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

Delayed finite-dimensional observer-based control of 2D linear parabolic PDEs

Pengfei Wang [email protected]    Emilia Fridman [email protected] School of Electrical Engineering, Tel-Aviv University, Tel-Aviv, Israel
Abstract

Recently, a constructive method was suggested for finite-dimensional observer-based control of 1D linear heat equation, which is robust to input/output delays. In this paper, we aim to extend this method to the 2D case with general time-varying input/output delays (known output delay and unknown input delay) or sawtooth delays (that correspond to network-based control). We use the modal decomposition approach and consider boundary or non-local sensing together with non-local actuation, or Neumann actuation with non-local sensing. To compensate the output delay that appears in the infinite-dimensional part of the closed-loop system, for the first time for delayed PDEs we suggest a vector Lyapunov functional combined with the recently introduced vector Halanay inequality. We provide linear matrix inequality (LMI) conditions for finding the observer dimension and upper bounds on delays that preserve the exponential stability. We prove that the LMIs are always feasible for large enough observer dimension and small enough upper bounds on delays. A numerical example demonstrates the efficiency of our method and show that the employment of vector Halanay’s inequality allows for larger delays than the classical scalar Halanay inequality for comparatively large observer dimension.

keywords:
2D parabolic PDEs, observer-based control, time delay, vector Halanay’s inequality.
thanks: This work was supported by Israel Science Foundation (grant no. 673/19) and by Chana and Heinrich Manderman Chair at Tel Aviv University.

,

1 Introduction

Finite-dimensional observer-based controllers for PDEs are attractive in applications. Such controllers were designed by the modal decomposition approach and have been extensively studied since the 1980s [2, 4, 5, 9, 10], where efficient bound estimate on the observer and controller dimensions is a challenging problem. In recent paper [13], the first constructive LMI-based method for finite-dimensional observer-based control of 1D parabolic PDEs was suggested, where the observer dimension was found from simple LMI conditions. The results in [13] were then extended to input/output delay robustness in [14, 15, 16], delayed PDEs [20] and delay compensation in [15, 19, 18, 21]. However, the results of [13, 14, 15, 16, 19, 18, 20, 21] were confined to 1D parabolic PDEs.

In recent years, control of high-dimensional PDEs became an active research area. Such systems have promising applications in engineering, water heating, metal rolling, sheet forming, medical imaging (see e.g. [26]) as well as in multi-agents deployment [29]. Sampled-data observers for 22D and NND heat equations with globally Lipschitz nonlinearities have been suggested in [1, 30]. Observer-based output-feedback controller for a linear parabolic NND PDEs was designed in [34]. In [12], the sampled-data control of 2D Kuramoto-Sivashinsky equation was explored. The results in [1, 12, 30, 34] were in rectangular domain and employed spatial decomposition approach where many sensors/actuators should be utilized.

The boundary state-feedback stabilization of NND parabolic PDEs was studied in [3, 27] by modal decomposition approach and in [26, 23] by backstepping method. Observer-based boundary control for NND parabolic PDEs under boundary measurement over cubes and balls was explored in [11, 33] by the backstepping method. In [6, 25], observer-based control via modal decomposition approach was designed for NND parabolic PDEs. Note that the observer designs in [6, 11, 25, 33] are in the from of PDEs. In [17], for the first time, the finite-dimensional observer-based control was studied for 2D and 3D parabolic PDEs under boundary actuation on an arbitrary subdomain and in-domain pointwise measurement. It was shown in [17] that the closed-loop system is stable provided the dimension of the controller is large enough. Note that the results in [6, 11, 17, 25, 33] are confined to observer-based controller design of NND delay-free PDEs. For NND parabolic PDEs, efficient finite-dimensional observer-based design with a quantitative bound on the observer as well as the input/output delay robustness remained open challenging problems.

In this paper, we aim to study finite-dimensional observer-based control of linear heat equation with input/output delays in Ω\Omega, an open and connected subset of 2\mathbb{R}^{2}. We consider either differentiable time-varying delays (unknown input delay and known output delay) or sawtooth delays (that correspond to network-based control). Based on modal decomposition approach, we consider the boundary or non-local sensing together with non-local actuation, or to Neumann actuation with non-local sensing. The novelty of this paper compared to existing results can be formulated as follows:

  • Compared with [6, 11, 17, 25, 33] for observer-based design of high-dimensional parabolic PDEs, we give efficient finite-dimensional observer-based design and provide LMI conditions for finding observer dimension and upper bounds of delays. We prove that the LMIs are always feasible for large enough observer dimension and small enough upper bounds on delays.

  • Differently from [14, 15, 16] for 1D parabolic PDEs where Lyapunov functional combined with classical scalar Halanay’s inequality (see P. 138 in [7]) was suggested, we construct vector Lyapunov functional combined with recently introduced vector Halanay’s inequality (see [24]). The latter allows to efficiently compensate the fast-varying output delay that appears in the infinite-dimensional part of the closed loop system essentially improving the upper bounds on delays in most of the numerical examples.

  • Compared with spatial decomposition approach suggested in [1, 12, 30, 34] for robust stabilization of NND parabolic PDEs, the modal decomposition approach in this paper allows for fewer actuators and sensors (including single boundary actuator or sensor).

Notations and preliminaries: For any bounded domain Ω2\Omega\subset\mathbb{R}^{2}, denote by L2(Ω)L^{2}(\Omega) the space of square integrable functions with inner product f,g=Ωf(x)g(x)dx\langle f,g\rangle=\int_{\Omega}f(x)g(x)\mathrm{d}x and induced norm fL22=f,f\|f\|_{L^{2}}^{2}=\langle f,f\rangle. H1(Ω)H^{1}(\Omega) is the Sobolev space of functions f:Ωf:\Omega\longrightarrow\mathbb{R} with a square integrable weak derivative. The norm defined in H1(Ω)H^{1}(\Omega) is fH12=fL22+fL22\|f\|^{2}_{H^{1}}=\|f\|^{2}_{L^{2}}+\|\nabla f\|^{2}_{L^{2}}, where f=[fx1,fx2]T\nabla f=[f_{x_{1}},f_{x_{2}}]^{\mathrm{T}} and fL22=Ω[(fx1)2+(fx2)2]dx\|\nabla f\|^{2}_{L^{2}}=\int_{\Omega}[(f_{x_{1}})^{2}+(f_{x_{2}})^{2}]\mathrm{d}x. The Euclidean norm is denoted by |||\cdot|. For Pn×nP\in\mathbb{R}^{n\times n}, P>0P>0 means that PP is symmetric and positive definite. The symmetric elements of a symmetric matrix will be denoted by *. For 0<Pn×n0<P\in\mathbb{R}^{n\times n} and xnx\in\mathbb{R}^{n}, we write |x|P2=xTPx|x|^{2}_{P}=x^{\mathrm{T}}Px. Denote \mathbb{N} by the set of positive integers.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open connected set. Following [32], we assume that either the boundary Ω\partial\Omega is of class C2C^{2} or Ω\Omega is a rectangular domain. Let Ω\partial\Omega be split into two disjoint parts Ω=ΓDΓN\partial\Omega=\Gamma_{D}\cup\Gamma_{N} such that ΓD\Gamma_{D} and ΓN\Gamma_{N} have non-zero Lebsgue measurement. Here subscripts D and N stand for Dirichlet and for Neumann boundary conditions respectively. Let

𝒜ϕ=Δϕ,𝒟(𝒜)={ϕ|ϕH2(Ω)HΓ1(Ω)},HΓ1(Ω)={ϕH1(Ω)|ϕ(x)=0forxΓD,ϕ𝐧(x)=0forxΓN},{\scriptsize\begin{array}[]{ll}\mathcal{A}\phi=-\Delta\phi,~{}\mathcal{D}(\mathcal{A})=\{\phi|\phi\in H^{2}(\Omega)\cap H^{1}_{\Gamma}(\Omega)\},\\ H^{1}_{\Gamma}(\Omega)=\{\phi\in H^{1}(\Omega)|\phi(x)=0~{}\mathrm{for}~{}x\in\Gamma_{D},\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\frac{\partial\phi}{\partial{\bf n}}(x)=0~{}\mathrm{for}~{}x\in\Gamma_{N}\},\vspace{-0.3cm}\end{array}} (1.1)

where 𝐧\frac{\partial}{\partial{\bf n}} is the normal derivative. It follows from [32, Proposition 3.2.12] that the eigenvalues {λn}n=1\{\lambda_{n}\}_{n=1}^{\infty} of 𝒜\mathcal{A} are real and we can repeat each eigenvalue according to its finite multiplicity to get

λ1<λ2λn,limnλn=.\begin{array}[]{ll}\lambda_{1}<\lambda_{2}\leq\dots\leq\lambda_{n}\leq\dots,~{}~{}\lim_{n\rightarrow\infty}\lambda_{n}=\infty.\vspace{-0.3cm}\end{array} (1.2)

We denote the corresponding eigenfunctions as {ϕn}n=1\{\phi_{n}\}_{n=1}^{\infty}. Differently from the 1D case where λN=O(N2)\lambda_{N}=O(N^{2}), NN\rightarrow\infty, for λN\lambda_{N}, we have the following estimate which will be used for the asymptotic feasibility of LMIs:

Lemma 1.1.

([31, Sec. 11.6]) For eigenvalues (1.2), we have limNλNN=4π|Ω|\lim_{N\rightarrow\infty}\frac{\lambda_{N}}{N}=\frac{4\pi}{|\Omega|}, where |Ω||\Omega| is the area of Ω\Omega.

Since 𝒜\mathcal{A} is strictly positive and diagonalizable, we have (see Proposition 3.4.8 in [32])

𝒟(𝒜12)={hL2(Ω)|n=1λn|h,ϕn|<}.\begin{array}[]{ll}\mathcal{D}(\mathcal{A}^{\frac{1}{2}})=\{h\in L^{2}(\Omega)|\sum_{n=1}^{\infty}\lambda_{n}|\langle h,\phi_{n}\rangle|<\infty\}.\vspace{-0.3cm}\end{array}

Following Remark 3.4.4 in [32], we can regard 𝒟(𝒜12)\mathcal{D}(\mathcal{A}^{\frac{1}{2}}) as the completion of 𝒟(𝒜)\mathcal{D}(\mathcal{A}) with respect to the norm f12=𝒜f,f=n=1λn|f,ϕn|2\|f\|_{\frac{1}{2}}=\sqrt{\langle\mathcal{A}f,f\rangle}=\sqrt{\sum_{n=1}^{\infty}\lambda_{n}|\langle f,\phi_{n}\rangle|^{2}}, f𝒟(𝒜)f\in\mathcal{D}(\mathcal{A}). For h𝒟(𝒜)h\in\mathcal{D}(\mathcal{A}), we have hL22=h,𝒜h=h122\|\nabla h\|_{L^{2}}^{2}=\langle h,\mathcal{A}h\rangle=\|h\|^{2}_{\frac{1}{2}}, which implies

hL22=n=1λnhn2.\begin{array}[]{ll}\|\nabla h\|_{L^{2}}^{2}=\sum_{n=1}^{\infty}\lambda_{n}h_{n}^{2}.\vspace{-0.3cm}\end{array} (1.3)

We have fL22C(Ω)fL22\|f\|^{2}_{L^{2}}\leq C(\Omega)\|\nabla f\|_{L^{2}}^{2}, f|ΓD=0f|_{\Gamma_{D}}=0 for some constant C(Ω)>0C(\Omega)>0 (see [8]), which together with (1.3) implies the equivalence of 12\|\cdot\|_{\frac{1}{2}} and H1\|\cdot\|_{H^{1}} subject to f(x)=0f(x)=0, xΓDx\in\Gamma_{D}. We have 𝒟(𝒜12)={hH1(Ω)|h(x)=0,xΓD}\mathcal{D}(\mathcal{A}^{\frac{1}{2}})=\{h\in H^{1}(\Omega)|h(x)=0,x\in\Gamma_{D}\}. Finally, density of 𝒟(𝒜)\mathcal{D}(\mathcal{A}) in 𝒟(𝒜12)\mathcal{D}(\mathcal{A}^{\frac{1}{2}}) yields that (1.3) holds for any h=L2(Ω)n=1hnϕn𝒟(𝒜12)h\overset{L^{2}(\Omega)}{=}\sum_{n=1}^{\infty}h_{n}\phi_{n}\in\mathcal{D}(\mathcal{A}^{\frac{1}{2}}).

Given a positive integer NN and hL2(Ω)h\in L^{2}(\Omega) satisfying h=L2n=1hnϕnh\overset{L^{2}}{=}\sum_{n=1}^{\infty}h_{n}\phi_{n}, where hn=h,ϕnh_{n}=\langle h,\phi_{n}\rangle, we denote hN2=n=N+1hn2\|h\|^{2}_{N}=\sum_{n=N+1}^{\infty}h^{2}_{n}. For ϕL2(Ω)\phi\in L^{2}(\Omega) and 𝐛=[b1,,bd]T(L2(Ω))d{\bf b}=[b_{1},\dots,b_{d}]^{\mathrm{T}}\in(L^{2}(\Omega))^{d}, we denote 𝐛,ϕ=[b1,ϕ,,bd,ϕ]T\langle{\bf b},\phi\rangle=[\langle b_{1},\phi\rangle,\dots,\langle b_{d},\phi\rangle]^{\mathrm{T}}.

Lemma 1.2.

(Vector Halanay’s Inequality [24]) Let Mn×nM\in\mathbb{R}^{n\times n} be a Metzler and Hurwitz matrix and Pn×nP\in\mathbb{R}^{n\times n} be a nonnegative matrix. Let τ=max{τ1,,τn}\tau=\max\{\tau_{1},\dots,\tau_{n}\} with τi>0\tau_{i}>0 and V=[V1,,Vn]T:[τ,)[0,)nV=[V_{1},\dots,V_{n}]^{\mathrm{T}}:[-\tau,\infty)\rightarrow[0,\infty)^{n} be C1C^{1} and

V˙(t)MV(t)+Psups[tτ,t]V(s),\begin{array}[]{ll}\dot{V}(t)\leq MV(t)+P\sup_{s\in[t-\tau,t]}V(s),\vspace{-0.3cm}\end{array}

where sups[tτ,t]V(s)=col{sups[tτi,t]Vi(s)}i=1n\sup_{s\in[t-\tau,t]}V(s)=\mathrm{col}\{\sup_{s\in[t-\tau_{i},t]}V_{i}(s)\}_{i=1}^{n}. If M+PM+P is Hurwitz, then |V(t)|Deδ0t|V(t)|\leq D\mathrm{e}^{-\delta_{0}t}, t0t\geq 0 for some δ0>0\delta_{0}>0 and D>0D>0.

2 Non-local actuation and measurement

2.1 System under consideration and controller design

Consider the following heat equation under delayed nonlocal actuation:

zt(x,t)=Δz(x,t)+qz(x,t)+𝐛T(x)u(tτu(t)),inΩ×(0,),z(x,t)=0,onΓD×(0,),z𝐧(x,t)=0,onΓN×(0,),z(,0)=z0()L2(Ω),\begin{array}[]{ll}z_{t}(x,t)=\Delta z(x,t)+qz(x,t)\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+{\bf b}^{\mathrm{T}}(x)u(t-\tau_{u}(t)),~{}\mathrm{in}~{}\Omega\times(0,\infty),\\ z(x,t)=0,~{}~{}\mathrm{on}~{}\Gamma_{D}\times(0,\infty),\\ \frac{\partial z}{\partial{\bf n}}(x,t)=0,~{}~{}\mathrm{on}~{}\Gamma_{N}\times(0,\infty),\\ z(\cdot,0)=z_{0}(\cdot)\in L^{2}(\Omega),\vspace{-0.3cm}\end{array} (2.1)

where qq\in\mathbb{R} is a constant reaction coefficient, τu(t)\tau_{u}(t) is a known input delay which is upper bounded by τM,u\tau_{M,u}. 𝐛=[b1,,bd]T(L2(Ω))d{\bf b}=[b_{1},\dots,b_{d}]^{\mathrm{T}}\in(L^{2}(\Omega))^{d}, u(t)=[u1(t),,ud(t)]Tu(t)=[u_{1}(t),\dots,u_{d}(t)]^{\mathrm{T}} is the control input to be designed later. Let δ>0\delta>0. From (1.2), it follows that there exists N0N_{0}\in\mathbb{N} such that

λn+q+δ<0,n>N0,-\lambda_{n}+q+\delta<0,~{}n>N_{0},\vspace{-0.3cm} (2.2)

where N0N_{0} is the number of modes used for the controller design. Let NN\in\mathbb{N}, NN0N\geq N_{0}, where NN will be the dimension of the observer. Let dd be the maximum of the geometric multiplicities of λn\lambda_{n}, n=1,,N0n=1,\dots,N_{0}. Assume the following delayed non-local measurement:

y(t)=𝐜,z(,tτy(t)),tτy(t)0,y(t)=0,tτy(t)<0,𝐜=[c1,,cd]T(L2(Ω))d,\begin{array}[]{ll}y(t)=\langle{\bf c},z(\cdot,t-\tau_{y}(t))\rangle,~{}~{}t-\tau_{y}(t)\geq 0,\\ y(t)=0,t-\tau_{y}(t)<0,~{}{\bf c}=[c_{1},\dots,c_{d}]^{\mathrm{T}}\in(L^{2}(\Omega))^{d},\vspace{-0.14cm}{}{}{}\end{array} (2.3)

where τy(t)\tau_{y}(t) is a known measurement delay which is upper bounded by τM,y\tau_{M,y}. The controller construction will follow [13] for 1D case (where only simple eigenvalues appear), but the single-input and single-output as in [13] are not applicable to the 2D case due to the existence of multiple eigenvalues (the system is uncontrollable and unobservable). Here we introduce multi-input u(t)u(t) and multi-output (2.3) with 𝐛{\bf b}, 𝐜{\bf c} satisfying Assumption 1 (see below) to manage with the controllability and observability.

We treat two classes of input/output delays: continuously differentiable delays and sawtooth delays that correspond to network-based control. For the case of continuously differentiable delays, we assume that τu(t)\tau_{u}(t) and τy(t)\tau_{y}(t) are lower bounded by τm>0\tau_{m}>0. This assumption is employed for well-posedness only. As in [14, 22], we assume that there exists a unique t[τm,min{τM,y,τM,u}]t_{*}\in[\tau_{m},\min\{\tau_{M,y},\tau_{M,u}\}] such that tτ(t)<0t-\tau(t)<0 if t<tt<t_{*} and tτ(t)0t-\tau(t)\geq 0 if ttt\geq t_{*} for τ(t){τu(t),τy(t)}\tau(t)\in\{\tau_{u}(t),\tau_{y}(t)\}. For the case of sawtooth delays, τy(t)\tau_{y}(t) and τu(t)\tau_{u}(t) are induced by two networks: from sensor to controller and from controller to actuator, respectively (see Section 7.5 in [7]). Henceforth the dependence of τy(t)\tau_{y}(t) and τu(t)\tau_{u}(t) on tt will be suppressed to shorten notations.

We present the solution to (2.1) as

z(x,t)=L2n=1zn(t)ϕn(x),zn(t)=z(,t),ϕn,\begin{array}[]{ll}z(x,t)\overset{L^{2}}{=}\sum_{n=1}^{\infty}z_{n}(t)\phi_{n}(x),~{}z_{n}(t)=\langle z(\cdot,t),\phi_{n}\rangle,\vspace{-0.3cm}\end{array} (2.4)

where {ϕn}n=1\{\phi_{n}\}_{n=1}^{\infty} are corresponding eigenfunctions of eigenvalues (1.2). Differentiating znz_{n} in (2.4) and applying Green’s first identity, we obtain

z˙n(t)=(λn+q)zn(t)+𝐛nTu(tτu),t0,zn(0)=z(,0),ϕn,𝐛n=𝐛,ϕnd.\begin{array}[]{ll}\dot{z}_{n}(t)=(-\lambda_{n}+q)z_{n}(t)+{\bf b}_{n}^{\mathrm{T}}u(t-\tau_{u}),~{}t\geq 0,\\ z_{n}(0)=\langle z(\cdot,0),\phi_{n}\rangle,~{}~{}{\bf b}_{n}=\langle{\bf b},\phi_{n}\rangle\in\mathbb{R}^{d}.\vspace{-0.3cm}\end{array} (2.5)

We construct a NN-dimensional observer of the form

z^(x,t)=n=1Nz^n(t)ϕn(x),N>N0,\begin{array}[]{ll}\hat{z}(x,t)=\sum_{n=1}^{N}\hat{z}_{n}(t)\phi_{n}(x),~{}~{}N>N_{0},\vspace{-0.25cm}\end{array} (2.6)

where z^n(t)\hat{z}_{n}(t) (1nN)(1\leq n\leq N) satisfy

z^˙n(t)=(λn+q)z^n(t)+𝐛nTu(tτu)ln[𝐜,z^(,tτy)y(t)],t>0,z^n(0)=0,t0,\begin{array}[]{ll}\dot{\hat{z}}_{n}(t)=(-\lambda_{n}+q)\hat{z}_{n}(t)+{\bf b}_{n}^{\mathrm{T}}u(t-\tau_{u})\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-l_{n}\left[\langle{\bf c},\hat{z}(\cdot,t-\tau_{y})\rangle-y(t)\right],~{}t>0,\\ \hat{z}_{n}(0)=0,~{}~{}t\leq 0,\vspace{-0.3cm}\end{array} (2.7)

with y(t)y(t) in (2.3), observer gains ln1×dl_{n}\in\mathbb{R}^{1\times d}, 1nN01\leq n\leq N_{0} being designed later and ln=01×dl_{n}=0_{1\times d} for N0<nNN_{0}<n\leq N.

Introduce the notations

A0=diag{λn+q}n=1N0,A1=diag{λn+q}n=N0+1N,𝐜n=𝐜,ϕn,𝐂0=[𝐜1,,𝐜N0],𝐂1=[𝐜N0+1,,𝐜N],𝐁0=[𝐛1,,𝐛N0]T,𝐁1=[𝐛N0+1,,𝐛N]T.\begin{array}[]{ll}A_{0}=\mathrm{diag}\{-\lambda_{n}+q\}_{n=1}^{N_{0}},A_{1}=\mathrm{diag}\{-\lambda_{n}+q\}_{n=N_{0}+1}^{N},\\ {\bf c}_{n}=\langle{\bf c},\phi_{n}\rangle,{\bf C}_{0}=[{\bf c}_{1},\dots,{\bf c}_{N_{0}}],{\bf C}_{1}=[{\bf c}_{N_{0}+1},\dots,{\bf c}_{N}],\\ {\bf B}_{0}=[{\bf b}_{1},\dots,{\bf b}_{N_{0}}]^{\mathrm{T}},~{}~{}{\bf B}_{1}=[{\bf b}_{N_{0}+1},\dots,{\bf b}_{N}]^{\mathrm{T}}.\vspace{-0.3cm}\end{array} (2.8)

We rewrite A0A_{0} as:

A0=diag{A~1,,A~p},A~j=diag{λj+q,,λj+q}nj×nj,λkλjiffkj,k,j=1,,p,\begin{array}[]{ll}A_{0}=\mathrm{diag}\{\tilde{A}_{1},\dots,\tilde{A}_{p}\},\\ \tilde{A}_{j}=\mathrm{diag}\{-\lambda_{j}+q,\dots,-\lambda_{j}+q\}\in\mathbb{R}^{n_{j}\times n_{j}},\\ \lambda_{k}\neq\lambda_{j}~{}~{}\mathrm{iff}~{}~{}k\neq j,~{}~{}k,j=1,\dots,p,\vspace{-0.3cm}\end{array} (2.9)

where n1,,npn_{1},\dots,n_{p} are positive integers such that n1++np=N0n_{1}+\dots+n_{p}=N_{0}. Clearly, njdn_{j}\leq d, j=1,,pj=1,\dots,p and there exists at least one ȷ{1,,p}\jmath\in\{1,\dots,p\} such that nȷ=dn_{\jmath}=d. According to the partition of (2.9), we rewrite 𝐁0{\bf B}_{0} and 𝐂0{\bf C}_{0} as

𝐁0=col{Bj}j=1p,Bjnj×d,𝐂0=[C1,,Cp],Cjd×nj.\begin{array}[]{ll}{\bf B}_{0}=\mathrm{col}\{B_{j}\}_{j=1}^{p},~{}~{}B_{j}\in\mathbb{R}^{n_{j}\times d},\\ {\bf C}_{0}=[C_{1},\dots,C_{p}],~{}~{}C_{j}\in\mathbb{R}^{d\times n_{j}}.\vspace{-0.3cm}\end{array}
Assumption 1.

Let rank(Bj)=nj\mathrm{rank}(B_{j})=n_{j} and rank(Cj)=nj\mathrm{rank}(C_{j})=n_{j}, j=1,,pj=1,\dots,p.

Lemma 2.1.

Under Assumption 1, the pair (A0,𝐁0)(A_{0},{\bf B}_{0}) is controllable and the pair (A0,𝐂0)(A_{0},{\bf C}_{0}) is observable.

{pf}

The proof is inspired by Lemma 7.2 of [25]. Assume that the pair (A0,𝐂0)(A_{0},{\bf C}_{0}) is not observable. By the Hautus test (see [32, Remark 1.5.2]), there exist 0νN00\neq\nu\in\mathbb{R}^{N_{0}} and j{1,,p}j\in\{1,\dots,p\} such that

A0ν=λjν,𝐂0ν=0.\begin{array}[]{ll}A_{0}\nu=\lambda_{j}\nu,~{}~{}~{}~{}{\bf C}_{0}\nu=0.\vspace{-0.3cm}\end{array} (2.10)

Without loss of generality, we suppose that ν=col{ν1,,νp}\nu=\mathrm{col}\{\nu_{1},\dots,\nu_{p}\}, where νj=[νj(1),,νj(nj)]T\nu_{j}=[\nu^{(1)}_{j},\dots,\nu^{(n_{j})}_{j}]^{\mathrm{T}}. Then (2.10) becomes A0νλjν=col{(λkλj)νk}k=1p=0A_{0}\nu-\lambda_{j}\nu=\mathrm{col}\{(\lambda_{k}-\lambda_{j})\nu_{k}\}_{k=1}^{p}=0 and k=1pCkνk=0\sum_{k=1}^{p}C_{k}\nu_{k}=0, which implies νk=0\nu_{k}=0 for kjk\neq j and Cjνj=0C_{j}\nu_{j}=0. Since rank(Cj)=nj\mathrm{rank}(C_{j})=n_{j}, we have νj=0\nu_{j}=0. This contradicts to the fact ν0\nu\neq 0. Therefore, pair (A0,𝐂0)(A_{0},{\bf C}_{0}) is observable. The controllability of (A0,𝐁0)(A_{0},{\bf B}_{0}) follows similarly.

Under Assumption 1, we can let L0=col{l1,,lN0}N0×dL_{0}=\mathrm{col}\{l_{1},\dots,l_{N_{0}}\}\in\mathbb{R}^{N_{0}\times d} and K0d×N0K_{0}\in\mathbb{R}^{d\times N_{0}} satisfy

Po(A0L0𝐂0)+(A0L0𝐂0)TPo<2δPo,\displaystyle P_{o}(A_{0}-L_{0}{\bf C}_{0})+(A_{0}-L_{0}{\bf C}_{0})^{\mathrm{T}}P_{o}<-2\delta P_{o}, (2.11a)
Pc(A0𝐁0K0)+(A0𝐁0K0)TPc2δPc,\displaystyle P_{c}(A_{0}-{\bf B}_{0}K_{0})+(A_{0}-{\bf B}_{0}K_{0})^{\mathrm{T}}P_{c}\leq-2\delta P_{c}, (2.11b)

for 0<Po,PcN0×N00<P_{o},P_{c}\in\mathbb{R}^{N_{0}\times N_{0}}. We propose a controller of the form

u(t)=K0z^N0(t),z^N0=[z^1,,z^N0]T.\begin{array}[]{ll}u(t)=-K_{0}\hat{z}^{N_{0}}(t),~{}~{}\hat{z}^{N_{0}}=[\hat{z}_{1},\dots,\hat{z}_{N_{0}}]^{\mathrm{T}}.\vspace{-0.3cm}\end{array} (2.12)

For well-posedness of closed-loop system (2.1), (2.7) with control input (2.12), we consider the state ξ(t)=col{z(,t),z^N(t)}\xi(t)=\mathrm{col}\{z(\cdot,t),\hat{z}^{N}(t)\}, where z^N(t)=col{z^n(t)}n=1N\hat{z}^{N}(t)=\mathrm{col}\{\hat{z}_{n}(t)\}_{n=1}^{N}. The closed-loop system can be presented as

ddtξ(t)+diag{𝒜,𝒜0}ξ(t)=[qz(,t)+f1(tτu)f2(tτu)+f3(tτy)],𝒜0=diag{A0,A1},f1(t)=𝐛T()K0z^N0(t),f2(t)=𝐁K0z^N0(t),𝐁=col{𝐁1,𝐁2},𝐂=[𝐂0,𝐂1],f3(t)=[L00(NN0)×d][𝐂z^N(t)𝐜,z(,t)],{\scriptsize\begin{array}[]{ll}\frac{\mathrm{d}}{\mathrm{d}t}\xi(t)+\mathrm{diag}\{\mathcal{A},\mathcal{A}_{0}\}\xi(t)={\tiny\left[\begin{array}[]{cc}qz(\cdot,t)+f_{1}(t-\tau_{u})\\ f_{2}(t-\tau_{u})+f_{3}(t-\tau_{y})\end{array}\right]},\\ \mathcal{A}_{0}=\mathrm{diag}\{-A_{0},-A_{1}\},~{}f_{1}(t)=-{\bf b}^{\mathrm{T}}(\cdot)K_{0}\hat{z}^{N_{0}}(t),\\ f_{2}(t)=-{\bf B}K_{0}\hat{z}^{N_{0}}(t),~{}{\bf B}=\mathrm{col}\{{\bf B}_{1},{\bf B}_{2}\},~{}{\bf C}=[{\bf C}_{0},{\bf C}_{1}],\\ f_{3}(t)=-{\tiny\left[\begin{array}[]{cc}L_{0}\\ 0_{(N-N_{0})\times d}\end{array}\right]}[{\bf C}\hat{z}^{N}(t)-\langle{\bf c},z(\cdot,t)\rangle],\vspace{-0.32cm}\end{array}} (2.13)

where 𝒜\mathcal{A} is defined in (1.1). We begin with continuously differentiable delays. By using Theorems 6.1.2 and 6.1.5 in [28] together with the step method on intervals [0,t][0,t_{*}], [t,(s+1)τm][t_{*},(s+1)\tau_{m}], [(s+1)τm,(s+2)τm][(s+1)\tau_{m},(s+2)\tau_{m}], \dots, where ss\in\mathbb{N} satisfies sτmt<(s+1)τms\tau_{m}\leq t_{*}<(s+1)\tau_{m} (see arguments similar to the well-posedness in Section 3 of [14]), we obtain that for any initial value ξ(0)=[z0(),0]T𝒟(𝒜)×N\xi(0)=[z_{0}(\cdot),0]^{\mathrm{T}}\in\mathcal{D}(\mathcal{A})\times\mathbb{R}^{N}, the closed-loop system (2.13) has a unique classical solution

ξC([0,),L2(Ω)×N)C1([0,)\J,L2(Ω)×N),ξ(t)𝒟(𝒜)×N,t0,{\scriptsize\begin{array}[]{ll}\xi\in C([0,\infty),L^{2}(\Omega)\times\mathbb{R}^{N})\cap C^{1}([0,\infty)\backslash J,L^{2}(\Omega)\times\mathbb{R}^{N}),\\ \xi(t)\in\mathcal{D}(\mathcal{A})\times\mathbb{R}^{N},~{}~{}\forall t\geq 0,\vspace{-0.35cm}\end{array}} (2.14)

where J={t,(s+1)τm,(s+2)τm,}J=\{t_{*},(s+1)\tau_{m},(s+2)\tau_{m},\dots\}. The well-posedness for sawtooth delays follows similarly.

2.2 Stability analysis and main results

Let en(t)=zn(t)z^n(t)e_{n}(t)=z_{n}(t)-\hat{z}_{n}(t), 1nN1\leq n\leq N be the estimation error. The last term on the right-hand side of (2.7) can be presented as

n=1N𝐜nz^n(tτy)y(t)=n=1N𝐜nen(tτy)ζ(tτy),ζ(t)=n=N+1𝐜nzn(t).\begin{array}[]{ll}\sum_{n=1}^{N}{\bf c}_{n}\hat{z}_{n}(t-\tau_{y})-y(t)\\ =-\sum_{n=1}^{N}{\bf c}_{n}e_{n}(t-\tau_{y})-\zeta(t-\tau_{y}),\\ \zeta(t)=\sum_{n=N+1}^{\infty}{\bf c}_{n}z_{n}(t).\vspace{-0.3cm}\end{array} (2.15)

From (2.5), (2.7), (2.15), the error system has the form

e˙n(t)=(λn+q)en(t)lni=1N𝐜iei(tτy)lnζ(tτy),1nN.\begin{array}[]{ll}\dot{e}_{n}(t)=(-\lambda_{n}+q)e_{n}(t)-l_{n}\sum_{i=1}^{N}{\bf c}_{i}e_{i}(t-\tau_{y})\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-l_{n}\zeta(t-\tau_{y}),~{}1\leq n\leq N.\vspace{-0.25cm}\end{array} (2.16)

Denote

z^NN0(t)=[z^N0+1(t),,z^N(t)]T,eN0(t)=[e1(t),,eN0(t)]T,eNN0(t)=[eN0+1(t),,eN(t)]T,X0(t)=col{z^N0(t),eN0(t)},𝒦0=[K0,0d×N0],F0=[A0𝐁0K0L0𝐂00A0L0𝐂0],0=col{L0,L0},ντu(t)=z^N0(t)z^N0(tτu),0=col{𝐁0,0N0×d},ντy(t)=X0(t)X0(tτy),𝒞0=[𝐂0,0d×N0].{\scriptsize\begin{array}[]{ll}\hat{z}^{N-N_{0}}(t)=[\hat{z}_{N_{0}+1}(t),\dots,\hat{z}_{N}(t)]^{\mathrm{T}},\\ e^{N_{0}}(t)=[e_{1}(t),\dots,e_{N_{0}}(t)]^{\mathrm{T}},\\ e^{N-N_{0}}(t)=[e_{N_{0}+1}(t),\dots,e_{N}(t)]^{\mathrm{T}},\\ X_{0}(t)=\mathrm{col}\{\hat{z}^{N_{0}}(t),e^{N_{0}}(t)\},~{}~{}~{}~{}~{}\mathcal{K}_{0}=[K_{0},0_{d\times N_{0}}],\\ {\tiny F_{0}=\left[\begin{array}[]{cccc}A_{0}-{\bf B}_{0}K_{0}&L_{0}{\bf C}_{0}\\ 0&A_{0}-L_{0}{\bf C}_{0}\end{array}\right]},~{}\mathcal{L}_{0}=\mathrm{col}\{L_{0},-L_{0}\},\\ \nu_{\tau_{u}}(t)=\hat{z}^{N_{0}}(t)-\hat{z}^{N_{0}}(t-\tau_{u}),~{}~{}\mathcal{B}_{0}=\mathrm{col}\{{\bf B}_{0},0_{N_{0}\times d}\},\\ \nu_{\tau_{y}}(t)=X_{0}(t)-X_{0}(t-\tau_{y}),~{}~{}~{}~{}~{}\mathcal{C}_{0}=[{\bf C}_{0},0_{d\times N_{0}}].\vspace{-0.3cm}\end{array}} (2.17)

From (2.16), we have eNN0(t)=eA1teNN0(0)e^{N-N_{0}}(t)=\mathrm{e}^{A_{1}t}e^{N-N_{0}}(0). By (2.7), (2.12), (2.16) and substituting eNN0(tτy)=eA1τyeNN0(t)e^{N-N_{0}}(t-\tau_{y})=\mathrm{e}^{-A_{1}\tau_{y}}e^{N-N_{0}}(t), we obtain the reduced-order closed-loop system

X˙0(t)=F0X0(t)+0K0ντu(t)0𝒞0ντy(t)\displaystyle\dot{X}_{0}(t)=F_{0}X_{0}(t)+\mathcal{B}_{0}K_{0}\nu_{\tau_{u}}(t)-\mathcal{L}_{0}\mathcal{C}_{0}\nu_{\tau_{y}}(t) (2.18a)
+0ζ(tτy)+0𝐂1eA1τyeNN0(t),\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}+\mathcal{L}_{0}\zeta(t-\tau_{y})+\mathcal{L}_{0}\mathbf{C}_{1}\mathrm{e}^{-A_{1}\tau_{y}}e^{N-N_{0}}(t),
z˙n(t)=(λn+q)zn(t)𝐛nT𝒦0X0(tτy),n>N,\displaystyle\dot{z}_{n}(t)=(-\lambda_{n}+q)z_{n}(t)-{\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X_{0}(t-\tau_{y}),~{}n>N, (2.18b)

where ζ(t)\zeta(t) is defined in (2.15). Note that ζ(t)\zeta(t) does not depend on z^NN0(t)\hat{z}^{N-N_{0}}(t) which satisfies

z^˙NN0(t)=A1z^NN0(t)𝐁1𝒦0X0(tτu),\begin{array}[]{ll}\dot{\hat{z}}^{N-N_{0}}(t)=A_{1}\hat{z}^{N-N_{0}}(t)-{\bf B}_{1}\mathcal{K}_{0}X_{0}(t-\tau_{u}),\end{array}\vspace{-0.3cm} (2.19)

and is exponentially decaying (since A1A_{1} defined in (2.8) is stable due to (2.2)) provided X0(t)X_{0}(t) is exponentially decaying. Therefore, for stability of (2.1) under the control law (2.12), it is sufficient to show the stability of the reduced-order system (2.18). The latter can be considered as a singularly perturbed system with the slow sate X0(t)X_{0}(t) and the fast infinite-dimensional state zn(t)z_{n}(t), n>Nn>N.

For exponential L2L^{2}-stability of the closed-loop system (2.18), we consider the following vector Lyapunov functional

V(t)=[V0(t),Vtail(t)]T,Vtail(t)=n=N+1zn2(t),V0(t)=VP(t)+Vy(t)+Vu(t)+Ve(t),VP(t)=|X0(t)|P2,Ve(t)=pe|eNN0(t)|2,Vy(t)=tτM,yte2δ(st)|X0(s)|Sy2ds,+τM,yτM,y0t+θte2δ(st)|X˙0(s)|Ry2dsdθ,Vu(t)=tτM,ute2δ(st)|𝒦0X0(s)|Su2ds+τM,uτM,u0t+θte2δ(st)|𝒦0X˙0(s)|Ru2dsdθ,{\scriptsize\begin{array}[]{ll}V(t)=[V_{0}(t),~{}~{}V_{\mathrm{tail}}(t)]^{\mathrm{T}},V_{\mathrm{tail}}(t)=\sum_{n=N+1}^{\infty}z^{2}_{n}(t),\\ V_{0}(t)=V_{P}(t)+V_{y}(t)+V_{u}(t)+V_{e}(t),\\ V_{P}(t)=|X_{0}(t)|_{P}^{2},~{}~{}~{}V_{e}(t)=p_{e}|e^{N-N_{0}}(t)|^{2},\\ V_{y}(t)=\int_{t-\tau_{M,y}}^{t}\mathrm{e}^{2\delta(s-t)}|X_{0}(s)|^{2}_{S_{y}}\mathrm{d}s,\\ ~{}~{}~{}~{}~{}~{}~{}+\tau_{M,y}\int_{-\tau_{M,y}}^{0}\int_{t+\theta}^{t}\mathrm{e}^{2\delta(s-t)}|\dot{X}_{0}(s)|^{2}_{R_{y}}\mathrm{d}s\mathrm{d}\theta,\\ V_{u}(t)=\int_{t-\tau_{M,u}}^{t}\mathrm{e}^{2\delta(s-t)}|\mathcal{K}_{0}X_{0}(s)|^{2}_{S_{u}}\mathrm{d}s\\ ~{}~{}~{}~{}~{}~{}~{}+\tau_{M,u}\int_{-\tau_{M,u}}^{0}\int_{t+\theta}^{t}\mathrm{e}^{2\delta(s-t)}|\mathcal{K}_{0}\dot{X}_{0}(s)|^{2}_{R_{u}}\mathrm{d}s\mathrm{d}\theta,\vspace{-0.2cm}\end{array}} (2.20)

where 0<P,Sy,Ry2N0×2N00<P,S_{y},R_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}} and 0<Su,Rud×d0<S_{u},R_{u}\in\mathbb{R}^{d\times d}. Here Vy(t)V_{y}(t) is used to compensate ντy(t)\nu_{\tau_{y}}(t), Vu(t)V_{u}(t) is used to compensate ντu(t)\nu_{\tau_{u}}(t), and Ve(t)V_{e}(t) is used to compensate eNN0(t)e^{N-N_{0}}(t). To compensate ζ(tτy)\zeta(t-\tau_{y}) we will use vector Halanay’s inequality and the following Cauchy-Schwarz inequality:

|ζ(t)|2𝐜N2n=N+1zn2(t),𝐜N2:=j=1dcjN2=n=N+1|𝐜n|2.\begin{array}[]{ll}|\zeta(t)|^{2}\leq\|{\bf c}\|^{2}_{N}\sum_{n=N+1}^{\infty}z^{2}_{n}(t),\\ \|{\bf c}\|^{2}_{N}:=\sum_{j=1}^{d}\|c_{j}\|^{2}_{N}=\sum_{n=N+1}^{\infty}|{\bf c}_{n}|^{2}.\vspace{-0.3cm}\end{array} (2.21)

As explained in Remark 2.1 below, compared to the classical Halanay’s inequality, the vector one allows to use smaller δ\delta in VyV_{y} and VuV_{u} in the stability analysis essentially improving results in the numerical examples for comparatively large NN.

Differentiation of Vtail(t)V_{\mathrm{tail}}(t) along (2.18b) gives

V˙tail(t)=n=N+12(λn+q)zn2(t)n=N+12zn(t)𝐛nT𝒦0X(tτu).\begin{array}[]{ll}\dot{V}_{\mathrm{tail}}(t)=\sum_{n=N+1}^{\infty}2(-\lambda_{n}+q)z^{2}_{n}(t)\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-\sum_{n=N+1}^{\infty}2z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u}).\vspace{-0.2cm}\end{array} (2.22)

Let α>0\alpha>0. Applying Young’s inequality we arrive at

n=N+12zn(t)𝐛nT𝒦0X(tτu)𝐛N2αXT(tτu)𝒦0T𝒦0X(tτu)+αn=N+1zn2(t),𝐛N2:=i=1dbiN2.\begin{array}[]{ll}-\sum_{n=N+1}^{\infty}2z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u})\\ \leq\frac{\|{\bf b}\|^{2}_{N}}{\alpha}X^{\mathrm{T}}(t-\tau_{u})\mathcal{K}^{\mathrm{T}}_{0}\mathcal{K}_{0}X(t-\tau_{u})\\ ~{}~{}+\alpha\sum_{n=N+1}^{\infty}z^{2}_{n}(t),~{}~{}\|{\bf b}\|^{2}_{N}:=\sum_{i=1}^{d}\|b_{i}\|_{N}^{2}.\vspace{-0.25cm}\end{array} (2.23)

From (2.22) and (2.23), we have

V˙tail(t)+[2λN+12qα]Vtail(t)𝐛N2α|𝒦0X(tτu)|2βV0(tτu)\begin{array}[]{ll}\dot{V}_{\mathrm{tail}}(t)+[2\lambda_{N+1}-2q-\alpha]V_{\mathrm{tail}}(t)\\ \leq\frac{\|{\bf b}\|^{2}_{N}}{\alpha}|\mathcal{K}_{0}X(t-\tau_{u})|^{2}\leq\beta V_{0}(t-\tau_{u})\vspace{-0.2cm}\end{array} (2.24)

provided

𝐛N2α𝒦0T𝒦0<βP.\begin{array}[]{ll}\frac{\|{\bf b}\|^{2}_{N}}{\alpha}\mathcal{K}^{\mathrm{T}}_{0}\mathcal{K}_{0}<\beta P.\end{array} (2.25)

Let β0=αβ\beta_{0}=\alpha\beta. By Schur complement, we find that (2.25) holds iff

[P𝒦0Tβ0𝐛N2I]<0.\begin{array}[]{ll}{\scriptsize\left[\begin{array}[]{ccc}-P&\mathcal{K}_{0}^{\mathrm{T}}\\ &-{\beta_{0}\over\|{\bf b}\|^{2}_{N}}I\end{array}\right]}<0.\vspace{-0.01cm}\end{array} (2.26)

Let

εy=e2δτM,y,θτy(t)=eN0(tτy)eN0(tτM,y),εu=e2δτM,u,θτu(t)=z^N0(tτu)z^N0(tτM,u).{\scriptsize\begin{array}[]{ll}\varepsilon_{y}=e^{-2\delta\tau_{M,y}},~{}\theta_{\tau_{y}}(t)=e^{N_{0}}(t-\tau_{y})-e^{N_{0}}(t-\tau_{M,y}),\\ \varepsilon_{u}=e^{-2\delta\tau_{M,u}},~{}\theta_{\tau_{u}}(t)=\hat{z}^{N_{0}}(t-\tau_{u})-\hat{z}^{N_{0}}(t-\tau_{M,u}).\vspace{-0.3cm}\end{array}}

Differentiation of V0(t)V_{0}(t) along (2.18a) gives

V˙0(t)+2δV0(t)X0T(t)[PF0+F0TP+2δP]X0(t)+2X0T(t)P[0K0ντu(t)0𝒞0ντy(t)+0ζ(tτy)]+2X0T(t)P0𝐂1eA1τyeNN0(t)+|X0(t)|Sy2εy|X0(t)ντy(t)θτy(t)|Sy2+τM,y2|X˙0(t)|Ry2εyτM,ytτM,yt|X˙0(s)|Ry2ds+|𝒦0X0(t)|Su2εu|𝒦0X0(t)K0ντu(t)K0θτu(t)|Su2+τM,u2|𝒦0X˙0(t)|Ru2εuτM,utτM,ut|𝒦0X˙0(s)|Ru2ds+2pe(eNN0(t))T[A1+δI]eNN0(t).{\scriptsize\begin{array}[]{ll}\dot{V}_{0}(t)+2\delta V_{0}(t)\leq X_{0}^{\mathrm{T}}(t)[PF_{0}+F_{0}^{\mathrm{T}}P+2\delta P]X_{0}(t)\\ +2X_{0}^{\mathrm{T}}(t)P[\mathcal{B}_{0}K_{0}\nu_{\tau_{u}}(t)-\mathcal{L}_{0}\mathcal{C}_{0}\nu_{\tau_{y}}(t)+\mathcal{L}_{0}\zeta(t-\tau_{y})]\\ +2X_{0}^{\mathrm{T}}(t)P\mathcal{L}_{0}\mathbf{C}_{1}\mathrm{e}^{-A_{1}\tau_{y}}e^{N-N_{0}}(t)\\ +|X_{0}(t)|^{2}_{S_{y}}-\varepsilon_{y}|X_{0}(t)-\nu_{\tau_{y}}(t)-\theta_{\tau_{y}}(t)|^{2}_{S_{y}}\\ +\tau^{2}_{M,y}|\dot{X}_{0}(t)|^{2}_{R_{y}}-\varepsilon_{y}\tau_{M,y}\int_{t-\tau_{M,y}}^{t}|\dot{X}_{0}(s)|^{2}_{R_{y}}\mathrm{d}s\\ +|\mathcal{K}_{0}X_{0}(t)|^{2}_{S_{u}}-\varepsilon_{u}|\mathcal{K}_{0}X_{0}(t)-K_{0}\nu_{\tau_{u}}(t)-K_{0}\theta_{\tau_{u}}(t)|^{2}_{S_{u}}\\ +\tau^{2}_{M,u}|\mathcal{K}_{0}\dot{X}_{0}(t)|^{2}_{R_{u}}-\varepsilon_{u}\tau_{M,u}\int_{t-\tau_{M,u}}^{t}|\mathcal{K}_{0}\dot{X}_{0}(s)|^{2}_{R_{u}}\mathrm{d}s\\ +2p_{e}(e^{N-N_{0}}(t))^{\mathrm{T}}[A_{1}+\delta I]e^{N-N_{0}}(t).\vspace{-0.3cm}\end{array}} (2.27)

Let Gy2N0×2N0G_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}} and Gud×dG_{u}\in\mathbb{R}^{d\times d} satisfy

[RyGyRy]0,[RuGuRu]0.{\tiny\left[\begin{array}[]{cc}R_{y}&G_{y}\\ &R_{y}\end{array}\right]\geq 0,~{}\left[\begin{array}[]{cc}R_{u}&G_{u}\\ &R_{u}\end{array}\right]}\geq 0.\vspace{-0.3cm} (2.28)

Applying Jensen’s and Park’s inequalities (see, e.g., [7, Section 3.6.3]), we obtain for ξy(t)=col{ντy(t),θτy(t)}\xi_{y}(t)=\mathrm{col}\{\nu_{\tau_{y}}(t),\theta_{\tau_{y}}(t)\}, ξu(t)=col{K0ντu(t),K0θτu(t)}\xi_{u}(t)=\mathrm{col}\{K_{0}\nu_{\tau_{u}}(t),K_{0}\theta_{\tau_{u}}(t)\},

τM,ytτM,yt|X˙0(s)|Ry2dsξyT(t)[RyGyRy]ξy(t),τM,utτM,ut|𝒦0X˙0(s)|Ru2dsξuT(t)[RuGuRu]ξu(t).{\scriptsize\begin{array}[]{ll}-\tau_{M,y}\int_{t-\tau_{M,y}}^{t}|\dot{X}_{0}(s)|^{2}_{R_{y}}\mathrm{d}s\leq\par-\xi^{\mathrm{T}}_{y}(t){\tiny\left[\begin{array}[]{cc}R_{y}&G_{y}\\ &R_{y}\end{array}\right]}\xi_{y}(t),\\ -\tau_{M,u}\int_{t-\tau_{M,u}}^{t}|\mathcal{K}_{0}\dot{X}_{0}(s)|^{2}_{R_{u}}\mathrm{d}s\leq-\xi^{\mathrm{T}}_{u}(t){\tiny\left[\begin{array}[]{cc}R_{u}&G_{u}\\ &R_{u}\end{array}\right]}\xi_{u}(t).\vspace{-0.3cm}\end{array}} (2.29)

Let η(t)=col{X0(t),ζ(tτy),ξy(t),ξu(t),eNN0(t)}\eta(t)=\mathrm{col}\{X_{0}(t),\zeta(t-\tau_{y}),\xi_{y}(t),\xi_{u}(t),e^{N-N_{0}}(t)\}. Substituting (2.29) into (2.27), we get for δ1>0\delta_{1}>0,

V˙0(t)+2δV0(t)2δ1Vtail(tτy)(2.21)V˙X(t)+2δVX(t)2δ1𝐜N2|ζ(tτy)|2ηT(t)Φη(t)0\begin{array}[]{ll}\dot{V}_{0}(t)+2\delta V_{0}(t)-2\delta_{1}V_{\mathrm{tail}}(t-\tau_{y})\\ \overset{\eqref{CSineqaa}}{\leq}\dot{V}_{X}(t)+2\delta V_{X}(t)-\frac{2\delta_{1}}{\|{\bf c}\|^{2}_{N}}|\zeta(t-\tau_{y})|^{2}\\ ~{}\leq\eta^{\mathrm{T}}(t)\Phi\eta(t)\leq 0\vspace{-0.3cm}\end{array} (2.30)

provided

Φ=[Φ0P0𝐂1eA1τy2pe(A1+δI)]+ΛT[τM,y2Ry+τM,u2𝒦0TRu𝒦0]Λ0,\begin{array}[]{ll}\Phi={\small\left[\begin{array}[]{c|c}\Phi_{0}&P\mathcal{L}_{0}\mathbf{C}_{1}\mathrm{e}^{-A_{1}\tau_{y}}\\ \hline\cr*&2p_{e}(A_{1}+\delta I)\end{array}\right]}\\ ~{}~{}~{}~{}~{}~{}+\Lambda^{\mathrm{T}}[\tau_{M,y}^{2}R_{y}+\tau_{M,u}^{2}\mathcal{K}^{\mathrm{T}}_{0}R_{u}\mathcal{K}_{0}]\Lambda\leq 0,\vspace{-0.3cm}\end{array} (2.31)

where

Φ0=[Ω0P02δ1𝐜N2IΩ1εySy00Ω2εu𝒦0TSu00Ωy0Ωu],Ω0=PF0+F0TP+2δP+(1εy)Sy+(1εu)𝒦0TSu𝒦0,Ω1=εySyP0𝒞0,Ω2=P0+εu𝒦0TSu,{\scriptsize\begin{array}[]{ll}\Phi_{0}={\tiny\left[\begin{array}[]{c|c|c}\begin{array}[]{cc}\Omega_{0}&P\mathcal{L}_{0}\\ &-\frac{2\delta_{1}}{\|{\bf c}\|^{2}_{N}}I\end{array}&\begin{array}[]{cc}\Omega_{1}&\varepsilon_{y}S_{y}\\ 0&0\end{array}&\begin{array}[]{cc}\Omega_{2}&\varepsilon_{u}\mathcal{K}_{0}^{\mathrm{T}}S_{u}\\ 0&0\end{array}\\ \hline\cr*&\Omega_{y}&0\\ \hline\cr*&*&\Omega_{u}\par\end{array}\right]},\\ \Omega_{0}=PF_{0}+F_{0}^{\mathrm{T}}P+2\delta P+(1-\varepsilon_{y})S_{y}+(1-\varepsilon_{u})\mathcal{K}^{\mathrm{T}}_{0}S_{u}\mathcal{K}_{0},\\ \Omega_{1}=\varepsilon_{y}S_{y}-P\mathcal{L}_{0}\mathcal{C}_{0},~{}~{}~{}\Omega_{2}=P\mathcal{B}_{0}+\varepsilon_{u}\mathcal{K}_{0}^{\mathrm{T}}S_{u},\end{array}}
Λ=[Λ0,0𝐂1eA1τy],Λ0=[F0,0,0𝒞0,0,0,0],ΩJ=[εJ(SJ+RJ)εJ(SJ+GJ)εJ(SJ+RJ)],J{y,u}.{\scriptsize\begin{array}[]{ll}\Lambda=[\Lambda_{0},~{}\mathcal{L}_{0}\mathbf{C}_{1}\mathrm{e}^{-A_{1}\tau_{y}}],~{}\Lambda_{0}=[F_{0},\mathcal{L}_{0},-\mathcal{L}_{0}\mathcal{C}_{0},0,\mathcal{B}_{0},0],\\ \Omega_{J}={\left[\begin{array}[]{cc}-\varepsilon_{J}(S_{J}+R_{J})&~{}-\varepsilon_{J}(S_{J}+G_{J})\\ &~{}-\varepsilon_{J}(S_{J}+R_{J})\end{array}\right]},J\in\{y,u\}.\vspace{-0.3cm}\end{array}} (2.32)

We now show the feasibility of (2.31) for large NN. Since A1+δI<0A_{1}+\delta I<0 due to (2.2), by Schur complement for pep_{e}\rightarrow\infty, we obtain that the feasibility of (2.31) holds iff

Φ0+Λ0T[τM,y2Ry+τM,u2𝒦0TRu𝒦0]Λ00.\begin{array}[]{ll}\Phi_{0}+\Lambda_{0}^{\mathrm{T}}[\tau_{M,y}^{2}R_{y}+\tau_{M,u}^{2}\mathcal{K}^{\mathrm{T}}_{0}R_{u}\mathcal{K}_{0}]\Lambda_{0}\leq 0.\vspace{-0.3cm}\end{array} (2.33)

From (2.24) and (2.30), we have

V˙(t)[2δ002λN+1+2q+1α]V(t)+[02δ100]V(tτy)+[00β0]V(tτu).\begin{array}[]{ll}\dot{V}(t)\leq{\scriptsize\left[\begin{array}[]{cc}-2\delta&0\\ 0&-2\lambda_{N+1}+2q+\frac{1}{\alpha}\end{array}\right]}V(t)\\ ~{}~{}~{}~{}~{}~{}~{}+{\scriptsize\left[\begin{array}[]{cc}0&2\delta_{1}\\ 0&0\end{array}\right]}V(t-\tau_{y})+{\scriptsize\left[\begin{array}[]{cc}0&0\\ \beta&0\end{array}\right]}V(t-\tau_{u}).\vspace{-0.3cm}\end{array} (2.34)

By vector Halanay’s inequality (see Lemma 1.2) we have

|V(t)|De2δ0t,t0\begin{array}[]{ll}|V(t)|\leq D\mathrm{e}^{-2\delta_{0}t},~{}~{}t\geq 0\vspace{-0.3cm}\end{array} (2.35)

for some δ0>0\delta_{0}>0 and D>0D>0, provided

[2δ2δ1β2λN+1+2q+α]isHurwitz.\begin{array}[]{ll}{\footnotesize\left[\begin{array}[]{ccc}-2\delta&2\delta_{1}\\ \beta&-2\lambda_{N+1}+2q+\alpha\end{array}\right]}\end{array}\mathrm{is}~{}~{}\mathrm{Hurwitz}.\vspace{-0.3cm} (2.36)

By Parseval’s equality, we obtain from (2.35) that

z(,t)L22+z(,t)z^(,t)L22D~eδ0t,t0\|z(\cdot,t)\|^{2}_{L^{2}}+\|z(\cdot,t)-\hat{z}(\cdot,t)\|^{2}_{L^{2}}\leq\tilde{D}\mathrm{e}^{-\delta_{0}t},~{}t\geq 0\vspace{-0.3cm} (2.37)

for some D~>0\tilde{D}>0. Recalling that β0=αβ\beta_{0}=\alpha\beta, we find that (2.36) holds iff

2(λN+1q+δ)+α<0,[2α(λN+1q)+δ1δβ0α1]<0.\begin{array}[]{ll}-2(\lambda_{N+1}-q+\delta)+\alpha<0,\\ {\footnotesize\left[\begin{array}[]{ccc}-2\alpha(\lambda_{N+1}-q)+\frac{\delta_{1}}{\delta}\beta_{0}&\alpha\\ &-1\end{array}\right]}<0.\vspace{-0.25cm}\end{array} (2.38)

For asymptotic feasibility of LMIs (2.26), (2.28), (2.33), and (2.38) with large NN and small τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0, let Si=0S_{i}=0, Gi=0G_{i}=0 for i{y,u}i\in\{y,u\}. Taking τM,y,τM,u0+\tau_{M,y},\tau_{M,u}\rightarrow 0^{+}, it is sufficient to show (2.26), (2.38) and

[PF0+F0TP+2δPP0P0𝒞0P02δ1𝐜N2I00Ry0Ru]<0.\begin{array}[]{ll}{\tiny\left[\begin{array}[]{cc|cc}PF_{0}+F_{0}^{\mathrm{T}}P+2\delta P&P\mathcal{L}_{0}&-P\mathcal{L}_{0}\mathcal{C}_{0}&P\mathcal{B}_{0}\\ &-\frac{2\delta_{1}}{\|{\bf c}\|^{2}_{N}}I&0&0\\ \hline\cr*&*&-R_{y}&0\\ &*&*&-R_{u}\end{array}\right]}<0.\vspace{-0.3cm}\end{array} (2.39)

Take α=δ=1\alpha=\delta=1, δ1=β0=N13\delta_{1}=\beta_{0}=N^{\frac{1}{3}}, Ry=NIR_{y}=NI, Ru=NIR_{u}=NI. Let 0<P2N0×2N00<P\in\mathbb{R}^{2N_{0}\times 2N_{0}} be the solution to the Lyapunov equation P(F0+δI)+(F0+δI)TP=IP(F_{0}+\delta I)+(F_{0}+\delta I)^{\mathrm{T}}P=-I. We have P=O(1)\|P\|=O(1), NN\to\infty. Substituting above values into (2.26), (2.38), (2.39) and using Schur complement and the fact that λN=O(N)\lambda_{N}=O(N) (see Lemma 1.1), 0=O(1)\|\mathcal{L}_{0}\|=O(1), 0=O(1)\|\mathcal{B}_{0}\|=O(1) for NN\rightarrow\infty, we obtain the feasibility of (2.26), (2.38) and (2.39) for large enough NN. Fixing such NN and using continuity, we have that (2.26), (2.28), (2.31) and (2.38) are feasible for small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0. Summarizing, we arrive at:

Theorem 2.1.

Consider (2.1) with control law (2.12) and measurement (2.3). For δ>0\delta>0, let N0N_{0}\in\mathbb{N} satisfy (2.2) and NN\in\mathbb{N} satisfy NN0N\geq N_{0}. Let Assumption 1 hold and L0L_{0}, K0K_{0} be obtained from (2.11). Given τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0 and δ1>0\delta_{1}>0, let there exist 0<P,Sy,Ry2N0×2N00<P,S_{y},R_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}}, 0<Su,Rud×d0<S_{u},R_{u}\in\mathbb{R}^{d\times d}, Gy2N0×2N0G_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}}, Gud×dG_{u}\in\mathbb{R}^{d\times d} and scalars α,β0>0\alpha,\beta_{0}>0 such that LMIs (2.26), (2.28), (2.33) with Φ0\Phi_{0} and Λ0\Lambda_{0} given in (2.32), and (2.38) hold. Then the solution z(x,t)z(x,t) to (2.1) subject to the control law (2.7), (2.12) and the corresponding observer z^(x,t)\hat{z}(x,t) given by (2.6) satisfy (2.37) for some D~>0\tilde{D}>0 and δ0>0\delta_{0}>0. Moreover, LMIs (2.26), (2.28), (2.33), and (2.38) are always feasible for large enough NN and small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0.

Remark 2.1.

Multiplying decision variables PP, SiS_{i}, RiR_{i}, GiG_{i} (i{y,u}i\in\{y,u\}) in (2.26), (2.28), (2.33) by δ1\delta_{1} and changing β0\beta_{0} in (2.26) and (2.38) to β0δ1{\beta_{0}\over\delta_{1}}, we find that the feasibility of LMIs (2.26), (2.28), (2.33), and (2.38) is independent of δ1>0\delta_{1}>0. The fact also holds true for Theorems 3.1 and 4.1 below. This is different from the classical Halanay inequality (see Remark 2.3 below) where δ1δ\delta_{1}\leq\delta should not be small to compensate ζ(tτy)\zeta(t-\tau_{y}). However, compared to the classical Halanay inequality, the vector one needs constraint (2.25) (i.e., (2.26) which is usually more difficult to meet for larger N0N_{0}) whose feasibility requires 𝐛N2\|{\bf b}\|^{2}_{N} or 1α\frac{1}{\alpha} to be very small. This together with (2.38) implies that NN should be very large.

Remark 2.2.

Note that for N0>1N_{0}>1, it is difficult to find efficient L0L_{0}, K0K_{0} from (2.11) (see numerical example in Section 4). Here for N0>1N_{0}>1 we can use the following steps to find more efficient L0L_{0} and K0K_{0}:
Step 1: We find L0L_{0} from the following inequality:

[Po(A0L0𝐂0)+(A0L0𝐂0)TPo+2δPoPoLo2δ𝐜N2I]<0.\begin{array}[]{ll}{\tiny\left[\begin{array}[]{c|cc}P_{o}(A_{0}-L_{0}{\bf C}_{0})+(A_{0}-L_{0}{\bf C}_{0})^{\mathrm{T}}P_{o}+2\delta P_{o}&-P_{o}L_{o}\\ \hline\cr*&-\frac{2\delta}{\|{\bf c}\|^{2}_{N}}I\end{array}\right]<0}.\vspace{-0.2cm}{}{}{}{}{}{}{}\end{array} (2.40)

The additional terms compared to (2.11) are from the compensation of infinite-tail term of closed-loop system.
Step 2: Based on the L0L_{0} obtained from (2.40), we design the controller gain K0d×N0K_{0}\in\mathbb{R}^{d\times N_{0}} from the delay-free case (i.e., τu0\tau_{u}\equiv 0 and τy0\tau_{y}\equiv 0). In this case, the closed-loop system (2.18) becomes

X˙0(t)=F0X0(t)+0ζ(t)+0𝐂1eNN0(t),z˙n(t)=(λn+q)zn(t)Bn𝒦0X0(t),n>N.\begin{array}[]{ll}\dot{X}_{0}(t)=F_{0}X_{0}(t)+\mathcal{L}_{0}\zeta(t)+\mathcal{L}_{0}\mathbf{C}_{1}e^{N-N_{0}}(t),\\ \dot{z}_{n}(t)=(-\lambda_{n}+q)z_{n}(t)-B_{n}\mathcal{K}_{0}X_{0}(t),~{}n>N.\vspace{-0.3cm}\end{array}

We consider vector Lyapunov function

V(t)=[V0(t),Vtail(t)]T,V0(t)=|z^N0(t)|Pz2+|eN0(t)|Pe2+pe|eNN0(t)|2,{\scriptsize\begin{array}[]{ll}V(t)=[V_{0}(t),~{}V_{\mathrm{tail}}(t)]^{\mathrm{T}},\\ V_{0}(t)=|\hat{z}^{N_{0}}(t)|_{P_{z}}^{2}+|e^{N_{0}}(t)|_{P_{e}}^{2}+p_{e}|e^{N-N_{0}}(t)|^{2},\vspace{-0.3cm}\end{array}} (2.41)

where 0<Pz,PeN0×N00<P_{z},P_{e}\in\mathbb{R}^{N_{0}\times N_{0}}, pe>0p_{e}>0 and Vtail(t)V_{\mathrm{tail}}(t) is defined in (2.20). By arguments similar to (2.22)-(2.38), we have (2.37) for some D~>0\tilde{D}>0 provided

1αK0TΛbK0<βPz,[ΦzPzL0𝐂0PzL0ΦePeL02δ1cN2I]<0,Φz=Pz(A0𝐁0K0)+(A0𝐁0K0)TPz+2δPz,Φe=Pe(A0L0𝐂0)+(A0L0𝐂0)TPe+2δPe,2δ+2λN+12qα>0,δ(2λN+12qα)βδ1>0.{\scriptsize\begin{array}[]{ll}\frac{1}{\alpha}K_{0}^{\mathrm{T}}\Lambda_{b}K_{0}<\beta P_{z},~{}~{}{\tiny\left[\begin{array}[]{c|c|c}\Phi_{z}&P_{z}L_{0}{\bf C}_{0}&P_{z}L_{0}\\ \hline\cr*&\Phi_{e}&-P_{e}L_{0}\\ \hline\cr*&*&-\frac{2\delta_{1}}{\|c\|^{2}_{N}}I\end{array}\right]}<0,\\ \Phi_{z}=P_{z}(A_{0}-{\bf B}_{0}K_{0})+(A_{0}-{\bf B}_{0}K_{0})^{\mathrm{T}}P_{z}+2\delta P_{z},\\ \Phi_{e}=P_{e}(A_{0}-L_{0}{\bf C}_{0})+(A_{0}-L_{0}{\bf C}_{0})^{\mathrm{T}}P_{e}+2\delta P_{e},\\ 2\delta+2\lambda_{N+1}-2q-\alpha>0,\\ \delta(2\lambda_{N+1}-2q-\alpha)-\beta\delta_{1}>0.\vspace{-0.33cm}\end{array}} (2.42)

Let β0=αβ\beta_{0}=\alpha\beta, Qz=Pz1Q_{z}=P_{z}^{-1} and Yz=K0QzY_{z}=K_{0}Q_{z}. By Schur complement, we find that (2.42) hold iff

[QzYzTβ0bN2]<0,[Φ~zL0𝐂0L0ΦePeL02δ1cN2I]<0,Φ~z=A0Qz+QzA0TB0YzYzTB0T+2δQz,2(λN+1q+δ)+α<0,[2α(λN+1q)+δ1δβ0α1]<0.{\scriptsize\begin{array}[]{ll}{\footnotesize\left[\begin{array}[]{ccc}-Q_{z}&Y_{z}^{\mathrm{T}}\\ &-\frac{\beta_{0}}{\|b\|_{N}^{2}}\end{array}\right]}<0,~{}~{}{\tiny\left[\begin{array}[]{c|c|c}\tilde{\Phi}_{z}&L_{0}{\bf C}_{0}&L_{0}\\ \hline\cr*&\Phi_{e}&-P_{e}L_{0}\\ \hline\cr*&*&-\frac{2\delta_{1}}{\|c\|^{2}_{N}}I\par\end{array}\right]}<0,\\ \tilde{\Phi}_{z}=A_{0}Q_{z}+Q_{z}A_{0}^{\mathrm{T}}-B_{0}Y_{z}-Y_{z}^{\mathrm{T}}B_{0}^{\mathrm{T}}+2\delta Q_{z},\\ -2(\lambda_{N+1}-q+\delta)+\alpha<0,\\ \left[\begin{array}[]{cc}-2\alpha(\lambda_{N+1}-q)+\frac{\delta_{1}}{\delta}\beta_{0}&\alpha\\ &-1\end{array}\right]<0.\vspace{-0.3cm}\end{array}} (2.43)

In particular, (2.43) are LMIs that depend on decision variables 0<Qz,PeN0×N00<Q_{z},P_{e}\in\mathbb{R}^{N_{0}\times N_{0}}, Yzd×N0Y_{z}\in\mathbb{R}^{d\times N_{0}} and scalars α,β0>0\alpha,\beta_{0}>0. If LMIs (2.43) hold, the controller gain is given by K0=Qz1YzK_{0}=Q_{z}^{-1}Y_{z}.

Remark 2.3.

(Stability analysis via classical Halanay’s inequality) Consider Lyapunov functional

V(t)=V0(t)+Vtail(t)V(t)=V_{0}(t)+V_{\mathrm{tail}}(t)\vspace{-0.25cm} (2.44)

with V0(t)V_{0}(t) and Vtail(t)V_{\mathrm{tail}}(t) in (2.20). To compensate ζ(tτy)\zeta(t-\tau_{y}), the following bound is used for 0<δ1<δ0<\delta_{1}<\delta:

2δ1suptτM,yθtV(θ)2δ1[VP(tτy)+Vtail(tτy)](2.21)2δ1|X0(t)ντy(t)|P22δ1𝐜N2|ζ(tτy)|2.{\scriptsize\begin{array}[]{ll}-2\delta_{1}\sup\limits_{t-\tau_{M,y}\leq\theta\leq t}V(\theta)\leq-2\delta_{1}[V_{P}(t-\tau_{y})+V_{\mathrm{tail}}(t-\tau_{y})]\\ \overset{\eqref{CSineqaa}}{\leq}-2\delta_{1}|X_{0}(t)-\nu_{\tau_{y}}(t)|_{P}^{2}-\frac{2\delta_{1}}{\|{\bf c}\|^{2}_{N}}|\zeta(t-\tau_{y})|^{2}.\vspace{-0.35cm}\end{array}} (2.45)

By arguments similar to (2.22), (2.27)-(2.30), (2.45), and the following Young inequality for α1,α2>0\alpha_{1},\alpha_{2}>0,

n=N+12zn(t)𝐛nT𝒦0X(tτu)α1𝐛N2|𝒦0X0(t)|2+α2𝐛N2|K0ντu(t)|2+(1α1+1α2)n=N+1zn2(t),{\scriptsize\begin{array}[]{ll}-\sum_{n=N+1}^{\infty}2z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u})\\ \leq\alpha_{1}\|{\bf b}\|^{2}_{N}|\mathcal{K}_{0}X_{0}(t)|^{2}+\alpha_{2}\|{\bf b}\|^{2}_{N}|K_{0}\nu_{\tau_{u}}(t)|^{2}\\ ~{}~{}~{}+(\frac{1}{\alpha_{1}}+\frac{1}{\alpha_{2}})\sum_{n=N+1}^{\infty}z^{2}_{n}(t),\vspace{-0.35cm}\end{array}} (2.46)

we have

V˙(t)+2δV(t)2δ1suptτM,yθtV(θ)0\begin{array}[]{ll}\dot{V}(t)+2\delta V(t)-2\delta_{1}\sup_{t-\tau_{M,y}\leq\theta\leq t}V(\theta)\leq 0\vspace{-0.3cm}\end{array} (2.47)

provided (2.28) and the following inequalities hold:

[λN+1+q+δ11diag{2α1,2α2}]<0,Φ0+Λ0T[τM,y2Ry+τM,u2𝒦0TRu𝒦0]Λ0<0,\begin{array}[]{ll}{\scriptsize\left[\begin{array}[]{c|cc}-\lambda_{N+1}+q+\delta&1~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}1\\ \hline\cr*&\mathrm{diag}\{-2\alpha_{1},-2\alpha_{2}\}\end{array}\right]<0},\\ \Phi_{0}+\Lambda_{0}^{\mathrm{T}}[\tau_{M,y}^{2}R_{y}+\tau_{M,u}^{2}\mathcal{K}^{\mathrm{T}}_{0}R_{u}\mathcal{K}_{0}]\Lambda_{0}<0,\vspace{-0.3cm}\end{array} (2.48)

where Λ0\Lambda_{0} is defined in (2.32) and

Φ0=[Ω0P02δ1𝐜N2IΩ1εySy00P0+εu𝒦0TSuεu𝒦0TSu00Ωy0Ωu],Ω0=PF0+F0TP+2(δδ1)P+(1εu)𝒦0TSu𝒦0+(1εy)Sy+α1𝐛N2𝒦0T𝒦0,Ω1=2δ1PP0𝒞0+εySy,Ωy=[2δ1Pεy(Sy+Ry)εy(Sy+Gy)εy(Sy+Ry)],Ωu=[α2𝐛N2Iεu(Su+Ru)εu(Su+Gu)εu(Su+Ru)].{\scriptsize\begin{array}[]{ll}\Phi_{0}={\tiny\left[\begin{array}[]{c|c|c}\begin{array}[]{cc}\Omega_{0}&P\mathcal{L}_{0}\\ &-\frac{2\delta_{1}}{\|{\bf c}\|^{2}_{N}}I\end{array}&\begin{array}[]{cc}\Omega_{1}&\varepsilon_{y}S_{y}\\ 0&0\end{array}&\begin{array}[]{cc}P\mathcal{B}_{0}+\varepsilon_{u}\mathcal{K}_{0}^{\mathrm{T}}S_{u}&\varepsilon_{u}\mathcal{K}_{0}^{\mathrm{T}}S_{u}\\ 0&0\end{array}\\ \hline\cr*&\Omega_{y}&0\\ \hline\cr*&*&\Omega_{u}\end{array}\right]},\\ \Omega_{0}=PF_{0}+F_{0}^{\mathrm{T}}P+2(\delta-\delta_{1})P+(1-\varepsilon_{u})\mathcal{K}^{\mathrm{T}}_{0}S_{u}\mathcal{K}_{0}\\ ~{}~{}~{}~{}~{}~{}+(1-\varepsilon_{y})S_{y}+\alpha_{1}\|{\bf b}\|^{2}_{N}\mathcal{K}_{0}^{\mathrm{T}}\mathcal{K}_{0},\\ \Omega_{1}=2\delta_{1}P-P\mathcal{L}_{0}\mathcal{C}_{0}+\varepsilon_{y}S_{y},\\ \Omega_{y}={\tiny\left[\begin{array}[]{cc}-2\delta_{1}P-\varepsilon_{y}(S_{y}+R_{y})&~{}~{}-\varepsilon_{y}(S_{y}+G_{y})\\ &~{}~{}-\varepsilon_{y}(S_{y}+R_{y})\end{array}\right]},\\ \Omega_{u}={\tiny\left[\begin{array}[]{cc}\alpha_{2}\|{\bf b}\|^{2}_{N}I-\varepsilon_{u}(S_{u}+R_{u})&~{}~{}-\varepsilon_{u}(S_{u}+G_{u})\\ &~{}~{}-\varepsilon_{u}(S_{u}+R_{u})\end{array}\right]}.\vspace{-0.3cm}\end{array}} (2.49)

Then classical Halanay’s inequality (see P. 138 in [7]) and (2.47) imply (2.37), where δ0>0\delta_{0}>0 is the unique solution of δ0=δδ1e2δ0τM,y\delta_{0}=\delta-\delta_{1}\mathrm{e}^{2\delta_{0}\tau_{M,y}}.

3 Non-local actuation and boundary measurement

Consider system (2.1) with 𝐛(H1(Ω))d{\bf b}\in(H^{1}(\Omega))^{d}, 𝐛(x)=0{\bf b}(x)=0 for xΓDx\in\Gamma_{D}. Let N0N_{0}\in\mathbb{N} satisfy (2.2), NN0N\geq N_{0}, and dd be the maximum of the geometric multiplicities of λn\lambda_{n}, n=1,,N0n=1,\dots,N_{0}. We assume the following delayed boundary measurement:

y(t)=ΓN𝐜(x)z(x,tτy))dx,tτy0,y(t)=0,tτy<0,𝐜=[c1,,cd]T(L2(ΓN))d.{\scriptsize\begin{array}[]{ll}y(t)=\int_{\Gamma_{N}}{\bf c}(x)z(x,t-\tau_{y}))\mathrm{d}x,~{}t-\tau_{y}\geq 0,\\ y(t)=0,t-\tau_{y}<0,~{}{\bf c}=[c_{1},\dots,c_{d}]^{\mathrm{T}}\in(L^{2}(\Gamma_{N}))^{d}.\vspace{-0.3cm}\end{array}} (3.1)

Note that (3.1) is actually a weighted averaged boundary measurement with 𝐜{\bf c} representing the weighted coefficient. We present the solution to (2.1) as (2.4) with znz_{n} satisfying (2.5). We construct a NN-dimensional observer of the form (2.6), where z^n(t)\hat{z}_{n}(t) (1nN)(1\leq n\leq N) satisfy

z^˙n(t)=(λn+q)z^n(t)+bnu(t)ln[i=1N𝐜iz^i(tτy)y(t)],t>0,z^n(0)=0,t0,𝐜i=ΓN𝐜(x)ϕi(x)dx,\begin{array}[]{ll}\dot{\hat{z}}_{n}(t)=(-\lambda_{n}+q)\hat{z}_{n}(t)+b_{n}u(t)\\ ~{}~{}~{}~{}~{}~{}~{}-l_{n}[\sum_{i=1}^{N}{\bf c}_{i}\hat{z}_{i}(t-\tau_{y})-y(t)],~{}t>0,\\ \hat{z}_{n}(0)=0,~{}t\leq 0,~{}~{}{\bf c}_{i}=\int_{\Gamma_{N}}{\bf c}(x)\phi_{i}(x)\mathrm{d}x,\vspace{-0.35cm}\end{array} (3.2)

with y(t)y(t) in (3.1) and observer gains {ln}n=1N\{l_{n}\}_{n=1}^{N}, ln1×dl_{n}\in\mathbb{R}^{1\times d}. In this section, all notations are the same as in Sec. 2 except of 𝐜n{\bf c}_{n} which are defined in (3.2). Let 𝐁0{\bf B}_{0} and 𝐂0{\bf C}_{0} satisfy Assumption 1. From Lemma 2.1, we let L0=col{l1,,lN0}N0×dL_{0}=\mathrm{col}\{l_{1},\dots,l_{N_{0}}\}\in\mathbb{R}^{N_{0}\times d} satisfy (2.11a). Define u(t)u(t) in (2.12) with K0d×N0K_{0}\in\mathbb{R}^{d\times N_{0}} satisfying (2.11b). By (2.12), (2.15), (2.16), (3.2), and X0(t)X_{0}(t) defined in (2.17), we obtain the closed-loop system (2.18).

Note that we need (2.21) to compensate ζ(tτy)\zeta(t-\tau_{y}) in (2.18a) by Halanay inequality. However, differently from the non-local measurement where n=N+1|𝐜n|2<\sum_{n=N+1}^{\infty}|{\bf c}_{n}|^{2}<\infty, for the boundary measurement with 𝐜n{\bf c}_{n} defined in (3.2), we do not have this property. Here we assume

n=N+1|𝐜n|2λnϱNϱ,\begin{array}[]{ll}\sum_{n=N+1}^{\infty}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}\leq\varrho_{N}\leq\varrho,\vspace{-0.35cm}\end{array} (3.3)

for some ϱN>0\varrho_{N}>0, where ϱ>0\varrho>0 is independent of NN. For ζ(t)\zeta(t) defined in (2.15), by Cauchy-Schwarz inequality, we have

|ζ(t)|2n=N+1|𝐜n|2λnn=N+1λnzn2(t)(3.3)ϱNn=N+1λnzn2(t).\begin{array}[]{ll}|\zeta(t)|^{2}\leq\sum_{n=N+1}^{\infty}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t)\\ \overset{\eqref{eq3.5}}{\leq}\varrho_{N}\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t).\vspace{-0.35cm}\end{array} (3.4)
Remark 3.1.

Note that (3.3) holds for rectangular domain Ω=(0,a1)×(0,a2)\Omega=(0,a_{1})\times(0,a_{2}) with the following boundary

Γ=ΓDΓN,ΓN={(x1,0),x1(0,a1)},ΓD={(0,x2),x2[0,a2]}{(a1,x2),x2[0,a2]}{(x1,a2),x1[0,a1]}.{\scriptsize\begin{array}[]{ll}\Gamma=\Gamma_{D}\cup\Gamma_{N},~{}~{}\Gamma_{N}=\{(x_{1},0),~{}x_{1}\in(0,a_{1})\},\\ \Gamma_{D}=\{(0,x_{2}),x_{2}\in[0,a_{2}]\}\cup\{(a_{1},x_{2}),x_{2}\in[0,a_{2}]\}\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}\cup\{(x_{1},a_{2}),x_{1}\in[0,a_{1}]\}.\vspace{-0.3cm}\end{array}} (3.5)

The eigenvalues and corresponding eigenfunctions of 𝒜\mathcal{A} (see (1.1)) are given by:

λm,k=π2[m2a12+(k12)2a22],m,k,ϕm,k(x1,x2)=2a1a2sin(mπx1a1)cos((k12)πx2a2).\begin{array}[]{ll}\lambda_{m,k}=\pi^{2}[\frac{m^{2}}{a_{1}^{2}}+\frac{(k-\frac{1}{2})^{2}}{a_{2}^{2}}],~{}~{}m,k\in\mathbb{N},\\ \phi_{m,k}(x_{1},x_{2})=\frac{2}{\sqrt{a_{1}a_{2}}}\sin(\frac{m\pi x_{1}}{a_{1}})\cos(\frac{(k-\frac{1}{2})\pi x_{2}}{a_{2}}).\vspace{-0.25cm}\end{array} (3.6)

We reorder the eigenvalues (3.6) to form a non-decreasing sequence (1.2) and denote the corresponding eigenfunctions as {ϕn}n=1\{\phi_{n}\}_{n=1}^{\infty}. Let the corresponding relationship between (3.6) and (1.2) be n(m,k)n\sim(m,k). We have |𝐜n|2=|𝐜m,k|2=j=1d|0a1cj(x1)ϕm,k(x1,0)dx1|2=2a2j=1dcj,m2|{\bf c}_{n}|^{2}=|{\bf c}_{m,k}|^{2}=\sum_{j=1}^{d}|\int_{0}^{a_{1}}c_{j}(x_{1})\phi_{m,k}(x_{1},0)\mathrm{d}x_{1}|^{2}=\frac{2}{a_{2}}\sum_{j=1}^{d}c_{j,m}^{2} where cj,m=0a1cj(x1)2a1sin(mπx1a1)dx1c_{j,m}=\int_{0}^{a_{1}}c_{j}(x_{1})\cdot\frac{\sqrt{2}}{\sqrt{a_{1}}}\sin(\frac{m\pi x_{1}}{a_{1}})\mathrm{d}x_{1} satisfying cjL2(ΓN)2=m=1cj,m2\|c_{j}\|^{2}_{L^{2}(\Gamma_{N})}=\sum_{m=1}^{\infty}c_{j,m}^{2}. Therefore, we have

n=1|𝐜n|2λn=2a2j=1dm,k=1cj,m2λm,kj=1dm=1cj,m2k=12a2(k12)2π2=a2j=1dcjL2(0,a1)2=:ϱ,\begin{array}[]{ll}\sum_{n=1}^{\infty}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}=\frac{2}{a_{2}}\sum_{j=1}^{d}\sum_{m,k=1}^{\infty}\frac{c_{j,m}^{2}}{\lambda_{m,k}}\\ \leq\sum_{j=1}^{d}\sum_{m=1}^{\infty}c_{j,m}^{2}\sum_{k=1}^{\infty}\frac{2a_{2}}{(k-\frac{1}{2})^{2}\pi^{2}}\\ =a_{2}\sum_{j=1}^{d}\|c_{j}\|^{2}_{L^{2}(0,a_{1})}=:\varrho,\vspace{-0.3cm}\end{array} (3.7)

where k=11(2k1)2=π28\sum_{k=1}^{\infty}\frac{1}{(2k-1)^{2}}=\frac{\pi^{2}}{8} 111The proof can be found on this website: https://daviddeley.com/pendulum/page17proof.htm. is used and ϱ\varrho is independent of NN. From (3.7) it follows

n=N+1|𝐜n|2λnϱn=1N|𝐜n|2λn=:ϱN.\begin{array}[]{ll}\sum_{n=N+1}^{\infty}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}\leq\varrho-\sum_{n=1}^{N}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}=:\varrho_{N}.\vspace{-0.3cm}\end{array} (3.8)

Taking into account (3.4), for exponential H1H^{1}-stability we consider the vector Lyapunov functional (2.20) with Vtail(t)V_{\mathrm{tail}}(t) therein replaced by

Vtail(t)=n=N+1λnzn2(t).\begin{array}[]{ll}V_{\mathrm{tail}}(t)=\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t).\vspace{-0.3cm}\end{array} (3.9)

Differentiation of Vtail(t)V_{\mathrm{tail}}(t) in (3.9) along (2.18b) gives

V˙tail(t)=n=N+12(λn+q)λnzn2(t)n=N+12λnzn(t)𝐛nT𝒦0X(tτu)n=N+12(λn+q+α)λnzn2(t)+1α𝐛N2|𝒦0X(tτu)|2\begin{array}[]{ll}\dot{V}_{\mathrm{tail}}(t)=\sum_{n=N+1}^{\infty}2(-\lambda_{n}+q)\lambda_{n}z^{2}_{n}(t)\\ ~{}~{}~{}-\sum_{n=N+1}^{\infty}2\lambda_{n}z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u})\\ \leq\sum_{n=N+1}^{\infty}2(-\lambda_{n}+q+\alpha)\lambda_{n}z^{2}_{n}(t)\\ ~{}~{}~{}+\frac{1}{\alpha}\|\nabla{\bf b}\|^{2}_{N}|\mathcal{K}_{0}X(t-\tau_{u})|^{2}\vspace{-0.45cm}\end{array} (3.10)

for some α>0\alpha>0, where 𝐛N2=j=1dbjN2=(1.3)j=1dn=N+1λnbj,ϕn2\|\nabla{\bf b}\|^{2}_{N}=\sum_{j=1}^{d}\|\nabla b_{j}\|^{2}_{N}\overset{\eqref{inequality00}}{=}\sum_{j=1}^{d}\sum_{n=N+1}^{\infty}\lambda_{n}\langle b_{j},\phi_{n}\rangle^{2}. By arguments similar to (2.22)-(2.38) and using (3.4), (3.10), we obtain

z(,t)H12+z(,t)z^(,t)H12D~eδ0t,t0\|z(\cdot,t)\|^{2}_{H^{1}}+\|z(\cdot,t)-\hat{z}(\cdot,t)\|^{2}_{H^{1}}\leq\tilde{D}\mathrm{e}^{-\delta_{0}t},~{}t\geq 0\vspace{-0.3cm} (3.11)

for some D~>0\tilde{D}>0 and δ0>0\delta_{0}>0 provided LMIs (2.26) (where 𝐛N2\|{\bf b}\|^{2}_{N} is changed to 𝐛N2\|\nabla{\bf b}\|^{2}_{N}), (2.28), (2.31) with Φ0\Phi_{0} (where 𝐜N2\|{\bf c}\|^{2}_{N} is changed to ϱN\varrho_{N}) and Λ0\Lambda_{0} given in (2.32), and (2.38) hold. The asymptotic feasibility of above LMIs for large enough NN and small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0 can be obtained by arguments similar to Theorem 2.1. Summarizing, we arrive at:

Theorem 3.1.

Consider (2.1) with control law (2.12) where 𝐛(H1(Ω))d{\bf b}\in(H^{1}(\Omega))^{d}, 𝐛(x)=0{\bf b}(x)=0 for xΓDx\in\Gamma_{D}, measurement (3.1), and z0𝒟(𝒜)z_{0}\in\mathcal{D}(\mathcal{A}). Given δ,δ1>0\delta,\delta_{1}>0, let N0N_{0}\in\mathbb{N} satisfy (2.2) and NN\in\mathbb{N} satisfy NN0N\geq N_{0}. Let Assumption 1 and (3.3) hold and L0L_{0}, K0K_{0} be obtained from (2.11). Given τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0, let there exist 0<P,Sy,Ry2N0×2N00<P,S_{y},R_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}}, 0<Su,Rud×d0<S_{u},R_{u}\in\mathbb{R}^{d\times d}, scalars α,β0>0\alpha,\beta_{0}>0, Gy2N0×2N0G_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}} and Gud×dG_{u}\in\mathbb{R}^{d\times d} such that LMIs (2.26) (where 𝐛N2\|{\bf b}\|^{2}_{N} is changed to 𝐛N2\|\nabla{\bf b}\|^{2}_{N}), (2.28), (2.31) with Φ0\Phi_{0} (where 𝐜N2\|{\bf c}\|^{2}_{N} is changed to ϱN\varrho_{N}) and Λ0\Lambda_{0} given in (2.32), and (2.38) hold. Then the solution z(x,t)z(x,t) to (2.1) subject to the control law (2.7), (2.12) and the corresponding observer z^(x,t)\hat{z}(x,t) given by (2.6) satisfy (3.11). Moreover, the above LMIs always hold for large enough NN and small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0.

Remark 3.2.

(Stability analysis via classical Halanay’s inequality) Consider Lyapunov functional (2.44) with V0(t)V_{0}(t) in (2.20) and Vtail(t)V_{\mathrm{tail}}(t) in (3.9). By arguments similar to (2.22)-(2.38) and using following bound for 0<δ1<δ0<\delta_{1}<\delta:

2δ1suptτM,yθtV(θ)2δ1[VP(tτy)+Vtail(tτy)](3.4)2δ1|X0(t)ντy(t)|P22δ1ϱN|ζ(tτy)|2,{\scriptsize\begin{array}[]{ll}-2\delta_{1}\sup\limits_{t-\tau_{M,y}\leq\theta\leq t}V(\theta)\leq-2\delta_{1}[V_{P}(t-\tau_{y})+V_{\mathrm{tail}}(t-\tau_{y})]\\ \overset{\eqref{CSineq}}{\leq}-2\delta_{1}|X_{0}(t)-\nu_{\tau_{y}}(t)|_{P}^{2}-\frac{2\delta_{1}}{\varrho_{N}}|\zeta(t-\tau_{y})|^{2},\vspace{-0.1cm}\end{array}}

and the following Young inequality for α1,α2>0\alpha_{1},\alpha_{2}>0,

n=N+12λnzn(t)𝐛nT𝒦0X(tτu)α1𝐛N2|𝒦0X0(t)|2+α2𝐛N2|K0ντu(t)|2+(1α1+1α2)n=N+1λnzn2(t),{\scriptsize\begin{array}[]{ll}-\sum_{n=N+1}^{\infty}2\lambda_{n}z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u})\\ \leq\alpha_{1}\|\nabla{\bf b}\|^{2}_{N}|\mathcal{K}_{0}X_{0}(t)|^{2}+\alpha_{2}\|\nabla{\bf b}\|^{2}_{N}|K_{0}\nu_{\tau_{u}}(t)|^{2}\\ ~{}~{}~{}+(\frac{1}{\alpha_{1}}+\frac{1}{\alpha_{2}})\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t),\vspace{-0.3cm}\end{array}}

we obtain (3.11) provided (2.28) and (2.48) hold with Λ0\Lambda_{0} in (2.32) and Φ0\Phi_{0}, Ωy\Omega_{y}, Ωu\Omega_{u} in (2.49) (where 𝐛N2\|{\bf b}\|^{2}_{N} and 𝐜N2\|{\bf c}\|^{2}_{N} are changed to 𝐛N2\|\nabla{\bf b}\|^{2}_{N} and ϱN\varrho_{N}, respectively).

4 Boundary actuation and non-local measurement

Consider the delayed Neumann actuation

zt(x,t)=Δz(x,t)+qz(x,t),inΩ×(0,),z(x,t)=0,onΓD×(0,),z𝐧(x,t)=𝐛T(x)u(tτu),onΓN×(0,),z(x,0)=z0(x),xΩ,\begin{array}[]{ll}z_{t}(x,t)=\Delta z(x,t)+qz(x,t),~{}\mathrm{in}~{}\Omega\times(0,\infty),\\ z(x,t)=0,~{}~{}\mathrm{on}~{}\Gamma_{D}\times(0,\infty),\\ \frac{\partial z}{\partial{\bf n}}(x,t)={\bf b}^{\mathrm{T}}(x)u(t-\tau_{u}),~{}~{}\mathrm{on}~{}\Gamma_{N}\times(0,\infty),\\ z(x,0)=z_{0}(x),~{}~{}x\in\Omega,\vspace{-0.3cm}\end{array} (4.1)

where 𝐛=[b1,,bd]T(L2(ΓN))d{\bf b}=[b_{1},\dots,b_{d}]^{\mathrm{T}}\in(L^{2}(\Gamma_{N}))^{d} and u(t)=[u1(t),,ud(t)]Tu(t)=[u_{1}(t),\dots,u_{d}(t)]^{\mathrm{T}} is the control input to be designed. Let N0N_{0}\in\mathbb{N} satisfy (2.2) and NN0N\geq N_{0}. Let dd be the maximum of the geometric multiplicities of λn\lambda_{n}, n=1,,N0n=1,\dots,N_{0}. We consider the delayed non-local measurement (2.3) with 𝐜(L2(Ω))d{\bf c}\in(L^{2}(\Omega))^{d}. We present the solution to (4.1) as (2.4) and obtain (2.5) with

𝐛n=ΓD𝐛(x)ϕn(x)dx\begin{array}[]{ll}{\bf b}_{n}=\int_{\Gamma_{D}}{\bf b}(x)\phi_{n}(x)\mathrm{d}x\vspace{-0.35cm}\end{array} (4.2)

In this section, all notations are the same as in Sec. 2 except of 𝐛n{\bf b}_{n} that are defined by (4.2). We construct a NN-dimensional observer of the form (2.6), where z^n(t)\hat{z}_{n}(t) satisfy (2.7). Let 𝐁0{\bf B}_{0} and 𝐂0{\bf C}_{0} satisfy Assumption 1. From Lemma 2.1, let L0=col{l1,,lN0}N0×dL_{0}=\mathrm{col}\{l_{1},\dots,l_{N_{0}}\}\in\mathbb{R}^{N_{0}\times d} satisfy (2.11a). Define u(t)u(t) in (2.12) with K0d×N0K_{0}\in\mathbb{R}^{d\times N_{0}} satisfying (2.11b).

For the well-posedness of closed-loop system (4.1) and (2.7), with control input (2.12), we introduce the change of variables

w(x,t)=z(x,t)𝐫T(x)u(tτu),w(x,t)=z(x,t)-{\bf r}^{\mathrm{T}}(x)u(t-\tau_{u}),\vspace{-0.3cm} (4.3)

where 𝐫(x)=[r1(x),,rd(x)]T{\bf r}(x)=[r_{1}(x),\dots,r_{d}(x)]^{\mathrm{T}} with rj(x)r_{j}(x), j=1,,dj=1,\dots,d being the solution to the following Laplace equation:

Δrj(x)=0,xΩ,rj(x)=0,xΓD,rj𝐧(x)=bj(x),xΓN.\begin{array}[]{ll}\Delta r_{j}(x)=0,~{}~{}x\in\Omega,\\ r_{j}(x)=0,~{}x\in\Gamma_{D},~{}~{}\frac{\partial r_{j}}{\partial{\bf n}}(x)=b_{j}(x),~{}x\in\Gamma_{N}.\vspace{-0.3cm}\end{array} (4.4)

Since bjL2(ΓN)b_{j}\in L^{2}(\Gamma_{N}), from [6, Lemma 2.1] we have rjL2(Ω)r_{j}\in L^{2}(\Omega). By (4.1), (4.3), and (4.4), we get the equivalent evolution equation:

w˙(t)+𝒜w(t)=qw(t)𝐫T()u˙(tτu)(1τ˙u)+q𝐫T()u(tτu),w(0)=z(,0).\begin{array}[]{ll}\dot{w}(t)+\mathcal{A}w(t)=qw(t)-{\bf r}^{\mathrm{T}}(\cdot)\dot{u}(t-\tau_{u})(1-\dot{\tau}_{u})\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+q{\bf r}^{\mathrm{T}}(\cdot)u(t-\tau_{u}),\\ w(0)=z(\cdot,0).\vspace{-0.35cm}\end{array} (4.5)

Define the state ξ(t)=col{w(t),z^N(t)}\xi(t)=\mathrm{col}\{w(t),\hat{z}^{N}(t)\}, where z^N(t)=[z^1(t),,z^N(t)]T\hat{z}^{N}(t)=[\hat{z}_{1}(t),\dots,\hat{z}_{N}(t)]^{\mathrm{T}}. By (2.7), (2.12), and (4.5), we present the closed-loop system as

ddtξ(t)+diag{𝒜,𝒜0}ξ(t)=[qw(t)+f1(tτu)f2(tτu)+f3(tτy)0(NN0)×1],f3(t)=L0[𝐂z^N(t)𝐜,w(,t)+𝐜,𝐫T()K0z^N0(tτu)],f1(t)=𝐫T()(1τ˙u)K0[A0z^N0(t)+f3(tτy)B0K0z^N0(tτu)]q𝐫T()z^N0(t),{\scriptsize\begin{array}[]{ll}\frac{\mathrm{d}}{\mathrm{d}t}\xi(t)+\mathrm{diag}\{\mathcal{A},\mathcal{A}_{0}\}\xi(t)={\tiny\left[\begin{array}[]{ccc}qw(t)+f_{1}(t-\tau_{u})\\ f_{2}(t-\tau_{u})+f_{3}(t-\tau_{y})\\ 0_{(N-N_{0})\times 1}\end{array}\right]},\\ f_{3}(t)=-L_{0}[{\bf C}\hat{z}^{N}(t)-\langle{\bf c},w(\cdot,t)\rangle+\langle{\bf c},{\bf r}^{\mathrm{T}}(\cdot)K_{0}\hat{z}^{N_{0}}(t-\tau_{u})\rangle],\\ f_{1}(t)={\bf r}^{\mathrm{T}}(\cdot)(1-\dot{\tau}_{u})K_{0}[A_{0}\hat{z}^{N_{0}}(t)+f_{3}(t-\tau_{y})\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}-B_{0}K_{0}\hat{z}^{N_{0}}(t-\tau_{u})]-q{\bf r}^{\mathrm{T}}(\cdot)\hat{z}^{N_{0}}(t),\vspace{-0.3cm}\end{array}} (4.6)

where 𝒜0\mathcal{A}_{0}, 𝐂{\bf C}, and f2(t)f_{2}(t) are defined in (2.13). By arguments similar to the well-posedness in Sec. 2, we obtain that (4.6) has a unique solution satisfying (2.14). From (4.3), it follows (4.1), subject to (2.7), (2.12), has a unique classical solution such that zC([0,),L2(Ω))C1((0,),L2(Ω))z\in C([0,\infty),L^{2}(\Omega))\cap C^{1}((0,\infty),L^{2}(\Omega)) and z(,t)H2(Ω)z(\cdot,t)\in H^{2}(\Omega) with z(x,t)=0z(x,t)=0, xΓDx\in\Gamma_{D} and 𝐧z(x,t)=𝐛T(x)u(tτu)\frac{\partial}{\partial{\bf n}}z(x,t)={\bf b}^{\mathrm{T}}(x)u(t-\tau_{u}), xΓNx\in\Gamma_{N}, for t[0,)t\in[0,\infty).

With notations (2.17), the closed-loop system has a form:

X˙0(t)=F0X0(t)0𝒞ντy(t)+𝒦0ντu(t)+0ζ(tτy),z˙n(t)=(λn+q)zn(t)𝐛nT𝒦0X0(tτu),n>N.{\scriptsize\begin{array}[]{ll}\dot{X}_{0}(t)=F_{0}X_{0}(t)-\mathcal{L}_{0}\mathcal{C}\nu_{\tau_{y}}(t)+\mathcal{B}\mathcal{K}_{0}\nu_{\tau_{u}}(t)+\mathcal{L}_{0}\zeta(t-\tau_{y}),\\ \dot{z}_{n}(t)=(-\lambda_{n}+q)z_{n}(t)-{\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X_{0}(t-\tau_{u}),~{}n>N.\vspace{-0.35cm}\end{array}} (4.7)

For non-local actuation case in Sec. 2, we employ Young’s inequality (2.23) to split the finite- and infinite-dimensional parts, where n=N+1|𝐛n|2<\sum_{n=N+1}^{\infty}|{\bf b}_{n}|^{2}<\infty is used. However, for the boundary actuation with 𝐛n{\bf b}_{n} defined in (4.2), we do not have such property. Here we assume

n=N+1|𝐛n|2λnρNρ,\begin{array}[]{ll}\sum_{n=N+1}^{\infty}\frac{|{\bf b}_{n}|^{2}}{\lambda_{n}}\leq\rho_{N}\leq\rho,\vspace{-0.3cm}\end{array} (4.8)

for some ρN>0\rho_{N}>0, where ρ>0\rho>0 is independent of NN. Then we use the following Young inequality for α>0\alpha>0:

n=N+12zn(t)𝐛nT𝒦0X0(tτu)1αn=N+1|𝐛n|2λn|𝒦0X0(tτu)|2+n=N+1αλnzn2(t)(4.8)ρNα|𝒦0X0(tτu)|2+n=N+1αλnzn2(t).{\scriptsize\begin{array}[]{ll}-\sum_{n=N+1}2z_{n}(t){\bf b}_{n}^{\mathrm{T}}\mathcal{K}_{0}X_{0}(t-\tau_{u})\\ \leq{1\over\alpha}\sum^{\infty}_{n=N+1}\frac{|{\bf b}_{n}|^{2}}{\lambda_{n}}|\mathcal{K}_{0}X_{0}(t-\tau_{u})|^{2}+\sum_{n=N+1}^{\infty}\alpha\lambda_{n}z^{2}_{n}(t)\\ \overset{\eqref{eq4.2}}{\leq}{\rho_{N}\over\alpha}|\mathcal{K}_{0}X_{0}(t-\tau_{u})|^{2}+\sum_{n=N+1}^{\infty}\alpha\lambda_{n}z^{2}_{n}(t).\vspace{-0.4cm}\end{array}} (4.9)
Remark 4.1.

Note that (4.8) holds for rectangular domain. Consider the rectangular domain introduced in Remark 3.1. Similar to estimates (3.7) and (3.8), we have n=N+1|𝐛n|2λnρn=1N|𝐛n|2λn=:ρN\sum_{n=N+1}^{\infty}\frac{|{\bf b}_{n}|^{2}}{\lambda_{n}}\leq\rho-\sum_{n=1}^{N}\frac{|{\bf b}_{n}|^{2}}{\lambda_{n}}=:\rho_{N} with ρ=a2j=1dbjL2(0,a1)2\rho=a_{2}\sum_{j=1}^{d}\|b_{j}\|^{2}_{L^{2}(0,a_{1})} which is independent of NN.

According to (4.9), we consider the following Cauchy-Schwarz inequality:

|ζ(t)|2n=N+1|𝐜n|2λnn=N+1λnzn2(t)𝐜N2λNn=N+1λnzn2(t),\begin{array}[]{ll}|\zeta(t)|^{2}\leq\sum_{n=N+1}^{\infty}\frac{|{\bf c}_{n}|^{2}}{\lambda_{n}}\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t)\\ \leq\frac{\|{\bf c}\|^{2}_{N}}{\lambda_{N}}\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t),\vspace{-0.3cm}\end{array} (4.10)

where 𝐜N2\|{\bf c}\|^{2}_{N} is defined in (2.21). Consider the vector Lyapunov functional (2.20) with Vtail(t)V_{\mathrm{tail}}(t) therein replaced by (3.9). By arguments similar to (2.24)-(2.38), (3.10), and using (4.9) and (4.10), we conclude that the solutions to (4.1), (2.7), (2.12) satisfy (3.11) for some D~>0\tilde{D}>0 and δ0>0\delta_{0}>0 provided (2.28), (2.33) with Φ0\Phi_{0}, Λ0\Lambda_{0} in (2.32) (where 𝐜N2\|{\bf c}\|_{N}^{2} is changed to 1λN𝐜N2\frac{1}{\lambda_{N}}\|{\bf c}\|^{2}_{N}), and the following inequalities hold:

[P𝒦0Tβ0ρNI]<0,[2α(λN+1q)+δ1δβ0α1]<0.\begin{array}[]{ll}{\tiny\left[\begin{array}[]{ccc}-P&\mathcal{K}_{0}^{\mathrm{T}}\\ &-\frac{\beta_{0}}{\rho_{N}}I\end{array}\right]}<0,{\tiny\left[\begin{array}[]{ccc}-2\alpha(\lambda_{N+1}-q)+\frac{\delta_{1}}{\delta}\beta_{0}&\alpha\\ &-1\end{array}\right]}<0.\vspace{-0.2cm}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}\end{array} (4.11)

The asymptotic feasibility of above LMIs for large enough NN and small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0 can be obtained by arguments similar to Theorem 2.1. Summarizing, we have:

Theorem 4.1.

Consider (4.1) with control law (2.12) and delayed non-local measurement (2.3). Given δ>0\delta>0, let N0N_{0}\in\mathbb{N} satisfy (2.2) and NN\in\mathbb{N} satisfy NN0N\geq N_{0}. Let Assumption 1 hold and L0N0×dL_{0}\in\mathbb{R}^{N_{0}\times d}, K0d×N0K_{0}\in\mathbb{R}^{d\times N_{0}} be obtained from (2.11). Given τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0, let there exist 0<P,Sy,Ry2N0×2N00<P,S_{y},R_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}}, 0<Su,Rud×d0<S_{u},R_{u}\in\mathbb{R}^{d\times d}, Gy2N0×2N0G_{y}\in\mathbb{R}^{2N_{0}\times 2N_{0}} and Gud×dG_{u}\in\mathbb{R}^{d\times d}, scalars α,β0>0\alpha,\beta_{0}>0 such that LMIs (2.28), (2.31) with Φ0\Phi_{0} and Λ0\Lambda_{0} given in (2.32) (where 𝐜N2\|{\bf c}\|^{2}_{N} is changed to 𝐜N2/λN\|{\bf c}\|^{2}_{N}/\lambda_{N}) and (4.11) hold. Then the solution z(x,t)z(x,t) to (4.1) subject to the control law (2.7), (2.12) and the corresponding observer z^(x,t)\hat{z}(x,t) given by (2.6) satisfy (3.11) for some D~>0\tilde{D}>0 and δ0>0\delta_{0}>0. Moreover, the above inequalities always hold for large enough NN and small enough τM,y,τM,u>0\tau_{M,y},\tau_{M,u}>0.

Remark 4.2.

(Stability analysis via classical Halanay’s inequality) Consider Lyapunov functional (2.44) with V0(t)V_{0}(t) in (2.20) and Vtail(t)V_{\mathrm{tail}}(t) in (3.9). By arguments similar to (2.22)-(2.38) and using following bound for 0<δ1<δ0<\delta_{1}<\delta:

2δ1suptτM,yθtV(θ)2δ1[VP(tτy)+Vtail(tτy)](4.10)2δ1|X0(t)ντy(t)|P22δ1λN𝐜N|ζ(tτy)|2,{\scriptsize\begin{array}[]{ll}-2\delta_{1}\sup\limits_{t-\tau_{M,y}\leq\theta\leq t}V(\theta)\leq-2\delta_{1}[V_{P}(t-\tau_{y})+V_{\mathrm{tail}}(t-\tau_{y})]\\ \overset{\eqref{Cauchy-Schwarz22}}{\leq}-2\delta_{1}|X_{0}(t)-\nu_{\tau_{y}}(t)|_{P}^{2}-\frac{2\delta_{1}\lambda_{N}}{\|{\bf c}\|_{N}}|\zeta(t-\tau_{y})|^{2},\end{array}}

and the following Young inequality for α1,α2>0\alpha_{1},\alpha_{2}>0,

n=N+12λnzn(t)𝐛nT𝒦0X(tτu)(4.8)α1ρN|𝒦0X0(t)|2+α2ρN|K0ντu(t)|2+(1α1+1α2)n=N+1λnzn2(t),{\scriptsize\begin{array}[]{ll}-\sum_{n=N+1}^{\infty}2\lambda_{n}z_{n}(t){\bf b}^{\mathrm{T}}_{n}\mathcal{K}_{0}X(t-\tau_{u})\\ \overset{\eqref{eq4.2}}{\leq}\alpha_{1}\rho_{N}|\mathcal{K}_{0}X_{0}(t)|^{2}+\alpha_{2}\rho_{N}|K_{0}\nu_{\tau_{u}}(t)|^{2}\\ ~{}~{}~{}+(\frac{1}{\alpha_{1}}+\frac{1}{\alpha_{2}})\sum_{n=N+1}^{\infty}\lambda_{n}z^{2}_{n}(t),\vspace{-0.3cm}\end{array}}

we obtain (3.11) provided (2.28) and (2.48) hold with Λ0\Lambda_{0} in (2.32) and Φ0\Phi_{0}, Ωy\Omega_{y}, Ωu\Omega_{u} in (2.49) (where 𝐛N2\|{\bf b}\|^{2}_{N} and 𝐜N2\|{\bf c}\|^{2}_{N} are changed to ρN\rho_{N} and 𝐜N2/λN\|{\bf c}\|^{2}_{N}/\lambda_{N}, respectively).

5 Numerical examples

In this section, we consider a rectangular domain Ω=(0,a1)×(0,a2)\Omega=(0,a_{1})\times(0,a_{2}) with a1=433a_{1}=\frac{4\sqrt{3}}{3}, a2=433a_{2}=\frac{4\sqrt{3}}{3} and boundary (3.5). We consider q=3q=3 which results in an unstable open-loop system with 1 unstable mode (in this case, N0=1N_{0}=1 and d=1d=1) and q=8.1q=8.1 which results in an unstable open-loop system with 3 unstable modes with λ1<λ2=λ3\lambda_{1}<\lambda_{2}=\lambda_{3} (in this case, N0=3N_{0}=3 and d=2d=2), respectively. We consider three cases corresponding to Sections 2, 3 and 4. For all cases we take τM,y=τM,u=τM\tau_{M,y}=\tau_{M,u}=\tau_{M}. In each case, functions 𝐛=b1{\bf b}=b_{1}, 𝐜=c1{\bf c}=c_{1} for d=1d=1 and 𝐛=[b1,b2]T{\bf b}=[b_{1},b_{2}]^{\mathrm{T}}, 𝐜=[c1,c2]T{\bf c}=[c_{1},c_{2}]^{\mathrm{T}} for d=2d=2 are chosen according to Table 1, where

f1(x)=20x1(x2x22)χ[0,a12]×[0,a22](x),f2(x)=x1(x2x22)χ[a12,3a14]×[a22,a2](x),f3(x)=(x12a1x1)(x23a2x22),f4(x)=(x2a2)sin(2πx1a1),f5(x1)=sin(2x1πa1)χ[0,a12],f6(x1)=sin(3x1πa1)χ[a13,2a13],g1(x)=χ[0,a1]×[0,a22](x),g2(x)=χ[a12,a1]×[0,a2](x),g3(x1)=0.2χ[0,a14](x1),g4(x1)=0.2χ[a14,a1](x1).{\scriptsize\begin{array}[]{ll}f_{1}(x)=20x_{1}(x_{2}-x_{2}^{2})\chi_{[0,\frac{a_{1}}{2}]\times[0,\frac{a_{2}}{2}]}(x),\\ f_{2}(x)=x_{1}(x_{2}-x_{2}^{2})\chi_{[\frac{a_{1}}{2},\frac{3a_{1}}{4}]\times[\frac{a_{2}}{2},a_{2}]}(x),\\ f_{3}(x)=(x_{1}^{2}-a_{1}x_{1})(x_{2}^{3}-a_{2}x_{2}^{2}),~{}f_{4}(x)=(x_{2}-a_{2})\sin(\frac{2\pi x_{1}}{a_{1}}),\\ f_{5}(x_{1})=\sin(\frac{2x_{1}\pi}{a1})\chi_{[0,\frac{a_{1}}{2}]},~{}~{}~{}~{}~{}~{}~{}~{}f_{6}(x_{1})=\sin(\frac{3x_{1}\pi}{a1})\chi_{[\frac{a_{1}}{3},\frac{2a_{1}}{3}]},\\ g_{1}(x)=\chi_{[0,a_{1}]\times[0,\frac{a_{2}}{2}]}(x),~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}g_{2}(x)=\chi_{[\frac{a_{1}}{2},a_{1}]\times[0,a_{2}]}(x),\\ g_{3}(x_{1})=0.2\chi_{[0,\frac{a_{1}}{4}]}(x_{1}),~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}g_{4}(x_{1})=0.2\chi_{[\frac{a_{1}}{4},a_{1}]}(x_{1}).\vspace{-0.3cm}\end{array}}

Here χ\chi is an indicator function. We see that f1,f2,g1,g2L1(Ω)f_{1},f_{2},g_{1},g_{2}\in L^{1}(\Omega), f3,f4H1(Ω)f_{3},f_{4}\in H^{1}(\Omega), f3(x)=f4(x)=0f_{3}(x)=f_{4}(x)=0 for xΓDx\in\Gamma_{D}, and g3,g4,f5,f6L2(ΓN)g_{3},g_{4},f_{5},f_{6}\in L^{2}(\Gamma_{N}). It can be checked that for each case, Assumption 1 is satisfied.

Table 1: Chosen gains L0L_{0} and K0K_{0}.
q=3q=3, N0=1N_{0}=1 Thm 2.1 Thm 3.1 Thm 4.1
b1b_{1} f1f_{1} f3f_{3} f5f_{5}
c1c_{1} g1g_{1} g3g_{3} g1g_{1}
L0L_{0} from (2.11a) 1.6349 2.1837 4.1634
K0K_{0} from (2.11b) 1.2696 47.3821 1.6349
q=8.1q=8.1, N0=3N_{0}=3 Thm 2.1 Thm 3.1 Thm 4.1
b1,b2b_{1},b_{2} f1,f2f_{1},f_{2} f3,f4f_{3},f_{4} f5,f6f_{5},f_{6}
c1,c2c_{1},c_{2} g1,g2g_{1},g_{2} g3,g4g_{3},g_{4} g1,g2g_{1},g_{2}
L0L_{0} from (2.40) [8.4286.0360.2950.4240.2040.150]{\tiny\left[\begin{array}[]{ccc}8.428&6.036\\ -0.295&-0.424\\ 0.204&0.150\end{array}\right]} [9.96458.1530.1610.4160.9270.188]{\tiny\left[\begin{array}[]{ccc}9.964&58.153\\ 0.161&-0.416\\ 0.927&-0.188\end{array}\right]} [7.1084.8410.1330.5250.7090.085]{\tiny\left[\begin{array}[]{ccc}7.108&4.841\\ -0.133&-0.525\\ 0.709&0.085\end{array}\right]}
δ=0.04\delta=0.04 δ=0.02\delta=0.02 δ=0.05\delta=0.05
K0K_{0} from (2.43) [5.2600.0290.0340.0940.2530.097]{\tiny\left[\begin{array}[]{ccc}5.260&0.029&-0.034\\ -0.094&0.253&-0.097\end{array}\right]} [11.0330.0260000.040]{\tiny\left[\begin{array}[]{ccc}11.033&0.026&0\\ 0&0&-0.040\end{array}\right]} [7.8860.2800.3858.4440.0390.518]{\tiny\left[\begin{array}[]{ccc}7.886&-0.280&0.385\\ -8.444&0.039&0.518\end{array}\right]}

For the case that q=3q=3 and N0=1N_{0}=1, the gains L0L_{0} and K0K_{0} are found from (2.11) with δ=1\delta=1 and are given in Table 1. The LMIs of Theorems 2.1, 3.1, and 4.1 as well as their counterparts by classical Halanay’s inequality (Remarks 2.3, 3.2, and 4.2) were verified, respectively, for N=2,,8N=2,\dots,8 to obtain maximal values of τM\tau_{M} (δ=δ1>0\delta=\delta_{1}>0 is chosen optimally) that preserve the feasibility of LMIs. The results are given in Table 2. From Table 2, it is seen that the vector Halanay inequality always leads to larger delays than the classical scalar Halanay inequality.

Table 2: Max τM\tau_{M} for feasibility of LMIs (q=3q=3, N0=1N_{0}=1): Theorems 2.1, 3.1, 4.1 (vector Halanay’s inequality) VS Remarks 2.3, 3.2, 4.2 (classical scalar Halanay’s inequality).
NN 2 3 4 5 6 7 8
δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M}
Thm 2.1 0.35 0.237 0.12 0.292 0.1 0.303 0.07 0.311 0.05 0.318 0.05 0.39 0.04 0.323
Rmk 2.3 1 0.196 1 0.225 1 0.236 0.95 0.247 0.9 0.256 0.8 0.259 0.7 0.264
Thm 3.1 0.48 0.137 0.45 0.175 0.3 0.248 0.25 0.259 0.2 0.272
Rmk 3.2 3 0.033 2.5 0.041 1.2 0.107 1.1 0.123 1.08 0.141
Thm 4.1 0.18 0.276 0.06 0.312 0.06 0.319 0.03 0.323 0.03 0.328 0.02 0.329 0.02 0.331
Rmk 4.2 0.9 0.222 0.8 0.257 0.6 0.266 0.6 0.275 0.5 0.281 0.4 0.285 0.3 0.291

For the case that q=8.1q=8.1 and N0=3N_{0}=3, we found that the L0L_{0} and K0K_{0} obtained from (2.11) were not efficient for the feasibility of LMIs of Theorems 2.1, 3.1, 4.1 and Remarks 2.3, 3.2, 4.2 even for τM,y=τM,u=0\tau_{M,y}=\tau_{M,u}=0. We design L0L_{0} (δ=δ1=0.01\delta=\delta_{1}=0.01, N=20N=20) and K0K_{0} (N=30N=30) from (2.40) and (2.43) in Remark 2.2 and give the values in Table 1. The LMIs of Theorems 2.1, 3.1, and 4.1 as well as their counterparts by classical Halanay’s inequality (Remarks 2.3, 3.2, and 4.2) were verified, respectively, for different NN to obtain maximal values of τM\tau_{M} (δ=δ1>0\delta=\delta_{1}>0 is chosen optimally) that preserve the feasibility of LMIs. The results are given in Table 3. From Table 3, it is seen that the vector Halanay inequality leads to larger delays than the classical scalar one for comparatively large NN, whereas for comparatively small NN, the classical scalar Halanay inequality leads to larger delays. This phenomenon corresponds to Remark 2.1.

Table 3: Max τM\tau_{M} for feasibility of LMIs (q=8.1q=8.1, N0=3N_{0}=3): Theorems 2.1, 3.1, 4.1 (Vector Halanay’s inequality) VS Remarks 2.3, 3.2, 4.2 (Classical Scalar Halanay’s inequality).
NN 20 2525 3030 3535 4040 4545
δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M}
Thm 2.1 0.051 0.0104 0.049 0.0342 0.048 0.0414 0.047 0.0454 0.045 0.0481 0.045 0.0502
Rmk 2.3 4.5 0.0267 4 0.0330 3 0.0357 3 0.0376 2.8 0.0395 2.5 0.0407
NN 3030 3535 4040 4545 5050 5555
δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M}
Thm 3.1 0.019 0.0168 0.018 0.0219 0.018 0.0271 0.017 0.0301 0.017 0.0311 0.017 0.0337
Rmk 3.2 6 0.0206 6 0.0215 5 0.0230 4.5 0.0238 4 0.0240 4 0.0254
NN 7 8 9 10 15 20
δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δopt\delta_{{\mathrm{opt}}} τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M} δ\delta τM\tau_{M}
Thm 4.1 0.15 0.0112 0.15 0.0242 0.15 0.0291 0.14 0.0467 0.12 0.0506
Rmk 4.2 7 0.0036 6 0.0106 5 0.0136 5 0.0151 2.5 0.0254 2 0.0311

For simulation of closed-loop systems studied in Sections 2, 3 and 4, we consider the case q=3q=3, N0=1N_{0}=1 and fix N=5N=5. Consider time-varying delays τy(t)=τM2[1+sin2t]\tau_{y}(t)=\frac{\tau_{M}}{2}[1+\sin^{2}t] and τu(t)=τM2[1+cos2t]\tau_{u}(t)=\frac{\tau_{M}}{2}[1+\cos^{2}t] (corresponding maximal values of τM\tau_{M} are chosen as 0.311, 0.175, and 0.323, respectively according to Table 3). We approximate the solution norm using 150 modes as z(,t)L22n=1150zn2(t)\|z(\cdot,t)\|^{2}_{L^{2}}\approx\sum_{n=1}^{150}z_{n}^{2}(t) and z(,t)L22n=1150λnzn2(t)\|\nabla z(\cdot,t)\|^{2}_{L^{2}}\approx\sum_{n=1}^{150}\lambda_{n}z_{n}^{2}(t). Take initial conditions z0(x)=x1(a1x1)cos(π2a2x2)z_{0}(x)=x_{1}(a_{1}-x_{1})\cos(\frac{\pi}{2a_{2}}x_{2}). The closed-loop systems (with the tail ODEs truncated after 150 modes) are simulated using MATLAB. The simulations are presented in Fig. 1. The numerical simulations validate the theoretical results. Stability of the closed-loop systems in simulations was preserved for τM=0.48\tau_{M}=0.48 for Theorem 2.1, τM=0.38\tau_{M}=0.38 for Theorem 3.1, and τM=0.42\tau_{M}=0.42 for Theorem 4.1, which may indicate that our approach is somewhat conservative in this example.

Refer to caption

Figure 1: Evolutions z(,t)L22\|z(\cdot,t)\|^{2}_{L^{2}} (Theorem 2.1), z(,t)L22\|\nabla z(\cdot,t)\|^{2}_{L^{2}} (Theorem 3.1), and z(,t)L22\|z(\cdot,t)\|^{2}_{L^{2}} (Theorem 4.1) VS tt.

6 Conclusions

We considered the finite-dimensional observer-based control of 2D linear heat equation with fast-varying unknown input and known output delays. To compensate the output delay that appears in the infinite-dimensional part of the closed-loop system, we suggested a vector Lyapunov functional combined with vector Halanay’s inequality. In the numerical examples, the vector Halanay inequality led to larger delays for larger dimensions of the observer that preserve the stability than the classical one. Improvements and extension of the results to various high-dimensional PDEs may be topics for future research.

References

  • [1] N. B. Am and E. Fridman. Network-based H\infty filtering of parabolic systems. Automatica, 50(12):3139–3146, 2014.
  • [2] M. J. Balas. Finite-dimensional controllers for linear distributed parameter systems: exponential stability using residual mode filters. Journal of Mathematical Analysis and Applications, 133(2):283–296, 1988.
  • [3] V. Barbu. Boundary stabilization of equilibrium solutions to parabolic equations. IEEE Transactions on Automatic Control, 58(9):2416–2420, 2013.
  • [4] P. D. Christofides. Nonlinear and robust control of PDE systems: Methods and applications to transport-reaction processes. Springer, 2001.
  • [5] R. Curtain. Finite-dimensional compensator design for parabolic distributed systems with point sensors and boundary input. IEEE Transactions on Automatic Control, 27(1):98–104, 1982.
  • [6] H. Feng, P.-H. Lang, and J. Liu. Boundary stabilization and observation of a weak unstable heat equation in a general multi-dimensional domain. Automatica, 138:110152, 2022.
  • [7] E. Fridman. Introduction to time-delay systems: Analysis and control. Springer, 2014.
  • [8] Glitch. Poincaré inequality for a subspace of H1(ω){H}^{1}(\omega). Mathematics Stack Exchange. URL (version: 2021-06-15): https://math.stackexchange.com/q/2051099.
  • [9] L. Grüne and T. Meurer. Finite-dimensional output stabilization for a class of linear distributed parameter systems–a small-gain approach. Systems & Control Letters, 164:105237, 2022.
  • [10] C. Harkort and J. Deutscher. Finite-dimensional observer-based control of linear distributed parameter systems using cascaded output observers. International Journal of Control, 84(1):107–122, 2011.
  • [11] L. Jadachowski, T. Meurer, and A. Kugi. Backstepping observers for linear PDEs on higher-dimensional spatial domains. Automatica, 51:85–97, 2015.
  • [12] W. Kang and E. Fridman. Sampled-data control of 2-D Kuramoto–Sivashinsky Equation. IEEE Transactions on Automatic Control, 67(3):1314–1326, 2021.
  • [13] R. Katz and E. Fridman. Constructive method for finite-dimensional observer-based control of 1-D parabolic PDEs. Automatica, 122:109285, 2020.
  • [14] R. Katz and E. Fridman. Delayed finite-dimensional observer-based control of 1-D parabolic PDEs. Automatica, 123:109364, 2021.
  • [15] R. Katz and E. Fridman. Delayed finite-dimensional observer-based control of 1D parabolic PDEs via reduced-order LMIs. Automatica, 142:110341, 2022.
  • [16] R. Katz and E. Fridman. Sampled-data finite-dimensional boundary control of 1D parabolic PDEs under point measurement via a novel ISS Halanay’s inequality. Automatica, 135:109966, 2022.
  • [17] H. Lhachemi, I. Munteanu, and C. Prieur. Boundary output feedback stabilisation for 2-D and 3-D parabolic equations. arXiv preprint arXiv:2302.12460, 2023.
  • [18] H. Lhachemi and C. Prieur. Predictor-based output feedback stabilization of an input delayed parabolic pde with boundary measurement. Automatica, 137:110115, 2022.
  • [19] H. Lhachemi and C. Prieur. Boundary output feedback stabilisation of a class of reaction–diffusion pdes with delayed boundary measurement. International Journal of Control, 96(9):2285–2295, 2023.
  • [20] H. Lhachemi and R. Shorten. Boundary output feedback stabilization of state delayed reaction–diffusion pdes. Automatica, 156:111188, 2023.
  • [21] H. Lhachemi and R. Shorten. Output feedback stabilization of an ode-reaction–diffusion pde cascade with a long interconnection delay. Automatica, 147:110704, 2023.
  • [22] K. Liu and E. Fridman. Delay-dependent methods and the first delay interval. Systems & Control Letters, 64:57–63, 2014.
  • [23] X. Liu and C. Xie. Boundary control of reaction–diffusion equations on higher-dimensional symmetric domains. Automatica, 114:108832, 2020.
  • [24] F. Mazenc, M. Malisoff, and M. Krstic. Vector extensions of Halanay’s inequality. IEEE Transactions on Automatic Control, 67(3):1453–1459, 2022.
  • [25] Y. Meng and H. Feng. Boundary stabilization and observation of a multi-dimensional unstable heat equation. arXiv preprint arXiv:2203.12847, 2022.
  • [26] T. Meurer. Control of higher–dimensional PDEs: Flatness and backstepping designs. Springer Science & Business Media, 2012.
  • [27] I. Munteanu. Boundary stabilization of parabolic equations. Springer, 2019.
  • [28] A. Pazy. Semigroups of linear operators and applications to partial differential equations, volume 44. Springer Science & Business Media, 1983.
  • [29] J. Qi, R. Vazquez, and M. Krstic. Multi-agent deployment in 3-D via PDE control. IEEE Transactions on Automatic Control, 60(4):891–906, 2015.
  • [30] A. Selivanov and E. Fridman. Delayed H\infty control of 2D diffusion systems under delayed pointlike measurements. Automatica, 109:108541, 2019.
  • [31] W. A. Strauss. Partial differential equations: An introduction. John Wiley & Sons, 2007.
  • [32] M. Tucsnak and G. Weiss. Observation and control for operator semigroups. Springer Science & Business Media, 2009.
  • [33] R. Vazquez and M. Krstic. Explicit output-feedback boundary control of reaction-diffusion PDEs on arbitrary-dimensional balls. ESAIM: Control, Optimisation and Calculus of Variations, 22(4):1078–1096, 2016.
  • [34] J.-W. Wang and J.-M. Wang. Dynamic compensator design of linear parabolic MIMO PDEs in nn-dimensional spatial domain. IEEE Transactions on Automatic Control, 66(3):1399–1406, 2021.