1. Introduction
Let be the n-dimensional space of real column vectors with any norm . Given , let denote the Banach space of continuous functions from into with the supremum norm for .
Consider the linear autonomous retarded functional differential equation
where
is a bounded linear functional and
is defined by
for
. According to the Riesz representation theorem,
L has the form
where
is a matrix function of bounded variation normalized such that
is left continuous on
and
. Equation (
1) includes as a special case the differential equation with a single delay
where
. By a solution of (
1), we mean a function
which is continuous on
, differentiable on
, and satisfies (
1) for
. It is well-known [
1] that, for every
, Equation (
1) has a unique solution with initial value
.
The characteristic equation of Equation (
1) has the form
where
is a complex variable and
I is the
identity matrix.
A solution
of (
1) is called oscillatory if all coordinate functions
have arbitrarily large zeros. Otherwise,
x is called nonoscillatory. As usual, the superscript
T indicates the transpose. If Equation (
4) has a real root
, then there exists
such that
and hence the corresponding eigensolution
is a nonoscillatory solution of (
1). Thus, the existence of a real root of Equation (
4) is sufficient for the existence of a nonoscillatory solution of (
1). One of the key results in the oscillation theory of functional differential equation [
2] states that the existence of a real root of Equation (
4) is not only sufficient, but also necessary for the existence of a nonoscillatory solution of (
1).
Theorem 1. ([
3], Theorem 4.1)
Equation (1) has a nonoscillatory solution if and only if Equation (4) has a real root. The aim of this paper is to give a similar characterization for the existence of a nontrivial nonnegative solution of Equation (
1). A solution
of (
1) is called nonnegative if all coordinate functions
are nonnegative on
, or, equivalently,
for
, where
is the nonnegative orthant, the set of those vectors in
which have nonnegative components. If there exist
and
such that
, then
is evidently a nontrivial nonnegative solution of Equation (
1). Therefore it is natural to ask whether the existence of
and
such that
is necessary for the existence of a nontrivial nonnegative solution of (
1). The following simple example shows that the answer in general is negative.
Consider the two-dimensional system
a special case of Equation (
3), where
and
. The characteristic matrix is given by
In this case, there exist no
and
such that
. Otherwise, we obtain
, a contradiction. On the other hand, Equation (
5) has the nontrivial nonnegative solution
given by
Note that the above nontrivial nonnegative solution
x is a small solution in the sense of the following definition [
1]. A solution
x of Equation (
1) is called a small solution if
The zero solution is always a small solution. The question is whether there exist initial conditions
which generate small solutions. Such solutions are called nontrivial small solutions. A linear autonomous ordinary differential equation cannot have a nontrivial small solution. The existence of nontrivial small solutions of Equation (
1) is a consequence of the fact that the phase space
C is infinite dimensional. As shown in ([
1], Chap. 7, Corollary 8.1), Equation (
1) has no nontrivial small solutions if and only if the exponential type of the characteristic function
is equal to
, or, equivalently, the system of eigenfunctions and generalized eigenfunctions of the generator of Equation (
1) is complete. As a corollary, we have that Equation (
3) has no nontrivial small solutions if and only if
.
In this paper, we will show that if we exclude the existence of nontrivial small solutions, then the “natural” sufficient condition
for some
and
is necessary for the existence of a nontrivial nonnegative solution of Equation (
1).
The main result and its proof are given in
Section 2. In
Section 3, we briefly mention some results which are relevant to our study.
2. Main Result
Our main result is the following theorem.
Theorem 2. Suppose that Equation (1) has no nontrivial small solutions. Then Equation (1) has a nontrivial nonnegative solution if and only if there exist and such that . Before we present the proof of Theorem 2, we recall some facts from the decomposition theory of linear autonomous functional differential equations given in ([
1], Chap. 7) and we establish two preliminary results.
It is known that Equation (
1) generates in
C a strongly continuous semigroup
, where
is a bounded linear operator, the so-called solution operator, defined by
for
and
,
being the unique solution of (
1) with initial value
. The infinitesimal generator
of this semigroup is defined by
whenever the limit exists in
C. It is known that
The spectrum
of the linear operator
is a point spectrum and it consists of the roots of Equation (
4). In each strip
, where
, Equation (
4) has only a finite number of roots. Furthermore, if
is a finite set of characteristic roots, then
C is decomposed by
into a direct sum
where
is the (realified) generalized eigenspace of
A associated with
and
is the complementary subspace of
C such that
for
. Thus, each
can be written uniquely as
From now on, let
denote the set of nonnegative functions in
C, i.e.,
The following lemma will play an important role in the proof of Theorem 2.
Lemma 1. Suppose that Equation (1) has no nontrivial small solutions. If x is a nontrivial nonnegative solution of Equation (1), then its Lyapunov exponent μ given byis finite, and if we letthen there exists such thatwhere is the generalized eigenspace of the infinitesimal generator A associated with . Proof. Let
be a nontrivial nonnegative solution of Equation (
1) with initial value
and Lyapunov exponent
. Then
for all
. By the assumptions,
x cannot be a small solution and hence
is finite (see [
1], Chap. 7, Theorem 6.1). Define
so that
. The generalized eigenspace
can be further decomposed into the direct sum
with
as in (
11) and
Thus, writing
,
,
and
for brevity, we have that
and hence
where
,
and
. All three subspaces
,
and
Q are invariant under the solution semigroup. The generalized eigenspaces
and
are finite-dimensional and the solutions starting from
and
can be extended backward to all
. As a consequence, on
and
the solution semigroup can be extended to a group. It is known that for every
there exists
such that the following exponential estimates hold:
where
m is the maximum of the ascents of the characteristic roots from
(see ([
1], Section 7.6) and ([
4], Equations (3.16) and (3.17))). Replacing
with
,
, in the last inequality and using the group property
on
, we find that
This implies that
. Otherwise, the last inequality, combined with (
15) and the previous exponential estimates, would imply that
, a contradiction. We claim that
≠ 0. Otherwise, by virtue of (
15), we have that
for
, which, together with the exponential estimate on
Q, implies that
, a contradiction. From the exponential estimate on
, we find that
Since
, this, together with the exponential estimate on
Q, implies that
Thus,
is a bounded sequence in the finite-dimensional and hence closed subspace
of
C. Therefore there exist
with
and a sequence
such that
It remains to show that
has the desired properties. Since
and
is invariant under the solution semigroup, we have that
=
for
. As shown before,
=
as
. Hence
From this, using the nonnegativity of
, the continuity and the semigroup property of
, we find that
for all
. ☐
Let X be a real Banach space. A subset is called a cone if the following three conditions hold:
- (i)
K is a nonempty, convex and closed subset of X,
- (ii)
for all , where ,
- (iii)
, where .
In the proof of Theorem 2, we will need the following result which gives a necessary and sufficient condition for the existence of a nontrivial orbit of a linear invertible map which lies in a given cone K. By an orbit starting from , we mean the sequence of iterates . As usual, , the identity on X.
Lemma 2. Let K be a cone in a finite-dimensional real Banach space X. Suppose that is a linear invertible operator. Then M has an orbit belonging to if and only if M has a positive eigenvalue with an eigenvector in K.
In the special case
, Lemma 2 was proved in ([
5], Theorem 3). Here we give a different argument which is valid in general finite-dimensional Banach spaces.
Proof of Lemma 2. If is an eigenvector of M corresponding to a positive eigenvalue , then . Thus, the orbit starting from v belongs to .
Now suppose that
M has an orbit starting from
which lies in
. In particular,
. Without loss of generality, we may (and do) assume that
. Otherwise, we replace
y with
and use the cone property (ii). Let
Evidently,
S is a convex closed subset of
X and
. Define an operator
by
The cone property (iii), the fact that
and the invertibility of
M imply that
F is well-defined. Evidently,
F is continuous on
S. The definition of
S and the cone properties (i) and (ii) imply that
⊂
S. By Brouwer’s fixed point theorem, there exists
S such that
=
v. Since
=
= 1, it follows that
Hence , where . Since and M is invertible, we have that and hence . Thus, is a positive eigenvalue of M and is a corresponding eigenvector. ☐
Now we can give a proof of Theorem 2 which follows similar lines as the proof of a Perron type theorem for positive solutions of a perturbed system of nonautonomous linear functional differential equations in [
6].
Proof of Theorem 2.
As noted before, if there exist
and
such that
, then
is a nontrivial nonnegative solution of (
1).
Now suppose that Equation (
1) has a nontrivial nonnegative solution
x. By Lemma 1, the Lyapunov exponent
is finite. Let
be the spectral set defined by (
11). As noted before, the associated generalized eigenspace
of (
1) is finite-dimensional and invariant under the solution semigroup
with infinitesimal generator
A given by (
7). Since
is finite-dimensional, it is a closed subspace of
C and therefore we can define the subspace semigroup
on
by
, the restriction of
to
([
7], Paragraph I.5.12). Its generator is
with domain
([
7], Paragraph II.2.3) and
. Since
, the generator
is bounded and therefore
for
([
8], Chap. I, Section 1.1). According to the spectral mapping theorem [
8], we have that
Define
. By Lemma 1, there exists
such that
for all
. Since
, we have that
is invertible for
. Hence
Evidently,
K is a cone in
. Let
be a sequence of positive numbers such that
as
. For every fixed
k, consider the linear operator
in
. By the semigroup property, we have that
for
. This, together with (
20), implies that
M has an orbit which belongs to
. By the application of Lemma 2, we conclude that
has a positive eigenvalue
with an eigenvector
. Without loss of generality, we may assume that
. Otherwise, we replace
with
which belongs to
K by the cone property (ii). By virtue of (
19),
for some characteristic root
z with
. From this, using the positivity of
, we find that
Hence
for
. Since
is a bounded sequence in the finite-dimension Banach space
, there exists a subsequence
of
such that the limit
exists in
. Evidently,
. Since
K is closed subset of
, we have that
. From (
21), we find that
for
. From this, letting
, using (
22) and the fact that
we obtain
Hence for . Since , we have that . Finally, implies that which is equivalent to . ☐
3. Discussion
The basic oscillation theorem for differential equations with constant coefficients and several delays was obtained by Arino and Győri [
9]. A generalization to a class of linear differential equations with distributed delays was given by Győri and Krisztin [
3]. Krisztin [
10] showed that linear functional differential equations of mixed type may have nonoscillatory solutions in spite of the nonexistence of a real root of the characteristic equation. Henry [
11] proved that small solutions of linear autonomous retarded functional differential equations must vanish after some time. Henry’s theorem was improved by Verduyn Lunel [
12]. Further information on small solutions and the completness of the eigenfunctions and generalized eigenfunctions of linear autonomous functional differential equations can be found in the monographs by Hale and Verduyn Lunel [
1] and Dieckman et al. [
13]. Small solutions for nonlinear equations were studied for the first time by Mallet Paret [
14], who showed that they do not exist on the attractor of certain nonlinear scalar delay differential equations. For later results, see the papers by Cao [
15], Arino [
16], Cooke and Verduyn Lunel [
17], Mallet Paret and Sell [
18], Braverman et al. [
19], Garab [
20] and the references therein.
The main result of this paper, Theorem 2, is closely related to Theorem 1.5 from our recent work [
6], which shows that if
K is a cone in
, then Equation (
1) has a positive solution with respect to the partial order induced by
K if and only if Equation (
4) has a real root with a positive eigenfunction. In the case
, Theorem 2 improves ([
6], Theorem 1.5) in the sense that while the latter theorem applies only to positive solutions, Theorem 2 provides a similar conclusion for the larger class of nontrivial nonnegative solutions. It should be noted that ([
6], Theorem 1.5) is a corollary of a more general Perron type theorem for positive solutions of a perturbed system of functional differential equation with a long proof. We believe that the above short proof of Theorem 2 can be of interest.