1. Introduction
Differential equations are essential mathematical tools employed to model and analyze a diverse array of phenomena across various scientific and engineering disciplines. Consequently, investigating the existence of solutions to differential equations has become a significant area of study. To address this, numerous mathematical theories have been applied, including phase space theories [
1,
2,
3,
4,
5], smooth theory [
6,
7,
8], operator methods [
9,
10,
11,
12], and critical point theory [
13,
14].
Let
H be a real Hilbert space, where we consider the self-adjoint operator equation:
where
A is a self-adjoint operator with domain
, and
F is a potential operator such that
,
and
.
Many problems can indeed be represented by the operator Equation (
1), such as Laplace’s equation on bounded domains with a Dirichlet boundary, periodic solutions of Hamiltonian systems [
15,
16], the Schrödinger equation [
17,
18,
19], periodic solutions of the wave equation [
14], resonant elliptic systems [
20,
21,
22], and others. For these problems, the calculus of variations indicates that the solutions of Equation (
1) correspond to the critical points of a functional on a Hilbert space ([
23], Chapter 1). Thus, finding solutions to Equation (
1) translates into finding the critical points of a functional. Infinite-dimensional Morse theory is articularly useful for obtaining critical points for the functional [
13,
24].
However, if the Palais—Smale (PS) condition does not hold, these methods become significantly more challenging [
21,
22]. To address this issue, saddle point reduction is a viable approach. The theory of saddle point reduction (also known as Lyapunov-Schmidt reduction in some literature) was established by Amann in 1979 [
25]. Since then, it has become an important tool in critical point theory and has been widely used to solve various boundary value problems [
13,
26,
27]. Therefore, the first step of our study is to perform a saddle point reduction.
According to operator spectral theory, the Hilbert space
H can be decomposed as follows:
where
for
,
for
, and
. Let
denote the spectrum of
A,
denote the eigenvalues with finite multiplicity, and
denote the essential spectrum of
A. We introduce two important conditions:
There is a constant such that and contain at most finitely many eigenvalues.
The operator
F is Gâteaux-differentiable in
H and satisfies
We define the following functional:
It is clear that
x is a critical point of
if and only if
x is a solution of Equation (
1). Let
x be a critical point of
g. We denote the
Morse index of
x by
, which is the dimension of the largest negative space of
. Similarly, we denote the positive Morse index of
x by
, which is the dimension of the largest positive space of
.
Let
be a projection operator, and we define an operator on
as follows:
Then
is a self-adjoint invertible Fredholm operator.
Without loss of generality, we fix a constant
C such that
Let
be the spectral family associated with
. Here are three projections on the space H defined as follows:
Then, we can decompose the space
H as follows.
where
, and
.
Using ([
13], Chapter IV, Theorem 2.1), we reduce the Equation (
1) to the finite dimensional case.
Proposition 1. Given assumptions and , there exist a functional and an injective map such that the following conditions are satisfied:
- 1.
The map u can be expressed as , where .
- 2.
The functional a adheres to the following: - 3.
z is a critical point of a if and only if is a critical point of g, which is equivalent to being a solution to the operator Equation (1).
Let
represent the
qth singular relative homology group of the topological pair
with coefficients in a field
. Consider
x as an isolated critical point of the function
g with
. The group defined by
which is known as the
qth critical group of
g at the point
x, where
.
Let us consider the following condition:
The spectrum of A consists solely of eigenvalues, meaning it is a point spectrum, and .
It is straightforward to deduce that condition
follows from
. Therefore, under conditions
and
, Proposition 1 remains valid. Additionally, with condition
in place, we can define the Morse index and critical groups for the functional Equation (
3).
This leads to a natural question: what is the relationship between the Morse index and the critical groups before and after applying the saddle point reduction? Our main result addresses this question as follows.
Theorem 1. Consider a real Hilbert space H, and let be the functional defined as in Equation (3). If the conditions and are satisfied, and is an isolated critical point of the reduced functional a, then the following results hold: - 1.
The Morse index of , denoted by , is equal to the Morse index of , denoted by , i.e., , where .
- 2.
The critical group at is related to the critical group of by for all .
Our exploration of the connection between the Morse index and the critical groups before and after saddle point reduction is driven by the study of multiple solutions in second-order Hamiltonian systems. In the third section of this paper, we apply our abstract results to asymptotically linear second-order Hamiltonian systems. These types of problems have garnered significant attention in recent years, as noted in [
28,
29,
30,
31].
Specifically, we focus on the following boundary value problem for second-order Hamiltonian systems:
where
satisfies
for some
, and it adheres to the linear growth condition:
where
is a constant, and
. We can assume without loss of generality that
. Additionally, we assume
, so
is a trivial solution of Equation (
7). Our goal is to find nontrivial
-periodic solutions to the system Equation (
7).
Consider
, a continuous symmetric matrix function that is
-periodic. We focus on the following eigenvalue problem:
subject to
-periodic boundary conditions. It is well established that there exists a complete sequence of distinct eigenvalues
such that
as
.
Next, we outline several key assumptions regarding the nonlinearity :
There exists a
-periodic continuous symmetric positive definite matrix function
with eigenvalue
for some
such that
where
as
.
There exists such that for all , and .
There exists a
-periodic continuous symmetric matrix function
with eigenvalue
for some
such that
where
as
.
There exists a constant
such that
for any
, and
There exists a constant
such that
and
for all
.
Define
With this notation, our second main result can be stated as follows:
Theorem 2. Suppose satisfies for some and the condition Equation (8). If assumptions , and are satisfied and if , then the system of Equation (7) has at least two nontrivial solutions in the following cases: - (i)
holds with ;
- (ii)
holds with .
2. Preliminaries
In this section, we provide the proof of Theorem 1. As a preliminary, we introduce several key definitions and lemmas, all of which are referenced from [
13].
First, we state a technical assumption—the Palais–Smale condition, which is frequently encountered in critical point theory. Consider H as a separable Hilbert space.
Definition 1. Let g be a functional defined on H. The functional g is said to be Fréchet-differentiable at if there exists a continuous linear map such that for any , there exists such thatfor all . The mapping L is typically denoted by . A critical point
u of
g is a point where
, that is,
for all
. The value of
g at
u is then referred to as a critical value of
g.
Let represent the set of functionals that are Fréchet-differentiable and whose Fréchet derivatives are continuous on H.
Definition 2. We say that a functional satisfies the Palais–Smale condition (denoted as (PS)) if any sequence for which is bounded, and as has a convergent subsequence.
Remark 1. The (PS) condition is a useful way to introduce some “compactness” into the functional g. Specifically, observe that (PS) implies thatmeaning that the set of critical points with critical value c is compact for any . The first lemma describes the critical group in terms of the Morse index.
Lemma 1 ([
13])
. Consider , and let u be a nondegenerate critical point of g with Morse index j. Then, . For a critical point, which may be degenerate, the following significant result is established:
Proposition 2 ([
32], Corollary 8.4)
. Let and be an isolated critical point with finite Morse index μ and nullity ν. If is a Fredholm operator, then for . Moreover, we have the following:- 1.
If , then .
- 2.
If , then .
The following lemma is known as the splitting theorem:
Lemma 2 ([
13,
33])
. Let U be a neighborhood of θ in a Hilbert space H, and let . Suppose θ is the only critical point of f, and denote by the Hessian at θ, with kernel N. If 0 is either an isolated point in the spectrum or not in , then there exist a ball () centered at θ, an origin-preserving local homeomorphism ϕ defined on , and a mapping such thatwhere , , and define the orthogonal projection onto the subspace N. Definition 3 ([
13])
. Consider a Banach space X and a connected Hausdorff space M. We define M as a Banach manifold, for (integer) and modeled on X, if the following conditions hold:- 1.
There exists a family of open coverings ;
- 2.
There exists a family of coordinate charts , where is a homeomorphism for each ;
- 3.
The transition maps are diffeomorphisms for all .
Each pair is called a chart, and the collection is referred to as an atlas.
Similarly, we can define (or ) maps between two Banach manifolds, as well as vector bundles over Banach manifolds. In particular, this includes the tangent bundle and the cotangent bundle .
Given a vector bundle , a section is a map such that . A section is said to be - (or ) -continuous if it is a (or ) map from M to E.
A Riemannian manifold
is metrizable, with the metric
d defined by the arc length of geodesics, which is in turn determined by the Riemannian metric
g:
As a metric space , the topology coincides with (or is equivalent to) the topology of the manifold.
Since the Riemannian metric is globally defined on , we shall introduce a Finsler structure on a Banach manifold in a similar manner.
Definition 4 ([
13])
. Let be a Banach vector bundle. A Finsler structure is a function that satisfies the following conditions:- 1.
is continuous;
- 2.
For every , the restriction is an equivalent norm on the fiber ;
- 3.
For any point and any neighborhood U of that trivializes the vector bundle E (i.e., ), there exists a neighborhood V of M with such that for all ,
Below is the definition of a characteristic submanifold and a theorem known as the Shifting Theorem:
Definition 5 ([
13])
. Let M be a Finsler manifold, and let be a functional. Consider a local parametrization Φ of M defined in an open neighborhood U of θ in , with . Suppose , where , and 0 is either an isolated point in the spectrum or not in . Here, is a function defined on N—the null space of L. We refer to as the characteristic submanifold of M for g at p with respect to the parametrization Φ. The following theorem relates the critical groups of f to those of . This is known as the Shifting Theorem:
Lemma 3 (Shifting Theorem, [
13,
33])
. Assume that the Morse index of g at p is j. Then, The following lemma addresses the relationship between isolated critical points and characteristic manifolds:
Lemma 4 ([
33])
. Let p be an isolated critical point of g, and let be a closed Hilbert submanifold of M such that contains the null space of the Hessian of g. Assume that for all . If is sufficiently small and characteristic for at p, then N is also characteristic for g. In particular, we have the following: We are now ready to provide the proof of Theorem 1.
Proof of Theorem 1. We divide our proof into two parts. First, we address part (i) of Theorem 1.
Let
. According to Proposition 1,
x is a critical point of
g, meaning that
. Let I denote the identity map on
H. It follows that
if and only if
Thus, in the decomposition Equation (
5), the matrix representation of
A can be written as
where
, and
. According to the definition of
P,
is invertible, and we have
The formal expression for
is
The second equation in Equation (
16) is satisfied if and only if
Differentiating both sides of Equation (
20) yields
where
denotes the identity map on
V. Based on Equation (
19), we obtain
Combining Equations (
21) and (
22), we obtain
Simple computations show that
since
is invertible by the choice of
C. Therefore, based on Equation (
24), we obtain
In the decomposition Equation (
5), we have
By direct computation, we find
Since
is invertible, we have
where
.
Lastly, we address part (ii) of Theorem 1. We split our proof into two cases.
Case 1. If
z is a nondegenerate critical point of
a, then according to (3°) of Proposition 1,
is a nondegenerate critical point of
g, and
. Consequently, according to Lemma 1, we have
Case 2. If
z is a degenerate critical point of
a with
, then from (3°) of Proposition 1,
is a degenerate critical point of
f, and
. According to Lemma 2, there exists a ball
centered at 0 with radius
and a local homeomorphism
defined on
with
such that
where
,
, and
constitute a function defined on
.
Let
N be the characteristic submanifold for
a at
z with respect to
. According to Lemma 3, we have
Now, consider the critical point
of
f. According to Lemma 2, there exists a local homeomorphism
. Let
be the characteristic submanifold for
g at
with respect to
. Then, according to Lemma 3, we have
Since the map
u defined in Proposition 1 is an injection,
is a closed Hilbert submanifold of
H. Thus, in accordance with the (3°) of Proposition 1, we have
Define
, and let
be a characteristic manifold for
. According to Lemma 4, we obtain
From Equation (
27), we obtain
Thus, according to Lemma 3, we have
We then have the following map between the topological pairs:
Clearly,
u is a homeomorphism with inverse
. Therefore, we have
Combining Equations (
30), (
31), (
33) and (
36), we obtain
This concludes the proof. □
3. Applications to Second-Order Hamiltonian Systems
In this section, we will utilize Morse theory and a critical point theorem to establish the existence of two nontrivial solutions for the second-order Hamiltonian system (Theorem 2).
Let
be the Hilbert space equipped with the norm
The corresponding inner product in
L is denoted by
. Now, consider the space
, which consists of vector functions from
to
with square-integrable first-order derivatives. The norm on
W is defined as
This norm turns
W into a Hilbert space, known as the Sobolev space of
-periodic functions, which is a dense subspace of
L.
Within the Hilbert space
L, we define the linear operator
as
The operator
T has a closed range, and its resolvent is compact. The spectrum of
T under the
L norm is given by
, consisting only of eigenvalues and indicating that it is a point spectrum.
We continue to denote by
A the self-adjoint operator on
L induced by
A, which is defined as
Similarly, the operator
B, self-adjoint on
L, is defined by
Now, define
. The eigenvalue problem
is equivalent to the eigenvalue problem
Consequently, the spectrum of the operator
is also a point spectrum.
Given the conditions
and
, we define a functional on the space
L as follows:
This functional satisfies
, and we have
It is important to note that
is Gâteaux-differentiable, and its Gâteaux derivative is given by
Furthermore, we have the estimate
where
is the constant defined in condition
.
We then introduce a new functional
h on
W defined by
Condition
ensures that
and that
is Gâteaux-differentiable. Consider the following equation in the space
L:
This Equation (
48) is the Euler equation of the functional
h on the space
L, which can be expressed as follows:
Therefore, the critical points of
h are solutions to the Equation (
48). Moreover, this Equation (
48) is equivalent to the
-periodic boundary value problem of the Hamiltonian system Equation (
7). Consequently, finding
-periodic solutions of the Hamiltonian system Equation (
48) is equivalent to finding the critical points of the functional
h in the space
W.
Remark 2. In this context, the operator corresponds to the operator A in the abstract operator Equation (1). Consider the projection operator
, and define an operator on
W as follows:
This operator
is a self-adjoint, invertible Fredholm operator. Let
denote the spectral family associated with
. We define the projections on the space
L by
With these projections, the Hilbert space
L can be orthogonally decomposed as
where
,
, and
.
By applying the saddle point reduction, Theorem 1, the Equation (
48) is reduced to
where
.
It is evident that 0 is a critical point for both
and
. According to Theorem 1, we have
where
.
To prove Theorem 2, we first introduce a key critical point theorem:
Proposition 3 ([
34], Theorem 2.1)
. Let X be a Banach space, and let satisfy the Palais–Smale (PS) condition. Assume that f is bounded from below. If for some , then f has at least three critical points. In the following section, we will verify that the second-order Hamiltonian system Equation (
7) meets the conditions of Proposition 3.
Lemma 5. Assuming conditions and are satisfied, the functional is bounded from below and meets the (PS) condition.
Proof. Consider . We define .
For any
, we introduce the function
A straightforward computation yields
It is evident that
Since
, there exists a constant
such that
According to condition
, we have
Thus, for any
, the functional
is concave along any line passing through
and thus achieves a global maximum at
. Specifically,
By
and the definition of
, we have
where
is the minimum eigenvalue of Equation (
44) within
. Therefore,
is bounded from below and satisfies the (PS) condition. □
Consider a Hilbert space
W and
. By the Sobolev inequality, there exists a constant
such that
This inequality implies that the embedding
is both continuous and compact. Consequently, to establish the existence of nontrivial critical points, it suffices to verify that the corresponding critical group is nontrivial.
Recall the condition
, which states that there exists an eigenvalue
. Given that
is positive definite for all
, consider the following weighted eigenvalue problem:
According to the spectral theory of compact self-adjoint operators, it is well established that there exists a complete sequence of distinct eigenvalues
such that
as
.
Define the subspaces as follows:
We can derive the following result through a straightforward calculation:
Lemma 6. - (i)
If condition is satisfied, then ;
- (ii)
If condition is satisfied, then .
Proof. We provide a complete proof for case here; the proof for the other case follows a very similar argument.
According to ([
35], Theorem 2.1), it is sufficient to verify that
h has a local linking with respect to the decomposition
under condition
. Specifically, there must exist a
such that
The argument for Equation (
65) is similar to that presented in [
36] (p. 24) and ([
21], Lemma 4.1).
Since
is an isolated eigenvalue, it is well known that there exists a positive number
such that
Based on Equations (
8) and (
11), there exists a constant
such that
For
with
, we can write that
, where
, and
. Define
Since
, there exists
such that
for all
.
For every
, we have
According to Equation (
67), for
, we obtain
This result also applies to
because, in this case,
Hence,
by our assumption
.
According to Equation (
61), we obtain
Now, let
such that
If
, from Equation (
70) we can deduce that
, since
. If
, then
, and
. By condition
again, we also have that
For
, according to Equation (
67), we obtain
provided that
.
Combining these results, we conclude that Equation (
65) holds with
. □
We are now ready to present the proof of Theorem 2.
Proof of Theorem 2. We will provide the full proof for the case where condition is met. The proof for the other case follows similarly.
According to Lemma 5, the functional
satisfies the (PS) condition and is bounded below. Given that 0 is a critical point of
h, it is also a critical point of
, and it holds that
. According to Lemma 6 and Theorem 1, we obtain
Consequently, according to Lemma 3, if
, there exist two additional critical points of
. Therefore,
h must have at least two nonzero critical points. This, in turn, implies the existence of at least two solutions to the Hamiltonian system Equation (
7). This completes the proof. □
4. Conclusions
Infinite dimensional Morse theory is a highly effective tool for analyzing multiple solution problems in nonlinear differential equations. One of its key concepts is the critical group for a functional f at an isolated critical point x. This critical group captures the local behavior of f near x. In many applications, critical groups help in distinguishing between critical points and, furthermore, can be used to identify new critical points through the Morse inequality. Consequently, studying the critical group is crucial.
On the other hand, the saddle point reduction (also known as the Lyapunov–Schmidt Reduction) is a powerful technique for finding the critical points of some functionals. Essentially, under certain conditions, this reduction yields a reduced functional on a subspace that is closely related to the study of the original functional, particularly concerning the relationship between critical points.
In this paper, we first establish the relationship between the Morse index and the critical group before and after the saddle point reduction. We then apply our abstract results to examine the existence of two nontrivial solutions for the second-order Hamiltonian system. This approach offers a new perspective for identifying nontrivial solutions in other boundary value problems.