
We explored the prescribed-time stability (PTSt) of impulsive piecewise smooth differential systems (IPSDS) based on the Lyapunov theory and set-valued analysis technology, allowing flexibility in selecting the settling time as desired. Furthermore, by developing a feedback controller, we employed the theoretical results to evaluate the synchronization behavior of impulsive piecewise-smooth network systems (IPSNS) within a prescribed time frame and obtained novel criteria to guarantee the synchronization objective. A numerical example was presented to validate the accuracy of the results.
Citation: Chenchen Li, Chunyan Zhang, Lichao Feng, Zhihui Wu. Note on prescribed-time stability of impulsive piecewise-smooth differential systems and application in networks[J]. Networks and Heterogeneous Media, 2024, 19(3): 970-991. doi: 10.3934/nhm.2024043
[1] | Martin Gugat, Mario Sigalotti . Stars of vibrating strings: Switching boundary feedback stabilization. Networks and Heterogeneous Media, 2010, 5(2): 299-314. doi: 10.3934/nhm.2010.5.299 |
[2] | Jun Guo, Yanchao Shi, Yanzhao Cheng, Weihua Luo . Projective synchronization for quaternion-valued memristor-based neural networks under time-varying delays. Networks and Heterogeneous Media, 2024, 19(3): 1156-1181. doi: 10.3934/nhm.2024051 |
[3] | Zhong-Jie Han, Gen-Qi Xu . Dynamical behavior of networks of non-uniform Timoshenko beams system with boundary time-delay inputs. Networks and Heterogeneous Media, 2011, 6(2): 297-327. doi: 10.3934/nhm.2011.6.297 |
[4] | Vincent Renault, Michèle Thieullen, Emmanuel Trélat . Optimal control of infinite-dimensional piecewise deterministic Markov processes and application to the control of neuronal dynamics via Optogenetics. Networks and Heterogeneous Media, 2017, 12(3): 417-459. doi: 10.3934/nhm.2017019 |
[5] | Samitha Samaranayake, Axel Parmentier, Ethan Xuan, Alexandre Bayen . A mathematical framework for delay analysis in single source networks. Networks and Heterogeneous Media, 2017, 12(1): 113-145. doi: 10.3934/nhm.2017005 |
[6] | Leandro M. Del Pezzo, Nicolás Frevenza, Julio D. Rossi . Convex and quasiconvex functions in metric graphs. Networks and Heterogeneous Media, 2021, 16(4): 591-607. doi: 10.3934/nhm.2021019 |
[7] | Lili Jia, Changyou Wang, Zongxin Lei . Synchronization of nonautonomous neural networks with Caputo derivative and time delay. Networks and Heterogeneous Media, 2023, 18(1): 341-358. doi: 10.3934/nhm.2023013 |
[8] | Matteo Novaga, Enrico Valdinoci . Closed curves of prescribed curvature and a pinning effect. Networks and Heterogeneous Media, 2011, 6(1): 77-88. doi: 10.3934/nhm.2011.6.77 |
[9] | Toyohiko Aiki, Kota Kumazaki . Uniqueness of solutions to a mathematical model describing moisture transport in concrete materials. Networks and Heterogeneous Media, 2014, 9(4): 683-707. doi: 10.3934/nhm.2014.9.683 |
[10] | Martino Bardi . Explicit solutions of some linear-quadratic mean field games. Networks and Heterogeneous Media, 2012, 7(2): 243-261. doi: 10.3934/nhm.2012.7.243 |
We explored the prescribed-time stability (PTSt) of impulsive piecewise smooth differential systems (IPSDS) based on the Lyapunov theory and set-valued analysis technology, allowing flexibility in selecting the settling time as desired. Furthermore, by developing a feedback controller, we employed the theoretical results to evaluate the synchronization behavior of impulsive piecewise-smooth network systems (IPSNS) within a prescribed time frame and obtained novel criteria to guarantee the synchronization objective. A numerical example was presented to validate the accuracy of the results.
Stability is a critical metric for assessing system responsiveness and control in the fields of automation, biology, aerospace, and engineering. The traditional stability [1] always studied the asymptotic tendency of a system over an infinite time horizon. However, as more strict stability concepts, finite-time stability (FTS) [2,3] and fixed-time stability (FxTS) [4] gradually attracted the attention of academics due to the growing need for control accuracy and response time. For instance, a novel FTS theory was proposed in [5], where system stability could be achieved within a finite time using a linear time-varying feedback controller. Although FTS has more stringent requirements than traditional stability, the settling time is limited by the system's initial values. FxTS is introduced to get around this restriction. Kong et al. [6] expanded on the FxTS theory and established stability criteria for fuzzy and discontinuous systems. However, the settling time in FxTS is limited due to the system's parameter values. PTSt, a more flexible stability, has gained scholars' interest. PTSt allows the settling time to be selected freely during the design stage and guarantees the system's stability within a prescribed time, which has excellent theoretical value and practical application prospects. For example, Holloway and Krstic [7] presented a novel approach that allows linear systems to be observed in a prescribed time. Zhou and Shi [8] investigated a class of nonlinear systems' prescribed-time stabilization problem and proposed a stabilizing strategy based on a linear time-varying feedback controller, which was extended by [9].
Piecewise-smooth differential systems (PSDS) consist of numerous subsystems, where each subsystem demonstrates smooth dynamical properties inside its state space but may exhibit discontinuous or abrupt behavior at the boundaries. These systems are often used in many real-world applications, including collision and friction phenomena in mechanical systems [10] and switching behavior in electrical circuits [11]. The researches on PSDS stability analysis are challenging, particularly in achieving system stability within a prescribed time. Samadi [12] focused on stability analysis and controller synthesis for a class of PSDS. Chen and Du [13] explored the stability and the response behavior of homoclinic loops in PSDS when subjected to small perturbations and derived the corresponding stability conditions. Samadi and Rodrigues [14] provided a systematic and generalized approach to the stability analysis of PSDS by introducing a unified dissipative framework. Glendinning and Jeffrey [15] provided a comprehensive introduction to piecewise smooth dynamical systems. Li et al. [16] developed two new lemmas related to PTSt in PSDS to achieve prescribed-time synchronization objectives. Although FTS and FxTS have made significant progress compared to traditional stability, they are limited by the system's initial conditions and parameter values. It is worth noting that although some progress has been made for PSDS stability, there is an urgent need for a stability concept that can control the convergence time more flexibly and should not be subject to the initial conditions and parameter limitations. This further emphasizes the need to study PTSt. From the preceding discussion, it is evident that the current research on the stability of PSDS primarily concentrates on the conventional stability theory, while few investigations are reported for PTSt. Filling this theoretical gap is our first motivation.
Many systems experience instantaneous changes, particularly impulse effects, during their dynamic processes. These changes occur in a very short time but significantly impact the systems. Studying such systems helps to simulate complex real-world phenomena accurately and improves the response speed and precision of control systems. Furthermore, the stability of impulsive systems follows into the spotlight. For example, Haddad et al. [17] proposed a mathematical model of impulsive systems and analyzed the stability using Lyapunov method. Nersesov and Haddad [18] developed vector Lyapunov functions theory for impulsive systems with large scale. Xi et al. [19] proposed a new analytical method, using Lyapunov function and characteristics of impulsive time-varying systems, to derive sufficient conditions for the uniform FTS. Xi et al. [20] solved practical FTS problem of nonlinear systems by delayed impulse control approach. Zhao et al. [21] investigated the FTS of linear time-varying singular systems with impulse effects. Jamal et al. [22] studied the FxTS of dynamical systems with impulse effects. Wang and Abdurahman [23] explored the FxTS of general impulsive systems and the synchronization problem of complex networks with hybrid impulses. Li et al. [24] analyzed the influence of impulse series on the stability of the system and derived fixed-time stability for impulsive systems. He et al. [25] addressed the problem of prescribed-time stabilization of nonlinear systems by impulse modulation. IPSDS are more complex and intriguing than traditional ones without piecewise-smooth feature, due to the mixed temporal characteristics. Li et al. [26] proposed a new PTSt theorem for IPSDS and demonstrated its application in network synchronization. In piecewise-smooth network systems with impulse effects, impulsive and piecewise properties complicate the synchronization problem. Traditional PTSy methods may be difficult to be applied directly or do not have the flexibility to deal with these complexities. Based on the above discussions, despite the significant impact of impulse effects on dynamic systems, further comprehensive research, particularly for PTSt, on IPSDS remains insufficient. Studying the PTSt of IPSDS is our further motivation rooted in the first motivation.
In modern engineering, network structures are widely found in various systems, such as social networks, biological networks, power networks, and communication networks. The nodes of networks are connected by edges, which form complex topologies to reflect real systems' complexity and diversity. Investigating the dynamic behavior for these networks, especially synchronization problem, is important for understanding and controlling the network systems. Coraggio et al. [27] proposed a new distributed discontinuous coupling method to synchronize network systems. Dieci and Elia [28] introduced and analyzed the principal stability function applicable to Filippov networks. The synchronization problem of complex networks has become one of the hot research topics, which contains different types of synchronization phenomena such as global synchronization, cluster synchronization. In particular, the prescribed-time synchronization (PTSy), as a special synchronization, has received much attention due to its ability to achieve synchronization within a specified time, see [29,30,31,32]. Detailly, Chen et al. [29] achieved PTSy of complex dynamic networks containing directed spanning trees by designing a suitable controller. Xu and Liu [30] proposed a series of sufficient conditions to ensure that the multi-weighted directed complex network can achieve PTSy by constructing an appropriate Lyapunov function. Yang et al. [31] derived the conditions for fixed-time synchronization of bidirectional associative memory memristive neural networks (BAMMNNs); Based on it, the PTSy of BAMMNNs was implemented. Tang et al. [32] proposed a prescribed-time controller that did not require fractional power and sign functions to achieve PTSy for a switched network. Although many synchronization results are addressed for complex networks, they mainly focus on network systems with continuous dynamic properties. For those network systems with impulse effects or piecewise-smooth characteristics, existing synchronization methods for PTSy show limitations in dealing with transient behavior or boundary behavior. The above discussions suggest that PTSy is an important problem to be solved in complex network systems with impulse effects and piecewise-smooth properties. The study on PTSy of IPSNS is our third motivation.
Based on the previous discussions, this work intends to investigate the PTSt of IPSDS, allowing flexibility in choosing the settling time as needed. Furthermore, the theoretical results are used to tackle the synchronization problem in IPSNS by importing a feedback controller. The significant contributions are as follows:
1) Employing the Lyapunov inequality and set-valued analysis technology, we construct a theorem on the PTSt of IPSDS, supplying sufficient criteria for the PTSt. Particularly, the setting time can be flexibly selected as required, regardless of the system's initial values and control parameters. This responds to our first and second motivations.
2) A feedback controller is designed to guarantee that the IPSNS can achieve synchronization within a prescribed time under the established stability theorem. This responds to our third motivation.
The rest of this paper is organized as follows: In Section 2, we introduce basic symbols, definitions, and lemmas. In Section 3, we present a PTSt theorem applicable to IPSDS. In Section 4, we design a feedback controller to achieve synchronization in IPSNS within a prescribed time. In Section 5, we provide an example to illustrate the correctness of the theoretical results. In Section 6, we conclude the paper.
Notations. Let R, R+, N, N+ denote the sets of real numbers, positive real numbers, natural numbers, and positive integers, respectively. Let Rn represent the n-dimensional column vector, Rn×m the set of n×m real matrices, In the identity matrix of dimension n. A>0(A<0) denotes positive (negative) definite matrix. AT represents the transpose of matrix A. ‖A‖=√λmax(ATA), where λmax(ATA) is the maximun eigenvalue of matrix ATA. Aς=12(A+AT). For any x=(x1, x2, ⋯, xn)T, y=(y1, y2, ⋯, yn)T, x∘y=(x1y1, x2y2, ⋯, xnyn)T. ⊗ represents the Kronecker produced.
Consider a series of non-empty, open, and disjoint finite sets P1, P2, ⋯, P𝓁 such that Q⊆⋃𝓁𝓀=1−P𝓀⊆Rn is a finite collection, and −Pi∩−Pj is a line. A differential system with impulse effects is concerned,
{˙x(t)=g(x(t)), t≠tkΔx(t)=αkx(t−), t=tkx(0)=x0, , | (1) |
where g:Q↦Rn is a vector field, αk is a constant. For k∈N, Δx(tk)=x(t+k)−x(t−k), in which x(t+k)=limt→t+kx(t), x(t−k)=limt→t−kx(t). The impulse time sequence {tk}k∈N satisfies 0=t0<t1<t2<⋯<tk⋯, limk→∞tk=+∞. If g(x) is a finite set of vector fields, i.e.,
g(x)=G𝓀(x), x∈P𝓀, |
then system (1) is IPSDS, where 𝓀=1, 2, ⋯, 𝓁, each vector field G𝓀(x) is smooth in x for any x∈P𝓀 and continuously extends on the boundary ∂P𝓀. Additionally, assume that the solution x(t) of system (1) is always right-discontinuous and has left-hand limits. Then, the solution x(t) can be defined as Filippov solutions.
Remark 1. We focus on one-dimensional discontinuity manifolds, which appear at the switching boundaries of PSDSs and play a crucial role in describing the systems' switching behavior and trajectory evolutions. However, except for these low-dimensional discontinuity manifolds, higher-order cases are also crucial in complex systems. Higher-order discontinuity manifolds involve more complex geometric structures and can lead to more intricate dynamic phenomena. Difonzo [33] systematically analyzed these higher-order manifolds using set-valued theory and geometric methods and introduced the "isochronous attainable manifolds", which can help one understand these manifolds in higher-dimensional systems. In the future, we will apply higher-order manifolds to more complex control systems to gain deeper insights into the overall dynamic behavior and switching mechanisms of such systems.
Definition 1. [34] A vector function x(t):R+↦Rn is a Filippov solution of system (1) if it is absolutely continuous on [tk, tk+1), k∈N, and for almost every t>0,
{x(t)∈G[g](x), t≠tkΔx(t)=αkx(t−), t=tkx(0)=x0, , |
where Filippov set-valued function G[g]:Rn↦A(Rn), with A(Rn) is the set of all subsets of Rn, is given by
G[g](x)=⋂ε>0⋂a(P)=0−co{g(B(x, ε))∖P}, |
a(·) denotes the Lebesgue measure, −co(·) represents the closed convex hull, B(x, ε)={x∈Rn:‖x‖≤ε}.
Definition 2. [35] Let ν:Rn↦R be a locally Lipschitz function. The set-valued Lie derivative LG[g]ν(x):Rn↦A(Rn) of ν concerning G[g], is given by
LG[g]ν(x)≜{b∈R:∃c∈G[g](x)such thatdTc=b, ∀d∈∂ν(x)}, |
where ∂ν(x)≜{lim𝓀→+∞∇ν(x𝓀):x𝓀→x, x𝓀∉P∪Ων} is the generalized gradient of ν at any x∈Rn, and Ων is the set of zero Lebesgue measures.
Definition 3. [2] System (1) is finite-time stable, if for ∀x0∈Rn, there exists a function 0≤T(x)<+∞, such that
{limt→T0‖x(t)‖=0, ‖x(t)‖≡0, t≥T0, |
where T0=T(x0) is the settling time function.
Definition 4. [36] Given a prescribed time constant ˆT>0, system (1) is prescribed-time stable, if it is finite-time stable and the settling time T0≤ˆT.
Lemma 1. [37] If x(t) be a Filippov solution of system (1), ν:Rn↦R be a locally Lipschitz regular function, then ˙ν(x(t)) exists and ˙ν(x(t))∈LG[g]ν(x(t)) almost everywhere.
We construct a PTSt theorem for IPSDS in which the settling time is freely chosen and unaffected by the system's initial values and control parameters.
Theorem 1. Given a prescribed time ˆT and x(0)∈B(x, ε), system (1) is prescribed-time stable, provided there exist a positive definite and regular locally Lipschitz function V:Rn↦Rn, K∞ -class functions λ, μ, and constants 0<γ≤1, p, q>0, 0<δ<1, 0<ϵ<1, such that
(C1)ψ(t)≤−βˆT(pVδ(x(t))−qV(x(t))),∀x(t)∈Rn∖{0},t≠tk,(C2)V(x(t))≤γV(x(t−)),∀x∈Rn∖{0},t=tk,(C3)λ(‖x(t)‖)≤V(x(t))≤μ(‖x(t)‖),(C4)maxk∈{1, 2, ⋯, M0}{tk−tk−1}≤ϵˆTM0, |
where ψ(t)∈LG[g]V(x(t)), β=ϵ(1−ϵ)2(1−δ)q, M0=min{ϱ∈N+|ϱ≥−lnπ−qβϵlnγ}, π=μ(ε)(qϵp)11−δ.
Proof. Based on Lemma 1, for almost every t>0, there exists
ψ(t)=˙V(x(t))∈LG[g]V(x(t)). |
From condition (C3), we can obtain that
‖x(t)‖≤λ−1(V(x(t))), |
V(x(0))≤μ(‖x(0)‖). |
From x(0)∈B(x, ε), then ‖x(0)‖≤ε. Combined with the definition of π, yields
V(x(0))≤μ(‖x(0)‖)≤μ(ε)=π(ϵpq)11−δ. |
Due to the uncertainty in π, we cannot determine the relationship between V(x(0)) and (ϵpq)11−δ. Therefore, we can analyze the PTSt of IPSDS (1) under the following two scenarios:
Case I: V(x(0))≤(ϵpq)11−δ. In this case, there is for ∀t≥0, V(x(t))≤(ϵpq)11−δ. There must be at least t≥0 such that V(x(t))>(ϵpq)11−δ, if this conclusion is false. Let ts=inf{t≥0|V(x(t))>(ϵpq)11−δ}, it follows that V(x(ts))=(ϵpq)11−δ. Thus, for t∈[0, ts], V(x(t))≤V(x(ts)). According to condition (C2), ts is not an impulse moment. Then we can get
˙V(x(ts))≥0. |
It follows from condition (C1) that
˙V(x(ts))≤−βˆT(pVδ(x(ts))−qV(x(ts)))=−βˆT[p−qV1−δ(x(ts))]Vδ(x(ts))=−(1−ϵ)pβˆTVδ(x(ts)). |
Letting ϖ=(1−ϵ)pβˆT, it can be inferred that ϖ>0, namely
˙V(x(ts))<0. |
This contradicts the above conclusion and, therefore, for ∀t≥0, V(x(t))≤(ϵpq)11−δ.
When t≠tk, it follows from condition (C1) that
˙V(x(t))≤−βˆT(pVδ(x(t))−qV(x(t)))=−βˆT[p−qV1−δ(x(t))]Vδ(x(t))=−(1−ϵ)pβˆTVδ(x(t)), |
from which we can get when t≠tk,
˙V(x(t))≤−ϖVδ(x(t)). | (2) |
Considering the impulse effect, divide the interval [0, t∞) into [0, t1), [t1, t2), ⋯, [tM−1, tM)⋯. Then, it can be deduced from (2) that
V1−δ(x(t))≤V1−δ(x(0))−ϖ(1−δ)t, t≠tk. | (3) |
Integrating on both sides of (2) from 0 to t, where t∈[0, t1), one has
V1−δ(x(t))≤V1−δ(x(0))−ϖ(1−δ)t. |
Hence, (3) holds for t∈[0, t1). Assume that (3) is correct for t∈[tM−2, tM−1), where M≥2. Subsequently, for t∈[tM−2, tM−1), there is
V1−δ(x(t))≤V1−δ(x(0))−ϖ(1−δ)t. |
From this, it can be easily inferred that V1−δ(x(t−M−1))≤V1−δ(x(0))−ϖ(1−δ)t−{M−1}. Because γ1−δ≤1, we can deduce from condition (C2) that
V1−δ(x(tM−1))≤γ1−δV1−δ(x(t−M−1))≤V1−δ(x(0))−ϖ(1−δ)tM−1. |
Integrating on both sides of (2) from tM−1 to t, where t∈[tM−1, tM), we have
V1−δ(x(t))≤V1−δ(x(tM−1))−ϖ(1−δ)(t−tM−1)≤V1−δ(x(0))−ϖ(1−δ)tM−1−ϖ(1−δ)(t−tM−1)=V1−δ(x(0))−ϖ(1−δ)t. |
Hence, (3) holds for t∈[tM−1, tM). In summary, (3) is deemed to be correct when t≠tk. Considering condition (C2), V1−δ(x(t))≤V1−δ(x(0))−ϖ(1−δ)t for ∀t≥0. Then, T0 satisfies
T0=V1−δ(x(0))ϖ(1−δ)≤((ϵpq)11−δ)1−δ·11−δ·ˆT(1−ϵ)pβ=ϵpq·11−δ·ˆT(1−ϵ)p·(1−ϵ)2(1−δ)qϵ=(1−ϵ)ˆT. |
We can then deduce that as t→T0, V(x(t))→0. By combining (2) and condition (C2), we have for ∀t≥0, ˙V(x(t))≤0. Therefore, as t≥T0, 0≤V(x(t))≤V(x(T0))→0. Consequently, as t≥T0, V(x(t))≡0, Since ‖x(t)‖≤λ−1(V(x(t))), T0≤(1−ϵ)ˆT and ˙V(x(t))≤0, for ∀t≥0, it can be inferred that ‖x(t)‖→0, t→ˆT; ‖x(t)‖≡0, t≥ˆT.
Thus, the IPSDS (1) is deemed to be prescribed-time stable when V(x(0))≤(ϵpq)11−δ.
Case II: V(x(0))>(ϵpq)11−δ. From condition (C1), we can obtain that when t≠tk,
˙V(x(t))≤qβˆTV(x(t)). | (4) |
Let M(t) represent the count of impulses within (0, t]. Then, it can be deduced from (4) that, when t≠tk,
V(x(t))≤exp{M(t)lnγ+qβˆTt}V(x(0)), | (5) |
Integrating on both sides of (4) from 0 to t, where t∈[0, t1), one has
V(x(t))≤exp{qβˆTt}V(x(0)). |
It is evident that when t∈[0, t1), M(t)=0. Hence, (5) holds for t∈[0, t1). Assume that (5) is correct for t∈[tM−2, tM−1), where M≥2. As t∈[tM−2, tM−1), M(t)=M−2, then we can get
V(x(t))≤exp{(M−2)lnγ+qβˆTt}V(x(0)). |
From this, it can be easily inferred that V(x(t−M−1))≤exp{(M−2)lnγ+qβˆTt−{M−1}}V(x(0)) . We can deduce from condition (C2) that
V(x(tM−1))≤γV(x(t−M−1))≤γexp{(M−2)lnγ+qβˆTtM−1}V(x(0))=exp{(M−1)lnγ+qβˆTtM−1}V(x(0)). |
Integrating on both sides of (4) from tM−1 to t, where t∈[tM−1, tM), we have
V(x(t))≤exp{qβˆT(t−tM−1)}V(x(tM−1))≤exp{qβˆT(t−tM−1)}exp{(M−1)lnγ+qβˆTtM−1}V(x(0))=exp{(M−1)lnγ+qβˆTt}V(x(0)). |
As t∈[tM−2, tM−1), M(t)=M−1, hence, (5) holds for t∈[tM−1, tM). In summary, (5) is deemed to be correct when t≠tk. Considering condition (C2), V(x(t))≤exp{M(t)lnγ+qβˆTt}V(x(0)) for ∀t≥0. Based on (5) and condition (C4), we can conclude that
V(x(tM0))≤exp{M0lnγ+qβˆTtM0}V(x(0))≤exp{−lnπ−qβϵlnγlnγ+qβˆTϵˆT}π(ϵpq)11−δ=exp{−lnπ}π(ϵpq)11−δ=(ϵpq)11−δ. |
From this, we can obtain V(x(tM0))≤(ϵpq)11−δ. According to Case I, T0 satisfies
T0=V1−δ(x(tM0))ϖ(1−δ)≤((ϵpq)11−δ)1−δ·11−δ·ˆT(1−ϵ)pβ≤ϵpq·11−δ·ˆT(1−ϵ)p·(1−ϵ)2(1−δ)qϵ=(1−ϵ)ˆT. |
We can then deduce that as t→tM0+T0, V(x(t))→0. By combining (2) and condition (C2), we have for ∀t≥tM0, ˙V(x(t))≤0. Therefore, as t≥tM0+T0, 0≤V(x(t))≤V(x(tM0+T0))→0. Consequently, V(x(t))≡0. From tM0≤ϵˆT and T0≤(1−ϵ)ˆT, tM0+T0≤ˆT. Similarly to Case I, it can be inferred that ‖x(t)‖→0, t→ˆT; ‖x(t)‖≡0, t≥ˆT.
Thus, the IPSDS (1) is deemed to be prescribed-time stable when V(x(0))>(ϵpq)11−δ. The proof of Theorem 1 is completed.
Remark 2. We introduce a constant gain ˆT in the nonlinear function of condition (C1) to accommodate PTSt. Although [16] also used a constant gain approach, it failed to consider the impact of the impulse effect.
Remark 3. The set-valued Lie derivative can handle discontinuities and piecewise smoothness in systems, whereas the conventional derivatives can be used only for smooth systems. By introducing the set-valued Lie derivative, we can well describe and analyze discontinuous changes in the systems. Therefore, for condition (C1) in Theorem 1, unlike [25,38], we employ the set-valued Lie derivative. This approach enables us to handle a wider range of functions and more complex systems than conventional ones.
Remark 4. The nonlinear Lyapunov inequality in [26] incorporated time-varying functions, which increased the difficulty in practical applications. In comparison, our nonlinear functions are easier to calculate and analyze.
Remark 5. Our nonlinear function of condition (C1) includes a positive linear term, enabling our nonlinear Lyapunov inequality to have a broader range of applicability compared to [39].
Remark 6. Research on systems with impulse effects [19,20,21,22,23,24] primarily focuses on analyzing finite-time/fixed-time stability. In contrast, our attention shifts to the PTSt of such systems. This paper broadens the understanding of stability and provides a detailed analytical framework and explicit time bounds for PTSt. For the PTSt problem of general systems, as discussed in [7,8,16,40], the analysis is typically straightforward since impulse effects are not considered. In this paper, we divide the system's time into series of impulsive intervals and analyze the system's stability within each interval. We also consider the impact of impulsive moments on the system to ensure that impulsive effects do not undermine stability. This piecewise analysis method guarantees that the system can achieve stability within the prescribed time, even with impulse effects.
Remark 7. Our method provides a framework for proving PTSt and effectively handles systems with impulse effects. The approach allows for the analysis of system behavior from different initial states based on the initial value V(x(0)) of the Lyapunov function. This flexibility makes the method applicable to a broader range of systems. Additionally, Theorem 1 gives a clear time bound within which the system state must converge to the equilibrium point, which is crucial for determining the maximum convergence time in practical applications. However, its complexity, potential conservatism, and dependence on specific conditions may limit practical applications of Theorem 1 in certain scenarios. For example, conditions (C1)–(C4) are complex, involving multiple parameters and inequalities, which may make it difficult to validate them in practical applications.
Remark 8. In this paper, we study the PTSt of IPSDS without considering the significant factors of stochasticity and fractional order. Work [41,42,43,44] have studied the stability of stochastic system with fractional order/impulse effects. These works motivate us to study the stability, particularly PTSt, of IPSDS with stochasticity/fractional order.
Consider the following IPSNS with N nodes
{˙xi(t)=g(xi(t))−m∑Nj=1ιijΠxj(t)+ui(t), t≠tkΔxi(t)=αkxi(t−), t=tk, | (6) |
where i=1, 2, ⋯, N, xi(t)=(xi1(t), xi2(t), ⋯, xiN(t))T denotes the i -th node's state vector, piecewise function g(xi(t))=(g1(xi1(t)), g2(xi2(t)), ⋯, gN(xiN(t)))T represents the i -th node's vector field, m stands for the coupling gain, Π=diag{τ1, τ2, ⋯, τN} denotes the internal coupling matrix, ui(t) indicates the i -th node's control input, −L=(ιij)N×N is its Laplace matrix.
Remark 9. |1+αk|≤1 holds for this section.
The synchronization target system of the IPSNS (6) is
˙z(t)=g(z(t)), | (7) |
where z(0)=z0, z(t)=(z1(t), z2(t), ⋯, zN(t))T.
Let ωi(t)=xi(t)−z(t) denote the synchronization error of the i -th node, i=1, 2, ⋯, N. Subtracting the synchronization target system (7) from the IPSNS (6), give the synchronization error system (SES),
{˙ωi(t)= g(ωi(t))−m∑Nj=1ιijΠωj(t)+ui(t), t≠tkΔωi(t)=αkωi(t−), t=tk, | (8) |
where g(ωi(t))=g(xi(t))−g(z(t)), ωi(t)=(ωi1(t), ωi2(t), ⋯, ωiN(t))T, i=1, 2, ⋯, N.
Definition 5. Given a prescribed time constant ˆT, the IPSNS (6) synchronizes with target system (7) within the prescribed time ˆT, if
{limt→ˆT‖xi(t)−z(t)‖=0, ‖xi(t)−z(t)‖≡0, t>ˆT, |
where i=1, 2, ⋯, N, namely
{limt→ˆT‖ωi(t)‖=0, ‖ωi(t)‖≡0, t>ˆT. |
This indicates that only if the IPSNS (6) and target system (7) synchronize within the prescribed time, the SES (8) will be prescribed-time stable.
We designed a feedback controller to attain the synchronization objective,
ui(t)=−ρiωi(t)−ϑisign(ωi(t))−βˆT{sign(ωi(t))∘[ωi(t)]2δ−1−sign(ωi(t))∘[ωi(t)]}, | (9) |
where ρi, ϑi, i=1, 2, ⋯, N are positive constants to be chosen, δ, β and ˆT are defined as in Theorem 1.
We incorporate a stacked vector to rewrite the SES (8):
{˙ω(t)=Λ(ω(t))+Σ(ω(t))+U(ω(t)), t≠tkΔωi(t)=αkωi(t−), t=tk, | (10) |
where Σ(ω(t))=−m(−L⊗Π)ω(t), ω(t)=(ωT1(t), ωT2(t), ⋯, ωTN(t))T, Λ(ω(t))=( gT(ω1(t)), gT(ω2(t)), ⋯, gT(ωN(t)))T, U(ω(t))=(uT1(t), uT2(t), ⋯, uTN(t))T.
Two rules are proposed to simplify the computation of the Filippov set-valued function,
(i) If g:Rn↦Rn is continuous at x∈R, then G[g](x)=g(x).
(ii) If g1, g2:Rn↦Rn are locally bounded at x∈R, then G[g1+g2](x)⊆G[g1](x)+G[g2](x).
Assumption 1. For ∀y1, y2∈Rn, there exist C, D∈Rn×n and a piecewise-smooth function g:Rn↦Rn such that
(y1−y2)T(−g(y1)−−g(y2))≤(y1−y2)TC(y1−y2)+(y1−y2)TDsign(y1−y2), |
holds, where −g(yi)∈G[g](yi), i=1, 2.
In this article, assume that x(tk)=x(t+k). Next, we propose the synchronization assertion under controller (9).
Theorem 2. Under Assumption 1, the IPSNS (6) synchronizes with the target system (7) within the prescribed time under controller ui(t), provided there exist matrices C, D∈RN×N such that
(i) IN⨂Cς−m−Lς⨂Π−ρ⨂IN≤0,
(ii) IN⨂D−ϑ⨂IN≤0,
(iii) maxk∈{1, 2, ⋯, M'0}{tk−tk−1}≤ϵˆTM'0,
where ρ=diag(ρ1, ρ2, ⋯, ρN), ϑ=diag(ϑ1, ϑ2, ⋯, ϑN), M'0=min{ϱ'∈N+|ϱ'≥−ln(ε2ϵ11−δ)−ϵ(1−δ)(1−ϵ)2lnγ}, ϵ, ε, and γ are defined in Theorem 1.
Proof. Consider
V(ω(t))=12ωT(t)ω(t)=12N∑i=1ωTi(t)ωi(t). |
Letting λ(ε)=μ(ε)=12ε2, we can confirm that λ and μ meet condition (C3) in Theorem 1. Now we will deduce the synchronization criteria from two scenarios: t≠tk and t=tk.
Case I: t≠tk. By Lemma 1, we can get
˙V(ω(t))=ωT(t)˙ω(t)∈LG[Ξ]V(ω(t)), |
where Ξ=Λ(ω(t))+Σ(ω(t))+U(ω(t)). From the criteria for calculating the set-valued function LG[Ξ]V(ω(t))⊆LG[Λ]+G[Σ]+G[U]V(ω(t)), and
LG[Λ]+G[Σ]+G[U]V(ω(t))=ωT(t){G[Λ](ω(t))+G[Σ](ω(t))+G[U](ω(t))}=N∑i=1ωTi(t){G[ g](ωi(t))−mN∑j=1ιijΠωj(t)−ρiωi(t)−ϑiG[sign](ωi(t))−βˆT{G[sign](ωi(t))∘[ωi(t)]2δ−1−G[sign](ωi(t))∘[ωi(t)]}}=N∑i=1ωTi(t)G[ g](ωi(t))−mN∑i=1N∑j=1ωTi(t)ιijΠωj(t)−N∑i=1ρiωTi(t)ωi(t)−N∑i=1ϑiωTi(t)G[sign](ωi(t))−βˆTN∑i=1ωTi(t){G[sign](ωi(t))∘[ωi(t)]2δ−1−G[sign](ωi(t))∘[ωi(t)]}, |
where G[ g](ωi(t))=G[g](xi(t))−G[g](z(t)), G[sign](ωi(t))={1, ωi(t)>0−1, ωi(t)<0[−1, 1], ωi(t)=0. From Assumption 1, we obtain that
N∑i=1ωTi(t)G[ g](ωi(t))=N∑i=1ωTi(t){G[g](xi(t))−G[g](z(t))}≤N∑i=1ωTi(t)Cωi(t)+N∑i=1ωTi(t)DG[sign](ωi(t)). |
Let
Γ1=N∑i=1ωTi(t)Cωi(t)−N∑i=1N∑j=1ωTi(t)ιijΠωj(t)−N∑i=1ρiωTi(t)ωi(t), |
Γ2=N∑i=1ωTi(t)DG[sign](ωi(t))−N∑i=1ϑiωTi(t)G[sign](ωi(t)), |
Γ3=−βˆTN∑i=1ωTi(t){G[sign](ωi(t))∘[ωi(t)]2δ−1−G[sign](ωi(t))∘[ωi(t)]}, |
Thus,
LG[Λ]+G[Σ]+G[U]V(ω(t))≤Γ1+Γ2+Γ3. |
From Aς=12(A+AT), it follows that
ωTi(t)Cωi(t)=12{ωTi(t)Cωi(t)+ωTi(t)CTωi(t)}=12{ωTi(t)(C+CT)ωi(t)}=ωTi(t){12(C+CT)}ωi(t)=ωTi(t)Cςωi(t). |
Then, Γ1, Γ2, Γ3 can be rewritten as
Γ1=ωT(t)(IN⨂Cς−m−Lς⨂Π−ρ⨂IN)ω(t)≤λmax(IN⨂Cς−m−Lς⨂Π−ρ⨂IN)ωT(t)ω(t), |
Γ2=ωT(t)(IN⨂D−ϑ⨂IN)G[sign](ω(t))≤λmax(IN⨂D−ϑ⨂IN)‖ω(t)‖1, |
Γ3=−βˆT(N∑i=1|ωi(t)|2δ−N∑i=1|ωi(t)|2)≤−βˆT{(N∑i=1|ωi(t)|2)δ−N∑i=1|ωi(t)|2}=−βˆT{2δVδ(ω(t))−2V(ω(t))}. |
From ˙V(ω(t))∈LG[Ξ]V(ω(t))⊆LG[Λ]+G[Σ]+G[U]V(ω(t)), we have
˙V(ω(t))≤Γ1+Γ2+Γ3≤λmax(IN⨂Cς−m−Lς⨂Π−ρ⨂IN)ωT(t)ω(t)+λmax(IN⨂D−ϑ⨂IN)‖ω(t)‖1−βˆT(2δVδ(ω(t))−2V(ω(t))). |
Based on conditions (i) and (ii), we can obtain that λmax(IN⨂Cς−m−Lς⨂Π−ρ⨂IN)≤0, λmax(IN⨂D−ϑ⨂IN)≤0. Therefore, the above equation can be reduced to
˙V(ω(t))≤−βˆT(2δVδ(ω(t))−2V(ω(t))). |
Case II: t=tk. From ω(tk)=(1+αk)ω(t−k), and |1+αk|≤1, we have
V(ω(tk))=12ωT(tk)ω(tk)#=12(1+αk)2ωT(t−k)ω(t−k)≤12ωT(t−k)ω(t−k)=V(ω(t−k)). |
Consequently, when t=tk, V(ω(tk))≤V(ω(t−k)).
Therefore, Theorem 2 is proven according to Theorem 1. The proof has been finished.
Remark 10. In contrast to [16], our assumption incorporates the sign function, enabling Theorem 2 to handle functions with discontinuities. Unlike [16], consider impulse effects and divide the proof of Theorem 2 into two distinct cases: t≠tk and t=tk.
Remark 11. Theorem 2 is based on Assumption 1. Different from the assumption in [26], our assumption does not include the matrix P.
Consider the following three-dimensional chaotic Sprott circuit system,
{˙xi(t)=g(xi(t))−m3∑j=1ιijΠxj(t)+ui(t), t≠tk, Δxi(t)=αkxi(t−), t=tk, |
where xi(t)=(xi1(t), xi2(t), xi3(t))T, i=1, 2, 3,
g(xi(t))=[010001−1−1−0.5]xi(t)+[00sign(xi(t))]. |
Let ˆT=3, m=1, −L=[2−1−1−12−1−1−12], Π=[10001.20001], αk=−0.3. Let ρ1=3.27, ρ2=2.64, ρ3=2.74, ϑ1=0.03, ϑ2=0.01, ϑ3=0.04, p=1.26, q=2, δ=0.33, ϵ=0.3, tk=0.15k, k=0, 1, 2, ⋯. Let C=[−0.300.89−0.810.29−1.15−2.94−0.79−1.071.44], D=[−0.83−0.530.52−0.98−2.00−0.02−1.160.96−0.03], it is easy to verify that C and D satisfy the conditions of Theorem 2. Choose the initial value x1(0)=(−3.81, −0.02, 4.60)T, x2(0)=(−1.60, 0.85, −2.76)T, x3(0)=(2.51, −2.45, 0.06)T. It is clear that Theorem 2's conditions are satisfied. Figure 1(a), (b) shows the trajectories of IPSNS (6) without and with controller (9), respectively. Figure 2 shows the evolution of the synchronization error of the IPSNS under the controller (9), indicating that the SES (10) converges to 0 within ˆT=3, i.e., the IPSNS (6) is synchronized with the target system (7) within ˆT=3.
We select the same initial values as previously mentioned and modify the control parameters to see if the control parameters would affect the settling time: ρ'1=3.27, ρ'2=2.64, ρ'3=2.74, ϑ'1=0.03, ϑ'2=0.01, ϑ'3=0.04. Figure 3(a) illustrates the results. We choose the same control parameters as previously mentioned and modify the initial values to see if the initial values would affect the settling time: x1'(0)=(2.63, −4.09, 2.44)T, x2'(0)=(2.14, −3.04, 4.35)T, x3'(0)=(−1.66, −4.39, 1.96)T. Figure 3(b) illustrates the results. Figure 3(a), (b) shows that the SES (10) converges to 0 within ˆT=3, i.e., the IPSNS (6) and the target system (7) are synchronized within ˆT=3, and the settling time is not affected by the control parameters and the initial values, although the control parameters and the initial values are changed.
To test whether the settling time can be set freely, we modify only the settling time to ˆT=5, keeping the previous parameter values unchanged. The results can be seen in Figure 4, which demonstrates that even with a change in the settling time, the IPSNS (6) and the target system (7) are synchronized within ˆT=5. This implies that the settling time can be adjusted as needed.
In this paper, the PTSt of IPSDS is investigated, where the settling time can be set arbitrarily, independent of the initial values and control parameters of the system. Furthermore, a feedback controller is developed and the theoretical results are used to derive a new criterion for ensuring the synchronization of the IPSNS within a prescribed time.
In the future, we can research the PTSt of IPSDS with the factors of time-varying/higher-order/ fractional order/stochasticity, and we can investigate the PTSt of IPSDS through the event-triggered mechanism. By studying the PTSt of IPSDS with these factors, we can realize more flexible and efficient time control in a wider range of practical applications. Compared with the traditional control methods, these methods can better adapt to dynamically changing environments and limited resource conditions and have a wide range of application prospects.
Chenchen Li: Investigation, Methodology, Writing, Validation; Chunyan Zhang: Investigation, Validation; Lichao Feng: Conceptualization, Investigation, Supervision, Funding acquisition; Zhihui Wu: Writing and Software.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This project is jointly supported by the Humanities and Social Science Fund of Ministry of Education of China (23YJAZH031), the Natural Science Foundation of Hebei Province of China (A2023209002, A2019209005), the Tangshan Science and Technology Bureau Program of Hebei Province of China (24130201C, 19130222g), and the Graduate Student Innovation Fund of North China University of Science and Technology (2024S15).
The authors declare there is no conflict of interest.
[1] | R. W. Brockett, Asymptotic stability and feedback stabilization, Differ. Geom. Control Theory, 27 (1983), 181–191. |
[2] |
S. P. Bhat, D. S. Bernstein, Finite-time stability of continuous autonomous systems, SIAM J. Control Optim., 38 (2000), 751–766. https://doi.org/10.1137/S0363012997321358 doi: 10.1137/S0363012997321358
![]() |
[3] | W. M. Haddad, Q. Hui, Dissipativity theory for discontinuous dynamical systems: Basic input, state, and output properties, and finite-time stability of feedback interconnections, Nonlinear Anal.: Hybrid Syst., 3 (2009), 551–564. https://doi.org/10.1016/j.nahs.2009.04.006 |
[4] |
A. Polyakov, Nonlinear feedback design for fixed-time stabilization of linear control systems, IEEE Trans. Autom. Control, 57 (2011), 2106–2110. https://doi.org/10.1109/TAC.2011.2179869 doi: 10.1109/TAC.2011.2179869
![]() |
[5] |
B. Zhou, Finite-time stability analysis and stabilization by bounded linear time-varying feedback, Automatica, 121 (2020), 109191. https://doi.org/10.1016/j.automatica.2020.109191 doi: 10.1016/j.automatica.2020.109191
![]() |
[6] |
F. C. Kong, Q. X. Zhu, T. W. Huang, New fixed-time stability lemmas and applications to the discontinuous fuzzy inertial neural networks, IEEE Trans. Fuzzy Syst., 29 (2020), 3711–3722. https://doi.org/10.1109/TFUZZ.2020.3026030 doi: 10.1109/TFUZZ.2020.3026030
![]() |
[7] |
J. Holloway, M. Krstic, Prescribed-time observers for linear systems in observer canonical form, IEEE Trans. Autom. Control, 64 (2019), 3905–3912. https://doi.org/10.1109/TAC.2018.2890751 doi: 10.1109/TAC.2018.2890751
![]() |
[8] |
B. Zhou, Y. Shi, Prescribed-time stabilization of a class of nonlinear systems by linear time-varying feedback, IEEE Trans. Autom. Control, 66 (2021), 6123–6130. https://doi.org/10.1109/TAC.2021.3061645 doi: 10.1109/TAC.2021.3061645
![]() |
[9] |
L. C. Feng, M. Y. Dai, N. Ji, Y. L. Zhang, L. P. Du, Prescribed-time stabilization of nonlinear systems with uncertainties/disturbances by improved time-varying feedback control, AIMS Math., 9 (2024), 23859–23877. https://doi.org/10.3934/math.20241159 doi: 10.3934/math.20241159
![]() |
[10] |
D. X. Cao, X. X. Zhou, X. Y. Guo, N. Song, Limit cycle oscillation and dynamical scenarios in piecewise-smooth nonlinear systems with two-sided constraints, Nonlinear Dyn., 112 (2024), 9887–9914. https://doi.org/10.1007/s11071-024-09589-6 doi: 10.1007/s11071-024-09589-6
![]() |
[11] |
M. R. Jeffrey, Smoothing tautologies, hidden dynamics, and sigmoid asymptotics for piecewise smooth systems, Chaos, 25 (2015), 103125. https://doi.org/10.1063/1.4934204 doi: 10.1063/1.4934204
![]() |
[12] | B. Samadi, Stability Analysis and Controller Synthesis for A Class of Piecewise Smooth Systems, Ph.D thesis, Concordia University, 2008. |
[13] |
S. Chen, Z. D. Du, Stability and perturbations of homoclinic loops in a class of piecewise smooth systems, Int. J. Bifurcation Chaos, 25 (2015), 1550114. https://doi.org/10.1142/S021812741550114X doi: 10.1142/S021812741550114X
![]() |
[14] |
B. Samadi, L. Rodrigues, A unified dissipativity approach for stability analysis of piecewise smooth systems, Automatica, 47 (2011), 2735–2742. https://doi.org/10.1016/j.automatica.2011.09.018 doi: 10.1016/j.automatica.2011.09.018
![]() |
[15] | P. Glendinning, M. R. Jeffrey, An Introduction to Piecewise Smooth Dynamics, Springer International Publishing, Switzerland, 2019. https://doi.org/10.1007/978-3-030-23689-2 |
[16] |
X. N. Li, H. Q. Wu, J. D. Cao, Prescribed-time synchronization in networks of piecewise smooth systems via a nonlinear dynamic event-triggered control strategy, Math. Comput. Simul., 203 (2023), 647–668. https://doi.org/10.1016/j.matcom.2022.07.010 doi: 10.1016/j.matcom.2022.07.010
![]() |
[17] |
W. M. Haddad, V. Chellaboina, N. A. Kablar, Non-linear impulsive dynamical systems, Part I: Stability and dissipativity, Int. J. Control, 74 (2001), 1631–1658. https://doi.org/10.1080/00207170110081705 doi: 10.1080/00207170110081705
![]() |
[18] |
S. G. Nersesov, W. M. Haddad, Control vector Lyapunov functions for large-scale impulsive dynamical system, Nonlinear Anal. Hybrid Syst., 1 (2007), 223–243. https://doi.org/10.1016/j.nahs.2006.10.006 doi: 10.1016/j.nahs.2006.10.006
![]() |
[19] |
Q. Xi, Z. L. Liang, X. D. Li, Uniform finite-time stability of nonlinear impulsive time-varying systems, Appl. Math. Modell., 91 (2021), 913–922. https://doi.org/10.1016/j.apm.2020.10.002 doi: 10.1016/j.apm.2020.10.002
![]() |
[20] |
Q. Xi, X. Z. Liu, X. D. Li, Practical finite-time stability of nonlinear systems with delayed impulsive control, IEEE Trans. Syst. Man Cybern.: Syst., 53 (2023), 7317–7325. https://doi.org/10.1109/TSMC.2023.3296481 doi: 10.1109/TSMC.2023.3296481
![]() |
[21] |
S. W. Zhao, J. T. Sun, L. Liu, Finite-time stability of linear time-varying singular systems with impulsive effects, Int. J. Control, 81 (2008), 1824–1829. https://doi.org/10.1080/00207170801898893 doi: 10.1080/00207170801898893
![]() |
[22] |
M. A. Jamal, R. Kumar, S. Mukhopadhyay, S. Das, Fixed-time stability of dynamical systems with impulsive effects, J. Franklin Inst., 359 (2022), 3164–3182. https://doi.org/10.1016/j.jfranklin.2022.02.016 doi: 10.1016/j.jfranklin.2022.02.016
![]() |
[23] |
Q. H. Wang, A. Abdurahman, Fixed-time stability analysis of general impulsive systems and application to synchronization of complex networks with hybrid impulses, Neurocomputing, 601 (2024), 128218. https://doi.org/10.1016/j.neucom.2024.128218 doi: 10.1016/j.neucom.2024.128218
![]() |
[24] |
H. F. Li, C. D. Li, T. W. Huang, D. Q. Ouyang, Fixed-time stability and stabilization of impulsive dynamical systems, J. Franklin Inst., 354 (2017), 8626–8644. https://doi.org/10.1016/j.jfranklin.2017.09.036 doi: 10.1016/j.jfranklin.2017.09.036
![]() |
[25] |
X. Y. He, X. D. Li, S. J. Song, Prescribed-time stabilization of nonlinear systems via impulsive regulation, IEEE Trans. Syst. Man Cybern.: Syst., 53 (2022), 981–985. https://doi.org/10.1109/TSMC.2022.3188874 doi: 10.1109/TSMC.2022.3188874
![]() |
[26] |
X. N. Li, H. Q. Wu, J. D. Cao, A new prescribed-time stability theorem for impulsive piecewise-smooth systems and its application to synchronization in networks, Appl. Math. Modell., 115 (2023), 385–397. https://doi.org/10.1016/j.apm.2022.10.051 doi: 10.1016/j.apm.2022.10.051
![]() |
[27] |
M. Coraggio, P. De Lellis, M. di Bernardo, Convergence and synchronization in networks of piecewise-smooth systems via distributed discontinuous coupling, Automatica, 129 (2021), 109596. https://doi.org/10.1016/j.automatica.2021.109596 doi: 10.1016/j.automatica.2021.109596
![]() |
[28] |
L. Dieci, C. Elia, Master stability function for piecewise smooth Filippov networks, Automatica, 152 (2023), 110939. https://doi.org/10.1016/j.automatica.2023.110939 doi: 10.1016/j.automatica.2023.110939
![]() |
[29] |
J. Chen, X. R. Li, X. Q. Wu, G. B. Shen, Prescribed-time synchronization of complex dynamical networks with and without time-varying delays, IEEE Trans. Network Sci. Eng., 9 (2022), 4017–4027. https://doi.org/10.1109/TNSE.2022.3191348 doi: 10.1109/TNSE.2022.3191348
![]() |
[30] |
L. L. Xu, X. W. Liu, Prescribed-time synchronization of multiweighted and directed complex networks, IEEE Trans. Autom. Control, 68 (2023), 8208–8215. https://doi.org/10.1109/TAC.2023.3292148 doi: 10.1109/TAC.2023.3292148
![]() |
[31] |
J. R. Yang, G. C. Chen, S. Zhu, S. P. Wen, J. H. Hu, Fixed/prescribed-time synchronization of BAM memristive neural networks with time-varying delays via convex analysis, Neural Networks, 163 (2023), 53–63. https://doi.org/10.1016/j.neunet.2023.03.031 doi: 10.1016/j.neunet.2023.03.031
![]() |
[32] |
Q. Tang, S. C. Qu, C. Zhang, Z. W. Tu, Y. T. Cao, Effects of impulse on prescribed-time synchronization of switching complex networks, Neural Networks, 174 (2024), 106248. https://doi.org/10.1016/j.neunet.2024.106248 doi: 10.1016/j.neunet.2024.106248
![]() |
[33] |
F. V. Difonzo, Isochronous attainable manifolds for piecewise smooth dynamical systems, Qual. Theory Dyn. Syst., 21 (2022), 6. https://doi.org/10.1007/s12346-021-00536-z doi: 10.1007/s12346-021-00536-z
![]() |
[34] | A. F. Filippov, Differential Equations with Discontinuous Righthand Sides: Control Systems, Springer Science & Business Media, Berlin, 2013. https://doi.org/10.1007/978-94-015-7793-9 |
[35] |
J. Cortes, Discontinuous dynamical systems, IEEE Control Syst. Mag., 28 (2008), 36–73. https://doi.org/10.1109/MCS.2008.919306 doi: 10.1109/MCS.2008.919306
![]() |
[36] |
Y. D. Song, Y. J. Wang, J. Holloway, M. Krstic, Time-varying feedback for regulation of normal-form nonlinear systems in prescribed finite time, Automatica, 83 (2017), 243–251. https://doi.org/10.1016/j.automatica.2017.06.008 doi: 10.1016/j.automatica.2017.06.008
![]() |
[37] |
A. Bacciotti, F. Ceragioli, Stability and stabilization of discontinuous systems and nonsmooth Lyapunov functions, ESAIM. Control. Optim. Calc. Var., 4 (1999), 361–376. https://doi.org/10.1051/cocv:1999113 doi: 10.1051/cocv:1999113
![]() |
[38] |
Y. J. Shen, X. H. Xia, Semi-global finite-time observers for nonlinear systems, Automatica, 44 (2008), 3152–3156. https://doi.org/10.1016/j.automatica.2008.05.015 doi: 10.1016/j.automatica.2008.05.015
![]() |
[39] |
Y. J. Shen, Y. H. Huang, Uniformly observable and globally Lipschitzian nonlinear systems admit global finite-time observers, IEEE Trans. Autom. Control, 54 (2009), 2621–2625. https://doi.org/10.1109/TAC.2009.2029298 doi: 10.1109/TAC.2009.2029298
![]() |
[40] |
P. J. Ning, C. C. Hua, K. Li, H. Li, A novel theorem for prescribed-time control of nonlinear uncertain time-delay systems, Automatica, 152 (2023), 111009. https://doi.org/10.1016/j.automatica.2023.111009 doi: 10.1016/j.automatica.2023.111009
![]() |
[41] |
L. C. Feng, L. Liu, J. D. Cao, L. Rutkowski, G. P. Lu, General decay stability for nonautonomous neutral stochastic systems with time-varying delays and Markovian switching, IEEE Trans. Cybern., 52 (2022): 5441–5453. https://doi.org/10.1109/TCYB.2020.3031992 doi: 10.1109/TCYB.2020.3031992
![]() |
[42] |
D. Chalishajar, D. Kasinathan, R. Kasinathan, R. Kasinathan, Exponential stability, T-controllability and optimal controllability of higher-order fractional neutral stochastic differential equation via integral contractor, Chaos, Solitons Fractals, 186 (2024), 115278. https://doi.org/10.1016/j.chaos.2024.115278 doi: 10.1016/j.chaos.2024.115278
![]() |
[43] |
D. Chalishajar, R. Kasinathan, Kasinathan R, S. Varshini, On solvability and optimal controls for impulsive stochastic integrodifferential varying-coefficient model, Automatika, 65 (2024), 1271–1283. https://doi.org/10.1080/00051144.2024.2361212 doi: 10.1080/00051144.2024.2361212
![]() |
[44] |
D. Chalishajar, R. Kasinathan, R. Kasinathan, Existence and Stability Results for Time-Dependent Impulsive Neutral Stochastic Partial Integrodifferential Equations with Rosenblatt Process and Poisson Jumps, Tatra Mt. Math. Publ., 00 (2024), 1–24. https://doi.org/10.2478/tmmp-2024-0002 doi: 10.2478/tmmp-2024-0002
![]() |