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

Hybrid Minimum-Seeking in Synergistic Lyapunov Functions: Robust Global Stabilization under Unknown Control Directions

Mahmoud Abdelgalil [email protected]    Jorge I. Poveda [email protected] Department of Electrical and Computer Engineering, University of California, San Diego, La Jolla, CA, USA.
Abstract

We study the problem of robust global stabilization in control-affine systems, focusing on dynamic uncertainties in the control directions and the presence of topological obstructions that prevent the existence of smooth global control Lyapunov functions. Building on a recently developed Lie-bracket averaging result for hybrid dynamic inclusions presented in [1], we propose a novel class of universal hybrid feedback laws that achieve robust global practical stability by identifying the minimum point of a set of appropriately chosen synergistic Lyapunov functions. As concrete applications of our results, we synthesize different hybrid high-frequency high-amplitude feedback laws for the solution of robust global stabilization problems on various types of manifolds under unknown control directions, as well as controllers for obstacle avoidance problems in vehicles characterized by kinematic models describing both holonomic and non-holonomic models. By leveraging Lie-bracket averaging for hybrid systems, we also show how the proposed hybrid minimum-seeking feedback laws can overcome lack of controllability during persistent (bounded) periods of time. Numerical simulation results are presented to illustrate the main results.

keywords:
Hybrid systems, Adaptive systems, Nonlinear Control, Stability and Stabilization.
thanks: This work was supported in part by the grants NSF ECCS CAREER 2305756 and AFOSR YIP: FA9550-22-1-0211. Corresponding Author: Jorge I. Poveda.

, ,

1 Introduction

Ensuring robustness in the face of uncertainty is a core challenge in controller design for autonomous systems. This challenge becomes significantly more difficult if the uncertainty lies in the control direction. For example, cyber-physical systems can experience significant damage when the control gain’s sign unexpectedly changes or vanishes. Such a scenario may occur without malicious interference, such as from internal software failures, or due to external spoofing attacks by adversarial agents [2]. Furthermore, the control direction may be time-varying and intermittently zero (indicating a lack of controllability) for bounded periods. In such cases, traditional identification or parameter estimation techniques are often inadequate due to insufficient excitation in the system’s trajectories. As a result, designing stabilizing feedback laws capable of addressing uncertainty in control directions and persistent periods of uncontrollability is essential for ensuring the resilience of autonomous systems operating in complex and dynamic environments.

Referred to as the problem of “stabilization with an unknown sign of the high-frequency gain”, or “stabilization under unknown control directions”, designing stabilizing feedback laws in the presence of uncertainty on control directions has a long history in the literature of robust adaptive control that dates back to the 1970s [3]. The first solution was proposed by Nussbaum [4], which became the common approach to solving this problem. However, it is well-known, e.g. see [5], that the Nussbaum approach can lead to poor performance if the control sign is dynamic. An alternative model-free approach for solving this difficult problem was introduced by Scheinker and Krstic in [5, 6]. The approach introduced in [5] is based on seeking the minimum of a suitably constructed control Lyapunov-like function VV, using a high-frequency time-varying oscillatory feedback law that relies only on real-time evaluations of this function. As shown in [5, 6], this model-free approach is impervious to persistent changes in the sign of the control gain, which makes it an ideal candidate for the design of stabilizing feedback laws operating in uncertain and adversarial environments. Since its conception in [5], the minimum-seeking approach has found numerous applications, including in the context of 2-D Vehicle Control [6], output regulation problems in nonlinear systems [7], etc, see [8, Sec. 6] for a recent survey on this subject.

A core theoretical development that enabled the results of [5, 6] was the emergence of Lie-bracket averaging for ODEs as a powerful framework for the analysis and design of model-free control and optimization algorithms [9, 10, 11, 12]. Nevertheless, since the feedback laws analyzed in [5, 6] are continuous and derive their stability properties from the stability of their Lie-bracket averaged systems (which are also continuous), their performance is inherently constrained by the standard limitations associated with continuous feedback laws [13]. In particular, when the system operates in topological spaces that are not contractible to a point, as in, e.g., smooth compact manifolds, robust global stabilization of a desired point is not possible via continuous feedback [14]. Such types of topological obstructions to robust global stabilization have been shown to emerge in different applications, including obstacle avoidance problems [15, 16], the attitude stabilization of rigid bodies[17, 18], stabilization problems in 𝕊1\mathbb{S}^{1}, synchronization problems [17], etc. When the sign in the control direction is static and known a priori, such stabilization problems can be tackled via synergistic hybrid control [19, 16, 20]. However, most existing results on synergistic hybrid control cannot cope with uncertainty in the control directions, particularly when such uncertainty is dynamic. Indeed, even the state-of-the-art results on synergistic hybrid feedback for uncertain systems [21] are inapplicable when the control gain sign is uncertain.

Motivated by the above challenges, the main contribution of this paper is to provide a solution for robust global stabilization problems under unknown control directions in spaces that are not contractible to a point. Specifically, the following are the main contributions of this paper:

1) We propose a novel class of hybrid and oscillatory feedback laws that leverage the existence of a collection of suitable local strong control Lyapunov functions (SCLFs) available to the plant. These local SCLFs are then used together to control the system when a global SCLF is unavailable. Building on existing results in model-based hybrid control [19, 18, 16] and using a recently introduced Lie-bracket averaging result for hybrid inclusions [1], we establish semi-global practical asymptotic stability for systems with unbounded operational sets, and global practical asymptotic stability for systems on smooth compact manifolds, even with unknown control directions that vanish intermittently.

2) Subsequently, we apply our results to various control problems with unknown control directions where a robustly globally stabilizing continuous control law does not exist even if the control direction is known, including stabilization problems on 𝕊1\mathbb{S}^{1}, 𝕊2\mathbb{S}^{2}, SO(3)\text{SO}(3), and stabilization problems on the plane with obstacle avoidance for both holonomic and non-holonomic vehicles.

3) Finally, we present different numerical examples to demonstrate the performance of the proposed controllers under different types of uncertainty in the control directions. Additionally, we compare their effectiveness against both non-hybrid approaches and hybrid non-adaptive methods. In all applications, we show how our results open the door for the design of “model-free” controllers based on existing well-posed model-based hybrid algorithms [22, 20, 16, 23], thus demonstrating how the synergistic use of hybrid control [20] and Lie-bracket averaging [1] can simultaneously overcome uncertainties in control directions and the topological obstructions to global stabilization.

To the best of our knowledge, the results presented in this paper provide the first theoretical link between the minimum-seeking approach introduced in [6], and the setting of synergistic hybrid control [19, 20].

The remainder of this manuscript is organized as follows. In Section 2, we present the preliminaries. Section 3 presents the main problem formulation and main theoretical result. Section 4 focuses on robust global stabilization problems on smooth manifolds. The proofs are presented in Section 5, followed by the conclusions and future work in Section 6.

2 Preliminaries

2.1 Notation

We use x,y=xy\langle x,y\rangle=x^{\top}y, to denote the inner product between any two vectors x,ynx,y\in\mathbb{R}^{n}. Given a compact set 𝒜n\mathcal{A}\subset\mathbb{R}^{n} and xnx\in\mathbb{R}^{n}, we use |x|𝒜:=minx~𝒜xx~2|x|_{\mathcal{A}}:=\min_{\tilde{x}\in\mathcal{A}}\|x-\tilde{x}\|_{2}. A set-valued mapping M:pnM:\mathbb{R}^{p}\rightrightarrows\mathbb{R}^{n} is outer semicontinuous (OSC) at zz if for each sequence {zi,si}(z,s)p×n\{z_{i},s_{i}\}\to(z,s)\in\mathbb{R}^{p}\times\mathbb{R}^{n} satisfying siM(zi)s_{i}\in M(z_{i}) for all i0i\in\mathbb{Z}_{\geq 0}, we have sM(z)s\in M(z). A mapping MM is locally bounded (LB) at zz if there exists an open neighborhood NzpN_{z}\subset\mathbb{R}^{p} of zz such that M(Nz)M(N_{z}) is bounded. The mapping MM is OSC and LB relative to a set KpK\subset\mathbb{R}^{p} if MM is OSC for all zKz\in K and M(K):=zKM(x)M(K):=\cup_{z\in K}M(x) is bounded. A function β:0×00\beta:\mathbb{R}_{\geq 0}\times\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0} is of class 𝒦\mathcal{K}\mathcal{L} if it is nondecreasing in its first argument, nonincreasing in its second argument, limr0+β(r,s)=0\lim_{r\to 0^{+}}\beta(r,s)=0 for each s0s\in\mathbb{R}_{\geq 0}, and limsβ(r,s)=0\lim_{s\to\infty}\beta(r,s)=0 for each r0r\in\mathbb{R}_{\geq 0}. For two (or more) vectors u,vnu,v\in\mathbb{R}^{n}, we write (u,v)=[u,v](u,v)=[u^{\top},v^{\top}]^{\top}. If x,y3x,y\in\mathbb{R}^{3}, we use [x]×[x]_{\times} to denote the the skew-symmetric matrix associated to xx and defined such that [x]×y=x×y[x]_{\times}y=x\times y, where ×\times indicates the cross product of vectors in 3\mathbb{R}^{3}. The notation \otimes denotes the Kronecker product. Let x=(x1,x2,,xn)nx=(x_{1},x_{2},\ldots,x_{n})\in\mathbb{R}^{n}. Given a Lipschitz continuous function f:nmf:\mathbb{R}^{n}\to\mathbb{R}^{m}, we use xif\partial_{x_{i}}f to denote the generalized Jacobian [24] of ff with respect to the variable xix_{i}. A map ff is said to be of class 𝒞k\mathcal{C}^{k} if it is kk-times continuously differentiable with the kkth-derivative being locally Lipschitz continuous. If CnC\subset\mathbb{R}^{n}, the notation TxCT_{x}C denotes the tangent cone [24] of CC at the point xx. When CC is an embedded Euclidean submanifold, TxCT_{x}C denotes the tangent space at xx, which is isomorphic to an affine linear subspace of n\mathbb{R}^{n}.

2.2 Hybrid Dynamical Systems

In this paper, our models are given by hybrid dynamical systems (HDS), as studied in [25]. Such systems are characterized by the following inclusions:

:{xC,x˙F(x)xD,x+G(x),\displaystyle\mathcal{H}:~{}~{}~{}\begin{cases}~{}~{}x\in C,&\dot{x}\hphantom{{}^{+}}\in F(x)\\ ~{}~{}x\in D,&x^{+}\in G(x),\end{cases} (1a)

where F:nnF:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{n} is called the flow map, G:nnG:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{n} is called the jump map, CnC\subset\mathbb{R}^{n} is called the flow set, and DnD\subset\mathbb{R}^{n} is called the jump set. We use =(C,F,D,G)\mathcal{H}=(C,F,D,G) to denote the data of the HDS \mathcal{H}. Purely continuous-time systems can be modeled as (1) by taking D={}D=\{\emptyset\}. Similarly, purely discrete-time systems can be modeled as (1) by taking C={}C=\{\emptyset\}. In this paper, we work with well-posed HDS that satisfy the following assumption [25, Assumption 6.5].

Assumption 1.

[Hybrid Basic Conditions] The sets C,DC,D are closed. The set-valued mapping FF is OSC, LB, and for each xCx\in C the set F(x)F(x) is convex and nonempty. The set-valued mapping GG is OSC, LB, and for each xDx\in D the set G(x)G(x) is nonempty. \square

Henceforth, all hybrid systems in this manuscript are assumed to satisfy the Hybrid Basic Conditions.

Solutions to system (1) are parameterized by a continuous-time index t0t\in\mathbb{R}_{\geq 0}, which increases continuously during flows, and a discrete-time index j0j\in\mathbb{Z}_{\geq 0}, which increases by one during jumps. Therefore, solutions to (1) are defined on hybrid time domains (HTDs). A set E0×0E\subset\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0} is called a compact HTD if E=j=0J1([tj,tj+1],j)E=\cup_{j=0}^{J-1}([t_{j},t_{j+1}],j) for some finite sequence of times 0=t0t1tJ0=t_{0}\leq t_{1}\ldots\leq t_{J}. The set EE is a HTD if for all (T,J)E(T,J)\in E, E([0,T]×{0,,J})E\cap([0,T]\times\{0,\ldots,J\}) is a compact HTD. The following definition formalizes the notion of solution to HDS of the form (1).

Definition 1.

A hybrid arc xx is a function defined on a HTD. In particular, x:dom(x)nx:\text{dom}(x)\to\mathbb{R}^{n} is such that x(,j)x(\cdot,j) is locally absolutely continuous for each jj such that the interval Ij:={t:(t,j)dom(x)}I_{j}:=\{t:(t,j)\in\text{dom}(x)\} has a nonempty interior. A hybrid arc x:dom(x)nx:\text{dom}(x)\to\mathbb{R}^{n} is a solution xx to the HDS (1) if x(0,0)CDx(0,0)\in C\cup D, and: 1) For all j0j\in\mathbb{Z}_{\geq 0} such that IjI_{j} has nonempty interior: x(t,j)Cx(t,j)\in C for all tIjt\in I_{j}, and x˙(t,j)F(x(t,j))\dot{x}(t,j)\in F(x(t,j)) for almost all tIjt\in I_{j}; 2) For all (t,j)dom(x)(t,j)\in\text{dom}(x) such that (t,j+1)dom(x)(t,j+1)\in\text{dom}(x): x(t,j)Dx(t,j)\in D and x(t,j+1)G(x(t,j))x(t,j+1)\in G(x(t,j)). A solution xx is said to be maximal if it cannot be further extended, and it is said to be complete if lengthdom(x)=\text{length}~{}\text{dom}(x)=\infty. \square

In this paper, we use the following standard stability notion:

Definition 2.

A compact set 𝒜n\mathcal{A}\subset\mathbb{R}^{n} is said to be uniformly globally asymptotically stable (UGAS) for the HDS (1) if there exists β𝒦\beta\in\mathcal{KL} such that each solution xx to (1) satisfies |x(t,j)|𝒜β(|x(0,0)|𝒜,t+j)|x(t,j)|_{\mathcal{A}}\leq\beta(|x(0,0)|_{\mathcal{A}},t+j), for all (t,j)dom(x)(t,j)\in\text{dom}(x). \square

In some cases, we consider HDS that depend on a small tunable parameter ε>0\varepsilon\in\mathbb{R}_{>0}, given by

ε:{Cx˙Fε(x)Dx+Gε(x).\displaystyle\mathcal{H}_{\varepsilon}:\begin{cases}C&\dot{x}\hphantom{{}^{+}}\in F_{\varepsilon}(x)\\ D&x^{+}\in G_{\varepsilon}(x).\end{cases} (2a)

For system (2), we use the following stability notion, which is standard in the literature [26, 6].

Definition 3.
\thlabel

defn:SPUAS For the HDS (2), a compact set 𝒜n\mathcal{A}\subset\mathbb{R}^{n} is said to be Semi-Globally Practically Asymptotically Stable (SGpAS) as ε0+\varepsilon\rightarrow 0^{+} if there exists β𝒦\beta\in\mathcal{KL} such that for each compact set KCDK\subset C\cup D and for each ν>0\nu>0, there exists ε>0\varepsilon^{*}>0 such that for all ε(0,ε]\varepsilon\in(0,\varepsilon^{*}], all solutions of (2) with x(0,0)Kx(0,0)\in K satisfy |x(t,j)|𝒜β(|x(0,0)|𝒜,t+j)+ν|x(t,j)|_{\mathcal{A}}\leq\beta(|x(0,0)|_{\mathcal{A}},t+j)+\nu, for all (t,j)dom(x)(t,j)\in\text{dom}(x). \square

If there are several parameters that need to be tuned sequentially, we use the notation (ε1,ε2,,ε)0+(\varepsilon_{1},\varepsilon_{2},\dots,\varepsilon_{\ell})\rightarrow 0^{+} to encode the sequence of parameter tuning, i.e. ε1\varepsilon_{1} is tuned, then ε2\varepsilon_{2} after fixing ε1\varepsilon_{1}, and so forth. Finally, note that when CDC\cup D is a compact set, SGpAS reduces to Uniform Global Practical Asymptotic Stability (UGpAS).

3 Strong V\nabla V-Stabilizability of Control-Affine HDS

In this section, we present the model of the systems under study, and a general result on model-free stabilization of control-affine HDS with unknown control directions.

3.1 Model

Consider the open-loop HDS:

{x𝒳C,x˙=f(x,θ,u)x𝒳D,x+G𝒳(x)\displaystyle\begin{cases}~{}x\in\mathcal{X}_{C},&\dot{x}\hphantom{{}^{+}}=f(x,\theta,u)\\ ~{}x\in\mathcal{X}_{D},&x^{+}\in G_{\mathcal{X}}(x)\end{cases} (3a)
where xnx\in\mathbb{R}^{n} is the main state, u:=(u1,u2,,ur)ru:=(u_{1},u_{2},\dots,u_{r})\in\mathbb{R}^{r} are the control inputs, 𝒳Cn\mathcal{X}_{C}\subset\mathbb{R}^{n} is the flow set, 𝒳Dn\mathcal{X}_{D}\subset\mathbb{R}^{n} is the jump set, G𝒳:nnG_{\mathcal{X}}:\mathbb{R}^{n}\rightrightarrows\mathbb{R}^{n} is the jump map, and f:n×m×rf:\mathbb{R}^{n}\times\mathbb{R}^{m}\times\mathbb{R}^{r} is the flow map, given by
f(x,θ,u)\displaystyle f(x,\theta,u) :=f0(x,θ)+i=1rfi(x,θ)ui.\displaystyle:=f_{0}(x,\theta)+\sum_{i=1}^{r}f_{i}(x,\theta)u_{i}. (3b)

In (3b), θΘm\theta\in\Theta\subset\mathbb{R}^{m} represents a vector of potentially time-varying unknown parameters that models the unknown control directions. The class of systems (3) generalizes the class of control-affine systems commonly studied in the literature on continuous-time systems modeled as ODEs [6] by allowing for jumps in the solutions. We can also consider the class of systems (3b) as an open-loop HDS in the sense of [27], where the flow set, jump set, and jump map have been designed, and it remains to design the continuous feedback laws uiu_{i}. We assume that (3) satisfies the Hybrid Basic Conditions.

The following regularity condition on (3) is needed to guarantee existence of maximal solutions and to rule out pathological behaviors, such as Zeno solutions.

Assumption 2.

For all i{1,2,,r}i\in\{1,2,\dots,r\} and all (x,θ)𝒳C×Θ(x,\theta)\in\mathcal{X}_{C}\times\Theta, the following holds

  1. 1.

    f0f_{0} and fif_{i} are 𝒞0\mathcal{C}^{0}, and fi(,θ)f_{i}(\cdot,\theta) is 𝒞1\mathcal{C}^{1}.

  2. 2.

    f0(x,θ)Tx𝒳Cf_{0}(x,\theta)\in T_{x}\mathcal{X}_{C}, and fi(x,θ)Tx𝒳Cf_{i}(x,\theta)\in T_{x}\mathcal{X}_{C}.

  3. 3.

    G(𝒳D)𝒳CG(\mathcal{X}_{D})\subset\mathcal{X}_{C}, and G(𝒳D)𝒳D=G(\mathcal{X}_{D})\cap\mathcal{X}_{D}=\emptyset.

The vector of unknown parameters θ\theta may be constant or time-varying. In the latter case, we model θ\theta as the solution of an exogenous HDS of the form

{θΘCθ˙Fe(θ)θΘDθ+Ge(θ),\displaystyle\begin{cases}~{}~{}\theta\in\Theta_{C}&\dot{\theta}\hphantom{{}^{+}}\in F_{e}(\theta)\\ ~{}~{}\theta\in\Theta_{D}&\theta^{+}\in G_{e}(\theta),\end{cases} (4)

where ΘC,ΘDΘ\Theta_{C},\Theta_{D}\subset\Theta. We remark that the case of static uncertainty θ˙=0\dot{\theta}=0 is trivially included in (4) by taking ΘC={Θ}\Theta_{C}=\{\Theta\}, ΘD=\Theta_{D}=\emptyset, and Fe(θ)={0}F_{e}(\theta)=\{0\}. As usual, we require that the HDS (4) satisfies the Hybrid Basic Conditions of Assumption 1. In addition, to guarantee existence of solutions to (4) from any initial condition in ΘCΘD\Theta_{C}\cup\Theta_{D}, we impose the following regularity assumption.

Assumption 3.

a) For each ϑΘC\vartheta\in\Theta_{C} there exists a neighborhood UU of ϑ\vartheta such that, for all θΘCU\theta\in\Theta_{C}\cap U we have Fe(θ)TθΘCF_{e}(\theta)\cap T_{\theta}\Theta_{C}\neq\emptyset; b) Ge(ΘD)ΘCG_{e}(\Theta_{D})\subset\Theta_{C}; d) and Ge(ΘD)ΘD=G_{e}(\Theta_{D})\cap\Theta_{D}=\emptyset.

Under Assumption 3, the HDS defined by (4) is sufficiently general to model a variety of complex behaviors. For example, the HDS (3) can model various classes of switching signals [25, Section 2.4], including those that have dwell-time bounds, which can be generated using hybrid automata [28, 29]. While item (a) in Assumption 3 rules out consecutive jumps of θ\theta, this assumption can be relaxed to consider systems that satisfy average dwell-time bounds. However, since they rarely occur in practice, we rule out consecutive jumps of the control direction.

Remark 1.

While θ\theta is treated in this paper as an uncertain parameter, the regularity properties on the dynamics of θ\theta are necessary so as to exclude pathological behaviors, e.g. Zeno solutions, purely discrete solutions, etc. In particular, by [30, Proposition 2.34], Assumption 3 guarantees that, due to the compactness of Θ\Theta, there exists t>0t_{\circ}>0 such that any solution θ\theta to (4) satisfies t¯jt¯jt>0\overline{t}_{j}-\underline{t}_{j}\geq t_{\circ}>0 for all jj in the domain of θ\theta, where t¯j:=sup{t0:(t,j)dom(θ)}\overline{t}_{j}:=\sup\{t\in\mathbb{R}_{\geq 0}~{}:~{}(t,j)\in\text{dom}(\theta)\}, t¯j:=inf{t0:(t,j)dom(θ)}\underline{t}_{j}:=\,\inf\,\{t\in\mathbb{R}_{\geq 0}~{}:~{}(t,j)\in\text{dom}(\theta)\}. \square

By interconnecting systems (3) and (4), we obtain the following open-loop HDS with unknown control directions:

:{ξC,ξ˙F(ξ,u)ξD,ξ+G(ξ),\displaystyle\mathcal{H}:~{}\begin{cases}~{}~{}\xi\in C,&\dot{\xi}\hphantom{{}^{+}}\in F(\xi,u)\\ ~{}~{}\xi\in D,&\xi^{+}\in G(\xi),\end{cases} (5a)
where ξ=(x,θ)\xi=(x,\theta), and where the sets CC and DD are given by
C\displaystyle C =𝒳C×ΘC,\displaystyle=\mathcal{X}_{C}\times\Theta_{C}, (5b)
D\displaystyle D =(𝒳C×ΘD)(𝒳D×ΘC)(𝒳D×ΘD),\displaystyle=(\mathcal{X}_{C}\times\Theta_{D})\cup(\mathcal{X}_{D}\times\Theta_{C})\cup(\mathcal{X}_{D}\times\Theta_{D}), (5c)
the flow map FF is given by
F(ξ,u)\displaystyle F(\xi,u) ={f(x,θ,u)}×Fe(θ),\displaystyle=\{f(x,\theta,u)\}\times F_{e}(\theta), (5d)
and the jump map GG is given by
G(ξ)={G𝒳(x)×Ge(θ)(x,θ)𝒳D×ΘD{x}×Ge(θ)(x,θ)𝒳C×ΘDG𝒳(x)×{θ}(x,θ)𝒳D×ΘC.\displaystyle G(\xi)=\begin{cases}G_{\mathcal{X}}(x)\times G_{e}(\theta)&(x,\theta)\in\mathcal{X}_{D}\times\Theta_{D}\\ \{x\}\times G_{e}(\theta)&(x,\theta)\in\mathcal{X}_{C}\times\Theta_{D}\\ G_{\mathcal{X}}(x)\times\{\theta\}&(x,\theta)\in\mathcal{X}_{D}\times\Theta_{C}.\end{cases} (5e)

In words, the jump map (5e) allows jumps in the system whenever xx, or θ\theta, are in their respective jump sets 𝒳D\mathcal{X}_{D} or ΘD\Theta_{D}, respectively. Note that, by construction, \mathcal{H} satisfies Assumption 1 whenever the individual elements of (3) and (4) satisfy the Hybrid Basic Conditions. In addition, the following Lemma is straightforward to verify.

Lemma 1.

Suppose that Assumptions 2 and 3 hold. Then, G(D)CG(D)\subset C, and G(D)D=G(D)\cap D=\emptyset.

3.2 SCLFs for HDS with Unknown Control Directions

To study the stabilization problem of system (5), and to simplify some expressions, in the sequel we let 𝒳=𝒳C𝒳D\mathcal{X}=\mathcal{X}_{C}\cup\mathcal{X}_{D}, Θ=ΘCΘD\Theta=\Theta_{C}\cup\Theta_{D}, and πx:n×mξ=(x,θ)xn\pi_{x}:\mathbb{R}^{n}\times\mathbb{R}^{m}\ni\xi=(x,\theta)\mapsto x\in\mathbb{R}^{n} be the canonical projection onto the first state. The following definition is inspired by the notion of “strong LgVL_{g}V stabilizability”, introduced in [6, 5] for input-affine ODEs.

Definition 3.1.

Let 𝒜𝒳\mathcal{A}\subset\mathcal{X} be a compact set. The 𝒞1\mathcal{C}^{1} function V:n0V:\mathbb{R}^{n}\rightarrow\mathbb{R}_{\geq 0} is said to be a Strong Control Lyapunov Function (SCLF) candidate with respect to 𝒜\mathcal{A} for \mathcal{H} if there exists γ>0\gamma>0 and class 𝒦\mathcal{K}_{\infty}-functions α1\alpha_{1}, α2\alpha_{2} such that:

  1. (a)

    For all x𝒳x\in\mathcal{X}, we have:

    α1(|x|𝒜)V(x)α2(|x|𝒜).\displaystyle\alpha_{1}(|x|_{\mathcal{A}})\leq\,V(x)\leq\alpha_{2}(|x|_{\mathcal{A}}). (6a)
  2. (b)

    For all ξC\xi\in C, we have:

    V˙(ξ):=V(x),f¯(ξ)0,\displaystyle\dot{V}(\xi):=\langle\nabla V(x),\bar{f}(\xi)\rangle\leq 0, (6b)

    where f¯\bar{f} is given by

    f¯(ξ)\displaystyle\bar{f}(\xi) :=f0(x,θ)γi=1rV(x),fi(x,θ)fi(x,θ).\displaystyle:=f_{0}(x,\theta)-\gamma\sum_{i=1}^{r}\langle\nabla V(x),f_{i}(x,\theta)\rangle f_{i}(x,\theta). (6c)
  3. (c)

    For all ξD\xi\in D, we have:

    ΔV(ξ):=maxgG(ξ)V(πx(g))V(πx(ξ))0.\displaystyle\Delta V(\xi):=\max_{g\in G(\xi)}V(\pi_{x}(g))-V(\pi_{x}(\xi))\leq 0. (6d)
Remark 3.2.

Definition 3.1 aims to generalize to HDS the “strong LgVL_{g}V stabilizability” property studied in [6, 5] for ODEs. Indeed, condition (6b) can be written in a more explicit form as

V(x),f0(x,θ)γi=1rV(x),fi(x,θ)20,\displaystyle\langle\nabla V(x),f_{0}(x,\theta)\rangle-\gamma\sum_{i=1}^{r}\langle\nabla V(x),f_{i}(x,\theta)\rangle^{2}\leq 0, (7)

for all (x,θ)C(x,\theta)\in C. If, in addition, for all (x,θ)C(x,\theta)\in C such that |x|𝒜=ϵ>0|x|_{\mathcal{A}}=\epsilon>0, the function VV satisfies the “strong small control property” [6, Sec. 3.1]:

limϵ0maxV(x),f0(x,θ)>0V(x),f0(x,θ)i=1rV(x),fi(x,θ)2<+,\displaystyle\lim_{\epsilon\rightarrow 0}\max_{\langle\nabla V(x),f_{0}(x,\theta)\rangle>0}\frac{\langle\nabla V(x),f_{0}(x,\theta)\rangle}{\sum_{i=1}^{r}\langle\nabla V(x),f_{i}(x,\theta)\rangle^{2}}<+\infty,

then, as shown in [5], it is possible to construct another SCLF candidate for which the non-increase inequality in (7) is strengthened to a strict decrease:

V(x),f0(x,θ)γi=1rV(x),fi(x,θ)2α(|x|𝒜),\displaystyle\langle\nabla V(x),f_{0}(x,\theta)\rangle-\gamma\sum_{i=1}^{r}\langle\nabla V(x),f_{i}(x,\theta)\rangle^{2}\leq-\alpha(|x|_{\mathcal{A}}),

for some positive definite function α\alpha. However, the new function will involve the expressions V(x),fi(x,θ)\langle\nabla V(x),f_{i}(x,\theta)\rangle, which necessitates the measurement of the unknown parameter θ\theta. Thus, we do not insist upon the strong small control property for VV so as to allow for some generality on how the unknown parameter θ\theta affects the dynamics. \square

Given a function VV that is a SCLF candidate with respect to 𝒜\mathcal{A} for \mathcal{H}, we introduce the following auxiliary HDS:

V:{ξC,ξ˙F¯(ξ)ξD,ξ+G(ξ)\displaystyle\mathcal{H}_{V}:\begin{cases}~{}~{}\xi\in C,&\dot{\xi}\hphantom{{}^{+}}\in\bar{F}(\xi)\\ ~{}~{}\xi\in D,&\xi^{+}\in G(\xi)\end{cases} (8)

wherein the flow map F¯\bar{F} is defined by

F¯(ξ)\displaystyle\bar{F}(\xi) :={f¯(ξ)}×Fe(θ),\displaystyle:=\{\bar{f}(\xi)\}\times F_{e}(\theta),

the map f¯\bar{f} is given by (6c), and the remaining data of V\mathcal{H}_{V} coincide with the data of \mathcal{H}. Note that the HDS V\mathcal{H}_{V} in (8) is obtained by closing the loop for the HDS \mathcal{H} in (5) with the following ideal feedback law:

ui(ξ)=γV(x),fi(ξ).\displaystyle u_{i}(\xi)=-\gamma\langle\nabla V(x),f_{i}(\xi)\rangle. (9)

By construction and inequalities (6), the compact set 𝒜×Θ\mathcal{A}\times\Theta is Lyapunov stable for V\mathcal{H}_{V} [30, Theorem 3.18]. However, implementing the ideal feedback law (9), which is ubiquitous in the literature, requires real-time measurements of the state xx and the control direction θ\theta, as well as complete knowledge of VV and fif_{i}. This level of information, however, is not accessible to the control law in (5), leading to the following observation:

Fact: For the HDS (5) with unknown control directions, the ideal feedback law (9) is not implementable. \square

Even though the SCLF-based control law uiu_{i} is not suitable for implementations, our main goal is to use information about the existence of an SCLF to design an implementable controller able to emulate uiu_{i} and stabilize system (5). However, since SCLF candidates only guarantee stability for the ideal closed-loop system V\mathcal{H}_{V}, additional structure might be needed to obtain asymptotic stability. To capture such structure, we introduce the notion of strong V\nabla V-stabilizability, which is also instrumental to conclude when VV is an actual SCLF (not just a “candidate”) with respect to 𝒜\mathcal{A} for \mathcal{H}.

Definition 3.3.

The HDS \mathcal{H} is said to be strongly V\nabla V-stabilizable if there exists a SCLF candidate VV with respect to 𝒜\mathcal{A} for \mathcal{H} such that 𝒜×Θ\mathcal{A}\times\Theta is UGAS for the ideal closed-loop system V\mathcal{H}_{V}. When \mathcal{H} is strongly V\nabla V-stabilizable, VV is said to be a Strong Control Lyapunov Function (SCLF) with respect to 𝒜\mathcal{A}  for \mathcal{H}.

Remark 3.4.

The topology of the problem under consideration may preclude the existence of a SCLF VV. Indeed, this is the case for stabilization problems defined on smooth compact manifolds, where (robust) global stabilization via continuous feedback is not possible [14]. However, we will show in Section 4 that this issue can be overcome by considering a collection of functions {Vq}q𝒬\{V_{q}\}_{q\in\mathcal{Q}}, indexed by a discrete state qq that acts as a logic mode to be selected in real time by the controller. In this case, the state xx will be decomposed into x=(p,q)x=(p,q), where pp belongs to the smooth manifold and qq belongs to a finite discrete set 𝒬\mathcal{Q}. It will be shown that the conditions of Definition 3.3 are satisfied by the function V(x)=Vq(p)V(x)=V_{q}(p), which can be used for the purpose of real-time control using hybrid feedback. \square

Remark 3.5.

Establishing that the open-loop HDS \mathcal{H} is strongly V\nabla V-stabilizable is an application-dependent task. A sufficient condition is to establish the existence of a positive definite function α\alpha such that the following inequalities hold along the solutions of the ideal closed-loop HDS V\mathcal{H}_{V}:

V˙(ξ)\displaystyle\dot{V}(\xi) α(|πx(ξ)|𝒜),\displaystyle\leq-\alpha(|\pi_{x}(\xi)|_{\mathcal{A}}), ξ\displaystyle\forall\xi C,\displaystyle\in C, (10a)
ΔV(ξ)\displaystyle\Delta V(\xi) α(|πx(ξ)|𝒜),\displaystyle\leq-\alpha(|\pi_{x}(\xi)|_{\mathcal{A}}), ξ\displaystyle\forall\xi D.\displaystyle\in D. (10b)

However, condition (10) place significant restrictions on the unknown parameter θ\theta. Thus, in lieu of (10), we only require that VV is a weak Lyapunov function [30] for V\mathcal{H}_{V} as described by the inequalities (6).Therefore, the weak decrease of VV may be used in conjunction with the properties of the solutions of V\mathcal{H}_{V} to certify that 𝒜×(ΘCΘD)\mathcal{A}\times(\Theta_{C}\cup\Theta_{D}) is UGAS for V\mathcal{H}_{V}. Among others, this allows to model the practical situation wherein temporary loss of control inhibits the ability of feedback to induce strict decrease of VV during flows and/or jumps. We further expand on this case in Section 4. \square

The previous facts and discussion motivate the main problem statement considered in this section:

Problem 1: Under the Assumption that the HDS (5) is strongly V\nabla V-stabilizable, design an implementable feedback law uu that renders UGAS the compact set 𝒜\mathcal{A}.

3.3 Robust Model-Free Stabilization via Oscillatory Hybrid Feedback Control

To solve Problem 1, we propose a class of hybrid model-free feedback controllers that require only real-time measurement or evaluations of the SCLF V(x)V(x). Specifically, for i{1,2,,r}i\in\{1,2,\dots,r\}, we propose the following control law:

uiε(V(x),η)=ε14πγTiκexp(κV(x)S)e1,ηi,\displaystyle u^{\varepsilon}_{i}(V(x),\eta)=\varepsilon^{-1}\sqrt{\frac{4\pi\gamma}{T_{i}\kappa}}\langle\exp(\kappa V(x)S)e_{1},\eta_{i}\rangle, (11a)
where VV is a SCLF with respect to 𝒜\mathcal{A} for \mathcal{H}, the constants κ,γ,ε>0\kappa,\gamma,\varepsilon\in\mathbb{R}_{>0} are tuning parameters, ηi𝕊1\eta_{i}\in\mathbb{S}^{1} for i{1,2,,r}i\in\{1,2,\dots,r\} is the state of the linear oscillator
ηi𝕊1,η˙i\displaystyle\eta_{i}\in\mathbb{S}^{1},~{}~{}\dot{\eta}_{i} =2πTi1ε2Sηi,\displaystyle=2\pi T_{i}^{-1}\varepsilon^{-2}S\eta_{i}, S\displaystyle S =(0110),\displaystyle=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}, (11b)
where 𝕊1={ηi2:ηi,12+ηi,22=1}\mathbb{S}^{1}=\{\eta_{i}\in\mathbb{R}^{2}:\eta_{i,1}^{2}+\eta_{i,2}^{2}=1\} is the unitary circle, and {T1,T2,,Tr}>0\{T_{1},T_{2},\dots,T_{r}\}\subset\mathbb{Q}_{>0} is a collection of constants satisfying TiTjT_{i}\neq T_{j} for all iji\neq j. For brevity, we will use
η˙\displaystyle\dot{\eta} =Λε(η),\displaystyle=\Lambda_{\varepsilon}(\eta), (11c)

to denote the collective (continuous-time) dynamics of the uncoupled oscillators η=(η1,η2,,ηr)𝕊1×𝕊1××𝕊1=:𝕋r\eta=(\eta_{1},\eta_{2},\dots,\eta_{r})\in\mathbb{S}^{1}\times\mathbb{S}^{1}\times\cdots\times\mathbb{S}^{1}=:\mathbb{T}^{r}, where the map Λε\Lambda_{\varepsilon} on the right hand side of (11c) is defined consistently with (11b).

By applying the proposed feedback law (11) to close the loop for \mathcal{H} in (5), the resulting closed-loop HDS has the state y=(ξ,η)y=(\xi,\eta) and dynamics

cl:{yC×𝕋r,y˙F^ε(y):=Fε(ξ,η)×Fe(θ)yD×𝕋r,y+G^(y):=G(ξ)×{η}\displaystyle\mathcal{H}_{cl}:\begin{cases}~{}y\in C\times\mathbb{T}^{r},&\dot{y}\hphantom{{}^{+}}\in\hat{F}_{\varepsilon}(y):=F_{\varepsilon}(\xi,\eta)\times F_{e}(\theta)\\ ~{}y\in D\times\mathbb{T}^{r},&y^{+}\in\hat{G}(y):=G(\xi)\times\{\eta\}\end{cases} (12a)

where, for each ε>0\varepsilon\in\mathbb{R}_{>0}, the set-valued map FεF_{\varepsilon} is given by

Fε(ξ,η)\displaystyle F_{\varepsilon}(\xi,\eta) ={fε(ξ,η))}×{Λε(η)}×Fe(θ),\displaystyle=\{f_{\varepsilon}(\xi,\eta))\}\times\{\Lambda_{\varepsilon}(\eta)\}\times F_{e}(\theta), (12b)
fε(ξ,η)\displaystyle f_{\varepsilon}(\xi,\eta) =f0(x,θ)+i=1rfi(x,θ)uiε(V(x),η).\displaystyle=f_{0}(x,\theta)+\sum_{i=1}^{r}f_{i}(x,\theta)u^{\varepsilon}_{i}(V(x),\eta). (12c)

Using 𝒜¯:=𝒜×Θ×𝕋r\bar{\mathcal{A}}:=\mathcal{A}\times\Theta\times\mathbb{T}^{r}, we can now state our first main result. The proof is presented in Section 5.

Theorem 2.

Let \mathcal{H} in (5) be strongly V\nabla V-stabilizable. Then, 𝒜¯\bar{\mathcal{A}} is SGpAS as ε0+\varepsilon\rightarrow 0^{+} for the closed-loop cl\mathcal{H}_{cl} in (12).

We also have the following corollary, which is of independent interest for problems defined on compact spaces.

Corollary 3.

Let \mathcal{H} in (5) be strongly V\nabla V-stabilizable. If 𝒳\mathcal{X} is compact, then 𝒜¯\bar{\mathcal{A}} is UGpAS as ε0+\varepsilon\rightarrow 0^{+} for the closed-loop HDS cl\mathcal{H}_{cl} in (12).

The proof of Theorem 2, presented in Section 5, exploits the Lie-Bracket averaging theorems for well-posed HDS introduced in [1]. However, in contrast to the results in [1], which consider time-varying HDS, in the present paper we modeled the complete dynamics as time-invariant HDS by using the dynamic oscillators (11c), which evolve in the compact set 𝕋r\mathbb{T}^{r}. Similar oscillators have been considered before for the study of hybrid extremum-seeking control via first-order averaging [28, 31]. However, Theorem 2 is the first result that uses time-invariant oscillators for the analysis of Lie-bracket-based averaging algorithms.

Remark 3.6.

The results of Theorem 2 extend the model-free control laws studied for ODEs in [6], [8, Sec. 6] and, more recently, in [7], to the framework of hybrid systems.

The following important corollary is a consequence of [25, Thm. 7.21] and cl\mathcal{H}_{cl} satisfying the Hybrid Basic Conditions of Assumption 1.

Corollary 4 (Robustness).

Consider the perturbed closed-loop HDS cld\mathcal{H}^{d}_{cl} obtained from (12) as follows:

cld:{y+d1C×𝕋r,y˙F^ε(y+d2)+d3,y+d4D×𝕋r,y+G^(y+d5)+d6,\displaystyle\mathcal{H}^{d^{*}}_{cl}:\begin{cases}~{}~{}y+d_{1}\in C\times\mathbb{T}^{r},&\dot{y}\hphantom{{}^{+}}\in\hat{F}_{\varepsilon}(y+d_{2})+d_{3},\\ ~{}~{}y+d_{4}\in D\times\mathbb{T}^{r},&y^{+}\in\hat{G}(y+d_{5})+d_{6},\end{cases}

where the signals di:dom(y)(CD)×𝕋rd_{i}:\text{dom}(y)\rightarrow(C\cup D)\times\mathbb{T}^{r} are measurable functions satisfying sup(t,j)dom(y)|di(t,j)|d\sup_{(t,j)\in\text{dom}(y)}|d_{i}(t,j)|\leq d^{*}, for all i{1,2,,6}i\in\{1,2,\dots,6\} and some d>0d^{*}\in\mathbb{R}_{>0}. Then, the perturbed HDS cld\mathcal{H}^{d}_{cl} renders the set 𝒜¯\bar{\mathcal{A}} SGpAS as (d,ε)0+(d^{*},\varepsilon)\rightarrow 0^{+}. Moreover, if 𝒳\mathcal{X} is compact, then the set 𝒜¯\bar{\mathcal{A}} is UGpAS as (d,ε)0+(d^{*},\varepsilon)\rightarrow 0^{+} for system cld\mathcal{H}^{d}_{cl}.

In the next section, we present different applications of Theorem 2 in the context of robust global stabilization of a point pp^{\star} on a smooth manifold \mathcal{M} that is not globally contractible, and under unknown switching control directions.

4 Minimum Seeking for Synergistic Potential Functions on Smooth Manifolds

Let \mathcal{M} be a smooth closed manifold properly embedded within, and equipped with the Riemannian metric of, an ambient Euclidean space np\mathbb{R}^{n_{p}}. Let pp\in\mathcal{M} denote the state of a plant with dynamics

p˙\displaystyle\dot{p} =i=1rbi(p)θiui,\displaystyle=\sum_{i=1}^{r}b_{i}(p)\theta_{i}u_{i}, (13)

where u=(u1,u2,,ur)u=(u_{1},u_{2},\dots,u_{r}) are the control inputs. The vector (θ1,θ2,,θr)(\theta_{1},\theta_{2},\dots,\theta_{r}) corresponds to unknown control gains that are allowed to dynamically switch between finitely many values, i.e. (θ1,θ2,,θr)r(\theta_{1},\theta_{2},\dots,\theta_{r})\in\mathcal{E}\subset\mathbb{R}^{r} and ||<+|\mathcal{E}|<+\infty. For example, without loss of generality, we may consider the normalized directions ={+1,0,1}r\mathcal{E}=\{+1,0,-1\}^{r}.

Remark 4.1.

For the purpose of illustration, in (13) we have specialized the form of uncertainty in the control directions to be linear. Nevertheless, we emphasize that general, potentially nonlinear, dependence on θ\theta is allowed as long as Assumptions 2 and 3 hold.

To satisfy Assumption 3, we impose a dwell time condition on the rate of switching of θ\theta. We also require that each control gain θi\theta_{i} does not vanish for an indefinite duration of time, since otherwise uniform (practical) asymptotic stability properties will be precluded. The two requirements are satisfied if the vector (θ1,θ2,,θr)(\theta_{1},\theta_{2},\dots,\theta_{r}) is governed by the HDS (4) with state θ=(θ1,θ2,,θr,θr+1,θr+2)ΘCΘD\theta=(\theta_{1},\theta_{2},\dots,\theta_{r},\theta_{r+1},\theta_{r+2})\in\Theta_{C}\cup\Theta_{D}, and the following data:

ΘC\displaystyle\Theta_{C} :=×[0,1]×[0,T],\displaystyle:=\mathcal{E}\times[0,1]\times[0,T_{\circ}], (14a)
ΘD\displaystyle\Theta_{D} :=×{1}×[0,T],\displaystyle:=\mathcal{E}\times\{1\}\times[0,T_{\circ}], (14b)
Fe(θ)\displaystyle F_{e}(\theta) :={0}×[0,χ1]×([0,χ2]𝕀b(θ)),\displaystyle:=\{0\}\times\left[0,\chi_{1}\right]\times(\left[0,\chi_{2}\right]-\mathbb{I}_{\mathcal{E}_{b}}(\theta)), (14c)
b\displaystyle\mathcal{E}_{b} :={θ|i{1,2,,r} s.t. θi=0},\displaystyle:=\{\theta~{}|~{}\exists i\in\{1,2,\dots,r\}\text{ s.t. }\theta_{i}=0\}, (14d)
G~e(θ)\displaystyle\tilde{G}_{e}(\theta) :=\{(θ1,θ2,,θr)},\displaystyle:=\mathcal{E}\backslash\{(\theta_{1},\theta_{2},\dots,\theta_{r})\}, (14e)
Ge(θ)\displaystyle G_{e}(\theta) :=G~e(θ)×{0}×{θr+2},\displaystyle:=\tilde{G}_{e}(\theta)\times\{0\}\times\{\theta_{r+2}\}, (14f)

for some T>0T_{\circ}\in\mathbb{R}_{>0}, χ1>0\chi_{1}\in\mathbb{R}_{>0}, χ2(0,1)\chi_{2}\in(0,1), and where 𝕀b()\mathbb{I}_{\mathcal{E}_{b}}(\cdot) is the classical indicator function on the subset b\mathcal{E}_{b}.

The construction (14) involves a hybrid automaton, similar to the one considered in [32, Prop. 1.1], with auxiliary state θr+10\theta_{r+1}\in\mathbb{R}_{\geq 0}, and a time-ratio monitor with auxiliary state θr+20\theta_{r+2}\in\mathbb{R}_{\geq 0}, similar to the one considered in [28, Lemma 7]. In particular, every time the condition θr+1=1\theta_{r+1}=1 is satisfied, θr+1\theta_{r+1} is reset to zero, and the vector of directions θ\theta is allowed to jump according to the rule θ+G~e(θ)\theta^{+}\in\tilde{G}_{e}(\theta), while θr+2\theta_{r+2} remains unchanged. The role of θr+2\theta_{r+2} is to model a monitor that decreases continuously whenever the current direction θ\theta has a component equal to zero, i.e., when θb\theta\in\mathcal{E}_{b} and therefore at least one of the entries in (13) has zero controllability. In this case, the condition θr+2=0\theta_{r+2}=0 will eventually occur, forcing θ\theta to jump out of the set b\mathcal{E}_{b} and into the set {1,1}r\{-1,1\}^{r}. Note that, unlike existing hybrid monitors for switching systems with unstable modes [28], the set {1,1}r\{-1,1\}^{r} also contains vectors θ\theta with negative signs, which could be destabilizing for system (13) under a feedback law designed for θ=𝟏r\theta=\mathbf{1}_{r}. In other words, in this paper, the time-ratio monitor prevents the system from spending too much time in modes with zero control directions but not in modes with potentially problematic negative control directions.

By construction, and by [32, Prop. 1.1], and [28, Lemma 7], each hybrid arc generated by the HDS defined by (4) with data (14) satisfies the following dwell-time and activation-time inequalities for any two times t2>t1t_{2}>t_{1} in its domain:

N(t1,t2)\displaystyle N_{\sharp}(t_{1},t_{2}) χ1(t2t1)+1,\displaystyle\leq\chi_{1}(t_{2}-t_{1})+1, (15a)
T(t1,t2)\displaystyle T_{\sharp}(t_{1},t_{2}) χ2(t2t1)+T,\displaystyle\leq\chi_{2}(t_{2}-t_{1})+T_{\circ}, (15b)

where N(t1,t2)N_{\sharp}(t_{1},t_{2}) is the total number of jumps during the time interval (t1,t2)(t_{1},t_{2}), and

T(t1,t2)=t1t2𝕀b(θ(t,j))dt,\displaystyle T_{\sharp}(t_{1},t_{2})=\int_{t_{1}}^{t_{2}}\mathbb{I}_{\mathcal{E}_{b}}(\theta(t,j))\,\text{d}t, (16)

is the total time that the unknown vector θ\theta spends in the set b\mathcal{E}_{b} during the same interval [28].

Next, we impose the following standard regularity condition on the control vector fields bib_{i} of system (13). This condition essentially guarantees that, in the ideal setting when θ=𝟏r\theta=\mathbf{1}_{r}, system (13) can actually be controlled by having vector fields bib_{i} that span the tangent space of the manifold \mathcal{M} everywhere. Such assumptions are standard in the literature related to model-free stabilization and optimization [33, 5]. Similar assumptions have also been used in the design of synergistic hybrid feedback laws that overcome topological obstructions to stability [16].

Assumption 4.

For all i{1,2,,r}i\in\{1,2,\dots,r\} and all pp\in\mathcal{M}, bib_{i} is 𝒞1\mathcal{C}^{1}, bi(p)Tpb_{i}(p)\in T_{p}\mathcal{M}, and there exists a constant λ>0\lambda>0 such that for all all vTpv\in T_{p}\mathcal{M}, it holds that i=1rbi(p),v2λv,v\textstyle\sum_{i=1}^{r}\langle b_{i}(p),v\rangle^{2}\geq\lambda\langle v,v\rangle. \square

Let VV be a 𝒞1\mathcal{C}^{1} function, and consider the subset

Crit(V):={p|V(p)(Tp)},\displaystyle\text{Crit}(V):=\{p\in\mathcal{M}~{}|~{}\nabla V(p)\in(T_{p}\mathcal{M})^{\perp}\}, (17)

where (Tp)np(T_{p}\mathcal{M})^{\perp}\subset\mathbb{R}^{n_{p}} denotes the orthogonal complement of the tangent space TpT_{p}\mathcal{M} with respect to the Riemmannian metric of the ambient Euclidean space np\mathbb{R}^{n_{p}}. With this notation, we recall the following definition, adapted from [19].

Definition 4.2.

For N1N\in\mathbb{N}_{\geq 1} and 𝒬={1,2,,N}\mathcal{Q}=\{1,2,\dots,N\}, a family of 𝒞1\mathcal{C}^{1} functions {Vq}q𝒬\{V_{q}\}_{q\in\mathcal{Q}}, Vq:0V_{q}:\mathcal{M}\rightarrow\mathbb{R}_{\geq 0}, is said to be a δ\delta-gap synergistic family of potential functions with respect to pp^{\star} if, for all q𝒬q\in\mathcal{Q}, the following holds:

  1. 1.

    VqV_{q} is positive definite with respect to pp^{\star};

  2. 2.

    c0\forall c\in\mathbb{R}_{\geq 0}, {p|Vq(p)c}\{p\in\mathcal{M}~{}|~{}V_{q}(p)\leq c\} is compact;

  3. 3.

    There exists an open neighborhood 𝒰qnp\mathcal{U}_{q}\subset\mathbb{R}^{n_{p}} of pp^{\star} such that 𝒰qCrit(Vq)=p\mathcal{U}_{q}\cap\text{Crit}(V_{q})=p^{\star};

  4. 4.

    There exists δ>0\delta\in\mathbb{R}_{>0} such that δ<Δ\delta<\Delta^{\star}, where Δ\Delta^{\star} is

    Δ=minq𝒬pCrit(Vq)Vq(p)maxq~𝒬Vq~(p).\displaystyle\Delta^{\star}=\min_{\begin{subarray}{c}q\in\mathcal{Q}\\ p\in\text{Crit}(V_{q})\end{subarray}}V_{q}(p)-\max_{\tilde{q}\in\mathcal{Q}}V_{\tilde{q}}(p). (18)

The construction of synergistic families of potential functions for stability problems typically requires qualitative information on the underlying manifold \mathcal{M} and the target point pp^{\star}\in\mathcal{M}. We make the assumption that such functions are available to us, and later we show how to satisfy this assumption in different applications.

Assumption 5.

The family of functions {Vq}q𝒬\{V_{q}\}_{q\in\mathcal{Q}}, 𝒬={1,2,,N}\mathcal{Q}=\{1,2,\dots,N\}, is a δ\delta-gap synergistic family of potential functions with respect to pp^{\star}.

To exploit the existence of synergistic potential functions, we introduce a logic state q𝒬q\in\mathcal{Q} into our controller, and we let x=(p,q)×𝒬x=(p,q)\in\mathcal{M}\times\mathcal{Q}, and define

𝒳C\displaystyle\mathcal{X}_{C} :={x×𝒬|μ(x)δ},\displaystyle:=\{x\in\mathcal{M}\times\mathcal{Q}~{}|~{}\mu(x)\leq\delta\}, (19a)
𝒳D\displaystyle\mathcal{X}_{D} :={x×𝒬|μ(x)δ},\displaystyle:=\{x\in\mathcal{M}\times\mathcal{Q}~{}|~{}\mu(x)\geq\delta\}, (19b)
G𝒳(x)\displaystyle G_{\mathcal{X}}(x) :={p}×{argminq~𝒬V(p,q~)},\displaystyle:=\{p\}\times\big{\{}\arg\min_{\tilde{q}\in\mathcal{Q}}V(p,\tilde{q})\big{\}}, (19c)
where the function μ:×𝒬0\mu:\mathcal{M}\times\mathcal{Q}\rightarrow\mathbb{R}_{\geq 0} is defined by
μ(x):=Vq(p)minq~𝒬Vq~(p).\displaystyle\mu(x):=V_{q}(p)-\textstyle\min_{\tilde{q}\in\mathcal{Q}}V_{\tilde{q}}(p). (19d)
Also, let f0f_{0} and fif_{i} for i{1,2,,r}i\in\{1,2,\dots,r\} be given by
f0(x,θ)\displaystyle f_{0}(x,\theta) =0,\displaystyle=0, fi(x,θ)\displaystyle f_{i}(x,\theta) :=θi(bi(p)0).\displaystyle:=\theta_{i}\begin{pmatrix}b_{i}(p)\\ 0\end{pmatrix}. (19e)

where the second entry in fif_{i} models the continuous-time dynamics of the logic state qq, which remains constant during flows, i.e., q˙=0\dot{q}=0. Finally, define the function V:×𝒬0V:\mathcal{M}\times\mathcal{Q}\rightarrow\mathbb{R}_{\geq 0} as follows:

V(x)=Vq(p).V(x)=V_{q}(p). (20)

Then, we have the following proposition that applies to any system of the form (13) that satisfies the previous assumptions. The proof is presented in Section 5.

Proposition 5.

Suppose that Assumptions 4 and 5 are satisfied, and VV is given by (20). Then, \mathcal{H} is strongly V\nabla V-stabilizable. \square

As a consequence of Theorem 2 and Proposition 5, we are able to conclude that, under the feedback law (11), the subset 𝒜×Θ×𝕋r\mathcal{A}\times\Theta\times\mathbb{T}^{r} is SGpAS as ε0+\varepsilon\rightarrow 0^{+} for the closed loop HDS cl\mathcal{H}_{cl} (12).

Refer to caption
Figure 1: Global stabilization of p=(0,1)p^{\star}=(0,1) on 𝕊1\mathbb{S}^{1} under unknown switching control directions. Left: trajectories of (21) under the proposed model-free feedback law (red color), and the non-hybrid feedback law [6] (blue color). Right: Evolution in time of θ1\theta_{1}.

Next, we present several novel concrete applications of the results in this section along with numerical simulation results. Henceforth, we take ={+1,0,1}r\mathcal{E}=\{+1,0,-1\}^{r}. Since the construction of synergistic families of potential functions is not the main contribution of our manuscript, we rely on existing results in the literature [19]. Instead, we focus on the role of the adaptive feedback law (11) and its ability to “emulate” the ideal control law (9) in the context of hybrid control using highly oscillatory feedback.

4.1 Robust Global Stabilization on 𝕊1\mathbb{S}^{1}

As the first example, we take =𝕊1\mathcal{M}=\mathbb{S}^{1}, r=1r=1, and we consider the control-affine system

p˙\displaystyle\dot{p} =b1(p)θ1u1,\displaystyle=b_{1}(p)\theta_{1}u_{1}, b1(p)\displaystyle b_{1}(p) =Sp,\displaystyle=Sp, (21)

where p𝕊1p\in\mathbb{S}^{1}, SS is the matrix defined in (11), θ1={+1,0,1}\theta_{1}\in\mathcal{E}=\{+1,0,-1\} is the unknown control gain, and u1u_{1} is the control input. The goal is to globally stabilize a point ϑ𝕊1\vartheta^{\star}\in\mathbb{S}^{1}. This problem was solved in [23] and [34] under the assumption of having constant and known control directions θ1:=1\theta_{1}:=1.

To globally stabilize ϑ𝕊1\vartheta^{\star}\in\mathbb{S}^{1} under unknown control directions, we consider the synergistic family of potential functions {W1,W2}\{W_{1},W_{2}\} defined by

Wq(p)\displaystyle W_{q}(p) :=WΦq(p),\displaystyle:=W\circ\Phi_{q}(p), W(p)\displaystyle W(p) :=1ϑ,p,\displaystyle:=1-\langle\vartheta^{\star},p\rangle, (22a)
where q𝒬={1,2}q\in\mathcal{Q}=\{1,2\}, and the maps Φq:𝕊1𝕊1\Phi_{q}:\mathbb{S}^{1}\rightarrow\mathbb{S}^{1} are
Φq(p):=exp((3/2q)W(p)S)p.\displaystyle\Phi_{q}(p):=\exp\left((3/2-q)W(p)S\right)p. (22b)

As shown in [23, 19], the family of functions {W1,W2}\{W_{1},W_{2}\} is a δ\delta-gap synergistic family of potential functions function with respect to ϑ\vartheta^{\star} for any δ(0,1)\delta\in(0,1). It follows from Proposition 5 that the function VV defined by V(p,q)=Wq(p)V(p,q)=W_{q}(p) is an SCLF with respect to 𝒜={p}×𝒬\mathcal{A}=\{p^{\star}\}\times\mathcal{Q} for \mathcal{H}, and therefore that 𝒜×Θ×𝕊1\mathcal{A}\times\Theta\times\mathbb{S}^{1} is SGpAS for cl\mathcal{H}_{cl}. Due to the compactness of 𝕊1\mathbb{S}^{1}, we can invoke Corollary 3 to conclude that 𝒜×Θ×𝕊1\mathcal{A}\times\Theta\times\mathbb{S}^{1} is UGpAS.

Numerical simulations results illustrating the performance of the proposed controller for this example are shown in Figure 1. To generate the results, we used γ=1\gamma=\sqrt{1}, κ=4\kappa=4, ε=1/4π0.28\varepsilon=1/\sqrt{4\pi}\approx 0.28. We also used δ=1/4\delta=1/4 for the synergistic family of potential functions {W1,W2}\{W_{1},W_{2}\}. The target point is p=(0,1)p^{\star}=(0,1). The right plot shows the evolution in time of the control direction θ\theta, which vanishes during bounded (but persistent) periods of time. Finally, to emphasize the robustness of the proposed feedback law, we added a small adversarial perturbation that locally stabilizes (in the absence of switching) the problematic critical point p=(0,1)p^{\sharp}=(0,-1). As shown in the figure, the proposed control law is not affected by the perturbation whereas a non-hybrid model-free feedback law is effectively trapped by the adversarial perturbation in the vicinity of the critical point pp^{\sharp}.

Refer to caption
Figure 2: Target-seeking with obstacle avoidance under unknown control gains. Left: trajectories of the vehicle under the proposed hybrid model-free feedback law (shown in red), and under the vanilla synergistic hybrid feedback (shown in blue) [16]. Right: Evolution in time of the control gain θ1\theta_{1}. The gain θ2=1\theta_{2}=1 is taken to be constant.

4.2 Robust Target Seeking with Obstacle Avoidance

Consider the problem of stabilizing a target position for a mobile vehicle moving in an obstructed planar domain. Let 𝒪2\mathcal{O}\subset\mathbb{R}^{2} denote the obstacle and let z=(z1,z2)2\𝒪z=(z_{1},z_{2})\in\mathbb{R}^{2}\backslash\mathcal{O} denote the position of the vehicle. The goal of the vehicle is to reach a target position z=(z1,z2)2\𝒪z^{\star}=(z_{1}^{\star},z_{2}^{\star})\in\mathbb{R}^{2}\backslash\mathcal{O} while avoiding the obstacle 𝒪\mathcal{O}. We assume that the motion of the vehicle is governed by the kinematic equations

z˙\displaystyle\dot{z} =i=12eiθiui,\displaystyle=\sum_{i=1}^{2}e_{i}\,\theta_{i}\,u_{i}, (23)

where u=(u1,u2)2u=(u_{1},u_{2})\in\mathbb{R}^{2} are the control input, and (θ1,θ2)={+1,0,1}r(\theta_{1},\theta_{2})\in\mathcal{E}=\{+1,0,-1\}^{r} are the unknown control gains. Since the obstacle 𝒪\mathcal{O} is bounded, there exists z𝒪2z_{\mathcal{O}}\in\mathbb{R}^{2} and d0>0d_{0}>0 such that 𝒪z𝒪+d𝔹\mathcal{O}\subset z_{\mathcal{O}}+d\,\mathbb{B} for all dd0d\geq d_{0}. To guarantee feasibility, we impose the following assumption on the target point zz^{\star}.

Assumption 6.

d>d0\exists~{}d^{\star}>d_{0} such that z2\(z𝒪+d𝔹)z^{\star}\in\mathbb{R}^{2}\backslash(z_{\mathcal{O}}+d^{\star}\,\mathbb{B}).

To avoid the obstacle, and following the ideas of [16], we consider the map φ:2\(z𝒪+d0𝔹))×𝕊1\varphi:\mathbb{R}^{2}\backslash(z_{\mathcal{O}}+d_{0}\,\mathbb{B}))\rightarrow\mathbb{R}\times\mathbb{S}^{1} given by

φ(z)\displaystyle\varphi(z) =(log(|zz𝒪|d),(zz𝒪)/|zz𝒪|),\displaystyle=(\log(|z-z_{\mathcal{O}}|-d^{\star}),(z-z_{\mathcal{O}})/|z-z_{\mathcal{O}}|), (24)

As shown in [16], φ\varphi is a well-defined diffeomorphism. The pushforward [35] of the kinematics of the vehicle under the diffeomorphism φ\varphi are given by

p˙\displaystyle\dot{p} =i=12bi(p)θiui,\displaystyle=\sum_{i=1}^{2}b_{i}(p)\theta_{i}u_{i}, bi(p)\displaystyle b_{i}(p) =Dφφ1(p)ei,\displaystyle=\text{D}\varphi\circ\varphi^{-1}(p)e_{i}, (25)

where p=(ρ,ϑ)×𝕊1p=(\rho,\vartheta)\in\mathbb{R}\times\mathbb{S}^{1}. Therefore, global stabilization of the target position zz^{\star} in 2\(p𝒪+d0𝔹))\mathbb{R}^{2}\backslash(p_{\mathcal{O}}+d_{0}\,\mathbb{B})) is equivalent to globally stabilizing the point p=(ρ,ϑ)p^{\star}=(\rho^{\star},\vartheta^{\star}) on the smooth manifold =×𝕊1\mathcal{M}=\mathbb{R}\times\mathbb{S}^{1}. However, the topology of \mathcal{M} prohibits global stabilization by continuous feedback since 𝕊1\mathbb{S}^{1} is a compact boundary-less manifold. With that in mind, we introduce the family of functions

Vq(p)\displaystyle V_{q}(p) :=12(ρρ)2+(eρeρ)2+11+Wq(ϑ),\displaystyle:=\frac{1}{2}(\rho-\rho^{\star})^{2}+\sqrt{(\text{e}^{\rho}-\text{e}^{\rho^{\star}})^{2}+1}-1+W_{q}(\vartheta),

where q𝒬={1,2}q\in\mathcal{Q}=\{1,2\}, and WqW_{q} are the functions defined in (22). As in the previous subsection, the family of functions {Wq}q𝒬\{W_{q}\}_{q\in\mathcal{Q}} is a δ\delta-gap synergistic Lyapunov function for ϑ\vartheta^{\star} on 𝕊1\mathbb{S}^{1} for any δ(0,1)\delta\in(0,1), which follows by [19]. Therefore, it is straightforward to show that the family of functions {V1,V2}\{V_{1},V_{2}\} is a δ\delta-gap synergistic family of potential functions with respect to pp^{\star}. Thus, from Proposition 5, the function VV defined by V(p,q)=Vq(p)V(p,q)=V_{q}(p) is an SCLF with respect to 𝒜={p}×𝒬\mathcal{A}=\{p^{\star}\}\times\mathcal{Q} for \mathcal{H}. By invoking Theorem 2, we conclude that 𝒜×Θ×𝕋2\mathcal{A}\times\Theta\times\mathbb{T}^{2} is SGpAS for the original hybrid system with unknown and dynamic control directions cl\mathcal{H}_{cl}.

To demonstrate the performance of the proposed controller compared to existing synergistic hybrid feedback controllers [16], we present numerical simulations in Figure 5. To generate the results, we used γ=2\gamma=2, δ=1/4\delta=1/4, κ=4\kappa=4, and ε=1/6π0.165\varepsilon=1/\sqrt{6\pi}\approx 0.165. The target position is z=(0,2)z^{\star}=(0,2), and an obstacle with radius d=1d=1 is centered at the origin, i.e. z𝒪=(0,0)z_{\mathcal{O}}=(0,0). Due to the special structure of the control vector fields in this example, we are able to use a single oscillator η1\eta_{1} with period T1=1T_{1}=1 and rely on a π/2\pi/2 phase shift to guarantee non-resonance between the inputs u1u_{1} and u2u_{2}. Figure 5 clearly indicates that the vanilla synergistic hybrid feedback fails to reach the target. This is to be expected since traditional synergistic hybrid controllers require knowledge of the control direction. By contrast, the proposed algorithm is able to reach the target despite the persistent switching in the control gain sign and the temporary loss of control.

Refer to caption
Figure 3: Global stabilization of p=(0,0,1)p^{\star}=(0,0,1) on 𝕊2\mathbb{S}^{2} under unknown persistently switching control directions. Left: trajectory corresponding to the proposed hybrid model-free feedback law (11) (shown in red), and the non-hybrid, model-free feedback law [6, 36] (shown in blue). Right: Evolution in time of the control gains (θ1,θ2,θ3)(\theta_{1},\theta_{2},\theta_{3}).

4.3 Robust Global Stabilization on 𝕊2\mathbb{S}^{2}

For the third example, we take =𝕊2\mathcal{M}=\mathbb{S}^{2}, r=3r=3, and we consider the control-affine system

p˙\displaystyle\dot{p} =i=13bi(p)θiui,\displaystyle=\sum_{i=1}^{3}b_{i}(p)\theta_{i}u_{i}, bi(p)\displaystyle b_{i}(p) =eip,eip,\displaystyle=e_{i}-\langle p,e_{i}\rangle p, (26)

where p𝕊2p\in\mathbb{S}^{2}, (θ1,θ2,θ3)={+1,0,1}3(\theta_{1},\theta_{2},\theta_{3})\in\mathcal{E}=\{+1,0,-1\}^{3} are the unknown control gains, and u=(u1,u2,u3)3u=(u_{1},u_{2},u_{3})\in\mathbb{R}^{3} are the control inputs. The goal is to globally stabilize an arbitrary point p𝕊2p^{\star}\in\mathbb{S}^{2}. To that end, we introduce the synergistic family of potential functions

Vq(p)\displaystyle V_{q}(p) :=WΦq(p),\displaystyle:=W\circ\Phi_{q}(p), W(p)\displaystyle W(p) :=1p,p,\displaystyle:=1-\langle p,p^{\star}\rangle, (27)

where q{1,2}q\in\{1,2\}, the maps Φq:𝕊2𝕊2\Phi_{q}:\mathbb{S}^{2}\rightarrow\mathbb{S}^{2} are defined by

Φq(p):=exp((3/2q)W(p)[p]×)p,\displaystyle\Phi_{q}(p):=\text{exp}\left((3/2-q)W(p)[p^{\star}_{\perp}]_{\times}\right)p, (28)
Refer to caption
Figure 4: Global stabilization of p=vec(I)p^{\star}=vec(I) on SO(3)\text{SO}(3) under unknown switching control directions (shown in the bottom plot). The red trajectory corresponds to the proposed model-free feedback law (11), whereas the blue trajectory corresponds to a non-hybrid model-free feedback law [6]. Both systems are subject to the same small bounded persistent disturbance, which disrupts the blue trajectory.

and p𝕊2p^{\star}_{\perp}\in\mathbb{S}^{2} is such that p,p=0\langle p^{\star}_{\perp},p^{\star}\rangle=0. As shown in [19], the family of functions {V1,V2}\{V_{1},V_{2}\} is a δ\delta-gap synergistic family of potential functions function with respect to pp^{\star} for any δ(0,1)\delta\in(0,1). It follows from Proposition 5 that the function VV defined by V(p,q)=Vq(p)V(p,q)=V_{q}(p) is an SCLF with respect to 𝒜={p}×𝒬\mathcal{A}=\{p^{\star}\}\times\mathcal{Q} for \mathcal{H}, and therefore that 𝒜×Θ×𝕊2\mathcal{A}\times\Theta\times\mathbb{S}^{2} is SGpAS for cl\mathcal{H}_{cl}. By invoking Corollary 3, we conclude that 𝒜×Θ×𝕋3\mathcal{A}\times\Theta\times\mathbb{T}^{3} is UGpAS.

Numerical simulations results are shown in Figure 3. To generate the results, we used γ=1\gamma=\sqrt{1}, κ=4\kappa=4, T1=3T_{1}=3, T2=2T_{2}=2, T3=1T_{3}=1, δ=1/5\delta=1/5, and ε=1/8π\varepsilon=1/\sqrt{8\pi}. The target point is p=(0,0,1)p^{\star}=(0,0,1), and p=(0,1,0)p_{\perp}^{\star}=(0,1,0). Finally, to emphasize the robustness of the proposed feedback law compared to the standard non-hybrid model-free controllers [33], we added a small adversarial perturbation that locally stabilizes the problematic critical point p=(0,0,1)p^{\sharp}=(0,0,-1) in the absence of switching. As shown in Figure 3, the proposed control law is not affected by the perturbation whereas a vanilla model-free feedback law is trapped by the adversarial perturbation in the vicinity of the critical point pp^{\sharp}.

4.4 Robust Global Stabilization on SO(3)\text{SO}(3)

Next, we consider the problem of globally stabilizing a desired attitude RSO(3)R^{\star}\in\text{SO}(3) for a rigid body. The kinematics of the rigid body are given by

R˙\displaystyle\dot{R} =i=13Re^iθiui,\displaystyle=\sum_{i=1}^{3}R\,\widehat{e}_{i}\,\theta_{i}\,u_{i}, (29)

where RSO(3)3×3R\in\text{SO}(3)\subset\mathbb{R}^{3\times 3} is the rotation matrix representing the attitude of the rigid body, u=(u1,u2,u3)3u=(u_{1},u_{2},u_{3})\in\mathbb{R}^{3} are the control inputs, and θi\theta_{i} are the unknown control gains. Since SO(3)\text{SO}(3) is a Lie Group, the problem of stabilizing any specific attitude RSO(3)R^{\star}\in\text{SO}(3) is equivalent to stabilizing the identity element, i.e. ISO(3)I\in\text{SO}(3). Therefore, without loss of generality, we only consider the case when R=IR^{\star}=I. We remark that, herein, we consider SO(3)\text{SO}(3) as an embedded submanifold of 3×3\mathbb{R}^{3\times 3} equipped with its Riemannian metric, i.e. the Frobenius norm.

Following [19], we introduce the family of functions

V~q(R)\displaystyle\tilde{V}_{q}(R) :=WΦq(R),\displaystyle:=W\circ\Phi_{q}(R), W(R)\displaystyle W(R) :=tr(A(IR)),\displaystyle:=\text{tr}(A(I-R)), (30)

where q𝒬={1,2}q\in\mathcal{Q}=\{1,2\}, AA is the matrix given by

A=3i=13ω~,eii=13ω~,eiei,\displaystyle A=\frac{3}{\sum_{i=1}^{3}\langle\tilde{\omega},e_{i}\rangle}\sum_{i=1}^{3}\langle\tilde{\omega},e_{i}\rangle e_{i}, (31)

ω~=(11,12,13)\tilde{\omega}=(11,12,13), the maps Φq:SO(3)SO(3)\Phi_{q}:\text{SO}(3)\rightarrow\text{SO}(3) are

Φq(R):=exp((32q)4W(R)[ω]×)R,\displaystyle\Phi_{q}(R):=\exp\left(\frac{(3-2q)}{4}W(R)[\omega]_{\times}\right)R, (32)

and ω=ω~/ω~𝕊2\omega=\tilde{\omega}/\lVert\tilde{\omega}\lVert\in\mathbb{S}^{2}. As shown in [19], the family of functions {V~1,V~2}\{\tilde{V}_{1},\tilde{V}_{2}\} is a δ\delta-gap synergistic family of potential functions with respect to II for any δ(0,1/2)\delta\in(0,1/2).

Next, let \mathcal{M} be the Euclidean submanifold

={vec(R)|RSO(3)}9.\displaystyle\mathcal{M}=\{{vec}(R)~{}|~{}R\in\text{SO}(3)\}\subset\mathbb{R}^{9}. (33)

By defining p=vec(R)p={vec}(R)\in\mathcal{M}, it follows that pp evolves according to the driftless control-affine system

p˙\displaystyle\dot{p} =i=13bi(p)θiui,\displaystyle=\sum_{i=1}^{3}b_{i}(p)\,\theta_{i}\,u_{i}, bi(p)\displaystyle b_{i}(p) =(e^iI)p\displaystyle=-(\widehat{e}_{i}\otimes I)p (34)

Since the map vec:SO(3){vec}:\text{SO}(3)\rightarrow\mathcal{M} is a diffeomorphism, the family of functions {V~1,V~2}\{\tilde{V}_{1},\tilde{V}_{2}\} can be pulled back (via the inverse of vecvec) to a δ\delta-gap synergistic family of potential functions with respect to p=vec(I)p^{\star}={vec}(I). More explicitly, the family of functions {V1,V2}\{V_{1},V_{2}\} defined by Vq(p)=V~qvec1(p)V_{q}(p)=\tilde{V}_{q}\circ vec^{-1}(p) is a δ\delta-gap synergistic family of potential functions with respect to p=vec(I)p^{\star}={vec}(I). Therefore, it follows from Proposition 5 that the function VV defined by V(p,q)=Vq(p)V(p,q)=V_{q}(p) is an SCLF with respect to 𝒜={p}×𝒬\mathcal{A}=\{p^{\star}\}\times\mathcal{Q} for the HDS \mathcal{H}, and therefore that 𝒜×Θ×𝕋3\mathcal{A}\times\Theta\times\mathbb{T}^{3} is SGpAS for cl\mathcal{H}_{cl}.

We now provide numerical simulations. The results of the simulations are shown in Figure 4. To generate the results, we used γ=1\gamma=\sqrt{1}, κ=4\kappa=4, T1=1T_{1}=1, T2=2T_{2}=2, T3=3T_{3}=3, δ=1/5\delta=1/5, and ε=1/12π\varepsilon=1/\sqrt{12\pi}. The target point is p=vec(I)=(1,0,0,0,1,0,0,0,1)p^{\star}=vec(I)=(1,0,0,0,1,0,0,0,1). To emphasize the robustness of the proposed feedback law compared to the non-hybrid model-free controllers [33], we added a small adversarial perturbation that locally stabilizes the (bad) critical point p=(1,0,0,0,1,0,0,0,1)p^{\sharp}=(-1,0,0,0,1,0,0,0,-1) in the absence of switching. As shown in Figure 4, the proposed control law is not affected by the perturbation whereas the vanilla model-free feedback law is trapped by the adversarial perturbation in the vicinity of the critical point pp^{\sharp}.

4.5 Robust Target Seeking with Obstacle Avoidance for a Nonholonomic Vehicle

Finally, we consider again the problem of robust global stabilization of a target position for a mobile vehicle moving in an obstructed planar domain. However, in contrast to the single integrator dynamics considered in Section 4.2, we now consider a nonholonomic vehicle model. Due to the nonholonomic nature of the vehicle, the control vector fields do not span the entire tangent space of the state space manifold, and therefore Proposition 5 is not applicable. Although we are still able to apply the model-free feedback law (11), the stability analysis is different from the examples considered hitherto. Thus, we treat this special case separately.

Let 𝒪2\mathcal{O}\subset\mathbb{R}^{2} denote the obstacle and z=(z1,z2)2\𝒪z=(z_{1},z_{2})\in\mathbb{R}^{2}\backslash\mathcal{O} denote the position of the vehicle. Suppose the vehicle is governed by the nonholonomic kinematic equations

z˙\displaystyle\dot{z} =θ1u1ψ,\displaystyle=\theta_{1}u_{1}\psi, ψ˙\displaystyle\dot{\psi} =u2Sψ\displaystyle=u_{2}S\psi (35)

where u=(u1,u2)2u=(u_{1},u_{2})\in\mathbb{R}^{2} are the control inputs, θ1\theta_{1}\in\mathcal{E}\subset\mathbb{R} is the unknown control gain, SS is the matrix defined in (11b), and ψ𝕊1\psi\in\mathbb{S}^{1} is a unit vector that indicates the forward direction of motion, i.e. the orientation, of the vehicle. This kinematic model has also been considered in [37, 15, 38, 39, 36]. The goal of the vehicle is to stabilize a known target position z=(z1,z2)2\𝒪z^{\star}=(z_{1}^{\star},z_{2}^{\star})\in\mathbb{R}^{2}\backslash\mathcal{O}, irrespective of the orientation ψ𝕊1\psi\in\mathbb{S}^{1}, while avoiding the obstacle 𝒪\mathcal{O}. We impose Assumption 6 on the position of the target point zz^{\star} relative to the obstacle.

Refer to caption
Figure 5: Simulation results for the problem of target-stabilization with obstacle avoidance for the nonholonomic kinematics (36).

Under the diffeomorphism defined by (24), the pushforward of the nonholonomic kinematics of the vehicle is given by

p˙\displaystyle\dot{p} =θ1u1Dφφ1(p)ψ,\displaystyle=\theta_{1}u_{1}\text{D}\varphi\circ\varphi^{-1}(p)\psi, ψ˙\displaystyle\dot{\psi} =u2Sψ,\displaystyle=u_{2}S\psi, (36)

where p=(ρ,ϑ)×𝕊1p=(\rho,\vartheta)\in\mathbb{R}\times\mathbb{S}^{1}. Therefore, the goal of globally stabilizing the subset {z}×𝕊1\{z^{\star}\}\times\mathbb{S}^{1} in (2\(z𝒪+d0𝔹)))×𝕊1\big{(}\mathbb{R}^{2}\backslash(z_{\mathcal{O}}+d_{0}\,\mathbb{B}))\big{)}\times\mathbb{S}^{1} is equivalent to globally stabilizing the compact subset {p}×𝕊1\{p^{\star}\}\times\mathbb{S}^{1} on the smooth manifold φ(2\(z𝒪+d0𝔹)))×𝕊1=×𝕊1\varphi\big{(}\mathbb{R}^{2}\backslash(z_{\mathcal{O}}+d_{0}\,\mathbb{B}))\big{)}\times\mathbb{S}^{1}=\mathcal{M}\times\mathbb{S}^{1}, where p=(ρ,ϑ)=φ(z)p^{\star}=(\rho^{\star},\vartheta^{\star})=\varphi(z^{\star}). In contrast to the single integrator kinematics considered in subsection 4.2, Proposition 5 is not applicable for a vehicle with the nonholonomic kinematics (36) since the control vector fields in (36) clearly do not span the entire tangent space of the manifold \mathcal{M} and so Assumption 4 is violated. Nevertheless, we will show that the structure of the control-affine system (36) permits the use of a slightly different model-free feedback law that stabilizes \mathcal{H}. To that end, we introduce the family of functions

Vq(μ)\displaystyle V_{q}(\mu) :=12(ρρ)2+(eρeρ)2+11+Wq(ϑ),\displaystyle:=\frac{1}{2}(\rho-\rho^{\star})^{2}+\sqrt{(\text{e}^{\rho}-\text{e}^{\rho^{\star}})^{2}+1}-1+W_{q}(\vartheta),

where q𝒬={1,2}q\in\mathcal{Q}=\{1,2\}, and WqW_{q} are the functions defined in (22). As in Section 4.2, the family of functions {V1,V2}\{V_{1},V_{2}\} is a δ\delta-gap synergistic family of potential functions with respect to μ=(ρ,ϑ)\mu^{\star}=(\rho^{\star},\vartheta^{\star}) on ×𝕊1\mathbb{R}\times\mathbb{S}^{1}. Let

μ:=(ρ,ϑ)×𝕊1,\mu:=(\rho,\vartheta)\in\mathbb{R}\times\mathbb{S}^{1},

p=(μ,ψ)×𝕊1×𝕊1=p=(\mu,\psi)\in\mathbb{R}\times\mathbb{S}^{1}\times\mathbb{S}^{1}=\mathcal{M}, x=(p,q)×𝒬x=(p,q)\in\mathcal{M}\times\mathcal{Q}, and define the function V:×𝒬0V:\mathcal{M}\times\mathcal{Q}\rightarrow\mathbb{R}_{\geq 0} as follows:

V(x)=Vq(μ).V(x)=V_{q}(\mu). (37)

Also, let r=1r=1, T1=1T_{1}=1, η=η1𝕊1\eta=\eta_{1}\in\mathbb{S}^{1}, and let the first control input u1u_{1} be given by the feedback law (11), i.e. we take u1u_{1} to be

u1ε(V(x),η)=ε14πγκexp(κV(x)S)e1,η1.\displaystyle u^{\varepsilon}_{1}(V(x),\eta)=\varepsilon^{-1}\sqrt{\frac{4\pi\gamma}{\kappa}}\langle\exp(\kappa V(x)S)e_{1},\eta_{1}\rangle. (38a)
Next, let the second control input u2u_{2} to be defined as:
u2ε=2πε1.\displaystyle u^{\varepsilon}_{2}=2\pi\varepsilon^{-1}. (38b)

Finally, define the compact sets 𝒜={μ}×𝕊1×𝒬\mathcal{A}=\{\mu^{\star}\}\times\mathbb{S}^{1}\times\mathcal{Q}, and 𝒜¯=𝒜×Θ\bar{\mathcal{A}}=\mathcal{A}\times\Theta. Then, we have the following proposition proved in Section 5.

Proposition 6.

𝒜¯×𝕊1\bar{\mathcal{A}}\times\mathbb{S}^{1} is SGpAS as ε0+\varepsilon\rightarrow 0^{+} for cl\mathcal{H}_{cl} with VV given by (37).

We conclude this section with a numerical simulation result illustrating the control of the non-holonomic vehicle (35). The resulting trajectories of the vehicle are shown in Figure 5. To generate the results, we used γ=2\gamma=2, δ=1/4\delta=1/4, κ=4\kappa=4, and ε=1/6π0.165\varepsilon=1/\sqrt{6\pi}\approx 0.165. The target position is z=(0,2)z^{\star}=(0,2), and we considered an obstacle with radius d=1d=1, centered at the origin, i.e. z𝒪=(0,0)z_{\mathcal{O}}=(0,0). The system is simulated under an additive adversarial perturbation designed to trap the trajectories of the vehicle behind the obstacle. The red trajectory corresponds to the proposed hybrid model-free feedback law, whereas the blue trajectory corresponds to a non-hybrid, model-free feedback law [6, 36]. The arrows indicate the direction of motion for the corresponding trajectory. The figure on the bottom depicts the magnitude of the adversarial perturbation which effectively traps the vanilla model-free feedback behind the obstacle. The figure on the top right depicts the control gain θ1\theta_{1} as a function of time. The figure on the top left depicts the obstructed planar domain wherein the vehicle operates.

5 Proofs

In this section, we present the proofs of our main results.

5.1 Proof of Theorem 2

Let τ0\tau\in\mathbb{R}_{\geq 0} and consider the HDS ~cl\widetilde{\mathcal{H}}_{cl} with state (ξ,η,τ)n×r×0(\xi,\eta,\tau)\in\mathbb{R}^{n}\times\mathbb{R}^{r}\times\mathbb{R}_{\geq 0} and dynamics:

{C×𝕋r×0,(ξ˙η˙τ˙)Fε(ξ,η)×{Λε(η)}×{ε2}D×𝕋r×0,(ξ+η+τ+)G(ξ)×{η}×{τ}\displaystyle\begin{cases}C\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0},&\begin{pmatrix}\dot{\xi}\\ \dot{\eta}\\ \dot{\tau}\end{pmatrix}\in F_{\varepsilon}(\xi,\eta)\times\{\Lambda_{\varepsilon}(\eta)\}\times\{\varepsilon^{-2}\}\\ D\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0},&\begin{pmatrix}\xi^{+}\\ \eta^{+}\\ \tau^{+}\end{pmatrix}\in G(\xi)\times\{\eta\}\times\{\tau\}\end{cases}

where FεF_{\varepsilon} is given by (12b), Λε(η)\Lambda_{\varepsilon}(\eta) is given by (11c), GG is given by (5e), and C,DC,D are given by (5b). System ~cl\widetilde{\mathcal{H}}_{cl} is a trivial dynamic extension of the closed-loop HDS cl\mathcal{H}_{cl}. Therefore, given an initial condition, any solution of the HDS cl\mathcal{H}_{cl} corresponds to some solution of the HDS ~cl\widetilde{\mathcal{H}}_{cl}. Let ζ:=(ζ1,ζ2,,ζr)𝕊1×𝕊1×𝕊1=𝕋r\zeta:=(\zeta_{1},\zeta_{2},\dots,\zeta_{r})\in\mathbb{S}^{1}\times\mathbb{S}^{1}\times\cdots\mathbb{S}^{1}=\mathbb{T}^{r} and consider the HDS ^cl\widehat{\mathcal{H}}_{cl} defined by

{C×𝕋r×0,(ξ˙ζ˙τ˙)F~ε(ξ,ζ,τ)×{0}×{ε2}D×𝕋r×0,(ξ+ζ+τ+)G(ξ)×{ζ}×{τ}\displaystyle\begin{cases}C\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0},&\begin{pmatrix}\dot{\xi}\\ \dot{\zeta}\\ \dot{\tau}\end{pmatrix}\in\tilde{F}_{\varepsilon}(\xi,\zeta,\tau)\times\{0\}\times\{\varepsilon^{-2}\}\\ D\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0},&\begin{pmatrix}\xi^{+}\\ \zeta^{+}\\ \tau^{+}\end{pmatrix}\in G(\xi)\times\{\zeta\}\times\{\tau\}\end{cases} (39)

where F~ε\tilde{F}_{\varepsilon} is defined by

F~ε(ξ,ζ,τ):=Fε(ξ,exp(Ωτ)ζ),\displaystyle\tilde{F}_{\varepsilon}(\xi,\zeta,\tau):=F_{\varepsilon}(\xi,\exp(\Omega\tau)\zeta),

and the matrix Ω\Omega is a block diagonal matrix of size 2r×2r2r\times 2r with rr diagonal blocks of size 2×22\times 2 such that the ii-th block is the matrix 2πTi1S2\pi T_{i}^{-1}S, where SS is the matrix defined in (11). Let (ξ,η,τ):dom(ξ,η,τ)n×𝕋r×0(\xi,\eta,\tau):\text{dom}(\xi,\eta,\tau)\rightarrow\mathbb{R}^{n}\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0} be any solution of ~cl\widetilde{\mathcal{H}}_{cl} and define the hybrid arc (ξ,ζ,τ):dom(ξ,η,τ)n×𝕋r×0(\xi,\zeta,\tau):\text{dom}(\xi,\eta,\tau)\rightarrow\mathbb{R}^{n}\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0} by

(ξ,ζ,τ):=(ξ,exp(Ωτ)η,τ).\displaystyle(\xi,\zeta,\tau):=(\xi,\exp(-\Omega\tau)\eta,\tau). (40)

Direct computation shows that

ζ˙\displaystyle\dot{\zeta} =exp(Ωτ)η˙τ˙exp(Ωτ)Ωη=0.\displaystyle=\exp(-\Omega\tau)\dot{\eta}-\dot{\tau}\exp(-\Omega\tau)\Omega\eta=0. (41)

Since, for any τ0\tau\in\mathbb{R}_{\geq 0} and any η\eta in 𝕋r\mathbb{T}^{r}, exp(Ωτ)η𝕋r\exp(-\Omega\tau)\eta\in\mathbb{T}^{r}, it follows that the hybrid arc (ξ,ζ,τ)(\xi,\zeta,\tau) defined by (41) is a solution of ^cl\widehat{\mathcal{H}}_{cl} with the same initial condition for ξ\xi and τ\tau, and the following initial condition for ζ\zeta:

ζ(0,0)=exp(Ωτ(0,0))η(0,0).\displaystyle\zeta(0,0)=\exp(-\Omega\tau(0,0))\eta(0,0). (42)

Hence, for any solution of the HDS ~cl\widetilde{\mathcal{H}}_{cl} there exists some solution of the HDS ^cl\widehat{\mathcal{H}}_{cl} such that (41) holds. In particular, the ξ\xi-component of both solutions coincide. We will use this equivalence to establish that every solution of ~cl\widetilde{\mathcal{H}}_{cl} satisfies suitable stability bounds by first establishing such bounds for every solution of the HDS ^cl\widehat{\mathcal{H}}_{cl}.

An explicit computation of F~ε\tilde{F}_{\varepsilon} shows that

F~ε(ξ,ζ,τ)={f~ε(ξ,ζ,τ)}×Fe(θ),\displaystyle\tilde{F}_{\varepsilon}(\xi,\zeta,\tau)=\{\tilde{f}_{\varepsilon}(\xi,\zeta,\tau)\}\times F_{e}(\theta), (43)

where FeF_{e} is given by (4), and

f~ε(ξ,ζ,τ)=f0(x,θ)+i=1rfi(x,θ)u^i(V(x),ζi,τ),\displaystyle\tilde{f}_{\varepsilon}(\xi,\zeta,\tau)=f_{0}(x,\theta)+\sum_{i=1}^{r}f_{i}(x,\theta)\hat{u}_{i}(V(x),\zeta_{i},\tau), (44)

and the functions u^i\hat{u}_{i} are given by

u^i(V,ζi,τ)\displaystyle\hat{u}_{i}(V,\zeta_{i},\tau) =ε14πγTiκexp(κVS)e1,exp(2πSτ/Ti)ζi.\displaystyle=\varepsilon^{-1}\sqrt{\frac{4\pi\gamma}{T_{i}\kappa}}\langle\exp(\kappa VS)e_{1},\exp(2\pi S\tau/T_{i})\zeta_{i}\rangle.

Now, since the constants {T1,T2,,Tr}\{T_{1},T_{2},\dots,T_{r}\} are rational, it follows that there exists T>0T\in\mathbb{R}_{>0} such that the map f~ε\tilde{f}_{\varepsilon} is TT-periodic in τ\tau. Moreover, by definition of the map f~ε\tilde{f}_{\varepsilon}, the HDS ^cl\widehat{\mathcal{H}}_{cl} belongs to the class of well-posed HDS with highly oscillatory flow maps which was analyzed in [1]. By applying the Lie-bracket averaging results in [1], we obtain that the Hybrid Lie-bracket averaged system ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}} corresponding to the HDS ^cl\widehat{\mathcal{H}}_{cl} is given by

^clave:{C×𝕋r,(ξ˙,ζ˙)F¯(ξ)×{0}D×𝕋r,(ξ,ζ)+G(ξ)×{ζ},\displaystyle\widehat{\mathcal{H}}_{cl}^{\text{ave}}:\begin{cases}~{}C\times\mathbb{T}^{r},&(\dot{\xi},\dot{\zeta})\hphantom{{}^{+}}\in\bar{F}(\xi)\times\{0\}\\ ~{}D\times\mathbb{T}^{r},&(\xi,\zeta)^{+}\in G(\xi)\times\{\zeta\}\end{cases}, (45)

wherein the flow map F¯\bar{F} is defined by

F¯(ξ)\displaystyle\bar{F}(\xi) :={f¯(x,θ)}×Fe(θ),\displaystyle:=\{\bar{f}(x,\theta)\}\times F_{e}(\theta),
f¯(ξ)\displaystyle\bar{f}(\xi) :=f0(x,θ)γi=1rV(x),fi(x,θ)fi(x,θ),\displaystyle:=f_{0}(x,\theta)-\gamma{\sum_{i=1}^{r}}\langle\nabla V(x),f_{i}(x,\theta)\rangle f_{i}(x,\theta),

which is independent of ζ\zeta. Thus, the HDS ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}} is nothing but a trivial dynamic extension of the HDS V\mathcal{H}_{V} obtained by adding ζ\zeta as a state with trivial flow and jump dynamics, i.e., ζ˙=0\dot{\zeta}=0 and ζ+=ζ\zeta^{+}=\zeta. From the assumptions of the theorem, namely that \mathcal{H} is strongly V\nabla V-stabilizable, it follows that the subset 𝒜×(ΘCΘD)\mathcal{A}\times(\Theta_{C}\cup\Theta_{D}) is UGAS for the HDS V\mathcal{H}_{V} defined in (8). It follows that the compact subset 𝒜¯\bar{\mathcal{A}} is UGAS for the HDS ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}} given by (45). From [1, Theorem 2], we obtain that there exists a class 𝒦\mathcal{KL} function β\beta such that for each compact subset K(CD)×𝕋rK\subset(C\cup D)\times\mathbb{T}^{r} and for each ν>0\nu>0, there exists ε>0\varepsilon^{*}>0 such that for all ε(0,ε]\varepsilon\in(0,\varepsilon^{*}] and for all solutions of ^cl\widehat{\mathcal{H}}_{cl} with (ξ(0,0),ζ(0,0))K(\xi(0,0),\zeta(0,0))\in K, the following inequality holds for all (t,j)dom(x,ζ,τ)(t,j)\in\text{dom}(x,\zeta,\tau):

|(ξ(t,j),ζ(t,j))|𝒜¯β(|(ξ(0,0),ζ(0,0))|𝒜¯,t+j)+ν.\displaystyle|(\xi(t,j),\zeta(t,j))|_{\bar{\mathcal{A}}}\leq\beta(|(\xi(0,0),\zeta(0,0))|_{\bar{\mathcal{A}}},t+j)+\nu.

With ε(0,ε]\varepsilon\in(0,\varepsilon^{*}], let (ξ,η,τ):dom(ξ,η,τ)n×𝕋r×0(\xi,\eta,\tau):\text{dom}(\xi,\eta,\tau)\rightarrow\mathbb{R}^{n}\times\mathbb{T}^{r}\times\mathbb{R}_{\geq 0} be any solution of the HDS ~cl\widetilde{\mathcal{H}}_{cl} such that (ξ(0,0),η(0,0))K(\xi(0,0),\eta(0,0))\in K. By construction, it follows that (ξ,η,τ)(\xi,\eta,\tau) satisfies

(ξ,η,τ)=(ξ,exp(Ωτ)ζ,τ),\displaystyle(\xi,\eta,\tau)=(\xi,\exp(\Omega\tau)\zeta,\tau), (46)

for some solution (ξ,ζ,τ)(\xi,\zeta,\tau) of the HDS ^cl\widehat{\mathcal{H}}_{cl} with (ξ,ζ)(0,0)K(\xi,\zeta)(0,0)\in K. Therefore, we obtain that there exists a class 𝒦\mathcal{KL} function β\beta such that for each compact subset K(CD)×𝕋rK\subset(C\cup D)\times\mathbb{T}^{r} and for each ν>0\nu>0, there exists ε>0\varepsilon^{*}>0 such that for all ε(0,ε]\varepsilon\in(0,\varepsilon^{*}] and for all solutions of ~cl\widetilde{\mathcal{H}}_{cl} with (ξ(0,0),η(0,0))K(\xi(0,0),\eta(0,0))\in K, the following inequality holds for all (t,j)dom(x,η,τ)(t,j)\in\text{dom}(x,\eta,\tau):

|(ξ(t,j),η(t,j))|𝒜¯β(|(ξ(0,0),η(0,0))|𝒜¯,t+j)+ν.\displaystyle|(\xi(t,j),\eta(t,j))|_{\bar{\mathcal{A}}}\leq\beta(|(\xi(0,0),\eta(0,0))|_{\bar{\mathcal{A}}},t+j)+\nu.

However, since ~cl\widetilde{\mathcal{H}}_{cl} is nothing but a trivial dynamic extension of the closed loop system cl\mathcal{H}_{cl}, it follows that the 𝒦\mathcal{KL} bound above is also true for every solution of cl\mathcal{H}_{cl} with (ξ(0,0),η(0,0))K(\xi(0,0),\eta(0,0))\in K. In particular, we may always take K=K1×𝕋rK=K_{1}\times\mathbb{T}^{r} for some compact K1(CD)K_{1}\subset(C\cup D) so that any initial condition η(0,0)\eta(0,0) for the state of the oscillators is admissible. This concludes the proof.

5.2 Proof of Proposition 5

When f0f_{0} and fif_{i} are as given in (19), direct computations show that f¯\bar{f}, defined in (6c), takes the form

f¯(ξ)=γi=1rθi2Vq(p),bi(p)(bi(p)0).\displaystyle\bar{f}(\xi)=-\gamma\sum_{i=1}^{r}\theta_{i}^{2}\langle\nabla V_{q}(p),b_{i}(p)\rangle\begin{pmatrix}b_{i}(p)\\ 0\end{pmatrix}.

As a result, if ξ=(x,θ)=((p,q),θ)\xi=(x,\theta)=((p,q),\theta), we have that

V˙(ξ)γi=1rθi2Vq(p),bi(p)20,\displaystyle\dot{V}(\xi)\leq-\gamma\sum_{i=1}^{r}\theta_{i}^{2}\langle\nabla V_{q}(p),b_{i}(p)\rangle^{2}\leq 0,

for all ξC\xi\in C. On the other hand, by construction we have that for all ξD\xi\in D:

ΔV(ξ)0,\displaystyle\Delta V(\xi)\leq 0,

Thus, VV is an SCLF candidate with respect to 𝒜\mathcal{A} for the HDS \mathcal{H} defined by (3)-(5), (14), and (19).

Since \mathcal{M} is equipped with the Riemannian metric of the ambient space and bi(p)Tpb_{i}(p)\in T_{p}\mathcal{M}, we have that

Vq(p),bi(p)=Vq(p),bi(p),\displaystyle\langle\nabla V_{q}(p),b_{i}(p)\rangle=\langle\nabla_{\mathcal{M}}V_{q}(p),b_{i}(p)\rangle,

where Vq(p)\nabla_{\mathcal{M}}V_{q}(p) is the unique orthogonal projection of Vq(p)\nabla V_{q}(p) onto the tangent space TpT_{p}\mathcal{M} with respect to the Riemannian metric of the ambient Euclidean space. Therefore, we obtain that

V˙(ξ)=γi=1rθi2Vq(p),bi(p)2.\displaystyle\dot{V}(\xi)=-\gamma\sum_{i=1}^{r}\theta_{i}^{2}\langle\nabla_{\mathcal{M}}V_{q}(p),b_{i}(p)\rangle^{2}.

On the other hand, by definition, if θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, then θi2>0\theta_{i}^{2}>0, i{1,2,,r}\forall i\in\{1,2,\dots,r\}. Since \mathcal{E} is a discrete finite set, there exists a constant λ1>0\lambda_{1}\in\mathbb{R}_{>0} such that, for all θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we have θi2>λ1\theta_{i}^{2}>\lambda_{1} i{1,2,,r}\forall i\in\{1,2,\dots,r\}.

Hence, for all ξC\xi\in C such that θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we have V˙(ξ)γλ1i=1rVq(p),bi(p)2\dot{V}(\xi)\leq-\gamma\lambda_{1}\sum_{i=1}^{r}\langle\nabla_{\mathcal{M}}V_{q}(p),b_{i}(p)\rangle^{2}. From Assumption 4, there exists a constant λ>0\lambda\in\mathbb{R}_{>0} such that, for all pp\in\mathcal{M} and all vTpv\in T_{p}\mathcal{M}, we have that i=1rv,bi(p)2λv,v\sum_{i=1}^{r}\langle v,b_{i}(p)\rangle^{2}\geq\lambda\langle v,v\rangle. Hence, for all ξC\xi\in C such that θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we have

V˙(ξ)γλ1λVq(p),Vq(p).\displaystyle\dot{V}(\xi)\leq-\gamma\lambda_{1}\lambda\langle\nabla_{\mathcal{M}}V_{q}(p),\nabla_{\mathcal{M}}V_{q}(p)\rangle.

By construction of the family {Vq}q𝒬\{V_{q}\}_{q\in\mathcal{Q}}, we have that

Vq(p)=0,p=p,\displaystyle\nabla_{\mathcal{M}}V_{q}(p)=0,\iff p=p^{\star},

for all (p,q)𝒳C×𝒬(p,q)\in\mathcal{X}_{C}\times\mathcal{Q}. It follows that there exists a positive definite function ρ\rho such that

Vq(p),Vq(p)ρ(|(p,q)|𝒜),\displaystyle\langle\nabla_{\mathcal{M}}V_{q}(p),\nabla_{\mathcal{M}}V_{q}(p)\rangle\geq\rho(|(p,q)|_{\mathcal{A}}),

for all (p,q)𝒳C(p,q)\in\mathcal{X}_{C}, which implies that for all ξC\xi\in C such that θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we have that

V˙(ξ)γλ1λρ(|x|𝒜).\displaystyle\dot{V}(\xi)\leq-\gamma\lambda_{1}\lambda\rho(|x|_{\mathcal{A}}).

Following similar steps as in [40], it can be shown that there exists a continuously differentiable 𝒦\mathcal{K}_{\infty} function α¯\bar{\alpha} such that the function V^=α¯V\hat{V}=\bar{\alpha}\circ V satisfies

α¯1(|x|𝒜)\displaystyle\bar{\alpha}_{1}(|x|_{\mathcal{A}})\leq V^(x)α¯2(|x|𝒜),\displaystyle\hat{V}(x)\leq\bar{\alpha}_{2}(|x|_{\mathcal{A}}), x\displaystyle\forall x 𝒳,\displaystyle\in\mathcal{X},
V^˙(ξ)0,\displaystyle\dot{\hat{V}}(\xi)\leq 0, ξ\displaystyle\forall\xi C,\displaystyle\in C,
Δ\displaystyle\Delta V^(ξ)0,\displaystyle\hat{V}(\xi)\leq 0, ξ\displaystyle\forall\xi D.\displaystyle\in D.

and, for all ξC\xi\in C such that θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we have

V^˙(ξ)V^(ξ).\displaystyle\dot{\hat{V}}(\xi)\leq-\hat{V}(\xi).

We now introduce the function V~(ξ)=V^(ξ)eθr+2\tilde{V}(\xi)=\hat{V}(\xi)\text{e}^{\theta_{r+2}}. Direct computations shows that, during flows, we have

V~˙(ξ)\displaystyle\dot{\tilde{V}}(\xi) =(V^˙(ξ)+θ˙r+2V^(ξ))eθr+2\displaystyle=(\dot{\hat{V}}(\xi)+\dot{\theta}_{r+2}\hat{V}(\xi))\text{e}^{\theta_{r+2}}

From (14), we have that θ˙r+2[0,χ2]𝕀b(θ)\dot{\theta}_{r+2}\in[0,\chi_{2}]-\mathbb{I}_{\mathcal{E}_{b}}(\theta). Therefore, if θb\theta\in\mathcal{E}_{b}, we have that

θ˙r+2(1χ2)V~˙(ξ)(1χ2)V~(ξ).\displaystyle\dot{\theta}_{r+2}\leq-(1-\chi_{2})\implies\dot{\tilde{V}}(\xi)\leq-(1-\chi_{2})\tilde{V}(\xi).

On the other hand, if θ(ΘCΘD)\b\theta\in(\Theta_{C}\cup\Theta_{D})\backslash\mathcal{E}_{b}, we also have that

θ˙r+2χ2V~˙(ξ)(1χ2)V~(ξ)\displaystyle\dot{\theta}_{r+2}\leq\chi_{2}\implies\dot{\tilde{V}}(\xi)\leq-(1-\chi_{2})\tilde{V}(\xi)

From (14), we have that (θr+2)+=θr+2(\theta_{r+2})^{+}=\theta_{r+2}, for all ξD\xi\in D. Thus, using the facts that θr+2[0,T]\theta_{r+2}\in[0,T_{\circ}] and |x|𝒜=|ξ|𝒜~|x|_{\mathcal{A}}=|\xi|_{\tilde{\mathcal{A}}} for all ξCD\xi\in C\cup D, we obtain

α¯1(|ξ|𝒜~)\displaystyle\bar{\alpha}_{1}(|\xi|_{\tilde{\mathcal{A}}})\leq V~(ξ)α¯2(|ξ|𝒜~)eT,\displaystyle\tilde{V}(\xi)\leq\bar{\alpha}_{2}(|\xi|_{\tilde{\mathcal{A}}})\text{e}^{T_{\circ}}, ξ\displaystyle\forall\xi CD,\displaystyle\in C\cup D, (47a)
V~˙(ξ)\displaystyle\dot{\tilde{V}}(\xi)\leq (1χ2)V~(ξ),\displaystyle-(1-\chi_{2})\tilde{V}(\xi), ξ\displaystyle\forall\xi C,\displaystyle\in C, (47b)
Δ\displaystyle\Delta V~(ξ)0,\displaystyle\tilde{V}(\xi)\leq 0, ξ\displaystyle\forall\xi D,\displaystyle\in D, (47c)

where 𝒜~=𝒜×(ΘCΘD)\tilde{\mathcal{A}}=\mathcal{A}\times(\Theta_{C}\cup\Theta_{D}). Because the set 𝒜~\tilde{\mathcal{A}} is Lyapunov stable for V\mathcal{H}_{V}, which is entailed by the conditions (47) [20, Theorem 3.18], it follows that any maximal solution to V\mathcal{H}_{V} starting in CDC\cup D is bounded. Thanks to Lemma 1, we obtain that, by invoking [20, Proposition 2.34], no nontrivial maximal solution to V\mathcal{H}_{V} starting in CDC\cup D is Zeno and that, for every maximal solution ξ\xi to V\mathcal{H}_{V} starting in CDC\cup D, there exists t>0t_{\circ}>0 such that t¯jt¯jt>0\overline{t}_{j}-\underline{t}_{j}\geq t_{\circ}>0, for all j1j\geq 1, where t¯j:=sup{t0|(t,j)dom(ξ)}\overline{t}_{j}:=\sup\{t\in\mathbb{R}_{\geq 0}~{}|~{}(t,j)\in\text{dom}(\xi)\}, and t¯j:=inf{t0|(t,j)dom(ξ)}\underline{t}_{j}:=\,\inf\,\{t\in\mathbb{R}_{\geq 0}~{}|~{}(t,j)\in\text{dom}(\xi)\}. In other words, no solution of V\mathcal{H}_{V} is discrete. By invoking [20, Theorem 3.19], we obtain that the compact subset 𝒜~\tilde{\mathcal{A}} is globally asymptotically stable in the sense of [20, Definition 3.1]. However, since 𝒜~\tilde{\mathcal{A}} is compact and V\mathcal{H}_{V} is a well-posed HDS, it follows from [20, Theorem 3.22] that the global asymptotic stability of 𝒜~\tilde{\mathcal{A}} is uniform, i.e. that 𝒜~\tilde{\mathcal{A}} is UGAS for V\mathcal{H}_{V} in the sense of Definition 2 which concludes the proof.

5.3 Proof of Proposition 6

The closed-loop HDS cl\mathcal{H}_{cl} is defined by (5) in conjunction with (4), (11), (12), (14), (19), (36), and (38). Let τ1,τ20\tau_{1},\tau_{2}\in\mathbb{R}_{\geq 0}, τ=(τ1,τ2)02=0×0\tau=(\tau_{1},\tau_{2})\in\mathbb{R}_{\geq 0}^{2}=\mathbb{R}_{\geq 0}\times\mathbb{R}_{\geq 0}, and define the HDS ~cl\widetilde{\mathcal{H}}_{cl} via the following data:

{C~×02,(ξ˙η˙τ˙)Fε(ξ,η)×{Λε(η)}×{(ε1,ε2)}D~×02,(ξ+η+τ+)G(ξ)×{η}×{τ}\displaystyle\begin{cases}~{}\tilde{C}\times\mathbb{R}^{2}_{\geq 0},&\begin{pmatrix}\dot{\xi}\\ \dot{\eta}\\ \dot{\tau}\end{pmatrix}\in F_{\varepsilon}(\xi,\eta)\times\{\Lambda_{\varepsilon}(\eta)\}\times\left\{(\varepsilon^{-1},\varepsilon^{-2})\right\}\\ ~{}\tilde{D}\times\mathbb{R}_{\geq 0}^{2},&\begin{pmatrix}\xi^{+}\\ \eta^{+}\\ \tau^{+}\end{pmatrix}\in G(\xi)\times\{\eta\}\times\{\tau\}\end{cases}

where C~=C×𝕋1\tilde{C}=C\times\mathbb{T}^{1}, D~=D×𝕋1\tilde{D}=D\times\mathbb{T}^{1}. It is clear that ~cl\widetilde{\mathcal{H}}_{cl} is a trivial dynamic extension of cl\mathcal{H}_{cl}. Next, similar to the proof of Theorem 2, we introduce the HDS ^cl\widehat{\mathcal{H}}_{cl} defined by

{C~×02(μ˙ζ˙τ˙)F^ε(μ,ζ,τ)×{0}×{(ε1,ε2)}D~×02(μ+ζ+τ+)G(μ)×{ζ}×{τ}\displaystyle\begin{cases}\tilde{C}\times\mathbb{R}^{2}_{\geq 0}&\begin{pmatrix}\dot{\mu}\\ \dot{\zeta}\\ \dot{\tau}\end{pmatrix}\in\widehat{F}_{\varepsilon}(\mu,\zeta,\tau)\times\{0\}\times\left\{(\varepsilon^{-1},\varepsilon^{-2})\right\}\\ \tilde{D}\times\mathbb{R}_{\geq 0}^{2}&\begin{pmatrix}\mu^{+}\\ \zeta^{+}\\ \tau^{+}\end{pmatrix}\in G(\mu)\times\{\zeta\}\times\{\tau\}\end{cases}

where ζ𝕊1\zeta\in\mathbb{S}^{1}, μ=(p,ϕ,q,θ)×𝕊1×𝒬×(ΘCΘD)\mu=(p,\phi,q,\theta)\in\mathcal{M}\times\mathbb{S}^{1}\times\mathcal{Q}\times(\Theta_{C}\cup\Theta_{D}), and the map F^ε\widehat{F}_{\varepsilon} is given by

F^ε(μ,ζ,τ)=Fε((p,exp(2πτ1S)ϕ,q,θ),exp(2πτ2S)η).\displaystyle\widehat{F}_{\varepsilon}(\mu,\zeta,\tau)=F_{\varepsilon}((p,\exp(2\pi\tau_{1}S)\phi,q,\theta),\exp(2\pi\tau_{2}S)\eta).

An explicit computation shows that F^ε(μ,ζ,τ)={f^ε(μ,ζ,τ)}×Fe(θ)\widehat{F}_{\varepsilon}(\mu,\zeta,\tau)=\{\widehat{f}_{\varepsilon}(\mu,\zeta,\tau)\}\times F_{e}(\theta), where the map f^ε\widehat{f}_{\varepsilon} is defined by

f^ε(μ,ζ,τ)=θ1b1(p,ϕ,τ1)u^1ε(V(x),ζ,τ2),\displaystyle\widehat{f}_{\varepsilon}(\mu,\zeta,\tau)=\theta_{1}b_{1}(p,\phi,\tau_{1})\hat{u}_{1}^{\varepsilon}(V(x),\zeta,\tau_{2}), (48)

where the vector field b1b_{1} is given by

f1(p,ϕ,τ1):=(Dφφ1(p)exp(2πτ1S)ϕ00),\displaystyle f_{1}(p,\phi,\tau_{1}):=\begin{pmatrix}\text{D}\varphi\circ\varphi^{-1}(p)\exp(2\pi\tau_{1}S)\phi\\ 0\\ 0\end{pmatrix}, (49)

and the function u^1ε\hat{u}_{1}^{\varepsilon} is given by

u^1ε(V,ζ,τ2)\displaystyle\hat{u}_{1}^{\varepsilon}(V,\zeta,\tau_{2}) =ε12γκexp(κVS)e1,exp(2πSτ2)ζ.\displaystyle=\varepsilon^{-1}\sqrt{\frac{2\gamma}{\kappa}}\langle\exp(\kappa VS)e_{1},\exp(2\pi S\tau_{2})\zeta\rangle.

It is clear that f^ε\widehat{f}_{\varepsilon} is 11-periodic in τ1\tau_{1} and τ2\tau_{2}. Moreover, the structure of the HDS ^cl\widehat{\mathcal{H}}_{cl} is the same as the structure of the class of well-posed HDS with highly oscillatory flow maps considered in [1]. By applying the recursive Lie-bracket averaging formulas in [1], it can be shown that the Hybrid Lie-bracket averaged system ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}} corresponding to ^cl\widehat{\mathcal{H}}_{cl} is the system defined by

{(μ,ζ)C~,(μ˙,ζ˙)F¯(μ)×{0}(μ,ζ)D~,(μ,ζ)+G(μ)×{ζ},\displaystyle\begin{cases}(\mu,\zeta)\in\tilde{C},&(\dot{\mu},\dot{\zeta})\hphantom{{}^{+}}\in\bar{F}(\mu)\times\{0\}\\ (\mu,\zeta)\in\tilde{D},&(\mu,\zeta)^{+}\in G(\mu)\times\{\zeta\},\end{cases}

where the map F¯(μ,ζ)={f¯(μ)}×Fe(θ)\bar{F}(\mu,\zeta)=\{\bar{f}(\mu)\}\times F_{e}(\theta),

f¯(μ)\displaystyle\bar{f}(\mu) =θ12γ2i=12bi(p),Vq(p)(bi(p)00),\displaystyle=-\frac{\theta_{1}^{2}\gamma}{2}\sum_{i=1}^{2}\langle b_{i}(p),\nabla V_{q}(p)\rangle\begin{pmatrix}b_{i}(p)\\ 0\\ 0\end{pmatrix}, (50)

and the maps bib_{i} are given by bi(p)=Dφφ1(p)eib_{i}(p)=\text{D}\varphi\circ\varphi^{-1}(p)e_{i}. It is clear that ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}} is a trivial dynamic extension of the HDS

V:{μC,μ˙F¯(μ)μD,μ+G(μ)\displaystyle\mathcal{H}_{V}:\begin{cases}~{}\mu\in C,&~{}\dot{\mu}\hphantom{{}^{+}}\in\bar{F}(\mu)\\ ~{}\mu\in D,&~{}\mu^{+}\in G(\mu)\end{cases}

Now, consider the function V(μ)=Vq(p)V(\mu)=V_{q}(p) and observe that

V˙(μ)=θ12γ2i=12bi(p),Vq(p)20,\displaystyle\dot{V}(\mu)=-\frac{\theta_{1}^{2}\gamma}{2}\sum_{i=1}^{2}\langle b_{i}(p),\nabla V_{q}(p)\rangle^{2}\leq 0, (51)

for all μD\mu\in D, and ΔV(μ)0\Delta V(\mu)\leq 0, for all μD\mu\in D. In addition, VV is positive definite with respect to 𝒜¯=𝒜×(ΘCΘD)\bar{\mathcal{A}}=\mathcal{A}\times(\Theta_{C}\cup\Theta_{D}) and, for every c0c\in\mathbb{R}_{\geq 0}, the set {μCD|V(μ)c}\{\mu\in C\cup D~{}|~{}V(\mu)\leq c\} is compact. Following similar steps to the proof of Proposition 5, it can be shown that 𝒜¯\bar{\mathcal{A}} UGAS for the HDS V\mathcal{H}_{V}, which implies that 𝒜¯×𝕊1\bar{\mathcal{A}}\times\mathbb{S}^{1} is UGAS for the HDS ^clave\widehat{\mathcal{H}}_{cl}^{\text{ave}}. Then, following similar steps as in the proof of Theorem 2, we obtain that 𝒜¯×𝕊1\bar{\mathcal{A}}\times\mathbb{S}^{1} is SGpAS for cl\mathcal{H}_{cl} as ε0+\varepsilon\rightarrow 0^{+}, which concludes the proof.

6 Conclusions

We studied the problem of robust global stabilization for a class of control-affine systems under dynamic uncertainty in the control directions and topological obstructions. By proposing a novel class of hybrid and highly-oscillatory feedback laws that seek the minimum of a family of synergistic Lyapunov functions, we provide a robust solution to this problem. Our method is particularly advantageous as it is model-free and can handle unknown and switching control gain signs, ensuring resilience in autonomous systems facing adversarial spoofing attacks. The practical relevance of our results is demonstrated through several concrete applications. Since our results allow to transfer stability properties from well-posed stable hybrid systems with known control directions to systems with unknown control directions, they open the door for the systematic development of new “model-free” controllers that can exploit the rich set of tools developed during the last two decades for hybrid systems. Numerical simulations validate the effectiveness and robustness of the proposed feedback law compared to existing techniques. Future research directions will focus on experimental validations of the proposed algorithms.

References

  • [1] M. Abdelgalil, J. I. Poveda, On Lie-bracket averaging for a class of hybrid dynamical systems with applications to model-free control and optimization, arXiv preprint arXiv:2308.15732 (2023).
  • [2] F. Pasqualetti, F. Dörfler, F. Bullo, Attack detection and identification in cyber-physical systems, IEEE Trans. on Automatic Control 58 (11) (2013) 2715–2729.
  • [3] A. Morse, Open problems in the theory of parameter adaptive control, in: 1978 IEEE Conference on Decision and Control including the 17th Symposium on Adaptive Processes, IEEE, 1979, pp. 1386–1387.
  • [4] R. D. Nussbaum, Some remarks on a conjecture in parameter adaptive control, Systems & control letters 3 (5) (1983) 243–246.
  • [5] A. Scheinker, M. Krstic, Minimum-seeking for CLFs: Universal semiglobally stabilizing feedback under unknown control directions, IEEE Trans. on Automatic Control 58 (2013) 1107–1122.
  • [6] A. Scheinker, M. Krstić, Model-free stabilization by extremum seeking, Springer, 2017.
  • [7] S. Wang, M. Guay, R. D. Braatz, Extremum-seeking regulator for a class of nonlinear systems with unknown control direction, IEEE Transactions on Automatic Control (2024).
  • [8] A. Scheinker, 100 years of extremum seeking: A survey, Automatica 161 (2024) 111481.
  • [9] H. Durr, M. S. Stankovic, C. Ebenbauer, K. H. Johansson, Lie bracket approximation of extremum seeking systems, Automatica 49 (6) (2013) 1538–1552.
  • [10] V. Grushkovskaya, A. Zuyev, C. Ebenbauer, On a class of generating vector fields for the extremum seeking problem: Lie bracket approximation and stability properties, Automatica 94 (2018) 151–160.
  • [11] R. Suttner, S. Dashkovskiy, Exponential stability for extremum seeking control systems, IFAC-PapersOnLine 50 (1) (2017) 15464–15470.
  • [12] R. Suttner, Extremum-seeking control for a class of mechanical systems, IEEE Transactions on Automatic Control 68 (2) (2022) 1200–1207.
  • [13] E. D. Sontag, Nonlinear Analysis, Differential Equations and Control, Kluwer Academic Publishers, 1999, Ch. Stability and Stabilization: discontinuities and the effect of disturbances.
  • [14] W. Jongeneel, E. Moulay, Topological Obstructions to Stability and Stabilization: History, Recent Advances and Open Problems, Springer Nature, 2023.
  • [15] J. I. Poveda, M. Benosman, A. R. Teel, R. G. Sanfelice, Coordinated Hybrid Source Seeking with Robust Obstacle Avoidance in Multi-vehicle Autonomous Systems, IEEE Transactions on Automatic Control Vol. 67 (No. 2) (2022) pp. 706–721.
  • [16] P. Casau, R. Cunha, R. G., C. Silvestre, Hybrid Control for Robust and Global Tracking on Smooth Manifolds, IEEE Transactions on Automatic Control 65 (2020) 1870–1885.
  • [17] C. G. Mayhew, R. G. Sanfelice, J. Sheng, M. Arcak, A. R. Teel, Quaternion-Based Hybrid Feedback for Robust Global Attitude Synchronization, IEEE Transactions on Automatic Control 57 (8) (2012) 2122–2127.
  • [18] C. G. Mayhew, R. G. Sanfelice, A. R. Teel, Synergistic lyapunov functions and backstepping hybrid feedbacks, in: Proceedings of the 2011 American control conference, IEEE, 2011, pp. 3203–3208.
  • [19] C. G. Mayhew, Hybrid Control for Topologically Constrained Systems, Ph.D Dissertation, University of California, Santa Barbara, 2010.
  • [20] R. Sanfelice, Hybrid Feedback Control, Princeton, 2021.
  • [21] P. Casau, R. G. Sanfelice, C. Silvestre, Robust synergistic hybrid feedback, IEEE Transactions on Automatic Control (2024).
  • [22] C. G. Mayhew, A. R. Teel, Synergistic Hybrid Feedback for Global Rigid-Body Attitude Tracking on SO(3), IEEE Transactions on Automatic and Control 58 (2013) 2730–2742.
  • [23] C. G. Mayhew, A. R. Teel, Hybrid control of planar rotations, in: Proceedings of the 2010 American Control Conference, IEEE, 2010, pp. 154–159.
  • [24] F. H. Clarke, Y. S. Ledyaev, R. J. Stern, P. R. Wolenski, Nonsmooth analysis and control theory, Vol. 178, Springer, 2008.
  • [25] R. Goebel, R. G. Sanfelice, A. R. Teel, Hybrid Dynamical Systems: Modeling, Stability, and Robustness, Princeton University Press, 2012.
  • [26] W. Wang, A. Teel, D. Nes̆ić, Analysis for a class of singularly perturbed hybrid systems via averaging, Automatica 48 (6) (2012).
  • [27] R. G. Sanfelice, On the existence of control lyapunov functions and state-feedback laws for hybrid systems, IEEE Transactions on Automatic Control 58 (12) (2013) 3242–3248.
  • [28] J. I. Poveda, A. R. Teel, A framework for a class of hybrid extremum seeking controllers with dynamic inclusions, Automatica 76 (2017) 113–126.
  • [29] S. Liu, A. Tanwani, Impulsive switching signals with functional inequalities: Stability analysis using hybrid systems framework, Automatica 171 (2025) 111928.
  • [30] R. G. Sanfelice, Hybrid feedback control, Princeton University Press, 2021.
  • [31] J. I. Poveda, N. Li, Robust Hybrid Zero-Order Optimization Algorithms with Acceleration via Averaging in Continuous Time, Automatica 123 (2021).
  • [32] C. Cai, A. R. Teel, R. Goebel, Smooth Lyapunov Functions for Hybrid Systems Part II: (Pre) Asymptotically Stable Compact Sets, IEEE Trans. Autom. Contr. 53 (3) (2008) 734–748.
  • [33] H. Durr, M. Stanković, K. H. Johansson, C. Ebenbauer, Extremum seeking on submanifolds in the Euclidian space, Automatica 50 (2014) 2591–2596.
  • [34] D. E. Ochoa, J. I. Poveda, Robust global optimization on smooth compact manifolds via hybrid gradient-free dynamics, Automatica 171 (2025) 111916.
  • [35] J. M. Lee, J. M. Lee, Smooth manifolds, Springer, 2012.
  • [36] H. Dürr, M. Krstic, A. Scheinker, C. Ebenbauer, Extremum seeking for dynamic maps using Liebrackets and singular perturbations, Automatica 83 (91-99) (2017).
  • [37] J. Cochran, M. Krstić, Nonholonomic Source Seeking With Tuning of Angular Velocity, IEEE Trans. on Autom. Control 54 (2009) 717–731.
  • [38] C. Zhang, D. Arnold, N. Ghods, A. Siranosian, M. Krstic, Source seeking with non-holonomic unicycle without position measurement and with tuning of forward velocity, Systems & control letters 56 (3) (2007) 245–252.
  • [39] J.-M. Yang, J.-H. Kim, Sliding mode control for trajectory tracking of nonholonomic wheeled mobile robots, IEEE Transactions on robotics and automation 15 (3) (1999) 578–587.
  • [40] C. M. Kellett, A compendium of comparison function results, Mathematics of Control, Signals, and Systems 26 (2014) 339–374.