This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Local gaps in three-dimensional periodic media

YuriΒ A. Godin and Boris Vainberg
The University of North Carolina at Charlotte,
Charlotte, NC 28223 USA
Email: ygodin@uncc.eduEmail: brvainbe@uncc.edu
Abstract

We consider the propagation of acoustic waves in a medium with a periodic array of small inclusions of arbitrary shape. The inclusion size aa is much smaller than the array period. We show that global gaps do not exist if aa is small enough. The notion of local gaps which depends on the choice of the wave vector π’Œ\bm{k} is introduced and studied. We determine analytically the location of local gaps for the Dirichlet and transmission problems.


Keywords: periodic media, Bloch waves, dispersion relation, asymptotic expansion, bandgaps.


1 Introduction

The discovery of photonic and phononic crystals, the creation and advancement of electromagnetic and acoustic metamaterials stimulated the development of efficient numerical methods for calculating their bandgap structure. The main tool for the evaluation of dispersion relations was the plane wave expansion method [1, 2, 3]. Later, other methods such as the finite-difference time-domain [4, 5, 6], the finite element [7, 8, 9] and boundary element methods [10, 11] employed. All these methods reduce the bandgap problem to the solution of an eigenvalue problem from which propagating frequency is determined. Analytic determination of bandgaps is a more challenging problem which becomes possible by introducing small or large parameters. This situation occurs in solid-state physics in the study of the energy bands of electrons when the periodic potential of crystals is small [12, 13]. Other examples are dealing with high contrast materials for inclusions and the matrix, such results were established in [14, 15, 16, 17].

The aim of standard bandgap analysis is the determination of absolute gaps consisting of frequencies for which no waves propagate in any direction. We will show the absence of absolute gaps in our problem and investigate local gaps defined by pairs (Ο‰,π’Œ0)(\omega,\bm{k}_{0}) for which waves with frequency Ο‰\omega do not propagate for small variations of the Bloch vector π’Œ0\bm{k}_{0}. A detailed definition of a local gap is given later.

We study the propagation of Bloch waves in a medium containing a periodic array of identical inclusions. The amplitude uu of the waves is governed by the equation

Δ​u+k2​u=0,\displaystyle\Delta u+k^{2}u=0, (1.1)

where k=Ο‰/ck=\omega/c is the wave number, Ο‰\omega is the time frequency, and cc is the wave propagation speed (different in the host medium and inclusions). Solution of (1.1) is sought in the form

u​(𝒙)=Φ​(𝒙)​eβˆ’iβ€‹π’Œβ‹…π’™,\displaystyle u({\bm{x}})=\Phi({\bm{x}})\mathrm{e}^{-\mathrm{i}\bm{k}\cdot{\bm{x}}},

where π’Œ=(k1,k2,k3)\bm{k}=(k_{1},k_{2},k_{3}) is the Bloch vector, and function Φ​(𝒙)\Phi({\bm{x}}) is periodic with the period of the lattice. The representation above is equivalent to

⟧eiβ€‹π’Œβ‹…π’™u(𝒙)⟦=0,\displaystyle\rrbracket\mathrm{e}^{\mathrm{i}\bm{k}\cdot{\bm{x}}}u({\bm{x}})\llbracket=0, (1.2)

where ⟧f⟦\rrbracket f\llbracket denotes the jump of ff and its gradient across the opposite sides of the cells of periodicity.

For simplicity, we assume that the fundamental cell of periodicity Ξ \Pi is a cube [βˆ’Ο€,Ο€]3[-\pi,\pi]^{3}. Denote by Ξ©\Omega the domain occupied by the inclusion in Ξ \Pi (see Figure 1). We select the origin for the coordinate system to be in Ξ©\Omega and suppose that Ξ©\Omega has a small size. Specifically, Ξ©=Ω​(a)\Omega=\Omega(a) is produced by compressing an aa-independent region Ξ©^\widehat{\Omega} of arbitrary shape using the factor aβˆ’1a^{-1}, 0<aβ‰ͺ10<a\ll 1. In other words, the transformation 𝒙→a​𝝃{\bm{x}}\to a{\bm{\xi}} maps Ω​(a)βŠ‚β„π’™3\Omega(a)\subset{\mathbb{R}}^{3}_{{\bm{x}}} into Ξ©^βŠ‚β„πƒ3\widehat{\Omega}\subset{\mathbb{R}}^{3}_{{\bm{\xi}}}. We assume that the boundary βˆ‚Ξ©\partial\Omega belongs to the class C1,Ξ²C^{1,\beta}, which means that the functions that define the boundary have first-order derivatives that belong to the HΓΆlder space with index Ξ²\beta.

x1x_{1}x3x_{3}x2x_{2}
Ξ©\OmegaΞ \PiBRB_{R}
Figure 1: The cell of periodicity Ξ \Pi with a small inclusion Ξ©\Omega of size aa. A ball BRβŠ‚Ξ B_{R}\subset\Pi of radius RR centered at the origin encloses the inclusion Ξ©\Omega.

We consider first the case when homogeneous Dirichlet conditions, which arise for some soft inclusions, are imposed on the boundary. The transmission problem is considered in Β§6. Then function uu in the fundamental cell of periodicity Ξ \Pi obeys the equation

Δ​u+k2​u\displaystyle\Delta u+k^{2}u =0,u∈H2​(Ξ βˆ–Ξ©)\displaystyle=0,\quad u\in H^{2}(\Pi\smallsetminus\Omega) (1.3)

and the boundary conditions

u|βˆ‚Ξ©\displaystyle\left.u\right|_{\partial\Omega} =0,⟧eiβ€‹π’Œβ‹…π’™u(𝒙)⟦=0.\displaystyle=0,\quad\rrbracket\mathrm{e}^{\mathrm{i}\bm{k}\cdot{\bm{x}}}u({\bm{x}})\llbracket=0. (1.4)

A non-trivial solution of (1.3),(1.4) with fixed Ξ©^\widehat{\Omega} exists not for all values of k=Ο‰/c,a,k=\omega/c,a, and π’Œ\bm{k}. The relation between the parameters Ο‰,a,π’Œ\omega,a,\bm{k}, for which there is a non-trivial solution, is called the dispersion relation. The problem (1.3),(1.4) is one-periodic in each component of π’Œ\bm{k}, and the dispersion relation is a periodic function defined in β„π’Œ3{\mathbb{R}}^{3}_{\bm{k}}.

The goal of this work is to find gaps i.e. the values of time frequency Ο‰=k​c\omega=kc such that (1.3),(1.4) has only trivial solutions for all Bloch vectors π’Œβˆˆβ„3\bm{k}\in{\mathbb{R}}^{3}. Thus, the waves with frequency Ο‰\omega do not propagate if Ο‰\omega belongs to a gap. These gaps are also called global or complete gaps. Geometrically, global gaps are empty intervals in the orthogonal projection of the dispersion surface onto the vertical Ο‰\omega-axis. Global gaps in Ο‰\omega correspond to the gaps in the spectrum Ξ»=k2\lambda=k^{2} of βˆ’Ξ”-\Delta in Ξ βˆ–Ξ©β€‹(a)\Pi\smallsetminus\Omega(a) with an arbitrary π’Œ\bm{k} in the boundary condition (1.4). We will also consider local gaps whose rigorous definition will be provided later in this section.

Consider the unperturbed problem. Observe that in the absence of inclusions, function eβˆ’iβ€‹π’Œβ‹…π’™\mathrm{e}^{-\mathrm{i}\bm{k}\cdot{\bm{x}}} satisfies (1.3),(1.4) with k=|π’Œ|k=|\bm{k}|, i.e. Ο‰=c​|π’Œ|\omega=c|\bm{k}|. For any vector π’Ž=(m1,m2,m3)\bm{m}=(m_{1},m_{2},m_{3}) with integer components, functions eβˆ’i​(π’Œβˆ’π’Ž)⋅𝒙\mathrm{e}^{-\mathrm{i}(\bm{k}-\bm{m})\cdot{\bm{x}}} are also satisfy the same boundary condition (1.4) and equation (1.3) with Ο‰=c​|π’Œβˆ’π’Ž|\omega=c|\bm{k}-\bm{m}|. The graph of the infinite-valued periodic in π’Œ\bm{k} dispersion function for the unperturbed problem consists of the set of all the cones Ο‰=c​|π’Œβˆ’π’Ž|\omega=c|\bm{k}-\bm{m}|, see Figure 2.

Refer to caption
Figure 2: Several cones forming the dispersion surface of the unperturbed problem in dimension two.

For some values of π’Œ\bm{k}, the unperturbed problem (1.3),(1.4) with k=|π’Œ|k=|\bm{k}| has solutions different from eβˆ’iβ€‹π’Œβ‹…π’™\mathrm{e}^{-\mathrm{i}\bm{k}\cdot{\bm{x}}}, and the solution space has dimension n>1n>1. All such solutions have the form eβˆ’i​(π’Œβˆ’π’Ž)⋅𝒙\mathrm{e}^{-\mathrm{i}(\bm{k}-\bm{m})\cdot{\bm{x}}} provided that |π’Œ|=|π’Œβˆ’π’Ž||\bm{k}|=|\bm{k}-\bm{m}| and π’Ž\bm{m} is a vector with integer components. The values of π’Œ\bm{k} for which n>1n>1 are called exceptional of order nn. Geometrically, nn refers to the number of points π’Ž\bm{m} in the integer lattice β„€3\mathbb{Z}^{3} (including the origin) belonging to the Ewald sphere [13] centered at π’Œ\bm{k} with radius |π’Œ||\bm{k}|, i.e. equation

|π’Œ|=|π’Œβˆ’π’Ž|\displaystyle|\bm{k}|=|\bm{k}-\bm{m}| (1.5)

has n>1n>1 solutions π’Žβ‰ πŸŽ\bm{m}\neq{\bm{0}} for exceptional π’Œ\bm{k}, and only one solution π’Ž=𝟎\bm{m}={\bm{0}} for non-exceptional π’Œ\bm{k}. Equation (1.5) can be written as

2β€‹π’Œβ‹…π’Ž=|π’Ž|2,π’Žβˆˆβ„€3βˆ–πŸŽ,\displaystyle 2\bm{k}\cdot\bm{m}=|\bm{m}|^{2},\quad\bm{m}\in{\mathbb{Z}}^{3}\smallsetminus{\bm{0}}, (1.6)

i.e. exceptional points belong to planes in β„π’Œ3{\mathbb{R}}^{3}_{\bm{k}} whose distance from the origin goes to infinity as |π’Ž|β†’βˆž|\bm{m}|\to\infty. An exceptional point π’Œ\bm{k} has order n>1n>1 if it belongs to nβˆ’1n-1 planes (1.6). One can also describe exceptional wave vectors as values of π’Œ\bm{k} for which n>1n>1 cones in Figure 2 interest each other. In particular, in the two-dimensional analog of our problem in Figure 2, the line A​BAB consists of exceptional points of order at least two.

The set of solutions of the problem (1.3),(1.4) remains the same if the wave vector is changed by a period of the reciprocal lattice. Thus it is enough to study local gaps related to one cone centered at the origin: C0:={Ο‰=c​|π’Œ|}C_{0}:=\{\omega=c|\bm{k}|\}. This will allow us to define local gaps related to other cones in Figure 2 using the periodicity of solutions in π’Œ\bm{k}.

Let us fix a wave vector π’Œ0\bm{k}_{0}. The local gap g​(π’Œ0)g(\bm{k}_{0}) corresponding to π’Œ0\bm{k}_{0} and associated with C0C_{0} consists of frequencies Ο‰=Ο‰a\omega=\omega_{a} for which Bloch waves do not propagate when wave vector π’Œ\bm{k} has the same direction as π’Œ0\bm{k}_{0} and |π’Œβˆ’π’Œ0||\bm{k}-\bm{k}_{0}| is sufficiently small. More precisely, Bloch waves with the frequency Ο‰a\omega_{a} do not exist when π’Œ=(1+Ξ΄)β€‹π’Œ0\bm{k}=(1+\delta)\bm{k}_{0} for sufficiently small aa-independent |Ξ΄||\delta|.

It is not expedient to consider frequencies Ο‰\omega if the point (Ο‰,π’Œ0)(\omega,\bm{k}_{0}) is located at a positive, aa-independent distance dd from the set of cones shown in Fig. 2 (dispersion surfaces of the unperturbed problem). Bloch waves with parameters (Ο‰,(1+Ξ΄)β€‹π’Œ0)(\omega,(1+\delta)\bm{k}_{0}) and Ξ΄<d/2\delta<d/2 do not propagate in the unperturbed problem and the problem with inclusions if aa is small enough. Thus, we introduce one more requirement in the definition of local gap g​(π’Œ0)g(\bm{k}_{0}). We assume that Ο‰=Ο‰aβ†’Ο‰0\omega=\omega_{a}\to\omega_{0} as aβ†’0a\to 0 where (Ο‰0,π’Œ0)∈C0(\omega_{0},\bm{k}_{0})\in C_{0}, i.e. Ο‰0=c​|π’Œ0|\omega_{0}=c|\bm{k}_{0}|.

Let Ο‰a\omega_{a} tends to Ο‰0=c​|π’Œ0|\omega_{0}=c|\bm{k}_{0}|. We will show that

  1. (a)

    If (Ο‰0,π’Œ0)(\omega_{0},\bm{k}_{0}) belongs to a cone C0C_{0} but not to the intersection of the cones, then g​(π’Œ0)=βˆ…g(\bm{k}_{0})=\varnothing, i. e. there are no local gap for wave vector π’Œ0\bm{k}_{0} and frequencies Ο‰\omega near Ο‰0\omega_{0}.

  2. (b)

    If (Ο‰0,π’Œ0)(\omega_{0},\bm{k}_{0}) belongs to an intersection of C0C_{0} with some other cone shown in Figure 2 then local gap g​(π’Œ0)g(\bm{k}_{0}) exists in a neighborhood of Ο‰0\omega_{0} for small aa if and only if |π’Œ0||π’Ž0|<22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}<\frac{\sqrt{2}}{2}, where π’Œ0,π’Ž0\bm{k}_{0},\bm{m}_{0} satisfy (1.5). The location of the gap will be specified. We do not consider the points (Ο‰0,π’Œ0)(\omega_{0},\bm{k}_{0}) that belong to the intersection of three or more cones and those for which |π’Œ0||π’Ž0|=22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}=\frac{\sqrt{2}}{2}.

  3. (c)

    We will show that global gaps do not exist in any fixed interval Ο΅<Ο‰<Ο΅βˆ’1\epsilon<\omega<\epsilon^{-1} of the time frequency Ο‰\omega if the size aa of the inclusion is small enough.

Note that the set of values of π’Œ\bm{k} omitted from the consideration in item (b) forms a one-dimensional manifold in ℝ3{\mathbb{R}}^{3}. Indeed, the set of exceptional vectors of multiplicity nβ©Ύ2n\geqslant 2 is given by the planes (1.6). Omitted from the consideration set of exceptional vectors of order higher than two are lines of intersections of those planes. Also, condition |π’Œ0||π’Ž0|=22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}=\frac{\sqrt{2}}{2} defines curves on planes (1.6), see Figure 3, where these curves are circles that are boundaries of shaded disks.

2 Outline of the approach

In [18, 19] we devised a new approach to determining the dispersion relation which allowed us to find the asymptotic expansion of the dispersion relation as aβ†’0a\to 0. The expansions have a form of power series in aa which are different for non-exceptional and exceptional wave vectors π’Œ\bm{k}. The expansion depends analytically on π’Œ\bm{k} when π’Œ\bm{k} is not exceptional, see a discussion below. To study gaps, we need to develop further these results in the present paper to obtain asymptotics in aa which is uniform in π’Œ\bm{k} in a neighborhood of exceptional points.

As before, we enclose the inclusion Ξ©\Omega in a ball BRβŠ‚Ξ B_{R}\subset\Pi of radius R>aR>a centered at the origin and split Ξ \Pi into the ball BRB_{R} and its complement Ξ βˆ–BR\Pi\smallsetminus B_{R} (see Figure 1). Next, we consider the following two problems in these two regions:

(Ξ”+k2)​u​(𝒙)\displaystyle\left(\Delta+k^{2}\right)u({\bm{x}}) =0,π’™βˆˆ(BRβˆ–Ξ©),u|βˆ‚Ξ©=0,u|r=R=ψ,\displaystyle=0,\quad{\bm{x}}\in(B_{R}\smallsetminus\Omega),\quad\left.u\right|_{\partial\Omega}=0,\quad\left.u\right|_{r=R}=\psi, (2.1)
(Ξ”+k2)​v​(𝒙)\displaystyle\left(\Delta+k^{2}\right)v({\bm{x}}) =0,π’™βˆˆΞ βˆ–BR,⟧eiβ€‹π’Œβ‹…π’™v(𝒙)⟦=0,v|r=R=ψ,\displaystyle=0,\quad{\bm{x}}\in\Pi\smallsetminus B_{R},\quad\left\rrbracket\mathrm{e}^{\mathrm{i}\bm{k}\cdot{\bm{x}}}v({\bm{x}})\right\llbracket=0,\quad\left.v\right|_{r=R}=\psi, (2.2)

where

k\displaystyle k =Ο‰/c=(1+Ξ΅)​|π’Œ|,\displaystyle=\omega/c=(1+\varepsilon)|\bm{k}|, (2.3)

Ξ΅\varepsilon will be determined below, u∈H2​(BRβˆ–Ξ©),v∈H2​(Ξ βˆ–BR)u\in H^{2}(B_{R}\smallsetminus\Omega),~v\in H^{2}(\Pi\smallsetminus B_{R}). The unperturbed problems with a=Ξ΅=0a=\varepsilon=0 are uniquely solvable for smooth enough ψ\psi for all RR except a countable set {Ri}\{R_{i}\}, and we choose Rβˆ‰{RiR\notin\{R_{i}}. Then solutions of (2.1),(2.2) defines two Dirichlet-to-Neumann operators \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon}, \mathsfbfit​Nπ’Œ,Ξ΅+{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}: H3/2​(βˆ‚BR)β†’H1/2​(βˆ‚BR)H^{3/2}(\partial B_{R})\to H^{1/2}(\partial B_{R}), i.e.

\mathsfbfit​Na,Ξ΅βˆ’β€‹Οˆ=βˆ‚uβˆ‚r|r=R,\mathsfbfit​Nπ’Œ,Ξ΅+β€‹Οˆ=βˆ‚vβˆ‚r|r=R.{\mathsfbfit N}^{-}_{a,\varepsilon}\psi=\left.\frac{\partial u}{\partial r}\right|_{r=R},\quad\quad{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}\psi=\left.\frac{\partial v}{\partial r}\right|_{r=R}. (2.4)

The existence of a non-trivial solution to the problem (1.3),(1.4) is equivalent to the existence of a zero eigenvalue (EZE) for the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}. This abbreviation pertains specifically to the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}. Hence, the dispersion relation is given by (2.3) with Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) determined from the EZE condition. In [18, 19], the reduction of the dispersion relation to the EZE problem enabled us to use regular perturbation techniques to study dispersion relations for each π’Œ\bm{k}.

Observe that the choice of RR and therefore the definition of operators \mathsfbfit​NΒ±\mathsfbfit N^{\pm} depend on π’Œ\bm{k}. Since k2k^{2} is an analytic function of π’Œ\bm{k} and since RR is chosen in such a way that the problem (2.2) is uniquely solvable when π’Œ=π’Œβ€²\bm{k}=\bm{k}^{\prime} is fixed and Ξ΅=0\varepsilon=0, the operator \mathsfbfit​Nπ’Œ,Ξ΅+{\mathsfbfit N}^{+}_{\bm{k},\varepsilon} is analytic in π’Œ\bm{k} and Ξ΅\varepsilon when |π’Œβˆ’π’Œβ€²|+|Ξ΅|β‰ͺ1|\bm{k}-\bm{k}^{\prime}|+|\varepsilon|\ll 1. We will call this property the local analyticity in π’Œβˆˆβ„3\bm{k}\in{\mathbb{R}}^{3} and small Ξ΅\varepsilon. Similarly, the operator \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon} is locally analytic in π’Œ\bm{k} and small Ξ΅\varepsilon. The smoothness of the operator \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon} in aa is not obvious, because problem (2.1) is a singular perturbation of the corresponding problem without the inclusion (Ξ©=βˆ…\Omega=\varnothing), and solutions of (2.1) do not have uniform limits as aβ†’0a\to 0. However, the solution of (2.1) is infinitely smooth in aa when aa is small and 𝒙{\bm{x}} is outside of a fixed neighborhood of the origin, and it was shown in [18, 19] that the interior DtN operator \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon} is infinitely differentiable in a,aβ‰ͺ1a,~a\ll 1. We summarize these properties in the Lemma.

Lemma 2.1.
  1. 1.

    The operator \mathsfbfit​Nπ’Œ,Ξ΅+{\mathsfbfit N}^{+}_{\bm{k},\varepsilon} is locally analytic in π’Œβˆˆβ„3\bm{k}\in{\mathbb{R}}^{3} and Ξ΅\varepsilon. The operator \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon} is locally analytic in π’Œ\bm{k} and Ξ΅\varepsilon and infinitely smooth in a,aβ‰ͺ1a,~a\ll 1.

  2. 2.

    The dispersion relation (2.3) for which (1.3),(1.4) have a nontrivial solution is given by the same function Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) for which the EZE condition holds.

The Lemma reduces the study of the dispersion relation to the EZE problem for a smooth operator function. Special matrix representations of operators \mathsfbfit​Nπ’Œ,Ξ΅+,\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon},~{\mathsfbfit N}^{-}_{a,\varepsilon} were constructed in [18, 19] that allowed us to solve the EZE problem and find the asymptotics of the dispersion relation Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) as aβ†’0a\to 0 which is analytic in π’Œ\bm{k} for non-exceptional π’Œ\bm{k} and is valid for fixed exceptional π’Œ\bm{k}. The gaps are related to exceptional points, and the study of gaps requires a uniform in π’Œ\bm{k} asymptotics of Ρ​(a,π’Œ)\varepsilon(a,\bm{k}). The main goal of the present paper is to construct a matrix representation of the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}, allowing to obtain uniform in π’Œ\bm{k} asymptotics of the dispersion relations near the exceptional points and use them to study the gaps.

The study of the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} is simplified if we subtract \mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{-}_{0,\varepsilon} from both terms of the difference above.

3 Necessary prerequisites

This section provides the results from [18, 19] needed to understand the goal and approach of the present paper.

The dispersion surface for the inclusionless (Ξ©=βˆ…,k=|π’Œ|\Omega=\varnothing,k=|\bm{k}|) problem (1.3), (1.4) consists of the cone C0:Ο‰=c​|π’Œ|C_{0}:~\omega=c|\bm{k}| and its shifts shown in Fig. 1,2. Since the solutions of (1.3), (1.4) and of the corresponding transmission problem are periodic in π’Œ\bm{k} and their dispersion surfaces are close to the surface for the inclusionless problem, it is enough to study the dispersion surface of the problems with inclusions only in a small neighborhood of C0C_{0}. The information about other parts of the surface can be obtained simply by the shift in π’Œ\bm{k}. Thus, similar to the approach used in [18, 19], we will assume first that (Ο‰,π’Œ)(\omega,\bm{k}) is located in a small neighborhood of C0C_{0}.
Non-exceptional wave vectors. Let π’Œ\bm{k} be a non-exceptional wave vector. Then solution space of the unperturbed problem (1.3),(1.4) (Ξ©=βˆ…,k=|π’Œ|\Omega=\varnothing,k=|\bm{k}|) is one-dimensional and consists of functions proportional to eβˆ’iβ€‹π’Œβ‹…π’™,π’™βˆˆΞ \mathrm{e}^{-\mathrm{i}\bm{k}\cdot{\bm{x}}},~{\bm{x}}\in\Pi. Respectively, the kernel of the operator \mathsfbfit​Nπ’Œ,0+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},0}-{\mathsfbfit N}^{-}_{0,\varepsilon} consists of functions proportional to Οˆβ€²=eβˆ’iβ€‹π’Œβ‹…π’™,π’™βˆˆβˆ‚BR{\psi^{\prime}}=\mathrm{e}^{-\mathrm{i}\bm{k}\cdot{\bm{x}}},~{\bm{x}}\in\partial B_{R} which is the restriction of the exponent on βˆ‚BR\partial B_{R}. We present the domain H32​(βˆ‚BR)H^{\frac{3}{2}}(\partial B_{R}) and the range H12​(βˆ‚BR)H^{\frac{1}{2}}(\partial B_{R}) of the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} as a direct sum of the one-dimensional space β„°\mathscr{E} of functions proportional to Οˆβ€²{\psi^{\prime}} and spaces β„°βŠ₯,d\mathscr{E}_{\bot,d},β„°βŠ₯,r\mathscr{E}_{\bot,r}, respectively, of functions orthogonal in L2​(βˆ‚Ξ©)L^{2}(\partial\Omega) to Οˆβ€²{\psi^{\prime}}. Here subindexes β€œdd” and β€œrr” stay for the domain and the range.

We represent each element ψ\psi in the domain and the range of operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} in the vector form ψ=(Οˆβ„°,ψβŠ₯),\psi=(\psi_{\mathscr{E}},\psi_{\bot}), where Οˆβ„°\psi_{\mathscr{E}} is the projection in L2​(βˆ‚BR)L^{2}(\partial B_{R}) of the function ψ\psi on Οˆβ€²{\psi^{\prime}}, and ψβŠ₯\psi_{\bot} is orthogonal to Οˆβ€²{\psi^{\prime}} in L2​(βˆ‚BR)L^{2}(\partial B_{R}), and use matrix representation of all the operators in the basis (Οˆβ„°,ψβŠ₯).(\psi_{\mathscr{E}},\psi_{\bot}). Due to Green’s formula, the operators \mathsfbfit​Nπ’Œ,Ξ΅+,{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}, \mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{-}_{a,\varepsilon} are symmetric and therefore

\mathsfbfit​Nπ’Œ,0+βˆ’\mathsfbfit​N0,Ξ΅βˆ’=(000𝑨),{\mathsfbfit N}^{+}_{\bm{k},0}-{\mathsfbfit N}^{-}_{0,\varepsilon}=\left(\begin{array}[]{cc}0&0\\[5.69054pt] 0&\bm{A}\end{array}\right), (3.1)

where 𝑨:β„°βŠ₯,dβ†’β„°βŠ₯,r\bm{A}:\mathscr{E}_{\bot,d}\to\mathscr{E}_{\bot,r} is an isomorphism.

Since operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon} depends smoothly on Ξ΅\varepsilon, we have

\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’=(C​Ρ+π’ͺ​(Ξ΅2)π’ͺ​(Ξ΅)π’ͺ​(Ξ΅)𝑨+π’ͺ​(Ξ΅)),{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon}=\left(\begin{array}[]{cc}C\varepsilon+{\cal O}(\varepsilon^{2})&{\cal O}(\varepsilon)\\[5.69054pt] {\cal O}(\varepsilon)&\bm{A}+{\cal O}(\varepsilon)\end{array}\right), (3.2)

where CC is a constant, and Ξ΅\varepsilon is defined in (2.3). In [18] it was shown that C=2​|π’Œ|2​|Ξ |C=2|\bm{k}|^{2}|\Pi|.

Matrix representation for the interior operator has the form

\mathsfbfit​Na,Ξ΅βˆ’βˆ’\mathsfbfit​N0,Ξ΅βˆ’=(4​π​q​a+π’ͺ​(a2+a​|Ξ΅|)π’ͺ​(a+|Ξ΅|)π’ͺ​(a+|Ξ΅|)O​(a+|Ξ΅|)),{\mathsfbfit N}^{-}_{a,\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon}=\left(\begin{array}[]{cc}4\pi qa+{\cal O}(a^{2}+a|\varepsilon|)&{\cal O}(a+|\varepsilon|)\\[5.69054pt] {\cal O}(a+|\varepsilon|)&O(a+|\varepsilon|)\end{array}\right), (3.3)

where q​aqa depends on the shape of βˆ‚Ξ©\partial\Omega and is equal to its electrical capacitance, i.e. total charge of βˆ‚Ξ©\partial\Omega when its potential is unity. In particular, q=1q=1 for a sphere. Thus

\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’=(Cβ€‹Ξ΅βˆ’4​π​q​a+π’ͺ​(a2+Ξ΅2)π’ͺ​(a+|Ξ΅|)π’ͺ​(a+|Ξ΅|)𝑨+π’ͺ​(a+|Ξ΅|)),C=2​|π’Œ|2​|Ξ |.{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}=\left(\begin{array}[]{cc}C\varepsilon-4\pi qa+{\cal O}(a^{2}+\varepsilon^{2})&{\cal O}(a+|\varepsilon|)\\[5.69054pt] {\cal O}(a+|\varepsilon|)&\bm{A}+{\cal O}(a+|\varepsilon|)\end{array}\right),\quad C=2|\bm{k}|^{2}|\Pi|. (3.4)

This matrix representation allows one to find Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) for which the kernel of the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} is not empty. Using (3.4), we write the equation (\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’)β€‹Οˆ=0({\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon})\psi=0 in the form

(2​Ρ​|π’Œ|2​|Ξ |βˆ’4​π​q​a+π’ͺ​(a2+Ξ΅2)π’ͺ​(a+|Ξ΅|)π’ͺ​(a+|Ξ΅|)𝑨+π’ͺ​(a+|Ξ΅|))​(Οˆβ„°ΟˆβŠ₯)=0.\left(\begin{array}[]{cc}2\varepsilon|\bm{k}|^{2}|\Pi|-4\pi qa+{\cal O}(a^{2}+\varepsilon^{2})&{\cal O}(a+|\varepsilon|)\\[5.69054pt] {\cal O}(a+|\varepsilon|)&\bm{A}+{\cal O}(a+|\varepsilon|)\end{array}\right)\left(\begin{array}[]{cc}\psi_{\mathscr{E}}\\[5.69054pt] \psi_{\bot}\end{array}\right)=0. (3.5)

Due to the invertibility of 𝑨\bm{A}, we solve equation above for ΟˆβŸ‚:ψβŠ₯=π’ͺ​(a+|Ξ΅|)β€‹Οˆβ„°\psi_{\perp}:~\psi_{\bot}={\cal O}(a+|\varepsilon|)\psi_{\mathscr{E}}, and reduce (3.5) to a simple equation for Οˆβ„°\psi_{\mathscr{E}}. This implies that function Ρ​(a,π’Œ)\varepsilon(a,\bm{k}) can be found by equating the top left element in (3.5) with a different remainder term to zero. Thus, the implicit function theorem implies the following statement:

Lemma 3.1.

Let π€β€²β‰ πŸŽ\bm{k}^{\prime}\neq{\bm{0}} be an arbitrary non-exceptional wave vector and kβ€²=|𝐀′|k^{\prime}=|\bm{k}^{\prime}|. Then

  1. 1.

    The dispersion relation Ο‰=c​(1+Ξ΅)​|π’Œ|\omega=c(1+\varepsilon)|\bm{k}| is uniquely defined in a neighborhood of the point (kβ€²,π’Œβ€²)βˆˆβ„4(k^{\prime},\bm{k}^{\prime})\in{\mathbb{R}}^{4} if aa is sufficiently small. Function Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) is infinitely smooth in a,π’Œa,\bm{k} and analytic in π’Œ\bm{k} when π’Œβ‰ πŸŽ\bm{k}\neq{\bm{0}}.

  2. 2.

    The following asymptotics holds

    Ξ΅=2​π​q​a|π’Œ|2​|Ξ |+O​(a2),aβ†’0,π’Œβ‰ πŸŽ.\varepsilon=\frac{2\pi qa}{|\bm{k}|^{2}|\Pi|}+O(a^{2}),\quad a\to 0,\quad\bm{k}\neq{\bm{0}}. (3.6)

Exceptional wave vectors of order two. Let the Bloch vector π’Œ=π’Œ0\bm{k}=\bm{k}_{0} in (1.4) be exceptional of multiplicity two, and therefore there exists a non-zero vector π’Ž\bm{m} with integer components such that |π’Œ0|=|π’Œ1||\bm{k}_{0}|=|\bm{k}_{1}| for π’Œ1=π’Œ0βˆ’π’Ž\bm{k}_{1}=\bm{k}_{0}-\bm{m}. Then solution space of the unperturbed problem (1.3),(1.4) is spanned by functions eβˆ’iβ€‹π’Œ0⋅𝒙,eβˆ’iβ€‹π’Œ1⋅𝒙,π’™βˆˆΞ \mathrm{e}^{-\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}},~~\mathrm{e}^{-\mathrm{i}\bm{k}_{1}\cdot{\bm{x}}},~{\bm{x}}\in\Pi. Respectively, the kernel β„°\mathscr{E} of the operator \mathsfbfit​Nπ’Œ,0+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},0}-{\mathsfbfit N}^{-}_{0,\varepsilon} is two-dimensional and consists of linear combinations of functions ψiβ€²=eβˆ’iβ€‹π’Œi⋅𝒙,π’™βˆˆβˆ‚BR,i=0,1\psi_{i}^{\prime}=\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}},~{\bm{x}}\in\partial B_{R},~i=0,1. Once again we present the domain and the range of the operator \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} as a direct sum of space β„°\mathscr{E} and its orthogonal in L2​(βˆ‚Ξ©)L^{2}(\partial\Omega) complements, and write all the operators in corresponding matrix form.

The formula (3.1) remains valid when π’Œ=π’Œ0\bm{k}=\bm{k}_{0} is an exceptional wave vector of order two if zeroes of the matrix on the right-hand side of (3.1) are understood as the zero elements of the first two rows and first two columns of the matrix. In addition, formulas (3.2) and (3.3) remain valid if constants CC and qq are replaced by C​𝑰C\bm{I} and q​𝑱q\bm{J}, respectively, where 𝑰\bm{I} and 𝑱\bm{J} are 2Γ—22\times 2 identity and all-ones matrices, respectively. The analogue of (3.4) has the following 2Γ—22\times 2 block 𝑺\bm{S} in the top left corner of the matrix:

𝑺=2​Ρ​|π’Œ0|2​|Ξ |β€‹π‘°βˆ’4​π​q​a​𝑱+π’ͺ​(a2+Ξ΅2),\displaystyle\bm{S}=2\varepsilon|\bm{k}_{0}|^{2}|\Pi|\bm{I}-4\pi qa\bm{J}+{\cal O}(a^{2}+\varepsilon^{2}), (3.7)

and therefore, (3.5) now takes the form

(𝑺+π’ͺ​(a2+Ξ΅2)π’ͺ​(a+|Ξ΅|)π’ͺ​(a+|Ξ΅|)𝑨+π’ͺ​(a+|Ξ΅|))​(Οˆβ„°ΟˆβŠ₯)=0.\left(\begin{array}[]{cc}\bm{S}+{\cal O}(a^{2}+\varepsilon^{2})&{\cal O}(a+|\varepsilon|)\\[5.69054pt] {\cal O}(a+|\varepsilon|)&\bm{A}+{\cal O}(a+|\varepsilon|)\end{array}\right)\left(\begin{array}[]{cc}\psi_{\mathscr{E}}\\[5.69054pt] \psi_{\bot}\end{array}\right)=0. (3.8)

The arguments after (3.5) imply that the dispersion relation at π’Œ=π’Œ0\bm{k}=\bm{k}_{0} is defined by the equation det𝑺~=0\det{\widetilde{\bm{S}}}=0 where 𝑺~{\widetilde{\bm{S}}} and 𝑺\bm{S} have the same asymptotics with different specific values of the remainders. Since the eigenvalues of 𝑱\bm{J} are distinct, the implicit function theorem can be applied to the equation det𝑺~=0\det{\widetilde{\bm{S}}}=0 for the variable Ξ΅/a\varepsilon/a. This leads to the existence of the following two roots:

Ξ΅=Ξ΅1=2​π​q​a|π’Œ0|2​|Ξ |+O​(a2),Ξ΅=Ξ΅2=O​(a2),aβ†’0.\varepsilon=\varepsilon_{1}=\frac{2\pi qa}{|\bm{k}_{0}|^{2}|\Pi|}+O(a^{2}),\quad\varepsilon=\varepsilon_{2}=O(a^{2}),~~a\to 0. (3.9)

In addition, the following important observation was made in [18]: there are no solutions to the problem (1.3),(1.4) with π’Œ=π’Œ0\bm{k}=\bm{k}_{0} propagating in one direction of either vector π’Œ0\bm{k}_{0} or π’Œ1\bm{k}_{1}. Each of the Bloch solutions with Ξ΅=Ξ΅i,i=0,1,\varepsilon=\varepsilon_{i},~i=0,1, is a cluster of waves propagating in both directions.

4 Asymptotics of the DtN operators in the neighborhoods of exceptional points

The results concerning non-exceptional points described above are valid for all non-exceptional points. In particular, the function (3.6) is analytic in π’Œ\bm{k}. On the contrary, formulas (3.9) were obtained for fixed exceptional points whose neighborhoods contained mostly non-exceptional points. We need to extend (3.1)-(3.4) in such a way that the corresponding asymptotics would be valid in the neighborhood of fixed wave vector π’Œ=π’Œ0\bm{k}=\bm{k}_{0} of multiplicity two.

We do not consider the whole neighborhood of π’Œ0\bm{k}_{0}. It will be sufficient to focus only on the values of π’Œ\bm{k} that belong to the ray through π’Œ0\bm{k}_{0}. Thus π’Œ=(1+Ξ΄)β€‹π’Œ0\bm{k}=(1+\delta)\bm{k}_{0}, Β |Ξ΄|β‰ͺ1|\delta|\ll 1. We will denote this segment by II. It contains one exceptional point π’Œ=π’Œ0\bm{k}=\bm{k}_{0}, and other points in Iβˆ–π’Œ0I\smallsetminus\bm{k}_{0} are non-exceptional if Ξ΄\delta is small enough. Matrix representation (3.1)-(3.4) in the previous section was based on one-dimensional space β„°\mathscr{E} when π’ŒβˆˆIβˆ–π’Œ0\bm{k}\in I\smallsetminus\bm{k}_{0} and two-dimensional space β„°\mathscr{E} when π’Œ=π’Œ0\bm{k}=\bm{k}_{0}. Our next goal is to obtain the dispersion relation for the entire interval II using the two-dimension space β„°\mathscr{E} spanned by functions ψ0\psi_{0} and ψ1\psi_{1}. This will allow us to obtain the asymptotics of the dispersion relation as aβ†’0a\to 0, π’ŒβˆˆI\bm{k}\in I, which is uniform in π’Œ\bm{k}.

We start with a matrix representation of \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon} for π’ŒβˆˆI\bm{k}\in I similar to (3.1)-(3.2). As mentioned at the end of the previous section, if Ξ΅=Ξ΄=0\varepsilon=\delta=0 then formula (3.1) remains valid for the two-dimensional space β„°\mathscr{E}. Since all the operators involved depend infinitely smoothly on small Ξ΅,Ξ΄\varepsilon,\delta and the space β„°\mathscr{E} is independent on Ξ΅,Ξ΄,\varepsilon,\delta, the matrix representation of \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon} for sufficiently small Ξ΅\varepsilon and Ξ΄\delta has the form

\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’=(𝑳​(Ξ΅,Ξ΄)+π’ͺ​(Ξ΅2+Ξ΄2)π’ͺ​(|Ξ΅|+|Ξ΄|)π’ͺ​(|Ξ΅|+|Ξ΄|)𝑨+π’ͺ​(|Ξ΅|+|Ξ΄|)),π’ŒβˆˆI,\displaystyle{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon}=\left(\begin{array}[]{cc}\bm{L}(\varepsilon,\delta)+{\cal O}(\varepsilon^{2}+\delta^{2})&{\cal O}(|\varepsilon|+|\delta|)\\[5.69054pt] {\cal O}(|\varepsilon|+|\delta|)&\bm{A}+{\cal O}(|\varepsilon|+|\delta|)\end{array}\right),\quad\bm{k}\in I, (4.3)

where 𝑳\bm{L} is a 2Γ—22\!\times\!2 matrix which is linear in Ξ΅\varepsilon and Ξ΄\delta, and the entries of the matrix on the right-hand side are infinitely smooth in Ξ΅\varepsilon and Ξ΄\delta. The next Lemma reveals the dependence of 𝑳\bm{L} on Ξ΅\varepsilon and Ξ΄\delta.

Lemma 4.1.

Matrix 𝐋\bm{L} has the form

𝑳=2​Ρ​|π’Œ0|2​|Ξ |​𝑰+δ​|Ξ |​[000|π’Ž0|2],\displaystyle\bm{L}=2\varepsilon|\bm{k}_{0}|^{2}|\Pi|\bm{I}+\delta|\Pi|\left[\begin{array}[]{cc}0&0\\[5.69054pt] 0&|\bm{m}_{0}|^{2}\end{array}\right], (4.6)

where the couple (𝐀0,𝐦0)(\bm{k}_{0},\bm{m}_{0}) satisfies (1.5), (1.6).

We will prove the lemma after we prove the following theorem.

Theorem 4.1.

The matrix representation of the operator \mathsfbfit​N𝐀,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon} has the form

\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’=(𝑺π’ͺ​(a+|Ξ΅|+|Ξ΄|)π’ͺ​(a+|Ξ΅|+|Ξ΄|)𝑨+π’ͺ​(a+|Ξ΅|+|Ξ΄|)),π’ŒβˆˆI,{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}=\left(\begin{array}[]{cc}\bm{S}&{\cal O}(a+|\varepsilon|+|\delta|)\\[5.69054pt] {\cal O}(a+|\varepsilon|+|\delta|)&~~\bm{A}+{\cal O}(a+|\varepsilon|+|\delta|)\end{array}\right),\quad\bm{k}\in I, (4.7)

where the 2Γ—22\times 2 upper left block is given by

𝑺=2​Ρ​|π’Œ0|2​|Ξ |β€‹π‘°βˆ’4​π​a​q​𝑱+δ​|Ξ |​[000|π’Ž0|2]+π’ͺ​(a2+Ξ΄2+Ξ΅2),a+|Ξ΅|+|Ξ΄|β‰ͺ1,\displaystyle\bm{S}=2\varepsilon|\bm{k}_{0}|^{2}|\Pi|\bm{I}-4\pi aq\bm{J}+\delta|\Pi|\left[\begin{array}[]{cc}0&0\\[5.69054pt] 0&|\bm{m}_{0}|^{2}\end{array}\right]+{\cal O}(a^{2}+\delta^{2}+\varepsilon^{2}),\quad a+|\varepsilon|+|\delta|\ll 1, (4.10)

where the couple (𝐀0,𝐦0)(\bm{k}_{0},\bm{m}_{0}) satisfies (1.5), (1.6) and qq is defined in (3.3).

Proof.

To obtain (4.7) we need to subtract the matrix representation of the operator \mathsfbfit​Nβˆ’:=\mathsfbfit​Na,Ξ΅βˆ’βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{-}:={\mathsfbfit N}^{-}_{a,\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon}, π’ŒβˆˆI\bm{k}\in I, from (4.3). \mathsfbfit​Nβˆ’{\mathsfbfit N}^{-} depends infinitely smoothly on a,Ξ΅,a,\varepsilon, and π’Œ\bm{k} and can be considered as an infinitely smooth operator-function of a,Ξ΅,a,\varepsilon, and Ξ΄\delta. Obviously, the operator is zero when a=0a=0, and therefore can be written as \mathsfbfit​Nβˆ’=a​\mathsfbfit​Nβ€²+O​(a​(a+|Ξ΅|+|Ξ΄|)){\mathsfbfit N}^{-}=a\mathsfbfit N^{\prime}+O(a(a+|\varepsilon|+|\delta|)), where operator \mathsfbfit​Nβ€²\mathsfbfit N^{\prime} does not depend on a,Ξ΅,Ξ΄a,\varepsilon,\delta. This formula, (4.3) and Lemma 4.6 will justify (4.7) if we show that the upper left 2Γ—22\times 2 block of the matrix representation for a​\mathsfbfit​Nβ€²a\mathsfbfit N^{\prime} is equal to 4​π​a​q​𝑱4\pi aq\bm{J}. This block of a​\mathsfbfit​Nβ€²a\mathsfbfit N^{\prime} coincides with the linear in aa part of a similar block of the matrix representation for \mathsfbfit​Nβˆ’{\mathsfbfit N}^{-} with Ξ΅=Ξ΄=0\varepsilon=\delta=0. The latter one was obtained in [19] when the matrix representation was studied for a fixed exceptional wave vector π’Œ=π’Œ0\bm{k}=\bm{k}_{0} of multiplicity two, see Β§2, and it equals 4​π​a​q​𝑱4\pi aq\bm{J}. ∎

Proof of Lemma 4.3.

Since 𝑳=Ρ​𝑳1+δ​𝑳2\bm{L}=\varepsilon\bm{L}_{1}+\delta\bm{L}_{2} and formula (4.3) when Ξ΄=0\delta=0 coincides with the matrix representation constructed in the previous section when π’Œ=π’Œ0\bm{k}=\bm{k}_{0}, we have 𝑳1=2​|π’Œ0|2​|Ξ |​𝑰\bm{L}_{1}=2|\bm{k}_{0}|^{2}|\Pi|\bm{I}. Thus we need to justify only the form of 𝑳2\bm{L}_{2} in (4.6) and therefore we set Ξ΅=0\varepsilon=0 in the calculations below.

The entries of 𝑳2={Li,j}\bm{L}_{2}=\{L_{i,j}\} are given by

Li,j=βˆ«βˆ‚BR(\mathsfbfit​Nπ’Œ,0+βˆ’\mathsfbfit​N0,Ξ΅βˆ’)​eβˆ’iβ€‹π’Œi⋅𝒙⋅eiβ€‹π’Œj⋅𝒙​dS+π’ͺ​(Ξ΄2),0β©½i,jβ©½1,\displaystyle L_{i,j}=\int_{\partial B_{R}}({\mathsfbfit N}^{+}_{\bm{k},0}-{\mathsfbfit N}^{-}_{0,\varepsilon})\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}\cdot\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S+{\cal O}(\delta^{2}),\quad 0\leqslant i,j\leqslant 1, (4.11)

where π’Œ0,π’Œ1\bm{k}_{0},\bm{k}_{1} are two related exceptional wave vectors of multiplicity two, i.e., π’Œ1=π’Œ0βˆ’π’Ž0\bm{k}_{1}=\bm{k}_{0}-\bm{m}_{0} and |π’Œ0|=|π’Œ1||\bm{k}_{0}|=|\bm{k}_{1}|.

Consider the solution u=uiu=u_{i} of the problem in Ξ βˆ–βˆ‚BR\Pi\smallsetminus\partial B_{R} which is a union of the exterior problem in Ξ βˆ–BR\Pi\smallsetminus B_{R} and the problem in BRB_{R} without the inclusion:

(Ξ”+|π’Œ|2)​ui​(𝒙)\displaystyle\left(\Delta+|\bm{k}|^{2}\right)u_{i}({\bm{x}}) =0,π’™βˆˆΞ ,π’™βˆ‰βˆ‚BR,ui|r=R=ψi=eβˆ’iβ€‹π’Œi⋅𝒙,i=0,1,\displaystyle=0,\quad{\bm{x}}\in\Pi,\quad{\bm{x}}\notin\partial B_{R},~~\left.u_{i}\right|_{r=R}=\psi_{i}=\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}},\quad i=0,1, (4.12)

with the Bloch boundary conditions on βˆ‚Ξ \partial\Pi

⟧eiβ€‹π’Œβ‹…π’™ui(𝒙)⟦\displaystyle\left\rrbracket\mathrm{e}^{\mathrm{i}{\bm{k}}\cdot{\bm{x}}}u_{i}({\bm{x}})\right\llbracket =0.\displaystyle=0. (4.13)

Here we replaced kk by |π’Œ||\bm{k}| since Ξ΅=0\varepsilon=0. We write (4.11) in terms of uiu_{i}:

Li,j=βˆ«βˆ‚BR(βˆ‚uiβˆ‚r|r=R+0βˆ’βˆ‚uiβˆ‚r|r=Rβˆ’0)​eiβ€‹π’Œj⋅𝒙​dS+π’ͺ​(Ξ΄2).\displaystyle L_{i,j}=\int_{\partial B_{R}}\left(\left.\frac{\partial u_{i}}{\partial r}\right|_{r=R+0}-\left.\frac{\partial u_{i}}{\partial r}\right|_{r=R-0}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S+{\cal O}(\delta^{2}). (4.14)

Similar to solutions of (2.1), (2.2), functions uiu_{i} are infinitely smooth function of Ξ΄\delta and therefore

ui​(𝒙)=eβˆ’iβ€‹π’Œi⋅𝒙+Ξ΄β‹…vi​(𝒙)+π’ͺ​(Ξ΄2).\displaystyle u_{i}({\bm{x}})=\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}+\delta\cdot v_{i}({\bm{x}})+{\cal O}(\delta^{2}). (4.15)

Then vi​(𝒙)v_{i}({\bm{x}}) are solutions of the following problems:

(Ξ”+|π’Œ0|2)​vi​(𝒙)\displaystyle\left(\Delta+|\bm{k}_{0}|^{2}\right)v_{i}({\bm{x}}) =βˆ’2​|π’Œ0|2​eβˆ’iβ€‹π’Œi⋅𝒙,π’™βˆˆΞ ,π’™βˆ‰βˆ‚BR,vi|r=R=0,\displaystyle=-2|\bm{k}_{0}|^{2}\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}},\quad{\bm{x}}\in\Pi,\quad{\bm{x}}\notin\partial B_{R},~~\left.v_{i}\right|_{r=R}=0, (4.16)
⟧eiβ€‹π’Œ0⋅𝒙vi(𝒙)+iπ’Œ0β‹…π’™βŸ¦\displaystyle\left\rrbracket\mathrm{e}^{\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}}v_{i}({\bm{x}})+\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\right\llbracket =0,i=0,1,\displaystyle=0,\quad i=0,1, (4.17)

and (4.14) takes the form

Li,j=βˆ«βˆ‚BR(βˆ‚viβˆ‚r|r=R+0βˆ’βˆ‚viβˆ‚r|r=Rβˆ’0)​eiβ€‹π’Œj⋅𝒙​dS+π’ͺ​(Ξ΄2).\displaystyle L_{i,j}=\int_{\partial B_{R}}\left(\left.\frac{\partial v_{i}}{\partial r}\right|_{r=R+0}-\left.\frac{\partial v_{i}}{\partial r}\right|_{r=R-0}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S+{\cal O}(\delta^{2}). (4.18)

In what follows we omit subscript β€œii” from v=viv=v_{i} and add subscripts β€œ++” and β€œβˆ’-” to the restriction of vv to the domains Ξ βˆ–BR\Pi\smallsetminus B_{R} and BRB_{R}, respectively.

To evaluate the integral βˆ«βˆ‚BRβˆ‚v+βˆ‚π’β€‹eiβ€‹π’Œj⋅𝒙​dS\displaystyle\int_{\partial B_{R}}\frac{\partial v_{+}}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S, where vv is solution of (4.16)-(4.17) in Ξ βˆ–BR\Pi\smallsetminus B_{R} we apply Green’s second identity to the functions v+​(𝒙)v_{+}({\bm{x}}) and eiβ€‹π’Œj⋅𝒙\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}} in the domain Ξ βˆ–BR\Pi\smallsetminus B_{R}, j=0,1j=0,1:

βˆ«Ξ βˆ–BR(Δ​v++|π’Œ0|2​v+)​eiβ€‹π’Œj⋅𝒙​dV\displaystyle\int_{\Pi\smallsetminus B_{R}}(\Delta v_{+}+|\bm{k}_{0}|^{2}v_{+})\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}V =βˆ«βˆ‚Ξ (βˆ‚v+βˆ‚π’β€‹eiβ€‹π’Œjβ‹…π’™βˆ’v+β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚π’)​dS\displaystyle=\int_{\partial\Pi}\left(\frac{\partial v_{+}}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-v_{+}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial{\bm{n}}}\right)\mathrm{d}S
βˆ’βˆ«βˆ‚BR(βˆ‚v+βˆ‚r​eiβ€‹π’Œjβ‹…π’™βˆ’v+β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚r)​dS.\displaystyle-\int_{\partial B_{R}}\left(\frac{\partial v_{+}}{\partial r}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-v_{+}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial r}\right)\mathrm{d}S. (4.19)

From here we have

βˆ«βˆ‚BRβˆ‚v+βˆ‚r​eiβ€‹π’Œj⋅𝒙​dS=βˆ«βˆ‚Ξ (βˆ‚v+βˆ‚π’β€‹eiβ€‹π’Œjβ‹…π’™βˆ’v+β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚π’)​dS+2​|π’Œ0|2β€‹βˆ«Ξ βˆ–BRei​(π’Œjβˆ’π’Œi)⋅𝒙​dV\displaystyle\int_{\partial B_{R}}\frac{\partial v_{+}}{\partial r}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S=\int_{\partial\Pi}\left(\frac{\partial v_{+}}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-v_{+}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial{\bm{n}}}\right)\mathrm{d}S+2|\bm{k}_{0}|^{2}\int_{\Pi\smallsetminus B_{R}}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V
=βˆ«βˆ‚Ξ (βˆ‚v+βˆ‚π’β€‹eiβ€‹π’Œjβ‹…π’™βˆ’v+β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚π’)​dS+2​|π’Œ0|2β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dVβˆ’2​|π’Œ0|2β€‹βˆ«BRei​(π’Œjβˆ’π’Œi)⋅𝒙​dV.\displaystyle=\int_{\partial\Pi}\left(\frac{\partial v_{+}}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-v_{+}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial{\bm{n}}}\right)\mathrm{d}S+2|\bm{k}_{0}|^{2}\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V-2|\bm{k}_{0}|^{2}\int_{B_{R}}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V. (4.20)

Evaluation of the similar integral for vβˆ’v_{-} yields

βˆ«βˆ‚BRβˆ‚vβˆ’βˆ‚r​eiβ€‹π’Œj⋅𝒙​dS\displaystyle\int_{\partial B_{R}}\frac{\partial v_{-}}{\partial r}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S =βˆ’2​|π’Œ0|2β€‹βˆ«BRei​(π’Œjβˆ’π’Œi)⋅𝒙​dV.\displaystyle=-2|\bm{k}_{0}|^{2}\int_{B_{R}}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V. (4.21)

Thus,

βˆ«βˆ‚BR(βˆ‚v+βˆ‚rβˆ’βˆ‚vβˆ’βˆ‚r)​eiβ€‹π’Œj⋅𝒙​dS=βˆ«βˆ‚Ξ (βˆ‚v+βˆ‚π’β€‹eiβ€‹π’Œjβ‹…π’™βˆ’v+β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚π’)​dS+2​|π’Œ0|2β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV.\displaystyle\int_{\partial B_{R}}\left(\frac{\partial v_{+}}{\partial r}-\frac{\partial v_{-}}{\partial r}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S=\int_{\partial\Pi}\left(\frac{\partial v_{+}}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-v_{+}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial{\bm{n}}}\right)\mathrm{d}S+2|\bm{k}_{0}|^{2}\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V. (4.22)

In the middle integral, we make the substitution

v+​(𝒙)=βˆ’iβ€‹π’Œ0⋅𝒙​eβˆ’iβ€‹π’Œ0⋅𝒙+v~​(𝒙).\displaystyle v_{+}({\bm{x}})=-\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\,\mathrm{e}^{-\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}}+\tilde{v}({\bm{x}}). (4.23)

Then v~​(𝒙)\tilde{v}({\bm{x}}) satisfies the homogeneous condition (4.17) and the middle integral in (4.22) with v~​(𝒙)\tilde{v}({\bm{x}}) instead of v+v_{+} vanishes, and (4.22) becomes

βˆ«βˆ‚BR(βˆ‚v+βˆ‚rβˆ’βˆ‚vβˆ’βˆ‚r)​eiβ€‹π’Œj⋅𝒙​dS\displaystyle\int_{\partial B_{R}}\left(\frac{\partial v_{+}}{\partial r}-\frac{\partial v_{-}}{\partial r}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\,\mathrm{d}S =βˆ’βˆ«βˆ‚Ξ (βˆ‚(iβ€‹π’Œ0⋅𝒙​eβˆ’iβ€‹π’Œi⋅𝒙)βˆ‚π’β€‹eiβ€‹π’Œjβ‹…π’™βˆ’iβ€‹π’Œ0⋅𝒙​eβˆ’iβ€‹π’Œiβ‹…π’™β€‹βˆ‚eiβ€‹π’Œjβ‹…π’™βˆ‚π’)​dS\displaystyle=-\int_{\partial\Pi}\left(\frac{\partial(\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\,\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}})}{\partial{\bm{n}}}\,\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}-\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\,\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}\,\frac{\partial\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}}{\partial{\bm{n}}}\right)\mathrm{d}S
+2​|π’Œ0|2β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV.\displaystyle+2|\bm{k}_{0}|^{2}\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V. (4.24)

Using again Green’s second identity we obtain

βˆ«βˆ‚BR(βˆ‚v+βˆ‚rβˆ’βˆ‚vβˆ’βˆ‚r)​eiβ€‹π’Œi⋅𝒙​dS\displaystyle\int_{\partial B_{R}}\left(\frac{\partial v_{+}}{\partial r}-\frac{\partial v_{-}}{\partial r}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}\,\mathrm{d}S
=βˆ’βˆ«Ξ (eiβ€‹π’Œj⋅𝒙​(Ξ”+|π’Œ0|2)​(iβ€‹π’Œ0⋅𝒙​eβˆ’iβ€‹π’Œi⋅𝒙)βˆ’iβ€‹π’Œ0⋅𝒙​eβˆ’iβ€‹π’Œi⋅𝒙​(Ξ”+|π’Œ0|2)​eiβ€‹π’Œj⋅𝒙)​dV\displaystyle=-\int_{\Pi}\left(\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\left(\Delta+|\bm{k}_{0}|^{2}\right)\left(\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\,\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}\right)-\mathrm{i}\bm{k}_{0}\cdot{\bm{x}}\,\mathrm{e}^{-\mathrm{i}\bm{k}_{i}\cdot{\bm{x}}}\,\left(\Delta+|\bm{k}_{0}|^{2}\right)\mathrm{e}^{\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}}\right)\mathrm{d}V
+2​(π’Œ0β‹…π’Œ0)β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV=βˆ’2​(π’Œiβ‹…π’Œ0)β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV+2​(π’Œ0β‹…π’Œ0)β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV\displaystyle+2(\bm{k}_{0}\cdot\bm{k}_{0})\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V=-2(\bm{k}_{i}\cdot\bm{k}_{0})\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V+2(\bm{k}_{0}\cdot\bm{k}_{0})\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V
=2​(π’Œ0βˆ’π’Œi)β‹…π’Œ0β€‹βˆ«Ξ ei​(π’Œjβˆ’π’Œi)⋅𝒙​dV={2​(π’Ž0β‹…π’Œ0)​|Ξ |,i=j=1,0,otherwise.\displaystyle=2(\bm{k}_{0}-\bm{k}_{i})\cdot\bm{k}_{0}\int_{\Pi}\mathrm{e}^{\mathrm{i}(\bm{k}_{j}-\bm{k}_{i})\cdot{\bm{x}}}\,\mathrm{d}V=\left\{\begin{array}[]{cl}2(\bm{m}_{0}\cdot\bm{k}_{0})|\Pi|,&i=j=1,\\[5.69054pt] 0,&\text{otherwise}.\end{array}\right. (4.27)

Thus, the latter formula with equality 2β€‹π’Ž0β‹…π’Œ0=|π’Ž0|22\bm{m}_{0}\cdot\bm{k}_{0}=|\bm{m}_{0}|^{2} and (4.18) implies (4.6). ∎

5 Local and global gaps in the Dirichlet problem

To find bands and gaps, it is necessary to construct a dispersion surface which we view as the graph of a periodic infinite-valued function of the Bloch vector π’Œ\bm{k} in the entire space. The arguments at the beginning of Β§3 imply that it is enough to study this surface only in a neighborhood of the cone C0:Ο‰=c​|π’Œ|C_{0}:~\omega=c|\bm{k}|.

The following statement justifies the item (a) from the introduction.

Theorem 5.1.

If (Ο‰0,𝐀0)(\omega_{0},\bm{k}_{0}) belongs to the cone C0C_{0} but not to the intersection of the cones of the dispersion surface of the unperturbed problem (see Figure 2), then g​(𝐀0)=βˆ…g(\bm{k}_{0})=\varnothing, i. e. there is no local gap for wave vector 𝐀0\bm{k}_{0} and frequencies Ο‰\omega near Ο‰0\omega_{0}.

Proof. We need to show that local gap g​(π’Œ0)g(\bm{k}_{0}) cannot contain points Ο‰=Ο‰a\omega=\omega_{a} such that Ο‰aβ†’Ο‰0=c​|π’Œ0|\omega_{a}\to\omega_{0}=c|\bm{k}_{0}| as aβ†’0a\to 0. Let us consider the segment Ο‰=Ο‰a,π’Œ=(1+Ξ΄)β€‹π’Œ0\omega=\omega_{a},~\bm{k}=(1+\delta)\bm{k}_{0} in ℝ4{\mathbb{R}}^{4}, where |Ξ΄|<Ξ΄0|\delta|<\delta_{0} with some aa-independent Ξ΄0>0\delta_{0}>0. This segment intersects the cone C0C_{0} when Ξ΄=Ξ΄a\delta=\delta_{a} where Ξ΄aβ†’0\delta_{a}\to 0 as aβ†’0a\to 0. The same remains true with a different Ξ΄a=Ξ΄aβ€²\delta_{a}=\delta_{a}^{\prime} if C0C_{0} is replaced by the dispersion surface of the problem with inclusions since the latter surface depends smoothly on π’Œ\bm{k} and aa when points π’Œ\bm{k} are not exceptional and aa is small, see Lemma 3.1. Hence there is a Bloch wave with parameters (Ο‰a,(1+Ξ΄aβ€²)β€‹π’Œ0),Ξ΄aβ€²β†’0(\omega_{a},(1+\delta_{a}^{\prime})\bm{k}_{0}),~\delta_{a}^{\prime}\to 0 as aβ†’0a\to 0, and therefore Ο‰aβˆ‰g​(π’Œ0)\omega_{a}\notin g(\bm{k}_{0}). ∎

We fix an interval (Ο΅,Ο΅βˆ’1(\epsilon,\epsilon^{-1}) of frequencies Ο‰\omega with an arbitrary small Ο΅>0\epsilon>0.

To prove item (b) we need the following Lemma concerning the dispersion surface located above the interval π’Œ=(1+Ξ΄)β€‹π’Œ0,|Ξ΄|<Ξ΄0\bm{k}=(1+\delta)\bm{k}_{0},~|\delta|<\delta_{0} with some aa-independent Ξ΄0>0\delta_{0}>0. For simplicity of the formulas below, we introduce in (4.10) new variables

a~=4​π​a​q|Ξ |,Ξ΄~=δ​|π’Ž0|22.\displaystyle\tilde{a}=\dfrac{4\pi aq}{|\Pi|},\quad{\tilde{\delta}}=\frac{\delta|\bm{m}_{0}|^{2}}{2}. (5.1)
Lemma 5.1.

There is Ξ΄0>0\delta_{0}>0 such that the dispersion surface in the neighborhood of the point (Ο‰0,𝐀0)∈C0,Ο‰0=c​|𝐀0|,(\omega_{0},\bm{k}_{0})\in C_{0},~\omega_{0}=c|\bm{k}_{0}|, above the interval 𝐀=(1+Ξ΄)​𝐀0\bm{k}=(1+\delta)\bm{k}_{0}, |Ξ΄|<|Ξ΄0||\delta|<|\delta_{0}|, is split into the two branches determined by

ω±/c\displaystyle\omega_{\pm}/c =|π’Œ0|+12​|π’Œ0|​(a~+ν​δ~Β±a~2+Ξ΄~2)+π’ͺ​(a~2+Ξ΄~2),|Ξ΄~|β©½Ξ΄0,a~β‰ͺ1,\displaystyle=|\bm{k}_{0}|+\frac{1}{2|\bm{k}_{0}|}\left({\tilde{a}}+\nu{\tilde{\delta}}\pm\sqrt{{\tilde{a}}^{2}+{\tilde{\delta}}^{2}}\right)+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}),\quad|\tilde{\delta}|\leqslant\delta_{0},~~\tilde{a}\ll 1, (5.2)

where

Ξ½\displaystyle\nu =4​|π’Œ0|2|π’Ž0|2βˆ’1β©Ύ0.\displaystyle=\frac{4|\bm{k}_{0}|^{2}}{|\bm{m}_{0}|^{2}}-1\geqslant 0. (5.3)

Proof. The dispersion relation near the point π’Œ0\bm{k}_{0} is determined by the equation k=(1+Ξ΅)​|π’Œ|k=(1+\varepsilon)|\bm{k}|, where Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) can be obtained from the EZE condition for the operator \mathsfbfit​N:=\mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​Na,Ξ΅βˆ’\mathsfbfit N:={\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{a,\varepsilon}, see Lemma 2.1. We studied this EZE problem in Β§3 when π’Œ=π’Œ0\bm{k}=\bm{k}_{0} by using the matrix representation (3.8) of the operator \mathsfbfit​N\mathsfbfit N, and this representation allowed us to reduce the calculation of the function Ξ΅=Ρ​(a,π’Œ)\varepsilon=\varepsilon(a,\bm{k}) to the equation det𝑺~=0,\det\widetilde{\bm{S}}=0, where 𝑺~\widetilde{\bm{S}} is a 2Γ—22\times 2 matrix that differs from the matrix 𝑺\bm{S} given by (3.7) only by the values of the remainder terms. The smoothness of the remainder terms of matrices 𝑺\bm{S} and 𝑺~\widetilde{\bm{S}} was established using Lemma 2.1.

Theorem 4.1 provides analogues of the matrix representation (3.8) and formula (3.7) on the whole interval π’Œ=(1+Ξ΄)β€‹π’Œ0\bm{k}=(1+\delta)\bm{k}_{0}, |Ξ΄|<|Ξ΄0||\delta|<|\delta_{0}|. The same arguments used to find Ξ΅\varepsilon for π’Œ=π’Œ0\bm{k}=\bm{k}_{0} in Β§3, now allow us to find Ξ΅\varepsilon on the whole interval. These arguments lead to

det[2​Ρ​|π’Œ|2​|Ξ |β€‹π‘°βˆ’4​π​a​q​𝑱+δ​|Ξ |​[000|π’Ž0|2]+π’ͺ​(a2+Ξ΄2+Ξ΅2)]=0,π’Œ=(1+Ξ΄)β€‹π’Œ0,\displaystyle\det\left[2\varepsilon|\bm{k}|^{2}|\Pi|\bm{I}-4\pi aq\bm{J}+\delta|\Pi|\left[\begin{array}[]{cc}0&0\\[5.69054pt] 0&|\bm{m}_{0}|^{2}\end{array}\right]+{\cal O}(a^{2}+\delta^{2}+\varepsilon^{2})\right]=0,\quad\bm{k}=(1+\delta)\bm{k}_{0}, (5.6)

where |Ξ΄|<|Ξ΄0||\delta|<|\delta_{0}| and the remainder term is infinitely smooth.

In new variables (5.1) equation (5.6) is equivalent to

det\displaystyle\det [Ξ΅~βˆ’a~+π’ͺ​(a~2+Ξ΄~2+Ξ΅~2)βˆ’a~+π’ͺ​(a~2+Ξ΄~2+Ξ΅~2)βˆ’a~+π’ͺ​(a~2+Ξ΄~2+Ξ΅~2)Ξ΅~βˆ’a~+2​δ~+π’ͺ​(a~2+Ξ΄~2+Ξ΅~2)]\displaystyle\left[\begin{array}[]{cc}\tilde{\varepsilon}-\tilde{a}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}+\tilde{\varepsilon}^{2})&-\tilde{a}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}+\tilde{\varepsilon}^{2})\\[5.69054pt] -\tilde{a}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}+\tilde{\varepsilon}^{2})&\tilde{\varepsilon}-\tilde{a}+2\tilde{\delta}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}+\tilde{\varepsilon}^{2})\end{array}\right] (5.8)
=Ξ΅~2+2​(Ξ΄~βˆ’a~)​Ρ~βˆ’2​δ~​a~+π’ͺ​(|a~|3+|Ξ΄~|3+|Ξ΅~|3)=0,\displaystyle=\tilde{\varepsilon}^{2}+2(\tilde{\delta}-\tilde{a})\tilde{\varepsilon}-2\tilde{\delta}\tilde{a}+{\cal O}(|\tilde{a}|^{3}+|\tilde{\delta}|^{3}+|\tilde{\varepsilon}|^{3})=0, (5.9)

where Ξ΅~=2​Ρ​|π’Œ|2\tilde{\varepsilon}=2\varepsilon|\bm{k}|^{2}. If the term |Ξ΅~|3|\tilde{\varepsilon}|^{3} in the remainder is omitted, then the roots of the simplified equation (5.9) are

Ξ΅~=a~βˆ’Ξ΄~Β±a~2+Ξ΄~2+π’ͺ​(a~2+Ξ΄~2)\displaystyle\tilde{\varepsilon}=\tilde{a}-\tilde{\delta}\pm\sqrt{\tilde{a}^{2}+\tilde{\delta}^{2}}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2}) (5.10)

The same is true without the simplification of the equation but there is a difficulty to justify this fact because the equation degenerates at the point of interest a~=Ξ΄~=0\tilde{a}=\tilde{\delta}=0, and we cannot apply the implicit function theorem to determine Ξ΅~\tilde{\varepsilon}. The same difficulty is present in equation (5.6) if we view 2​Ρ​|π’Œ|2​|Ξ |2\varepsilon|\bm{k}|^{2}|\Pi| as the eigenvalue of the remaining terms of the matrix (5.6). Then the zero eigenvalue of the matrix has multiplicity two when a~=Ξ΄~=0\tilde{a}=\tilde{\delta}=0. Therefore we cannot guarantee a smooth dependence of the eigenvalue on the parameters. In our case the situation is even worse since the remainder depends on the eigenvalue.

We will proceed as follows. We rewrite (5.9) in the form

(Ξ΅~βˆ’a~+Ξ΄~)2=a~2+Ξ΄~2+π’ͺ​(|a~|3+|Ξ΄~|3+|Ξ΅~|3),\displaystyle(\tilde{\varepsilon}-\tilde{a}+\tilde{\delta})^{2}=\tilde{a}^{2}+\tilde{\delta}^{2}+{\cal O}(|\tilde{a}|^{3}+|\tilde{\delta}|^{3}+|\tilde{\varepsilon}|^{3}), (5.11)

divide this equation by a~2+Ξ΄~2\tilde{a}^{2}+\tilde{\delta}^{2}, introduce polar coordinates (Ο±,Ο†)(\varrho,\varphi) in the plane (a~,Ξ΄~)(\tilde{a},\tilde{\delta}):

Ο±=a~2+Ξ΄~2,cos⁑φ=a~/Ο±,sin⁑φ=Ξ΄~/Ο±,\displaystyle\varrho=\sqrt{\tilde{a}^{2}+\tilde{\delta}^{2}},\quad\cos\varphi=\tilde{a}/\varrho,\quad\sin\varphi=\tilde{\delta}/\varrho, (5.12)

and replace Ξ΅~\tilde{\varepsilon} by Ξ·~=Ξ΅~/a~2+Ξ΄~2\tilde{\eta}=\tilde{\varepsilon}/\sqrt{\tilde{a}^{2}+\tilde{\delta}^{2}}. In the new variables (5.11) reads

(Ξ·~βˆ’cos⁑φ+sin⁑φ)2=1+π’ͺ​(|a~|+|Ξ΄~|+ϱ​|Ξ·~|).\displaystyle(\tilde{\eta}-\cos\varphi+\sin\varphi)^{2}=1+{\cal O}(|\tilde{a}|+|\tilde{\delta}|+\varrho|\tilde{\eta}|). (5.13)

Now we can apply the implicit function theorem and solve (5.13) for Ξ·~\tilde{\eta}. This leads to (5.10). Since Ξ΅~=2​Ρ​|π’Œ|2\tilde{\varepsilon}=2\varepsilon|\bm{k}|^{2} and Ρ​|π’Œ|\varepsilon|\bm{k}| here can be replaced by Ο‰cβˆ’|π’Œ|\frac{\omega}{c}-|\bm{k}|, see (2.3), equation (5.10) can be rewritten as

Ο‰cβˆ’|π’Œ|=12​|π’Œ|​[a~βˆ’Ξ΄~Β±a~2+Ξ΄~2+π’ͺ​(a~2+Ξ΄~2)].\displaystyle\frac{\omega}{c}-|\bm{k}|=\frac{1}{2|\bm{k}|}[\tilde{a}-\tilde{\delta}\pm\sqrt{\tilde{a}^{2}+\tilde{\delta}^{2}}+{\cal O}(\tilde{a}^{2}+\tilde{\delta}^{2})].

To complete the proof of the lemma, it remains to use the relation between π’Œ\bm{k} and π’Œ0\bm{k}_{0}:

π’Œ=(1+Ξ΄)β€‹π’Œ0=(1+2​δ~/|π’Ž0|2)β€‹π’Œ0.\bm{k}=(1+\delta)\bm{k}_{0}=(1+2\tilde{\delta}/|\bm{m}_{0}|^{2})\bm{k}_{0}.

∎

Theorem 5.2.

Let 𝐀0\bm{k}_{0} be an exceptional point of order two and (𝐀0,𝐦0)(\bm{k}_{0},\bm{m}_{0}) satisfy (1.5).

  1. 1.

    If |π’Œ0||π’Ž0|<22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}<\frac{\sqrt{2}}{2}, then the local gap g​(π’Œ0)g(\bm{k}_{0}) exists in a neighborhood of Ο‰0=c​|π’Œ0|\omega_{0}=c|\bm{k}_{0}| for small enough aa and consists of frequencies Ο‰=Ο‰a\omega=\omega_{a} for which

    |π’Œ0|+a~β€‹Ξ½βˆ’+π’ͺ​(a2)<Ο‰/c<|π’Œ0|+a~​ν++π’ͺ​(a2),Ξ½Β±=1Β±1βˆ’Ξ½2,\displaystyle\displaystyle|\bm{k}_{0}|+\tilde{a}\nu_{-}+{\cal O}(a^{2})<\omega/c<|\bm{k}_{0}|+\tilde{a}\nu_{+}+{\cal O}(a^{2}),\quad\nu_{\pm}=1\pm\sqrt{1-\nu^{2}}, (5.14)

    where a~\tilde{a} is given by (5.1) and Ξ½\nu is given by (5.3).

  2. 2.

    If |π’Œ0||π’Ž0|>22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}>\frac{\sqrt{2}}{2} and aa is small enough, then g​(π’Œ0)β‰ βˆ…g(\bm{k}_{0})\neq\varnothing.

Remark 1.

The location of the Bloch vector on the surface of the first Brillouin zone for which waves cannot propagate is shown in Figure 3 by shaded unit disks.

0.50.50.50.50.50.5k1k_{1}k2k_{2}k3k_{3}
Figure 3: Shaded unit discs on the faces of the first Brillouin zone show the location of the Bloch vector where waves cannot propagate.

Proof. Local gap g​(π’Œ0)g(\bm{k}_{0}) consists of frequencies Ο‰=Ο‰a\omega=\omega_{a} such that Ο‰aβ†’c​|π’Œ0|\omega_{a}\to c|\bm{k}_{0}| as aβ†’0a\to 0 and the interval lβˆˆβ„4l\in\mathbb{R}^{4} defined by equations Ο‰=Ο‰a,π’Œ=(1+Ξ΄~)β€‹π’Œ0,|Ξ΄~|<Ξ΄~0,\omega=\omega_{a},~\bm{k}=(1+\tilde{\delta})\bm{k}_{0},~|\tilde{\delta}|<\tilde{\delta}_{0}, with some aa-independent value of Ξ΄~0>0\tilde{\delta}_{0}>0 does not intersect the branches of the dispersion surface described in Lemma (5.1).

We start the proof with the second part. Let |π’Œ0||π’Ž0|>22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}>\frac{\sqrt{2}}{2}. Then for Ξ½\nu defined in Lemma (5.1) we have Ξ½>1\nu>1. Using the inequality a~2+Ξ΄~2>|Ξ΄~|\sqrt{\tilde{a}^{2}+\tilde{\delta}^{2}}>|\tilde{\delta}|, we obtain the following estimates for function Ο‰=Ο‰+\omega=\omega_{+} defined in (5.2) with small, aa-independent Ξ΄~0>0\tilde{\delta}_{0}>0 and aβ†’0a\to 0:

Ο‰+c|Ξ΄~=Ξ΄~0βˆ’|π’Œ0|β©Ύ(Ξ½+1)​δ~0/2>0,Ο‰+c|Ξ΄~=βˆ’Ξ΄~0βˆ’|π’Œ0|β©½(βˆ’Ξ½+1)​δ~0/2<0.\left.\frac{\omega_{+}}{c}\right|_{\tilde{\delta}=\tilde{\delta}_{0}}-|\bm{k}_{0}|\geqslant(\nu+1)\tilde{\delta}_{0}/2>0,\quad\left.\frac{\omega_{+}}{c}\right|_{\tilde{\delta}=-\tilde{\delta}_{0}}-|\bm{k}_{0}|\leqslant(-\nu+1)\tilde{\delta}_{0}/2<0.

Hence, the range of the function Ο‰+/cβˆ’|π’Œ0|,|Ξ΄~|<Ξ΄~0,\omega_{+}/c-|\bm{k}_{0}|,~|\tilde{\delta}|<\tilde{\delta}_{0}, contains the segment [(βˆ’Ξ½+1)​δ~0/2,(Ξ½+1)​δ~0/2][(-\nu+1)\tilde{\delta}_{0}/2,(\nu+1)\tilde{\delta}_{0}/2] if aa is small enough. Thus, for any Ο‰a\omega_{a} close enough to c​|π’Œ0|c|\bm{k}_{0}|, there exists Ξ΄~a\tilde{\delta}_{a} such that |Ξ΄~a|<Ξ΄~0|\tilde{\delta}_{a}|<\tilde{\delta}_{0} and Ο‰+=Ο‰a\omega_{+}=\omega_{a}. In other words, the ray ll intersect the graph of the dispersion surface Ο‰+\omega_{+} when Ξ΄~=Ξ΄~a,Ο‰+=Ο‰a\tilde{\delta}=\tilde{\delta}_{a},~\omega_{+}=\omega_{a}, the Bloch wave propagates with the frequency Ο‰a\omega_{a} and π’Œ=(1+Ξ΄~a)​|π’Œ0|\bm{k}=(1+\tilde{\delta}_{a})|\bm{k}_{0}|, and therefore g​(π’Œ0)β‰ βˆ…g(\bm{k}_{0})\neq\varnothing. Note that the same result can be obtained using Ο‰βˆ’\omega_{-} instead of Ο‰+\omega_{+}.

Let now |π’Œ0||π’Ž0|<22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}<\frac{\sqrt{2}}{2}, and therefore Ξ½<1\nu<1. Denote by ω±0\omega_{\pm}^{0} functions ω±\omega_{\pm} defined in (5.2) with the remainder terms omitted. The minimum of Ο‰+0/cβˆ’|π’Œ0|\omega_{+}^{0}/c-|\bm{k}_{0}| is attained at Ξ΄~=βˆ’a~​ν1βˆ’Ξ½2\tilde{\delta}=-\frac{\tilde{a}\nu}{\sqrt{1-\nu^{2}}} with the value Ο‰+0/cβˆ’|π’Œ0|=a~​ν+\omega_{+}^{0}/c-|\bm{k}_{0}|=\tilde{a}\nu_{+}. Similarly, the maximum of Ο‰βˆ’0/cβˆ’|π’Œ0|\omega_{-}^{0}/c-|\bm{k}_{0}| is attained at Ξ΄~=a~​ν1βˆ’Ξ½2\tilde{\delta}=\frac{\tilde{a}\nu}{\sqrt{1-\nu^{2}}} with the value Ο‰βˆ’0/cβˆ’|π’Œ0|=a~β€‹Ξ½βˆ’\omega_{-}^{0}/c-|\bm{k}_{0}|=\tilde{a}\nu_{-}. Thus, the gap between the ranges of these two functions is the interval (a~β€‹Ξ½βˆ’,a~​ν+)(\tilde{a}\nu_{-},~\tilde{a}\nu_{+}). If |Ξ΄~|β©½C​a~|\tilde{\delta}|\leqslant C\tilde{a} with a large, but fixed aa-independent constant CC, then (5.2) implies that the gap between ranges of functions ω±/cβˆ’|π’Œ0|\omega_{\pm}/c-|\bm{k}_{0}| is the interval I:=(a~β€‹Ξ½βˆ’+O​(a2),a~​ν++O​(a2))I:=(\tilde{a}\nu_{-}+O(a^{2}),~\tilde{a}\nu_{+}+O(a^{2})). When C​a~β©½|Ξ΄~|β©½Ξ΄~0C\tilde{a}\leqslant|\tilde{\delta}|\leqslant\tilde{\delta}_{0} and Ξ΄~0\tilde{\delta}_{0} is small enough, the gap between functions ω±/cβˆ’|π’Œ0|\omega_{\pm}/c-|\bm{k}_{0}| can be estimated through Ξ΄~|\tilde{\delta}| and this gap includes interval II. Hence, II is the exact interval such that the segment ll introduced at the beginning of the proof does not intersect the dispersion surface when Ο‰a/cβˆ’|π’Œ0|∈I\omega_{a}/c-|\bm{k}_{0}|\in I. This completes the proof of the theorem. ∎

Finally, we prove the theorem on the absence of global gaps under small perturbations.

Theorem 5.3.

Global gaps do not exist in any fixed interval Ο΅<Ο‰<Ο΅βˆ’1\epsilon<\omega<\epsilon^{-1} of the time frequency Ο‰\omega if the size aa of the inclusion is sufficiently small.

Proof. We will provide two proofs: the first one is geometrical and the second is shorter but more formal. For clarity, we explain the first proof in the two-dimensional case shown in Figure 2.

For Ο‰0\omega_{0} to be in a global gap, the plane Ο‰=Ο‰0\omega=\omega_{0} must not intersect the dispersion surface of the problem with inclusions. Hence the surface in Figure 2 must break when a>0a>0 and create a gap allowing the plane Ο‰=Ο‰0\omega=\omega_{0} to go through without intersection with the surface. It could happen only near the intersection of C0C_{0} with other cones where the dispersion surface could split and move apart when a>0a>0. This splitting takes place, and the line of intersection of the cones where the splitting occurs is shown in Figure 2 by the dashed curve. Its projection on β„π’Œ2{\mathbb{R}}^{2}_{\bm{k}} is the line A​BAB consisting of exceptional points of multiplicity two given by (1.6). However, the plane Ο‰=Ο‰0\omega=\omega_{0} intersects the cone C0C_{0} along the circle Ο‰0=c​|π’Œ|\omega_{0}=c|\bm{k}| and the splitting when π’Œ\bm{k} is on the circle may occur only in a small neighborhood of exceptional points near the circle. The plane intersects the dispersion surface in the neighborhood of other points on the circle and therefore Ο‰0\omega_{0} does not belong to a global gap.

An alternative proof is based on Theorem 5.1. The plane Ο‰=Ο‰0\omega=\omega_{0} interests the cone C0C_{0} at the sphere |π’Œ|2=Ο‰02/c2|\bm{k}|^{2}=\omega_{0}^{2}/c^{2} which contains mostly non-exceptional values of π’Œ\bm{k}. For such non-exceptional point π’Œ=π’Œ0\bm{k}=\bm{k}_{0}, the set g​(π’Œ0)=βˆ…g(\bm{k}_{0})=\varnothing. According to Theorem 5.1, Ο‰0\omega_{0} cannot belong to the local gap g​(π’Œ0)g(\bm{k}_{0}) and therefore to a global as well. ∎

6 Local and global gaps in the transmission problem

The amplitude uu of the Bloch solution of the transmission problem satisfies the equation

Δ​u+kΒ±2​u=0,π’™βˆˆβ„βˆ–βˆ‚Ξ©,\displaystyle\Delta u+k^{2}_{\pm}u=0,\quad{\bm{x}}\in{\mathbb{R}}\smallsetminus\partial\Omega, (6.1)

with the boundary conditions on βˆ‚Ξ©\partial\Omega

⟦u(𝒙)⟧=0,⟦1ϱ​(𝒙)βˆ‚u​(𝒙)βˆ‚π’βŸ§=0,\displaystyle\llbracket u({\bm{x}})\rrbracket=0,\quad\left\llbracket\frac{1}{\varrho({\bm{x}})}\frac{\partial u({\bm{x}})}{\partial{\bm{n}}}\right\rrbracket=0, (6.2)

where kΒ±=Ο‰/cΒ±k_{\pm}=\omega/c_{\pm}, cΒ±=1/γ±​ϱ±c_{\pm}=1/\sqrt{\gamma_{\pm}\varrho_{\pm}} is the speed of the wave propagation in the medium and inclusion, respectively, Ξ³Β±\gamma_{\pm} is the adiabatic bulk compressibility modulus, and ϱ​(𝒙)\varrho({\bm{x}}) is the mass density. Here and on the subscript Β±\pm refers to the value of the quantity outside/inside of the inclusion. The brackets βŸ¦β‹…βŸ§\llbracket\cdot\rrbracket denote the jump of the enclosed quantity across βˆ‚Ξ©\partial\Omega.

The analysis of gaps in the transmission problem is similar to that in the Dirichlet problem. In particular, Lemma 4.1 about matrix representation of \mathsfbfit​Nπ’Œ,Ξ΅+βˆ’\mathsfbfit​N0,Ξ΅βˆ’{\mathsfbfit N}^{+}_{\bm{k},\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon} remains valid with the same representation of matrix 𝑳\bm{L}. Also, Theorem 5.1 about the absence of local gaps near simple points (Ο‰0,π’Œ0)(\omega_{0},\bm{k}_{0}) of the cone C0C_{0} still holds for the transmission problem.

While the exterior operators of the Dirichlet and transmission problems coincide, the interior operators have a different form. For the exceptional wave vector π’Œ0\bm{k}_{0} of order nn such that its integer-valued shifts π’Œi=π’Œ0βˆ’π’Ži\bm{k}_{i}=\bm{k}_{0}-\bm{m}_{i}, π’Žiβˆˆβ„€3,0β©½iβ©½nβˆ’1\bm{m}_{i}\in{\mathbb{Z}}^{3},~0\leqslant i\leqslant n-1 are also satisfy (1.1), (1.2), the asymptotics of the quadratic form

((\mathsfbfit​Na,Ξ΅βˆ’βˆ’\mathsfbfit​N0,Ξ΅βˆ’)β€‹Οˆi,ψj),ψj=eβˆ’iβ€‹π’Œj⋅𝒙,0β©½i,jβ©½nβˆ’1,\displaystyle\left(\left({\mathsfbfit N}^{-}_{a,\varepsilon}-{\mathsfbfit N}^{-}_{0,\varepsilon}\right)\psi_{i},\psi_{j}\right),\quad\psi_{j}=\mathrm{e}^{-\mathrm{i}\bm{k}_{j}\cdot{\bm{x}}},\quad 0\leqslant i,j\leqslant n-1, (6.3)

is given by the matrix 𝑴\bm{M} obtained in [18]. For simplicity, we provide the form of this matrix in the case of spherical inclusions

Mi,j=|Ξ |​|π’Œ02|​(Ξ±+Ξ²β€‹π’Œ^iβ‹…π’Œ^j)​f+π’ͺ​(a4+a3​|Ξ΅|),0β©½i,jβ©½nβˆ’1,\displaystyle M_{i,j}=|\Pi||\bm{k}_{0}^{2}|\left(\alpha+\beta\,\hat{\bm{k}}_{i}\cdot\hat{\bm{k}}_{j}\right)f+{\cal O}\left(a^{4}+a^{3}|\varepsilon|\right),\quad 0\leqslant i,j\leqslant n-1, (6.4)

where

Ξ±=1βˆ’Ξ³βˆ’Ξ³+,Ξ²=3β€‹Οƒβˆ’1Οƒ+2,Οƒ=Ο±+Ο±βˆ’.\displaystyle\alpha=1-\frac{\gamma_{-}}{\gamma_{+}},\quad\beta=3\,\frac{\sigma-1}{\sigma+2},\quad\sigma=\frac{\varrho_{+}}{\varrho_{-}}. (6.5)

Here π’Œ^=π’Œ/|π’Œ|\hat{\bm{k}}=\bm{k}/|\bm{k}| and ff is the volume fraction of the inclusions. Matrix 𝑴\bm{M} plays the role of the matrix 4​π​a​q​𝑱4\pi aq\bm{J} in (4.10) for the Dirichlet problem. Therefore, if π’Œ0\bm{k}_{0} is an exceptional vector of multiplicity two, the 2Γ—22\times 2 matrix 𝑺\bm{S} in (4.10) has the form

𝑺\displaystyle\bm{S} =2​Ρ​|π’Œ0|2​|Ξ |β€‹π‘°βˆ’f​|π’Œ0|2​|Ξ |​[Ξ±+Ξ²Ξ±+Ξ²β€‹π’Œ^0β‹…π’Œ^1Ξ±+Ξ²β€‹π’Œ^0β‹…π’Œ^1Ξ±+Ξ²]+δ​|Ξ |​[000|π’Ž0|2]\displaystyle=2\varepsilon|\bm{k}_{0}|^{2}|\Pi|\bm{I}-f|\bm{k}_{0}|^{2}|\Pi|\left[\begin{array}[]{cc}\alpha+\beta&\alpha+\beta\,\hat{\bm{k}}_{0}\cdot\hat{\bm{k}}_{1}\\[5.69054pt] \alpha+\beta\,\hat{\bm{k}}_{0}\cdot\hat{\bm{k}}_{1}&\alpha+\beta\end{array}\right]+\delta|\Pi|\left[\begin{array}[]{cc}0&0\\[5.69054pt] 0&|\bm{m}_{0}|^{2}\end{array}\right] (6.10)
+π’ͺ​(a4+a3​|Ξ΅|+Ξ΄2+Ξ΅2).\displaystyle+{\cal O}(a^{4}+a^{3}|\varepsilon|+\delta^{2}+\varepsilon^{2}). (6.11)

Equating the leading term of the determinant of 𝑺\bm{S} to zero we obtain a quadratic equation for Ξ΅\varepsilon

4​|π’Œ0|2​Ρ2+4​Ρ​(Ξ΄~βˆ’f​|π’Œ0|2​(Ξ±+Ξ²))+f2​|π’Œ0|2​((Ξ±+Ξ²)2βˆ’(Ξ±+Ξ²β€‹π’Œ^0β‹…π’Œ^1)2)βˆ’2​δ~​f​(Ξ±+Ξ²)=0.\displaystyle 4|\bm{k}_{0}|^{2}\varepsilon^{2}+4\varepsilon\left(\tilde{\delta}-f|\bm{k}_{0}|^{2}(\alpha+\beta)\right)+f^{2}|\bm{k}_{0}|^{2}\left((\alpha+\beta)^{2}-(\alpha+\beta\,\hat{\bm{k}}_{0}\cdot\hat{\bm{k}}_{1})^{2}\right)-2\tilde{\delta}f(\alpha+\beta)=0. (6.12)

The analog of Lemma 5.1 is

Lemma 6.1.

There is Ξ΄0>0\delta_{0}>0 such that the dispersion surface in the neighborhood of the point (Ο‰0,𝐀0)∈C0,Ο‰0=c​|𝐀0|,(\omega_{0},\bm{k}_{0})\in C_{0},~\omega_{0}=c|\bm{k}_{0}|, above the interval 𝐀=(1+Ξ΄)​𝐀0\bm{k}=(1+\delta)\bm{k}_{0}, |Ξ΄|<|Ξ΄0||\delta|<|\delta_{0}|, is split into the two branches determined by

ω±/c\displaystyle\omega_{\pm}/c =|π’Œ0|​(1+12​(Ξ±+Ξ²)​f)\displaystyle=|\bm{k}_{0}|\left(1+\frac{1}{2}(\alpha+\beta)f\right)
+12​|π’Œ0|​(ν​δ~Β±(Ξ±+Ξ²β€‹π’Œ^0β‹…π’Œ^1)2​|π’Œ0|4​f2+Ξ΄~2)+π’ͺ​(a4+Ξ΄~2),|Ξ΄~|β©½Ξ΄0,aβ‰ͺ1,\displaystyle+\frac{1}{2|\bm{k}_{0}|}\left(\nu\tilde{\delta}\pm\sqrt{(\alpha+\beta\,\hat{\bm{k}}_{0}\cdot\hat{\bm{k}}_{1})^{2}|\bm{k}_{0}|^{4}f^{2}+\tilde{\delta}^{2}}\right)+{\cal O}(a^{4}+\tilde{\delta}^{2}),\quad|\tilde{\delta}|\leqslant\delta_{0},~~a\ll 1, (6.13)

where Ξ΄~\tilde{\delta} and Ξ½\nu are given by (5.1) and (5.3). This expression is similar to (5.2) and we can use the same analysis to find the location of local gaps. The result is given by

Theorem 6.1.

Let 𝐀0\bm{k}_{0} be an exceptional point of order two and (𝐀0,𝐦0)(\bm{k}_{0},\bm{m}_{0}) satisfy (1.5).

  1. 1.

    If |π’Œ0||π’Ž0|<22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}<\frac{\sqrt{2}}{2}, then the local gap g​(π’Œ0)g(\bm{k}_{0}) exists in a neighborhood of Ο‰0=c​|π’Œ0|\omega_{0}=c|\bm{k}_{0}| for small enough aa and consists of frequencies Ο‰=Ο‰a\omega=\omega_{a} for which

    |π’Œ~0|βˆ’ΞΌ2​|π’Œ0|​1βˆ’Ξ½2+π’ͺ​(a4)<Ο‰/c<|π’Œ~0|+ΞΌ2​|π’Œ0|​1βˆ’Ξ½2+π’ͺ​(a4),\displaystyle|\tilde{\bm{k}}_{0}|-\frac{\mu}{2|\bm{k}_{0}|}\sqrt{1-\nu^{2}}+{\cal O}(a^{4})<\omega/c<|\tilde{\bm{k}}_{0}|+\frac{\mu}{2|\bm{k}_{0}|}\sqrt{1-\nu^{2}}+{\cal O}(a^{4}), (6.14)

    where |π’Œ~0|=|π’Œ0|​(1+12​(Ξ±+Ξ²)​f)|\tilde{\bm{k}}_{0}|=|\bm{k}_{0}|\left(1+\frac{1}{2}(\alpha+\beta)f\right), ΞΌ=|Ξ±+Ξ²β€‹π’Œ^0β‹…π’Œ^1|​|π’Œ0|2​f\mu=|\alpha+\beta\,\hat{\bm{k}}_{0}\cdot\hat{\bm{k}}_{1}||\bm{k}_{0}|^{2}f, and Ξ½\nu is given by (5.3).

  2. 2.

    If |π’Œ0||π’Ž0|>22\displaystyle\frac{|\bm{k}_{0}|}{|\bm{m}_{0}|}>\frac{\sqrt{2}}{2} and aa is small enough, then g​(π’Œ0)β‰ βˆ…g(\bm{k}_{0})\neq\varnothing.

Note that the existence of absolute gaps has been shown numerically for cubic arrays of air bubbles in water [20] but no absolute gaps were found for cubic arrays of rigid spheres in air [21].

7 Conclusions

We considered the propagation of acoustic waves in a medium containing a simple cubic array of small inclusions of arbitrary shape. We show that global gaps do not exist in each fixed interval of the time frequency if the size of inclusions is small enough. The notion of local gaps defied by wave vectors π’Œ\bm{k} is introduced and studied. We calculate the location of local gaps analytically for the Dirichlet and transmission problems. The width of the local bandgaps in the Dirichlet problem is proportional to the size of the inclusions while in the transmission problem problem, it is proportional to the volume fraction of the inclusions. Although the bandgaps in the two problems were different, they exist for the same values of the Bloch vector π’Œ\bm{k}. In particular, local gaps exist on the boundary of the first Brillouin zone which is the unit cube in our scaling if π’Œ\bm{k} belongs to the unit disks on the faces of the cube.

References

  • [1] Kushwaha MS, Halevi P, MartΓ­nez G, Dobrzynski L, Djafari-Rouhani B. 1994 Theory of acoustic band structure of periodic elastic composites. Physical Review B 49, 2313–2322.
  • [2] Johnson S, Joannopoulos J. 2001 Block-iterative frequency-domain methods for Maxwell’s equations in a planewave basis. Optics Express 8, 173–190.
  • [3] Wu T, Huang Z, Lin S. 2004 Surface and bulk acoustic waves in two-dimensional phononic crystal consisting of materials with general anisotropy. Physical Review B 69.
  • [4] Fan S, Villeneuve P, Joannopoulos J. 1996 Large omnidirectional band gaps in metallodielectric photonic crystals. Physical Review B 54, 11245–11251.
  • [5] Tanaka Y, Tomoyasu Y, Tamura S. 2000 Band structure of acoustic waves in phononic lattices: Two-dimensional composites with large acoustic mismatch. Physical Review B 62, 7387–7392.
  • [6] Taflove A, Hagness SC. 2005 Computational Electrodynamics: The Finite-Difference Time-Domain Method. Artech House Publishers, 3rd edition.
  • [7] Axmann W, Kuchment P. 1999 An efficient finite element method for computing spectra of photonic and acoustic band-gap materials - I. Scalar case. Journal of Computational Physics 150, 468–481.
  • [8] Dobson D. 1999 An efficient method for band structure calculations in 2D photonic crystals. Journal of Computational Physics 149, 363–376.
  • [9] Giani S, Graham IG. 2012 Adaptive finite element methods for computing band gaps in photonic crystals. Numerische Mathematik 121, 31–64.
  • [10] Joannopoulos JD, Johnson SG, Winn JN, Meade RD. 2011 Photonic Crystals: Molding the Flow of Light. Princeton, NJ: Princeton University Press.
  • [11] Li FL, Wang YS, Zhang C, Yu GL. 2013 Bandgap calculations of two-dimensional solid-fluid phononic crystals with the boundary element method. Wave Motion 50, 525–541.
  • [12] Ashcroft NW, Mermin ND. 1976 Solid state physics. New York: Holt, Rinehart and Winston.
  • [13] Kittel C. 2004 Introduction to Solid State Physics. Wiley 8th edition.
  • [14] Figotin A, Kuchment P. 1996 Band-gap structure of spectra of periodic dielectric and acoustic media. 2. Two-dimensional photonic crystals. SIAM Journal on Applied Mathematics 56, 1561–1620.
  • [15] Ammari H, Kang H, Lee H. 2009 Asymptotic Analysis of High-Contrast Phononic Crystals and a Criterion for the Band-Gap Opening. Archive for Rational Mechanics and Analysis 193, 679–714.
  • [16] Lipton R, Viator, Jr. R. 2017 Creating band gaps in periodic media. Multiscale Modeling and Simulation 15, 1612–1650.
  • [17] Lipton R, Viator, Jr. R, Bolanos SJ, Adili A. 2022 Bloch waves in high contrast electromagnetic crystals. ESAIM-Mathematical Modelling and Numerical Analysis 56, 1483–1519.
  • [18] Godin YA, Vainberg B. 2022 Clusters of Bloch waves in three-dimensional periodic media. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 478, 20220519.
  • [19] Godin YA, Vainberg B. 2023 Propagation and dispersion of Bloch waves in periodic media with soft inclusions. eprint arXiv: 2207.08296.
  • [20] Kafesaki M, Penciu RS, Economou EN. 2000 Air Bubbles in Water: A Strongly Multiple Scattering Medium for Acoustic Waves. Phys. Rev. Lett. 84, 6050–6053.
  • [21] Kushwaha M, Djafari-Rouhani B, Dobrzynski L, Vasseur J. 1998 Sonic stop-bands for cubic arrays of rigid inclusions in air. European Physical Journal B 3, 155–161.