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

The band-gap structure of the spectrum in a periodic medium of masonry type

  • Received: 01 August 2019 Revised: 01 April 2020 Published: 12 July 2020
  • Primary: 35J57; Secondary: 35P99, 47B25

  • We consider the spectrum of a class of positive, second-order elliptic systems of partial differential equations defined in the plane R2. The coefficients of the equation are assumed to have a special form, namely, they are doubly periodic and of high contrast. More precisely, the plane R2 is decomposed into an infinite union of the translates of the rectangular periodicity cell Ω0, and this in turn is divided into two components, on each of which the coefficients have different, constant values. Moreover, the second component of Ω0 consist of a neighborhood of the boundary of the cell of the width h and thus has an area comparable to h, where h>0 is a small parameter.

    Using the methods of asymptotic analysis we study the position of the spectral bands as h0 and in particular show that the spectrum has at least a given, arbitrarily large number of gaps, provided h is small enough.

    Citation: Günter Leugering, Sergei A. Nazarov, Jari Taskinen. The band-gap structure of the spectrum in a periodic medium of masonry type[J]. Networks and Heterogeneous Media, 2020, 15(4): 555-580. doi: 10.3934/nhm.2020014

    Related Papers:

    [1] 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
    [2] 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
    [3] Nurehemaiti Yiming . Spectral distribution and semigroup properties of a queueing model with exceptional service time. Networks and Heterogeneous Media, 2024, 19(2): 800-821. doi: 10.3934/nhm.2024036
    [4] 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
    [5] Michele Gianfelice, Marco Isopi . On the location of the 1-particle branch of the spectrum of the disordered stochastic Ising model. Networks and Heterogeneous Media, 2011, 6(1): 127-144. doi: 10.3934/nhm.2011.6.127
    [6] 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
    [7] Zhong-Jie Han, Gen-Qi Xu . Spectrum and dynamical behavior of a kind of planar network of non-uniform strings with non-collocated feedbacks. Networks and Heterogeneous Media, 2010, 5(2): 315-334. doi: 10.3934/nhm.2010.5.315
    [8] Maria Cameron . Computing the asymptotic spectrum for networks representing energy landscapes using the minimum spanning tree. Networks and Heterogeneous Media, 2014, 9(3): 383-416. doi: 10.3934/nhm.2014.9.383
    [9] Denis Mercier . Spectrum analysis of a serially connected Euler-Bernoulli beams problem. Networks and Heterogeneous Media, 2009, 4(4): 709-730. doi: 10.3934/nhm.2009.4.709
    [10] Delfina Gómez, Sergey A. Nazarov, Eugenia Pérez . Spectral stiff problems in domains surrounded by thin stiff and heavy bands: Local effects for eigenfunctions. Networks and Heterogeneous Media, 2011, 6(1): 1-35. doi: 10.3934/nhm.2011.6.1
  • We consider the spectrum of a class of positive, second-order elliptic systems of partial differential equations defined in the plane R2. The coefficients of the equation are assumed to have a special form, namely, they are doubly periodic and of high contrast. More precisely, the plane R2 is decomposed into an infinite union of the translates of the rectangular periodicity cell Ω0, and this in turn is divided into two components, on each of which the coefficients have different, constant values. Moreover, the second component of Ω0 consist of a neighborhood of the boundary of the cell of the width h and thus has an area comparable to h, where h>0 is a small parameter.

    Using the methods of asymptotic analysis we study the position of the spectral bands as h0 and in particular show that the spectrum has at least a given, arbitrarily large number of gaps, provided h is small enough.



    Let Ω0=(l1,l1)×(l2,l2) be a rectangle containing another, deformed rectangle Ωh with curved sides, Fig. 1, a,

    Ωh={xΩ0:lj+hHj(x3j)<xj<ljhH+j(x3j),j=1,2} (1)
    Figure 1.  The original (a) and limit (b) periodicity cell.

    where h>0 is a small parameter and H±j are positive profile functions, which are smooth in the variable x3j[lj,lj]. We will treat the plane R2 as paved by the shifts

    Ω0(θ)={x=(x1,x2):(x1l1θ1,x2l2θ2)Ω0},θ=(θ1,θ2)Z2,Z={0,±1,±2,} (2)

    of the periodicity cell Ω0. In the plane, we consider the spectral problem for an elliptic system of second-order differential equations

    Lh(x,)uh(x)=λhBh(x)uh(x),xR2, (3)

    and its variational form

    ah(uh,vh;R2)=λhbh(uh,vh;R2)vhH1(R2)J. (4)

    Let us explain the notation. The number λh is a spectral parameter and uh=(uh1,,uhJ) is a column of functions so that stands for the transposition. In (3), Lh(x,) is a J×J-matrix of differential operators,

    Lh(x,)=D()Ah(x)D(), (5)

    where D() is a N×J-matrix of first-order homogeneous differential operators with constant (complex) coefficients and D()=¯D() is the adjoint of D(), while Ah and Bh are Hermitian matrix functions of sizes N×N and J×J, respectively. These matrices are assumed piecewise constant and of high contrast, depending on the small parameter h and the subdomain Ωh,

    Ah(x)=A,Bh(x)=BforxΩh,Ah(x)=hαAA,Bh(x)=hαBBforxΓh=Ω0¯Ωh. (6)

    where A, A and B, B are positive definite, constant matrices. Moreover, Ah and Bh are extended periodically to the plane R2:

    Ah(x1l1θ1,x2l2θ2)=Ah(x),  Bh(x1l1θ1,x2l2θ2)=Bh(x)

    for all xΩ0 and θZ2.

    Furthermore, in (4) ah and bh are Hermitian sesquilinear forms,

    ah(uh,vh;Ξ)=(AhD()uh,D()vh)Ξ,bh(uh,vh;Ξ)=(Bhuh,vh)Ξ, (7)

    where (,)Ξ is the natural scalar product in the Lebesgue space L2(Ξ)m, which is either scalar (m=1), or vectorial (m>1). In (4), H1(Ξh) stands for the standard Sobolev space, while the superscript J indicates the number of the components of the test vector functions vh; this superscript is omitted in the notation of norms and scalar products.

    As a consequence of our assumptions, the operator (5) is formally self-adjoint and the forms (7) are positive.

    The main goal of our paper is to describe the asymptotic behavior of the spectrum of the problem (4) when h+0. For the formal asymptotic procedure, we assume that the exponents in (6) satisfy

    αB+2>αA>1. (8)

    In the framework of elastic materials, see Example 1.2, this means that the material in Ωh is much harder and much heavier than in Γh, provided αB(1,0). This kind of structure appears in many natural and man-made elastic composites, e.g. quartzite and brick masonry.

    We restrict ourselves to treating only the main asymptotic term in the expansion of eigenvalues. Thus, the formal asymptotic structures, which are derived in Section 2, are sufficient for the justification of the asymptotics only under the additional assumptions

    αA<2,αB+1>αA. (9)

    These will be accepted in Section 3 in order to avoid the construction of higher order asymptotic terms and to simplify the proofs (see Section 2.4 for a generalization of the presented results).

    We assume that the matrix D(ξ) is algebraically complete [24]: there exists a positive integer ρDN={1,2,3,} such that, for any row p(ξ)=(p1(ξ),,pJ(ξ)) of homogeneous polynomials of degree ρρD, one can find a row q(ξ)=(q1(ξ),,qN(ξ)) of polynomials satisfying

    p(ξ)=q(ξ)D(ξ)ξR2. (10)

    Furthermore, the form ah in (7) possesses the polynomial property [17], namely, there exists a finite dimensional space P of polynomials in x such that for any domain ΞR2 there holds the equivalence

    uH1(Ξ)J,D()u(x)=0,xΞuP|Ξ. (11)

    In other words, the quadratic energy form ah degenerates only for some polynomials. From (10) it follows that all polynomials in P are of degree at most ρD1.

    As proved in [24,Thm 3.7.7], the property (10) assures the Korn inequality

    uh;H1(Ω0)2cD(D()uh;L2(Ω0)2+uh;L2(Ω0)2),

    where the constant cD is independent of uH1(Ω0), so that the sum

    uh,vhh=ah(uh,vh;R2)+bh(uh,vh;R2) (12)

    is a scalar product in the Hilbert space H=H1(R2)J. We introduce the positive definite, symmetric and continuous, therefore, self-adjoint, operator Th by the identity

    Thuh,vhh=bh(uh,vh;R2)uh,vhH, (13)

    which reduces the problem (4) to the abstract equation

    Thuh=τhuhinH (14)

    with the new spectral parameter

    τh=(1+λh)1. (15)

    The spectrum Sh of the operator Th is contained to the segment [0,1] where th=1 is the norm of Th, see [20,Remark 1]. The set

    σh={λh:(1+λh)1Sh} (16)

    is regarded as the spectrum of the problem (4).

    The structure of the spectrum (16) is described by the Floquet–Bloch–Gelfand- (FBG-) theory, see, e.g. [11,25,12], which yields for it the band-gap structure

    σh=nNβhm. (17)

    Here, the bands

    βhm=[βhm,βhm+]={Λhm(η)|η=(η1,η2)Y} (18)

    are formed by the eigenvalue sequence

    0Λh1(η)Λh2(η)Λhm(η)+ (19)

    of the model problem in the periodicity cell

    ah(Uh(;η),Vh(;η);Ω0)=Λh(η)bh(Uh(;η),Vh(;η);Ω0)Vh(;η)H(η)=H1η(Ω0)J (20)

    depending on the Floquet parameter η=(η1,η2), which is the Gelfand dual variable belonging to the rectangle

    Y=[0,1/l1)×[0,1/l2).

    In (20), H1η(Ω0) is the subspace of functions UhH1(Ω0) subject to the quasi-periodicity conditions

    Uh(l1,x2;η)=e2πiη1l1Uh(l1,x2;η),|x2|<l2,Uh(x1,l2;η)=e2πiη2l2Uh(x1,l2;η),|x1|<l1. (21)

    Using an argument similar to (13)–(15) and recalling the compactness of the embedding H1(Ω0)L2(Ω0), we conclude that the spectrum of the model problem (20) is discrete and consists of the positive monotone unbounded sequence (19), cf. [3,Thm 10.1.5,10.2.2], while the corresponding eigenvectors Uh(m)(;η)H(η) can be subject to the normalization and orthogonality conditions

    bh(Uh(m),Uh(n))=δm,n, (22)

    where δm,n is the Kroneker symbol. Moreover, the functions YηΛhm(η) are continuous and periodic with the periods 1/lj in ηj so that the bands (18) indeed are compact intervals.

    The variational problem (20) is obtained from (4) by the FBG-transform [5]. All objects related to the model problem are denoted by capital letters; in particular, {Λh(η),Uh(x;η)} is a new notation for eigenpairs.

    The spectral bands (18) may overlap, but between them there can also exist gaps, i.e., nonempty open intervals γhm=]βhm,βhm+1[ which are free of the spectrum. In the paper [20] it was proved that the spectrum of the problem (4) has at least one open gap of width O(1). In what follows, we will describe the asymptotic structure of the low-frequency range {λhσh:λhconst} of the spectrum (17). The results imply the existence of a large number of open gaps, the geometric characteristics of which will also be described asymptotically.

    Let us list some concrete problems in mathematical physics which have the properties assumed above. Other important examples will be discussed in Sections 4.4 and 4.5.

    Example 1.1. Let J=1 and N=2. Then D()= and (5) is a scalar elliptic second-order differential operator in the divergence form. Clearly, P=R in (11). The problem (3) describes e.g. a heterogeneous acoustic medium with thin high-conductive streaks.

    Example 1.2. Let J=2, N=3 and

    D()=(1021/220221/21). (23)

    It is known that in this case ρD=2 in the property (10), cf. [24,§3.7], [17,Example 1.12]. In the Voigt–Mandel notation of elasticity, u=(u1,u2) is the displacement vector, D()u and AD()u are the strain and stress columns and A is a real, symmetric and positive definite 3×3-matrix of elastic moduli. Furthermore, B=diag{b,b} and b>0 is the mass density of the elastic material. The space of polynomials (11)

    P={u:u1(x)=c1c0x2,u2(x)=c2+c0x1,cpR} (24)

    consists of rigid motions.

    The problem (3) describes elastic composites, some of which were already mentioned in Section 1.1.

    The band-gap structure of the spectrum of an elliptic equation

    Aε(x)uε(x)=λεuε(x),xRd,d2,

    with highly contrasting coefficients

    Aε(x)={ε1A,xωA,xΩ¯ω (25)

    which are 1-periodic in all coordinates xj, was first investigated in the paper [6], where the existence of a non-empty spectral gap was proved; see also [7]. Notice that the coefficients of d×d-matrix (25) become large in an interior subdomain ω of the unit open cube Ω, ¯ωΩ (ε>0 is a small parameter). A similar problem was considered in [28], and there it was in addition shown that the number of open gaps grows unboundedly when ε+0.

    The subdomain ω is fixed in the definition (25), but the papers [20,1] deal with the situation described in Section 1.1: the subdomain (1) of the high contrast covers the whole periodicity cell Ωh in the limit h+0.

    In the present paper we employ an asymptotic method which differs quite significantly from the analysis used in [6,7,28,22,2] and [20,1].

    First of all, we rewrite the problem (20) in differential form. In view of (5) and (6), it consists of two systems

    ¯D()AD()Uh(x;η)=Λh(η)BUh(x;η),xΩh, (26)
    hαA¯D()AD()Uh(x;η)=Λh(η)hαBBUh(x;η),xΓh, (27)

    coupled by the transmission conditions

    Uh(x;η)=Uh(x;η),xΣh=Ωh, (28)
    ¯D(νh(x))AD()Uh(x;η)=hαA¯D(νh(x))AD()Uh(x;η),xΣh, (29)

    where Uh and Uh are the restrictions of Uh onto Ωh and Γh, respectively, and νh(x) is the exterior normal unit vector on the boundary Σh of Ωh, i.e. the interface. The sides of the rectangle Ω0 are supplied, in the terminology of [14], with the stable quasi-periodicity conditions (21) as well as the intrinsic conditions

    1Uh(l1,x2;η)=e2πiη1l11Uh(l1,x2;η),|x2|<l2,2Uh(x1,l2;η)=e2πiη2l22Uh(x1,l2;η),|x1|<l1, (30)

    where j=/xj, j=1,2.

    Since the right-hand-side of (29) includes the small coefficient hαA, the passing to the limit h0 turns Σh into Σ0=Ω0 (see (1)) and leads to the limit problem

    ¯D()AD()U0(x)=Λ0BU0(x),xΩ0, (31)
    ¯D(ν0(x))AD()U0(x)=0,xΣ0=Ω0. (32)

    Notice that the Floquet parameter η does not appear in this problem because the quasi-periodicity conditions (21), (30) are isolated from the interior part Ωh by a thin "shim" Γh. The variational formulation of this problem is written as the integral identity [14]

    a(U0,V0;Ω0)=Λ0b(U0,V0;Ω0)V0H1η(Ω0)J. (33)

    The spectrum of the problem (31), (32) or (33) is discrete, as it consists of the eigenvalues

    0Λ01Λ02Λ0m+, (34)

    while the corresponding vector eigenfunctions U0(m)H1(Ω0)J can be subject to the orthogonality and normalization conditions

    (BU0(m),U0(p))Ω0=δm,p,m,pN. (35)

    We emphasize that the multiplicity of the null eigenvalue in (34) equals dimP, by (11).

    Let {Λ0,U0} be an eigenpair of the limit problem (31), (32). Due to the boundary condition (32), the eigenfunction U0 leaves only a small discrepancy in the intrinsic transmission conditions (29), but the discrepancy in the stable transmission condition (28) is of order 1=h0. To compensate the latter, we need to construct a boundary layer W in the thin bordering Γh. We consider two vertical curved strips

    Γh1±={x:x2(l2,l2),l1>±x1>l1hH±1(x2)}

    and denote by W1 the restriction of W onto Γh1=Γh1Γh1+. Notice that due to the periodicity, the set Γh1 can be identified with

    {x:|x2|<l2,l1hH+(x2)<x1<l1hH(x2)}. (36)

    As always in the asymptotic theory of elliptic problems in thin domains, we introduce the rapid variable

    ζ=h1(x1l1)inΓh1+,ζ=h1(x1+l1)inΓh1, (37)

    so that, according to (1) and (36),

    ζΥ1(y)=(H+1(y),H1(y))

    while y=x2 is still a slow variable. The boundary layer term W1 depends on the coordinate couple (ζ,y). Inside the set Γh2=Γh2Γh2+ we define the corresponding coordinates and the boundary layer term W2 analogously.

    We have

    D()=h1D(e(1)ζ)+D(e(2)y),e(j)=(δ1,j,δ2,j).

    Thus, the left-hand side of (27) takes the form

    hαA2¯D(e(1)ζ)AD(e(1)ζ)W1(ζ,y)+, (38)

    where dots stand for higher-order terms that are inessential in our asymptotic analysis, and the right-hand side is

    Λ0hαBBW1(ζ,y)+= (39)

    In other words, by (8), the expression (39) is much smaller than (38). Thus, the boundary layer term W1 satisfies the problem

    ¯D(e(1)ζ)AD(e(1)ζ)W1(ζ,y)=0,ζΥ(y){0}, (40)
    W1(H1(y),y)=U0(l1,y),W1(H+1(y),y)=U0(l1,y), (41)

    together with the following transmission conditions at the point ζ=0

    W1(0,y)=e2πiη1l1W1(+0,y),ζW1(0,y)=e2πiη1l1ζW1(+0,y) (42)

    coming from (30) and (37).

    It follows from (40)–(42) that

    W1(ζ,y)={C0(y)+C1(y)ζforζ>0,e2πiη1l1(C0(y)+C1(y)ζ)forζ<0, (43)

    where the coefficient columns C0(y) and C1(y) can be found from the linear system

    C0(y)+C1(y)H1(y)=U0(l1,y),
    C0(y)C1(y)H+1(y)=e2πiη1l1U0(l1,y).

    Thus,

    C1(y)=H1(y)1[U0]1(y;η1), (44)

    where

    Hj(x3j)=H+j(x3j)+Hj(x3j)[U0]j(x3j;ηj)=U0(x)|xj=lje2πiηjljU0(x)|xj=lj. (45)

    We will not need an explicit expression for the coefficient C0(y).

    Let us return to the transmission conditions. By (1), the normal vector on Σh1± equals

    νh(x)=(1+h2|2H±1(x2)|2)1/2(±1,h2H±1(x2))=±e(1)+O(h), (46)

    hence, the main asymptotic term of the right-hand side of (29), calculated for the boundary layer term W1(ζ,x2), reads as

    hαA1¯D(±e(1))AD(e(1))ζW1(ζ,x2)|ζ=H±1(x2)==hαA1A(1)1H1(x2){e2πiη1l1[U0]1(x2;η1)atΣh1+,[U0]1(x2;η1)atΣh1, (47)

    where Σhj±=ΣhΓhj± and A(1) is a Hermitian and positive definite J×J- matrix; we denote

    A(j)=¯D(e(j))AD(e(j)),j=1,2. (48)

    The coefficient hαA1 is small, since its exponent is positive in view of our assumption (8). We thus readily accept the asymptotic ansätze

    Uh(x;η)=U0(x)+hαA1U(x;η)+, (49)
    Λh(η)=Λ0+hαA1Λ(η)+ (50)

    The correction term satisfies the problem

    ¯D()AD()U(x;η)Λ0BU(x;η)=Λ(η)BU0(x),xΩ0, (51)
    ¯D(±e(j))AD()U(x;η)=Fj±(x;η),xj=±lj,|x3j|<l3j, (52)

    where the data of the boundary conditions is taken from (47) and a similar formula for j=2 so that

    Fj+(x3j;η)=A(j)1Hj(x3j)e2πiηjlj[U0]j(x3j;ηj)Fj(x3j;η)=A(j)1Hj(x3j)[U0]j(x3j;ηj),j=1,2. (53)

    The variational formulation of the problem (49), (52) reads as

    a(U,V;Ω0)Λ0b(U,V;Ω0)=Λ(η)(BU,V)Ω0+j=1,2±(Fj±,V)Σ0j±
    V0H(η)=H1η(Ω0)J,

    where Σ0j± denote the sides of the rectangle Ω0.

    First, we assume that Λ0=Λ0m is a simple eigenvalue of the problem (31) and hence, by the Fredholm alternative, the problem (51), (52) gets one compatibility condition, namely

    Λm(η)=Λm(η)(BU0(m),U0(m))Ω0=(L()UΛ0mBU,U0(m))Ω0=
    =j=1,2±l3jl3j¯U0(m)(x)¯D(±e(j))AD()U(x;η)|xj=±ljdx3j=
    =j=1,2l3jl3jHj(x3j)1(e2πiηjlj¯U0(m)(x)|xj=lj¯U0(m)(x)|xj=lj)×
    ×A(j)[U0(m)]j(x3j;ηj)dx3j=
    =j=1,2l3jl3jHj(y)1¯[U0(m)]j(y;ηj)A(j)[U0(m)]j(y;ηj)dy,

    and, therefore,

    Λm(η)=J(U0(m),U0(m);η) (54)

    where the Hermitian sesquilinear form J is defined by

    J(U0,V0;η)=j=1,2l3jl3jHj(y)1¯[V0]j(y;ηj)A(j)[U0]j(y;ηj)dy, (55)

    see the notation in (45).

    Second, if Λ0=Λ0m is an eigenvalue of multiplicity ϰm and U0(m),,U0(m+ϰm1) are the related, orthonormalized eigenfunctions, similar calculations show that the correction terms in the ansatz (50) for the eigenvalues Λ0m(η),,Λ0m+ϰm1(η) are nothing but eigenvalues of the Hermitian ϰm×ϰm-matrix Mm with entries

    Mmpq(η)=J(U0(p),U0(q);η),p,q,=m,,m+ϰm1. (56)

    Moreover, the eigenvectors U0(m),,U0(m+ϰm1) can be fixed such that the matrix M(η) with the entries (56) becomes diagonal with eigenvalues

    Λm(η)Λm+1(η)Λm+ϰm1(η). (57)

    In Section 3, we will prove the following error estimates for the asymptotics constructed above. However, one additional albeit reasonable assumption, (68), on the eigenfunction U0(m)H2(Ω0) has to be made; see Section 3.1 for details.

    Theorem 2.1. Let the assumptions (8), (9) and (68) hold true. Then, for any eigenvalue Λ0m of multiplicity ϰm of the limit problem (33),

    Λ0m1<Λ0m==Λ0m+ϰm1<Λ0m+ϰm, (58)

    there exist positive hm and cm such that, for h(0,hm], the sequence (19) contains the eigenvalues Λhn(η),,Λhn+ϰm1(η) of the problem (20) satisfying the estimates

    |Λhn+l(η)Λ0m+lhαA1Λm+l(η)|cmhαA1+δAB,l=0,,ϰm1, (59)

    where

    δAB=min{1αA/2,αB+23αA/2,(αA1)/2,(αB+1)/2,2αA}>0 (60)

    and Λm+l(η) are the corrections terms constructed in Section 2.4, see (54) and (57). Moreover, n=m, i.e. formula (59) includes the asymptotic relationship between the corresponding entries of the eigenvalue sequencies (19) and (34).

    The number (60) is positive because of the restrictions (8) and (9) so that Theorem 2.1 indeed confirms the asymptotic form (50) of the eigenvalues, including the formulas (54)–(57) for the correction term.

    Since the J×J-matrix (48) is positive definite, the matrix Mm is positive. The latter means that the numbers (54) and (57) are non-negative. Thus, formulas (50) and (59) imply that, for a small h>0, the inequalities

    Λhm(η)Λ0m(η),mN,ηY,

    are valid. In other words, the appearence of the flexible thin frame Γh shifts the dispersion surfaces upwards.

    The following description of the spectral bands is an important consequence of Theorem 2.1, and it will have further applications in Section 4.

    Theorem 2.2. If the hypotheses of Theorem 2.1 are true, then the endpoints of the spectral band (18) have the asymptotic form

    |βhm+l±Λ0mhαA1βm+l±|cmhδAB,l=0,,ϰm1, (61)

    where h(0,hm],

    βm+l=minηYΛm+l(η),βm+l+=maxηYΛm+l(η)

    and the correction terms in (61) come from the formulas (54), (55) and (57).

    Relation (61) shows in particular that in the situation (58) there appears an open spectral gap of width

    Λ0mΛ0m1+hαA1(βmβml+)+O(hδAB).

    between the bands βhm1 and βhm. To make conclusions on gaps between the bands βhm,, βhm+ϰ1 we need to put the formulas (56)–(57) into a more conrete form, see Section 3.3.

    The Kondratiev theory of elliptic problems in domains with corners and conical points of the boundary (see the key works [8,15,16,17] and, e.g., the monographs [21,10]) shows that the solution of the problem (31), (32) is of the form

    U0(x)=4j=1χj(x)(Pj(xxj)+Kjk=1CkjΨkj(rj,φj))+˜U0(x), (62)

    where Pjk are some vector polynomials, (rj,φj) are polar coordinates centered at the corner points xj (see Fig. 1.b), φj(0,π/2) in Ω0,

    x1=(l1,l2),x2=(l1,l2),x3=(l1,l2),x4=(l1,l2), (63)

    and χj is a smooth cut-off function supported and equal to 1 in a small neighborhood of xj. Moreover, Ckj is a constant coefficient and Ψkj is a power-logarithmic solution,

    Ψkj(rj,φj)=rμkjjψkj(φj;lnrj) (64)

    where ψkj is a polynomial of degree deg ψkj in lnrj with coefficients, which are smooth functions in the angular variable φj[0,π/2]. The number Kj in (62) can be fixed such that ˜U0H2(Ω0)J and in this way the exponents of power-law solutions (64) satisfy the inequalities

    0<μkj<1+δ0for someδ0>0. (65)

    In view of (65), the polynomials Pj can be reduced to linear vector functions, i.e. (64) with μkj=1 and deg ψkj=0

    Lemma 3.1. In (62) we have

    D()Pj=0PjP. (66)

    Proof. By the above conclusion, we can write

    Pj(x)=Pj(1)x1+Pj(2)x2+Pj(0), (67)

    and here the constant term Pj(0) satisfies (66). The boundary conditions (32) on the sides of the rectangle Ω0 imply

    0=¯D(e(k))A(D(e(1))Pj(1)+D(e(2))Pj(2)),k=1,2,

    and, hence,

    0=¯(D()Pj(x))AD()Pj(x).

    We obtain the formula (66), since A is Hermitian and positive definite.

    To simplify the justification scheme, which would otherwise become too cumbersome, we assume that for some δ0>0 we have in (62)

    Kj=0. (68)

    In other words, non-constant and non-linear power-logarithmic terms (64) are assumed not to exist in the representation (62). Consequently, we have U0H2(Ω0)J. Moreover, Hölder estimates in domains with corner and conical points on the boundary (see the original paper [16] and also the monographs [21,10]) yield the bounds

    |U0(x)|c0,|U0(x)|c1,|nU0(x)|cnrδ0n+1,n=2,3,, (69)

    where r=min{rj|j=1,,4} is the distance of the rectangle Ω0 to the vertices and δ0>0 is the number in (65).

    In Section 4.2, we will consider examples of scalar equations and elasticity systems which meet all the assumptions made above.

    Let us then turn to the equations for the correction term U. The right-hand side of the system (51) is sufficiently smooth but the data (53) of the Neumann conditions (52) does not vanish at the corner points (63) of Ω0. Again, the Kondratiev theory provides for the solution U the decomposition

    U(x)=4j=1χj(x)(Pj(xxj)+CjΨj(rj,φj))+˜U(x), (70)

    where ˜UH2(Ω2)J, the term Pj is a linear vector function in x, and Ψj is of the form (64), where μj=1 and ψj is a linear function of lnrj. Therefore, χjΨjH2(Ω2)J, if a logarithmic term exists. Furthermore, Ψj(rj,φj) is a polynomial in x if and only if, in the Neumann conditions (52), the data (53) frozen at sides of the rectangle Ω0 in the corner point xj can be compensated by a linear vector function (67). This remark allows us to verify the following property.

    Lemma 3.2. The terms Ψj(rj,φj) are always of the form (67) with some Pj(k)</italic><italic>∈Cj, k=0,1,2. Moreover, Pj(0)=0 if and only if P=CJ.

    Proof. Considering the vertex x1 and searching for the linear vector (67), we need to solve the system of 2J algebraic equations

    ¯D(e1)AD(e1)P1(1)+¯D(e1)AD(e2)P1(2)=F1+(l2;η),¯D(e2)AD(e1)P1(1)+¯D(e2)AD(e2)P1(2)=F2(l1;η). (71)

    The (2J×2J)-matrix of this system is Hermitian and, therefore, the system is uniquely solvable provided the homogeneous system (71) has only the trivial solution. According to Lemma 3.1, the latter is true provided the polynomial subspace P in (11) does not contain a non-trivial linear vector function. This completes the proof, since P is invariant with respect to the coordinate translation xx+x0.

    Using the notation of (69), we derive the following estimates for the decomposion (70) :

    |U(x)|c0,|U(x)|c1(1+|lnr|),|2U(x)|c2r1. (72)

    In spite of the singularities these will be sufficient to justify our asymptotic formulas for the eigenvalues (19).

    Similarly to (12), we introduce the scalar product

    Uh,Vhh,η=ah(Uh,Vh;Ωh)+bh(Uh,Vh;Ωh) (73)

    in the space

    H(η)={Uh(;η)H1(Ω0)J:(21)is satisfied}

    and define the self-adjoint and positive operator Th(η) in H(η) by the identity

    Th(η)Uh(;η),Vh(;η)h,η=ah(Uh(;η),Vh(;η);Ωh)+bh(Uh(;η),Vh(;η);Ωh)
    Uh(;η),Vh(;η)H(η).

    In this way, the variational formulation

    ah(Uh(;η),Vh(;η);Ωh)=Λh(η)bh(Uh(;η),Vh(;η);Ωh)Uh(;η)H(η)

    of the problem (26)–(30), (21) turns into the abstract equation in H(η),

    Th(η)Uh(;η)=τh(η)Uh(;η)

    where

    τh(η)=(1+Λh(η))1. (74)

    The operator Th(η) is compact because of the compact embedding H1(Ω0)L2(Ω0) in the bounded domain Ω0, and its spectrum forms the monotone positive sequence (see, e.g., [3,Theorem 10.1.5,10.2.2])

    τh1(η)τh2(η)τhn(η)+0,

    which turns into the sequence (19) by the formula Λh(η)=τh(η)11, cf. (74).

    The following assertion is known as lemma on "near eigenvalues and eigenvectors", cf. [27], and it follows immediately from the spectral decomposition of resolvent, see [3,Ch.6].

    Lemma 3.3. Let Uh(η)H(η) and th(η)(0,+) be such that

    Uh(η);H(η)=1,Th(η)Uh(η)th(η)Uh(η);H(η)=:ϵ(0,th(η)). (75)

    Then at least one eigenvalue τhn(η) of the operator Th(η) satisfies the inequality

    |th(η)τhn(η)|ϵ.

    Moreover, for every ϵ(ϵ,th(η)) one can find coefficients ahM(η),,ahM+X1(η) such that

    Uh(η)M+X1q=Mahq(η)Uh(q)(η);H(η)2ϵϵ,M+X1q=M|ahq(η)|2=1 (76)

    where τhM(η),,τhM+X1(η) are all the eigenvalues of Th(η) contained in the interval [th(η)ϵ,th(η)+ϵ] (the numbers M and X may dependent on h), and Uh(M)(η),,Uh(M+X1)(η)H(η) are the corresponding eigenvectors subject to the orthogonality and normalization conditions

    Uh(p)(η),Uh(q)(η)h,η=δp,q. (77)

    Naturally, our next task is the construction of a proper approximate eigenpair {th(η),Uh(η)}. This will be based on the formal asymptotic analysis in Section 2.

    Let Λ0m be an eigenvalue of the limit problem (31), (32) with multiplicity ϰ=ϰm and let the corresponding eigenvectors U0(m),, U0(m+ϰ1) be subject to the normalization and orthogonality conditions (35). We take

    thm(η)=(1+Λ0m+hαA1Λ(η))1 (78)

    as an approximate eigenvalue of the operator Th(η) in (74). Here, =m,,m+ϰ1 and Λ(η) is the correction term constructed in Section 2.1.

    The approximate eigenvectors are defined in Ωh as

    uh()=U0()(x)+hαA1U()(x). (79)

    However, the definition becomes much more complicated inside the thin frame Γh and it involves several smooth cut-off functions. First of all, we select a cut-off function whose support covers almost the whole cell:

    Xh=1inΩh,Xh=0in a neighborhood ofΩ0; (80)
    0Xh1,|Xh|cXh1. (81)

    Then, the function χhj±C(¯Ω0) is defined such that the support is contained in ¯Γhj± and

    χhj±(x)=1for|xj|<lj2ϱh,χhj±(x)=0for|xj|>ljϱh,

    where ϱ>0 is chosen such that suppχhj±(Γh3jΓh3j+)=. Finally, Xh1 is supported in a small neighborhood V1 of the corner point x1 of the rectangle Ω0, vanishes in the vicinity of Σ0 and equals 1χh2χh2+ near the curve ΣhV1. The relations (81) hold true for all these cut-off-functions.

    Completing the definition (78) of uh(), we set in Γh1+

    uh()(x)=wh()(x)+hαA1uh()(x),wh()(x)=(Xh1(x)+Xh2(x))U0()(x)+χhj+(x)W1(h1(x1l1),x2),uh()(x)=Xh(x)U()(x). (82)

    where U() is the correction term in the ansatz (49), i.e. the solution of the problem (51), (52) with data (53). The definitions (82) can clearly be extended to Γh1, Γh2± and thus to the whole frame Γh.

    Owing to the boundary conditions (41), relations (42) and the definitions of the cut-off functions, we conclude that (82) satisfies the quasi-periodicity conditions (21) and has the same trace on Σh=Ωh as the approximate solution (79).

    We proceed to evaluate the scalar product uh(),uh(n)h,η, see (73). The inequality

    |a(uh(),uh(n);Ωh)+b(uh(),uh(n);Ωh)δ,n(1+Λ0m)|ch (83)

    follows from the formulas (79) for uh(), (35), (33) and (69) for the eigenfunction U0() and (72) for the correction term U() as well as the fact that area of Γh is of order h.

    Now, to derive the formulas

    |uh(),uh(n)h,ηδ,n(1+Λ0m)|chmin{1,αA1},uh();H(η)cm>0, (84)

    we recall (6) and write the estimate

    |hαAa(uh(),uh(n);Γh)+hαBb(uh(),uh(n);Γh)|c(hαA(1+(1+hαA1|lnh|)h1)2h1+hαBh1)chαA1.

    Here, in the central expression, the factor h1 comes from the differentiation of the cut-off functions and the boundary layer term (cf. (81) and (37)). The factor h1 is the order of the area of Γh and hαA1|lnh| results from the gradient estimate (72) of hαA1U(). Notice that the exponent of h in the bound (84) equals αA1, owing to (9).

    Let us estimate the value ϵ=ϵ in (75) for the number (78) and the vector

    Uh()=uh();H(η)1uh()H(η). (85)

    We have

    ϵ=ThUh()th()Uh();H(η)=sup|ThUh()th()Uh(),vhh,η|==uh();H(η)1th()sup|ah(uh(),vh;Ωh)(Λ0m+hαA1Λ)bh(uh(),vh;Ωh)| (86)

    where the supremum is calculated over the unit ball in H(η). Since the first two factors on the right-hand side of (86) are uniformly bounded in h, see (68) and (84), it suffices to consider the expression

    ah(uh(),vh;Ωh)(Λ0m+hαA1Λ(η))bh(uh(),vh;Ωh)=Ih+IhA+IhB+IhA+IhB+IhΣ (87)

    where

    Ih=((L(Λ0m+hαA1Λ(η)))(U0m+hαA1U()),vh)ΩhIhA=hαA(Lwh(),vh)Γh,  IhB=hαB(Λ0m+hαA1Λ(η)(Bwh(),vh)ΓhIhA=h2αA1a(uh(),vh;Γh), IhB=hαB+αA1(Λ0m+hαA1Λ(η))b(uh(),vh;Γh)IhΣ=(D(νh)(AD()uh()hαAAD()wh()),vh)Σh. (88)

    For the first term Ih we have, according to (31) and (52),

    |Ih|=h2(αA1)|Λ(η)(U()),vh)Ωh|ch2(αA1). (89)
    Figure 2.  Additional geometric objects.

    We will need some additional inequalities for the processing of the other terms in (88). To this end, we introduce eight geometric figures depicted in Fig. 3, namely four squares hk attached to the vertices xk (k=1,,4) having side lengths ϱh, and four rectangles Πhj± with short and long sides of lengths ρh and l3j2ϱh, respectively. Here, the coefficient ϱ>0 of the parameter h is chosen such that the union of the eight sets contains Γh and a ch-neighborhood of Σh=Ωh for some c>0 which we fix now. In the following we will use the symmetry to replace some estimates over the set Γh by corresponding estimates only over the rectangle Πh1+ and the squares h1 and h2. We denote h1,2=Πh1+h1h2.

    Figure 3.  Pavements of different shapes.

    Lemma 3.4. If vhH(η), then there holds the inequalities

    vh;L2(hk)chhαA/2vh;H(η), (90)
    vh;L2(Πhj±)ch1/2hαA/2vh;H(η), (91)

    where c is independent of vh and h(0,h].

    Proof. Let h1={x:r(0,d1),ϕ(0,ϕ1)}Ωh be the sector which is shown in Fig. 2.a and which contains the small triangle h1 (marked with black in Fig. 3.a) inside the square h1. The estimate

    r1|lnr|1vh;L2(h1)cvh;H1(h1) (92)

    is a consequence of the classical one-dimensional Hardy inequality

    d0r2|Vh(r)|2drcd0(|dVhdr(r)|+|Vh(r)|2)rdr

    integrated in the angular variable ϕ. Taking into account the weight factor on the left-hand side of (92), we can write

    h2vh;L2(h1)2c|lnh|2vh;H1(Ωh)2

    Then, we apply the Poincaré ineqality

    h2vh;L2(h1)2c(vh;L2(h1)2+h2vh;L2(h1)2)

    which can be easily derived by a coordinate dilation, and obtain the desired inequality (90) as follows:

    h2vh;L2(h1)2c(a(vh,vh;Γh)+|lnh|2(a(vh,vh;Ωh)+b(vh,vh;Ωh))
    cmax{hαA,|lnh|2}vh;H(η)2=chαAvh;H(η)2.

    To derive (91), we employ the Newton–Leibnitz formula and write

    vh(l1ϱh+t),x2)=vh(0,x2)+l1ϱh+t0vhx1(x1,x2)dx1.

    We estimate

    l1l1ϱh|vh(t,x2)|2dt2l1l1ϱh|vh(0,x2)|2dx+2l1l1l1ϱhl10|vhx1(x1,x2)|2dx1dt
    2ϱh|vh(0,x2)|2+2l1ϱhl10|vhx1(x1,x2)|2dx1.

    It suffices to integrate x2 over the interval (l2+ϱh,l2ϱh) and to apply the standard trace inequality.

    Since vh;H(η)=1 in (86), we can use in the following calculations the fact that the left-hand sides of (90) and (91) are uniformly bounded in h. Moreover, we denote by IhA() and IhB() the quantities where the inner products of the expressions IhA and IhB, respectively, are taken over the set h1,2Γh instead of Γh. For IhA() we obtain the bound

    |IhA()|chαA((2U0(m);L2(h1,2)+h1U0(m);L2(h1,2)++h2U0(m);L2(h1,2))vh;L2(h1,2)+L0(χ1+W1);L2(Πh1+)vh;L2(γh1+))chαA((hδ1+h1+h2)h1hhαA/2+(h1/2h1h1/2+h1h2h1h)hαA/2)chαA/2. (93)

    Here, the norms h2ppU0(m);L2(h1,2), p=0,1,2, came from the differentiation of the product XhkU0() and (81); they were estimated by using (69) and taking into account the area O(h2) of hk. In a similar way we used (40) and (43), (44), (69) to get

    L0(χ1+W1)=χ1+L0W1+[L0,χ1+]W1,
    |χ1+L0W1|ch1,supp(χ1+L0W1)γh1+h1,2,
    |[L0,χ1+]W1|ch2,supp([L0,χ1+]W1)h1,2.

    Finally, inequalities (90) and (91) for vh together with bounds for the areas of the above-mentioned supports were used to complete the estimate (93). A much simpler consideration yields the estimate

    |IhB()|chαB(U0(m);L2(h1,2)vh;L2(h1,2)+W1;L2(γh1+vh;L2(γh1+))chαB(h1h+h1/2h1/2)hαA/2ch1+αBαA/2. (94)

    Notice that the exponents of the bounds in (93) and (94) are included in the formula (60).

    In view of symmetry, the same estimates hold when h1,2 is replaced by the unions of the other rectangles hk and Πhj± (see the explanation above Lemma 3.4). Since the frame Γh is covered by these sets, we obtain

    |IhA|chαA/2 ,   |IhB|ch1+αBαA/2. (95)

    To treat the fourth and fifth terms IhA and IhB on the right-hand side of (87), we apply formulas (72) and the basic estimates following directly from definitions (73) and (6),

    vh;L2(Γh)chαA/2,vh;L2(Γh)chαB/2.

    Accordingly, we obtain

    |IA|ch2αA1(h1U();L2(Γh)+U();L2(Γh))vh;L2(Γh)ch2αA1((h1h1/2+|lnh|h1/2)hαA/2ch3(αA1)/2,|IB|chαB+αA1h1/2hαB/2chαA1+(αB+1)/2, (96)

    Exponents in both bounds are included in (60).

    Let us consider the last term IhΣ in (88). Here, our assumption on the smoothness properties in Section 3.1 plays an important role. Since the vector function U0()C2(Ω0) satisfies the homogeneous Neumann conditions (32), we have

    |D()U0()(x)|ch,xΣh,|(D(νh)AD()U0(),vh)Σh|chvh;L2(Σh)chvh;H1(Ωh)chvh;H(η)ch. (97)

    Furthermore, the boundary condition (52), (53), formula (46) for the normal vector νh and the relations (72) for U() imply that, for xΣh,

    |D(νh)AD()U()(x)D(νh)AD()W1(H+(x2),x2)|chr1 (98)

    where the last singular factor r1 is caused by the second derivatives of U(). Now we use the known weighted trace inequality

    r1/2|lnr|1vh;L2(Σh)c(vh;L2(Ωh)+r1|lnr|1vh;L2(Ωh)),

    cf. (92) and (97), (98), to obtain

    |IhΣ|c(h+hαA1hΣhr1|vh(x)|ds)ch(1+hαA1(r+h)1vh;L2(Σh))ch(1+hαA1|lnh|)ch3(αA1)/2, (99)

    where we also applied the inequalities rr+ch, c>0 on Σh, see Fig. 2.a, and h|lnh|ch(αA1)/2, due to (9).

    Combining the estimates (89), (95), (96), (99) and recalling the definition (60) yield for the number (86) the estimates

    ϵcmhαA1+δAB,=m,,m+ϰm1, (100)

    which according to Lemma 3.3 means that the operator Th(η) has an eigenvalue τhn()(η) related to the "almost eigenvalues" (78) by

    |τhn()(η)th(η)|chαA1+δAB.

    If all eigenvalues (57) of the matrix Mn(η) are simple, in particular, Λ0m is a simple eigenvalue in the sequence (34), then the distance of any two points thm(η),,thm+ϰm1(η) is at least ChαA1+δAB, and thus there exist ϰm different eigenvalues τhn(m)(η),,τhn(m+ϰm1)(η). However, in the case

    Λm1(η)<Λm(η)==Λm+x1(η)<Λm+x(η)

    with x>1, Lemma 3.3 would not prevent that some of the eigenvalues τhn(m)(η),, τhn(m+x1)(η) might coincide. We show that they can be chosen to be different from each other.

    Taking into account the second assertion in Lemma 3.3, we set ϵ=max{ϵm,,ϵm+x1} and

    ϵ=Tϵ, (101)

    where T is fixed to be large enough, and denote by τhM,,τhM+X1 all eigenvalues of the operator Th in the interval

    υhm=[thm(η)ϵ,thm(η)+ϵ]. (102)

    For each =m,,m+x1, Lemma 3.3 gives a column of coefficients ah()=(ahM,, ahM+X1) such that the relations (76) are valid for the vector function (85). We denote by Sh the linear combination

    Sh=ahMUhM++ahM+X1UhM+X1

    and write

    Uh,Uhpδ,p=UhSh,Uhp+Sh,UhpShp+Sh,Shpδ,p.

    Inequalities (84), (76), (100) and conditions (77) yield

    |¯ah(p)ah()δ,p|=|Sh,Shpδ,p|c(ϵ+ϵ+ϵϵ1)
    c(hαA1max{cm,,cm+x1}+T1).

    In other words, the columns ahm,,ahm+x1CX are almost orthonormalized for small h and big T. The latter situation can only happen in the case Xx, when the interval (102) contains at least x eigenvalues of the operator Th(η) which meet the estimate

    |τhn+q(η)thm(η)|TcmhαA1+δAB,q=1,,x, (103)

    see (101), (102), (100). Definition (78) and relation (74) between the spectral parameters, turn formula (103) into the desired inequality (59).

    Finally, the assertion on the equality of n and m in Theorem 2.1 follows from a standard convergence theorem which we formulate as the next lemma and prove in the next section.

    Lemma 3.5. The entries of the eigenvalue sequencies (19) and (34) are related by ΛhmΛ0m, as h+0.

    To prove Lemma 3.5, let {Λhm(η),Uhm(;η)} be an eigenpair of the problem (20) for some ηY. We fix this Floquet parameter and suppress it in the notation from now on. In Section 3.3 it was proved that in the vicinity of each eigenvalue Λ0p, p=1,,m, of the limit problem (33) there exists an eigenvalue Λhn(p) of the problem (20) and n(p1)n(p2) for p1p2, and we have

    ΛhmΛhn(m)Λ0m+cmhαA1Cm.

    We normalize the eigenvector Uh(m) by

    bh(Uh(m),Uh(m);Ω0)=1.

    The integral identity (20) and formulas (6), (7) yield the implication

    ah(Uh(m),Uh(m);Ω0)cmUh(m);H1(Ω(ϱ))cm, (104)

    where the number ϱ>0 is chosen such that the rectangular domain

    Ω(ϱ)={xΩ0:|xj|<lj(1ϱh))}

    is contained in the domain Ωh.

    According to (104), the vector function

    Ω0xUh(x)=Uh(m)((1ϱh)1x) (105)

    has a uniformly bounded H1(Ω0)-norm and, hence, there exists a positive sequence {hn}nN tending to 0 such that

    ΛhnmΛ0,UhnU0 weakly inH1(Ω0) and stronly inL2(Ω0) as n+. (106)

    We take an arbitrary test function VC(¯Ω0), extend it smoothly outside the square Ω0 and insert into the integral identity (20) the function

    Vh=XhVh, (107)

    where Vh(x)=V((1ϱh)1x). Notice that the function (107) vanishes in the vicinity of Ω0, cf. (80), and therefore it belongs to H(η).

    We have

    0=ah(Uh(m),Vh;Ω0)Λhmbh(Uh(m),Vh;Ω0)==a(Uh(m),Vh;Ω(ϱ))Λhmb(Uh(m),Vh;Ω(ϱ))++a(Uh(m),Vh;Ω0Ω(ϱ))Λhmb(Uh(m),Vh;Ω0Ω(ϱ)). (108)

    Using (105) and (107), (80) we observe that

    a(Uh(m),Vh;Ω(ϱ))=Ω(ϱ)¯D(x)Uh((1ϱh)1x)AD(x)Vh((1ϱh)1x)dx=
    =a(Uh,V;Ω0)a(U0,V;Ω0),
    Λhb(Uh,Vh;Ω(ϱ))=Λh(1ϱh)2b(Uh,Vh;Ω0)Λ0b(U0,V;Ω0).

    To process the remaining terms in (108) we write

    |ah(Uh,Vh;Ω0Ω(ϱ))|c(xUh;L2(ΩhΩ(ϱ))xVh;L2(ΩhΩ(ϱ))+
    +hαAxUh;L2(Γh)xVh;L2(Γh))c(xUh;L2(ΩhΩ(ϱ))h1/2+
    +hαA/2xUh;L2(Γh)hαA/2h1/2(1+h1))c(h1/2+h1+(1+αA)/2)0,
    Λh|bh(Uh,Vh;Ω0Ω(kh))|c(Uh;L2(ΩhΩ(kh))Vh;L2(ΩhΩ(kh))+
    +hαBUh;L2(Γh)Vh;L2(Γh))c(h1/2+h(αB+1)/2)0.

    Thus, passing to the limit hn+0 yields the integral identity for the limit problem,

    a(U0,V;Ω0)Λ0b(U0,V;Ω0)=0    VC(¯Ω0).

    By a completion argument, the test function space can be changed here to be H1(Ω)J.

    Hence, to conclude that {Λ0,U0} is an eigenpair of the problem (31), (32), it is sufficient to verify that U0 is non-zero. To this end we write

    1=bh(Uh(m),Uh(m);Ω0)==b(Uh(m),Uh(m);Ω(ϱ))+b(Uh(m),Uh(m);ΩhΩ(ϱ))+hαBb(Uh(m),Uh(m);Γh). (109)

    Lemma 3.4 gives estimates for the last two terms:

    b(Uh(m),Uh(m);ΩhΩ(ϱ))chUh(m);H1(Ωh)2ChUh(m);H2=Ch,
    hαBb(Uh(m),Uh(m);Γh)hαB+1αAUh(m);H2

    These the upper bounds tend to 0 as h+0, see (9). Recalling (105) yields

    b(Uh(m),Uh(m);Ω(ϱ))=(1ϱh)2b(Uh,Uh;Ω0),

    and we thus obtain b(U0,U0;Ω0)=1 by passing to the limit h+0 in (109).

    Now Lemma 3.5 can be proved in a standard way. Namely, if Uh and Uh are two different eigenvectors of the problem (20) and they are orthogonal in the sense that bh(Uh,Uh)=0, see (22), then the orthogonality b(U0,U0)=0 follows from the above calculations. In this way, supposing n>m would contradict our way to compose the eigenvalue sequences (19) and (34). But we recall that the inequality nm was already verified in Section 3.3. Thus, the identity n=m holds, and this completes the proofs of Lemma 3.5 as well as Theorem 2.1.

    Let us derive exactly the correction terms in the eigenvalue asymptotics (59) for the problems mentioned in Examples 1.1 and 1.2.

    Example 4.1. Let L=Δ and L=Δ, cf. Example 1.1, and let 1/2=l1>l2. Then

    Λ01=0,U0(1)=(2l2)1/2,
    Λ02=π2,U0(2)=l1/22sin(πx1).

    The asymptotic formulas (59), (54) and the definitions (55), (45) show that

    Λh1(η)=0+hαA1j=1,21lj(1cos(2πηjlj))l3jl3jdyHj(y)+O(hαA1+δAB),
    Λh2(η)=π2+hαA12l2((1+cos(2πη1l1))l2l2dx2H1(x2)+
    +(1cos(2πη2l2))l1l1sin2(πx1)dx1H2(x1))+O(hαA1+δAB).

    Example 4.2. Assuming that the frame Γh is made of a homogeneous isotropic material, we have the following 3×3-matrix

    A=(λ+2μλ0λλ+2μ0002μ),

    where λ0 and μ>0 are the Lamé constants, cf. Example 1.2. We choose the orthonormalized basis

    U0(j)(x)=12(l1l2)1/2ej,j=1,2,
    U03(x)=(34)1/2(l1l2)1/2(l21+l22)1/2(x2e1x1e2)

    in the polynomial space P of rigid motions.

    We have

    Λh1(θ)=0+hαA12l1l2((λ+2μ)(1cos(2πl1η1)l2l2dx2H1(x2)+
    +μ(1cos(2πl2η2)l1l1dx1H2(x1))+O(hαA1+δAB),
    Λh2(θ)=0+hαA12l1l2(μ(1cos(2πl1η1)l2l2dx2H1(x2)+
    +(λ+2μ)(1cos(2πl2η2)l1l1dx1H2(x1))+O(hαA1+δAB),
    Λh3(θ)=0+4hαA13l1l2(l21+l22)(l2l2((λ+2μ)(1cos(2πl1η1)x22+
    +μ(1+cos(2πl1η1)l21)dx2H1(x2)+
    +l1l1((λ+2μ)(1cos(2πl2η2)x21+μ(1+cos(2πl2η2)l22)dx1H2(x1))+O(hαA1+δAB).

    The next two examples show that the restrictions introduced in Section 3.1 are relevant in certain problems of mathematical physics.

    Example 4.3. An appropriate affine transform converts the problem (31), (32) into the spectral Neumann problem for the Laplace operator in a parallelogram with angles ϕ(0,π/2] and πϕ[π/2,π). As known for example by [21,Ch. 2], the worst singularity of an eigenfunction of this problem in is

    rπ/(πϕ)cosπφπϕ,πφπϕ=1+δ0,δ0=ϕπϕ>0.

    Example 4.4. According to [13], an affine transform can be used to reduce the stationary elasticity problem (31), (32) with Λ0=0, in Ω0, to the particular case of an orthotropic elastic parallelogram with the elastic symmetry axis of rank 4. This means that the rigidity matrix A is of the form

    A=(a11a120a21a22000a33),a11=a22.

    Singularities at corner points for such orthotropic materials have been computed in, e.g., [26]. However, the inequality μ>1 for the positive singularity exponents in (69) at the tops of the convex angles of the parallelogram has been proved in [9] so that the elasticity problem in Example 1.2 also satisfies our assumption (68).

    Until now we have restricted ourselves to two-dimensional problems, in order to simplify formulas and the justification scheme in Section 3. However, our formal asymptotic analysis would apply also in the multi-dimensional cases without notable changes. Namely, the periodicity cell Ω0={xRd:|xj|<lj,j=1,,d}, d3, can composed of the curved parallelepiped

    Ωh={xΩ0:lj+hHj(x(j))<xj<ljhH+j(x(j)),j=1,,d} (110)

    and the surrounding box-shaped frame Γh=Ω0Ωh. Here, the notation is similar to that in (1) except that x(j)=(x1,,xj1,xj+1,,xd). On the other hand, stating smoothness properties of the eigenvectors, which we used in Section 3, would become much more complicated in higher dimensions, due to the many edges and corners of the boundary of the curved parallelepiped (110). Thus, the justification of the asymptotics might require some additional assumptions.

    Other shapes of the periodicity cells like the honeycomb structure in Fig. 3.a, can be used to cover the plane, and they can be treated with the same asymptotic tools. Another example of a non-rectangular tiling with the periodicity cell in Fig. 3.b requires a modification of our asymptotic procedure, because of the inward obtuse angle. However, the main difficulty is caused in formulas (43)–(45) by the strong corner singularities on the short sides of the curved rectangle Γh+1. The resulting strengthening of the singularities of the solutions may seriously reduce the accuracy of our asymptotic formulas (cf. Section 3.1 and the error estimates in Theorems 2.1 and 2.2).

    The influence of these corner singularities may be compensated by constructing two-dimensional boundary layers (cf. [4,18] for the Poisson equation and [19] for general elliptic problems and the elasticity system). It should be mentioned that using the same scheme as in Section 2.2 one can up to some extend find higher order asymptotic terms in Ωh and in Γhj± outside small neighborhoods of the vertices of Ω0, but it is not possible to specify them completely without the two-dimensional boundary layers. This was the very reason for introducing the assumption (9) which helps to avoid the problem.

    It looks that near concave corner points the required, acceptable approximation can be achieved by constructing two dimensional boundary layers (cf. [19]).

    There is no obstacle to treat strongly curved thin frames in the geometric situation of Fig. 3.c.

    Example 4.5. Let J=3, N=5 and

    D()=(1021/22000221/210000012)=(DM()O2×2O1×3), (111)

    where DM() is the matrix (23) and Om×n is the null matrix of size m×n. Furthermore, let

    A=(AMMAMEAEMAEE) ,   B=(b000b0000), b>0. (112)

    Here, AMM and AEE are symmetric and positive definite matrices of elastic and di-electric moduli, respectively, while there is no restriction on the piezoelectric matrix AME=(AEM) of size 3×2. Finally, u=(uM,uE), uM=(u1,u2) is the displacement vector and uE is the electric potential.

    Due to the minus sign of AEE, the symmetric matrix A is not positive definite and therefore the Hermitian sesquilinear form Ah in (7) does not satisfy condition (11). However, thanks to the right lower null entry of the matrix B, (112), and the Dirichlet condition

    uEh(x)=0,  xΣh, (113)

    on the insulator surface, all necessary conditions are actually satisfied and we conclude in particular that the space of polynomials P can be chosen to be

    P={p=(pM,pE):pMPM, pE=0}, (114)

    where PM is the space (24) of mechanical rigid motions.

    We consider the composite plane ΩhΓh, where Ωh is the union over θZ2 of the identical piezoelectric inclusions Ωh(θ) (cf. (2)), which are connected by thin paddings of pure elastic solid insulator Γh. The boundary value problem consists of the system of differential equations

    D()AD()uh(x)=λhBuh(x) ,  xΩh, (115)
    DM()ADM()uh(x)=λhBuh(x) ,  xΓh, (116)

    the transmission conditions

    DM(νh(x))AD()uh(x)=DM(νh(x))AD()uh(x), xΣh (117)

    and the Dirichlet condition (113) on the union Σh of the contours Σh(θ)=Ωh(θ), θZ2. The notation in (116) and on the right-hand side of (117) is the same as in Example 1.2. We emphasize that condition (117) means that the traction is continuous on the contact surface. Moreover, (113) describes the fact that the electric potential is constant on the surface of the insulator, and this constant can be set to zero because the set ¯Γh=θZ2¯Γh(θ) is connected.

    The limit problem in the rectangle Ω=(1,1)×(2,2) consists of the system (31) including the matrices (111) and (112), the quasiperiodicity condition (21), (30) and the boundary conditions, cf. (32) and (117), (113), and

    DM()AD()U(x)=0,  UE(x)=0,  xΣ. (118)

    Although the matrix A in (112) is not positive definite, the spectrum of the problem is discrete and consists of the monotone non-negative unbounded sequence (34). For example in the paper [23] one can find a procedure for reducing the weak formulation of this piezoelectricity problem to a problem with a positive self-adjoint operator in H1(Ω)UM; the reduction uses the specific structure of the matrix B in (112). The corresponding eigenfunctions U(1),U(2),H1(Ω))3 can be subject to the normalization and orthogonality conditions (35), which read as

    (bUM(m),UM(p))Ω=δm,p,  m,pN,

    with the positive constant b of (112).

    Since the material in Γh is purely elastic, the forms of the second limit problem (40)–(42) and its solution (43) are kept unchanged. Thus, our calculation of the correction term in the eigenvalue ansatz (45) does not need modifications, and also the final formulas (56), (57) remain unchanged, if the formulas (48) and (55) are understood as

    A(j)=¯DM(e(j))ADM(e(j)),J(U,V;η)=j=1,23j3jHj(y)1¯[VM]j(y;ηj)A(j)[VM]j(y;ηj)dy

    with the notation (45) preserved as such.



    [1] Effects of Rayleigh waves on the essential spectrum in perturbed doubly periodic elliptic problems. Integral Equations Operator Theory (2017) 88: 373-386.
    [2] F. L. Bakharev and J. Taskinen, Bands in the spectrum of a periodic elastic waveguide, Z. Angew. Math. Phys., 68 (2017), 27 pp. doi: 10.1007/s00033-017-0846-0
    [3] M. Sh. Birman and M. Z. Solomyak, Spectral Theory of Self-adjoint Operators in Hilbert Space, D. Reidel Publishing Co., Dordrecht, 1987.
    [4] Asymptotic expansion of the solution of an interface problem in a polygonal domain with thin layer. Asymptot. Anal. (2006) 50: 121-173.
    [5] Expansion in characteristic functions of an equation with periodic coefficients. Dokl. Akad. Nauk SSSR (1950) 73: 1117-1120.
    [6] Spectral properties of periodic media in the large coupling limit. Comm. Partial Differential Equations (2000) 25: 1445-1470.
    [7] R. Hempel and O. Post, Spectral gaps for periodic elliptic operators with high contrast: An overview, Progress in analysis, Vol. I, II (Berlin, 2001), 577–587.
    [8] Boundary-value problems for elliptic equations in domains with conical or angular points. Trudy Moskov. Mat. Obšč. (1967) 16: 209-292.
    [9] Spectral properties of the operator bundles generated by elliptic boundary-value problems in a cone. translation in Funct. Anal. Appl. (1988) 22: 114-121.
    [10] V. A. Kozlov, V. G. Maz'ya and J. Rossmann, Elliptic Boundary Value Problems in Domains With Point Singularities, Amer. Math. Soc., Providence RI, 1997.
    [11] Floquet theory for partial differential equations. Uspekhi Mat. Nauk (1982) 37: 3-52.
    [12] P. A. Kuchment, Floquet Theory for Partial Differential Equations, Birkhäuser Verlag, Basel, 1993. doi: 10.1007/978-3-0348-8573-7
    [13] Artificial boundary conditions on polyhedral truncation surfaces for three-dimensional elasticity systems. Comptes Rendus Mécanique (2004) 332: 591-596.
    [14] J.–L. Lions and E. Magenes, Non-Homogeneus Boundary Value Problems and Applications, Die Grundlehren der mathematischen Wissenschaften, Band 181. Springer-Verlag, New York-Heidelberg, 1972.
    [15] On the coefficients in the asymptotics of solutions of elliptic boundary value problems in domains with conical points. Math. Nachr. (1977) 76: 29-60.
    [16] Estimates in Lp and Hölder classes and the Miranda-Agmon maximum principle for solutions of elliptic boundary value problems in domains with singular points on the boundary. Math. Nachr. (1978) 81: 25-82.
    [17] The polynomial property of self-adjoint elliptic boundary-value problems and the algebraic description of their attributes. translation in Russian Math. Surveys (1999) 54: 947-1014.
    [18] Asymptotic behavior of the solution and the modeling of the Dirichlet problem in an angular domain with rapidly oscillating boundary. translation in St. Petersburg Math. J. (2008) 19: 297-326.
    [19] S. A. Nazarov, The Neumann problem in angular domains with periodic boundaries and parabolic perturbations of the boundaries, Tr. Mosk. Mat. Obs., 69(2008), 182–241; translation in Trans. Moscow Math. Soc., (2008), 153–208. doi: 10.1090/S0077-1554-08-00173-8
    [20] Gap in the essential spectrum of an elliptic formally self-adjoint system of differential equation. translation in Differ. Equ. (2010) 46: 730-741.
    [21] S. A. Nazarov and B. A. Plamenevsky, Elliptic Problems in Domains with Piecewise Smooth Boundaries, Walter de Gruyter & Co., Berlin, 1994. doi: 10.1515/9783110848915.525
    [22] Essential spectrum of a periodic elastic waveguide may contain arbitrarily many gaps. Appl. Anal. (2010) 89: 109-124.
    [23] Spectral gaps for periodic piezoelectric waveguides. Z. Angew. Math. Phys. (2015) 66: 3017-3047.
    [24] J. Nečas, Les Méthodes in Théorie Des Équations Elliptiques, Masson et Cie, Éditeurs, Paris; Academia, Éditeurs, Prague, 1967.
    [25] M. M. Skriganov, Geometric and arithmetic methods in the spectral theory of multidimensional periodic operators, Trudy Mat. Inst. Steklov., 171 (1985), 122 pp.
    [26] Eigenfunctions of the 2-dimensional anisotropic elasticity operator and algebraic equivalent materials. ZAMM Z. Angew. Math. Mech. (2008) 88: 100-115.
    [27] M. I. Vishik and L. A. Lyusternik, Regular degeneration and boundary layer for linear differential equations with a small parameter, Uspehi Mat. Nauk (N.S.), 12 1957, 3–122.
    [28] On gaps in the spectrum of some elliptic operators in divergent form with periodic coefficients. translation in St. Petersburg Math. J. (2005) 16: 773-790.
  • This article has been cited by:

    1. Ali Sili, On the limit spectrum of a degenerate operator in the framework of periodic homogenization or singular perturbation problems, 2022, 360, 1778-3569, 1, 10.5802/crmath.263
  • Reader Comments
  • © 2020 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(1843) PDF downloads(323) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog