
The current manuscript focuses on the solution and analysis of space and time fractional coupled Schrödinger system that belongs to a class of evolution equations. These systems encounter in different fields related to plasma waves, optics, and quantum physics. The fractional He-Laplace approach is proposed for the series form solutions of fractional systems. This approach contains hybrid of Laplace transform and homotopy perturbation along with Caputo fractional derivative. The current study provide new results on time and space fractional coupled Schrödinger systems which are not captured in existing literature. Reliability of proposed algorithm in both time and space fractional scenarios is observed through residual error concept throughout fractional domains. The effect of fractional parameters on wave profiles are analyzed numerically and graphically as 2D and 3D illustrations. Analysis reveals that proposed algorithm is suitable for non-linear time-space fractional systems encountering in different fields of sciences.
Citation: Mubashir Qayyum, Efaza Ahmad, Hijaz Ahmad, Bandar Almohsen. New solutions of time-space fractional coupled Schrödinger systems[J]. AIMS Mathematics, 2023, 8(11): 27033-27051. doi: 10.3934/math.20231383
[1] | Kunting Yu, Yongming Li . Adaptive fuzzy control for nonlinear systems with sampled data and time-varying input delay. AIMS Mathematics, 2020, 5(3): 2307-2325. doi: 10.3934/math.2020153 |
[2] | Guijun Xing, Huatao Chen, Zahra S. Aghayan, Jingfei Jiang, Juan L. G. Guirao . Tracking control for a class of fractional order uncertain systems with time-delay based on composite nonlinear feedback control. AIMS Mathematics, 2024, 9(5): 13058-13076. doi: 10.3934/math.2024637 |
[3] | Fang Zhu, Pengtong Li . Adaptive fuzzy consensus tracking control of multi-agent systems with predefined time. AIMS Mathematics, 2025, 10(3): 5307-5331. doi: 10.3934/math.2025245 |
[4] | Qike Zhang, Tao Xie, Wenxiang Fang . Fixed/predefined-time generalized synchronization for stochastic complex dynamical networks with delays. AIMS Mathematics, 2024, 9(3): 5482-5500. doi: 10.3934/math.2024266 |
[5] | Xuelian Jin . Exponential stability analysis and control design for nonlinear system with time-varying delay. AIMS Mathematics, 2021, 6(1): 102-113. doi: 10.3934/math.2021008 |
[6] | Chao Ma, Tianbo Wang, Wenjie You . Master-slave synchronization of Lurie systems with time-delay based on event-triggered control. AIMS Mathematics, 2023, 8(3): 5998-6008. doi: 10.3934/math.2023302 |
[7] | Miao Xiao, Zhe Lin, Qian Jiang, Dingcheng Yang, Xiongfeng Deng . Neural network-based adaptive finite-time tracking control for multiple inputs uncertain nonlinear systems with positive odd integer powers and unknown multiple faults. AIMS Mathematics, 2025, 10(3): 4819-4841. doi: 10.3934/math.2025221 |
[8] | Gaoran Wang, Weiwei Sun, Shuqing Wang . Stabilization of wind farm integrated transmission system with input delay. AIMS Mathematics, 2021, 6(9): 9177-9193. doi: 10.3934/math.2021533 |
[9] | Qiaoping Li, Yu Chen, Lingyuan Ma . Predefined-time control of chaotic finance/economic system based on event-triggered mechanism. AIMS Mathematics, 2023, 8(4): 8000-8017. doi: 10.3934/math.2023404 |
[10] | Weiwei Sun, Mengyang Qiu, Xinyu Lv . H∞ filter design for a class of delayed Hamiltonian systems with fading channel and sensor saturation. AIMS Mathematics, 2020, 5(4): 2909-2922. doi: 10.3934/math.2020188 |
The current manuscript focuses on the solution and analysis of space and time fractional coupled Schrödinger system that belongs to a class of evolution equations. These systems encounter in different fields related to plasma waves, optics, and quantum physics. The fractional He-Laplace approach is proposed for the series form solutions of fractional systems. This approach contains hybrid of Laplace transform and homotopy perturbation along with Caputo fractional derivative. The current study provide new results on time and space fractional coupled Schrödinger systems which are not captured in existing literature. Reliability of proposed algorithm in both time and space fractional scenarios is observed through residual error concept throughout fractional domains. The effect of fractional parameters on wave profiles are analyzed numerically and graphically as 2D and 3D illustrations. Analysis reveals that proposed algorithm is suitable for non-linear time-space fractional systems encountering in different fields of sciences.
Actual nonlinear systems are subject to a wide variety of unknown disturbances, parameter uncertainties and delay phenomena [1,2]. Robust control and adaptive control methods have been proven successful at addressing these issues [3,4,5,6,7,8,9,10,11,12,13,14]. For nonlinear time-delay systems (NTDSs) with variable powers, [4] solved the adaptive robust tracking control problem. Taking into account the disturbance and time-varying delay of nonlinear Lipschitz systems, [5] designed a class of delay-dependent H∞ dynamic observers. The literature [6] investigated the global adaptive state feedback control issue of stochastic NTDSs with unknown disturbances and time-dependent delays. A new adaptive tracking controller with minimum learning parameters was designed to solve an adaptive tracking issue in finite time caused by full-state constrained NTDSs under input saturation [8]. The Lyapunov-Krasovskii functional method was applied to investigate the adaptive robust control problem on finite time of NTDSs subject to uncertainty and disturbances [10]. For NTDSs with the triangular form, the robust output tracking control problem is solved by neural network [11]. For stochastic NTDSs with partially known jump rates, [12] investigated the continuous gain-scheduled robust fault detection issue. The above results developed several good infinite-time and finite-time methods. Although the finite-time control scheme presented above achieves faster convergence and greater robustness than the infinite-time scheme, its convergence time is determined by giving the initial state. Namely, its convergence time has an uncertain upper bound.
As a result, a predefined-time control method is presented. As is well known, a closed-loop system (CLS) operated under a predefined-time strategy will maintain its convergence time uncorrelated with its initial state, and the upper bound of its convergence time is set beforehand, which improves the control accuracy of the CLS and compensates for the limitations of the finite-time control scheme. Consequently, the predefined-time control approach has received considerable focus from scholars, and some excellent results have been produced [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33] on the mechanical system, time-varying system and uncertain system. Meanwhile, many effective predefined-time control methods have also been proposed like optimal control [16], time-varying feedback [17,18,19], backstepping [20], high-gain approach [26] and so on. Among them, the Lyapunov-like theorem proposed in [25] ensured that such dynamic systems reached stability in a predefined time and also constituted a framework based on predefined-time stability analysis. By introducing time transformations with terminal time terms, [26] investigated the stability in a predefined time of the nonlinear uncertain system. With a new scale-free backstepping method, [28] solved the predefined-time mean-square stability as well as the inverse optimality issues of nonlinear stochastic systems. By designing a double time-varying gain, [29] presented a new predefined-time adaptive control method and solved the predefined-time control problems involved in triangular NTDSs having unknown parameters. Even so, there are few predefined-time control results of NTDSs except [29]. Particularly, as far as the authors know, no predefined-time control result has been published regarding the NTDSs with external disturbances and parameter uncertainty simultaneously, which inspired the present work.
As is well known, it is a very difficult task to study nonlinear systems due to lack of an effective research tool. Recently, the Hamiltonian method, as an important nonlinear research tool, has received much attention in the academic community and has been used to solve succesfully a large number of nonlinear problems. In fact, one of the advantages of using the Hamiltonian function method is that the Hamiltonian function in a Hamiltonian system can be chosen as a candidate Lyapunov function, which effectively overcomes the difficulty of constructing Lyapunov functions. Moreover, several well-established techniques for the design of Hamiltonian systems are given in the literature [34,35,36]. Among them, the authors also develop many strategies to obtain their Hamiltonian forms for given nonlinear systems. These methods are presented so that one can easily convert the system under study into its Hamiltonian version. However, the fact that the powers of the Hamiltonian functions in the literature [9,14,37,38] are the same for each state implies that the results have greater limitations.
In the paper, we investigate a general class of NTDSs with uncertain parameter and external disturbance, and propose several predefined-time H∞ and adaptive robust control results by applying the Hamiltonian function method with different powers. Here are some of the main contributions of this paper: (1) Contrary to the available results of infinite-time control as well as finite-time control, our predefined-time controller is capable of ensuring the CLS converges to zero within any given predefined constant number, and its convergence time cannot be affected by the initial state, making the system behavior better determined. (2) In contrast with the previous NTDS control results using the Hamiltonian approach (with the same power Hamiltonian form), this paper develops several more general results with different power Hamiltonian function forms, implying that these results have broader applications. (3) Compared with the recent predefined-time control result on the delay system in [29] (considering only the triangular form and unknown parameter), this paper examines a general class of NTDSs subject to uncertainties and disturbances and designs its predefined-time H∞ controller and adaptive robust controller by applying different power Hamiltonian forms. As expected, the results are more in line with reality.
The remaining sections of the paper are listed below. In Section 2, we present several preliminaries. Section 3 gives our main findings, Section 4 illustrates a case of river pollution that supports our new result and Section 5 is the conclusion.
Consider the system
˙x=f(t,x,x(t−h),ρ), | (2.1) |
where f:R+×Rn×Rn→Rn indicates a nonlinear function, x∈Rn represents the system state, h>0 is the constant delay, ρ denotes an uncertain constant parameter with appropriate dimension, and t indicates time variable defined on [t0,∞), where t0∈R+∪{0}, x0=x(t0) is the initial condition.
In some practical cases, it would be desirable if the system (2.1) could achieve its original point in a given time Ta, where Ta is a constant and can be defined beforehand.
Assume that M(⊂Rn) is a non-empty set. If any solution x(t,x0) reaches M in t≤t0+Ta, then the predefined time Ta gets defined.
Lemma 2.1. Suppose that V(t,xt):=V is a continuous radial unbounded function with V(t,0)=0 such that
˙V≤−1qTaexp(Vq)V1−q,x∈M, | (2.2) |
holds, then M is predefined-time attractive, where xt is a delay function segment defined as xt:=x(t+ϱ),ϱ∈[−h,0] with h>0, the constant 0<q≤1 and Ta is the predefined time.
Proof. From Eq (2.2), it is easy to obtain:
V(t,xt)≤[ln(1t−t0Ta+exp(−(Vx0)q))]1q. | (2.3) |
To give the upper boundary of t, using V(t,0)=0, we can obtain
t≤t0+Ta[1−exp(−(Vx0)q)],∀x0∈Rn, | (2.4) |
where Vx0 denotes the initial value of V. From Eq (2.4), and 0<exp(−(Vx0)q)≤1, we have
t≤t0+Ta,∀x0∈Rn, | (2.5) |
implying Ta is the predefined time. Therefore, it has been proved.
Lemma 2.2. [39] For real matrices ϝa,ϝb and a positive real number τ, the following equation holds:
ϝTaϝb+ϝTbϝa⩽τϝTaϝa+τ−1ϝTbϝb. | (2.6) |
Lemma 2.3. [35] When a scalar function g(x) has continuous partial derivatives of order n and g(0)=0(x∈Rn), then g(x) may be characterized as g(x)=l1(x)x1+⋯+ln(x)xn with li(x)(i=1,2,⋯,n) also being scalar functions.
Assuming k(x)(∈Rn) is a smooth function and k(0)=0, then k(x):=M(x)x=
[l11(x)l12(x)⋯l1n(x)l21(x)l22(x)⋯l2n(x)⋮⋮⋮⋮ln1(x)ln2(x)⋯lnn(x)][x1x2⋯xn]. | (2.7) |
Lemma 2.4. [13] For any given real number d≥1, the inequality holds:
nd−1d(n∑i=1|xi|)1d≥n∑i=1|xi|1d≥(n∑i=1|xi|)1d. | (2.8) |
Consider the following NTDS
˙x=[J(x)−R(x)]∇xH(x)+T(x)∇˜xH(˜x)+g1(x)u+g2(x)ω, | (3.1) |
where the state vector is x(t)(∈Ω), Ω represents the bounded convex neighborhood of the zero point inside the space Rn, inverse symmetric structure matrix J(x)(∈Rn×n) and positive definite symmetric matrix R(x)(∈Rn×n) are given, ˜x:=x(t−h), Hamiltonian function H(x) reaches its minimum when x=0 that is H(0)=0, the gradient vector of H(x) is denoted by ∇TxH(x), T(x)∈Rn×n with T(0)=0, g1(x) and g2(x) are the weighted matrices, u represents system input, ω denotes outside disturbance, and φ(η) denotes a vector-valued initial value function. Moreover, assume that g1(x) has full column rank.
To present several predefined-time control results on the Hamiltonian system (3.1), we assume that
H(x)=n∑i=1(x2i)αi2αi−1(αi>1,i=1,⋯,n) | (3.2) |
and y=gT1(x)∇xH(x) is the system output signal.
Remark 3.1. Note that the Hamiltonian function of the present paper has different powers, which is different from the existing results on the Hamiltonian system [9,14,37,38]. It should be mentioned as well that we presume the Hamiltonian function takes the form (3.2) so as to get certain predefined-time control outcomes on the Hamiltonian system (3.1). Obviously, a real system cannot satisfy the form. For a specific system, we have the following methods to develop its Hamiltonian form: the orthogonal decomposition method [40], Hamiltonian functional method [40], the energy shaping approach [40], the vector field decomposition method [41] and so on. In this section, we only present a theory result. Once the theory result is obtained, then one can design a suitable controller and apply the energy shaping approach to convert the general form to the form (3.2). Please refer to Section 4 for more details.
Assumption 3.1. The disturbance ω satisfies
Λ={ω∈Rq:μ2∫+∞0ωT(t)ω(t)dt≤1}, | (3.3) |
where μ>0 is a given positive constant.
For the purpose of analysis, the set Ω is selected in the following form:
Ω:={x:αTjx⩽1,j=1,2,⋯,n}, | (3.4) |
where αj(j=1,2,⋯,n) are known constant vectors, which describe n edges of Ω.
We begin by defining the predefined-time H∞ stabilization issue.
The predefined-time H∞ stabilization is to: Design a controller u=a(x) so that the CLS is stable for a predefined time when ω disappears. Meanwhile, when ω∈Λ is not zero, the zero-state response (φ(η)=0, η∈[−h,0]) meets
∫t0∥z(s)∥2ds≤γ2∫t0∥ω(s)∥2ds,0<t<∞, | (3.5) |
where positive constant γ represents the disturbance suppression level, while z denotes the penalty signal described by
z=h(x)y=h(x)g1T(x)∇xH(x), | (3.6) |
with the weighted matrix h(x) having the appropriate dimension.
Now, the main results of our study are presented.
Theorem 3.1. Considering the system (3.1) under Assumption 3.1, for Ta>0 and γ>0, if
(1) constant matrices L<0, P>0 and constant numbers ϵ>0, 0<r≤γ2 exist, the following inequalities hold:
−2R(x)+ϵ−1In+r−1g2(x)g2T(x)−1γ2g1(x)g1T(x)≤L, | (3.7) |
2∇xTH(x)∇xH(x)P−ϵTT(x)T(x)≥0, | (3.8) |
(2) real numbers s>0 and μ>0 exist and
[2s−δγ22μ2−sαTj−sαjIn]⩾0,j=1,2,⋯,n, | (3.9) |
then a predefined-time H∞ controller of the NTDS (3.1) could be constructed as follows:
g1(x)u=−1σTaΦ(x)−∇xH(x)∇˜xTH(˜x)P∇˜xH(˜x)+g1(x)v, | (3.10) |
v=−12[hT(x)h(x)+1γ2Im]g1T(x)∇xH(x), | (3.11) |
where δ:=max{c2αmin−22αmin−1,c2αmax−22αmax−1} with c:=max{‖x‖:x∈Ω}, \sigma : = W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}} with H(x)\leq W when x \in \Omega , \alpha_{min } represents the minimum value of \alpha_{i} , \alpha_{max } represents the maximum value of \alpha_{i} , and \Phi(x) is the predefined-time stabilizing function that follows
\begin{equation} \begin{aligned} \Phi(x) = 2^{\frac{1-3\alpha_{min}}{\alpha_{min}}}\dfrac{\alpha_{min}(2\alpha_{max}-1)^{2}}{\alpha_{max}^{2}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}] \nabla_{x}H(x).\end{aligned} \end{equation} | (3.12) |
Proof. To prove the system (3.1) is a predefined-time H_\infty stabilization, we first show (3.5) holds for \omega \neq 0 , and then prove its stability in a predefined time when \omega disappears.
The Lyapunov function is constructed in the following form:
\begin{equation} V(x) = 2H(x). \end{equation} | (3.13) |
Based on Lemma 2.2, one can derive
\begin{equation} 2\nabla_{x}^\mathrm{T}H(x)g_{2}(x)\omega\leq r\omega^\mathrm{T}\omega +r^{-1}\nabla_{x}^\mathrm{T}H(x)g_{2}(x)g_{2}^\mathrm{T}(x)\nabla_{x}H(x). \nonumber \end{equation} |
Computing the derivative of V(x) and using {\nabla_{x}}^\mathrm{T}H(x)J(x)\nabla_{x}H(x) = 0 , one can obtain
\begin{equation} \begin{aligned} \dot{V}(x) \leq&-2\nabla_{x}^\mathrm{T}H(x)R(x)\nabla_{x}H(x)+2\nabla_{x}^\mathrm{T}H(x)T(x)\nabla_{\tilde{x}}H(\tilde{x})+2\nabla_{x}^\mathrm{T}H(x)g_{1}(x) u+2\nabla_{x}^\mathrm{T}H(x)g_{2}(x)\omega\\ \leq&-2\nabla_{x}^\mathrm{T}H(x)R(x)\nabla_{x}H(x)+2\nabla_{x}^\mathrm{T}H(x)T(x)\nabla_{\tilde{x}}H(\tilde{x})+2\nabla_{x}^\mathrm{T}H(x)g_{1}(x) u+r\omega^\mathrm{T}\omega\\ &+r^{-1}\nabla_{x}^\mathrm{T}H(x)g_{2}(x)g_{2}^\mathrm{T}(x)\nabla_{x}H(x)\\ \leq& {\nabla_{x}}^\mathrm{T}H(x)[-2R(x)+r^{-1}g_{2}(x)g_{2}^\mathrm{T}(x)+\epsilon^{-1}I_{n}]\nabla_{x}H(x)+\nabla_{\tilde{x}}^\mathrm{T}H(\tilde{x})[-2\nabla_{x}^\mathrm{T}H(x)\nabla_{x}H(x)P\\ &+\epsilon T^\mathrm{T}(x)T(x)]\nabla_{\tilde{x}}H(\tilde{x})-2\nabla_{x}^\mathrm{T}H(x)\dfrac{1}{\sigma T_{a}}\Phi(x)+2\nabla_{x}^\mathrm{T}H(x)g_{1}(x)v+ r\parallel \omega \parallel^{2}\\ \leq& \nabla_{x}^\mathrm{T}H(x)[-2R(x)+\epsilon^{-1}I_{n}+r^{-1}g_{2}(x)g_{2}^\mathrm{T}(x)-\frac{1}{\gamma^{2}}g_{1}(x)g_{1}^\mathrm{T}(x)]\nabla_{x}H(x)\\ &+\nabla_{\tilde{x}}^\mathrm{T}H(\tilde{x})[-2\nabla_{x}^\mathrm{T}H(x)\nabla_{x}H(x)P+\epsilon T^\mathrm{T}(x)T(x)]\nabla_{\tilde{x}}H(\tilde{x})-2\nabla_{x}^\mathrm{T}H(x)\dfrac{1}{\sigma T_{a}}\Phi(x)\\ &-\parallel z \parallel^{2}+ r\parallel \omega \parallel^{2}\\ \leq& -\frac{2^{\frac{1-2\alpha_{min}}{\alpha_{min}}}}{\sigma T_{a}}\dfrac{\alpha_{min}(2\alpha_{max}-1)^{2}}{\alpha_{max}^{2}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}]\nabla_{x}^\mathrm{T}H(x)\nabla_{x}H(x)-\parallel z \parallel^{2}+ r\parallel \omega \parallel^{2}. \end{aligned} \end{equation} | (3.14) |
Noting that \frac{2 \alpha_{i}}{2 \alpha_{i}-1} = \frac{1}{1-\frac{1}{2 \alpha_{i}}} is a decreasing function, we have
\begin{equation} \begin{aligned} {\nabla_{x}}^\mathrm{T}H(x)\nabla_{x}H(x) = \sum\limits_{i = 1}^{n}\left(\frac{2 \alpha_{i}}{2 \alpha_{i}-1}\right)^{2}\left(x_{i}^{2}\right)^{\frac{1}{2 \alpha_{i}-1}} \geq\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2} \sum\limits_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{1}{2 \alpha_{i}-1}}. \end{aligned} \end{equation} | (3.15) |
Next, when x \in \Omega , let H(x)\leq W . Since H(x)\leq W , we can obtain that \dfrac{1}{W}\sum_{i = 1}^{n}\left(x_{i}\right)^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}} < 1 . From that and Lemma 2.4, we get
\begin{equation} \begin{aligned} \sum\limits_{i = 1}^{n}\left(\frac{2 \alpha_{i}}{2 \alpha_{i}-1}\right)^{2}\left(x_{i}^{2}\right)^{\frac{1}{2 \alpha_{i}-1}} & \geq\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2} \sum\limits_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{1}{2 \alpha_{i}-1}} \\ & = \left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2} \sum\limits_{i = 1}^{n}\left[\left(x_{i}^{2}\right)^{\frac{\alpha_{i}}{2 \alpha_{i}-1}}\right]^{\frac{1}{\alpha_{i}}}\\ &\geq\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2}\left[\sum\limits_{i = 1}^{n}\left(x_{i}\right)^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}\right]^{\frac{1}{\alpha_{i }}}\\ & = \left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2}W^{\frac{1}{\alpha_{i}}}\left[\dfrac{1}{W}\sum\limits_{i = 1}^{n}\left(x_{i}\right)^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}\right]^{\frac{1}{\alpha_{i }}}\\ &\geq\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2}W^{\frac{1}{\alpha_{i}}}\left[\dfrac{1}{W}\sum\limits_{i = 1}^{n}\left(x_{i}\right)^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}\right]^{\frac{1}{\alpha_{min }}}\\ & = \left(\frac{2 \alpha_{max }}{2 \alpha_{max}-1}\right)^{2}W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}}(H(x))^{\frac{1}{\alpha_{min }}}.\\ \end{aligned} \end{equation} | (3.16) |
Note that the monotonicity of W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}} is related to the value range of W . When 0 < W\leq1 , W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}} is an increasing function of \alpha_i , then W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}}\geq W^{\frac{1}{\alpha_{min}}-\frac{1}{\alpha_{min }}} = 1 and if W > 1 , W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}} is a decreasing function on \alpha_i , we have W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}}\geq W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}} . Combining these two cases, we have W^{\frac{1}{\alpha_{i}}-\frac{1}{\alpha_{min }}}\geq\min\{1, \; W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}}\} = W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}} : = \sigma with \sigma being a positive constant. Substituting it into Eq (3.16), we have \sum_{i = 1}^{n}\left(\frac{2 \alpha_{i}}{2 \alpha_{i}-1}\right)^{2}\left(x_{i}^{2}\right)^{\frac{1}{2 \alpha_{i}-1}}\geq\sigma\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2}(H(x))^{\frac{1}{\alpha_{min }}} , namely,
\begin{equation} \nabla_{x}^{T} H(x) \nabla_{x} H(x) \geq\sigma\left(\frac{2 \alpha_{max }}{2 \alpha_{max }-1}\right)^{2}(H(x))^{\frac{1}{\alpha_{min }}} . \end{equation} | (3.17) |
Substituting (3.17) into (3.14), one can get
\begin{equation} \begin{aligned} \dot{V}(x)\leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}](2H(x))^{\frac{1}{\alpha_{min}}}-\parallel z \parallel^{2}+ r\parallel \omega \parallel^{2}. \end{aligned} \end{equation} | (3.18) |
Now, we prove (3.5) holds.
By letting \Gamma(t, x) = V(x)+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds , we next indicate that \Gamma(t, x)\leq0 .
Noting that r \leq \gamma^{2} , we obtain
\begin{equation} \begin{aligned} \dot{\Gamma}(t, x) = &\dot{V}(x)+\parallel z \parallel^{2}-\gamma^{2}\parallel \omega \parallel^{2}\\ \leq&-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}](2H(x))^{\frac{1}{\alpha_{min}}}\\ &-\parallel z \parallel^{2}+ r\parallel \omega \parallel^{2}+\parallel z \parallel^{2}-\gamma^{2}\parallel \omega \parallel^{2}\\ \leq&(r-\gamma^{2})\parallel \omega \parallel^{2}\leq0. \end{aligned} \end{equation} | (3.19) |
Using the condition of zero-state response and integrating (3.19) over 0 and t result in
\begin{equation} V(x)+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds\leq 0. \end{equation} | (3.20) |
Since V(x)\geq0 , one obtains
\begin{equation} \int_{0}^{t}\parallel z \parallel^{2}ds\leq\gamma^{2}\int_{0}^{t}\parallel \omega(s) \parallel^{2})ds. \end{equation} | (3.21) |
In addition, noting that \frac{\alpha_{i}}{2 \alpha_{i}-1} < 1 and x^{T} x = \sum_{i = 1}^{n}\left(x_{i}\right)^{2} , we have x^{T}x = c^{2} \sum_{i = 1}^{n}\left(\frac{x_{i}}{c}\right)^{2} \leq c^{2} \sum_{i = 1}^{n}\left[\left(\frac{x_{i}}{c}\right)^{2}\right]^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = \frac{c^{2}}{c^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}} \sum_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} H(x) holds on \Omega .
The monotonicity of c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is related to the value range of c . When 0 < c\leq1 , c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is a decreasing function of \alpha_i , then c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq c^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}} and if c > 1 , c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is an increasing function of \alpha_i , we have c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq c^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}} . Combining these two equations, we can reach the conclusion that c^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq\max\{c^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}}, \; c^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}}\} : = \delta , from which we have x^{T}x\leq\delta H(x) .
Noting Assumption 3.1 and (3.20), we have x^{T}x\leq \delta H(x) = \frac{1}{2}\delta V\left(t, x\right) \leq \delta \frac{\gamma^{2}}{2} \int_{0}^{T}\|w(s)\|^{2} d s \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} , namely, \|x\|^{2} \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} .
Afterward, it is shown that x(t) \in \Omega holds when \forall t > 0 , \varphi = 0 , \omega \in \Lambda . With (3.4), it must be demonstrated that
\begin{equation} x^{T} x-\frac{\delta \gamma^{2}}{2 \mu^{2}} \leq 0, \text { s.t. } 2-2 \alpha_{j}^{T} x \geq 0 (j = 1, \cdots, n). \end{equation} | (3.22) |
From [42], we have
\begin{equation} \zeta^{T}\left[\begin{array}{cc} 2 s-\frac{\delta \gamma^{2}}{2 \mu^{2}}\; \; & -s \alpha_{j}^{T} \\ -s \alpha_{j} & I_{n} \end{array}\right] \zeta \geq 0(j = 1, 2, \cdots, n), \end{equation} | (3.23) |
which indicates that (3.9) holds with free scalar s > 0 introduced by the S-procedure and \zeta = \left[1, x^{T}\right]^{T} .
Thus, we conclude that x(t) remains in \Omega for all t > 0 , \varphi = 0 , \omega\in \Lambda .
Next, we demonstrate the stability in a predefined time for the CLS (3.1) when \omega disappears.
Noting (3.18), we get
\begin{equation} \begin{aligned} \dot{V}(x)\leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}](2H(x))^{\frac{1}{\alpha_{min}}}-\parallel z \parallel^{2}\\ \leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2H(x))^{1-\frac{1}{\alpha_{min}}}](2H(x))^{\frac{1}{\alpha_{min}}}\\ = &-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[V^{1-\frac{1}{\alpha_{min}}}]V^{\frac{1}{\alpha_{min}}}. \end{aligned} \end{equation} | (3.24) |
It implies Lemma 2.1 holds. Thus, it has been proved.
In Theorem 3.1, the controller (3.10) contains the delay term. Next, we present a controller without containing the delay term for the NTDS (3.1).
Theorem 3.2. Consider the NTDS (3.1) under Assumption 3.1. For the presented T_{a} > 0 and \gamma > 0 , if
(1) constant matrices L < 0, P > 0 and constant numbers \epsilon > 0 , r > 0 exist such that e^{\left(h \lambda_{\max }\{P\}\right) \chi^{2}}r\leq \gamma^{2} ,
\begin{equation} \begin{aligned} 2H(x)P-2R(x)+\epsilon^{-1}I_{n}+r^{-1}g_{2}(x)g_{2}^\mathrm{T}(x)-\frac{1}{\gamma^{2}}g_{1}(x)g_{1}^\mathrm{T}(x)\leq L < 0, \end{aligned} \end{equation} | (3.25) |
\begin{equation} -\epsilon T^\mathrm{T}(x)T(x)+2H(x)P\geq 0, \end{equation} | (3.26) |
(2) real numbers s > 0 and \mu > 0 exist, and
\begin{equation} \left[\begin{array}{cc} 2 s-\frac{\delta \gamma^{2}}{2 \mu^{2}}\; \; \; & -s \alpha_{j}^{\mathrm{T}} \\ -s \alpha_{j} & I_{n} \end{array}\right] \geqslant 0, j = 1, 2, \cdots, n, \end{equation} | (3.27) |
where \delta: = \max\{\chi^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}}, \; \chi^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}}\} with \chi: = \max \left\{\left\|x_{t}\right\|: x \in \Omega\right\} , then a predefined-time H_\infty controller of the NTDS (3.1) could be constructed as follows:
\begin{equation} g_{1}(x)u = -\dfrac{1}{\sigma T_{a}}\Phi(x)+g_{1}(x)v, \end{equation} | (3.28) |
\begin{equation} v = -\frac{1}{2}[h^\mathrm{T}(x)h(x)+\frac{1}{\gamma^{2}}I_{m}]g_{1}^\mathrm{T}(x)\nabla_{x}H(x), \end{equation} | (3.29) |
where \sigma : = W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}} with H(x)\leq W when x \in \Omega , \alpha_{min } and \alpha_{max } are similar to these in Theorem 3.1, and \Phi(x) is the predefined-time stabilizing function that follows
\begin{equation} \begin{aligned} \Phi(x) = &2^{\frac{1-3\alpha_{min}}{\alpha_{min}}}\dfrac{\alpha_{min}(2\alpha_{max}-1)^{2}}{\alpha_{max}^{2}(\alpha_{min}-1)}G(t)^{\dfrac{1-\alpha_{min }}{\alpha_{min }}}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}]\nabla_{x}H(x) \end{aligned} \end{equation} | (3.30) |
and G(t): = e^{\int_{t-h}^{t} \nabla_{x}^{T} H(x(s)) P \nabla_{x} H(x(s)) d s} .
Proof. To prove the system (3.1) has predefined-time H_\infty stability, we first show
\begin{equation} \int_{0}^{t}\parallel z \parallel^{2}ds\leq\gamma^{2}\int_{0}^{t}\parallel \omega(s) \parallel^{2})ds \end{equation} | (3.31) |
holds for \omega \neq 0 .
Construct the Lyapunov function as:
\begin{equation} \begin{aligned} V(t, x_t) = 2 e^{\int_{t-h}^{t} \nabla_{x}^{T} H(x(s)) P \nabla_{x} H(x(s)) d s} H(x): = 2 G(t) H(x). \end{aligned} \end{equation} | (3.32) |
Computing the derivative of V(t, x_{t}) and using \nabla_{x}^\mathrm{T}H(x)J(x)\nabla_{x}H(x) = 0 , one can obtain
\begin{equation} \begin{aligned} \dot{V}(t, x_{t})\leq&G(t)[2\nabla_{x}^\mathrm{T}H(x)H(x)P\nabla_{x}H(x)-2\nabla_{\tilde{x}}^\mathrm{T}H(\tilde{x})H(x)P\nabla_{\tilde{x}}H(\tilde{x})-2\nabla_{x}^\mathrm{T}H(x)R(x)\nabla_{x}H(x)\\ &+2\nabla_{x}^\mathrm{T}H(x)T(x)\nabla_{\tilde{x}}H(\tilde{x})+2\nabla_{x}^\mathrm{T}H(x)g_{1}(x) u+r{\omega}^\mathrm{T}\omega+r^{-1}\nabla_{x}^\mathrm{T}H(x)g_{2}(x){g_{2}}^\mathrm{T}(x)\nabla_{x}H(x)]\\ \leq&G(t)[ {\nabla_{x}}^\mathrm{T}H(x)[2H(x)P-2R(x)+\epsilon^{-1}I_{n}+r^{-1}g_{2}(x){g_{2}}^\mathrm{T}(x)]\nabla_{x}H(x)\\ &-{\nabla_{\tilde{x}}}^\mathrm{T}H(\tilde{x})[2H(x)P-\epsilon{T}^\mathrm{T}(x)T(x)]\nabla_{\tilde{x}}H(\tilde{x})-2\nabla_{x}^\mathrm{T}H(x)\dfrac{1}{\sigma T_{a}}\Phi(x)\\ &+2{\nabla_{x}}^\mathrm{T}H(x)g_{1}(x)v+ r\parallel \omega \parallel^{2}]\\ = & G(t)[{\nabla_{x}}^\mathrm{T}H(x)[2H(x)P-2R(x)+\epsilon^{-1}I_{n} +r^{-1}g_{2}(x){g_{2}}^\mathrm{T}(x)-\frac{1}{\gamma^{2}}g_{1}(x){g_{1}}^\mathrm{T}(x)]\nabla_{x}H(x)\\ &+{\nabla_{\tilde{x}}}^\mathrm{T}H(\tilde{x})[-2H(x)P+\epsilon{T}^\mathrm{T}(x)T(x)]\nabla_{\tilde{x}}H(\tilde{x}) -2\nabla_{x}^\mathrm{T}H(x)\dfrac{1}{\sigma T_{a}}\Phi(x)-\parallel z \parallel^{2}+ r\parallel \omega \parallel^{2}]\\ \leq& -\frac{2^{\frac{1-2\alpha_{min}}{\alpha_{min}}}}{\sigma T_{a}}\dfrac{\alpha_{min}(2\alpha_{max}-1)^{2}}{\alpha_{max}^{2}(\alpha_{min}-1)}G(t)^{\dfrac{1}{\alpha_{min }}}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}]{\nabla_{x}}^\mathrm{T}H(x)\nabla_{x}H(x)\\ &-G(t)\parallel z \parallel^{2}+ G(t) r\parallel \omega \parallel^{2}. \end{aligned} \end{equation} | (3.33) |
From that and using (3.17), one obtains
\begin{equation} \begin{aligned} \dot{V}(t, x_{t})\leq &-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}-G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2}. \end{aligned} \end{equation} | (3.34) |
Now, we prove (3.5) holds. Do so by letting \Gamma_{1}(t, x) = V(t, x_{t})+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds . We next indicate that \Gamma_{1}(t, x)\leq0 .
Assuming that \left\|x_{t}\right\| \leq \chi , it follows that G(t) \leq e^{\left(h \lambda_{\max }\{P\}\right) \chi^{2}} . Substituting \dot{V}(t, x_{t}) into \dot{\Gamma}_{1}(t, x) , and noting that e^{\left(h \lambda_{\max }\{P\}\right) \chi^{2}}r \leq \gamma^{2}, G(t)\geq1 , one gets
\begin{equation} \begin{aligned} \dot{\Gamma}_{1}(t, x) = &\dot{V}(t, x_{t})+\parallel z \parallel^{2}-\gamma^{2}\parallel \omega \parallel^{2}\\ \leq&-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ &-G(t)\parallel z \parallel^{2}+ G(t)r\parallel \omega \parallel^{2}+\parallel z \parallel^{2}-\gamma^{2}\parallel \omega \parallel^{2}\\ \leq&-(G(t)-1)\parallel z \parallel^{2}+(e^{\left(h \lambda_{\max }\{P\}\right) \chi^{2}}r-\gamma^{2})\parallel \omega \parallel^{2}\leq0. \end{aligned} \end{equation} | (3.35) |
The condition of zero-state response and the integration of \dot{\Gamma}_{1}(t, x) over 0 and t result in
\begin{equation} V(t, x_{t})+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds\leq 0, \end{equation} | (3.36) |
and under V(t, x_{t})\geq0 , one obtains
\begin{equation} \int_{0}^{t}\parallel z \parallel^{2}ds\leq\gamma^{2}\int_{0}^{t}\parallel \omega(s) \parallel^{2})ds. \end{equation} | (3.37) |
Using \frac{\alpha_{i}}{2 \alpha_{i}-1} < 1 as well as x^{T} x = \sum_{i = 1}^{n}\left(x_{i}\right)^{2} , we have x^{T} x = \chi^{2} \sum_{i = 1}^{n}\left(\frac{x_{i}}{\chi}\right)^{2} \leq \chi^{2} \sum_{i = 1}^{n}\left[\left(\frac{x_{i}}{\chi}\right)^{2}\right]^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = \frac{\chi^{2}}{\chi^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}} \sum_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} H(x) holds on \Omega .
The monotonicity of \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is related to the value range of \chi . When 0 < \chi\leq1 , \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is a decreasing function, then \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq \chi^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}} and if \chi > 1 , \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} is an increasing function, we have \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq \chi^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}} . Combining these two equations, we can conclude that \chi^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}}\leq\max\{\chi^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}}, \; \chi^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}}\} : = \delta , from which we have x^{T}x\leq\delta H(x) .
Noting Assumption 3.1, Eq (3.36) and G(t)\geq1 , one has x^{T}x\leq \delta H(x)\leq G(t)\delta H(x) = \frac{1}{2}\delta V\left(t, x_{t}\right) \leq \delta \frac{\gamma^{2}}{2} \int_{0}^{T}\|\omega(s)\|^{2} d s \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} , namely, \|x\|^{2} \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} .
Afterward, it is shown that x(t) \in \Omega holds when \forall t > 0 , \varphi = 0 , \omega \in \Lambda . With (3.4), it must be demonstrated that
\begin{equation} x^{T} x-\frac{\delta \gamma^{2}}{2 \mu^{2}} \leq 0, \text { s.t. } 2-2 \alpha_{j}^{T} x \geq 0 (j = 1, \cdots, n). \end{equation} | (3.38) |
From [42], we have
\begin{equation} \xi^{T}\left[\begin{array}{cc} 2 s-\frac{\delta \gamma^{2}}{2 \mu^{2}}\; \; & -s \alpha_{j}^{T} \\ -s \alpha_{j} & I_{n} \end{array}\right] \xi \geq 0(j = 1, 2, \cdots, n), \end{equation} | (3.39) |
which indicates that (3.27) holds with free scalar s > 0 introduced by the S-procedure and \xi = \left[1, x^{T}\right]^{T} . Thus, we conclude that x(t) remains in \Omega when all t > 0 , \varphi = 0 , \omega\in \Lambda .
As the remainder of the proof is similar to Theorem 3.1, it is omitted.
Now, we present an adaptive robust stabilization result with external disturbance and uncertainty.
Consider the NTDS
\begin{equation} \dot{x} = [J(x, p)-R(x, p)]\nabla_{x}H_{1}(x, p)+T(x)\nabla_{\tilde{x}}H_{1}(\tilde{x})+g_{1}(x) u+g_{2}(x)\omega, \end{equation} | (3.40) |
where p represents constant bounded uncertainty, inverse symmetric structure matrix J(x, p) and symmetric matrix R(x, p) are given, J (x, 0) = J (x) , R (x, 0) = R (x) , J (x, p) = J (x) + \Delta J (x, p) , R (x, p) = R (x) + \Delta R (x, p) , Hamiltonian function H_{1}(x) , which is smooth, reaches its minimum when x = 0 that is H_{1}(0) = 0 , H_{1} (x, 0) = H_{1} (x) , \nabla_{x}H_{1} (x, p) = \nabla_{x}H_{1}(x) + \Delta H_{1} (x, p) , u \in \mathbb{R}^{m_{1}} represents system input, \omega \in \mathbb{R}^{q} indicates the external interference satisfying \int_{0}^{\infty}\omega^\mathrm{T}(t)\omega(t)dt < \infty , and \varphi(\eta) denotes a vector-valued initial value function. Moreover, assume that g_{1} (x) has full column rank.
Assumption 3.2. [14,43,44,45] Assume \phi(x) satisfies
\begin{equation} [J(x, p)-R(x, p)]\Delta H_{1}(x, p) = g_{1}(x)\phi^\mathrm{T}(x)\theta \end{equation} | (3.41) |
for \forall x\in\Omega , where \theta \in R^{m_{2}} indicates the constant vector determined by the uncertain parameter p and assume the constant k > 0 exists such that \|\theta\|\leq k .
Under Assumption 3.2, noting that J (x, p) = J (x) + \Delta J (x, p) , R (x, p) = R (x) + \Delta R (x, p) , the system (3.40) can be transformed into:
\begin{equation} \begin{aligned} \dot{x}(t) = [J(x)-R(x)] \nabla_{x} H_{1}(x)+T(x) \nabla_{\tilde{x}} H_{1}(\tilde{x}) +g_{1}(x) u+g_{2}(x) \omega+g_{1}(x) \phi^{\mathrm{T}}(x) \theta+G(x, p), \end{aligned} \end{equation} | (3.42) |
where G(x, p): = [\triangle J(x, p)-\triangle R(x, p)] \nabla_{x} H_{1}(x) .
Based on the NTDS (3.40) and Lemma 2.3, we can easily know that matrix M(x) \in \mathbb{R}^{n \times n} as well as positive constant \varpi exist such that \nabla_{x}H_{1}(x) = M (x) x , and
\begin{equation} \parallel G (x, p) \parallel ^{2} \leq \varpi\parallel x \parallel ^{2}, x \in \Omega, \end{equation} | (3.43) |
where
\begin{equation} \begin{aligned} \varpi = : \lambda_{\max }\{{M}^\mathrm{T}(x)[\Delta J(x, p)-\Delta R(x, p)] ^\mathrm{T} [\Delta J (x, p) - \Delta R (x, p)] M (x)\}. \end{aligned} \end{equation} | (3.44) |
To investigate the predefined-time stabilizing issue, we need to convert the Hamiltonian function H_{1}(x) into:
\begin{equation} H(x) = \sum\limits_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{\alpha_{i}}{2 \alpha_{i}-1}}(\alpha_{i} > 1, i = 1, \cdots, n) . \end{equation} | (3.45) |
To do this, the controller u is designed as
\begin{equation} g_{1}(x)u = [J(x)-R(x)] \nabla_{x} H_{2}(x)+T(x) \nabla_{\tilde{x}} H_{2}(\tilde{x})+[\iota I_{n}-T(x)]\nabla_{\tilde{x}} H(\tilde{x}) +g_{1}(x) v, \end{equation} | (3.46) |
where v represents a new input, \iota > 0 , and \iota^{2} = dH(x) with d being a positive constant, H_{2}(x) : = H(x)-H_{1}(x), v = v_{1}+v_{2} with v_1 and v_2 being determined later.
Substituting (3.46) into (3.42), system (3.42) is expressed as
\begin{equation} \begin{aligned} \dot{x}(t) = [J(x)-R(x)] \nabla_{x} H(x)+\iota I_{n}\nabla_{\tilde{x}} H(\tilde{x}) +g_{1}(x) v+g_{2}(x)\omega+g_{1}(x) \phi^{\mathrm{T}}(x) \theta+G(x, p). \end{aligned} \end{equation} | (3.47) |
Consider the system (3.47), set \gamma > 0 as disturbance suppression level, and choose penalty signal z = h(x)g_{1}^\mathrm{T}(x)\nabla_{x}H(x) , where h(x) denotes the weight matrix of the appropriate dimension.
Theorem 3.3. Consider the NTDS (3.40) under Assumptions 3.1 and 3.2. For the presented T_{a} > 0 and \gamma > 0 , if
(1) constant matrices L , P > 0 , Q\geq\frac{1}{2}k^{2}I_{n} and constant numbers r > 0 , b > 0 exist such that e^{\left(h \lambda_{\max }\{P\}\right) a^{2}}r\leq \gamma^{2} ,
\begin{equation} \begin{aligned} 2H(x)P-2R(x)+r^{-1}g_{2}(x){g_{2}}^\mathrm{T}(x) -\frac{1}{\gamma^{2}}g_{1}(x){g_{1}}^\mathrm{T}(x)+2I_{n}+\varpi{N}^\mathrm{T}(x)N(x)\leq L < -b^{-1}I_{n}, \end{aligned} \end{equation} | (3.48) |
\begin{equation} 2H(x)P-\iota^{2} I_{n}\geq 0 , \end{equation} | (3.49) |
\begin{equation} \nabla_{x}^{\mathrm{T}} H(x)\nabla_{x} H(x)-b\phi(x) g_{1}^{\mathrm{T}}(x) g_{1}(x)\phi^{\mathrm{T}}(x)\geq 0 , \end{equation} | (3.50) |
where k > 0 , \varpi and \iota are given in Assumption 3.2, (3.44) and (3.46), respectively, N(x): = \operatorname{Diag}\left\{(\frac{2 \alpha_{1}-1}{2 \alpha_{1}})x_{1}^{\frac{2 \alpha_{1}-2}{2 \alpha_{1}-1}}, (\frac{2 \alpha_{2}-1}{2 \alpha_{2}})x_{2}^{\frac{2 \alpha_{2}-2}{2 \alpha_{2}-1}}, \cdots, (\frac{2 \alpha_{n}-1}{2 \alpha_{n}})x_{n}^{\frac{2 \alpha_{n}-2}{2 \alpha_{n}-1}}\right\} , and
(2) real numbers s > 0 and \mu > 0 exist, and
\begin{equation} \left[\begin{array}{cc} 2 s-\frac{\delta \gamma^{2}}{2 \mu^{2}}\; \; \; & -s \alpha_{j}^{\mathrm{T}} \\ -s \alpha_{j} & I_{n} \end{array}\right] \geqslant 0, j = 1, 2, \cdots, n, \end{equation} | (3.51) |
where \delta: = \max\{a^{\frac{2 \alpha_{min}-2}{2 \alpha_{min}-1}}, \; a^{\frac{2 \alpha_{max}-2}{2 \alpha_{max}-1}}\} with a: = \max \left\{\left\|x_{t}\right\|: x \in \Omega\right\} , then a predefined-time adaptive robust controller of the NTDS (3.47) could be constructed as
\begin{equation} g_{1}(x)v_{1} = -\frac{1}{\sigma T_{a}}\Phi(x), \end{equation} | (3.52) |
\begin{equation} \begin{aligned} g_{1}(x)v_{2} = &-\frac{1}{2}g_{1}(x)[{h}^\mathrm{T}(x)h(x)+\frac{1}{\gamma^{2}}I_{m}]{g_{1}}^\mathrm{T}(x)\nabla_{x}H(x)-Q\nabla_{x}H(x), \end{aligned} \end{equation} | (3.53) |
where \sigma : = \min\{1, \; W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}}\} with H(x)\leq W when x \in \Omega , \alpha_{min } and \alpha_{max } are similar to these in Theorem 3.1, and \Phi(x) is given in (3.30).
Proof. V(t, x_{t}) is set as follows:
\begin{equation} \begin{aligned} V(t, x_{t}) = 2 e^{\int_{t-h}^{t} \nabla_{x}^{T} H(x(s)) P \nabla_{x} H(x(s)) d s} H(x): = 2 G(t) H(x). \end{aligned} \end{equation} | (3.54) |
When we compute the derivative of V(t, x_{t}) and use {\nabla_{x}}^\mathrm{T}H(x)J(x)\nabla_{x}H(x) = 0 , one can obtain
\begin{equation} \begin{aligned} \dot{V}\left(t, x_{t}\right)\leqslant &G(t)[2\nabla_{x}^\mathrm{T}H(x)H(x)P\nabla_{x}H(x)-2\nabla_{\tilde{x}}^\mathrm{T}H(\tilde{x})H(x)P\nabla_{\tilde{x}}H(\tilde{x})\\ & -2 \nabla_{x}^{\mathrm{T}} H(x) R \nabla_{x} H(x)+\nabla_{x}^{\mathrm{T}} H(x) \nabla_{x} H(x)+\iota^{2}\nabla_{\tilde{x}}^{\mathrm{T}} H(\tilde{x}) \nabla_{\tilde{x}} H(\tilde{x})-2\nabla_{x}^{\mathrm{T}} H(x)\frac{1}{\sigma T_{a}}\Phi(x)\\ &-\parallel z \parallel^{2}-\dfrac{1}{\gamma^{2}} \nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) g_{1}^{\mathrm{T}}(x) \nabla_{x} H(x)+r^{-1} \nabla_{x}^{\mathrm{T}} H(x) g_{2}(x) g_{2}^{\mathrm{T}}(x) \nabla_{x} H(x)+r \omega^{\mathrm{T}}\omega \\ &+\nabla_{x}^{\mathrm{T}} H(x) \nabla_{x} H(x)+G^{\mathrm{T}}(x, p) G(x, p) +2 \nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) \phi^{\mathrm{T}}(x)\theta-2\nabla_{x}^{\mathrm{T}} H(x)Q\nabla_{x} H(x)]. \end{aligned} \end{equation} | (3.55) |
There exists a matrix N(x): = \operatorname{Diag}\left\{(\frac{2 \alpha_{1}-1}{2 \alpha_{1}})x_{1}^{\frac{2 \alpha_{1}-2}{2 \alpha_{1}-1}}, (\frac{2 \alpha_{2}-1}{2 \alpha_{2}})x_{2}^{\frac{2 \alpha_{2}-2}{2 \alpha_{2}-1}}, \cdots, (\frac{2 \alpha_{n}-1}{2 \alpha_{n}})x_{n}^{\frac{2 \alpha_{n}-2}{2 \alpha_{n}-1}}\right\} such that x: = N(x)\nabla_{x}H(x) holds, and using (3.43), one can get
\begin{equation} \begin{aligned} G^{\mathrm{T}}(x, p) G(x, p) \leq \varpi x^{\mathrm{T}} x : = \varpi{\nabla_{x}}^\mathrm{T}H(x){N}^\mathrm{T}(x)N(x)\nabla_{x}H(x). \end{aligned} \end{equation} | (3.56) |
Substituting (3.56) and (3.17) into (3.55), we have
\begin{equation} \begin{aligned} \dot{V}(t, x_{t}) \leqslant &G(t)[ \nabla_{x}^{\mathrm{T}} H(x) L \nabla_{x} H(x)-\parallel z \parallel^{2}-\nabla_{\tilde{x}}^{\mathrm{T}} H(\tilde{x}) [2H(x)P-\iota^{2} I_{n}]\nabla_{\tilde{x}} H(\tilde{x})\\ &-2\nabla_{x}^{\mathrm{T}} H(x)\frac{1}{\sigma T_{a}}\Phi(x) +r \omega^{\mathrm{T}} \omega +2 \nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) \phi^{\mathrm{T}}(x)\theta-2\nabla_{x}^{\mathrm{T}} H(x)Q\nabla_{x} H(x)]\\ \leq& G(t)\nabla_{x}^{\mathrm{T}} H(x) L \nabla_{x} H(x) -\frac{2^{\frac{1-2\alpha_{min}}{\alpha_{min}}}}{\sigma T_{a}}\dfrac{\alpha_{min}(2\alpha_{max}-1)^{2}}{\alpha_{max}^{2}(\alpha_{min}-1)}G(t)^{\dfrac{1}{\alpha_{min }}}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}]\\&\times\nabla_{x}^\mathrm{T}H(x)\nabla_{x}H(x) -G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2}+2 G(t)\nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) \phi^{\mathrm{T}}(x)\theta\\&-2G(t)\nabla_{x}^{\mathrm{T}} H(x)Q\nabla_{x} H(x)\\ \leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}} +G(t)\nabla_{x}^{\mathrm{T}} H(x) L \nabla_{x} H(x)\\&-G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2} +2 G(t)\nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) \phi^{\mathrm{T}}(x)\theta-2G(t)\nabla_{x}^{\mathrm{T}} H(x)Q\nabla_{x} H(x). \end{aligned} \end{equation} | (3.57) |
Noting that \parallel \theta\parallel^{2}\leq k^{2} and Q\geq\frac{1}{2}k^{2}I_{n} , one can obtain that -2Q\leq-k^{2}I_{n}\leq-\theta^{\mathrm{T}}\theta I_{n} , which is
\begin{equation} -2\nabla_{x}^{\mathrm{T}} H(x)Q\nabla_{x} H(x)\leq-\theta^{\mathrm{T}}\nabla_{x}^{\mathrm{T}} H(x)\nabla_{x} H(x)\theta. \end{equation} | (3.58) |
Substituting (3.58) into (3.57), and noting (3.48) and (3.50), one can get
\begin{equation} \begin{aligned} \dot{V}(t, x_{t})\leq&-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ &+G(t)[\nabla_{x}^{\mathrm{T}} H(x) L \nabla_{x} H(x)+2\nabla_{x}^{\mathrm{T}} H(x) g_{1}(x) \phi^{\mathrm{T}}(x)\theta-\theta^{\mathrm{T}}\nabla_{x}^{\mathrm{T}} H(x)\nabla_{x} H(x)\theta]\\ &-G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2}\\ \leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ &+G(t)[\nabla_{x}^{\mathrm{T}} H(x) [L+b^{-1}I_{n}] \nabla_{x} H(x)\\&-\theta^{\mathrm{T}}[\nabla_{x}^{\mathrm{T}} H(x)\nabla_{x} H(x)-b\phi(x) g_{1}^{\mathrm{T}}(x) g_{1}(x)\phi^{\mathrm{T}}(x)]\theta]\\ &-G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2}\\ \leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ &-G(t)\parallel z \parallel^{2}+G(t) r\parallel \omega \parallel^{2}. \end{aligned} \end{equation} | (3.59) |
First, we prove (3.5) holds by letting D\left(t, x\right) = V(t, x_{t})+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds . Then we indicate that D\left(t, x\right)\leq0 .
Substituting (3.59) into \dot{D}\left(t, x\right) , we get
\begin{equation} \begin{aligned} \dot{D}\left(t, x\right) = &\dot{V}(t, x_{t})+\parallel z \parallel^{2}-\gamma^{2}\parallel \omega \parallel^{2}\\ \leq&-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ &+(G(t)r-\gamma^{2})\parallel \omega \parallel^{2}+(-G(t)+1)\parallel z \parallel^{2}.\\ \end{aligned} \end{equation} | (3.60) |
Assuming that \left\|x_{t}\right\| \leq a , it follows that G(t) \leq e^{\left(h \lambda_{\max }\{P\}\right) a^{2}} , from which we obtain \dot{D}\left(t, x\right)\leq -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}+(e^{\left(h \lambda_{\max }\{P\}\right) a^{2}}r-\gamma^{2})\parallel \omega \parallel^{2}+(-G(t)+1)\parallel z \parallel^{2}.
Noting that e^{\left(h \lambda_{\max }\{P\}\right) a^{2}}r \leq \gamma^{2} and G(t)\geq1 , one gets
\begin{equation} \begin{aligned} \dot{D}\left(t, x\right) \leq&-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}} < 0. \end{aligned} \end{equation} | (3.61) |
The condition of zero-state response and the integration of \dot{D}\left(t, x\right) over 0 and t result in
\begin{equation} V(t, x_{t})+\int_{0}^{t}(\parallel z(s) \parallel^{2}-\gamma^{2}\parallel \omega(s) \parallel^{2})ds\leq 0, \end{equation} | (3.62) |
and with V(t, x_{t})\geq0 , we get
\begin{equation} \int_{0}^{t}\parallel z \parallel^{2}ds\leq\gamma^{2}\int_{0}^{t}\parallel \omega(s) \parallel^{2})ds. \end{equation} | (3.63) |
Furthermore, using \frac{\alpha_{i}}{2 \alpha_{i}-1} < 1 and x^{T} x = \sum_{i = 1}^{n}\left(x_{i}\right)^{2} , it is easy to obtain that x^{T} x = a^{2} \sum_{i = 1}^{n}\left(\frac{x_{i}}{a}\right)^{2} \leq a^{2} \sum_{i = 1}^{n}\left[\left(\frac{x_{i}}{a}\right)^{2}\right]^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = \frac{a^{2}}{a^{\frac{2 \alpha_{i}}{2 \alpha_{i}-1}}} \sum_{i = 1}^{n}\left(x_{i}^{2}\right)^{\frac{\alpha_{i}}{2 \alpha_{i}-1}} = a^{\frac{2 \alpha_{i}-2}{2 \alpha_{i}-1}} H(x) holds on \Omega .
Similar to Theorem 3.1, we can get x^{T}x\leq\delta H(x) .
Noting Assumption 3.1 and Eq (3.62), we have x^{T}x\leq \delta H(x)\leq G(t)\delta H(x) = \frac{1}{2}\delta V\left(t, x_{t}\right) \leq \delta \frac{\gamma^{2}}{2}\\\times \int_{0}^{T}\|\omega(s)\|^{2} d s \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} , namely, \|x\|^{2} \leq \frac{\delta \gamma^{2}}{2 \mu^{2}} .
Afterward, it is shown that x(t) \in \Omega holds when \forall t > 0 , \varphi = 0 , \omega \in \Lambda . With (3.4), it must be demonstrated that
\begin{equation} x^{T} x-\frac{\delta \gamma^{2}}{2 \mu^{2}} \leq 0, \text { s.t. } 2-2 \alpha_{j}^{T} x \geq 0 (j = 1, \cdots, n). \end{equation} | (3.64) |
From [42], we have
\begin{equation} \varsigma^{T}\left[\begin{array}{cc} 2 s-\frac{\delta \gamma^{2}}{2 \mu^{2}}\; \; \; & -s \alpha_{j}^{T} \\ -s \alpha_{j} & I_{n} \end{array}\right] \varsigma \geq 0(j = 1, 2, \cdots, n), \end{equation} | (3.65) |
which indicates that (3.64) holds with free scalar s > 0 introduced by the S-procedure and \zeta = \left[1, x^{T}\right]^{T} . Thus, we conclude that x(t) remains in \Omega for all t > 0 , \varphi = 0 , \omega\in \Lambda .
Next, we demonstrate the stability in the predefined time for the NTDS (3.47) when \omega disappears.
Noting (3.59), one obtains
\begin{equation} \begin{aligned} \dot{V}(t, x_{t})\leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}-G(t)\parallel z \parallel^{2}\\ \leq& -\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[(2G(t)H(x))^{1-\frac{1}{\alpha_{min}}}](2G(t)H(x))^{\frac{1}{\alpha_{min}}}\\ = &-\frac{\alpha_{min}}{T_{a}(\alpha_{min}-1)}\exp[V^{1-\frac{1}{\alpha_{min}}}]V^{\frac{1}{\alpha_{min}}}, \end{aligned} \end{equation} | (3.66) |
which implies that Lemma 2.1 holds when \omega = 0 . Thus, it has been proved.
Take into account the following pollution control system with the single delay for rivers [46,47]:
\begin{equation} \dot{x} = \left(A_{0}\right. \left.+\Delta A_{0}(t)\right) x+\left(A_{1}+\Delta A_{1}(t)\right) \tilde{x}+B u(t)+\bar{\omega}(t), \end{equation} | (4.1) |
where A_{1} = \beta_{0} \eta_{2} I_{2} , \Delta A_{1}(t) = \Delta \beta \eta_{2} I_{2} , B = \operatorname{Diag}\left\{\eta_{1}, 1\right\}, \begin{array}{c} A_{0} = \left[\begin{array}{cc} -k_{10}-\eta_{1}-\eta_{2} & 0 \\ -k_{30} & -k_{20}-\eta_{1}-\eta_{2} \end{array}\right], \Delta A_{0}(t) = \left[\begin{array}{cc} -\Delta k_{1}(t) & 0 \\ -\Delta k_{3}(t) & -\Delta k_{2}(t) \end{array}\right]. \end{array}
From Ref. [47], we are aware that \triangle A_{0}(t) = B \Psi(t) , \bar{\omega}(t) = B \omega(t) and constants \rho_{i}(i = 1, 2) > 0 , where \|\Psi(t)\| \leq \sqrt{\rho_{1}}, \; \|\omega(t)\| \leq \sqrt{\rho_{2}} . Ref. [46] explains the physical meaning of the above parameters (for more details, please see Ref. [46]). Furthermore, assume that x = \left[x_{1}, x_{2}\right]^{T} \in \Omega = \left\{\left(x_{1}, x_{2}\right):\left|x_{1}\right| \leq 2.5, \left|x_{2}\right| \leq 2.5\right\} and \triangle A_{0}(t) = \left[\begin{array}{ll}p \; & 0 \\ p \; & p\end{array}\right] .
First, the Hamiltonian form of system (4.1) is expressed as follows:
\begin{equation} \begin{aligned} \dot{x} = (J-R) \nabla_{x} H_{a}(x)+T \nabla_{\tilde{x}} H_{a}(\tilde{x})+g_{1} u(t) +g_{2}\omega(t)+\Delta A_{0}(t) x, \end{aligned} \end{equation} | (4.2) |
where H_{a}(x) = 0.5\left(x_{1}^{2}+x_{2}^{2}\right) , T = A_{1}+\triangle A_{1}(t) , g_{1} = g_{2} = B and J-R = : A_{0}. Besides, since \|G(x, p)\| = 0 , we can easily see that (3.43) holds on \Omega with \varpi = 0 .
In the following, we can obtain
\begin{equation} \begin{aligned} \triangle A_{0}(t) x = \left[x_{1}, x_{1}+x_{2}\right]^{T} p = B\left[\frac{1}{\eta_{1}} x_{1}, x_{1}+x_{2}\right]^{T} p = g_{1}(x) \phi^{T}(x) \theta, \end{aligned} \end{equation} | (4.3) |
where \theta = p and \phi(x) = \left[\frac{1}{\eta_{1}} x_{1}, x_{1}+x_{2}\right] . Substituting (4.3) into (4.2), one gets
\begin{equation} \begin{aligned} \dot{x} = {[J-R] \nabla_{x} H_{a}(x)+T \nabla_{\tilde{x}} H_{a}(\tilde{x})+g_{1}(x) u } +g_{1}(x) \omega+g_{1}(x) \phi^{T}(x) \theta. \end{aligned} \end{equation} | (4.4) |
Next, design the controller in the following manner:
\begin{equation} \begin{aligned} g_{1}(x) u = [J-R] \nabla_{x} H_{b}(x)+T \nabla_{\tilde{x}} H_{b}(\tilde{x}) +\left[\iota I_{n}-T\right] \nabla_{\tilde{x}} H(\tilde{x})+g_{1}(x) v, \end{aligned} \end{equation} | (4.5) |
where v represents a new input, \iota: = (dH(x))^{\frac{1}{2}} = (0.0001H(x))^{\frac{1}{2}} , and H_{b}(x): = H(x)- H_{a}(x) with H(x) = x_{1}^{\frac{4}{3}}+x_{2}^{\frac{6}{5}}(\alpha_{1} = 2, \alpha_{2} = 3) . Noting that H(x)\leq W when x \in \Omega and selecting W = 8 , one can obtain \sigma : = \min\{1, \; W^{\frac{1}{\alpha_{max}}-\frac{1}{\alpha_{min }}}\} = \dfrac{\sqrt{2}}{2} . As a result, the system (4.4) can be expressed as follows:
\begin{equation} \begin{aligned} \dot{x} = [J-R] \nabla_{x} H(x)+\iota I_{n} \nabla_{\tilde{x}} H(\tilde{x})+g_{1}(x) v +g_{1}(x) \omega+g_{1}(x) \phi^{T}(x) \theta. \end{aligned} \end{equation} | (4.6) |
In addition, set \gamma = 0.19 as disturbance suppression level, and choose penalty signal z = h(x) g_{1}^{T} \nabla_{x} H(x) = \frac{4 \eta_{1}}{3} x_{1}^{\frac{1}{3}}+ \frac{6}{5} x_{2}^{\frac{1}{5}} , where h(x) = [1,1] denotes the weight matrix of the appropriate dimension.
Now, we demonstrate that all conditions of Theorem 3.3 hold for the system by choosing k_{10} = 19 , k_{20} = 5 , k_{30} = 10 , \beta_{0} = 0.6 , \triangle\beta = 0.02 , \eta_{1} = 0.3 , \eta_{2} = 0.7 , and then we can obtain J(x) = \left[\begin{array}{cc} 0 \; & 5 \\ -5 \; & 0 \end{array}\right], R(x) = \left[\begin{array}{cc} 20 \; & 5 \\ 5 \; & 6 \end{array}\right] , T = 0.434 I_{2} .
Meanwhile, set the parameters as follows: T_{a} = 0.1s and 0.05s , respectively, k = 4 , b = 0.1661 , r = 0.035 , P = 0.15 I_{2} , Q = 10 I_{2} . From x: = N(x)\nabla_{x}H(x) , we have N = \operatorname{Diag}\left\{\dfrac{3}{4}x_{1}^{\frac{2}{3}}, \dfrac{5}{6}x_{2}^{\frac{4}{5}}\right\} . Then, condition (1) holds for 0 < h < 0.5 .
Selecting \alpha_{j} = 0.1(j = 1, 2) , s = 10 and \mu = 0.06 , the equations \alpha_{j}\left|x_{j}\right| \leq 1 and (3.51) are valid. It follows that condition (2) of Theorem 3.3 holds.
For this system, Theorem 3.3 is fulfilled when 0 < h < 0.5 . Based on Theorem 3.3, we design a adaptive robust controller in the predefined time for the system (4.2) as
\begin{equation} \left[\begin{array}{c} u_1 = -139.0735 x_{1}^{\frac{1}{3}}-0.6 x_{2}^{\frac{1}{5}}+66.6667 x_{1}\\-1.4467\tilde{x}_{1}+0.0444(\sqrt{x_{1}^{\frac{4}{3}}+x_{2}^{\frac{6}{5}}})\tilde{x}_{1}^{\frac{1}{3}} \\ -\dfrac{6.1728 x_{1}^{\frac{1}{3}}}{ T_{a}\sqrt{\exp(0.1333x_{1}^{\frac{2}{3}}+0.108x_{2}^{\frac{2}{5}})}} \\ \times\exp(\sqrt{(2x_{1}^{\frac{4}{3}}+2x_{2}^{\frac{6}{5}})\exp(0.1333x_{1}^{\frac{2}{3}}+0.108x_{2}^{\frac{2}{5}})}), \\ u_2 = -13.5333x_{1}^{\frac{1}{3}}-36.4205 x_{2}^{\frac{1}{5}}+10x_{1}+6x_{2}\\-0.434\tilde{x}_{2}+0.012(\sqrt{x_{1}^{\frac{4}{3}}+x_{2}^{\frac{6}{5}}})\tilde{x}_{2}^{\frac{1}{5}} \\ -\dfrac{1.6667 x_{2}^{\frac{1}{5}}}{T_{a} \sqrt{\exp(0.1333x_{1}^{\frac{2}{3}}+0.108x_{2}^{\frac{2}{5}})}}\\ \times\exp(\sqrt{(2x_{1}^{\frac{4}{3}}+2x_{2}^{\frac{6}{5}})\exp(0.1333x_{1}^{\frac{2}{3}}+0.108x_{2}^{\frac{2}{5}})})\\ \end{array}\right]. \end{equation} | (4.7) |
Choose h = 0.5 and p = 0.5 . To assess how robust the controller (4.7) is against disturbances from outside, a disturbance with magnitude [120, \; 120]^{T} is introduced into our system in the time interval [0.01s \sim 0.03s] . Figures 1–4 are the simulation results.
Applying the predefined-time adaptive robust controller (4.7) to the system (4.2) with \varphi = (1.9, 0.5) , T_{a} = 0.05s and T_{a} = 0.1s , the simulation outcomes can be seen in Figure 1 as well as 2, and the behaviors of the corresponding control inputs are shown in Figures 5 and 6, indicating the system converges to zero within 0.05s and 0.1s . In addition, when the infinite-time controller designed in [34] is applied to the system (4.6) with the same initial conditions and disturbance, the result is shown in Figure 3, and the behavior of the corresponding control inputs is shown in Figure 7, and it is obvious that the system states converge to zero in 0.35s .
In addition, we also give a comparison with the finite-time controller from [14] with the same initial conditions and disturbance, Figure 4 illustrates the simulation result, and Figure 8 shows the behavior of the corresponding control inputs. According to Figure 4, the states stabilize within 0.15s .
Comparing Figures 1 and 2, Figures 3 and 4, the predefined-time, infinite-time and finite-time control schemes were applied to the same system (4.6) under the same initial conditions, the same applied perturbations and the same parameters, respectively, and the comparison results are shown in Table 1. Through Table 1, we can find that the system converges faster under predefined-time control. Moreover, under the same disturbances, the CLS with the above-mentioned predefined-time controller has a smaller amplitude and returns to equilibrium rapidly after the perturbation ends. Consequently, the controller presented in this paper is shown that is very effective for robust stabilization calming.
Total system convergence time | Convergence time of the system after the disappearance of the disturbance | System amplitude after adding the same disturbance | |
Predefined-time control with T=0.05s | 0.039s | 0.009s | 0.6 |
Predifined-time control with T=0.1s | 0.045s | 0.015s | 0.8 |
Finite-time control | 0.1s | 0.07s | 1.55 |
Infinite-time control | 0.31s | 0.28s | 1.58 |
Throughout the work, we have studied the predefined-time control problem for NTDSs based on the Hamiltonian function approach. By choosing a suitable Hamiltonian form of different powers and constructing appropriate Lyapunov functions, we have presented two corresponding predefined-time control results, namely, H_\infty control and adaptive robust control ones, which have guaranteed the systems to be quickly calibrated in a predefined time and have shortened the calibration time and improved the control accuracy. The simulation results have indicated that the predefined-time adaptive robust scheme presented here has quicker convergence and greater robustness over existing infinite-time and finite-time control schemes. Moreover, unlike the results of existing Hamiltonian systems (where the powers are the same), this paper uses a more realistic form of different power Hamiltonian functions, implying its wider application.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This article received support from the National Natural Science Foundation of China (G61773015), the Natural Science Foundation of Shandong Province (ZR2023MF019), the Research Leader Studio of Jinan (202228119).
All the authors declare no conflict of interest.
[1] |
S. Nadeem, W. Fuzhang, F. M. Alharbi, F. Sajid, N. Abbas, A. S. El-Shafay, et al., Numerical computations for Buongiorno nano fluid model on the boundary layer flow of viscoelastic fluid towards a nonlinear stretching sheet, Alex. Eng. J., 61 (2022), 1769–1778. https://doi.org/10.1016/j.aej.2021.11.013 doi: 10.1016/j.aej.2021.11.013
![]() |
[2] |
M. Qayyum, S. Afzal, E. Ahmad, M. B. Riaz, Fractional modeling and analysis of unsteady squeezing flow of Casson nanofluid via extended He-Laplace algorithm in Liouville-Caputo sense, Alex. Eng. J., 73 (2023), 579–591. https://doi.org/10.1016/j.aej.2023.05.010 doi: 10.1016/j.aej.2023.05.010
![]() |
[3] |
A. E. Aboanber, A. A. Nahla, A. M. El-Mhlawy, O. Maher, An efficient exponential representation for solving the two-energy group point telegraph kinetics model, Ann. Nucl. Energy, 166 (2022), 108698. https://doi.org/10.1016/j.anucene.2021.108698 doi: 10.1016/j.anucene.2021.108698
![]() |
[4] |
C. Villa, A. Gerisch, M. A. J. Chaplain, A novel nonlocal partial differential equation model of endothelial progenitor cell cluster formation during the early stages of vasculogenesis, J. Theor. Biol., 534 (2022), 110963. https://doi.org/10.1016/j.jtbi.2021.110963 doi: 10.1016/j.jtbi.2021.110963
![]() |
[5] |
I. Ahmad, H. Ahmad, P. Thounthong, Y. Chu, C. Cesarano, Solution of multi-term time-fractional PDE models arising in mathematical biology and physics by local meshless method, Symmetry, 12 (2020), 1195. https://doi.org/10.1016/j.heliyon.2023.e16522 doi: 10.1016/j.heliyon.2023.e16522
![]() |
[6] |
O. D. Adeyemo, C. M. Khalique, Lie group classification of generalized variable coefficient Korteweg-de Vries equation with dual power-law nonlinearities with linear damping and dispersion in quantum field theory, Symmetry, 14 (2022), 83. https://doi.org/10.3390/sym14010083 doi: 10.3390/sym14010083
![]() |
[7] |
S. Afzal, M. Qayyum, M. B. Riaz, A. Wojciechowski, Modeling and simulation of blood flow under the influence of radioactive materials having slip with MHD and nonlinear mixed convection, Alex. Eng. J., 69 (2023), 9–24. https://doi.org/10.1016/j.aej.2023.01.013 doi: 10.1016/j.aej.2023.01.013
![]() |
[8] |
L. Guo, H. Wu, T. Zhou, Normalizing field flows: Solving forward and inverse stochastic differential equations using physics-informed flow models, J. Comput. Phys., 461 (2022), 111202. https://doi.org/10.1016/j.jcp.2022.111202 doi: 10.1016/j.jcp.2022.111202
![]() |
[9] |
S. P. Joseph, New traveling wave exact solutions to the coupled Klein-Gordon system of equations, PDE Appl. Math., 5 (2022), 100208. https://doi.org/10.1016/j.padiff.2021.100208 doi: 10.1016/j.padiff.2021.100208
![]() |
[10] |
M. Farman, A. Akgül, S. Askar, T. Botmart, A. Ahmad, H. Ahmad, Modeling and analysis of fractional order Zika model, AIMS Math., 7 (2022), 3912–3938. https://doi.org/10.3934/math.2022216 doi: 10.3934/math.2022216
![]() |
[11] | J. Liouville, Mémoire sur quelques questions de géométrie et de mécanique, et sur un nouveau genre de calcul pour résoudre ces questions, J. Éc. Polytech. Math., 1832. |
[12] | K. S. Miller, B. Ross, An Introduction to the Fractional Calculus and Fractional Differential Equations, Hoboken: Wiley, 1993. |
[13] | B. Riemann, Versuch einer allgemeinen auffassung der integration und differentiation, Gesammelte Werke, 62 (1876), 1876. |
[14] | M. Caputo, Elasticita e Dissipazione, Bologna: Zanichelli, 1969. |
[15] | H. Ahmad, N. Alam, M. Omri, New computational results for a prototype of an excitable system, Results Phys., 28 (2021). https://doi.org/10.1016/j.rinp.2021.104666 |
[16] |
M. Qayyum, E. Ahmad, S. Afzal, T. Sajid, W. Jamshed, A. Musa, et al., Fractional analysis of unsteady squeezing flow of casson fluid via homotopy perturbation method, Sci. Rep., 12 (2022), 18406. https://doi.org/10.1038/s41598-022-23239-0 doi: 10.1038/s41598-022-23239-0
![]() |
[17] |
K. S. Nisar, K. Logeswari, V. Vijayaraj, H. M. Baskonus, C. Ravichandran, Fractional order modeling the gemini virus in capsicum annuum with optimal control, Fractal Fract., 6 (2022), 61. https://doi.org/10.3390/fractalfract6020061 doi: 10.3390/fractalfract6020061
![]() |
[18] |
A. Yusuf, S. Qureshi, U. T. Mustapha, S. S. Musa, T. A. Sulaiman, Fractional modeling for improving scholastic performance of students with optimal control, Int. J. Appl. Comput. Math., 8 (2022), 37. https://doi.org/10.1007/s40819-021-01177-1 doi: 10.1007/s40819-021-01177-1
![]() |
[19] |
H. Hassani, J. A. Tenreiro Machado, Z. Avazzadeh, E. Naraghirad, S. Mehrabi, Optimal solution of the fractional-order smoking model and its public health implications, Nonlinear Dynam., 108 (2022), 2815–2831. https://doi.org/10.1007/s11071-022-07343-4 doi: 10.1007/s11071-022-07343-4
![]() |
[20] |
M. Qayyum, E. Ahmad, S. T. Saeed, H. Ahmad, S. Askar, Homotopy perturbation method-based soliton solutions of the time-fractional (2+1)-dimensional wu-zhang system describing long dispersive gravity water waves in the ocean, Front. Phys., 11 (2023), 1178154. https://doi.org/10.3389/fphy.2023.1178154 doi: 10.3389/fphy.2023.1178154
![]() |
[21] |
C. Wang, X. Zhou, X. Shi, Y. Jin, Variable fractional order sliding mode control for seismic vibration suppression of uncertain building structure, J. Vib. Eng. Tech., 10 (2021), 299–312. https://doi.org/10.1007/s42417-021-00377-9 doi: 10.1007/s42417-021-00377-9
![]() |
[22] |
I. M. Batiha, S. A. Njadat, R. M. Batyha, A. Zraiqat, A. Dababneh, S. Momani, Design fractional-order PID controllers for single-joint robot arm model, Int. J. Adv. Soft Comput. Appl., 14 (2022), 97–114. https://doi.org/10.15849/ijasca.220720.07 doi: 10.15849/ijasca.220720.07
![]() |
[23] |
M. H. Derakhshan, Existence, uniqueness, Ulam-Hyers stability and numerical simulation of solutions for variable order fractional differential equations in fluid mechanics, J. Appl. Math. Comput., 68 (2021), 403–429. https://doi.org/10.1007/s12190-021-01537-6 doi: 10.1007/s12190-021-01537-6
![]() |
[24] |
A. Cardone, D. Conte, R. D'Ambrosio, B. Paternoster, Multivalue collocation methods for ordinary and fractional differential equations, Mathematics, 10 (2022), 185. https://doi.org/10.3390/math10020185 doi: 10.3390/math10020185
![]() |
[25] |
N. A. Shah, A. Wakif, E. R. El-Zahar, T. Thumma, S. J. Yook, Heat transfers thermodynamic activity of a second-grade ternary nanofluid flow over a vertical plate with Atangana-Baleanu time-fractional integral, Alex. Eng. J., 61 (2022), 10045–10053. https://doi.org/10.1016/j.aej.2022.03.048 doi: 10.1016/j.aej.2022.03.048
![]() |
[26] |
N. P. Dong, H. V. Long, N. L. Giang, The fuzzy fractional SIQR model of computer virus propagation in wireless sensor network using Caputo Atangana–Baleanu derivatives, Fuzzy Sets Syst., 429 (2022), 28–59. https://doi.org/10.1016/j.fss.2021.04.012 doi: 10.1016/j.fss.2021.04.012
![]() |
[27] |
A. Din, F. M. Khan, Z. U. Khan, A. Yusuf, T. Munir, The mathematical study of climate change model under nonlocal fractional derivative, PDE Appl. Math., 5 (2022), 100204. https://doi.org/10.1016/j.padiff.2021.100204 doi: 10.1016/j.padiff.2021.100204
![]() |
[28] |
Y. Gurefe, Y. Pandir, T. Akturk, Analysis of exact solutions of a mathematical model by new function method, Cumhuriyet Sci. J., 43 (2022), 703–707. https://doi.org/10.17776/csj.1083033 doi: 10.17776/csj.1083033
![]() |
[29] |
M. R. Ahamed Fahim, P. R. Kundu, M. E. Islam, M. A. Akbar, M. S. Osman, Wave profile analysis of a couple of (3+1)-dimensional nonlinear evolution equations by sine-Gordon expansion approach, J. Ocean Eng. Sci., 7 (2022), 272–279. https://doi.org/10.1016/j.joes.2021.08.009 doi: 10.1016/j.joes.2021.08.009
![]() |
[30] |
B. Ghanbari, Employing Hirota's bilinear form to find novel lump waves solutions to an important nonlinear model in fluid mechanics, Results Phys., 29 (2021), 104689. https://doi.org/10.1016/j.rinp.2021.104689 doi: 10.1016/j.rinp.2021.104689
![]() |
[31] | W. Razzaq, A. Zafar, H. M. Ahmed, W. B. Rabie, Construction solitons for fractional nonlinear Schrodinger equation with \beta-time derivative by the new sub-equation method, J. Ocean Eng. Sci., in press, 2022. https://doi.org/10.1016/j.joes.2022.06.013 |
[32] |
G. Akram, M. Sadaf, S. Arshed, F. Sameen, Bright, dark, kink, singular and periodic soliton solutions of Lakshmanan-Porsezian-Daniel model by generalized projective riccati equations method, Optik, 241 (2021), 167051. https://doi.org/10.1016/j.ijleo.2021.167051 doi: 10.1016/j.ijleo.2021.167051
![]() |
[33] |
S. Liao, Homotopy analysis method: A new analytical technique for nonlinear problems, Commun. Nonlinear Sci. Numer. Simul., 2 (1997), 95–100. https://doi.org/10.1016/s1007-5704(97)90047-2 doi: 10.1016/s1007-5704(97)90047-2
![]() |
[34] |
S. Afzal, M. Qayyum, G. Chambashi, Heat and mass transfer with entropy optimization in hybrid nanofluid using heat source and velocity slip: a hamilton–crosser approach, Sci. Rep., 13 (2023), 12392. https://doi.org/10.1038/s41598-023-39176-5 doi: 10.1038/s41598-023-39176-5
![]() |
[35] |
T. Hayat, K. Muhammad, S. Momani, Melting heat and viscous dissipation in flow of hybrid nanomaterial: a numerical study via finite difference method, J. Therm. Anal. Calorime., 147 (2021), 6393–6401. https://doi.org/10.1007/s10973-021-10944-7 doi: 10.1007/s10973-021-10944-7
![]() |
[36] |
H. Ahmad, M. N. Khan, I. Ahmad, M. Omri, M. F. Alotaibi, A meshless method for numerical solutions of linear and nonlinear time-fractional Black-Scholes models, AIMS Math., 8 (2023), 19677–19698. https://doi.org/10.3934/math.20231003 doi: 10.3934/math.20231003
![]() |
[37] |
J. H. He, Homotopy perturbation technique, Comput. Meth. Appl. Mech. Eng., 178 (1999), 257–262. https://doi.org/10.1016/s0045-7825(99)00018-3 doi: 10.1016/s0045-7825(99)00018-3
![]() |
[38] |
R. Amin, K. Shah, H. Ahmad, A. H. Ganie, A. Abdel-Aty, T. Botmart, Haar wavelet method for solution of variable order linear fractional integro-differential equations, AIMS Math., 7 (2022), 5431–5443. https://doi.org/10.3934/math.2022301 doi: 10.3934/math.2022301
![]() |
[39] |
G. Singh, I. Singh, New laplace variational iterative technique to solve twodimensional Schr¨odinger equation, Mater. Today Proc., 62 (2022), 3995–4000. https://doi.org/10.1016/j.matpr.2022.04.585 doi: 10.1016/j.matpr.2022.04.585
![]() |
[40] |
M. Croci, G. R. de Souza, Mixed-precision explicit stabilized Runge-Kutta methods for single- and multi-scale differential equations, J. Comput. Phys., 464 (2022), 111349. https://doi.org/10.1016/j.jcp.2022.111349 doi: 10.1016/j.jcp.2022.111349
![]() |
[41] |
M. Aslam, M. Farman, H. Ahmad, T. N. Gia, A. Ahmad, S. Askar, Fractal fractional derivative on chemistry kinetics hires problem, AIMS Math., 7 (2022), 1155–1184. https://doi.org/10.3934/math.2022068 doi: 10.3934/math.2022068
![]() |
[42] |
J. H. He, M. L. Jiao, K. A. Gepreel, Y. Khan, Homotopy perturbation method for strongly nonlinear oscillators, Math. Comput. Simul., 204 (2023), 243–258. https://doi.org/10.1016/j.matcom.2022.08.005 doi: 10.1016/j.matcom.2022.08.005
![]() |
[43] |
M. Qayyum, E. Ahmad, M. B. Riaz, J. Awrejcewicz, Improved soliton solutions of generalized fifth order time-fractional KdV models: Laplace transform with homotopy perturbation algorithm, Universe, 8 (2022), 563. https://doi.org/10.3390/universe8110563 doi: 10.3390/universe8110563
![]() |
[44] | Y. Pandir, T. A\ddot{g}ir, Genisletilmis deneme denklemi yöntemi ile k\ddot{u}bik lineer olmayan Schrödinger denkleminin yeni tam ccöz\ddot{u}mleri, Afyon Kocatepe Uni. J. Sci. Eng., 20 (2020), 582–588. |
[45] |
K. J. Wang, G. D. Wang, Variational theory and new abundant solutions to the (1+2)-dimensional chiral nonlinear Schrödinger equation in optics, Phys. Letters A, 412 (2021), 127588. https://doi.org/10.1016/j.physleta.2021.127588 doi: 10.1016/j.physleta.2021.127588
![]() |
[46] |
M. Al-Smadi, O. A. Arqub, S. Momani, Numerical computations of coupled fractional resonant Schrödinger equations arising in quantum mechanics under conformable fractional derivative sense, Phys. Scripta, 95 (2020), 075218. https://doi.org/10.1088/1402-4896/ab96e0 doi: 10.1088/1402-4896/ab96e0
![]() |
[47] |
S. F. Tian, X. F. Wang, T. T. Zhang, W. H. Qiu, Stability analysis, solitary wave and explicit power series solutions of a (2 + 1)-dimensional nonlinear Schrödinger equation in a multicomponent plasma, Int. J. Numer. Meth. Heat Fluid Flow, 31 (2021), 1732–1748. https://doi.org/10.1108/hff-08-2020-0517 doi: 10.1108/hff-08-2020-0517
![]() |
[48] |
D. F. Li, J. L. Wang, J. W. Zhang, Unconditionally convergent l1-galerkin FEMs for nonlinear time-fractional Schrödinger equations, SIAM J. Sci. Comput., 39 (2017), A3067–A3088. https://doi.org/10.1137/16m1105700 doi: 10.1137/16m1105700
![]() |
[49] |
W. Q. Yuan, C. J. Zhang, D. F. Li, Linearized fast time-stepping schemes for time–space fractional Schrödinger equations, Phys. D Nonlinear Phenomena, 454 (2023), 133865. https://doi.org/10.1016/j.physd.2023.133865 doi: 10.1016/j.physd.2023.133865
![]() |
[50] |
K. Hosseini, E. Hincal, S. Salahshour, M. Mirzazadeh, K. Dehingia, B. J. Nath, On the dynamics of soliton waves in a generalized nonlinear Schrödinger equation, Optik, 272 (2023), 170215. https://doi.org/10.1016/j.ijleo.2022.170215 doi: 10.1016/j.ijleo.2022.170215
![]() |
[51] |
N. A. Kudryashov, Method for finding optical solitons of generalized nonlinear Schrödinger equations, Optik, 261 (2022), 169163. https://doi.org/10.1016/j.ijleo.2022.169163 doi: 10.1016/j.ijleo.2022.169163
![]() |
[52] |
W. Q. Yuan, D. F. Li, C. J. Zhang, Linearized transformed l1 galerkin FEMs with unconditional convergence for nonlinear time fractional Schrödinger equations, Numer. Math. Theory Meth. Appl., 16 (2023), 348–369. https://doi.org/10.4208/nmtma.oa-2022-0087 doi: 10.4208/nmtma.oa-2022-0087
![]() |
[53] |
T. Y. Han, Z. Li, X. Zhang, Bifurcation and new exact traveling wave solutions to time-space coupled fractional nonlinear Schrödinger equation, Phys. Letters A, 395 (2021), 127217. https://doi.org/10.1016/j.physleta.2021.127217 doi: 10.1016/j.physleta.2021.127217
![]() |
[54] |
P. F. Dai, Q. B. Wu, An efficient block Gauss-Seidel iteration method for the space fractional coupled nonlinear Schrödinger equations, Appl. Math. Letters, 117 (2021), 107116. https://doi.org/10.1016/j.aml.2021.107116 doi: 10.1016/j.aml.2021.107116
![]() |
[55] |
K. S. Nisar, S. Ahmad, A. Ullah, K. Shah, H. Alrabaiah, M. Arfan, Mathematical analysis of SIRD model of COVID-19 with Caputo fractional derivative based on real data, Results Phys., 21 (2021), 103772. https://doi.org/10.1016/j.rinp.2020.103772 doi: 10.1016/j.rinp.2020.103772
![]() |
[56] |
T. Bakkyaraj, Lie symmetry analysis of system of nonlinear fractional partial differential equations with Caputo fractional derivative, Eur. Phys. J. Plus, 135 (2020), 126. https://doi.org/10.1140/epjp/s13360-020-00170-9 doi: 10.1140/epjp/s13360-020-00170-9
![]() |
[57] |
N. H. Tuan, H. Mohammadi, S. Rezapour, A mathematical model for COVID-19 transmission by using the Caputo fractional derivative, Chaos Solitons Fract., 140 (2020), 110107. https://doi.org/10.1016/j.chaos.2020.110107 doi: 10.1016/j.chaos.2020.110107
![]() |
1. | Dongmei Xia, Kaiyuan Chen, Lin Sun, Guojin Qin, Research on reachable set boundary of neutral system with various types of disturbances, 2025, 20, 1932-6203, e0317398, 10.1371/journal.pone.0317398 | |
2. | Xiaotong Liu, Renming Yang, Predefined-time adaptive robust simultaneous stabilisation control for two nonlinear singular time-delay systems, 2025, 2330-7706, 1, 10.1080/23307706.2025.2454530 |
Total system convergence time | Convergence time of the system after the disappearance of the disturbance | System amplitude after adding the same disturbance | |
Predefined-time control with T=0.05s | 0.039s | 0.009s | 0.6 |
Predifined-time control with T=0.1s | 0.045s | 0.015s | 0.8 |
Finite-time control | 0.1s | 0.07s | 1.55 |
Infinite-time control | 0.31s | 0.28s | 1.58 |
Total system convergence time | Convergence time of the system after the disappearance of the disturbance | System amplitude after adding the same disturbance | |
Predefined-time control with T=0.05s | 0.039s | 0.009s | 0.6 |
Predifined-time control with T=0.1s | 0.045s | 0.015s | 0.8 |
Finite-time control | 0.1s | 0.07s | 1.55 |
Infinite-time control | 0.31s | 0.28s | 1.58 |