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

Asymptotic structure of the spectrum in a Dirichlet-strip with double periodic perforations

  • Received: 01 October 2018 Revised: 01 May 2019
  • Primary: 35B27, 35P05, 47A55, 35J25, 47A10; Secondary: 35P10, 35P15, 47A75

  • We address a spectral problem for the Dirichlet-Laplace operator in a waveguide $ \Pi^ \varepsilon $. $ \Pi^ \varepsilon$ is obtained from repsilon an unbounded two-dimensional strip $ \Pi $ which is periodically perforated by a family of holes, which are also periodically distributed along a line, the so-called "perforation string". We assume that the two periods are different, namely, $ O(1) $ and $ O( \varepsilon) $ respectively, where $ 0< \varepsilon\ll 1 $. We look at the band-gap structure of the spectrum $ \sigma^ \varepsilon $ as $ \varepsilon\to 0 $. We derive asymptotic formulas for the endpoints of the spectral bands and show that $ \sigma^ \varepsilon $ has a large number of short bands of length $ O( \varepsilon) $ which alternate with wide gaps of width $ O(1) $.

    Citation: Sergei A. Nazarov, Rafael Orive-Illera, María-Eugenia Pérez-Martínez. Asymptotic structure of the spectrum in a Dirichlet-strip with double periodic perforations[J]. Networks and Heterogeneous Media, 2019, 14(4): 733-757. doi: 10.3934/nhm.2019029

    Related Papers:

    [1] Sergei A. Nazarov, Rafael Orive-Illera, María-Eugenia Pérez-Martínez . Asymptotic structure of the spectrum in a Dirichlet-strip with double periodic perforations. Networks and Heterogeneous Media, 2019, 14(4): 733-757. doi: 10.3934/nhm.2019029
    [2] Valeria Chiado Piat, Sergey S. Nazarov, Andrey Piatnitski . Steklov problems in perforated domains with a coefficient of indefinite sign. Networks and Heterogeneous Media, 2012, 7(1): 151-178. doi: 10.3934/nhm.2012.7.151
    [3] Luis Caffarelli, Antoine Mellet . Random homogenization of fractional obstacle problems. Networks and Heterogeneous Media, 2008, 3(3): 523-554. doi: 10.3934/nhm.2008.3.523
    [4] Günter Leugering, Sergei A. Nazarov, Jari Taskinen . The band-gap structure of the spectrum in a periodic medium of masonry type. Networks and Heterogeneous Media, 2020, 15(4): 555-580. doi: 10.3934/nhm.2020014
    [5] 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
    [6] Chaoqun Huang, Nung Kwan Yip . Singular perturbation and bifurcation of diffuse transition layers in inhomogeneous media, part I. Networks and Heterogeneous Media, 2013, 8(4): 1009-1034. doi: 10.3934/nhm.2013.8.1009
    [7] Martin Heida . Stochastic homogenization on perforated domains III–General estimates for stationary ergodic random connected Lipschitz domains. Networks and Heterogeneous Media, 2023, 18(4): 1410-1433. doi: 10.3934/nhm.2023062
    [8] Patrizia Donato, Florian Gaveau . Homogenization and correctors for the wave equation in non periodic perforated domains. Networks and Heterogeneous Media, 2008, 3(1): 97-124. doi: 10.3934/nhm.2008.3.97
    [9] 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
    [10] Oleh Krehel, Toyohiko Aiki, Adrian Muntean . Homogenization of a thermo-diffusion system with Smoluchowski interactions. Networks and Heterogeneous Media, 2014, 9(4): 739-762. doi: 10.3934/nhm.2014.9.739
  • We address a spectral problem for the Dirichlet-Laplace operator in a waveguide $ \Pi^ \varepsilon $. $ \Pi^ \varepsilon$ is obtained from repsilon an unbounded two-dimensional strip $ \Pi $ which is periodically perforated by a family of holes, which are also periodically distributed along a line, the so-called "perforation string". We assume that the two periods are different, namely, $ O(1) $ and $ O( \varepsilon) $ respectively, where $ 0< \varepsilon\ll 1 $. We look at the band-gap structure of the spectrum $ \sigma^ \varepsilon $ as $ \varepsilon\to 0 $. We derive asymptotic formulas for the endpoints of the spectral bands and show that $ \sigma^ \varepsilon $ has a large number of short bands of length $ O( \varepsilon) $ which alternate with wide gaps of width $ O(1) $.



    In this paper we consider a spectral problem for the Laplace operator in an unbounded strip $ \Pi \equiv (-\infty, \infty)\times (0, H)\subset \mathbb{R}^2 $ periodically perforated by a family of holes, which are also periodically distributed along a line, the so-called "perforation string". The perforated domain $ \Pi^ \varepsilon $ is obtained by removing the double periodic family of holes $ \overline{\omega^ \varepsilon} $ from the strip $ \Pi $, cf. Figure 1, a), (4)-(6). The diameter of the holes and the distance between them in the string is $ O( \varepsilon) $, while the distance between two perforation strings is $ 1 $. $ \varepsilon\ll 1 $ is a small positive parameter. A Dirichlet condition is prescribed on the whole boundary $ \partial \Pi^ \varepsilon $. We study the band-gap structure of the essential spectrum of the problem as $ \varepsilon\to 0 $.

    Figure 1. 

    a) The perforated strip $ \Pi^ \varepsilon $ is obtained by removing the double periodic family of holes $ \overline{\omega^ \varepsilon} $ from the strip $ \Pi\equiv(-\infty, \infty)\times(0, H) $. The periodicities $ 1 $ and $ \varepsilon H $ come from the width of he periodicity cell $ \varpi^ \varepsilon $ and the distance between two consecutive holes in the perforation string. b) The periodicity cell $ \varpi^ \varepsilon $ is obtained by removing a periodic family of holes of diameter $ O( \varepsilon) $ from $ \varpi^0\equiv (-1/2, 1/2)\times(0, H) $. It contains one perforation string

    .

    We provide asymptotic formulas for the endpoints of the spectral bands and show that these bands collapse asymptotically at the points of the spectrum of the Dirichlet problem in a rectangle obtained by gluing the lateral sides of the periodicity cell. These formulas show that the spectrum has spectral bands of length $ O( \varepsilon) $ that alternate with gaps of width $ O(1) $. In fact, there is a large number of spectral gaps and their number grows indefinitely when $ \varepsilon\to +0 $.

    It should be emphasized that waveguides with periodically perturbed boundaries have been the subject of research in the last decade: let us mention e.g. [34], [21], [22], [2] and [3] and the references therein. However the type of singular perturbation that we study in our paper has never been addressed. We consider a waveguide perforated by a periodic perforation string, which implies using a combination of homogenization methods and spectral perturbation theory.

    As usual in waveguide theory, we first apply the Gelfand transform (cf. [6], [30], [33], [26], [11] and (11)) to convert the original problem, cf. (7), into a family of spectral problems depending on the Floquet-parameter $ \eta\in[-\pi, \pi] $ posed in the periodicity cell $ \varpi^ \varepsilon $ (cf. (13)-(16) and Fig. 1, b). Each one of these problems has a discrete spectrum, cf. (18), which describe the spectrum $ \sigma^ \varepsilon $ as the union of the spectral bands, cf. (20) and (9). One of the main distinguishing features of this paper is that each problem constitutes itself a homogenization problem with one perforation string. As a consequence, in the stretched coordinates, cf. (30), there appears a boundary value problem in an unbounded strip $ \Xi $ which contains the unit hole $ \omega $ (cf. (2), (31)-(33) and Fig. 2).

    Figure 2. 

    The strip $ \Xi $ with the hole $ \omega $. $ \Xi $ is involved with the unit cell for the homogenization problem (13)-(16)

    .

    The above mentioned homogenization spectral problems have different boundary conditions from those considered in the literature (cf. [5], [14] and [16] for an extensive bibliography). Obtaining convergence for their spectra, correcting terms and precise bounds for discrepancies (cf. (10)), as $ \varepsilon\to 0 $, prove essential for our analysis. We use matched asymptotic expansions methods, homogenization theory and basic techniques from the spectral perturbation theory.

    Let

    $ Π={x=(x1,x2):x1R,x2(0,H)} $ (1)

    be a strip of width $ H>0 $. Let $ \omega $ be a domain in the plane $ {\mathbb R}^2 $ which is bounded by a simple closed contour $ \partial \omega $ which, for simplicity, we assume to be of class $ C^\infty $, and that has the compact closure

    $ ¯ω=ωωϖ0, $ (2)

    where $ \varpi^0 $ is a rectangle, the "limit periodicity" cell in $ \Pi $,

    $ ϖ0=(1/2,1/2)×(0,H)Π. $ (3)

    We also introduce the strip $ \Pi^ \varepsilon $ (see Figure 1, a) perforated by the holes

    $ ωε(j,k)={x:ε1(x1j,x2εkH)ω}with jZ,k{0,,N1}, $ (4)

    where $ \varepsilon = 1/N $ is a small positive parameter, and $ N\in {\mathbb N} $ is a big natural number that we will send to $ \infty $. The period of the perforation along the $ x_1 $-axis in the domain

    $ Πε=ΠjZN1k=0¯ωε(j,k) $ (5)

    is made equal to 1 by rescaling, and similarly, the period is made equal to $ \varepsilon H $ in the $ x_2 $-direction. The periodicity cell in $ \Pi^ \varepsilon $ takes the form

    $ \varpi^ \varepsilon = \varpi^0 \setminus\bigcup\limits_{k = 0}^{N-1}\overline{ \omega^ \varepsilon(0, k)}, $

    (see b) in Figure 1). For brevity, we shall denote by $ \omega^ \varepsilon $ the union of all the holes in (4), namely,

    $ ωε=jZN1k=0ωε(j,k), $ (6)

    while $ \omega $ is referred to as the "unit hole", cf. (2).

    In the domain (5) we consider the Dirichlet spectral problem

    $ {Δuε(x)=λεuε(x),xΠε,uε(x)=0,xΠε. $ (7)

    The variational formulation of problem (7) refers to the integral identity

    $ (uε,v)Πε=λε(uε,v)ΠεvH10(Πε), $ (8)

    where $ \left( \cdot, \cdot\right)_{\Pi^ \varepsilon} $ is the scalar product in the space $ L^2(\Pi^ \varepsilon) $, and $ H^1_0(\Pi^ \varepsilon) $ denotes the completion, in the topology of $ H^1(\Pi^ \varepsilon) $, of the space of the infinitely differentiable functions which vanish on $ \partial \Pi^ \varepsilon $ and have a compact support in $ \overline{\Pi^ \varepsilon} $. Since the bi-linear form on the left of (8) is positive, symmetric and closed in $ H_0^1(\Pi^ \varepsilon) $, the problem (8) is associated with a positive self-adjoint unbounded operator $ {\mathcal A}^ \varepsilon $ in $ L^2(\Pi^ \varepsilon) $ with domain $ H^1_0(\Pi^ \varepsilon)\cap H^2(\Pi^ \varepsilon) $ (see Ch. 10 in [1]).

    Problem (7) gets a positive cutoff value $ \lambda^ \varepsilon_\dagger $ and, therefore, its spectrum $ \sigma^ \varepsilon \subset [\lambda^ \varepsilon_\dagger, \infty) $ (cf. (20) and Remark 5). It is known, see e.g. [30], [33], [11] and [26], that $ \sigma^ \varepsilon $ has the band-gap structure

    $ σε=nNBεn, $ (9)

    where $ B^ \varepsilon_n $ are closed connected bounded segments in the real positive axis. The segments $ B^ \varepsilon_n $ and $ B^ \varepsilon_{n+1} $ may intersect but also they can be disjoint so that a spectral gap becomes open between them. Recall that a spectral gap is a non empty interval which is free of the spectrum but has both endpoints in the spectrum.

    In Section 2 we address the setting of the Floquet parametric family of problems (13)-(16), obtained by applying the Gelfand transform (11) to the original problem (7). They are homogenization spectral problems in a perforated domain, the periodicity cell $ \varpi^ \varepsilon $, with quasi-periodicity conditions (15)-(16) on the lateral sides of $ \varpi^ \varepsilon $. Obviously, each problem of the parametric family (13)-(16) depends on the Floquet-parameter $ \eta $, cf. (11), (19) and (20). For a fixed $ \eta\in[-\pi, \pi] $, the problem has the discrete spectrum $ \Lambda_i^ \varepsilon(\eta), \, i = 1, 2, \cdots $, cf. (18). Section 2.2 contains a first approach to the eigenpairs (i.e., eigenvalues and eigenfunctions) of this problem via the homogenized problem, cf. (27). To get this homogenized problem, we use the energy method combined with techniques from the spectral perturbation theory. We show that its eigenvalues $ \Lambda_i^0, \, i = 1, 2, \cdots $ do not depend on $ \eta $, since they constitute the spectrum of the Dirichlet problem in $ \upsilon = (0, 1)\times(0, H) $, cf. (24). In particular, Theorem 2.1 shows that

    $ \Lambda_i^ \varepsilon(\eta)\to \Lambda_i^0 \, \quad \mbox{ as }\quad \varepsilon\to 0, \quad \forall \eta \in [-\pi, \pi], \quad i = 1, 2, \cdots. $

    However, this result does not give information on the spectral gaps.

    Using the method of matched asymptotic expansions for the eigenfunctions of the homogenization problems (cf. Section 4) we are led to the unit cell boundary value problem (31)-(33), the so-called local problem, that is, a problem to describe the boundary layer phenomenon. Section 3 is devoted to the study of this stationary problem for the Laplace operator, which is independent of $ \eta $ and it is posed in an unbounded strip $ \Xi $ which contains the unit hole $ \omega $. Its two solutions, with a polynomial growth at the infinity, play an important role when determining correctors for the eigenvalues $ \Lambda_i^ \varepsilon(\eta) $, $ i = 1, 2, \cdots $. Further specifying, the definition and the properties of the so-called polarization matrix $ p(\Xi) $, which depend on the "Dirichlet hole" $ \omega $, cf. (38) and Section 3.1, are related with the first term of the Fourier expansion of certain solutions of the unit cell problem (cf. (39) and (42)). The correctors $ \varepsilon\Lambda_i^1(\eta) $ depend on the polarization matrix and the eigenfunctions of the homogenized problem, and we prove that for sufficiently small $ \varepsilon $,

    $ |Λεi(η)Λ0iεΛ1i(η)|ciε3/2, $ (10)

    with some $ c_i>0 $ independent of $ \eta $. These bounds are obtained in Section 5, see Theorems 5.1 and 5.2 depending on the multiplicity of the eigenvalues of (24). $ \Lambda^{1}_{i}(\eta) $ is a well determined function of $ \eta $ (see formulas (61), (62), (68), (69), (71) and Remarks 3 and 4); it is identified by means of matched asymptotic expansions in Section 4.

    As a consequence, we deduce that the bands $ B_i^ \varepsilon = \{\Lambda_{i}^ \varepsilon(\eta), \, \eta \in [-\pi, \pi]\} $ are contained in intervals

    $ \left [\Lambda^{0}_{i}+ \varepsilon B^i_- -c_i \varepsilon^{3/2} \, , \, \Lambda^{0}_{i}+ \varepsilon B^i_+ + c_i \varepsilon^{3/2}\right], $

    of length $ O( \varepsilon) $, where $ B^i_- , \, B^i_+ $ are also well determined values for each eigenvalue $ \Lambda^{0}_{i} $ of (24) (cf. Corollaries 5.1 and 5.2 depending on the multiplicity). All of this together gives that for each $ i $ such that $ \Lambda_i^0 <\Lambda_{i+1}^0 $, cf. (23), the spectrum $ \sigma^ \varepsilon $ opens a gap of width $ O(1) $ between the corresponding spectral bands $ B^ \varepsilon_{i } $ and $ B^ \varepsilon_{i +1} $.

    Dealing with the precise length of the band, we note that the results rely on the fact that the elements of the antidiagonal of the polarization matrix do not vanish (cf. (70)-(75)), but this is a generic property for many geometries of the unit hole $ \omega $ (see, e.g., (47) and (51)). Also note, that for simplicity, we have considered that $ \omega $ has a smooth boundary but most of the results hold in the case where $ \omega $ has a Lipschitz boundary or even when $ \omega $ is a vertical crack, cf. Section 3.1.

    Summarizing, Section 2 addresses some asymptotics for the spectrum of the Floquet-parameter family of spectral problems; Section 3 considers the unit cell problem; Section 4 deals with the asymptotic expansions; in Section 5.1, we formulate the main asymptotic results of the paper, while the proofs are performed in Section 5.2.

    In this section, we deal with the setting of the Floquet-parameter dependent spectral problems and the limit behavior of their spectra, cf. Sections 2.1 and 2.2, respectively.

    The Floquet-Bloch-Gelfand transform (FBG-transform, in short)

    $ uε(x)Uε(x;η)=12πnZeinηuε(x1+n,x2), $ (11)

    see [6] and, e.g., [30], [33], [11], [26] and [4], converts problem (7) into a $ \eta $-parametric family of spectral problems in the periodicity cell

    $ ϖε={xΠε:|x1|<1/2} $ (12)

    see Figure 1, b. Note that $ x\in \Pi^ \varepsilon $ on the left of (11), while $ x\in \varpi^ \varepsilon $ on the right. For each $ \eta\in [-\pi, \pi] $, the spectral problem of the family is defined by the equations

    $ ΔUε(x;η)=Λε(η)Uε(x;η),xϖε, $ (13)
    $ Uε(x;η)=0,xΓε, $ (14)
    $ Uε(1/2,x2;η)=eiηUε(1/2,x2;η),x2(0,H), $ (15)
    $ Uεx1(12,x2;η)=eiηUεx1(12,x2;η),x2(0,H), $ (16)

    where $ \Gamma^ \varepsilon = {\partial \varpi^ \varepsilon} \cap \partial \Pi^ \varepsilon $, $ \eta $ is the dual variable, i.e., the Floquet-parameter, while $ \Lambda^ \varepsilon(\eta) $ and $ U^ \varepsilon(\cdot;\eta) $ denote the spectral parameter and an eigenfunction, respectively. If no confusion arises, they can be denoted by $ \Lambda^ \varepsilon $ and $ U^ \varepsilon $, respectively. Conditions (15)-(16) are the quasi-periodicity conditions on the lateral sides $ \{\pm \frac{1}{2}\}\times(0, H) $ of $ \varpi^ \varepsilon $.

    The variational formulation of the spectral problem (13)-(16) reads:

    $ (Uε,V)ϖε=Λε(Uε,V)ϖεVH1,ηper(ϖε;Γε), $ (17)

    where $ H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $ is a subspace of $ H^1(\varpi^ \varepsilon) $ of functions which satisfy the quasi-periodicity condition (15) and vanish on $ \Gamma^ \varepsilon $. In view of the compact embedding $ H^1(\varpi^ \varepsilon)\subset L^2(\varpi^ \varepsilon) $, the positive, self-adjoint operator $ {\mathcal A}^ \varepsilon(\eta) $ associated with the problem (17) has the discrete spectrum constituting the monotone unbounded sequence of eigenvalues

    $ 0<Λε1(η)Λε2(η)Λεm(η), $ (18)

    which are repeated according to their multiplicities (see Ch. 10 in [1] and Ch. 13 in [30]). The eigenfunctions are assumed to form an orthonormal basis in $ L^2(\varpi^ \varepsilon) $.

    The function

    $ η[π,π]Λεm(η) $ (19)

    is continuous and $ 2\pi $-periodic (see, e.g., Ch. 7 of [9]). Consequently, the sets

    $ Bεm={Λεm(η):η[π,π]} $ (20)

    are closed, connected and bounded intervals of the real positive axis $ \overline{ {\mathbb R}_+} $. Results (9) and (20) for the spectrum of the operator $ {\mathcal A}^ \varepsilon(\eta) $ and the boundary value problem (7) are well-known in the framework of the FBG-theory (see the above references). As a consequence of our results, we show that in our problem, depending on the geometry of the unit hole, and for certain lower frequency range of the spectrum, the spectral band (20) does not reduce to a point (cf. (72), (47), (70) and (74)).

    A first approach to the asymptotics for eigenpairs of (13)-(16) is given by the following convergence result, that we show adapting standard techniques in homogenization and spectral perturbation theory: see, e.g., Ch. 3 in [27] for a general framework and [14] for its application to spectral problems in perforated domains with different boundary conditions. Let us recall $ \varpi^0 $ which coincides with $ \varpi^ \varepsilon $ at $ \varepsilon = 0 $ (cf. (12), and (3)) and contains the perforation string

    $ ωε(0,0),,ωε(0,N1)ϖ0. $ (21)

    Theorem 2.1. Let the spectral problem (13)-(16) and the sequence of eigenvalues (18). Then, for any $ \eta\in[-\pi, \pi] $, we have the convergence

    $ Λεm(η)Λ0m,as ε0, $ (22)

    where

    $ 0<Λ01<Λ02Λ0m,as m, $ (23)

    are the eigenvalues, repeated according to their multiplicities, of the Dirichlet problem

    $ ΔU0(x)=Λ0U0(x),xυ,υ(0,1)×(0,H)U0(x)=0,xυ. $ (24)

    Proof. First, for each fixed $ m $, we show that there are two constants $ C, C_m $ such that

    $ 0<CΛεm(η)Cmη[π,π]. $ (25)

    To obtain the lower bound in (25), it suffices to consider (17) for the eigenpair $ (\Lambda^\varepsilon, \, U^ \varepsilon) $ with $ \Lambda^\varepsilon\equiv \Lambda_1^\varepsilon(\eta) $ and apply the the Poincaré inequality in $ H^1(\varpi^0) $ once that $ U^ \varepsilon $ is extended by zero in $ \overline \omega^ \varepsilon $. To get $ C_m $ in (25) we use the minimax principle,

    $ Λεm(η)=minEεmH1,ηper(ϖε;Γε)maxVEεm,V0(V,V)ϖε(V,V)ϖε, $

    where the minimum is computed over the set of subspaces $ E_m^{ \varepsilon} $ of $ H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $ with dimension $ m $. Indeed, let us take a particular $ E_m^{ \varepsilon} $ that we construct as follows. Consider the eigenfunctions corresponding to the $ m $ first eigenvalues of the mixed eigenvalue problem in the rectangle $ (1/4, 1/2)\times (0, H) $, with Neumann condition on the part of the boundary $ \{1/2\}\times (0, H) $, and Dirichlet condition on the rest of the boundary. Extend these eigenfunctions by zero for $ x \in [0, 1/4]\times (0, H) $, and by symmetry for $ x \in [-1/2, 0]\times (0, H) $. Finally, multiplying these eigenfunctions by $ e^{i\eta x_1} $ gives $ E_m^{ \varepsilon} $ and the rigth hand side of (25).

    Hence, for each $ \eta $ and $ m $, we can extract a subsequence, still denoted by $ \varepsilon $ such that

    $ Λεm(η)Λ0m(η),Uεm(;η)U0m(;η) in H1(ϖ0)weak, as ε0, $ (26)

    for a certain positive $ \Lambda^0_m (\eta) $ and a certain function $ U^0_m(\cdot;\eta) \in H^{1, \eta}_{per} (\varpi^0) $, both of which, in principle, can depend on $ \eta $. Obviously, $ U_m^0(\cdot;\eta) $ vanish on the lower and upper bases of $ \varpi^0 $. Also, we use the Poincaré inequality in $ \varpi^0\supset\omega $, cf. (3),

    $ \Vert U; L^2(\varpi^0\setminus {\overline \omega})\Vert \leq C \Vert \nabla U; L ^2 (\varpi^0\setminus {\overline \omega})\Vert \quad \forall U\in H^1(\varpi^0\setminus {\overline \omega}), \quad U = 0 \mbox{ on } \partial \omega, $

    and we deduce

    $ \varepsilon^{-1} \Vert U_m^ \varepsilon (\cdot;\eta); L^2( \{\vert x_1\vert \leq { \varepsilon}/{2}\} \cap \varpi^0 ) \Vert^2 \leq C \varepsilon \Vert \nabla U_m^ \varepsilon (\cdot;\eta) ; L^2( \{\vert x_1\vert \leq { \varepsilon}/{2}\} \cap \varpi^0 ) \Vert^2 . $

    Now, taking limits as $ \varepsilon\to 0 $, we get $ U_m^0(\cdot, \eta) = 0 $ on $ \{0\}\times(0, H) $ (cf., e.g., [16] and (25)). Hence, we identify ($ \Lambda_m^0(\eta) $, $ U^0_m(\cdot;\eta) $) with an eigenpair of the following problem:

    $ ΔU0m(x;η)=Λ0m(η)U0m(x;η),x1{(1/2,0)(0,1/2)},x2(0,H),U0m(x;η)=0 for x2{0,H},x1(1/2,1/2) and x1=0,x2(0,H),U0m(1/2,x2;η)=eiηU0m(1/2,x2;η),x2(0,H),U0mx1(1/2,x2;η)=eiηU0mx1(1/2,x2;η),x2(0,H), $ (27)

    where the differential equation has been obtained by taking limits in the variational formulation (17) for $ V\in \mathcal{ C}_0^\infty ((-1/2, 0)\times (0, H)) $ and for $ V\in \mathcal{ C}_0^\infty ((0, 1/2)\times (0, H)) $.

    Now, from the orthonormality of $ U_m^ \varepsilon(\cdot;\eta) $ in $ L^2(\varpi^ \varepsilon) $, we get the orthonormality of $ U_m^0(\cdot, \eta) $ in $ L^2(\varpi^0) $. Also, an argument of diagonalization (cf., e.g., Ch. 3 in [27]) shows the convergence of the whole sequence of eigenvalues (18) towards those of (27) with conservation of the multiplicity, and that the set $ \{U_m^0(\cdot; \eta)\}_{m = 1}^\infty $ forms a basis of $ L^2(\varpi^0) $.

    In addition, extending by $ \eta $-quasiperiodicity the eigenfunctions $ U^0_m(\cdot;\eta) $,

    $ u0m(x;η)={U0m(x;η),x1(0,1/2),eiηU0m(x11,x2;η),x1(1/2,1), $ (28)

    we obtain a smooth function in the rectangle $ \upsilon $, and moreover that the pair ($ \Lambda_m^0(\eta) $, $ U^0_m(\cdot, \eta) $) satisfies (24). In addition, the orthogonality of $ \{U^0_m(\cdot; \eta)\}_{m = 1}^\infty $ in $ L^2(\varpi^0) $ implies that the extended functions $ \{u^0_m(\cdot; \eta)\}_{m = 1}^\infty $ in (28) form an orthogonal basis in $ L^2(\upsilon) $, cf. also (55), and we have proved that $ \Lambda_m^0(\eta) $ coincides with $ \Lambda_m^0 $ in the sequence (23) for any $ \eta\in [-\pi, \pi] $. Consequently, the result of the theorem holds.

    Remark 1. Note that the eigenpairs of (24) can be computed explicitly

    $ Λ0np=π2(n2+p2H2),U0np(x)=2Hsin(nπx1)sin(pπx2/H),p,nN. $ (29)

    The eigenvalues $ \Lambda^0_{n p} $ are numerated with two indexes and must be reordered in the sequence (23); the corresponding eigenfunctions $ U^0_{n p} $ are normalized in $ L^2(\upsilon) $. Also, we note that if $ H^2 $ is an irrational number all the eigenvalues are simple.

    In this section, we study the properties of certain solutions of the boundary value problem in the unbounded strip $ \Xi $, cf. (31)-(33) and Figure 2. This problem, the so-called unit cell problem, is involved with the homogenization problem (13)-(16) and the periodical distribution of the openings in the periodicity cell $ \varpi^ \varepsilon $, but it remains independent of the Floquet-parameter.

    In order to obtain a corrector for the approach to the eigenpairs of (13)-(16) given by Theorem 2.1, we introduce the stretched coordinates

    $ ξ=(ξ1,ξ2)=ε1(x1,x2εkH). $ (30)

    which transforms each opening of the string $ \omega^\varepsilon(0, k) $ into the unit opening $ \omega $. Then, we proceed as usual in two-scale homogenization when boundary layers arise (cf., e.g. [28], [18], [32] and [24]): assuming a periodic dependence of the eigenfunctions on the $ \xi_2 $-variable, cf. (34), we make the change (30) in (13)-(16), and take into account (22), to arrive at the unit cell problem. This problem consists of the Laplace equation

    $ ΔξW(ξ)=0,ξΞ, $ (31)

    with the periodicity conditions

    $ W(ξ1,H)=W(ξ1,0),Wξ2(ξ1,H)=Wξ2(ξ1,0),ξ1R, $ (32)

    and the Dirichlet condition on the boundary of the hole $ \overline{\omega} $

    $ W(ξ)=0,ξω. $ (33)

    Regarding (31)-(33), it should be noted that, for any $ \Lambda^ \varepsilon \leq C $, we have

    $ \Delta_x+\Lambda^ \varepsilon = \varepsilon^{-2} (\Delta_\xi + \varepsilon^2 \Lambda^ \varepsilon), $

    and $ \varepsilon^2\Lambda^ \varepsilon \leq C \varepsilon^2 $ while the main part $ \Delta_\xi $ is involved in (31). Also, the boundary condition (33) is directly inherited from (14), while the periodicity conditions (32) have no relation to the original quasi-periodicity conditions (15)-(16), but we need them to support the standard asymptotic ansatz

    $ w(x2)W(ε1x), $ (34)

    for the boundary layer. Here, $ w $ is a sufficiently smooth function in $ x_2\in[0, H] $ and $ W $ is $ H $-periodic in $ \xi_2 = \varepsilon^{-1}x_2 $.

    It is worth recalling that, according to the general theory of elliptic problems in domains with cylindrical outlets to infinity, cf., e.g., Ch. 5 in [26], problem (31)-(33) has just two solutions with a linear polynomial growth as $ \xi_1\to \pm \infty $. Here, we search for these two solutions $ W^\pm(\xi) $ by setting $ \pm 1 $ for the constants accompanying $ \xi_1 $ (cf. Proposition 1). In order to do it, let us consider a fixed positive $ R $ such that

    $ ¯ω(R,R)×(0,H) $ (35)

    and define the cut-off functions $ \chi_\pm\in C^\infty( {\mathbb R}) $ as follows

    $ χ±(y)={1, for ±y>2R,0, for ±y<R, $ (36)

    where the subindex $ \pm $ represent the support in $ \pm \xi_1\in [0, \infty) $.

    Proposition 3.1. There are two normalized solutions of (31)-(33) in the form

    $ W±(ξ)=±χ±(ξ1)ξ1+τ=±χτ(ξ1)pτ±+˜W±(ξ),ξΞ, $ (37)

    where the remainder $ \widetilde{W}^\pm(\xi) $ gets the exponential decay rate $ O(e^{-|\xi_1|2\pi/H}) $, and the coefficients $ p_{\tau\pm}\equiv p_{\tau\pm}{ (\Xi)} $, with $ \tau = \pm $, which are independent of $ R $ and compose a $ 2\times 2 $-polarization matrix

    $ p(Ξ)=(p++(Ξ)p+(Ξ)p+(Ξ)p(Ξ)). $ (38)

    Proof. The existence of two linearly independent normalized solutions $ W^\pm $ of (31)-(33) with a linear polynomial behavior $ \pm\xi_1+p_{\pm\pm} $, as $ \pm \xi_1 \to \infty $, is a consequence of the Kondratiev theory [10] (cf. Ch. 5 in [26] and Sect. 3 [20]). Each solution has a linear growth in one direction and stabilizes towards a constant $ p_{\mp\pm} $ in the other direction. In addition, it lives in an exponential weighted Sobolev space which guarantees that, substracting the linear part, the remaining functions have a gradient in $ (L^2(\Xi))^2 $.

    Let us consider the functions

    $ ˆW±(ξ)=W±(ξ)χ±(ξ1)ξ1, $ (39)

    which, obviously, satisfy (32), (33) and

    $ ΔξˆW±(ξ)=F±(ξ),ξΞ, $ (40)

    with $ F^\pm(\xi) = F^\pm(\xi_1) = \pm \Delta(\chi_\pm(\xi_1)\xi_1) = \pm (\partial_{\xi_1}^2\chi_\pm\, \xi_1 + 2 \partial_{\xi_1}\chi_\pm) $. By construction, $ F^\pm $ has a compact support in $ \pm \xi_1\in [ R, 2R] $.

    Let $ C^\infty_{c\, per}(\overline{\Xi}) $ be the space of the infinitely differentiable $ H $-periodic functions, vanishing on $ \partial \omega $, with compact support in $ \overline{\Xi} $. Let us denote by $ \mathcal{H} $ the completion of $ C^\infty_{c\, per}(\overline{\Xi}) $ in the norm

    $ \|W, \mathcal{H}\| = \|\nabla_y W ;L^2(\Xi)\|. $

    The variational formulation of (40), (32) and (33) reads: to find $ \widehat W^\pm \in \mathcal{H} $ satisfying the integral identity

    $ (yˆW±,yV)Ξ=(F±,V)ΞVH. $ (41)

    Since $ supp(F^\pm) $ is compact, we can apply the Poincaré inequality to the elements of $ \{V\in H^1([-2R, 2R])\times(0, H)): \, V\vert_{\partial \omega} = 0\} $, to derive that the right hand side of (41) defines a linear continuous functional on $ \mathcal{H} $. In addition, the left-hand side of the integral identity (41) implies a norm in the Hilbert space $ { \mathcal{H }} $, and consequently, the Riesz representation theorem assures that the problem (41) has a unique solution $ \widehat W\in { \mathcal{H }} $ satisfying (41).

    In addition, since for each $ \tau $, $ \tau = \pm $, function $ \widehat W^\tau (\xi) $ in (39) is harmonic for $ \vert \xi_1 \vert>2R $ with gradient in $ L^2 (( -\infty, -2R)\times(0, H))\cap L^2 ((2R, +\infty)\times(0, H)) $, the Fourier method (cf., e.g. [13] and [26]) ensures that

    $ \widehat W^\tau (\xi) = c^\tau_\pm+ O(e^{-(\pm\xi_1) 2\pi/H}) \mbox{ as } \pm\xi_1\to +\infty , $

    where the constants $ c^\tau_\pm $ are defined by

    $ cτ±=limT1HH0ˆWτ(±T,ξ2)dξ2=limT1HH0(Wτ(±T,ξ2)τδτ,±T)dξ2. $ (42)

    Obviously, $ c^+_{\pm} $ ($ c^-_\pm $ respectively) are independent of $ R $ and they provide all the constants appearing in (37); namely, $ c^\tau_\pm = p_{\tau{\pm }}{ (\Xi)} $. Hence, the result of the proposition holds.

    In this section, we detect certain properties of the matrix $ p(\Xi) $. This matrix represents an integral characteristics of the "Dirichlet hole" $ \overline{\omega} $ in the strip $ \Pi $. Its definition is quite analogous to the classical polarization tensor in the exterior Dirichlet problem, see Appendix G in [29]. Let us refer to [23] for further properties of matrix $ p(\Xi) $ as well as for examples on its dependence on the shape and dimensions of the hole.

    Proposition 3.2. The matrix $ p(\Xi)+ R\, {\mathbb I} $ is symmetric and positive, where $ {\mathbb I} $ stands for the $ 2\times 2 $ unit matrix and $ R $ given in (35).

    Proof. We represent (37) in the form

    $ W±(ξ)=W±0(ξ)+{±ξ1R,±ξ1>R,0,±ξ1<R. $ (43)

    The function $ W^\pm_0 $ still satisfies the periodicity condition of (32) and the homogeneous Dirichlet condition (33) but remains harmonic in $ \Xi\setminus\Upsilon^\pm(R) $, $ \Upsilon^\pm(R) = \{\xi\in\Xi: \: \pm \xi_1 = R\} $, and its derivative has a jump on the segment $ \Upsilon^\pm(R) $, namely

    $ [W_0^\pm]_\pm(\xi_2) = 0, \quad \left[\frac{\partial W_0^\pm}{\partial|\xi_1|}\right]_\pm(\xi_2) = -1, \quad \xi_2\in (0, H), $

    where $ [\phi]_\pm(\xi_2) = \phi(\pm R \pm 0, \xi_2)-\phi(\pm R \mp 0, \xi_2) $.

    In what follows, we write the equations for $ \tau = \pm $. Since $ \Delta W_0^\pm = 0 $, we multiply it with $ W_0^\tau $ and apply the Green formula in $ (\Xi\setminus\Upsilon^\pm(R)) \cap \{|\xi_1| <T\} $. Finally, we send $ T $ to $ +\infty $ and get

    $ H0Wτ0(±R,ξ2)dξ2=H0Wτ0(±R,ξ2)[W±0|ξ1|]±(ξ2)dξ2=(ξWτ0,ξW±0)Ξ. $ (44)

    On the other hand, on account of (43) and the definition of $ W^\tau $, we have

    $ W^\tau_0(\pm R , \xi_2) = W^\tau(\pm R , \xi_2) \quad \mbox { and } \quad \left[\frac{\partial W^\tau}{\partial|\xi_1|}\right]_\pm(\xi_2) = 0. $

    Consequently, we can write

    $ H0Wτ0(±R,ξ2)dξ2=H0Wτ(±R,ξ2)[W±0|ξ1|]±(ξ2)dξ2=H0(Wτ(±R,ξ2)[W±0|ξ1|]±(ξ2)W±0(±R,ξ2)[Wτ|ξ1|]±(ξ2))dξ2, $

    and using again the Green formula for $ W^\tau $ and $ W_0^\pm $, in a similar way to (44) we get

    $ H0Wτ0(±R,ξ2)dξ2=+limTH0(Wτ(τT,ξ2)W±0|ξ1|(τT,ξ2)W±0(τT,ξ2)Wτ|ξ1|(τT,ξ2))dξ2=H(pτ±(Ξ)+δτ,±R). $ (45)

    Here, we have used the following facts: $ \partial/\partial|\xi_1| $ is the outward normal derivative at the end of the truncated domain $ \{\xi\in{\Xi}:\, |\xi_1|<R\} $, the function $ W^\tau_0 $ is smooth near $ \Upsilon^\pm(R) $, the derivative $ \partial W_0^\pm/\partial|\xi_1| $ decays exponentially and, according to (37) and (43), the function $ W^\pm_0 $ admits the representation when $ \pm \xi_1>2R $ (cf. (37))

    $ W^\pm_0(\xi) = \chi_\pm(\xi_1)\left(p_{\pm\pm}+R\right) + \chi_\mp(\xi_1)p_{\mp\pm}+\widetilde{W}^\pm(\xi). $

    Considering (44) and (45) we have shown the equality for the Gram matrix

    $ \left(\nabla_\xi W^\tau_0, \nabla_\xi W^\pm_0\right)_\Xi = H\left(p_{\tau\pm}(\Xi) + \delta_{\tau, \pm}R\right), $

    which gives the symmetry and the positiveness of the matrix $ p(\Xi)+ R\, {\mathbb I}\, $.

    Let us note that our results above apply for Lipschitz domains or even cracks as it was pointed out in Section 2.1. Now, we get the following results in Propositions 3.3 and 3.4 depending on whether $ \omega $ is an open domain in the plane with a positive measure $ mes_2(\omega) $, or it is a crack with $ mes_2(\omega) = 0 $.

    Proposition 3.3. Let $ \omega $ be such that $ mes_2(\omega) >0 $. Then, the coefficients of the polarization matrix $ p(\Xi) $ satisfy

    $ H\left(2p_{+-}-p_{++}-p_{–}\right) > mes_2(\omega). $

    Proof. We consider the linear combination

    $ W_0(\xi) = W^+(\xi)-W^-(\xi)-\xi_1 = \chi_+(\xi_1)\left(p_{++}-p_{+-}\right)-\chi_-(\xi_1)\left(p_{–}-p_{-+}\right)+\widetilde{W}_0(\xi). $

    It satisfies

    $ -\Delta_\xi W_0(\xi) = 0, \, \xi\in\Xi, \qquad W_0(\xi) = -\xi_1, \, \xi\in\partial\omega, $

    with the periodicity conditions in the strip, and $ \widetilde{W_0}(\xi) = \widetilde{W}^+(\xi)-\widetilde{W}^-(\xi) $ gets the exponential decay rate $ O(e^{-|\xi_1|2\pi/H}) $. Considering the equations $ \Delta W_0 = 0 $ and $ \Delta (W_0+\xi_1) = 0 $ in $ \Xi\cap \{\vert \xi_1\vert <T\} $, and $ \Delta \xi_1 = 0 $ in $ \omega $, we apply the Green formula taking into account the boundary condition for $ W_0 $. Then, taking limits as $ T\to \infty $, we have

    $ 0<W0;L2(Ξ)2+mes2(ω)=ωξ1ν(ξ1)dν+ωW0ν(W0(ξ))dν=ωξ1ν(ξ1+W0(ξ))dν=ω(νξ1(ξ1+W0(ξ))ξ1ν(ξ1+W0(ξ)))dν=limT±±H0W0(±T,ξ2)dξ2=H(p+++pp+p+). $

    Remark 2. observe that for a hole $ \omega $, which is symmetric with respect to the $ x_1 $-axis, the matrix $ p(\Xi) $ becomes symmetric with respect to the anti-diagonal, namely,

    $ p++=p. $ (46)

    Indeed, this is due to the fact that each one of the two normalized solutions in (37) are related with each other by symmetry. Also, we note that, on account of Proposition 3.2, the symmetry $ p_{+-} = p_{-+} $ holds for any shape of the hole $ \omega $.

    Proposition 3.4. Let $ \omega $ be the crack $ {\omega} = \{\xi\in {\mathbb R}^2:\: \xi_1 = 0, \:\xi_2\in(h, H-h)\} $, where $ h< {H}/{2} $. Then,

    $ p+=p+>0. $ (47)

    In addition, $ p_{–} = p_{++} = p_{-+} = p_{+-}. $

    Proof. First, let us note that due to the symmetry $ W^+(\xi_1, \xi_2) = W^-(-\xi_1, \xi_2) $, and the construction (43) when $ R = 0 $ reads

    $ W(ξ1,ξ2)={ξ1+W(ξ1,ξ2),ξ1<0,W(ξ1,ξ2),ξ1>0. $ (48)

    where $ W^*(\xi_1, \xi_2) $ is the function defined in $ \Pi^+ = \{\xi: \: \xi_1>0, \, \xi_2\in(0, H)\} $ satisfying the periodicity condition (32) and equations

    $ ΔξW(ξ)=0, for ξΠ+,W(0,ξ2)=0, for ξ2(h,Hh),ξ1W(0,ξ2)=1/2, for ξ2(0,h)(Hh,H). $ (49)

    Indeed, denoting by $ \widetilde W^* $ the extension of $ W^* $ to $ \Pi^- = \{\xi : \: \xi_1<0, \, \xi_2\in(0, H)\} $, in order to verify the representation (48), it suffices to verify that the jump of $ \widetilde W^* $ and its the derivative of through $ \Upsilon(0) = \{\xi\in\Xi: \: \xi_1 = 0\} $ is given by

    $ [\widetilde W^* ] (0, \xi_2) = 0, \quad \left[\frac{\partial\widetilde W^* }{\partial \xi_1}\right] (0, \xi_2) = -1, $

    and hence, the function on the right hand side of (48) is a harmonic function in $ \Xi $.

    Now, considering (49), integrating by parts on $ (0, T)\times (0, H) $, and taking limits as $ T\to +\infty $ provide

    $ \int\limits_{\Upsilon(0)} W^*(0, \xi_2)d\xi_2 = \lim\limits_{T\to\infty}\int\limits^H_0 W^*(T, \xi_2)d\xi_2 = H p_{-+}(\Xi). $

    Similarly, from (49), we get

    $ 0 = -\int\limits_{\Pi^+} W^*(\xi)\Delta_\xi W^*(\xi) d\xi = \int\limits_{\Pi^+} |\nabla_\xi W^*(\xi)|^2 d\xi -\frac{1}{2}\int\limits_{\Upsilon(0)} W^*(0, \xi_2)d\xi_2. $

    Therefore, we deduce

    $ H2p+(Ξ)=Π+|ξW(ξ)|2dξ>0 $ (50)

    and from the symmetry of $ p(\Xi) $ (cf. Proposition 3.2), we obtain (47).

    Also, from the definition (48), we have $ p_{–}(\Xi) = p_{-+}(\Xi), $ and (cf. (46)) all the elements of the polarization matrix $ p(\Xi) $ coincide. Thus, the proposition is proved.

    From Proposition 3.4, note that when $ \omega $ is a vertical crack, the inequality in Proposition 3.3 must be replaced by $ H\left(2p_{+-}-p_{++}-p_{–}\right) = mes_2(\omega) = 0. $ Also, we observe that in order to get property (47) for a domain $ \omega $ with a smooth boundary, we may apply asymptotic results on singular perturbation boundaries (cf. [7], Ch. 3 in [8] and Ch. 5 in [17]) which guarantee that for thin ellipses

    $ ¯ω={ξ:δ2ξ21+(ξ2H/2)2τ2},τ=H/2h, $ (51)

    (47) holds true, for a small $ \delta>0 $.

    In this section we construct asymptotic expansions for the eigenpairs ($ \Lambda^{ \varepsilon}_{m} (\eta) $, $ U^{ \varepsilon}_{m}(\cdot;\eta) $) of problem (13)-(16) on the periodicity cell $ \varpi^{ \varepsilon} $. The parameters $ m\in \mathbb{N} $ and $ \eta \in [-\pi, \pi] $ are fixed in this analysis. In Sections 4.1-4.2 we consider the case in which the eigenvalue $ \Lambda^{0}_{m} $ of (24) is simple. Note that for many values of $ H $, all the eigenvalues are simple (cf. Remark 1). Section 4.3 contains the asymptotic ansatz for the eigenpairs case where $ \Lambda^{0}_{m} $ is an eigenvalue of (24) of multiplicity $ \kappa_m\geq 2 $.

    Let $ \Lambda^{0}_{m} $ be a simple eigenvalue in sequence (23) and let $ U^{0}_{m} $ be the corresponding eigenfunction of problem (24) normalized in $ L^2(\upsilon) $. Then, on account of the Theorem 2.1, for the eigenvalue $ \Lambda^{ \varepsilon}_{m} $ of problem (13)-(16) we consider the asymptotic ans$ {\rm \ddot{a }} $tze

    $ Λεm=Λ0m+εΛ1m(η)+. $ (52)

    To construct asymptotics of the corresponding eigenfunctions $ U^{ \varepsilon}_{m}(x;\eta) $, we employ the method of matched asymptotic expansions, see, e.g., the monographs [35] and [8], and the papers [32], [18] and [24] where this method has been applied to homogenization problems. Namely, we take

    $ Uεm(x;η)=U0m(x;η)+εU1m(x;η)+ $ (53)

    as the outer expansion, and

    $ Uεm(x;η)=ε±wm±(x2;η)W±(xε)+ $ (54)

    as the inner expansion near the perforation string, cf. (4) and (21).

    Above, $ U^{0}_{m}(x;\eta) $ is built from the eigenfunction $ U^{0}_{m} $ of (24) by formula

    $ U0m(x;η)={U0m(x),x1(0,1/2),eiηU0m(x1+1,x2),x1(1/2,0), $ (55)

    $ W^\pm $ are the solutions (37) to problem (31)-(33), while the functions $ U^{1}_{m} $, $ w^m_\pm $ and the number $ \Lambda^{1}_{m}(\eta) $ are to be determined applying matching principles, cf. Section 4.2. Note that near the perforation string, cf. (4), (21), the Dirichlet condition satisfied by $ U^{0}_{m}(x;\eta) $ implies that the term accompanying $ \varepsilon^0 $ in the inner expansion vanishes (see, e.g., [24]); this is why the first order function in (54) is $ \varepsilon $. Also, above and in what follows, the ellipses stand for higher-order terms, inessential in our formal analysis.

    First, let us notice that $ U^{0}_{m}\in C^\infty(\overline{\upsilon}) $, and the Taylor formula applied in the outer expansion (53) yields

    $ Uεm(x;η)=0+x1U0mx1(0,x2)+εU1m(+0,x2;η)+,x1>0,Uεm(x;η)=0+x1eiηU0mx1(1,x2)+εU1m(0,x2;η)+,x1<0, $ (56)

    where, for second formula (56), we have used (55).

    The inner expansion (54) is processed by means of decompositions (37). We have

    $ Uεm(x;η)=εwm+(x2;η)(ξ1+p++)+εwm(x2;η)p++,ξ1>0,Uεm(x;η)=εwm(x2;η)(ξ1+p)+εwm+(x2;η)p++,ξ1<0. $ (57)

    Recalling relationship between $ x_1 $ and $ \xi_1 $, we compare coefficients of $ \varepsilon $ and $ x_1 = \varepsilon \xi_1 $ on the right-hand sides of (56) and (57). As a result, we identify $ w^m_\pm $ by

    $ wm+(x2;η)=U0mx1(0,x2),wm(x2;η)=eiηU0mx1(1,x2), $ (58)

    and also obtain the equalities

    $ U1m(+0,x2;η)=τ=±wmτ(x2;η)pτ+,U1m(0,x2;η)=τ=±wmτ(x2;η)pτ. $ (59)

    Formulas (58) define coefficients of the linear combination (54) while formulas (59) are the boundary conditions for the correction term in (53). Moreover, inserting ans$ {\rm \ddot{a }} $tze (52) and (53) into (13)-(14), we derive that

    $ {ΔxU1m(x;η)Λ0mU1m(x;η)=Λ1m(η)U0m(x;η),xϖ0,x10,U1m(x1,H;η)=U1m(x1,0;η)=0,x1(1/2,0)(0,1/2), $ (60)

    and the quasi-periodic conditions with $ \eta $ (cf. (15)-(16)).

    Since $ U^0_m(x;\eta) $ is defined by (55), $ (\Lambda^{0}_{m}, \, U^{0}_{m}(x)) $ is an eigenpair of (24), and $ \|U^{0}_{m}; L^2(\upsilon)\| = \|U^{0}_{m}(\cdot;\eta); L^2(\varpi^0)\| = 1 $, we multiply by $ U^0_m(x;\eta) $ in the differential equation of (60), integrate by parts and obtain

    $ ϖ0Λ1m(η)U0m(x;η)¯U0m(x;η)dx=H0U1m(0,x2;η)¯U0mx1(0,x2;η)dx2H0U1m(+0,x2;η)¯U0mx1(+0,x2;η)dx2. $

    Thus, by (55) and (59), the only compatibility condition in (60) (recall that $ \Lambda^{0}_{m} $ is a simple eigenvalue) converts into

    $ Λ1m(η)=H0¯Bm(x2;η)p(Ξ)Bm(x2;η)dx2 $ (61)

    where

    $ Bm(x2;η)=(U0mx1(0,x2),eiηU0mx1(1,x2))TC2, $ (62)

    and it determines uniquely the second term of the ansatz (52). Here and in what follows, the top index $ T $ indicates the transpose vector.

    Also, from (53), (54) and (57) the composite expansion approaching $ U^ \varepsilon_{m}(x;\eta) $ in the whole domain $ \varpi_0 $ reads

    $ Uεm(x;η)U0m(x;η)+εU1m(x;η)+ετ=±wmτ(x2;η)Wτ(xε)(εwm±(x2;η)(ε1|x1|+p±±)+εwm(x2;η)p±),±x10. $ (63)

    We address the case where $ \Lambda^{0}_{m} $ is an eigenvalue of (24) with multiplicity $ \kappa_m\geq 2 $. Let us consider $ \Lambda^{0}_{m} = \cdots = \Lambda^{0}_{m+\kappa_m-1} $ in the sequence (23) and the corresponding eigenfunctions $ U^{0}_{m}, \cdots , $ $ U^{0}_{m+\kappa_m-1} $ which are orthonormal in $ L^2(\upsilon) $. On account of Theorem 2.1 there are $ \kappa_m $ eigenvalues of problem (13)-(16), which we denote by $ \Lambda^ \varepsilon_{m+l}( \eta) $, $ l = 0, \cdots, \kappa_m-1 $, satisfying

    $ Λεm+l(η)Λ0m+l as ε0, for l=0,,κm1. $ (64)

    Let $ U^ \varepsilon_{m+l}(\cdot;\eta) $, $ l = 0, \cdots, \kappa_m-1 $, be the corresponding eigenfunctions among the set of the eigenfunctions which form an orthonormal basis in $ L^2(\varpi^ \varepsilon) $, cf. (17).

    Following Section 4.1, for each $ l = 0, \cdots, \kappa_m-1 $, we take the ansatz for $ \Lambda^ \varepsilon_{m+l}( \eta) $

    $ Λεm+l=Λ0m+εΛ1m+l(η)+, $ (65)

    the outer expansion for $ U^ \varepsilon_{m+l}(\cdot;\eta) $

    $ Uεm+l(x;η)=U0m+l(x;η)+εU1m+l(x;η)+, $ (66)

    and the inner expansion

    $ Uεm(x;η)=ε±wm+l±(x2;η)W±(xε)+, $ (67)

    where the terms $ \Lambda^{1}_{m+l}(\eta) $, $ U^{1}_{m+l}(x;\eta) $ and $ w^{m+l}_\pm(x_2;\eta) $ have to be determined by the matching procedure, cf. Section 4.2, while $ U^{0}_{m+l}(x;\eta) $ is constructed from $ U^{0}_{m+l}(x ) $ replacing $ U^{0}_{m} $ by $ U^{0}_{m+l} $ in formula (55), and $ W^\pm $ are the solutions (37) to problem (31)-(33).

    By repeating the reasoning in Section 4.2, we obtain formulas for the above mentioned terms in (65), (66) and (67) by replacing index $ m $ by $ m+l $ in (56)-(62), while we realize that the compatibility condition for each $ \Lambda_{m+l}^1(\eta) $ is satisfied. Indeed, multiplying by $ U^0_{m+l'}(x;\eta) $, $ l' = 0, \cdots, \kappa_m-1 $ in the partial differential equation satisfied by $ U^1_{m+l'}(x;\eta) $ (cf. (60))

    $ -\Delta_x U^{1}_{m+l}(x;\eta)-\Lambda^{0}_{m} U^{1}_{m+l}(x;\eta) = \Lambda^{1}_{m+l}(\eta) U^{0}_{m+l}(x;\eta), \quad x\in \varpi^0, \, x_1\not = 0, $

    and integrating by parts, we obtain

    $ ϖ0Λ1m+l(η)U0m+l(x;η)¯U0m+l(x;η)dx=H0(U0m+lx1(0,x2),eiηU0m+lx1(1,x2))p(Ξ)×(U0m+lx1(0,x2),eiηU0m+lx1(1,x2))Tdx2. $

    Since the eigenfunctions $ U^0_{m+l} $ and $ U^0_{m+l'} $ have been computed (cf. the explicit formulas (29) in Remark 1), we conclude now that

    $ \int\limits^H_0 \frac{\partial {U^{0}_{m+l' }}}{\partial x_1}(x_1^*, x_2)\frac{\partial U^{0}_{m+l }}{\partial x_1}(x_1^*, x_2)dx_2 = 0, \mbox{ with } x_1^*\in \{0, 1\}, \quad l\not = l', $

    and, hence, for each $ l = 0, \cdots, \kappa_m-1 $, the $ \kappa_{m} $ compatibility conditions to be satisfied by the pairs ($ \Lambda_{m+l}^1(\eta) $, $ U^1_{m+l}(x;\eta) $), cf. (60), provide $ \Lambda_{m+l}^1(\eta) $ given by

    $ Λ1m+l(η)=H0¯Bm+l(x2;η)p(Ξ)Bm+l(x2;η)dx2, $ (68)

    where $ B_{m+l}(x_2;\eta) $ is defined by

    $ Bm+l(x2;η)=(U0m+lx1(0,x2),eiηU0m+lx1(1,x2))T. $ (69)

    Therefore we have determined completely all the terms in the asymptotic ans$ {\rm \ddot{a }} $tze (65), (66) and (67) for $ l = 0, \cdots, \kappa_m-1 $.

    In this section, we justify the results obtained by means of matched asymptotic expasions in Section 4. Since the case in which all the eigenvalues of the Dirichlet problem (24) are simple can be a generic property, we first consider this case, cf. Theorem 5.1 and Corollary 5.1, and then the case in which these eigenvalues have a multiplicity greater than $ 1 $, cf. Theorem 5.2 and Corollary 5.2. We state the results in Section 5.1 while we perform the proofs in Section 5.2.

    Theorem 5.1. Let $ m\in {\mathbb N} $, let $ \Lambda^{0}_{m} $ be a simple eigenvalue of the Dirichlet problem (24), and let $ \Lambda^{1}_{m}(\eta) $ be defined in (61) and (62). There exist positive $ \varepsilon_m $ and $ c_m $ independent of $ \eta $ such that, for any $ \varepsilon\in(0, \varepsilon_m] $, the eigenvalue $ \Lambda^{ \varepsilon}_{m}(\eta) $ of problem (13)-(16) meets the estimate

    $ |Λεm(η)Λ0mεΛ1m(η)|cmε3/2 $ (70)

    and there are no other different eigenvalues in the sequence (18) satisfying (70).

    Theorem 5.1 shows that $ \varepsilon \Lambda^{1}_{m}(\eta) $ provides a correction term for $ \Lambda^{ \varepsilon}_{m}(\eta) $ improving the approach to $ \lambda^0_m $ shown in Theorem 2.1. In particular, it justifies the asymptotic ansatz (52) and formula (61). This corrector depends on the polarization matrix $ p(\Xi) $, which is given by the coefficients $ p_{\tau\pm}\equiv p_{\tau\pm}{ (\Xi)} $, with $ \tau = \pm $, in the decomposition (37), and on the eigenfunction $ U^{0}_{m} $ of problem (24), which corresponds to $ \Lambda^{0}_{m} $ and is normalized in $ L^2(\upsilon) $ (cf. (61) and (62)).

    In order to detect the gaps between consecutive spectral bands (20) it is worthy writing formulas

    $ Λ1m(η)=B0(m)+B1(m)cos(η), with B0(m)=H0(p++|U0mx1(0,x2)|2+p|U0mx1(1,x2)|2)dx2,B1(m)=2p+H0U0mx1(0,x2)U0mx1(1,x2)dx2, $ (71)

    which are obtained from (61) and (62). Formula (29) demonstrates that

    $ B_0(m) = \left(p_{++}+p_{–}\right)\int\limits^H_0\left|\frac{\partial U^{0}_{m }}{\partial x_1}(0, x_2)\right|^2dx_2, $

    and that the integral in $ B_1(m) $ does not vanish. We note that $ B_1(m) = 0 $ only in the case when $ p_{+-} = 0 $; if so, $ p(\Xi) $ is diagonal and the solutions of (37), $ W^\pm $, decay exponentially when $ \xi_1\to\mp\infty $, respectively. However, we have given examples of cases where $ p_{+-}\not = 0 $ (cf. (47) and (51)).

    Remark 3. Let us consider that the eigenvalue $ \Lambda^{0}_{m} $ coincides with $ \Lambda^{0}_{n{ q }} $ in formula (29) for certain natural $ n $ and $ {q} $. Then, we obtain

    $ B_0(m) = 2\left(p_{++}+p_{–}\right)n^2\pi^2, \quad B_1(m) = (-1)^n 4 p_{+-} n^2\pi^2, $

    and, consequently,

    $ Λ1m(η)=2(p+++p)n2π2+(1)n4p+n2π2cos(η). $ (72)

    Corollary 5.1. Under the hypothesis of Theorem 5.1, the endpoints $ B^{ \varepsilon\pm}(m) $ of the spectral band (20) satisfy the relation

    $ |Bε±(m)Λ0mε(B0(m)±|B1(m)|)|cmε3/2. $ (73)

    Hence, the length of the band $ B^ \varepsilon_m $ is $ 2 \varepsilon|B_1(m)|+ O( \varepsilon^{3/2}) $.

    Note that for the holes such that the polarization matrix (38) satisfies $ p_{+-}\not = 0 $, asymptotically, the bands $ B^ \varepsilon_m $ have the precise length $ 2 \varepsilon|B_1(m)|+ O( \varepsilon^{3/2}) $ and they cannot reduce to a point, namely to the point $ \Lambda_m^0 + \varepsilon B_0(m) $ (cf. (71) and Remark 3). Also note that if $ p_{+-} = 0 $, Theorem 5.1 still provides a correction term for $ \Lambda^ \varepsilon_m(\eta) $ which however does not depend on $ \eta $ (cf. (70), (71) and Remark 3), the width of the band being $ O( \varepsilon^{3/2}) $. Although the length of the band is shorter than in the cases where $ p_{+-}\not = 0 $, bounds in Corollary 5.1 may not be optimal (cf. Remark 3) and further information on the corrector depending on $ \eta $ can be obtained by constructing higher-order terms in the asymptotic ansatz (53).

    Theorem 5.2. Let $ m\in {\mathbb N} $, let $ \Lambda^{0}_{m} $ be an eigenvalue of the Dirichlet problem (24) with multiplity $ \kappa_m>1 $. Let $ \Lambda^{1}_{m+l}(\eta) $ defined in (68) and (69) for $ l = 0, \cdots, \kappa_m-1 $. There exist positive $ \varepsilon_m $ and $ c_m $ independent of $ \eta $ such that, for any $ \varepsilon\in(0, \varepsilon_m] $, and for each $ l = 0, \cdots, \kappa_m-1 $, at least one eigenvalue $ \Lambda^{ \varepsilon}_{m+l_0}(\eta) $ of problem (13)-(16) satisfying (64) meets the estimate

    $ |Λεm+l0(η)Λ0mεΛ1m+l(η)|cmε3/2. $ (74)

    In addition, when $ l\in \{0, 1, \cdots, \kappa_m-1\} $, the total multiplicity of the eigenvalues in (18) satisfying (74) is $ \kappa_m $.

    Corollary 5.2. Under the hypothesis in Theorem 5.2, the spectral bands $ B^ \varepsilon_{m+l} $ associated with $ \Lambda^{ \varepsilon}_{m+l }(\eta) $, for $ l = 0, \cdots, \kappa_m-1 $, cf. (20), are contained in the interval

    $ [Λ0m+εmin0lκm1η[π,π]Λ1m+l(η)cmε3/2,Λ0m+εmax0lκm1η[π,π]Λ1m+l(η)+cmε3/2]. $ (75)

    Hence, the length of the the bands $ B^ \varepsilon_{m+l} $ are $ O( \varepsilon) $ but they may not be disjoint.

    Remark 4. Under the hypothesis of Theorem 5.2, it may happen that, for $ l = 0, \cdots, \kappa_m-1 $, only the eigenvalue $ \Lambda^{ \varepsilon}_{m+l_0}(\eta) $ in the sequence (18) satisfies (74). This depends on the polarization matrix $ p(\Xi) $. As a matter of fact, it can be shown by contradiction under the assumption that for two different $ l $ the functions $ \Lambda^1_{m+l}(\eta) $ do not intersect at any point $ \eta $, cf. (71) and (72). For instance, this follows for $ \omega $ with $ p_{+-}(\Xi) = 0 $.

    Remark 5. Notice that the positive cutoff value $ \lambda^ \varepsilon_\dagger $ such that the spectrum $ \sigma^ \varepsilon \subset [\lambda^ \varepsilon_\dagger, \infty) $ is bounded from above by a positive constant, cf. (9), (20) and (25). In addition, from Theorem 5.2 (cf. Remark 1), we have proved that $ \lambda^ \varepsilon_\dagger \to \pi^2( 1+H^{-2}) $ as $ \varepsilon\to 0 $.

    In this section we prove the results of Theorems 5.1 and 5.2 and of their respective corollaries.

    Proof of Theorem 5.1. Let us fix $ \eta $ in $ [-\pi, \pi] $. Let us endow the space $ H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $, with the scalar product $ \langle U^ \varepsilon, V^ \varepsilon\rangle = (\nabla U^ \varepsilon, \nabla V^ \varepsilon)_{\varpi^ \varepsilon}+ ( U^ \varepsilon, V^ \varepsilon)_{\varpi^ \varepsilon}, $ and the positive, symmetric and compact operator $ T^ \varepsilon(\eta) $,

    $ Tε(η)Uε,Vε=(Uε,Vε)ϖεUε,VεH1,ηper(ϖε;Γε). $ (76)

    The integral identity (17) for problem (13)-(16) can be rewritten as the abstract equation

    $ T^ \varepsilon(\eta) U^ \varepsilon(\cdot;\eta) = \tau^ \varepsilon(\eta) U^ \varepsilon(\cdot;\eta) \quad \hbox{in } H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon), $

    with the new spectral parameter

    $ τε(η)=(1+Λε(η))1. $ (77)

    Since $ T^ \varepsilon(\eta) $ is compact (cf. e.g, Section I.4 in [31] and Ⅲ.9 in [1]), its spectrum consists of the point $ \tau = 0 $, the essential spectrum, and of the discrete spectrum $ \{\tau^{ \varepsilon}_{m}(\eta) \}_{m\in {\mathbb N }} $ which, in view of (18) and (77), constitutes the infinitesimal sequence of positive eigenvalues

    $ \left\{\tau^{ \varepsilon}_{m}(\eta) = (1 +\Lambda^{ \varepsilon}_{m}(\eta) )^{-1}\right\}_{m\in {\mathbb N }}. $

    For the point

    $ tεm(η)=(1+Λ0m+εΛ1m(η))1, $ (78)

    cf. (52) and (61), we construct a function $ \mathcal{U}^{ \varepsilon}_{m} \in H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $ such that

    $ Uεm;H1,ηper(ϖε;Γε)cm, $ (79)
    $ Tε(η)Uεmtεm(η)Uεm;H1,ηper(ϖε;Γε)Cmε3/2, $ (80)

    where $ c_m $ and $ C_m $ are some positive constants independent of $ \varepsilon\in(0, \varepsilon_m] $, with $ \varepsilon_m>0 $. These inequalities imply the estimate for the norm of the resolvent operator $ (T^ \varepsilon(\eta) -t^{ \varepsilon}_{ m}(\eta))^{-1} $

    $ \|(T^ \varepsilon(\eta) -t^{ \varepsilon}_{m}(\eta))^{-1}; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon)\to H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) \|\geq \mathbf{c}_m^{-1} \varepsilon^{-3/2}, $

    with $ \mathbf{c}_m = c_m^{-1} C_m>0 $. According to the well-known formula for self-adjoint operators

    $ \operatorname{dist}(t^{ \varepsilon}_{m}(\eta), \sigma(T^ \varepsilon(\eta)) = \|(T^ \varepsilon(\eta) -t^{ \varepsilon}_{m}(\eta))^{-1}; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon)\to H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) \|^{-1} $

    supported by the spectral decomposition of the resolvent (cf., e.g., Section V.5 in [9] and Ch. 6 in [1]), we deduce that the closed segment

    $ [t^{ \varepsilon}_{m}(\eta)- \mathbf{c}_m \varepsilon^{3/2}, t^{ \varepsilon}_{m}(\eta) + \mathbf{c}_m \varepsilon^{3/2}] $

    contains at least one eigenvalue $ \tau^{ \varepsilon}_{p}(\eta) $ of the operator $ T^ \varepsilon(\eta) $. Since the eigenvalues of $ T^ \varepsilon(\eta) $ satisfy (77) and we get the definition (78), we derive that

    $ |(1+Λεp(η))1(1+Λ0m+εΛ1m(η))1|cmε3/2. $ (81)

    Then, simple algebraic calculations (cf. (81) and (25)) show that, for a $ \varepsilon_m>0 $, the estimate

    $ |Λεp(η)Λ0mεΛ1m(η)|Cmε3/2 $ (82)

    is satisfied with a constant $ \mathcal{C}_m $ independent of $ \varepsilon\in(0, \varepsilon_m] $. Due to the convergence with conservation of the multiplicity (22), $ p = m $ in (82) and this estimate becomes (70).

    To conclude with the proof of Theorem 5.1, there remains to present a function $ \mathcal{U}^{ \varepsilon}_{m}\in H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $ enjoying restrictions (79) and (80). In what follows, we construct $ \mathcal{U}^{ \varepsilon}_{ m} $ using (63) suitably modified with the help of cut-off functions with "overlapping supports", cf. [19], Ch. 2 in [17] and others. We define

    $ Vεmout(x;η)=U0m(x;η)+εU1m(x;η), $ (83)

    where $ U^0_m $ satisfies (55) and $ U^1_m $ is the solution of (60) satisfying the boundary conditions (15)-(16) and (59). Similarly, we define

    $ Vεmin(x;η)=ε±wm±(x2;η)W±(ε1x), $ (84)

    and

    $ Vεmmat(x;η)=εwm±(x2;η)(ε1|x1|+p±±)+εwm(x2;η)p±,±x1>0, $ (85)

    with $ w^m_\pm $ defined in (58), and $ W^\pm $ and matrix $ p(\Xi) $ in Proposition 1 (cf. (38)). Finally, we set

    $ Uεm(x;η)=Xε(x1)Vεmout(x;η)+X(x1)Vεmin(x;η)Xε(x1)X(x1)Vεmmat(x;η), $ (86)

    where $ X^ \varepsilon $ and $ \mathcal{X} $ are two cut-off functions, both smoothly dependent on the $ x_1 $ variable, such that

    $ Xε(x1)={1, for |x1|>2Rε,0, for |x1|<Rε, and X(x1)={1, for |x1|<1/6,0, for |x1|>1/3. $ (87)

    Note that (85) takes into account components in both expressions (83) and (84), but the last subtrahend in $ \mathcal{U}^{ \varepsilon}_{m} $ compensates for this duplication. In further estimations, term (85) will be joined to either $ \mathcal{V}^{ \varepsilon m}_{out} $ or $ \mathcal{V}^{ \varepsilon m}_{in} $ in order to obtain suitable bounds.

    First, let us show that $ \mathcal{U}^{ \varepsilon}_{m} \in H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $. Indeed, the function defined in (86) enjoys the conditions (15)-(16) and (14). This is due to the fact that $ \mathcal{U}^{ \varepsilon}_{m} = \mathcal{V}^{ \varepsilon m}_{out} $ near the sides $ \{x_1 = \pm1/2, \, x_2\in (0, H)\} $ and the quasi-periodicity conditions (15)-(16) are verified by both terms in (83). Also, $ \mathcal{U}^{ \varepsilon}_{m} = \mathcal{V}^{ \varepsilon m}_{in} $ near the perforation string (21) so that the Dirichlet conditions are fulfilled on the boundary of the perforation string $ \Gamma^ \varepsilon \cap \varpi^0 $ because $ W^\pm $ satisfy (33). Finally, formulas (58) and (29) assure that $ w^m_\pm(H;\eta) = w^m_\pm(0;\eta) = 0 $ and hence the Dirichlet condition is met on $ \Gamma^ \varepsilon\cap\partial\varpi^0 $ as well.

    First of all, we recall (83) and (87) to derive

    $ Uεm;H1,ηper(ϖε;Γε)Uεm;L2((1/3,1/2)×(0,H))=Vεmout;L2((1/3,1/2)×(0,H))U0m;L2((1/3,1/2)×(0,H))εU1m;L2((1/3,1/2)×(0,H))c>0, $

    for a small $ \varepsilon>0 $. Thus, (79) is fulfilled.

    Using (76) and (78), we have

    $ Tε(η)Uεmtεm(η)Uεm;H1,ηper(ϖε;Γε)=sup|Tε(η)Uεmtεm(η)Uεm,Wε|=(1+Λ0m+εΛ1m(η))1sup|(Uεm,Wε)ϖε(Λ0m+εΛ1m(η))(Uεm,Wε)ϖε|, $ (88)

    where the supremum is computed over all $ \mathcal{W}^{ \varepsilon}\in H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) $ such that

    $ \| \mathcal{W}^{ \varepsilon}; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon)\|\leq 1. $

    Taking into account the Dirichlet conditions on $ \partial\omega^ \varepsilon $ we use the Poincaré and Hardy inequalities, namely, for a fixed $ T $ such that $ \omega\subset \Pi_T\equiv\Pi\cap \{y_1<T \} $,

    $ \int_{ \Pi_T\setminus\overline{\omega} } \vert U \vert^2\, dy \leq C_{T}\int_{ \Pi_T\setminus\overline{\omega} } \vert \nabla_y U \vert^2 dy \quad \forall U \in H^1( \Pi_T\setminus\overline{\omega} ) , \, U = 0 \mbox{ on } \partial \omega, $

    where $ C_{T} $ is a constant independent of $ U $, and

    $ \int_0^\infty \frac{1}{t^2} z(t)^2\, dt\leq 4 \int_0^\infty \left\vert \frac{dz}{dt}(t)\right\vert^2\, dt \quad \forall z\in C^1[0, \infty), \, z(0) = 0. $

    Then, we have

    $ (ε+|x1|)1Wε;L2(ϖε)cWε;L2(ϖε)c. $ (89)

    Clearly, from (71), $ (1+ \Lambda^{0}_{m}+ \varepsilon\Lambda^{1}_{m}(\eta))^{-1}\leq 1 $ for a small $ \varepsilon>0 $ independent of $ \eta $, and there remains to estimate the last supremum in (88). We integrate by parts, take the Dirichlet and quasi-periodic conditions into account and observe that

    $ |(Uεm,Wε)ϖε(Λ0m+εΛ1m(η))(Uεm,Wε)ϖε|=|(ΔUεm+(Λ0m+εΛ1m(η))Uεm,Wε)ϖε|. $

    On the basis of (83)-(86) we write

    $ ΔUεm+(Λ0m+εΛ1m(η))Uεm=Xε(ΔU0m+Λ0mU0m+ε(ΔU1m+Λ0mU1m+Λ1mU0m)+ε2Λ1mU1m)+[Δ,Xε](VεmoutVεmmat)+X(ΔVεminXεΔVεmmat)+[Δ,X](VεminVεmmat)+(Λ0m+εΛ1m)X(VεminXεVεmmat)=:Sε1+Sε2+Sε3+Sε4+Sε5. $ (90)

    Here, $ [\Delta, \chi] = 2\nabla\chi \cdot\nabla+ \Delta \chi $ is the commutator of the Laplace operator with a function $ \chi $, and the equality $ [\Delta, X^ \varepsilon \mathcal{X}] = [\Delta, \mathcal{X}]+[\Delta, X^ \varepsilon], $ which is valid due to the position of supports of functions in (87), is used when distributing terms originated by the last subtrahend in (86). Let us estimate the scalar products $ (S^ \varepsilon_k , \mathcal{W}^ \varepsilon)_{\varpi^ \varepsilon} $ for $ S_k^ \varepsilon $ in (90).

    Considering $ S^ \varepsilon_1 $, because of (27), (60), (87) and (71), we have that in fact $ S^ \varepsilon_1 = \varepsilon^2 X^ \varepsilon\Lambda^{1}_{m} U^{1}_{m} $, hence

    $ \left|(S^ \varepsilon_1, \mathcal{W}^ \varepsilon)_{\varpi^ \varepsilon}\right|\leq \varepsilon^2 \Lambda^{1}_{m}(\eta) \|U^{1}_{m} ; L^2(\varpi^ \varepsilon)\|\| \mathcal{W}^{ \varepsilon}; L^2(\varpi^ \varepsilon)\|\leq C_m \varepsilon^2. $

    As regards $ S^ \varepsilon_2 $, we take into account that the supports of the functions $ \partial_{x_1} X^ \varepsilon $ and $ \Delta X^ \varepsilon $ belong to the adherence of the thin domain $ \varpi^ \varepsilon_{ \varepsilon R} = \{x\in\overline{\varpi^ \varepsilon}:\: |x_1|\in ( \varepsilon R, 2 \varepsilon R)\} $, cf. (87). Thus, the error in the Taylor formula up to the second term, and relations (58), (59) and (85) provide

    $ |Vεmout(x;η)Vεmmat(x;η)|c(|x1|2+ε|x1|),|Vεmoutx1(x;η)Vεmmatx1(x;η)|c(|x1|+ε),±x1[εR,2εR]. $

    Above, we have also used the smoothness of the function $ U_m^1 $ which holds on account that $ V = U_m^1e^{-{\rm i} \eta y_1} $ is a periodic function in the $ y_1 $ variable, solution of an elliptic problem with constant coefficients (cf. (60), (15)-(16) and (91)), and therefore it is smooth. Then, we make use of the weighted inequality (89) and write

    $ |(Sε2,Wε)ϖε|Sε2;L2(ϖεεR)Wε;L2(ϖεεR)cε(ε+|x1|)1Wε;L2(ϖε)×(H02εRεR(1ε2|Vεmoutx1Vεmmatx1|2+1ε4|VεmoutVεmmat|2)d|x1|dx2)12c(1ε2ε2+1ε4ε4)12(mes2ϖεεR)12ε(ε+|x1|)1Wε;L2(ϖε)cε32. $

    Dealing with $ S^ \varepsilon_3 $, we match the definitions of the cut-off functions $ \chi_\pm $ and $ X^ \varepsilon $ such that $ X^ \varepsilon(x_1) = \chi_\pm(x_1/ \varepsilon) $ for $ \pm x_1>0 $ (cf. (36)). Recalling formulas (37), (84) and (85), we write

    $ ΔVεmin(x;η)Xε(x1)ΔVεmmat(x;η)=2±wm±x2(x2;η)W±ξ2(y)+ε±2wm±x22(x2;η)˜W±(y), $

    when $ \pm x_1>0 $, respectively. Note that $ W^\pm $ are harmonics and both, $ \partial W^\pm/\partial \xi_2 $ and $ \widetilde{W}^\pm $ decay exponentially as $ |\xi_1|\to\infty $, see Proposition 3.1. Thus,

    $ |(Sε3,Wε)ϖε|c((ε+|x1|)W±ξ2;L2(ϖε)+ε(ε+|x1|)˜W±;L2(ϖε))1ε+|x1|Wε;L2(ϖε)c(1/20(ε+t)2e2δt/εdt)12Wε;L2(ϖε)cε32. $

    Above, obviously, we take the positive constant $ \delta $ to be $ 2\pi /H $, cf. Proposition 3.1, and we note that the last integral has been computed to obtain the bound. With the same argument on the exponential decay of $ \mathcal{V}^{ \varepsilon m}_{in}- X^ \varepsilon \mathcal{V}^{ \varepsilon m}_{mat} $, one derives that

    $ \left|(S^ \varepsilon_5, \mathcal{W}^ \varepsilon)_{\varpi^ \varepsilon}\right|\leq c \varepsilon^\frac{3}{2}. $

    Moreover, the supports of the coefficients $ \partial_{ x_1} \mathcal{X} $ and $ \Delta \mathcal{X} $ in the commutator $ [\Delta, \mathcal{X}] $ are contained in the set $ \overline{\varpi^ \varepsilon } \cap \{1/6<|x_1|<1/3\} $ while the above-mentioned decay brings the estimate

    $ \left|(S^ \varepsilon_4, \mathcal{W}^ \varepsilon)_{\varpi^ \varepsilon}\right|\leq c{e^{-2\delta/(3 \varepsilon) }}. $

    Revisiting the obtained estimates we find the worst bound $ c \varepsilon^{3/2} $, and this shows (80).

    The fact that the constants $ \varepsilon_m $ and $ c_m $ of the statement of the theorem are independent of $ \eta $ follows from the independence of $ \eta $ of the above inequalities throughout the proof. Indeed, we use formulas (55) and (71) for the boundedness of $ U_m^0 $ and $ \Lambda_m^1(\eta) $, while we note that the fact that $ \|U^{1}_{m} ; H^1(\varpi^ \varepsilon)\| $ is bounded by a constant independent of $ \eta $ follows from the definition of the solution of (60) with the quasi-periodic boundary conditions (15)-(16). Further specifying, the change $ V = U_m^1e^{-{\rm i} \eta y_1} $ converts the Laplacian into the differential operator

    $ (y1+iη)(y1+iη)2y22, $ (91)

    and therefore, performing this change in (60), gives the solution $ V\in H^{1 }_{per} (\varpi^0) $. Then, as a consequence of the variational formulation of the problem for $ V $ in the set of spaces $ L^2 (\varpi_0)\subset H^{1 }_{per} (\varpi^0) $, the bound of $ \|U^{1}_{m} ; H^1(\varpi^ \varepsilon)\| $ independently of $ \eta\in [-\pi, \pi] $ holds true. Hence, the proof of Theorem 5.1 is completed.

    Proof of Corollary 5.1. Due to the continuity of the function (19), the maximum and minimum of $ \Lambda_m^ \varepsilon(\eta) $ for $ \eta\in [-\pi, \pi] $ are achieved at two points $ \eta^\pm_{ \varepsilon, m}\in [-\pi, \pi] $. Thus, the endpoints $ B^{ \varepsilon\pm}(m) $ of the spectral band (20) are given by $ B^{ \varepsilon\pm}(m) = \Lambda_m^ \varepsilon(\eta^\pm_{ \varepsilon, m}) $.

    In order to show (73) for the maximum $ B^{ \varepsilon+}(m) $, we consider $ \eta = \eta^+ $ to be $ \pi $ or $ -\pi $ in such a way that $ \Lambda^1(\eta^+) = B_0(m) +|B_1(m)| $. Since (70) is satisfied for $ \eta = \eta^\pm_{ \varepsilon, m} $ and for $ \eta = \pm \pi $, we write

    $ \Lambda^{0}_{m}+ \varepsilon B_0(m) + \varepsilon |B_1(m)|-c_m \varepsilon^{3/2} \leq \Lambda_m^{ \varepsilon}(\eta^+) \leq \Lambda^{0}_{m}+ \varepsilon B_0(m) + \varepsilon |B_1(m)| + c_m \varepsilon^{3/2} $

    and

    $ \Lambda^{0}_{m} + \varepsilon \Lambda^1_m(\eta^+_{ \varepsilon, m}) -c_m \varepsilon^{3/2} \leq\Lambda_m^{ \varepsilon}(\eta^+_{ \varepsilon, m}) \leq \Lambda^{0}_{m}+ \varepsilon \Lambda^1_m(\eta^+_{ \varepsilon, m}) + c_m \varepsilon^{3/2}. $

    Consequently, from (71), we derive

    $ Λ0m+εB0(m)+ε|B1(m)|cmε3/2Λεm(η+)Λεm(η+ε,m)Λ0m+εB0(m)+ε|B1(m)|+cmε3/2, $

    which gives (73) for $ B^{ \varepsilon+}(m) $.

    We proceed in a similar way for the minimum $ B^{ \varepsilon-}(m) $ and $ \eta = \eta^- $ such that $ \Lambda^1(\eta^-) = B_0(m) -|B_1(m)| $ and we obtain (73). Obviously, this implies that $ B^{ \varepsilon\pm}(m) $ belong to the interval

    $ \left [\Lambda^{0}_{m}+ \varepsilon B_0(m) - \varepsilon |B_1(m)|-c_m \varepsilon^{3/2} \, , \, \Lambda^{0}_{m}+ \varepsilon B_0(m) + \varepsilon |B_1(m)| + c_m \varepsilon^{3/2}\right] $

    Therefore, the whole band $ B_m^ \varepsilon $ is contained in the interval above whose length is $ 2 \varepsilon |B_1(m)| + O( \varepsilon^{3/2}) $ and the corollary is proved.

    Proof of Theorem 5.2. This proof holds exactly the same scheme of Theorem 5.1. Indeed, for each $ l = 0, \cdots, \kappa_m-1 $, we follow the reasoning in (76)-(82) and we deduce (cf. (81) and (25)) that for each $ l $, and for a $ \varepsilon_{m, l}>0 $, the estimate

    $ |Λεp(η)Λ0mεΛ1m+l(η)|Cm,lε3/2 $ (92)

    is satisfied for a certain natural $ p{ \equiv p(l)} $ and $ \mathcal{C}_{m, l} $ independent of $ \varepsilon\in(0, \varepsilon_{m, l}] $. Due to the convergence with conservation of the multiplicity (64), the only possible eigenvalues $ \Lambda^{ \varepsilon}_{p}(\eta) $ of problem (13)-(16) satisfying (92) are the set $ \{ \Lambda^{ \varepsilon}_{m+l}(\eta) \}_{l = 0, \cdots, \kappa_m-1} $. Then, it suffices to show that there are $ \kappa_m $ linearly independent eigenfunctions associated with the eigenvalues $ \{ \Lambda^{ \varepsilon}_{p(l)}(\eta) \}_{l = 0, \cdots, \kappa_m-1} $ in (92), to deduce the result of the theorem.

    We use a classical argument of contradiction (cf. [15] and [25]). We consider the set of functions $ \{ \mathcal{U}^{ \varepsilon}_{m+l}(x;\eta)\}_{l = 0, \cdots, \kappa_m-1} $, constructed in (86), and we verify that they satisfy the almost orthogonality conditions

    $ Uεm+l;H1,ηper(ϖε;Γε)˜cm and |Uεm+l,Uεm+l|˜Cmε1/2, with ll, $ (93)

    for certain constants $ \widetilde c_m $ and $ \widetilde C_m $. Indeed, the first inequality above is a consequence of (79), for $ l = 0, \cdots, \kappa_m-1 $, while the second one follows from the orthogonality of the set of eigenfunctions $ \{U^0_{m+l}(x, \eta)\}_{l = 0, \cdots, \kappa_m-1} $ and the definitions (83)-(87).

    Then, we define $ \widetilde{\mathcal{U }}^ \varepsilon_{m+l} = \mathcal{U}^ \varepsilon_{m+l} \, \| \mathcal{U}^{ \varepsilon}_{m+l} ; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon)\|^{-1} $ and consider $ \mathcal{W}^ \varepsilon_{m+ l } $ the projection of $ T^ \varepsilon(\eta)\widetilde{\mathcal{U }}^{ \varepsilon}_{m+l} -t^{ \varepsilon}_{m}(\eta)\widetilde{\mathcal{U }}^{ \varepsilon}_{m+l} $ in the space of the eigenfunctions of $ T^ \varepsilon(\eta) $ associated with all the eigenvalues $ \{ \Lambda^{ \varepsilon}_{p(l)}(\eta)\}_{l = 0, \cdots, \kappa_m-1} $ in (92) for certain constants $ \mathcal{C}_{m, l} $, and more precisely, satisfying

    $ |(1+Λεp(l)(η))1(1+Λ0m+εΛ1m+l(η))1|˜cmε3/2, $ (94)

    for a constant $ \widetilde{\mathbf{c }}_m $ that we shall set later in the proof, cf. (81) and definitions (76)-(78) for the operator $ T^ \varepsilon(\eta) $ and the "almost eigenvalue" $ t^{ \varepsilon}_{m}(\eta) $. Then, we show

    $ ˜Wεm+l˜Uεm+l;H1,ηper(ϖε;Γε)˜Cm, $ (95)

    where $ \widetilde{\mathcal{W }}^ \varepsilon_{m+l} = { \mathcal{W}^ \varepsilon_{m+l }}\| \mathcal{W}^{ \varepsilon}_{m+l} ; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon)\|^{-1} $, and $ \widetilde{\mathcal{C }}_m = 2\widetilde{\mathbf{c }}_m^{-1} \max_{0\leq l\leq \kappa_m-1} \mathcal{C}_{l, m} $. This is due to the fact that

    $ \left \| \, \widetilde{\mathcal{U }}^ \varepsilon_{m+l} - \mathcal{W}^ \varepsilon_{m+l} ; H^{1, \eta}_{per} (\varpi^ \varepsilon;\Gamma^ \varepsilon) \right \|\leq \widetilde{\mathbf{c }}_m^{-1} \max\limits_{0\leq l\leq \kappa_m-1} \mathcal{C}_{l, m} , $

    and some straightforward computation (cf., eg., Lemma 1 in Ch. 3 of [27]). Now, from (93) and (95) and straightforward computations we obtain

    $ |˜Wεm+l,˜Wεm+l|5˜Cm with ll, $ (96)

    and this allows us to assert that set $ \{ \widetilde{\mathcal{W }}^ \varepsilon_{m+l}\}_{l = 0, \cdots, \kappa_m-1} $ defines $ \kappa_m $ linearly independent functions. Indeed, to prove it, we proceed by contradiction, by assuming that there are constants $ \alpha_l^ \varepsilon $ different from zero such that

    $ \sum\limits_{l = 0}^{\kappa_m-1} \alpha_l^ \varepsilon \widetilde{\mathcal{W }}^ \varepsilon_{m+l} = 0. $

    Let us consider $ \alpha^{*, \varepsilon} = \max_{0\leq l\leq \kappa_m-1} \vert \alpha_l^ \varepsilon \vert $ and assume, without any restriction that $ \alpha^{*, \varepsilon} = \alpha_0^ \varepsilon $. Then, we write

    $ \langle \widetilde{\mathcal{W }}^ \varepsilon_{m}, \widetilde{\mathcal{W }}^ \varepsilon_{m} \rangle \leq \sum\limits_{l = 1}^{\kappa_m-1} \left\vert \frac{\alpha_l^ \varepsilon}{\alpha_0^ \varepsilon} \right\vert \left\vert \langle \widetilde{\mathcal{W }}^ \varepsilon_{m+l}, \widetilde{\mathcal{W }}^ \varepsilon_{m} \rangle \right \vert \leq (\kappa_m-1) 5\widetilde{\mathcal{C }}_m . $

    Now, setting $ (\kappa_m-1)5\widetilde{\mathcal{C }}_m <1 $ gives a contradiction, since the left hand side takes the value $ 1 $, cf. (96). In this way, we also have fixed $ \widetilde{\mathbf{c}}_m $ in (94).

    Thus, $ \{ \widetilde{\mathcal{W }}^ \varepsilon_{m+l}\}_{l = 0, \cdots, \kappa_m-1} $ define $ \kappa_m $ linearly independent functions, which obviously implies that they are associated with $ \kappa_m $ eigenvalues; hence the set of eigenvalues $ \{ \Lambda^{ \varepsilon}_{p(l)}(\eta) \}_{l = 0, \cdots, \kappa_m-1} $ coincides with $ \{ \Lambda^{ \varepsilon}_{m+l}(\eta) \}_{l = 0, \cdots, \kappa_m-1} $ and this concludes the proof of the theorem.



    [1]

    M. S. Birman and M. Z. Solomjak, Spectral Theory of Selfadjoint Operators in Hilbert Spaces, Mathematics and its Applications (Soviet Series). D. Reidel Publishing Co., Dordrecht, 1987.

    [2] Gap opening and split band edges in waveguides coupled by a periodic system of small windows. Math. Notes (2013) 93: 660-675.
    [3]

    D. I. Borisov and K. V. Pankrashkin, Quantum waveguides with small periodic perturbations: Gaps and edges of Brillouin zones, Journal of Physics A: Mathematical and Theoretical, 46 (2013), 235203, 18 pp.

    [4]

    C. Conca, J. Planchard and M. Vanninathan, Fluids and Periodic Structures, RAM: Research in Applied Mathematics, 38. John Wiley & Sons, Ltd., Chichester, Masson, Paris, 1995.

    [5] A strange term coming from nowhere. Topics in the Mathematical Modelling of Composite Materials, Progr. Nonlinear Differential Equations Appl., Birkhäuser, Boston (1997) 31: 45-93.
    [6] Expansion in characteristic functions of an equation with periodic coefficients. Doklady Akad. Nauk SSSR(N.S.) (1950) 73: 1117-1120.
    [7] A boundary value problem for the elliptic equation of second order in a domain with a narrow slit. 1. The two-dimensional case. Math. USSR-Sb. (1976) 28: 459-480.
    [8]

    A. M. Il'in, Matching of Asymptotic Expansions of Solutions of Boundary Value Problems, Translations of Mathematical Monographs, 102. American Mathematical Society, Providence, RI, 1992.

    [9]

    T. Kato, Perturbation Theory for Linear Operators, Die Grundlehren der mathematischen Wissenschaften, Band 132 Springer-Verlag New York, Inc., New York, 1966.

    [10] Boundary value problems for elliptic equations in domains with conical or angular points. Trudy Moskov. Mat. Obshch. (1967) 16: 209-292.
    [11]

    P. Kuchment, Floquet Theory for Partial Differential Equations, Operator Theory: Advances and Applications, 60. Birkhäuser Verlag, Basel, 1993.

    [12]

    N. S. Landkof, Foundations of Modern Potential Theory, Die Grundlehren der Mathematischen Wissenschaften, Band 180. Springer-Verlag, New York-Heidelberg, 1972.

    [13]

    D. Leguillon and E. Sánchez-Palencia, Computations of Singular Solutions in Elliptic Problems and Elasticity, John Wiley & Sons, Ltd., Chichester, Masson, Paris, 1987.

    [14]

    M. Lobo, O. A. Oleinik, E. Perez and T. A. Shaposhnikova, On homogenization of solutions of boundary value problems in domains, perforated along manifolds, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 25 (1998), 611–629.

    [15]

    M. Lobo and E. Pérez, On the local vibrations for systems with many concentrated masses near the boundary, C.R. Acad. Sci. Paris, Ser. IIb, 324 (1997), 323-329.

    [16]

    V. A. Marchenko and E. Y. Khruslov, Homogenization of Partial Differential Equations, Progress in Mathematical Physics, 46. Birkhäuser Boston, Inc., Boston, MA, 2006.

    [17]

    V. Maz'ya, S. Nazarov and B. Plamenevskij, Asymptotic Theory of Elliptic Boundary Value Problems in Singularly Perturbed Domains, Vol. I, Operator Theory: Advances and Applications, 111. Birkhäuser Verlag, Basel, 2000.

    [18] Problème d'écrans perforés pour l'équation de Laplace. RAIRO. Modél. Math. Anal. Numér. (1985) 19: 33-63.
    [19] Asymptotic conditions at a point, self-adjoint extensions of operators and the method of matched asymptotic expansions. Proceedings of the St. Petersburg Mathematical Society, Vol. V, Amer. Math. Soc. Transl. Ser. 2, Amer. Math. Soc., Providence, RI (1999) 193: 77-125.
    [20] The polynomial property of self-adjoint elliptic boundary-value problems and the algebraic description of their attributes. Russ. Math. Surveys (1999) 54: 947-1014.
    [21] Opening of a gap in the continuous spectrum of a periodically perturbed waveguide. Mathematical Notes (2010) 87: 738-756.
    [22] Asymptotic behavior of spectral gaps in a regularly perturbed periodic waveguide. Vestnik St. Petersburg Univ. Mathematics (2013) 46: 89-97.
    [23]

    S. A. Nazarov, R. Orive-Illera and M.-E. Pérez-Martínez, On the polarization matrix for a perforated strip, in Integral Methods in Science and Engineering: Analytic Treatment and Numerical Approximations, Birkhauser, N.Y., (2019), 267–281.

    [24] New asymptotic effects for the spectrum of problems on concentrated masses near the boundary. Comptes Rendues de Mécanique (2009) 337: 585-590.
    [25] On multi-scale asymptotic structure of eigenfunctions in a boundary value problem with concentrated masses near the boundary. Rev. Mat. Complut. (2018) 31: 1-62.
    [26]

    S. A. Nazarov and B. A. Plamenevskii, Elliptic Problems in Domains with Piecewise Smooth Boundaries, De Gruyter Expositions in Mathematics, 13. Walter de Gruyter & Co., Berlin, 1994.

    [27]

    O. A. Oleinik, A. S. Shamaev and G. A. Yosifia, Mathematical Problems in Elasticity and Homogenization, Studies in Mathematics and its Applications, 26. North-Holland Publishing Co., Amsterdam, 1992.

    [28] Higher order asymptotics of solutions of problems on the contact of periodic structures. Mat. Sb. (N.S.) (1979) 110(152): 505-538.
    [29]

    G. Polya and G. Szegö, Isoperimetric Inequalities in Mathematical Physics, Annals of Mathematics Studies, no. 27, Princeton University Press, Princeton, N. J., 1951.

    [30] (1978) Methods of Modern Mathematical Physics. IV. Analysis of Operators. New York-London: Academic Press.
    [31]

    J. Sanchez-Hubert and E. Sánchez-Palencia, Vibration and Coupling of Continuous Systems. Asymptotic Methods, Springer-Verlag, Berlin, 1989.

    [32] Un problème d'ecoulement lent d'un fluide incompressible au travers d'une paroi finement perforée. Homogenization Methods: Theory and Applications in Physics, Collect. Dir. Études Rech. Élec. France, Eyrolles, Paris (1985) 57: 371-400.
    [33]

    M. M. Skriganov, Geometric and arithmetic methods in the spectral theory of multidimensional periodic operators, Trudy Mat. Inst. Steklov., 171 (1985), 122 pp.

    [34] Band spectrum of the Laplacian on a slab with the Dirichlet boundary condition on a grid. Kyushu J. Math. (2003) 57: 87-116.
    [35]

    M. Van Dyke, Perturbation Methods in Fluid Mechanics, Applied Mathematics and Mechanics, Vol. 8 Academic Press, New York-London, 1964.

  • This article has been cited by:

    1. Delfina Gómez, Sergei A. Nazarov, Rafael Orive-Illera, María-Eugenia Pérez-Martínez, Spectral gaps in a double-periodic perforated Neumann waveguide, 2023, 131, 18758576, 385, 10.3233/ASY-221776
    2. D. Gómez, S. A. Nazarov, R. Orive-Illera, M.-E. Pérez-Martínez, Remark on Justification of Asymptotics of Spectra of Cylindrical Waveguides with Periodic Singular Perturbations of Boundary and Coefficients, 2021, 257, 1072-3374, 597, 10.1007/s10958-021-05506-z
    3. Sergei A. Nazarov, Jari Taskinen, Band-gap structure of the spectrum of the water-wave problem in a shallow canal with a periodic family of deep pools, 2023, 36, 1139-1138, 263, 10.1007/s13163-021-00420-z
    4. D. Gómez, S. A. Nazarov, R. Orive-Illera, M. -E. Pérez-Martínez, 2022, Chapter 7, 978-3-031-07170-6, 95, 10.1007/978-3-031-07171-3_7
    5. Delfina Gómez, Sergei A. Nazarov, Rafael Orive‐Illera, María‐Eugenia Pérez‐Martínez, Asymptotic stability of the spectrum of a parametric family of homogenization problems associated with a perforated waveguide, 2023, 0025-584X, 10.1002/mana.202100589
    6. D. Gómez, S. A. Nazarov, M.-E. Pérez-Martínez, Pointwise Fixation along the Edge of a Kirchhoff Plate, 2023, 1072-3374, 10.1007/s10958-023-06862-8
  • Reader Comments
  • © 2019 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(3173) PDF downloads(381) Cited by(6)

Figures and Tables

Figures(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog