Loading [MathJax]/extensions/MathZoom.js
Linear Active Disturbance Rejection Control for Nonaffine Strict-Feedback Nonlinear Systems | IEEE Journals & Magazine | IEEE Xplore

Linear Active Disturbance Rejection Control for Nonaffine Strict-Feedback Nonlinear Systems


A linear active disturbance rejection controller for a class of uncertain, nonaffine, strict-feedback nonlinear systems is proposed. A reduced-order linear extended state...

Abstract:

This paper presents a linear active disturbance rejection controller (LADRC) for nonaffine nonlinear systems with strict-feedback form. The proposed LADRC results in a cl...Show More

Abstract:

This paper presents a linear active disturbance rejection controller (LADRC) for nonaffine nonlinear systems with strict-feedback form. The proposed LADRC results in a closed-loop system with a three-time-scale structure, in which a reduced-order linear extended state observer estimates unmeasured states and total disturbance in the fastest time scale and dynamic inversion based on sector conditions is used to deal with nonaffine inputs in the intermediate time scale while the system dynamics evolves in the slowest time scale. The singular perturbation method is used to analyze the stability and performance of the system. The effectiveness of the proposed LADRC is demonstrated through simulation studies and experimental validation on a linear motor servo system with nonaffine uncertainties.
A linear active disturbance rejection controller for a class of uncertain, nonaffine, strict-feedback nonlinear systems is proposed. A reduced-order linear extended state...
Published in: IEEE Access ( Volume: 7)
Page(s): 120030 - 120040
Date of Publication: 27 August 2019
Electronic ISSN: 2169-3536

Funding Agency:

References is not available for this document.

CCBY - IEEE is not the copyright holder of this material. Please follow the instructions via https://creativecommons.org/licenses/by/4.0/ to obtain full-text articles and stipulations in the API documentation.
SECTION I.

Introduction

Control of uncertain nonlinear systems is one of the most important issues in control theory. Among many different methods developed in the past few decades, the active disturbance rejection control (ADRC), first proposed by Han [1], is an efficient method to deal with uncertain nonlinear systems. For its ability in dealing with a vast range of uncertainties, simplicity in engineering implementation, and superior performance in practice, ADRC is becoming an emerging technology in control engineering [2]. The last two decades have witnessed ADRC’s success in many industrial applications, including speed control of induction motor [3] and permanent magnet synchronous motor [4], fan control in server [5], air-fuel ratio control in gasoline engines [6], hydraulic servo systems [7]. In recent years, the theoretical research has been gradually developed. In [8], [9], the characteristics of the linear ADRC (LADRC), first proposed by Gao [10], are comprehensively studied both in time-domain and frequency-domain. In [11]–​[13], the convergence of nonlinear extended state observer (NLESO) and nonlinear ADRC (NLADRC) are proved, respectively. Several modified ADRCs for nonlinear uncertain systems with time delay are studied and compared with theoretical analysis in [14].

Most of the existing ADRC studies consider that the considered systems are affine-in-control, i.e., the control inputs appear linearly in the control systems, especially in the known nominal models. However, there are many practical control systems, such as the inverted pendulum [15], and flight control system [16], [17], which cannot be expressed in an affine-in-control form. Thus, it is necessary to study how to design ADRC for uncertain nonaffine nonlinear systems.

Various approaches for nonaffine uncertain nonlinear systems have been proposed recently [18]–​[25]. An output feedback controller is proposed for uncertain, single-input-single-output nonaffine input systems using neural network observers together with dynamic inversion in [18]. Using neural networks to approximate the nonaffine nominal model and uncertainties, the stability of the closed-loop system can be proved [19], [20]. All of these methods based on neural networks require heavy computations and good prior knowledge of the system. To deal with heavy computation burdens, an extended high-gain observer (EHGO) is used to estimate system states and uncertainties instead of neural networks. And dynamic inversion is used to deal with the nonaffine inputs and input uncertainties based on the estimates provided by EHGO in [17], [21]. Based on the control design method for feedback linearization systems in [21], an ADRC design method, combining NLESO and dynamic inversion, is proposed for uncertain nonaffine strict-feedback nonlinear systems in [22]. In [23], an indirect dynamic inversion approach is developed for uncertain nonaffine strict-feedback nonlinear systems. However, the model uncertainties estimation provided by EHGO is not used and information of initial states are required in the indirect dynamic inversion approach. In [24], an observer-based adaptive fuzzy method is proposed to control an uncertain nonaffine feedback linearization system. Similarly, an adaptive fuzzy controller is designed for a class of uncertain nonaffine feedback linearization systems based on the passivity theorem in [25].

In classical ADRC framework, the choice of nominal value b_{0} of control gain b has a great impact on the performance of the closed-loop systems. In [26], the stability of the closed-loop system can only be guaranteed with the condition \frac {b}{b_{0}} \in \left ({{0,2} }\right) be satisfied. The range of b_{0} is extended to \frac {b}{b_{0}} \in \left ({{0,2 + \frac {2}{n}} }\right) in [8], [9]. Compared with these results in [26] and [8], [9], the advantage of dynamic inversion is that only the control direction, i.e., the sign of control gain b , is needed. This idea is successfully applied to the stabilization of the inverted pendulum on a cart to solve the uncertainties of control gain in [27].

This paper presents a LADRC design method for a class of uncertain nonaffine strict-feedback nonlinear systems. Reduced-order linear extended state observer (RLESO) and dynamic inversion are combined together with a multi-time-scale structure to deal with model uncertainties and nonaffine inputs. Numerical simulation and experimental results demonstrate the effectiveness of the proposed method. The contributions of this paper can be summarized as follows.

  1. We extend the applications of dynamic inversion from feedback linearization systems in [21], [24], [25] to a more general nonaffine systems with strict feedback form.

  2. Compared with the indirect dynamic inversion approach for strict-feedback nonlinear systems in [23], a novel design method is proposed, where the uncertainties estimation of the system is adopted effectively and the requirements of the initial states can be removed. The stability and performance analysis for the multi-time-scale structure is rigorously studied via the singular perturbation method.

  3. The proposed LADRC provides an extra ADRC design method to handle the difficulty of the choice of nominal value of control gain in the classical ADRC framework.

This paper is organized as follows. In Section II, the problem is formulated for a class of uncertain, nonaffine, nonlinear systems. In Section III, the LADRC design is presented. Section IV provides the stability and performance analysis. Comparative simulation and experimental results with the existing approaches are presented in Section V. Concluding remarks are provided in Section VI.

SECTION II.

Problem Statements

Consider a class of uncertain nonaffine nonlinear systems with strict-feedback form:\begin{align*} {\dot x_{i}}=&{x_{i + 1}} + {f_{i}}({x_{1}}, \cdots,{x_{i}}), \\&i = 1,2, \cdots,n - 1,\quad {x_{i}}(0) = {x_{i0}}, \\ {\dot x_{n}}=&{\bar f_{n}}(x,z,w,u),\quad {x_{n}}(0) = {x_{n0}}, \\ \dot z=&{f_{0}}(x,z,w),\quad z(0) = {z_{0}}, \\ y=&{x_{1}}\tag{1}\end{align*}

View SourceRight-click on figure for MathML and additional features. where x = {[{x_{1}},{x_{2}}, \cdots,{x_{n}}]^{T}} \in {D_{x}} \subset {\mathbb {R}^{n}} denotes the system states, z \in {D_{z}} \subset {\mathbb {R}^{m}} is the states of internal dynamics, u \in {D_{u}} \subset \mathbb {R} denotes the control input, the domains {D_{x}},{D_{z}},{D_{u}} contain their respective origins. y is the measured output, w \in \mathbb {R} denotes external disturbances. Functions {f_{i}}(\cdot):{\mathbb {R}^{i}} \to \mathbb {R},i = 1,2, \cdots,n - 1 are known, and {f_{0}}(\cdot) is unknown, the function {\bar f_{n}}(\cdot):{\mathbb {R}^{n + m + 2}} \to \mathbb {R} can be expressed as \begin{equation*}{\bar f_{n}}(x,z,w,u) = {f_{n}}(x,u) + {\sigma _{f_{n}}}(x,z,w,u),\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where {f_{n}}:{\mathbb {R}^{n + 1}} \to \mathbb {R} is known, and {\sigma _{f_{n}}}:{\mathbb {R}^{n + m + 2}} \to \mathbb {R} is unknown.

The goal of this paper is to design the control input u that can make the system (1) follow a target system. The target system is defined by the following dynamics:\begin{align*} {\dot r_{i}}=&{r_{i + 1}} + {f_{i}}({x_{1}}, \cdots,{x_{i}}), \\&i = 1,2, \cdots,n - 1,\quad {r_{i}}(0) = {r_{i0}}, \\ {\dot r_{n}}=&{f_{rn}}(t),\quad {r_{n}}(0) = {r_{n0}},\tag{2}\end{align*}

View SourceRight-click on figure for MathML and additional features. where {f_{rn}}(t) belongs to a compact set {D_{f}} \subset \mathbb {R} , and r = {[{r_{1}},{r_{2}}, \cdots,{r_{n}}]^{T}} belongs to a compact set {D_{r}} \subset {D_{x}} .

Let \begin{align*} {e_{i}}=&{x_{i}} - {r_{i}},\quad e = {\left [{ {\begin{array}{cccc} {e_{1}}&\quad {e_{2}}&\quad \cdots &\quad {e_{n}} \end{array}} }\right]^{T}},\\ A=&\left [{ {\begin{array}{cccc} 0&\quad 1&\quad {}&\quad {}\\ \vdots &\quad {}&\quad \ddots &\quad {}\\ 0&\quad {}&\quad {}&\quad 1\\ 0&\quad 0&\quad \cdots &\quad 0 \end{array}} }\right],\quad B = \left [{ {\begin{array}{c} 0\\ \vdots \\ 0\\ 1 \end{array}} }\right],\end{align*}

View SourceRight-click on figure for MathML and additional features. the error dynamics are \begin{align*} {\dot e_{i}}=&{e_{i + 1}},i = 1,2, \cdots,n - 1, \\ {\dot e_{n}}=&- Le + F(x,z,w,u,{f_{rn}}(t)),\tag{3}\end{align*}
View SourceRight-click on figure for MathML and additional features.
where the error gain matrix L = {[{\ell _{1}},{\ell _{2}}, \cdots,{\ell _{n}}]} \in {\mathbb {R}^{n}} can be chosen that (A - BL) is Hurwitz, F(x,z,w,u,{f_{rn}}(t)) is \begin{equation*}F(x,z,w,u,{f_{rn}}(t)) = {\bar f_{n}}(x,z,w,u) - {f_{rn}}(t) + Le.\end{equation*}
View SourceRight-click on figure for MathML and additional features.

The following mild assumptions are made for the system (1) and error dynamics (3), and they are common in the relevant literature.

Assumption 1:

w(t) belongs to a known compact set D_{w} \subset \mathbb {R} , and \dot w(t) is bounded.

Remark 1:

Assumption 1 is a rational assumption, which is also presented in [26].

Assumption 2:

The functions {f_{i}}(\cdot),i = 1, \cdots,n,{\sigma _{f_{n}}}(\cdot), {f_{0}}(\cdot) are continuously differentiable. Besides, for any given compact set in {D_{x}} \times {D_{z}} \times {D_{u}} , there exist {k_{i}} \in \mathbb {R},i = 1,2, \cdots,n,{k_{u}} \in \mathbb {R} satisfying:\begin{align*}&\hspace {-2pc}\left |{ {f_{i}({x_{1}} + {\delta _{1}}, \cdots,{x_{i}} + {\delta _{i}}) - {f_{i}}({x_{1}}, \cdots,{x_{i}})} }\right | \\&\le {k_{i}}\sum \limits _{j = 1}^{i} {\left |{ {\delta _{j}} }\right |},\quad \forall {\delta _{j}} \in \mathbb {R},\\&\hspace {-2pc}\left |{ {f_{n}({x_{1}} + {\delta _{1}}, \cdots,{x_{n}} + {\delta _{n}},u) - {f_{n}}({x_{1}}, \cdots,{x_{n}},u)} }\right |\\&\le {k_{n}}\sum \limits _{j = 1}^{n} {\left |{ {\delta _{j}} }\right |},\quad \forall {\delta _{j}} \in \mathbb {R}, \\&\hspace {-2pc}\left |{ {{\sigma _{f_{n}}}(x,z,w,u + {\delta _{u}}) - {\sigma _{f_{n}}}(x,z,w,u)} }\right | \\&\le {k_{u}}\left |{ {\delta _{u}} }\right |,\quad \forall {\delta _{u}} \in \mathbb {R}.\end{align*}

View SourceRight-click on figure for MathML and additional features.

Remark 2:

Assumption 2 is a Lipschitz-like condition, which guarantees the existence of solutions for (1).

Assumption 3:

There exists a positive definite function {V_{z}} such that for all x \in {D_{x}}, z \in {D_{z}}, w \in {D_{w}} \begin{align*} {\alpha _{1}}(\left \|{ z }\right \|)\le&{V_{z}}(z) \le {\alpha _{2}}(\left \|{ z }\right \|), \\ {\dot V_{z}}\le&0,\quad { {for}}\;\left \|{ z }\right \| \ge {\alpha _{3}}(\left \|{ {x,w} }\right \|),\end{align*}

View SourceRight-click on figure for MathML and additional features. where {\alpha _{1}}(\cdot), {\alpha _{2}}(\cdot) and {\alpha _{3}}(\cdot) are class \kappa functions.

Remark 3:

Assumption 3 implies that the z -subsystem in (1) is bounded-input-bounded-state stable and if {\left \|{ x }\right \|} is bounded in a finite time, the states of the internal dynamics are also bounded. It leads our attention to the trajectory of x .

Assumption 4:

There is a unique continuously differentiable function \phi (x,z,w,{f_{rn}}(t)) such that u_{r} = \phi (x,z,w,{f_{rn}}(t)) satisfies \begin{equation*}F(x,z,w,{u_{r}},{f_{rn}}(t)) = 0,\end{equation*}

View SourceRight-click on figure for MathML and additional features. and its derivative {\dot u_{r}} \stackrel { \Delta } = {\phi _{d}}(x,z,w,{f_{rn}}(t),{\dot f_{rn}}(t)) is bounded on compact sets of {D_{x}} \times {D_{z}} \times {D_{u}} . Furthermore, there’s a known continuously differentiable K(x,{f_{rn}}(t)) , for any (x,z,u) \in {D_{x}} \times {D_{z}} \times {D_{u}} , satisfying \begin{equation*}sK(x,{f_{rn}}(t))F(x,z,w,s + {u_{r}},{f_{rn}}(t)) \ge \beta {s^{2}},\quad \beta > 0,\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where s = u - {u_{r}} , K(x,{f_{rn}}(t)) satisfies \begin{equation*}\left |{ {K(x + {\delta _{x}},{f_{rn}}(t)) - K(x,{f_{rn}}(t))} }\right | \le {k_{b}}\left \|{ {\delta _{x}} }\right \|,\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where {k_{b}} \ge 0 .

Remark 4:

Assumption 4 guarantees the uniform controllability of the system (1), and the requirement of K(x,{f_{rn}}(t)) in Assumption 4 is a sector condition. An available method is selecting K as the generalized direction of {\bar f_{n}}(\cdot) with respect to u , i.e., K = sign\left ({{\frac {{\partial {\bar f_{n}}}}{\partial u}(x,z,w,u)} }\right) . In this case, according to the mean value theorem there exists a constant k_{u}^ {*} such that \begin{align*}&\hspace {-2pc}{sK\left [{ {\bar f_{n}(x,z,w,s + {u_{r}}) - {f_{rn}}(t) + Le} }\right]} \\=&sign\left ({{\frac {{\partial {\bar f_{n}}}}{\partial u}(x,z,w,u)} }\right)s{{[}}k_{u}^{*}s + {\bar f_{n}}(x,z,w,{u_{r}}) \\&-\, {f_{rn}}(t) + Le{{]}} \\\ge&\left |{ {k_{u}^{*}} }\right |{s^{2}},\end{align*}

View SourceRight-click on figure for MathML and additional features. Therefore, this selection of K(x,{f_{rn}}(t)) satisfies Assumption 4.

SECTION III.

LADRC Design

A novel LADRC design method for the system (1) will be introduced, and we use dynamic inversion to deal with nonaffine inputs and input uncertainties.

Let \begin{align*} F(x,z,w,u,{f_{rn}}(t))=&{\bar f_{n}}(x,z,w,u) - {f_{rn}}(t) + Le \\=&{f_{n}}(x,u) + {\sigma _{f_{n}}}(x,z,w,u)\\&- \,{f_{rn}}(t) + Le \\\stackrel { \Delta } =&{F_{c}}(x,{\sigma _{f_{n}}},u,{f_{rn}}(t)).\end{align*}

View SourceRight-click on figure for MathML and additional features. If the system states x and {\sigma _{f_{n}}}(x,z,w,u) are known, the dynamic inversion [18], [21], [28] can be expressed as follows:\begin{equation*} \mu \dot u = - K(x,{f_{rn}}(t)){F_{c}}(x,{\sigma _{f_{n}}},u,{f_{rn}}(t)),\quad u(0) = {u_{0}},\tag{4}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where \mu is a small positive number.

The closed-loop system can be expressed as \begin{align*} \dot e=&(A - BL)e + B{F_{c}}(x,{\sigma _{f_{n}}},s + u_{r},{f_{rn}}(t)), \\ \dot z=&{f_{0}}(x,z,w), \\ \mu \dot s=&- K{F_{c}}(x,{\sigma _{f_{n}}},s + u_{r},{f_{rn}}(t)) - \mu {\phi _{d}}.\tag{5}\end{align*}

View SourceRight-click on figure for MathML and additional features.

Let \begin{equation*}{V_{s}} = \frac {1}{2}{s^{2}},\quad {V_{e}} = {e^{T}}{P_{e}}e,\end{equation*}

View SourceRight-click on figure for MathML and additional features. where P_{e} is the solution of the Lyapunov equation P_{e}(A - BL) + {(A - BL)^{T}}P_{e} = - I and I is the identity matrix. From Assumption 4, \begin{align*} {\dot V_{s}}\le&- \left ({{\frac {\beta }{\mu }} }\right){s^{2}} + \left |{ s }\right |\left |{ {\phi _{d}} }\right |, \\ {\dot V_{e}}=&- {e^{T}}e + 2{e^{T}}{P_{e}}B{F_{c}}(x,{\sigma _{f_{n}}},s + u_{r},{f_{rn}}(t)), \\\le&- {\left \|{ e }\right \|^{2}} + {k_{e}}\left |{ s }\right |\left \|{ e }\right \|,\end{align*}
View SourceRight-click on figure for MathML and additional features.
on the compact sets, for sufficiently small \mu . Therefore, there exist constants {b_{1}} > 0,{b_{2}} > 0,{c_{0}} > 0 such that \begin{equation*}{\Omega _{b}} = \left \{{ {V_{e} \le {b_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {b_{2}}} }\right \}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
is a compact, positively invariant set with respect to (5).

In output feedback control and {\sigma _{f_{n}}}(x,z,w,u) is unknown, the system states x and {\sigma _{f_{n}}}(x,z,w,u) are estimated by the following RLESO: \begin{align*} {\dot z_{i}}=&- \frac {{{\alpha _{i - 1}}}}{{{\varepsilon ^{i - 1}}}}\left ({{z_{2} + {f_{1}}({x_{1}})} }\right) + {z_{i + 1}} \\&+\, {f_{i}}({x_{1}},{\hat x_{2}}, \cdots,{\hat x_{i}}) + \left ({{\frac {\alpha _{i}}{\varepsilon ^{i}} - \frac {{\alpha _{1}{\alpha _{i - 1}}}}{\varepsilon ^{i}}} }\right){x_{1}}, \\ {\hat x_{i}}=&{z_{i}} + \frac {{{\alpha _{i - 1}}}}{{{\varepsilon ^{i - 1}}}}{x_{1}}, \\&i = 2,3, \cdots,n,\quad {z_{i}}(0) = - {\beta _{i - 1}}{x_{1}}(0), \\ {\dot z_{n + 1}}=&- \frac {\alpha _{n}}{\varepsilon ^{n}}\left ({{z_{2} + {f_{1}}({x_{1}})} }\right) - \frac {{\alpha _{1}{\alpha _{n}}}}{{{\varepsilon ^{n + 1}}}}{x_{1}}, \\ {\hat \sigma _{f_{n}}}=&{z_{n + 1}} + \frac {\alpha _{n}}{\varepsilon ^{n}}{x_{1}},\quad {z_{n + 1}}(0) = - {\beta _{n}}{x_{1}}(0),\tag{6}\end{align*}

View SourceRight-click on figure for MathML and additional features. where the parameters {\alpha _{1}},{\alpha _{2}}, \cdots,{\alpha _{n}} are chosen such that \begin{equation*}{s^{n}} + {\alpha _{1}}{s^{n-1}} + \cdots + {\alpha _{n-1}}s + {\alpha _{n}}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
is Hurwitz, and \varepsilon is a small positive number. Especially, when n=1 , RLESO in (6) can be simplified as:\begin{align*} {\dot z_{2}}=&- \frac {\alpha _{1}}{\varepsilon }\left ({{z_{2} + {f_{1}}({x_{1}})} }\right) - \frac {\alpha _{1}^{2}}{\varepsilon ^{2}}{x_{1}}, \\ {\hat \sigma _{f_{1}}}=&{z_{2}} + \frac {\alpha _{1}}{\varepsilon }{x_{1}},\quad {z_{2}}(0) = - {\beta _{1}}{x_{1}}(0).\tag{7}\end{align*}
View SourceRight-click on figure for MathML and additional features.

Since the dynamic inversion uses the estimates provided by the RLESO, \varepsilon is smaller than \mu to make the EHGO faster than the dynamic inversion. Therefore, the parameters \varepsilon and \mu are chosen such that 0 < \varepsilon \ll \mu \ll 1 . Using the RLESO in (6) together with the dynamic inversion in (4), the output feedback control is designed as \begin{align*}&\hspace {-1pc}\mu \dot u = - K({\hat x_{s}},{f_{rn}}(t))\left [{ {f_{n}({\hat x_{s}},u) + {\hat \sigma _{f_{n}s}} - {f_{rn}}(t) + L{\hat e_{s}}} }\right], \\&\quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad u(0) = {u_{0}},\tag{8}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{align*} {\hat x_{s}}=&{\left [{ {x_{1},{M_{2}}sat\left ({{\frac {\hat x_{2}}{M_{2}}} }\right), \cdots,{M_{n}}sat\left ({{\frac {\hat x_{n}}{M_{n}}} }\right)} }\right]^{T}}, \\ {\hat \sigma _{f_{n}s}}=&{M_\sigma }sat\left ({{\frac {{{\hat \sigma _{f_{n}}}}}{M_\sigma }} }\right),\\ \hat e_{s}=&\hat x_{s} - r,\end{align*}
View SourceRight-click on figure for MathML and additional features.
the saturation function sat(\cdot) is defined as sat(\xi) = \min \left \{{ {1,\left |{ \xi }\right |} }\right \}sign(\xi) and is used to prevent peaking from degrading the system performance.

SECTION IV.

Main Results

In this section, we will prove that in the presence of uncertainties, the output feedback control (6) and (8) can force the system (1) to follow the target system (2).

A. n \gt 1

Consider fast variables \begin{equation*} {\eta _{i}} = \frac {{x_{i} - {\hat x_{i}}}}{{{\varepsilon ^{n + 1 - i}}}},i = 2, \cdots,n,\quad {\eta _{n + 1}} = {\sigma _{f_{n}}} - {\hat \sigma _{f_{n}}},\tag{9}\end{equation*}

View SourceRight-click on figure for MathML and additional features. the closed-loop system can be expressed as a standard singularly perturbed form \begin{align*} \dot e=&(A - BL)e + {\delta _{f}} \\&+\, B\left [{ {\bar f_{n}(x,z,w,s + {u_{r}}) - {f_{rn}}(t) + Le} }\right], \\ \dot z=&{f_{0}}(z,x,w), \\ \mu \dot s=&- K\left [{ {f_{n}({\hat x_{s}},s + {u_{r}}) + {\hat \sigma _{f_{n}s}} - {f_{rn}}(t) + L\hat e} }\right] - \mu {\phi _{d}}, \\ \varepsilon \dot \eta=&\Lambda \eta + {\delta _{f/\varepsilon }} + {B}(\varepsilon /\mu){\psi _{1}},\tag{10}\end{align*}
View SourceRight-click on figure for MathML and additional features.
where \begin{align*} \Lambda=&\left [{ {\begin{array}{ccccc} { - {\alpha _{1}}}&\quad 1&\quad 0&\quad \cdots &\quad 0\\ { - {\alpha _{2}}}&\quad 0&\quad 1&\quad \ddots &\quad \vdots \\ \vdots &\quad \vdots &\quad \ddots &\quad \ddots &\quad 0\\ { - {\alpha _{n-1}}}&\quad 0&\quad \cdots &\quad 0&\quad 1\\ { - {\alpha _{n}}}&\quad 0&\quad 0&\quad \cdots &\quad 0 \end{array}} }\right], \\ {\delta _{f}}=&\left [{ {\begin{array}{c} {0}\\ {\left ({{f_{2}({x_{1}},{x_{2}}) - {f_{2}}({x_{1}},{\hat x_{2}})} }\right)}\\ \vdots \\ {\left ({{{f_{n - 1}}({x_{1}}, \cdots,{x_{n - 1}}) - {f_{n - 1}}({x_{1}}, \cdots,{\hat x_{n - 1}})} }\right)}\\ 0 \end{array}} }\right], \\ {\delta _{f/\varepsilon }}=&\left [{ {\begin{array}{c} {\dfrac {1}{{{\varepsilon ^{n - 2}}}}\left ({{f_{2}({x_{1}},{x_{2}}) - {f_{2}}({x_{1}},{\hat x_{2}})} }\right)}\\ {\dfrac {1}{{{\varepsilon ^{n - 3}}}}\left ({{f_{3}({x_{1}},{x_{2}},{x_{3}}) - {f_{3}}({x_{1}},{\hat x_{2}},{\hat x_{3}})} }\right)}\\ \vdots \\ {\left ({{f_{n}(x, s + u_{r}) - {f_{n}}({\hat x}, s + u_{r})} }\right)}\\ 0 \end{array}} }\right], \\ {\psi _{1}}=&- \frac {{\partial {\sigma _{f_{n}}}}}{\partial u}K{F_{cs}} + \mu \left \{{ {\frac {{\partial {\sigma _{f_{n}}}}}{\partial x}\left [{ {Ax + f(x,z,s + {u_{r}})} }\right]} }\right \} \\&+ \,\mu \left \{{ {\frac {{\partial {\sigma _{f_{n}}}}}{\partial z}{f_{0}}(x,z) + \frac {{\partial {\sigma _{f_{n}}}}}{\partial w}\dot w} }\right \}, \\ {F_{cs}}=&{f_{n}}({\hat x_{s}},s + {u_{r}}) + {\hat \sigma _{f_{n}s}} - {f_{rn}}(t) + L\hat e.\end{align*}
View SourceRight-click on figure for MathML and additional features.

According to Assumption 4 and (9), we have:\begin{align*} \left \|{ {\delta _{f}} }\right \|\le&{\varepsilon ^{2}}\sum \limits _{j = 2}^{n - 1} {j{k_{j}}} \left \|{ \eta }\right \|, \\ \left \|{ {{\delta _{f/\varepsilon }}} }\right \|\le&\varepsilon \sum \limits _{j = 2}^{n} {j{k_{j}}} \left \|{ \eta }\right \|.\tag{11}\end{align*}

View SourceRight-click on figure for MathML and additional features.

We are going to analyze the stability of the boundary layer and reduced systems in the closed-loop system (10). Since the z -subsystem with input x is bounded-input-bounded-state stable, we are focusing on the e/s/\eta subsystems. The stability analysis of each subsystem will be shown one by one using the time-scale structure of the closed-loop system (10).

Firstly, by considering \eta -subsystem as the fast system and the e/s -subsystems as the slow system, the boundary layer system can be expressed as\begin{equation*} \varepsilon \dot \eta = \Lambda \eta,\quad \eta (0) = {\eta _{0}}.\tag{12}\end{equation*}

View SourceRight-click on figure for MathML and additional features. Since \Lambda is Hurwitz, the boundary layer system (12) is exponentially stable at \eta = 0 .

After the fast variable \eta reaches its quasi-steady state, \eta = 0 , the reduced system is obtained by setting \eta = 0 for e/s -subsystems. In the reduced system, s -subsystem is the fast system and the e -subsystem is slow, the boundary layer system can be obtained as \begin{equation*} \mu \dot s = - KF(x,z,w,s + {u_{r}},{f_{rn}}(t)),\quad s(0) = {s_{0}},\tag{13}\end{equation*}

View SourceRight-click on figure for MathML and additional features. according to Assumption 4, we have \begin{equation*}\mu {\dot V_{s}} = - sKF \le - \beta {s^{2}}.\end{equation*}
View SourceRight-click on figure for MathML and additional features.
Therefore, the boundary layer system in (13) will be exponentially stable at s = 0 . Meanwhile, the reduced system is \begin{equation*} \dot e = (A - BL)e,\quad e(0) = {e_{0}}.\tag{14}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
Since (A - BL) is Hurwitz, e -subsystem will be exponentially stable at e = 0 .

Let \begin{equation*}{\Omega _{a}} = \left \{{ {V_{e} \le {a_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {a_{2}}} }\right \},\end{equation*}

View SourceRight-click on figure for MathML and additional features. where \begin{equation*}0 < {a_{1}} < {b_{1}},\quad 0 < {a_{2}} < {b_{2}},\end{equation*}
View SourceRight-click on figure for MathML and additional features.
and {Q_{x}} \subset {\mathbb {R}^{n}} and {Q_{s}} \subset \mathbb {R} be any compact subsets containing their respective origins.

Theorem 1:

Under Assumptions 1–​4, suppose the trajectory (e,z,s,\hat x,\hat \sigma) of the closed-loop system (10) starts from ({e_{0}},{z_{0}},{s_{0}},{\hat x_{0}},{\hat \sigma _{0}}) \in {\Omega _{a}} \times {Q_{x}} \times {Q_{s}} , then there exists a constant \varsigma \in \left]{ {0{{,}}1} }\right [ such that for \max \left \{{ {\mu,(\varepsilon /\mu)} }\right \} < \varsigma ,

  1. all trajectories of (10) are bounded.

  2. for any given positive number a , \begin{equation*}\lim \limits _{\mu,(\varepsilon /\mu) \to 0} \left \|{ {x - \hat x} }\right \| = 0,\quad \lim \limits _{\mu,(\varepsilon /\mu) \to 0} \left |{ {u - {u_{r}}} }\right | = 0,\end{equation*}

    View SourceRight-click on figure for MathML and additional features. hold uniformly on \left [{ {a,\infty } }\right [ , and \begin{equation*}\lim \limits _{\mu,(\varepsilon /\mu) \to 0,t \to \infty } \left \|{ {x - {r}} }\right \| = 0.\end{equation*}
    View SourceRight-click on figure for MathML and additional features.

Proof:

The proof of Theorem 1 is given in Appendix A.

B. n = 1

Define \begin{equation*}{e_{1}} = {x_{1}} - {r_{1}},\quad \eta = {\sigma _{f_{1}}} - {\hat \sigma _{f_{1}}},\end{equation*}

View SourceRight-click on figure for MathML and additional features. the closed-loop system can be expressed as a standard singularly perturbed form \begin{align*} {\dot e_{1}}=&- \ell {e_{1}} + \left [{ {\bar f_{1}({x_{1}},z,w,s + {u_{r}}) - {f_{r1}}(t) + \ell {e_{1}}} }\right], \\ \dot z=&{f_{0}}(z,x,w), \\ \mu \dot s=&- K\left [{ {f_{1}({x_{1}},s + {u_{r}}) + {\hat \sigma _{f_{1}s}} - {f_{r1}}(t) + \ell {e_{1}}} }\right] - \mu {\phi _{d}}, \\ \varepsilon \dot \eta=&- {\alpha _{1}}\eta + (\varepsilon /\mu){\psi _{1}},\tag{15}\end{align*}
View SourceRight-click on figure for MathML and additional features.
where \begin{align*} {\psi _{1}}=&- \frac {{\partial {\sigma _{f_{1}}}}}{\partial u}K{F_{cs}} + \mu \left \{{ {\frac {{\partial {\sigma _{f_{1}}}}}{{\partial {x_{1}}}}{\bar f_{1}}({x_{1}},z,w,s + {u_{r}})} }\right \} \\&+ \,\mu \left \{{ {\frac {{\partial {\sigma _{f_{1}}}}}{\partial z}{f_{0}}(x,z,w) + \frac {{\partial {\sigma _{f_{1}}}}}{\partial w}\dot w} }\right \}, \\ {F_{cs}}=&{f_{1}}({x_{1}},s + {u_{r}}) + {\hat \sigma _{f_{1}s}} - {f_{r1}}(t) + \ell {e_{1}}.\end{align*}
View SourceRight-click on figure for MathML and additional features.

The closed-loop system (15) can be viewed as a special case of (10) in Subsection IV-A. Therefore, the stability and performance analysis results in Subsection IV-A is also applicable to (15).

SECTION V.

Simulation and Experiment

A. Numerical Simulations

In order to verify the effectiveness of the proposed LADRC, consider the modified Van der Pol oscillator [23] \begin{align*} {\dot x_{1}}=&{x_{2}} + {f_{1}}({x_{1}}),\quad {x_{1}}(0) = 0, \\ {\dot x_{2}}=&{f_{2}}(x,u) + {\sigma _{f_{2}}}(x,z,w,u),\quad {x_{2}}(0) = 0, \\ \dot z=&{f_{0}}(z,x),\quad z(0) = 0.5, \\ y=&{x_{1}} + n(t),\tag{16}\end{align*}

View SourceRight-click on figure for MathML and additional features. where n(t) is measurement noises, {f_{1}}({x_{1}}) = 0.2\sin ({x_{1}}) and {f_{2}}(x,u) = - {x_{1}} + 0.01u are known, {f_{0}}(z,x) = - z + x_{1}^{2} + x_{2}^{2} is unknown, and {\sigma _{f_{2}}}(x,z,w,u) is \begin{align*}&\hspace{-0.5pc}{\sigma _{f_{2}}}(x,z,w,u) = (1 - x_{1}^{2}){x_{2}} + \tanh ({x_{1}} + z + u + 3) \\&\qquad \qquad \qquad \qquad \qquad \quad \qquad \qquad \,\,+\, \tanh (u - 3) + w,\end{align*}
View SourceRight-click on figure for MathML and additional features.
where w = 0.1\sin (t) denotes external disturbances.

The target system of (16) is \begin{align*} {\dot r_{1}}=&{r_{2}} + {f_{1}}({x_{1}}),\quad {r_{1}}(0) = 0.1, \\[3pt] {\dot r_{2}}=&{f_{r2}}(t),\quad {r_{2}}(0) = 0,\tag{17}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{align*} {f_{r2}}(t)=&- 25{r_{1}} - 10{r_{2}} + {u_{r}}(t), \\[3pt] {u_{r}}(t)=&{e^{ - 0.2t}}\sin (t + 2) + {e^{ - 0.2t}}\cos (2t + 10) \\[3pt]&+\, {e^{ - 0.2t}}\sin (3t + 5).\end{align*}
View SourceRight-click on figure for MathML and additional features.

To make the system (16) follow the target system (17), the proposed LADRC can be designed as \begin{align*} {\dot z_{2}}=&- \frac {\alpha _{1}}{\varepsilon }\left ({{z_{2} + 0.2\sin ({x_{1}})} }\right) + {z_{3}} - {x_{1}} + 0.01u_{s} \\[3pt]&+\, \left ({{\frac {\alpha _{2}}{\varepsilon ^{2}} - \frac {\alpha _{1}^{2}}{\varepsilon ^{2}}} }\right){x_{1}}, \\ {\hat x_{2}}=&{z_{2}} + \frac {\alpha _{1}}{\varepsilon }{x_{1}},\quad {z_{2}}(0) = - \frac {\alpha _{1}}{\varepsilon }{x_{1}}(0), \\[3pt] {\dot z_{3}}=&- \frac {\alpha _{2}}{\varepsilon ^{2}}\left ({{z_{2} + 0.2\sin ({x_{1}})} }\right) - \frac {{\alpha _{1}{\alpha _{2}}}}{\varepsilon ^{3}}{x_{1}}, \\[3pt] {\hat \sigma _{f_{2}}}=&{z_{3}} + \frac {\alpha _{2}}{\varepsilon ^{2}}{x_{1}},\quad {z_{3}}(0) = - \frac {\alpha _{2}}{\varepsilon ^{2}}{x_{1}}(0), \\[3pt] \mu \dot u=&- K\left [{ { - {x_{1}} + 0.01u_{s} + {\hat \sigma _{f_{2}s}} - {f_{r2}}(t) + L{\hat e_{s}}} }\right], \\&u(0) = 0,\tag{18}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{align*} L{\hat e_{s}}=&{\ell _{1}}({x_{1}} - {r_{1}}) + {\ell _{2}}({\hat x_{2s}} - {r_{2}}), \\ {\hat x_{2s}}=&{M_{2}}sat\left ({{\frac {\hat x_{2}}{M_{2}}} }\right),\quad {\hat \sigma _{f_{2}s}} = {M_\sigma }sat\left ({{\frac {{{\hat \sigma _{f_{2}}}}}{M_\sigma }} }\right), \\ u_{s}=&{M_{u}}sat\left ({\frac {u}{M_{u}}}\right).\end{align*}
View SourceRight-click on figure for MathML and additional features.

The parameters of the proposed LADRC in (18) are given in Table 1.

TABLE 1 Parameters of the Proposed LADRC in (18)
Table 1- 
Parameters of the Proposed LADRC in (18)

The indirect dynamic inversion in [23] is introduced to be compared with the proposed LADRC in (18) and can be designed as \begin{align*} {{\dot {\hat x}}_{1}}=&{\hat x_{2}} + 0.2\sin ({\hat x_{1}}) + \frac {\alpha _{1}}{\varepsilon }({x_{1}} - {\hat x_{1}}),\quad {\hat x_{1}}(0) = 0, \\ {{\dot {\hat x}}_{2}}=&{\hat \sigma _{f_{2}}}(t) + \frac {\alpha _{2}}{\varepsilon ^{2}}({x_{1}} - {\hat x_{1}}),\quad {\hat x_{2}}(0) = 0, \\ {{\dot {\hat \sigma } }_{f_{2}}}=&\frac {\alpha _{3}}{\varepsilon ^{3}}({x_{1}} - {\hat x_{1}}),\quad {\hat \sigma _{f_{2}}}(0) = 0, \\ \dot \chi=&- \frac {1}{\mu }K\left [{ { - {f_{r2}}(t) + {\ell _{1}}({\hat x_{1s}} - {r_{1}}) + {\ell _{2}}({\hat x_{2s}} - {r_{2}})} }\right], \\&\chi (0) = 0, \\ u=&\chi - \frac {1}{\mu }K\left ({{\hat x_{2s}(t)} - {x}_{2}(0) }\right),\tag{19}\end{align*}

View SourceRight-click on figure for MathML and additional features. where the parameters are given in Table 2. One thing to be noticed is that the initial state x_{2}(0) is needed.

TABLE 2 Parameters of the Indirect Dynamic Inversion in (19)
Table 2- 
Parameters of the Indirect Dynamic Inversion in (19)

The nonlinear active disturbance rejection control (NLADRC) in [22] is also introduced to be compared with the proposed LADRC in (18) and can be designed as \begin{align*} {{\dot {\hat x}}_{1}}=&{\hat x_{2}} + 0.2\sin ({\hat x_{1}}) + \frac {\alpha _{1}}{\varepsilon }({x_{1}} - {\hat x_{1}}) + \varepsilon \varphi \left ({{\frac {{x_{1} - {\hat x_{1}}}}{\varepsilon ^{2}}} }\right), \\&{\hat x_{1}}(0) = 0{{,}} \\ {{\dot {\hat x}}_{2}}=&{\hat \sigma _{f_{2}}}(t) + \frac {\alpha _{2}}{\varepsilon ^{2}}({x_{1}} - {\hat x_{1}}),\quad {\hat x_{2}}(0) = 0, \\ {{\dot {\hat \sigma }}_{f_{2}}}=&\frac {\alpha _{3}}{\varepsilon ^{3}}({x_{1}} - {\hat x_{1}}),\quad {\hat \sigma _{f_{2}}}(0) = 0, \\ \mu \dot u=&- K\left [{ { - {\hat x_{1s}} + 0.01{u_{s}} + {\hat \sigma _{f_{2}s}} - {f_{r2}}(t) + L{\hat e_{s}}} }\right], \\&u(0) = 0,\tag{20}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{align*} \varphi (v)=&\begin{cases} { - \dfrac {1}{4},}&{v < - \dfrac {\pi }{2},}\\ {\dfrac {1}{4}\sin (v),}&{ - \dfrac {\pi }{2} \le v \le \dfrac {\pi }{2},}\\ {\dfrac {1}{4},}&{v > \dfrac {\pi }{2},} \end{cases}\\ L{\hat e_{s}}=&{\ell _{1}}({\hat x_{1s}} - {r_{1}}) + {\ell _{2}}({\hat x_{2s}} - {r_{2}}),\\ {\hat x_{1s}}=&{M_{1}}sat\left ({{\frac {\hat x_{1}}{M_{1}}} }\right),\quad {\hat x_{2s}} = {M_{2}}sat\left ({{\frac {\hat x_{2}}{M_{2}}} }\right),\\ {\hat \sigma _{f_{2}s}}=&{M_\sigma }sat\left ({{\frac {{{\hat \sigma _{f_{2}}}}}{M_\sigma }} }\right),\quad {u_{s}} = {M_{u}}sat\left ({{\frac {u}{M_{u}}} }\right),\end{align*}
View SourceRight-click on figure for MathML and additional features.
and the parameters take the same value as those of the indirect dynamic inversion in Table 2.

In the absence of measurement noises, i.e., n(t)=0 , simulation results of the proposed LADRC in (18) are almost indistinguishable from those of the indirect dynamic inversion in (19) and the NLADRC in (20), so they are neglected here. In the presence of measurement noises, n(t) = 2 \times {10^{ - 6}} \times N(t) , where N(t) is the standard Gaussian noise generated by the MATLAB program command randn.

Figs. 1 and 2 show the numerical simulation results of the proposed LADRC, the indirect dynamic inversion, and the NLADRC, respectively. Three performance indices, maximum tracking error {e_{\max }} , root-mean-square error {{RMSE}} , and generalized energy consumption {E} \begin{align*} {e_{\max }}=&\max \left \{{ {\left |{ {x_{1}(t) - {r_{1}}(t)} }\right |} }\right \}, \\ {{RMSE}}=&\sqrt {\frac {1}{t}\int _{0}^{t} {{{\left ({{x_{1}(\tau) - {r_{1}}(\tau)} }\right)}^{2}}d\tau } },\\ E=&\int _{0}^{t} {\left |{ {u(\tau)} }\right |d\tau },\end{align*}

View SourceRight-click on figure for MathML and additional features. are introduced and tabulated in Table 3. From Fig. 1 and Table 3, it is shown that these three approaches can achieve good estimation and regulation of x_{1} . According to Fig. 2, it is obvious that the oscillation of the control law of the proposed LADRC in (18) is smallest among the three approaches. This feature makes the proposed LADRC more advantageous in engineering applications, which will be shown in the next experimental results.

TABLE 3 Performance Indices of the Proposed LADRC in (18), the Indirect Dynamic Inversion in (19), and the NLADRC in (20) in Simulations
Table 3- 
Performance Indices of the Proposed LADRC in (18), the Indirect Dynamic Inversion in (19), and the NLADRC in (20) in Simulations
FIGURE 1. - Estimation and regulation of 
$x_{1}$
 by the proposed LADRC in (18), the indirect dynamic inversion in (19) and the NLADRC in (20) in simulations.
FIGURE 1.

Estimation and regulation of x_{1} by the proposed LADRC in (18), the indirect dynamic inversion in (19) and the NLADRC in (20) in simulations.

FIGURE 2. - Control inputs by the proposed LADRC in (18), the indirect dynamic inversion in (19) and the NLADRC in (20) in simulations.
FIGURE 2.

Control inputs by the proposed LADRC in (18), the indirect dynamic inversion in (19) and the NLADRC in (20) in simulations.

B. Experiments

To show the practical effectiveness of the proposed LADRC, we perform experiments in the linear motor servo system shown in Fig. 3. The linear motor servo system with nonaffine input and uncertainties can be expressed as \begin{align*} {\dot x_{1}}=&{x_{2}}, \\ {\dot x_{2}}=&{\sigma _{f_{2}}}(x,w,u), \\ y=&{x_{1}},\tag{21}\end{align*}

View SourceRight-click on figure for MathML and additional features. where x_{1} is the position of the mover, y is the measured output, u is the control voltage, and unknown {\sigma _{f_{2}}}(x,w,u) is \begin{equation*}{\sigma _{f_{2}}}(x,w,u) = bu + \arctan (u) + d(x),\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where the parameter b > 0 is relative with the load. Since it is difficult to obtain the precise value of b as the working conditions vary, we assume that absolute value of b is unknown here. \arctan (u) is a nonaffine uncertainty term added through software, d(x) denotes the external disturbances, e.g., the friction between the mover and the racks.

FIGURE 3. - Experimental platform of the linear motor servo system.
FIGURE 3.

Experimental platform of the linear motor servo system.

The target system of (21) is \begin{align*} {\dot r_{1}}=&{r_{2}},\quad {r_{1}}(0) = {l_{0}}, \\ {\dot r_{2}}=&{f_{r2}}(t),\quad {r_{2}}(0) = {v_{0}},\tag{22}\end{align*}

View SourceRight-click on figure for MathML and additional features. where {f_{r2}}(t) is a bounded command signal generated by the LSPB (Linear Segment with Parabolic Blend) method (refer to [29] for specific principles of LSPB).

The proposed LADRC for (21) can be designed as \begin{align*} {\dot z_{2}}=&- \frac {\alpha _{1}}{\varepsilon }{z_{2}} + {z_{3}} + \left ({{\frac {\alpha _{2}}{\varepsilon ^{2}} - \frac {\alpha _{1}^{2}}{\varepsilon ^{2}}} }\right){x_{1}}, \\ {\hat x_{2}}=&{z_{2}} + \frac {\alpha _{1}}{\varepsilon }{x_{1}},\quad {z_{2}}(0) = - \frac {\alpha _{1}}{\varepsilon }{x_{1}}(0), \\ {\dot z_{3}}=&- \frac {\alpha _{2}}{\varepsilon ^{2}}{z_{2}} - \frac {{\alpha _{1}{\alpha _{2}}}}{\varepsilon ^{3}}{x_{1}}, \\ {\hat \sigma _{f_{2}}}=&{z_{3}} + \frac {\alpha _{2}}{\varepsilon ^{2}}{x_{1}},\quad {z_{3}}(0) = - \frac {\alpha _{2}}{\varepsilon ^{2}}{x_{1}}(0), \\ \mu \dot u=&- K\left [{ {{\hat \sigma _{f_{2}s}} - {f_{r2}}(t) + {\ell _{1}}({x_{1}} - {r_{1}}) + {\ell _{2}}({\hat x_{2s}} - {r_{2}})} }\right], \\&u(0) = 0,\tag{23}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{align*} {\hat x_{2s}}=&{M_{2}}sat\left ({{\frac {\hat x_{2}}{M_{2}}} }\right),\quad {\hat \sigma _{f_{2}s}} = {M_\sigma }sat\left ({{\frac {{{\hat \sigma _{f_{2}}}}}{M_\sigma }} }\right), \\ {u_{s}}=&{M_{u}}sat\left ({{\frac {u}{M_{u}}} }\right).\end{align*}
View SourceRight-click on figure for MathML and additional features.

The indirect dynamic inversion in [23] for (21) can be expressed as \begin{align*} {{\dot {\hat x}}_{1}}=&{\hat x_{2}} + \frac {\alpha _{1}}{\varepsilon }({x_{1}} - {\hat x_{1}}),\quad {\hat x_{1}}(0) = 0, \\ {{\dot {\hat x}}_{2}}=&{\hat \sigma _{f_{2}}}(t) + \frac {\alpha _{2}}{\varepsilon ^{2}}({x_{1}} - {\hat x_{1}}),\quad {\hat x_{2}}(0) = 0, \\ {{\dot {\hat \sigma } }_{f_{2}}}=&\frac {\alpha _{3}}{\varepsilon ^{3}}({x_{1}} - {\hat x_{1}}),\quad {\hat \sigma _{f_{2}}}(0) = 0, \\ \dot \chi=&- \frac {1}{\mu }K\left [{ { - {f_{r2}}(t) + {\ell _{1}}({\hat x_{1s}} - {r_{1}}) + {\ell _{2}}({\hat x_{2s}} - {r_{2}})} }\right], \\&\chi (0) = 0, \\ u=&\chi - \frac {1}{\mu }K\left ({{{\hat x_{2s}}(t) - {x_{2}}(0)} }\right),\tag{24}\end{align*}

View SourceRight-click on figure for MathML and additional features. where the initial state x_{2}(0) = 0 is needed.

In addition, the conventional PID controller is introduced and can be expressed as \begin{equation*} u = {k_{p}}{e_{1}} + {k_{d}}{\dot e_{1}} + {k_{i}}\int _{0}^{t} {e_{1}(\tau)d\tau },\tag{25}\end{equation*}

View SourceRight-click on figure for MathML and additional features. where {e_{1}} = {r_{1}} - {x_{1}} and since numerical differentiation usually introduces significant high-frequency noise in velocity measurements, we estimate \dot e_{1} from e_{1} by utilizing a derivative filter given by \begin{equation*} {{DF}}(s) = \frac {\omega _{c}^{2}s}{{s^{2} + 2\zeta {\omega _{c}}s + \omega _{c}^{2}}},\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where \zeta and \omega _{c} are the damping ratio and the cutoff frequency of the filter, respectively. In our experiments, \zeta = 0.707,{\omega _{c}} = 80 .

In order to compare these three approaches fairly, their parameters are tuned to produce similar tracking performance without the nonaffine term as shown in Fig. 4. The designed control parameters of the LADRC in (23) are chosen in Table 4, and the parameters of the indirect dynamic inversion in (24) are given in Table 5. The parameters of the PID in (25) are selected in Table 6.

TABLE 4 Parameters of the Proposed LADRC in (23)
Table 4- 
Parameters of the Proposed LADRC in (23)
TABLE 5 Parameters of the Indirect Dynamic Inversion in (24)
Table 5- 
Parameters of the Indirect Dynamic Inversion in (24)
TABLE 6 Parameters of the PID in (25)
Table 6- 
Parameters of the PID in (25)
FIGURE 4. - Experimental results with the proposed LADRC in (23), the indirect dynamic inversion in (24), and the PID in (25) that are tuned to achieve similar performance without the nonaffine uncertainty term.
FIGURE 4.

Experimental results with the proposed LADRC in (23), the indirect dynamic inversion in (24), and the PID in (25) that are tuned to achieve similar performance without the nonaffine uncertainty term.

Fig. 5 shows experimental results of the proposed LADRC, the indirect dynamic inversion, and the PID, respectively. Three performance indices, maximum tracking error {e_{\max }} , root-mean-square error {{RMSE}} , and generalized energy consumption {E} tabulated in Table 7. From Fig. 5 and Table 7, the maximum tracking error {e_{\max }} and the generalized energy consumption {E} of the proposed LADRC is the smallest among these three controllers, but its {{RMSE}} is a little larger than that of the indirect dynamic inversion and much less than that of the PID. Therefore, it can be concluded that the proposed LADRC in (23) shows overall a better performance than other two controllers, i.e., the indirect dynamic inversion in (24), and the PID in (25).

TABLE 7 Performance Indices of the Proposed LADRC in (23), the Indirect Dynamic Inversion in (24), and the PID in (25)
Table 7- 
Performance Indices of the Proposed LADRC in (23), the Indirect Dynamic Inversion in (24), and the PID in (25)
FIGURE 5. - Performance of the proposed LADRC in (23), the indirect dynamic inversion in (24), and the PID in (25) with the nonaffine uncertainty term in experiments.
FIGURE 5.

Performance of the proposed LADRC in (23), the indirect dynamic inversion in (24), and the PID in (25) with the nonaffine uncertainty term in experiments.

SECTION VI.

Conclusion

In this paper, we propose a LADRC design method for the output feedback control of a class of uncertain, nonaffine, strict-feedback nonlinear systems. A RLESO is designed to estimate the unmeasured states and uncertainties, and dynamic inversion is used to deal with the nonaffine inputs with sector conditions for the inputs. The singular perturbation method is used to analyze the stability and performance of the closed-loop system, and the comparison simulation and experimental results with the existing methods demonstrates the effectivpeness of the proposed LADRC.

Although the considered system in this paper is nonaffine, for affine nonlinear systems, the proposed LADRC provides an extra ADRC design method when the choice of nominal value of control gain is difficult in the classical ADRC framework.

Appendix

Proof of Theorem 1

The main steps of the proof are outlined in three steps:

  1. Firstly, all trajectories (e,z,s,\eta) starting from {\Omega _{a}} \times {Q_\eta } , i.e., ({e_{0}},{z_{0}},{s_{0}},{\eta _{0}}) \in {\Omega _{a}} \times {Q_\eta } , will enter the set {\Omega _{b}} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} in finite time.

  2. Secondly, all trajectories (e,z,s,\eta) starting from {\Omega _{b}} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} will enter the set \left \{{ {V_{e} \le {b_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} in finite time.

  3. Lastly, all trajectories (e,z,s,\eta) starting from \left \{{ {V_{e} \le {b_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} will enter the set \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} .

In the first step, since (e,z,s,\hat x,\hat \sigma) start from ({e_{0}}, {z_{0}},{s_{0}},{\hat x_{0}},{\hat \sigma _{0}}) \in {\Omega _{a}} \times {Q_{x}} \times {Q_{s}} , there exist a compact set {Q_\eta } such that {\eta _{0}} \in {Q_\eta } . From the definition (9), \eta (0) = o(1/{\varepsilon ^{n}}) . According to Assumption 1, Assumption 2, and (11), there exist positive constants {l_{1}},{l_{2}} > 0 such that \begin{equation*} \left \|{ {{\delta _{f/\varepsilon }} + {B}(\varepsilon /\mu){\psi _{1}}} }\right \| \le \varepsilon {l_{1}}\left \|{ \eta }\right \| + \frac {\varepsilon }{\mu }{l_{2}},\tag{26}\end{equation*}

View SourceRight-click on figure for MathML and additional features. where {l_{1}} = \sum \limits _{j = 1}^{n} {j k_{j}} .

Let \begin{equation*}{V_\eta } = {\eta ^{T}}{P_\eta }\eta,\end{equation*}

View SourceRight-click on figure for MathML and additional features. where {P_\eta } is the solution of the Lyapunov function {P_\eta }\Lambda + {\Lambda ^{T}}{P_\eta } = - I . Based on (10), we have \begin{align*} \varepsilon {\dot V_\eta }=&- {\left \|{ \eta }\right \|^{2}} + 2{\eta ^{T}}{P_\eta }\left ({{{\delta _{f/\varepsilon }} + {B_{1}}(\varepsilon /\mu){\psi _{1}}} }\right) \\\le&- {\left \|{ \eta }\right \|^{2}} + 2{l_{3}}\left \|{ \eta }\right \|\left ({{\varepsilon {l_{1}}\left \|{ \eta }\right \| + \frac {\varepsilon }{\mu }{l_{2}}} }\right) \\=&- {\left \|{ \eta }\right \|^{2}} + \varepsilon {l_{4}}{\left \|{ \eta }\right \|^{2}} + \frac {\varepsilon }{\mu }{l_{5}}\left \|{ \eta }\right \|,\tag{27}\end{align*}
View SourceRight-click on figure for MathML and additional features.
where {l_{3}} \stackrel { \wedge } = {\lambda _{\max }}({P_\eta }),{l_{4}} \stackrel { \wedge } = 2{l_{1}}{l_{3}},{l_{5}} \stackrel { \wedge } = 2{l_{2}}{l_{3}} .

When \varepsilon \le \frac {1}{{2{l_{4}}}} , \begin{equation*}\varepsilon {\dot V_\eta } \le - \frac {1}{2}{\left \|{ \eta }\right \|^{2}} + \frac {\varepsilon }{\mu }{l_{5}}\left \|{ \eta }\right \|,\end{equation*}

View SourceRight-click on figure for MathML and additional features. then, there exist positive constants {l_{6}},{l_{7}} > 0 such that \begin{equation*} \varepsilon {\dot V_\eta } \le - {l_{6}}{V_\eta },\quad for~{V_\eta } \ge {\left ({{\frac {\varepsilon }{\mu }} }\right)^{2}}{l_{7}},\tag{28}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where \begin{equation*}{l_{6}} < \frac {1}{{2{l_{3}}}},\quad {l_{7}} = {\left ({{\frac {{2\sqrt {l_{3}} {l_{5}}}}{{1 - 2{l_{6}}{l_{3}}}}} }\right)^{2}}.\end{equation*}
View SourceRight-click on figure for MathML and additional features.

Taking into consideration that there is {l_{8}} \in {\mathbb {R}^ {+} } such that \left \|{ {\eta (0)} }\right \| \le {l_{8}}/{\varepsilon ^{n}} , so

  1. if \eta (0) \in \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \} , then \begin{equation*}\eta (t) \in \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \},\quad t \in \left [{ {0,\infty } }\right [,\end{equation*}

    View SourceRight-click on figure for MathML and additional features.

  2. if \eta (0) \notin \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \} , then according to (28), we have:\begin{equation*} {V_\eta }(t) \le {V_\eta }(0){e^{ - \frac {l_{6}}{\varepsilon }t}} \le \frac {l_{9}}{{{\varepsilon ^{2n}}}}{e^{ - \frac {l_{6}}{\varepsilon }t}},\tag{29}\end{equation*}

    View SourceRight-click on figure for MathML and additional features. where {l_{9}} = {l_{3}}l_{8}^{2} .

To estimate the time {T_{1}}(\varepsilon) that \eta enters the set \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \} , let {V_\eta } = {\left ({{\frac {\varepsilon }{\mu }} }\right)^{2}}{l_{7}} , then according to (29), we have \begin{equation*}\frac {l_{9}}{{{\varepsilon ^{2n}}}}{e^{ - \frac {l_{6}}{\varepsilon }{T_{1}}}} = {\left ({{\frac {\varepsilon }{\mu }} }\right)^{2}}{l_{7}} \Leftrightarrow {T_{1}} = \frac {\varepsilon }{l_{6}}\ln \left ({{\frac {{l_{9}{\mu ^{2}}}}{{l_{7}{\varepsilon ^{2(n + 1)}}}}} }\right).\end{equation*}

View SourceRight-click on figure for MathML and additional features. By L’hopital’s rule, it can be shown \begin{align*} \lim \limits _{\varepsilon \to 0} {T_{1}}(\varepsilon)=&\lim \limits _{\varepsilon \to 0} \frac {\varepsilon }{l_{6}}\ln \left ({{\frac {{l_{9}{\mu ^{2}}}}{{l_{7}{\varepsilon ^{2(n + 1)}}}}} }\right) \\\le&\frac {{\ln ({l_{9}}/{l_{7}})}}{l_{6}}\lim \limits _{\varepsilon \to 0} \varepsilon - \frac {2(n + 1)}{l_{6}}\lim \limits _{\varepsilon \to 0} \varepsilon \ln \varepsilon \\=&0.\tag{30}\end{align*}
View SourceRight-click on figure for MathML and additional features.

To summarize, we have \begin{equation*} {T_{1}}(\varepsilon) = \begin{cases} {0,}&{\eta (0)\; \in \left \{{ {V_\eta \le {{\left ({{\dfrac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \},}\\ {0,}&{\varepsilon \to 0,\eta (0)\; \notin \left \{{ {V_\eta \le {{\left ({{\dfrac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \}.} \end{cases}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

Because {a_{1}} < {b_{1}},{a_{2}} < {b_{2}} , (e,z,s) -subsystems are bounded uniformly in \varepsilon , there is time {T_{0}} such that (e,z,s) \in {\Omega _{b}},\;{{for}}\;t \in \left [{ {0,{T_{0}}} }\right] . By choosing \varepsilon small enough, we can have {T_{1}}(\varepsilon) = \frac {1}{2}{T_{0}} .

In the second step, the s -subsystem in (10) is a perturbation of the s -subsystem in (13) with the perturbation term bounded by {l_{10}}\left \|{ \eta }\right \| + {l_{11}}\mu for {l_{10}},{l_{11}} > 0 . Because \eta (t)\; \in \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \},t \in \left [{ {T_{1}(\varepsilon),\infty } }\right [ , \begin{equation*} {l_{10}}\left \|{ \eta }\right \| + {l_{11}}\mu \le {l_{12}}\lambda,\tag{31}\end{equation*}

View SourceRight-click on figure for MathML and additional features. where \begin{equation*}{l_{12}} = {l_{10}}\sqrt {\frac {l_{7}}{{{\lambda _{\min }}({P_\eta })}}} + {l_{11}},\quad \lambda = \max \left \{{ {\mu,\frac {\varepsilon }{\mu }} }\right \}.\end{equation*}
View SourceRight-click on figure for MathML and additional features.

Based on (31), we have \begin{align*}&\hspace {-1pc} \mu {\dot V_{s}} = - s(KF - \mu {\phi _{d}}) \le - \beta {s^{2}} + \lambda {l_{12}}s \le - {l_{13}}{V_{s}}, \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad for\;{V_{s}} \ge {l_{14}}{\lambda ^{2}}.\tag{32}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{equation*}{l_{13}} \in \left]{ {0,2\beta } }\right [,{l_{14}} = \frac {1}{2}{\left ({{\frac {{{l_{12}}}}{{\beta - {l_{13}}/2}}} }\right)^{2}}.\end{equation*}
View SourceRight-click on figure for MathML and additional features.

Next we estimate the time T_{2} that s(t) enters \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} .

  1. If s(0)\; \in \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} , then \begin{equation*}s(t)\; \in \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \},\quad t \in \left [{ {T_{1}(\varepsilon),\infty } }\right [.\end{equation*}

    View SourceRight-click on figure for MathML and additional features.

  2. If s(0)\; \notin \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} , based on (32), we have \begin{equation*}{V_{s}}(t) \le {V_{s}}(0){e^{ - \frac {{{l_{13}}}}{\mu }t}}.\end{equation*}

    View SourceRight-click on figure for MathML and additional features.

    • If \lambda = \mu \ge \frac {\varepsilon }{\mu } , let {V_{s}} = {l_{14}}{\mu ^{2}} , then \begin{align*} \displaystyle {V_{s}}(0){e^{ - \frac {{{l_{13}}}}{\mu }{T_{2}}}}=&{\mu ^{2}}{l_{14}} \Leftrightarrow \\ \displaystyle \quad {T_{2}}(\mu)=&\frac {\mu }{{{l_{13}}}}\ln \left ({{\frac {V_{s}(0)}{{{l_{14}}{\mu ^{2}}}}} }\right),\end{align*}

      View SourceRight-click on figure for MathML and additional features. by L’hopital’s rule, it can be shown \begin{equation*}\lim \limits _{\mu \to 0} {T_{2}}(\mu) = \lim \limits _{\mu \to 0} \frac {\mu }{{{l_{13}}}}\ln \left ({{\frac {V_{s}(0)}{{{l_{14}}{\mu ^{2}}}}} }\right) = 0.\end{equation*}
      View SourceRight-click on figure for MathML and additional features.

    • If \lambda = \frac {\varepsilon }{\mu } \ge \mu , let {V_{s}} = {l_{14}}{\left ({{\frac {\varepsilon }{\mu }} }\right)^{2}} , then \begin{align*} {V_{s}}(0){e^{ - \frac {{{l_{13}}}}{\mu }{T_{2}}}}=&{l_{14}}{\left ({{\frac {\varepsilon }{\mu }} }\right)^{2}} \Leftrightarrow \\ {T_{2}}(\mu)=&\frac {\mu }{{{l_{13}}}}\ln \left ({{\frac {{V_{s}(0){\mu ^{2}}}}{{{l_{14}}{\varepsilon ^{2}}}}} }\right),\end{align*}

      View SourceRight-click on figure for MathML and additional features. by L’hopital’s rule, \begin{align*} \lim \limits _{\mu \to 0} {T_{2}}(\mu)=&\lim \limits _{\mu \to 0} \frac {\mu }{{{l_{13}}}}\ln \left ({{\frac {{V_{s}(0){\mu ^{2}}}}{{{l_{14}}{\varepsilon ^{2}}}}} }\right) \\\le&\lim \limits _{\mu \to 0} \frac {\mu }{{{l_{13}}}}\ln \left ({{\frac {V_{s}(0)}{{{l_{14}}{\mu ^{2}}}}} }\right) = 0.\end{align*}
      View SourceRight-click on figure for MathML and additional features.

To summarize, we have \begin{equation*}{T_{2}}(\mu) = \begin{cases} {0,}&{s(0)\; \in \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \},}\\ {0,}&{\mu \to 0,s(0)\; \notin \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \}.} \end{cases}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

By choosing \mu small enough, we can ensure that (e,z) still stays in \left \{{ {V_{e} \le {b_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} . Therefore, after the time interval {T_{1}}(\varepsilon) + {T_{2}}(\mu) , (e,z,s,\eta) will enter \left \{{ {V_{e} \le {b_{1}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} which is positively invariant for sufficiently small \varepsilon and \mu .

Lastly, the x -subsystem in (10) is a perturbation of the x -subsystem in (14) with the perturbation term bounded by {\varepsilon ^{2}}\sum \limits _{j = 1}^{n - 1} {j k_{j}} \left \|{ \eta }\right \| + {k_{u}}\left |{ s }\right | . Because \begin{align*}&\hspace{-0.8pc}\eta (t)\; \in \left \{{ {V_\eta \le {{\left ({{\frac {\varepsilon }{\mu }} }\right)}^{2}}{l_{7}}} }\right \},\quad s(t)\; \in \;\left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \}, \\&\qquad \qquad \qquad \qquad \qquad \qquad \quad ~~ t \in \left [{ {T_{1}(\varepsilon) + {T_{2}}(\mu),\infty } }\right [\end{align*}

View SourceRight-click on figure for MathML and additional features. we can obtain \begin{equation*}{\varepsilon ^{2}}\sum \limits _{j = 1}^{n - 1} {j k_{j}} \left \|{ \eta }\right \| + {k_{u}}\left |{ s }\right | < \lambda {l_{15}},\end{equation*}
View SourceRight-click on figure for MathML and additional features.
where \begin{equation*}{l_{15}} = \left ({{\sum \limits _{j = 1}^{n - 1} {j k_{j}} \sqrt {\frac {l_{7}}{{{\lambda _{\min }}({P_\eta })}}} + {k_{u}}\sqrt {2{l_{14}}} } }\right).\end{equation*}
View SourceRight-click on figure for MathML and additional features.

The derivative of {V_{e}} can be expressed as \begin{align*}&\hspace{-1.2pc} {\dot V_{e}} \le - {\left \|{ e }\right \|^{2}} + 2{\lambda _{\max }}({P_{e}})\lambda {l_{15}}\left \|{ e }\right \| \le - {l_{16}}{V_{e}}, \\&\qquad \qquad \qquad \qquad \qquad \qquad \qquad for\;{V_{e}} \ge {l_{17}}{\lambda ^{2}},\tag{33}\end{align*}

View SourceRight-click on figure for MathML and additional features. where \begin{equation*}{l_{16}} \in \left]{ {0,\frac {1}{{{\lambda _{\max }}({P_{e}})}}} }\right [,{l_{17}} = \frac {{4\lambda _{\max }^{3}({P_{e}})l_{15}^{2}}}{{{{(1 - {l_{16}}{\lambda _{\max }}({P_{e}}))}^{2}}}}.\end{equation*}
View SourceRight-click on figure for MathML and additional features.

Next we estimate the time T_{3} that e(t) enters the set \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \} .

  1. If e(0) \in \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \} , then \begin{equation*}e(t) \in \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \},\quad t \in \left [{ {T_{1}(\varepsilon) + {T_{2}}(\mu),\infty } }\right [.\end{equation*}

    View SourceRight-click on figure for MathML and additional features.

  2. If e(0) \notin \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \} , based on (33), we have \begin{equation*}{V_{e}}(t) \le {V_{e}}(0){e^{ - {l_{16}}t}}.\end{equation*}

    View SourceRight-click on figure for MathML and additional features. Let {V_{e}} = {l_{17}}{\lambda ^{2}} , we obtain \begin{equation*}{V_{e}}(0){e^{ - {l_{16}}{T_{3}}}} = {l_{17}}{\lambda ^{2}} \Leftrightarrow {T_{3}}(\lambda) = \frac {1}{{{l_{16}}}}\ln \left ({{\frac {V_{e}(0)}{{{l_{17}}{\lambda ^{2}}}}} }\right).\end{equation*}
    View SourceRight-click on figure for MathML and additional features.

To summarize, we have \begin{equation*}{T_{3}}(\lambda) = \begin{cases} {0,}&{e(0) \in \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \},}\\ {\infty,}&{\lambda \to 0,e(0) \notin \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \},} \end{cases}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

Therefore, (e,z,s,\eta) will enter the positively invariant set \left \{{ {V_{e} \le {l_{17}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_{z} \le {c_{0}}} }\right \} \times \left \{{ {V_{s} \le {l_{14}}{\lambda ^{2}}} }\right \} \times \left \{{ {V_\eta \le {l_{7}}{(\varepsilon /\mu)^{2}}} }\right \} in the time {T_{1}}(\varepsilon) + {T_{2}}(\mu) + {T_{3}}(\lambda) for sufficiently small \varepsilon , \mu , and \lambda . Q.E.D.

References is not available for this document.

References

References is not available for this document.