Abstract
The Laplace-Beltrami operator (LBO) on a sphere with a cut arises when considering the problem of wave scattering by a quarter-plane. Recent methods developed for sound-soft (Dirichlet) and sound-hard (Neumann) quarter-planes rely on an a priori knowledge of the spectrum of the LBO. In this article we consider this spectral problem for more general boundary conditions, including Dirichlet, Neumann, real and complex impedance, where the value of the impedance varies like , being the distance from the vertex of the quarter-plane and being constant, and any combination of these. We analyse the corresponding eigenvalues of the LBO, both theoretically and numerically. We show in particular that when the operator stops being self-adjoint, its eigenvalues are complex and are contained within a sector of the complex plane, for which we provide analytical bounds. Moreover, for impedance of small enough modulus , the complex eigenvalues approach the real eigenvalues of the Neumann case.
1. Introduction
Scattering (or diffraction) problems involve studying the field resulting from a wave incident upon an obstacle. This can for example be an acoustic or an electromagnetic wave. In general, these are complicated time-dependent problems, but often a hypothesis of time harmonicity can be made (the time dependency is simply a factor , where is the frequency of the incident wave) and the wave equation reduces to the Helmholtz equation. A class of scattering problems, the canonical scattering problems, is particularly important. It derives from studying scattering by simple obstacles with particular characteristics such as sharp edges or corners, often of infinite size. Although ‘simple’, these geometries can be used to evaluate the scattered field from more complicated geometries with high-frequency incident waves. The understanding of such canonical geometries represents the building blocks for Keller's Geometrical Theory of Diffraction (GTD) ( 1 ). There is a long history of mathematicians working on such canonical problems; one of the first was Sommerfeld, leading to his famous solution to the half-plane problem ( 2 ) and to the creation of the field of Mathematical Theory of Diffraction. Since then, some very ingenious mathematical methods have been developed to tackle such problems, the most famous being the Wiener-Hopf ( 3 ) and the Sommerfeld-Malyuzhinets ( 4 ) techniques. However, despite tremendous efforts in this field, some canonical problems remain open, in the sense that they have no clear analytical solution or rapidly convergent fast hybrid numerical-analytical scheme. In particular, this is the case for the canonical problem of scattering by a quarter-plane ( Fig. 1 ).

Fig. 1
The scattering problem for a quarter-plane
In the last 50 years, the quarter-plane problem has attracted a lot of attention, and different approaches have been used. By considering the quarter-plane as a degenerated elliptic cone, the field can be expressed as a spherical-wave multipole series ( 5–7 ), but these series are poorly convergent in the far-field. A review of this approach and attempts to accelerate the series convergence are described in ( 8 ). Another approach ( 9 , 10 ), elegantly based on the Wiener–Hopf technique in two complex variables, has been used, however it led to a solution that has proved to be erroneous ( 11 , 12 ). It is also worth mentioning an unusual and interesting approach in ( 13 ), where diffraction theory and statistical analysis are used in order to obtain the corner diffraction coefficient for a restricted set of incidence and observation direction.
A different way of considering this problem (in fact the more general problem of arbitrarily shaped cones), based on the use of spherical Green's functions has been introduced in ( 14 , 15 ) and led to an integral formula for the spherical diffraction coefficient. However, this solution is not valid for all incidence-observation directions and requires a numerical treatment and some regularisation of Abel-Poisson type in order to be evaluated ( 16 ). Building on this type of approach, a hybrid numerical-analytical method, which partially solves the acoustic quarter-plane problem in the Dirichlet case has been introduced in ( 17 ) and ( 18 ). The main advantage of this method compared to the one mentioned previously is that in this case the formulae giving the diffraction coefficient, the Modified Smyshlyaev Formulae (MSF) are ‘naturally convergent’ in the sense that they do not require special treatment to regularise or accelerate their convergence. The method is based on edge and spherical edge Green's functions and on the theory of embedding formulae, introduced in ( 19 ) and further developed in ( 20 ) for example. This method has been extensively described, adapted to the Neumann case and implemented in ( 21 ). The solution obtained, though valid for all observer directions over a wide range of incidence directions, remain partial in general, since for a range of incidence directions, there exists observation points for which the MSF are not valid. However, the range of validity of the MSF is bigger than that of any other method, going beyond the first singularity regions caused by the primary edge diffracted waves. A reason behind the limits of the MSF validity is the existence of secondary edge diffracted waves in the quarter-plane problem. In ( 22 ), a rigorous definition of the notion of the diffraction coefficient for the quarter-plane problem was given and the first analytical expression for these secondary edge waves generated during the diffraction process was derived. More recently, the Sommerfeld-Malyuzhinets technique was applied to the quarter-plane problem in the acoustic ( 23 ) and the electromagnetic ( 24 ) case, giving the overall structure of the diffracted field and recovering the results of ( 22 ) regarding the secondary diffracted waves.
The motivation of the present work is two-fold. On one hand, it has been shown for example in ( 21 ) that the derivation and the numerical evaluation of the MSF relies strongly on a detailed knowledge of the spectrum of the Laplace-Beltrami operator (LBO) on the unit sphere with a cut. On the other hand, it seems that most of the work regarding the quarter-plane problem has been carried out for Dirichlet or Neumann boundary conditions and that more general boundary conditions should now be considered. The case of constant impedance conditions on the surface of the quarter-plane cannot be treated in a similar way as the more usual Neumann and Dirichlet conditions. The reason behind this difference, is that the usual approach relies on the fact that such problems are separable in spherical coordinates. However, as shall be seen later, the problem of constant impedance on the face of the quarter-plane is not separable anymore. Important work regarding constant impedance for the diffraction by an arbitrary shaped cone has been carried out in ( 25–27 ), and alternative formulations have been provided. However, the problem becomes separable if one assumes that the value of the impedance depends on the radial variable and can be written as for some constant . Therefore, it is reasonable to believe that in that case, an embedding (or MSF) approach should work. Once separated, the problem reduces to that of finding the eigenvalues of the LBO on the unit sphere, with constant impedance boundary imposed on a cut. Hence, in order to make progress in this direction, it is first important to shed some light on the eigenvalues of the LBO with impedance conditions imposed on a cut. Impedance boundary conditions in canonical scattering problems have led to some very interesting developments, such has the Malyuzhinets technique in the case of the impedance wedge and impedance half-plane ( 28 , 29 ) or the Wiener–Hopf–Hilbert method for the impedance half-plane ( 30 ). There has also been some work on the spectrum of the LBO on a sphere with a cut, such as ( 31 ), ( 32 ), ( 33 ), ( 34 ) and ( 21 ), but again, to the authors' knowledge, solely in the Dirichlet and Neumann cases.
In the present work, the Laplace-Beltrami eigenvalue problem on a sphere with a cut will be approached both theoretically and numerically for a wide range of boundary conditions including Dirichlet, Neumann, real and complex impedance, and any combination of these. One of the most interesting features comes from the fact that a change in boundary conditions can lead to a big theoretical change. Indeed, the LBO with Dirichlet or Neumann boundary condition is a self-adjoint operator, for which the spectral theory is well understood and developed ( 35 ), while for complex impedance boundary conditions this ceases to be the case and the operator becomes non-self-adjoint, resulting in a rich behaviour of the eigenvalues and necessitating a different theoretical approach ( 36 , 37 ).
The rest of this article is structured as follows. The problem is presented in Section 2 . To this end, the problem of scattering by a quarter-plane is formulated in Section 2.1 and its link with the LBO eigenvalue problem is explained in Section 2.2 . In Section 2.3 , we provide a short description of the impedance boundary conditions and their implementation.
In Section 3 , we focus our attention on a set of boundary conditions, said to be of type I, corresponding to the LBO being self-adjoint. In Section 3.1 , we briefly summarise the relevant theoretical results in spectral theory of self-adjoint differential operators and, after having described important function sets in Section 3.2 , we apply them to the case of the LBO in Section 3.3 , showing that in this case, as expected, there is an infinite set of real positive discrete eigenvalues. Moreover, the relative position of the eigenvalues on the real line is investigated for different sets of boundary conditions of type I.
In Section 4 , we focus our attention on a set of boundary conditions, said to be of type II, corresponding to the LBO being non-self-adjoint. The breakdown of self-adjointness is explained in Section 4.1 . The appropriate theoretical framework for non-self-adjoint operators is presented in Section 4.2 and applied to the Laplace-Beltrami operator in Section 4.3 , showing that in this case there is an infinite discrete set of complex eigenvalues that are contained within a sector of the complex plane. In Section 4.4 , we show how to find an estimate of such sector.
Section 5 is dedicated to the numerical evaluation of the eigenvalues of the LBO. The numerical method used (surface finite elements) is briefly described in Section 5.1 and the results are presented, starting with the self-adjoint case (boundary conditions of type I) in Section 5.2 and following with the non-self-adjoint case (boundary conditions of type II) in Section 5.3 . We show that in this case, the complex eigenvalues are indeed contained within a sector of the complex plane, and that for small or large enough impedance, we may recover the self-adjoint eigenvalues. An explanation as why this is true on a theoretical level is provided in Appendix B.
2. Presentation of the problem
2.1 The quarter-plane problem
Throughout the article we assume that the problem is harmonic in time with frequency
and a time dependency proportional to
. As illustrated in
Fig. 1 , consider an incident acoustic plane wave
on a quarter-plane
. Let us call
and
the upper and lower surfaces of the plane sector respectively. The problem of scattering can be summarised as follows. We can write
, where the four quantities
(total field),
(incident field),
(reflected field) and
(scattered field) satisfy the Helmholtz equation
and
is the wavenumber of the homogeneous media surrounding the quarter-plane.
In order for the problem to be well-posed, some conditions need to be satisfied.
Using the geometric theory of diffraction (
22 ) or the Sommerfeld integrals (
23 ), it has been shown in the case of Dirichlet and Neumann boundary conditions that the far-field behaviour of the scattered field could be written as
the subscript
refers to the spherical wave emanating from the vertex, the subscript
refers to the different (primary and secondary) conical waves emanating from the edges and the subscript
refers to the wave reflected by the illuminated surface of the quarter plane as explained in details in (
21 ) and (
22 ).
should satisfy the edge conditions, that is the energy of the system should remain bounded as we approach the edge. This can be expressed mathematically as near the edges, where represents the distance to the edge. Physically, this also means that no sources should be located at the edges.
should satisfy the vertex condition, that is the energy of the system should remain bounded as we approach the vertex. This can be expressed mathematically by as , where represents the distance to the vertex.
The scattered field should satisfy a radiation condition. In other word, it should not exhibit any sources at infinity and should only consist of outgoing waves in the far-field. This radiation condition is an extension of the original Sommerfeld radiation condition ( 38 ), that is valid for infinite domains exhibiting edges or vertices. This extension follows the work by Rellich ( 39 ) and later Levine ( 40 ), as well as some uniqueness proof by Jones ( 41 ). For a comprehensive history of the evolution of the radiation condition, see ( 42 ).
Finally, and most importantly, has to satisfy some boundary conditions on and . These can for example be the Dirichlet boundary conditions ( ), the Neumann boundary conditions ( , where is a unit normal vector to ) or varying Robin boundary conditions ( , being a specified constant).
The structure of the conical waves is well understood (at least in the case of Dirichlet or Neumann boundary conditions), and some analytical expressions of their far-field structures are given in (
22 ) and (
23 ). The spherical wave is less understood. In order to reduce the problem, it is useful to introduce the diffraction coefficient
, where
represents the direction of observation and
the direction of incidence of
. The far-field behaviour of
can be represented by
The evaluation of the diffraction coefficient
has been the subject of many studies. A recent way of approaching the problem, leading to a partial resolution of the problem in the case of Dirichlet (
17 ) and Neumann (
21 ), is to use the theory of embedding formulae to obtain an integral expression of
. A spherical version of the embedded formulae, the so-called Modified Smyshlyaev formulae (MSF), are somehow easier to evaluate numerically, using the so-called coordinate equation (
18 ,
21 ). Thanks to this method, the diffraction coefficient is now easily computable for many pairs
. However, as emphasised in (
21 ), because of the secondary diffracted waves there are still some regions where it is complicated to evaluate
, and for this reason the problem of scattering by a quarter-plane is still considered as an open mathematical problem. In both cases (Dirichlet and Neumann), the definition and the numerical evaluation of the MSF relies heavily on knowledge of the eigenvalues of a certain linear differential operator: the LBO. Hence if one envisages the evaluation of the diffraction coefficient
for boundary conditions different from the pure Dirichlet or Neumann boundary conditions, it is important to be able to compute a priori the eigenvalues of the LBO associated with these boundary conditions. In the next Section, we shall introduce the LBO in more detail.
2.2 The LBO
Seeking a spherically separable solution to the Helmholtz equation (
2.1 ) of the form
, where
are the usual spherical coordinates (see
Fig. 2 (a)) leads to an equation for
,
, with solution
where
and
are constants,
and
are the Bessel function and the Hankel function of type 1 of order
respectively, and
is defined by
The separation of variables also leads to an equation for
:
where
is the LBO defined by
This can be interpreted as the restriction of the spherical Laplace operator to the surface of the unit sphere. It is also useful to introduce
and
that represents the restriction of the spherical gradient operator and the spherical divergence operator to the surface of the sphere. More precisely, for a scalar function
and a vector field
, we have
so that we have
. In fact, because of the geometry of the quarter plane, (
2.4 ) is actually defined on the surface of a sphere with a slit (or cut) of length
corresponding to the intersection between the unit sphere and the quarter-plane. As shall be seen in the next section, for the boundary conditions considered in this article, the boundary condition equation is also separable. Hence, solving the problem on the unit sphere with a cut is enough. Let us call
the corresponding domain (see
Fig. 2 (b)) and
and
the upper and lower part of the slit such that
.

Fig. 2
(a) Spherical coordinates and (b) definition of , and
The aim of this work is the rigorous study of the eigenvalue problem (
2.4 ), and that can be reformulated as
As already emphasised at the end of the previous section and shown in detail in (
21 ), knowledge of the eigenvalues is crucial to the definition and evaluation of the MSF that involve complicated contours surrounding the said eigenvalues. Remember that
and
are linked by (
2.3 ), and even if they are different, we may refer to both of them as eigenvalues of the problem (
2.7 ). Moreover, the first eigenvalue is also useful in order to understand the behaviour of the total field at the vertex of the quarter-plane. Indeed, it can be shown (see (
34 )) that
as
tends to zero, where
, corresponding to the first eigenvalue of the LBO, takes different values for different boundary conditions.
Many authors ( 21 , 31–34 ) have successfully considered this problem with different techniques (for example, WKB or shooting methods) for the case of pure Dirichlet boundary conditions (Dirichlet on both and ) and pure Neumann boundary conditions (Neumann on both and ). In this article, we would like to extend this knowledge to a broader range of boundary conditions. As we shall see in the following sections, the choice of boundary conditions may have a significant impact on the theory and the numerical methods necessary for the evaluation of the eigenvalues.
2.3 On the Robin boundary condition
Let us briefly summarise the definition and physical relevance of the Robin boundary conditions. Consider a locally reacting surface boundary surrounded by a homogeneous two- or three-dimensional space of density
, speed of sound
and wavenumber
. We can define the acoustic field by the pressure
and the acoustic velocity
. Then the
acoustic impedance or
normal acoustic impedance is defined as the ratio between the pressure and the normal fluid velocity at a point on the surface:
where we consider
as the normal pointing into the surface.
Let us now consider the potential
such that
Assuming that the acoustic field is harmonic with frequency
, we have
and we can write
Now using (
2.10 ) and (
2.9 ) into (
2.8 ), we obtain
Let us now introduce some notations/vocabulary. It is common to introduce the
characteristic acoustic impedance of a medium defined by
. Note that this is totally independent of the boundary of the domain. One can then define the
specific acoustic impedance by
. Looking at (
2.11 ), it is also useful to define the
specific acoustic admittance by
. In general,
and
can be complex. It is quite useful to rewrite (
2.11 ) in terms of
, which gives:
where we have used the notation
. Physical boundaries have to be either passive or absorbent, that is that the energy flux which flows across a surface element
of the boundary over an acoustic period
has to be positive. As emphasised in (
4 ), using the energy flux density vector (or acoustic Poynting vector), it can be shown that for a surface to be either passive or absorbent, we need to have
. As it turns out, this condition is often required to prove uniqueness results of scattering problem on a surface with boundary conditions (
2.12 ) (see for example (
4 ,
43 )).
In order to simplify the argument in the following sections, we shall introduce the quantity
defined by
, so that the impedance boundary condition on a surface can be rewritten as
Note that the absorbent condition (
) translates in a condition on
that is
. In the rest of this article, we shall refer to
as the impedance of our surface. Now, writing
, this condition reduces to
where
is now considered as a two-dimensional vector in the
basis. Hence, despite the fact that the Helmholtz equation is separable, this is not the case for the boundary equation if
is constant. However, if one considers the case when
for some constant
, then the boundary equation becomes separable and reduces to
Choosing such a varying impedance implies that the surface of the quarter-plane in that case is sound-soft at the vertex, and gradually becomes harder as we move away from the vertex. Note that uniqueness theorems for this type of boundary condition are studied in (
41 ) and mentioned in (
40 ). In what follows, when referring to impedance or Robin boundary conditions, we will refer to that situation.
3. Self-adjoint operators and boundary conditions of type I
Before presenting the spectral theory of linear operators, let us specify the boundary conditions that will be considered in this section, the boundary conditions of type I.
A suitably smooth functiononis said to satisfy boundary conditions of type I if it satisfies any one of the following boundary conditions onand :
The boundary conditions do not necessarily have to be the same on each side (or) of the slit. The orientation of the normals is specified in Fig. 3 (a) .

Fig. 3
Orientation of the normals and (a) and of the cut (b)
As we shall see in Section 3.3 , the LBO is well behaved for these boundary conditions, in the sense that it is self-adjoint. Let us start by summarising the theoretical framework appropriate to the study of such operators.
3.1 Theoretical framework
We first recall some key definitions and results from ( 35 ) and ( 36 ). A linear operator on a Hilbert space consists of a pair , where is a dense linear subset of for the Hilbert norm . is the domain of definition of the linear map and is called the domain of the linear operator . A linear operator on is said to be symmetric if for any in , we have . It is said to be non-negative if for any , . If and are two linear operators on such that and for all , we say that is an extension of and is a restriction of , and we write .
Given a linear operator on , we can define its adjoint operator , with domain , and with the condition that . We say that a linear operator on is self-adjoint if is symmetric and .
The resolvent set is defined as being the set of all such that is invertible ( being the identity operator) and the resolvent operator is bounded. A complex number is said to be an eigenvalue of if there exists a non-zero in Dom such that . The spectrum of , , is defined by . In the finite dimensional case, consists purely of eigenvalues. However, in the infinite dimensional case (of interest here), this is not necessarily the case. The spectrum of can be decomposed into the discrete spectrum of , (the set of all isolated eigenvalues of finite multiplicity) and the essential spectrum of , (the non-discrete part of the spectrum).
The main theoretical result that will be used in this section is the following spectral theorem.
Spectral theorem [Cor. 4.2.3 in ( 35 )]. Let T be an unbounded self-adjoint non-negative linear operator on . The following are equivalent :
(1) (Compact resolvent) The resolvent operatoris compact .
(2) The operatorhas empty essential spectrum .
(3) There exists a complete orthonormal set of eigenvectorsofwith corresponding eigenvaluessatisfyingastends to infinity .
In order to make some mathematical progress, it is very useful to think in terms of sesquilinear forms. In this paragraph, we shall recall a few facts and definitions about these. A sesquilinear form
defined on a subspace
of a Hilbert space
is a map
that is linear in its first argument and conjugate-linear in its second, which means that for any
, and any
, we have
where
denotes the complex conjugate. Concepts similar to the linear operators apply to sesquilinear forms. In particular, we say that a sesquilinear form
is densely defined if
is dense in
. We say that a sesquilinear form
is non-negative if for any
in
, we have
. A non-negative sesquilinear form
is said to be closed if the normed space
is complete, where
is the norm associated with
defined for
by
If this is the case,
is called the Hilbert space associated with
. We can define extensions of sesquilinear forms similarly to the operator case. If
and
are two non-negative sesquilinear forms such that
and for any
in
, we have
, we say that
is an extension of
. A sesquilinear form
is said to be closable if it admits a closed extension. The smallest closed extension of a closable form
is called the closure of
. For a closed sesquilinear form
, a finite dimensional subspace
, we can introduce the functional
as follows:
This allows us to define a quantity
defined for each integer
by
Note that when there is no ambiguity, we may just write
and
instead of
and
. Let us now state some useful results that link linear operators and sesquilinear forms.
Representation theorem [Thm. VI.2.6 in (
36 ), Thm. 4.4.2 in (
35 ), Thm. B.1.6 in (
44 )].
A closed non-negative sesquilinear formwith domain acting in a Hilbert space H gives rise to a non-negative self-adjoint operator with domaindefined byand such that for any ,
we have .
[Direct corollary of Thm. 4.5.2 in ( 35 )]. Letbe a closed non-negative sesquilinear form on a Hilbert space , andits associated self-adjoint operator (resulting from Theorem 3.3). Ifas , thenhas empty essential spectrum. And so by Theorem 3.2 , also has compact resolvent and an ordered discrete set of eigenvaluesand these correspond exactly to .
Finally, as a direct consequence of the definitions of and , it is possible to deduce the following theorem that will prove very useful in Section 3.3 .
Letandbe two closed sesquilinear forms. Then
(1) Ifand , then for any integer ,
(2) Ifandwe have , then .
3.2 Important function sets on
Let us start by defining different sets of smooth functions on
Let us also define different sets of compactly supported functions, where
refers to the support of a function
.
where what is meant by such a neighbourhood is described in
Fig. 4 .

Fig. 4
Illustration of typical neighbourhood in the definition of (a) and (b)
Let us also define the space
as the Lebesgue space of square integrable functions on
.
is a Hilbert space for the inner product
and its associated norm
defined by
where
. It is a classic result that
is dense in
(for the norm
). Let us now introduce the classic Sobolev space
, also sometimes referred to as
, as
where here functions and derivatives are understood in the distributional sense. Note that
is a Hilbert space for the inner product
and the associated norm
defined by
It is clear that
and
are all included in
. Hence, it is now possible to define the spaces
,
and
as the closure in
(for the norm
) of the sets
and
respectively. It is important to note that
3.3 The LBO as a self-adjoint operator for boundary conditions of type I
Consider the linear operator
acting in the Hilbert space
and formally defined by
as in (
2.5 ) with boundary conditions of type I. Let
be initially defined on the domain
defined by
The appropriate notation for each possible combination of boundary condition is specified in
Table 1 .
Table 1 The operator for boundary conditions of type I
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Dirichlet | Dirichlet | |
Neumann | Neumann | |
Dirichlet | Neumann | |
Neumann | Dirichlet | |
Real positive Robin | Dirichlet | |
Dirichlet | Real positive Robin | |
Real positive Robin | Neumann | |
Neumann | Real positive Robin | |
Real positive Robin | Real positive Robin | |
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Dirichlet | Dirichlet | |
Neumann | Neumann | |
Dirichlet | Neumann | |
Neumann | Dirichlet | |
Real positive Robin | Dirichlet | |
Dirichlet | Real positive Robin | |
Real positive Robin | Neumann | |
Neumann | Real positive Robin | |
Real positive Robin | Real positive Robin | |
Table 1 The operator for boundary conditions of type I
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Dirichlet | Dirichlet | |
Neumann | Neumann | |
Dirichlet | Neumann | |
Neumann | Dirichlet | |
Real positive Robin | Dirichlet | |
Dirichlet | Real positive Robin | |
Real positive Robin | Neumann | |
Neumann | Real positive Robin | |
Real positive Robin | Real positive Robin | |
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Dirichlet | Dirichlet | |
Neumann | Neumann | |
Dirichlet | Neumann | |
Neumann | Dirichlet | |
Real positive Robin | Dirichlet | |
Dirichlet | Real positive Robin | |
Real positive Robin | Neumann | |
Neumann | Real positive Robin | |
Real positive Robin | Real positive Robin | |
Because of the inclusion , and the fact that is dense in , it is automatic to see that is dense in , the Hilbert space is acting on. Hence is a well-defined linear operator on .
Now, using Green's identity on
(See Appendix A), we can write
When boundary conditions are applied, the right-hand side simplifies and the results are presented in
Table 2 , where what is meant by the line integrals along
and
is specified in Appendix A by (
A.6 ).
Table 2 Green's identity applied to for boundary conditions of type I
Operator
. |
. |
. |
---|
, , , , | | |
, | | |
, | | |
| | |
Operator
. |
. |
. |
---|
, , , , | | |
, | | |
, | | |
| | |
Table 2 Green's identity applied to for boundary conditions of type I
Operator
. |
. |
. |
---|
, , , , | | |
, | | |
, | | |
| | |
Operator
. |
. |
. |
---|
, , , , | | |
, | | |
, | | |
| | |
Since in this section, when considering Robin boundary conditions, we only consider real impedances
,
, and that as such, using
Table 2 we have that
which means that
is a symmetric linear operator. Now, using
in
Table 2 , and the fact that
, it is easy to see that
,
, which means that
is a non-negative linear operator. Note that the operator
gives rise to a sesquilinear form
with
defined in
Table 3 for each type of boundary conditions.
Table 3 Definition of the sesquilinear form arising from for boundary conditions of type I
Sesquilinear form
. |
. | Associated norm
. |
---|
, , , | | |
, | | |
, | | |
| | |
Sesquilinear form
. |
. | Associated norm
. |
---|
, , , | | |
, | | |
, | | |
| | |
Table 3 Definition of the sesquilinear form arising from for boundary conditions of type I
Sesquilinear form
. |
. | Associated norm
. |
---|
, , , | | |
, | | |
, | | |
| | |
Sesquilinear form
. |
. | Associated norm
. |
---|
, , , | | |
, | | |
, | | |
| | |
Now it is possible to define a closed sesquilinear form that is an extension of , that is, the formal definition of and its associated norm is the same as that given for in Table 3 , but its domain is bigger. The domain for different boundary conditions is given in Table 4 .
Table 4 Domain of the closed extension for boundary conditions of type I
Sesquilinear form
. | Closed extension
. |
. |
---|
| , | |
, | , | |
, | , | |
, , , | , , , | |
Sesquilinear form
. | Closed extension
. |
. |
---|
| , | |
, | , | |
, | , | |
, , , | , , , | |
Table 4 Domain of the closed extension for boundary conditions of type I
Sesquilinear form
. | Closed extension
. |
. |
---|
| , | |
, | , | |
, | , | |
, , , | , , , | |
Sesquilinear form
. | Closed extension
. |
. |
---|
| , | |
, | , | |
, | , | |
, , , | , , , | |
The fact that
is closed follows directly for boundary conditions of type I not involving Robin conditions, since for these (see
Table 3 ), we have
and by definition the associated
defined as a closure for
is complete for that norm. In order to show that
is closed in the Robin case as well, we need to show that there exists a constant
such that
. This proof will be omitted at this point, but will be shown to be a direct consequence of the analysis in
Section 4.3 , as will be explained in Remark 4.9. This implies that the norms
and
are equivalent and hence it is clear that
is the closure of
. Now, by Theorem 3.3, the closed sesquilinear form
gives rise to a non-negative self-adjoint operator
with domain
defined by
and such that for any
, we have
. It is easy to see that
is an extension of
. We shall refer to
as the Friedrich's extension of
, and when referring to the LBO with boundary conditions of type I, we will from now on mean
.
For boundary conditions of type I, the self-adjoint LBOhas an empty essential spectrum and a discrete set of real positive eigenvaluesenumerated in increasing order and repeated in accordance with multiplicity. In particular, ifandare the corresponding eigenvalues for the pure Dirichlet and pure Neumann operatorsand ,
then we have thatwhereis the set of non-zero natural numbers .
Let be the non-negative self-adjoint LBO with some boundary condition of type I and let its associated closed sesquilinear form.
Let us start by using the fact that in ( 31 ) it has been shown that the pure Neumann operator has eigenvalues (and hence also the associated , denoted ) that tend to infinity and so by Theorem 3.4 has compact resolvent and a discrete set of real positive eigenvalues enumerated in increasing order and repeated in accordance with multiplicity.
Now let us remark that we have
(see
Table 4 ), and that for any
we have
(see
Table 3 , and the fact that
), hence by Theorem 3.5
(ii) we have for any
that
Note also that
(see
Table 4 ) and that
(see
Table 3 ). Hence, by Theorem 3.5
(i) , we have for any
that
The last two cases remaining are treated in a similar way by noting that
and
(see
Table 4 ) and that for any
and
, we have
and
(see
Table 3 ). Hence, by Theorem 3.5
(ii) , we have for any
that
and
and so by using (
3.3 ), we have that
All the different cases have been treated and so, using the fact that
as
and the inequalities (
3.2 ),(
3.3 ) and (
3.4 ), we have that
as
and so by Theorem 3.4,
has compact resolvent and a discrete set of real positive eigenvalues
enumerated in increasing order and repeated in accordance with multiplicity corresponding exactly to
. From the inequalities (
3.2 ),(
3.3 ) and (
3.4 ), we also know that for any
and that along the way we have proved that
and
In order to conclude the proof, we just need to show the second part of the inequality (
3.1 ), and we shall proceed in a similar way. In order to do that, let us note that
(see
Table 4 ) and that moreover we have
. This comes from the fact that the line integrals that may occur in the definition of
(see
Table 3 ) are always zero when
. Hence, by Theorem 3.5 (i), we have that
which completes the proof. Note also that with the same reasoning, and realising that
and
, we can show that
. Similarly, we have that
. Moreover, using the fact that
and
, we can conclude by Theorem 3.5 (ii) that
. Finally, using the fact that
and
, we can conclude by Theorem 3.5 (i) that
.
And so we can summarise our findings by
▪
Note of course that the eigenvalues related to the cases with Robin boundary conditions do depend on
and
, and with the exact same method as that used in the proof of Theorem 3.6, using Theorem 3.5 (ii), we can show that
and
are strictly increasing functions of
, that
and
are strictly increasing functions of
and that
is a ‘strictly increasing function of
and
’ in the sense that if
and
, then
. It results from the boundedness and the strict increasing character of these eigenvalues, that they converge to a finite limit as
or
. And that we have
as can be verified by assuming
‡ continuity of the eigenvalues as functions of
and setting
We can also infer that
in a similar way. For
, divide the Robin boundary condition through by
to get a condition of the type
. The equality can then be verified by assuming continuity of the eigenvalues as functions of
and setting
.
4. Non-self-adjoint operators and boundary conditions of type II
4.1 Breakdown of self-adjointness for boundary conditions of type II
Let us now consider the same problem as in the previous section, but with a different type of boundary conditions.
A suitably smooth function onis said to satisfy boundary conditions of type II if it satisfies the complex Robin boundary condition on at least one ofor :
The condition on the other face can either be of type I or type II .
Note that to be physically relevant, as mentioned at the end of Section 2.3 , we should have . However, mathematically, the sign of is not relevant to obtain the eigenvalues. In fact if is solution to the eigenvalue problem associated with , then is solution to the eigenvalue problem associated with . Hence we can restrict our study to , the cases with negative imaginary part can be directly obtained by symmetry.
Consider the linear operator
acting in the Hilbert space
and formally defined by
as in (
2.5 ) with boundary conditions of type II. Let
be initially defined on the domain
defined by
Note that once again
is dense in
and so
is a well-defined linear operator. The appropriate notation for each possible combination of boundary condition of type II is specified in in
Table 5 .
Table 5 The operator for boundary conditions of type II
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Complex Robin | Dirichlet | |
Dirichlet | Complex Robin | |
Complex Robin | Neumann | |
Neumann | Complex Robin | |
Complex Robin | Complex Robin | |
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Complex Robin | Dirichlet | |
Dirichlet | Complex Robin | |
Complex Robin | Neumann | |
Neumann | Complex Robin | |
Complex Robin | Complex Robin | |
Table 5 The operator for boundary conditions of type II
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Complex Robin | Dirichlet | |
Dirichlet | Complex Robin | |
Complex Robin | Neumann | |
Neumann | Complex Robin | |
Complex Robin | Complex Robin | |
Boundary condition on
. | Boundary condition on
. | Operator
. |
---|
Complex Robin | Dirichlet | |
Dirichlet | Complex Robin | |
Complex Robin | Neumann | |
Neumann | Complex Robin | |
Complex Robin | Complex Robin | |
As in Section 3.3 , we can use Green's identity on in order to obtain expressions for and , a summary of which is displayed in Table 6 .
Table 6 Green's identity applied to for boundary conditions of type II
Operator
. |
. |
. |
---|
, | | |
, | | |
| | |
Operator
. |
. |
. |
---|
, | | |
, | | |
| | |
Table 6 Green's identity applied to for boundary conditions of type II
Operator
. |
. |
. |
---|
, | | |
, | | |
| | |
Operator
. |
. |
. |
---|
, | | |
, | | |
| | |
Note however that this time we have and so, as such, we have and the operator is not symmetric, and hence it will not be possible to find a self-adjoint extension of . Note that we also lose the non-negative property here since is not even real. Hence the theoretical framework of the previous section cannot be used in this case and a new approach must be taken.
4.2 Theoretical framework
Let us consider again a densely defined linear operator on a Hilbert space . The numerical range of , , is defined by s.t. and . We say that is m-accretive if and 1 belongs to . We say that is quasi-m-accretive if there exists such that is m-accretive. is said to be sectorial with vertex and semi-angle , if its numerical range is included in a sector of the complex plane with vertex and semi-angle , as illustrated in Fig. 5 . Finally, we say that a linear operator is m-sectorial if it is sectorial and quasi-m-accretive.

Fig. 5
Typical sector of the complex plane with vertex and semi-angle
In order to get some information about the spectrum of a non-self-adjoint linear operator, it is, once again, very useful to think in terms of sesquilinear forms. The numerical range of a sesquilinear form
is defined by
. We say that
is accretive if
and sectorial with vertex
and semi-angle
if its numerical range is included in a sector of the complex plane with vertex
and semi-angle
. We say that a sectorial sesquilinear form
is closed if the normed space
is complete, where
is the norm associated with
defined for
by
If this is the case,
is called the Hilbert space associated with
. A sectorial sesquilinear form
is said to be closable if it admits a closed extension. The smallest closed extension of a closable form
is called the closure of
.
Let us now present three important results from ( 36 ), that we will use in this article to link m-sectorial linear operators and sectorial sesquilinear forms
[Thm. VI.2.1 in ( 36 )]. Letbe a densely defined, closed, sectorial sesquilinear form in . Then there exists an m-sectorial linear operatorsuch that
(1)and
(2) The resulting operatoris uniquely determined by this construction
Theorem 4.3 is important in the sense that it allows us to associate an m-sectorial operator to a given (closed and sectorial) sesquilinear form. The following theorem makes use of this association and will prove extremely important when trying to describe the non-self-adjoint operators arising in Section 4.3 .
[Thm. VI.3.4 in ( 36 )]. Letbe a densely defined, closed, accretive and sectorial sesquilinear form. Letbe the m-sectorial linear operator associated with . Letbe a sesquilinear form bounded with respect toin the sense thatand for , we have , whereandare non-negative real numbers and .
Then the sesquilinear formdefined byis also closed and sectorial. Letbe the m-sectorial operator associated with . Ifhas compact resolvent, thenhas compact resolvent too .
At this stage, it is useful to remark that from the definitions it is clear that a non-negative self-adjoint operator with compact resolvent is an m-sectorial operator and a closed non-negative sesquilinear form is a closed, accretive and sectorial sesquilinear form.
[Thm. III.6.29 in ( 36 )]. Letbe an m-sectorial operator with compact resolvent. Then the spectrum of , , consists entirely of isolated eigenvalues with finite multiplicities and is included in its numerical range , . In order to avoid confusion with the eigenvalues described in Section 3.3 , we shall denote these eigenvalues .
The last part of this theorem, , is not directly included in Theorem 6.29 in ( 36 ), however it is relatively straightforward. Indeed, let , and a corresponding eigenvector in . We have and so . Now let . It is clear that , that and that . Hence, by definition of the numerical range, . ▪
4.3 The LBO as m-sectorial operator for boundary conditions of type II
Note that the operator defined in Section 4.1 for boundary conditions of type II gives rise to a densely defined sesquilinear form with defined in Table 7 for each type of boundary conditions.
Table 7 Expression of the sesquilinear form arising from and the sesquilinear form for boundary conditions of type II
Sesquilinear form
. |
. | Sesquilinear form
. |
. |
---|
, | | , | |
, | | , | |
| | | |
Sesquilinear form
. |
. | Sesquilinear form
. |
. |
---|
, | | , | |
, | | , | |
| | | |
Table 7 Expression of the sesquilinear form arising from and the sesquilinear form for boundary conditions of type II
Sesquilinear form
. |
. | Sesquilinear form
. |
. |
---|
, | | , | |
, | | , | |
| | | |
Sesquilinear form
. |
. | Sesquilinear form
. |
. |
---|
, | | , | |
, | | , | |
| | | |
Let us now consider the extension of defined as in Table 7 , but with a bigger domain that contains and given in Table 8 .
Table 8 Domain of the extension for boundary conditions of type II
Sesquilinear form
. | Extension
. |
. |
---|
| | |
| | |
, , | , , | |
Sesquilinear form
. | Extension
. |
. |
---|
| | |
| | |
, , | , , | |
Table 8 Domain of the extension for boundary conditions of type II
Sesquilinear form
. | Extension
. |
. |
---|
| | |
| | |
, , | , , | |
Sesquilinear form
. | Extension
. |
. |
---|
| | |
| | |
, , | , , | |
It is now possible to define two new sesquilinear forms
and
with
, such that
The exact expression of
is given in
Table 7 for different boundary conditions of type II. The crucial point at this stage is to realise that
, by its definition and its domain
corresponds exactly to one of the sesquilinear forms
,
or
defined in
Section 3.3 , as specified in
Table 9 . But we know from
Section 3.3 that
,
and
are closed and non-negative sesquilinear forms, and so, by Remark 4.5, they are also closed, accretive and sectorial forms and hence, so is
. Hence we can conclude that
is associated uniquely with an m-sectorial linear operator
in the sense of Theorem 4.3. Moreover, we know from
Section 3.3 that
,
and
are associated in the same way with the non-negative self-adjoint operators with compact resolvent
,
and
. By Remark 4.5, we know that
,
and
are also m-sectorial, and so, by uniqueness,
is exactly equal to one of these operators as specified in
Table 9 and has compact resolvent.
Table 9 Correspondence between and as well as between and
Sesq. Form
. | Form
. | Operator
. | Operator
. |
---|
| | | |
| | | |
, , | | , | |
Sesq. Form
. | Form
. | Operator
. | Operator
. |
---|
| | | |
| | | |
, , | | , | |
Table 9 Correspondence between and as well as between and
Sesq. Form
. | Form
. | Operator
. | Operator
. |
---|
| | | |
| | | |
, , | | , | |
Sesq. Form
. | Form
. | Operator
. | Operator
. |
---|
| | | |
| | | |
, , | | , | |
We are now almost in a position to apply the Theorem 4.4 to . The last remaining hypothesis that we need to verify is that is bounded with respect to .
is bounded with respect to
We will focus on the most general case of Robin conditions on both sides of the cut (the other boundary conditions of type II can be treated in exactly the same way), so in this proof we shall focus only on
and
and will drop the superscript for the duration of the proof only. In order to prove that
is bounded with respect to
, we need to show that there exists two positive real constants
and
, with
such that
for
belonging to a dense subset of
. Here we shall choose
to be in the dense subset
defined in (
4.1 ). Before embarking into the proof, let us state an intermediate lemma deriving from ((
36 ), Eqn. IV.(1.19)) and from the fact that for any real
and
, we have
.
Letbe an open segment ofwith .
Then for any ,
anyin the closed segmentand any ,
we have :
We are now well equipped to start the proof. Let us try to bound
by first noting that
where
Let us try to bound
first. Making use of the lemma 4.8, by choosing
,
and
, for
, we can show that
Hence we can use (
4.3 ) to show that
Now noting that when
, we have
, we can deduce that
because the region
and
described in
Fig. 6 (a) is a subset of
. It is useful to introduce the coefficients
and
,
Similarly, but this time using an intermediate region
and
, also described in
Fig. 6 (b), and choosing
,
and
, we can also show that

Fig. 6
Illustration of the intermediate domains (a) and (b)
Hence, using (
4.4 ) and (
4.6 ) in (
4.2 ), we obtain
and note that we can always have
upon choosing
big enough. Hence by choosing
we have proved that
is bounded with respect to
. ▪
Note that if
and
are real and positive as in
Section 3.3 , the inequality (
4.7 ) implies that we have
and hence upon choosing
, we have
, as requested in
Section 3.3 .
To summarise, we now know that , with , where is a densely defined closed accretive and sectorial sesquilinear form associated to a linear operator that is m-sectorial and has compact resolvent. Moreover the sesquilinear form is bounded with respect to . Hence, we can apply the Theorem 4.4 to conclude that the sesquilinear form is sectorial and closed and is associated with an m-sectorial operator that has compact resolvent § . From now on, when referring to the LBO with boundary conditions of type II, we will mean . This allows us to obtain the main result of this section:
For boundary conditions of type II, the spectrum of the LBOconsists entirely of isolated eigenvalues with finite multiplicities denoted . Moreover, there existsandsuch that the eigenvalues are contained in the sector of the complex plane with vertexand semi-angle .
We know that is m-sectorial, so its numerical range is contained in a sector of the complex plane with vertex and semi-angle . Moreover, has compact resolvent and so by Theorem 4.6, its spectrum consists entirely of isolated eigenvalues with finite multiplicities, and we have . ▪
4.4 Estimating an appropriate sector
So far, we know that the eigenvalues of the LBO lie within a sector of the complex plane, but we do not have a constructive method of finding the characteristic of this sector. In order to do so, let us start by noting that, following exactly the same technique as that used in the proof of Lemma 4.7, we obtain
where
Finding a pair
such that
is included in the sector with vertex
and semi-angle
is equivalent to finding a pair
such that we have
In order to obtain an inequality of the type (
4.12 ), we need to work with the real part of
. Let us proceed as follows:
Hence, by choosing
to be
and noting that
, (
4.14 ) becomes
which implies that
Hence, provided that the following condition is satisfied
we are sure that (
4.12 ) holds. Let us now try to find an appropriate
that will make (
4.13 ) hold. Note that from (
4.16 )we have
Now, we shall assume that
and so we have
So by choosing
such that
the inequality (
4.13 ) holds, provided that the following condition
is satisfied. Note that if
Cond2 is satisfied, then
Cond1 is automatically satisfied.
Cond2 results in the following condition on the choice of
:
It is always possible to find such
. We therefore have a constructive way of finding
and
. For a given pair of impedances we can compute
and
using (
4.11 ). Pick
such that
Cond3 is satisfied and choose
and
according to (
4.15 ) and (
4.20 ). It is then possible to look for the best possible
, that is
as close to zero as possible (remember that with (
4.15 ),
is always negative). This can be done by optimising the choice of
and
in order to minimise the value of
. In particular, one may realise that
is strictly increasing with
for a given
, and that for a given
,
is always minimum when
is such that
. Note that we do not claim here to have the smallest sector possible, but this is not the focus of the present work.
5. Numerical method and results
5.1 Numerical method
In order to compute the eigenvalues we have used a numerical method based on surface finite element (see for example ( 45 )). In particular, we used the open-source library Deal.II ( 46 ). This particular library has the capability of dealing with computation on two dimensional surfaces embedded in a three-dimensional space (Riemannian manifolds, typical characteristic of ). For our purpose, we shall use a combination of step-38 and step-36 of the tutorial problems. Once the mass and stiffness matrices are constructed (see Appendix C), the eigenvalue problem is solved using the PETSc and SLEPc libraries ( 47 , 48 ).
One of the drawbacks of Deal.II is the fact that it can only deal with quadrilateral meshes for such cases. In order to mesh the sphere with the slit, we have made used of the open-source meshing software Gmsh ( 49 ). Meshing the sphere with a slit is not a trivial issue, and the procedure used here was first to mesh the sphere with a diamond hole, and then to force the two extremities of the hole to be the same point. Hence, even if and are at the same geometrical location, they do not have the same mesh points so that we can prescribe different boundary conditions on and . We use the finite element space of continuous, piecewise polynomials of degree 2 in each coordinate direction. It is worth noting that other innovative methods such as ( 50 ), based on the properties of stereographic projections and fast multipole methods, have been developed to solve the Laplace-Beltrami equation on a sphere with ‘islands’ subject to Dirichlet boundary conditions.
5.2 Results for boundary conditions of type I
The first thing to do is to consider the mixed Dirichlet–Neumann problem, that is Dirichlet boundary condition on one face and Neumann boundary condition on the other. This is the simplest example that could not be dealt with by the simple method proposed in ( 21 ) to solve the pure Dirichlet and pure Neumann cases. These three cases, pure Dirichlet, pure Neumann and mixed Dirichlet–Neumann, do not depend on any impedance parameter and are represented by plain horizontal lines in Fig. 7 . The next step is to apply Robin boundary conditions with real. For these conditions, the LBO is also non-negative and self-adjoint, so we expect to obtain real positive eigenvalues. Figure 7 shows the results with Robin boundary conditions on one face and Neumann boundary conditions on the other with hollow squares , Robin boundary condition on one face and Dirichlet on the other with plain squares ( ) and Robin boundary conditions on both faces with circles . For plotting purposes, we have chosen , but cases when can also be dealt with in a similar way. The eigenvalues are plotted against and presented in Fig. 7 , while some of the normalised eigenfunctions are plotted in Fig. 8 . The results presented in Fig. 7 illustrate the smooth transition from pure Neumann to pure Dirichlet via Robin conditions, the smooth transition from pure Neumann to mixed Dirichlet–Neumann via Robin-Neumann conditions and the smooth transition from Dirichlet–Neumann to pure Dirichlet via Dirichlet-Robin boundary conditions. As predicted in Section 3.3 and Remark 3.7, the eigenvalues are increasing functions of and the inequalities ( 3.5 ) are satisfied.

Fig. 7
The first four eigenvalues of the LBO for different boundary conditions of type I

Fig. 8
Four typical normalised eigenfunctions for different boundary conditions of type I: 1st eigenfunction for pure Dirichlet (top left), 2nd eigenfunction for pure Neumann (top right), 5th eigenfunction for mixed Neumann-Dirichlet (bottom left) and 2nd eigenfunction for real Robin conditions on both faces with
It is interesting to note that contrary to the pure Dirichlet and the pure Neumann eigenfunctions, the mixed Neumann-Dirichlet eigenfunctions do not exhibit symmetry (or antisymmetry) across the equatorial plane (see Fig. 8 ). This breakdown in symmetry can also be observed in the Robin case when different impedances are being used on each side of the slit. This is an important observation since the symmetry properties of the pure Neumann and pure Dirichlet eigenfunctions were essential for the method developed in ( 21 ) to work.
In order to have an idea of the precision with which our eigenvalues were computed in the Robin case, we performed a mesh sensitivity analysis by performing these computations for successive mesh refinements. The usual exponential convergence from above is obtained, as shown in Fig. 9 .

Fig. 9
Mesh sensitivity analysis for Robin boundary conditions on both faces, with . The mesh has been refined three times and the logarithm (with base 10) of the error relative to the results of the third refinement, , is plotted against the number of refinements. The superscript denotes the results of the third refinement. The results were recorded with six significant digits, so when the point of the second refinement is missing, this means that the exact same result (up to six digit accuracy) was obtained for the second and the third refinement
5.3 Results for boundary conditions of type II
For these computations we use a very similar method to that presented in the previous section, that is a surface finite element method using the libraries Deal.II, PETSc and SLEPc and a similar mesh. However, at the time of performing the computations the use of complex numbers was not yet fully supported in Deal.II. Hence some work had to be carried out in order to solve the problem with boundary conditions of type II, leading to a partial implementation of complex numbers within the Deal.II library. For plotting purposes, and in order to match with Section 4.3 , we are solving the problem for and ( , ), that is, for Robin–Robin ( , circles in Figs 10 and 11 ) and Neumann–Robin ( , hollow squares) boundary conditions of type II. We then rewrite as and let vary between and . For completeness, we show that it is possible to deal with the case and compute the Robin–Robin eigenvalues ( , plain squares) when and and let vary between and . The results for the location of the eigenvalues in are presented in Fig. 10 . The pure Neumann ( ), Dirichlet–Neumann ( ) and pure Dirichlet ( ) are represented as single points on these graphs. Once again, we observe a smooth transition from pure Neumann to pure Dirichlet in the case of Robin–Robin conditions, and a smooth transition from pure Neumann to Dirichlet–Neumann in the case of Neumann-Robin conditions. The main difference between these results and the results presented in Section 5.2 is that this time the smooth transition occurs through the upper half of the complex plane.
![The first four eigenvalues of the LBO for different boundary conditions of type II, when α=c(1+i) and c∈[10−2,103] . In the Robin–Robin case (circles) we have α±=α , and in the Neumann-Robin case (hollow squares), we have α+=0 and α−=α . In the Robin–Robin case (plain squares), we have α+=c(1+i) and α−=c(1/2+3i)](https://oup.silverchair-cdn.com/oup/backfile/Content_public/Journal/qjmam/69/3/10.1093_qjmam_hbw008/3/m_hbw008f10.jpeg?Expires=1748027309&Signature=qrnsGFgS28fM-I4RqgC6GGns3zD2AuzdU~OEKCyMu8uAvtTnlaDWS7uh~IRbWRnvUYyb~E0Lex-P5WCuUCwve0npW8sNDRTM~1sF9EDZyeAf2h4FBKFtipEi1sr2HXF4KKHNOkcSlieVNmQtRq9N0QiCnTtPgfKluJ~fehgQbZ~GoixAEuk04X67eixr~2YmclAFFBmbUBrsR2ZLrcbKMJ2Fv11pGYZcqhxs~0cPy0FN1mD30E8WQzVtGuQTFWTTBO9YnmyHmSeuye~fn9tY5RKrE5OKOpEr4wW4gJdrerSk~lcfej7QvJmwc4Qam7RZ1Ql-1P9csfIJ~5iF9T2iGw__&Key-Pair-Id=APKAIE5G5CRDK6RD3PGA)
Fig. 10
The first four eigenvalues of the LBO for different boundary conditions of type II, when and . In the Robin–Robin case (circles) we have , and in the Neumann-Robin case (hollow squares), we have and . In the Robin–Robin case (plain squares), we have and

Fig. 11
Illustration of the proximity of the Robin and pure Neumann eigenvalues in the case of a pure Neumann eigenvalue of multiplicity 1 (a) and a pure Neumann eigenvalue of multiplicity 2 (b)
In particular, we observe that when is small, the eigenvalues are very close to the pure Neumann case. This can actually be predicted using the perturbation theory of linear operators ( 36 ), as presented in Appendix B. In particular, we can show that if is a pure Neumann eigenvalue of multiplicity and is separated from the other pure Neumann eigenvalues by a distance , and represents the ball of radius around , then if is smaller than a certain value there are a finite number of -Robin–Robin eigenvalues with total multiplicity that lie within . This is illustrated in Fig. 11 , where two cases are considered. The case of , which has a pure Neumann eigenvalue of multiplicity 1, and the case of and , that have the same pure Neumann eigenvalue (equal to 2) and hence represent an eigenvalue of multiplicity 2. In both cases, we have .
Moving towards the left on the blue and orange curves in Fig. 11 corresponds to reducing the absolute value of . Hence we can see from Fig. 11 (a) that in the case of pure Neumann eigenvalue of multiplicity 1, for small enough, there is a single Robin–Robin eigenvalue inside , while in the case of the eigenvalue of multiplicity 2 in Fig. 11 (b), for small enough, we always have two Robin–Robin eigenvalues inside , meaning that the finite number of Robin–Robin eigenvalues inside has a total multiplicity of 2, as predicted by the theory in Appendix B.
6. Concluding remarks
The LBO arising from the problem of diffraction by a quarter-plane has been carefully studied, numerically and theoretically, from a spectral point of view. Two types of boundary conditions have been considered, the boundary conditions of type I for which the operator is self-adjoint, and the boundary conditions of type II for which the operator ceases to be self-adjoint but remains m-sectorial. In the case of boundary conditions of type I, we have shown that the spectrum of the operator is an infinite set of isolated real positive eigenvalues with finite multiplicity. The relative position of these eigenvalues for different boundary conditions of type I has been studied. In the case of boundary conditions of type II, we have shown that the spectrum of the operator is an infinite set of isolated complex eigenvalues with finite multiplicity, and that these eigenvalues are contained in a sector of the complex plane. A constructive way of obtaining such a sector has been described. Note that although the theory in Section 4 has been developed for any value of the impedance parameters , most of the numerical results have been given for impedance parameters with a positive real part. This is due to the fact that it becomes difficult to order the eigenvalues for large negative values of this real part. However, our sector estimate gives a good way of reducing the size of the region where we need to search.
Of course, the qualitative results obtained in this study are valid for any length of the cut between 0 and , corresponding to the problem of diffraction by a plane sector of arbitrary angle. These qualitative results should also hold for arbitrary holes on the sphere with mixed boundary conditions, corresponding to the problem of a cone with arbitrary cross section. This knowledge will prove useful when trying to evaluate the diffraction coefficient of such scattering problems, when mixed boundary conditions are being used.
Acknowledgements
The authors are grateful to E. B. Davies, A. C. Hansen, T. D. Young and R. Nürnberg for helpful discussions and comments regarding the subject of this article. R. C. Assier would like to acknowledge the support by UK EPSRC (EP/N013719/1).
References
1.
,
Geometrical theory of diffraction
,
J. Opt. Soc. Am.
52
(
1962
)
116
–
130
.
2.
,
Mathematische Theorie der Diffraction
,
Mathematische Annalen
47
(
1896
)
317
–
374
.
3.
,
A brief historical perspective of the Wiener-Hopf technique
,
J. Eng. Math.
59
(
2007
)
351
–
358
.
4.
,
Diffraction Theory: The Sommerfeld-Malyuzhinets Technique
(
Alpha Science Intl Ltd
2007
).
5.
,
Diffraction by an elliptic cone
,
Commun. Pure Appl. Math.
14
(
1961
)
49
–
68
.
6.
,
Diffraction by a quarter plane, exact solution, and some numerical results
,
IEEE Trans. Antennas Propag.
AP-22
(
1974
)
500
–
503
.
7.
,
Corner diffraction coefficients for the quarter plane
,
IEEE Trans. Antennas Propag.
39
(
1991
)
976
–
984
.
8.
,
Spherical-multipole analysis of electromagnetic and acoustical scattering by a semi-infinite elliptic cone
,
IEEE Antennas Propag. Mag.
38
no.
2
(
1996
)
33
–
44
.
9.
,
Diffraction by a quarter-plane
,
Arch. Ration. Mech. Anal.
8
(
1961
)
139
–
158
.
10.
,
Note on the diffraction at a corner
,
Arch. Ration. Mech. Anal.
19
(
1965
)
62
–
70
.
11.
,
Some solved and unsolved canonical problems of diffraction theory
,
Lect. Notes Math.
1285
(
1987
)
320
–
336
.
12.
,
On Radlow's quarter-plane diffraction solution
,
Radio Sci.
42
(
2007
)
1
–
10
.
13.
,
Diffraction of a plane wave by a sector with Dirichlet or Neumann boundary conditions
,
IEEE Trans. Antennas. Propag.
53
(
2005
)
711
–
718
.
14.
,
Diffraction by conical surfaces at high-frequencies
,
Wave Motion
12
(
1990
)
329
–
339
.
15.
,
On the diffraction of high-frequency waves by a cone of arbitrary shape
,
Wave Motion
21
(
1995
)
203
–
207
.
16.
,
On evaluation of the diffraction coefficients for arbitrary ‘nonsingular’ directions of a smooth convex cone
,
SIAM J. Appl. Math.
60
(
2000
)
536
–
573
.
17.
,
Modified Smyshlyaev's formulae for the problem of diffraction of a plane wave by an ideal quarter-plane
,
Wave Motion
41
(
2005
)
79
–
93
.
18.
,
Coordinate equations for a problem on a sphere with a cut associated with diffraction by an ideal quarter-plane
,
Q. Jl Mech. Appl. Math.
58
(
2005
)
289
–
308
.
19.
,
Diffraction by a finite strip
,
Q. Jl Mech. Appl. Math.
35
(
1982
)
103
–
124
.
20.
,
Embedding formulae in diffraction theory
,
Proc. R. Soc.
A
459
(
2003
)
2475
–
2496
.
21.
,
On the diffraction of acoustic waves by a quarter-plane
,
Wave Motion
49
(
2012
)
64
–
82
.
22.
,
Precise description of the different far fields encountered in the problem of diffraction of acoustic waves by a quarter-plane
,
IMA J. Appl. Math.
77
(
2012
)
605
–
625
.
23.
,
Scattering of acoustic waves by a sector
,
Wave Motion
50
(
2013
)
739
–
762
.
24.
,
Electromagnetic scattering by a plane angular sector: I. Diffraction coefficients of the spherical wave from the vertex
,
Wave Motion
55
(
2015
)
10
–
34
.
26.
,
Diffraction of scalar waves by an impedance cone of arbitrary cross-section
,
Wave Motion
33
(
2001
)
155
–
181
.
27.
,
Advance Theory of the Diffraction by a Semi-Impedance Cone
(
Alpha Science International Ltd
2014
).
28.
,
Excitation, reflection and emission of surface waves from a wedge with given face impedances
,
Soviet Phys. Doklady
3
(
1958
)
752
–
755
.
29.
,
Scattering of Waves by Wedges and Cones with Impedance Boundary Conditions
. In
Mario Boella Series on Electromagnetism in Information & Communication
.
SciTech-IET
,
Edison, NJ
,
2013
.
30.
,
On the diffraction problem of a half plane with different face impedances
,
Can. J. Phys.
62
(
1984
)
853
–
860
.
31.
,
The eigenvalues of the Laplacian on a sphere with boundary conditions specified on a segment of a great circle
,
J. Math. Phys.
38
(
1997
)
1623
–
1649
.
32.
,
The numerical solution of two-parameter eigenvalue problems in ordinary differential equations with an application to the problem of diffraction by a plane angular sector
,
J. Inst. Maths Applics
14
(
1974
)
9
–
22
.
33.
,
Singularities at the tip of a plane angular sector
,
J. Math. Phys.
40
(
1999
)
1087
.
35.
,
Spectral Theory and Differential Operators
. (
Cambridge University Press
1996
).
36.
,
Perturbation Theory for Linear Operators
(
Springer
,
Berlin
1976
).
37.
Linear Operators and their Spectra
(
Cambridge University Press
2007
).
38.
,
Die Greenshe Funktion der Schwingungsgleichung
,
Jber. Deutsch. Math.-Verein.
21
(
1912
)
309
–
353
.
39.
,
Über das asymptotische verhalten der lösungen von Δ
u
+
u
= 0 in unendlichen gebieten
,
Jber. Deutsch. Math.-Verein.
53
(
1943
)
57
–
64
.
40.
,
A uniqueness theorem for the reduced wave equation
,
Commun. Pure Appl. Math.
17
(
1964
)
147
–
176
.
41.
,
The eigenvalues of ∇
2
u
+ λ
u
= 0 when the boundary conditions are given on semi-infinite domains
,
Proc. Cambridge Philosophical Soc.
49
(
1953
)
668
–
684
.
42.
,
Eighty years of Sommerfeld's radiation condition
,
Historia Mathematica
19
(
1992
)
385
–
401
.
44.
,
Analysis on Fractals
(
Cambridge University Press
2001
).
45.
,
Surface finite elements for parabolic equations
,
J. Comput. Math.
25
(
2007
)
385
–
407
.
46.
,
The
deal.II
library, version 8.4
.
J. Numer. Math.
24
(
2016
) to appear.
48.
,
SLEPc: A scalable and flexible toolkit for the solution of eigenvalue problems
.
ACM Trans. Math. Software
31
(
2005
)
351
–
362
.
49.
,
Gmsh: a three-dimensional finite element mesh generator with built-in pre- and post-processing facilities
,
Int. J. Numer. Meth. Eng.
79
(
2009
)
1309
–
1331
.
50.
,
Fast integral equation methods for the Laplace-Beltrami equation on the sphere
,
Adv. Comput. Math.
40
(
2014
)
577
–
596
.
APPENDIX A: GREEN'S IDENTITY FOR
A.1 Green's identity on a sphere with a hole
Consider the surface of a sphere with a hole that represents a certain curved surface
, such that the boundary of the surface,
is oriented as described in
Fig. A.1 (b). In order to apply Stokes' theorem on
, it needs to be oriented, that is, we need to make a choice of normal that is compatible with the right hand rule and the orientation of
. For this to work, we need to choose the normal
of
as in
Fig. A.1 (a), that is
. Let us now consider a vector field
. We can apply Stokes' theorem to
to get
Let us now consider another vector field
. In our case, it can be shown that the divergence of
restricted to
, denoted
, can be expressed as
Now, we can apply Stokes' theorem to the vector field
defined by
and so (
A.1 ) becomes
Now, let us parametrise
by a parameter
say so that
is defined by
in the
basis. Note that because we are on the sphere, we have
. The left-hand side (LHS) of (
A.3 ) can be rewritten as
Upon writing
(representing a vector field tangent to
), and using the vector calculus identity
, we obtain
Note that the quantity
is tangent to
and perpendicular to
, so it represents the normal to
within
, let us call this quantity
, so that we have
and the LHS of (
A.3 ) becomes
Now using the definition of the divergence in (
A.2 ), (
A.3 ) becomes
This is an equivalent of the divergence theorem for this curved surface
. Note that in this case,
is oriented towards the centre of the hole as shown in the
Fig. A.1 (b).

Fig. A.1
Choice of normal to so that it is compatible with the orientation of in view of the application of Stokes' theorem (a) and orientation of normal and tangent vectors and (b)
Note that even if
and
t are three-dimensional vectors, they only depend on
and
and their
component is always zero. Hence they can also be considered as two-dimensional vectors in the
basis, and so, for a function
defined on
, the quantity
makes sense, and using the vector identity
for any
and the divergence theorem derived above, we have Green's identity
A.2 Application to
Let us now consider the cut
as a degenerate hole, where the lower and upper boundaries of the hole have collapsed onto each other. The normal
remains well defined on
, apart from the end points. And in fact this normal is constant on
, where it is denoted
and on
, where it is denoted
as shown on
Fig. 3 (a). In fact, we clearly have
and
. We can now parametrise
and
with the position vectors
and
and the parameter
and
by
for
and
for
. Hence for
, the Green's identity becomes
where the line integrals are defined by
and
In particular, for any function
, here is what is meant by a line integral:
APPENDIX B: PERTURBATION THEORY APPLIED TO THE LBO
The aim of this appendix is to show that for impedances of small enough modulus, the eigenvalues of the LBO with boundary conditions of type II are close to the eigenvalues of a given LBO with boundary conditions of type I. In particular, we will show that for small enough impedances,
and
are close to
,
is close to
and
is close to
. However, here we shall only focus on the most general case of
, the other cases being very similar. In order to do so, we shall apply the perturbation theory for linear operators. Let us start by defining a few important notions. Let consider
, and a family
of sesquilinear forms acting on a Hilbert space
. We say that the family
is a holomorphic family of forms if (i)
is sectorial, densely defined and closed with domain independent of
,
and (ii) for each
, the function of
defined by
is a holomorphic function for
. Also, we need to define the operator norm
as follows. Let
be an operator acting on
with domain
, then the operator norm of
is given by
The main result of this section will be a consequence of the following theorem:
Let us consider the LBO with Robin boundary conditions of type II on both sides of the cut. Let the two closed sesquilinear forms
and
with
be defined as in
Section 4.3 for
by
Recall that
is also symmetric and non-negative and associated with a self-adjoint operator
. Let us define the family of forms
for
by
and
. Note that for
,
is associated with a LBO with boundary conditions of type II,
with impedances
and
. Recall that we have proven in
Section 4.3 that
is relatively bounded with respect to
with
, for
and
given in (
4.8 ). Hence we can apply Theorem B.1, with
,
and
,
,
and
for
to conclude that the family
forms a holomorphic family of forms and is associated with a family of m-sectorial operators
for
(note from (
4.5 ) that
can be made as big as possible, so this is in fact true for all
). Moreover, since
is symmetric, we know that for any
, the resolvent
exists and is a convergent power series for
Now, let
be an eigenvalue of
. As has been shown in
Section 3.3 , we know that
is real, positive, isolated and has a finite multiplicity, say
. Let
be such that
for all
. Let
be the circle of centre
and radius
so that
and let
, we shall note
the resulting ball. According to ((
36 ), V. Eqn. (4.9)), because
is self-adjoint, we know that
where we have used the fact that for
,
Hence, provided we have
, the bound (
B.1 ) is satisfied, and
is well defined for all
,
. We can then define the well studied spectral projection operator
by
Note that since
,
separates its spectrum, and we can write
, with
where
and
. Moreover (see (
37 ), Thm. 11.1.5), we know that
is idempotent and that the restriction of
to the range of
(denoted
) has spectrum
. And hence, if
, then
is made of finitely many eigenvalues with total multiplicity
. But because the projections
depend analytically on
, we have
by ((
37 ), Lemma 1.5.5), and so we have proved the following theorem:
APPENDIX C: FINITE ELEMENT IMPLEMENTATION
In order to obtain the finite element formulation of our eigenvalue problem, we want to express our solution
as a finite linear combination of shape functions
,
,
where
are constant. Let us now consider one particular shape function
and take the inner product of it with (
2.4 ),
. Using Green's identity (
A.5 ) for
, this becomes
where
. Using the sum decomposition (
C.1 ) and the linearity of the inner products in their first argument, we get
So if we define the
matrices
and
by
the expansion (
C.2 ) can be rewritten as
which, upon defining the vector
becomes
, which is a typical generalised eigenvalue problem. The matrix
is called the
stiffness matrix , while the matrix
is called the
mass matrix . Note that
and
being complex in general, the matrix
A and hence the vector
and eigenvalues
can be complex.
Published by Oxford University Press 2016.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (
http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.