}(\mathbb{R}^d) , hence the Bloch trasform for H^{-1}(\mathbb{R}^d) , that is, (7.7) applies. Hence,

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_1l^{\kappa, \varepsilon}_4(\xi) = -4i\kappa\xi_k\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)\kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k}(x)\overline{\phi^\rho_1\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)}\,dx\\+4\kappa\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)\kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k^2}(x)\frac{\partial\overline{\phi^{\rho}_1}}{\partial x_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\,dx. \end{align} (8.33)

    The analysis of the first term is the same as that of \mathcal{B}_1^{\kappa, \varepsilon}l_3^{\kappa, \varepsilon} . For the second term, observe that \kappa\frac{\partial^2 u^ \varepsilon}{\partial x_k^2}(x) is bounded in L^2(\Omega) and \frac{\partial{\phi^{\rho}_1}}{\partial x_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) is bounded in L^2_\sharp( \varepsilon Y) uniformly in \rho (see Remark 6.4). Hence, their product is bounded in L^1(K) . As a result, the integral is a Fourier transform of a sequence bounded in \rho and \varepsilon . However, the presence of \kappa in front of the integral causes the expression to go to zero.

    Finally, passing to the limit in (8.11) as \varepsilon\to 0 by applying equations (8.12), (8.13), (8.14), (8.26), (8.28), (8.30), and (8.32), we get:

    \begin{align} a_{kl}^*\xi_k\xi_l\widehat{\psi_o u^*}(\xi) = \widehat{\psi_0 f}-\left(\frac{\partial \psi_0}{\partial x_k}(x)\sigma^*_k(x)\right)^{\bf\widehat{}}(\xi)-i\xi_ka^*_{kl}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi). \end{align} (8.34)

    In this regime, the convergence proofs for \mathcal{B}_1^{\kappa, \varepsilon}l_1^{\kappa, \varepsilon} , \mathcal{B}_1^{\kappa, \varepsilon}l_2^{\kappa, \varepsilon} , \mathcal{B}_1^{\kappa, \varepsilon}l_3^{\kappa, \varepsilon} , \mathcal{B}_1^{\kappa, \varepsilon}g^{ \varepsilon} , \mathcal{B}_1^{\kappa, \varepsilon}(\psi_0 f) are the same as in the earlier regime. Therefore, we will only look at the remaining convergences.

    We expand the first Bloch eigenvalue about \eta = 0 in \lambda^{\kappa, \varepsilon}_1(\xi)\mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon) to write

    \begin{align*} \left(\frac{1}{2}\frac{\partial^2\lambda^\rho_1}{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t + \frac{ \varepsilon^2}{4!}\partial^{e_s+e_t+e_u+e_v}_0\lambda_1^\rho\xi_s\xi_t\xi_u\xi_v + O( \varepsilon^4) \right) \mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon). \end{align*}

    The fourth order derivative is of order \rho^2 by Theorem 6.5. The derivatives of \lambda^\rho_1(\eta) of order greater than 4 are bounded uniformly in \rho (see Theorem 6.5). Hence, their contribution is O( \varepsilon^4) . Hence, we can write the above as

    \begin{align*} \left(\frac{1}{2}\frac{\partial^2\lambda^\rho_1}{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t + O( \varepsilon^2\rho^2) + O( \varepsilon^4) \right) \mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon)\\ = \left(\frac{1}{2}\frac{\partial^2\lambda^\rho_1}{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t + O(\kappa^2) + O( \varepsilon^4) \right) \mathcal{B}^{\kappa, \varepsilon}_1(\psi_0 u^ \varepsilon) \end{align*}

    Now, we can pass to the limit \kappa, \varepsilon\to 0 in L^2_{\rm{loc}}(\mathbb{R}^d_\xi) -weak by applying Lemma 7.2 to obtain:

    \begin{align*} e_s\cdot A^{hom}e_t{\partial\eta_s\partial\eta_t}(0)\xi_s\xi_t\widehat{\psi_o u^*}(\xi). \end{align*}
    \begin{align*} \lim\limits_{ \varepsilon\to 0}\,\chi_{ \varepsilon^{-1}U}(\xi)\mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi) = -i\xi_ka^*_{kl}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi) \end{align*}

    We shall prove this in the following steps.

    Step 1. As before, we have

    \begin{align} \mathcal{B}^{\kappa, \varepsilon}_1h^{ \varepsilon}(\xi) = -i\xi_k\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x)\overline{\phi^\rho_1\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)}\,dx\\+\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x)\frac{\partial\overline{\phi^{\rho}_1}}{\partial x_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\,dx. \end{align} (8.35)

    Step 2. As before, the first term on RHS of (8.35) is the Bloch transform of the expression -i\xi_k\frac{\partial \psi_0}{\partial x_l}(x)a^{ \varepsilon}_{kl}(x)u^ \varepsilon(x) which converges weakly to

    \begin{align*} -i\xi_k\mathcal{M}_Y(a_{kl})\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right) = -i\xi_k a^*_{kl}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right) \end{align*}

    where the last equality is due to Definition 5.3.

    Step 3. Now we shall prove that the second term on RHS of (8.35) goes to zero. As before, we make use of analyticity of first Bloch eigenfunction with respect to the dual parameter \eta near 0 . We have the following power series expansion in H^1_\sharp(Y) for \phi_1^\rho(\eta) about \eta = 0 :

    \begin{align*} \phi_1^\rho(y;\eta) = \phi_1^\rho(y;0)+\eta_s\frac{\partial \phi^\rho_1}{\partial \eta_s}(y;0)+\gamma^\rho(y;\eta). \end{align*}

    We know that \gamma^\rho(y;0) = 0 and (\partial \gamma^\rho/\partial \eta_s)(y;0) = 0 , therefore, \gamma^\rho(\cdot;\eta) = O(|\eta|^2) in L^\infty(U;H^1_\sharp(Y)) . We also have (\partial\gamma^\rho/\partial y_k)(\cdot;\eta) = O(|\eta|^2) in L^\infty(U;L^2_\sharp(Y)) . These orders are uniform in \rho by Theorem 6.5. Now,

    \begin{align*} \phi_1^{\kappa, \varepsilon}(x;\xi) = \phi_1^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) = \phi_1^\rho\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon\xi_s\frac{\partial \phi^\rho_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+\gamma^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right). \end{align*}

    Differentiating the last equation with respect to x_k , we obtain

    \begin{align} \frac{\partial}{\partial x_k}\phi_1^\rho\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) = \xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^\rho_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon^{-1}\frac{\partial \gamma^\rho}{\partial y_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right). \end{align} (8.36)

    For \xi belonging to the set \{\xi: \varepsilon\xi\in U{\rm{ and }}|\xi|\leq M\} , we have

    \begin{align} \frac{\partial \gamma^\rho}{\partial y_k}(\cdot; \varepsilon\xi) = O(| \varepsilon\xi|^2) = \varepsilon^2O(|\xi|^2)\leq CM^2 \varepsilon^2. \end{align} (8.37)

    As a consequence,

    \begin{align*} \varepsilon^{-2}\frac{\partial \gamma^\rho}{\partial y_k}(x/ \varepsilon; \varepsilon\xi)\in L^\infty_{\rm{loc}}(\mathbb{R}^d_\xi;L^2_\sharp( \varepsilon Y)). \end{align*}

    The second term on the RHS of (8.35) is given by

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\frac{\partial}{\partial x_k}\left(\overline{\phi^{\rho}_1}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\right)\,dx. \end{align} (8.38)

    Substituting (8.36) in (8.38), we obtain

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\biggl[\xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^{\rho}_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon^{-1}\frac{\partial \gamma^\rho}{\partial y_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\biggr]\,dx. \end{align} (8.39)

    In the last expression, the term involving \gamma^\rho goes to zero as \varepsilon\to 0 in view of (8.37).

    The other term also goes to zero as \rho\to\infty due to strong convergence of u^ \varepsilon and the fact that

    \frac{\partial}{\partial x_k}\left(\frac{\partial\phi^{\rho}_1}{\partial\eta_s}(x/ \varepsilon;0)\right) = O\left(\frac{1}{\rho^2}\right),

    as shown in Theorem 6.5.

    The analysis is the same as before, however the uniform-in- \rho boundedness of \frac{\partial{\phi^{\rho}_1}}{\partial x_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right) in L^2_\sharp( \varepsilon Y) is due to Theorem 6.5.

    Hence, for the regime \rho\to\infty , we also recover (8.34).

    In this regime, all the convergence proofs are the same as in the regime \rho\to\theta\in(0,\infty) except for the convergence of \mathcal{B}_1^{\kappa, \varepsilon}h^{ \varepsilon} , which we prove below.

    For this limit, all steps except the third are the same, hence we only explain the part of Step 3 which differs from the regime \rho\to\theta\in(0,\infty) . We begin with the following equation which was earlier labelled as (8.22).

    \begin{align} \chi_{ \varepsilon^{-1}U}(\xi)\int_K e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)a_{kl}\left(\frac{x}{ \varepsilon}\right)u^ \varepsilon(x)\biggl[\xi_s\frac{\partial}{\partial y_k}\frac{\partial \phi^{\rho}_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)+ \varepsilon^{-1}\frac{\partial \gamma^\rho}{\partial y_k}\left(\frac{x}{ \varepsilon}; \varepsilon\xi\right)\biggr]\,dx. \end{align} (8.40)

    In the last expression, the term involving \gamma^\rho goes to zero as \varepsilon\to 0 in view of (8.19), whereas the other term has the following limit as \rho\to 0 :

    \begin{align} \mathcal{M}_Y\left(a_{kl}(y)\frac{\partial \chi^0_s}{\partial y_k}(y)\right)\xi_s\int_{\mathbb{R}^d}e^{-ix\cdot\xi}\frac{\partial\psi_0}{\partial x_l}(x)u^*(x)\,dx. \end{align} (8.41)

    To see this, we write

    \begin{align*} \frac{\partial}{\partial y_k}\frac{\partial \phi^{\rho}_1}{\partial \eta_s}\left(\frac{x}{ \varepsilon};0\right)& = \frac{\partial \chi^{\rho}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)\\ & = \underbrace{\frac{\partial \chi^0_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)}_{I}+\underbrace{\left[\frac{\partial \chi^{\rho}_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)-\frac{\partial \chi^B_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)\right]}_{II}+\underbrace{\left[\frac{\partial \chi^0_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)-\frac{\partial \chi^B_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right)\right]}_{III}. \end{align*}

    The first term I is responsible for (8.41) due to strong convergence of u^ \varepsilon in L^2(K) and weak convergence of a_{kl}\frac{\partial \chi^0_s}{\partial y_k}\left(\frac{x}{ \varepsilon}\right) .

    For the expressions II and III , we make use of Lemma 5.2. Indeed, we obtain II = O(\rho)+O(\varkappa) and III = O(\varkappa) . Hence, their contribution to the limit in (8.40) as \rho\to 0 is O(\varkappa) .

    This completes the modification required for the regime \rho\to 0 . Instead of (8.34), we obtain

    \begin{align*} a_{kl}^*\xi_k\xi_l\widehat{\psi_o u^*}(\xi) = \widehat{\psi_0 f}-\left(\frac{\partial \psi_0}{\partial x_k}(x)\sigma^*_k(x)\right)^{\bf\widehat{}}(\xi)-i\xi_ka^*_{kl}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)^{\bf\widehat{}}(\xi) + O(\varkappa). \end{align*}

    However, since \varkappa>0 is an arbitrary positive number in Lemma 5.2, we also recover (8.34) for the regime \rho\to 0 .

    Taking the inverse Fourier transform in the equation (8.34), we obtain the following:

    \begin{align} (\mathcal{A}^{hom}(\psi_0 u^*)(x)) = \psi_0 f-\frac{\partial \psi_0}{\partial x_k}(x)\sigma^*_k(x)-a^*_{kl}\frac{\partial}{\partial x_k}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right), \end{align} (8.42)

    where the operator \mathcal{A}^{hom} is defined in (8.3). At the same time, calculating using Leibniz rule, we have:

    \begin{align} (\mathcal{A}^{hom}(\psi_0 u^*)(x)) = (\psi_0(x)\mathcal{A}^{hom}u^*(x))&-a^*_{kl}\frac{\partial}{\partial x_k}\left(\frac{\partial \psi_0}{\partial x_l}(x)u^*(x)\right)\\ &-a_{kl}^*\frac{\partial\psi_0}{\partial x_k}(x)\frac{\partial u^*}{\partial x_l}(x) \end{align} (8.43)

    Using equations (8.42) and (8.43), we obtain

    \begin{align} \psi_0(x)\left(\mathcal{A}^{hom}u^*-f\right)(x) = \frac{\partial\psi_0}{\partial x_k}\left[a_{kl}^*\frac{\partial u^*}{\partial x_l}(x)-\sigma_k^*(x)\right]. \end{align} (8.44)

    Let \omega be a unit vector in \mathbb{R}^d , then \psi_0(x)e^{ix\cdot\omega}\in\mathcal{D}(\Omega) . On substituting in the above equation, we get, for all k = 1,2,\ldots,d and for all \psi_0\in\mathcal{D}(\Omega) ,

    \begin{align} \psi_0(x)\left[a_{kl}^*\frac{\partial u^*}{\partial x_l}(x)-\sigma_k^*(x)\right] = 0. \end{align} (8.45)

    Let x_0 be an arbitrary point in \Omega and let \psi_0(x) be equal to 1 near x_0 , then for a small neighbourhood of x_0 :

    \begin{align} \;{\rm{ for }}\; k = 1,2,\ldots,d,\; \left[a_{kl}^*\frac{\partial u^*}{\partial x_l}(x)-\sigma_k^*(x)\right] = 0 \end{align} (8.46)

    However, x_0\in\Omega is arbitrary, so that

    \begin{align} \mathcal{A}^{hom}u^* = f{\rm{ and }}\sigma^*_k(x) = a_{kl}^*\frac{\partial u^*}{\partial x_l}(x). \end{align} (8.47)

    Thus, we have obtained the limit equation in the physical space. This finishes the proof of Theorem 8.1.

    The proof of the qualitative homogenization theorem only requires the first Bloch transform. It is not clear whether the higher Bloch modes make any contribution to the homogenization limit. In this section, we show that they do not. We know that Bloch decomposition is the isomorphism L^2(\mathbb{R}^d)\cong L^2(Y^{'}/ \varepsilon;\ell^2(\mathbb{N})) which is reflected in the inverse identity (7.4). For simplicity, take \Omega = \mathbb{R}^d and consider the equation \mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon = f in \mathbb{R}^d which is equivalent to

    \begin{align*} \mathcal{B}^{\kappa, \varepsilon}_m \mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon(\xi) = \mathcal{B}^{\kappa, \varepsilon}_mf(\xi)\quad\forall m\geq 1,\forall\,\xi\in \varepsilon^{-1}Y^{'}. \end{align*}

    We claim that one can neglect all the equations corresponding to m\geq 2 .

    Proposition 9.1. Let

    v^{\kappa, \varepsilon}(x) = \int_{ \varepsilon^{-1}Y^{'}}\sum\limits_{m = 2}^{\infty}\mathcal{B}^{\kappa, \varepsilon}_m u^ \varepsilon(\xi)\phi_m^{\kappa, \varepsilon}(x;\xi)e^{ix\cdot\xi}\,d\xi,

    then ||v^{\kappa, \varepsilon}||_{L^2(\mathbb{R}^d)}\leq c \varepsilon , where c does not depend on \rho and \varepsilon .

    Proof. Due to boundedness of the sequence (u^ \varepsilon) in H^2(\mathbb{R}^d) , we have

    \begin{align} \int_{\mathbb{R}^d}\mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon\,\overline{u^ \varepsilon}\leq C. \end{align} (9.1)

    However, by Plancherel Theorem (7.6), we have

    \begin{align*} \int_{\mathbb{R}^d} \mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon\, \overline{u^ \varepsilon} = \sum\limits_{m = 1}^{\infty}\int_{ \varepsilon^{-1}Y^{'}}\left(\mathcal{B}^{\kappa, \varepsilon}_m\mathcal{A}^{\kappa, \varepsilon} u^ \varepsilon\right)(\xi)\,\overline{\mathcal{B}^{\kappa, \varepsilon}_mu^ \varepsilon(\xi)}\,d\xi\leq C \end{align*}

    Using (7.8), we have

    \begin{align*} \sum\limits_{m = 1}^{\infty}\int_{ \varepsilon^{-1}Y^{'}}\lambda^{\kappa, \varepsilon}_m(\xi)|{\mathcal{B}^{\kappa, \varepsilon}_mu^ \varepsilon(\xi)}|^2\,d\xi\leq C. \end{align*}

    Now, by Lemma 4.2

    \begin{align} \lambda_m^\rho(\eta)\geq\alpha\lambda_2^N > 0\quad\forall\, m\geq 2\quad\forall\,\eta\in Y^{'}, \end{align} (9.2)

    where \lambda^N_2 is the second eigenvalue of Laplacian on Y with Neumann boundary condition on \partial Y . Since \lambda^{\kappa, \varepsilon}_m(\xi) = \varepsilon^{-2}\lambda^{\rho}_m( \varepsilon\xi) , we obtain

    \begin{align*} \sum\limits_{m = 2}^{\infty}\int_{ \varepsilon^{-1}Y^{'}}|{\mathcal{B}^{\kappa, \varepsilon}_mu^ \varepsilon(\xi)}|^2\,d\xi\leq C \varepsilon^2. \end{align*}

    By Parseval's identity (7.5), the LHS equals ||v^{\kappa, \varepsilon}||^2_{L^2(\mathbb{R}^d)} . This completes the proof.

    The author gratefully acknowledges the careful examination of the editor and the anonymous referees. The author would also like to thank Ayan Roychowdhury for a discussion about formation of shear bands.

    [1] Bieber T, Akdis C, Lauener R, et al. (2016) Global allergy forum and 3rd davos declaration 2015: atopic dermatitis/eczema: challenges and opportunities toward precision medicine. Allergy 71: 588–592. doi: 10.1111/all.12857
    [2] Ring J (2012) Davos declaration: allergy as a global problem. Allergy 67: 141–143. doi: 10.1111/j.1398-9995.2011.02770.x
    [3] Schaub B, Lauener R, Von ME (2006) The many faces of the hygiene hypothesis. J Allergy Clin Immun 117: 969–977. doi: 10.1016/j.jaci.2006.03.003
    [4] Roduit C, Wohlgensinger J, Frei R, et al. (2011) Prenatal animal contact and gene expression of innate immunity receptors at birth are associated with atopic dermatitis. J Allergy Clin Immun 127: 179–185. doi: 10.1016/j.jaci.2010.10.010
    [5] Gilles S, Mariani V, Bryce M, et al. (2009) Pollen allergens do not come alone: pollen associated lipid mediators (PALMS) shift the human immune systems towards a T(H)2-dominated response. Allergy Asthma Cl Im 5: 3–8. doi: 10.1186/1710-1492-5-3
    [6] Wimmer M, Alessandrini F, Gilles S, et al. (2015) Pollen-derived adenosine is a necessary cofactor for ragweed allergy. Allergy 70: 944–954. doi: 10.1111/all.12642
    [7] Traidl-Hoffmann C, Jakob T, Behrendt H (2009) Determinants of allergenicity. J Allergy Clin Immun 123: 558–566. doi: 10.1016/j.jaci.2008.12.003
    [8] Gilles S, Fekete A, Zhang X, et al. (2011) Pollen metabolome analysis reveals adenosine as a major regulator of dendritic cell-primed T(H) cell responses. J Allergy Clin Immun 127: 1–9. doi: 10.1016/j.jaci.2010.11.027
    [9] Papadopoulos NG, Agache I, Bavbek S, et al. (2012) Research needs in allergy: an EAACI position paper, in collaboration with EFA. Clin Transl Allergy 2: 21–45. doi: 10.1186/2045-7022-2-21
    [10] Cecchi L, D'Amato G, Ayres JG, et al. (2010) Projections of the effects of climate change on allergic asthma: the contribution of aerobiology. Allergy 65: 1073–1081.
    [11] Damialis A, Kaimakamis E, Konoglou M, et al. (2017) Estimating the abundance of airborne pollen and fungal spores at variable elevations using an aircraft: how high can they fly? Sci Rep 7: 44535–44545. doi: 10.1038/srep44535
    [12] Buters JT, Kasche A, Weichenmeier I, et al. (2008) Year-to-year variation in release of Bet v 1 allergen from birch pollen: evidence for geographical differences between West and South Germany. Int Arch Allergy Imm 145: 122–130.
    [13] Jochner S, Markevych I, Beck I, et al. (2015) The effects of short-and long-term air pollutants on plant phenology and leaf characteristics. Environ Pollut 206: 382–389. doi: 10.1016/j.envpol.2015.07.040
    [14] Beck I, Jochner S, Gilles S, et al. (2013) High environmental ozone levels lead to enhanced allergenicity of birch pollen. Plos One 8: e80147–e80153. doi: 10.1371/journal.pone.0080147
    [15] Obersteiner A, Gilles S, Frank U, et al. (2016) Pollen-associated microbiome correlates with pollution parameters and the allergenicity of pollen. Plos One 11: e0149545–e0149560. doi: 10.1371/journal.pone.0149545
    [16] von Mutius E (2016) The microbial environment and its influence on asthma prevention in early life. J Allergy Clin Immun 137: 680–689. doi: 10.1016/j.jaci.2015.12.1301
    [17] Xing H, Goedert JJ, Pu A, et al. (2016) Allergy associations with the adult fecal microbiota: analysis of the American gut project. Ebiomedicine 3: 172–179. doi: 10.1016/j.ebiom.2015.11.038
    [18] Kong HH, Andersson B, Clavel T, et al. (2017) Performing skin microbiome research: a method to the madness. J Invest Dermatol 137: 561–568. doi: 10.1016/j.jid.2016.10.033
    [19] Kong HH, Segre JA (2017) The molecular revolution in cutaneous biology: investigating the skin microbiome. J Invest Dermatol 137: e119–e122. doi: 10.1016/j.jid.2016.07.045
    [20] Eyerich K, Pennino D, Scarponi C, et al. (2009) IL-17 in atopic eczema: linking allergen-specific adaptive and microbial-triggered innate immune response. J Allergy Clin Immun 123: 59–66. doi: 10.1016/j.jaci.2008.10.031
    [21] Lehtimäki J, Karkman A, Laatikainen T, et al. (2017) Patterns in the skin microbiota differ in children and teenagers between rural and urban environments. Sci Rep 7: 45651–45661. doi: 10.1038/srep45651
    [22] Peters EM (2016) Stressed skin?-a molecular psychosomatic update on stress-causes and effects in dermatologic diseases. J Dtsch Dermatol Ges 14: 233–252.
    [23] Peters EMJ, Michenko A, Kupfer J, et al. (2014) Mental stress in atopic dermatitis-neuronal plasticity and the cholinergic system are affected in atopic dermatitis and in response to acute experimental mental stress in a randomized controlled pilot study. Plos One 9: 1–17.
    [24] Ring J, Akdis C, Lauener R, et al. (2014) Global allergy forum and second davos declaration 2013 allergy: barriers to cure-challenges and actions to be taken. Allergy 69: 978–982. doi: 10.1111/all.12406
  • This article has been cited by:

    1. Nikolai N. Nefedov, Lutz Recke, A common approach to singular perturbation and homogenization II: Semilinear elliptic systems, 2024, 0022247X, 129099, 10.1016/j.jmaa.2024.129099
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5084) PDF downloads(1052) Cited by(1)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog