This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Observer-based Event-triggered Boundary Control of a Class of Reaction-Diffusion PDEs

Bhathiya Rathnayake, Mamadou Diagne, Nicolás Espitia, and Iasson Karafyllis B. Rathnayake is with the Department of Electrical, Computer, and Systems Engineering, Rensselaer Polytechnic Institute, New York, 12180, USA. Email: rathnb@rpi.eduM. Diagne is with the Department of Mechanical, Aerospace, and Nuclear Engineering, Rensselaer Polytechnic Institute, New York, 12180, USA. Email: diagnm@rpi.eduN. Espitia is with CRIStAL UMR 9189 CNRS - Centre de Recherche en Informatique Signal et Automatique de Lille - CNRS, Centrale Lille, Univ. Lille, F-59000 Lille, France. Email: nicolas.espitia-hoyos@univ-lille.frI. Karafyllis is with the Department of Mathematics, National Technical University of Athens, Greece. Email: iasonkar@central.ntua.gr
Abstract

This paper presents an observer-based event-triggered boundary control strategy for a class of reaction-diffusion PDEs with Robin actuation. The observer only requires boundary measurements. The control approach consists of a backstepping output feedback boundary controller, derived using estimated states, and a dynamic triggering condition, which determines the time instants at which the control input needs to be updated. It is shown that under the proposed observer-based event-triggered boundary control approach, there is a minimal dwell-time between two triggering instants independent of initial conditions. Furthermore, the well-posedness and the global exponential convergence of the closed-loop system to the equilibrium point are established. A simulation example is provided to validate the theoretical developments.

Index Terms:
Backstepping control design, event-triggered control, linear reaction-diffusion systems, output feedback.

I Introduction

Event-triggered control (ETC) is a control implementation technique that closes the feedback loop only if an event indicates that the control input error exceeds an appropriate threshold. When an event occurs, the controller computes the control value and transmits it to the actuator, completing the feedback path. Therefore, unlike periodic sampled-data control [1, 2], ETC only requires the control input to be updated aperiodically (only when needed). For networked and embedded control systems, the periodic computation and transmission of the control inputs are sometimes not desirable due to task scheduling limitations and bandwidth constraints in the communication [3, 4, 5, 6]. Despite the developments in aperiodic sampled-data control [7, 8, 9, 10], they lack explicit criteria for selecting appropriate sampling schedules. ETC, on the other hand, provides a rigorous resource-aware method of implementing the control laws into digital platforms [11, 12, 13].

In general, ETC consists of two main components: a feedback control law that stabilizes the system and an event-triggered mechanism, which determines when the control value has to be computed and sent toward the actuator. In the literature, one can find two main approaches to the control law design: emulation (e.g. [11]) and co-design (e.g. [5]). The former requires a pre-designed continuous feedback controller applied to the plant in a Zero-Order-Hold fashion between two event times. In co-design,  the feedback control law and the event-triggered mechanism are simultaneously designed to obtain the desired stability properties. An important property that every ETC design should possess is the non-existence of Zeno behavior[13]; otherwise, it will lead to the triggering of an infinite number of control updates over a finite period, making the design infeasible for digital implementation. Usually, ETC designs are ensured to be Zeno-free by showing the existence of a guaranteed lower bound for the time between two consecutive events, known as the minimal dwell-time.

In the case of finite-dimensional systems, ETC has grown to be a mature field of research. During the past decade, many significant results related to ETC have been reported on systems described by linear and nonlinear  ODEs in both full-state and output feedback settings (see [14, 11, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]). Recently, there has been a growing interest in periodic ETC [16, 22, 23], which employs a sampled-data event-triggered mechanism instead of the usual continuous-time event trigger. With this, the system states and the triggering condition need to be monitored and evaluated only at the sampling instants, making ETC more realistic for digital implementation. Despite some recent developments such as in [26, 27, 28, 29, 30, 31, 32, 33], ETC strategies for PDE systems have not reached the level of maturity seen by finite-dimensional systems yet. Only recently, even the notions of solutions of linear hyperbolic and parabolic PDEs under sampled-data control have been clarified [8, 9].

One can find several recent works that employ event-triggered boundary control strategies based on emulation on hyperbolic and parabolic PDEs [27, 29, 31, 32]. The authors of [27] propose an output feedback event-triggered boundary controller for 1-D  linear hyperbolic systems of conservation laws, using Lyapunov techniques. By utilizing a dynamic triggering condition,  an event-based backstepping boundary controller is designed for a coupled 2×22\times 2 hyperbolic system in [29]. This work is extended in [31] to obtain an event-triggered output feedback boundary controller for a similar system. In the case of parabolic PDEs, [32] is the first work that reports an event-based boundary control design. Using ISS properties and small gain arguments, the authors propose a full state feedback backstepping ETC strategy for a reaction-diffusion system with constant parameters and Dirichlet boundary conditions.

The importance of event-triggered output feedback control cannot be stressed enough as the use of full state measurements is either impossible or prohibitively expensive for many practical applications. Some studies such as [26, 34, 35] report several event-triggered output feedback designs for parabolic PDEs. All these works, however, rely on in-domain control and distributed observation. Event-based boundary control of parabolic PDEs with boundary sensing only is quite challenging, and no prior results have been reported. The possibility of avoiding Zeno behavior in ETC of general parabolic PDE systems with boundary actuation and boundary observation is not known. The actuation type is critical as both Dirichlet and Neumann actuation pose a severe impediment in establishing a minimal dwell-time and hence well-posedness and convergence results, due to unbounded local terms. However, we have identified that a class of reaction-diffusion equations with Robin actuation is conducive for event-triggered boundary control with boundary sensing. This paper proposes an observer-based event-triggered backstepping boundary controller for the class of PDEs mentioned above using emulation. The observer only requires boundary observation. The main contributions are as follows:

  • We consider a class of reaction-diffusion systems with Robin actuation. We perform emulation on an observer-based backstepping boundary control design and propose a dynamic triggering condition under which Zeno behavior is excluded. It is proved the existence of a minimal-dwell-time independent of the initial conditions.

  • We prove the well-posedness of the closed-loop system and its global exponential convergence to the equilibrium point in L2L^{2}-sense.

The paper is organized as follows. Section 2 introduces the class of linear reaction-diffusion system and the continuous-time output feedback boundary control. Section 3 presents the observer-based event-triggered boundary control and some properties. In Section 4, we discuss the main results of this paper. We provide a numerical example in Section 5 to illustrate the results and conclude the paper in Section 6.

Notation: +\mathbb{R}_{+} is the nonnegative real line whereas \mathbb{N} is the set of natural numbers including zero. By C0(A;Ω)C^{0}(A;\Omega), we denote the class of continuous functions on AnA\subseteq\mathbb{R}^{n}, which takes values in Ω\Omega\subseteq\mathbb{R}. By Ck(A;Ω)C^{k}(A;\Omega), where k1k\geq 1, we denote the class of continuous functions on AA, which takes values in Ω\Omega and has continuous derivatives of order kk. L2(0,1)L^{2}(0,1) denotes the equivalence class of Lebesgue measurable functions f:[0,1]f:[0,1]\rightarrow\mathbb{R} such that f=(01|f(x)|2)1/2<\|f\|=\big{(}\int_{0}^{1}|f(x)|^{2}\big{)}^{1/2}<\infty. Let u:[0,1]×+u:[0,1]\times\mathbb{R}_{+}\rightarrow\mathbb{R} be given. u[t]u[t] denotes the profile of uu at certain t0t\geq 0, i.e., (u[t])(x)=u(x,t),\big{(}u[t]\big{)}(x)=u(x,t), for all x[0,1]x\in[0,1]. For an interval I+,I\subseteq\mathbb{R}_{+}, the space C0(I;L2(0,1))C^{0}\big{(}I;L^{2}(0,1)\big{)} is the space of continuous mappings Itu[t]L2(0,1)I\ni t\rightarrow u[t]\in L^{2}(0,1). Im(),I_{m}(\cdot), and Jm()J_{m}(\cdot) with mm being an integer respectively denote the modified Bessel and (nonmodified) Bessel functions of the first kind.

II Observer-Based Backstepping Boundary Control and Emulation

Let us consider the following 1-D reaction-diffusion system with constant coefficients:

ut(x,t)\displaystyle u_{t}(x,t) =εuxx(x,t)+λu(x,t),\displaystyle=\varepsilon u_{xx}(x,t)+\lambda u(x,t), (1a)
ux(0,t)\displaystyle u_{x}(0,t) =0,\displaystyle=0, (1b)
ux(1,t)+qu(1,t)\displaystyle u_{x}(1,t)+qu(1,t) =U(t),\displaystyle=U(t), (1c)

and the initial condition u[0]L2(0,1),u[0]\in L^{2}(0,1), where ε,λ>0,\varepsilon,\lambda>0, u:[0,1]×[0,)u:[0,1]\times[0,\infty)\rightarrow\mathbb{R} is the system state and U(t)U(t) is the control input.

Assumption 1:

q>(λ+ε)/2ε.q>(\lambda+\varepsilon)/2\varepsilon.

Remark 1:

It should be mentioned that the solution of the PDE system defined as (1) with U(t)=0U(t)=0 (zero input), where ε,q>0\varepsilon,q>0 and λ\lambda\in\mathbb{R} are constants, satisfies the estimate

u[t]e(λεμ2)tu[0],\|u[t]\|\leq e^{(\lambda-\varepsilon\mu^{2})t}\|u[0]\|, (2)

for all t0t\geq 0 and u[0]L2(0,1),u[0]\in L^{2}(0,1), where μ\mu is the unique solution of the transcendental equation μtan(μ)=q\mu\tan(\mu)=q in the interval (0,π/2)(0,\pi/2) [36]. The estimate (2) guarantees the global exponential stability of the zero-input system in L2L^{2}-norm when λ<εμ2\lambda<\varepsilon\mu^{2}. Further, an eigenfunction expansion of its solution shows that the system is unstable when λ>εμ2\lambda>\varepsilon\mu^{2}. Since μ\mu is in the interval (0,π/2)(0,\pi/2), this implies instability when λ>επ2/4\lambda>\varepsilon\pi^{2}/4, no matter what q>0q>0 is.

We propose an observer for the system (1) using u(0,t)u(0,t) as the available measurement/output. Note that the output is anticollocated with the input. The observer consists of a copy of the system (1) with output injection terms, which is stated as follows:

u^t(x,t)=εu^xx(x,t)+λu^(x,t)+p1(x)(u(0,t)u^(0,t)),\displaystyle\begin{split}\hat{u}_{t}(x,t)&=\varepsilon\hat{u}_{xx}(x,t)+\lambda\hat{u}(x,t)\\ &\qquad+p_{1}(x)\big{(}u(0,t)-\hat{u}(0,t)\big{)},\end{split} (3a)
u^x(0,t)\displaystyle\hat{u}_{x}(0,t) =p10(u(0,t)u^(0,t)),\displaystyle=p_{10}\big{(}u(0,t)-\hat{u}(0,t)\big{)}, (3b)
u^x(1,t)+qu^(1,t)\displaystyle\hat{u}_{x}(1,t)+q\hat{u}(1,t) =U(t),\displaystyle=U(t), (3c)

and the initial condition u^[0]L2(0,1)\hat{u}[0]\in L^{2}(0,1). Here, the function p1(x)p_{1}(x) and the constant p10p_{10} are observer gains to be determined. Let us denote the observer error by u~(x,t)\tilde{u}(x,t), which is defined as

u~(x,t):=u(x,t)u^(x,t).\tilde{u}(x,t):=u(x,t)-\hat{u}(x,t). (4)

By subtracting (3) from (1), one can see that u~(x,t)\tilde{u}(x,t) satisfies the following PDE:

u~t(x,t)=εu~xx(x,t)+λu~(x,t)p1(x)u~(0,t),\displaystyle\begin{split}\tilde{u}_{t}(x,t)&=\varepsilon\tilde{u}_{xx}(x,t)+\lambda\tilde{u}(x,t)-p_{1}(x)\tilde{u}(0,t),\end{split} (5a)
u~x(0,t)\displaystyle\tilde{u}_{x}(0,t) =p10u~(0,t),\displaystyle=-p_{10}\tilde{u}(0,t), (5b)
u~x(1,t)+qu~(1,t)\displaystyle\tilde{u}_{x}(1,t)+q\tilde{u}(1,t) =0.\displaystyle=0. (5c)
Proposition 1:

Subject to Assumption 1 and the invertible backstepping transformation

u~(x,t)=w~(x,t)0xP(x,y)w~(y,t)𝑑y,\tilde{u}(x,t)=\tilde{w}(x,t)-\int_{0}^{x}P(x,y)\tilde{w}(y,t)dy, (6)

where

P(x,y)=qλ/ελ/ε+q2×0xyeqτ/2I0(λ(2xy)(xyτ)/ε)×sinh(λ/ε+q22τ)dτλε(1y)I1(λ((1y)2(1x)2)/ε)λ((1y)2(1x)2)/ε,\begin{split}P(x,y)=&\frac{q\lambda/\varepsilon}{\sqrt{\lambda/\varepsilon+q^{2}}}\\ &\times\int_{0}^{x-y}e^{-q\tau/2}I_{0}\Big{(}\sqrt{\lambda(2-x-y)(x-y-\tau)/\varepsilon}\Big{)}\\ &\hskip 35.0pt\times\sinh\Big{(}\frac{\sqrt{\lambda/\varepsilon+q^{2}}}{2}\tau\Big{)}d\tau\\ &-\frac{\lambda}{\varepsilon}(1-y)\frac{I_{1}\Big{(}\sqrt{\lambda\big{(}(1-y)^{2}-(1-x)^{2}\big{)}/\varepsilon}\Big{)}}{\sqrt{\lambda\big{(}(1-y)^{2}-(1-x)^{2}\big{)}/\varepsilon}},\end{split} (7)

for 0yx10\leq y\leq x\leq 1, then, the system (5) with the gains p1(x)p_{1}(x) and p10p_{10} chosen as

p1(x)=εPy(x,0),p10=P(0,0)=λ2ε,\displaystyle p_{1}(x)=\varepsilon P_{y}(x,0),\hskip 20.0ptp_{10}=P(0,0)=-\frac{\lambda}{2\varepsilon}, (8)

gets transformed to the following globally L2L^{2}-exponentially stable observer error target system

w~t(x,t)\displaystyle\tilde{w}_{t}(x,t) =εw~xx(x,t),\displaystyle=\varepsilon\tilde{w}_{xx}(x,t), (9a)
w~x(0,t)\displaystyle\tilde{w}_{x}(0,t) =0,\displaystyle=0, (9b)
w~x(1,t)\displaystyle\tilde{w}_{x}(1,t) =qw~(1,t).\displaystyle=-q\tilde{w}(1,t). (9c)

Proof: See Appendix A.

The inverse transformation of (6) can be shown to be as follows:

w~(x,t)=u~(x,t)+0xQ(x,y)u~(y,t)𝑑y,\tilde{w}(x,t)=\tilde{u}(x,t)+\int_{0}^{x}Q(x,y)\tilde{u}(y,t)dy, (10)

where Q(x,y)Q(x,y) is

Q(x,y)=qλ/ελ/ε+q2×0xyeqτ/2J0(λ(2xy)(xyτ)/ε)×sinh(λ/ε+q22τ)dτλε(1y)J1(λ((1y)2(1x)2)/ε)λ((1y)2(1x)2)/ε,\begin{split}Q(x,y)=&\frac{q\lambda/\varepsilon}{\sqrt{-\lambda/\varepsilon+q^{2}}}\\ &\times\int_{0}^{x-y}e^{-q\tau/2}J_{0}\Big{(}\sqrt{\lambda(2-x-y)(x-y-\tau)/\varepsilon}\Big{)}\\ &\hskip 35.0pt\times\sinh\Big{(}\frac{\sqrt{-\lambda/\varepsilon+q^{2}}}{2}\tau\Big{)}d\tau\\ &-\frac{\lambda}{\varepsilon}(1-y)\frac{J_{1}\Big{(}\sqrt{\lambda\big{(}(1-y)^{2}-(1-x)^{2}\big{)}/\varepsilon}\Big{)}}{\sqrt{\lambda\big{(}(1-y)^{2}-(1-x)^{2}\big{)}/\varepsilon}},\end{split} (11)

for 0yx10\leq y\leq x\leq 1.

Proposition 2:

The invertible backstepping transformation

w^(x,t)=u^(x,t)0xK(x,y)u^(y,t)𝑑y,\hat{w}(x,t)=\hat{u}(x,t)-\int_{0}^{x}K(x,y)\hat{u}(y,t)dy, (12)

where

K(x,y)=λεxI1(λ(x2y2)/ε)λ(x2y2)/ε,K(x,y)=-\frac{\lambda}{\varepsilon}x\frac{I_{1}\big{(}\sqrt{\lambda(x^{2}-y^{2})/\varepsilon}\big{)}}{\sqrt{\lambda(x^{2}-y^{2})/\varepsilon}}, (13)

for 0yx1,0\leq y\leq x\leq 1, and a control law U(t)U(t) chosen as

U(t)=01(rK(1,y)+Kx(1,y))u^(y,t)𝑑y,U(t)=\int_{0}^{1}\Big{(}rK(1,y)+K_{x}(1,y)\Big{)}\hat{u}(y,t)dy, (14)

map the system (3) with the gains p1(x)p_{1}(x) and p10p_{10} chosen as in (8), into the following target system:

w^t(x,t)\displaystyle\hat{w}_{t}(x,t) =εw^xx(x,t)+g(x)w~(0,t),\displaystyle=\varepsilon\hat{w}_{xx}(x,t)+g(x)\tilde{w}(0,t), (15a)
w^x(0,t)\displaystyle\hat{w}_{x}(0,t) =λ2εw~(0,t),\displaystyle=-\frac{\lambda}{2\varepsilon}\tilde{w}(0,t), (15b)
w^x(1,t)\displaystyle\hat{w}_{x}(1,t) =rw^(1,t),\displaystyle=-r\hat{w}(1,t), (15c)

with

g(x)=p1(x)λ2K(x,0)0xK(x,y)p1(y)𝑑y,g(x)=p_{1}(x)-\frac{\lambda}{2}K(x,0)-\int_{0}^{x}K(x,y)p_{1}(y)dy, (16)

and

r=qλ2ε.r=q-\frac{\lambda}{2\varepsilon}. (17)

Proof: See Appendix B.

The inverse transformation of (12) can be shown to be as follows:

u^(x,t)=w^(x,t)+0xL(x,y)w^(y,t)𝑑y,\hat{u}(x,t)=\hat{w}(x,t)+\int_{0}^{x}L(x,y)\hat{w}(y,t)dy, (18)

where

L(x,y)=λεxJ1(λ(x2y2)/ε)λ(x2y2)/ε,L(x,y)=-\frac{\lambda}{\varepsilon}x\frac{J_{1}\big{(}\sqrt{\lambda(x^{2}-y^{2})/\varepsilon}\big{)}}{\sqrt{\lambda(x^{2}-y^{2})/\varepsilon}}, (19)

for 0yx10\leq y\leq x\leq 1.

Proposition 3:

Subject to Assumption 1, the closed-loop system which consists of the plant (1) and the observer (3) with the continuous-time control law (14), is globally exponentially stable in L2L^{2}-sense.

Proof: See Appendix C.

II-A Emulation of the Observer-based Backstepping Boundary Control

We strive to stabilize the closed-loop system containing the plant (1) and the observer (3) while sampling the continuous-time controller U(t)U(t) given by (14) at a certain sequence of time instants (tj)j(t_{j})_{j\in\mathbb{N}}. These time instants will be given a precise characterization later on based on an event trigger. The control input is held constant between two successive time instants and is updated when a certain condition is met. Therefore, we define the control input for t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N} as

Uj:=U(tj)=01(rK(1,y)+Kx(1,y))u^(y,tj)𝑑y.U_{j}:=U(t_{j})=\int_{0}^{1}\Big{(}rK(1,y)+K_{x}(1,y)\Big{)}\hat{u}(y,t_{j})dy. (20)

Accordingly, the boundary conditions (1c) and (3c) are modified, respectively, as follows:

ux(1,t)+qu(1,t)=Uj,u_{x}(1,t)+qu(1,t)=U_{j}, (21)
u^x(1,t)+qu^(1,t)=Uj.\hat{u}_{x}(1,t)+q\hat{u}(1,t)=U_{j}. (22)

The deviation between the continuous-time control law and its sampled counterpart, referred to as the input holding error, is defined as follows:

d(t):=01(rK(1,y)+Kx(1,y))(u^(y,tj)u^(y,t))𝑑y.\begin{split}d(t):=\int_{0}^{1}\Big{(}rK(1,y)+K_{x}(1,y)\Big{)}\big{(}\hat{u}(y,t_{j})-\hat{u}(y,t)\big{)}dy.\end{split} (23)

for t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N}. It can be shown that the backstepping transformation (12), applied on the system (3a),(3b),(22) between tjt_{j} and tj+1t_{j+1}, yields the following target system, valid for t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N}:

w^t(x,t)\displaystyle\hat{w}_{t}(x,t) =εw^xx(x,t)+g(x)w~(0,t),\displaystyle=\varepsilon\hat{w}_{xx}(x,t)+g(x)\tilde{w}(0,t), (24a)
w^x(0,t)\displaystyle\hat{w}_{x}(0,t) =λ2εw~(0,t),\displaystyle=-\frac{\lambda}{2\varepsilon}\tilde{w}(0,t), (24b)
w^x(1,t)\displaystyle\hat{w}_{x}(1,t) =rw^(1,t)+d(t),\displaystyle=-r\hat{w}(1,t)+d(t), (24c)

where g(x)g(x) and rr are given by (16) and (17), respectively.

It is straightforward to see that the observer error system u~\tilde{u} for t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N} under the modified boundary conditions (21) and (22) will still be the same as (5). Therefore, the application of the backstepping transformation (6) on u~\tilde{u} between tjt_{j} and tj+1t_{j+1} yields the following observer error target system

w~t(x,t)\displaystyle\tilde{w}_{t}(x,t) =εw~xx(x,t),\displaystyle=\varepsilon\tilde{w}_{xx}(x,t), (25a)
w~x(0,t)\displaystyle\tilde{w}_{x}(0,t) =0,\displaystyle=0, (25b)
w~x(1,t)\displaystyle\tilde{w}_{x}(1,t) =qw~(1,t),\displaystyle=-q\tilde{w}(1,t), (25c)

valid for t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N}.

II-B Well-posedness Issues

Proposition 4:

For every given initial data u[tj],u^[tj]L2(0,1)u[t_{j}],\hat{u}[t_{j}]\in L^{2}(0,1), there exist unique mappings u,u^C0([tj,tj+1];L2(0,1))C1((tj,tj+1)×[0,1])u,\hat{u}\in C^{0}([t_{j},t_{j+1}];L^{2}(0,1))\cap C^{1}((t_{j},t_{j+1})\times[0,1]) with u[t],u^[t]C2([0,1])u[t],\hat{u}[t]\in C^{2}([0,1]) which satisfy (1b),(3b),(20)-(22) for t(tj,tj+1]t\in(t_{j},t_{j+1}] and (1a), (3a) for t(tj,tj+1]t\in(t_{j},t_{j+1}], x(0,1)x\in(0,1).

Proof: The initial condition w~[tj]\tilde{w}[t_{j}] for the system (25) can be uniquely determined by using (4) and (10) once u[tj]u[t_{j}] and u^[tj]\hat{u}[t_{j}] are given. Therefore, from the straightforward application of Theorem 4.11 in [37], we can show that there exist unique mappings u,w~C0([tj,tj+1];L2(0,1))C1((tj,tj+1)×[0,1])u,\tilde{w}\in C^{0}([t_{j},t_{j+1}];L^{2}(0,1))\cap C^{1}((t_{j},t_{j+1})\times[0,1]) with u[t],w~[t]C2([0,1])u[t],\tilde{w}[t]\in C^{2}([0,1]) which satisfy (1b),(20),(21),(25b),(25c) for t(tj,tj+1]t\in(t_{j},t_{j+1}] and (1a),(25a) for t(tj,tj+1]t\in(t_{j},t_{j+1}], x(0,1)x\in(0,1). Further, due to (4) and the transformation (6), there also exists a unique mapping u^C0([tj,tj+1];L2(0,1))C1((tj,tj+1)×[0,1])\hat{u}\in C^{0}([t_{j},t_{j+1}];L^{2}(0,1))\cap C^{1}((t_{j},t_{j+1})\times[0,1]) with u^[t]C2([0,1])\hat{u}[t]\in C^{2}([0,1]) which satisfy (3b),(20),(22) for t(tj,tj+1]t\in(t_{j},t_{j+1}]and (3a) for t(tj,tj+1]t\in(t_{j},t_{j+1}], x(0,1)x\in(0,1).

III Observer-based Event-triggered Boundary Control

Refer to caption
Figure 1: Event-triggered observer-based closed-loop system.

Let us now present the observer-based event-triggered boundary control approach considered in this work. It consists of two components: 1) an event-triggered mechanism which decides the time instants at which the control value needs to be sampled/updated and 2) the observer-based backstepping output feedback controller. The structure of the closed-loop system consisting of the plant, the observer-based controller, and the event trigger is illustrated in Fig. 1. The event-triggering condition is based on the square of the input holding error d(t)d(t) and a dynamic variable m(t)m(t) which depends on the information of the systems (24) and (25).

Definition 1:

Let η,ρ,β1,β2,β3>0\eta,\rho,\beta_{1},\beta_{2},\beta_{3}>0. The observer-based event-triggered boundary control strategy consists of two components.

  1. 1.

    (The event-trigger) For some jj\in\mathbb{N}, the event times tj0t_{j}\geq 0 with t0=0t_{0}=0 form a finite increasing sequence via the following rules:

    • if {t+|t>tjd2(t)>m(t)}=\{t\in\mathbb{R}_{+}|t>t_{j}\wedge d^{2}(t)>-m(t)\}=\emptyset then the set of the times of the events is {t0,,tj}.\{t_{0},\ldots,t_{j}\}.

    • if {t+|t>tjd2(t)>m(t)}\{t\in\mathbb{R}_{+}|t>t_{j}\wedge d^{2}(t)>-m(t)\}\neq\emptyset then the next event time is given by:

      tj+1=inf{t+|t>tjd2(t)>m(t)},t_{j+1}=\inf\{t\in\mathbb{R}_{+}|t>t_{j}\wedge d^{2}(t)>-m(t)\}, (26)

      where d(t)d(t) is given by

      d(t)=01(rK(1,y)+Kx(1,y))×(u^(y,tj)u^(y,t))dy,\begin{split}d(t)=\int_{0}^{1}\Big{(}&rK(1,y)+K_{x}(1,y)\Big{)}\\ &\times\big{(}\hat{u}(y,t_{j})-\hat{u}(y,t)\big{)}dy,\end{split} (27)

      for all t[tj,tj+1)t\in[t_{j},t_{j+1}) and m(t)m(t) satisfies the ODE

      m˙(t)=ηm(t)+ρd2(t)β1w^[t]2β2|w^(1,t)|2β3|w~(0,t)|2,\begin{split}\dot{m}(t)=&-\eta m(t)+\rho d^{2}(t)-\beta_{1}\|\hat{w}[t]\|^{2}\\ &-\beta_{2}|\hat{w}(1,t)|^{2}-\beta_{3}|\tilde{w}(0,t)|^{2},\end{split} (28)

      for all t(tj,tj+1)t\in(t_{j},t_{j+1}) with m(t0)=m(0)<0m(t_{0})=m(0)<0 and m(tj)=m(tj)=m(tj+)m(t_{j}^{-})=m(t_{j})=m(t_{j}^{+}).

  2. 2.

    (The control action) The output boundary feedback control law

    Uj=01(rK(1,y)+Kx(1,y))u^(y,tj)𝑑y,U_{j}=\int_{0}^{1}\Big{(}rK(1,y)+K_{x}(1,y)\Big{)}\hat{u}(y,t_{j})dy, (29)

    for all t[tj,tj+1),jt\in[t_{j},t_{j+1}),j\in\mathbb{N}.

In Definition 1, it is worth noting that the initial condition for m(t)m(t) in each time interval has been chosen such that m(t)m(t) is time-continuous.

Proposition (4) allows us to define the solution of the closed-loop system under the observer-based event-triggered boundary control (26)-(29) in the interval [0,F)[0,F), where F=limj(tj)F=\lim_{j\rightarrow\infty}(t_{j}).

Lemma 1:

Under the definition of the observer-based event-triggered boundary control (26)-(29), it holds that d2(t)m(t)d^{2}(t)\leq-m(t) and m(t)<0,m(t)<0, for t[0,F)t\in[0,F) where F=limj(tj)F=\lim_{j\rightarrow\infty}(t_{j}).

Proof: From Definition 1, the events are triggered to guarantee d2(t)m(t),t[0,F)d^{2}(t)\leq-m(t),t\in[0,F). This inequality in combination with (28) yields:

m˙(t)(η+ρ)m(t)β1w^[t]2β2|w^(1,t)|2β3|w~(0,t)|2,\begin{split}\dot{m}(t)\leq-(\eta+\rho)m(t)&-\beta_{1}\|\hat{w}[t]\|^{2}\\ &-\beta_{2}|\hat{w}(1,t)|^{2}-\beta_{3}|\tilde{w}(0,t)|^{2},\end{split} (30)

for t(tj,tj+1),j.t\in(t_{j},t_{j+1}),j\in\mathbb{N}. Thus, considering the time-continuity of m(t)m(t), we can obtain the following estimate:

m(t)m(tj)e(η+ρ)(ttj)tjte(η+ρ)(tτ)(β1w^[τ]2+β2|w^(1,τ)|2+β3|w~(0,τ)|2)dτ,\begin{split}&m(t)\leq m(t_{j})e^{-(\eta+\rho)(t-t_{j})}\\ &-\int_{t_{j}}^{t}e^{-(\eta+\rho)(t-\tau)}\big{(}\beta_{1}\|\hat{w}[\tau]\|^{2}\\ &\hskip 30.0pt+\beta_{2}|\hat{w}(1,\tau)|^{2}+\beta_{3}|\tilde{w}(0,\tau)|^{2}\big{)}d\tau,\end{split} (31)

for t[tj,tj+1],jt\in[t_{j},t_{j+1}],j\in\mathbb{N}. From Definition 1, we have that m(t0)=m(0)<0m(t_{0})=m(0)<0. Therefore, it follows from (31) that m(t)<0m(t)<0 for all t[0,t1]t\in[0,t_{1}]. Again using (31) on [t1,t2][t_{1},t_{2}], we can show that m(t)<0m(t)<0 for all t[t1,t2]t\in[t_{1},t_{2}]. Applying the same reasoning successively to the future intervals, it can be shown that m(t)<0m(t)<0 for t[0,F)t\in[0,F).

Lemma 2:

For d(t)d(t) given by (27), it holds that

d˙2(t)ρ1d2(t)+α1w^[t]2+α2|w^(1,t)|2+α3|w~(0,t)|2,\dot{d}^{2}(t)\leq\rho_{1}d^{2}(t)+\alpha_{1}\|\hat{w}[t]\|^{2}+\alpha_{2}|\hat{w}(1,t)|^{2}+\alpha_{3}|\tilde{w}(0,t)|^{2}, (32)

for some ρ1,α1,α2,α3>0,\rho_{1},\alpha_{1},\alpha_{2},\alpha_{3}>0, for all t(tj,tj+1),j.t\in(t_{j},t_{j+1}),j\in\mathbb{N}.

Proof: From (27), we can show for t(tj,tj+1),jt\in(t_{j},t_{j+1}),j\in\mathbb{N}

d˙(t)=01k(y)u^t(y,t)𝑑y,\dot{d}(t)=-\int_{0}^{1}k(y)\hat{u}_{t}(y,t)dy, (33)

where

k(y)=rK(1,y)+Kx(1,y).k(y)=rK(1,y)+K_{x}(1,y). (34)

Using (3a) on (33) and integrating by parts twice in the interval (tj,tj+1),j(t_{j},t_{j+1}),j\in\mathbb{N}, we can show that

d˙(t)=ε01k(y)u^yy(y,t)𝑑yλ01k(y)u^(y,t)𝑑y01k(y)p1(y)𝑑yu~(0,t)=εk(1)u^x(1,t)+εk(0)u^x(0,t)+εdk(x)dx|x=1u^(1,t)εdk(x)dx|x=0u^(0,t)ε01d2k(y)dy2u^(y,t)𝑑yλ01k(y)u^(y,t)𝑑y01k(y)p1(y)𝑑yu~(0,t).\begin{split}&\dot{d}(t)=-\varepsilon\int_{0}^{1}k(y)\hat{u}_{yy}(y,t)dy-\lambda\int_{0}^{1}k(y)\hat{u}(y,t)dy\\ &\hskip 28.0pt-\int_{0}^{1}k(y)p_{1}(y)dy\tilde{u}(0,t)\\ &=-\varepsilon k(1)\hat{u}_{x}(1,t)+\varepsilon k(0)\hat{u}_{x}(0,t)+\varepsilon\frac{dk(x)}{dx}\Big{|}_{x=1}\hat{u}(1,t)\\ &\hskip 12.0pt-\varepsilon\frac{dk(x)}{dx}\Big{|}_{x=0}\hat{u}(0,t)-\varepsilon\int_{0}^{1}\frac{d^{2}k(y)}{dy^{2}}\hat{u}(y,t)dy\\ &\hskip 12.0pt-\lambda\int_{0}^{1}k(y)\hat{u}(y,t)dy-\int_{0}^{1}k(y)p_{1}(y)dy\tilde{u}(0,t).\end{split} (35)

Furthermore, using (3b),(22), (27), and (29), we can show that

d˙(t)=εk(1)d(t)+(εqk(1)+εdk(x)dx|x=1)u^(1,t)01(εd2k(y)dy2+εk(1)k(y)+λk(y))u^(y,t)𝑑y(λk(0)2+01k(y)p1(y)𝑑y)u~(0,t).\begin{split}\dot{d}(t)=&-\varepsilon k(1)d(t)+\Big{(}\varepsilon qk(1)+\varepsilon\frac{dk(x)}{dx}\Big{|}_{x=1}\Big{)}\hat{u}(1,t)\\ &-\int_{0}^{1}\Big{(}\varepsilon\frac{d^{2}k(y)}{dy^{2}}+\varepsilon k(1)k(y)+\lambda k(y)\Big{)}\hat{u}(y,t)dy\\ &-\Big{(}\frac{\lambda k(0)}{2}+\int_{0}^{1}k(y)p_{1}(y)dy\Big{)}\tilde{u}(0,t).\end{split} (36)

It is worth mentioning that above we have used the fact that dk(x)/dx|x=0=0,dk(x)/dx\Big{|}_{x=0}=0, which can be shown using (34) and (13). Using Young’s and Cauchy-Schwarz inequalities on (18), we also can show that

u^[t]2(1+(010xL2(x,y)𝑑y𝑑x)1/2)2w^[t]2,\|\hat{u}[t]\|^{2}\leq\bigg{(}1+\Big{(}\int_{0}^{1}\int_{0}^{x}L^{2}(x,y)dydx\Big{)}^{1/2}\bigg{)}^{2}\|\hat{w}[t]\|^{2}, (37)
u^2(1,t)2w^2(1,t)+201L2(1,y)𝑑yw^[t]2.\hat{u}^{2}(1,t)\leq 2\hat{w}^{2}(1,t)+2\int_{0}^{1}L^{2}(1,y)dy\|\hat{w}[t]\|^{2}. (38)

Using Young’s and Cauchy-Schwarz inequalities repeatedly on (36) along with (37) and (38), we can show that

d˙2(t)ρ1d2(t)+α1w^[t]2+α2|w^(1,t)|2+α3|w~(0,t)|2,\begin{split}\dot{d}^{2}(t)\leq&\rho_{1}d^{2}(t)+\alpha_{1}\|\hat{w}[t]\|^{2}+\alpha_{2}|\hat{w}(1,t)|^{2}\\ &+\alpha_{3}|\tilde{w}(0,t)|^{2},\end{split} (39)

where

ρ1\displaystyle\rho_{1} =6ε2k2(1),\displaystyle=6\varepsilon^{2}k^{2}(1), (40)
α1=3(1+(010xL2(x,y)𝑑y𝑑x)1/2)2×01(εd2k(y)dy2+εk(1)k(y)+λk(y))2dy+6(εqk(1)+εdk(x)dx|x=1)201L2(1,y)𝑑y,\displaystyle\begin{split}\alpha_{1}&=3\bigg{(}1+\Big{(}\int_{0}^{1}\int_{0}^{x}L^{2}(x,y)dydx\Big{)}^{1/2}\bigg{)}^{2}\\ &\quad\times\int_{0}^{1}\Big{(}\varepsilon\frac{d^{2}k(y)}{dy^{2}}+\varepsilon k(1)k(y)+\lambda k(y)\Big{)}^{2}dy\\ &\quad+6\Big{(}\varepsilon qk(1)+\varepsilon\frac{dk(x)}{dx}\Big{|}_{x=1}\Big{)}^{2}\int_{0}^{1}L^{2}(1,y)dy,\end{split} (41)
α2\displaystyle\alpha_{2} =6(εqk(1)+εdk(x)dx|x=1)2,\displaystyle=6\Big{(}\varepsilon qk(1)+\varepsilon\frac{dk(x)}{dx}\Big{|}_{x=1}\Big{)}^{2}, (42)
α3\displaystyle\alpha_{3} =6(λk(0)2+01k(y)p1(y)𝑑y)2.\displaystyle=6\Big{(}\frac{\lambda k(0)}{2}+\int_{0}^{1}k(y)p_{1}(y)dy\Big{)}^{2}. (43)

IV Main Results

Theorem 1:

Under the observer-based event-triggered boundary control in Definition 1, with β1,β2,β3\beta_{1},\beta_{2},\beta_{3} chosen as

β1=α1/(1σ),β2=α2/(1σ),β3=α3/(1σ),\beta_{1}=\alpha_{1}/(1-\sigma),\hskip 5.0pt\beta_{2}=\alpha_{2}/(1-\sigma),\hskip 5.0pt\beta_{3}=\alpha_{3}/(1-\sigma), (44)

where α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3} given by (41)-(43) and σ(0,1)\sigma\in(0,1), there exists a minimal dwell-time τ>0\tau>0 between two triggering times, i.e., there exists a constant τ>0\tau>0 such that tj+1tjτ,t_{j+1}-t_{j}\geq\tau, for all jj\in\mathbb{N}, which is independent of the initial conditions and only depends on the system and control parameters.

Proof: From Lemma 1, we have that d2(t)(1σ)m(t)σm(t),d^{2}(t)\leq-(1-\sigma)m(t)-\sigma m(t), where σ(0,1)\sigma\in(0,1) and m(t)<0m(t)<0 for t[0,F)t\in[0,F), where F=limj(tj)F=\lim_{j\rightarrow\infty}(t_{j}). Let us define the function

ψ(t):=d2(t)+(1σ)m(t)σm(t).\psi(t):=\frac{d^{2}(t)+(1-\sigma)m(t)}{-\sigma m(t)}. (45)

Note that ψ(t)\psi(t) is continuous in [tj,tj+1)[t_{j},t_{j+1}). A lower bound for the inter-execution times is given by the time it takes for the function ψ\psi to go from ψ(tj)\psi(t_{j}) to ψ(tj+1)=1,\psi(t_{j+1}^{-})=1, where ψ(tj)<0\psi(t_{j})<0, which holds since d(tj)=0d(t_{j})=0. Here tj+1t_{j+1}^{-} is the left limit at t=tj+1t=t_{j+1}. So, by the intermediate value theorem, there exists a tj>tjt_{j}^{{}^{\prime}}>t_{j} such that ψ(tj)=0\psi(t^{\prime}_{j})=0 and ψ(t)[0,1]\psi(t)\in[0,1] for t[tj,tj+1]t\in[t_{j}^{{}^{\prime}},t_{j+1}^{-}]. The time derivative of ψ\psi on [tj,tj+1)[t_{j}^{{}^{\prime}},t_{j+1}) is given by

ψ˙(t)=2d(t)d˙(t)+(1σ)m˙(t)σm(t)m˙(t)m(t)ψ(t).\dot{\psi}(t)=\frac{2d(t)\dot{d}(t)+(1-\sigma)\dot{m}(t)}{-\sigma m(t)}-\frac{\dot{m}(t)}{m(t)}\psi(t). (46)

From Young’s inequality, we have that

ψ˙(t)d2(t)+d˙2(t)+(1σ)m˙(t)σm(t)m˙(t)m(t)ψ(t).\dot{\psi}(t)\leq\frac{d^{2}(t)+\dot{d}^{2}(t)+(1-\sigma)\dot{m}(t)}{-\sigma m(t)}-\frac{\dot{m}(t)}{m(t)}\psi(t). (47)

Using Lemma 2 and (28), we can show that

ψ˙(t)(1+ρ1+(1σ)ρ)d2(t)σm(t)+(1σ)ησ+ηψ(t)+σρd2(t)σm(t)ψ(t)+(α1(1σ)β1)w^[t]2σm(t)+(α2(1σ)β2)|w^(1,t)|2σm(t)+(α3(1σ)β3)|w~(0,t)|2σm(t)+β1w^[t]2+β2|w^(1,t)|2+β3|w~(0,t)|2m(t)ψ(t).\begin{split}&\dot{\psi}(t)\leq\frac{\Big{(}1+\rho_{1}+(1-\sigma)\rho\Big{)}d^{2}(t)}{-\sigma m(t)}+\frac{(1-\sigma)\eta}{\sigma}\\ &+\eta\psi(t)+\frac{\sigma\rho d^{2}(t)}{-\sigma m(t)}\psi(t)+\frac{\big{(}\alpha_{1}-(1-\sigma)\beta_{1}\big{)}\|\hat{w}[t]\|^{2}}{-\sigma m(t)}\\ &+\frac{\big{(}\alpha_{2}-(1-\sigma)\beta_{2}\big{)}|\hat{w}(1,t)|^{2}}{-\sigma m(t)}\\ &+\frac{\big{(}\alpha_{3}-(1-\sigma)\beta_{3}\big{)}|\tilde{w}(0,t)|^{2}}{-\sigma m(t)}\\ &+\frac{\beta_{1}\|\hat{w}[t]\|^{2}+\beta_{2}|\hat{w}(1,t)|^{2}+\beta_{3}|\tilde{w}(0,t)|^{2}}{m(t)}\psi(t).\end{split} (48)

Let us choose β1,β2,β3\beta_{1},\beta_{2},\beta_{3} as in (44), where α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3} are given by (41)-(43), respectively. Also, note that the last term in the right hand side of (48) is negative. Therefore, we have

ψ˙(t)(1+ρ1+(1σ)ρ)d2(t)σm(t)+(1σ)ησ+ηψ(t)+σρd2(t)σm(t)ψ(t).\begin{split}\dot{\psi}(t)\leq&\frac{\Big{(}1+\rho_{1}+(1-\sigma)\rho\Big{)}d^{2}(t)}{-\sigma m(t)}+\frac{(1-\sigma)\eta}{\sigma}\\ &+\eta\psi(t)+\frac{\sigma\rho d^{2}(t)}{-\sigma m(t)}\psi(t).\end{split} (49)

We also can write that

ψ˙(t)(1+ρ1+(1σ)ρ)(d2(t)+(1σ)m(t))σm(t)+(1σ)ησ+ηψ(t)+σρd2(t)+(1σ)m(t)σm(t)ψ(t)+(1+ρ1+(1σ)ρ)(1σ)σ+ρ(1σ)ψ(t).\begin{split}&\dot{\psi}(t)\leq\Big{(}1+\rho_{1}+(1-\sigma)\rho\Big{)}\frac{\Big{(}d^{2}(t)+(1-\sigma)m(t)\Big{)}}{-\sigma m(t)}\\ &+\frac{(1-\sigma)\eta}{\sigma}+\eta\psi(t)+\sigma\rho\frac{d^{2}(t)+(1-\sigma)m(t)}{-\sigma m(t)}\psi(t)\\ &+\Big{(}1+\rho_{1}+(1-\sigma)\rho\Big{)}\frac{(1-\sigma)}{\sigma}+\rho(1-\sigma)\psi(t).\end{split} (50)

We can rewrite (50) as

ψ˙(t)a1ψ2(t)+a2ψ(t)+a3,\dot{\psi}(t)\leq a_{1}\psi^{2}(t)+a_{2}\psi(t)+a_{3}, (51)

where

a1\displaystyle a_{1} =σρ>0,\displaystyle=\sigma\rho>0, (52)
a2\displaystyle a_{2} =1+ρ1+2(1σ)ρ+η>0,\displaystyle=1+\rho_{1}+2(1-\sigma)\rho+\eta>0, (53)
a3\displaystyle a_{3} =(1+ρ1+(1σ)ρ+η)1σσ>0.\displaystyle=\big{(}1+\rho_{1}+(1-\sigma)\rho+\eta\big{)}\frac{1-\sigma}{\sigma}>0. (54)

By the Comparison principle, it follows that the time needed for ψ\psi to go from ψ(tj)=0\psi(t_{j}^{{}^{\prime}})=0 to ψ(tj+1)=1\psi(t_{j+1}^{-})=1 is at least

τ=011a1s2+a2s+a3𝑑s.\tau=\int_{0}^{1}\frac{1}{a_{1}s^{2}+a_{2}s+a_{3}}ds. (55)

Therefore, tj+1tjτt_{j+1}-t_{j}^{{}^{\prime}}\geq\tau. As tj+1tjtj+1tjt_{j+1}-t_{j}\geq t_{j+1}-t_{j}^{{}^{\prime}}, we can conclude that tj+1tjτt_{j+1}-t_{j}\geq\tau. Thus, τ\tau can be considered a lower bound for the minimal dwell-time. Note that τ\tau is independent of initial conditions and only depends on system and control parameters.

Corollary 1:

For every given initial data u[0],u^[0]L2(0,1)u[0],\hat{u}[0]\in L^{2}(0,1), there exist unique mappings u,u^C0(+;L2(0,1))C1(I×[0,1])u,\hat{u}\in C^{0}(\mathbb{R_{+}};L^{2}(0,1))\cap C^{1}(I\times[0,1]) with u[t],u^[t]C2([0,1])u[t],\hat{u}[t]\in C^{2}([0,1]) which satisfy (1b),(3b),(20)-(22) for all t>0t>0 and (1a), (3a) for all t>0t>0, x(0,1)x\in(0,1), where I=+\{tj0,j}I=\mathbb{R_{+}}\text{\textbackslash}\{t_{j}\geq 0,j\in\mathbb{N}\}. The increasing sequence {tj0,j}\{t_{j}\geq 0,j\in\mathbb{N}\} is determined by the set of rules given in Definition 1.

Proof: This is a straightforward consequence of Proposition 4 and Theorem 4.10 in [37]. The solutions are constructed iteratively between consecutive triggering times.

Theorem 2:

Let η>0\eta>0 be a design parameter, σ(0,1)\sigma\in(0,1), and g(x)g(x) and rr be given by (16) and (17), respectively, while β1,β2,β3\beta_{1},\beta_{2},\beta_{3} are chosen according to (44). Further, subject to Assumption 1, let us choose parameters B,κ1,κ2,κ3>0B,\kappa_{1},\kappa_{2},\kappa_{3}>0 such that

B(εmin{r12,12}ε2κ15λ8κ2g2κ3)2β1β2>0,\begin{split}B\bigg{(}\varepsilon\min\Big{\{}r-\frac{1}{2},\frac{1}{2}\Big{\}}-\frac{\varepsilon}{2\kappa_{1}}&-\frac{5\lambda}{8\kappa_{2}}-\frac{\|g\|^{2}}{\kappa_{3}}\bigg{)}\\ &-2\beta_{1}-\beta_{2}>0,\end{split} (56)

AA such that

Aεmin{q12,12}5λκ2B85κ3B45β32>0,A\varepsilon\min\Big{\{}q-\frac{1}{2},\frac{1}{2}\Big{\}}-\frac{5\lambda\kappa_{2}B}{8}-\frac{5\kappa_{3}B}{4}-\frac{5\beta_{3}}{2}>0, (57)

and ρ\rho as

ρ=εκ1B2.\rho=\frac{\varepsilon\kappa_{1}B}{2}. (58)

Then, the closed-loop system which consists of the plant (1a),(1b),(21) and the observer (3a),(3b),(22) with the event-triggered boundary controller (26)-(29) has a unique solution and globally exponentially converges to zero, i.e., u[t]+u^[t]0\|u[t]\|+\|\hat{u}[t]\|\rightarrow 0 as t.t\rightarrow\infty.

Proof: From Corollary 1, the existence and the uniqueness of solutions to the plant (1a),(1b),(21) and the observer (3a),(3b),(22) are guaranteed. Now let us show that the closed-loop system is globally L2L^{2}-exponentially convergent to zero.

Let us choose the following candidate Lyapunov function noting that m(t)<0m(t)<0 for t0t\geq 0:

V=A201w~2(x,t)𝑑x+B201w^2(x,t)𝑑xm(t).V=\frac{A}{2}\int_{0}^{1}\tilde{w}^{2}(x,t)dx+\frac{B}{2}\int_{0}^{1}\hat{w}^{2}(x,t)dx-m(t). (59)

Here w^\hat{w} and w~\tilde{w} are the systems described by (24) and (25), respectively. We can show that for t(tj,tj+1),jt\in(t_{j},t_{j+1}),j\in\mathbb{N}

V˙=Aεqw~2(1,t)Aε01w~x2(x,t)𝑑xrεBw^2(1,t)+εBd(t)w^(1,t)+λB2w^(0,t)w~(0,t)εB01w^x2(x,t)𝑑x+B01g(x)w^(x,t)𝑑xw~(0,t)m˙(t).\begin{split}\dot{V}=&-A\varepsilon q\tilde{w}^{2}(1,t)-A\varepsilon\int_{0}^{1}\tilde{w}^{2}_{x}(x,t)dx\\ &-r\varepsilon B\hat{w}^{2}(1,t)+\varepsilon Bd(t)\hat{w}(1,t)+\frac{\lambda B}{2}\hat{w}(0,t)\tilde{w}(0,t)\\ &-\varepsilon B\int_{0}^{1}\hat{w}_{x}^{2}(x,t)dx+B\int_{0}^{1}g(x)\hat{w}(x,t)dx\tilde{w}(0,t)\\ &-\dot{m}(t).\end{split} (60)

From Young’s and Cauchy-Schwarz inequalities, we can write that

εBd(t)w^(1,t)εB2κ1w^2(1,t)+εκ1B2d2(t),\varepsilon Bd(t)\hat{w}(1,t)\leq\frac{\varepsilon B}{2\kappa_{1}}\hat{w}^{2}(1,t)+\frac{\varepsilon\kappa_{1}B}{2}d^{2}(t), (61)
λB2w^(0,t)w~(0,t)λB4κ2w^2(0,t)+λκ2B4w~2(0,t),\frac{\lambda B}{2}\hat{w}(0,t)\tilde{w}(0,t)\leq\frac{\lambda B}{4\kappa_{2}}\hat{w}^{2}(0,t)+\frac{\lambda\kappa_{2}B}{4}\tilde{w}^{2}(0,t), (62)
B01g(x)w^(x,t)𝑑xw~(0,t)g2B2κ3w^[t]2+κ3B2w~2(0,t).\begin{split}B\int_{0}^{1}g(x)\hat{w}(x,t)dx\tilde{w}(0,t)\leq&\frac{\|g\|^{2}B}{2\kappa_{3}}\|\hat{w}[t]\|^{2}+\frac{\kappa_{3}B}{2}\tilde{w}^{2}(0,t).\end{split} (63)

Therefore, using (61)-(63),(28),(58), we can write (60) as

V˙Aεqw~2(1,t)Aεw~x[t]2rεBw^2(1,t)+εB2κ1w^2(1,t)+λB4κ2w^2(0,t)+λκ2B4w~2(0,t)εBw^x[t]2+g2B2κ3w^[t]2+κ3B2w~2(0,t)+β1w^[t]2+β2w^2(1,t)+β3w~2(0,t)+ηm(t).\begin{split}\dot{V}\leq&-A\varepsilon q\tilde{w}^{2}(1,t)-A\varepsilon\|\tilde{w}_{x}[t]\|^{2}-r\varepsilon B\hat{w}^{2}(1,t)\\ &+\frac{\varepsilon B}{2\kappa_{1}}\hat{w}^{2}(1,t)+\frac{\lambda B}{4\kappa_{2}}\hat{w}^{2}(0,t)+\frac{\lambda\kappa_{2}B}{4}\tilde{w}^{2}(0,t)\\ &-\varepsilon B\|\hat{w}_{x}[t]\|^{2}+\frac{\|g\|^{2}B}{2\kappa_{3}}\|\hat{w}[t]\|^{2}+\frac{\kappa_{3}B}{2}\tilde{w}^{2}(0,t)\\ &+\beta_{1}\|\hat{w}[t]\|^{2}+\beta_{2}\hat{w}^{2}(1,t)\\ &+\beta_{3}\tilde{w}^{2}(0,t)+\eta m(t).\end{split} (64)

From Agmon’s and Young’s inequalities, we have that

w~2(0,t)\displaystyle\tilde{w}^{2}(0,t) w~2(1,t)+w~[t]2+w~x[t]2,\displaystyle\leq\tilde{w}^{2}(1,t)+\|\tilde{w}[t]\|^{2}+\|\tilde{w}_{x}[t]\|^{2}, (65)
w^2(0,t)\displaystyle\hat{w}^{2}(0,t) w^2(1,t)+w^[t]2+w^x[t]2.\displaystyle\leq\hat{w}^{2}(1,t)+\|\hat{w}[t]\|^{2}+\|\hat{w}_{x}[t]\|^{2}. (66)

Therefore, we can show from (64) that

V˙(Aεqλκ2B4κ3B2β3)w~2(1,t)(Aελκ2B4κ3B2β3)w~x[t]2+(λκ2B4+κ3B2+β3)w~[t]2(rεBεB2κ1λB4κ2β2)w^2(1,t)(εBλB4κ2)w^x[t]2+(λB4κ2+g2B2κ3+β1)w^[t]2+ηm(t).\begin{split}\dot{V}\leq&-\Big{(}A\varepsilon q-\frac{\lambda\kappa_{2}B}{4}-\frac{\kappa_{3}B}{2}-\beta_{3}\Big{)}\tilde{w}^{2}(1,t)\\ &-\Big{(}A\varepsilon-\frac{\lambda\kappa_{2}B}{4}-\frac{\kappa_{3}B}{2}-\beta_{3}\Big{)}\|\tilde{w}_{x}[t]\|^{2}\\ &+\Big{(}\frac{\lambda\kappa_{2}B}{4}+\frac{\kappa_{3}B}{2}+\beta_{3}\Big{)}\|\tilde{w}[t]\|^{2}\\ &-\Big{(}r\varepsilon B-\frac{\varepsilon B}{2\kappa_{1}}-\frac{\lambda B}{4\kappa_{2}}-\beta_{2}\Big{)}\hat{w}^{2}(1,t)\\ &-\Big{(}\varepsilon B-\frac{\lambda B}{4\kappa_{2}}\Big{)}\|\hat{w}_{x}[t]\|^{2}\\ &+\Big{(}\frac{\lambda B}{4\kappa_{2}}+\frac{\|g\|^{2}B}{2\kappa_{3}}+\beta_{1}\Big{)}\|\hat{w}[t]\|^{2}+\eta m(t).\end{split} (67)

From Poincaré Inequality, we have that

w~x[t]2\displaystyle-\|\tilde{w}_{x}[t]\|^{2} 12w~2(1,t)14w~[t]2,\displaystyle\leq\frac{1}{2}\tilde{w}^{2}(1,t)-\frac{1}{4}\|\tilde{w}[t]\|^{2}, (68)
w^x[t]2\displaystyle-\|\hat{w}_{x}[t]\|^{2} 12w^2(1,t)14w^[t]2.\displaystyle\leq\frac{1}{2}\hat{w}^{2}(1,t)-\frac{1}{4}\|\hat{w}[t]\|^{2}. (69)

Furthermore, we have from (56) and (57) that

Aελκ2B4κ3B2β3>0, and εBλB4κ2>0.A\varepsilon-\frac{\lambda\kappa_{2}B}{4}-\frac{\kappa_{3}B}{2}-\beta_{3}>0,\text{ and }\varepsilon B-\frac{\lambda B}{4\kappa_{2}}>0. (70)

Therefore, using (67)-(70), we can show that

V˙(Aε(q12)λκ2B8κ3B4β32)w~2(1,t)(Aε45λκ2B165κ3B85β34)w~[t]2(εB(r12)εB2κ1λB8κ2β2)w^2(1,t)(εB45λB16κ2g2B2κ3β1)w^[t]2+ηm(t).\begin{split}\dot{V}\leq&-\Big{(}A\varepsilon(q-\frac{1}{2})-\frac{\lambda\kappa_{2}B}{8}-\frac{\kappa_{3}B}{4}-\frac{\beta_{3}}{2}\Big{)}\tilde{w}^{2}(1,t)\\ &-\Big{(}\frac{A\varepsilon}{4}-\frac{5\lambda\kappa_{2}B}{16}-\frac{5\kappa_{3}B}{8}-\frac{5\beta_{3}}{4}\Big{)}\|\tilde{w}[t]\|^{2}\\ &-\Big{(}\varepsilon B(r-\frac{1}{2})-\frac{\varepsilon B}{2\kappa_{1}}-\frac{\lambda B}{8\kappa_{2}}-\beta_{2}\Big{)}\hat{w}^{2}(1,t)\\ &-\Big{(}\frac{\varepsilon B}{4}-\frac{5\lambda B}{16\kappa_{2}}-\frac{\|g\|^{2}B}{2\kappa_{3}}-\beta_{1}\Big{)}\|\hat{w}[t]\|^{2}+\eta m(t).\end{split} (71)

From (56) and (57), we can show that

Aε(q12)λκ2B8κ3B4β32\displaystyle A\varepsilon(q-\frac{1}{2})-\frac{\lambda\kappa_{2}B}{8}-\frac{\kappa_{3}B}{4}-\frac{\beta_{3}}{2} >0,\displaystyle>0, (72)
εB(r12)εB2κ1λB8κ2β2\displaystyle\varepsilon B(r-\frac{1}{2})-\frac{\varepsilon B}{2\kappa_{1}}-\frac{\lambda B}{8\kappa_{2}}-\beta_{2} >0.\displaystyle>0. (73)

Thus, using (71)-(73), we can obtain that

V˙(Aε45λκ2B165κ3B85β34)w~[t]2(εB45λB16κ2g2B2κ3β1)w^[t]2+ηm(t).\begin{split}&\dot{V}\leq-\Big{(}\frac{A\varepsilon}{4}-\frac{5\lambda\kappa_{2}B}{16}-\frac{5\kappa_{3}B}{8}-\frac{5\beta_{3}}{4}\Big{)}\|\tilde{w}[t]\|^{2}\\ &-\Big{(}\frac{\varepsilon B}{4}-\frac{5\lambda B}{16\kappa_{2}}-\frac{\|g\|^{2}B}{2\kappa_{3}}-\beta_{1}\Big{)}\|\hat{w}[t]\|^{2}+\eta m(t).\end{split} (74)

Again, we have from (56) and (57) that

b1\displaystyle b_{1} =Aε45λκ2B165κ3B85β34>0,\displaystyle=\frac{A\varepsilon}{4}-\frac{5\lambda\kappa_{2}B}{16}-\frac{5\kappa_{3}B}{8}-\frac{5\beta_{3}}{4}>0, (75)
b2\displaystyle b_{2} =εB45λB16κ2g2B2κ3β1>0.\displaystyle=\frac{\varepsilon B}{4}-\frac{5\lambda B}{16\kappa_{2}}-\frac{\|g\|^{2}B}{2\kappa_{3}}-\beta_{1}>0. (76)

Therefore, we have for t(tj,tj+1),jt\in(t_{j},t_{j+1}),j\in\mathbb{N} that

V˙ϱV,\dot{V}\leq-\varrho V, (77)

where

ϱ=min{b1,b2,η}max{A/2,B/2,1}.\varrho=\frac{\min\big{\{}b_{1},b_{2},\eta\big{\}}}{\max\big{\{}A/2,B/2,1\big{\}}}. (78)

Concentrating on this time interval, we can show that V(tj+1)eϱ(tj+1tj+)V(tj+)V(t^{-}_{j+1})\leq e^{-\varrho(t_{j+1}^{-}-t_{j}^{+})}V(t^{+}_{j}). Here tj+t_{j}^{+} and tjt_{j}^{-} are the right and left limits of t=tjt=t_{j}. Since V(t)V(t) is continuous (as m(t),w^[t]m(t),\|\hat{w}[t]\|,w~[t]\|\tilde{w}[t]\| are continuous), we have that V(tj+1)=V(tj+1)V(t^{-}_{j+1})=V(t_{j+1}) and V(tj+)=V(tj),V(t_{j}^{+})=V(t_{j}), and therefore,

V(tj+1)eϱ(tj+1tj)V(tj).V(t_{j+1})\leq e^{-\varrho(t_{j+1}-t_{j})}V(t_{j}). (79)

Hence, for any t0t\geq 0 in t[tj,tj+1),j,t\in[t_{j},t_{j+1}),j\in\mathbb{N}, we can also obtain the following:

V(t)eϱ(ttj)V(tj)eϱ(ttj)×eϱ(tjtj1)V(tj1)eϱ(ttj)×i=1i=jeϱ(titi1)V(0)=eϱtV(0).\begin{split}V(t)&\leq e^{-\varrho(t-t_{j})}V(t_{j})\\ &\leq e^{-\varrho(t-t_{j})}\times e^{-\varrho(t_{j}-t_{j-1})}V(t_{j-1})\\ &\leq\cdots\leq\\ &\leq e^{-\varrho(t-t_{j})}\times\prod_{i=1}^{i=j}e^{-\varrho(t_{i}-t_{i-1})}V(0)\\ &=e^{-\varrho t}V(0).\end{split} (80)

Thus, recalling that m(0)<0m(0)<0 from Definition 1, we have that

A2w~[t]2+B2w^[t]2m(t)eϱt(A2w~[0]2+B2w^[0]2m(0)).\begin{split}\frac{A}{2}\|\tilde{w}[t]\|^{2}&+\frac{B}{2}\|\hat{w}[t]\|^{2}-m(t)\\ &\leq e^{-\varrho t}\Big{(}\frac{A}{2}\|\tilde{w}[0]\|^{2}+\frac{B}{2}\|\hat{w}[0]\|^{2}-m(0)\Big{)}.\end{split} (81)

As m(t)<0m(t)<0, we also have that

A2w~[t]2+B2w^[t]2eϱt(A2w~[0]2+B2w^[0]2m(0)).\begin{split}\frac{A}{2}\|\tilde{w}[t]\|^{2}&+\frac{B}{2}\|\hat{w}[t]\|^{2}\\ &\leq e^{-\varrho t}\Big{(}\frac{A}{2}\|\tilde{w}[0]\|^{2}+\frac{B}{2}\|\hat{w}[0]\|^{2}-m(0)\Big{)}.\end{split} (82)

This implies that the systems w^\hat{w} and w~\tilde{w} given by (24) and (25), respectively, are globally exponentially convergent in L2L^{2}-sense to zero. Using (6) and (18), we can show that u~[t]2P~2w~[t]2\|\tilde{u}[t]\|^{2}\leq\tilde{P}^{2}\|\tilde{w}[t]\|^{2} and u^[t]2L~2w^[t]2\|\hat{u}[t]\|^{2}\leq\tilde{L}^{2}\|\hat{w}[t]\|^{2}, respectively. Here P~=(1+(010xP2(x,y)𝑑y𝑑x)1/2)2\tilde{P}=\bigg{(}1+\Big{(}\int_{0}^{1}\int_{0}^{x}P^{2}(x,y)dydx\Big{)}^{1/2}\bigg{)}^{2} and L~=(1+(010xL2(x,y)𝑑y𝑑x)1/2)2\tilde{L}=\bigg{(}1+\Big{(}\int_{0}^{1}\int_{0}^{x}L^{2}(x,y)dydx\Big{)}^{1/2}\bigg{)}^{2}. Further from (4), we can show that u[t]22u~[t]2+2u^[t]2\|u[t]\|^{2}\leq 2\|\tilde{u}[t]\|^{2}+2\|\hat{u}[t]\|^{2}. Therefore, we can obtain from (82) that

min{AP~,BL~}u[t]24eϱt(A2w~[0]2+B2w^[0]2m(0)),\min\Big{\{}\frac{A}{\tilde{P}},\frac{B}{\tilde{L}}\Big{\}}\|u[t]\|^{2}\leq 4e^{-\varrho t}\Big{(}\frac{A}{2}\|\tilde{w}[0]\|^{2}+\frac{B}{2}\|\hat{w}[0]\|^{2}-m(0)\Big{)}, (83)
BL~u^[t]22eϱt(A2w~[0]2+B2w^[0]2m(0)).\frac{B}{\tilde{L}}\|\hat{u}[t]\|^{2}\leq 2e^{-\varrho t}\Big{(}\frac{A}{2}\|\tilde{w}[0]\|^{2}+\frac{B}{2}\|\hat{w}[0]\|^{2}-m(0)\Big{)}. (84)

Therefore, we can derive the following estimate:

u[t]+u^[t]12(A2w~[0]2+B2w^[0]2m(0))min{AP~,BL~}eϱt2,\begin{split}\|u[t]\|&+\|\hat{u}[t]\|\\ &\leq\sqrt{\frac{12\Big{(}\frac{A}{2}\|\tilde{w}[0]\|^{2}+\frac{B}{2}\|\hat{w}[0]\|^{2}-m(0)\Big{)}}{\min\Big{\{}\frac{A}{\tilde{P}},\frac{B}{\tilde{L}}\Big{\}}}}e^{-\frac{\varrho t}{2}},\end{split} (85)

which implies that u[t]+u^[t]0\|u[t]\|+\|\hat{u}[t]\|\rightarrow 0 as tt\rightarrow\infty.

Remark 2:

In Theorem 2, we have established the global exponential convergence of the closed-loop system to the equilibrium point. It follows from (82) that we could have obtained global exponential stability if we chose m(0)=0m(0)=0. However, if m(0)=0m(0)=0, then m(t)0m(t)\leq 0 (this can be shown by following the same arguments in the proof of Lemma 1). Then, the function ψ(t)\psi(t) in (45) is not defined when m(t)=0m(t)=0. Therefore, the existence of a minimal-dwell time cannot be proved easily by following the same arguments as in the proof of Theorem 1. Hence, m(0)m(0) has to be chosen strictly negative.

Remark 3:

The parameter η>0\eta>0 characterizes the decay rate of m(t)m(t) governed by (28). Thus, η\eta may be used to adjust the sampling speed of the event-triggered mechanism. The larger η\eta, the faster is the sampling speed. We consider σ(0,1)\sigma\in(0,1) as a free parameter that can be tuned appropriately such that the conditions for guaranteeing a minimal dwell-time are met.

Remark 4:

We remark that if a periodic sampling scheme where the control value is periodically updated in a sampled-and-hold manner is to be used to stabilize the plant (1) and the observer (3), one can choose a sampling period TT upper bounded by the minimal dwell-time τ\tau (55). It will ensure the closed-loop system’s global exponential convergence because the relation (80) is guaranteed to hold for all TτT\leq\tau. However, one should expect τ\tau to be very small and conservative as the coefficients a1,a2,a_{1},a_{2}, and a3a_{3} (52)-(54) are usually large. This issue, on the other hand, reinforces the motivation for ETC, that is sample and update only when required.

V Numerical Simulations

We consider a reaction-diffusion system with ε=0.1;λ=0.25;q=2.3\varepsilon=0.1;\lambda=0.25;q=2.3 and the initial conditions u[0]=10x2(x1)2u[0]=10x^{2}(x-1)^{2} and u^[0]=5x2(x1)2+5x3(x1)3\hat{u}[0]=5x^{2}(x-1)^{2}+5x^{3}(x-1)^{3}. For numerical simulations, both plant and observer states are discretized with uniform step size of h=0.0062h=0.0062 for the space variable. Time discretization was done using the implicit Euler scheme with a step size Δt=h\Delta t=h.

The parameters for the event-trigger mechanism are chosen as follows: m(0)=104,η=1 or 100m(0)=-10^{-4},\eta=1\text{ or }100 and σ=0.1\sigma=0.1. It can be shown using (41)-(43) that α1=4.14;α2=2.07;α3=3.3\alpha_{1}=4.14;\alpha_{2}=2.07;\alpha_{3}=3.3. Therefore, from (44), we can obtain β1=4.6;β2=2.3;β3=3.7\beta_{1}=4.6;\beta_{2}=2.3;\beta_{3}=3.7. Finding that g2=0.0297\|g\|^{2}=0.0297, let us choose κ1=2.1;κ2=312.5;κ3=59.4\kappa_{1}=2.1;\kappa_{2}=312.5;\kappa_{3}=59.4 and B=460B=460 to satisfy (56). Then, from (58), we can obtain ρ=48.3\rho=48.3.

Fig. 2 shows the zero-input response of the plant and it is clear that the system is unstable. Fig. 3(a) shows the response of the pant under ETC with η=1\eta=1 and Fig. 3(b) shows the evolution of u[t],u^[t]\|u[t]\|,\|\hat{u}[t]\| and u~[t]\|\tilde{u}[t]\|. The evolution of the control input when η=1\eta=1 is presented in Fig. 4(a) along with the corresponding continuous-time control input. The behavior of the functions associated with the triggering condition (26) for the case of η=1\eta=1 is depicted in Fig. 4(b). Once the trajectory d2(t)d^{2}(t) reaches the trajectory m(t)-m(t), an event is generated, the control input is updated and d(t)d(t) is reset to zero. Fig. 5 compares the ETC control input for η=1\eta=1 and η=100\eta=100, and it can be seen that η=100\eta=100 results in faster sampling than η=1\eta=1.

Finally, we conduct simulations for 100100 different initial conditions u[0]=x2(x1)2sin(nπx),n=1,,100u[0]=x^{2}(x-1)^{2}\sin(n\pi x),n=1,\ldots,100 and u^[0]=2u[t]\hat{u}[0]=2u[t] on a time frame of 150 s150\text{ }s. Next, we compute the inter-execution times between two events and compare the cases for slow and fast sampling, i.e., η=1\eta=1 and η=100\eta=100, respectively. Fig. 6 shows the density of the inter-execution times plotted in logarithmic scale, and it can be stated that when η\eta is smaller, the inter-executions times are larger and the sampling is less often. For η=1\eta=1 and η=100\eta=100, the inter-execution times are typically in the range 0.1 s10 s0.1\text{ }s-10\text{ }s. For the system under consideration, the minimal dwell time τ\tau calculated using (55) when η=1\eta=1 and η=100\eta=100 are respectively 2.2×103 s2.2\times 10^{-3}\text{ }s and 7.2×104 s7.2\times 10^{-4}\text{ }s. Therefore, the fact that an analogous sampled-data controller guaranteeing exponential convergence has to be implemented using these small and conservative sampled periods manifests the importance and the need of ETC.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: Results for open-loop plant with ε=0.1,λ=0.25,q=2.3\varepsilon=0.1,\lambda=0.25,q=2.3 and u[0]=10x2(x1)2u[0]=10x^{2}(x-1)^{2}. (a) u(x,t).u(x,t). (b) u[t].\|u[t]\|.
Refer to caption
(a)
Refer to caption
(b)
Figure 3: Results for the event-triggered closed-loop system with ε=0.1,λ=0.25,q=2.3,m(0)=104,η=1\varepsilon=0.1,\lambda=0.25,q=2.3,m(0)=-10^{-4},\eta=1, u[0]=10x2(x1)2u[0]=10x^{2}(x-1)^{2} and u^[0]=5x2(x1)2+5x3(x1)3\hat{u}[0]=5x^{2}(x-1)^{2}+5x^{3}(x-1)^{3} (a) u(x,t)u(x,t). (b) u[t],u^[t] and u~[t].\|u[t]\|,\|\hat{u}[t]\|\text{ and }\|\tilde{u}[t]\|.
Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a) The event-triggered control (ETC) input for the system considered in Fig. 3 along with the corresponding continuous-time control (CTC) input. (b) Trajectories involved in the triggering condition (26) for the system in Fig. 3.
Refer to caption
Figure 5: Comparison of ETC input U(t)U(t) for different η\eta: η=1 and η=100\eta=1\text{ and }\eta=100, for the same system considered in Fig. 3.
Refer to caption
Figure 6: Density of the inter-execution times (logarithmic scale) computed for 100 different initial conditions: u[0]=x2(x1)2sin(nπx),n=1,,100u[0]=x^{2}(x-1)^{2}\sin(n\pi x),n=1,\ldots,100 and u^[0]=2u[t]\hat{u}[0]=2u[t] on a time frame of 150 s150\text{ }s, for the same system considered in Fig. 3 with η=1\eta=1 and η=100\eta=100.

VI Conclusion

This paper has proposed an event-triggered output feedback boundary control strategy for a class of reaction-diffusion systems with Robin boundary actuation. We have used a dynamic event triggering condition to determine when the control value needs to be updated. Under the proposed strategy, we have proved the existence of a minimal-dwell time between two updates independent of initial conditions, which excludes Zeno behavior. Further, we have shown the well-posedness of the closed-loop system and its global L2L^{2}-exponential convergence to the equilibrium point.

We may consider periodic event-triggered boundary control (PETC) of reaction-diffusion systems under full-state and output feedback settings in future work. The idea is to evaluate the triggering condition periodically and to decide, at every sampling instant, whether the feedback loop needs to be closed. PETC is highly desirable as it not only guarantees a minimal dwell-time equal to the sampling period but also provides a more realistic approach toward digital implementations while reducing the utilization of computational resources.

Appendix A
Proof of Proposition 1

Considering Remark 1, it can be deduced that the solution of target system (9) satisfies w~[t]eεμ2tw~[0]\|\tilde{w}[t]\|\leq e^{-\varepsilon\mu^{2}t}\|\tilde{w}[0]\|, which implies the global exponential stability in L2L^{2}-sense.

Let us show the gain kernel P(x,y)P(x,y) and the observer gains p1(x)p_{1}(x) and p10p_{10}, which transform (5) into (9) via (6), are given by (7) and (8), respectively. This proves Proposition 1.

Taking the time derivative of (6) along the solutions of (5) and applying integration by parts twice, we can show that

u~t(x,t)=w~t(x,t)εP(x,x)w~x(x,t)+εP(x,0)w~x(0,t)+εPy(x,x)w~(x,t)εPy(x,0)w~(0,t)ε0xPyy(x,y)w~(y,t)𝑑y.\begin{split}\tilde{u}_{t}(x,t)=&\tilde{w}_{t}(x,t)-\varepsilon P(x,x)\tilde{w}_{x}(x,t)\\ &+\varepsilon P(x,0)\tilde{w}_{x}(0,t)+\varepsilon P_{y}(x,x)\tilde{w}(x,t)\\ &-\varepsilon P_{y}(x,0)\tilde{w}(0,t)-\varepsilon\int_{0}^{x}P_{yy}(x,y)\tilde{w}(y,t)dy.\end{split} (86)

Differentiating (6) w.r.t xx and using Leibnitz differentiation rule, we can obtain that

u~x(x,t)=w~x(x,t)P(x,x)w~(x,t)0xPx(x,y)w~(y,t)𝑑y,\begin{split}\tilde{u}_{x}(x,t)=\tilde{w}_{x}(x,t)-P(x,x)\tilde{w}(x,t)-\int_{0}^{x}P_{x}(x,y)\tilde{w}(y,t)dy,\end{split} (87)
u~xx(x,t)=w~xx(x,t)dP(x,x)dxw~(x,t)P(x,x)w~x(x,t)Px(x,x)w~(x,t)0xPxx(x,y)w~(y,t)𝑑y.\begin{split}\tilde{u}_{xx}(x,t)=&\tilde{w}_{xx}(x,t)-\frac{dP(x,x)}{dx}\tilde{w}(x,t)-P(x,x)\tilde{w}_{x}(x,t)\\ &-P_{x}(x,x)\tilde{w}(x,t)-\int_{0}^{x}P_{xx}(x,y)\tilde{w}(y,t)dy.\end{split} (88)

Therefore, from (5a),(6),(86) and (88), we can show that

0=(p1(x)εPy(x,0))w~(0,t)(λ2εdP(x,x)dx)w~(x,t)+0x(εPxx(x,y)εPyy(x,y)+λP(x,y))w~(y,t)𝑑y.\begin{split}&0=\big{(}p_{1}(x)-\varepsilon P_{y}(x,0)\big{)}\tilde{w}(0,t)-\bigg{(}\lambda-2\varepsilon\frac{dP(x,x)}{dx}\bigg{)}\tilde{w}(x,t)\\ &+\int_{0}^{x}\bigg{(}\varepsilon P_{xx}(x,y)-\varepsilon P_{yy}(x,y)+\lambda P(x,y)\bigg{)}\tilde{w}(y,t)dy.\end{split} (89)

Let us choose Pxx(x,y)Pyy(x,y)=λεP(x,y),dP(x,x)dx=λ2ε, and p1(x)=εPy(x,0)P_{xx}(x,y)-P_{yy}(x,y)=-\frac{\lambda}{\varepsilon}P(x,y),\frac{dP(x,x)}{dx}=\frac{\lambda}{2\varepsilon},\text{ and }p_{1}(x)=\varepsilon P_{y}(x,0) so that (89) is valid for any w~\tilde{w}. Further, let us choose p10=P(0,0)p_{10}=P(0,0) so that the boundary conditions (5b) and (9b) are satisfied, and choose P(1,1)=0P(1,1)=0 and Px(1,y)=qP(1,y)P_{x}(1,y)=-qP(1,y) so that the conditions (5c) and (9c) are met. Therefore, the gain kernel P(x,y)P(x,y) in (6) as a whole should satisfy the following PDE:

Pxx(x,y)Pyy(x,y)\displaystyle P_{xx}(x,y)-P_{yy}(x,y) =λεP(x,y),\displaystyle=-\frac{\lambda}{\varepsilon}P(x,y), (90a)
Px(1,y)\displaystyle P_{x}(1,y) =qP(1,y),\displaystyle=-qP(1,y), (90b)
P(x,x)\displaystyle P(x,x) =λ2ε(x1).\displaystyle=\frac{\lambda}{2\varepsilon}(x-1). (90c)

It can be shown that the change of variables x=1y¯x=1-\bar{y} and y=1x¯y=1-\bar{x} on (90) leads to the same PDE (108-110) in [38] to which the explicit solution has been obtained. Therefore, the solution to (90) can be shown to be given by (7). Above we have obtained p1(x)=εPy(x,0)p_{1}(x)=\varepsilon P_{y}(x,0) and p10=P(0,0)p_{10}=P(0,0), which are the same as stated in Proposition 1.

Appendix B
Proof of Proposition 2

Let us show that the gain kernel K(x,y)K(x,y) and the control law U(t)U(t), which transform (3) into (15)-(17) via (12), are given by (13) and (14), respectively. This proves Proposition 2.

Taking the time derivative of (12) along the solutions of (3) and applying integration by parts twice, we can show that

w^t(x,t)=u^t(x,t)λ0xK(x,y)u^(y,t)𝑑y0xK(x,y)p1(y)𝑑yw~(0,t)εK(x,x)u^x(x,t)+εK(x,0)u^x(0,t)+εKy(x,x)u^(x,t)εKy(x,0)u^(0,t)ε0xKyy(x,y)u^(y,t)𝑑y,\begin{split}&\hat{w}_{t}(x,t)=\hat{u}_{t}(x,t)-\lambda\int_{0}^{x}K(x,y)\hat{u}(y,t)dy\\ &-\int_{0}^{x}K(x,y)p_{1}(y)dy\tilde{w}(0,t)-\varepsilon K(x,x)\hat{u}_{x}(x,t)\\ &+\varepsilon K(x,0)\hat{u}_{x}(0,t)+\varepsilon K_{y}(x,x)\hat{u}(x,t)\\ &-\varepsilon K_{y}(x,0)\hat{u}(0,t)-\varepsilon\int_{0}^{x}K_{yy}(x,y)\hat{u}(y,t)dy,\end{split} (91)

Differentiating (12) w.r.t xx and using Leibnitz differentiation rule, we can obtain that

w^x(x,t)=u^x(x,t)K(x,x)u^(x,t)0xKx(x,y)u^(y,t)𝑑y,\begin{split}\hat{w}_{x}(x,t)=\hat{u}_{x}(x,t)-K(x,x)\hat{u}(x,t)-\int_{0}^{x}K_{x}(x,y)\hat{u}(y,t)dy,\end{split} (92)
w^xx(x,t)=u^xx(x,t)dK(x,x)dxu^(x,t)K(x,x)u^x(x,t)Kx(x,x)u^(x,t)0xKxx(x,y)u^(y,t)𝑑y.\begin{split}\hat{w}_{xx}(x,t)=&\hat{u}_{xx}(x,t)-\frac{dK(x,x)}{dx}\hat{u}(x,t)-K(x,x)\hat{u}_{x}(x,t)\\ &-K_{x}(x,x)\hat{u}(x,t)-\int_{0}^{x}K_{xx}(x,y)\hat{u}(y,t)dy.\end{split} (93)

Therefore, from (8),(12),(15a),(16),(91) and (93), we can show

0=(λ+2εdK(x,x)dx)u^(x,t)dyεKy(x,0)u^(0,t)+0x(εKxx(x,y)εKyy(x,y)λK(x,y))u^(y,t)𝑑y.\begin{split}&0=\Big{(}\lambda+2\varepsilon\frac{dK(x,x)}{dx}\Big{)}\hat{u}(x,t)dy-\varepsilon K_{y}(x,0)\hat{u}(0,t)\\ &+\int_{0}^{x}\bigg{(}\varepsilon K_{xx}(x,y)-\varepsilon K_{yy}(x,y)-\lambda K(x,y)\bigg{)}\hat{u}(y,t)dy.\end{split} (94)

Let us choose Kxx(x,y)Kyy(x,y)=λεK(x,y),Ky(x,0)=0, and, dK(x,x)dx=λ2εK_{xx}(x,y)-K_{yy}(x,y)=\frac{\lambda}{\varepsilon}K(x,y),K_{y}(x,0)=0,\text{ and, }\frac{dK(x,x)}{dx}=-\frac{\lambda}{2\varepsilon} so that (94) holds for any u^\hat{u}. Further, let us choose K(0,0)=0,K(0,0)=0, so that the boundary conditions (3b) and (15b) are met, and choose U(t)=01(rK(1,y)+Kx(1,y))u^(y,t)𝑑yU(t)=\int_{0}^{1}\Big{(}rK(1,y)+K_{x}(1,y)\Big{)}\hat{u}(y,t)dy so that the boundary conditions (3c) and (15c) are satisfied. Therefore, the gain kernel K(x,y)K(x,y) in (12) as a whole should satisfy the following PDE:

Kxx(x,y)Kyy(x,y)\displaystyle K_{xx}(x,y)-K_{yy}(x,y) =λεK(x,y),\displaystyle=\frac{\lambda}{\varepsilon}K(x,y), (95a)
Ky(x,0)\displaystyle K_{y}(x,0) =0,\displaystyle=0, (95b)
K(x,x)\displaystyle K(x,x) =λ2εx.\displaystyle=-\frac{\lambda}{2\varepsilon}x. (95c)

The solution to (95) is given by (13)[39]. Further, the control law obtained above is the same as (14).

Appendix C
Proof of Proposition 3

Subject to Assumption 1, let us choose parameters δ1,δ2>0\delta_{1},\delta_{2}>0 such that

εmin{r12,12}5λ8δ1g2δ20,\varepsilon\min\Big{\{}r-\frac{1}{2},\frac{1}{2}\Big{\}}-\frac{5\lambda}{8\delta_{1}}-\frac{\|g\|^{2}}{\delta_{2}}\geq 0, (96)

and H>0H>0 such that

Hεmin{q12,12}5λδ185δ240.H\varepsilon\min\Big{\{}q-\frac{1}{2},\frac{1}{2}\Big{\}}-\frac{5\lambda\delta_{1}}{8}-\frac{5\delta_{2}}{4}\geq 0. (97)

Here g(x)g(x) and rr are given by (16) and (17), respectively. Note that r>1/2r>1/2 due to Assumption 1. Then, let us consider the following Lyapunov candidate

𝒱=H201w~2(x,t)𝑑x+1201w^2(x,t)𝑑x,\mathcal{V}=\frac{H}{2}\int_{0}^{1}\tilde{w}^{2}(x,t)dx+\frac{1}{2}\int_{0}^{1}\hat{w}^{2}(x,t)dx, (98)

where w~\tilde{w} and w^\hat{w} are the systems described by (9) and(15), respectively. We can show that

𝒱˙=Hεqw~2(1,t)Hε01w~x2(x,t)𝑑xrεw^2(1,t)+λ2w^(0,t)w~(0,t)ε01w^x2(x,t)𝑑x+01g(x)w^(x,t)𝑑xw~(0,t).\begin{split}&\dot{\mathcal{V}}=-H\varepsilon q\tilde{w}^{2}(1,t)-H\varepsilon\int_{0}^{1}\tilde{w}^{2}_{x}(x,t)dx\\ &-r\varepsilon\hat{w}^{2}(1,t)+\frac{\lambda}{2}\hat{w}(0,t)\tilde{w}(0,t)\\ &-\varepsilon\int_{0}^{1}\hat{w}_{x}^{2}(x,t)dx+\int_{0}^{1}g(x)\hat{w}(x,t)dx\tilde{w}(0,t).\end{split} (99)

From Young’s and Cauchy-Schwarz inequalities, we can obtain that

λ2w^(0,t)w~(0,t)λ4δ1w^2(0,t)+λδ14w~2(0,t),\frac{\lambda}{2}\hat{w}(0,t)\tilde{w}(0,t)\leq\frac{\lambda}{4\delta_{1}}\hat{w}^{2}(0,t)+\frac{\lambda\delta_{1}}{4}\tilde{w}^{2}(0,t), (100)
01g(x)w^(x,t)𝑑xw~(0,t)g22δ2w^[t]2+δ22w~2(0,t).\begin{split}\int_{0}^{1}g(x)\hat{w}(x,t)dx\tilde{w}(0,t)\leq&\frac{\|g\|^{2}}{2\delta_{2}}\|\hat{w}[t]\|^{2}+\frac{\delta_{2}}{2}\tilde{w}^{2}(0,t).\end{split} (101)

Therefore, using (100) and (101), we can write (99) as

𝒱˙Hεqw~2(1,t)Hεw~x[t]2rεw^2(1,t)+λ4δ1w^2(0,t)+λδ14w~2(0,t)εw^x[t]2+g22δ2w^[t]2+δ22w~2(0,t).\begin{split}\dot{\mathcal{V}}\leq&-H\varepsilon q\tilde{w}^{2}(1,t)-H\varepsilon\|\tilde{w}_{x}[t]\|^{2}-r\varepsilon\hat{w}^{2}(1,t)\\ &+\frac{\lambda}{4\delta_{1}}\hat{w}^{2}(0,t)+\frac{\lambda\delta_{1}}{4}\tilde{w}^{2}(0,t)\\ &-\varepsilon\|\hat{w}_{x}[t]\|^{2}+\frac{\|g\|^{2}}{2\delta_{2}}\|\hat{w}[t]\|^{2}+\frac{\delta_{2}}{2}\tilde{w}^{2}(0,t).\end{split} (102)

From Agmon’s and Young’s inequalities, we have that

w~2(0,t)w~2(1,t)+w~[t]2+w~x[t]2,\tilde{w}^{2}(0,t)\leq\tilde{w}^{2}(1,t)+\|\tilde{w}[t]\|^{2}+\|\tilde{w}_{x}[t]\|^{2}, (103)
w^2(0,t)w^2(1,t)+w^[t]2+w^x[t]2.\hat{w}^{2}(0,t)\leq\hat{w}^{2}(1,t)+\|\hat{w}[t]\|^{2}+\|\hat{w}_{x}[t]\|^{2}. (104)

Therefore, we can show using (102) that

𝒱˙(Hεqλδ14δ22)w~2(1,t)(Hελδ14δ22)w~x[t]2+(λδ14+δ22)w~[t]2(rελ4δ1)w^2(1,t)(ελ4δ1)w^x[t]2+(λ4δ1+g22δ2)w^[t]2.\begin{split}\dot{\mathcal{V}}\leq&-\Big{(}H\varepsilon q-\frac{\lambda\delta_{1}}{4}-\frac{\delta_{2}}{2}\Big{)}\tilde{w}^{2}(1,t)\\ &-\Big{(}H\varepsilon-\frac{\lambda\delta_{1}}{4}-\frac{\delta_{2}}{2}\Big{)}\|\tilde{w}_{x}[t]\|^{2}\\ &+\Big{(}\frac{\lambda\delta_{1}}{4}+\frac{\delta_{2}}{2}\Big{)}\|\tilde{w}[t]\|^{2}-\Big{(}r\varepsilon-\frac{\lambda}{4\delta_{1}}\Big{)}\hat{w}^{2}(1,t)\\ &-\Big{(}\varepsilon-\frac{\lambda}{4\delta_{1}}\Big{)}\|\hat{w}_{x}[t]\|^{2}+\Big{(}\frac{\lambda}{4\delta_{1}}+\frac{\|g\|^{2}}{2\delta_{2}}\Big{)}\|\hat{w}[t]\|^{2}.\end{split} (105)

From Poincaré Inequality, we have that

w~x[t]212w~2(1,t)14w~[t]2,-\|\tilde{w}_{x}[t]\|^{2}\leq\frac{1}{2}\tilde{w}^{2}(1,t)-\frac{1}{4}\|\tilde{w}[t]\|^{2}, (106)
w^x[t]212w^2(1,t)14w^[t]2.-\|\hat{w}_{x}[t]\|^{2}\leq\frac{1}{2}\hat{w}^{2}(1,t)-\frac{1}{4}\|\hat{w}[t]\|^{2}. (107)

Furthermore, we have from (96) and (97) that

Hελδ14δ22>0 and ελ4δ1>0.H\varepsilon-\frac{\lambda\delta_{1}}{4}-\frac{\delta_{2}}{2}>0\text{ and }\varepsilon-\frac{\lambda}{4\delta_{1}}>0. (108)

Therefore, using (105)-(108), we can show that

𝒱˙(Hε(q12)λδ18δ24)w~2(1,t)(Hε45λδ1165δ28)w~[t]2(ε(r12)λ8δ1)w^2(1,t)(ε45λ16δ1g22δ2)w^[t]2.\begin{split}\dot{\mathcal{V}}\leq&-\Big{(}H\varepsilon(q-\frac{1}{2})-\frac{\lambda\delta_{1}}{8}-\frac{\delta_{2}}{4}\Big{)}\tilde{w}^{2}(1,t)\\ &-\Big{(}\frac{H\varepsilon}{4}-\frac{5\lambda\delta_{1}}{16}-\frac{5\delta_{2}}{8}\Big{)}\|\tilde{w}[t]\|^{2}\\ &-\Big{(}\varepsilon(r-\frac{1}{2})-\frac{\lambda}{8\delta_{1}}\Big{)}\hat{w}^{2}(1,t)\\ &-\Big{(}\frac{\varepsilon}{4}-\frac{5\lambda}{16\delta_{1}}-\frac{\|g\|^{2}}{2\delta_{2}}\Big{)}\|\hat{w}[t]\|^{2}.\end{split} (109)

From (96) and (97), we have that

Hε(q12)λδ18δ24>0 and ε(r12)λ8δ1>0,H\varepsilon(q-\frac{1}{2})-\frac{\lambda\delta_{1}}{8}-\frac{\delta_{2}}{4}>0\text{ and }\varepsilon(r-\frac{1}{2})-\frac{\lambda}{8\delta_{1}}>0, (110)

Thus, it follows from (109) that

𝒱˙(Hε45λδ1165δ28)w~[t]2(ε45λ16δ1g22δ2)w^[t]2.\begin{split}\dot{\mathcal{V}}\leq&-\Big{(}\frac{H\varepsilon}{4}-\frac{5\lambda\delta_{1}}{16}-\frac{5\delta_{2}}{8}\Big{)}\|\tilde{w}[t]\|^{2}\\ &-\Big{(}\frac{\varepsilon}{4}-\frac{5\lambda}{16\delta_{1}}-\frac{\|g\|^{2}}{2\delta_{2}}\Big{)}\|\hat{w}[t]\|^{2}.\end{split} (111)

Again, we have from (96) and (97) that

ϑ1=Hε45λδ1165δ28>0,\vartheta_{1}=\frac{H\varepsilon}{4}-\frac{5\lambda\delta_{1}}{16}-\frac{5\delta_{2}}{8}>0, (112)
ϑ2=ε45λ16δ1g22δ2>0.\vartheta_{2}=\frac{\varepsilon}{4}-\frac{5\lambda}{16\delta_{1}}-\frac{\|g\|^{2}}{2\delta_{2}}>0. (113)

Therefore, (111) can be written as

𝒱˙min{ϑ1,ϑ2}max{H/2,1/2}𝒱.\dot{\mathcal{V}}\leq-\frac{\min\big{\{}\vartheta_{1},\vartheta_{2}\big{\}}}{\max\big{\{}H/2,1/2\big{\}}}\mathcal{V}. (114)

Hence, from standard arguments, we can state that the closed-loop system which consists of the plant (1) and the observer (3) with the continuous-time control law (14), is globally exponentially stable in L2L^{2}-sense.

References

  • [1] T. Chen and B. A. Francis, “Input-output stability of sampled-data systems,” IEEE Transactions on Automatic Control, vol. 36, no. 1, pp. 50–58, 1991.
  • [2] K. J. Åström and B. Wittenmark, Computer-controlled systems: theory and design. Courier Corporation, 2013.
  • [3] G. C. Walsh and H. Ye, “Scheduling of networked control systems,” IEEE control systems magazine, vol. 21, no. 1, pp. 57–65, 2001.
  • [4] X. Wang and M. D. Lemmon, “Event-triggering in distributed networked control systems,” IEEE Transactions on Automatic Control, vol. 56, no. 3, pp. 586–601, 2011.
  • [5] C. Peng and T. C. Yang, “Event-triggered communication and h\infty control co-design for networked control systems,” Automatica, vol. 49, no. 5, pp. 1326–1332, 2013.
  • [6] C. Nowzari, E. Garcia, and J. Cortés, “Event-triggered communication and control of networked systems for multi-agent consensus,” Automatica, vol. 105, pp. 1–27, 2019.
  • [7] L. Hetel, C. Fiter, H. Omran, A. Seuret, E. Fridman, J.-P. Richard, and S. I. Niculescu, “Recent developments on the stability of systems with aperiodic sampling: An overview,” Automatica, vol. 76, pp. 309–335, 2017.
  • [8] I. Karafyllis and M. Krstic, “Sampled-data boundary feedback control of 1-D linear transport PDEs with non-local terms,” Systems & Control Letters, vol. 107, pp. 68–75, 2017.
  • [9] ——, “Sampled-data boundary feedback control of 1-D parabolic PDEs,” Automatica, vol. 87, pp. 226–237, 2018.
  • [10] W. Kang and E. Fridman, “Distributed sampled-data control of kuramoto–sivashinsky equation,” Automatica, vol. 95, pp. 514–524, 2018.
  • [11] W. Heemels, K. H. Johansson, and P. Tabuada, “An introduction to event-triggered and self-triggered control,” in 2012 IEEE 51st IEEE Conference on Decision and Control (CDC). IEEE, 2012, pp. 3270–3285.
  • [12] Z. Yao and N. H. El-Farra, “Resource-aware model predictive control of spatially distributed processes using event-triggered communication,” in 52nd IEEE conference on decision and control. IEEE, 2013, pp. 3726–3731.
  • [13] M. Lemmon, “Event-triggered feedback in control, estimation, and optimization,” in Networked control systems. Springer, 2010, pp. 293–358.
  • [14] P. Tabuada, “Event-triggered real-time scheduling of stabilizing control tasks,” IEEE Transactions on Automatic Control, vol. 52, no. 9, pp. 1680–1685, 2007.
  • [15] M. C. F. Donkers and W. P. M. H. Heemels, “Output-based event-triggered control with guaranteed \mathcal{L}_{\infty}-gain and improved and decentralized event-triggering,” IEEE Transactions on Automatic Control, vol. 57, no. 6, pp. 1362–1376, 2012.
  • [16] W. H. Heemels, M. Donkers, and A. R. Teel, “Periodic event-triggered control for linear systems,” IEEE Transactions on Automatic Control, vol. 58, no. 4, pp. 847–861, 2013.
  • [17] N. Marchand, S. Durand, and J. F. G. Castellanos, “A general formula for event-based stabilization of nonlinear systems,” IEEE Transactions on Automatic Control, vol. 58, no. 5, pp. 1332–1337, 2013.
  • [18] P. Tallapragada and N. Chopra, “On event triggered tracking for nonlinear systems,” IEEE Transactions on Automatic Control, vol. 58, no. 9, pp. 2343–2348, 2013.
  • [19] R. Postoyan, P. Tabuada, D. Nešić, and A. Anta, “A framework for the event-triggered stabilization of nonlinear systems,” IEEE Transactions on Automatic Control, vol. 60, no. 4, pp. 982–996, 2015.
  • [20] A. Girard, “Dynamic triggering mechanisms for event-triggered control,” IEEE Transactions on Automatic Control, vol. 60, no. 7, pp. 1992–1997, 2015.
  • [21] M. Abdelrahim, R. Postoyan, J. Daafouz, and D. Nešić, “Stabilization of nonlinear systems using event-triggered output feedback controllers,” IEEE Transactions on Automatic Control, vol. 61, no. 9, pp. 2682–2687, 2016.
  • [22] D. P. Borgers, R. Postoyan, A. Anta, P. Tabuada, D. Nešić, and W. Heemels, “Periodic event-triggered control of nonlinear systems using overapproximation techniques,” Automatica, vol. 94, pp. 81–87, 2018.
  • [23] W. Wang, R. Postoyan, D. Nešić, and W. Heemels, “Periodic event-triggered control for nonlinear networked control systems,” IEEE Transactions on Automatic Control, vol. 65, no. 2, pp. 620–635, 2020.
  • [24] P. Zhang, T. Liu, and Z.-P. Jiang, “Event-triggered stabilization of a class of nonlinear time-delay systems,” IEEE Transactions on Automatic Control, 2020.
  • [25] ——, “Systematic design of robust event-triggered state and output feedback controllers for uncertain nonholonomic systems,” IEEE Transactions on Automatic Control, 2020.
  • [26] A. Selivanov and E. Fridman, “Distributed event-triggered control of diffusion semilinear PDEs,” Automatica, vol. 68, pp. 344–351, 2016.
  • [27] N. Espitia, A. Girard, N. Marchand, and C. Prieur, “Event-based control of linear hyperbolic systems of conservation laws,” Automatica, vol. 70, pp. 275–287, 2016.
  • [28] N. Espitia, A. Tanwani, and S. Tarbouriech, “Stabilization of boundary controlled hyperbolic PDEs via lyapunov-based event triggered sampling and quantization,” in 2017 IEEE 56th Annual Conference on Decision and Control (CDC). IEEE, 2017, pp. 1266–1271.
  • [29] N. Espitia, A. Girard, N. Marchand, and C. Prieur, “Event-based boundary control of a linear 2x2 hyperbolic system via backstepping approach,” IEEE Transactions on Automatic Control, vol. 63, no. 8, pp. 2686–2693, 2018.
  • [30] J.-W. Wang, “Observer-based boundary control of semi-linear parabolic PDEs with non-collocated distributed event-triggered observation,” Journal of the Franklin Institute, vol. 356, no. 17, pp. 10 405–10 420, 2019.
  • [31] N. Espitia, “Observer-based event-triggered boundary control of a linear 2×\times 2 hyperbolic systems,” Systems & Control Letters, vol. 138, p. 104668, 2020.
  • [32] N. Espitia, I. Karafyllis, and M. Krstic, “Event-triggered boundary control of constant-parameter reaction-diffusion PDEs: a small-gain approach,” in 2020 American Control Conference (ACC). IEEE, 2020, pp. 3437–3442.
  • [33] M. Diagne and I. Karafyllis, “Event-triggered control of a continuum model of highly re-entrant manufacturing system,” arXiv preprint arXiv:2004.01916, 2020.
  • [34] D. Xue and N. H. El-Farra, “Output feedback-based event-triggered control of distributed processes with communication constraints,” in 2016 IEEE 55th Conference on Decision and Control (CDC). IEEE, 2016, pp. 4296–4301.
  • [35] F. Ge and Y. Chen, “Observer-based event-triggered control for semilinear time-fractional diffusion systems with distributed feedback,” Nonlinear Dynamics, vol. 99, no. 2, pp. 1089–1101, 2020.
  • [36] A. D. Polyanin and V. E. Nazaikinskii, Handbook of linear partial differential equations for engineers and scientists. CRC press, 2015.
  • [37] I. Karafyllis and M. Krstic, Input-to-state stability for PDEs. Springer, 2019.
  • [38] A. Smyshlyaev and M. Krstic, “Closed-form boundary state feedbacks for a class of 1-D partial integro-differential equations,” IEEE Transactions on Automatic Control, vol. 49, no. 12, pp. 2185–2202, 2004.
  • [39] M. Krstic and A. Smyshlyaev, Boundary control of PDEs: A course on backstepping designs. SIAM, 2008.