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

Sliding Mode Observers for Set-valued Lur’e Systems with Uncertainties Beyond Observational Range

Samir Adly   Jun Huang  Ba Khiet Le Laboratoire XLIM, Université de Limoges, 123 Avenue Albert Thomas, 87060 Limoges CEDEX, FranceEmail: samir.adly@unilim.frSchool of Mechanical and Electrical Engineering, Soochow University, ChinaEmail: cauchyhot@163.comOptimization Research Group, Faculty of Mathematics and Statistics, Ton Duc Thang University, Ho Chi Minh City, VietnamE-mail: lebakhiet@tdtu.edu.vn
Abstract

In this paper, we introduce a new sliding mode observer for Lur’e set-valued dynamical systems, particularly addressing challenges posed by uncertainties not within the standard range of observation. Traditionally, most of Luenberger-like observers and sliding mode observer have been designed only for uncertainties in the range of observation. Central to our approach is the treatment of the uncertainty term which we decompose into two components: the first part in the observation subspace and the second part in its complemented subspace. We establish that when the second part converges to zero, an exact sliding mode observer for the system can be obtained. In scenarios where this convergence does not occur, our methodology allows for the estimation of errors between the actual state and the observer state. This leads to a practical interval estimation technique, valuable in situations where part of the uncertainty lies outside the observable range. Finally, we show that our observer is also a TT- observer as well as a strong HH^{\infty} observer.

Keywords. Sliding mode observer; set-valued Lur’e systems; uncertainties outside the
observation subspace

AMS Subject Classification. 28B05, 34A36, 34A60, 49J52, 49J53, 93D20

1 Introduction

Differential inclusion serves as an extension of ordinary differential equations, playing a pivotal role as a mathematical model within applied mathematics and control theory. A set-valued Lur’e dynamical system, a form of differential inclusion, incorporates interconnection feedback, broadening its scope to encompass various models frequently used in the analysis of nonsmooth dynamical systems. Such models include evolution variational inequalities, projected dynamical systems, relay systems, and complementarity systems. Set-valued Lur’e dynamical systems have been the subject of thorough investigations in recent decades (we refer, for example, to [1, 2, 3, 7, 8, 9, 12, 14, 20, 22, 27, 28] and the references therein). An observer is a mathematical model used to reconstruct complete information about the state variables from partially measurable data. Observers design for set-valued Lur’e systems represents a significant practical challenge, alongside other fundamental issues like well-posedness and stability analysis, which have received extensive attention in the literature. Numerous references addressing observer design for set-valued Lur’e systems are available, with a predominant focus on two methodological approaches: Luenberger-like techniques [11, 15, 16, 17, 27, 28] and sliding mode methodologies [4, 21].

The Luenberger-like observers for set-valued Lur’e systems was initially explored by Brogliato-Heemels in [11], which is an extension of the work of Arcak-Kokotovic for the nonlinear case [5]. Brogliato-Heemels considered the differential inclusion

{x˙=Ax+Bw+Gu,w(Cx),y=Fx,\left\{\begin{array}[]{lll}\;\dot{x}=Ax+Bw+Gu,\\ &&\\ \;w\in-\mathcal{F}(Cx),\\ &&\\ \;y=Fx,\end{array}\right. (1)

and proposed the observer

{x~˙=(ALF)x~+Bw~+Ly+Gu,w~((CKF)x~+Ky)\left\{\begin{array}[]{lll}\;\dot{\tilde{x}}=(A-LF)\tilde{x}+B\tilde{w}+Ly+Gu,\\ &&\\ \;\tilde{w}\in-\mathcal{F}\Big{(}(C-KF)\tilde{x}+Ky\Big{)}\\ \end{array}\right. (2)

where A,B,C,G,F,KA,B,C,G,F,K are matrices with appropriate dimensions, \mathcal{F} is a maximal monotone set-valued operator and yy is the partial measurable output. Later, it was extended by Huang et al in [15] to incorporate uncertainty as follows

{x˙=Ax+Bw+f1(x,u)+f2(x,u)θ,w(Cx),y=Fx,\left\{\begin{array}[]{lll}\;\dot{x}=Ax+Bw+f_{1}(x,u)+f_{2}(x,u)\theta,\\ &&\\ \;w\in-\mathcal{F}(Cx),\\ &&\\ \;y=Fx,\end{array}\right. (3)

where θ\theta is an unknown constant. The proposed observer is

{x~˙=(ALF)x~+Bw~+f1(x~,u)+f2(x~,u)θ~+Ly,w~((CKF)x~+Ky)θ~˙=h(x~,u)(yFx~),\left\{\begin{array}[]{lll}\;\dot{\tilde{x}}=(A-LF)\tilde{x}+B\tilde{w}+f_{1}(\tilde{x},u)+f_{2}(\tilde{x},u)\tilde{\theta}+Ly,\\ &&\\ \;\tilde{w}\in-\mathcal{F}\Big{(}(C-KF)\tilde{x}+Ky\Big{)}\\ &&\\ \dot{\tilde{\theta}}=h(\tilde{x},u)(y-F\tilde{x}),\\ \end{array}\right. (4)

where the range condition f2T(x,u)P=h(x,u)Ff_{2}^{T}(x,u)P=h(x,u)F is satisfied for some symmetric positive definite matrix PP. The key idea in [15] is to approximate the unknown by using an additional ODE, aiming to achieve observer convergence with an unknown rate. On the other hand, B. K. Le [21] proposed a sliding mode observer for a more general class of set-valued Lur’e systems, which extends the nonlinear case considered by Xiang-Su-Chu [29], see also [26] for a survey. Following this, Adly-Le applied the technique to system (3) in [4], offering clear advantages over Luenberger-like methods (4). These advantages include improved assumptions and achieving exponential convergence for the sliding mode observer without the need to solve additional ordinary differential equations. Moreover, the authors proposed a new, efficient smoothing approximation for the sliding mode technique to reduce the chattering effect using time guiding functions.

Nonetheless, a significant limitation of existing results using Luenberger-like or sliding mode observers (we refer, for example, to [4, 15, 21]) is the necessity of the range condition for uncertainties, meaning that uncertainties must fall within the range of P1FTP^{-1}F^{T}. An alternative approach involves constructing HH^{\infty} observers. However, HH^{\infty} observers only offer estimations for the entire process under zero initial conditions, rather than providing estimations at specific time instances (see, for instance, [17]). This limitation prompts us to introduce a sliding mode observer to attain time-specific estimations for the Lur’e system in the presence of uncertainties that do not fall within the observation range, as follows:

{x˙=Ax+Bω+f(x,u)+ξ(t),ω(Cx),y=Fx,\left\{\begin{array}[]{lll}\;\dot{x}=Ax+B\omega+f(x,u)+\xi(t),\\ &&\\ \;\omega\in-\mathcal{F}(Cx),\\ &&\\ \;y=Fx,\end{array}\right. (5)

where An×n,Bn×m,Cm×n,Fp×nA\in{\mathbb{R}}^{n\times n},B\in{\mathbb{R}}^{n\times m},C\in{\mathbb{R}}^{m\times n},F\in{\mathbb{R}}^{p\times n} are given matrix, xnx\in{\mathbb{R}}^{n} is the state variable, :mm\mathcal{F}:{\mathbb{R}}^{m}\rightrightarrows{\mathbb{R}}^{m} is a maximal monotone operator, uru\in{\mathbb{R}}^{r} is the control input and ypy\in{\mathbb{R}}^{p} is the measurable output. The nonlinear functions ff is Lipschitz continuous and ξl\xi\in{\mathbb{R}}^{l} is unknown. The model in (5) is quite general since in the right-hand side of the ODE, one has the linear AxAx, the nonlinear f(x,u)f(x,u), the set-valued B(Cx)B\mathcal{F}(Cx) and the general unknown ξ\xi. The idea is to decompose ξ\xi into 2 parts: ξ=ξ1+ξ2\xi=\xi_{1}+\xi_{2} where ξ1=P1FTξ\xi_{1}=P^{-1}F^{T}\xi is the projection of ξ\xi onto the observation subspace Im(P1FT){\rm Im}(P^{-1}F^{T}) and ξ2\xi_{2} is in the complement of the observation subspace. We show that if ξ2\xi_{2} is bounded, we can obtain a useful estimation for the state of the original system. Interval estimation is practical and unavoidable if the uncertainty ξ\xi is general and not in the range of observation. We refer here some interval observers results for some specific systems [18, 23]. In addition, if ξ2\xi_{2} tends to vanish when the time is large, we obtain the exact sliding mode observer, i.e., the observer state converges to the original state asymptotically. Finally, we show that our proposed observer is a TT- observer, a notion that we introduce to obtain a time instant estimation for the error. We also provide a mild additional condition such that our observer is a strong HH^{\infty} observer, an extension of HH^{\infty} observers which does not require the zero initial condition. TT- observers and strong HH^{\infty} observers allow us to obtain the total process and time instant estimations of the errors. In practice, there is a preference for the insights provided by the characteristics of strong HH^{\infty} observers over those of HH^{\infty} observers.

The paper is structured as follows: In Section 2, we revisit established definitions and essential results for our subsequent analyses. Section 3 presents our proposal for a sliding mode observer for the original system, which simultaneously functions as a TT- observer as well as a strong HH^{\infty} observer. In Section 4, we give several numerical examples to validate the theoretical findings. The paper concludes in Section 5, where we offer our closing remarks and explore potential future directions.

2 Notations and mathematical background

We denote the scalar product and the corresponding norm of Euclidean spaces by ,\langle\cdot,\cdot\rangle and \|\cdot\| respectively. A matrix Pn×nP\in{\mathbb{R}}^{n\times n} is called positive definite, written P>0P>0, if there exists a>0a>0 such that

Px,xax2,xn.\langle Px,x\rangle\geq a\|x\|^{2},\;\;\forall\;x\in{{\mathbb{R}}^{n}}.

The Sign functions in n{\mathbb{R}}^{n} is defined by

Sign(x)={xxifx0𝔹ifx=0,{\rm Sign}(x)=\left\{\begin{array}[]{l}\frac{x}{\|x\|}\;\;\;\;{\rm if}\;\;\;\;x\neq 0\\ \\ \mathbb{B}\;\;\;\;\;\;\;{\rm if}\;\;\;\;\;x=0,\end{array}\right.

where 𝔹\mathbb{B} denotes the closed unit ball in n{\mathbb{R}}^{n}.

A set-valued mapping :mm\mathcal{F}:{\mathbb{R}}^{m}\rightrightarrows{\mathbb{R}}^{m} is called 𝑚𝑜𝑛𝑜𝑡𝑜𝑛𝑒\it{monotone} if for all x,ymx,y\in{\mathbb{R}}^{m} and x(x),y(y)x^{*}\in\mathcal{F}(x),y^{*}\in\mathcal{F}(y), one has

xy,xy0.\langle x^{*}-y^{*},x-y\rangle\geq 0.

Furthermore, \mathcal{F} is called 𝑚𝑎𝑥𝑖𝑚𝑎𝑙𝑚𝑜𝑛𝑜𝑡𝑜𝑛𝑒\it{maximal\;monotone} if there is no monotone operator 𝒢\mathcal{G} such that the graph of \mathcal{F} is contained strictly in the graph of 𝒢\mathcal{G} (see, e.g, [6, 13]).

Let us recall a general version of Gronwall’s inequality [25].

Lemma 1.

Let T>0T>0 be given and a(),b()L1([0,T];)a(\cdot),b(\cdot)\in L^{1}([0,T];{\mathbb{R}}) with b(t)0b(t)\geq 0 for almost all t[0,T].t\in[0,T]. Let the absolutely continuous function w:[0,T]+w:[0,T]\to{\mathbb{R}}_{+} satisfy:

(1α)w(t)a(t)w(t)+b(t)wα(t),a.e.t[0,T],(1-\alpha)w^{\prime}(t)\leq a(t)w(t)+b(t)w^{\alpha}(t),\;\;a.e.\;t\in[0,T], (6)

where 0α<10\leq\alpha<1. Then for all t[0,T]t\in[0,T]:

w1α(t)w1α(0)exp(0ta(τ)𝑑τ)+0texp(sta(τ)𝑑τ)b(s)𝑑s.w^{1-\alpha}(t)\leq w^{1-\alpha}(0){\rm exp}\Big{(}\int_{0}^{t}a(\tau)d\tau\Big{)}+\int_{0}^{t}{\rm exp}\Big{(}\int_{s}^{t}a(\tau)d\tau\Big{)}b(s)ds. (7)

3 Sliding mode observer and the convergence analysis

In this section, we propose a sliding mode observer for the system (5) under the following assumptions:

Assumption 1: The set-valued operator :mm\mathcal{F}:{\mathbb{R}}^{m}\rightrightarrows{\mathbb{R}}^{m} is a maximal monotone operator.

Assumption 2: The functions f:n×rn,(x,u)f(x,u)f:{\mathbb{R}}^{n}\times{\mathbb{R}}^{r}\to{\mathbb{R}}^{n},\;{(x,u)\mapsto f(x,u)} is LfL_{f} - Lipschitz continuous w.r.t xx, i.e.:

f(x1,u)f(x2,u)Lfx1x2.\|f(x_{1},u)-f(x_{2},u)\|\leq L_{f}\|x_{1}-x_{2}\|.

Assumption 3: There exist ϵ>0\epsilon>0, Pn×n>0P\in{\mathbb{R}}^{n\times n}>0, Ln×pL\in{\mathbb{R}}^{n\times p} and Km×pK\in{\mathbb{R}}^{m\times p} such that

P(ALF)+(ALF)TP+2LfPI+2ϵI0,\displaystyle P(A-LF)+(A-LF)^{T}P+2L_{f}\|P\|I+2\epsilon I\leq 0, (8)
BTP=CKF.\displaystyle B^{T}P=C-KF. (9)

Assumption 4: The unknown ξ\xi is continuous and can be decomposed as ξ(t)=P1FTξ1(t)+ξ2(t)\xi(t)=P^{-1}F^{T}\xi_{1}(t)+\xi_{2}(t), where P1FTξ1(t)P^{-1}F^{T}\xi_{1}(t) is the projection of ξ\xi onto Im(P1FT){\rm Im}(P^{-1}F^{T}). The terms ξ1\xi_{1} and ξ2\xi_{2} are unknown, bounded by known continuous positive functions κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t) respectively, i.e., for all t0t\geq 0, we have

|ξ1(t)|κ1(t),|ξ2(t)|κ2(t).|\xi_{1}(t)|\leq\kappa_{1}(t),\;\;|\xi_{2}(t)|\leq\kappa_{2}(t).
Remark 1.

i) Assumption 4 is quite general, since it does not require the uncertainty in the range of P1FTP^{-1}F^{T}, i.e., the range of observation.
ii) Suppose that Assumption 4 holds, then we have FP1ξ2=0FP^{-1}\xi_{2}=0, i.e., ξ2\xi_{2} is in the kernel of FP1FP^{-1}.

The proposed sliding mode observer for (5) is

{x~˙=Ax~+Bω~Ley+f1(x~,u)P1FT(κ1Sign(ey)+κ3eyey2+δ),ω~(Cx~Key),y~=Fx~,\left\{\begin{array}[]{lll}\;\dot{\tilde{x}}=A\tilde{x}+B\tilde{\omega}-Le_{y}+f_{1}(\tilde{x},u)-P^{-1}F^{T}(\kappa_{1}{\rm Sign}(e_{y})+\frac{\kappa_{3}e_{y}}{\|e_{y}\|^{2}+\delta}),\\ &&\\ \;\tilde{\omega}\in-\mathcal{F}(C\tilde{x}-Ke_{y}),\\ &&\\ \;\tilde{y}=F\tilde{x},\end{array}\right. (10)

where

e:=x~x,ey:=Fe,e:=\tilde{x}-{x},\;\;e_{y}:=Fe, (11)

for some given small δ>0\delta>0. Since the nonlinear ff and the unknown ξ\xi are continuous, we can obtain the existence and uniqueness of the original system (5) and the approximate observer (10) under some additional mild conditions (see, e.g., [4, 11, 15]). In the following, we show that if ξ2\xi_{2} is bounded, we obtain a useful estimation for the error te(t)t\mapsto e(t). On the other hand, if ξ2\xi_{2} tends to vanish when the time is large, we have an exact observer indeed.

Theorem 2.

Suppose that Assumptions 1–4 hold. Let V(t)=Pe(t),e(t).V(t)=\langle Pe(t),e(t)\rangle. Then

V(t)V(0)exp(ϵtλmax)+Pλmin0texp(ϵ(st)λmax)κ2(t)𝑑s,\sqrt{V(t)}\leq\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\int_{0}^{t}{\rm exp}(\frac{\epsilon(s-t)}{\lambda_{max}}){\|\kappa_{2}(t)\|}ds, (12)

where λmax\lambda_{max} and λmin\lambda_{min} are the largest and smallest eigenvalues of PP respectively. In addition

  1. (a)

    If κ2()\kappa_{2}(\cdot) is bounded by some k>0k>0, then

    e(t)λmaxPkλminϵ+exp(ϵtλmax)(V(0)λminλmaxPkϵλmin).\|e(t)\|\leq\frac{\lambda_{max}\|P\|k}{{\lambda_{min}}\epsilon}+{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})(\sqrt{\frac{V(0)}{\lambda_{min}}}-\frac{\lambda_{max}\|P\|k}{\epsilon{\lambda_{min}}}). (13)

    Additionally, we introduce the attractive set Ω:=[0,kPϵ]\Omega:=[0,\frac{k\|P\|}{\epsilon}]. Consequently, the error norm e(t)\|e(t)\| converges to any neighborhood of Ω\Omega in finite time and remains within that vicinity.

  2. (b)

    If κ2(t)0\kappa_{2}(t)\to 0 as tt\to\infty, then e(t)0\|e(t)\|\to 0 as tt\to\infty, i.e., (10) is indeed an exact observer of (5).

  3. (c)

    If κ2(t)keat\|\kappa_{2}(t)\|\leq ke^{-at} for all t0t\geq 0, then e(t)e(t) converges exponentially to zero.

  4. (d)

    If κ2L2(0,)\kappa_{2}\in L^{2}(0,\infty), then

    V(t)\displaystyle\sqrt{V(t)} \displaystyle\leq V(0)exp(ϵtλmax)+Pλmin0tκ22(s)𝑑sλmax2ϵ(1exp(2ϵtλmax))\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\sqrt{\int_{0}^{t}\kappa^{2}_{2}(s)ds}\sqrt{\frac{\lambda_{max}}{2\epsilon}(1-{\rm exp}(\frac{-2\epsilon t}{\lambda_{max}}))}
    \displaystyle\leq V(0)exp(ϵtλmax)+Pκ2L2(0,)λminλmax2ϵ(1exp(2ϵtλmax)).\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|\|\kappa_{2}\|_{L^{2}(0,\infty)}}{\sqrt{\lambda_{min}}}\sqrt{\frac{\lambda_{max}}{2\epsilon}(1-{\rm exp}(\frac{-2\epsilon t}{\lambda_{max}}))}.
Proof.

From (5) and (10), we have

e˙\displaystyle\dot{e} \displaystyle\in (ALF)eB(ω~ω)+f(x~,u)f(x,u)\displaystyle(A-LF)e-B(\tilde{\omega}-{\omega})+f(\tilde{x},u)-f({x},u) (14)
\displaystyle- P1FT(κ1Sign(ey)+κ3eyey2+δ)(P1FTξ1(t)+ξ2(t)).\displaystyle P^{-1}F^{T}(\kappa_{1}{\rm Sign}(e_{y})+\frac{\kappa_{3}e_{y}}{\|e_{y}\|^{2}+\delta})-(P^{-1}F^{T}\xi_{1}(t)+\xi_{2}(t)).

With V(t)=Pe(t),e(t)V(t)=\langle Pe(t),e(t)\rangle, one has

12dVdt\displaystyle\frac{1}{2}\frac{dV}{dt} =\displaystyle= Pe˙,e=P(ALF)e+PB(ω~ω),e\displaystyle\langle P\dot{e},e\rangle=\langle P(A-LF)e+PB(\tilde{\omega}-{\omega}),e\rangle (15)
+\displaystyle+ P(f1(x~,u)f1(x,u)),e(κ1(t)ξ1(t))eyPξ2(t),e\displaystyle\langle P(f_{1}(\tilde{x},u)-f_{1}({x},u)),e\rangle-(\kappa_{1}(t)-\|\xi_{1}(t)\|){\|e_{y}\|}-\langle P\xi_{2}(t),e\rangle
\displaystyle- κ3ey2ey2+δ.\displaystyle\frac{\kappa_{3}\|e_{y}\|^{2}}{\|e_{y}\|^{2}+\delta}.

From (9) and the monotonicity of \mathcal{F}, we have

PB(ω~ω),e=ω~ω),BTPe=ω~ω,(CKF)e0.\displaystyle\langle PB(\tilde{\omega}-{\omega}),e\rangle=\langle\tilde{\omega}-{\omega}),B^{T}Pe\rangle=\langle\tilde{\omega}-{\omega},(C-KF)e\rangle\leq 0. (16)

On the other hand, using Assumption 2, we obtain

P(f1(x~,u)f1(x,u)),eL1Pe2.\displaystyle\langle P(f_{1}(\tilde{x},u)-f_{1}({x},u)),e\rangle\leq L_{1}\|P\|\|e\|^{2}. (17)

Thus from (8), (14)-(17), we deduce that

12dVdtϵe2+Pκ2(t)e(t)ϵλmaxV+Pκ2(t)λminV.\frac{1}{2}\frac{dV}{dt}\leq-\epsilon\|e\|^{2}+\|P\|\|\kappa_{2}(t)\|\|e(t)\|\leq-\frac{\epsilon}{\lambda_{max}}V+\frac{\|P\|\|\kappa_{2}(t)\|}{\sqrt{\lambda_{min}}}\sqrt{V}. (18)

Using Lemma 1 with α=1/2\alpha=1/2, we have

V(t)V(0)exp(ϵtλmax)+Pλmin0texp(ϵ(st)λmax)κ2(s)𝑑s.\sqrt{V(t)}\leq\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\int_{0}^{t}{\rm exp}(\frac{\epsilon(s-t)}{\lambda_{max}}){\|\kappa_{2}(s)\|}ds.

(a) If κ2()\kappa_{2}(\cdot) is bounded by some k>0k>0, then

λmine(t)V(t)\displaystyle\sqrt{\lambda_{min}}\|e(t)\|\leq\sqrt{V(t)} \displaystyle\leq V(0)exp(ϵtλmax)+Pkλmin0texp(ϵ(st)λmax)𝑑s.\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|k}{\sqrt{\lambda_{min}}}\int_{0}^{t}{\rm exp}(\frac{\epsilon(s-t)}{\lambda_{max}})ds.
=\displaystyle= V(0)exp(ϵtλmax)+λmaxPkϵλmin(1exp(ϵtλmax))\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\lambda_{max}\|P\|k}{\epsilon\sqrt{\lambda_{min}}}\big{(}1-{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})\big{)}
=\displaystyle= λmaxPkϵλmin+exp(ϵtλmax)(V(0)λmaxPkϵλmin).\displaystyle\frac{\lambda_{max}\|P\|k}{\epsilon\sqrt{\lambda_{min}}}+{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})(\sqrt{V(0)}-\frac{\lambda_{max}\|P\|k}{\epsilon\sqrt{\lambda_{min}}}).

Thus, one obtains (13)(\ref{estia}). On the other hand, from (18) we have

dVdtϵe2+kPe.\frac{dV}{dt}\leq-\epsilon\|e\|^{2}+k\|P\|\|e\|.

If ekPϵ+ρ\|e\|\geq\frac{k\|P\|}{\epsilon}+\rho for some ρ>0\rho>0, one has dVdtρϵe\frac{dV}{dt}\leq-\rho\epsilon\|e\|. Hence classically e\|e\| converges to any neighborhood of Ω\Omega in finite time and stay there.

(b) In particular, κ2()\kappa_{2}(\cdot) is bounded by some k>0k>0. Note that from (13)(\ref{estia}), we imply that

e(t)λmaxPkλminϵ+exp(ϵtλmax)V(0)λmin,t0.\|e(t)\|\leq\frac{\lambda_{max}\|P\|k}{{\lambda_{min}}\epsilon}+{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})\sqrt{\frac{V(0)}{\lambda_{min}}},\;\;\forall t\geq 0. (19)

In particular, e(t)e(t) is bounded and hence V(e(t))V(e(t)) is bounded by some VV^{*}. Let T>0T>0 be given, Tn=nTT_{n}=nT and

cn:=supt[Tn,+)|κ2(t)|,n1.c_{n}:=\sup_{t\in[T_{n},+\infty)}|\kappa_{2}(t)|,\;n\geq 1.

Then (cn)(c_{n}) is bounded by kk and converges to zero since κ2(t)0\kappa_{2}(t)\to 0 as tt\to\infty. Using (19), by replacing 0 by TnT_{n}, for all tTnt\geq T_{n}, we obtain

e(t)λmaxPcnλminϵ+exp(ϵ(tTn)λmax)Vλmin.\|e(t)\|\leq\frac{\lambda_{max}\|P\|c_{n}}{{\lambda_{min}}\epsilon}+{\rm exp}(\frac{-\epsilon(t-T_{n})}{\lambda_{max}})\sqrt{\frac{V^{*}}{\lambda_{min}}}. (20)

For given δ>0\delta>0, we can choose n0n_{0} such that λmaxPcn0λminϵδ\frac{\lambda_{max}\|P\|c_{n_{0}}}{{\lambda_{min}}\epsilon}\leq\delta. Then

e(t)δ+exp(ϵ(tTn0)λmax)Vλmin,tTn0.\|e(t)\|\leq\delta+{\rm exp}(\frac{-\epsilon(t-T_{n_{0}})}{\lambda_{max}})\sqrt{\frac{V^{*}}{\lambda_{min}}},\;\;\forall\;t\geq T_{n_{0}}. (21)

Thus

lim supte(t)δ.\limsup_{t\to\infty}\|e(t)\|\leq\delta.

Since δ\delta is arbitrary, we must have

lim supte(t)0,\limsup_{t\to\infty}\|e(t)\|\leq 0,

and thus

limte(t)=0.\lim_{t\to\infty}\|e(t)\|=0.

(c) If κ2(t)keat\|\kappa_{2}(t)\|\leq ke^{-at} for all t0t\geq 0, then

V(t)\displaystyle\sqrt{V(t)} \displaystyle\leq V(0)exp(ϵtλmax)+Pkλmin0texp(ϵ(st)λmaxas)𝑑s\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|k}{\sqrt{\lambda_{min}}}\int_{0}^{t}{\rm exp}(\frac{\epsilon(s-t)}{\lambda_{max}}-as)ds (22)
=\displaystyle= V(0)exp(ϵtλmax)+λmaxPkλmin(ϵaλmax)(eatexp(ϵtλmax)),\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\lambda_{max}\|P\|k}{\sqrt{\lambda_{min}}(\epsilon-a\lambda_{max})}(e^{-at}-{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})), (23)

if ϵaλmax\epsilon\neq a\lambda_{max}. Thus the error e(t)e(t) converges to zero with exponential rate. The case ϵ=aλmax\epsilon=a\lambda_{max} is trivial.

(d) If k2L2(0,)k_{2}\in L^{2}(0,\infty), then from (12), one has

V(t)\displaystyle\sqrt{V(t)} \displaystyle\leq V(0)exp(ϵtλmax)+Pλmin0tκ22(s)𝑑s0texp(2ϵ(st)λmax)𝑑s\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\sqrt{\int_{0}^{t}\kappa^{2}_{2}(s)ds}\sqrt{\int_{0}^{t}{\rm exp}(\frac{2\epsilon(s-t)}{\lambda_{max}})ds} (24)
=\displaystyle= V(0)exp(ϵtλmax)+Pλmin0tκ22(s)𝑑sλmax2ϵ(1exp(2ϵtλmax))\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\sqrt{\int_{0}^{t}\kappa^{2}_{2}(s)ds}\sqrt{\frac{\lambda_{max}}{2\epsilon}(1-{\rm exp}(\frac{-2\epsilon t}{\lambda_{max}}))}
\displaystyle\leq V(0)exp(ϵtλmax)+Pκ2L2(0,)λminλmax2ϵ(1exp(2ϵtλmax)).\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|\|\kappa_{2}\|_{L^{2}(0,\infty)}}{\sqrt{\lambda_{min}}}\sqrt{\frac{\lambda_{max}}{2\epsilon}(1-{\rm exp}(\frac{-2\epsilon t}{\lambda_{max}}))}.

The proof is thereby completed. ∎

Remark 2.

i) In the first case, we have the interval estimation for the error, i.e., the original state x(t)x(t) belongs to some known interval centered at the approximate observer x~(t)\tilde{x}(t). In the remaining cases, one obtains the exact observer. It covers the case where the unknown ξ\xi vanishes after some time instant T0>0T_{0}>0 as considered in some examples in [17]. Note that with our observer, we only require that the unknown part ξ2\xi_{2} in the complement of the observation space tends to vanish (see Figure 3, Example 2).
ii) The term κ3eyey2+δ\frac{\kappa_{3}e_{y}}{\|e_{y}\|^{2}+\delta} makes eye_{y} small very fast and remains small. Note that if ey2δ\|e_{y}\|^{2}\geq\delta, then ey2ey2+δ12\frac{e^{2}_{y}}{\|e_{y}\|^{2}+\delta}\geq\frac{1}{2}. Thus, if we choose κ3(t)>(Pκ2(t)+ρ)22ϵ\kappa_{3}(t)>\frac{(\|P\|\|\kappa_{2}(t)\|+\rho)^{2}}{2\epsilon} for some ρ>0\rho>0 and ey2δ\|e_{y}\|^{2}\geq\delta, then

12dVdt\displaystyle\frac{1}{2}\frac{dV}{dt} =\displaystyle= Pe˙,e=P(ALF)e+PB(ω~ω),e+P(f1(x~,u)f1(x,u)),e\displaystyle\langle P\dot{e},e\rangle=\langle P(A-LF)e+PB(\tilde{\omega}-{\omega}),e\rangle+\langle P(f_{1}(\tilde{x},u)-f_{1}({x},u)),e\rangle (25)
\displaystyle- (κ1(t)ξ1(t))eyκ3(t)ey2ey2+δPξ2(t),e\displaystyle(\kappa_{1}(t)-\|\xi_{1}(t)\|){\|e_{y}\|}-\frac{\kappa_{3}(t)e^{2}_{y}}{\|e_{y}\|^{2}+\delta}-\langle P\xi_{2}(t),e\rangle
\displaystyle\leq ϵe2κ3(t)2+Pκ2(t)eρe.\displaystyle-\epsilon\|e\|^{2}-\frac{\kappa_{3}(t)}{2}+\|P\|\|\kappa_{2}(t)\|\|e\|\leq-\rho\|e\|.

Then e(t)\|e(t)\| decreases very fast in finite time such that ey(t)2<δ\|e_{y}(t)\|^{2}<\delta.
iii) Suppose that Pk2()Pk_{2}(\cdot) is bounded by kk^{\prime}, then we can improve the set Ω\Omega by the new attractive set Ω=[0,kϵ]\Omega^{\prime}=[0,\frac{k^{\prime}}{\epsilon}].
iv) The same result is obtained if the unknown ξ(t)\xi(t) is replaced by ξ(t,x,u)\xi(t,x,u).

Let us recall the definition of HH^{\infty} observer (see, e.g., [17]) and introduce the notion of strong HH^{\infty} observer as well as TT-observer.

Definition 1.

If the error te(t)t\mapsto e(t) is asymptotically stable for ξ0\xi\equiv 0 and eL2(0,)μξL2(0,)\|e\|_{L^{2}(0,\infty)}\leq\mu\|\xi\|_{L^{2}(0,\infty)} for some μ>0\mu>0 under zero initial condition (i.e., e(0)=0e(0)=0) for non-zero ξL2(0,)\xi\in L^{2}(0,\infty), then (10) is said an HH^{\infty} observer for the system (5).

Definition 2.

The observer (10) is called a strong HH^{\infty} observer for the system (5) if the error te(t)t\mapsto e(t) is asymptotically stable for ξ0\xi\equiv 0 and for non-zero ξL2(0,)\xi\in L^{2}(0,\infty), there exists μ>0\mu>0 such that

eL2(0,)C+μξL2(0,),\|e\|_{L^{2}(0,\infty)}\leq C+\mu\|\xi\|_{L^{2}(0,\infty)}, (26)

where CC is a non-negative constant that vanishes under zero initial condition.

Definition 3.

The observer (10) is called a TT- observer for the system (5) if the error te(t)t\mapsto e(t) is asymptotically stable for ξ0\xi\equiv 0 and for non-zero ξL2(0,)\xi\in L^{2}(0,\infty), there exists μ>0\mu>0 such that

e(t)γ(t)+μ0tξ2(s)𝑑s,t0,\|e(t)\|\leq\gamma(t)+\mu\sqrt{\int_{0}^{t}\xi^{2}(s)ds},\;\;\forall\;t\geq 0, (27)

where γ(t)\gamma(t) vanishes under zero initial condition and converges exponentially to zero for the remaining case.

Remark 3.

It is easy to see that a strong HH^{\infty} observer is also an HH^{\infty} observer while TT- observer provides a time instant estimation for the error.

Assumption 5: There exists some μ>0\mu>0 such that

(ΩPP2μϵI)0\displaystyle\left(\begin{array}[]{ccc}\Omega&\;\;-P\\ \\ -P&\;\;-2\mu\epsilon I\end{array}\right)\leq 0 (31)

where Ω:=P(ALF)+(ALF)TP+2LfPI+2ϵI\Omega:=P(A-LF)+(A-LF)^{T}P+2L_{f}\|P\|I+2\epsilon I and ϵ>0\epsilon>0 is in Assumption 4.

Theorem 3.

Suppose that Assumptions 1–4 hold. Then (10) is a TT - observer for the system (5). In addition, if Assumption 5 holds, then (10) is a strong HH^{\infty} observer for the system (5).

Proof.

From Theorem 2, if the uncertainty ξ0\xi\equiv 0, then the error te(t)t\mapsto e(t) converges to zero exponentially. For non-zero ξL2(0,)\xi\in L^{2}(0,\infty), using Theorem 2-d, we have

λmine(t)\displaystyle\lambda_{min}\|e(t)\| \displaystyle\leq V(t)V(0)exp(ϵtλmax)+Pλmin0tξ22(s)𝑑sλmax2ϵ(1exp(2ϵtλmax))\displaystyle\sqrt{V(t)}\leq\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\sqrt{\int_{0}^{t}\xi^{2}_{2}(s)ds}\sqrt{\frac{\lambda_{max}}{2\epsilon}(1-{\rm exp}(\frac{-2\epsilon t}{\lambda_{max}}))} (32)
\displaystyle\leq V(0)exp(ϵtλmax)+Pλminλmax2ϵ0tξ2(s)𝑑s,\displaystyle\sqrt{V(0)}{\rm exp}(\frac{-\epsilon t}{\lambda_{max}})+\frac{\|P\|}{\sqrt{\lambda_{min}}}\sqrt{\frac{\lambda_{max}}{2\epsilon}}\sqrt{\int_{0}^{t}\xi^{2}(s)ds},

which deduces that (10) is a TT - observer for the system (5). If Assumption 5 is satisfied, from (15)-(17), with the same V(t)=Pe(t),e(t)V(t)=\langle Pe(t),e(t)\rangle, we have

12dVdt\displaystyle\frac{1}{2}\frac{dV}{dt} =\displaystyle= Pe˙,e\displaystyle\langle P\dot{e},e\rangle (33)
=\displaystyle= P(ALF)e+PB(ω~ω),e\displaystyle\langle P(A-LF)e+PB(\tilde{\omega}-{\omega}),e\rangle
+\displaystyle+ P(f1(x~,u)f1(x,u)),eκ1(t)eyκ3ey2ey2+δPξ,e\displaystyle\langle P(f_{1}(\tilde{x},u)-f_{1}({x},u)),e\rangle-\kappa_{1}(t){\|e_{y}\|}-\frac{\kappa_{3}\|e_{y}\|^{2}}{\|e_{y}\|^{2}+\delta}-\langle P\xi,e\rangle
\displaystyle\leq P(ALF)e,e+LfPe2Pξ,e\displaystyle\langle P(A-LF)e,e\rangle+L_{f}\|P\|\|e\|^{2}-\langle P\xi,e\rangle
\displaystyle\leq ϵ(e2+μξ2).\displaystyle\epsilon(-\|e\|^{2}+\mu\|\xi\|^{2}).

Integrating both side from 0 to \infty and note that VV is non-negative, we obtain that eL2(0,)V(0)2ϵ+μξL2(0,)\|e\|_{L^{2}(0,\infty)}\leq\frac{V(0)}{2\epsilon}+\mu\|\xi\|_{L^{2}(0,\infty)}. It means that (10) is a strong HH^{\infty} observer for the system (5). ∎

Remark 4.

i) Assumption 5 implies that Ω=P(ALF)+(ALF)TP+2LfPI+2ϵI0\Omega=P(A-LF)+(A-LF)^{T}P+2L_{f}\|P\|I+2\epsilon I\leq 0, which is consistent to Assumption 4. In Ω\Omega, we use the term ϵI\epsilon I instead of II as in [17], which is easier to have the non-positiveness of Ω\Omega.
ii) The positive number μ\mu exists, for example, if Ω2ϵI\Omega\leq-2\epsilon I, we can choose μ>P24ϵ2\mu>\frac{\|P\|^{2}}{4\epsilon^{2}}.

4 Numerical examples

Example 1.

First we consider the system (5) with

A=(6  4  078  00  07),B=(463),f(x,u)=(u+2sinx22u+3cosx1u+4sinx3)A=\left(\begin{array}[]{ccc}-6&\;\;4&\;\;0\\ \\ 7&\;\;-8&\;\;0\\ \\ 0&\;\;0&\;\;-7\end{array}\right),B=\left(\begin{array}[]{ccc}4\\ \\ 6\\ \\ -3\end{array}\right),f(x,u)=\left(\begin{array}[]{ccc}u+2\sin x_{2}\\ \\ 2u+3\cos x_{1}\\ \\ -u+4\sin x_{3}\end{array}\right)
C=(8  63),F=(1  0  0),u=5sint.C=\left(\begin{array}[]{ccc}8&\;\;6&\;\;-3\end{array}\right),F=\left(\begin{array}[]{ccc}1&\;\;0&\;\;0\end{array}\right),\;\;u=5\sin t.

Suppose that the unknown ξ(t,x,u)=(25cosx14sint)\xi(t,x,u)=\left(\begin{array}[]{ccc}2\\ \\ 5\cos x_{1}\\ \\ 4\sin t\end{array}\right) and

(x)={sign(x)(3|x|+6) if x0,[6,6] if x=0.\mathcal{F}(x)=\left\{\begin{array}[]{lll}{\rm sign}(x)(3|x|+6)&\mbox{ if }&x\neq 0,\\ &&\\ \;[-6,6]&\mbox{ if }&x=0.\end{array}\right.

Then L=4L=4 and k=41k=\sqrt{41}. Then Assumptions 1–4 are satisfied with

P=(1  0  00  1  00  0  1),L=(0110),ϵ=2,K=4.P=\left(\begin{array}[]{ccc}1&\;\;0&\;\;0\\ \\ 0&\;\;1&\;\;0\\ \\ 0&\;\;0&\;\;1\end{array}\right),\;\;\;L=\left(\begin{array}[]{ccc}0\\ \\ 11\\ \\ 0\end{array}\right),\;\;\;\epsilon=2,\;\;\;K=4.

One can observe that the total error e\|e\| converges to any neighborhood of the set Ω=[0,41/2]\Omega=[0,\sqrt{41}/2] in finite time while the observed error e1e_{1} tends to zero very fast. Consequently, the observer state x~\tilde{x} using (10) provides a good estimation for the original state xx.

Refer to caption
Figure 1: Errors using the approximate sliding mode observer (10)
Example 2.

Next let us consider the rotor system with friction as in [10, 19]. Let x1=θuθlx_{1}=\theta_{u}-\theta_{l}, x2=θ˙ux_{2}=\dot{\theta}_{u} and x3=θ˙lx_{3}=\dot{\theta}_{l} where θu\theta_{u} and θl\theta_{l} are the angular positions of the upper and lower discs, respectively, we have

x˙1=x2x3x˙2=kmJuukθJux11JuTfu(x2)x˙3=kθJlx11JlTfl(x3),\begin{split}\dot{x}_{1}&=x_{2}-x_{3}\\ \dot{x}_{2}&=\frac{k_{m}}{J_{u}}u-\frac{k_{\theta}}{J_{u}}x_{1}-\frac{1}{J_{u}}T_{fu}(x_{2})\\ \dot{x}_{3}&=\frac{k_{\theta}}{J_{l}}x_{1}-\frac{1}{J_{l}}T_{fl}(x_{3}),\end{split} (34)

where uu is the input voltage to the power amplifier of the motor, Tfu(x2)=bupx2T_{fu}(x_{2})=b_{up}x_{2} and

Tfl(x3)={[Tsl+T1(121+eω1|x3|)+T2(121+eω1|x3|)]sign(x3)+blx3,x30[Tsl,Tsl],x3=0.T_{fl}(x_{3})=\left\{\begin{array}[]{lr}$[$T_{sl}+T_{1}(1-\frac{2}{1+e^{\omega_{1}|x_{3}|}})\\ +T_{2}(1-\frac{2}{1+e^{\omega_{1}|x_{3}|}})$]${\rm sign}(x_{3})+b_{l}x_{3},&x_{3}\neq 0\\ $[$-T_{sl},\leavevmode\nobreak\ T_{sl}$]$,&x_{3}=0.\end{array}\right. (35)

Then (34) can be written as

x˙1=x2x3x˙2=kmJuukθJux1bupJux2x˙3=kθJlx11JlTfl(x3).\begin{split}\dot{x}_{1}&=x_{2}-x_{3}\\ \dot{x}_{2}&=\frac{k_{m}}{J_{u}}u-\frac{k_{\theta}}{J_{u}}x_{1}-\frac{b_{up}}{J_{u}}x_{2}\\ \dot{x}_{3}&=\frac{k_{\theta}}{J_{l}}x_{1}-\frac{1}{J_{l}}T_{fl}(x_{3}).\end{split} (36)

The estimation of the parameters are given in [19, Table 1]. Substituting the values into the functions, we have

Tfu(x2)=2.2247x2,T_{fu}(x_{2})=2.2247x_{2},
Tfl(x3)={[0.1642+0.0603(121+e5.7468|x3|)0.2267(121+e0.2941|x3|)]sign(x3)+0.0109x3,x30[0.1642, 0.1642],x3=0.T_{fl}(x_{3})=\left\{\begin{array}[]{lr}$[$0.1642+0.0603(1-\frac{2}{1+e^{5.7468|x_{3}|}})\\ -0.2267(1-\frac{2}{1+e^{0.2941|x_{3}|}})$]${\rm sign}(x_{3})+0.0109x_{3},&x_{3}\neq 0\\ $[$-0.1642,\leavevmode\nobreak\ 0.1642$]$,&\;\;x_{3}=0.\end{array}\right.

We can rewrite (36) as follows

{x˙=Ax+Bω+GuωTfl(Cx)y=Fx,\left\{\begin{array}[]{lr}\dot{x}=Ax+B\omega+Gu\\ \omega\in-T_{fl}(Cx)\\ y=Fx,\end{array}\right. (37)

where A=[011kθJubupJu0kθJl00]A=\left[\begin{array}[]{ccc}0&1&-1\\ -\frac{k_{\theta}}{J_{u}}&-\frac{b_{up}}{J_{u}}&0\\ \frac{k_{\theta}}{J_{l}}&0&0\\ \end{array}\right], G=[0kmJu0]G=\left[\begin{array}[]{c}0\\ \frac{k_{m}}{J_{u}}\\ 0\end{array}\right], B=[001Jl]B=\left[\begin{array}[]{c}0\\ 0\\ \frac{1}{J_{l}}\end{array}\right], C=[0 0 1]C=[0\leavevmode\nobreak\ 0\leavevmode\nobreak\ 1], F=[1 0 0]F=[1\leavevmode\nobreak\ 0\leavevmode\nobreak\ 0]. Substituting the estimated values of the parameters, we have

A=[0110.15264.668802.230100],G=[08.38410],B=[0030.6748].A=\left[\begin{array}[]{ccc}0&1&-1\\ -0.1526&-4.6688&0\\ 2.2301&0&0\\ \end{array}\right],\quad G=\left[\begin{array}[]{c}0\\ 8.3841\\ 0\end{array}\right],\quad B=\left[\begin{array}[]{c}0\\ 0\\ 30.6748\end{array}\right].

Since Tfl(x3)T_{fl}(x_{3}) is not monotone, we use the loop transformation T~fl(Cx)=Tfl(Cx)mCx\tilde{T}_{fl}(Cx)=T_{fl}(Cx)-mCx, where m=0.021m=-0.021. Then

T~fl(x3)={[0.1642+0.0603(121+e5.7468|x3|)0.2267(121+e0.2941|x3|)]sign(x3)+0.0319x3,x30[0.1642, 0.1642],x3=0.\tilde{T}_{fl}(x_{3})=\left\{\begin{array}[]{lr}$[$0.1642+0.0603(1-\frac{2}{1+e^{5.7468|x_{3}|}})\\ -0.2267(1-\frac{2}{1+e^{0.2941|x_{3}|}})$]${\rm sign}(x_{3})+0.0319x_{3},&x_{3}\neq 0\\ $[$-0.1642,\leavevmode\nobreak\ 0.1642$]$,&x_{3}=0.\end{array}\right.

Under the transformation, we have a new system

{x˙=A¯x+Bω+Guω(Cx)y=Fx,\left\{\begin{array}[]{lr}\dot{x}=\bar{A}x+B\omega+Gu\\ \omega\in-\mathcal{F}(Cx)\\ y=Fx,\end{array}\right. (38)

where

A¯=AmBC=[0110.15264.668802.230100.6442],G=[08.38410],\bar{A}=A-mBC=\left[\begin{array}[]{ccc}0&1&-1\\ -0.1526&-4.6688&0\\ 2.2301&0&0.6442\\ \end{array}\right],\quad G=\left[\begin{array}[]{c}0\\ 8.3841\\ 0\end{array}\right],
B=[0030.6748],C=[0 0 1],F=[1 0 0].\quad B=\left[\begin{array}[]{c}0\\ 0\\ 30.6748\end{array}\right],\quad C=[0\leavevmode\nobreak\ 0\leavevmode\nobreak\ 1],\quad F=[1\leavevmode\nobreak\ 0\leavevmode\nobreak\ 0].
(λ)={[0.1642+0.0603(121+e5.7468|λ|)0.2267(121+e0.2941|λ|)]sign(λ)+0.0319λ,λ0[0.1642, 0.1642],λ=0.\mathcal{F}(\lambda)=\left\{\begin{array}[]{lr}$[$0.1642+0.0603(1-\frac{2}{1+e^{5.7468|\lambda|}})\\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ -0.2267(1-\frac{2}{1+e^{0.2941|\lambda|}})$]${\rm sign}(\lambda)+0.0319\lambda,&\lambda\neq 0\\ $[$-0.1642,\leavevmode\nobreak\ 0.1642$]$,&\lambda=0.\end{array}\right.

Here we choose u=2u=2. Solving the following LMIs:

(A¯LF)TP+P(A¯LF)=QBTP=CKF,\begin{split}(\bar{A}-LF)^{T}P+P(\bar{A}-LF)&=-Q\\ B^{T}P=C-KF,\end{split} (39)

we obtain

P=[0.29580.04170.06000.04170.028600.060000.0326],Q=[0.13270.00640.01580.00640.18370.01830.01580.01830.0780],P=\left[\begin{array}[]{ccc}0.2958&0.0417&0.0600\\ 0.0417&0.0286&0\\ 0.0600&0&0.0326\\ \end{array}\right],\leavevmode\nobreak\ Q=\left[\begin{array}[]{ccc}-0.1327&-0.0064&-0.0158\\ -0.0064&-0.1837&0.0183\\ 0.0158&0.0183&-0.0780\end{array}\right],
L=[3.30691.214012.2290],K=1.8392,ϵ=0.0714.L=\left[\begin{array}[]{c}3.3069\\ -1.2140\\ -12.2290\end{array}\right],\leavevmode\nobreak\ K=-1.8392,\leavevmode\nobreak\ \epsilon=0.0714.

Now suppose that the system (38) is influenced by an uncertainty ξ(t,x)\xi(t,x), which is described as follows:

{x˙=A¯x+Bω+Gu+ξ(t,x)ω(Cx)y=Fx.\left\{\begin{array}[]{lr}\dot{x}=\bar{A}x+B\omega+Gu+\xi(t,x)\\ \omega\in-\mathcal{F}(Cx)\\ y=Fx.\end{array}\right. (40)

First we take the unknown ξ(t,x)=ξ(1)=(14sinx2cost)\xi(t,x)=\xi^{(1)}=\left(\begin{array}[]{ccc}1\\ \\ 4\sin x_{2}\\ \\ \cos t\end{array}\right).

Refer to caption
Figure 2: Errors using (10) applied to Example 2 with the uncertainty ξ(1)\xi^{(1)}

Using the approximate sliding mode observer (10), we can see that observation error e1e_{1} converges to zero very fast and the total error module e(t)\|e(t)\| belongs to the attractive set Ω=[0,7.75]\Omega^{\prime}=[0,7.75] in finite time (Figure 2). The numerical result is consistent with Theorem 2 when the uncertainty is chosen randomly. Next we consider the unknown

ξ(t,x)=ξ(2)=(16.055223.409229.5495)+(ete2te1.5t),\xi(t,x)=\xi^{(2)}=\left(\begin{array}[]{ccc}16.0552\\ \\ -23.4092\\ \\ -29.5495\end{array}\right)+\left(\begin{array}[]{ccc}e^{-t}\\ \\ e^{-2t}\\ \\ e^{-1.5t}\end{array}\right),

which has 2 parts: the first part in Im(P1FT){\rm Im}(P^{-1}F^{T}) and the second part vanishes when the time is large. Then the sliding mode observer (10) is indeed the observer of the original system (Figure 3).

Refer to caption
Figure 3: Errors using (10) applied to Example 2 with the uncertainty ξ(2)\xi^{(2)}

5 Conclusions

In this paper, we introduce a sliding mode observer, which also functions as a TT- observer and a strong HH^{\infty} observer, designed for a general category of set-valued Lur’e dynamical systems. Importantly, our approach accommodates scenarios where the uncertainty does not fall within the range of observation. Consequently, we are able to obtain accurate state estimations for the original systems. Furthermore, in cases where the unobservable portion of uncertainty tends to vanish when the time is large, we achieve exact observers. It would be interesting to investigate strategies that enhance the robustness of our proposed observer, enabling it to handle increasingly complex and dynamic uncertainties or disturbances. Exploring ways to incorporate adaptive control strategies into our observer framework would be a promising direction, facilitating its adaptation to evolving system conditions. Optimizing and reducing the size of the attractive set could be a valuable pursuit. The practical application and validation of our observer in real-world systems, such as autonomous vehicles, robotics, or industrial processes, present good opportunities for future exploration. Extending our observer framework to address the challenges posed by set-valued uncertainties would further broaden its applicability and impact. These open questions require further investigation and are beyond the scope of the current paper. They will be the focus of a new research project.

References

  • [1] V. Acary, O. Bonnefon, B. Brogliato, Nonsmooth Modeling and Simulation for Switched Circuits, Lecture Notes in Electrical Engineering Vol 69. Springer Netherlands, 2011
  • [2] S. Adly, A. Hantoute, B. K. Le, Nonsmooth Lur’e Dynamical Systems in Hilbert Spaces, Set-Valued Var Anal 24(1), 13–35, 2016
  • [3] S. Adly, A. Hantoute, B. K. Le, Maximal Monotonicity and Cyclic-Monotonicity Arising in Nonsmooth Lur’e Dynamical Systems”, J Math Anal Appl 448(1), 691–706, 2017
  • [4] S. Adly, B. K. Le, Sliding Mode Observer for Set-valued Lur’e Systems and Chattering Removing, Nonlinear Analysis: Hybrid Systems, Vol. 50, November 2023, 101406.
  • [5] M. Arcak, P. Kokotovic, Nonlinear observers: A circle criterion design and robustness analysis. Automatica 37, 1923–1930, 2001
  • [6] J. Aubin , A. Cellina, Differential Inclusions: Set-valued Maps and Viability Theory. Berlin: Springer Verlag, 1984
  • [7] B. Brogliato, Absolute stability and the Lagrange-Dirichlet theorem with monotone multivalued mappings, Syst Control Lett 51 (5), 343–353, 2004
  • [8] B. Brogliato, D. Goeleven, Well-posedness, stability and invariance results for a class of multivalued Lur’e dynamical systems, Nonlinear Anal Theory Methods Appl 74, 195–212, 2011
  • [9] B. Brogliato, R. Lozano, B. Maschke, O. Egeland, Dissipative Systems Analysis and Control, Springer Nature Switzerland AG, 3rd Edition, 2020
  • [10] J. C. A. de Bruin, A. Doris, N. van de Wouw, W. P. M. H. Heemels, H. Nijmeijer, Control of mechanical motion systems with non-collocation of actuation and friction: A Popov criterion approach for input-to-state stability and set-valued nonlinearities, Automatica, Volume 45, Issue 2, 2009, Pages 405–415.
  • [11] B. Brogliato, W. P. M. H. Heemels, Observer Design for Lur’e Systems With Multivalued Mappings: A Passivity Approach, IEEE Trans Automat Contr 54(8), 1996–2001, 2009
  • [12] B. Brogliato, A. Tanwani, Dynamical Systems Coupled with Monotone Set-Valued Operators: Formalisms, Applications, Well-Posedness, and Stability, SIAM Rev., 62(1), 3–129, 2020
  • [13] H. Brezis, Opérateurs Maximaux Monotones et Semi-groupes de Contractions dans les Espaces de Hilbert, Math. Studies 5, North-Holland American Elsevier, 1973
  • [14] M. K. Camlibel, J. M. Schumacher, Linear passive systems and maximal monotone mappings, Math Program 157(2), 397–420, 2016
  • [15] J. Huang , Z. Han , X. Cai , L. Liu, Adaptive full-order and reduced-order observers for the Lur’e differential inclusion system, Commun Nonlinear Sci Numer Simul 16, 2869–2879, 2011
  • [16] J. Huang , Z. Han , X. Cai, Note on observer for Lur’e differential inclusion systems, IET Control Theory & Applications, vol. 5 (17), 2011
  • [17] J. Huang, W. Zhang, M. Shi, L. Chen, L. Yu, HH^{\infty} Observer design for singular one-sided Lur’e differential inclusion system, J Franklin Inst, 3305–3321, 2017
  • [18] J. Huang , X. Ma, X. Zhao, H. Che, L. Chen, Interval observer design method for asynchronous switched systems, IET Control Theory & Applications, 2020
  • [19] A. L. Juloski, N. Mihajlovic, W. P. M. H. Heemels, N. van de Wouw and H. Nijmeijer, Observer design for an experimental rotor system with discontinuous friction, 2006 American Control Conference, Minneapolis, MN, USA, 2006
  • [20] B. K. Le, On a class of Lur’e dynamical systems with state-dependent set-valued feedback, Set-Valued Var Anal 28, 537–557, 2020
  • [21] B. K. Le, Sliding Mode Observers for Time-Dependent Set-Valued Lur’e Systems Subject to Uncertainties, J Optim Theory Appl 194, 290–305, 2022
  • [22] A.I. Lur’e, V.N. Postnikov. On the theory of stability of control systems. Appl. Math. Mech. 8(3), (1944) (in Russian)
  • [23] F. Mazenca, O. Bernard, Interval observers for linear time-invariant systems with disturbances, Automatica 47 (2011) 140–147
  • [24] Y. Shtessel, C. Edwards, L. Fridman and A. Levant, Sliding Mode Control and Observation, Birkhauser, 2013
  • [25] R. E. Showalter, Monotone Operators in Banach Spaces and Nonlinear Partial Differential Equations, Amer. Math. Society, Providence, RI, 1997
  • [26] S. Spurgeon, Sliding mode observers - a survey. Int. J. Syst. Sci. 39(8), 751–764, 2008
  • [27] A. Tanwani, B. Brogliato, C. Prieur, Stability and observer design for Lur’e systems with multivalued, non-monotone, time-varying nonlinearities and state jumps, SIAM J. Control Opti., Vol. 52, No. 6, 3639–3672, 2014
  • [28] A. Tanwani, B. Brogliato, C. Prieur, Well-Posedness and Output Regulation for Implicit Time-Varying Evolution Variational Inequalities, SIAM J. Control Opti., Vol. 56(2), 751–781, 2018
  • [29] J. Xiang, H. Su, J. Chu: On the design of Walcott-Zak sliding mode observer, Proc. Amer. Contr. Conf. Portland, OR, pp. 2451–2456, 2005