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

Integrated Adaptive Control and Reference Governors for Constrained Systems with State-Dependent Uncertainties

Pan Zhao1,∗,  Ilya Kolmanovsky2,  Naira Hovakimyan1 This work is supported by AFOSR, NASA and NSF under the NRI grant #1830639, CPS grant #1932529, and AI Institute: Planning grant #2020289.1P. Zhao and N. Hovakimyan are with the Department of Mechanical Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA. Email: {panzhao2, nhovakim}@illinois.edu. Corresponding author: P. Zhao.2I. Kolmanovsky is with the Department of Aerospace Engineering, University of Michigan, Ann Arbor, MI 48109, USA. Email: [email protected].
Abstract

This paper presents an adaptive reference governor (RG) framework for a linear system with matched nonlinear uncertainties that can depend on both time and states, subject to both state and input constraints. The proposed framework leverages an 1{\mathcal{L}_{1}} adaptive controller (1{\mathcal{L}_{1}}AC) that estimates and compensates for the uncertainties, and provides guaranteed transient performance, in terms of uniform bounds on the error between actual states and inputs and those of a nominal (i.e., uncertainty-free) system. The uniform performance bounds provided by the 1{\mathcal{L}_{1}}AC are used to tighten the pre-specified state and control constraints. A reference governor is then designed for the nominal system using the tightened constraints, and guarantees robust constraint satisfaction. Moreover, the conservatism introduced by the constraint tightening can be systematically reduced by tuning some parameters within the 1{\mathcal{L}_{1}}AC. Compared with existing solutions, the proposed adaptive RG framework can potentially yield less conservative results for constraint enforcement due to the removal of uncertainty propagation along a prediction horizon, and improved tracking performance due to the inherent uncertainty compensation mechanism. Simulation results for a flight control example illustrate the efficacy of the proposed framework.

Index Terms:
Constrained Control; Robust Adaptive control; Uncertain Systems; Reference Governor

I Introduction

There has been a growing interest in developing control methods that can handle state and/or input constraints. Examples of such constraints include actuator magnitude and rate limits, bounds imposed on process variables to ensure safe and efficient system operation, and collision/obstacle avoidance requirements. There are several choices for a control practitioner when dealing with constraints. One choice is to adopt the model predictive control (MPC) framework [1, 2], in which the state and input constraints can be incorporated into the optimization problem for computing the control signals. Another route is to augment a well-designed nominal controller, that already achieves high performance for small signals, with constraint handling capability that protects the system against constraint violations in transients for large signals. The second route is attractive to practitioners who are interested in preserving an existing/legacy controller or are concerned with the computational cost, tuning complexity, stability, robustness, certification issues, and/or other requirements satisfactorily addressed by the existing controller. The reference governor (RG) is an example of the second approach. As its name suggests, RG is an add-on scheme for enforcing pointwise-in-time state and control constraints by modifying the reference command to a well-designed closed-loop system. The RG acts like a pre-filter that, based on the current value of the desired reference command r(t)r(t) and of the states (measured or estimated) x(t)x(t), generates a modified reference command v(t)v(t) which avoids constraint violations. Since its advent, variants of RGs have been proposed for both linear and nonlinear systems. See the survey paper [3] and references therein. While RG has been extensively studied for systems for which exact dynamic models are available, the design of RG for uncertain systems, i.e., systems with unknown parameters, state-dependent uncertainties, unmodelled dynamics and/or external disturbances, has been less addressed.

I-A Related Work

Robust Approaches: As mentioned in [3], the RG can be straightforwardly modified to handle unmeasured set-bounded disturbances by taking into account all possible realizations of the disturbances when determining the maximal output admissible set [4]. For uncertain systems, various robust or tube MPC schemes have also been proposed [5, 6, 7, 8, 9, 10, 11] and summarized in [12], most of which consider parametric uncertainties and bounded disturbances with only a few exceptions (e.g., [10, 11]) that consider state-dependent uncertainties. However, robust approaches often lead to conservative results when the disturbances are large.

Adaptive and uncertainty compensation based approaches could potentially achieve less conservative results than robust approaches. Along these lines, various adaptive MPC strategies with performance guarantees have been proposed for systems with unknown parameters [13, 14, 15] and state-dependent uncertainties [16, 17]. In particular, [15] uses an 1{\mathcal{L}_{1}} adaptive controller [18] to compensate for matched parametric uncertainties so that the uncertain plant behaves close to a nominal model, and uses robust MPC to handle the error between the combined system, consisting of the uncertain plant and the adaptive controller, and the nominal model. To the best of our knowledge, all of the existing adaptive MPC solutions, including [15] involve propagation of uncertainties along a prediction horizon. Reference [19] merged a Lyapunov function based RG with a disturbance cancelling controller based on an input observer to achieve non-conservative treatment of uncertainties. Unfortunately, a bound on the rate of change of the disturbance is needed for the design, which is often difficult to obtain when the disturbance is dependent on states. Additionally, input constraints were not considered in that work.
State-dependent uncertainties (SDUs): If a system is affected by SDUs, and the states are limited to a compact set, it is always possible to bound the SDU with a worst-case value and to apply the robust approaches (e.g., robust or tube MPC [5, 6, 7]) developed for bounded disturbances. However, by accounting for the state dependence, one can improve performance and reduce conservatism, as demonstrated in robust MPC solutions in [20, 11]. Adaptive MPC solutions which account for SDUs have been proposed in [16, 17]. These solutions essentially rely on computing the uncertainty or state bounds along the prediction horizon using the Lipschitz proprieties of SDUs, and solving a robust MPC problem, using the computed bounds.

I-B Contributions

The contributions of this paper are as follows. Firstly, for constrained control under uncertainties, we develop an 1{\mathcal{L}_{1}}-RG framework for linear systems with matched nonlinear uncertainties that could depend on both time and states, and with both input and state constraints. Our adaptive robust RG framework leverages an 1{\mathcal{L}_{1}} adaptive controller (1{\mathcal{L}_{1}}AC) to estimate and compensate for the uncertainties, and to guarantee uniform bounds on the error between actual states and inputs and those of a nominal (i.e., uncertainty-free) closed-loop system. These uniform bounds characterize tubes in which actual states and control inputs are guaranteed to stay despite the uncertainties. A reference governor designed for the nominal system with constraints tightened using these uniform bounds guarantees robust constraint satisfaction in the presence of uncertainties. Additionally, we show that these uniform bounds on state and input errors, and thus the conservatism induced by constraint tightening can be arbitrarily reduced in theory by tuning the filter bandwidth and estimation sample time parameters of the 1{\mathcal{L}_{1}}AC. Secondly, as a separate contribution to 1{\mathcal{L}_{1}} adaptive control, we propose a novel scaling technique that allows deriving separate tight uniform bounds on each state and adaptive control input, as opposed to a single bound for all states, or adaptive control inputs in existing 1{\mathcal{L}_{1}}AC solutions [18]. The ability to provide such separate tight bounds makes an 1{\mathcal{L}_{1}}AC particularly attractive to be integrated with an RG for simultaneous constraint enforcement and improved trajectory tracking. Thirdly, we validate the efficacy of the proposed 1{\mathcal{L}_{1}}-RG framework on a flight control example and we compare it with both baseline and robust RG solutions in simulations.

Compared to existing literature, in particular, robust/adaptive MPC, 1{\mathcal{L}_{1}}-RG has the following novel aspects:

  • Thanks to the uncertainty compensation and transient performance guarantees available for the 1{\mathcal{L}_{1}}AC, 1{\mathcal{L}_{1}}-RG, (under suitable assumptions,) does not require uncertainty propagation along the prediction horizon. This uncertainty propagation is generally required in all existing robust and adaptive MPC approaches, and incurs conservatism, which is avoided by 1{\mathcal{L}_{1}}-RG.

  • 1{\mathcal{L}_{1}}-RG simultaneously improves tracking performance and enforces the constraints, while existing robust/disturbance-observer-based RG or robust/adaptive MPC solutions except a few such as [15, 19], focus on constraint satisfaction only.

  • Within 1{\mathcal{L}_{1}}-RG, the uniform bounds on the state and input errors (used for constraint tightening) and thus the conservatism induced by constraint tightening can be made arbitrarily small, which cannot be achieved by existing methods.

  • 1{\mathcal{L}_{1}}-RG is able to handle uncertainties that can nonlinearly depend on both time and states. Such a case has not been considered by previous adaptive MPC solutions that are based on uncertainty compensation. For instance, the solution in [15], which also leverages an 1{\mathcal{L}_{1}}AC, only treats parametric uncertainties and state constraints.

The paper is structured as follows. Section II formally states the problem. Section III provides an overview of the proposed solution and discusses preliminaries related to RG and 1{\mathcal{L}_{1}}AC design. Section IV introduces a scaling technique to derive separate and tight performance bounds for an 1{\mathcal{L}_{1}}AC, while Section V presents synthesis and performance analysis of the proposed 1{\mathcal{L}_{1}}-RG framework. Section VI includes validation of the proposed 1{\mathcal{L}_{1}}-RG framework on a flight control problem in simulations.

Notations: Let \mathbb{R}, +\mathbb{R}_{+} and +\mathbb{Z}_{+} denote the set of real, non-negative real, and non-negative integer numbers, respectively. n\mathbb{R}^{n} and m×n\mathbb{R}^{m\times n} denote the nn-dimensional real vector space and the set of real mm by nn matrices, respectively. i\mathbb{Z}_{i} and 1n\mathbb{Z}_{1}^{n} denote the integer sets {i,i+1,}\{i,i+1,\cdots\} and {1,2,,n}\{1,2,\cdots,n\}, respectively. InI_{n} denotes an identity matrix of size nn, and 0 is a zero matrix of a compatible dimension. \left\lVert\cdot\right\rVert and \left\lVert\cdot\right\rVert_{\infty} denote the 22-norm and \infty-norm of a vector or a matrix, respectively. The \mathcal{L}_{\infty}- and truncated \mathcal{L}_{\infty}-norm of a function x:+nx:\mathbb{R}_{+}\rightarrow\mathbb{R}^{n} are defined as xsupt0x(t)\left\lVert x\right\rVert_{\mathcal{L}_{\infty}}\triangleq\sup_{t\geq 0}\left\lVert x(t)\right\rVert_{\infty} and x[0,T]sup0tTx(t)\left\lVert x\right\rVert_{\mathcal{L}_{\infty}^{[0,T]}}\triangleq\sup_{0\leq t\leq T}\left\lVert x(t)\right\rVert_{\infty}, respectively. The Laplace transform of a function x(t)x(t) is denoted by x(s)𝔏[x(t)]x(s)\triangleq\mathfrak{L}[x(t)]. For a vector xx, xix_{i} denotes the iith element of xx. Given a positive scalar ρ\rho, Ω(ρ){zn:zρ}\Omega(\rho)\triangleq\{z\in\mathbb{R}^{n}:\left\lVert z\right\rVert_{\infty}\leq\rho\} denotes a high dimensional ball set of radius ρ\rho and centered at the origin, while its dimension nn can be deduced from the context. For a high-dimensional set 𝒳\mathcal{X}, int(𝒳)\textup{int}(\mathcal{X}) denotes the interior of 𝒳\mathcal{X} and 𝒳i\mathcal{X}_{i} denotes the projection of 𝒳\mathcal{X} onto the iith coordinate. For given sets 𝒳,𝒴n\mathcal{X},{\mathcal{Y}}\subset\mathbb{R}^{n}, 𝒳𝒴{x+y:x𝒳,y𝒴}\mathcal{X}\oplus{\mathcal{Y}}\triangleq\{x+y:x\in\mathcal{X},y\in{\mathcal{Y}}\} is the Minkowski set sum and 𝒳𝒴{z:z+y𝒳,y𝒴}\mathcal{X}\ominus{\mathcal{Y}}\triangleq\{z:z+y\in\mathcal{X},\forall y\in{\mathcal{Y}}\} is the Pontryagin set difference.

II Problem statement

Consider an uncertain linear system represented by

{x˙(t)=Ax(t)+B(u(t)+f(t,x(t))),y(t)=Cx(t),x(0)=x0,\left\{\begin{aligned} \dot{x}(t)&=Ax(t)+B(u(t)+f(t,x(t))),\hfill\\ y(t)&=Cx(t),\ x(0)=x_{0},\\ \end{aligned}\right. (1)

where x(t)nx(t)\in\mathbb{R}^{n}, u(t)mu(t)\in\mathbb{R}^{m} and y(t)my(t)\in\mathbb{R}^{m} are the state, input and output vectors, respectively, x0nx_{0}\in\mathbb{R}^{n} is the initial state vector, f(t,x(t))mf(t,x(t))\in\mathbb{R}^{m} denotes the uncertainty that can depend on both time and states, and A,B,A,~{}B, and CC are matrices of compatible dimensions. We want to design a control law for u(t)u(t) such that the output vector y(t)y(t) tracks a reference signal r(t)r(t) while satisfying the specified state and control constraints:

x(t)𝒳,u(t)𝒰,t0,\begin{gathered}x(t)\in\mathcal{X},\quad u(t)\in\mathcal{U},\quad\forall t\geq 0,\hfill\end{gathered} (2)

where 𝒳n\mathcal{X}\subset\mathbb{R}^{n} and 𝒰m{\mathcal{U}}\subset\mathbb{R}^{m} are pre-specified convex and compact sets with 0 in the interior. Note that 2 can also represent constraints on some of the states and/or inputs.

Suppose a baseline controller is available and achieves desired performance for the nominal (i.e., uncertainty-free) system given a small desired reference command r(t)r(t) to track. To enforce state and input constraints 2 for the nominal system with larger signals, one can simply leverage the conventional RG, which will generate a modified reference command v(t)v(t) based on r(t)r(t). In such a case, the baseline controller can be selected as

ub(t)=Kxx(t)+Kvv(t),u_{\textup{b}}(t)=K_{x}x(t)+K_{v}v(t), (3)

where KxK_{x} and KvK_{v} are feedback and feedforward gains. For both improved tracking performance and constraint enforcement in the presence of the uncertainty f(t,x)f(t,x), we leverage an 1{\mathcal{L}_{1}}AC. To this end, we adopt a compositional control law:

u(t)=ub(t)+ua(t),u(t)=u_{\textup{b}}(t)+u_{\textup{a}}(t), (4)

where ua(t)u_{\textup{a}}(t) is the vector of the adaptive control inputs designed to cancel f(t,x)f(t,x). With 3, the uncertain system 1 can be rewritten as

{x˙(t)=Amx(t)+Bvv(t)+B(ua(t)+f(t,x(t))),y(t)=Cx(t),x(0)=x0,\left\{\begin{aligned} \dot{x}(t)&={A_{m}}x(t)+{B_{v}}v(t)+B(u_{\textup{a}}(t)+f(t,x(t))),\\ y(t)&=Cx(t),\ x(0)=x_{0},\end{aligned}\right. (5)

where AmA+BKx{A_{m}}\triangleq A+B{K_{x}} is a Hurwitz matrix and BvBKv{B_{v}}\triangleq B{K_{v}}.

The problem to be tackled can be stated as follows: Given an uncertain system 1, a baseline controller 3 and a desired reference signal r(t)r(t), design a RG (for determining v(t)v(t)) and the 1{\mathcal{L}_{1}}AC for ua(t)u_{\textup{a}}(t) such that the output signal y(t)y(t) tracks r(t)r(t) whenever possible, while the state and input constraints 2 are satisfied. We make the following assumption on the uncertainty.

Assumption 1.

Given a compact set 𝒵{\mathcal{Z}}, there exist known positive constants Lfj,𝒵L_{f_{j},{\mathcal{Z}}}, lfj,𝒵l_{f_{j},{\mathcal{Z}}} and bfj,𝒵b_{f_{j},{\mathcal{Z}}} (j1mj\in\mathbb{Z}_{1}^{m}) such that for any x,z𝒵x,z\in{\mathcal{Z}} and t,τ0t,\tau\geq 0, the following inequalities hold for each j1mj\in\mathbb{Z}_{1}^{m}:

|fj(t,x)fj(τ,z)|\displaystyle\left\lvert f_{j}(t,x)-f_{j}(\tau,z)\right\rvert Lfj,𝒵xz+lfj,𝒵|tτ|,\displaystyle\leq L_{f_{j},{\mathcal{Z}}}\left\lVert x-z\right\rVert_{\infty}+l_{f_{j},{\mathcal{Z}}}\left\lvert t-\tau\right\rvert, (6a)
|fj(t,x)|\displaystyle\left\lvert f_{j}(t,x)\right\rvert bfj,𝒵,\displaystyle\leq b_{f_{j},{\mathcal{Z}}}, (6b)

where fj(t,x)f_{j}(t,x) denotes the iith element of f(t,x)f(t,x).

Remark 1.

Assumption 1 indicates that in the compact set 𝒵{\mathcal{Z}}, fj(t,x)f_{j}(t,x) is Lipschitz continuous with respect to xx with a known Lipschitz constant Lfj,𝒵L_{f_{j},{\mathcal{Z}}}, has a bounded rate of variation lfj,𝒵l_{f_{j},{\mathcal{Z}}} with respect to tt, and is uniformly bounded by a constant bfj,𝒵b_{f_{j},{\mathcal{Z}}}.

In fact, given the local Lipschitz constant Lfj,𝒵L_{f_{j},{\mathcal{Z}}} and the bounded rate of variation lfj,𝒵l_{f_{j},{\mathcal{Z}}}, a uniform bound for fj(t,x)f_{j}(t,x) in 𝒵{\mathcal{Z}} can always be derived if the bound on fj(t,x)f_{j}(t,x^{\ast}) for an arbitrary xx^{\ast} in 𝒵{\mathcal{Z}} and any t0t\geq 0 is known. For instance, assuming we know |fj(t,0)|b0i\left\lvert f_{j}(t,0)\right\rvert\leq b^{i}_{0}, from 6a, we have that |fj(t,x)fj(t,0)|Lfj,𝒵x\left\lvert f_{j}(t,x)-f_{j}(t,0)\right\rvert\leq L_{f_{j},{\mathcal{Z}}}\left\lVert x\right\rVert_{\infty}, which immediately leads to |fj(t,x)|b0i+Lfj,𝒵maxx𝒳x\left\lvert f_{j}(t,x)\right\rvert\leq b_{0}^{i}+L_{f_{j},{\mathcal{Z}}}\max_{x\in\mathcal{X}}\left\lVert x\right\rVert_{\infty}, for any x𝒵x\in{\mathcal{Z}} and t0t\geq 0. In practice, some prior knowledge about the uncertainty (e.g., fjf_{j} depends on only a few instead of all states) may be leveraged to obtain a tighter bound than the preceding one, derived using the Lipschitz continuity and triangular inequalities. This motivates the assumption on the uniform bound in 6b.

Under the conditions in Assumption 1, we immediately obtain that for any x,z𝒵x,z\in{\mathcal{Z}} and t,τ0t,\tau\geq 0,

f(t,x)f(τ,z)\displaystyle\left\lVert f(t,x)-f(\tau,z)\right\rVert_{\infty} Lf,𝒵xz+lf,𝒵|tτ|,\displaystyle\leq{L_{f,{\mathcal{Z}}}}\left\lVert x-z\right\rVert_{\infty}+l_{f,{\mathcal{Z}}}\left\lvert t-\tau\right\rvert, (7a)
f(t,x)\displaystyle\left\lVert f(t,x)\right\rVert_{\infty} bf,𝒵,\displaystyle\leq{b_{f,{\mathcal{Z}}}}, (7b)

where

Lf,𝒵=maxj1mLfj,𝒵,lf,𝒵=maxj1mlfj,𝒵,bf,𝒵=maxj1mbfj,𝒵.L_{f,{\mathcal{Z}}}=\max_{j\in\mathbb{Z}_{1}^{m}}L_{f_{j},{\mathcal{Z}}},\quad l_{f,{\mathcal{Z}}}=\max_{j\in\mathbb{Z}_{1}^{m}}l_{f_{j},{\mathcal{Z}}},\quad b_{f,{\mathcal{Z}}}=\max_{j\in\mathbb{Z}_{1}^{m}}b_{f_{j},{\mathcal{Z}}}. (8)
Remark 2.

Our choice of making assumptions on fj(t,x)f_{j}(t,x) instead of on f(t,x)f(t,x) as in 8 facilitates deriving an individual bound on each state and on each adaptive input (see Section IV for details).

Remark 3.

In principle, given the uniform bound on f(t,x)f(t,x) in 7b obtained from Assumption 1, constraints can be enforced via robust RG or robust MPC approaches that handle bounded disturbances, as discussed in Section I-A. However, when this bound is large, robust approaches can yield overly conservative performance.

III Overview and Preliminaries

In this section, we first present an overview of the proposed 1{\mathcal{L}_{1}}-RG framework and then introduce some preliminary results that provides a foundation for the 1{\mathcal{L}_{1}}-RG framework.

III-A Overview of the 1{\mathcal{L}_{1}}-RG Framework

Figure 1 depicts the proposed 1{\mathcal{L}_{1}}-RG framework. As shown in Fig. 1, 1{\mathcal{L}_{1}}-RG is comprised of two integrated components. The first one is an 1{\mathcal{L}_{1}}AC designed to compensate for the uncertainty f(t,x)f(t,x) and to guarantee uniform bounds on the errors between actual states and inputs, and those of the nominal closed-loop system:

x˙n(t)\displaystyle\dot{x}_{\textup{n}}(t) =Amxn(t)+Bvv(t),xn(0)=x0,\displaystyle={A_{m}}x_{\textup{n}}(t)+{B_{v}}v(t),\ x_{\textup{n}}(0)=x_{0}, (9a)
un(t)\displaystyle u_{\textup{n}}(t) =Kxxn(t)+Kvv(t).\displaystyle=K_{x}x_{\textup{n}}(t)+K_{v}v(t). (9b)

The second component is a RG designed for the nominal system 9a with tightened constraints computed using the uniform bounds guaranteed by the 1{\mathcal{L}_{1}}AC.

Refer to caption
Figure 1: Diagram of the proposed 1{\mathcal{L}_{1}}-RG framework

More formally, we will design the 1{\mathcal{L}_{1}}AC to ensure

x(t)xn(t)𝒳~,u(t)un(t)𝒰~,t0,\displaystyle x(t)-x_{\textup{n}}(t)\in\tilde{\mathcal{X}},\quad u(t)-u_{\textup{n}}(t)\in\tilde{\mathcal{U}},\quad\forall t\geq 0, (10)

where x(t)x(t) and u(t)u(t) are the vectors of states and of the total control inputs of the closed-loop system 5:

u(t)un(t)=Kx(x(t)xn(t))+ua(t),\displaystyle u(t)-u_{\textup{n}}(t)=K_{x}(x(t)-x_{\textup{n}}(t))+u_{\textup{a}}(t), (11)

where u(t)u(t) is given by 4 and un(t)u_{\textup{n}}(t) by 9b, and 𝒳~\tilde{\mathcal{X}} and 𝒰~\tilde{\mathcal{U}} are some pre-computed hyperrectangular sets dependent on the properties of f(t,x)f(t,x) and of the 1{\mathcal{L}_{1}}AC. The details will be given in Theorem 3 in Section III-C. Define

𝒳n𝒳𝒳~,𝒰n𝒰𝒰~.\mathcal{X}_{\textup{n}}\triangleq\mathcal{X}\ominus\tilde{\mathcal{X}},\quad{\mathcal{U}}_{\textup{n}}\triangleq{\mathcal{U}}\ominus\tilde{\mathcal{U}}. (12)

Then for robust constraint enforcement, one just needs to design a RG for the nominal system 9 with tightened constraints given by

xn(t)𝒳n,un(t)𝒰n,t0.x_{\textup{n}}(t)\in\mathcal{X}_{\textup{n}},\ u_{\textup{n}}(t)\in{\mathcal{U}}_{\textup{n}},\quad\forall t\geq 0. (13)

III-B Reference Governor Design for a Nominal System

We now introduce the RG for the nominal system 9 to enforce the constraints 13. We use the discrete-time RG approach of [3] that uses a discrete-time model:

{xn(k+1)=A^mxn(k)+B^vv(k),x(0)=x0,un(k)=Kxxn(k)+Kvv(k),\left\{\begin{aligned} \mathrm{x_{\textup{n}}}(k+1)&=\hat{A}_{m}\mathrm{x_{\textup{n}}}(k)+\hat{B}_{v}\mathrm{v}(k),\ \mathrm{x}(0)=x_{0},\\ \mathrm{u}_{\textup{n}}(k)&=K_{x}\mathrm{x_{\textup{n}}}(k)+K_{v}\mathrm{v}(k),\end{aligned}\right. (14)

where xn(k)\mathrm{x_{\textup{n}}}(k), v(k)\mathrm{v}(k) and un(k)\mathrm{u_{\textup{n}}}(k) denotes the vectors of states, of reference command inputs, and of nominal control inputs, respectively, and A^m\hat{A}_{m} and B^v\hat{B}_{v} are computed from AmA_{m} and BvB_{v} in 9 assuming a sampling time, TdT_{d}. When doing the discretization, we ensure that the discrete-time system 14 has the same states as the continuous-time system at all sampling instants. This can be achieved by using the zero-order hold discretization, since

v(t)=v(kTd),t[kTd,(k+1)Td),v(t)=\mathrm{v}(kT_{d}),\quad\forall t\in[kT_{d},(k+1)T_{d}), (15)

which indicates that v(t)v(t) is piecewise constant. The constraints 13 are imposed in discrete-time as

xn(k)𝒳^n,un(k)𝒰^n,k+,\mathrm{x_{\textup{n}}}(k)\in\hat{\mathcal{X}}_{\textup{n}},\ \mathrm{u_{\textup{n}}}(k)\in\hat{\mathcal{U}}_{\textup{n}},\quad\forall k\in\mathbb{Z}_{+}, (16)

where 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} are tightened versions of 𝒳n\mathcal{X}_{\textup{n}} and 𝒰n{\mathcal{U}}_{\textup{n}}, respectively, introduced to avoid inter-sample constraint violations, and are defined by

𝒳^n\displaystyle\hat{\mathcal{X}}_{\textup{n}} 𝒳n{zn:zν(Td)},\displaystyle\triangleq\mathcal{X}_{\textup{n}}\ominus\left\{z\in\mathbb{R}^{n}:\left\lVert z\right\rVert_{\infty}\leq\nu(T_{d})\right\}, (17a)
𝒰^n\displaystyle\hat{\mathcal{U}}_{\textup{n}} 𝒰n{zm:zKxν(Td)},\displaystyle\triangleq{\mathcal{U}}_{n}\ominus\left\{z\in\mathbb{R}^{m}:\left\lVert z\right\rVert_{\infty}\leq\left\lVert K_{x}\right\rVert_{\infty}\nu(T_{d})\right\}, (17b)

while

ν(Td)maxτ[0,Td]eAmτInmaxx𝒳n,v𝒱x+Am1Bvv,\nu(T_{d})\!\triangleq\!\max_{\tau\in[0,T_{d}]}\!\left\lVert e^{A_{m}\tau}\!\!-\!I_{n}\right\rVert_{\infty}\!\max_{x\in\mathcal{X}_{n},v\in\mathcal{V}}\!\left\lVert x\!+\!A_{m}^{-1}B_{v}v\right\rVert_{\infty}\!, (18)

with 𝒱\mathcal{V} denoting the set of all possible reference commands output by the RG.

The following lemma formally guarantees that no inter-sample constraint violations will happen for the continuous-time system 9 when the constraints for the discrete-time system 21 are satisfied at all sampling instants.

Lemma 1.

Consider the continuous-time system 9 and its discrete-time counterpart 14 that has the same states as 9 at all sampling instants. If for the discrete-time system 14,

xn(k)𝒳^n,un(k)𝒰^n,k+,\mathrm{x_{\textup{n}}}(k)\in\hat{\mathcal{X}}_{\textup{n}},\ \mathrm{u_{\textup{n}}}(k)\in\hat{\mathcal{U}}_{\textup{n}},\quad k\in\mathbb{Z}_{+}, (19)

with 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} defined in 17, then 13 holds for the continuous-time system 9.

Proof.

See Section A-A. ∎

Remark 4.

From 18 and 17, we can see that 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} are close to 𝒳n\mathcal{X}_{\textup{n}} and 𝒰n{\mathcal{U}}_{\textup{n}}, respectively, when TdT_{d} is small. For practical implementation, inter-sample constraint violations may not be a big concern when TdT_{d} is small. Under such case, we can simply set 𝒳^n=𝒳n\hat{\mathcal{X}}_{\textup{n}}=\mathcal{X}_{\textup{n}} and 𝒰^n=𝒰n\hat{\mathcal{U}}_{\textup{n}}={\mathcal{U}}_{\textup{n}}.

Define

ync(k)[xn(k)Kxxn(k)+Kvv(k)]=[InKx]Ccxn(k)+[0Kv]Dcv(k).\mathrm{y}_{\textup{n}}^{c}(k)\!\triangleq\!\begin{bmatrix}\mathrm{x_{\textup{n}}}(k)\\ K_{x}\mathrm{x_{\textup{n}}}(k)\!+\!K_{v}\mathrm{v}(k)\end{bmatrix}\!=\!\underbrace{\begin{bmatrix}I_{n}\\ K_{x}\end{bmatrix}}_{\triangleq\ C_{c}}\mathrm{x_{\textup{n}}}(k)\!+\!\underbrace{\begin{bmatrix}0\\ K_{v}\end{bmatrix}}_{\triangleq\ D_{c}}\mathrm{v}(k). (20)

Then, the constraints 16 can be rewritten as

ync(k)𝒴n𝒳^n×𝒰^n,k+,\mathrm{y}_{\textup{n}}^{c}(k)\in{\mathcal{Y}}_{\textup{n}}\triangleq\hat{\mathcal{X}}_{\textup{n}}\times\hat{\mathcal{U}}_{\textup{n}},\quad\forall k\in\mathbb{Z}_{+}, (21)

where ×\times denotes the cross product.

Remark 5.

In case there are no constraints on certain states and/or inputs, one can remove the rows of CcC_{c} and DcD_{c} defined in 20 corresponding to these states and/or inputs, and adjust the sets 𝒳^n\hat{\mathcal{X}}_{\textup{n}}, 𝒰^n\hat{\mathcal{U}}_{\textup{n}} and 𝒴n{\mathcal{Y}}_{\textup{n}} accordingly.

Similar to most RG schemes, the RG scheme we adopt here computes at each time instant a command v(k)\mathrm{v}(k) such that, if it is constantly applied from the time instant kk onward, the ensuing output will always satisfy the constraints. More formally, we define the maximal output admissible set OO_{\infty} [21] as the set of all states xn\mathrm{x_{\textup{n}}} and inputs v\mathrm{v}, such that the predicted response from the initial state xn\mathrm{x_{\textup{n}}} and with a constant input v\mathrm{v} satisfies the constraints 21, i.e.,

O{(v,xn):y^nc(k|v,xn)𝒴n,k+},O_{\infty}\triangleq\{(\mathrm{v},\mathrm{x_{\textup{n}}}):\hat{\mathrm{y}}_{\textup{n}}^{c}(k|\mathrm{v},\mathrm{x_{\textup{n}}})\in{\mathcal{Y}}_{n},\ \forall k\in\mathbb{Z}_{+}\}, (22)

where for system 14 the output prediction y^nc(k|xn,v)\hat{\mathrm{y}}_{\textup{n}}^{c}(k|\mathrm{x_{\textup{n}}},\mathrm{v}) is given by

y^nc(k|v,xn)=CcA^mkxn+Ccj=1kA^mj1B^vv+Dcv=CcA^mkxn+Cc(InA^m)1(InA^mk)B^vv+Dcv.\displaystyle\hat{\mathrm{y}}_{\textup{n}}^{c}(k|\mathrm{v},\mathrm{x_{\textup{n}}})\!=~{}C_{c}\hat{A}_{m}^{k}\mathrm{x_{\textup{n}}}\!+\!C_{c}\sum_{j=1}^{k}\hat{A}_{m}^{j-1}\hat{B}_{v}\mathrm{v}\!+\!\!D_{c}\mathrm{v}=C_{c}\hat{A}_{m}^{k}\mathrm{x_{\textup{n}}}\!+\!C_{c}(I_{n}\!-\!\!\hat{A}_{m})^{\!-1}(I_{n}\!-\!\!\hat{A}_{m}^{k})\hat{B}_{v}\mathrm{v}\!+\!\!D_{c}\mathrm{v}. (23)

Define O~\tilde{O}_{\infty} as a slightly tightened version of OO_{\infty} obtained by constraining the command v\mathrm{v} so that the associated steady-state output y¯nc=(Dc+Cc(InA^m)1B^v)v\bar{\mathrm{y}}_{\textup{n}}^{c}=(D_{c}+C_{c}(I_{n}\!-\!\hat{A}_{m})^{-1}\hat{B}_{v})\mathrm{v} satisfies constraints with a nonzero (typically small) margin ϵ>0\epsilon>0, i.e.,

O~=OOϵ,{\tilde{O}}_{\infty}=O_{\infty}\cap O^{\epsilon}, (24)

where Oϵ{(v,xn):y¯nc(1ϵ)𝒴n}.O^{\epsilon}\triangleq\{(\mathrm{v},\mathrm{x_{\textup{n}}}):\bar{\mathrm{y}}_{\textup{n}}^{c}\in(1-\epsilon){\mathcal{Y}}_{\textup{n}}\}. Clearly, O~{\tilde{O}}_{\infty} can be made arbitrarily close to OO_{\infty} by decreasing ϵ\epsilon. Based on the currently available state xn(k)\mathrm{x_{\textup{n}}}(k) at an instant kk, the RG computes v(k)\mathrm{v}(k) so that

(v(k),xn(k))O~.(\mathrm{v}(k),\mathrm{x_{\textup{n}}}(k))\in\tilde{O}_{\infty}. (25)

It is proven in [21] that if A^m\hat{A}_{m} is Schur, (A^m,Cc)(\hat{A}_{m},C_{c}) is observable, 𝒴n{\mathcal{Y}}_{\textup{n}} is compact with 0 in the interior, and ϵ>0\epsilon>0 is sufficiently small, then the set O~{\tilde{O}}_{\infty} is finitely determined, i.e., there exists a finite index kk^{\star} such that

O~=O~k={(v,xn):y^nc(k|v,xn)𝒴n,k=0,1,,k}Oϵ.{\tilde{O}}_{\infty}\!=\tilde{O}_{k^{\star}}=\!\{(\mathrm{v},\mathrm{x_{\textup{n}}})\!:\hat{\mathrm{y}}_{\textup{n}}^{c}(k|\mathrm{v},\mathrm{x_{\textup{n}}})\!\in\!{\mathcal{Y}}_{n},\ k=0,1,\dots,k^{\star}\}\cap O^{\epsilon}. (26)

Moreover, O~{\tilde{O}}_{\infty} is positively invariant, which means that if (v(k),xn(k))O~(\mathrm{v}(k),\mathrm{x_{\textup{n}}}(k))\in{\tilde{O}}_{\infty} and v(k)\mathrm{v}(k) is applied to the system at time kk, then (v(k),xn(k+1))O~(\mathrm{v}(k),\mathrm{x_{\textup{n}}}(k+1))\in{\tilde{O}}_{\infty}. Furthermore, if 𝒴n{\mathcal{Y}}_{\textup{n}} is convex, then O~{\tilde{O}}_{\infty} is also convex.

Remark 6.

The process of computing kk^{\star} involves computing sets O~k\tilde{O}_{k} for k=1,2,k=1,2,\dots, and checking the condition O~k=O~k+1\tilde{O}_{k}=\tilde{O}_{k+1}; kk^{\star} is the minimum kk for which this condition holds.

The proposed 1{\mathcal{L}_{1}}-RG framework can leverage most of existing RG schemes developed for uncertainty-free systems. As an illustration and demonstration in Section VI, we choose the scalar RG introduced in [22, 23]. The scalar RG computes at each time instant kk a command v(k)\mathrm{v}(k) which is the best approximation of the desired set-point r(k)\mathrm{r}(k) along the line segment connecting v(k1)\mathrm{v}(k-1) and r(k)\mathrm{r}(k) that ensures (v(k),xn(k))O~(\mathrm{v}(k),\mathrm{x_{\textup{n}}}(k))\in{\tilde{O}}_{\infty}. More specifically, the scalar RG solves at each discrete time kk, the following optimization problem:

κ(k)=maxκ[0,1]\displaystyle\kappa(k)=\max_{\kappa\in[0,1]} κ\displaystyle\kappa (27a)
s.t. v=v(k1)+κ(r(k)r(k1)),\displaystyle\mathrm{v}=\mathrm{v}(k-1)+\kappa(\mathrm{r}(k)-\mathrm{r}(k-1)), (27b)
(v,xn(k))O~,\displaystyle(\mathrm{v},\mathrm{x_{\textup{n}}}(k))\in{\tilde{O}}_{\infty}, (27c)

where κ(k)\kappa(k) is a scalar adjustable bandwidth parameter and v(k)=v(k1)+κ(k)(r(k)v(k1))\mathrm{v}(k)=\mathrm{v}(k-1)+\kappa(k)(\mathrm{r}(k)-\mathrm{v}(k-1)) is the modified reference command to be applied to the system. If there is no danger of constraint violation, κ(k)=1\kappa(k)=1 and v(k)=r(k)\mathrm{v}(k)=\mathrm{r}(k) so that the RG does not interfere with the desired operation of the system. If v(k)=r(k)\mathrm{v}(k)=\mathrm{r}(k) would cause a constraint violation, the value of κ(k)\kappa(k) is decreased by the RG. In the extreme case, κ(k)=0\kappa(k)=0, v(k)=v(k1)\mathrm{v}(k)=\mathrm{v}(k-1), which means that the RG momentarily isolates the system from further variations of the reference command for constraint enforcement. Due to the positive invariance of O~{\tilde{O}}_{\infty}, v(k)=v(k1)\mathrm{v}(k)=\mathrm{v}(k-1) always satisfies the constraints, which ensures recursive feasibility under the condition that at t=0t=0 a command v(0)\mathrm{v}(0) is known such that (v(0),xn(0))O~\left(\mathrm{v}(0),\mathrm{x_{\textup{n}}}(0)\right)\in{\tilde{O}}_{\infty}. Response properties of the scalar RG, including conditions for the finite-time convergence of v(k)\mathrm{v}(k) to r(k)\mathrm{r}(k) are detailed in [23].

III-C 1{\mathcal{L}_{1}} Adaptive Control Design and Uniform Performance Bounds

We now present an 1{\mathcal{L}_{1}}AC that guarantees the bounds in (10), without considering the state and control constraints in 2. We first recall some basic definitions and facts from control theory, and introduce some definitions and lemmas.

Definition 1.

[24, Section III.F] For a stable proper MIMO system (s)\mathcal{H}(s) with input u(t)m\mathrm{u}(t)\in\mathbb{R}^{m} and output y(t)p\mathrm{y}(t)\in\mathbb{R}^{p}, its 1{\mathcal{L}_{1}} norm is defined as

(s)1supx(0)=0,u1y.\left\lVert\mathcal{H}(s)\right\rVert_{\mathcal{L}_{1}}\triangleq\sup_{\mathrm{x}(0)=0,\left\lVert\mathrm{u}\right\rVert_{\mathcal{L}_{\infty}}\leq 1}{\left\lVert\mathrm{y}\right\rVert_{\mathcal{L}_{\infty}}}.\vspace{-2mm} (28)

The following lemma follows directly from Definition 1.

Lemma 2.

For a stable proper MIMO system (s)\mathcal{H}(s) with states x(t)n\mathrm{x}(t)\in\mathbb{R}^{n}, inputs u(t)m\mathrm{u}(t)\in\mathbb{R}^{m} and outputs y(t)p\mathrm{y}(t)\in\mathbb{R}^{p}, under zero initial states, i.e., x(0)=0\mathrm{x}(0)=0, we have y[0,τ](s)1u[0,τ]\left\lVert\mathrm{y}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq\left\lVert\mathcal{H}(s)\right\rVert_{\mathcal{L}_{1}}\left\lVert\mathrm{u}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}, for any τ0\tau\geq 0. Furthermore, for any matrix Tq×p\mathrm{T}\in\mathbb{R}^{q\times p}, we have T(s)T(s)\left\lVert\mathrm{T}\mathcal{H}(s)\right\rVert_{\mathcal{L}_{\infty}}\leq\left\lVert\mathrm{T}\right\rVert_{\infty}\left\lVert\mathcal{H}(s)\right\rVert_{\mathcal{L}_{\infty}}.

A unique feature of an 1\mathcal{L}_{1}AC is a low-pass filter 𝒞(s){\mathcal{C}}(s) (with DC gain 𝒞(0)=Im{\mathcal{C}}(0)=I_{m}) that decouples the estimation loop from the control loop, thereby allowing for arbitrarily fast adaptation without sacrificing the robustness [18]. For simplicity, we can select 𝒞(s){\mathcal{C}}(s) to be a first-order transfer function matrix

𝒞(s)=diag(𝒞1(s),,𝒞m(s)),𝒞j(s)kfj(s+kfj),j1m,{\mathcal{C}}(s)=\textup{diag}({\mathcal{C}}_{1}(s),\dots,{\mathcal{C}}_{m}(s)),\ {\mathcal{C}}_{j}(s)\triangleq\frac{k_{f}^{j}}{(s+k_{f}^{j})},\ j\in\mathbb{Z}_{1}^{m}, (29)

where kfjk^{j}_{f} (j1mj\in\mathbb{Z}_{1}^{m}) is the bandwidth of the filter for the jjth input channel. We now introduce a few notations that will be used later:

xm(s)\displaystyle\mathcal{H}_{xm}(s) (sInAm)1B,xv(s)(sInAm)1Bv,\displaystyle\!\triangleq\!(sI_{n}\!-\!A_{m})^{-1}\!B,\ \mathcal{H}_{xv}(s)\!\triangleq\!(sI_{n}\!-\!A_{m})^{-1}\!B_{v}, (30a)
𝒢xm(s)\displaystyle\mathcal{G}_{xm}(s) xm(s)(Im𝒞(s)),\displaystyle\!\triangleq\!\mathcal{H}_{xm}(s)(I_{m}-{\mathcal{C}}(s)),\ (30b)

where Am,BvA_{m},B_{v} correspond to system 9 and BB to 1. Also, letting xin(t)x_{\textup{in}}(t) be the state of the system x˙in(t)=Amxin(t),xin(0)=x0,\dot{x}_{\textup{in}}(t)=A_{m}x_{\textup{in}}(t),\ x_{\textup{in}}(0)=x_{0}, we have xin(s)(sInAm)1x0x_{\textup{in}}(s)\triangleq(sI_{n}-A_{m})^{-1}x_{0}. Defining ρins(sInAm)11maxx0𝒳0x0\rho_{\textup{in}}\triangleq\left\lVert s(sI_{n}-A_{m})^{-1}\right\rVert_{\mathcal{L}_{1}}\max_{x_{0}\in\mathcal{X}_{0}}\left\lVert x_{0}\right\rVert_{\infty}, and further considering that AmA_{m} is Hurwitz and 𝒳0\mathcal{X}_{0} is compact, we have xinρin\left\lVert x_{\textup{in}}\right\rVert_{\mathcal{L}_{\infty}}\leq\rho_{\textup{in}} according to Lemma 2.

III-C1 1{\mathcal{L}_{1}} adaptive control architecture

For stability guarantees, the filter 𝒞(s){\mathcal{C}}(s) in (29) needs to ensure that there exists a positive constant ρr\rho_{r} and a (small) positive constant γ1\gamma_{1} such that

𝒢xm(s)1bf,𝒳r<ρrxv(s)1\displaystyle\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}}<\rho_{r}-\left\lVert\mathcal{H}_{xv}(s)\right\rVert_{\mathcal{L}_{1}} vρin,\displaystyle\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}-\rho_{\textup{in}}, (31a)
𝒢xm(s)1Lf,𝒳a\displaystyle\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}L_{f,\mathcal{X}_{a}} <1,\displaystyle<1, (31b)

where

ρ\displaystyle\rho ρr+γ1,\displaystyle\triangleq\rho_{r}+\gamma_{1}, (32)
𝒳r\displaystyle\mathcal{X}_{r} Ω(ρr),𝒳aΩ(ρ).\displaystyle\triangleq\Omega(\rho_{r}),\ \mathcal{X}_{a}\triangleq\Omega(\rho). (33)
Remark 7.

We will show in Lemma 3 and Theorem 1 that ρr\rho_{r} and ρ\rho are actually uniform bounds on the states of a non-adaptive reference system (defined in (43)) and of the adaptive system, respectively.

Remark 8.

Note that 𝒢xm(s)10\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}\rightarrow 0, when the bandwidth of the filter 𝒞(s){\mathcal{C}}(s) goes to infinity, i.e., kfjk_{f}^{j}\rightarrow\infty for all j1mj\in\mathbb{Z}_{1}^{m}. Furthermore, bf,Ω(ρr)b_{f,\Omega(\rho_{r})} can be bounded using the Lipschitz property 7a of f(t,x)f(t,x) in Ω(ρr)\Omega(\rho_{r}), and Lf,Ω(ρ)L_{f,\Omega(\rho)} is bounded given any ρ>0\rho>0. Therefore, 31 can always be satisfied under a sufficiently high bandwidth for 𝒞(s){\mathcal{C}}(s).

A typical 1{\mathcal{L}_{1}}AC is comprised of three elements, namely a state predictor, an adaptive law and a low-pass filtered control law. For the system 5, the state predictor is defined by

x^˙(t)=Amx(t)+Bvv(t)+B(ua(t)+σ^1(t))+Bσ^2(t)+Aex~(t),x^(0)=x0,\dot{\hat{x}}(t)=A_{m}x(t)+B_{v}v(t)+B(u_{\textup{a}}(t)+\hat{\sigma}_{1}(t))+B^{\perp}\hat{\sigma}_{2}(t)+A_{e}\tilde{x}(t),\ \hat{x}(0)=x_{0}, (34)

where x~(t)=x^(t)x(t)\tilde{x}(t)=\hat{x}(t)-x(t) is the prediction error, AeA_{e} is a Hurwitz matrix, Bn×(nm)B^{\perp}\in\mathbb{R}^{n\times(n-m)} is an arbitrary matrix satisfying BB=0B^{\perp}B=0 and rank([BB])=n\textup{rank}\left(\left[B\ B^{\perp}\right]\right)=n, and σ^1(t)\hat{\sigma}_{1}(t) and σ^2(t)\hat{\sigma}_{2}(t) are estimated matched and unmatched disturbances, respectively. The estimates σ^1(t)\hat{\sigma}_{1}(t) and σ^2(t)\hat{\sigma}_{2}(t) are updated by the following piecewise-constant adaptive law (similar to that in [18, Section 3.3]):

{[σ^1(t)σ^2(t)]=[σ^1(iT)σ^2(iT)],t[iT,(i+1)T),[σ^1(iT)σ^2(iT)]=[BB]1Φ1(T)eAeTx~(iT),\left\{\begin{aligned} &\begin{bmatrix}\hat{\sigma}_{1}(t)\\ \hat{\sigma}_{2}(t)\end{bmatrix}&&=\begin{bmatrix}\hat{\sigma}_{1}(iT)\\ \hat{\sigma}_{2}(iT)\end{bmatrix},\quad t\in[iT,(i+1)T),\\ &\begin{bmatrix}\hat{\sigma}_{1}(iT)\\ \hat{\sigma}_{2}(iT)\end{bmatrix}&&=-\!\left[B\ B^{\perp}\right]^{-1}\Phi^{-1}(T)e^{A_{e}T}\tilde{x}(iT),\end{aligned}\right. (35)

where TT is the estimation sampling time and Φ(T)Ae1(eAeTIn)\Phi(T)\triangleq A_{e}^{-1}\left(e^{A_{e}T}\!-I_{n}\right). Finally, the control law is given by

ua(s)=𝒞(s)𝔏[σ^1(t)].u_{\textup{a}}(s)=-{\mathcal{C}}(s)\mathfrak{L}\left[\hat{\sigma}_{1}(t)\right]. (36)

The control law 36 tries to cancel the estimated (matched) uncertainty within the bandwidth of the filter 𝒞(s){\mathcal{C}}(s). Additionally, unmatched uncertainty estimate (σ^2(t)\hat{\sigma}_{2}(t)) appears in 34 and 35, although the system dynamics 5 contains only matched uncertainty. This is due to the adoption of the piecewise-constant adaptive law, which may produce nonzero value for σ^2(t)\hat{\sigma}_{2}(t). However, a non-zero σ^2(t)\hat{\sigma}_{2}(t) will not cause an issue either for implementation or for performance guarantee. Additionally, it is possible to prove that limT0σ^2(t)=0\lim_{T\rightarrow 0}\hat{\sigma}_{2}(t)=0 for any t0t\geq 0 [25], i.e., the estimated unmatched uncertainty will be close to zero when TT is small.

III-C2 Uniform performance bounds

We first define some constants:

α¯0(T)\displaystyle\bar{\alpha}_{0}(T) 0TeAe(Tτ)B𝑑τ,\displaystyle\triangleq\int_{0}^{T}\left\lVert e^{A_{e}(T-\tau)}B\right\rVert_{\infty}d\tau, (37a)
α¯1(T)\displaystyle\bar{\alpha}_{1}(T) maxt[0,T]eAet,\displaystyle\triangleq\max_{t\in[0,T]}\left\lVert e^{A_{e}t}\right\rVert_{\infty}, (37b)
α¯2(T)\displaystyle\bar{\alpha}_{2}(T) maxt[0,T]0teAe(tτ)Φ1(T)eAeT𝑑τ,\displaystyle\triangleq\max_{t\in[0,T]}\int_{0}^{t}\left\lVert e^{A_{e}(t-\tau)}\Phi^{-1}(T)e^{A_{e}T}\right\rVert_{\infty}d\tau, (37c)
γ0(T)\displaystyle\gamma_{0}(T) bf,𝒳aα¯0(T)(a¯1(T)+a¯2(T)+1),\displaystyle\triangleq b_{f,\mathcal{X}_{a}}\bar{\alpha}_{0}(T)\left(\bar{a}_{1}(T)+\bar{a}_{2}(T)+1\right), (37d)

where α0(T),α1(T)\alpha_{0}(T),~{}\alpha_{1}(T) and α2(T)\alpha_{2}(T) are defined in 37a, 37b and 37c, respectively. Clearly, bf,𝒳ab_{f,\mathcal{X}_{a}} for a compact set 𝒳a\mathcal{X}_{a} and limT0α¯1(T)=0\lim_{T\rightarrow 0}\bar{\alpha}_{1}(T)=0 are bounded, and limT0α¯0(T)=0\lim_{T\rightarrow 0}\bar{\alpha}_{0}(T)=0. By using Taylor series expansion of eAeTe^{A_{e}T}, one can show that limT00TΦ1(T)𝑑τ\lim_{T\rightarrow 0}\int_{0}^{T}\left\lVert\Phi^{-1}(T)\right\rVert_{\infty}d\tau is bounded, which implies that limT0α¯2(T)\lim_{T\rightarrow 0}\bar{\alpha}_{2}(T) is bounded. As a result, we have

limT0γ0(T)=0.\lim_{T\rightarrow 0}\gamma_{0}(T)=0. (38)

Further define

ρur\displaystyle\rho_{ur}\triangleq 𝒞(s)1bf,𝒳r,\displaystyle\left\lVert{\mathcal{C}}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}}, (39)
γ2\displaystyle\gamma_{2}\triangleq 𝒞(s)1Lf,𝒳aγ1+𝒞(s)B(sInAe)1γ0(T),\displaystyle\left\lVert{\mathcal{C}}(s)\right\rVert_{\mathcal{L}_{1}}\!L_{f,\mathcal{X}_{a}}\gamma_{1}\!+\!\left\lVert{\mathcal{C}}(s)B^{\dagger}(sI_{n}\!-\!A_{e})\right\rVert_{\mathcal{L}_{1}}\!\gamma_{0}(T), (40)
ρua\displaystyle\rho_{u_{\textup{a}}}\triangleq ρur+γ2,\displaystyle\rho_{ur}+\gamma_{2}, (41)

where γ1\gamma_{1} is introduced in 32. Due to 38 and 31b, we can always select a small enough T>0T>0 such that

xm(s)𝒞(s)B(sInAe)11𝒢xm(s)1Lf,𝒳aγ0(T)<γ1,\frac{\left\lVert\mathcal{H}_{xm}(s){\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\right\rVert_{\mathcal{L}_{1}}}{1-\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}L_{f,\mathcal{X}_{a}}}\gamma_{0}(T)<\gamma_{1}, (42)

where 𝒳a\mathcal{X}_{a} is defined in 33 and BB^{\dagger} is the pseudo-inverse of BB.

Following the convention for performance analysis of an 1{\mathcal{L}_{1}}AC[18], we introduce the following reference system:

x˙r(t)\displaystyle\dot{x}_{\textup{r}}(t) =Amxr(t)+Bvv(t)+B(ur(t)+f(t,xr(t))),\displaystyle\!=\!{A_{m}}x_{\textup{r}}(t)\!+\!{B_{v}}v(t)\!+\!B({u_{\textup{r}}}(t)\!+\!f(t,x_{\textup{r}}(t))),\hfill (43a)
ur(s)\displaystyle{u_{\textup{r}}}(s) =𝒞(s)𝔏[f(t,xr(t))],xr(0)=x0,\displaystyle\!=\!-{\mathcal{C}}(s)\mathfrak{L}\left[f(t,x_{\textup{r}}(t))\right],\quad x_{\textup{r}}(0)\!=\!x_{0}, (43b)

Clearly, the control law in the reference system 43 partially cancels the uncertainty f(t,xr(t)))f(t,x_{\textup{r}}(t))) within the bandwidth of the filter 𝒞(s){\mathcal{C}}(s). Moreover, the control law depends on the true uncertainties and is thus not implementable. The reference system is introduced to help characterize the performance of the adaptive closed-loop system, which will be done in four sequential steps: (i) establishing the bounds on the states and inputs of the reference system (Lemma 3); (ii) quantifying the difference between the states and inputs of the adaptive system and those of the reference system (Theorem 1); (iii) quantifying the difference between the states and inputs of the reference system and those of the nominal system (Lemma 5); (iv) based on the results from (ii) and (iii), quantifying the difference between the states and inputs of the adaptive system and those of the nominal system (Theorem 2).

The proofs of these lemmas and theorems mostly follow the typical 1{\mathcal{L}_{1}}AC analysis procedure [18], and are included in appendices for completeness.

For notation brevity, we define:

η(t)f(t,x(t)),ηr(t)f(t,xr(t)).\eta(t)\triangleq f(t,x(t)),\quad\eta_{\textup{r}}(t)\triangleq f(t,x_{\textup{r}}(t)). (44)

To provide an overview, Table I summarizes the different (error) systems involved in this section and their related theorems/lemmas, the uniform bounds, the 1{\mathcal{L}_{1}}AC parameters and conditions.

TABLE I: An overview of different (error) systems involved in Section III-C, and their related theorem/lemma, uniform bounds, 1{\mathcal{L}_{1}}AC parameters and conditions
(Error) System Theorem/Lemma Uniform Bounds on States and Inputs 1{\mathcal{L}_{1}}AC Parameters Conditions
1 Nominal system 9 Lemma 2 xnρin+xv1v\left\lVert x_{\textup{n}}\right\rVert_{\mathcal{L}_{\infty}}\!\leq\!\rho_{\textup{in}}+\left\lVert\mathcal{H}_{xv}\right\rVert_{\mathcal{L}_{1}}\left\lVert v\right\rVert_{\infty} N/A N/A
2 Reference system 43 Lemma 3 xr<ρr\left\lVert x_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}}\!<\!\rho_{r}, ur<ρur\left\lVert u_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}}\!<\!\rho_{ur} 𝒞(s){\mathcal{C}}(s) 31a
3 Diff. b/t reference and adaptive systems Theorem 1 xrxγ1,uruaγ2\left\lVert x_{\textup{r}}\!-\!x\right\rVert_{\mathcal{L}_{\infty}}\!\leq\!\gamma_{1},\ \left\lVert u_{\textup{r}}\!-\!u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}}\!\leq\!\gamma_{2} Ae,T,𝒞(s)A_{e},~{}T,~{}{\mathcal{C}}(s) 42 and 31a
4 Diff. b/t reference and nominal systems Lemma 5 xrxn𝒢xm1bf,𝒳r\left\|x_{\textup{r}}\!-\!x_{\textup{n}}\right\|_{{\mathcal{L}}{{}_{\infty}}}\!\leq\!\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}} 𝒞(s){\mathcal{C}}(s) 31a
5 Adaptive system: 5 and the 1{\mathcal{L}_{1}}AC Theorem 1 x<ρ\left\lVert x\right\rVert_{\mathcal{L}_{\infty}}<\rho, ua<ρu\left\lVert u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}}<\rho_{u} Ae,T,𝒞(s)A_{e},~{}T,~{}{\mathcal{C}}(s) 42 and 31a
6 Diff. b/t adaptive and nominal systems Theorem 2 xxnρ~\left\lVert x\!-\!x_{\textup{n}}\right\rVert_{\mathcal{L}_{\infty}}\!\leq\!\tilde{\rho} Ae,T,𝒞(s)A_{e},~{}T,~{}{\mathcal{C}}(s) 42 and 31a

The proofs for Lemmas 3, 4, 1 and 5 are given in Sections A-B, A-C, A-D and A-E.

Lemma 3.

For the closed-loop reference system in (43) subject to Assumption 1 and the stability condition in (31a), we have

xr\displaystyle\left\lVert x_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}} <ρr,\displaystyle<\rho_{r}, (45)
ur\displaystyle\left\lVert u_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}} <ρur,\displaystyle<\rho_{ur}, (46)

where ρr\rho_{r} is introduced in 31a, and ρur\rho_{ur} is defined in 39.

From 5 and 34, the prediction error dynamics are given by

x~˙(t)\displaystyle\dot{\tilde{x}}(t) =Aex~(t)+B(σ^1(t)f(t,x(t)))+Bσ^2(t).\displaystyle=A_{e}\tilde{x}(t)+B\left(\hat{\sigma}_{1}(t)-f(t,x(t))\right)+B^{\perp}\hat{\sigma}_{2}(t). (47)

The following lemma establishes a bound on the prediction error under the assumption that the actual states and adaptive inputs are bounded.

Lemma 4.

Given the uncertain system (5) subject to Assumption 1, the state predictor (34) and the adaptive law (35), if

x[0,τ]ρ,ua[0,τ]ρua,\left\lVert x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq\rho,\quad\left\lVert u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq\rho_{u_{\textup{a}}}, (48)

with ρ\rho and ρua\rho_{u_{\textup{a}}} defined in 41 and 32, respectively, then

x~[0,τ]γ0(T).\displaystyle\left\lVert\tilde{x}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq\gamma_{0}(T). (49)
Theorem 1.

Given the uncertain system (5) subject to Assumption 1 and the reference system (43) subject to the conditions 31b and 31a with a constant γ1>0\gamma_{1}>0, with the 1{\mathcal{L}_{1}}AC defined via 34, 35 and 36 subject to the sample time constraint (42), we have

x\displaystyle\left\lVert x\right\rVert_{\mathcal{L}_{\infty}} ρ,\displaystyle\leq\rho, (50a)
ua\displaystyle\left\lVert u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}} ρua,\displaystyle\leq\rho_{u_{\textup{a}}}, (50b)
xrx\displaystyle\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}} γ1,\displaystyle\leq\gamma_{1}, (50c)
urua\displaystyle\left\lVert u_{\textup{r}}-u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}} γ2,\displaystyle\leq\gamma_{2}, (50d)

where ρ\rho,γ2\gamma_{2} and ρua\rho_{u_{\textup{a}}} are defined in 32, 40 and 41, respectively.

Remark 9.

For an arbitrarily small γ1>0\gamma_{1}>0, one can always find a small enough TT such that the constraint 42 is satisfied. According to 40, γ2\gamma_{2} depends on γ1\gamma_{1} and γ0(T)\gamma_{0}(T), and can be made arbitrarily small by reducing γ1\gamma_{1} and TT. Thus, by reducing TT, both γ1\gamma_{1} and γ2\gamma_{2} can be made arbitrarily small, which indicates that the difference between the inputs and states of the adaptive system and those of the reference system can be made arbitrarily small from Theorem 1.

Lemma 5.

Given the reference system (43) and the nominal system (9a), subject to Assumption 1, and the condition 31a, we have

xrxn\displaystyle{\left\|{{x_{\textup{r}}}-x_{\textup{n}}}\right\|_{{\mathcal{L}}{{}_{\infty}}}} 𝒢xm1bf,𝒳r\displaystyle\leq\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}} (51)
Remark 10.

When the bandwidth of the filter 𝒞(s){\mathcal{C}}(s) goes to infinity, 𝒢xm1\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}} and thus xrxn{\left\|{{x_{\textup{r}}}-x_{\textup{n}}}\right\|_{{\mathcal{L}}{{}_{\infty}}}} go to 0. This indicates that the difference between the states of the reference system and those of the nominal system can be made arbitrarily small by increasing the filter bandwidth. However, a high-bandwidth filter allows for high-frequency control signals to enter the system under fast adaptation (corresponding to small TT), compromising the robustness. Thus, the filter presents a trade-off between robustness and performance. More details about the role and design of the filter can be found in [18].

From Theorem 1, Lemma 5 and application of the triangle inequality, we can obtain uniform bounds on the error between the actual system 5 and the nominal system 9a, formally stated in the following theorem. The proof is straightforward and thus omitted.

Theorem 2.

Given the uncertain system (5) subject to Assumption 1, the nominal system (9a), and the 1{\mathcal{L}_{1}}AC defined via 34, 35 and 36 subject to the conditions 31b and 31a with a constant γ1>0\gamma_{1}>0 and the sample time constraint (42), we have

xxn\displaystyle\left\lVert x-x_{\textup{n}}\right\rVert_{\mathcal{L}_{\infty}} ρ~,\displaystyle\leq\tilde{\rho}, (52)
ua\displaystyle\left\lVert u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}} ρua,\displaystyle\leq\rho_{u_{\textup{a}}}, (53)

where ρua\rho_{u_{\textup{a}}} is defined in 41, and

ρ~\displaystyle\tilde{\rho} 𝒢xm(s)1bf,𝒳r+γ1.\displaystyle\triangleq\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}}+\gamma_{1}. (54)
Remark 11.

From Remarks 10 and 9, by decreasing TT and increasing the bandwidth of the filter 𝒞(s){\mathcal{C}}(s), one can make (i) the states of the adaptive system arbitrarily close to those of the nominal system; and (ii) the adaptive inputs ua(t)u_{\textup{a}}(t) arbitrarily close to f(t,x)f(t,x), i.e., the true uncertainty, since f(t,xr)f(t,x_{\textup{r}}) is arbitrarily close to f(t,x)f(t,x) when the error between x(t)x(t) and xr(t)x_{\textup{r}}(t) is arbitrarily small.

IV 1{\mathcal{L}_{1}}AC with Separate Bounds for States and Inputs

In Section III-C, we presented an 1{\mathcal{L}_{1}}AC that guarantees uniform bounds on the states and adaptive control inputs of the adaptive system with respect to the nominal system, without consideration of the constraints 2. However, as can be seen from Theorem 2, the uniform bound on x(t)xn(t)x(t)-x_{\textup{n}}(t) or ua(t)u_{\textup{a}}(t) is represented by the vector-\infty norm, which always leads to the same bound for all the states, xixn,i(t)x_{i}-x_{\textup{n},i}(t) (i1ni\in\mathbb{Z}_{1}^{n}), or all the adaptive inputs, ua,ju_{\textup{a},j} (j1mj\in\mathbb{Z}_{1}^{m}). The use of vector-\infty norms may lead to conservative bounds for some specific states or adaptive inputs, making it impossible to satisfy the constraints 2 or leading to significantly tightened constraints for the RG design. To reduce such conservatism, this section will present a scaling technique to derive an individual bound for each xi(t)xn,i(t)x_{i}(t)-x_{\textup{n},i}(t) (i1ni\in\mathbb{Z}_{1}^{n}) and ua,j(t)u_{\textup{a},j}(t) (j1mj\in\mathbb{Z}_{1}^{m}).

From Theorem 2, one can see that the bound on x(t)xn(t)x(t)-x_{\textup{n}}(t) (or ua(t)u_{\textup{a}}(t)) consists of two parts: the first part is γ1\gamma_{1} (or γ2\gamma_{2}) that can be made arbitrarily small by reducing TT (see Remark 9), while the second part is a bound on xr(t)xn(t)x_{\textup{r}}(t)-x_{\textup{n}}(t) (or ur(t)u_{\textup{r}}(t)). Next, we will derive an individual bound for each xi(t)xr,t(t)x_{i}(t)-x_{\textup{r},t}(t) (or ur,j(t)u_{\textup{r},j}(t)).
Derive Separate Bounds for States via Scaling: For deriving an individual bound for each xi(t)xr,t(t)x_{i}(t)-x_{\textup{r},t}(t), we introduce the following coordinate transformations for the reference system 43 and the nominal system 9a for each i1ni\in\mathbb{Z}_{1}^{n}:

{xˇr=Txixr,xˇn=Txixn,Aˇmi=TxiAm(Txi)1,Bˇi=TxiB,Bˇvi=TxiBv,\left\{\begin{aligned} \check{x}_{\textup{r}}&=\mathrm{T}_{x}^{i}x_{\textup{r}},\quad\check{x}_{\textup{n}}=\mathrm{T}_{x}^{i}x_{\textup{n}},\\ \check{A}_{m}^{i}&=\mathrm{T}_{x}^{i}A_{m}(\mathrm{T}_{x}^{i})^{-1},\\ \check{B}^{i}&=\mathrm{T}_{x}^{i}B,\quad\check{B}^{i}_{v}=\mathrm{T}_{x}^{i}B_{v},\end{aligned}\right. (55)

where Txi>0\mathrm{T}_{x}^{i}\!>\!0 is a diagonal matrix that satisfies

Txi[i]\displaystyle\mathrm{T}_{x}^{i}[i] =1, 0<Txi[k]1,ki,\displaystyle=1,\ 0<\mathrm{T}_{x}^{i}[k]\leq 1,\ \forall k\neq i, (56)

with Txi[k]\mathrm{T}_{x}^{i}[k] denoting the kkth diagonal element. Under the transformation 55, the reference system 43 is converted to

{xˇ˙r(t)=Aˇmixˇr(t)+Bˇviv(t)+Bˇi(ur(t)+fˇ(t,xˇr(t))),ur(s)=𝒞(s)𝔏[fˇ(t,xˇr(t))],xˇ(0)=Txix0,\left\{\begin{aligned} \dot{\check{x}}_{\textup{r}}(t)&\!=\!\check{A}_{m}^{i}\check{x}_{\textup{r}}(t)\!+\!\check{B}_{v}^{i}v(t)\!+\!\check{B}^{i}(u_{\textup{r}}(t)\!+\!\!\check{f}(t,\check{x}_{\textup{r}}(t))),\\ {u_{\textup{r}}}(s)&\!=\!-{\mathcal{C}}(s)\mathfrak{L}\left[\check{f}(t,\check{x}_{\textup{r}}(t))\right],\ \check{x}(0)\!=\!\mathrm{T}_{x}^{i}x_{0},\end{aligned}\right. (57)

where

fˇ(t,xˇr(t))=f(t,xr(t)))=f(t,(Txi)1xˇr(t))).\check{f}(t,\check{x}_{\textup{r}}(t))=f(t,x_{\textup{r}}(t)))=f(t,(\mathrm{T}_{x}^{i})^{-1}\check{x}_{\textup{r}}(t))). (58)

Given a set 𝒵{\mathcal{Z}}, define

𝒵ˇ{zˇn:(Txi)1zˇ𝒵}.\check{\mathcal{Z}}\triangleq\{\check{z}\in\mathbb{R}^{n}:(\mathrm{T}_{x}^{i})^{-1}\check{z}\in{\mathcal{Z}}\}. (59)

Similar to 30, for the transformed reference system 57, we have

xˇmi(s)\displaystyle\mathcal{H}_{\check{x}m}^{i}(s) (sInAˇmi)1Bˇi=Txixm(s),\displaystyle\triangleq(sI_{n}\!-\!\check{A}_{m}^{i})^{-1}\!\check{B}^{i}=\mathrm{T}_{x}^{i}\mathcal{H}_{xm}(s), (60a)
xˇvi(s)\displaystyle\mathcal{H}_{\check{x}v}^{i}(s) (sInAˇmi)1Bˇvi=Txixv(s),\displaystyle\triangleq(sI_{n}\!-\!\check{A}_{m}^{i})^{-1}\!\check{B}^{i}_{v}=\mathrm{T}_{x}^{i}\mathcal{H}_{xv}(s), (60b)
𝒢xˇmi(s)\displaystyle\mathcal{G}_{\check{x}m}^{i}(s) xˇmi(s)(Im𝒞(s))=Txi𝒢xm(s),\displaystyle\triangleq\mathcal{H}_{\check{x}m}^{i}(s)(I_{m}-{\mathcal{C}}(s))=\mathrm{T}_{x}^{i}\mathcal{G}_{xm}(s), (60c)

where xm,xv,𝒢xm\mathcal{H}_{xm},~{}\mathcal{H}_{xv},~{}\mathcal{G}_{xm} are defined in 30. By applying the transformation 55 to the nominal system 9a, we obtain

{xˇ˙n(t)=Aˇmixˇn(t)+Bˇviv(t),xˇn(0)=Txix0,yn(t)=Cˇxˇn(t).\left\{\begin{aligned} \dot{\check{x}}_{\textup{n}}(t)&=\check{A}_{m}^{i}{\check{x}}_{\textup{n}}(t)+\check{B}^{i}_{v}v(t),\ {\check{x}}_{\textup{n}}(0)\!=\!\mathrm{T}_{x}^{i}x_{0},\\ y_{\textup{n}}(t)&=\check{C}{\check{x}}_{\textup{n}}(t).\end{aligned}\right. (61)

Letting xˇin(t){\check{x}}_{\textup{in}}(t) be the state of the system xˇ˙in(t)=Aˇmixˇin(t)\dot{\check{x}}_{\textup{in}}(t)=\check{A}_{m}^{i}{\check{x}}_{\textup{in}}(t) with xˇin(0)=xˇn(0)=Txix0{\check{x}}_{\textup{in}}(0)=\check{x}_{\textup{n}}(0)=\mathrm{T}_{x}^{i}x_{0}, we have xˇin(s)(sInAˇmi)1xˇin(0)=Txi(sInAm)1x0{\check{x}}_{\textup{in}}(s)\triangleq(sI_{n}-\check{A}_{m}^{i})^{-1}{\check{x}}_{\textup{in}}(0)=\mathrm{T}_{x}^{i}(sI_{n}-A_{m})^{-1}x_{0}. Defining

ρˇinisTxi(sInAm)11maxx0𝒳0x0,\check{\rho}_{\textup{in}}^{i}\triangleq\left\lVert s\mathrm{T}_{x}^{i}(sI_{n}-A_{m})^{-1}\right\rVert_{\mathcal{L}_{1}}\max_{x_{0}\in\mathcal{X}_{0}}\left\lVert x_{0}\right\rVert_{\infty}, (62)

and further considering Lemma 2, we have xˇinρˇini\left\lVert{\check{x}}_{\textup{in}}\right\rVert_{\mathcal{L}_{\infty}}\leq\check{\rho}_{\textup{in}}^{i}. Similar to 31a, for the transformed reference system 57, consider the following condition:

𝒢xˇmi(s)1bfˇ,𝒳ˇr<ρˇrixˇvi(s)1vρˇini,\displaystyle\left\lVert\mathcal{G}_{\check{x}m}^{i}(s)\right\rVert_{\mathcal{L}_{1}}b_{\check{f},\check{\mathcal{X}}_{r}}<\check{\rho}_{r}^{i}-\left\lVert\mathcal{H}_{\check{x}v}^{i}(s)\right\rVert_{\mathcal{L}_{1}}\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}-\check{\rho}_{\textup{in}}^{i}, (63)

where 𝒳r\mathcal{X}_{r} is defined in 33 and 𝒳ˇr\check{\mathcal{X}}_{r} is defined according to 59 and ρˇri\check{\rho}_{r}^{i} is a positive constant to be determined. Then we have the following result.

Lemma 6.

Consider the reference system (43) subject to Assumption 1, the nominal system 9a, the transformed reference system 57 and transformed nominal system 61 obtained by applying 55 with any Txi\mathrm{T}_{x}^{i} satisfying 56. Suppose that 31a holds with some constants ρr\rho_{r} and v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}. Then, there exists an constant ρˇriρr\check{\rho}_{r}^{i}\leq\rho_{r} such that 63 holds with the same v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}. Furthermore,

|xr,i(t)|\displaystyle\left\lvert x_{\textup{r},i}(t)\right\rvert ρˇri,t0,\displaystyle\leq\check{\rho}_{r}^{i},\ \forall t\geq 0, (64)
|xr,i(t)xn,i(t)|\displaystyle\left\lvert x_{\textup{r,i}}(t)-x_{\textup{n},i}(t)\right\rvert 𝒢xˇm(s)1bf,𝒳r,t0,\displaystyle\leq\left\lVert\mathcal{G}_{\check{x}m}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}},\ \forall t\geq 0, (65)

where we re-define

𝒳r{zn:|zi|ρˇri,i1n}.\mathcal{X}_{r}\triangleq\left\{z\in\mathbb{R}^{n}:\left\lvert z_{i}\right\rvert\leq\check{\rho}_{r}^{i},i\in\mathbb{Z}_{1}^{n}\right\}. (66)
Proof.

For any Txi\mathrm{T}_{x}^{i} satisfying 56 with an arbitrary i1ni\in\mathbb{Z}_{1}^{n}, we have Txi=1\left\lVert\mathrm{T}_{x}^{i}\right\rVert_{\infty}=1. Therefore, under the transformation 55, considering 60 and 62 and Lemma 2, we have

xˇmi(s)\displaystyle\left\lVert\mathcal{H}_{\check{x}m}^{i}(s)\right\rVert_{\mathcal{L}_{\infty}} Txixm(s)=xm(s),\displaystyle\!\leq\!\left\lVert\mathrm{T}_{x}^{i}\right\rVert_{\infty}\left\lVert\mathcal{H}_{xm}(s)\right\rVert_{\mathcal{L}_{\infty}}\!=\!\left\lVert\mathcal{H}_{xm}(s)\right\rVert_{\mathcal{L}_{\infty}}\!, (67a)
xˇvi(s)\displaystyle\left\lVert\mathcal{H}_{\check{x}v}^{i}(s)\right\rVert_{\mathcal{L}_{\infty}} Txixv(s)=xv(s),\displaystyle\!\leq\!\left\lVert\mathrm{T}_{x}^{i}\right\rVert_{\infty}\left\lVert\mathcal{H}_{xv}(s)\right\rVert_{\mathcal{L}_{\infty}}\!=\!\left\lVert\mathcal{H}_{xv}(s)\right\rVert_{\mathcal{L}_{\infty}}\!, (67b)
𝒢xˇmi(s)\displaystyle\left\lVert\mathcal{G}_{\check{x}m}^{i}(s)\right\rVert_{\mathcal{L}_{\infty}} Txi𝒢xm(s)=𝒢xm(s),\displaystyle\!\leq\!\left\lVert\mathrm{T}_{x}^{i}\right\rVert_{\infty}\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{\infty}}\!=\!\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{\infty}}\!, (67c)
ρˇini\displaystyle\check{\rho}_{\textup{in}}^{i} Txiρin=ρin.\displaystyle\!\leq\!\left\lVert\mathrm{T}_{x}^{i}\right\rVert_{\infty}\rho_{\textup{in}}\!=\!\rho_{\textup{in}}. (67d)

It follows from Lemma 3 that xr(t)𝒳rx_{\textup{r}}(t)\in\mathcal{X}_{r} for any t0t\geq 0, which, together with 55, implies xˇr(t)𝒳ˇr\check{x}_{\textup{r}}(t)\in\check{\mathcal{X}}_{r} for any t0t\geq 0, where 𝒳ˇr\check{\mathcal{X}}_{r} is defined via 59. Considering 58 and 59, for any compact set 𝒳r\mathcal{X}_{r}, we have

bfˇ,𝒳ˇr=bf,𝒳r.b_{\check{f},\check{\mathcal{X}}_{r}}=b_{f,\mathcal{X}_{r}}. (68)

Now suppose that constants ρr\rho_{r} and v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}} satisfy 31a. Then, due to 68 and 67, with ρˇri=ρr\check{\rho}_{r}^{i}=\rho_{r} and the same v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}, 63 is satisfied.

Additionally, if 63 holds, by applying Lemma 3 to the transformed reference system 57, we obtain that xˇrρˇri\left\lVert\check{x}_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}}\leq\check{\rho}_{r}^{i}, implying that |xˇr,i(t)|ρˇri\left\lvert\check{x}_{\textup{r},i}(t)\right\rvert\leq\check{\rho}_{r}^{i} for any t0t\geq 0. Since xˇr,i(t)=xr,i(t)\check{x}_{\textup{r},i}(t)=x_{\textup{r},i}(t) due to the constraint 56 on Txi\mathrm{T}_{x}^{i}, we have 64. Equation 64 is equivalent to xr(t)𝒳rx_{\textup{r}}(t)\in\mathcal{X}_{r} for any t0t\geq 0, with the re-definition of 𝒳r\mathcal{X}_{r} in 66. Following the proof of Lemma 5, one can obtain xˇrxˇn𝒢xˇm1bfˇ,𝒳ˇr=𝒢xˇm(s)1bf,𝒳ˇr\left\|{\check{x}_{\textup{r}}-\check{x}_{\textup{n}}}\right\|_{{\mathcal{L}}{{}_{\infty}}}\leq\left\lVert\mathcal{G}_{\check{x}m}\right\rVert_{\mathcal{L}_{1}}b_{\check{f},\check{\mathcal{X}}_{r}}=\left\lVert\mathcal{G}_{\check{x}m}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\check{\mathcal{X}}_{r}}, where the equality is due to 68. Further considering xˇr,i(t)=xr,i(t)\check{x}_{\textup{r},i}(t)=x_{\textup{r},i}(t) and xˇn,i(t)=xn,i(t)\check{x}_{\textup{n},i}(t)=x_{\textup{n},i}(t) due to the constraint 56 on Txi\mathrm{T}_{x}^{i}, we have 65. ∎

Remark 12.

Lemma 3 and Lemma 5 imply |xr,i(t)|ρr\left\lvert x_{\textup{r},i}(t)\right\rvert\leq\rho_{r} and |xr,i(t)xn,i(t)|𝒢xm1bf,Ω(ρr)\left\lvert x_{\textup{r},i}(t)-x_{\textup{n},i}(t)\right\rvert\leq\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}}b_{f,\Omega(\rho_{r})}, respectively, for all i1ni\in\mathbb{Z}_{1}^{n} and t0t\geq 0. Lemma 6 indicates that by applying the coordinate transformation 55 and leveraging the condition 63 for the transformed system 57, one can obtain a tighter bound on xr,i(t)x_{\textup{r},i}(t) than ρr\rho_{r} and a tighter bound on |xr,i(t)xn,i(t)|\left\lvert x_{\textup{r},i}(t)-x_{\textup{n},i}(t)\right\rvert than 𝒢xm(s)1bf,Ω(ρr)\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\Omega(\rho_{r})}.

Derive Separate Bounds for Adaptive Inputs: From 43b and the structure with 𝒞(s){\mathcal{C}}(s) 29, we can obtain

ur,j(s)=𝒞j(s)𝔏[fj(t,xr(t))],j1m.u_{\textup{r},j}(s)=-{\mathcal{C}}_{j}(s)\mathfrak{L}\left[f_{j}(t,x_{\textup{r}}(t))\right],\quad\forall j\in\mathbb{Z}_{1}^{m}. (69)

Therefore, given a set 𝒳r\mathcal{X}_{r} such that xr(t)𝒳rx_{\textup{r}}(t)\!\in\!\mathcal{X}_{r} for any t0t\!\geq\!0, from Assumptions 1 and 2 we get

|ur,j(t)|𝒞j(s)1bfj,𝒳r,t0,j1m.\left\lvert u_{\textup{r},j}(t)\right\rvert\leq\left\lVert{\mathcal{C}}_{j}(s)\right\rVert_{\mathcal{L}_{1}}b_{f_{j},\mathcal{X}_{r}},\quad\forall t\geq 0,\ \forall j\in\mathbb{Z}_{1}^{m}. (70)

With the preceding preparations, we are ready to derive an individual bound for xi(t)xn,i(t)x_{i}(t)-x_{\textup{n},i}(t) (i1ni\in\mathbb{Z}_{1}^{n}) and uj(t)un,j(t)u_{j}(t)-u_{\textup{n},j}(t) (j1mj\in\mathbb{Z}_{1}^{m}), as stated in the following theorem.

Theorem 3.

Consider the uncertain system (5) subject to Assumption 1, the nominal system (9a), and the 1{\mathcal{L}_{1}}AC defined via 34, 35 and 36 subject to the conditions 31b and 31a with constants ρr\rho_{r} and γ1>0\gamma_{1}>0 and the sample time constraint (42). Suppose that for each i1ni\in\mathbb{Z}_{1}^{n}, 63 holds with a constant ρˇri\check{\rho}_{r}^{i} for the transformed reference system 57 obtained by applying 55. Then, we have

x(t)xn(t)𝒳~{zn:|zi|ρ~i,i1n},\displaystyle{x(t)-x_{\textup{n}}(t)}\!\in\!\tilde{\mathcal{X}}\!\triangleq\!\left\{z\!\in\!\mathbb{R}^{n}\!:\!\left\lvert z_{i}\right\rvert\!\leq\!\tilde{\rho}^{i},\ i\in\mathbb{Z}_{1}^{n}\right\}\!,\ t0,\displaystyle\forall t\!\geq\!0, (71a)
ua(t)𝒰a{zm:|zj|ρuaj,j1m},\displaystyle{u_{\textup{a}}(t)}\!\in\!{\mathcal{U}}_{\textup{a}}\!\triangleq\!\left\{z\!\in\!\mathbb{R}^{m}\!:\!\left\lvert z_{j}\right\rvert\!\leq\!\rho_{u_{\textup{a}}}^{j},\ \!j\!\in\!\mathbb{Z}_{1}^{m}\right\}\!,\ t0,\displaystyle\forall t\!\geq\!0, (71b)
u(t)un(t)𝒰~{zm:|zj|ρ~uj,j1n},\displaystyle u(t)-u_{\textup{n}}(t)\!\in\!\tilde{\mathcal{U}}\!\triangleq\!\left\{z\!\in\!\mathbb{R}^{m}\!:\!\left\lvert z_{j}\right\rvert\!\leq\!\tilde{\rho}_{u}^{j},\ j\!\in\!\mathbb{Z}_{1}^{n}\right\}\!,\ t0,\displaystyle\forall t\!\geq\!0, (71c)

where

ρi\displaystyle\rho^{i} ρˇri+γ1,ρ~i𝒢xˇmi(s)1bf,𝒳r+γ1,\displaystyle\!\triangleq\!\check{\rho}_{r}^{i}\!+\!\gamma_{1},\ \tilde{\rho}^{i}\!\triangleq\!\left\lVert\mathcal{G}_{\check{x}m}^{i}(s)\right\rVert_{\mathcal{L}_{1}}\!\!b_{f,\mathcal{X}_{r}}\!+\!\gamma_{1}, (72a)
ρuaj\displaystyle\rho_{u_{\textup{a}}}^{j} 𝒞j(s)1bfj,𝒳r+γ2,ρ~ujρuaj+i=1n|Kx[j,i]|ρ~i,\displaystyle\!\triangleq\!\left\lVert{\mathcal{C}}_{j}(s)\right\rVert_{\mathcal{L}_{1}}b_{f_{j},\mathcal{X}_{r}}\!+\!\gamma_{2},\ \tilde{\rho}_{u}^{j}\!\triangleq\!\rho_{u_{\textup{a}}}^{j}\!+\!\sum_{i=1}^{n}\left\lvert K_{x}[j,i]\right\rvert\tilde{\rho}^{i}, (72b)

with 𝒳r\mathcal{X}_{r} defined in 66, and C[j,i]C[j,i] denoting the (j,i)(j,i) element of CC.

Proof.

For each i1ni\in\mathbb{Z}_{1}^{n}, Lemma 6 implies |xr,i(t)|ρˇri\left\lvert x_{\textup{r},i}(t)\right\rvert\leq\check{\rho}_{r}^{i} and |xr,i(t)xn,i(t)|𝒢xˇmi(s)1bf,𝒳r\left\lvert x_{\textup{r,i}}(t)-x_{\textup{n},i}(t)\right\rvert\leq\left\lVert\mathcal{G}_{\check{x}m}^{i}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}} for all t0t\geq 0. On the other hand, Theorem 1 indicates that |xr,i(t)xi(t)|γ1\left\lvert x_{\textup{r},i}(t)-x_{i}(t)\right\rvert\leq\gamma_{1} for any t0t\geq 0 and any i1ni\in\mathbb{Z}_{1}^{n}. Therefore, 71a is true. On the other hand, Theorem 1 indicates that |ur,j(t)ua,j(t)|γ2\left\lvert u_{\textup{r},j}(t)-u_{\textup{a},j}(t)\right\rvert\leq\gamma_{2} for any t0t\geq 0 and any j1mj\in\mathbb{Z}_{1}^{m}, which, together with 70, leads to 71b. Finally, 71c follows from 11, 71b and 71a. The proof is complete. ∎

Remark 13.

Theorem 3 provides a way to derive an individual bound on xi(t)x_{i}(t), and xi(t)xn,i(t)x_{i}(t)-x_{\textup{n},i}(t) for each i1ni\in\mathbb{Z}_{1}^{n} and on uj(t)un,j(t)u_{j}(t)-u_{\textup{n},j}(t) for each j1mj\in\mathbb{Z}_{1}^{m} via coordinate transformations. Additionally, similar to the arguments in Remark 11, by decreasing TT and increasing the bandwidth of the filter 𝒞(s){\mathcal{C}}(s), one can make ρ~i\tilde{\rho}^{i} (i1ni\in\mathbb{Z}_{1}^{n}) arbitrarily small, i.e., making the states of the adaptive system arbitrarily close to those of the nominal system, and make the bounds on ua,j(t)u_{\textup{a},j}(t) and uj(t)un,j(t)u_{j}(t)-u_{\textup{n},j}(t) arbitrarily close to the bound on the true uncertainty fj(t,x)f_{j}(t,x) for x𝒳ax\in\mathcal{X}_{a}, for each j1mj\in\mathbb{Z}_{1}^{m}.

According to Theorems 2 and 3, the procedure for designing an 1{\mathcal{L}_{1}}AC with separate bounds on states and adaptive inputs can be summarized in Algorithm 1.

Algorithm 1 Designing an 1{\mathcal{L}_{1}}AC with separate bounds
1:uncertain system 5 subject to Assumption 1, initial parameters AeA_{e}, 𝒞(s){\mathcal{C}}(s) and TT to define an 1{\mathcal{L}_{1}}AC, γ1\gamma_{1}, 𝒳0\mathcal{X}_{0}, v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}, tol
2:procedure DecideFilterUncertBnd(𝒞(s){\mathcal{C}}(s),γ1\gamma_{1},𝒳0\mathcal{X}_{0},v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}})
3:     while condition 31a or 31b does not hold do
4:         Increase the bandwidth of 𝒞(s){\mathcal{C}}(s) \triangleright See Remark 8.
5:     end while\triangleright ρr\rho_{r}, 𝒳r=Ω(ρr)\mathcal{X}_{r}\!=\!\Omega(\rho_{r}) and bf,𝒳rb_{f,\mathcal{X}_{r}} will be computed.
6:end procedure
7:Set bf,𝒳rold=bf,𝒳rb_{f,\mathcal{X}_{r}}^{old}=b_{f,\mathcal{X}_{r}}
8:procedure DeriveSepStateBnds(bf,𝒳rb_{f,\mathcal{X}_{r}},γ1\gamma_{1},𝒞(s){\mathcal{C}}(s),𝒳0\mathcal{X}_{0},v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}})
9:     for i=1,,ni=1,\dots,n do
10:         Select Txi\mathrm{T}_{x}^{i} satisfying​ 56 and apply the transformation​ 55
11:         Evaluate 60 and compute ρˇini\check{\rho}_{\textup{in}}^{i} according to 62
12:         Compute ρˇri\check{\rho}_{r}^{i} that satisfies 63 \triangleright Such a ρˇriρr\check{\rho}_{r}^{i}\leq\rho_{r} is guaranteed to exist from Lemma 6
13:         Set ρi=ρˇri+γ1\rho^{i}=\check{\rho}_{r}^{i}+\gamma_{1}, ρ~i=𝒢xˇmi(s)1bf,𝒳r+γ1\tilde{\rho}^{i}=\left\lVert\mathcal{G}^{i}_{\check{x}m}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}}+\gamma_{1}
14:     end for
15:end procedure
16:Set 𝒳r={zn:|zi|ρˇri}\mathcal{X}_{r}\!=\!\left\{\!z\!\in\!\mathbb{R}^{n}\!:\!\left\lvert z_{i}\right\rvert\!\leq\!\check{\rho}_{r}^{i}\right\} and update bf,𝒳rb_{f,\mathcal{X}_{r}}
17:if bf,𝒳roldbf,𝒳r>tolb_{f,\mathcal{X}_{r}}^{old}-b_{f,\mathcal{X}_{r}}>\textup{tol} then
18:     Set bf,𝒳rold=bf,𝒳rb_{f,\mathcal{X}_{r}}^{old}=b_{f,\mathcal{X}_{r}} and go to step 8
19:end if
20:Set ρ=maxi1nρi\rho=\max_{i\in\mathbb{Z}_{1}^{n}}\rho^{i} and compute 𝒳a\mathcal{X}_{a} via 33
21:procedure DeriveSepInputBnds(𝒳r,{ρ~i}i1n,γ2,𝒞(s)\mathcal{X}_{r},\{\tilde{\rho}^{i}\}_{i\in\mathbb{Z}_{1}^{n}},\gamma_{2},{\mathcal{C}}(s))
22:     for j=1,,mj=1,\dots,m do
23:         Compute ρuaj\rho_{u_{\textup{a}}}^{j} and ρ~uj\tilde{\rho}_{u}^{j} according to 72b
24:     end for
25:end procedure
26:procedure DecideSampleTime(AeA_{e},𝒞(s){\mathcal{C}}(s),TT,𝒳a\mathcal{X}_{a})
27:     while constraint 42 does not hold do
28:         Decrease TT \triangleright Small TT can enforce 42 due to 38.
29:     end while
30:end procedure
31:An 1{\mathcal{L}_{1}}AC defined by 36, 35 and 34 with parameters AeA_{e} and 𝒞(s){\mathcal{C}}(s) and TT, ρi\rho^{i} and ρ~i\tilde{\rho}^{i} for i1ni\in\mathbb{Z}_{1}^{n}, ρuaj\rho_{u_{\textup{a}}}^{j} and ρ~uj\tilde{\rho}_{u}^{j} for j1mj\in\mathbb{Z}_{1}^{m}
Remark 14.

One can try different Txi\mathrm{T}_{x}^{i} in step 10 of Algorithm 1 and select the one that yields the tightest bound for the iith state.

Remark 15.

The conditions 42 and 31 can be quite conservative for some systems, due to the frequent use of inequalities related to the 1{\mathcal{L}_{1}} norm (stated in Lemma 2), Lipschitz continuity and matrix/vector norms. As a result, the bandwidth of the filter 𝒞(s){\mathcal{C}}(s) computed via 31 could be unnecessarily high, while the sample time TT computed via 42 under a given γ1\gamma_{1} could be unnecessarily small. Based on our experience, assuming that some bounds ρ~i\tilde{\rho}^{i} (i1ni\in\mathbb{Z}_{1}^{n}) and ρ~uj\tilde{\rho}_{u}^{j} (j1mj\in\mathbb{Z}_{1}^{m}) satisfying 71 are derived under a specific filter 𝒞(s){\mathcal{C}}^{\star}(s) and TT^{\star} that satisfy 42 and 31, those bounds will most likely be respected in simulations even if we decrease the bandwidth of 𝒞(s){\mathcal{C}}^{\star}(s) by 131\sim 3 times and/or increase TT^{\star} by 1101\sim 10 times.

V 1{\mathcal{L}_{1}}-RG: Adaptive Reference Governor for Constrained Control Under Uncertainties

Leveraging the uniform bounds on state and input errors guaranteed by the 1{\mathcal{L}_{1}}AC, we now integrate the 1{\mathcal{L}_{1}}AC and the RG introduced in Section III-B to synthesize the 1{\mathcal{L}_{1}}-RG framework for simultaneously enforcing the constraints 2 and improving the tracking performance.

V-A 1{\mathcal{L}_{1}}-RG Design

We first make the following assumption.

Assumption 2.

𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} defined by 12, 17 and 71 are nonempty. Furthermore, there exists a known command v(0)v(0) such that

(v(0),x0)O~,(v(0),x_{0})\in{\tilde{O}}_{\infty}, (73)

where O~{\tilde{O}}_{\infty} is defined in 26.

Remark 16.

Considering 26, 73 implies x0𝒳^nx_{0}\in\hat{\mathcal{X}}_{\textup{n}} and un(0)=Kxx0+Kvv(0)𝒰^nu_{\textup{n}}(0)=K_{x}x_{0}+K_{v}v(0)\in\hat{\mathcal{U}}_{\textup{n}} (since xn(0)=x0x_{\textup{n}}(0)=x_{0}) where 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}}, according to 17, are tightened versions of 𝒳n\mathcal{X}_{\textup{n}} and 𝒰n{\mathcal{U}}_{\textup{n}} that are defined in 12. From Remark 13, with a sufficiently high bandwidth for 𝒞(s){\mathcal{C}}(s) and sufficiently small TT, one can make 𝒳n\mathcal{X}_{\textup{n}} arbitrarily close to 𝒳\mathcal{X}, and make 𝒰~\tilde{\mathcal{U}} arbitrarily close to the bound set for the true uncertainty in 𝒳\mathcal{X}. Additionally, as mentioned in Remark 4, 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} are close to 𝒳n\mathcal{X}_{\textup{n}} and 𝒰n{\mathcal{U}}_{\textup{n}}, respectively, when TdT_{d} is small. As a result, with a sufficiently high bandwidth for 𝒞(s){\mathcal{C}}(s), and sufficiently small TT and TdT_{d}, Assumption 2 roughly states that the initial state stays in 𝒳\mathcal{X}, and the constraint set 𝒰{\mathcal{U}} is sufficiently large to ensure enough control authority for tracking an initial reference command v(0)v(0) and additionally for compensating the uncertainty in 𝒳\mathcal{X}.

Under the preceding assumption, the design procedure for 1{\mathcal{L}_{1}}-RG is summarized in Algorithm 2. Compared to step 3 of Algorithm 1, we additionally constrain xr(t)x_{\textup{r}}(t) and x(t)x(t) to stay in 𝒳\mathcal{X} for all t0t\geq 0 in step 4 of Algorithm 2. Such constraints can potentially limit the size of uncertainties that need to be compensated and significantly reduce the conservatism of the proposed solution.

Algorithm 2 1{\mathcal{L}_{1}}-RG Design
1:An continuous-time uncertain system 5 subject to Assumption 1, constraint sets 𝒳\mathcal{X} and 𝒰{\mathcal{U}} as in 2, 𝒳0\mathcal{X}_{0}, 𝒱\mathcal{V} (admissible set for v(t)v(t)), baseline control law in 3, initial parameters AeA_{e}, 𝒞(s){\mathcal{C}}(s) and TT to define an 1{\mathcal{L}_{1}}AC, γ1\gamma_{1}, TdT_{d} and ϵ\epsilon for RG design, tol
2:procedure 1{\mathcal{L}_{1}}AC-DesignUnderConstraints
3:     Compute v\left\lVert v\right\rVert_{\mathcal{L}_{\infty}} given 𝒱\mathcal{V}
4:     while 31a with 𝒳r=Ω(ρr)𝒳\mathcal{X}_{r}=\Omega(\rho_{r})\cap\mathcal{X} or 31b with 𝒳a=Ω(ρr+γ1)𝒳\mathcal{X}_{a}=\Omega(\rho_{r}+\gamma_{1})\cap\mathcal{X} does not hold with any ρr\rho_{r} do
5:         Increase the bandwidth of 𝒞(s){\mathcal{C}}(s) \triangleright See Remark 8.
6:     end while\triangleright 𝒳r\mathcal{X}_{r} and bf,𝒳rb_{f,\mathcal{X}_{r}} will be computed.
7:     Set bf,𝒳rold=bf,𝒳rb_{f,\mathcal{X}_{r}}^{old}=b_{f,\mathcal{X}_{r}}
8:     Run DeriveSepStateBnds of Algorithm 1 with bf,𝒳rb_{f,\mathcal{X}_{r}}, and obtain ρˇri\check{\rho}_{r}^{i} ρi\rho^{i} and ρ~i\tilde{\rho}^{i} for i1ni\in\mathbb{Z}_{1}^{n}
9:     Set 𝒳r={zn:|zi|ρˇri}𝒳\mathcal{X}_{r}\!=\!\left\{\!z\!\in\!\mathbb{R}^{n}\!:\!\left\lvert z_{i}\right\rvert\!\leq\!\check{\rho}_{r}^{i}\right\}\cap\mathcal{X} and update bf,𝒳rb_{f,\mathcal{X}_{r}}
10:     if bf,𝒳roldbf,𝒳r>tolb_{f,\mathcal{X}_{r}}^{old}-b_{f,\mathcal{X}_{r}}>\textup{tol} then
11:         Set bf,𝒳rold=bf,𝒳rb_{f,\mathcal{X}_{r}}^{old}=b_{f,\mathcal{X}_{r}} and go to step 8
12:     end if
13:     Set 𝒳a={zn:|zi|ρi,i1n}𝒳\mathcal{X}_{a}=\{z\in\mathbb{R}^{n}:\left\lvert z_{i}\right\rvert\leq\rho^{i},\ i\in\mathbb{Z}_{1}^{n}\}\cap\mathcal{X}
14:     Run DeriveSepInputBnds of Algorithm 1 with 𝒞(s){\mathcal{C}}(s) from step 6 and 𝒳r\mathcal{X}_{r} from step 9, and obtain ρuaj\rho_{u_{\textup{a}}}^{j} and ρ~uj\tilde{\rho}_{u}^{j} for j1mj\in\mathbb{Z}_{1}^{m}
15:     Run DecideSampleTime of Algorithm 1 with 𝒞(s){\mathcal{C}}(s) from step 6 and 𝒳a\mathcal{X}_{a} from step 13, and obtain TT
16:     Compute 𝒳~\tilde{\mathcal{X}} and 𝒰~\tilde{\mathcal{U}} with {ρ~i}i1n\{\tilde{\rho}^{i}\}_{i\in\mathbb{Z}_{1}^{n}} and {ρ~uj}j1m\{\tilde{\rho}_{u}^{j}\}_{j\in\mathbb{Z}_{1}^{m}} via 71
17:end procedure
18:procedure RG-Design
19:     Compute 𝒳n\mathcal{X}_{\textup{n}} and 𝒰n{\mathcal{U}}_{\textup{n}} with 𝒳\mathcal{X}𝒰{\mathcal{U}}𝒳~\tilde{\mathcal{X}} and 𝒰~\tilde{\mathcal{U}} via 12
20:     Formulate the nominal discrete-time model 14 with the sample time TdT_{d}
21:     Compute 𝒳^n\hat{\mathcal{X}}_{\textup{n}} and 𝒰^n\hat{\mathcal{U}}_{\textup{n}} via 17 \triangleright With a small TdT_{d}, one may set 𝒳^n=𝒳n\hat{\mathcal{X}}_{\textup{n}}\!=\!\mathcal{X}_{\textup{n}}, 𝒰^n=𝒰n\hat{\mathcal{U}}_{\textup{n}}\!=\!{\mathcal{U}}_{\textup{n}} for practical implementation.
22:     Compute the set O~{\tilde{O}}_{\infty} dependent on ϵ\epsilon according to 26
23:end procedure
24:An 1{\mathcal{L}_{1}}-RG consisting of a RG designed for the nominal system 9a and an 1{\mathcal{L}_{1}}AC to compensate for uncertainties

We are ready to state the guarantees regarding tracking performance and constraint enforcement provided by 1{\mathcal{L}_{1}}-RG.

Theorem 4.

Consider an uncertain system (5) subject to Assumption 1 and the state and control constraints in 2. Suppose that an 1{\mathcal{L}_{1}}AC (defined by 36, 35 and 34) and a RG are designed by following Algorithm 2. If Assumption 2 hold, then, under the baseline control law 3 and the 1{\mathcal{L}_{1}}-RG consisting of the compositional control law 4, the 1{\mathcal{L}_{1}}AC and the RG for computing the reference command v(t)v(t) according to 15 and 27, we have

x(t)int(𝒳),u(t)int(𝒰),\displaystyle x(t)\in\textup{int}(\mathcal{X}),\ u(t)\in\textup{int}({\mathcal{U}}),\quad t0,\displaystyle\forall t\geq 0, (74)
x(t)xn(t)int(𝒳~),\displaystyle x(t)-x_{\textup{n}}(t)\in\textup{int}(\tilde{\mathcal{X}}),\quad t0,\displaystyle\forall t\geq 0, (75)
y(t)yn(t){zm:|zi|ρ~yj},\displaystyle y(t)-y_{\textup{n}}(t)\in\{z\in\mathbb{R}^{m}:\left\lvert z_{i}\right\rvert\leq\tilde{\rho}_{y}^{j}\},\quad t0,\displaystyle\forall t\geq 0, (76)

where xn(t)x_{\textup{n}}(t) and yn(t)y_{\textup{n}}(t) are the states and outputs of the nominal system 9 under the reference command input v(t)v(t), and

ρ~yji=1n|C[j,i]|ρ~i,j1m.\tilde{\rho}_{y}^{j}\!\triangleq\!\sum_{i=1}^{n}\left\lvert C[j,i]\right\rvert\tilde{\rho}^{i},\quad\forall j\in\mathbb{Z}_{1}^{m}. (77)
Proof.

Equation 73 in Assumption 2 implies (v(0),xn(0))O~(\mathrm{v}(0),\mathrm{x}_{\textup{n}}(0))\in{\tilde{O}}_{\infty} (due to xn(0)=x0\mathrm{x_{\textup{n}}}(0)=x_{0}), and un(0)=Kxxn(0)+Kvv(0)𝒰^n\mathrm{u}_{\textup{n}}(0)=K_{x}\mathrm{x_{\textup{n}}}(0)+K_{v}\mathrm{v}(0)\in\hat{\mathcal{U}}_{\textup{n}}. Thus, the reference command v(k)\mathrm{v}(k) produced by 27 ensures xn(k)𝒳^n\mathrm{x}_{\textup{n}}(k)\in\hat{\mathcal{X}}_{\textup{n}} and un(k)𝒰^n\mathrm{u}_{\textup{n}}(k)\in\hat{\mathcal{U}}_{\textup{n}} for all k+k\in\mathbb{Z}_{+}, which, due to Lemma 1, implies

xn(t)𝒳n,un(t)𝒰n,t0.x_{\textup{n}}(t)\in\mathcal{X}_{\textup{n}},\ u_{\textup{n}}(t)\in{\mathcal{U}}_{\textup{n}},\quad\forall t\geq 0. (78)

Compared to 3, 16 and 20 of Algorithm 1, we restrain 𝒳r\mathcal{X}_{r} and 𝒳a\mathcal{X}_{a} to be subsets of 𝒳\mathcal{X} in 4, 9 and 13 of Algorithm 2. As a result, if 74 and

xr(t)𝒳,t0,x_{\textup{r}}(t)\in\mathcal{X},\quad\forall t\geq 0, (79)

jointly hold, condition 75 holds according to Theorem 3, while 65 holds according to Lemma 6.

We next prove 74 and 79 by contradiction. Assume 74 or 79 do not hold. The initial condition 73 implies that x0𝒳nint(𝒳)x_{0}\in\mathcal{X}_{\textup{n}}\subset\textup{int}(\mathcal{X}) and un(0)𝒰nint(𝒰)u_{\textup{n}}(0)\in{\mathcal{U}}_{\textup{n}}\subset\textup{int}({\mathcal{U}}). As a result, we have x(0)𝒳x(0)\in\mathcal{X}, xr(0)𝒳x_{\textup{r}}(0)\in\mathcal{X} and u(0)𝒰u(0)\in{\mathcal{U}}. Since x(t)x(t), xr(t)x_{\textup{r}}(t) and u(t)u(t) are continuous, there must exist a time instant τ\tau, such that

x(t)\displaystyle x(t) int(𝒳),xr(t)int(𝒳) and u(t)int(𝒰),t[0,τ)\displaystyle\!\in\!\textup{int}(\mathcal{X}),\ x_{\textup{r}}(t)\!\in\!\textup{int}(\mathcal{X})\textup{ and }u(t)\!\in\!\textup{int}({\mathcal{U}}),\ \forall t\!\in\![0,\tau) (80a)
x(τ)\displaystyle x(\tau) bnd(𝒳) or xr(τ)bnd(𝒳) or u(τ)bnd(𝒰).\displaystyle\!\in\!\textup{bnd}(\mathcal{X})\textup{ or }x_{\textup{r}}(\tau)\!\in\!\textup{bnd}(\mathcal{X})\textup{ or }u(\tau)\!\in\!\textup{bnd}({\mathcal{U}}). (80b)

Now consider the interval [0,τ][0,\tau]. According to Lemma 6, due to 80 and the definitions in 71a and 72a, we have xr(t)xn(t){zn:|zi|𝒢xˇm(s)1bf,𝒳r}int(𝒳~)x_{\textup{r}}(t)-x_{\textup{n}}(t)\in\{z\in\mathbb{R}^{n}:\left\lvert z_{i}\right\rvert\leq\left\lVert\mathcal{G}_{\check{x}m}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\mathcal{X}_{r}}\}\subset\textup{int}(\tilde{\mathcal{X}}), which, together with 12 and 78, implies

xr(t)int(𝒳),t[0,τ].x_{\textup{r}}(t)\in\textup{int}(\mathcal{X}),\quad\forall t\in[0,\tau]. (81)

Similarly, according to Theorem 3, due to 80 and the definition in 71, we have x(t)xn(t)int(𝒳~)x(t)-x_{\textup{n}}(t)\in\textup{int}(\tilde{\mathcal{X}}), and u(t)un(t)int(𝒰~)u(t)-u_{\textup{n}}(t)\in\textup{int}(\tilde{\mathcal{U}}), for any t[0,τ]t\in[0,\tau], which, together with 12 and 78, implies

x(t)int(𝒳),u(t)int(𝒰),t[0,τ].x(t)\in\textup{int}(\mathcal{X}),\ u(t)\in\textup{int}({\mathcal{U}}),\quad\forall t\in[0,\tau]. (82)

Both 81 and 82 contradict 80b, which proves 74 and 79. By applying the inference right before 82 again for t0t\geq 0, we obtain 75, which, together with yj(t)yn,j(t)=i=1nC[j,i](xi(t)xn,i(t))y_{j}(t)-y_{\textup{n},j}(t)=\sum_{i=1}^{n}C[j,i]\left(x_{i}(t)-x_{\textup{n},i}(t)\right), leads to 76.∎

VI Simulation Results

We now apply 1{\mathcal{L}_{1}}-RG to the longitudinal dynamics of an F-16 aircraft. The model was adapted from [26] with slight modifications to remove the actuator dynamics, in which the state vector x(t)=[γ(t),q(t),α(t)]x(t)=[\gamma(t),q(t),\alpha(t)]^{\top} consists of the flight path angle, pitch rate and angle of attack, and the control input vector u(t)=[δe(t),δf(t)]u(t)=[\delta_{e}(t),\delta_{f}(t)] includes the elevator deflection and flaperon deflection. The output vector is y(t)=[θ(t),γ(t)]y(t)=[\theta(t),\gamma(t)]^{\top}, where θ(t)=γ(t)+α(t)\theta(t)=\gamma(t)+\alpha(t) is the pitch angle; the reference input vector is r(t)=[θc(t),γc(t)]r(t)=[\theta_{c}(t),\gamma_{c}(t)]^{\top}, where θc\theta_{c} and γc\gamma_{c} are the commanded pitch angle and flight path angle, respectively. The system is subject to state and control constraints:

|α(t)|4 deg,|δe(t)|25 deg,|δf(t)|22 deg,\left\lvert\alpha(t)\right\rvert\leq 4\!\textup{ deg},\ \left\lvert\delta_{e}(t)\right\rvert\leq 25\!\textup{ deg},\ \left\lvert\delta_{f}(t)\right\rvert\leq 22\!\textup{ deg}, (83)

where the state constraint can also be represented as x(t)𝒳[103,103]×[103,103]×[4,4]x(t)\in\mathcal{X}\triangleq[-10^{3},10^{3}]\times[-10^{3},10^{3}]\times[-4,4] following the convention in 2. Furthermore, we assume

r10,x(0)𝒳0=Ω(0.1).\left\lVert r\right\rVert_{\mathcal{L}_{\infty}}\leq 10,\quad x(0)\in\mathcal{X}_{0}=\Omega(0.1). (84)

The open-loop dynamics are given by

x˙\displaystyle\dot{x} =[00.00671.3400.86943.200.9931.34]x+[0.1690.25217.31.580.1690.252](u+f(t,x)),\displaystyle\!=\!\begin{bmatrix}0&0.0067&1.34\\ 0&-0.869&43.2\\ 0&0.993&-1.34\end{bmatrix}\!x+\begin{bmatrix}0.169&0.252\\ -17.3&-1.58\\ -0.169&-0.252\end{bmatrix}\!(u\!+\!f(t,x)), (85)

where f(t,x)=[0.8sin(0.4πt)0.1α2,0.10.2α]f(t,x)=[-0.8\sin(0.4\pi t)-0.1\alpha^{2},0.1-0.2\alpha]^{\top} is the uncertainty dependent on both time and α\alpha. The feedback and feedforward gains of the baseline controller 3 are selected to be Kx=[3.25,0.891,7.12;6.10,0.898,10.0]K_{x}=[3.25,0.891,7.12;-6.10,-0.898,-10.0] and Kv=[3.93,0.679;2.57,3.53]K_{v}=[-3.93,0.679;2.57,3.53]. Via simple calculations, we can see that f(t,x)𝒲=[2.4,2.4]×[0.9,0.9]f(t,x)\in\mathcal{W}=[-2.4,2.4]\times[-0.9,0.9] when x𝒳x\in\mathcal{X} holds.

VI-A 1{\mathcal{L}_{1}}-RG Design

It can be verified that given any set 𝒵{\mathcal{Z}}, Lf1,𝒵=0.2maxα𝒵3|α|L_{f_{1},{\mathcal{Z}}}=0.2\max_{\alpha\in{\mathcal{Z}}_{3}}\left\lvert\alpha\right\rvert, Lf2,𝒵=0.2L_{f_{2},{\mathcal{Z}}}=0.2, bf1,𝒵=0.8+0.1maxα𝒵3α2b_{f_{1},{\mathcal{Z}}}=0.8+0.1\max_{\alpha\in{\mathcal{Z}}_{3}}\alpha^{2}, bf2,𝒵=0.1+0.2maxα𝒵3|α|b_{f_{2},{\mathcal{Z}}}=0.1+0.2\max_{\alpha\in{\mathcal{Z}}_{3}}\left\lvert\alpha\right\rvert satisfy Assumption 1. For design of the 1{\mathcal{L}_{1}}AC in 36, 35 and 34, we select Ae=10I3A_{e}=-10I_{3} and parameterize the filter as 𝒞(s)=kfs+kfI2{\mathcal{C}}(s)=\frac{k_{f}}{s+k_{f}}I_{2}, where kf>0k_{f}>0 denotes the bandwidth for both input channels. Table II lists the bounds on xi(t)xn,i(t)x_{i}(t)-x_{\textup{n},i}(t) and uj(t)un,j(t)u_{j}(t)-u_{\textup{n},j}(t) theoretically computed by applying Algorithm 2 under different 𝒞(s){\mathcal{C}}(s) and TT with and without using the scaling technique in Section IV. When applying the scaling technique, we set Txi[k]=0.01\mathrm{T}_{x}^{i}[k]=0.01 for each i,k13i,k\in\mathbb{Z}_{1}^{3} and kik\neq i, which satisfies 56. Several observations can be made from Table II. First, by increasing the filter bandwidth kfk_{f} and decreasing TT, we are able to obtain a smaller γ1\gamma_{1} satisfying 42 and achieve tighter bounds for all states and inputs. In fact, if kf=103k_{f}=10^{3} and T=107T=10^{-7}, then ρ~u1\tilde{\rho}_{u}^{1} ρ~u2\tilde{\rho}_{u}^{2} are fairly close to the bounds on f1(t,x)f_{1}(t,x) and f2(t,x)f_{2}(t,x) for x𝒳x\in\mathcal{X}, respectively, which is consistent with Remark 13. Additionally, with scaling, we could significantly reduce ρ~1\tilde{\rho}^{1} and ρ~3\tilde{\rho}^{3}, the bounds on γ(t)γn(t)\gamma(t)-\gamma_{\textup{n}}(t) and α(t)αn(t)\alpha(t)-\alpha_{\textup{n}}(t), and ρ~u1\tilde{\rho}_{u}^{1} and ρ~u2\tilde{\rho}_{u}^{2}, the bounds on δe(t)δe,n(t)\delta_{e}(t)-\delta_{e,\textup{n}}(t) and δf(t)δf,n(t)\delta_{f}(t)-\delta_{f,\textup{n}}(t). Moreover, with Tx3\mathrm{T}_{x}^{3}, we can verify that the condition 63 holds with bf,𝒳rb_{f,\mathcal{X}_{r}} as long as v<1.868\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}<1.868. As mentioned in Remark 15, the conditions 42 and 31 and the resulting bounds ρ~i\tilde{\rho}^{i} and ρ~uj\tilde{\rho}_{u}^{j} could be conservative. As a result, a larger reference command can potentially be allowed in a practical implementation while keeping x(t)x(t) to stay in 𝒳\mathcal{X}, as demonstrated in the following simulations.

TABLE II: Performance bounds obtained under different filter bandwidth and sample time TT with and without (W/O) scaling
Var. Config. kf=200,T=105k_{f}=200,\ T=10^{-5} kf=103,T=107k_{f}=10^{3},\ T=10^{-7}
W/O scaling With scaling W/O scaling With scaling
γ1/bf,𝒳r\gamma_{1}~{}/~{}b_{f,\mathcal{X}_{r}} 0.01/2.400.01~{}/~{}2.40 2×104/2.402\times 10^{-4}~{}/~{}2.40
[ρ~1,ρ~2,ρ~3][\tilde{\rho}^{1},\tilde{\rho}^{2},\tilde{\rho}^{3}] .41[1,1,1].41[1,1,1] [.015,.41,.038][.015,.41,.038] .043[1,1,1].043[1,1,1] [.12,4.3,.35]102[.12,4.3,.35]10^{-2}
[ρ~u1,ρ~u2][\tilde{\rho}_{u}^{1},\tilde{\rho}_{u}^{2}] [8.15,9.02][8.15,9.02] [4.20,2.85][4.20,2.85] [2.94,1.69][2.94,1.69] [2.51,1.03][2.51,1.03]

Following Algorithm 2, we used the bounds ρ~3\tilde{\rho}^{3}, ρ~u1\tilde{\rho}_{u}^{1} and ρ~u2\tilde{\rho}_{u}^{2} obtained for the case when kf=200k_{f}=200 and T=105T=10^{-5}, to tighten the original constraints 83 and then used the tightened constraints to design the RG, for which we chose Td=0.005T_{d}=0.005. Considering that TdT_{d} was small, we did not consider inter-sample constraint violations and simply set 𝒳^n=𝒳n\hat{\mathcal{X}}_{\textup{n}}=\mathcal{X}_{\textup{n}} and 𝒰^n=𝒰n\hat{\mathcal{U}}_{\textup{n}}={\mathcal{U}}_{\textup{n}} instead of 17. For comparisons, we also designed a robust RG (RRG) that treats the uncertainty f(t,x)f(t,x) as a bounded disturbance w(t)𝒲w(t)\in\mathcal{W}, where 𝒲\mathcal{W} is introduced below 85. RRG design also uses OO_{\infty} set (defined in 22 for RG design); however, the prediction of the output, which corresponds to y^c(k|v,x)\hat{\mathrm{y}}_{c}(k|v,x) for RG design, becomes a set-valued one taking into account all possible realizations of the disturbance w(t)w(t) (see [3] for details). We additionally designed a standard RG by simply ignoring the uncertainty f(t,x)f(t,x).

Refer to caption
Figure 2: Tracking performance under RG, RRG and 1{\mathcal{L}_{1}}-RG
Refer to caption
Figure 3: Trajectories of constrained state and inputs under RG, RRG and 1{\mathcal{L}_{1}}-RG. Green dash-dotted lines illustrate the constraints specified in 83.
Refer to caption
Figure 4: Actual and estimated uncertainties under 1{\mathcal{L}_{1}}-RG. The symbol fjf_{j} (σ^i\hat{\sigma}_{i}) denotes the iith element of ff (σ^\hat{\sigma}), for i=1,2i=1,2.

VI-B Simulation Results

As mentioned in Remark 15, the value of TT theoretically computed according to 42 is often unnecessarily small. For the subsequent simulations, we simply adopted an estimation sample time of 1 millisecond, i.e., T=0.001T=0.001 s. As one can see in the subsequent simulation results, all the bounds derived in Section VI-A for kf=200k_{f}=200 and T=105T=10^{-5} still hold. The reference command r(t)r(t) was set to be [9,6.5][9,6.5] deg for t[0,7.5]t\in[0,7.5] s, and [0,0][0,0] deg for t[7.5,15]t\in[7.5,15] s. The results are shown in Figs. 4, 3 and 2. In terms of constraint enforcement, Fig. 3 shows that both RRG and 1{\mathcal{L}_{1}}-RG successfully enforced all the constraints, while violation of the constraints on the state α(t)\alpha(t) and the input δf(t)\delta_{f}(t) happened under RG. However, from Fig. 2, one can see that the RRG was quite conservative, leading to a large difference between the modified reference and original reference commands and subsequently large tracking errors for both θ(t)\theta(t) and γ(t)\gamma(t) throughout the simulation. In comparison, the modified reference command under RG reached the original reference command, leading to better tracking performance. Finally, 1{\mathcal{L}_{1}}-RG yielded the best tracking performance, driving both θ(t)\theta(t) and γ(t)\gamma(t) very close to their commanded values at steady state. While noticeable under RG and RRG, the uncertainty-induced swaying in the outputs at steady state was negligible under 1{\mathcal{L}_{1}}-RG, thanks to the active compensation of the uncertainty by the 1{\mathcal{L}_{1}}AC. From Fig. 4, one can see that the estimation of the uncertainty within the 1{\mathcal{L}_{1}}-RG was quite accurate.

Refer to caption
Figure 5: Adaptive control inputs and theoretical bounds
Refer to caption
Figure 6: Trajectories of states of the uncertain system (x(t)x(t)) under 1{\mathcal{L}_{1}}-RG and of the nominal system (xn(t)x_{\textup{n}}(t)) under the same command v(t)v(t) and their differences. The actual-nominal state errors and bounds for γ(t)\gamma(t) and α\alpha are scaled by 10 for a clear illustration.

We next check whether the derived uniform bounds on the errors in states, x(t)xn(t)x(t)-x_{\textup{n}}(t), and on the adaptive inputs, ua(t)u_{\textup{a}}(t), hold in the simulation. It can be seen from Fig. 5 that the bounds on both ua,1(t)u_{\textup{a},1}(t) and ua,2(t)u_{\textup{a},2}(t) were respected in the simulation and moreover are fairly tight. Figure 6 reveals that all actual states under 1{\mathcal{L}_{1}}-RG were fairly close to their nominal counterparts, and moreover, the bound on xi(t)xn,i(t)x_{i}(t)-x_{\textup{n},i}(t) for each i13i\in\mathbb{Z}_{1}^{3} was respected. Note that xn(t)x_{\textup{n}}(t) in Fig. 6 was produced by applying the same reference command v(t)v(t) yielded by 1{\mathcal{L}_{1}}-RG to the nominal system 9.

VII Conclusion

In this paper, we developed 1{\mathcal{L}_{1}}-RG, an adaptive reference governor (RG) framework, for control of linear systems with time- and state-dependent uncertainties subject to both state and input constraints. At the core of 1{\mathcal{L}_{1}}-RG is an 1{\mathcal{L}_{1}} adaptive controller that provides guaranteed uniform bounds on the errors between states and inputs of the uncertain system and those of a nominal (i.e., uncertainty-free) system. With such uniform error bounds for constraint tightening, a RG designed for the nominal system with tightened constraints guarantees the satisfaction of the original constraints by the actual states and inputs. Simulation results validate the efficacy and advantages of the proposed approach.

In the future, we will address unmatched uncertainties following [25], and extend the proposed framework to the nonlinear setting leveraging the results in [27, 28]. Additionally, we would like to extend the proposed solution to adaptive MPC.

References

  • [1] E. F. Camacho and C. B. Alba, Model predictive control. Springer Science & Business Media, 2013.
  • [2] J. B. Rawlings, D. Q. Mayne, and M. M. Diehl, Model Predictive Control: Theory, Computation, and Design, 2nd Ed. Nob Hill Publishing, 2020.
  • [3] E. Garone, S. Di Cairano, and I. Kolmanovsky, “Reference and command governors for systems with constraints: A survey on theory and applications,” Automatica, vol. 75, pp. 306–328, 2017.
  • [4] I. Kolmanovsky and E. G. Gilbert, “Theory and computation of disturbance invariant sets for discrete-time linear systems,” Mathematical Problems in Engineering, vol. 4, no. 4, pp. 317–367, 1998.
  • [5] E. C. Kerrigan, Robust constraint satisfaction: Invariant sets and predictive control. PhD thesis, University of Cambridge, 2001.
  • [6] W. Langson, I. Chryssochoos, S. Raković, and D. Q. Mayne, “Robust model predictive control using tubes,” Automatica, vol. 40, no. 1, pp. 125–133, 2004.
  • [7] S. Rakovic, Robust control of constrained discrete time systems: Characterization and implementation. PhD thesis, University of London, 2005.
  • [8] D. Q. Mayne, S. V. Raković, R. Findeisen, and F. Allgöwer, “Robust output feedback model predictive control of constrained linear systems,” Automatica, vol. 42, no. 7, pp. 1217–1222, 2006.
  • [9] D. Q. Mayne, E. C. Kerrigan, E. Van Wyk, and P. Falugi, “Tube-based robust nonlinear model predictive control,” International Journal of Robust and Nonlinear Control, vol. 21, no. 11, pp. 1341–1353, 2011.
  • [10] J. Köhler, R. Soloperto, M. A. Müller, and F. Allgöwer, “A computationally efficient robust model predictive control framework for uncertain nonlinear systems,” IEEE Transactions on Automatic Control, vol. 66, no. 2, pp. 794–801, 2020.
  • [11] B. T. Lopez, J.-J. E. Slotine, and J. P. How, “Dynamic tube MPC for nonlinear systems,” in Proceedings of American Control Conference, pp. 1655–1662, 2019.
  • [12] B. Kouvaritakis and M. Cannon, Model Predictive Control: Classical, Robust and Stochastic. Advanced Textbooks in Control and Signal Processing, Springer, London, 2015.
  • [13] K. Zhang and Y. Shi, “Adaptive model predictive control for a class of constrained linear systems with parametric uncertainties,” Automatica, vol. 117, p. 108974, 2020.
  • [14] V. Adetola, D. DeHaan, and M. Guay, “Adaptive model predictive control for constrained nonlinear systems,” Systems & Control Letters, vol. 58, no. 5, pp. 320–326, 2009.
  • [15] K. Pereida, L. Brunke, and A. P. Schoellig, “Robust adaptive model predictive control for guaranteed fast and accurate stabilization in the presence of model errors,” International Journal of Robust and Nonlinear Control, vol. 31, no. 18, pp. 8750–8784, 2021.
  • [16] X. Wang, L. Yang, Y. Sun, and K. Deng, “Adaptive model predictive control of nonlinear systems with state-dependent uncertainties,” Int. J. Robust Nonlinear Control, vol. 27, no. 17, pp. 4138–4153, 2017.
  • [17] M. Bujarbaruah, S. H. Nair, and F. Borrelli, “A semi-definite programming approach to robust adaptive MPC under state dependent uncertainty,” in European Control Conference, pp. 960–965, IEEE, 2020.
  • [18] N. Hovakimyan and C. Cao, 1\mathcal{L}_{1} Adaptive Control Theory: Guaranteed Robustness with Fast Adaptation. Philadelphia, PA: Society for Industrial and Applied Mathematics, 2010.
  • [19] T. Polóni, U. Kalabić, K. McDonough, and I. Kolmanovsky, “Disturbance canceling control based on simple input observers with constraint enforcement for aerospace applications,” in IEEE Conference on Control Applications, pp. 158–165, IEEE, 2014.
  • [20] G. Pin, D. M. Raimondo, L. Magni, and T. Parisini, “Robust model predictive control of nonlinear systems with bounded and state-dependent uncertainties,” IEEE Transactions on Automatic Control, vol. 54, no. 7, pp. 1681–1687, 2009.
  • [21] E. G. Gilbert and K. T. Tan, “Linear systems with state and control constraints: The theory and application of maximal output admissible sets,” IEEE Transactions on Automatic control, vol. 36, no. 9, pp. 1008–1020, 1991.
  • [22] A. Bemporad and E. Mosca, “Nonlinear predictive reference filtering for constrained tracking,” in Proceedings of European Control Conference, pp. 1720–1725, 1995.
  • [23] E. G. Gilbert, I. Kolmanovsky, and K. T. Tan, “Discrete-time reference governors and the nonlinear control of systems with state and control constraints,” International Journal of Robust and Nonlinear control, vol. 5, no. 5, pp. 487–504, 1995.
  • [24] C. Scherer, P. Gahinet, and M. Chilali, “Multiobjective output-feedback control via LMI optimization,” IEEE Transactions on Automatic Control, vol. 42, no. 7, pp. 896–911, 1997.
  • [25] P. Zhao, S. Snyder, N. Hovakimyana, and C. Cao, “Robust adaptive control of linear parameter-varying systems with unmatched uncertainties,” arXiv:2010.04600, 2021.
  • [26] K. M. Sobel and E. Y. Shapiro, “A design methodology for pitch pointing flight control systems,” Journal of Guidance, Control, and Dynamics, vol. 8, no. 2, pp. 181–187, 1985.
  • [27] A. Lakshmanan, A. Gahlawat, and N. Hovakimyan, “Safe feedback motion planning: A contraction theory and 1\mathcal{L}_{1}-adaptive control based approach,” in Proceedings of 59th IEEE Conference on Decision and Control (CDC), pp. 1578–1583, 2020.
  • [28] P. Zhao, A. Lakshmanan, K. Ackerman, A. Gahlawat, M. Pavone, and N. Hovakimyan, “Tube-certified trajectory tracking for nonlinear systems with robust control contraction metrics,” IEEE Robotics and Automation Letters, pp. 1–1, 2022.

Appendix A Proofs

A-A Proof of Lemma 1

Proof.

Since the continuous-time system 9 has the same states as the discrete-time system 14 at all sampling instants, if 19 holds for 14, then we have

xn(kTd)𝒳^n𝒳n,un(kTd)𝒰^n𝒰n,k+,x_{\textup{n}}(kT_{d})\in\hat{\mathcal{X}}_{\textup{n}}\subset\mathcal{X}_{\textup{n}},\ u_{\textup{n}}(kT_{d})\in\hat{\mathcal{U}}_{\textup{n}}\subset{\mathcal{U}}_{\textup{n}},\quad\forall k\in\mathbb{Z}_{+}, (86)

for 9. Next we analyze the behavior of 9 between adjacent sampling instants. Towards this end, consider any t=kTd+τt=k^{\ast}T_{d}+\tau for some k+k^{\ast}\in\mathbb{Z}_{+} and τ[0,Td)\tau\in[0,T_{d}). From 9, we have xn(t)=xn(kTd+τ)=eAmτxn(kTd)+kTdkTd+τeAm(kTd+τξ)Bvv(τ)𝑑ξ=eAmτxn(kTd)+kTdkTd+τeAm(kTd+τξ)𝑑ξBvv(kTd)=eAmτxn(kTd)+Am1(eAmτIn)Bvv(kTd),x_{\textup{n}}(t)=x_{\textup{n}}(k^{\ast}T_{d}+\tau)=~{}e^{A_{m}\tau}x_{\textup{n}}(k^{\ast}T_{d})+\int_{k^{\ast}T_{d}}^{k^{\ast}T_{d}+\tau}e^{A_{m}(k^{\ast}T_{d}+\tau-\xi)}B_{v}v(\tau)d\xi=e^{A_{m}\tau}x_{\textup{n}}(k^{\ast}T_{d})+\int_{k^{\ast}T_{d}}^{k^{\ast}T_{d}+\tau}e^{A_{m}(k^{\ast}T_{d}+\tau-\xi)}d\xi B_{v}v(k^{\ast}T_{d})=e^{A_{m}\tau}x_{\textup{n}}(k^{\ast}T_{d})+A_{m}^{-1}\left(e^{A_{m}\tau}-I_{n}\right)B_{v}v(k^{\ast}T_{d}), where the third equality is due to the fact that v(kTd+τ)=v(kTd)v(k^{\ast}T_{d}+\tau)=v(k^{\ast}T_{d}) for all τ[0,Td)\tau\in[0,T_{d}). As a result, we have xn(t)xn(kTd)=xn(kTd+τ)xn(kTd)=(eAmτIn)(xn(kTd)+Am1Bvv(kTd))x_{\textup{n}}(t)-x_{\textup{n}}(k^{\ast}T_{d})=x_{\textup{n}}(k^{\ast}T_{d}+\tau)-x_{\textup{n}}(k^{\ast}T_{d})=\left(e^{A_{m}\tau}-I_{n}\right)\left(x_{\textup{n}}(k^{\ast}T_{d})+A_{m}^{-1}B_{v}v(k^{\ast}T_{d})\right). Thus, we have

xn(t)xn(kTd)\displaystyle\left\lVert x_{\textup{n}}(t)-x_{\textup{n}}(k^{\ast}T_{d})\right\rVert_{\infty} ν(Td),\displaystyle\leq\nu(T_{d}), (87a)
un(t)un(kTd)\displaystyle\left\lVert u_{\textup{n}}(t)-u_{\textup{n}}(k^{\ast}T_{d})\right\rVert_{\infty} Kxν(Td),\displaystyle\leq\left\lVert K_{x}\right\rVert_{\infty}\nu(T_{d}), (87b)

where ν(Td)\nu(T_{d}) defined in 18, while 87b is due to the fact that un(t)un(kTd)=Kx(xn(t)xn(kTd))u_{\textup{n}}(t)-u_{\textup{n}}(k^{\ast}T_{d})=K_{x}\left(x_{\textup{n}}(t)-x_{\textup{n}}(k^{\ast}T_{d})\right). Considering 87, 86 and 17, we have xn(t)𝒳nx_{\textup{n}}(t)\in\mathcal{X}_{\textup{n}} and un(t)𝒳nu_{\textup{n}}(t)\in\mathcal{X}_{\textup{n}} for all t0t\geq 0. The proof is complete. ∎

A-B Proof of Lemma 3

Proof.

Rewriting the dynamics of the reference system in (43) in the Laplace domain yields

xr(s)=𝒢xm(s)𝔏[f(t,xr(t))]+xv(s)v(s)+xin(s).x_{\textup{r}}(s)=\mathcal{G}_{xm}(s)\mathfrak{L}\left[f(t,x_{\textup{r}}(t))\right]+\mathcal{H}_{xv}(s)v(s)+x_{\textup{in}}(s). (88)

Therefore, from Lemma 2, for any ξ>0\xi>0, we have

xr[0,ξ]𝒢xm(s)1ηr[0,ξ]+xv(s)1v+xin,\left\lVert x_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\xi]}}\leq\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}\left\lVert\eta_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\xi]}}+\left\lVert\mathcal{H}_{xv}(s)\right\rVert_{\mathcal{L}_{1}}\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}+\left\lVert x_{\textup{in}}\right\rVert_{\mathcal{L}_{\infty}}, (89)

where ηr(t)\eta_{\textup{r}}(t) is defined in 44. If 45 is not true, since xr(t)x_{\textup{r}}(t) is continuous and xr(0)<ρr\left\lVert x_{\textup{r}}(0)\right\rVert_{\infty}<\rho_{r}, there exists a τ>0\tau\!>\!0 such that

xr(t)<ρr,t[0,τ),andxr(τ)=ρr,\left\lVert x_{\textup{r}}(t)\right\rVert_{\infty}<\rho_{r},\ \forall t\in[0,\tau),\ \textup{and}\ \left\lVert x_{\textup{r}}(\tau)\right\rVert_{\infty}=\rho_{r}, (90)

which implies xr(t)Ω(ρr)x_{\textup{r}}(t)\in\Omega(\rho_{r}) for any tt in [0,τ][0,\tau]. Further considering 7b that results from Assumption 1, we have

ηr[0,τ]bf,Ω(ρr).\left\lVert\eta_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq b_{f,\Omega(\rho_{r})}. (91)

Plugging the preceding inequality into 89 leads to

ρr𝒢xm(s)1bf,Ω(ρr)+xv(s)1v+ρin,\rho_{r}\leq\left\lVert\mathcal{G}_{xm}(s)\right\rVert_{\mathcal{L}_{1}}b_{f,\Omega(\rho_{r})}+\left\lVert\mathcal{H}_{xv}(s)\right\rVert_{\mathcal{L}_{1}}\left\lVert v\right\rVert_{\mathcal{L}_{\infty}}+\rho_{\textup{in}}, (92)

which contradicts the condition (31a). Therefore, (45) is true. Equation (46) immediately follows from (45) and 43. ∎

A-C Proof of Lemma 4

Proof.

Due to 48, we have x(t)Ω(ρ)x(t)\in\Omega(\rho) for any tt in [0,τ][0,\tau]. Further considering 7b that results from Assumption 1, we have

f(t,x(t))=η(t)bf,Ω(ρ),t[0,τ].\left\lVert f(t,x(t))\right\rVert_{\infty}=\left\lVert\eta(t)\right\rVert_{\infty}\leq b_{f,\Omega(\rho)},\quad\forall t\in[0,\tau]. (93)

From (47), for any 0t<T0\leq t<T and i0i\in\mathbb{Z}_{0}, we have

x~(iT+t)=\displaystyle\tilde{x}(iT+t)= eAetx~(iT)+iTiT+teAe(iT+tξ)[BB][σ^1(iT)σ^2(iT)]𝑑ξiTiT+teAe(iT+tξ)Bη(ξ)𝑑ξ\displaystyle~{}e^{A_{e}t}\tilde{x}(iT)+\int_{iT}^{iT+t}e^{A_{e}(iT+t-\xi)}[B\ B^{\perp}]\begin{bmatrix}\hat{\sigma}_{1}(iT)\\ \hat{\sigma}_{2}(iT)\end{bmatrix}d\xi-\int_{iT}^{iT+t}e^{A_{e}(iT+t-\xi)}B\eta(\xi)d\xi
=\displaystyle= eAetx~(iT)+0teAe(tξ)[BB][σ^1(iT)σ^2(iT)]𝑑ξ0teAe(tξ)Bη(iT+ξ)𝑑ξ.\displaystyle~{}e^{A_{e}t}\tilde{x}(iT)+\int_{0}^{t}e^{A_{e}(t-\xi)}[B\ B^{\perp}]\begin{bmatrix}\hat{\sigma}_{1}(iT)\\ \hat{\sigma}_{2}(iT)\end{bmatrix}d\xi-\int_{0}^{t}e^{A_{e}(t-\xi)}B\eta(iT+\xi)d\xi. (94)

Considering the adaptive law (35), the preceding equality implies

x~((i+1)T)=0TeAe(Tξ)Bη(iT+ξ)𝑑ξ.\displaystyle\tilde{x}((i+1)T)\!=\!-\!\int_{0}^{T}\!e^{A_{e}(T-\xi)}B\eta(iT\!+\!\xi)d\xi. (95)

Therefore, for any i0i\in\mathbb{Z}_{0} with (i+1)Tτ(i+1)T\leq\tau, we have

x~((i+1)T)\displaystyle\left\lVert\tilde{x}((i+1)T)\right\rVert_{\infty} 0TeAe(Tξ)Bη(iT+ξ)𝑑ξα¯0(T)bf,Ω(ρ),\displaystyle\leq\!\int_{0}^{T}\!\left\lVert e^{A_{e}(T-\xi)}B\right\rVert_{\infty}\left\lVert\eta(iT\!+\!\xi)\right\rVert_{\infty}d\xi\leq\bar{\alpha}_{0}(T)b_{f,\Omega(\rho)}, (96)

where α¯0(T)\bar{\alpha}_{0}(T) is defined in 37a, and the last inequality is due to 93. Since x~(0)=0\tilde{x}(0)=0, we therefore have

x~(iT)α¯0(T)bf,Ω(ρ)γ0(T),iTτ,i0.\left\lVert\tilde{x}(iT)\right\rVert_{\infty}\leq\bar{\alpha}_{0}(T)b_{f,\Omega(\rho)}\leq\gamma_{0}(T),\;\forall iT\leq\tau,i\in\mathbb{Z}_{0}. (97)

Now consider any t(0,T]t\in(0,T] such that iT+tτiT+t\leq\tau with i0i\in\mathbb{Z}_{0}. From (94) and the adaptive law 35, we have

x~(iT+t)\displaystyle\left\lVert\tilde{x}(iT+t)\right\rVert_{\infty}\leq eAetx~(iT)+0teAe(tξ)Φ1(T)eAeTx~(iT)𝑑ξ\displaystyle\left\lVert e^{A_{e}t}\right\rVert_{\infty}\left\lVert\tilde{x}(iT)\right\rVert_{\infty}+\int_{0}^{t}\left\lVert e^{A_{e}(t-\xi)}\Phi^{-1}(T)e^{A_{e}T}\right\rVert_{\infty}\left\lVert\tilde{x}(iT)\right\rVert_{\infty}d\xi
+0teAe(tξ)Bη(iT+ξ)𝑑ξ\displaystyle+\int_{0}^{t}\left\lVert e^{A_{e}(t-\xi)}B\right\rVert_{\infty}\left\lVert\eta(iT+\xi)\right\rVert_{\infty}d\xi
\displaystyle\leq (α¯1(T)+α¯2(T)+1)α¯0(T)bf,Ω(ρ)=γ0(T),\displaystyle\left(\bar{\alpha}_{1}(T)+\bar{\alpha}_{2}(T)+1\right)\bar{\alpha}_{0}(T)b_{f,\Omega(\rho)}=\gamma_{0}(T), (98)

where α¯i(T)\bar{\alpha}_{i}(T) (i=0,1,2i=0,1,2) are defined in 37a, 37b and 37c, and the last inequality is partially due to the fact that 0teAe(tξ)B𝑑ξ0TeAe(Tξ)B𝑑ξ=α¯0(T)\int_{0}^{t}\left\lVert e^{A_{e}(t-\xi)}B\right\rVert_{\infty}d\xi\leq\int_{0}^{T}\left\lVert e^{A_{e}(T-\xi)}B\right\rVert_{\infty}\!d\xi\!=\!\bar{\alpha}_{0}(T). Equations 97 and 98 imply (49). ∎

A-D Proof of Theorem 1

Proof.

We first prove 50c and 50d by contradiction. Assume (50c) or (50d) do not hold. Since xr(0)x(0)=0<γ1\left\lVert x_{\textup{r}}(0)-x(0)\right\rVert_{\infty}=0<\gamma_{1} and ur(0)ua(0)=0<γ2\left\lVert u_{\textup{r}}(0)-u_{\textup{a}}(0)\right\rVert_{\infty}=0<\gamma_{2}, and x(t)x(t), ua(t)u_{\textup{a}}(t), xr(t)x_{\textup{r}}(t) and ur(t)u_{\textup{r}}(t) are all continuous, there must exist an instant τ\tau such that

xr(τ)x(τ)=γ1 or ur(τ)u(τ)=γ2,\left\lVert x_{\textup{r}}(\tau)-x(\tau)\right\rVert_{\infty}=\gamma_{1}\textup{ or }\left\lVert u_{\textup{r}}(\tau)-u(\tau)\right\rVert_{\infty}=\gamma_{2}, (99)

while

xr(t)x(t)<γ1,ur(t)u(t)<γ2,t[0,τ).\left\lVert x_{\textup{r}}(t)\!-\!x(t)\right\rVert_{\infty}\!<\!\gamma_{1},\ \left\lVert u_{\textup{r}}(t)\!-\!u(t)\right\rVert_{\infty}\!<\!\gamma_{2},\ \forall t\in[0,\tau). (100)

This implies that at least one of the following equalities hold:

xrx[0,τ]=γ1,urua[0,τ]=γ2.\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}=\gamma_{1},\quad\left\lVert u_{\textup{r}}-u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}=\gamma_{2}. (101)

Note that xrρr<ρ\left\lVert x_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}}\leq\rho_{r}<\rho according to Lemma 3 and xρr+γ1=ρ\left\lVert x\right\rVert_{\mathcal{L}_{\infty}}\leq\rho_{r}+\gamma_{1}=\rho from 101. Further considering 7a that results from Assumption 1, we have that

f(t,xr(t))f(t,x(t))Lf,Ω(ρ)xrx[0,τ],t[0,τ].\left\lVert f(t,x_{\textup{r}}(t))\!-\!f(t,x(t))\right\rVert_{\infty}\!\leq\!L_{f,\Omega(\rho)}\!\left\lVert x_{\textup{r}}\!-\!x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\!,\ \forall t\!\in\![0,\tau]. (102)

The control laws in 36 and 43 indicate

ur(s)ua(s)=𝒞(s)𝔏[f(t,xr)σ^1(s)]=𝒞(s)𝔏[f(t,x)f(t,xr)]+𝒞(s)(σ^1(s)𝔏[f(t,x)]).\displaystyle u_{\textup{r}}(s)-u_{\textup{a}}(s)=-{\mathcal{C}}(s)\mathfrak{L}\left[f(t,x_{\textup{r}})-\hat{\sigma}_{1}(s)\right]={\mathcal{C}}(s)\mathfrak{L}\left[f(t,x)\!-\!f(t,x_{\textup{r}})\right]+{\mathcal{C}}(s)(\hat{\sigma}_{1}(s)\!-\!\mathfrak{L}\left[f(t,x)\right]). (103)

Equation (47) indicates that

σ^1(s)𝔏[f(t,x)])=B(sInAe)x~(s).\hat{\sigma}_{1}(s)-\mathfrak{L}\left[f(t,x)\right])=B^{\dagger}(sI_{n}-A_{e})\tilde{x}(s). (104)

Considering (5), 36 and 104, we have

x(s)=𝒢xm(s)𝔏[f(t,x)]+xv(s)v(s)+xin(s)xm(s)𝒞(s)B(sInAe)x~(s),x(s)=\mathcal{G}_{xm}(s)\mathfrak{L}\left[f(t,x)\right]+\mathcal{H}_{xv}(s)v(s)+x_{\textup{in}}(s)-\mathcal{H}_{xm}(s){\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\tilde{x}(s), (105)

which, together with 88, implies

xr(s)x(s)=𝒢xm(s)𝔏[f(t,xr)f(t,x)]+xm(s)𝒞(s)B(sInAe)x~(s).x_{\textup{r}}(s)-x(s)=\mathcal{G}_{xm}(s)\mathfrak{L}\left[f(t,x_{\textup{r}})-f(t,x)\right]+\mathcal{H}_{xm}(s){\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\tilde{x}(s). (106)

Therefore, further considering (102) and Lemma 4, we have

xrx[0,τ]\displaystyle\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}\leq 𝒢xm1Lf,Ω(ρ)xrx[0,τ]+xm(s)𝒞(s)B(sInAe)1γ0(T).\displaystyle\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}}L_{f,\Omega(\rho)}\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}+\!\left\lVert\mathcal{H}_{xm}(s){\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\right\rVert_{\mathcal{L}_{1}}\!\gamma_{0}(T).

The preceding equation, together with 31b, leads to

xrx[0,τ]\displaystyle\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}} xm(s)𝒞(s)B(sInAe)11𝒢xm1Lf,Ω(ρ)γ0(T),\displaystyle\!\leq\!\frac{\left\lVert\mathcal{H}_{xm}(s){\mathcal{C}}(s)B^{\dagger}(sI_{n}\!-\!A_{e})\right\rVert_{\mathcal{L}_{1}}}{1-\left\lVert\mathcal{G}_{xm}\right\rVert_{\mathcal{L}_{1}}L_{f,\Omega(\rho)}}\gamma_{0}(T), (107)

which, together with the sample time constraint 42, indicates that

xrx[0,τ]<γ1.\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}<\gamma_{1}. (108)

On the other hand, it follows from 102, 103, 104 and 108 that

urua[0,τ]\displaystyle\left\lVert u_{\textup{r}}-u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}} 𝒞(s)1Lf,Ω(ρ)xrx[0,τ]+𝒞(s)B(sInAe)1x~[0,τ]\displaystyle\leq\left\lVert{\mathcal{C}}(s)\right\rVert_{\mathcal{L}_{1}}L_{f,\Omega(\rho)}\left\lVert x_{\textup{r}}-x\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}+\left\lVert{\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\right\rVert_{\mathcal{L}_{1}}\left\lVert\tilde{x}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}
<𝒞(s)1Lf,Ω(ρ)γ1+𝒞(s)B(sInAe)1γ0(T).\displaystyle<\left\lVert{\mathcal{C}}(s)\right\rVert_{\mathcal{L}_{1}}L_{f,\Omega(\rho)}\gamma_{1}+\left\lVert{\mathcal{C}}(s)B^{\dagger}(sI_{n}-A_{e})\right\rVert_{\mathcal{L}_{1}}\gamma_{0}(T).

Further considering the definition in 40, we have

urua[0,τ]<γ2.\left\lVert u_{\textup{r}}-u_{\textup{a}}\right\rVert_{\mathcal{L}_{\infty}^{[0,\tau]}}<\gamma_{2}. (109)

Note that 108 and 109 contradict the equalities in 101, which proves 50c and 50d. The bounds in 50a and 50b follow directly from 50c, 50d, 45 and 46 and the definitions of ρ\rho and ρua\rho_{u_{\textup{a}}} in 32 and 41. The proof is complete. ∎

A-E Proof of Lemma 5

Proof.

From 9a and 43, we have

xr(s)xn(s)=Gxm(s)𝔏[f(t,xr)]=Gxm(s)𝔏[ηr(t)].x_{\textup{r}}(s)-x_{\textup{n}}(s)=G_{xm}(s)\mathfrak{L}\left[f(t,x_{\textup{r}})\right]=G_{xm}(s)\mathfrak{L}\left[\eta_{\textup{r}}(t)\right]. (110)

According to Lemma 3, we have xr(t)Ω(ρr)x_{\textup{r}}(t)\in\Omega(\rho_{r}) for any t0t\geq 0. Further considering 7b that results from Assumption 1, we have ηrbf,Ω(ρr)\left\lVert\eta_{\textup{r}}\right\rVert_{\mathcal{L}_{\infty}}\leq b_{f,\Omega(\rho_{r})}, which, together with 110, leads to 51. ∎