1. Introduction and Main Results
The study of the structure of solution sets began with Peano in 1890. He proposed the existence theorem of the solution set. Later, Kneser generalized Peano’s theorem that the solution set was not only nonempty, but also compact and connected. In 1942, Aronszajn further improved the theorem by showing that the solution set was even an
-set. Today, the topological structure of the solution set continues to arouse research enthusiasm. It is important not only from the viewpoint of academic interest, but also has a wide range of applications in practice. As is known to all, the topological structure of the solution set to differential inclusions is linked to obstacle problems, processes of controlled heat transfer, describing hybrid systems with dry friction and others (see, e.g., [
1,
2,
3] and references therein).
Some scholars have conducted numerous studies on the topological structure of the solution set to differential equations or inclusions, especially Gabor, who presented many results on the topological structure of the solution set and developed effective techniques for dealing with the structure of fixed point sets (see [
4,
5,
6]). Concerning other related results, we refer readers to Wojciech [
7], Zhou et al. [
8], Cheng et al. [
9], Grniewicz [
10], Andres [
11] and Djebali [
12,
13]. However, all the above results are for integer-order differential systems. It is known that fractional differential systems can better describe practical problems than integer differential systems, especially in describing the memory and hereditary nature of various processes and materials. Based on the background of the above research, are there similar conclusions for fractional differential systems?
In recent years, some results have been obtained on the structure of the solution set of fractional differential systems. For Riemann–Liouville fractional derivatives, Bugajwska-Kasprzak studied two Aronszajn-type theorems for some initial value problems; see [
14]; Ziane [
15] used a condensing map and Chalco-Cano et al. [
16] used the extended Kneser’s theorem to research an initial value problem for nonlinear fractional differential equations. For Caputo fractional derivatives, Wang et al. [
17] illustrated that a fractional control problem was approximately controllable by studying the structure of its solution set; Hoa et al. [
18] used Krasnosel’skiǐ-type operators to prove the
-property of the solution set of a fractional neutral evolution equation, and used the inverse limit method to obtain the same result on the half-line. There are also many studies on the stability of fractional differential systems. A. Singh et al. [
19] discussed the asymptotic stability of stochastic differential equations of fractional order
in Banach spaces. For more details, see the research articles [
20,
21].
Nevertheless, most of the papers describe research on the fractional differential equation on finite intervals, with related results about fractional differential inclusion on infinite interval being very rare. Motivated by the above consideration, we study the following fractional differential inclusion problem:
where
denotes the Caputo fractional derivative with order
and is a matrix operator;
is a set-valued function,
Specifically, we consider the following three cases:
:
m is a determined constant;
is a single-valued function.
We use to denote a single-valued function.
:
m is infinite;
is a single-valued function.
where
is a single-valued function.
:
m is infinite;
is a set-valued function.
where
is a set-valued function, and
is defined in
Section 4.
The first contribution of this paper is to investigate the
-property of the solution set to a fractional differential inclusion with time delay defined on the infinite interval; it is embodied in the proof of the solution set to problem (
2), which is an
-set. Furthermore, to the best of our knowledge, there are many studies on the nonlocal problem concerning fractional differential inclusion, but few involve the nonlocal function being set-valued. Stimulated by this consideration, we study the topological structure of the solution set to the nonlocal problem of a fractional differential inclusion, for cases of the nonlocal set-valued function with convex or nonconvex value, and this is the second contribution of this paper. It is reflected in the proof of the
-property of the solution set to problem (
3).
Throughout the paper, denotes the sup-norm of Banach space , where stands for the Euclidean norm in , and denotes the norm of .
In order to put forward our main results, we will first present all the hypotheses that we need for this paper.
Hypothesis 1 (
H1).
is an integrable matrix function, which satisfies , for a.e. .
Hypothesis 2 (
H2).
is an upper carathéodory function with compact and convex value:
(i) for every there exists a function such thatwhere denotes the Hausdoff metric, satisfies is a constant defined in Section 3. (ii) there exists an integrable function and a continuous function with for all such thatfor all Hypothesis 3 (
H3).
The fractional equation has a unique solution for the Cauchy problem, where for .
Hypothesis (H4). is u.s.c with convex and compact value, and satisfies
(i) for everywhere is the unique continuous solution of the integral equation in the form (ii) If is relatively compact in , then is relatively compact in .
Hypothesis 5 (
H5).
For all and , then , where .
Hypothesis 6 (
H6).
is l.s.c with closed value, and holds.
Hypothesis 7 (
H7).
, and there exists a constant , such that where in the case of .
The main results are stated as follows:
We first study the topological structure of the solution set to a fractional differential inclusion with time delay on compact intervals.
Theorem 1. If the hypotheses (H1)–(H3) are satisfied, the solution set to the inclusion problem (1) is an -set. Then, by means of the inverse limit method, we generalize the -property of the solution set to the inclusion to noncompact intervals.
Theorem 2. If the hypotheses (H1)–(H3) are satisfied, the solution set to the inclusion problem (2) is an -set. When the nonlocal function is set-valued and with convex value, we study the existence theorem of a solution to the nonlocal problem.
Theorem 3. If the hypotheses (H1)–(H4) are satisfied, the nonlocal problem (3) has at least one solution. Further, we investigate the topological structure of the solution set to the nonlocal problem.
Theorem 4. If the hypotheses (H1)–(H5) are satisfied, the solution set to the nonlocal problem (3) is an -set. Changing the convex condition of the nonlocal function to the unconvex condition, we still obtain the property of the solution set to the nonlocal problem.
Theorem 5. If the hypotheses (H1)–(H3), (H5)–(H7) are satisfied, the solution set to the nonlocal problem (3) is an -set. The rest of this paper is organized as follows. In
Section 2, we present some definitions and lemmas for the fractional calculus and topological structure of the solution set. In
Section 3,
Section 4,
Section 5,
Section 6 and
Section 7, we complete the proofs of Theorems 1–5, respectively, and
Section 8 presents the conclusions.
2. Preliminaries
This section provides some properties of fractional calculus and some notions of the topological structure of the solution set that will be needed in our analysis. For more results on fractional calculus, we refer readers to [
22,
23,
24,
25,
26], and for the topological structure of the solution set, we refer the interested readers to [
4,
6,
27].
Definition 1. The Riemann–Liouville fractional integral of order for a function f is defined aswhere is the Gamma function. Definition 2. The Caputo fractional derivative of order for a function f can be written aswhere There is an important property of the Caputo fractional derivative, which will be used in our following proof:
If
and
, then
The following lemmas are crucial in our research:
Lemma 1. [28] Assume that is a vector, where are continuous differentiable functions for all , and is a positive definite matrix. Then, for any Lemma 2. [29](Gronwall inequality) Suppose that , is a nondecreasing function on and is a nonnegative, nondecreasing continuous function defined on , and suppose that is nonnegative and integrable on withon this interval. Then,where , is the single-parameter Mittag–Leffler function. Let
X be a Housdorff topological space. For
, the Hausdorff metric is obtained by
According to the metric, let
be a multivalued map with bounded value; if there exists a constant
such that
then
F is called Lipschitzian, and if the constant
,
F is called a contraction.
Let U be a nonempty subset of X, and suppose that there exists a retraction such that is the identity map; then, U is called a retract of X. Clearly, a retract is closed.
Definition 3. [30] Let X be a metric space and U be a closed subset of X. For every metric space Y and a closed set , (i) U is called an absolute retract (AR space), if each continuous map can be extended to a continuous function .
(ii) U is called an absolute neighborhood retract (ANR space), if there exists a neighborhood , such that the continuous map can extend to be a continuous map .
From the definition, it is easy to see that the AR space contains the ANR space. Furthermore, in a Fréchet space, a retract of the convex set must be an AR space (see [
4]). In particular, each Banach space is an AR space. Space
is an AR space, where
is an arbitrary interval.
Definition 4. Let X be a metric space and U be a subset of X. U is said to be contractible, if there exist a continuous function and a point , such that and for all .
Definition 5. A subset U of a metric space X is called an -set, if there exists a decreasing sequence of absolute retracts satisfying Specifically, if every is compact, U is called a compact set. If every is symmetric, U is called a symmetric -set.
We can find that if a set
U is a compact
-set, then it must be nonempty, compact and connected. Thus, the following hierarchy for nonempty subsets of a metric space is true:
and all the above inclusions are proper.
Now, we present some useful facts connected with the semicontinuity of the set-valued map.
Definition 6. A set-valued map is called upper semicontinuous (u.s.c.) provided that, for every open subset , the set is open in X.
Definition 7. A set-valued map is called lower semicontinuous (l.s.c.) if, for every open subset , the set is open in X.
Proposition 1. [27] A set-valued map is u.s.c. if and only if, for every closed set , the set is a closed subset of X. Proposition 2. [27] Let be metric spaces and Z be a compact set, if a set-valued map , with the closed graph, such that , then is u.s.c. Definition 8. Let be metric spaces; a set-valued map is called an -map, if is u.s.c and is an -set for every .
Clearly, if a set-valued map with contractible value is u.s.c., it can be seen as an
-map. Let
and
be two set-valued maps; the composition
is defined in the following form:
for each
. We obtain the proposition as follows:
Proposition 3. [27] Let and be two u.s.c. maps with compact values; then, the composition is an u.s.c. map with compact values. denotes the fixed point set of , and when is set-valued and contractional, is more complex. Thus, the topological property of is a question worth researching.
Lemma 3. [31] Let X be a Banach space, and Y be a closed, convex subset of X; if set-valued map is a contraction with compact, convex value, then is a nonempty, compact AR space. Lemma 4. [27] Let be an u.s.c. map with compact value and A be a compact subset of X. Then, is compact. What follows is a fixed point theorem due to Górniewicz and Lassonde [
32] (Corollary 4.3), which plays an important role in our proof.
Lemma 5. Let Y be an ANR space. Suppose that can be factorized as , where are -maps, are ANR spaces, and are AR spaces. If there exists a compact subset satisfying , then α admits a fixed point.
Lemma 6.(Bressan–Colombo continuous selection theorem) Let X be a measurable and separable Banach space, and be a finite measure space. Suppose that is a set-valued map with closed decomposable values and l.s.c. Then, H has a continuous selection.
4. Proof of Theorem 2
In this section, we will study the -property of the solution set to the fractional differential inclusion defined on the half-line. In order to study the problem on an infinite interval, we recall some related knowledge of the inverse system.
The system
can be known as an inverse system, where
is a set denoted for the relation ≤, for all
;
is a metric space, and for all
with
,
is a continuous function.
is defined as the limit of inverse systems
S, in the form of
For more details, we refer readers to [
4,
5]. The following lemmas are useful for our research.
Lemma 7. ([4] Theorem 3.9) Let be an inverse system, and be a limit map derived by a family , where , and if all , are -sets, then is an -set too. Lemma 8. [35] Let be an inverse system. For each , if is nonempty and compact (relatively compact), then is also nonempty and compact (relatively compact). Before the proof, we need to present some notations.
For any
and
, we define a projection
, in the form of
Let
, and it is easy to see that
is an inverse system, and its limit is isometrically homeomorphic to
; then, for convenience, we express
Then,
is an inverse system and its limit can be represented as
Moreover,
is an inverse system with
and its limit is written as
where
is the separated locally convex space, which can be composed of all locally Bocher integrable components from
to
endowed with a family of seminorms
, defined by
.
It is easy to see that for with ; thus, the family is a map from to . As a result, for all with , the family induces a limit map , such that In what follows, we demonstrate the topological structure of the solution set on the infinite intervals.
Proof. From the proof of Theorem 1, we see that the fixed point set of set-valued map
is the solution set to problem (
1), and it is an
-set.
What follows is to indicate that the family
is a map from
to
. Since
, we only need to show that, for
,
where
is defined in (
9) with
n instead of
m. In the case of
, it is obviously true. In the case of
, it is clear that
Thus, we only need to explain the reverse inclusion. For
and
, we set
, where
is the characteristic function. It is obvious that
, which means that
Combining (
12) and (
13), we have
Therefore, the map
induces the limit map
, such that
. Consequently, the fixed point set of the map
is the solution set to inclusion problem (
2). As
is an
-set, invoking Lemma 7, for each
,
is an
-set too. The proof is completed. □