2.1. Problems of Type G: (Linear) Operator- and Matrix-Valued Functions
A good point to start a presentation of nonlinear eigenvalue problems of the type in Equation (
3) is perhaps R.E.L. Turner’s paper [
6] of 1968. Given a complex Hilbert space
H, rather than considering the spectrum of a single bounded linear operator
A acting in
H, he considers for
operators of the form
where
are given
, with
; thus if
we are back to the familiar
considered in linear spectral theory. The
spectrum of
is defined as the set of those
for which
fails to be a homeomorphism of
H onto itself. In particular, a point
such that
is not injective, i.e., such that the nullspace
, is an
eigenvalue of
. Note that, in the case
, these definitions of spectrum and eigenvalues of
yield what we usually call the spectrum and eigenvalues
of A. The new point of view is that the spectrum is now attributed to the (polynomial)
function of
into
defined putting
In the case
, the spectrum so defined consists very simply of the zeroes of the polynomial
G itself. Now recall (see e.g., [
14] or [
15]) that, if
A is compact, self-adjoint and nonnegative, then:
The spectrum of
A is at most countable and consists of a finite or infinite decreasing sequence of non-negative eigenvalues
:
If the sequence is infinite, then .
The eigenvectors associated with the eigenvalues form an orthonormal basis of H.
Turner first generalizes this to operators as in Equation (
12) where
A is compact, self-adjoint and nonegative,
is self-adjoint and non-negative for
and
A belongs to the Schatten class
(i.e., its eigenvalues
satisfy the condition
) for some
. Another basic fact concerning the spectrum of an operator
A as above is the
variational characterization of its positive eigenvalues
: indeed,
where
V is a vector subspace of
H, and
denotes the family of all vector subspaces of dimension
n.
Turner generalizes the variational principle as follows. For
, let
be the unique non-negative zero of the polynomial
[
6]. Note that, in the case
, as
, we have
so that the function
Z is the usual
Rayleigh quotient of A, of which the eigenvalues are
extremal values as shown by Equation (
14). Then, under the stated assumptions on
A and
, if moreover the eigenvectors of
, corresponding to non-negative eigenvalues, form a basis for
H, then the variational principles in Equations (
14) and (
15) hold replacing
with
.
Finally, we have by definition of
that
Results similar to those of Turner, and practically at the same time, were obtained by K.P. Hadeler in [
16,
17]. He considered several-parameter dependent operators of the form
with
bounded self-adjoint for
, and in connection with the variational property of their eigenvalues introduced the general concept of
Rayleigh functional of a matrix function as follows. Let
be a differentiable mapping of the real interval
to the set
of real symmetric matrices of order
n. Then, a Rayleigh functional of
T is a continuous real-valued function
p on
such that
for all
and
.
The last is a definiteness condition that can be replaced by
, and is plainly satisfied in the basic case
, where
. Thus, looking at Equations (
16) and (
17), we see that this is a sensible extension of the definition and properties of the Rayleigh quotient.
The results of Turner and Hadeler indicated above were developed and improved by, among others, H. Langer. For instance, in [
18], studying combinations
of bounded self-adjoint operators of the form of Equation (
12) considered by Turner, he assumed that, for each nonzero vector
x, the polynomial
has only real roots
Under this assumption he showed that the ranges of the functions are intervals, called spectral zones, whose interiors do not overlap.
A systematization of the spectral theory (that is, of the properties of eigenvalues and eigenvectors) of
polynomial operator pencils, as had been named the families
where
is a spectral parameter, and
are linear operators in a Hilbert space, was given by A.S. Markus in his book [
19]. Among others, he considered in depth the problem of the
factorization of pencils, which in the simplest case consists in representing a quadratic pencil
in the form
The importance of many results in [
19] is due to the fact that they hold for the more general case of
holomorphic(i.e., analytic) operator-valued functions, namely operators
expressed as the sum of convergent power series in
:
For an updated reference reviewing the spectral properties of self-adjoint analytic operator functions, and in particular the factorization problem, see [
10]. On the other hand, for further work on the variational characterization of eigenvalues as well as for the development of the theory of Rayleigh functionals, the interested reader can look for instance into the quite recent papers by Binding, Eschwé and Langer [
20], Hasanov [
21], Voss [
22], and Schwetlick and Schreiber [
23], and the references therein.
Let us now add some more specific indication for the case in which
, so that the function
G appearing in Equation (
3) takes its values in the space
of
real or complex matrices. We shall stress the finite-dimensionality of the ambient space
E using the letter
M rather than
G, and often the letter
v rather than
x for the vectors of
E. A well known reference book for the matter is the one by Gohberg, Lancaster and Rodman [
24], and the very Introduction to this book explains to us that problems of the form
where
appear naturally when dealing with linear systems of higher order ordinary differential equations (ODE) with constant coefficients:
where
for
. Indeed, looking for solutions of the form
of Equation (
21) leads to the equation
which—as long as
, and putting
—is equivalent to Equation (
20) with
Thus,
is a nontrivial solution of Equation (
21) if and only if
is an eigenvalue of Equation (
20), i.e., it is a zero of the
characteristic equation
and
. More generally, the function
is a solution of Equation (
21) if and only if the vectors
satisfy the relations
Such a set of vectors
is called a
Jordan chain of length
for the matrix function
, corresponding to the eigenvalue
and
starting with the eigenvector
. The above definitions extend from matrix polynomials as in Equation (
23) to any analytic matrix function
. It is good to see the explicit form of Equation (
26), which is
If
in Equation (
23), we have
; and if moreover
, then
. In this case,
, while
for all
, so that the above equalities reduce to (putting
)
and are those defining an
ordinary Jordan chain for the matrix
A corresponding to
and
, used to represent
A in its Jordan canonical form and in particular to construct a basis of the
generalized eigenspace associated with
. We recall that this is defined as
where
p is the least integer such that
, and that the dimension
of
is equal to the
algebraic multiplicity of
, that is, the multiplicity of the eigenvalue as a zero of the characteristic polynomial
. We say that
is
semisimple if
in Equation (
29)—that is, if the algebraic multiplicity coincides with the
geometric multiplicity of
, defined as
—and that
is
simple if they are both equal to 1.
These familiar concepts from Linear Algebra, concerning the basic case
, need to be extended to analytic matrix functions
. To this purpose, we quote from [
25]; see also ([
26], Chapter 7).
Let be an eigenvector corresponding to an eigenvalue . The maximal length of a Jordan chain starting at is called the multiplicity of and denoted by . An eigenvalue is said to be normal if it is an isolated eigenvalue and the multiplicity of each corresponding eigenvector is finite.
Suppose that
is a normal eigenvalue. Then, a corresponding
canonical system of Jordan chains
is defined by the following rules:
- (1)
The vectors form a basis of (and so ).
- (2)
is a Jordan chain of the maximal length .
- (3)
Once that the vectors () have been chosen, then pick an eigenvector linearly independent from and form a Jordan chain of the maximal length .
A canonical system is not defined uniquely; however, the numbers do not depend on the choice of Jordan chains and are called partial multiplicities of the eigenvalue . The number is the algebraic multiplicity of the eigenvalue .
The next statement—which is based on results found in [
27]—proves that these definitions are a coherent generalization of the usual ones.
Proposition 1. An eigenvalue is a zero of of multiplicity .
Based on Proposition 1, the definitions of simple and semisimple eigenvalue carry over to the case of matrix polynomials and more generally to analytic matrix functions. For instance, one may check that the matrix function
considered in [
8] has
as a double (i.e., of algebraic multiplicity 2), nonsemisimple (i.e., of geometric multiplicity 1) eigenvalue, with Jordan chain
for any
. This example also shows that in the nonlinear case, generalized eigenvectors do not need to be linearly independent. Indeed, in the construction (and notation) recalled above, the generating vectors
of the system of Jordan chain are chosen to be linearly independent, but it is not necessarily so for the vectors in each corresponding chain, generated by the rules given by the system in Equation (
27).
An especially important source for the study of NLEVP are the Delay Differential Equations (DDE), or systems of them. For instance, in [
26] is considered the so-called Wright equation
where
. The objective is to determine the periodic orbits (if any) of Equation (
32). To do this, one must first look at the linearized equation of Equation (
32) near
, which is
Solutions
of this exist iff
satisfies the characteristic equation
For
, this has
as a simple purely imaginary root, corresponding to the periodic solution
. Studying the properties of these nonlinear eigenvalues, that is of the roots
of Equation (
34) as a function of
, and using deep topological and functional-analytic results from [
26], it is possible to demonstrate that Equation (
32) has a Hopf bifurcation at
, and that for
every Equation (
32) has a nonconstant periodic solution with period close to 4. Finally, the authors show that for
, there is a periodic solution of Equation (
32) of period
p.
One can also consider systems of DDE, for instance
whose
characteristic matrix is precisely that displayed in Equation (
30). The general form of a system of
N delay differential equations, with delays
is
with
, and the corresponding characteristic matrix is
More general forms of Equation (
35) are considered in
Section 3.
2.2. Problems of Type K: Nonlinear Operators and Bifurcation
Throughout this Section
E will be a real Banach space, of finite or infinite dimension. Originally, bifurcation theory deals with the local study of Equation (
1) near a point
, and studies precisely the conditions under which from the given point
of the line
of the
trivial solutions of Equation (
1), there
bifurcates a branch of nontrivial solutions, that is, of solutions
with
. Of course, the basic situation that comes to one’s mind is the case
, with
an eigenvalue of the linear operator
A, the “branch” being here the special subset
of
. The interesting case is when
F depends in a less obvious way from
and
x; an easy example of what we mean is given for instance by the equation
in which the parabola
bifurcates at the point
from the line of the trivial solutions. For a motivating introduction to the theory, and a discussion of some important physical problems that fall in this context, an excellent source is the old review paper by Stackgold [
28].
The previous “naif” idea of bifurcation needs to be made both more precise and more general, and this is done by saying that
is a bifurcation point for Equation (
1) if any neighborhood of
in
contains nontrivial solutions of Equation (
1). For this definition to make sense, it is enough that
F be defined in an open set
with
, and this is what we assume from now on. For the next step, we further assume that
F is differentiable at the point
, so that
F can be linearized near that point as
where the remainder term
R satisfies
Some more regularity on F yields immediately a necessary condition for bifurcation:
Theorem 1. Suppose that F is of class in a neighborhood of . If is a homeomorphism of E onto itself, then cannot be a bifurcation point for Equation (
1).
Proof. The assumption implies, via the Implicit Function Theorem, that there is a neighborhood
of
such that, for any
, there is a
unique such that
. As by assumption
for any
, we must have
for
, so that there is no nontrivial solution to Equation (
1) in the neighborhood
of
. ☐
For simplicity, we shall henceforth consider only the special case
where
and
A is of class
near
. Here,
, and we have a more explicit form of the remainder term in the linearized form in Equation (
36) of
F: for we can write
with
as
, so that Equation (
37) yields
and comparing this with Equation (
36) we see that
in this special case. Resuming, the equation we want to study is
with
and
A of class
near
, and can be written as
where
and
as
. The necessary condition for bifurcation implicitly stated in Theorem 1 can now be rephrased as follows:
The standard case considered in the literature is when is in the point spectrum of , and we formalize this more precisely under the form of a basic assumption, which is plainly satisfied if :
H0. is an isolated eigenvalue of and is a Fredholm operator of index zero.
Let us recall (see, e.g., [
14]) that a bounded linear operator
L between two Banach spaces
E and
F is said to be a
Fredholm operator if its nullspace
has finite dimension and its range
is closed and has finite codimension; in this case, the
index of
L,
, is defined as
Thus, if
, then any linear operator is Fredholm of index zero. From the Riesz–Schauder theory of such operators (see e.g., [
15]), it is known that also the nullspaces
(
) are finitedimensional, and that they stabilize for
j sufficiently large; with reference to the case
, this means that there exists a least integer
p such that
, and moreover one has
It follows in particular that the
algebraic multiplicity of
is finite, where in general this is is defined—consistently with the definition recalled in
Section 2.1 for the case
—as the dimension of the subspace
In the following, when speaking of multiplicity of an eigenvalue, we refer to the algebraic multiplicity. We recall that this coincides with the geometric multiplicity when T is a self-adjoint operator in a Hilbert space.
Remark 1. The assumption H0 alone is not sufficient to guarantee that an eigenvalue of the “linear part” T of A at 0 is a bifurcation point for A. To see this, consider the example ([29], Chapter 11) given by the systemHere, and we have (in our notations) , and . Multiplying the first equation by y, the second by x and subtracting the second from the first, we obtain , showing that Equation (
40)
has no nontrivial solution whatsoever. One way of seeing this is that the two-dimensional eigenspace associated with is completely destroyed by the addition of the perturbing term B. Three typical situations are then considered, each of them guaranteeing bifurcation from , and described by the following assumptions, respectively:
H1. is a simple eigenvalue of .
H2. A is compact and is an eigenvalue of odd multiplicity of .
H3. A is a gradient operator in a Hilbert space and is an isolated eigenvalue of finite multiplicity of .
These assumptions call immediately for some explanation. In fact, it could be noted at once that both H2 and H3 are a strengthening of H0. However, to proceed with some order, in the remaining part of this subsection, we shall give a precise statement for each of the three bifurcation results roughly indicated above, preceded by a comment on the respective assumption, and followed by an indication of the proof.
Thus, starting with
H2, we recall that, if
A is compact, then the linear operator
is a compact, too [
30]. Therefore
H0 is redundant in this case, as it is a basic spectral property of any such operator [
14].
Theorem 2. If H2 is satisfied, then is a bifurcation point for Equation (
38).
Moreover, it is a global
bifurcation point in the following sense: if S denotes the closure in of the set of nontrivial solutions of Equation (
38),
then has a connected
subset containing , and which is either unbounded
in or contains a point with an eigenvalue of odd multiplicity of T. Proof. The proof relies on the
Leray–Schauder degree. Roughly speaking, this is a topological tool to detect the fixed points of a compact map and can be briefly introduced as follows (see, for instance, [
3] (Part I) for a complete presentation). Suppose we have a continuous compact map
C of
E into itself, a bounded open set
with
, and suppose that
for
. Then, there exists an integer, denoted
and called the (Leray–Schauder)
degree of relative to the set Ω
and to the point 0, having the following properties:
- (i)
If , then there exists an such that .
- (ii)
.
- (iii)
If and has no zeroes in , then
- (iv)
Suppose
are compact maps, and put
If
for
and
, then
- (v)
If
C is a
linear compact map and
is injective, then
where
is the number of eigenvalues >1 of
C, each counted with its algebraic multiplicity.
To prove that
is a bifurcation point, it is enough to show that for any sufficiently small
, there exists a solution
of Equation (
38) with
and
. Thus, let
be the open ball centered at
and with radius
r; we consider the degree of various maps with respect to this neighborhood of 0. Precisely, assume for instance
and write
,
, for
near
. Consider thus the equivalent equation
and let
vary in an interval
containing as interior point
and no other
characteristic values(as are named the reciprocals of the nonzero eigenvalues) of
except
. Assume by way of contradiction that
for
and
; then using the
Homotopy invariance Property (iv) with
we would have
On the other hand, for small
, using again Property (iv) we have
because
is homotopic to
on
; indeed, since the latter operator is a homeomorphism and since
as
, we have (diminishing
r if necessary)
for some
and for all
. Similarly,
However, using Property (v), we have
where
,
h an odd integer (the algebraic multiplicity of
); therefore the two degrees in Equation (
45) are different, contradicting the previous equalities in Equations (
42)–(
44). This proves that
for some
and some
, and therefore that there is bifurcation from
for the Equation (
41), or equivalently from
for Equation (
38). The proof that under the stated assumptions the bifurcation has a
global character, in the sense described by the statement of Theorem 2, requires the much deeper topological analysis performed by P.H. Rabinowitz in his famous paper [
31]. ☐
We now go on to comment assumption
H3, and to briefly discuss the corresponding bifurcation result. For the next definition, and the statements following it, see for instance [
2] or [
30].
Definition 1. Let H be a real Hilbert space with scalar product denoted . An operator is said to be a gradient (
or potential) operator
if there exists a differentiable functional such that One then writes
; the functional
a—the
potential of
A—is uniquely determined by the requirement that
, and is explicitly given by the formula
A bounded linear operator is a gradient if and only if it is self-adjoint. Moreover, if a gradient operator A is differentiable at a point , then is self-adjoint.
Theorem 3. If H3 is satisfied, then is a bifurcation point of Equation (
38).
Moreover, for each sufficiently small, Equation (
38)
has at least two
distinct solutions such that . Proof. The proof makes use of the
Lyapunov–Schmidt method (see, for instance, ([
3], Chapter 2) or ([
29]), Chapter 11) which allows to reduce the infinite-dimensional problem in Equation (
38) to a problem in the finite-dimensional space
. Indeed, consider the equivalent form Equation (
39) of Equation (
38), and rewrite it as
where
and
. Now recalling that
, the assumption
H3 implies that
H is the orthogonal sum
Then, letting
denote the orthogonal projections of
H onto
and
, respectively, we have
and using this in Equation (
48), we obtain the equivalent system
The restriction
of
L to
is a homeomorphism of
onto itself. A standard application of the Implicit Function Theorem, together with the condition
as
, then allows to solve the
complementary equation, Equation (
52) in the form
with
, where
and
v belong to suitably small neighborhoods
J and
V of
and
, respectively in
and in
. Replacing this in Equation (
51) first yields
whence, applying once more the Implicit Function Theorem, one can recover
as a function of
v,
for
v in a neighborhood
of 0 in
. Finally, putting
and replacing this in Equation (
51), one is left with the finite-dimensional equation (the
bifurcation equation)
Any solution
,
, of this equation will give rise to a solution
,
,
of the original Equation (
48), and the continuity (in fact,
regularity) of the maps
will ensure that this solution
stays into a given small neighborhood of
in
provided that
v is small enough. Thus, proving bifurcation from
for Equation (
38)—or equivalently, bifurcation from
for Equation (
48)—reduces to prove that Equation (
57) has solutions
of arbitrarily small norm.
Remark 2. The Lyapunov–Schmidt reduction can be applied more generally, and with minor modifications, in a Banach space E whenever the basic assumption H0 (i.e., that is Freholm of index zero) holds and is supplemented by the transversality conditionwhich is plainly satisfied when T is self-adjoint, as the two subspaces in Equation (
58)
are then orthogonal. Note that Equation (
58)
is in general equivalent to , and thus to the fact that the algebraic and geometric multiplicities of coincide. H0 and Equation (
58)
imply the direct decomposition of E into (closed) subspaces as in Equation (
49),
and therefore allow for the same reduction on taking for the (continuous) projections associated with Equation (
49).
Returning to the proof of Theorem 3, we let now come in the assumption that the whole
A, and therefore also its “nonlinear part”
B, is a gradient. Here, we bound ourselves to give the main idea of the particularly clear demonstration provided by C. Stuart [
32]. Thus, let
f be such that
, and consider the reduced functional
defined putting
Moreover, for small
put
Then,
is the level set of the
functional
g, and is compact because it is a closed and bounded subset of the finite dimensional space
. Thus,
attains its minimum and its maximum on
, and if
is such an extremal point we have, by the Lagrange multiplier’s rule,
Performing the computations of
and
by the definitions in Equations (
59) and (
60), and using the fact that
satisfies the complementary equation, Equation (
52), one checks that
and that Equation (
57) is satisfied. ☐
We finally come to H1. Unlike H2 and H3, in general H1 is independent from H0, and must be supplemented with it to guarantee bifurcation. Of course, when E is finite dimensional, H0 does not play any role, and indeed H1 can in this case be viewed as a special case of H2, because any continuous map is then compact.
Theorem 4. If H0 and H1 are satisfied, then is a bifurcation point of Equation (
38).
Moreover, if A is of class in a neighborhood of , then near the solution set of Equation (
38)
consists of the trivial solutions and of a
curve
with and for . Finally, if , then as The statement of Theorem 4 means that near
, the solution set of Equation (
38) is topologically equivalent to the “cross”
As to the proof, this goes for a first part along the same lines used to prove the previous Theorem 3, that is, using the Lyapounov–Schmidt decomposition in the sense indicated in Remark 2. What is specific here is that, since
, one ends with an equation in
; a further nontrivial application of the Implicit Function Theorem then leads to the result: see, for instance, ([
3], Chapter 2).
2.3. A Very Special Nonlinear Problem: The p-Laplace Equation
Let
be a bounded open set in
, let
, and let
E be the Sobolev space
, equipped with the norm
That this is actually a norm in
, equivalent to the standard one of
, is a consequence of
Poincaré’s inequality(see e.g., [
14]), stating that
for some
and for all
. Let
be the dual space of
E. A (weak) solution of the
p-Laplace Equation (
5) is a function
such that
where
(it will soon be clear that
is not an eigenvalue of Equation (
5)) and
are defined by duality via the equations
where
and
denotes the duality pairing between
E and
.
The proof of the existence of countably many eigenvalues and eigenfunctions of Equation (
65) relies on the Lusternik–Schnirelmann (LS) theory of critical points for an even functional on a symmetric manifold. Complete presentations of this theory, in both finite and infinite dimensional spaces, can be found, among others, in [
3,
4,
29,
33,
34]. Theorem 5 below is essentially a simplified version of Theorem A in [
35], save that with respect to [
35] we have for expository convenience interchanged the roles of the operators
A and
B. Thus, let
E be a real, infinite dimensional, uniformly convex Banach space with dual
, and consider the problem
where
are continuous gradient operators with potentials
, respectively:
. Definition 1 of gradient operator extends of course to mappings of
E into
replacing the scalar product with the duality pairing.
Suppose that
for
; then, the eigenvectors of Equation (
67) satisfying a normalization condition
(
), are precisely the constrained critical points of
a on the level set
The additional key assumptions that we make on
A and
B are as follows:
are odd(that is, for , and similarly for B).
A is non-negative (that is, for ) and strongly sequentially continuous (that is, if converges weakly to , then converges strongly to in ).
B is
strongly monotone in the following sense: there exist constants
and
such that, for all
,
By the above assumptions on
B,
is
symmetric (that is,
) and
sphere-like, in the sense that each ray through the origin hits
in exactly one point. If
is compact and symmetric, then the
genus of
K, denoted
, is defined as
If
V is a subspace of
E with
, then
. For
put
Theorem 5. Let be as above. Suppose moreover that implies . For and , putwhere is as in Equation (69). Then Moreover, as , and if then is attained and is critical value of a on : thus, there exist and such thatand Here are a few indications for the Proof of Theorem 5:
(i) The sequence
is non-decreasing because, for any
, we have
as shown by Equation (
69). (ii) In addition,
because
contains all sets of the form
,
. (iii) The proof that
as
, together with a lot of related information, can be found for instance in [
34]. (iv) Finally, the assumption that
whenever
, together with the stated continuity properties of
A and
B, ensures that
a satisfies the crucial
Palais–Smale(PS)
condition on
at any level
, needed to prove the final (and most important) assertion of the Theorem via the standard deformation methods of Critical Point Theory; see for this any of the above cited references.
Of special importance—with reference to the the
p-Laplace equation—is the case in which
A and
B have the additional property of being
positively homogeneous of the same degree
, meaning that
for
and
, and similarly for
B. In this case, we have from Equation (
47)
so that
implies
. Moreover, the use of Equation (
74) in Equations (
72) and (
73) yields at once the relation
. In fact, here,
is
independent of : to see this, it is convenient to re-parameterize the level sets on putting for
As
a and
b are
p-homogeneous, it follows that
, that each
is the image of the corresponding set in
under the map
, and that
. By these remarks, we thus have the equalities
showing as expected that
is independent of
R, and precisely that
where
. From Theorem 5, we then get immediately the following statements about
:
;
as ; and
if , then there exists (that is, ) such that ; in particular, .
Remark 3. The situation just described contains as a more special case that of two linear operators A and B, in which the above formulae hold with . Suppose in particular that A acts in a real Hilbert space H and ; then is the unit sphere in H, while A is a compact, self-adjoint, non-negative linear operator (strong sequential continuity and compactness are equivalent properties for a linear operator acting in a reflexive Banach space, see e.g., [15]). Then, Equation (
75)
and the statements following this formula yield a good part of the familiar spectral properties of such operators: indeed it is not hard to see that the LS variational characterization in Equation (
75)
of reduces in this case to the classical Courant’s minimax principle expressed by Equation (
15),
so that the sequence in Equation (
75)
of the LS eigenvalues of A coincides with the decreasing sequence of all the eigenvalues of A, each counted with its multiplicity. Returning finally to the
p-Laplacian, it is now a matter of applying the above information to the operators
defined in Equation (
66). One can check (see [
36,
37], for instance) that they satisfy all the requirements for the application of Theorem 5. Moreover, they are evidently positively homogeneous of degree
, and finally
is (strictly)
positive, for
This implies that each of the numbers
defined in Equation (
75) for the pair
is strictly positive, whence it follows—using the last statement of Theorem 5—that the eigenvalue problem in Equation (
65) for the
p-Laplacian possesses an infinite sequence of eigenvalues
, each given by
where
Setting
, this finally proves the properties of Equation (
5) stated in the Introduction, and in particular Equation (
6).
Remark 4. For the very special properties owned by the first eigenvalue in the sequence in Equation (
6)
and by the associated eigenfunctions, see for instance [37]. Anyway, it follows by our discussion that is the best constant in Poincaré’s inequality, Equation (
64):
To conclude this section, let us remark that the study and research in problems related to the
p-Laplacian has grown enormously in the last decades, and even remaining in the strict context of a “spectral theory” for Equation (
5), one should at least mention the following relevant points: (i) the problem of the
asymptotic distribution of the LS eigenvalues (along the classical Weyl’s law for the Laplacian); (ii) the question of the existence of
other eigenvalues outside the LS sequence; and (iii) the
Fredholm alternative for perturbed non-homogeneous versions of Equation (
5). For information on these issues, we refer the reader to [
37,
38,
39] and to the recent and very clear review paper [
36]. Related material can be found in [
40].