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

Bifurcations of thresholds in essential spectra of elliptic operators under localized non-Hermitian perturbations

D. I. Borisov1, D. A. Zezyulin2111Corresponding author, M. Znojil3
Abstract

We consider the operator

=2xd2onω×\mathcal{H}=\mathcal{H}^{\prime}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}\quad\text{on}\quad\omega\times\mathds{R}

subject to the Dirichlet or Robin condition, where a domain ωd1\omega\subseteq\mathds{R}^{d-1} is bounded or unbounded. The symbol \mathcal{H}^{\prime} stands for a second order self-adjoint differential operator on ω\omega such that the spectrum of the operator \mathcal{H}^{\prime} contains several discrete eigenvalues Λj\Lambda_{j}, j=1,,mj=1,\ldots,m. These eigenvalues are thresholds in the essential spectrum of the operator \mathcal{H}. We study how these thresholds bifurcate once we add a small localized perturbation ε(ε)\varepsilon\mathcal{L}(\varepsilon) to the operator \mathcal{H}, where ε\varepsilon is a small positive parameter and (ε)\mathcal{L}(\varepsilon) is an abstract, not necessarily symmetric operator. We show that these thresholds bifurcate into eigenvalues and resonances of the operator \mathcal{H} in the vicinity of Λj\Lambda_{j} for sufficiently small ε\varepsilon. We prove effective simple conditions determining the existence of these resonances and eigenvalues and find the leading terms of their asymptotic expansions. Our analysis applies to generic non-self-adjoint perturbations and, in particular, to perturbations characterized by the parity-time (𝒫𝒯\mathcal{PT}) symmetry. Potential applications of our result embrace a broad class of physical systems governed by dispersive or diffractive effects. As a case example, we employ our findings to develop a scheme for a controllable generation of non-Hermitian optical states with normalizable power and real part of the complex-valued propagation constant lying in the continuum. The corresponding eigenfunctions can be interpreted as an optical generalization of bound states embedded in the continuum. For a particular example, the persistence of asymptotic expansions is confirmed with direct numerical evaluation of the perturbed spectrum.

1 Institute of Mathematics, Ufa Federal Research Center, Russian Academy of Sciences, Ufa, Russia,
&
Bashkir State University, Ufa, Russia,
&
University of Hradec Králové, Hradec Králové, Czech Republic
borisovdi@yandex.ru

2 ITMO University, Saint-Petersburg 197101, Russia
d.zezyulin@gmail.com

3 The Czech Academy of Sciences, Nuclear Physics Institute, Hlavní 130, 25068 Řež, Czech Republic
&
University of Hradec Králové, Hradec Králové, Czech Republic,
&
Institute of System Science, Durban University of Technology, P.O. Box 1334, Durban 4000, South Africa
znojil@ujf.cas.cz

1 Introduction

Physical context and motivation.

Physics of non-Hermitian Hamiltonians is attracting steadily growing attention both on the fundamental level in the development of complex formulations of quantum mechanics [2], [48], [47] and in several applied and experimental fields, such as optics and photonics, Bose-Einstein condensates of atoms or exciton-polaritons, acoustics, and in other areas where diffractive or dispersive effects are governed by Schröginder-like elliptic operators, see [70], [40], [24], [43], [22] for recent reviews. Prominent examples of essentially non-Hermitian phenomena that were introduced in mathematical literature long ago but entered various areas of physics much more recently include exceptional points [35] and spectral singularities [53], [61], [64]. In particular, unusual effects associated with exceptional points are being extensively discussed in optics and photonics (e.g. [30], [46], [57]), whereas spectral singularities [53], [61], [64] are now understood to play an important role in wave scattering [49], [50] and are used to implement coherent perfect absorption of electromagnetic [66], sound [45], and matter [52] waves. Another prototypical behavior, which is forbidden in Hermitian physics but is of the utmost importance in non-Hermitian systems, is the transition from the entirely real spectrum of eigenvalues to a partially complex one.

In a real-world system, the non-Hermiticity usually corresponds to the presence of an energy gain or absorption, which creates an effective complex potential for propagating waves [51]. An especially interesting situation, where a judicious balance between amplification and losses results in rich physics, corresponds to the so-called parity-time (𝒫𝒯\mathcal{PT}) symmetric systems famous for their property to robustly preserve reality of all eigenvalues in spite of the absence of Hermiticity [3, 2]. Tuning a control parameter of a 𝒫𝒯\mathcal{PT}-symmetric system, one can realize various qualitative changes in its spectral structure, and these changes are typically associated with rich and intriguing behaviors. The simplest of those behaviors, which was observed in a series of experiments [46, 57], corresponds to the collision of a pair of real discrete eigenvalues in an exceptional point with a subsequent splitting in a complex-conjugate pair. In systems with a continuous spectrum, the situation is further enriched. In particular, the bifurcation of an isolated eigenvalue from the bottom of the essential spectrum can be accompanied by a so-called jamming anomaly, i.e., a non-monotonous dependence of the energy flux through the gain-to-loss interface on the parameter characterizing the strength of the non-Hermiticity [1]. The bifurcations of a complex-conjugate pair of eigenvalues from an internal point in the essential spectrum are even more interesting and have recently been discussed as an unconventional mechanism of 𝒫𝒯\mathcal{PT}-symmetry breaking [67], [28], [41], [42] distinctively different from the better studied 𝒫𝒯\mathcal{PT}-symmetry breaking through an exceptional point. Complex eigenfunctions associated with bifurcated eigenvalues are L2L^{2}-integrable, and the real parts of these eigenvalues belong to the continuous spectrum. This enables a physical interpretation of such eigenfunctions in terms of non-Hermitian generalizations [44], [34] of bound states embedded in the continuum, well-known in quantum mechanics [56], [63], [58], [60], optics, and other fields [32]. It should be noticed at the same time that most of the activity devoted to non-Hermitian optical bound states in the continuum is being carried out for one-dimensional systems. For multi-dimensional geometries, most of the available results are numerical in nature [34]. From the practical point of view, it is also important that eigenfunctions associated with emerging from the essential spectrum eigenvalues are extremely weakly localized in the vicinity of the bifurcation, which hinders their efficient numerical evaluation. Naturally, this problem is even more pronounced in multi-dimensional geometries, where many more computational resources are necessary to approximate the eigenfunctions. Therefore, any analytical information on the properties of such states is highly desirable.

Mathematical context.

The phenomenon that a small localized perturbation of a self-adjoint differential operator can generate discrete eigenvalues from the edges in the essential spectrum is known for about a hundred of years. Its rigorous mathematical study was initiated by classical works by B. Simon, M. Klaus, R. Blankenbecler, M. L. Goldberger [62], [37], [4], [38] and since that time, hundreds of papers on this subject were written. While classical works were devoted to the Schrödinger operator on an axis and plane perturbed by a small localized potentials, in further works the studies were made for plenty of other models, like waveguide-like structures, see, for instance, [23], [21], [54], [15], for periodic operators, see, for instance, [69], for operators with distant perturbations [31] and many others. All these works treated symmetric perturbations of self-adjoint operators, and the perturbed operators were self-adjoint as well.

Non-symmetric perturbations of self-adjoint operators were studied in essentially less details. In [25], [26] there was considered the Laplacian on the axis perturbed by a small abstract localized operator, which was not assumed to be symmetric. The main result was sufficient conditions ensuring the existence and absence of the emerging eigenvalues from the bottom of the essential spectrum and if they exist, the leading terms in the asymptotic expansions for the emerging eigenvalues were found. These results were essentially extended in [8], [9]. Here an unperturbed operator was an arbitrary periodic self-adjoint operator on the line [8] or on the plane [9]. A perturbation was a small abstract operator not necessarily symmetric and localized in a much weaker sense than in [25], [26]. The structure of the spectra of such operators was studied in details. Qualitative properties like stability of the essential spectrum, the countability of the point spectrum, the absence of the residual spectrum, the existence of embedded eigenvalues were addressed. Sufficient conditions ensuring the existence and absence of the eigenvalues emerging from edges of internal gaps in the essential spectrum were established and if they exist, the leading terms in their asymptotic expansions were obtained. Eigenvalues emerging from the bottom of the essential spectrum were also studied in [12], [13], [14] for waveguides with 𝒫𝒯\mathcal{PT}-symmetric Robin-type boundary condition. In [12], a planar waveguide was considered with a locally perturbed coefficient in the 𝒫𝒯\mathcal{PT}-symmetric boundary condition. In [13], [14], similar two- and three-dimensional waveguides were considered and the perturbation was a small width of these waveguides. The main obtained results were sufficient conditions ensuring the existence of the emerging eigenvalues and the leading terms of their asymptotic expansions. We also mention work [27], where the Dirichlet or Neumann Laplacian in a multi-dimensional cylinder was considered and it was perturbed by a small localized non-symmetric perturbations. The eigenvalues bifurcating from the bottom and the internal thresholds in the essential spectrum were studied. There were obtained certain sufficient conditions ensuring the existence of such eigenvalues and the leading terms of their asymptotic expansions were calculated. However, there was a gap in calculations in [27], which made the results of this work true but incomplete. Namely, while working with operators providing meromorphic continuations for the resolvent in the vicinity of internal thresholds, the author of [27] considered only one continuation, while, as we show in the present work, even in our more general setting two continuations exist and complex eigenvalues are the poles just for one of these continuations. This is why the results of [27] described, roughly speaking, only half of emerging eigenvalues.

The bifurcations of the thresholds in the essential spectrum can be also studied for perturbations of non-self-adjoint operators provided the spectral structure of the limiting operator is known in sufficiently great details. As examples, we mention works [20], [65], where an evolutionary nonlinear Schrödinger equation was considered with both linear and nonlinear perturbations. The linearization of this equation on solitary wave solutions gave rise to a spectral problem for a linear non-self-adjoint operator. The essential spectral of such operator consists of two real semi-axes; the main results of [20], [65] provided conditions, under which the end-points of the essential spectrum bifurcated into eigenvalues. If the latter existed, their two-terms asymptotic expansions were calculated.

An important feature of the eigenvalues emergence is that usually the total multiplicity of the emerging eigenvalues does not exceed the multiplicity of edge in the essential spectrum from which they emerge. The multiplicity of the edge is to be treated in the sense of some appropriate generalized eigenfunctions. However, there were found examples, when this commonly believed rule failed. The earliest work on this subject we know is paper [29], where the Schrödinger operator on 3\mathds{R}^{3} perturbed by a small localized potential was considered. It was found that in certain cases, an nn-multiple bottom of the essential spectrum can generate nn eigenvalues and nn anti-bound states or 2n2n resonances. In [16], a similar phenomenon was found for the Dirichlet Laplacian in a pair of three-dimensional layers coupled by a window, when the perturbation was a small variation of the window shape. Very recently we succeeded to find an even more impressive example of infinitely many eigenvalues and/or resonances emerging from the bottom of an essential spectrum. This was done in papers [17], [18], where we considered a one-dimensional Schrödinger operator on the axis with two complex localized potentials, the supports of which were separated by a large distance. It was found that as this distance increases, more and more resonances and eigenvalues appear in the vicinity of the bottom of the essential spectrum, while the multiplicity of this bottom is at most one. The location and asymptotic behaviour of these emerging eigenvalues and resonances were analyzed in details.

Emerging eigenvalues were also studied not only in the context of classical eigenvalue problems, but also for more complicated operator pencils. In [11], there was considered a special quadratic operator pencil on the line with a special small periodic 𝒫𝒯\mathcal{PT}-symmetric perturbation. The structure of the gaps in the essential spectrum and complex eigenvalues in the vicinities of the edges of these gaps were analyzed in great details. In [10], there was considered a similar quadratic operator pencil with a special small localized 𝒫𝒯\mathcal{PT}-symmetric potential and there were studied eigenvalues emerging from thresholds in the essential spectrum. Sufficient existence conditions were established and the leading terms of the asymptotic expansions of the emerging eigenvalues were obtained.

Model and results.

In the present paper we carry out a rigorous analysis of bifurcations of isolated eigenvalues and resonances from the essential spectrum of a multi-dimensional operator under a small localized general abstract perturbation. Namely, we consider a self-adjoint operator of the form

=2xd2onω×\mathcal{H}=\mathcal{H}^{\prime}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}\quad\text{on}\quad\omega\times\mathds{R}

subject to the Dirichlet or Robin condition, where ωd1\omega\subseteq\mathds{R}^{d-1} is some domain, which can be both bounded and unbounded and \mathcal{H}^{\prime} is a self-adjoint second order differential operator on ω\omega subject to the same boundary condition as \mathcal{H}. We assume that the spectrum of the operator \mathcal{H}^{\prime} contains several discrete eigenvalues Λ1Λ2Λm\Lambda_{1}\leqslant\Lambda_{2}\leqslant\ldots\leqslant\Lambda_{m} below the essential spectrum. Then the essential spectrum of the operator \mathcal{H} is the half-line [Λ1,+)[\Lambda_{1},+\infty) and the mentioned eigenvalues become thresholds in this essential spectrum. We add a small localized perturbation to the operator \mathcal{H}. This perturbation reads as ε(ε)\varepsilon\mathcal{L}(\varepsilon), where ε\varepsilon is a small positive parameter and (ε)\mathcal{L}(\varepsilon) is an abstract not necessarily symmetric operator acting from a weighted Sobolev space W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) into a weighted Lebesgue space L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx). Exact definitions of these spaces will be given in the next section and now we just say that these weights and the operator (ε)\mathcal{L}(\varepsilon) are designed so that this operator maps exponentially growing functions into exponentially decaying ones. The latter fact is exactly how we understand the localization of this operator in a generalized sense.

The main result of our paper describes how the thresholds Λj\Lambda_{j} in the essential spectrum of the operator \mathcal{H} bifurcate under the presence of the perturbation ε(ε)\varepsilon\mathcal{L}(\varepsilon). We show that if the bottom of the essential spectrum Λ1\Lambda_{1} is an mm-multiple eigenvalue of the operator \mathcal{H}^{\prime}, then there can be at most mm eigenvalues and resonances of the operator \mathcal{H} in the vicinity of Λ1\Lambda_{1} for sufficiently small ε\varepsilon. The vicinity of an internal threshold Λj>Λ1\Lambda_{j}>\Lambda_{1} in the essential spectrum being an mm-multiple eigenvalue of the operator \mathcal{H}^{\prime} can contain at most 2m2m eigenvalues and resonances of the operator \mathcal{H}. The eigenvalues and resonances are identified via an analysis of the poles of an appropriate meromorphic continuation of the resolvent in the vicinity of each threshold Λj\Lambda_{j}, j1j\geqslant 1. Each such pole generates either an eigenvalue or a resonance, and we provide simple sufficient conditions allowing one to identify whether a considered pole is an eigenvalue or a resonance. We also construct two-terms asymptotic expansions for the emerging eigenvalues and resonances.

Applications to specific models.

While our result is rather general and applies to a broad range of physical models, where elliptic operators play the prominent role, we will exemplify applications of the work using some particular physical models. Namely, we discuss various examples of the unperturbed operator \mathcal{H} and of the perturbation (ε)\mathcal{L}(\varepsilon). Then we consider models of two- and three-dimensional waveguides and models of two- and three-dimensional quantum oscillators. As a perturbation, we choose a small localized complex potential. Such choice is motivated by physical models of optical waveguides filled with a homogeneous medium, when the refractive index of the waveguide is locally modulated. This creates a small, generally, non-Hermitian perturbation in the form of an effective complex-valued optical potential. In particular, this potential can be 𝒫𝒯\mathcal{PT}-symmetric. Another physical model motivating the above examples is a two-dimensional Bose-Einstein condensate trapped in a harmonic potential in one dimension and without any trapping in the second dimension. The nonlinear interactions between particles of the condensate are assumed to be negligible such that its evolution can be described by the linear Schrödinger operator. As a perturbation, a localized non-Hermitian defect such as a localized dissipation serves. The similar approach can be applied to a three-dimensional condensate, where a localized perturbation can trigger formation of fully localized structures with internal vorticity.

For such examples we show that given an internal threshold of a multiplicity nn, by tuning appropriately the perturbing potential, we can make the threshold to bifurcate into nn pairs of complex-conjugated eigenvalues. This example demonstrates that the total multiplicity of the emerging eigenvalues can exceed the multiplicity of the internal threshold.

Organization of the paper.

The rest of this paper is organized as follows. In Section 2 we elaborate rigorous mathematical formulation of the problem, and then present and discuss the main results which are formulated in several theorems. Section 3 is dedicated to examples, including a case study of optical bound states in the continuum emerging under a small 𝒫𝒯\mathcal{PT}-symmetric perturbation. Sections 4 and 5 contain the proofs of theorems.

2 Problem and results

2.1 Problem

Let x=(x1,,xd1)x^{\prime}=(x_{1},\ldots,x_{d-1}), x=(x,xd)x=(x^{\prime},x_{d}) be Cartesian coordinates in d1\mathds{R}^{d-1} and d\mathds{R}^{d}, respectively, where d2d\geqslant 2, and ωd1\omega\subseteq\mathds{R}^{d-1} be an arbitrary domain. The domain ω\omega can be bounded or unbounded, the case ω=d1\omega=\mathds{R}^{d-1} is also possible. If the boundary of the domain ω\omega is non-empty, we assume that ωC2\partial\omega\in C^{2}. We let Ω:=ω×\Omega:=\omega\times\mathds{R} and we suppose that the domain ω\omega is such that

uL2(ω)CuW21(ω)\|u\|_{L_{2}(\partial\omega)}\leqslant C\|u\|_{W_{2}^{1}(\omega)}

for all uW21(ω)u\in W_{2}^{1}(\omega) with a constant CC independent of uu. This inequality implies that

uL2(Ω)CuW21(Ω)\|u\|_{L_{2}(\partial\Omega)}\leqslant C\|u\|_{W_{2}^{1}(\Omega)}

for all uW21(Ω)u\in W_{2}^{1}(\Omega) with a constant CC independent of uu. This means that on the boundary of the domain Ω\Omega, the traces of the functions in W21(Ω)W_{2}^{1}(\Omega) are well-defined and the trace operator is bounded. This fact is employed below in definitions of various operators and sesquilinear forms without explicit mentioning.

By Aij=Aij(x)A_{ij}=A_{ij}(x^{\prime}), Aj=Aj(x)A_{j}=A_{j}(x^{\prime}), A0=A0(x)A_{0}=A_{0}(x^{\prime}), i,j=1,,d1i,j=1,\ldots,d-1, we denote real functions defined on ω¯\overline{\omega} and with the following smoothness: Aij,AjC1(ω¯)A_{ij},A_{j}\in C^{1}(\overline{\omega}), A0C(ω¯)A_{0}\in C(\overline{\omega}). The functions AijA_{ij} satisfy the usual uniform ellipticity condition, that is, Aij=AjiA_{ij}=A_{ji} and

i,j=1d1Aij(x)ξiξj¯c0i=1d1|ξi|2for allxω¯,ξi,\sum\limits_{i,j=1}^{d-1}A_{ij}{(x^{\prime})}\xi_{i}\overline{\xi_{j}}\geqslant c_{0}\sum\limits_{i=1}^{d-1}|\xi_{i}|^{2}\quad\text{for all}\quad x^{\prime}\in\overline{\omega},\quad\xi_{i}\in\mathds{C},

where c0>0c_{0}>0 is a positive constant independent of xx^{\prime} and ξi\xi_{i}. The functions AijA_{ij} and AjA_{j} are assumed to be uniformly bounded on ω¯\overline{\omega}, while for A0A_{0} only an uniform lower bound is supposed. By i\mathrm{i} we denote the imaginary unit.

In terms of the introduced functions we define an operator

=i,j=1d1xiAijxj2xd2+ij=1d1(Ajxj+xjAj)+A0inΩ\mathcal{H}=-\sum\limits_{i,j=1}^{d-1}\frac{\partial\ }{\partial x_{i}}A_{ij}\frac{\partial\ }{\partial x_{j}}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial\ }{\partial x_{j}}+\frac{\partial\ }{\partial x_{j}}A_{j}\right)+A_{0}\quad\text{in}\quad\Omega (2.1)

subject to the Dirichlet condition or Robin condition:

u=0onΩoru𝝂au=0onΩ.u=0\quad\text{on}\quad\partial\Omega\qquad\text{or}\qquad\frac{\partial u}{\partial\boldsymbol{\nu}}-au=0\quad\text{on}\quad\partial\Omega. (2.2)

In the case of Robin condition, the conormal derivative is defined as

u𝝂:=i,j=1d1Aijνiuxjij=1d1Ajνju+νduxd,\frac{\partial u}{\partial\boldsymbol{\nu}}:=\sum\limits_{i,j=1}^{d-1}A_{ij}\nu_{i}\frac{\partial u}{\partial x_{j}}-\mathrm{i}\sum\limits_{j=1}^{d-1}A_{j}\nu_{j}{u}+\nu_{d}\frac{\partial u}{\partial x_{d}},

where ν=(ν1,,νd)\nu=(\nu_{1},\ldots,\nu_{d}) is the unit outward normal to Ω\partial\Omega and a=a(x)a=a(x^{\prime}) is a real function defined on Ω\partial\Omega. We assume that aC(Ω)a\in C(\partial\Omega) and that this function is uniformly bounded on Ω\partial\Omega. We define a chosen boundary operator in (2.2) by \mathcal{B}, that is, u=u\mathcal{B}u=u or u=u𝝂au\mathcal{B}u=\frac{\partial u}{\partial\boldsymbol{\nu}}-au.

Rigorously we introduce the operator \mathcal{H} as follows. In the space L2(Ω)L_{2}(\Omega) we define a sesquilinear form

𝔥(u,v):=\displaystyle\mathfrak{h}(u,v):= i,j=1d1(Aijuxj,vxi)L2(Ω)+(uxd,vxd)L2(Ω)+ij=1d1(Ajuxj,v)L2(Ω)\displaystyle\sum\limits_{i,j=1}^{d-1}\left(A_{ij}\frac{\partial u}{\partial x_{j}},\frac{\partial v}{\partial x_{i}}\right)_{L_{2}(\Omega)}+\left(\frac{\partial u}{\partial x_{d}},\frac{\partial v}{\partial x_{d}}\right)_{L_{2}(\Omega)}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial u}{\partial x_{j}},v\right)_{L_{2}(\Omega)}
ij=1d1(u,Ajvxj)L2(Ω)+(A0u,v)L2(Ω)\displaystyle-\mathrm{i}\sum\limits_{j=1}^{d-1}\left(u,A_{j}\frac{\partial v}{\partial x_{j}}\right)_{L_{2}(\Omega)}+(A_{0}u,v)_{L_{2}(\Omega)}

on the domain 𝔇(𝔥):=W̊21(Ω)L2(Ω,(1+|A0|)dx)\mathfrak{D}(\mathfrak{h}):=\mathring{W}_{2}^{1}(\Omega)\cap L_{2}(\Omega,(1+|A_{0}|)dx) if the Dirichlet condition is chosen in (2.2), and

𝔥(u,v):=\displaystyle\mathfrak{h}(u,v):= i,j=1d1(Aijuxj,vxi)L2(Ω)+(uxd,vxd)L2(Ω)+ij=1d1(Ajuxj,v)L2(Ω)\displaystyle\sum\limits_{i,j=1}^{d-1}\left(A_{ij}\frac{\partial u}{\partial x_{j}},\frac{\partial v}{\partial x_{i}}\right)_{L_{2}(\Omega)}+\left(\frac{\partial u}{\partial x_{d}},\frac{\partial v}{\partial x_{d}}\right)_{L_{2}(\Omega)}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial u}{\partial x_{j}},v\right)_{L_{2}(\Omega)}
ij=1d1(u,Ajvxj)L2(Ω)+(A0u,v)L2(Ω)(au,v)L2(Ω)\displaystyle-\mathrm{i}\sum\limits_{j=1}^{d-1}\left(u,A_{j}\frac{\partial v}{\partial x_{j}}\right)_{L_{2}(\Omega)}+(A_{0}u,v)_{L_{2}(\Omega)}-(au,v)_{L_{2}(\partial\Omega)}

on the domain 𝔇(𝔥):=W21(Ω)L2(Ω,(1+|A0|)dx)\mathfrak{D}(\mathfrak{h}):=W_{2}^{1}(\Omega)\cap L_{2}(\Omega,(1+|A_{0}|)dx). Here W̊21(Ω)\mathring{W}_{2}^{1}(\Omega) is a subspace of the space W21(Ω)W_{2}^{1}(\Omega) consisting of the functions with a zero trace on Ω\partial\Omega. Given a positive function ϕ\phi on Ω\Omega, by L2(Ω,ϕdx)L_{2}(\Omega,\phi dx) we denote a weighted space formed by the functions in L2,loc(Ω)L_{2,loc}(\Omega) with a finite norm L2(Ω,ϕdx)\|\cdot\|_{L_{2}(\Omega,\phi dx)} defined as

uL2(Ω,ϕdx)2=Ω|u(x)|2ϕ𝑑x.\|u\|_{L_{2}(\Omega,\phi\,dx)}^{2}=\int\limits_{\Omega}|u(x)|^{2}\phi\,dx.

Thanks to the above assumptions on the functions AijA_{ij}, AjA_{j}, A0A_{0} and aa, the form 𝔥\mathfrak{h} is closed, symmetric and lower-semibounded. The self-adjoint operator in L2(Ω)L_{2}(\Omega) associated with this form is exactly the operator \mathcal{H}.

We introduce one more weighted space W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) as a subspace of W2,loc2(Ω)W_{2,loc}^{2}(\Omega) formed by the functions with finite norms W22(Ω,eϑ|xd|dx)\|\cdot\|_{W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx)}, is defined as follows:

uW22(Ω,eϑ|xd|dx)2=Ωα+2|α|2|αu(x)|2eϑ|xd|dx.\|u\|_{W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx)}^{2}=\int\limits_{\Omega}\sum\limits_{\begin{subarray}{c}\alpha\in\mathds{Z}_{+}^{2}\\ |\alpha|\leqslant 2\end{subarray}}|\partial^{\alpha}u(x)|^{2}e^{-{\vartheta}|x_{d}|}dx.

Here ϑ>0\vartheta>0 is some fixed constant. By ε\varepsilon we denote a small positive parameter and the symbols 1\mathcal{L}_{1}, 2\mathcal{L}_{2}, 3=3(ε)\mathcal{L}_{3}=\mathcal{L}_{3}(\varepsilon) stand for operators mapping the space W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) into L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx). These operators are assumed to be bounded; the operator 3\mathcal{L}_{3} is bounded uniformly in ε\varepsilon. We stress that the operators 1\mathcal{L}_{1}, 2\mathcal{L}_{2}, 3\mathcal{L}_{3} are not supposed to be necessarily symmetric.

The main object of our study is a perturbed operator

ε=+ε(ε),(ε):=1+ε2+ε23(ε)\mathcal{H}_{\varepsilon}=\mathcal{H}+\varepsilon\mathcal{L}(\varepsilon),\qquad\mathcal{L}(\varepsilon):=\mathcal{L}_{1}+\varepsilon\mathcal{L}_{2}+\varepsilon^{2}\mathcal{L}_{3}(\varepsilon)

on 𝔇()\mathfrak{D}(\mathcal{H}). The operator ε\mathcal{H}_{\varepsilon} is well-defined since

𝔇()W22(Ω)W22(Ω,eϑ|xd|dx).\mathfrak{D}(\mathcal{H})\subseteq W_{2}^{2}(\Omega)\subseteq W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx).

Moreover, it is clear that the operator (ε)\mathcal{L}(\varepsilon) is relatively bounded with respect to the operator \mathcal{H} and this is why, for sufficiently small ε\varepsilon, the operator ε\mathcal{H}_{\varepsilon} is closed.

Our main aim is to study the behaviour of the eigenvalues of the operator ε\mathcal{H}_{\varepsilon} emerging from certain internal points in its essential spectrum. We denote the latter by σess()\operatorname{\sigma_{ess}}(\cdot) and define it in terms of a characteristic sequences. Namely, a point λ\lambda belongs to an essential spectrum σess(𝒜)\operatorname{\sigma_{ess}}(\mathcal{A}) of some operator 𝒜\mathcal{A} if there exists a bounded noncompact sequence ud𝔇(𝒜)u_{d}\in\mathfrak{D}(\mathcal{A}) such that

infdud>0and(𝒜λ)ud0,d.\inf\limits_{d}\|u_{d}\|>0\qquad\text{and}\qquad(\mathcal{A}-\lambda)u_{d}\to 0,\quad d\to\infty.

In order to describe the essential spectrum σess(ε)\operatorname{\sigma_{ess}}(\mathcal{H}_{\varepsilon}), we introduce two auxiliary operators \mathcal{H}^{\prime} and 0\mathcal{H}_{0}. The former is a self-adjoint operator in L2(d1)L_{2}(\mathds{R}^{d-1}) associated with a lower-semibounded symmetric sesquilinear form

𝔥(u,v):=\displaystyle\mathfrak{h}^{\prime}(u,v):= i,j=1d1(Aijuxj,vxi)L2(ω)+ij=1d1(Ajuxj,v)L2(ω)\displaystyle\sum\limits_{i,j=1}^{d-1}\left(A_{ij}\frac{\partial u}{\partial x_{j}},\frac{\partial v}{\partial x_{i}}\right)_{L_{2}(\omega)}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial u}{\partial x_{j}},v\right)_{L_{2}(\omega)}
ij=1d1(u,Ajvxj)L2(ω)+(A0u,v)L2(ω)\displaystyle-\mathrm{i}\sum\limits_{j=1}^{d-1}\left(u,A_{j}\frac{\partial v}{\partial x_{j}}\right)_{L_{2}(\omega)}+(A_{0}u,v)_{L_{2}(\omega)}

on the domain 𝔇(𝔥):=W̊21(ω)L2(ω,(1+|A0|)dx)\mathfrak{D}(\mathfrak{h}^{\prime}):=\mathring{W}_{2}^{1}(\omega)\cap L_{2}(\omega,(1+|A_{0}|)dx^{\prime}) if the Dirichlet condition is chosen in (2.2) and

𝔥(u,v):=\displaystyle\mathfrak{h}^{\prime}(u,v):= i,j=1d1(Aijuxj,vxi)L2(ω)+ij=1d1(Ajuxj,v)L2(ω)\displaystyle\sum\limits_{i,j=1}^{d-1}\left(A_{ij}\frac{\partial u}{\partial x_{j}},\frac{\partial v}{\partial x_{i}}\right)_{L_{2}(\omega)}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial u}{\partial x_{j}},v\right)_{L_{2}(\omega)}
ij=1d1(u,Ajvxj)L2(ω)+(A0u,v)L2(ω)(au,v)L2(ω)\displaystyle-\mathrm{i}\sum\limits_{j=1}^{d-1}\left(u,A_{j}\frac{\partial v}{\partial x_{j}}\right)_{L_{2}(\omega)}+(A_{0}u,v)_{L_{2}(\omega)}-(au,v)_{L_{2}(\partial\omega)}

on the domain 𝔇(𝔥):=W21(ω)L2(ω,(1+|A0|)dx)\mathfrak{D}(\mathfrak{h}^{\prime}):=W_{2}^{1}(\omega)\cap L_{2}(\omega,(1+|A_{0}|)dx^{\prime}) if the Robin condition is chosen in (2.2). This is the operator

=i,j=1d1xiAijxj+ij=1d1(Ajxj+xjAj)+A0inω\mathcal{H}^{\prime}=-\sum\limits_{i,j=1}^{d-1}\frac{\partial\ }{\partial x_{i}}A_{ij}\frac{\partial\ }{\partial x_{j}}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial\ }{\partial x_{j}}+\frac{\partial\ }{\partial x_{j}}A_{j}\right)+A_{0}\quad\text{in}\quad\omega

subject to the Dirichlet condition or Robin condition:

u=0onωoru𝝂=auonΩ,u𝝂:=i,j=1d1Aijνiuxjij=1d1Ajνj.u=0\quad\text{on}\quad\partial\omega\qquad\text{or}\qquad\frac{\partial u}{\partial\boldsymbol{\nu}^{\prime}}=au\quad\text{on}\quad\partial\Omega,\qquad\frac{\partial u}{\partial\boldsymbol{\nu}^{\prime}}:=\sum\limits_{i,j=1}^{d-1}A_{ij}\nu_{i}\frac{\partial u}{\partial x_{j}}-\mathrm{i}\sum\limits_{j=1}^{d-1}A_{j}\nu_{j}.

The operator 0\mathcal{H}_{0} is a one-dimensional Schrödinger operator

0:=d2dxd2\mathcal{H}_{0}:=-\frac{d^{2}\ }{dx_{d}^{2}}

in L2()L_{2}(\mathds{R}) on the domain W22()W_{2}^{2}(\mathds{R}). Its spectrum is pure essential and coincides with [0,+)[0,+\infty). We assume that there exists a constant c0c_{0} such that the spectrum of the operator \mathcal{H}^{\prime} below this constant consists of finitely many discrete eigenvalues, which we denote by Λj\Lambda_{j} and we arrange them in an ascending order counting multiplicities:

Λ1Λ2Λm<c0.\Lambda_{1}\leqslant\Lambda_{2}\leqslant\ldots\leqslant\Lambda_{m}<c_{0}.

The associated orthonormalized in L2(d1)L_{2}(\mathds{R}^{d-1}) eigenfunctions are denoted by ψj=ψj(x)\psi_{j}=\psi_{j}(x^{\prime}), j=1,,mj=1,\ldots,m.

The essential spectrum of the operator ε\mathcal{H}_{\varepsilon} is described in the following lemma.

Lemma 2.1.

The essential spectrum of the operator ε\mathcal{H}_{\varepsilon} coincides with that of the operator \mathcal{H} for all sufficiently small ε\varepsilon and is given by the identity:

σess(ε)=σess()=σ()=[Λ1,+).\operatorname{\sigma_{ess}}(\mathcal{H}_{\varepsilon})=\operatorname{\sigma_{ess}}(\mathcal{H})=\operatorname{\sigma}(\mathcal{H})=[\Lambda_{1},+\infty).

where σ()\operatorname{\sigma}(\cdot) denotes a spectrum of an operator.

According to this lemma, the points Λj\Lambda_{j}, j=1,,mj=1,\ldots,m, belong to the essential spectrum of the operator ε\mathcal{H}_{\varepsilon}. The point Λ1\Lambda_{1} is the bottom of such spectrum, while other points Λj\Lambda_{j} are internal thresholds.

2.2 Main results

Our results describe a meromorphic continuation of the resolvent of the operator ε\mathcal{H}_{\varepsilon} in the vicinity of the points Λj\Lambda_{j} as well as eigenvalues and resonances emerging from these points due to the presence of the perturbation ε(ε)\varepsilon\mathcal{L}(\varepsilon). Before presenting our main results, we introduce some auxiliary constants and notations.

By BδB_{\delta} we denote a ball of radius δ\delta centered at the origin in the complex plane. We fix p{1,,m}p\in\{1,\ldots,m\} and assume that Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1} is an nn-multiple eigenvalue of the operator \mathcal{H}^{\prime}, where n1n\geqslant 1. Then we consider a new complex parameter kk ranging in a small neighbourhood of the origin and we introduce auxiliary functions:

Kj(k):=iΛpΛjk2asj<p,\displaystyle K_{j}(k):=-\mathrm{i}\sqrt{\Lambda_{p}-\Lambda_{j}-k^{2}}\qquad\text{as}\quad j<p,
Kj(k):=kasj=p,,p+n1,\displaystyle K_{j}(k):=k\hphantom{1\sqrt{\Lambda_{p}-\Lambda_{j}-k^{2}}}\qquad\text{as}\quad j=p,\ldots,p+n-1,
Kj(k):=ΛjΛp+k2asjp+n.\displaystyle K_{j}(k):=\sqrt{\Lambda_{j}-\Lambda_{p}+k^{2}}\hphantom{-\mathrm{i}}\qquad\text{as}\quad j\geqslant p+n.

Hereinafter the branch of the square root is fixed by the condition 1=1\sqrt{1}=1 with the branch cut along the negative real semi-axis. Given R>0R>0, we let Ω±R:=Ω{x:±xd>R}\Omega_{\pm}^{R}:=\Omega\cap\{x:\pm x_{d}>R\}.

Now we are in position to formulate our first main result.

Theorem 2.1.

Fix p{1,,m}p\in\{1,\ldots,m\}, and τ{1,+1}\tau\in\{-1,+1\} and let Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1} be an nn-multiple eigenvalue of the operator \mathcal{H}^{\prime}, where n1n\geqslant 1. For all sufficiently small ε\varepsilon, the resolvent (εΛp+k2)1(\mathcal{H}_{\varepsilon}{-}\Lambda_{p}+k^{2})^{-1} admits a meromorphic continuation with respect to a complex parameter kk ranging in a sufficiently small neighbourhood of the origin. Namely, there exists a bounded operator

ε,τ(k):L2(Ω,eϑ|xd|dx)W22(Ω,eϑ|xd|dx)\mathcal{R}_{\varepsilon,\tau}(k):\,L_{2}(\Omega,e^{{\vartheta}|x_{d}|}\,dx)\to W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}\,dx)

meromorphic with respect to complex kBδk\in B_{\delta} for a sufficiently small fixed δ\delta independent of ε\varepsilon. If p=1p=1, then the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) is independent of the choice of τ\tau and for Rek>0\operatorname{Re}k>0, this operator coincides with the resolvent (εΛ1+k2)1(\mathcal{H}_{\varepsilon}-\Lambda_{1}+k^{2})^{-1} restricted on L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}\,dx). If p>1p>1, then the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) does depend on the choice of τ\tau and for Rek>0\operatorname{Re}k>0 and τImk2<0\tau\operatorname{Im}k^{2}<0, this operator coincides with the resolvent (εΛp+k2)1(\mathcal{H}_{\varepsilon}-\Lambda_{p}+k^{2})^{-1} restricted on L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}\,dx).

For all fL2(Ω,eϑ|xd|dx)f\in L_{2}(\Omega,e^{{\vartheta}|x_{d}|}\,dx), the function uε:=ε,τ(k)fu_{\varepsilon}:=\mathcal{R}_{\varepsilon,\tau}(k)f solves the boundary value problem

(i,j=1d1xiAijxj2xd2+ij=1d1(Ajxj+xjAj)+A0+ε(ε)Λp+k2)uε=finΩ,uε=0onΩ,\begin{gathered}\left(-\sum\limits_{i,j=1}^{d-1}\frac{\partial\ }{\partial x_{i}}A_{ij}\frac{\partial\ }{\partial x_{j}}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial\ }{\partial x_{j}}+\frac{\partial\ }{\partial x_{j}}A_{j}\right)+A_{0}+\varepsilon\mathcal{L}(\varepsilon)-\Lambda_{p}+k^{2}\right)u_{\varepsilon}=f\quad\text{in}\quad\Omega,\\ \mathcal{B}u_{\varepsilon}=0\quad\text{on}\quad\partial\Omega,\end{gathered} (2.3)

and for sufficiently large xdx_{d} it can be represented as follows:

uε(x,k)=j=1muε,j±(xd,k)ψj(x)+uε,±(x,k),±xd>R,u_{\varepsilon}(x,k)=\sum\limits_{j=1}^{m}u_{\varepsilon,j}^{\pm}(x_{d},k)\psi_{j}(x^{\prime})+u_{\varepsilon,\bot}^{\pm}(x,k),\qquad\pm x_{d}>R, (2.4)

where RR is some fixed number, uε,j±L2(I±,eϑxddxd)u_{\varepsilon,j}^{\pm}\in L_{2}(I_{\pm},e^{{\mp\vartheta}x_{d}}dx_{d}) are some meromorphic in kBδk\in B_{\delta} functions, I+:=(R,+)I_{+}:=(R,+\infty), I:=(,R)I_{-}:=(-\infty,-R), possessing the asymptotic behavior

uε,j±(xd)=eτKj(k)|xd|(Cε,j±(k)+O(eϑ~|xd|)),\displaystyle u_{\varepsilon,j}^{\pm}(x_{d})=e^{-\tau K_{j}(k)|x_{d}|}\big{(}C_{\varepsilon,j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)},\quad |xd|,j=1,,pn+1,\displaystyle|x_{d}|\to\infty,\qquad j=1,\ldots,p-n+1, (2.5)
uε,j±(xd)=eKj(k)|xd|(Cε,j±(k)+O(eϑ~|xd|)),\displaystyle u_{\varepsilon,j}^{\pm}(x_{d})=e^{-K_{j}(k)|x_{d}|}\big{(}C_{\varepsilon,j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)}, |xd|,j=p,,m,\displaystyle|x_{d}|\to\infty,\qquad j=p,\ldots,m,

Cε,j±(k)C_{\varepsilon,j}^{\pm}(k) are some meromorphic in kBδk\in B_{\delta} functions, 0<ϑ~<ϑ0<{\tilde{\vartheta}<\vartheta} is some fixed constant independent of kk and xx, and uε,±W22(ΩR±)u_{\varepsilon,{\bot}}^{\pm}\in W_{2}^{2}(\Omega_{R}^{\pm}) are some functions meromorphic in kBδk\in B_{\delta} and obeying the identities

(uε,±(,xd),ψj)L2(ω)=0(u_{\varepsilon,\bot}^{\pm}(\cdot,x_{d}),\psi_{j})_{L_{2}({\omega})}=0 (2.6)

for almost each xdI±x_{d}\in I_{\pm} and for each j=1,,mj=1,\ldots,m.

If kεBδk_{\varepsilon}\in B_{\delta} is a pole of the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k), for k=kεk=k_{\varepsilon}, problem (2.3) with f=0f=0 has a non-trivial solution ψε\psi_{\varepsilon} in W2,loc2(Ω)W_{2,loc}^{2}(\Omega), which satisfies a representation similar to (2.4):

ψε(x)=j=1mϕε,j±(xd,k)ψj(x)+ψε,±(x,k),±xd>R,\psi_{\varepsilon}(x)=\sum\limits_{j=1}^{m}\phi_{\varepsilon,j}^{\pm}(x_{d},k)\psi_{j}(x^{\prime})+\psi_{\varepsilon,\bot}^{\pm}(x,k),\qquad\pm x_{d}>R, (2.7)

where RR is some fixed number, ϕε,j±L2(I±,eϑxddxd)\phi_{\varepsilon,j}^{\pm}\in L_{2}(I_{\pm},e^{{\mp\vartheta}x_{d}}dx_{d}) are functions with the asymptotic behavior

ϕε,j±(xd)=eτKj(k)|xd|(cε,j±(k)+O(eϑ~|xd|)),\displaystyle\phi_{\varepsilon,j}^{\pm}(x_{d})=e^{-\tau K_{j}(k)|x_{d}|}\big{(}c_{\varepsilon,j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)},\quad |xd|,j=1,,pn+1,\displaystyle|x_{d}|\to\infty,\qquad j=1,\ldots,p-n+1, (2.8)
ϕε,j±(xd)=eKj(k)|xd|(cε,j±(k)+O(eϑ~|xd|)),\displaystyle\phi_{\varepsilon,j}^{\pm}(x_{d})=e^{-K_{j}(k)|x_{d}|}\big{(}c_{\varepsilon,j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)}, |xd|,j=p,,m,\displaystyle|x_{d}|\to\infty,\qquad j=p,\ldots,m,

cε,j±(k)c_{\varepsilon,j}^{\pm}(k) are some constants, and ψε±W22(ΩR±)\psi_{\varepsilon}^{\pm}\in W_{2}^{2}(\Omega_{R}^{\pm}) are some functions obeying the identities

(ψε,±(,xd),ψj)L2(ω)=0(\psi_{\varepsilon,\bot}^{\pm}(\cdot,x_{d}),\psi_{j})_{L_{2}({\omega})}=0 (2.9)

for almost each xdI±x_{d}\in I_{\pm} and for each j=1,,mj=1,\ldots,m.

We define a subspace LL^{\bot} in L2(Ω)L_{2}(\Omega) as a set of functions vL2(Ω)v\in L_{2}(\Omega) such that

(v(,xd),ψj)L2(ω)=0(v(\cdot,x_{d}),\psi_{j})_{L_{2}(\omega)}=0

for almost each xdx_{d}\in\mathds{R} and for all j=1,,mj=1,\ldots,m. The space LL^{\bot} is a Hilbert one. By \mathcal{H}^{\bot} we denote the restriction of the operator \mathcal{H} on 𝔇()L\mathfrak{D}(\mathcal{H})\cap L^{\bot}. The following lemma will be proved in Section 4.1.

Lemma 2.2.

The space LL^{\bot} is invariant for the operator \mathcal{H}^{\bot}, that is, this operator maps 𝔇()L\mathfrak{D}(\mathcal{H})\cap L^{\bot} into LL^{\bot}. This is an unbounded self-adjoint operator in LL^{\bot} and its spectrum is located in [c0,+)[c_{0},+\infty).

The above lemma means that the resolvent (Λp)1(\mathcal{H}^{\bot}-\Lambda_{p})^{-1} is well-defined for all p=1,,mp=1,\ldots,m as an operator from L2(L)L_{2}(L^{\bot}) into 𝔇()L\mathfrak{D}(\mathcal{H})\cap L^{\bot}. As above, we fix p{1,,m}p\in\{1,\ldots,m\} and assume that Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1}, where n1n\geqslant 1, and in terms of the latter resolvent, we introduce an auxiliary operator mapping L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) into W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{{-\vartheta}|x_{d}|}dx):

(𝒢p,τf)(x):=j=1p1ψj(x)2τKj(0)ΩeτKj(0)|xdyd|ψj(y)¯f(y)𝑑y12j=pp+n1ψj(x)Ω|xdyd|ψj(y)¯f(y)𝑑y+j=p+nmψj(x)2Kj(0)ΩeKj(0)|xdyd|ψj(y)¯f(y)𝑑y+((Λp)1f)(x),\displaystyle\begin{aligned} (\mathcal{G}_{p,\tau}f)(x):=&\sum\limits_{j=1}^{p-1}\frac{\psi_{j}(x^{\prime})}{2\tau K_{j}(0)}\int\limits_{\Omega}e^{-\tau K_{j}(0)|x_{d}-y_{d}|}\overline{\psi_{j}(y^{\prime})}f(y)\,dy\\ &-\frac{1}{2}\sum\limits_{j=p}^{p+n-1}\psi_{j}(x^{\prime})\int\limits_{\Omega}|x_{d}-y_{d}|\overline{\psi_{j}(y^{\prime})}f(y)\,dy\\ &+\sum\limits_{j=p+n}^{m}\frac{\psi_{j}(x^{\prime})}{2K_{j}(0)}\int\limits_{\Omega}e^{-K_{j}(0)|x_{d}-y_{d}|}\overline{\psi_{j}(y^{\prime})}f(y)\,dy+((\mathcal{H}^{\bot}-\Lambda_{p})^{-1}f^{\bot})(x),\end{aligned} (2.10)
f(x):=f(x)j=1mfj(xd)ψj(x).\displaystyle f^{\bot}(x):=f(x)-\sum\limits_{j=1}^{m}f_{j}(x_{d})\psi_{j}(x^{\prime}). (2.11)

As above, here τ{1,+1}\tau\in\{-1,+1\}. In the case p=1p=1, the first sum in the above definition is missing and the operator 𝒢p,τ\mathcal{G}_{p,\tau} becomes independent of the choice of τ\tau.

We define the matrix M1\mathrm{M}_{1} with entries

M1ij:=12Ωψi+p1¯1ψj+p1𝑑x,i,j=1,,n,M_{1}^{ij}:=-\frac{1}{2}\int\limits_{\Omega}\overline{\psi_{i+p-1}}\mathcal{L}_{1}\psi_{j+p-1}\,dx,\qquad i,j=1,\ldots,n, (2.12)

where ii counts the rows and jj does the columns in the matrix M1\mathrm{M}_{1}; since the operator 1\mathcal{L}_{1} acts from W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-\vartheta|x_{d}|}dx) into L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{\vartheta|x_{d}|}dx), the functions 1ψj+p1\mathcal{L}_{1}\psi_{j+p-1} belong to L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{\vartheta|x_{d}|}dx) and this obviously ensures the convergence of the integrals in the above identity.

By μi\mu_{i}, i=1,,Ni=1,\ldots,N, we denote different eigenvalues of the matrix M1\mathrm{M}_{1} of multiplicities q1,,qNq_{1},\ldots,q_{N}. It is clear that NnN\leqslant n and q1++qN=nq_{1}+\ldots+q_{N}=n.

Theorem 2.2.

Fix p{1,,m}p\in\{1,\ldots,m\}, τ{1,+1}\tau\in\{-1,+1\} and let Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1} be an nn-multiple eigenvalue of the operator \mathcal{H}^{\prime}, where n1n\geqslant 1. There are exactly NN poles, counting their orders, of the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) converging to zero as ε+0\varepsilon\to+0. These poles, denoted by kij(ε)k_{ij}(\varepsilon), have the asymptotic behavior

kij(ε)=εμi+O(ε1+1qi),i=1,,N,j=1,,qi.k_{ij}(\varepsilon)=\varepsilon\mu_{i}+O\big{(}\varepsilon^{1+\frac{1}{q_{i}}}\big{)},\qquad i=1,\ldots,N,\quad j=1,\ldots,q_{i}. (2.13)

Asymptotic expansion (2.13) for the poles kijk_{ij} can be specified in more details and this will be done in terms of one more matrix M2,τ\mathrm{M}_{2,\tau} with entries

M2,τij:=12Ωψi+p1¯(21𝒢p,τ1)ψj+p1𝑑xi,j=1,,n,M_{2,\tau}^{ij}:=\frac{1}{2}\int\limits_{\Omega}\overline{\psi_{i+p-1}}(\mathcal{L}_{2}-\mathcal{L}_{1}\mathcal{G}_{p,\tau}\mathcal{L}_{1})\psi_{j+p-1}\,dx\qquad i,j=1,\ldots,n, (2.14)

where ii counts the rows and jj does the columns in the matrix M2,τ\mathrm{M}_{2,\tau}. Due to the definition of the operators 1\mathcal{L}_{1}, 2\mathcal{L}_{2} , the second term in the integrand in (2.14) belongs to L2(,eϑ|xd|dxd)L_{2}(\mathds{R},e^{\vartheta|x_{d}|}dx_{d}) and this ensures the convergence of the integral. We denote

Qi,τ(z):=εdet(zEM1+εM2,τ)|ε=0.Q_{i,\tau}(z):=\frac{\partial\ }{\partial\varepsilon}\det\big{(}z\mathrm{E}-\mathrm{M}_{1}+\varepsilon\mathrm{M}_{2,\tau}\big{)}\bigg{|}_{\varepsilon=0}. (2.15)

We stress that if Λp=Λ1\Lambda_{p}=\Lambda_{1}, the matrix M2,τM_{2,\tau} and the function Qi,τQ_{i,\tau} become independent of τ\tau.

Theorem 2.3.

Under the assumptions of Theorem 2.2, we fix i{1,,N}i\in\{1,\ldots,N\}. If Qi,τ(z)Q_{i,\tau}(z) vanishes identically, then

kij(ε)=εμi+O(ε1+2qi),i=1,,N,j=1,,qi.k_{ij}(\varepsilon)=\varepsilon\mu_{i}+O\big{(}\varepsilon^{1+\frac{2}{q_{i}}}\big{)},\qquad i=1,\ldots,N,\quad j=1,\ldots,q_{i}. (2.16)

If Qi,τQ_{i,\tau} is not identically zero, then there exists a fixed non-negative integer ri,τ<qir_{i,\tau}<q_{i} such that

γi,τ:=ri,τ!j=1jiN(μiμj)qjdri,τQi,τdzri,τ(μi)0.\gamma_{i,\tau}:=\frac{r_{i,\tau}!}{\prod\limits_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{N}(\mu_{i}-\mu_{j})^{q_{j}}}\frac{d^{r_{i,\tau}}Q_{i,\tau}}{dz^{r_{i,\tau}}}(\mu_{i})\neq 0. (2.17)

If 2ri,τqi2r_{i,\tau}\geqslant q_{i}, then

kij(ε)=εμi+O(ε1+1ri,τ),i=1,,N,j=1,,qi.k_{ij}(\varepsilon)=\varepsilon\mu_{i}+O\big{(}\varepsilon^{1+\frac{1}{r_{i,\tau}}}\big{)},\qquad i=1,\ldots,N,\quad j=1,\ldots,q_{i}. (2.18)

If 2ri,τqi12r_{i,\tau}\leqslant q_{i}-1, then exactly ri,τr_{i,\tau} poles kijk_{ij}, j=1,,ri,τj=1,\ldots,r_{i,\tau} have the asymptotic behavior

kij(ε)=εμi+O(ε1+1ri,τ),i=1,,N,j=1,,ri,τ,k_{ij}(\varepsilon)=\varepsilon\mu_{i}+O\big{(}\varepsilon^{1+\frac{1}{r_{i,\tau}}}\big{)},\qquad i=1,\ldots,N,\quad j=1,\ldots,r_{i,\tau}, (2.19)

while other poles kijk_{ij}, j=ri,τ+1,,qij=r_{i,\tau}+1,\ldots,q_{i}, have the asymptotic behavior

kij(ε)=εμi+ε1+1qiri,τ(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)+O(ε1+2qiri,τ),k_{ij}(\varepsilon)=\varepsilon\mu_{i}+\varepsilon^{1+\frac{1}{q_{i}-r_{i,\tau}}}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}+O\big{(}\varepsilon^{1+\frac{2}{q_{i}-r_{i,\tau}}}\big{)}, (2.20)

where the branch of the fractional power z1qiri,τz^{\frac{1}{q_{i}-r_{i,\tau}}} is fixed by the condition 11qiri,τ=11^{\frac{1}{q_{i}-r_{i,\tau}}}=1 with the branch cut along the negative real semi-axis.

We give some definitions before we formulate our next result. A pole kεBδk_{\varepsilon}\in B_{\delta} of an operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) corresponds to an eigenvalue Λpkε2\Lambda_{p}-k_{\varepsilon}^{2} of an operator ε\mathcal{H}_{\varepsilon} if an associated nontrivial solution ψε\psi_{\varepsilon} to (2.3), (2.7) belongs to W22(Ω)W_{2}^{2}(\Omega). Otherwise it corresponds to a resonance Λpkε2\Lambda_{p}-k_{\varepsilon}^{2}.

Our further results provide conditions allowing to determine whether resonances or eigenvalues are associated with the poles described in two previous theorems. We first present the main result on the poles emerging from the bottom of the essential spectrum.

Theorem 2.4.

Let p=1p=1 and make the assumptions of Theorems 2.22.3. If Reμi>0\operatorname{Re}\mu_{i}>0, then the poles kijk_{ij}, j=1,,qij=1,\ldots,q_{i} correspond to the eigenvalues λij(ε)=Λpkij2(ε)\lambda_{ij}(\varepsilon)=\Lambda_{p}-k_{ij}^{2}(\varepsilon) with the asymptotic behavior

λij(ε)=Λpε2μi2+O(ε2+1αi)\lambda_{ij}(\varepsilon)=\Lambda_{p}-\varepsilon^{2}\mu_{i}^{2}+O\big{(}\varepsilon^{2+\frac{1}{\alpha_{i}}}\big{)} (2.21)

with j=1,,qij=1,\ldots,q_{i}, where

αi:={qi2ifQi,τvanishes identically,ri,τif2ri,τqi.\alpha_{i}:=\left\{\begin{aligned} &\frac{q_{i}}{2}\qquad\hphantom{\tau}\text{if}\quad Q_{i,\tau}\ \text{vanishes identically},\\ &\,r_{i,\tau}\qquad\text{if}\quad 2r_{i,\tau}\geqslant q_{i}.\end{aligned}\right. (2.22)

If 2ri,τqi12r_{i,\tau}\leqslant q_{i}-1, then the eigenvalues λij\lambda_{ij} still have asymptotic behavior (2.21) with αi=ri,τ\alpha_{i}=r_{i,\tau} for j=1,,ri,τj=1,\ldots,r_{i,\tau}, while the asymptotic behaviors for the other eigenvalues read as

λij(ε)=Λpε2μi22ε2+1qiri,τ(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)+O(ε2+2qiri,τ),j=ri,τ+1,,qi.\lambda_{ij}(\varepsilon)=\Lambda_{p}-\varepsilon^{2}\mu_{i}^{2}-2\varepsilon^{2+\frac{1}{q_{i}-r_{i,\tau}}}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}+O(\varepsilon^{2+\frac{2}{q_{i}-r_{i,\tau}}}),\qquad j=r_{i,\tau}+1,\ldots,q_{i}. (2.23)

If Reμi<0\operatorname{Re}\mu_{i}<0, then the poles kijk_{ij}, j=1,,qij=1,\ldots,q_{i} correspond to the resonances λij(ε)=Λpkij2(ε)\lambda_{ij}(\varepsilon)=\Lambda_{p}-k_{ij}^{2}(\varepsilon) with asymptotic expansions (2.21), (2.22), (2.23).

Let Reμi=0\operatorname{Re}\mu_{i}=0, Qi,τQ_{i,\tau} be not identically zero and 2ri,τqi12r_{i,\tau}\leqslant q_{i}-1. As j=ri,τ+1,,qij=r_{i,\tau}+1,\ldots,q_{i}, if

Re(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)>0,\operatorname{Re}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}>0, (2.24)

then the pole kijk_{ij} corresponds to an eigenvalue, while if

Re(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)<0,\operatorname{Re}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}<0, (2.25)

the pole kijk_{ij} corresponds to a resonance. The asymptotic expansion for this eigenvalue/resonance is given by (2.23) if μi0\mu_{i}\neq 0 and in the case μi=0\mu_{i}=0 it reads as

λij(ε)=Λpε2+2qiri,τ(γi,τ)2qiri,τe4πiqiri,τ(jri,τ)+O(ε2+3qiri,τ).\lambda_{ij}(\varepsilon)=\Lambda_{p}-\varepsilon^{2+\frac{2}{q_{i}-r_{i,\tau}}}(-\gamma_{i,\tau})^{\frac{2}{q_{i}-r_{i,\tau}}}e^{\frac{4\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}+O\big{(}\varepsilon^{2+\frac{3}{q_{i}-r_{i,\tau}}}\big{)}. (2.26)

The next theorem concerns the poles emerging from the internal thresholds in the essential spectrum. Given p>1p>1 such that Λp>Λ1\Lambda_{p}>\Lambda_{1}, and τ{1,+1}\tau\in\{-1,+1\}, by kij,τ=kij,τ(ε)k_{ij,\tau}=k_{ij,\tau}(\varepsilon) we redenote the corresponding poles kijk_{ij} of the operator ε,τ\mathcal{R}_{\varepsilon,\tau} described in Theorems 2.22.3.

Theorem 2.5.

Let Λp>Λ1\Lambda_{p}>\Lambda_{1} and fix i{1,,N}i\in\{1,\ldots,N\}, j{1,,qi}j\in\{1,\ldots,q_{i}\}, τ{1,+1}\tau\in\{-1,+1\}. Let

Reμi>0\displaystyle\operatorname{Re}\mu_{i}>0 (2.27)
or
Reμi=0,Qi,τ0,2ri,τqi1,j{ri+1,,qi},Re(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)>0\displaystyle\begin{gathered}\operatorname{Re}\mu_{i}=0,\qquad Q_{i,\tau}\not\equiv 0,\qquad 2r_{i,\tau}\leqslant q_{i}-1,\\ j\in\{r_{i}+1,\ldots,q_{i}\},\qquad\operatorname{Re}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}>0\end{gathered} (2.30)

and

τImμi<0\displaystyle\tau\operatorname{Im}\mu_{i}<0 (2.31)
or
Imμi=0,Qi,τ0,2ri,τqi1,j{ri+1,,qi},τIm(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)<0.\displaystyle\begin{gathered}\operatorname{Im}\mu_{i}=0,\qquad Q_{i,\tau}\not\equiv 0,\qquad 2r_{i,\tau}\leqslant q_{i}-1,\\ j\in\{r_{i}+1,\ldots,q_{i}\},\qquad\tau\operatorname{Im}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}<0.\end{gathered} (2.34)

Then the pole kij,τ(ε)k_{ij,\tau}(\varepsilon) corresponds to an eigenvalue λij,τ(ε)=Λpkij,τ2(ε)\lambda_{ij,\tau}(\varepsilon)=\Lambda_{p}-k_{ij,\tau}^{2}(\varepsilon) with asymptotic expansions (2.21), (2.22), (2.23) if μi0\mu_{i}\neq 0 and asymptotic expansion (2.26) if μi=0\mu_{i}=0.

Let

Reμi<0\displaystyle\operatorname{Re}\mu_{i}<0 (2.35)
or
Reμi=0,Qi,τ0,2ri,τqi1,j{ri+1,,qi},Re(γi,τ)1qiri,τe2πiqiri,τ(jri,τ)<0.\displaystyle\begin{gathered}\operatorname{Re}\mu_{i}=0,\qquad Q_{i,\tau}\not\equiv 0,\qquad 2r_{i,\tau}\leqslant q_{i}-1,\\ j\in\{r_{i}+1,\ldots,q_{i}\},\qquad\operatorname{Re}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}<0.\end{gathered} (2.38)

Then the pole kij,τ(ε)k_{ij,\tau}(\varepsilon) corresponds to a resonance λij,τ(ε)=Λpkij,τ2(ε)\lambda_{ij,\tau}(\varepsilon)=\Lambda_{p}-k_{ij,\tau}^{2}(\varepsilon) with asymptotic expansions (2.21), (2.22), (2.23) if μi0\mu_{i}\neq 0 and asymptotic expansion (2.26) if μi=0\mu_{i}=0.

Let μi\mu_{i} be a simple eigenvalue of the matrix M1\mathrm{M}_{1} with an associated eigenvector ei:=(ei,1,,ei,n)\mathrm{e}_{i}:=(\mathrm{e}_{i,1},\ldots,\mathrm{e}_{i,n}), j=1j=1, condition (2.27) or (2.30) hold, and

τImμi<0\displaystyle\tau\operatorname{Im}\mu_{i}<0 (2.39)
or
Imμi=0,Qi,τ0,ri,τ=0,τImγi,τ<0,\displaystyle\begin{gathered}\operatorname{Im}\mu_{i}=0,\qquad Q_{i,\tau}\not\equiv 0,\qquad r_{i,\tau}=0,\qquad\tau\operatorname{Im}\gamma_{i,\tau}<0,\end{gathered} (2.41)

and let there exist s{1,,p1}s\in\{1,\ldots,p-1\} such that

t=1nΩeKt(0)xdψs(x)¯1ei,tψtp+1𝑑x0ort=1nΩeKt(0)xdψs(x)¯1ei,tψtp+1𝑑x0.\sum\limits_{t=1}^{n}\int\limits_{\Omega}e^{-K_{t}(0)x_{d}}\overline{\psi_{s}(x^{\prime})}\mathcal{L}_{1}\mathrm{e}_{i,t}\psi_{t-p+1}\,dx\neq 0\quad\text{or}\quad\sum\limits_{t=1}^{n}\int\limits_{\Omega}e^{K_{t}(0)x_{d}}\overline{\psi_{s}(x^{\prime})}\mathcal{L}_{1}\mathrm{e}_{i,t}\psi_{t-p+1}\,dx\neq 0. (2.42)

Then the pole ki1,τ(ε)k_{i1,\tau}{(\varepsilon)} corresponds to a resonance λi,τ(ε)=Λpki1,τ2(ε)\lambda_{i,\tau}{(\varepsilon)}=\Lambda_{p}-k_{i1,\tau}^{2}(\varepsilon) with the asymptotic behavior

λi,τ(ε)=Λpε2μi2+O(ε4)\lambda_{i,\tau}(\varepsilon)=\Lambda_{p}-\varepsilon^{2}\mu_{i}^{2}+O(\varepsilon^{4})

if Qi,τQ_{i,\tau} vanishes identically, and

λi,τ(ε)=Λpε2(μiεγi,τ)2+O(|μi|ε4+ε5)\lambda_{i,\tau}(\varepsilon)=\Lambda_{p}-\varepsilon^{2}(\mu_{i}-\varepsilon\gamma_{i,\tau})^{2}+O\big{(}|\mu_{i}|\varepsilon^{4}+\varepsilon^{5}\big{)}

otherwise.

2.3 Discussion of the results

In this subsection we discuss the main results formulated in Theorems 2.12.22.32.42.5. The first of them, Theorem 2.1, describes a meromorphic continuation of the resolvent of the perturbed operator. This continuation is local and is constructed in the vicinity of the points Λp\Lambda_{p}, p=1,,mp=1,\ldots,m. The point Λ1\Lambda_{1} is the bottom of the essential spectrum, see Lemma 2.1 and in vicinity of this point just one meromorphic continuation is possible. It is introduced as a solution to problem (2.3) with a specified behaviour at infinity, see (2.4), in terms of an auxiliary spectral parameter kk. The right hand side in the equation in (2.3) is not in the class of compactly supported functions as it is usually assumed for meromorphic continuations, but an element of a wider space L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|\,dx}). Here the presence of the weight eϑ|xd|e^{{\vartheta}|x_{d}|} means that the elements of latter space in certain sense decays exponentially as xd±x_{d}\to\pm\infty, namely they are represented as f=eϑ|xd|2f~f=e^{-\frac{{\vartheta}|x_{d}|}{2}}\tilde{f}, where f~L2(Ω)\tilde{f}\in L_{2}(\Omega). The final operator providing the meromorphic continuation is ε,τ\mathcal{R}_{\varepsilon,\tau} and for Λ1\Lambda_{1} it is independent of τ\tau.

In the vicinity of internal thresholds Λp>Λ1\Lambda_{p}>\Lambda_{1} in the essential spectrum, there are two different meromorphic continuations given by the operators ε,1\mathcal{R}_{\varepsilon,-1} and ε,+1\mathcal{R}_{\varepsilon,+1}. The former describes a meromorphic continuation from the lower complex half-plane into the upper one, while the latter does from the upper half-plane into the lower one. In the theory of self-adjoint operators, usually only the latter continuation from the upper half-plane into the lower one is studied since it is physically meaning and it arises while considering a corresponding Cauchy problem for an evolutionary Schrödinger equation. However, since our perturbing operator is not assumed to be symmetric, the operator is not necessary self-adjoint. As a result, it can possess complex eigenvalues in the vicinity of the threshold Λp\Lambda_{p}. These eigenvalues are the poles of the resolvent of the perturbed operator. And as we shall see below, once we continue meromorphically the resolvent from the upper half-plane into the lower one, the eigenvalues in the lower half-plane can become ‘‘invisible’’ for the continuation in the sense that this continuation has no poles at such eigenvalues. A similar situation can hold once we continue analytically the resolvent from the lower half-plane into the upper one. A clear explanation of this phenomenon is due to representations (2.4), (2.5), (2.6) and (2.7), (2.8), (2.9). Namely, as Λp>Λ1\Lambda_{p}>\Lambda_{1}, the functions ϕε,j±\phi_{\varepsilon,j}^{\pm}, j=1,,p1j=1,\ldots,p-1, in (2.8) behave at infinity as ϕε,j±(xd)eτKj(k)|xd|\phi_{\varepsilon,j}^{\pm}(x_{d})\sim e^{-\tau K_{j}(k)|x_{d}|}. In view of obvious identities

Kj(k)=iΛpΛj+i2ΛpΛjk2+O(k4),k0,K_{j}(k)=-\mathrm{i}\sqrt{\Lambda_{p}-\Lambda_{j}}+\frac{\mathrm{i}}{2\sqrt{\Lambda_{p}-\Lambda_{j}}}k^{2}+O(k^{4}),\qquad k\to 0, (2.43)

the exponents eτKj(k)|xd|e^{-\tau K_{j}(k)|x_{d}|} decay only if τImk2<0\tau\operatorname{Im}k^{2}<0. Depending on τ\tau, the latter condition means that in general only the eigenvalues either in the upper or lower complex half-plane can serve as poles of the meromorphic continuation of the resolvent of the operator ε\mathcal{H}_{\varepsilon}. This is a main reason why we deal with both meromorphic continuations, in contrast to the case of self-adjoint operators with symmetric perturbations.

Theorems 2.22.3 describe the poles of the meromorphic continuations of the resolvent in the vicinity of the thresholds Λp\Lambda_{p} in the essential spectrum. The first theorem states that in the vicinity of an nn-multiple threshold Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1} there exist exactly nn poles of the operator ε,τ\mathcal{R}_{\varepsilon,\tau} counting their orders. We stress that here we count the orders of the poles and not their multiplicities, that is, not the number of associated linear independent solutions to problem (2.3) with f=0f=0, k=kij(ε)k=k_{ij}(\varepsilon). The multiplicity of each pole does not exceed its order; this can be shown by the technique used in the proofs of Lemmata 6.2, 6.3 in [6] and Lemmata 6.2, 6.3 in [7]. However, in general, the multiplicities and the orders coincide only if the perturbation (ε)\mathcal{L}(\varepsilon) is symmetric. The reason is that in the general case of a non-symmetric perturbation, in a certain matrix controlling the structure of the poles kijk_{ij}, a non-diagonal Jordan block can arise and this gives rise to adjoint vectors instead of the eigenvectors, see Section 5 and the calculations involving matrix Mε,τ\mathrm{M}_{\varepsilon,\tau}. Of course, the multiplicity of each pole kij,τk_{ij,\tau} is at least one. In particular, if all poles kij,τk_{ij,\tau} are different for a fixed τ\tau, the total multiplicity is equal to nn. Theorem 2.2 provides leading terms in the asymptotic expansions for the poles kij,τk_{ij,\tau}, while Theorem 2.3 specifies these expansions. In some cases it just improves the estimate for the error terms, see (2.16), (2.18), (2.19), while in some cases, a next-to-leading term in the expansions can be found, see (2.20). Theorems 2.22.3 treat a general case, when the eigenvalues of the matrix M1\mathrm{M}_{1} are of arbitrary multiplicities and no extra assumptions are made for the matrix M2,τ\mathrm{M}_{2,\tau}. In an important particular case, when μi\mu_{i} is a simple eigenvalue of the matrix M1\mathrm{M}_{1}, we have qi=1q_{i}=1 and ri,τ=0r_{i,\tau}=0. In this case there exists just one pole ki1,τk_{i1,\tau} with asymptotic behavior (2.13) and expansion (2.16), (2.20) can be applied, which yields that

ki1,τ(ε)=εμiε2γi,τ+O(ε3).k_{i1,\tau}(\varepsilon)=\varepsilon\mu_{i}-\varepsilon^{2}\gamma_{i,\tau}+O(\varepsilon^{3}).

If n=1n=1, that is, Λp\Lambda_{p} is a simple eigenvalue of the operator \mathcal{H}^{\prime}, the above expansions can be specified as follows:

ki1,τ(ε)=ε2Ωψp¯1ψp𝑑xε22Ωψp¯(21𝒢p,τ1)ψp𝑑x+O(ε3).k_{i1,\tau}(\varepsilon)=-\frac{\varepsilon}{2}\int\limits_{\Omega}\overline{\psi_{p}}\mathcal{L}_{1}\psi_{p}\,dx-\frac{\varepsilon^{2}}{2}\int\limits_{\Omega}\overline{\psi_{p}}(\mathcal{L}_{2}-\mathcal{L}_{1}\mathcal{G}_{p,\tau}\mathcal{L}_{1})\psi_{p}\,dx+O(\varepsilon^{3}). (2.44)

We also observe that since the operator ε,τ\mathcal{R}_{\varepsilon,\tau} is independent of τ\tau if p=1p=1, in the general situation there are only nn poles in the vicinity of the bottom Λ1\Lambda_{1} of the essential spectrum. In the vicinity of internal thresholds Λp>Λ1\Lambda_{p}>\Lambda_{1}, the operators ε,τ\mathcal{R}_{\varepsilon,\tau} depend on τ\tau and this is why there are 2n2n poles in the vicinity of Λp\Lambda_{p}. In particular, if Λp\Lambda_{p} is an nn-multiple eigenvalue of the operator \mathcal{H}^{\prime}, there can be 2n2n different simple eigenvalues of the operator ε\mathcal{H}_{\varepsilon} converging to Λp\Lambda_{p}, see examples in Subsection 3.3.

The above discussed poles of the operators ε,τ\mathcal{R}_{\varepsilon,\tau} correspond either to the eigenvalues or resonances depending on the behavior of the associated non-trivial solutions. This behaviour is completely described by formulae (2.7), (2.8) and we just need to identify whether the function ψε\psi_{\varepsilon} decays exponentially at infinity or not. In the former case we deal with an eigenvalue, otherwise with a resonance. As we see, the functions ϕε,j±(xd)\phi_{\varepsilon,j}^{\pm}(x_{d}) decay exponentially as jp+nj\geqslant p+n no matter how the corresponding pole looks like. But for j=p,,p+n1j=p,\ldots,p+n-1 these functions behave at infinity as ϕε,j±(xd,k)ekε|xd|\phi_{\varepsilon,j}^{\pm}(x_{d},k)\sim e^{-k_{\varepsilon}|x_{d}|}. These functions decay exponentially as Rekε>0\operatorname{Re}k_{\varepsilon}>0, is periodic as Rekε=0\operatorname{Re}k_{\varepsilon}=0 and grows exponentially as Rekε<0\operatorname{Re}k_{\varepsilon}<0. If Λp>Λ1\Lambda_{p}>\Lambda_{1}, we also have to control the behavior of the functions ϕε,j±(xd)\phi_{\varepsilon,j}^{\pm}(x_{d}) with j=1,,p1j=1,\ldots,p-1. This is easily done by identities (2.43): the functions ϕε,j±(xd)\phi_{\varepsilon,j}^{\pm}(x_{d}), j=1,,p1j=1,\ldots,p-1, decay exponentially if τImkε2<0\tau\operatorname{Im}k_{\varepsilon}^{2}<0, are periodic if Imkε2=0\operatorname{Im}k_{\varepsilon}^{2}=0 and grow exponentially if τImkε2>0\tau\operatorname{Im}k_{\varepsilon}^{2}>0. All discussed conditions can be checked by means of asymptotic expansions provided by Theorems 2.22.3 for a given pole. And exactly this is done in the proof of Theorems 2.42.5. Conditions in Theorem 2.4 are aimed at checking the sign of the real part of a given pole and proving at the same time that at least one of the coefficients cε,j±c_{\varepsilon,j}^{\pm}, j=1,,nj=1,\ldots,n, in (2.8) is non-zero. Similar conditions (2.27), (2.30), (2.31), (2.34), (2.35), (2.38) ensure that the functions ϕε,j±\phi_{\varepsilon,j}^{\pm}, j=1,,p1j=1,\ldots,p-1 decay exponentially, that is, τImkε2<0\tau\operatorname{Im}k_{\varepsilon}^{2}<0, while for j=p,,p+n1j=p,\ldots,p+n-1, these functions demonstrate either an exponential decay or an exponential growth. Conditions (2.39), (2.41), (2.42) describe a more gentle situation. Namely, here the real part of the pole is negative and the functions ϕε,j±\phi_{\varepsilon,j}^{\pm}, j=p,,p+n1j=p,\ldots,p+n-1, decay exponentially. However, τImkε2>0\tau\operatorname{Im}k_{\varepsilon}^{2}>0 and this means that the functions ϕε,j±\phi_{\varepsilon,j}^{\pm}, j=1,,p1j=1,\ldots,p-1, can grow exponentially. This is true, once we guarantee that at least one of the coefficients cε,j±c_{\varepsilon,j}^{\pm}, j=1,,p1j=1,\ldots,p-1, is non-zero. This is indeed the case thanks to condition (2.42). The asymptotic expansions for the eigenvalues and the resonances provided in Theorems 2.42.5 are implied immediately by the formula λε=Λpkε2\lambda_{\varepsilon}=\Lambda_{p}-k_{\varepsilon}^{2} relating the eigenvalues/resonances with a pole kεk_{\varepsilon} and the asymptotic expansions for the poles stated in Theorems 2.22.3.

Let us briefly discuss the main ideas underlying our main results. First, we rather straightforwardly construct the meromorphic continuation for the resolvent of the unperturbed operator. Namely, we find explicitly the projection of the solution to problem (2.3) on the eigenfunctions ψj\psi_{j}, j=1,,mj=1,\ldots,m, and study then the properties of the coefficients in this projection and of the remaining orthogonal part in the solution. Once such continuation is constructed, for proving our main results, we apply an approach being a modification of the technique suggested in [5], [8], [25], [26]. The idea is to regard the perturbation as a right-hand side and to apply then the meromorphic continuation of the unperturbed operator. After some simple calculations this leads us to an operator equation with a certain finite rank perturbation. Resolving this equation, we rather easily succeed to construct the meromorphic continuation for the perturbed operator and identify its poles as solutions to a nonlinear eigenvalue problem for some explicitly calculated matrix depending also on the small parameter. Analysing then this problem by means of methods from the theory of complex functions, we study the existence of the poles and their asymptotic behavior.

Although we restrict ourselves by considering second order differential operators, our approach can be also adapted for certain operators of higher order. However, general higher order operators can have a richer spectral structure of the edges in the essential spectrum and there can be more complicated scenarios of their bifurcations under perturbations, see, for instance, [55]. This is why the case of second order operators deserves a separate study, what is done in the present work. Our results are of general nature and are applicable to wide classes of unperturbed operators and perturbations. In the next section we discuss some possible examples of both unperturbed operators and perturbations as well as some specific examples motivated by physical models.

3 Examples

In this section we provide examples demonstrating our main results.

3.1 Unperturbed operator

Here we discuss some examples of the unperturbed operator, namely, of the operator \mathcal{H}. This is a general self-adjoint second order differential operator and it includes such classical operators as a Schrödinger operator:

=Δ+A0,A0=A0(x),\mathcal{H}=-\Delta+A_{0},\qquad A_{0}=A_{0}(x^{\prime}),

a magnetic Schrödinger operator:

=(ix+A)22xd2+A0,A=(A1,,Ad1),Aj=Aj(x),A0=A0(x),\mathcal{H}=(\mathrm{i}\nabla_{x^{\prime}}+{A})^{2}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}+A_{0},\qquad{A=(A_{1},\ldots,A_{d-1})},\qquad{A_{j}=A_{j}(x^{\prime})},\qquad A_{0}=A_{0}(x^{\prime}),

a Schrödinger operator with metric:

=i,j=1d1xiAijxj2xd2+A0,Aij=Aij(x),A0=A0(x).\mathcal{H}=-\sum\limits_{i,j=1}^{d-1}\frac{\partial\ }{\partial x_{i}}A_{ij}\frac{\partial\ }{\partial x_{j}}-\frac{\partial^{2}\ }{\partial x_{d}^{2}}+A_{0},\qquad A_{ij}=A_{ij}(x^{\prime}),\qquad A_{0}=A_{0}(x^{\prime}).

All these operators are considered in a tubular domain Ω=ω×\Omega=\omega\times\mathds{R}. If ω=d1\omega=\mathds{R}^{d-1}, then the domain Ω\Omega becomes an entire space d\mathds{R}^{d}. If ω\omega is a bounded domain, not necessary connected, then Ω\Omega is an infinite cylinder, which is to be regarded as a quantum waveguide if the Dirichlet condition is imposed on its boundary and as an acoustic waveguide if the boundary is subject to the Neumann condition. Further examples of unbounded domains ω\omega are also possible. For instance, if ω\omega is the half-space ω:={x:xj>0}\omega:=\{x^{\prime}:\,x_{j}>0\} in d1\mathds{R}^{d-1} for some j=1,,d1j=1,\ldots,d-1, the domain Ω\Omega becomes the half-space {x:xj>0}\{x:\,x_{j}>0\} in d\mathds{R}^{d}. We can also consider a more complicated domain ω:={x:xd1<h(x1,,xd2)}\omega:=\{x^{\prime}:x_{d-1}<h(x_{1},\ldots,x_{d-2})\} for some smooth function hh, then Ω={x:xd1<h(x1,,xd2),xd}\Omega=\{x:\,x_{d-1}<h(x_{1},\ldots,x_{d-2}),x_{d}\in\mathds{R}\}.

3.2 Perturbation

In this subsection we discuss possible examples of the perturbing operator (ε)\mathcal{L}(\varepsilon). The first example is a second order differential operator:

(ε)=i,j=1nΥij(x,ε)2xixj+j=1nΥj(x,ε)xj+Υ0(x,ε).\mathcal{L}(\varepsilon)=\sum\limits_{i,j=1}^{n}{\Upsilon}_{ij}(x,\varepsilon)\frac{\partial^{2}\ }{\partial x_{i}\partial x_{j}}+\sum\limits_{j=1}^{n}{\Upsilon}_{j}(x,\varepsilon)\frac{\partial\ }{\partial x_{j}}+{\Upsilon}_{0}(x,\varepsilon). (3.1)

Here Υij,Υj,Υ0L(Ω){\Upsilon}_{ij},\,{\Upsilon}_{j},\,{\Upsilon}_{0}\in L_{\infty}(\Omega) are some functions, not necessarily real-valued, satisfying the representations

Υ(x,ε)=Υ(1)(x)+εΥ(2)(x)+ε2Υ(3)(x,ε),=ij,j,0,{\Upsilon}_{\natural}(x,\varepsilon)={\Upsilon}_{\natural}^{(1)}(x)+\varepsilon{\Upsilon}_{\natural}^{(2)}(x)+\varepsilon^{2}{\Upsilon}_{\natural}^{(3)}(x,\varepsilon),\qquad\natural=ij,j,0, (3.2)

where Υ(s)L(Ω){\Upsilon}_{\natural}^{(s)}\in L_{\infty}(\Omega) are some functions obeying the estimates:

Υ(s)e2ϑ|xd|L(Ω)<C,\|{\Upsilon}_{\natural}^{(s)}e^{2{\vartheta}|x_{d}|}\|_{L_{\infty}(\Omega)}<C,

and ϑ\vartheta and CC are some fixed positive constant independent of ε\varepsilon. In this case the operators s\mathcal{L}_{s} read as

s:=i,j=1nΥij(s)2xixj+j=1nΥj(s)xj+Υ0(s),s=1,2,3.\mathcal{L}_{s}:=\sum\limits_{i,j=1}^{n}{\Upsilon}_{ij}^{(s)}\frac{\partial^{2}\ }{\partial x_{i}\partial x_{j}}+\sum\limits_{j=1}^{n}{\Upsilon}_{j}^{(s)}\frac{\partial\ }{\partial x_{j}}+{\Upsilon}_{0}^{(s)},\qquad s=1,2,3.

Particular cases of this example are small potential, small magnetic field, small metric.

The second example is an integral operator of the form

((ε)u)(x,ε)=ΩJ(x,y,ε)u(y)𝑑y,(\mathcal{L}(\varepsilon)u)(x,\varepsilon)=\int\limits_{\Omega}{J}(x,y,\varepsilon)u(y)\,dy,

where JL2(Ω×Ω){J}\in L_{2}(\Omega\times\Omega) is some kernel, not necessarily real-valued and symmetric and satisfying the representation:

J(x,y,ε)=J1(x,y)+εJ2(x,y)+ε2J3(x,y,ε),{J}(x,y,\varepsilon)={J}_{1}(x,y)+\varepsilon{J}_{2}(x,y)+\varepsilon^{2}{J}_{3}(x,y,\varepsilon),

where LiL_{i} are some functions obeying the estimates:

Ω×Ω|Ji|2evt(|xd|+|yd|)𝑑x𝑑y<C,i=1,2,3,\int\limits_{\Omega\times\Omega}|{J}_{i}|^{2}e^{{vt}(|x_{d}|+|y_{d}|)}\,dx\,dy<C,\qquad i=1,2,3,

and vt{vt} and CC are some fixed positive constant independent of ε\varepsilon.

The third example is a localized δ\delta-interaction with a complex-valued density. Namely, let SΩS\subset\Omega be a manifold of codimension 11 and of smoothness C3C^{3}. We assume that it is compact and has no edge. The perturbed operator in question is

ε=Δ+εβδ(xS),\mathcal{H}_{\varepsilon}=-\Delta+\varepsilon\beta\delta(x-S),

which acts as εu=Δu\mathcal{H}_{\varepsilon}u=-\Delta u on the domain 𝔇(ε)\mathfrak{D}(\mathcal{H}_{\varepsilon}) formed by the functions uW22(ΩS)W21(Ω)u\in W_{2}^{2}(\Omega\setminus S)\cap W_{2}^{1}(\Omega) obeying the boundary condition u=0\mathcal{B}u=0 on Ω\partial\Omega and the boundary conditions

[u]S=0,[uν]S=εβu.[u]_{S}=0,\qquad\left[\frac{\partial u}{\partial\nu}\right]_{S}=\varepsilon\beta u.

Here [u]S[u]_{S} denotes the jump of the function on SS, namely,

[u]S:=limt0+(u(+tν)u(tν)),[u]_{S}:={\lim\limits_{t\to 0+}\big{(}u(\cdot+t\nu)-u(\cdot-t\nu)\big{)},}

and ν\nu is the unit normal to SS directed outside the domain enveloped by SS. By β\beta we denote some complex-valued function defined on SS, uniformly bounded and belonging to C2(S)C^{2}(S) . Such operator does not satisfy our assumptions for (ε)\mathcal{L}(\varepsilon) since now the perturbation changes the domain. However, it is possible to reduce the perturbed operator to another one obeying needed assumptions and having the same eigenvalues and resonances. Namely, thanks to the made assumptions on the manifold SS, in a small vicinity of the SS we can introduce a new variable ρ\rho being the distance from a point to SS measured along the normal ν\nu. This variable is well-defined at least in the neighbourhood {x:dist(x,S)<ρ0}\{x:\,\operatorname{dist}(x,S)<\rho_{0}\} of SS, where ρ0\rho_{0} is some fixed number. Let χ=χ(ρ)\chi=\chi(\rho) be an infinitely differentiable function such that χ(ρ)=|ρ|2\chi(\rho)=\frac{|\rho|}{2} as |ρ|<ρ03|\rho|<\frac{\rho_{0}}{3} and χ(ρ)=0\chi(\rho)=0 as |ρ|>2ρ03|\rho|>\frac{2\rho_{0}}{3}. By 𝒰ε\mathcal{U}_{\varepsilon} we denote the multiplication operator 𝒰εu:=(1+εβχ)1u\mathcal{U}_{\varepsilon}u:=(1+\varepsilon\beta\chi)^{-1}u. It is straightforward to check that this operator maps the domain of the operator ε\mathcal{H}_{\varepsilon} onto the space {uW22(Ω):u=0onΩ}\{u\in W_{2}^{2}(\Omega):\,\mathcal{B}u=0\ \text{on}\ \partial\Omega\}. This space serves as the domain for an operator ~ε:=𝒰εε𝒰ε1\tilde{\mathcal{H}}_{\varepsilon}:=\mathcal{U}_{\varepsilon}\mathcal{H}_{\varepsilon}\mathcal{U}_{\varepsilon}^{-1}. It is straightforward to confirm that the differential operator for the latter operator reads as

~ε=Δ+ε(ε),(ε):=2(1+εβχ)1βχ(1+εβχ)1Δβχ\tilde{\mathcal{H}}_{\varepsilon}=-\Delta+\varepsilon\mathcal{L}(\varepsilon),\qquad\mathcal{L}(\varepsilon):=-2(1+\varepsilon\beta\chi)^{-1}\nabla\beta\chi\cdot\nabla-(1+\varepsilon\beta\chi)^{-1}\Delta\beta\chi

and we see that a first order differential operator (ε)\mathcal{L}(\varepsilon) is a particular case of operator (3.1). It is also clear that the operators ε\mathcal{H}_{\varepsilon} and ~ε\tilde{\mathcal{H}}_{\varepsilon} have the same eigenvalues and resonances since the operator 𝒰ε\mathcal{U}_{\varepsilon} does not change the behavior of the functions at infinity. Hence, we can study the eigenvalues and resonances of the operator ~ε\tilde{\mathcal{H}}_{\varepsilon} and transfer then the results to the operator ε\mathcal{H}_{\varepsilon}.

Our fourth example is a geometric perturbation. Namely, let ω\omega have a non-empty boundary, then the same is true for Ω\Omega. By Γ\Gamma we denote a bounded subset of the boundary Ω\partial\Omega. Let ρ\rho be a distance to a point measured along the outward normal to Ω\partial\Omega and hC2(Ω)h\in C^{2}(\partial\Omega) be some real function defined on Γ\Gamma and compactly supported in Γ\Gamma. Then we consider a domain Ωε\Omega_{\varepsilon} obtained by a small variation of the part Γ\Gamma of the boundary Ω\partial\Omega. Namely, Ωε\Omega_{\varepsilon} is a domain with the following boundary

Ωε:=(ΩΓ){x:ρ=εh}.\partial\Omega_{\varepsilon}:=(\partial\Omega\setminus\Gamma)\cup\{x:\,\rho=\varepsilon h\}.

In such domain we consider an operator with differential expression (2.1) subject to the Dirichlet boundary condition or Neumann condition. We assume that all the coefficients in the differential expression depend on xx^{\prime} only and are infinitely differentiable. Such perturbed operator does not fit our scheme since here the domain Ωε\Omega_{\varepsilon} depends on ε\varepsilon. However, as in the previous example, it is possible to transform such operator to another one fitting our assumptions. Namely, let χ=χ(x)\chi=\chi(x) be an infinitely differentiable cut-off function equalling to one in some fixed sufficiently small dd-dimensional neighbourhood of Γ\Gamma and vanishing outside some bigger neighbourhood. In this bigger neighbourhood we introduce local coordinates (P,ρ)(P,\rho), where PΩP\in\partial\Omega. A point xx is recovered from (P,ρ)(P,\rho) by measuring the distance ρ\rho along the outward normal to Ω\partial\Omega at the point PP. Then we define a mapping 𝒫\mathcal{P} by the following rule: for each point xx, we find corresponding (P,ρ)(P,\rho) and the action of the mapping is a point corresponding to (P,ρεh(P))(P,\rho-\varepsilon h(P)). We introduce new coordinates by the formula x~:=x(1χ(x))+εχ(x)𝒫(x)\tilde{x}:=x(1-\chi(x))+\varepsilon\chi(x)\mathcal{P}(x). It is easy to see that these coordinates are well-defined provided ε\varepsilon is small enough and after passing to these new coordinates, the domain Ωε\Omega_{\varepsilon} transforms into Ω\Omega, while the operator ε\mathcal{H}_{\varepsilon} becomes +(ε)\mathcal{H}+\mathcal{L}(\varepsilon), where (ε)\mathcal{L}(\varepsilon) is some second order differential operator of form (3.1) with compactly supported coefficients obeying (3.2).

3.3 Emerging poles for particular models

In this section we apply Theorems 2.42.5 to some simple two- and three-dimensional operators motivated by an interesting physical background.

3.3.1 Planar waveguide

The first model is an infinite planar waveguide modeled by the Dirichlet Laplacian. Namely, we let d=2d=2, ω:=(0,π)\omega:=(0,\pi), and =d2dx22\mathcal{H}^{\prime}=-\frac{d^{2}\ }{dx_{2}^{2}} subject to the Dirichlet boundary condition. Then Ω:={x: 0<x1<π}\Omega:=\{x:\,0<x_{1}<\pi\} is an infinite strip and =Δ\mathcal{H}=-\Delta is the Dirichlet Laplacian in Ω\Omega. As a perturbation, we choose a complex-valued potential of the form (ε):=V1+εV2\mathcal{L}(\varepsilon):=V_{1}+\varepsilon V_{2}, where Vi=Vi(x)V_{i}=V_{i}(x) are some continuous compactly supported complex-valued functions. The operator \mathcal{H}^{\prime} has a purely discrete spectrum formed by simple eigenvalues Λp:=p2\Lambda_{p}:=p^{2}, pp\in\mathds{N}, and the associated eigenfunctions normalized in L2(0,π)L_{2}(0,\pi) are ψp(xd):=2πsinpx1\psi_{p}(x_{d}):=\frac{\sqrt{2}}{\sqrt{\pi}}\sin px_{1}.

This situation models a slab optical waveguide of a finite width, where the cladding in direction x1x_{1} imposes zero boundary conditions at x1=0x_{1}=0 and x1=πx_{1}=\pi. Assuming that the waveguide is infinite in the second direction, a paraxial diffraction of an incident beam can be described using the normalized equation in the form (see e.g. [36])

izΦ+ΔΦεVε(x)Φ=0,\mathrm{i}\partial_{z}\Phi+\Delta\Phi-\varepsilon V_{\varepsilon}(x)\Phi=0, (3.3)

where Φ(x1,x2,z)\Phi(x_{1},x_{2},z) corresponds to complex amplitude of the electrical field, the optical potential Vε(x)V_{\varepsilon}(x) describes a weak localized modulation of the complex-valued refractive index, and zz is the direction of propagation of the pulse. For stationary modes Φ=eiλzψ\Phi=e^{-\mathrm{i}\lambda z}\psi, where λ-\lambda has the meaning of propagation constant, equation (3.3) reduces to the eigenvalue problem in the above described planar waveguide for the equation

Δψ+εVεψ=λψ.-\Delta\psi+\varepsilon V_{\varepsilon}\psi=\lambda\psi.

This is exactly the mathematical model we formulated above once we let Vε=V1+εV2V_{\varepsilon}=V_{1}+\varepsilon V_{2}. Let us consider the bifurcation of the thresholds p2p^{2} under the presence of a small localized potential VεV_{\varepsilon}.

For p=1p=1, there is just one pole kεk_{\varepsilon} and according formula (2.44), its asymptotic behavior reads as

kε=επΩV1(x)sin2x1dxε2πΩ(V2(x)sin2x1V1(x)U(x)sinx1)𝑑x+O(ε3),k_{\varepsilon}=-\frac{\varepsilon}{\pi}\int\limits_{\Omega}V_{1}(x)\sin^{2}x_{1}\,dx-\frac{\varepsilon^{2}}{\pi}\int\limits_{\Omega}\big{(}V_{2}(x)\sin^{2}x_{1}-V_{1}(x)U(x)\sin x_{1}\big{)}\,dx+O(\varepsilon^{3}), (3.4)

where U:=𝒢1,τ(V1sinx1)U:=\mathcal{G}_{1,\tau}(V_{1}\sin x_{1}). This function is given by formula (2.10). The term U:=((Λ1)1f)(x)U^{\bot}:=((\mathcal{H}^{\bot}-\Lambda_{1})^{-1}f^{\bot})(x) with f=V1sinx1f=V_{1}\sin x_{1} and ff^{\bot} defined by (2.11) solves the boundary value problem

(Δ1)U=V1sinx12πsinx10πV1(t1,x2)sin2t1dt1inΩ,U=0onΩ.(-\Delta-1)U^{\bot}=V_{1}\sin x_{1}-\frac{2}{\pi}\sin x_{1}\int\limits_{0}^{\pi}V_{1}(t_{1},x_{2})\sin^{2}t_{1}\,dt_{1}\quad\text{in}\quad\Omega,\qquad U^{\bot}=0\quad\text{on}\quad\partial\Omega.

This problem can be solved explicitly by the separation of variables and this gives the final formula for UU:

U(x)=\displaystyle U(x)= 1πsinx1Ω|x2t2|V1(t1,t2)sin2t1dt\displaystyle-\frac{1}{\pi}\sin x_{1}\int\limits_{\Omega}|x_{2}-t_{2}|V_{1}(t_{1},t_{2})\sin^{2}t_{1}\,dt
+j=2sinjx1πj21Ωej21|x2t2|V1(t1,t2)sint1sinjt1dt.\displaystyle+\sum\limits_{j=2}^{\infty}\frac{\sin jx_{1}}{\pi\sqrt{j^{2}-1}}\int\limits_{\Omega}e^{-\sqrt{j^{2}-1}|x_{2}-t_{2}|}V_{1}(t_{1},t_{2})\sin t_{1}\sin jt_{1}\,dt.

Hence,

Ω(V2(x)\displaystyle\int\limits_{\Omega}\big{(}V_{2}(x) sin2x1V1(x)U(x)sinx1)dx\displaystyle\sin^{2}x_{1}-V_{1}(x)U(x)\sin x_{1}\big{)}\,dx
=\displaystyle= ΩV2(x)sin2x1dx+1πΩ2|x2t2|V1(x)V1(t)sin2t1sin2x1dtdx\displaystyle\int\limits_{\Omega}V_{2}(x)\sin^{2}x_{1}\,dx+\frac{1}{\pi}\int\limits_{\Omega^{2}}|x_{2}-t_{2}|V_{1}(x)V_{1}(t)\sin^{2}t_{1}\sin^{2}x_{1}\,dt\,dx
j=21πj21Ω2ej21|x2t2|V1(t)V1(x)sint1sinjt1sinx1sinjx1dtdx.\displaystyle-\sum\limits_{j=2}^{\infty}\frac{1}{\pi\sqrt{j^{2}-1}}\int\limits_{\Omega^{2}}e^{-\sqrt{j^{2}-1}|x_{2}-t_{2}|}V_{1}(t)V_{1}(x)\sin t_{1}\sin jt_{1}\sin x_{1}\sin jx_{1}\,dt\,dx.

Now we apply Theorem 2.4 and we see that if

ReΩV1(x)sin2x1dx<0\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}x_{1}\,dx<0

or

ReΩV1(x)sin2x1dx=0andReΩ(V2(x)sin2x1V1(x)U(x)sinx1)𝑑x<0,\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}x_{1}\,dx=0\quad\text{and}\quad\operatorname{Re}\int\limits_{\Omega}\big{(}V_{2}(x)\sin^{2}x_{1}-V_{1}(x)U(x)\sin x_{1}\big{)}\,dx<0,

then the pole kεk_{\varepsilon} corresponds to an eigenvalue λε=1kε2\lambda_{\varepsilon}=1-k_{\varepsilon}^{2}. And if

ReΩV1(x)sin2x1dx>0\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}x_{1}\,dx>0

or

ReΩV1(x)sin2x1dx=0andReΩ(V2(x)sin2x1V1(x)U(x)sinx1)𝑑x>0,\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}x_{1}\,dx=0\quad\text{and}\quad\operatorname{Re}\int\limits_{\Omega}\big{(}V_{2}(x)\sin^{2}x_{1}-V_{1}(x)U(x)\sin x_{1}\big{)}\,dx>0,

then the pole kεk_{\varepsilon} corresponds to a resonance λε=1kε2\lambda_{\varepsilon}=1-k_{\varepsilon}^{2}. The asymptotic expansion for this eigenvalue/resonance is given by (2.23), (2.26) but it is more straightforward to find it by (3.4) and the above formula for λε\lambda_{\varepsilon}.

We proceed to the case p>1p>1. Here we again apply formula (2.44) to obtain

kε,τ=επΩV1(x)sin2px1dxε2πΩ(V2(x)sin2px1V1(x)Uτ(x)sinpx1)𝑑x+O(ε3),k_{\varepsilon,\tau}=-\frac{\varepsilon}{\pi}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx-\frac{\varepsilon^{2}}{\pi}\int\limits_{\Omega}\big{(}V_{2}(x)\sin^{2}px_{1}-V_{1}(x)U_{\tau}(x)\sin px_{1}\big{)}\,dx+O(\varepsilon^{3}), (3.5)

where Uτ:=𝒢τV1sinpx1U_{\tau}:=\mathcal{G}_{\tau}V_{1}\sin px_{1} is given by formula (2.10). The term Uτ:=((Λp)1f)(x)U_{\tau}^{\bot}:=((\mathcal{H}^{\bot}-\Lambda_{p})^{-1}f^{\bot})(x) with f=V1sinpx1f=V_{1}\sin px_{1} and ff^{\bot} defined by (2.11) solves the boundary value problem

(Δp2)Uτ=V1sinpx12πsinpx10πV1(t1,x2)sin2pt1dt1inΩ,Uτ=0onΩ.(-\Delta-p^{2})U_{\tau}^{\bot}=V_{1}\sin px_{1}-\frac{2}{\pi}\sin px_{1}\int\limits_{0}^{\pi}V_{1}(t_{1},x_{2})\sin^{2}pt_{1}\,dt_{1}\quad\text{in}\quad\Omega,\qquad U_{\tau}^{\bot}=0\quad\text{on}\quad\partial\Omega.

The solution is again given by the separation of variables and a final formula for UτU_{\tau} reads as

Uτ(x)=j=1p1iτsinjx1πp2j2eiτp2j2|x2t2|Uj(t2)𝑑t2\displaystyle U_{\tau}(x)=\sum\limits_{j=1}^{p-1}\frac{\mathrm{i}\tau\sin jx_{1}}{\pi\sqrt{p^{2}-j^{2}}}\int\limits_{\mathds{R}}e^{\mathrm{i}\tau\sqrt{p^{2}-j^{2}}|x_{2}-t_{2}|}U_{j}(t_{2})\,dt_{2} (3.6)
1πsinpx1|x2t2|Up(t2)𝑑t2\displaystyle\hphantom{U_{\tau}(x)=\sum\limits_{j=1}^{p-1}}-\frac{1}{\pi}\sin px_{1}\int\limits_{\mathds{R}}|x_{2}-t_{2}|U_{p}(t_{2})\,dt_{2}
+j=p+1sinjx1πj2p2ej2p2|x2t2|Uj(t2)𝑑t2.\displaystyle\hphantom{U_{\tau}(x)=\sum\limits_{j=1}^{p-1}}+\sum\limits_{j=p+1}^{\infty}\frac{\sin jx_{1}}{\pi\sqrt{j^{2}-p^{2}}}\int\limits_{\mathds{R}}e^{-\sqrt{j^{2}-p^{2}}|x_{2}-t_{2}|}U_{j}(t_{2})\,dt_{2}.
Uj(x2):=0πV1(t1,x2)sinpt1sinjt1dt,jp,\displaystyle U_{j}(x_{2}):=\int\limits_{0}^{\pi}V_{1}(t_{1},x_{2})\sin pt_{1}\sin jt_{1}\,dt,\qquad j\neq p,

Then we get:

ΩV1(x)Uτ(x)sinpx1dx=\displaystyle\int\limits_{\Omega}V_{1}(x)U_{\tau}(x)\sin px_{1}\,dx= j=1p12iτeiτp2j2|x2t2|πp2j2Uj(x2)Uj(t2)𝑑x2𝑑t2\displaystyle\sum\limits_{j=1}^{p-1}\int\limits_{\mathds{R}^{2}}\frac{\mathrm{i}\tau e^{\mathrm{i}\tau\sqrt{p^{2}-j^{2}}|x_{2}-t_{2}|}}{\pi\sqrt{p^{2}-j^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dx_{2}\,dt_{2} (3.7)
1π2|x2t2|Up(x2)Up(t2)𝑑x2𝑑t2\displaystyle\hphantom{\sum\limits_{j=1}^{p-1}}-\frac{1}{\pi}\int\limits_{\mathds{R}^{2}}|x_{2}-t_{2}|U_{p}(x_{2})U_{p}(t_{2})\,dx_{2}\,dt_{2}
+j=p+12ej2p2|x2t2|πj2p2Uj(x2)Uj(t2)𝑑t2𝑑x2.\displaystyle+\sum\limits_{j=p+1}^{\infty}\int\limits_{\mathds{R}^{2}}\frac{e^{-\sqrt{j^{2}-p^{2}}|x_{2}-t_{2}|}}{\pi\sqrt{j^{2}-p^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}.

Now we can apply Theorem 2.5 for τ=+1\tau=+1 and τ=1\tau=-1 and to determine whether the poles kε,τk_{\varepsilon,\tau} correspond to eigenvalues or resonances. As we see, in a general situation we can have two eigenvalues or two resonances or one eigenvalue and one resonance. Let us show that each of these situations is possible.

First of all we observe that in notations of Theorem 2.5 we have N=1N=1, qi=1q_{i}=1, ri=0r_{i}=0,

μ1=1πΩV1(x)sin2px1dx,\displaystyle\mu_{1}=-\frac{1}{\pi}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx,
(γi,τ)1qiri,τeπiqiri,τ(jri,τ)=1πΩ(V2(x)sin2px1V1(x)U1,τ(x)sinpx1)𝑑x.\displaystyle(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{\pi\mathrm{i}}{q_{i}-r_{i,\tau}}(j-r_{i,\tau})}=-\frac{1}{\pi}\int\limits_{\Omega}\big{(}V_{2}(x)\sin^{2}px_{1}-V_{1}(x)U_{1,\tau}(x)\sin px_{1}\big{)}\,dx.

Assume now that V1V_{1} is a complex-valued potential such that

ReΩV1(x)sin2px1dx>0.\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx>0.

Then condition (2.35) is satisfied and both poles kε,τk_{\varepsilon,\tau}, τ={1,+1}\tau=\{-1,+1\}, correspond to resonances.

If

ReΩV1(x)sin2px1dx<0,ImΩV1(x)sin2px1dx0,\operatorname{Re}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx<0,\qquad\operatorname{Im}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx\neq 0,

then conditions (2.27), (2.31) are satisfied with

τ=sgnImΩV1(x)sin2px1dx\tau=\operatorname{sgn}\operatorname{Im}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx

and for such τ\tau, the pole kε,τk_{\varepsilon,\tau} corresponds to an eigenvalue. If in addition,

Ωeip2s2x2V1(x)sinsx1sinpx1dx0orΩeip2s2x2V1(x)sinsx1sinpx1dx0,\int\limits_{\Omega}e^{-\mathrm{i}\sqrt{p^{2}-s^{2}}x_{2}}V_{1}(x)\sin sx_{1}\sin px_{1}\,dx\neq 0\quad\text{or}\quad\int\limits_{\Omega}e^{\mathrm{i}\sqrt{p^{2}-s^{2}}x_{2}}V_{1}(x)\sin sx_{1}\sin px_{1}\,dx\neq 0,

for some s{1,,p1}s\in\{1,\ldots,p-1\}, then conditions (2.39), (2.42) are satisfied and the pole kε,τk_{\varepsilon,\tau} with

τ=sgnImΩV1(x)sin2px1dx\tau=-\operatorname{sgn}\operatorname{Im}\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx

corresponds to a resonance.

In order to realize a situation with two eigenvalues, we assume that V2=0V_{2}=0 and consider a special class of 𝒫𝒯\mathcal{PT}-symmetric potentials V1V_{1}. Namely, we suppose that

V1(x)=W1(x)+iW2(x),V_{1}(x)=W_{1}(x)+\mathrm{i}W_{2}(x), (3.8)

where W1W_{1}, W2W_{2} are real-valued compactly supported potentials with certain parity:

W1(x1,x2)=W1(x1,x2),W2(x1,x2)=W2(x1,x2).W_{1}(x_{1},-x_{2})=W_{1}(x_{1},x_{2}),\qquad W_{2}(x_{1},-x_{2})=-W_{2}(x_{1},x_{2}). (3.9)

These assumptions yield that the operator ε\mathcal{H}_{\varepsilon} is 𝒫𝒯\mathcal{PT}-symmetric (or partially 𝒫𝒯\mathcal{PT}-symmetric using the terminology from [68]). They also imply immediately that

ΩV1(x)sin2px1dx=ΩW1(x)sin2px1dx\int\limits_{\Omega}V_{1}(x)\sin^{2}px_{1}\,dx=\int\limits_{\Omega}W_{1}(x)\sin^{2}px_{1}\,dx

and we assume that

ΩW1(x)sin2px1dx<0.\int\limits_{\Omega}W_{1}(x)\sin^{2}px_{1}\,dx<0. (3.10)

It follows from assumptions (3.9) and the definition of the functions UjU_{j} in (3.6) that these functions are given by the formulae

Uj(x2):=W1,j(x2)+iW2,j(x2),Ws,j(x2):=0πWs(t1,x2)sinpt1sinjt1dt,s=1,2,U_{j}(x_{2}):=W_{1,j}(x_{2})+\mathrm{i}W_{2,j}(x_{2}),\qquad W_{s,j}(x_{2}):=\int\limits_{0}^{\pi}W_{s}(t_{1},x_{2})\sin pt_{1}\sin jt_{1}\,dt,\qquad s=1,2,

and the functions W1,jW_{1,j} are even, while W2,jW_{2,j} are odd. As jp+1j\geqslant p+1, by making the change of the variables x2x2x_{2}\mapsto-x_{2}, t2t2t_{2}\mapsto-t_{2} in the integrals in the second sum in (3.7), we get:

2ej2p2|x2t2|πj2p2Uj(x2)Uj(t2)𝑑t2𝑑x2=2ej2p2|x2t2|πj2p2Uj(x2)¯Uj(t2)¯𝑑t2𝑑x2\int\limits_{\mathds{R}^{2}}\frac{e^{-\sqrt{j^{2}-p^{2}}|x_{2}-t_{2}|}}{\pi\sqrt{j^{2}-p^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}=\int\limits_{\mathds{R}^{2}}\frac{e^{-\sqrt{j^{2}-p^{2}}|x_{2}-t_{2}|}}{\pi\sqrt{j^{2}-p^{2}}}\overline{U_{j}(x_{2})}\overline{U_{j}(t_{2})}\,dt_{2}\,dx_{2}

and hence,

Imj=p+12ej2p2|x2t2|πj2p2Uj(x2)Uj(t2)𝑑t2𝑑x2=0.\operatorname{Im}\sum\limits_{j=p+1}^{\infty}\int\limits_{\mathds{R}^{2}}\frac{e^{-\sqrt{j^{2}-p^{2}}|x_{2}-t_{2}|}}{\pi\sqrt{j^{2}-p^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}=0.

In the same way we confirm that

Imj=1p12τsinτp2j2|x2t2|πp2j2Uj(x2)Uj(t2)𝑑t2𝑑x2=0,\displaystyle\operatorname{Im}\sum\limits_{j=1}^{p-1}\int\limits_{\mathds{R}^{2}}\frac{\tau\sin\tau\sqrt{p^{2}-j^{2}}|x_{2}-t_{2}|}{\pi\sqrt{p^{2}-j^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}=0,
Im1π2|x2t2|Up(x2)Up(t2)𝑑x2𝑑t2=0.\displaystyle\operatorname{Im}\frac{1}{\pi}\int\limits_{\mathds{R}^{2}}|x_{2}-t_{2}|U_{p}(x_{2})U_{p}(t_{2})\,dx_{2}\,dt_{2}=0.

Hence, by two above identities and (3.7),

ImΩ\displaystyle\operatorname{Im}\int\limits_{\Omega} V1(x)Uτ(x)sinpx1dx=Imj=1p12iτcosτp2j2|x2t2|πτp2j2Uj(x2)Uj(t2)𝑑t2𝑑x2\displaystyle V_{1}(x)U_{\tau}(x)\sin px_{1}\,dx=\operatorname{Im}\sum\limits_{j=1}^{p-1}\int\limits_{\mathds{R}^{2}}\frac{\mathrm{i}\tau\cos\tau\sqrt{p^{2}-j^{2}}|x_{2}-t_{2}|}{\pi\tau\sqrt{p^{2}-j^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}
=\displaystyle= τRej=1p12cosp2j2|x2t2|πp2j2Uj(x2)Uj(t2)𝑑t2𝑑x2\displaystyle\tau\operatorname{Re}\sum\limits_{j=1}^{p-1}\int\limits_{\mathds{R}^{2}}\frac{\cos\sqrt{p^{2}-j^{2}}|x_{2}-t_{2}|}{\pi\sqrt{p^{2}-j^{2}}}U_{j}(x_{2})U_{j}(t_{2})\,dt_{2}\,dx_{2}
=\displaystyle= τj=1p12cosp2j2(x2t2)πp2j2(W1,j(x2)W1,j(t2)W2,j(x2)W2,j(t2))𝑑t2𝑑x2.\displaystyle\tau\sum\limits_{j=1}^{p-1}\int\limits_{\mathds{R}^{2}}\frac{\cos\sqrt{p^{2}-j^{2}}(x_{2}-t_{2})}{\pi\sqrt{p^{2}-j^{2}}}\big{(}W_{1,j}(x_{2})W_{1,j}(t_{2})-W_{2,j}(x_{2})W_{2,j}(t_{2})\big{)}\,dt_{2}\,dx_{2}.

By straightforward calculations, for an arbitrary compactly supported function W(x2)W(x_{2}) we obtain:

2cosp2j2(x2t2)W(x2)W(t2)𝑑t2𝑑x2=\displaystyle\int\limits_{\mathds{R}^{2}}\cos\sqrt{p^{2}-j^{2}}(x_{2}-t_{2})W(x_{2})W(t_{2})\,dt_{2}\,dx_{2}= (W(x2)cosp2j2x2dx2)2\displaystyle\left(\int\limits_{\mathds{R}}W(x_{2})\cos\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}
+(W(x2)sinp2j2x2dx2)2.\displaystyle+\left(\int\limits_{\mathds{R}}W(x_{2})\sin\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}.

Hence, by two latter identities and (3.9),

τImΩV1(x)Uτ(x)sinpx1dx=1πj=1p11p2j2\displaystyle\tau\operatorname{Im}\int\limits_{\Omega}V_{1}(x)U_{\tau}(x)\sin px_{1}\,dx=\frac{1}{\pi}\sum\limits_{j=1}^{p-1}\frac{1}{\sqrt{p^{2}-j^{2}}} ((W1,j(x2)cosp2j2x2dx2)2\displaystyle\left(\left(\int\limits_{\mathds{R}}W_{1,j}(x_{2})\cos\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}\right.
(W2,j(x2)sinp2j2x2dx2)2).\displaystyle\left.\hphantom{\Bigg{(}}-\left(\int\limits_{\mathds{R}}W_{2,j}(x_{2})\sin\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}\right).

The latter formula implies that the sign of its left hand side is the same for both τ{1,+1}\tau\in\{-1,+1\} and we can make this sign being 1-1 by choosing appropriately W2W_{2} once we fix W1W_{1} satisfying (3.10), namely, we can satisfy the condition

j=1p11p2j2\displaystyle\sum\limits_{j=1}^{p-1}\frac{1}{\sqrt{p^{2}-j^{2}}} ((W1,j(x2)cosp2j2x2dx2)2\displaystyle\left(\left(\int\limits_{\mathds{R}}W_{1,j}(x_{2})\cos\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}\right. (3.11)
(W2,j(x2)sinp2j2x2dx2)2)<0\displaystyle\left.\hphantom{\Bigg{(}}-\left(\int\limits_{\mathds{R}}W_{2,j}(x_{2})\sin\sqrt{p^{2}-j^{2}}x_{2}\,dx_{2}\right)^{2}\right)<0

For instance, this can be done by letting W2=αW~2W_{2}=\alpha\tilde{W}_{2} with a sufficient large α\alpha, where W~2=W~2(x)\tilde{W}_{2}=\tilde{W}_{2}(x) is a real odd compactly supported function such that

j=1p11p2j2(ΩW~2(x)sinpx1sinjx1sinp2j2x2dx)2>0.\sum\limits_{j=1}^{p-1}\frac{1}{\sqrt{p^{2}-j^{2}}}\left(\int\limits_{\Omega}\tilde{W}_{2}(x)\sin px_{1}\sin jx_{1}\sin\sqrt{p^{2}-j^{2}}x_{2}\,dx\right)^{2}>0.

Once conditions (3.10), (3.11) hold, Theorem 2.5 states that both poles kε,τk_{\varepsilon,\tau}, τ{1,+1}\tau\in\{-1,+1\} correspond to the eigenvalues located in the vicinity of the internal threshold Λp\Lambda_{p}.

The above analytic results, namely, the discussed asymptotic expansions, approximate well the true eigenvalues and resonances for sufficiently small ε\varepsilon. In order to demonstrate how small ε\varepsilon is to be chosen, we make some numerical computations.

For numerics we use

W1(x)=j=13ajsinjx1cosx22,W2(x)=j=13bjsinjx1sinx2,W_{1}(x)=-\sum_{j=1}^{3}a_{j}\sin jx_{1}\cos\frac{x_{2}}{2},\qquad W_{2}(x)=\sum_{j=1}^{3}b_{j}\sin jx_{1}\sin x_{2},

where aja_{j} and bjb_{j} are real coefficients, and we additionally let Wi(x)0W_{i}(x)\equiv 0 as |x2|>π|x_{2}|>\pi, i=1,2i=1,2. The corresponding eigenvalue problem is approximated using a second-difference numerical scheme with Dirichlet boundary conditions at x1=0x_{1}=0 and x1=πx_{1}=\pi and a decay condition at x2±x_{2}\to\pm\infty. In order to achieve a numerically efficient approximation of decay condition at x2±x_{2}\to\pm\infty, a quasi-equidistant grid [33] is used with a step size gradually increasing towards x2±x_{2}\to\pm\infty. For small ε0.2\varepsilon\lessapprox 0.2, where the localization of the eigenfunctions in x2x_{2}-direction is extremely weak, and an adequate approximation of the decay condition as x2±x_{2}\to\pm\infty is practically impossible, we use the Neumann condition in order to approximate slowly decaying oscillating tails of the eigenfunctions:

x2Ψ(x1,±X0)=0,whereX01.\partial_{x_{2}}\Psi(x_{1},\pm X_{0})=0,\quad\text{where}\quad X_{0}\gg 1.
Refer to caption
Figure 1: (a) Eigenvalue Λ1(ε)\Lambda_{1}(\varepsilon) emerging from the bottom of the spectrum computed using only the first term in expansion (3.4) [blue curve] and two terms in (3.4) [green curve]. Red points connected by red lines are obtained from direct numerical solution of the eigenvalue problem. (b) The real and imaginary parts of the eigenfunction at ε=0.1\varepsilon=0.1 plotted as a function of x2x_{2} for x1=π/2x_{1}=\pi/2. Here a1=1a_{1}=1, a3=4a_{3}=4, b1=0.5b_{1}=0.5 and all other coefficients are zero.
Refer to caption
Figure 2: The real and imaginary parts of the eigenvalue Λ2(ε)\Lambda_{2}(\varepsilon) emerging from an internal threshold in the essential spectrum computed using only the first term in expansion (3.5) [blue curve] and two terms in (3.5) [green curve]. Red points connected by red lines are obtained from direct numerical evaluation of the spectrum. (c) The real and imaginary parts of the eigenfunction at ε=0.3\varepsilon=0.3 plotted as a function of x2x_{2} for x1=π/4x_{1}=\pi/4. Two insets show the plots of the squared amplitude of the eigenfunction |Ψ|2=(ReΨ)2+(ImΨ)2|\Psi|^{2}=(\mathrm{Re\,}\Psi)^{2}+(\mathrm{Im\,}\Psi)^{2}, using larger domains in the horizontal axes and linear (left inset) and logarithmic (right inset) scales in the vertical axes. (d) Full plot of the modulus of the eigenfunction. In all panels a1=1a_{1}=1, b2=3b_{2}=3, and all other coefficients aia_{i} and bib_{i} are zero. Only the eigenvalue with a positive imaginary part is shown. There also exists a complex-conjugate eigenvalue with a negative imaginary part and a 𝒫𝒯\mathcal{PT}-conjugate eigenfunction.

In Fig. 1 we plot the dependencies Λ1(ε)\Lambda_{1}(\varepsilon) obtained from the asymptotic expansions, when only the leading term is taken into account and both terms are used, for the particular set of parameters a1=1a_{1}=1, b1=0.1b_{1}=0.1, a2=b2=0a_{2}=b_{2}=0. For ε0.2\varepsilon\lessapprox 0.2, the agreement between the analytical predictions and numerical results is rather good, while for large values of ε\varepsilon it is only qualitative. In Fig. 2 we plot the same data for two eigenvalues bifurcating from the internal threshold with p=2p=2; here only the eigenvalue with positive imaginary part is shown. The following set of parameters is used: a1=1a_{1}=1, b1=0b_{1}=0, a2=0a_{2}=0, and b2=3b_{2}=3. Again, the numerical results are in a good agreement with asymptotic expansions for weak perturbations, and in a qualitative agreement for stronger ones. Real part of the eigenvalue is found to decrease with the increase of ε\varepsilon. Nevertheless, for all values of ε\varepsilon shown in Fig. 2 the real part of eigenvalue Λ2(ε)\Lambda_{2}(\varepsilon) is larger than the lower edge of the essential spectrum. Therefore, in the optical context, the corresponding eigenfunction, whose three-dimensional modulus plot is shown in Fig. 2(d), indeed represents a non-Hermitian generalization of a bound state in the continuum.

3.3.2 Two-dimensional Bose-Einstein condensate with parabolic trapping

In the previous example a choice of ω\omega was not really important and the discussed results are of a more general nature. For instance, we can choose the operator \mathcal{H}^{\prime} being a quantum harmonic oscillator. Namely, let ω=\omega=\mathds{R} and

:=d2dx12+x12on.\mathcal{H}^{\prime}:=-\frac{d^{2}\ }{dx_{1}^{2}}+x_{1}^{2}\quad\text{on}\quad\mathds{R}.

Then Ω=2\Omega=\mathds{R}^{2} and the operator \mathcal{H} becomes

=Δ+x12on2.\mathcal{H}=-\Delta+x_{1}^{2}\quad\text{on}\quad\mathds{R}^{2}.

As a perturbation, we choose (ε)=V1\mathcal{L}(\varepsilon)=V_{1}, where V1V_{1} is a complex-valued compactly supported potential on 2\mathds{R}^{2}. Here

Λp=2p+1,ψp(x1)=ex2/22pp!πHp(x1),Hn(t):=(1)net2dndtnet2,\Lambda_{p}=2p+{1},\qquad\psi_{p}(x_{1})=\frac{e^{-x^{2}/2}}{\sqrt{2^{p}p!\sqrt{\pi}}}H_{p}\left(x_{1}\right),\qquad H_{n}(t):=(-1)^{n}e^{t^{2}}\frac{d^{n}\ }{dt^{n}}e^{-t^{2}},

i.e., HpH_{p} are Hermite polynomials, and p=0,1,p=0,1,\ldots. Then all calculations and results from previous example can be easily reproduced, just in all formulae the functions 2πsinjx1\frac{\sqrt{2}}{\sqrt{\pi}}\sin jx_{1} are to be replaced by the functions ψj(x1)\psi_{j}(x_{1}) introduced above.

The described situation corresponds to a two-dimensional cloud of Bose-Einstein condensate with a parabolic confinement in x1x_{1} direction. Assuming that the interparticle interactions are negligible, that is, the condensate is effectively linear, we can model its dynamics by the Schröginer-like equation, which in the theory of Bose-Einstein condensates is known as Gross-Pitaevskii equation [59]:

itΦ+ΔΦx12ΦεVε(x)Φ=0,\mathrm{i}\partial_{t}\Phi+\Delta\Phi-x_{1}^{2}\Phi-\varepsilon V_{\varepsilon}(x)\Phi=0, (3.12)

where Φ(x1,x2,t)\Phi(x_{1},x_{2},t) stands for the macroscopic wavefunction of the condensate. Again, for stationary states in the form Φ=eiλtΨ\Phi=e^{-\mathrm{i}\lambda t}\Psi, where λ\lambda has the meaning of the chemical potential, the problem is reduced to the described eigenvalue problem. Positive and negative imaginary parts of the perturbation correspond to the injecting the particles from an external source and absorption of the particles, respectively.

3.3.3 Three-dimensional circular waveguide and three-dimensional Bose-Einstein condensate

Here we consider two examples of a three-dimensional waveguide and a three-dimensional quantum oscillator, which extend the above examples of the planar waveguide and the harmonic oscillator adduced above.

In first example we deal with a circular waveguide assuming that ω={x=(x1,x2):|x|<1}\omega=\{x^{\prime}=(x_{1},x_{2}):\,|x^{\prime}|<1\} and the operator \mathcal{H}^{\prime} is introduced as the Schrödinger operator with a radially symmetric potential subject to the Dirichlet condition:

=Δx+V0inω,V0=V0(|x|).\mathcal{H}^{\prime}=-\Delta_{x^{\prime}}+V_{0}\quad\text{in}\quad\omega,\qquad V_{0}=V_{0}(|x^{\prime}|).

Then the domain Ω\Omega is a straight cylinder along the axis x3x_{3} with the cross-section ω\omega and

=Δ+V0(|x|)inΩ\mathcal{H}=-\Delta+V_{0}(|x^{\prime}|)\quad\text{in}\quad\Omega

subject to the homogeneous Dirichlet condition. The operator \mathcal{H}^{\prime} has a purely discrete spectrum and thanks to the assumed radial symmetricity for the potential V0V_{0}, it possesses double eigenvalues Λp=Λp+1>Λ1\Lambda_{p}=\Lambda_{p+1}>\Lambda_{1} such that the associated eigenfunctions, orthonormalized in L2(ω)L_{2}(\omega), read as ψp(x)=Ψ(|x|)cossθ\psi_{p}(x^{\prime})=\Psi(|x^{\prime}|)\cos s\theta, ψp+1(x)=Ψ(|x|)sinsθ\psi_{p+1}(x^{\prime})=\Psi(|x^{\prime}|)\sin s\theta, where ss is some fixed integer number, Ψ\Psi is some real function, and θ\theta is a polar angle associated with xx^{\prime}. Such eigenvalues are degenerate internal thresholds in the essential spectrum. We define then a 𝒫𝒯\mathcal{PT}-symmetric potential V1V_{1} by formula (3.8) and in addition, we assume that W1W_{1} is even in x1x_{1}. Then it is easy to see that the corresponding matrix M1\mathrm{M}_{1} is diagonal:

M1=(μ100μ2),μ1:=12ΩW1ψp2𝑑x,μ2:=12ΩW1ψp+12𝑑x.\mathrm{M}_{1}=\begin{pmatrix}\mu_{1}&0\\ 0&\mu_{2}\end{pmatrix},\qquad\mu_{1}:=-\frac{1}{2}\int\limits_{\Omega}W_{1}\psi_{p}^{2}\,dx,\qquad\mu_{2}:=-\frac{1}{2}\int\limits_{\Omega}W_{1}\psi_{p+1}^{2}\,dx.

Then we assume that μ1μ2\mu_{1}\neq\mu_{2} and μ1<0\mu_{1}<0, μ2<0\mu_{2}<0; these conditions can be easily satisfied by choosing appropriately W1W_{1}. Since for both μi\mu_{i} we have qi=1q_{i}=1, for each corresponding pole we can apply asymptotic expansion (2.44):

ki,τ(ε)=εμi+ε22ΩV1(x)Ui,τ(x)ψi1+p𝑑x+O(ε3),k_{i,\tau}(\varepsilon)=\varepsilon\mu_{i}+\frac{\varepsilon^{2}}{2}\int\limits_{\Omega}V_{1}(x)U_{i,\tau}(x)\psi_{i-1+p}\,dx+O(\varepsilon^{3}),

where Ui,τ:=𝒢1,τV1ψi1+pU_{i,\tau}:=\mathcal{G}_{1,\tau}V_{1}\psi_{i-1+p} are solutions to the boundary value problem

(ΔΛp)Ui,τ=V1ψi1+pψi1+pΩV1(t,x3)ψp2(t)𝑑t𝑑x3inΩ,Ui,τ=0onΩ,(-\Delta-\Lambda_{p})U_{i,\tau}=V_{1}\psi_{i-1+p}-\psi_{i-1+p}\int\limits_{\Omega}V_{1}(t^{\prime},x_{3})\psi_{p}^{2}(t^{\prime})\,dt^{\prime}{dx_{3}}\quad\text{in}\quad\Omega,\qquad U_{i,\tau}=0\quad\text{on}\quad\partial\Omega,

and can be found by a separation of variables similar to (3.6). Then we can reproduce calculations from the first example and obtain that choosing appropriately the function W2W_{2}, we can satisfy conditions (2.27), (2.34) for all poles ki,τk_{i,\tau}, i=1,2i=1,2, τ{1,+1}\tau\in\{-1,+1\} and this means that these poles correspond to eigenvalues. In other words, this means that for appropriately chosen W1W_{1} and W2W_{2}, we can generate four different simple eigenvalues of the operator ε\mathcal{H}_{\varepsilon} in the vicinity of the double eigenvalue Λp\Lambda_{p} of the operator \mathcal{H}^{\prime} serving as an internal threshold in the essential spectrum.

A similar situation can be realized also in other three-dimensional models. For instance, we can let ω=2\omega=\mathds{R}^{2} and =Δx+V0(|x|)\mathcal{H}^{\prime}=-\Delta_{x^{\prime}}+V_{0}(|x^{\prime}|), where V0=V0(t)V_{0}=V_{0}(t) is some function growing unboundedly at infinity. Such operator \mathcal{H}^{\prime} again can have double eigenvalues with the eigenfunctions of form Ψ(|x|)cossθ\Psi(|x^{\prime}|)\cos s\theta and Ψ(|x|)sinsθ\Psi(|x^{\prime}|)\sin s\theta. Linear combinations Ψ(|x|)cossθ±iΨ(|x|)sinsθ\Psi(|x^{\prime}|)\cos s\theta\pm\mathrm{i}\Psi(|x^{\prime}|)\sin s\theta correspond to vortex states with integer ss being the vorticity or topological charge. Therefore, in the context of Bose-Einstein condensates, this mechanism can be potentially applied for generation of localized in all three spatial dimensions vortex rings (see e.g. [19]).

We also observe that in both discussed three-dimensional examples the operator \mathcal{H}^{\prime} can have not only double eigenvalues, but also ones of higher multiplicities nn. And in such cases, it is possible to find a 𝒫𝒯\mathcal{PT}-symmetric potential V1V_{1} generating 2n2n eigenvalues of the operator ε\mathcal{H}_{\varepsilon} in the vicinity of the considered multiple eigenvalue.

4 Meromorphic continuation

In this section we prove Theorem 2.1 on the meromorphic continuation of the resolvent of the operator ε\mathcal{H}_{\varepsilon}. First we prove some auxiliary statements in a separate subsection and then we prove the theorem.

4.1 Auxiliary lemmata

In this subsection we prove three auxiliary lemmata, which will be employed in the proof of Theorem 2.1. The first statement is Lemma 2.1.

Proof of Lemma 2.1.

Reproducing literally the proof of Lemma 2.3 in [9], one can check easily the identity σess(ε)=σess()\operatorname{\sigma_{ess}}(\mathcal{H}_{\varepsilon})=\operatorname{\sigma_{ess}}(\mathcal{H}). The operator \mathcal{H} can be represented as a sum of tensor products =+0\mathcal{H}=\mathcal{H}^{\prime}\otimes\mathcal{I}+\mathcal{I}\otimes\mathcal{H}_{0}. And since

σess(0)=σ(0)=[0,+),infσ()=Λ1,\operatorname{\sigma_{ess}}(\mathcal{H}_{0})=\operatorname{\sigma}(\mathcal{H}_{0})=[0,+\infty),\qquad\inf\operatorname{\sigma}(\mathcal{H}^{\prime})=\Lambda_{1},

we immediately get:

σ()=σess()=[Λ1,+).\operatorname{\sigma}(\mathcal{H})=\operatorname{\sigma_{ess}}(\mathcal{H})=[\Lambda_{1},+\infty).

This completes the proof. ∎

The next statement is Lemma 2.2.

Proof of Lemma 2.2.

For each u𝔇()Lu\in\mathfrak{D}(\mathcal{H})\cap L^{\bot}, each ψj\psi_{j} and almost each xdx_{d}\in\mathds{R} we have:

((u)(,xd),ψj)L2(ω)=\displaystyle\big{(}(\mathcal{H}^{\bot}u)(\cdot,x_{d}),\psi_{j}\big{)}_{L_{2}(\omega)}= 𝔥(u(,xd),ψj)(2uxd2(,xd),ψj)L2(ω)\displaystyle\mathfrak{h}^{\prime}\big{(}u(\cdot,x_{d}),\psi_{j}\big{)}-\left(\frac{\partial^{2}u}{\partial x_{d}^{2}}(\cdot,x_{d}),\psi_{j}\right)_{L_{2}(\omega)}
=\displaystyle= Λj(u(,xd),ψj)L2(ω)d2dxd2(u(,xd),ψj)L2(ω)=0.\displaystyle-\Lambda_{j}\big{(}u(\cdot,x_{d}),\psi_{j}\big{)}_{L_{2}(\omega)}-\frac{d^{2}\ }{dx_{d}^{2}}\big{(}u(\cdot,x_{d}),\psi_{j}\big{)}_{L_{2}(\omega)}=0.

Hence, the operator \mathcal{H}^{\bot} maps 𝔇()L\mathfrak{D}(\mathcal{H})\cap L^{\bot} into LL^{\bot}. This is an unbounded self-adjoint operator in LL^{\bot} associated with the restriction of the form 𝔥\mathfrak{h} on 𝔇(𝔥)L\mathfrak{D}(\mathfrak{h})\cap L^{\bot}. Moreover, for each u𝔇(𝔥)Lu\in\mathfrak{D}(\mathfrak{h})\cap L^{\bot} we have:

𝔥(u,u)=\displaystyle\mathfrak{h}(u,u)= 𝔥(u(,xd),u(,xd))𝑑xd+Ω|uxd|2𝑑x𝔥(u(,xd),u(,xd))𝑑xd\displaystyle\int\limits_{\mathds{R}}\mathfrak{h}^{\prime}\big{(}u(\cdot,x_{d}),u(\cdot,x_{d})\big{)}\,dx_{d}+\int\limits_{\Omega}\left|\frac{\partial u}{\partial x_{d}}\right|^{2}\,dx\geqslant\int\limits_{\mathds{R}}\mathfrak{h}^{\prime}\big{(}u(\cdot,x_{d}),u(\cdot,x_{d})\big{)}\,dx_{d}
\displaystyle\geqslant c0du(,xd)L2(ω)2𝑑xd=c0uL2(Ω)2.\displaystyle c_{0}\int\limits_{\mathds{R}^{d}}\|u(\cdot,x_{d})\|_{L_{2}(\omega)}^{2}\,dx_{d}=c_{0}\|u\|_{L_{2}(\Omega)}^{2}.

Hence, the spectrum of the operator \mathcal{H}^{\bot} is located in [c0,+)[c_{0},+\infty). The proof is complete. ∎

The third lemma provides a meromorphic continuation for the resolvent of the unperturbed operator.

Lemma 4.1.

Fix p{1,,m}p\in\{1,\ldots,m\} and τ{1,+1}\tau\in\{-1,+1\} and let Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1} be an nn-multiple eigenvalue of the operator \mathcal{H}^{\prime}, where n1n\geqslant 1. There exists a sufficiently small fixed δ>0\delta>0 such that for all complex kBδk\in B_{\delta} and all fL2(Ω,eϑ|xd|dx)f\in L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) the boundary value problem

(i,j=1d1xiAijxj+ij=1d1(Ajxj+xjAj)+A0Λp+k2)u=finΩ,u=0onΩ,\begin{gathered}\left(-\sum\limits_{i,j=1}^{d-1}\frac{\partial\ }{\partial x_{i}}A_{ij}\frac{\partial\ }{\partial x_{j}}+\mathrm{i}\sum\limits_{j=1}^{d-1}\left(A_{j}\frac{\partial\ }{\partial x_{j}}+\frac{\partial\ }{\partial x_{j}}A_{j}\right)+A_{0}-\Lambda_{p}+k^{2}\right)u=f\quad\text{in}\quad\Omega,\\ \mathcal{B}u=0\quad\text{on}\quad\partial\Omega,\end{gathered} (4.1)

is solvable in W2,loc2(Ω)W_{2,loc}^{2}(\Omega) and possesses a solution, which can be represented as

u=𝒜1,τ(k)f,u=\mathcal{A}_{1,\tau}(k)f, (4.2)

where 𝒜1,τ\mathcal{A}_{1,\tau} is a linear operator mapping L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) into W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx). This operator is bounded and meromorphic in kBδk\in B_{\delta}. It has the only pole in BδB_{\delta}, which is at zero and simple:

𝒜1,τ(k)=1k𝒜2+𝒢p,τ+k𝒜4,τ(k),\displaystyle\mathcal{A}_{1,\tau}(k)=\frac{1}{k}\mathcal{A}_{2}+\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k), (4.3)
𝒜2f:=j=pp+n1ψjjf,jf:=12Ωψj(x)¯f(x)𝑑x,\displaystyle\mathcal{A}_{2}f:=\sum\limits_{j=p}^{p+n-1}\psi_{j}\ell_{j}f,\qquad\ell_{j}f:=\frac{1}{2}\int\limits_{\Omega}\overline{\psi_{j}(x^{\prime})}f(x)\,dx, (4.4)

where 𝒜4,τ(k):L2(Ω,eϑ|xd|dx)W22(Ω,eϑ|xd|dx)\mathcal{A}_{4,\tau}(k):\,L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx)\to W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) is an operator bounded uniformly in kBδ¯k\in\overline{B_{\delta}} and holomorphic in kBδk\in B_{\delta}, and, we recall, 𝒢p,τ\mathcal{G}_{p,\tau} is the operator defined in (2.10).

For sufficiently large xdx_{d}, the solution uu given by (4.2) can be represented as

u(x,k)=j=1muj±(xd,k)ψj(xd)+u±(x,k),±xd>R,u(x,k)=\sum\limits_{j=1}^{m}u_{j}^{\pm}(x_{d},k)\psi_{j}(x_{d})+u_{\bot}^{\pm}(x,k),\qquad\pm x_{d}>R, (4.5)

where RR is some fixed number, uj±L2(I±,eϑxddxd)u_{j}^{\pm}\in L_{2}(I_{\pm},e^{{\mp\vartheta}x_{d}}dx_{d}) are some meromorphic in kBδk\in B_{\delta} functions possessing the asymptotic behavior

uj±(xd)=eτKj(k)|xd|(Cj±(k)+O(eϑ~|xd|)),\displaystyle u_{j}^{\pm}(x_{d})=e^{-\tau K_{j}(k)|x_{d}|}\big{(}C_{j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)},\quad |xd|,j=1,,p1,\displaystyle|x_{d}|\to\infty,\qquad j=1,\ldots,p-1, (4.6)
uj±(xd)=eKj(k)|xd|(Cj±(k)+O(eϑ~|xd|)),\displaystyle u_{j}^{\pm}(x_{d})=e^{-K_{j}(k)|x_{d}|}\big{(}C_{j}^{\pm}(k)+O({e}^{-{\tilde{\vartheta}}|x_{d}|})\big{)}, |xd|,j=p,,m,\displaystyle|x_{d}|\to\infty,\qquad j=p,\ldots,m,

Cj±(k)C_{j}^{\pm}(k) are some meromorphic in kBδk\in B_{\delta} functions, 0<ϑ~<ϑ0<{\tilde{\vartheta}}<{\vartheta} is some fixed constant independent of kk and xx, and u±W22(ΩR±)u_{\bot}^{\pm}\in W_{2}^{2}(\Omega_{R}^{\pm}) are some functions meromorphic in kBδk\in B_{\delta} and obeying the identities

(u±(,xd),ψj)L2(Ω)=0(u_{\bot}^{\pm}(\cdot,x_{d}),\psi_{j})_{L_{2}(\Omega)}=0 (4.7)

for almost each xdI±x_{d}\in I_{\pm} and for each j=1,,mj=1,\ldots,m.

Proof.

The functions

fj(xd):=(f(,xd),ψj)L2(ω)f_{j}(x_{d}):=\big{(}f(\cdot,x_{d}),\psi_{j}\big{)}_{L_{2}(\omega)}

obviously belong to L2(,eϑ|xd|dxd)L_{2}(\mathds{R},e^{{\vartheta}|x_{d}|}dx_{d}) since fL2(Ω,eϑ|xd|dx)f\in L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx). It is straightforward to check that the functions

uj(xd,k):=12τKj(k)eτKj(k)|xdyd|fj(yd)𝑑yd\displaystyle u_{j}(x_{d},k):=\frac{1}{2\tau K_{j}(k)}\int\limits_{\mathds{R}}e^{-\tau K_{j}(k)|x_{d}-y_{d}|}f_{j}(y_{d})\,dy_{d}\quad asj<p1,\displaystyle\text{as}\quad j<p-1, (4.8)
uj(xd,k):=12Kj(k)eKj(k)|xdyd|fj(yd)𝑑yd\displaystyle u_{j}(x_{d},k):=\frac{1}{2K_{j}(k)}\int\limits_{\mathds{R}}e^{-K_{j}(k)|x_{d}-y_{d}|}f_{j}(y_{d})\,dy_{d}\quad asjp,\displaystyle\text{as}\quad j\geqslant p,

solve the equations

uj′′+(ΛjΛp+k2)uj=fjin,j=1,,m.-u_{j}^{\prime\prime}+(\Lambda_{j}-\Lambda_{p}+k^{2})u_{j}=f_{j}\quad\text{in}\quad\mathds{R},\qquad j=1,\ldots,m.

Employing these facts, we seek a solution to problem (4.1) as

u(x,k)=j=1muj(xd,k)ψj(x)+u(x,k)u(x,k)=\sum\limits_{j=1}^{m}u_{j}(x_{d}^{\prime},k)\psi_{j}(x^{\prime})+u_{\bot}(x,k) (4.9)

and for uu_{\bot} we immediately get problem (4.1) with ff replaced by the function ff^{\bot} defined in (2.11). We see easily that fLf^{\bot}\in L^{\bot}.

Thanks to Lemma 2.2, the resolvent (Λp+k2)1(\mathcal{H}^{\bot}-\Lambda_{p}+k^{2})^{-1} is well-defined for all sufficiently small complex kBδk\in B_{\delta} provided δ\delta is small enough. This resolvent is holomorphic in kBδk\in B_{\delta} as an operator in L2(Ω)L_{2}(\Omega) and is uniformly bounded in kBδ¯k\in\overline{B_{\delta}}. It can be expanded via the standard Neumann series:

(Λp+k2)1=j=0(k2)j((Λp)1)j+1.(\mathcal{H}^{\bot}-\Lambda_{p}+k^{2})^{-1}=\sum\limits_{j=0}^{\infty}(-k^{2})^{j}\big{(}(\mathcal{H}^{\bot}-\Lambda_{p})^{-1}\big{)}^{j+1}. (4.10)

This expansion implies that the resolvent (Λp+k2)1(\mathcal{H}^{\bot}-\Lambda_{p}+k^{2})^{-1} is also holomorphic as an operator from L2(Ω)L_{2}(\Omega) into W22(Ω)W_{2}^{2}(\Omega). We define then

u:=(Λp+k2)1fu_{\bot}:=(\mathcal{H}^{\bot}-\Lambda_{p}+k^{2})^{-1}f^{\bot}

and this obviously gives a solution to problem (4.1). We denote the operator mapping ff into the described solution by 𝒜1,τ(k)\mathcal{A}_{1,\tau}(k) and let us show that it possesses all stated properties.

First of all we observe that the introduced operator is independent of the choice of τ\tau as Λp=Λ1\Lambda_{p}=\Lambda_{1} since in this case the functions uju_{j} with Λj<Λp\Lambda_{j}<\Lambda_{p} are missing and the above constructions become independent of τ\tau.

Just by the embeddings L2(Ω,eϑ|xd|dx)L2(Ω)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx)\subset L_{2}(\Omega) and W22(Ω)W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega)\subset W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx), the resolvent (Λp+k2)1(\mathcal{H}^{\bot}-\Lambda_{p}+k^{2})^{-1} is a bounded operator from L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) into W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) holomorphic in kBδk\in B_{\delta} for sufficiently small δ\delta and bounded uniformly in kBδ¯k\in\overline{B_{\delta}}. An operator mapping ff into j=1muj(xd,k)ψj(x)\sum\limits_{j=1}^{m}u_{j}(x_{d},k)\psi_{j}(x^{\prime}) is given explicitly and by straightforward calculations we can check that this is also a bounded operator from L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) into W22(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx) holomorphic in kBδk\in B_{\delta}. The calculations are based on estimates of the following kind:

ujψjL2(Ω,eϑ|xd|dx)2=\displaystyle\|u_{j}\psi_{j}\|_{L_{2}(\Omega,e^{-{\vartheta}|x_{d}|}dx)}^{2}= ujL2(,eϑ|xd|dx)2\displaystyle\|u_{j}\|_{L_{2}(\mathds{R},e^{-{\vartheta}|x_{d}|}dx)}^{2}
\displaystyle\leqslant fjL2(,eϑ|xd|dxd)22|Kj|22e|xdyd|τReKj(k)a(|xd|+|yd|)𝑑xd𝑑yd\displaystyle\frac{\|f_{j}\|_{L_{2}(\mathds{R},e^{{\vartheta}|x_{d}|}dx_{d})}^{2}}{2|K_{j}|^{2}}\int\limits_{\mathds{R}^{2}}e^{-|x_{d}-y_{d}|\tau\operatorname{Re}K_{j}(k)-a(|x_{d}|+|y_{d}|)}\,dx_{d}\,dy_{d}
\displaystyle\leqslant C(k)fL2(Ω,eϑ|xd|dx)2,\displaystyle C(k)\|f\|_{L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx)}^{2},

where C(k)C(k) is some constant independent of ff. Since the functions ff belong to L2(Ω,eϑ|xd|dxd)L_{2}(\Omega,e^{\vartheta|x_{d}|}dx_{d}), the functions uju_{j} can be expanded into the Laurent series with respect to the small parameter kk; in particular, this means that the functionals j:L2(Ω,eϑ|xd|dx)\ell_{j}:L_{2}(\Omega,e^{\vartheta|x_{d}|}dx)\to\mathds{C} are well-defined and bounded. Substituting these expansions and (4.10) into (4.9), we immediately get representation (4.3), (4.4), (2.10).

In view of formula (4.9), in order to prove representation (4.5), (4.6), (4.7), it is sufficient to analyze the behavior of the functions uju_{j} at infinity. This can be done easily by an obvious identity

uj(xd,k)eτKj(k)xd2τKj(k)e±τKj(k)ydfj(yd)𝑑yd=u~j±(xd,k),±xd>0,\displaystyle u_{j}(x_{d},k)-\frac{e^{\mp\tau K_{j}(k)x_{d}}}{2\tau K_{j}(k)}\int\limits_{\mathds{R}}e^{\pm\tau K_{j}(k)y_{d}}f_{j}(y_{d})\,dy_{d}=\tilde{u}_{j}^{\pm}(x_{d},k),\qquad\pm x_{d}>0, (4.11)
u~j±(xd,k):=12Kj(k)xd±(eKj(k)(xdyd)eKj(k)(xdyd))fj(yd)𝑑yd,\displaystyle\tilde{u}_{j}^{\pm}(x_{d},k):=\frac{1}{2K_{j}(k)}\int\limits_{x_{d}}^{\pm\infty}\Big{(}e^{K_{j}(k)(x_{d}-y_{d})}-e^{-K_{j}(k)(x_{d}-y_{d})}\Big{)}f_{j}(y_{d})\,dy_{d},

and an estimate

|u~j±(xd,k)|\displaystyle|\tilde{u}_{j}^{\pm}(x_{d},k)|\leqslant 12|Kj(k)|(xd±|eKj(k)(xdyd)eKj(k)(xdyd)|2ea|yd|dyd,)12fjL2(,eϑ|xd|dxd)\displaystyle\frac{1}{2|K_{j}(k)|}\left(\int\limits_{x_{d}}^{\pm\infty}\Big{|}e^{K_{j}(k)(x_{d}-y_{d})}-e^{-K_{j}(k)(x_{d}-y_{d})}\Big{|}^{2}e^{-a|y_{d}|}\,dy_{d},\right)^{\frac{1}{2}}\|f_{j}\|_{L_{2}(\mathds{R},e^{{\vartheta}|x_{d}|}dx_{d})} (4.12)
\displaystyle\leqslant C(k)eϑ~|yd|fjL2(,eϑ|xd|dxd),±xd>0,\displaystyle C(k)e^{-{\tilde{\vartheta}}|y_{d}|}\|f_{j}\|_{L_{2}(\mathds{R},e^{{\vartheta}|x_{d}|}dx_{d})},\qquad\pm x_{d}>0,

where C(k)>0C(k)>0 and 0<ϑ~<ϑ0<{\tilde{\vartheta}<\vartheta} are some constants independent of xdx_{d} and ff. The proof is complete. ∎

4.2 Meromorphic continuation for the resolvent of the perturbed operator

This section is devoted to the proof of Theorem 2.1. We begin with rewriting the operator equation

(εΛp+k2)uε=f,(\mathcal{H}_{\varepsilon}-\Lambda_{p}+k^{2})u_{\varepsilon}=f,

as a boundary value problem (2.3). Then we denote

gε:=fε(ε)uεg_{\varepsilon}:=f-\varepsilon\mathcal{L}(\varepsilon)u_{\varepsilon}

and rewrite (2.3) as problem (4.1) with the function ff replaced by gεg_{\varepsilon}. According Lemma 4.1, such problem is solvable and there exists a solution given by formula (4.2):

uε=𝒜1,τ(k)gε.u_{\varepsilon}=\mathcal{A}_{1,\tau}(k)g_{\varepsilon}. (4.13)

We substitute this formula into corresponding boundary value problem (2.3) and this leads to an operator equation in the space L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx):

gε+ε(ε)𝒜1,τ(k)gε=f.g_{\varepsilon}+\varepsilon\mathcal{L}(\varepsilon)\mathcal{A}_{1,\tau}(k)g_{\varepsilon}=f.

Then we substitute representation (4.3), (4.4), (2.10) into this equation:

gε+εkj=pp+n1(jgε)(ε)ψj+ε(ε)(𝒢p,τ+k𝒜4,τ(k))gε=f.g_{\varepsilon}+\frac{\varepsilon}{k}\sum\limits_{j=p}^{p+n-1}\big{(}\ell_{j}g_{\varepsilon}\big{)}\mathcal{L}(\varepsilon)\psi_{j}+\varepsilon\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}g_{\varepsilon}=f. (4.14)

Thanks to the properties of the operators 𝒢p,τ\mathcal{G}_{p,\tau} and 𝒜4,τ(k)\mathcal{A}_{4,\tau}(k) described in Lemma 4.1, the operator (ε)(𝒢p,τ+k𝒜4,τ(k))\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)} is bounded uniformly in ε\varepsilon and kk as an operator in L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) and is holomorphic in kBδk\in B_{\delta}. Hence, for sufficiently small ε\varepsilon, the operator (+ε(ε)(𝒢p,τ+k𝒜4,τ(k)))(\mathcal{I}+\varepsilon\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}) is boundedly invertible and the inverse is holomorphic in kBδk\in B_{\delta} as an operator in L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{\vartheta|x_{d}|}dx); hereinafter the symbol \mathcal{I} stands for the identity mapping. We denote

𝒜5,τε(k):=(+ε(ε)(𝒢p,τ+k𝒜4,τ(k)))1\mathcal{A}_{5,\tau}^{\varepsilon}(k):=\big{(}\mathcal{I}+\varepsilon\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}\big{)}^{-1} (4.15)

and apply this operator to equation (4.14). This gives:

gε+εkj=pp+n1(jgε)𝒜5,τε(k)(ε)ψj=𝒜5,τε(k)f.g_{\varepsilon}+\frac{\varepsilon}{k}\sum\limits_{j=p}^{p+n-1}\big{(}\ell_{j}g_{\varepsilon}\big{)}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{j}=\mathcal{A}_{5,\tau}^{\varepsilon}(k)f. (4.16)

Our next step is to find jgε\ell_{j}g_{\varepsilon}. Once we do this, we shall be able to find gεg_{\varepsilon} from equation (4.14) as

gε=εkj=pp+n1(jgε)𝒜5,τε(k)(ε)ψj+𝒜5,τε(k)fg_{\varepsilon}=-\frac{\varepsilon}{k}\sum\limits_{j=p}^{p+n-1}\big{(}\ell_{j}g_{\varepsilon}\big{)}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{j}+\mathcal{A}_{5,\tau}^{\varepsilon}(k)f (4.17)

and recover the function uεu_{\varepsilon} by formula (4.13).

We apply the functionals i\ell_{i} to equation (4.16) with all ii such that Λi=Λp\Lambda_{i}=\Lambda_{p}. As a result, we arrive to a system of linear equations for the vector gε:=(jgε)j=p,,p+n1\ell g_{\varepsilon}:=(\ell_{j}g_{\varepsilon})_{j=p,\ldots,p+n-1}; this is a vector-column. The system reads as

(E+εkMε,τ(k))gε=Fε(k).\left(\mathrm{E}+\frac{\varepsilon}{k}\mathrm{M}_{\varepsilon,\tau}(k)\right)\ell g_{\varepsilon}=F_{\varepsilon}(k). (4.18)

Here E\mathrm{E} is the unit matrix and Mε,τ(k)\mathrm{M}_{\varepsilon,\tau}(k) is a square matrix with entries Mεij(k):=i𝒜5,τε(k)(ε)ψjM_{\varepsilon}^{ij}(k):=\ell_{i}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{j}, where ii counts the rows and jj does the columns in the matrix Mε,τ(k)\mathrm{M}_{\varepsilon,\tau}(k), while the symbol Fε(k)F_{\varepsilon}(k) denotes a vector column with coordinates i𝒜5,τε(k)f\ell_{i}\mathcal{A}_{5,\tau}^{\varepsilon}(k)f, i=p,,p+n1i=p,\ldots,p+n-1. As it was shown in the proof of Lemma 4.1, the functionals i:L2(Ω,eϑ|xd|dx)\ell_{i}:L_{2}(\Omega,e^{\vartheta|x_{d}|}dx)\to\mathds{C} are bounded and since the operator 𝒜5,τε(k)\mathcal{A}_{5,\tau}^{\varepsilon}(k) is a bounded one in L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{\vartheta|x_{d}|}dx), the introduced matrix Mε,τ(k)\mathrm{M}_{\varepsilon,\tau}(k) and vector Fε(k)F_{\varepsilon}(k) are well-defined.

In view of the holomorphy of the operator 𝒜5,τε\mathcal{A}_{5,\tau}^{\varepsilon}, the entries of the matrix Mε,τ(k)\mathrm{M}_{\varepsilon,\tau}(k) are holomorphic in kk and this implies that the matrix Mε,τ(k)\mathrm{M}_{\varepsilon,\tau}(k) is holomorphic in kBδk\in B_{\delta}. Hence, the determinant of the matrix kE+εMε,τ(k)k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k) is a holomorphic in kk function and therefore, the matrix (kE+εMε,τ(k))1(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k))^{-1} is well-defined as a meromorphic in kk matrix. This allows us to solve system (4.18):

gε=k(kE+εMε,τ(k))1Fε(k)=k(jε(k)f)j=p,,p+n1,jgε=kjε(k)f,\ell g_{\varepsilon}=k(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k))^{-1}F_{\varepsilon}(k)=k\big{(}\ell_{j}^{\varepsilon}(k)f\big{)}_{j=p,\ldots,p+n-1},\qquad\ell_{j}g_{\varepsilon}=k\ell_{j}^{\varepsilon}(k)f, (4.19)

where jε(k)\ell_{j}^{\varepsilon}(k) are functionals on L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}dx) meromorphic in kBδk\in B_{\delta}. Substituting these formulae into (4.17), we can find gεg_{\varepsilon} and determine uεu_{\varepsilon} by formula (4.13):

uε=\displaystyle u_{\varepsilon}= j=pp+n1(jεf)(ψjε(𝒢p,τ+k𝒜4,τ(k))𝒜5,τε(k)(ε)ψj)+(𝒢p,τ+k𝒜4,τ(k))𝒜5,τε(k)f\displaystyle\sum\limits_{j=p}^{p+n-1}\big{(}\ell_{j}^{\varepsilon}f\big{)}\Big{(}\psi_{j}-\varepsilon\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{j}\Big{)}+\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}\mathcal{A}_{5,\tau}^{\varepsilon}(k)f (4.20)
=\displaystyle= j=pp+n1(jεf)(+ε(𝒢p,τ+k𝒜4,τ(k))(ε))1ψj+(𝒢p,τ+k𝒜4,τ(k))𝒜5,τε(k)f=:ε,τ(k)f.\displaystyle\sum\limits_{j=p}^{p+n-1}\big{(}\ell_{j}^{\varepsilon}f\big{)}\big{(}\mathcal{I}+\varepsilon\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}\mathcal{L}(\varepsilon)\big{)}^{-1}\psi_{j}+\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)\big{)}\mathcal{A}_{5,\tau}^{\varepsilon}(k)f=:\mathcal{R}_{\varepsilon,\tau}(k)f.

The function uεu_{\varepsilon} solves problem (2.3) and satisfies representation (2.4), (2.5), (2.6). The latter are implied by similar representations (4.5), (4.6), (4.7) and formula (4.13). Since for sufficiently small kk the functions KjK_{j}, Λj<Λp\Lambda_{j}<\Lambda_{p}, satisfy asymptotic identities (2.43), the functions eτKj(k)|xd|e^{-\tau K_{j}(k)|x_{d}|}, Λj<Λp\Lambda_{j}<\Lambda_{p} and ek|xd|e^{-k|x_{d}|} decay exponentially at infinity as Rek>0\operatorname{Re}k>0 and τImk2<0\tau\operatorname{Im}k^{2}<0. Hence, under the same conditions, we have uε=(εΛp+k2)1fu_{\varepsilon}=(\mathcal{H}_{\varepsilon}-\Lambda_{p}+k^{2})^{-1}f and this means that the operator ε,τ\mathcal{R}_{\varepsilon,\tau} provides a meromorphic continuation of the resolvent (εΛp+k2)1(\mathcal{H}_{\varepsilon}-\Lambda_{p}+k^{2})^{-1}.

We observe that the introduced operator ε,τ\mathcal{R}_{\varepsilon,\tau} is meromorphic in kBδk\in B_{\delta} as an operator from L2(Ω,eϑ|xd|dx)L_{2}(\Omega,e^{{\vartheta}|x_{d}|}\,dx) into W22(Ω)L2(Ω,eϑ|xd|dx)W_{2}^{2}(\Omega)\cap L_{2}(\Omega,e^{-{\vartheta}|x_{d}|}\,dx) due to formulae (4.13), (4.17), (4.19) and representation (4.3), (4.4), (2.10).

According to formula (4.20), the poles of the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) defined in this formula coincide with the poles of the functionals jε(k)\ell_{j}^{\varepsilon}(k). Let at least one of these functionals have a pole at a point kεBδk_{\varepsilon}\in B_{\delta}. In view of the definition of the functionals in (4.19), this means that the matrix (kεE+εMε,τ(kε))(k_{\varepsilon}\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k_{\varepsilon})) is degenerate and by the Cramer’s rule, system (4.18) with Fε(k)=0F_{\varepsilon}(k)=0, k=kεk=k_{\varepsilon} has a non-trivial solution, which we denote by lε=(ljε)j:Λj=Λpl^{\varepsilon}=(l^{\varepsilon}_{j})_{j:\,\Lambda_{j}=\Lambda_{p}}. We observe that by the definition,

kεliε=εj=pp+n1ljεi𝒜5,τε(k)(ε)ψj.k_{\varepsilon}l_{i}^{\varepsilon}=-\varepsilon\sum\limits_{j=p}^{p+n-1}l_{j}^{\varepsilon}\ell_{i}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{j}. (4.21)

Employing these identities and (4.3), (4.4), it is straightforward to check that the formula

ψε:=\displaystyle\psi_{\varepsilon}:= εlimkkε𝒜1,τ(k)i=1nliε𝒜5,τε(k)(ε)ψi+p1\displaystyle-\varepsilon\lim\limits_{k\to k_{\varepsilon}}\mathcal{A}_{1,\tau}(k)\sum\limits_{i=1}^{n}l_{i}^{\varepsilon}\mathcal{A}_{5,\tau}^{\varepsilon}(k)\mathcal{L}(\varepsilon)\psi_{i+p-1} (4.22)
=\displaystyle= i=1nliεψi+p1ε(𝒢p,τ+kε𝒜4,τ(kε))i=1nliε𝒜5,τε(kε)(ε)ψi+p1\displaystyle\sum\limits_{i=1}^{n}l_{i}^{\varepsilon}\psi_{i+p-1}-\varepsilon\big{(}\mathcal{G}_{p,\tau}+k_{\varepsilon}\mathcal{A}_{4,\tau}(k_{\varepsilon})\big{)}\sum\limits_{i=1}^{n}l_{i}^{\varepsilon}\mathcal{A}_{5,\tau}^{\varepsilon}(k_{\varepsilon})\mathcal{L}(\varepsilon)\psi_{i+p-1}
=\displaystyle= i=1nliε(+ε(𝒢p,τ+kε𝒜4,τ(kε))(ε))1ψi+p1\displaystyle\sum\limits_{i=1}^{n}l_{i}^{\varepsilon}\big{(}\mathcal{I}+\varepsilon\big{(}\mathcal{G}_{p,\tau}+k_{\varepsilon}\mathcal{A}_{4,\tau}(k_{\varepsilon})\big{)}\mathcal{L}(\varepsilon)\big{)}^{-1}\psi_{i+p-1}

defines a non-trivial solution to problem (2.3) with f=0f=0, k=kεk=k_{\varepsilon}. Thanks to formulae (4.9), (4.8), (4.12), this function can be also represented as

ψε=j=1mψj(x)φj(xd)+(Λp+kε2)1f,\displaystyle\psi_{\varepsilon}=\sum\limits_{j=1}^{m}\psi_{j}(x^{\prime})\varphi_{j}(x_{d})+(\mathcal{H}^{\bot}-\Lambda_{p}+k_{\varepsilon}^{2})^{-1}f^{\bot}, (4.23)
φj(xd):=12τKj(kε)ΩeτKj(kε)|xdyd|f(y)ψj(y)¯𝑑y\displaystyle\varphi_{j}(x_{d}):=\frac{1}{2\tau K_{j}(k_{\varepsilon})}\int\limits_{\Omega}e^{-\tau K_{j}(k_{\varepsilon})|x_{d}-y_{d}|}f(y)\overline{\psi_{j}(y^{\prime})}\,dy asj<p1,\displaystyle\text{as}\quad j<p-1,
φj(xd):=12Kj(kε)ΩeKj(kε)|xdyd|f(y)ψj(y)¯𝑑y\displaystyle\varphi_{j}(x_{d}):=\frac{1}{2K_{j}(k_{\varepsilon})}\int\limits_{\Omega}e^{-K_{j}(k_{\varepsilon})|x_{d}-y_{d}|}f(y)\overline{\psi_{j}(y^{\prime})}\,dy asjp+n1,\displaystyle\text{as}\quad j\geqslant p+n-1,
φj(xd):=12kεΩekε|xdyd|f(y)ψj(y)¯𝑑y\displaystyle\varphi_{j}(x_{d}):=\frac{1}{2k_{\varepsilon}}\int\limits_{\Omega}e^{-k_{\varepsilon}|x_{d}-y_{d}|}f(y)\overline{\psi_{j}(y^{\prime})}\,dy asj=p,,p+n1,kε0,\displaystyle\text{as}\quad j=p,\ldots,p+n-1,\quad k_{\varepsilon}\neq 0,
φj(xd):=12Ω|xdyd|f(y)ψj(y)¯𝑑y\displaystyle\varphi_{j}(x_{d}):=-\frac{1}{2}\int\limits_{\Omega}|x_{d}-y_{d}|f(y)\overline{\psi_{j}(y^{\prime})}\,dy asj=p,,p+n1,kε=0,\displaystyle\text{as}\quad j=p,\ldots,p+n-1,\quad k_{\varepsilon}=0,

with

f:=εi=1nliε𝒜5,τε(kε)(ε)ψi+p1f:=-\varepsilon\sum\limits_{i=1}^{n}l_{i}^{\varepsilon}\mathcal{A}_{5,\tau}^{\varepsilon}(k_{\varepsilon})\mathcal{L}(\varepsilon)\psi_{i+p-1}

and ff^{\bot} defined by (2.11). We also observe that as kε=0k_{\varepsilon}=0, it follows from (4.21) that

(f(,xd),ψj)L2(Ω)=0(f(\cdot,x_{d}),\psi_{j})_{L_{2}(\Omega)}=0

for almost each xdI±x_{d}\in I_{\pm} and for each j=1,,mj=1,\ldots,m. Representations (4.22), (4.23) and relations (4.11), (4.12) imply representation (2.7), (2.8), (2.9). This completes the proof of Theorem 2.1.

5 Poles of meromorphic continuations

In this section we study the asymptotic behaviour of poles of the operators ε,τ\mathcal{R}_{\varepsilon,\tau} and we prove Theorems 2.22.3. As it was shown in the proof of Theorem 2.1 in the previous section, the poles of the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) coincides with poles of the matrix (kE+εMε,τ(k))1(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k))^{-1}. The orders of the mentioned poles of the operator ε,τ(k)\mathcal{R}_{\varepsilon,\tau}(k) and the orders of the zeroes of det(kE+εMε,τ(k))\det(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k)) obviously coincide.

5.1 Proof of Theorem 2.2

Let us study the solvability of the equation

det(kE+εMε,τ(k))=0\det(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k))=0 (5.1)

in BδB_{\delta}. It is straightforward to check that

det(kE+εMε,τ(k))=kn+P(k,ε),P(k,ε):=i=1nεikniPi(k,ε),\det(k\mathrm{E}+\varepsilon\mathrm{M}_{\varepsilon,\tau}(k))=k^{n}+P(k,\varepsilon),\qquad P(k,\varepsilon):=\sum\limits_{i=1}^{n}\varepsilon^{i}k^{n-i}P_{i}(k,\varepsilon),

where Pi(k,ε)P_{i}(k,\varepsilon) are some functions holomorphic in kBδk\in B_{\delta} and uniformly bounded in kBδ¯k\in\overline{B_{\delta}} and sufficiently small ε\varepsilon. We fix arbitrary δ(0,δ)\delta^{\prime}\in(0,\delta) and for sufficiently small ε\varepsilon we have an uniform estimate:

|P(k,ε)|CεonBδ.|P(k,\varepsilon)|\leqslant C\varepsilon\quad\text{on}\quad\partial B_{\delta^{\prime}}. (5.2)

Since |kn|=(δ)n|k^{n}|=(\delta^{\prime})^{n} on Bδ\partial B_{\delta^{\prime}} and the function kknk\mapsto k^{n} has the only zero in BδB_{\delta} of order nn at the origin, estimate (5.2) allows us to apply the Rouché theorem and to conclude that equation (5.1) has exactly nn zeroes counting their orders and these zeroes converge to the origin as ε+0\varepsilon\to+0.

Our next step is to show that all zeroes of equation (5.1) are located in a ball BbεB_{b\varepsilon} for some fixed bb. For arbitrary bb and sufficiently small ε\varepsilon such that bε<db\varepsilon<d, by the definition of the function PP we have:

|P(k,ε)|Cεn(1+b++bn1),|k|n=εnbnonBbε.|P(k,\varepsilon)|\leqslant C\varepsilon^{n}(1+b+\ldots+b^{n-1}),\qquad|k|^{n}=\varepsilon^{n}b^{n}\quad\text{on}\quad\partial B_{b\varepsilon}.

Then we choose bb large enough and ε\varepsilon small enough so that

bn>C(1+b++bn1),bε<δ,b^{n}>C(1+b+\ldots+b^{n-1}),\qquad b\varepsilon<\delta,

and applying the Rouché theorem once again, we see that the zeroes of equation (5.1) are located inside the ball BbεB_{b\varepsilon}. This fact allows us to seek the zeroes of equation (5.1) as k=zεk=z\varepsilon, where |z|<b|z|<b for all sufficiently small ε\varepsilon.

The operator 𝒜5,τε\mathcal{A}_{5,\tau}^{\varepsilon} defined in (4.15) can be expanded into the standard Neumann series and this yields:

𝒜5,τε(k)=ε(ε)(𝒢p,τ+k𝒜4,τ(k))+ε2((ε)(𝒢p,τ+k𝒜4,τ(k)))2𝒜5,τε(k).\mathcal{A}_{5,\tau}^{\varepsilon}(k)=\mathcal{I}-\varepsilon\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k))+\varepsilon^{2}(\mathcal{L}(\varepsilon)\big{(}\mathcal{G}_{p,\tau}+k\mathcal{A}_{4,\tau}(k)))^{2}\mathcal{A}_{5,\tau}^{\varepsilon}(k). (5.3)

In view of the above identity, the change k=zεk=z\varepsilon, |z|<b|z|<b, and the definition of the matrix Mε,τ\mathrm{M}_{{\varepsilon,\tau}}, we can represent the latter as

Mε,τ(k)=M1+εM~2(z,ε),\mathrm{M}_{\varepsilon,\tau}(k)=-\mathrm{M}_{1}+\varepsilon\tilde{\mathrm{M}}_{2}(z,\varepsilon), (5.4)

where M~2(z,ε)\tilde{\mathrm{M}}_{2}(z,\varepsilon) is some uniformly bounded matrix holomorphic in zz and we recall that the matrix M1\mathrm{M}_{1} is defined in (2.12). Both these matrices are well-defined since they arise as the terms in the expansion for Mε,τ\mathrm{M}_{\varepsilon,\tau}.

Then we can rewrite equation (5.1) as

det(zEM1+εM~2(z,ε))=0.\det\big{(}z\mathrm{E}-\mathrm{M}_{1}+\varepsilon\tilde{\mathrm{M}}_{2}(z,\varepsilon)\big{)}=0. (5.5)

We recall that μj\mu_{j}, j=1,,Nj=1,\ldots,N, are different eigenvalues of the matrix M1\mathrm{M}_{1} of the multiplicities qjq_{j} and hence,

det(zEM1)=j=1N(zμj)qj.\det(z\mathrm{E}-\mathrm{M}_{1})=\prod\limits_{j=1}^{N}(z-\mu_{j})^{q_{j}}.

This allows us to represent equation (5.5) as

i=1N(zμi)qi+εP~(z,ε)=0,\prod\limits_{i=1}^{N}(z-\mu_{i})^{q_{i}}+\varepsilon\tilde{P}(z,\varepsilon)=0, (5.6)

where P~\tilde{P} is some function holomorphic in zz and bounded uniformly in zz and ε\varepsilon:

|P~(z,ε)|C.|\tilde{P}(z,\varepsilon)|\leqslant C. (5.7)

By Tr(z0)T_{r}(z_{0}) we denote a ball of radius rr in the complex plane centered at the point z0z_{0}. We introduce the balls Tb~ε1qi(μi)T_{\tilde{b}\varepsilon^{\frac{1}{q_{i}}}}(\mu_{i}), where b~\tilde{b} is some fixed number. Then for sufficiently small ε\varepsilon, each ball Tb~ε1qi(μi)T_{\tilde{b}\varepsilon^{\frac{1}{q_{i}}}}(\mu_{i}) contains the only zero of the function zj=1N(zμj)qjz\mapsto\prod\limits_{j=1}^{N}(z-\mu_{j})^{q_{j}}, this zero is μi\mu_{i} and the order of this zero is qiq_{i}. On the boundary of each ball Tb~ε1qi(μi)T_{\tilde{b}\varepsilon^{\frac{1}{q_{i}}}}(\mu_{i}), for sufficiently small ε\varepsilon, an obvious estimate holds true:

|j=1N(zμj)qj|Cb~qiε,\left|\prod\limits_{j=1}^{N}(z-\mu_{j})^{q_{j}}\right|\geqslant C\tilde{b}^{q_{i}}\varepsilon, (5.8)

where CC is some fixed constant independent of ε\varepsilon and b~\tilde{b}. Hence, in view of estimate (5.7), we can choose b~\tilde{b} large enough and ε\varepsilon small enough and apply the Rouché theorem to the left hand side of equation (5.6). This yields that each ball Tb~ε1qi(μi)T_{\tilde{b}\varepsilon^{\frac{1}{q_{i}}}}(\mu_{i}) contains exactly qiq_{i} zeroes of equation (5.6), which we denote by zij=zij(ε)z_{ij}=z_{ij}(\varepsilon). These zeroes satisfy the asymptotic identities

zij(ε)=μi+O(ε1qi),j=1,,qi.z_{ij}(\varepsilon)=\mu_{i}+O(\varepsilon^{\frac{1}{q_{i}}}),\qquad j=1,\ldots,q_{i}. (5.9)

Returning back to equation (5.1), we see that it has exactly nn zeroes counting their orders and these zeroes obey asymptotic expansion (2.13). The proof is complete.

5.2 Proof of Theorem 2.3

Identity (5.3) allows us to specify the form of the matrix Mε,τ\mathrm{M}_{\varepsilon,\tau} in more details than in (5.4). Namely, this identity implies that

Mε,τ(k)=M1+εM2τ+ε2M~3(z,ε),\mathrm{M}_{\varepsilon,\tau}(k)=-\mathrm{M}_{1}+\varepsilon\mathrm{M}_{2}^{\tau}+\varepsilon^{2}\tilde{\mathrm{M}}_{3}(z,\varepsilon),

where M~3(z,ε)\tilde{\mathrm{M}}_{3}(z,\varepsilon) is some uniformly bounded matrix holomorphic in zz and we recall that the matrix M2\mathrm{M}_{2} is defined in (2.14). All these matrices are well-defined as the terms in the expansion for Mε,τ\mathrm{M}_{\varepsilon,\tau}. Equation (5.5) is replaced by a more detailed one:

det(zEM1+εM2τ+ε2M~3(z,ε))=0.\det\big{(}z\mathrm{E}-\mathrm{M}_{1}+\varepsilon\mathrm{M}_{2}^{\tau}+\varepsilon^{2}\tilde{\mathrm{M}}_{3}(z,\varepsilon)\big{)}=0. (5.10)

We fix i{1,,N}i\in\{1,\ldots,N\} and we are going to study the asymptotic behavior of the zeroes zij(ε)z_{ij}(\varepsilon), j=1,,qij=1,\ldots,q_{i}. Let S\mathrm{S} be a matrix reducing the matrix M1\mathrm{M}_{1} to its Jordan canonical form, which we denote by J:=SM1S1\mathrm{J}:=\mathrm{S}\mathrm{M}_{1}\mathrm{S}^{-1}. Then equation (5.10) can be rewritten as

det(zEJ+εSM2τS1+ε2SM~3(z,ε)S1)=0.\det\big{(}z\mathrm{E}-\mathrm{J}+\varepsilon\mathrm{S}\mathrm{M}_{2}^{\tau}\mathrm{S}^{-1}+\varepsilon^{2}\mathrm{S}\tilde{\mathrm{M}}_{3}(z,\varepsilon)\mathrm{S}^{-1}\big{)}=0.

Taking into consideration the structure of the matrix J\mathrm{J}, we rewrite this equation as follows:

j=1N(zμj)qj+ε(zμi)ri,τY1(z)+ε2Y2(z,ε)=0,\prod\limits_{j=1}^{N}(z-\mu_{j})^{q_{j}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}Y_{1}(z)+\varepsilon^{2}Y_{2}(z,\varepsilon)=0, (5.11)

where Y1Y_{1}, Y2Y_{2} are some functions holomorphic in zz and bounded uniformly in ε\varepsilon and zz and ri,τ<qir_{i,\tau}<q_{i} is some non-negative integer number. These functions arise as some polynomials in zz, ε\varepsilon and the entries of the matrices SM2S1\mathrm{S}\mathrm{M}_{2}\mathrm{S}^{-1} and SM~3(z,ε)S1\mathrm{S}\tilde{\mathrm{M}}_{3}(z,\varepsilon)\mathrm{S}^{-1}. It is clear that (zμi)ri,τY1(z)=Qi,τ(z)(z-\mu_{i})^{r_{i,\tau}}Y_{1}(z)=Q_{i,\tau}(z), where QiQ_{i} is defined in (2.15). And if Y1Y_{1} is not identically zero, then

Y1(zi)=γi,τj=1jiN(μiμj)qj0,Y_{1}(z_{i})=\gamma_{i,\tau}\prod\limits_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{N}(\mu_{i}-\mu_{j})^{q_{j}}\neq 0, (5.12)

where γi,τ\gamma_{i,\tau} is defined in (2.17). By Tr(z0)T_{r}(z_{0}) we denote a ball of radius rr in the complex plane centered at the point z0z_{0}.

We first consider the case, when Y1Y_{1} vanishes identically. Then the term εY1\varepsilon Y_{1} disappears in equation (5.11). Here we consider the ball Tb~ε2qi(μi)T_{\tilde{b}\varepsilon^{\frac{2}{q_{i}}}}(\mu_{i}) and for sufficiently small ε\varepsilon, on the boundary of this ball we have the estimates

|j=1N(zμj)qj|Cb~qiε2,|ε2Y2(z,ε)|C~ε2,\Big{|}\prod\limits_{j=1}^{N}(z-\mu_{j})^{q_{j}}\Big{|}\geqslant C\tilde{b}^{q_{i}}\varepsilon^{2},\qquad|\varepsilon^{2}Y_{2}(z,\varepsilon)|\leqslant\tilde{C}\varepsilon^{2},

where CC and C~\tilde{C} are some fixed constants independent of ε\varepsilon and b~\tilde{b}. Then we apply the Rouché theorem proceeding as in (5.6), (5.7), (5.8), (5.9) and we arrive immediately to (2.16).

Now we proceed to the case, when Y1Y_{1} is not identically zero. Here we divide equation (5.11) by j=1jiN(zμj)qj\prod\limits_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{N}(z-\mu_{j})^{q_{j}} and in view of (5.12) we get:

(zμi)qi+ε(zμi)ri,τγi,τ+ε(zμi)ri,τ+1Y3(z)+ε2Y4(z,ε)=0,(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau}+\varepsilon(z-\mu_{i})^{r_{i,\tau}+1}Y_{3}(z)+\varepsilon^{2}Y_{4}(z,\varepsilon)=0, (5.13)

where Y3Y_{3}, Y3Y_{3} are some functions holomorphic in zz in the vicinity of the point μi\mu_{i} and bounded uniformly in ε\varepsilon and zz.

Suppose that qi2ri,τq_{i}\leqslant 2r_{i,\tau} and consider the ball Tb~ε1ri,τ(μi)T_{\tilde{b}\varepsilon^{\frac{1}{r_{i,\tau}}}}(\mu_{i}). For sufficiently small ε\varepsilon, on the boundary of this ball we have

|(zμi)qi+ε(zμi)ri,τγi,τ|\displaystyle|(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau}|\geqslant |(zμi)qi|ε|γi,τ||(zμi)ri,τ|\displaystyle|(z-\mu_{i})^{q_{i}}|-\varepsilon|\gamma_{i,\tau}||(z-\mu_{i})^{r_{i,\tau}}| (5.14)
=\displaystyle= b~qiε22ri,τqiri,τε2|γi,τ|b~ri,τb~qi2ε22ri,τqiri,τ\displaystyle\tilde{b}^{q_{i}}\varepsilon^{2-\frac{2r_{i,\tau}-q_{i}}{r_{i,\tau}}}-\varepsilon^{2}|\gamma_{i,\tau}|\tilde{b}^{r_{i,\tau}}\geqslant\frac{\tilde{b}^{q_{i}}}{2}\varepsilon^{2-\frac{2r_{i,\tau}-q_{i}}{r_{i,\tau}}}

and

|ε(zμi)ri,τ+1Y3(z)+ε2Y4(z,ε)|C(b~ri,τ+1ε2+1ri,τ+ε2)<Cε2,|\varepsilon(z-\mu_{i})^{r_{i,\tau}+1}Y_{3}(z)+\varepsilon^{2}Y_{4}(z,\varepsilon)|\leqslant C(\tilde{b}^{r_{i,\tau}+1}\varepsilon^{2+\frac{1}{r_{i,\tau}}}+\varepsilon^{2})<C\varepsilon^{2}, (5.15)

where CC is some constant independent of ε\varepsilon and b~\tilde{b}. Since qiri,τri,τ+1q_{i}-r_{i,\tau}\geqslant r_{i,\tau}+1, it is also straightforward to check that for sufficiently small ε\varepsilon and sufficiently large b~\tilde{b}, all zeroes of the function z(zμi)qi+ε(zμi)ri,τγi,τz\mapsto(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau} belong to the ball Tb~ε1ri,τ(μi)T_{\tilde{b}\varepsilon^{\frac{1}{r_{i,\tau}}}}(\mu_{i}). Taking this fact and estimates (5.14), (5.15) into consideration, we apply the Rouché theorem and arrive at asymptotic expansion (2.18).

Suppose that qi2ri,τ1q_{i}-2r_{i,\tau}\geqslant 1. The zeroes of the function z(zμi)qi+ε(zμi)ri,τγi,τz\mapsto(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau} can be found explicitly. The first of them is a ri,τr_{i,\tau}-multiple zero z0=μiz_{0}=\mu_{i} and qiri,τq_{i}-r_{i,\tau} simple zeroes

zj(ε):=μiε1qiri,τ(γi,τ)1qiri,τe2πiqiri,τj,j=1,,qiri,τ.z_{j}(\varepsilon):=\mu_{i}-\varepsilon^{\frac{1}{q_{i}-r_{i,\tau}}}(-\gamma_{i,\tau})^{\frac{1}{q_{i}-r_{i,\tau}}}e^{\frac{2\pi\mathrm{i}}{q_{i}-r_{i,\tau}}j},\qquad j=1,\ldots,q_{i}-r_{i,\tau}. (5.16)

First we consider a ball Tb~ε1ri,τ(μi)T_{\tilde{b}\varepsilon^{\frac{1}{r_{i,\tau}}}}(\mu_{i}) and we observe that under our assumption, qiri,τri,τ+1q_{i}-r_{i,\tau}\geqslant r_{i,\tau}+1. Hence, for sufficiently small ε\varepsilon, the ball Tb~ε1ri,τ(μi)T_{\tilde{b}\varepsilon^{\frac{1}{r_{i,\tau}}}}(\mu_{i}) does not contain the zeroes zj(ε)z_{j}(\varepsilon). Moreover, on the boundary of this ball, the estimate holds true:

|(zμi)qi+ε(zμi)ri,τγi,τ|\displaystyle|(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau}|\geqslant ε|γi,τ||(zμi)ri,τ||(zμi)qi|\displaystyle\varepsilon|\gamma_{i,\tau}||(z-\mu_{i})^{r_{i,\tau}}|-|(z-\mu_{i})^{q_{i}}|
=\displaystyle= b~ri,τε2|γi,τ|b~qiε2+qi2ri,τri,τ>|γi,τ|2b~ri,τε2.\displaystyle\tilde{b}^{r_{i,\tau}}\varepsilon^{2}|\gamma_{i,\tau}|-\tilde{b}^{q_{i}}\varepsilon^{2+\frac{q_{i}-2r_{i,\tau}}{r_{i,\tau}}}>\frac{|\gamma_{i,\tau}|}{2}\tilde{b}^{r_{i,\tau}}\varepsilon^{2}.

We also observe that estimate (5.15) remains true in the considered case. Then we apply Rouché theorem to the ball Tb~ε1ri,τ(μi)T_{\tilde{b}\varepsilon^{\frac{1}{r_{i,\tau}}}}(\mu_{i}) and conclude that this ball contains exactly ri,τr_{i,\tau} zeroes of equation (5.13) counting their orders. This implies asymptotic expansion (2.19).

Now we consider another ball Tb~ε2qiri,τ(zj(ε))T_{\tilde{b}\varepsilon^{\frac{2}{q_{i}-r_{i,\tau}}}}(z_{j}(\varepsilon)) for zjz_{j} defined in (5.16) with some fixed jj. For sufficiently small ε\varepsilon, this ball does not contain the origin and other points zsz_{s} with sjs\neq j. On the boundary of this ball, we have the following estimate:

|(zμi)qi+ε(zμi)ri,τγi,τ|\displaystyle|(z-\mu_{i})^{q_{i}}+\varepsilon(z-\mu_{i})^{r_{i,\tau}}\gamma_{i,\tau}|\geqslant Cεri,τqiri,τ|(zμi)qiri,τ+εγi,τ|\displaystyle C\varepsilon^{\frac{r_{i,\tau}}{q_{i}-r_{i,\tau}}}|(z-\mu_{i})^{q_{i}-r_{i,\tau}}+\varepsilon\gamma_{i,\tau}| (5.17)
=\displaystyle= Cεri,τqiri,τs=1qiri,τ|zzs|Cb~εqi+1qiri,τ,\displaystyle C\varepsilon^{\frac{r_{i,\tau}}{q_{i}-r_{i,\tau}}}\prod\limits_{s=1}^{q_{i}-r_{i,\tau}}|z-z_{s}|\geqslant C\tilde{b}\varepsilon^{\frac{q_{i}+1}{q_{i}-r_{i,\tau}}},

where CC is some constant independent of ε\varepsilon and b~\tilde{b}. We observe that under our assumptions

qi+1qiri,τ2.\frac{q_{i}+1}{q_{i}-r_{i,\tau}}\leqslant 2.

In view of the latter inequality, on the boundary of the ball Tb~ε2qiri,τ(zj(ε))T_{\tilde{b}\varepsilon^{\frac{2}{q_{i}-r_{i,\tau}}}}(z_{j}(\varepsilon)), the estimates

|ε(zμi)ri,τ+1Y3(z)+ε2Y4(z,ε)|C(εqi+1qiri,τ+ε2)<Cεqi+1qiri,τ|\varepsilon(z-\mu_{i})^{r_{i,\tau}+1}Y_{3}(z)+\varepsilon^{2}Y_{4}(z,\varepsilon)|\leqslant C(\varepsilon^{\frac{q_{i}+1}{q_{i}-r_{i,\tau}}}+\varepsilon^{2})<C\varepsilon^{\frac{q_{i}+1}{q_{i}-r_{i,\tau}}}

hold true, where CC is some constant independent of ε\varepsilon and b~\tilde{b}. This estimate and (5.17) allow us to apply the Rouché theorem and we conclude that the ball Tb~ε2qiri,τ(zj(ε))T_{\tilde{b}\varepsilon^{\frac{2}{q_{i}-r_{i,\tau}}}}(z_{j}(\varepsilon)) contains exactly one zero of equation (5.11). This gives asymptotic expansion (2.20). The proof is complete.

6 Emerging resonances and eigenvalues

In this section we determine whether the poles of the operators ε,τ\mathcal{R}_{\varepsilon,\tau} described in the previous section correspond to the eigenvalues or resonances of the operator ε\mathcal{H}_{\varepsilon}. In this way, we shall prove Theorems 2.42.5.

We fix p{1,,m}p\in\{1,\ldots,m\} and assume that Λp==Λp+n1\Lambda_{p}=\ldots=\Lambda_{p+n-1}, where n1n\geqslant 1 is a multiplicity of the eigenvalue Λp\Lambda_{p} of the operator \mathcal{H}^{\prime}. In what follows, we analyze the nature of poles kij(ε)k_{ij}(\varepsilon) from Theorem 2.2 with asymptotic behavior (2.13). In order to do this, we analyze the behavior of the associated nontrivial solutions to problem (2.3) with f=0f=0.

6.1 Bottom of the spectrum

In this subsection we assume that p=1p=1 and we prove Theorem 2.4. In the considered case, the meromorphic continuation of the resolvent given the operator ε,τ\mathcal{R}_{\varepsilon,\tau} is independent of τ\tau and the same is true for the corresponding poles kij(ε)k_{ij}(\varepsilon) from Theorem 2.2 with asymptotic behavior (2.13). The behavior of the associated nontrivial solutions to problem (2.3) with f=0f=0 is provided by representation (2.7). In this representation, the first sum over j=1,,p1j=1,\ldots,p-1 is obviously missing and the leading term at infinity is the sum over j=p,,p+n1j=p,\ldots,p+n-1.

We fix one of the poles kij(ε)k_{ij}(\varepsilon) with some i{1,,n}i\in\{1,\ldots,n\}, j=1,,qij=1,\ldots,q_{i}. It follows from (4.21), (4.22), (4.23), (4.3), (4.4) that the coefficients cs,ε±c_{s,\varepsilon}^{\pm}, s=1,,ns=1,\ldots,n, in representation (2.7) coincide with lsεl_{s}^{\varepsilon} determined by (4.21). Since we deal with a non-trivial solution of system (4.21), this implies that at least one of the coefficients cs,ε±c_{s,\varepsilon}^{\pm} is non-zero. Then representation (2.7) for a non-trivial solution to problem (2.3) defined in (4.22) and formulae (2.43) imply that if

Rekij(ε)>0,\operatorname{Re}k_{ij}(\varepsilon)>0, (6.1)

then a non-trivial solution ψε\psi_{\varepsilon} decays exponentially at infinity and thus, is an eigenfunction, while the opposite inequality

Rekij(ε)0\operatorname{Re}k_{ij}(\varepsilon)\leqslant 0 (6.2)

ensures that ψε\psi_{\varepsilon} is not in W22(Ω)W_{2}^{2}(\Omega).

Asymptotic expansions for kijk_{ij} established in Theorems 2.22.3 allow us to check effectively the above conditions. Namely, if Reμi>0\operatorname{Re}\mu_{i}>0, this ensures inequality (6.1) for sufficiently small ε\varepsilon. Then the poles kijk_{ij}, j=1,,qij=1,\ldots,q_{i} correspond to the eigenvalues λij(ε)=Λpkij2(ε)\lambda_{ij}(\varepsilon)=\Lambda_{p}-k_{ij}^{2}(\varepsilon) with asymptotic expansions (2.21), (2.22), (2.23) as described in the formulation of Theorem 2.4. If Reμi<0\operatorname{Re}\mu_{i}<0, this guarantees inequality (6.2) for sufficiently small ε\varepsilon and the poles kijk_{ij} correspond to the resonances λij(ε)=Λpkij2(ε)\lambda_{ij}(\varepsilon)=\Lambda_{p}-k_{ij}^{2}(\varepsilon) with same asymptotic expansions (2.21), (2.22), (2.23).

If Reμi=0\operatorname{Re}\mu_{i}=0, in order to understand which of conditions (6.1), (6.2) is realized, we need to check the next term in the asymptotic expansion for kijk_{ij}. According Theorem 2.3, this can be done under an additional assumption 2ri,τqi12r_{i,\tau}\leqslant q_{i}-1 and only for poles kij,τk_{ij,\tau} with j=ri+1,,qij=r_{i}+1,\ldots,q_{i}. Then asymptotic expansion (2.20) imply that under condition (2.24), the pole kijk_{ij}, j=ri,τ+1,,qij=r_{i,\tau}+1,\ldots,q_{i} corresponds to an eigenvalue, while under condition (2.25) the pole kijk_{ij} corresponds to a resonance. The asymptotic behavior for this eigenvalue/resonance is given by (2.23) if μi0\mu_{i}\neq 0 and it is given by (2.26) if μi=0\mu_{i}=0.

6.2 Internal thresholds in the spectrum

In this subsection we study the nature of the poles emerging from internal points Λp\Lambda_{p} in the essential spectrum and we prove Theorem 2.5. As in the previous subsection, here we again analyze the behavior at infinity of non-trivial solutions ψε\psi_{\varepsilon} associated with poles kijk_{ij} and this analysis will be based on representation (2.7). However, there are important differences. The matter is that now the representation involves also a sum over j=1,,p1j=1,\ldots,p-1. Its terms can decay or grow at infinity. As it has been mentioned in the proof of Theorem 2.1, according identity (2.43), the functions eτKj(k)|xd|e^{-\tau K_{j}(k)|x_{d}|}, j=1,,p1j=1,\ldots,p-1, decay exponentially at infinity if τImk2<0\tau\operatorname{Im}k^{2}<0 and grow exponentially or vary periodically as τImk20\tau\operatorname{Im}k^{2}\geqslant 0. This means that apart of the sign of Rekij\operatorname{Re}k_{ij}, we should also control the sign of τImkij2\tau\operatorname{Im}k_{ij}^{2}. One more point is that now we have two meromorphic continuations of the resolvent, the operators ε,τ\mathcal{R}_{\varepsilon,\tau}, τ{1,+1}\tau\in\{-1,+1\}, respectively, two sets of their poles kij,τ(ε)k_{ij,\tau}(\varepsilon) converging to zero as ε+0\varepsilon\to+0. We observe that according Theorems 2.22.3, the first terms in asymptotic expansions for poles kij,τk_{ij,\tau} are the eigenvalues μi\mu_{i} of the matrix M1\mathrm{M}_{1} and they are independent of τ\tau. In fact, we can see the influence of τ\tau on the asymptotic behavior of kijk_{ij} only in formulae (2.20), for j=ri,τ+1,,qij=r_{i,\tau}+1,\ldots,q_{i} under an additional assumption 2ri,τqi12r_{i,\tau}\leqslant q_{i}-1.

We choose i{1,,N}i\in\{1,\ldots,N\}, j{1,,qi}j\in\{1,\ldots,q_{i}\}, τ{1,+1}\tau\in\{-1,+1\} and consider the pole kij,τ(ε)k_{ij,\tau}(\varepsilon). As in the above proof of Theorem 2.4, it is easy to see that at least one of the coefficients cs,ε±c_{s,\varepsilon}^{\pm}, s=1,,ns=1,\ldots,n, in representation (2.7) for the associated non-trivial functions ψε\psi_{\varepsilon} is non-zero. Then it follows from asymptotic expansions (2.16), (2.18), (2.19), (2.20) that conditions (2.27), (2.30), (2.31), (2.34) ensure that Rekij,τ(ε)>0\operatorname{Re}k_{ij,\tau}(\varepsilon)>0 and τImkij,τ(ε)<0\tau\operatorname{Im}k_{ij,\tau}(\varepsilon)<0 and hence, the pole kij,τ(ε)k_{ij,\tau}(\varepsilon) corresponds to an eigenvalue λij,τ(ε)=Λpkij,τ2(ε)\lambda_{ij,\tau}(\varepsilon)=\Lambda_{p}-k_{ij,\tau}^{2}(\varepsilon). The stated asymptotic behavior for this eigenvalue is implied by (2.16), (2.18), (2.19), (2.20). In the same way, conditions (2.35), (2.38) ensure that Rekij,τ(ε)<0\operatorname{Re}k_{ij,\tau}(\varepsilon)<0 and hence, the pole kij,εk_{ij,\varepsilon} corresponds to a resonance λij,τ(ε)=Λpkij,τ2(ε)\lambda_{ij,\tau}(\varepsilon)=\Lambda_{p}-k_{ij,\tau}^{2}(\varepsilon) and it possesses the stated behavior.

If μi\mu_{i} is a simple eigenvalue of the matrix M1\mathrm{M}_{1}, then qi=1q_{i}=1. In this case, a non-trivial solution lεl_{\varepsilon} to system (4.21) associated with ki1,τ(ε)k_{i1,\tau}(\varepsilon) converges to the vector ei\mathrm{e}_{i}. Indeed, it follows from representation (5.4) that the corresponding zero zij,τ(ε)z_{ij,\tau}(\varepsilon) of equation (5.5) is an eigenvalue of the matrix M1εM~2(zij,τ(ε),ε)\mathrm{M}_{1}-\varepsilon\tilde{\mathrm{M}}_{2}(z_{ij,\tau}(\varepsilon),\varepsilon) and lεl_{\varepsilon} is an associated eigenvector. And since zij,τ(ε)z_{ij,\tau}(\varepsilon) converges to μi\mu_{i} and both these eigenvalues are simple, the eigenvector associated with ki1,τ(ε)k_{i1,\tau}(\varepsilon) converges to lεl_{\varepsilon} as well. Now by representation (4.22), (4.23), definition (2.10) of the operator 𝒢p,τ\mathcal{G}_{p,\tau} and formula (4.11) we conclude that the coefficients cs,ε±c_{s,\varepsilon}^{\pm} in representation (2.7) satisfy the identities:

cs,ε±=εt=1nΩeKt(0)xdψs(x)¯1ei,tψtp+1𝑑x+o(ε).c_{s,\varepsilon}^{\pm}=\varepsilon\sum\limits_{t=1}^{n}\int\limits_{\Omega}e^{\mp K_{t}(0)x_{d}}\overline{\psi_{s}(x^{\prime})}\mathcal{L}_{1}\mathrm{e}_{i,t}\psi_{t-p+1}\,dx+o(\varepsilon).

Hence, by condition (2.42), at least one of the coefficients cs,ε±c_{s,\varepsilon}^{\pm} is non-zero. As above, it is easy to see that one of conditions (2.27) or (2.30) and one of conditions (2.39) or (2.41) ensures that the functions eKs(kij,τ(ε))|xd|e^{-K_{s}(k_{ij,\tau}(\varepsilon))|x_{d}|} grows exponentially at infinity and hence, the same is true for the function ψε\psi_{\varepsilon}. Therefore, the pole kij,τ(ε)k_{ij,\tau}(\varepsilon) corresponds to a resonance. The asymptotic expansion for this resonance can be established as above. The proof of Theorem 2.5 is complete.

Acknowledgments

The authors thank the referees for useful remarks. The research by D.I.B. and D.A.Z. is supported by the Russian Science Foundation (Grant No. 20-11-19995).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • [1] I.V. Barashenkov, D.A. Zezyulin and V.V. Konotop, Jamming anomaly in 𝒫𝒯\mathcal{PT}-symmetric systems, New J. Phys. 18 (2016) 075015.
  • [2] C.M. Bender, Making sense of non-Hermitian Hamiltonians, Rep. Prog. Phys. 70 (2007) 947–1018.
  • [3] C.M. Bender and S. Boettcher, Real spectra in non-Hermitian Hamiltonians having 𝒫𝒯\mathcal{PT}-symmetry, Phys. Rev. Lett. 80 (1998) 5243–5246.
  • [4] R. Blankenbecler, M. L. Goldberger, B. Simon, The bound states of weakly coupled long-range one-dimensional quantum Hamiltonians, Ann. Physics, 108(1977) 69–78.
  • [5] D. Borisov, Discrete spectrum of a pair of non-symmetric waveguides coupled by a window, Sb. Math., 197 (2006) 475–504.
  • [6] D. Borisov, Asymptotic behaviour of the spectrum of a waveguide with distant perturbation, Math. Phys. Anal. Geom. 10 (2007) 155–196.
  • [7] D.I. Borisov, Distant perturbations of the Laplacian in a multi-dimensional space, Ann. H. Poincaré, 8 (2007) 1371–1399.
  • [8] D.I. Borisov, R.R. Gadyl’shin, On spectrum of periodic operator with a small localized perturbation, Izv. Math. 72 (2008) 659–688.
  • [9] D. Borisov, On spectrum of two-dimensional periodic operator with small localized perturbation, Izv. Math. 75 (2011) 471–505.
  • [10] D. Borisov, S. Dmitriev, On the spectral stability of kinks in 2D Klein-Gordon model with parity-time-symmetric perturbation, Stud. Appl. Math. 138 (2017) 317–342.
  • [11] D. Borisov, G. Cardone, On a one-dimensional quadratic operator pencil with a small periodic perturbation, ESAIM: COCV 26 (2020) id 21.
  • [12] D. Borisov, D. Krejčiřík, 𝒫𝒯\mathcal{PT}-symmetric waveguide, Integ. Equat. Oper. Theo. 62 (2008) 489–515.
  • [13] D. Borisov, Emergence of eigenvalue for 𝒫𝒯\mathcal{PT}-symmetric operator in thin strip, Math. Notes. 98 (2015) 827–883.
  • [14] D.I. Borisov, M. Znojil, On eigenvalues of a 𝒫𝒯\mathcal{PT}-symmetric operator in a thin layer, Sb. Math. 208 (2017) 173–199.
  • [15] D. Borisov, The spectrum of two quantum layers coupled by a window, J. Phys. A: Math. Theor. 40 (2007) 5045–5066.
  • [16] D. Borisov, Perturbation of threshold of essential spectrum for waveguide with window. I. Decaying resonance solutions, J. Math. Sci. 205 (2015) 19–54.
  • [17] D.I. Borisov, D.A. Zezyulin, Spacing gain and absorption in a simple PT-symmetric model: spectral singularities and ladders of eigenvalues and resonances, J. Phys. A: Math. Theor. 52 (2019) id 445202.
  • [18] D.I. Borisov, D.A. Zezyulin, Sequences of closely spaced resonances and eigenvalues for bipartite complex potentials, Appl. Math. Lett. 100 (2020) id 106049.
  • [19] L. D. Carr and J. Brand, Multidimensional solitons: theory, in Emergent nonlinear phenomena in Bose-Einstein condensates: theory and experiment, ed. by P. G. Kevrekidis, D. J. Frantzeskakis, and R. Carretero-González (Springer, Berlin, 2008).
  • [20] S. Cuccagna, D. Pelinovsky, Bifurcations from the endpoints of the essential spectrum in the linearized nonlinear Schrödinger problem, J. Math. Phys. 46 (2005) 053520.
  • [21] P. Duclos and P. Exner, Curvature-induced bound state in quantum waveguides in two and three dimensions, Rev. Math. Phys. 7 (1995) 73–102.
  • [22] R. El-Ganainy, K.G. Makris, M. Khajavikhan, Z.H. Musslimani, S. Rotter and D.N. Christodoulides, Non-Hermitian physics and PT symmetry, Nat. Phys. 14 (2018) 11–19.
  • [23] P. Exner, P. Šeba, M. Tater, D. Vaněk, Bound states and scattering in quantum waveguides coupled laterally through a boundary window, J. Math. Phys. 37 (1996) 4867–4887.
  • [24] L. Feng, R. El-Ganainy, and L. Ge, Non-hermitian photonics based on parity-time symmetry, Nat. Photonics 11 (2017) 752–762.
  • [25] R.R. Gadyl’shin, Local perturbations of the Schrödinger operator on the axis, Theor. Math. Phys. 132 (2002) 976–-982.
  • [26] R. R. Gadyl’shin, Local perturbations of the Schrödinger operator on the plane, Theor. Math. Phys. 138 (2004), 33–44.
  • [27] R.R. Gadyl’shin. On regular and singular perturbations of acoustic and quantum waveguides, C.R. Méch. 332 (2004) 647–652.
  • [28] S. Garmon, M. Gianfreda, and N. Hatano, Bound states, scattering states, and resonant states in 𝒫𝒯\mathcal{PT}-symmetric open quantum systems, Phys. Rev. A 92, (2015) id 022125.
  • [29] F. Gesztesy, H. Holden, A unified approach to eigenvalues and resonances of Schrödinger operators using Fredholm determinants, J. Math. Anal. Appl. 123 (1987) 181–198.
  • [30] W.D. Heiss, The physics of exceptional points, J. Phys. A: Math. Theor. 45 (2011) id 444016.
  • [31] R. Høegh-Krohn, M. Mebkhout, The 1r\frac{1}{r} expansion for the critical multiple well problem, Comm. Math. Phys. 91 (1983) 65–73.
  • [32] C. W. Hsu, B. Zhen, A. Douglas Stone, J. D. Joannopoulos M. Soljačić, Bound states in the continuum, Nat. Rev. Mater. 1(2016), 16048.
  • [33] N.N. Kalitkin, A.B. Alshin, Ye.A. Alshina, B.V. Rogov, Calculations On Quasi-Equidistant Grids, (Fizmatlit 2005). (in Russian)
  • [34] Y. V. Kartashov, C. Milián, V. V. Konotop, and L. Torner, Bound states in the continuum in a two-dimensional PT-symmetric system, Opt. Lett. 43, (2018) 575–578.
  • [35] T. Kato, Perturbation Theory for Linear Operators, (Springer-Verlag, 1976).
  • [36] Yu. S. Kivshar, G. P. Agrawal, Optical Solitons: From Fibers to Photonic Crystals, (Academic Press, 2003).
  • [37] M. Klaus, On the bound state of Schrödinger operators in one dimension, Ann. Phys. 108 (1977), 288–-300.
  • [38] M. Klaus, B. Simon, Coupling constant thresholds in nonrelativistic quantum mechanics. I. Short-range two-body case, Ann. Phys. 130 (1980) 251–281.
  • [39] K. Knopp, Theory of functions. Part II. (Dover Publ. 1947).
  • [40] V. V. Konotop, J. Yang, D. A. Zezyulin, Nonlinear waves in 𝒫𝒯\mathcal{PT}-symmetric systems, Rev. Mod. Phys. 88 (2016) id 035002.
  • [41] V. V. Konotop and D. A. Zezyulin, Phase transition through the splitting of self-dual spectral singularity in optical potentials, Opt. Lett. 42 (2017) 5206–5209.
  • [42] V. V. Konotop and D. A. Zezyulin, Spectral singularities of odd-𝒫𝒯\mathcal{PT}-symmetric potentials, Phys. Rev. A 99 (2019) 013823.
  • [43] S. Longhi, Parity-time symmetry meets photonics: A new twist in non-hermitian optics. Europhys. Lett. 120 (2017) id 64001.
  • [44] S. Longhi, Bound states in the continuum in 𝒫𝒯\mathcal{PT}-symmetric optical lattices, Opt. Lett. 39 (2014) 1697–1700.
  • [45] C. Meng, X. Zhang, S. T. Tang, M. Yang, Z. Yang, Acoustic coherent perfect absorbers as sensitive null detectors, Sci. Rep. 7 (2017) id 43574
  • [46] M.-A. Miri and A. Alú, Exceptional points in optics and photonics, Science 363 (2019) eaar7709.
  • [47] N. Moiseyev, Non-Hermitian Quantum Mechanics (Cambridge Univ. Press, 2011).
  • [48] A. Mostafazadeh, Pseudo-hermitian Representation of Quantum Mechanics. Int. J. Geom. Methods Mod. Phys. 7 (2010) 1191–1306.
  • [49] A. Mostafazadeh, Spectral singularities of complex scattering potentials and infinite reflection and transmission coefficients at real energies Phys. Rev. Lett. 102 (2009) id 220402.
  • [50] A. Mostafazadeh, Physics of spectral singularities, Trends in Mathematics 145 (Springer International Publishing, 2015).
  • [51] J. G. Muga, J. P. Palao, B. Navarro, and I. L. Egusquiza, Complex absorbing potentials, Phys. Rep. 395 (2004) 357–426.
  • [52] A. Müllers, B. Santra, C. Baals, J. Jiang, J. Benary, R. Labouvie, D. A. Zezyulin, V. V. Konotop, H. Ott, Coherent perfect absorption of nonlinear matter waves. Sci. Adv. 4 (2018) eaat6539.
  • [53] M. A. Naimark, Investigation of the spectrum and the expansion in eigenfunctions of a nonselfadjoint operator of the second order on a semi-axis, Tr. Mosk. Mat. Obs. 3 (1954) 181–270.
  • [54] S.A. Nazarov, Almost standing waves in a periodic waveguide with a resonator and near-threshold eigenvalues, St. Petersburg Math. J. 28 (2017) 377–410.
  • [55] S.A. Nazarov, Threshold resonances and virtual levels in cylindrical and periodic waveguides, Izv. Math. 84 (2020), to appear. DOI: 10.1070/IM8928
  • [56] J. von Neumann and E. Wigner, Über merkwürdige diskrete eigenwerte, Phys. Z. 30 (1929) 467–470.
  • [57] Ş. K. Özdemir, S. Rotter, F. Nori, and L. Yang, Parity-time symmetry and exceptional points in photonics, Nature Materials 18 (2019) 783–798.
  • [58] J. Pappademos, U. Sukhatme, and A. Pagnamenta, Bound states in the continuum from supersymmetric quantum mechanics, Phys. Rev. A 48 (1993) 3525–3532.
  • [59] L.P. Pitaevskii and S. Stringari, Bose-Einstein Condensation, (Clarendon Press, 2003).
  • [60] M. Robnik, A simple separable Hamiltonian having bound states in the continuum, J. Phys. A: Math. Gen. 19 (1986) 3845–3848.
  • [61] J. Schwartz, Some non-selfadjoint operators, Comm. Pure Appl. Math. 13 (1960) 609–639.
  • [62] B. Simon, The bound state of weakly coupled Schrödinger operators in one and two dimensions, Ann. Phys. 97 279–288 (1976).
  • [63] F. H. Stillinger and D. R. Herrick, Bound states in the continuum, Phys. Rev. A 11 (1975) 446–454.
  • [64] B. Vainberg, On the analytical properties of the resolvent for a certain class of operator-pencils, Math. USSR Sbornik 6 (1968) 241–273.
  • [65] V. Vougalter, On threshold eigenvalues and resonances for the linearized NLS equation, Math. Model. Nat. Phenom. 5 (2010), no. 4, 448-469.
  • [66] W. Wan, Y. Chong, L. Ge, H. Noh, A. D. Stone, H. Cao, Time-reversed lasing and interferometric control of absorption, Science 331 (2011) 889–892.
  • [67] J. Yang, Classes of non-parity-time-symmetric optical potentials with exceptional-point-free phase transitions, Opt. Lett. 42 (2017) 4067–4070.
  • [68] J. Yang, Partially 𝒫𝒯\mathcal{PT}-symmetric optical potentials with all-real spectra and soliton families in multidimensions, Opt. Lett. 39 (2014) 1133–1136.
  • [69] V. A. Zheludev, Perturbations of the spectrum of the Schrödinger operator with a complex periodic potential, Spectral Theory, Topics Math. Phys. 3 (1969) 25–41.
  • [70] A. A. Zyablovsky, A.P. Vinogradov, A.A. Pukhov, A.V. Dorofeenko and A.A. Lisyansky, 𝒫𝒯\mathcal{PT}-symmetry in optics, Physics – Uspekhi (2014) 57 1063–1082.