
We investigated synchronization of dynamic systems with mixed delays and delayed impulses. Using impulsive control method and the average impulsive interval approach, several Lyapunov sufficient conditions were given for ensuring synchronization in terms of impulsive perturbation and impulsive control, respectively. The derived conditions indicated that delays in continuous dynamical systems were flexible under impulsive perturbation and were not strictly dependent on the size of impulsive delays, and they may have a potential impact on synchronization of the considered system. In addition, applying the proposed concepts of average positive impulsive estimation and average impulsive estimation, we integrated the information in impulsive delay into the rate coefficient to eliminate the limitation of having the same threshold at each impulse point, while the impulsive delay maintained the synchronization effect. This was an improvement on the previous results obtained. Finally, we provided two numerical examples to illustrate the validity of our results.
Citation: Biwen Li, Qiaoping Huang. Synchronization of time-delay systems with impulsive delay via an average impulsive estimation approach[J]. Mathematical Biosciences and Engineering, 2024, 21(3): 4501-4520. doi: 10.3934/mbe.2024199
[1] | Haitao Zhu, Xinrui Ji, Jianquan Lu . Impulsive strategies in nonlinear dynamical systems: A brief overview. Mathematical Biosciences and Engineering, 2023, 20(2): 4274-4321. doi: 10.3934/mbe.2023200 |
[2] | Linni Li, Jin-E Zhang . Input-to-state stability of stochastic nonlinear system with delayed impulses. Mathematical Biosciences and Engineering, 2024, 21(2): 2233-2253. doi: 10.3934/mbe.2024098 |
[3] | A. Vinodkumar, T. Senthilkumar, S. Hariharan, J. Alzabut . Exponential stabilization of fixed and random time impulsive delay differential system with applications. Mathematical Biosciences and Engineering, 2021, 18(3): 2384-2400. doi: 10.3934/mbe.2021121 |
[4] | Yilin Tu, Jin-E Zhang . Event-triggered impulsive control for input-to-state stability of nonlinear time-delay system with delayed impulse. Mathematical Biosciences and Engineering, 2025, 22(4): 876-896. doi: 10.3934/mbe.2025031 |
[5] | Xueyan Yang, Xiaodi Li, Qiang Xi, Peiyong Duan . Review of stability and stabilization for impulsive delayed systems. Mathematical Biosciences and Engineering, 2018, 15(6): 1495-1515. doi: 10.3934/mbe.2018069 |
[6] | Zixiao Xiong, Xining Li, Ming Ye, Qimin Zhang . Finite-time stability and optimal control of an impulsive stochastic reaction-diffusion vegetation-water system driven by Lˊevy process with time-varying delay. Mathematical Biosciences and Engineering, 2021, 18(6): 8462-8498. doi: 10.3934/mbe.2021419 |
[7] | Xinli Hu, Yansheng Liu, Jianhong Wu . Culling structured hosts to eradicate vector-borne diseases. Mathematical Biosciences and Engineering, 2009, 6(2): 301-319. doi: 10.3934/mbe.2009.6.301 |
[8] | Zeyan Yue, Sheng Wang . Dynamics of a stochastic hybrid delay food chain model with jumps in an impulsive polluted environment. Mathematical Biosciences and Engineering, 2024, 21(1): 186-213. doi: 10.3934/mbe.2024009 |
[9] | Hai-Feng Huo, Fan Wu, Hong Xiang . On threshold dynamics for periodic and time-delayed impulsive systems and application to a periodic disease model. Mathematical Biosciences and Engineering, 2022, 19(1): 836-854. doi: 10.3934/mbe.2022038 |
[10] | Jian Meng, Bin Zhang, Tengda Wei, Xinyi He, Xiaodi Li . Robust finite-time stability of nonlinear systems involving hybrid impulses with application to sliding-mode control. Mathematical Biosciences and Engineering, 2023, 20(2): 4198-4218. doi: 10.3934/mbe.2023196 |
We investigated synchronization of dynamic systems with mixed delays and delayed impulses. Using impulsive control method and the average impulsive interval approach, several Lyapunov sufficient conditions were given for ensuring synchronization in terms of impulsive perturbation and impulsive control, respectively. The derived conditions indicated that delays in continuous dynamical systems were flexible under impulsive perturbation and were not strictly dependent on the size of impulsive delays, and they may have a potential impact on synchronization of the considered system. In addition, applying the proposed concepts of average positive impulsive estimation and average impulsive estimation, we integrated the information in impulsive delay into the rate coefficient to eliminate the limitation of having the same threshold at each impulse point, while the impulsive delay maintained the synchronization effect. This was an improvement on the previous results obtained. Finally, we provided two numerical examples to illustrate the validity of our results.
When control systems produce a state change at one discrete instant suddenly, system evolutions exhibit impulsive behavior, and such systems can also be called impulsive dynamical systems, which belong to a special class of hybrid systems. Impulsive systems have attracted the attention of many researchers due to their important applications in various fields such as network control systems [1], financial interest rate adjustment [2], and pharmaceutical management [3]. It is well known that delays are inevitable in signal transmission and impulsive input, so the study of impulsive delay is very necessary. Furthermore, time delay systems have been widely studied in many fields, such as neural networks [4]. Recently, impulsive time-delay systems have been studied in some studies [5,6,7,8,9]. Specially, reference [9] obtained an implicit function related time delay and synchronization rate of impulsive systems. This function was used to reveal the potential influence of delay in continuous dynamical systems on the synchronization of the systems. However, there are few studies on synchronization of hybrid time-delay systems, which is a topic worth investigating.
In addition, impulsive control is an effective method of regulating the synchronization of time-delay systems, as it has a simple structure and makes it possible to change the state of the systems in a fraction of a second and transmit information at discrete times. This feature of impulsive control reduces communication costs to a certain extent and has been used effectively in secure communication, epidemiological model control, and biomedical systems [10,11,12,13]. Unstable or asynchronous systems tend to be stable or synchronous under impulsive control. For instance, reference [14] proposed a class of event-triggered impulsive controllers with time delay for system stability. The synchronization of linear dynamical systems with impulsive delay was discussed in [15] by pinning control. However, the interaction between the system time delay and the impulsive delay can increase the difficulty of analyzing stability and synchronization, and there are fewer relevant studies. The literature [16] concluded that the impulsive delay had a beneficial effect on the stability of time-delay systems through a Razumikhin-type inequality with impulsive delay. Based on the idea of average impulsive delay (AID), reference [17] addressed the influence of impulsive delay on time-delay systems. However, the impulsive control in most researches usually does not take into account delay, i.e., Δz(tu)=z(tu)−z(t−u)=Auz(t−u) [18,19], or just consider "pure" delay, that is, Δz(tu)=f(z(tu−ρu)−) [16,17,20]. We consider a more universal impulsive delay: z(tu)=Duz(t−u)+Euz((tu−ρu)−). Furthermore, many related works imposed strict limitations on impulsive size or impulsive interval length [17,19,21]. For instance, reference [21] required that input delays must be small enough, whereas time delays may even be longer than impulsive intervals. Reference [19] considered more general input delays, but required that the length of impulsive interval is subject to a common upper or lower bound. Therefore, obtaining relaxing criterion for the synchronization of impulsive time-delay systems is crucial.
On the other hand, if the time-delay system is stable or synchronous but the impulses are unstable, we refer to this situation as an impulsive perturbation [22,23]. However, in the vast majority of current studies on impulsive perturbation problems usually in the Lyapunov sense, requires that V(tu)≤exp{σ}V((tu−ρu)−) with σ>0 [17,20]. However, due to the fact that impulsive delays may not be invariable forever, it is more practical to set a flexible σu rather than setting the same threshold at every impulsive moment. Additionally impulsive delays may affect the characteristics of original system [20,24]. It is natural to consider the question: Can we extract the delay information of the impulse term and then integrate it into the rate coefficient σu so to ensure the influence of impulsive perturbation even though a number of σu<0?
Based on the above proposed problems, we employ the Razumikhin-type Lyapunov function to give sufficient conditions for synchronization of the impulsive system with mixed time delays, and the key contributions of this paper are as follows:
1). Differ from [9] proposed an implicit function, we consider the relationship between the system time delay and the rate of synchronization through a display inequality ϵ=q1exp{ρλαˉt−pρ∗+ˆσ∗+ˉt}+q2exp{μλαˉt−pρ∗+ˆσ∗+ˉt} which shows that the system time delay has a positive or negative effect on synchronization in this paper.
2). The size relationship between impulsive delay and impulsive interval length has no strict limitation, which was restricted in [16,20]. In addition, the time delay in continuous dynamics is smaller or greater than impulsive delay.
3). Through the presented concept of average impulsive estimation (AIE), the effect of impulsive perturbation is ensured by integrating the delay information obtained in impulse into impulsive estimation σu, even if some σu are negative. Furthermore, the limitation of having an universal threshold for impulsive estimates at every impulsive point is eliminated. Compared with recent relevant studies [16,17,20,25], the results in this paper are less conservative.
Let R and R+ denote the set of real numbers and positive real numbers, respectively. Denote the set of positive integers by Z+, the set of nonnegative integers by Z0+, the set of n-dimensional real-valued vectors by Rn, and the set of n×m-dimensional real matrices by Rn×m. ‖⋅‖ denotes the vector Euclidean norm. Let PCv:=PC([−v,0],Rn) is the set of piece-wise right-continuous function ϕ:[−v,0]→R+ with the norm ‖ϕ‖v:=sup−v≤θ≤0‖ϕ(θ)‖. D+ℏ(⋅) denotes the upper-right Dini-derivative of ℏ(⋅).
Consider a class of delayed system involved finite distributed delay:
{˙ξ(t)=Cξ(t)+Af(ξ(t))+Bˉf(ξ(t−ρ))+ˆB∫μ0k(s)ξ(t−ρ)ds+J,t≥t0,ξ(t0+s)=ψ(s),s∈[−v,0], | (2.1) |
where ξ(t)=(ξ1(t),⋯,ξn(t))T∈Rn is the state vector, C, A, B, ˆB∈Rn×n are constant system matrices, f(⋅), ˉf(⋅)∈Rn represent the neuron activation functions and f(0)≡0 and ˉf(0)≡0. ρ and μ denote time delay and distributed delay, respectively. k(⋅):[0,μ]→R+ is the delay kernel satisfying ∫μ0k(s)ds=1, ∫μ0sk(s)ds<+∞. v=max{ρ,μ}. Let the initial condition is ψ∈PCv. Denote external input by J.
Consider Eq (2.1) as drive system, then corresponding response system with delayed impulsive control can be
{˙ζ(t)=Cζ(t)+Af(ζ(t))+Bˉf(ζ(t−ρ))+ˆB∫μ0k(s)ζ(t−ρ)ds+J,t≥t0≠tu,ζ(t)−ζ(t−)=H(t),t=tu,ζ(t0+s)=φ(s),s∈[−v,0], | (2.2) |
in which {tu} is a strictly increasing sequence such that limu→+∞tu=+∞ and t0 is the initial time. ζ(t−) and ζ(t+) denote the left limit and right limit at time t, respectively. For this paper, let ζ(t) is right-continuous at every tu, that is, ζ(t+u)=ζ(tu). ρu, u∈Z+, is impulse input delays satisfying 0≤ρu≤tu−tu−1, ρ0=0 and ρu≤v. Let synchronization error is z(t)=ζ(t)−ξ(t) and impulsive control input is H(tu)=Duz(t−u)+Euz((tu−ρu)−)−z(t−u), such that z(tu) can be expressed by
z(tu)=hu(z(0),z((−τu)−))=ζ(tu)−ξ(tu)=ζ(tu)−ζ(t−u)+ζ(t−u)−ξ(tu)=ζ(t−u)+Duz(t−u)+Euz((tu−ρu)−)−z(t−u)−ξ(t−u)=Duz(t−u)+Euz((tu−ρu)−)−z(t−u)+z(t−u)=Duz(t−u)+Euz((tu−ρu)−). |
Then it is easy to obtain error system as follows:
{˙z(t)=Cz(t)+Ag(z(t))+Bˉg(z(t−ρ))+ˆB∫μ0k(s)z(t−ρ)ds,t≠tu,z(t)=Duz(t−)+Euz((t−ρu)−),t=tu,z(t0+s)=φ(s)−ψ(s),s∈[−v,0], | (2.3) |
where g(z(t))=f(ζ(t))−f(ξ(t)), ˉg(z(t−ρ))=ˉf(ζ(t−ρ))−ˉf(ξ(t−ρ)).
Definition 1. [26] Suppose that there are positive numbers N0 and ˉt, such that
ˆt−ˇtˉt−N0≤N(ˆt,ˇt)≤ˆt−ˇtˉt+N0, | (2.4) |
where N(ˆt,ˇt) is the number of impulsive times {tu} occurring in (ˆt,ˇt], ˇt>ˆt>t0. Then ˉt is the average impulsive interval (AII) of impulsive instant sequence {tu} and N0 is the chatter bound.
Definition 2. [27] Suppose that there exist positive numbers ˆρ0 and ρ∗ such that
N(t,t0)∑j=1ρj≤ρ∗N(t,t0)+ˆρ0, | (2.5) |
where N(t,t0) is the number of impulses on the interval (t0,t], then ρ∗ is the AID of impulsive delay sequence {ρu}.
Let H[{tu,ρu}] is the class consisting of impulse time sequence {tu} satisfying AII condition Eq (2.4) and impulsive delay sequence {ρu} satisfying AID condition Eq (2.5).
Definition 3. [28] Response system (2.2) can achieve exponential synchronization with drive system (2.1) if there exist positive scalars χ, λ satisfy
‖δ(t)‖≤χ(‖φ−ψ‖vexp(−λ(t−t0))),∀t≥t0. | (2.6) |
Definition 4. [29] Function V:[t0−v,+∞)×Rn⟶R+ belong to the class V∗ when following conditions are met:
(1). V is continuous on each set [tu−1,tu)×Rnz and lim(t,z)→(t−u,z)V(t,z)=V(t−u,z) exists;
(2). V(t,z) is locally Lipschitz in z and V(t,0)≡0, ∀t∈R+;
(3). V(t,z) satisfies l1‖z‖α≤V(t,z)≤l2‖z‖α, where l1,l2,α are positive scalars.
If V∈V∗ is a locally Lipschitz function, then D+V(t,z(0)) along with the state trajectory of system (2.3) is defined by
D+V(t,z(0))=limr→0+sup1r[V(t+r,z(0)+rf)−V(t,z(0))], |
in which (t,z)∈[t0,+∞)×PCv.
Assumption 1: For any s∈R, z∈R, there exist Lipschitz constants ri>0 and ˉri>0, such that
|fi(s)−fi(z)|≤ri|s−z|,|ˉfi(s)−ˉfi(z)|≤¯ri|s−z|, |
where i=1,2,⋅⋅⋅,n and fi(0)=ˉfi(0)=0.
In this section, we obtain some sufficient conditions for exponential synchronization of systems (2.1) and (2.2).
{From impulsive perturbations point of view, we establish some criteria for exponential synchronization.}
Theorem 1: Considering system (2.3) under Assumption 1. Suppose that there exists a function V∈V∗, scalars p>0, c with p>c>0, q1=∑ni=1rimaxj|gij|, q2=∑ni=1ˉrimaxj|ˆgij|, ϵ=q1exp{cρ}+q2exp{cμ}, σu=ln(∑ni=1maxj|d(u)ij|+∑ni=1maxj|e(u)ij|), Γu=∑ni=1maxj|d(u)ij|exp{σu}, ˆΓu=∑ni=1maxj|e(u)ij|exp{σu}, ˆσ0+>0, and σ∗+>0 such that for every t∈R+, we have
D+V(t,z(0))≤−pV(t,z(0)),wheneverq1V(t−ρ,z(−ρ))+q2∫μ0k(s)V(t−s,z(−s))ds≤ϵV(t,z(0)), t≠tu, | (3.1) |
V(tu,hu(z(0),z((−τu)−)))≤exp{σu}(ΓuV(t−u,z(0−))+ˆΓuV((tu−ρu)−,z((−ρu)−))), | (3.2) |
N(t,t0)∑j=0,σj>0σj≤σ∗+N(t,t0)+ˆσ0+, | (3.3) |
where N(t,t0) is the same as in Definition 1, and
−cˉt+pρ∗+σ∗+<0. | (3.4) |
Then drive system (2.1) can achieve exponential synchronization with response system (2.2) over the class H[{tu,ρu}].
Proof: Define V(t)=V(t,z(t))=‖z(t)‖=∑ni=1|zi(t)|, and V0=supu∈[t0−v,t0]V(u).
The proof is divided into the next three steps.
Step 1: We firstly need to prove that for some t∈[t0,tu) one has
V(t)≤ΩuV0exp(−c(t−t0)), t∈[t0,tu), u∈Z+, | (3.5) |
where Ωu=exp{∑u−1j=0,pρj+σj>0(pρj+σj)}.
In order to prove Eq (3.5), we construct the function
Λ(t)={V(t)exp{c(t−t0)}, t≥t0,V(t), t0−v≤t≤t0. |
From Eq (3.5), it yields that
Λ(t)≤ΩuV0,t∈[t0,tu), u∈Z+. | (3.6) |
We will show that Eq (3.6) holds for u=1, i.e.,
Λ(t)≤Ω1V0=V0, t∈[t0,t1). | (3.7) |
It is easy for us to get Λ(t)≤V0 for t∈[t0−v,t0], which indicates that Λ(t0)≤V0. If (3.7) is not true, then there exists t∗∈(t0,t1) such that Λ(t∗)>V0, Λ(t)≤V0 for t∈(t0−v,t∗) and D+Λ(t)|t=t∗≥0. Obviously, Λ(t∗)>Λ(t) for t∈(t∗−μ,t∗), then we have
q1V(t∗−ρ)+q2∫μ0k(s)V(t∗−s)ds=n∑i=1rimaxj|bij|V(t∗−ρ)+n∑i=1ˉrimaxj|ˆbij|∫μ0k(s)V(t∗−s)ds<n∑i=1rimaxj|bij|exp{cρ}V(t∗)+n∑i=1ˉrimaxj|ˆbij|∫μ0k(s)exp{cs}V(t∗)ds≤n∑i=1rimaxj|bij|exp{cρ}V(t∗)+n∑i=1ˉrimaxj|ˆbij|∫μ0k(s)exp{cμ}V(t∗)ds=n∑i=1rimaxj|bij|exp{cρ}V(t∗)+n∑i=1ˉrimaxj|ˆbij|exp{cμ}V(t∗)=ϵV(t∗). |
From Eq (3.1) we can obtain
D+V(t)|t=t∗≤−pV(t∗). | (3.8) |
Hence, there is
D+Λ(t)|t=t∗=[D+V(t)|t=t∗+cV(t∗)]exp{c(t∗−t0)}≤(c−p)V(t∗)exp{c(t∗−t0)}<0, |
which contradicts D+Λ(t)|t=t∗≥0. Then, provided that (3.6) is true when t∈[t0,tm), 1≤m≤u−1, that is Λ(t)≤ΩmV0, t∈[t0,tm). Next, we will show that (3.6) is true when t∈[t0,tm+1), i.e., we need to prove that Λ(t)≤Ωm+1V0, for t∈[tm,tm+1).
When t=tm, from (3.2) we can get
Λ(tm)=V(tm)exp{c(tm−t0)}=n∑i=1|δi(t)|exp{c(tm−t0)}≤exp{σm}(n∑i=1maxj|d(u)ij|exp(σu)V(t−m)+n∑i=1maxj|e(u)ij|exp(σu)V((tm−ρm)−))exp{c(tm−t0))}=exp{σm}(ΓmV(t−m)+ˆΓm(V(tm−ρm)−))exp{c(tm−t0)}≤exp{σm}(ΓmV(t−m)exp{c(tm−t0)}+ˆΓmV((tm−ρm)−)exp{c(tm−ρm−t0)}exp{cρm})≤exp{σm}(ΓmΩmV0+ˆΓmΩmV0exp(cρm))≤exp{σm+pρm}ΩmV0≤Ωm+1V0. |
Therefore, Eq (3.6) holds for t=tm. Provided that Eq (3.6) is not true for t∈(tm,tm+1), then there exists t∗∈(tm,tm+1) has Λ(t∗)>Ωm+1V0, Λ(t)≤Ωm+1V0 for t∈(t0−v,t∗) and D+Λ(t)|t=t∗≥0. Similar with the argument used in Eq (3.7), we can obtain D+Λ(t)|t=t∗<0, it contradicts D+Λ(t)|t=t∗≥0. Hence, we can found Eq (3.6) holds through using mathematical induction, in which t∈[t0,tu), which implies Eq (3.5) holds for t∈[t0,tu), u∈Z+.
Step 2: According to Eq (3.5), condition Eqs (2.4), (2.5) and (3.3), we have
V(t)≤V0exp{−c(t−t0)}⋅exp{u−1∑j=0,pρj+σj>0(pρj+σj)}≤V0exp{−c(t−t0)+pu−1∑j=0ρj+u−1∑j=0,σj>0σj}≤V0exp{−c(t−t0)+p(ρ∗(u−1)+ˆρ0)+σ∗+(u−1)+ˆσ0+}≤V0exp{−c(t−t0)+p(ρ∗N(t,t0)+ˆρ0)+σ∗+N(t,t0)+ˆσ0+}≤V0exp{−c(t−t0)+p(ρ∗(t−t0)ˉt+N0ρ∗+ˆρ0)+σ∗+(t−t0)ˉt+N0σ∗++ˆσ0+}≤V0exp{pN0ρ∗+pˆρ0+N0σ∗++ˆσ0+}exp{(−c+pρ∗+σ∗+ˉt)(t−t0)}, | (3.9) |
where t≥t0.
Step 3: Based on condition Eqs (3.4), (3.9) and Assumption l1‖z‖α≤V(t,z)≤l2‖z‖α, which can derive that
‖z(t)‖≤χ(‖φ−ψ‖vexp(−λ(t−t0))), ∀t≥t0, |
where χ=(l2l1exp{pN0ρ∗+pˆρ0+N0σ∗++ˆσ0+})1α, λ=cˉt−pρ∗−ˆσ∗+αˉt, which implies the system (2.2) can be exponentially synchronized with System (2.1) over the class H[{tu,ρu}].
Remark 1: System under consideration with hybrid delayed impulses is discussed in Theorem 1 using the Lyapunov–Razumikhin method. Condition Eq (3.1) describes the continuous evolution of the considered system, and it follows from a>0 that the continuous dynamics are stabilizing. Condition Eq (3.2) overviews the impulsive effect. In the case where Γu=0, ˆΓu=1, Eq (3.2) simplifies to U(tu,h(z))≤exp{d}U(π,z), where π=tu−ρu [20]. In the case where Γu=1, ˆΓu=0, Eq (3.2) simplifies to U(tu,hu(z(0)))≤exp{d}U(t−u,z(0)) [7].
Remark 2: According to the derivation condition ϵ=q1exp{cρ}+q2exp{cμ} of Theorem 1, it can be learnt that the parameter c will increase with the decrease of ρ and μ. Furthermore, the synchronization rate λ=cˉt−pρ∗−ˆσ∗+αˉt will increase. Therefore we have ϵ=q1exp{ρλαˉt−pρ∗+ˆσ∗+ˉt}+q2exp{μλαˉt−pρ∗+ˆσ∗+ˉt}. This means that, in some cases, system delays ρ might have potentially negative impact on the synchronization between the systems (2.1) and (2.2).
Remark 3: According to condition Eq (3.2), σu is called as impulsive estimate. In order to analyze the function of the impulsive estimation sequence {σu}, the concepts of average positive impulsive estimation (APIE) and AIE are introduced in this paper.
Assuming the existence of scalars ˆσ0+>0 and σ∗+>0 satisfying condition Eq (3.3), σ∗+ is called as APIE.
Supposing that there are some ˆσ0>0 and σ∗>0 such that
σ∗N(t,t0)−ˆσ0≤N(t,t0)∑j=1σj≤σ∗N(t,t0)+ˆσ0, | (3.10) |
where N(t,t0) is given in Definition 2, thus σ∗ is referred to as AIE of the impulsive estimation sequence {σu}. When σu>0, there is σ∗=σ∗+. Since t0 does not act as an impulse point, we assume that σ0=0.
Remark 4: Actually, there exist some results about impulsive delays. In [30,31], the time delays had to have strict upper and lower bounds, or to be smaller than the length of impulse interval [16,20]. Even if [19,21,32] relaxed impulsive delays, impulsive interval should meet that infu∈Z+{tu−tu−1}≥ϱ or supu∈Z+{tu−tu−1}≤ϱ for some ϱ≥0. The length of impulsive interval is flexible in Theorem 1.
When the delays of the continuous dynamics are not to be considered, systems (2.1) and (2.2) can be represented as
{˙ξ(t)=Cξ(t)+Af(ξ(t))+Bˉf(ξ(t))+J,t≠tu,ξ(t0+s)=ψ(s),s∈[−v,0], | (3.11) |
{˙ζ(t)=Cζ(t)+Af(ζ(t))+Bˉf(ζ(t))+J,t≠tu,ζ(t)−ζ(t−)=H(t),t=tu,ζ(t0+s)=φ(s),s∈[−v,0], | (3.12) |
and the resulting error system for systems (1) and (2) follows as
{˙z(t)=Cz(t)+Ag(z(t))+Bˉg(z(t)),t≠tu,z(t)=Duz(t−)+Euz((t−ρu)−),t=tu,z(t0+s)=φ(s)−ψ(s),s∈[−v,0], | (3.13) |
Corollary 1: Suppose that Assumption 1 holds. If there exists a function V∈V∗, scalars p>0, c with p>c>0, σu=ln(∑ni=1maxj|d(u)ij|+∑ni=1maxj|e(u)ij|), σu with pρu+σu>0, σ∗>0 satisfying condition Eq (3.10), Γu=∑ni=1maxj|d(u)ij|exp(σu), ˆΓu=∑ni=1maxj|e(u)ij|exp(σu). In that case the exponential synchronization of systems (3.11) and (3.12) can be achieved while Eq (3.1) holds with ϵ=q1=q2=0 and Eq (3.2) holds, and
−cˉt+pρ∗+σ∗<0. |
Remark 5: Corollary 1 offers some criteria for exponential synchronization between system (2.1) and system (2.2) from the point of view of impulsive perturbation, which lowers the limitation on σu>0. Most of previous works [17,20,25] need σu<0 in the impulsive control case and σu>0 in the impulsive perturbation case. Corollary 1 presents condition pρu+σu>0 which makes σu is flexible. If σu>0, pρu+σu>0 always holds. If σu<0, we just need ρu>−σup which ensures above condition to hold. It's worth noting that, in the impulsive perturbation problem, the smaller σu is, ρu must be larger to compensate.
In this subsection, from the perspective of impulsive control, we establish a number of criteria of exponential synchronization based on the concepts of AID and AIE. Moreover, we assume that tu−tu−1≥ρ>ρu and tu−tu−1≥η>ρu, u∈Z+.
Theorem 2: Considering system (2.3) under Assumption 1. Suppose that there exists a function V∈V∗, scalars b1=maxicii+∑ni=1maxj,j≠i|cij|+∑ni=1maxj|aij|, b2=∑ni=1rimaxj|gij| and b3=∑ni=1ˉrimaxj|ˆgij|, σu=−ln(∑ni=1maxj|d(u)ij|+∑ni=1maxj|e(u)ij|) with ˉσ=supu∈Z+σu>0, γ>0 such that γ>b1+b2exp{ˉσ}+b3exp{ˉσ}, Γu=∑ni=1maxj|d(u)ij|exp{σu}, ˆΓu=∑ni=1maxj|e(u)ij|exp{σu}, the impulsive estimation sequence {σu} satisfies the condition Eq (3.10), for every t>0, the following inequalities hold:
D+V(t,z(0))≤b1V(t,z(0))+b2V(t−ρ,z(−ρ))+b3∫μ0k(s)V(t−s,z(−s))ds,t≠tu, |
V(tu,hu(−z(0),z(−τu))≤exp(σu)(ΓuV(t−u,z(0))+ˆΓuV((tu−ρu)−,z(−ρu))), |
γˉt−σ∗<0. |
Then, drive system (2.1) can achieve exponential synchronization with response system (2.2) over the class H[{tu,ρu}].
Proof: Define V(t)=V(t,z(t))=‖z(t)‖=∑ni=1|zi(t)|, and V0=supu∈[t0−v,t0]V(u), such that
D+V(t)=D+‖z(t)‖=D+n∑i=1|zi(t)|≤(maxicii+n∑i=1maxj,j≠i|cij|+n∑i=1maxj|aij|)V(t)+n∑i=1rimaxj|gij|V(t−ρ)+n∑i=1ˉrimaxj|ˆgij|∫μ0k(s)V(t−s)ds=b1V(t,z(0))+b2V(t−ρ,z(−ρ))+b3∫μ0k(s)V(t−s,z(−s))ds, t≠tu. |
The proof is divided into three steps.
Step 1 : We will demonstrate that
\begin{align} \begin{aligned} V(t)\le \Omega _ u V_0\exp\left\lbrace \gamma (t-t_0)\right\rbrace , \end{aligned} \end{align} | (3.14) |
where \Omega _ u = \exp\left \{ -\sum_{j = 0}^{ u}\sigma _j \right \} .
First, we need show that the following two situations for u\in \mathcal{Z} ^0_+ .
(ⅰ). If t^\bigtriangleup\in\left[t_0, t_1 \right) , one has
\begin{align} \begin{aligned} \Theta (s)\le \Theta (t^\bigtriangleup), \ \ t_0-v\le s\le t^\bigtriangleup. \end{aligned} \end{align} | (3.15) |
(ⅱ). If t^\bigtriangleup\in\left[t_ u, t_{ u-1} \right) , u\in Z_+ , one can obtain that
\begin{align} \begin{aligned} \Theta (s)\le \Theta (t^\bigtriangleup), \ \ t_ u\le s\le t^\bigtriangleup, \end{aligned} \end{align} | (3.16) |
and
\begin{align} \begin{aligned} \Theta (s)\exp\left\lbrace \gamma(t_ u-t_{ u-1})\right\rbrace \le \Theta (t^\bigtriangleup)\exp\left\lbrace \bar\sigma\right\rbrace , \ \ t_{ u-1}\le s\le t_ u. \end{aligned} \end{align} | (3.17) |
Then D^+\Theta (t)|_{t = t^\bigtriangleup} < 0 , where
{\Theta}(t) = \left\{ \begin{aligned} &{V}\exp\left\lbrace -\gamma(t-t_ u)\right\rbrace , \ \ t\in\left [ t_ u, t_{ u+1}\right ) , u\in \mathcal{Z} _+, \\ &{V(t)}, \ \ t_0-v\le t\le t_0. \end{aligned} \right. |
Construct an auxiliary function with \iota > 0
{\Theta}_\iota(t) = \left\{ \begin{aligned} &{V}\exp\left\lbrace -(\gamma\!+\iota)(t-t_ u)\right\rbrace , &&t\in\left [ t_ u, t_{ u+1}\right ) , u\in \mathcal{Z} _+, \\ &{V(t)}, \ \ t_0-v\le t\le t_0. \end{aligned} \right. |
Without loss of generality, we assume \rho\geq\eta . First, provided that Eq (3.15) holds.
If t^\bigtriangleup-\eta\geq t^\bigtriangleup-\rho\geq t_0 , based on Eq (3.16), one has
\begin{aligned} &\exp\left\lbrace \iota(t^\bigtriangleup-t_0)\right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup} \\ & = \left( D^+V (t)|_{t = t^\bigtriangleup}-(\gamma+\iota)V(t^\bigtriangleup) \right) \exp\left\lbrace -\gamma (t^\bigtriangleup-t_0)\right\rbrace \\ &\leq(b_1-\gamma -\iota )V(t^\bigtriangleup)\exp\left\lbrace -\gamma(t^\bigtriangleup-t_0) \right\rbrace +b_2V(t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma (t^\bigtriangleup-\rho-t_0)\right\rbrace \exp\left\lbrace -\gamma \rho\right\rbrace \\ &\quad +b_3\int_{0}^{\mu } k(s)V(t^\bigtriangleup-s)\exp(-\gamma (t^\bigtriangleup-s-t_0))\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma -\iota )\Theta (t^\bigtriangleup)+b_2\Theta (t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma \rho\right\rbrace +b_3\int_{0}^{\mu } k(s)\Theta (t^\bigtriangleup-s)\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma-\iota+b_2+b_3)\Theta(t^\bigtriangleup). \end{aligned} |
If t^\bigtriangleup-\rho < t_0 < t^\bigtriangleup-\eta or t^\bigtriangleup-\rho\leq t^\bigtriangleup-\eta < t_0 , we can derive similarly that
\begin{aligned} &\exp\left\lbrace \iota(t^\bigtriangleup-t_0) \right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup} \\ &\leq(b_1-\gamma -\iota )V(t^\bigtriangleup)\exp\left\lbrace -\gamma(t^\bigtriangleup-t_0) \right\rbrace +b_2V(t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma (t^\bigtriangleup-t_0)\right\rbrace \\ &\quad +b_3\int_{0}^{\mu } k(s)V(t^\bigtriangleup-s)ds\exp\left\lbrace -\gamma (t^\bigtriangleup-t_0)\right\rbrace \\ &\leq(b_1-\gamma-\iota+b_2+b_3)\Theta(t^\bigtriangleup). \end{aligned} |
Next, suppose that Eqs (3.16) and (3.17) hold.
If t^\bigtriangleup-\eta\ge t^\bigtriangleup-\rho\ge t_ u , using Eq (3.16) we can found
\begin{aligned} &\exp\left\lbrace \iota(t^\bigtriangleup-t_ u)\right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup} \\ &\leq(b_1-\gamma -\iota )V(t^\bigtriangleup)\exp\left\lbrace -\gamma(t^\bigtriangleup-t_ u)\right\rbrace +b_2V(t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma (t^\bigtriangleup-\rho-t_ u)\right\rbrace \exp\left\lbrace -\gamma \rho\right\rbrace \\ &\quad +b_3\int_{0}^{\mu } k(s)V(t^\bigtriangleup-s)\exp\left\lbrace -\gamma (t^\bigtriangleup-s-t_ u)\right\rbrace \exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma -\iota )\Theta (t^\bigtriangleup)+b_2\Theta (t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma \rho\right\rbrace +b_3\int_{0}^{\mu } k(s)\Theta (t^\bigtriangleup-s)\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma-\iota+b_2+b_3)\Theta(t^\bigtriangleup). \end{aligned} |
If t^\bigtriangleup-\rho < t_ u\leq t^\bigtriangleup-\eta , by the fact t_{ u-1}\leq t^\bigtriangleup-\rho < t_ u , Eqs (3.16) and (3.17), it leads to
\begin{aligned} &\exp\left\lbrace \iota(t^\bigtriangleup-t_0) \right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup} \\ &\leq(b_1-\gamma -\iota )V(t^\bigtriangleup)\exp\left\lbrace -\gamma(t^\bigtriangleup-t_ u) \right\rbrace +b_2V(t^\bigtriangleup-\rho)\exp\left\lbrace -\gamma (t^\bigtriangleup-\rho-t_{ u-1})\right\rbrace \exp\left\lbrace \gamma(t_ u-t_{ u-1}-\rho)\right\rbrace \\ &\quad +b_3\int_{0}^{\mu } k(s)V(t^\bigtriangleup-s)\exp\left\lbrace -\gamma (t^\bigtriangleup-s-t_ u)\right\rbrace \exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma -\iota )\Theta (t^\bigtriangleup)+b_2\Theta (t^\bigtriangleup-\rho)\exp\left\lbrace \gamma(t_ u-t_{ u-1}-\rho)\right\rbrace +b_3\int_{0}^{\mu } k(s)\Theta (t^\bigtriangleup-s)\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma-\iota+b_2\exp\left\lbrace \bar\sigma\right\rbrace +b_3)\Theta(t^\bigtriangleup). \end{aligned} |
If t^\bigtriangleup-\rho \leq t^\bigtriangleup-\eta < t_ u , because of t_{ u-1}\leq t^\bigtriangleup-\rho < t_ u , Eqs (3.16) and (3.17), one has
\begin{aligned} &\exp\left\lbrace \iota(t^\bigtriangleup-t_0) \right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup} \\ &\leq(b_1-\gamma -\iota )\Theta(t^\bigtriangleup)+b_2\Theta(t^\bigtriangleup-\rho)\exp\left\lbrace \gamma(t_ u-t_{ u-1}-\rho)\right\rbrace \\ &\quad+b_3\int_{0}^{\mu } k(s)V(t^\bigtriangleup-s)\exp\left\lbrace -\gamma (t^\bigtriangleup-s-t_ u)\right\rbrace )\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\leq(b_1-\gamma -\iota )\Theta (t^\bigtriangleup)+b_2\Theta (t^\bigtriangleup-\rho)\exp\left\lbrace \gamma(t_ u-t_{ u-1}-\rho)\right\rbrace +b_3\int_{0}^{t^\bigtriangleup-t_ u} k(s)\Theta (t^\bigtriangleup-s)\exp\left\lbrace -\gamma s\right\rbrace ds \\ &\quad+b_3\int_{t^\bigtriangleup-t_ u}^{\eta} k(s)\Theta (t^\bigtriangleup-s)\exp\left\lbrace \gamma (t_ u-t_{ u-1}-s)\right\rbrace ds \\ &\leq(b_1-\gamma-\iota+b_2\exp\left\lbrace \bar\sigma\right\rbrace) \Theta(t^\bigtriangleup)+b_3\exp\left\lbrace \bar\sigma\right\rbrace \Theta(t^\bigtriangleup)\left( \int_{0}^{t^\bigtriangleup-t_ u} k(s)ds+\int_{t^\bigtriangleup-t_ u}^{\eta} k(s)ds\right) \\ &\leq(b_1-\gamma-\iota+b_2\exp\left\lbrace \bar\sigma\right\rbrace +b_3\exp\left\lbrace \bar\sigma\right\rbrace )\Theta(t^\bigtriangleup). \end{aligned} |
According to the above situations and \gamma > b_1 +b_2\exp(\bar\sigma)+b_3\exp(\bar\sigma) , one can obtain that
\begin{aligned} \exp\left\lbrace \iota (t^\bigtriangleup-t_ u)\right\rbrace D^+\Theta _\iota (t)|_{t = t^\bigtriangleup}&\le ( b_1-\gamma -\iota+b_2\exp\left\lbrace \bar\sigma\right\rbrace +b_3\exp\left\lbrace \bar\sigma\right\rbrace) \Theta (t^\bigtriangleup) \\ & < -\iota\Theta (t^\bigtriangleup). \end{aligned} |
It can be further deduced that
\begin{aligned} D^+\Theta (t) |_{t = t^\bigtriangleup} = &\exp\left\lbrace \iota (t^\bigtriangleup-t_ u)\right\rbrace D^+\Theta_\iota (t) |_{t = t^\bigtriangleup}+\iota \exp\left\lbrace \iota (t^\bigtriangleup-t_ u)\right\rbrace \Theta_\iota (t) |_{t = t^\bigtriangleup} \\ < &\iota\Theta (t^\bigtriangleup)-\iota\Theta (t^\bigtriangleup) \\ = &0. \end{aligned} |
Then, we shall show that
\begin{align} \begin{aligned} \Theta (t)\le \Omega _ u V_0\exp(\gamma (t_ u-t_0)), t\in\left [ t_ u, t_{ u-1}\right ) , u\in \mathcal{Z} ^+. \end{aligned} \end{align} | (3.18) |
We can easily get \Theta(t)\le V_0 when t\in\left[t_0-v, t_0\right] , so that \Theta(t_0)\le V_0 . Suppose Eq (3.18) is false for u = 0, then there exists t_\bigtriangledown\in(t_0, t_1) makes \Theta(t_\bigtriangledown) > V_0 , \Theta(t)\leq V_0 for t\in (t_0-v, t_\bigtriangledown) and D^+\Theta(t)|_{t = t_\bigtriangledown}\geq0 , which is contrary to D^+\Theta(t)|_{t = t_\bigtriangledown} < 0 in previous discussion. Provided that (3.18) holds for u\leq U , next we will prove that Eq (3.18) holds for u = U+1 . Based on \Theta(t)\leq\Omega_ U V_0exp\left\lbrace \gamma(t_ U-t_0)\right\rbrace , t\in\left[t_ U, t_{ U+1}\right) , we have
\begin{aligned} \Theta (t_{ U+1}) = &V(t_{ U+1}) \\ \leq&\exp\left\lbrace -\sigma _{ U+1}\right\rbrace \left( \Gamma _{ U+1}V(t^-_{ U+1})+\hat\Gamma _{ U+1}V((t_{ U+1}-\rho_{ U+1})^-)\right) \\ \leq&\exp\left\lbrace -\sigma _{ U+1}\right\rbrace \left( \Gamma _{ U+1}\Omega _ U V_0\exp\left\lbrace \gamma (t_{ U+1}-t_0)\right\rbrace +\hat\Gamma _{ U+1}\Omega _ U V_0\exp\left\lbrace \gamma (t_{ U+1}-\rho_{ U+1}-t_0)\right\rbrace \right) \\ \leq&\exp\left\lbrace -\sigma _{ U+1}\right\rbrace \Omega _ U V_0\exp\left\lbrace \gamma (t_{ U+1}-t_0)\right\rbrace \\ \leq&\Omega _{ U+1}V_0\exp\left\lbrace \gamma (t_{ U+1}-t_0)\right\rbrace . \end{aligned} |
Therefore, Eq (3.18) holds for t = t_{ U+1} . Suppose that there are some t\in \left(t_{ U+1}, t_{ U+2} \right) which leads to \Theta(t)\geq\Omega_ U V_0\exp\left\lbrace \gamma(t_ u-t_0)\right\rbrace , through the continuity of V(t) in \left(t_{ U+1}, t_{ U+2} \right) , we can get \hat t\in\left(t_{ U+1}, t_{ U+2} \right) such that \Theta(\hat t) = \Omega_{ U+1}V_0\exp\left\lbrace \gamma(t_{ U+1}-t_0)\right\rbrace , \Theta(t) < \Omega_{ U+1}V_0\exp\left\lbrace \gamma(t_{ U+1}-t_0)\right\rbrace , t\in\left(t_{ U+1}, \hat t \ \right) and D^+\Theta(t)|_{t = \hat t}\geq0 . When s\in\left[t_{ U}, t_{ U+1} \right) , it leads to
\begin{aligned} \Theta& (s)\exp\left\lbrace \gamma(t_{ U+1}-t_ U)\right\rbrace \\\leq&\Omega _{ U}V_0\exp\left\lbrace \gamma (t_{ U}-t_0)\right\rbrace \exp\left\lbrace \gamma (t_{ U+1}-t_ U)\right\rbrace \\ = &\Omega _{ U}V_0\exp\left\lbrace \gamma (t_{ U+1}-t_0)\right\rbrace \\ = &\Omega _{ U+1}\exp\left\lbrace \sigma_{ U+1}\right\rbrace V_0\exp\left\lbrace \gamma (t_{ U+1}-t_0)\right\rbrace \\ = &\Theta(\hat t)\exp\left\lbrace \bar\sigma\right\rbrace . \end{aligned} |
Thus, it follows from Eqs (3.16) and (3.17) that D^+\Theta(t)|_{t = \hat t} < 0 , which is a contradiction with D^+\Theta(t)|_{t = \hat t}\geq0 .
Then we can conclude that Eq (3.18) is true through using mathematical induction.
Step 2 : From Eqs (3.8) and (3.10), it can be derived that
\begin{align} \begin{aligned} V(t)\leq&V_0exp\left\lbrace -\sum\limits_{j = 0}^{N(t, t_0)}\sigma _j \right\rbrace \exp\left\lbrace \gamma (t-t_0)\right\rbrace \\ \leq&V_0\exp\left\lbrace -(\sigma ^*N(t, t_0)-\hat\sigma^0)\right\rbrace \exp\left\lbrace \gamma (t-t_0)\right\rbrace \\ \leq&V_0\exp\left\lbrace -\left( \sigma ^*\left( \frac{t-t_0}{\bar t}-N_0 \right) -\hat\sigma^0\right) \right\rbrace \exp\left\lbrace \gamma (t-t_0)\right\rbrace \\ \leq&V_0\exp\left\lbrace \left( \gamma -\frac{\sigma ^*}{\bar t} \right) (t-t_0 )\right\rbrace \exp\left\lbrace \sigma ^*N_0+\hat \sigma ^0\right\rbrace . \end{aligned} \end{align} | (3.19) |
Based on condition \gamma\bar t-\sigma^* < 0 , inequality (3.19) and Assumption l_1\left \| z \right \|^\alpha \le V(t, z)\le l_2\left \| z \right \|^\alpha , one has
\begin{aligned} \left \| z (t)\right \| \le \chi \left \| \varphi -\psi \right \|_v\exp(-\lambda (t-t_0)) , \ \ t\ge t_0, \end{aligned} |
where \chi = \left(\frac{l_2}{l_1}\exp\left\lbrace \sigma ^*N_0+\hat \sigma ^0\right\rbrace \right) ^{\frac{1}{\alpha } } , \lambda = \frac{\sigma^*-\gamma\bar t}{\alpha \bar t} , which implies the system (2.2) can be exponentially synchronized with system (2.1) over the class \mathscr{H}\left [\left \{ t_m, \rho_m \right \} \right] .
Remark 6: Theorem 2 presents some criteria for exponential synchronization between systems (2.1) and (2.2) from the point of view of impulsive control. Compared to previous studies [16,17,18,20], which have a common threshold for \sigma > 0 at each impulsive point, the results in this paper are less conservative and the rate coefficient \sigma_ u is flexible here through the proposed concept of AIE.
In this section, we give illustrative examples to show the effectiveness of the obtained results.
Example 1: We consider the error system (2.3) with parameters as
C = \begin{bmatrix} -0.9 &0.01 \\ 0.02&-0.9 \end{bmatrix}, \ A = \begin{bmatrix} 0.11 &-0.15 \\ -0.2&0.1 \end{bmatrix}, B = \begin{bmatrix} 0.1 &0 \\ 0&0.1 \end{bmatrix}, \hat B = \begin{bmatrix} 0.29 &-0.31 \\ -0.32&0.28 \end{bmatrix}, |
k(s) = \frac{1}{\mu } , 0\leq\rho_ u\leq\bar\rho , v: = \max\left \{ \bar\rho, \rho, \mu \right \} , f(\cdot) = \bar f(\cdot) = 0.3tanh(\cdot) , \rho = 0.1 , \mu = 0.4 , t_ u = 0.7 u and
\begin{aligned} &\rho_ u = &\left\{ \begin{aligned} &{0}, && u = 3 m-2, \\ &{0.1}, && u = 3 m-1, \\ &{1.1}, && u = 3 m, \end{aligned} \right. \end{aligned} |
where m\in\mathcal{Z} _+ , the matrices D_ u , E_ u can be chosen by
\ \ \ \ D_ u = \begin{bmatrix} 0.1 & 0 \\ 0 & 0.1 \end{bmatrix} , E_ u = \begin{bmatrix} 0.85 & 0 \\ 0 & 0.85 \end{bmatrix} , when \rho_ u = 0 ;
\ \ \ \ D_ u = \begin{bmatrix} 0.28 & 0 \\ 0 & 0.28 \end{bmatrix} , E_ u = \begin{bmatrix} 0.85 & 0 \\ 0 & 0.85 \end{bmatrix} , when \rho_ u = 0.1 ;
\ \ \ \ D_ u = \begin{bmatrix} 0.35 & 0 \\ 0 & 0.35 \end{bmatrix} , E_ u = \begin{bmatrix} 0.45 & 0 \\ 0 & 0.45 \end{bmatrix} , when \rho_ u = 1.1 .
Let V(t) = \left \| z (t) \right \| in system (2.3), c = 0.2 , \epsilon = 0.7519 , it can be derived that p = 0.2641 , and
\begin{aligned} &\sigma_ u = &\left\{ \begin{aligned} &{-0.0512}, && u = 3 m-2, \\ &{0.1222}, && u = 3 m-1, \\ &{-0.2231}, && u = 3 m. \end{aligned} \right. \end{aligned} |
Therefore, we can get \bar t = 0.7 , \rho^* = 0.4 , \sigma_+^* = 0.0407 . Furthermore it can be obtained that -c\bar t+p\rho^*+\sigma_+^* = -0.08166 < 0 . By Theorem 1, drive system (2.1) can achieve exponential synchronization with response system (2.2) over the class \mathscr{H}\left [\left \{ t_ u, \rho_ u \right \} \right] , see Figure 1.
If the delays \rho , \mu and the rate coefficient \rho are chosen as \rho = 0.8 , \mu = 1.4 and \rho = 0.3122 . We can find that all conditions in Theorem 1 are satisfied so that drive system (2.1) can achieve exponential synchronization with response system (2.2), see Figure 2. According to Remark 3, synchronization rate drops as the delay \rho or \mu grows, which agrees well with the simulation result in Figure 2.
If D_ u , E_ u are selected as
\ \ \ \ \ D_ u = \begin{bmatrix} 0.36 & 0 \\ 0 & 0.35 \end{bmatrix} , E_ u = \begin{bmatrix} 1.31 & 0 \\ 0 & 1.46 \end{bmatrix} , when \rho_ u = 0 ;
\ \ \ \ D_ u = \begin{bmatrix} 0.51 & 0 \\ 0 & 0.51 \end{bmatrix} , E_ u = \begin{bmatrix} 0.86 & 0 \\ 0 & 0.86 \end{bmatrix} , when \rho_ u = 0.1 ;
\ \ \ \ D_ u = \begin{bmatrix} 0.31 & 0 \\ 0 & 0.31 \end{bmatrix} , E_ u = \begin{bmatrix} 0.82 & 0 \\ 0 & 0.82 \end{bmatrix} , when \rho_ u = 1.1 ,
where u\in Z_+ , then by calculation we have
\begin{aligned} &\sigma_ u = &\left\{ \begin{aligned} &{0.5988}, && u = 3 m-2, \\ &{0.3148}, && u = 3 m-1, \\ &{0.1222}, && u = 3 m, \end{aligned} \right. \end{aligned} |
and \sigma_ u = 0.3452 . Under this circumstance, -c\bar t+p\rho^*+\sigma_+^* = 0.2228 > 0 , this is contrary to condition Eq (3.4). Therefore, system (2.2) may not be able to achieve exponential synchronization with system (2.1), see Figure 3.
Example 2: We consider the following error system:
\begin{align} \begin{aligned} \begin{cases} \dot{ z}(t) = 0.3 z (t) +0.03g( z (t))+0.08\int_{0}^{\mu }k(s) z (t-\rho )ds, \ \ t\neq t_ u, \\ z(t _ u) = d_ u z (t_ u^{-} )+e_ u z ((t_ u- \rho _ u)^{-} ), \ \ t = t_ u, \\ \delta(t_0+s) = \varphi(s)-\psi(s) , \ \ s\in \left [ -v, 0 \right ], \end{cases} \end{aligned} \end{align} | (4.1) |
where k(s) = \frac{1}{1-e^{-\mu }} e^{-s} , \rho = 0.8 , \mu = 0.4 , t_ u = u and \tau_ u , c_ u , d_ u can be selected by
\begin{aligned} &\rho_ u = &\left\{ \begin{aligned} &{0}, && u = 3 m-2, \\ &{0.1}, && u = 3 m-1, \\ &{0.9}, && u = 3 m, \end{aligned} \right. \end{aligned}\ \ \ \begin{aligned} &d_ u = &\left\{ \begin{aligned} &{0.15}, &&\rho_ u = 0, \\ &{0.1}, &&\rho_ u = 0.1, \\ &{0.15}, &&\rho_ u = 0.9, \end{aligned} \right. \end{aligned} \ \ \begin{aligned} &e_ u = &\left\{ \begin{aligned} &{0.5}, &&\rho_ u = 0, \\ &{0.3}, &&\rho_ u = 0.1, \\ &{0.9}, &&\rho_ u = 0.9. \end{aligned} \right. \end{aligned} |
Let V(t) = \left \| z (t) \right \| in system (4.1), there are
\begin{aligned} &\sigma_ u = &\left\{ \begin{aligned} &{0.4307}, && u = 3 m-2, \\ &{0.916}, && u = 3 m-1, \\ &{-0.0487}, && u = 3 m, \end{aligned} \right. \end{aligned} |
and \sigma^* = 0.4326 , \bar t = 1 . Then we can choose \gamma = 0.42 > 0.3+0.03+0.08 makes condition \gamma\bar t-\sigma^* = -0.0126 < 0 is established. From Theorem 1, drive system (2.1) can achieve exponential synchronization with response system (2.2) over the class \mathscr{H}\left [\left \{ t_ u, \rho_ u \right \} \right] , see Figure 4 (solid red line). It is clear that, owing to the idea of AIE, it is not necessary that \sigma_ u is positive for every u\in \mathcal{Z} _+ .
If d_ u , e_ u are
\begin{aligned} &d_ u = &\left\{ \begin{aligned} &{0.44}, &&\rho_ u = 0, \\ &{0.5}, &&\rho_ u = 0.1, \\ &{0.4}, &&\rho_ u = 0.9, \end{aligned} \right. \end{aligned}\ \ \ \ \begin{aligned} &e_ u = &\left\{ \begin{aligned} &{0.4}, &&\rho_ u = 0, \\ &{0.4}, &&\rho_ u = 0.1, \\ &{0.5}, &&\rho_ u = 0.9, \end{aligned} \right. \end{aligned} |
then it can be figured out that
\begin{aligned} &\sigma_ u = &\left\{ \begin{aligned} &{0.1743}, && u = 3 m-2, \\ &{0.1053}, && u = 3 m-1, \\ &{0.0943}, && u = 3 m, \end{aligned} \right. \end{aligned} |
therefore, \sigma^* = 0.1246 . Under this circumstance, \gamma\bar t-\sigma^* = 0.2954 > 0 , which contradicts one of the conditions of Theorem 2, i.e., \gamma\bar t-\sigma^* < 0 . Consequently, system (2.2) may not be able to achieve exponential synchronization with system (2.1), see Figure 4 (dotted blue line).
In this paper, we have explored the issue of synchronization for a class of impulsive systems with discrete and distributed delay. Sufficient Lyapunov conditions for the synchronization of the considered system under impulse perturbation and impulse control are established, respectively. It is worth noting that the concepts of AIE and APIE, which are presented in this paper, make impulsive estimation more flexible and relax the constraint of a common threshold. Theoretical results show that time delay size of continuous dynamics is variable and does not have a strict magnitude relationship with impulsive delay. In addition, the obtained display inequalities indicate that time delay of a continuous system may have a potential effect on synchronization. However, from the perspective of impulsive control, there exists a limitation between system time delay size and impulsive interval length, which is an issue to be discussed in subsequent work.
The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors declare that there are no conflicts of interest.
[1] |
S. Dashkovskiy, P. J. Feketa, Input-to-state stability of impulsive systems and their networks, Nonlinear Anal. Hybrid Syst., 26 (2017), 190–200. https://doi.org/10.1016/j.nahs.2017.06.004 doi: 10.1016/j.nahs.2017.06.004
![]() |
[2] |
G. Bertola, W. J. Runggaldier, K. Yasuda, On classical and restricted impulse stochastic control for the exchange rate, Appl. Math. Optim., 74 (2016), 423–454. https://doi.org/10.1007/s00245-015-9320-6 doi: 10.1007/s00245-015-9320-6
![]() |
[3] |
C. Treesatayapun, Impulsive optimal control for drug treatment of influenza A virus in the host with impulsive-axis equivalent model, Inf. Sci., 576 (2021), 122–139. https://doi.org/10.1016/j.ins.2021.06.051 doi: 10.1016/j.ins.2021.06.051
![]() |
[4] |
T. Ensari, S. Arik, Global stability analysis of neural networks with multiple time varying delays, IEEE Trans. Autom. Control, 50 (2005), 1781–1785. https://doi.org/10.1109/TAC.2005.858634 doi: 10.1109/TAC.2005.858634
![]() |
[5] |
M. A. Davo, A. Banos, F. Gouaisbaut, S. Tarbouriech, A. Seuret, Stability analysis of linear impulsive delay dynamical systems via looped-functionals, Automatica, 81 (2017), 107–114. https://doi.org/10.1016/j.automatica.2017.03.029 doi: 10.1016/j.automatica.2017.03.029
![]() |
[6] |
W. Chen, K. Zhang, X. Lu, A refined discretized timer-dependent Lyapunov functional for impulsive delay systems, Automatica, 134 (2021), 109929. https://doi.org/10.1016/j.automatica.2021.109929 doi: 10.1016/j.automatica.2021.109929
![]() |
[7] |
Z. Tang, J. H. Park, J. Feng, Impulsive effects on quasi-synchronization of neural networks with parameter mismatches and time-varying delay, IEEE Trans. Neural Networks Learn. Syst., 29 (2017), 908–919. https://doi.org/10.1109/TNNLS.2017.2651024 doi: 10.1109/TNNLS.2017.2651024
![]() |
[8] |
W. Ren, J. Xiong, Stability analysis of impulsive switched time-delay systems with state-dependent impulses, IEEE Trans. Autom. Control, 64 (2019), 3928–3935. https://doi.org/10.1109/TAC.2018.2890768 doi: 10.1109/TAC.2018.2890768
![]() |
[9] |
J. Lu, B. Jiang, W. Zheng, Potential impacts of delay on stability of impulsive control systems, IEEE Trans. Autom. Control, 67 (2021), 5179–5190. https://doi.org/10.1109/TAC.2021.3120672 doi: 10.1109/TAC.2021.3120672
![]() |
[10] |
A. Khadra, X. Z. Liu, X. Shen, Impulsively synchronizing chaotic systems with delay and applications to secure communication, Automatica, 41 (2005), 1491–1502. https://doi.org/10.1016/j.automatica.2005.04.012 doi: 10.1016/j.automatica.2005.04.012
![]() |
[11] |
K. P. Wijaya, J. P. Chavez, T. Gotz, A dengue epidemic model highlighting vertical-sexual transmission and impulsive control strategies, Appl. Math. Modell., 95 (2021), 279–296. https://doi.org/10.1016/j.apm.2021.02.008 doi: 10.1016/j.apm.2021.02.008
![]() |
[12] |
F. Cacace, V. Cusimano, P. Palumbo, Optimal impulsive control with application to antiangiogenic tumor therapy, IEEE Trans. Control Syst. Technol., 28 (2018), 106–117. https://doi.org/10.1109/TCST.2018.2861410 doi: 10.1109/TCST.2018.2861410
![]() |
[13] |
P. S. Rivadeneira, C. H. Moog, X. Zhao, F. E. Alsaadi, Observability criteria for impulsive control systems with applications to biomedical engineering processes, Automatica, 55 (2015), 125–131. https://doi.org/10.1016/j.automatica.2015.02.042 doi: 10.1016/j.automatica.2015.02.042
![]() |
[14] |
X. Li, W. Liu, S. Gorbachev, J. Cao, Event-Triggered impulsive control for Input-to-State stabilization of nonlinear Time-Delay systems, IEEE Trans. Cybern., 2023. https://doi.org/10.1109/TCYB.2023.3270487 doi: 10.1109/TCYB.2023.3270487
![]() |
[15] |
X. Li, K. Zhang, Synchronization of linear dynamical networks on time scales: Pinning control via delayed impulses, Automatica, 72 (2016), 147–152. https://doi.org/10.1016/j.automatica.2016.06.001 doi: 10.1016/j.automatica.2016.06.001
![]() |
[16] |
X. Li, P. Li, Stability of time-delay systems with impulsive control involving stabilizing delays, Automatica, 124 (2021), 109336. https://doi.org/10.1016/j.automatica.2020.109336 doi: 10.1016/j.automatica.2020.109336
![]() |
[17] |
B. Jiang, J. Lu, Y. Liu, Exponential stability of delayed systems with average-delay impulses, SIAM J. Control Optim., 58 (2020), 3763–3784. https://doi.org/10.1137/20M1317037 doi: 10.1137/20M1317037
![]() |
[18] |
X. Li, S. Song, Stabilization of delay systems: delay-dependent impulsive control, IEEE Trans. Autom. Control, 62 (2016), 406–411. https://doi.org/10.1109/TAC.2016.2530041 doi: 10.1109/TAC.2016.2530041
![]() |
[19] |
X. Liu, K. Zhang, Input-to-state stability of time-delay systems with delay-dependent impulses, IEEE Trans. Autom. Control, 65 (2019), 1676–1682. https://doi.org/10.1109/TAC.2019.2930239 doi: 10.1109/TAC.2019.2930239
![]() |
[20] |
X. Li, S. Song, J. Wu, Exponential stability of nonlinear systems with delayed impulses and applications, IEEE Trans. Autom. Control, 64 (2019), 4024–4034. https://doi.org/10.1109/TAC.2019.2905271 doi: 10.1109/TAC.2019.2905271
![]() |
[21] |
W. Chen, W. Zheng, Exponential stability of nonlinear time-delay systems with delayed impulse effects, Automatica, 47 (2011), 1075–1083. https://doi.org/10.1016/j.automatica.2011.02.031 doi: 10.1016/j.automatica.2011.02.031
![]() |
[22] |
T. Huang, C. Li, S. Duan, J. A. Starzyk, Robust exponential stability of uncertain delayed neural networks with stochastic perturbation and impulse effects, IEEE Trans. Neural Networks Learn. Syst., 23 (2012), 866–875. https://doi.org/10.1109/TNNLS.2012.2192135 doi: 10.1109/TNNLS.2012.2192135
![]() |
[23] |
J. Lu, D. W. C. Ho, J. Cao, J. Kurths, Exponential synchronization of linearly coupled neural networks with impulsive disturbances, IEEE Trans. Neural Networks, 6 (2018), 52570–52581. https://doi.org/10.1109/TNN.2010.2101081 doi: 10.1109/TNN.2010.2101081
![]() |
[24] |
X. Li, X. Zhang, S. Song, Effect of delayed impulses on input-to-state stability of nonlinear systems, Automatica, 76 (2017), 378–382. https://doi.org/10.1016/j.automatica.2016.08.009 doi: 10.1016/j.automatica.2016.08.009
![]() |
[25] |
X. Lv, J. Cao, X. Li, M. Abdel-Aty, U. A. Al-Juboori, Synchronization analysis for complex dynamical networks with coupling delay via event-triggered delayed impulsive control, IEEE Trans. Cybern., 51 (2020), 5269–5278. https://doi.org/10.1109/TCYB.2020.2974315 doi: 10.1109/TCYB.2020.2974315
![]() |
[26] |
N. Wang, X. Li, J. Lu, M. F. E. Alsaadi, Unified synchronization criteria in an array of coupled neural networks with hybrid impulses, Neural Networks, 101 (2018), 25–32. https://doi.org/10.1016/j.neunet.2018.01.017 doi: 10.1016/j.neunet.2018.01.017
![]() |
[27] |
X. Li, N. Wang, J. Lou, J. Lu, Global \mu-synchronization of impulsive pantograph neural networks, Neural Networks, 131 (2020), 78–92. https://doi.org/10.1016/j.neunet.2020.07.004 doi: 10.1016/j.neunet.2020.07.004
![]() |
[28] |
Z. Guan, Z. Liu, G. Feng, Y. Wang, Synchronization of complex dynamical networks with time-varying delays via impulsive distributed control, IEEE Trans. Circuits Syst. I Regul. Pap., 57 (2010), 2182–2195. https://doi.org/10.1109/TCSI.2009.2037848 doi: 10.1109/TCSI.2009.2037848
![]() |
[29] |
W. Chen, W. Zheng, Input-to-state stability and integral input-to-state stability of nonlinear impulsive systems with delays, Automatica, 45 (2009), 1481–1488. https://doi.org/10.1016/j.automatica.2009.02.005 doi: 10.1016/j.automatica.2009.02.005
![]() |
[30] |
W. Chen, J. Chen, X. Lu, Effects of impulse delays on Lp-stability of a class of nonlinear time-delay systems, J. Franklin. Inst., 357 (2020), 7983–8007. https://doi.org/10.1016/j.jfranklin.2020.05.028 doi: 10.1016/j.jfranklin.2020.05.028
![]() |
[31] |
W. Chen, Z. Ruan, W. X. Zheng, Stability and L_2 -gain analysis for linear time-delay systems with delayed impulses: An augmentation-based switching impulse approach, IEEE Trans. Autom. Control, 64 (2019), 4209–4216. https://doi.org/10.1109/TAC.2019.2893149 doi: 10.1109/TAC.2019.2893149
![]() |
[32] |
W. Chen, J. Wen, X. Lu, S. Niu, New stability criteria for linear impulsive systems with interval impulse-delay, J. Franklin. Inst., 358 (2021), 6775–6797. https://doi.org/10.1016/j.jfranklin.2021.07.011 doi: 10.1016/j.jfranklin.2021.07.011
![]() |
1. | Miao Zhang, Bole Li, Weiqiang Gong, Shuo Ma, Qiang Li, Matrix measure-based exponential stability and synchronization of Markovian jumping QVNNs with time-varying delays and delayed impulses, 2024, 9, 2473-6988, 33930, 10.3934/math.20241618 |