Processing math: 100%
Research article

Approximate controllability of second-order neutral stochastic differential evolution systems with random impulsive effect and state-dependent delay

  • Received: 27 July 2024 Revised: 19 September 2024 Accepted: 25 September 2024 Published: 14 October 2024
  • MSC : 34F05, 34K45, 34K50

  • In this paper, we have discussed a class of second-order neutral stochastic differential evolution systems, based on the Wiener process, with random impulses and state-dependent delay. The system is an extension of impulsive stochastic differential equations, since its random effect is not only from stochastic disturbances but also from the random sequence of the impulse occurrence time. By using the cosine operator semigroup theory, stochastic analysis theorem, and the measure of noncompactness, the existence of solutions was obtained. Then, giving appropriate assumptions, the approximate controllability of the considered system was inferred. Finally, two examples were given to illustrate the effectiveness of our work.

    Citation: Chunli You, Linxin Shu, Xiao-bao Shu. Approximate controllability of second-order neutral stochastic differential evolution systems with random impulsive effect and state-dependent delay[J]. AIMS Mathematics, 2024, 9(10): 28906-28930. doi: 10.3934/math.20241403

    Related Papers:

    [1] Rajesh Dhayal, Muslim Malik, Syed Abbas . Solvability and optimal controls of non-instantaneous impulsive stochastic neutral integro-differential equation driven by fractional Brownian motion. AIMS Mathematics, 2019, 4(3): 663-683. doi: 10.3934/math.2019.3.663
    [2] Ramkumar Kasinathan, Ravikumar Kasinathan, Dumitru Baleanu, Anguraj Annamalai . Well posedness of second-order impulsive fractional neutral stochastic differential equations. AIMS Mathematics, 2021, 6(9): 9222-9235. doi: 10.3934/math.2021536
    [3] Shuo Ma, Jiangman Li, Qiang Li, Ruonan Liu . Adaptive exponential synchronization of impulsive coupled neutral stochastic neural networks with Lévy noise and probabilistic delays under non-Lipschitz conditions. AIMS Mathematics, 2024, 9(9): 24912-24933. doi: 10.3934/math.20241214
    [4] Yakufu Kasimu, Gulijiamali Maimaitiaili . Non-fragile ${H_\infty }$ filter design for uncertain neutral Markovian jump systems with time-varying delays. AIMS Mathematics, 2024, 9(6): 15559-15583. doi: 10.3934/math.2024752
    [5] Zhifeng Lu, Fei Wang, Yujuan Tian, Yaping Li . Lag synchronization of complex-valued interval neural networks via distributed delayed impulsive control. AIMS Mathematics, 2023, 8(3): 5502-5521. doi: 10.3934/math.2023277
    [6] Yueli Huang, Jin-E Zhang . Asymptotic stability of impulsive stochastic switched system with double state-dependent delays and application to neural networks and neural network-based lecture skills assessment of normal students. AIMS Mathematics, 2024, 9(1): 178-204. doi: 10.3934/math.2024011
    [7] Krishnan Kavitha, Velusamy Vijayakumar, Kottakkaran Sooppy Nisar, Anurag Shukla, Wedad Albalawi, Abdel-Haleem Abdel-Aty . Existence and controllability of Hilfer fractional neutral differential equations with time delay via sequence method. AIMS Mathematics, 2022, 7(7): 12760-12780. doi: 10.3934/math.2022706
    [8] Shyam Sundar Santra, Palash Mondal, Mohammad Esmael Samei, Hammad Alotaibi, Mohamed Altanji, Thongchai Botmart . Study on the oscillation of solution to second-order impulsive systems. AIMS Mathematics, 2023, 8(9): 22237-22255. doi: 10.3934/math.20231134
    [9] Sivajiganesan Sivasankar, Ramalingam Udhayakumar, Arumugam Deiveegan, Reny George, Ahmed M. Hassan, Sina Etemad . Approximate controllability of Hilfer fractional neutral stochastic systems of the Sobolev type by using almost sectorial operators. AIMS Mathematics, 2023, 8(12): 30374-30404. doi: 10.3934/math.20231551
    [10] Zihan Li, Xiao-Bao Shu, Fei Xu . The existence of upper and lower solutions to second order random impulsive differential equation with boundary value problem. AIMS Mathematics, 2020, 5(6): 6189-6210. doi: 10.3934/math.2020398
  • In this paper, we have discussed a class of second-order neutral stochastic differential evolution systems, based on the Wiener process, with random impulses and state-dependent delay. The system is an extension of impulsive stochastic differential equations, since its random effect is not only from stochastic disturbances but also from the random sequence of the impulse occurrence time. By using the cosine operator semigroup theory, stochastic analysis theorem, and the measure of noncompactness, the existence of solutions was obtained. Then, giving appropriate assumptions, the approximate controllability of the considered system was inferred. Finally, two examples were given to illustrate the effectiveness of our work.



    As an essential topic in modern control theory, controllability plays an important role in designing control systems. Approximate controllability is frequently considered by researchers because it is relatively easier to realize than exact controllability, especially in infinite dimensional systems. Therefore, approximate controllability problems for various types of control systems have been investigated in many articles. Additionally, second-order differential equations have attracted more attention due to their applications in physics, mechanics, and engineering [8,9]. Recently, there are several investigations about approximate controllability to second-order abstract differential equations. For example, the approximate controllability of second-order differential equations with finite delay and the impulsive integro-differential equations have been discussed in [16]. A set of sufficient conditions for the evolution of second-order nonlocal neutral differential inclusions to be approximately controllable have been established in [28]. Palanisamy et al. [21] studied the following second-order neutral stochastic differential:

    {d[x(t)f(t,xt)]=[Ax(t)+Bu(t)]dt+g(t,xt)dW(t)+Zh(t,xt,η)˜N(dt,dη),tJ:=[0,b],x0=φB,x(0)=ξ.

    Authors constructed a Cauchy sequence by means of the range condition, and then obtained the sufficient conditions of its approximate controllability.

    In many science engineering fields, differential equations with delay are usually used to simulate dynamic systems. Application of delay differential equations in the field of biological sciences have been explored by Rihan in his monograph [23]. Bellen et al. [4] established a numerical scheme to analyze the stability of differential equations with time delay. Some researchers have noticed that the time delay may not always be a constant; it may change with the state of the system. Differential equations with state-dependent delay arise from applications and have attracted increasing attention from scholars. Bˊelair [3] considered the population model with state-dependent delay. Hernˊandez et al. [11] discussed the existence, uniqueness, and approximate controllability of solutions of first-order differential equations based on state-dependent delay. Ravichandran et al. [22] combined the fixed point theorem and resolvent operator to deduce the exact controllability of solutions of neutral integro-differential equations with state-dependent delay, and further deduced the continuous dependence of the system. The scholar of [19,20] discussed the stability and existence of periodic solutions of state-dependent delay differential equations, respectively.

    On the other hand, impulse perturbations are ubiquitous in natural phenomena. Stochastic impulsive differential equations have attracted considerable attention in current research. For the approximate controllability problem of second-order neutral stochastic impulsive systems, we refer the readers to [14,26] and the references therein. Very recently, Huang et al. [13] considered the following second-order neutral impulsive stochastic equations with state-dependent delay and a Poisson jump:

    {d[x(t)F(t,xt)]=[Ax(t)+f(t,xt)+Bu(t)]dt+σ(t,xρ(t,xt))dW(t)+Uh(t,x(t),ν)˜N(dt,dυ),tJ=[0,T],ttk,x(tk)=Ik(xtk),x(tk)=˜Ik(xtk),k=1,2,,n,x0=ϕB,x(0)=x1H,

    where the history xt:(,0]H,xt(θ)=x(t+θ),t0, belongs to the phase space B, ρ:J×BH, Ik,˜Ik:BH(k=1,2,), and Δx(tk) represents the jump of the function x at tk. Using Sadovskii's fixed point theorem, Lipschitz continuity, and phase space theory, the existence of a system solution was proved, and then the sufficient conditions of the approximate controllability to the system were established.

    However, systems with determining impulse occurrence time may not adequately describe the characteristics of some complex phenomena in real life. It is significant to study the systems with the influence of random impulses, which means its impulse occurrence time is a group of random sequences. So, the system with random impulses fairly differs from that with determining impulses. In recent years, there are several articles devoted to the existence, uniqueness, and other quantitative and qualitative properties of mild solutions of random impulsive differential equations. Guo et al. [10] obtained the existence of mild solutions of first-order Hamiltonian stochastic impulsive differential equations by using a variational method and Legendre transformation. Jose et al. [15] deduced the existence of solutions of integro-differential equations with random impulses through the Banach fixed point theorem and appropriate estimation. Existence and Hyers-Ulam stability of stochastic functional differential equations with random impulses and finite delays were investigated in [17]. In [25], the following second-order neutral random impulsive stochastic equation was considered:

    {d[x(t)g(t,xt)]=[Ax(t)+f(t,xt)]dt+σ(t,xt)dW(t),t>t0,tξk,x(ξk)=bk(τk)x(ξk),x(ξk)=bk(τk)x(ξk),k=1,2,,xt0=ϕ,x(t0)=ψ.

    The form of the solution of the equation was derived by Laplace transformation, and then the existence of the solution was verified by noncompact measures and M¨onch's fixed point theory. Then, the exponential stability was given accordingly. Yang et al. [33] proved the existence of solutions of random impulsive partial differential equations by using noncompact semigroup theory. Recently, the existence of upper and lower solutions to second-order random impulsive differential equations with a boundary value problem has been considered in [18]. In the latest research [5,30,31], random impulses have been introduced into network models, chaotic systems, and so on. Stability, control, and its application issues have been studied.

    With the continuous advancement of the theory of random impulsive differential systems, great progress has been made, but there is still much space for research on the approximate controllability of random impulsive differential systems. Based on the previous cited works of [13,21,25], we study the existence and approximate controllability of mild solutions to second-order neutral stochastic differential equations with random impulsive and state-dependent delay, as follows:

    d[x(t)g(t,xt)]=Ax(t)dt+Bu(t)dt+f(t,xρ(t,xt))dt+η(t,xρ(t,xt))dW(t),tJ=[0,T]{ξk},k=1,2,, (1.1)
    x(ξk)=qk(εk)x(ξk),x(ξk)=qk(εk)x(ξk), (1.2)
    x0=φ,x(0)=ψ, (1.3)

    where x() takes value in a Hilbert space X with the norm . A is the infinitesimal generator of a strongly continuous cosine operator C(t) on X. u() is the control function and u()L2(J,L2(Ω,U)). B is a bounded linear operator from U to X. Suppose K is another Hilbert space, and W(t) is a given Kvalued Wiener process with a finite trace Q. The functions f, g:J×BX, and η:J×BLQ(K,X), where LQ(K,X) is the space of all Q-Hilbert-Schmidt operators. Function xt:(,0]X, xt(s)=x(t+s) belongs to some phase space B, and xρ(t,xt) stands for time delay depending on the state ρ(t,xt), where ρ:J×B(,T] is a continuous function. x(ξk) represents the left limit of x(ξk). Suppose ξ0=0 and {ξk} is an increasing sequence, that is, 0=ξ0<ξ1<<ξk<, satisfying ξk=ξk1+εk,(k=1,2,). {εk} is a sequence of random variables mutually independent from Ω to Dk=(0,dk), where 0<dk<. qk maps Dk into R for each k=1,2,. Assume that φB and ψX are independent with εk.

    The main motivations and contributions in this paper are as follows:

    (ⅰ) We consider the existence and approximate controllability problem to a class of second-order impulsive stochastic differential equations with state-dependent delay, which is a more realistic abstract wave equation involving the Wiener process and random impulses sequences. As far as we know, there are very few studies of such issues.

    (ⅱ) To prove the main result, we employ evolution operator theory, stochastic analysis skills, the inequality technique, the Ascoli-Azela theorem, and combine the measure of noncompactness under a stochastic case. Then corresponding sufficient conditions of existence of a mild solution result have been established.

    (ⅲ) We further discuss the approximate controllability of Eqs (1.1)–(1.3) based on the main technique in [21]. Compared with [21], we considered the random impulsive effect and state-dependent delay. We extend the corresponding conditions to the random impulse system. This method is different from [13]. We also briefly analyzed the conclusion of the approximate controllability of the mild solution to the system under nonlocal conditions.

    (ⅳ) Two examples are given to show the effectiveness of the results.

    The framework of this paper is as follows: In Section 2, we give some notation and preparatory knowledge adopted from [2,6,7,27] and so on. In Section 3, some assumptions are given to verify the existence of solutions of differential systems. In Section 4, we study the approximate controllability of random impulsive neutral stochastic differential equations, and give proper assumptions on the premise that the corresponding equations are approximately controllable, and then obtain sufficient conditions for the approximate controllability of the system. Section 5 proves the approximate controllability of second-order differential equations under nonlocal conditions. We give examples to verify the theoretical results of this paper in Section 6.

    Let (Ω,F,{Ft}t0,P) be a complete probability space with flow, ωΩ. K and X are both real separable Hilbert spaces. Let Q:KK be a symmetric nonnegative trace family operator with Tr(Q)=n=1λn<, where {λn}n=1 is a nonnegative eigenvalue sequence of operator Q. Let {fn}n=1 be a set of complete orthogonal bases in space K, and then Qfn=λnfn. Assume βn(t) is a sequence of real-valued one-dimensional standard Wiener process defined on (Ω,F,{Ft}t0,P). W(t)=n=1λnβn(t)fn(t) is called a Q-Wiener process. Assume Ft=σ(W(s):0st). Denote L(K,X) as the space of all bounded operators from K to X. An operator φL(K,X) is called a Q-Hilbert Schmidt operator if φQ<, where Q is defined by

    φ2Q=Tr(φQφ)=n=1λnφfn2.

    Let LQ(K,X) denote the space of all Q-Hilbert Schmidt operators. The completion LQ(K,X) of L(K,X) with respect to the topology induced by the norm φ2Q=<φ,φ> is a Hilbert space. Assume L2(Ω,X) is the set of all strongly measurable and mean integrable X-valued random variables with norm xL2=(Ex2)1/2, where E stands for expectation define as E(x)=Ωx(ω)dP, and then L2(Ω,X) is a Banach space. The subset L20(Ω,X) is defined as L20(Ω,X)={xL2(Ω,X):x isF0-measurable}.

    The family of bounded linear operators {C(t),tR} is called a strongly continuous cosine family, if

    (ⅰ) C(s+t)+C(st)=2C(s)C(t) for all s, tR;

    (ⅱ) C(0)=I;

    (ⅲ) C(t)x is continuous in t on R for each fixed xX.

    The strongly continuous sine family {S(t),tR} associated with the cosine family is defined by S(t)x=t0C(s)xds,xX,tR. For more details on the theory of the cosine function of operators, one can see [27].

    Referring to [12], the axioms of phase space B can be established.

    Definition 2.1. [12] Assume phase space B consists of all F0adapted functions from (,0] to X with seminorm B, and then the following axiomatic conditions hold:

    1) If x:(,γ+a]X,a>0, such that xγB and x|[γ,γ+a]C([γ,γ+a];X), then for every t[γ,γ+a], the following conditions hold:

    (a) xtB;

    (b) x(t)KxtB;

    (c) xtBR(tγ)supγstEx(s)+T(t+γ)ExγB;

    where K>0 is a constant, R(),T():[0,+)[1,+), R() is continuous, T() is locally bounded, and then K,R(), and T() have no concern with x().

    2) The space B is complete.

    We denote DPC([a,b],L2(Ω,X)) as the set of all piecewise continuous functions, with a first derivative, mapping the interval [a,b] to L2(Ω,X) and Ft-adapted processes. If xDPC((,T],L2(Ω,X)), then x is continuous as tξk, x(ξk)=x(ξk), and x(ξ+k) exists, k=1,2,,n. Then (DPC,DPC) is a Banach space with norm

    xDPC=suptJ(xt2B)1/2,

    where the estimate of xtB is given by the following lemma.

    Lemma 2.1. [32] Let x:(,T]X be an Ft-adapted process such that F0-adapted process x0=φ(t)L20(Ω,B),x|JDPC(J,L2(Ω,X)), and then

    xsBTmEφB+Rmsup0sTEx(s),

    where Tm=suptJT(t) and Rm=suptJR(t).

    Lemma 2.2. [7] Note map m:JX is an arbitrary LQ(K,X)-valued predictable process, and then for every tJ,p2, the following inequality holds:

    Et0m(s)dW(s)pCp(t0(Em(s)pQ)2pds)p2,

    where Cp=(p(p1)2)p2.

    Now, we introduce the definition and properties of the non-compactness measure used in the theoretical proof of this paper.

    Definition 2.2. [2] β() represents the Hausdorff non-compactness measure (NCM), which is defined on bounded subset D of the Banach space by

    β(D)=inf{ϵ>0,D has a finite ϵnet in X}.

    Lemma 2.3. [2] There exists nonempty bounded subsets C,DX, where X is a real separable space, and then the following properties hold:

    (i) β(D)=0 iff D is pre-compact on X.

    (ii) β(D)=β(¯D)=β(conv(D)), where ¯D and conv(D) are for the closure and convex hull of D, respectively.

    (iii) If CD, then β(C)β(D).

    (iv) β({κ}D)=β(D), for all κX.

    (v) β(C+D)β(C)+β(D), where C+D={κ+ι;κC,ιD}.

    (vi) β(CD)max{β(C),β(D)}.

    (vii) β(μD)|μ|β(D),μR.

    Lemma 2.4. [2] For bounded and equicontinuous set DL2(Ω,X), β(D) is continuous on J, and β(D)=suptJβ(D(t)).

    Lemma 2.5. [2] Suppose sequence {xn}n=1 relating to Bochner integrable functions maps J to L2(Ω,X), and then D={xn}n=1 is a bounded and countable set, and β(D(t)) is a Lebegue integral on L2(Ω,X), which satisfies

    β({t0xn(s)ds:n1})2t0β(D(s))ds.

    Lemma 2.6. [24] If the set DLp(J,LQ(K,H)), W(t) is a Q-Wiener process, then for any p2, t[0,T], Hausdorff NCM β satisfies

    β(t0D(s)dW(s))Tp2(p1)Tr(Q)β(D(t)),

    where

    t0D(s)dW(s)={t0u(s)dW(s):for all uD,t[0,T]}.

    Remark 2.1. Specially, when p=2,

    β(t0D(s)dW(s))Tr(Q)t0β(D(s))ds.

    Lemma 2.7. [6] Let αR+, m() be nonnegative continuous function. If there is

    u(t)α+t0m(s)u(s)ds,fort[0,T],

    then,

    u(t)αet0m(s)ds.

    In this section, the existence of mild solutions of evolution systems (1.1)–(1.3) will be derived.

    Definition 3.1. [25] An Ft-adapted process x:(,T]X is a mild solution of systems (1.1)–(1.3) if xt,xρ(t,xt)B, x|JDPC(J,L2(Ω,X)), and

    (ⅰ) x0=φ(t)L20(Ω,B) for t(,0];

    (ⅱ) x(0)=ψ(t)L20(Ω,X) for tJ;

    (ⅲ) The function g(t,xt) is continuous and f(t,xρ(t,xt)) and η(t,xρ(t,xt)) are integrable. For given T(0,), x(t) satisfies:

    x(t)=+k=0[ki=1qi(εi)C(t)φ(0)+ki=1qi(εi)S(t)[ψg(0,φ)]+ki=1kj=iqj(εj)ξiξi1C(ts)g(s,xs)ds+tξkC(ts)g(s,xs)ds+ki=1kj=iqj(εj)ξiξi1S(ts)(Bu(s)+f(s,xρ(s,xs)))ds+tξkS(ts)(Bu(s)+f(s,xρ(s,xs)))ds+ki=1kj=iqj(εj)ξiξi1S(ts)η(s,xρ(s,xs))dW(s)+tξkS(ts)η(s,xρ(s,xs))dW(s)ds]δ[ξk,ξk+1)(t),t[0,T],

    where kj=iqj(εj)=qk(εk)qk1(εk1)qi(εi), and δA(t) is the index function, that is,

    δA(t)={0,tA,1,tA.

    To acquire the desired results, we give the following assumptions:

    (H1): For every v1,v2B,tJ, there exists positive constant L1 such that

    Eg(t,v1)g(t,v2)2+Ef(t,v1)f(t,v2)2+Eη(t,v1)η(t,v2)2Q3L1v1v22B,

    where L1=suptJE{g(t,0)2,f(t,0)2,Tr(Q)η(t,0)2}.

    (H2): E{maxi,kkj=iqj(εj)} is uniformly bounded, and there is ˜M>0 such that

    E{maxi,kkj=iqj(εj)2}˜M,

    for every εj,j=1,2,.

    (H3): C(t) and S(t) are continuous in the uniform operator topology for every t>0, and there exists constants M1,M2>0 such that

    suptR+C(t)2M1,suptR+S(t)2M2.

    (H4): We assume function ρ:J×B(,T] is continuous. Function tφt maps set (ρ)={ρ(s,ς)0,ρ(s,ς):(s,ς)J×B} to B. There exists a continuous and bounded function Dφ:(ρ)(0,) such that

    φtBDφ(t)φB,t(ρ).

    (H5): The functions g,f:J×BL2(Ω,X) and η:J×BLQ(K,L2(Ω,X)) have the following properties:

    (a) The functions g(t,),f(t,):BL2(Ω,X), and η(t,):BLQ(K,L2(Ω,X)) are continuous for every tJ, and for vB, g(,v),f(,v):JL2(Ω,X), and η(,v):JLQ(K,L2(Ω,X)) are measurable.

    (b) There is integrable function nhL1(J,R+) and continuous nondecreasing function Ph:R+R+ such that

    Eh(t,v)2nh(t)Ph(Ev2B),limrinfPh(r)r=0.

    (c) There exists LhL1(J,R+) and any bounded set DL2(Ω,X) such that Hausdorff NCM β satisfies:

    β(h(t,D))Lh(t)suptJβ(D),suptJLh(t)=ˉLh<,

    where it is effective for functions g,f, and η to replace h in (b) and (c).

    Lemma 3.1. [1] Let xDPC((,T],X), such that x0=φ and x(0)=ψ, and then

    xρ(s,xs)B(Tm+ˉDφ)φB+Rmsup{Ex(θ):θ[0,max{0,s}]},s(ρ)J,

    where ˉDφ=sup{Dφ(t):t(ρ)}.

    Theorem 3.1. If hypotheses (H1)–(H5) are satisfied, then evolution systems (1.1)–(1.3) have at least one mild solution.

    Proof: Define the function z:(,T]X by

    z(t)={φ(t),t(,0],+k=0[ki=1qi(εi)C(t)φ(0)]δ[ξk,ξk+1)(t),tJ.

    Then, we denote the function ˉx that satisfies Definition 3.1, and can be decomposed as ˉx=x(t)+z(t) for tJ. From Lemmas 2.1 and 3.1, it is easy to get:

    ˉxt2B=xt+zt2B2xt2B+2zt2B2R2msup0stEx(s)2+c1, (3.1)

    where c1=4R2m˜MM1Eφ(0)2+4T2mEφ2B;

    ˉxρ(s,xs)2B2xρ(s,xs)2B+2zt2B2R2msup0stEx(s)2+c2, (3.2)

    where c2=4R2m˜MM1Eφ(0)2+4(Tm+ˉDφ)2Eφ2B.

    Let Y={xDPC:x(0)=0} be a space endowed with a uniform convergence topology. Denote Br(0,Y)={xY:Ex2r} for r>0. Define operator θ:YY, such that (θx)(t)=0 as t(,0], and

    (θx)(t)=+k=0[ki=1qi(εi)S(t)[ψg(0,φ)]+ki=1kj=iqj(εj)ξiξi1C(ts)g(s,ˉxs)ds+tξkC(ts)g(s,ˉxs)ds+ki=1kj=iqj(εj)ξiξi1S(ts)(Bu(s)+f(s,ˉxρ(s,ˉxs)))ds+tξkS(ts)(Bu(s)+f(s,ˉxρ(s,ˉxs)))ds+ki=1kj=iqj(εj)ξiξi1S(ts)η(s,ˉxρ(s,ˉxs))dW(s)+tξkS(ts)η(s,ˉxρ(s,ˉxs))dW(s)]δ[ξk,ξk+1)(t),t[0,T].

    Now, we show that operator θ has a fixed point by the following steps.

    Step 1. We first prove that there exists an r such that θ maps Br into Br. Without loss of generality, let E(θx)(t)2>r, and then

    r<Ex(t)210ki=1qi(εi)2M2[Eψ2+g(0,φ)2]+5[ki=1kj=iqj(εj)ξiξi1C(ts)Eg(s,ˉxs)ds+tξkC(ts)Eg(s,ˉxs)ds]2+5[ki=1kj=iqj(εj)ξiξi1S(ts)EBu(s)ds+tξkS(ts)EBu(s)ds]2+5[ki=1kj=iqj(εj)ξiξi1S(ts)Ef(s,ˉxρ(s,ˉxs))ds+tξkS(ts)Ef(s,ˉxρ(s,ˉxs))ds]2+5[ki=1kj=iqj(εj)ξiξi1S(ts)Eη(s,ˉxρ(s,ˉxs))dW(s)+tξkS(ts)Eη(s,ˉxρ(s,ˉxs))dW(s)]210ki=1qi(εi)2M2[Eψ2+g(0,φ)2]+5max{1,˜M}Tt0C(ts)2Eg(s,ˉxs)2ds+5max{1,˜M}Tt0S(ts)2EBu2ds+5max{1,˜M}Tt0S(ts)2Ef(s,ˉxρ(s,ˉxs))2ds+5max{1,˜M}Tr(Q)t0S(ts)2Eη(s,ˉxρ(s,ˉxs))2Qds10˜MM2[Eψ2+L1(1+φ2B)]+5max{1,˜M}M1TngL1(J,X)Pg(2R2mr+c1)+5max{1,˜M}M2TEBu(t)2L2(J,X)+5max{1,˜M}M2TnfL1(J,X)Pf(2R2mr+c2)+5max{1,˜M}M2Tr(Q)nηL1(J,X)Pη(2R2mr+c2).

    Both sides of the above formula are divided by r at the same time, and it is not difficult to find

    10˜MM2[Eψ2+L1(1+φ2B)]r+5max{1,˜M}M2TEBu(t)2L2(J,X)r=0,asr.

    Then, there is

    15max{1,˜M}M1T2ngL1(J,X)Pg(2R2mr+c1)r+5max{1,˜M}M2T2nfL1(J,X)Pf(2R2mr+c2)r+5max{1,˜M}M2Tr(Q)TnηL1(J,X)Pη(2R2mr+c2)r,

    where

    limrPg(2R2mr+c1)r=limrPg(2R2mr+c1)2R2mr+c12R2mr+c1r=0.

    Similarly, limrPf(2R2mr+c2)r=limrPη(2R2mr+c2)r=0. Thus, 10, which is obviously contradictory. Accordingly, there exists an r>0 such that θ(Br)Br.

    Step 2. θ:YY is continuous. Assume {xn}+n=0Y such that xnx, as n. Let control function u() is continuous, and then

    E(θxn)(t)(θx)(t)24[ki=1kj=iqj(εj)ξiξi1C(ts)Eg(s,ˉxns)g(s,ˉxs)ds+tξkC(ts)Eg(s,ˉxns)g(s,ˉxs)ds]2+4[ki=1kj=iqj(εj)ξiξi1S(ts)EB(uˉxn(s)uˉx(s))ds+tξkS(ts)EB(uˉxn(s)uˉx(s))ds]2+4[ki=1kj=iqj(εj)ξiξi1S(ts)Ef(s,ˉxnρ(s,ˉxns)f(s,ˉxρ(s,ˉxs))ds+tξkS(ts)Ef(s,ˉxnρ(s,ˉxns))f(s,ˉxρ(s,ˉxs))ds]2+4[ki=1kj=iqj(εj)ξiξi1S(ts)Eη(s,ˉxnρ(s,ˉxns)η(s,ˉxρ(s,ˉxs))dW(s)+tξkS(ts)Eη(s,ˉxnρ(s,ˉxns))η(s,ˉxρ(s,ˉxs))dW(s)]2,

    where xnx implies ˉxnˉx. Since B is a bounded linear operator and uˉxn(s)uˉx(s)0 as ˉxnˉx0, then B(uˉxn(s)uˉx(s))0 as ˉxnˉx. In view of the continuity of g(t,),f(t,), and η(t,), we have

    E(θxn)(t)(θx)(t)20.

    Step 3. We prove that θ(Br) is equicontinuous on every [ξk,ξk+1),(k=1,2,). Denote

    r1=2R2mr+c1,r2=2R2mr+c2.

    Let ξkt1<t2<ξk+1, and then as t1t2,

    (θx)(t1)(θx)(t2)+k=0[ki=1kj=iqj(εj)ξiξi1[C(t1s)C(t2s)]g(s,ˉxs)ds+t1ξk[C(t1s)C(t2s)]g(s,ˉxs)dst2t1C(t2s)g(s,ˉxs)ds+ki=1kj=iqj(εj)ξiξi1[S(t1s)S(t2s)]Bu(s)ds +t1ξk[S(t1s)S(t2s)]Bu(s)dst2t1S(t2s)Bu(s)ds+ki=1kj=iqj(εj)ξiξi1[S(t1s)S(t2s)]f(s,ˉxρ(s,ˉxs))ds+t1ξk[S(t1s)S(t2s)]f(s,ˉxρ(s,ˉxs))dst2t1S(t2s)f(s,ˉxρ(s,ˉxs))ds+ki=1kj=iqj(εj)ξiξi1[S(t1s)S(t2s)]η(s,ˉxρ(s,ˉxs))dW(s)+t1ξk[S(t1s)S(t2s)]η(s,ˉxρ(s,ˉxs))dW(s)t2t1S(t2s)η(s,ˉxρ(s,ˉxs))dW(s)]δ[ξk,ξk+1)(t2),

    and then,

    E(θx)(t1)(θx)(t2)28max{1,˜M}TL1(1+r1)t10C(t1s)C(t2s)2ds+8(t2t1)M1TL1(1+r1)+8max{1,˜M}EBu(s)2L2(J,X)t10S(t1s)S(t2s)2ds+8(t2t1)M2EBu(s)2L2(J,X)+8max{1,˜M}TL1(1+r2)t10S(t1s)S(t2s)2ds+8(t2t1)M2TL1(1+r2)+8max{1,˜M}Tr(Q)L1(1+r2)t10S(t1s)S(t2s)2ds+8(t2t1)M2Tr(Q)L1(1+r2).

    By the continuity of C(t) and S(t), E(θx)(t1)(θx)(t2)20 as t1t2, which means that θ is equicontinuous.

    Step 4. Let O={xm}m=1. We demonstrate O(t)={xm(t)|xmBr(J),m=1,2,} is relatively compact. Let xm+1=θxm,m=0,1,2,. According to the properties of the Hausdorff NCM in Lemma 2.3, we have

    β(O)=β({xm}m=0)=β({x0}{xm}m=1)=β({xm}m=1).

    Subsequently,

    β({xm(t)}m=1)=β({θxm(t)}m=0)β[{ki=1kj=iqj(εj)ξiξi1C(ts)g(s,ˉxms)ds+tξkC(ts)g(s,ˉxms)ds}m=0]+β[{ki=1kj=iqj(εj)ξiξi1S(ts)f(s,ˉxmρ(s,ˉxms))ds+tξkS(ts)f(s,ˉxmρ(s,ˉxms))ds}m=0]+β[{ki=1kj=iqj(εj)ξiξi1S(ts)η(s,ˉxmρ(s,ˉxms))dW(s)+tξkS(ts)η(s,ˉxmρ(s,ˉxms))dW(s)}m=0]2max{1,˜M}M1Lg(t)t0β({ˉxms(s)}m=0)ds+2max{1,˜M}M2Lf(t)t0β({ˉxmρ(s,ˉxms)(s)}m=0)ds+2max{1,˜M}M2Lη(t)Tr(Q)t0β({ˉxmρ(s,ˉxms)(s)}m=0)ds2max{1,˜M}M1ˉLgRmt0β({xm(s)}m=0)ds+2max{1,˜M}M2ˉLfRmt0β({xm(s)}m=0)ds+2max{1,˜M}M2ˉLηTr(Q)Rmt0β({xm(s)}m=0)ds=At0β({xm(s)}m=0)ds,

    where A=2max{1,˜M}Rm(M1ˉLg+M2ˉLf+M2ˉLηTr(Q)).

    We acquire β(O(t))At0β(O(s))ds. Due to Lemma 2.7, we have β(O(t))0, and then we can deduce that β(O(t))=0, which implies O(t) is relatively compact. Combining Steps 1–3, O is uniformly bounded and equicontinuous. Thus, β(O)=suptJβ(O(t)) and O is relatively compact. From the Ascoli-Azela Theorem, there apparently exists a convergent subsequence of {xm}m=0 and ˆx such that limmxm=ˆx. In addition, operator θ is continuous, and then,

    ˆx=limmxm=limmθxm1=θ(limmxm1)=θˆx.

    Therefore, ˆxBr(0,Y) is called the fixed point of θ, which is also the mild solution of systems (1.1)–(1.3).

    In this section, we deduce the approximate controllability of systems (1.1)–(1.3).

    Definition 4.1. [21] Let x(T,u) be a mild solution of evolution systems (1.1)–(1.3) corresponding to the control u at terminal time T. Set

    R(T)= { x(T,u):u()L2(J,L2(Ω,U) } 

    denotes the reachable set of the systems (1.1)–(1.3) at terminal time T. If ¯R(T)=L2(Ω,X), then systems (1.1)–(1.3) are said to be approximately controllable on J.

    Now, define the Nemytskil operator Γ:DPC(J,X)L2(J,X) related to the nonlinear function f by

    Γf(x)(t)=f(t,xρ(t,xt)).

    Definition 4.2. Define Ξ and bounded linear operators

    ϕ:L2(J,LQ(K,X))L2(Ω,X),Φ:L2(J,X)L2(Ω,X),

    and then,

    Ξ=+k=0[ki=1qi(εi)C(T)φ(0)+ki=1qi(εi)S(T)[ψg(0,φ)]+ki=1kj=iqj(εj)ξiξi1C(Ts)g(s,xs)ds+TξkC(Ts)g(s,xs)ds]δ[ξk,ξk+1)(T),Φh1=+k=0[ki=1kj=iqj(εj)ξiξi1S(Ts)h1(s)ds+TξkS(Ts)h1(s)ds]δ[ξk,ξk+1)(T),ϕh2=+k=0[ki=1kj=iqj(εj)ξiξi1S(Ts)h2(s)dW(s)+TξkS(Ts)h2(s)dW(s)]δ[ξk,ξk+1)(T),

    where h1L2(J,X) and h2L2(J,LQ(K,X)).

    Similarly, we give the following assumptions to verify the approximate controllability of systems (1.1)–(1.3).

    (H6): Systems (1.1*)–(1.3*) denote systems corresponding to (1.1)–(1.3) with f=0 and η=0, and RT(0,0) is the reachable set of the systems (1.1*)–(1.3*) at terminal time T. Systems (1.1*)–(1.3*) are approximately controllable, i.e., ¯RT(0,0)=L2(Ω,X).

    (H7): (ⅰ) For every ε>0, h1L2(J,X), and h2L2(J,LQ(K,X)), there exists a control function uL2(J,L2(Ω,U)) such that

    EΦh1+ϕh2ΦBu2<ε.

    (ⅱ) EBu2L2(J,X)C(Eh12L2(J,X)+Eh22L2(J,X)),

    where C is a constant independent of h1 and h2.

    (ⅲ) 2CL1R2mTL<1,

    where

    L=[1(˜L1+˜L2+˜L3)]1˜L,˜L1=4max{1,˜M}M1L1T2R2m,
    ˜L2=4max{1,˜M}M2L1T2R2m,˜L3=4max{1,˜M}M2TL1Tr(Q)R2m,
    ˜L=4max{1,˜M}M2T.

    Lemma 4.1. Any mild solution of systems (1.1)–(1.3) satisfies the following inequality if hypotheses (H1)–(H5) hold:

    sup0tTEx1x22LEBu1Bu22L2(J,X),

    where xm(m=1,2) is the solution of systems (1.1)–(1.3) related to control um(m=1,2).

    Proof: xm has the following form

    xm=+k=0[ki=1qi(εi)C(t)φ(0)+ki=1qi(εi)S(t)[ψg(0,φ)]+ki=1kj=iqj(εj)ξiξi1C(ts)g(s,xms)ds+tξkC(ts)g(s,xms)ds+ki=1kj=iqj(εj)ξiξi1S(ts)(Bum(s)+f(s,xmρ(s,xms)))ds+tξkS(ts)(Bum(s)+f(s,xmρ(s,xms)))ds+ki=1kj=iqj(εj)ξiξi1S(ts)η(s,xmρ(s,xms))dW(s)+tξkS(ts)η(s,xmρ(s,xms))dW(s)]δ[ξk,ξk+1)(t).

    So, for x1,x2X, we obtain that

    sup0tTEx1x22E{+k=0[ki=1kj=iqj(εj)ξiξi1C(ts)g(s,x1s)g(s,x2s)ds+tξkC(ts)g(s,x1s)g(s,x2s)ds+ki=1kj=iqj(εj)ξiξi1S(ts)(Bu1(s)Bu2(s)+f(s,x1ρ(s,x1s))f(s,x2ρ(s,x2s)))ds+tξkS(ts)(Bu1(s)Bu2(s)+f(s,x1ρ(s,x1s))f(s,x2ρ(s,x2s)))ds+ki=1kj=iqj(εj)ξiξi1S(ts)η(s,x1ρ(s,x1s))η(s,x2ρ(s,x2s))dW(s)+tξkS(ts)η(s,x1ρ(s,x1s))η(s,x2ρ(s,x2s))dW(s)]δ[ξk,ξk+1)(t)}24max{1,˜M}M1L1T2R2msup0tTEx1x22+4max{1,˜M}M2TEBu1Bu22L2(J,X)+4max{1,˜M}M2L1T2R2msup0tTEx1x22+4max{1,˜M}M2TL1Tr(Q)R2msup0tTEx1x22=(˜L1+˜L2+˜L3)sup0tTEx1x22+˜LEBu1Bu22L2(J,X).

    Therefore,

    sup0tTEx1x22[1(˜L1+˜L2+˜L3)]1˜LEBu1Bu22L2(J,X)=LEBu1Bu22L2(J,X).

    The proof is complete.

    Now, we prove the approximate controllability of systems (1.1)–(1.3).

    Theorem 4.1. Suppose that Lemma 4.1 and hypotheses (H6)(H7) hold. Then, systems (1.1)–(1.3) are approximately controllable.

    Proof: We can obtain the equivalent condition of approximate controllability of systems (1.1)–(1.3) from Definition 4.1.

    For any desired state of the terminal ωX, ε>0, if there exists a control function uεL2(J,L2(Ω,U)) such that the mild solution of systems (1.1)–(1.3) satisfy:

    EωΞΦΓf(xε)ϕΓη(xε)Φ(Buε)2<ε,

    where xε=x(,uε), then systems (1.1)–(1.3) are approximately controllable.

    Due to ¯RT(0,0)¯R(T), let ω¯RT(0,0), and we construct a sequence that converges to ω. According to (H6), systems (1.1*)–(1.3*) are approximately controllable. So, for ε>0, there exists uL2(J,L2(Ω,U)) and nZ+ such that

    E(ωΞΦ(Bu))2ε2n+4. (4.1)

    Let x1L2(Ω,X) be a mild solution of systems (1.1)–(1.3) under control u1. Because of (H7)(i), there exists u2L2(J,L2(Ω,U)) such that

    EΦ((Bu)Γf(x1))ϕΓη(x1)Φ(Bu2)2<ε2n+4. (4.2)

    Combining (4.1) and (4.2), we have

    EωΞΦΓf(x1)ϕΓη(x1)Φ(Bu2)22EωΞΦ(Bu)2+2EΦ((Bu)Γf(x1))ϕΓη(x1)Φ(Bu2)2<ε2n+2. (4.3)

    By using (H7)(i) again, there exists control function v2 such that

    EΦ(Γf(x2)Γf(x1))+ϕ(Γη(x2)Γη(x1))Φ(Bv2)2<ε2n+3. (4.4)

    Based on hypothesis (H7)(ii) and Lemma 4.1, we have

    EBv22L2(J,X)C(EΓf(x2)()Γf(x1)()2L2(J,X)+EΓη(x2)()Γη(x1)()2L2(J,X))C(T0Ef(,x2ρ(s,x2s))f(,x1ρ(s,x1s))2ds+T0Eη(,x2ρ(s,x2s))η(,x1ρ(s,x1s))2Qds)2CR2mL1Tsup0tTEx1x222CR2mL1TLEBu1Bu22L2(J,X).

    Set u3=u2v2, and combine (4.3) and (4.4),

    EωΞΦΓf(x2)ϕΓη(x2)Φ(Bu3)22EωΞΦΓf(x1)ϕΓη(x1)Φ(Bu2)2+2EΦ(Γf(x2)Γη(x1))+ϕ(Γη(x2)Γη(x1))Φ(Bv2)2<(12n+1+12n+2)ε. (4.5)

    By mathematical induction, we construct un+1=unvnL2(J,L2(Ω,U)) satisfying

    EωΞΦΓf(xn)ϕΓη(xn)Φ(Bun+1)2<(123+124++12n+2)ε<14ε, (4.6)

    and

    EBunBun+12L2(J,X)2CR2mL1TLEBun1Bun2L2(J,X). (4.7)

    Due to (H7)(iii), we infer that {Bun}n=1 is Cauchy and convergent. Then, for ε>0, there exists positive integer number N, n>N such that

    EΦB(un)ΦB(un+1)2<ε4.

    Hence,

    EωΞΦΓf(xn)ϕΓη(xn)Φ(Bun)22EωΞΦΓf(xn)ϕΓη(xn)Φ(Bun+1)2+2EΦ(Bun)Φ(Bun+1)2<ε.

    In summary, systems (1.1)–(1.3) are approximately controllable.

    In this section, we study the approximate controllability of second-order stochastic differential equations with nonlocal conditions.

    d[x(t)g(t,xt)]=Ax(t)dt+Bu(t)dt+f(t,xρ(t,xt))dt+η(t,xρ(t,xt))dW(t),tJ=[0,T],tξk,k=1,2,, (5.1)
    x(ξk)=qk(εk)x(ξk),x(ξk)=qk(εk)x(ξk), (5.2)
    x0=φ+H1(x),x(0)=ψ+H2(x). (5.3)

    In order to get the result, it is necessary to give some properties of functions H1 and H2.

    (H8): H1,H2 are continuous and compact, and satisfy the following conditions.

    (a) For any x,yB,

    EH1(x)H1(y)2N1xy2B,EH2(x)H2(y)2N2xy2B.

    (b) There are integrable functions nH1,nH2L1(J,R+), and continuous nondecreasing functions PH1,PH2:R+R+, such that

    EH1(x)2nH1PH1(Ex2B),limrinfPH1(r)r=0,EH2(x)2nH2PH2(Ex2B),limrinfPH2(r)r=0.

    Definition 5.1. An Ft-adapted process x:(,T]X is a mild solution of systems (5.1)–(5.3), if xt,xρ(t,xt)B, x|JDPC(J,L2(Ω,X)), and

    (ⅰ) x0=φ(t)+H1(x)L20(Ω,B) for t(,0];

    (ⅱ) x(0)=ψ(t)+H2(x)L20(Ω,X) for tJ;

    (ⅲ) The function g(t,xt) is continuous and f(t,xρ(t,xt)) and η(t,xρ(t,xt)) are integrable. For given T(0,), x(t) satisfies

    x(t)={φ,(,0],+k=0[ki=1qi(εi)C(t)[φ(0)+H1(x)]+ki=1qi(εi)S(t)[ψ+H2(x)g(0,φ)]+ki=1kj=iqj(εj)ξiξi1C(ts)g(s,xs)ds+tξkC(ts)g(s,xs)ds+ki=1kj=iqj(εj)ξiξi1S(ts)(Bu(s)+f(s,xρ(s,xs)))ds+tξkS(ts)(Bu(s)+f(s,xρ(s,xs)))ds+ki=1kj=iqj(εj)ξiξi1S(ts)η(s,xρ(s,xs))dW(s)+tξkS(ts)η(s,xρ(s,xs))dW(s)]δ[ξk,ξk+1)(t),t[0,T],

    and then, the Ft-adapted stochastic process x:(,T]X is called a mild solution to systems (5.1)–(5.3).

    Theorem 5.1. If (H1)(H8) are established, referring to the proof process of Theorems 3.1 and 4.1, then evolution systems (5.1)–(5.3) are approximately controllable.

    Example 1. In order to verify the abstract conclusions, we give the following hyperbolic wave equations with impulse at random moments:

    t[z(t,y)ttπ0c1(st,τ,y)z(s,τ)dτds]=2y2z(t,y)+Bu(t,y)+tc2(st)z(sρ1(t)ρ2(|z(t)|,y))ds (6.1)
    +b(t,tc3(st)z(sρ1(t)ρ2(|z(t)|,y))ds)dβ(t)dt,tξk,z(ξk,y)=p(k)εkz(ξk,y),z(ξk,y)=p(k)εkz(ξk,y),t=ξk, (6.2)
    z(t,0)=z(t,π)=0,t[0,1], (6.3)
    z(t,y)=φ(t,y),r<t0,0yπ,r(0,), (6.4)
    tz(0,y)=ψ(y), (6.5)

    where ρ1:[0,)[0,) and ρ2:[0,)[0,) are continuous functions. c1,c2, and c3 are suitable functions. β(t) denotes a standard cylindrical Wiener process in Hilbert space K=L2([0,π]) defined on a stochastic space (Ω,F,P). Let εk be a random variable defined on Dk(0,dk), where 0<dk<+, for k=1,2,. Suppose εi and εj are independent of each other as ij for i,j=1,2,. ξ0=t0=0 and ξk=ξk1+εk for k=1,2,. p is a function regarding k.

    Let Z=K=L2([0,π]) and define operator A:D(A)ZZ as Ax=x, where

    D(A)={zZ:z,zareabsolutelycontinuous,zZ,z(0)=z(π)=0}.

    Operator A has a discrete spectrum, and its eigenvalue is n2 and en=2πeinz,nZ. {C(t):tR} is a family of strongly continuous cosine operators, and A is its infinitesimal generator. Then

    C(t)z=n=1cos(nt)(z,en)en,zZ.

    The correlative sine family S(t) is given by

    S(t)z=n=11nsin(nt)(z,en)en,zZ.

    It is easy to infer S(t)21 and C(t)21. Hence, C(t) and S(t) are uniformly bounded for tR.

    B is a phase space with norm μB=supθ0μ(θ), and define B as a set of bounded and uniformly continuous functions from (,0] to Z. Define the functions g,f:J×BL2(Ω,Z),η:J×BLQ(K,L2(Ω,Z)),ρ:J×B(0,),zt:JL2(Ω,Z), and qk(εk),

    g(t,μ)(y)=tπ0c1(s,τ,y)μ(s,τ)dτds,f(t,μ)(y)=0c2(s)μ(s,y)ds,η(t,μ)(y)=b(t,0c3(s)μ(s,y)ds),ρ(s,μ)=ρ1(t)ρ2(|μ(s,y)|),zt=z(tr),qk(εk)=p(k)εk.

    In this way, we can rewrite the equations (6.1)–(6.5) in the form of (1.1)–(1.3). In order to get controllable results, we need to make the following assumptions.

    (a) Continuous functions c1,c2,c3:RR satisfy:

    Lc1=0π0c1(st,τ)dτds<,Lc2=0c2(s)ds<,Lc3=0c3(s)ds<.

    (b) E{maxi,kkj=ip(j)εj2}<.

    (c) Dφ:(ρ)φt is continuous, and

    φtBDφ(t)φB,t(ρ).

    (d) For (t,μ)(,0]×B,

    Eg(t,μ)2=E[π0(0π0c1(st,τ,y)μ(s)(y)dτds)2dy]hg(t)Gg(Eμ2B),Ef(t,μ)2=E[π0(0c2(st)μ(s)(y))2dy]hf(t)Gf(Eμ2B),Eη(t,μ)2=Eπ0(b(t,G(μ)(y)))2dyhη(t)Gη(Eμ2B),

    where G(μ)(y)=0c3(s)μ(s,y)ds, hg,hf and hη are integral, and Gg,Gf and Gη are all positive continuous nondecreasing functions.

    Under the above conditions, for every t[0,1], μ1,μ2B,

    Eg(t,μ1)g(t,μ2)2=E[π0(0π0c1(st,τ,y)[μ1(s)(y)μ2(s)(y)]dτds)2dy](0π0c1(st,τ)[μ1(s)μ2(s)]dτds)2(0π0c1(st,τ)dτds)2μ1μ22Blgμ1μ22B.

    For arbitrary bounded set DB,

    β(g(t,D))lgsupθ0β(D).

    Similarly, Ef(t,μ1)f(t,μ2)2lgμ1μ22B, and for arbitrary bounded set DB,

    β(f(t,D))lfsupθ0β(D).

    Suppose b follows the Lipschitz condition:

    Eb(t,γ1)b(t,γ2)2lγγ1γ22.

    Let G be bounded,

    Eη(t,μ1)η(t,μ2)2Q=Tr(Q)Eπ0b(t,G(μ1)(y))b(t,G(μ2)(y))2QdyTr(Q)lγGμ1μ22B.

    For arbitrary bounded set DB,

    β(η(t,D))Tr(Q)lγGsupθ0β(D).

    Assume L1=max{lg,lf,Tr(Q)lγG}, and then conditions (H1) and (H5) hold.

    Define U={u:u=n=2unenwithn=2u2n<}. B:UL2(Ω,X) and Bu=2u2e1+n=2unen. Then assumption (H7) holds. For a more detailed explanation, see [34]. Then, as the related systems with f=0 and η=0 are approximately controllable, based on Theorem 4.1, systems (6.1)–(6.5) are approximately controllable.

    Example 2. We then provide a numerical example to further prove the feasibility of the theoretical results.

    ddt[x(t)etsin(x(tr))5+et]=x(t)+Bucos(x(tex(tr)))8+etsin(x(tex(tr)))7dβ(t)dt,tξk, (6.6)
    x(ξk)=21kτkx(ξk),x(ξk)=21kτkx(ξk),t=ξk, (6.7)
    x(t)=cost,tx(t)=sint,r<t0,r(0,+), (6.8)

    where β(t) denotes a standard one-dimensional Wiener process. Let τk be a random variable following the exponential distribution. We assume A=1,T=30 and J=[0,30].

    We choose the state of terminal time T as x(T)=5. For every h1L2(J,X),h2L2(J,LQ(K,X)), let

    Φ1h1=+k=0[ki=1kj=i21jτjξiξi1S(30s)h1(s)ds+30ξkS(30s)h1(s)ds]δ[ξk,ξk+1)(30),ϕ1h2=+k=0[ki=1kj=i21jτjξiξi1S(30s)h2(s)dβ(s)+30ξkS(30s)h2(s)dβ(s)]δ[ξk,ξk+1)(30).

    Let B=1, and we choose control function u as Φ1u=ϕ1h1+Φ1h2.

    Figure 1 shows a sample path of the systems (6.6)–(6.8) with Bu=0, and it can see that the systems (6.6)–(6.8) are not equal to 5 at t=30. Figure 2 shows a sample path of the systems (6.6)–(6.8) under control u. It can be seen that the state value of systems (6.6)–(6.8) is very close to 5 and the error is very small.

    Figure 1.  A sample path of the systems (6.6)–(6.8) with Bu=0.
    Figure 2.  A sample path of the systems (6.6)–(6.8).

    In this paper, we pay attention to the existence and approximate controllability of mild solutions to systems (1.1)–(1.3), which can be abstracted from the second-order stochastic wave equation and extended to more general random impulses cases. To obtain the result of existence, we applied evolution operator theory, stochastic analysis skills, and the measure of noncompactness. Then, under some appropriate conditions, the approximate controllability was established. Further, we considered relevant conclusions under the nonlocal conditions. At the end of this paper, two examples were given to show the effectiveness of the results. Our work may generalize some existing results on this topic.

    Stochastic differential systems with random impulsive effect have applications in many practical problems, and there are many relative problem worth studying. In recent reference [29], Vinodkumar et al. discussed the existence, uniqueness, and stability of solutions of fractional differential equations with random impulses. As we know, the literature related to approximate controllability of fractional stochastic differential systems with random impulses remains very limited. In later work, we will continue to consider approximate controllability of fractional impulsive stochastic differential systems under the interference of various random factors such as random sequence, fractional Brownian motion, or Rosenblatt process.

    Chunli You: conceptualization, methodology, investigation, and writing-original draft; Linxin Shu: methodology, project administration, and writing-review and editing; Xiao-Bao Shu: resources, supervision, technical support. All authors have read and approved the final version of the manuscript for publication.

    This work is supported by the Jiangxi Provincial Department of Education Science and Technology Foundation (No. GJJ2201122).

    The authors declare that there are no conflicts of interest.



    [1] B. Ahmad, Instability of impulsive hybrid state dependent delay differential systems, Vietnam J. Math., 35 (2007), 285–298.
    [2] J. Banas, K. Goebel, Measure of noncompactness in Banach Space, New York: Mercel Dekker, 1980.
    [3] J. Bélair, Population models with state-dependent delays, In: Mathematical population dynamics, Boca Raton: CRC Press, 1991,165–176. https://doi.org/10.1201/9781003072706-13
    [4] A. Bellen, N. Guglielmi, M. Zennaro, Numerical stability of nonlinear delay differential equations of neutral type, J. Comput. Appl. Math., 125 (2000), 251–263. https://doi.org/10.1016/S0377-0427(00)00471-4 doi: 10.1016/S0377-0427(00)00471-4
    [5] A. J. G. Bento, J. J. Oliveira, C. M. Silva, Existence and stability of a periodic solution of a general difference equation with applications to neural networks with a delay in the leakage terms, Commun. Nonlinear Sci., 126 (2023), 107429. https://doi.org/10.1016/j.cnsns.2023.107429 doi: 10.1016/j.cnsns.2023.107429
    [6] C. Corduneanu, Principles of differential and integral equations, Boston: Allyn and Bacon, 1971.
    [7] G. Da Prato, J. Zabczyk, Stochastic equations in infinite dimensions, Cambridge: Cambridge University Press, 1992. http://doi.org/10.1017/CBO9780511666223
    [8] J. Q. Duan, W. Wang, Effective dynamics of stochastic partial differential equations, Amsterdam: Elsevier, 2014. https://doi.org/10.1016/C2013-0-15235-X
    [9] W. E. Fitzgibbon, Global existence and boundedness of solutions to the extensible beam equation, SIAM J. Math. Anal., 13 (1982), 739–745. https://doi.org/10.1137/0513050 doi: 10.1137/0513050
    [10] Y. Guo, X. B. Shu, Q. B. Yin, Existence of solutions for first-order Hamiltonian random impulsive differential equations with Dirichlet boundary conditions, Discrete Cont. Dyn. B, 27 (2022), 4455–4471. https://doi.org/10.3934/dcdsb.2021236 doi: 10.3934/dcdsb.2021236
    [11] E. Hernández, J. H. Wu, A. Chadha, Existence, uniqueness and approximate controllability of abstract differential equations with state-dependent delay, J. Differ. Equations, 269 (2020), 8701–8735. https://doi.org/10.1016/j.jde.2020.06.030 doi: 10.1016/j.jde.2020.06.030
    [12] Y. Hino, S. Murakami, T. Naito, Functional differential equations with infinite delay, Heidelberg: Springer, 1991. http://doi.org/10.1007/bfb0084432
    [13] H. Huang, Z. Wu, X. F. Su, Approximate controllability of second-order impulsive neutral stochastic differential equations with state-dependent delay and Poisson jumps, J. Inequal. Appl., 2023 (2023), 53. https://doi.org/10.1186/s13660-023-02959-5 doi: 10.1186/s13660-023-02959-5
    [14] M. Johnson, V. Vijayakumar, A. Shukla, K. S. Nisar, B. Hazarika, Existence and approximate controllability results for second-order impulsive stochastic neutral differential systems, Appl. Anal., 103 (2024), 481–505. https://doi.org/10.1080/00036811.2023.2196293 doi: 10.1080/00036811.2023.2196293
    [15] S. A. Jose, W. Yukunthorn, J. E. N. Valdes, H. Leiva, Some existence, uniqueness and stability results of nonlocal random impulsive integro-differential equations, Appl. Math. E-Notes, 20 (2020), 481–492.
    [16] A. Kumar, R. K. Vats, A. Kumar, Approximate controllability of second-order non-autonomous system with finite delay, J. Dyn. Control Syst., 26 (2020), 611–627. https://doi.org/10.1007/s10883-019-09475-0 doi: 10.1007/s10883-019-09475-0
    [17] S. Li, L. X. Shu, X. B. Shu, F. Xu, Existence and Hyers-Ulam stability of random impulsive stochastic functional differential equations with finite delays, Stochastics, 91 (2019), 857–872. https://doi.org/10.1080/17442508.2018.1551400 doi: 10.1080/17442508.2018.1551400
    [18] Z. H. Li, X. B. Shu, F. Xu, The existence of upper and lower solutions to second order random impulsive differential equation with boundary value problem, AIMS Mathematics, 5 (2020), 6189–6210. https://doi.org/10.3934/math.2020398 doi: 10.3934/math.2020398
    [19] P. Magal, O. Arino, Existence of periodic solutions for a state dependent delay differential equation, J. Differ. Equations, 165 (2000), 61–95. https://doi.org/10.1006/jdeq.1999.3759 doi: 10.1006/jdeq.1999.3759
    [20] J. Mallet-Paret, R. D. Nussbaum, Stability of periodic solutions of state-dependent delay-differential equations, J. Differ. Equations, 250 (2011), 4085–4103. https://doi.org/10.1016/j.jde.2010.10.023 doi: 10.1016/j.jde.2010.10.023
    [21] M. Palanisamy, R. Chinnathambi, Approximate controllability of second-order neutral stochastic differential equations with infinite delay and Poisson jumps, J. Syst. Sci. Complex., 28 (2015), 1033–1048. http://doi.org/10.1007/s11424-015-3075-7 doi: 10.1007/s11424-015-3075-7
    [22] C. Ravichandran, N. Valliammal, J. J. Nieto, New results on exact controllability of a class of fractional neutral integro-differential systems with state-dependent delay in Banach spaces, J. Franklin I., 356 (2019), 1535–1565. https://doi.org/10.1016/j.jfranklin.2018.12.001 doi: 10.1016/j.jfranklin.2018.12.001
    [23] F. A. Rihan, Delay differential equations and applications to biology, Singapore: Springer, 2021. https://doi.org/10.1007/978-981-16-0626-7
    [24] L. X. Shu, X. B. Shu, J. Z. Mao, Approximate controllability and existence of mild solutions for Riemann-Liouville fractional stochastic evolution equations with nonlocal conditions of order 1<α<2, Fract. Calc. Appl. Anal., 22 (2019), 1086–1112. https://doi.org/10.1515/fca-2019-0057 doi: 10.1515/fca-2019-0057
    [25] L. X. Shu, X. B. Shu, Q. X. Zhu, F. Xu, Existence and exponential stability of mild solutions for second-order neutral stochastic functional differential equation with random impulses, J. Appl. Anal. Comput., 11 (2021), 59–80. https://doi.org/10.11948/20190089 doi: 10.11948/20190089
    [26] S. Sivasankar, R. Udhayakumar, A note on approximate controllability of second-order neutral stochastic delay integro-differential evolution inclusions with impulses, Math. Method. Appl. Sci., 45 (2022), 6650–6676. https://doi.org/10.1002/mma.8198 doi: 10.1002/mma.8198
    [27] C. C. Travis, G. F. Webb, Cosine families and abstract nonlinear second order differential equations, Acta Mathematica Academiae Scientiarum Hungaricae, 32 (1978), 75–96. http://doi.org/10.1007/bf01902205 doi: 10.1007/bf01902205
    [28] V. Vijayakumar, R. Udhayakumar, C. Dineshkumar, Approximate controllability of second order nonlocal neutral differential evolution inclusions, IMA J. Math. Control I., 38 (2021), 192–210. https://doi.org/10.1093/imamci/dnaa001 doi: 10.1093/imamci/dnaa001
    [29] A. Vinodkumar, K. Malar, M. Gowrisankar, P. Mohankumar, Existence, uniqueness and stability of random impulsive fractional differential equations, Acta Math. Sci., 36 (2016), 428–442. http://doi.org/10.1016/S0252-9602(16)30010-8 doi: 10.1016/S0252-9602(16)30010-8
    [30] A. Vinodkumar, T. Senthilkumar, S. Hariharan, J. Alzabut, Exponential stabilization of fixed and random time impulsive delay differential system with applications, Math. Biosci. Eng., 18 (2021), 2384–2400. https://doi.org/10.3934/mbe.2021121 doi: 10.3934/mbe.2021121
    [31] A. Vinodkumar, T. Senthilkumar, H. Işık, S. Hariharan, N. Gunasekaran, An exponential stabilization of random impulsive control systems and its application to chaotic systems, Math. Method. Appl. Sci., 46 (2023), 3237–3254. https://doi.org/10.1002/mma.8688 doi: 10.1002/mma.8688
    [32] Z. M. Yan, X. X. Yan, Existence of solutions for impulsive partial stochastic neutral integrodifferential equations with state-dependent delay, Collect. Math., 64 (2013), 235–250. https://doi.org/10.1007/s13348-012-0063-2 doi: 10.1007/s13348-012-0063-2
    [33] Q. C. Yang, D. Wu, X. B. Shu, Existence and stability results of mild solutions for random impulsive stochastic partial differential equations with noncompact semigroups, Stochastics, 95 (2023), 168–190. https://doi.org/10.1080/17442508.2022.2056415 doi: 10.1080/17442508.2022.2056415
    [34] H. X. Zhou, Approximate controllability for a class of semilinear abstract equations, SIAM J. Control Optim., 21 (1983), 551–565. https://doi.org/10.1137/0321033 doi: 10.1137/0321033
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(421) PDF downloads(26) Cited by(0)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog