Eigenvalues of elliptic operators with density
Abstract.
We consider eigenvalue problems for elliptic operators of arbitrary order subject to Neumann boundary conditions on bounded domains of the Euclidean -dimensional space. We study the dependence of the eigenvalues upon variations of mass density and in particular we discuss the existence and characterization of upper and lower bounds under both the condition that the total mass is fixed and the condition that the -norm of the density is fixed. We highlight that the interplay between the order of the operator and the space dimension plays a crucial role in the existence of eigenvalue bounds.
Key words and phrases:
High order elliptic operators, eigenvalues, mass densities, eigenvalue bounds, Weyl eigenvalue asymptotics2010 Mathematics Subject Classification:
Primary: 35P15. Secondary: 35J40, 35P20.1. Introduction
We consider the eigenvalue problem
(1.1) |
on a connected bounded open subset of , where is a positive function bounded away from zero and infinity and where we impose Neumann boundary conditions on . Under suitable regularity assumptions on the boundary of (e.g., if has a Lipschitz boundary) it is standard to prove that problem (1.1) admits an increasing sequence of non-negative eigenvalues of finite multiplicity
where denotes the dimension of the space of polynomials of degree at most in .
In this paper we will prove a few results on the dependence of the eigenvalues upon variation of . In particular we will consider the problem of finding upper bounds for among all positive and bounded densities satisfying suitable constraints. We shall also consider the issue of lower bounds, which also presents some interesting features.
Keeping in mind important problems for the Laplace and the biharmonic operators in linear elasticity (see e.g., [16]) we shall think of the weight as a mass density of the body and we shall refer to the quantity as to the total mass of . In fact, when the eigenvalues describe the vibrations of a non-homogeneous membrane with free edge when (see e.g., [28, § 9]) and of a non-homogeneous plate with free edge when (see [9, 12]).
Relevant questions on the dependence of the eigenvalues upon are whether it is possible to minimize or maximize the eigenvalues under the assumption that the total mass is fixed, or whether it is possible to have uniform upper or lower bounds for the eigenvalues (i.e., bounds which depend only on the total mass, the dimension and the eigenvalue index) under the same constraint, and which have the correct behavior in as described by the Weyl’s asymptotic law.
Most of the existing literature treats the case of the Laplace operator with Dirichlet boundary conditions. In particular, we recall the famous result of Krein [31] on the eigenvalues of the Dirichlet Laplacian in one dimension (fixed string) which completely answers the questions raised above. In fact he finds sharp upper and lower bounds which depend only on for all the eigenvalues of the Laplacian on the string upon densities for which is fixed (see Remark 3.24). We refer also to the extensive work of Banks and collaborators for generalizations and extensions of Krein’s results (see [2, 3, 5, 6, 26] and the references therein). We mention also [21, § 5] which contains a detailed analysis of the eigenvalues of Sturm-Liouville problems with Dirichlet conditions with density (and also other types of weight). In particular, in [21, § 5], the authors provide estimates (upper and lower bounds) under various type of linear and non-linear constraints on the weights. Existence of minimizers and maximizers under mass constraint in higher dimensions for the Dirichlet Laplacian has been investigated in [11, 17, 18, 19, 24], where the authors impose the additional constraint that admissible densities are uniformly bounded from below and above by some fixed constants. We refer to [28, § 9] and to the references therein for further discussions on eigenvalue problems for inhomogeneous strings and membranes with fixed edges.
As for Neumann boundary conditions, much less is known. Very recently the problem of finding uniform upper bounds for the Neumann eigenvalues of the Laplacian with weights has been solved (for ) by Colbois and El Soufi [14] in the more general context of Riemannian manifolds, by exploiting a general result of decomposition of a metric measure space by annuli (see [27], see also [29]). The authors have not considered the case and, in fact, as we shall see in the present paper, upper bounds with mass constraint do not exists in dimension one.
There are very few results for what concerns higher order operators. We recall [4, 35], where the authors consider the case of the biharmonic operator in one dimension with intermediate boundary conditions (hinged rod) and [7], where the author considers the case of the biharmonic operator with Dirichlet conditions in dimension one (clamped rod) and two (clamped plate). We also refer to [21, § 7.9] where it is possible to find some estimates for the eigenvalues of elliptic operators of order with density subject to Dirichlet boundary conditions. We refer again to [28, § 11] for a more detailed discussion on eigenvalue problems for inhomogeneous rods and plates with hinged and clamped edges. Up to our knowledge, there are no results in the literature on the existence and characterization of upper and lower bounds with respect to mass densities for higher order operators subject to Neumann boundary conditions (already for the biharmonic operator or the Laplacian in dimension one).
Finally, we refer to [32] where the authors prove continuity and differentiability results for the dependence of the eigenvalues of a quite wide class of higher order elliptic operators and homogeneous boundary conditions upon variation of the mass density and in most of the cases (except, again, that of Neumann boundary conditions), they establish a maximum principle for extremum problems related to mass density perturbations which preserve the total mass. We remark that in [32] partial results are obtained in the case of Neumann boundary conditions only for the Laplace operator.
In this paper we shall primarily address the issue of finding upper bounds for the eigenvalues of the polyharmonic operators with Neumann boundary conditions which are consistent with the power of in the Weyl’s asymptotic formula (see (2.9)), among all densities which satisfy a suitable constraint. In particular, we consider two very natural constraints: and . This second constraint arises naturally since it is well-known (see e.g., [23]) that if we set , then . This means that describes the asymptotic distribution of the eigenvalues of problem (1.1) (and in particular implies the Weyl’s law (2.9)). Most of the literature mentioned above considers only the fixed mass constraint.
In view of the physical interpretation of problem (1.1) when and or , it is very natural to ask whether it is possible to redistribute a fixed amount of mass on a string (of fixed length) or on a membrane (of fixed shape) such that all the eigenvalues become arbitrarily large when the body is left free to move, or, on the contrary, if there exists uniform upper bounds for all the eigenvalues. As highlithed in [14], uniform upper bounds with mass constraint exist if . In this paper, by using the techniques of [14] we prove that if , uniform upper bounds exist (see Theorem 3.4), namely we prove that if
(1.2) |
where depends only on and . Surprisingly, in lower dimensions, uniform upper bounds do not hold. In fact we find explicit examples of densities with fixed mass and arbitrarily large eigenvalues (see Theorem 3.29). In this case, however, we are able to find upper bounds which depend also on (see Theorem 3.22), namely we prove that if
(1.3) |
where again depends only on and and the exponent of is sharp. We remark that this inequality holds when for , and it is the analogue of the upper bounds (3.25) proved by Krein [31] for the Dirichlet Laplacian on an interval (up to a universal constant). We note that in order to prove that certain eigenvalue bounds under some natural constraints do not hold, one has to provide counterexamples. It is then natural to ask whether it is possible to find ‘weaker’ bounds which include the correct quantities that explain the counterexamples. This is the case of the bounds (1.3).
We note that the interplay between the dimension of the space and the order of the operator plays a crucial role in the existence of uniform upper bounds for the eigenvalues of problem (1.1) under mass constraint. We can summarize our first result in this way:
“If there exist uniform upper bounds with mass constraint for all the eigenvalues of (1.1), while if we can always redistribute a fixed amount of mass such that all the eigenvalues of (1.1) become arbitrarily large”.
As for the the non-linear constraint , in view of the fact that , it is natural to ask whether upper bounds of the form
(1.4) |
hold. We will call bounds of the form (1.4) “Weyl-type bounds”. Clearly, for inequality (1.4) is equivalent to (1.2). For we are able to find densities with fixed -norm and which produce arbitrarily large eigenvalues (see Theorem 4.16). However, we are able to prove upper bounds for all the eigenvalues which involve both and (see Theorem 4.11), namely we prove that if then
(1.5) |
where depends only on and and the exponent of is sharp. Since (1.4) holds for we are led to conjecture that it must hold for any . We are still not able to prove (1.4) for , and actually it seems to be a quite difficult issue. However we can prove the weaker inequality
(1.6) |
We leave the validity of (1.4) for as an open question. We refer to Remark 4.10 where we discuss relevant examples in support of the validity of our conjecture. In particular, we note that if (1.4) holds true for , when and we would find uniform upper bounds for the eigenvalues of the Neumann Laplacian under the constraint that . We can summarize our second result as follows:
“If , we can always find a density with fixed such that all the eigenvalues of (1.1) are arbitrarily large, while we have uniform Weyl-type upper bounds when . We conjecture the existence of uniform Weyl-type upper bounds when ”.
We also mention [28, § 9.2.3] where it is considered a spectral optimization problem for the Dirichlet Laplacian with the non-linear constraint , where and (see also [21, § 5]).
We have also considered the issue of lower bounds and we have found that ‘surprisingly’ the interplay between the space dimension and the order of the operator plays a fundamental role also in the existence of lower bounds. In fact we are able to prove the following facts (see Theorems 5.1,5.4,5.13 and 5.15):
“If there exists a positive constant which depends only on and such that the first positive eigenvalue of problem (1.1) is bounded from below by , while if , for all we can always redistribute a fixed amount of mass such that the first eigenvalues of (1.1) are arbitrarily close to zero”
and
“If there exists a positive constant which depends only on and such that the first positive eigenvalue of problem (1.1) is bounded from below by , while if for all we can always find densities with fixed -norm such that the first eigenvalues of (1.1) are arbitrarily close to zero”.
We note that lower bounds for the first eigenvalue under one of the two constraints exist in the case that upper bounds with the same constraint do not exist. We remark that the situation is very different if we consider for example the issue of the minimization of the eigenvalues of (1.1) with among all bounded domains with fixed measure: it is standard to prove that there exist domains with fixed volume and such that the first eigenvalues can be made arbitrarily close to zero, in any dimension .
Finally we remark that all the results of this paper can be adapted to the more general eigenvalue problem
with Neumann boundary conditions, where is defined by
and is an elliptic operator of order , under suitable assumptions on the domain and the coefficients of . We refer to [32] for a detailed description of eigenvalue problems for higher order elliptic operators with density (see also [21, § 7]).
The present paper is organized as follows: Section 2 is dedicated to some preliminaries. In Section 3 we consider the problem of finding uniform upper bounds with mass constraint. In particular in Subsection 3.1 we prove uniform upper bounds (1.2) for , in Subsection 3.2 we prove upper bounds (1.3) for and in Subsection 3.3 we provide counterexamples to uniform upper bounds in dimension . In Section 4 we investigate the existence of upper bounds with the non-linear constraint . In particular in Subsections 4.1 and 4.2 we prove upper bounds (1.6) and (1.5), respectively, while in Subsection 4.3 we provide counterexamples to uniform upper bounds (1.4) for . In Subsection 4.1 we state the open question whether bounds of the form (1.4) hold if . In Section 5 we consider lower bounds and in particular we discuss how the constraint, the space dimension and the order of the operator influence their existence. At the end of the paper we have two appendices, Appendix A and Appendix B. In Appendix A we discuss Neumann boundary conditions for higher order operators and develop some basic spectral theory for such operators. In Appendix B we prove some useful functional inequalities which are crucial in the proof of Theorem 3.29 in Subsection 3.3.
2. Preliminaries and notation
Let be a bounded domain (i.e., an open connected bounded set) of . By we shall denote the standard Sobolev space of functions in with weak derivatives up to order in , endowed with its standard norm defined by
for all , where
In what follow we will use the standard multi-index notation. Hence, for , , we shall denote by the quantity . Moreover, for , and . For , , we will write . For a function of class , we write . Finally, for a function and , we shall write to denote the -th derivative of with respect to .
In the sequel we shall assume that the domain is such that the embedding of into is compact (which is ensured, for example, if is a bounded domain with Lipschitz boundary). By we shall denote the subset of of those functions such that .
We shall consider the following eigenvalue problem:
(2.1) |
in the unknowns (the eigenfunction), (the eigenvalue), where
We note that problem (2.1) is the weak formulation of the following eigenvalue problem:
(2.2) |
in the unknowns and . Here are uniquely defined ‘complementing’ boundary operators of degree at most (see [25] for details), which we will call Neumann boundary conditions (see Appendix A.1).
Example 2.3.
If , and (2.2) is the classical formulation of the Neumann eigenvalue problem for the Laplace operator, namely
(2.4) |
in the unknowns and , while if we have the Neumann eigenvalue problem for the biharmonic operator, namely
(2.5) |
in the unknowns and . Here denotes the tangential divergence operator on (we refer to [20, § 7] for more details on tangential operators).
In Appendix A.1 we discuss in more detail boundary conditions for problems (2.2) and (2.5) and, more in general, Neumann boundary conditions for the polyharmonic operators.
It is standard to prove (see Theorem A.5) that the eigenvalues of (2.1) are non-negative, have finite multiplicity and consist of a sequence diverging to of the form
where
The eigenfunctions associated with the eigenvalue are the polynomials of degree at most in (the dimension of the space spanned by the polynomials of degree at most in is exactly ). We note that we have highlithed the dependence of the eigenvalues upon the density , which is the main object of study of the present paper.
By standard spectral theory, we deduce the validity of the following variational representation of the eigenvalues (see [16, § IV] for more details):
Theorem 2.6.
Let be a bounded domain in such that the embedding is compact. Let . Then for all we have
(2.7) |
We conclude this section by recalling the asymptotic behavior of the eigenvalues as , which is described by the Weyl’s law.
Theorem 2.8.
Let be a bounded domain in with Lipschitz boundary. Let . Then
(2.9) |
as .
3. Upper bounds with mass constraint
In this section we consider the problem of finding uniform upper bounds for the -th eigenvalue among all mass densities which preserve the mass (that is, among all such that ), and which show the correct growth in the power of with respect to the Weyl’s law (2.9). In particular, in Subsection 3.1 we prove that such bounds exist if (see Theorem 3.4), while in Subsection 3.3 we will give counter-examples in dimension (see Theorem 3.29). Moreover, in Subsection 3.2 we establish upper bounds in the case which involve also a suitable power of (see Theorem 3.22) which turns out to be sharp.
3.1. Uniform upper bounds with mass constraint for
In this subsection we will prove the existence of uniform upper bounds for with respect to mass preserving densities.
The main tool which we will use is a result of decomposition of a metric measure space by annuli (see [27, Theorem 1.1]). We recall it here for the reader’s convenience.
Let be a metric space. By an annulus in we mean any set of the form
where and . By we denote
We are ready to state the following theorem (see [27, Theorem 11]):
Theorem 3.1.
Let be a metric space and be a Radon measure on it. Assume that the following properties are satisfied:
-
i)
there exists a constant such that any metric ball of radius can be covered by at most balls of radius ;
-
ii)
all metric balls in are precompact sets;
-
iii)
the measure is non-atomic.
Then for any integer there exist a sequence of annuli in such that, for any
and the annuli are pairwise disjoint. The constant depends only on .
Corollary 3.2.
We are now ready to state the main result of this section.
Theorem 3.4.
Let be a bounded domain in , , such that the embedding is compact. Let . Then for every we have
(3.5) |
where is a constant which depends only on and .
Remark 3.6.
Inequality (3.5) says that there exists a uniform upper bound for all the eigenvalues with respect to those densities which give the same mass . We note that the quantity is an average density, i.e., the total mass over the total volume of . Moreover, from (3.5) it follows that
for all densities and bounded domains (with compact) such that .
Proof of Theorem 3.4.
The proof is based on the general method described by Grigor’yan, Netrusov and Yau in [27] (see Theorem 3.1; see also [14, 15] for the case of the Laplace operator). In particular, we will build a suitable family of disjointly supported test functions with controlled Rayleigh quotient.
Let . Let be defined by
(3.7) |
The coefficients , are uniquely determined by the equations
(3.8) |
We note that (3.8) can be written as
(3.9) |
which is a system of equations in unknowns , . It is standard to see that (3.9) admits an unique non-zero solution. Moreover, the coefficients , depend only on .
We note that by construction . Let now be an annulus in . We define a function supported on and such that on and on by setting
(3.10) |
By construction, the restriction of this function to belongs to the Sobolev space . Now we exploit Theorem 3.1 with endowed with the Euclidean distance, and the measure given by for all measurable . The hypothesis of Theorem 3.1 are clearly satisfied. Hence, for each index we find annuli such that are disjoint and
where depends only on . Since we have annuli , , we can choose of such annuli, say such that
(3.11) |
for all . To each of such annuli, we associate a function defined by
(3.12) |
We have then built a family of disjointly supported functions, which we relabel as , and whose restriction to belong to the space .
From the min-max principle (2.7) it follows that
(3.13) |
where is the subspace of generated by (and which has dimension ). Since the space is generated by disjointly supported functions, it is standard to prove that (3.13) is equivalent to the following:
(3.14) |
This means that it is sufficient to have a control on the Rayleigh quotient of each of the generating functions in order to bound . It is standard to see that, if is given by , for some function of one real variable, then for all with
(3.15) |
where depends only on , and . From (3.15) it follows then that there exists a constant which depends only on and such that
(3.16) |
Through the rest of the proof we will denote by a positive constant which depends only on and and which can be eventually re-defined line by line.
By standard approximation of functions in by smooth functions (see [22, § 5.3]), from (3.16) we deduce that if is given by , for some function of one real variable, then
(3.17) |
for all . We are now ready to estimate the right-hand side of (3.14). For the denominator we have
(3.18) |
This follows from the fact that and on and from Theorem 3.1. For the numerator, since , we have
(3.19) |
From (3.11), we have that . Moreover, from (3.7), (3.17) and standard calculus we have that
(3.20) |
From (3.18), (3.19) and (3.20) we have that
(3.21) |
for all . From (3.14) and (3.21) we deduce the validity of (3.5). This concludes the proof. ∎
3.2. Upper bounds with mass constraint for
We note that the proof of Theorem 3.4 does not work in the case . Indeed, as we will see in Subsection 3.3 (see Theorem 3.29), bounds of the form (3.5) do not hold if . In this subsection we prove upper bounds for the eigenvalues which involve also a suitable power of , namely, we prove the following theorem:
Theorem 3.22.
Let be a bounded domain in , , such that the embedding is compact. Let . Then for every we have
(3.23) |
where depends only on and .
Remark 3.24.
We recall the following well-known result by Krein [31] which states that in the case of the equation on with Dirichlet boundary conditions, we have
(3.25) |
which is the analogous of (3.23) for the Laplace operator () in dimension . Actually, inequality (3.25) is sharp, i.e., for all there exists such that the equality holds in (3.25) when .
Proof of Theorem 3.22.
In order to prove (3.23) we exploit in more detail Theorem 3.1 and Corollary 3.2. As in the proof of Theorem 3.4, for any measurable , let . By following the same lines of the proof of Theorem 3.4, we find for each , annuli such that the annuli are disjoint, for all , where depends only on , and moreover for all .
Let and denote the inner and outer radius of , respectively ( denotes the radius of if is a ball). Associated to each annulus we construct a test function supported on and such that on , and which satisfies
and
where depends only on and (see (3.7), (3.10), (3.12) and (3.17)). In what follows we shall denote by a positive constant which depends only on and and which can be eventually re-defined line by line. Then, if is an annulus of inner radius or a ball of radius , we have
(3.26) |
for all .
We note that Corollary 3.2 provides an estimate for the inner radius of the annuli given by the decomposition of Theorem 3.1 (and, respectively, an estimate of the radius of the ball in the case that the decomposition of the space produces a ball). In particular, if is the inner radius of , we have for all
where and . Let . If , then
where denotes the volume of the unit ball in . This means that
for all . Hence, if , then and therefore
(3.27) |
for all . From (3.26) and (3.27) it follows that
for all , which implies (3.23) by (2.7) and by the fact that are disjointly supported (see also (3.13) and (3.14)). This concludes the proof.
∎
3.3. Non-existence of uniform upper bounds for and sharpness of the exponent of in (3.23)
In this subsection we will prove that if , there exist families such that as and for all as , and moreover we will provide the rate of divergence to of the eigenvalues with respect to the parameter . This means that in dimension we can redistribute a bounded amount of mass in in such a way that all the positive eigenvalues become arbitrarily large. This is achieved, for example, by concentrating all the mass at one point of . Moreover, the families considered in this subsection will provide the sharpness of the power of in (3.23).
Through all this subsection, will be a bounded domain in with Lipschitz boundary. Assume without loss of generality that and let be such that for all (all the result of this section hold true if we substitute with any other ). For all let be defined by
(3.28) |
for some , which we fix once for all and which can be chosen arbitrarily close to zero. We have the following theorem:
Theorem 3.29.
Let be a bounded domain in , , with Lipschitz boundary. Let be defined by (3.28) for all . Then
-
i)
;
-
ii)
for all , , there exists which depends only on , , and such that (up to subsequences) (for all ).
From Theorem 3.29 it follows that for and
Moreover, since , it follows that the power of in (3.23) is sharp.
We remark that the proof of Theorem 3.29 requires some precise inequalities for function in and which we prove in Lemmas B.5 and B.13 of the Appendix B.
Proof of Theorem 3.29.
The proof of point is a standard computation. In fact
We prove now . In order to simplify the notation, from now on we will denote an eigenvalue simply as . The proof of is divided into two steps. In the first step we will prove that there exists a positive constant which depends only on , , and such that . In the second step we will prove that there exists a positive constant which depends only on , , and such that . This yields the result (up to choosing a suitable subsequence of ).
Step 1. We note that , hence for all ,
(3.30) |
By taking the minimum and the maximum into (3.30), by (2.7) we have that for all . Hence .
Step 2. We introduce the function defined by
(3.31) |
We note that is an eigenvalue of (2.1) with if and only if is an eigenvalue of problem (2.1) with , where is defined by (3.31).
We prove now that for all there exist which depends only on and such that for all . This implies the existence of constants such that .
We recall that the first positive eigenvalue is . From the min-max principle (2.7) we have
where
We argue by contradiction. Assume that as . Let be an eigenfunction associated with normalized by . Clearly (see formula (3.31)) and since as , we have that the sequence is bounded in . Then there exists such that, up to subsequences, in and in as by the compactness of the embedding .
Since , it is standard (see e.g., [22, § 5.8]) to prove that in for all , and hence for some constants . This means that is a polymonial of degree at most . Moreover, from Lemma B.13, point and , it follows that for all , where if is odd, and if is even. Hence for all . Then
(3.32) |
Moreover, from Lemma B.13, points and it follows that for all which implies along with (3.32) that .
Again, from Lemma B.13, and , it follows that as , which yields the contradiction. This concludes the proof.
∎
4. Weyl-type upper bounds
In this section we investigate the existence of uniform upper bounds for which are compatible with the Weyl’s law (2.9), namely we look for uniform upper bounds of the form
(4.1) |
where the constant depends only on and . Actually we will not prove (4.1), but a weaker form involving also in the case (see Theorem 4.2). We remark that in the case , the bounds (4.1) hold, in fact this is already contained in Theorem 3.4. Moreover, we shall prove upper bounds involving in the case (see Theorem 4.11) in which the power of turns out to be sharp (see Theorem 4.16). In particular this implies that bounds of the form (4.1) do not hold if . We will be left with the open question (see Remark 4.10) of the existence of bounds of the form (4.1) in the case .
4.1. Upper bounds for
In this subsection we prove upper bounds of the form (4.1) in the case involving a certain power of , namely, we prove the following theorem:
Theorem 4.2.
Let be a bounded domain in , , such that the embedding is compact. Let . Then there exists a constant which depends only on and such that for all it holds
(4.3) |
Proof.
First, we remark that (4.3) with has already been proved in Theorem 3.4. Hence, from now on we let .
The proof is very similar to that of Theorem 3.4. It differs from the choice of the measure in Theorem 3.1. In fact, we exploit Theorem 3.1 with endowed with the Euclidean distance and defined by for all measurable .
The hypothesis of Theorem 3.1 are clearly satisfied. Then for each index we find metric annuli such that are disjoint,
where depends only on , and such that
(4.4) |
for all .
As in the proof of Theorem 3.4, for all , we define a function supported on and such that on and on by setting
where is given by (3.7). By construction, the restriction of this function to belongs to the Sobolev space . In order to simplify the notation, we will set .
We have then built a family of disjointly supported functions belonging to the space . We estimate now the Rayleigh quotient of for all . For the denominator we have
(4.5) |
This follows from the fact that and on , from Hölder’s inequality and from Theorem 3.1. For the numerator we have
(4.6) |
where is a constant which depends only on and . From now on we shall denote by a positive constant which depends only on and and which can be eventually re-defined line by line. Assume now that is a ball of center and radius . From (4.5) and (4.6) we have that
(4.7) |
Assume now that is a proper annulus (i.e., ). From Corollary 3.2 it follows that for all
where and . As in the proof of Theorem 3.22, we see that
(4.8) |
where the second inequality follows from (4.4). By combining (4.5), (4.6) and (4.8) we obtain
(4.9) |
By combining (4.7) and (4.9) and by the fact that for all , we obtain (4.3) thanks to (2.7) (see also (3.13) and (3.14)). This concludes the proof.
Remark 4.10.
From Theorem 4.2 it naturally arises the question whether bounds of the form (4.1) hold in the case . We conjecture an affirmative answer. In fact, in order to produce a family of densities such that as , a necessary condition is that for almost every (otherwise we will find a subset of positive measure where for all and construct suitable test functions supported in which can be used to prove upper bounds for all the eigenvalues independent of as is done in Theorem 3.4). Hence, concentration phenomena are the right candidates in order to produce the blow-up of the eigenvalues. We may think to very simple toy models, like concentration around a point or in a neighborhood of the boundary (or in general, in a neighborhood of submanifolds contained in ). It is possible, for example, to show that if we concentrate all the mass in a single point, then the eigenvalues remain bounded (one can adapt the same arguments used in the proof of Theorem 3.29 or explicitly construct test functions for the Rayleigh quotient). If we concentrate all the mass in a neighborhood of the boundary, it is possible to prove that as (see Theorem 5.15 here below). These two types of concentration are somehow the extremal cases of mass concentration around submanifolds contained in .
4.2. Upper bounds for
In this subsection we prove upper bounds for the eigenvalues which involve a suitable power of . We have the following theorem:
Theorem 4.11.
Let be a bounded domain in , , such that the embedding is compact. Let . Then there exists a constant such that for all
(4.12) |
4.3. Non-existence of Weyl-type upper bounds for and sharpness of the exponent of in (4.12)
In this subsection we prove that for there exist sequences such that as , and for all as , and we also provide the rate of divergence to of the eigenvalues with respect to . This means that if bounds of the form (4.1) do not hold. This result can be achieved, for example, by concentrating all the mass in a neighborhood of the boundary. Thus, mass densities with fixed -norm and which concentrate on particular submanifolds may produce blow-up of the eigenvalues if . Moreover the families of densities considered in this subsection will provide the sharpness of the power of in (4.12).
Through all this subsection will be a bounded domain in of class . Let
(4.14) |
be the -tubular neighborhood of . Since is of class it follows that there exist such that for all , each point in has a unique nearest point on (see e.g., [30]). For all let be defined by
(4.15) |
We have the following theorem:
Theorem 4.16.
Let be a bounded domain in , with , of class . Let be defined by (4.15) for all . Then
-
i)
;
-
ii)
for all , , there exists which depends only on , , and such that .
In order to prove Theorem 4.16 we will exploit a result on the convergence of the Neumann eigenvalues of the polyharmonic operator to the corresponding Steklov eigenvalues.
The weak formulation of the polyharmonic Steklov eigenvalue problem reads:
(4.17) |
in the unknowns (the eigenfunction), (the eigenvalue). It is standard to prove that the eigenvalues of (4.17) are non-negative and of finite multiplicity and are given by
We refer e.g., to [9] for a more detailed discussion on the Steklov eigenvalue problem for the biharmonic operator. We have the following theorem:
Theorem 4.18.
We refer to [33, 34] for the proof of Theorem 4.18 in the case of the Laplace operator and to [9] for the proof of Theorem 4.18 in the case of the biharmonic operator, and for more information on the convergence of Neumann eigenvalues to Steklov eigenvalues via mass concentration to the boundary. We remark that the proof of Theorem 4.18 for all values of follows exactly the same lines as the proof of the case and .
Proof of Theorem 4.16.
We start from point . It is standard to see that . The first summand goes to zero as . For the second summand we note that since is of class , it is standard to prove that as . This concludes the proof of point .
We consider now point . We note that for all , , where
We note that is an eigenvalue of (2.1) with if and only if is an eigenvalue of problem (2.1) with . Problem (2.1) with admits an increasing sequence of non-negative eigenvalues of finite multiplicity given by
∎
5. Lower bounds
In this last section we shall discuss the issue of the lower bounds. In many situations (e.g., shape optimization problems) the problem of minimization of the eigenvalues leads to trivial solutions in the case of Neumann boundary conditions. Nevertheless, the eigenvalue problems with density which we have considered in this paper show an interesting behavior with respect to lower bounds, both if we fix the total mass or the -norm of the density. In the first case, we are able to show that there exist densities which preserve the total mass for which the -th eigenvalue can be made arbitrarily close to zero if (which is the case when upper bounds with mass constraint exist). This is stated in Theorem 5.4. On the contrary, if , the first positive eigenvalue is uniformly bounded from below by a positive constant which depends only on , and divided by the total mass (in this case we recall that upper bounds with mass constraint do not exist). This is stated in Theorem 5.1.
When we choose as a constraint the -norm of the density, we see that exactly the opposite happens: if we find densities with prescribed -norm such that the -th eigenvalue can be made arbitrarily close to zero (in this case we have conjectured the existence of upper bounds of the form (4.1)), see Theorem 5.15; if , then the first positive eigenvalue is uniformly bounded from below by a positive constant which depends only on , and divided by the -norm of the density, and in this case we recall that upper bounds with constraint do not exist. This is stated in Theorem 5.13.
We present now the precise statements and the corresponding proofs of such phenomena.
We start with the following theorem concerning lower bounds with mass constraint:
Theorem 5.1.
Let be a bounded domain in , , with Lipschitz boundary. Then there exists a positive constant which depends only on , and such that for every
(5.2) |
Densities which preserve the total mass and produce arbitrarily small eigenvalues can be given, for example, by concentrating all the mass around distinct points of . For all let us fix once for all points and a number such that and are disjoint. For , we will write . Let defined by
(5.3) |
We have the following theorem:
Theorem 5.4.
Let be a bounded domain in , with , such that the embedding is compact. Let be defined by (5.3) for all and . Then
-
i)
;
-
ii)
if ;
-
iii)
if ,
where are positive constants which depend only on and respectively.
Proof.
We start with point . We have
which yields the result.
We prove now . Let and let us fix . Let , and be as in the definition of in (5.3). Let . Associated to each we construct a function which is supported in and such that on in the following way:
where
The coefficients , are uniquely determined by imposing , , for all and depend only on (see also (3.7), (3.8) and (3.9) in the proof of Theorem 3.4).
Now we estimate the Rayleigh quotient of , for all . We start from the numerator. As in the proof of Theorem 3.4 (see (3.19) and (3.20)) we have that
(5.5) |
where depends only on and and can be eventually re-defined through the rest of the proof. For the denominator we have
(5.6) |
From (5.5), (5.6) and the min-max principle (2.7) and from the fact that is a set of disjointly supported functions, it follows that
(see also (3.13), (3.14) in the proof of Theorem 3.4). This concludes the proof of .
Consier now . Let . Again, let us fix . Let , and be as in the definition of (5.3) (we note that admissible values for depend on and ). Associated to each we construct for all a function which is supported in in the following way:
where
and
where the coefficients and are uniquely determined by imposing for all . Moreover (possibly choosing a smaller value for ), it is standard to prove that there exist positive constants and which depend only on and (and hence on , and ) such that , for all and , and that and . In particular (possibly re-defining the constants and and choosing a smaller value for ), we have that for all .
From the min-max principle (2.7) and from the fact that is a set of disjointly supported functions, it follows that
(5.7) |
(see also (3.13), (3.14) in the proof of Theorem 3.4). It remains then to estimate the Rayleigh quotient of all the function . We have for the denominator
(5.8) |
From now on will denote a positive constat which depends only on and . For the numerator, we have, since the functions are radial with respect to (see also (3.17))
(5.9) |
Since and , we have
(5.10) |
Moreover, , hence
(5.11) |
From (5.9), (5.10) and (5.11) we have that
(5.12) |
By combining (5.8) and (5.12), from (5.7) we deduce that . This concludes the proof for the case and of the theorem. ∎
We consider now lower bounds with the non-linear constraint and . We have the following theorem:
Theorem 5.13.
Let be a bounded domain in , , with Lipschitz boundary. Then there exists a positive constant which depends only on , and such that for every
(5.14) |
Densities with prescribed -norm and which made the -th eigenvalue arbitrarily small in dimension are, for example, densities which explode in a -tubular neighborhood of the boundary of . This is contained in the following theorem.
Appendices
Appendix A Eigenvalues of polyharmonic operators
In this section we shall present some basics of spectral theory for the polyharmonic operators. In particular, we will discuss Neumann boundary conditions, mainly for the Laplace and the biharmonic operator. Then we will characterize the spectrum of the polyharmonic operators subjet to Neumann boundary conditions by exploiting classical tools of spectral theory for compact self-adjoint operators. We refer to [21, 32] and to the references therein for a discussion on eigenvalue problems for general elliptic operators of order with density subject to homogeneous boundary conditions.
A.1. Neumann boundary conditions
Neumann boundary conditions are usually called ‘natural’ boundary conditions. This is well understood for the Laplace operator. In fact, assume that is a classical solution of (2.4). If we multiply the equation by a test function and integrate both sides of the resulting identity over , thanks to Green’s formula we obtain:
Hence (2.1) with holds for all when is a solution of (2.2). We can relax our hypothesis on and just require that and that (2.1) holds for all . Hence (2.1) is the weak formulation of the Neumann eigenvalue problem for the Laplace operator. We note that the boundary condition in (2.2) arises naturally and is not imposed a priori with the choice of a subspace of in the weak formulation (as in the case of for Dirichlet conditions): if a weak solution of (2.1) for exists and is sufficiently smooth, then it solves in the classical sense and satisfies the Neumann boundary condition .
Let us consider now more in detail the case of the biharmonic operator. Assume that is a classical solution of problem (2.5). We multiply the equation by a test function and apply the biharmonic Green’s formula (see [1, Lemma 8.56]). We obtain:
where denotes the tangential component of . Hence (2.1) with holds for all when is a solution of problem (2.5) (we remark that if then ). We can relax our hypothesis on and just require that and that (2.1) holds for all . This is exactly the weak formulation of the Neumann eigenvalue problem for the biharmonic operator. We note again that the two boundary conditions in (2.5) arise naturally and are not imposed a priori: if a weak solution of (2.1) exists and is sufficiently smooth, then it satisfies the two Neumann boundary conditions. We also remark that if is sufficiently regular, e.g., if it is of class with and is continuous, then a weak solution of (2.1) with is actually a classical solution of (2.5) (see [25, § 2]). The choice of the whole space in the weak formulation (2.1) contains the information on the boundary conditions in (2.2).
It is natural then to consider problem (2.1) for any as the weak formulation of an eigenvalue problem for the polyharmonic operator with Neumann boundary conditions. In the case of a generic value of it is much more difficult to write explicitly the boundary operators (this is already extremely involved for ). If moreover is sufficiently regular and is continuous, then weak solutions of (2.1) are actually classical solution of (2.2), and the bounday conditions are uniquely determined and arise naturally from the choice of the whole space (see [25] for further discussions on higher order elliptic operators and eigenvalue problems).
A.2. Characterization of the spectrum
The aim of this subsection is to prove that problem (2.1) admits an increasing sequence of non-negative eigenvalues of finite multiplicity diverging to , and to provide some additional information on the spectrum. To do so, we will reduce problem (2.1) to an eigenvalue problem for a compact self-adjoint operator on a Hilbert space.
We define first the following (equivalent) problem: find and such that
(A.1) |
Clearly the eigenfuctions of (A.1) coincide with the eigenfunctions of (2.1), while all the eigenvalues of (2.1) are given by , where is an eigenvalue of (A.1).
We consider the operator as a map from to its dual defined by
The operator is a continuous isomorphism between and . In fact it follows immediately that there exist such that
(A.2) |
Next we denote by the canonical embedding of into and by the embedding of into , defined by
Let be the operator from to itself defined by . Problem (A.1) is then equivalent to
in the unknows , . We now consider the space endowed with the bilinear form
(A.3) |
From (A.2) it follows that (A.3) is a scalar product on whose induced norm is equivalent to the standard one. We denote by the space endowed with the scalar product defined by (A.3). Then we can state the following theorem:
Theorem A.4.
Let be a bounded domain in such that the embedding is compact. Let . The operator is a compact self-adjoint operator in , whose eigenvalues coincide with the reciprocals of the eigenvalues of problem (A.1) for all .
The proof of Theorem A.4 is standard, hence we omit it (see e.g., [8, § IX]). As a consequence of Theorem A.4 we have the following:
Theorem A.5.
Let be a bounded domain in such that the embedding is compact. Let . Then the set of the eigenvalues of (2.1) is contained in and consists of the image of a sequence increasing to . The eigenvalue has multiplicity and the eigenfunctions corresponding to the eigenvalue are the polynomials of degree at most in . Each eigenvalue has finite multiplicity. Moreover the space has a Hilbert basis of eigenfunctions of problem (2.1).
Proof.
We note that , hence by standard spectral theory it follows that the eigenvalues of are positive and bounded and form an infinite sequence converging to zero. Moreover to each eigenvalue is possible to associate an eigenfunction such that is a orthonormal basis of .
From Theorem A.4 it follows that the eigenvalues of (A.1) form a sequence of real numbers increasing to which is given by and that the space has a Hilbert basis of eigenfunctions of (A.1). The eigenvalues of (2.1) are given by for all , where are the eigenvalues of (A.1) and the eigenfunctions associated with coincide with the eigenfunctions associated with . Moreover, given an eigenvalue of (2.1) and a corresponding eigenfunction , we have that
thus . Finally, if , then , thus is a polynomial of degree at most in . The eigenspace associated with the eigenvalue has dimension and coincides with the space of the polynomials of degree at most in . This concludes the proof.
Appendix B A few useful functional inequalities
In this section we will prove some useful functional inequalities which are crucial in the proof of the results of Subsection 3.3, in particular of Theorem 3.29. Since we think that they are interesting on their own, we shall provide here all the details of the proofs. Through this section will be a bounded domain in with Lipschitz boundary. We start this section by recalling the standard Sobolev embeddings.
Theorem B.1.
Let be a bounded domain with Lipschitz boundary. Let and assume that .
-
i)
If then , where
Moreover there exists a positive constant which depends only on and such that
(B.2) -
ii)
If then and
(B.3) the constant depending only on and .
-
iii)
If there exist constants which depend only on and such that
(B.4)
We refer to [10, § 4.6-4.7] and to [22, § 5.6.3] for the proof ot points and of Theorem B.1. We refer to [13, Theorem 1.1] for the proof of (B.4) (see also [10, § 4.7]).
From Theorem B.1 it follows that if then a function is (equivalent to) a function of class . In particular if is odd, we can write for some and a function is (equivalent to) a function of class . If is even, we can write for some and a function is (equivalent to) a function of class for any .
Assume now that a function has all its partial derivatives up to the -th order vanishing at a point . Then the integral of over (where is such that ) can be controlled by if is odd, and by if is even. This is proved in the following lemma, where without loss of generality we set .
Lemma B.5.
Let be a bounded domain in , , with Lipschitz boundary. Assume that and let be such that . Let . Then there exists a positive constant which depends only on and such that
-
i)
if , ;
-
ii)
if , .
Proof.
We start by proving . Let for some . Actually, we will prove for a function . The result for a function will follow from standard approximation of functions in the space by smooth functions (see [10, § 2.3] and [22, § 5.3]). Let then . Through the rest of the proof we shall denote by a positive constant which depends only on and and which we can eventually re-define line by line. From the standard Sobolev embedding theorem, it follows that . From Taylor’s Theorem it follows that
(B.6) |
We consider now the absolute value of the expression in the right-hand side of (B.6) and integrate over each integral which appears in the sum. We have
(B.7) |
where in the last line we have used the Sobolev inequality (B.3) for functions in with .
Next, we estimate the quantity
for a function . First we note that there exits such that
then
which implies
(B.8) |
where the constant depends only on and (see also (B.2)).
Consider now (B.6). We take the squares of both sides and integrate over . We have
(B.9) |
where the last inequality follows from (B.7) and (B.8) and the constant depends only on and . Since inequality (B.9) holds for all , by standard approximation of functions by smooth functions, we conclude that it holds for all . This proves .
Consider now . Let for some . Again, we shall prove for a function . The result for a function will follows from standard approximation of functions in the space by smooth functions.
We prove first the following inequality:
(B.10) |
for all (the constant depending only on and ). In order to prove (B.10) we will use the exponential inequality (B.4) which describes the limiting behavior of the Sobolev inequality (B.3) when for functions in . Let then and let be the constants appearing in (B.4). We have
where in the first inequality we have used the concavity of the logarithm and Jensen’s inequality and in the third inequality we have applied (B.4). Inequality (B.10) is now proved.
Let now . From the Sobolev inequality (B.2), it follows that for all . From Taylor’s Theorem (see also (B.6)) it follows it follows then
(B.11) |
We estimate the integrals appearing in the right-hand side of (B.11)
(B.12) |
where the last inequality follows from (B.10) applied with . From (B.11) and (B.12), immediately follows. This ends the proof of the lemma. ∎
Assume now that is such that for all , that is, for a fixed
for all . In view of Lemma B.5 we expect that the quantities , and can be bounded by the and a suitable power of . In particular we expect that all these quantities vanish as . The aim of the next lemma is to prove that this is exactly what happens. We shall also provide the correct powers of in the estimates which are crucial in the proof of Theorem 3.29.
Lemma B.13.
Let be a bounded domain in , , with Lipschitz boundary. Assume that and let be such that . Let be fixed. Let be such that for all with . Then there exists a positive constant which depends only on and such that
-
i)
for all with , if , ;
-
ii)
for all with , if , ;
-
iii)
if , ;
-
iv)
if , ;
-
v)
for all with , if , ;
-
vi)
for all with , if , .
Proof.
Let be such that for all with . We start by proving and . We recall that by Sobolev inequality (B.2), if , while for all if . We have, for , that
(B.14) |
From (B.14) and from the fact that it follows that
(B.15) |
We compute now . It is convenient to pass to the spherical coordinates , where denotes the unit sphere in endowed with the dimensional volume element . With respect to these new variables, we write the coordinate functions as , where are the standard spherical harmonics of degree in . We have then
where
(B.16) |
From (B.15) it follows that for all
(B.17) |
Clearly
(B.18) |
where depends only on and . Moreover, from Lemma B.5 and Hölder’s inequality it follows that
(B.19) |
if , while
(B.20) |
if . From (B.17), (B.18), (B.19), (B.20) and since , we deduce that
if , while
if . Since for all , by definition (see (B.16)), necessairily inequalities and must hold with a constant which depends only on and .
Consider now . We have
where in the second inequality we have used point of Lemma B.5 and in the last inequality we have used point of the present lemma and the fact that
This concludes the proof of . Point is proved exactly as point , by using point of Lemma B.5 and point of the present lemma.
We consider now point . Let , with and let such that . We have
(B.21) |
From Lemma B.5 point and from Hölder’s inequality, we have that
(B.22) |
Moreover, from point of the present lemma we have that for all
(B.23) |
The proof of follows by combining (B.21) with (B.22) and (B.23). The proof of is identical to that of and follows form point of Lemma B.5 and point of the present lemma. This concludes the proof.
∎
References
- [1] M. J. Arrieta and P. D. Lamberti. Higher order elliptic operators on variable domains. stability results and boundary oscillations for intermediate problems. Online preprint ArXiv:1502.04373v2, 2015.
- [2] D. Banks. Bounds for the eigenvalues of some vibrating systems. Pacific J. Math., 10:439–474, 1960.
- [3] D. Banks. Upper bounds for the eigenvalues of some vibrating systems. Pacific J. Math., 11:1183–1203, 1961.
- [4] D. O. Banks. Bounds for the eigenvalues of nonhomogeneous hinged vibrating rods. J. Math. Mech., 16:949–966, 1967.
- [5] D. O. Banks. Lower bounds for the eigenvalues of a vibrating string whose density satisfies a Lipschitz condition. Pacific J. Math., 20:393–410, 1967.
- [6] D. C. Barnes and D. O. Banks. Asymptotic bounds for the eigenvalues of vibrating systems. J. Differential Equations, 7:497–508, 1970.
- [7] P. R. Beesack. Isoperimetric inequalities for the nonhomogeneous clamped rod and plate. J. Math. Mech., 8:471–482, 1959.
- [8] H. Brezis. Analyse fonctionnelle. Collection Mathématiques Appliquées pour la Maîtrise. [Collection of Applied Mathematics for the Master’s Degree]. Masson, Paris, 1983. Théorie et applications. [Theory and applications].
- [9] D. Buoso and L. Provenzano. A few shape optimization results for a biharmonic Steklov problem. J. Differential Equations, 259(5):1778–1818, 2015.
- [10] V. I. Burenkov. Sobolev spaces on domains, volume 137 of Teubner-Texte zur Mathematik [Teubner Texts in Mathematics]. B. G. Teubner Verlagsgesellschaft mbH, Stuttgart, 1998.
- [11] S. Chanillo, D. Grieser, M. Imai, K. Kurata, and I. Ohnishi. Symmetry breaking and other phenomena in the optimization of eigenvalues for composite membranes. Comm. Math. Phys., 214(2):315–337, 2000.
- [12] L. M. Chasman. An isoperimetric inequality for fundamental tones of free plates. Comm. Math. Phys., 303(2):421–449, 2011.
- [13] A. Cianchi. Moser-Trudinger inequalities without boundary conditions and isoperimetric problems. Indiana Univ. Math. J., 54(3):669–705, 2005.
- [14] B. Colbois and A. El Soufi. Spectrum of the laplacian with weights. Online preprint ArXiv:1606.04095 [math.DG], 2016.
- [15] B. Colbois, A. El Soufi, and A. Savo. Eigenvalues of the Laplacian on a compact manifold with density. Comm. Anal. Geom., 23(3):639–670, 2015.
- [16] R. Courant and D. Hilbert. Methods of mathematical physics. Vol. I. Interscience Publishers, Inc., New York, N.Y., 1953.
- [17] S. J. Cox. Extremal eigenvalue problems for the Laplacian. In Recent advances in partial differential equations (El Escorial, 1992), volume 30 of RAM Res. Appl. Math., pages 37–53. Masson, Paris, 1994.
- [18] S. J. Cox and J. R. McLaughlin. Extremal eigenvalue problems for composite membranes. I,. Appl. Math. Optim., 22(2):153–167, 1990.
- [19] S. J. Cox and J. R. McLaughlin. Extremal eigenvalue problems for composite membranes, II. Appl. Math. Optim., 22(1):169–187, 1990.
- [20] M. C. Delfour and J.-P. Zolésio. Shapes and geometries, volume 22 of Advances in Design and Control. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, second edition, 2011. Metrics, analysis, differential calculus, and optimization.
- [21] Y. Egorov and V. Kondratiev. On spectral theory of elliptic operators, volume 89 of Operator Theory: Advances and Applications. Birkhäuser Verlag, Basel, 1996.
- [22] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2010.
- [23] J. Fleckinger and M. L. Lapidus. Eigenvalues of elliptic boundary value problems with an indefinite weight function. Trans. Amer. Math. Soc., 295(1):305–324, 1986.
- [24] S. Friedland. Extremal eigenvalue problems defined for certain classes of functions. Arch. Rational Mech. Anal., 67(1):73–81, 1977.
- [25] F. Gazzola, H.-C. Grunau, and G. Sweers. Polyharmonic boundary value problems, volume 1991 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2010. Positivity preserving and nonlinear higher order elliptic equations in bounded domains.
- [26] R. D. Gentry and D. O. Banks. Bounds for functions of eigenvalues of vibrating systems. J. Math. Anal. Appl., 51:100–128, 1975.
- [27] A. Grigor’yan, Y. Netrusov, and S.-T. Yau. Eigenvalues of elliptic operators and geometric applications. In Surveys in differential geometry. Vol. IX, volume 9 of Surv. Differ. Geom., pages 147–217. Int. Press, Somerville, MA, 2004.
- [28] A. Henrot. Extremum problems for eigenvalues of elliptic operators. Frontiers in Mathematics. Birkhäuser Verlag, Basel, 2006.
- [29] N. Korevaar. Upper bounds for eigenvalues of conformal metrics. J. Differential Geom., 37(1):73–93, 1993.
- [30] S. G. Krantz and H. R. Parks. Distance to hypersurfaces. J. Differential Equations, 40(1):116–120, 1981.
- [31] M. G. Krein. On certain problems on the maximum and minimum of characteristic values and on the Lyapunov zones of stability. Amer. Math. Soc. Transl. (2), 1:163–187, 1955.
- [32] P. Lamberti and L. Provenzano. A maximum principle in spectral optimization problems for elliptic operators subject to mass density perturbations. Eurasian Math. J., 4(3):70–83, 2013.
- [33] P. Lamberti and L. Provenzano. Viewing the Steklov eigenvalues of the Laplace operator as critical Neumann eigenvalues. In V. V. Mityushev and M. V. Ruzhansky, editors, Current Trends in Analysis and Its Applications, Trends in Mathematics, pages 171–178. Springer International Publishing, 2015.
- [34] P. D. Lamberti and L. Provenzano. Neumann to Steklov eigenvalues: asymptotic and monotonicity results. Proc. Roy. Soc. Edinburgh Sect. A, 147(2):429–447, 2017.
- [35] B. Schwarz. Some results on the frequencies of nonhomogeneous rods. J. Math. Anal. Appl., 5:169–175, 1962.