Abstract
We consider a class of nonautonomous stochastic evolution equations in real separable Hilbert spaces. We establish a new composition theorem for square-mean almost automorphic functions under non-Lipschitz conditions. We apply this new composition theorem as well as intermediate space techniques, Krasnoselskii fixed point theorem, and Banach fixed point theorem to investigate the existence of square-mean almost automorphic mild solutions. Some known results are generalized and improved.
1. Introduction
The concept of almost periodicity is of great importance in probability for investigating stochastic processes [1–3]. The basic results on the almost periodic functions and their applications to deterministic differential equations may refer to [4, 5] and references therein. The concept of almost automorphy introduced initially by Bochner [6] is an important generalization of the classical almost periodicity. Since then, there has been an intense interest in studying several extensions of this concept such as asymptotic almost automorphy, -almost automorphy, and Stepanov-like almost automorphy (see [5, 7–9] and references therein). Much of the motivation has come from mathematical physics, mathematical biology, and various fields of science and engineering [10–12].
Besides, it should be pointed out that noise or stochastic perturbation is unavoidable and omnipresent in nature as well as that in man-made systems. Therefore, we must import the stochastic effects into the investigation of differential systems. In fact, the existence of almost periodic solutions for stochastic differential systems has been thoroughly investigated (see [13–17] and reference therein) while the existence of almost automorphic solutions for stochastic version has been in growing state. More precisely, in [18], the concept of square-mean almost automorphic process was introduced and investigated. Particularly, such a concept was utilized to study the existence and stability of square-mean almost automorphic mild solutions for a class of stochastic differential equations of the form in a Hilbert space; Chang et al. [19] extended the results in [18] to nonautonomous stochastic differential equations in Hilbert spaces; in [20], the square-mean pseudo almost automorphic process and its application to (1.2) were investigated; in [21], existence and exponential stability of almost automorphic mild solutions were considered to a class of stochastic differential equations with finite delay of the form one can also see [22, 23] for the existence of square-mean almost automorphic mild solutions of stochastic differential equations.
In this paper, we consider a general setting; that is, we make extensive use of intermediate space techniques to investigate the existence of square-mean almost automorphic mild solutions to the class of abstract nonautonomous neutral stochastic evolution equations of the form where is a family of closed linear operators whose corresponding analytic semigroup is exponential dichotomy, , , are bounded operators, is a -Brownian motion defined on a probability space with a filtration , and , , and are jointly continuous functions to be specified later.
The rest of this paper is organized as follows. In Section 2 we present some basic notations and preliminary results. Section 3 is devoted to the study of existence of almost automorphic mild solutions for systems (1.3).
2. Preliminaries
For more details on this section, we refer to Da Prato et al. [24], Diagana [8], and Fu-Liu [18]. Throughout this paper, we assume that , are real separable Hilbert spaces and is supposed to be a filtered complete probability space. Denote by the Banach space of all -valued random variables such that endowed with the norm . If are Banach spaces, we denote by the Banach spaces of bounded linear operators from to equipped with natural operator norm; when , this is simply denoted by . Furthermore, denotes the space of all -Hilbert-Schmidt operators from to with the norm
For , is a family of closed linear operators (not necessarily densely defined) satisfying the so-called Acquistapace-Terreni conditions (ATCs for short; see Lemma 2.3). If is a linear operator on , then the symbols stand, respectively, for the domain, resolvent set, spectrum, kernel, and range of . We also set for all and for a projection .
Definition 2.1. A family of bounded linear operators on associated with is said to be an evolution family of operators if the following conditions hold: (i) for all , such that ; (ii), for ; (iii) is strongly continuous, for ; (iv) and
Definition 2.2 (see [8]). One says that an evolution family is exponential dichotomy (or hyperbolic) if there are projections , , being uniformly bounded and strongly continuous in and constants and such that (1); (2)the restriction of is invertible, and we set ; (3) and , for , .
If has an exponential dichotomy, then the operator family
is called Green's function corresponding to and . If for , then is exponentially stable.
The following lemma holds by [25].
Lemma 2.3. If satisfy the ATCs; that is, there exists a positive constant such that the operator , satisfying for , and constants , , with , then there exists a unique evolution family on .
Definition 2.4 (see [26]). A linear operator (not necessarily densely defined) is said to be sectorial if the following hold.
There exist constants , and such that
Let be a sectorial operator on and . Define the real interpolation space
it is a Banach space endowed with the norm . Given a family of linear operators , , for , we set with the corresponding norms.
The following estimates for the evolution family appeared in [8] are useful.
Lemma 2.5. For , and , there exist some constants , such that
Throughout the rest of this paper, we assume that the following conditions on and hold:()ATCs are satisfied and the evolution family generated by has an exponential dichotomy with constants and dichotomy projections for . Moreover, for each and the following holds: ()there exists with such that
for all , with uniform equivalent norms. And there exist constants , such that
Lemma 2.6 (see [8]). Under the above assumptions, there exist constants such that
We recall some basic definitions and results of square-mean almost automorphic processes (see [18, 19]).
Let be a Banach space and its -space.
Definition 2.7. A stochastic process is said to be stochastically continuous if
Definition 2.8 (see [18, 21]). A stochastically continuous stochastic process is said to be square-mean almost automorphic if for every sequence of real numbers there exists a subsequence such that
This is equivalent to that there exists a stochastic process such that, for each ,
Denote by the collection of all the square-mean almost automorphic processes . It is a Banach space equipped with the usual sup-norm
Lemma 2.9 (see [18]). If are square-mean almost automorphic processes, one has (i) is square-mean almost automorphic; (ii) is square-mean almost automorphic for every scalar ; (iii)there exists a positive constant such that .
Let , be Banach spaces, and , , their corresponding -spaces, respectively.
Definition 2.10 (compare with [18, 21]). A function , , which is jointly continuous, is said to be square-mean almost automorphic in for each , if, for every sequence of real numbers , there exists a subsequence such that for each and .
This is equivalent to that there exists a function such that, for each and ,
We need the following composition of square-mean almost automorphic processes.
Lemma 2.11. Suppose that is square-mean almost automorphic in , and assume that satisfies where is a concave nondecreasing function from to such that , and . Then for any square-mean almost automorphic process , the stochastic process given by is square-mean almost automorphic.
Proof. Since and are square-mean almost automorphic processes, for every sequence of real numbers , there exist a subsequence and some functions , such that, for each , , Let . Then we have by Jensen's inequality, it follows that noting that is concave and , we deduce that Similarly, we can prove that ; this completes the proof.
Lemma 2.12 (see [22]). Let and assume . Then .
The consideration is mainly based on the following fixed point theorem of Krasnoselskii (see [27]).
Lemma 2.13. Let be a closed, bounded, and convex subset of a Banach space . Let and be operators, defined on satisfying the conditions:(a) when ;(b)the operator is a contraction;(c)the operator is continuous and is contained in a compact set.
Then the equation has a solution in .
3. Existence of Square-Mean Almost Automorphic Mild Solutions
Firstly, we present the definition of mild solution for system (1.3).
Definition 3.1. An continuous stochastic function is called a mild solution of (1.3) provided that the function is integrable on , is integrable on for each , and satisfies the following stochastic integral equation:
In order to obtain our main results, we need the following assumptions. (), are bounded linear operators, and we set (). For any sequence of real numbers , there exists a subsequence such that, for each , one can find such that whenever , , , where is integrable. () Let . is square-mean almost automorphic in , and there exists a small such that for all and . (), are square-mean almost automorphic in , and, for each , , where is a concave nondecreasing function such that , for and . () For any , there exist a constant and nondecreasing continuous functions such that, for all and with ,
Remark 3.2. Functions such as and satisfy assumption (); in particular, we see that the Lipschitz condition is a special case of the proposed assumptions.
Throughout the rest of this paper, we denote by the sup-norm of the space . Let , be the operators defined by
Lemma 3.3. Under assumptions ()–(), the operators , , defined above map into itself.
Proof. Let . By Lemma 2.12, as , . And hence, by Lemma 2.11. In particular, . Let us show that . Indeed, since , for every sequence of real numbers , there exists a subsequence such that, for each , From , for any , one can find such that whenever , , . Thus, Using (3.13) and condition , one has For , we use Lemma 2.6 to get Combing this estimates with (3.12), one obtains which implies that . By a similar argument, we can show that .
Lemma 3.4. Under assumptions , and , the operators , , defined above map into itself.
Proof . Let . By Lemmas 2.11 and 2.12 it follows that . Particularly, . We now show that . Since , for every sequence of real numbers , there exists a subsequence such that, for each ,
For any , by making changes of variables we have
an elementary inequality shows that
Using Lemma 2.5, one has
a straightforward computation yields
where is a constant satisfying .
For , applying Proposition 4.4 in [4], we have
Combing this estimates with (3.18), we get
which implies that whenever . Similarly, we can verify that .
Lemma 3.5. Under assumptions and , the operators , , defined above map into itself.
Proof. Let . Using Lemmas 2.11 and 2.12, we get . Particularly, . We now show that . Since , for every sequence of real numbers , there exists a subsequence such that, for each ,
Note that the process for each is also a Brownian motion and has the same distribution as . For any , similar argument as above, we have
From (3.25), we immediately get
which implies that . Similarly, we can show that whenever .
Consider the nonlinear operators , , on defined by
for each . In view of Lemmas 3.3, 3.4, and 3.5, it follows that , , map into itself. In what follows, we will prove that , , satisfy all the conditions in Lemma 2.13.
Lemma 3.6. Under assumptions ()–(), the operator defined above is a contraction provided that
Proof. Let . By using condition and assumptions , we have Now, using Lemma 2.6 together with Hölder's inequality, we obtain Similarly, Thus, The proof is completed.
Lemma 3.7. Under assumptions and , the operator defined above is continuous; moreover, its image is contained in a compact set.
Proof. Let , for some . It is obvious that is a closed bounded convex subset of . We begin with the continuity of . Let be a sequence with ; that is, . Using the estimates in Lemma 2.5, we get
by the continuity of , , and Lebesgue’s dominated convergence theorem, it follows that
Similarly, it is easy to show that
Applying the isometry inequality, we obtain
by the continuity of , , and Lebesgue’s dominated convergence theorem yields
Similarly, it is easy to show that
Thus,
which implies that as .
Next, we show that is contained in a compact set. In fact, by the Ascoli-Arzela theorem, it suffices to show that maps into a equicontinuous family. Let be arbitrary and .
An analogue argument as Lemma 4.8 in [13], we have
For the first term on the right-hand side of (3.41), we have
and for the second term, we get
Combing these estimates with (3.41), it follows that there exists a positive constant such that
Similar computation can show that there exists a positive constant such that
As to , we have
For the first term on the right-hand side of (3.46), we have
using condition , Hölder's inequality together with isometry inequality yields
where is a positive constant depending on .
The second term is straightforward; we have
where .
Therefore,
Similarly, we can deduce that there exist some constants such that
Since
combing the evaluations above, we conclude that the right-hand side of (3.52) tends to zero independent of as . This completes the proof.
Theorem 3.8. Assume that assumptions ()–() are satisfied, and one further assumes that . Then the system (1.3) has a square-mean almost automorphic mild solution which can be expressed as .
Proof. Define an operator on by
From Lemma 3.3 to Lemma 3.5, it is easy to see that maps into itself. To complete the proof, it suffices to show that, for some closed bounded convex subset of , we have
Let be fixed. By it follows that there exist a positive constant and nondecreasing continuous functions such that, for all and with ,
Thus, for all , ,
where .
Let . A standard computation involving assumptions –, Lemma 2.5, and Hölder's inequality, we can deduce that
using and (3.56), we further derive that
Note that, for sufficiently small, we can choose such that
Let
It is easy to see that is a closed bounded convex subset of . Moreover, for all ,
Therefore, . By Lemmas 3.6 and 3.7 together with Krasnoselskii fixed point theorem we conclude that there exists a square-mean almost automorphic mild solution to (1.3). This completes the proof.
Now, we give another main result by Banach fixed point theorem. We require the following assumptions. (), are square-mean almost automorphic in and there exist some constants such that for each , ,
Theorem 3.9. Under assumptions ()–() and , (1.3) has a unique square-mean almost automorphic mild solution provided that
Proof . Let be the operator defined by (3.53). From Lemma 3.3 to Lemma 3.5, it is easy to see that maps into itself. To complete the proof, it suffices to show that is a contractive map and has a unique fixed point. To this end, let . By a similar argument as above we can deduce that which implies that Hence, by the Banach fixed point principle, has a unique fixed point which is obviously the square-mean almost automorphic mild solution to (1.3). The proof is completed.
Remark 3.10. The results of Theorem 3.9 can be applied to the existence of square-mean almost automorphic mild solutions to the example in [14].
Remark 3.11. If , is densely defined and the evolution family generated by is exponentially stable (that is, ), the existence of square-mean almost automorphic mild solutions has been studied in [23] by Banach fixed point theorem; if is the infinitesimal generator of an analytic semigroup of linear operators, the existence of square-mean almost automorphic mild solutions has been studied in [22] by Banach fixed point theorem. In other words, the results in [22, 23] have been generalized and improved.
Acknowledgment
The Project is sponsored by the NSFC (11171062), the Innovation Program of Shanghai Municipal Education Commission (12ZZ063), the Natural Science Foundation of Anhui Province (1208085MA11), and the NSF of Anhui Educational Committee (KJ2011A139).