Next Article in Journal
Topologies of Bihyperbolic Numbers
Next Article in Special Issue
Optimization of Robust LMI-Control Systems for Unstable Vertical Plasma Position in D-Shaped Tokamak
Previous Article in Journal
Study of Fractional Differential Equations Emerging in the Theory of Chemical Graphs: A Robust Approach
Previous Article in Special Issue
Approximate Controllability of Semilinear Stochastic Generalized Systems in Hilbert Spaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Controllability of Stochastic Delay Systems Driven by the Rosenblatt Process

1
Department of Mathematical Sciences, College of Sciences, Princess Nourah Bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
School of Mathematics, Harbin Institute of Technology, Harbin 150001, China
3
Department of Mathematics, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
*
Author to whom correspondence should be addressed.
Mathematics 2022, 10(22), 4223; https://doi.org/10.3390/math10224223
Submission received: 14 September 2022 / Revised: 13 October 2022 / Accepted: 27 October 2022 / Published: 11 November 2022
(This article belongs to the Special Issue Dynamics and Control Theory with Applications)

Abstract

:
In this work, we consider dynamical systems of linear and nonlinear stochastic delay-differential equations driven by the Rosenblatt process. With the aid of the delayed matrix functions of these systems, we derive the controllability results as an application. By using a delay Gramian matrix, we provide sufficient and necessary criteria for the controllability of linear stochastic delay systems. In addition, by employing Krasnoselskii’s fixed point theorem, we present some necessary criteria for the controllability of nonlinear stochastic delay systems. Our results improve and extend some existing ones. Finally, an example is given to illustrate the main results.

1. Introduction

Due to its effective modeling in numerous fields of science and engineering, including economics, biology, physics, medicine, finance, fluid dynamics, etc., stochastic delay-differential equations and their applications have received a great deal of attention; see, for instance, [1,2,3,4,5,6,7,8,9,10]. The concept of the controllability of systems is one of the most fundamental and important concepts in contemporary control theory, which involves figuring out the control parameters that directly a control system’s solutions from their initial state to their final state using the set of permissible controls, where initial and final states may vary across the entire space. The representation of time-delay system solutions has recently received attention. The seminal study in [11,12], in particular, yielded several novel results in the representation of the solutions, stability, and controllability of time-delay systems; see, for example, [13,14,15,16,17,18,19,20,21] and the references therein.
The Hermite process of order one is known as the fractional Brownian motion, while the Hermite process of order two is known as the Rosenblatt process. Rosenblatt first proposed the following distribution for x 0
Z U x = D U R 2 0 x ϑ t 1 + 1 + U / 2 ϑ t 2 + 1 + U / 2 d ϑ d J t 1 d J t 2 ,
where U 0 , 1 2 and J t , t U are a standard Brownian motion. The process of Z U 1 is known as the ‘1 non-Gaussian limiting distribution’ (Rosenblatt distribution); for more details see [22]. The Rosenblatt process is a non-Gaussian process with many interesting properties, such as the stationarity of the increments, long-range dependence, and self-similarity. Therefore, it is important to study a new class of stochastic differential equations driven by the Rosenblatt process. Shen and Ren [23] investigated the existence and uniqueness of the mild solution for neutral stochastic partial differential equations with finite delay driven by the Rosenblatt process in a real separable Hilbert space. Maejima and Tudor [24] presented a technique for constructing self-similar processes in the second Wiener chaos using limit theorems. Shen et al. [25] used fixed point theory to examine controllability and stability in functional nonlinear neutral stochastic systems with delay driven by the Rosenblatt process; furthermore, we refer the reader to [26,27] for further details on the Rosenblatt process.
Elshenhab and Wang [14] have established a novel formula to solve the following linear delay-differential systems
z x + Ξ z x ω = g x , x 0 , z x Π x , z x Π x , ω x 0 ,
of the form
z x = H ω Ξ x ω Π 0 + M ω Ξ x ω Π 0 Ξ ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ + 0 x M ω Ξ x ω ϑ g ϑ d ϑ ,
where H ω Ξ x and M ω Ξ x are the delayed matrix functions defined by
H ω Ξ x : = Θ , < x < ω ,   I , ω x < 0 ,     I Ξ x 2 2 ! , 0 x < ω ,     I Ξ x 2 2 ! + Ξ 2 x ω 4 4 ! + + ( 1 ) ς Ξ ς x ς 1 ω 2 ς 2 ς ! ,     ς 1 ω x < ς ω ,
and
M ω Ξ x : = Θ , < x < ω ,   I x + ω , ω x < 0 ,     I x + ω Ξ x 3 3 ! , 0 x < ω ,     I x + ω Ξ x 3 3 ! + Ξ 2 x ω 5 5 ! + + ( 1 ) ς Ξ ς x ς 1 ω 2 ς + 1 2 ς + 1 ! , ς 1 ω x < ς ω ,
respectively, where ς = 0 , 1 , 2 , , and I and Θ are the n × n identity and null matrix, respectively.
Based on [14], they [17] used the solutions (2) of (1) to obtain the controllability of the linear delay control systems
z x + Ξ z x ω = Λ u ( x ) , x : = 0 , x 1 , z x Π x , z x Π x , ω x 0 ,
and the controllability of the corresponding nonlinear delay control systems
z x + Ξ z x ω = g x , z x + Λ u ( x ) , x , z x Π x , z x Π x , ω x 0 ,
where ω > 0 is a delay, Ξ R n × n and Λ R n × m are any matrices, x 1 > n 1 ω , z x R n , Π C ω , 0 , R n , and g C × R n , R n are given functions, and u x R m shows control vector.
In light of the preceding work, we investigate the controllability of stochastic linear delay-differential systems driven by the Rosenblatt process,
z x + Ξ z x ω = Λ u ( x ) + Δ ¯ x d Z H x , x : = 0 , x 1 , z x Π x , z x Π x , ω x 0 ,
in addition to the controllability of the corresponding stochastic nonlinear delay differential systems driven by the Rosenblatt process
z x + Ξ z x ω = Λ u ( x ) + Δ x , z x d Z H x , x , z x Π x , z x Π x , ω x 0 ,
where the state vector z x R n , Ξ R n × n , and Λ R n × m are any matrices, Π C ω , 0 , R n and u x R m show control vectors, and Δ ¯ C , T R n is the Thorin class, symbolized by T R n , and is the smallest distribution class on R n that comprises all Gamma distributions and is closed under convolution and weak convergence. Let z · take value in the separable Hilbert space R n with inner product · , · and norm · . Z H x is a Rosenblatt process with parameter H 1 2 , 1 on an another real separable Hilbert space K , · K , · , · K . Moreover, assume Δ C × R n , L 2 0 , where L 2 0 = L 2 Q 1 2 K , R n .
The following is how the rest of this paper is structured: In Section 2, we provide some introductions, fundamental notations, and definitions, as well as some relevant lemmas. In Section 3, using a delay Gramian matrix, we obtain the sufficient and necessary conditions for the controllability of (7). In Section 4, by applying Krasnoselskii’s fixed point theorem, we establish sufficient conditions for the controllability of (8). Finally, to illustrate the theoretical findings, we provide numerical examples.

2. Preliminaries

Throughout the paper, let Ω , F , P be the complete probability space with probability measure P on Ω with a filtration F x | x generated by Z H s | s 0 , x . Let X and Y be two Banach spaces and L b X , Y be the space of bounded linear operators from X to Y and Q L b X , X represent a non-negative self-adjoint trace class operator on X . Let L 2 0 = L 2 Q 1 2 X , Y be the space of all Q-Hilbert–Schmidt operators from Q 1 2 X into Y , equipped with the norm
φ L 2 0 2 = φ Q 1 2 2 = Tr φ Q φ T .
Now, for some 1 < e < , let L e Ω , F x 1 , R n be the Hilbert space of all F x 1 -measurable eth-integrable variables with values in R n with norm z L e e = E z x e , where the expectation E is defined by E z = Ω z d P . Let L F e , R n be the Banach space of all functions g : R n that are Bochner integrable normed by g L F e , R n - and F x 1 -measurable processes with values in R n . Let F : = C ω , 0 , L e Ω , F x 1 , P , R n be the Banach space of all eth-integrable and F x 1 -adapted processes ϕ endowed with the norm ϕ C = sup x ω , 0 E ϕ x e 1 / e . Additionally, we denote C , L e Ω , F x 1 , P , R n as the Banach space of continuous function from L e Ω , F x 1 , P , R n equipped with the norm z C = sup x E z x e 1 / e for a norm · on R n . We describe a space
C 1 , L e Ω , F x 1 , P , R n = z C , L e Ω , F x 1 , P , R n : z C , L e Ω , F x 1 , P , R n ,
and we use the matrix norm (column sum)
Ξ = max i = 1 n a i 1 , i = 1 n a i 2 , , i = 1 n a i n ,
where Ξ : R n R n . Furthermore, we let
Π C = sup s ω , 0 E Π s e 1 / e and Π C = sup s ω , 0 E Π s e 1 / e .
The Wiener–Ito multiple integral of order k with respect to the standard Wiener process G ρ ρ R is defined by
Z H k x = c H , k R k 0 x j = 1 k ϑ ρ j + 1 2 + 1 H k d ϑ d G ρ 1 d G ρ k ,
where c H , k is a normalizing constant such that E Z H k 1 2 = 1 and ρ + = max ρ , 0 . The process Z H k x x 0 is called the Hermite process. If k = 1 , the Hermite process given by (9) is the fBm with Hurst parameter H 1 2 , 1 ; furthermore, the process is not Gaussian for k = 2 . Moreover, for k = 2 , the Hermite process given by (9) is called the Rosenblatt process.
We provide some fundamental concepts and lemmas used in this work.
 Lemma 1 
([28]). If σ : L 2 0 satisfies
0 x 1 σ ϑ L 2 0 2 d ϑ < ,
then, for a, b with b > a , we have
E 0 x σ ϑ d Z H ϑ 2 2 H x 2 H 1 0 x σ ϑ L 2 0 2 d ϑ .
 Definition 1 
([29]). If there is a control function u L F e , R m such that (7) or (8) have a solution z : ω , x 1 R n with z x = Π x , z x = Π x satisfies z x 1 = z 1 for all x ω , 0 , z 1 R n , then, the systems (7) or (8) are controllable on = 0 , x 1 .
 Definition 2 
([30]). The two-parameter Mittag-Leffler function is provided by
E α , γ x = r = 0 x r Γ α r + γ , α , γ > 0 , x C .
In the case of γ = 1 , then
E α , 1 x = E α x = r = 0 x r Γ α r + 1 , α > 0 .
 Lemma 2 
([17] (Lemmas 1 and 2)). For any x m 1 ω , m ω , m = 1 , 2 , . . . , we have
H ω Ξ x E 2 Ξ x 2 ,
and
M ω Ξ x x + ω E 2 , 2 Ξ x + ω 2 .
 Lemma 3 
(Krasnoselskii’s fixed point theorem [31]). Let M be a closed, bounded, and convex subset of a real Banach space K , and let J 1 and J 2 be operators on M satisfying the following conditions:
 (i) 
J 1 x + J 2 z M for x, z M ;
 (ii) 
J 1 is compact and continuous;
 (iii) 
J 2 is a contraction mapping.
Then, there exists m M such that m = J 1 m + J 2 m .
We define the operator Q x 1 L b L F e , R m , L e Ω , F x 1 , R n as
Q x 1 u = 0 x 1 M ω Ξ x 1 ω ϑ Λ u ϑ d ϑ ,
and its adjoint operator Q x 1 T L b L e Ω , F x 1 , R n , L F e , R m is defined as
Q x 1 T u = Λ T M ω Ξ T x 1 ω x E t | F x .
Consider the linear controllability operator
Γ ω x 1 · = Q x 1 Q x 1 T · = 0 x 1 M ω Ξ x 1 ω ϑ Λ Λ T M ω Ξ T x 1 ω ϑ E · | F ϑ d ϑ ,
and the delayed Gramian matrix W ω M 0 , x 1 L b R n , R n defined by
W ω M 0 , x 1 = 0 x 1 M ω Ξ x 1 ω ϑ Λ Λ T M ω Ξ T x 1 ω ϑ d ϑ .
Here, T denotes the transpose.

3. Controllability of Linear Stochastic Delay Systems

In this section, we derive the controllability results of (7) using the delayed Gramian matrix W ω M 0 , x 1 defined by (10).
 Theorem 1. 
The stochastic system (7) is controllable if and only if W ω M 0 , x 1 is positive definite.
 Proof. Sufficiency. 
Assume that W ω M 0 , x 1 is positive definite; then, it is invertible. Consequently, for any finite terminal conditions z 1 , z 1 R n , we can determine the related control input u x as
u x = Λ T M ω Ξ T x 1 ω x W ω M 1 0 , x 1 β ,
where
β = z 1 H ω Ξ x 1 ω Π 0 M ω Ξ x 1 ω Π 0 + Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ .
Applying the Formula (2), the solution of (7) can be expressed as
z x = H ω Ξ x ω Π 0 + M ω Ξ x ω Π 0 Ξ ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ + 0 x M ω Ξ x ω ϑ Λ u ( ϑ ) d ϑ + 0 x M ω Ξ x ω ϑ Δ ¯ ϑ d Z H ϑ .
From (13), the solution z x 1 of (7) can be expressed as
z x 1 = H ω Ξ x 1 ω Π 0 + M ω Ξ x 1 ω Π 0 Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Λ u ( ϑ ) d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ .
Substituting (11) into (14), we obtain
z x 1 = H ω Ξ x 1 ω Π 0 + M ω Ξ x 1 ω Π 0 Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Λ Λ T M ω Ξ T x 1 ω ϑ d ϑ × W ω M 1 0 , x 1 β + 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ .
From (10), (12), and (15), we obtain
z x 1 = H ω Ξ x 1 ω Π 0 + M ω Ξ x 1 ω Π 0 Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ + β + 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ . = z 1 .
We can see from (3), (4), and (13) that the boundary conditions z x Π x , z x Π x , ω x 0 are fulfilled. Therefore, (7) is controllable.
Necessity. Let (7) be controllable. Assume, for the sake of a contradiction, that W ω M 0 , x 1 is not positive definite and there exists at least a nonzero vector ρ R n such that ρ T W ω M 0 , x 1 ρ = 0 , which leads to
0 = ρ T W ω M 0 , x 1 ρ = 0 x 1 ρ T M ω Ξ x 1 ω ϑ Λ Λ T M ω Ξ T x 1 ω ϑ ρ d ϑ = 0 x 1 ρ T M ω Ξ x 1 ω ϑ Λ ρ T M ω Ξ x 1 ω ϑ Λ T d ϑ = 0 x 1 ρ T M ω Ξ x 1 ω ϑ Λ d ϑ .
Hence,
ρ T M ω Ξ x 1 ω ϑ Λ = 0 , , 0 : = 0 T , for all ϑ ,
where 0 denotes the n dimensional zero vector. From Definition 1, there is a control function u 1 x that steers the initial state to z 1 = 0 at x = x 1 . Then,
z x 1 = H ω Ξ x 1 ω Π 0 + M ω Ξ x 1 ω Π 0 Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Λ u 1 ( ϑ ) d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ = 0 .
Similarly, there is a control function u 2 x that steers the initial state to z 1 = ρ at x = x 1 . Thus,
z x 1 = H ω Ξ x 1 ω Π 0 + M ω Ξ x 1 ω Π 0 Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Λ u 2 ( ϑ ) d ϑ + 0 x 1 M ω Ξ x 1 ω ϑ Δ ¯ ϑ d Z H ϑ = ρ .
Combining (17) with (18), we have
ρ = 0 x 1 M ω Ξ x 1 ω ϑ Λ u 2 ϑ u 1 ϑ d ϑ
Multiplying (19) by ρ T and using (16), we have ρ T ρ = 0 . This is a contradiction to ρ 0 . Thus, W ω M 0 , x 1 is positive definite. This completes the proof. □
 Corollary 1. 
Theorem 1 is satisfied in the case of Ξ = Ξ 2 in (7).
 Corollary 2. 
In the case of Ξ being a nonsingular n × n matrix, and Ξ = Ξ 2 in (7), the system (7) is controllable if and only if W ω 0 , x 1 is nonsingular, where W ω 0 , x 1 is defined as
W ω 0 , x 1 = Ξ 1 0 x 1 sin ω Ξ x 1 ω ϑ Λ Λ T sin ω Ξ T x 1 ω ϑ d ϑ ,
and sin ω Ξ x and cos ω Ξ x are known as the delayed matrices of the sine and cosine types, respectively, which are given in [12].
Proof. 
When Ξ = Ξ 2 and using the definition of H ω Ξ x and M ω Ξ x , we have
H ω Ξ 2 x = cos ω Ξ x , M ω Ξ 2 x = Ξ 1 sin ω Ξ x ,
which implies that
W ω M 0 , x 1 = 0 x 1 M ω Ξ 2 x 1 ω ϑ Λ Λ T M ω Ξ 2 T x 1 ω ϑ d ϑ = 0 x 1 Ξ 1 sin ω Ξ x 1 ω ϑ Λ Λ T sin ω Ξ T x 1 ω ϑ Ξ 1 T d ϑ = Ξ 1 0 x 1 sin ω Ξ x 1 ω ϑ Λ Λ T sin ω Ξ T x 1 ω ϑ d ϑ Ξ T 1 = W ω 0 , x 1 Ξ T 1 .
Hence,
W ω 0 , x 1 = W ω M 0 , x 1 Ξ T .
From the conclusion of Theorem 1, we determine that W ω M 0 , x 1 is invertible. Thus, from (20), we find that W ω 0 , x 1 is also invertible. This completes the proof. □

4. Controllability of Nonlinear Stochastic Delay Systems

In this section, we present sufficient conditions of the controllability of (8).
The following hypotheses are made:
 (S1) 
The function Δ : × R n L 2 0 is continuous, and there exists a constant L Δ L q , R + and q > 1 such that
E Δ x , z 1 Δ x , z 2 L 2 0 e L Δ x z 1 z 2 e , for all x , z 1 , z 2 R n .
Let e 2 , and sup x E Δ x , 0 L 2 0 e = N Δ < .
 (S2) 
The linear stochastic delay system (7) is controllable on .
Under the assumption of ( S 2 ) , for some η > 0 , we have E Γ ω x 1 z , z η E z e for all z L e Ω , F x 1 , R n (see [32] [Lemma 2]). Furthermore, Γ ω x 1 1 e 1 / η : = N 1 (see [33]) and set N : = max W ω M ϑ , x 1 e : ϑ .
 Theorem 2. 
Let ( S 1 ) and ( S 2 ) be satisfied. Then, the nonlinear stochastic system (8) is controllable on Υ if there exists a constant τ e > 0 such that
N 2 1 + 5 e 1 N N 1 < 1 ,
where
N 2 : = 5 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 1 q e p + 1 1 p E 2 , 2 Ξ x 1 2 e L Δ L q , R + ,
and 1 p + 1 q = 1 , p, q > 1 .
Proof. 
We consider the following set before beginning to prove this theorem
B = z F : z F e = sup x ω , x 1 E z x e ,
for each postive number . Let x 0 , x 1 . Applying Formula (2), the solution of (8) can be expressed as
z x = H ω Ξ x ω Π 0 + M ω Ξ x ω Π 0 Ξ ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ + 0 x M ω Ξ x ω ϑ Λ u z ( ϑ ) d ϑ + 0 x M ω Ξ x ω ϑ Δ ϑ , z ϑ d Z H ϑ ,
and its control function u z is defined as
u z x = Λ T M ω Ξ T x 1 ω x × E Γ ω x 1 1 z 1 H ω Ξ x 1 ω Π 0 M ω Ξ x 1 ω Π 0 + Ξ ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ 0 x 1 M ω Ξ x 1 ω ϑ Δ ϑ , z ϑ d Z H ϑ | F x
for x . Additionally, we define the following operators L 1 and L 2 on B of the form
L 1 z x = H ω Ξ x ω Π 0 + M ω Ξ x ω Π 0 Ξ ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ + 0 x M ω Ξ x ω ϑ Λ u z ϑ d ϑ ,
L 2 z x = 0 x M ω Ξ x ω ϑ Δ ϑ , z ϑ d Z H ϑ .
At this point, it is evident that B is a closed, bounded, and convex set of F . Therefore, there are three essential steps to our proof:
Step 1. We prove that there exists a > 0 such that L 1 z + L 2 ρ B for all z, ρ B .
Using (23) and (24), we obtain
L 1 z + L 2 ρ F e = sup x ω , x 1 E L 1 z + L 2 ρ x e 5 e 1 H ω Ξ x ω e E Π 0 e + M ω Ξ x ω e E Π 0 e + Ξ e E ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ e + E 0 x M ω Ξ x ω ϑ Λ u z ϑ d ϑ e + E 0 x M ω Ξ x ω ϑ Δ ϑ , ρ ϑ d Z H ϑ e = n = 1 5 I n ,
for each x and z, ρ B . From Lemma 2, we have
I 1 = 5 e 1 H ω Ξ x ω e E Π 0 e 5 e 1 E 2 Ξ x ω 2 e E Π C e ,
I 2 = 5 e 1 M ω Ξ x ω e E Π 0 e 5 e 1 x E 2 , 2 Ξ x 2 e E Π C e ,
I 3 = 5 e 1 Ξ e E ω 0 M ω Ξ x 2 ω ϑ Π ϑ d ϑ e 5 e 1 Ξ e ω e 1 E Π C e ω 0 M ω Ξ x 2 ω ϑ e d ϑ 5 e 1 Ξ e ω e x E 2 , 2 Ξ x 2 e E Π C e ,
I 4 = 5 e 1 E 0 x M ω Ξ x ω ϑ Δ ϑ , ρ ϑ d Z H ϑ e = 5 e 1 E 0 x M ω Ξ x ω ϑ Δ ϑ , ρ ϑ d Z H ϑ 2 e / 2 ,
and by employing Lemma 1, the Kahane–khintchine inequality, and Hölder’s inequality, there exists a constant τ e such that
I 4 5 e 1 τ e E 0 x M ω Ξ x ω ϑ Δ ϑ , ρ ϑ d Z H ϑ 2 e / 2 5 e 1 τ e 2 H x 2 H 1 0 x E M ω Ξ x ω ϑ Δ ϑ , ρ ϑ L 2 0 2 d ϑ e / 2 5 e 1 τ e 2 H x 2 H 1 e / 2 0 x E M ω Ξ x ω ϑ Δ ϑ , ρ ϑ L 2 0 2 d ϑ e / 2 5 e 1 τ e 2 H x 2 H 1 e / 2 × 0 x E M ω Ξ x ω ϑ Δ ϑ , ρ ϑ L 2 0 2 e / 2 d ϑ 2 / e 0 x d ϑ e 2 e e / 2 5 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 0 x E M ω Ξ x ω ϑ Δ ϑ , ρ ϑ L 2 0 e d ϑ ,
and by employing Lemma 2 and ( S 1 ) , we obtain
I 4 5 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e E Δ ϑ , ρ ϑ L 2 0 e d ϑ 5 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 × 2 e 1 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e E Δ ϑ , ρ ϑ Δ ϑ , 0 L 2 0 e d ϑ + 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e E Δ ϑ , 0 L 2 0 e d ϑ 5 e 1 2 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e L Δ ϑ ρ ϑ e d ϑ + N Δ 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e d ϑ 5 e 1 2 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 { ρ F e 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e L Δ ϑ d ϑ + x 1 e + 1 N Δ e + 1 ( E 2 , 2 ( Ξ x 1 2 ) ) e }
Moreover, from ( S 1 ) and the Hölder inequality, we have
0 x x ϑ E 2 , 2 Ξ x ϑ 2 e L Δ ϑ d ϑ 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e p d ϑ 1 p 0 x L Δ q ϑ d ϑ 1 q E 2 , 2 Ξ x 1 2 e 0 x x ϑ e p d ϑ 1 p 0 x L Δ q ϑ d ϑ 1 q x 1 e + 1 p e p + 1 1 p E 2 , 2 Ξ x 1 2 e L Δ L q , R + .
Substituting (27) into (26), we obtain
I 4 5 e 1 2 e 1 τ e ( 2 H ) e / 2 x 1 e H 1 × x 1 e + 1 p e p + 1 1 p E 2 , 2 Ξ x 1 2 e L Δ L q , R + + x 1 e + 1 N Δ e + 1 E 2 , 2 Ξ x 1 2 e = 2 e 1 N 2 + 10 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 N Δ e + 1 E 2 , 2 Ξ x 1 2 e .
Furthermore, using (22), we obtain
I 5 = 5 e 1 E 0 x M ω Ξ x ω ϑ Λ u z ϑ d ϑ e 5 e 1 W ω M 0 , x 1 e × Γ ω x 1 1 e 5 e 1 E z 1 e + H ω Ξ x 1 ω e E Π 0 e + M ω Ξ x 1 ω e E Π 0 e + Ξ e E ω 0 M ω Ξ x 1 2 ω ϑ Π ϑ d ϑ e + E 0 x 1 M ω Ξ x ω ϑ Δ ϑ , z ϑ d Z H ϑ e 5 2 e 1 N N 1 E z 1 e + θ x 1 + 2 5 e 1 N 2 ,
where
θ x : = E 2 Ξ x ω 2 e E Π C e + x E 2 , 2 Ξ x 2 e E Π C e + Ξ e ω e x E 2 , 2 Ξ x 2 e E Π C e + 2 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 N Δ e + 1 E 2 , 2 Ξ x 1 2 e .
From I 1 to I 5 , (25) becomes
L 1 z + L 2 ρ F e 5 e 1 E 2 Ξ x ω 2 e E Π C e + x E 2 , 2 Ξ x 2 e E Π C e + Ξ e ω e E Π C e x E 2 , 2 Ξ x 2 e + 2 5 e 1 N 2 + 2 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 N Δ e + 1 E 2 , 2 Ξ x 1 2 e + 5 e 1 N N 1 E z 1 e + θ x 1 + 2 5 e 1 N 2 5 e 1 θ x 1 1 + 5 e 1 N N 1 + 5 e 1 N N 1 E z 1 e + 2 5 e 1 N 2 1 + 5 e 1 N N 1 .
Thus, some are sufficiently large, and from (21), we have L 1 z + L 2 ρ B .
Step 2. We prove L 1 : B F is a contraction.
Using (22), we obtain
E L 1 z x L 1 ρ x e = E 0 x M ω Ξ x ω ϑ Λ u z ϑ u ρ ϑ d ϑ e W ω M 0 , x 1 e Γ ω x 1 1 e × E 0 x 1 M ω Ξ x ω ϑ Δ ϑ , ρ ϑ Δ ϑ , z ϑ d Z H ϑ e τ e N N 1 ( 2 H ) e / 2 x 1 e H 1 × 0 x E M ω Ξ x ω ϑ Δ ϑ , ρ ϑ Δ ϑ , z ϑ L 2 0 e d ϑ τ e N N 1 ( 2 H ) e / 2 x 1 e H 1 E z ρ F e 0 x x ϑ E 2 , 2 Ξ e x ϑ 2 e L Δ ϑ d ϑ τ e N N 1 ( 2 H ) e / 2 x 1 e H + 1 1 q e p + 1 1 p E 2 , 2 Ξ x 1 2 e L Δ L q , R + E z ρ F e N N 1 N 2 5 e 1 E z ρ F e μ E z ρ F e ,
for each x and z, ρ B , where μ : = N N 1 N 2 / 5 e 1 . We may deduce from (21), noting μ < 1 , that L 1 is a contraction mapping.
Step 3. We prove L 2 : B F is a continuous compact operator.
Firstly, we show that L 2 is continuous. Let z n be a sequence such that z n z as n in B . Hence, using (24) and Lebesgue’s dominated convergence theorem, we obtain
E L 2 z n x L 2 z x e τ e ( 2 H ) e / 2 x 1 e H 1 0 x M ω Ξ x ω ϑ e E Δ ϑ , z n ϑ Δ ϑ , z ϑ L 2 0 e d ϑ τ e ( 2 H ) e / 2 x 1 e H 1 0 x x ϑ E 2 , 2 Ξ x ϑ 2 e L Δ ϑ E z n ϑ z ϑ e d ϑ 0 , as n ,
for each x . Thus L 2 : B F is continuous.
After that, we prove that L 2 is uniformly bounded on B . For each x , z B , we obtain
L 2 z F e = sup x E L 2 z x e sup x E 0 x M ω Ξ x ω ϑ Δ ϑ , z ϑ d Z H ϑ e 2 5 e 1 N 2 + 2 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 N Δ e + 1 E 2 , 2 Ξ x 1 2 e ,
which leads to L 2 being uniformly bounded on B .
It remains to be shown that L 2 is equicontinuous. For each x 2 , x 3 , 0 < x 2 < x 3 x 1 and z B , using (24), we obtain
L 2 z x 3 L 2 z x 2 = 0 x 3 M ω Ξ x 3 ω ϑ Δ ϑ , z ϑ d Z H ϑ 0 x 2 M ω Ξ x 2 ω ϑ Δ ϑ , z ϑ d Z H ϑ = Ψ 1 + Ψ 2 ,
where
Ψ 1 = x 2 x 3 M ω Ξ x 3 ω ϑ Δ ϑ , z ϑ d Z H ϑ ,
and
Ψ 2 = 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ Δ ϑ , z ϑ d Z H ϑ .
Thus,
E L 2 z x 3 L 2 z x 2 e = E Ψ 1 + Ψ 2 e 2 e 1 E Ψ 1 e + E Ψ 2 e .
Now, we can check Ψ i 0 as x 2 x 3 , i = 1 , 2. For Ψ 1 , we obtain
E Ψ 1 e = E x 2 x 3 M ω Ξ x 3 ω ϑ Δ ϑ , z ϑ d Z H ϑ e τ e ( 2 H ) e / 2 x 3 x 2 e H 1 x 2 x 3 E M ω Ξ x ω ϑ Δ ϑ , z ϑ L 2 0 e d ϑ 2 e 1 τ e ( 2 H ) e / 2 x 3 x 2 e H 1 × z F e x 2 x 3 x ϑ E 2 , 2 Ξ x ϑ 2 e L Δ ϑ d ϑ + x 3 x 2 e + 1 N Δ e + 1 E 2 , 2 Ξ x 3 2 e 0 , as x 2 x 3 .
For Ψ 2 , we obtain
E Ψ 2 e = E 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ Δ ϑ , z ϑ d Z H ϑ e τ e ( 2 H ) e / 2 x 2 e H 1 × 0 x 2 E M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ Δ ϑ , z ϑ L 2 0 e d ϑ 2 e 1 τ e ( 2 H ) e / 2 x 2 e H 1 × 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ e L Δ ϑ d ϑ + N Δ 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ e d ϑ 2 e 1 τ e ( 2 H ) e / 2 x 2 e H 1 × L Δ L q , R + × 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ e p 1 / p d ϑ + N Δ 0 x 2 M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ e d ϑ
From (4), we know that M ω Ξ x is uniformly continuous for x . Hence,
M ω Ξ x 3 ω ϑ M ω Ξ x 2 ω ϑ 0 , as x 2 x 3 .
Therefore, we have Ψ i 0 as x 2 x 3 , i = 1 , 2, which implies that, using (28),
E L 2 z x 3 L 2 z x 2 e 0 , as x 2 x 3 ,
for all z B . As a result, L 2 is compact on B by applying the Arzelà–Ascoli theorem. Thus, L 1 + L 2 has a fixed point z on B via Krasnoselskii’s fixed point theorem (Lemma 3). Moreover, z is also a solution of (8), and L 1 z + L 2 z x 1 = z 1 . This indicates that u z steers the system (8) from z 0 to z 1 in finite time x 1 , implying that (8) is controllable on . This completes the proof. □
 Corollary 3. 
Theorem 2 is satisfied in the case of Ξ = Ξ 2 in (8).
 Corollary 4. 
When Ξ is a nonsingular n × n matrix and Ξ = Ξ 2 in (8). Theorem 2 coincides with Theorem 3.2 in [16].
Proof. 
Since M ω Ξ 2 x = Ξ 1 sin ω Ξ x . From ( S 1 ) and the Hölder inequality, we have
0 x M ω Ξ 2 x ω ϑ L Δ ϑ d ϑ = Ξ 1 0 x Ξ M ω Ξ 2 x ω ϑ L Δ ϑ d ϑ Ξ 1 0 x sin ω Ξ x ω ϑ L Δ ϑ d ϑ Ξ 1 0 x sinh Ξ x ϑ L Δ ϑ d ϑ Ξ 1 0 x sinh Ξ x ϑ p d ϑ 1 p 0 x L Δ q ϑ d ϑ 1 q = Ξ 1 0 x exp Ξ p x ϑ 2 p d ϑ 1 p 0 x L Δ q ϑ d ϑ 1 q = Ξ 1 1 2 p Ξ p exp Ξ p x 1 1 p L Δ L q , R + .
and
0 x M ω Ξ 2 x ω ϑ Δ ϑ , 0 d ϑ = Ξ 1 0 x sin ω Ξ x ω ϑ Δ ϑ , 0 d ϑ N Δ Ξ 1 0 x sinh Ξ x ϑ d ϑ = N Δ Ξ 1 Ξ cosh Ξ x 1 .
Similar to how the Theorem 2 was proved at Ξ = Ξ 2 , and as a result of (29) and (30), Theorem 3.2 in [16] yields the same result. This completes the proof. □
 Remark 1. 
We note that Theorem 1 extends Theorem 3.1 in [16] by choosing matrix Ξ, which is an arbitrary, not necessarily squared, matrix, and Corollaries 2 and 4 coincide with Theorems 3.1 and 3.2 in [16]. Therefore, our results in Corollaries 1–4 extend and improve Theorems 3.1 and 3.2 in [16] by removing the condition that Ξ is a nonsingular matrix.

5. An Example

Consider the following linear stochastic delay system:
z x + Ξ z x 0.5 = Λ u ( x ) + Δ ¯ x d Z H x , for x : = 0 , 1 , z x Π x , z x Π x for 0.5 x 0 ,
where
Ξ = 0 1 0 0 , Λ = 1 2 , Δ ¯ x = x e x 4 x e x 4 ,
and
Π x = 2 x x , Π x = 2 1 .
Generating the corresponding delay Gramian matrix of (31) using (10), we have
W 0.5 M 0 , 1 = 0 1 M 0.5 Ξ 0.5 ϑ Λ Λ T M 0.5 Ξ T 0.5 ϑ d ϑ = : O 1 + O 2 ,
where
O 1 = 0 0.5 M 0.5 Ξ 0.5 ϑ Λ Λ T M 0.5 Ξ T 0.5 ϑ d ϑ ,
for 0.5 ϑ 0 , 0.5 ,
O 2 = 0.5 1 M 0.5 Ξ 0.5 ϑ Λ Λ T M 0.5 Ξ T 0.5 ϑ d ϑ ,
for 0.5 ϑ 0.5 , 0 , where
H 0.5 Ξ x : = Θ , < x < 0.5 ,   I , 0.5 x < 0 ,     I Ξ x 2 2 , 0 x < 0.5 , I Ξ x 2 2 + Ξ 2 x 0.5 4 4 ! , 0.5 x < 1 ,
and
M 0.5 Ξ x : =     Θ , < x < 0.5 ,   I x + 0.5 , 0.5 x < 0 ,     I x + 0.5 Ξ x 3 3 ! , 0 x < 0.5 , I x + 0.5 Ξ x 3 3 ! + Ξ 2 x 0.5 5 5 ! , 0.5 x < 1 ,
Next, we can calculate that
O 1 = 0.28242 0.57396 0.57396 1.1667 , O 2 = 4.1667 × 10 2 8.3333 × 10 2 8.3333 × 10 2 0.16667 .
Then, we obtain
W 0.5 M 0 , 1 = O 1 + O 2 = 0.32409 0.65729 0.65729 1.3334 ,
and
W 0.5 M 1 0 , 1 = 11962.865 5897.01 5897.01 2907.638 .
Therefore, we see that W 0.5 M 0 , 1 is positive definite. Therefore, by Theorem 1, the system (31) is controllable on 0 , 1 , which implies that the assumption ( S 2 ) is satisfied.
Furthermore, consider the corresponding nonlinear stochastic delay system of (31) as follows:
z x + Ξ z x 0.5 = Λ u ( x ) + Δ x , z x d Z H x , for x : = 0 , 1 , z x Π x , z x Π x for 0.5 x 0 ,
where
Δ x , z x = x e x 4 z 1 x x e x 4 z 2 x .
Next, by selecting e = p = q = 2 , we obtain
E Δ x , z Δ x , ρ L 2 0 2 = x e x 4 2 z 1 x ρ 1 x 2 + z 2 x ρ 2 x 2 = x e 2 x 16 z ρ L 2 0 2 .
for all x , z x , and ρ x R 2 . We set L Δ x = x exp ( 2 x ) / 16 such that L Δ L 2 , R + in ( S 1 ) , so we have
L Δ L 2 , R + = 0 1 ϑ exp ( 2 ϑ ) 16 2 d ϑ 1 2 = 0.00964 .
Then, by choosing τ e = 0.23 and H = 0.75 , we obtain
N 2 : = 5 e 1 τ e ( 2 H ) e / 2 x 1 e H + 1 1 q e p + 1 1 p E 2 , 2 Ξ x 1 2 e L Δ L q , R + = 0.01 .
Furthermore, we have
E W 0.5 M 0 , 1 z , z = 0.32409 z 1 2 0.65729 z 2 2 0.65729 z 1 2 1.3334 z 2 2 η E z 2 ,
where 0 < η 0.32409 , and thus N 1 = 3.0856 and N = 3.98 . Finally, we calculate that
N 2 1 + 5 e 1 N N 1 = 0.62403 < 1 ,
which implies that all the conditions of Theorem 2 are met. Therefore, the system (32) is controllable.
 Remark 2. 
We see that the controllability of systems (31) and (32) cannot determined using Theorems 3.1 and 3.2 in [16] because the matrix Ξ is singular and the square of matrix Ξ is used in [16] instead of Ξ; furthermore, systems (31) and (32) are considered with matrix Ξ instead of Ξ 2 .

6. Conclusions

In this work, first via a delay Gramian matrix and the solutions of these systems, we developed some sufficient and necessary criteria for the controllability of linear stochastic delay systems. Moreover, by employing Krasnoselskii’s fixed point theorem and the solutions of these systems, we derived some sufficient criteria of controllability of nonlinear stochastic delay systems. As a result, the results apply to any matrix, not necessarily just those that are nonsingular and squared.

Author Contributions

Conceptualization, B.A., X.W. and A.M.E.; data curation, B.A., X.W. and A.M.E.; formal analysis, B.A., X.W. and A.M.E.; software, A.M.E.; supervision, X.W.; validation, B.A., X.W. and A.M.E.; visualization, B.A., X.W. and A.M.E.; writing—original draft, A.M.E.; writing—review and editing, B.A., X.W. and A.M.E.; investigation, X.W. and A.M.E.; methodology, B.A., X.W. and A.M.E.; funding acquisition, B.A. All authors have read and agreed to the published version of the manuscript.

Funding

Princess Nourah bint Abdulrahman University Researchers Supporting Project number (PNURSP2022R216), Princess Nourah bint Abdulrahman University, Riyadh, Saudi Arabia.

Data Availability Statement

Not applicable.

Acknowledgments

The authors sincerely appreciate the editor and anonymous referees for their careful reading and helpful comments to improve this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rajivganthi, C.; Thiagu, K.; Muthukumar, P.; Balasubramaniam, P. Existence of solutions and approximate controllability of impulsive fractional stochastic differential systems with infinite delay and Poisson jumps. Appl. Math. 2015, 60, 395–419. [Google Scholar] [CrossRef] [Green Version]
  2. Muthukumar, P.; Rajivganthi, C. Approximate controllability of stochastic nonlinear third-order dispersion equation. Internat. J. Robust Nonlinear Control 2014, 24, 585–594. [Google Scholar] [CrossRef]
  3. Ahmed, H.M. Semilinear neutral fractional stochastic integro-differential equations with nonlocal conditions. J. Theoret. Probab. 2015, 28, 667–680. [Google Scholar] [CrossRef]
  4. El-Borai, M.M.; EI-Nadi, K.E.S.; Fouad, H.A. On some fractional stochastic delay differential equations. Comput. Math. Appl. 2010, 59, 1165–1170. [Google Scholar] [CrossRef] [Green Version]
  5. Da Prato, G.; Zabczyk, J. Stochastic Equations in Infinite Dimensions; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  6. Diop, M.A.; Ezzinbi, K.; Lo, M. Asymptotic stability of impulsive stochastic partial integrodifferential equations with delays. Stochastics 2014, 86, 696–706. [Google Scholar] [CrossRef]
  7. Sakthivel, R.; Revathi, P.; Ren, Y. Existence of solutions for nonlinear fractional stochastic differential equations. Nonlinear Anal. 2013, 81, 70–86. [Google Scholar] [CrossRef]
  8. Agarwal, R.P.; Bazighifan, O.; Ragusa, M.A. Nonlinear neutral delay differential equations of fourth-order: Oscillation of solutions. Entropy 2021, 23, 129. [Google Scholar] [CrossRef]
  9. Choucha, A.; Boulaaras, S.M.; Ouchenane, D.; Alkhalaf, S.; Mekawy, I.; Abdalla, M. On the system of coupled nondegenerate Kirchhoff equations with distributed delay:global existence an exponential decay. J. Funct. Spaces 2021, 2021, 5577277. [Google Scholar]
  10. Petrovic, A.; Milosevic, M. The truncated Euler-Maruyama method for highly non-linear neutral stochastic differential equations with time-dependent delay. Filomat 2021, 35, 2457–2484. [Google Scholar] [CrossRef]
  11. Khusainov, D.Y.; Shuklin, G.V. Linear autonomous time-delay system with permutation matrices solving. Stud. Univ. Zilina. Math. Ser. 2003, 17, 101–108. [Google Scholar]
  12. Khusainov, D.Y.; Diblík, J.; Růžičková, M.; Lukáčová, J. Representation of a solution of the Cauchy problem for an oscillating system with pure delay. Nonlinear Oscil. 2008, 11, 276–285. [Google Scholar] [CrossRef]
  13. Elshenhab, A.M.; Wang, X.T. Representation of solutions for linear fractional systems with pure delay and multiple delays. Math. Meth. Appl. Sci. 2021, 44, 12835–12850. [Google Scholar] [CrossRef]
  14. Elshenhab, A.M.; Wang, X.T. Representation of solutions of linear differential systems with pure delay and multiple delays with linear parts given by non-permutable matrices. Appl. Math. Comput. 2021, 410, 1–13. [Google Scholar] [CrossRef]
  15. Elshenhab, A.M.; Wang, X.T. Representation of solutions of delayed linear discrete systems with permutable or nonpermutable matrices and second-order differences. RACSAM Rev. R. Acad. Cienc. Exactas Fís. Nat. Ser. A Mat. 2022, 116, 1–15. [Google Scholar]
  16. Sathiyaraj, T.; Wang, J.; O’Regan, D. Controllability of stochastic nonlinear oscillating delay systems driven by the Rosenblatt distribution. Proc. Roy. Soc. Edinburgh Sect. A 2021, 151, 217–239. [Google Scholar] [CrossRef]
  17. Elshenhab, A.M.; Wang, X.T. Controllability and Hyers–Ulam stability of differential systems with pure delay. Mathematics 2022, 10, 1248. [Google Scholar] [CrossRef]
  18. Elshenhab, A.M.; Wang, X.T.; Mofarreh, F.; Bazighifan, O. Exact solutions and finite time stability of linear conformable fractional systems with pure delay. CMES 2022, 134, 1–14. [Google Scholar] [CrossRef]
  19. Elshenhab, A.M.; Wang, X.T.; Bazighifan, O.; Awrejcewicz, J. Finite-time stability analysis of linear differential systems with pure delay. Mathematics 2022, 10, 1539. [Google Scholar] [CrossRef]
  20. Liang, C.; Wang, J.; O’Regan, D. Controllability of nonlinear delay oscillating systems. Electron. J. Qual. Theory Differ. Equ. 2017, 2017, 1–18. [Google Scholar] [CrossRef]
  21. Elshenhab, A.M.; Wang, X.T.; Cesarano, C.; Almarri, B.; Moaaz, O. Finite-Time Stability Analysis of Fractional Delay Systems. Mathematics 2022, 10, 1883. [Google Scholar] [CrossRef]
  22. Rosenblatt, M. Independence and dependence. In Proceedings of the 4th Berkeley Symposium on Mathematical Statistics and Probability, Volume 1: Contributions to the Theory of Statistics, Berkeley, CA, USA, 20 June–30 July 1960; University of California Press: Berkeley, CA, USA, 1961; pp. 431–443. [Google Scholar]
  23. Shen, G.J.; Ren, Y. Neutral stochastic partial differential equations with delay driven by Rosenblatt process in a Hilbert space. J. Korean Stat. Soc. 2015, 4, 123–133. [Google Scholar] [CrossRef]
  24. Maejima, M.; Tudor, C.A. Selfsimilar processes with stationary increments in the second Wiener chaos. Probab. Math. Stat. 2012, 32, 167–186. [Google Scholar]
  25. Shen, G.; Sakthivel, R.; Ren, Y.; Li, M. Controllability and stability of fractional stochastic functional systems driven by Rosenblatt process. Collect. Math. 2020, 71, 63–82. [Google Scholar] [CrossRef]
  26. Maejima, M.; Tudor, C.A. On the distribution of the Rosenblatt process. Stat. Probab. Lett. 2013, 83, 1490–1495. [Google Scholar] [CrossRef]
  27. Tudor, C.A. Analysis of the Rosenblatt process. ESAIM Probab. Stat. 2008, 12, 230–257. [Google Scholar] [CrossRef] [Green Version]
  28. Lakhel, E.H.; McKibben, M. Controllability for time-dependent neutral stochastic functional differential equations with Rosenblatt process and impulses. Int. J. Control Autom. Syst. 2019, 17, 286–297. [Google Scholar] [CrossRef]
  29. Sharma, J.P.; George, R.K. Controllability of matrix second order systems: A trigonometric matrix approach. Electron. J. Diff. Equ. 2007, 80, 1–14. [Google Scholar]
  30. Kilbas, A.A.; Srivastava, H.M.; Trujillo, J.J. Theory and Applications of Fractional Differential Equations; Elsevier Science BV: Amsterdam, The Netherlands, 2006. [Google Scholar]
  31. Smart, D.R. Fixed Point Theorems; University Press: Cambridge, UK, 1980. [Google Scholar]
  32. Mahmudov, N.I.; Zorlu, S. Controllability of non-linear stochastic systems. Int. J. Control 2003, 76, 95–104. [Google Scholar] [CrossRef]
  33. Klamka, J. Stochastic controllability of linear systems with state delays. Int. J. Appl. Math. Comput. 2007, 17, 5–13. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Almarri, B.; Wang, X.; Elshenhab, A.M. Controllability of Stochastic Delay Systems Driven by the Rosenblatt Process. Mathematics 2022, 10, 4223. https://doi.org/10.3390/math10224223

AMA Style

Almarri B, Wang X, Elshenhab AM. Controllability of Stochastic Delay Systems Driven by the Rosenblatt Process. Mathematics. 2022; 10(22):4223. https://doi.org/10.3390/math10224223

Chicago/Turabian Style

Almarri, Barakah, Xingtao Wang, and Ahmed M. Elshenhab. 2022. "Controllability of Stochastic Delay Systems Driven by the Rosenblatt Process" Mathematics 10, no. 22: 4223. https://doi.org/10.3390/math10224223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop