Home Global Perturbation of Nonlinear Eigenvalues
Article Open Access

Global Perturbation of Nonlinear Eigenvalues

  • Julián López-Gómez EMAIL logo and Juan Carlos Sampedro ORCID logo
Published/Copyright: April 2, 2021

Abstract

This paper generalizes the classical theory of perturbation of eigenvalues up to cover the most general setting where the operator surface 𝔏:[a,b]×[c,d]Φ0(U,V), (λ,μ)𝔏(λ,μ), depends continuously on the perturbation parameter, μ, and holomorphically, as well as nonlinearly, on the spectral parameter, λ, where Φ0(U,V) stands for the set of Fredholm operators of index zero between U and V. The main result is a substantial extension of a classical finite-dimensional theorem of T. Kato (see [T. Kato, Perturbation Theory for Linear Operators, 2nd ed., Class. Math., Springer, Berlin, 1995, Chapter 2, Section 5]).

MSC 2010: 47A55; 47H11; 35J25

1 Introduction

The problem of the global perturbation of eigenvalues, whose origins go back to Rayleigh in the 19th century, has shown to have a wide range of applications in Mathematical Physics and Nonlinear Partial Differential Equations. Essentially, given a family T(μ) of linear operators on a Banach space U on 𝕂{,} whose dependence on μ is holomorphic, the main classical problem consists in ascertaining whether, or not, the eigenvalues of T(μ) also define holomorphic functions of μ. In other words, should it be true that the values λ(μ) for which

(1.1) T ( μ ) - λ ( μ ) I U GL ( U )

also vary holomorphically with respect to the parameter μ? Through this paper, IU and GL(U) stand for the identity map and the linear group of U, respectively.

In the finite-dimensional context, X=N for some N and (1.1) can be expressed equivalently in the form

(1.2) det ( T ( μ ) - λ ( μ ) I N ) = 0 ,

which is an algebraic equation of degree N in λ with holomorphic coefficients. Thus, by the theory of Algebraic Riemann Surfaces [17, 18], [11, Chapter 1, Section 8], the simple roots of (1.2) are branches of holomorphic functions with respect to μ. Therefore, for any algebraically simple eigenvalue λ0 of T(μ0), there is an holomorphic function, λ(μ), defined on a neighborhood of μ0, Ωμ0, such that λ(μ0)=λ0 and λ(μ) is an eigenvalue of T(μ) for all μΩμ0. Adopting a global perspective, there are, at most, finitely many branching points on each compact subset of where the number of distinct eigenvalues of T(μ) is less than N, i.e., points μ with multiple roots λ(μ) of (1.2). Actually, if we denote by the set of these branching points, the set of eigenvalues λ(μ) can be viewed as a Riemann surface of N-sheets with ramification points . Hence, as soon as μ0, there are N holomorphic local functions λi(μ), i{1,,N}, of eigenvalues of T(μ) such that

σ ( T ( μ ) ) = { λ 1 ( μ ) , λ 2 ( μ ) , , λ N ( μ ) } .

These classical finite-dimensional results are covered, e.g., by [2, 5, 13, 16, 28]. The monograph of Kato [16] being paradigmatic from the point of view of the applications of perturbation theory in Mathematical Physics.

In the more general context when U is a complex Banach space and T(μ) is a compact operator for μ in a neighborhood of μ0, a similar perturbation result holds at simple eigenvalues of T(μ0). Indeed, for any simple eigenvalue λ0 of T(μ0), there exists a holomorphic function, λ(μ), in a neighborhood of μ0, Ωμ0, such that λ(μ0)=λ0 and λ(μ) is a simple eigenvalue of T(μ) for all μΩμ0. Moreover, adopting a global perspective, any finite set of isolated eigenvalues of T(μ) behaves much like in the finite-dimensional context (see, e.g., [15, 29, 31]). Most of these classical results were collected by T. Kato in his influential monograph [16].

Subsequently, we will denote by c(U) the set of compact perturbations of IU. Summarizing the previous discussion, classical perturbation theory deals with operator surfaces 𝔏:Ω¯×c(U) of the special form

𝔏 ( λ , μ ) = T ( μ ) - λ I U , λ Ω ¯ , μ ,

where Ω is an open neighborhood of λ0 in , and T(μ) is assumed to be holomorphic in μ with 𝔏(Ω×)GL(U), i.e., T(μ) cannot admit any eigenvalue on the boundary of Ω, Ω. Under these circumstances, the perturbation problem consists in establishing the existence of perturbed eigenvalues from any given λ0σ(T(μ0)). Most of the technical tools available to deal with this problem are algebraic and analytical, and rely upon the linear structure of the operators involved in the formulation of the perturbation problem as well as, very strongly, on the holomorphy with respect to μ. However, in many applications to Real World problems, one cannot enjoy the holomorphy with respect to μ and the parameters are real, instead of being complex. Although Kato [16] also dealt with the general case when T(μ) depends continuously on the perturbation parameter μ, his analysis only covered the finite-dimensional setting (see [16, Chapter 2, Section 5]). More precisely, according to [16], for any given finite-dimensional operator surface 𝔏:Ω¯×[c,d](N) of the, very special, form

𝔏 ( λ , μ ) = T ( μ ) - λ I U , ( λ , μ ) Ω ¯ × [ c , d ] ,

continuous in μ[c,d], with 𝔏(Ω×[c,d])GL(N), and for every λ0σ(T(μ0)) with μ0(c,d), there exists a continuous function λ(μ) of eigenvalues of T(μ) defined on a neighborhood of μ0 and such that λ(μ0)=λ0. Naturally, as T. Kato worked without any differentiability assumption on μ, his proof relies on topological techniques; in particular, on [16, Chapter 2, Theorem 5.2]. Our main goal in this paper is to deliver a generalized version of Katos’s result valid for a wide class of Fredholm surfaces 𝔏(λ,μ) through the algebraic multiplicity of Esquinas and López-Gómez [10, 9, 19].

To appreciate the strength, versatility and generality of the main theorem of this paper, the reader should note that classical spectral theory focuses attention on the eigenvalues of linear operator curves of the form T-λIU, where the dependence on the spectral parameter λ also is holomorphic, for as it is linear. This is a serious shortage of the existing theory from the point of view of its applications, because many engineering problems involve an intricate nonlinear dependence with respect to the spectral parameter λ and, as a byproduct, classical spectral theory cannot be applied. Actually, T. Kato in [16, Chapter 7, Section 6] tried to generalize the classical theory to cover the simplest prototypes of operator surfaces of the form

(1.3) 𝔏 ( λ , μ ) = T ( μ ) - A ( μ ) λ with  A ( μ ) GL ( U , V ) for every  μ .

However, since the spectrum of these operators differs from the classical spectrum, new problems arose. Following Kato’s words (see [16, pp. 416–417]):

“In applications one often encounters problems of a more general form

(1.4) T u = λ A u

where T and A are operators in U or, more generally, operators from U to another Banach space V. There are several ways of dealing with this general type of eigenvalue problem, etc. But it is not clear what is meant by an isolated eigenvalue of (1.4) or the algebraic multiplicity of such an eigenvalue.”

These open problems were solved some time later by Esquinas and López-Gómez [10, 9, 19], enhancing the birth of nonlinear spectral theory, which will play a central role in this paper. In Section 2 we will collect all the necessary results to read comfortably this paper. In particular, we will shortly remember how the classical spectrum is replaced by the most general generalized spectrum and the classical eigenvalues by the generalized eigenvalues. Problem (1.4) has been widely studied in recent years trying to understand the behavior of such eigenvalues in more general situations where the operator family A(μ) needs not to be invertible (see, e.g., [2, 3, 6, 7, 8, 27]).

As already commented above, this paper sharpens, very substantially, [16, Theorem 5.2 of Chapter 2] to cover the most general case when 𝔏:Ω¯×[c,d]Φ0(U,V) is an operator surface depending continuously on the perturbation parameter μ and analytically on the spectral parameter λ, where Φ0(U,V) stands for the set of Fredholm operators of index zero between U and V. Naturally, this setting covers the case of operators surfaces in Φ0(U,V) of the form

(1.5) 𝔏 ( λ , μ ) = T 0 ( μ ) + T 1 ( μ ) λ + T 2 ( μ ) λ 2 + = n = 0 T n ( μ ) λ n ,

where Tn𝒞([c,d],(U,V)) for each integer n0. The huge gap between the framework of this paper and the classical setting becomes apparent by simply comparing the operators involved in the formulations of (1.3) and (1.5). In particular:

  1. Instead of compact perturbations of the identity map, we are dealing with Fredholm operators in Φ0(U,V), which are the more natural generalizations of the finite-dimensional ones, for as they respect the dimension formula. In particular, compact perturbations of identity operators are Fredholm of index zero by Fredholm’s Theorem, i.e., c(U,V)Φ0(U,V) if UV.

  2. Instead of imposing the analyticity with respect to μ of 𝔏(λ,μ), we only require continuity, which was a real challenge in perturbation theory and it seems an important advance.

  3. We allow the dependence on the spectral parameter λ to be of general analytic type, rather than linear, T-λIU, as in classical perturbation theory, which provides us with a huge flexibility to apply our perturbation theorem in applied sciences and engineering, where the dependence on λ can be nonlinear.

In particular, this type of Fredholm operator surfaces covers nonlinear elliptic eigenvalue problems as those described next. Let Ω be a bounded domain of N of class 𝒞2 whose boundary consists of two disjoint open and closed subsets, Γ0 and Γ1, and consider the next family of elliptic operators

( λ , μ ) := - div ( A ( λ , μ , x ) ) + b ( λ , μ , x ) ,

where the dependence of

A = ( a i j ( λ , μ , ) ) i j N sym ( W 1 , ( Ω ) ) and b ( λ , μ , ) L ( Ω , N )

is analytic in λ[a,b] and continuous in μ[c,d]. For every p[1,+], W2,p(Ω) stands for the Sobolev space of the functions uLp(Ω) such that DαuLp(Ω), in a distributional sense, for every αN with |α|2. Let 𝔅:𝒞(Γ0)𝒞(Γ1)𝒞(Ω) be the boundary operator defined by

𝔅 u = { u on  Γ 0 , ν u + β ( x ) u on  Γ 1 ,

where ν=A𝐧 is the conormal vector field, i.e., 𝐧 is the normal outward vector field of Ω, and β𝒞(Γ1) satisfies either β0 if Γ0=, or β0 if Γ0. Then, setting

W 𝔅 2 , p ( Ω ) := { u W 2 , p ( Ω ) : 𝔅 u = 0 } ,

we can consider, as soon as p>N, the operator surface 𝔏(λ,μ):W𝔅2,p(Ω)Lp(Ω) defined by

(1.6) 𝔏 ( λ , μ ) u := ( λ , μ ) u + c ( λ , μ , x ) u , u W 𝔅 2 , p ( Ω ) ,

where c(λ,μ,)L(Ω) for every (λ,μ)[a,b]×[c,d] and it is analytic in λ and continuous in μ. Since (1.6) satisfies the general requirements of the abstract perturbation results of this paper, we can apply them in this rather general context where the dependence on the spectral parameter λ is nonlinear. No previous perturbation result seems to be available under these general assumptions.

The main theorems of this paper, Theorems 3.1 and 4.1, can be summarized into the next one, where χ and Σ stand for the generalized algebraic multiplicity and the generalized spectrum of Esquinas and López-Gómez [10, 9, 19], respectively.

Theorem 1.1.

Let U and V be two Banach spaces and let L:[a,b]×[c,d]Φ0(U,V) be an operator surface in Φ0(U,V) varying continuously with respect to μ[c,d] and analytically with respect to λ[a,b] and such that

(1.7) 𝔏 ( { a , b } × [ c , d ] ) GL ( U , V ) , χ [ 𝔏 ( , c ) , [ a , b ] ] = λ Σ ( 𝔏 ( , c ) ) [ a , b ] χ [ 𝔏 ( , c ) , λ ] 2 + 1 .

Then there exists a compact and connected subset C of [a,b]×[c,d] satisfying

  1. 𝒞 Σ ( 𝔏 ) ,

  2. 𝒞 links Σ ( 𝔏 ( , c ) ) × { c } to Σ ( 𝔏 ( , d ) ) × { d } .

Suppose, in addition, that χ[L(,μ),λ] is odd for all elements (λ,μ)C. Then there exists a continuous curve γ:[c,d][a,b] such that

  1. ( γ ( μ ) , μ ) 𝒞 for all μ [ c , d ] ,

  2. γ ( c ) Σ ( 𝔏 ( , c ) ) and γ ( d ) Σ ( 𝔏 ( , d ) ) .

Furthermore, one can choose the curve γ to be

(1.8) γ ( μ ) := min { λ [ a , b ] : ( λ , μ ) 𝒞 } for every  μ [ c , d ] .

Any of the two assumptions in (1.7) are imperative for the validity of the theorem. Indeed, if the first requirement of (1.7) fails, then the (generalized) eigenvalues of 𝔏(λ,μ) might be lost through the lateral boundary of the rectangle [a,b]×[c,d], through {a,b}×[c,d]. And Section 3 ends with a simple example establishing the necessity of the oddity of χ[𝔏(,c),[a,b]]. When, in addition, χ[𝔏(,μ),λ] is odd for all (λ,μ)𝒞, then the pathological structures shown in Figure 2 cannot occur. Thus, the curve γ defined by (1.8) is continuous.

The distribution of this paper is the following one. Section 2 collects the main ingredients of the theory of [10, 9, 19], which are necessary to prove Theorem 1.1. In Section 3, we deliver the proof of the first part of Theorem 1.1. In Section 4, we prove the second part of Theorem 1.1. In Section 5 we adapt Theorem 1.1 to a complex setting. Finally, Section 6 delivers an interesting application of Theorem 1.1 to the intricate bi-parametric weighted eigenvalue problem

{ u = α - λ 1 + λ 2 | x | 2 u in  Ω , 𝔅 u = 0 on  Ω .

Our condition to guarantee that γ(μ) is a curve in the setting of Theorem 1.1 seems far from optimal, but it should not be forgotten that we are within the classical topological dichotomy of connected and path connected spaces where sometimes things are more involved that they look. Theorem 1.1 is providing us with the simplest sufficient condition to guarantee the existence of a path joining two extremal points in a component 𝒞.

2 Nonlinear Spectral Theory

In this section we give a brief self-contained introduction to nonlinear spectral theory. Classical spectral theory deals with special curves 𝔏𝒞(Ω,c(U,V)) of the form

𝔏 ( λ ) = λ I U - K , λ Ω ,

where U and V are two 𝕂-Banach spaces such that UV, 𝕂{,}, Ω is an open domain of 𝕂, c(U,V) stands for the space of linear continuous operators that are compact perturbations of the identity map, IU, and K is a given compact operator. Adopting a geometrical perspective, classical spectral theory studies the intersections of the straight line 𝔏(λ) with the space of singular operators

𝒮 ( U , V ) := ( U , V ) \ GL ( U , V ) ,

where GL(U,V) denotes the space of invertible operators. In this context, λ0Ω is said to belong to the spectrum of the straight line 𝔏(λ)=λIU-K if 𝔏(λ0)𝒮(U,V), i.e., if λ0σ(K), the spectrum of K. Note that, in particular, these linear paths lie in the set of Fredholm operators of index zero, denoted in this paper by Φ0(U,V). More generally, nonlinear spectral theory deals with general continuous paths in Φ0(U,V), 𝔏𝒞(Ω,Φ0(U,V)), generalizing the classical theory not only because it deals with arbitrary continuous curves, not necessarily straight lines, but also because these paths can lie in Φ0(U,V) remaining outside c(U,V)Φ0(U,V). In this general context, given a Fredholm path, 𝔏𝒞(Ω,Φ0(U,V)), the generalized spectrum of 𝔏, Σ(𝔏), is defined through

Σ ( 𝔏 ) := { λ Ω : 𝔏 ( λ ) GL ( U , V ) } .

and each λΣ(𝔏) is called a generalized eigenvalue. In the remaining, when we refer to an eigenvalue, we will refer to a generalized eigenvalue. As in the classical case, Σ(𝔏) also consists of the intersection points of the curve 𝔏(λ) with the singular manifold 𝒮(U,V)Φ0(U,V). Naturally, for every compact operator K,

Σ ( λ I U - K ) = σ ( K ) .

A pivotal concept in nonlinear spectral theory is the concept of generalized algebraic multiplicity, which assigns to every pair (𝔏,λ0) with 𝔏𝒞(Ω,Φ0(U,V)) and λ0Σ(𝔏), an integer χ[𝔏,λ0][1,+] such that, whenever dimU=dimV<,

χ [ 𝔏 , λ 0 ] = ord λ 0 det 𝔏 ( λ ) .

The algebraic multiplicity of a generalized eigenvalue χ[𝔏,λ0] can be seen as the intersection index of the curve 𝔏 with the singular surface 𝒮(U,V) at λ0 as the authors have recently shown in [23]. Although a number of constructions of χ[𝔏,λ0] had been done by Göhberg and Sigal [12], Magnus [24], Ize [14], Esquinas and López-Gómez [10], Esquinas [9] and Rabier [26], it was not until 2001 that López-Gómez [19, Chapters 4 and 5] characterized whether all these generalized algebraic multiplicities are finite through the pivotal concept of algebraic eigenvalue. For any 𝔏(λ)𝒞(Ω,Φ0(U,V)), a given λ0ΩΣ(𝔏) is said to be a κ-algebraic eigenvalue if there exist ε>0 and C>0 such that 𝔏(λ) is an isomorphism whenever 0<|λ-λ0|<ε, and

(2.1) 𝔏 - 1 ( λ ) < C | λ - λ 0 | κ if  0 < | λ - λ 0 | < ε ,

with κ the minimal integer for which (2.1) holds. Throughout this paper, the set of κ-algebraic eigenvalues of 𝔏(λ) is denoted by Algκ(𝔏), and the set of algebraic eigenvalues of 𝔏 is defined by

Alg ( 𝔏 ) := κ Alg κ ( 𝔏 ) .

It turns out that the multiplicities of [24, 10, 9] are well defined if, and only if, the path 𝔏(λ) is of class 𝒞r for some r1 and λ0Algκ(𝔏) for some 1κr. Moreover, by [19, Theorems 4.4.1 and 4.4.4], when 𝔏 is analytic and 𝔏(Ω) contains some invertible operator, then Σ(𝔏) is discrete, and every λ0Σ(𝔏) is an algebraic eigenvalue of 𝔏. Thus, through this paper, for any given domain Ωλ0 containing λ0, it is convenient to denote by 𝒜λ0(Ωλ0,Φ0(U,V)) the set of curves 𝔏𝒞r(Ωλ0,Φ0(U,V)) such that λ0Algκ(𝔏)Ωλ0 with 1κr for some r. According to [19, Chapter 4], the generalized algebraic multiplicity χ[𝔏,λ0] is well defined if and only if 𝔏𝒜λ0(Ωλ0,Φ0(U,V)).

Short time later, in 2004, the theory of algebraic multiplicities for 𝒞-Fredholm paths was axiomatized by Mora-Corral [25] by establishing that, modulus a normalization condition, given an open connected neighborhood of λ0 in 𝕂, denoted by Ωλ0, the algebraic multiplicity χ is the unique map

χ [ , λ 0 ] : 𝒞 ( Ω λ 0 , Φ 0 ( U ) ) [ 0 , ]

satisfying the product formula

χ [ 𝔏 𝔐 , λ 0 ] = χ [ 𝔏 , λ 0 ] + χ [ 𝔐 , λ 0 ]

for all 𝔏,𝔐𝒞(Ωλ0,Φ0(U)), regardless whether, or not, λ0 is singular. These findings were collected by López-Gómez and Mora-Corral in [21], where, in addition, the theory of Göhberg and Sigal [12] was substantially generalized to a non-holomorphic framework, and the existence of the local Smith form was established in the class 𝒜λ0(Ωλ0,Φ0(U,V)) through the lengths of the Jordan chains of 𝔏(λ).

A key stone in the development of the results of this paper is the behavior of the algebraic multiplicity under appropriate operator perturbations. Precisely, an homotopy H:Ω¯×[0,1]Φ0(U,V) is said to be a 𝒞-homotopy if H(Ω×[0,1])GL(U,V). According to [22, Theorem 4.8], if 𝔏:[a,b]×[0,1]Φ0(U,V) is a 𝒞-homotopy with analytic μ-sections, then signχ[𝔏(,μ),[a,b]] is constant for all μ[0,1], where

χ [ 𝔏 , [ a , b ] ] λ Σ ( 𝔏 ) ( a , b ) χ [ 𝔏 , λ ]

and

sign χ [ 𝔏 , [ a , b ] ] ( - 1 ) χ [ 𝔏 , [ a , b ] ] = { + 1 if  χ [ 𝔏 , [ a , b ] ] 2 , - 1 if  χ [ 𝔏 , [ a , b ] ] 2 + 1 .

More generally, for every integer n0, we will denote

sign n ( - 1 ) n .

According to [21, Theorem 8.4.3], the next complex counterpart of this result can be easily obtained. If Ω is a subdomain of and 𝔏:Ω¯×[0,1]Φ0(U,V) is a 𝒞-homotopy with holomorphic μ-sections, then χ[𝔏(,μ),Ω] is constant for all μ[0,1], where, rather naturally, χ[𝔏,Ω] stands for

χ [ 𝔏 , Ω ] λ Σ ( 𝔏 ) Ω χ [ 𝔏 , λ ] .

Adopting this point of view, given an operator surface 𝔏:Ω¯×[c,d]Φ0(U,V) continuous in μ[a,b] and analytic in λΩ, the problem that we are addressing in this paper can be viewed as the problem of analyzing the global topological properties of the intersection of 𝔏(Ω¯×[c,d]) with the singular manifold 𝒮(U,V).

3 Perturbation of Eigenvalues

In this section, we give an optimal condition in terms of χ to guarantee the existence of a spectral continuum connecting the spectra of two given operators. Throughout the rest of this paper, the curve 𝔏μ will stand for

𝔏 μ ( λ ) 𝔏 ( λ , μ ) ,

regardless the value of μ. The main result of this section reads as follows.

Theorem 3.1.

Let U and V be two Banach spaces and let L:[a,b]×[c,d]Φ0(U,V) be a C-homotopy with analytic μ-sections, μ[c,d], such that

(3.1) χ [ 𝔏 c , [ a , b ] ] 2 + 1 .

Then there exists a compact and connected subset C of [a,b]×[c,d] such that

  1. 𝒞 Σ ( 𝔏 ) := { ( λ , μ ) [ a , b ] × [ c , d ] : 𝔏 μ ( λ ) GL ( U , V ) } ,

  2. 𝒞 connects Σ ( 𝔏 c ) × { c } and Σ ( 𝔏 d ) × { d } , i.e.,

    𝒞 [ Σ ( 𝔏 c ) × { c } ] 𝑎𝑛𝑑 𝒞 [ Σ ( 𝔏 d ) × { d } ] .

Proof.

By the definition of 𝒞-homotopy, for every μ[c,d],

𝔏 μ ( a ) , 𝔏 μ ( b ) GL ( U , V ) .

Thus, by [19, Theorems 4.4.1 and 4.4.4], Σ(𝔏μ)[a,b] is finite for all μ[c,d]. Thus, for every μ[c,d], the number of connected components, 𝒞, of Σ(𝔏) such that (λ,μ)𝒞 for some λ(a,b) is, at most, finite; possibly empty. Throughout the proof of this theorem, we will denote by 𝒞μ the union of these components if some, otherwise 𝒞μ=.

Note that

Σ ( 𝔏 μ ) [ a , b ] = Σ ( 𝔏 μ ) ( a , b ) ,

because 𝔏μ(a),𝔏μ(b)GL(U,V). According to (3.1), Σ(𝔏c)(a,b) and hence 𝒞c. We will also consider

𝒞 c , d 𝒞 c 𝒞 d ,

which consists of the components of Σ(𝔏) reaching the top, or the bottom, of the rectangle [a,b]×[c,d]. For every μ[c,d], 𝒞μ is closed because it is a finite union of closed sets. Indeed, as 𝒮(U,V)=Φ0(U,V)\GL(U,V) is closed in Φ0(U,V), the set

Σ ( 𝔏 ) = 𝔏 - 1 ( 𝒮 ( U , V ) )

is also closed and hence, each connected component is compact, for as it is also bounded.

The proof of the theorem will follow after a series of technical lemmas. The first one establishes the existence of an open isolating neighborhood, 𝒰𝒞, for every μ[c,d] and every component 𝒞𝒞μ, i.e., an open subset 𝒰𝒞 of (a,b)×[c,d] such that

𝒞 𝒰 𝒞 and 𝒰 𝒞 Σ ( 𝔏 ) = .

The set 𝒰𝒞 is said to be an open isolating η-neighborhood of 𝒞 if, in addition,

𝒞 𝒰 𝒞 𝒞 + B ( 0 , η ) .

Lemma 3.2.

For sufficiently small η>0, μ[c,d] and CCμ, there exists an open isolating η-neighborhood, UC, of the component C.

Proof.

Indeed, fix μ[c,d] such that 𝒞μ and 𝒞𝒞μ, and consider the open η-neighborhood

𝒰 η = 𝒞 + B ( 0 , η )

for sufficiently small η>0 so that 𝒰η(a,b)×[c,d]; η exists because 𝒞 is compact and do not intersect {a,b}×[c,d]. If 𝒰ηΣ(𝔏)=, then we can take 𝒰𝒞=𝒰η and hence, we are done. Suppose that

𝒰 η Σ ( 𝔏 ) ,

and consider the compact set M:=𝒰η¯Σ(𝔏) and the compact disjoint subsets A:=𝒞 and B:=𝒰ηΣ(𝔏). By the maximality of 𝒞, no subcontinuum can connect A with B. Thus, by [19, Lemma 3.2.3], going back to Whyburn [30, Chapter 1], there are two compact disjoint subsets MA and MB of M such that AMA, BMB and M=MAMB. Since MA𝒰η, given a sufficiently small ρ satisfying 0<ρ<d(MA,MB), it becomes apparent that

𝒰 𝒞 = M A + B ( 0 , ρ )

provides us with an open isolating η-neighborhood of 𝒞. Indeed, by construction, we have 𝒞𝒰𝒞𝒰η and 𝒰𝒞Σ(𝔏)=. This concludes the proof. ∎

Naturally, for every μ[c,d] and sufficiently small η>0, the open set

𝒰 μ , η = 𝒞 𝒞 μ 𝒰 𝒞 ,

where the union goes for every component 𝒞𝒰𝒞, provides us with an open isolating η-neighborhood of 𝒞μ, i.e., it is an open subset of (a,b)×[c,d] such that

𝒞 μ 𝒰 μ , η 𝒞 μ + B ( 0 , η ) and 𝒰 μ , η Σ ( 𝔏 ) = .

And the open set 𝒰c,d,η:=𝒰c,η𝒰d,η provides us with an open isolating η-neighborhood of 𝒞c,d, of course.

The next lemma establishes that signχ[𝔏μ,λ] is locally constant at any (λ0,μ0)Σ(𝔏).

Lemma 3.3.

For every (λ0,μ0)Σ(L), there exist ε,δ>0 such that

  1. ( { λ 0 ± ε } × [ μ 0 - δ , μ 0 + δ ] ) Σ ( 𝔏 ) = , and

  2. sign χ [ 𝔏 μ , [ λ 0 - ε , λ 0 + ε ] ] is constant for all μ [ μ 0 - δ , μ 0 + δ ] .

Proof.

Since λ0Σ(𝔏μ0) and Σ(𝔏μ0) is finite, there exists ε>0 such that λ0±εΣ(𝔏μ0). Moreover, as 𝔏 is continuous and GL(U,V) is an open subset of Φ0(U,V), the set

𝔏 - 1 ( GL ( U , V ) ) = ( [ a , b ] × [ c , d ] ) \ Σ ( 𝔏 )

is open; equivalently, Σ(𝔏) is closed. Thus, there exists δ>0 such that

(3.2) ( { λ 0 ± ε } × [ μ 0 - δ , μ 0 + δ ] ) Σ ( 𝔏 ) = .

This ends the proof of part (a).

By (3.2), the restriction of the homotopy 𝔏(λ,μ) to the rectangle

[ λ 0 - ε , λ 0 + ε ] × [ μ 0 - δ , μ 0 + δ ]

is a 𝒞-homotopy with analytic μ-sections. Therefore, owing to [22, Theorem 4.8], part (b) holds. ∎

The next result establishes that the components of Σ(𝔏) which do not reach the top and the bottom of the rectangle [a,b]×[c,d], i.e., those outside 𝒞c,d, do not alter the sign of the global algebraic multiplicity.

Lemma 3.4.

Let Uc,d,η be an open isolating η-neighborhood of Cc,d. Then, for sufficiently small η>0 and every μ[c,d],

(3.3) λ Σ ( 𝔏 μ ) \ 𝒰 c , d , η χ [ 𝔏 μ , λ ] 2 .

Proof.

Pick a μ0[c,d]. Since Σ(𝔏μ0) is finite, also Σ(𝔏μ0)\𝒰c,d,η is finite, possibly empty. If it is empty, then (3.3) holds. So, suppose that, for some integer n1,

Σ ( 𝔏 μ 0 ) \ 𝒰 c , d , η = { ( λ 1 , μ 0 ) , ( λ 2 , μ 0 ) , , ( λ n , μ 0 ) } .

By Lemma 3.3, we can isolate each (λi,μ0) with a small rectangle

λ i ( ε i , δ i ) = [ λ i - ε i , λ i + ε i ] × [ μ 0 - δ i , μ 0 + δ i ] , i { 1 , , n } ,

whose μ-sections do not intersect Σ(𝔏) at λ=λi±εi. By setting

δ 0 := min { δ i : i { 1 , 2 , , n } } ,

it is apparent that, for every δ(0,δ0), the rectangle λi(εi,δ) has also disjoint μ-sections with Σ(𝔏) at λ=λi±εi. Let us consider the union of these rectangles

n , δ := i = 1 n λ i ( ε i , δ )

as well as the entire rectangle

δ := [ a , b ] × [ μ 0 - δ , μ 0 + δ ] n , δ .

Figure 1 sketches the construction carried out in an admissible situation where n=2. We claim that, as illustrated by Figure 1, for sufficiently small δ>0,

(3.4) ( δ \ n , δ ) ( Σ ( 𝔏 ) \ 𝒰 c , d , η ) = .

On the contrary, assume that there exist a sequence of values δ, {δk}n1, with limkδk=0, and a sequence of points, {(λk,μk)}k1, such that, for every k1,

(3.5) ( λ k , μ k ) ( δ k \ n , δ k ) ( Σ ( 𝔏 ) \ 𝒰 c , d , η ) .

Figure 1

The construction of n,δ in the case where n=2.

As {(λk,μk)}k1 is bounded, by the Bolzano–Weierstrass theorem, we can assume, without loss of generality, that

lim k ( λ k , μ k ) = ( λ ω , μ ω )

for some (λω,μω)[a,b]×[c,d]. By (3.5), (λk,μk)δk for all k1. Thus,

μ 0 - δ k μ k μ 0 + δ k .

Hence, letting k, yields μω=μ0. Moreover, since

( λ k , μ k ) n , δ k = i = 1 n λ i ( ε i , δ k ) ,

it becomes apparent that, for every k1,

λ k i = 1 n [ λ i - ε i , λ i + ε i ] .

So, letting k yields to

(3.6) λ ω i = 1 n ( λ i - ε i , λ i + ε i ) .

On the other hand, since Σ(𝔏)\𝒰c,d,η is a closed subset of [a,b]×[c,d] and (λk,μk)Σ(𝔏)\𝒰c,d,η for all k1, letting k, we also find that

( λ ω , μ ω ) = ( λ ω , μ 0 ) Σ ( 𝔏 μ 0 ) \ 𝒰 c , d , η = { ( λ 1 , μ 0 ) , ( λ 2 , μ 0 ) , , ( λ n , μ 0 ) } ,

which entails λω=λi for some i{1,,n} and contradicts (3.6). Therefore, there exists δ0>0 such that (3.4) holds for all δ(0,δ0).

According to (3.4), we have that, for every μ[μ0-δ,μ0+δ],

Σ ( 𝔏 μ ) \ 𝒰 c , d , η n , δ

and hence,

(3.7) χ ( μ ) i = 1 n χ [ 𝔏 μ , [ λ i - ε i , λ i + ε i ] ] = λ Σ ( 𝔏 μ ) \ 𝒰 c , d , η χ [ 𝔏 μ , μ ] .

On the other hand, by Lemma 3.3 (b), δ can be shortened, if necessary, so that

sign χ [ 𝔏 μ , [ λ i - ε i , λ i + ε i ] ]

is constant (either 1, or -1) for all μ[μ0-δ,μ0+δ] and i{1,2,,n}. Thus, also

sign i = 1 n χ [ 𝔏 μ , [ λ i - ε i , λ i + ε i ] ] = i = 0 n sign χ [ 𝔏 μ , [ λ i - ε i , λ i + ε i ] ]

is constant for all μ[μ0-δ,μ0+δ]. Therefore, it follows from (3.7) that

sign λ Σ ( 𝔏 μ ) \ 𝒰 c , d , η χ [ 𝔏 μ , μ ]  is constant for all  μ [ μ 0 - δ , μ 0 + δ ] .

On the other hand, since Σ(𝔏)\𝒰c,d,η is compact and, by hypothesis, does not intersect

( [ a , b ] × { c } ) ( [ a , b ] × { d } ) ,

necessarily

μ α min { μ : Σ ( 𝔏 μ ) \ 𝒰 c , d , η } > c

(see Figure 1). Suppose that

Σ ( 𝔏 μ α ) \ 𝒰 c , d , η = { ( λ 1 , μ α ) , ( λ 2 , μ α ) , , ( λ m , μ α ) } .

By the discussion just carried out, there exists δ>0 such that the function

ζ ( μ ) ( - 1 ) λ Σ ( 𝔏 μ ) \ 𝒰 c , d , η χ [ 𝔏 μ , λ ] , μ [ c , d ] ,

is constant for all μ[μα-δ,μα+δ]. Consequently, since by the definition of μα

Σ ( 𝔏 μ ) \ 𝒰 c , d , η = for all  μ [ μ α - δ , μ α ) ,

it becomes apparent that

(3.8) ζ ( μ ) = 1 for all  μ [ μ α - δ , μ α + δ ] .

As, owing to Lemma 3.3 (b), the sign of ζ(μ) is locally constant in μ[c,d], it becomes apparent that ζ(μ) is continuous. In particular, ζ([c,d]){-1,1} must be connected. Thus, (3.8) implies that ζ([c,d])={1}. Therefore, χ(μ)2 for all μ[c,d]. The proof is completed. ∎

Naturally, thanks to (3.1), almost rephrasing the proof of Lemma 3.4 one can obtain the following counterpart of Lemma 3.4.

Lemma 3.5.

Suppose (3.1), and let Uc,η,Ud,η and Uc,d,η be isolating η-neighborhoods of Cc,Cd and Cc,d, respectively, with sufficiently small η>0. Then, for every μ[c,d],

λ Σ ( 𝔏 μ ) 𝒰 χ [ 𝔏 μ , λ ] 2 + 1

for every U{Uc,η,Ud,η,Uc,d,η}.

Proof.

By (3.1), 𝒞c and by the homotopy invariance of the sign of the multiplicity, χ[𝔏d,[a,b]]2+1, which implies that 𝒞d. Let 𝒰{𝒰c,η,𝒰d,η,𝒰c,d,η}. Since the argumentation is the same in each case, we will suppose without loss of generality that 𝒰=𝒰c,η. First, we will show that

(3.9) Σ ( 𝔏 μ ) 𝒰 for all  μ [ c , d ] .

By (3.1), we have that

Σ ( 𝔏 c ) = Σ ( 𝔏 c ) 𝒰 .

Hence, the set of values of μ[c,d] for which Σ(𝔏μ)𝒰 is nonempty. Let μ0[c,d] be any of these numbers μ. Then, since Σ(𝔏μ0) is finite, we have

Σ ( 𝔏 μ 0 ) 𝒰 = { ( λ 1 , μ 0 ) , ( λ 2 , μ 0 ) , , ( λ n , μ 0 ) }

for some integer n1 and some λi(a,b), i{1,,n}. By Lemma 3.3, each (λi,μ0) can be isolated with a small rectangle

( ε i , δ ) = [ λ i - ε i , λ i + ε i ] × [ μ 0 - δ , μ 0 + δ ] ,

whose μ-sections do not intersect Σ(𝔏) at λ=λi±εi. Naturally, the interval [μ0-δ,μ0+δ] should be inter-exchanged with [c,c+δ] when μ=c, or with [d-δ,d] if μ=d.

Arguing as in the proof of Lemma 3.4, δ can be shortened, if necessary, so that

( δ \ n , δ ) ( Σ ( 𝔏 ) 𝒰 ) = .

Thus, for sufficiently small δ>0,

Σ ( 𝔏 μ ) 𝒰 n , δ if  | μ - μ 0 | δ .

Consequently, owing to [22, Theorem 4.8], we can infer that, for every μ[μ0-δ,μ0+δ], the sign of

(3.10) i = 1 n χ [ 𝔏 μ , [ λ i - ε i , λ i + ε i ] ] = λ Σ ( 𝔏 μ ) 𝒰 χ [ 𝔏 μ , μ ]

is constant. In particular, combining (3.1) with (3.10) yields to

Σ ( 𝔏 μ ) 𝒰 for all  μ [ c , c + δ ] .

Now, let μ*(c,d] be the supremum of the set of values of μ~(c,d] such that

Σ ( 𝔏 μ ) 𝒰 for all  μ [ c , μ ~ ] .

If μ*=d, then (3.9) gets shown. So, suppose that μ*<d. Then, by the previous analysis, there exists δ>0 such that the sign of (3.10) is constant for every μ[c,μ*+δ], which contradicts the definition of μ* and ends the proof of (3.9). We have actually shown that the sign of (3.10) is constant for all μ[c,d]. By (3.1), the proof is complete. ∎

Thanks to the proof of Lemma 3.5, we have that 𝒞c and 𝒞d are nonempty. By construction, to complete the proof of Theorem 3.1, it suffices to show that

𝒞 c 𝒞 d .

On the contrary, assume that

(3.11) 𝒞 c 𝒞 d = .

Then, since 𝒞c and 𝒞d are compact, we have that

μ c max { μ : ( λ , μ ) 𝒞 c  for some  λ ( a , b ) } < d , μ d min { μ : ( λ , μ ) 𝒞 d  for some  λ ( a , b ) } > c ,

and we have three different possibilities according to the relative positions of μc and μd.

Suppose that μc<μd. Then there exists μ(μc,μd) such that

Σ ( 𝔏 μ ) 𝒞 c , d = .

As this contradicts (3.9) for a sufficiently small η-neighborhood, necessarily μcμd.

Suppose that μc=μd, and set

Σ ( 𝔏 μ c ) 𝒞 d = { ( a 1 , μ c ) , ( a 2 , μ c ) , , ( a n , μ c ) } ,
Σ ( 𝔏 μ c ) 𝒞 c = { ( b 1 , μ c ) , ( b 2 , μ c ) , , ( b m , μ c ) } .

By Lemma 3.3, we have that

χ [ 𝔏 μ c , a i ] , χ [ 𝔏 μ c , b i ] 2

for all i{1,,n} and j{1,,m}. Thus,

(3.12) λ Σ ( 𝔏 μ c ) 𝒞 c , d χ [ 𝔏 μ c , λ ] 2 .

As shortening η, if required, one can choose 𝒰c,d,η such that

Σ ( 𝔏 μ c ) 𝒰 c , d , η = Σ ( 𝔏 μ c ) 𝒞 c , d ,

by Lemma 3.5, (3.12) is impossible.

Therefore, μc>μd. By equation (3.11), for sufficiently small η, we can assume that the open isolating η-neighborhoods of 𝒞c and 𝒞d, denoted by 𝒰c,η and 𝒰d,η are disjoint. Thus,

(3.13) λ Σ ( 𝔏 μ c ) 𝒰 c , d , η χ [ 𝔏 μ c , λ ] = λ Σ ( 𝔏 μ c ) 𝒰 d , η χ [ 𝔏 μ c , λ ] + λ Σ ( 𝔏 μ c ) 𝒰 c , η χ [ 𝔏 μ c , λ ] .

On the other hand, by Lemma 3.5,

λ Σ ( 𝔏 μ c ) 𝒰 c , d , η χ [ 𝔏 μ c , λ ] , λ Σ ( 𝔏 μ c ) 𝒰 d , η χ [ 𝔏 μ c , λ ] , λ Σ ( 𝔏 μ c ) 𝒰 c , η χ [ 𝔏 μ c , λ ] 2 + 1 ,

which contradicts (3.13). Therefore, 𝒞d𝒞c, which ends the proof of Theorem 3.1. ∎

The condition χ[𝔏c,[a,b]]2+1 cannot be removed from the statement of Theorem 3.1. Indeed, consider the operator surface

𝔏 : [ a , b ] × [ c , d ] Φ 0 ( H 2 ( 0 , 1 ) H 0 1 ( 0 , 1 ) , L 2 ( 0 , 1 ) )

defined by

𝔏 ( λ , μ ) u = - u ′′ - f ( λ , μ ) u , u U H 2 ( 0 , 1 ) H 0 1 ( 0 , 1 ) ,

where f:[a,b]×[c,d] is a polynomial. Since the eigenvalues of -d2dx2 in (0,1) under Dirichlet boundary conditions are given by (κπ)2, κ1, it is apparent that

Σ ( 𝔏 ) = κ { ( λ , μ ) [ a , b ] × [ c , d ] : f ( λ , μ ) = ( κ π ) 2 } .

Thus, by choosing

[ a , b ] × [ c , d ] = [ - 3 π 2 , 3 π 2 ] × [ - 3 π 2 , 3 π 2 ] and f ( λ , μ ) = μ 2 - λ 2 ,

it is easily realized that

Σ ( 𝔏 ) = { ( λ , μ ) [ - 3 π 2 , 3 π 2 ] × [ - 3 π 2 , 3 π 2 ] : μ 2 - λ 2 = π 2 } .

Observe that Σ(𝔏) consists of the two connected components

Σ ( 𝔏 ) ( [ - 3 π 2 , 3 π 2 ] × [ - 3 π 2 , 0 ) ) , Σ ( 𝔏 ) ( [ - 3 π 2 , 3 π 2 ] × ( 0 , 3 π 2 ] ) ,

which are disjoint because, for sufficiently small μ,

( [ - 3 π 2 , 3 π 2 ] × { μ } ) Σ ( 𝔏 ) = .

On the other hand,

Σ ( 𝔏 3 π 2 ) × { 3 π 2 } = { ( - 5 π 2 , 3 π 2 ) , ( 5 π 2 , 3 π 2 ) } ,
Σ ( 𝔏 - 3 π 2 ) × { - 3 π 2 } = { ( - 5 π 2 , - 3 π 2 ) , ( 5 π 2 , - 3 π 2 ) } ,

and therefore

(3.14) χ [ 𝔏 ± 3 π 2 , [ - 3 π 2 , 3 π 2 ] ] = χ [ 𝔏 ± 3 π 2 , - 3 π 2 ] + χ [ 𝔏 ± 3 π 2 , 3 π 2 ] = 1 + 1 = 2 2 .

Figure 2 shows the two components of Σ(𝔏), 𝒞c and 𝒞d, for this particular example. As a consequence of (3.14), 𝒞c𝒞d=, which cannot happen under condition (3.1) as guaranteed by Theorem 3.1.

Figure 2

An admissible Σ(𝔏) when (3.1) fails.

4 Spectral Continuum Curves

The main result of this section is the next abstract theorem, which provides us with a criterium to guarantee the existence of a continuous curve within a continuum.

Theorem 4.1.

Let C be a compact connected subset of [a,b]×[c,d] satisfying

  1. 𝒞 ( { a , b } × [ c , d ] ) = , and 𝒞 ( [ a , b ] × { g } ) for each g { c , d } ,

  2. 𝒞 μ = 𝒞 ( [ a , b ] × { μ } ) is finite for every μ [ c , d ] ,

  3. for every ( λ 0 , μ 0 ) 𝒞 , there exist ε , δ > 0 such that

    𝒞 ( [ λ 0 - ε , λ 0 + ε ] × { μ } ) for every  μ [ μ 0 - δ , μ 0 + δ ] .

Then the function γ:[c,d][a,b] defined by

(4.1) γ ( μ ) := min { λ [ a , b ] : ( λ , μ ) 𝒞 } for all  μ [ c , d ]

is continuous.

By applying Theorem 4.1 to the special case when 𝒞 is the continuum whose existence was proved in Theorem 3.1, the next complement of Theorem 3.1 can be derived. Note that assumption (c) is a consequence of the local 𝒞-homotopy invariance of the sign of the multiplicity and our assumption on 𝒞 that

(4.2) χ [ 𝔏 μ , λ ] 2 + 1 for all  ( λ , μ ) 𝒞 .

Under this assumption, also in terms of the generalized algebraic multiplicity, there exists a continuous curve, not only a continuum, in Σ(𝔏) that links the spectrums of the operators 𝔏c and 𝔏d. Precisely, the next result holds.

Corollary 4.2.

Let U,V be a pair of Banach spaces and let L:[a,b]×[c,d]Φ0(U,V) be a C-homotopy with analytic μ-sections such that χ[Lc,[a,b]]2N+1. Let C be the continuum (compact and connected subset) of Σ(L) linking Σ(Lc)×{c} with Σ(Ld)×{d}, whose existence is guaranteed by Theorem 3.1. Suppose, in addition, that (4.2) holds. Then, the curve γ:[c,d][a,b] defined by (4.1) satisfies:

  1. ( γ ( μ ) , μ ) 𝒞 for every μ [ c , d ] ,

  2. γ ( c ) Σ ( 𝔏 c ) and γ ( d ) Σ ( 𝔏 d ) .

Our proof of the continuity is based on the next technical result.

Lemma 4.3.

Under the assumptions of Theorem 4.1, pick a μ0(c,d). Then, for every ε,δ>0, there exist μ1(μ0-δ,μ0) and μ2(μ0,μ0+δ) such that

γ ( μ 1 ) , γ ( μ 2 ) ( γ ( μ 0 ) - ε , γ ( μ 0 ) + ε ) .

In particular, there exist an increasing sequence {μ1,n}nN in (μ0-δ,μ0) and a decreasing sequence {μ2,n}nN in (μ0,μ0+δ) such that

lim n μ 1 , n = μ 0 , lim n γ ( μ 1 , n ) = γ ( μ 0 ) ,
lim n μ 2 , n = μ 0 , lim n γ ( μ 2 , n ) = γ ( μ 0 ) .

Proof.

Suppose that it is not true for some μ0(c,d). Then there exist ε0,δ0>0 such that

(4.3) γ ( μ ) ( γ ( μ 0 ) - ε 0 , γ ( μ 0 ) + ε 0 ) for all  μ ( μ 0 , μ 0 + δ 0 ) .

Moreover, δ0 can be shortened, if necessary, so that

(4.4) γ ( μ ) γ ( μ 0 ) + ε 0 for all  μ ( μ 0 , μ 0 + δ 0 ) .

Indeed, by (4.3), should not be this true, for every ρ(0,δ0) there would exist a μ(μ0,μ0+ρ) such that γ(μ)γ(μ0)-ε0. Thus, there exists a decreasing sequence {μn}n1 such that

(4.5) lim n μ n = μ 0 , γ ( μ n ) γ ( μ 0 ) - ε 0 , n 1 .

Moreover, since 𝒞 is compact and, by the definition of γ,

( γ ( μ n ) , μ n ) 𝒞 for all  n 1

along some subsequence, relabeled by n1, we have that

lim n ( γ ( μ n ) , μ n ) = ( α , μ 0 ) 𝒞

for some α[a,b]. Furthermore, letting n in (4.5) yields to

α γ ( μ 0 ) - ε 0 < γ ( μ 0 ) ,

which contradicts the definition of γ. Therefore, (4.4) holds. Consequently, by the definition of γ, we find that

[ ( γ ( μ 0 ) - ε 0 , γ ( μ 0 ) + ε 0 ) × ( μ 0 , μ 0 + δ 0 ) ] 𝒞 = .

As this contradicts assumption (c) of Theorem 4.1, it becomes apparent that, for every ε,δ>0, there exists μ2(μ0,μ0+δ) such that γ(μ2)(γ(μ0)-ε,γ(μ0)+ε).

To prove the existence of μ1(μ0-δ,μ0) such that γ(μ1)(γ(μ0)-ε,γ(μ0)+ε), we also argue by contradiction. Suppose that there exist μ0(c,d) and ε0,δ0>0 such that

γ ( μ ) ( γ ( μ 0 ) - ε 0 , γ ( μ 0 ) + ε 0 ) for all  μ ( μ 0 - δ 0 , μ 0 ) .

Then, arguing as above, we find that, shortening δ0>0 if necessary,

γ ( μ ) γ ( μ 0 ) + ε 0 for all  μ ( μ 0 - δ 0 , μ 0 ) .

So, by the definition of γ, we obtain that

[ ( γ ( μ 0 ) - ε 0 , γ ( μ 0 ) + ε 0 ) × ( μ 0 - δ 0 , μ 0 ) ] 𝒞 = .

As this contradicts assumption (c) of Theorem 4.1, the proof is completed. ∎

Now, we are ready to complete the proof of Theorem 4.1.

Proof of Theorem 4.1.

Suppose that there exists μ0(c,d) such that γ is not continuous at μ0. There are two options:

  1. Either limμμ0γ(μ), or limμμ0γ(μ), does not exist; or

  2. the lateral limits at μ0,

    γ μ 0 - lim μ μ 0 γ ( μ ) and γ μ 0 + lim μ μ 0 γ ( μ ) ,

    do exist, but γμ0-γμ0+.

Suppose that option (i) occurs, and, more precisely, that limμμ0γ(μ) does not exist, i.e.,

(4.6) α lim inf μ μ 0 γ ( μ ) < β lim sup μ μ 0 γ ( μ ) .

Then there are two sequences {μn±}n such that

(4.7) lim n μ n ± = μ 0 , lim n γ ( μ n - ) = α , lim n γ ( μ n + ) = β .

Let us denote by 𝒮μ0 the set of limit points of 𝒞 on [a,b]×{μ0}. Since 𝒞 is closed, it follows that

𝒮 μ 0 = 𝒞 ( [ a , b ] × { μ 0 } ) .

Clearly, α,β𝒮μ0 and 𝒮μ0𝒞μ0. In particular, 𝒮μ0 must be finite. Moreover, by (4.6) and (4.7), we can choose an increasing sequence, {μn}n, such that

(4.8) lim n μ n = μ 0 , γ ( μ 2 n ) < min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } for all  n 1

and

(4.9) lim n γ ( μ 2 n ) = α < lim n γ ( μ 2 n + 1 ) = β .

We claim that, for every integer n1, there exists a subcontinuum of 𝒞, say 𝒟n, such that (μ2n,γ(μ2n))𝒟n,

𝒟 n 𝒬 n [ a , min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } ] × [ μ 2 n - 1 , μ 2 n + 1 ] ,

and

(4.10) [ γ ( μ 2 n ) , min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } ] 𝒫 λ ( 𝒟 n ) ,

where 𝒫λ stands for the λ-projection operator, 𝒫λ(λ,u):=λ (see Figure 3). Indeed, first we consider the portion of the component 𝒞 on the rectangle 𝒬n,

𝒟 := 𝒞 𝒬 n .

Then we take the connected component of 𝒟 containing (μ2n,γ(μ2n)) and denote it by 𝒟n. Let us check that 𝒟n satisfies the required properties. By our general assumptions and the definition of γ, it is easily seen that 𝒟n cannot intersect the set

( { a } × [ μ 2 n - 1 , μ 2 n + 1 ] ) ( [ a , γ ( μ 2 n - 1 ) ) × { μ 2 n - 1 } ) ( [ a , γ ( μ 2 n + 1 ) ) × { μ 2 n + 1 } ) .

Suppose that, in addition, contrarily to the situation sketched on Figure 3, the component 𝒟n does not intersect the segment

{ min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } } × [ μ 2 n - 1 , μ 2 n + 1 ] .

Then 𝒟nint𝒬n. Thus, since 𝒟n is compact, we can isolate it with an open neighborhood within the interior of the rectangle 𝒬n. As this is impossible, because 𝒞 is connected and 𝒟n𝒞, it becomes apparent that there exists μ0(μ2n-1,μ2n+1) such that

( γ ( μ 2 n ) , μ 2 n ) , ( min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } , μ 0 ) 𝒟 n

for every n1. Therefore, since 𝒟n is connected, (4.10) holds. This shows that 𝒟n satisfies the desired requirements. In particular, for every

λ n ( γ ( μ 2 n ) , min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } )

there exists μn0[μ2n-1,μ2n+1] such that

( λ n , μ n 0 ) 𝒟 n 𝒞 .

Finally, pick a ρ(α,β) and ε>0. Then, owing to (4.8) and (4.9), there exists an integer n01 such that

ρ ( γ ( μ 2 n ) , min { γ ( μ 2 n - 1 ) , γ ( μ 2 n + 1 ) } ) , | μ 2 n - 1 - μ 2 n + 1 | < ε 2 , | μ 2 n - 1 - μ 0 | < ε 2

for all nn0. Hence, by (4.10), there exists μn0[μ2n-1,μ2n+1] such that (ρ,μn0)𝒞. By construction,

| ( ρ , μ n 0 ) - ( ρ , μ 0 ) | = | μ n 0 - μ 0 | | μ n 0 - μ 2 n - 1 | + | μ 2 n - 1 - μ 0 | < ε 2 + ε 2 = ε .

Consequently,

lim n ( ρ , μ n 0 ) = ( ρ , μ 0 ) 𝒞

and hence, ρ𝒮μ0. As this holds true for all ρ(α,β), it becomes apparent that [α,β]𝒮μ0𝒞μ0. But this is impossible, because 𝒞μ0 is finite. This contradiction shows the existence of limμμ0γ(μ). Similarly, one can establish the existence of limμμ0γ(μ). Therefore, option (i) cannot occur.

Figure 3

Construction of 𝒟n.

Since limμμ0γ(μ) and limμμ0γ(μ) exist, as we are assuming that γ is not continuous at μ0, necessarily γμ0-γμ0+. Without loss of generality, assume that γμ0-γ(μ0). Then every increasing sequence {μn}n such that

lim n μ n = μ 0

satisfies

lim n γ ( μ n ) = γ μ 0 - γ ( μ 0 ) .

According to Lemma 4.3, this is impossible. Therefore, γ must be continuous on (c,d). This argument can be adapted, very easily, to establish the continuity of γ at μ{c,d}. So, we will omit any further technical detail here. This ends the proof of Theorem 4.1. ∎

The main idea to establish the continuity of γ(μ) in the proof of Theorem 4.1 has consisted in discarding, based on assumptions (a)–(c), the two pathological situations sketched in Figure 4, according to whether options (i) or (ii) are satisfied.

Figure 4 
          Options (i) and (ii) in the proof of Theorem 4.1
Figure 4

Options (i) and (ii) in the proof of Theorem 4.1

5 A Perturbation Theorem in

It is not difficult to rewrite the proof of Theorem 3.1 in order to derive a complex counterpart of it by using the invariance of the algebraic multiplicity under perturbations instead of the invariance of the associated parity of the multiplicity as in the real case. Proceeding in this way, it is straightforward to obtain the following:

Theorem 5.1.

Let U,V be a pair of Banach spaces, Ω an open domain of C, and

𝔏 : Ω ¯ × [ c , d ] Φ 0 ( X , Y )

a C-homotopy with holomorphic μ-sections such that χ[Lc,Ω]2N+1. Then χ[Lμ,Ω] is constant for all μ[c,d], and there exists a compact connected subset C in Ω¯×[c,d], linking Σ(Lc)×{c} to Σ(Ld)×{d}, such that CΣ(L).

Based on Theorem 5.1, it is possible to give another criterium to prove the existence of a continuous curve inside the continuum 𝒞. It has the advantage that, in the practical situations, it is far more computable than the one given by Corollary 4.2.

Theorem 5.2.

Let U,V be two Banach spaces, let

Γ = { z = ( x , y ) : x [ a 1 , b 1 ] , y [ a 2 , b 2 ] } ,

and let L:Γ×[c,d]Φ0(U,V) be a C-homotopy with holomorphic μ-sections such that χ[Lc,Ω]2N+1. Let C be a compact connected subset of Σ(L) linking Σ(Lc)×{c} to Σ(Ld)×{d}, whose existence has been established by Theorem 5.1. If C(ΓR)×[c,d], then there exists a continuous curve γ:[c,d]ΓR such that

  1. ( γ ( μ ) , μ ) 𝒞 for each μ [ c , d ] ,

  2. γ ( c ) Σ ( 𝔏 c ) and γ ( d ) Σ ( 𝔏 d ) .

Moreover, γ can be chosen to be

(5.1) γ ( μ ) = min { λ Γ : ( λ , μ ) 𝒞 } , μ [ c , d ] .

The proof consists in applying Corollary 4.2 to the connected component 𝒞 whose existence has been established by Theorem 5.1. Observe that the assumption (c) of Theorem 4.1 is a consequence of the local 𝒞-homotopy invariance of χ.

Essentially, Theorem 5.2 establishes that, as soon as Σ(𝔏), the existence of the curve γ is guaranteed. Actually, the only pathological case where the curve (5.1) might not be continuous is the case illustrated by the second picture of Figure 4, as the first one is excluded by the finiteness of the μ-sections. The second case cannot happen neither because the whole multiplicity is invariant under perturbations, not only the parity as in the real case. Figure 5 illustrates this situation. Although the branch with multiplicity 2 escapes from the real plane, the multiplicity of the branch remains 2, however in the real plane the branch disappears and hence χ=0 with invariant parity. When Σ(𝔏), this cannot happen and the branches cannot escape of the real plane, which contradicts the second pathological case of Figure 4.

Figure 5

Complex structure of the spectrum.

6 An Intricate Weighted Eigenvalue Problem

In this section, we will apply the abstract theory to a class of linear weighted eigenvalue problems of non-standard type. Those problems arise in applications to engineering and applied sciences where the particular dependence on the parameter λ depends of the nature of the problem. Precisely, we will analyze the eigenvalue problem

(6.1) { u = α - λ 1 + λ 2 | x | 2 u in  Ω , 𝔅 u = 0 on  Ω ,

where α is a positive constant, λ is regarded as a real parameter,

u := - div ( A ( x ) u ) + b ( x ) u

for every uW𝔅2,p(Ω), with p>N, ANsym(W1,(Ω)), bL(Ω), and the boundary operator 𝔅 is defined as in Section 1. Under these assumptions, 𝔏1=0 in Ω, 𝔅10 on Ω, and hence, the constant function h1 provides us with a positive strict supersolution of the tern (𝔏,𝔅,Ω). Thus, thanks to [1, Theorem 2.4], the principal eigenvalue of (𝔏,𝔅,Ω), denoted by σ1(𝔏,𝔅,Ω) through this paper, is positive, and (𝔏,𝔅,Ω) satisfies the strong maximum principle.

Moreover, by the compactness of the resolvent operator -1 (see, e.g., [20, Theorem 4.9]), the spectrum of (,𝔅,Ω), i.e., the set of eigenvalues σ for which (-σ,𝔅,Ω) is not invertible, subsequently denoted by Σ(,𝔅,Ω), consists of a sequence of eigenvalues, {σn}n1, such that

lim n | σ n | = + .

By the generalized Krein–Rutman theorem (see [20, Theorem 6.3]), σ1=σ1[𝔏,𝔅,Ω]>0 is algebraically simple and strictly dominant in the sense that

σ 1 < Re σ n for all  n 2 .

Moreover, it is the unique eigenvalue with a positive eigenfunction (see [20, Chapter 7]).

Subsequently, for any given α>0 such that αΣ(,𝔅,Ω), we consider the operator valued surface

𝔏 : [ 0 , α ] × [ 0 , 1 ] Φ 0 ( W 𝔅 2 , p ( Ω ) , L p ( Ω ) )

defined, for every uW𝔅2,p(Ω), by

𝔏 ( λ , μ ) u := u - α - λ 1 + μ λ 2 | x | 2 u , ( λ , μ ) [ 0 , α ] × [ 0 , 1 ] ,

where μ[0,1] plays the role of a homotopy parameter. Since αΣ(,𝔅,Ω) and σ1(,𝔅,Ω)>0, we have that, for every μ[0,1]:

  1. 𝔏 ( 0 , μ ) = - α GL ( W 𝔅 2 , p ( Ω ) , L p ( Ω ) ) ,

  2. 𝔏 ( α , μ ) = GL ( W 𝔅 2 , p ( Ω ) , L p ( Ω ) ) .

Therefore, 𝔏(λ,μ) is a 𝒞-homotopy with analytic μ-sections. Now, consider the operator curves

𝔏 0 , 𝔏 1 : [ 0 , α ] Φ 0 ( W 𝔅 2 , p ( Ω ) , L p ( Ω ) )

defined by

𝔏 0 ( λ ) := 𝔏 ( λ , 0 ) = + λ - α , 𝔏 1 ( λ ) := 𝔏 ( λ , 1 ) = - α - λ 1 + λ 2 | x | 2 ,

and set

ρ 2 := { min { ( Σ ( , 𝔅 , Ω ) \ { σ 1 } ) } if  ( Σ ( , 𝔅 , Ω ) \ { σ 1 } ) , if  ( Σ ( , 𝔅 , Ω ) \ { σ 1 } ) = .

Then the following result holds.

Theorem 6.1.

The next assertions are true:

  1. If α ( 0 , σ 1 ) , then ( 6.1 ) cannot admit an eigenvalue λ * [ 0 , α ] .

  2. If α ( σ 1 , ρ 2 ) , then the eigenvalue problem ( 6.1 ) admits an eigenvalue λ * [ 0 , α ] . Moreover, λ perturbs continuously from σ 1 , in the sense that there exists a continuum 𝒞 [ 0 , α ] × [ 0 , 1 ] of eigenvalues of the problem

    { u = α - λ 1 + μ λ 2 | x | 2 u in  Ω , 𝔅 u = 0 on  Ω ,

    linking ( λ , μ ) = ( λ , 1 ) to ( λ , μ ) = ( σ 1 , 0 ) .

Proof.

Suppose α(0,σ1) and λ[0,α]. Then, by the monotonicity of the principal eigenvalue with respect to the potential (see [4, Proposition 3.3]), it is apparent that

σ 1 ( - α - λ 1 + λ 2 | x | 2 , 𝔅 , Ω ) σ 1 - sup x Ω α - λ 1 + λ 2 | x | 2 σ 1 - α + λ > λ 0 .

Thus,

σ 1 ( 𝔏 1 ( λ ) ) > 0 for all  λ [ 0 , α ] .

As σ1(𝔏1(λ)) is strictly dominant, this implies that

𝔏 1 ( [ 0 , α ] ) GL ( W 𝔅 2 , p ( Ω ) , L p ( Ω ) )

and ends the proof of part (a).

Suppose σ1<α<ρ2. Then λ*(0):=α-σ1(0,α) provides us with an eigenvalue of 𝔏0(λ) in (0,α). Its next real eigenvalue, if it exists, is α-ρ2<0. Thus, λ*(0) is the unique eigenvalue of 𝔏0(λ) in [0,α]. Moreover, since λ*(0) is algebraically simple,

χ [ 𝔏 0 , [ 0 , α ] ] = χ [ 𝔏 0 , λ * ( 0 ) ] = dim Ker ( - σ 1 ) = 1 .

Hence, by the invariance of the parity of the multiplicity, we have that

sign χ [ 𝔏 1 , [ 0 , α ] ] 2 + 1 .

Therefore, 𝔏1 admits an eigenvalue λ(1)[0,α], which provides us with the desired eigenvalue of (6.1). The existence of the continuum linking λ*(0) to λ*(1) is a direct consequence of Theorem 3.1. ∎

Subsequently, we will focus our attention to the particular case when

( , 𝔅 , Ω ) = ( - Δ , 𝔇 , Ω ) ,

where 𝔇 is the Dirichlet boundary operator, i.e., Γ1=. Then problem (6.1) becomes

(6.2) { - Δ u = α - λ 1 + λ 2 | x | 2 u in  Ω , u = 0 on  Ω .

In this particular case, by the symmetry of -Δ,

Σ ( - Δ , 𝔇 , Ω ) = { σ n : n 1 }

consists of semisimple eigenvalues, and we can refine Theorem 6.1 up to get:

Theorem 6.2.

The following assertions are true:

  1. Suppose α ( 0 , σ 1 ) . Then ( 6.2 ) cannot admit an eigenvalue λ * [ 0 , α ] .

  2. Suppose α ( σ 1 , σ 2 ) . Then problem ( 6.2 ) admits an eigenvalue λ [ 0 , α ] .

  3. More generally, suppose that α ( σ k - 1 , σ k ) for some k 3 such that

    (6.3) i = 1 k - 1 dim Ker ( - Δ + σ i ) 2 + 1 .

    Then problem ( 6.2 ) admits an eigenvalue λ [ 0 , α ] .

Moreover, in cases (b) and (c), the existing eigenvalues λ must perturb from σ1, i.e., there exists a continuum C[0,α]×[0,1] of eigenvalues of the problem

(6.4) { - Δ u = α - λ 1 + μ λ 2 | x | 2 u in  Ω , u = 0 on  Ω ,

linking (λ,μ)=(λ,1) to (λ,μ)=(σ1,0).

When the spatial dimension, N, equals one, then all the eigenvalues σi are simple and hence, (6.3) holds if, and only if, k2.

Proof.

Parts (a) and (b) are direct consequences of Theorem 6.1. Suppose that α(σk-1,σk) for some k3 satisfying (6.3). Then, since

χ [ 𝔏 0 , [ 0 , α ] ] = χ [ + λ - α , [ 0 , α ] ] = i = 1 k - 1 χ [ + λ - α , σ j ] = i = 1 k - 1 dim Ker ( - Δ - σ i ) 2 + 1 ,

by the invariance of the parity of the multiplicity, we can infer that

χ [ 𝔏 1 , [ 0 , α ] ] 2 + 1 ,

which ends the proof of part (c). ∎


Communicated by Fabio Zanolin


Award Identifier / Grant number: PGC2018-097104-B-IOO

Funding source: Eusko Jaurlaritza

Award Identifier / Grant number: PRE2019_1_0220

Funding statement: The first author has been supported by the Research Grant PGC2018-097104-B-IOO of the Spanish Ministry of Science, Technology and Universities and by the Institute of Interdisciplinar Mathematics of Complutense University. The second author has been supported by PhD Grant PRE2019_1_0220 of the Basque Country Government.

References

[1] H. Amann and J. López-Gómez, A priori bounds and multiple solutions for superlinear indefinite elliptic problems, J. Differential Equations 146 (1998), no. 2, 336–374. 10.1006/jdeq.1998.3440Search in Google Scholar

[2] H. Baumgärtel, Analytic Perturbation Theory for Matrices and Operators, Oper. Theory Adv. Appl. 15, Birkhaüser, Basel, 1985. 10.1515/9783112721810Search in Google Scholar

[3] P. Benevieri, A. Calamai, M. Furi and M. P. Pera, On the persistence of the eigenvalues of a perturbed Fredholm operator of index zero under nonsmooth perturbations, Z. Anal. Anwend. 36 (2017), no. 1, 99–128. 10.4171/ZAA/1581Search in Google Scholar

[4] S. Cano-Casanova and J. López-Gómez, Properties of the principal eigenvalues of a general class of non-classical mixed boundary value problems, J. Differential Equations 178 (2002), no. 1, 123–211. 10.1006/jdeq.2000.4003Search in Google Scholar

[5] F. Chatelin, Eigenvalues of Matrices, Classics Appl. Math. 71, Society for Industrial and Applied Mathematics, Philadelphia, 2012. 10.1137/1.9781611972467Search in Google Scholar

[6] R. Chiappinelli, Isolated connected eigenvalues in nonlinear spectral theory, Nonlinear Funct. Anal. Appl. 8 (2003), no. 4, 557–579. Search in Google Scholar

[7] R. Chiappinelli, M. Furi and M. P. Pera, Normalized eigenvectors of a perturbed linear operator via general bifurcation, Glasg. Math. J. 50 (2008), no. 2, 303–318. 10.1017/S0017089508004217Search in Google Scholar

[8] R. Chiappinelli, M. Furi and M. P. Pera, Topological persistence of the unit eigenvectors of a perturbed Fredholm operator of index zero, Z. Anal. Anwend. 33 (2014), no. 3, 347–367. 10.4171/ZAA/1516Search in Google Scholar

[9] J. Esquinas, Optimal multiplicity in local bifurcation theory. II. General case, J. Differential Equations 75 (1988), no. 2, 206–215. 10.1016/0022-0396(88)90136-2Search in Google Scholar

[10] J. Esquinas and J. López-Gómez, Optimal multiplicity in local bifurcation theory. I. Generalized generic eigenvalues, J. Differential Equations 71 (1988), no. 1, 72–92. 10.1016/0022-0396(88)90039-3Search in Google Scholar

[11] O. Forster, Lectures on Riemann Surfaces, Grad. Texts in Math. 81, Springer, New York, 1981. 10.1007/978-1-4612-5961-9Search in Google Scholar

[12] I. C. Gohberg and E. I. Sigal, An operator generalization of the logarithmic residue theorem and Rouché’s theorem, Mat. Sb. (N.S.) 84(126) (1971), 607–629; translation in Math. USSR Sb. 13 (1971), 603–625. Search in Google Scholar

[13] A. Greenbaum, R.-C. Li and M. L. Overton, First-order perturbation theory for eigenvalues and eigenvectors, SIAM Rev. 62 (2020), no. 2, 463–482. 10.1137/19M124784XSearch in Google Scholar

[14] J. Ize, Bifurcation theory for Fredholm operators, Mem. Amer. Math. Soc. 7 (1976), no. 174, 1–128. 10.1090/memo/0174Search in Google Scholar

[15] T. Kato, On the perturbation theory of closed linear operators, J. Math. Soc. Japan 4 (1952), 323–337. 10.2969/jmsj/00430323Search in Google Scholar

[16] T. Kato, Perturbation Theory for Linear Operators, 2nd ed., Class. Math., Springer, Berlin, 1995. 10.1007/978-3-642-66282-9Search in Google Scholar

[17] K. Knopp, Elements of the Theory of Functions. I, Dover, New York, 1945. Search in Google Scholar

[18] K. Knopp, Elements of the Theory of Functions. II, Dover, New York, 1947. Search in Google Scholar

[19] J. López-Gómez, Spectral Theory and Nonlinear Functional Analysis, Chapman & Hall/CRC Res. Notes in Math. 426, Chapman & Hall/CRC, Boca Raton, 2001. 10.1201/9781420035506Search in Google Scholar

[20] J. López-Gómez, Linear Second Order Elliptic Operators, World Scientific, Singapore, 2013. 10.1142/8664Search in Google Scholar

[21] J. López-Gómez and C. Mora-Corral, Algebraic Multiplicity of Eigenvalues of Linear Operators, Oper. Theory Adv. Appl. 177, Birkhäuser, Basel, 2007. Search in Google Scholar

[22] J. López-Gómez and J. C. Sampedro, Algebraic multiplicity and topological degree for Fredholm operators, Nonlinear Anal. 201 (2020), Article ID 112019. 10.1016/j.na.2020.112019Search in Google Scholar

[23] J. López-Gómez and J. C. Sampedro, New analytical and geometrical aspects of the algebraic multiplicity, preprint (2020), https://arxiv.org/abs/2012.05671. 10.1016/j.jmaa.2021.125375Search in Google Scholar

[24] R. J. Magnus, A generalization of multiplicity and the problem of bifurcation, Proc. Lond. Math. Soc. (3) 32 (1976), no. 2, 251–278. 10.1112/plms/s3-32.2.251Search in Google Scholar

[25] C. Mora-Corral, On the uniqueness of the algebraic multiplicity, J. Lond. Math. Soc. (2) 69 (2004), no. 1, 231–242. 10.1112/S002461070300471XSearch in Google Scholar

[26] P. J. Rabier, Generalized Jordan chains and two bifurcation theorems of Krasnoselskii, Nonlinear Anal. 13 (1989), no. 8, 903–934. 10.21236/ADA192353Search in Google Scholar

[27] D. G. Rakhimov, Perturbation of Fredholm eigenvalues of linear operators, Differ. Equ. 53 (2017), no. 5, 607–616. 10.1134/S0012266117050044Search in Google Scholar

[28] F. Rellich, Perturbation Theory of Eigenvalue Problems, Gordon and Breach Science, New York, 1969. Search in Google Scholar

[29] B. Sz-Nagy, Perturbations des transformations linéaires fermées, Acta Sci. Math. (Szeged) 14 (1951), 125–137. Search in Google Scholar

[30] G. T. Whyburn, Topological Analysis, Princeton University, Princeton, 1958. Search in Google Scholar

[31] F. Wolf, Analytic perturbation of operators in Banach spaces, Math. Ann. 124 (1952), 317–333. 10.1007/BF01343573Search in Google Scholar

Received: 2021-02-07
Revised: 2021-03-11
Accepted: 2021-03-11
Published Online: 2021-04-02
Published in Print: 2021-05-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 15.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2021-2127/html?lang=en
Scroll to top button