1. Introduction
We study the existence and localization of positive solutions for the system
subject to the Sturm–Liouville boundary conditions (
7).
The novelties in this paper are in two directions. On the one hand, we allow the functions () to be discontinuous with respect to the unknown over some time-dependent sets, see Definitions 1 and 2. On the other hand, in order to localize the solutions of the system, we shall establish a multivalued vector version of Krasnosel’skiĭ’s fixed point theorem which allows different asymptotic behaviors in the nonlinearities and , see Remark 3.
The existence of discontinuities in the functions
or
makes impossible to apply directly the standard fixed point theorems in cones for compact operators since the integral operator corresponding to the differential problem is not necessarily continuous. In order to avoid this difficulty, we regularize the possibly discontinuous operator obtaining an upper semicontinuous multivalued one. Then we look for fixed points of this multivalued mapping that are proved to be Carathéodory solutions for the differential system. In the case of scalar problems, similar ideas appear in the papers [
1,
2,
3].
This approach of using set-valued analysis in the study of discontinuous problems is a classical one, see [
4]. Nevertheless, the regularization is usually made in the nonlinearities transforming the problem into a differential inclusion and the solutions are often given in the sense of the set-valued analysis (Krasovskij and Filippov solutions [
5,
6]), see e.g., [
7,
8]. Similar ideas are also used in the papers [
5,
9] where there are provided some sufficient conditions for the Krasovskij solutions to be Carathéodory solutions. Recently, second-order scalar discontinuous problems have been investigated by using variational methods [
10,
11,
12]. However, in these papers there are not considered time-dependent discontinuity sets. Observe also that a lot of existence results for discontinuous differential problems are based on monotonicity hypotheses on their nonlinear parts, see [
13], but such assumptions are not necessary in our approach.
Going from scalar discontinuous problems to systems of discontinuous equations is not trivial and it makes possible to consider two different notions for the discontinuity sets. The first approach (see Definition 1 and Theorem 3) allows to study the discontinuities in each variable independently. For instance, it guarantees the existence of a positive solution for the following particular system
subject to the Sturm–Liouville boundary conditions, where
is the Heaviside step function given by
see Example 1. Notice that the nonlinearities in this example are discontinuous at
for each
and at
for every
. Moreover, the first nonlinearity has a superlinear behavior and the second one has a sublinear one. Our second approach allows to study functions which are discontinuous over time-dependent curves in
and the conditions imposed to these curves are local, see Definition 2 and Theorem 4. In particular, we establish the existence of a positive solution for the system
subject to the Sturm–Liouville boundary conditions.
As mentioned above, our results rely on fixed point theory for multivalued operators in cones. We finish this introductory part by recalling the version of Krasnosel’skiĭ’s fixed point theorem for set-valued maps given by Fitzpatrick–Petryshyn [
14].
Theorem 1. Let X be a Fréchet space with a cone . Let d be a metric on X and let , , and usc and condensing. Suppose there exists a continuous seminorm p such that is p-bounded. Moreover, suppose that F satisfies:
- 1.
There is some with and such that for any and ;
- 2.
for any and .
Then F has a fixed point with .
In the case of a Banach space
and of an operator
under the hypotheses of the previous theorem, we obtain the existence of a fixed point
for
F such that
, where
denotes a norm in
, for example,
. Then
and
, but it is not possible to obtain a lower bound for the norm of every component. This fact motivates the use of a vector version of Krasnosel’skiĭ’s fixed point theorem. Such a version was introduced in [
15] for single-valued operators. Another advantage of the vector approach is that it allows different behaviors in each component of the system.
2. Multivalued Vector Version of Krasnosel’skiĭ’s Fixed Point Theorem
In the sequel, let be a Banach space, two cones and the corresponding cone of . For , , , we denote
The following fixed point theorem is an extension of the vector version of Krasnosel’skiĭ’s fixed point theorem given in [
15,
16] to the class of upper semicontinuous (usc, for short) multivalued mappings.
Theorem 2. Let with , and for . Assume that , , is an usc map with nonempty closed and convex values such that is compact, and there exist , , such that for each the following conditions are satisfied:Then N has a fixed point in K, that is, , with for . Proof. We shall consider the four possible combinations of compression-expansion conditions for and .
Assume first that
for both
(compression for
and
). Then
and
for
. Denote
and define the map
given, for
, by
where
for
.
The map
is usc (the composition of usc maps is usc, see [
17], Theorem 17.23) and
is relatively compact since its values belong to the compact set
. Then Kakutani’s fixed point theorem implies that there exists
such that
.
It remains to prove that
. It is clear that
since
for
. Assume
and
. If
, then
so
what contradicts (
2) for
. Analogously, we can obtain contradictions for any other point
, as done in [
15,
16] for single-valued maps.
Assume that
(compression for
) and
(expansion for
). Let
(
) be given by
Notice that the map
is in case a, and thus
has a fixed point
. Further, the point
u defined as
and
is a fixed point of the operator
N.
The case
(expansion for
) and
(compression for
) is similar to the previous one by taking the map
defined as
The case
for
(expansion for
and
) reduces to case a, if we consider the map
where
is defined by (
4) and
, by (
3).
Therefore, the proof is over. □
Remark 1 (Multiplicity).
Although we are interested in fixed points for the operator N satisfying that both components are nonzero, if we replace conditions (1) and (2) in Theorem 2 by the following ones:then we can achieve multiplicity results. Indeed, if for or , then the operator N has one additional fixed point such that and with . Furthermore, if for , then N has three nontrivial fixed points. Such cases are considered in the paper [18] in connection with -Laplacian systems. Our purpose is to apply Theorem 2 to a multivalued regularization of a discontinuous system of single-valued operators associated to a system of differential equations with discontinuous nonlinearities. Our aim is to obtain new existence and localization results for such kind of problems.
In order to do that, we need the following definitions and results.
Let
U be a relatively open subset of the cone
and
,
, an operator not necessarily continuous. We associate to the operator
T the following multivalued map
given by
where
,
denotes the closure of the set
U with the relative topology of
K and
means closed convex hull. The map
is called the closed-convex envelope of
and it satisfies the following properties, see [
2].
Proposition 1. Let be the closed-convex envelope of an operator . The following properties are satisfied:
- 1.
If T maps bounded sets into relatively compact sets, then assumes compact values and it is usc;
- 2.
If is relatively compact, then is relatively compact too.
Remark 2. The following two statements are equivalent:
- (a)
();
- (b)
for every and every there exist and a finite family of vectors and coefficients () such that and
3. Positive Solutions of Discontinuous Systems
We study the existence and localization of positive solutions for the following second-order coupled differential system
for
, with the following boundary conditions
for
, where
and
for
. Assume that, for
,
- (H1)
, for a.e. and ;
- (H2)
satisfies that
- (i)
are measurable whenever ;
- (ii)
for each
there exists
such that
Notice that condition is satisfied if is measurable for all constants and if is continuous for a.a. t, which is not necessarily the case in this paper.
Let
be the space of continuous functions defined on
I endowed with the usual norm
and let
P be the cone of all nonnegative functions of
X. A positive solution to (
6)–(
7) is a function
with
,
(
) such that
u satisfies (
6) for a.a.
and the boundary conditions (
7). The existence of positive solutions to problems (
6)–(
7) is equivalent to the existence of fixed points of the integral operator
,
, given by
where
are the corresponding Green’s functions which are explicitly given by
Denote
then it is possible to check the following inequalities:
Consider in
X the cones
and
defined as
and the corresponding cone
in
. Then,
. Indeed, for
and
,
Hence, for every and .
Therefore, it must be clear that we intend to apply Theorem 2 in a subset of
K to the multivalued operator
associated to the discontinuous operator
T. Later, we shall provide conditions about the functions
(
) which guarantee that
, where
stands for the set of fixed points of the mapping
S. As a consequence, we obtain some results concerning the existence of positive solutions for system (
6)–(
7).
Let us introduce some notations. For
with
and
, we let
,
(
) and
Also, denote
for
.
Lemma 1. Assume that there exist with , , and such thatThen, for each , the following conditions are satisfied:where and are constant functions equal to 1. Moreover, the map defined as in (5) has at least one fixed point in . Proof. First, observe that if
, then
and if
for some
, and
, then
for all
and
Now we prove (
10) for
. Assume that
and let us see that
for
. First, we shall show that given a family of vectors
and numbers
such that
(
), then
what implies that
. Indeed, if not, taking the supremum for
,
a contradiction. Notice that if
, then it is the limit of a sequence of functions satisfying the previous inequality and thus, as a limit, it satisfies
which is also a contradiction since
. Therefore,
for
.
In order to prove (
11) for
, assume that
and
for some family of vectors
and numbers
such that
(
) and some
. Then for
, we have
so
, a contradiction. Hence,
. As before,
because in that case we arrive to the inequality
for
. Therefore,
.
Similarly, it is possible to prove conditions (
10) and (
11) for
.
To finish, the conclusion is obtained by applying Theorem 2 to the operator . □
Remark 3 (Asymptotic conditions)
The existence of with , , and satisfying (9) is guaranteed, in the autonomous case, by the following sufficient conditions: - (a)
has a superlinear behavior and , a sublinear one, that is, - (b)
Both and have a superlinear behavior, that is, - (b)
Both and have a sublinear behavior, that is,
Remark 4. If and are monotone in both variables, it is possible to specify the numbers and (), so in this case, conditions (9) only depend on the behavior of the functions at four points in , see [15,16]. Note that Lemma 1 gives us sufficient conditions for the existence of a fixed point in
of the multivalued operator
. Hence, it remains to provide hypothesis on the functions
(
) which imply
in order to obtain a solution for the system (
6)–(
7). Observe also that no continuity hypotheses were required to the functions
until now.
The following definition introduces some curves where we allow the functions
to be discontinuous in each variable. The idea of using such curves can be found in some recent papers for second-order discontinuous scalar problems [
1,
2,
3] and, in some sense, it recalls the notion of time-depending discontinuity sets from [
9].
Definition 1. We say that , , is an inviable discontinuity curve with respect to the first variable if there exist and for a.e. such that eitheror Similarly, we say that , , is an inviable discontinuity curve with respect to the second variable if there exist and for a.e. such that eitheror Now we state some technical results that we need in the proof of the condition
. Their proofs can be found in [
3]. In the sequel,
m denotes the Lebesgue measure in
.
Lemma 2 ([
3], Lemma 4.1).
Let , , and let , a.e., and a.e. in . For every measurable set with there is a measurable set with such that for every we have Corollary 1 ([
3], Corollary 4.2)
Let , , and let be such that a.e. in . For every measurable set with there is a measurable set with such that for all we have We shall also need the following lemma, see [
2], Lemma 3.11.
Lemma 3. If , almost everywhere, then the setis closed in endowed with the maximum norm topology. Moreover, if for all and uniformly in , then there exists a subsequence which tends to u in the norm.
Now we are ready to present the following existence and localization result for the differential system (
6)–(
7).
Theorem 3. Suppose that the functions and () satisfy conditions , and
- (H3)
There exist inviable discontinuity curves with respect to the first variable, , and inviable discontinuity curves with respect to the second variable, , such that for each and for a.e. the function is continuous on
Moreover, assume that there exist with , , and such that Then system (6)–(7) has at least one solution in . Proof. The operator
,
, given by (
8) is well-defined and the hypotheses
and
imply that
is relatively compact as an immediate consequence of the Ascoli–Arzelá theorem. Moreover, by
and
, there exist functions
(
) such that
Therefore,
, where
for
, which by virtue of Lemma 3 is a closed and convex subset of
. Then, by `convexification’,
, where
is the multivalued map associated to
T defined as in (
5).
By Lemma 1, the multivalued map has a fixed point in . Hence, if we show that all the fixed points of the operator are fixed points of T, the conclusion is obtained. To do so, we fix an arbitrary function and we consider three different cases.
Case 1: for all . Let us prove that T is continuous at u, which implies that , and therefore the relation gives that .
The assumption implies that for a.a.
the mappings
and
are continuous at
. Hence if
in
then
which, along with (
14), yield
in
, so
T is continuous at
u.
Case 2: for some . In this case we can prove that , and thus .
To this aim, first, we fix some notation. Let us assume that for some
we have
and there exist
and
,
for a.a.
, such that (
13) holds with
replaced by
. (The proof is similar if we assume (
12) instead of (
13), so we omit it.)
We denote
, and we deduce from Lemma 2 that there is a measurable set
with
such that for all
we have
By Corollary 1 there exists
with
such that for all
we have
Let us now fix a point
. From (
15) and (
16) we deduce that there exist
and
,
sufficiently close to
so that the following inequalities are satisfied for all
:
and for all
:
Finally, we define a positive number
and we are ready to prove that
. It suffices to prove the following claim:
Claim: let
be given by our assumptions over
as Definition 1 shows, and let
, where
is as in (
21). For every finite family
and
(
), with
, we have
.
Let
and
be as in the Claim and, for simplicity, denote
. For a.a.
we have
On the other hand, for every
and every
we have
and then the assumptions on
ensure that for a.a.
we have
Now for
we compute
hence
provided that
. Therefore, by integration we obtain
So, if
, then
. Otherwise, if
, then we have
and thus
, too.
Similar computations in the interval instead of show that if then we have for all and this also implies . The claim is proven.
Case 3: for some . In this case it is possible to prove that . The details are similar to those in Case 2, with obvious changes, so we omit them. □
Remark 5. Observe that Definition 1 allows to study the discontinuities of the functions independently in each variable and , as shown in condition .
In addition, a continuum set of discontinuity points is possible: for instance, the function may be discontinuous at the point for all provided that the constant function is an inviable discontinuity curve with respect to the first variable. This fact improves the ideas given in [5] for first-order autonomous systems where “only” a countable set of discontinuity points are allowed. Remark 6. Notice that conditions (12) and (13) are not local in the last variable. However, the conditionimplies that any constant function stands for an inviable discontinuity curve with respect to the first variable (since condition (13) holds). Moreover, any function with strictly positive second derivative is always an inviable discontinuity curve with respect to the variable without any additional condition on . Now we illustrate our existence result by some examples.
Example 1. Consider the coupled systemsubject to the boundary conditions (7) (replacing and by x and y, respectively) where and H denotes the Heaviside function. The existence of numbers and in the conditions of (9) is guaranteed by Remark 3 since is a superlinear function and is a sublinear function. On the other hand, the function is continuous on and the constant function stands for an inviable curve with respect to the first variable. Indeed,hence (13) holds with . Moreover, the constant function is an inviable curve with respect to the second variable, according to Remark 6 since Similarly, the function satisfies the hypothesis in Theorem 3, so the system (7)–(24) has at least one positive solution. Example 2. Consider the systemsubject to the boundary conditions (7), where and . Now, for a.a. , the function , whereis continuous on and the curve is inviable with respect to the first variable. Indeed, (13) is satisfied with , since On the other hand, the curve is inviable with respect to the variable y, according to Remark 6, since and .
Therefore, Theorem 3 ensures the existence of one positive solution for problem (7)–(25). Nevertheless, the conditions of Definition 1 are too strong for functions
which are discontinuous at a single isolated point
or, more generally, over a curve
for
. This is the motivation for another definition of the notion of discontinuity curves. This notion will be a generalization of the admissible curves presented in [
2] for one equation.
Definition 2. We say that , (), is an admissible discontinuity curve for the differential equation if one of the following conditions holds:
- (a)
for a.e. (then we say γ is viable for the differential equation),
- (b)
There exist and for a.e. such that eitheror In this case we say that γ is inviable.
Similarly, we can define admissible discontinuity curves for .
Theorem 4. Suppose that the functions and () satisfy conditions , and
- ()
There exist admissible discontinuity curves for the first differential equation , , such that for a.e. the function is continuous on ;
- ()
There exist admissible discontinuity curves for the second differential equation , , such that for a.e. the function is continuous on .
Moreover, assume that there exist with , , and such that Then the differential system (6)–(7) has at least one solution in . Proof. Notice that in virtue of Lemma 1 it is sufficient to show that . Reasoning as in the proof of Theorem 3, if we fix a function , we have to consider three different cases.
Case 1: for all . Then T is continuous at u.
Case 2: or for some or inviable. Then . The proof follows the ideas from Case 2 in Theorem 3.
Case 3:
or
only for viable curves. Then the relation
implies
. In this case the idea is to show that
u is a solution of the differential system. The proof is analogus to that of the equivalent case in [
2], Theorem 3.12 or [
3], Theorem 4.4, so we omit it here. □
Remark 7. Notice that, in the case of a function which is discontinuous at a single point , Definition 2 requires that one of the following two conditions holds:
- (i)
for a.e. ;
- (ii)
there exist and , for a.e. such that
In particular, for (ii), it suffices that there exist such that To finish, we present two simple examples which fall outside of the applicability of Theorem 3, but which can be studied by means of Theorem 4.
Example 3. Consider the problemsubject to the boundary conditions (7). It is clear that and have a sublinear behavior, see Remark 3.
The function is continuous on and the constant function is an inviable admissible discontinuity curve for the differential equation since for all and all ; and .
Therefore, Theorem 4 guarantees the existence of a positive solution for problem (7)–(26). Example 4. Consider the following systemsubject to the boundary conditions (7). The nonlinearities of the system have again a sublinear behavior. Now, the function is continuous on and the constant function is a viable admissible discontinuity curve for the differential equation.
Hence, by application of Theorem 4, one obtains that the system (7)–(27) has at least one positive solution.