Next Article in Journal
Research on Air Confrontation Maneuver Decision-Making Method Based on Reinforcement Learning
Next Article in Special Issue
A Nonlinearly Modulated Logistic Map with Delay for Image Encryption
Previous Article in Journal
Decentralized Load Shedding Method Based on Voltage Stability Margin Index Using Synchrophasor Measurement Technology
Previous Article in Special Issue
Event-Triggered and Memory-Based Sliding Mode Variable Structure Control for Memristive Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Stability Analysis of Linear Systems under Time-Varying Samplings by a Non-Standard Discretization Method

School of Automation, Hangzhou Dianzi University, Hangzhou 310018, Zhejiang, China
*
Author to whom correspondence should be addressed.
Submission received: 19 September 2018 / Revised: 17 October 2018 / Accepted: 17 October 2018 / Published: 27 October 2018
(This article belongs to the Special Issue Recent Developments on Time-Delay Systems and Its Applications)

Abstract

:
This paper is concerned with the stability of linear systems under time-varying sampling. First, the closed-loop sampled-data system under study is represented by a discrete-time system using a non-standard discretization method. Second, by introducing a new sampled-date-based integral inequality, the sufficient condition on stability is formulated by using a simple Lyapunov function. The stability criterion has lower computational complexity, while having less conservatism compared with those obtained by a classical input delay approach. Third, when the system is subject to parameter uncertainties, a robust stability criterion is derived for uncertain systems under time-varying sampling. Finally, three examples are given to show the effectiveness of the proposed method.

1. Introduction

Recently, sampled-data systems have been widely applied in digital control systems and networked control systems [1,2,3,4,5]. More and more attention has been focused on the stability analysis and synthesis of sampled-data systems [6,7,8,9,10,11,12,13,14]. The sampled-data systems are often those including continuous-time state and discrete-time control, simultaneously [15]. Consider the following linear system:
x ˙ ( t ) = A x ( t ) + B u ( t ) , x ( 0 ) = x 0
where x ( t ) R n and u ( t ) R m are the state vector and input vector, respectively; x 0 is the initial condition of the system (1); A and B are known parameter matrices of appropriate dimensions. The following sampled-data control law is assumed by a zero-order holder for the system (1):
u ( t ) = K x ( t k ) , t [ t k , t k + 1 )
where K is the given controller gain of (2) and t k are the sampling instants satisfying 0 = t 0 < t 1 < < t k < . Let t k + 1 t k = h ( k ) , k = 1 , 2 , . x ( t k ) are the state vectors at the instants t k .
Substituting (2) into (1) yields:
x ˙ ( t ) = A x ( t ) + A 1 x ( t k ) , x ( 0 ) = x 0
for t [ t k , t k + 1 ) , k = 1 , 2 , , where A 1 = B K .
Recalling some existing results reported in the literature, there are mainly three approaches to dealing with stability analysis and synthesis of the sampled-data system (3) based on linear matrix inequality (LMI) techniques.
  • The first approach is called an input delay approach [7]. The input delay approach is very popular in the analysis of sampled-data systems. This approach has been applied by constructing time-independent Lyapunov–Krasovskii functionals or Razumikhin-type functions to derive stability criteria for linear sampled-data systems under constant/time-varying sampling. The idea of the input delay approach is to represent x ( t k ) as x ( t k ) = x ( t ( t t k ) ) . By introducing an artificial delay τ ( t ) = t t k for t [ t k , t k + 1 ) , k = 1 , 2 , , the system (3) is thus modeled as the following time-delay system:
    x ˙ ( t ) = A x ( t ) + A 1 x ( t τ ( t ) ) , x ( 0 ) = x 0
    Clearly, τ ( t ) is a piecewise function with its time-derivative being one, i.e., τ ˙ ( t ) = 1 for t t k . Notice that τ ( t ) = t t k t k + 1 t k = h ( k ) . Therefore, a stability criterion of (4) can be obtained in terms of LMIs using the Razumikhin method or the Lyapunov–Krasovskii functional method [16,17,18]. The input delay approach is developed by introducing time-dependent Lyapunov–Krasovskii functionals [5].
  • The second approach is the so-called impulsive model approach [5,19]. The impulsive model approach is to model a sampled-data system as an impulsive system. By choosing a piecewise time-dependent Lyapunov–Krasovskii functional or a discontinuous Lyapunov–Krasovskii functional, less stability criteria can be derived [20].
    It should be mentioned that, although some less conservative stability criteria can be derived using the above two approaches, the chosen Lyapunov–Krasovskii functionals are commonly complicated. Since the obtained LMIs require more scalar decision variables, the total numerical complexity of the stability criteria is definitely much higher.
  • The third approach is a discrete-time approach [3,21,22,23,24,25], by which a sampled-data system is equivalently transformed into a finite-dimensional discrete-time system, where inter-sampling information of the systems can be maintained. The discrete-time approach assumes the sampling period to be a constant, i.e., h ( k ) = h , where h is a positive constant. Under such an assumption, the system (3) is often represented as the following form by using a standard discretization technique [3,24,25,26]:
    x ( t k + 1 ) = G ( h ) x ( t k ) ,
    where G ( h ) = e A h + 0 h e A r A 1 d r . In this situation, one can draw a conclusion that the system (3) is asymptotically stable if and only if there exists a real matrix P > 0 such that:
    G T ( h ) P G ( h ) P < 0 .
    Thus, a maximum allowable constant sampling h m a x of h can been obtained such that (6) holds. However, if the above assumption is not satisfied, that is the sampling is not uniform, the standard discretization approach can hardly be utilized to represent the system (3) as (5) [25,26]. As a result, the discrete-time approach based on a standard discretization technique may not be applicable in this case.
This paper focuses on the stability of sampled-data systems under time-varying sampling. Different from the three aforementioned approaches, a non-standard discretization method is introduced to model the system (3) as a discrete-time system. By establishing a new sampled-data-based integral inequality, the sufficient conditions on the stability and robust stability of sampled-data systems under time-varying sampling are proposed, which are of much lower computational complexity. Finally, the effectiveness of the proposed method is demonstrated by three numerical examples.

2. Main Results

In this section, we introduce a non-standard discretization method to model the system (3) under time-varying sampling as a discrete-time system. Notice that the asymptotical stability of (3) is equivalent to that of the following discrete-time system:
x ( t k + 1 ) = x ( t k ) + t k t k + 1 [ A x ( s ) + A 1 x ( t k ) ] d s , x ( 0 ) = x 0
which can be rewritten as:
x ( t k + 1 ) = ( I + h ( k ) A 1 ) x ( t k ) + t k t k + 1 A x ( s ) d s , x ( 0 ) = x 0
Clearly, inter-sampling information of the system (3) is maintained by (8).
Remark 1.
It is clear that the discrete-time system (8) is different from the one in (5). In fact, the system (5) is obtained by a standard discretization technique, while the system (8) is not. Thus, we refer to the discretization technique used in (8) as a non-standard discretization method. It should be pointed out that the discrete-time system (8) allows the sampling to be non-uniform, while only uniform sampling applies in the system (5).
Replacing t k + 1 and t k with t in (7), respectively, one has:
x ( t ) = x ( t k ) + t k t X ( s ) d s = x ( t k + 1 ) t t k + 1 X ( s ) d s
for t [ t k , t k + 1 ) , k = 1 , 2 , , where X ( s ) = A x ( s ) + A 1 x ( t k ) .

2.1. A New Sampled-Data-Based Integral Inequality

In the following, we first introduce a new sampled-data-based integral inequality, which is useful in the derivative of our main results.
Lemma 1.
For a given matrix Q > 0 of appropriate dimension, we have:
1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s + 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s 3 4 h ( k ) t k t k + 1 X T ( u ) Q X ( u ) d u
Proof. 
1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s + 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s = 1 h ( k ) t k t k + 1 [ t k s X ( u ) d u ] T Q [ t k s X ( u ) d u ] d s + 1 h ( k ) t k t k + 1 [ s t k + 1 X ( u ) d u ] T Q [ s t k + 1 X ( u ) d u ] d s 1 h ( k ) t k t k + 1 ( s t k ) t k s X T ( u ) Q X ( u ) d u d s + 1 h ( k ) t k t k + 1 ( t k + 1 s ) s t k + 1 X T ( u ) Q X ( u ) d u d s
From the integral region of the above two integrals, which are given in Figure 1, we can obtain that:
1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s + 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s 1 h ( k ) t k t k + 1 ( s t k ) t k s X T ( u ) Q X ( u ) d u d s + 1 h ( k ) t k t k + 1 ( t k + 1 s ) s t k + 1 X T ( u ) Q X ( u ) d u d s = 1 h ( k ) t k t k + 1 u t k + 1 ( s t k ) X T ( u ) Q X ( u ) d s d u + 1 h ( k ) t k t k + 1 t k u ( t k + 1 s ) X T ( u ) Q X ( u ) d s d u = 1 2 h ( k ) t k t k + 1 [ h 2 ( k ) ( u t k ) 2 ] X T ( u ) Q X ( u ) d u + 1 2 h ( k ) t k t k + 1 [ h 2 ( k ) ( t k + 1 u ) 2 ] X T ( u ) Q X ( u ) d u = 1 2 h ( k ) t k t k + 1 [ 2 h 2 ( k ) ( u t k ) 2 ( t k + 1 u ) 2 ] X T ( u ) Q X ( u ) d u 3 4 h ( k ) t k t k + 1 X T ( u ) Q X ( u ) d u
In the process of the above enlargements, we use the following fact:
( u t k ) 2 ( t k + 1 u ) 2 = 2 ( u t k + t k + 1 2 ) 2 1 2 ( t k + 1 t k ) 2 1 2 h 2 ( k ) .
Thus, the proof is completed. □
Lemma 1 presents an integral inequality for the sum of two integral terms. Since these integral terms are related to sample-data information, Lemma 1 is called a sampled-data-based integral inequality, which plays an important role in the proof of the main results.

2.2. A Stability Criterion

Then, based on Lemma 1, we state the following result.
Proposition 1.
For given scalars h M h m > 0 , the system (3) is asymptotically stable if there exist some matrices P > 0 , Q > 0 , M 1 , M 2 of appropriate dimensions such that the following LMI holds for h ( k ) = h M and h ( k ) = h m , simultaneously.
Ξ = Ξ 11 Ξ 12 Ξ 13 h ( k ) A 1 T Q Ξ 12 T Ξ 22 Ξ 23 0 Ξ 13 T Ξ 23 T 2 Q h ( k ) A T Q h ( k ) Q A 1 0 h ( k ) Q A 4 3 Q < 0
where:
Ξ 11 = P Q M 1 T ( I + h ( k ) A 1 ) ( I + h ( k ) A 1 ) T M 1 Ξ 12 = M 1 T ( I + h ( k ) A 1 ) T M 2 Ξ 13 = Q h ( k ) M 1 T A Ξ 22 = P Q + M 2 T + M 2 Ξ 23 = Q h ( k ) M 2 T A
Proof. 
Choose the Lyapunov function candidate V ( t k ) for (8) as V ( t k ) = x T ( t k ) P x ( t k ) , where P > 0 . The forward difference of V ( t k ) can be calculated as:
Δ V ( t k ) = x T ( t k + 1 ) P x ( t k + 1 ) x T ( t k ) P x ( t k ) = x T ( t k + 1 ) P x ( t k + 1 ) x T ( t k ) P x ( t k ) + 1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s + 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s 1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s
where Q > 0 . By Lemma 1, we have:
Δ V ( t k ) x T ( t k + 1 ) P x ( t k + 1 ) x T ( t k ) P x ( t k ) + 3 4 h ( k ) t k t k + 1 X T ( u ) Q X ( u ) d u 1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s = x T ( t k + 1 ) P x ( t k + 1 ) x T ( t k ) P x ( t k ) + 3 4 h ( k ) t k t k + 1 X T ( s ) Q X ( s ) d s 1 h ( k ) t k t k + 1 [ x ( s ) x ( t k ) ] T Q [ x ( s ) x ( t k ) ] d s 1 h ( k ) t k t k + 1 [ x ( t k + 1 ) x ( s ) ] T Q [ x ( t k + 1 ) x ( s ) ] d s + 2 x T ( t k ) M 1 T + x T ( t k + 1 ) M 2 T x ( t k + 1 ) ( I + h ( k ) A 1 ) x ( t k ) t k t k + 1 A x ( s ) d s = 1 h ( k ) t k t k + 1 x T ( t k ) x T ( t k + 1 ) x T ( s ) Ξ ^ x ( t k ) x ( t k + 1 ) x ( s ) d s
where M 1 and M 2 are two free-weighting n × n matrices,
Ξ ^ = Ξ ^ 11 Ξ ^ 12 Ξ ^ 13 Ξ ^ 12 T Ξ ^ 22 Ξ ^ 23 Ξ ^ 13 T Ξ ^ 23 T Ξ ^ 33
with:
Ξ ^ 11 = P Q + 3 4 h 2 ( k ) A 1 T Q A 1 M 1 T ( I + h ( k ) A 1 ) ( I + h ( k ) A 1 ) T M 1 Ξ ^ 12 = M 1 T ( I + h ( k ) A 1 ) T M 2 Ξ ^ 13 = 3 4 h 2 ( k ) A 1 T Q A + Q h ( k ) M 1 T A Ξ ^ 22 = P Q + M 2 T + M 2 Ξ ^ 23 = Q h ( k ) M 2 T A Ξ ^ 33 = 3 4 h 2 ( k ) A T Q A 2 Q
If the inequality (9) holds for h ( k ) = h M and h ( k ) = h m , simultaneously, then by the Schur complement, one has Ξ ^ < 0 , which means that there exists a constant λ > 0 such that Δ V ( t k ) λ x T ( t k ) x ( t k ) . Therefore, the system (8) is asymptotically stable. This completes the proof. □
Remark 2.
Proposition 1 is obtained based on a simple Lyapunov function rather than a complicated Lyapunov–Krasovskii functional involved if using the input delay approach and the impulsive model approach. Moreover, from (10), one can find that two free-weighting n × n matrices M 1 and M 2 are introduced to build the relationship among x ( t k ) , x ( t k + 1 ) and x ( s ) .
Remark 3.
For Proposition 1, the total number of scalar decision variables is M = 2 × n ( n + 1 ) 2 + 2 n 2 = 3 n 2 + n , and the total row size of the LMIs is L = 5 n . The numerical complexity of Proposition 1 is proportional to L M 3 [27]. From Table 1, one can find that the stability criterion provided by Proposition 1 requires less scalar decision variables and a lesser row size of the LMIs compared with some existing stability criteria.

2.3. A Robust Stability Criterion

If there exist time-varying norm-bounded parameter uncertainties in (1), then (1) becomes:
x ˙ ( t ) = ( A + Δ A ( t ) ) x ( t ) + ( B + Δ B ( t ) ) u ( t ) , x ( 0 ) = x 0
where Δ A ( t ) and Δ B ( t ) represent norm-bounded parameter uncertainties of the form:
Δ A ( t ) Δ B ( t ) = D F ( t ) [ E E b ]
where D, E and E b are known real constant matrices with compatible dimensions; the unknown time-varying matrix F ( t ) R i 1 × i 2 satisfies:
F T ( t ) F ( t ) I   for   all   t 0 .
Similar to (3), the closed-loop system of (11) is given by:
x ˙ ( t ) = ( A + Δ A ( t ) ) x ( t ) + ( A 1 + Δ A 1 ( t ) ) x ( t k ) , x ( 0 ) = x 0
where Δ A 1 ( t ) = Δ B ( t ) K .
Extending Proposition 1 to the uncertain system (14) gives the following robust stability criterion.
Proposition 2.
For given scalars h M h m > 0 , the system (14) is robustly stable if there exist some matrices P > 0 , Q > 0 , R > 0 , M 1 , M 2 of appropriate dimensions such that the following LMI holds for h ( k ) = h m and h ( k ) = h M , simultaneously.
Ψ = Ψ 11 Ψ 12 Ψ 13 h ( k ) A 1 T Q h ( k ) M 1 T D Ψ 12 T Ψ 22 Ψ 23 0 Ψ 25 Ψ 13 T Ψ 23 T Ψ 33 h ( k ) A T Q 0 h ( k ) Q A 1 0 h ( k ) Q A 4 3 Q h ( k ) Q D h ( k ) D T M 1 Ψ 25 T 0 h ( k ) D T Q h ( k ) R < 0
where:
Ψ 11 = P Q M 1 T ( I + h ( k ) A 1 ) ( I + h ( k ) A 1 ) T M 1 + h ( k ) E b T R E b Ψ 12 = M 1 T ( I + h ( k ) A 1 ) T M 2 Ψ 22 = P Q + M 2 T + M 2 Ψ 13 = Q h ( k ) M 1 T A + h ( k ) E b T R E Ψ 23 = Q h ( k ) M 2 T A Ψ 25 = h ( k ) M 2 T D Ψ 33 = 2 Q + h ( k ) E T R E
Proof. 
Replacing A and A 1 in Proposition 1 with A + Δ A ( t ) and A 1 + Δ A 1 ( t ) , respectively, one can see that the system (14) is robustly stable if the following inequality holds:
Ξ + h ( k ) Υ 1 T F ( t ) Υ 2 + h ( k ) Υ 2 T F T ( t ) Υ 1 < 0
where:
Υ 1 = [ D T M 1 D T M 2 0 D T Q T ] , Υ 2 = [ E b 0 E 0 ] .
Then, the above matrix inequality is inferred by:
Ξ + h ( k ) Υ 1 T R 1 Υ 1 + h ( k ) Υ 2 T R Υ 2 < 0
for any matrix R > 0 of appropriate dimensions. By the Schur complement, (16) is equivalent to LMI (15). Thus, if the inequality (15) is satisfied, then so is (16), leading to robust stability of the system (14), which completes the proof. □

3. Numerical Examples

Example 1.
Consider the following much-studied problem [29]:
x ˙ ( t ) = x ( t k ) , t k t < t k + 1 , k = 0 , 1 , 2 ,
The system (17) remains stable for all constant samplings less than two and becomes unstable for samplings greater than two [5]. By Proposition 1 in [5] and Proposition 1 in this paper, h m a x = 2 is found, but less scalar decision variables are needed by Proposition 1 in this paper from Table 1.
Example 2.
To illustrate the proposed stability criterion of linear systems with a constant/time-varying sampling, we consider a sampled-data system as follows: [30]
x ˙ ( t ) = 0 1 0 0.1 x ( t ) + 0 0.1 K x ( t k )
where K = 3.75 11.5 . If we transfer the above system to the following system with an input delay:
x ˙ ( t ) = 0 1 0 0.1 x ( t ) + 0 0.1 K x ( t τ ( t ) )
where τ ( t ) = t t k , then the system (19) becomes unstable for τ ( t ) > 1.167144. This is because the analytical delay limit for the stability of (19) can be calculated as τ a n a l y t i c a l = 1.167144 for constant delay τ ( t ) = τ . Therefore, if we transfer (18) to a delay system (19), then we cannot guarantee the stability for samplings greater than 1.167144. Figure 2 shows the trajectories of states of (19) with constant delay τ = 1.167144. Moreover, it was found by applying an eigenvalue-based analysis that the system (18) remains stable for all constant samplings less than 1.729 and becomes unstable for samplings greater than 1.729 [28].
To show the effectiveness of the proposed method, we now consider two cases of sampling.
Under Case I, the sampling is a constant. Applying Proposition 1, the maximum allowable sampling period can be obtained as h m a x = 1.6962, which means that the system (18) is asymptotically stable for the constant sampling h = 1.6962, which is demonstrated by Figure 3. To further show the effectiveness of the proposed method, we make a comparison with some existing results. The obtained maximum allowable sampling periods h by some existing methods in [5,12,19,31] are listed in Table 2. Moreover, the total numbers of scalar decision variables M and the total row size L of the corresponding LMIs are also given in this table. From this table, one can see that Proposition 1 outperforms [5,19,31] from both the maximum allowable period h and the computational complexity. Although the obtained maximum h by Proposition 1 is smaller than that by [12,28], the computational complexity of Proposition 1 is less than that in [12,28]. Therefore, as a tradeoff between the maximum sampling period h and the complexity, Proposition 1 is more effective than [12,28].
Under Case II, the sampling is time-varying. Then, by Proposition 1, we calculate the maximum allowable sampling interval that retains the stability of the system (18). For different h m , the obtained results are listed in Table 3. However, these results cannot be derived using the standard discretization method [3,24,25,26].
Example 3.
Let us consider the uncertain sampled-data system (11) with matrices:
A = 0 1 0 0 0 0 3.3235 0.0212 0.0184 0.0030 5.3449 0.8819 0 0 0 1 0 0 0.0184 0.0030 118.1385 0.11188 5.3465 0.8822 0 0 0 0 0 1 0.0114 0.0019 0.0114 0.0019 3.3501 0.5454 , B = 0 0.003445 0 0.00344628 0 0.00213 T , D = 0 0 0 0 0 0 0 0.1 0 0 0 0 0 0 0 0 0 0 0 0 0 0.1 0 0 0 0 0 0 0 0 0 0 0 0 0 0.1 , E = 0.2 0 0 0 0 0 0 0.2 0 0 0 0 0 0 0.2 0 0 0 0 0 0 0.2 0 0 0 0 0 0 0.2 0 0 0 0 0 0 0.2 , E b = 0 0 0 0 0 0 , K = 10,100 5034 3634 22,432 5426 7681 ,
borrowed from the offshore steel jacket platforms [32]. We then apply Proposition 2 to obtain upper bounds h M for different lower bounds h m of samplings. The results are given in Table 4, which show the effectiveness of the given robust stability criterion for uncertain sampled-data systems.

4. Conclusions

The stability problem has been studied in this paper for linear systems under time-varying sampling. First, by using a non-standard discretization method, the sampled-data system has been represented by a linear discrete-time system. Then, by introducing a new sampled-data-based integral inequality, several sufficient conditions on stability and robust stability for sampled-data systems and the uncertain sampled-data systems, respectively, have been provided in terms of LMIs. These criteria are of lower computational complexity since less scalar decision variables and smaller row sizes of the LMIs are required. Finally, the effectiveness of the proposed criteria has been demonstrated by three numerical examples.

Author Contributions

X.J. proposed the sampled-data-based integral inequality and stability criteria, designed the experiments and wrote the paper; Z.Y. performed the experiments; J.W. analyzed the data.

Funding

The research work was partially supported by the Key project of Natural Science Foundation of Zhejiang Province of China (Grant No. LZ13F030001) and the National Natural Science Foundation of the People’s Republic of China under Grant 61673148, 61603118.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hetel, L.; Fiter, C.; Omran, H.; Seuret, A.; Fridman, E.; Richard, J.-P.; Niculescu, S. Recent developments on the stability of systems with aperiodic sampling: An overview. Automatica 2017, 76, 309–335. [Google Scholar] [CrossRef]
  2. Zhang, B.-L.; Han, Q.-L.; Zhang, X.-M. Recent advances in vibration control of offshore platforms. Nonlinear Dyn. 2017, 89, 755–771. [Google Scholar] [CrossRef]
  3. Koichi, K.; Kunihiko, H. An optimization-based approach to sampled-data control of networked control systems with multiple delays. Appl. Math. Comput. 2014, 247, 786–794. [Google Scholar]
  4. Zhang, X.-M.; Han, Q.-L.; Yu, X. Survey on recent advances in networked control systems. IEEE Trans. Ind. Inf. 2016, 12, 1740–1752. [Google Scholar] [CrossRef]
  5. Fridman, E. A refined input delay approach to sampled-data control. Automatica 2010, 46, 421–427. [Google Scholar] [CrossRef]
  6. Xiao, S.-P.; Lian, H.; Teo, K.; Zeng, H.-B.; Zhang, X.-H. A new Lyapunov functional approach to sampled-data synchronization control for delayed neural networks. J. Franklin Inst. 2018. to be published. [Google Scholar] [CrossRef]
  7. Fridman, E.; Seuret, A.; Richard, J.-P. Robust sampled-data stabilization of linear systems: An input delay approach. Automatica 2004, 40, 1441–1446. [Google Scholar] [CrossRef]
  8. Lee, T.; Park, J. Improved criteria for sampled-data synchronization of chaotic Lur’e systems using two new approaches. Nonlinear Anal. Hybrid Syst. 2017, 24, 132–145. [Google Scholar] [CrossRef]
  9. Zhang, X.; Han, Q. Network-based H filtering using a logic jumping-like trigger. Automatica 2013, 49, 1428–1435. [Google Scholar] [CrossRef]
  10. Ge, X.; Han, Q.-L.; Zhang, X.-M. Achieving cluster formation of multi-agent systems under aperiodic sampling and communication delays. IEEE Trans. Ind. Electron. 2018, 65, 3417–3426. [Google Scholar] [CrossRef]
  11. Zhang, X.; Han, Q. A decentralized event-triggered dissipative control scheme for systems with multiple sensors to sample the system outputs. IEEE Trans. Cybernet. 2016, 46, 2745–2757. [Google Scholar] [CrossRef] [PubMed]
  12. Seuret, A. A novel stability analysis of linear systems under asynchronous sampling. Automatica 2012, 48, 177–182. [Google Scholar] [CrossRef]
  13. Zhang, B.-L.; Han, Q.-L.; Zhang, X.-M. Event-triggered H reliable control for offshore structures in network environments. J. Sound Vib. 2016, 368, 1–21. [Google Scholar] [CrossRef]
  14. Zhang, X.-M.; Han, Q.-L. Event-triggered H control for a class of nonlinear networked control systems using novel integral inequalities. Int. J. Robust Nonlinear Control 2017, 27, 679–700. [Google Scholar] [CrossRef]
  15. Briat, C. Stability analysis and stabilization of stochastic linear impulsive, switched and sampled-data systems under dwell-time constraints. Automatica 2016, 74, 279–287. [Google Scholar] [CrossRef] [Green Version]
  16. Du, Z.; Qin, Z.; Ren, H.; Lu, Z. Fuzzy robust H sampled-data control for uncertain nonlinear systems with time-varying delay. Int. J. Fuzzy Syst. 2017, 19, 1417–1429. [Google Scholar] [CrossRef]
  17. Lee, T.; Wu, Z.-G.; Park, J. Synchronization of a complex dynamical network with coupling time-varying delays via sampled-data control. Appl. Math. Comput. 2012, 219, 1354–1366. [Google Scholar] [CrossRef]
  18. Tsai, S.-H.; Ko, K.-W.; Jen, C.-Y.; Li, S.-Y. Sampled-data control for polynomial fuzzy systems with time-varying delay. In Proceedings of the 2017 IEEE International Conference on System, Man, and Cybernetice, Banff Center, AB, Canada, 5–8 October 2017; pp. 2963–2967. [Google Scholar]
  19. Naghshtabrizi, P.; Hespanha, J.P.; Teel, A.R. Exponential stability of impulsive systems with application to uncertain sampled-data systems. Syst. Control Lett. 2008, 57, 378–385. [Google Scholar] [CrossRef] [Green Version]
  20. Chen, W.-H.; Zheng, W.X. Input-to-state stability for networked control systems via an improved impulsive system approach. Automatica 2011, 47, 789–796. [Google Scholar] [CrossRef]
  21. Bamieh, B.; Pearson, J.; Francis, A. A lifting technique for linear periodic systems. Syst. Control Lett. 2014, 17, 79–88. [Google Scholar] [CrossRef]
  22. Briat, C. Convex conditions for robust stability analysis and stabilization of linear aperiodic impulsive and sampled-data systems under dwell-time constraints. Automatica 2013, 49, 3449–3457. [Google Scholar] [CrossRef]
  23. Oishi, Y.; Fujioka, H. Stability and stabilization of aperiodic sampled-data control systems using robust linear matrix inequalities. Automatica 2010, 46, 1327–1333. [Google Scholar] [CrossRef] [Green Version]
  24. Jiang, X. On sampled-data fuzzy control design approach for T-S model-based fuzzy systems by using discretization approach. Inf. Sci. 2015, 296, 307–314. [Google Scholar] [CrossRef]
  25. Suh, Y. Stability and stabilization of nonuniform sampling systems. Automatica 2008, 44, 3222–3226. [Google Scholar] [CrossRef]
  26. Sun, H.; Chen, J.; Sun, J. Stochastic optimal control for sampled-data system under stochastic sampling. IET Control Theory Appl. 2018, 12, 1553–1560. [Google Scholar] [CrossRef]
  27. Han, Q.-L. A discrete delay decomposition approach to stability of linear retarded and neutral systems. Automatica 2009, 45, 1948–1952. [Google Scholar] [CrossRef]
  28. Seuret, A.; Briat, C. Stability analysis of uncertain sampled-data systems with incremental delay using looped-functionals. Automatica 2015, 55, 274–278. [Google Scholar] [CrossRef]
  29. Papachristodoulou, A.; Peet, M.; Niculescu, S. Stability analysis of linear systems with time-varying delays: delay uncertainty and quenching. In Proceedings of the 46th IEEE CDC, New Orleans, LA, USA, 12–14 December 2007; pp. 2117–2122. [Google Scholar]
  30. Naghshtabrizi, P.; Hespanha, J.; Teel, A. On the robust stability and stabilization of sampled-data systems: A hybrid system approach. In Proceedings of the 45th IEEE Conference on Decision and Control, San Diego, CA, USA, 13–15 December 2006; pp. 4873–4878. [Google Scholar]
  31. Zhang, C.-K.; Jiang, L.; He, Y.; Wu, H.; Wu, M. Stability analysis for control systems with aperiodically sampled data using an augmented Lyapunov functional method. IET Control Theory Appl. 2012, 7, 1219–1226. [Google Scholar] [CrossRef]
  32. Sivaranjani, K.; Pakkiyappan, R.; Lakshmanan, S.; Lim, C.P. Robust stochastic sampled-data control for offshore steel jacket platforms with non-linear perturbations. IMA J. Math. Control Inf. 2017, 34, 337–357. [Google Scholar] [CrossRef]
Figure 1. The integral regions.
Figure 1. The integral regions.
Electronics 07 00278 g001
Figure 2. The trajectories of states x 1 and x 2 of (19) with constant delay h = 1.167144 .
Figure 2. The trajectories of states x 1 and x 2 of (19) with constant delay h = 1.167144 .
Electronics 07 00278 g002
Figure 3. The trajectories of states x 1 and x 2 of (18) with constant sampling 1.6962 .
Figure 3. The trajectories of states x 1 and x 2 of (18) with constant sampling 1.6962 .
Electronics 07 00278 g003
Table 1. The total number of scalar decision variables and the total row size of LMIs.
Table 1. The total number of scalar decision variables and the total row size of LMIs.
Method[19][5][12][9][28]Proposition 1
M 4 n 2 + n 8 n 2 + n 5 n 2 + 2 n 5 n 2 + 2 n 10.5 n 2 + 3.5 n 3 n 2 + n
L 7 n 10 n 11 n 7 n 9n 5 n
L M 3 7 n ( 4 n 2 + n ) 3 10 n ( 8 n 2 + n ) 3 11 n ( 5 n 2 + 2 n ) 3 7 n ( 5 n 2 + 2 n ) 3 9 n ( 10.5 n 2 + 3.5 n ) 3 5 n ( 3 n 2 + n ) 3
Table 2. h, M and L of LMIs.
Table 2. h, M and L of LMIs.
Method[19][5][12][31][28]Proposition 1
h 1.3277 1.69 1.723 1.69 1.72 1.6962
M 183424244914
L 142022141810
L M 3 81,648786,080304,128193,5362,117,68227,440
Table 3. Maximum bounds of h M for different h m by Proposition 1.
Table 3. Maximum bounds of h M for different h m by Proposition 1.
h m 0.0010.010.10.51.01.51.69
h M 1.60661.61801.65071.68681.69371.69571.6962
Table 4. Maximum bounds of h M for different h m by Proposition 2.
Table 4. Maximum bounds of h M for different h m by Proposition 2.
h m 0.10.110.120.130.14
h M 0.12930.14080.14670.14930.1499

Share and Cite

MDPI and ACS Style

Jiang, X.; Yin, Z.; Wu, J. Stability Analysis of Linear Systems under Time-Varying Samplings by a Non-Standard Discretization Method. Electronics 2018, 7, 278. https://0-doi-org.brum.beds.ac.uk/10.3390/electronics7110278

AMA Style

Jiang X, Yin Z, Wu J. Stability Analysis of Linear Systems under Time-Varying Samplings by a Non-Standard Discretization Method. Electronics. 2018; 7(11):278. https://0-doi-org.brum.beds.ac.uk/10.3390/electronics7110278

Chicago/Turabian Style

Jiang, Xiefu, Zongming Yin, and Jinjing Wu. 2018. "Stability Analysis of Linear Systems under Time-Varying Samplings by a Non-Standard Discretization Method" Electronics 7, no. 11: 278. https://0-doi-org.brum.beds.ac.uk/10.3390/electronics7110278

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

Article Metrics

Back to TopTop