LaTex2Web logo

Documents Live, a web authoring and publishing system

If you see this, something is wrong

Collapse and expand sections

To get acquainted with the document, the best thing to do is to select the "Collapse all sections" item from the "View" menu. This will leave visible only the titles of the top-level sections.

Clicking on a section title toggles the visibility of the section content. If you have collapsed all of the sections, this will let you discover the document progressively, from the top-level sections to the lower-level ones.

Cross-references and related material

Generally speaking, anything that is blue is clickable.

Clicking on a reference link (like an equation number, for instance) will display the reference as close as possible, without breaking the layout. Clicking on the displayed content or on the reference link hides the content. This is recursive: if the content includes a reference, clicking on it will have the same effect. These "links" are not necessarily numbers, as it is possible in LaTeX2Web to use full text for a reference.

Clicking on a bibliographical reference (i.e., a number within brackets) will display the reference.

Speech bubbles indicate a footnote. Click on the bubble to reveal the footnote (there is no page in a web document, so footnotes are placed inside the text flow). Acronyms work the same way as footnotes, except that you have the acronym instead of the speech bubble.

Discussions

By default, discussions are open in a document. Click on the discussion button below to reveal the discussion thread. However, you must be registered to participate in the discussion.

If a thread has been initialized, you can reply to it. Any modification to any comment, or a reply to it, in the discussion is signified by email to the owner of the document and to the author of the comment.

Table of contents

First published on Friday, May 23, 2025 and last modified on Friday, May 23, 2025 by François Chaplais.

Contributions to Robust Observer Design for Discrete-Time Systems

Gia Quoc Bao TRAN Mines Paris, Université PSL, France

Observer Design for Hybrid Systems
Ph.D. defense

TRAN Gia Quoc Bao
Mines Paris, Université PSL
With Pauline Bernard, Florent Di Meglio (Mines Paris), Ricardo G. Sanfelice (UCSC, USA) \( 22^{\text{nd}}\) November \( 2024\)

Slider 1. Observer Design for Hybrid Systems - PhD defense

1 Preamble

 

1.1 Acknowledgments

Working on this immensely challenging Ph.D. topic has been a rewarding journey, and I am deeply grateful to those who have supported and guided me along the way.

First and foremost, I would like to express my deepest gratitude to my supervisors, Dr. Pauline Bernard and Dr. Florent Di Meglio. The rigorous academic standards you upheld have pushed me to strive for excellence and have equipped me with the skills necessary to overcome the numerous challenges encountered along the way. Your dedicated mentorship has been instrumental in my professional and personal development. I could not have asked for better advisors, and this thesis would not have been possible without your invaluable contributions.

I am also immensely thankful to my collaborators, whose contributions have profoundly enriched this work. Dr. Ricardo G. Sanfelice, thank you for your collaboration and for welcoming me during my two-month visit to Santa Cruz to work on control of hybrid systems. Dr. Vincent Andrieu and Dr. Daniele Astolfi, thank you for inviting me to Lyon for the productive two-day collaboration and the subsequent follow-up work on discrete-time observers. I also extend my thanks to Dr. Thanh Phong Pham and Dr. Olivier Sename, who have been my partners on robust observers since before my Ph.D. started, as well as Dr. Thach Ngoc Dinh for our joint efforts on interval observers. I am deeply grateful to each of you for your invaluable contributions.

To the members of my Ph.D. jury, Dr. Romain Postoyan, Dr. Antoine Girard, Dr. Nathan van de Wouw, Dr. Angelo Alessandri, and Dr. Hyungbo Shim, thank you for taking the time to review my work, attend my defense, and provide constructive feedback. Your critical insights and thoughtful comments are instrumental in refining this dissertation.

I would also like to extend my heartfelt gratitude to my labmates, in France and the US, whose companionship and intellectual vigor have profoundly enriched this journey. Special thanks to the dedicated students at Mines Paris who joined me in their internships, namely Zikang Zhu, Aymeric Cardot, Shang Liu, Sergio Garcia, and Clément Boutaric. Thanks to Valentin Alleaume, Pauline’s new Ph.D. student, for his collaboration. Our results together are reflected in this thesis. The shared memories have made my time in the labs unforgettable.

To my family, words cannot adequately express my profound gratitude for your unwavering love and support. I hope I can see you more often in the future. To my friends, thank you for the much-needed distractions when the weight of academia became overwhelming.

Let me acknowledge those who have supported me in more indirect ways: those who provided constructive feedback on my work—including Dr. Lorenzo Marconi and Dr. Lucas Brivadis, those who funded my US stay—Fondation Sciences Mathématiques de Paris and Fondation Mines Paris, and all whom I failed to list here. I am grateful for having you on my journey.

Last but not least, I thank myself for the perseverance and dedication put into this work.

1.2 Résumé

En théorie du contrôle, les observateurs sont des algorithmes conçus pour estimer en temps réel l’état d’un système (dynamique) à partir de ses sorties mesurées et des entrées connues, à des fins incluant le contrôle ou le diagnostic. Bien qu’il existe une abondante littérature sur les observateurs pour les systèmes dynamiques non linéaires en temps continu et discret, la plupart des résultats pour les systèmes hybrides—présentant à la fois un comportement en temps continu et des sauts discontinus selon l’état—sont limités à des cas spécifiques tels que les systèmes commutés ou les systèmes en temps continu avec des sorties échantillonnées, et ne se généralisent pas facilement aux systèmes hybrides généraux. Ce travail de thèse vise à combler cette lacune en développant plusieurs cadres théoriques pour la synthèse d’estimateurs d’état robustes pour cette grande classe de systèmes, avec diverses applications en robotique et dynamique non lisse. Comme étape clé vers cet objectif, la première partie conçoit de nouveaux observateurs robustes pour les systèmes non linéaires en temps discret. En effet, les observateurs hybrides reposent souvent en partie sur des observateurs en temps discret, mais peu de méthodes générales de synthèse pour les systèmes discrets non linéaires existent dans la littérature. Ensuite, comme la conception des observateurs hybrides varie selon que les instants de saut sont connus\( /\) détectés ou non, nous examinons ces deux cas dans les parties suivantes. Cette thèse en trois parties est détaillée comme suit.

La première partie se concentre sur les observateurs robustes pour les systèmes non linéaires en temps discret. Après une revue de la littérature, nous introduisons une nouvelle version en temps discret de l’observateur classique à grand gain, bien connu en temps continu. Cet observateur fonctionne sous la condition de constructibilité du système, une version moins stricte de l’observabilité pour laquelle la distingabilité en temps rétrograde est suffisante. Nous présentons ensuite un nouvel observateur Kravaris-Kazantzis$\slash$Luenberger[KKL] pour les systèmes discrets non linéaires à temps variant. Cette méthode consiste à chercher un changement de coordonnées transformant la dynamique en un filtre stable de la sortie, et inverser cette transformation pour retrouver une estimée de l’état. Les deux observateurs fonctionnent lorsqu’ils sont accélérés suffisamment rapidement, exhibent des phénomènes similaires à ceux de l’observateur grand gain en temps continu, et sont robustes face aux incertitudes. Ils sont ensuite appliqués à un moteur synchrone à aimants permanents et comparés à la version discrétisée de leurs équivalents en temps continu.

La deuxième partie aborde la conception d’observateurs pour les systèmes hybrides avec des temps de saut connus. Les sauts étant détectés, l’observateur est synchronisé avec le système, c’est-à-dire que ses sauts sont déclenchés en même temps que ceux du système. Dans le cas linéaire, nous proposons un observateur systématique de type Kalman, combinant l’observabilité du flot et des sauts grâce à une seule matrice de covariance liée à une nouvelle version hybride du Gramien d’observabilité. Une autre approche décompose l’état du système en une partie instantanément observable à partir de la sortie et des dynamiques de flot, et une partie observable par la combinaison du flot et des sauts. Pour la première, un observateur à grand gain est construit, couplé avec un observateur basé sur les sauts pour la seconde. Cette idée est généralisée aux systèmes hybrides non linéaires. Nous donnons des conditions suffisantes basées sur Lyapunov pour coupler un observateur à grand gain basé sur le flot avec un observateur basé sur les sauts. Nous proposons différentes méthodes de synthèse pour ce dernier, incluant des approaches basées sur des inégalités matricielles linéaires, des observateurs à horizon glissant ou du KKL. Le point fort est l’utilisation d’une sortie fictive, décrivant comment la seconde partie de l’état impacte la première lors des sauts. En effet, cette dernière devient visible au cours de l’intervalle de flot suivant, rendant la seconde partie de l’état détectable via l’interaction entre le flot et les sauts. Des applications en robotique sont ensuite présentées, incluant l’estimation de biais dans une unité de mesure inertielle et l’estimation de l’état et des incertitudes dans différentes configurations d’un robot marcheur bipède.

La troisième partie explore la conception d’observateurs pour les systèmes hybrides avec des temps de saut inconnus. En supposant que la sortie soit identique avant et après les sauts (ce qui rend les sauts indétectables), nous proposons une approche systématique basée sur KKL, transformant le système hybride en un filtre stable en temps continu de la sortie, construisant un observateur en temps continu, et récupérant l’estimation en inversant cette transformation, sans avoir à détecter les instants de saut. Nous fournissons des conditions suffisantes pour que cette transformation soit bien définie et injective en dehors de l’ensemble de saut, permettant la construction de son inverse à gauche pour l’utiliser dans l’observateur. La convergence asymptotique en dehors des temps de saut, avec une robustesse face aux incertitudes, est alors démontrée, dans une métrique non euclidienne prenant en compte l’indistinguabilité des états avant et après les sauts. Comme application, nous étudions l’estimation de la force de friction dans le phénomène de stick-slip, fréquent dans le forage pétrolier. Dans ce cas, la transformation inverse est approximée à l’aide d’un réseau de neurones entraîné sur des données de simulations hors ligne.

Enfin, la conclusion et les perspectives de travaux futurs sont discutées à la fin de cette thèse.


Mots clés : observateurs, systèmes hybrides, systèmes nonlinéaires, robotique, stick-slip

1.3 Abstract

In control theory, observers are algorithms designed to estimate in real time the state of a (dynamical) system from its measured outputs and known inputs, for different purposes such as control or diagnosis. While there is extensive literature on observer designs for nonlinear continuous- and discrete-time dynamical systems, most existing results for hybrid systems—those whose solutions can exhibit both continuous-time behavior and discontinuous jumps depending on the state value—are limited to specific classes such as switched\( /\) impulsive systems or continuous-time systems with sampled outputs and hence are lacking for general hybrid systems. This Ph.D. work aims to address this gap by developing several theoretical frameworks for robust state estimation of this large class of systems, together with various applications in robotics and non-smooth dynamics. As a fundamental step towards this objective, since hybrid observer design often relies on robust discrete-time observers for which few general methods exist in the literature, the first part of this work designs new robust observers for nonlinear discrete-time systems. Next, since hybrid observer designs differ greatly depending on whether the jump times of the systems are known\( /\) detected or not, we individually study these two cases in the subsequent parts. This three-part dissertation is presented in more detail as follows.

The first part of this dissertation focuses on robust observers for nonlinear discrete-time systems. After some literature review, we first present a novel discrete-time version of the classical high-gain observer, which is well-known in continuous time. This observer is shown to work under the constructibility of the system, a less stringent counterpart of observability for which backward distinguishability is a sufficient condition. Then, we present a novel Kravaris-Kazantzis$\slash$Luenberger[KKL] observer for nonlinear time-varying discrete-time systems, relying on transforming the system into some stable filter of the output for which an observer is readily designed and then recovering the estimate by left-inverting this transformation. Both observers are shown to work when pushed sufficiently fast, exhibit phenomena similar to the continuous-time high-gain design, and be robust against uncertainties. Both are then applied to the Permanent Magnet Synchronous Motor[PMSM] and compared with the discretized version of their continuous-time counterparts.

The second part of this dissertation addresses observer designs for hybrid systems with known jump times. Since the solution’s jumps are detected, the observer is synchronized with the system, i.e., its jumps are triggered at the same time as those of the system. In the case of linear dynamics and output maps, we propose a systematic Kalman-like observer, which blends observability from the combination of flows and jumps thanks to a single covariance matrix linked to a newly defined hybrid version of the observability Gramian. Another approach is to decompose the system’s state into a part that is instantaneously observable from the flow output and dynamics, and a part that is not. For the former, a high-gain observer is built, which we couple with a jump-based observer for the latter. This decomposition idea is then generalized to hybrid systems with nonlinear maps. In this general setting, we give Lyapunov-based sufficient conditions to couple a high-gain flow-based observer with a jump-based one. We then propose different jump-based observers for the non-instantaneously-observable part of the state. These include Linear Matrix Inequality[LMI]-, Moving Horizon Observer[MHO]-, and KKL-based schemes. The highlight of this coupling of observers is the use of a fictitious output, which describes the way the second part of the state impacts the first one at jumps. Indeed, the latter becomes visible during the next flow interval, which can make the second part of the state detectable\( /\) observable via this interaction between flows and jumps. Applications in robotics are presented next, including bias estimation in an Inertial Measurement Unit[IMU], or state and uncertainty estimation in different configurations of a bipedal walking robot.

The third part of this dissertation explores observer designs for hybrid systems with unknown jump times. Based on the assumption that the output is the same before and after the jumps (thus making the jumps undetectable), we present a systematic KKL-based design approach of transforming the hybrid system into a continuous-time stable filter of the output, building for this a continuous-time observer, and then recovering the estimate by inverting this transformation, without detecting the jump times. The main work is to give sufficient conditions to have this transformation well-defined and importantly, injective outside the jump set so that its left inverse can be constructed to use in the observer. Asymptotic convergence in the original coordinates outside of the jump times, with robustness against uncertainties, is then shown, in a non-Euclidean metric taking into account the indistinguishability of the states before and after the jumps. As an application, we study the friction force estimation problem in the stick-slip phenomenon commonly encountered in oil drilling. In this case, the inverse transformation is approximated using a Neural Network[NN] strategically trained on data obtained through offline simulations.

Finally, the conclusion and future work are discussed at the end of this dissertation.


Keywords : observers, hybrid systems, nonlinear systems, robotics, stick-slip

1.4 Introduction

Ce chapitre donne un aperçu des résultats de ce travail de doctorat. Nous introduisons les observateurs, décrivons les défis associés à la conception d’observateurs pour les systèmes hybrides, puis présentons un aperçu des principales contributions de cette thèse.

1.4.1 Context and Challenges

In control theory, (state) observers—sometimes referred to as estimators or filters—are indispensable tools for estimating in real time the state of dynamical systems when direct measurement of the state variables is impossible or impractical [1]. As illustrated in Figure 7, these algorithms leverage the system’s measured outputs (obtained thanks to sensors) and known inputs, potentially utilizing their history as well, to reconstruct in real time the current state, which contains critical information about the system’s current condition. In certain contexts, the system’s unknown parameters or uncertainties can be modeled as an extra state component, so state estimation schemes typically encompass parameter or unknown input estimation schemes (thus, in this dissertation we sometimes use these concepts interchangeably in our considered applications). The estimated state then serves multiple purposes, such as enabling precise feedback control, facilitating continuous system monitoring, and supporting fault detection and diagnosis. By providing a reliable estimate of the state, observers allow for implementing control strategies that would otherwise be infeasible, enhance the understanding of system behavior, and help identify and mitigate issues that could lead to system failures. The importance of observers is underscored by their widespread application in various fields, namely robotics [2, 3, 4], vehicles [5, 6], electrical machines [7, 8], power systems [9], etc. The ability to infer the system’s state from the measured information is related to the notion of observability of the said system, and types of observers may require different versions of this notion, which is linked to the desired property of the estimation error (i.e., the difference between the real state and its estimate provided by the observer), such as asymptotic convergence to zero, tunability of convergence rate, and robustness against uncertainties (inputs disturbances, measurement noise, modeling errors, etc.). Overall, research on various types of observers for different classes of systems, as well as their properties and applications, has long been a topic of significant interest in control theory.

<span data-controller="mathjax">General scheme of using an observer for state estimation in a dynamical system.</span>
Figure 7. General scheme of using an observer for state estimation in a dynamical system.

In this work, we design observers for general hybrid systems defined in the framework of [10], specifically those whose state exhibits both continuous and discontinuous evolution over time (also known as flows and jumps, respectively), depending on the current state value and\( /\) or exogenous inputs. As explained later in Section 1.5, this class of systems is exceptionally broad, encompassing various non-smooth systems such as impulsive differential equations [11], switched systems [12], hybrid automata [13], etc. Consequently, our findings are applicable to these particular classes, as will be illustrated in our examples and applications throughout this dissertation. Hybrid systems are incredibly versatile and find applications in numerous fields, including sample-and-hold control [14], robotics [15, 16], power systems [10, Example 1.3], etc. Besides, hybrid techniques can be exploited for controlling constrained systems [17], or to improve the performance of observers designed for continuous-time systems [18, 19, 20, 21]. Our work not only broadens the applicability of hybrid systems but also pushes the boundaries of what can be achieved in terms of control and observation in complex, real-world scenarios.

The primary challenges encountered in this Ph.D. work when designing observers for hybrid systems are:

  • Challenge \( \mathbf{1}\) : The time domains of hybrid solutions are typically unknown before their initialization [22], unlike in continuous- or discrete-time systems. Consequently, the estimation error generally cannot be defined in the traditional sense, i.e., by comparing the true state and the estimate at the same hybrid time. Even if we forget about the hybrid nature of the time domains, this Euclidean distance cannot converge to zero if the jump times of the system and observer are not exactly synchronized, and new notions of observers and convergence thus have to be developed. This major issue has also been encountered in feedback stabilization or stability analysis of hybrid systems or synchronization, where generalized distances have been proposed [23, 24, 25]. A pertinent example is the stiction phenomenon encountered in oil drilling, discussed in Section 4.3, when the end of the drill lying deep underground exhibits stick-slip behavior, constituting a switched system with state jumps and whose jump times (the instants when we switch from stick to slip and vice-versa) are undetectable via overground sensors, especially when the friction forces—the parameters for those transitions—are unknown. This challenge disappears when the jump times of the system are known\( /\) detected and the observer’s jump times can be triggered at the same time as those of the system;
  • Challenge \( \mathbf{2}\) : Even in this latter favorable case where jump times are known, the observability of hybrid systems can arise from flows, jumps, or a nonlinear combination of both as seen in Section 3 later. Academic examples and applications in Section 3 show that a hybrid system whose flow and jump parts are both observable can be non-observable and vice-versa. As a result, observers need to strategically gather and integrate information from these sources to estimate the otherwise hidden state components. An important example is the estimation of unknown restitution coefficients\( /\) biases (see Section 3.4) that cannot be measured by any physical sensors but that become visible thanks to impacts and is estimated during flows, leveraging the hybrid nature of the system;
  • Challenge \( \mathbf{3}\) : There are very few results on robust discrete-time observers. This difficulty arises because designing hybrid observers often requires coupling observers that estimate different parts of the state, generally combining a continuous-time observer with a discrete-time one. This combination is made possible under the robustness of each observer, treating the estimation error linked to the part not being estimated by the considered observer as a disturbance (see Section 3.3). At the start of this Ph.D., robust nonlinear observers were available in continuous time [26, 27, 28, 29, 30], but not in discrete time.

Therefore, the design of observers for hybrid systems presents a highly nonlinear problem, contributing significantly to its complexity. This nonlinearity arises not only from the form of the system’s maps but also from the solutions’ unpredictable time domain due to flow\( /\) jump conditions, which must be handled with utmost caution. However, this complexity also opens up intriguing theoretical and practical problems that can lead to significant advancements in control theory and its applications, which will be thoroughly addressed in this dissertation.

1.4.2 Contributions and Dissertation Outline

First, we tackle Challenge \( \mathbf{3}\) in Section 2, by proposing observers in discrete time that are robust against uncertainties. After an exhaustive review of existing related results in Section 2.1, we present, in Section 2.2 and Section 2.3 respectively, the discrete-time counterpart of the renowned high-gain observer in continuous time [26, 27], and the Kravaris-Kazantzis$\slash$Luenberger[KKL] observer for nonlinear time-varying systems. Besides being robust, the designed schemes do not assume any specific structures of the system’s dynamics and output maps, making them applicable to a large class of nonlinear systems. Both are then applied to a strategically discretized Permanent Magnet Synchronous Motor[PMSM] in Section 2.4, demonstrating significant advantages over observers designed in continuous time and then discretized. These results represent Contribution \( \mathbf{1}\) of this Ph.D.

To address Challenge \( \mathbf{1}\) and Challenge \( \mathbf{2}\) with hybrid systems, our strategy as illustrated in Figure 8 is to consider two main cases: when the jump times of the system’s solutions, i.e., the times at which discontinuous jumps occur, are known and when they are not. For a broad class of hybrid systems including mechanical systems with impacts such as [31], the solution’s jump times can be inferred from the system’s measured output (e.g., in a bouncing ball as in Example 1 when the measured ball position is at the ground) or detected thanks to sensors. For some other classes of hybrid systems, such as mechanical systems exhibiting stick-slip (see Section 4.3), the detection of solutions’ jump times is challenging, thus classifying them into the second main case of unknown jump times.

In the first main case, knowing when the real solution jumps (modulo some possible delays that would be handled by the observer’s robustness) allows the observer to be synchronized with the system, meaning that the estimate is triggered to jump with this solution. This synchronization ensures that the real solution and the estimate share the same hybrid time domain, making the estimation error well-defined and simplifying stability analysis. This critical case is presented in Section 3, with the primary approach being to combine observability of different components of the state as will be explained right next. Within this part, after a thorough literature review in Section 3.1, we consider cases of hybrid systems with known jump times, increasing in nonlinearity in their dynamics and output maps:

  1. Linear maps (see Section 3.2): We propose here two alternative approaches:
    1. A Kalman-like observer: This algorithm systematically blends observability from the combination of flows and jumps using a single covariance matrix, facilitating straightforward design and implementation without requiring observability checks. It relies on our newly defined versions of the observability Gramian and Uniform Complete Observability[UCO] for hybrid systems. Applications to various classes of systems, including the Inertial Measurement Unit[IMU]—see Section 3.4, are presented;
    2. Observability decomposition: Exploiting the structure of the maps, we linearly decompose the system’s state into a part that is instantaneously observable from the flow output (i.e., belonging to the observable subspace) and a part that is not. For the former, a high-gain Kalman-like observer [27] with jump resets is built, which is coupled with a jump-based observer designed using a Linear Matrix Inequality[LMI] or the KKL framework for the latter. It is shown that, besides the real jump output, we can exploit in the jump-based observer a special kind of measurement arising from the flow-jump combination to correct its estimate. This idea addresses Challenge \( \mathbf{2}\) and will be made clearer right next, referred to as the fictitious measurement.
  2. Nonlinear maps (see Section 3.3): When the system’s maps are nonlinear, the strategy is to perform observability decomposition as said above, but here in a nonlinear way. More precisely, we rely on the combination of i) a high-gain flow-based observer, i.e., one that estimates the instantaneously observable part of the state during flows using the flow output, and ii) a jump-based observer for the rest of the state, either using the jump output or a fictitious one resulting from the flow-jump combination, when this part of the state interacts with the instantaneously observable one at jumps. Indeed, at the end of each flow interval (right before the jump), the estimation error of the instantaneously observable part of the state has been reduced thanks to the high-gain observer, and at the jump, the other parts of the state may interfere with this one, causing an error that then gets corrected in the next flow interval. This measurement is thus fictitious, arising from the hybrid nature of the system and allowing us to estimate state components that are not measurable thanks to physical sensors. It makes the special output considered earlier for the case of linear maps clearer and more nonlinear. While the structure of the flow-based observer is fixed to be the ones in [26, 27], the other one depends on the structures of the system’s maps, and so we consider two subcases of nonlinear maps:
    1. The jump maps are affine in the non-instantaneously-observable states: The jump-based observer can either be built thanks to LMIs or KKL as above, but more intricate dependence on the history of the instantaneously observable part of the state and the inputs must be considered. These methods are applied to restitution coefficient estimation in a bipedal walking robot (a highly nonlinear mechanical system with impacts) in Section 3.4;
    2. The maps are fully nonlinear: The jump-based part is fully nonlinear and can either be a Moving Horizon Observer[MHO] or a KKL observer. These methods are then applied to parameter estimation in mechanical systems with impacts in Section 3.4.

    These two cases of systems with nonlinear maps constitute Contribution \( \mathbf{3}\) of this Ph.D.

On the other hand, when the solution’s jump times are unknown, assuming that the measured output is continuous at jumps, we propose to glue the time domain, i.e., to transform the hybrid system into some continuous-time system that does not jump, build an observer for it, and then inverting this transformation (when we can) to retrieve the estimate—see Section 4. This technique builds on the general gluing concept in [32], but here it becomes systematic thanks to the KKL paradigm with some target continuous-time dynamics taking a fixed form of a linear stable filter of the measured output, for which an observer in these new coordinates can always be designed. After a comprehensive review of existing observer designs for hybrid systems with unknown jump times in Section 4.1, the theoretical contributions, including the definition of the transformation into the target dynamics and its regularity, are presented in Section 4.2. The injectivity of this transformation outside of the jump set, achieved under backward distinguishability and some regularity of the hybrid system’s maps, ensures that it can be inverted to recover the estimate in the original coordinates. The robust convergence of the estimate outside of the jump times is then demonstrated. In Section 4.3, this gluing observer is applied to the stiction phenomenon found in oil drilling, when the bottom part of the drill exhibits some stick-slip behavior, allowing us to estimate unknown friction forces deep underground using overground sensors. We also propose a systematic implementation scheme using a Neural Network[NN] trained with data obtained from offline simulations. These results represent Contribution \( \mathbf{4}\) of this Ph.D. that deals with Challenge \( \mathbf{1}\) .

<span data-controller="mathjax">An overview of observer designs for hybrid systems in this dissertation.</span>
Figure 8. An overview of observer designs for hybrid systems in this dissertation.

All in all, this work proposes a fairly complete framework of observer design for general hybrid systems, with the following main contributions:

  • Contribution \( \mathbf{1}\) : Robust observer design for nonlinear (time-varying) discrete-time systems, with application to a PMSM;
  • Contribution \( \mathbf{2}\) : Observer design for hybrid systems with linear maps and known jump times, with application to an IMU;
  • Contribution \( \mathbf{3}\) : Observer design for hybrid systems with nonlinear maps and known jump times, with application to mechanical systems with impacts (including walking robots);
  • Contribution \( \mathbf{4}\) : Observer design for hybrid systems with unknown jump times, with application to the stiction phenomenon.

Additionally, during my Ph.D., I have had the opportunity to collaborate with external researchers on various aspects of continuous- and discrete-time observers. These collaborations have led to the following contributions to the field of observer design, which are not included in this dissertation:

  • In [33, 34]: Systematic interval observer design for non-autonomous linear systems and autonomous nonlinear discrete-time systems;
  • In [35, 36]: Robust observer design for continuous-time descriptor systems, with application to vehicle suspensions.

1.4.3 Other Aspects of This Ph.D. Work

1.4.3.1 Student Supervision

Throughout my Ph.D., I have had the privilege of co-supervising several master’s students alongside my advisors, Prof. Pauline Bernard and Prof. Florent Di Meglio, during their research internships at Mines Paris - PSL. The students and their research projects, listed by the order in which their results appear in this dissertation, are as follows:

  • Clément Boutaric, Kalman-Like Observer for Hybrid Systems with Linear Maps and Known Jump Times, November \( 2022\) - February \( 2023\) , related results included in Section 3.2;
  • Aymeric Cardot, Observer Design for Hybrid Systems with Fully Nonlinear Maps and Known Jump Times, November \( 2023\) - February \( 2024\) , related results included in Chapters 3.3 and 3.4;
  • Shang Liu, Hybrid Observer Design for a Three-Link Biped Robot, November \( 2023\) - February \( 2024\) , related results included in Section 3.4;
  • Sergio Garcia, Gluing KKL Observer for Hybrid Systems with Unknown Jump Times, November \( 2023\) - February \( 2024\) , related results included in Section 4.2;
  • Zikang Zhu, NN-Based Implementation of Gluing KKL Observer for Hybrid Systems with Unknown Jump Times, June - September \( 2024\) , related results included in Section 4.3.

Close to the end of my Ph.D., I had the pleasure of collaborating with Valentin Alleaume, a new Ph.D. student working with my supervisors. Our collaboration is reflected in several joint papers.

1.4.3.2 Honors and Awards

During my Ph.D., I have been honored with the following awards and recognitions:

1.4.3.3 Short Courses

During my Ph.D., I have participated in the following technical courses:

  • \( 44\) th International Summer School of Automatic Control, Grenoble, France, August \( 2023\) ;
  • Introduction to Nonlinear Systems and Control (Instructor: Prof. Hassan Khalil), International Graduate School on Control, European Embedded Control Institute, Paris-Saclay, France, May \( 2022\) .

1.4.4 Publications

The publications of this Ph.D. are listed below.

Journal articles

Gia Quoc Bao Tran, Valentin Alleaume, Pauline Bernard, Ricardo G. Sanfelice. "Existence and Regularity of KKL Gluing Change of Coordinates for Hybrid Systems". Will be submitted to a journal, content taken from Section 4.2.

Gia Quoc Bao Tran, Valentin Alleaume, Pauline Bernard, Florent Di Meglio. "Dealing with Indistinguishability in Gluing KKL Observer Design for Hybrid Systems with Unknown Jump Times". Will be submitted to a journal, content taken from Section 4.2 and Section 4.3.

Gia Quoc Bao Tran, Pauline Bernard, Ricardo G. Sanfelice. "Observer Design for Hybrid Systems with Known Jump Times". Will be submitted to a journal, content taken from Section 3.3 and Section 3.4.

Gia Quoc Bao Tran, Pauline Bernard, Ricardo G. Sanfelice. "Observer Design for Hybrid Systems with Partially Affine Forms and Known Jump Times: Applications a Walking Robots". Submitted to Automatica.
Link: https://hal.science/hal-04885099

Gia Quoc Bao Tran, Pauline Bernard. "Arbitrarily Fast Robust KKL Observer for Nonlinear Time-Varying Discrete Systems". IEEE Transactions on Automatic Control, Volume \( 69\) , Issue \( 3\) , March \( 2024\) .
DOI: 10.1109/TAC.2023.3328833

Book chapters

Gia Quoc Bao Tran, Pauline Bernard, Lorenzo Marconi. "Observer Design for Hybrid Systems with Linear Maps and Known Jump Times". Hybrid and Networked Dynamical Systems (Editors: Romain Postoyan, Paolo Frasca, Elena Panteley, Luca Zaccarian), pp. \( 115\) -\( 154\) , Lecture Notes in Control and Information Sciences, Volume \( 493\) , Springer, March \( 2024\) .
DOI: 10.1007/978-3-031-49555-7_6

Conference papers and extended abstracts

Gia Quoc Bao Tran, Sergio Garcia, Pauline Bernard, Florent Di Meglio, Ricardo G. Sanfelice. "Towards Gluing KKL Observer for Hybrid Systems with Unknown Jump Times". \( 63\) rd IEEE Conference on Decision and Control (CDC), Milan, Italy, December \( 2024\) .
Link: https://hal.science/hal-04685195

Gia Quoc Bao Tran, Pauline Bernard, Vincent Andrieu, Daniele Astolfi. "Constructible Canonical Form and High-Gain Observer in Discrete Time". \( 63\) rd IEEE Conference on Decision and Control (CDC), Milan, Italy, December \( 2024\) .
Link: https://hal.science/hal-04685007

Gia Quoc Bao Tran, Pauline Bernard, Florent Di Meglio, Ricardo G. Sanfelice. "Observer Design for Hybrid Systems: Applications to Robotics and Hybrid Dynamics". Extended abstract at \( 11\) th European Nonlinear Dynamics Conference (ENOC), Delft, Netherlands, July \( 2024\) .

Gia Quoc Bao Tran, Pauline Bernard. "Kalman-Like Observer for Hybrid Systems with Linear Maps and Known Jump Times". \( 62\) nd IEEE Conference on Decision and Control (CDC), Singapore, December \( 2023\) .
DOI: 10.1109/CDC49753.2023.10383629

Gia Quoc Bao Tran, Pauline Bernard, Ricardo G. Sanfelice. "Coupling Flow and Jump Observers for Hybrid Systems with Known Jump Times". \( 22\) nd IFAC World Congress, Yokohama, Japan, July \( 2023\) .
DOI: 10.1016/j.ifacol.2023.10.522

Gia Quoc Bao Tran, Pauline Bernard, Florent Di Meglio, Lorenzo Marconi. "Observer Design based on Observability Decomposition for Hybrid Systems with Linear Maps and Known Jump Times". \( 61\) st IEEE Conference on Decision and Control (CDC), Cancun, Mexico, December \( 2022\) .
DOI: 10.1109/CDC51059.2022.9993225

Publications from external collaborations during this Ph.D. (\( \star\) denotes co-first authorship)

Gia Quoc Bao Tran, Thanh-Phong Pham, Olivier Sename. "Reduced-Order Observer for a Class of Uncertain Descriptor NLPV Systems: Application to the Vehicle Suspension". Submitted to a journal.

Thach Ngoc Dinh\( ^\star\) , Gia Quoc Bao Tran\( ^\star\) . "Systematic Interval Observer Design for Linear Systems". Submitted to a journal.
DOI: 10.48550/arXiv.2405.06445

Thach Ngoc Dinh\( ^\star\) , Gia Quoc Bao Tran\( ^\star\) . "A Unified KKL-Based Interval Observer for Nonlinear Discrete-Time Systems". IEEE Control Systems Letter (jointly presented at \( 63\) rd IEEE CDC), Volume \( 8\) , April \( 2024\) .
DOI: 10.1109/LCSYS.2024.3394324

Gia Quoc Bao Tran, Thanh-Phong Pham, Olivier Sename. "Fault Estimation Observers for the Vehicle Suspension with a Varying Chassis Mass". \( 12\) th IFAC Symposium on Fault Detection, Supervision and Safety for Technical Processes (SAFEPROCESS), Ferrara, Italy, June \( 2024\) .
DOI: 10.1016/j.ifacol.2024.07.207

Seminars, talks, and posters

Poster at \( 63\) rd IEEE Conference on Decision and Control. "Observer Design for Hybrid Systems". Milan, Italy, December \( 2024\) .

Seminar at Division of Systems and Control, Uppsala University. "Observer Design for Hybrid Systems". Uppsala, Sweden, October \( 2024\) .

Seminar at Division of Decision and Control Systems, KTH Royal Institute of Technology. "Observer Design for Hybrid Systems". Stockholm, Sweden, October \( 2024\) .

Poster at Mines Paris Ph.D. Student Day. "Observer Design for Hybrid Systems". Paris, France, June \( 2024\) .

Seminar at LAGEPP (CNRS and Université Lyon \( 1\) ). "Observer Design for Hybrid Systems". Villeurbanne, France, February \( 2024\) .

Poster at \( 62\) nd IEEE Conference on Decision and Control. "Observer Design for Hybrid Systems". Singapore, December \( 2023\) .

Talk at French Observer Group Day (Journée du GT SynObs). "Kalman-Like Observer for Hybrid Systems with Linear Maps and Known Jump Times". Paris, France, December \( 2023\) .

Student talk at \( 44\) th International Summer School of Automatic Control. "Observer Design for Hybrid Systems with Known Jump Times". Grenoble, France, August \( 2023\) .

Seminar at Hybrid Systems Laboratory, University of California Santa Cruz. "Observer Design for Hybrid Systems with Known Jump Times". Santa Cruz, California, USA, May \( 2023\) .

Poster at \( 5\) th NorCal Control Workshop, University of California Berkeley. "Observer Design for Hybrid Systems". Berkeley, California, USA, April \( 2023\) .

Talk at French Hybrid System Technical Committee Reunion (Réunion du CT SDH). "Coupling Observers for Hybrid Systems with Known Jump Times". Paris, France, February \( 2023\) .

Poster at \( 9\) th Heidelberg Laureate Forum. "Observer Design for Hybrid Systems by Observability Decomposition". Heidelberg, Germany, September \( 2022\) .

Seminar at Hanoi University of Science and Technology (HUST). "Arbitrarily Fast Robust KKL Observer for Nonlinear Time-Varying Discrete Systems". Hanoi, Vietnam, August \( 2022\) .

Talk at French Observer Group Day (Journée du GT SynObs). "Arbitrarily Fast Robust KKL Observer for Nonlinear Time-Varying Discrete Systems". Paris, France, June \( 2022\) .

1.5 Notations and Preliminaries

\chaptermark{Notations and Preliminaries}

Ce chapitre contient les préliminaires nécessaires à la compréhension du reste de cette thèse. Nous commençons par un aperçu des concepts mathématiques connexes, puis nous présentons les systèmes dynamiques et les observateurs, ainsi que leurs propriétés, en temps continu et discret. Ensuite, nous décrivons les systèmes hybrides et leurs sous-classes, suivis d’une discussion sur leurs propriétés en lien avec la conception d’observateurs hybrides. Les notations utilisées dans cette thèse sont également introduites et classées dans leurs sections respectives.

1.5.1 Mathematical Definitions and Properties

In this section, we introduce the mathematical preliminaries and the related notations used throughout this dissertation. The notations are organized into specific sections to facilitate easy reference for the reader.

1.5.1.1 Sets and Topology

Notations. Denote \( \emptyset\) as the empty set. Let \( \mathbb{R}\) (resp., \( \mathbb{Z}\) ) denote the set of real numbers (resp., integers, i.e., \( \{…,-2,-1,0, 1, 2, …\}\) ). Let \( \mathbb{N}\) denote the set of natural numbers, i.e., \( \{0, 1, 2, …\}\) . Let \( \mathbb{R}_{\geq a}\) (resp., \( \mathbb{R}_{> a}\) ) denote the interval \( [a,+\infty)\) (resp., \( (a,+\infty)\) ), and similarly for \( \mathbb{Z}_{\geq a}\) and \( \mathbb{N}_{\geq a}\) . Similarly, we denote \( \mathbb{R}_{\leq a}\) (resp., \( \mathbb{R}_{< a}\) ), as well as \( \mathbb{Z}_{\leq a}\) and \( \mathbb{N}_{\leq a}\) . Denote \( \mathbb{C}\) as the set of complex numbers. Let \( \Re(z)\) and \( \Im(z)\) be the real and imaginary parts of \( z \in \mathbb{C}\) , respectively. Given \( \rho > 0\) , define \( \mathbb{C}_\rho = \{\lambda \in \mathbb{C}: \Re(\lambda) < -\rho\}\) . Denote \( \mathbb{R}^{m \times n}\) (resp., \( §_{> 0}^{n}\) ) as the set of real \( (m \times n)\) - (resp., symmetric positive definite real \( (n \times n)\) -) dimensional matrices. The sequence \( a_1,a_2,…\) (of numbers, vectors, sets, functions, etc.) is denoted as \( (a_n)_{n\in \mathbb{N}}\) . For some set \( S\) , let \( S^{\mathbb{N}}\) be the set of sequences whose elements belong to \( S\) . The closure of a set \( S\) , denoted as \( \rm{cl}(S)\) , is the smallest closed superset of \( S\) , while the interior of \( S\) , denoted as \( \rm{int}(S)\) , is the largest open subset contained within \( S\) . Denote \( \mathbb{B}\) (resp., \( \mathring{\mathbb{B}}\) ) as the closed (resp., open) unit ball and \( § + \delta\mathbb{B}\) (resp., \( § + \delta\mathring{\mathbb{B}}\) ) as the set of points within (resp., strictly smaller than) a distance \( \delta > 0\) from any point in \( §\) . For \( x\in \mathbb{R}^n\) , \( B_{r}(x)\) denotes the open ball of radius \( r > 0\) centered at \( x\) . For a non-empty set \( S \subseteq \mathbb{R}^n\) , let \( d_S(x):= \inf_{y\in S}|x-y|\) be the distance from \( x \in \mathbb{R}^n\) to \( S\) .

A metric space is an ordered pair \( (M, d)\) where \( M\) is a set and \( d\) is a metric on \( M\) , satisfying the following conditions: i) \( d(x,y) \geq 0\) for all \( (x,y) \in M \times M\) and \( d(x,y) = 0 \Leftrightarrow x = y\) , ii) \( d(x,y) = d(y,x)\) for all \( (x,y) \in M \times M\) , and iii) \( d(x,z) \leq d(x,y) + d(y,z)\) for all \( (x,y,z) \in M \times M \times M\) (the triangle inequality). If \( M\) is a normed vector space, a possible metric is \( d(x,y):= |x-y|\) for \( (x,y) \in M \times M\) , where \( |\cdot|\) denotes the corresponding norm.

A subset \( S\) of a metric space \( (M,d)\) is said to be open if, for each \( x \in S\) , there exists \( \epsilon > 0\) such that any \( y \in M\) with \( d(x,y) < \epsilon\) is also in \( S\) . A closed subset of \( (M,d)\) is one whose complement in \( (M,d)\) is open. A set \( S\) in a metric space \( (M, d)\) is bounded if there exists \( M \geq 0\) such that \( d(x, y) \leq M\) for all \( (x,y) \in S \times S\) . In a finite-dimensional Euclidean space, a set is compact if it is both closed and bounded.

The Lebesgue measure is a function \( \mu\) that assigns a non-negative number to each measurable set, generalizing the notions of length, area, and volume. A property is said to hold for almost any (or almost every) point in a set \( S\) if the set of points where the property does not hold has Lebesgue measure zero, i.e., this property holds for all points in \( S\) except for a subset of \( S\) with Lebesgue measure zero.

1.5.1.2 Matrices

A vector-induced matrix norm is defined based on a vector norm. Given a vector norm \( |\cdot|\) on \( \mathbb{R}^n\) , the induced norm \( \|A\|\) of a matrix \( A \in \mathbb{R}^{m \times n}\) is defined as:

\[ \begin{equation} \|A\| = \sup_{|x| \neq 0} \frac{|Ax|}{|x|} = \sup_{|x| = 1} |Ax|. \end{equation} \]

(1)

In this dissertation, the \( 2\) -norm (or spectral norm) is primarily used for matrices, which is induced by the Euclidean vector norm. It is defined as \( \|A\|_2 = \sqrt{\lambda_{\max}(A^H A)}\) where \( \lambda_{\max}(A^H A)\) is the largest eigenvalue of \( A^H A\) , and \( A^H\) is the conjugate transpose of \( A\) .

A matrix \( A \in \mathbb{R}^{m \times n}\) is termed left-invertible if there exists a matrix \( B \in \mathbb{R}^{n \times m}\) such that \( BA = \rm{Id}_n\) , where \( \rm{Id}_n\) denotes the \( n \times n\) identity matrix. The matrix \( B\) is referred to as the left inverse of \( A\) , denoted as \( A^*\) . For \( A\) to be left-invertible, it must satisfy the rank condition \( \rm{rank}(A) = n\) , indicating that \( A\) has full column rank and therefore \( m \geq n\) . In this dissertation, we often deal with matrices that vary with inputs, necessitating that this left invertibility is uniform across all inputs. Specifically, the matrix \( u \mapsto A(u) \in \mathbb{R}^{m \times n}\) , defined for an input \( u \in \mathcal{U}\) (where \( \mathcal{U}\) is a possibly unbounded set), is said to be uniformly left-invertible if there exists a constant \( c > 0\) such that:

\[ \begin{equation} (A(u))^\top A(u) \geq c \rm{Id}_n, ~~ \forall u \in \mathcal{U}. \end{equation} \]

(2)

A matrix \( A \in \mathbb{R}^{n \times n}\) is invertible if it is both left-invertible and right-invertible, meaning that there exists a matrix \( B \in \mathbb{R}^{n \times n}\) such that \( BA = AB = \rm{Id}_n\) . The matrix \( B\) is called the inverse of \( A\) , denoted as \( A^{-1}\) . The rank condition for \( A\) to be invertible is \( \rm{rank}(A) = n\) , implying that \( A\) has full rank. The notion of uniform invertibility is also defined using (2), with \( m = n\) .

The Kronecker product of two matrices \( A \in \mathbb{R}^{m \times n}\) and \( B \in \mathbb{R}^{p \times q}\) , denoted as \( A \otimes B\) , is defined as:

\[ \begin{equation} A \otimes B = \begin{bmatrix} a_{11} B & a_{12} B & \cdots & a_{1n} B \\ a_{21} B & a_{22} B & \cdots & a_{2n} B \\ \vdots & \vdots & \ddots & \vdots \\ a_{m1} B & a_{m2} B & \cdots & a_{mn} B \end{bmatrix} \in \mathbb{R}^{mp \times nq}. \end{equation} \]

(3)

Notations. Let \( \sigma_{\min}(A):=\sqrt{\lambda_{\min}(A^\top A)}\) (resp., \( \sigma_{\max}(A):=\sqrt{\lambda_{\max}(A^\top A)}\) ) denote the smallest (resp., largest) singular value of matrix \( A\) , where \( \lambda_{\min}\) (resp., \( \lambda_{\max}\) ) represents the minimal (resp., maximal) eigenvalue. Let \( |\cdot|\) be the Euclidean norm and \( \|\cdot\|\) the induced matrix norm which coincides with \( \sigma_{\max}(\cdot)\) . Denote \( A^\top\) as the transpose of matrix \( A\) and \( A^H\) as its conjugate transpose. Let \( A^\bot\) be the orthogonal complement of matrix \( A\) satisfying \( A^\bot A = 0\) and \( A^\bot (A^\bot)^\top > 0\) , and \( A^\dagger\) be the Moore-Penrose inverse of \( A\)  [37]; if \( A\) is square and invertible, then \( A^\dagger = A^{-1}\) . Let \( \rm{diag}(\lambda_1, \lambda_2, …, \lambda_n)\) be the diagonal matrix with entries \( \lambda_i\) , \( i = 1,2,…,n\) . Denote \( \rm{Id}_n\) (resp., \( 0_{m\times n}\) ) as the \( (n\times n)\) -dimensional identity matrix (resp., \( (m \times n)\) -dimensional zero matrix), with dimensions omitted when contextually clear. Let \( A \otimes B\) denote the Kronecker product of matrices \( A\) and \( B\) . For \( x = (x_1,x_2,x_3) \in \mathbb{R}^3\) , denote

\[ [x]_\times = \begin{pmatrix} 0 & -x_3 & x_2 \\ x_3 & 0 & -x_1 \\ -x_2 & x_1 & 0 \end{pmatrix} \in \mathbb{R}^{3 \times 3}. \]

The symbol \( \star\) in the matrix inequalities denotes the symmetric block, while in some long derivations, \( \star^\top P x\) denotes \( x^\top P x\) . A square matrix is Hurwitz (resp., Schur) if all of its eigenvalues lie strictly on the left-half complex plan (resp., strictly inside the unit ball centered at the origin of the complex plan).

1.5.1.3 Functions

A function (or map\( /\) mapping) \( f: M \to N\) between metric spaces \( (M, d_M)\) and \( (N, d_N)\) is continuous at a point \( x \in M\) if, for every \( \epsilon > 0\) , there exists \( \delta > 0\) such that \( d_M(x, y) < \delta\) implies \( d_N(f(x), f(y)) < \epsilon\) . The function \( f\) is continuous on \( S \subseteq M\) if it is continuous at every point in \( S\) ; if \( S = M\) , it is simply called continuous. If the same \( \delta > 0\) works for every \( \epsilon > 0\) , then \( f\) is uniformly continuous at \( x\) .

A real-valued function \( f\) defined on an interval \( I \subseteq \mathbb{R}\) is absolutely continuous on \( I\) if, for each \( \epsilon > 0\) , there exists \( \delta > 0\) such that for any finite collection of disjoint sub-intervals \( (x_k,y_k)\) of \( I\) with \( x_k < y_k \in I\) , if

\[ \begin{equation} \sum_{k=1}^N(y_k-x_k) < \delta, \end{equation} \]

(4)

then

\[ \begin{equation} \sum_{k=1}^N|f(y_k)-f(x_k)| < \epsilon. \end{equation} \]

(5)

It is locally absolutely continuous on \( I\) if, for every closed sub-interval \( I^\prime \subseteq I\) , the restriction of \( f\) to \( I^\prime\) is absolutely continuous.

Let us now recall comparison functions [38, Section 4.4]. A function \( \alpha: \mathbb{R}_{\geq 0} \to \mathbb{R}_{\geq 0}\) is of class-\( \mathcal{K}\) if it is continuous, strictly increasing, and \( \alpha(0) = 0\) . A class-\( \mathcal{K}\) function belongs to class-\( \mathcal{K}_\infty\) if it is also unbounded, i.e., \( \lim_{r \to +\infty}\alpha(r) = +\infty\) . A function \( \beta: \mathbb{R}_{\geq 0} \times \mathbb{R}_{\geq 0} \to \mathbb{R}_{\geq 0}\) is of class-\( \mathcal{KL}\) if i) for each fixed \( s\) , \( r \mapsto \beta(r,s)\) is of class-\( \mathcal{K}\) , and ii) for each fixed \( r\) , \( s \mapsto \beta(r,s)\) is decreasing and \( \lim_{s \to +\infty}\beta(r,s) = 0\) .

With these comparison functions at hand, we uniformize the notion of continuity as follows. A function \( f\) is \( \mathcal{K}\) -continuous on a compact set \( \mathcal{X}\) if there exists a class-\( \mathcal{K}\) function \( \rho\) such that

\[ \begin{equation} |f(x_a) - f(x_b)| \leq \rho(|x_a - x_b|), ~~ \forall (x_a,x_b) \in \mathcal{X} \times \mathcal{X}. \end{equation} \]

(6)

If \( f\) is continuous on \( \mathcal{X}\) and \( \mathcal{X}\) is compact, then \( f\) is \( \mathcal{K}\) -continuous on this set [39, Lemmas A.6 and A.9]. If \( \rho\) takes the form \( \rho(s) = Ls\) for some \( L > 0\) , then \( f\) is said to be Lipschitz on \( \mathcal{X}\) with Lipschitz constant \( L\) . A function \( f\) is locally Lipschitz on \( \mathcal{X}\) if \( f\) is Lipschitz on each compact subset of \( \mathcal{X}\) , but with possibly different Lipschitz constants depending on the compact subset. In this dissertation, we often consider multi-variable functions that are uniformly continuous with respect to one of its arguments, while uniformity also holds in the rest. For example, if \( f\) also has \( u \in \mathcal{U}\) as its extra argument besides \( x\) , this corresponds to the fact that there exists a class-\( \mathcal{K}\) function \( \rho\) such that

\[ \begin{equation} |f(x_a,u) - f(x_b,u)| \leq \rho(|x_a - x_b|), ~~ \forall (x_a,x_b) \in \mathcal{X} \times \mathcal{X}, \forall u \in \mathcal{U}. \end{equation} \]

(7)

Here, the function \( \rho\) is the same for all \( u \in \mathcal{U}\) , and we say that \( f\) is uniformly continuous in \( x\) on \( \mathcal{X}\) , uniformly in \( u\) on \( \mathcal{U}\) . The same holds for the Lipschitz case when \( f\) is Lipschitz in \( x\) on \( \mathcal{X}\) , uniformly in \( u\) on \( \mathcal{U}\) , and we simply say that \( f\) is uniformly Lipschitz when no confusion is possible.

The notion of injectivity can be seen as a counterpart to continuity, as explained next. A function \( f\) is injective on a set \( \mathcal{X}\) if \( f(x_a) = f(x_b)\) implies \( x_a = x_b\) for all \( (x_a,x_b) \in \mathcal{X} \times \mathcal{X}\) . A function \( f\) is uniformly injective on a set \( \mathcal{X}\) if there exists a class-\( \mathcal{K}\) function \( \rho\) such that

\[ \begin{equation} |f(x_a) - f(x_b)| \geq \rho(|x_a - x_b|), ~~ \forall (x_a,x_b) \in \mathcal{X} \times \mathcal{X}. \end{equation} \]

(8)

Similarly to the above, using [39, Lemmas A.6 and A.9], if \( f\) is injective on \( \mathcal{X}\) and \( \mathcal{X}\) is compact, then \( f\) is uniformly injective on this set. If \( \rho\) takes the form \( \rho(s) = Ls\) for some \( L > 0\) , then \( f\) is said to be Lipschitz injective on \( \mathcal{X}\) . If this Lipschitz injectivity is uniform with respect to the other arguments \( f\) may have, then \( f\) is Lipschitz injective in \( x\) on \( \mathcal{X}\) , uniformly in the other arguments on their respective sets, and we simply say that \( f\) is uniformly Lipschitz injective on \( \mathcal{X}\) when no confusion is possible.

A function \( f: \mathbb{R}^n \to \mathbb{R}^m\) where \( m \geq n\) , which is uniformly injective on \( \mathcal{X} \subseteq \mathbb{R}^n\) , is left-invertible on \( f(\mathcal{X})\) , i.e., there exists a function \( g: f(\mathcal{X}) \to \mathbb{R}^n\) such that \( g(f(x)) = x\) for all \( x \in \mathcal{X}\) . The function \( g\) is called a left inverse of \( f\) on \( f(\mathcal{X})\) , denoted as \( f^*\) . A function \( f: \mathbb{R}^n \to \mathbb{R}^n\) that is uniformly injective on \( \mathcal{X} \subseteq \mathbb{R}^n\) is invertible on \( \mathcal{X}\) , i.e., there exists a function \( g: f(\mathcal{X}) \to \mathbb{R}^n\) such that \( g(f(x)) = x\) for all \( x \in S\) and \( f(g(z)) = z\) for all \( z \in f(\mathcal{X})\) . The function \( g\) is called an inverse of \( f\) on \( f(\mathcal{X})\) , denoted as \( f^{-1}\) . These are more general notions of (left-)inverses than the ones defined using matrices above, which are linear.

A function \( f\) is bounded on a set \( \mathcal{X}\) if there exists \( M \geq 0\) such that

\[ \begin{equation} |f(x)| \leq M, ~~ \forall x \in \mathcal{X}. \end{equation} \]

(9)

A function \( f\) is locally bounded on \( \mathcal{X}\) if \( f\) is bounded on each compact subset of \( \mathcal{X}\) , but with possibly different bounds depending on the compact subset. Similarly to the above, if this boundedness is uniform with respect to the other arguments \( f\) may have, then \( f\) is uniformly bounded on \( \mathcal{X}\) .

A function is continuously differentiable or \( C^1\) if its derivative exists and is continuous. More generally, a function is \( C^k\) if it is continuously differentiable of order \( k \in \mathbb{N}\) , i.e., its derivative up to order \( k\) is continuous. It may happen that \( k = +\infty\) , in which case the function is infinitely differentiable. The gradient of a continuously differentiable function \( V: \mathbb{R}^n \to \mathbb{R}\) at \( x = (x_1,x_2,…,x_n) \in \mathbb{R}^n\) is

\[ \begin{equation} \nabla V(x) = \left(\frac{\partial V}{\partial x_1}(x),\frac{\partial V}{\partial x_2}(x),\ldots,\frac{\partial V}{\partial x_n}(x)\right) \in \mathbb{R}^n, \end{equation} \]

(10)

where \( \frac{\partial V}{\partial x_i}(x)\) , for each \( i \in \{1,2,…,n\}\) , is the partial derivative of \( V\) with respect to \( x_i\) , evaluate at \( x\) .

A set-valued function assigns to each element in the domain a set of elements in the codomain. A set-valued function \( F: \mathbb{R}^n \rightrightarrows \mathbb{R}^m\) is outer semi-continuous (or in some literature, upper semi-continuous) at \( x \in \mathbb{R}^n\) if for any sequence \( (x_n)_{n \in \mathbb{N}}\) converging to \( x\) and any sequence \( (z_n)_{n \in \mathbb{N}}\) where \( z_n \in F(x_n)\) converging to \( z \in \mathbb{R}^m\) , we have \( z \in F(x)\) .

Notations. For two functions \( f\) and \( g\) , \( f \circ g\) is their composition, namely for all \( x\) in the domain of \( g\) , \( g(x)\) is in the domain of \( f\) and \( (f \circ g)(x) = f(g(x))\) . Given a function \( (x,u) \mapsto f(x,u)\) that is locally Lipschitz in \( x\) on \( \mathcal{X}\) , uniformly in \( u\) on \( \mathcal{U}\) , define \( f_\rm{{sat}}\) as a map that is globally Lipschitz in \( x\) , uniformly in \( u\) on \( \mathcal{U}\) , and agrees with \( f\) on \( \mathcal{X} \times \mathcal{U}\) ; such a map is guaranteed to exist by [40, Corollary 1]. Let \( \nabla V(x)\) denote the gradient of a continuously differentiable scalar function \( V\) at \( x\) . Let \( \|\cdot\|_{C^1}\) be the \( C^1\) -norm of \( C^1\) functions on the compact set \( \mathcal{X}\) , defined as \( \|f\|_{C^1} = \max_{x\in\mathcal{X}}|f(x)| + \max_{x\in\mathcal{X}}\left|\frac{\partial f}{\partial x}(x)\right|\) .

1.5.2 Dynamical Systems and Observers in Continuous and Discrete Time

This section introduces the fundamental concepts of dynamical systems in continuous and discrete time. Understanding these systems is crucial for comprehending hybrid systems, which are the main focus of this work. Additionally, we define observers for these dynamical systems.

1.5.2.1 Dynamical Systems with Inputs and Outputs

In continuous time, a dynamical system with input and (measured) output is described by the following equations:

\[ \begin{equation} \dot{x}(t) = f(x(t),u(t)) + v(t), ~~ y(t) = h(x(t),u(t)) + w(t), \end{equation} \]

(11)

where \( x(t) \in \mathbb{R}^{n_x}\) represents the state and \( \dot{x}(t) = \frac{dx}{dt}(t)\) is its time derivative, \( u(t) \in \mathbb{R}^{n_u}\) is the known input, \( y(t) \in \mathbb{R}^{n_y}\) is the measured output, and \( (v(t),w(t))\) are the unknown additive disturbances and measurement noise, respectively, all at continuous time \( t \in \mathbb{R}_{\geq 0}\) . While it is possible to consider a nonlinear dependence on these unknown inputs, doing so may render \( (f,h)\) as functions of \( x\) not well-defined. From now on, when unnecessary, the dependence on time \( t\) is not explicitly written.

Similarly, in discrete time, a dynamical system with input and output is described by the following equations:

\[ \begin{equation} x_{k+1} = f(x_k,u_k) + v_k, ~~ y_k = h(x_k,u_k) + w_k, \end{equation} \]

(12)

where \( x_k \in \mathbb{R}^{n_x}\) , \( u_k \in \mathbb{R}^{n_u}\) , \( y_k \in \mathbb{R}^{n_y}\) , and \( (v_k,w_k)\) represents the state, known input, measured output, and unknown additive disturbances and measurement noise, respectively, all at discrete time \( k \in \mathbb{N}\) .

Systems (11) and system (12) are non-autonomous due to the inputs \( (u(t),v(t))\) (and \( (u_k,v_k)\) ) as functions of time. This representation covers time-varying systems, which is a special case when \( u(t) = t\) (resp., \( u_k = k\) ). When \( (u(t),v(t))\) (resp., \( (u_k,v_k)\) ) are absent, these systems are said to be autonomous. These systems, described by differential and difference equations, are dynamical because their current state depends on the entire history of its past values and inputs. Indeed, initialized as \( x(0)\) and with input trajectories \( t \mapsto (u(t),v(t))\) , any solution to system (11), evaluated at time \( t \in \mathbb{R}_{\geq 0}\) , if defined, verifies

\[ \begin{equation} x(t) = x_0 + \int_{0}^t (f(x(s),u(s)) + v(s))ds. \end{equation} \]

(13)

Similarly, initialized as \( x_0\) and with input trajectories \( k \mapsto (u_k,v_k)\) , any solution to system (12), evaluated at time \( k \in \mathbb{N}\) , if defined, verifies

\[ \begin{equation} x_k = f(f(\ldots (f(x_0,u_0) + v_0)\ldots,u_{k-2}) + v_{k-2},u_{k-1}) + v_{k-1}. \end{equation} \]

(14)

Nonlinear phenomena such as finite-time escape [38, Chapter 1] or solutions leaving the flow set of a state-constrained system [41] can make solutions, for some initial conditions and input trajectories, defined only up to a certain time. Then, a solution \( t \mapsto x(t)\) (resp., \( k \mapsto x_k\) ) is (forward) complete if its time domain is the whole \( \mathbb{R}_{\geq 0}\) (resp., \( \mathbb{N}\) ). In the context of Kravaris-Kazantzis$\slash$Luenberger[KKL] observers considered later in Chapters 2.3 and 4.2, we often need to consider solutions in backward or negative time, leading to the notion of backward completeness of solutions. Note that in this dissertation, exogenous inputs are defined at all times, even though solutions may not be. Next, we denote:

  • \( \mathcal{X}_0 \subseteq \mathbb{R}^{n_x}\) as the set of initial conditions \( x(0)\) (resp., \( x_0\) ) of interest for system (11) (resp., system (12));
  • \( \mathfrak{U}\) as the set of input trajectories \( t \mapsto u(t)\) (resp., \( k \mapsto u_k\) ) of interest for system (11) (resp., system (12));
  • \( \mathcal{U} \subseteq \mathbb{R}^{n_u}\) as the set of the values that \( u(t)\) (resp., \( u_k\) ) can take along a trajectory in \( \mathfrak{U}\) , for \( t \in \mathbb{R}_{\geq 0}\) (resp., \( k \in \mathbb{N}\) );
  • \( \mathcal{V}\) as the set of input trajectories \( t \mapsto v(t)\) (resp., \( k \mapsto v_k\) ) that system (11) (resp., system (12)) is considered to have;
  • \( \mathcal{W}\) as the set of input trajectories \( t \mapsto w(t)\) (resp., \( k \mapsto w_k\) ) that system (11) (resp., system (12)) is considered to have.

The knowledge of \( \mathcal{X}_0\) is usually not needed for observer design but is used in the statement of theoretical results to describe the set of solutions of interest. Here, we clarify the difference between the sets \( \mathfrak{U}\) and \( \mathcal{U}\) for the following reason. Some of the properties of solutions are linked to the considered inputs or even their derivatives, up to a high enough order, as trajectories of time; these include completeness (e.g., only some controlled solutions can be complete), invariance (not escaping from some bounded set), and some observability conditions like instantaneous observability [42, Assumption 3] or Uniform Complete Observability[UCO] [43] (e.g., only solutions subject to some persistently exciting control inputs are observable). On the other hand, some properties such as regularity of the system’s maps or observability conditions such as quadratic detectability (typically used for observer design using a Linear Matrix Inequality[LMI]—see for instance [44]) are point-wise and hence not depending on the history or variation of the inputs. Thus, inputs are classified either based on trajectories using the set \( \mathfrak{U}\) , or values using \( \mathcal{U}\) . With these definitions at hand, we make the following basic assumptions on the solutions to systems (11) and (12) of interest.

Assumption 1

The solutions to system (11) (resp., system (12)) initialized in \( \mathcal{X}_0\) and with inputs in \( (\mathfrak{U},\mathcal{V},\mathcal{W})\) , are complete and remain in \( \mathcal{X} \subseteq \mathbb{R}^{n_x}\) for all \( t \in \mathbb{R}_{\geq 0}\) (resp., \( k \in \mathbb{N}\) ).

Notations. For a given input \( t \mapsto u(t)\) to the dynamics \( \dot{x} = f(x,u)\) that admit unique solutions (e.g., when \( f\) is locally Lipschitz in \( x\) , uniformly in \( u\) ), let \( \Psi_{f(\cdot,u)}(x_0,t,\tau)\) denote the associated flow operator from initial value \( x_0\) at initial time \( t\) evaluated after \( \tau\) time unit(s); based on (13), it verifies

\[ \begin{equation} \Psi_{f(\cdot,u)}(x_0,t,\tau) = x_0 + \int_{t}^{t+\tau} f(x(s),u(s))ds. \end{equation} \]

(15)

The duration \( \tau\) can be negative, in which case we integrate \( f\) backward. If \( f\) is autonomous, then this notation reduces to \( \Psi_f(x_0,\tau)\) . If \( f\) takes the form \( f(x,u) = A(u)x + B(u)\) , then \( \Psi_{f(\cdot,u)}(x_0,t,\tau)\) is linear in \( x_0\) , taking the form

\[ \begin{equation} \Psi_{f(\cdot,u)}(x_0,t,\tau) = \Phi_{A(u)}(t+\tau,t)x_0 + \int_{t}^{t+\tau}\Phi_{A(u)}(t+s,t)B(u(s))ds, \end{equation} \]

(16)

where \( \Phi_{A(u)}(t,t^\prime) \in \mathbb{R}^{n_x \times n_x}\) is the state transition matrix from time \( t^\prime \in \mathbb{R}\) to time \( t \in \mathbb{R}\) of the system \( \dot{x} = A(u) x\) , such that solutions to this system verify \( x(t) = \Phi_{A(u)}(t,t^\prime) x(t^\prime)\) and such that \( \Phi(t,t) = \rm{Id}\) for all \( t \in \mathbb{R}\) . If moreover \( f\) is Linear Time-Invariant[LTI], i.e., \( f(x) = Ax + B\) for some constant matrices \( (A,B)\) , then

\[ \begin{equation} \Psi_f(x_0,\tau) = e^{A\tau}x_0 + \int_{0}^{\tau}e^{A(\tau-s)}Bds, \end{equation} \]

(17)

i.e., as (16) with \( \Phi_A(\tau) = e^{A\tau}\) , which is now time-invariant, where the matrix exponential is \( e^{A} = \displaystyle\sum_{k=0}^{+\infty} \frac{(A)^k}{k!} \in \mathbb{R}^{n_x \times n_x}\) for \( A \in \mathbb{R}^{n_x \times n_x}\) (this expression is still true if \( A\) has complex components).

1.5.2.2 Observer Design for Continuous- and Discrete-Time Systems

In this section, we introduce observers for continuous- and discrete-time systems and discuss their properties. Observers, as illustrated in Figure 7, are algorithms designed to estimate the states of dynamical systems from their known inputs and measured outputs. In continuous time, an observer for system (11) takes the general form

\[ \begin{equation} \dot{\hat{z}} = \mathcal{F}(\hat{z},y,u,t), ~~ \hat{x} = \Upsilon(\hat{z},y,u,t), \end{equation} \]

(18)

where \( \hat{z} \in \mathbb{R}^{n_z}\) is the observer state. Similarly, for system (11) in discrete time, a general observer is

\[ \begin{equation} \hat{z}_{k+1} = \mathcal{F}(\hat{z}_k,y_k,u_k,k), ~~ \hat{x}_k = \Upsilon(\hat{z}_k,y_k,u_k,k). \end{equation} \]

(19)

Designing an observer involves determining the functions \( (\mathcal{F},\Upsilon)\) and an initialization set \( \mathcal{Z}_0 \subseteq \mathbb{R}^{n_z}\) such that the estimation error satisfies certain desired properties. The types of observers thus depend on these properties, namely observer (18) (resp., observer (19)), is an

  • Asymptotic observer, if in the absence of disturbance \( t \mapsto v(t)\) (resp., \( k \mapsto v_k\) ) and noise \( t \mapsto w(t)\) (resp., \( k \mapsto w_k\) ), for any solutions to system (11) (resp., system (12)) initialized in \( \mathcal{X}_0\) with \( t \mapsto u(t) \in \mathfrak{U}\) (resp., \( k \mapsto u_k \in \mathfrak{U}\) ) and to observer (18) (resp., observer (19)) initialized in \( \mathcal{Z}_0\) and fed with \( t \mapsto (u(t),y(t))\) (resp., \( k \mapsto (u_k,y_k)\) ), we have

    \[ \begin{equation} \lim_{t \to +\infty} |x(t) - \hat{x}(t)| = 0, \text{~ or resp., ~} \lim_{k \to +\infty} |x_k - \hat{x}_k| = 0. \end{equation} \]

    (20)

    Note that this property and the next ones require the system’s solutions of interest to be complete, which is guaranteed by Assumption 1;

  • Asymptotically stable observer, if there exists a class-\( \mathcal{KL}\) function \( \beta\) such that in the absence of disturbance \( t \mapsto v(t)\) (resp., \( k \mapsto v_k\) ) and noise \( t \mapsto w(t)\) (resp., \( k \mapsto w_k\) ), for any solutions to system (11) (resp., system (12)) initialized in \( \mathcal{X}_0\) with \( t \mapsto u(t) \in \mathfrak{U}\) (resp., \( k \mapsto u_k \in \mathfrak{U}\) ) and to observer (18) (resp., observer (19)) initialized in \( \mathcal{Z}_0\) and fed with \( t \mapsto (u(t),y(t))\) (resp., \( k \mapsto (u_k,y_k)\) ), we have

    \[ \begin{multline} |x(t) - \hat{x}(t)| \leq \beta(|x(0) - \hat{x}(0)|,t), ~~ \forall t \geq 0, \\ \text{~ or resp., ~} |x_k - \hat{x}_k| \leq \beta(|x_0 - \hat{x}_0|,k), ~~ \forall k \in \mathbb{N}. \end{multline} \]

    (21)

    These properties ensure not only asymptotic convergence but also the stability of the estimation error with respect to its initial condition. Specifically, the estimation error remains arbitrarily small at all times if its initial value is small enough. In certain contexts, especially in discrete time, (21) can be stated with respect to the initial condition but only holds from a certain time, i.e.,

    \[ \begin{multline} |x(t) - \hat{x}(t)| \leq \beta(|x(0) - \hat{x}(0)|,t), ~~ \forall t \geq t_0, \\ \text{~ or resp., ~} |x_k - \hat{x}_k| \leq \beta(|x_0 - \hat{x}_0|,k), ~~ \forall k \in \mathbb{N}_{\geq k_0}, \end{multline} \]

    (22)

    for some certain \( t_0 \geq 0\) (resp., \( k_0 \in \mathbb{N}\) ), then we say that the observer is asymptotically stable (with respect to its initial condition) after a certain time. If the same map \( \beta\) can be used to express the stability of the estimation error with respect to its value at any time in addition to the one at the initial time, namely

    \[ \begin{multline} |x(t) - \hat{x}(t)| \leq \beta(|x(t_0) - \hat{x}(t_0)|,t-t_0), ~~ \forall t \geq t_0, \\ \text{~ or resp., ~} |x_k - \hat{x}_k| \leq \beta(|x_{k_0} - \hat{x}_{k_0}|,k-k_0), ~~ \forall k \in \mathbb{N}_{\geq k_0}. \end{multline} \]

    (23)

    for any \( t_0 \geq 0\) (resp., \( k_0 \in \mathbb{N}\) ), then this observer is uniformly asymptotically stable. If the function \( \beta\) takes the form \( \beta(r,s) = c r e^{-\lambda s}\) for some \( c \geq 0\) and \( \lambda > 0\) , then this observer is exponentially stable. Similarly to the above, if the parameters \( (c,\lambda)\) are uniform with respect to treating the estimation error at any time as the new initial condition, then the observer is uniformly exponentially stable. In this dissertation, we frequently achieve high-gain results, where the observer can be designed to ensure any desired convergence rate \( \lambda > 0\) for the estimation error. However, a large \( \lambda\) (indicating fast convergence) may result in a larger \( c\) , which means that the estimation error is amplified during the transient; this is known as the peaking phenomenon. Such an observer is said to be arbitrarily fast;

  • ISS observer, if there exist a class-\( \mathcal{KL}\) function \( \beta\) and a class-\( \mathcal{K}_\infty\) function \( \alpha\) such that in the presence of disturbance \( t \mapsto v(t)\) (resp., \( k \mapsto v_k\) ) in \( \mathcal{V}\) and noise \( t \mapsto w(t)\) (resp., \( k \mapsto w_k\) ) in \( \mathcal{W}\) , for any solutions to system (11) (resp., system (12)) initialized in \( \mathcal{X}_0\) with \( t \mapsto u(t) \in \mathfrak{U}\) (resp., \( k \mapsto u_k \in \mathfrak{U}\) ) and to observer (18) (resp., observer (19)) initialized in \( \mathcal{Z}_0\) and fed with \( t \mapsto (u(t),y(t))\) (resp., \( k \mapsto (u_k,y_k)\) ), we have

    \[ \begin{multline} |x(t) - \hat{x}(t)| \leq \beta(|x(0) - \hat{x}(0)|,t) + \alpha\left(\sup_{s \in [0,t]}|v(s)| + \sup_{s \in [0,t]}|w(s)|\right), ~ \forall t \geq 0, \\ \text{or resp., } |x_k - \hat{x}_k| \leq \beta(|x_0 - \hat{x}_0|,k) + \alpha\left(\sup_{s \in \{0,1,\ldots,k\}}|v_s| + \sup_{s \in \{0,1,\ldots,k\}}|w_s|\right), ~ \forall k \in \mathbb{N}. \end{multline} \]

    (24)

    Condition (24) is based on the traditional definition of Input-to-State Stability[ISS] as presented in [45]. This condition, typically achieved when the estimation error is exponentially stable, implies that the estimation error converges asymptotically to a neighborhood around zero, where the size of this neighborhood is determined by the magnitude of the disturbance and noise. A more recent definition, referred to in this dissertation as robust stability, which introduces a forgetting factor to penalize past disturbances and noise, is proposed in [46]. This approach refines the original ISS definition by replacing (24) with the following conditions:

    \[ \begin{multline} |x(t) - \hat{x}(t)| \leq \beta(|x(0) - \hat{x}(0)|,t) + \sup_{s\in[0,t]}\beta_v(|v(s)|,t-s)+\sup_{s\in[0,t]}\beta_w(|w(s)|,t-s), ~ \forall t \geq 0, \\ \text{or resp., } \\ |x_k - \hat{x}_k| \leq \beta(|x_0 - \hat{x}_0|,k) + \max_{q\in\{0,1,\ldots,k-1\}}\beta_v(|v_q|,k-q-1)+\max_{q\in\{0,1,\ldots,k-1\}}\beta_w(|w_q|,k-q-1), ~ \forall k \in \mathbb{N}, \end{multline} \]

    (25)

    for some class-\( \mathcal{KL}\) functions \( \beta_v\) and \( \beta_w\) . This property is utilized throughout this dissertation whenever it is achieved.

There are other types of observers such as finite-time observers (where the estimation error becomes zero within a finite time in the absence of disturbances and noise) of which a case is sliding mode observers [47], interval observers [48, 49], etc.

In general, observer design involves transforming the given system into a suitable target form in a new set of coordinates, known as the \( z\) -coordinates, for which we know how to design an observer (via the map \( \mathcal{F}\) of observers (18) and (19)), giving us some good properties in these new coordinates such as convergence or stability as discussed earlier, and then we recover the estimate in the original \( x\) -coordinates (via the map \( \Upsilon\) of observers (18) and (19)), relying on an injectivity property of the transformation. The stronger the injectivity of this transformation in terms of uniformity, the more robust the properties that can be transferred from the \( z\) -coordinates to the \( x\) -coordinates. For instance, to recover asymptotic convergence in the \( x\) -coordinates, the transformation only needs to be injective uniformly in time. However, to bring back (arbitrarily fast) exponential stability or robustness of the estimation error, the transformation must be Lipschitz injective uniformly with respect to time. This general idea of nonlinear observer design is recapped in [39] for continuous-time systems. Notice from observers (18) and (19) that the transformation from the \( x\) - to the \( z\) -coordinates is not used for observer implementation; it allows us to design the observer maps and thus plays a role in the theoretical analysis only. Therefore, in some contexts such as a discrete-time Moving Horizon Observer[MHO] [50], this transformation may be implicit and we only need to know a way to recover the estimate from an appropriately designed observer in the \( z\) -coordinates consisting of storing the past outputs; this idea is elaborated in Theorem 1. For nonlinear observers, the target \( z\) -coordinates satisfy \( n_z \geq n_x\) , with equality holding for linear observers. The case \( n_z < n_x\) corresponds to a special class called reduced-order observers that is not considered in this dissertation.

Now, we briefly discuss a very important continuous-time observer called the high-gain observer. This design comes in two primary versions: the classical high-gain observer [26] and the Kalman-like high-gain observer [27]. Let us discuss the former. For brevity, we consider the case where the continuous-time system (11) is autonomous, (i.e., \( u\) is absent) and has a one-dimensional output (i.e., \( n_y = 1\) ). These designs require this system to be instantaneously observable (or in some literature, differentially observable[51, Definition 1.2], implying that there exists an injective immersion into new \( z\) -coordinates of the observable canonical form [52, 1]

\[ \begin{equation} \dot{z} = Az + B\varphi(z) + \nu, ~~ y = Cz + w, \end{equation} \]

(26)

where \( \nu\) is an image of \( v\) through the immersion that typically depends on \( z\) , \( \varphi\) is a locally Lipschitz function, and the matrices

\[ A = \begin{pmatrix} 0_{(n_z - 1)\times 1} & \rm{Id}_{n_z - 1} \\ 0 & 0_{1\times (n_z - 1)} \end{pmatrix}, ~~ B = \begin{pmatrix} 0_{(n_z - 1)\times 1} \\ 1 \end{pmatrix}, ~~ C = \begin{pmatrix} 1 & 0_{1\times (n_z - 1)} \end{pmatrix}. \]

System (26) with such an \( A\) matrix has a canonical form and is always observable. A classical high-gain observer for system (26) is given by

\[ \begin{equation} \dot{\hat{z}} = A\hat{z} + B\varphi_{\rm{sat}}(\hat{z}) + D(\ell)K(y - C\hat{z}), \end{equation} \]

(27)

where \( \varphi_\rm{{sat}}\) is obtained from \( \varphi\) as defined in Section 1.5.1.3, \( K = (k_1,k_2,…,k_{n_z})\) is chosen such that \( A - KC\) is Hurwitz (defined in Section 1.5.1.2) and \( D(\ell) = \rm{diag}(\ell,\ell^2,…,\ell^{n_z})\) , with \( \ell > 0\) an observer parameter to be chosen. Because \( \varphi\) is locally Lipschitz, with such a \( \varphi_\rm{{sat}}\) , we deduce that there exists \( L > 0\) such that

\[ \begin{equation} |\varphi(z) - \varphi_{\rm{sat}}(\hat{z})| \leq L|z - \hat{z}|, ~~ \forall (z,\hat{z}) \in \mathcal{Z} \times \mathbb{R}^{n_z}, \end{equation} \]

(28)

where \( \mathcal{Z}\) is the image of \( \mathcal{X}\) in Assumption 1 via the immersion, which is compact when \( \mathcal{X}\) is compact. The following lemma, whose proof can be found in works such as [53], establishes the fundamental properties of the classical high-gain observer.

Lemma 1 (Classical high-gain observer in continuous time)

Consider system (26) and observer (27). Let \( K\) be such that \( A - KC\) is Hurwitz. There exist \( \ell_1^\star > 0\) , \( \lambda>0\) , \( \overline{b}_o>0\) , and rational positive functions \( \underline{b}_o\) , \( b_\nu\) , \( b_w\) such that for all \( \ell > \ell^\star_1\) , there exists a function \( V_\ell: \mathbb{R}^{n_z}\times \mathbb{R}^{n_z}\) such that:

  1. For all \( (u,z,\hat{z}) \in \mathcal{U} \times \mathbb{R}^{n_z} \times \mathbb{R}^{n_z}\) ,

    \[ \underline{b}_o(\ell) |z-\hat{z}|^2 \leq V_\ell(z,\hat{z}) \leq \overline{b}_o |z - \hat{z}|^2; \]
  2. For all \( (u,z,\hat{z},\nu,w) \in \mathcal{U} \times \mathcal{Z} \times \mathbb{R}^{n_z}\times \mathbb{R}^{n_z} \times \mathbb{R}\) ,

    \[ \begin{multline*} \langle \nabla V_\ell(z,\hat{z}), (A-D(\ell)KC)(z-\hat{z}) + B(\varphi(z) - \varphi_{\rm{sat}}(\hat{z})) + \nu - D(\ell)Kw\rangle \\ \leq -\ell \lambda V_\ell(z,\hat{z})+b_\nu(\ell)|\nu|^2 + b_w(\ell) |w|^2. \end{multline*} \]

From Lemma 1, we deduce that if \( \ell\) in observer (27) is chosen sufficiently large, there exist rational positive functions \( c_\nu\) and \( c_w\) such that

\[ \begin{equation} |z(t)-\hat{z}(t)| \leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}}e^{-\frac{\ell \lambda}{2} t}|z(0)-\hat{z}(0)| + c_\nu(\ell) \sup_{s \in [0,t]}|\nu(s)| + c_w(\ell) \sup_{s \in [0,t]}|w(s)|, ~~ \forall t \geq 0. \end{equation} \]

(29)

Similarly, there is a Kalman-like version of the high-gain observer for systems of the form (26) whose \( A\) matrix depends on \( (u,y)\) still following a canonical form—see [27]. In short, it uses a time-varying \( \mathbb{R}^{n_z} \times \mathbb{R}^{n_z}\) covariance matrix and provides, under the same observability, the same results as in Lemma 1 and (29). Two crucial properties of high-gain observers can thus be highlighted:

  1. In the absence of disturbance \( \nu\) and noise \( w\) , the estimation error is exponentially stable, with an arbitrarily fast convergence rate achievable by increasing \( \ell\) . This property is achieved thanks to system (26) being observable over any arbitrarily short time interval, which is the strongest observability condition that exists in continuous time and not discrete time. However, making \( \ell\) large also reduces \( b_o(\ell)\) , potentially leading to large transient errors, a phenomenon known as peaking that is common in high-gain designs;
  2. In the presence of \( (\nu,w)\) , the estimation is ISS and even robustly stable, with gains that can be amplified as \( \ell\) increases (for more details on how \( (c_\nu(\ell),c_w(\ell))\) vary with \( \ell\) , see [54]).

These properties, including at least asymptotic convergence, may then be brought back to the \( x\) -coordinates thanks to the “sufficient” injectivity of the transformation, as explained above. In this dissertation, for hybrid systems with known jump times, we use a combination of observers to estimate different parts of the state: a high-gain observer for the instantaneously observable part and another observer for the remaining components. This combination is facilitated by Lyapunov analysis, with the two properties of high-gain observers being critical for the analysis—see Section 3.3. Some other variations of high-gain designs include the low-power high-gain observers [52] or sliding mode high-gain observers [47].

Observers for nonlinear discrete-time systems are reviewed in greater detail in Section 2.1.

1.5.3 Hybrid Systems

In this section, we introduce hybrid systems with their modeling, explain their scope and applications, and discuss some of their properties related to observer design.

1.5.3.1 Definitions

A hybrid (dynamical) system with input and output is described by

\[ \begin{equation} \dot{x} \in F(x,u_c) ~ (x,u_c) \in C ~ y_c = h_c(x,u_c), ~ ~ x^+ \in G(x,u_d) ~ (x,u_d) \in D ~ y_d = h_d(x,u_d), \end{equation} \]

(30)

where \( x \in \mathbb{R}^{n_x}\) is the state, \( u_c \in \mathbb{R}^{n_{u,c}}\) (resp., \( u_d \in \mathbb{R}^{n_{u,d}}\) ) is the exogenous flow (resp., jump) input, \( y_c \in \mathbb{R}^{n_{y,c}}\) (resp., \( y_d \in \mathbb{R}^{n_{y,d}}\) ) is the flow (resp., jump) output, \( C \subseteq \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}}\) (resp., \( D \subseteq \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,d}}\) ) is the flow (resp., jump) set, \( F:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \rightrightarrows \mathbb{R}^{n_x}\) (resp., \( G:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,d}} \rightrightarrows \mathbb{R}^{n_x}\) ) is the set-valued flow (resp., jump) map (often called inclusions), and \( h_c:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \rightrightarrows \mathbb{R}^{n_{y,c}}\) (resp., \( h_d:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \rightrightarrows \mathbb{R}^{n_{y,d}}\) ) is the flow (resp., jump) output map. As further clarified in Definition 3, the flow input \( u_c\) is a continuous-time signal, while the jump input \( u_d\) is a sequence of values. The state \( x\) of system (30) evolves as a continuous function of time following \( F\) when itself and some flow input \( u_c\) remain in \( C\) , and it jumps following \( G\) when itself and some jump input \( u_d\) belong to the set \( D\) . This state \( x\) thus exhibits both continuous- and discrete-time behavior, resulting in the unique properties of hybrid systems described in Section 1.5.3.2. When the inclusions consist of exactly one element (i.e., singletons), (30) becomes

\[ \begin{equation} \dot{x} = f(x,u_c) ~ (x,u_c) \in C ~ y_c = h_c(x,u_c), ~ ~ x^+ = g(x,u_d) ~ (x,u_d) \in D ~ y_d = h_d(x,u_d), \end{equation} \]

(31)

with flow and jump maps \( f:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \to \mathbb{R}^{n_x}\) and \( g: \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,d}} \to \mathbb{R}^{n_x}\) , respectively. While some classes of systems require inclusions, e.g., those with timers (see Example 16) or those encountered when describing friction [55], the majority of hybrid systems encountered in this dissertation take the form of (31) and we henceforth lay special emphasis on it. An autonomous hybrid system (30) (resp., (31)), i.e., with \( (u_c,u_d)\) absent, is shortly called a hybrid system with data \( (C,F,D,G)\) (resp., \( (C,f,D,g)\) ). Assumption 2 lists the hybrid basic conditions that ensure a hybrid system is well-posed.

Assumption 2

[Hybrid basic assumptions [10]] The hybrid system with data \( (C, F, D, G)\) satisfies:

  1. \( C\) and \( D\) are closed sets in \( \mathbb{R}^{n_x}\) ;
  2. \( F: \mathbb{R}^{n_x} \rightrightarrows \mathbb{R}^{n_x}\) is an outer semi-continuous set-valued map, locally bounded on \( C\) , and such that \( F(x)\) is non-empty and convex for each \( x \in C\) ;
  3. \( G: \mathbb{R}^{n_x} \rightrightarrows \mathbb{R}^{n_x}\) is an outer semi-continuous set-valued map, locally bounded on \( D\) , and such that \( F(x)\) is non-empty for each \( x \in D\) .

To better illustrate hybrid systems, we introduce the following classical academic example that is revisited throughout this dissertation in various configurations.

Example 1 (Bouncing ball)

Consider a ball bouncing vertically on a flat surface, with position denoted as \( x_1\) and velocity as \( x_2\) . While this ball is in the air (i.e., when \( x_1 \geq 0\) ), its flow dynamics read \( \dot{x}_1 = x_2\) and \( \dot{x}_2 = -c_fx_2^{d_f}-a_g\) where \( a_g = 9.81\) m\( /\) s\( ^2\) is the gravitational acceleration. The constants \( c_f > 0\) (in s\( ^{d_f-2}\) m\( ^{1-d_f}\) ) and \( d_f > 0\) model the nonlinear friction when the ball is in the air. When this ball hits the ground (when \( x_1 = 0\) and \( x_2 \leq 0\) ), its post-impact position remains zero as \( x_1^+ = x_1\) and its velocity reverses direction and may decrease in magnitude according to \( x_2^+ = - c_r x_2 + u_d\) , where \( c_r \in (0,1]\) is a restitution coefficient and \( u_d\) is an impulse input applied at each jump. We measure the position \( x_1\) at all times using a camera. This system can be modeled as a hybrid system of the form (31):

\[ \begin{align} \begin{pmatrix} \dot{x}_1 \\ \dot{x}_2 \end{pmatrix} & = \begin{pmatrix} x_2 \\ c_fx_2^{d_f}-a_g \end{pmatrix} & x& \in \{x \in \mathbb{R}^2: x_1 \geq 0\}, & y_c &= x_1, \\ \begin{pmatrix} x_1^+ \\ x_2^+ \end{pmatrix} & = \begin{pmatrix} x_1 \\ -c_r x_2 + u_d\end{pmatrix} & x &\in \{x \in \mathbb{R}^2: x_1 = 0, x_2 \leq 0\}, & y_d &= x_1. \\\end{align} \]

(32.a)

In this scenario, jumps can be detected when the measurement equals zero, making this a hybrid system with known jump times.

Hybrid systems, as described by (30), provide a framework capable of expressing both continuous- and discrete-time behavior. Besides systems that require a hybrid representation such as the one given by (30) or an equivalent formulation as described later (examples include mechanical systems with impacts such as the bouncing ball in Example 1 and walking robots as in Section 3.4, or the stick-slip behavior studied in Section 4.3), the versatility of hybrid systems allows them to encompass a wide variety of system classes, including:

  • Ordinary continuous- and discrete-time systems (11)\( /\) (12), possibly with constraints: These are straightforward special cases of (30), occurring when \( D = \emptyset\) and \( C = \mathbb{R}^{n_x}\) , and vice-versa. The general formulation can also handle systems with constraints [41] by replacing \( \mathbb{R}^{n_x}\) with a bounded subset;
  • Impulsive differential equations [56]: Impulses are modeled as jumps in the solution, occurring at specific times dictated either by an exogenous input or a state-dependent variable;
  • Switched systems [57]\( /\) hybrid automata [13, 58]: These are continuous-time systems where the state evolves continuously but the derivatives change discontinuously across different modes. For the former, exogenous modes are incorporated into the flow input \( u_c\) . For the latter where the mode changes depending on the state, to fall into the scope of (30), the system adds a discrete state \( q\) with \( \dot{q} = 0\) , incorporating the mode-switching dynamics in \( G\) or \( g\) , and uses \( C\) and \( D\) for switching conditions;
  • Continuous-time systems with sporadic (multi-rate) measurements: These systems are effectively modeled with \( G\) or \( g\) being an identity map and using a timer for each output, to account for the intermittent nature of the measurements. Each timer counts down to zero; when it hits zero, the corresponding output is taken and this timer resets to a certain value modeling the measurement rate. See Example 16 for a case study.

Given the broad range of systems that hybrid models can represent, observers designed for hybrid systems can be applied across these classes and this idea is illustrated using examples throughout this dissertation. Variations of the model (30) that are other formulations of general hybrid systems include:

  • Switched systems with state jumps [59, 60]: Normally, solutions to switched systems, e.g., in [57] are continuous (but not their derivatives), but by allowing the state to jump at the switches triggered by an input, we may cover most of the systems covered by (30). This form is particularly suited to describing systems where jumps are triggered externally and time-dependently, rather than by the system state;
  • Systems with modes, guard sets, and reset maps [61, 62, 63]: This representation differs from (30) in the sense that they separate the state into components that evolve continuously in time and those that switch (called the modes). It is thus more appropriate for systems where the states can be distinctly divided in this manner, as opposed to hybrid systems like the bouncing ball in Example 1, where the hybrid behavior is more intertwined.

Let us now return to the forms (30)\( /\) (31), and introduce the concept of hybrid time domains and solutions. While the time domain for solutions to continuous-time systems is typically \( \mathbb{R}_{\geq 0}\) and for discrete-time systems is \( \mathbb{N}\) , or subsets of these in the case of incomplete solutions, the hybrid nature of hybrid systems necessitates defining solutions on two-component hybrid time domains, as follows.

Definition 1 ((Compact) hybrid time domain [10])

A set \( E\subseteq\mathbb{R}_{\geq 0}\times\mathbb{N}\) is a compact hybrid time domain if (when not empty),

\[ \begin{equation} E = \bigcup_{j=0}^{J-1}\left([t_{j},t_{j+1}]\times \{j\}\right), \end{equation} \]

(33)

for some \( J \in \mathbb{N}\) and a finite sequence of times \( 0=t_0\leq … \leq t_{J - 1} \leq t_J\) in \( \mathbb{R}_{\geq 0}\) . A set \( E\subset \mathbb{R}_{\geq 0} \times \mathbb{N}\) is a hybrid time domain if it is the union of a non-decreasing sequence of compact hybrid time domains, namely, \( E\) is the union of compact hybrid time domains \( E_j\) such that \( E_0 \subset E_1 … \subset E_j \subset…\) .

We see from Definition 1 that each jump of the solution is instantaneous, i.e., it lasts zero units of ordinary time. This definition is used throughout this dissertation. In Section 4.2, for analysis purposes, we consider hybrid solutions in backward time and define the backward version of hybrid time domains, thereby extending Definition 1.

Given an initial condition and input trajectories, a solution to a hybrid system, defined formally in Definition 3, is described as a hybrid arc \( x: \mathbb{R}_{\geq 0} \times \mathbb{N} \to \mathbb{R}^{n_x}\) defined on a hybrid time domain, as follows.

Definition 2 (Hybrid arc [10])

A function \( \rm{dom} x \to \mathbb{R}^{n_x}\) , defined on a hybrid time domain \( \rm{dom} x\) , is a hybrid arc if \( x(\cdot,j)\) is locally absolutely continuous for each \( j\) .

Notations. For a hybrid arc \( (t,j) \mapsto x(t,j)\) , we denote \( \rm{dom} x\) its domain [10, Definition 2.6], \( \rm{dom}_tx\) (resp., \( \rm{dom}_j x\) ) the domain’s projection on the ordinary time (resp., jump) component, for \( j \in \mathbb{N}\) , \( t_j(x)\) the unique time such that \( (t_j(x),j)\in \rm{dom} x\) and \( (t_j(x),j-1) \in \rm{dom} x\) , and \( \mathcal{T}_j(x):=\{t\in \rm{dom}_t(x):(t,j)\in \rm{dom} x\}\) the \( j\) -th flow interval. The mention of \( x\) is omitted when no confusion is possible. Given a hybrid arc \( \phi\) defined on \( \rm{dom} \phi\) , let \( \phi_{|_{\mathcal{D}}}\) be the restriction of \( \phi\) to \( \mathcal{D} \subset \rm{dom} \phi\) . Let \( X(x,t_{j}(x),j)\) denote the solution to an autonomous hybrid system initialized as \( x\) and at hybrid time \( (t_{j}(x),j)\) that can be negative (i.e., after its \( j\) -th jump)—see about backward hybrid systems in Section 4.2. The graph of the hybrid arc \( x\) is the set

\[ \begin{equation} \rm{graph} x = \{(t,j,\xi)\in \mathbb{R}_{\geq 0} \times \mathbb{N} \times \mathbb{R}^{n_x}, (t,j) \in \rm{dom} x, \xi = x(t,j)\} \subseteq \mathbb{R}_{\geq 0} \times \mathbb{N} \times \mathbb{R}^{n_x}. \end{equation} \]

(34)

With these tools at hand, we now define solutions to system (31). This definition extends [10, Definition 2.6], which only covers the autonomous case, by incorporating exogenous flow and jump inputs seen as continuous- and discrete-time signals respectively, and it is used throughout this dissertation. In the literature, definitions of solutions to hybrid systems with inputs are given in [64, 65, 66] for other classes of inputs.

Definition 3 (Solutions to system (31))

The hybrid arc \( (t,j) \mapsto x(t,j)\) is solution to system (31) with flow input \( t \mapsto u_c(t)\) defined on \( \mathbb{R}_{\geq 0}\) and jump input \( (u_{d,j})_{j \in \mathbb{N}}\) if:

  • \( (x(0,0),u_c(0)) \in \rm{cl}(C)\) or \( (x(0,0),u_{d,0}) \in D\) ;
  • For all \( j \in \rm{dom}_j x\) such that \( \rm{int}(\mathcal{T}_j(x)) \neq \emptyset\) , we have \( (x(t,j),u_c(t)) \in C\) for all \( t \in \rm{int}(\mathcal{T}_j(x))\) and \( \dot{x}(t,j) = f(x(t,j),u_c(t))\) for almost all \( t \in \mathcal{T}_j(x)\) ;
  • For all \( (t,j) \in \rm{dom} x\) such that \( (t,j+1) \in \rm{dom} x\) , we have \( (x(t,j),u_{d,j}) \in D\) and \( x(t,j+1) = g(x(t,j),u_{d,j})\) .

Now, let us define the terms related to the types of hybrid solutions found in this dissertation.

Definition 4 (Types of hybrid solutions [10])

A solution \( x\) to system (31) with inputs \( t \mapsto u_c(t)\) defined on \( \mathbb{R}_{\geq 0}\) and \( (u_{d,j})_{j \in \mathbb{N}}\) is:

  1. Non-trivial, if \( \rm{dom} x\) contains at least two points;
  2. Maximal, if there does not exist another solution \( x^\prime\) to system (31) with the same inputs \( (u_c,u_d)\) , such that \( \rm{dom} x\) is a proper subset of \( \rm{dom} x^\prime\) and \( x(t,j) = x^\prime(t,j)\) for all \( (t,j) \in \rm{dom} x\) ;
  3. Complete, if it is maximal and \( \rm{dom} x\) is unbounded;
  4. \( t\) -complete (resp., \( j\) -complete), if it is complete and \( \sup \rm{dom}_t x = +\infty\) (resp., \( \sup \rm{dom}_j x = +\infty\) );
  5. Zeno, if it is \( j\) -complete and \( \sup \rm{dom}_t x < +\infty\) .

Note that in Definition 3, similar to the case of continuous- and discrete-time systems in Section 1.5.2.1, the hybrid arc can stop being defined after some time (i.e., being incomplete), but the inputs are given and complete, i.e., they are defined at all times.

Now we introduce some key notions and tools that are useful for hybrid observer design later.

Definition 5 (Uniform Global pre-Asymptotic Stability[UGpAS] [10])

A closed set \( S \subseteq \mathbb{R}^{n_x}\) is said to be uniformly globally pre-asymptotically stable (UGpAS) for system (30)\( /\) (31) if:

  • (Stability) There exists a class-\( \mathcal{K}_\infty\) function \( \rho\) such that any solution \( x\) to system (30)\( /\) (31) satisfies \( d_S(x(t,j)) \leq \rho(d_S(x(0,0)))\) for all \( (t,j) \in \rm{dom} x\) ;
  • (Pre-attractivity) For each \( \epsilon > 0\) and \( r > 0\) , there exists \( T > 0\) such that, for any solution \( x\) to system (30)\( /\) (31) with \( d_S(x(0,0)) \leq r\) , \( (t,j) \in \rm{dom} x\) and \( t+j \geq T\) imply \( d_S(x(t,j)) \leq \epsilon\) .

Using comparison functions as done when defining asymptotic stability in continuous and discrete time above, we deduce that Definition 5 is equivalent to the existence of a class-\( \mathcal{KL}\) function \( \beta\) such that for any solution \( x\) to system (30)\( /\) (31), we have

\[ \begin{equation} d_S(x(t,j)) \leq \beta(d_S(x(0,0)),t+j), ~~ \forall (t,j) \in \rm{dom} x. \end{equation} \]

(35)

This latter definition is used throughout this dissertation. The consideration of sets generalizes the notions of asymptotic stability introduced in Section 1.5.2.2, which is particularly relevant for hybrid systems with unknown jump times considered in Section 4. The term pre-attractive, as opposed to attractive, highlights the possibility of a maximal solution that is not complete, though it may be bounded. This distinction separates conditions for completeness (closely related to existence) from those for stability and attractivity, and asymptotic convergence holds when \( t+j\) tends to infinity, i.e., a complete solution. If solutions to system (30)\( /\) (31) are complete, then we omit the prefix “pre-” and get Uniform Global Asymptotic Stability[UGAS] of the set \( S\) for system (30)\( /\) (31). Similarly to Section 1.5.2.2, if the function \( \beta\) takes the form \( \beta(r,s) = c r e^{-\lambda s}\) for some \( c \geq 0\) and \( \lambda > 0\) , then we achieve Uniform Global pre-Exponential Stability[UGpES]; if \( \lambda\) can be made arbitrarily large, then the convergence is said to be arbitrarily fast.

The following Lyapunov conditions for hybrid systems, inspired from [10], which are the sufficient conditions for the UGpAS of the set \( S\) for system (31), provide a valuable tool for stability analysis and are utilized throughout this dissertation, possibly in slightly different variations.

Lemma 2 (Lyapunov conditions for UGpAS)

A closed set \( S \subseteq \mathbb{R}^{n_x}\) is UGpAS for system (31) if there exists a function \( V: \mathbb{R}^{n_x} \to \mathbb{R}_{\geq 0}\) that is \( C^1\) on \( C \cup D\) and class-\( \mathcal{K}_\infty\) functions \( \underline{\alpha}\) , \( \overline{\alpha}\) , \( \alpha_c\) , and \( \alpha_d\) such that:

  • (Boundedness) For all \( x \in \mathbb{R}^{n_x}\) and \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) such that \( (x,u_c) \in C\) or \( (x,u_d) \in D\) ,

    \[ \begin{equation} \underline{\alpha}(d_S(x)) \leq V(x) \leq \overline{\alpha}(d_S(x)); \end{equation} \]

    (36)

  • (Decrease during flows) For all \( x \in \mathbb{R}^{n_x}\) and \( u_c \in \mathcal{U}_c\) such that \( (x,u_c) \in C\) ,

    \[ \begin{equation} \langle\nabla V(x),f(x,u_c)\rangle \leq -\alpha_c(V(x)); \end{equation} \]

    (37)

  • (Decrease at jumps) For all \( x \in \mathbb{R}^{n_x}\) and \( u_d \in \mathcal{U}_d\) such that \( (x,u_d) \in D\) ,

    \[ \begin{equation} V(g(x,u_d)) - V(x) \leq -\alpha_d(V(x)). \end{equation} \]

    (38)

If the class-\( \mathcal{K}_\infty\) functions in Lemma 2 are linear, we can show UGpES of \( S\) , which can result in robustness against uncertainties. The same property can also be shown with \( \underline{\alpha}\) , \( \overline{\alpha}\) , and only \( \alpha_c\) (resp., \( \alpha_d\) ) being linear, under a dwell-time (resp., reverse dwell-time) condition, as in [22] and defined properly later in this dissertation. As shown later in this dissertation, such conditions are useful for bringing stability and convergence thanks to flows to make up for the lack of these at jumps, and vice-versa, achieving pre-asymptotic stability from the flow-jump combination.

Notations. For a function \( V:\mathbb{R}^{n_\eta}\to \mathbb{R}\) and a hybrid system with state \( \eta \in \mathbb{R}^{n_\eta}\) and input \( u\) , flow dynamics \( \dot{\eta}=f(\eta,u)\) and jump dynamics \( \eta^+=g(\eta,u)\) , we denote \( \dot{V}(\eta,u) :=\langle \nabla V(\eta),f(\eta,u)\rangle\) (where \( \langle a,b \rangle\) denotes the cross product of vectors \( a\) and \( b\) ) the derivative of \( V\) along the flows and \( V^+(\eta,u) :=V(g(\eta,u))\) the value of \( V\) after a jump.

1.5.3.2 Towards Observer Design

In this dissertation, our goal is to design observers that estimate the state of hybrid systems using their known inputs and outputs, similar to the approach taken for continuous- and discrete-time systems discussed in Section 1.5.2.2. To begin, we analyze some key properties of hybrid systems to understand how they influence the observer design problem:

• Non-uniqueness of system representations

In continuous- or discrete-time systems, changing the representation (or the maps) requires a change of state variables. However, in hybrid systems, the coupling of continuous- and discrete-time dynamics, along with the flow and jump conditions, can lead to non-uniqueness in the system maps even with the same state variables. Since observer design relies heavily on these system maps, this non-uniqueness can present challenges in designing accurate observers. In certain cases, however, it plays a valuable role in analysis, revealing that some conditions imposed on hybrid systems, such as the invertibility of the jump dynamics, may be sufficient but not strictly necessary. For an example of this, refer to Example 15;

Example 2 (Non-unique representation of a bouncing ball)

Consider the bouncing ball system in Example 1. With the same state variables, this system can be described not only by the form (32), but also by any representation where \( x_1^+ = x_1\) is replaced with \( x_1^+ = p(x_1)\) where \( p\) is any function satisfying \( p(0) = 0\) . Indeed, all such choices of \( p\) define hybrid systems that yield the same solutions because the discrete dynamics occur only when \( x_1 = 0\) .

• Non-uniqueness of solutions

In continuous-time systems (11), given an initial condition, the resulting solution is unique if \( f\) is locally Lipschitz around this initial condition, uniformly in the time \( t\) (as per the Cauchy-Lipschitz theorem). In discrete-time systems (12), the solution is unique as long as \( f\) is a singleton. In hybrid systems (31), different solutions may arise from the same initial condition due to inclusions. For example, the timer

\[ \begin{equation} \dot{\tau} \in [1, 1.5] ~ \tau \in [0, 10], ~~ \tau^+ \in [0, 2] ~ \tau \in [8, 10], \end{equation} \]

(39)

can evolve at any rate within \( [1, 1.5]\) during flows and reset to any value in \( [0, 2]\) after each jump. Even when the flow map is Lipschitz and the jump map is a singleton, solutions may still be non-unique if the flow and jump set overlap (\( C \cap D \neq \emptyset\) ), i.e., the solution can choose either to flow or jump when in this intersection. For instance, the timer

\[ \begin{equation} \dot{\tau} = 1 ~ \tau \in [0, 10], ~~ \tau^+ = 0 ~ \tau \in [8, 10], \end{equation} \]

(40)

can either reset to zero or continue counting to ten when it is in the interval \( [8, 10]\) . If solutions are non-unique, determining the initial condition is not equivalent to determining the current state, which is the objective of observer design. However, non-unique solutions are permissible as long as they remain distinguishable from the measured output. The following conditions ensure the uniqueness of solutions to an autonomous hybrid system;

Lemma 3 (Conditions for uniqueness of hybrid solutions [10])

A hybrid system with data \( (C,F,D,G)\) has unique solutions if and only if:

  1. For \( x^\star \in C\) , there exists a unique solution to the differential inclusion \( \dot{x} \in F(x)\) satisfying \( x(0) = x^\star\) and \( x(t) \in C\) ;
  2. For \( x^\star \in D\) , \( G(x^\star)\) is a singleton;
  3. For \( x^\star \in C \cap D\) , there are no non-trivial solutions to the differential inclusion \( \dot{x} \in F(x)\) satisfying \( x(0) = x^\star\) and \( x(t) \in C\) .

• Dependence of jump times on initial conditions (and inputs)

In continuous-time systems (11), the time domain of solutions is (a subset of) \( \mathbb{R}_{\geq 0}\) , and for discrete-time systems (12), it is \( \mathbb{N}\) . These domains are generally the same for the solutions of interest and are also shared with the estimate, making it straightforward to define the estimation error by comparing the real solution with the estimate at the same time. However, in hybrid systems, because the flow and jump conditions depend on the state and inputs via the sets \( C\) and \( D\) , the jump times and thus the time domain of solutions—see Definition 1—depend greatly on the initial condition and the inputs. This domain \( \rm{dom} x\) thus becomes solution-exclusive.

Example 3 (Jump times of a bouncing ball)

Consider the bouncing ball system in Example 1 with some parameters. In two scenarios where the ball is initialized at the same height but with different starting velocities, it touches the ground at different times and rebounds at different speeds, resulting in distinct time domains as illustrated in Example 3.

Figure 10
Figure 11
Figure 9. Solutions to system (32) with \( x(0,0) = (1,0)\) (left) and \( x(0,0) = (1,3)\) (right).

Because the initial condition and the inputs are unknown when designing observers (only a set of initial conditions and sets of considered solutions are known), we cannot ensure that the estimate provided by the observer shares the same time domain as the system’s solution. Consequently, directly comparing them at the same time becomes difficult, complicating the usual definition of estimation error and observer convergence [67];

• Zeno solutions

Recall from Definition 4 that a (maximal) solution is Zeno if it is not \( t\) -complete but is \( j\) -complete. This phenomenon is illustrated in the following example.

Example 4 (Zeno behavior in a bouncing ball)

Consider the bouncing ball system in Example 1, with \( c_f = 0\) and any \( d_f\) , with a coefficient of restitution \( c_r \in (0,1)\) , which excludes the case where \( c_r = 1\) (corresponding to a perfectly elastic collision), and with \( u_d = 0\) (no energy injected at jumps). Let the velocity immediately after the first impact\( /\) jump be \( v_1 > 0\) . After solving the flow part of system (32), we determine the flow time between the \( j\) -th and \( (j+1)\) -th jump, for \( j \in \mathbb{N}_{\geq 1}\) , to be \( \frac{2v_j}{a_g}\) , where \( v_j = c_r v_{j-1} = ... = c_r^{j-1} v_1\) is the velocity right after the \( j\) -th jump. Consequently, the solution’s jump times (or total flow time since initialization) are given by

\[ \begin{equation} t_j = t_1 + \sum_{q=1}^{j-1}\frac{2v_q}{a_g} = t_1 + \frac{2v_1}{a_g} \sum_{q=1}^{j-1} c_r^{q-1} = t_1 + \frac{2v_1}{a_g} \frac{1 - c_r^{j-1}}{1 - c_r}, ~~ j \in \mathbb{N}_{\geq 2}, \end{equation} \]

(41)

for \( c_r \in (0, 1)\) . Therefore, for \( c_r \in (0, 1)\) , we have \( \sup \rm{dom}_j x = +\infty\) but the maximal solutions are not \( t\) -complete since \( \sup \rm{dom}_t x = t_1 + \frac{2v_1}{a_g} \frac{1}{1 - c_r}\) , indicating the Zeno behavior. Suppose still \( c_r \in (0,1)\) but we have a constant input \( u_d > 0\) at jumps. In that case, this injects energy into this system at each jump and the solution converges to a periodic trajectory with \( x_2^+\) approaching \( \frac{u_d}{1+c_r}\) , so no more Zeno solutions.

The Zeno phenomenon is against physics in the sense that an infinite number of events (jumps) can happen in a finite amount of time. It arises here because, in the mathematical model (30), it is supposed that each jump takes zero time. In observer design, besides the jumps which may bring us observability, we need to have a sufficient amount of flow time for observability and convergence. For instance, the high-gain observer (27) which is used to estimate part of the system’s state, converges asymptotically and thus needs infinite flow time. This is why Zeno can be an obstruction and we may need to assume some conditions on the time domains of hybrid solutions of interest to ensure some uniform amount of flowing and jumping.

Summary

Due to the considerations mentioned above, this dissertation focuses on hybrid systems with well-defined maps, specifically omitting the case of (30) of inclusions to work only on observer designs for systems of the form (31). Uniqueness of solutions is assumed whenever needed, especially in Section 4.2 when this is supposed to hold in backward time. We then divide these hybrid systems into two main groups, namely those whose jump times are known as treated in Section 3 and those with unknown jump times, addressed in Section 4. The first case assumes some mechanisms that allow us to detect when the solution jumps, such as impact sensors in walking robots, or some conditions related to the jump that can be checked if satisfied using the measured output, for instance, the bouncing ball in Example 1 where a jump occurs when the measurement equals zero. While it is true that this detection can be imperfect or delayed, such imperfections could be managed by the observer’s robustness provided that the maximal delay is smaller than the smallest possible time between successive jumps of the system solution, resulting in practical convergence—convergence of the estimation error to a bounded neighborhood of zero outside of the delays—as studied in [22, Section 6]. With this in mind, in Section 3 of this dissertation, we assume that the jump times are exactly detected to focus ourselves on the observer design problem. Similarly to Section 1.5.2.1, we denote:

  • \( \mathcal{X}_0 \subseteq \mathbb{R}^{n_x}\) as the set of initial conditions \( x(0,0)\) of interest for system (31);
  • \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ) as the set of input trajectories \( t \mapsto u_c(t)\) , called \( \mathfrak{u}_c\) (resp., \( j \mapsto u_{d,j}\) , called \( \mathfrak{u}_d\) ) of interest for system (31);
  • \( \mathcal{U}_c \subseteq \mathbb{R}^{n_{u,c}}\) (resp., \( \mathcal{U}_d \subseteq \mathbb{R}^{n_{u,d}}\) ) as the set of the values that \( u_c(t)\) (resp., \( u_{d,j}\) ) can take along a trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) (resp., \( \mathfrak{u}_d \in \mathfrak{U}_d\) ), for \( t \in \mathbb{R}_{\geq 0}\) (resp., \( j \in \mathbb{N}\) ).

Because the jump times of solutions to system (31) are known, we build for this system a synchronized observer of the general form

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{\zeta}}&=& \mathcal{F}(\hat{\zeta},y_c,u_c,t,j) &\text{when~(31) flows}\\ \hat{\zeta}^+&=&\mathcal{G}(\hat{\zeta},y_d,u_d,t,j) &\text{when~(31) jumps} \end{array} \right. \end{equation} \]

(42.a)

with the estimate \( \hat{x}\) obtained from a solution \( (t,j) \mapsto \hat{\zeta}(t,j)\) to dynamics (42.a) as

\[ \begin{equation} \hat{x}(t,j)= \Upsilon(\hat{\zeta}(t,j),t,j), \end{equation} \]

(42.b)

where \( \hat{\zeta} \in \mathbb{R}^{n_\zeta}\) is the observer state; \( \mathcal{F}\) , \( \mathcal{G}\) , and \( \Upsilon\) are the observer dynamics and output maps designed such that the dependence of \( \Upsilon\) on time \( (t,j)\) is only through the inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) and each maximal solution \( (x,\hat{\zeta})\) to the cascade (31)-(42) initialized in \( \mathcal{X}_0 \times \mathcal{Z}_0\) for some appropriate \( \mathcal{Z}_0 \subseteq \mathbb{R}^{n_\zeta}\) and with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in \mathfrak{U}_c\times \mathfrak{U}_d\) is complete and verifies at least asymptotic convergence

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x (= \rm{dom} \hat{x})}}|x(t,j)-\hat{x}(t,j)| = 0. \end{equation} \]

(43)

Indeed, (42.a) ensures that the estimate \( \hat{x}\) shares the same time domain as the system solution \( x\) , allowing them to be directly compared at the same hybrid time. Using the language of sets, we see that condition (43) is equivalent to the asymptotic convergence to a set where the estimate equals the true state, i.e.,

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x (= \rm{dom} \hat{x})}}d_S((x(t,j),\hat{x}(t,j))) = 0, \end{equation} \]

(44.a)

where

\[ \begin{equation} S:=\{(x,\hat{x}) \in \mathbb{R}^{n_x}\times\mathbb{R}^{n_x}: x = \hat{x}\}. \end{equation} \]

(44.b)

When jump times are known, we prefer writing (43) over (44) since it is more direct and easily related to similar properties in continuous and discrete time. In Section 3, we assume certain properties of the solutions of interest, including completeness, and design various observer structures (42) based on the form of system (31), to guarantee (43) and potentially additional properties such as robustness or arbitrarily fast convergence rate as introduced in Section 1.5.2.2.

In the second scenario of unknown jump times addressed in Section 4, synchronization between the system and observer as in (42.a) becomes impossible, making asymptotic convergence more challenging to define. Examples of such hybrid systems whose jump times cannot be detected include the stick-slip behavior encountered in oil drilling, as discussed in Section 4.3. In this dissertation, we focus on the autonomous form of (31)

\[ \begin{equation} \dot{x} = f(x) ~ x \in C, ~ ~ x^+ = g(x) ~ x \in D ~ y = h(x), \end{equation} \]

(45)

with the output \( y\) assumed to be continuous through the jumps, i.e., \( h(x) = h(g(x))\) for all \( x \in D\) , preventing jump detection through measurement alone. Building on the concept of gluing the time domain initiated in [32], we transform system (45) into continuous-time dynamics of the form

\[ \begin{equation} \dot{z} = Az + By, \end{equation} \]

(46)

of dimension \( n_z \geq n_x\) , for some \( A\) Hurwitz. In these new coordinates, the state \( z\) evolves as a continuous-time function, effectively “gluing” the hybrid time domain as the method’s name suggests. This eliminates the need for jump detection in observer design, unlike in the previous case. An exponentially stable observer for system (47) is then obtained by running the same dynamics from any initial condition in \( \mathbb{R}^{n_z}\) , specifically \( \dot{\hat{z}} = A\hat{z} + By\) , as the estimation error verifies \( \dot{z} - \dot{\hat{z}} = A(z - \hat{z})\) with \( A\) being Hurwitz. Then, by ensuring that the transformation is injective (and hence left-invertible) thanks to some distinguishability condition and a sufficiently large dimension \( n_z\) , we can recover the estimate in the original \( x\) -coordinates. Formally, an observer for system (45) would take the form

\[ \begin{equation} \dot{\hat{z}} = A\hat{z} + By, ~~ \hat{x} = \Upsilon(\hat{z}), \end{equation} \]

(47)

where \( \Upsilon\) is the left inverse of the transformation. In this case, the estimate \( \hat{x}\) is a continuous-time signal instead of a hybrid one, and so convergence is not defined in the sense of (43) by comparing the solution and the estimate at the same hybrid time, but as the convergence of \( (x(t,j),\hat{x}(t))\) in time to the indistinguishable set

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x}} d_{S_\mathcal{I}}((x(t,j),\hat{x}(t))) = 0, \end{equation} \]

(48.a)

where

\[ \begin{multline} S_\mathcal{I} := \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times \mathbb{R}^{n_x} : x=\hat{x} \} \cup \{(x,\hat{x}) \in D\times \mathbb{R}^{n_x} : g(x)=\hat{x} \}\\ \cup \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times D : x=g(\hat{x}) \}\cup \{(x,\hat{x}) \in D\times D : g(x)=g(\hat{x})\}. \end{multline} \]

(48.b)

The set \( S_\mathcal{I}\) consists of points that generate solutions that cannot be distinguished based on the output \( y\) . Therefore, convergence to this set is the best achievable outcome with observer (47). Because \( y\) is continuous at jumps, points that either are the same, are one jump apart, or jump to the same point, are indistinguishable from each other and hence they belong to this set. The formulations and results are made more precise in Section 4.2.

1.5.4 Conclusion

In this chapter, we have introduced the foundational concepts necessary for understanding the material in this dissertation. Key notations used throughout are presented in their respective sections. Moving forward, in Section 2, we begin by discussing the discrete-time observers developed during this Ph.D. work. Then, in Section 3, we propose observers for hybrid systems with known jump times, leveraging the discrete-time tools introduced earlier. Finally, in Section 4, we address observers for hybrid systems with unknown jump times.

2 Contributions to Robust Observer Design for Discrete-Time Systems

2.1 Review of Observer Designs for Discrete-Time Systems

 

Ce chapitre passe en revue les méthodes de synthèse d’observateurs existantes pour les systèmes en temps discret. Nous commençons par les observateurs pour les systèmes linéaires, incluant les méthodes basées sur les Inégalités Matricielles Linéaires qui reposent sur la détectabilité et les approches de type Kalman qui exploitent l’observabilité complète uniforme. Pour les systèmes non linéaires, nous explorons une gamme de méthodes, allant des techniques de linéarisation locale, telles que le filtre de Kalman étendu, à des approches comme les estimateurs dits “dead-beat” et les observateurs à horizon glissant, qui nécessitent le stockage des sorties passées. Nous introduisons ensuite deux nouvelles approches grand gain qui constituent deux contributions de cette thèse et seront détaillées dans les deux chapitres suivants: l’une basée sur une forme normale de constructibilité qui ressemble à ce qui est fait pour l’observateur grand gain continu classique, et l’autre basée sur la méthode de KKL. Ces synthèses globales offrent une vitesse de convergence arbitrairement rapide (en un sens à préciser) ainsi qu’une robustesse face aux incertitudes.

2.1.1 Overview

Consider a nonlinear (time-varying) discrete-time system given by

\[ \begin{equation} x_{k+1} = f_k(x_k), ~~ y_k = h_k(x_k), \end{equation} \]

(49)

where \( x_k \in \mathbb{R}^{n_x}\) and \( y_k \in \mathbb{R}^{n_y}\) are the state and the measured output at discrete time \( k\in \mathbb{N}\) ; \( f_k: \mathbb{R}^{n_x}\times\mathbb{R}^{n_y} \to \mathbb{R}^{n_x}\) and \( h_k: \mathbb{R}^{n_x} \to \mathbb{R}^{n_y}\) are the dynamics and output maps. The time dependence of \( (f_k,h_k)_{k \in \mathbb{N}}\) may capture their dependence on some exogenous inputs \( k \mapsto u_k\) , seen as known functions of time. The goal here is to build for system (49) an algorithm that estimates online its state \( x_k\) , initialized in a set \( \mathcal{X}_0 \subset \mathbb{R}^{n_x}\) characterizing the solutions of interest, from its known outputs \( y_k\) and inputs in \( (f_k,h_k)\)  [1], which we call an asymptotic observer as formalized next in Definition 6.

Definition 6 (Asymptotic observer for system (49))

The system

\[ \begin{equation} \hat{z}_{k+1} = \mathcal{F}_k(\hat{z}_k,y_k), ~~ \hat{x}_k = \Upsilon_k(\hat{z}_k,y_k), \end{equation} \]

(50)

with the map sequences \( (\mathcal{F}_k)_{k \in \mathbb{N}}\) and \( (\Upsilon_k)_{k \in \mathbb{N}}\) , where \( \mathcal{F}_k: \mathbb{R}^{n_z} \times \mathbb{R}^{n_y} \to \mathbb{R}^{n_z}\) and \( \Upsilon_k: \mathbb{R}^{n_z} \times \mathbb{R}^{n_y} \to \mathbb{R}^{n_x}\) for some \( n_z \in \mathbb{N}\) and some initialization set \( \mathcal{Z}_0 \subset \mathbb{R}^{n_z}\) , designed such that along any solution \( k \mapsto (x_k,\hat{z}_k)\) to the cascade (49)-(50) initialized in \( \mathcal{X}_0 \times \mathcal{Z}_0\) , we have

\[ \begin{equation} \lim_{k \to +\infty} |x_k - \hat{x}_k| = 0, \end{equation} \]

(51)

is an asymptotic observer for system (49).

From Definition 6, we see that the (discrete-time) observer design problem consists in finding the observer map sequences \( (\mathcal{F}_k,\Upsilon_k)_{k \in \mathbb{N}}\) and the initialization set \( \mathcal{Z}_0\) to achieve at least the asymptotic convergence of the estimation error (51). The observer state \( z_k \in \mathbb{R}^{n_z}\) can be \( x_k\) in certain linear observers (in which case we say that the observer is designed in the original \( x\) -coordinates). It can also contain extra state variables that help the observer work, such as the covariance matrix in Kalman(-like) designs, or especially in nonlinear observers where we typically need to go to new coordinates of higher dimensions, as discussed in Section 1.5. Note that from an application point of view, the implementation of observers, which may require storing a history of the inputs in \( z_k\) , is computation-wise lighter in discrete time since this is finite-dimensional, contrary to continuous time when this storage becomes infinite-dimensional.

The existence of an asymptotic observer requires the following property from system (49).

Definition 7 (Asymptotic detectability of system (49))

System (49) is asymptotically detectable if any of its two solutions \( k \mapsto x_{a,k}\) and \( k \mapsto x_{b,k}\) with the same \( (f_k,h_k)_{k \in \mathbb{N}}\) and the same outputs, i.e.,

\[ \begin{equation} y_{a,k} = y_{b,k}, ~~ \forall k \in \mathbb{N}, \end{equation} \]

(52)

verify

\[ \begin{equation} \lim_{k \to +\infty}|x_{a,k} - x_{b,k}| = 0. \end{equation} \]

(53)

This property assumes the completeness of solutions to (49) as introduced in Section 1.5.2.1, which is trivial for discrete-time systems as long as the maps are well-defined but may not hold for continuous-time or hybrid systems. The necessity of asymptotic detectability for the existence of an asymptotic observer is now established in Lemma 4.

Lemma 4 (Asymptotic detectability is necessary for an asymptotic observer)

If there exists an asymptotic observer (50) for system (49), then system (49) is asymptotically detectable (by Definition 7).

Proof. Assume that there exists an asymptotic observer (50) for system (49). Consider two solutions \( k \mapsto x_{a,k}\) and \( k \mapsto x_{b,k}\) to system (49) initialized in \( \mathcal{X}_0\) with equal outputs, i.e., \( y_{a,k} = y_{b,k}\) for all \( k \in \mathbb{N}\) . Consider \( k \mapsto \hat{z}_k\) defined on \( \mathbb{N}\) such that \( k \mapsto (x_{a,k},\hat{z}_k)\) is solution to the cascade (49)-(50) with output \( k \mapsto \hat{x}_{a,k}\) . Since \( y_{a,k} = y_{b,k}\) for all \( k \in \mathbb{N}\) , \( k \mapsto (x_{b,k},\hat{z}_k)\) is also solution to the cascade (49)-(50) with output \( k \mapsto \hat{x}_{b,k}\) such that \( \hat{x}_{a,k}=\hat{x}_{b,k}\) for all \( k \in \mathbb{N}\) . By Definition 6, we have both \( \lim_{k \to +\infty} |x_{a,k}-\hat{x}_{a,k}| = 0\) and \( \lim_{k \to +\infty} |x_{b,k}-\hat{x}_{b,k}| = 0\) . Therefore, we have \( \lim_{k \to +\infty} |x_{a,k}-x_{b,k}| = 0\) , implying that system (49) is asymptotically detectable. \( \blacksquare\)

In other words, detectability is defined in such a way that it is the necessary condition for observer existence. In terms of observability conditions in discrete time, the notions of observability [68, 69] and constructibility [70, 71] (or reconstructibility in some literature [72]) differ mainly in the invertibility of the dynamics. To be more precise, given a finite sequence of outputs (and possibly known inputs), the former corresponds to uniquely determining from these the initial state—see Definition 10, and the latter typically means determining the final state (see Definition 9 and also the backward distinguishability in Definition 11). If we manage to retrieve the initial state, we can always proceed with the system’s dynamics to reach the final state, indicating that observability classically implies constructibility. However, when the dynamics lack invertibility, the initial state may not be recoverable from the final state. It is crucial to emphasize that the observer does not necessarily require this capability, as its primary task is to estimate the current state. For linear systems, [73] distinguishes these notions and proposes a constructible canonical form with a linear output. For nonlinear systems, [70] introduced constructibility as a local property. The recent work [71] proposes in the nonlinear setting a notion of backward observability in the same spirit as constructibility but written mainly in terms of the observation space of the linearized systems (from a differential geometry point of view). An important note is that the notion of instantaneous observability of continuous-time systems, which is based on the successive derivatives of the output determining uniquely the state, does not exist in discrete time. Thus, the convergence of the estimation, even in finite time, cannot happen arbitrarily fast from the initial time as ensured by the continuous-time high-gain observer [53], but only after a certain time when we have gathered sufficient outputs.

In the next sections, we review discrete-time observer designs and point out their respective observability conditions. Existing results are summarized in Table 1 at the end of this chapter. We also draw important conclusions for the observers developed later in Section 2.2 and Section 2.3.

2.1.2 Global Linear Designs

Let us first consider a linear (time-varying) discrete-time system

\[ \begin{equation} x_{k+1} = F_k x_k + u_k, ~~ y_k = H_k x_k, \end{equation} \]

(54)

where \( x_k \in \mathbb{R}^{n_x}\) is the state, \( u_k \in \mathbb{R}^{n_x}\) is the control input, \( y_k \in \mathbb{R}^{n_y}\) is the measured output, and \( (F_k,H_k)\) are known matrices of appropriate dimensions, all at discrete time \( k\) . Benefiting from this form’s linearity, many observers have been proposed in the literature as reviewed next.

2.1.2.1 LMI-Based Observers

If the pair \( (F_k,H_k)\) is quadratically detectable [44] uniformly in time, namely if there exists \( k \mapsto L_k\) , some \( 0 \leq a < 1\) , and \( Q \in §_{>0}^{n_x}\) such that the matrix inequality

\[ \begin{equation} (F_k - L_k H_k)^\top Q (F_k - L_k H_k) \leq a Q, ~~ \forall k \in \mathbb{N}, \end{equation} \]

(55)

holds, then an asymptotic observer for system (54) would be

\[ \begin{equation} \hat{x}_{k+1} = F_k \hat{x}_k + u_k + L_k(y_k - H_k \hat{x}_k). \end{equation} \]

(56)

This takes the structure of observer (50) with \( \hat{z}_k = \hat{x}_k\) , usually referred to as the Luenberger observer for linear systems, which is the discrete-time version of [74]. Actually, (55) is not a Linear Matrix Inequality[LMI] since it is not linear in its variables, but it may be transformed into one thanks to Schur’s lemma. While detectability alone is not sufficient for observer design in nonlinear systems, it becomes adequate for linear systems due to the well-defined structure of the maps, allowing the construction of the observer gains. Note that while it allows achieving convergence, detectability is not sufficient for tuning the convergence rate, for which we need stronger properties such as observability as in the next section.

In practice, LMI-based observers are frequently extended to nonlinear systems, as seen in [75] or the work by the author in [35], leveraging the linearity in the dynamics to incorporate additional useful properties, such as disturbance decoupling in the estimation error [76]. Some LMI-based approaches have been developed for discrete-time normal forms with Lipschitz nonlinearities as in [77]. The main limitation of these methods is that, besides having to assume that the nonlinear part of the dynamics is Lipschitz, they require that detectability comes from the linear part of the dynamics, and so they do not work when it is the nonlinear part that allows us to estimate the state. Moreover, they typically do not guarantee the solvability of the LMIs.

2.1.2.2 Kalman(-Like) Designs

For the linear form (54), the celebrated Kalman observer was introduced in the early 60s by Kalman and Bucy [78] as an optimal filter. In a stochastic context and under an Uniform Complete Observability[UCO] assumption, it minimizes the covariance of the estimation error in the presence of Gaussian dynamics and measurement noise. Its appeal lies in its systematic design and easiness of tuning, which is linked to the (assumedly known) covariance of those disturbances. The Kalman filter was later extended to discrete-time systems [79] and various other contexts [80], becoming widely used in industrial applications. The discrete-time Kalman filter, consisting of prediction, namely proceeding from the estimate with the system’s dynamics, and correction of the estimate by comparing it with the measurement, has been shown to exhibit asymptotic stability “in the large”, namely convergence of the estimation error after several discrete steps instead of after each step, within the stochastic and deterministic frameworks in [79, 81], respectively.

On the other hand, in the early \( 90\) s, an alternative Kalman-like observer was developed [82, 83], optimizing in a deterministic setting the ability of the estimate to explain the past output history, with a certain forgetting factor and weighting, describing the confidence in the output measurement. This design was extended to discrete-time systems in [43], taking the form

\[ \begin{align} \hat{x}_{k+1} &= F_k\hat{x}_k + u_k + K_k(y_k-H_k\hat{x}_k)\\ P_{k+1} &= \frac{1}{\gamma} F_k(\rm{Id}-K_kH_k)P_kF_k^\top \\\end{align} \]

(57.a)

with

\[ \begin{equation} K_k = P_kH_k^\top (H_k P_k H_k^\top+R_k)^{-1}, \end{equation} \]

(57.b)

where \( \gamma \in (0,1]\) is a forgetting factor and \( k\mapsto R_k\) is a uniformly bounded positive definite weighting matrix, both serving as design parameters. Observer (57) takes the structure (50) with \( \hat{z}_k = (\hat{x}_k,P_k)\) . This observer works globally but must be initialized with \( P_0\) positive definite, so \( \mathcal{Z}_0 = \mathbb{R}^{n_x} \times §_{>0}^{n_x}\) . The forgetting factor allows for speeding up the convergence rate when we push it close to zero, and a form of (57) without this is proposed in [81].

Both the Kalman and Kalman-like observers rely on the following observability condition.

Definition 8 (Uniform Complete Observability[UCO] of system (54))

System (54) is UCO if there exist \( m \in \mathbb{N}_{>0}\) and \( c_o > 0\) such that for all \( k \in \mathbb{N}_{\geq m}\) ,

\[ \begin{equation} \sum_{j = k-m}^{k-1} (F^{-1}_{k-1})^\top (F^{-1}_{k-2})^\top\ldots (F^{-1}_j)^\top H_j^\top H_j F^{-1}_j \ldots F^{-1}_{k-2} F^{-1}_{k-1}\geq c_o \rm{Id} > 0. \end{equation} \]

(58)

Indeed, the analysis of both types relies on the Lyapunov function \( V(x_k,\hat{x}_k,P_k) = (x_k - \hat{x}_k)^\top P_k^{-1} (x_k - \hat{x}_k)\) , and UCO ensures the uniform lower boundedness of the observability Gramian, which in turn leads to that of \( P_k^{-1}\) . The covariance matrix \( P_k\) here serves to gather observability from the outputs, and thus this boundedness signifies the uniform sufficiency of information needed to estimate the state, related to the notion of persistence of excitation [84]. A significant advantage of Kalman(-like) designs is that they are systematic, meaning that they can be implemented without having to check the system’s UCO, unlike in LMI-based designs where we have to solve the LMI, i.e., check the detectability, in order to obtain the observer’s gain. The difference between the Kalman-like observer and the Kalman one mainly lies in the absence of noise on the dynamics which facilitates a Lyapunov stability analysis by linking the Lyapunov matrix directly to the observability Gramian. The Kalman-like observer in [43] thus provides exponential stability with an arbitrarily fast rate after a certain time, achievable with a (quadratic) strict Lyapunov function, unlike in [79, 81], where the Lyapunov function decreases over a certain finite number of steps. Note also that for these Kalman(-like) designs, while the invertibility of the system dynamics typically appears in the analysis such as (58), it is rather a sufficient condition and is not required for observer implementation.

2.1.3 Local Linearization-Based Techniques

Now, we come back to the observer designs for the general nonlinear form (49), starting with techniques based on linearization, aiming to benefit from the structure or properties of linear systems to build nonlinear observers.

Substantial literature has been focusing on transforming system (49) via an immersion into a (higher-order) form that is linear in both the dynamics and output, modulo output injection:

\[ \begin{equation} z_{k+1} = A(k,y_k) z_k + B(k,y_k), ~~ y_k = C_k z_k. \end{equation} \]

(59)

We can then try to design an asymptotic observer for this new form, using methods in Section 2.1.2, and either come back with the estimate by left-inverting the transformation or write back the observer in the original coordinates. The work [85] builds an observer using the \( m\) past outputs and necessitates the full rank of the Jacobian of an observability map derived from future outputs of the form

\[ \begin{equation} x \mapsto \mathcal{O}^{fw}_k(x) = \begin{pmatrix} h_k(x) \\ (h_{k+1} \circ f_k)(x) \\ \ldots \\ (h_{k+(m-1)} \circ f_{k+(m-2)} \circ \ldots \circ f_k)(x) \end{pmatrix}. \end{equation} \]

(60)

This results in a global observer in the transformed coordinates (59), but only a local observer when written in the original coordinates. Other studies such as [86, 87, 88, 89], relying on differential geometry, apply local transformations to obtain the form (59). These techniques, however, face two key limitations: i) The transformations may only exist locally, preventing global results, and ii) The transformed linear form is not always guaranteed to be observable.

In later years, attempts have been made to develop for system (49) the nonlinear version of the Kalman observer, called the Extended Kalman Filter[EKF], whose estimate propagates following the system’s nonlinear dynamics but whose covariance matrix \( P_k\) evolves based on the Jacobians of the dynamics and output maps evaluated along the estimate provided by the same observer, which are \( \left(\frac{\partial f_k}{\partial x}(\hat{x}_k),\frac{\partial h_k}{\partial x}(\hat{x}_k)\right)\) . This ensures only local convergence, assuming the UCO condition in Definition 8 holds on these linearizations of the dynamics and output along the estimate [90, 91, 92]. We also have no guarantee that the linearized dynamics and output are UCO, even if the original system (49) is observable in the nonlinear sense. Even if satisfied, this kind of assumption typically introduces a loop in the analysis, since the estimation error must remain small to guarantee observability along the estimate, which is in turn needed to keep the error small. This loop is broken in [90, 93] but the analysis remains inherently local. Note also that those papers do not mention any explicit stability\( /\) robustness guarantees. On the other hand, we have not documented any nonlinear version of the discrete-time Kalman-like observer [43].

Therefore, though applicable to nonlinear systems, linearization-based techniques face major issues of localness (which hinders robustness) and possible loss of observability from linearization.

2.1.4 Output-Inverting Estimators

These methods involve storing a sufficient number of past outputs in a moving window of size \( m\in \mathbb{N}_{\geq 1}\) and determining the state from them, either by directly inverting this data as in dead-beat observers or through online optimization as in a Moving Horizon Observer[MHO]. In the absence of uncertainties, finite-time convergence is achieved after \( m\) steps. For these algorithms to work, it suffices that the map between a state and its \( m\) past outputs is uniformly Lipschitz injective as in our Definition 12. But works such as [69, Definitions 3 and 4] rather write this condition with the map made of future outputs (60). These observers also do not require invertibility of the system’s dynamics.

2.1.4.1 Dead-Beat Observers

These rely on the left inversion of the (forward) observability map comprising future outputs, such as with Newton algorithms [94, 95], providing instantaneous estimation as soon as enough output information is gathered, but lacking filtering effects against measurement noise. In fact, very few works in the literature propose global nonlinear discrete-time observer designs that exhibit robustness against disturbances and measurement noise, e.g., [96, 50] and in Section 2.2 and Section 2.3. Note also that though these dead-beat observers can work under a mild constructibility condition that allows determining the current state, the papers [94, 95] assume the stronger observability condition.

2.1.4.2 Moving Horizon Observers

The MHO [97, 98, 99, 96, 50] can be viewed as a more computationally demanding but also more robust variant of dead-beat observers. These minimize the estimation error using an observability map made of the \( m\) current and future outputs, relying on the injectivity of this map with respect to the state from \( m\) steps ago, thus requiring observability instead of constructibility\( /\) backward distinguishability. Specifically, these algorithms involve solving an online optimization problem to find \( \hat{x}_{k - m | k}\) , the estimate made at time \( k\) of the state \( m\) steps ago, that minimizes a cost function of the form

\[ \begin{equation} J_k(\hat{x}_{k - m | k}) = \mu \left|\hat{x}_{k - m | k} - \bar{x}_{k - m}\right|^2 + \sum_{q = k - m}^k \left|y_q - h_q(\hat{x}_{q|k})\right|^2 \end{equation} \]

(61.a)

for some \( \mu \geq 0\) , where \( \bar{x}_{k - m}\) is an initial guess of the true state value \( m\) steps in the past, subject to the \( m\) recursive constraints

\[ \begin{equation} \hat{x}_{q+1|k} = f_q(\hat{x}_{q|k}), ~~ q \in \{k - m,k - m + 1,\ldots,k-1\}, \end{equation} \]

(61.b)

that imposes the system’s dynamics on the values \( \hat{x}_{q|k}\) to compute the \( m\) past estimated outputs. Once \( \hat{x}_{k - m | k}\) is obtained from this optimization, the current state can be estimated by propagating with the system’s dynamics. This observer takes the form (50), with \( (\mathcal{F}_k)_{k \in \mathbb{N}}\) serving to store the past outputs in \( z_k\) and \( (\Upsilon_k)_{k \in \mathbb{N}}\) the optimizer. Some versions of the cost function [50] include terms that allow for estimating the system’s disturbances. These exhibit robustness against modeling uncertainties or numerical errors [100]. Recently, they have been shown to have some kind of Input-to-State Stability[ISS] against disturbances in [50, Corollary 1], resulting in one of the first robust designs for nonlinear discrete-time systems. Similar to dead-beat observers, while the MHO can function under constructibility, the works [97, 98, 99] all assume observability, which is a stronger condition.

2.1.5 Novel High-Gain Designs in this Dissertation

These designs are the contributions of this dissertation in Section 2.2 and Section 2.3, and are included here for an exhaustive literature review.

The high-gain mechanism, well-known in continuous time [26, 53], refers to an observer that works when its dynamics are pushed sufficiently fast. As discussed in Section 1.5.2.2, in continuous time, such designs are known for their two key properties: arbitrarily fast convergence and robustness against disturbances\( /\) noise. However, when this Ph.D. research began in \( 2021\) , no such designs had been documented in discrete time, as noted in [101]. Recently, the MHO has been shown to exhibit similar properties, with [50] proving its robustness. So far, we only know of two discrete-time high-gain designs as follows.

2.1.5.1 “Classical” High-Gain Structure

We propose in Section 2.2 below a discrete-time observer whose structure resembles that of the classical continuous-time high-gain observer (26). This observer gives exponential stability of the estimation error when pushed sufficiently fast, enabling robustness to be stated, and moreover, the convergence rate can be arbitrary. Note that arbitrarily fast convergence from the initial time cannot be achieved in discrete time due to the absence of instantaneous observability. More importantly, our high-gain observer exploits constructibility or backward distinguishability, which are among the weakest observability conditions.

2.1.5.2 KKL Observers

The Kravaris-Kazantzis$\slash$Luenberger[KKL] observer is of particular interest in the nonlinear setting thanks to its requiring a fairly mild backward distinguishability—see in Definition 13. This observer consists of transforming system (49) into a stable filter of the output

\[ \begin{equation} z_{k+1} = Az_k + B y_k, \end{equation} \]

(62)

where \( A\) is Schur. An observer readily exists in these target \( z\) -coordinates by running system (62) fed with \( y_k\) from any initial condition, and inverting this transformation recovers the estimate in the original coordinates. This paradigm thus fits the general structure (50), but with a particular form in the \( z\) -coordinates and with \( (\Upsilon_k)_{k \in \mathbb{N}}\) being a left inverse of the transformation, which can be obtained if the latter is (uniformly) injective. In [102], for autonomous systems, this injectivity is established under a mild backward distinguishability condition, constituting a systematic observer that applies to a large class of systems. Unfortunately, while injectivity without uniformity in time is sufficient in the time-invariant setting, this is not the case for non-autonomous systems, which require a strong and uniform kind of injectivity. In Section 2.3, we propose the KKL design for non-autonomous systems, which provides a systematic arbitrarily fast robust observer design under a stronger uniform Lipschitz backward distinguishability condition on the system.

Notably, although they both rely on transforming the given dynamics into linear dynamics with output injection, the crucial distinction between KKL designs and linearization-based techniques reviewed in Section 2.1.3 is that the former does not require a linear output map in the new coordinates (62) (we do not even need its expression), leading to much more generic global results as the class of systems where the method is applicable is much wider. Compared to the MHO discussed in Section 2.1.4.2, the KKL observer is more systematic but is more limited in the computation of the transformation (for which learning-based methods are studied, such as [103]) and the tuning of robustness.

2.1.6 Conclusion

This chapter comprehensively reviews discrete-time observer designs and their properties (observability, robustness, etc.), as summarized in Table 1. Notably, there is limited literature on global designs that assume constructibility or backward distinguishability while also ensuring robustness against uncertainties. In Section 2.2 and Section 2.3, we introduce the novel discrete-time high-gain and KKL observers. These findings play a key role in the development of hybrid observers in this dissertation, as many designs in Section 3 leverage the robustness of these nonlinear discrete-time observers to effectively integrate with continuous-time observers.

Table 1. Summary of discrete-time observer results.
Linearity of systemObservability conditionLocalness of resultsConvergence typeRobustness
LMI-based (Section 2.1.2.1)Linear, Lipschitz nonlinearDetectability of linear partGlobalAsymptoticYes
Kalman(-like) (Section 2.1.2.2)LinearUCOGlobalAsymptoticYes
Immersion into linear form (Section 2.1.3)NonlinearObservability of target formLocalAsymptoticNot guaranteed
EKF (Section 2.1.3)NonlinearUCO of linearization along the estimateLocalAsymptoticNot guaranteed
Dead-beat (Section 2.1.4.1)Nonlinear(Re)constructibility or observabilityGlobalFinite-timeNo results
MHO (Section 2.1.4.2)Nonlinear(Re)constructibility or observabilityGlobalFinite-timeYes
“Classical” high-gain (Section 2.1.5.1)Nonlinear(Re)constructibility or backward distinguishabilityGlobalFinite-time or asymptoticYes
KKL (Section 2.1.5.2)NonlinearBackward distinguishabilityGlobalFinite-time or asymptoticYes

2.2 Constructible Canonical Form and High-Gain Observer in Discrete Time

 

Ce chapitre présente une nouvelle forme triangulaire, que l’on prouve être canonique pour les systèmes constructibles en temps discret. Pour cette forme, nous proposons un observateur qui ressemble à l’observateur grand gain bien connu en temps continu. L’observateur proposé en temps discret est exponentiellement stable lorsque son gain est choisi suffisamment petit, et a une robustesse face aux perturbations et au bruit de mesure. De plus, nous explorons des méthodes pour transformer des systèmes généraux en temps discret en cette forme constructible, sous des conditions de constructibilité et de distinguabilité en temps rétrograde, et récupérer la convergence asymptotique de l’erreur d’estimation dans les coordonnées originales.

2.2.1 Introduction

The content of this chapter has been published in [104]. We first introduce in Section 2.2.2 a novel triangular form, which is proven to be canonical for constructible discrete-time systems. For this form, we then propose in Section 2.2.3 an observer that mirrors the well-known high-gain observer in continuous time. The proposed discrete-time observer demonstrates exponential stability when its dynamics are sufficiently fast, as well as robustness against disturbances and measurement noise. Additionally, we explore methods to transform general discrete-time systems into this constructible form, under conditions of constructibility and backward distinguishability, and recover asymptotic convergence of the estimation error in the original coordinates. This observer will be applied to a Permanent Magnet Synchronous Motor[PMSM] in Section 2.4.

2.2.2 On Notion of Constructibility

In this chapter, we introduce and analyze the notion of constructibility of nonlinear (time-varying) discrete-time systems. Consider systems of the form

\[ \begin{equation} x_{k+1} = f_k(x_k,y_k), ~~ y_k = h_k(x_k), \end{equation} \]

(63)

where \( x_k \in \mathbb{R}^{n_x}\) and \( y_k \in \mathbb{R}^{n_y}\) are the state and the measured output at discrete time \( k\in \mathbb{N}\) ; \( f_k: \mathbb{R}^{n_x}\times\mathbb{R}^{n_y} \to \mathbb{R}^{n_x}\) and \( h_k: \mathbb{R}^{n_x} \to \mathbb{R}^{n_y}\) are the dynamics and output maps. The time dependence of \( f_k\) and \( h_k\) may capture their dependence on inputs \( u_k\) , seen as known functions of time. The goal here is to build for system (63) an asymptotic observer as defined in Definition 6.

2.2.2.1 Constructibility and Constructible Canonical Form

This chapter is based on the following definition.

Definition 9 (Constructibility of order \( m\) of system (63))

System (63) is constructible (or in some literature, reconstructible) of order \( m\) if there exist \( m \in \mathbb{N}\) and a map sequence \( (\Psi_k)_{k \in \mathbb{N}_{\geq m}}\) such that for all solutions \( k\mapsto x_k\) to system (63), we have

\[ \begin{equation} x_k = \Psi_k(y_{k-1},\ldots,y_{k-m}), ~~ k \in \mathbb{N}_{\geq m}. \end{equation} \]

(64)

In the linear context, this property is well-known to be weaker than observability when the dynamics are not invertible [73]. Observability, as defined in Definition 10, would instead require \( x_k\) to be a function of the future outputs (or \( x_{k-m}\) function of \( (y_{k-1},…,y_{k-m})\) in (64)), which is easily checked with the Kalman criterion but is not necessary for observer design. In the nonlinear context, constructibility notions are studied in [71], but from the local point of view of differential geometry, and in [70, Proposition \( 1\) ] as a local property. It is also related to the notion of backward distinguishability exploited in KKL designs in [102] or Definition 11. But otherwise, in general, observability notions are used instead, by exploiting the “observability map” gathering the outputs as a function of the initial state (instead of final), as reviewed in Section 2.1.

Definition 10 (Observability of order \( m\) of system (63))

System (63) is observable of order \( m\) if there exist \( m \in \mathbb{N}\) and a map sequence \( (\Phi_k)_{k \in \mathbb{N}}\) such that for all solutions \( k\mapsto x_k\) to system (63), we have

\[ \begin{equation} x_k = \Phi_k(y_k,y_{k+1},\ldots,y_{k+(m-1)}), ~~ k \in \mathbb{N}. \end{equation} \]

(65)

Remark 1

Note that if, in Definition 9, we write instead \( x_k = \Psi_k^\prime(y_k,…,y_{k-(m-1)})\) for all \( k \in \mathbb{N}_{\geq m-1}\) (as in [70]), it is implied that (64) holds with \( \Psi_k = f_{k-1} \circ \Psi_{k-1}^\prime\) . Also, observability implies constructibility, as stated in Lemma 7. The converse of these statements is true if each \( f_k\) is invertible.

Lemma 5 below shows that the constructibility of system (63) is necessary and sufficient for it to be transformed into what we call a constructible canonical form (67).

Lemma 5 (Constructibility equals transformation into a constructible form)

The following statements are equivalent:

  1. System (63) is constructible of order \( m\) ;
  2. There exist map sequences \( (\mathcal{T}_k)_{k \in \mathbb{N}_{\geq m}}\) , \( (\varphi_{i,k})_{k \in \mathbb{N}}\) with \( i \in \{1,2,…,m\}\) , and \( (\theta_k)_{k \in \mathbb{N}_{\geq m}}\) such that for all solutions \( k \mapsto x_k\) to system (63), we have

    \[ \begin{equation} x_k = \mathcal{T}_k(z_k), ~~ k \in \mathbb{N}_{\geq m}, \end{equation} \]

    (66)

    with \( k \mapsto z_k\) solution for all \( k \in \mathbb{N}_{\geq m}\) to the dynamics

    \[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} z_{1,k+1} & = & \varphi_{1,k}(y_k)\\ z_{2,k+1} & = & \varphi_{2,k}(z_{1,k},y_k)\\ \ldots&&\\ z_{i,k+1} & = & \varphi_{i,k}(z_{1,k},\ldots,z_{i-1,k},y_k)\\ \ldots &&\\ z_{m,k+1} & = & \varphi_{m,k}(z_{1,k},\ldots,z_{m-1,k},y_k), \end{array}\right. \end{equation} \]

    (67.a)

    with the measured output

    \[ \begin{equation} y_k = \theta_k(z_k). \end{equation} \]

    (67.b)

Proof. First, if Item (II) holds, then by Definition 9, there exists \( (\Psi_k)_{k \in \mathbb{N}_{\geq m}}\) and we define for all \( k \in \mathbb{N}_{\geq m}\) the maps \( \mathcal{T}_k = \Psi_k\) and \( \theta_k = h_k \circ \mathcal{T}_k\) , and for all \( k \in \mathbb{N}\) the maps \( \varphi_{i,k}\) , \( i \in \{1,2,…,m\}\) as \( \varphi_{1,k} = \rm{Id}\) , \( \varphi_{2,k}(z_1,y) = z_1\) , \( \varphi_{3,k}(z_1,z_2,y) = z_2\) , etc., which means Item (II) implies Item (III). On the other hand, if Item (III) holds, then because we have for all \( k \in \mathbb{N}_{\geq m}\) , \( z_{1,k} = \varphi_{1,k-1}(y_{k-1})\) , \( z_{2,k} = \varphi_{2,k-1}(z_{1,k-1},y_{k-1}) = \varphi_{2,k-1}(\varphi_{1,k-2}(y_{k-2}),y_{k-1})\) , etc., we have \( x_k = \mathcal{T}_k(z_k) = \mathcal{T}_k(\varphi_{1,k-1}(y_{k-1}),\varphi_{2,k-1}(\varphi_{1,k-2}(y_{k-2}),y_{k-1}),…)\) for all \( k \in \mathbb{N}_{\geq m}\) , which is a function of only \( (y_{k-1},y_{k-2},…,y_{k-m})\) and that corresponds to \( \Psi_k\) in Definition 9, so Item (III) implies Item (II). \( \blacksquare\)

Compared to [88, Section 4] and [71] where the constructible forms are linear (modulo output injection), our form (67) comes with as much nonlinearity in the dynamics and output as possible, allowing us to consider a large class of systems.

The following examples illustrate that we can rely on constructibility to transform the system into the constructible canonical form (67).

Example 5 (Using constructibility to transform into constructible form)

Consider the system in [71, Section I], with dynamics and output:

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} x_{1,k+1} & = & u_k\\ x_{2,k+1} & = & x_{3,k}\\ x_{3,k+1} & = & x_{1,k} + x_{2,k} u_k \end{array}\right. ~~ y_k = x_{3,k}, \end{equation} \]

(68)

where \( u_k\) is some known input. This system is not observable (see in [71, Example \( 1\) ]), but it is constructible because \( x_k\) for all \( k \in \mathbb{N}_{\geq 3}\) can be expressed in the past \( y_k\) and \( u_k\) . We see that \( x_k = (z_{2,k},z_{1,k},z_{3,k})\) , where \( z_k\) follows dynamics of the form (67):

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} z_{1,k+1} & = & y_k\\ z_{2,k+1} & = & u_k\\ z_{3,k+1} & = & z_{1,k} u_k + z_{2,k} \end{array}\right. ~~ y_k = z_{3,k}. \end{equation} \]

(69)

Example 6 (Using constructibility to transform into constructible form (bis))

Consider the system inspired from [71, Example \( 3\) ] with dynamics and output:

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} x_{1,k+1} & = & x_{1,k}x_{2,k}^2 + x_{3,k}u_k\\ x_{2,k+1} & = & x_{3,k}^2u_k^2\\ x_{3,k+1} & = & x_{1,k} \end{array}\right. ~~ y_k = x_{1,k} + u_k, \end{equation} \]

(70)

where \( u_k\) is some known input. This system is not observable (see in [71, Example \( 3\) ] for similar reasoning), but it is constructible. We see that \( x_k = (z_{3,k},z_{2,k},z_{1,k})\) , where \( z_k\) follows dynamics of the form (67):

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} z_{1,k+1} & = & y_k - u_k\\ z_{2,k+1} & = & z_{1,k}^2u_k^2\\ z_{3,k+1} & = & (y_k - u_k)z_{2,k}^2 + z_{1,k} u_k \end{array}\right. ~~ y_k = z_{3,k} + u_k. \end{equation} \]

(71)

However, such transformations are situational since it is not clear how they can be obtained from constructibility. We then propose in the next part a constructive way, when possible, to find this transformation from the system’s maps.

2.2.2.2 Transformation into a Constructible Canonical Form

We introduce the following definition.

Definition 11 (Backward distinguishability of order \( m\) of system (63))

System (63) is backward distinguishable of order \( m\) if there exist \( m \in \mathbb{N}\) and a sequence of maps \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}_{\geq m}}\) with

\[ \begin{equation} \mathcal{O}^{bw}_k(x) =(h^{bw}_{-1,k}(x), h^{bw}_{-2,k}(x), \ldots, h^{bw}_{-m,k}(x)), \end{equation} \]

(72.a)

such that for all \( k \in \mathbb{N}_{\geq m}\) ,

\[ \begin{equation} h^{bw}_{-i,k} (x_k) = y_{k-i}, ~~ \forall i \in \{1,2,\ldots,m\}, \end{equation} \]

(72.b)

along any solution \( k \mapsto x_k\) to system (63), and \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}_{\geq m}}\) is injective.

Remark 2

Definition 11 is the constructibility counterpart of the observability condition assumed in [94, 97, 95, 98, 68, 69, 99]. If \( f_k\) is independent of \( y_k\) and each map \( f_k\) is invertible with the corresponding inverse function \( f^{-1}_k\) defined on \( \mathbb{R}^{n_x}\) , we can define \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}_{\geq m}}\) based on the inverses as in [102] or Definition 12.

Lemma 6 (Backward distinguishability implies transformation into a constructible form)

If system (63) is backward distinguishable of order \( m\) , then the variable \( z_k = \mathcal{O}^{bw}_k(x_k) \in \mathbb{R}^{n_z}\) where \( n_z = mn_y\) , along the solutions \( k \mapsto x_k\) to system (63), for all \( k \in \mathbb{N}_{\geq m}\) , is solution to the dynamics

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} z_{1,k+1} & = & y_k\\ z_{2,k+1} & = & z_{1,k}\\ \ldots&&\\ z_{i,k+1} & = & z_{i-1,k}\\ \ldots &&\\ z_{m,k+1} & = & z_{m-1,k}, \end{array}\right. \end{equation} \]

(73.a)

with the measured output

\[ \begin{equation} y_k = h_k(\mathcal{O}^{bw,-1}_k(z_k)), \end{equation} \]

(73.b)

where \( \mathcal{O}^{bw,-1}_k\) is the left inverse of \( \mathcal{O}^{bw}_k\) on \( \mathbb{R}^{n_x}\) . Moreover,

\[ \begin{equation} x_k = \mathcal{O}^{bw,-1}_k(z_k), ~~ \forall k \in \mathbb{N}_{\geq m}. \end{equation} \]

(74)

Proof. First, by definition of \( h_{-1,k}\) , along the solutions \( k \mapsto x_k\) to system (63), we have for all \( k \in \mathbb{N}_{\geq m}\) , \( z_{1,k+1} =h_{-1,k+1}(x_{k+1})= y_k\) . Then, for each \( i \in \{2,…,m\}\) , by definition of \( h_{-i,k}\) , along the solutions \( k \mapsto x_k\) to system (63), we have for all \( k \in \mathbb{N}_{\geq m}\) , \( z_{i,k+1} = h_{-i,k+1}(x_{k+1}) = h_{k-(i-1)}(x_{k-(i-1)}) = h_{-(i-1),k}(x_k) = z_{i-1,k}\) , concluding the proof. \( \blacksquare\)

This property is used in Section 2.4 to design an observer for a discretized PMSM. The form (73), obtained through backward distinguishability, is a particular case of the form (67) and was also obtained under constructibility in the proof of Lemma 5. The difference is that in (72), the past outputs are expressed as an injective function of \( x_k\) (and thus \( z_k\) is a function of \( x_k\) in (73) and vice-versa), while in (64), \( x_k\) is directly written as a function of the past outputs (and thus we only have \( x_k\) as a function of \( z_k\) in (67)). In other words, backward distinguishability is sufficient, but not necessary, for constructibility. This is shown in Lemma 7 and is illustrated in Example 7.

Example 7 (Backward distinguishability is sufficient but not necessary for constructibility)

Consider the system with dynamics and output

\[ \begin{equation} x_{k+1} = x_k^2, ~~ y_k = x_k. \end{equation} \]

(75)

We have \( x_k = y_{k-1}^2\) for all \( k \in \mathbb{N}_{\geq 1}\) so that this system is constructible. However, given a current state \( x_k\) at a time \( k \in \mathbb{N}_{\geq 1}\) , there exist two corresponding past outputs \( y_{k-1}\) and \( -y_{k-1}\) , so that we cannot write the map (72) and this system is therefore not backward distinguishable.

The relationships among the introduced notions and properties are captured in Lemma 7.

Lemma 7 (Relationships between observability conditions and observer design)

Consider system (63). We have the following implications among its properties:

\[ \begin{array}{@{}c@{~~}c@{~~}c@{}} \text{Observability} & & \\ (\text{I}1){\big\Downarrow} & & \\ \text{Constructibility}& \xLeftarrow{(\text{I}2)} & \text{Backward distinguishability}\\ (\text{I}3)\big\Downarrow& & \\ \text{Existence of an asymptotic observer}& \xRightarrow{(\text{I}4)} & \text{Asymptotic detectability} \end{array} \]

And, the converse of implication (I\( 1\) ) holds if the dynamics of system (63) are invertible.

Proof. We consider system (63) and prove each implication individually.
For (I\( 1\) ): If system (63) is observable, from (65), we have \( x_k =F_k(x_{k-m}) = (F_k \circ \Phi_{k-m}) (y_{k-m},y_{k-(m-1)},…,y_{k-1})\) for all \( k \in \mathbb{N}_{\geq m}\) , where \( F_k = (f_{k-1} \circ f_{k-2} \circ … \circ f_{k-m})\) . From Definition 9, system (63) is constructible. Now for the converse, assume that the dynamics of system (63) are invertible, so each \( f_k\) is invertible as \( f_k^{-1}\) for all \( k \in \mathbb{N}\) . If system (63) is constructible, from (64), we have \( x_k =F_k^\prime(x_{k+(m-1)}) = (F_k^\prime \circ \Psi_{k+(m-1)}) (y_{k+(m-1)},y_{k+(m-2)},…,y_k)\) for all \( k \in \mathbb{N}\) , where \( F_k^\prime = (f_k^{-1} \circ f_{k+1}^{-1} \circ … \circ f_{k+(m-2)}^{-1})\) . From Definition 10, system (63) is observable.
For (I\( 2\) ): Assume that system (63) is backward distinguishable. First, we use Lemma 6 to show that this implies the existence of a transformation into a particular constructible form (73). Then, we use Lemma 5 to show that this constructible form is equivalent to system (63) being constructible.
For (I\( 3\) ): If system (63) is constructible, by Lemma 5, (67)-(66) is an asymptotic observer for it.
For (I\( 4\) ): See in the proof of Lemma 4. \( \blacksquare\)

2.2.3 High-Gain Observer for a Constructible Form

2.2.3.1 Observer Design

Consider a system in the constructible canonical form (67) where \( z_k \in \mathbb{R}^{n_z}\) is the state, and \( y_k\) is the measured output in \( \mathbb{R}^{n_y}\) . For system (67), we propose the following observer:

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{z}_{1,k+1} & = & \bar\varphi_{1,k}(y_k)+ \gamma^m k_1 (y_k - \bar \theta_k(\hat{z}_k))\\ \hat{z}_{2,k+1} & = & \bar{\varphi}_{2,k}(\hat{z}_{1,k},y_k)+ \gamma^{m-1} k_2 (y_k - \bar{\theta}_k(\hat{z}_k))\\ \ldots&&\\ \hat{z}_{i,k+1} & = & \bar{\varphi}_{i,k}(\hat{z}_{1,k},\ldots,\hat{z}_{i-1,k},y_k)+ \gamma^{m-i+1} k_i (y_k - \bar{\theta}_k(\hat{z}_k))\\ \ldots &&\\ \hat{z}_{m,k+1} & = & \bar{\varphi}_{m,k}(\hat{z}_{1,k},\ldots,\hat{z}_{m-1,k},y_k)+ \gamma k_m(y_k - \bar{\theta}_k(\hat{z}_k)), \end{array}\right. \end{equation} \]

(76)

where \( k_i \in \mathbb{R}\) , \( i \in \{1,2,…,m\}\) and \( \gamma \in [0,1]\) are design parameters to be selected, which may be \( 0\) , and the maps \( \bar\varphi_{i,k}\) , \( i \in \{2,…,m\}\) and \( \bar\theta_k\) are such that Item (IV) of Assumption 3 below holds.

Assumption 3

Assume that:

  1. There exist sets \( \mathcal{Z}_0 \subset \mathbb{R}^{n_z}\) , \( \mathcal{Y} \subset \mathbb{R}^{n_y}\) , \( \mathcal{V} \subset \mathbb{R}^{n_z}\) , and \( \mathcal{Z} \subset \mathbb{R}^{n_z}\) such that the solutions \( k \mapsto z_k\) to system (67) of interest are initialized in \( \mathcal{Z}_0\) , have outputs \( y_k \in \mathcal{Y}\) and possibly disturbance \( v_k \in \mathcal{V}\) for all \( k \in \mathbb{N}\) , and remain in \( \mathcal{Z}\) for all \( k \in \mathbb{N}\) ;
  2. The maps \( \varphi_{i,k}\) , \( \bar\varphi_{i,k}\) , \( i \in \{1,2,…,m\}\) and \( \theta_k\) , \( \bar\theta_k\) are such that there exist \( L_{z,i}, L_{y,i} > 0\) (for each \( i \in \{1,2,…,m\}\) ) and \( L_\theta > 0\) such that for all \( (z,\hat{z},k,y,v) \in \mathcal{Z} \times \mathbb{R}^{n_z} \times \mathbb{N} \times \mathcal{Y}\times \mathbb{R}^{n_y}\) ,

    \[ \begin{align} |\varphi_{i,k}(z,y)-\bar\varphi_{i,k}(\hat{z},y + v)| &\leq L_{z,i}|z - \hat{z}| + L_{y,i}|v|, \\ |\theta_k(z) - \bar\theta_k(\hat{z})| &\leq L_\theta |z - \hat{z}|. \\\end{align} \]

    (77.a)

Remark 3

In the case where \( \varphi_{i,k}\) (resp., \( \theta_k\) ) is globally Lipschitz with respect to \( z_k\) (uniformly in \( k \in \mathbb{N}\) ), we can take \( \bar\varphi_{i,k} = \varphi_{i,k}\) (resp., \( \bar\theta_k = \theta_k\) ). In another case where the set \( \mathcal{Z}\) is compact and the map \( \varphi_{i,k}\) is locally Lipschitz with respect to \( z_k\) , uniformly in \( k \in \mathbb{N}\) and \( y_k \in \mathcal{Y}\) (resp., the map \( \theta_k\) is locally Lipschitz with respect to \( z_k\) , uniformly in \( k \in \mathbb{N}\) ), we take \( \bar\varphi_{i,k}\) (resp., \( \bar\theta_k\) ) as a bounded map that coincides with \( \varphi_{i,k}\) (resp., \( \theta_k\) ) for all \( z_k \in \mathcal{Z} + c\) , for some \( c > 0\) . This way, Item (IV) of Assumption 3 is satisfied.

Because system (67) is constructible for all \( k \in \mathbb{N}_{\geq m}\) , we can neglect the correction terms in observer (76), i.e., pick \( \gamma = 0\) for an instantaneous convergence after \( m\) time steps. However, the price to pay is the absence of filtering effects against disturbances and noise. On the other hand, the observer structure (76) resembles the famous high-gain design in continuous time (26) discussed in Section 1.5.2.2, the difference being that: i) the triangularity constraint is lower diagonal, ii) all the maps may be nonlinear, and iii) there are no constraints on the choice of the scalars \( k_i\) , \( i \in \{1,2,…,m\}\) .

Remark 4

Further generalizations of system (67) and correspondingly observer (76) can be realized:

  • The correction term in observer (76) can be replaced by \( \Upsilon_k(y_k) - \bar\Upsilon_k(\bar\theta_k(\hat{z}_k))\) with \( \Upsilon_k\) any map that is locally Lipschitz in \( y_k \in \mathcal{Y}\) , uniformly in \( k \in \mathbb{N}\) , and \( \bar\Upsilon_k\) a globally Lipschitz map;
  • Dependence on the history of inputs \( k \mapsto u_k\) and \( k \mapsto y_k\) on a window can be considered via

    \[ \begin{equation} \overline{u}_{k+1} = A_u \overline{u}_k + B_u u_k, ~~ \overline{y}_{k+1} = A_y \overline{y}_k + B_y y_k, \end{equation} \]

    (78)

    for some matrices \( (A_u,B_u,A_y,B_y)\) of appropriate dimensions. The Lipschitzness of the maps as in Item (IV) of Assumption 3 must then be uniform in these.

Theorem 1 shows the exponential stability of the estimation error with an arbitrarily fast rate, achievable by pushing \( \gamma\) smaller.

Theorem 1 (Exponential stability with an arbitrarily fast rate)

Under Assumption 3, for any choice of \( k_i\) , \( i \in \{1,2,…,m\}\) , there exists \( \gamma^\star > 0\) such that any solution \( k \mapsto z_k\) to system (67) initialized in \( \mathcal{Z}_0\) with \( y_k \in \mathcal{Y}\) for all \( k \in \mathbb{N}\) and any solution \( k \mapsto \hat{z}_k\) to observer (76) with \( 0 < \gamma < 1\) , initialized in \( \mathbb{R}^{n_z}\) and fed with \( y_k\) in (67.b) verify:

\[ \begin{equation} |z_k - \hat{z}_k| \leq \frac{1}{\gamma^{m-1}}\left(\frac{\gamma}{\gamma^\star}\right)^k|z_0 - \hat{z}_0|, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(79)

Proof. This technical proof has been moved to Section 6.1.1.1 to facilitate reading. It involves two main steps: i) Obtain the estimation error dynamics and re-scale these, and ii) Bound nonlinear terms by their Lipschitzness and push \( \gamma\) to dominate these terms. \( \blacksquare\)

Remark 5

Notice that, unlike the continuous-time high-gain observer, this observer is arbitrarily fast only after \( m\) steps. Our design differs from those in [88, Section 4] and [71], which are linear modulo output injection. Here, we try to be as general as possible by allowing \( z_{i,k+1}\) to depend on not only \( z_{i-1,k}\) but also the whole \( (z_{1,k},…,z_{i-1,k})\) , and the output \( y_k\) to be nonlinear in \( z_k\) .

2.2.3.2 Robustness of the Observer

Consider system (67) with disturbance \( v_k \in \mathbb{R}^{n_z}\)

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}l@{}} z_{1,k+1} & = & \varphi_{1,k}(y_k) &{}+ v_{1,k}\\ z_{2,k+1} & = & \varphi_{2,k}(z_{1,k},y_k) &{} + v_{2,k}\\ \ldots&&\\ z_{i,k+1} & = & \varphi_{i,k}(z_{1,k},\ldots,z_{i-1,k},y_k) &{}+ v_{i,k}\\ \ldots &&\\ z_{m,k+1} & = & \varphi_{m,k}(z_{1,k},\ldots,z_{m-1,k},y_k) &{}+ v_{m,k}, \end{array}\right. \end{equation} \]

(80.a)

and measurement noise \( w_k \in \mathbb{R}^{n_y}\) added to the output

\[ \begin{equation} y_k + w_k. \end{equation} \]

(80.b)

The disturbance \( v_{i,k}\) could also model the non-Lipschitzness of \( \varphi_{i,k}\) . Theorem 2 below shows the robustness of observer (76) under the considered disturbance and noise in system (80).

Theorem 2 (Robustness against disturbances and measurement noise)

Under Assumption 3, for any choice of \( k_i\) , \( i \in \{1,2,…,m\}\) , there exists \( \gamma^\star > 0\) such that any solution \( k \mapsto z_k\) to system (80) initialized in \( \mathcal{Z}_0\) with \( y_k \in \mathcal{Y}\) and \( v_k \in \mathcal{V}\) for all \( k \in \mathbb{N}\) and any solution \( k \mapsto \hat{z}_k\) to observer (76) with \( 0< \gamma < 1\) , initialized in \( \mathbb{R}^{n_z}\) and fed with \( y_k + w_k\) in (80.b) verify for each \( i \in \{1,2,…,m\}\) , for all \( k \in \mathbb{N}\) , and for all \( j \in \{0,…,k-1\}\) :

\[ \begin{multline} |z_{i,k} - \hat{z}_{i,k}| \leq \frac{1}{\gamma^{i-1}}\left(\frac{\gamma}{\gamma^\star}\right)^k |z_0-\hat{z}_0| +\sum_{j = 0}^{k-1}\left(\frac{\gamma}{\gamma^\star}\right)^{k-1-j}\sum_{q = 1}^m \gamma^{q-i}|v_{q,j}|\\ +\sum_{j = 0}^{k-1}\left(\frac{\gamma}{\gamma^\star}\right)^{k-1-j}\sum_{q = 1}^m\left(\gamma^{q-i}L_{y,q}+\gamma^{m-i+1} |c_q|\right)|w_j|. \end{multline} \]

(81)

Proof. This technical proof is an adaptation of that of Theorem 1 taking into account the uncertainties. It has been moved to Section 6.1.1.2 to facilitate reading. \( \blacksquare\)

From Theorem 2, we see that with \( 0 < \gamma < \min\{1,\gamma^\star\}\) :

  • The estimation error is robustly stable with respect to disturbance \( v_k\) and noise \( w_k\) , in the sense of [46, Definition \( 2.3\) ], which differs from the classical Input-to-State Stability[ISS] in [45] by the exponentially penalization of the past values of \( (v_k,w_k)\) thanks to the forgetting factor \( \left(\frac{\gamma}{\gamma^\star}\right)^{k-1-j}\) ;
  • Similarly to the high-gain design in continuous time (see e.g., [53]), the effects of \( v_{q,j}\) (past disturbance on line \( q\) ) on \( \tilde{z}_{i,k}\) (current estimation error on line \( i\) ) can be either magnified or reduced depending on \( q\) with respect to \( i\) , because of the coefficient \( \gamma^{q-i}\) . Note that, however, the impact of the disturbance on the last line \( (v_{m,j})\) can only be reduced (for \( \gamma\) sufficiently small) on the lines \( i<m\) , which does not give practical convergence, contrary to continuous time;
  • In the proof of Theorem 2, it is not evident that we can attenuate the disturbance and noise by choosing \( k_i\) in observer (76) since this proof is done conservatively for the general case of nonlinear maps \( \varphi_{i,k}\) . However, for the specific form (73) widely used in Moving Horizon Observer[MHO] schemes, by picking \( k_1 < 0\) and \( k_i=0\) for \( i\neq0\) , we get a penalization factor in front of the noise.

2.2.3.3 Asymptotic Convergence in the Original \( x\) -Coordinates

In Section 2.2.2, we have seen that constructible systems (63) can be linked to constructible form (67) via some map (66), at least after \( m\) discrete steps. Then, an observer (76) can be designed, and we now study conditions to recover the asymptotic convergence in the \( x\) -coordinates. For this, the following assumption is made.

Assumption 4

There exist a closed set \( \mathcal{Z}\) and \( m\in \mathbb{N}\) such that system (63) is constructible of order \( m\) , with \( k \mapsto z_k\) in Lemma 6 such that \( z_k \in \mathcal{Z}\) for all \( k \in \mathbb{N}\) , and with \( (\mathcal{T}_k)_{k \in \mathbb{N}_{\geq m}}\) uniformly continuous on \( \mathcal{Z}\) for all \( k \in \mathbb{N}_{\geq m}\) , i.e., there exists a class-\( \mathcal{K}\) function \( \rho\) such that for all \( (z_a,z_b) \in \mathcal{Z} \times \mathcal{Z}\) and for all \( k \in \mathbb{N}_{\geq m}\) ,

\[ \begin{equation} |\mathcal{T}_k(z_a) - \mathcal{T}_k(z_b)| \leq \rho(|z_a - z_b|). \end{equation} \]

(82)

Remark 6

Assumption 4 is satisfied if either:

  • System (63) is uniformly constructible of order \( m\) , i.e., the map sequence \( (\Psi_k)_{k \in \mathbb{N}_{\geq m}}\) in (64) is uniformly continuous; or
  • System (63) has solutions remaining in \( \mathcal{X}\) and is uniformly backward distinguishable of order \( m\) on \( \mathcal{X}\) , i.e., the map sequence \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}_{\geq m}}\) in (72) is uniformly injective on \( \mathcal{X}\) for all \( k \in \mathbb{N}_{\geq m}\) . More precisely, there exists a class-\( \mathcal{K}\) function \( \rho\) such that \( |\mathcal{O}^{bw}_k(x_a) - \mathcal{O}^{bw}_k(x_b)| \geq \rho(|x_a - x_b|) \) for all \( (x_a,x_b) \in \mathcal{X} \times \mathcal{X}\) and for all \( k \in \mathbb{N}_{\geq m}\) .

In these cases, we have \( n_z = m n_y\) .

The asymptotic convergence is then recovered in the \( x\) -coordinates as follows.

Lemma 8 (Asymptotic convergence in the \( x\) -coordinates)

Under Assumption 4, any solution \( k \mapsto x_k\) to system (63) and any solution \( k \mapsto \hat{z}_k\) to observer (76) with \( 0 < \gamma < \min\{1,\gamma^\star\}\) and fed with \( y_k\) in (63) verify \( \lim_{k\to+\infty}|x_k - \hat{x}_k| = 0, \) where \( \hat{x}_k = \bar{\mathcal{T}}_k(\hat{z}_k)\) , with \( k \in \mathbb{N}_{\geq m}\) , and \( (\bar{\mathcal{T}}_k)_{k \in \mathbb{N}_{\geq m}}\) is a sequence of extensions of \( (\mathcal{T}_k)_{k \in \mathbb{N}_{\geq m}}\) that is uniformly continuous on \( \mathbb{R}^{n_z}\) .

Proof. Based on (82) and using [40], we can extend \( (\mathcal{T}_k)_{k \in \mathbb{N}_{\geq m}}\) into \( (\bar{\mathcal{T}}_k)_{k \in \mathbb{N}_{\geq m}}\) such that there exists \( c_1 > 0\) such that for all \( (z_a,z_b) \in \mathbb{R}^{n_z} \times \mathbb{R}^{n_z}\) and for all \( k \in \mathbb{N}_{\geq m}\) ,

\[ \begin{equation} |\bar{\mathcal{T}}_k(z_a) - \bar{\mathcal{T}}_k(z_b)| \leq c_1\rho(|z_a - z_b|). \end{equation} \]

(83)

Along any solution \( k \mapsto x_k\) to system (63) and any solution \( k \mapsto \hat{z}_k\) to observer (76) with \( 0 < \gamma < \min\{1,\gamma^\star\}\) and fed with \( y_k\) in (63), we have

\[ \begin{align*} |x_k - \hat{x}_k| = |\bar{\mathcal{T}}_k(z_k) - \bar{\mathcal{T}}_k(\hat{z}_k)|\leq c_1 \rho(|z_k - \hat{z}_k|), \end{align*} \]

and we have \( \lim_{k\to+\infty}|z_k - \hat{z}_k| = 0\) for all \( k \in \mathbb{N}_{\geq m}\) thanks to Theorem 1 when \( 0 < \gamma < \min\{1,\gamma^\star\}\) , which concludes this proof. \( \blacksquare\)

Combining the ingredients in Lemma 6Theorem 1, and Lemma 8, we arrive at a constructive observer design for general nonlinear systems of the form (63), under constructibility or backward distinguishability.

2.2.4 Conclusion

We propose a constructible canonical form in discrete time and a corresponding high-gain observer, with a structure resembling that of the well-known continuous-time version. This design exhibits robust exponential stability if the rate is pushed fast enough. We also show how a system can be transformed into a constructible form under constructibility and backward distinguishability, and how convergence can be recovered in the original coordinates. In Section 2.4, we apply this observer to a PMSM and compare it with other designs.

Future work is to understand better the potential gain of performance in terms of noise\( /\) disturbance attenuation by strategically picking the parameters \( k_i\) , \( i \in \{1,2,…,m\}\) . The MHO, which resembles our form where (73) is used to store the past outputs but with an optimizer replacing (74), offers a straightforward way to tune robustness through its forgetting factor. It could be interesting to run in parallel or combine these designs since robustness lies in the \( z\) -coordinates in our case and in the observer output for MHOs. We will also figure out results relevant to the continuous-time high-gain observer, including, for instance, the case of connection with a feedback controller [53], the high-gain Kalman-like design [27], and potentially a peaking-free design [52].

Acknowledgments. We thank Vincent Andrieu and Daniele Astolfi, researchers at LAGEPP-CNRS, for their collaborations in this chapter.

2.3 KKL Observer in Discrete Time

 

Cette chapitre présente la synthèse d’un observateur Kravaris-Kazantzis$\slash$Luenberger[KKL] pour les systèmes nonlinéaires à temps discret temps-variants. Nous donnons d’abord des conditions suffisantes pour l’existence d’une suite de fonctions \( (T_k)_{k \in \mathbb{N}}\) transformant la dynamique du système en un filtre exponentiellement stable de la sortie dans d’autres coordonnées cibles, où un observateur est directement conçu. Ensuite, nous prouvons que sous une hypothèse de distinguabilité uniforme et Lipschitz en temps rétrograde, les applications \( (T_k)_{k \in \mathbb{N}}\) deviennent uniformément Lipschitz-injectives après un certain temps si la dynamique cible est suffisamment rapide. Cela conduit à un observateur à temps discret arbitrairement rapide après un certain temps, qui présente des similitudes avec le célèbre observateur à grand gain pour les systèmes en temps continu. La stabilité entrée-état de l’erreur d’estimation par rapport aux incertitudes, aux perturbations d’entrée et au bruit de mesure est ensuite démontrée. Ensuite, sous la distinguabilité en temps rétrograde qui est plus faible, nous montrons l’injectivité des applications \( (T_k)_{k \in \mathbb{N}}\) après un certain temps pour un choix générique de la dynamique du filtre cible. Des exemples illustrent l’observateur proposé.

2.3.1 Introduction

Among many existing observer design methods [1], the Kravaris-Kazantzis$\slash$Luenberger[KKL] observers [105, 106, 42, 102] are of interest in nonlinear observer design thanks to their beautiful theory revolving around Coron’s Lemma [107, 105]. They consist of transforming the system dynamics (of dimension \( n_x\) ) into an exponentially stable filter of the output in some new coordinates (referred to as the target coordinates, of dimension \( n_z\geq n_x\) ), where an observer readily exists, and inverting this transformation to recover the estimate of the state in the original coordinates under its uniform injectivity. Initially, David Luenberger proposed this method for Linear Time-Invariant[LTI] continuous-time systems in [74]—he showed that an invertible linear transformation into a stable filter of the output always exists as long as the given system is observable and the eigenvalues of the filter are picked different from those of the system. Several attempts were then made to extend this theory to nonlinear continuous-time systems. The existence of a nonlinear transformation was first considered in [108, 109, 110] in the analytic context and around an equilibrium point. Then, the localness was dropped following another perspective in [111] where a global existence result was proposed based on a strong observability assumption which unfortunately did not indicate the necessary dimension of the filter. This problem was solved in [105] by proving the existence of an injective transformation under a mild backward distinguishability condition, for complex-valued filters of dimension \( n_x+1\) , with almost any choice of \( n_x+1\) distinct complex eigenvalues and recently in [112] for almost any real diagonalizable filter of dimension \( 2n_x+1\) , both applied to each output. Stronger uniform injectivity results were also obtained under differential observability conditions, in the case where the eigenvalues of the filter are pushed sufficiently fast [106]. In parallel, this KKL paradigm was also developed for non-autonomous continuous-time systems [42] and for autonomous discrete-time systems [102], under similar backward distinguishability and differential observability conditions. Existing KKL observer results for various system classes are reviewed in Table 2 at the end of this chapter.

Finding an explicit and implementable expression of this transformation, and more importantly, of its left inverse, is possible in particular contexts such as parameter identification [113] or state\( /\) parameter estimation for electrical machines [114, 115]. When an implementable expression for the transformation or its left inverse is not available, numerical approximation methods based on a Neural Network[NN] are being developed as in [103, 116, 117, 118, 119], but essentially for autonomous systems. The computation of those maps in the time-varying setting still remains a challenge, except for particular forms, such as linear dynamics with polynomial output studied in Example 9.

Note that the KKL observer shares some similarities with the Koopman operator [120], which also involves mappings into linear(-like) forms and has recently been explored for control purposes [121]. While the former is exact in the sense that the transformation is constructed systematically and deterministically for a specific system, the latter relies on an infinite-dimensional linear operator into an infinite-dimensional linear form, which can be approximated using a finite number of dimensions by Dynamic Mode Decomposition[DMD] [122], and hence is inherently inexact. DMD, exploited for control in [123], is a data-driven algorithm that finite-dimensionally approximates the Koopman operator by analyzing snapshots of system trajectories to identify dominant modes, eigenvalues, and dynamics. Unlike the KKL approach, a systematic method to determine a sufficiently high-dimensional target space for the Koopman operator has not yet been documented, as the required dimensionality depends on the complexity of the nonlinear system, the choice of observables, and the desired approximation accuracy. Meanwhile, the application of the KKL observer for stabilization has been examined in [124].

The content of this chapter has been published in [101]. It presents the KKL observer design for nonlinear time-varying discrete-time systems. After stating the problem in Section 2.3.2, we give sufficient conditions on the existence of a sequence of functions \( (T_k)_{k \in \mathbb{N}}\) transforming the given system dynamics into an exponentially stable filter of the output in some other target coordinates, where an observer is directly designed. Then, in Section 2.3.4, we prove that under uniform Lipschitz backward distinguishability, the maps \( (T_k)_{k \in \mathbb{N}}\) become uniformly Lipschitz injective after a certain time if the target dynamics have an appropriate dimension and are pushed sufficiently fast. This leads to an observer combining two main features. First, it provides an arbitrarily fast convergence of the estimation error in the system coordinates, as soon as allowed by the distinguishability condition. Second, this KKL design allows us to filter the output and provides after that time robust stability of the estimation error in the sense of [46], with an explicit strict Input-to-State Stability[ISS] Lyapunov function. Such a design may thus be seen as a discrete-time counterpart of the celebrated high-gain observer for continuous-time systems [26], which as far as we know does not exist for discrete-time systems (apart from discretizations of continuous-time high-gain observers [125]). Next, in Section 2.3.5, under the milder backward distinguishability, we show the injectivity of the maps \( (T_k)_{k \in \mathbb{N}}\) after a certain time for a generic choice of the target filter dynamics. Section 2.3.6 analyzes the case of linear time-varying systems and makes the link with the existing Kalman design, then an application of our observer to a discretized Permanent Magnet Synchronous Motor[PMSM] will be presented in Section 2.4.

2.3.2 Problem Statement

Consider the nonlinear time-varying discrete-time system

\[ \begin{equation} x_{k+1} = f_k(x_k), ~~ y_k = h_k(x_k), \end{equation} \]

(84)

where \( f_k: \mathbb{R}^{n_x} \to \mathbb{R}^{n_x}\) and \( h_k: \mathbb{R}^{n_x} \to \mathbb{R}^{n_y}\) are the dynamics and output maps, \( x_k \in \mathbb{R}^{n_x}\) is the state, and \( y_k \in \mathbb{R}^{n_y}\) is the output at discrete time \( k\) .

Remark 7

Any system of the form

\[ \begin{equation} x_{k+1} = f_k(x_k,u_k), ~~ y_k = h_k(x_k,u_k), \end{equation} \]

(85)

where the input \( u_k \in \mathbb{R}^{n_u}\) is a known trajectory of time, can be put into form (84) with the maps \( (f_k,h_k)_{k\in \mathbb{N}}\) depending on a particular sequence of inputs \( (u_k)_{k \in \mathbb{N}}\) . The results of this chapter thus depend on this sequence of inputs, but some can be made uniform with respect to a family of \( (u_k)_{k \in \mathbb{N}}\) , if the corresponding assumptions also hold uniformly in the inputs.

Assumption 5

The solutions to system (84) of interest, initialized in a set \( \mathcal{X}_0\) , remain in a compact set \( \mathcal{X} \supseteq \mathcal{X}_0\) in positive time.

The KKL observer design consists in seeking a sequence of nonlinear maps \( (T_k)_{k \in \mathbb{N}}\) , with \( T_k : \mathbb{R}^{n_x} \to \mathbb{R}^{n_z}\) , transforming the dynamics (84) into a stable discrete-time filter of the output, i.e., such that \( z_k = T_k(x_k)\) verifies

\[ \begin{equation} z_{k+1} = A z_k + B y_k, \end{equation} \]

(86)

where \( A \in \mathbb{R}^{n_z \times n_z}\) is Schur and \( B \in \mathbb{R}^{n_z \times n_y}\) such that \( (A, B)\) is controllable. In other words, we look for \( (T_k)_{k \in \mathbb{N}}\) satisfying

\[ \begin{equation} T_{k+1}(x_{k+1}) = A T_k(x_k) + B h_k(x_k), ~~ \forall k \in \mathbb{N}, \end{equation} \]

(87)

along solutions to system (84) remaining in \( \mathcal{X}\) . A sufficient condition for that is to have

\[ \begin{equation} (T_{k+1} \circ f_k)(x) = A T_k(x) + B h_k(x), ~~ \forall x\in \mathcal{X}: f_k(x)\in \mathcal{X}, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(88)

The observer in the \( z\) -coordinates is then made of a simple filter of the output

\[ \begin{equation} \hat{z}_{k+1} = A \hat{z}_k + B y_k, \end{equation} \]

(89)

since the estimation error then verifies \( (z_{k+1}-\hat{z}_{k+1}) = A (z_k - \hat{z}_k)\) , which is exponentially stable. The following Theorem 3 then shows that if the sequence \( (T_k)_{k \in \mathbb{N}}\) to (88) is uniformly injective after a certain time (as in (91) below), it admits a sequence of left inverses \( (T^*_k)_{k \in \mathbb{N}}\) , with \( T^*_k:\mathbb{R}^{n_z}\to \mathbb{R}^{n_x}\) , such that the observer

\[ \begin{equation} \hat{z}_{k+1} = A \hat{z}_k + B y_k, ~~ \hat{x}_k = T^*_k(\hat{z}_k), \end{equation} \]

(90)

initialized in \( T_0(\mathcal{X})\) , provides an asymptotic estimate \( \hat{x}_k \in \mathbb{R}^{n_x}\) of \( x_k\) , with an asymptotic stability property of the estimation error after a certain time (as in (92) below). The goal of this chapter is then to provide sufficient conditions to guarantee the existence of such a sequence of maps \( (T_k)_{k \in \mathbb{N}}\) .

Theorem 3 (Recovering asymptotic stability in the \( x\) -coordinates)

Assume there exists \( (T_k)_{k \in \mathbb{N}}\) satisfying (88) with \( T_0\) continuous on \( \mathcal{X}\) and \( (T_k)_{k \in \mathbb{N}}\) is uniformly injective after a time, i.e., there exist a concave class-\( \mathcal{K}\) function \( \rho\) and \( k^\star \in \mathbb{N}\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{equation} |x_a - x_b| \leq \rho(|T_k(x_a) - T_k(x_b)|). \end{equation} \]

(91)

Then, there exists \( (T^*_k)_{k \in \mathbb{N}}\) and a class-\( \mathcal{KL}\) function \( \beta\) such that for any solution \( k\mapsto x_k\) to system (84) with \( x_0 \in \mathcal{X}_0\) and any solution \( k\mapsto \hat{z}_k\) to observer (90) with \( \hat{z}_0\in T_0(\mathcal{X})\) and input \( y_k=h_k(x_k)\) , we have for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ \begin{equation} |x_k - \hat{x}_k| \leq \beta(|x_0 - \hat{x}_0|, k). \end{equation} \]

(92)

Remark 8

In this section, the concavity assumption of \( \rho\) is not restrictive because we will achieve, in Theorem 5, uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) characterized by a linear \( \rho\) . In general, this assumption can also be dropped if there exists a compact set \( \mathcal{Z} \subset \mathbb{R}^{n_z}\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) , \( T_k(\mathcal{X}) \subseteq \mathcal{Z}\) .

Proof. From the uniform injectivity of \( (T_k)_{k \in \mathbb{N}}\) in (91), there exists a sequence of left inverse maps \( (T^{-1}_k)_{k \in \mathbb{N}}: T_k(\mathcal{X}) \to \mathbb{R}^{n_x}\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

  • For all \( x \in \mathcal{X}\) , \( T^{-1}_k(T_k(x)) = x\) ;
  • For all \( (z_a, z_b) \in T_k(\mathcal{X}) \times T_k(\mathcal{X})\) , \( |T^{-1}_k(z_a) - T^{-1}_k(z_b)| \leq \rho(|z_a - z_b|)\) .

Applying [40] component-wise, we can extend \( (T^{-1}_k)_{k \in \mathbb{N}}\) into a sequence of left inverse maps \( (T^*_k)_{k \in \mathbb{N}}: \mathbb{R}^{n_z} \to \mathbb{R}^{n_x}\) such that there exists \( c_1 > 0\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

  • For all \( x \in \mathcal{X}\) , \( T^*_k(T_k(x)) = x\) ;
  • For all \( (z_a, z_b) \in \mathbb{R}^{n_z} \times \mathbb{R}^{n_z}\) , \( |T^*_k(z_a) - T^*_k(z_b)| \leq c_1 \rho(|z_a - z_b|)\) .

It follows that for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ \begin{align*} |x_k - \hat{x}_k| &= |T^*_k(T_k(x_k)) - T^*_k(\hat{z}_k)|\\ &\leq c_1 \rho(|T_k(x_k) - \hat{z}_k|) \\ &\leq c_1 \rho(c_2c_3^k |T_0(x_0) - \hat{z}_0|), \end{align*} \]

for some \( c_2 > 0\) and \( c_3 \in (0, 1)\) thanks to the exponential stability in the \( z\) -coordinates given by \( (z_{k+1}-\hat{z}_{k+1}) = A (z_k - \hat{z}_k)\) . Pick \( \hat{x}_0 \in \mathcal{X}\) such that \( \hat{z}_0=T_0(\hat{x}_0)\) . Because \( T_0\) is continuous on the compact set \( \mathcal{X}\) , it is also uniformly continuous on \( \mathcal{X}\) , meaning that there exists a class-\( \mathcal{K}\) function \( \rho_0\) such that for any \( x_0 \in \mathcal{X}_0\) and \( \hat{x}_0 \in \mathcal{X}\) , \( |T_0(x_0) - \hat{z}_0| = |T_0(x_0) - T_0(\hat{x}_0)| \leq \rho_0(|x_0 - \hat{x}_0|)\) . Finally, for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ |x_k - \hat{x}_k| \leq c_1 \rho(c_2c_3^k \rho_0(|x_0 - \hat{x}_0|)), \]

which is a class-\( \mathcal{KL}\) function in \( |x_0 - \hat{x}_0|\) and \( k\) . \( \blacksquare\)

The uniform injectivity of \( (T_k)_{k \in \mathbb{N}}\) as in (91) is thus sufficient to guarantee asymptotic stability of the estimation error. The following academic example shows that it is not necessary, but the injectivity of each map \( T_k\) alone, without uniformity in \( k\) , can sometimes be insufficient to ensure convergence.

Example 8 (Uniform injectivity is sufficient but not necessary)

Consider the first-order time-varying system

\[ \begin{equation} x_{k+1} = x_k, ~~ y_k = h_k x_k, \end{equation} \]

(93)

where \( h_k\in \mathbb{R}\) . We see that the output enables us to reconstruct the constant state \( x_k\) as soon as \( h_k \neq 0\) for some \( k\) . Let us try to build a KKL observer. Thanks to the dynamics being linear, we look for a transformation of the form \( T_k (x) = m_k x\) , where \( (m_k)_{k \in \mathbb{N}}\) is a sequence of scalars to be found so that (88) holds. Picking \( \lambda\in (0,1)\) , this is achieved if for all \( k \in \mathbb{N}\) ,

\[ \begin{equation} m_{k+1} = \lambda m_k + h_k, \end{equation} \]

(94)

to which the solution is \( m_k = \lambda^k m_0 + \sum_{j=0}^{k-1}\lambda^{k-j-1}h_j\) for some initial \( m_0\) . As long as \( m_0\neq 0\) , the \( m_k\) are always non-zero for \( k>0\) so that each \( T_k\) is injective. However, if \( h_k\) vanishes asymptotically, \( m_k\) decays to zero as \( k\) increases, and the sequence \( (T_k)_{k \in \mathbb{N}}\) is not uniformly injective. We get

\[ \begin{align*} |x_k - \hat{x}_k| & = \frac{1}{m_k}|z_k - \hat{z}_k| \\ & = \frac{ \lambda^k}{m_k}|z_0 - \hat{z}_0| \\ & = \frac{ \lambda^k}{\lambda^k m_0 + \sum_{j=0}^{k-1}\lambda^{k-j-1}h_j}|h_0 x_0 - h_0\hat{x}_0| \\ & = \frac{ h_0}{m_0 + \sum_{j=0}^{k-1}\frac{h_j}{\lambda^{j+1}}}|x_0 - \hat{x}_0|. \end{align*} \]

Consider a first case where for some \( k^\star \in \mathbb{N}_{>0}\) ,

\[ \begin{equation} h_k = \left\{\begin{array}{@{}l@{~~}l@{}} 1 & \text{if } k \leq k^\star \\ 0 & \text{if } k > k^\star, \end{array}\right. \end{equation} \]

(95)

then, \( |x_k - \hat{x}_k|\) does not converge to zero. The reason is that even though each map \( T_k\) is injective at each \( k\) , \( (T_k)_{k \in \mathbb{N}}\) becomes less and less injective over time. Consider another case where \( h_k = h_0 \epsilon^k\) for some constants \( h_0 \neq 0\) and \( \epsilon \in (0, 1)\) , so the system is instantaneously observable at each \( k\) , but “less and less” over time. We have

\[ \begin{align*} \begin{split} |x_k - \hat{x}_k|&=\frac{h_0}{m_0 + \frac{h_0}{\lambda}\sum_{j=0}^{k-1}\left(\frac{\epsilon}{\lambda}\right)^j}|x_0 - \hat{x}_0|\\ & =\frac{h_0}{m_0 + \frac{h_0}{\epsilon - \lambda} \left(\left(\frac{\epsilon}{\lambda}\right)^k - 1\right)}|x_0 - \hat{x}_0|, \end{split} \end{align*} \]

so that if we choose \( \lambda < \epsilon\) , the estimation error converges to zero asymptotically. Furthermore, if we initialize \( (m_k)_{k \in \mathbb{N}}\) as \( m_0 = \frac{h_0}{\epsilon - \lambda} > 0\) (note that \( (h_k)_{k \in \mathbb{N}}\) is known), we even get exponential stability of the estimation error as

\[ \begin{equation} |x_k - \hat{x}_k| = (\epsilon - \lambda)\left(\frac{\lambda}{\epsilon}\right)^k |x_0 - \hat{x}_0|. \end{equation} \]

(96)

This estimation can also be made arbitrarily fast by keeping pushing \( \lambda\) smaller. Therefore, the uniform injectivity of \( (T_k)_{k \in \mathbb{N}}\) is a sufficient condition according to Theorem 3, but it is not necessary. Convergence, stability, as well as other properties, could still happen without uniformity in \( k\) , but it is not guaranteed.

In this chapter, we provide sufficient conditions to guarantee:

  • In Section 2.3.3, existence of \( (T_k)_{k \in \mathbb{N}}\) satisfying (88);
  • In Section 2.3.4, uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) after a certain time;
  • In Section 2.3.5, injectivity of each \( T_k\) after a certain time.

Actually, in Section 2.3.4, we achieve a stronger asymptotic property than (92): we show the exponential stability of the estimation error in the \( x\) -coordinates, namely, there exist \( c_1 > 0\) , \( c_2 \in (0, 1)\) , and \( k^\star\in \mathbb{N}\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ \begin{equation} |x_k - \hat{x}_k| \leq c_1 c_2^k |x_0 - \hat{x}_0|. \end{equation} \]

(97)

Such a property is achieved by strengthening the uniform injectivity of \( (T_k)_{k \in \mathbb{N}}\) in (91) into uniform Lipschitz injectivity and the continuity of \( T_0\) into Lipschitz continuity (with \( \rho\) and \( \rho_0\) linear). This stronger result enables us to obtain a discrete-time observer with arbitrarily fast robust convergence as soon as allowed by the distinguishability property. More precisely, for any desired convergence rate \( c_2^\star \in (0, 1)\) , there exists a choice of \( (A,B)\) such that (97) is satisfied with \( c_2 \leq c_2^\star\) . Also, such a design allows for robustness against disturbances\( /\) uncertainties and filtering of measurement noise.

2.3.3 Existence of \( (T_k)_{k \in \mathbb{N}}\) Satisfying (88)

This part studies the sufficient conditions for the existence of \( (T_k)_{k \in \mathbb{N}}\) satisfying (88). It is established under the following assumption.

Assumption 6

For all \( k \in \mathbb{N}\) , \( f_k\) is invertible and its inverse function \( f_k^{-1}\) is defined on \( \mathbb{R}^{n_x}\) .

Remark 9

While invertibility is for now required globally, since the solutions of interest are known to remain in \( \mathcal{X}\) , it may be possible to modify the maps \( (f_k)_{k \in \mathbb{N}}\) (and so \( (f^{-1}_k)_{k \in \mathbb{N}}\) ) outside of the set \( \mathcal{X}\) , while still keeping the observability property mentioned below (see Section 2.3.4.4).

Such an assumption is common in observer designs for discrete-time systems, both nonlinear [94, 89, 102] and linear [79, 90, 43, 81, 126], and concerns a wide class of systems. For instance, discrete-time dynamics that are discretizations of continuous-time dynamics take the form \( x_{k+1}= x_k + \Delta t_k \Phi(x_k, t_k)\) , which is close to identity for sufficiently small sampling times \( \Delta t_k\) , and therefore invertible. The physical meaning of this assumption is that a given current state has only one possible past. Such invertibility of the dynamics allows us to go back and forth in discrete time and access states at different times, according to

\[ \begin{align}x_{k+n} &= (f_{k+n-1} \circ f_{k+n-2} \circ \ldots \circ f_k) (x_k), \\ x_{k-n}& = (f^{-1}_{k-n} \circ f^{-1}_{k-(n-1)} \circ \ldots \circ f^{-1}_{k-1}) (x_k),\\\end{align} \]

(98.a)

for \( k,n\in \mathbb{N}\) . Note that the ability to cope with the non-invertibility of the dynamics has been studied in the linear context [127]. Under this invertibility assumption, Theorem 4 gives existence results for the function sequence \( (T_k)_{k \in \mathbb{N}}\) .

Theorem 4 (Existence of \( (T_k)_{k\in \mathbb{N}}\) satisfying (88))

Under Assumption 6, given any \( T_0: \mathbb{R}^{n_x} \to \mathbb{R}^{n_z}\) , the sequence \( (T_k)_{k\in \mathbb{N}}\) such that each \( T_k: \mathbb{R}^{n_x} \to \mathbb{R}^{n_z}\) is given by

\[ \begin{equation} T_k (x) = A^k (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1})(x) + \sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x), \end{equation} \]

(99)

verifies (88). Conversely, verifying (88) for all \( k\in \mathbb{N}\) implies (99) for all \( k\in \mathbb{N}\) and for all \( x\in \mathcal{X}\) such that \( (f^{-1}_{k-1-p}\circ f^{-1}_{k-p}\circ … \circ f^{-1}_{k-1})(x)\in \mathcal{X}\) for all \( 0\leq p \leq k-1\) .

Proof. To start, notice that under Assumption 6, (88) is verified for all \( k\in \mathbb{N}\) if and only if for all \( k \in \mathbb{N}_{>0}\) ,

\[ \begin{equation} T_k(x) = A (T_{k-1} \circ f^{-1}_{k-1})(x) + B (h_{k-1} \circ f^{-1}_{k-1})(x), ~~ \forall x \in \mathcal{X} : f^{-1}_{k-1}(x)\in \mathcal{X}. \end{equation} \]

(100)

Notice that for all \( k \in \mathbb{N}_{>0}\) , \( T_k\) defined in (99) satisfies (100) analytically. We next show by induction that having (100) for all \( k\in\mathbb{N}_{>0}\) implies verifying (99) for all \( k\in\mathbb{N}_{>0}\) and for all \( x \in \mathcal{X}\) such that \( (f^{-1}_{k-1-p}\circ f^{-1}_{k-p}\circ … \circ f^{-1}_{k-1})(x)\in \mathcal{X}\) for all \( 0\leq p \leq k-1\) . This is trivial for \( k = 1\) . Then, assume having (100) up to rank \( k\in \mathbb{N}_{>0}\) implies verifying (99) at rank \( k\) for all \( x \in \mathcal{X}\) such that \( (f^{-1}_{k-1-p}\circ f^{-1}_{k-p}\circ … \circ f^{-1}_{k-1})(x)\in \mathcal{X}\) for all \( 0\leq p \leq k-1\) . We next show it at rank \( k+1\) . Let \( x \in \mathcal{X}\) such that \( (f^{-1}_{k-p}\circ f^{-1}_{k-p+1}\circ … \circ f^{-1}_{k})(x)\in \mathcal{X}\) for all \( 0\leq p \leq k\) . Then, \( f^{-1}_{k}(x) \in \mathcal{X}\) and \( (f^{-1}_{k-1-p}\circ f^{-1}_{k-p}\circ … \circ f^{-1}_{k-1})(f^{-1}_{k}(x))\in \mathcal{X}\) for all \( 0\leq p \leq k-1\) . Having (100) up to rank \( k+1\) thus implies (99) at rank \( k\) applied to \( f^{-1}_{k}(x)\) and thus

\[ \begin{align*} T_{k+1}(x)&= A (T_k \circ f^{-1}_k)(x) + B (h_k \circ f^{-1}_k)(x)\\ & A\Bigg(A^k (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1})(f^{-1}_k(x))\\ &~~{}+ \sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(f^{-1}_k(x))\Bigg) + B (h_k \circ f^{-1}_k)(x)\\ &= A A^k (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1} \circ f^{-1}_k)(x)\\ &~~{} + \left( A\sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1} \circ f^{-1}_k)(x) + B (h_k \circ f^{-1}_k)(x)\right) \\ &=A^{k+1} (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1} \circ f^{-1}_k)(x)\\ &~~{} + \sum_{j=0}^{k+1-1}A^{k+1-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1} \circ f^{-1}_k)(x), \end{align*} \]

which is (99) at rank \( k+1\) at the point \( x\) . \( \blacksquare\)

Example 9 (Systems with linear dynamics and polynomial output)

Consider the class of systems of the form (84) with linear dynamics and polynomial output (as in [102, Section III] but here with time-varying matrices)

\[ \begin{equation} x_{k+1} = F_k x_k, ~~ y_k = H_k P_d(x_k), \end{equation} \]

(101)

where \( (F_k)_{k \in \mathbb{N}} \in \mathbb{R}^{n_x \times n_x}\) and \( (H_k)_{k \in \mathbb{N}} \in \mathbb{R}^{n_y \times n_d}\) are sequences of matrices and \( P_d: \mathbb{R}^{n_x} \to \mathbb{R}^{n_d}\) is a vector of \( n_d\) monomials with degrees less than or equal to \( d\) . We then look for \( (T_k)_{k \in \mathbb{N}}\) of the form

\[ \begin{equation} T_k(x) = M_k P_d(x). \end{equation} \]

(102)

Since \( P_d(F_k x)\) contains polynomials of \( x\) of order less than or equal to \( d\) , there exists \( (D_k)_{k \in \mathbb{N}} \in \mathbb{R}^{n_d \times n_d}\) such that for all \( k \in \mathbb{N}\) ,

\[ \begin{equation} P_d(F_k x) = D_k P_d(x). \end{equation} \]

(103)

Therefore, we have \( T_{k+1}(x_{k+1}) = M_{k+1}P_d(x_{k+1}) = M_{k+1} P_d(F_k x_k) = M_{k+1} D_k P_d(x_k)\) and (88) holds if

\[ \begin{equation} M_{k+1}D_k = A M_k + B H_k. \end{equation} \]

(104)

If \( (D_k)_{k \in \mathbb{N}}\) is invertible for all \( k \in \mathbb{N}\) , it can be proven by mathematical induction that (104) admits the unique solution

\[ M_k = A^k M_0 \prod_{j=0}^{k-1}D_j^{-1} + \sum_{j=0}^{k-1}A^{k-j-1}BH_j \prod_{q=j}^{k-1}D_q^{-1}, \]

for all \( k \in \mathbb{N}_{>0}\) , initialized as \( M_0\) . So \( (T_k)_{k \in \mathbb{N}}\) is of the form

\[ \begin{equation} T_k(x) = \left(A^k M_0 \prod_{j=0}^{k-1}D_j^{-1} + \sum_{j=0}^{k-1}A^{k-j-1}BH_j \prod_{q=j}^{k-1}D_q^{-1}\right)P_d(x). \end{equation} \]

(105)

The particular case where the system is fully linear, namely with \( P_d(\cdot)\) identity, is detailed below in Section 2.3.6.

Now that the existence of \( (T_k)_{k \in \mathbb{N}}\) has been shown, we next provide sufficient conditions guaranteeing its injectivity.

2.3.4 Arbitrarily Fast Robust Discrete-Time Observer

This part shows that the uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) is obtained after a certain time under uniform Lipschitz backward distinguishability if the target dynamics are pushed sufficiently fast. This leads to an arbitrarily fast robust discrete-time observer as soon as allowed by distinguishability.

2.3.4.1 Uniform Lipschitz Injectivity of \( (T_k)_{k \in \mathbb{N}}\)

In this part, \( A\) is chosen of the form \( \gamma \tilde{A}\) with \( \tilde{A}\) Schur, and \( \gamma \in (0, 1]\) sufficiently small to ensure uniformly Lipschitz injectivity of \( (T_k)_{k\in \mathbb{N}}\) after a certain time. This is done under the following distinguishability conditions.

Definition 12 (Uniform Lipschitz backward distinguishability of system (84))

System (84) is uniformly Lipschitz backward distinguishable on a set \( \mathcal{X}\) if for each \( i \in \{1, 2, …, n_y\}\) , there exists \( m_i \in \mathbb{N}_{>0}\) such that for all \( k \in \mathbb{N}_{\geq \overline{m}}\) where \( \overline{m}: = \max_{i \in \{1, 2, …, n_y\}} m_i\) , the sequence of backward distinguishability maps \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}}\) defined as

\[ \begin{equation} \mathcal{O}^{bw}_k(x) = \left( \mathcal{O}^{bw}_{1,k}(x), \mathcal{O}^{bw}_{2,k}(x), \ldots, \mathcal{O}^{bw}_{n_y,k}(x) \right), \end{equation} \]

(106.a)

where \( \mathcal{O}^{bw}_{i,k}(x) \in \mathbb{R}^{m_i}\) is defined as

\[ \begin{equation} \mathcal{O}^{bw}_{i,k}(x) = \begin{pmatrix} (h_{i,k-1} \circ f^{-1}_{k-1})(x) \\ (h_{i,k-2} \circ f^{-1}_{k-2} \circ f^{-1}_{k-1})(x) \\ \ldots \\ (h_{i,k-(m_i-1)} \circ f^{-1}_{k-(m_i-1)} \circ \ldots \circ f^{-1}_{k-1})(x) \\ (h_{i,k-m_i} \circ f^{-1}_{k-m_i} \circ f^{-1}_{k-(m_i-1)}\circ \ldots \circ f^{-1}_{k-1})(x) \end{pmatrix}, \end{equation} \]

(106.b)

is uniformly Lipschitz injective on \( \mathcal{X}\) , i.e., there exists \( c_o > 0\) such that for all \( k \in \mathbb{N}_{\geq \overline{m}}\) and for all \( (x_a,x_b)\in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{equation} |\mathcal{O}^{bw}_k(x_a) - \mathcal{O}^{bw}_k(x_b)| \geq c_o|x_a - x_b|. \end{equation} \]

(107)

Intuitively, the concatenation of a sufficient number \( m_i\) of the past outputs determines uniquely and uniformly the current state (and equivalently the trajectory as well). Equivalent kinds of uniform observability are assumed in [106, Theorem \( 4.1\) ] and [42, Theorem \( 2\) ] for autonomous and time-varying continuous-time systems respectively, leading to similar results with arbitrarily fast convergence of the estimation error.

Remark 10

While the condition in Definition 12 is what is required later for the proof, in practice it is not always easy to obtain the closed forms of the inverse maps of \( f_k\) in \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}}\) . Actually, this condition is satisfied with \( m_i = m\) for all \( i \in \{1, 2, …, n_y\}\) if both of the following conditions are satisfied.

  • There exists \( m \in \mathbb{N}_{>0}\) such that there exists \( c_o^\prime > 0\) such that for all \( k \in \mathbb{N}\) and for all \( (x_a,x_b)\in \mathcal{X} \times \mathcal{X}\) ,

    \[ |\mathcal{O}^{fw}_k(x_a) - \mathcal{O}^{fw}_k(x_b)| \geq c_o^\prime|x_a - x_b|, \]

    where the sequence of forward distinguishability maps \( (\mathcal{O}^{fw}_k)_{k \in \mathbb{N}}\) is defined as

    \[ \mathcal{O}^{fw}_k(x) = \left( \mathcal{O}^{fw}_{1,k}(x), \mathcal{O}^{fw}_{2,k}(x), \ldots, \mathcal{O}^{fw}_{n_y,k}(x)\right), \]

    where \( \mathcal{O}^{fw}_{i,k}(x) \in \mathbb{R}^m\) is defined as

    \[ \mathcal{O}^{fw}_{i,k}(x) = \begin{pmatrix} h_{i,k}(x) \\ (h_{i,k+1} \circ f_k)(x) \\ \ldots \\ (h_{i,k+(m-2)} \circ f_{k+(m-3)} \circ \ldots \circ f_k)(x) \\ (h_{i,k+(m-1)} \circ f_{k+(m-2)} \circ f_{k+(m-3)} \circ \ldots \circ f_k)(x) \end{pmatrix}; \]
  • The sequence of inverses \( (f^{-1}_k)_{k \in \mathbb{N}}\) is uniformly Lipschitz injective, i.e., there exists \( \underline{c}_f > 0\) such that for all \( k \in \mathbb{N}\) and for all \( (x_a, x_b) \in \mathbb{R}^{n_x} \times \mathbb{R}^{n_x}\) ,

    \[ |f^{-1}_k(x_a) - f^{-1}_k(x_b)| \geq \underline{c}_f|x_a - x_b|. \]

Indeed, from the two conditions above, we have for all \( k \in \mathbb{N}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{align*} |\mathcal{O}^{bw}_k(x_a) - \mathcal{O}^{bw}_k(x_b)| &= |\mathcal{O}^{fw}_{k-m}((f^{-1}_{k-m} \circ f^{-1}_{k-(m-1)} \circ \ldots \circ f^{-1}_{k-1})(x_a)) \\ &~~{}- \mathcal{O}^{fw}_{k-m}((f^{-1}_{k-m} \circ f^{-1}_{k-(m-1)} \circ \ldots \circ f^{-1}_{k-1})(x_b))| \\ & \geq c_o^\prime |(f^{-1}_{k-m} \circ f^{-1}_{k-(m-1)} \circ \ldots \circ f^{-1}_{k-1})(x_a) - (f^{-1}_{k-m} \circ f^{-1}_{k-(m-1)} \circ \ldots \circ f^{-1}_{k-1})(x_b)|\\ & \geq c_o^\prime \underline{c}_f^m |x_a - x_b|:= c_o|x_a - x_b|, \end{align*} \]

by letting \( c_o = c_o^\prime \underline{c}_f^m\) . Checking uniform Lipschitz backward distinguishability using \( (\mathcal{O}^{fw}_k)_{k \in \mathbb{N}}\) is much more convenient than \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}}\) since the forward maps \( (f_k)_{k \in \mathbb{N}}\) are available.

For our uniform Lipschitz injectivity result, we make the following assumptions.

Assumption 7

Assume that:

  1. The sequences \( (f^{-1}_k)_{k \in \mathbb{N}}\) and \( (h_k)_{k \in \mathbb{N}}\) are uniformly Lipschitz, i.e., there exist positive scalars \( c_f,c_h\) such that for all \( k \in \mathbb{N}\) and for all \( (x_a,x_b) \in \mathbb{R}^{n_x} \times \mathbb{R}^{n_x}\) ,

    \[ \begin{align} |f^{-1}_{k}(x_a) - f^{-1}_{k}(x_b)| \leq c_f |x_a - x_b|, \\ |h_k(x_a) - h_k(x_b)| \leq c_h |x_a - x_b|; \\\end{align} \]

    (108.a)

  2. System (84) is uniformly Lipschitz backward distinguishable on \( \mathcal{X}\) for some \( m_i \in \mathbb{N}_{>0}\) , \( i \in \{1, 2, …, n_y\}\) .

Remark 11

In Assumption 7Item (V) requires global uniform Lipschitzness of the sequences \( (f^{-1}_k)_{k \in \mathbb{N}}\) and \( (h_k)_{k \in \mathbb{N}}\) . Its relaxation into uniform Lipschitzness over a compact set is analyzed in Section 2.3.4.4. Note that for a linear time-varying system, Item (V) is reduced to uniform boundedness of the dynamics and output matrices (see Section 2.3.6).

The following theorem shows the uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) after a certain time.

Theorem 5 (Uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) after a certain time)

Suppose Assumptions 56, and 7 hold. Define \( n_z = \sum_{i = 1}^{n_y} m_i\) . Consider a globally Lipschitz map \( T_0:\mathbb{R}^{n_x}\to \mathbb{R}^{n_z}\) , and for each \( i\in \{1, 2, …,n_y\}\) , a controllable pair \( (\tilde{A}_i, \tilde{B}_i)\in \mathbb{R}^{m_i\times m_i}\times \mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur. Then, there exists \( 0 < \gamma^\star \leq 1\) such that for any \( 0 < \gamma < \gamma^\star\) , there exists \( k^\star \in \mathbb{N}\) such that the sequence \( (T_k)_{k \in \mathbb{N}}\) defined in (99) with

\[ \begin{align} A &=\gamma \tilde{A} = \gamma \rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_y}) \in \mathbb{R}^{n_z \times n_z}, \\ B& = \rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_y}) \in \mathbb{R}^{n_z \times n_y}, \\\end{align} \]

(109.a)

and initialized as \( T_0\) , is uniformly Lipschitz injective on \( \mathcal{X}\) for all \( k \in \mathbb{N}_{\geq k^\star}\) , where \( \gamma^\star\) and \( k^\star \in \mathbb{N}_{\geq \overline{m}}\) are defined in the proof. More precisely, there exists \( c > 0\) (independent of \( \gamma)\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) , we have

\[ \begin{equation} |T_k(x_a) - T_k(x_b)| \geq c \gamma^{\overline{m}-1} |x_a - x_b|, \end{equation} \]

(110)

where \( \overline{m} = \max_{i \in \{1, 2, …, n_y\}} m_i\) .

Proof. This highly technical proof has been moved to Section 6.1.2.1 to facilitate reading. It involves three main steps: i) Express \( T_k\) as a sum of three parts, the first one linked to the initial condition \( T_0\) (that may not be injective), the second one linked to all the outputs from the initial time up to \( \overline{m}\) steps ago, and the last one linked to the closest \( \overline{m}\) outputs (that gives us injectivity), ii) Upper-bound the first two parts thanks to Lipschitzness and lower-bound the last part thanks to uniform Lipschitz backward distinguishability, and iii) Push \( \gamma\) sufficiently small to make the last part dominate the first two parts, bringing injectivity. \( \blacksquare\)

Remark 12

This is a high-gain result in discrete time since we have to push the (discrete) dynamics sufficiently fast, namely take \( \gamma\) sufficiently small, to guarantee uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) . However, as \( \gamma\) is picked closer to zero, the coefficient \( \frac{1}{c\gamma^{\overline{m}-1}}\) quantifying the injectivity of \( (T_k)_{k \in \mathbb{N}}\) in (110) increases, making \( (T_k)_{k \in \mathbb{N}}\) “less (but still) uniformly Lipschitz injective”. We also observe from the proof of Theorem 5 in Section 6.1.2.1 that:

  • If \( c_o\) is close to zero, i.e., system (84) is “less uniformly Lipschitz backward distinguishable”, the upper bound on \( \gamma\) is reduced, which means we have to pick \( \gamma\) closer to zero to guarantee uniform Lipschitz injectivity of \( (T_k)_{k\in \mathbb{N}}\) ;
  • As \( \gamma\) is picked closer to zero, \( k^\star\) approaches \( \overline{m}\) , which means \( (T_k)_{k \in \mathbb{N}}\) becomes uniformly Lipschitz injective right after we have uniform Lipschitz backward distinguishability, namely in \( \overline{m}\) steps. Also, the discontinuity of \( k^\star\) in \( c_T\) reflects the time dependence of the injectivity of \( (T_k)_{k \in \mathbb{N}}\) . Indeed, if \( c_T = 0\) , then the uniform Lipschitz injectivity of \( (T_k)_{k \in \mathbb{N}}\) is achieved as soon as we get uniform Lipschitz backward distinguishability, so it is independent of time. For \( c_T > 0\) , we will have to wait for some time until the terms \( (\mathcal{I}_k)_{k \in \mathbb{N}}\) become dominated. Therefore, this injectivity is time-dependent.

Example 10 (Case of non-uniform backward distinguishability)

Consider the system in Example 8. We have

\[ \begin{equation} \mathcal{O}^{bw}_k(x) = \begin{pmatrix} h_{k-1}x \\ h_{k-2}x \\ \ldots \\ h_{k-m}x \end{pmatrix} = \begin{pmatrix} h_{k-1} \\ h_{k-2} \\ \ldots \\ h_{k-m} \end{pmatrix} x := \mathcal{H}^{bw}_k x, \end{equation} \]

(111)

which is not uniformly Lipschitz injective since

\[ |\mathcal{O}^{bw}_k(x_a) - \mathcal{O}^{bw}_k(x_b)| = \|\mathcal{H}^{bw}_k\| |x_a - x_b| \]

and \( \|\mathcal{H}^{bw}_k\|\) cannot be lower bounded by any positive constant uniformly in \( k\) for any \( m\) . Therefore, this example does not fall into the context of Theorem 5.

2.3.4.2 Arbitrarily Fast Observer Design

According to the proof of Theorem 3, once \( (T_k)_{k \in \mathbb{N}}\) has become uniformly Lipschitz injective on \( \mathcal{X}\) following Theorem 5, there exists a sequence of left inverse maps \( (T^*_k)_{k \in \mathbb{N}}: \mathbb{R}^{n_z} \to \mathbb{R}^{n_x}\) and \( c^\prime > 0\) such that

\[ \begin{align} T^*_k(T_k(x))& = x, ~~ \forall k \in \mathbb{N}_{\geq k^\star}, ~~ \forall x \in \mathcal{X},\\ |T^*_k(z_a) - T^*_k(z_b)| &\leq \frac{c^\prime}{c \gamma^{\overline{m}-1}}|z_a - z_b|,~~ \forall k \in \mathbb{N}_{\geq k^\star}, ~~ \forall (z_a, z_b) \in \mathbb{R}^{n_z} \times \mathbb{R}^{n_z}. \\\end{align} \]

(112.a)

Exploiting Lipschitzness, the result of Theorem 3 can thus be strengthened as follows, obtaining exponential asymptotic stability of the estimation error in the \( x\) -coordinates, and an arbitrarily fast discrete-time observer.

Corollary 1 (Arbitrarily fast observer in discrete time)

Under the assumptions of Theorem 5, consider \( A\) and \( B\) of the form (109) with \( 0<\gamma<\gamma^\star\) , \( (T_k)_{k \in \mathbb{N}}\) , and \( k^\star\) provided by Theorem 5. Then, there exist \( (T_k^*)_{k\in \mathbb{N}}\) and \( \overline{c} > 0\) such that for any solution \( k\mapsto x_k\) to system (84) with \( x_0 \in \mathcal{X}_0\) and any solution \( k\mapsto \hat{z}_k\) to observer (90) with \( \hat{z}_0\in T_0(\mathcal{X})\) and input \( y_k=h_k(x_k)\) , we have

\[ \begin{equation} |x_k - \hat{x}_k| \leq \frac{\overline{c}(\gamma \|\tilde{A}\|)^k}{\gamma^{\overline{m}-1}}|x_0 - \hat{x}_0|, ~~ \forall k \in \mathbb{N}_{\geq k^\star}. \end{equation} \]

(113)

Corollary 1 shows that observer (90) can be made arbitrarily fast after \( (T_k)_{k \in \mathbb{N}}\) has become uniformly Lipschitz injective, by picking \( \gamma\) closer to zero. Indeed, compared with (97), the estimation error in the \( x\) -coordinates is exponentially stable with \( c_1 = \frac{\overline{c}}{ \gamma^{\overline{m}-1}}\) and \( c_2 = \gamma \|\tilde{A}\|<1\) (according to the proof of Theorem 5). For any desired convergence rate \( c_2^\star \in (0, 1)\) , by picking \( 0 < \gamma < \min\left\{\frac{c_2^\star}{\|\tilde{A}\|}, \gamma^\star\right\}\) , we achieve \( 0<c_2 \leq c_2^\star\) . Note that this typically increases \( c_1\) , because if \( c_2 \leq c_2^\star\) then \( c_1 \geq c_1^\star = \frac{\overline{c} \|\tilde{A}\|^{\overline{m}-1}}{ (c_2^\star)^{\overline{m}-1}}. \) We seem to recover a discrete-time version of the well-known peaking behavior in continuous-time high-gain designs [26]. Note though that since \( k^\star \in \mathbb{N}_{\geq \overline{m}}\) , (113) can actually be re-written as

\[ \begin{equation} |x_k - \hat{x}_k| \leq \overline{c}^\prime\gamma(\gamma \|\tilde{A}\|)^{k-\overline{m}}|x_0 - \hat{x}_0|, ~~ \forall k \in \mathbb{N}_{\geq k^\star}, \end{equation} \]

(114)

indicating that the peaking is over after \( k^\star\) . This observer is illustrated in Section 2.3.6.

Remark 13

While we assume in Item (V) of Assumption 7 that the maps \( (f^{-1}_k)_{k \in \mathbb{N}}\) and \( (h_k)_{k \in \mathbb{N}}\) are uniformly Lipschitz, namely the Lipschitz constants \( c_f\) and \( c_h\) are the same for all \( k\) , we can instead consider sequences of Lipschitz constants \( (c_{f,k})_{k \in \mathbb{N}}\) and \( (c_{h,k})_{k \in \mathbb{N}}\) providing that there are positive scalars \( \overline{c}_f\) and \( \overline{c}_h\) such that for all \( k \in \mathbb{N}\) , \( c_{f,k} \leq \overline{c}_f\) and \( c_{h,k} \leq \overline{c}_h\) . Item (V) of Assumption 7 then holds with \( c_f = \overline{c}_f\) and \( c_h = \overline{c}_h\) . These upper bounds prevent an asymptotic loss of Lipschitzness (when \( (c_{f,k})_{k \in \mathbb{N}}\) and \( (c_{h,k})_{k \in \mathbb{N}}\) diverge to infinity). Similarly, in Item (VI) of Assumption 7, we can consider a sequence \( (c_{o,k})_{k \in \mathbb{N}}\) lower bounded by \( \underline{c}_o > 0\) (to prevent an asymptotic loss of observability). Indeed, this allows us to update dynamically \( \gamma\in (0, 1]\) at each iteration \( k\) , as follows (see Section 6.1.2.1 for these constants)

\[ \begin{multline*} 0 < \gamma_k < \mu \gamma^\star_k = \mu\min\bigg\{\frac{1}{\|\tilde{A}\|}, \frac{1}{\max_{i \in \{1, 2, \ldots, n_y\}} \|\tilde{A}_i\| c_{f,k}},\\ \frac{c_c c_{o,k}}{\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_{f,k}c_c c_{o,k} + \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\|c_{h,k} c_{f,k}\max_{i \in \{1, 2, \ldots, n_y\}}((\|\tilde{A}_i\|c_{f,k})^{m_i})}\bigg\}, \end{multline*} \]

for some constant \( \mu \in (0, 1)\) . The role of \( \mu\) is to prevent \( (\gamma_k)_{k \in \mathbb{N}}\) from converging asymptotically to \( (\gamma^\star_k)_{k \in \mathbb{N}}\) , which cannot converge to zero thanks to the upper bounds \( \overline{c}_f\) and \( \overline{c}_h\) . Indeed, this could prevent convergence\( /\) injectivity. The interest of allowing \( \gamma\) to vary is that, at some time when we have a lot of observability (large \( c_{o,k}\) ) or Lipschitzness (small \( c_{f,k}\) or \( c_{h,k}\) ), we can afford to let \( \gamma_k\) increase while still keeping convergence, thus decreasing the noise amplification (see Section 2.3.4.3) caused by a too fast observer (see Section 2.4 for illustrations). Finally, we can pick a time-varying target filter in the \( z\) -coordinates, provided that the properties are uniform with respect to this variation. For instance, it was observed on a continuous-time motor [128], without any rigorous proof, that performance can be improved if the eigenvalues of the filter are adapted to the motor speed.

Remark 14

If \( T_0\) is taken constant (or even identically zero) meaning that \( c_T = 0\) , then for any initial condition \( x_0 \in \mathcal{X}_0\) of the system and \( \hat{z}_0\) of the observer, we have \( \hat{z}_0 = T_0(x_0)\) . This leads to \( \hat{z}_k = T_k(x_k)\) for all \( k \in \mathbb{N}\) and so \( \hat{x}_k = x_k\) for all \( k \in \mathbb{N}_{\geq k^\star}\) . Therefore, we have finite-time convergence.

2.3.4.3 Robust and Input-to-State Stability of the Estimation Error

In this part, we study the robust stability (in the sense of [46]) and ISS properties [45] of the observer given by Corollary 1. Suppose the system has dynamics (84) added with some disturbance\( /\) uncertainty \( v_k\) and a measurement with noise \( w_k\) :

\[ \begin{equation} x_{k+1} = f_k(x_k) + v_k, ~~ y_k = h_k(x_k) + w_k. \end{equation} \]

(115)

Then, if the pair \( (f_k)_{k\in \mathbb{N}}\) , \( (h_k)_{k\in \mathbb{N}}\) verifies the conditions of Theorem 5, we know that there exists a sequence of left inverses \( (T_k^*)_{k\in \mathbb{N}}\) for \( k \in \mathbb{N}_{\geq k^\star}\) that verifies (112). However, in practice, following for instance [103], such maps are only approximately known. Theorem 6 then shows the robustness of the estimation error in the \( x\) -coordinates with respect to all those uncertainties (since \( \gamma \|\tilde{A}\|<1\) ).

Theorem 6 (Robust stability of the estimation error against uncertainties)

Under the assumptions of Theorem 5, consider \( A\) and \( B\) of the form (109) with \( 0<\gamma<\gamma^\star\) , \( (T_k)_{k \in \mathbb{N}}\) , and \( k^\star\) provided by Theorem 5, and \( (T_k^*)_{k \in \mathbb{N}}\) provided by Corollary 1. Consider an approximation \( (\tilde{T}_k^*)_{k \in \mathbb{N}}\) of \( (T_k^*)_{k \in \mathbb{N}}\) and \( \delta > 0\) such that

\[ \begin{equation} |\tilde{T}_k^*(z) - T_k^*(z)|\leq \delta, ~~ \forall k \in \mathbb{N}, \forall z\in \mathbb{R}^{n_z}. \end{equation} \]

(116)

Then, there exist positive scalars \( \overline{c}\) , \( \overline{c}_v\) , and \( \overline{c}_w\) (independent of \( \gamma\) ) such that for any solution to system (115) with \( x_0 \in \mathcal{X}_0\) and any solution to

\[ \begin{equation} \hat{z}_{k+1} = \gamma \tilde{A} \hat{z}_k + B y_k, ~~ \hat{x}_k = \tilde{T}^*_k(\hat{z}_k), \end{equation} \]

(117)

initialized as \( \hat{z}_0=T_0(\hat{x}_0)\in T_0(\mathcal{X})\) , we have for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ \begin{equation} |x_k - \hat{x}_k| \leq \frac{\overline{c}(\gamma \|\tilde{A}\|)^k}{\gamma^{\overline{m}-1}}|x_0 - \hat{x}_0|\\ +\frac{1}{\gamma^{\overline{m}-1}}\sum_{j=0}^{k-1}(\gamma \|\tilde{A}\|)^{k-j-1}(\overline{c}_v |v_j| + \overline{c}_w |w_j|) + \delta. \end{equation} \]

(118)

Proof. This highly technical proof has been moved to Section 6.1.2.2 to facilitate reading. It involves two main steps: i) Show that \( (T_k)_{k \in \mathbb{N}}\) provided by Theorem 5 is uniformly Lipschitz, and ii) Exhibit bounds on the uncertainties in the \( z\) -coordinates then bring these back to the \( x\) -coordinates. \( \blacksquare\)

Remark 15

Theorem 6 shows that the estimation error in the \( x\) -coordinates is robustly stable with respect to the disturbance\( /\) uncertainty \( v_k\) as well as the noise \( w_k\) and it is ISS with respect to the approximation error \( \delta\) . The former property, defined in [46], is stronger than the ISS one defined in [45].

Note that it is the exponential stability (rather than asymptotic stability) of the estimation error that provides the ISS with respect to disturbances and measurement noise. We also see from (118) that accelerating the convergence by pushing \( \gamma\) closer to zero will worsen the effect of the disturbances and noise, but not that of the approximation of the inverse transformation.

2.3.4.4 Saturating the Inverse Maps to Relax Assumption 7

In Item (V) of Assumption 7, we require that the map sequences \( (f^{-1}_k)_{k \in \mathbb{N}}\) and \( (h_k)_{k \in \mathbb{N}}\) are globally uniformly Lipschitz, which is due to the fact that we do not have backward invariance of the sequence on \( \mathcal{X}\) . Here, we would like to study how to relax that into a local requirement on a certain bounded set, without losing Item (VI) of Assumption 7.

Let us assume that, given the \( m_i\) of Item (VI) of Assumption 7, there exists a large enough \( \sigma_d > 0\) such that for all \( x \in \mathcal{X}\) and for all \( k \in \mathbb{N}_{\geq \overline{m}}\) where \( \overline{m}: = \max_{i \in \{1, 2, …, n_y\}} m_i\) , all the pre-images \( f^{-1}_{k-1}(x)\) , \( (f^{-1}_{k-2} \circ f^{-1}_{k-1})(x)\) , up to \( (f^{-1}_{k-\overline{m}} \circ f^{-1}_{k-(\overline{m}-1)} \circ … \circ f^{-1}_{k-1})(x)\) are in \( \mathcal{X} + \sigma_d\mathbb{B}\) . This means that we can change \( (f_k^{-1})_{k \in \mathbb{N}}\) as we want outside of \( \mathcal{X}+\sigma_d\mathbb{B}\) without altering Item (VI) of Assumption 7 (and without altering the system dynamics on the set \( \mathcal{X}\) where the solutions of interest evolve).

Now, for any \( \sigma_c > \sigma_d\) , let us consider a saturating function \( \chi: \mathbb{R}^{n_x} \to \mathbb{R}\) defined as

\[ \begin{equation} \chi(x) = \left\{\begin{array}{@{}l@{~~}l@{}} 1 & \text{if } x \in \mathcal{X} + \sigma_d \mathbb{B} \\ g(x) & \text{if } x \in (\mathcal{X} + \sigma_c \mathbb{B}) \setminus (\mathcal{X} + \sigma_d \mathbb{B}) \\ 0 & \text{if } x \notin \mathcal{X} + \sigma_c \mathbb{B}, \end{array}\right. \end{equation} \]

(119)

where \( g\) is any locally Lipschitz function such that \( \chi\) is locally Lipschitz. We then define \( (f^{\dagger}_k)_{k \in \mathbb{N}}: \mathbb{R}^{n_x} \to \mathbb{R}^{n_x}\) as

\[ \begin{equation} f^{\dagger}_k(x) = \chi(x) f^{-1}_k(x) + (1 - \chi(x))x. \end{equation} \]

(120)

The set

\[ \begin{equation} \mathcal{I} = (\mathcal{X} + \sigma_c\mathbb{B}) \cup \left(\bigcup\limits_{k \in \mathbb{N}}f^{\dagger}_k(\mathcal{X} + \sigma_c\mathbb{B}) \right) \subset \mathbb{R}^{n_x} \end{equation} \]

(121)

illustrated in Figure 12 is backward invariant with respect to \( (f^{\dagger}_k)_{k \in \mathbb{N}}\) . Indeed, pick any \( x\in \mathcal{I}\) and any \( k \in \mathbb{N}\) . Then, either \( x \in \mathcal{X} + \sigma_c\mathbb{B}\) and thus \( f^{\dagger}_{k}(x) \in \mathcal{I}\) , or \( x \notin \mathcal{X} + \sigma_c\mathbb{B}\) and then \( \chi(x) = 0\) and \( f^{\dagger}_{k}(x) = x \in \mathcal{I}\) . It follows that all the requirements of global uniform Lipschitzness of \( (f^{-1}_k)_{k \in \mathbb{N}}\) , \( (h_k)_{k \in \mathbb{N}}\) , and \( T_0\) as in Assumption 7 can be replaced by uniform Lipschitzness on this backward invariant set \( \mathcal{I}\) , by replacing \( (f^{-1}_k)_{k \in \mathbb{N}}\) with \( (f^{\dagger}_k)_{k \in \mathbb{N}}\) defined in (120) in all the equations. Similarly, in Remark 10, we can check uniform Lipschitz backward distinguishability using \( (\mathcal{O}^{fw}_k)_{k \in \mathbb{N}}\) instead of \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}}\) if \( (f^{-1}_k)_{k \in \mathbb{N}}\) is uniformly Lipschitz injective on \( \mathcal{I}\) . Actually, even the invertibility of each \( f_k\) as in Assumption 6 may only be required on \( \mathcal{I}\) .

&lt;span data-controller=&quot;mathjax&quot;&gt;Illustration of the backward invariant set I) (the union of all).&lt;/span&gt;
Figure 12. Illustration of the backward invariant set \( \mathcal{I}\) (the union of all).

In particular, \( \mathcal{I}\) is bounded if and only if the sequence of sets \( (f^{\dagger}_k(\mathcal{X} + \sigma_c\mathbb{B}))_{k \in \mathbb{N}}\) is uniformly bounded, which is guaranteed if \( (f^{-1}_k)_{k \in \mathbb{N}}\) is uniformly bounded on \( \mathcal{X} + \sigma_c\mathbb{B}\) . In this case, all those assumptions become much more favorable.

Remark 16

In the case of a discretization, \( f_k(x) = x + \Delta t \Phi(x, t_k)\) , where either \( \Delta t\) is very small or the function \( \Phi\) is uniformly bounded (like in our application to the PMSM in Section 2.4), then the maps \( (f^{\dagger}_k)_{k \in \mathbb{N}}\) are close to identity and there is a good chance that the sets \( (f^{\dagger}_k(\mathcal{X} + \sigma_c\mathbb{B}))_{k \in \mathbb{N}}\) should be close to \( \mathcal{X} + \sigma_c\mathbb{B}\) , which is known, and that \( \mathcal{I}\) should be bounded.

2.3.5 Injectivity from Backward Distinguishability

In this part, we show the injectivity of \( (T_k)_{k \in \mathbb{N}}\) after a certain time from non-uniform and non-Lipschitz backward distinguishability only. Note that, as illustrated in Section 2.3.2, non-uniform injectivity can sometimes be insufficient to guarantee the asymptotic convergence of the observer.

Definition 13 (Backward distinguishability of system (84))

System (84) is backward distinguishable on a set \( \mathcal{X}\) after time \( k^\star\) if there exist an open set \( \mathcal{O}\) containing \( \rm{cl}(\mathcal{X})\) and \( k^\star\in \mathbb{N}\) such that for each \( k \in \mathbb{N}_{\geq k^\star}\) , for all \( (x_a,x_b)\in \mathcal{O} \times \mathcal{O}\) with \( x_a \neq x_b\) , there exists a \( j_k \in \{0, 1, …, k-1\}\) such that

\[ \begin{equation} (h_{j_k} \circ f^{-1}_{j_k} \circ f^{-1}_{j_k + 1} \circ \ldots \circ f^{-1}_{k-1})(x_a) \neq (h_{j_k} \circ f^{-1}_{j_k} \circ f^{-1}_{j_k + 1} \circ \ldots \circ f^{-1}_{k-1})(x_b). \end{equation} \]

(122)

In other words, this means that given two different states at a time \( k\) , there exists at least one instant in the past where their corresponding outputs have been different. Note that this is much lighter than the uniform Lipschitz backward distinguishability of Section 2.3.4—no uniformity of the sequence \( (j_k)_{k \in \mathbb{N}}\) is required with respect to \( k\) nor to the pair \( (x_a, x_b)\) . Therefore, this is one of the weakest forms of observability we may consider. Contrary to the version of backward distinguishability with an order \( m\) in Definition 11, this condition requires the maps \( f_k\) to be invertible. For our injectivity result, we then make the following assumptions.

Assumption 8

Assume that:

  1. For all \( k \in \mathbb{N}\) , the functions \( f^{-1}_k\) and \( h_k\) are \( C^1\) ;
  2. There exists \( k^\star \in \mathbb{N}\) such that system (84) is backward distinguishable on \( \mathcal{X}\) after time \( k^\star\) (following Definition 13).

Theorem 7 then gives injectivity results for \( (T_k)_{k \in \mathbb{N}}\) , with \( T_0=0\) and for a generic choice of \( (A,B)\) of sufficient dimension. Its proof is based on the generalized Coron’s Lemma developed recently in [112].

Theorem 7 (Injectivity of \( (T_k)_{k \in \mathbb{N}}\) after a certain time)

Under Assumptions 56, and 8, there exists a set \( \mathcal{M}\) of zero Lebesgue measure in \( \mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}\) such that for any pair \( (\tilde{A}, \tilde{B}) \in (\mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}) \setminus \mathcal{M}\) with \( \tilde{A}\) Schur and any \( k \in \mathbb{N}_{\geq k^\star}\) , the sequence of functions \( (T_k)_{k \in \mathbb{N}}\) defined in (99) for

\[ \begin{flalign} A &= \rm{Id}_{n_y} \otimes \tilde{A} \in \mathbb{R}^{(2n_x+1)n_y \times(2n_x+1)n_y}, \\ B& = \rm{Id}_{n_y} \otimes \tilde{B} \in \mathbb{R}^{(2n_x+1)n_y \times n_y}, \end{flalign} \]

(123.a)

and initialized as \( T_0=0\) , is injective on \( \mathcal{X}\) after time \( k^\star\) .

Remark 17

Actually, the pair \( (\tilde{A},\tilde{B})\) is chosen controllable and with \( \tilde{A}\) diagonalizable, which is true for almost any such pair in \( \mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}\) . Note also that the distinguishability property in Definition 13 was shown to be generic for \( 2n_x+1\) outputs in [129] (and the references therein) when the number of outputs is larger than the number of inputs.

Proof. This highly technical proof has been moved to Section 6.1.2.3 to facilitate reading. It involves three main steps: i) Express \( T_k\) as a map of eigenvalues of \( \tilde{A}\) , ii) Apply the generalized Coron’s Lemma [112, Lemma B.\( 3\) ] at each \( k\) to show that the set of eigenvalues making \( T_k\) non-injective has zero Lebesgue measure, and iii) Show this for almost all \( \tilde{A}\) and independently of time. It is interesting that this injectivity result is proven differently from the continuous-time case in [42], due to the different nature of time. Indeed, the continuous time \( t\) belongs to the open uncountable set \( [0, +\infty)\) , so the result in [42, Theorem \( 3\) ] is proven with Coron’s Lemma applied only once to a set \( \Upsilon\) that contains time. However, the discrete time \( k\) belongs to \( \mathbb{N}\) , which is not open but countable, so the generalized Coron’s Lemma is here applied separately at each instant \( k\) , and the result is then obtained for the whole time domain by the countable union of zero-measure sets. \( \blacksquare\)

Example 11 (Case of non-uniform backward distinguishability (bis))

Consider the system in Example 8. It verifies the backward distinguishability condition in Item (VIII) of Assumption 8 as long as there exists \( k \in \mathbb{N}\) such that \( h_k\neq 0\) . Therefore, Theorem 7 applies with \( T_0 = 0\) (so \( m_0 = 0\) ): there exists a sequence of injective maps \( (T_k)_{k \in \mathbb{N}}\) (from a certain time) transforming the dynamics into a form (86).

This result only ensures the injectivity of each map \( T_k\) after a certain time, without any uniformity in \( k\) , which may impair convergence, as seen in Example 8. However, we saw in Example 8 that injectivity alone can still suffice in some cases. Therefore, if we initialize \( T_0 = 0\) , the observer may still work under backward distinguishability only, which is a very mild observability condition.

Remark 18

In general, solutions to (88) taking the form (99) is written as \( T_k(x) = \mathcal{I}_k(x) + \mathcal{T}_k(x)\) where \( \mathcal{I}_k(x) = A^k (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ … \circ f^{-1}_{k-1})(x)\) and \( \mathcal{T}_k(x) = \sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ … \circ f^{-1}_{k-1})(x)\) . In Theorem 7, we prove the injectivity of \( (T_k)_{k \in \mathbb{N}}\) for all \( k \in \mathbb{N}_{\geq k^\star}\) assuming \( T_0 = 0\) , namely the injectivity of \( (\mathcal{T}_k)_{k \in \mathbb{N}}\) . Therefore, it is advised to initialize \( (T_k)_{k \in \mathbb{N}}\) such that \( T_0\) is identically zero, if possible. In a stronger case, if \( (\mathcal{T}_k)_{k \in \mathbb{N}}\) is uniformly injective, i.e., there exist a class-\( \mathcal{K}\) function \( \kappa\) and \( l > 0\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ |\mathcal{T}_k(x_a) - \mathcal{T}_k(x_b)| \geq l \kappa(|x_a - x_b|), ~~ |T_0(x_a) - T_0(x_b)| \leq \kappa(|x_a - x_b|), \]

then

\[ \begin{align*} \begin{split} |T_k(x_a) - T_k(x_b)| & = |\mathcal{I}_k(x_a) - \mathcal{I}_k(x_b) + \mathcal{T}_k(x_a) - \mathcal{T}_k(x_b)| \\ & \geq |\mathcal{T}_k(x_a) - \mathcal{T}_k(x_b)| - |\mathcal{I}_k(x_a) - \mathcal{I}_k(x_b)| \\ & \geq (l - 2 \|A\|^k)\kappa(|x_a - x_b|), \end{split} \end{align*} \]

which implies that \( (T_k)_{k \in \mathbb{N}}\) becomes uniformly injective after a certain time. This, as seen in Theorem 3, is sufficient for an asymptotic observer assuming that the inverse map of \( \kappa\) is concave, whose dynamics, unfortunately, cannot be assigned arbitrarily fast.

2.3.6 Particular Case of Linear Time-Varying Systems

While our KKL observer in this chapter works for nonlinear systems, we would like to illustrate it in the linear case and make the link with existing Kalman(-like) designs. Consider a linear time-varying discrete-time system of the form

\[ \begin{equation} x_{k+1} = F_k x_k, ~~ y_k = H_k x_k. \end{equation} \]

(124)

A linear transformation \( x_k \mapsto z_k = T_k x_k\) into (86) can be found with the sequence of matrices \( (T_k)_{k \in \mathbb{N}}\) satisfying

\[ \begin{equation} T_{k+1} F_k = A T_k + BH_k, ~~ \forall k \in \mathbb{N}, \end{equation} \]

(125)

initialized as \( T_0\) . Under invertibility of the sequence \( (F_k)_{k \in \mathbb{N}}\) , it is defined by the closed form

\[ \begin{equation} T_k = A^k T_0 \prod_{j=0}^{k-1}F_j^{-1} + \sum_{j=0}^{k-1}A^{k-j-1}BH_j \prod_{q=j}^{k-1}F_q^{-1}, \end{equation} \]

(126)

for all \( k \in \mathbb{N}_{>0}\) . Then, provided each \( T_k\) is full-rank, and thus left-invertible (see below), the KKL observer takes the form

\[ \begin{equation} \left\{ \begin{aligned} \hat{z}_{k+1}&= A \hat{z}_k + B y_k \\ T_{k+1}&= AT_kF_k^{-1} + B H_k F_k^{-1} \end{aligned} \right. ~~ ~~ \hat{x}_k=T_k^* \hat{z}_k, \end{equation} \]

(127)

where \( T_k^*\) is a left inverse of \( T_k\) .

System (124) is uniformly Lipschitz backward distinguishable (see Definition 12) if and only if there exists \( m \in \mathbb{N}_{>0}\) such that there exists \( c_o > 0\) such that for all \( k \in \mathbb{N}_{\geq m}\) , the backward distinguishability matrix

\[ \begin{equation} \mathcal{O}^{bw}_{k} = \begin{pmatrix} H_{k-1} F^{-1}_{k-1} \\ H_{k-2} F^{-1}_{k-2} F^{-1}_{k-1} \\ \ldots \\ H_{k-(m-1)} F^{-1}_{k-(m-1)} \ldots F^{-1}_{k-1} \\ H_{k-m} F^{-1}_{k-m} F^{-1}_{k-(m-1)} \ldots F^{-1}_{k-1} \end{pmatrix} \end{equation} \]

(128)

verifies

\[ \begin{equation} (\mathcal{O}^{bw}_{k})^\top \mathcal{O}^{bw}_{k} \geq c_o \rm{Id} > 0. \end{equation} \]

(129)

Alternatively, under uniform boundedness of \( (F_k)_{k \in \mathbb{N}}\) , we can use the forward version similar to the one in Remark 10. According to Theorem 5, under (129) and the uniform boundedness of \( (F_k^{-1}, H_k)_{k \in \mathbb{N}}\) , picking \( A\) sufficiently fast of dimension \( m\) , there exist \( c_t > 0\) and \( k^\star \in \mathbb{N}\) such that for all \( k \in \mathbb{N}_{\geq k^\star}\) , \( T_k^\top T_k \geq c_t \rm{Id} > 0\) , namely \( (T_k)_{k \in \mathbb{N}}\) is (uniformly) left-invertible for \( k\) sufficiently large. Therefore, (127) is implementable and provides arbitrarily fast robust exponentially stable estimation for (124).

Interestingly, (129) coincides with the Uniform Complete Observability[UCO] condition required by the Kalman(-like) observers in Definition 8 (see also [79, Condition (\( 13\) )][43, Assumption \( 2\) -\( 3\) ]), and [81, Definition \( 3\) ]). It is thus interesting to compare those designs. In terms of dimensions, the complexity of the Kalman(-like) filter is \( \frac{n_x(n_x+1)}{2} + n_x\) , while that of the KKL observer is \( (mn_y)^2 + mn_y\) with \( mn_y \geq n_x\) (or \( \sum_{i=1}^{n_y}m_i\) instead of \( mn_y\) if the observability multiplicities \( m_i\) are considered in (129)). Therefore, the Kalman(-like) filter is advantageous in dimension compared to the KKL observer. However, the advantage of the latter (besides being applicable in the nonlinear context) is that there exists a strict ISS Lyapunov function \( V_k: \mathbb{R}^{n_x} \times \mathbb{R}^{n_x} \to \mathbb{R}_{\geq 0}\) of the quadratic form

\[ \begin{equation} V_k(x_k,\hat{x}_k) = (x_k - \hat{x}_k)^\top T_k^\top P T_k (x_k - \hat{x}_k), \end{equation} \]

(130)

where \( P \in \mathbb{R}^{n_z \times n_z}\) is a positive definite solution to \( A^\top P A - P < 0\) , and verifying

\[ \begin{equation} \alpha x_k^\top x_k \leq V_k(x_k), ~~ \forall k \in \mathbb{N}_{\geq k^\star}, \end{equation} \]

(131)

for some \( \alpha > 0\) independent of \( k\) . Exponential ISS of the estimation error can thus be proven with an explicit quadratic Lyapunov function, unlike the discrete-time Kalman filter [79, 81] whose Lyapunov function is not strict. The discrete-time Kalman-like observer [43] on the other hand, which works under the same observability condition, also has a strict Lyapunov function and provides arbitrarily fast exponential stability of the estimation error. It thus shares the same features as KKL but is restricted to linear systems. Note that when \( n_z = n_x\) , the linear KKL observer may be equivalently written in the original \( x\) -coordinates with an innovation term similar to [79, 43, 81], and the matrix \( T_k^\top P T_k\) in \( V_k\) plays the role of the inverse of the covariance matrix in the Kalman(-like) designs.

2.3.7 Conclusion

This chapter presents the KKL observer design for nonlinear time-varying discrete-time systems. After giving the closed form of the transformation \( (T_k)_{k \in \mathbb{N}}\) into an exponentially stable filter of the measurement, we show how the uniform Lipschitz injectivity of this transformation is achieved after a certain time under uniform Lipschitz backward distinguishability if the target dynamics are sufficiently fast. This results in an arbitrarily fast discrete-time observer after a sufficient number of iterations corresponding to the order of backward distinguishability, that is ISS with respect to uncertainties. For linear systems, this provides an alternative to the discrete-time Kalman filter, with an explicit quadratic ISS Lyapunov function. We also show how non-uniform injectivity of the transformations is achieved under backward distinguishability, a mild observability condition, which in some cases is enough for observer design.

The application of this observer to a PMSM with sampled inputs in Section 2.4 illustrates the efficiency of designing a discrete-time KKL observer for an appropriate faithful discrete model of the system, instead of discretizing a continuous-time KKL observer designed for the continuous model. Other examples with nonlinear dynamics would typically require us to approximate \( (T_k)_{k \in \mathbb{N}}\) using numerical tools, in which case the robustness results in Section 2.3.4.3 become useful. Indeed, closed-form expressions such as (99) are generally unavailable, apart from particular classes of systems. To address this, numerical tools need to be developed as in [103, 117]. Another open question is how to obtain a uniform injectivity result possibly without Lispchitzness and arbitrarily fast convergence, typically through a uniform non-Lipschitz distinguishability property.

Acknowledgments. We thank Vincent Andrieu and Lucas Brivadis, researchers at LAGEPP-CNRS and L\( 2\) S-CNRS, respectively, for their useful remarks and fruitful discussions.

Table 2. Summary of KKL observer results.
Class of systems and referencesSystem dynamicsObserver in the \( z\) -coordinatesEquation to solve for the transformationClosed form of the transformationObservability assumption
Linear time-invariant continuous [74]\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}\dot{x}& = &F x \\ y &=& H x\end{array}\]\( \dot{z} = Az+By\) \( A\) Hurwitz, \( z \in \mathbb{R}^{n_x}\)\[ TF = AT + BH\]\[ T = \int_{0}^{+\infty}e^{At}BHe^{-Ft}dt\] (when defined)Observability of \( (F,H)\)
Linear time-invariant discrete (inferred from [74])\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}x_{k+1} &=& F x_k \\ y_k &=& H x_k\end{array}\]\( z_{k+1} =Az_k + By_k\) \( A\) Schur, \( z \in \mathbb{R}^{n_x}\)\[ TF = AT + BH\]\[ T = \sum_{i=0}^{+\infty}A^iBHF^{-(i+1)}\] (when defined)Observability of \( (F,H)\)
Nonlinear time-invariant continuous [105, 112]\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}\dot{x}&=& f(x) \\ y&=&h(x)\end{array}\]\( \dot{z} = Az+By\) \( A\) Hurwitz, \( z \in \mathbb{R}^{n_z}\) , \( n_z \geq n_x\)\[ \frac{\partial T}{\partial x}(x) f(x) = AT(x) + Bh(x)\]\[ T(x) = \int_{-\infty}^0 e^{-As}B h(X(x,s))ds\]Backward distinguishability
Nonlinear time-invariant discrete [102]\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}x_{k+1} &=& f(x_k) \\ y_k &=& h(x_k)\end{array}\]\( z_{k+1} =Az_k + By_k\) \( A\) Schur, \( z \in \mathbb{R}^{n_z}\) , \( n_z \geq n_x\)\[ T(f(x)) = AT(x) + Bh(x)\]\[ \begin{multline*} T(x)= \\ \sum_{i=0}^{+\infty}A^i B h((\underbrace{f^{-1} \circ f^{-1} \circ … \circ f^{-1}}_{i+1 \text{ time(s)}})(x)) \end{multline*} \]Backward distinguishability
Nonlinear non-autonomous continuous [42]\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}\dot{x} &=& f(x,u) \\ y &=& h(x,u)\end{array}\]\( \dot{z} = Az+By\) \( A\) Hurwitz, \( z \in \mathbb{R}^{n_z}\) , \( n_z \geq n_x\)\[ \frac{\partial T}{\partial x}(x,t) f(x,u(t)) + \frac{\partial T}{\partial t}(x,t) \] \[ = AT(x,t) + B h(x,u(t))\]\[ T(x,t) = \int_{0}^t e^{A(t-s)}Bh(X(x,t;s;u))ds\]Uniform Lipschitz differential observability or Backward distinguishability
Nonlinear non-autonomous discrete [101]\[ \begin{array}{@{}r@{\;}c@{\;}l@{}}x_{k+1} &=& f_k(x_k) \\ y_k &=& h_k(x_k)\end{array}\]\( z_{k+1} =Az_k + By_k\) \( A\) Schur, \( z \in \mathbb{R}^{n_z}\) , \( n_z \geq n_x\)\[ \begin{multline*}T_{k+1}( f_k(x)) \\ = A T_k(x) + B h_k(x)\end{multline*} \]\[ T_k (x) = A^k (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ … \circ f^{-1}_{k-1})(x) \] \[+ \sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ … \circ f^{-1}_{k-1})(x)\]Uniform Lipschitz backward distinguishability Backward distinguishability

2.4 Application to an Electrical Machine

 

Dans ce chapitre, nous appliquons les deux nouveaux observateurs en temps discret introduits dans cette thèse à un moteur synchrone à aimants permanents et tirons des enseignements clés de leur mise en œuvre. Le système est discrétisé en utilisant deux méthodes: un schéma d’Euler simple et une approche plus avancée qui intègre la dynamique de rotation du système. Nous comparons les performances des deux observateurs avec leurs homologues en temps continu discrétisés, en montrant qu’il est préférable de concevoir et d’implémenter les observateurs directement en temps discret plutôt que de discrétiser des observateurs déjà conçus. De plus, nous comparons les observateurs conçus pour des modèles obtenus par différentes méthodes de discrétisation, en soulignant que la prise en compte de la physique du système améliore considérablement la précision de la discrétisation.

2.4.1 Introduction

The content of this chapter has been published in [101, 104]. We apply the two novel discrete-time observers introduced in this dissertation to the Permanent Magnet Synchronous Motor[PMSM] and derive key insights from their implementation. In Section 2.4.2, we present the system modeling and discretize it using two approaches: a basic Euler scheme and a more advanced approach that incorporates the system’s rotational dynamics. The application of the high-gain observer from Section 2.2 is detailed in Section 2.4.3, while the implementation of the Kravaris-Kazantzis$\slash$Luenberger[KKL] observer from Section 2.3 is described in Section 2.4.4. We compare the performance of the two observers with their discretized continuous-time counterparts, demonstrating that designing and implementing observers directly in discrete time yields better results than discretizing pre-designed continuous-time observers. Additionally, we compare observers designed for models derived from different discretization techniques, highlighting that accounting for the system’s physics significantly improves discretization accuracy. A comparative analysis of these two new designs, particularly under measurement noise, is provided in Section 2.4.5. Finally, Section 2.4.6 summarizes the key insights from these applications.

2.4.2 System Modeling and Discretization

Consider a PMSM with the continuous-time model [8]

\[ \begin{equation} \dot{x} = u - Ri, ~~ y = |x-Li|^2 - \Phi^2 = 0, \end{equation} \]

(132)

where \( x \in \mathbb{R}^2\) represents the electromagnetic flux (in Vs). The known inputs are the voltages \( u\) (in V) and currents \( i\) (in A), both in \( \mathbb{R}^2\) . The system parameters are the resistance \( R = 1.45\) (\( \Omega\) ), the inductance \( L = 0.0121\) (H), and the flux \( \Phi = 0.1994\) (Vs), which are constant. In the PMSM’s salient mode, the output \( y\) is always zero. Although the system dynamics are linear, its quadratic output map makes observer design highly challenging, requiring a rigorous approach [114, 8, 115]. Also, given that \( u\) and \( i\) are known trajectories of time, system (132) can be seen as a time-varying system without inputs.

We aim to estimate the state of system (132). However, it is not advisable to design and implement continuous-time observers for this system since the input signals \( u\) and \( i\) are only known at specific sampling times and limited computations are possible, typically related to Pulse-Width Modulation[PWM]. Moreover, in certain schemes like the high-gain observer, the trajectories of the known inputs and their higher derivatives need to be stored for implementation (see in (135) below). In continuous time, this is infinite-dimensional and thus necessitates storing only their samples. Given these challenges, a discrete-time observer is suitable in this context.

To design discrete-time observers, we then need to properly discretize the PMSM in (132) with a small sampling rate \( \tau > 0\) . This could be done in several ways and we herein consider and compare two discretization methods. First, a simple Euler discretized model of system (132) would be

\[ \begin{equation} x_{k+1} = x_k + \tau (u_k - Ri_k), ~~ y_k = |x_k-Li_k|^2 - \Phi^2 = 0. \end{equation} \]

(133)

This scheme, though straightforward, lacks the adaptability to the PMSM’s rotation speed and so leads to big discretization errors, especially at high rotating speeds. A more refined method, which incorporates the system’s rotational dynamics, is given by [8]:

\[ \begin{equation} x_{k+1} = x_k + \tau \Omega_k(u_k - Ri_k)\rm{sinc}(\varphi_k):= x_k + g_k, ~~ y_k = |x_k-Li_k|^2 - \Phi^2 = 0, \end{equation} \]

(134.a)

where \( \Omega_k = \begin{pmatrix}\cos(\varphi_k) & -\sin(\varphi_k) \\ \sin(\varphi_k) & \cos(\varphi_k)\end{pmatrix}\) , and \( \varphi_k = \frac{\omega_k \tau}{2}\) where

\[ \begin{equation} \omega_k =\rm{sign}((u_k - u_{k-1})^\top u_{k-1}) \frac{|u_k - u_{k-1}|}{\tau |u_k|} \end{equation} \]

(134.b)

is an approximation of the rotation speed of the motor, assuming that this speed does not vary too fast, and the map \( \rm{sinc}\) is defined as

\[ \begin{equation} \rm{sinc}(x) = \left\{ \begin{array}{ll} \displaystyle\frac{\sin(x)}{x}, & \text{if } x \neq 0,\\ 1, & \text{if } x = 0. \end{array}\right. \end{equation} \]

(134.c)

This scheme takes into account the physics of the system and so is much more precise compared to the general Euler one (133). Notably, when \( \omega_k = 0\) for all \( k \in \mathbb{N}\) (no rotation), from (134) we recover Euler’s discretized version in (133). We highlight that the more appropriate discretization schemes such as system (134) stem from physics, so they only make sense for physical systems like system (132) and not mathematical ones like observers. Therefore, if we want to discretize a continuous-time observer, we typically cannot have any intuition but naively use Euler’s method.

Let us illustrate our observer designs in Section 2.2 and Section 2.3 by constructing and comparing for system (132) different observers:

  • In Section 2.4.3, we focus on the high-gain observer: i) First, we build it in continuous time and then discretize it using the Euler method, and ii) Next, we design it directly in discrete time (as described in Section 2.2) based on system (134), a strategically discretized model of system (132);
  • In Section 2.4.4, we examine the KKL observer: i) First, we construct it in continuous time and then discretize it; ii) Then, we build it directly in discrete time (following Section 2.3) using both the naive Euler discretization (133) and the strategically discretized model (134) of system (132);
  • In Section 2.4.5, we compare the discrete-time high-gain and KKL observer, both designed for model (134), in the presence of measurement noise.

Intuitively, the paths should be equivalent for sufficiently small sampling times \( \tau\) . However, for a PMSM discretized at the PWM level, discretization errors are significant at high speeds and we illustrate here the great interest of following the second path instead of the first one, i.e., directly designing a discrete-time observer rather than discretizing an already designed continuous-time observer. Indeed, it offers the crucial advantage of using an appropriate discretization, adapted to the physics of the system, which is not the case in the first path where physical insight is much trickier to exploit for the observer discretization.

Throughout this chapter, we consider a unique simulation scenario, where the PMSM’s rotating speed gradually increases to reach its maximum value after around \( 12\) seconds, and then decreases. The voltage and current trajectories that result in this regime are shown in Figure 13. Also, since the curves are similar between the two state components, in all our figures, we show the estimation results for only the first component, namely \( x_1\) . Last, to ensure a fair comparison, the observer estimates are compared against the true continuous-time trajectories in all cases.

&lt;span data-controller=&quot;mathjax&quot;&gt;Voltage and current inputs in the simulation scenario considered in&amp;#160;chp_pmsm.&lt;/span&gt;
Figure 13. Voltage and current inputs in the simulation scenario considered in Section 2.4.

2.4.3 Application of the High-Gain Observer in Section 2.2

In this section, we apply the high-gain observer developed in Section 2.2 to the PMSM described in Section 2.4.2. Implementation-wise, this design is more handy than the KKL one in Section 2.3, asking for the same or even lighter observability conditions and does not necessarily require invertibility of the dynamics.

2.4.3.1 Continuous-Time High-Gain Design and its Euler Discretization

First, we attempt to design for system (132) a continuous-time high-gain observer [26]. Note that this requires derivating the output, which includes the voltage\( /\) current inputs, up to a certain order, thus can terribly magnify the noise in these trajectories as well as noise due to non-saliency of the PMSM, limiting the observer’s performance. To see this, notice that the function describing the output and its next two time derivatives along the solutions to system (132), given by

\[ \begin{equation} (y,\dot{y},\ddot{y})\left(x,u,i,\dot{u},\frac{di}{dt},\frac{d^2i}{dt^2}\right) = \begin{pmatrix}|x-Li|^2 - \Phi^2 \\ 2\eta^\top(x-Li) \\ 2\dot{\eta}^\top(x-Li) + 2\eta^\top\eta \end{pmatrix}, \end{equation} \]

(135)

where \( \eta = u - Ri + L\frac{di}{dt}\) , is uniformly Lipschitz injective if there exists \( c_\eta > 0\) such that

\[ \begin{equation} \begin{pmatrix} \eta^\top \\ \dot{\eta}^\top \end{pmatrix}^\top \begin{pmatrix} \eta^\top \\ \dot{\eta}^\top \end{pmatrix} \geq c_\eta \rm{Id} > 0. \end{equation} \]

(136)

It can be shown that this property holds if the motor speed is uniformly bounded away from zero [8]. Exploiting this, we perform the uniformly injective transformation

\[ \begin{align} z_1 &= y = |x-Li|^2 - \Phi^2, \\ z_2 &= \dot{y} = 2\eta^\top(x-Li), \\ z_3 &= \ddot{y} = 2\dot{\eta}^\top(x-Li) + 2\eta^\top\eta, \\\end{align} \]

(137.a)

which puts system (132) into an observable canonical form in the \( z\) -coordinates. A continuous-time high-gain observer for system (132) would then be of the form

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{z}}_1 & = & \hat{z}_2 + \ell k_1 (\Phi^2 - |\hat{x}-Li|^2)\\ \dot{\hat{z}}_2 & = & \hat{z}_3 + \ell^2 k_2 (\Phi^2 - |\hat{x}-Li|^2) \\ \dot{\hat{z}}_3 & = & 2\ddot{\eta}^\top(\phi(\hat{z})-Li) + 2\dot{\eta}^\top\left(u - Ri -L\frac{di}{dt}\right) + 4\dot{\eta}^\top\eta + \ell^3 k_3 (\Phi^2 - |\hat{x}-Li|^2)\\ &:=&\phi_3(\hat{z})+ \ell^3 k_3 (\Phi^2 - |\hat{x}-Li|^2), \end{array}\right. \end{equation} \]

(138.a)

with the output

\[ \begin{equation} \hat{x} = \phi(\hat{z}), \end{equation} \]

(138.b)

where \( \phi\) is a globally Lipschitz left inverse of (137), which depends on \( u\) , \( i\) , and their derivatives (which may introduce noise); \( (k_1,k_2,k_3)\) and \( \ell\) are observer parameters, with \( \ell > 0\) having to be chosen sufficiently large. However, as pointed out in Section 2.4.2, we must implement a discrete-time observer. Without any specific physical guidelines about the way observer (138) should be discretized, we use a naive Euler discretization scheme with a given sampling period \( \tau = 10^{-3}\) (s), leading to

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{z}_{1,k+1} & = & \hat{z}_{1,k} + \tau(\hat{z}_{2,k} + \ell k_1 (\Phi^2 - |\hat{x}_k-Li_k|^2))\\ \hat{z}_{2,k+1} & = & \hat{z}_{2,k} + \tau(\hat{z}_{3,k} + \ell^2 k_2 (\Phi^2 - |\hat{x}_k-Li_k|^2)) \\ \hat{z}_{3,k+1} & = & \hat{z}_{3,k} + \tau(\phi_{3,k}(\hat{z}_k) + \ell^3 k_3 (\Phi^2 - |\hat{x}_k-Li_k|^2)), \end{array}\right. \end{equation} \]

(139.a)

with the output

\[ \begin{equation} \hat{x}_k = \phi_k(\hat{z}_k). \end{equation} \]

(139.b)

We then implement observer (139) for the PMSMFigure 14 shows an ineffective estimation performance, especially at high speeds, due to the lack of precision of the observer discretization, a phenomenon also seen below with the KKL observer, as well as noise amplification.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results of the high-gain observer designed in continuous time then discretized&amp;#160;(eq:ch6_pmsm_hg_d).&lt;/span&gt;
Figure 14. Estimation results of the high-gain observer designed in continuous time then discretized (139).

2.4.3.2 Discrete-Time High-Gain Design

In this part, we propose strategically discretizing the PMSM model first, and then designing and implementing the discrete-time high-gain observer of Section 2.2. Note that system (134) does not follow a discrete-time triangular form (67) because \( x_{1,k+1}\) depends on \( x_{1,k}\) , so we deploy the transformation in Lemma 6. Based on the knowledge from continuous time in (137), we conjecture that the maps \( (\mathcal{O}^{bw}_k)_{k \in \mathbb{N}}\) of order \( 3\) should be uniformly Lipschitz injective if \( \tau\) is sufficiently small. We then perform the change of variables

\[ \begin{align} z_{1,k} &= |x_k - g_{k-1} - Li_{k-1}|^2 - \Phi^2, \end{align} \]

(140.a)

\[ \begin{align} z_{2,k} &= |x_k - g_{k-2} - g_{k-1} -Li_{k-2}|^2 - \Phi^2, \end{align} \]

(140.b)

\[ \begin{align} z_{3,k} &= |x_k - g_{k-3} - g_{k-2} - g_{k-1} -Li_{k-3}|^2 - \Phi^2, \\ \\ \\\end{align} \]

(140.c)

with \( g_k\) defined in system (134), depending only on the inputs. Then, we implement observer (76) and obtain the results in Figure 15 with visibly better accuracy compared to Figure 14. From this, we draw an important lesson: instead of designing and then discretizing a continuous-time observer, we should properly discretize the system based on its physics and then build a discrete-time observer. Note also that storing the past samples of the inputs as done in this discrete-time design is finite-dimensional and no derivatives need to be computed.

Unfortunately, with \( k_1 = k_2 =k_3 = 1\) in observer (76), we need to select an exceedingly small value of \( 7 \cdot 10^{-5}\) for \( \gamma\) to make the observer work. This necessity arises from the large Lipschitz constant of the inverse map of (140). The reason behind this lies in the fact that with a low sampling rate of \( \tau = 10^{-3}\) (s), past outputs closely resemble each other, resulting in a poorly conditioned transformation. This does not allow exploring the filtering properties of the high-gain observer in this particular example. A way to improve this is to store more past outputs, namely take more dimensions in \( z_k\) like in MHO[97, 98, 99, 96, 50]. Another way could be to increase the sampling period to make the outputs different enough, risking the deviation of the discretized model from the real one hence requiring an even more accurate discretization scheme.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results of the high-gain observer designed and implemented in discrete time&amp;#160;(eq:ch4_obsZ).&lt;/span&gt;
Figure 15. Estimation results of the high-gain observer designed and implemented in discrete time (76).

In Figure 16, we compare the estimation errors given by the discrete-time high-gain observer for model (134) but with different choices of \( \gamma\) . It is seen that the smaller \( \gamma\) , the faster the convergence, with \( \gamma = 0\) corresponding to instantaneous estimation after time \( m\) . However, a smaller \( \gamma\) results in higher sensitivity to numerical discretization noise. Moreover, the choice of \( \gamma\) seems to have little effect in the region of too high rotating speeds (around \( 12\) -\( 15\) seconds), because the discretized model becomes less accurate and deviates from the true dynamics of the PMSM. The high-gain observer, regardless of \( \gamma\) , is anyway unable to effectively address the model inaccuracies introduced by discretization.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation errors of the discrete-time high-gain observer&amp;#160;(eq:ch4_obsZ), design for model&amp;#160;(eq:ch6_pmsm_bernard) with three different choices of ) .&lt;/span&gt;
Figure 16. Estimation errors of the discrete-time high-gain observer (76), design for model (134) with three different choices of \( \gamma\) .

2.4.4 Application of the KKL Observer in Section 2.3

In this section, we apply the KKL observer developed in Section 2.3 to the PMSM described in Section 2.4.2. First, if the map (135) is uniformly Lipschitz injective (guaranteed when the PMSM’s speed is uniformly bounded away from zero) as discussed in Section 2.4.3, a continuous-time KKL observer with a sufficiently fast continuous-time pair \( (A,B)\) of dimension \( 3\) can be designed for system (132) as demonstrated in [8, Section V.A]. However, as analyzed in Section 2.4.2, the observer must be implemented in discrete time, and without specific physical insight for observer discretization, we have to naively adopt Euler’s method.

2.4.4.1 Discrete-Time KKL Observer with Euler’s Method

We now focus on designing discrete-time KKL observers for different discretized models of the PMSM, starting with the Euler one in (133). Let us verify the assumptions needed for observer design, more particularly those required by Theorem 5.

  • Assumption 5: The solutions to system (133), when injected with sinusoidal inputs (see Figure 13), are also sine waves, so they remain in a compact set in positive time;
  • Assumption 6: The dynamics map of system (133), with \( (u_k)_{k \in \mathbb{N}}\) and \( (i_k)_{k \in \mathbb{N}}\) known, is invertible. Note that this is not necessarily required for the high-gain design in Section 2.4.3;
  • Assumption 7: First, uniform Lipschitzness of the inverse dynamics and output maps of system (133) holds since the inputs \( (u_k)_{k \in \mathbb{N}}\) and \( (i_k)_{k \in \mathbb{N}}\) are uniformly bounded and solutions remain in a compact set. Second, the uniform Lipschitz backward distinguishability is very hard to check analytically in discrete time because it involves the inversion of the dynamics. Similar to Section 2.4.3, we use its continuous-time version related to (136) to conjecture that the equivalent property holds in discrete time if the sampling period \( \tau\) is sufficiently small.

Guided by Example 9 and the knowledge that a KKL observer of dimension \( 3\) exists in continuous time, we look for a transformation of the form

\[ \begin{equation} z_k = T_k(x_k) = a_k |x_k|^2 + b_k x_k + c_k \in \mathbb{R}^3, \end{equation} \]

(141)

where

\[ \begin{align*} a_k & = \begin{pmatrix} a_{1,k} & a_{2,k} & a_{3,k} \end{pmatrix}^\top \in \mathbb{R}^3, \\ b_k & = \begin{pmatrix} b_{1,k} & b_{2,k} & b_{3,k} \end{pmatrix}^\top \in \mathbb{R}^{3 \times 2}, \\ c_k & = \begin{pmatrix} c_{1,k} & c_{2,k} & c_{3,k} \end{pmatrix}^\top \in \mathbb{R}^3. \end{align*} \]

Note that each \( b_{i,k}\) , \( i = 1,2,3\) is a vector in \( \mathbb{R}^2\) . With \( A\in \mathbb{R}^{3\times 3}\) Schur and the pair \( (A, B)\) controllable, we get that \( z_k\) is solution to system (86) if \( (a_k,b_k,c_k)_{k\in\mathbb{N}}\) evolve following

\[ \begin{align} \begin{split} a_{k+1} &= A a_k + B,\\ b_{k+1} &= A b_k -2\tau a_{k+1}(u_k-Ri_k)^\top- 2LBi_k^\top,\\ c_{k+1} &= A c_k -\tau^2 a_{k+1}|u_k - Ri_k|^2 - \tau b_{k+1}(u_k-Ri_k)+ B(L^2|i_k|^2 - \Phi^2). \end{split} \end{align} \]

(142)

Note that \( a_k\) can be picked constant equal to \( (\rm{Id}-A)^{-1}B\) . Because \( y_k = 0\) for all \( k\) , \( z_k\) converges to zero exponentially fast and it is straightforward to pick for the observer, e.g., the particular solution \( \hat{z}_k = 0\) for all \( k \in \mathbb{N}\) . Then, the estimate is obtained by solving \( T_k(\hat{x}_k)=\hat{z}_k=0\) , namely

\[ \begin{equation} \hat{x}_k = -\begin{pmatrix}a_{1,k} b_{2,k} - a_{2,k} b_{1,k} \\ a_{1,k} b_{3,k} - a_{3,k} b_{1,k} \end{pmatrix}^{-1}\begin{pmatrix}a_{1,k} c_{2,k} - a_{2,k} c_{1,k} \\ a_{1,k} c_{3,k} - a_{3,k} c_{1,k} \end{pmatrix}. \end{equation} \]

(143)

2.4.4.2 Discrete-Time KKL Observer with Rotation Correction

We now design the KKL observer in Section 2.3 but for the model (134), where the discretization scheme incorporates the system’s rotating dynamics. Note that system (134), with the inputs \( (u_k)_{k \in \mathbb{N}}\) and \( (i_k)_{k \in \mathbb{N}}\) being sinusoidal, satisfies all the assumptions required by Theorem 5. Keeping the same pair \( (A, B)\) , we get this time different dynamics of \( (a_k,b_k,c_k)_{k\in\mathbb{N}}\)

\[ \begin{align} \begin{split} a_{k+1} &= A a_k + B,\\ b_{k+1} & = A b_k -2\tau a_{k+1}(\Omega_k(u_k - Ri_k)\rm{sinc}(\varphi_k))^\top - 2LBi_k^\top,\\ c_{k+1} &= A c_k -\tau^2 a_{k+1}\rm{sinc}^2(\varphi_k)|\Omega_k(u_k - Ri_k)|^2 \\ &~~{}- \tau b_{k+1}\Omega_k(u_k - Ri_k)\rm{sinc}(\varphi_k)+ B(L^2|i_k|^2 - \Phi^2), \end{split} \end{align} \]

(144)

with \( a_k\) still possibly constant equal to \( (\rm{Id}-A)^{-1}B\) , and the estimate is still obtained with (143). In Figure 17, the estimation errors with respect to the continuous-time trajectory of system (132) are compared among: i) The continuous-time KKL observer from [8] with \( A_c = -\rm{diag}(10, 44, 80)\) discretized at \( \tau = 10^{-3}\) (s) using Euler’s scheme; ii) The discrete-time KKL observer (142)-(143); and iii) The discrete-time KKL observer with rotation correction (144)-(143), for \( A = e^{\tau A_c}\) , \( B = (1,1,1)\) , and \( \gamma = 1\) .

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation errors of: i) A continuous-time &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;KKL&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Kravaris-Kazantzis$$Luenberger&lt;/span&gt;&lt;/span&gt; observer designed following&amp;#160;bernardPMSM and discretized using Euler&amp;rsquo;s method, ii) Observer&amp;#160;(eq:ch6_pmsm_dynamics_discrete)-(eq:ch6_pmsm_xhat), and iii) Observer&amp;#160;(eq:ch6_pmsm_dynamics_discrete_r)-(eq:ch6_pmsm_xhat).&lt;/span&gt;
Figure 17. Estimation errors of: i) A continuous-time KKL observer designed following [8] and discretized using Euler’s method, ii) Observer (142)-(143), and iii) Observer (144)-(143).

Similar to Section 2.4.3, we see that it is better not to discretize an observer already designed so as not to introduce numerical errors, and that the discretization scheme has a huge impact on the precision of the discretized model, thus determining the observer’s performance.

In Figure 18, we compare the estimation errors given by the discrete-time KKL observer for model (134) but with different choices of \( \gamma\) . Similar to Section 2.4.3, it is observed that a smaller \( \gamma\) gives a faster convergence, but a more serious amplification of numerical noise, which is coherent with the robustness results in Theorem 6. However, in the region of too high rotating speeds, the three choices of \( \gamma\) tend to perform the same, since the discretized model becomes less appropriate, which is something the observers cannot deal with. Last, it is interesting to notice that in this application case, as we choose for the observer in the \( z\) -coordinates \( \hat{z}_k = 0\) for all \( k \in \mathbb{N}\) , it is indeed the transformation \( (T_k)_{k \in \mathbb{N}}\) that serves to provide the estimation.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation errors of the discrete-time &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;KKL&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Kravaris-Kazantzis$$Luenberger&lt;/span&gt;&lt;/span&gt; observer&amp;#160;(eq:ch6_pmsm_dynamics_discrete_r)-(eq:ch6_pmsm_xhat), design for system&amp;#160;(eq:ch6_pmsm_bernard) with A = e^{ A_c}) and B = (1,1,1)) where A_c = -diag(10, 44, 80)) , with three different choices of ) .&lt;/span&gt;
Figure 18. Estimation errors of the discrete-time KKL observer (144)-(143), design for system (134) with \( A = e^{\tau A_c}\) and \( B = (1,1,1)\) where \( A_c = -\rm{diag}(10, 44, 80)\) , with three different choices of \( \gamma\) .

2.4.5 High-Gain vs. KKL Observer

In this section, we compare the performance of the two discrete-time observers, both designed for system (134) whose rotating dynamics are taken into account, in the presence of random noise of zero mean and \( 0.001\) standard deviation in both the voltage and current inputs. From the estimation errors in Figure 19, we see that the KKL design outperforms the high-gain one. At the beginning and the end of the simulation scenario, the high-gain observer becomes more susceptible to noise because of the bad conditioning of the inverse of (140), when the motor speed is low making the successive outputs take very close values. The KKL observer performs well since its transformation is computed dynamically using (144) and therefore has a built-in filtering effect against noise by keeping memories of all the past outputs, while that of the high-gain observer is not dynamical and only uses a moving window made of \( m\) past inputs and outputs (see Lemma 6). Another advantage of the KKL observer in this application is that its \( \gamma\) does not have to be pushed exceedingly small, making it less susceptible to noise. Therefore, another lesson to draw here is to do KKL when we can and only use the high-gain observer when the KKL one is not implementable (such as when its transformation into the target coordinates cannot be easily computed). Note again that while both designs exploit backward distinguishability in this PMSM application, the advantages of the high-gain observer in Section 2.2 are that it can still work under a weaker constructibility condition and it does not necessarily require the invertibility of the system’s dynamics.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation errors of the discrete-time high-gain and &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;KKL&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Kravaris-Kazantzis$$Luenberger&lt;/span&gt;&lt;/span&gt; observers, both designed for system&amp;#160;(eq:ch6_pmsm_bernard), in the presence of noise.&lt;/span&gt;
Figure 19. Estimation errors of the discrete-time high-gain and KKL observers, both designed for system (134), in the presence of noise.

2.4.6 Conclusion

In this chapter, we apply the two novel discrete-time observers developed in this dissertation to a PMSM discretized using two different methods: a standard Euler scheme and a more sophisticated approach that captures the system’s rotational dynamics. Through performance comparisons between the two observers and their discretized continuous-time versions, the following important insights are drawn from these applications:

  • It can be better to design a discrete-time observer from a discretized model than to discretize a continuous-time observer already designed, so as not to introduce discretization errors. Note that the possibilities of following these two routes of observer design have been studied in [130], which provides sufficient conditions for each to work;
  • The numerical errors due to incorrect discretization can be considerably reduced by taking into account the system’s physics in the discrete-time model;
  • Between the high-gain observer in Section 2.2 and the KKL one in Section 2.3, we should go with the latter when we can, to benefit from its superior filtering capability, and only use the former when the KKL one is not implementable.

3 Observer Design for Hybrid Systems with Known Jump Times

3.1 Review of Observer Designs for Hybrid Systems with Known Jump Times

 

Dans ce chapitre, nous passons en revue les conceptions d’observateurs pour des systèmes hybrides dont les instants de saut, c’est-à-dire les moments où les solutions sautent, sont connus ou détectés. Dans ce contexte, nous définissons d’abord la détectabilité (pré-)asymptotique de ces systèmes comme une condition nécessaire à l’existence d’un observateur asymptotique synchronisé avec le système. Étant donné que l’observateur est synchronisé, l’estimation partage le même domaine temporel que la solution du système, ce qui permet de les comparer au même temps hybride. Nous examinons ensuite les conceptions d’observateurs pour cette large classe de systèmes, y compris les classes particulières des systèmes impulsifs/commutés et des systèmes en temps continu avec des mesures sporadiques. Enfin, nous tirons des conclusions.

3.1.1 Overview

Hybrid systems are widely studied and have many applications, e.g., impulsive systems, walking robots, biology, etc. [10]. However, the observer design problem for this class of systems is still largely unsolved mainly because the time domain of each hybrid solution typically depends on its initial condition and is thus unknown to the observer. Hence, the time domain of the system and observer solutions typically differ, making both design and analysis of convergence challenging [67]. Actually, very few general results exist apart from [32] assuming the existence of a gluing function, or [131] assuming the flow dynamics are differentially observable, as reviewed later in Section 4.1. In the favorable case considered in this Section 3 where the solutions’ jump times, namely the times at which discrete events appear, are known or detected, the observer’s jumps can instead be synchronized accordingly, leading to the system and observer solutions having the same time domain. Consider a hybrid system of the form

\[ \begin{equation} \mathcal{H} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{~~}l@{}} \dot{x} &=& f(x,u_c)&(x,u_c) \in C&y_c = h_c(x,u_c)\\ x^+ &=& g(x,u_d)& (x,u_d) \in D& y_d = h_d(x,u_d) \end{array} \right. \end{equation} \]

(145)

where \( x \in \mathbb{R}^{n_x}\) is the state, \( u_c \in \mathbb{R}^{n_{u,c}}\) (resp., \( u_d \in \mathbb{R}^{n_{u,d}}\) ) is the exogenous flow (resp., jump) input that may include the continuous time \( t\) (resp., discrete time \( j\) ), \( y_c \in \mathbb{R}^{n_{y,c}}\) (resp., \( y_d \in \mathbb{R}^{n_{y,d}}\) ) is the flow (resp., jump) output, \( C \subseteq \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}}\) (resp., \( D \subseteq \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,d}}\) ) is the flow (resp., jump) sets, \( f:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \to \mathbb{R}^{n_x}\) (resp., \( g: \mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,d}} \to \mathbb{R}^{n_x}\) ) is the flow (resp., jump) map, and \( h_c:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \to \mathbb{R}^{n_{y,c}}\) (resp., \( h_d:\mathbb{R}^{n_x}\times\mathbb{R}^{n_{u,c}} \to \mathbb{R}^{n_{y,d}}\) ) is the flow (resp., jump) output map. See Definition 3 for the definitions of the inputs and solutions to system (145).

The goal of this Section 3 is to design an asymptotic observer for system \( \mathcal{H}\) in (145) (as defined below in (146)), assuming we know: i) when the plant’s jumps occur, ii) the output(s) \( y_c\) during flows and\( /\) or \( y_d\) at jumps, iii) the exogenous signals \( u_c\) and \( u_d\) , and iv) possibly some information about the admitted flow lengths (i.e., the time between two consecutive jumps). Since in practice, we may be interested in estimating only certain trajectories of “physical interest”, we denote in what follows \( \mathcal{X}_0\) a set containing the initial conditions of the trajectories to be estimated and \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ) a set of locally bounded continuous (resp., discrete) inputs of interest, defined on \( \mathbb{R}_{\geq 0}\) (resp., \( \mathbb{N}\) ). We then denote \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) as the set of maximal solutions to system \( \mathcal{H}\) initialized in \( \mathcal{X}_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , as defined in Section 1.5.3. Since the jump times of system (145) are known, following [22], a synchronized asymptotic observer can be designed for this system, with dynamics of the form

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{z}}&=& \mathcal{F}(\hat{z},y_c,u_c) &\text{when~(145) flows}\\ \hat{z}^+&=&\mathcal{G}(\hat{z},y_d,u_d) &\text{when~(145) jumps} \end{array} \right. \end{equation} \]

(146.a)

with the estimate \( \hat{x}\) obtained from a solution \( (t,j) \mapsto \hat{z}(t,j)\) to dynamics (146.a) as

\[ \begin{equation} \hat{x}(t,j)= \Upsilon(\hat{z}(t,j),t,j), \end{equation} \]

(146.b)

where \( \hat{z} \in \mathbb{R}^{n_z}\) is the observer state; \( \mathcal{F}\) , \( \mathcal{G}\) , and \( \Upsilon\) are the observer dynamics and output maps designed together with an initialization set \( \mathcal{Z}_0 \subseteq \mathbb{R}^{n_z}\) such that the dependence of \( \Upsilon\) on time \( (t,j)\) is only through the inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) and each maximal solution \( (x,\hat{z})\) to the cascade (145)-(146) initialized in \( \mathcal{X}_0 \times \mathcal{Z}_0\) and with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in \mathfrak{U}_c\times \mathfrak{U}_d\) is complete and verifies

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x (= \rm{dom} \hat{x})}}|x(t,j)-\hat{x}(t,j)| = 0. \end{equation} \]

(147)

That the dynamics (146.a) are triggered synchronously with system (145) makes \( \hat{z}\) and \( \hat{x}\) share the same time domain as \( x\) , allowing them to be compared at the same hybrid time as in (147).

Remark 19

As shown in [22, Section 6], a dwell time after jump \( j_m\) typically ensured the robustness of the observer against delays in jump detection, namely when the jumps of observer (146) are not triggered simultaneously with those of system (145), but slightly later. More precisely, the semi-global practical stability of the estimation error could be obtained over the time intervals after such delays. Robustness against delays in jump detection but in the context of constant parameter estimation in hybrid systems is established in [132, Chapter 7]. Keeping these in mind, in the whole Section 3, we will assume that the jump times are perfectly known and observers are triggered without any delay, giving exact synchronization of the observer with the system, to focus on observer design for different forms of system (145).

While sufficient conditions providing observers of the form (146) are provided in Section 3.2 and Section 3.3, let us start by defining and presenting some necessary conditions for their existence. We define the (pre-)asymptotic detectability of system (145) with known jump times.

Definition 14 ((Pre-)asymptotic detectability of system (145))

System (145) with known jump times is pre-asymptotically detectable on \( \mathcal{X}_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) if any two complete solutions \( x_a\) and \( x_b\) in \( §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) with the same inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) such that \( \rm{dom} x_a = \rm{dom} x_b := \mathcal{D}\) and whose flow outputs \( y_{a,c}\) , \( y_{b,c}\) and jump outputs \( y_{a,d}\) , \( y_{b,d}\) satisfy

\[ \begin{align} y_{a,c}(t,j) &= y_{b,c}(t,j), &\forall& t \in \rm{int}(\mathcal{T}_j(x_a))=\rm{int}(\mathcal{T}_j(x_b)), \forall j \in \rm{dom}_j x_a = \rm{dom}_j x_b, \end{align} \]

(148.a)

\[ \begin{align} y_{a,d}(t,j)&= y_{b,d}(t,j), &\forall &(t,j) \in \rm{dom} \mathcal{D} \text{ such that } (t,j+1) \in \mathcal{D}, \end{align} \]

(148.b)

verify

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \mathcal{D}}} |x_a(t,j)-x_b(t,j)| = 0. \end{equation} \]

(149)

If, in addition, each solution in \( §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) is complete, then system (145) is said to be asymptotically detectable. The set \( \mathcal{X}_0\) (resp., \( \mathfrak{U}_c \times \mathfrak{U}_d\) ) may be omitted if the property holds for any initial condition in \( \mathcal{X}_0 = \rm{cl}(C) \cup D\) (resp., any input).

Remark 20

In the case where the domain of solutions is such that all flow intervals have non-empty interior, i.e., if no successive jumps can occur, provided that \( t \mapsto y_c(t,j)\) is continuous during flows for all \( j\) in the \( j\) -domain, (148.a) simplifies to equality for all \( (t,j)\) in the domain. On the contrary, when a solution admits consecutive jumps, condition (148.a) is required only on the flow intervals with a non-empty interior since it holds vacuously on the others. In other words, the equality of \( y_c\) is only required when the system is flowing. Last, for some hybrid systems \( \mathcal{H}\) , the knowledge of the jump times determines uniquely the solution (or its initial condition). In that case, asymptotic detectability in the sense of Definition 14 holds vacuously because there does not exist a pair of distinct solutions with the same time domain and the same outputs.

Lemma 9 below shows that this asymptotic detectability property is a necessary condition for the design of a synchronized asymptotic observer (146).

Lemma 9 (Asymptotic detectability is necessary for an asymptotic observer)

If there exists a synchronized asymptotic observer (146) for system (145) initialized in \( \mathcal{X}_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) , then system (145) is asymptotically detectable on \( \mathcal{X}_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) (following Definition 14).

Proof. First, by the definition of a synchronized asymptotic observer for system (145) initialized in \( \mathcal{X}_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) given in (146) above, all solutions in \( §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) are complete. Consider two complete solutions \( x_a\) and \( x_b\) in \( §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) with the same inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) such that \( \rm{dom} x_a = \rm{dom} x_b := \mathcal{D}\) , whose respective flow outputs \( y_{a,c}\) , \( y_{b,c}\) and jump outputs \( y_{a,d}\) , \( y_{b,d}\) satisfy (148). Consider a hybrid arc \( \hat{z}\) defined on \( \rm{dom} x_a\) such that \( (x_a,\hat{z})\) is solution to the cascade (145)-(146) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) and output \( \hat{x}_a\) . According to (148), \( (x_b,\hat{z})\) is also solution to the cascade (145)-(146) with the same inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) and output \( \hat{x}_b\) such that \( \hat{x}_a(t,j)=\hat{x}_b(t,j)\) for all \( (t,j)\in \mathcal{D}\) (since the dependence of \( \Upsilon\) on time is only through the inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) and not the outputs \( (y_c,y_d)\) ). By the definition of a synchronized asymptotic observer given in (146) above, we have both \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \mathcal{D}}} |x_a(t,j)-\hat{x}_a(t,j)| = 0\) , and \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \mathcal{D}}} |x_b(t,j)-\hat{x}_b(t,j)| = 0\) . Therefore, we have (147) by the triangle inequality. \( \blacksquare\)

In fact, the detectability and observability of hybrid systems have been quite extensively studied and defined in the literature. Roughly speaking, (asymptotic) detectability as in Definition 14 means that any complete solutions producing the same output must asymptotically converge to each other—it is thus an incremental notion—while observability means that the output determines uniquely the solution (possibly over a given time window). Usually, those properties constitute necessary conditions for the existence of an observer depending on its properties (convergence, stability, tunability, robustness, etc.)—see [1]. In [67], detectability is defined for general hybrid systems with nonlinear maps and unknown jump times in a way that makes this property necessary for the existence of an asymptotic observer. Re-parameterization of the solutions is therefore needed to compare them on a common hybrid time domain. In the context of known jump times of this Section 3, detectability in Definition 14 reduces to the incremental convergence of solutions with the same time domain and the same output. Stronger notions of detectability could be similarly defined when asking more of the observer, for instance, stability properties as in [133]. When the dynamics and output maps are linear, detectability (resp., observability) in turn reduces to zero detectability (resp., zero observability) as considered in [134, 135]. Also thanks to linearity in the maps, algebraic observability certificates are developed in [11, 136, 137, 138, 134, 135]. Note that detectability\( /\) observability notions have also been developed for hybrid automata, i.e., systems with both continuous states and discrete, event-triggered ones (e.g., [139, 140, 141]), but using somewhat different vocabulary.

In those works, criteria are exhibited to determine whether the system’s full state is detectable\( /\) observable or not. However, in hybrid systems, state components may exhibit different kinds of observability properties, associated with the flow and\( /\) or jump output(s) or even hidden inside the flow-jump coupling. Such information is interesting when designing an observer since different observation strategies may then be developed and combined depending on the source of observability. It has been suggested from [142] that for linear time-varying systems, components with different observability properties can be separated from each other by means of decomposition. In the context of output regulation and internal model design, [143] extends these ideas to hybrid systems with linear maps and periodic jumps with outputs during flows only. The goal of an internal model is indeed to extract and model the dynamics that are able to generate the outputs. In that way, observability decompositions are relevant to extracting the dynamics “seen from the outputs”. Indeed, it is seen that a part of the dynamics is instantaneously observable during flows from the flow output, while a part of the non-observable dynamics becomes visible in the observable states at jumps. The rest of the dynamics, called the invisible dynamics, may be discarded for internal model design. This idea, however, is still limited in [143] to the case of periodic jumps and flow output only.

In this Section 3, since the jump times are known, the flow and jump outputs \( y_c\) and \( y_d\) are distinguished, unlike in the context of unknown jump times when a single output is measured at all times. Due to the hybrid nature of the system, detectability\( /\) observability can either come during flows from the flow output \( y_c\) , come at jumps thanks to the jump output \( y_d\) , or emerge as a result of the flow-jump combination. As illustrated in Example 12, a hybrid system with both undetectable flow and jump parts can actually be detectable (and vice-versa), signifying that detectability is not only a property of the maps \( (f,h_c,g,h_d)\) but the sets \( (C,D)\) as well.

Example 12 (Detectability of a hybrid system)

Consider the hybrid system with dynamics

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{x}_1 &=& x_2 \\ \dot{x}_2 &=& 0 \\ \dot{x}_3 &=& 0 \end{array}\right. ~~ y_c=x_1,~~ ~~ \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}}x_1^+ &=& x_1 \\ x_2^+ &=& x_2 \\ x_3^+ &=& x_1\end{array}\right. ~~ y_d=x_1, \end{equation} \]

(150)

and some flow and jump sets \( (C,D)\) . We see that \( x_3\) is undetectable from the flow part of system (150), and \( x_2\) is undetectable from the jump part. So, if solutions only flow or jump, then they are not detectable. However, if the sets \( (C,D)\) make flows and jumps persist, then the full state becomes detectable by combining the information from \( y_c\) and \( y_d\) .

Example 12 illustrates the complexity of hybrid systems in terms of detectability. After discussing detectability\( /\) observability of hybrid systems (with known jump times), we review existing literature on observer designs for these. We start with two important classes, namely impulsive\( /\) switched systems and continuous-time systems with sampled measurements, then consider general hybrid systems of the form (145).

3.1.2 Observers for Special Classes of Hybrid Systems

3.1.2.1 Designs for Impulsive\( /\) Switched Systems

In the literature, impulsive systems take the form

\[ \begin{align} \dot{x}(t)& = f(x(t),u(t)),& t &\neq t_k, \\ x(t_k^+) &= g(x(t_k^-),t_k,k),& t &= t_k,\\ y(t) &= h(x(t),u(t)),&& \\\end{align} \]

(151.a)

for \( t_1 < t_2 < … < t_k < …\) a given sequence of times when impulses take place, making the state jump instantly from \( x(t_k^-)\) to \( x(t_k^+)\) . On the other hand, switched systems are modeled by

\[ \begin{equation} \dot{x}(t) = f_\sigma(x(t),u(t)), ~~ y(t) = h_\sigma(x(t),u(t)), \end{equation} \]

(152)

where the exogenous switching signal \( \sigma\) determines the active mode. These are thus continuous-time systems with discrete events such as switches or impulses occurring at specific time instants that are generally known and therefore constitute an important large class of hybrid systems with known jump times with a wide range of applications, e.g., power electronics [144]. Their observability\( /\) determinability is analyzed (possibly with algebraic certificates) in [11, 136, 137, 138, 134]. In [145], switching observers are designed using a Linear Matrix Inequality[LMI] for both continuous- and discrete-time linear systems assuming the full detectability of each mode, while the switching signal is completely known. In the linear context, [134] suggests a change of coordinates based on the Kalman decomposition to extract the observable components of each individual mode, which are then estimated during each interval using a Luenberger observer. [12] develops parallel observers estimating the observable part of each mode under pre-determined transitions of bounded varying intervals, relying on the determinability gained after a high enough number of switches. These two works thus exploit observability gained by accumulating information from individual non-observable sub-systems under persistent switching. However, so far, results for switched systems mostly concern those with linear maps.

3.1.2.2 Designs for Continuous-Time Systems with Sampled Measurements

Oftentimes, for various reasons such as bandwidth limitations, communication safety, or simply the availability of the measurements, the output of a continuous-time system is only taken at specific time instants, leading to systems of the form

\[ \begin{equation} \dot{x}(t) = f(x(t),u(t)), ~~ y(t_k) = h(x(t_k)), \end{equation} \]

(153)

for \( t_1 < t_2 < … < t_k < …\) a sequence of times when the measurement is performed. This is a particular class of hybrid systems (145) with an identity jump map and known jump times. Some observer designs consist of adapting existing continuous-time observers, usually with constraints on the maximal sampling period. For instance, [146] adapts the high-gain observer design into a continuous-discrete extended Kalman filter where exponential stability is achieved if the gain belongs to an interval that contracts when the constant sampling period increases; and [147] proposes an impulsive Luenberger observer for linear systems with quantized measurements that guarantees stability if the varying quantizer transition inter-arrival time is taken sufficiently small. LMI-based impulsive and sample-and-hold designs with varying sampling times are developed in [148, 149] respectively, taking into account the Lipschitzness of the system nonlinearity. Moreover, a sample-and-hold observer is designed in [150] with a time-varying adaptation gain that is reset to \( 1\) at each sampling instant and decreases in between sampling events. On the other hand, hybrid designs with correction at sampling times only have also been developed using LMIs, with both a polytopic approach [151] and a grid-based one [152]. Similarly to Section 3.1.2.1, results for this class of systems are mostly restricted to the linear context.

In nonlinear settings, attempts have been made in [153], supposing that a continuous-time observer has been designed as if the output were continuous, by proposing a hybrid observer consisting of the said observer and an output predictor for the time interval between two consecutive measurements. The work [154] does this by letting the estimate evolve with the system’s dynamics and correcting it each time a measurement arrives, using gains obtained by LMIs.

3.1.3 Observers for General Hybrid Systems (145)

As shown in [22], many of those designs can be unified under the hybrid framework of [10], assuming that some information is known about the time elapsed between successive jumps. Currently, existing results typically assume at least one of the following properties:

  • Observability of the flow dynamics or jump dynamics via Lyapunov\( /\) LMI-based sufficient conditions, such as [155] or [22, Section 3], but without constructive observability-based criteria to check their solvability;
  • Observability of the full state during flows from the flow output only, leading to flow-based observers, essentially exploiting the continuous-time high-gain observer design and an average dwell time, i.e., persistent flowing; see for example [22, Section 4];
  • Observability of the full state from the jump output only thanks to the combination of flows and jumps, leading to jump-based observers, essentially exploiting discrete-time observers on an equivalent discrete-time system sampled at the jumps and a persistence of jumps; see for instance [22, Section 5] or [156, 151, 152].

As contributions of this dissertation, in this Section 3, we propose several novel observers for hybrid systems with known jump times, relying on the combination of flows and jumps. We consider the cases when the maps are linear and nonlinear in Section 3.2 and Section 3.3, respectively. In the linear context, we first propose in Section 3.2.2 a systematic Kalman-like observer design under a Gramian-based observability property that gathers and combines observability from both flows and jumps. Then, in Section 3.2.3, we were inspired by the idea of decomposing the state into components of different observability conditions [143] to propose observers for hybrid systems with linear maps, without requiring the periodicity of the jumps. The designs in this part exploit an observability decomposition of the system state, where a part is instantaneously observable from the flow output, while the rest is (at least) detectable from the jump output as well as the combination of flows and jumps. Indeed, as observed in [143] and illustrated in Example 13, the interaction between flows and jumps may render visible, in the flow output, state components that are not observable through the flow dynamics. This special source of observability constitutes a hidden measurement that we refer to in the subsequent chapters as the fictitious output. This output proves to be very handy for estimating unmeasurable parameters, i.e., those that cannot be measured by any physical sensors, such as sensor biases or restitution coefficients, and only become observable thanks to their interaction with the rest of the state components that are either measured or estimated sufficiently fast. This is illustrated with our applications in Section 3.4 to mechanical systems with impacts.

Example 13 (Fictitious output in a hybrid system)

Consider the hybrid system with dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{x}_1 &=& x_2 \\ \dot{x}_2 &=& 0 \\ \dot{x}_3 &=& 0 \end{array} \right. ~~ y_c=x_1, ~~ ~~ \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} x_1^+ &=& x_1 \\ x_2^+ &=& x_3 \\ x_3^+ &=& x_3 \end{array} \right. ~~ y_d=x_1, \end{equation} \]

(154)

and some flow and jump sets that make flows and jumps persist. We see that \( (x_1,x_2)\) are instantaneously observable from \( y_c\) (and \( \dot{y}_c\) ) during flows, but \( x_3\) is not detectable from the flow or the jump part (unlike in Example 12 where it is detectable from the jump part). However, notice that since \( x_2^+\) can be estimated arbitrarily fast knowing \( \dot{y}_c\) during flows, it is as if we have an extra output \( \dot{y}_c = x_3\) after each jump. This output is fictitious and it stems from the flow-jump combination, when a part of the state interacts with another one that is instantaneously observable, making it visible in the next flow interval.

The same design spirit is pursued in Section 3.3 when the maps are nonlinear. In such a context, we propose a general framework of still estimating arbitrarily fast a part of the state using \( y_c\) , then estimate the rest from \( y_d\) as well as its interaction with the first part at jumps. These observers are coupled under Lyapunov conditions exhibited based on the robustness of the continuous-time high-gain observer [53] and the novel discrete-time designs in Section 2.

3.1.4 Conclusion

In this chapter, we first define and discuss the (pre-)asymptotic detectability of hybrid systems with known jump times. We then review observer designs for this large class of systems, including the special classes of impulsive\( /\) switched systems and continuous-time systems with sporadic measurements. In the following, observers will be designed for hybrid systems with linear maps in Section 3.2 and nonlinear maps in Section 3.3, followed by their applications to mechanical systems with impacts in Section 3.4.

3.2 Observer Design for Hybrid Systems with Linear Maps

 

Ce chapitre développe des observateurs pour les systèmes hybrides avec des dynamiques et sorties linéaires, dont les instants de saut sont connus. Nous définissons et analysons la détectabilité (pré-)asymptotique et l’observabilité complète uniforme de cette classe de systèmes, puis présentons deux approches différentes pour la conception d’observateurs. La première repose sur un observateur synchronisé de type Kalman qui intègre l’observabilité provenant à la fois du flot et des sauts. Des exemples, incluant des systèmes à commutation et des systèmes en temps continu avec des mesures sporadiques (à multi-fréquences), illustrent nos méthodes. La deuxième consiste à décomposer l’état en parties avec des propriétés d’observabilité différentes et à coupler des observateurs estimant chacune de ces parties, en exploitant une mesure fictive supplémentaire provenant de la combinaison des flux et des sauts. Ces observateurs sont basés sur une Inégalité Matricielle Linéaire (LMI) ou le paradigme de Kravaris-Kazantzis$\slash$Luenberger[KKL]. Une comparaison de ces méthodes est présentée dans un tableau à la fin du chapitre.

3.2.1 Introduction

The content of this chapter has been published in [157, 158].

Consider a hybrid system with linear maps defined as

\[ \begin{equation} \mathcal{H}\left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}c@{~~}l@{}} \dot{x}&=&Fx + u_c & (x, u_c) \in C & y_c=H_c x\\ x^+&=&Jx+u_d & (x, u_d)\in D & y_d = H_d x \end{array} \right. \end{equation} \]

(155)

where \( x \in \mathbb{R}^{n_x}\) is the state, \( C\) and \( D\) are the flow and jump sets, \( y_c \in \mathbb{R}^{n_{y,c}}\) and \( y_d \in \mathbb{R}^{n_{y,d}}\) are the outputs known during the flow intervals and at the jump times respectively, \( u_c \in \mathbb{R}^{n_x}\) and \( u_d \in \mathbb{R}^{n_x}\) are known exogenous inputs, as well as the dynamics matrices \( F, J \in \mathbb{R}^{n_x \times n_x}\) and the output matrices \( H_c \in \mathbb{R}^{n_{y,c} \times n_x}\) , \( H_d \in \mathbb{R}^{n_{y,d} \times n_x}\) which are all known and possibly time-varying.

Remark 21

Models of the form (155) encompass not only hybrid systems with linear maps described in the setting of [10], but also switched and\( /\) or impulsive systems with linear maps where the active mode is treated as an exogenous signal making \( (F, J, H_c, H_d)\) time-varying (see [145, 156, 134] among many others and Example 14 below), and continuous-time systems with sporadic or multi-rate sampled outputs where “jumps” correspond to sampling events, \( J=\rm{Id}\) , \( u_d=0\) , \( y_c=0\) , and \( y_d\) the outputs available at the sampling event [148, 151, 152]. See [10] and Example 16 below for some examples of those classes of systems set in the framework of system (155).

This chapter aims to design an asymptotic observer for system (155), assuming that its jump times are known or detected. In practical applications, the objective may be to estimate specific trajectories of “physical interest”, initialized in some set \( \mathcal{X}_0\subset \mathbb{R}^{n_x}\) and with exogenous terms \( (F,J,H_c,H_d,u_c,u_d)\) , as trajectories of time, contained in some set \( \mathfrak{U}\) of the trajectories of interest. We then denote \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) as the set of those maximal solutions of interest. Because we look for an asymptotic observer, we assume that maximal solutions are complete as follows.

Assumption 9

Given \( \mathcal{X}_0\) and \( \mathfrak{U}\) , each solution in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) is complete.

Since the jump times of solutions to system (155) are known, it is natural to strive for a synchronized asymptotic observer of the form (146). The knowledge of the jump times is not only used to trigger the observer jumps at the same time as those of the system, but it can also be used to design the observer maps \( \mathcal{F}\) and \( \mathcal{G}\) . The manner in which this information is utilized depends on the type of observer gains:

  • Gains that are computed offline (for example via matrix inequalities), according to all possible lengths of flow intervals in between jumps, i.e., they depend on each individual flow length, not the particular sequence of them. This is the path taken by [145, 159, 151, 152, 155, 22] and Section 3.2.3;
  • Or, gains that are computed online along the time domain of each solution of interest, i.e., they depend on the sequence of flow lengths in each particular solution. This is the path taken by all Kalman-like approaches in [156, 160] and Section 3.2.2.

In the former case, the design requires some information about the possible duration of flow intervals between successive jumps in each solution of interest (at least after a certain time) as formalized next. In the context of switched (resp., sampled) systems, this corresponds to information about the possible switching (resp., sampling) rates.

Definition 15 (Set of flow lengths of a hybrid arc)

For a closed subset \( \mathcal{I}\) of \( [0,+\infty)\) and some \( j_m \in \mathbb{N}\) , we say that a hybrid arc \( (t,j) \mapsto x(t,j)\) has flow lengths within \( \mathcal{I}\) after jump \( j_m\) if

  • \( 0 \leq t - t_j(x) \leq \sup\mathcal{I}\) for all \( (t,j) \in \rm{dom} x\) ;
  • \( t_{j+1}(x) - t_j(x) \in \mathcal{I}\) holds for all \( j \in \mathbb{N}_{\geq j_m}\) if \( \sup \rm{dom}_j x = +\infty\) , and for all \( j \in \{j_m,j_m+1, …, \sup \rm{dom}_j x - 1\}\) otherwise.

In brief, \( \mathcal{I}\) contains all the possible lengths of the flow intervals between successive jumps, at least after some time. The first item is to bound the length of the flow intervals not covered by the second item, namely possibly the first ones before \( j_m\) , and the last one, which is \( \rm{dom}_t x \cap [t_{J(x)}, +\infty)\) where \( t_{J(x)}\) is the time when the last jump happens (when defined). If \( \mathcal{I}\) is unbounded, the system may admit (eventually) continuous solutions and the observer should correct the estimate at least during flows, while \( 0 \in \mathcal{I}\) means the hybrid arc can jump more than once at the same time instance or have flow lengths going to zero (including (eventually) discrete and Zeno solutions) and the observer should reduce the estimation error primarily at jumps.

From this, one may design either:

  • A flow-based observer with an innovation term during flows only, leveraging the observability of the full state during flows from \( y_c\) when \( 0\notin \mathcal{I}\)  [145, 22];
  • A jump-based observer with an innovation term at jumps only, exploiting the detectability of the full state via the combination of flows and jumps from \( y_d\) available at the jumps only when \( \mathcal{I}\) is bounded [156, 151, 152, 160, 22];
  • An observer with innovation terms during both flows and jumps, leveraging the observability from both \( y_c\) and \( y_d\) and the combination of flows and jumps: this is done via a hybrid Kalman-like approach or an observability decomposition as in this chapter, or Lyapunov-based LMIs in [155, 22].

This chapter focuses on the third case, where the full state is not necessarily instantaneously observable during flows and is not observable from the jump output alone. We propose two routes for observer design: one based on hybrid Gramian conditions and the other on an observability decomposition. The latter decomposes the state in two parts: the first one is instantaneously observable through the flow output \( y_c\) , while the second one must be detectable from an extended jump output featuring the available jump output \( y_d\) and an additional fictitious one, describing how the non-observable states impact the observable ones at jumps and become visible through \( y_c\) . While the Gramian-based analysis has led in Section 3.2.2 to a systematic hybrid Kalman-like design, we demonstrate in Section 3.2.3 how the observability decomposition enables the design of observers composed of:

  • A Kalman-like high-gain flow-based observer of the state components that are instantaneously observable from \( y_c\) ;
  • A jump-based observer for the remaining components, derived from a discrete-time LMI-based (resp., KKL-based) observer in Section 3.2.3.3 (resp., Section 3.2.3.4).

As shown in Lemma 9, the existence of an asymptotic and synchronized observer for system (155) requires it to be asymptotically detectable, in the sense of Definition 14. For observer design, one typically requires stronger observability assumptions depending on the class of observers and the required observer properties [1]. Observability typically means that the equality of the outputs in (148), possibly over a large enough time window, implies that the solutions \( x_a\) and \( x_b\) are actually the same or said differently, there does not exist any pair of distinct solutions with the same time domain and the same outputs in the sense of (148). Actually, in the context of observer design, a more relevant property is the ability to determine uniquely the current state from the knowledge of the past outputs over a certain time window \( \Delta > 0\) , which is typically called backward distinguishability as in Section 2.3 or constructibility [73]. In other words, for all \( (t,j)\) in the domain such that \( t+j\geq \Delta\) , the equality of the outputs \( (y_c,y_d)\) along \( x_a\) and \( x_b\) at all past times \( (t^\prime,j^\prime)\) in the domain such that \( 0\leq (t+j)-(t^\prime+j^\prime)\leq \Delta\) implies that \( x_a(t,j)=x_b(t,j)\) (see later in Section 3.2.2.1). For continuous-time systems, this property is equivalent to observability over a time window because of the uniqueness of solutions in forward and backward time. However, in discrete-time or hybrid systems, when the jump maps are not invertible, the two properties can diverge: a system could be constructible without being observable (see Section 2.2 or [73, Section 2.3.3] for detailed discussions on those notions).

3.2.2 Hybrid Kalman-Like Observer

Assuming the jump times of the solutions \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) are exactly known or detected—for instance from discontinuities in the output, impact sensors, or because they are triggered by the user or the availability of the sensor in the sampled-data case—and exploiting the linearity of the maps of system \( \mathcal{H}\) , we propose a systematic design of a synchronized hybrid Kalman-like observer of the form

\[ \begin{equation} \hat{\mathcal{H}}\left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{x}} &=& F\hat{x} + u_c + PH_c^\top R_c^{-1}(y_c-H_c\hat{x}) \\ \dot{P} &=& \lambda P + FP + PF^\top -PH_c^\top R_c^{-1} H_{c}P \end{array}\right\}\text{when \( \mathcal{H}\) flows}\\ \\ \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{x}^+ &=& J\hat{x} + u_d + J K(y_d-H_d\hat{x}) \\ P^+&=& \gamma^{-1} J(\rm{Id}-KH_d)PJ^\top \end{array}\right\} \text{when \( \mathcal{H}\) jumps} \end{array}\right. \end{equation} \]

(156.a)

with

\[ \begin{equation} K = PH_d^\top (H_d PH_d^\top+R_d)^{-1}, \end{equation} \]

(156.b)

where \( \lambda \geq 0\) and \( \gamma \in (0,1]\) are design parameters, \( R_c \in §_{>0}^{n_{y,c}}\) and \( R_d \in §_{>0}^{n_{y,d}}\) are (possibly time-varying) weighting matrices such that there exist positive scalars \( \underline{c}_{R_c}\) , \( \overline{c}_{R_c}\) , \( \underline{c}_{R_d}\) , and \( \overline{c}_{R_d}\) such that for all \( (t,j) \in \rm{dom} x\) , we have

\[ \begin{align} \underline{c}_{R_c}\rm{Id} \leq R_c(t,j) \leq \overline{c}_{R_c} \rm{Id}, \end{align} \]

(156.c)

\[ \begin{align} \underline{c}_{R_d}\rm{Id}\leq R_d(t_{j+1},j) \leq \overline{c}_{R_d}\rm{Id}. \end{align} \]

(156.d)

Observer \( \hat{\mathcal{H}}\) in (156) gathers in a common setting the continuous- and discrete-time Kalman-like observers of [82, 83] and [43, 81]. The difference compared to the continuous- and discrete-time Kalman designs [78, 79] mainly lies in the absence of the \( Q\) -covariance matrices, commonly describing the confidence in the dynamics. They are here replaced by forgetting factors \( \lambda\) and \( \gamma\) , which allows us to: i) Make the dynamics of \( P^{-1}\) linear and explicitly solvable, with a direct link to the so-called observability Gramian; and ii) Obtain a quadratic strict Lyapunov function. Note that in the jump-based part, the computation steps of the discrete-time Kalman filter [81] are gathered here into a single jump map. It combines i) Correction and ii) Prediction, instead of the contrary, since the output available to compute \( \hat{x}^+\) is its value before the jump, namely \( H_d x\) instead of \( H_d x^+\) . This justifies the presence of \( J\) in front of \( K\) in the jump-based correction term. In the classical Kalman notations, this means that our \( (\hat{x},P)\) corresponds to \( (\hat{x},P)(k|k-1)\) instead of \( (\hat{x},P)(k|k)\) , which is consistent with the use of \( P(k|k-1)\) in the Lyapunov function in [81]. Finally, note that adding the Kalman \( Q\) -parameters in (156) would preserve the decrease of the Lyapunov function but would make its lower-boundedness more intricate to prove (see [81]).

Remark 22

In many hybrid systems, especially switched systems and continuous-time systems with sporadic measurements, observability is acquired by the combination of flows with \( (F,H_c)\) and jumps with \( (J,H_d)\) . Therefore, the direct coupling of classical continuous- and discrete-time linear observers relying on the observability of each pair separately typically does not work. Here, we instead design a single unified algorithm, automatically gathering observability from both flows and jumps via a shared covariance matrix.

The goal of this section is two-fold. First, we establish conditions ensuring asymptotic convergence of (156) without imposing any further constraint on the forgetting factors \( \lambda\geq 0\) and \( \gamma \in (0,1]\) , i.e., all maximal solutions \( (x,\hat{x},P)\) to the cascade \( \mathcal{H} - \hat{\mathcal{H}}\) initialized in \( \mathcal{X}_0 \times \mathbb{R}^{n_x} \times §_{>0}^{n_x}\) with inputs \( \mathfrak{U}\) are complete and verify

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x (= \rm{dom} \hat{x})}}|x(t,j)-\hat{x}(t,j)| = 0. \end{equation} \]

(157)

In a second step, conditions for exponential stability of the estimation error with an arbitrarily fast rate as well as robustness against disturbances are derived when \( \lambda> 0\) and\( /\) or \( \gamma \in (0,1)\) . Classically, the asymptotic convergence of the Kalman(-like) observer is shown for continuous-time and discrete-time systems under the so-called UCO condition [78, 83, 43]. This condition enforces uniform and persistent invertibility of the observability Gramian, quantifying the richness of the information provided by the output on a certain time window. We extend those notions and objects in the next section in the hybrid context.

3.2.2.1 Hybrid Definitions of the Observability Gramian and Uniform Complete Observability

To define the required observability concepts for our hybrid Kalman-like observer, we begin by introducing the notions of the Gramian and observability needed for this observer. Let us assume the following; recall that \( F\) and \( J\) are allowed to be time-varying.

Assumption 10

For all solutions \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) and for all \( j \in \rm{dom}_j x\) , the map \( t \mapsto F(t,j)\) is locally bounded on \( \mathcal{T}_j\) , and the matrix \( J(t_{j+1},j)\) is invertible if \( j+1\in \rm{dom}_j x\) .

Remark 23

Assuming the invertibility of each \( J(t_{j+1},j)\) may seem restrictive in the hybrid context. But as seen in Example 15, thanks to the non-uniqueness of representation in hybrid systems as studied in Section 1.5.3.2, it is sometimes possible to rewrite \( J\) satisfying this assumption. Note though that inverting \( J\) is not necessary to implement observer (156) and is needed for analysis only, much like in the discrete-time Kalman literature [43, 81]Example 15 below is a case where the observer works without this condition and the analysis might be adaptable as suggested in [127].

Under Assumption 10, solutions \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) are unique in both forward and backward time, so we can define hybrid transition matrices of system \( \mathcal{H}\) in (155). More precisely, given a solution \( x\in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) with \( u_c = 0\) and \( u_d = 0\) , for all hybrid times \( ((t^\prime,j^\prime),(t,j))\in \rm{dom} x \times \rm{dom} x\) , we have

\[ \begin{equation} x(t,j) = \Phi_{F,J}((t,j),(t^\prime,j^\prime))x(t^\prime,j^\prime), \end{equation} \]

(158.a)

where \( \Phi_{F,J}\) is defined as

\[ \begin{equation} \Phi_{F,J}((t,j),(t^\prime,j^\prime)):=\phi_F(t,t_{j+1})\left(\prod_{k=j}^{j^\prime+1} \phi_F(t_{k+1},t_k)J(t_k,k-1)\right)\phi_F(t_{j^\prime+1},t^\prime), \end{equation} \]

(158.b)

if \( t \geq t^\prime\) and \( j \geq j^\prime\) , and, if the jump matrix \( J\) is invertible at the jump times,

\[ \begin{equation} \Phi_{F,J}((t,j),(t^\prime,j^\prime)):= \phi_F(t,t_{j})\left(\prod_{k=j+1}^{j^\prime} \phi_F(t_{k-1},t_k)(J(t_k,k-1))^{-1}\right)\phi_F(t_{j^\prime},t^\prime), \end{equation} \]

(158.c)

if \( t \leq t^\prime\) and \( j \leq j^\prime\) , with the domain of \( F\) and \( J\) inherited from \( \mathcal{D}\) , where \( \phi_F\) denotes the continuous-time transition matrix associated with \( F\) , i.e., describing solutions to \( \dot{x}=Fx\) .

Now, consider a pair of solutions \( x_a\) and \( x_b\) in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) with the same inputs \( (F,J,H_c,H_d,u_c,u_d)\) such that \( \rm{dom} x_a = \rm{dom} x_b:=\mathcal{D}\) . Then, for all hybrid times \( ((t^\prime,j^\prime),(t,j))\in \mathcal{D} \times \mathcal{D}\) , we have

\[ \begin{equation} x_a(t,j)-x_b(t,j) = \Phi_{F,J}((t,j),(t^\prime,j^\prime))(x_a(t^\prime,j^\prime)-x_b(t^\prime,j^\prime)). \end{equation} \]

(159)

By summing and integrating squares, it follows that the equality of the outputs \( (y_c,y_d)\) along \( x_a\) and \( x_b\) between time \( (t^\prime,j^\prime)\in\mathcal{D}\) and a later time \( (t,j)\in\mathcal{D}\) is equivalent to

\[ \begin{equation} (x_a(t^\prime,j^\prime)-x_b(t^\prime,j^\prime))^\top \mathcal{G}^{fw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) (x_a(t^\prime,j^\prime)-x_b(t^\prime,j^\prime)) = 0, \end{equation} \]

(160)

or, assuming the invertibility of \( J\) at the jump times,

\[ \begin{equation} (x_a(t,j)-x_b(t,j))^\top \mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) (x_a(t,j)-x_b(t,j)) = 0, \end{equation} \]

(161)

where \( \mathcal{G}^{fw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j))\) (resp., \( \mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j))\) ) is the observability Gramian (resp., backward observability Gramian) between those times as defined next.

Definition 16 (Forward and backward observability Gramians)

The forward observability Gramian and backward observability Gramian of a quadruple \( (F,J,H_c,H_d)\) defined on a hybrid time domain \( \mathcal{D}\) , between time \( (t^\prime,j^\prime) \in \mathcal{D}\) and a later time \( (t,j) \in \mathcal{D}\) , are defined respectively as

\[ \begin{multline} \mathcal{G}^{fw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) =\int_{t^\prime}^{t_{j^\prime+1}}\star^\top \Psi_c((s,j^\prime),(t^\prime,j^\prime))ds+\sum_{k=j^\prime+1}^{j-1} \int_{t_k}^{t_{k+1}} \star^\top \Psi_c((s,k),(t^\prime,j^\prime))ds \\ + \sum_{k=j^\prime}^{j-1}\star^\top \Psi_d((t_{k+1},k),(t^\prime,j^\prime)) + \int_{t_{j}}^{t}\star^\top \Psi_c((s,j),(t^\prime,j^\prime))ds, \end{multline} \]

(162.a)

and, when \( J\) is invertible at the jump times,

\[ \begin{multline} \mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) =\int_{t^\prime}^{t_{j^\prime+1}}\star^\top \Psi_c((s,j^\prime),(t,j))ds+\sum_{k=j^\prime+1}^{j-1} \int_{t_k}^{t_{k+1}} \star^\top \Psi_c((s,k),(t,j))ds \\ + \sum_{k=j^\prime}^{j-1}\star^\top \Psi_d((t_{k+1},k),(t,j)) + \int_{t_{j}}^{t}\star^\top \Psi_c((s,j),(t,j))ds, \end{multline} \]

(162.b)

where

\[ \begin{align} \Psi_c((s,k),(t,j)) &= H_c(s,k)\Phi_{F,J}((s,k),(t,j)), \\ \Psi_d((t_{k+1},k),(t,j)) &=H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t,j)), \\\end{align} \]

(163.a)

with all the jump times determined from \( \mathcal{D}\) .

According to (160), we deduce that the observability between times \( (t^\prime,j^\prime)\) and \( (t,j)\) , namely the ability to reconstruct the initial state \( x(t^\prime,j^\prime)\) from the knowledge of the future output until \( (t,j)\) , is equivalent to the positive definiteness of the observability Gramian over this period. On the other hand, when \( J\) is invertible at the jump times, the ability to reconstruct the current state \( x(t,j)\) from the knowledge of the past output until \( (t^\prime,j^\prime)\) , i.e. the backward distinguishability or constructibility, is characterized by the positive definiteness of the backward observability Gramian over this period according to (161). Actually, in that case, both notions are actually equivalent but the backward observability Gramian tends to appear more naturally in the analysis of observers.

Remark 24

The backward Gramian (162.b) characterizes the ability to reconstruct \( x(t,j)\) from the knowledge of the past output. This form naturally comes up in the analysis (see the proof of Theorem 8), but we also define a forward Gramian, characterizing the ability to reconstruct \( x(t^\prime,j^\prime)\) from the knowledge of the future output. They are equivalent under the capability to go forward and backward in time, namely Assumption 10. Note also that unlike for purely continuous-time or discrete-time linear systems, the inputs \( (u_c,u_d)\) involved in the hybrid dynamics (155) may impact the observability properties since they may change the domain of the solutions and thus the Gramian.

For observer design, the invertibility of the backward observability Gramian is typically assumed to be uniform, leading to the following hybrid Uniform Complete Observability[UCO], extending the classical UCO condition of the Kalman and Bucy’s filter [78].

Definition 17 (Uniform Complete Observability[UCO])

The quadruple \( (F,J,H_c,H_d)\) defined on a hybrid time domain \( \mathcal{D}\) is UCO with data \( (\Delta,\mu)\) if there exists \( \Delta > 0\) and \( \mu > 0\) such that for all \( ((t^\prime,j^\prime),(t,j)) \in \mathcal{D} \times \mathcal{D}\) verifying \( (t-t^\prime) + (j-j^\prime) \geq \Delta\) ,

\[ \begin{equation} \mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \geq \mu \rm{Id}. \end{equation} \]

(164)

In this section, we exploit these newly defined notions to show three main results: i) The estimation error converges asymptotically to zero for any choice of \( \lambda \geq 0\) , \( \gamma \in (0,1]\) , under boundedness of the matrices and UCO along the considered solution only (Section 3.2.2.2); ii) It is exponentially stable with an arbitrarily fast rate for appropriate choices of \( \lambda\) and \( \gamma\) if these requirements hold uniformly with respect to solutions (Section 3.2.2.3); and iii) It is robustly stable (in the sense of [46]) with respect to flow\( /\) jump input disturbances and measurement noise (Section 3.2.2.4).

3.2.2.2 Asymptotic Convergence from Uniform Complete Observability

In this part, we provide sufficient conditions for asymptotic convergence of the synchronized hybrid Kalman-like observer for a given solution \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) , thanks to some boundedness and UCO assumptions, made along that particular solution only.

Assumption 11

For each solution \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) , assume that:

  1. There exist \( c_F \geq 0\) , \( c_{H_c} \geq 0\) , \( c_{H_d} \geq 0\) , \( c_J > 0\) , and \( c_{J^{-1}} > 0\) such that for all \( (t,j) \in \rm{dom} x\) , we have (if \( j+1\in\rm{dom}_j x\) )

    \[ \begin{equation} \begin{array}{@{}r@{\;}c@{\;}l@{~}r@{\;}c@{\;}l@{~}r@{\;}c@{\;}l@{}} \|F(t,j)\| &\leq& c_F,& \|J(t_{j+1},j)\| & \leq& c_J, & \|(J(t_{j+1},j))^{-1}\|& \leq& c_{J^{-1}}, \\ &&& \|H_c(t,j)\|& \leq& c_{H_c}, & \|H_d(t_{j+1},j)\| &\leq &c_{H_d}; \end{array} \end{equation} \]

    (165)

  2. There exists a pair of positive scalars \( (\Delta,\mu)\) such that the quadruple \( (F,J,H_c,H_d)\) defined on the time domain of \( x\) is UCO with this data.

Theorem 8 below then shows that all solutions in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) can be estimated by observer (156). Note that for asymptotic convergence only (without stability guarantees), no uniformity with respect to solutions or time domains is required, but only along the time domain of each particular solution.

Theorem 8 (Asymptotic convergence under UCO)

Under Assumptions 910, and 11, for any \( \lambda \geq 0\) and any \( \gamma \in (0,1]\) , any maximal solution \( (x,\hat{x},P)\) to the cascade \( \mathcal{H}-\hat{\mathcal{H}}\) initialized in \( \mathcal{X}_0 \times \mathbb{R}^{n_x} \times §_{>0}^{n_x}\) with inputs in \( \mathfrak{U}\) and \( (R_c,R_d)\) satisfying (156.c) and (156.d) for some \( (\underline{c}_{R_c},\overline{c}_{R_c},\underline{c}_{R_d},\overline{c}_{R_d})\in \mathbb{R}^{4}_{> 0}\) is complete and verifies (157).

Proof. Consider a solution \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) . By Assumption 9, it is complete. In the rest of this proof, all variables are evolving on \( \rm{dom} x\) and so are complete. Consider \( (t,j)\mapsto \Pi(t,j)\) with \( \Pi(0,0) \in §_{>0}^{n_x}\) and dynamics

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\Pi} &=& -\lambda \Pi -\Pi F - F^\top \Pi + H_{c}^\top R_c^{-1} H_{c} \\ \Pi^+ &=& \gamma(J^{-1})^\top(\Pi+H_d^\top R_d^{-1}H_d)J^{-1}. \end{array}\right. \end{equation} \]

(166)

Because \( J\) is invertible at jumps from Assumption 10, \( \Pi\) is well defined. It can be proven using mathematical induction that the closed form of \( \Pi(t,j)\) for all \( (t,j) \in \rm{dom} x\) is

\[ \begin{multline} \Pi(t,j) = e^{-\lambda t}\gamma^j(\Phi_{F,J}((0,0),(t,j)))^\top \Pi(0,0)\Phi_{F,J}((0,0),(t,j))\\ +\sum_{k=0}^{j-1} \int_{t_k}^{t_{k+1}} e^{-\lambda(t-s)}\gamma^{j-k}\star^\top \Psi_{c^\prime}((s,k),(t,j))ds \\ +\sum_{k=0}^{j-1}e^{-\lambda(t-t_{k+1})}\gamma^{j-k}\star^\top \Psi_{d^\prime}((t_{k+1},k),(t,j))+ \int_{t_{j}}^{t}e^{-\lambda(t-s)}\star^\top \Psi_{c^\prime}((s,j),(t,j))ds, \end{multline} \]

(167.a)

where

\[ \begin{align} \Psi_{c^\prime}((s,k),(t,j)) & = R_c^{-\frac{1}{2}}(s,k) \Psi_c((s,k),(t,j)), \\ \Psi_{d^\prime}((t_{k+1},k),(t,j)) & = R_d^{-\frac{1}{2}}(t_{k+1},k)\Psi_d((t_{k+1},k),(t,j)), \\\end{align} \]

(167.b)

(with \( \Psi_c\) and \( \Psi_d\) defined in Definition 16). Now, we show that \( \Pi\) is uniformly lower-bounded along \( \rm{dom} x\) . First use Gronwall’s inequality to show that \( \|\phi_F(t,t^\prime)\| \leq e^{c_F |t-t^\prime|}\) , then it follows that for any \( ((t^\prime,j^\prime),(t,j)) \in \rm{dom} x \times \rm{dom} x\) with \( t^\prime\leq t\) and \( j^\prime\leq j\) , we have \( \|\Phi_{F,J}((t,j),(t^\prime,j^\prime))\| \leq e^{c_F(t-t^\prime)}c_J^{j-j^\prime}\) . Because

\[ \Phi_{F,J}((t,j),(t^\prime,j^\prime))\Phi_{F,J}((t^\prime,j^\prime),(t,j))=\rm{Id}, \]

this implies that

\[ \star^\top\Phi_{F,J}((t^\prime,j^\prime),(t,j)) \geq e^{-2c_F(t-t^\prime)}c_J^{-2(j-j^\prime)}\rm{Id}. \]

Then, for any \( (t,j) \in \rm{dom} x\) such that \( t+j\leq \Delta\) , we have

\[ \begin{align*} \Pi(t,j)& \geq {} e^{-\lambda t} \gamma^j e^{-2c_F t} c_J^{-2j} \Pi(0,0) \\ &\geq (e^{-\lambda} \gamma e^{-2c_F} \max\{1, c_J\}^{-2})^\Delta \Pi(0,0)\\ &\geq c_{\Pi,1}\rm{Id}, \end{align*} \]

for some \( c_{\Pi,1} > 0\) . Next, for any \( (t,j) \in \rm{dom} x\) such that \( t+j\geq \Delta\) , we can always pick \( (t^\prime,j^\prime) \in \rm{dom} x\) (before \( (t,j)\) ) such that \( \Delta \leq (t-t^\prime)+(j-j^\prime)\leq \Delta+1\) and from Assumption 11, we have

\[ \begin{align*} \Pi(t,j) & \geq e^{-\lambda(t-t^\prime)}\gamma^{j-j^\prime} \min\left\{\frac{1}{\overline{c}_{R_c}},\frac{1}{\overline{c}_{R_d}}\right\}\mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \\ &\geq (e^{-\lambda}\gamma)^{\Delta+1}\min\left\{\frac{1}{\overline{c}_{R_c}},\frac{1}{\overline{c}_{R_d}}\right\}\mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \\ &\geq\mu (e^{-\lambda}\gamma)^{\Delta+1}\min\left\{\frac{1}{\overline{c}_{R_c}},\frac{1}{\overline{c}_{R_d}}\right\} \rm{Id}\\ &:= c_{\Pi,2} \rm{Id}. \end{align*} \]

Therefore, for all \( (t,j) \in \rm{dom} x\) , we have

\[ \begin{equation} \Pi(t,j) \geq \min\{c_{\Pi,1},c_{\Pi,2}\}\rm{Id} := c_\Pi \rm{Id}, \end{equation} \]

(168)

which means that \( \Pi\) is uniformly lower-bounded and thus uniformly invertible on \( \rm{dom} x\) . Let us now study the dynamics of \( W := \Pi^{-1}\) , which is well defined and belongs to \( §_{>0}^{n_x}\) . During flows, we have

\[ \begin{align*} \dot{W} & = {}-W \dot{\Pi} W \\ & = {} -W (-\lambda \Pi -\Pi F - F^\top \Pi + H_{c}^\top R_c^{-1} H_{c}) W \\ & = {}-W (-\lambda W^{-1} -W^{-1} F - F^\top W^{-1} + H_{c}^\top R_c^{-1} H_{c}) W \\ & = {}\lambda W + FW + WF^\top - WH_{c}^\top R_c^{-1} H_{c}W. \end{align*} \]

At jumps, using Woodbury matrix identity, we have

\[ \begin{align*} W^+ & = (\Pi^+)^{-1}\\ &= (\gamma (J^{-1})^\top (\Pi + H_{d}^\top R_d^{-1}H_{d})J^{-1})^{-1}\\ &=\gamma^{-1} J (W^{-1} + H_{d}^\top R_d^{-1}H_{d})^{-1}J^\top\\ &=\gamma^{-1} J (W- WH_d^\top(H_dWH_d^\top + R_d)^{-1}H_dW)J^\top\\ &=\gamma^{-1} J (\rm{Id}- WH_d^\top(H_dWH_d^\top + R_d)^{-1}H_d)WJ^\top. \end{align*} \]

Therefore, \( W\) follows the same dynamics as \( P\) in (156). So if \( W(0,0)=P(0,0)\) then \( W(t,j) = P(t,j)\) for all \( (t,j) \in \rm{dom} x\) . This means that \( P=\Pi^{-1}\) (with \( \Pi(0,0)=(P(0,0))^{-1}\) ) and that \( P\) is invertible at all times (but not necessarily uniformly along \( \rm{dom} x\) since \( P\) may go to \( 0\) asymptotically if \( \Pi\) is not uniformly upper-bounded). Therefore, the estimation error \( \tilde{x} := x - \hat{x}\) follows the dynamics

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\tilde{x}} &=& (F-\Pi^{-1}H_c^\top R_c^{-1} H_c)\tilde{x} &:= \tilde{F} \tilde{x} \\ \tilde{x}^+ &=& J(\rm{Id}-KH_d)\tilde{x}&:= \tilde{J} \tilde{x}, \end{array}\right. \end{equation} \]

(169)

where \( K = \Pi^{-1}H_d^\top (H_d \Pi^{-1}H_d^\top+R_d)^{-1}\) . Consider the Lyapunov function \( V(\tilde{x},\Pi)=\tilde{x}^\top \Pi \tilde{x}\) . For all \( (t,j) \in \rm{dom} x\) , \( V(\tilde{x}(t,j),\Pi(t,j)) \geq c_\Pi |\tilde{x}(t,j)|^2 \) , so Theorem 8 is proven if we show that \( V\) asymptotically converges to \( 0\) . Let us study the dynamics of \( V\) along (169) and (166). During flows, we have

\[ \begin{align*} \dot{V} &={} \tilde{x}^\top[(F-\Pi^{-1}H_{c}^\top R_c^{-1} H_{c})^\top \Pi + \dot{\Pi} + \Pi(F-\Pi^{-1}H_{c}^\top R_c^{-1} H_{c})]\tilde{x} \\ &={}\tilde{x}^\top(-\lambda \Pi - H_{c}^\top R_c^{-1} H_{c})\tilde{x} \notag \\ &={}-\lambda V -\tilde{x}^\top H_c^\top R_c^{-1} H_c\tilde{x}\\ &\leq{}-\lambda V - \frac{1}{\overline{c}_{R_c}}\tilde{x}^\top H_c^\top H_c\tilde{x}. \end{align*} \]

Using Woodbury matrix identity yields

\[ \begin{align*} K &= PH_{d}^\top(R_d^{-1} - R_d^{-1}H_{d}(P^{-1}+H_{d}^\top R_d^{-1}H_{d})^{-1}H_{d}^\top R_d^{-1}) \\ & =PH_{d}^\top R_d^{-1} -PH_{d}^\top R_d^{-1}H_{d}(P^{-1}+H_{d}^\top R_d^{-1}H_{d})^{-1}H_{d}^\top R_d^{-1} \\ & =PH_{d}^\top R_d^{-1} -P((P^{-1} + H_{d}^\top R_d^{-1}H_{d})-P^{-1}) (P^{-1}+H_{d}^\top R_d^{-1}H_{d})^{-1}H_{d}^\top R_d^{-1}\\ & =(P^{-1}+H_{d}^\top R_d^{-1}H_{d})^{-1}H_{d}^\top R_d^{-1}\\ & =(\Pi+H_{d}^\top R_d^{-1}H_{d})^{-1}H_{d}^\top R_d^{-1}. \end{align*} \]

At jumps, thanks to the newly obtained expression of \( K\) and Woodbury matrix identity, we have

\[ \begin{align*} V^+ &= \gamma \tilde{x}^\top (\rm{Id}-KH_d)^\top(\Pi+H_d^\top R_d^{-1}H_d)(\rm{Id}-KH_d) \tilde{x}\\ &=\gamma \tilde{x}^\top (\rm{Id}-KH_d)^\top(\Pi+H_d^\top R_d^{-1}H_d) (\rm{Id}-(\Pi+H_d^\top R_d^{-1}H_d)^{-1}H_d^\top R_d^{-1}H_d)\tilde{x}\\ &=\gamma \tilde{x}^\top (\rm{Id}-KH_d)^\top\Pi \tilde{x}\\ &=\gamma \tilde{x}^\top (\rm{Id}-\Pi^{-1}H_d^\top (H_d PH_d^\top+R_d)^{-1}H_d)^\top\Pi \tilde{x}\\ &= \gamma V- \gamma\tilde{x}^\top H_d^\top(H_d\Pi^{-1}H_d^\top + R_d)^{-1}H_{d}\tilde{x}\\ & \leq \gamma V- \gamma\tilde{x}^\top H_d^\top\left(\frac{c_{H_d}^2}{c_\Pi} + \overline{c}_R\right)^{-1}H_{d}\tilde{x}. \end{align*} \]

We see that \( V\) decreases strictly and exponentially to \( 0\) if \( \lambda > 0\) and \( \gamma \in (0,1)\) . We next show that actually, thanks to UCO, it converges in-the-large even for \( \lambda = 0\) and \( \gamma = 1\) . In this case, we have

\[ \begin{align*} \dot{V} &\leq {}-\frac{1}{\overline{c}_{R_c}} \tilde{x}^\top H_c^\top H_c\tilde{x} &:= {}-c_c\tilde{x}^\top H_c^\top H_c\tilde{x},\\ V^+ - V &\leq{}-\frac{c_\Pi}{c_{H_d}^2 + c_\Pi \overline{c}_R}\tilde{x}^\top H_d^\top H_{d}\tilde{x}&:= {}- c_d\tilde{x}^\top H_d^\top H_{d}\tilde{x}, \end{align*} \]

and thus, for all \( ((t^\prime,j^\prime),(t,j)) \in \rm{dom} x \times \rm{dom} x\) , we have

\[ \begin{equation} V(t,j) \leq V(t^\prime,j^\prime) - \tilde{V}, \end{equation} \]

(170.a)

where

\[ \begin{equation} \tilde{V} = c_c\int_{t^\prime}^{t_{j^\prime+1}}\star^\top H_c(s,j^\prime) \tilde{x}(s,j^\prime)ds + c_c\sum_{k=j^\prime+1}^{j-1} \tilde{\mathcal{G}}_F(k) +c_d\sum_{k=j^\prime}^{j-1}\tilde{\mathcal{G}}_J(k) + c_c\int_{t_{j}}^{t}\star^\top H_c(s,j) \tilde{x}(s,j)ds, \end{equation} \]

(170.b)

with \( \tilde{\mathcal{G}}_F\) and \( \tilde{\mathcal{G}}_J\) defined as

\[ \begin{align} \tilde{\mathcal{G}}_F(k) &=\int_{t_k}^{t_{k+1}} \star^\top H_c(s,k)\tilde{x}(s,k)ds, \end{align} \]

(170.c)

\[ \begin{align} \tilde{\mathcal{G}}_J(k) &=\star^\top H_d(t_{k+1},k)\tilde{x}(t_{k+1},k). \end{align} \]

(170.d)

Applying Lemma 22 in the Appendix with \( \Delta_m=\Delta+1\) , \( K_c = \Pi^{-1}H_c^\top R_c^{-1}\) , and \( K_d=JK= J\Pi^{-1}H_d^\top (H_d \Pi^{-1}H_d^\top+R_d)^{-1}\) , which are indeed upper-bounded by \( c_{H_c}(c_\Pi \underline{c}_{R_c})^{-1}\) and \( c_J c_{H_d}(c_\Pi \underline{c}_{R_d})^{-1}\) respectively, there exists \( c_\mathcal{G}>0\) such that for all \( ((t^\prime,j^\prime),(t,j)) \in \rm{dom} x \times \rm{dom} x\) such that \( \Delta \leq (t-t^\prime) + (j-j^\prime) \leq \Delta+1\) , we have

\[ \begin{align*} \tilde{V} &\geq \min\{c_c,c_d\} c_\mathcal{G} (\tilde{x}(t,j))^\top \mathcal{G}^{bw}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j))\tilde{x}(t,j)\\ & \geq \min\{c_c,c_d\} c_\mathcal{G} \mu |\tilde{x}(t,j)|^2 , \end{align*} \]

exploiting the UCO property in Assumption 11. We finally conclude that there exists \( c_V>0\) such that for any \( ((t^\prime,j^\prime),(t,j)) \in \rm{dom} x \times \rm{dom} x\) verifying \( \Delta \leq (t-t^\prime) + (j-j^\prime) \leq \Delta+1\) , we have

\[ V(t,j)\leq V(t^\prime,j^\prime) - c_V |\tilde{x}(t,j)|^2. \]

It remains to show that \( \tilde{x}\) converges asymptotically to \( 0\) using contradiction, similar to [81]. Assume that \( \tilde{x}\) does not converge to \( 0\) . Then, there exists \( \epsilon > 0\) such that for any \( (t^\prime,j^\prime) \in \rm{dom} x\) , we can always find (exploiting the completeness of \( x\) ) \( (t,j) \in \rm{dom} x\) such that \( (t-t^\prime)+(j-j^\prime) \geq \Delta\) and \( |\tilde{x}(t,j)| \geq \epsilon\) . Hence we have \( V(t,j) \leq V(t^\prime,j^\prime) -c_V \epsilon^2\) . By repeating this process, still thanks to the completeness of \( x\) , \( V\) becomes negative after a finite amount of time, which contradicts its definition. Therefore, by contradiction, \( \tilde{x}\) converges asymptotically to \( 0\) . \( \blacksquare\)

Example 14 (Switched system)

Inspired by [134, Example 1], consider a switched system with linear maps

\[ \begin{equation} \dot{x} = A_i x, ~~ y = C_i x, \end{equation} \]

(171.a)

characterized by two modes \( i \in \{1,2\}\) as

\[ \begin{equation} A_1 = \begin{pmatrix} 0 & 0\\ 0& 0 \end{pmatrix}, ~~ C_1 = \begin{pmatrix} 1 & 0 \end{pmatrix}, ~~ A_2 = \begin{pmatrix} \epsilon & 1\\ -1& \epsilon \end{pmatrix}, ~~ C_2 = \begin{pmatrix} 0 & 0 \end{pmatrix}, \end{equation} \]

(171.b)

and triggered such that the time between two successive switches cannot be shorter than some \( \delta > 0\) . As pointed out in [134], neither \( (A_1,C_1)\) nor \( (A_2,C_2)\) is observable, but the switching order \( 1 \to 2 \to 1\) allows us to determine the initial condition unless the times elapsed in-between switches are multiples of \( \pi\) , which corresponds to a singular switching signal. Because some linear combination of the matrices of the two modes is observable, UCO could be obtained even without a dwell time, but then convergence does not have enough flow time to happen. A hybrid Kalman-like observer (156) is then designed for system (171), leading to a much simpler observer than in [134]. Asymptotic convergence of the estimation error is shown in Figure 20-Left for \( \lambda = 0\) and \( \gamma = 1\) . Convergence could be significantly accelerated by taking \( \lambda>0\) and\( /\) or \( \gamma\) in \( (0,1)\) , for which exponential stability with an arbitrarily fast rate and robustness are shown in the next sections. On the other hand, Figure 20-Right shows the observer estimate with a \( \pi\) -periodic switching signal for which UCO does not hold. The estimation error \( \tilde{x}_1\) goes to \( 0\) during mode \( 1\) because \( x_1\) is measured; in mode \( 2\) , it gets affected by the estimation error \( \tilde{x}_2\) , which cannot contract because \( x_2\) is not observable.

Figure 21
Figure 22
Figure 20. State estimation in a switched system, with \( \lambda = 0\) and \( \gamma = 1\) , under an observable switching signal (left) and a singular one (right).

3.2.2.3 Exponential Stability of the Estimation Error with an Arbitrarily Fast Rate

In this part, we show that under some extra uniformity in the boundedness and observability, the estimation error is exponentially stable with an arbitrarily fast convergence rate. For this, Assumption 11 is strengthened into the following.

Assumption 12

Assume as in Assumption 11, but all scalars therein are now the same for all solutions \( x \in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) .

Theorem 9 then shows the exponential stability of the estimation error with respect to the initial condition at any desired convergence rate.

Theorem 9 (Arbitrarily fast exponential stability after a time)

Under Assumptions 910, and 12, for any \( (\underline{c}_{R_c},\overline{c}_{R_c},\underline{c}_{R_d},\overline{c}_{R_d})\in \mathbb{R}^{4}_{> 0}\) , there exists a map \( c:\mathbb{R}_{\geq 0}\to \mathbb{R}_{\geq 0}\) such that for any \( \lambda^\prime > 0\) , the choice \( \lambda=2\lambda^\prime\) and \( \gamma= e^{-2\lambda^\prime}\) is such that any maximal solution \( (x,\hat{x},P)\) to the cascade \( \mathcal{H}-\hat{\mathcal{H}}\) initialized in \( \mathcal{X}_0 \times \mathbb{R}^{n_x} \times §_{>0}^{n_x}\) with inputs in \( \mathfrak{U}\) and \( (R_c,R_d)\) satisfying (156.c) and (156.d) with \( (\underline{c}_{R_c},\overline{c}_{R_c},\underline{c}_{R_d},\overline{c}_{R_d})\) , is complete and verifies

\[ \begin{equation} |x(t,j) - \hat{x}(t,j)| \leq c(\|\Pi(0,0)\|) |x(0,0) - \hat{x}(0,0)|e^{-\lambda^\prime(t+j-(\Delta+1))}, ~~ (t,j) \in \rm{dom} x. \end{equation} \]

(172)

Proof. First, adapting the steps leading to (168) in the proof of Theorem 8 to the particular choice of \( \lambda\) and \( \gamma\) , \( \Pi\) is uniformly lower-bounded by \( e^{-2\lambda^\prime(\Delta+1)} \underline{c}(\|\Pi(0,0)\|)\) for some \( \underline{c}:\mathbb{R}_{\geq 0}\to \mathbb{R}_{\geq 0}\) depending only on the uniform quantities in Assumption 12 and \( \overline{c}_{R_c},\overline{c}_{R_d}\) . Second, from the proof of Theorem 8, we have \( \dot{V} \leq -\lambda V\) and \( V^+\leq \gamma V\) along (169) and (166), which translates to \( V(t,j) \leq e^{-\lambda t} \gamma^j V(0,0)\leq e^{-2\lambda^\prime (t+j)}V(0,0)\) . Then (172) holds. \( \blacksquare\)

Remark 25

Note from (172) that the gain with respect to the initial error is proportional to \( e^{\lambda^\prime(\Delta+1)}\) , which increases with the choice of the rate \( \lambda^\prime\) , characterizing the peaking phenomenon typically encountered in high-gain designs. While an arbitrarily fast exponential rate is achieved in (172) at all times, arbitrarily fast convergence of the estimation error can only be achieved after \( t+j=\Delta+1\) . This is explained by the necessity of achieving observability (see the UCO condition in Definition 17). Note finally that (172) can also easily be achieved by pushing only \( \lambda\) (resp., \( \gamma\) ) under a dwell time (resp., reverse dwell time) (see [22]), by bringing stability and rate from flows to jumps and vice-versa.

Note that the asymptotic stability of the estimation error typically ensures robustness properties with respect to delays in the jump triggering of the observer, when the jump times are not perfectly known. For instance, in the autonomous context, [22, Theorem 6.4] shows the semi-global practical stability outside of the delay intervals assuming a dwell time, boundedness of solutions, and the hybrid basic conditions.

Example 15 (Spiking neuron)

The spiking behavior of a cortical neuron may be modeled with state \( \eta = (\eta_1,\eta_2) \in \mathbb{R}^2\) as

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\eta}& =& \begin{pmatrix} 0.04 \eta_1^2 + 5\eta_1 + 140 - \eta_2 + I_{\text{ext}}\\ a(b\eta_1 - \eta_2) \end{pmatrix} & \text{when \( \eta_1 \leq v_m\) } \\ \eta^+ &=& \begin{pmatrix} c\\ \eta_2 + d \end{pmatrix} & \text{when \( \eta_1 = v_m\) } \end{array} \right. \end{equation} \]

(173)

where \( \eta_1\) is the membrane potential, \( \eta_2\) is the recovery variable, and \( I_{\text{ext}}\) represents the (constant) synaptic current or injected DC current [161]. We pick here the parameters as \( I_{\text{ext}} = 10\) , \( a = 0.02\) , \( b = 0.2\) , \( c = -55\) , \( d = 4\) , and \( v_m = 30\) (all in appropriate units), thus characterizing the neuron type and its firing pattern [161]. The jump times of the solutions to system (173) are detected from the discontinuities of the output \( y_c = \eta_1\) . Since \( y_c=\eta_1\) is known during flows, we treat \( 0.04 \eta_1^2 + 140+ I_{\text{ext}}\) as a known term that can be compensated using output injection with \( u_c = (0.04 y_c^2 + 140+ I_{\text{ext}}, 0)\) . On the other hand, we assume \( d\) is unknown and seek to estimate online \( (\eta_1,\eta_2,d)\) . Note that \( d\) is not observable during flow, but it becomes observable from the combination of flows and jumps as noticed in [22]. We thus remodel system (173) into the form (155) with \( x = (x_1, x_2, x_3) = (\eta_1, \eta_2, d) \in \mathbb{R}^3\) , matrices

\[ F = \begin{pmatrix}5 & -1 & 0 \\ ab & -a & 0 \\ 0 & 0 & 0\end{pmatrix}, ~~ H_c = \begin{pmatrix}1 & 0 &0\end{pmatrix}, ~~ J = \begin{pmatrix}0 & 0 & 0 \\ 0 & 1 & 1 \\ 0 & 0 & 1\end{pmatrix}, ~~ H_d = \begin{pmatrix}0 & 0 &0\end{pmatrix}, \]

and \( u_c = (0.04 y_c^2 + I_\rm{ ext},0,0)\) , \( u_d = (c,0,0)\) known exogenous terms that can be perfectly compensated using output injection. Note that \( J\) is not invertible and does not verify Assumption 10. A possibility is to notice that because the jump map of system (173) is only active when \( \eta_1 = v_m\) , it can be rewritten as

\[ \begin{equation} \eta_1^+ = \eta_1 - v_m + c, \end{equation} \]

(174)

while preserving the same hybrid system. It would then be cast in the form (155) with \( u_d=(- v_m + c,0,0)\) and the invertible matrix \( J = \begin{pmatrix} 1 & 0 & 0 \\ 0 & 1 & 1 \\ 0 & 0 & 1 \end{pmatrix}\) , thus satisfying Assumption 10. However, for the sake of illustration, we show in Figure 23 the results of a simulation using the non-invertible formulation. This suggests that the invertibility of \( J\) might only be for theoretical analysis and it is not necessary to implement observer (156). See [127] for more details.

&lt;span data-controller=&quot;mathjax&quot;&gt;State and parameter estimation in a spiking neuron. Last figure: Comparison of the 2) -norm of the estimation error for increasing ^) (yellow&amp;mdash;nominal case, blue and red&amp;mdash;higher values of ^) with worse peaking).&lt;/span&gt;
Figure 23. State and parameter estimation in a spiking neuron. Last figure: Comparison of the \( 2\) -norm of the estimation error for increasing \( \lambda^\prime\) (yellow—nominal case, blue and red—higher values of \( \lambda^\prime\) with worse peaking).

3.2.2.4 Robustness against Disturbances and Measurement Noise

Dealing with uncertainties such as input disturbances and measurement noise is an asset of Kalman(-like) observers for a robust and practical design. Consider system (155) with flow\( /\) jump input disturbances \( v_c \in \mathbb{R}^{n_x}\) , \( v_d \in \mathbb{R}^{n_x}\) and measurement noise \( w_c \in \mathbb{R}^{n_{y,c}}\) , \( w_d \in \mathbb{R}^{n_{y,d}}\) as

\[ \begin{equation} \mathcal{H}_d\left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}c@{~~}c} \dot{x}&=&Fx + u_c +v_c& (x, u_c) \in C & y_c=H_c x + w_c\\ x^+&=&Jx + u_d + v_d& (x, u_d)\in D & y_d = H_d x + w_d \end{array} \right. \end{equation} \]

(175)

Theorem 10 shows that the estimate provided by the cascade of system \( \mathcal{H}_d\) in (175) with observer \( \hat{\mathcal{H}}\) in (156) is robustly stable with respect to the uncertainties in the sense of [46, Definition 2] (extended to hybrid systems), which is stronger than the classical Input-to-State Stability[ISS] defined in [162] by an increasing penalty of past uncertainties.

Theorem 10 (Robust stability of the estimation error)

Under Assumptions 910, and 12, there exist \( \lambda^\star>0\) and \( 0<\gamma^\star\leq 1\) such that for any \( \Pi_0\in §_{>0}^{n_x}\) , any \( (\underline{c}_{R_c},\overline{c}_{R_c},\underline{c}_{R_d},\overline{c}_{R_d})\in \mathbb{R}^{4}_{> 0}\) , any \( \lambda > \lambda^\star\) , and any \( 0 < \gamma < \gamma^\star\) , any maximal solution to the cascade \( \mathcal{H}_d-\hat{\mathcal{H}}\) initialized in \( \mathcal{X}_0 \times \mathbb{R}^{n_x} \times \{\Pi_0\}\) with inputs in \( \mathfrak{U}\) and \( (R_c,R_d)\) satisfying (156.c) and (156.d) for \( (\underline{c}_{R_c},\overline{c}_{R_c},\underline{c}_{R_d},\overline{c}_{R_d})\) is complete and robustly stable with respect to the uncertainties \( (v_c,w_c,v_d,w_d)\) .

Proof. Following the proof of (168) in Theorem 8, \( \Pi\) is uniformly lower-bounded by \( (e^{-\lambda}\gamma)^{\Delta+1} \underline{c}\) with \( \underline{c}\) depending only on the parameters of Assumption 12, \( \overline{c}_{R_c},\overline{c}_{R_d}\) and \( \Pi_0\) . On the other hand, since for any \( ((t^\prime,j^\prime),(t,j)) \in \rm{dom} x \times \rm{dom} x\) with \( t^\prime<t\) and \( j^\prime<j\) , we have \( \|\Phi_{F,J}((t^\prime,j^\prime),(t,j))\| \leq e^{c_F(t-t^\prime)}(c_{J^{-1}})^{j-j^\prime}\) , we get from (167.a) and the triangle inequality:

\[ \begin{multline*} \|\Pi(t,j)\| \leq e^{(2c_F-\lambda)t}(\gamma c_{J^{-1}}^2)^j\| \Pi_0\| +c_{H_c}^2\sum_{k=0}^{j-1} (\gamma c_{J^{-1}}^2)^{j-k}\int_{t_k}^{t_{k+1}} e^{(2c_F-\lambda)(t-s)}ds \\ +c_{H_d}^2\sum_{k=0}^{j-1}e^{(2c_F-\lambda)(t-t_k)}(\gamma c_{J^{-1}}^2)^{j-k}+ c_{H_c}^2\int_{t_{j}}^{t}e^{(2c_F-\lambda)(t-s)}ds. \end{multline*} \]

Therefore, pick \( \lambda^\star_0 > 2c_F\) and \( 0 < \gamma^\star_0 < c_{J^{-1}}^{-2}\) with \( \gamma^\star_0 \leq 1\) . Then, there exists \( \overline{c} > 0\) depending only on the uniform quantities in Assumption 12, \( \Pi_0\) , and \( \lambda^\star_0\) , \( \gamma^\star_0\) such that for any \( \lambda > \lambda^\star_0\) and for any \( 0 < \gamma < \gamma^\star_0\) , \( \Pi \leq \overline{c} \rm{Id}\) . Now, in the presence of disturbances and noise, the estimation error \( \tilde{x} := x - \hat{x}\) has the dynamics

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{x}} &=& (F-\Pi^{-1}H_c^\top R_c^{-1} H_c)\tilde{x} + v_c - \Pi^{-1}H_c^\top R_c^{-1}w_c \\ \tilde{x}^+ &=& J(\rm{Id}-\Pi^{-1}H_d^\top (H_d \Pi^{-1}H_d^\top+R_d)^{-1}H_d)\tilde{x}+v_d - J\Pi^{-1}H_d^\top (H_d \Pi^{-1}H_d^\top+R_d)^{-1}w_d. \end{array}\right. \end{equation} \]

(176)

Consider the Lyapunov function \( V(\tilde{x},\Pi)=\tilde{x}^\top \Pi \tilde{x}\) . Let us study the dynamics of \( V\) along (176) and (166). During flows, thanks to Cauchy-Schwartz and Young’s inequalities as well as the uniform upper bounds of the matrices, there exist \( \sigma_1 > 0\) and \( \sigma_2 > 0\) (independent of \( \lambda\) and \( \gamma\) ) such that we have for any \( \lambda > \lambda^\star_0\) and for any \( 0 < \gamma < \gamma^\star_0\) ,

\[ \begin{align*} \dot{V} &= [\tilde{x}^\top(F-\Pi^{-1}H_{c}^\top R_c^{-1} H_{c})^\top + v_c^\top - w_c^\top R_c^{-1} H_c \Pi^{-1}] \Pi \tilde{x} + \tilde{x}^\top(-\lambda \Pi -\Pi F - F^\top \Pi + H_{c}^\top R_c^{-1} H_{c})\tilde{x} \\ &~~{} + \tilde{x}^\top \Pi[(F-\Pi^{-1}H_{c}^\top R_c^{-1} H_{c})\tilde{x} + v_c - \Pi^{-1}H_c^\top R_c^{-1} w_c] \\ &={}-\lambda V -\tilde{x}^\top H_c^\top R_c^{-1} H_c\tilde{x}+2 \tilde{x}^\top \Pi v_c - 2 \tilde{x}^\top H_c^\top R_c^{-1}w_c\\ &\leq {}-\frac{\lambda}{3} V + \frac{\sigma_1}{\lambda}|v_c|^2 + \frac{\sigma_2}{\lambda( e^{-\lambda}\gamma)^{\Delta + 1}}|w_c|^2. \end{align*} \]

At jumps, in a similar way, there exist \( \sigma_3 > 0\) , \( \sigma_4 > 0\) , and \( \sigma_5 > 0\) (independent of \( \lambda\) and \( \gamma\) ) such that we have for any \( \lambda > \lambda^\star_0\) and for any \( 0 < \gamma < \gamma^\star_0\) ,

\[ \begin{align*} V^+ &= \gamma V- \gamma\tilde{x}^\top H_d^\top(H_d\Pi^{-1}H_d^\top + R_d)^{-1}H_{d}\tilde{x}+ 2\gamma\tilde{x}^\top \Pi J^{-1} v_d - 2\gamma\tilde{x}^\top H_d^\top(H_d \Pi^{-1} H_d^\top + R_d)^{-1} w_d \\ &~~{}+\gamma v_d^\top (J^{-1})^\top(\Pi+H_d^\top R_d^{-1}H_d)J^{-1}v_d-2\gamma v_d^\top (J^{-1})^\top H_d^\top R_d^{-1} w_d \\ &~~{}+ \gamma w_d^\top R_d^{-1}H_d(\Pi + H_d^\top R_d^{-1} H_d)^{-1} H_d^\top R_d^{-1} w_d\\ & \leq 3\gamma V + \gamma \sigma_3 |v_d|^2 + \gamma \left(\frac{\sigma_4}{(e^{-\lambda}\gamma)^{\Delta + 1}} + \sigma_5 \right) |w_d|^2. \end{align*} \]

Therefore, for any \( \lambda > \lambda^\star_0\) and any \( 0 < \gamma < \min\left\{\gamma^\star_0, \frac{1}{3}\right\}\) , we have

\[ \begin{align*} \dot{V} & \leq -\lambda_c V + \alpha_c |d_c|^2, \\ V^+ & \leq \gamma_d V + \alpha_{d} |d_d|^2, \end{align*} \]

where \( \lambda_c = \frac{\lambda}{3} > 0\) , \( \gamma_d = 3\gamma \in (0,1)\) , and

\[ \begin{align*} \alpha_c & = 2\max\left\{\frac{\sigma_1}{\lambda}, \frac{\sigma_2}{\lambda(e^{-\lambda}\gamma)^{\Delta + 1}}\right\}, & \alpha_d = 2\max\left\{\gamma \sigma_3, \gamma \left(\frac{\sigma_4}{(e^{-\lambda}\gamma)^{\Delta + 1}} + \sigma_5 \right)\right\}, \\ |d_c|^2 &= \max\{|v_c|^2,|w_c|^2\}, &|d_d|^2 = \max\{|v_d|^2,|w_d|^2\}. \end{align*} \]

This means that \( \tilde{x}\) satisfies for some \( \kappa_1 > 0\) and \( \kappa_2 > 0\) ,

\[ \begin{multline} |\tilde{x}(t,j)|^2 \leq \kappa_1 \bigg(e^{-\lambda_c t}\gamma_d^j\kappa_2|\tilde{x}(0,0)|^2 +\alpha_c\sum_{k=0}^{j-1} \int_{t_k}^{t_{k+1}}e^{-\lambda_c (t-s)}\gamma_d^{j-k} |d_c(s,k)|^2 ds \\ + \alpha_d\sum_{k=0}^{j-1} e^{-\lambda_c (t-t_{k+1})}\gamma_d^{j-k} |d_d(t_{k+1},k)|^2\bigg) + \int_{t_{j}}^{t}e^{-\lambda_c (t-s)} |d_c(s,j)|^2 ds, ~~ (t,j) \in \rm{dom} x. \end{multline} \]

(177)

Taking the square root of both sides, we obtain robust stability (with a penalty of past uncertainties) according to [46] (but extended for a hybrid system). \( \blacksquare\)

Example 16 (Continuous-time system with multi-rate outputs)

The goal of this academic example is to illustrate how the setting of continuous-time systems with multi-rate sporadic outputs can be cast in our framework and thus it admits a systematic hybrid Kalman-like observer. Consider a vehicle with position \( x_1\) , velocity \( x_2\) , and acceleration \( x_3\) . We measure \( x_3\) with a fast rate of 50 (Hz), so this can be seen as a continuous output, which however contains a lot of high-frequency noise. We then measure \( x_1\) thanks to a less noisy GPS at the rate of 1 (Hz). This makes the system observable already; however, to illustrate that our method covers systems with multi-rate sampled outputs, let us assume that we also measure \( x_2\) sporadically from every 1.5 (s) to every 2 (s). This system is written in hybrid form, with state \( x = (x_1,x_2,x_3)\) , input \( u_c\) , and two additional timers \( \tau_1\) , \( \tau_2\) as

\[ \begin{equation} \left\{ \begin{array}{@{}l@{~~}l} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{x} &=& (x_2,x_3,u_c)\\ \dot{\tau}_1 &=&-1 \\ \dot{\tau}_2 &=&-1 \end{array} \right\}& \text{when \( \left\{ \begin{array}{l}\tau_1 \in [0,1] \\ \tau_2 \in [0,2]\end{array}\right. \) }\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} x^+ &=& x\\ \tau_1^+ &=& \left\{\begin{array}{@{}l@{~~}l@{}} 1, &\text{if \( \tau_1 = 0\) }\\ \tau_1, & \text{if \( \tau_1 \neq 0\) } \end{array} \right.\\ \tau_2^+ &\in& \left\{\begin{array}{@{}l@{~~}l@{}} [1.5, 2], &\text{if \( \tau_2 = 0\) }\\ \{\tau_2\}, & \text{if \( \tau_2 \neq 0\) } \end{array} \right. \end{array} \right\} &\text{when \( \left[ \begin{array}{@{\:}l@{}} \tau_1 = 0 \\ \tau_2 = 0 \end{array} \right.\) } \end{array} \right. \end{equation} \]

(178.a)

with the outputs

\[ \begin{equation} y_c = x_3, ~~ y_d = \left\{\begin{array}{@{}l@{~~}l@{}} (x_1,0), &\text{if \( \tau_1 = 0\) and \( \tau_2 \neq 0\) }\\ (0,x_2), & \text{if \( \tau_2 = 0\) and \( \tau_1 \neq 0\) }\\ (x_1,x_2), & \text{if \( \tau_1 = \tau_2 = 0\) } \end{array} \right., \end{equation} \]

(178.b)

and initialized with \( \tau_1(0,0) = 1\) and \( \tau_2(0,0) \in [1.5, 2]\) . Then, this system fits into the form (155) and admits the Kalman-like observer (156), where the state estimate is continuously updated with the continuous-time measurement of \( x_3\) via a continuous-time Kalman filter, and it is corrected sporadically, each time a discrete-time output (\( x_1\) , or \( x_2\) , or both) shows up, by adapting the corresponding output matrix \( H_d\) depending whether only \( x_1\) , only \( x_2\) , or both are available at that time. This design works no matter the frequency of the different outputs, as long as the information gathered along the way is sufficient, namely if UCO is satisfied. Unlike in other designs in the literature, this one does not require us to discretize the dynamics, nor to run parallel filters. Figure 24-Right shows a scenario where we have a fixed sampling period of 1.5 (s) for \( x_2\) , measurement noise (with high frequency and amplitude \( \approx 3\) in \( y_c\) , low frequency and amplitude \( \approx 1\) in \( y_d\) as shown in Figure 24-Left), and the input \( u_c = 0.01\) (m\( /\) s\( ^3\) ) is an unknown bias (assumed \( 0\) in the observer, so that \( v_c = u_c\) ).

Figure 25
Figure 26
Figure 24. Left: Sensor noise used for simulation. Right: State estimation in a car with multi-rate sampled outputs (with input disturbances and noise).

We have provided a systematic hybrid Kalman-like observer for general hybrid systems with linear maps and known jump times, based on UCO. Its implementation is straightforward and applies directly to a wide class of systems including switched as well as continuous-time sampled systems with sporadic\( /\) multiple rates. Its complexity is the same as for a continuous- or discrete-time Kalman filter, with dimension \( n_x+n_x^2\) , with the same covariance matrix shared among flows and jumps.

Now, another route could be to follow our past work [163] through an observability decomposition: we could combine a continuous Kalman-like observer—estimating only the part of the state that is instantaneously observable during flows from \( y_c\) —with a discrete Kalman-like observer for the rest of the state, thus possibly reducing the observer dimension by splitting the covariance matrix. However, the possibility of decomposition is not guaranteed for time-varying systems and may not verify the necessary decoupling conditions. Future directions include properly taking into account, in the covariance matrix, errors in the jump triggering and eventually developing a Kalman(-like) observer for hybrid systems with unknown jump times.

In what follows, we attempt to analyze more precisely the observability\( /\) detectability of the system by decomposing the state according to the different sources of observability. Beyond a finer comprehension, this allows for the design of observers when UCO is not satisfied, for instance under mere detectability properties, or to design observers of smaller dimensions through decoupling (see Table 3 for a comparison).

3.2.3 Observability Decomposition

In the case where the full state is instantaneously observable during flows via the flow output \( y_c\) and the system admits an average dwell time, a high-gain flow-based observer (using only \( y_c\) ) can be designed (see [22, Section 4]); and when the full state is observable from the jump output \( y_d\) only, a jump-based observer based on an equivalent discrete-time system can be designed if the jumps are persistent (see [22, Section 5]). We are thus interested here in the case where observability rather comes from the combination of flows and jumps and\( /\) or the combination of \( (y_c,y_d)\) . The idea of the decomposition is thus to isolate state components that are instantaneously observable during flows from \( y_c\) , from other ones that become visible thanks to \( y_d\) or the combination of flows and jumps. It follows that both the flow and jump outputs may need to be fully exploited to reconstruct the state and that neither (eventually) continuous nor discrete\( /\) Zeno trajectories are allowed: both flows and jumps need to be persistent at least after a certain time, unlike in Section 3.2.2.1, as assumed next.

Assumption 13

There exists \( j_m \in \mathbb{N}\) such that solutions have flow lengths within a compact set \( \mathcal{I} \subseteq [\tau_{m},\tau_{M}]\) after jump \( j_m\) (see Definition 15), where \( \tau_m > 0\) .

Assumption 13 means that, for all solutions \( x\in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) , the hybrid arc \( (t,j)\mapsto (x(t,j),t-t_j)\) is solution after some time to the hybrid system

\[ \begin{equation} \mathcal{H}_\tau \left\{ \begin{array}{@{}l@{~~}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{x}&=&F x+u_c\\ \dot{\tau}&=&1 \end{array} \right\} (x, \tau)\in C_\tau & y_c = H_c x\\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} x^+&=&J x+u_d \\ \tau^+ &=& 0 \end{array} \right\} (x, \tau)\in D_\tau & y_d =H_d x \end{array} \right. \end{equation} \]

(179.a)

with the flow and jump sets

\[ \begin{equation} C_\tau = \mathbb{R}^{n_x}\times [0,\tau_M],~~ D_\tau = \mathbb{R}^{n_x}\times \mathcal{I}, \end{equation} \]

(179.b)

where \( \tau\in \mathbb{R}\) is a timer keeping track of the time elapsed since the previous jump. In other words, along a solution \( (t,j)\mapsto x(t,j)\) , the signal \( \tau\) is given by \( \tau(t,j) = t-t_j(x)\) . Note that \( \mathcal{H}_\tau\) admits (after the first \( j_m\) jumps) a larger set of solutions than \( §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) since the information of the flow and jump sets are replaced by the knowledge of flow lengths in \( \mathcal{I}\) only (as long as the inputs \( (F,J,H_c,H_d,u_c,u_d)\) are defined along the time domains of those extra solutions). But, as discussed in Section 3.2.1, when the observer contains gains that are computed offline, based on the knowledge of the possible flow lengths only, it is actually designed for \( \mathcal{H}_\tau\) instead of \( \mathcal{H}\) and it is thus the detectability\( /\) observability of \( \mathcal{H}_\tau\) that is relevant. In that case, the design depends only implicitly on the sets \( \mathcal{X}_0\) , \( \mathfrak{U}\) , \( C\) , and \( D\) through the choice of \( \mathcal{I}\) satisfying Assumption 13.

In view of observer design and motivated by [143], we start by proposing a change of variables decomposing the state \( x\) of \( \mathcal{H}_\tau\) into components associated with different types of observability. In order to guarantee the existence of the decomposition, we assume in the next section that the flow pair \( (F,H_c)\) is constant. However, the subsequent results of this section still hold with \( (F,H_c)\) varying, as long as the transformation into the decomposition form exists and is invertible uniformly in time as explained in Remark 26.

3.2.3.1 Observability from \( y_c\) during Flows

Assume \( (F,H_c)\) is constant. Let the (flow) observability matrix be

\[ \begin{equation} \mathcal{O} := \text{row}(H_c,H_c F, \ldots, H_cF^{n_x-1}), \end{equation} \]

(180)

and assume it is of rank \( n_{o}:=\dim \rm{Img} \mathcal{O}< {n_x}\) . Consider a basis \( (v_i)_{1 \leq i \leq n_x}\) of \( \mathbb{R}^{n_x}\) such that \( (v_i)_{1 \leq i \leq n_o}\) is a basis of the observable subspace and \( (v_i)_{n_o+1 \leq i \leq n_x}\) is a basis of the non-observable subspace \( \ker \mathcal{O}\) . Then, we define the invertible matrix \( \mathcal{D} := \begin{pmatrix} \mathcal{D}_o & \mathcal{D}_{no} \end{pmatrix}\) where

\[ \begin{align} \mathcal{D}_o &:= \begin{pmatrix} v_1 & \ldots & v_{n_o} \end{pmatrix} \in \mathbb{R}^{n_x \times n_o}, \\ \mathcal{D}_{no} &:= \begin{pmatrix} v_{n_o+1} & \ldots & v_{n_x} \end{pmatrix} \in \mathbb{R}^{n_x \times n_{no}}, \\\end{align} \]

(181.a)

which by definition satisfies for all \( \tau \geq 0\) ,

\[ \begin{equation} \mathcal{O} \mathcal{D}_{no} = 0, ~~ H_c e^{F \tau} \mathcal{D}_{no} = 0. \end{equation} \]

(182)

We denote \( \mathcal{V} := \mathcal{D}^{-1}\) which we decompose consistently into two parts \( \mathcal{V} =: \begin{pmatrix} \mathcal{V}_o \\ \mathcal{V}_{no} \end{pmatrix}\) , so that \( \mathcal{V}_o x\) represents the part of the state that is instantaneously observable during flows (see [164, Theorem 6.O6]).

A first idea could be to stop the decomposition here and design a sufficiently fast high-gain observer for \( \mathcal{V}_o x\) while estimating the rest of the state \( \mathcal{V}_{no} x\) through \( y_d\) and detectability. However, as noticed in [143, Proposition 6], the fact that \( \mathcal{V}_{o} x\) and \( \mathcal{V}_{no} x\) possibly interact with each other during flows prevents achieving stability by further pushing the high gain. The case of such a decomposition where \( \mathcal{V}_o x\) and \( \mathcal{V}_{no} x\) evolve independently during flows is exploited in a more general context in Section 3.3. Here, because the maps are linear, we can go further and solve this possible coupling by more efficiently decoupling the state components as follows.

Indeed, the estimation of any state that is not instantaneously observable during flows needs to take into account the combination of flows and jumps. That is why it is relevant to exhibit explicitly this combination via the change of coordinates

\[ \begin{equation} x \mapsto z = \begin{pmatrix}z_o\\ z_{no} \end{pmatrix} = \mathcal{V} e^{-F \tau} x = \begin{pmatrix}\mathcal{V}_o \\ \mathcal{V}_{no} \end{pmatrix}e^{-F \tau}x, \end{equation} \]

(183)

whose inverse transformation is

\[ \begin{equation} x = e^{F \tau} \mathcal{D} z = e^{F \tau} (\mathcal{D}_o z_o + \mathcal{D}_{no}z_{no}), \end{equation} \]

(184)

and which, according to (182), transforms \( \mathcal{H}_\tau\) into

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{z}_o &=& G_{o}(\tau)u_c \\ \dot{z}_{no} &=& G_{no}(\tau)u_c \\ \dot{\tau} &=& 1 \\ \\ z_o^+ &=& J_o(\tau) z_o + J_{ono}(\tau)z_{no} + \mathcal{V}_o u_d\\ z_{no}^+ &=& J_{noo}(\tau)z_o + J_{no}(\tau) z_{no} + \mathcal{V}_{no}u_d \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(185.a)

with the flow and jump sets

\[ \begin{equation} \mathbb{R}^{n_{o}}\times \mathbb{R}^{n_{no}}\times [0,\tau_M], ~~ \mathbb{R}^{n_{o}}\times \mathbb{R}^{n_{no}}\times \mathcal{I}, \end{equation} \]

(185.b)

and the measurements

\[ \begin{equation} y_c = H_{co}(\tau)z_o, ~~ y_d = H_{do}(\tau)z_o + H_{dno}(\tau)z_{no}, \end{equation} \]

(185.c)

where

\[ \begin{align} G_o(\tau)& = \mathcal{V}_o e^{-F \tau}, & G_{no}(\tau) & = \mathcal{V}_{no} e^{-F \tau}, &J_o(\tau) &= \mathcal{V}_o J e^{F \tau} \mathcal{D}_o, \\ J_{ono}(\tau) &= \mathcal{V}_o J e^{F \tau} \mathcal{D}_{no}, &J_{noo}(\tau) &= \mathcal{V}_{no} J e^{F \tau} \mathcal{D}_o, &J_{no}(\tau)& = \mathcal{V}_{no} J e^{F \tau} \mathcal{D}_{no}, \\ H_{co}(\tau) &= H_c e^{F \tau} \mathcal{D}_o, & H_{do}(\tau) &= H_d e^{F \tau} \mathcal{D}_o, & H_{dno}(\tau)& = H_d e^{F \tau} \mathcal{D}_{no}. \\\end{align} \]

(185.d)

This idea of bringing at the jumps the whole combination of flows and jumps is similar to the so-called equivalent discrete-time system exhibited in [22] for jump-based observer designs. Notice that by definition and thanks to linearity, the observability decomposition through \( \mathcal{V}\) ensures that the flow dynamics of \( z_o\) and \( y_c\) are totally independent of \( z_{no}\) , which only impacts \( z_o\) at jumps. In other words, the whole dependence of the observable part on the non-observable part via flows and jumps has been gathered at the jumps. Besides, \( z_o\) is by definition instantaneously observable from \( y_c\) . More precisely, for any \( \delta > 0\) , there exists \( \alpha > 0\) such that the observability Gramian of the continuous-time pair \( (0, H_{co}(\tau))\) satisfies

\[ \begin{equation} \int_{t}^{t + \delta} (H_{co}(s))^\top H_{co}(s) ds = \int_{t}^{t + \delta} \mathcal{D}_o^\top e^{F^\top s} H_c^\top H_c e^{F s} \mathcal{D}_o ds \geq \alpha \rm{Id}, ~~ \forall t \geq 0. \end{equation} \]

(186)

Indeed, this Gramian corresponds to the observability Gramian of the pair \( (F, H_c)\) projected onto the observable subspace. This condition is thus related to the uniform complete observability of the continuous-time pair \( (0, H_{co}(\tau))\) in the Kalman literature [78] (continuous-time version of the one in Definition 17), but here with an arbitrarily small window \( \delta\) . Since \( z_o\) is observable via \( y_c\) , we propose to estimate \( z_o\) sufficiently fast during flows to compensate for the interaction with \( z_{no}\) at jumps. Then, intuitively from system (185), information about \( z_{no}\) may be drawn from two sources: the jump output \( y_d\) and the part of \( z_{no}\) impacting \( z_o\) at jumps, namely \( J_{ono}(\tau)z_{no}\) , which may become “visible” in \( z_o\) , via \( y_c\) during the following flow interval. This is illustrated in Example 17 below. Actually, we show next in Section 3.2.3.2 that the detectability of \( z_{no}\) comes from these two sources of information only.

Example 17

Consider a hybrid system of form (155) with state \( x = (x_1,x_2,x_3,x_4)\) , \( u_c=0\) , \( u_d=0\) , and the matrices

\[ \begin{equation} F = \begin{pmatrix} 0 & -1 & 0 & 0 \\ 1 & 0 & 0 & 0 \\ 0 & 0 & 0 & -2 \\ 0 & 0 & 2 & 0 \end{pmatrix}, ~ H_c = \begin{pmatrix} 1 & 0 & 0 & 0 \end{pmatrix}, ~ J = \begin{pmatrix} 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}, ~ H_d = \begin{pmatrix} 0 & 0 & 1 & 0 \end{pmatrix}, \end{equation} \]

(187)

with random flow lengths varying in some compact set \( \mathcal{I} \subset \left(0, \frac{\pi}{2}\right)\) . It can be seen that only \( x_1\) and \( x_2\) are instantaneously observable during flows from \( y_c\) , but \( x_3\) impacts \( x_1\) at jumps (or \( y_d\) ) and \( x_4\) impacts \( x_3\) during flows. Therefore, we may hope to estimate the full state. In order to decouple those various impacts and analyze detectability more easily, we proceed with the change of variables (183). We obtain

\[ \begin{equation} z_o = \begin{pmatrix} \cos(\tau) & \sin(\tau) & 0 & 0 \\ -\sin(\tau) & \cos(\tau) &0 & 0 \end{pmatrix}x,~~ z_{no} = \begin{pmatrix} 0 & 0&\cos(2\tau) & \sin(2\tau) \\ 0 & 0& -\sin(2\tau) & \cos(2\tau) \end{pmatrix}x, \end{equation} \]

(188)

and form (185) with the matrices \( J_o(\tau) = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix}\) , \( J_{ono}(\tau) = \begin{pmatrix} \cos(2\tau) & -\sin(2\tau) \\ 0 & 0\end{pmatrix}\) , \( J_{noo}(\tau) = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix}\) , \( J_{no}(\tau) = \begin{pmatrix} \cos(2\tau) & -\sin(2\tau) \\ \sin(2\tau) & \cos(2\tau) \end{pmatrix}\) , \( H_{co}(\tau) = \begin{pmatrix} \cos(\tau) & -\sin(\tau) \end{pmatrix}\) , \( H_{do}(\tau) = \begin{pmatrix} 0 & 0 \end{pmatrix}\) , and \( H_{dno}(\tau) = \begin{pmatrix} \cos(2\tau) & -\sin(2\tau) \end{pmatrix}\) . It can be seen that in this case, the terms \( J_{ono}(\tau)z_{no}\) and \( H_{dno}(\tau)z_{no}\) contain the same information on \( z_{no}\) and both should be able to let us estimate this part.

Remark 26

In what follows, a varying pair \( (F,H_c)\) can be considered as long as the transformation into the form (185) satisfying (186) exists and is invertible uniformly in time. This can be done with the transition matrix of \( F\) replacing the exponential form, if the observable subspace remains the same at all times. In that case, the jump matrices \( J_o\) , \( J_{ono}\) , \( J_{noo}\) , \( J_{no}\) , \( H_{co}\) , \( H_{do}\) , and \( H_{dno}\) are (discrete) known inputs that are no longer functions of \( \tau\) only, but of the (discrete) jump index, which is not considered in this section. Similarly, \( J\) could vary at each jump as long as every related condition in the rest of this section holds uniformly in \( u_d\) .

3.2.3.2 Detectability Analysis for System (185)

We first provide a more specific characterization of the (pre-)asymptotic detectability of system (185) in the case of zero inputs \( (u_c,u_d)\) . Indeed, we are going to see in Theorem 12 that this detectability is relevant to characterize that of the initial system \( \mathcal{H}\) .

Lemma 10 (Zero (pre-)asymptotic detectability with known jump times)

System (185) with known jump times and zero inputs \( (u_c,u_d)\) is pre-asymptotically detectable if and only if any of its complete solutions \( (z,\tau)\) with zero inputs \( (u_c,u_d)\) and flow and jump outputs satisfying

\[ \begin{align} y_c(t,j)& = 0, & \forall t\in \rm{int}(\mathcal{T}_j(z)), \forall j \in \rm{dom}_j z, \\ y_d(t_j,j-1)& = 0, & \forall j \in \rm{dom}_j z: j \geq 1, \\\end{align} \]

(189.a)

verifies

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z}} z(t,j)= 0. \end{equation} \]

(190)

Proof. First, assume system (185) is pre-asymptotically detectable. Let \( (z,\tau)\) be a complete solution to system (185) with zero inputs \( (u_c,u_d)\) and with outputs satisfying (189). Notice that the hybrid arc \( (z^\prime,\tau)\) , with \( \rm{dom} z^\prime=\rm{dom} z\) and \( z^\prime\) constantly zero, is also solution to system (185) (thanks to linearity in the maps and the inputs \( (u_c,u_d)\) being zero). It can be seen that this solution is complete and also satisfies (189). By the pre-asymptotic detectability of system (185), we have \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z}} |z(t,j)-z^\prime(t,j)|= 0\) , which implies that \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z}} z(t,j)= 0\) . Second, let us prove the converse. For that, assume that any complete solution to system (185) with zero inputs \( (u_c,u_d)\) and with outputs verifying (189) is such that \( z\) converges to zero. Consider two complete solutions \( (z_a, \tau_a)\) and \( (z_b, \tau_b)\) to system (185) with \( \rm{dom} (z_a, \tau_a) = \rm{dom} (z_b, \tau_b)\) , with zero inputs \( (u_c,u_d)\) , and with outputs satisfying (148). By definition of the flow set, \( z_a\) and \( z_b\) jump at least once. Because the timers are both reset to zero at jumps, we have \( \tau_a(t,j) = \tau_b(t,j) := \tau(t,j)\) for all \( (t,j) \in \rm{dom} z_a\) such that \( t \geq t_1\) and \( j \in \mathbb{N}_{\geq 1}\) . By removing \( [0,t_1]\times\{0\}\) from the time domain, we see that \( (z_a - z_b, \tau)\) is a complete solution to system (185) with outputs verifying (189) (thanks to system (185) having linear maps in \( z\) and the flow and jump sets being independent of \( z\) ). Therefore, by assumption, we have \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z_a}} (z_a - z_b)(t,j)= 0\) , which implies the (pre-)asymptotic detectability of system (185). \( \blacksquare\)

Note that the equivalence of the incremental detectability as in Definition 14 with the zero detectability as in Lemma 10 is classical for linear continuous-time or discrete-time systems, but it is not automatic for hybrid systems with linear maps due to the flow\( /\) jump conditions. Here, it holds only because:

  • The flow and jump conditions in system (185) do not depend on \( z\) but only on \( \tau\) ;
  • \( \tau\) is determined uniquely after the first jump by the time domain of solutions;
  • The inputs \( (u_c,u_d)\) are removed, thus avoiding a restriction of solutions to system (185) due to a mismatch of time domains.

Theorem 11 (Detectability from a fictitious output)

Assume that \( 0 \notin \mathcal{I}\) and \( \mathcal{I}\) is compact. Then, the following three statements are equivalent:

  1. The hybrid system (179) with zero inputs \( (u_c,u_d)\) and known jump times is asymptotically detectable;
  2. The hybrid system (185) with zero inputs \( (u_c,u_d)\) and known jump times is asymptotically detectable;
  3. The discrete-time system defined as

    \[ \begin{equation} z_{no,k+1} = J_{no}(\tau_k)z_{no,k}, ~~ y_k = H_{d,\rm ext}(\tau_k)z_{no,k}, \end{equation} \]

    (191)

    where \( H_{d,\rm{ext}}(\tau_k) = \begin{pmatrix}H_{dno}(\tau_k)\\ J_{ono}(\tau_k)\end{pmatrix}\) , with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) , is asymptotically detectable.

Proof. First, notice that all maximal solutions to systems (179) and (185) are complete because their dynamics maps are linear and the flow and jump conditions do not depend on the state but only on the timer. Notice also that since \( 0\notin \mathcal{I}\) , consecutive jumps cannot happen, so that condition (148.a) is equivalent to

\[ \begin{equation} y_{a,c}(t,j) = y_{b,c}(t,j), ~ \forall (t,j) \in \rm{dom} x_a=\rm{dom} x_b, \end{equation} \]

(192)

following Remark 20. Because systems (179) and (185) are the same system modulo a uniformly invertible change of variables, Item (IX) and Item (X) are equivalent. Then, let us prove that Item (X) implies Item (XI). So assume Item (X) holds and consider a solution \( (z_{no,k})_{k\in\mathbb{N}}\) to system (191) with input \( (\tau_k)_{k\in\mathbb{N}}\) in \( \mathcal{I}\) , such that \( y_k = 0\) for all \( k \in \mathbb{N}\) . We want to show that \( (z_{no,k})_{k\in\mathbb{N}}\) asymptotically goes to zero. For that, we build and analyze the complete solution \( z = (z_o,z_{no})\) to system (185) initialized as \( z_o(0,0) = 0\) , \( z_{no}(0,0) = z_{no,0}\) , and \( \tau(0,0) = 0\) with jumps verifying \( \tau(t_j,j-1) = \tau_{j-1} \in \mathcal{I}\) for all \( j \in \mathbb{N}_{\geq 1}\) and zero inputs \( (u_c,u_d)\) . It follows from the fact that i) \( z_o\) and \( z_{no}\) are constant during flows, ii) \( y_c\) is independent of \( z_{no}\) , and iii) \( y_k = 0\) for all \( k \in \mathbb{N}\) , that for all \( j \in \mathbb{N}\) ,

\[ \begin{align*} z_o(t,j) & = 0, & \forall t \in [t_j, t_{j+1}], \\ z_{no}(t,j) & = z_{no,j}, & \forall t \in [t_j, t_{j+1}], \\ y_c(t,j) & = 0 , & \forall t \in [t_j, t_{j+1}], \\ z_o(t_{j+1},j+1) & = J_{ono}(\tau_j)z_{no,j} = 0,\\ z_{no}(t_{j+1},j+1) & = J_{no}(\tau_j)z_{no,j} = z_{no,j+1},\\ y_d(t_{j+1},j)& = H_{dno}(\tau_j)z_{no,j} = 0. \end{align*} \]

By Item (X) and Lemma 10, this implies \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z}}z_{no}(t,j) = 0\) . Because this solution \( z_{no}\) coincides with \( (z_{no,k})_{k\in\mathbb{N}}\) at the jumps, we deduce \( \lim_{\substack{k\to+\infty\\ k\in\mathbb{N}}}z_{no,k} = 0\) , Item (XI). Finally, let us prove that Item (XI) implies Item (X). Consider a complete solution \( (z,\tau) = (z_o, z_{no},\tau)\) to system (185) with zero inputs \( (u_c,u_d)\) and such that (189) holds. By definition of the jump set, for all \( j \in \rm{dom}_j z\) , \( \tau_{j}:=\tau(t_{j+1},j)\in \mathcal{I}\) . Since \( 0\notin \mathcal{I}\) , the solution admits a dwell time and, because we look for an asymptotic property, we may assume without any loss of generality that the solution starts with a flow, possibly overlooking the first part of the domain (with \( j=0\) ) in case of a jump at time \( 0\) . Since \( z_o\) is instantaneously observable during flows according to (186), \( y_c=0\) implies that \( z_o\) is zero during each flow interval. Next, as \( 0 \notin \mathcal{I}\) , there is no more than one jump at each jump time so that \( z_o(t,j) = 0\) for all \( (t,j) \in \rm{dom} z\) . Besides, since \( \mathcal{I}\) is compact, \( \rm{dom}_j z=\mathbb{N}\) and from (185.a), we then have \( J_{ono}(\tau(t_j,j-1))z_{no}(t_j,j-1) = 0\) for all \( j \in \rm{dom}_j z\) such that \( j\geq 1\) . Therefore, since \( z_{no}\) is constant during flows and \( y_d(t_j,j-1)=0\) , for all \( j \in \rm{dom}_j z\) such that \( j\geq 1\) , we have for all \( j\in \mathbb{N}\) ,

\[ \begin{align*} z_{no}(t_{j+1},j+1) = J_{no}(\tau(t_{j+1},j))z_{no}(t_{j+1},j)&= J_{no}(\tau_j) z_{no}(t_j,j), \\ H_{dno}(\tau(t_{j+1},j))z_{no}(t_{j+1},j) = H_{dno}(\tau_j)z_{no}(t_j,j)& = 0, \\ J_{ono}(\tau(t_{j+1},j))z_{no}(t_{j+1},j) = J_{ono}(\tau_j)z_{no}(t_j,j)& = 0. \end{align*} \]

By considering the sequence \( (z_{no}(t_j,j))_{j\in \mathbb{N}}\) solution to system (191) with input \( (\tau_j)_{j\in \mathbb{N}}\) in \( \mathcal{I}\) and applying Item (XI)., we obtain \( \lim_{\substack{j\to+\infty\\ j\in\mathbb{N}}}z_{no}(t_j,j) = 0\) . Since \( z_{no}\) is constant during flows, we have \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} z}} = 0\) , implying Item (X) according to Lemma 10. \( \blacksquare\)

We thus conclude that (at least when the inputs \( (u_c,u_d)\) are zero) the asymptotic detectability of system (179) requires \( z_{no}\) to be asymptotically detectable through the output made of the measured output \( y_d\) and the fictitious one \( J_{ono}(\tau)z_{no}\) , which describes how \( z_{no}\) impacts \( z_o\) at jumps. We insist also that the detectability of \( z_{no}\) comes from the combination of flows and jumps and not due to jumps alone since the useful information contained in the flow dynamics and output is gathered at the jumps via the transformation (183).

It follows from this analysis that the design of an asymptotic observer for \( \mathcal{H}\) with gains computed offline from the knowledge of \( \mathcal{I}\) only (without any special consideration of \( \mathcal{X}_0,\mathfrak{U}, C\) , or \( D\) ), namely an observer for system (179), without considering the possible restrictions of time domains by the inputs \( (u_c,u_d)\) , requires the asymptotic detectability of the discrete-time system (191). This is because any time domain with flow lengths in \( \mathcal{I}\) is a priori possible. This is done in Section 3.2.3.3 with an LMI-based design. On the other hand, if we consider the more precise problem of observer design with time domains restricted to that of solutions in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) , we end up with the following sufficient condition for asymptotic detectability of \( \mathcal{H}\) .

Theorem 12 (Detectability from an equivalent discrete-time system)

Suppose Assumptions 9 and 13 hold. Then, \( \mathcal{H}\) initialized in \( \mathcal{X}_0\) with inputs in \( \mathfrak{U}\) is asymptotically detectable if for each \( x\in §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) , the discrete-time system (191), with input \( (\tau_k)_{k\in \mathbb{N}}\) defined as \( \tau_k=t_{k+1}(x)-t_k(x)\) for all \( k \in \mathbb{N}\) , is asymptotically detectable.

Proof. Pick solutions \( x_a\) and \( x_b\) in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) with the same inputs \( (F,J,H_c,H_d,u_c,u_d)\) , such that \( \rm{dom} x_a = \rm{dom} x_b\) , and with outputs \( y_{a,c}\) , \( y_{b,c}\) , \( y_{a,d}\) , \( y_{b,d}\) satisfying (148). By Assumptions 9 and 13, these solutions are complete with \( \rm{dom}_j x_a = \rm{dom}_j x_b=\mathbb{N}\) . By Assumption 13, for all \( j \in \mathbb{N}_{\geq j_m}\) , \( \tau_{j}:=t_{j+1}(x_a) - t_j(x_a) = t_{j+1}(x_b) - t_j(x_b)\in \mathcal{I}\) . Since \( 0\notin \mathcal{I}\) , the solutions admit a dwell time after the first \( j_m\) jumps. Since we look for an asymptotic property, we may discard the first part of the solutions with \( j<j_m\) , and assume without any loss of generality that they start with a flow interval and have a dwell time. Then, condition (148.a) is equivalent to (192) following Remark 20. Consider the hybrid signals \( (z_a,z_b,\tau)\) defined for all \( (t,j) \in \rm{dom} x_a\) as

\[ \begin{align*} \tau(t,j) &= t-t_j, \\ \begin{pmatrix}z_{a,o}(t,j)\\ z_{a,no} (t,j)\end{pmatrix} &= \begin{pmatrix}\mathcal{V}_o \\ \mathcal{V}_{no} \end{pmatrix} e^{-F \tau(t,j)}x_a(t,j), \\ \begin{pmatrix}z_{b,o}(t,j)\\ z_{b,no}(t,j) \end{pmatrix} &= \begin{pmatrix}\mathcal{V}_o \\ \mathcal{V}_{no} \end{pmatrix}e^{-F \tau(t,j) }x_b(t,j) . \end{align*} \]

We see that both \( (z_a,\tau)\) and \( (z_b,\tau)\) are solutions to system (185) with the same \( \tau\) -dependent matrices \( G_o\) , \( G_{no}\) , \( J_o\) , \( J_{ono}\) , \( J_{noo}\) , \( J_{no}\) , \( H_{co}\) , \( H_{do}\) , \( H_{dno}\) and the same inputs \( (u_c,u_d)\) . Since \( z_{a,o}\) and \( z_{b,o}\) are instantaneously observable during flows according to (186), \( y_{a,c}=y_{b,c}\) implies that \( z_{a,o} = z_{b,o}\) during each flow interval. Next, as \( 0 \notin \mathcal{I}\) , there is no more than one jump at each jump time so that \( z_{a,o}(t,j) = z_{b,o}(t,j)\) for all \( (t,j) \in \rm{dom} x_a\) . Besides, since \( u_d\) is the same for both solutions, from (185.a), we have \( J_{ono}(\tau(t_j,j-1))z_{a,no}(t_j,j-1) = J_{ono}(\tau(t_j,j-1))z_{b,no}(t_j,j-1)\) for all \( j \in \mathbb{N}_{\geq 1}\) . Therefore, since \( z_{a,no}\) and \( z_{b,no}\) evolve in the same way during flows (with the same \( u_c\) ) and \( y_{a,d}(t_j,j-1)=y_{b,d}(t_j,j-1)\) , for all \( j \in \mathbb{N}_{\geq 1}\) , by defining \( \tilde{z}_{no} = z_{a,no} - z_{b,no}\) , we have that \( \tilde{z}_{no}\) is constant during flows and for all \( j\in \mathbb{N}\) ,

\[ \begin{align*} \tilde{z}_{no}(t_{j+1},j+1) = J_{no}(\tau(t_{j+1},j))\tilde{z}_{no}(t_{j+1},j)&= J_{no}(\tau_j) \tilde{z}_{no}(t_j,j), \\ H_{dno}(\tau(t_{j+1},j))\tilde{z}_{no}(t_{j+1},j) = H_{dno}(\tau_j)\tilde{z}_{no}(t_j,j)& = 0, \\ J_{ono}(\tau(t_{j+1},j))\tilde{z}_{no}(t_{j+1},j) = J_{ono}(\tau_j)\tilde{z}_{no}(t_j,j)& = 0. \end{align*} \]

By considering the sequence \( (\tilde{z}_{no}(t_j,j))_{j\in \mathbb{N}}\) solution to system (191) with input \( (\tau_j)_{j\in \mathbb{N}}\) in \( \mathcal{I}\) and using the asymptotic detectability of system (191) for the particular sequence \( (\tau_k)_{k \in \mathbb{N}}\) generated by \( x_a\) and \( x_b\) , we obtain \( \lim_{\substack{j\to+\infty\\ j\in\mathbb{N}}}\tilde{z}_{no}(t_j,j) = 0\) . Since \( \tilde{z}_{no}\) is constant during flows, we have \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x_a}}\tilde{z}_{no}(t,j)=0\) , which means that \( \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x_a}}(z_{a,no}(t,j)-z_{b,no}(t,j)) = 0\) , implying that \( \mathcal{H}\) initialized in \( \mathcal{X}_0\) with inputs in \( \mathfrak{U}\) is asymptotically detectable according to Lemma 10. \( \blacksquare\)

Unlike Theorem 11Theorem 12 does not give a necessary condition for detectability (and thus observer design). The reason is that the flow and jump conditions of \( \mathcal{H}\) are not taken into account in system (191). But it still suggests us to build observers for \( \mathcal{H}\) under observability\( /\) detectability conditions on system (191) for the flow length sequences \( (\tau_k)_{k\in \mathbb{N}}\) appearing in \( §_{\mathcal{H}}(\mathcal{X}_0,\mathfrak{U})\) . This is done in Sections 3.2.3.3 and 3.2.3.4 through an LMI-based and a KKL-based design.

Example 18

Consider the system in Example 17. It is possible to check, by computing the (time-varying) observability matrix of the pair \( (J_{no}(\tau_k),H_{dno}(\tau_k))\) , that system (191) is observable for any sequence \( (\tau_k)_{k\in\mathbb{N}}\) as long as \( \sin(\tau_k+\tau_{k+1})\neq 0\) at some \( k\in \mathbb{N}\) , which is the case for \( \tau_k\in \mathcal{I}\) since \( \mathcal{I}\subset \left(0,\frac{\pi}{2}\right)\) . This implies in particular that system (191) is asymptotically detectable. Actually, even if \( H_d = 0\) , i.e., no output is available at jumps, the pair \( (J_{no}(\tau_k), J_{ono}(\tau_k))\) is also observable using the same arguments. This means that \( z_{no}\) is actually observable through the fictitious measurement of \( z_o\) at jumps. We see from this example that thanks to the flow-jump coupling, by using \( z_o\) as a fictitious measurement, state components that are not observable during flows from the flow output may become observable via jumps even without any additional measurements at jumps (hidden dynamics), only through the way they impact the observable ones.

3.2.3.3 LMI-Based Observer Design Using Observability Decomposition

Inspired by the detectability analysis of Theorem 11, we propose a first observer design under the following detectability assumption.

Assumption 14

Given \( \tau_M\) and \( \mathcal{I}\) defined in Assumption 13, there exist \( Q_{no} \in §_{> 0}^{n_{no}}\) , \( L_{dno}:[0,\tau_M]\to \mathbb{R}^{n_{no} \times n_{y,d}}\) bounded on \( [0,\tau_M]\) and continuous on \( \mathcal{I}\) , and \( K_{no} \in \mathbb{R}^{n_{no} \times n_{o}}\) such that

\[ \begin{equation} \star^\top Q_{no} \left(J_{no}(\tau) - \begin{pmatrix} L_{dno}(\tau) & K_{no} \end{pmatrix}\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}\right)- Q_{no} < 0, ~~ \forall \tau \in \mathcal{I}. \end{equation} \]

(193)

We refer the reader to Remark 27 for constructive methods to solve (193), where it is shown that the solvability of (193) in \( Q_{no}\) and \( \begin{pmatrix} L_{dno}(\tau) & K_{no} \end{pmatrix}\) is equivalent to that of a reduced LMI involving \( Q_{no}\) only. Consistently with Theorem 11Assumption 14 requires the detectability of the pair \( \left(J_{no}(\tau),\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}\right)\) for each frozen \( \tau \in \mathcal{I}\) . But it is actually stronger because it further requires \( Q_{no}\) and \( K_{no}\) to be independent of \( \tau\) . It corresponds to a stronger version of the quadratic detectability of system (191) defined in [44]. Actually, the detectability of Assumption 14 allows us to build an observer for any sequence of flow lengths \( (\tau_k)_{k \in \mathbb{N}} \in \mathcal{I}\) and thus requires the detectability of the discrete-time pair for any such sequences, which is still consistent with the result of Theorem 11. Note that the reason why \( K_{no}\) is required to be independent of \( \tau\) is that it is used to carry out another change of variables in the proof of Theorem 13 below, allowing us to exhibit the fictitious output in the analysis.

3.2.3.3.1 LMI-Based Observer Design in the \( z\) -Coordinates

Because the flow output matrix \( H_{co}(\cdot)\) varies and satisfies the observability condition (186), we design a flow-based Kalman-like observer of \( z_o\) during flows using \( y_c\)  [27]. Its advantage over a Kalman observer is that it admits a strict Lyapunov function, allowing for direct robust Lyapunov analysis. Besides, it provides a direct relationship between the Lyapunov matrix and the observability Gramian. Then, as suggested by the detectability analysis, \( z_{no}\) should be estimated thanks to both \( y_d\) and its interaction with \( z_o\) at jumps via a fictitious output. The latter is not available for injection in the observer, but it becomes visible through \( z_o\) after the jump, and thus through \( y_c\) during flows. This justifies correcting the estimate of \( z_{no}\) during flows with \( y_c\) , via the gain \( K_{no}\) . The dynamics of the observer are then given by

\[ \begin{equation} \left\{\begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{z}}_o &=& G_o(\tau)u_c + P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}(y_c - H_{co}(\tau)\hat{z}_o) \\ \dot{\hat{z}}_{no} & = & G_{no}(\tau)u_c + K_{no}P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}(y_c-H_{co}(\tau)\hat{z}_o)\\ \dot{P} &= &{}-\lambda P + (H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau) \\ \dot{\tau} &=& 1\end{array}\right\} \text{when}~(155) \text{ flows} \\ \\ \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{z}_o^+ &=& J_o(\tau) \hat{z}_o + J_{ono}(\tau) \hat{z}_{no} +\mathcal{V}_o u_d \\ \hat{z}_{no}^+ &=& J_{noo}(\tau) \hat{z}_o+J_{no}(\tau)\hat{z}_{no} +\mathcal{V}_{no}u_d\\ &&{}+L_{dno}(\tau)(y_d-H_{do}(\tau)\hat{z}_{o}-H_{dno}(\tau)\hat{z}_{no}) \\ P^+ &=& P_0\\ \tau^+ &=& 0 \end{array}\right\} \text{when}~(155) \text{ jumps} \end{array} \right. \end{equation} \]

(194)

with \( P_0\in §_{> 0}^{n_o}\) , \( K_{no}\) and \( L_{dno}\) given by Assumption 14, and where \( \tau \mapsto R(\tau) \in §_{>0}^{n_{y,c}}\) is a positive definite weighting matrix that is defined and is continuous on \( [0, \tau_M]\) to be chosen arbitrarily. The estimate is then recovered by using (184) on \( \hat{z}\) with the Global Exponential Stability[GES] of the estimation error as stated next.

Theorem 13 (LMI-based observer design for system (155) in the \( z\) -coordinates)

Under Assumptions 913, and 14, given any \( P_0\in §_{> 0}^{n_o}\) , there exists \( \lambda^\star>0\) such that for any \( \lambda > \lambda^\star\) , there exist \( \rho^\prime>0\) and \( \lambda^\prime>0\) such that for any solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and any solution \( (\hat{z},P,\tau)\) to observer (194) with \( P(0,0)=P_0\) , and \( \tau(0,0)=0\) , \( (\hat{z},P,\tau)\) is complete and we have

\[ \begin{equation} |x(t, j) - \hat{x}(t, j)| \leq \rho^\prime |x(0,0) - \hat{x}(0,0)| e^{-\lambda^\prime(t+j)}, ~~ \forall (t, j)\in \rm{dom} x, \end{equation} \]

(195)

with \( \hat{x}\) obtained by \( \hat{x} = e^{F \tau} \mathcal{D}\hat{z}\) with \( \mathcal{D}\) defined in (184).

Proof. Consider a solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and a solution \( (\hat{z},P,\tau)\) to observer (194) with \( P(0,0)=P_0\) and \( \tau(0,0)=0\) . By Assumption 9, it is complete and so is \( (\hat{z},P,\tau)\) . Following (183), define

\[ \begin{equation} z(t,j) := \mathcal{V} e^{-F \tau(t,j)} x(t,j), ~~ \forall (t,j)\in \rm{dom} x, \end{equation} \]

(196)

and consider the estimation error \( \tilde{z} = (\tilde{z}_o, \tilde{z}_{no}) = (z_o - \hat{z}_o, z_{no} - \hat{z}_{no})\) . Because the flow lengths of \( x\) are in \( \mathcal{I}\) by Assumption 13 after the first \( j_m\) jumps only, the proof consists of two parts: first, we use Lemma 23 to show the exponential convergence of \( \tilde{z}\) starting at hybrid time \( (t_{j_m},j_m)\) by putting the estimation error dynamics into the appropriate form, and then, we analyze the behavior of the estimation error before \( (t_{j_m},j_m)\) using Lemma 24. Consider first the solution \( (\tilde{z},P,\tau)\) starting from \( (t_{j_m},j_m)\) . According to Assumption 13 and since the observer’s jumps are synchronized with those of \( \mathcal{H}\) , \( (\tilde{z},P,\tau)\) is solution to

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{z}}_o &=&{} -P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}H_{co}(\tau)\tilde{z}_o \\ \dot{\tilde{z}}_{no} & = &{} -K_{no}P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}H_{co}(\tau)\tilde{z}_o\\ \dot{P} &= &{}-\lambda P + (H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau) \\ \dot{\tau} &=& 1\\ \\ \tilde{z}_o^+ &=& J_o(\tau) \tilde{z}_o + J_{ono}(\tau) \tilde{z}_{no} \\ \tilde{z}_{no}^+ &=& (J_{noo}(\tau)-L_{dno}(\tau)H_{do}(\tau)) \tilde{z}_o+(J_{no}(\tau)-L_{dno}(\tau)H_{dno}(\tau))\tilde{z}_{no}\\ P^+ &=& P_0 \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(197.a)

with the flow and jump sets

\[ \begin{equation} \mathbb{R}^{n_o}\times \mathbb{R}^{n_{no}} \times\mathbb{R}^{n_o \times n_o}\times [0,\tau_M], ~~ \mathbb{R}^{n_o}\times \mathbb{R}^{n_{no}} \times\mathbb{R}^{n_o \times n_o}\times \mathcal{I}. \end{equation} \]

(197.b)

We next perform the change of variables

\[ \begin{equation} \tilde{\eta} = \tilde{z}_{no} - K_{no}\tilde{z}_o, \end{equation} \]

(198)

which transforms the estimation error system (197) into

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{z}}_o &=& {}-P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau)\tilde{z}_o \\ \dot{\tilde{\eta}} & = & 0 \\ \dot{P} &= &{}-\lambda P + (H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau) \\ \dot{\tau} &=& 1 \\ \\ \tilde{z}_o^+ &=& \overline{J}_o(\tau) \tilde{z}_o + J_{ono}(\tau) \tilde{\eta} \\ \tilde{\eta}^+ & = & (\overline{J}_{noo}(\tau)-L_{dno}(\tau)\overline{H}_{do}(\tau)) \tilde{z}_o+J_{\eta}(\tau)\tilde{\eta}\\ P^+&=&0 \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(199)

with the same flow and jump sets where \( \overline{J}_o(\tau) = J_o(\tau) + J_{ono}(\tau)K_{no}\) , \( \overline{J}_{noo}(\tau) = J_{noo}(\tau) + {J}_{no}(\tau)K_{no} - K_{no}J_o(\tau) - K_{no}J_{ono}(\tau)K_{no}\) , \( \overline{H}_{do}(\tau) = H_{do}(\tau) + H_{dno}(\tau)K_{no}\) , and \( J_\eta(\tau) = J_{no}(\tau) -L_{dno}(\tau)H_{dno}(\tau) - K_{no}J_{ono}(\tau)\) , with the flow set \( \mathbb{R}^{n_{o}}\times \mathbb{R}^{n_{no}}\times [0,\tau_M]\) and the jump set \( \mathbb{R}^{n_{o}}\times \mathbb{R}^{n_{no}}\times \mathcal{I}\) . From Assumption 14, \( J_\eta(\tau)\) is Schur for all \( \tau \in \mathcal{I}\) , and more precisely, there exists \( Q_\eta\in§_{>0}^{n_{no}}\) such that

\[ \begin{equation} (J_\eta(\tau))^\top Q_\eta J_\eta(\tau) - Q_\eta < 0, ~~ \forall \tau \in \mathcal{I}. \end{equation} \]

(200)

Using Lemma 23, we proceed to prove the GES of the estimation error \( (\tilde{z}_o,\tilde{\eta})\) with respect to the value \( (\tilde{z}_o,\tilde{\eta})(t_{j_m}, j_m)\) . Then, the GES with respect to \( (\tilde{z}_o,\tilde{\eta})(0,0)\) is proven using Lemma 24. Last, because the transformations (183) and (198) are linear with \( \tau \mapsto e^{F\tau}\) bounded with a strictly positive lower bound on the compact set \( [0, \tau_M]\) , we obtain (195) observing that \( (\tilde{z}_o, \tilde{\eta}) = \begin{pmatrix}\mathcal{V}_o \\ \mathcal{V}_{no} - K_{no}\mathcal{V}_o \end{pmatrix}e^{-F \tau}(x-\hat{x})\) , concluding the proof. \( \blacksquare\)

Remark 27

Applying Schur’s lemma and then the elimination lemma [165] to (193), we see that \( Q_{no} \in §_{>0}^{n_{no}}\) exists only if there exists a solution to the LMI

\[ \begin{equation} \begin{pmatrix}\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}^{\bot\top} Q_{no}\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}^\bot & \star \\ Q_{no}J_{no}(\tau)\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}^\bot & Q_{no} \end{pmatrix}> 0, ~~ \forall \tau \in \mathcal{I}. \end{equation} \]

(201)

If such a \( Q_{no}\) is obtained, the gains \( L_{dno}(\cdot)\) and \( K_{no}\) are then found by using (193) with \( Q_{no}\) known. If \( \mathcal{I}\) is has infinitely many points, then there is an infinite number of LMIs to solve. Actually, it is worth noting that the exponential term \( e^{F \tau}\) contained in all the \( \tau\) -dependent matrices in (201) can be expanded using residue matrices [151], as

\[ \begin{equation} e^{F \tau} = \sum_{i=1}^{\sigma_r} \sum_{j=1}^{m_i^r} R_{ij} e^{\lambda_i \tau} \frac{\tau^{j-1}}{(j-1)!} + \sum_{i=1}^{\sigma_c} \sum_{j=1}^{m_i^c} 2e^{\Re(\lambda_i) \tau} (\Re(R_{ij}) \cos(\Im(\lambda_i) \tau) - \Im(R_{ij}) \sin(\Im(\lambda_i) \tau) ) \frac{\tau^{j-1}}{(j-1)!}, \end{equation} \]

(202)

where \( \sigma_r\) and \( \sigma_c\) are the numbers of distinct real eigenvalues and complex conjugate eigenvalue pairs; \( m_i^r\) and \( m_i^c\) are the multiplicity of the real eigenvalue \( \lambda_i\) and of the complex conjugate eigenvalue pair \( \lambda_i, \lambda_i^*\) in the minimal polynomial of \( F\) ; \( R_{ij} \in \mathbb{R}^{n_x \times n_x}\) are matrices corresponding to the residues associated to the partial fraction expansion of \( (s\rm{Id} - F)^{-1}\) . This in turn allows \( e^{F \tau}\) to be written as a finite sum of matrices affine in \( N\) scalar functions \( \beta_{ij} = e^{\lambda_i \tau} \tau^{j-1}\) , \( \gamma_{ij} = e^{\Re(\lambda_i) \tau} \cos(\Im(\lambda_i) \tau) \tau^{j-1}\) , and \( \gamma^*_{ij} = e^{\Re(\lambda_i) \tau} \sin(\Im(\lambda_i) \tau) \tau^{j-1}\) . It then implies that (201) can be solved in a polytopic approach, i.e., the LMIs are satisfied for all \( \tau \in \mathcal{I}\) compact if they are satisfied at the finite number \( 2^N\) of vertices of the polytope formed by these scalar functions when \( \tau\) varies in \( \mathcal{I}\) . Alternatively, the LMIs can be solved in a grid-based approach followed by post-analysis of the solution’s stability as in [44], possibly with a theoretical proof extended from [152].

3.2.3.3.2 LMI-Based Observer Design in the \( x\) -Coordinates

In this section, we show that the observer can equivalently be implemented directly in the original \( x\) -coordinates, with dynamics given by

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{x}}&=& F \hat{x}+u_c+¶ H^\top_c (R(\tau))^{-1}(y_c-H_c\hat{x})\\ \dot{¶} &=& \lambda ¶ + F¶ +¶ F^\top - ¶ H_c^\top (R(\tau))^{-1} H_c ¶\\ \dot{\tau} &=& 1\end{array}\right\} \text{when}~(155) \text{ flows} \\ \\ \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{x}^+ &=& J\hat{x}+u_d+\mathcal{D}_{no}L_{dno}(\tau) (y_d-H_d\hat{x}) \\ ¶^+ &=& (\mathcal{D}_o+\mathcal{D}_{no}K_{no}) P_0^{-1} (\mathcal{D}_o+\mathcal{D}_{no}K_{no})^\top\\ \tau^+ &=& 0 \end{array}\right\} \text{when}~(155) \text{ jumps} \end{array} \right. \end{equation} \]

(203)

The GES of the estimation error is proven in Theorem 14.

Theorem 14 (LMI-based observer design for system (155) in the \( x\) -coordinates)

Under Assumptions 913, and 14, given any \( P_0\in §_{> 0}^{n_o}\) , there exists \( \lambda^\star>0\) such that for any \( \lambda > \lambda^\star\) , there exist \( \rho^\prime>0\) and \( \lambda^\prime>0\) such that for any solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and for any solution \( (\hat{x},¶,\tau)\) to observer (203) with \( ¶(0,0) = (\mathcal{D}_o+\mathcal{D}_{no}K_{no}) P_0^{-1} (\mathcal{D}_o+\mathcal{D}_{no}K_{no})^\top\) , and \( \tau(0,0)=0\) , \( (\hat{x},¶,\tau)\) is complete and we have

\[ \begin{equation} |x(t,j)-\hat{x}(t, j)| \leq \rho^\prime |x(0,0)-\hat{x}(0, 0)| e^{-\lambda^\prime(t+j)}, ~~ \forall (t, j)\in \rm{dom} x. \end{equation} \]

(204)

Proof. Pick a solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and a solution \( (\hat{\zeta},¶,\tau)\) to observer (203) with \( ¶(0,0)=(\mathcal{D}_o+\mathcal{D}_{no}K_{no}) P_0^{-1} (\mathcal{D}_o+\mathcal{D}_{no}K_{no})^\top\) and \( \tau(0,0)=0\) . Consider \( (\hat{z},P,\underline{\tau})\) solution to observer (194), with \( \hat{z}(0,0)=\mathcal{V} \hat{x}(0,0)\) , \( P(0,0)=P_0\) , and \( \underline{\tau}(0,0)=0\) . First, notice from its dynamics that \( \underline{\tau}=\tau\) . Then, applying Theorem 13, we get

\[ \begin{equation} |x(t, j) - \underline{\hat{x}}(t, j)| \leq \rho^\prime |x(0,0) - \underline{\hat{x}}(0,0)| e^{-\lambda^\prime(t+j)}, ~~ \forall (t, j)\in \rm{dom} x, \end{equation} \]

(205)

where \( \underline{\hat{x}}=e^{F \tau} \mathcal{D}\hat{z}\) . The proof consists in showing that \( \underline{\hat{x}}=\hat{x}\) , thus obtaining (204). Observe that

\[ \begin{equation} \dot{\underline{\hat{x}}}= F \underline{\hat{x}}+u_c+L_c(P,\tau) (y_c-H_c\underline{\hat{x}}), \end{equation} \]

(206.a)

during flows and

\[ \begin{equation} \underline{\hat{x}}^+ = J\underline{\hat{x}}+u_d+L_d(\tau) (y_d-H_d\underline{\hat{x}}), \end{equation} \]

(206.b)

at jumps where

\[ \begin{align} L_c(P,\tau) &= e^{F \tau} \mathcal{D} \begin{pmatrix} P^{-1} (H_{co}(\tau))^\top (R(\tau))^{-1}\notag\\ K_{no} P^{-1} (H_{co}(\tau))^\top (R(\tau))^{-1} \end{pmatrix}\\ & = e^{F \tau} (\mathcal{D}_o+ \mathcal{D}_{no}K_{no}) P^{-1} \mathcal{D}_o^\top e^{F^\top \tau} H_c^\top (R(\tau))^{-1}, \\\end{align} \]

(207.a)

and

\[ \begin{equation} L_d(\tau) = \mathcal{D} \begin{pmatrix} 0 \\ L_{dno}(\tau) \end{pmatrix} = \mathcal{D}_{no}L_{dno}(\tau). \end{equation} \]

(207.b)

From (182), \( L_c(P,\tau)\) can be rewritten as \( L_c(P,\tau)=\underline{¶} H_c^\top (R(\tau))^{-1}\) where

\[ \begin{equation} \underline{¶} = e^{F \tau}(\mathcal{D}_o+\mathcal{D}_{no}K_{no}) P^{-1} (\mathcal{D}_o+\mathcal{D}_{no}K_{no})^\top e^{F^\top \tau}. \end{equation} \]

(208)

Calculating \( \dot{\underline{¶}}\) while noting that the time derivative of \( P^{-1}\) is \( -P^{-1} \dot{P} P^{-1}\) , we obtain the same flow\( /\) jump dynamics as \( ¶\) in observer (203). Besides, \( \underline{¶}(0,0)=¶(0,0)\) , so \( \underline{¶}=¶\) thanks to the uniqueness of solutions. We deduce that \( \underline{\hat{x}}\) follows the same dynamics as \( \hat{x}\) , and since

\[ \underline{\hat{x}}(0,0)=\mathcal{D}^\prime\hat{z}(0,0) = \hat{x}(0,0), \]

we have \( \underline{\hat{x}}=\hat{x}\) , which concludes the proof. \( \blacksquare\)

It is interesting to see that the observability provided at jumps by the fictitious output in the non-observable subspace \( \mathcal{D}_{no}\) is stored into \( ¶\) at jumps. This allows the use of \( y_c\) to correct the estimate in the non-observable subspace during flows, while the Riccati dynamics of \( ¶\) instead excites only the observable directions provided by (186). In terms of implementation, observer (203) is of a larger dimension than that in observer (194) but it allows us to avoid the online inversion of the change of variables. Actually, observer (203) has the same dimension and the same flow dynamics as the Kalman-like observer (156) proposed in Section 3.2.2. The difference lies in i) the jump dynamics, which here contains a priori gains \( K_{no}\) and \( L_{dno}\) instead of dynamic gains computed online via \( ¶\) , and ii) the quadratic detectability assumption (14) which replaces the UCO in Section 3.2.2.

Remark 28

Denote \( L_d(\tau) = \mathcal{D}_{no}L_{dno}(\tau)\) and observe that \( L_{dno}(\tau) = \mathcal{V}_{no} L_d(\tau)\) . Similarly, denote \( K_{no}^\prime = \mathcal{D}_{no}K_{no}\) and observe that \( K_{no} = \mathcal{V}_{no} K_{no}^\prime\) . The conditions in Assumption 14 for the design of \( L_{dno}(\cdot)\) and \( K_{no}\) are equivalent to solving for \( L_{d}(\cdot)\) and \( K_{no}^\prime\) directly in the \( x\) -coordinates and for all \( \tau \in \mathcal{I}\) ,

\[ \begin{align} \star^\top Q_{no} \mathcal{V}_{no} (J-L_{d}(\tau)H_{d}-K_{no}^\prime\mathcal{V}_o J)e^{F\tau}\mathcal{D}_{no} - Q_{no} <0, \\ \mathcal{V}_o L_{d}(\tau) =0, \\ \mathcal{V}_o K_{no}^\prime =0. \\\end{align} \]

(209.a)

Actually, these are projections of more conservative LMIs (where variables have the full dimensions corresponding to the plant) onto the observable subspaces.

Example 19

Consider the spiking neuron in Example 15. The solutions to this system are known to have a dwell time with flow lengths remaining in a compact set \( \mathcal{I} = [\tau_m, \tau_M]\) where \( \tau_m > 0\) . In the jump map of system (173), we assume that \( c\) and \( d\) are unknown and include them in the state to be estimated along with \( (\eta_1,\eta_2)\) . In [22], a jump-based observer for this extended state is proposed using the knowledge of both \( y_c=\eta_1\) during flows (for output injection) and \( y_d = \eta_1\) at jumps. We show here that, using the decomposition in this paper, we can design a flow-based observer using the knowledge of \( y_c=\eta_1\) during flows only, although the flow dynamics are not observable. In other words, we take \( y_d=0\) , which comes back to not using any output injection at jumps. This was not possible via the designs in [22]. We remodel system (173) extended with \( (c,d)\) into the form (155) with \( x = (x_1, x_2, x_3, x_4) = (\eta_1, \eta_2, c, d) \in \mathbb{R}^4\) and the matrices

\[ \begin{equation} F = \begin{pmatrix} 5 & -1 & 0 & 0 \\ ab & -a & 0 &0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \end{pmatrix}, ~ H_c = \begin{pmatrix} 1 & 0 &0 & 0 \end{pmatrix}, ~ J = \begin{pmatrix} 0 & 0 & 1 & 0 \\ 0 & 1 & 0 & 1 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}, ~ H_d = \begin{pmatrix} 0 & 0 &0 & 0 \end{pmatrix}. \end{equation} \]

(210)

We see that \( x_1\) and \( x_2\) are instantaneously observable from \( y_c\) and following (183), we get \( z_o = \begin{pmatrix} \mu_1(\tau) & \mu_2(\tau) & 0 &0\\ \mu_3(\tau) & \mu_4(\tau)&0 & 0 \end{pmatrix}x\in \mathbb{R}^2\) , \( z_{no} = \begin{pmatrix} 0&0&1 & 0 \\ 0&0&0&1 \end{pmatrix}x = (x_3,x_4)\in \mathbb{R}^2\) , and the form (185) with matrices \( J_o(\tau) = \begin{pmatrix} 0 & 0 \\ \mu_5(\tau) & \mu_6(\tau) \end{pmatrix}\) , \( J_{ono}(\tau) = \begin{pmatrix} 1 &0 \\ 0 & 1\end{pmatrix}\) , \( J_{noo}(\tau) = \begin{pmatrix} 0 & 0 \\ 0 & 0 \end{pmatrix}\) , \( J_{no}(\tau) = \begin{pmatrix} 1 &0 \\ 0 & 1\end{pmatrix}\) , \( H_{co}(\tau) = \begin{pmatrix} \mu_7(\tau) & \mu_8(\tau)\end{pmatrix}\) , \( H_{do}(\tau) = \begin{pmatrix} 0 & 0 \end{pmatrix}\) , and \( H_{dno}(\tau) = \begin{pmatrix} 0 & 0 \end{pmatrix}\) , where \( \mu_i\) , \( i = 1, 2, …, 8\) are known exponential functions of \( \tau\) . We see that for any \( \tau \in \mathcal{I}\) , \( z_{no}\) cannot be seen from \( y_d\) because \( H_{dno}(\tau) = 0\) , but it can be accessed through \( z_o\) via \( J_{ono}(\tau)\) (hidden dynamics). Solving (193), we obtain \( K_{no} = \begin{pmatrix} 1 &0 \\ 0 & 1\end{pmatrix}\) with any \( Q_{no} \in §_{>0}^{1}\) and any \( L_{dno}(\cdot)\) . Let us then take \( L_{dno} = 0\) . In this application, we see that \( (x_3,x_4)\) is estimated thanks to its interaction with \( z_o\) at jumps, the latter being estimated during flows, namely we exploit the hidden observability analyzed in Theorem 11. Then, an LMI-based observer as in observer (194) or (203), with the mentioned gains \( K_{no}\) and \( L_{dno}\) , a weighting matrix taken as \( R = \rm{Id}\) , and a large enough \( \lambda\) , can be designed for this system.

3.2.3.4 KKL-Based Observer Design Using Observability Decomposition

The idea of this section is to replace the LMI-based design of the observer for the \( z_{no}\) part with a systematic KKL-based one. To do that, following Section 2.3, we exploit a discrete-time KKL observer design for the discrete-time linear time-varying system (191), for which observability is assumed as suggested by Theorem 12. The main reasons for relying on this discrete-time design, as opposed to a discrete-time Kalman(-like) design, for instance, are two-fold:

  • Compared to a Kalman design [79], it admits a strict Input-to-State Stability[ISS] Lyapunov function, allowing for an interconnection with the high-gain flow-based observer of \( z_o\) ;
  • Compared to a Kalman-like design [43], its gain with respect to the fictitious output in system (191) is constant, allowing us to reproduce in the analysis a similar change of variables as (198) in the LMI-based design of Theorem 13.

For this method, we make the following assumption.

Assumption 15

Given a subset \( \mathcal{I} \subset [0,+\infty)\) , the matrix \( J_{no}(\tau)\) is uniformly invertible for all \( \tau\in \mathcal{I}\) , i.e., there exists \( s_J > 0\) such that \( \|(J_{no}(\tau))^{-1}\| \leq s_J\) for all \( \tau \in \mathcal{I}\) .

Remark 29

Contrary to discrete-time systems, the jump map of a hybrid system has little chance of being invertible since it is not a discretization of some continuous-time dynamics. However, here thanks to the transformation (183), the flow dynamics are somehow merged with the jumps and thus it is reasonable to expect \( J_{no}(\tau)\) to be invertible for all \( \tau\in \mathcal{I}\) . On the other hand, the ability to deal with the non-invertibility of the dynamics has been studied in [127] (in the linear context). With hybrid systems, it can even be coped with using the non-uniqueness of system representation as analyzed in Section 1.5.3.2, as done in Example 15. Note that while simulations suggest that the invertibility of the jump dynamics is typically not necessary in the Kalman-like design in Section 3.2.2, and may only be for theoretical analysis, it is needed here to implement observer (221) below, namely to compute \( T^+\) correctly.

3.2.3.4.1 Discrete-Time KKL Observer Design for System (191)

Consider the discrete-time system (191) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) . Following the KKL spirit, we look for a transformation \( (T_k)_{k \in \mathbb{N}}\) such that in the new coordinates \( \eta_k := T_k z_{no,k}\) , system (191) follows the dynamics

\[ \begin{equation} \eta_{k+1} = \gamma A \eta_k + B y_k, \end{equation} \]

(211)

where \( A \in \mathbb{R}^{n_\eta \times n_\eta}\) is Schur, \( B \in \mathbb{R}^{n_\eta \times n_{d, \rm{ext}}}\) where \( n_{d, \rm{ext}}:=n_{y,d} + n_o\) such that the discrete-time pair \( (A,B)\) is controllable, and \( \gamma \in (0,1]\) is a design parameter. It then follows that the transformation \( (T_k)_{k \in \mathbb{N}}\) must be such that for every \( (\tau_k)_{k \in \mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) ,

\[ \begin{equation} T_{k+1} J_{no}(\tau_k) = \gamma AT_k + BH_{d,\rm ext}(\tau_k), \end{equation} \]

(212)

with \( H_{d,\rm{ext}}(\cdot)\) defined in system (191) in Theorem 11. The interest of this form is that an observer for system (191) in the \( \eta\) -coordinates is a simple filter of the output

\[ \begin{equation} \hat{\eta}_{k+1} = \gamma A \hat{\eta}_k + B y_k, \end{equation} \]

(213)

making the estimation error \( \tilde{\eta}_k = \eta_k - \hat{\eta}_k\) verify

\[ \begin{equation} \tilde{\eta}_{k+1} = \gamma A \tilde{\eta}_k, \end{equation} \]

(214)

and thus is exponentially stable. Then, if \( (T_k)_{k \in \mathbb{N}}\) is uniformly left-invertible, the estimate defined by \( \hat{z}_{no,k} = T_k^* \hat{\eta}\) , where \( (T^*_k)_{k \in \mathbb{N}}\) is a bounded sequence of left inverses of \( (T_k)_{k \in \mathbb{N}}\) verifying \( T^*_k T_k = \rm{Id}\) for all \( k \in \mathbb{N}_{\geq k^\star}\) for some \( k^\star \in \mathbb{N}\) , is such that for any \( (\tau_k)_{k\in\mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) , there exist \( c_1 > 0\) and \( c_2 \in (0,1)\) such that for any initial conditions \( z_{no,0}\) and \( \hat{z}_{no,0}\) and for all \( k \in\mathbb{N}\) ,

\[ \begin{equation} |z_{no,k} - \hat{z}_{no,k}| \leq c_1 c_2^k |z_{no,0}-\hat{z}_{no,0}|. \end{equation} \]

(215)

From Corollary 1, we know that this is possible under the uniform backward distinguishability of system (191) as defined next.

Definition 18 (Uniform backward distinguishability of system (191))

Given \( (\tau_k)_{k \in \mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) , system (191) is uniformly backward distinguishable if for each \( i \in \{1, 2, …, n_{d,\rm{ext}}\}\) its output dimension, there exists \( m_i \in \mathbb{N}_{>0}\) such that there exists \( \alpha_m>0\) such that for all \( k \in \mathbb{N}_{\geq \overline{m}}\) with \( \overline{m} := \max_{i \in \{1,2,…,n_{d,\rm{ext}}\}} m_i\) , the backward distinguishability matrix sequence \( (\mathcal{O}^{bw}_k)_{k\in\mathbb{N}}\) defined as

\[ \begin{equation} \mathcal{O}^{bw}_k = ( \mathcal{O}^{bw}_{1,k}, \mathcal{O}^{bw}_{2,k}, \ldots, \mathcal{O}^{bw}_{n_{d,\rm ext},k} ) \in \mathbb{R}^{(\sum_{i=1}^{n_{d, \rm ext}} m_i) \times n_{no}}, \end{equation} \]

(216.a)

where

\[ \begin{equation} \mathcal{O}^{bw}_{i,k} :=\begin{pmatrix*}[l] H_{d,\text{ext},i}(\tau_{k-1}) (J_{no}(\tau_{k-1}))^{-1} \\ H_{d,\text{ext},i}(\tau_{k-2}) (J_{no}(\tau_{k-2}))^{-1} (J_{no}(\tau_{k-1}))^{-1} \\ \ldots \\ H_{d,\text{ext},i}(\tau_{k-(m_i-1)}) (J_{no}(\tau_{k-(m_i-1)}))^{-1}\ldots (J_{no}(\tau_{k-1}))^{-1}\\ H_{d,\text{ext},i}(\tau_{k-m_i}) (J_{no}(\tau_{k-m_i}))^{-1}(J_{no}(\tau_{k-(m_i-1)}))^{-1}\ldots (J_{no}(\tau_{k-1}))^{-1} \end{pmatrix*}, \end{equation} \]

(216.b)

where \( H_{d,\text{ext},i}(\cdot)\) denotes the \( i^{\text{th}}\) row of the extended output matrix \( H_{d,\text{ext}}(\cdot)\) of system (191), has full rank and satisfies

\[ \begin{equation} (\mathcal{O}^{bw}_k)^\top \mathcal{O}^{bw}_k \geq \alpha_m \rm{Id} > 0. \end{equation} \]

(217)

Note that the forward version \( (\mathcal{O}^{fw}_k)_{k\in\mathbb{N}}\) of \( (\mathcal{O}^{bw}_k)_{k\in\mathbb{N}}\) , which is much easier to compute, can be considered as in Remark 10, under the additional assumptions that the \( m_i\) are the same for \( i\in \{1,2,…,n_{d, \rm{ext}}\}\) and there exists \( c_J>0\) such that \( ((J_{no}(\tau_k))^{-1})^\top (J_{no}(\tau_k))^{-1}\geq c_J \rm{Id}\) for all \( k \in \mathbb{N}\) . Lemma 11, which is a particular case of Theorem 4 and Theorem 5, states the existence and uniform left-invertibility of \( (T_k)_{k\in\mathbb{N}}\) solving (212).

Lemma 11 (Discrete-time KKL observer design for system (191))

Consider \( n_\eta\in \mathbb{N}_{>0}\) , \( \gamma > 0\) , and a pair \( (A,B)\in \mathbb{R}^{n_\eta \times n_\eta}\times\mathbb{R}^{n_\eta\times n_{d,\rm{ext}}}\) . Under Assumption 15, for any \( (\tau_k)_{k \in \mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) and any \( T_0 \in \mathbb{R}^{n_\eta \times n_{no}}\) , the transformation \( (T_k)_{k\in\mathbb{N}}\) initialized as \( T_0\) and satisfying (212) uniquely exists under the following closed form

\[ \begin{equation} T_k = (\gamma A)^k T_0 \prod_{p=0}^{k-1}(J_{no}(\tau_p))^{-1} + \sum_{p=0}^{k-1}(\gamma A)^{k-p-1}BH_{d,\rm ext}(\tau_p) \prod_{r=p}^{k-1}(J_{no}(\tau_r))^{-1}. \end{equation} \]

(218)

Moreover, for any \( (\tau_k)_{k \in \mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) such that system (191) is uniformly backward distinguishable for some \( m_i\in \mathbb{N}_{> 0}\) , \( i \in \{1,2,…,n_{d, \rm{ext}}\}\) and any controllable pairs \( (\tilde{A}_i,\tilde{B}_i) \in \mathbb{R}^{m_i \times m_i}\times\mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur, \( i \in \{1,2,…,n_{d, \rm{ext}}\}\) , there exists \( 0<\gamma^\star \leq 1\) such that for any \( 0 < \gamma < \gamma^\star\) , there exist \( k^\star \in \mathbb{N}_{>0}\) , \( \underline{c}_T>0\) , and \( \overline{c}_T>0\) such that \( (T_k)_{k\in\mathbb{N}}\) in (218) with

\[ \begin{align} A &= \rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_{d, \rm ext}}) \in \mathbb{R}^{n_\eta \times n_\eta}, \end{align} \]

(219.a)

\[ \begin{align} B& = \rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_{d,\rm ext}) \in \mathbb{R}^{n_\eta \times n_{d, \rm ext}}}, \\ \\\end{align} \]

(219.b)

where \( n_\eta := \sum_{i=1}^{n_{d, \rm{ext}}}m_i\) , verifies \( T_k^\top T_k \geq \underline{c}_T \rm{Id}\) for all \( k \in \mathbb{N}_{\geq k^\star}\) and \( \|T_k\| \leq \overline{c}_T\) for all \( k\in \mathbb{N}\) .

In other words, for \( \gamma\) sufficiently small, \( (T_k)_{k\in\mathbb{N}}\) is uniformly left-invertible and upper-bounded after \( k^\star\) . Note that the dependence of \( \gamma^\star\) , \( \underline{c}_T\) , \( \overline{c}_T\) , and \( k^\star\) on \( (\tau_k)_{k \in \mathbb{N}}\) and \( T_0\) is only through \( \alpha_m\) and the \( m_i\) coming from the uniform backward distinguishability and the upper bounds of \( T_0\) and \( (J_{no}(\cdot))^{-1}\) . Note also that the (non-uniform) injectivity of \( T\) can be obtained from non-uniform distinguishability conditions, as seen in Example 8, which may suffice in some cases to ensure the convergence of the KKL observer.

Proof. These results are the particular case of Theorem 4 and Theorem 5. Note that by continuity, \( \tau\mapsto J_{no}(\tau)\) is uniformly invertible on \( \mathcal{I}\) because \( \mathcal{I}\) is compact, and \( \tau \mapsto H_{dno}(\tau)\) is uniformly bounded on the compact set \( [0, \tau_M]\) . \( \blacksquare\)

Remark 30

Interestingly, Definition 18, whose nonlinear version is defined in Definition 12, coincides with the UCO condition required by the discrete-time Kalman(-like) filter (see [79, Condition (13)][43, Assumption 3], or [81, Definition 3]), on the pair \( \left(J_{no}(\tau),\begin{pmatrix} H_{dno}(\tau) \\ J_{ono}(\tau) \end{pmatrix}\right)\) , which is the discrete-time version of that in Definition 17 above, i.e., there exist \( m_i \in \mathbb{N}_{> 0}\) and \( c_o > 0\) such that for all \( (\tau_k)_{k \in \mathbb{N}}\) with \( \tau_k \in \mathcal{I}\) for all \( k \in \mathbb{N}\) , we have for all \( k \in \mathbb{N}_{\geq \overline{m}}\) ,

\[ \begin{equation} \sum_{i=1}^{n_{d, \rm ext}}\sum_{p = k-m_i}^{k-1} \star^\top H_{d,\text{ext},i}(\tau_p) (J_{no}(\tau_p))^{-1} \ldots (J_{no}(\tau_{k-2}))^{-1} (J_{no}(\tau_{k-1}))^{-1} \geq c_o \rm{Id} > 0. \end{equation} \]

(220)

It is thus interesting to compare these two discrete-time observers. In terms of dimensions, the complexity of the Kalman filter is \( \frac{n_{no}(n_{no}+1)}{2} + n_{no}\) , while that of the KKL observer is \( \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right)n_{no} + \sum_{i=1}^{n_{d, \rm{ext}}}m_i\) . Therefore, the Kalman filter is advantageous in dimension compared to the KKL observer. However, the advantage of the latter (besides being applicable in the nonlinear context) is that there exists a strict ISS Lyapunov function of quadratic form that allows us to prove exponential ISS, unlike the discrete-time Kalman filter [79] whose Lyapunov function is not strict. This advantage is then exploited in the next part, where we design a KKL-based observer to estimate the \( z_{no}\) part in the hybrid system (185) while the estimation error in \( z_o\) is seen as a disturbance. Note that a discrete-time Kalman-like observer [43] could seem like a possible alternative to the KKL-based one since it also exhibits a strict Lyapunov function. However, the gain multiplied with the fictitious output in the observer must be constant during flows for us to perform the analysis (see \( K_{no}\) in (193)), which is not the case in a Kalman-like observer (unless the pair \( (J_{no}(\tau),H_{dno}(\tau))\) at the jump times is UCO, so without the need for the fictitious output). This is ensured in KKL design since it relies on a transformation into a linear time-invariant form (see below in the proof of Theorem 15).

Next, in Section 3.2.3.4.2, we exploit this section’s results for the hybrid system (185).

3.2.3.4.2 KKL-Based Observer Design for System (185)

The KKL-based observer we propose for system (185) has the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{z}}_o &=& G_o(\tau)u_c+P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}(y_c - H_{co}(\tau)\hat{z}_o) \\ \dot{\hat{\eta}} & = & (TG_{no}(\tau) - B_{ono}G_o(\tau))u_c \\ \dot{P} &= &{}-\lambda P + (H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau) \\ \dot{T} & = & 0 \\ \dot{\tau} &=& 1\end{array}\right\} \text{when}~(155) \text{ flows} \\ \\ \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{z}_o^+ &=& J_o(\tau) \hat{z}_o + J_{ono}(\tau)\hat{z}_{no} + \mathcal{V}_o u_d \\ \hat{\eta}^+ &=& (T^+ J_{noo}(\tau) + \gamma A B_{ono} - B_{ono}J_o(\tau))\hat{z}_o + \gamma A \hat{\eta}\\ &&{} + (T^+ \mathcal{V}_{no}-B_{ono}\mathcal{V}_o) u_d + B_{dno}(y_d - H_{do}(\tau)\hat{z}_o) \\ P^+ &=& P_0\\ T^+ & = & (\gamma AT + B_{dno}H_{dno}(\tau)+ B_{ono}J_{ono}(\tau))\rm{sat}_{s_J}(J_{no}^{\dagger}(\tau)) \\ \tau^+ &=& 0 \end{array}\right\} \text{when}~(155) \text{ jumps} \end{array} \right. \end{equation} \]

(221.a)

with

\[ \begin{equation} \hat{z}_{no} = \rm{sat}_{s_T}(T^\dagger)(\hat{\eta} + B_{ono}\hat{z}_o), \end{equation} \]

(221.b)

with \( P_0\in §_{> 0}^{n_o}\) , \( \tau \mapsto R(\tau) \in §_{>0}^{n_{y,c}}\) is a positive definite weighting matrix that is defined and is continuous on \( [0, \tau_M]\) to be chosen only for design purpose, \( \gamma \in (0,1]\) , and \( (A,B_{dno},B_{ono})\) are design parameters to be chosen. Following similar reasoning as in Section 3.2.3.3.1, those dynamics are picked so that \( T\) coincides with the \( (T_k)_{k\in\mathbb{N}}\) in Section 3.2.3.4.1 at jumps and so that the corresponding discrete-time KKL estimation error dynamics (214) appear after a certain change of coordinates, modulo some errors on \( z_o\) (see (229) below). The difficulty comes here from the fact that the discrete-time output \( y_k\) in the discrete-time KKL dynamics (211) is not fully available at jumps since it contains the fictitious output.

Assumption 16

Given \( j_m\) defined in Assumption 13, there exist \( m_i \in \mathbb{N}_{>0}\) for each \( i=1,2,…,n_{d, \rm{ext}}\) and \( \alpha_m > 0\) such that for every complete solution \( x \in §_\mathcal{H}(\mathcal{X}_0, \mathfrak{U})\) , the sequence of flow lengths \( (\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) where \( \tau_j = t_{j+1} - t_j\) is such that system (191), scheduled with that \( (\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) , is uniformly backward distinguishable with the parameters \( m_i\) and \( \alpha_m\) following Definition 18.

Theorem 15 (KKL-based observer design for system (155) in the \( \eta\) -coordinates)

Suppose Assumptions 91315, and 16 hold. Define \( n_\eta := \sum_{i=1}^{n_{d, \rm{ext}}}m_i\) and consider for each \( i \in \{1,2,…,n_{d, \rm{ext}}\}\) a controllable pair \( (\tilde{A}_i,\tilde{B}_i) \in \mathbb{R}^{m_i \times m_i}\times\mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur. Let

\[ \begin{align} A &= \rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_{d, \rm ext}}) \in \mathbb{R}^{n_\eta \times n_\eta}, \end{align} \]

(222.a)

\[ \begin{align} B_{dno}& = \rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_{y,d}}) \in \mathbb{R}^{n_\eta \times n_{y,d}},\\ B_{ono}& = \rm{diag}(\tilde{B}_{n_{y,d}+1}, \tilde{B}_{n_{y,d}+2}, \ldots, \tilde{B}_{n_{d, \rm ext}}) \in \mathbb{R}^{n_\eta \times n_o}. \\ \\\end{align} \]

(222.b)

Given any \( \lambda^\prime > 0\) , any \( P_0\in §_{> 0}^{n_o}\) , and any \( T_0\in \mathbb{R}^{n_\eta \times n_{no}}\) , there exists \( 0<\gamma^\star \leq 1\) such that there exists \( \lambda^\star>0\) such that for any \( 0<\gamma < \gamma^\star\) and for any \( \lambda > \lambda^\star\) , there exist \( \overline{j} \in \mathbb{N}_{>0}\) , saturation levels \( s_T>0\) , \( s_J > 0\) , and scalar \( \rho^\prime>0\) such that for any solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and any solution \( (\hat{z}_o,\hat{\eta},P,T,\tau)\) to observer (221) with \( P(0,0)=P_0\) , \( T(0,0)=T_0\) , \( \tau(0,0)=0\) , the chosen \( (A,B_{dno},B_{ono})\) , \( \rm{sat}_{s_T}\) at level \( s_T\) , and \( \rm{sat}_{s_J}\) at level \( s_J\) , \( (\hat{z}_o,\hat{\eta},P,T,\tau)\) is complete and we have

\[ \begin{equation} |x(t, j) - \hat{x}(t, j)| \leq \rho^\prime |x(t_{\overline{j}},\overline{j}) - \hat{x}(t_{\overline{j}},\overline{j})| e^{-\lambda^\prime(t+j)}, ~~ \forall (t, j)\in \rm{dom} x:j \geq \overline{j}, \end{equation} \]

(223)

with \( \hat{x}\) obtained by \( \hat{x} = e^{F \tau} \mathcal{D}\hat{z}\) with \( \mathcal{D}\) defined in (184).

The parameter \( \overline{j}\) is related to \( j_m\) in Assumption 15 and to the number of jumps needed to get the uniform left-injectivity of \( (T_k)_{k\in\mathbb{N}}\) in Lemma 11.

Proof. First, according to Assumption 13, the flow lengths of solutions in \( §_\mathcal{H}(\mathcal{X}_0, \mathfrak{U})\) are in the compact set \( [0,\tau_M]\) , so there exists \( \overline{c}_{T,m}>0\) such that for every solution \( x \in §_\mathcal{H}(\mathcal{X}_0, \mathfrak{U})\) and any solution \( (\hat{z}_o,\hat{\eta},P,T,\tau)\) to observer (221) with \( P(0,0)=P_0\) , \( T(0,0)=T_0\) , and \( \tau(0,0)=0\) , we have \( \|T(t,j)\|\leq \overline{c}_{T,m}\) for all \( (t,j)\in \rm{dom} x\) such that \( j\leq j_m\) . Then, from Assumptions 1315, and 16, and according to Lemma 11 starting from jump \( j_m\) , there exists \( 0<\gamma_0^\star\leq 1\) such that for any \( 0<\gamma < \gamma_0^\star\) , there exist \( j^\star \in \mathbb{N}_{>0}\) , \( \underline{c}_T>0\) , and \( \overline{c}_T>0\) such that for every solution \( x \in §_\mathcal{H}(\mathcal{X}_0, \mathfrak{U})\) , the solution \( (T_j)_{j\in \mathbb{N}_{\geq j_m}}\) to (212) with \( \tau_j = t_{j+1} - t_j\) , initialized at any \( T_{j_m}\) verifying \( \|T_{j_m}\|\leq \overline{c}_{T,m}\) , is uniformly left-invertible for all \( j\in\mathbb{N}_{\geq j_m+j^\star}\) and uniformly bounded for all \( j\in\mathbb{N}_{\geq j_m}\) , i.e.,

\[ \begin{equation} T_j^\top T_j \geq \underline{c}_T \rm{Id}, ~~ \forall j\in\mathbb{N}_{\geq j_m+j^\star}, ~~ \|T_j\| \leq \overline{c}_T, ~~ \forall j \in \mathbb{N}_{\geq j_m}. \end{equation} \]

(224)

It follows that \( (T^{\dagger}_j)^\top T^\dagger_j \leq (1/\underline{c}_T) \rm{Id}\) , for all \( j\in \mathbb{N}_{\geq j_m+j^\star}\) , and there exists a saturation level \( s_T > 0\) such that \( \rm{sat}_{s_T}(T^\dagger_j)=T^\dagger_j\) for all \( j\in \mathbb{N}_{\geq j_m+j^\star}\) . Pick \( 0<\gamma<\gamma_0^\star\) and consider a solution \( x \in §_\mathcal{H}(\mathcal{X}_0,\mathfrak{U})\) and a solution \( (\hat{z}_o,\hat{\eta},P,T,\tau)\) to observer (221) with \( P(0,0)=P_0\) , \( T(0,0) =T_0\) , \( \tau(0,0)=0\) , the chosen \( (A,B_{dno},B_{ono})\) , and the saturation level \( s_T\) . Following (183), define

\[ \begin{equation} z(t,j) := \mathcal{V} e^{-F \tau(t,j)} x(t,j), ~~ \forall (t,j)\in \rm{dom} x, \end{equation} \]

(225)

and consider the estimation error \( \tilde{z} = (\tilde{z}_o, \tilde{z}_{no}) = (z_o - \hat{z}_o, z_{no} - \hat{z}_{no})\) . As justified above, \( \|T(t_{j_m},j_m)\|\leq \overline{c}_{T,m}\) . Since \( \dot{T} = 0\) during flows, the sequence \( (T(t_j,j))_{j\in \mathbb{N}_{\geq j_m}}\) coincides with the sequence \( (T_j)_{j\in \mathbb{N}_{\geq j_m}}\) solution to (212) with \( \tau_j = t_{j+1} - t_j\) for all \( j\in \mathbb{N}_{\geq j_m}\) . Therefore,

\[ \begin{equation} (T(t,j))^\top T(t,j) \geq \underline{c}_T \rm{Id}, ~~ \forall(t,j)\in \rm{dom} x: j \geq j_m+j^\star. \end{equation} \]

(226)

We now use Corollary 2 to show exponential convergence of \( \tilde{z}\) starting at hybrid time \( (t_{j_m+j^\star},j_m+j^\star)\) by putting the estimation error dynamics into the appropriate form. In order to exploit the KKL design, we define

\[ \begin{equation} \eta(t,j) = T(t,j) z_{no}(t,j) - B_{ono}z_o(t,j), ~~ \forall (t,j)\in \rm{dom} x. \end{equation} \]

(227)

Notice that \( \eta\) verifies \( \dot{\eta} = (TG_{no}(\tau) - B_{ono}G_o(\tau))u_c\) during flows. From Assumptions 13 and 15, after hybrid time \( (t_{j_m},j_m)\) we get \( \rm{sat}_{s_J}(J_{no}^\dagger(\tau)) = J_{no}^\dagger(\tau)\) and \( J_{no}^\dagger(\tau)J_{no}(\tau) = \rm{Id}\) so that at jumps,

\[ \begin{equation} \eta^+ = (T^+ J_{noo}(\tau) + \gamma A B_{ono} - B_{ono}J_o(\tau))z_o + \gamma A \eta + (T^+ \mathcal{V}_{no}-B_{ono}\mathcal{V}_o) u_d + B_{dno}H_{dno}(\tau)z_{no}. \end{equation} \]

(228)

Given the dynamics of \( \hat{\eta}\) in observer (221), the estimation error \( \tilde{\eta} := \eta - \hat{\eta}\) verifies \( \dot{\tilde{\eta}} = 0\) during flows and at jumps (after hybrid time \( (t_{j_m},j_m)\) ),

\[ \begin{equation} \tilde{\eta}^+ = (T^+ J_{noo}(\tau) + \gamma A B_{ono} - B_{ono}J_o(\tau)- B_{dno}H_{dno}(\tau))\tilde{z}_o + \gamma A \tilde{\eta}, \end{equation} \]

(229)

which is a contracting dynamics in \( \tilde{\eta}\) . After \( (t_{j_m+j^\star},j_m+j^\star)\) , we have i) \( T^\dagger T = \rm{Id}\) so that \( z_{no} = T^\dagger \eta + T^\dagger B_{ono} z_o\) , and ii) \( \rm{sat}_{s_T}(T^\dagger)=T^\dagger\) so that \( \hat{z}_{no} = T^\dagger \hat\eta + T^\dagger B_{ono} \hat z_o\) . Therefore, after \( (t_{j_m+j^\star},j_m+j^\star)\) , \( (\tilde{z}_o,\tilde{\eta},P,\tau)\) is solution to

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{z}}_o &=& {}-P^{-1}(H_{co}(\tau))^\top (R(\tau))^{-1}H_{co}(\tau)\tilde{z}_o \\ \dot{\tilde{\eta}} & = & 0 \\ \dot{P} &= &{}-\lambda P + (H_{co}(\tau))^\top (R(\tau))^{-1} H_{co}(\tau) \\ \dot{\tau} &=& 1 \\ \\ \tilde{z}_o^+ &=& \overline{J}_o(\tau) \tilde{z}_o + J_{ono}(\tau) T^\dagger \tilde{\eta} \\ \tilde{\eta}^+ & = & (T^+J_{noo}(\tau) + \gamma A B_{ono} - B_{ono}J_o(\tau)- B_{dno}H_{dno}(\tau))\tilde{z}_o + \gamma A \tilde{\eta}\\ P^+ &=& P_0 \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(230.a)

where \( \overline{J}_o(\tau) = J_o(\tau) + J_{ono}T^\dagger B_{ono}\) , with \( T^+\) seen as a uniformly bounded input, and with the flow and jump sets

\[ \begin{equation} \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times \mathbb{R}^{n_o \times n_o}\times [0,\tau_M], ~~ \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times \mathbb{R}^{n_o \times n_o}\times \mathcal{I}. \end{equation} \]

(230.b)

Since \( A\) is Schur, let \( Q_\eta \in §_{>0}^{n_\eta}\) be a solution to the inequality \( A^\top Q_\eta A - Q_\eta < 0\) . Using Corollary 2, we prove that there exist \( \lambda^\star > 0\) and \( 0 < \gamma^\star \leq 1\) such that we have the arbitrarily fast GES of the estimation error \( (\tilde{z}_o,\tilde{\eta})\) with respect to the value \( (\tilde{z}_o,\tilde{\eta})(t_{j_m+j^\star},j_m+j^\star)\) when \( \lambda > \lambda^\star\) and \( 0 < \gamma < \gamma^\star\) . Then the GES in the \( z\) -coordinates with respect to \( (\tilde{z}_o,\tilde{z}_{no})(t_{j_m+j^\star},j_m+j^\star)\) is obtained thanks to the uniform left-invertibility of \( T\) . Last, the arbitrarily fast GES is recovered in the \( x\) -coordinates after hybrid time \( (t_{\overline{j}},\overline{j})\) where \( \overline{j} = j_m + j^\star\) by seeing that \( \tilde{z} = \mathcal{V} e^{-F \tau}\tilde{x}\) with \( \tau \in [0, \tau_M]\) . \( \blacksquare\)

Remark 31

Note that it is the rate of convergence that can be arbitrarily fast and not the convergence time since we must anyway wait for \( j_m\) jumps for the flow lengths to be in \( \mathcal{I}\) and, more importantly, for the \( \overline{m}\) jumps giving us uniform backward distinguishability (see Definition 18). Furthermore, speeding up the rate may make \( T\) poorer conditioned, thus increasing the bound \( \rho^\prime\) , which is the well-known peaking phenomenon. This type of result is typical in high-gain KKL designs (see Section 2.3). Note though that this arbitrarily fast convergence rate is an advantage compared to the LMI-based design in Section 3.2.3.3 where the rate is fixed once the LMI is solved: Corollary 2 does not apply in that case because the parameters \( a\) and \( Q_\eta\) in (389) are not independent, i.e., \( Q_\eta\) is not such that (389) holds for any \( a > 0\) .

Similarly to the LMI-based design in Section 3.2.3.3, this KKL-based one in this context of linear maps can also be written in the \( z\) - and the original \( x\) -coordinates. However, due to the \( \eta\) -coordinates potentially having a higher dimension, we typically need to add fictitious states to equalize dimensions and properly extend the KKL matrix. That has been done in our published work [158, Sections 6.4.3 and 6.4.4] and is omitted in this dissertation due to space constraints.

Example 20

Consider the same system as in Example 19. First, with \( J_{no}(\tau) = \begin{pmatrix} 1 &0 \\ 0 & 1\end{pmatrix}\) for all \( \tau \in \mathcal{I}\) , Assumption 15 is satisfied. Since \( H_{dno}(\tau) = 0\) for all \( \tau \in \mathcal{I}\) , we discard the jump output and only consider the fictitious one described by the matrix \( J_{ono}(\tau)\) . We then see that with \( m_1 = m_2 = 1\) , for all sequences of flow lengths \( (\tau_j)_{j \in \mathbb{N}}\) ,

\[ \begin{equation} \mathcal{O}^{bw}_j = J_{ono}(\tau_j) (J_{no}(\tau_j))^{-1} = \rm{Id}_2 \end{equation} \]

(231)

satisfies \( (\mathcal{O}^{bw}_j)^\top\mathcal{O}^{bw}_j = \rm{Id}_2 > 0\) for all \( j \in \mathbb{N}_{\geq \max\{m_1,m_2\}} = 1\) . Therefore, Assumption 16 is satisfied and we can design a KKL-based observer where \( T\) is of dimension \( 2 \times 2\) . Let us take \( A = \rm{diag}(0.1, 0.2)\) , an empty \( B_{dno}\) , and \( B_{ono} = \begin{pmatrix} 1 & 0 \\ 0 & 1 \end{pmatrix}\) . Then, a KKL-based observer as in observer (221) with the mentioned \( (A,B_{dno},B_{ono})\) , a weighting matrix taken as \( R = \rm{Id}\) , a large enough \( \lambda\) , and a small enough \( \gamma\) , can be designed for this system.

3.2.4 Conclusion

This section presents and discusses new results on observer design for general hybrid systems with linear maps and known jump times. After defining and discussing the hybrid uniform complete observability condition, we present the hybrid Kalman-like observer in Section 3.2.2. We then propose in Section 3.2.3 a decomposition of the state of a hybrid system with linear maps and known jump times into a part that is instantaneously observable during flows and a part that is not. A thorough analysis of the asymptotic detectability of the second part is performed, where we show that this part can actually be detectable from an extended output made of the jump output and a fictitious one thanks to the flow-jump combination. A high-gain Kalman-like observer with resets at jumps is proposed to estimate the first part, while two different jump-based algorithms are proposed for the second one. Several examples are provided to illustrate the methods.

A comparison among the mentioned designs, namely the Kalman-like observer in Section 3.2.2 and the two designs depending on observability decomposition in Section 3.2.3, is presented in Table 3 at the end of this chapter. While the KKL-based design requires stronger conditions and has a larger dimension than the LMI-based one, it can provide an arbitrarily fast convergence rate of the estimate (achieved after a certain time). Compared to the Kalman-like design in Section 3.2.2, the LMI-based observer in Section 3.2.3.3 has a smaller dimension, whereas the KKL-based one in Section 3.2.3.4 can be bigger or smaller depending on \( n_o\) versus \( n_{no}\) . Therefore, for a particular application, if \( n_o\) is large compared to \( n_{no}\) , going through a decomposition is advantageous dimension-wise. Note also that while the Kalman-like observer can easily deal with time-varying matrices in the system dynamics and output, the decomposition method must additionally assume the (uniform) existence of the transformation into system (185), for example, to solve the LMI (193) for the gains \( L_{dno}(\cdot)\) and \( K_{no}\) . Furthermore, even though the invertibility of \( J_{no}(\tau)\) for all \( \tau \in \mathcal{I}\) assumed for the KKL-based method may seem lighter than the invertibility of \( J\) at the jump times assumed in the Kalman-like approach in Assumption 10 (since \( J_{no}(\tau)\) has a smaller dimension than \( J\) ), it may turn out stronger since \( J_{no}^{\dagger}(\tau)\) is used in the KKL implementation (see observer (221)), while \( J^{-1}\) is used only for analysis in the Kalman-like design, not in the implementation (see observer (156)).

Next, in Section 3.3, we study hybrid systems with nonlinear maps (and known jump times) by finding transformations into coordinates of possibly higher dimensions and Lyapunov-based sufficient conditions to couple different observers Section 6.2.2.1.

Acknowledgments. We thank Clément Boutaric, student at Mines Paris - PSL, and Lorenzo Marconi, professor at University of Bologna, for their collaborations in Section 3.2.2 and Section 3.2.3, respectively.

Table 3. Comparison of observer designs for hybrid systems with linear maps and known jump times. All the conditions here are sufficient conditions.
Kalman-like observer (156) (Section 3.2.2)LMI-based observer (194) (Section 3.2.3.3)KKL-based observer (221) (Section 3.2.3.4)
Time domainNo assumptionPersistence of both flows and jumps (Assumption 13)
Hybrid modelInvertibility of \( J\) at the jump times (Assumption 10)The pair \( (F,H_c)\) constantThe pair \( (F,H_c)\) constant, invertibility of \( J_{no}(\tau)\) for all \( \tau \in \mathcal{I}\) (Assumption 15)
ObservabilityUniform complete observability of system (155) (Definition 17)Instantaneous observability of \( z_o\) and a form of quadratic detectability of system (191) for \( \tau \in \mathcal{I}\) (Assumption 14)Instantaneous observability of \( z_o\) and uniform backward distinguishability of system (191) for the \( (\tau_j)_{j \in \mathbb{N}}\) generated by the time domain of \( x\) (Assumption 16)
Dimension\( n_x + \frac{n_x(n_x+1)}{2}\)\( n_x+\frac{n_o(n_o+1)}{2}+1\)\( n_o + \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right) + \frac{n_o(n_o+1)}{2} + \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right)n_{no} + 1\)
Convergence rateArbitrarily fast, achieved after a certain time (Theorem 9)Determined by the solution of a matrix inequality (Assumption 14)Arbitrarily fast, achieved after a certain time (Theorem 15)
ISS Lyapunov functionYes (Theorem 10)YesYes

3.3 Observer Design for Hybrid Systems with Nonlinear Maps

 

Nous proposons de nouvelles conceptions d’observateurs pour les systèmes hybrides avec des dynamiques et sorties non linéaires et des instants de saut connus, en nous basant sur une décomposition de l’état en plusieurs parties présentant des propriétés d’observabilité différentes. Nous supposons que l’état du système considéré peut être décomposé en deux parties, la première étant indépendante de la seconde pendant les phases de flot. On suppose que la première partie est instantanément observable pendant les flots à partir de la sortie continue, tandis que la seconde partie doit être soit détectable, constructible ou distinguable en temps rétrograde via la combinaison du flot et des sauts, à partir de la sortie discrète ainsi qu’une sortie fictive décrivant comment cette seconde partie impacte la première lors des sauts, et devient ainsi visible à partir de la sortie continue dans l’intervalle de flot suivant. Une analyse de la détectabilité montre la nécessité de la détectabilité à partir de cette sortie étendue pour l’existence d’un observateur. Un observateur est ensuite conçu pour estimer chaque partie, à savoir un observateur grand gain basé sur les flots utilisant la sortie continue pour la première partie de l’état et un observateur basé sur les sauts par LMI\( /\) MHO\( /\) KKL, utilisant une sortie de saut étendue pour la seconde partie de l’état. L’indisponibilité de la sortie fictive est traitée implicitement dans les observateurs basés sur les sauts. La convergence ou stabilité exponentielle globale de l’erreur d’estimation dans les coordonnées d’origine est prouvée par une analyse de Lyapunov. Nos contributions théoriques sont illustrées à travers des exemples académiques.

3.3.1 Introduction

Half of this chapter has already been submitted to a journal. The other half will also be submitted to a journal. Some related preliminary results that have been published in [166] are put in Section 6.2.2.1.

In Section 3.2.3, we exploit an observability decomposition of the system state, where a part is instantaneously observable from the flow output, while the rest is (at least) detectable from the jump output as well as the combination of flows and jumps. Indeed, as observed in [143], the interaction between flows and jumps may render visible, in the flow output, state components that are not observable through the flow dynamics. In this chapter, we extend these ideas in the context of observer design for hybrid systems with nonlinear maps. Our main contributions are as follows. First, the state is assumed to be decomposed into two parts, where the first one is instantaneously observable during flows from the flow output. Assuming both persistent flowing and jumping, at least after a certain time, we show via a detectability analysis that the second part of the state needs to be detectable at jumps from an extended output made of the jump output and a fictitious one, obtained from the jump dynamics of the first part that becomes observable via the flow output during the subsequent flow interval (hidden dynamics exhibited in [143]). This in turn allows us to explicitly construct and couple observers estimating each mentioned part of the state depending on their observability properties, namely an arbitrarily fast high-gain flow-based observer exploiting the flow output for the first part and a jump-based observer exploiting the (fictitious) extended output for the second one. Input-to-state properties of each observer are exploited to handle the interconnections at jumps. More precisely, for a wide class of hybrid systems with nonlinear maps, we propose two designs for the jump-based part of the observer: a Linear Matrix Inequality[LMI]-based design exploiting a detectability property, and a KKL-based one inspired by the Kravaris-Kazantzis$\slash$Luenberger[KKL] method such as [105, 42] or in Section 2.3 under an observability property. Thanks to general Lyapunov-based conditions to couple observers in Section 6.2.2.1, a high-gain-like result is obtained, where the Global Exponential Stability[GES] of the estimation error is achieved in the original coordinates if the high-gain flow-based observer is pushed sufficiently fast. Then, we propose observer designs for the case where the maps are fully nonlinear, still using the high-gain observer as the flow-based observer but with either a Moving Horizon Observer[MHO] under constructibility or a KKL observer under backward distinguishability as the jump-based observer.

In Section 3.4, the observers proposed in this chapter are applied for estimating state and uncertainties in mechanical systems with impacts, including a bouncing ball and a bipedal walking robot.

3.3.2 Problem Formulation

In this chapter, we generalize the decomposition idea in Section 3.2.3 for hybrid systems with nonlinear maps. To this end, we assume that system (145) with fully nonlinear maps has been put via some uniformly injective transformation into a standard form for which we will consider throughout this chapter to design observers, then the estimate can be recovered in the \( x\) -coordinates thanks to this injectivity. For the sake of presentation, we will not repeat this highly technical transformation process that is similar to Theorem 3 (which is in discrete time), but go straight to the considered target form.

Consider a hybrid system

\[ \begin{equation} \mathcal{H}\left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\xi}_o &=& f_o(\xi_o, u_c) \\ \dot{\xi}_{no} &=& f_{no}(\xi_o, \xi_{no}, u_c) \end{array} \right\} (\xi,u_c)\in C\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \xi_o^+ &=& g_o(\xi_o, \xi_{no}, u_d) \\ \xi_{no}^+ &=& g_{no}(\xi_o, \xi_{no}, u_d) \end{array} \right\} (\xi,u_d)\in D \end{array} \right. \end{equation} \]

(232.a)

with state \( \xi = (\xi_o,\xi_{no}) \in \mathbb{R}^{n_\xi}\) where \( n_\xi = n_o + n_{no}\) , with flow and jump maps \( f = (f_o,f_{no})\) and \( g = (g_o,g_{no})\) , flow and jump sets \( C\) and \( D\) , and flow and jump outputs

\[ \begin{equation} y_c = h_c(\xi_o,u_c), ~~ y_d = h_d(\xi_o,\xi_{no},u_d), \end{equation} \]

(232.b)

respectively. The jump times of each solution to system (232) are assumed to be detected without delay (the robustness with respect to delays in jump detection is discussed in Remark 19). Note that the outputs \( y_c\) and \( y_d\) may either come from sensor measurements or other known information about the state, for instance, the flow and jump conditions, which encodes the fact that \( (\xi,u_c) \in C\) and \( (\xi,u_d)\in D\) , respectively. Solutions and inputs to system (232) are defined in Definition 3.

Remark 32

Model (232) covers time-varying systems [167, 168] (by including the times \( t\) , \( j\) in the inputs or the state), impulsive and switched systems with state jumps as in [59] (by treating the switching signal as a known input), and continuous-time systems with (possibly multi-rate) sampled outputs, with fast sampled outputs treated as \( y_c\) and other sporadic outputs treated as \( y_d\) (by considering the sampling events as jumps triggering the availability of \( y_d\) )—see for instance Example 16 for this modeling.

The goal of this chapter is to design an asymptotic observer for system \( \mathcal{H}\) in (232) (as defined below in (235)), assuming that its jump times are known, based on the idea of observability decomposition and the coupling of flow- and jump-based observers Section 3.2.3. Since in practice, we may be interested in estimating only certain trajectories of “physical interest”, we denote in what follows \( \Xi_0\) a set containing the initial conditions of the trajectories to be estimated and \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ) a set of locally bounded continuous (resp., discrete) inputs of interest, defined on \( \mathbb{R}_{\geq 0}\) (resp., \( \mathbb{N}\) ). We then denote \( §_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) as the set of maximal solutions to \( \mathcal{H}\) initialized in \( \Xi_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , as defined in Definition 3. Let us denote \( \mathcal{U}_c \subseteq \mathbb{R}^{n_{u,c}}\) (resp., \( \mathcal{U}_d \subseteq \mathbb{R}^{n_{u,d}}\) ) as the set of points that the inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) (resp., \( \mathfrak{u}_d \in \mathfrak{U}_d\) ) take values in. Following [22], our design only requires us to have information about the possible duration of flow intervals between successive jumps in the trajectories of interest as defined next.

Definition 19 (Flow lengths of a hybrid arc)

For a closed subset \( \mathcal{I}\) of \( \mathbb{R}_{\geq 0}\) and some \( j_m \in \mathbb{N}\) , we say that a hybrid arc \( (t,j) \mapsto x(t,j)\) has flow lengths within \( \mathcal{I}\) after jump \( j_m \in \mathbb{N}\) if:

  1. \( 0 \leq t - t_j(x) \leq \sup\mathcal{I}\) for all \( (t,j) \in \rm{dom} x\) ;
  2. \( t_{j+1}(x) - t_j(x) \in \mathcal{I}\) holds for all \( j \in \mathbb{N}_{\geq j_m}\) if \( \sup \rm{dom}_j x = +\infty\) , and for all \( j \in \{j_m, j_m + 1, …, \sup \rm{dom}_j x - 1\}\) otherwise.

In brief, \( \mathcal{I}\) contains all the possible lengths of the flow intervals between successive jumps, at least after some time. In Definition 19Item (XII) is to bound the length of the flow intervals not covered by Item (XIII), namely possibly the first one, which is \( [0, t_1]\) , and the last one, which is \( \rm{dom}_t x \cap [t_{J(x)}, +\infty)\) , where \( t_{J(x)}\) is the time when the last jump happens (when the number of jumps is finite). For some classes of systems where the jump times of solutions are determined by exogenous inputs and thus are independent of trajectories, such as those in Remark 32, \( j_m\) can be zero.

As noticed in [22], the observer type to use depends on \( \mathcal{I}\) . In particular, [22] studies how to design:

  • A flow-based observer with an innovation term during flows only, exploiting the observability of the full state during flows obtained from measuring \( y_c\) , when \( 0\notin \mathcal{I}\) , i.e., \( \min\mathcal{I}>0\) , and the solutions exhibit a dwell time after some time;
  • A jump-based observer with an innovation term at jumps only, exploiting the detectability or observability\( /\) distinguishability of the full state via the combination of flows and jumps obtained from \( y_d\) available at the jumps only, when \( \mathcal{I}\) is bounded, i.e., when the solutions exhibit persistent jumps.

In this chapter, we tackle the general case where some state components may be observable during flows while others may be observable via the combination of flows and jumps. More precisely, we will assume that the component \( \xi_o\) of the state \( \xi\) in system (232) is instantaneouly observable from \( y_c\) during flows, while the rest of the state \( \xi_{no}\) draws observability from the coupling of flows and jumps. It follows that both flows and jumps need to be exploited to reconstruct the state and that neither (eventually) continuous nor discrete\( /\) Zeno solutions are allowed: both flows and jumps need to be persistent, as assumed next. Moreover, because we are interested in asymptotic observers (in the sense that convergence holds asymptotically in hybrid time—see in (235) below), we assume completeness of solutions.

Assumption 17

For system (232), assume that:

  1. There exist \( j_m \in \mathbb{N}\) and compact sets \( \Xi_o\subset \mathbb{R}^{n_o}\) , \( \Xi_{no}\subset \mathbb{R}^{n_{no}}\) , \( \Xi_0 \subset \Xi_o\times\Xi_{no} \subset \mathbb{R}^{n_\xi}\) , and \( \mathcal{I}\subset\mathbb{R}_{>0}\) such that each solution in \( §_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) is complete, remains in \( \Xi:=\Xi_o\times\Xi_{no}\) , and has flow lengths within \( \mathcal{I}\) after jump \( j_m\) (see Definition 19);
  2. The flow pair \( (f_o,h_c)\) is independent of \( \xi_{no}\) and is instantaneously observable, in the sense that, for any \( \mathfrak{u}_c\in \mathfrak{U}_c\) , the knowledge of \( t\mapsto h_c(\xi_o(t),u_c(t))\) for an arbitrarily short amount of time uniquely determines the solution \( t\mapsto \xi_o(t)\) to \( \dot{\xi}_o=f_o(\xi_o,u_c)\) as long as \( (\xi_o(t),\mathfrak{u}_c(t))\in \{(\xi_o,u_c) \in \mathbb{R}^{n_o}\times\mathcal{U}_c: \exists \xi_{no} \in \mathbb{R}^{n_{no}}: (\xi_o,\xi_{no},u_c) \in C\}\) .

Remark 33

Assuming \( \min\mathcal{I}>0\) is not only to ensure an infinite amount of flows in the maximal solutions. Indeed, \( \sup \rm{dom}_t x = +\infty\) could happen even for solutions with flow lengths vanishing to zero. It is the ratio\( /\) balance between flows and jumps that is important to ensure that the estimation error contracting during flows decreases sufficiently to compensate for its possible increase at jumps, and vice versa. In that sense, this persistence of flows could be relaxed into an average-dwell-time condition with \( \tau_m\) replaced by the average dwell time by appropriately modifying the Lyapunov analysis as in [169, Proposition IV.1] or [22, Theorem 3.1].

Example 21 (Modeling a spiking neuron with unknown parameters in the form (232))

Based on [161], the spiking behavior of a cortical neuron with state \( s = (s_1,s_2) \in \mathbb{R}^2\) can be modeled as the hybrid system

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{~~}l@{}} \dot{s} &=&\begin{pmatrix} 0.04 s_1^2 + 5s_1 + 140 - s_2 + I_{\text{ext}}\\ a(bs_1 - s_2) \end{pmatrix}, & \text{when } s_1 \leq v_m, & y_c = s_1, \\ s^+ &=& \begin{pmatrix} c\\ s_2 + d \end{pmatrix},& \text{when } s_1 = v_m, & y_d = s_1, \end{array}\right. \end{equation} \]

(233)

where \( s_1\) is the membrane potential in mV, \( s_2\) is the recovery variable, and \( I_{\text{ext}}\) represents the (constant) synaptic current or injected DC current [161]. The time scale is in milliseconds and the parameters are \( I_{\text{ext}} = 10\) , \( a = 0.02\) , \( b = 0.2\) , \( c = -55\) , \( d = 4\) , and \( v_m = 30\) (all in appropriate units), characterizing the neuron type and its firing pattern [161]. The solutions to this system are known to have a dwell time with flow lengths remaining in a compact set \( \mathcal{I} = [\tau_m, \tau_M]\) where \( \tau_m > 0\) , and the jump times can be detected from the jumps of the output \( y_c = s_1\) . Let us assume the parameters \( a\) , \( b\) , \( c\) , and \( d\) are all unknown constants. Since \( c\) is the value of \( s_1\) right after the jumps, it is directly obtained by the value of \( y_c\) after the jumps. We would like to estimate the rest of the parameters by modeling them as extra state components. The extended system with state \( x = (s, a, b, d)\) is

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{~~}l@{}} \dot{x} &=& \begin{pmatrix} 0.04x_1^2 + 5x_1 - x_2 + I_{\rm ext}^\prime\\ x_3 (x_4x_1 - x_2)\\ 0\\ 0\\ 0 \end{pmatrix}, & \text{when } x_1 \leq v_m, & y_c = x_1, \\ x^+ &=& \begin{pmatrix} c\\ x_2 + x_5\\ x_3\\ x_4\\ x_5 \end{pmatrix},& \text{when } x_1 = v_m, & y_d = x_1, \end{array}\right. \end{equation} \]

(234)

where \( I_\rm{ ext}^\prime = 140 + I_\rm{ ext}\) . We see that the flow output \( y_c\) and its first derivative \( \dot{y}_c\) determine \( (x_1,x_2)\) uniquely. Then, the following derivatives of \( (x_1,x_2)\) verify

\[ \begin{pmatrix} \dot{x}_2 \\ \ddot{x}_2 \end{pmatrix} = \begin{pmatrix} x_3(x_4x_1 - x_2)\\ x_3(x_4 \dot{x}_1 - \dot{x}_2) \end{pmatrix}= \begin{pmatrix} x_1 & - x_2\\ \dot{x}_1 & -\dot{x}_2 \end{pmatrix}\begin{pmatrix} x_3x_4\\ x_3 \end{pmatrix}, \]

so that \( (x_1,x_2,x_3,x_4)\) are instantaneously observable from \( y_c\) as long as the matrices \( \begin{pmatrix} x_1 & -x_2 \\ \dot{x}_1 & - \dot{x}_2 \end{pmatrix}\) is uniformly invertible along solutions (since \( x_3\) is known to be constant and non-zero along solutions, letting us recover \( x_4\) from \( (x_3x_4,x_3)\) ). Moreover, the flow dynamics of \( (x_1,x_2,x_3,x_4)\) and the output \( y_c\) are independent of \( x_5\) . We thus obtain a hybrid system of the form (232). Note that \( x_5\) is not detectable from \( y_d\) , but it becomes visible via its interaction with \( x_2\) at jumps (hidden dynamics). These ideas will be exploited in the next part to design an observer for systems of the form (232) based on splitting the state into \( (\xi_o,\xi_{no})\) and coupling different observers estimating each of these parts.

Thus, our goal is to design an asymptotic observer for system (232), as introduced right next, assuming we know: i) the exogenous signals \( u_c\) and \( u_d\) , ii) the output(s) \( y_c\) during flows and\( /\) or \( y_d\) at jumps, iii) when the plant’s jumps occur, and iv) some information about the possible flow lengths, as in Item (XIV) of Assumption 17. Since the jump times of the system are known, following [22], a synchronized asymptotic observer has dynamics of the form

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{\zeta}}&=& \mathcal{F}(\hat{\zeta},y_c,u_c) &\text{when~(232) flows}\\ \hat{\zeta}^+&=&\mathcal{G}(\hat{\zeta},y_d,u_d) &\text{when~(232) jumps} \end{array} \right. \end{equation} \]

(235.a)

with the estimate \( \hat{\xi}\) obtained from a solution \( (t,j) \mapsto \hat{\zeta}(t,j)\) to dynamics (235.a) as

\[ \begin{equation} \hat{\xi}(t,j)= \Upsilon(\hat{\zeta}(t,j),t,j), \end{equation} \]

(235.b)

where \( \hat{\zeta} \in \mathbb{R}^{n_\zeta}\) is the observer state; \( \mathcal{F}\) , \( \mathcal{G}\) , and \( \Upsilon\) are the observer dynamics and output maps designed together with an initialization set \( \mathcal{Z}_0 \subseteq \mathbb{R}^{n_\zeta}\) such that the dependence of \( \Upsilon\) on time \( (t,j)\) is only through the inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) and each maximal solution \( (\xi,\hat{\zeta})\) to the cascade (232)-(235) initialized in \( \Xi_0 \times \mathcal{Z}_0\) and with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in \mathfrak{U}_c\times \mathfrak{U}_d\) is complete and verifies

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} \xi (= \rm{dom} \hat{\xi})}}|\xi(t,j)-\hat{\xi}(t,j)| = 0. \end{equation} \]

(236)

Remark 34

As seen in Lemma 9, asymptotic detectability in Definition 14 is necessary for the existence of a synchronized asymptotic observer. However, some of the observer designs in this chapter only exploit the system dynamics and output maps as well as the flow lengths given by Item (XIV) of Assumption 17, so that the system solutions in \( §_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) are seen, after jump \( j_m\) , as the \( \xi\) component of solutions to the extended hybrid system

\[ \begin{equation} \mathcal{H}^\tau \left\{ \begin{array}{@{}l@{~~}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\xi}_o&=&f_o(\xi_o,u_c)\\ \dot{\xi}_{no}&=&f_{no}(\xi_o,\xi_{no},u_c)\\ \dot{\tau}&=&1 \end{array}\right\} (\xi,u_c, \tau)\in C^\tau & y_c = h_c(\xi_o,u_c)\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \xi_o^+&=&g_o(\xi_o,\xi_{no},u_d) \\ \xi_{no}^+&=&g_{no}(\xi_o,\xi_{no},u_d)\\ \tau^+ &=& 0 \end{array} \right\} (\xi, u_d, \tau)\in D^\tau & y_d =h_d(\xi_o,\xi_{no},u_d) \end{array} \right. \end{equation} \]

(237.a)

with the flow and jump sets (with \( \tau_M:=\max\mathcal{I}\) )

\[ \begin{equation} C^\tau = \mathbb{R}^{n_\xi}\times \mathbb{R}^{n_{u,c}} \times [0,\tau_M],~~ D^\tau = \mathbb{R}^{n_\xi}\times \mathbb{R}^{n_{u,d}} \times \mathcal{I}, \end{equation} \]

(237.b)

initialized in \( \Xi_0 \times [0,\tau_M]\) with the same inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c \times \mathfrak{U}_d\) . In this case, we are actually building an observer for system \( \mathcal{H}^\tau\) after jump \( j_m\) , which is a hybrid system with richer behavior relative to system \( \mathcal{H}\) that contains all solutions generated by those maps and flow lengths, regardless of the state flow\( /\) jump sets of system \( \mathcal{H}\) . The detectability of system \( \mathcal{H}^\tau\) is then necessary, which is stronger than the detectability of system \( \mathcal{H}\) .

3.3.2.1 Detectability from Fictitious Measurement

The observability condition stated in Item (XV) of Assumption 17 allows us to consider a high-gain observer for the pair \( (f_o,h_c)\) during flows, estimating \( \xi_o\) arbitrarily fast from the knowledge of \( y_c\) and exhibiting Input-to-State Stability[ISS] with respect to estimation errors in \( \xi_{no}\) affecting the estimate of \( \xi_o\) at jumps. Then, we show how to estimate \( \xi_{no}\) via a jump-based observer from the knowledge of \( y_d\) and of the estimate of \( \xi_o\) . Our approach is illustrated in Figure 27.

&lt;span data-controller=&quot;mathjax&quot;&gt;Illustration of our decomposition-based observer. The known inputs (u_c,u_d)) are neglected for clarity.&lt;/span&gt;
Figure 27. Illustration of our decomposition-based observer. The known inputs \( (u_c,u_d)\) are neglected for clarity.

Section 3.3.3 of this chapter thus provides conditions on the maps \( f_o\) , \( f_{no}\) , \( h_c\) , and \( h_d\) defining system (232) for which an observer (235) satisfying (236) can be designed. Our designs rely on the remark that, under Item (XV) of Assumption 17, the existence of an observer requires another equivalent detectability condition.

Lemma 12 (Asymptotic detectability from fictitious measurement)

Assume that:

  1. Each solution \( \xi\in§_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) is \( j\) -complete and verifies \( t_{j+1}(\xi) - t_j(\xi) > 0\) for all \( j \in\rm{dom}_j \xi\) (\( = \mathbb{N}\) );
  2.  Item (XV) of Assumption 17 holds;
  3. The dynamics \( \dot{\xi}_o=f_o(\xi_o,u_c)\) admit unique solutions in backward time in the set \( \{\xi_o \in \mathbb{R}^{n_o}: \exists (\xi_{no},u_c) \in \mathbb{R}^{n_{no}}\times\mathcal{U}_c: (\xi_o,\xi_{no},u_c) \in \rm{cl}(C)\}\) for \( \mathfrak{u}_c\in \mathfrak{U}_c\) ;
  4. For each solution \( \xi\in§_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) , \( t \mapsto y_c(t,j)\) is continuous during flows for all \( j \in \rm{dom}_j \xi\) .

Then, the asymptotic detectability of system \( \mathcal{H}\) in (232) on \( \Xi_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) is equivalent to the asymptotic detectability of the system with dynamics (232.a), no flow output, and only the extended jump output \( (y_d,\xi_o,g_o(\xi_o, \xi_{no}, u_d))\) , on \( \Xi_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , where \( g_o\) is defined in system (232).

Proof. Since both considered systems have the same dynamics and \( y_d\) appears in the jump output in both cases, we only need to prove that the knowledge of \( y_c\) during flows is equivalent to the knowledge of \( \xi_o\) and \( g_o(\xi_o,\xi_{no},u_d)\) at jumps, in the sense of (148). By Item (XVI) of Lemma 12, each solution in \( §_{\mathcal{H}}(\Xi_0,\mathfrak{U}_c \times \mathfrak{U}_d)\) has non-zero flow lengths and infinitely many jumps. From Remark 20 combined with Item (XVIII) of Lemma 12, this means the condition (148.a) is to be checked for all \( (t,j)\) in the domain. Since by Item (XV) of Assumption 17, the \( \xi_o\) component of the solution is instantaneously observable from \( y_c\) , for any two considered solutions \( \xi_a\) and \( \xi_b\) to system \( \mathcal{H}\) in (232) with the same inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) with \( \rm{dom} \xi_a = \rm{dom} \xi_b=:\mathcal{D}\) , having \( h_c(\xi_{a,o}(t,j),\mathfrak{u}_c(t)) = h_c(\xi_{b,o}(t,j),\mathfrak{u}_c(t))\) for all \( (t,j) \in \mathcal{D}\) is equivalent to having \( \xi_{a,o}(t,j) = \xi_{b,o}(t,j)\) for all \( (t,j) \in \mathcal{D}\) (because i) the flow lengths are non-zero, ii) \( (\xi_{o,a},\mathfrak{u}_c)\) and \( (\xi_{o,b},\mathfrak{u}_c)\) remain in \( C\) in the interior of the flow intervals by the definition of solutions, and iii) \( t \mapsto \xi_{o,a}(t,j)\) and \( t \mapsto \xi_{o,b}(t,j)\) are continuous during flows for all \( j \in \rm{dom}_j \xi_a = \rm{dom}_j \xi_b\) ). First, assume we have \( \xi_{a,o}(t,j) = \xi_{b,o}(t,j)\) for all \( (t,j) \in \mathcal{D}\) . Then, it implies in particular that, at the jump times, i.e., for all \( (t,j)\in \mathcal{D}\) such that \( (t,j+1)\in \mathcal{D}\) , we have \( \xi_{a,o}(t,j) = \xi_{b,o}(t,j)\) and \( \xi_{a,o}(t,j+1) = g_o(\xi_{a}(t,j),\mathfrak{u}_d(j)) = g_o(\xi_{b}(t,j),\mathfrak{u}_d(j))=\xi_{b,o}(t,j+1)\) . On the other hand, assume that for all \( (t,j)\in \mathcal{D}\) such that \( (t,j+1)\in \mathcal{D}\) , we have \( \xi_{a,o}(t,j) = \xi_{b,o}(t,j)\) and \( g_o(\xi_{a}(t,j),\mathfrak{u}_d(j)) = g_o(\xi_{b}(t,j),\mathfrak{u}_d(j))\) . Then, by backward uniqueness of solutions to \( f_o\) in Item (XVII) of Lemma 12, it implies that \( \xi_{a,o}(t,j) = \xi_{b,o}(t,j)\) for all \( (t,j) \in \mathcal{D}\) . \( \blacksquare\)

Note that if \( \Xi_0 = \rm{cl}(C)\cup D\) and the sets \( \mathfrak{U}_c\) and \( \mathfrak{U}_d\) are invariant with respect to the discarding of time domains, i.e., truncating the beginning of any input in \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ) still gives us an input in \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ), then Item (XVI) of Lemma 12 can be replaced by Item (XIV) of Assumption 17. In other words, Lemma 12 holds under Assumption 17, the backward uniqueness of solutions to \( \dot{\xi}_o = f_o(\xi_o,u_c)\) for each given input, and the continuity of the flow output. The proof follows similarly by seeing \( (t_{j_m},j_m)\) as the new initial time, i.e., by discarding the part of solutions before the first \( j_m\) jumps.

With Lemma 12 in mind, we proceed to estimate \( \xi_o\) from \( y_c\) sufficiently fast during flows using a high-gain flow-based observer, then estimate \( \xi_{no}\) via a jump-based observer exploiting an observability property of an appropriately defined discrete-time system with extended jump output \( (y_d,g_o(\xi_o, \xi_{no}, u_d))\) and known input \( \xi_o\) . Note that this extended output is fictitious, in the sense that it is not available at jumps. However, it can be used in the analysis by exploiting the interaction between \( \xi_{no}\) and \( \xi_o\) at jumps and the fact that \( \xi_o\) is estimated arbitrarily fast during flows from \( y_c\) . Using \( \xi_o\) as a jump output instead of \( g_o(\xi_o, \xi_{no}, u_d)\) could also be possible depending on its relationship with \( \xi_{no}\) , which may be implicit; if exploitable, it would constitute a more accessible fictitious output.

3.3.2.2 High-Gain Flow-Based Observer for \( \xi_o\)

We first describe the class of considered flow-based observers for \( \xi_o\) , which will be common to all the designs developed in this chapter. More precisely, we assume available a high-gain flow-based observer of the form

\[ \begin{equation} \dot{\hat{\xi}}_o = \hat{f}_{o,\ell}(\hat{\xi}_o, p, \tau, y_c,u_c), \end{equation} \]

(238.a)

with jump dynamics assigning \( \hat{\xi}_o^+\) to be chosen later and with possibly the extra state \( p\) synchronized with the flows and jumps of system (232), following the dynamics (see more in Example 22 below)

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{p} &=& \varphi_{c,\ell}(p, \tau, y_c, u_c)\\ p^+ &=& \varphi_{d,\ell}(p, \tau, y_d, u_d) \end{array} \right. \end{equation} \]

(238.b)

that satisfy the following assumption. Note that the potential dependence of (238) on the timer \( \tau\) comes from system (237) as analyzed in Remark 34.

Assumption 18

Suppose Assumption 17 holds and with \( \mathcal{I}\) from its Item (XIV), define \( \tau_M:=\max \mathcal{I}\) . There exist maps \( \hat{f}_{o,\ell}\) , \( \varphi_{c,\ell}\) , and \( \varphi_{d,\ell}\) , and \( \ell_1^\star > 0\) such that the following holds:

  1. There exist sets \( ¶_{0},¶_{c},¶_{d}\subseteq \mathbb{R}^{n_p}\) such that for any \( \ell>\ell_1^\star\) , for any solution \( \xi\) to system (232) initialized in \( \Xi_0\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in \mathfrak{U}_c\times \mathfrak{U}_d\) and outputs \( (y_c,y_d)\) , any solution \( p\) to the synchronized hybrid system (238.b) with inputs \( (\mathfrak{u}_c,y_c,\mathfrak{u}_d,y_d)\) , \( \tau\) defined as \( \tau(t,j)=t-t_j(\xi)\) , and initialized in \( ¶_{0}\) , is such that for all \( (t,j)\in \rm{dom} \xi\) , we have \( p(t,j)\in ¶_{c}\) and for all \( j > 0\) , \( p(t_j(\xi),j-1)\in ¶_{d}\) ;
  2. There exist scalars \( \lambda_c>0\) , \( \overline{b}_o>0\) , and a rational function \( \underline{b}_o>0\) such that for all \( \ell > \ell^\star_1\) , there exists a function \( V_{o,\ell}: \mathbb{R}^{n_o}\times \mathbb{R}^{n_o} \times \mathbb{R}^{n_p} \times \mathbb{R}\to \mathbb{R}\) such that:
    1. For all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , \( \xi= (\xi_o,\xi_{no})\in \mathbb{R}^{n_\xi}\) such that \( (\xi,u_c)\in C\) or \( (\xi,u_d)\in D\) , \( \hat{\xi}= (\hat{\xi}_o,\hat{\xi}_{no})\in \mathbb{R}^{n_\xi}\) , \( p\in ¶_{c}\cup¶_{d}\) , and \( \tau \in [0,\tau_M]\) ,

      \[ \underline{b}_o(\ell) |\xi_o-\hat{\xi}_o|^2 \leq V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) \leq \overline{b}_o |\xi_o - \hat{\xi}_o|^2; \]
    2. For all \( u_c \in \mathcal{U}_c\) , \( \xi\in \mathbb{R}^{n_\xi}\) such that \( (\xi,u_c)\in C\) , \( \hat{\xi}\in \mathbb{R}^{n_\xi}\) , \( p\in ¶_{c}\) , and \( \tau \in [0,\tau_M]\) ,

      \[ \dot{V}_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau,u_c) \leq -\ell \lambda_c V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau), \]

      where \( \dot{V}_{o,\ell}\) is the Lie derivative of \( {V}_{o,\ell}\) along the flow vector field

      \[ (f_o(\xi_o, u_c),\hat{f}_{o,\ell}(\hat{\xi}_o, p, \tau, y_c,u_c),\varphi_{c,\ell}(p,\tau,y_c,u_c),1). \]

Examples of observers verifying this assumption include the classical high-gain observer [26] (without \( p\) ) as well as its Kalman-like version [27] as illustrated in the next example.

Example 22 (High-gain observers satisfy Assumption 18)

We illustrate here how high-gain observers meet the requirements in Assumption 18. In a first case, assume that system (232) is such that \( f_o\) and \( h_c\) take the following triangular form decoupled from \( \xi_{no}\) :

\[ \begin{equation} \dot{\xi}_o = A \xi_o + \Phi(\xi_o,u_c), ~~ y_c = \xi_{o,1} = H \xi_o, \end{equation} \]

(239)

with \( A\) , \( H\) , and \( \Phi\) being of triangular observable form [26], where \( \Phi\) is locally Lipschitz with respect to \( \xi_o\) , uniformly with respect to \( u_c\) . A classical high-gain observer [26] is

\[ \begin{equation} \dot{\hat{\xi}}_o = A \hat{\xi}_o + \Phi(\rm{sat}_o(\hat{\xi}_o),u_c) + \ell\mathcal{L}(\ell)K(y_c - H\hat{\xi}_o), \end{equation} \]

(240)

where \( \ell > 0\) , \( \mathcal{L}(\ell) = \rm{diag}(1, \ell,…,\ell^{n_o-1})\) , and \( K = (k_1,k_2,…,k_{n_o})\) chosen independently of \( \ell\) such that \( A - KH\) is Hurwitz. In this case, there is no extra observer state \( p\) , so that Item (XIX) of Assumption 18 holds vacuously. It can then be checked that the Lyapunov function

\[ V_o(z_o,\hat{z}_o) = (z_o-\hat{z}_o)^\top (\mathcal{L}(\ell))^{-1} P (\mathcal{L}(\ell))^{-1}(z_o-\hat{z}_o), \]

where the constant matrix \( P = P^\top > 0\) is a solution to

\[ P(A - KH) + (A - KH)^\top P \leq -aP, \]

for some \( a > 0\) , verifies the two conditions in Item (XX) of Assumption 18. In a second case, if \( f_o\) and \( h_c\) are such that

\[ \begin{equation} \dot{\xi}_o = A(u_c,y_c) \xi_o + \Phi(\xi_o,u_c), ~~ y_c = H(u_c) \xi_o, \end{equation} \]

(241)

still with \( A\) , \( H\) , and \( \Phi\) also being of triangular observable form but varying in \( (u_c,y_c)\) , and \( \Phi\) locally Lipschitz, a Kalman-like high-gain observer [27] is

\[ \begin{equation} \dot{\hat{\xi}}_o =A(u_c,y_c) \hat{\xi}_o + \Phi(\rm{sat}_o(\hat{\xi}_o),u_c) + \ell\mathcal{L}(\ell)P^{-1}(H(u_c))^\top (y_c - H(u_c)\hat{\xi}_o), \end{equation} \]

(242)

where the covariance matrix \( P \in \mathbb{R}^{n_o \times n_o}\) is dynamically updated along the solution following

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{P} &=& -\ell(\mu P - (A(u_c,y_c))^\top P - PA(u_c,y_c) + (H(u_c))^\top H(u_c))\\ P^+ & = & P_0, \end{array}\right. \end{equation} \]

(243)

and initialized with \( P_0 > 0\) , for \( \mu > 0\) large enough. In this case, the extra observer state in (238) is \( p = P\) , with its flow and jump dynamics \( (\varphi_{c,\ell},\varphi_{d,\ell})\) given by (243). The observability property in Item (XV) of Assumption 17 is linked to the existence of \( \ell^\star>0\) and \( \alpha>0\) such that, for any \( \ell>\ell^\star\) , for all \( t>\frac{1}{\ell}\) ,

\[ \begin{equation} \int_{t-\frac{1}{\ell}}^t \star^\top \underbrace{H(u_c(s))\psi_{A(u_c,y_c)}(s,t)}_{\star}ds \geq \frac{\alpha}{\ell}((\mathcal{L}(\ell))^{-1})^2, \end{equation} \]

(244)

where \( \psi_{A(u_c,y_c)}(s,t)\) is the transition matrix of the linear continuous-time system \( \dot{v} = A(u_c,y_c)v\) from time \( t\) to time \( s\) along each \( (u_c,y_c)\) trajectory. Under this observability and with \( \mu\) sufficiently large, from the dwell time and because \( P\) is reset to a constant at jumps, there exists uniform lower and upper bounds \( 0 < \underline{c}_P \leq \overline{c}_P\) for \( P\) that are independent of \( \ell\) . So, Item (XIX) of Assumption 18 holds for \( P\) in this design, namely \( p=P\) initialized in \( ¶_0=\{P_0\}\) remains in

\[ ¶_c = ¶_d := \left\{P \in \mathbb{R}^{n_o \times n_o}: \underline{c}_P \rm{Id} \leq P(t,j) \leq \overline{c}_P \rm{Id}, \forall (t,j) \in \rm{dom} \xi\right\}. \]

It can then be checked that the Lyapunov function

\[ V_o(z_o,\hat{z}_o,P) = (z_o-\hat{z}_o)^\top (\mathcal{L}(\ell))^{-1} P (\mathcal{L}(\ell))^{-1}(z_o-\hat{z}_o), \]

verifies the conditions in Item (XX) of Assumption 18.

Example 23 (High-gain observer for \( \xi_o\) of the spiking neuron)

Consider the spiking neuron in Example 21. We have checked that \( (x_1,x_2,x_3,x_4)\) are instantaneously observable from \( y_c\) . Now, we want to design for this part of the state a high-gain flow-based observer. Therefore, we consider the change of coordinates

\[ \xi_o = \begin{pmatrix} x_1\\ 0.04x_1^2+5x_1-x_2\\ -x_3(x_4x_1 - x_2)\\ -x_3x_4(0.04x_1^2+5x_1-x_2+I_{\rm ext}^\prime) + x_3^2(x_4x_1-x_2) \end{pmatrix}, ~~ \xi_{no} = x_5, \]

where \( \xi_o\) is built from \( (x_1,\dot{x}_1,-\dot{x}_2,-\ddot{x}_2)\) . If the matrix \( \begin{pmatrix} x_1 & -x_2 \\ \dot{x}_1 & - \dot{x}_2 \end{pmatrix}\) is uniformly invertible along solutions as mentioned in Example 21, i.e., there exists some \( c > 0\) such that

\[ \begin{pmatrix} x_1 & -x_2 \\ \dot{x}_1 & - \dot{x}_2 \end{pmatrix}^\top \begin{pmatrix} x_1 & -x_2 \\ \dot{x}_1 & - \dot{x}_2 \end{pmatrix} \geq c \rm{Id} \]

along solutions, then this change of coordinates is Lipschitz injective (also because \( x_3\) is constant and non-zero along solutions), and thus, there exists a Lipschitz function \( \Theta\) such that

\[ \begin{pmatrix} x_1 \\ x_2 \\ x_3 \\ x_4 \end{pmatrix} = \begin{pmatrix} \Theta_1(\xi_o) \\ \Theta_2(\xi_o) \\ \Theta_3(\xi_o) \\ \Theta_4(\xi_o) \end{pmatrix} = \Theta(\xi_o). \]

It can then be checked that \( \dot{\xi}_o\) and \( y_c\) are of the triangular form (239), with some \( \Phi\) Lipschitz. Thus, we can build for \( \xi_o\) a high-gain observer using \( y_c\) , as in (240).

Now, we see that the observer designs in this chapter combine two ingredients: a high-gain flow-based observer for \( \xi_o\) and jump-based observers for \( \xi_{no}\) . In the next section, we start by detailing the Lyapunov-based conditions that the jump-based observer should satisfy in order to be coupled with the high-gain flow-based observer. Then in the subsequent sections, we use these to provide different designs of the jump-based observer, depending on the structure and observability properties of the system.

3.3.3 Observer Design for Affine Structures

In this section, we rely on Theorem 23 and Theorem 24, namely general Lyapunov conditions for coupling observers, to design asymptotic observers for the class of hybrid systems (232) where \( f_{no}\) and \( g=(g_o,g_{no})\) are affine with respect to \( \xi_{no}\) , namely of the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}l@{}} \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\xi}_o &=& f_o(\xi_o, u_c) \\ \dot{\xi}_{no} &=& F_{no}\xi_{no}+U_{cno} \end{array} & \bigg.\bigg\} (\xi,u_c)\in C\\ \\ \begin{array}{@{}r@{\;}c@{\;}l@{}} \xi_o^+ &=& J_o(\xi_o, u_d) + J_{ono}(\xi_o,u_d)\xi_{no}\\ \xi_{no}^+ &=& J_{no}(\xi_o, u_d)\xi_{no} + J_{noo}(\xi_o, u_d) \end{array}& \bigg.\bigg\} (\xi,u_d)\in D \end{array} \right. \end{equation} \]

(245.a)

with the outputs

\[ \begin{equation} y_c = h_c(\xi_o,u_c), ~~ y_d = H_{do}(\xi_o,u_d) + H_{dno}(\xi_o,u_d)\xi_{no}, \end{equation} \]

(245.b)

where still state \( \xi=(\xi_o,\xi_{no}) \in \mathbb{R}^{n_o}\times \mathbb{R}^{n_{no}}\) , and \( F_{no}\) and \( U_{cno}\) are constant matrices. Model (245) covers a wide class of hybrid systems, including hybrid mechanical systems with uncertain parameters at impacts, namely where the state \( \xi_o\) (typically containing positions and velocities) is observable during flows and some unknown (constant) parameters \( \xi_{no}\) affect \( \xi_o\) at jumps in an affine way. These parameters typically become detectable from \( y_c\) through the way they affect \( \xi_o\) at jumps, namely from the fictitious output \( J_{ono}(\xi_o,u_d)\xi_{no}\) , which corresponds to the part of \( g_o(\xi_o,\xi_{no},u_d)\) (see in system (232)) that is not measured. Examples of hybrid systems that fit into the form (245) range from a bouncing ball in Section 3.4.3 to spiking neurons [161] (see in the examples throughout this chapter) and walking robots [31], as studied in Section 3.4.4. Note that system (245) is here a particular case of system (232) with those dynamics and sets, so Assumption 17 and Assumption 18, if assumed, refer to system (245) in this section.

Following Item (XIV) of Assumption 17, without loss of generality, we assume that \( C\subseteq \Xi_o\times \Xi_{no}\times \mathcal{U}_c\) and \( D \subseteq \Xi_o\times \Xi_{no}\times \mathcal{U}_d\) . Denote the projected flow\( /\) jump sets:

\[ \begin{align} C_o & := \{\xi_o \in \mathbb{R}^{n_o}: \exists (\xi_{no},u_c) \in \Xi_{no}\times\mathcal{U}_c: (\xi_o,\xi_{no},u_c) \in C\},\\ C_{no} & := \{\xi_{no} \in \mathbb{R}^{n_{no}}: \exists (\xi_o,u_c) \in \Xi_{o}\times \mathcal{U}_c: (\xi_o,\xi_{no},u_c) \in C\},\\ D_o & := \{\xi_o \in \mathbb{R}^{n_o}: \exists (\xi_{no},u_d) \in \Xi_{no}\times\mathcal{U}_d: (\xi_o,\xi_{no},u_d) \in D\}, \\ D_{no} & := \{\xi_{no} \in \mathbb{R}^{n_{no}}: \exists (\xi_o,u_d) \in \Xi_{o}\times \mathcal{U}_d: (\xi_o,\xi_{no},u_d) \in D\}. \\\end{align} \]

(246.a)

We then make the following assumptions on the system dynamics and output maps.

Assumption 19

The maps \( f_o\) , \( J_o\) , \( J_{ono}\) , \( J_{no}\) , \( J_{noo}\) , \( H_{do}\) , and \( H_{dno}\) are locally Lipschitz with respect to \( \xi_o\) , uniformly in \( (u_c,u_d) \in \mathcal{U}_c\times\mathcal{U}_d\) . The maps \( J_{no}\) , \( J_{ono}\) , and \( H_{dno}\) are locally bounded with respect to \( \xi_o\) , uniformly in \( u_d \in \mathcal{U}_d\) .

Example 24 (Checking Assumption 19 for the spiking neuron)

Consider the spiking neuron in Example 21. With the choice of \( \xi_o\) and \( \xi_{no}\) as in Example 23, we obtain the form (245) with \( F_{no} = 0\) , \( U_{cno} = 0\) , and the matrices

\[ \begin{align*} J_o(\xi_o) &= \begin{pmatrix} c\\ 0.04c^2 + 5c - \Theta_2(\xi_o)\\ -\Theta_3(\xi_o) (c\Theta_4(\xi_o)- \Theta_2(\xi_o))\\ (\Theta_3(\xi_o))^2 (c \Theta_4(\xi_o) -\Theta_2(\xi_o)) - \Theta_3(\xi_o)\Theta_4(\xi_o)(0.04c^2 + 5c - \Theta_2(\xi_o) + I_{\rm ext}^\prime) \end{pmatrix}, \\ J_{ono}(\xi_o) &= \begin{pmatrix} 0\\ -1\\ \Theta_3(\xi_o)\\ -(\Theta_3(\xi_o))^2 + \Theta_3(\xi_o) \Theta_4(\xi_o) \end{pmatrix}, \end{align*} \]

with the map \( \Theta\) described in Example 23, \( J_{no}(\xi_o) = 1\) , \( J_{noo}(\xi_o) = 0\) , \( H_{do}(\xi_o) = \xi_{o,1}\) , and \( H_{dno}(\xi_o) = 0\) . Thanks to the Lipschitzness of \( \Theta\) in Example 23, these maps are both locally Lipschitz and locally bounded in \( \xi_o\) , so Assumption 19 is verified for this system.

To estimate the full state \( \xi = (\xi_o,\xi_{no})\) of system (245), we propose to combine the high-gain flow-based observer (238) for \( \xi_o\) with a jump-based observer for \( \xi_{no}\) . For this, inspired by the detectability analysis of Lemma 12, we will rely on the observability\( /\) detectability properties of an equivalent (time-varying) discrete-time system, modeling the dynamics of \( k\mapsto\xi_{no}(t_k,k)\) sampled after each jump, with output made of \( H_{dno}(\xi_o,u_d)\xi_{no}\) appearing in \( y_d\) as well as \( J_{ono}(\xi_o,u_d)\xi_{no}\) affecting \( \xi_o\) at jumps, both sampled at the jump times. More precisely, we consider

\[ \begin{equation} \xi_{no,k+1} = \mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k) \xi_{no,k}, \end{equation} \]

(247.a)

with the extended output

\[ \begin{equation} y_k = \begin{pmatrix} \mathcal{H}_{dno}(\xi_{o,k},u_{d,k},\tau_k)\\ \mathcal{J}_{ono}(\xi_{o,k},u_{d,k},\tau_k)\end{pmatrix} \xi_{no,k} \in \mathbb{R}^{n_{y,d,\rm ext}}, \end{equation} \]

(247.b)

where \( n_{y,d,\rm{ext}} := n_{y,d} + n_o\) and

\[ \begin{align} \mathcal{J}_{no}(\xi_{o},u_{d},\tau) &:=J_{no}(\xi_{o},u_{d})e^{F_{no}\tau}, \\ \mathcal{H}_{dno}(\xi_{o},u_{d},\tau) &:= H_{dno}(\xi_{o},u_{d})e^{F_{no}\tau},\\ \mathcal{J}_{ono}(\xi_{o},u_{d},\tau) &:= J_{ono}(\xi_{o},u_{d})e^{F_{no}\tau}, \\\end{align} \]

(247.c)

with inputs \( (\xi_{o,k},u_{d,k},\tau_k) \in (\Xi_o \cap D_o) \times \mathcal{U}_d \times \mathcal{I}\) for all \( k \in \mathbb{N}\) , where \( \tau_k\) models the length of the \( k\) -th flow interval. System (247) can be interpreted as a time-varying discrete-time system, with dynamics and output of the form \( \xi_{no,k+1} = \mathcal{A}(k) \xi_{no,k}\) and \( y_k = \mathcal{C}(k) \xi_{no,k}\) . From Lemma 12, we see that the possibility of designing a jump-based observer for system (245) relies on the detectability of the equivalent discrete-time system (247).

3.3.3.1 Observer Design Based on Quadratic Detectability

3.3.3.1.1 Observer Construction

In this part, we propose a design combining a high-gain flow-based observer with an LMI-based jump-based one. More precisely, the observer we propose for system (245) takes the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{\xi}}_o &=& \hat{f}_{o,\ell}(\hat{\xi}_o, p, \tau, y_c,u_c) \\ \dot{\hat{\xi}}_{no} &=& F_{no}\hat{\xi}_{no} + U_{cno}+e^{F_{no}\tau}K_d\frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,u_c)}(\hat{\xi}_o,t,-\tau)\\ \dot{p}&=&\varphi_{c,\ell}(p, \tau, y_c, u_c) \\ \dot{\tau}&=&1 \end{array}\right\} \text{when}~(245) \text{ flows} \\ \\ \left.\begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{\xi}_o^+ &=& J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no}) \\ \hat{\xi}_{no}^+ &=& J_{no}(\rm{sat}_o(\hat{\xi}_o), u_d)\hat{\xi}_{no} + J_{noo}(\rm{sat}_o(\hat{\xi}_o), u_d)\\ &&{}+ L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (y_d - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)-H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d)\hat{\xi}_{no})\\ &&{} - K_d J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d) (\hat{\xi}_{no} - \rm{sat}_{no}(\hat{\xi}_{no})) \\ p^+ & = & \varphi_{d,\ell}(p, \tau, y_d, u_d)\\ \tau^+ &=&0 \end{array}\right\} \text{when}~(245) \text{ jumps} \end{array} \right. \end{equation} \]

(248)

with \( \hat{f}_{o,\ell}\) being a high-gain observer as described in (238) to be defined, with gain \( \ell > 0\) and possibly extra states \( p \in \mathbb{R}^{n_p}\) with dynamics \( (\varphi_{c,\ell},\varphi_{d,\ell})\) depending on \( \ell\) , \( f_{o,\rm{sat}}\) being a map that is globally Lipschitz with respect to \( \xi_o\) , uniformly in \( u_c\in \mathcal{U}_c\) , and equal to \( f_o\) on \( \Xi_o\times \mathcal{U}_c\) (guaranteed to exist by Assumption 19 and [40, Corollary 1]), \( \Psi_{f_{o,\rm{sat}}}\) denoting the flow operator associated with \( f_{o,\rm{sat}}\) as defined in the Notations above, \( \rm{sat}_{no}\) being a bounded map such that \( \rm{sat}_{no}(\xi_{no}) = \xi_{no}\) for all \( \xi_{no} \in (\Xi_{no}\cap D_{no}) + \overline{c}_{no}\mathbb{B}\) for some \( \overline{c}_{no} > 0\) and \( K_d\) and \( L_d\) being gains and \( \rm{sat}_o\) being a bounded saturation function to be designed. Note that all the terms in this observer have an explicit expression except for \( \frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\) , for which we propose an alternative expression and an approximation algorithm in Section 3.3.3.1.2.

The high-gain observer (238) for \( \xi_o\) , verifying Assumption 18, combined with i) the decoupling of \( \xi_o\) from \( \xi_{no}\) during flows, ii) the saturation of the impact of \( \hat{\xi}_{no}\) on \( \hat{\xi}_o\) at jumps through \( \rm{sat}_{no}\) , and iii) the dwell time (after jump \( j_m\) ), allows us to make the estimation error \( \xi_o(t_{j+1}(\xi),j) - \hat{\xi}_o(t_{j+1}(\xi),j)\) at jumps arbitrarily small by further choosing \( \ell\) large. Then, we make the following assumption to design a jump-based observer for \( \xi_{no}\) .

Assumption 20

There exist \( \overline{c}_o > 0\) , \( 0 \leq a < 1\) , a symmetric positive definite matrix \( Q \in \mathbb{R}^{n_{no}\times n_{no}}\) , and gains \( K_d\in \mathbb{R}^{n_{no}\times n_o}\) and \( (\xi_o,u_d,\tau) \mapsto L_d(\xi_o,u_d,\tau)\in \mathbb{R}^{n_{no}\times n_{y,d}}\) bounded on \( ((\Xi_o\cap D_o) + \overline{c}_o \mathbb{B}) \times \mathcal{U}_d \times \mathcal{I}\) such that for all \( (\xi_o,u_d,\tau) \in ((\Xi_o \cap D_o) + \overline{c}_o \mathbb{B})\times\mathcal{U}_d\times\mathcal{I}\) ,

\[ \begin{equation} (\Phi(\xi_o,u_d,\tau))^\top Q \Phi(\xi_o,u_d,\tau) \leq a Q, \end{equation} \]

(249.a)

where

\[ \begin{equation} \Phi(\xi_o,u_d,\tau)=\left(\mathcal{J}_{no}(\xi_o,u_d,\tau) - \begin{pmatrix} L_d(\xi_o,u_d,\tau) & K_d\end{pmatrix}\begin{pmatrix} \mathcal{H}_{dno}(\xi_o,u_d,\tau) \\ \mathcal{J}_{ono}(\xi_o,u_d,\tau) \end{pmatrix}\right). \end{equation} \]

(249.b)

Note that if \( \Phi\) is continuous and the sets \( \Xi_o \cap D_o\) and \( \mathcal{U}_d\) are compact, then satisfying

\[ \begin{equation} (\Phi(\xi_o,u_d,\tau))^\top Q \Phi(\xi_o,u_d,\tau) < Q \end{equation} \]

(250)

on the compact set \( (\xi_o,u_d,\tau) \in ((\Xi_o \cap D_o) + \overline{c}_o \mathbb{B})\times\mathcal{U}_d\times\mathcal{I}\) for some \( \overline{c}_o > 0\) implies the existence of \( 0 \leq a < 1\) such that (249) holds. Assumption 20 deals with the detectability of the discrete-time system (247) with inputs \( (\xi_o,u_d,\tau)\) . Contrary to \( L_d\) which may depend on \( (\xi_o,u_d,\tau)\) , \( K_d\) is required to be constant to perform the analysis (see in the proof of Theorem 16). This extra requirement is similar to the one we made in Assumption 14 for hybrid systems with linear maps. It is stronger than the notion of quadratic detectability [44] by the constant nature of \( K_d\) . Constructive methods to solve (249) are described later in Section 3.3.3.1.2.

Theorem 16 (Asymptotic stability of the LMI-based observer (248))

Suppose Assumptions 171819, and 20 hold. Consider \( \overline{c}_o\) , \( K_d\) , and \( L_d\) defined in Assumption 20, and a bounded map \( \rm{sat}_o\) such that \( \rm{sat}_o(\hat{\xi}_o) = \hat{\xi}_o\) for all \( \hat{\xi}_o \in (\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}\) . Then, there exists \( \ell^\star>0\) such that for any \( \ell > \ell^\star\) , there exist \( \rho>0\) and \( \lambda>0\) such that any maximal solution to the cascade (245)-(248) initialized in \( \Xi_0 \times \mathbb{R}^{n_\xi}\times ¶_{0}\times\{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) is complete and verifies

\[ \begin{equation} |\xi(t,j) - \hat{\xi}(t,j)| \leq \rho |\xi(0,0) - \hat{\xi}(0,0)| e^{-\lambda(t+j)},~~ \forall (t,j) \in \rm{dom} \xi:j \geq j_m. \end{equation} \]

(251)

Proof. This highly technical proof has been moved to Section 6.2.2.2 to facilitate reading. It involves three main steps: i) Define new coordinates \( (z_{no},\hat{z}_{no})\) , replacing \( (\xi_{no},\hat{\xi}_{no})\) and obtain the state dynamics in those new coordinates, ii) Show that the Lyapunov conditions in Theorem 23 hold after jump \( j_m\) and obtain exponential stability of the estimation error in the new coordinates with respect to time \( (t_{j_m},j_m)\) , where \( j_m\) comes from Item (XIV) of Assumption 17, and iii) Recover the exponential stability in the \( \xi\) -coordinates with respect to the initial time. \( \blacksquare\)

3.3.3.1.2 Constructive Methods for Implementing Observer (248)

A drawback in observer (248) lies in the computation of the correction term \( \frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\) in \( \dot{\hat{\xi}}_{no}\) . A first solution to avoid this term is to implement the observer in the new coordinates defined in the proof of Theorem 16, where \( \hat{z}_{no}\) replaces the observer state \( \hat{\xi}_{no}\) with dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{z}}_{no} & = &0\\ \hat{z}^+_{no} &=& \Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) \left(\hat{z}_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)+\displaystyle\int_{0}^{\tau}e^{-F_{no}s}U_{cno}ds\right)\\ &&{}+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)-K_dJ_o(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &&{}+ L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (y_d-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)), \end{array} \right. \end{equation} \]

(252)

as shown in (430) and \( \hat{\xi}_{no}\) is recovered from \( \hat{z}_{no}\) at all times with (429.a), as the observer’s output. In this case, a numerical scheme must be used to integrate backward \( f_{o,\rm{sat}}\) and compute the term \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\) in (429.a). Note though that, as shown in the proof of Theorem 16, this term estimates the value of \( \xi_o\) at the beginning of the flow interval, i.e., \( \xi_o(t_j,j)\) , which is by essence constant during flows: it is thus sufficient to update its estimate only “from time to time”, as allowed by the available computational power, at a frequency independent of the integration steps of the observer, and at the jump times for observer implementation. This is rendered possible by the fact that \( \hat{\xi}_{no}\) is used in the dynamics of observer (248) only at the jump times to compute \( \hat{\xi}_o^+\) .

When \( (\xi_o,t)\mapsto f_{o,\rm{sat}}(\xi_o,\mathfrak{u}_c(t))\) is continuous and \( C^1\) with respect to \( \xi_o\) , a second option to implement (248) is to notice by Lemma 26 in Section 6.2.2.6, that observer (248) can equivalently be written and implemented with

\[ \begin{equation} \dot{\hat{\xi}}_{no} = F_{no}\hat{\xi}_{no} + U_{cno}-e^{F_{no}\tau}K_d\frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau) \left(f_{o,\rm{sat}}(\hat{\xi}_o,u_c) - \hat{f}_{o,\ell}(\hat{\xi}_o, p, \tau, y_c,u_c) \right). \end{equation} \]

(253)

Intuitively, the update of \( \hat{\xi}_{no}\) during flows comes from the mismatch between \( f_{o,\rm{sat}}\) and \( \hat{f}_{o,\ell}\) (given in observer (248)) at \( \hat{\xi}_o\) , which typically remains zero only when \( \hat{\xi}_o=\xi_o\) : if an estimation error in \( \xi_{no}\) triggers an estimation error in \( \xi_o^+\) , it becomes visible through \( y_c\) and corrected during the subsequent flow interval.

Then, we exploit Lemma 27 in Section 6.2.2.6 to propose an approximation strategy of the gain \( \frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau)\) in (253), when it cannot be exactly computed. Indeed, Lemma 27 in Section 6.2.2.6 shows that, when \( (\xi_o,t)\mapsto f_{o,\rm{sat}}(\xi_o,\mathfrak{u}_c(t))\) is continuous and \( C^1\) with respect to \( \xi_o\) , for any given \( (\hat{\xi}_o,t)\in \mathbb{R}^{n_o}\times \mathbb{R}_{\geq 0}\) , the pair

\[ \left(\Psi_{f_{o,\rm{sat}}}(\hat{\xi}_o,t,-\tau), \frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau) \right) \]

is obtained by integrating backward, during \( \tau\) unit(s) of time, the system

\[ \begin{equation} \frac{d}{ds} \zeta(s) = f_{o,\rm{sat}}(\zeta(s),\mathfrak{u}_c(s)),~~ \frac{d}{ds} \Phi_o(s) = \frac{\partial f_{o,\rm{sat}}}{\partial \xi_o}(\zeta(s),\mathfrak{u}_c(s)) \Phi_o(s), \end{equation} \]

(254)

from final condition \( (\zeta(t),\Phi_o(t))=(\hat{\xi}_o, \rm{Id})\) at time \( t\) . In other words, the gain \( \frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau)\) in (253) is approximated by executing Algorithm 1 at each time step, based on a forward Euler discretization of (254). While we have found in some simulations that this scheme works with a small enough step size—see Section 3.4.4, any other time discretization scheme of (254) (with possibly varying steps) can be used, depending on the properties of the differential equation (254) (e.g., stiffness, invariance), and the available computational power. Note that the gain \( \frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau)\) in (253) appears only in front of a term that vanishes when \( \hat{\xi}_o\) converges to \( \xi_o\) (when \( \hat{f}_{o,\ell}\) is properly designed), so that if the approximation is precise enough, the asymptotic stability of the estimation error can still be possible. Actually, in many cases, the right sign of the gain may still suffice.


Algorithm 1 Approximation of the gain \( \frac{\partial \Psi_{f_{o,\rm{sat}}}}{\partial \xi_o}(\hat{\xi}_o,t,-\tau)\) in (253)
1.Require:
2.Initialize \( \zeta = \hat{\xi}_o\) and \( \Phi_o = \rm{Id}\)
3.for \( k = 1 : \left\lfloor \frac{\tau}{\Delta} \right\rfloor\) do
4.end for
5.Output \( \Phi_o\)

Now, concerning the choice of the gains \( (L_d(\cdot),K_d)\) in Assumption 20, according to Lemma 29 in Section 6.2.2.6, we see that \( Q = Q^\top > 0\) and \( a > 0\) satisfying a strict version of (249) for some \( (L_d(\cdot),K_d)\) exist only if they are solution to

\[ \begin{equation} \begin{pmatrix}\left(\begin{pmatrix} \mathcal{H}_{dno}(\xi_o,u_d,\tau)\\ \mathcal{J}_{ono}(\xi_o,u_d,\tau) \end{pmatrix}^\bot\right)^\top Q\begin{pmatrix} \mathcal{H}_{dno}(\xi_o,u_d,\tau)\\ \mathcal{J}_{ono}(\xi_o,u_d,\tau) \end{pmatrix}^\bot & \star \\ Q \mathcal{J}_{no}(\xi_o,u_d,\tau)\begin{pmatrix} \mathcal{H}_{dno}(\xi_o,u_d,\tau)\\ \mathcal{J}_{ono}(\xi_o,u_d,\tau) \end{pmatrix}^\bot & aQ \end{pmatrix}> 0, ~ \forall (\xi_o,u_d,\tau) \in (\Xi_o \cap D_o)\times\mathcal{U}_d\times\mathcal{I}. \end{equation} \]

(255)

This is an LMI in \( (Q,aQ)\) , treating \( aQ\) as a new variable. If such \( (a,Q)\) are obtained, the gains \( L_d(\cdot)\) and \( K_d\) can then be found by using (249) with \( Q\) and \( a\) known.

If \( ((\Xi_o \cap D_o) + \overline{c}_o \mathbb{B})\times\mathcal{U}_d\times\mathcal{I}\) has infinitely many points, then there is an infinite number of LMIs to solve. Actually, it is worth noting that the exponential term \( e^{F_{no}\tau}\) contained in all the \( \tau\) -dependent matrices in (255) can be expanded using residue matrices [151], as

\[ \begin{equation} e^{F_{no}\tau} = \sum_{i=1}^{\sigma_r} \sum_{j=1}^{m_i^r} R_{ij} e^{\lambda_i \tau} \frac{\tau^{j-1}}{(j-1)!} + \sum_{i=1}^{\sigma_c} \sum_{j=1}^{m_i^c} 2e^{\Re(\lambda_i) \tau} (\Re(R_{ij}) \cos(\Im(\lambda_i) \tau) - \Im(R_{ij}) \sin(\Im(\lambda_i) \tau) ) \frac{\tau^{j-1}}{(j-1)!}, \end{equation} \]

(256)

where \( \sigma_r\) and \( \sigma_c\) are the numbers of distinct real eigenvalues and complex conjugate eigenvalue pairs; \( m_i^r\) and \( m_i^c\) are the multiplicity of the real eigenvalue \( \lambda_i\) and of the complex conjugate eigenvalue pair \( \lambda_i, \lambda_i^*\) in the minimal polynomial of \( F_{no}\) ; \( R_{ij} \in \mathbb{R}^{n_{no} \times n_{no}}\) are matrices corresponding to the residues associated to the partial fraction expansion of \( (s\rm{Id} - F_{no})^{-1}\) . This in turn allows \( e^{F_{no} \tau}\) to be written as a finite sum of matrices affine in a finite number of scalar functions \( \beta_{ij} = e^{\lambda_i \tau} \tau^{j-1}\) , \( \gamma_{ij} = e^{\Re(\lambda_i) \tau} \cos(\Im(\lambda_i) \tau) \tau^{j-1}\) , and \( \gamma^*_{ij} = e^{\Re(\lambda_i) \tau} \sin(\Im(\lambda_i) \tau) \tau^{j-1}\) . Then, it implies that if \( \Xi_o \cap D_o\) and \( \mathcal{U}_d\) are compact, (249) or (255) can be solved in a polytopic approach, i.e., the LMIs are satisfied for all \( (\xi_o,u_d,\tau) \in ((\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}) \times \mathcal{U}_d \times \mathcal{I}\) compact if they are satisfied at the finite number of vertices of the polytope formed by these scalar functions when \( (\xi_o,u_d,\tau)\) takes values in \( ((\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}) \times \mathcal{U}_d \times \mathcal{I}\) . Alternatively, the matrix inequalities can be solved in a grid-based approach assuming a particular structure of \( L_d\) followed by post-analysis of the solution’s stability as in [44], possibly with a theoretical proof extended from [152].

Example 25 (LMI-based observer design for the spiking neuron)

Consider the spiking neuron in Example 21, with the choice of \( \xi_o\) and \( \xi_{no}\) as in Example 23. Let us now design observer (248) for this system, exploiting the affine structure (245) detailed in Example 24. Given the expression of \( J_{ono}(\xi_o)\) , we see that the information provided by the second fictitious output is \( -\xi_{no}\) , which directly determines \( \xi_{no}\) , so that we may discard the jump output \( y_d\) by choosing \( L_d = 0\) and pick a gain of the form \( K_d=\begin{pmatrix} 0 & K_d^\prime & 0 &0 \end{pmatrix}\) selecting the second fictitious output only. Indeed, \( K_d^\prime\) is picked by solving (249), which reduces to, for all \( \xi_o\) in a suitable range,

\[ (1 - K_d J_{ono}(\xi_o))^\top Q (1 - K_d J_{ono}(\xi_o)) < Q. \]

This condition is verified for \( K_d^\prime \in (-2,0)\) with \( Q=1\) . We pick \( K_d^\prime = -1\) and \( K = (14, 71, 154, 120)\) for the jump-based observer, and pick \( \ell = 4\) for the high-gain flow-based observer (240). Observer (248) is implemented in the new coordinates where \( \hat{z}_{no}\) replaces the observer state \( \hat{\xi}_{no}\) with dynamics (252) and \( \hat{\xi}_{no}\) is recovered from \( \hat{z}_{no}\) at all times with (429.a), as the observer’s output. Computing the term \( \Psi_{f_{o,\rm{sat}}}(\hat{\xi}_o,t,-\tau)\) in (429.a) involves solving backward at each time step the continuous-time dynamics with map \( f_{o,\rm{sat}}\) fed with the estimate. In this chapter, we approximate this term using a simple Euler scheme with time step \( 0.001\) (s). Simulation results for the states \( x = (s_1,s_2,a,b,d)\) and the estimation errors \( (\tilde{s}_1,\tilde{s}_2,\tilde{a},\tilde{b},\tilde{d})\) are shown in Figure 28.

&lt;span data-controller=&quot;mathjax&quot;&gt;State and parameter estimation in a spiking neuron based on quadratic detectability.&lt;/span&gt;
Figure 28. State and parameter estimation in a spiking neuron based on quadratic detectability.

3.3.3.2 Observer Design Based on Uniform Backward Distinguishability

The idea of this section is to replace the LMI-based design of the observer for \( \xi_{no}\) in system (245) with a KKL-based one, updating the gains \( K_d\) and \( L_d\) dynamically along the solution, instead of solving (249). To do that, we rely on the KKL observer design proposed in Section 2.3 applied to the equivalent discrete-time system (247).

3.3.3.2.1 Discrete-Time KKL Observer Design for System (247)

Consider the discrete-time system (247) for a given sequence \( (\xi_{o,k},u_{d,k},\tau_k)_{k\in \mathbb{N}}\) . Following the spirit of the KKL observer, we look for a transformation \( (T_k)_{k \in \mathbb{N}}\) such that, in the new coordinates \( \zeta_k:= T_k \xi_{no,k} \in \mathbb{R}^{n_\zeta}\) (with \( n_\zeta \in \mathbb{N}\) to be defined later), system (247) follows the dynamics

\[ \begin{equation} \zeta_{k+1} = \gamma A \zeta_k + B y_k, \end{equation} \]

(257)

where \( A \in \mathbb{R}^{n_\zeta \times n_\zeta}\) is Schur, \( B \in \mathbb{R}^{n_\zeta \times n_{y,d,\rm{ext}}}\) (recall that \( n_{y,d,\rm{ext}} = n_{y,d} + n_o\) ) is such that the pair \( (A,B)\) is controllable, and \( \gamma \in (0,1]\) is a design parameter. Then, it follows that the transformation \( (T_k)_{k \in \mathbb{N}}\) must be such that for the given sequence \( (\xi_{o,k},u_{d,k},\tau_k)_{k \in \mathbb{N}}\) ,

\[ \begin{equation} T_{k+1} \mathcal{J}_k = \gamma AT_k + B\mathcal{H}_k, \end{equation} \]

(258)

where we use the abbreviation \( \mathcal{J}_k = \mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k)\) and \( \mathcal{H}_k = \begin{pmatrix} \mathcal{H}_{dno}(\xi_{o,k},u_{d,k},\tau_k) \\ \mathcal{J}_{ono}(\xi_{o,k},u_{d,k},\tau_k)\end{pmatrix} \) . The interest of this form is that the observer for system (257) in the \( \zeta\) -coordinates is a simple filter of the output, namely,

\[ \begin{equation} \hat{\zeta}_{k+1} = \gamma A \hat{\zeta}_k + B y_k, \end{equation} \]

(259)

making the estimation error \( \tilde{\zeta}_k = \zeta_k - \hat{\zeta}_k\) verify

\[ \begin{equation} \tilde{\zeta}_{k+1} = \gamma A \tilde{\zeta}_k, \end{equation} \]

(260)

and thus exponentially stable because \( \gamma A\) is Schur. Then, if \( (T_k)_{k \in \mathbb{N}}\) is uniformly left-invertible after some discrete time \( k^\star \in \mathbb{N}\) , the estimate defined by \( \hat{\xi}_{no,k} = T_k^* \hat{\zeta}_k\) , where \( (T^*_k)_{k \in \mathbb{N}}\) is a bounded sequence of left inverses of \( (T_k)_{k \in \mathbb{N}}\) verifying \( T^*_k T_k = \rm{Id}\) for all \( k \in \mathbb{N}_{\geq k^\star}\) , is such that the estimation error \( \xi_{no,k} - \hat{\xi}_{no,k}\) is exponentially stable and converges to zero after \( k^\star\) . From Corollary 1, we know that this is possible when the following property holds.

Definition 20 (Uniform backward distinguishability of system (247))

System (247), fed with \( (\xi_{o,k},u_{d,k},\tau_k)_{k \in \mathbb{N}}\) such that the matrix \( \mathcal{J}_k\) is invertible for all \( k\in \mathbb{N}\) , is uniformly backward distinguishable if, there exists \( \alpha>0\) and for each \( i \in \{1, 2, …, n_{y,d,\rm{ext}}\}\) , there exists \( m_i \in \mathbb{N}_{>0}\) such that, for all \( k \in \mathbb{N}_{\geq \overline{m}}\) where \( \overline{m} := \max_{i \in \{1,2,…,n_{d, \rm{ext}}\}} m_i\) , the backward distinguishability matrix sequence \( (\mathcal{O}^{bw}_k)_{k\in\mathbb{N}}\) defined as

\[ \begin{equation} \mathcal{O}^{bw}_k = ( \mathcal{O}^{bw}_{1,k}, \mathcal{O}^{bw}_{2,k}, \ldots, \mathcal{O}^{bw}_{n_{y,d,\rm ext},k} ) \in \mathbb{R}^{\left(\sum_{i=1}^{n_{y,d,\rm ext}} m_i\right) \times n_{no}}, \end{equation} \]

(261.a)

where

\[ \begin{equation} \mathcal{O}^{bw}_{i,k} := \begin{pmatrix*}[l] \mathcal{H}_{i,k-1} \mathcal{J}_{k-1}^{-1} \\ \mathcal{H}_{i,k-2} \mathcal{J}_{k-2}^{-1}\mathcal{J}_{k-1}^{-1} \\ \ldots \\ \mathcal{H}_{i,k-(m_i-1)}\mathcal{J}_{k-(m_i-1)}^{-1}\ldots \mathcal{J}_{k-1}^{-1}\\ \mathcal{H}_{i,k-m_i} \mathcal{J}_{k-m_i}^{-1}\mathcal{J}_{k-(m_i-1)}^{-1}\ldots \mathcal{J}_{k-1}^{-1} \end{pmatrix*}, \end{equation} \]

(261.b)

where \( \mathcal{H}_{i,k}\) denotes the \( i^{\text{th}}\) row of \( \mathcal{H}_k\) , has full rank and satisfies \( (\mathcal{O}^{bw}_k)^\top \mathcal{O}^{bw}_k \geq \alpha \rm{Id} > 0\) .

The property in Definition 20 is equivalent to the fact that

\[ \begin{equation} \sum_{i = 1}^{n_{y,d,\rm{ext}}}\sum_{j = k-m_i}^{k-1} (\mathcal{J}^{-1}_{k-1})^\top (\mathcal{J}^{-1}_{k-2})^\top\ldots (\mathcal{J}^{-1}_j)^\top \mathcal{H}_{i,j}^\top \mathcal{H}_{i,j} \mathcal{J}^{-1}_j \ldots \mathcal{J}^{-1}_{k-2} \mathcal{J}^{-1}_{k-1}\\ \geq \alpha \rm{Id} > 0, \end{equation} \]

(262)

so that, when all observability indexes \( m_i\) are equal, it coincides with Kalman’s uniform complete observability (see [79, Condition (13)][43, Assumption 2-3], and [81, Definition 3]). This property can also be checked using the forward observability matrix as in Remark 10, which is much easier to compute, when all the \( m_i\) are the same and there exists \( c_{\mathcal{J}}>0\) such that \( (\mathcal{J}_k^{-1})^\top\mathcal{J}_k^{-1} \geq c_{\mathcal{J}} \rm{Id}\) for all \( k \in \mathbb{N}\) . Note that a discrete-time Kalman-like observer [43, 81] could seem like a possible alternative to the KKL one since it requires the same observability condition as we have just shown and exhibits a strict Lyapunov function. However, the gain that is multiplied with the fictitious output in the observer must be constant during flows for us to perform the analysis (similar to \( K_d\) in (249)), which is not the case in a Kalman-like observer (unless the pair \( (\mathcal{J}_{no}(\xi_o,u_d,\tau),\mathcal{H}_{dno}(\xi_o,u_d,\tau))\) at the jump times along the solution is UCO, without the need for the fictitious output). This constancy of the gain is ensured in the KKL-based design since it relies on a transformation into a time-invariant form with a chosen constant gain \( B_{ono}\) (see below in the proof of Theorem 17). But indeed, if UCO is guaranteed from \( y_d\) and without the fictitious output, one could use the Kalman-like observer [43, 81] in place of the KKL-based one for a more systematic design.

Lemma 13, which is a particular case of Theorem 4 and Theorem 5, then states the existence of \( (T_k)_{k\in\mathbb{N}}\) that is uniformly bounded and uniformly left-invertible after some time, assuming i) uniform invertibility of the matrix sequence \( (\mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k))_{k \in \mathbb{N}}\) , ii) uniform boundedness of \( (\mathcal{H}_{dno}(\xi_{o,k},u_{d,k},\tau_k))_{k \in \mathbb{N}}\) and \( (\mathcal{J}_{ono}(\xi_{o,k},u_{d,k},\tau_k))_{k \in \mathbb{N}}\) , and iii) uniformly backward distinguishability of the discrete-time system (247). Note that the injectivity of each \( T_k\) without uniformity in time can be obtained from (non-uniform) backward distinguishability conditions Section 2.3.5, which may suffice in some cases to ensure convergence of the KKL observer as seen in Example 8, but is not sufficient in general.

Lemma 13 (KKL transformation for system (247))

Consider \( c_{\mathcal{J}^{-1}_{no}} > 0\) , \( c_{\mathcal{H}_{dno}} > 0\) , \( c_{\mathcal{J}_{ono}} > 0\) , \( \alpha > 0\) , \( c_{T,0} > 0\) , and for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) , a positive integer \( m_i\) and a controllable pair \( (\tilde{A}_i,\tilde{B}_i) \in \mathbb{R}^{m_i \times m_i}\times\mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur. Define \( n_\zeta = \sum_{i=1}^{n_{y,d,\rm{ext}}}m_i\) . For any \( k^\star_m\in \mathbb{N}_{\geq \overline{m} + 1}\) where \( \overline{m} = \max_{i \in \{1,2,…,n_{d, \rm{ext}}\}} m_i\) , there exists \( 0<\gamma^\star \leq 1\) such that for any \( 0 < \gamma < \gamma^\star\) , there exist \( \underline{c}_T(\gamma)>0\) , \( \overline{c}_T(\gamma)>0\) , and an integer \( k^\star \leq k^\star_m\) such that for any \( T_0 \in \mathbb{R}^{n_\zeta \times n_{no}}\) verifying \( \|T_0\| \leq c_{T,0}\) and for any sequence \( (\xi_{o,k},u_{d,k},\tau_k)_{k \in \mathbb{N}}\) such that:

  1. The matrix \( \mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k)\) is invertible and \( \|(\mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k))^{-1}\| \leq c_{\mathcal{J}^{-1}_{no}}\) for all \( k \in \mathbb{N}\) ;
  2. \( \|\mathcal{H}_{dno}(\xi_{o,k},u_{d,k},\tau_k)\|\leq c_{\mathcal{H}_{dno}}\) and \( \|J_{ono}(\xi_{o,k},u_{d,k},\tau_k)\|\leq c_{\mathcal{J}_{ono}}\) for all \( k \in \mathbb{N}\) ;
  3. System (247) fed with \( (\xi_{o,k},u_{d,k},\tau_k)_{k \in \mathbb{N}}\) is uniformly backward distinguishable with parameters \( \alpha\) and \( m_i\) (as in Definition 20);

we have that the (unique) sequence \( (T_k)_{k\in\mathbb{N}}\) initialized as \( T_0\) and satisfying (258) with

\[ \begin{align} A &= \rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_{y,d,\rm ext}}) \in \mathbb{R}^{n_\zeta \times n_\zeta}, \\ B& = \rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_{y,d,\rm ext}}) \in \mathbb{R}^{n_\zeta \times n_{y,d,\rm ext}}, \\\end{align} \]

(263.a)

verifies \( \|T_k\| \leq \overline{c}_T(\gamma)\) for all \( k\in \mathbb{N}\) and \( T_k ^\top T_k \geq (\underline{c}_T(\gamma))^2 \rm{Id}\) for all \( k\in \mathbb{N}_{\geq k^\star}\) .

Proof. \( (T_k)_{k\in\mathbb{N}}\) is defined uniquely from \( T_0\) by (258) and the assumed invertibility of \( \mathcal{J}_{no}(\xi_{o,k},u_{d,k},\tau_k)\) . The rest is a particular case of Theorem 4 and Theorem 5. \( \blacksquare\)

Next, in Section 3.3.3.2.2, we exploit the results from this section for system (245).

3.3.3.2.2 KKL-Based Observer Design for System (245)

Inspired by the developments in the previous section, we make the following assumption.

Assumption 21

Given \( j_m\) defined in Item (XIV) of Assumption 17, assume that there exist \( c_{\mathcal{J}^{-1}_{no}} > 0\) , \( \alpha > 0\) , and \( m_i \in \mathbb{N}_{>0}\) for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) , such that along every complete solution \( \xi \in §_\mathcal{H}(\Xi_0, \mathfrak{U}_c \times \mathfrak{U}_d)\) , the sequences \( (\xi_{o,j},u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) such that \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) for each \( j \in \mathbb{N}_{\geq j_m}\) , we have:

  1. The matrix sequence \( (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))_{j \in \mathbb{N}_{\geq j_m}}\) is uniformly invertible, i.e., we have that \( \|(\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1}\| \leq c_{\mathcal{J}^{-1}_{no}}\) for all \( j \in \mathbb{N}_{\geq j_m}\) ;
  2. System (247) fed with \( (\xi_{o,j},u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) is uniformly backward distinguishable with parameters \( \alpha\) and \( m_i\) for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) .

Remark 35

Contrary to discrete-time systems obtained from discretizing a physical system, the jump map of a hybrid system may not be invertible since it is not a discretization of some continuous-time dynamics. However, here \( \mathcal{J}_{no}\) combines flow and jump dynamics, i.e., models the dynamics of the hybrid system sampled at jumps. Thus, it is reasonable to expect its invertibility along solutions. Besides, in many applications of jump parameter\( /\) uncertainty estimation (see in the examples along this chapter and in Section 3.4), \( \xi_{no}\) is constant making \( F_{no} = 0\) and \( J_{no}(\xi_o,u_d) = \rm{Id}\) , so \( \mathcal{J}_{no}(\xi_o,u_d,\tau) = \rm{Id}\) at all times. Note though that, while the invertibility of the dynamics is sufficient but typically not necessary for linear Kalman(-like) designs in [127] or Section 3.2.2 (this assumption is generally used for analysis only), it is here required to implement the observer (see the dynamics \( \hat{T}^+\) in observer (264) below), so it plays a much more crucial role.

A difficulty in exploiting the discrete-time KKL observer (259) is that the discrete-time output \( y_k\) is not fully available at jumps since it contains fictitious outputs. The KKL-based observer we propose for system (245) has the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{\xi}}_o &=& \hat{f}_{o,\ell}(\hat{\xi}_o, p, \tau, y_c, u_c) \\ \dot{\hat{\eta}} & = & 0\\ \dot{p}&=&\varphi_{c,\ell}(p, \tau, y_c, u_c) \\ \dot{\hat{T}} & = & 0 \\ \dot{\tau} &=& 1 \end{array}\right\} \text{when}~(245) \text{ flows}\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{\xi}_o^+ &=& J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no}) \\ \hat{\eta}^+ & = & \gamma A \hat{\eta}+ \gamma A B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau) \\ &&{}+ \gamma A \hat{T} \displaystyle \int_0^{\tau}e^{-F_{no}s}U_{cno}ds+B_{dno} \left(y_d - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)\right)\\ &&{}+\hat{T}^+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)- B_{ono}J_o(\rm{sat}_o(\hat{\xi}_o),u_d) \\ p^+ & = & \varphi_{d,\ell}(p, \tau, y_d, u_d)\\ \hat{T}^+ & = & \big(\gamma A\hat{T} + B_{dno}\mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\\ &&{}+ B_{ono}\mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\big)\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)) \\ \tau^+ &=& 0 \end{array} \right\} \text{when}~(245) \text{ jumps} \end{array} \right. \end{equation} \]

(264.a)

with

\[ \begin{equation} \hat{\xi}_{no} = e^{F_{no}\tau} \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})\left(\hat{\eta} + B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\right) + \int_0^\tau e^{F_{no}(\tau-s)}U_{cno}ds, \end{equation} \]

(264.b)

with \( \hat{f}_{o,\ell}\) being a high-gain observer as described in (248), \( (\mathcal{J}_{no},\mathcal{H}_{dno},\mathcal{J}_{ono})\) defined in system (247), \( \rm{sat}_o\) being a bounded map such that \( \rm{sat}_o(\hat{\xi}_o) = \hat{\xi}_o\) for all \( \hat{\xi}_o \in (\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}\) for some \( \overline{c}_o > 0\) to be designed, \( \rm{sat}_{no}\) and \( f_{o,\rm{sat}}\) fixed as in observer (248), \( \ell > 0\) , \( \gamma \in (0,1]\) and \( (A,B_{dno},B_{ono})\in \mathbb{R}^{n_\eta \times n_\eta} \times \mathbb{R}^{n_\eta \times n_{y,d}} \times \mathbb{R}^{n_\eta \times n_o}\) (for some dimension \( n_\eta \in \mathbb{N}\) to be defined later) being design parameters to be chosen, \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}\) and \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) being extended left inverse maps defined as

\[ \begin{equation} M \mapsto \rm{inv}_{\underline{c}_M}(M) = \left\{\begin{array}{@{}l@{~~}l@{}} M^\dagger & \text{ if } \sigma_{\min}(M) \geq \underline{c}_M \\ \frac{\sigma_{\min}(M)}{\underline{c}_M}M^\dagger & \text{ if } 0 < \sigma_{\min}(M) \leq \underline{c}_M \\ 0 & \text{ if } \sigma_{\min}(M) = 0, \end{array}\right. \end{equation} \]

(264.c)

for some fixed saturation level \( \underline{c}_{\hat{\mathcal{J}}_{no}} < \frac{1}{ c_{\mathcal{J}^{-1}_{no}}}\) with \( c_{\mathcal{J}^{-1}_{no}}\) given in Item (XXI) of Assumption 21 and some saturation level \( \underline{c}_{T}(\gamma)\) to be designed. Those dynamics are picked so that, modulo some estimation errors on \( \xi_o\) , \( \hat{T}\) coincides at jumps with \( (T_k)_{k\in\mathbb{N}}\) studied in Section 3.3.3.2.1 and the corresponding discrete-time KKL error dynamics (260) appear in some way after a certain change of coordinates.

Theorem 17 (Asymptotic convergence of the KKL-based observer (264))

Suppose Assumptions 171819, and 21 hold. Define \( n_\eta = n_\zeta = \sum_{i=1}^{n_{y,d,\rm{ext}}}m_i\) (with \( m_i\) coming from Item (XXII) of Assumption 21) and consider, for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) , a controllable pair \( (\tilde{A}_i,\tilde{B}_i) \in \mathbb{R}^{m_i \times m_i}\times\mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur. Let

\[ \begin{align} A &= \rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_{y,d,\rm ext}}) \in \mathbb{R}^{n_\eta \times n_\eta}, \\ B_{dno}& = \rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_{y,d}}) \in \mathbb{R}^{n_\eta \times n_{y,d}},\\ B_{ono}& =\rm{diag}(\tilde{B}_{n_{y,d}+1}, \tilde{B}_{n_{y,d}+2}, \ldots, \tilde{B}_{n_{y,d,\rm ext}}) \in \mathbb{R}^{n_\eta \times n_o}. \\\end{align} \]

(265.a)

There exists \( \overline{c}_o>0\) in the definition of \( \rm{sat}_o\) such that, given any \( T_0\in \mathbb{R}^{n_\eta \times n_{no}}\) , there exist \( 0<\gamma^\star \leq 1\) and \( \ell^\star>0\) such that for any \( 0 < \gamma < \gamma^\star\) and for any \( \ell > \ell^\star\) , there exists \( \underline{c}_{T}(\gamma)>0\) such that any maximal solution to the cascade (245)-(264) initialized in \( \Xi_0 \times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{T_0\}\times\{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) , is complete and verifies (236).

Remark 36

Note that in this KKL-based design, contrary to the LMI-based one in Theorem 16, the stability of the estimation error with respect to its initial conditions cannot be stated. This is because the KKL estimate \( \hat{\xi}_{no}\) is fed back to \( \hat{\xi}_o^+\) which will then enter \( \hat{T}^+\) and eventually \( \hat{\eta}^+\) : even if the estimates are initialized exactly at the right states, the non-injectivity of \( \hat{T}\) before a certain time could still trigger an error in \( \hat{\xi}_{no}\) , which would propagate in the whole estimate. However, the asymptotic convergence in (236) still guarantees the estimation of \( \xi\) . Actually, exponential stability of \( (\xi,T)\) , with any arbitrarily fast convergence rate, can be achieved by selecting \( \ell\) large and \( \gamma\) small, but only after a certain number of jumps and with respect to the estimation error at that later time—see (455) below.

Proof. This highly technical proof has been moved to Section 6.2.2.3 to facilitate reading. It involves three main steps: i) Choose \( \overline{c}_o\) (for the definition of \( \rm{sat}_o\) ), \( \gamma^\star\) , \( \ell^\star\) , and \( \underline{c}_{T}(\gamma)\) allowing us to guarantee some preliminary bounds and injectivity properties on the maps and variables, ii) Define a change of coordinates for the \( \xi_{no}\) state into some target \( \eta\) -coordinates, and obtain the state dynamics in those new coordinates (along solutions), and iii) Define a Lyapunov function and apply Theorem 24 to obtain exponential stability in the new coordinates after a certain time and retrieve asymptotic convergence in the \( \xi\) -coordinates. \( \blacksquare\)

Example 26 (KKL-based observer design for the spiking neuron)

Consider the spiking neuron in Example 21, with the choice of \( \xi_o\) and \( \xi_{no}\) as in Example 23. Let us now design observer (264) for this system, exploiting the affine structure (245) detailed in Example 24. Since \( J_{no}(\xi_o) = 1\) and \( F_{no} = 0\) , \( \mathcal{J}_{no}(\xi_o,\tau) = 1\) is always uniformly invertible. The pair \( (\mathcal{J}_{no}(\xi_o),\mathcal{J}_{ono}(\xi_o))=(J_{no}(\xi_o),J_{ono}(\xi_o))\) is uniformly backward distinguishable with \( m = 1\) , so \( \eta\) is of dimension one. Realizing that we only need the second line of \( \mathcal{J}_{ono}(\xi_o)\) for distinguishability, we then choose \( A = 0.5\) which is Schur, an empty \( B_{dno}\) , and \( B_{ono} = \begin{pmatrix} 0 & 1 & 0 & 0 \end{pmatrix}\) making the pair \( (A,B_{ono})\) controllable. Note that since \( \mathcal{J}_{no}\) and \( B_{ono}\mathcal{J}_{ono}\) are independent of \( \xi_o\) , \( T\) and \( \hat{T}\) (see in the proof of Theorem 17) are the same in this case. This choice also means that, as in the previous section, we choose to estimate \( \xi_{no}\) only via its interaction with the second component of \( \xi_o\) at jumps. Take \( \gamma = 0.001\) to ensure a fast convergence after the first jump. We take the same \( K\) and \( \ell\) as in Example 25 for the high-gain flow-based observer (240). Like in Example 25, we approximate the term \( \Psi_{f_{o,\rm{sat}}}(\hat{\xi}_o,t,-\tau)\) in (264.b) using a Euler scheme with time step \( 0.001\) (s). Simulation results for the states \( x = (s_1,s_2,a,b,d)\) and the estimation errors \( (\tilde{s}_1,\tilde{s}_2,\tilde{a},\tilde{b},\tilde{d})\) are shown in Figure 29.

&lt;span data-controller=&quot;mathjax&quot;&gt;State and parameter estimation in a spiking neuron based on uniform backward distinguishability.&lt;/span&gt;
Figure 29. State and parameter estimation in a spiking neuron based on uniform backward distinguishability.

Table 4 then summarizes and compares the two designs proposed in this section. Compared to the LMI-based observer, the KKL-based one seems to require stronger conditions, but it is more systematic, namely, we can implement it without checking observability, and it provides arbitrarily fast convergence (after a certain time).

Table 4. Comparison of observer designs for hybrid systems (245) with affine structures and known jump times. All the conditions here are sufficient conditions.
LMI-based observer (248) (Section 3.3.3.1)KKL-based observer (264) (Section 3.3.3.2)
Time domainPersistence of both flows and jumps (Item (XIV) of Assumption 17)
Hybrid modelNo assumptionInvertibility of the matrices \( (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))_{j \in \mathbb{N}_{\geq j_m}}\) along solutions (Item (XXI) of Assumption 21)
ObservabilityInstantaneous observability of \( \xi_o\) and a form of quadratic detectability of system (247) (Assumption 20)Instantaneous observability of \( \xi_o\) and uniform backward distinguishability of system (247) along solutions (Item (XXII) of Assumption 21)
State dimension\( n_\xi + n_p +1\)\( n_o + n_p + \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right) + \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right)n_{no} + 1\)
Convergence rateDetermined by the solution to a matrix inequalityArbitrarily fast, achieved after a certain time

In Section 3.4.4, the LMI-based and KKL-based observers proposed in this section will be applied for estimating state and uncertainties in a bipedal walking robot.

3.3.4 Observer Design for System (232)

In this part, we propose an observer design for system (232) with fully nonlinear maps instead of being affine in \( \xi_{no}\) as in system (245). For that, we make the next assumption on the maps of system (232).

Assumption 22

The maps \( f = (f_o,f_{no})\) , \( g=(g_o,g_{no})\) , and \( h_d\) are locally Lipschitz with respect to \( \xi\) , uniformly in \( (u_c,u_d) \in \mathcal{U}_c\times\mathcal{U}_d\) (see in Assumption 19 for explanations). The maps \( g\) and \( h_d\) are locally bounded with respect to \( \xi\) , uniformly in \( u_d \in \mathcal{U}_d\) (see in Assumption 19 for explanations).

3.3.4.1 Conditions for the Jump-Based Observer Estimating \( \xi_{no}\)

Still supposing that \( \xi_o\) is estimated from \( y_c\) thanks to a high-gain flow-based observer satisfying the conditions in Assumption 18, we now study the conditions that the observer estimating \( \xi_{no}\) (from \( y_d\) and the fictitious output) should satisfy to constitute a full-state observer.

Similar to what has been done above with system (247), inspired from [22, Section V], we start by considering an equivalent discrete-time system, describing the evolution of the \( \xi_{no}\) component of solutions to system (232) discretized after each jump, namely, the dynamics of \( \xi_{no,k}=\xi_{no}(t_k,k)\) for \( k\in \mathbb{N}\) . The computation of \( \xi_{no,k+1}\) as a function of \( \xi_{no,k}\) depends on:

  • The component \( \xi_o\) of the solution over the interval \( [t_k,t_{k+1}]\) ;
  • The continuous-time input \( \mathfrak{u}_c\) over the interval \( [t_k,t_{k+1}]\) ;
  • The discrete-time input \( \mathfrak{u}_d\) used at the jump at time \( t_{k+1}\) , namely \( u_{d,k}=\mathfrak{u}_d(k)\) ;
  • The time \( t_k\) (because of the inputs making this dependence non-autonomous);
  • The flow length \( \tau_k = t_{k+1}-t_k\) .

Because the computation of \( \xi_{no,k+1}\) in the hybrid system takes place at hybrid time \( (t_{k+1},k)\) , i.e., before a jump, \( \tau_k\) is available in \( \tau(t_{k+1},k)\) , and the available estimate of \( \xi_o\) is \( \hat{\xi}_o(t_{k+1},k)\) , which is the outcome of the high-gain flow-based observer over the period \( [t_k,t_{k+1}]\) . This estimate has the merit of being much more reliable than the initial estimate \( \hat{\xi}_o(t_{k},k)\) , which is impacted by the estimation error of \( \hat{\xi}_{no}(t_k,k-1)\) . Indeed, the estimation error on \( \hat{\xi}_o(t_{k+1},k)\) can be made arbitrarily small by pushing the high-gain parameter. For those reasons, we choose to parameterize the dependence of \( \xi_{no,k+1}\) in \( \xi_o\) via its value before the jump, i.e., \( \xi_{o,k}=\xi_{o}(t_{k+1},k)\) . All this leads us to considering the discrete-time system

\[ \begin{equation} \xi_{no,k+1} = \mathcal{G}_k(\xi_{no,k}), ~~ y_k = \mathcal{H}_k(\xi_{no,k}), \end{equation} \]

(266.a)

with the dynamics map \( \mathcal{G}_k:\mathbb{R}^{n_{no}} \to \mathbb{R}^{n_{no}} \) defined as

\[ \begin{equation} \mathcal{G}_k(\xi_{no}) := g_{no}(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_{o,k},t_k+\tau_k,-\tau_k),\xi_{no}),t_k,\tau_k),u_{d,k}), \end{equation} \]

(266.b)

and the extended output map \( \mathcal{H}_k: \mathbb{R}^{n_{no}} \to \mathbb{R}^{n_{y,d,\rm{ext}}}\) defined as

\[ \begin{equation} \mathcal{H}_k(\xi_{no}) =(\mathcal{H}_{d,k}(\xi_{no}),\mathcal{H}_{o,k}(\xi_{no})), \end{equation} \]

(266.c)

with \( n_{y,d,\rm{ext}} = n_{y,d} + n_o\) , where \( \mathcal{H}_{d}\) models the output \( y_d\) available at jumps while \( \mathcal{H}_{o}\) models the fictitious output \( \xi_o^+\) , i.e., \( \xi_o(t_{k+1},k+1)\) that becomes “visible” during the next interval of flow through \( y_c\) like in system (247), namely

\[ \begin{align} \mathcal{H}_{d,k}(\xi_{no}) &:= h_d(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_{o,k},t_k+\tau_k,-\tau_k), \xi_{no}),t_k,\tau_k),u_{d,k}),\\ \mathcal{H}_{o,k}(\xi_{no}) &:= g_o(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_{o,k},t_k+\tau_k,-\tau_k), \xi_{no}),t_k,\tau_k),u_{d,k}), \\\end{align} \]

(266.d)

with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (u_k)_{k\in \mathbb{N}}\) where \( u_k :=(\xi_{o,k},u_{d,k},t_k,\tau_k) \in (\Xi_o \cap D_o) \times \mathcal{U}_d \times \mathbb{R}_{\geq 0} \times \mathcal{I} \subset \mathbb{R}^{n_{u,d,\rm{ ext }}}\) where \( n_{u,d,\rm{ ext }} := n_o+n_{u,d}+1+1\) for all \( k \in \mathbb{N}\) . As illustrated in Figure 30, system (266) is formed by looking at the dynamics of \( \xi_{no}(t_k,k)\) after each jump, with \( \xi_{no}(t_{k+1},k+1)\) computed from \( \xi_{no}(t_{k},k)\) (red arrows) based on the value of \( (\xi_o,u_d,\tau)\) right before the jump and of \( t_k\) (in blue). Therefore, \( \mathcal{G}_k\) and \( \mathcal{H}_k\) are obtained by flowing backward to the beginning of the flow interval using \( f_{o,\rm{sat}}\) to compute \( \xi_o(t_k,k)\) from the input \( \xi_{o,k}=\xi_o(t_{k+1},k)\) , and then, flowing forward with \( f(\cdot,\mathfrak{u}_c)\) from the obtained \( \xi_o(t_k,k)\) and the discrete state \( \xi_{no,k}=\xi_{no}(t_k,k)\) to recover the value of \( (\xi_o,\xi_{no})\) before the jump.

&lt;span data-controller=&quot;mathjax&quot;&gt;Illustration of the equivalent discrete-time system&amp;#160;(eq:ch9_sys_dis_fn). The state is colored in red while the inputs are in blue.&lt;/span&gt;
Figure 30. Illustration of the equivalent discrete-time system (266). The state is colored in red while the inputs are in blue.

Remark 37

In the case where \( f_{no}\) is identically zero, i.e., \( \xi_{no}\) is constant during flows, the equivalent discrete-time system (266) is simplified to

\[ \begin{equation} \xi_{no,k+1} = g_{no}(\xi_{o,k},\xi_{no,k},u_{d,k}), ~~ y_k = \begin{pmatrix} h_d(\xi_{o,k}, \xi_{no,k},u_{d,k}) \\ g_o(\xi_{o,k},\xi_{no,k},u_{d,k}) \end{pmatrix}, \end{equation} \]

(267)

with inputs \( (\xi_{o,k},u_{d,k})_{k\in \mathbb{N}}\) .

From Lemma 12, the possibility of designing an observer for system (232) is linked to the observability analysis of system (266). Similarly to how system (247) is used to design observers for system (245) in the previous sections, we assume that a discrete-time observer is available for system (266) in order to build an observer for system (232). More precisely, inspired by MHO[97, 98, 99, 96, 50] and KKL observers as in [102] or Section 2.3, we assume that solutions to system (266) can be expressed, at least after a certain time, as a function of solutions to a system of the form

\[ \begin{equation} \zeta_{k+1} = A \zeta_k + B y_k, \end{equation} \]

(268)

with \( A \in \mathbb{R}^{n_\zeta \times n_\zeta}\) Schur and \( B \in \mathbb{R}^{n_\zeta \times n_{y,d,\rm{ext}}}\) . Indeed, a discrete-time observer is then derived by running the dynamics (268) from any initial condition, as detailed next. In order to remain as little conservative as possible, but still take into account possible errors in the estimation of \( \xi_o\) feeding system (266), we only assume this for the input \( \mathfrak{u}_c \in \mathfrak{U}_c\) and input sequences \( (u_k)_{k \in \mathbb{N}}\) encountered in system (232), modulo some arbitrarily small estimation errors on \( \xi_{o,k}\) . All in all, we will state a general observer design for \( \xi_{no}\) in system (232), starting with some assumptions on the equivalent discrete-time system (266).

Assumption 23

Given \( j_m\in \mathbb{N}\) and \( \Xi_{no}\subset \mathbb{R}^{no}\) defined in Item (XIV) of Assumption 17, assume that:

  1. Expression of solutions to system (266) as function of solutions to system (268) after a certain time: There exist \( n_\zeta \in \mathbb{N}\) , \( A \in \mathbb{R}^{n_\zeta \times n_\zeta}\) Schur, \( B \in \mathbb{R}^{n_\zeta \times n_{y,d,\rm{ext}}}\) , \( j_m^\star \in \mathbb{N}_{\geq j_m}\) , \( \overline{c}_o > 0\) , and a compact set \( \Xi_{\zeta} \subset \mathbb{R}^{n_\zeta}\) , such that for any \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c\times \mathfrak{U}_d\) and for any \( (\hat{\xi}_{o,j},t_j,\tau_j)_{j \in \mathbb{N}}\) for which there exists a complete solution \( \xi \in §_\mathcal{H}(\Xi_0,(\mathfrak{u}_c,\mathfrak{u}_d))\) with jump times \( (t_j)_{j \in \mathbb{N}}\) verifying

    \[ \begin{equation} \tau_j = t_{j+1} - t_j, ~~ |\xi_o(t_{j+1},j) - \hat{\xi}_{o,j}| \leq \overline{c}_o, ~~ \forall j \in \mathbb{N}_{\geq j_m}, \end{equation} \]

    (269)

    there exists a sequence \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) of maps \( \hat{\mathcal{T}}_j:\mathbb{R}^{n_\zeta}\to \mathbb{R}^{n_{no}}\) such that any solution \( j \mapsto \xi_{no,j}\) to system (266) remaining in \( \Xi_{no}\) and fed with \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j \in \mathbb{N}} :=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) (where \( u_{d,j} = \mathfrak{u}_d(j)\) for all \( j \in \mathbb{N}\) ) can be written as

    \[ \begin{equation} \xi_{no,j} = \hat{\mathcal{T}}_j(\zeta_j), ~~ j \in \mathbb{N}_{\geq j_m^\star}, \end{equation} \]

    (270)

    for some solution \( j \mapsto \zeta_j\) to the dynamics (268) remaining in \( \Xi_{\zeta}\) for all \( j \in \mathbb{N}\) ;

  2. Uniform Lipschitzness of the function sequence: There exists \( c_{\hat{\mathcal{T}}} > 0\) such that, each map sequence \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) considered in the previous item is uniformly Lipschitz on \( \mathbb{R}^{n_\zeta}\) with constant \( c_{\hat{\mathcal{T}}}\) , i.e., for all \( j \in \mathbb{N}_{\geq j_m^\star}\) and for all \( (\zeta_a,\zeta_b) \in \mathbb{R}^{n_\zeta} \times \mathbb{R}^{n_\zeta}\) ,

    \[ \begin{equation} |\hat{\mathcal{T}}_j(\zeta_a) - \hat{\mathcal{T}}_j(\zeta_b)| \leq c_{\hat{\mathcal{T}}}|\zeta_a - \zeta_b|; \end{equation} \]

    (271)

  3. Dependence of \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) on the history of \( (\hat{\xi}_{o,j})_{j\in \mathbb{N}}\) along solutions: There exist \( n_\Gamma \in \mathbb{N}\) , \( A^\prime \in \mathbb{R}^{n_\Gamma \times n_\Gamma}\) Schur, \( B^\prime \in \mathbb{R}^{n_\Gamma}\) , \( c_\Gamma > 0\) , such that each map \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) considered in Item (XXIII) verifies

    \[ \begin{equation} |\mathcal{T}_j(\zeta_j) - \hat{\mathcal{T}}_j(\zeta_j)| \leq c_\Gamma |\hat{\Gamma}_j|, ~~ \forall j \in \mathbb{N}_{\geq j_m^\star}, \end{equation} \]

    (272)

    where \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) and \( j \mapsto \zeta_j\) are the map sequence and solutions to system (268) considered in Item (XXIII) when feeding system (266) with \( (u_j)_{j \in \mathbb{N}}:=(\xi_o(t_{j+1},j),u_{d,j}, t_j,\tau_j)_{j\in \mathbb{N}}\) instead of \( (\hat{u}_j)_{j \in \mathbb{N}}\) , and where \( j \mapsto \hat{\Gamma}_j \in \mathbb{R}^{n_\Gamma}\) is solution to

    \[ \begin{equation} \hat{\Gamma}_{j+1} = A^\prime \hat{\Gamma}_j + B^\prime |\xi_o(t_{j+1},j)-\hat{\xi}_{o,j}|. \end{equation} \]

    (273)

Remark 38

Under Item (XXIII) and Item (XXIV) of Assumption 23, given any inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_k)_{k \in \mathbb{N}}\) as in Item (XXIII), an observer for the discrete-time system (266), fed by those inputs, can be obtained by implementing from any initial conditions the filter

\[ \begin{equation} \hat{\zeta}_{k+1} = A \hat{\zeta}_k + B y_k, \end{equation} \]

(274.a)

and computing, at least after a certain time, the estimate

\[ \begin{equation} \hat{\xi}_{no,k} = \hat{\mathcal{T}}_k(\hat{\zeta}_k), ~~ k \in \mathbb{N}_{\geq j^\star}. \end{equation} \]

(274.b)

Indeed, \( \zeta_{k} - \hat{\zeta}_{k}\) given by (268) and (274.a) is exponentially stable and the estimation error \( \xi_{no,k} - \hat{\xi}_{no,k}\) asymptotically converges to zero according to (271) in view of (270) and (274.b). In the next section, Item (XXV) of Assumption 23 will be used to combine, in the hybrid setting of system (232), a jump-based observer for \( \xi_{no}\) , based on (274), with the high-gain flow-based observer for \( \xi_o\) . Indeed, the role of Item (XXV) is to take into account the fact that the real system follows (266) fed with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (u_k)_{k \in \mathbb{N}}\) , while in the observer, only estimates of \( (\xi_{o,k})_{k \in \mathbb{N}}\) are available: it thus quantifies the error made on \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) when using \( (\hat{\xi}_{o,k})_{k \in \mathbb{N}}\) instead of \( (\xi_{o,k})_{k \in \mathbb{N}}\) as input. In particular, it says that \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) may depend on the history of the inputs, but in an ISS way. Thus, Assumption 23 gives general conditions for the part of the observer used to estimate \( \xi_{no}\) .

Notice that both the MHO [97, 98, 99, 96, 50] and the KKL observers [102] or in Section 2.3 fall into the framework of Assumption 23. Indeed, for the former, the observer in the \( \zeta\) -coordinates is a simple storing of the past \( n_\zeta\) outputs with

\[ \begin{equation} A = \begin{pmatrix} 0 & 0 & \ldots & 0 &0\\ 1 & 0 & \ldots & 0 &0\\ \ldots & \ldots & \ldots & \ldots&\ldots\\ 0 & \ldots & 1 & 0 & 0\\ 0 & \ldots & 0 & 1 & 0 \end{pmatrix} \otimes \rm{Id}_{n_{y,d,{\rm ext }}}, ~~ B = \begin{pmatrix} 1 \\ 0 \\ \ldots\\ 0 \\ 0 \end{pmatrix}\otimes\rm{Id}_{n_{y,d,{\rm ext }}}, \end{equation} \]

(275)

and \( (\hat{\mathcal{T}}_k)_{k \in \mathbb{N}_{\geq j_m^\star}}\) represents the nonlinear optimization solvers estimating \( \xi_{no}\) from the stored outputs \( \hat{\zeta}\) , after a certain time. This relies on a so-called constructibility property, which is lighter than the observability frequently encountered in the MHO literature—see Section 2.2. The way MHOs may satisfy Assumption 23 is studied in Section 3.3.4.3. On the other hand, for other controllable pairs \( (A,B)\) without any particular structure, we directly recover the paradigm of KKL observers, where \( (\hat{\mathcal{T}}_k)_{k \in \mathbb{N}_{\geq j^\star}}\) denotes the left inverse of the transformation mapping solutions to system (266) into solutions to system (268), with dynamics picked sufficiently fast to ensure the injectivity of this transformation after a certain time. This is done under a so-called backward distinguishability condition. The way KKL observers may satisfy Assumption 23 is studied in Section 3.3.4.4. Note that in the MHO setting, the dependence of each \( \hat{\mathcal{T}}_k\) on the inputs \( \mathfrak{u}_c\) and \( (\hat{u}_k)_{k \in \mathbb{N}}\) is only within a moving window of size \( n_\zeta\) , while in the KKL setting, each \( \hat{\mathcal{T}}_k\) depends (implicitly) on the full past of the inputs.

Remark 39

The variable \( \hat{\Gamma}_j\) characterizes the closeness of the map sequences \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) and \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) when the inputs \( (\xi_{o,j})_{j \in \mathbb{N}}\) and \( (\hat{\xi}_{o,j})_{j \in \mathbb{N}}\) are close. Therefore, it plays the role of analysis in the Lyapunov proof, similar to \( T - \hat{T}\) in the proof of Theorem 17, and is not used for observer design. This analysis is needed here because, in this fully nonlinear context, the link between \( \xi_{no}\) and the new coordinates \( \zeta\) is via a sequence of nonlinear maps \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) instead of a linear matrix like \( \hat{T}\) in observer (264) that could be directly integrated into the state of the observer and in the Lyapunov function (451).

3.3.4.2 Hybrid Observer Design for System (232)

Given \( n_\zeta \in \mathbb{N}\) and \( (A,B) \in \mathbb{R}^{n_\zeta \times n_\zeta}\times \mathbb{R}^{n_\zeta \times n_{y,d,\rm{ext}}}\) in Assumption 23, and denoting \( B_d\) the first \( n_{y,d}\) columns of \( B\) and \( B_o\) the other \( n_o\) columns so that \( B = \begin{pmatrix}B_d & B_o\end{pmatrix}\) , consider for system (232) the observer

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{\xi}}_o &= & \hat{f}_{o,\ell}(\hat{\xi}_o,p, \tau,y_c,u_c)\\ \dot{\hat{\eta}} & = &0 \\ \dot{p}&=&\varphi_{c,\ell}(p, \tau, y_c,u_c)\\ \dot{\tau}&=&1 \end{array}\right\} \text{when}~(232) \text{ flows}\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{\xi}_o^+&=&g_o(\rm{sat}_o(\hat{\xi}_o),\rm{sat}_{no}(\hat{\xi}_{no}),u_d)\\ \hat{\eta}^+&=& A \hat{\eta} + B_dy_d+ AB_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\\ p^+&=&\varphi_{d,\ell}(p, \tau, y_d,u_d)\\ \tau^+&=&0 \end{array} \right\} \text{when}~(232) \text{ jumps} \end{array} \right. \end{equation} \]

(276.a)

with

\[ \begin{equation} \hat{\xi}_{no}= \Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c),no}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau),\hat{\mathcal{T}}_j(\hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))),t_j,\tau), \end{equation} \]

(276.b)

with \( \hat{f}_{o,\ell}\) being a high-gain observer as described in (248) with gain \( \ell > 0\) to be defined, \( \rm{sat}_o\) being a bounded map to be defined, \( \rm{sat}_{no}\) and \( f_{o,\rm{sat}}\) fixed as in observer (248), \( f_\rm{{sat}}\) being a map that is globally Lipschitz with respect to \( \xi\) , uniformly in \( u_c\in \mathcal{U}_c\) , and equal to \( f\) on \( \Xi\times \mathcal{U}_c\) (guaranteed to exist by Assumption 22 and [40, Corollary 1]), where \( \Xi := \{\xi=(\xi_o,\xi_{no})\in\mathbb{R}^{n_\xi}:\xi_o \in \Xi_o,\xi_{no}\in\Xi_{no}\}\) , \( \Psi_{f_\rm{{sat}}(\cdot,\mathfrak{u}_c),no}\) being the \( \xi_{no}\) -component of \( \Psi_{f_\rm{{sat}}(\cdot,\mathfrak{u}_c)}\) , \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) picked arbitrarily before \( j_m^\star\) and coming from Assumption 23 after \( j_m^\star\) with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j \in \mathbb{N}}=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) , where \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) for all \( j \in \mathbb{N}\) . An obstruction here arises from the fact that \( (\xi_{o,j})_{j \in \mathbb{N}}\) is unknown and we thus have to rely on \( (\hat{\xi}_{o,j})_{j \in \mathbb{N}}\) to schedule the observer.

Theorem 18 (Asymptotic convergence of observer (276))

Suppose Assumptions 171822, and 23 hold. Consider \( n_\zeta\) , \( (A,B)\) and \( \bar{c}_o\) given by Assumption 23. Let \( B_d\) be the first \( n_{y,d}\) columns of \( B\) and \( B_o\) be the remaining \( n_o\) columns, i.e., \( B = \begin{pmatrix} B_d & B_o\end{pmatrix}\) . Consider a bounded map \( \rm{sat}_o\) such that \( \rm{sat}_o(\hat{\xi}_o) = \hat{\xi}_o\) for all \( \hat{\xi}_o \in (\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}\) . Then, there exists \( \ell^\star>0\) such that for any \( \ell > \ell^\star\) , any maximal solution to the cascade (232)-(276) initialized in \( \Xi_0 \times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{0\}\) (with \( n_\eta = n_\zeta\) ) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\in\mathfrak{U}_c\times\mathfrak{U}_d\) , is complete and verifies (236).

Proof. This highly technical proof has been moved to Section 6.2.2.4 to facilitate reading. It involves three main steps: i) Choose \( \overline{c}_o\) (for the definition of \( \rm{sat}_o\) ) allowing us to guarantee some preliminary bounds and injectivity properties on the maps and variables, ii) Study the state dynamics of the \( \eta\) -coordinates (along solutions) and link these with \( \xi_{no}\) , and iii) Define a Lyapunov function and apply Theorem 23 to obtain exponential stability in the new coordinates after a certain time and retrieve asymptotic convergence in the \( \xi\) -coordinates. \( \blacksquare\)

With Theorem 18, we have shown that observers of the form (276) will work under Assumption 23. Next, we study specific cases of observers that fit into this framework, namely the MHO in Section 3.3.4.3 and the KKL observer in Section 3.3.4.4.

3.3.4.3 Case 1: Moving Horizon Observer for \( \xi_{no}\)

The idea of the MHO is to assume constructibility of the discrete-time system as in Section 2.2, namely the fact that the current state is expressible as a function of past inputs and outputs over a finite window, as in Assumption 24. Because the input \( (\xi_{o,k})\) is not exactly known in the observer and constructibility may depend on the inputs, we assume this property holds for any input \( (\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) , sufficiently close to \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) generated by the hybrid system (232).

Assumption 24

Given \( j_m\) and set \( \Xi_{no}\) defined in Item (XIV) of Assumption 17, there exist \( \overline{m}\in \mathbb{N}_{>0}\) , \( j_m^\star\in\mathbb{N}_{\geq\max\{j_m,\overline{m}\}}\) , \( \overline{c}_o>0\) , \( L_\Omega>0\) and for each \( \mathfrak{u}_c\in \mathfrak{U}_c\) , a Lipschitz map \( \Omega^{\mathfrak{u}_c}: \mathbb{R}^{\overline{m}n_{y,d,\rm{ ext }}} \times \mathbb{R}^{\overline{m}n_{u,\rm{ ext }}} \to \mathbb{R}^{n_{no}}\) with \( n_{u,\rm{ ext }}:=n_o+n_{u_d}+1+1\) and Lipschitz constant \( L_\Omega\) , such that along every complete solution \( \xi \in §_\mathcal{H}(\Xi_0, \mathfrak{U}_c \times \mathfrak{U}_d)\) , defining the sequences \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) as \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) for all \( j \in \mathbb{N}\) , we have that solutions to system (266) evolving in \( \Xi_{no}\) , with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j \in \mathbb{N}}:=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) such that \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_o\) for all \( j \in \mathbb{N}_{\geq j_m}\) , verify

\[ \begin{equation} \xi_{no,k} = \Omega^{\mathfrak{u}_c}(y_{k-1},y_{k-2},\ldots,y_{k-\overline{m}},\hat{u}_{k-1},\hat{u}_{k-2},\ldots,\hat{u}_{k-\overline{m}}), ~~ \forall k \in \mathbb{N}_{\geq j_m^\star}. \end{equation} \]

(277)

Remark 40

Contrary to the property in Definition 20 where for each component of the extended output, we have a different number of times \( m_i\) that we need to go to the past, here for the ease of presentation, we have assumed that this number of times is the same \( \overline{m}\) for all components. In practice, depending on the specific system, we can indeed store only the output components and the past inputs that are necessary for estimating \( \xi_{no}\) . Note that observability (or uniform forward distinguishability), namely the uniform injectivity of the forward map sequence

\[ \begin{equation} \xi_{no} \mapsto \mathcal{O}^{fw}_k(\xi_{no}) = \begin{pmatrix*}[l] (\mathcal{H}_{k-1} \circ \mathcal{G}_{k-2} \ldots \circ \mathcal{G}_{k-\overline{m}}) (\xi_{no}) \\ \ldots \\ (\mathcal{H}_{k-(\overline{m}-1)} \circ \mathcal{G}_{k-\overline{m}})(\xi_{no})\\ \mathcal{H}_{k-\overline{m}}(\xi_{no}) \end{pmatrix*} \in \mathbb{R}^{\overline{m}n_{y,d,\rm ext}}, \end{equation} \]

(278)

implies constructibility in Assumption 24. Therefore, if system (266) fed with \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j\in\mathbb{N}}\) is observable in this sense, then \( \Omega^{\mathfrak{u}_c}\) is obtained by first left-inverting (278) to obtain \( \xi_{no,k-\overline{m}}\) , then apply the jump dynamics (266) fed with \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j\in\mathbb{N}}\) to this to recover the current \( \xi_{no,k}\) . This condition is typically used in MHO[97, 98, 99, 96, 50].

Under this property, we fall in the framework of Assumption 23 as will be proven next, where:

  • \( \zeta_j\) represents the storage of the \( \overline{m}\) most recent past outputs \( (y_{j-1},y_{j-2},…,y_{j-\overline{m}})\) , thus of dimension \( n_\zeta = \overline{m}n_{y,\rm{ ext }}\) , corresponding to the pair \( (A,B)\) given in (275);
  • \( \mathcal{T}_j\) corresponds to \( \Omega^{\mathfrak{u}_c}(\cdot, u_{j-1},u_{j-2},…,u_{j-\overline{m}})\) , with \( (u_j)_{j \in \mathbb{N}}:=(\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) namely the mapping between the \( \overline{m}\) past outputs and \( \xi_{no}\) , whose time dependence is through the memory of the past \( \overline{m}\) inputs only;
  • \( \hat{\mathcal{T}}_j\) corresponds to \( \Omega^{\mathfrak{u}_c}(\cdot, \hat{u}_{j-1},\hat{u}_{j-2},…,\hat{u}_{j-\overline{m}})\) , with \( (\hat{u}_j)_{j \in \mathbb{N}}:=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) containing the estimated \( \hat{\xi}_{o,j}\) instead of \( \xi_{o,j}\) .

Proposition 1 then shows that the MHO falls into the scope of Assumption 23.

Proposition 1 (The MHO-based observer satisfies Assumption 23)

Suppose Item (XIV) of Assumption 17Assumption 22, and Assumption 24 hold. Then, Assumption 23 holds with \( n_\zeta\) , \( (A,B)\) in (275), \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}_{\geq \overline{m}}}\) as described above, and \( \overline{c}_o\) coming from Assumption 24.

Proof. Let us check each item of Assumption 23. First, with \( (A,B)\) in (275), it follows that \( \zeta_j = (y_{j-1},y_{j-2},…,y_{j-\overline{m}})\) for all \( j \in \mathbb{N}_{\geq \overline{m}}\) . So Item (XXIII) is directly obtained from Assumption 24, with the compact set \( \Xi_\zeta\) obtained from the uniform boundedness of \( (y_j)_{j\in \mathbb{N}}\) for \( (\xi_{o,j},\xi_{no,j})\in \Xi_{o}\times \Xi_{no}\) , \( \tau_j\in [0,\tau_M]\) , by global Lipschitzness of \( f_\rm{{sat}}\) and local boundedness of \( h_d\) and \( g_o\) uniformly in \( u_d\) according to Assumption 22. Next, Item (XXIV) is contained in Assumption 24. Last, for Item (XXV), by the Lipschitzness of \( \Omega^{\mathfrak{u}_c}\) , by definition of \( \mathcal{T}_j\) and \( \hat{\mathcal{T}_j}\) above, we get that, for any trajectory \( (\zeta_j)_{j\in \mathbb{N}}\) ,

\[ |\mathcal{T}_j(\zeta_j)-\hat{\mathcal{T}}_j(\zeta_j)| \leq L_\Omega |(\xi_{o,j-1},\xi_{o,j-2}, \ldots,\xi_{o,j-\overline{m}})-(\hat{\xi}_{o,j-1},\hat{\xi}_{o,j-2},\ldots,\hat{\xi}_{o,j-\overline{m}})|, ~~ \forall j \in \mathbb{N}_{\geq \overline{m}}. \]

Then, by taking the dynamics (273) of \( (\hat{\Gamma}_j)_{j \in \mathbb{N}}\) with

\[ A^\prime = \begin{pmatrix} 0 & 0 & \ldots & 0 &0\\ 1 & 0 & \ldots & 0 &0\\ \ldots & \ldots & \ldots & \ldots&\ldots\\ 0 & \ldots & 1 & 0 & 0\\ 0 & \ldots & 0 & 1 & 0 \end{pmatrix}, ~~ B^\prime = \begin{pmatrix} 1 \\ 0 \\ \ldots\\ 0 \\ 0 \end{pmatrix}, \]

we get that \( \hat{\Gamma}_j = (|\xi_{o,j-1}-\hat{\xi}_{o,j-1}|,|\xi_{o,j-2}-\hat{\xi}_{o,j-2}|,…,|\xi_{o,j-\overline{m}}-\hat{\xi}_{o,j-\overline{m}}|)\) for all \( j \in \mathbb{N}_{\geq \overline{m}}\) , and so this item follows. This concludes the proof. \( \blacksquare\)

By applying Theorem 18, we thus obtain observer (276) for system (232) where:

  • \( \tau(t,j)=t-t_j\) is the time elapsed since the previous jump;
  • At each time \( (t,j)\) , the high-gain estimate \( \hat{\xi}_o(t,j)\) can be used to compute, through backward integration by \( f_{o,\rm{sat}}\) , the value \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\) which is an estimate of \( \xi_o^+\) at the previous jump, namely \( \xi_o(t_j,j)\) which is the “fictitious output” \( \mathcal{H}_{o,j-1}\) in the discrete-time system (266). This estimate of the fictitious measurement gets better and better during flows, as the high-gain observer converges and “forgets” its wrong initial condition;
  • At the end of the flow interval at time \( (t_{j+1},j)\) , this is where the estimate of the previous fictitious measurement \( \mathcal{H}_{o,j-1}\) is the most reliable and we are thus ready to store it in \( \hat{\eta}^+\) , through \( AB_o\) , namely in the second memory block of \( \hat{\eta}\) ; on the other hand, the current \( y_d\) is obtained, corresponding to \( \mathcal{H}_{d,j}\) , which is going to be stored in the first memory block of \( \hat{\eta}^+\) through \( B_{d}\) ;
  • At any time, \( \hat{\eta}(t,j)\) thus contains the memory of \( (\mathcal{H}_{d,j-1},…, \mathcal{H}_{d,j-\overline{m}})\) , while it only contains the memory of \( (\mathcal{H}_{o,j-2},…, \mathcal{H}_{o,j-\overline{m}})\) . That is why we compute in (276.b) the quantity \( \hat{\zeta} = \hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\) which adds to the first memory block the best available estimate of the missing \( \mathcal{H}_{o,j-1}\) , namely \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\) . Then, \( \hat{\zeta}\) truly contains (an estimate of) \( (y_{j-1},…, y_{j-\overline{m}})\) , on which we can apply \( \hat{\mathcal{T}}_j\) , which provides an estimate of \( \xi_{no,j}\) , namely of \( \xi_{no}(t_j,j)\) . That is why in (276.b), a forward flow with \( f_\rm{{sat}}\) is then implemented to recover an estimate of \( \xi_{no}(t,j)\) ;
  • As explained above, \( \hat{\mathcal{T}}_j=\Omega^{\mathfrak{u}_c}(\cdot, \hat{u}_{j-1},\hat{u}_{j-2},…,\hat{u}_{j-\overline{m}})\) , with \( (\hat{u}_j)_{j \in \mathbb{N}}:=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) . Since, by definition, \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) , this can be implemented by running storage of those input signals in the observer at the jump times, namely

    \[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{\mathcal{M}}}_u &=& 0 & \text{when}~(232) \text{ flows} \\ \hat{\mathcal{M}}_{u}^+ &=& A_u \hat{\mathcal{M}}_{u} + B_u \begin{pmatrix} \hat{\xi}_o \\ u_d \\ t \\ \tau \end{pmatrix} & \text{when}~(232) \text{ jumps} \end{array} \right. \end{equation} \]

    (279.a)

    with

    \[ \begin{equation} A_u = \begin{pmatrix} 0 & 0 & \ldots & 0 &0\\ 1 & 0 & \ldots & 0 &0\\ \ldots & \ldots & \ldots & \ldots&\ldots\\ 0 & \ldots & 1 & 0 & 0\\ 0 & \ldots & 0 & 1 & 0 \end{pmatrix}\otimes \rm{Id}_{n_{u,{\rm ext }}}, ~~ B_u = \begin{pmatrix} 1 \\ 0 \\ \ldots\\ 0 \\ 0 \end{pmatrix}\otimes \rm{Id}_{n_{u,{\rm ext }}}, \end{equation} \]

    (279.b)

    so that, after \( \overline{m}\) jumps, \( \mathcal{M}_u(t,j)\) becomes exactly \( (\hat{u}_{j-1},\hat{u}_{j-2},…,\hat{u}_{j-\overline{m}})\) , and \( \hat{\mathcal{T}}_j\) simply writes as

    \[ \begin{equation} \hat{\mathcal{T}}_j(\cdot) = \Omega^{\mathfrak{u}_c}(\cdot, \hat{\mathcal{M}}_u(t,j)). \end{equation} \]

    (280)

In Section 3.4.3, the MHO-based observer proposed in this section will be applied for estimating restitution parameters in a bouncing ball.

3.3.4.4 Case 2: KKL Observer for \( \xi_{no}\)

In this section, we show that Assumption 23 can be verified through the (discrete-time) KKL paradigm applied to the discrete-time system (266). The main idea of KKL observer design as in Section 2.3 is to look for a time-varying change of coordinates, namely a sequence of maps \( (T_{\gamma,k})_{k\in \mathbb{N}}\) (depending on a design parameter \( \gamma \in (0,1]\) ), such that the image \( \zeta_k = T_{\gamma,k}(\xi_{no,k}) \in \mathbb{R}^{n_\zeta}\) , with some dimension \( n_\zeta \in \mathbb{N}_{\geq n_{no}}\) to be chosen, is solution to the dynamics (268) for some chosen \( (A(\gamma),B)\) . An observer for \( \zeta_k\) is then (274). If \( (T_{\gamma,k})_{k \in \mathbb{N}}\) is uniformly Lipschitz injective on \( \Xi_{no}\) after some discrete time \( j_m^\star \in \mathbb{N}\) , a bounded and uniformly Lipschitz sequence of left inverses \( (T_{\gamma,k}^*)_{k \in\mathbb{N}_{\geq j_m^\star}}\) of \( (T_{\gamma,k})_{k \in\mathbb{N}_{\geq j_m^\star}}\) can then be constructed on the whole \( \mathbb{R}^{n_\zeta}\) , verifying

\[ \begin{equation} T_{\gamma,k}^*(T_{\gamma,k}(\xi_{no})) = \xi_{no}, ~~ \forall k \in \mathbb{N}_{\geq j_m^\star},\forall \xi_{no} \in \Xi_{no} . \end{equation} \]

(281)

Then, the estimate defined by \( \hat{\xi}_{no,k} = T_{\gamma,k}^* (\hat{\zeta}_k)\) with \( (\hat{\zeta}_k)\) any solution to (274), is such that the estimation error \( |\xi_{no,k} - \hat{\xi}_{no,k}|\) is exponentially stable and converges to zero after \( j_m^\star\) . This injectivity is typically ensured by choosing \( A\) in (268) with sufficiently fast eigenvalues, namely of the form \( A(\gamma) = \gamma \tilde{A}\) where \( \tilde{A}\) is Schur and \( 0<\gamma\leq 1\) is pushed sufficiently close to zero. From there, an observer for the hybrid system (232) takes the form (276) with the pair \( (A(\gamma),B)\) and the map sequence \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) being \( (\hat{T}_{\gamma,j}^*)_{j \in \mathbb{N}_{\geq j_m}}\) after \( j_m^\star\) , obtained via the KKL observer applied to the discrete-time system (266) with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{u}_j)_{j \in \mathbb{N}}=(\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) defined along solutions as \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) for all \( j \in \mathbb{N}\) .

In order to obtain the right dynamics (268) in the \( \zeta\) -coordinates, the map sequence \( (T_{\gamma,k})_{k\in\mathbb{N}}\) is typically chosen recursively such that, along the inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (u_k)_{k \in \mathbb{N}}=(\xi_{o,k},u_{d,k},t_k,\tau_k)_{k \in \mathbb{N}}\) of interest, we have, at least after some time \( j_m\) ,

\[ \begin{multline} T_{\gamma,k+1}(\mathcal{G}_k(\xi_{no})) = A(\gamma)T_{\gamma,k}(\xi_{no}) + B_d\mathcal{H}_{d,k}(\xi_{no})+ B_o \mathcal{H}_{o,k}(\xi_{no}), \\ \forall k \in \mathbb{N}_{\geq j_m},\forall \xi_{no} \in \Xi_{no}: \mathcal{G}_k(\xi_{no}) \in \Xi_{no}. \end{multline} \]

(282)

As shown in Section 2.3, the existence and injectivity of this map sequence are typically obtained under the invertibility of \( \mathcal{G}_k\) and the backward distinguishability of the discrete-time system (266), as assumed next.

Assumption 25

Assume that \( g_o\) and \( h_d\) are globally Lipschitz in \( \xi\) , uniformly in \( u_d\) . Assume that there exist \( c_{\mathcal{G}^{-1}} > 0\) , \( \alpha > 0\) , \( \delta > 0\) , and \( m_i \in \mathbb{N}_{>0}\) for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) , such that along every complete solution \( \xi \in §_\mathcal{H}(\Xi_0, \mathfrak{U}_c \times \mathfrak{U}_d)\) , defining the sequences \( (u_j)_{j \in \mathbb{N}}=(\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) such that \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j\) for all \( j \in \mathbb{N}\) , we have:

  1. The map sequence \( (\mathcal{G}_j)_{j \in \mathbb{N}_{\geq j_m}}\) defined in (266) is uniformly Lipschitz invertible on \( \mathbb{R}^{n_{no}}\) , i.e., the inverse map \( \mathcal{G}_j^{-1}\) of \( \mathcal{G}_j\) is defined on \( \mathbb{R}^{n_{no}}\) and is such that \( |\mathcal{G}_j^{-1}(\xi_{no,a}) - \mathcal{G}_j^{-1}(\xi_{no,b})| \leq c_{\mathcal{G}^{-1}} |\xi_{no,a} - \xi_{no,b}|\) for all \( j \in \mathbb{N}_{\geq j_m}\) and for all \( (\xi_{no,a},\xi_{no,b}) \in \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_{no}}\) ;
  2. System (266) with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (u_j)_{j \in \mathbb{N}_{\geq j_m}}\) is uniformly Lipschitz backward distinguishable on \( \Xi_{no} + \delta \mathbb{B}\) after \( j_m\) , with parameters \( \alpha\) and \( m_i\) for each \( i \in \{1,2,…,n_{y,d,\rm{ext}}\}\) , i.e., for all \( k \in \mathbb{N}_{\geq j_m+\overline{m}}\) where \( \overline{m} := \max_{i \in \{1,2,…,n_{d, \rm{ext}}\}} m_i\) , the backward distinguishability map sequence \( (\mathcal{O}^{bw}_k)_{k\in\mathbb{N}}\) defined as

    \[ \begin{equation} \xi_{no} \mapsto \mathcal{O}^{bw}_k(\xi_{no}) = ( \mathcal{O}^{bw}_{1,k}(\xi_{no}), \mathcal{O}^{bw}_{2,k}(\xi_{no}), \ldots, \mathcal{O}^{bw}_{n_{y,d,\rm ext},k}(\xi_{no})) \in \mathbb{R}^{\sum_{i=1}^{n_{y,d,\rm ext}} m_i}, \end{equation} \]

    (283.a)

    where

    \[ \begin{equation} \mathcal{O}^{bw}_{i,k}(\xi_{no}) := \begin{pmatrix*}[l] (\mathcal{H}_{i,k-1} \circ \mathcal{G}_{k-1}^{-1}) (\xi_{no}) \\ (\mathcal{H}_{i,k-2} \circ \mathcal{G}_{k-2}^{-1} \circ \mathcal{G}_{k-1}^{-1})(\xi_{no}) \\ \ldots \\ (\mathcal{H}_{i,k-(m_i-1)} \circ \mathcal{G}_{k-(m_i-1)}^{-1}\circ\ldots\circ \mathcal{G}_{k-1}^{-1})(\xi_{no})\\ (\mathcal{H}_{i,k-m_i} \circ \mathcal{G}_{k-m_i}^{-1} \circ \mathcal{G}_{k-(m_i-1)}^{-1} \circ\ldots \circ\mathcal{G}_{k-1}^{-1})(\xi_{no}) \end{pmatrix*}, \end{equation} \]

    (283.b)

    where \( \mathcal{H}_{i,k}\) denotes the \( i^{\text{th}}\) row of \( \mathcal{H}_k\) , is uniformly Lipschitz injective on \( \Xi_{no} + \delta \mathbb{B}\) with constant \( \alpha\) , i.e., it satisfies

    \[ \begin{multline} |\mathcal{O}^{bw}_k(\xi_{no,a}) - \mathcal{O}^{bw}_k(\xi_{no,b})| \geq \alpha |\xi_{no,a} - \xi_{no,b}|, \\ \forall k \in \mathbb{N}_{\geq j_m+ \overline{m}}, \forall (\xi_{no,a},\xi_{no,b}) \in (\Xi_{no} + \delta \mathbb{B}) \times (\Xi_{no}+ \delta \mathbb{B}). \end{multline} \]

    (284)

Remark 41

Contrary to discrete-time systems, the jump map \( g_{no}\) of a hybrid system may not be invertible since it is not a discretization of some continuous-time dynamics. However, here \( (\mathcal{G}_j)_{j \in \mathbb{N}_{\geq j_m}}\) combines flow and jump dynamics, i.e., models the dynamics of the hybrid system sampled at jumps. It is thus reasonable to expect its invertibility along solutions after time \( j_m\) when we have the dwell time.

Note that in Assumption 25, the properties of uniform invertibility and uniform backward distinguishability of the discrete-time equivalent system are assumed to hold when this system is fed with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) along the hybrid solutions of interest, and not along the estimate provided by the observer, which differs in the value of \( (\hat{\xi}_{o,j})_{j \in \mathbb{N}}\) compared to \( (\xi_{o,j})_{j \in \mathbb{N}}\) . In the case of affine structures in Section 3.3.3, we also assume similar conditions along solutions (see Assumption 21), then manage to transfer these to the discrete-time equivalent system fed with the estimate, by the power of the continuous-time high-gain observer to converge arbitrarily fast (see the proof of Theorem 17). Now, we would like to exploit similar possibilities in the case of fully nonlinear maps. However, in this trickier case, the KKL transformation is no longer linear in \( \xi_{no}\) , making the conditions that this transformation must verify like (282) depend on \( \xi_{no}\) and so we must pay attention to the sets where they hold. Thus, to carry on with the technical proof, we need to exhibit more clearly the dependence of \( \mathcal{G}_j\) and \( \mathcal{O}^{bw}_j\) on the variables and inputs. More specifically, we define, whenever possible, \( \mathcal{G}^{\mathfrak{u}_c}\) , \( \mathcal{G}^{-1,\mathfrak{u}_c}\) , \( \mathcal{H}^{\mathfrak{u}_c}\) , and \( \mathcal{O}^{bw,\mathfrak{u}_c}\) , for \( u= (\xi_o,u_d,t,\tau)\) and \( u_{-1},u_{-2},…,u_{-\overline{m}}\) all in \( \mathbb{R}^{n_o} \times \mathbb{R}^{n_{u,d}}\times \mathbb{R}_{\geq 0} \times \mathbb{R}_{\geq 0}\) , as

\[ \begin{align} \mathcal{G}^{\mathfrak{u}_c}(u,\xi_{no})&:=g_{no}(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t+\tau,-\tau),\xi_{no}),t,\tau),u_d), \end{align} \]

(285.a)

\[ \begin{align} \mathcal{H}^{\mathfrak{u}_c}(u,\xi_{no})&:=\begin{pmatrix} h_d(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t+\tau,-\tau), \xi_{no}),t,\tau),u_d)\\ g_o(\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c)}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t+\tau,-\tau), \xi_{no}),t,\tau),u_d) \end{pmatrix},\\ \mathcal{G}^{-1,\mathfrak{u}_c}(u,\xi_{no})&:= (\mathcal{G}^{\mathfrak{u}_c}(u,\cdot))^{-1}(\xi_{no}),\\ \mathcal{O}^{bw,\mathfrak{u}_c}(u_{-1},u_{-2},\ldots, u_{-\overline{m}},\xi_{no})&:=\begin{pmatrix} \mathcal{O}^{bw,\mathfrak{u}_c}_1(u_{-1},u_{-2},\ldots, u_{-m_1},\xi_{no})\\ \mathcal{O}^{bw,\mathfrak{u}_c}_2(u_{-1},u_{-2},\ldots, u_{-m_2},\xi_{no})\\ \ldots\\ \mathcal{O}^{bw,\mathfrak{u}_c}_{n_{y,d,\rm ext}}(u_{-1},u_{-2},\ldots, u_{-m_{n_{y,d,\rm ext}}},\xi_{no})\end{pmatrix}, \end{align} \]

(285.b)

where

\[ \begin{equation} \mathcal{O}^{bw,\mathfrak{u}_c}_i(u_{-1},u_{-2},\ldots, u_{-m_i},\xi_{no}) := \begin{pmatrix*}[l] (\mathcal{H}_{i,-1} \circ \mathcal{G}_{-1}^{-1}) (\xi_{no}) \\ (\mathcal{H}_{i,-2} \circ \mathcal{G}_{-2}^{-1} \circ \mathcal{G}_{-1}^{-1})(\xi_{no}) \\ \ldots \\ (\mathcal{H}_{i,-m_i} \circ \mathcal{G}_{-m_i}^{-1} \circ \mathcal{G}_{-(m_i-1)}^{-1} \circ\ldots \circ\mathcal{G}_{-1}^{-1})(\xi_{no}) \end{pmatrix*}, \end{equation} \]

(285.c)

where \( \mathcal{H}_{i,-q}(\xi_{no})\) denotes the \( i^{\text{th}}\) row of \( \mathcal{H}^{\mathfrak{u}_c}(u_{-q},\xi_{no})\) and \( \mathcal{G}_{-q}^{-1}(\xi_{no}) = \mathcal{G}^{-1,\mathfrak{u}_c}(u_{-q},\xi_{no})\) , for \( q \in \{1,2,…,m_i\}\) . For the subsequent proof, the following additional regularity conditions are assumed for the maps that we have just defined.

Assumption 26

Assume that:

  1. There exists an open set \( \mathcal{V}_{\mathcal{G}}\) containing \( \Xi_o \times \mathcal{U}_d \times \mathbb{R}_{\geq 0} \times \mathbb{R}_{\geq 0} \times \Xi_{no}\) such that the map \( \mathcal{G}^{\mathfrak{u}_c}\) defined in (285.a) is \( C^2\) with respect to \( (\xi_o,\xi_{no})\) (where \( \xi_o\) is the first component of \( u=(\xi_o,u_d,t,\tau)\) ), with the maps \( \mathcal{G}^{\mathfrak{u}_c}\) , \( \frac{\partial \mathcal{G}^{\mathfrak{u}_c}}{\partial \xi_o}\) , \( \frac{\partial \mathcal{G}^{\mathfrak{u}_c}}{\partial \xi_{no}}\) , and \( \frac{\partial^2 \mathcal{G}^{\mathfrak{u}_c}}{\partial \xi_o \partial \xi_{no}}\) all bounded on \( \mathcal{V}_{\mathcal{G}}\) , uniformly for all \( \mathfrak{u}_c \in \mathfrak{U}_c\) ;
  2. There exists an open set \( \mathcal{V}_{\mathcal{O}}\) containing \( (\Xi_o \times \mathcal{U}_d \times \mathbb{R}_{\geq 0} \times \mathbb{R}_{\geq 0})^{\overline{m}} \times \Xi_{no}\) such that the map \( \mathcal{O}^{bw,\mathfrak{u}_c}\) defined in (285.b) is \( C^2\) with respect to \( (\xi_{o,-1},…, \xi_{o,-\overline{m}},\xi_{no})\) (where \( \xi_{o,-q}\) is the first component of \( u_{-q}=(\xi_{o,-q},u_{d,-q},t_{-q},\tau_{-q})\) ), with the maps \( \mathcal{O}^{bw,\mathfrak{u}_c}\) , \( \frac{\partial \mathcal{O}^{bw,\mathfrak{u}_c}}{\partial \xi_{o,-q}}\) , \( \frac{\partial \mathcal{O}^{bw,\mathfrak{u}_c}}{\partial \xi_{no}}\) , and \( \frac{\partial^2 \mathcal{O}^{bw,\mathfrak{u}_c}}{\partial \xi_{o,-q} \partial \xi_{no}}\) all bounded on \( \mathcal{V}_{\mathcal{G}}\) , for each \( q\in \{1,…,\overline{m}\}\) , uniformly for all \( \mathfrak{u}_c \in \mathfrak{U}_c\) .

Proposition 2 then shows that the KKL observer falls into the scope of Assumption 23.

Proposition 2 (The KKL-based observer satisfies Assumption 23)

Suppose Assumptions 17182225, and 26 hold. Define \( n_\eta = n_\zeta = \sum_{i=1}^{n_{d,\text{ext}}}m_i\) (with \( m_i\) coming from Item (XXVII) of Assumption 25) and consider for each \( i \in \{1,2,…,n_{d,\text{ext}}\}\) , a controllable pair \( (\tilde{A}_i,\tilde{B}_i) \in \mathbb{R}^{m_i \times m_i}\times\mathbb{R}^{m_i}\) with \( \tilde{A}_i\) Schur and invertible. Let \( T_{j_m}: \mathbb{R}^{n_{no}} \to \mathbb{R}^{n_\eta}\) globally Lipschitz and

\[ \begin{align} A(\gamma) &= \gamma\rm{diag}(\tilde{A}_1, \tilde{A}_2, \ldots, \tilde{A}_{n_{\text{ext}}}) \in \mathbb{R}^{n_\eta \times n_\eta}, \\ B_d& =\rm{diag}(\tilde{B}_1, \tilde{B}_2, \ldots, \tilde{B}_{n_{y,d}}) \in \mathbb{R}^{n_\eta \times n_{y,d}},\\ B_o& =\rm{diag}(\tilde{B}_{n_{y,d}+1}, \tilde{B}_{n_{y,d}+2}, \ldots, \tilde{B}_{n_{\text{ext}}}) \in \mathbb{R}^{n_\eta \times n_o}. \\\end{align} \]

(286.a)

There exists \( 0<\gamma^\star \leq 1\) such that for any \( 0<\gamma<\gamma^*\) , Assumption 23 holds with the pair \( (A(\gamma),B)\) where \( B = \begin{pmatrix} B_d & B_o \end{pmatrix}\) .

Proof. This highly technical proof has been moved to Section 6.2.2.5 to facilitate reading. It involves three main steps: i) Choose \( \overline{c}_o\) (for the definition of \( \rm{sat}_o\) ) allowing us to guarantee some preliminary bounds and injectivity properties on the maps and variables, ii) Apply the results in Theorem 4 and Theorem 7 starting from \( j_m\) to get the KKL maps, and iii) Check the conditions of Assumption 23. \( \blacksquare\)

Example 27 (System (232) with Polynomial Extended Output)

Consider the following form of system (232):

\[ \begin{equation} \mathcal{H}\left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\xi}_o &=& f_o(\xi_o, u_c) \\ \dot{\xi}_{no} &=& F_{noo}(\xi_o, u_c) + F_{no}(\xi_o, u_c)\xi_{no} \end{array} \right\} (\xi,u_c)\in C\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \xi_o^+ &=& J_o(\xi_o, u_d) + J_{ono}(\xi_o, u_d)P_{ono}(\xi_{no}) \\ \xi_{no}^+ &=& J_{noo}(\xi_o,u_d) + J_{no}(\xi_o,u_d)\xi_{no} \end{array} \right\} (\xi,u_d)\in D \end{array} \right. \end{equation} \]

(287.a)

with the flow and jump outputs

\[ \begin{equation} y_c = h_c(\xi_o,u_c), ~~ y_d = H_{do}(\xi_o,u_d) + H_{dno}(\xi_o,u_d)P_{dno}(\xi_{no}), \end{equation} \]

(287.b)

where \( P_{ono}\) and \( P_{dno}\) are vectors of monomials of \( \xi_{no}\) of order less than or equal to some \( m \in \mathbb{N}\) . Similar to the form (245), the jump dynamics of \( \xi_{no}\) are still required to be affine as in system (245), but the dependence of \( \xi_o^+\) and \( y_d\) on \( \xi_{no}\) is now allowed to be polynomial, which renders the considered form more nonlinear and hence wider, covering the form (245) as a special case when the polynomial is first-order. Because the flow dynamics of \( \xi_{no}\) are affine in \( \xi_{no}\) , the flow operator along \( f\) is also affine in \( \xi_{no}\) and we denote its \( \xi_{no}\) component as

\[ \begin{equation} \Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c),no}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t+\tau,-\tau), \xi_{no}),t,\tau) = \phi_{1,\mathfrak{u}_c}(\xi_o,t,\tau)\xi_{no}+ \phi_{2,\mathfrak{u}_c}(\xi_o,t,\tau). \end{equation} \]

(288)

In this setting, the discrete-time equivalent system (266) becomes

\[ \begin{equation} \xi_{no,k+1} = \mathcal{J}_{no,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) \xi_{no,k} +\mathcal{B}_{dno,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k), \end{equation} \]

(289.a)

where

\[ \begin{align} \mathcal{J}_{no,\mathfrak{u}_c}(\xi_{o},u_{d},t,\tau)& :=J_{no}(\xi_{o},u_{d})\phi_{1,\mathfrak{u}_c}(\xi_o,t,\tau), \\ \mathcal{B}_{dno,\mathfrak{u}_c}(\xi_o,u_d,t,\tau) &:= J_{noo}(\xi_o,u_d) + J_{no}(\xi_o,u_d)\phi_{2,\mathfrak{u}_c}(\xi_o,t,\tau), \\\end{align} \]

(289.b)

with the output

\[ \begin{equation} y_k = \begin{pmatrix} H_{do}(\xi_{o,k},u_{d,k})+H_{dno}(\xi_{o,k},u_{d,k})P_{dno}(\phi_{1,\mathfrak{u}_c}(\xi_{o,k},t_k,\tau_k)\xi_{no,k}+ \phi_{2,\mathfrak{u}_c}(\xi_{o,k},t_k,\tau_k))\\ J_o(\xi_{o,k},u_{d,k}) + J_{ono}(\xi_{o,k},u_{d,k})P_{ono}(\phi_{1,\mathfrak{u}_c}(\xi_{o,k},t_k,\tau_k)\xi_{no,k}+ \phi_{2,\mathfrak{u}_c}(\xi_{o,k},t_k,\tau_k))\end{pmatrix} \in \mathbb{R}^{n_{y,d,\rm ext}}, \end{equation} \]

(289.c)

where \( n_{y,d,\rm{ext}} := n_{y,d} + n_o\) , with inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,k},u_{d,k},t_k,\tau_k) \in (\Xi_o \cap D_o) \times \mathcal{U}_d \times \mathbb{R}_{\geq 0} \times \mathcal{I}\) for all \( k \in \mathbb{N}\) . Inspired by Example 9, when the system is linear in the dynamics and polynomial in the output, we look for the sequence \( (T_{k,\gamma})_{k \in \mathbb{N}}\) of the form

\[ \begin{equation} T_{\gamma,k}(\xi_{no}) = M_{\gamma,k} P(\xi_{no}) + N_{\gamma,k}, \end{equation} \]

(290)

where \( P\) is a vector of all monomials up to order \( m\) and we then want to find the dynamics of \( (M_{\gamma,k},N_{\gamma,k})_{k \in \mathbb{N}}\) to later implement them as observer states, thus somehow computing \( (T_{\gamma,k})_{k \in \mathbb{N}}\) dynamically. First, there exist matrix-valued functions \( H_{dno}^\prime\) and \( J_{ono}^\prime\) such that for all \( k \in \mathbb{N}\) and for all \( \xi_{no} \in \Xi_{no}\) ,

\[ \begin{align*} H_{dno}^\prime(\xi_{o,k},u_{d,k})P(\xi_{no}) & = H_{dno}(\xi_{o,k},u_{d,k})P_{dno}(\xi_{no}),\\ J_{ono}^\prime(\xi_{o,k},u_{d,k})P(\xi_{no}) & = J_{ono}(\xi_{o,k},u_{d,k})P_{ono}(\xi_{no}). \end{align*} \]

To obtain the dynamics of \( (M_{\gamma,k},N_{\gamma,k})_{k \in \mathbb{N}}\) , notice that since the dependence of the dynamics on \( \xi_{no}\) is linear, there exists a matrix-valued function \( Q_{\mathfrak{u}_c}\) (depending on \( \mathfrak{u}_c \in \mathfrak{U}_c\) ), such that for all \( k \in \mathbb{N}\) and for all \( \xi_{no} \in \Xi_{no}\) ,

\[ \begin{equation} P(\mathcal{J}_{no,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) \xi_{no} +\mathcal{B}_{dno,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k)) = Q_{\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k)P(\xi_{no}). \end{equation} \]

(291)

Applying (282) to this particular case, we have that for all \( \xi_{no} \in \Xi_{no}\) such that \( \mathcal{J}_{no,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) \xi_{no} +\mathcal{B}_{dno,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) \in \Xi_{no}\) and for all \( k \in \mathbb{N}\) ,

\[ \begin{multline} M_{\gamma,k+1} P(\mathcal{J}_{no,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) \xi_{no} +\mathcal{B}_{dno,\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k)) = A(\gamma) M_{\gamma,k} P(\xi_{no}) \\ + (B_d H_{dno}^\prime(\xi_{o,k},u_{d,k}) + J_{ono}^\prime(\xi_{o,k},u_{d,k}))P(\xi_{no}). \end{multline} \]

(292)

Using (291) and equalizing both sides for all \( \xi_{no}\) , we get that for all \( k \in \mathbb{N}\) ,

\[ \begin{equation} M_{\gamma,k+1} Q_{\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k) = A(\gamma) M_{\gamma,k} + B_d H_{dno}^\prime(\xi_{o,k},u_{d,k}) + B_o J_{ono}^\prime(\xi_{o,k},u_{d,k}). \end{equation} \]

(293)

Assuming that the matrix \( Q_{\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k)\) is invertible for all \( \mathfrak{u}_c \in \mathfrak{U}_c\) and for all \( k \in \mathbb{N}\) , we obtain the dynamics of \( (M_{\gamma,k})_{k \in \mathbb{N}}\) as

\[ \begin{equation} M_{\gamma,k+1} = (A(\gamma) M_{\gamma,k} + B_d H_{dno}^\prime(\xi_{o,k},u_{d,k}) + B_o J_{ono}^\prime(\xi_{o,k},u_{d,k}))(Q_{\mathfrak{u}_c}(\xi_{o,k},u_{d,k},t_k,\tau_k))^{-1}. \end{equation} \]

(294)

Similarly, the dynamics of \( (N_k)_{k \in \mathbb{N}}\) are obtained as

\[ \begin{equation} N_{\gamma,k+1} = A(\gamma) N_{\gamma,k} + B_d H_{do}(\xi_{o,k},u_{d,k}) + B_o J_o(\xi_{o,k},u_{d,k}). \end{equation} \]

(295)

The hybrid observer for system (287) will then be

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{\xi}}_o &= & \hat{f}_{o,\ell}(\hat{\xi}_o,p, \tau,y_c,u_c)\\ \dot{\hat{\eta}} & = &0 \\ \dot{p}&=&\varphi_{c,\ell}(p, \tau, y_c,u_c)\\ \dot{\hat{M}}_{\gamma}&=&0\\ \dot{\hat{N}}_{\gamma}&=&0\\ \dot{\tau}&=&1 \end{array}\right\} \text{when}~(287) \text{ flows}\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{\xi}_o^+&=&g_o(\rm{sat}_o(\hat{\xi}_o),\rm{sat}_{no}(\hat{\xi}_{no}),u_d)\\ \hat{\eta}^+&=& A(\gamma) \hat{\eta} + B_dy_d+ A(\gamma)B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\\ p^+&=&\varphi_{d,\ell}(p, \tau, y_d,u_d)\\ \hat{M}_{\gamma}^+&=&(A(\gamma) \hat{M}_{\gamma} + B_d H_{dno}^\prime(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &&~{}+ B_o J_{ono}^\prime(\rm{sat}_o(\hat{\xi}_o),u_d))\rm{inv}_{c_{\hat{Q}}}(Q_{\mathfrak{u}_c}(\rm{sat}_o(\hat{\xi}_o),u_d,t,\tau))\\ \hat{N}_{\gamma}^+&=&A(\gamma) \hat{N}_{\gamma} + B_d H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)+B_o J_o(\rm{sat}_o(\hat{\xi}_o),u_d)\\ \tau^+&=&0 \end{array} \right\} \text{when}~(287) \text{ jumps} \end{array} \right. \end{equation} \]

(296.a)

with the saturation maps \( \rm{inv}_{c_{\hat{Q}}}\) defined as in (264.c) for some levels \( c_{\hat{Q}} > 0\) from the invertibility of \( (Q_{\mathfrak{u}_c}(\xi_{o,j},u_{d,j},t_j,\tau_j))_{j \in \mathbb{N}}\) , and the same output map (276.b) with \( \hat{\mathcal{T}}_j\) being the left inverse of \( \hat{T}_j\) defined as

\[ \begin{equation} \hat{T}_j(\xi_{no}) := \hat{M}_\gamma(t,j) P(\xi_{no}) + \hat{N}_\gamma(t,j). \end{equation} \]

(296.b)

Table 5 then summarizes and compares the two designs proposed in this section. We see that the MHO-based observer is more advantageous implementation-wise than the KKL-based one, while the latter should have more filtering effect thanks to its transformation being updated dynamically, which keeps track of all the past outputs and inputs. This has been illustrated in Section 2.4.5 for a discrete-time system.

Table 5. Comparison of observer designs for hybrid systems (232) with fully nonlinear maps and known jump times. All the conditions here are sufficient conditions.
MHO-based observer (Section 3.3.4.3)KKL-based observer (Section 3.3.4.4)
Time domainPersistence of both flows and jumps (Item (XIV) of Assumption 17)
Hybrid modelNo assumptionInvertibility of the maps \( (\mathcal{G}_j)_{j \in \mathbb{N}}\) along estimates, or along solutions (Item (XXVI) of Assumption 25) plus some regularity conditions (Item (XXVIII) of Assumption 26)
ObservabilityInstantaneous observability of \( \xi_o\) and uniform Lipschitz constructibility of system along estimates (266) (Assumption 24)Instantaneous observability of \( \xi_o\) and uniform Lipschitz backward distinguishability of system (266) along estimates, or along solutions (Item (XXVII) of Assumption 25) plus some regularity conditions (Item (XXIX) of Assumption 26)
State dimension\( n_o + n_p +n_\eta+1\)\( n_o + n_p + \left(\sum_{i=1}^{n_{d, \rm{ext}}}m_i\right) + 1\)
Computation of left inverseOptimizing a cost function made of finite past outputsLeft-inverting the KKL transformation
Convergence rateArbitrarily fast, achieved after a certain timeArbitrarily fast, achieved after a certain time

3.3.5 Conclusion

We propose novel observer designs for hybrid systems with nonlinear maps and known jump times based on decomposing the state into two parts: the first part is independent of the second one during flows. The first part is assumed to be instantaneously observable during flows from the flow output and the second one is required to be either detectable, constructible, or backward distinguishable via the combination of flows and jumps, from the jump output as well as a fictitious output describing how this second part impacts the first one at jumps and thus becomes in turn visible from the flow output in the subsequent flow interval. A detectability analysis shows the necessity of detectability from this extended output for the existence of an observer. An observer is then designed to estimate each part, namely a high-gain flow-based observer using the flow output for the first part and either an LMI\( /\) MHO\( /\) KKL-based jump-based observer using an extended jump output for the second one, depending on the linearity of the system’s maps and corresponding to each of the mentioned observability conditions. Global exponential convergence or GES of the estimation error in the original coordinates is proven using Lyapunov analysis. While our theoretical contributions are illustrated using academic examples of a spiking neuron, in Section 3.4, the proposed observers are applied to estimate state and impact uncertainties in mechanical systems: a bouncing ball and a walking robot.

Acknowledgments. We thank Aymeric Cardot, student at Mines Paris - PSL, for his collaboration in this chapter.

3.4 Application to Mechanical Systems with Impacts

 

Dans ce chapitre, nous appliquons les observateurs développés dans Chapitre 3.2 et Chapitre 3.3 à divers systèmes mécaniques avec impacts, qui représentent une classe importante de systèmes hybrides avec des instants de saut connus. Nous commençons par tirer parti de l’observateur hybride de type Kalman présenté au Chapitre 3.2 pour estimer les biais dans une IMU. Ensuite, nous utilisons un observateur basé sur MHO, comme discuté au Chapitre 3.3, pour estimer les paramètres liés au coefficient de restitution d’une balle rebondissante. Enfin, nous mettons en œuvre à la fois des observateurs basés sur LMI et KKL du Chapitre 3.3 pour estimer les coefficients de restitution ainsi que la pente inconnue du terrain dans un robot bipède. Nos applications couvrent un large éventail de conditions d’observabilité, permettant l’estimation de paramètres autrement non mesurables grâce à la sortie fictive qui émerge lorsque ces paramètres interagissent avec d’autres composants d’état lors des sauts.

3.4.1 Introduction

Section 3.4.2 of this chapter (without the figures) has been published in [158, Example 6.2.6], while the rest will be submitted to a journal as applications of the theoretical results in Section 3.3.

In this chapter, we apply the observers designed in Section 3.2 and Section 3.3 to several mechanical systems with impacts:

  • In Section 3.4.2: The hybrid Kalman-like observer proposed in Section 3.2.2 is applied for estimating states and unknown biases in an Inertial Measurement Unit[IMU], consisting of a gyroscope and an accelerometer. We consider first the case where the bias only appears in the accelerometer, then the more complicated case where both this and the gyroscope are biased, necessitating another hybrid observer forming a cascade with the Kalman-like one. This observer relies on the Uniform Complete Observability[UCO] of the whole state from the combination of flow and jump outputs;
  • In Section 3.4.3: The Moving Horizon Observer[MHO]-based observer proposed in Section 3.3.4 is applied for estimating states and unknown impact parameters in a bouncing ball. We present a model of the restitution coefficient containing several parameters following a fully nonlinear form. We then illustrate how an MHO-based observer can be used to estimate these, providing a strategically designed jump input that excites the observability of these components. This observer requires instantaneous observability from the physical states (position and velocity), as well as constructibility of the rest;
  • In Section 3.4.4: The Linear Matrix Inequality[LMI]-based and Kravaris-Kazantzis$\slash$Luenberger[KKL]-based observers in Section 3.3.3 are applied to a walking robot for estimating states and unknown restitution coefficients when one leg hits the ground, as well as potentially an unknown slope of the terrain that the robot is walking on. As a proof of concept, we model the two-link walking robot, stabilized by a controller that is assumed robust against the uncertainties, as a hybrid system whose maps are nonlinear but affine in the restitution coefficients, allowing for observers in Section 3.3.3 to be designed. We also illustrate through simulations, without rigorous proof, that the scheme still works if the slope and the restitution coefficients vary in time, suggesting that online estimation of the terrain slope can be done online when the robot is walking through a realistic terrain. These observers require instantaneous observability from the physical states (angular positions and velocities), as well as quadratic detectability or respectively backward distinguishability of the rest.

Note that mechanical systems with impacts, especially walking robots, constitute an important class of hybrid systems with known jump times. Thanks to available impact detectors or even by simply exploiting the output, the jumps can be effectively detected in these systems.

The highlights of our work lie not only in the consideration of the solution’s discontinuities but also in the special kind of observability brought by the fictitious output, allowing estimating otherwise non-measurable variables. Indeed, the sensor biases in Section 3.4.2 as well as the jump parameters in Section 3.4.3 and Section 3.4.4 cannot be measured by any physical sensors. In our studies, they become detectable\( /\) observable thanks to their interaction at jumps with the rest of the state components, which are instantaneously observable during flows, thus allowing them to be indirectly “seen” in the subsequent flow interval. This story of the fictitious measurement has been formalized and exploited for observer design in Section 3.2 as well as Section 3.3, and the applications in this chapter help illustrate this interesting hybrid phenomenon. Our applications also have a wide range of observability conditions, reflecting the richness of our results in Section 3.2 and Section 3.3.

3.4.2 Altitude Estimation in a Pendulum with IMU Biases

In this section, we apply the hybrid Kalman-like observer in Section 3.2.2 to a pendulum equipped with an IMU. We estimate both the physical state of the former and biases in the latter.

3.4.2.1 System Modeling and Observability Analysis

As illustrated in Figure 31, consider a pendulum equipped with an IMU and bouncing on a vertical wall, with angular position \( \boldsymbol{\theta} \in \mathbb{R}^3\) . The IMU contains a gyroscope measuring its angular velocity \( \boldsymbol{\omega} \in \mathbb{R}^3\) and an accelerometer measuring its proper acceleration (linear acceleration minus gravity) \( \boldsymbol{y_a} \in \mathbb{R}^3\) in the IMU frame, modulo an unknown constant bias \( \boldsymbol{b_a} \in \mathbb{R}^3\) . We treat both the cases of a bias-free and biased gyroscope. We assume that the pendulum tilt \( \boldsymbol{t} \in \mathbb{R}^3\) is measured at the jump times, i.e., that the impacts are detected and the wall is vertical. The linear velocity \( \boldsymbol{v} \in \mathbb{R}^3\) in the sensor frame can be deduced from the gyroscope measurement via kinematics [3]. We also assume the velocity magnitude is reduced by an unknown constant restitution coefficient \( c \in (0,1]\) at each impact.

&lt;span data-controller=&quot;mathjax&quot;&gt;Illustration of a pendulum equipped with an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;IMU&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Inertial Measurement Unit&lt;/span&gt;&lt;/span&gt; bouncing on a vertical wall.&lt;/span&gt;
Figure 31. Illustration of a pendulum equipped with an IMU bouncing on a vertical wall.

With these in mind, we model the system in hybrid form with state \( x = (\boldsymbol{t},\boldsymbol{v},\boldsymbol{b}_{\boldsymbol{a}},c)\) as (see [3] for the flow dynamics)

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\boldsymbol{t}} &=& -[\boldsymbol{\omega}]_\times \boldsymbol{t} \\ \dot{\boldsymbol{v}} &=& -[\boldsymbol{\omega}]_\times \boldsymbol{v} + \boldsymbol{y_a} - \boldsymbol{b_a} + a_g \boldsymbol{t} \\ \dot{\boldsymbol{b}}_{\boldsymbol{a}} &=& 0 \\ \dot{c} &=& 0 \end{array} \right. ~~ y_c = \boldsymbol{v}, ~~ \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \boldsymbol{t}^+ &=& \boldsymbol{t} \\ \boldsymbol{v}^+ &=& - c \boldsymbol{v} \\ \boldsymbol{b}^+_{\boldsymbol{a}} &=& \boldsymbol{b}_{\boldsymbol{a}}\\ c^+ &=& c \end{array} \right. ~~ y_d=(\boldsymbol{v},\boldsymbol{t}), \end{equation} \]

(297)

where \( [\boldsymbol{\omega}]_\times = \begin{pmatrix} 0 & -\omega_3 & \omega_2 \\ \omega_3 & 0 & -\omega_1 \\ -\omega_2 & \omega_1 & 0 \end{pmatrix}\) and \( a_g = 9.81\) m\( /\) s\( ^2\) is the gravitational acceleration, with the flow and jump sets depending on the wall configuration. This system takes the form (155) where \( F = \begin{pmatrix} -[\boldsymbol{\omega}]_\times & 0 & 0 & 0 \\ a_g \rm{Id} & -[\boldsymbol{\omega}]_\times & -\rm{Id} & 0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \end{pmatrix}\) , \( J = \begin{pmatrix} \rm{Id} & 0 & 0 & 0 \\ 0 & 0 & 0 & -y_{d,\boldsymbol{v}} \\ 0 & 0 & \rm{Id} & 0 \\ 0 & 0 & 0 & 1 \end{pmatrix}\) , \( u_c = \begin{pmatrix} 0 \\ \boldsymbol{y_a} \\ 0 \\ 0 \end{pmatrix}\) , and \( u_d=0\) , with \( y_{d,\boldsymbol{v}}\) the \( \boldsymbol{v}\) -component of \( y_d\) . We would like to estimate both \( \boldsymbol{b_a}\) and \( c\) as well as the states. Assuming the system is persistently not at rest (with external excitation if \( c<1\) ), it is observable over each period of time containing at least one jump and a flow interval, because:

  • \( \boldsymbol{t}\) is available at jumps and its dynamics are independent of the other state components, making it observable, independently of the input signal \( [\boldsymbol{\omega}]_\times\) ;
  • \( \boldsymbol{v}\) and \( \boldsymbol{b_a}\) are both instantaneously observable during flows from \( y_c\) once \( \boldsymbol{t}\) is known, also independently of \( [\boldsymbol{\omega}]_\times\) ;
  • \( c\) is observable at jumps by seeing \( \boldsymbol{v}\) as a known input, because \( \boldsymbol{v}\) is measured.

It follows that if the pendulum velocity is uniformly lower-bounded (thanks to an appropriate input in the mechanical system, whose effects are in fact contained in \( \boldsymbol{y_a}\) in the IMU frame), the observability Gramian computed over a time window \( \Delta\) larger than the maximal length of flow intervals would be uniformly positive definite, and so this system is UCO—see Definition 17.

3.4.2.2 Kalman-Like Observer Design for System (297)

3.4.2.2.1 Case 1: Bias in Accelerometer Only

In this first scenario, we consider that only the accelerometer is biased, and the gyroscope gives perfect measurements. Because in this setting, the pendulum remains in a vertical plane, the states are simplified so that \( \boldsymbol{\theta}\) and \( \boldsymbol{\omega}\) each has one dimension, and \( \boldsymbol{t}\) , \( \boldsymbol{v}\) , \( \boldsymbol{b_a}\) each has two dimensions (corresponding to the horizontal and vertical axes). Assuming the boundedness of \( F\) and \( J\) , a Kalman-like observer (156) fed with \( (y_c,y_d, \boldsymbol{\omega},\boldsymbol{y_a})\) can be designed for system (297). Simulation results in Figure 32 illustrate an effective estimation of the states and the uncertainties.

&lt;span data-controller=&quot;mathjax&quot;&gt;State, bias, and restitution coefficient estimation in an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;IMU&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Inertial Measurement Unit&lt;/span&gt;&lt;/span&gt; with only accelerometer bias and without gyroscope bias.&lt;/span&gt;
Figure 32. State, bias, and restitution coefficient estimation in an IMU with only accelerometer bias and without gyroscope bias.
3.4.2.2.2 Case 2: Bias in Both Accelerometer and Gyroscope

Now, consider the case where there is an unknown constant bias \( \boldsymbol{b_g} \in \mathbb{R}^3\) in the gyroscope measurement (when restricted to a vertical plane, this dimension reduces to one). As a result, \( [\boldsymbol{\omega}]_\times\) in system (297) is replaced by \( [\boldsymbol{\omega_m} - \boldsymbol{b_g}]_\times\) , where \( \boldsymbol{\omega_m}\) is the biased measurement. Assume an estimate \( \hat{\boldsymbol{b}}_{\boldsymbol{g}}\) of \( \boldsymbol{b_g}\) is available such that \( \hat{\boldsymbol{b}}_{\boldsymbol{g}}-\boldsymbol{b_g}\) asymptotically vanishes. Then, the dynamics (297) may be rewritten as

\[ \begin{equation} \dot{x} = \hat{F} x + (F - \hat{F})x, \end{equation} \]

(298)

where \( [\boldsymbol{\omega_m} - \hat{\boldsymbol{b}}_{\boldsymbol{g}}]_\times\) replaces \( [\boldsymbol{\omega_m} - \boldsymbol{b_g}]_\times\) in \( \hat{F}\) . Consider the previous Kalman-like observer, but designed with the known \( \hat{F}\) instead of \( F\) . According to our observability analysis above, the observability of the quadruple \( (F,J,H_c,H_d)\) does not depend on \( F\) and therefore still holds for \( (\hat{F},J,H_c,H_d)\) . The robust stability of the Kalman-like observer (see Theorem 10 in Section 3.2) then ensures that the estimation error converges asymptotically to zero because:

  • The UCO condition holds with \( \hat{F}\) replacing \( F\) ;
  • The “disturbance” \( (F - \hat{F})x\) vanishes asymptotically thanks to \( \hat{\boldsymbol{b}}_{\boldsymbol{g}}\) converging to \( \boldsymbol{b_g}\) and the boundedness of \( x\) .

In other words, as illustrated in Figure 33, we only need to find an asymptotic estimate of \( \boldsymbol{b_g}\) and feed it to the hybrid Kalman-like observer of \( x\) . For that, notice that the pendulum position \( \boldsymbol{\theta}\) verifies the hybrid dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\boldsymbol{\theta}} &=& \boldsymbol{\omega_m} - \boldsymbol{b_g} \\ \dot{\boldsymbol{b}}_{\boldsymbol{g}} &=& 0 \end{array} \right. ~~ y_c^\prime = 0, ~~ \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \boldsymbol{\theta}^+ &=& \boldsymbol{\theta} \\ \boldsymbol{b_g}^+ &=& \boldsymbol{b_g} \end{array} \right. ~~ y_d^\prime=\boldsymbol{\theta}, \end{equation} \]

(299)

with the flow and jump sets depending on the wall configuration. In other words, \( \boldsymbol{\theta}\) has continuous-time dynamics, but with sampled measurement at each impact obtained via the impact condition. Because (299) has linear maps and only the jump output, we can design a jump-based observer with a constant gain using LMIs on the equivalent discrete-time system sampled at the jumps (see [22, Corollary \( 5.2\) ]).

&lt;span data-controller=&quot;mathjax&quot;&gt;Cascade of two observers in the case of both accelerometer and gyroscope biases; another observer estimates the gyroscope bias b_g) and feeds it to the Kalman-like observer. Flow outputs are colored in blue and jump outputs in red.&lt;/span&gt;
Figure 33. Cascade of two observers in the case of both accelerometer and gyroscope biases; another observer estimates the gyroscope bias \( \boldsymbol{b_g}\) and feeds it to the Kalman-like observer. Flow outputs are colored in blue and jump outputs in red.

All in all, still if the pendulum velocity is uniformly lower-bounded away from zero and \( (F,J)\) upper-bounded, an observer of \( x\) is obtained by the cascade of a jump-based observer of \( \boldsymbol{b_g}\) and a Kalman-like one for system (297) fed with \( [\boldsymbol{\omega_m} - \hat{\boldsymbol{b}}_{\boldsymbol{g}}]_\times\) instead of \( [\boldsymbol{\omega}]_\times\) . Simulation results in Figure 34 illustrate that following our approach, all the biases and the restitution coefficient can be effectively estimated.

&lt;span data-controller=&quot;mathjax&quot;&gt;State, bias, and restitution coefficient estimation in an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;IMU&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Inertial Measurement Unit&lt;/span&gt;&lt;/span&gt; with both gyroscope and accelerometer biases. Top two figures: Estimation by the jump-based observer; bottom four figures: Estimation by the Kalman-like observer.&lt;/span&gt;
Figure 34. State, bias, and restitution coefficient estimation in an IMU with both gyroscope and accelerometer biases. Top two figures: Estimation by the jump-based observer; bottom four figures: Estimation by the Kalman-like observer.

3.4.3 Restitution Parameter Estimation in a Bouncing Ball

In this section, we apply the MHO-based observer in Section 3.3.4 to a bouncing ball, to estimate the parameters in its restitution coefficient, following a fully nonlinear form.

3.4.3.1 System Modeling and Observability Analysis

Consider a ball with position \( s_1\) and velocity \( s_2\) , bouncing vertically on a flat surface. While this ball is in the air (when \( s_1 \geq 0\) ), its flow dynamics read \( \dot{s}_1 = s_2\) and \( \dot{s}_2 = -c_fs_2^{d_f}-a_g\) where \( a_g = 9.81\) m\( /\) s\( ^2\) is the gravitational acceleration. We take the constants \( c_f = 0.01\) s\( ^{d_f-2}\) m\( ^{1-d_f}\) and \( d_f = 2\) , corresponding to nonlinear friction when the ball is in the air. When this ball hits the ground (when \( s_1 = 0\) and \( s_2 \leq 0\) ), we assume that its post-impact position remains zero as \( s_1^+ = s_1\) and its velocity gets reversed in direction and possibly reduced in magnitude according to \( s_2^+ = - c_r(s_2) s_2 + u_d\) , where \( u_d\) is some strategically designed jump input described later, and with a restitution coefficient \( c_r\) depending on \( s_2\) at the impacts following the form

\[ \begin{equation} c_r(s_2) = \frac{a}{1+e^{b(|s_2| - c)}} + d \in (0,1], \end{equation} \]

(300)

where \( (a,b,c,d)\) are unknown constant parameters of appropriate units. Illustrated in Figure 35, the interest of the form (300) is that when the impact velocity is too large, some irreversible damage is caused, which reduces the restitution. Motivated by this, we pick a restitution coefficient model that takes a high value when the velocity is low, and the other way around, with some reasonable transition in the middle. We measure the position \( s_1\) at all times using a camera, and the solutions’ jumps can be detected when this measurement equals zero. Let us try to estimate the restitution parameters by adding them to the state, leading to the following extended hybrid system, with state \( x = (s_1,s_2,a,b,c,d)\) ,

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{~~}l@{}} \dot{x} &=&\begin{pmatrix} x_2 \\ -c_fx_2^{d_f}-a_g \\ 0_{4 \times 1} \end{pmatrix}, & \text{when } x_1 \geq 0, & y_c = x_1, \\ x^+ &=& \begin{pmatrix} x_1\\ -c_r(x_2,x_3,x_4,x_5,x_6)x_2 + u_d\\ (x_3,x_4,x_5,x_6) \end{pmatrix},& \text{when} \left\{\begin{array}{@{}l@{}} x_1 \geq 0 \\ x_2 \leq 0 \end{array}\right., & y_d = x_1. \end{array}\right. \end{equation} \]

(301)

Let us analyze the observability of system (301). We see that \( (x_1,x_2)\) are instantaneously observable from \( y_c\) during flows, so following the form (232), let \( \xi_o = (x_1,x_2)\) . Now, exploiting the instantaneous observability of \( x_2\) , we qualitatively analyze the \( (a,b,c,d)\) part.

&lt;span data-controller=&quot;mathjax&quot;&gt;Restitution coefficient c_r(s_2)) as function of s_2 = x_2) with a = 0.65) , b = 0.65) (s /) m), c = 14.5) (m /) s), and d = 0.2) .&lt;/span&gt;
Figure 35. Restitution coefficient \( c_r(s_2)\) as function of \( s_2 = x_2\) with \( a = 0.65\) , \( b = 0.65\) (s\( /\) m), \( c = 14.5\) (m\( /\) s), and \( d = 0.2\) .

See from Figure 35 that, \( a\) and \( d\) determine the lower and upper bounds of the sigmoid-like graph of \( c_r\) , while \( b\) and \( c\) determine its shape. As a result, when the value of \( |x_2|\) at jumps lies in the left and right zones, i.e., when roughly \( |x_2| = |s_2| \leq 5\) (m\( /\) s) and \( |x_2| = |s_2| \geq 25\) (m\( /\) s), knowing \( u_d\) , we can recover \( a\) and \( d\) from two different \( (x_2,x_2^+)\) pairs because roughly \( \frac{-(x_2^+-u_d)}{x_2} = c_r(x_2) \approx a + d\) in the left zone and \( \frac{-(x_2^+-u_d)}{x_2} = c_r(x_2) \approx d\) in the right zone. Similarly, when \( |x_2|\) at jumps lies in the middle zone, i.e., when roughly \( 10 \leq |x_2| = |s_2| \leq 20\) (m\( /\) s), knowing \( u_d\) , we should be able to recover \( b\) and \( c\) from another two different \( (x_2,x_2^+)\) pairs. This means that, given \( x_2\) being instantaneously observable from \( y_c\) , \( \xi_{no} = (x_3,x_4,x_5,x_6)\) should be (re)constructible from the fictitious measurement \( x_2^+\) and inputs \( (x_2,u_d)\) at jumps, provided that the value of \( x_2\) at jumps sufficiently “browses" throughout the different zones and at least four different past values of \( (x_2,u_d,x_2^+)\) are gathered at jumps. This analysis shows that the equivalent discrete-time system taking dynamics simplified from (267), which are

\[ \begin{equation} \xi_{no,k+1} = \xi_{no,k}, ~~ y_k = g_o(\xi_{o,k},\xi_{no,k},u_{d,k}), \end{equation} \]

(302)

due to the constant nature of \( \xi_{no,k}\) , is constructible provided that the inputs \( (\xi_{o,k},u_{d,k})_{k\in\mathbb{N}}\) make its solution regularly visit the different zones.

3.4.3.2 MHO-Based Observer Design for System (301)

3.4.3.2.1 Case 1: Observability Guaranteed by Varying Jump Input

We then design an MHO-based observer for the \( \xi_{no}\) part of this system. Given that the number of unknowns is four, we expect that the window size should be five or higher and we here choose it as \( n_\eta = 14\) , which is considerably larger. In general, we must also store the most recent \( n_\eta\) past pieces of the jump inputs that are necessary for the implementation of the inverse map sequence \( (\hat{\mathcal{T}_j}_{j \in \mathbb{N}})\) , namely the most recent \( n_\eta\) intervals of the continuous-time trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) (which is infinite-dimensional) and the most recent \( n_\eta\) values of the sequences \( (\hat{\xi}_o(t_{j+1},j),u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) . In this particular case, since the flow part is time-invariant with \( a_g\) constant, we only need to store the most recent \( n_\eta\) values of the sequences \( (\hat{\xi}_o(t_{j+1},j),u_{d,j})_{j \in \mathbb{N}}\) in the variable \( \hat{\mathcal{M}}_u\) as in (279). We see that \( \hat{\eta}\) serves to store \( \hat{\xi}_o(t_j,j)\) (after each jump), so with \( (\hat{\eta},\hat{\mathcal{M}}_u)\) , we can solve the second line of \( x^+\) in (301) to get \( (a,b,c,d)\) . As analyzed above, the constructibility of \( (a,b,c,d)\) can be achieved by appropriately choosing the input sequence \( \mathfrak{u}_d\) , which we propose to increase in time and reset after a period. The resulting \( x_2\) trajectory is shown in Figure 36 and we see that the pre-impact velocity periodically visits all three zones analyzed above, leading to a well-observable trajectory.

Estimation results are shown in Figure 36. We see that the resolution for \( (a,b,c,d)\) happens after we have stored enough past outputs in accordance with the window size, namely after around \( 14\) jumps.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation of states and parameters in a bouncing ball by coupling a high-gain observer with an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;MHO&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Moving Horizon Observer&lt;/span&gt;&lt;/span&gt;-based observer, in an observable setting.&lt;/span&gt;
Figure 36. Estimation of states and parameters in a bouncing ball by coupling a high-gain observer with an MHO-based observer, in an observable setting.
3.4.3.2.2 Case 2: Observability Poorly Guaranteed by Constant Jump Input

On the other hand, if \( u_d\) remains constant as \( u_d = 2\) , the resulting \( x_2\) is also periodic, but it does not visit the different zones, which corresponds to a poorly observable mode. As a result, \( \xi_{no}\) is hard to observe in this case and the estimation, still using an MHO-based observer of size fourteen, becomes impaired as shown in Figure 37. This example highlights the importance of observability from the fictitious measurement provided by the inputs.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation of states and parameters in a bouncing ball by coupling a high-gain observer with an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;MHO&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Moving Horizon Observer&lt;/span&gt;&lt;/span&gt;-based observer, in a poorly observable setting.&lt;/span&gt;
Figure 37. Estimation of states and parameters in a bouncing ball by coupling a high-gain observer with an MHO-based observer, in a poorly observable setting.

3.4.4 Uncertainty Estimation in a Walking Robot

In this section, as a proof of concept, we apply the LMI-based and KKL-based observers in Section 3.3.3 to the compass model of a bipedal walking robot, to estimate the physical states and especially the unknown restitution coefficients as well as the walking terrain’s slope.

3.4.4.1 Hybrid Modeling of a Bipedal Walking Robot

In this part, we derive the swing and impact dynamics of the bipedal walking robot presented in [31], then formulate these into a hybrid system.

3.4.4.1.1 Swing Phase Dynamics and Controller

The considered compass model following [31] of a bipedal robot walking on a slope \( \phi\) has state \( (\theta,\omega)\) with angular position \( \theta = (\theta_1,\theta_2) \in \mathbb{R}^2\) defined with respect to axes perpendicular to the inclined walking ground, angular velocity \( \omega = (\omega_1,\omega_2) \in \mathbb{R}^2\) , and torque input \( u \in \mathbb{R}^2\) . This configuration is illustrated in Figure 38. The parameters are from [31], namely \( m_H = 3\) (kg), \( m = 0.4\) (kg), \( a = 0.215\) (m), \( b = 0.465\) (m), and \( l = a+b = 0.680\) (m).

&lt;span data-controller=&quot;mathjax&quot;&gt;Configuration of a two-link walking robot (redrawn from&amp;#160;asanoBiped).&lt;/span&gt;
Figure 38. Configuration of a two-link walking robot (redrawn from [31, Fig. 1]).

The flow (swing) dynamics satisfy

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\:}c@{\:}l@{}} \dot{\theta}&=& \omega \\ \dot{\omega} &=& (M(\theta))^{-1}(u-C(\theta,\omega)\omega - G(\theta,\phi)), \end{array} \right. \end{equation} \]

(303)

where, with gravitational acceleration \( a_g = 9.81\) m\( /\) s\( ^2\) , we obtain the mass matrix, Coriolis matrix, and gravity matrix respectively as [170, 31]

\[ \begin{align*} M(\theta) &= \begin{pmatrix} m_H l^2 + ml^2 + ma^2 & -mbl \cos(\theta_1 - \theta_2) \\ -mbl \cos(\theta_1 - \theta_2) & mb^2 \end{pmatrix},\\ C(\theta,\omega)&=\begin{pmatrix} 0 & -mbl \sin(\theta_1 - \theta_2) \omega_2 \\ mbl \sin(\theta_1 - \theta_2) \omega_1 & 0 \end{pmatrix},\\ G(\theta,\phi)&=\begin{pmatrix} -(m_Hl + ml + ma)\sin (\theta_1+\phi) \\ mb \sin(\theta_2+\phi) \end{pmatrix}\frac{a_g}{\cos\phi}. \end{align*} \]

Since we are interested in designing observers to estimate variables including the slope, we assume the availability of a controller that is robust to this unknown slope, as follows.

Assumption 27

Assume that:

  • The slope \( \phi\) is constant;
  • The (state-feedback) controller \( u\) stabilizes the system to a periodic walking gait, for some small values of \( \phi\) .

Under Assumption 27, we can choose for simplicity and ease of presentation, during flows only, the state-feedback energy tracking controller as in [171], given by

\[ \begin{equation} u(\theta,\omega,\phi) = \left\{\begin{array}{@{}l@{~~}l@{}} \displaystyle-\frac{\lambda_E(E(\theta,\omega,\phi)-E_d)}{(\mu + 1)\omega_1 - \omega_2}\begin{pmatrix} \mu+1\\ -1 \end{pmatrix} & \text{if } |(\mu + 1)\omega_1 - \omega_2| \geq \underline{\omega} \\ 0 & \text{if } |(\mu + 1)\omega_1 - \omega_2| < \underline{\omega} \end{array}\right. \end{equation} \]

(304.a)

which exponentially stabilizes the total (kinetic \( +\) potential) energy obtained under the virtual gravity

\[ \begin{equation} E(\theta,\omega,\phi) = \frac{1}{2}\omega^\top M(\theta) \omega + (m_H l + ml + ma)a_g\cos(\theta_1-\phi)\cos\phi - mba_g\cos(\theta_2-\phi)\cos\phi \end{equation} \]

(304.b)

to a given value \( E_d > 0\) at the rate of \( \lambda_E > 0\) , while \( \mu > 0\) resembles a prescribed ratio between the two torque components. Here, we choose \( E_d = 22\) (J), \( \lambda_E = 10\) (J\( /\) s), and \( \mu = 10\) as in [31]. All of our simulations below, both in the presence and absence of \( \phi\) , are performed with this controller.

Remark 42

Since the focus of our work is observer design, similar to [172, 3], we suppose the availability of a controller that is robust against uncertainties. One controller that we have found in the literature is (304), which supposes knowing the energy hence the slope. The fact that the slope is integrated into the controller is only for stabilizing the system, allowing us to start the forthcoming analyses; this slope is still assumed unknown and is later estimated by our observers. This controller can be replaced with any other more complex state-feedback law that is able to work under a slight slope; in fact, even the same expression as (304) but assuming some reasonable value of \( \phi\) may also be used.

3.4.4.1.2 Derivation of the Impact Dynamics

In this work, because we are interested in estimating impact uncertainties, we choose here to consider the impact dynamics of [173] instead of the simpler one in [31]. Because the angular positions are with respect to the axes that are perpendicular to the tilted walking terrain, an impact happens when all of the following conditions hold:

\[ \begin{equation} \theta_1 + \theta_2 = 0, ~ \omega_1 \geq 0, ~ \omega_2 \geq 0, ~ \theta_1 > 0. \end{equation} \]

(305)

At the impact, the angular positions are swapped when the legs swap their roles:

\[ \begin{align} \theta_1^+&= \theta_2, \\ \theta_2^+ &= \theta_1. \\\end{align} \]

(306.a)

We are now ready to derive the post-impact angular velocities. For that, the following assumptions are made.

Assumption 28

Assume that:

  • Slipping and rebounds (of the legs) occur only at impacts and not during swings;
  • The horizontal and vertical velocities of the tip of the swing leg when it touches the ground are both zero.

The relation of post-impact angular velocities with the pre-impact ones and the uncertainties are given in Lemma 14.

Lemma 14 (Post-impact angular velocities of a two-link walking robot)

Under Assumption 28, the post-impact angular velocity \( \omega^+=(\omega_1^+,\omega_2^+)\) is given by:

\[ \begin{equation} \omega^+=J_s\begin{pmatrix} M_e(\theta) & -(E_J(\theta))^\top \\ E_J(\theta) & 0 \end{pmatrix}^{-1}\begin{pmatrix} M_e(\theta)(\omega,0,0) \\ \delta \end{pmatrix}, \end{equation} \]

(307.a)

where \( \delta = (\delta_t,\delta_n)\) are respectively the slipping constant and rebound coefficient, \( J_s = \begin{pmatrix} 1&0&0&0&0&0\\ 0&1&0&0&0&0 \end{pmatrix}\) , the Jacobian \( E_J(\theta) = \begin{pmatrix} l_1 \cos\theta_1 & -l_1 \cos \theta_2 & 1 & 0 \\ -l_1 \sin \theta_1 & l_1 \sin \theta_2 & 0 & 1 \end{pmatrix}\) , and

\[ \begin{equation} M_e(\theta) = \begin{pmatrix} M_{e,11} & M_{e,12}(\theta) & M_{e,13}(\theta) & M_{e,14}(\theta)\\ \star & M_{e,22} & M_{e,23}(\theta) & M_{e,24}(\theta)\\ \star & \star & M_{e,33} & 0 \\ \star & \star & \star & M_{e,44} \end{pmatrix}, \end{equation} \]

(307.b)

where \( M_{e,11} = m_H l_1^2 + m_1 l_1^2 + m_1 a_1^2\) , \( M_{e,12}(\theta) = -m_1b_1l_1 \cos(\theta_1 - \theta_2)\) , \( M_{e,13}(\theta) = (m_H l_1 + m_1 l_1 + m_1 a_1)\cos\theta_1\) , \( M_{e,14}(\theta) = -(m_H l_1 + m_1 l_1 + m_1 a_1)\sin\theta_1\) , \( M_{e,22} = m_1b_1^2\) , \( M_{e,23}(\theta) = -m_1b_1\cos\theta_2\) , \( M_{e,24}(\theta) = m_1b_1 \sin\theta_2\) , and \( M_{e,33} = m_H + 2m_1\) .

Proof. First, we derive the extended mass matrix \( M_e(\theta)\) . Define coordinates attached to the tip of the stance leg, with \( x\) -axis in parallel with the walking surface and \( y\) -axis perpendicular to this surface. Denote the position of the tip of the stance leg as \( (x_t,y_t)\) and its velocity as \( (\dot{x}_t,\dot{y}_t)\) . From the configuration in Figure 38, the positions of the masses (denoted “\( 1\) ", “\( H\) ", and “\( 2\) " from left to right) are

\[ \begin{align} p_1 & = a(\sin\theta_1,\cos\theta_1) + (x_t,y_t),\\ p_H & =l(\sin\theta_1,\cos\theta_1) + (x_t,y_t),\\ p_2 & = l(\sin\theta_1,\cos\theta_1)+b(-\sin\theta_2,-\cos\theta_2) + (x_t,y_t). \\\end{align} \]

(308.a)

It follows that their velocities are:

\[ \begin{align} v_1 & = a(\cos\theta_1,-\sin\theta_1)\dot{\theta}_1 + (\dot{x}_t,\dot{y}_t),\\ v_H & =l(\cos\theta_1,-\sin\theta_1)\dot{\theta}_1 + (\dot{x}_t,\dot{y}_t),\\ v_2 & = l(\cos\theta_1,-\sin\theta_1)\dot{\theta}_1+b(-\cos\theta_2,\sin\theta_2)\dot{\theta}_2 + (\dot{x}_t,\dot{y}_t). \\\end{align} \]

(309.a)

Denote \( K\) as the total kinetic energy of this system. We have:

\[ \begin{align*} 2K & = mv_1^2 + m_Hv_H^2 + mv_2^2\\ & =m(a^2\dot{\theta}_1^2 + \dot{x}_t^2 + \dot{y}_t^2 + 2a\dot{\theta}_1(\dot{x}_t\cos\theta_1-\dot{y}_t\sin\theta_1))\\ & ~{}+m_H(l^2\dot{\theta}_1^2 + \dot{x}_t^2 + \dot{y}_t^2 + 2l\dot{\theta}_1(\dot{x}_t\cos\theta_1-\dot{y}_t\sin\theta_1))\\ & ~{}+m(l^2\dot{\theta}_1^2 + b^2\dot{\theta}_2^2 + \dot{x}_t^2 + \dot{y}_t^2+ 2l\dot{\theta}_1(\dot{x}_t\cos\theta_1-\dot{y}_t\sin\theta_1)\\ & ~{}+ 2b\dot{\theta}_2(-\dot{x}_t\cos\theta_2+\dot{y}_t\sin\theta_2)- 2bl\dot{\theta}_1\dot{\theta}_2\cos(\theta_1-\theta_2)). \end{align*} \]

By seeing that this kinetic energy also verifies \( 2K = (\theta_1,\theta_2,\dot{x}_t,\dot{y}_t)^\top M_e(\theta)(\theta_1,\theta_2,\dot{x}_t,\dot{y}_t)\) , we obtain the expression of \( M_e(\theta)\) in Lemma 14. Following [173], assuming the absence of slipping and rebound at the impacts, we can write the post-impact dynamics as

\[ \begin{equation} \begin{pmatrix} \omega_e^+ \\ F \end{pmatrix} =\begin{pmatrix} M_e(\theta) & -(E_J(\theta))^\top \\ E_J(\theta) & 0 \end{pmatrix}^{-1}\begin{pmatrix} M_e(\theta)\omega_e \\ 0 \end{pmatrix}, \end{equation} \]

(310)

where \( \omega_e = (\omega_1,\omega_2,v_x,v_y)\) , with \( v_x\) and \( v_y\) the horizontal and vertical velocity of the tip of the swing leg when it touches the ground, supposing they are both zero as in Assumption 28, with the Jacobian and the extended mass matrix as in Lemma 14. Here, \( F = (F_t,F_n) \in \mathbb{R}^2\) is the contact force and the impact uncertainty \( \delta = (\delta_t,\delta_n) \in \mathbb{R}^2\) are unknown constants to estimate along with \( (\theta,\omega)\) and \( \phi\) . However, we assume instead the presence of uncertainties \( \delta = (\delta_t,\delta_n)\) and integrate these in the impact dynamics, replacing the zeros in \( (M_e(\theta)\omega_e,0)\) in (310), hence obtaining the results. The matrix \( J_s\) is to take out the post-impact angular velocities. \( \blacksquare\)

3.4.4.1.3 Hybrid System Formulation

Now, let us formulate the bipedal robot into the hybrid system framework [10]. First, choosing the state as \( x = (\theta,\omega,\delta)\) to estimate later, we define the state space (the set where the admissible state values remain):

\[ S = \left\{x \in \mathbb{R}^6: \theta \in \left(-\frac{\pi}{2},\frac{\pi}{2}\right)^2,\omega \in \mathbb{R}^2,\delta \in [0,0.2]^2\right\}. \]

Note that dealing with uncertainty issues (impacts, slipping, friction, etc.) is of research interest in the robotics community [174, 175, 176].

Assumption 29

We can measure angular position \( \theta\) during swing phase, despite slope \( \phi\) .

Remark 43

Assumption 29 is possible when \( y_c\) is obtained through encoders placed at the ankles and the hip, assuming the foot is flat on the ground and the robot is rigid.

Following the formulation in Section 3.4.4.1.1 and Section 3.4.4.1.2, the system dynamics are

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}c@{~~}c@{}} \dot{x}&=&f(x) & x \in S \setminus D & y_c=\theta\\ x^+&=&g(x) & x\in D & y_d = 0 \end{array} \right. \end{equation} \]

(311.a)

where \( y_c\) and \( y_d\) are respectively flow and jump outputs, with the jump set

\[ \begin{equation} D = \left\{x \in S, \theta_1 + \theta_2 = 0, \omega_1 \geq 0,\omega_2\geq 0, \theta_1 > 0\right\}, \end{equation} \]

(311.b)

the flow map

\[ \begin{equation} f(x) = (\omega, (M(\theta))^{-1}(u(\theta,\omega,\phi)-C(\theta,\omega)\omega - G(\theta,\phi)),0,0), \end{equation} \]

(311.c)

and jump map

\[ \begin{equation} g(x) = (\theta_2,\theta_1,g_1(\theta)\omega + g_2(\theta)\delta,\delta), \end{equation} \]

(311.d)

with \( g_1\) and \( g_2\) functions of only \( \theta\) obtained from (307). Here, \( y_d\) is taken equal to zero to encode the fact that we would like to avoid using sensor outputs at the impact due to a loss of reliability (apart from the required jump detection obtained by impact sensors). We see that \( (\theta,\omega)\) are independent of \( \delta\) during flows and depend on \( \delta\) in an affine way at jumps. In the next part, we analyze the observability of system (311) and design corresponding observers for it.

3.4.4.2 Observability Analysis and Observer Design for System (311)

Before designing observers, let us analyze the observability of system (311), starting with the physical states \( (\theta,\omega)\) and the slope \( \phi\) .

3.4.4.2.1 Instantaneous Observability of \( \phi\) from \( y_c\) during Flows

First, Lemma 15 shows the instantaneous observability of \( (\theta,\omega,\phi)\) from \( y_c = \theta\) during flows.

Lemma 15 (Instantaneous observability of \( \phi\) from \( y_c\) during flows)

Under Assumption 29, with \( u\) known, \( (\theta,\omega,\phi)\) such that \( \cos\phi \neq 0\) is instantaneously observable from \( y_c\) on \( S\) .

Proof. First, we get instantaneously \( \theta = y_c\) and \( \omega = \dot{y}_c\) . From (303), we get

\[ \begin{align*} M(y_c)\ddot{y}_c & = M(\theta)\dot{\omega}\\ & = u-C(\theta,\omega)\omega - G(\theta,\phi)\\ &= u-C(y_c,\dot{y}_c)\dot{y}_c - G(y_c,\phi). \end{align*} \]

It follows that the first lines of both sides are equal, giving us

\[ -(m_Hl + ml + ma)\sin(\theta_1+\phi)\frac{a_g}{\cos\phi} = \begin{pmatrix} 1 & 0 \end{pmatrix}(u-C(y_c,\dot{y}_c)\dot{y}_c - M(y_c)\ddot{y}_c). \]

By seeing that \( \sin(\theta_1+\phi) = \sin\theta_1\cos\phi + \cos\theta_1\sin\phi\) and \( \theta_1 = y_{c,1}\) , we have (since \( \cos\phi \neq 0\) on \( S\) ), \( \sin y_{c,1} + \cos y_{c,1}\tan\phi = \frac{-1}{a_g(m_Hl + ml + ma)}\begin{pmatrix} 1 & 0 \end{pmatrix}(u-C(y_c,\dot{y}_c)\dot{y}_c - M(y_c)\ddot{y}_c)\) , which means, since \( \cos y_{c,1} \neq 0\) on \( S\) ,

\[ \begin{equation} \phi = \arctan \left(\frac{1}{\cos y_{c,1}}\left(\frac{-1}{a_g(m_Hl + ml + ma)}\begin{pmatrix} 1 & 0 \end{pmatrix}(u-C(y_c,\dot{y}_c)\dot{y}_c - M(y_c)\ddot{y}_c) - \sin y_{c,1}\right)\right). \end{equation} \]

(312)

Because \( \phi\) is determined uniquely from \( (y_c,\dot{y}_c,\ddot{y}_c,u)\) , we have the result. \( \blacksquare\)

Lemma 15 tells us that \( \phi\) can be estimated along with \( (\theta,\omega)\) during flows using \( y_c\) by an arbitrarily fast high-gain observer. With this in mind, we perform the injective transformation

\[ \begin{equation} (\theta,\omega,\phi) \mapsto \xi_o = \begin{pmatrix} \theta \\ \omega \\ -(M(\theta))^{-1}G(\theta,\phi) \end{pmatrix} \in \mathbb{R}^6. \end{equation} \]

(313)

Then, with \( \xi_{no} = \delta\) , \( \xi = (\xi_o,\xi_{no})\) is solution to hybrid dynamics of the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\xi}_o &=& f_o(\xi_o, u) \\ \dot{\xi}_{no} &=& 0 \end{array} \right\} \xi_o \in C_o\\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \xi_o^+ &=& J_o(\xi_o) + J_{ono}(\xi_o)\xi_{no}\\ \xi_{no}^+ &=& \xi_{no} \end{array} \right\} \xi_o \in D_o \end{array} \right. \end{equation} \]

(314)

with \( C_o\) and \( D_o\) obtained from \( C\) and \( D\) in system (311), the flow outputs \( y_c = H_c\xi_o\) where \( H_c = \begin{pmatrix} \rm{Id} & 0_{2\times2} \end{pmatrix}\) , and an empty jump output. This takes the form (245) with \( F_{no} = 0_{2\times 2}\) , \( U_{cno} = 0\) , \( J_o(\xi_o) = (\theta_2,\theta_1,g_1(\theta)\omega)\in \mathbb{R}^2\times \mathbb{R}^2\) , \( J_{ono}(\xi_o) = (0_{2\times 2}, g_2(\theta))\in \mathbb{R}^{4\times 2}\) , \( J_{no}(\xi_o) = \rm{Id}_2\) , and \( J_{noo}(\xi_o) = 0 \in \mathbb{R}^2\) . Solutions of interest are known to be bounded in physical compact sets. Moreover, the maps are both locally Lipschitz and locally bounded in \( \xi_o\) . We can also show that the matrices to be inverted, like those in (303) and (310), are uniformly invertible. The map \( f_o\) is of triangular form and we can build for \( \xi_o\) a classical high-gain observer [26] with gain \( \ell\) during flows, with a copy of dynamics and saturation functions \( \rm{sat}_o\) and \( \rm{sat}_{no}\) at jumps (obtained by simulating the system solutions of interest):

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{\xi}}_o & = &\hat{f}_{o,\ell}(\hat{\xi}_o,y_c,u) & \text{when~(314) flows}\\ \hat{\xi}_o^+ & = & J_o(\rm{sat}_o(\hat{\xi}_o)) + J_{ono}(\rm{sat}_o(\hat{\xi}_o))\rm{sat}_{no}(\hat{\xi}_{no}) & \text{when~(314) jumps}, \end{array}\right. \end{equation} \]

(315)

then the estimate \( \hat{\phi}\) of \( \phi\) is function of \( \hat{\xi}_o\) , recovered as an output of the observer. It follows that the extra observer state \( p\) is empty. Note that when \( \phi = 0\) , i.e., walking on flat terrain, we let \( \xi_o = (\theta,\omega)\) thanks to this being instantaneously observable and decoupled from \( \delta\) during flows.

Next, knowing that \( \xi_o\) is estimated by a high-gain flow-based observer, we propose two observer designs for the \( \xi_{no} = \delta\) part, namely the LMI-based design in Section 3.4.4.2.2 and the KKL-based one in Section 3.4.4.2.3. We will consider the absence and presence of the slope \( \phi\) .

3.4.4.2.2 LMI-Based Observer Design for System (311)

We first design a jump-based observer estimating \( \xi_{no}\) , based on quadratic detectability in Assumption 20. To do this, we need to solve (249) for this system. Because there is no jump output \( y_d\) , we choose \( L_d = 0\) . Then, the gain \( K_d \in \mathbb{R}^{2\times 4}\) is found by solving

\[ \begin{equation} \left(\rm{Id}_2 - K_d \begin{pmatrix} 0_{2\times2}\\ g_2(\theta) \end{pmatrix}\right)^\top Q \left(\rm{Id}_2 - K_d \begin{pmatrix} 0_{2\times2}\\ g_2(\theta) \end{pmatrix}\right) < Q, \end{equation} \]

(316)

for \( \theta\in \mathbb{R}^2\) in a subset of the jump set \( D\) , namely with \( \theta_1 = -\theta_2\) and \( \theta_1>0\) . In fact, we solve (316) in a simple and practical way as follows. By looking along the solutions of interest, we see that \( \theta_1\) is around \( 0.2\) (rad) at the impacts and thus design \( K_d = \begin{pmatrix} 0_{2\times 2} & K_d^\prime \end{pmatrix}\) satisfying (316) for \( \theta_1 = 0.2\) (rad) by pole placement at zero. This is made possible by the fact that the matrix \( g_2(\theta)\) is invertible at \( \theta = (0.2,-0.2)\) (rad), so that indeed the pair \( (\rm{Id}_2,g_2(\theta))\) is observable at that point and \( K_d^\prime\) can be designed. Then, thanks to the stability margin in the unit circle, we should have the eigenvalues of \( \rm{Id}_2 - K_d J_{ono}(\theta)\) , for \( \theta\) in some small interval around \( \theta_1 = -\theta_2 = 0.2\) (rad), remain within the unit circle. To illustrate this, we show in Figure 39 the largest magnitude of the eigenvalues of \( \rm{Id}_2 - K_d J_{ono}(\theta)\) obtained by taking \( 200\) random samples in \( \theta_1 = -\theta_2 \in [0.17,0.23]\) (rad). It is seen that these magnitudes remain under \( 1\) , so the eigenvalues lie safely inside the unit circle and thus we have a reason to believe that (316) holds for all \( \theta_1 = -\theta_2 \in [0.17,0.23]\) (rad) (although it is not guaranteed that the same \( Q\) holds for all such \( \theta\) ). Of course, we can solve (316) in more complicated but rigorous ways—see Section 3.3.3.1.2.

&lt;span data-controller=&quot;mathjax&quot;&gt;Maximum magnitude of eigenvalues of Id_2 - K_d J_{ono}()) with 200) values of ) randomly sampled around _1 = -_2 [0.17,0.23]) (rad).&lt;/span&gt;
Figure 39. Maximum magnitude of eigenvalues of \( \rm{Id}_2 - K_d J_{ono}(\theta)\) with \( 200\) values of \( \theta\) randomly sampled around \( \theta_1 = -\theta_2 \in [0.17,0.23]\) (rad).

With \( K_d\) and the high-gain observer at hand, we are now ready to implement the hybrid observer (248). However, following the recommendations in Section 3.3.3.1.2, in order to avoid the computation of the Jacobian of the flow operator, we choose here to implement the observer dynamics in the \( z\) -coordinates, as in the proof of Theorem 16, where \( \hat{z}_{no}\) replaces \( \hat{\xi}_{no}\) with dynamics (430) and \( \hat{\xi}_{no}\) is recovered from \( \hat{z}_{no}\) with (429.a). Actually, exploiting the fact that \( \xi_{no}\) is constant in this application and that the estimate \( \hat{\xi}_{no}\) is only needed at the jumps in (430) for the observer, we propose to obtain \( \hat{\xi}_{no}\) using (429.a) not at all times, but at some (possibly varying) rate, to allow for a sporadic but more precise computation. In our case, we use (429.a) to get \( \hat{\xi}_{no}\) after each \( 0.0001\) (s) and hold its value in between the samples.

First, in the case \( \phi = 0\) and hence \( \xi_o = (\theta,\omega)\) , simulation results are shown in Figure 40. Before the first impact, even though we can estimate \( \xi_o\) due to its being instantaneously observable, we cannot recover \( \delta=\xi_{no}\) from the fictitious output. Indeed, no impact has occurred yet, so the information about \( \xi_{no}\) has not entered \( \xi_o\) yet and so it is not available. The convergence starts right after the first impact, when \( \xi_o\) , which now contains the \( \xi_{no}\) information, becomes observable again from \( y_c\) during the next flow interval. This illustrates the idea of the fictitious output, an interesting phenomenon brought by the hybrid nature of the system.

&lt;span data-controller=&quot;mathjax&quot;&gt;State and impact uncertainty estimation in a bipedal robot using an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;LMI&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Linear Matrix Inequality&lt;/span&gt;&lt;/span&gt;-based observer when = 0) .&lt;/span&gt;
Figure 40. State and impact uncertainty estimation in a bipedal robot using an LMI-based observer when \( \phi = 0\) .

Then, in the presence of a small slope of \( \phi = 0.03\) (rad), we see that the controller is still able to stabilize the robot to its walking gait. Figure 41 shows the estimation results, where \( \phi\) is recover as the output of the observer. Of course, bigger values of the uncertainties \( (\phi,\delta)\) can be considered given a more robust controller. The feedback of \( \hat{\phi}\) by the observer into the controller (304) might also be done under a kind of hybrid separation principle based on the idea of [177], without any rigorous proof.

&lt;span data-controller=&quot;mathjax&quot;&gt;State, slope, and impact uncertainty estimation in a bipedal robot using an &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;LMI&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Linear Matrix Inequality&lt;/span&gt;&lt;/span&gt;-based observer when 0) .&lt;/span&gt;
Figure 41. State, slope, and impact uncertainty estimation in a bipedal robot using an LMI-based observer when \( \phi \neq 0\) .
3.4.4.2.3 KKL-Based Observer Design for System (311)

Now, we exploit uniform backward distinguishability of the pair \( (J_{no},J_{ono}(\theta))\) as in Definition 20 and implement the KKL-based design (264). A first difference with the quadratic detectability-based design in the previous section is that, because \( \xi_{no}\) is constant, i.e., its dynamics matrix \( J_{no}\) is identity, solving (316) requires, in particular, the observability of the pair \( (\rm{Id}_2,g_2(\theta))\) , i.e., invertibility of the matrix \( g_2(\theta)\) , at each fixed \( \theta\) . In this section, uniform backward distinguishability instead draws observability from a certain number \( m\) of past outputs. It thus only requires the concatenation of all \( g_2(\theta)\) encountered in the considered window to be left-invertible (instead of each \( g_2(\theta)\) ). This advantage disappears however in the case of periodic solutions and we simply pick \( m_1=m_2=0\) (no information provided by the first two components of \( J_{ono}(\theta)\) ), and \( m_3=m_4=1\) (\( g_2(\theta)\) directly invertible), which gives us \( n_\eta = \sum_{i=1}^4 m_i = 2\) . A second important advantage is that the KKL-based design (264) of this section is systematic and does not require offline computation of the gains as in (316). The price to pay is its higher complexity.

The high-gain flow-based observer for \( \xi_o\) is the same as above. For the jump-based observer estimating \( \xi_{no}\) , since \( n_\eta = 2\) and given the choice of the \( m_i\) , we choose \( A = \rm{diag}(0.1,0.2)\) , \( B_{ono} = \begin{pmatrix} 0_{2 \times 2} & \rm{Id}_2 \end{pmatrix}\) , an empty \( B_{dno}\) , and \( \gamma = 0.4\) . This means we use only the last two components of \( J_{ono}(\theta)\) , which is \( g_2(\theta)\) , as the new output matrix for observer implementation. We still recover \( \hat{\xi}_{no}\) using (429.a) at the rate of \( 0.0001\) (s). First, in the case \( \phi = 0\) and hence \( \xi_o = (\theta,\omega)\) , simulation results are shown in Figure 42. Like in the last section, convergence of \( \hat{\delta}=\hat{\xi}_{no}\) only starts after the first impact when \( \xi_{no}\) has become visible from the fictitious output.

Now, we explore the case of varying slope and impact coefficients. We assume a scenario where the slope suddenly increases from \( 0.01\) to \( 0.02\) (rad) after \( 3\) seconds, \( \delta_t\) decreases from \( 0.2\) to \( 0.1\) (m\( /\) s\( ^2\) ) after \( 5\) seconds, and \( \delta_n\) increases from \( 0.1\) to \( 0.2\) (m/s\( ^2\) ) at the same time. The controller (304) is still able to work with this scenario. From Figure 43, we see that as \( \phi\) is instantaneously observable through \( y_c\) , any change in \( \phi\) is immediately noticed by the observer, and the estimate \( \hat{\phi}\) is updated right away. On the other hand, \( \delta\) is observable from the fictitious output which comes from the flow-jump combination. Therefore, the change in \( \delta\) can only be detected at the subsequent jump, and \( \hat{\delta}\) remains corrected with the outdated output until that jump when it gets corrected again. This means any change in \( \delta\) that is within a single flow interval will never be seen. Note that the variations we consider for \( \phi\) and \( \delta\) are consistent with practice—the slope may change quite fast and our observers can capture this instantaneously, while the impact uncertainties typically depend on the materials of the robot leg and the ground hence should not vary unless the walking terrain changes, so there is essentially no reason for these to vary faster than a flow interval (one step of the robot).

&lt;span data-controller=&quot;mathjax&quot;&gt;State and impact uncertainty estimation in a bipedal robot using a &lt;span class=&quot;acronym&quot; data-controller=&quot;footnote&quot; data-action=&quot;click-&gt;footnote#showHide&quot;&gt;KKL&lt;span class=&quot;footnoteText&quot; style=&quot;display:none&quot; data-footnote-target=&quot;footnoteText&quot;&gt;Kravaris-Kazantzis$$Luenberger&lt;/span&gt;&lt;/span&gt;-based observer when = 0) .&lt;/span&gt;
Figure 42. State and impact uncertainty estimation in a bipedal robot using a KKL-based observer when \( \phi = 0\) .
Figure 44
Figure 45
Figure 43. Estimation of \( \phi\) (left) and \( \delta\) (right) using a KKL-based observer, when these uncertainties vary in time.\label{fig:ch10_robot_kkl_slopev}

3.4.5 Conclusion

In this chapter, we apply the observers designed in Section 3.2 and Section 3.3 to mechanical systems with impacts, exploiting the fictitious measurement to estimate unmeasurable parameters.

Acknowledgments. We thank Aymeric Cardot and Shang Liu, students at Mines Paris - PSL, for their collaborations in Section 3.4.3 and Section 3.4.4, respectively.

4 Observer Design for Hybrid Systems with Unknown Jump Times

4.1 Review of Observer Designs for Hybrid Systems with Unknown Jump Times

 

Ce chapitre passe en revue les techniques d’observation disponibles pour les systèmes hybrides avec des instants de saut inconnus, c’est-à-dire ceux dont les sauts ne peuvent pas être détectés par les mesures, compliquant ainsi les définitions des observateurs et de la convergence. La notion de détectabilité ainsi que la convergence asymptotique vers un ensemble dans ce contexte sont discutées. Les résultats étudiés couvrent une large gamme d’approches, incluant des observateurs localisant les modes pour les systèmes commutés avec des commutations asynchrones, le filtre de Kalman à sauts (salted Kalman filter), des conceptions semi-globales grand gain basées sur l’estimation en boucle ouverte autour des instants de saut, et ainsi que des techniques impliquant le “collage” des sauts des systèmes hybrides. Des idées clés sont présentées en conclusion.

4.1.1 Overview

Here and for the rest of Section 4, we focus on autonomous hybrid systems of the form

\[ \begin{equation} \dot{x} = f(x), ~ x \in C, ~~ x^+ = g(x), ~ x \in D, ~~ y = h(x), \end{equation} \]

(317)

where \( x \in \mathbb{R}^{n_x}\) is the state and \( y\) is the output measured at all times. The objective is to design an observer for such systems, estimating \( x\) from the knowledge of \( y\) . Recall that when the jump times of solutions to system (317) are known or can be detected, as explored in Section 3, an observer can be constructed to jump synchronously with the system, thereby facilitating convergence and stability analysis and enabling a vast literature of designs as surveyed in Section 3.1.

In an intermediary scenario where the jump times of system (317) are not perfectly known but only approximately known, one approach consists of using an observer with known jump times and relying on the robustness of this design with respect to small errors in jump detection and triggering. An observer with non-synchronized jumps is designed for billiard-type systems in [178] but still requires jump detection in the system. The work [22, Section 6] establishes practical convergence of the estimation error when the jump detection is delayed, achieved outside of the solution’s jump times when the delay is smaller than an assumed dwell time. The more recent work in [132, Chapter 7] addresses parameter estimation in hybrid systems under jump detection delays, providing an upper bound on the estimation error under a dwell-time condition.

In this Section 4, we rather concentrate on the case where the jump times of system (317) are unknown, either because they are completely undetectable or because the knowledge of them is too approximate or noisy. In this setting, the definition of an observer and its convergence, or even of observability, is no longer straightforward because the observer jumps are not synchronized with those of the system. In other words, the observer cannot converge to the system state in standard distances around the jump times because of asynchronous discontinuities [67]. Such a phenomenon is also encountered in the context of contraction and trajectory tracking [179, 180, 181]. Jump times typically become unknown when there is no jump detector and the output \( y\) remains continuous, namely taking the same value before and after the jumps, rendering them undetectable. Such situations commonly arise in friction phenomena, where the system may either stick or slip depending on its external force relative to an unknown friction threshold. Even in applications where a jump detector is supposed to be available (like contact sensors in mechanical systems with impacts considered in Section 3.4), its output may be so noisy that the user may want to explore the possibility of designing an observer using this information as little as possible. In cases such as these, observer design should avoid any jump time information, and very few techniques exist as surveyed below.

Regarding the detectability of hybrid systems with unknown jump times, [67] relies on \( j\) -reparameterization of the hybrid time domain to define asymptotic detectability and observers. Specifically, given a complete observer solution \( \hat{z}\) with output \( \hat{x}\) , the paper proposes \( j\) -reparameterizing the true system solution \( x\) , possibly defined on \( \rm{dom} x \neq \rm{dom} \hat{z}\) , into the hybrid arc \( x^r\) that is defined on \( \rm{dom} \hat{z} = \rm{dom} \hat{x}\) and so can be compared with the estimate at the same hybrid time, defining the asymptotic convergence to the set

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} \hat{x} (= \rm{dom} x^r)}} d_§((x^r(t,j),\hat{x}(t,j))) = 0, \end{equation} \]

(318)

where \( §\) is the diagonal set where the estimate matches the (reparameterized) solution:

\[ \begin{equation} § = \{(x,\hat{x}) \in \mathbb{R}^{n_x} \times \mathbb{R}^{n_x}: x = \hat{x}\}. \end{equation} \]

(319)

Even then, \( d_§\) is typically not the right distance to compare solutions, because of peaking, i.e., a non-reducible error around the jump times due to asynchronous discontinuities [67]. This phenomenon was reported in the context of observation [178] as well as output feedback and tracking [23]. Therefore, when the estimate does not synchronize asymptotically with the system solution, provided that only one jump can happen at a time, we can only hope to converge asymptotically to the set

\[ \begin{multline} \mathcal{A} := \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times \mathbb{R}^{n_x} : x=\hat{x} \} \cup \{(x,\hat{x}) \in D\times \mathbb{R}^{n_x} : g(x)=\hat{x} \}\\ \cup \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times D : x=g(\hat{x}) \}\cup \{(x,\hat{x}) \in D\times D : g(x)=g(\hat{x})\}, \end{multline} \]

(320)

containing points that either are the same, or are one jump away from each other, or jump once to the same point. In a worse case when multiple \( N\) jumps can happen at the same time, the set \( \mathcal{A}\) can be generalized accordingly to contain points that are at most \( N\) jumps away from each other, as well as those jumping at most \( N\) times to the same point. Note that when jump times are known, asymptotic detectability is simplified to the notion in Definition 14, where solutions are compared directly within the same time domain, eliminating the need for parameterization as in (318).

In the following sections, we review the very scarce existing literature on observer designs for hybrid systems when the jump times are unknown.

4.1.2 Mode Location Observers for Switched Systems

Switched systems represent an important class of hybrid systems, described by

\[ \begin{equation} \dot{x} = f_\sigma(x,u), ~~ y = h_\sigma(x,u), \end{equation} \]

(321)

where the switching signal \( \sigma\) determines the active mode. In the case of asynchronous switching, where the observer cannot switch at the same instants as the system, this becomes a problem of designing observers for hybrid systems with unknown jump times. For such scenarios, mode-location observers, which aim to determine the current mode, have been proposed, but primarily in the linear context.

In [183], robust LMI-based designs are proposed for discrete-time switched systems with linear maps, also subject to disturbances and noise, guaranteeing the \( \mathcal{H}_\infty\) performance. The work in [139] addresses switched systems modeled as hybrid systems with discrete-time modes and continuous-time states by running parallel observers and monitoring the residual output error. A similar approach is taken in [184], focusing on switched systems with linear maps subject to disturbances and measurement noise, identifying the exact active mode and approximate state. An interval observer-based method to estimate the switching signal of switched systems with linear maps is proposed in [185]. Additionally, [186], still in the linear context, establishes the stability of the Kalman filter under uncertain switching times and thereby presents an optimization-based method for joint estimation of both the state and switching signal, inspired by the interacting multiple model-extended Viterbi algorithm.

4.1.3 Self-Triggered Designs for Hybrid Systems

For hybrid systems (317) with unknown jump times, since a synchronized observer like (146) is infeasible, an intriguing approach is to build instead a self-triggered hybrid observer of the form

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l@{}} \dot{\hat{z}}&=& \mathcal{F}(\hat{z},y,t,j) & (\hat{z},y) \in \mathcal{C}\\ \hat{z}^+&=&\mathcal{G}(\hat{z},y,t,j) &(\hat{z},y) \in \mathcal{D} \end{array} \right. ~~ ~~\hat{x}= \mathcal{T}(\hat{z},t,j), \end{equation} \]

(322)

with the maps \( \mathcal{F}\) , \( \mathcal{G}\) , and \( \mathcal{T}\) , and especially the sets \( \mathcal{C}\) and \( \mathcal{D}\) designed appropriately so that the observer state \( \hat{z}\) decides its jumps. This is quite common in the Kalman literature, with the so-called salted Kalman filter, taking into account the uncertainties at the jump times in the covariance matrix using the so-called saltation matrix, but without proof of convergence [187, 188].

Another approach is to rely on the continuous-time high-gain observer. The preliminary work [189], realizing that keeping correcting the estimate close to the jump times using the measurement could be counterproductive, proposes a self-triggered observer (322) consisting of: i) a high-gain observer that corrects the estimate arbitrarily fast during flows, and ii) a mechanism that, depending only on the estimate and measurement, disconnects the correction term near the jumps and lets the observer flow with the system’s dynamics until reaching the jump set (open-loop estimation). This mechanism is integrated into the map \( \mathcal{G}\) and sets \( \mathcal{C}\) , \( \mathcal{D}\) of observer (322), relying on dwell time and instantaneous observability of the flow part to establish a local result. Then, the developed version [131] achieves a semi-global design by first using a continuous-time high-gain observer, then after a certain time (depending only on the estimate), switching to the available local hybrid observer.

The primary limitation of this high-gain approach is its dependence on the high-gain observer’s ability to converge arbitrarily fast, which requires that the entire state be instantaneously observable from \( y\) during flows, i.e., requiring full observability from \( (f,h)\) . This unfortunately neglects the observability provided by the jump dynamics through \( g\) , excluding many hybrid systems where only part of the state is observable, while the rest is detectable only via interactions with other states, as discussed throughout Section 3 with the fictitious measurement.

4.1.4 Technique of “Gluing” the Jumps of Hybrid Systems

The idea of “gluing” the hybrid time domain was formalized in [190] for systems whose outputs are continuous at jumps (hence rendering jump detection impossible), based on transforming the hybrid dynamics (317) into continuous-time dynamics:

\[ \begin{equation} \dot{z} = f^\prime(z), ~~ y = h^\prime(z). \end{equation} \]

(323)

This change of coordinates is thus called a “gluing” transformation since it effectively glues the time domain of hybrid solutions. Then, if an observer \( \mathcal{F}^\prime\) can be designed in the new coordinates and if the transformation is injective outside of the jump set, the state estimate is obtained without requiring jump detection by running this continuous-time observer and left-inverting the transformation:

\[ \begin{equation} \dot{\hat{z}} = \mathcal{F}^\prime(\hat{z},y), ~~ \hat{x}(t) = \mathcal{T}(\hat{z}(t),y(t)). \end{equation} \]

(324)

Since \( \hat{x}\) is now a continuous-time signal instead of a hybrid one like in observer (322), it can be compared with the true solution \( x\) without having to parameterize \( x\) as in Section 4.1.1. Because points that are one jump away from each other are “glued” together, the change of coordinates is not injective on \( D\) , which results in asymptotic convergence except during increasingly smaller intervals around the jump times. The later work [191] develops this idea specifically for mechanical systems with impacts by providing constructive ways to find the gluing transformation for this class of systems, benefiting from their polynomial forms. To better explain this gluing concept, we now illustrate it in Example 28.

Example 28 (Gluing transformation for the bouncing ball)

Consider the bouncing ball in Example 1 with \( c_r = 1\) and \( u_d = 0\) so that \( x_2^+ = -x_2\) , and assume that we do not detect its jump times. Realizing that before and after the jumps, the measured position \( x_1\) is zero and the velocity \( x_2\) only switches signs, we pick \( z = (x_1,x_2^2,x_1x_2)\) as suggested in [190, 191]. With this choice, as illustrated in Figure 46, even if \( x\) is discontinuous, the dynamics (323) of \( z\) are continuous in time and we may build for \( z\) a continuous-time observer (324) without any jump detection, under some appropriate form and observability condition. Moreover, we see that this transformation is injective outside the jump set, allowing recovery of the estimate in the \( x\) -coordinates except around the jumps.

&lt;span data-controller=&quot;mathjax&quot;&gt;Gluing transformation for the bouncing ball based on&amp;#160;KimChoShaShiSeo.&lt;/span&gt;
Figure 46. Gluing transformation for the bouncing ball based on [190].

In the advanced version [32], the existence of such a gluing change of coordinates that is injective except on the jump set is shown to exist for a broad class of hybrid systems with appropriate manifold structure of the flow and jump sets. However, the main limitations of the gluing technique proposed in [32] are

  • There are no constructive methods to find the transformation, except for mechanical systems with impacts [191];
  • Even if we can go to the continuous-time form (323), it is not guaranteed that an observer \( \mathcal{F}^\prime\) can be designed for this, due to insufficient observability or the structure of \( (f^\prime,h^\prime)\) .

Our preliminary work in [192], further developed in Section 4.2, systematizes this “gluing” approach. It addresses the two limitations outlined above by proposing a fixed form in the \( z\) -coordinates for which an observer can always be designed, and by guaranteeing the existence of an injective transformation under some regularity conditions and a mild backward distinguishability assumption. When jumps are separated from each other, the gluing technique can exhibit asymptotic convergence of \( (x(t,j),\hat{x}(t))\) to the set \( \mathcal{A}\) in (320) as seen in Section 4.2.5. This is also stated in [32] but in another fashion.

4.1.5 Conclusion

This chapter reviews observers for hybrid systems with unknown jump times, namely those whose solutions jump at unknown instants, complicating the definitions of observers and convergence. The notion of detectability as well as asymptotic convergence to a set in this setting is discussed. The surveyed results cover a wide range of approaches, including mode-locating observers for switched systems with asynchronous switching, the salted Kalman filter, semi-global high-gain designs based on open-loop estimation around jump times, and techniques that involve “gluing” the jumps in hybrid systems.

For the class of impulsive systems, the work [193] proposes a novel viewpoint, appropriately discretizing the dynamics following the Schatzman–Paoli scheme [194] and designing a dead-beat discrete-time observer for the resulting system. This allows them to bypass the time domain mismatch issue by defining the estimation error in discrete time. However, this work is still limited to a class of impulsive systems whose maps after discretization are linear, and the link between the observability of the original system and the discretized one is not clear.

In addition to the reviewed literature, recent work in [4] builds on existing Kalman filtering techniques to introduce a framework for simultaneous state estimation and impact detection in legged robots modeled as switched systems, albeit without theoretical guarantees. Since legged robots represent an important class of hybrid systems, this work suggests the possibility of generalizing to simultaneous state and jump time estimation in hybrid systems, a concept reminiscent of the approach used for switched systems in Section 4.1.2, and which will be examined in more detail in the conclusion of this dissertation.

The remainder of Section 4 will focus on developing the “gluing” technique introduced in Section 4.1.4 into a systematic method, grounded in the Kravaris-Kazantzis$\slash$Luenberger[KKL] observer paradigm. Theoretical contributions will be presented in Section 4.2, followed by an application of this method to the stick-slip phenomenon in Section 4.3.

4.2 Gluing KKL Observer for Hybrid Systems

 

Dans ce chapitre, nous proposons une synthèse d’observateur systématique et novatrice pour les systèmes hybrides avec des instants de saut inconnus, en utilisant le paradigme de l’observateur Kravaris-Kazantzis$\slash$Luenberger[KKL] et le concept de “collage” des sauts. Nous démontrons l’existence d’une transformation qui convertit un système hybride avec des sorties continues aux instants de saut en un filtre stable en temps continu de cette sortie, pour lequel un observateur peut toujours être conçu sans nécessiter la détection des sauts. Sous les hypothèses de distingabilité en temps rétrograde et de régularité de cette transformation, nous prouvons que cette transformation est injective et donc inversible à gauche en dehors de l’ensemble de saut, permettant ainsi de récupérer l’estimation dans les coordonnées d’origine. Nous démontrons alors la convergence de cette estimée au sens d’une distance à un ensemble, modélisant la possibilité d’un saut d’avance ou de retard autour des instants de sauts. La robustesse vis à vis des incertitudes de modèle, de mesure, ou de transformation, est aussi démontrée au sens de cette distance. Nous décrivons également une méthode d’apprentissage systématique de l’inverse de la transformation à partir de données générées par des simulations hors ligne. Nos contributions théoriques sont illustrées à travers des exemples académiques impliquant la balle rebondissante.

4.2.1 Introduction

A preliminary part of this chapter has been published in [192]; the full version will be submitted to a journal.

As reviewed in Section 4.1, the concept of gluing the time domain of hybrid systems, namely transforming them via some special transformation called the gluing function into continuous-time dynamics for observer design without jump detection, has been proposed in [190] for hybrid systems with unknown jump times. The journal version [32] of that work shows the existence of an injective gluing function for hybrid systems whose outputs are continuous in time, but without providing constructive methods to build the transformation or ensuring that an observer can be designed for the continuous-time system in the new coordinates. Some constructive techniques have been proposed in [191] specifically for the class of mechanical systems with impacts, leveraging their polynomial structure, though these results are still limited in scope.

In this chapter, building on the theory of the Kravaris-Kazantzis$\slash$Luenberger[KKL] observer [105, 112], we propose utilizing the KKL transformation as a gluing function for general hybrid systems of the form [10]

\[ \begin{equation} \dot{x}=f(x)~ x\in C,~~~~ x^{+}=g(x)~ x\in D,~~~~ y=h(x), \end{equation} \]

(325)

with state \( x\in\mathbb{R}^{n_x}\) , where \( C,D \subset \mathbb{R}^{n_x}\) are the flow and jump sets, \( f,g:\mathbb{R}^{n_x}\to \mathbb{R}^{n_x}\) , and \( h:\mathbb{R}^{n_x} \to \mathbb{R}^{n_y}\) are the flow, jump, and output maps, respectively, such that the output is continuous at jumps. Thus, the jumps are not immediately visible in the output, and pairs of points \( (x,g(x))\) for \( x\) in \( D\) are not distinguishable. Nonetheless, in the spirit of [32], we show that there exists a map \( T:\rm{cl}(C)\cup D\to \mathbb{R}^{n_z}\) , such that the image by \( T\) of solutions to system (325) follows the continuous-time dynamics

\[ \begin{equation} \dot{z} = A z + B y, \end{equation} \]

(326)

for some pairs \( (A,B)\in \mathbb{R}^{n_z\times n_z}\times \mathbb{R}^{n_z \times n_y}\) where \( A\) is Hurwitz, and \( T\) is injective on \( \rm{int}(C\setminus D)\) , under a mild backward distinguishability of the system outside of the jump set and some regularity conditions. Then, a trivial observer for the target form (326) is obtained by running system (326) from any initial condition and an estimate for \( x\) can be recovered by a left inversion of \( T\) , with asymptotic convergence except in smaller and smaller intervals around the jump times. This approach addresses the outlined limitations of [32] by offering a constructive method to derive the transformation into continuous-time dynamics while ensuring observer design in the new coordinates. Also, the available tools for the numerical approximation of \( T\) and its left inverse [103, 119] extend to this hybrid framework with some adaptation, taking into account the loss of injectivity of \( T\) on \( D\) , thus providing a systematic observer design for hybrid systems with unknown jump times. Models of the form (325) encompass dynamical systems with state-triggered changes, which includes state-triggered switched systems or hybrid automata, with state \( x = (x_c,q)\) where \( x_c\) is the physical state and \( q\in \mathbb{N}\) encodes the modes, \( f=(f_q,0)\) the continuous dynamics in each mode, and \( g\) and \( D\) the transitions from each mode to the others, that can also depend on \( x_c\) . The theoretical assumptions and results are illustrated throughout this chapter by considering the classical bouncing ball. In Section 4.3, this gluing KKL observer will be applied to estimate the state as well as friction forces in the stiction phenomenon.

After setting the technical background and assumptions of this chapter in Section 4.2.2, we propose in Section 4.2.3 a systematic change of coordinates into the continuous-time system (326), then study its injectivity on \( C\setminus D\) in Section 4.2.4. The possibilities of returning to the initial \( x\) -coordinates are discussed in Section 4.2.5 by revisiting the convergence result of [32] under milder assumptions, where the observer’s robustness against uncertainties is also established. Section 4.2.7 proposes a systematic method for implementing this gluing observer. Finally, conclusions are drawn in Section 4.2.8.

4.2.2 Technical Assumptions

In [10, Definition 2.6], solutions to system (325) are defined in forward positive hybrid time, namely on a hybrid time domain subset of \( \mathbb{R}_{\geq 0}\times \mathbb{N}\) . In this chapter, we need to consider solutions defined both in forward and backward time, namely on a hybrid time domain subset of \( \mathbb{R}\times \mathbb{Z}\) . Thus, we generalize the notion of a hybrid time domain.

Definition 21 ((Compact) hybrid time domain in forward and backward time)

A subset \( E\subset\mathbb{R}\times\mathbb{Z}\) is a compact hybrid time domain if, denoting \( E_{\geq 0} = E \cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) and \( E_{\leq 0} = E \cap (\mathbb{R}_{\leq 0}\times\mathbb{Z}_{\leq 0})\) , we have (when not empty)

\[ \begin{align} E_{\geq 0} &= \bigcup_{j=0}^{J_M-1}\left([t_{j},t_{j+1}]\times \{j\}\right), \\ E_{\leq 0}& = \bigcup_{j=J_m}^{-1}\left([t_{j},t_{j+1}]\times \{j+1\}\right), \\\end{align} \]

(327.a)

for some integers \( J_m\leq 0\) , \( J_M\geq 0\) and a finite sequence of times \( t_{J_m}\leq t_{J_m+1}\leq … \leq t_0=0\leq … \leq t_{J_M-1}\leq t_{J_M}\) in \( \mathbb{R}\) . A set \( E\subset \mathbb{R} \times \mathbb{Z}\) is a hybrid time domain if it is the union of a non-decreasing sequence of compact hybrid time domains, namely, \( E\) is the union of compact hybrid time domains \( E_j\) such that \( … \subset E_{j-1}\subset E_j \subset E_{j+1}…\) .

This definition corresponds to that of [10, Definition 2.6] when \( J_m=0\) . More generally, it coincides with the notion of hybrid time domain with memory introduced in [195], but where \( E_{\leq 0}\) is rather written as

\[ \begin{equation} E_{\leq 0} = \bigcup_{k=1}^{K}\left([s_{k},s_{k-1}]\times \{-k+1\}\right), \end{equation} \]

(328)

with the convention that \( s_k = t_{-k}\) . Inspired from [196, Definitions 6 and 8], we now define solutions to system (325) in forward and backward time.

Definition 22 (Backward version of system (325))

Given system (325), we define its backward version as

\[ \begin{align} \dot{x}&=-f(x) &x&\in C,\\ x^{+}&\in \{x^\prime\in D: x=g(x^\prime)\} & x&\in g(D),\\ y&=h(x).&& \\\end{align} \]

(329.a)

Then, \( \phi:\rm{dom}\phi\to\mathbb{R}^{n_x}\) is solution to system (325) if \( \rm{dom}\phi\) is a hybrid time domain and if, denoting \( \mathcal{D}_{\geq0} = \rm{dom} \phi \cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) and \( \mathcal{D}_{\leq0} = \rm{dom} \phi \cap (\mathbb{R}_{\leq 0}\times\mathbb{Z}_{\leq 0})\) , we have that:

  1. \( \phi_{fw}:=\phi_{|\mathcal{D}_{\geq0}}\) is solution to system (325) on \( \mathcal{D}_{fw}:=\mathcal{D}_{\geq0}\) in the sense of [10, Definition 2.6];
  2. \( \phi_{bw}\) defined on \( \mathcal{D}_{bw}:=-\mathcal{D}_{\leq0}\) as

    \[ \begin{equation} \phi_{bw}(t,j) = \phi(-t,-j), ~~ \forall (t,j) \in \mathcal{D}_{bw}, \end{equation} \]

    (330)

    is a solution to system (329) in the sense of [10, Definition 2.6].

Remark 44

Another way of defining solutions without using [10] would be to say that: \( \phi:\rm{dom}\phi\to\mathbb{R}^{n_x}\) is a solution to system (325) if \( \rm{dom}\phi\) is a hybrid time domain, the map \( t\mapsto \phi(t,j)\) is locally absolutely continuous on each interval \( \mathcal{T}_j=\{t:(t,j)\in\rm{dom}\phi\}\) for all \( j\in\rm{dom}_j \phi\) , and it satisfies the following properties:

  • (a) \( \phi(0,0)\in \rm{cl}(C)\cup D\) ;
  • (b)

    For each \( j\in\rm{dom}_j \phi\) such that \( \rm{int}(\mathcal{T}_j)\neq 0\) , we have

    \[ \begin{equation} \begin{aligned} \dot{\phi}(t,j)&=f(\phi(t,j)),~~&&\text{for almost all } t\in \mathcal{T}_j,\\ \phi(t,j)&\in C,&&\text{for all } t\in \rm{int}(\mathcal{T}_j)\setminus \{0\}; \end{aligned} \end{equation} \]

    (331)

  • (c)

    For each \( (t,j)\in\rm{dom}\phi\) such that \( (t,j+1)\in\rm{dom}\phi\) , we have

    \[ \begin{equation} \begin{aligned} \phi(t,j+1)&=g(\phi(t,j)),\\ \phi(t,j)&\in D. \end{aligned} \end{equation} \]

    (332)

After introducing new definitions related to backward hybrid solutions, we generalize the properties of forward hybrid solutions discussed in Section 1.5. Additionally, the reader is encouraged to revisit the notations in Section 1.5.3 before proceeding. A solution \( \phi\) to system (325) is maximal if there does not exist any other solution \( \phi^\prime\) to system (325) such that \( \rm{dom} \phi\) is a strictly proper subset of \( \rm{dom} \phi^\prime\) and \( \phi=\phi^\prime\) on \( \rm{dom} \phi\) . Besides, \( \phi\) is forward (resp., backward) complete if \( \rm{dom} \phi \cap (\mathbb{R}_{\geq 0}\times \mathbb{Z}_{\geq 0})\) (resp., \( \rm{dom} \phi \cap (\mathbb{R}_{\leq 0}\times \mathbb{Z}_{\leq 0})\) ) is unbounded, \( \phi\) is \( t\) -forward (resp., \( t\) -backward) complete if \( \rm{dom}_t \phi \cap \mathbb{R}_{\geq t}\) (resp., \( \rm{dom}_t \phi \cap \mathbb{R}_{\leq t}\) ) is unbounded (and similarly, for \( j\) -forward and backward completeness). Last, \( \phi\) is Zeno in forward (resp., backward) time if it is forward (resp., backward) complete and \( \sup \rm{dom}_t \phi\) (resp., \( \inf \rm{dom}_t \phi\) ) is bounded.

We are interested in estimating the state of system (325) from the knowledge of the measurement \( y\) . To that end, we make the following assumptions.

Assumption 30

For system (325), assume that:

  1. There exist sets \( \mathcal{X}_0\subset\mathcal{X}\subset \rm{cl}(C)\cup D\) such that all maximal solutions to system (325) initialized in \( \mathcal{X}_0\) are \( t\) -forward complete and remain in \( \mathcal{X}\) in forward time;
  2. For every \( x\in \rm{cl}(C)\cup D\) , there exists a unique maximal solution \( \phi\) to system (325) such that \( \phi(0,0)=x\) ;
  3. The maps \( f\) , \( g\) , and \( h\) are continuous, and

    \[ \begin{equation} h(g(x)) = h(x), ~~ \forall x\in D. \end{equation} \]

    (333)

In Assumption 30, the sets \( \mathcal{X}_0\) and \( \mathcal{X}\) in Item (XXX) could be \( \rm{cl}(C)\cup D\) if no extra information is available about the solutions of interest, but sometimes we know from physical knowledge that they remain within some bounds that can be encoded in \( \mathcal{X}\) , from a certain set of initial conditions \( \mathcal{X}_0\) that may be unknown. The \( t\) -forward completeness condition of the maximal solutions of interest is simply because we propose in this chapter an asymptotic observer design exploiting (326). With Item (XXXII), we say that the output \( y\) does not change at jumps, which in turn means that jumps in the solutions cannot be detected by jumps of the measurement. Item (XXXI) allows us to define uniquely an output map as follows.

Definition 23 (\( t^{-}(x)\) , \( t^{+}(x)\) , and \( Y(x,t)\))

Suppose Item (XXXI) and Item (XXXII) of Assumption 30 hold. For every \( x\in \rm{cl}(C)\cup D\) , given the maximal solution \( \phi\) to system (325) initialized as \( x\) , define

\[ \begin{equation} t^{-}(x):=\inf \rm{dom}_t \phi, ~~ t^{+}(x):=\sup \rm{dom}_t \phi, \end{equation} \]

(334)

and \( Y(x,\cdot):(t^{-}(x),t^{+}(x)) \to\mathbb{R}\) as follows

\[ \begin{equation} Y(x,t)=h(\phi(t,j(t))), ~~ \forall t: (t,j(t))\in\rm{dom} \phi, \end{equation} \]

(335)

where \( j(t) = \min\{j^\prime: (t,j^\prime) \in \rm{dom} \phi\}\) .

Remark 45

With (333) in Item (XXXII) of Assumption 30, \( j(t)\) could be replaced by any \( j\) such that \( (t,j)\in \rm{dom}\phi\) in (335). The map \( Y(x,\cdot)\) is well-defined and continuous for all \( x\in \rm{cl}(C)\cup D\) thanks to (333). Indeed, for any \( t\in \rm{dom}_t\phi\) , \( h(\phi(t,j^\prime))=h(g^{j^\prime-j}(\phi(t_j,j)))=h(\phi(t_j,j))\) for any \( j<j^\prime\) such that \( (t,j)\) and \( (t,j^\prime)\) in \( \rm{dom} \phi\) .

We now propose a KKL-based approach to estimate the state of system (325), which does not require jump detection.

4.2.3 Change of Coordinates into Continuous-Time Dynamics

Following the KKL approach [112] and the gluing idea in [32], we wish to find a map \( T:\rm{cl}(C)\cup D\to\mathbb{R}^{n_z}\) such that i) the image of solutions to system (325) under \( T\) follows continuous-time dynamics of the form \( \dot z = A z + By\) , for appropriate matrices \( A\) and \( B\) , and such that ii) its restriction to the set \( C\setminus D\) is injective, in order to reconstruct from \( z\) an estimate \( \hat{x}\) of \( x\) that converges in the \( x\) -coordinates, at least outside of the jump times.

4.2.3.1 Definition of \( T\)

In order to define \( T\) , we make the following assumption.

Assumption 31

There exists \( \rho>0\) such that for every \( x\in \rm{cl}(C)\cup D\) , the map \( Y(x,\cdot)\) introduced in Definition 23 verifies:

  1. If \( t^{-}(x)\neq -\infty\) , \( \lim_{s\to t^{-}(x)}Y(x,s)\) exists and is finite;
  2. If \( t^{-}(x)= -\infty\) , \( s \mapsto e^{\rho s}Y(x,s)\) is integrable on \( \mathbb{R}_{\leq 0}\) .

The next two lemmas provide sufficient conditions to satisfy this assumption when the maps describing the backward hybrid system (329) are affine or polynomially bounded and exhibit an average dwell time in backward time, or when the solutions of interest are known to remain in a compact set in forward time.

Lemma 16 (Sufficient conditions for Item (XXXIV) of Assumption 31)

Consider system (325). Item (XXXIV) of Assumption 31 holds if the following conditions are satisfied:

  • The map \( g\) is invertible on \( D\) ;
  • There exist non-negative scalars \( (c_f,d_f,c_g,d_g,c_h,d_h,p_h)\) such that:

    \[ \begin{align} |f(x)| &\leq c_f|x| + d_f, & \forall x &\in \rm{cl}(C), \end{align} \]

    (336.a)

    \[ \begin{align} |g^{-1}(x)|& \leq c_g|x| + d_g, & \forall x &\in g(D), \end{align} \]

    (336.b)

    \[ \begin{align} |h(x)|& \leq c_h|x|^{p_h} + d_h, & \forall x &\in \rm{cl}(C) \cup D; \end{align} \]

    (336.c)

  • There exists \( \tau_m > 0\) such that for any \( t\) -backward complete solution \( x\) to system (325), there exists \( N \in \mathbb{N}\) such that we have \( |j| \leq \frac{1}{\tau_m} |t| + N\) for all \( (t,j) \in \rm{dom} x \cap (\mathbb{R}_{\leq 0}\times\mathbb{Z}_{\leq 0})\) .

Proof. This technical proof has been moved to Section 6.3.1 to facilitate reading. \( \blacksquare\)

Example 29 (Application of Lemma 16 to a bouncing ball)

Consider a bouncing ball with height \( x_1\) , velocity \( x_2\) , and restitution coefficient \( c > 0\) , described by

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l} \dot{x}& =& ( x_2, - dx_2^p - a_g), & \text{if } x_1 \geq 0\\ x^+ &=& (x_1, -c x_2 + \mu), & \text{if } x_1 = 0 \text{ and } x_2 \leq 0, \end{array}\right. \end{equation} \]

(337)

where \( a_g = 9.8\) (m\( /\) s\( ^2\) ) is the gravitational acceleration, \( c > 0\) is a restitution coefficent, \( \mu > 0\) is some constant jump input, \( d > 0\) and \( p \in \mathbb{R}_{\geq 0}\) are friction parameters, and with output \( y = x_1\) . Firstly, maximal solutions to system (337) initialized in \( \mathcal{X}_0 = \mathbb{R}_{\geq 0}\times \mathbb{R}\) are both \( t\) - and \( j\) -forward complete and remain in \( \mathcal{X} = \mathcal{X}_0\) in forward time. Secondly, considering the backward version of system (337) given by

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{~~}l} \dot{x}& =& ( -x_2, dx_2^p +a_g), & \text{if } x_1 \geq 0\\ x^+ &=& \left(x_1, \frac{-x_2 + \mu}{c}\right), & \text{if } x_1 = 0 \text{ and } x_2 \geq \mu, \end{array}\right. \end{equation} \]

(338)

we deduce that maximal solutions to system (337) are unique (both in forward and backward time). Thirdly, (333) holds. Therefore, system (337) satisfies Assumption 30. If \( p = 1\) and \( c \leq 1\) , all maximal solutions to (337) in backward time are either \( t\) -backward complete or with a bounded and closed time domain (indeed, no Zeno can happen in system (338) and solutions either end with flow at \( x=(0,x_2)\) with \( 0\leq x_2<\mu\) , or end with one jump to \( (0,0)\) ). For the former type, thanks to linearity in the maps and with \( a_g\) being constant, this system satisfies the conditions in Lemma 16 and Item (XXXIV) of Assumption 31 holds. For the latter, Item (XXXIII) of Assumption 31 holds by continuity. Therefore, Assumption 31 is satisfied in this case. However, with \( p > 1\) , we may have a finite-time escape in backward time, so solutions may not be \( t\) -backward complete and \( Y(x,s)\) may not have a finite limit as \( s \to t^-(x)\) , so Assumption 31 may not hold. This problem is addressed in Example 30.

Lemma 16 guarantees that Item (XXXIV) of Assumption 31 is satisfied for a wide class of systems whose solutions do not grow faster than some uniform exponential rate. Actually, if solutions are bounded in backward time, then the whole Assumption 31 holds directly, as long as \( h\) is locally Lipschitz, as stated next.

Lemma 17 (Sufficient conditions for Assumption 31)

Suppose Assumption 30 holds and \( h\) is locally Lipschitz. If any solution \( \phi\) to system (325) is bounded in backward time, i.e., on \( \rm{dom} \phi\cap (\mathbb{R}_{\leq 0} \times \mathbb{Z}_{\leq 0})\) , then Assumption 31 holds.

Proof. This technical proof has been moved to Section 6.3.2 to facilitate reading. \( \blacksquare\)

Remark 46

If Assumption 30 holds with \( \mathcal{X}\) which is compact and \( h\) is locally Lipschitz, the hybrid system can always be modified to satisfy Assumption 31 by changing the flow and jump sets to \( C_\mathcal{X}:=C \cap (\mathcal{X} + \delta_d)\) and \( D_\mathcal{X}:=D \cap (\mathcal{X} + \delta_d)\) respectively for any \( \delta_d > 0\) . Indeed, this change does not affect the solutions of interest, which evolve in \( \mathcal{X}\) in positive time; Assumption 30 holds unchanged, and the solutions are freely bounded in backward time. The only constraint on this change will be related to distinguishability—see Remark 48.

Example 30 (Application of Lemma 17 to the bouncing ball)

Consider the bouncing ball in Example 29. If \( p = 1\) , \( c > 1\) , and \( \mu = 0\) , solutions are Zeno and bounded in backward time. From Lemma 17Assumption 31 holds. More precisely, Item (XXXIV) holds vacuously since there is no \( t\) -backward complete solution, and for any \( x \in \rm{cl}(C) \cup D\) , \( Y(x,s)\) converges to \( 0\) when \( s \to t^-(x)\) , so Item (XXXIII) holds too. On the other hand, if \( p > 1\) , \( c < 1\) , and \( \mu > 0\) , there may be a finite-time escape in backward time. If solutions are known to remain in some compact set \( \mathcal{X}\) in forward time, by modifying the maps as in Remark 46, we guarantee that Assumption 31 holds. However, if \( p > 1\) and \( c > 1\) , solutions do not remain in any compact set \( \mathcal{X}\) in forward time and may have a finite-time escape in backward time, so neither Lemma 16 nor Lemma 17 holds and we cannot guarantee that Assumption 31 holds.

With \( \rho\) defined in Assumption 31, consider \( n_z\in \mathbb{N}\) , \( A\in\mathbb{R}^{n_z\times n_z}\) such that \( A+\rho \rm{Id}\) is Hurwitz, and \( B\in\mathbb{R}^{n_z \times n_y}\) . We define \( T:\rm{cl}(C)\cup D \to \mathbb{R}^{n_z}\) as

\[ \begin{equation} T(x)=\int_{-\infty}^{0}e^{-As}B\breve{Y}(x,s)ds, \end{equation} \]

(339)

where, for every \( x\in \rm{cl}(C)\cup D\) and \( s\in\mathbb{R}\) ,

\[ \begin{equation} \breve{Y}(x,s)= \left\{ \begin{array}{ll} Y(x,s),&\text{if } s>t^{-}(x),\\ \displaystyle\lim_{\tau\to t^{-}(x)}Y(x,\tau),&\text{otherwise}. \end{array}\right. \end{equation} \]

(340)

Remark 47

While for every \( x\in \rm{cl}(C)\cup D\) , \( s\mapsto Y(x,s)\) is defined on \( (t^{-}(x),t^+(x))\) only, \( s \mapsto \breve{Y}(x,s)\) is defined and continuous on \( \mathbb{R}_{\leq 0}\) according to Assumption 31. Besides, still under Assumption 31, for all \( x\in \rm{cl}(C)\cup D\) , the function \( s\mapsto e^{-As}B\Breve{Y}(x,s)\) is integrable on \( \mathbb{R}_{\leq 0}\) . Thus, it is a continuous extension of \( Y(x,\cdot)\) , allowing us to define \( T\) as in (339).

4.2.3.2 Continuous-Time Dynamics in the \( z\) -Coordinates

Now we prove that the image by \( T\) of solutions to system (325) satisfies (326).

Lemma 18 (Continuous-time stable filter in the \( z\) -coordinates)

Suppose Assumption 30 and Assumption 31 hold. For any maximal solution \( \phi\) to system (325) initialized in \( \mathcal{X}_0\) , there exist a \( C^1\) map \( z:\mathbb{R}_{\geq 0} \to \mathbb{R}^{n_z}\) and a continuous map \( y:\mathbb{R}_{\geq 0} \to \mathbb{R}^{n_y}\) such that

\[ \begin{equation} T(\phi(t,j)) = z(t), ~~ h(\phi(t,j)) = y(t), ~~ \forall (t,j)\in \rm{dom} \phi : t \geq 0, \end{equation} \]

(341)

and

\[ \begin{equation} \dot{z}(t) = A z(t) + By(t), ~~ \forall t \in \mathbb{R}_{\geq 0}, \end{equation} \]

(342)

where \( T\) is defined in (339) with \( A\in\mathbb{R}^{n_z\times n_z}\) such that \( A+\rho \rm{Id}\) is Hurwitz for \( \rho\) in Item (XXXIV) of Assumption 31, and \( B\in\mathbb{R}^{n_z}\) .

Proof. Let \( \phi\) be a maximal solution to system (325). The map \( y:\mathbb{R}_{\geq 0} \to\mathbb{R}^{n_y}\) defined by \( y(t)=h(\phi(t,j))\) for all \( (t,j)\in \rm{dom} \phi\) , \( t\geq 0\) , is well-defined and continuous according to Item (XXXI) and Item (XXXII) of Assumption 30. Then, we exploit Lemma 34 in Section 6.3 to deduce that

\[ \begin{equation} T(g(x)) = T(x), ~~ \forall x \in D. \end{equation} \]

(343)

It follows that there exists a continuous map \( z:\mathbb{R}_{\geq 0} \to \mathbb{R}^{n_z}\) such that \( z(t)=T(\phi(t,j))\) for all \( (t,j)\in \rm{dom} \phi\) with \( t\geq 0\) . Let us show that \( z\) is \( C^1\) . Consider first \( t\in \mathbb{R}_{\geq 0}\) such that \( t\in(t_j,t_{j+1})\) for some \( j\in\rm{dom}_j \phi\) and a scalar \( \Delta \neq 0\) small enough such that \( t+\Delta\in(t_j,t_{j+1})\) . Exploiting the uniqueness of solutions from Item (XXXI) of Assumption 30 and the fact that for all \( x\in \rm{cl}(C)\cup D\) and \( s\in [0,\Delta]\) , we have \( \breve{Y}(x,s)=Y(x,s)\) , we get

\[ \begin{align*} z(t+\Delta)&=\int_{-\infty}^{0}e^{-As}B\breve{Y}(\phi(t+\Delta,j),s)ds\\ &=\int_{-\infty}^{0}e^{-As}B\breve{Y}(\phi(t,j),s+\Delta)ds\\ &=\int_{-\infty}^{\Delta}e^{-A(s^\prime-\Delta)}B\breve{Y}(\phi(t,j),s^\prime)ds^\prime\\ &=e^{A\Delta}\left(\int_{-\infty}^{0}e^{-As^\prime}B\breve{Y}(\phi(t,j),s^\prime)ds^\prime+\int_{0}^{\Delta}e^{-As^\prime}B\breve{Y}(\phi(t,j),s^\prime)ds^\prime\right)\\ &=e^{A\Delta}z(t)+e^{A\Delta}\int_{0}^{\Delta}e^{-As}BY(\phi(t,j),s)ds. \end{align*} \]

Re-arranging the terms, we get

\[ \begin{equation} \frac{z(t+\Delta)-z(t)}{\Delta}=\frac{(e^{A\Delta}-I)}{\Delta}z(t)+\frac{1}{\Delta}e^{A\Delta}\int_{0}^{\Delta}e^{-As}BY(\phi(t,j),s)ds. \end{equation} \]

(344)

Taking the limit as \( \Delta\to 0\) , we obtain that \( z\) is differentiable at time \( t\) and

\[ \begin{align*} \dot{z}(t) &=Az(t)+BY(\phi(t,j),0)\\ &=Az(t)+Bh(\phi(t,j))\\ & =Az(t)+By(t). \end{align*} \]

Now consider any \( t>0\) . By the \( t\) -forward completeness in Item (XXX) of Assumption 30, there exist \( j_1, j_2\in\rm{dom}_j \phi\) such that \( (t-\Delta,j_1)\in \rm{dom} \phi\) and \( (t+\Delta,j_2)\in \rm{dom} \phi\) , for all \( \Delta>0\) sufficiently small. Reproducing the same computations as before, we get

\[ \begin{align*} \lim_{\Delta\to 0^-} \frac{z(t+\Delta)-z(t)}{\Delta} &= A z(t) + B Y(\phi(t,j_1),0)\\ & = A z(t) + B y(t),\\ \lim_{\Delta\to 0^+} \frac{z(t+\Delta)-z(t)}{\Delta} &= A z(t)+ B Y(\phi(t,j_2),0) \\ &= A z(t) + B y(t). \end{align*} \]

Similarly, at \( t=0\) , reasoning with \( \Delta>0\) , we get that \( z\) is continuously differentiable on \( \mathbb{R}_{\geq 0}\) and moreover verifies (342). \( \blacksquare\)

In Lemma 18, we have shown the existence of a map \( T\) transforming system (325) into the continuous-time dynamics (342). It follows that for any solution \( \phi\) to system (325), implementing (342) and

\[ \begin{equation} \dot{\hat{z}}(t) = A \hat{z}(t) + By(t) \end{equation} \]

(345)

fed with the measured output \( y\) , from any initial condition, gives us

\[ \begin{equation} \lim_{t+j\to +\infty} |T(\phi(t,j)) - \hat{z}(t) | = 0, \end{equation} \]

(346)

namely, \( \hat{z}(t)\) provides an asymptotic estimate of \( T(\phi(t,j))\) without any jump detection. The great advantage of this approach compared to [32] is that an observer is directly available in the \( z\) -coordinates, given the specific target form of the dynamics (342). Moreover, the change of coordinates given by \( T\) is guaranteed to exist, with a systematic approach for constructing a numerical model of it (see Section 4.2.7).

Example 31 (Transforming the bouncing ball into continuous-time dynamics)

Consider the bouncing ball in Example 29 with \( d = 0.01\) (m\( ^{-1}\) ), \( p = 2\) , \( c = 0.8\) , and \( \mu = 2\) (m\( /\) s). Exploiting [32] only, we are not able to find an analytic gluing function \( T\) for this system. Instead, we follow the KKL route of this chapter. Taking advantage of the system’s low dimension, we propose to approximate \( T\) using a look-up table. To do this, we simulate the interconnection (325)-(342) from any initial conditions in \( \mathcal{X}_0 \times \mathbb{R}^{n_z}\) , where \( n_z = 3\) with \( A=\rm{diag}(-1,-2,-3)\) and \( B=(1,1,1)\) . After a long enough time compared to the eigenvalues of \( A\) , \( z\) approximates \( T(x)\) and we can form a look-up table consisting of the \( (x,T(x))\) pairs by storing these points taken from a large number of simulations. In Figure 47, we show (left) the data points of the first component of \( z = (z_1,z_2,z_3)\) as a function of \( x=(x_1,x_2)\) taken from the look-up table and plot, along a particular system solution \( (t,j)\mapsto x(t,j)\) , the value \( z_1(t)\) in the look-up table (approximating \( T(x(t,j))\) ) such that its corresponding \( x\) component matches most \( x(t,j)\) in the Euclidean norm. It is confirmed that the values of \( z_1\) before and after each jump are the same, which is consistent with the continuity of \( z\) at the jumps. On the right, we compare \( z_1(t)\) from the look-up table and \( \hat{z}_1(t)\) obtained by running observer (345) from some arbitrary initial condition, showing the convergence in the \( z\) -coordinates. Conditions to deduce an estimate in the \( x\) -coordinates are given in the next section.

Figure 48
Figure 49
Figure 47. Left: Data points \( (x,z_1)\) in look-up table (blue), and \( t\mapsto (x(t,j),z_1(t))\) along a solution (red-yellow), where \( z_1(t)\) is fetched as the closest point in the look-up table; Right: \( t\mapsto z_1(t)\) from look-up table obtained from \( t\mapsto x(t,j)\) vs. \( t\mapsto \hat{z}_1(t)\) solution to observer (345).

4.2.4 Injectivity of \( T\)

As it is usually done in KKL theory, we now exploit a distinguishability property to study the injectivity of \( T\) , and thus the ability to reconstruct \( \phi\) from the knowledge of \( \hat{z}\) . Of course, due to (343), \( T\) is not injective on \( D\) unless the map \( g\) is identity, and we thus focus on \( C\setminus D\) .

Assumption 32

Any distinct points \( x_a,x_b\) in \( \rm{int}(C\setminus D)\) are backward distinguishable, i.e., there exists \( t\in(\max\{t^{-}(x_a),t^{-}(x_b)\},0]\) such that \( Y(x_a,t)\neq Y(x_b,t)\) , with \( Y\) introduced in Definition 23.

Remark 48

If the flow and jump sets of system (325) are modified according to Remark 46 to satisfy Assumption 31, \( \delta_d\) needs to be picked such that backward distinguishability is preserved, i.e., the moment where \( Y(x_a,t)\neq Y(x_b,t)\) , for distinct points \( x_a\) , \( x_b\) in \( \rm{int}(C_\mathcal{X}\setminus D_\mathcal{X})\) , has to happen before solutions starting from \( x_a\) and \( x_b\) leave \( \mathcal{X}+\delta_d\) in backward time.

Using Assumption 32, we now exploit the tools developed in [112] (and the references therein) to show that the map \( T\) is injective on \( \rm{int}(C\setminus D)\) .

Lemma 19 (Injectivity of \( T\))

Suppose Assumptions 3031, and 32 hold. Assume that, for all \( \lambda \in \mathbb{C}_\rho\) with \( \rho\) from Assumption 31, the map

\[ \begin{equation} x \mapsto T_0(\lambda,x) := \int_{-\infty}^0 e^{-\lambda s} \breve{Y}(x,s)ds \end{equation} \]

(347)

is \( C^1\) on \( \rm{int}(C\setminus D)\) ; moreover, for all \( \lambda \in \mathbb{R}_\rho\) and for all \( k \in \mathbb{N}\) , the map \( x \mapsto \frac{\partial^k T_0}{\partial \lambda^k}(\lambda,x) \) exists and is \( C^1\) on \( \rm{int}(C\setminus D)\) . Define \( m_0=2n_x+1\) . For almost any pair of matrices \( (A_0,B_0)\in\mathbb{R}^{m_0\times m_0}\times\mathbb{R}^{m_0}\) with \( A_0+\rho I\) Hurwitz for \( \rho\) in Item (XXXIV) of Assumption 31, the map \( T:C\cup D\to\mathbb{R}^{n_z}\) defined in (339) with \( A=A_0\otimes I_{n_y}\) and \( B=B_0\otimes I_{n_y}\) is injective on \( \rm{int}(C\setminus D)\) .

Proof. Similarly to [112, Appendix B.1], we show that the result is equivalent to showing that for all \( l \in \{0,1,…,n_x\}\) and for almost all \( (\lambda_1,\lambda_2,…, \lambda_{2n_x-l+1}) \in \Omega_{l,\rho}\) , with \( \Omega_{l,\rho} = \mathbb{C}_\rho^l \times \mathbb{R}_\rho\) , the map

\[ \begin{equation} x\mapsto T_{\rm{diag}}(x) = (T_0(\lambda_1,x),T_0(\lambda_2,x),\ldots,T_0(\lambda_{2n_x-l+1},x)) \end{equation} \]

(348)

is injective on \( \rm{int}(C\setminus D)\) . Now we adapt [112, Appendix B.2.3]. Since \( \rm{int}(C\setminus D)\) is open, we use \( \Upsilon = \{(x_a,x_b) \in \rm{int}(C\setminus D) \times \rm{int}(C\setminus D): x_a \neq x_b \}\) an open subset of \( \mathbb{R}^{2n_x}\) , and define the same \( \Theta_i\) , the same \( g_i(\lambda,x_a,x_b)=T_0(\lambda,x_a)-T_0(\lambda,x_b)=\int_{-\infty}^0 e^{-(\lambda+\rho)s}\Delta(x_a,x_b,s) ds \) with

\[ \begin{equation} \Delta(x_a,x_b,s) = e^{\rho s}(\breve{Y}(x_a,s)-\breve{Y}(x_b,s)). \end{equation} \]

(349)

From Assumption 32, for all \( (x_a,x_b) \in \Upsilon\) , by the definition of \( \breve{Y}\) in (340), there exists \( s \leq 0\) such that \( \Delta(x_a,x_b,s) \neq 0\) . By properties of the Laplace transform and continuity of \( s\mapsto \breve{Y}(x,s)\) , we deduce that for all \( (x_a,x_b)\in \Upsilon,\) \( \lambda\mapsto g_i(\lambda,x_a,x_b)\) cannot be identically zero on \( \Omega_{\ell,\rho}\) . Moreover, we can check the regularity conditions of [112, Lemma B.3]: i) for all \( x \in \rm{int}(C\setminus D)\) , \( T_0(\cdot,x)\) is holomorphic on \( \mathbb{C}_\rho\) and is \( C^\infty\) on \( \mathbb{R}_\rho\) , and ii) we have the required regularity of \( T_0\) with respect to \( x\) by assumption. Applying the generalized Coron’s lemma [112, Lemma B.3], we get the results. \( \blacksquare\)

Remark 49

For now, we assume that \( T\) has the necessary regularity to guarantee its injectivity. Note that if the regularity of \( \frac{\partial^k T_0}{\partial \lambda^k}\) is not guaranteed, we can still achieve injectivity of \( T_\rm{{diag}}\) for almost any \( (\lambda_1,\lambda_2,…, \lambda_{n_x+1}) \in \mathbb{C}_\rho^{n_x+1}\) as in [105, Theorem 3], for \( A_0 = \rm{diag}(\lambda_1,…,\lambda_{n_x+1})\) and \( B_0 = (1,…,1) \in \mathbb{R}^{n_x+1}\) .

The reason for considering \( \rm{int}(C\setminus D)\) instead of \( C\setminus D\) or \( \rm{cl}(C)\setminus D\) is that an open set is needed to apply Coron’s lemma, the key tool of [105, 112]. A way around this would be to manage to define \( T\) on a larger open set containing \( C\setminus D\) , but this typically requires us to extend \( Y\) outside of \( \rm{cl}(C)\cup D\) while preserving its regularity. This injectivity is a stepping stone towards the injectivity of \( T\) on \( C\setminus D\) which allows for the construction of an appropriate left inverse of \( T\) and the reconstruction of an estimate \( \hat{x}\) of \( x\) , with robust asymptotic convergence outside the jump times, namely in an appropriate metric taking into account the indistinguishability at jumps (see [178, 23]). The injectivity property of \( T\) on \( C\setminus D\) is thus a key property of the design. While the backward distinguishability property is a very mild assumption in observer design, the regularity of \( T\) assumed in Lemma 19 may seem out of reach given the hybrid nature of system (325). Note that, as explained in Remark 49, injectivity can still be shown when the regularity of \( \frac{\partial^k T_0}{\partial \lambda^k}\) is not guaranteed, for a generic choice of \( n_x+1\) complex eigenvalues, leading to a real filter of dimension \( n_y(2n_x +2)\) instead of \( n_y(2n_x +1)\) . It follows that the key assumption lies in the \( C^1\) regularity of \( T\) and we show in Section 4.2.6 that, surprisingly, such property is actually ensured under checkable sufficient conditions on the hybrid system (325).

Remark 50

It is interesting to note that, when \( n_y=1\) , the generic dimension \( n_z=2n_x+1\) providing injectivity of \( T\) according to Lemma 19 corresponds to the dimension of the gluing function guaranteed to exist in [32, Remark 2] through proper embedding arguments and under additional smoothness and manifold assumptions on the data. This dimension is conservative, to guarantee a generic result impervious to the data of the hybrid system, but it does not seem to be a necessary condition, as illustrated in the examples.

Back to our estimation problem, we are interested in reconstructing \( x(t,j)\) from the knowledge of \( T(x(t,j))\) . The injectivity of \( T\) on \( \rm{int}(C\setminus D)\) suggests it is possible except around the jump times. To formalize the notion of convergence, we use the concept of the gluing function introduced in [32].

Example 32 (Gluing KKL observer for the bouncing ball)

Consider the bouncing ball in Example 29 with parameters of Example 31. Since the whole state is instantaneously observable during flows, Assumption 32 is satisfied. We take \( n_z = 3\) which is seen to give injectivity of \( T\) on \( C\setminus D\) , run observer (345), and recover the estimate in the \( x\) -coordinates by the look-up table built in Example 31. In Figure 50, convergence is recovered in the \( x\) -coordinates, but outside of the jump times when we cannot distinguish between the \( x\) before and after the jumps. Note that for systems of large dimensions, the look-up table approach does not give satisfactory performance due to memory limitations and the curse of dimensionality. Therefore, a systematic learning-based approach is proposed in Section 4.2.7.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results for the bouncing ball system.&lt;/span&gt;
Figure 50. Estimation results for the bouncing ball system.

4.2.5 Robust Convergence in the \( x\) -Coordinates

4.2.5.1 Asymptotic Convergence of the Estimation Error Outside the Jump Times

A gluing function is essentially a function that transforms the hybrid system (325) into continuous-time dynamics, by “gluing” the jump set \( D\) with its image \( g(D)\) , while preserving injectivity on the rest of the domain. The key gluing properties are thus as follows.

Definition 24 (Gluing function for system (325))

A function \( T:\rm{cl}(C)\cup D\to\mathbb{R}^{n_z}\) , with \( n_z\geq n_x\) , is called a gluing function for system (325) if it satisfies:

  1. \( T(x)=T(g(x))\) for all \( x\in D\) ;
  2. \( T\) is injective on \( \rm{cl}(C)\setminus D\) .

Define the set

\[ \begin{equation} \mathcal{A} = \mathcal{A}_0 \cup \mathcal{A}_1 \cup \mathcal{A}_2 \cup \mathcal{A}_3, \end{equation} \]

(350)

with

\[ \begin{align*} \mathcal{A}_0 &= \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times \mathbb{R}^{n_x} : x=\hat{x} \}, \\ \mathcal{A}_1 &= \{(x,\hat{x}) \in D\times \mathbb{R}^{n_x} : g(x)=\hat{x} \}, \\ \mathcal{A}_2 &= \{(x,\hat{x}) \in \mathbb{R}^{n_x}\times D : x=g(\hat{x}) \}, \\ \mathcal{A}_3 &= \{(x,\hat{x}) \in D\times D : g(x)=g(\hat{x}) \}. \end{align*} \]

Note that under uniqueness of solutions in backward time, \( g\) is injective on \( D\) and we have \( \mathcal{A}_3\subset \mathcal{A}_0\) . Next, Lemma 20 shows that the points in \( \mathcal{A}\) are indistinguishable through \( T\) , i.e., they have the same image via \( T\) , and conversely, all the points that are indistinguishable through \( T\) are in \( \mathcal{A}\) .

Lemma 20 (Points in \( \mathcal{A}\) are indistinguishable through \( T\))

Assume that \( g(D)\subset \rm{cl}(C)\setminus D\) and let \( T:\rm{cl}(C)\cup D\to\mathbb{R}^{n_z}\) be a gluing function in the sense of Definition 24. Then, for any \( (x_a,x_b)\in (\rm{cl}(C)\cup D) \times (\rm{cl}(C)\cup D)\) , we have

\[ \begin{equation} T(x_a)=T(x_b) ~ \Longleftrightarrow ~ (x_a,x_b)\in \mathcal{A}. \end{equation} \]

(351)

Proof. Pick \( x_a\) and \( x_b\) in \( \rm{cl}(C)\cup D\) and assume \( T(x_a)=T(x_b)\) . Firstly, if both \( x_a\) and \( x_b\) are in \( \rm{cl}(C)\setminus D\) , we have \( (x_a,x_b)\in \mathcal{A}_0\) from Item (XXXVI). Secondly, if only one is in \( \rm{cl}(C)\setminus D\) , let us say it is \( x_b\) without loss of generality, then \( x_a\in D\) , and by Item (XXXV), we have \( T(x_a)=T(g(x_a))\) . Since \( g(D)\subset \rm{cl}(C)\setminus D\) , we have \( g(x_a)\in \rm{cl}(C)\setminus D\) and \( x_b=g(x_a)\) according to Item (XXXVI). Therefore, we have \( (x_a,x_b)\in \mathcal{A}_1\) . Thirdly, if both \( x_a\) and \( x_b\) are in \( D\) , we have \( T(g(x_a))=T(g(x_b))\) by Item (XXXV), and since both \( g(x_a)\) and \( g(x_b)\) are in \( g(D)\subset \rm{cl}(C)\setminus D\) , we get \( g(x_a)=g(x_b)\) according to Item (XXXV), which gives \( (x_a,x_b)\in \mathcal{A}_3\) . Finally, the other direction of the equivalence is obvious from Item (XXXVI). \( \blacksquare\)

Before formulating the left inverse of \( T\) , we make the following assumption.

Assumption 33

For system (325), assume that:

  1. The set \( \mathcal{X}\) defined in Item (XXX) of Assumption 30, in which the solutions of interest remain in forward time, is compact. Moreover, it verifies \( g(\mathcal{X}\cap D)\subset \mathcal{X}\) ;
  2. \( g(D)\subset \rm{cl}(C)\setminus D\) .

Remark 51

In Assumption 33Item (XXXVII) allows us to have uniform injectivity and continuity properties along solutions thanks to the compactness of \( \mathcal{X}\) . It implies in particular that solutions are bounded in forward time. Moreover, the condition \( g(\mathcal{X}\cap D)\subset \mathcal{X}\) will allow defining a left inverse of \( T\) that takes values in \( \mathcal{X}\) , which in turn will ensure the estimate is in \( \mathcal{X}\) at all times, and uniformity properties can be used to show its convergence. Note that this latter assumption \( g(\mathcal{X}\cap D)\subset \mathcal{X}\) can always be verified by inflating \( \mathcal{X}\) , taking \( \mathcal{X} \cup \rm{cl}(g(\mathcal{X}\cap D))\) instead of \( \mathcal{X}\) . Indeed, \( \rm{cl}(g(\mathcal{X}\cap D))\subset g(\mathcal{X})\) is bounded by the continuity of \( g\) and is closed, so it is compact; besides, \( g(\mathcal{X}\cap D)\subset \rm{cl}(C)\) by Item (XXXVIII) so that \( \rm{cl}(g(\mathcal{X}\cap D))\) is indeed a compact subset of \( \rm{cl}(C)\cup D\) which can be added to \( \mathcal{X}\) to verify Item (XXXVII). On the other hand, Item (XXXVIII) ensures that no consecutive jumps can happen and is instrumental in defining the set \( \mathcal{A}\) to which solutions converge.

As done in [32], with \( \mathcal{X}\) being compact, \( T(\mathcal{X})\) is closed if \( T\) is continuous, and we introduce a projection map \( \Pi_{T(\mathcal{X})}:\mathbb{R}^{n_z}\to T(\mathcal{X})\) that satisfies

\[ \begin{equation} \Pi_{T(\mathcal{X})}(z)\in\left\{z^\prime:\underset{z^\prime\in T(\mathcal{X})}{\rm{argmin}}|z-z^\prime|\right\}, ~~ \forall z\in\mathbb{R}^{n_z}. \end{equation} \]

(352)

We know from Item (XXXVI) of Definition 24 that the restriction of \( T\) to \( \rm{cl}(C)\setminus D\) is injective, so \( T|_\rm{{cl}(C)\setminus D}\) admits a left inverse on \( T(\rm{cl}(C)\setminus D)\) . But actually, by Item (XXXV) of Definition 24, we have \( T(D)=T(g(D))\) , and by Item (XXXVIII) of Assumption 33, we have \( g(D)\subset\rm{cl}(C)\setminus D\) , so that we get \( T(D)\subset T(\rm{cl}(C)\setminus D)\) and

\[ T(\rm{cl}(C)\setminus D)= T(\rm{cl}(C)\setminus D)\cup T(D) = T(\rm{cl}(C)\cup D). \]

It follows that the left inverse of \( T|_\rm{{cl}(C)\setminus D}\) is defined on \( T(\mathcal{X})\subset T(\rm{cl}(C)\cup D)\) . We thus define \( T^* : \mathbb{R}^{n_z} \to \rm{cl}(C)\setminus D\) as

\[ \begin{equation} T^*(z) = T\big|_{\rm{cl}(C)\setminus D}^{-1}\left(\Pi_{T(\mathcal{X})}(z)\right), \end{equation} \]

(353)

which verifies

\[ \begin{equation} T^*(T(x)) = x, ~~ \forall x\in \mathcal{X} \setminus D, \end{equation} \]

(354)

and

\[ \begin{equation} T^*(T(x)) = g(x), ~~ \forall x\in \mathcal{X} \cap D. \end{equation} \]

(355)

We then get the following results.

Theorem 19 (Asymptotic convergence\( /\) stability in the \( x\) -coordinates)

Suppose Item (XXX) of Assumption 30, and Assumption 33 hold. Pick \( n_z\in \mathbb{N}\) , \( (A,B)\in \mathbb{R}^{n_z\times n_z}\times \mathbb{R}^{n_z}\) such that \( A\) is Hurwitz. Assume there exists \( T:\rm{cl}(C)\cup D\to \mathbb{R}^{n_z}\) a continuous gluing function in the sense of Definition 24 such that the conclusion of Lemma 18 holds. Then, for any solution \( x\) to system (325) initialized in \( \mathcal{X}_0\) and any solution to

\[ \begin{equation} \dot{\hat{z}} = A\hat{z} + By, ~~ \hat{x} = T^*(\hat{z}), \end{equation} \]

(356)

initialized in \( \mathbb{R}^{n_z}\) where \( y\) is the corresponding output in system (325), we have

\[ \begin{equation} \lim_{\substack{t+j \to +\infty\\ (t,j) \in \rm{dom} x \cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})}} d_\mathcal{A}((x(t,j),\hat{x}(t))) = 0. \end{equation} \]

(357)

Moreover, there exists a class-\( \mathcal{KL}\) function \( \beta\) such that for any solution \( x\) to system (325) initialized in \( \mathcal{X}_0\) and any solution to observer (356) initialized in \( T(\mathcal{X})\) and fed with the corresponding output \( y\) of system (325), we have

\[ \begin{equation} d_\mathcal{A}((x(t,j),\hat{x}(t))) \leq \beta(d_\mathcal{A}((x(0,0),\hat{x}(0))),t), ~~ \forall (t,j) \in \rm{dom} x \cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0}), \end{equation} \]

(358)

with \( \hat{x}(0)\in \mathcal{X}\) such that \( \hat{z}(0)=T(\hat{x}(0))\) .

Theorem 19 shows that the distance to the indistinguishability set \( \mathcal{A}\) is asymptotically convergent as \( t \to +\infty\) for any \( \hat{z}(0) \in \mathbb{R}^{n_z}\) and is moreover asymptotically stable if \( \hat{z}(0) \in T(\mathcal{X})\) .

Proof. This highly technical proof has been moved to Section 6.3.3 to facilitate reading. It involves showing that \( T^*\) defined in (353) takes values in \( \mathcal{X}\setminus D\) and bringing convergence\( /\) stability from the \( z\) - to the \( x\) -coordinates using comparison functions. \( \blacksquare\)

This result proves the asymptotic convergence of the estimate \( \hat{x}\) given by (356) to the real solution \( x\) of system (325), except around the jump set where the estimate might be a jump ahead or behind \( x\) .

Remark 52

Alternatively, we could have produced similar results using the Lyapunov function \( V: (\rm{cl}(C)\cup D)\times \mathbb{R}^{n_z} \to \mathbb{R}_{\geq 0}\) defined as

\[ \begin{equation} V(x,\hat{z}) = (T(x) - \hat{z})^\top Q (T(x) - \hat{z}), \end{equation} \]

(359)

with \( Q = Q^\top > 0\) solution to \( A^\top Q + QA \leq -\lambda Q\) for some \( \lambda > 0\) , which are guaranteed to exist because \( A\) is Hurwitz. By Lemma 18, this function verifies \( \dot{V}(x,\hat{z}) \leq -\lambda V(x,\hat{z})\) along solutions to the interconnection of the system and observer (325)-(356). Then, we can show the asymptotic convergence of \( (x,\hat{x})\) to the set \( \mathcal{A}\) by the existence of a class-\( \mathcal{K}\) function \( \underline{\alpha}\) such that

\[ \begin{equation} \underline{\alpha}(d_\mathcal{A}(x,\hat{x})) \leq V(x,\hat{z}), ~~ \forall (x,\hat{z}) \in \mathcal{X} \times \mathbb{R}^{n_z}, \end{equation} \]

(360)

where \( \hat{x} = T^*(\hat{z})\) . This existence is proven by reproducing the arguments in Section 6.3.3, but this time in an algebraic way instead of along solutions. Next, the stability of the distance to \( \mathcal{A}\) with respect to its initial value can be shown by the existence of a class-\( \mathcal{K}\) function \( \overline{\alpha}\) such that

\[ \begin{equation} V(x,\hat{z}) \leq \overline{\alpha}(d_\mathcal{A}(x,\hat{x})), ~~ \forall (x,\hat{z}) \in \mathcal{X} \times T(\mathcal{X}), \end{equation} \]

(361)

where \( \hat{x} = T^*(\hat{z})\) . Combining these, we can show the asymptotic stability of the distance from \( (x,\hat{x})\) to \( \mathcal{A}\) in Theorem 19 using Lyapunov’s approach.

Remark 53

In our preliminary work [192], we assume moreover that \( C\) is closed and:

  1. There exist smooth maps \( r_{D}:\mathbb{R}^{n_x}\to\mathbb{R}\) and \( r_{g(D)}:\mathbb{R}^{n_x}\to\mathbb{R}\) satisfying that

    \[ \begin{align} D&=\{x\in C:r_{D}(x)=0\}, \\ g(D)&=\{x\in C:r_{g(D)}(x)=0\}, \\ C&\subset\{x\in\mathbb{R}^{n_x} : r_{D}(x)\leq 0 \text{ and }r_{g(D)}(x)\geq 0\}; \\\end{align} \]

    (362.a)

  2. We have

    \[ \begin{equation} \left\{ \begin{array}{ll} \langle\nabla r_{D}(x), f(x) \rangle>0,&\forall x\in D,\\ \langle\nabla r_{g(D)}(x), f(x)\rangle >0,&\forall x\in g(D). \end{array}\right. \end{equation} \]

    (363)

These conditions allow us to state the results similar to [32], as follows. There exists a class-\( \mathcal{K}\) function \( \alpha\) and \( \epsilon^\star > 0\) such that for any \( 0<\epsilon<\epsilon^\star\) , there exists \( t_\epsilon\geq 0\) such that for any solution \( x\) to system (325) initialized in \( \mathcal{X}_0\) and any solution to observer (356) initialized in \( \mathbb{R}^{n_z}\) with \( y\) the output of system (325), we have

\[ \begin{equation} |x(t,j)-\hat{x}(t)|<\epsilon,~~\forall (t,j)\in \rm{dom} x: t \geq t_\epsilon, t\in\tau_{\alpha}(\epsilon), \end{equation} \]

(364)

where \( \tau_{\alpha}(\epsilon)=\mathbb{R}_{\geq 0}\setminus\underset{j\in\rm{dom}_j x}{\bigcup} [t_j - \alpha(\epsilon), t_j + \alpha(\epsilon)]\) . These additional assumptions guarantee that a solution \( x\) to system (325) cannot stay inside of \( D\) or \( g(D)\) during flows and also forbid the solution from leaving \( \rm{cl}(C)\cup D\) after a jump. It follows that the solutions only pass through the jump set around the (spaced) jump times, and the converge of \( \hat{x}\) to \( x\) is shown outside some intervals around the jump times \( t_j\) , whose length tends to zero as time goes to infinity.

4.2.5.2 Robustness Against Uncertainties

In this part, we assume the uncertainty-free system (325) satisfies all the conditions of Theorem 19, and we consider the corresponding gluing function \( T\) as well as its left inverse \( T^*\) defined in (353). We next study the effects of uncertainties in the observer implementation and performance, namely: i) additive disturbances of the dynamics, \( t \mapsto v_c(t)\) defined on \( \mathbb{R}_{\geq 0}\) and \( (v_{d,j})_{j \in \mathbb{N}}\) respectively during flows and at jumps , ii) additive noise \( t \mapsto w(t)\) defined on \( \mathbb{R}_{\geq 0}\) in the output \( y\) , and iii) imprecise approximation of the inverse map \( T^*\) .

More specifically, consider sets \( \mathfrak{V}_c\times\mathfrak{V}_d\times\mathfrak{W}\) containing the considered disturbances and noise trajectories \( (v_c,v_d,w)\) and system (325) added with those disturbances and noise

\[ \begin{equation} \dot{x}=f(x) + v_c~ x\in C,~~~~ x^{+}=g(x) + v_{d,j}~ x\in D,~~~~ y=h(x) + w. \end{equation} \]

(365)

Since in practice, \( T^*\) typically is not available analytically and is only approximately known via numerical methods, with some error bound \( \delta_T > 0\) , i.e., we assume to know a map \( \tilde{T}^*\) taking value in \( \mathcal{X}\) such that

\[ \begin{equation} |T^*(z) - \tilde{T}^*(z)| \leq \delta_T, ~~ \forall z \in \mathbb{R}^{n_z}. \end{equation} \]

(366)

We consider the observer designed for system (325), but with \( T^*\) replaced by \( \tilde{T}^*\) and the output \( y\) coming from system (365) instead, i.e., we implement

\[ \begin{equation} \dot{\hat{z}} = A\hat{z} + By, ~~ \hat{x} = \tilde{T}^*(\hat{z}). \end{equation} \]

(367)

In order to analyze the performance of observer (367) for system (365), we make the following assumption.

Assumption 34

All maximal solutions to system (365) initialized in \( \mathcal{X}_0\) and with disturbances and noise in \( \mathfrak{V}_c\times\mathfrak{V}_d\times\mathfrak{W}\) , are \( t\) -forward complete and remain in \( \mathcal{X}\) in forward time.

Theorem 20 then shows the robustness of the estimation error in the \( x\) -coordinates with respect to all those uncertainties and in terms of the distance to the set \( \mathcal{A}\) , highlighting the practicality of our results.

Theorem 20 (Robust stability in the \( x\) -coordinates)

Suppose all the conditions of Theorem 19 hold for system (325) and Assumption 34 holds for system (365). Assume that \( T\) in Theorem 19 is \( C^1\) on \( \mathcal{X}\) . There exist positive scalars \( (\lambda,c_1,c_2,c_3,c_4)\) and class-\( \mathcal{K}\) functions \( \alpha\) , \( \alpha^\prime\) such that for any solution \( x\) to system (365) initialized in \( \mathcal{X}_0\) with disturbances and noise in \( \mathfrak{V}_c\times\mathfrak{V}_d\times\mathfrak{W}\) and any solution to observer (367) initialized with \( \hat{z}(0) \in T(\mathcal{X})\) where \( y\) is the output of system (365), we have for all \( (t,j)\in \rm{dom} x \cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{multline} d_\mathcal{A}((x(t,j),\hat{x}(t)))\leq \alpha\bigg(2c_1e^{-\lambda t}\alpha^\prime\left(d_\mathcal{A}((x(0,0),\hat{x}(0)))\right) + 2\max_{s\in [0,t]}e^{-\lambda (t-s)}\left(c_2|v_c(s)| + c_3w(s) \right) \\ + 2c_4\sum_{k = 0}^{j-1}e^{-\lambda (t-t_k)}|v_{d,k}| + L_T\delta_T \bigg), \end{multline} \]

(368)

where \( L_T > 0\) is the Lipschitz constant of \( T\) on \( \mathcal{X}\) and \( \hat{x}(0)\in \mathcal{X}\) such that \( \hat{z}(0)=T(\hat{x}(0))\) .

In Theorem 20, we deduce the robustness of the estimate in terms of the distance \( d_\mathcal{A}\) . Thus, the uncertainties may have a consequence to make \( \hat{x}\) oscillate around \( x\) and \( g(x)\) around the jump times. This robustness comes with the penalization of past disturbances and noise by a forgetting factor, guaranteeing robust stability of the estimation error if the uncertainties vanish after a time. This is the hybrid version of the notion defined in [46] for discrete-time systems, which is stronger than the classical Input-to-State Stability[ISS] in [45].

Proof. This highly technical proof has been moved to Section 6.3.4 to facilitate reading. It involves carefully adapting the proof of Theorem 19 taking into account the uncertainties. \( \blacksquare\)

Example 33 (Robustness of the observer on the bouncing ball)

Consider the bouncing ball in Example 29. Consider some additive noise added to the measured output \( y\) fed to the observer. The dataset \( (x,T(x))\) is still obtained by simulating the system without uncertainties, hence the map \( T\) is still the same, which is injective. Simulation results are shown in Figure 51, illustrating the robustness with respect to the measurement noise.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results for the bouncing ball system in the presence of measurement noise.&lt;/span&gt;
Figure 51. Estimation results for the bouncing ball system in the presence of measurement noise.

4.2.6 Regularity of \( T\)

In this section, we provide sufficient conditions for \( T\) to be \( C^1\) on any open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) . As seen with Lemma 19, this regularity property is constructive towards the injectivity of \( T\) . Towards this goal, we start by providing sufficient conditions to guarantee a \( C^1\) property of the jump times and the values of the solutions after each jump.

Assumption 35

The flow map \( f\) is \( C^1\) and there exists \( \delta>0\) such that the inverse \( g^{-1}\) of the jump map is defined and \( C^1\) on \( g(D)+\delta \mathbb{B}\) . Moreover, define

\[ \begin{equation} \mathfrak{D} := \left\{x\in \rm{cl}(C)\cup D\cup g(D) : t_{-1}(x) \text{ exists} \right\} \end{equation} \]

(369)

where \( t_{-1}(x)\) refers to the first backward jump time of the maximal solution to system (325) initialized as \( x\) , uniquely defined under Item (XXXI) of Assumption 30. Then, \( t_{-1}\) , when defined, is locally the restriction of a \( C^1\) map, namely, for any \( x\in \mathfrak{D}\) , there exist \( \delta_x >0\) and a \( C^1\) map \( t_{-1,x}: \mathbb{B}(x,\delta_x) \to \mathbb{R}\) such that \( t_{-1,x}=t_{-1}\) on \( \mathbb{B}(x,\delta_x)\cap \mathfrak{D}\) .

Remark 54

First, the existence of \( g^{-1}\) on \( g(D)\) typically is imposed by the uniqueness of solutions in backward time assumed in Item (XXXI) of Assumption 30. It is here strengthened slightly outside of \( g(D)\) to properly define the \( C^1\) property. Then, the fact that \( t_{-1}\) , when defined, is locally the restriction of a \( C^1\) map can either be checked analytically when this map is computable, or under the following sufficient condition: there exists a \( C^1\) map \( \omega: \mathbb{R}^{n_x} \to \mathbb{R}\) characterizing the first jump in backward time, namely such that for all \( x_0 \in \mathfrak{D}\) , there exist neighborhoods \( U\subset \mathbb{R}^{n}\) and \( V\subset \mathbb{R}\) of \( x_0\) and \( t_{-1}(x_0)\) respectively such that for all \( x\in U\cap \mathfrak{D}\) and for all \( s\in V\) ,

\[ \begin{equation} \omega(\Psi_f(x,s)) = 0 ~ \Longleftrightarrow ~ s=t_{-1}(x), \end{equation} \]

(370)

and the following transversality condition holds

\[ \begin{equation} \frac{\partial \omega}{\partial x}(x) f(x) \neq 0, ~~ \forall x \in g(D). \end{equation} \]

(371)

Indeed, pick any \( x\in \mathfrak{D}\) . Then, by definition of \( t_{-1}\) , \( \Psi_f(x,t_{-1}(x))\in g(D)\) and by the \( C^1\) property of \( f\) , there exist neighborhoods \( U\) and \( V\) of \( x\) and \( t_{-1}(x)\) such that \( \Psi_f\) is defined and \( C^1\) on \( U\times V\) . By the transversality condition (371) and from the implicit function theorem, there exists a \( C^1\) map \( \mathfrak{T}:U\to V\) such that for \( (x,s)\in U\times V\) \( \omega(\Psi_f(x,s)) = 0\) is equivalent to \( s=\mathfrak{T}(x)\) . By (370), \( t_{-1}\) is locally the restriction of \( \mathfrak{T}\) to \( \mathfrak{D}\) .

Lemma 21 (\( C^1\) of \( x\mapsto t_j(x)\) and \( x\mapsto X(x,t_{j}(x),j)\))

Under Assumption 35, for any open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) and for any \( j\in \mathbb{Z}_{< 0}\) , if the maps \( x\mapsto t_j(x)\) and \( x\mapsto X(x,t_{j}(x),j)\) are defined on \( \mathcal{O}\) , then they are \( C^1\) on \( \mathcal{O}\) .

Proof. Pick an open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) . Notice that if \( x\mapsto t_j(x)\) or \( x\mapsto X(x,t_{j}(x),j)\) is defined on \( \mathcal{O}\) for some \( j\in \mathbb{Z}_{< 0}\) , then \( x\mapsto t_{j^\prime}(x)\) and \( x\mapsto X(x,t_{j^\prime}(x),j^\prime)\) are defined on \( \mathcal{O}\) for all \( j^\prime \in \mathbb{Z}_{> j} \cap \mathbb{Z}_{< 0}\) . We can thus proceed by induction on \( j\) . From [197, Lemma 4.2.2], because \( f\) is \( C^1\) , the map \( (x,s) \mapsto \Psi_f(x,s)\) is also \( C^1\) on its (open) domain of definition. First, if \( x\mapsto t_{-1}(x)\) is defined on \( \mathcal{O}\) , then it is \( C^1\) on \( \mathcal{O}\) by Assumption 35. Second, if \( x\mapsto X(x,t_{-1}(x),-1)\) is defined on \( \mathcal{O}\) , then \( X(x,t_{-1}(x),-1)=g^{-1}(\Psi_f(x,t_{-1}(x)))\) on \( \mathcal{O}\) with \( \Psi_f(x,t_{-1}(x))\in g(D)\) , so from Assumption 35, \( x\mapsto X(x,t_{-1}(x),-1)\) is \( C^1\) on \( \mathcal{O}\) . Third, if \( x\mapsto t_{-2}(x)\) is defined on \( \mathcal{O}\) , then \( t_{-2}(x) = t_{-1}(x) + t_{-1}(g^{-1}(\Psi_f(x,t_{-1}(x))))\) on \( \mathcal{O}\) which is thus also \( C^1\) on \( \mathcal{O}\) by Assumption 35. By induction, it follows that all the maps \( x \mapsto t_j(x)\) and \( x \mapsto X(x,t_j(x),j)\) , for \( j \in \mathbb{Z}_{\leq 0}\) as long as defined, are \( C^1\) on \( \mathcal{O}\) . \( \blacksquare\)

Example 34 (Application of Lemma 21 to the bouncing ball)

Consider the bouncing ball in Example 29, but with \( d = 0\) for simplicity. Consider an open set \( \mathcal{O} \subset \rm{int}(C\cup g(D))\) on which the maps \( x\mapsto t_j(x)\) and \( x\mapsto X(x,t_{j}(x),j)\) are defined for \( j\in \mathbb{Z}_{\leq 0}\) with \( j_m\leq j \leq 0\) for some \( j_m \in \mathbb{Z}_{\leq 0}\cup\{-\infty\}\) . For this system, we have that \( f\) is \( C^1\) , and \( g\) is invertible with \( g^{-1}\) being \( C^1\) . The jump condition is characterized by the map \( \omega\) defined as \( \omega(x)=x_1\) , which is \( C^1\) . Indeed, the map \( t_{-1}\) is obtained by solving

\[ \begin{equation} \omega(\Psi_{f}(x,s)) = -a_g\frac{s^2}{2} + x_2 s + x_1 = 0, \end{equation} \]

(372)

for a negative time \( s\) , giving us

\[ \begin{equation} t_{-1}(x) = \frac{1}{a_g} \left(x_2 - \sqrt{x_2^2+2a_g x_1}\right) <0. \end{equation} \]

(373)

This is \( C^1\) on \( \mathcal{O}\) because \( x_1\) and \( x_2\) are not zero at the same time in \( \mathcal{O}\) . Therefore, Assumption 35 holds following Remark 54. Moreover, since \( d = 0\) , we saw in Example 29 that backward solutions are either \( t\) - and \( j\) -complete, or have a bounded and closed time domain (ending with a flow or jump at \( (0,x_2)\) with \( x_2\in [0,\mu]\) ). From Lemma 21, we deduce that all the maps \( x \mapsto t_j(x)\) and \( x \mapsto X(x,t_j(x),j)\) , as long as they are defined, are \( C^1\) on \( \mathcal{O}\) . Let us now test the \( C^1\) analytically. Based on the backward system (338), we then have

\[ \begin{equation} X(x,t_{-1}(x),0) = \begin{pmatrix} -a_g\frac{(t_{-1}(x))^2}{2} + x_2 t_{-1}(x) + x_1 \\ -a_g t_{-1}(x) + x_2 \end{pmatrix} = \begin{pmatrix} 0 \\ \sqrt{x_2^2+2a_g x_1} \end{pmatrix}, \end{equation} \]

(374)

and then, if \( \sqrt{x_2^2+2a_g x_1} \geq \mu\) ,

\[ \begin{equation} X(x,t_{-1}(x),-1) = \begin{pmatrix} 0 \\ \frac{-\sqrt{x_2^2+2a_g x_1} + \mu}{c} \end{pmatrix}, \end{equation} \]

(375)

which is \( C^1\) on \( \mathcal{O}\) because \( t_{-1}\) is \( C^1\) on \( \mathcal{O}\) , again because \( x_1\) and \( x_2\) are not zero at the same time in \( \mathcal{O}\) . This is then used to deduce that \( t_{-2}\) defined as

\[ \begin{align*} t_{-2}(x)& = t_{-1}(x) + t_{-1}(X(x,t_{-1}(x),-1)) \\ &= t_{-1}(x) + \frac{1}{a_g} \left(\frac{-\sqrt{x_2^2+2a_g x_1} + \mu}{c}- \sqrt{\left(\frac{-\sqrt{x_2^2+2a_g x_1} + \mu}{c}\right)^2}\right) \\ &= t_{-1}(x) + \frac{1}{a_g} \left(\frac{-\sqrt{x_2^2+2a_g x_1} + \mu}{c}- \left|\frac{-\sqrt{x_2^2+2a_g x_1} + \mu}{c}\right|\right) \\ & =t_{-1}(x) + \frac{2}{a_g c}\left( -\sqrt{x_2^2+2a_g x_1} + \mu\right) < 0 \end{align*} \]

is \( C^1\) on \( \mathcal{O}\) . Repeating this computation and reasoning for \( j = -2,-3,…\) , we get analytically that all the maps \( x \mapsto t_j(x)\) and \( x \mapsto X(x,t_j(x),j)\) , for all \( j \in \mathbb{Z}_{\leq 0}\) as long as these maps are defined, are all affine functions of the velocity before the first backward jump \( \sqrt{x_2^2+2a_g x_1}\) and are thus \( C^1\) on \( \mathcal{O}\) .

From Lemma 21 and assuming that \( h\) is also \( C^1\) , we know that for any \( j\in \mathbb{Z}_{\leq 0}\) as long as \( x\mapsto t_j(x)\) and \( x\mapsto X(x,t_{j}(x),j)\) are defined and for any \( s\in [t_{j-1}(x),t_j(x)]\) , the Jacobian matrices \( x\mapsto \frac{\partial t_{j}}{\partial x}(x)\) and \( x\mapsto \frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{j}(x),j),s)))\) are well-defined and bounded on any compact subset of any open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) . However, to further prove the regularity of \( T\) , we need additional assumptions on the uniformity of those bounds with respect to \( j\) and \( s\) . Because the proofs are performed on a neighborhood of a point in \( \mathcal{O}\) , in the following we will consider two cases, where the solutions starting from a neighborhood of any \( x \in \mathcal{O}\) are either all \( j\) -backward incomplete with a locally uniform number of jumps (Case 1), or all \( j\) -backward complete (Case 2).

4.2.6.1 Case 1: Solutions from \( \mathcal{O}\) are \( j\) -Backward Incomplete

We start with the case where solutions from \( \mathcal{O}\) are \( j\) -backward incomplete. We dissociate in the next assumptions the case where solutions are \( t\) -backward complete, and the case where solutions are besides also \( t\) -backward incomplete, leading to a compact hybrid time domain.

Assumption 36

Given an open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) , assume that for any \( x \in \mathcal{O}\) :

  1. There exist a neighborhood \( \mathcal{C} \subset \mathcal{O}\) of \( x\) and \( J_m \in \mathbb{Z}_{\leq 0}\) such that solutions \( \phi\) to system (325) initialized in \( \mathcal{C}\) are \( t\) -backward complete, i.e., \( \inf \rm{dom}_t \phi = -\infty\) , and are \( j\) -backward incomplete with \( \min\rm{dom}_j \phi = J_m\) ;
  2. There exist \( \rho > 0\) such that for any compact set \( \mathcal{C} \subset \mathcal{O}\) , there exist \( c > 0\) such that for all \( x \in \mathcal{C}\) and for all \( s \in \mathbb{R}_{\leq 0}\) , we have \( |Y(x,s)| \leq c e^{-\rho s}\) ;
  3. There exist \( \rho_h,\rho_f,\rho^\prime \geq 0\) such that for any compact set \( \mathcal{C} \subset \mathcal{O}\) , there exist \( c_h,c_f \geq 0\) such that for all \( x \in \mathcal{C}\) and for all \( s \in (-\infty,t_{J_m}(x)]\) , the following partial derivatives are well-defined and:

    \[ \left|\frac{\partial h}{\partial x} (X(x,s, J_m))\right| \leq c_h e^{-\rho_h s}, ~ \left|\frac{\partial f}{\partial x} (X(x,s, J_m))\right| \leq \rho^\prime, ~ \left|f(X(x,s, J_m))\right| \leq c_f e^{-\rho_f s}. \]

Remark 55

In Assumption 36Item (XLIII) should be checkable given the maps of system (325). In particular, this item holds automatically with \( \rho_h=\rho_f=0\) if the solutions initialized in \( \mathcal{O}\) remain in a compact set in backward time.

Theorem 21 then gives sufficient conditions for \( T\) to be \( C^1\) on open subsets of \( \rm{cl}(C) \cup D\) , when solutions from this subset are \( j\) -backward incomplete.

Theorem 21 (\( C^1\) of \( T\) on \( \mathcal{O}\) in Case \( \mathbf{1}\))

Suppose Assumptions 3031, and 35 hold. Assume that \( h\) is \( C^1\) . For any open set \( \mathcal{O} \subset \rm{cl}(C)\cup D\) satisfying Assumption 36, there exists \( \rho_0 > 0\) such that for any pair of matrices \( (A,B)\) with \( A+ \rho_0\rm{Id}\) Hurwitz, the map \( T: \rm{cl}(C)\cup D\to\mathbb{R}^{n_z}\) defined in (339) is \( C^1\) on \( \mathcal{O}\) .

Proof. This highly technical proof has been moved to Section 6.3.5 to facilitate reading. It involves two main steps: i) Split \( T\) into a part stopping at \( t_{J_m(x)}\) and another part from \( t_{J_m(x)}\) to \( -\infty\) and show that the first one is \( C^1\) , and ii) Show that the second is \( C^1\) by making it converge uniformly under the \( C^1\) -norm when \( A\) is pushed sufficiently Hurwitz. \( \blacksquare\)

4.2.6.2 Case 2: Solutions from \( \mathcal{O}\) are \( j\) -Backward Complete

In the case where solutions from \( \mathcal{O}\) are \( j\) -backward complete, we start by giving Assumption 37.

Assumption 37

Given any open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) , assume that:

  1. Solutions to system (325) initialized in \( \mathcal{O}\) are \( j\) -backward complete;
  2. There exist \( \tau_m > 0\) and \( N \in \mathbb{N}\) such that for any solution \( \phi\) to system (325) initialized in \( \mathcal{O}\) , we have \( |j| \leq \frac{1}{\tau_m} |t| + N\) for all \( (t,j) \in \rm{dom} \phi \cap (\mathbb{R}_{\leq 0}\times\mathbb{Z}_{\leq 0})\) ;

Moreover, assume that there exist \( \rho,\rho_t,\rho_t^\prime,\rho_X,\rho_X^\prime,\rho_h,\rho_h^\prime,\rho^\prime \geq 0\) such that for any compact set \( \mathcal{C} \subset \mathcal{O}\) , there exist \( c,c_t,c_X,c_h \geq 0\) such that for all \( x \in \mathcal{C}\) , the following partial derivatives are well-defined and:

  1. For all \( s \in \mathbb{R}_{\leq 0}\) , we have \( |Y(x,s)| \leq c e^{-\rho s}\) ;
  2. For all \( j \in \mathbb{Z}_{\leq 0}\) , we have \( \left|\frac{\partial t_{j}}{\partial x}(x)\right| \leq c_t e^{- \rho_t j -\rho_t^\prime t_j(x)}\) ;
  3. For all \( j \in \mathbb{Z}_{\leq 0}\) , we have \( \left|\frac{\partial }{\partial x}(X(x,t_j(x),j))\right| \leq c_X e^{-\rho_X j - \rho_X^\prime t_j(x)}\) ;
  4. For all \( j \in \mathbb{Z}_{\leq 0}\) and for all \( s \in [t_{j-1}(x),t_j(x)]\) , we have \( \left|\frac{\partial h}{\partial x}(X(x,s,j))\right| \leq c_h e^{-\rho_h j - \rho_h^\prime t_j(x)}\) ;
  5. For all \( j \in \mathbb{Z}_{\leq 0}\) and for all \( s \in [t_{j-1}(x),t_j(x)]\) , we have \( \left|\frac{\partial f}{\partial x}(X(x,s,j))\right| \leq \rho^\prime\) .

Remark 56

In Assumption 37Item (XLV) is about the existence of a uniform average dwell time in backward time, which differs from the third condition of Lemma 16 by the extra uniformity of \( N\) with respect to \( x\) . Note that from Lemma 21, the maps \( x \mapsto t_j(x)\) and \( x \mapsto X(x,t_j(x),j)\) are \( C^1\) on the compact set \( \mathcal{C}\) and so the maps \( x \mapsto \frac{\partial t_{j}}{\partial x}(x)\) and \( x \mapsto \frac{\partial}{\partial x}(X(x,t_j(x),j))\) , are continuous thus bounded on \( \mathcal{C}\) for each \( j \in \mathbb{Z}_{\leq 0}\) ; Item (XLVII) and Item (XLVIII) are thus mainly about the dependence of these bounds on \( j\in \mathbb{Z}_{\leq 0}\) and on \( t_j \in \rm{dom}_t \phi \cap \mathbb{R}_{\leq 0}\) , which is allowed to grow exponentially but with uniform rates \( \rho_t,\rho_t^\prime\) and \( \rho_X,\rho_X^\prime\) .

Using Assumption 37, we now show that the map \( T\) is \( C^1\) on \( \mathcal{O}\) .

Theorem 22 (\( C^1\) of \( T\) on \( \mathcal{O}\) in Case \( \mathbf{2}\))

Suppose Assumptions 3031, and 35 hold. Assume that \( h\) is \( C^1\) . For any open set \( \mathcal{O} \subset \rm{cl}(C) \cup D\) satisfying Assumption 37, there exists \( \rho_0 > 0\) such that for any pair of matrices \( (A,B)\) with \( A+ \rho_0\rm{Id}\) Hurwitz, the map \( T: \rm{cl}(C)\cup D\to\mathbb{R}^{n_z}\) defined in (339) is \( C^1\) on \( \mathcal{O}\) .

Proof. This technical proof has been moved to Section 6.3.6 to facilitate reading. It involves two main steps: i) Express \( T\) as a function sequence and show that each function is \( C^1\) , and ii) Show that the sequence converges uniformly under the \( C^1\) -norm when \( A\) is pushed sufficiently Hurwitz. \( \blacksquare\)

4.2.7 A Systematic Implementation Method

To facilitate the practical implementation of our observer, we propose a systematic method to approximate the left inverse of \( T\) , which serves to recover \( \hat{x}(t)\) from \( \hat{z}(t)\) in (356). As demonstrated in Example 32, a simple look-up table may suffice for low-dimensional cases. However, when dealing with dimensions of three or more, the curse of dimensionality becomes an issue, necessitating more advanced methods such as using a Neural Network[NN]. It is important to note that, even if a dataset of points \( (x,T(x))\) is available, the lack of injectivity on \( D\) can degrade NN training performance. Specifically, points \( (x, g(x))\) for \( x \in D\) have the same image through \( T\) . When training a single NN to learn the inverse map, the learning algorithm tends to map the point \( T(x) = T(g(x))\) to a point that lies between \( x\) and \( g(x)\) , which could be far from both, thus degrading efficiency. It follows that standard numerical methods developed for the continuous-time or discrete-time KKL observer [117, 118] fail to apply directly. To address this, we recommend training two separate NNs: one with data containing \( x\) and the other with data containing \( g(x)\) , for \( x \in D\) . This ensures that each NN focuses solely on its respective data, considerably improving learning efficiency.

Our systematic implementation method, based on splitting the data, is illustrated in Figure 52 (via the bouncing ball) and involves the following steps:

  • (Data generation) Following the approach in [118], simulate the concatenation (325)-(326) from a large number of initial conditions and discard the first part of the solutions \( (x,z)\) corresponding to the transient phase of the filter (326), which is determined by \( A\) . The remaining points \( (x(t, j), z(t, j))\) approximate \( (x(t, j), T(x(t, j)))\) after the transient phase, thanks to Lemma 18 and the contracting nature of the \( z\) -dynamics;
  • (Classification into two groups) Store the data points \( (x,T(x))\) in two distinct datasets, corresponding to the part of the solution in the first and second half of the flow intervals. More precisely, all the points \( (x(t,j),z(t,j))\) belonging to the first half of each flow interval, namely with \( t-t_j \leq \frac{t_{j+1}-t_j}{2}\) , are stored in one dataset; all the points \( (x(t,j),z(t,j))\) belonging to the second half of each flow interval, namely with \( t-t_j > \frac{t_{j+1}-t_j}{2}\) , are stored in the other dataset. As the jumps are assumed not to occur twice at the same instant, the dataset taken from the second half of the flow intervals, corresponding to points before the jumps, includes \( x \in D\) but not \( g(x)\) , and similarly, the dataset from the first half, corresponding to points after the jumps, does not contain \( x \in D\) but only \( g(x)\) , effectively separating these data;
  • (Network structure) After appropriately pre-processing the data (normalization, outlier removal, etc.), train two regression NNs, one on each dataset to learn two inverse maps giving \( \hat{x}\) from \( \hat{z}\) , and a classification NN using both \( z\) -datasets, with a label on each point \( z\) indicating whether it belongs to the first or second half of flow intervals. During implementation, the classifier determines from \( \hat{z}(t)\) whether we are in the first or second half of the flow interval (i.e., after or before a jump), and the corresponding regressor is applied as the inverse map.
&lt;span data-controller=&quot;mathjax&quot;&gt;A systematic implementation approach based on splitting the data.&lt;/span&gt;
Figure 52. A systematic implementation approach based on splitting the data.

The classifier, trained using the combined dataset, is permitted to make mistakes due to the following reasons:

  • In the middle of the flow interval, both datasets yield nearly identical inverse maps. Therefore, the classifier’s choice of regressor has little impact on overall accuracy, with the outputs of both regressors being almost the same;
  • At the jumps, where the points \( x\) and \( g(x)\) for \( x \in D\) are indistinguishable, either one could be selected without improving the results, as distinguishing between them is inherently impossible.

Example 35 (Systematic gluing KKL observer for the bouncing ball)

As done in Example 32, consider the bouncing ball in Example 29 with parameters of Example 31. This time, we apply the systematic implementation scheme outlined in this section, replacing the look-up table with two simple NNs. We see from Figure 53 that our method works well for this system, with enhanced performance compared to Example 32. This method will be applied in Section 4.3 to a much more complicated and higher-dimensional problem of stick-slip.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results for the bouncing ball system using our systematic implementation scheme.&lt;/span&gt;
Figure 53. Estimation results for the bouncing ball system using our systematic implementation scheme.

4.2.8 Conclusion

In this chapter, we propose a systematic and novel observer design for hybrid systems with unknown jump times, using the KKL observer paradigm and the concept of “gluing” the jumps. We demonstrate the existence of a transformation that converts a hybrid system with continuous outputs at jump times into a continuous-time stable filter of this output, for which an observer can always be designed without requiring jump detection. Under the assumptions of backward distinguishability and regularity of this transformation, we prove that the transformation is injective and therefore left-invertible outside the jump set, allowing the recovery of the estimate in the original coordinates. We then show the convergence of this estimate in terms of distance to a set modeling the possibility of an early or delayed jump around the jump times. The robustness of the observer against uncertainties is also demonstrated in terms of this distance. We also provide a systematic method for learning the inverse transformation from data generated by offline simulations. Our theoretical contributions are illustrated through academic examples involving the bouncing ball. In Section 4.3, this gluing KKL observer will be applied for state and parameter estimation in the stiction phenomenon.

Acknowledgments. We thank Valentin Alleaume, Ph.D. student at Mines Paris - PSL, for his collaboration in the theoretical part. We also thank Sergio Garcia and Zikang Zhu, students at Mines Paris - PSL, for their collaborations in the theoretical and numerical parts, respectively.

4.3 Application to the Stick-Slip Phenomenon

 

Dans ce chapitre, nous appliquons l’observateur Kravaris-Kazantzis$\slash$Luenberger[KKL] à “collage” ou “gluing” pour l’estimation d’état et de paramètres dans le phénomène de stick-slip, fréquemment rencontré dans le forage pétrolier. Nous modélisons d’abord ce phénomène non-lisse comme un système hybride autonome avec des instants de saut inconnus. Après une analyse de l’observabilité, nous concevons un observateur KKL à collage pour ce système, en utilisant un réseau de neurones pour approximer la transformation inverse de KKL à l’aide de données obtenues à partir de simulations hors ligne. Nous fournissons également des recommandations pour un prétraitement efficace des données et l’apprentissage de la fonction inverse, aboutissant à une synthèse systématique.

4.3.1 Introduction

A preliminary part of this chapter has been published in [192]; the full version will be submitted to a journal.

In this chapter, we apply the gluing Kravaris-Kazantzis$\slash$Luenberger[KKL] observer developed in Section 4.2 for state and parameter estimation in a system exhibiting the stick-slip phenomenon commonly encountered in oil drilling. First, we model these non-smooth dynamics as an autonomous seven-dimensional hybrid system with unknown jump times, for which classical gluing techniques like [191, 32] face difficulty. Following an observability analysis, we design a gluing KKL observer for this system following the systematic implementation method proposed in Section 4.2.7, utilizing a Neural Network[NN] to approximate the inverse KKL transformation using data obtained from offline simulations.

4.3.2 Hybrid Modeling of a Rotary Drilling System with Dry Friction

We study the stick-slip phenomenon encountered, e.g., in rotary drilling [198]. In this process, a hole is created several kilometers into the ground by a drill bit connected to the surface actuators by a series of pipes called the drill string. Only surface real-time measurements are usually available, and the estimation of downhole conditions is of paramount importance to improve efficiency and reduce failure. Illustrated in Figure 54, a simplified model of the rotational dynamics consists of two masses (top and bottom) connected by a torsional spring of stiffness \( k\) . The equations of motion and the output read

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\Delta \theta} &=& \omega_1 - \omega_2\\ \dot{\omega}_1 & =& -k\Delta \theta +u \\ \dot{\omega}_2 &=& \gamma k \Delta \theta + \text{Dry Friction} \end{array}\right. ~~ y = \omega_1, \end{equation} \]

(376)

where \( \omega_1\) (resp., \( \omega_2\) ) is the top (resp., bottom) velocity of the bit string, and \( \Delta \theta\) models the distortion of the string. At the top, \( \omega_1\) is measured and regulated to \( \omega_\rm{ ref}\) via a Proportional-Integral[PI] controller \( u=-k_P(\omega_1-\omega_\rm{ ref})-k_I \eta\) where \( \dot{\eta} = \omega_1-\omega_\rm{ ref}\) . The parameters that we use correspond to a \( 2700\) -meter long inclined well and read \( \gamma = 3.25\) , \( k = 0.08\) (s\( ^{-1}\) ), \( k_P = 1.87\)  (s\( ^{-1}\) ), \( k_I = 7\) (s\( ^{-2}\) ), and \( \omega_\rm{ ref} = 1\) (rad\( /\) s).

&lt;span data-controller=&quot;mathjax&quot;&gt;Illustration of the drilling system (figure taken and modified from &lt;a href=&quot;https://www.softtorque.com/products/&quot; target=&quot;_blank&quot;&gt;https://www.softtorque.com/products/&lt;/a&gt;).&lt;/span&gt;
Figure 54. Illustration of the drilling system (figure taken and modified from https://www.softtorque.com/products/).

High-gain PI controllers and large static-to-dynamic friction ratios typically induce undesirable stick-slip limit cycles, i.e., periodic trajectories alternating stick and slip phases. To model the dry friction between the string and the walls, we use a two-parameter stiction model which cannot be seen as a differential inclusion and requires some switching\( /\) hybrid logic [199, Section 4.2]. This model has been seen to reproduce industrial data [200]. The stick-slip phenomenon is described as follows. Consider the static and dynamic friction forces \( F_s\) and \( F_d\) that are unknown but are assumed to belong to \( [0,10]\) (N) and \( [0,5]\) (N), respectively, with the constraint that \( F_d\leq F_s\) . While \( \omega_2> 0\) , the friction equals \( -F_d\) , and when the velocity decreases to \( 0\) (with \( \dot{\omega}_2<0\) and thus \( \gamma k \Delta\theta\leq F_d\) ), \( \omega_2\) may either stick with \( \omega_2=0\) , if the external force \( \gamma k \Delta\theta\in [-F_s,F_d]\) is not high enough to win over friction, or slip with \( \omega_2<0\) if \( \gamma k \Delta\theta \leq -F_s\) . This happens symmetrically for \( \omega_2<0\) . Then, once it has stuck, \( \omega_2\) may slip again with \( \omega_2> 0\) (resp., \( \omega_2<0\) ) only if the external force \( \gamma k \Delta\theta\) overcomes static friction, i.e., becomes larger than \( F_s\) (resp., smaller than \( -F_s\) ). Since the mode \( q\) depends on the rest of the states, this constitutes a hybrid automaton or switched system with unknown switching times.

All in all, these dynamics can be modeled using a hybrid system of the form (325) with state

\[ \begin{equation} x=(\Delta \theta, \omega_{1}, \omega_{2},q,\eta,F_s,F_d), \end{equation} \]

(377)

where \( q\) is a logic variable that is \( 0\) in the stick phase, \( 1\) in the forward slip phase, and \( -1\) in the backward slip phase, with the flow dynamics

\[ \begin{equation} \dot{x}= \left\{ \begin{array}{ll}f_{\pm 1}(x), & \text { if } x \in C_{1} \cup C_{-1} \\ f_{0}(x), & \text { if } x \in C_{0},\end{array}\right. \end{equation} \]

(378.a)

where

\[ \begin{align*} f_{\pm 1}(x) & = (\omega_1-\omega_2,-k \Delta \theta-k_P(\omega_1-\omega_{\rm ref})-k_I \eta, \gamma k \Delta \theta-qF_d,0,\omega_1-\omega_{\rm ref},0,0),\\ f_{0}(x) &=(\omega_1-\omega_2,-k \Delta \theta-k_P(\omega_1-\omega_{\rm ref})-k_I \eta, 0,0,\omega_1-\omega_{\rm ref},0,0), \end{align*} \]

with the flow sets

\[ \begin{align*} C_{0}&=\left[-\frac{F_s}{\gamma k}, \frac{F_s}{\gamma k}\right] \times \mathbb{R} \times\{0\} \times\{0\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0}, \\ C_{1}&=\mathbb{R} \times \mathbb{R} \times[0,+\infty) \times\{1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0},\\ C_{-1}&=\mathbb{R} \times \mathbb{R} \times(-\infty, 0] \times\{-1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0},\\ \end{align*} \]

with the jump dynamics

\[ \begin{equation} x^+ =\left\{ \begin{array}{ll} g_{0\pm 1}(x), & \text {if } x \in D_{0\pm 1} \\ g_{\pm 10}(x), & \text {if } x \in D_{\pm 10} \\ g_{1-1}(x), & \text {if } x \in D_{1-1} \\ g_{-11}(x), & \text {if } x \in D_{-11}, \end{array}\right. \end{equation} \]

(378.b)

where

\[ \begin{align*} g_{0\pm 1}(x) & = (\Delta \theta,\omega_1,\omega_2, \rm{sign}(\Delta \theta),\eta,F_s,F_d),\\ g_{\pm 10}(x) & = (\Delta \theta,\omega_1,0,0,\eta,F_s,F_d),\\ g_{1-1}(x) & = (\Delta \theta,\omega_1,\omega_2, -1,\eta,F_s,F_d),\\ g_{-11}(x) & = (\Delta \theta,\omega_1,\omega_2, 1,\eta,F_s,F_d), \end{align*} \]

with the map \( \rm{sign}\) defined as

\[ \rm{sign}(x) = \left\{ \begin{array}{ll} -1, & \text{if } x < 0\\ 0, & \text{if } x = 0\\ 1, & \text{if } x > 0 \end{array}\right. \]

with the jump sets

\[ \begin{align*} D_{0\pm 1}&=\left(\mathbb{R} \setminus\left(-\frac{F_s}{\gamma k}, \frac{F_s}{\gamma k}\right)\right) \times \mathbb{R} \times\{0\} \times\{0\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0},\\ D_{ \pm 10}&=\left(\left[-\frac{F_s}{\gamma k}, \frac{F_d}{\gamma k}\right] \times \mathbb{R} \times\{0\} \times\{1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0}\right)\\ &~~ \cup \left(\left[-\frac{F_d}{\gamma k}, \frac{F_s}{\gamma k}\right] \times \mathbb{R} \times\{0\} \times\{-1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0}\right), \\ D_{1-1}&=\left(-\infty,-\frac{F_s}{\gamma k}\right] \times \mathbb{R} \times\{0\} \times\{1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0},\\ D_{-11}&=\left[\frac{F_s}{\gamma k},+\infty\right) \times \mathbb{R} \times\{0\} \times\{-1\} \times \mathbb{R}\times \mathbb{R}_{\geq 0}\times \mathbb{R}_{\geq 0}, \end{align*} \]

and the measured output

\[ \begin{equation} y=h(x)=\omega_1. \end{equation} \]

(378.c)

Note that for this complicated \( 7\) -dimensional hybrid system, finding gluing functions based on [191, 32] is very difficult. In the next section, we analyze the observability of system (378) and design a gluing KKL observer as in Section 4.2 for this system exploiting NNs.

4.3.3 Gluing KKL Observer Design for System (378)

4.3.3.1 Observability Analysis for System (378)

First, let us analyze the observability of system (378). Notice that \( \eta\) is known from controller design and thus can be seen as an extra measurement. By taking the successive derivatives of \( y = \omega_1\) and knowing \( \eta\) , we deduce that:

  • \( (\Delta \theta,\omega_2)\) are instantaneously observable during flows, in both the stick and slip modes;
  • \( F_d\) is instantaneously observable during flows but only in the slip mode;
  • \( F_s\) is only visible when the solution switches from stick to slip. Notably, since this switching is based on the controller storing enough force to overcome \( F_s\) , the known \( \eta\) actually contains more information about \( F_s\) than \( y=\omega_1\) does.

With this analysis, we infer that system (378) is backward distinguishable along solutions exhibiting stick-slip, with the information of \( F_s\) “hidden” but present in the full output trajectory, so Assumption 32 is satisfied for this system.

We observe that maximal solutions are \( t\) -forward complete and the output is continuous at jumps, so that Item (XXXII) of Assumption 30 holds. However, solutions are non-unique, neither in forward nor in backward time, so Item (XXXI) of Assumption 30 does not hold. Keeping in mind that these are sufficient conditions, we still proceed to use the KKL-based gluing approach presented in Section 4.2 to estimate the full state, including the forces \( (F_s,F_d)\) .

4.3.3.2 Observer Design and Implementation

To implement the observer, we follow the systematic approach proposed in Section 4.2.7, summarized step-by-step as follows.

  1. Selection of observer parameters:
    • Pick \( n_z\) sufficiently large to guarantee the injectivity of \( T\) , but sufficiently small to avoid a sparse dataset. Here, given that we have \( 7\) state components with \( 2\) measurements, there are \( 5\) unknown state components and we pick \( n_z=6\) , so one additional dimension. Note that the \( z\) -dimension in Lemma 19 is a sufficient condition and we typically do not need as many;
    • Pick appropriate eigenvalues in the \( z\) -coordinates: fast eigenvalues filter less and preserve instantaneous information; on the other hand, slow eigenvalues preserve long-term information only. Here, because we have seen from Section 4.3.3.1 that we need both instantaneous and long-term information, we pick \( A=-\rm{diag}(0.05,0.06,1,2,3,4)\) to preserve the information of the stick-slip phenomenon, and \( B=(1,1,…,1)\) , seeing only \( \omega_1\) as output instead of \( (\omega_1,\eta)\) .
  2. Generation and pre-processing of the datasets:
    • Run simulations of the concatenation (378)-(326) during \( 300\) seconds from a large number of initial conditions to compute the \( (x,T(x))\) pairs [103]. The first \( 100\) seconds of each new trajectory should be discarded to get past the transient of \( z\) , ensuring that \( z\) has approximately converged to \( T(x)\) . To efficiently sample the state space, one is presented with a choice: either run a lot of simulations from a lot of initial conditions or store a great number of points in each one of fewer simulations. While the latter method is more economical in terms of resources, it is efficient only if each solution explores a large part of the state space. Here on the contrary, \( (F_s,F_d)\) are constant in each simulation, so a large number of simulations is necessary to diversify their values. For our example, we store \( 100\) data points from each of \( 10000\) simulations with initial conditions for system (378) taken randomly from the set \( [-5,5]\times [0,2] \times [0,2]\times \{0,1\}\times [-2,2]\times [0,10] \times [0,5]\) (with the constraint that \( F_d\leq F_s\) ) and with the \( z\) -dynamics (326) always initialized at zero;
    • Remove the outliers from the dataset to improve the learning. In our case, we discard the solutions not exhibiting stick-slip, believing that this phenomenon should happen most of the time and accepting a wrong estimation otherwise. This is done by discarding solutions that do not jump at least twice during the last \( 200\) seconds of each simulation (i.e., after discarding the first \( 100\) seconds). We also remove all the points for which \( q = -1\) , which occurs much less frequently than \( q = 0\) and \( q = 1\) . Backward rotation of the drill bit does occur, but very rarely, and we accept that the NNs would make mistakes when \( q = -1\) happens, in exchange for more accurate learning;
    • As proposed in Section 4.2.7, split the data into two parts corresponding to the first and second halves of the flow intervals (after and before the jumps), to avoid ineffective learning due to non-injectivity of \( T\) in \( D\) ;
    • Split each dataset into training and testing sets, and normalize both sets component-wise (here we use the \( z\) -score method) but with normalization parameters taken from the training set only.
  3. Designing and training the NNs:
    • Note that while the observer uses the measurement \( y=\omega_1\) in the dynamics \( \dot{\hat{z}} = A\hat{z} + By\) , we also integrate the knowledge of \( (\omega_1,\eta)\) into the observer by feeding their instantaneous values in the inverse maps (the NNs) as inputs. All in all, our NNs recover five unknown components of \( x\) , namely \( (\Delta\theta,\omega_2,q,F_s,F_d)\) , from \( (\hat{z},\omega_1,\eta)\) ;
    • Training separate neural networks for specific state components may improve performance. In particular, it allows us to not use the data splitting for the continuous state components \( (\Delta \theta, \omega_2, F_s, F_d)\) , for which a single NN could be trained on the full dataset without injectivity issues. The data splitting is essentially meaningful here for the discontinuous \( q\) . In our case, for the classifier (deciding from \( \hat{z}(t)\) if we are in the first or second half of the flow interval and thus which regressors to use next), we use a \( 4\) -layer Multi-Layer Perceptron[MLP]. For each dataset, we use a \( 3\) -layer MLP to compute \( (\Delta \theta,\omega_2)\) , another \( 3\) -layer MLP for \( q\) , which is in fact a classifier since we know that \( q \in \{-1,0,1\}\) , and two \( 3\) -layer MLPs for each of \( F_s\) and \( F_d\) . Each layer contains from \( 32\) to \( 64\) neurons.
  4. Implementation and post-processing:
    • Un-normalize the output of the MLPs. For constant state components such as \( (F_s,F_d)\) , we can apply a moving average to filter numerical errors.

The results of a simulation of the observer (356) (with \( T^*\) provided by the obtained MLPs) are given in Figure 55, where \( (\hat{F}_s,\hat{F}_q)\) are filtered to take into account their constant nature. Given the complexity of this problem, these results are considered satisfactory.

&lt;span data-controller=&quot;mathjax&quot;&gt;Estimation results for the stick-slip system.&lt;/span&gt;
Figure 55. Estimation results for the stick-slip system.

In this system, non-stick-slip solutions are indistinguishable, because many different large values of \( F_s\) result in the same “sticking” trajectory. Therefore, we have chosen to consider only solutions exhibiting stick-slip by removing from the datasets those that almost do not jump. Shortly, we intend to build a classifier to tell from \( \hat{z}\) , \( \omega_1\) , and especially \( \eta\) if stick-slip happens. Another future development is to estimate \( F_s\) from a filter of \( \eta\) with slow complex eigenvalues since as analyzed in Section 4.3.3.1 it could be better visible from the amplitude and frequency of the oscillations of the integrator \( \eta\) , which can be more effectively extracted using this filter.

4.3.4 Conclusion

This chapter presents the application of the gluing KKL observer developed in Section 4.2 to the highly challenging problem of stick-slip. The stiction phenomenon encountered in oil drilling is modeled as a \( 7\) -dimensional hybrid system with unknown jump times, with complicated observability conditions of the state components. We detail the methods used to generate datasets and design NNs to learn the KKL inverse map from these data, leading to satisfactory estimation results. Future work will focus on dealing with non-stick-slip solutions and enhancing the estimation of \( F_s\) by leveraging \( \eta\) and strategically adapting the eigenvalues of \( A\) .

Acknowledgments. We thank Zikang Zhu, student at Mines Paris - PSL, for his helpful improvement of the machine learning part in this chapter.

5 Conclusion and Perspectives

Conclusion and Near-Future Developments of this Dissertation

This dissertation proposes novel observer designs for estimating the state of general hybrid systems, with applications in mechanical systems with impacts (such as an Inertial Measurement Unit[IMU] and a walking robot) and the stick-slip phenomenon encountered in oil drilling.

In Section 3, the state estimation problem has been answered for (most) hybrid systems with known jump times, of the form (31). Indeed, we address different structures of this form, from the simpler case of linear maps in Section 3.2 to nonlinear ones in Section 3.3, and propose the corresponding synchronized observers (those whose jumps are triggered at the same time as those of the real solution). These designs range from a systematic Kalman-like observer that gathers observability from the flow-jump combination to more complex ones built by combining a high-gain observer with a jump-based one, leveraging a fictitious measurement derived from the flow-jump interaction. With a broad range of observability conditions and map structures, these designs are believed to encompass most hybrid systems with known jump times. Note that when the jump detection (and hence observer triggering) is delayed to a tolerable extent, we could still achieve practical convergence of the estimation error as pointed out in [22, Section 6]. In Section 3.4, applications to mechanical systems with impacts, including IMUs and walking robots, where many constant parameters such as biases and restitution coefficients, which are normally not measurable using sensors, are estimated thanks to fictitious outputs arising from their interaction with the other states, illustrate our approaches. Therefore, Section 3 of this dissertation constitutes a fairly complete picture of observer design for hybrid systems with known jump times, both theory-wise and application-wise. To improve the application side of our results, we would like to explore more robust controllers for walking robots that reduce sensitivity to uncertainties (such as unknown road slopes and restitution coefficients at impacts). One potential solution is designing a trajectory for an ideal walking gait, unaffected by uncertainties, and then creating a controller that tracks this reference trajectory. This added robustness would allow us to consider and estimate larger uncertainties in the system.

Many of the observers proposed in Section 3 draw on robust discrete-time designs, effectively building on our results developed in Section 2, including the new high-gain observer and KKL observer (in discrete time) presented in Section 2.2 and Section 2.3, respectively. In Section 2.4, these designs are applied to the Permanent Magnet Synchronous Motor[PMSM] discretized in a way that strategically leverages the system’s rotating dynamics. Therefore, Section 2 contains notable novel contributions to discrete-time observer design, both theory- and application-wise.

On the other hand, in Section 4, we have solved the observer design problem for autonomous hybrid systems with fully nonlinear maps and unknown jump times—those of the form (45)—whose outputs are continuous at jumps. We propose in Section 4.2 a systematic observer design by gluing the jump times following the Kravaris-Kazantzis$\slash$Luenberger[KKL] paradigm, leveraging a weak backward distinguishability condition. Application to friction force estimation in the stick-slip phenomenon in oil drilling, with the help of a Neural Network[NN] to approximate the observer using data obtained from offline simulations, is then discussed in Section 4.3, where constructive implementation methods are also proposed. Therefore, Section 4 of this dissertation develops systematic and implementable observers for a fairly large class of hybrid systems with unknown jump times. The obstruction when generalizing our results to non-autonomous hybrid systems lies in the unavailability of KKL results; indeed, backward distinguishability becomes insufficient in this setting when a stronger injectivity property of the gluing transformation is needed. Another obstruction is implementation-wise, we do not have efficient numerical methods to do KKL with time dependence that is not known in advance unless we model this dependence as a “class” of possible inputs (with constant derivatives, sinusoidal form, etc.) and do a functional design. As an immediate continuation of Section 4.3, we would like to improve the performance of our NNs by possibly integrating the system’s physics into the learning.

Open Questions and Perspectives

We outline four key long-term questions that emerge from the problems addressed in this dissertation. These will be explored in our future research beyond the scope of this Ph.D. work.

Understanding Better the High-Gain Observer in Section 2.2

In Section 2.2, we propose a high-gain observer in discrete time that retains the two essential properties of its well-known continuous-time counterpart outlined in Section 1.5.2.2, namely an arbitrarily fast convergence rate and robustness against uncertainties. Our observer structure (76) mirrors that of its continuous-time equivalent (27), with the correction term’s gain \( \gamma\) —of increasing order—playing a similar role to \( \ell\) in (27) that is to speed up the observer and allowing for dominating the nonlinearities; additionally, parameters \( k_i\) , \( i \in \{1,2,…,m\}\) (where \( m\) is the \( z\) -coordinates’ dimension obtained using constructibility), must be chosen. However, while in continuous time, we know that \( K\) in observer (27) must be chosen such that \( A - KC\) is Hurwitz, this clarity of the choice of \( K = (k_1,k_2,…,k_m)\) is lacking in Section 2.2 since the \( A\) matrix here is vacuously Schur. It appears that any choice of \( K\) works, and it only impacts the threshold \( \gamma^\star\) below which \( \gamma\) must be chosen to ensure exponential stability in the \( z\) -coordinates. Unfortunately, our PMSM application in Section 2.4 does not shed light on the role of \( K\) , as it requires pushing \( \gamma\) to a very small value (of order \( 10^{-5}\) ), rendering the effect of \( K\) negligible. Open Question \( \mathbf{1}\) : Can the choice of \( K = (k_1,k_2,…,k_m)\) in observer (27) be optimized for better estimation performance?

Continuous Differentiability of the Gluing Transformation \( T\) in Section 4.2

In Section 4.2, a critical step in proving the injectivity of the gluing transformation \( T\) , which allows it to be left-invertible and makes the observer functional, hinges on the \( C^1\) of \( T\) . This regularity is required by the generalized Coron’s lemma [112, Lemma B.3], the key tool for establishing injectivity under weak backward distinguishability. While we show in this dissertation that the \( C^1\) of \( T\) can be achieved under certain conditions regarding that of the hybrid system’s maps and sets, an open problem here is finding an alternative version of Coron’s lemma that does not require \( C^1\) , forgoing using this property to prove the injectivity of \( T\) .

It is important to note that Coron’s lemma provides sufficient conditions for the injectivity of \( T\) under backward distinguishability, but these conditions are not necessary. They are primarily used for theoretical convergence proofs. Consequently, our observer (356) can still be implemented using NNs to learn the left inverse of \( T\) from data, without checking if \( T\) is \( C^1\) .

Open Question \( \mathbf{2}\) : Can we relax the generalized Coron’s lemma [112, Lemma B.3] into a similar condition that does not require continuous differentiability?

Simultaneous State and Jump Time Estimation for Hybrid Systems

While there are several observer design methods for hybrid systems with unknown jump times as reviewed in Section 4.1, in Section 4.2 of this dissertation, the way we treat a fairly large class of these systems is by gluing the jumps. Inspired by the idea of combing observers in Section 3 and of the recent robotics-focused work [4], we could imagine another novel design paradigm consisting in strategically combining:

  • An observer estimating the system’s state assuming the jump times are known as in Section 3 while being robust to inexact jump triggering, similarly to [22, Section 6];
  • An observer estimating the jump times from the state. For many hybrid systems, using \( D\) —a form of the jump condition, we can express the time until the next jump as a function of the state, often called the time-to-impact function in the literature [201, 202]. E.g., in the bouncing ball in Example 1 with \( c_f = 0\) , from initial state \( x = (x_1,x_2)\) , this time is

    \[ t_{\rm impact}(x) = \frac{1}{a_g} \left(x_2 \pm \sqrt{x_2^2+2a_g x_1}\right) > 0. \]

    In this case, we can use these functions to estimate the jump times from the state. Note that probability-based impact detectors are available in robots [203, 204].

The key idea is to feed the estimates from each observer to the other one to refine both. The state observer assumes accurate jump times, and the jump time estimator assumes accurate state knowledge. The two observers can be designed simultaneously under Lyapunov conditions as done in Section 3, creating a form of separation principle between state and jump time estimation.

Open Question \( \mathbf{3}\) : Can we leverage Lyapunov theory to design simultaneously a state observer and a jump time estimator for hybrid systems with unknown jump times?

Besides, the idea of discretizing hybrid systems for observer design without jump detection, inspired by [193] which restricts to linear impulsive systems, is very promising. This allows considering systems with inputs and an output \( y\) that jumps, for which gluing is infeasible.

Output-Feedback Control of Hybrid Systems

During my Ph.D., I have had the opportunity to visit Prof. Ricardo G. Sanfelice at the University of California Santa Cruz (UCSC), USA in April and May \( 2023\) , to explore output-feedback control of hybrid systems. As discussed in Section 1.4, one of the purposes of observer design is to use the estimate provided by the observer to control the system by feeding it to a state-feedback controller, and here we would like to do the same for hybrid systems as a follow-up of this Ph.D. work. However, the well-known separation principle, one that allows for the independent design of a controller and an observer and then combining them, is typically restricted to linear systems, and hybrid systems are inherently nonlinear due to the flow-jump combination (even if they have linear maps). The work [177] attempts to establish a form of the separation principle for nonlinear continuous-time systems, relying on Lyapunov theory to combine:

  • An arbitrarily fast high-gain observer that allows us to reconstruct the state at any precision and in any short amount of time, possibly with significant peaking during the transient;
  • A state-feedback controller that is fed with the estimate by the high-gain observer and is appropriately saturated to prevent the controlled system from exploding during the transient.

As a major step towards hybrid output-feedback control of hybrid systems, we first generalize these existing results for hybrid systems. However, several challenges arise:

  • Completeness of solutions: In continuous- and discrete-time systems, solutions that are bounded are complete, and this property plays a crucial role in [177] when showing asymptotic convergence of both the estimation error and the tracking error. Lyapunov functions can be exploited to show the boundedness of solutions, leading to their completeness. However, hybrid systems can have bounded but incomplete solutions, such as when they leave the set \( C \cup D\) . A way to address this is to define pre-boundedness, where solutions of interest do not have to be complete but are bounded if they are. This is analogous to pre-asymptotic detectability as discussed in Definition 14, and we can study conditions for this new property;
  • Time domain of solutions: As seen throughout this dissertation, unlike solutions to continuous- and discrete-time systems, each hybrid solution can have its own time domain. This presents difficulties when comparing an ideal solution—generated by a state-feedback controller assuming full-state measurement—with one produced by the output-feedback controller. Such comparisons are essential in [177] but are more complex for hybrid systems;
  • Unavailability of arbitrarily fast discrete-time observers: As analyzed in Section 2.1, instantaneous observability is unique for continuous-time systems and does not exist in discrete time, and thus discrete-time observers can only converge arbitrarily fast after a certain time. In [177], the ability of the observer to make the estimation error converge right from the initialization is very important. In hybrid systems, if the solutions start by jumping, for some first jumps we can never reduce the estimation error because of the lack of observability, which obstructs the convergence of the controlled trajectory.

During my visit to UCSC, I worked on some of these problems and made preliminary progress. However, these results are not included in this dissertation as they remain part of a long-term research effort on hybrid output-feedback control, which will be pursued in future studies.

Open Question \( \mathbf{4}\) : Can we establish a version of the separation principle for hybrid systems?

6 Appendix

6.1 Proofs of\specialChar{160}\Cref{part_discrete}

 

Cette annexe contient les preuves de la Partie 2.

6.1.1 Proofs of Section 2.2

6.1.1.1 Proof of Theorem 1

Along the solutions \( k \mapsto z_k\) to system (67) initialized in \( \mathcal{Z}_0\) with \( y_k \in \mathcal{Y}\) for all \( k \in \mathbb{N}\) and the solutions \( k \mapsto \hat{z}_k\) to observer (76) initialized in \( \mathbb{R}^{n_z}\) and fed with \( y_k\) in (67.b), the estimation error \( \tilde{z}_k := z_k - \hat{z}_k\) verifies

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \tilde{z}_{1,k+1} & = & {}- \gamma^m k_1(\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \tilde{z}_{2,k+1} & = & \Delta \varphi_{2,k}(z_{1,k},\hat{z}_{1,k},y_k)- \gamma^{m-1}k_2 (\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \ldots&&\\ \tilde{z}_{i,k+1} & = & \Delta \varphi_{i,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{i-1,k},\hat{z}_{i-1,k},y_k)- \gamma^{m-i+1} k_i (\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \ldots &&\\ \tilde{z}_{m,k+1} & = & \Delta \varphi_{m,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{m-1,k},\hat{z}_{m-1,k},y_k)- \gamma k_m (\theta_k(z_k) - \bar\theta_k(\hat{z}_k)), \end{array}\right. \end{equation} \]

(379.a)

where for each \( i \in \{2,…,m\}\) ,

\[ \begin{equation} \Delta \varphi_{i,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{i-1,k},\hat{z}_{i-1,k},y_k) = \varphi_{i,k}(z_{1,k},\ldots,z_{i-1,k},y_k)-\bar\varphi_{i,k}(\hat{z}_{1,k},\ldots,\hat{z}_{i-1,k},y_k). \end{equation} \]

(379.b)

Define the re-scaled estimation error \( \varepsilon_k\) where \( \varepsilon_{1,k} = \tilde{z}_{1,k}, …, \varepsilon_{i,k} = \gamma^{i-1}\tilde{z}_{i,k},…, \varepsilon_{m,k} = \gamma^{m-1}\tilde{z}_{m,k}\) . Since \( 0 < \gamma < 1\) , we get \( |\varepsilon_k| \leq |\tilde{z}_k|\) and \( |\tilde{z}_k| \leq \frac{1}{\gamma^{m-1}} |\varepsilon_k|\) for all \( k \in \mathbb{N}\) . Along the solutions to system (67) and observer (76), \( \varepsilon_k\) verifies

\[ \begin{equation} \left\{\begin{array}{@{}r@{\;}c@{\;}l@{}} \varepsilon_{1,k+1} & = &{}- \gamma^m k_1(\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \varepsilon_{2,k+1} & = & \gamma \Delta \varphi_{2,k}(z_{1,k},\hat{z}_{1,k},y_k)- \gamma^m k_2 (\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \ldots&&\\ \varepsilon_{i,k+1} & = & \gamma^{i-1} \Delta \varphi_{i,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{i-1,k},\hat{z}_{i-1,k},y_k)- \gamma^m k_i (\theta_k(z_k) - \bar\theta_k(\hat{z}_k))\\ \ldots &&\\ \varepsilon_{m,k+1} & = & \gamma^{m-1} \Delta \varphi_{m,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{m-1,k},\hat{z}_{m-1,k},y_k)- \gamma^m k_m(\theta_k(z_k) - \bar\theta_k(\hat{z}_k)). \end{array}\right. \end{equation} \]

(380)

Thanks to Item (IV) of Assumption 3, there exists \( c_N > 0\) such that for each \( i \in \{2,…,m\}\) , for all \( (z_k,\hat{z}_k,k,y_k) \in \mathcal{Z} \times \mathbb{R}^{n_z} \times \mathbb{N} \times \mathcal{Y}\) ,

\[ \begin{align*} \gamma^{i - 1}|\Delta \varphi_{i,k}(z_{1,k},\hat{z}_{1,k},\ldots,z_{i-1,k},\hat{z}_{i-1,k},y_k)| &\leq \gamma^{i-1} L_{z,i} \sum_{j=1}^{i-1}|z_{j,k} - \hat{z}_{j,k}|\\ &\leq \gamma^{i-1}L_{z,i}\frac{1}{\gamma^{i-2}}\sum_{j=1}^{i-1}|\varepsilon_{j,k}|\\ &\leq \gamma^{i-1}L_{z,i}\frac{1}{\gamma^{i-2}}c_N|\varepsilon_k|\\ &\leq \gamma \max_{i \in \{2,\ldots,m\}}L_{z,i}c_N|\varepsilon_k|, \end{align*} \]

and for any \( i \in \{1,2,…,m\}\) and \( (z_k,\hat{z}_k,k) \in \mathcal{Z} \times \mathbb{R}^{n_z} \times \mathbb{N}\) ,

\[ \gamma^m |k_i||\theta_k(z_k) - \bar\theta_k(\hat{z}_k)| \leq \gamma^m |k_i| L_\theta |z_k - \hat{z}_k| \leq \gamma^m \max_{i \in \{1,2,\ldots,m\}} |k_i| L_\theta \frac{1}{\gamma^{m-1}} |\varepsilon_k|\leq \gamma \max_{i \in \{1,2,\ldots,m\}} |k_i| L_\theta |\varepsilon_k|. \]

It follows that there exists \( c > 0\) independent of \( \gamma\) such that

\[ \begin{equation} |\varepsilon_{k+1}| \leq \gamma c |\varepsilon_k|. \end{equation} \]

(381)

So, we have \( |\varepsilon_k| \leq (\gamma c)^k |\varepsilon_0|\) for all \( k \in \mathbb{N}\) . Taking \( \gamma^\star = \frac{1}{c}\) , we have for all \( k \in \mathbb{N}\) , \( |\varepsilon_k| \leq \left(\frac{\gamma}{\gamma^\star}\right)^k |\varepsilon_0|. \) We obtain that \( |\tilde{z}_k| \leq \frac{1}{\gamma^{m-1}}|\varepsilon_k|\leq \frac{1}{\gamma^{m-1}}\left(\frac{\gamma}{\gamma^\star}\right)^k |\varepsilon_0| \leq \frac{1}{\gamma^{m-1}}\left(\frac{\gamma}{\gamma^\star}\right)^k|\tilde{z}_0| \) for all \( k \in \mathbb{N}\) , concluding the proof.

6.1.1.2 Proof of Theorem 2

Along the solutions \( k \mapsto z_k\) to system (80) initialized in \( \mathcal{Z}_0\) with \( y_k \in \mathcal{Y}\) and \( v_k \in \mathcal{V}\) for all \( k \in \mathbb{N}\) and the solutions \( k \mapsto \hat{z}_k\) to observer (76) fed with \( y_k + w_k\) in (80.b) instead of \( y_k\) , the \( i^\rm{ th}\) line of dynamics (380) now becomes

\[ \begin{multline} \varepsilon_{i,k+1} = \gamma^{i-1} (\varphi_{i,k}(z_{1,k},\ldots,z_{i-1,k},y_k) + v_{i,k} -\bar\varphi_{i,k}(\hat{z}_{1,k},\ldots,\hat{z}_{i-1,k},y_k+w_k))\\ {}- \gamma^m k_i (\theta_k(z_k) + w_k - \bar\theta_k(\hat{z}_k)). \end{multline} \]

(382)

Based on the proof of Theorem 1, thanks to Item (IV) of Assumption 3, there exists \( c_N > 0\) such that for each \( i \in \{1,2,…,m\}\) , for all \( (z_k,\hat{z}_k,k,y_k,v_k,w_k) \in \mathcal{Z} \times \mathbb{R}^{n_z} \times \mathbb{N} \times \mathcal{Y} \times \mathcal{V} \times \mathbb{R}^{n_y}\) ,

\[ \begin{multline*} \gamma^{i - 1}|\varphi_{i,k}(z_{1,k},\ldots,z_{i-1,k},y_k) + v_{i,k} -\bar\varphi_{i,k}(\hat{z}_{1,k},\ldots,\hat{z}_{i-1,k},y_k+w_k)| \\ \leq \gamma L_{z,i}c_N|\varepsilon_k| + \gamma^{i-1}|v_{i,k}| + \gamma^{i-1}L_{y,i}|w_k|, \end{multline*} \]

and similarly, for each \( i \in \{1,2,…,m\}\) , for all \( (z_k,\hat{z}_k,k,w_k) \in \mathcal{Z} \times \mathbb{R}^{n_z} \times \mathbb{N} \times \mathbb{R}^{n_y}\) ,

\[ \gamma^m |k_i||\theta_k(z_k) + w_k - \bar\theta_k(\hat{z}_k)| \leq \gamma |k_i| L_\theta |\varepsilon_k| + \gamma^m |k_i| |w_k|. \]

Then, we have

\[ |\varepsilon_{i,k+1}| \leq \gamma (L_{z,i}c_N + |k_i|L_\theta) |\varepsilon_k| + \gamma^{i-1} |v_{i,k}| + (\gamma^{i-1}L_{y,i}+\gamma^m |k_i|)|w_k|, \]

and (381) becomes

\[ |\varepsilon_{k+1}| \leq \gamma \sum_{i = 1}^m(L_{z,i}c_N + |k_i|L_\theta) |\varepsilon_k| + \sum_{i = 1}^m\gamma^{i-1} |v_{i,k}| + \sum_{i = 1}^m(\gamma^{i-1}L_{y,i}+\gamma^m |k_i|)|w_k|. \]

Take \( \gamma^\star = \frac{1}{\sum_{i = 1}^m L_{z,i}c_N + |k_i|L_\theta}\) . It then follows that for all \( k \in \mathbb{N}_{\geq 0}\) ,

\[ |\varepsilon_k| \leq \left(\frac{\gamma}{\gamma^\star}\right)^k |\varepsilon_0| + \sum_{j = 0}^{k-1}\left(\frac{\gamma}{\gamma^\star}\right)^{k-1-j}\sum_{i = 1}^m\gamma^{i-1} |v_{i,j}|+\sum_{j = 0}^{k-1}\left(\frac{\gamma}{\gamma^\star}\right)^{k-1-j}\sum_{i = 1}^m(\gamma^{i-1}L_{y,i}+\gamma^m |k_i|)|w_j|. \]

Note that \( |\varepsilon_{i,k}| \leq |\varepsilon_k|\) for all \( k \in \mathbb{N}\) , so we obtain the same results for \( |\varepsilon_{i,k}|\) . Finally, realizing that \( z_{i,k}-\hat{z}_{i,k} = \frac{1}{\gamma^{i-1}}\varepsilon_{i,k}\) , we get (81).

6.1.2 Proofs of Section 2.3

6.1.2.1 Proof of Theorem 5

First, pick \( \gamma\in (0,1]\) small enough to ensure that \( \gamma\|\tilde{A}\|<1\) . Consider a solution \( (T_k)_{k \in \mathbb{N}}\) to (88) for \( (A,B)\) given in (109). Then,

\[ T_k(x) = \left(T_{1,k}(x), T_{2,k}(x), \ldots, T_{i,k}(x), \ldots, T_{n_y,k}(x)\right), \]

where for each \( i\in \{1,2,…,n_y\}\) , \( (T_{i,k})_{k \in \mathbb{N}}\) is solution to (88) with \( (A,B)\) replaced by \( (\gamma \tilde{A}_i,\tilde{B}_i)\) . Therefore, Theorem 4 applies to each \( (T_{i,k})_{k \in \mathbb{N}}\) . It follows that for each \( i \in \{1,2,…,n_y\}\) , for all \( k \in \mathbb{N}_{\geq m_i}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) , \( T_{i,k}(x_a) - T_{i,k}(x_b)\) can be written as the sum of three parts

\[ T_{i,k}(x_a) - T_{i,k}(x_b) = \left(\mathcal{I}_{i,k}(x_a) - \mathcal{I}_{i,k}(x_b)\right) + \left(\mathcal{T}_{i,k}(x_a) - \mathcal{T}_{i,k}(x_b)\right) + \left(\mathcal{R}_{i,k}(x_a) - \mathcal{R}_{i,k}(x_b)\right), \]

where

\[ \begin{align*} \mathcal{I}_{i,k}(x_a) - \mathcal{I}_{i,k}(x_b)& = (\gamma \tilde{A}_i)^k \left((T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1})(x_a) - (T_0 \circ f^{-1}_0 \circ f^{-1}_1 \circ \ldots \circ f^{-1}_{k-1})(x_b)\right), \\ \mathcal{R}_{i,k}(x_a) - \mathcal{R}_{i,k}(x_b)& = \sum_{j=0}^{k-m_i-1}(\gamma \tilde{A}_i)^{k-j-1}\tilde{B}_i\big((h_{i,j} \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_a) \\ &~~{}- (h_{i,j} \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_b)\big),\\ \mathcal{T}_{i,k}(x_a) - \mathcal{T}_{i,k}(x_b)& = \sum_{j=k-m_i}^{k-1}(\gamma \tilde{A}_i)^{k-j-1}\tilde{B}_i\big((h_{i,j} \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_a) \\ &~~{}- (h_{i,j} \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_b)\big)\\ & = \mathcal{D}_i(\gamma)\mathcal{C}_i (\mathcal{O}^{bw}_{i,k}(x_a) - \mathcal{O}^{bw}_{i,k}(x_b)), \end{align*} \]

where \( \mathcal{D}_i(\gamma) = \rm{diag}(1, \gamma, \gamma^2, …, \gamma^{m_i-1})\) and \( \mathcal{C}_i = \begin{pmatrix} \tilde{B}_i & \tilde{A}_i \tilde{B}_i & \tilde{A}_i^2 \tilde{B}_i & … & \tilde{A}_i^{m_i}\tilde{B}_i \end{pmatrix}\) is the controllability matrix of the pair \( (\tilde{A}_i,\tilde{B}_i)\) . Now, we will establish bounds on each of the three parts. As \( (T_k)_{k \in \mathbb{N}}\) is initialized globally Lipschitz, there exists \( c_T \in \mathbb{R}_{\geq 0}\) such that for all \( (x_a,x_b) \in \mathbb{R}^{n_x} \times \mathbb{R}^{n_x}\) , \( |T_0(x_a) - T_0(x_b)| \leq c_T |x_a - x_b|\) . Exploiting Item (V) of Assumption 7, we thus have for each \( i \in \{1,2,…,n_y\}\) , for all \( k \in \mathbb{N}_{\geq m_i}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ |\mathcal{I}_{i,k}(x_a) - \mathcal{I}_{i,k}(x_b)| \leq c_T (\gamma \|\tilde{A}_i\|c_f)^k |x_a - x_b|. \]

Then, for \( \gamma\) such that \( \gamma\max_{i \in \{1, 2, …, n_y\}}\|\tilde{A}_i\|c_f < 1\) , exploiting Item (V) of Assumption 7, we have for each \( i \in \{1,2,…,n_y\}\) , for all \( k \in \mathbb{N}_{\geq m_i}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{align*} |\mathcal{R}_{i,k}(x_a) - \mathcal{R}_{i,k}(x_b)| &\leq \sum_{j=0}^{k-m_i-1}(\gamma \|\tilde{A}_i\|)^{k-j-1} \|\tilde{B}_i\| c_h c_f^{k-j}|x_a - x_b|\\ & = \|\tilde{B}_i\| c_h c_f \frac{(\gamma\|\tilde{A}_i\|c_f)^{m_i}}{1 - \gamma\|\tilde{A}_i\|c_f}\left(1 - (\gamma\|\tilde{A}_i\|c_f)^{k-m_i-1}\right)|x_a - x_b|\\ & \leq \|\tilde{B}_i\| c_h c_f \frac{(\gamma\|\tilde{A}_i\|c_f)^{m_i}}{1 - \gamma\|\tilde{A}_i\|c_f}|x_a - x_b|. \end{align*} \]

Since the pairs \( (\tilde{A}_i, \tilde{B}_i)\in \mathbb{R}^{m_i\times m_i}\times \mathbb{R}^{m_i}\) are controllable, there exists \( c_c > 0\) such that \( \|\mathcal{C}_i^{-1}\| \leq \frac{1}{c_c}\) for each \( i \in \{1, 2, …, n_y\}\) . Next, we deduce that for each \( i \in \{1,2,…,n_y\}\) , for all \( k \in \mathbb{N}_{\geq m_i}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{split} |\mathcal{T}_{i,k}(x_a) - \mathcal{T}_{i,k}(x_b)| \geq \gamma^{m_i-1} c_c |\mathcal{O}^{bw}_{i,k}(x_a) - \mathcal{O}^{bw}_{i,k}(x_b)|. \end{split} \]

Therefore, for each \( i \in \{1,2,…,n_y\}\) , for all \( k \in \mathbb{N}_{\geq m_i}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{align*} \begin{split} |T_{i,k}(x_a) - T_{i,k}(x_b)| &= |(\mathcal{I}_{i,k}(x_a) - \mathcal{I}_{i,k}(x_b)) + (\mathcal{T}_{i,k}(x_a)- \mathcal{T}_{i,k}(x_b)) + (\mathcal{R}_{i,k}(x_a) - \mathcal{R}_{i,k}(x_b))| \\ &\geq |\mathcal{T}_{i,k}(x_a) - \mathcal{T}_{i,k}(x_b)| - |\mathcal{R}_{i,k}(x_a) - \mathcal{R}_{i,k}(x_b)| - |\mathcal{I}_{i,k}(x_a) - \mathcal{I}_{i,k}(x_b)| \\ & \geq \gamma^{m_i-1} c_c |\mathcal{O}^{bw}_{i,k}(x_a) - \mathcal{O}^{bw}_{i,k}(x_b)| \\ &~~{}- \|\tilde{B}_i\| c_h c_f \frac{(\gamma\|\tilde{A}_i\|c_f)^{m_i}}{1 - \gamma\|\tilde{A}_i\|c_f}|x_a - x_b|- c_T (\gamma \|\tilde{A}_i\|c_f)^k |x_a - x_b|\\ &\geq \gamma^{m_i-1} \bigg(c_c |\mathcal{O}^{bw}_{i,k}(x_a) - \mathcal{O}^{bw}_{i,k}(x_b)|\\ &~~{} - \|\tilde{B}_i\| c_h c_f \frac{\gamma(\|\tilde{A}_i\|c_f)^{m_i}}{1 - \gamma\|\tilde{A}_i\|c_f}|x_a - x_b|- c_T \gamma^{k-m_i+1}(\|\tilde{A}_i\|c_f)^k |x_a - x_b|\bigg). \end{split} \end{align*} \]

Now, since \( \gamma \in (0,1]\) and thanks to Item (VI) of Assumption 7, if we concatenate the outputs, there exists a constant \( c_N > 0\) (depending on the chosen norms only) such that for all \( k \in \mathbb{N}_{\geq \overline{m}}\) and for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) , we have

\[ \begin{multline*} |T_k(x_a) - T_k(x_b)|\geq c_N \gamma^{\overline{m}-1} \bigg(c_c c_o - \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\| c_h c_f \frac{\gamma \max_{i \in \{1, 2, \ldots, n_y\}}((\|\tilde{A}_i\|c_f)^{m_i})}{1 - \gamma\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f} \\ - c_T \gamma^{k-\overline{m}+1}\left(\max_{i \in \{1, 2, \ldots, n_y\}} \|\tilde{A}_i\|c_f\right)^k\bigg) |x_a - x_b|. \end{multline*} \]

If we select \( \gamma \in (0,1]\) such that

\[ \begin{multline*} 0 < \gamma < \gamma^\star = \min\bigg\{\frac{1}{\|\tilde{A}\|},\frac{1}{\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\| c_f}, \\ \frac{c_c c_o}{\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_fc_c c_o + \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\|c_h c_f\max_{i \in \{1, 2, \ldots, n_y\}}((\|\tilde{A}_i\|c_f)^{m_i})}\bigg\}, \end{multline*} \]

then with \( \gamma\) fixed, for all \( k \in \mathbb{N}_{\geq k^\star}\) where \( k^\star = \overline{m}\) if \( c_T = 0\) and

\[ k^\star = \max\left\{\overline{m}, \left\lfloor \frac{(\overline{m} - 1)\ln\gamma + \ln\tilde{c} - \ln c_T}{\ln\gamma + \ln(\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f)} + 1\right\rfloor \right\}, \]

where \( \tilde{c} = c_c c_o - \max_{i \in \{1, 2, …, n_y\}}\|\tilde{B}_i\| c_h c_f \frac{\gamma\max_{i \in \{1, 2, …, n_y\}}((\|\tilde{A}_i\|c_f)^{m_i})}{1 - \gamma\max_{i \in \{1, 2, …, n_y\}}\|\tilde{A}_i\|c_f}\) if \( c_T > 0\) , there exists a constant \( c > 0\) (independent of \( \gamma\) and \( k\) , because it is picked according to \( \gamma^\star\) and \( k^\star\) ) such that for all \( (x_a, x_b) \in \mathcal{X} \times \mathcal{X}\) , we have (110).

6.1.2.2 Proof of Theorem 6

First, we prove that \( (T_k)_{k \in \mathbb{N}}\) provided by Theorem 5 is uniformly Lipschitz. Indeed, from Item (V) of Assumption 7, we have for all \( k\in \mathbb{N}\) and for all \( (x_a, x_b) \in \mathbb{R}^{n_x} \times \mathbb{R}^{n_x}\) ,

\[ \begin{align*} |T_k(x_a) - T_k(x_b)| &\leq c_T (\gamma \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f)^k |x_a - x_b| \\ &~~ {}+ \sum_{j=0}^{k-1}(\gamma \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|)^{k-j-1} \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\| c_h c_f^{k-j}|x_a - x_b|\\ & \leq c_T |x_a - x_b| + \max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\| c_h c_f \frac{1 - (\gamma\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f)^{k-1}}{1 - \gamma\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f}|x_a - x_b|\\ & \leq \left(c_T + \frac{\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{B}_i\| c_h c_f}{1 - \gamma\max_{i \in \{1, 2, \ldots, n_y\}}\|\tilde{A}_i\|c_f}\right)|x_a - x_b| := c_L |x_a - x_b|, \end{align*} \]

since \( \gamma \max_{i \in \{1, 2, …, n_y\}}\|\tilde{A}_i\|c_f<1\) according to the proof of Theorem 5. We now prove the robust stability and ISS properties. Consider a solution to system (115) with \( x_0 \in \mathcal{X}_0\) and a solution to observer (117) with \( z_0\in T_0(\mathcal{X})\) . Denoting \( z_k=T_k(x_k)\) , we write the dynamics in the \( z\) -coordinates as

\[ \begin{align*} \begin{split} z_{k+1} & = T_{k+1}(f_k(x_k) + v_k) \\ & = T_{k+1}(f_k(x_k)) + T_{k+1}(f_k(x_k) + v_k) - T_{k+1}(f_k(x_k)) \\ & = \gamma \tilde{A} T_k(x_k) + Bh_k(x_k) + T_{k+1}(f_k(x_k) + v_k) - T_{k+1}(f_k(x_k)) \\ & = \gamma \tilde{A} T_k(x_k) + B(y_k - w_k) + T_{k+1}(f_k(x_k) + v_k) - T_{k+1}(f_k(x_k)) \\ & = \gamma \tilde{A} z_k + B y_k + T_{k+1}(f_k(x_k) + v_k) - T_{k+1}(f_k(x_k))- Bw_k . \end{split} \end{align*} \]

Because \( (T_k)_{k \in \mathbb{N}}\) is uniformly Lipschitz, for all \( k\in \mathbb{N}\) ,

\[ |T_{k+1}(f_k(x) + v_k) - T_{k+1}(f_k(x))| \leq c_L|v_k|. \]

According to (117), we get for all \( k \in \mathbb{N}_{>0}\) ,

\[ z_k - \hat{z}_k = (\gamma \tilde{A})^k(z_0 - \hat{z}_0) + \sum_{j=0}^{k-1}(\gamma \tilde{A})^{k-j-1}(T_{j+1}(f_j(x_j) + v_j) - T_{j+1}(f_j(x_j))-Bw_j ). \]

Therefore, we have for all \( k \in \mathbb{N}_{\geq k^\star}\) ,

\[ \begin{align*} |x_k - \hat{x}_k|& = |T^*_k (z_k) - \tilde{T}^*_k(\hat{z}_k)|\\ &\leq |T^*_k (z_k) - T^*_k(\hat{z}_k)| + \delta \\ & \leq \frac{c^\prime}{c \gamma^{\overline{m}-1}}|z_k - \hat{z}_k|+ \delta \\ & \leq \frac{c^\prime(\gamma \|\tilde{A}\|)^k}{c \gamma^{\overline{m}-1}}|z_0 - \hat{z}_0| \\ & ~~{} + \frac{c^\prime}{c \gamma^{\overline{m}-1}} \sum_{j=0}^{k-1}(\gamma \|\tilde{A}\|)^{k-j-1}|T_{j+1}(f_j(x_j) + v_j) - T_{j+1}(f_j(x_j)) - Bw_j|+ \delta \\ & \leq \frac{c^\prime(\gamma \|\tilde{A}\|)^k}{c \gamma^{\overline{m}-1}}|T_0(x_0) - T_0(\hat{x}_0)| + \frac{c^\prime}{c \gamma^{\overline{m}-1}}\sum_{j=0}^{k-1}(\gamma\|\tilde{A}\|)^{k-j-1}(c_L |v_j| + \|B\||w_j|)+ \delta. \end{align*} \]

Since \( \gamma \|\tilde{A}\|<1\) (according to the proof of Theorem 5), this concludes the proof.

6.1.2.3 Proof of Theorem 7

Recall that the set \( \mathcal{M}_{ND}\) of pairs of real matrices \( (\tilde{A},\tilde{B})\) where \( \tilde{A}\) is non-diagonalizable in \( \mathbb{C}\) and the set \( \mathcal{M}_{NC}\) of uncontrollable pairs of real matrices \( (\tilde{A},\tilde{B})\) are both of zero measure in \( \mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}\) . Indeed, they are the zero locus of non-identically zero polynomials: the discriminant of the characteristic polynomial for the former and the determinant of the controllability matrix for the latter. Now, consider \( (\tilde{A},\tilde{B})\) in \( \mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}\) controllable with \( \tilde{A}\) Schur and diagonalizable in \( \mathbb{C}\) . Consider the maps \( (T_k)_{k\in \mathbb{N}}\) defined in (99) with \( T_0=0\) , which can be written as

\[ \begin{equation} T_k(x) = \sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x), \end{equation} \]

(383)

with \( (A,B)\) defined in (123). Define

\[ \tilde{A}_{\rm real} = \rm{diag}(\Lambda_1, \Lambda_2, \ldots, \Lambda_l, \lambda_{l+1}, \lambda_{l+2}, \ldots, \lambda_{2n_x-l+1}), ~~ \tilde{B}_{\rm real}= \begin{pmatrix} B_1 \\ B_2 \\ \ldots \\ B_{2n_x-l+1} \end{pmatrix}, \]

where

\[ \Lambda_i = \begin{pmatrix} \Re{\lambda_i} & -\Im(\lambda_i) \\ \Im(\lambda_i) & \Re(\lambda_i) \end{pmatrix}, ~~ B_i = \left\{ \begin{array}{@{}l@{~~}l@{}} \begin{pmatrix} 1 \\ 0 \end{pmatrix}& i \in \{1, 2, \ldots, l\} \\ 1 & i \in \{l+1, l+2, \ldots, 2n_x-l+1\}, \end{array}\right. \]

where \( l \in \{0, 1, …, n_x\}\) is the number of complex non-real eigenvalues of \( \tilde{A}\) (that come in pairs of conjugates since \( \tilde{A}\) is real). As shown in [112, Appendix B.\( 1\) ], there exists an invertible matrix \( \tilde{P}\in\mathbb{R}^{(2n_x+1)\times(2n_x+1)}\) such that

\[ \tilde{A}_{\rm real} = \tilde{P}^{-1}\tilde{A} \tilde{P}, ~~ \tilde{B}_{\rm real} = \tilde{P}^{-1}\tilde{B}. \]

First, since \( \tilde{P}\) is invertible, the injectivity of the maps \( (T_k)_{k \in \mathbb{N}}\) in (383) is implied by the injectivity of the maps \( (T_\rm{{ real},k})_{k \in \mathbb{N}}\) defined as

\[ T_{{\rm real},k}(x) = (\rm{Id}_{n_y} \otimes \tilde{P}^{-1}) T_k(x). \]

We have

\[ \begin{align*} T_{{\rm real},k}(x) &=(\rm{Id}_{n_y} \otimes \tilde{P}^{-1}) T_k(x)\\ &=(\rm{Id}_{n_y} \otimes \tilde{P}^{-1})\sum_{j=0}^{k-1}A^{k-j-1}B(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x) \\ &=(\rm{Id}_{n_y} \otimes \tilde{P}^{-1})\sum_{j=0}^{k-1}(\rm{Id}_{n_y} \otimes \tilde{A})^{k-j-1}(\rm{Id}_{n_y} \otimes \tilde{B})(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x) \\ &= \sum_{j=0}^{k-1}(\rm{Id}_{n_y} \otimes (\tilde{P}^{-1}\tilde{A}\tilde{P}))^{k-j-1}(\rm{Id}_{n_y} \otimes (\tilde{P}^{-1}\tilde{B}))(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x)\\ &= \sum_{j=0}^{k-1}A_{\rm real}^{k-j-1}B_{\rm real}(h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x), \end{align*} \]

with the pair \( (A_\rm{ real},B_\rm{ real})\) defined as

\[ A_{\rm real} = \rm{Id}_{n_y} \otimes \tilde{A}_{\rm real}, ~~ B_{\rm real} = \rm{Id}_{n_y} \otimes \tilde{B}_{\rm real}. \]

Second, we prove the injectivity of the maps \( (T_\rm{{ real},k})_{k \in \mathbb{N}}\) . Define now the open sets \( \Upsilon = \{(x_a, x_b) \in \mathcal{O} \times \mathcal{O}: x_a \neq x_b\}\) and \( \Lambda_l = (B_1(0))^l \times (-1,1)^{2n_x-l+1}\) . For \( \lambda\in \mathbb{C}\) , define the map \( \mathcal{T}_{\lambda,k}\) as

\[ T_{\lambda,k}(x) = \sum_{j=0}^{k-1}\lambda^{k-j-1} (h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x). \]

With the structure of \( \tilde{A}_\rm{ real}\) and \( \tilde{B}_\rm{ real}\) , the functions \( (T_\rm{{ real},k})_{k \in \mathbb{N}}\) can be written up to a permutation as

\[ T_{{\rm real},k}(x) = ( \Re(T_{\lambda_1,k}(x)),\Im(T_{\lambda_1,k}(x)),\ldots, \Re(T_{\lambda_l,k}(x)),\Im(T_{\lambda_l,k}(x)), T_{\lambda_{l+1},k}(x), \ldots, T_{\lambda_{2n_x -l+ 1},k}(x)). \]

It follows that proving the injectivity of \( T_\rm{{ real},k}\) for some \( (\lambda_1, \lambda_2, …, \lambda_{2n_x-l+1}) \in \Lambda_l\) is equivalent to proving the injectivity of

\[ T_{{\rm complex},k}(x) = ( T_{\lambda_1,k}(x), T_{\lambda_2,k}(x),\ldots, T_{\lambda_{2n_x -l+ 1},k}(x)). \]

We now prove that this is guaranteed for all \( k \in \mathbb{N}_{\geq k^\star}\) and for almost any choice of \( (\lambda_1, \lambda_2, …, \lambda_{2n_x-l+1}) \in \Lambda_l\) in the Lebesgue measure sense. For that, we define the sets

\[ \Theta_i = \left\{ \begin{array}{@{}l@{~~}l@{}} B_1(0)& i \in \{1, 2, \ldots, l\} \\ (-1, 1) & i \in \{l+1, l+2, \ldots, 2n_x-l+1\}, \end{array}\right. \]

the counters

\[ p_i = \left\{ \begin{array}{@{}l@{~~}l@{}} 2& i \in \{1, 2, \ldots, l\} \\ 1 & i \in \{l+1, l+2, \ldots, 2n_x-l+1\}, \end{array}\right. \]

and the functions \( g_{i,k}: \Upsilon \times \Theta_i \to \mathbb{R}\) for \( i \in \{1, 2, …, l\}\) and \( g_{i,k}: \Upsilon \times \Theta_i \to \mathbb{C}\) for \( i \in \{l+1, l + 2, …, 2n_x-l+1\}\) , by

\[ \begin{align*} g_{i,k}((x_a, x_b), \lambda) &= \mathcal{T}_{\lambda,k}(x_a) - \mathcal{T}_{\lambda,k}(x_b)\\ &= \sum_{j=0}^{k-1}\lambda^{k-j-1} \left((h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_a) - (h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ \ldots \circ f^{-1}_{k-1})(x_b)\right). \end{align*} \]

Now, we check the conditions for the generalized Coron’s Lemma in [112, Lemma B.\( 3\) ]. For any \( l \in \{0, 1, …, n_x\}\) ,

  • For \( i \in \{1, 2, …, l\}\) , we see that \( g_{i,k}((x_a, x_b), \cdot)\) is holomorphic on \( B_1(0)\) for all \( (x_a, x_b) \in \Upsilon\) . From the chain rule, under Item (VII) of Assumption 8, for all \( \lambda \in B_1(0)\) , \( g_{i,k}(\cdot, \lambda)\) is \( C^1\) on \( \Upsilon\) for each \( \lambda \in B_1(0)\) because it is a finite composition of \( C^1\) functions;
  • For \( i \in \{l+1, l + 2,…, 2n_x-l+1\}\) , as \( g_{i,k}((x_a, x_b), \cdot)\) is a polynomial, it is \( C^\infty\) on \( (-1, 1)\) for all \( (x_a, x_b) \in \Upsilon\) . From the chain rule, under Item (VII) of Assumption 8, for all \( \lambda \in (-1, 1)\) and for all \( j \in \mathbb{N}\) , the maps \( \frac{\partial^j g_{i,k}}{\partial \lambda^j}(\cdot, \lambda)\) are \( C^1\) on \( \Upsilon\) because they are finite compositions of \( C^1\) functions.

We then show that under Item (VIII) of Assumption 8, for all \( i\in \{1, 2, …,2n_x-l+1\}\) , \( g_{i,k}((x_a, x_b), \cdot)\) cannot be identically zero on \( \Theta_i\) . Take \( (x_a, x_b) \in \Upsilon\) and take \( k \in \mathbb{N}_{\geq k^\star}\) and assume \( g_{i,k}((x_a, x_b), \lambda) = 0\) for all \( \lambda \in \Theta_i\) . By uniqueness of polynomials, for all \( j \in \{0, 1, …, k-1\}\) , we have \( (h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ … \circ f^{-1}_{k-1})(x_a) = (h_j \circ f^{-1}_j \circ f^{-1}_{j+1} \circ … \circ f^{-1}_{k-1})(x_b)\) , which contradicts Item (VIII) of Assumption 8. From the generalized Coron’s Lemma [112] applied at each \( l\in \{0,1,…,n_x\}\) and each \( k \in \mathbb{N}_{\geq k^\star}\) , since \( \sum_{i = 1}^{2n_x - l + 1} p_i = 2n_x + 1\) , the set

\[ \mathcal{E}_{l,k} = \bigcup\limits_{(x_a,x_b) \in \Upsilon}\left\{(\lambda_1, \lambda_2, \ldots, \lambda_{2n_x-l+1}) \in \Lambda_l \, | \, \forall i \in \{1, 2, \ldots, 2n_x-l+1\}, g_{i,k}((x_a, x_b), \lambda_i) = 0\right\}, \]

which is literally the set of eigenvalues in \( \Lambda_l\) making \( T_\rm{{ complex},k}\) (at each \( k\) ) non-injective, has zero Lebesgue measure. Then, from [112, Lemma B.\( 2\) ], the set

\[ \mathcal{M}_{l,k} = \{(\tilde{A}, \tilde{B}) \in \mathbb{R}^{(2n_x+1)\times(2n_x+1)} \times \mathbb{R}^{2n_x+1}| \tilde{A} \text{ has the eigenvalues in } \mathcal{E}_{l,k}\} \]

also has zero measure. Now, recall that the countable union of infinitely many zero Lebesgue measure sets also has zero Lebesgue measure [205]. Therefore, the set

\[ \mathcal{M} = \mathcal{M}_{ND} \cup \mathcal{M}_{NC} \cup \bigcup\limits_{\substack{k \in \mathbb{N} \\ l \in \{0, 1, \ldots, n_x\}}}\mathcal{M}_{l,k} \]

also has zero Lebesgue measure.

6.2 Technical Lemmas and Proofs of\specialChar{160}\Cref{part_known}

 

Cette annexe contient les lemmes techniques et les preuves de la Partie 3.

6.2.1 Technical Lemmas and Proofs of Section 3.2

6.2.1.1 Technical Lemmas and Proofs of Section 3.2.2

Lemma 22 (Uniform lower-boundedness of the observability Gramian)

Consider a quadruple \( (F,J,H_c,H_d)\) defined on a hybrid time domain \( \mathcal{D}\) and verifying the boundedness condition of Assumption 11. Pick \( (K_c,K_d)\) defined and uniformly upper-bounded on \( \mathcal{D}\) . Then, for any \( \Delta_m>1\) , there exists \( c_\mathcal{G}>0\) such that any hybrid arc \( \tilde{x}\) defined on \( \mathcal{D}\) and verifying the linear dynamics \( \dot{\tilde{x}} = (F-K_cH_c) \tilde{x}\) during flows and \( \tilde{x}^+ = (J-K_dH_d)\tilde{x}\) at jumps, verifies for all \( ((t^\prime,j^\prime),(t,j))\in \mathcal{D}\times \mathcal{D}\) with \( (t-t^\prime)+(j-j^\prime)\leq \Delta_m\) ,

\[ \begin{equation} \tilde{\mathcal{G}}((t^\prime,j^\prime),(t,j))\geq c_\mathcal{G} (\tilde{x}(t,j))^\top\mathcal{G}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j))\tilde{x}(t,j), \end{equation} \]

(384.a)

where

\[ \begin{multline} \tilde{\mathcal{G}}((t^\prime,j^\prime),(t,j)) = \int_{t^\prime}^{t_{j^\prime+1}}\star^\top H_c(s,j^\prime) \tilde{x}(s,j^\prime)ds \\ + \sum_{k=j^\prime+1}^{j-1} \tilde{\mathcal{G}}_F(k) +\sum_{k=j^\prime}^{j-1}\tilde{\mathcal{G}}_J(k) + \int_{t_{j}}^{t}\star^\top H_c(s,j) \tilde{x}(s,j)ds, \end{multline} \]

(384.b)

with \( \tilde{\mathcal{G}}_F\) and \( \tilde{\mathcal{G}}_J\) defined in (170.c) and (170.d).

Remark 57

Define \( \tilde{F} := F-K_cH_c\) and \( \tilde{J}:= J-K_dH_d\) . If \( \tilde{J}\) is invertible at all times, then Lemma 22 is equivalent to the fact that for all \( ((t^\prime,j^\prime),(t,j))\in \mathcal{D}\times \mathcal{D}\) with \( (t-t^\prime)+(j-j^\prime)\leq \Delta_m\) , we have

\[ \mathcal{G}_{(\tilde{F},\tilde{J},H_c,H_d)}((t^\prime,j^\prime),(t,j))\geq c_\mathcal{G} \mathcal{G}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)). \]

Proof. This proof resembles that of [206, Theorem 3] but in the case where \( F\neq 0\) and \( J\neq 0\) , and that of [81] extended to the hybrid case. The key idea is to consider the terms \( -K_cH_c\tilde{x}\) and \( -K_dH_d\tilde{x}\) in the dynamics of \( \tilde{x}\) as flow\( /\) jump inputs respectively, so that

\[ \begin{multline} \tilde{x}(t,j) = \Phi_{F,J}((t,j),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime) -\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((t,j),(s,j^\prime))ds - \sum_{k = j^\prime+1}^{j-1}\int_{t_k}^{t_{k+1}} \Psi_A((t,j),(s,k))ds\\ -\sum_{k = j^\prime}^{j-1}\Psi_B((t,j),(t_{k+1},k))-\int_{t_j}^t \Psi_A((t,j),(s,j))ds, \end{multline} \]

(385.a)

where \( A:=K_cH_c\) , \( B:= K_dH_d\) , and

\[ \begin{align} \Psi_A((t,j),(t^\prime,j^\prime)) &= \Phi_{F,J}((t,j),(t^\prime,j^\prime))A(t^\prime,j^\prime)\tilde{x}(t^\prime,j^\prime), \\ \Psi_B((t,j),(t^\prime,j^\prime)) &= \Phi_{F,J}((t,j),(t^\prime,j^\prime))B(t^\prime,j^\prime)\tilde{x}(t^\prime,j^\prime). \\\end{align} \]

(385.b)

Since \( (t-t^\prime) + (j-j^\prime) \leq \Delta_m\) , we deduce the upper bounds on \( \Psi_A\) and \( \Psi_B\) as

\[ \begin{align*} |\Psi_A((t,j),(t^\prime,j^\prime))| &\leq c_A|H_c(t^\prime,j^\prime) \tilde{x}(t^\prime,j^\prime)|,\\ |\Psi_B((t,j),(t^\prime,j^\prime))| &\leq c_B|H_d(t^\prime,j^\prime) \tilde{x}(t^\prime,j^\prime)|, \end{align*} \]

with scalars \( c_A\) and \( c_B\) independent of \( (t,j,t^\prime,j^\prime)\) given by

\[ \begin{align*} c_A &= c_{K_c}(e^{c_F}\max\{c_J,c_{J^{-1}}\})^{\Delta_m}, \\ c_B &= c_{K_d}(e^{c_F}\max\{c_J,c_{J^{-1}}\})^{\Delta_m}. \end{align*} \]

Let us now lower-bound each term in \( \tilde{\mathcal{G}}\) . First, using (385.a) with \( (s,k)\) replacing \( (t,j)\) , we get

\[ \begin{multline*}\tilde{\mathcal{G}}_F(k) = \int_{t_k}^{t_{k+1}} \bigg|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\\ - H_c(s,k)\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}} \Psi_A((s,k),(u,q))du- H_c(s,k)\sum_{q = j^\prime}^{k-1}\Psi_B((s,k),(t_{q+1},q))\\ -H_c(s,k)\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((s,k),(u,j^\prime))du-H_c(s,k)\int_{t_k}^s \Psi_A((s,k),(u,k))du\bigg|^2ds. \end{multline*} \]

Using \( |a-b|^2 \geq \frac{\rho}{1+\rho}|a|^2-\rho|b|^2\) for some \( \rho > 0\) , we get

\[ \begin{multline*} \tilde{\mathcal{G}}_F(k) \geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \left|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\right|^2ds \\ - \rho c_{H_c}^2 \int_{t_k}^{t_{k+1}}\bigg|\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}} \Psi_A((s,k),(u,q))du + \sum_{q = j^\prime}^{k-1}\Psi_B((s,k),(t_{q+1},q)) \\ + \int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((s,k),(u,j^\prime))du+ \int_{t_k}^s \Psi_A((s,k),(u,k))du\bigg|^2ds. \end{multline*} \]

Applying \( \left|\sum_{i=1}^Na_i\right|^2 \leq N\sum_{i=1}^N|a_i|^2\) (obtained from Cauchy-Schwartz inequality), we get

\[ \begin{multline*} \tilde{\mathcal{G}}_F(k) \geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2ds- \rho c_{H_c}^2(2(k-j^\prime)+1)\times \\ \times\int_{t_k}^{t_{k+1}}\bigg[\sum_{q = j^\prime+1}^{k-1}\bigg|\int_{t_q}^{t_{q+1}} \Psi_A((s,k),(u,q))du\bigg|^2 + \sum_{q = j^\prime}^{k-1}\big|\Psi_B((s,k),(t_{q+1},q))\big|^2 \\ +\bigg|\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((s,k),(u,j^\prime))du\bigg|^2+\bigg|\int_{t_k}^s \Psi_A((s,k),(u,k)) du\bigg|^2\bigg]ds. \end{multline*} \]

Using the triangle and Cauchy-Schwartz inequalities, we get

\[ \begin{align*} \tilde{\mathcal{G}}_F(k)& \geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2ds- \rho c_{H_c}^2(2(k-j^\prime)+1)\times \\ &~~\times\int_{t_k}^{t_{k+1}}\bigg[\sum_{q = j^\prime+1}^{k-1}(t_{q+1}-t_q)\int_{t_q}^{t_{q+1}}\big| \Psi_A((s,k),(u,q))\big|^2du+ \sum_{q = j^\prime}^{k-1}\big|\Psi_B((s,k),(t_{q+1},q))\big|^2 \\ &~~+(t_{j^\prime+1}-t^\prime)\int_{t^\prime}^{t_{j^\prime+1}}\big| \Psi_A((s,k),(u,j^\prime))\big|^2du+(s-t_k)\int_{t_k}^s \big|\Psi_A((s,k),(u,k))\big|^2du\bigg]ds\\ &\geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2ds- \rho c_{H_c}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1) \times \\ &~~\times\int_{t_k}^{t_{k+1}}\bigg[\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}}\big| \Psi_A((s,k),(u,q))\big|^2du + \sum_{q = j^\prime}^{k-1}\big|\Psi_B((s,k),(t_{q+1},q))\big|^2 \\ &~~{}+\int_{t^\prime}^{t_{j^\prime+1}}\big| \Psi_A((s,k),(u,j^\prime))\big|^2du+\int_{t_k}^s \big|\Psi_A((s,k),(u,k))\big|^2du\bigg]ds. \end{align*} \]

Using the bounds on \( \Psi_A\) and \( \Psi_B\) , we get

\[ \begin{align*} \tilde{\mathcal{G}}_F(k) & \geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2ds \\ & ~~ {}- \rho c_{H_c}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\max\{c_A^2,c_B^2\}\int_{t_k}^{t_{k+1}}\bigg[\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}}\big| H_c(u,q) \tilde{x}(u,q)\big|^2du \\ &~~ {}+ \sum_{q = j^\prime}^{k-1}\big|H_d(t_{q+1},q)\tilde{x}(t_{q+1},q)\big|^2 +\int_{t^\prime}^{t_{j^\prime+1}} \big|H_c(u,j^\prime) \tilde{x}(u,j^\prime)\big|^2du +\int_{t_k}^s \big|H_c(u,k) \tilde{x}(u,k)\big|^2du\bigg]ds \\ &\geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2ds\\ &~~{} - \rho c_{H_c}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\max\{c_A^2,c_B^2\}(t_{k+1} - t_k)\times \\ &~~\times\bigg[\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}}\big| H_c(u,q) \tilde{x}(u,q)\big|^2du + \sum_{q = j^\prime}^{k-1}\big|H_d(t_{q+1},q)\tilde{x}(t_{q+1},q)\big|^2 \\ &~~{}+\int_{t^\prime}^{t_{j^\prime+1}} \big|H_c(u,j^\prime) \tilde{x}(u,j^\prime)\big|^2du +\int_{t_k}^{t_{k+1}} \big|H_c(u,k) \tilde{x}(u,k)\big|^2du\bigg] \\ &\geq \frac{\rho}{1+\rho}\int_{t_k}^{t_{k+1}} \big|H_c(s,k)\Phi_{F,J}((s,k),(t,j))\tilde{x}(t,j)\big|^2ds \\ &~~- \rho c_{H_c}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\max\{c_A^2,c_B^2\} (t_{k+1} - t_k)\tilde{\mathcal{G}}. \end{align*} \]

Similarly to \( \tilde{\mathcal{G}}_F(k)\) , we get for the two separate integral terms of \( \tilde{\mathcal{G}}\) ,

\[ \begin{align*} \int_{t^\prime}^{t_{j^\prime+1}}\star^\top H_c(s,j^\prime) \tilde{x}(s,j^\prime)ds &\geq \frac{\rho}{1+\rho}\int_{t^\prime}^{t_{j^\prime+1}} \big|H_c(s,j^\prime)\Phi_{F,J}((s,j^\prime),(t,j))\tilde{x}(t,j)\big|^2ds\\ &~~{} - \rho c_{H_c}^2(t_{j^\prime+1} - t^\prime)\max\{c_A^2,c_B^2\} (t_{j^\prime+1} - t^\prime)\tilde{\mathcal{G}}, \\ \int_{t_{j}}^{t}\star^\top H_c(s,j) \tilde{x}(s,j)ds &\geq \frac{\rho}{1+\rho}\int_{t_j}^{t} \big|H_c(s,j)\Phi_{F,J}((s,j),(t,j))\tilde{x}(t,j)\big|^2ds \\ &~~{} - \rho c_{H_c}^2(2(j-j^\prime)+1)(t-t^\prime+1)\max\{c_A^2,c_B^2\}(t - t_j)\tilde{\mathcal{G}}. \end{align*} \]

For \( \tilde{\mathcal{G}}_J(k)\) , using (385.a) with \( (t_{k+1},k)\) replacing \( (t,j)\) , we get

\[ \begin{align*} \tilde{\mathcal{G}}_J(k) &= \bigg|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime) - H_d(t_{k+1},k)\sum_{q = j^\prime+1}^{k-1}\int_{t_q}^{t_{q+1}} \Psi_A((t_{k+1},k),(u,q))du \\ &~~- H_d(t_{k+1},k)\sum_{q = j^\prime}^{k-1}\Psi_B((t_{k+1},k),(t_{q+1},q)) -H_d(t_{k+1},k)\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((t_{k+1},k),(u,j^\prime))du \\ &~~-H_d(t_{k+1},k)\int_{t_k}^{t_{k+1}} \Psi_A((t_{k+1},k),(u,k))du\bigg|^2\\ &= \bigg|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime) - H_d(t_{k+1},k)\sum_{q = j^\prime+1}^{k}\int_{t_q}^{t_{q+1}} \Psi_A((t_{k+1},k),(u,q))du \\ &~~-H_d(t_{k+1},k)\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((t_{k+1},k),(u,j^\prime))du- H_d(t_{k+1},k)\sum_{q = j^\prime}^{k-1}\Psi_B((t_{k+1},k),(t_{q+1},q))\bigg|^2. \end{align*} \]

Using \( |a-b|^2 \geq \frac{\rho}{1+\rho}|a|^2-\rho|b|^2\) for the same \( \rho\) , we get

\[ \begin{multline*} \tilde{\mathcal{G}}_J(k) \geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2 - \rho c_{H_d}^2\bigg|\sum_{q = j^\prime+1}^{k}\int_{t_q}^{t_{q+1}} \Psi_A((t_{k+1},k),(u,q))du\\ +\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((t_{k+1},k),(u,j^\prime))du+ \sum_{q = j^\prime}^{k-1}\Psi_B((t_{k+1},k),(t_{q+1},q))\bigg|^2. \end{multline*} \]

Applying \( \left|\sum_{i=1}^Na_i\right|^2 \leq N\sum_{i=1}^N|a_i|^2\) , we get

\[ \begin{multline*} \tilde{\mathcal{G}}_J(k) \geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2 \\ - \rho c_{H_d}^2(2(k-j^\prime)+1)\bigg[\sum_{q = j^\prime+1}^{k}\bigg|\int_{t_q}^{t_{q+1}} \Psi_A((t_{k+1},k),(u,q))du\bigg|^2 \\ +\bigg|\int_{t^\prime}^{t_{j^\prime+1}} \Psi_A((t_{k+1},k),(u,j^\prime))du\bigg|^2+\sum_{q = j^\prime}^{k-1}\big|\Psi_B((t_{k+1},k),(t_{q+1},q))\big|^2\bigg]. \end{multline*} \]

Using the triangle and Cauchy-Schwartz inequality, we get

\[ \begin{align*} \tilde{\mathcal{G}}_J(k) &\geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2\\ &~~{}- \rho c_{H_d}^2(2(k-j^\prime)+1)\bigg[\sum_{q = j^\prime+1}^{k}(t_{q+1}-t_q)\int_{t_q}^{t_{q+1}} \big|\Psi_A((t_{k+1},k),(u,q))\big|^2du \\ & ~~{}+(t_{j^\prime+1}-t^\prime)\int_{t^\prime}^{t_{j^\prime+1}} \big|\Psi_A((t_{k+1},k),(u,j^\prime))\big|^2du+\sum_{q = j^\prime}^{k-1}\big|\Psi_B((t_{k+1},k),(t_{q+1},q))\big|^2\bigg] \\ &\geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2\\ &~~ {}- \rho c_{H_d}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\bigg[\sum_{q = j^\prime+1}^{k}\int_{t_q}^{t_{q+1}} \big|\Psi_A((t_{k+1},k),(u,q))\big|^2du \\ &~~{}+\int_{t^\prime}^{t_{j^\prime+1}} \big|\Psi_A((t_{k+1},k),(u,j^\prime))\big|^2du+\sum_{q = j^\prime}^{k-1}\big|\Psi_B((t_{k+1},k),(t_{q+1},q))\big|^2\bigg]. \end{align*} \]

Using the bounds on \( \Psi_A\) and \( \Psi_B\) , we get

\[ \begin{align*} \tilde{\mathcal{G}}_J(k) &\geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t^\prime,j^\prime))\tilde{x}(t^\prime,j^\prime)\big|^2 \\ &~~{}- \rho c_{H_d}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\max\{c_A^2,c_B^2\}\bigg[\sum_{q = j^\prime+1}^{k}\int_{t_q}^{t_{q+1}} \big|H_c(u,q) \tilde{x}(u,q)\big|^2du \\ &~~{}+\int_{t^\prime}^{t_{j^\prime+1}} \big|H_c(u,j^\prime)\tilde{x}(u,j^\prime)\big|^2du +\sum_{q = j^\prime}^{k-1}\big|H_d(t_{q+1},q)\tilde{x}(t_{q+1},q)\big|^2\bigg] \\ &\geq \frac{\rho}{1+\rho}\big|H_d(t_{k+1},k)\Phi_{F,J}((t_{k+1},k),(t,j))\tilde{x}(t,j)\big|^2 - \rho c_{H_d}^2(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1) \max\{c_A^2,c_B^2\}\tilde{\mathcal{G}}. \end{align*} \]

Now let us lower-bound \( \tilde{\mathcal{G}}\) by summing the obtained inequalities. Since \( (t-t^\prime) + (j-j^\prime) \leq \Delta_m\) , we get

\[ \begin{align*} \tilde{\mathcal{G}} &\geq \frac{\rho}{1+\rho}(\tilde{x}(t,j))^\top\mathcal{G}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \tilde{x}(t,j)- \rho \max\{c_A^2,c_B^2\} \{c_{H_c}^2,c_{H_d}^2\}\tilde{\mathcal{G}} \times\\ &~~{}\times\bigg[\sum_{k=j^\prime+1}^{j-1}(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)(t_{k+1}-t_k)+\sum_{k=j^\prime}^{j-1}(2(k-j^\prime)+1)(t_{k+1}-t^\prime+1)\\ &~~{}+ (2(j-j^\prime)+1)(t-t^\prime+1)(t-t_j) + (t_{j^\prime+1}-t^\prime)^2\bigg]\\ &\geq \frac{\rho}{1+\rho}(\tilde{x}(t,j))^\top\mathcal{G}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \tilde{x}(t,j)- \rho \max\{c_A^2,c_B^2\} \{c_{H_c}^2,c_{H_d}^2\}\tilde{\mathcal{G}} \times\\ & ~~\times[(j-j^\prime-1)(2(j-j^\prime)-1)(t_j-t^\prime+1)(t_j - t_{j^\prime+1})+ (j-j^\prime)(2(j-j^\prime)-1)(t_j-t^\prime+1)\\ & ~~+(2(j-j^\prime)+1)(t-t^\prime+1)(t-t_j) + (t_{j^\prime+1}-t^\prime)^2]\\ &\geq \frac{\rho}{1+\rho}(\tilde{x}(t,j))^\top\mathcal{G}_{(F,J,H_c,H_d)}((t^\prime,j^\prime),(t,j)) \tilde{x}(t,j) - \rho \max\{c_A^2,c_B^2\} \{c_{H_c}^2,c_{H_d}^2\}\tilde{\mathcal{G}} \times\\ & ~~\times[\Delta_m^2(2\Delta_m-1)(\Delta_m+1)+\Delta_m(2\Delta_m-1)(\Delta_m+1)+(2\Delta_m+1)(\Delta_m+1)\Delta_m + \Delta_m^2], \end{align*} \]

and thus, the result follows. \( \blacksquare\)

6.2.1.2 Technical Lemmas and Proofs of Section 3.2.3

Consider a hybrid system of the form

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{z}}_o &=&{} -P^{-1}(H_{c,o}(\tau))^\top (R(\tau))^{-1}H_{c,o}(\tau)\tilde{z}_o \\ \dot{\tilde{\eta}} & = & 0 \\ \dot{P} &= &{}-\lambda P + (H_{c,o}(\tau))^\top (R(\tau))^{-1} H_{c,o}(\tau) \\ \dot{\tau} &=& 1 \\ \\ \tilde{z}_o^+ &=& M_o(u,\tau) \tilde{z}_o + M_{o\eta}(u,\tau) \tilde{\eta} \\ \tilde{\eta}^+ & = & M_{\eta o}(u,\tau) \tilde{z}_o+M_{\eta}(\tau)\tilde{\eta}\\ P^+ & = & P_0 \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(386.a)

where \( u \in \mathfrak{U}\) is the input, with the flow and jump sets

\[ \begin{equation} \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times \mathbb{R}^{n_o \times n_o} \times [0,\tau_M], ~~ \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times \mathbb{R}^{n_o \times n_o} \times \mathcal{I}, \end{equation} \]

(386.b)

where \( \mathcal{I}\) is a compact subset of \( [\tau_m,\tau_M]\) for some positive \( \tau_m\) and \( \tau_M\) .

Lemma 23 (Exponential stability of the estimation error)

Assume that:

  1. \( H_{c,o}\) is continuous on \( [0,\tau_M]\) and such that the pair \( (0, H_{c,o}(\tau))\) satisfies (186);
  2. \( R\) is continuous on \( [0,\tau_M]\) and \( R(\tau) > 0\) for all \( \tau \in [0,\tau_M]\) ;
  3. \( M_o\) , \( M_{o\eta}\) , and \( M_{\eta o}\) are bounded on \( \mathfrak{U} \times \mathcal{I}\) ;
  4. \( M_\eta\) is continuous on \( \mathcal{I}\) and there exists \( Q_\eta \in §_{>0}^{n_\eta}\) such that

    \[ \begin{equation} (M_\eta(\tau))^\top Q_\eta M_\eta(\tau) - Q_\eta < 0, ~~ \forall \tau \in \mathcal{I}. \end{equation} \]

    (387)

Then for any \( P_0 \in §_{>0}^{n_o}\) , there exists \( \lambda^\star > 0\) such that for any \( \lambda > \lambda^\star\) , there exist \( \rho_1>0\) and \( \lambda_1>0\) such that any maximal solution \( (\tilde{z}_o, \tilde{\eta}, P, \tau)\) to system (386) with \( P(0,0) = P_0\) , \( \tau(0,0) = 0\) , and \( u\in \mathfrak{U}\) , is complete and verifies

\[ \begin{equation} |(\tilde{z}_o,\tilde{\eta})(t,j)| \leq \rho_1 e^{-\lambda_1(t+j)}|(\tilde{z}_o,\tilde{\eta})(0,0)|, ~~ \forall (t,j) \in \rm{dom} (\tilde{z}_o, \tilde{\eta}, P, \tau). \end{equation} \]

(388)

Proof. First, due to the compactness of \( \mathcal{I}\) and Item (LIV) of Lemma 23, there exists \( a > 0\) such that for all \( \tau \in \mathcal{I}\) , (387) is strengthened into

\[ \begin{equation} (M_\eta(\tau))^\top Q_\eta M_\eta(\tau) - Q_\eta \leq -aQ_\eta, ~~ \forall \tau \in \mathcal{I}. \end{equation} \]

(389)

Since \( P(0,0)=P_0\) , \( P^+=P_0\) , and \( \tau(0,0)=0\) , the component \( (t,j)\mapsto P(t,j)\) of the solution to system (386) can actually be written as a closed form of the component \( (t,j)\mapsto \tau(t,j)\) by defining

\[ \begin{equation} \mathbb{P}(\tau) = e^{-\lambda \tau} P_0 + \int_{0}^{\tau} e^{-\lambda (\tau-s)} (H_{c,o}(s))^\top (R(s))^{-1} H_{c,o}(s) ds, \end{equation} \]

(390)

namely \( P(t,j)=\mathbb{P}(\tau(t,j))\) for all \( (t,j)\in \rm{dom} x\) . Note that since \( P_0>0\) , \( \mathbb{P}(\tau)\) is invertible for all \( \tau\in [0,\tau_M]\) . It follows that the maximal solution is complete and \( (\tilde{z}_o, \tilde{\eta},\tau)\) is solution to

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{z}}_o &=& {}-\mathbb{P}^{-1}(\tau)(H_{c,o}(\tau))^\top (R(\tau))^{-1} H_{c,o}(\tau)\tilde{z}_o \\ \dot{\tilde{\eta}} & = & 0 \\ \dot{\tau} &=& 1 \\ \\ \tilde{z}_o^+ &=& M_o(u,\tau) \tilde{z}_o + M_{o\eta}(u,\tau) \tilde{\eta} \\ \tilde{\eta}^+ & = & M_{\eta o}(u,\tau) \tilde{z}_o+M_{\eta}(\tau)\tilde{\eta} \\ \tau^+ &=& 0, \end{array} \right. \end{equation} \]

(391)

with the flow set \( \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times [0,\tau_M]\) and the jump set \( \mathbb{R}^{n_o} \times \mathbb{R}^{n_\eta} \times \mathcal{I}\) . Consider the Lyapunov function

\[ \begin{equation} V(\tilde{z}_o, \tilde{\eta}, \tau) = e^{\frac{\lambda}{2} \tau} \tilde{z}_o^\top \mathbb{P}(\tau) \tilde{z}_o + ke^{-\epsilon \tau} \tilde{\eta}^\top Q_{\eta} \tilde{\eta}, \end{equation} \]

(392)

where \( k>0\) and \( \epsilon>0\) . We have for all \( \tau \in [\tau_m, \tau_M] \supseteq \mathcal{I}\) ,

\[ \begin{align} e^{\frac{\lambda }{2}\tau} \mathbb{P}(\tau)&\geq e^{\frac{\lambda }{2}\tau} \int_{\frac{3}{4}\tau}^{\tau} e^{-\lambda (\tau-s)} (H_{c,o}(s))^\top (R(s))^{-1} H_{c,o}(s) ds \notag \end{align} \]

(393)

\[ \begin{align} &\geq r_m e^{\frac{\lambda }{4}\tau} \int_{\frac{3}{4}\tau}^{\tau} \mathcal{D}_o^\top e^{F^\top s} H_c^\top H_c e^{F s} \mathcal{D}_o ds \notag \\ &\geq r_m e^{\frac{\lambda }{4}\tau_m} \int_{\frac{3}{4}\tau}^{\frac{3}{4}\tau+\frac{1}{4}\tau_m} \mathcal{D}_o^\top e^{F^\top s} H_c^\top H_c e^{F s} \mathcal{D}_o ds \notag \\ &\geq e^{\frac{\lambda }{4}\tau_m} r_m\alpha \rm{Id} \notag\\ &:= e^{\frac{\lambda }{4}\tau_m}\lambda_m \rm{Id}, \\\end{align} \]

(394)

where \( r_m>0\) is a lower bound of the continuous map \( R\) on the compact set \( [0,\tau_M]\) (thanks to Item (LII) of Lemma 23), \( \alpha >0\) (independent of \( \lambda\) ) is obtained by applying (186) with \( \delta = \frac{\tau_m}{4}\) , and \( \lambda_m:= r_m\alpha\) . On the other hand, from (390), for all \( \tau \in [0, \tau_m]\) , \( \mathbb{P}(\tau) \geq e^{-\lambda \tau_m} P_0\) , so \( e^{{\frac{\lambda }{2}\tau}}\mathbb{P}(\tau) \geq e^{-\lambda \tau_m} P_0\) . Besides, as \( \mathbb{P}\) is continuous on the compact set \( [0, \tau_M]\) , there exists \( p_M>0\) such that \( \mathbb{P}(\tau) \leq p_M \rm{Id}\) for all \( \tau\in [0, \tau_M]\) . It then follows that there exist \( \underline{\rho}>0\) and \( \overline{\rho}>0\) defined as

\[ \begin{align} \underline{\rho} &= \min\left\{\rm{eig}\left(e^{-\lambda \tau_m} P_0\right), e^{\frac{\lambda }{4}\tau_m} \lambda_m, \rm{eig}\left(ke^{-\epsilon\tau_M} Q_\eta\right) \right\}, \\ \overline{\rho} &= \max\left\{e^{\frac{\lambda}{2} \tau_M}p_M, \rm{eig}\left(ke^{-\epsilon \tau_m} Q_\eta\right)\right\}, \\\end{align} \]

(395.a)

such that

\[ \begin{equation} \underline{\rho} |(\tilde{z}_o, \tilde{\eta})|^2 \leq V(\tilde{z}_o, \tilde{\eta}, \tau) \leq \overline{\rho} |(\tilde{z}_o, \tilde{\eta})|^2, ~~ \forall (\tilde{z}_o, \tilde{\eta}) \in \mathbb{R}^n, \forall \tau \in [0, \tau_M]. \end{equation} \]

(396)

During flows, for all \( (\tilde{z}_o, \tilde{\eta}) \in \mathbb{R}^n\) and \( \tau \in [0, \tau_M]\) ,

\[ \begin{align} \dot{V}& = e^{\frac{\lambda}{2} \tau} \tilde{z}_o^\top \left(\frac{\lambda}{2}\mathbb{P}(\tau) - 2(H_{c,o}(\tau))^\top (R(\tau))^{-1} H_{c,o}(\tau) + \dot{\mathbb{P}}(\tau) \right) \tilde{z}_o -\epsilon ke^{-\epsilon \tau} \tilde{\eta}^\top Q_\eta \tilde{\eta} \notag \end{align} \]

(397)

\[ \begin{align} & = e^{\frac{\lambda}{2} \tau} \tilde{z}_o^\top \left(-\frac{\lambda}{2}\mathbb{P}(\tau) - (H_{c,o}(\tau))^\top (R(\tau))^{-1} H_{c,o}(\tau)\right) \tilde{z}_o -\epsilon ke^{-\epsilon \tau} \tilde{\eta}^\top Q_\eta \tilde{\eta} \notag\\ & \leq -\frac{\lambda}{2} e^{\frac{\lambda}{2} \tau} \tilde{z}_o^\top \mathbb{P}(\tau)\tilde{z}_o - \epsilon ke^{-\epsilon \tau} \tilde{\eta}^\top Q_\eta \tilde{\eta}\notag\\ &\leq -\min\left\{\frac{\lambda}{2}, \epsilon\right\}V. \\\end{align} \]

(398)

At jumps, for all \( (\tilde{z}_o, \tilde{\eta}) \in \mathbb{R}^n\) , \( u \in \mathfrak{U}\) , and \( \tau \in \mathcal{I}\) ,

\[ \begin{multline} V^+ - V= \star^\top P_0 (M_o(u,\tau) \tilde{z}_o + M_{o\eta}(u,\tau) \tilde{\eta}) - e^{\frac{\lambda}{2} \tau} \tilde{z}_o^\top \mathbb{P}(\tau) \tilde{z}_o \\ +k \star^\top Q_\eta (M_{\eta o}(u,\tau) \tilde{z}_o+M_{\eta}(\tau)\tilde{\eta})- ke^{-\epsilon \tau} \tilde{\eta}^\top Q_\eta \tilde{\eta}. \end{multline} \]

(399)

From Young’s inequality, (393), (389), and Item (LIII) and Item (LIV) of Lemma 23, there exist non-negative constants \( c_i, i = 1, 2, …, 5\) independent of \( (\lambda,k,\epsilon)\) such that for any \( \kappa>0\) , for all \( (\tilde{z}_o, \tilde{\eta}) \in \mathbb{R}^n\) , \( u \in \mathfrak{U}\) , and \( \tau \in \mathcal{I}\) ,

\[ \begin{equation} V^+ - V \leq \left(c_1 + k c_2 + \kappa c_3- e^{\frac{\lambda}{4}\tau_{m}}\lambda_m\right)\tilde{z}_o^\top \tilde{z}_o - \left(k\left(a - \left(1 - e^{-\epsilon\tau_{M}}\right)\right) - c_4 - \frac{k^2 c_5}{\kappa}\right)\tilde{\eta}^\top Q_\eta \tilde{\eta}. \end{equation} \]

(400)

We now show that this quantity can be made negative definite by successively picking the degrees of freedom. For the \( \tilde{\eta}\) part, \( \exists \epsilon^\star > 0\) such that \( 0 < \epsilon < \epsilon^\star \implies a - \left(1 - e^{-\epsilon\tau_{M}}\right)> 0\) , \( \exists k^\star > 0\) such that \( k > k^\star \implies k\left(a - \left(1 - e^{-\epsilon\tau_{M}}\right)\right) - c_4> 0\) , and \( \exists \kappa^\star\) such that \( \kappa > \kappa^\star \implies k\left(a - \left(1 - e^{-\epsilon\tau_{M}}\right)\right) - c_4 - \frac{k^2c_5}{\kappa} > 0\) . Then, for the \( \tilde{z}_{o}\) part, \( \exists \lambda^\star > 0\) such that \( \lambda > \lambda^\star \implies c_1 + k c_2 + \kappa c_3 - e^{\frac{\lambda}{4}\tau_{m}}\lambda_m <0\) . We deduce that for any \( \lambda>\lambda^\star\) , there exist \( a_c>0\) and \( a_d>0\) such that for all \( (\tilde{z}_o, \tilde{\eta}) \in \mathbb{R}^n\) ,

\[ \begin{align} \dot{V} &\leq -a_c V, ~~ \forall \tau \in [0, \tau_M], \\ V^+ - V&\leq -a_d V, ~~ \forall u \in \mathfrak{U}, \forall \tau \in \mathcal{I}. \\\end{align} \]

(401.a)

From (396) and (401), we conclude according to [10, Definition 7.29 and Theorem 7.30] that the set \( \mathcal{A} = \{(\tilde{z}_o, \tilde{\eta}, \tau) \in \mathbb{R}^{n_o} \times \mathbb{R}^{n_{no}} \times [0,\tau_M]: \tilde{z}_o = 0, \tilde{\eta} = 0\}\) is GES for system (391). \( \blacksquare\)

Corollary 2 (Arbitrarily fast exponential stability of the estimation error)

Let us now consider system (386) with \( M_\eta(\tau)\) replaced by \( \gamma M_\eta(\tau)\) for \( \gamma \in (0,\gamma^\star_0]\) . Under the same assumptions as in Lemma 23, for any \( \lambda_c > 0\) and any \( P_0 \in §_{>0}^{n_o}\) , there exists \( \gamma^\star > 0\) such that there exists \( \lambda^\star>0\) such that for any \( 0<\gamma < \gamma^\star\) and for any \( \lambda > \lambda^\star\) , there exists \( \rho_c > 0\) such that any maximal solution \( (\tilde{z}_o, \tilde{\eta}, P, \tau)\) of the new system (386), with \( P(0,0) = P_0\) , \( \tau(0,0) = 0\) , and \( u\in \mathfrak{U}\) , is complete and verifies

\[ \begin{equation} |(\tilde{z}_o,\tilde{\eta})(t,j)| \leq \rho_c e^{-\lambda_c(t+j)}|(\tilde{z}_o,\tilde{\eta})(0,0)|, ~~ \forall (t,j) \in \rm{dom} (\tilde{z}_o, \tilde{\eta}, P, \tau). \end{equation} \]

(402)

Proof. This is a modification of the proof of Lemma 23. Consider the Lyapunov function in (392). First, let us show with an appropriate choice of \( \epsilon\) that for \( \lambda\) sufficiently large and \( \gamma\) sufficiently small, we have for some \( a_d > 0\) ,

\[ \begin{equation} \dot{V} \leq -2 \lambda_c\left(\frac{1}{\tau_m}+1\right) V, ~~ V^+ < V. \end{equation} \]

(403)

Following the same analysis as in the proof of Lemma 23, we obtain that during flows, \( \dot{V} \leq-\min\left\{\frac{\lambda}{2}, \epsilon\right\}V\) for all \( \tau\in[0,\tau_M]\) , and at jumps thanks to (387), for all \( u \in \mathfrak{U}\) and \( \tau \in \mathcal{I}\) ,

\[ \begin{equation} V^+ - V \leq \left(c_1 + k c_2 + \kappa c_3- e^{\frac{\lambda}{4}\tau_{m}}\lambda_m\right)\tilde{z}_o^\top \tilde{z}_o - \left(k\left(e^{-\epsilon\tau_{M}} - \gamma^2\right) - c_4 - \frac{k^2 c_5}{\kappa}\right)\tilde{\eta}^\top Q_\eta \tilde{\eta}. \end{equation} \]

(404)

Let us pick \( \epsilon = 2\lambda_c\left(\frac{1}{\tau_m}+1\right)\) and define \( \lambda_0^\star := 4\lambda_c\left(\frac{1}{\tau_m}+1\right)\) . Then, the first item in (403) holds as soon as \( \lambda > \lambda_0^\star\) . Now define \( \gamma^\star := \min\left\{\gamma_0^\star,\sqrt{e^{-\epsilon\tau_{M}}}\right\}\) . For any \( 0 < \gamma < \gamma^\star\) , we have \( e^{-\epsilon\tau_{M}} - \gamma^2>e^{-\epsilon\tau_{M}} - (\gamma^\star)^2>0\) , then \( k\) , \( \kappa\) , and \( \lambda\) are successively picked (based on \( \gamma^\star\) ) as in the proof of Lemma 23. The final \( \lambda^\star\) is the larger one between this and \( \lambda_0^\star\) . For any \( \lambda>\lambda^\star\) and for any \( 0<\gamma<\gamma^\star\) , we obtain (403). Second, we deduce (402) from (403) and the dwell time condition. From (403), we get \( V(t,j) \leq e^{-2\lambda_c\left(\frac{1}{\tau_m}+1\right)t}V(0,0)\) for all \( (t,j) \in \rm{dom} (\tilde{z}_o, \tilde{\eta}, P, \tau)\) . Since the flow lengths of solutions to system (386) are at least \( \tau_m > 0\) for all \( j\geq 1\) , we have \( j \leq \frac{t}{\tau_m}+1\) so that \( t \geq \frac{t+j-1}{\frac{1}{\tau_m} + 1}\) , for all \( (t,j) \in \rm{dom} (\tilde{z}_o, \tilde{\eta}, P, \tau)\) . Therefore, \( V(t,j) \leq e^{2\lambda_c} e^{-2\lambda_c(t+j)}V(0,0)\) , for all \( (t,j) \in \rm{dom} (\tilde{z}_o, \tilde{\eta}, P, \tau)\) , implying (402). \( \blacksquare\)

Lemma 24 (Boundedness in finite time)

Consider a hybrid system with state \( \eta \in \mathbb{R}^{n_\eta}\) and input \( u\in \mathfrak{U} \subset \mathbb{R}^{n_u}\) :

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{~~}l} \dot{\eta}&=&M_c(u)\eta & (\eta,u)\in C \\ \eta^+&=&M_d(u)\eta & (\eta, u)\in D \end{array} \right. \end{equation} \]

(405)

with \( M_c, M_d: \mathbb{R}^{n_u} \to \mathbb{R}^{n_\eta \times n_\eta}\) . Consider positive scalars \( m_c\) , \( \rho_d\) , and \( \tau_M\) , as well as \( j_m \in \mathbb{N}\) . Then, there exists \( \rho > 0\) such that for any solution \( (\eta,u)\) to system (405) with flow lengths in \( [0,\tau_M]\) and such that \( \|M_c(u(t,j))\|\leq m_c\) and \( \|M_d(u(t,j))\|\leq \rho_d\) for all \( u \in \mathfrak{U}\) and \( (t,j)\in \rm{dom} x\) , we have \( j_m\in \rm{dom}_j \eta\) and

\[ \begin{equation} |\eta(t_{j_m},j_m)| \leq \rho |\eta(0,0)|. \end{equation} \]

(406)

Proof. First, the flow lengths of solutions to system (405) are in \( [0,\tau_m]\) and maximal solutions are both \( t\) - and \( j\) -complete. During flows, the evolution of \( \eta\) is characterized by the transition matrix \( \Psi_{M_c(u)}\) as

\[ \begin{equation} \eta(t,j) = \Psi_{M_c(u), u \in \mathfrak{U}}(t,t_{j-1})\eta(t_{j-1},j-1). \end{equation} \]

(407)

If \( M_c\) is uniformly bounded for \( u \in \mathfrak{U}\) , there exists \( \rho_c > 0\) such that for all \( (t,j) \in \rm{dom} \eta\) and all \( u \in \mathfrak{U}\) , \( |\eta(t,j)| \leq \rho_c |\eta(t_{j-1},j-1)|\) . Next, we have for all \( j \in \rm{dom}_j \eta\) , \( |\eta(t_j,j)| \leq \rho_d|\eta(t_j,j-1)|\) . Therefore, for any \( j_m \in \mathbb{N}_{>0}\) , we have \( |\eta(t_{j_m},j_m)| \leq \rho_c^{j_m} \rho_d^{j_m-1} |\eta(0,0)|\) , which is (406) by seeing that \( \rho = \rho_c^{j_m} \rho_d^{j_m-1}\) . \( \blacksquare\)

6.2.2 Technical Lemmas and Proofs of Section 3.3

6.2.2.1 Lyapunov-Based Sufficient Conditions for Coupling Observers

This appendix presents preliminary results that have been published in [166]. These are sufficient conditions to couple a flow-based observer with a jump-based one using Lyapunov analysis, which are used throughout Section 3.3 to design observers for systems of the form (408).

Consider a hybrid system of the form

\[ \begin{equation} \left\{ \begin{array}{@{}l@{~~}l@{}} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{z}_o &=& f_o(z_o, u_c) \\ \dot{z}_{no} &=& f_{no}(z_o, z_{no}, u_c) \end{array} \right\} (z,u_c)\in C_z & y_c = h_o(z_o,u_c)\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} z_o^+ &=& g_o(z_o, z_{no}, u_d) \\ z_{no}^+ &=& g_{no}(z_o, z_{no}, u_d) \end{array} \right\} (z,u_d)\in D_z & y_d = h_{no}(z_o,z_{no},u_d), \end{array} \right. \end{equation} \]

(408)

with state \( z = (z_o,z_{no}) \in \mathbb{R}^{n_z}\) , where \( z_o \in \mathbb{R}^{n_o}\) and \( z_{no} \in \mathbb{R}^{n_{no}}\) . Denote \( \mathcal{Z}_0 \subseteq \mathbb{R}^{n_z}\) as the set of initial conditions of interest. Denote \( \mathfrak{U}_c\) (resp., \( \mathfrak{U}_d\) ) as the set of flow (resp., jump) input trajectories of interest, and \( \mathcal{U}_c\) and \( \mathcal{U}_d\) as the sets of values these considered inputs can take. The following assumption is made.

Assumption 38

For system (408), we assume that:

  • There exists a compact set \( \mathcal{I} \subset \mathbb{R}_{>0}\) such that each maximal solution to system (408) initialized in \( \mathcal{Z}_0\) and with inputs in \( \mathfrak{U}_c\times\mathfrak{U}_d\) have flow lengths within \( \mathcal{I}\) ;
  • The flow pair \( (f_o,h_o)\) is independent of \( z_{no}\) and is instantaneously observable on \( C_z\) for any flow inputs in \( \mathfrak{U}_c\) .

This observability condition allows us to consider a high-gain observer of the pair \( (f_o,h_o)\) during flows, which estimates \( z_o\) arbitrarily fast from the knowledge of \( y_c\) and exhibits Input-to-State Stability[ISS] with respect to errors in \( z_{no}\) affecting the estimate of \( z_o\) at jumps. Then, we propose to estimate \( z_{no}\) via a jump-based observer from the knowledge of \( y_d\) as well as the estimate of \( z_o\) . More precisely, our observer takes the form

\[ \begin{equation} \left\{ \begin{array}{@{}l} \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{z}}_o &=& \hat{f}_{o,\ell}(\hat{z}_o, p, \tau, y_c, u_c) \\ \dot{\hat{z}}_{no} &=& \hat{f}_{no}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_c, u_c) \\ \dot{p}&=&\varphi_{c,\ell}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_c, u_c)\\ \dot{\tau}&=&1 \end{array} \right\} \text{when~(408) flows}\\ \\ \left. \begin{array}{@{}r@{\;}c@{\;}l@{}} \hat{z}_o^+ &=& \hat{g}_o(\hat{z}_o, \hat{z}_{no}, p, \tau, y_d, u_d) \\ \hat{z}_{no}^+ &=& \hat{g}_{no}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_d, u_d) \\ p^+ & = & \varphi_{d,\ell}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_d, u_d)\\ \tau^+ &=&0 \end{array} \right\} \text{when~(408) jumps} \end{array} \right. \end{equation} \]

(409)

where \( \hat{z}_o \in \mathbb{R}^{n_o}\) , \( \hat{z}_{no} \in \mathbb{R}^{n_{no}}\) , \( \hat{z} = (\hat{z}_o,\hat{z}_{no}) \in \mathbb{R}^{n_z}\) , \( p \in \mathbb{R}^{n_p}\) might contain additional observer states (see Example 22) and \( \tau\) is a timer keeping track of the time elapsed since the previous jump. This timer evolves in \( [0,\max\mathcal{I}]\) during flows and is in \( \mathcal{I}\) at jumps according to Assumption 38. The map \( \hat{f}_{o,\ell}\) corresponds to a high-gain observer, possibly with extra states contained in \( p\) . The maps \( \hat{f}_{no}\) , \( \varphi_{c,\ell}\) , \( \hat{g}_o\) , \( \hat{g}_{no}\) , and \( \varphi_{d,\ell}\) are to be designed.

We now present intermediary technical results, namely sufficient Lyapunov-based conditions for coupling:

  • An arbitrarily fast high-gain flow-based observer estimating \( \xi_o\) , designed using continuous-time observer theory with a Lyapunov function \( V_{o,\ell}\) depending on the gain \( \ell\) of the observer. This function is evaluated along the dynamics of the estimation error \( z_o - \hat{z}_o\) :

    \[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{z}_o - \dot{\hat{z}}_o &=& f_o(z_o, u_c) - \hat{f}_{o,\ell}(\hat{z}_o, p, \tau, y_c, u_c)\\ z_o^+ -\hat{z}_o^+ &=& g_o(z_o, z_{no}, u_d)-\hat{g}_o(\hat{z}_o, \hat{z}_{no}, p, \tau, y_d, u_d); \end{array} \right. \end{equation} \]

    (410)

  • A jump-based observer estimating \( \xi_{no}\) , designed using discrete-time observer theory with a Lyapunov function \( V_{no}\) . This function is evaluated along the dynamics of the estimation error \( z_{no} - \hat{z}_{no}\) :

    \[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{z}_{no} - \dot{\hat{z}}_{no} &=& f_{no}(z_o, z_{no}, u_c) - \hat{f}_{no}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_c, u_c)\\ z_{no}^+ -\hat{z}_{no}^+ &=& g_{no}(z_o, z_{no}, u_d)-\hat{g}_{no}(\hat{z}_o, \hat{z}_{no}, p, \tau, y_d, u_d). \end{array} \right. \end{equation} \]

    (411)

This is done by studying conditions that \( V_{o,\ell}\) and \( V_{no}\) should satisfy during flows and at jumps to achieve this coupling. These Lyapunov-based conditions are general ones that are applied for designing observers throughout Section 3.3. We typically require some ISS properties from each observer with respect to the estimation error coming from the other one.

6.2.2.1.1 Conditions for Exponential Stability

Theorem 23 below gives sufficient conditions to couple a high-gain flow-based observer and a jump-based one.

Theorem 23 (Lyapunov conditions for exponential stability)

Suppose Assumption 38 holds and define \( \tau_M:=\max \mathcal{I}\) . Consider the cascade (408)-(409) and sets \( ¶_0,¶_c,¶_d\subseteq \mathbb{R}^{n_p}\) such that each solution \( (z,\hat{z},p,\tau)\) initialized in \( \mathcal{Z}_0\times \mathbb{R}^{n_z}\times ¶_0\times \{0\}\) with inputs in \( \mathfrak{U}_c\times \mathfrak{U}_d\) is such that \( p(t,j)\in ¶_c\) during flows and \( p(t,j)\in ¶_d\) at jumps. Assume that there exist a function \( V_{no}: \mathbb{R}^{n_{no}}\times \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_p} \times \mathbb{R}\to \mathbb{R}\) , scalars \( \ell_0>0\) , \( \underline{b}_{no}>0\) , \( \overline{b}_{no}>0\) , \( \lambda_c>0\) , \( a_c\) , \( c_{noo}\geq 0\) , \( d_{ono}\geq 0\) , \( a_d\) and rational functions \( \underline{b}_o>0\) , \( \overline{b}_o>0\) , \( d_o\geq 0\) , \( d_{noo}\geq 0\) such that, for any \( \ell>\ell_0\) , there exists function \( V_{o,\ell}: \mathbb{R}^{n_o}\times \mathbb{R}^{n_o} \times \mathbb{R}^{n_p} \times \mathbb{R}\to \mathbb{R}\) such that:

  1. (Uniform boundedness) For all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , \( z = (z_o,z_{no}) \in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) or \( (z,u_d)\in D_z\) , \( \hat{z}= (\hat{z}_o,\hat{z}_{no})\in \mathbb{R}^{n_z}\) , \( p \in ¶_c\cup¶_d\) , and \( \tau \in [0,\tau_M]\) ,

    \[ \begin{align} \underline{b}_{o}(\ell) |z_o-\hat{z}_o|^2 &\leq V_{o,\ell}(z_o,\hat{z}_o,p,\tau) \leq \overline{b}_{o}(\ell) |z_o - \hat{z}_o|^2, \end{align} \]

    (412.a)

    \[ \begin{align} \underline{b}_{no} |z_{no}-\hat{z}_{no}|^2 &\leq V_{no}(z_{no},\hat{z}_{no},p,\tau) \leq \overline{b}_{no} |z_{no}-\hat{z}_{no}|^2; \\ \\\end{align} \]

    (412.b)

  2. (Flow-based conditions) For all \( u_c \in \mathcal{U}_c\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c\) , and \( \tau \in [0,\tau_M]\) ,

    \[ \begin{align} \dot{V}_{o,\ell}(z,\hat{z},p,\tau,u_c) & \leq -\ell \lambda_c V_{o,\ell}(z_o,\hat{z}_o,p,\tau), \end{align} \]

    (413.a)

    \[ \begin{align} \dot{V}_{no}(z,\hat{z},p,\tau,u_c) & \leq a_c V_{no}(z_{no},\hat{z}_{no},p,\tau)+ c_{noo} V_{o,\ell}(z_o,\hat{z}_o,p,\tau); \\ \\\end{align} \]

    (413.b)

  3. (Jump-based conditions) For all \( u_d \in \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_d\) , and \( \tau \in \mathcal{I}\) ,

    \[ \begin{align} V_{o,\ell}^+(z,\hat{z},p,\tau,u_d)&\leq d_o(\ell) V_{o,\ell}(z_o,\hat{z}_o,p,\tau)+ d_{ono}|z_{no}-\hat{z}_{no}|^2, \end{align} \]

    (414.a)

    \[ \begin{align} V_{no}^+(z,\hat{z},p,\tau,u_d)&\leq e^{a_d} V_{no}(z_{no},\hat{z}_{no},p,\tau) + d_{noo}(\ell) V_{o,\ell}(z_o, \hat{z}_o,p,\tau). \\ \\\end{align} \]

    (414.b)

  4. (Overall decay of \( V_{no}\) ) \( a_c \tau_M + a_d < 0\) .

Then, there exists \( \ell^\star\geq \ell_0\) such that for any \( \ell>\ell^\star\) , there exist \( \rho>0\) and \( \lambda>0\) such that any solution \( (z,\hat{z},p,\tau)\) to the cascade (408)-(409) initialized in \( \mathcal{Z}_0\times \mathbb{R}^{n_z}\times ¶_0\times \{0\}\) with inputs in \( \mathfrak{U}_c\times \mathfrak{U}_d\) verifies

\[ \begin{equation} |z(t,j) - \hat{z}(t,j)| \leq \rho e^{-\lambda(t+j)}|z(0,0) - \hat{z}(0,0)|, ~~ \forall (t,j) \in \rm{dom} z = \rm{dom} \hat{z}. \end{equation} \]

(415)

Note that \( V_{no}\) does not need to decay during both flows and jumps, namely we do not need both \( a_c < 0\) and \( a_d < 0\) . It only has to decay from the combination of flows and jumps, characterized by the last item of Theorem 23. Notice that \( \hat{f}_{o,\ell}\) , \( \hat{g}_o\) , and the flow-based parts of \( \varphi_{c,\ell}\) and \( \varphi_{d,\ell}\) can be chosen first to guarantee the existence of \( V_{o,\ell}\) satisfying the inequalities involving it in Theorem 23 for some \( \ell_0>0\) , \( \underline{b}_{o}>0\) , \( \overline{b}_o>0\) , \( \lambda_c>0\) , \( d_o>0\) , and \( d_{ono}\geq 0\) , independently of \( \hat{f}_{no}\) , \( \hat{g}_{no}\) , and \( V_{no}\) , which can be designed in a second step.

Proof. Consider the Lyapunov function \( W_\ell: \mathbb{R}^{n_z}\times\mathbb{R}^{n_z} \times \mathbb{R}^{n_p} \times \mathbb{R}\to\mathbb{R}\) defined as

\[ \begin{equation} W_\ell(z,\hat{z},p,\tau) = e^{\frac{\lambda_c \ell \tau}{2}} V_{o,\ell}(z_o,\hat{z}_o,p,\tau)+ r e^{-\varepsilon\tau} e^{-a_c(\tau-\tau_M)} V_{no}(z_{no},\hat{z}_{no},p,\tau), \end{equation} \]

(416)

where \( r > 0\) and \( \varepsilon > 0\) are analysis parameters. The role of \( e^{\frac{\lambda_c \ell \tau}{2}}\) is to bring convergence from flows to jumps, while that of \( e^{-\varepsilon\tau}\) is indeed to bring contraction from jumps to flows; \( r\) is tuned to ensure negativity at jumps despite the interactions between these components. First, we have for all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) or \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c \cup ¶_d\) , and \( \tau \in [0,\tau_M]\) ,

\[ \begin{equation} \underline{\rho}_{o}(\ell) |z_o - \hat{z}_{o}|^2 + \underline{\rho}_{no}|z_{no}-\hat{z}_{no}|^2 \leq W_\ell(z,\hat{z},\tau) \leq \overline{\rho}_{o}(\ell) |z_o-\hat{z}_{o}|^2 + \overline{\rho}_{no}|z_{no}-\hat{z}_{no}|^2, \end{equation} \]

(417)

where \( \underline{\rho}_{o}(\ell) = \underline{b}_o(\ell)\) , \( \underline{\rho}_{no} = \underline{b}_{no}r e^{-\varepsilon\tau_M}\) , \( \overline{\rho}_{o}(\ell) = \overline{b}_o (\ell)e^{\frac{\lambda_c\ell\tau_M}{2}}\) , and \( \overline{\rho}_{no} = \overline{b}_{no}r e^{a_c\tau_M}\) . During flows, for all \( u_c \in \mathcal{U}_c\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c\) , and \( \tau \in [0,\tau_M]\) ,

\[ \begin{equation} \dot W_\ell(z,\hat{z},p,\tau,u_c) \leq - \min \left\{\frac{\lambda_c \ell}{2} -r c_{noo}, \varepsilon \right\} W_\ell(z,\hat{z},p,\tau). \end{equation} \]

(418)

At jumps, for all \( u_d \in \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_d\) , and \( \tau \in \mathcal{I}\) ,

\[ \begin{multline} W_\ell^+(z,\hat{z},p,\tau,u_d) - W_\ell(z,\hat{z},p,\tau) \leq-\left(e^{\frac{\lambda_c \ell\tau_m}{2}}\underline{b}_{o}(\ell) -d_o(\ell)\overline{b}_{o}(\ell)- r d_{noo}(\ell) \overline{b}_{o}(\ell)\right)|z_o-\hat{z}_{o}|^2 \\ - \left( r \overline{b}_{no}(e^{-\varepsilon\tau_M} - e^{a_c\tau_M+a_d}) -d_{ono}\right) |z_{no}-\hat{z}_{no}|^2, \end{multline} \]

(419)

where \( \tau_m:=\min \mathcal{I} > 0\) . Then, we choose successively

\[ 0<\varepsilon < -\frac{a_c\tau_M+a_d}{\tau_M}, ~~ r > \frac{d_{ono}}{\overline{b}_{no}(e^{-\varepsilon\tau_M} - e^{a_c\tau_M+a_d})}, \]

and finally \( \ell\) sufficiently large to have both

\[ e^{\frac{\lambda_c \ell\tau_m}{2}}\underline{b}_{o}(\ell) >d_o(\ell)\overline{b}_{o}(\ell)+ r d_{noo}(\ell) \overline{b}_{o}(\ell), ~~ \frac{\lambda_c \ell}{2}>r c_{noo}, \]

which is possible because exponential growth wins over a rational one. Then, Theorem 23 directly follows from [65]. \( \blacksquare\)

Remark 58

Note that while similar results might be obtained via a small-gain methodology as in [169], we choose here to follow an explicit Lyapunov proof. With the high-gain designs in Example 22, an error \( z_o-\hat{z}_o\) appearing in \( \hat{f}_{no}\) would typically make \( c_{noo}\) in Theorem 23 depend on \( \ell\) , which is not allowed. Therefore, we may need to require \( f_{no}\) and \( \hat{f}_{no}\) to be independent of \( z_o\) and \( \hat{z}_o\) . This can always be done by taking \( z_{no}\) containing \( z_o\) , meaning that \( z_o\) is somehow estimated twice. This obstruction to handle coupling during flows with a high-gain design was similarly noticed in the linear output regulation context [143, Proposition 6]. Note also that with the existing high-gain designs in Example 22, \( \overline{b}_{o}\) does not depend on \( \ell\) (and the same for Theorem 24 below).

6.2.2.1.2 Conditions for Arbitrarily Fast Exponential Stability

Under certain conditions, as shown in Theorem 24 below, the coupling of an arbitrarily fast high-gain flow-based observer and an arbitrarily fast jump-based one that has an ISS property can actually result in arbitrarily fast exponential stability of the estimation error. For this, the jump-based observer should have another gain \( \gamma\) that is to be pushed closer to zero to speed up the convergence of \( V_{no,\gamma}\) , which now depends on this parameter.

Theorem 24 (Lyapunov conditions for arbitrarily fast exponential stability)

Suppose Assumption 38 holds and define \( \tau_M:=\max \mathcal{I}\) . Consider the cascade (408)-(409) and sets \( ¶_0,¶_c,¶_d\subseteq \mathbb{R}^{n_p}\) such that each solution \( (z,\hat{z},p,\tau)\) initialized in \( \mathcal{Z}_0\times \mathbb{R}^{n_z}\times ¶_0\times \{0\}\) with inputs in \( \mathfrak{U}_c\times \mathfrak{U}_d\) is such that \( p(t,j)\in ¶_c\) during flows and \( p(t,j)\in ¶_d\) at jumps. Assume there exist scalars \( \ell_0>0\) , \( 0<\gamma_0 \leq 0\) , \( \lambda_c>0\) , \( c_{no} \geq 0\) , and \( \lambda_d\geq 0\) , rational functions \( \underline{b}_o>0\) and \( \overline{b}_o>0\) , functions \( \underline{b}_{no}>0\) and \( \overline{b}_{no}>0\) , functions \( c_{noo}\geq 0\) and \( d_{ono} \geq 0\) , and functions \( d_o\geq 0\) and \( d_{noo}\geq 0\) rational in their first argument such that, for any \( \ell\geq\ell_0\) and for any \( 0<\gamma\leq\gamma_0\) , there exist functions \( V_{o,\ell}: \mathbb{R}^{n_o}\times \mathbb{R}^{n_o} \times \mathbb{R}^{n_p} \times \mathbb{R}\to \mathbb{R}\) and \( V_{no,\gamma}: \mathbb{R}^{n_{no}}\times \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_p} \times \mathbb{R}\to \mathbb{R}\) such that:

  1. (Uniform boundedness) For all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , \( z= (z_o,z_{no})\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) or \( (z,u_d)\in D_z\) , \( \hat{z}= (\hat{z}_o,\hat{z}_{no})\in \mathbb{R}^{n_z}\) , \( p\in ¶_c\cup¶_d\) , and \( \tau \in [0,\tau_M]\) ,

    \[ \begin{align} \underline{b}_{o}(\ell) |z_o-\hat{z}_o|^2 &\leq V_{o,\ell}(z_o,\hat{z}_o,p,\tau) \leq \overline{b}_{o}(\ell) |z_o - \hat{z}_o|^2,\\ \underline{b}_{no}(\gamma) |z_{no}-\hat{z}_{no}|^2 &\leq V_{no,\gamma}(z_{no},\hat{z}_{no},p,\tau) \leq \overline{b}_{no}(\gamma) |z_{no}-\hat{z}_{no}|^2; \\\end{align} \]

    (420.a)

  2. (Flow-based conditions) For all \( u_c \in \mathcal{U}_c\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c\) , and \( \tau \in [0,\tau_M]\) ,

    \[ \begin{align} \dot{V}_{o,\ell}(z,\hat{z},p,\tau,u_c) & \leq -\ell \lambda_c V_{o,\ell}(z_o,\hat{z}_o,p,\tau),\\ \dot{V}_{no,\gamma}(z,\hat{z},p,\tau,u_c) & \leq c_{no}V_{no,\gamma}(z_{no},\hat{z}_{no},p,\tau)+ c_{noo}(\gamma) V_{o,\ell}(z_o,\hat{z}_o,p,\tau); \\\end{align} \]

    (421.a)

  3. (Jump-based conditions) For all \( u_d \in \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_d\) , and \( \tau \in \mathcal{I}\) ,

    \[ \begin{align} V_{o,\ell}^+(z,\hat{z},p,\tau,u_d)&\leq d_o(\ell,\gamma) V_{o,\ell}(z_o,\hat{z}_o,p,\tau) + d_{ono}(\gamma)|z_{no}-\hat{z}_{no}|^2, \\ V_{no,\gamma}^+(z,\hat{z},p,\tau,u_d)&\leq \gamma e^{-\lambda_d} V_{no,\gamma}(z_{no},\hat{z}_{no},p,\tau) + d_{noo}(\ell,\gamma) V_{o,\ell}(z_o, \hat{z}_o,p,\tau). \\\end{align} \]

    (422.a)

Then, for any \( \lambda>0\) , there exists \( 0<\gamma^\star\leq\gamma_0\) such that there exists \( \ell^\star\geq\ell_0\) such that for any \( 0<\gamma<\gamma^\star\) and for any \( \ell>\ell^\star\) , there exist \( \rho>0\) such that for any solution \( (z,\hat{z},p,\tau)\) to the cascade (408)-(409) initialized in \( \mathcal{Z}_0\times \mathbb{R}^{n_z}\times ¶_0\times \{0\}\) with inputs in \( \mathfrak{U}_c\times \mathfrak{U}_d\) verifies (415).

Note that the dependence of the functions on \( \gamma\) is arbitrary.

Proof. Consider the following Lyapunov function \( W_{\ell,\gamma}: \mathbb{R}^{n_z}\times\mathbb{R}^{n_z}\times\mathbb{R}^{n_p}\times\mathbb{R}\to\mathbb{R}\) (with \( r > 0\) and \( \varepsilon > 0\) ):

\[ \begin{equation} W_{\ell,\gamma}(z,\hat{z},p,\tau) = e^{\frac{\lambda_c \ell \tau}{2}} V_{o,\ell}(z_o,\hat{z}_o,p,\tau) + r e^{-\varepsilon\tau} V_{no,\gamma}(z_{no},\hat{z}_{no},p,\tau). \end{equation} \]

(423)

First, we have for all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) or \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c \cup ¶_d\) , and \( \tau \in [0,\tau_M]\) ,

\[ \begin{equation} \underline{\rho}_{o}(\ell) |z_o-\hat{z}_{o}|^2 + \underline{\rho}_{no}(\gamma)|z_{no}-\hat{z}_{no}|^2 \leq W_{\ell,\gamma}(z,\hat{z},p,\tau) \leq \overline{\rho}_{o}(\ell) |z_o-\hat{z}_{o}|^2 + \overline{\rho}_{no}(\gamma)|z_{no}-\hat{z}_{no}|^2, \end{equation} \]

(424)

where \( \underline{\rho}_{o}(\ell) = \underline{b}_o(\ell)\) , \( \underline{\rho}_{no}(\gamma) = \underline{b}_{no}(\gamma)r e^{-\varepsilon\tau_M}\) , \( \overline{\rho}_{o}(\ell) = \overline{b}_o (\ell)e^{\frac{\lambda_c\ell\tau_M}{2}}\) , and \( \overline{\rho}_{no}(\gamma) = \overline{b}_{no}(\gamma)r\) . During flows, for all \( u_c\in \mathcal{U}_c\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_c)\in C_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_c\) , and \( \tau \in [0,\tau_M]\) ,

\[ \begin{equation} \dot W_{\ell,\gamma}(z,\hat{z},p,\tau,u_c) \leq - \min \left\{\frac{\lambda_c \ell}{2} -r c_{noo}(\gamma), \varepsilon - c_{no} \right\} W_{\ell,\gamma}(z,\hat{z},p,\tau). \end{equation} \]

(425)

At jumps, for all \( u_d \in \mathcal{U}_d\) , \( z\in \mathbb{R}^{n_z}\) such that \( (z,u_d)\in D_z\) , \( \hat{z}\in \mathbb{R}^{n_z}\) , \( p\in ¶_d\) , and \( \tau \in \mathcal{I}\) ,

\[ \begin{multline} W_{\ell,\gamma}^+(z,\hat{z},p,\tau,u_d) - W_{\ell,\gamma}(z,\hat{z},p,\tau) \leq -\left(e^{\frac{\lambda_c \ell\tau_m}{2}}\underline{b}_{o}(\ell) -d_o(\ell,\gamma)\overline{b}_{o}(\ell)- r d_{noo}(\ell,\gamma) \overline{b}_{o}(\ell)\right)|z_o-\hat{z}_{o}|^2 \\ - \left( r \overline{b}_{no}(\gamma)(e^{-\varepsilon\tau_M} - \gamma e^{-\lambda_d}) - d_{ono}(\gamma)\right) |z_{no}-\hat{z}_{no}|^2, \end{multline} \]

(426)

with \( \tau_m:=\min \mathcal{I} > 0\) . Then, given any \( \lambda>0\) , we choose successively

\[ \varepsilon > c_{no} + 2\lambda\left(\frac{1}{\tau_m}+1\right), ~~ 0< \gamma< e^{\lambda_d-\varepsilon\tau_m}, ~~ r > \frac{d_{ono}(\gamma)}{\overline{b}_{no}(\gamma)(e^{-\varepsilon\tau_M} - \gamma e^{-\lambda_d})}, \]

and finally \( \ell\) sufficiently large to have both

\[ e^{\frac{\lambda_c \ell\tau_m}{2}}\underline{b}_{o}(\ell) >d_o(\ell,\gamma)\overline{b}_{o}(\ell)+ r d_{noo}(\ell,\gamma) \overline{b}_{o}(\ell), ~~ \frac{\lambda_c \ell}{2}>r c_{noo}(\gamma) + 2\lambda\left(\frac{1}{\tau_m}+1\right), \]

which is possible because exponential growth wins over a rational one. We then obtain during flows and at jumps respectively,

\[ \begin{align*} \dot{W}_{\ell,\gamma}(z,\hat{z},p,\tau)& \leq -2 \lambda\left(\frac{1}{\tau_m}+1\right)W_{\ell,\gamma}(z,\hat{z},p,\tau),\\ W_{\ell,\gamma}^+(z,\hat{z},p,\tau) - W_{\ell,\gamma}(z,\hat{z},p,\tau) &\leq -a_{d,o} |z_o-\hat{z}_o|^2 - a_{d,no}|z_{no}-\hat{z}_{no}|^2, \end{align*} \]

for some \( a_{d,o} > 0\) and \( a_{d,no}>0\) . That implies that

\[ W_{\ell,\gamma}(z(t,j),\hat{z}(t,j),p(t,j),\tau(t,j)) \leq e^{-2\lambda\left(\frac{1}{\tau_m}+1\right) t}W_{\ell,\gamma}(z(0,0),\hat{z}(0,0),p(0,0),\tau(0,0)), \]

for all \( (t,j) \in \rm{dom} z = \rm{dom}\hat{z}\) . From Assumption 38, we have \( j \leq \frac{t}{\tau_m}+1\) or \( t \geq \frac{t+j-1}{\frac{1}{\tau_m} + 1}\) for all \( (t,j) \in \rm{dom} z = \rm{dom}\hat{z}\) . Therefore, we have

\[ W_{\ell,\gamma}(z(t,j),\hat{z}(t,j),p(t,j),\tau(t,j)) \leq e^{2\lambda} e^{-2\lambda(t+j)}W_{\ell,\gamma}(z(0,0),\hat{z}(0,0),p(0,0),\tau(0,0)), \]

for all \( (t,j) \in \rm{dom} z = \rm{dom}\hat{z}\) . Then Theorem 24 directly follows from [65]. \( \blacksquare\)

Remark 59

The conclusion of Theorem 24 differs from Theorem 23 in the sense that \( \lambda\) can be achieved arbitrarily, instead of being defined by the observer parameter(s). From the proof of Theorem 24, it is seen that \( \rho\) increases as \( \lambda\) increases, characterizing the peaking phenomenon typically encountered in high-gain designs when we push convergence arbitrarily fast. Discrete-time observers satisfying the conditions in Theorem 24 include [43] and our designs in Section 2.2 and Section 2.3, where the former requires linearity in the dynamics and output.

6.2.2.2 Proof of Theorem 16

First, exploiting exponential decrease over rational growth, given \( \overline{c}_o\) from Assumption 20 and using the compactness of \( \Xi_o\) , let \( \ell_2^\star\geq \ell_1^\star\) (defined in Item (XX) of Assumption 18) such that, for all \( \ell > \ell_2^\star\) ,

\[ \begin{equation} \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\leq \overline{c}_o, \end{equation} \]

(427)

where \( \overline{b}_o\) , \( \underline{b}_o\) , \( \lambda_c\) are defined in Item (XX) of Assumption 18, \( \tau_m:=\min \mathcal{I}>0\) , and \( M_o > 0\) is a bound of \( J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no})\) (uniform in \( u_d \in \mathcal{U}_d\) ) obtained from the definitions of \( \rm{sat}_{o}\) , \( \rm{sat}_{no}\) , and Assumption 19. The proof then consists of three main parts: i) Define new coordinates \( (z_{no},\hat{z}_{no})\) , replacing \( (\xi_{no},\hat{\xi}_{no})\) and obtain the state dynamics in those new coordinates, ii) Show that the Lyapunov conditions in Theorem 23 hold after jump \( j_m\) and obtain exponential stability of the estimation error in the new coordinates with respect to time \( (t_{j_m},j_m)\) , where \( j_m\) comes from Item (XIV) of Assumption 17, and iii) Recover the exponential stability in the \( \xi\) -coordinates with respect to the initial time.

Let us begin with the first part of this proof and consider the transformation \( (\xi_o,\xi_{no},\hat{\xi}_o,\hat{\xi}_{no},p,\tau,t,j)\mapsto (\xi_o,z_{no},\hat{\xi}_o,\hat{z}_{no},p,\tau)\) with

\[ \begin{align} z_{no} &= \Psi_{f_{no}}(\xi_{no},-\tau)- K_d \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau), \end{align} \]

(428.a)

\[ \begin{align} \hat{z}_{no} &= \Psi_{f_{no}}(\hat{\xi}_{no},-\tau) - K_d \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau), \\\end{align} \]

(428.b)

where

\[ \begin{equation} \Psi_{f_{no}}(\xi_{no},\tau) = e^{F_{no}\tau}\xi_{no}+ \int_{0}^{\tau}e^{F_{no}(\tau-s)}U_{cno}ds. \end{equation} \]

(428.c)

In the new coordinates, the dynamics of \( (\xi_o,\hat{\xi}_o,p,\tau)\) are obtained by replacing \( \xi_{no}\) and \( \hat{\xi}_{no}\) respectively with

\[ \begin{align} \xi_{no} &= \Psi_{f_{no}}(z_{no} + K_d\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o,t,-\tau),\tau), \\ \hat{\xi}_{no} &=\Psi_{f_{no}}(\hat{z}_{no} + K_d\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau),\tau), \end{align} \]

(429.a)

and considering the extended inputs \( \mathfrak{u}_{c,\rm{ ext }}(t)=(\mathfrak{u}_c(t),t)\in \mathcal{U}_{c,\rm{ ext }}\) and \( \mathfrak{u}_{d,\rm{ ext }}(j)=(\mathfrak{u}_d(j),t_{j+1})\in \mathcal{U}_{d,\rm{ ext }}\) , with \( \mathcal{U}_{c,\rm{ ext }}=\mathcal{U}_c\times \mathbb{R}_{\geq 0}\) and \( \mathcal{U}_{d,\rm{ ext }}=\mathcal{U}_d\times \mathbb{R}_{\geq 0}\) . Concerning the dynamics of \( z_{no}\) and \( \hat{z}_{no}\) , we start by showing that \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o,t,-\tau)\) is constant along solutions to the cascade (245)-(248) initialized in \( \Xi_0\times \mathbb{R}^{n_\xi}\times ¶_{0}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) . To do that, pick \( \ell > \ell^\star_2\) and pick a solution \( (\xi,\hat{\xi},p,\tau)\) to the cascade (245)-(248) initialized in \( \Xi_0\times \mathbb{R}^{n_\xi}\times ¶_{0}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) (which is complete thanks to Item (XIV) of Assumption 17 and given the dynamics of observer (248) which do not allow finite-time escape thanks to saturation functions and the local boundedness of the maps of system (245) as in Assumption 19). In the following, we refer to the jump times of this solution as \( t_j\) instead of \( t_j(\xi)\) to ease the notations. Since the solution component \( \xi_o\) flows according to \( f_o\) with input \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for each \( j\in \rm{dom}_j \xi\) and for each \( s\in [0,t-t_j]\) , we have

\[ \Psi_{f_{o}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-s) = \xi_o(t-s,j). \]

Since the trajectory \( t \mapsto \xi_o(t,j)\) remains in \( \Xi_o\) and \( \tau\) is initialized as \( \tau(0,0) = 0\) , we have for all \( (t,j) \in \rm{dom} \xi_o\) ,

\[ \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j))=\Psi_{f_{o}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)). \]

In addition, by the definition of the dynamics, \( \tau\) initialized as \( \tau(0,0) = 0\) is the time elapsed since the previous jump, namely \( \tau(t,j)=t-t_j\) for all \( j \in \rm{dom}_j \xi\) . Therefore, exploiting again that \( \xi_o\) evolves according to \( f_o\) with input \( \mathfrak{u}_c\) , we have \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)) = \Psi_{f_o(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-(t-t_j)) = \xi_o(t_j,j)\) for all \( j\in \rm{dom}_j \xi\) and \( t\in [t_j,t_{j+1}]\) . Hence, \( t \mapsto \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j))\) is constant during flow intervals. We also deduce that \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\) remains in the compact set \( \Xi_o\) at all times. Similarly, since \( \xi_{no}\) evolves during flows along with the vector field \( \xi_{no} \mapsto f_{no}(\xi_{no})=F_{no}\xi_{no}+U_{cno}\) , the quantity \( \Psi_{f_{no}}(\xi_{no},-\tau)\) is constant during flow intervals and remains in the compact set \( \Xi_{no}\) at all times. Then, the compact set

\[ \mathcal{Z}_{no} := \{z_{no} \in \mathbb{R}^{n_{no}}: \exists (\xi_o,\xi_{no}) \in \Xi_o \times \Xi_{no}, z_{no} = \xi_{no} - K_d\xi_o\}, \]

is such that along solutions to the cascade (245)-(248) initialized in \( \Xi_0\times \mathbb{R}^{n_\xi}\times ¶_{0}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , the image \( (t,j) \mapsto z_{no}(t,j)\) defined in (428.a) remains in \( \mathcal{Z}_{no}\) at all times. Solutions to the cascade (245)-(248) that are initialized in \( \Xi_0\times \mathbb{R}^{n_\xi}\times ¶_{0}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) are such that the variable \( z_{no}\) takes the dynamics (using that \( \tau^+=0\) and so \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t^+,-\tau^+) = \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t,0) = \xi_o^+\) ),

\[ \begin{align*} \dot{z}_{no} & = {}-F_{no} e^{-F_{no}\tau}\xi_{no} + e^{-F_{no}\tau}\dot{\xi}_{no} - e^{-F_{no}\tau}U_{cno}\\ &={}-F_{no} e^{-F_{no}\tau}\xi_{no} + e^{-F_{no}\tau}(F_{no}\xi_{no}+U_{cno})- e^{-F_{no}\tau}U_{cno}\\ &=0, \\ z_{no}^+ & = \xi_{no}^+ - K_d \xi_o^+\\ &=J_{no}(\xi_o, u_d)\xi_{no} + J_{noo}(\xi_o, u_d)-K_d(J_o(\xi_o, u_d) + J_{ono}(\xi_o,u_d)\xi_{no})\\ &=(J_{no}(\xi_o, u_d) - K_d J_{ono}(\xi_o,u_d))\xi_{no}+J_{noo}(\xi_o,u_d)-K_dJ_o(\xi_o,u_d)\\ &= \phi(\xi_o,u_d,\tau) \left(z_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)+ \int_{0}^{\tau}e^{-F_{no}s}U_{cno}ds\right)+J_{noo}(\xi_o,u_d)-K_dJ_o(\xi_o,u_d), \end{align*} \]

where \( \phi(\xi_o,u_d,\tau) = (J_{no}(\xi_o,u_d) - K_d J_{ono}(\xi_o,u_d))e^{F_{no}\tau}\) , and the variable \( \hat{z}_{no}\) takes the dynamics

\[ \begin{align} \dot{\hat{z}}_{no} & = {}-F_{no} e^{-F_{no}\tau}\hat{\xi}_{no} + e^{-F_{no}\tau}\dot{\hat{\xi}}_{no} -K_d\frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau) - e^{-F_{no}\tau}U_{cno}\\ &={}-F_{no} e^{-F_{no}\tau}\hat{\xi}_{no} + e^{-F_{no}\tau}(F_{no}\hat{\xi}_{no} + U_{cno}+e^{F_{no}\tau}K_d\frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\nonumber\\ &~~{} -K_d\frac{d}{dt}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau) - e^{-F_{no}\tau}U_{cno}\nonumber\\ & = 0, \\ \hat{z}_{no}^+ &=\hat{\xi}_{no}^+ - K_d \hat{\xi}_o^+\nonumber\\ &=J_{no}(\rm{sat}_o(\hat{\xi}_o), u_d)\hat{\xi}_{no} + J_{noo}(\rm{sat}_o(\hat{\xi}_o), u_d)+ L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (y_d - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)\nonumber\\ &~~{}-H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d)\hat{\xi}_{no}) - K_d J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d) (\hat{\xi}_{no} - \rm{sat}_{no}(\hat{\xi}_{no}))\nonumber\\ &~~{} - K_d(J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no})) \nonumber\\ &=(J_{no}(\rm{sat}_o(\hat{\xi}_o), u_d)-L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d)-K_dJ_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d))\hat{\xi}_{no}\nonumber\\ &~~{} + J_{noo}(\rm{sat}_o(\hat{\xi}_o), u_d) - K_d J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (y_d - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)\nonumber\\ & = \Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) \left(\hat{z}_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)+\int_{0}^{\tau}e^{-F_{no}s}U_{cno}ds\right)\nonumber\\ &~~{}+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)-K_dJ_o(\rm{sat}_o(\hat{\xi}_o),u_d)+ L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (y_d-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)), \\\end{align} \]

where \( \Phi\) is defined in Assumption 20. The flow and jump sets are subsets of \( \Xi_o\times \mathcal{Z}_{no}\times \mathcal{U}_c\) and \( \Xi_o\times \mathcal{Z}_{no}\times \mathcal{U}_d\) , respectively. Now, we deduce the estimation error dynamics. For brevity, let us denote

\[ \begin{align*} \Upsilon &= z_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)+\int_0^{\tau} e^{-F_{no}s}U_{cno}ds,\\ \hat{\Upsilon} &= \hat{z}_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)+\int_{0}^{\tau}e^{-F_{no}s}U_{cno}ds. \end{align*} \]

Then, we see that

\[ \begin{align*} y_d &= H_{do}(\xi_o,u_d) + H_{dno}(\xi_o,u_d)e^{F_{no}\tau}\Upsilon,\\ z_{no}^+ & = \phi(\xi_o,u_d,\tau) \Upsilon+J_{noo}(\xi_o,u_d)-K_dJ_o(\xi_o,u_d),\\ \hat{z}_{no}^+ & = \Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) \hat{\Upsilon}+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)-K_dJ_o(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &~~{}+ L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (H_{do}(\xi_o,u_d)-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d) + H_{dno}(\xi_o,u_d)e^{F_{no}\tau}\Upsilon). \end{align*} \]

Define the estimation error \( \tilde{z}_{no}:= z_{no} - \hat{z}_{no}\) . Then, we get

\[ \begin{multline*} \tilde{z}_{no}^+ = \phi(\xi_o,u_d,\tau) \Upsilon - \Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) \hat{\Upsilon} \\ - L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) H_{dno}(\xi_o,u_d)e^{F_{no}\tau}\Upsilon +(J_{noo}(\xi_o,u_d) - J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)) \\ -K_d(J_o(\xi_o,u_d) -J_o(\rm{sat}_o(\hat{\xi}_o),u_d)) -L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (H_{do}(\xi_o,u_d)-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)). \end{multline*} \]

Now, add and subtract both the terms \( \phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) \Upsilon\) and \( L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d)e^{F_{no}\tau}\Upsilon\) to get

\[ \begin{multline*} \tilde{z}_{no}^+ =\Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (\Upsilon - \hat{\Upsilon}) +(\phi(\xi_o,u_d,\tau)-\phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\\ - L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)(H_{dno}(\xi_o,u_d) - H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d))e^{F_{no}\tau}) \Upsilon\\ +(J_{noo}(\xi_o,u_d) - J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)) -K_d(J_o(\xi_o,u_d) -J_o(\rm{sat}_o(\hat{\xi}_o),u_d))\\ -L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (H_{do}(\xi_o,u_d)-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)). \end{multline*} \]

Now, see that \( \Upsilon - \hat{\Upsilon} = \tilde{z}_{no} + K_d(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\) and use the expression of \( \Upsilon\) . As a result, the estimation error \( \tilde{z}_{no}\) takes the dynamics

\[ \begin{align*} \dot{\tilde{z}}_{no} & = 0, \\ \tilde{z}_{no}^+ & ={}\Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\tilde{z}_{no} +\Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)K_d(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\\ &~~{} + (\phi(\xi_o,u_d,\tau) - \phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) - L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)(H_{dno}(\xi_o,u_d) - H_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d))e^{F_{no}\tau})\times\\ &~~{}\times\left(z_{no}+K_d\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)+\int_0^{\tau} e^{-F_{no}s}U_{cno}ds\right)\\ &~~{}+(J_{noo}(\xi_o,u_d)-J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d))-K_d(J_o(\xi_o,u_d)-J_o(\rm{sat}_o(\hat{\xi}_o),u_d))\\ &~~{}-L_d(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) (H_{do}(\xi_o,u_d)-H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)). \end{align*} \]

Let us move to the second part of this proof. First, from i) Item (XX) of Assumption 18, ii) the fact that \( t_{j+1} - t_j \geq \tau_m\) according to Item (XIV) of Assumption 17, and iii) the choice of \( \ell > \ell^\star_2\) satisfying (427), we have for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{align*} |\xi_o(t_{j+1},j) - \hat{\xi}_o(t_{j+1},j)|& \leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} |\xi_o(t_j,j) - \hat{\xi}_o(t_j,j)| e^{-\ell \frac{\lambda_c}{2}\tau_m} \\ &\leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\\ & \leq \overline{c}_o. \end{align*} \]

Then, by the definition of \( \rm{sat}_o\) and since \( \xi_o(t_{j+1},j)\in \Xi_o\cap D_o\) , we have, for all \( j\in \mathbb{N}_{\geq j_m}\) ,

\[ \rm{sat}_o(\hat{\xi}_o(t_{j+1},j)) = \hat{\xi}_o(t_{j+1},j). \]

Hence, the condition in Assumption 20 is satisfied with \( \rm{sat}_o(\hat{\xi}_o)\) replacing \( \xi_o\) , making \( \Phi(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\) Schur and so the dynamics of \( \tilde{z}_{no}\) are contracting (since the input \( u_d \in \mathfrak{U}_d\) takes values in \( \mathcal{U}_d\) ). Consider the Lyapunov function

\[ \begin{equation} V_{no}(z_{no},\hat{z}_{no}) = (z_{no}-\hat{z}_{no})^\top Q (z_{no}-\hat{z}_{no}), \end{equation} \]

(431)

with \( Q\) in (249). Denote \( \underline{c}_Q > 0\) and \( \overline{c}_Q > 0\) , respectively, as the minimum and maximum eigenvalues of \( Q\) . Then, we have for all \( (z_{no},\hat{z}_{no}) \in \mathcal{Z}_{no}\times\mathbb{R}^{n_{no}}\) ,

\[ \begin{equation} \underline{c}_Q |z_{no}-\hat{z}_{no}|^2 \leq V_{no}(z_{no},\hat{z}_{no}) \leq \overline{c}_Q |z_{no}-\hat{z}_{no}|^2. \end{equation} \]

(432)

For all \( (\xi_o,z_{no})\in (\Xi_o \cap C_o)\times \mathcal{Z}_{no}\) , \( (\hat{\xi}_o,\hat{z}_{no})\in \mathbb{R}^{n_o} \times \mathbb{R}^{n_{no}}\) and for all \( (\tau,t,u_c)\in [0,\tau_M]\times \mathbb{R}_{\geq 0}\times \mathcal{U}_c\) , we have

\[ \dot{V}_{no}(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no},\tau,t,u_c) = 0, \]

along the respective flow dynamics. By the global Lipschitzness of \( f_{o,\rm{sat}}\) with respect to \( \xi_o\) , uniformly with respect to \( u_c\) , Lemma 25 in Section 6.2.2.6 allows us to show that \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) . With \( \rm{sat}_o\) defined as in Theorem 16 and \( \rm{sat}_{no}\) defined in observer (248), there exist \( L_o > 0\) and \( L_{no} > 0\) such that

\[ \begin{align} |\xi_o - \rm{sat}_o(\hat{\xi}_o)| & \leq L_o |\xi_o - \hat{\xi}_o|, &\forall (\xi_o,\hat{\xi}_o) &\in (\Xi_o \cap D_o) \times \mathbb{R}^{n_o}, \end{align} \]

(433.a)

\[ \begin{align} |\xi_{no} - \rm{sat}_{no}(\hat{\xi}_{no})| &\leq L_{no} |\xi_{no} - \hat{\xi}_{no}|, & \forall (\xi_{no},\hat{\xi}_{no}) & \in (\Xi_{no} \cap D_{no}) \times \mathbb{R}^{n_{no}}. \end{align} \]

(433.b)

Indeed, if \( \hat{\xi}_o \in (\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}\) then \( \rm{sat}_o(\hat{\xi}_o) = \hat{\xi}_o\) and the property holds with \( L_o = 1\) ; on the other hand, if \( \hat{\xi}_o \notin (\Xi_o \cap D_o) + \overline{c}_o \mathbb{B}\) then \( |\xi_o - \hat{\xi}_o| \geq \overline{c}_o\) and thus \( |\xi_o - \rm{sat}_o(\hat{\xi}_o)| = |\rm{sat}_o(\xi_o) - \rm{sat}_o(\hat{\xi}_o)| \leq 2 M_o^\prime \leq \frac{2 M_o^\prime}{\overline{c}_o}|\xi_o - \hat{\xi}_o|\) , where \( M_o^\prime > 0\) is the bound of \( \rm{sat}_o\) ; therefore, take \( L_o = \max\left\{1,\frac{2 M_o^\prime}{\overline{c}_o}\right\}\) . The proof of (433.b) follows similarly. Thanks to Assumption 19Assumption 20, and Young’s inequality on the cross terms, there exist \( c_1\in[0,1)\) , \( c_2>0\) , \( c_3 > 0\) , and \( c_4>0\) such that for any \( \kappa>0\) , for all \( (\xi_o,z_{no})\in (\Xi_o\cap D_o)\times \mathcal{Z}_{no}\) , for all \( (\hat{\xi}_o,\hat{z}_{no})\in \mathbb{R}^{n_o} \times \mathbb{R}^{n_{no}}\) , and for all \( (\tau,u_c,u_d)\in \mathcal{I}\times \mathcal{U}_c\times\mathcal{U}_d\) , we have

\[ \begin{align*} V_{no}^+(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no},\tau,t,u_c,u_d)&\leq \left(c_1 + \frac{c_2}{\kappa}\right) V_{no}(z_{no},\hat{z}_{no})+ (c_3\kappa+c_4) |\xi_o - \hat{\xi}_o|^2 \\ & \leq\left(c_1 + \frac{c_2}{\kappa}\right) V_{no}(z_{no},\hat{z}_{no})+ \frac{c_3\kappa+c_4}{\underline{b}_o(\ell)}V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau), \end{align*} \]

along the jump dynamics of \( \tilde{z}_{no}\) , where the latter inequality is obtained from Item (XX) of Assumption 18. Pick \( \kappa\) large enough so that \( c_1 + \frac{c_2}{\kappa}\in[0,1)\) and so \( V_{no}\) satisfies the second item of each condition of Theorem 23 (note that \( (c_3\kappa+c_4/)/(\underline{b}_o(\ell)) \) is rational in \( \ell\) because so is \( \underline{b}_o(\ell)\) ). Using

\[ \xi_{no} - \hat{\xi}_{no} = e^{F_{no}\tau}(z_{no} - \hat{z}_{no}) + K_d(\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o,t,-\tau) -\Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)), \]

and since \( \tau\) remains in \( [0,\tau_M]\) and \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) , using Young’s inequality, we deduce that there exist \( c_5>0\) , \( c_6>0\) , \( c_7 > 0\) , and \( c_8 > 0\) such that for all \( (\xi_o,z_{no},\hat{\xi}_o,\hat{z}_{no},\tau,u_c,t)\in \Xi_o\times \mathcal{Z}_{no}\times\mathbb{R}^{n_o}\times \mathbb{R}^{n_{no}}\times [0,\tau_M]\times \mathcal{U}_c\times \mathbb{R}_{\geq 0}\) , \( (\xi_{no},\hat{\xi}_{no})\) defined in (429) verifies

\[ \begin{equation} |\xi_{no}-\hat{\xi}_{no}|\leq c_5 |\xi_o-\hat{\xi}_o| + c_6 |z_{no}-\hat{z}_{no}|, ~~ |z_{no}-\hat{z}_{no}|\leq c_7 |\xi_o-\hat{\xi}_o| + c_8 |\xi_{no}-\hat{\xi}_{no}|. \end{equation} \]

(434)

Then, there exist \( c_9 > 0\) and \( c_{10} > 0\) such that

\[ \begin{equation} |\xi - \hat{\xi}| \leq c_9 |(\xi_o,z_{no}) - (\hat{\xi}_o,\hat{z}_{no})|, ~~ |(\xi_o,z_{no}) - (\hat{\xi}_o,\hat{z}_{no})| \leq c_{10} |\xi - \hat{\xi}|. \end{equation} \]

(435)

On the other hand, we have

\[ \begin{align} \xi_o^+ - \hat{\xi}_o^+ &= J_o(\xi_o, u_d) - J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\xi_o,u_d)\xi_{no} - J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no})\notag\\ & = J_o(\xi_o, u_d) - J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + (J_{ono}(\xi_o,u_d)- J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d))\xi_{no} \notag\\ &~~{}+ J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d) (\xi_{no} - \rm{sat}_{no}(\hat{\xi}_{no})). \end{align} \]

(436)

Because \( \xi_{no}\) is bounded and thanks to (433.b), there exist \( c_{11}>0\) and \( c_{12}>0\) and from these, \( d_o(\ell):=(\overline{b}_o c_{11})/(\underline{b}_o(\ell)) > 0 \) (rational in \( \ell\) because so is \( \underline{b}_o\) ) and \( d_{ono}:=\overline{b}_oc_{12} > 0\) , such that \( V_{o,\ell}\) in Item (XX) of Assumption 18 satisfies

\[ \begin{align*} V_{o,\ell}^+(\xi,\hat{\xi},p,\tau,u_d)&\leq\overline{b}_o|\xi_o^+ - \hat{\xi}_o^+|^2\\ &\leq\overline{b}_o c_{11} |\xi_o - \hat{\xi}_o|^2 + \overline{b}_oc_{12} |\xi_{no} - \hat{\xi}_{no}|^2\\ & \leq \frac{\overline{b}_o c_{11}}{\underline{b}_o(\ell)} V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) + \overline{b}_oc_{12}|\xi_{no}-\hat{\xi}_{no}|^2\\ & \leq d_o(\ell)V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) + d_{ono}|\xi_{no}-\hat{\xi}_{no}|^2, \end{align*} \]

where \( V_{o,\ell}^+\) denotes the jump of \( V_{o,\ell}\) along

\[ (J_o(\xi_o, u_d)+ J_{ono}(\xi_o,u_d)\xi_{no},J_o(\rm{sat}_o(\hat{\xi}_o),u_d)+J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no}))), \]

according to Item (XX) of Assumption 18. This inequality, along with Item (XX) of Assumption 18, shows that \( V_{o,\ell}\) satisfies the inequalities involving it in Theorem 23 in the new coordinates. Since \( \varphi_{c,\ell}\) and \( \varphi_{d,\ell}\) are independent of \( (\hat{\xi}_o,\hat{z}_{no})\) , for any \( \ell > \ell^\star_1\) , solutions \( (\xi_o,z_{no},\hat{\xi}_o,\hat{z}_{no},p,\tau)\) to the cascade (245)-(248) initialized in \( \Xi_0\times \mathbb{R}^{n_\xi}\times ¶_{0}\times \{0\}\) with inputs \( (u_{c,\rm{ ext }},u_{d,\rm{ ext }})\in\mathfrak{U}_{c,\rm{ ext }}\times \mathfrak{U}_{d,\rm{ ext }}\) are such that \( p(t,j) \in ¶_{c}\) during flows and \( p(t,j) \in ¶_{d}\) at jumps. Applying Theorem 23 starting from hybrid time \( (t_{j_m},j_m)\) with \( z_o = \xi_o\) and with \( z_{no}\) defined above, we deduce that there exists \( \ell^\star_3 \geq \ell^\star_2\) such that for any \( \ell > \ell^\star_3\) , there exist \( \rho_1(\ell) > 0\) and \( \lambda(\ell) > 0\) such that

\[ \begin{multline} |(\xi_o,z_{no})(t,j) - (\hat{\xi}_o,\hat{z}_{no})(t,j)| \leq \rho_1(\ell) |(\xi_o,z_{no})(t_{j_m},j_m) - (\hat{\xi}_o,\hat{z}_{no})(t_{j_m},j_m)| e^{-\lambda(\ell)(t+j)},\\ \forall (t,j) \in \rm{dom} \xi: j \geq j_m. \end{multline} \]

(437)

Now, we move to the last part of this proof, providing an exponential decreasing bound of the estimation error \( (\xi_o,z_{no})(t_{j_m},j_m) - (\hat{\xi}_o,\hat{z}_{no})(t_{j_m},j_m)\) compared to its value at time \( (0,0)\) . Consider the Lyapunov function

\[ \begin{equation} W(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no}) = V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) + V_{no}(z_{no},\hat{z}_{no}), \end{equation} \]

(438)

which verifies for all \( (u_c,u_d) \in \mathcal{U}_c \times \mathcal{U}_d\) , for all \( \xi= (\xi_o,\xi_{no})\in \mathbb{R}^{n_\xi}\) such that \( (\xi,u_c)\in C\) or \( (\xi,u_d)\in D\) , for all \( \hat{\xi}= (\hat{\xi}_o,\hat{\xi}_{no})\in \mathbb{R}^{n_\xi}\) , for all \( p\in ¶_{c}\cup¶_{d}\) , for all \( \tau \in [0,\tau_M]\) , and for all \( (z_{no},\hat{z}_{no}) \in \mathcal{Z}_{no}\times\mathbb{R}^{n_{no}}\) ,

\[ \underline{b}_o(\ell) |\xi_o - \hat{\xi}_o|^2 + \underline{c}_Q |z_{no} - \hat{z}_{no}|^2 \leq W(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no}) \leq \overline{b}_o |\xi_o - \hat{\xi}_o|^2 + \overline{c}_Q |z_{no} - \hat{z}_{no}|^2. \]

For any \( \ell > \ell^\star_1\) , for all \( (\xi_o,z_{no})\in (\Xi_o \cap C_o)\times \mathcal{Z}_{no}\) , for all \( (\hat{\xi}_o,\hat{z}_{no})\in \mathbb{R}^{n_o} \times \mathbb{R}^{n_{no}}\) , and for all \( (\tau,t,u_c)\in [0,\tau_M]\times \mathbb{R}_{\geq 0}\times \mathcal{U}_c\) , we have

\[ \begin{align*} \dot{W}(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no},p,\tau,t,u_c) & = -\ell \lambda_c V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau)\leq 0, \end{align*} \]

along the respective flow dynamics. Thanks to boundedness in solutions and the observer jump map brought by Assumption 19 and the saturation functions, there exists \( c_{13}(\ell) > 0\) such that before jump \( j_m\) , for all \( (\xi_o,z_{no})\in (\Xi_o\cap D_o)\times \mathcal{Z}_{no}\) , for all \( (\hat{\xi}_o,\hat{z}_{no})\in \mathbb{R}^{n_o} \times \mathbb{R}^{n_{no}}\) , and for all \( (\tau,u_d)\in [0,\tau_M]\times \mathcal{U}_d\) ,

\[ W^+(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no},\tau,t,u_d) \leq c_{13} (\ell)W(\xi_o,\hat{\xi}_o,z_{no},\hat{z}_{no}), \]

along the respective jump dynamics. So there exists \( c_{14}(\ell) > 0\) such that

\[ W(\xi_o(t_{j_m},j_m),\hat{\xi}_o(t_{j_m},j_m),z_{no}(t_{j_m},j_m),\hat{z}_{no}(t_{j_m},j_m)) \leq c_{14}(\ell) W(\xi_o(0,0),\hat{\xi}_o(0,0),z_{no}(0,0),\hat{z}_{no}(0,0)). \]

Consequently, there exists \( c_{15}(\ell) > 0\) such that

\[ \begin{equation} |(\xi_o,z_{no})(t_{j_m},j_m) - (\hat{\xi}_o,\hat{z}_{no})(t_{j_m},j_m)| \leq c_{15}(\ell)|(\xi_o,z_{no})(0,0) - (\hat{\xi}_o,\hat{z}_{no})(0,0)|. \end{equation} \]

(439)

From (435), (437), and (439), we deduce that for all \( (t,j) \in \rm{dom} \xi\) such that \( j \geq j_m\) ,

\[ \begin{align*} |\xi(t,j) - \hat{\xi}(t,j)| & \leq c_9 |(\xi_o,z_{no})(t,j) - (\hat{\xi}_o,\hat{z}_{no})(t,j)|\\ & \leq c_9 \rho_1 (\ell)|(\xi_o,z_{no})(t_{j_m},j_m) - (\hat{\xi}_o,\hat{z}_{no})(t_{j_m},j_m)| e^{-\lambda(\ell)(t+j)}\\ & \leq c_9 \rho_1(\ell) c_{15}(\ell) |(\xi_o,z_{no})(0,0) - (\hat{\xi}_o,\hat{z}_{no})(0,0)| e^{-\lambda(\ell)(t+j)}\\ & \leq c_9 \rho_1(\ell) c_{15}(\ell)c_{10} |\xi(0,0) - \hat{\xi}(0,0)| e^{-\lambda(\ell)(t+j)}, \end{align*} \]

and so Theorem 16 follows.

6.2.2.3 Proof of Theorem 17

This proof consists of three main parts: i) Choose \( \overline{c}_o\) (for the definition of \( \rm{sat}_o\) ), \( \gamma^\star\) , \( \ell^\star\) , and \( \underline{c}_{T}(\gamma)\) allowing us to guarantee some preliminary bounds and injectivity properties on the maps and variables, ii) Define a change of coordinates for the \( \xi_{no}\) state into some target \( \eta\) -coordinates, and obtain the state dynamics in those new coordinates (along solutions), and iii) Define a Lyapunov function and apply Theorem 24 to obtain exponential stability in the new coordinates after a certain time and retrieve asymptotic convergence in the \( \xi\) -coordinates.

Let us begin with the first preliminary part of this proof. We start by choosing \( \overline{c}_o\) . Along every complete solution \( \xi \in §_\mathcal{H}(\Xi_0, \mathfrak{U}_c \times \mathfrak{U}_d)\) , according to Item (XIV) of Assumption 17 and Item (XXI) of Assumption 21, denoting \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1}-t_j\) , we have \( \xi_{o,j}\in \Xi_o\) for all \( j \in \mathbb{N}\) and

\[ \|(\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1}\| = \sqrt{\lambda_{\max}(((\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1})^\top (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1})} \leq c_{\mathcal{J}^{-1}_{no}}, \]

for all \( j \in \mathbb{N}_{\geq j_m}\) . It follows that \( ((\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1})^\top (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^{-1}\leq c_{\mathcal{J}^{-1}_{no}}^2 \rm{Id}\) and thus we have \( (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))^\top \mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j)\geq \frac{1}{c_{\mathcal{J}^{-1}_{no}}^2} \rm{Id}\) for all \( j \in \mathbb{N}_{\geq j_m}\) . Applying Lemma 30 in Section 6.2.2.6 after jump \( j_m\) with \( M\) therein being \( \mathcal{J}_{no}\) , the sequences \( (\xi_j)_{j \in \mathbb{N}}\) and \( (u_j)_{j \in \mathbb{N}}\) therein being \( (\xi_{o,j})_{j \in \mathbb{N}_{\geq j_m}}\) and \( (u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) respectively, the scalar \( c\) therein being \( \frac{1}{c_{\mathcal{J}^{-1}_{no}}^2}\) , and the scalar \( c^\prime\) therein being \( \underline{c}_{\hat{\mathcal{J}}_{no}}^2\) for the chosen \( \underline{c}_{\hat{\mathcal{J}}_{no}} < \frac{1}{c_{\mathcal{J}^{-1}_{no}}}\) in the definition of \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}\) , we deduce that there exists \( \overline{c}_{o,1} > 0\) such that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_{o,1}\) for all \( j\in\mathbb{N}_{\geq j_m}\) , then \( (\mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j))^\top \mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j)\geq \underline{c}_{\hat{\mathcal{J}}_{no}}^2 \rm{Id}\) for all \( j \in \mathbb{N}_{\geq j_m}\) , which implies that \( (\mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j))_{j \in \mathbb{N}_{\geq j_m}}\) is uniformly invertible and is such that \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j)) = (\mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j))^{-1}\) for all \( j \in \mathbb{N}_{\geq j_m}\) .
Next, from Assumption 19 and the compactness of \( \mathcal{I}\) , the maps \( \mathcal{J}_{no}\) , \( \mathcal{H}_{dno}\) , and \( \mathcal{J}_{ono}\) are locally Lipschitz with respect to \( \xi_o\) , uniformly in \( (u_d,\tau) \in \mathcal{U}_d \times\mathcal{I}\) , and so are the maps \( \mathcal{O}^{bw}_j\) defined in (261) for all \( j \in \mathbb{N}_{\geq j_m + \overline{m}}\) . Applying Lemma 30 in Section 6.2.2.6 after jump \( j_m + \overline{m}\) with \( M\) therein being \( \mathcal{O}^{bw}_j\) , the sequence \( (\xi_j)_{j \in \mathbb{N}}\) therein being \( (\xi_{o,j-1},\xi_{o,j-2},…,\xi_{o,j-\overline{m}})_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) , the sequence \( (u_j)_{j \in \mathbb{N}}\) therein being \( (u_{d,j-1},u_{d,j-2},…,u_{d,j-\overline{m}},\tau_{j-1},\tau_{j-2},…,\tau_{j-\overline{m}})_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) , the scalar \( c\) therein being \( \alpha\) , and the scalar \( c^\prime\) therein being any \( 0 < \alpha^\prime < \alpha\) , we deduce that there exists \( \overline{c}_{o,2} > 0\) such that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_{o,2}\) for all \( j\in\mathbb{N}_{\geq j_m + \overline{m}}\) , then \( (\hat{\mathcal{O}}^{bw}_j)_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) defined as in (261) but fed with \( (\hat{\xi}_{o,j},u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) is such that each \( \hat{\mathcal{O}}^{bw}_j\) has full rank and satisfies \( (\hat{\mathcal{O}}^{bw}_j)^\top \hat{\mathcal{O}}^{bw}_j \geq \alpha^\prime \rm{Id} > 0\) for all \( j \in \mathbb{N}_{\geq j_m + \overline{m}}\) .
Pick \( \overline{c}_o = \min\{\overline{c}_{o,1},\overline{c}_{o,2}\}\) . It follows that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_o\) for all \( j\in\mathbb{N}_{\geq j_m}\) , then Item (XXI) of Assumption 21 holds after jump \( j_m\) for \( (\mathcal{J}_{no}(\hat{\xi}_{o,j},u_{d,j},\tau_j))_{j \in \mathbb{N}_{\geq j_m}}\) in place of \( (\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))_{j \in \mathbb{N}_{\geq j_m}}\) with \( 1/(\underline{c}_{\hat{\mathcal{J}}_{no}})\) replacing \( c_{\mathcal{J}^{-1}_{no}}\) , and Item (XXII) of Assumption 21 holds after jump \( j_m + \overline{m}\) for system (247) fed with \( (\hat{\xi}_{o,j},u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) in place of \( (\xi_{o,j},u_{d,j},\tau_j)_{j \in \mathbb{N}_{\geq j_m + \overline{m}}}\) with the same \( m_i\) , \( i \in \{1,2,…,n_{d,\text{ext}}\}\) , and with \( \alpha^\prime\) replacing \( \alpha\) .
Now fix \( T_0\in \mathbb{R}^{n_\xi \times n_{no}}\) . We next choose a first upper bound \( \gamma_1^\star\) for \( \gamma\) and \( \underline{c}_{T}(\gamma)\) . Thanks to Item (XIV) of Assumption 17 and Assumption 19, there exists \( \overline{c}_{T,m} > 0\) such that along any maximal solution \( (\xi,\hat{\xi}_o,\hat{\eta}, p, \hat{T}, \tau)\) to the cascade (245)-(264) initialized in \( \Xi_0\times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{T_0\}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , with \( \overline{c}_o\) in \( \rm{sat}_o\) fixed in the previous step, with any \( \gamma \in (0,1]\) , any \( \ell > \ell_1^\star\) , and any \( \underline{c}_{T}(\gamma)\) , any solutions \( (T_j)_{j\in \mathbb{N}}\) and \( (\hat{T}_j)_{j\in \mathbb{N}}\) to

\[ \begin{align} T_{j+1}&=(\gamma A T_j + B_{dno}\mathcal{H}_{dno}(\xi_{o,j},u_{d,j},\tau_j) + B_{ono}\mathcal{J}_{ono}(\xi_{o,j},u_{d,j},\tau_j) ) \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j)), \\ \hat{T}_{j+1} &=(\gamma A \hat{T}_j + B_{dno}\mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_{o,j}), u_{d,j}, \tau_j)\notag\\ &~~{} + B_{ono}\mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_{o,j}), u_{d,j}, \tau_j) )\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_{o,j}), u_{d,j}, \tau_j)), \\\end{align} \]

(440.a)

both initialized as \( T_0\) , where \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , \( \tau_j = \tau(t_{j+1},j)\) , and \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , are such that

\[ \begin{equation} \|T_j\|\leq \overline{c}_{T,m}, ~~ \|\hat{T}_j\|\leq \overline{c}_{T,m},~~ \forall j \in \mathbb{N}_{\leq j_m}. \end{equation} \]

(441)

Now, let us study what happens after jump \( j_m\) . First, observe that, if \( |\xi_{o}(t_{j+1},j) - \hat{\xi}_{o}(t_{j+1},j)| \leq \overline{c}_o\) for all \( j \in \mathbb{N}_{\geq j_m}\) , \( (T_j)_{j\in \mathbb{N}}\) and \( (\hat{T}_j)_{j\in \mathbb{N}}\) are solution to (258) after jump \( j_m\) , fed respectively with \( (\xi_{o,j}, u_{d,j}, \tau_j)_{j \in \mathbb{N}}\) and \( (\rm{sat}_o(\hat{\xi}_{o,j}), u_{d,j}, \tau_j)_{j \in \mathbb{N}}\) . From Assumption 19, since \( \tau_j \in \mathcal{I}\) after jump \( j_m\) , there exist \( c_{\mathcal{H}_{dno}} > 0\) and \( c_{\mathcal{J}_{ono}} > 0\) such that all these matrix sequences are uniformly bounded for all \( j \in \mathbb{N}_{\geq j_m}\) :

\[ \begin{align*} \|\mathcal{H}_{dno}(\xi_{o,j},u_{d,j},\tau_j)\|\leq c_{\mathcal{H}_{dno}}, \\ \|\mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_{o,j}),u_{d,j},\tau_j)\|\leq c_{\mathcal{H}_{dno}},\\ \|J_{ono}(\xi_{o,j},u_{d,j},\tau_j)\|\leq c_{\mathcal{J}_{ono}}, \\ \|J_{ono}(\rm{sat}_o(\hat{\xi}_{o,j}),u_{d,j},\tau_j)\|\leq c_{\mathcal{J}_{ono}}. \end{align*} \]

Then, according to Lemma 13 starting from \( j_m\) (i.e., with \( \overline{c}_{T,m}\) playing the role of \( \overline{c}_{T,0}\) ), for any \( j_m^\star \in \mathbb{N}_{\geq j_m + \overline{m} + 1}\) , there exists \( 0<\gamma^\star_1\leq 1\) such that for all \( 0<\gamma < \gamma^\star_1\) , there exist \( \underline{c}_{T}(\gamma)>0\) and \( \overline{c}_T(\gamma)>0\) such that for any maximal solution \( (\xi,\hat{\xi}_o,\hat{\eta}, p, \hat{T}, \tau)\) to the cascade (245)-(264) initialized in \( \Xi_0\times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{T_0\}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) , and verifying \( |\xi_{o}(t_{j+1},j) - \hat{\xi}_{o}(t_{j+1},j)| \leq \overline{c}_o\) for all \( j \in \mathbb{N}_{\geq j_m}\) , the solutions \( (T_j)_{j\in \mathbb{N}}\) and \( (\hat{T}_j)_{j\in \mathbb{N}}\) are both uniformly left-invertible for all \( j \in \mathbb{N}_{\geq j_m^\star}\) and uniformly bounded for all \( j \in \mathbb{N}_{\geq j_m}\) , i.e.,

\[ \begin{equation} \begin{array}{@{}l@{}} T_j^\top T_j \geq (\underline{c}_{T}(\gamma))^2 \rm{Id}, ~~ \forall j \in \mathbb{N}_{\geq j_m^\star}, ~~ \|T_j\| \leq \overline{c}_T(\gamma), ~~ \forall j \in \mathbb{N}_{\geq j_m}, \\ \hat{T}_j^\top \hat{T}_j \geq (\underline{c}_{T}(\gamma))^2 \rm{Id}, ~~ \forall j \in \mathbb{N}_{\geq j_m^\star}, ~~ \|\hat{T}_j\| \leq \overline{c}_T(\gamma), ~~ \forall j \in \mathbb{N}_{\geq j_m}. \end{array} \end{equation} \]

(442)

Finally, we pick a lower bound \( \ell_2^\star\) for \( \ell\) . Exploiting exponential decrease over rational growth, given \( \overline{c}_o\) picked above and using the compactness of \( \Xi_o\) , let \( \ell_2^\star\geq \ell_1^\star\) (defined in Item (XX) of Assumption 18) such that, for all \( \ell >\ell_2^\star\) ,

\[ \begin{equation} \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\leq \overline{c}_o, \end{equation} \]

(443)

where \( \overline{b}_o\) , \( \underline{b}_o\) , \( \lambda_c\) are defined in Item (XX) of Assumption 18, \( \tau_m:=\min \mathcal{I}>0\) , and \( M_o > 0\) is a bound of \( J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no})\) (uniform in \( u_d \in \mathcal{U}_d\) ) obtained from the definitions of \( \rm{sat}_o\) , \( \rm{sat}_{no}\) , and Assumption 19.
Now, pick \( 0<\gamma<\gamma^\star_1\) and \( \ell > \ell^\star_2\) . Consider a solution \( (\xi,\hat{\xi}_o,\hat{\eta}, p, \hat{T}, \tau)\) to the cascade (245)-(264) initialized in \( \Xi_0\times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{T_0\}\times \{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c \times \mathfrak{U}_d\) (which is complete thanks to Item (XIV) of Assumption 17 and given the dynamics of observer (264) which do not allow finite-time escape thanks to saturation functions and the local boundedness of the maps of system (245) as in Assumption 19), with the chosen \( (A,B_{dno},B_{ono})\) , the saturation maps \( \rm{sat}_o\) , \( \rm{sat}_{no}\) , and the inverse maps \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}\) , and \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) . In the following, we refer to the jump times of this solution as \( t_j\) instead of \( t_j(\xi)\) to ease the notations. First, from i) Item (XX) of Assumption 18, ii) the fact that \( t_{j+1} - t_j \geq \tau_m\) according to Item (XIV) of Assumption 17, and iii) the choice of \( \ell > \ell^\star_2\) satisfying (443), we have for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{align*} |\xi_o(t_{j+1},j) - \hat{\xi}_o(t_{j+1},j)|& \leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} |\xi_o(t_j,j) - \hat{\xi}_o(t_j,j)| e^{-\ell \frac{\lambda_c}{2}\tau_m} \\ &\leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\\ & \leq \overline{c}_o. \end{align*} \]

Then, by the definition of \( \rm{sat}_o\) and since \( \xi_o(t_{j+1},j)\in \Xi_o\cap D_o\) , we have, for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \rm{sat}_o(\hat{\xi}_o(t_{j+1},j)) = \hat{\xi}_o(t_{j+1},j). \]

Over the solution’s time domain, we finally introduce the hybrid arc \( T\) with the same dimension as \( \hat{T}\) and initialized as \( T(0,0) = T_0\) , with dynamics \( \dot{T} = 0\) during flows and at jumps given by

\[ \begin{equation} T^+ =(\gamma A T + B_{dno}\mathcal{H}_{dno}(\xi_o,u_d,\tau) + B_{ono}\mathcal{J}_{ono}(\xi_o,u_d,\tau))\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\xi_o,u_d,\tau)). \end{equation} \]

(444)

Since \( \dot{T} = 0\) and \( \dot{\hat{T}} = 0\) during flows, \( (T(t_j,j))_{j\in \mathbb{N}}\) and \( (\hat{T}(t_j,j))_{j\in \mathbb{N}}\) coincide for all \( j\in \mathbb{N}\) respectively with \( (T_j)_{j\in \mathbb{N}}\) and \( (\hat{T}_j)_{j\in \mathbb{N}}\) studied above, with \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1} - t_j =\tau(t_{j+1},j)\) for all \( j \in \mathbb{N}_{\geq j_m}\) . Therefore, since \( 0<\gamma<\gamma_1^\star\) and \( |\xi_o(t_{j+1},j) - \hat{\xi}_o(t_{j+1},j)|\leq \overline{c}_o\) for all \( j\in \mathbb{N}_{\geq j_m}\) , we deduce from above that

\[ \begin{align} (T(t,j))^\top T(t,j) &\geq (\underline{c}_{T}(\gamma))^2 \rm{Id}, & \forall (t,j) \in \rm{dom} \xi_o: j\geq j_m^\star,\\ (\hat{T}(t,j))^\top \hat{T}(t,j) &\geq (\underline{c}_{T}(\gamma))^2 \rm{Id}, & \forall (t,j) \in \rm{dom} \xi_o: j\geq j_m^\star. \\\end{align} \]

(445.a)

It follows that for all \( (t,j) \in \rm{dom} \xi_o\) with \( j\geq j_m^\star\) , the map \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) defined as in (264.c) is such that \( \rm{inv}_\underline{{c}_{T}(\gamma)}(T(t,j)) = (T(t,j))^\dagger\) and \( \rm{inv}_\underline{{c}_{T}(\gamma)}(\hat{T}(t,j)) = (\hat{T}(t,j))^\dagger\) , which are left inverses of \( T(t,j)\) and \( \hat{T}(t,j)\) , respectively. Note that we also recover from above that for all \( (t,j)\in \rm{dom} \xi_o\) , \( \|T(t,j)\|\) and \( \|\hat{T}(t,j)\|\) are upper-bounded by \( \overline{c}_{T,m}\) if \( j\leq j_m\) and by \( \overline{c}_T(\gamma)\) if \( j\geq j_m\) .

Now, we go to the second part of this proof. To exploit the results in Section 3.3.3.2.1, we define, over the time domain of the considered solution, the hybrid arc \( \eta:\rm{dom} \xi \to \mathbb{R}^{n_\eta}\) as

\[ \begin{equation} \eta(t,j) = T(t,j) \Psi_{f_{no}} (\xi_{no}(t,j),-\tau(t,j)) - B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)), \end{equation} \]

(446)

where \( \Psi_{f_{no}} (\cdot,\tau)\) is defined as in (428.c). Arguing in a similar way as in the proof of Theorem 16 above, we notice that

\[ \begin{align*} \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)) &= \Psi_{f_o(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-(t-t_j)) = \xi_o(t_j,j)\in \Xi_o, \\ \Psi_{f_{no}} (\xi_{no}(t,j),-\tau(t,j))&=\xi_{no}(t_j,j)\in \Xi_{no}, \end{align*} \]

for all \( j\in \rm{dom}_j \xi\) and \( t\in [t_j,t_{j+1}]\) . We deduce that \( \eta\) remains at all times in the compact set

\[ \begin{equation} \Xi_\eta(\gamma):=\{\eta \in \mathbb{R}^{n_\eta}:\exists (\xi_o,T,\xi_{no}) \in \Xi_o \times ¶_{T}(\gamma) \times \Xi_{no}, \eta = T \xi_{no} - B_{ono}\xi_o \}, \end{equation} \]

(447)

with \( ¶_{T}(\gamma) := \{T \in \mathbb{R}^{n_\eta \times n_{no}}:\|T\| \leq \max\{\overline{c}_{T,m},\overline{c}_T(\gamma)\}\}\) where \( \overline{c}_{T,m}\) is in (441) and \( \overline{c}_T(\gamma)\) is in (442). We deduce also that \( \eta\) verifies \( \dot{\eta} = 0\) during flows. From Item (XIV) of Assumption 17Item (XXI) of Assumption 21, and the choice of \( \underline{c}_{\hat{\mathcal{J}}_{no}}\) , we have after time \( (t_{j_m},j_m)\) , \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\xi_{o,j},u_{d,j},\tau_j))=(\mathcal{J}_{no}(\xi_o,u_d,\tau))^{-1}\) so that at jumps (using that \( \tau^+=0\) and so \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t^+,-\tau^+) = \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t,0) = \xi_o^+\) ),

\[ \begin{align*} \eta^+ &= T^+ \xi_{no}^+ - B_{ono}\xi_o^+\\ &= T^+\left(\mathcal{J}_{no}(\xi_o,u_d,\tau)e^{-F_{no}\tau}\xi_{no} + J_{noo}(\xi_o,u_d)\right) - B_{ono}\left(J_o(\xi_o,u_d) + J_{ono}(\xi_o,u_d)\xi_{no}\right)\\ &= \left(\gamma A T + B_{dno}\mathcal{H}_{dno}(\xi_o,u_d,\tau) + B_{ono}\mathcal{J}_{ono}(\xi_o,u_d,\tau)\right) e^{-F_{no}\tau}\xi_{no} +T^+J_{noo}(\xi_o,u_d) \\ &~~{} - B_{ono}\left(J_o(\xi_o,u_d) + J_{ono}(\xi_o,u_d)\xi_{no}\right)\\ &= \gamma A \eta + \gamma A B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau) + \gamma A T \int_0^{\tau}e^{-F_{no}s}U_{cno}ds+ B_{dno}H_{dno}(\xi_o,u_d)\xi_{no} \\ &~~{}+T^+J_{noo}(\xi_o,u_d) - B_{ono}J_o(\xi_o,u_d)\\ &= \gamma A \eta + \gamma A B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau) + \gamma A T \int_0^{\tau}e^{-F_{no}s}U_{cno}ds+ B_{dno}(y_d - H_{do}(\xi_o,u_d)) \\ &~~{}+T^+J_{noo}(\xi_o,u_d) - B_{ono}J_o(\xi_o,u_d). \end{align*} \]

On the other hand, from (446), (445) and the expression of \( \Psi_{f_{no}}\) , we can express \( \xi_{no}\) as a function of \( (\xi_o,\eta,T,\tau,t)\) and input \( \mathfrak{u}_c\) after time \( (t_{j_m^\star}, j_m^\star)\) as

\[ \begin{align} \xi_{no}&= \Psi_{f_{no}} \left(T^\dagger(\eta + B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)),\tau\right) \nonumber \end{align} \]

(448)

\[ \begin{align} &= e^{F_{no}\tau} T^\dagger(\eta + B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)) + \int_0^\tau e^{F_{no}(\tau-s)}U_{cno}ds, \\\end{align} \]

(449)

which is known to be in \( \Xi_{no}\) . With \( \hat{\eta}\) and \( \hat{T}\) of dynamics as in the observer (264), the estimation errors \( \tilde{\eta} := \eta - \hat{\eta}\) and \( \tilde{T} : = T - \hat{T}\) verify \( \dot{\tilde{\eta}} = 0\) and \( \dot{\tilde{T}} = 0\) during flows and at jumps, after time \( (t_{j_m},j_m)\) ,

\[ \begin{align*} \tilde{T}^+ &= \gamma A (T (\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \hat{T} \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)))+ B_{dno}(\mathcal{H}_{dno}(\xi_o,u_d,\tau)(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger \\ &~~ {} - \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)))\\ &~~ {} + B_{ono}(\mathcal{J}_{ono}(\xi_o,u_d,\tau)(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger- \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))) \\ &= \gamma A \tilde{T} \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)) + \gamma A T((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)))\\ &~~ {} + B_{dno} ((\mathcal{H}_{dno}(\xi_o,u_d,\tau) - \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger\\ &~~ {} + \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) ((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))) \\ & ~~{} + B_{ono} ((\mathcal{J}_{ono}(\xi_o,u_d,\tau) - \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger\\ &~~ {} + \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))), \end{align*} \]

and

\[ \begin{align*} \tilde{\eta}^+ &= \gamma A \tilde{\eta} + \gamma A \tilde{T} \int_0^{\tau}e^{-F_{no}s}U_{cno}ds+ \gamma A B_{ono}(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)) \\ &~~{}-B_{dno}(H_{do}(\xi_o,u_d) - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)) +T^+J_{noo}(\xi_o,u_d)\\ &~~{} - \hat{T}^+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)- B_{ono}(J_o(\xi_o,u_d) - J_o(\rm{sat}_o(\hat{\xi}_o),u_d))\\ &= \gamma A \tilde{\eta} + \gamma A \tilde{T} \int_0^{\tau}e^{-F_{no}s}U_{cno}ds+ \gamma A B_{ono}(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)) \\ &~~{}-B_{dno}(H_{do}(\xi_o,u_d) - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d)) +\tilde{T}^+J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &~~{}+T^+(J_{noo}(\xi_o,u_d)-J_{noo} (\rm{sat}_o(\hat{\xi}_o),u_d))- B_{ono}(J_o(\xi_o,u_d) - J_o(\rm{sat}_o(\hat{\xi}_o),u_d)). \end{align*} \]

Plugging the expression of \( \tilde{T}^+\) above into the one in \( \tilde{\eta}^+\) , we obtain

\[ \begin{align} \tilde{\eta}^+ & = \gamma A \tilde{\eta} + \gamma A \tilde{T} u_1 + \gamma v_1 + w_1,\\ \tilde{T}^+ & = \gamma A \tilde{T} U_2 + \gamma V_2 + W_2, \\\end{align} \]

(450.a)

where

\[ \begin{align*} u_1 &= \int_0^{\tau}e^{-F_{no}s}U_{cno}ds + \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d),\\ v_1 &= AB_{ono}(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\\ &~~{}+AT((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)))J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d),\\ w_1 & = B_{dno}(((\mathcal{H}_{dno}(\xi_o,u_d,\tau) - \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger\\ &~~{}+ \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)))J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &~~ {}-(H_{do}(\xi_o,u_d) - H_{do}(\rm{sat}_o(\hat{\xi}_o),u_d))) \\ &~~{} + B_{ono}(((\mathcal{J}_{ono}(\xi_o,u_d,\tau) - \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger\\ &~~{} + \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))) J_{noo}(\rm{sat}_o(\hat{\xi}_o),u_d)\\ &~~{}-(J_o(\xi_o,u_d) - J_o(\rm{sat}_o(\hat{\xi}_o),u_d)))+T^+(J_{noo}(\xi_o,u_d)-J_{noo} (\rm{sat}_o(\hat{\xi}_o),u_d)),\\ U_2 &= \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)),\\ V_2 &= AT ((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))),\\ W_2 & = B_{dno} ((\mathcal{H}_{dno}(\xi_o,u_d,\tau) - \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger\\ &~~{}+ \mathcal{H}_{dno}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau) ((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))) \\ & ~~{} + B_{ono} ( (\mathcal{J}_{ono}(\xi_o,u_d,\tau) - \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))(\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger \\ &~~{}+ \mathcal{J}_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau)((\mathcal{J}_{no}(\xi_o,u_d,\tau))^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\mathcal{J}_{no}(\rm{sat}_o(\hat{\xi}_o),u_d,\tau))), \end{align*} \]

with the expression of \( T^+\) from (444).

This is the third part of this proof. We are ready to start a Lyapunov analysis in the new coordinates, with \( \eta\) replacing \( \xi_{no}\) in view of applying Theorem 24 after time \( (t_{j_m^\star},j_m^\star)\) . To place ourselves in the framework of Section 6.2.2.1, we define \( z = (z_o,z_{no})\) where \( z_o = \xi_o\) and \( z_{no} = (\eta,T)\) , as well as \( \hat{z} = (\hat{z}_o,\hat{z}_{no})\) where \( \hat{z}_o = \hat{\xi}_o\) and \( \hat{z}_{no} = (\hat{\eta},\hat{T})\) . We also consider the extended inputs \( (\mathfrak{u}_{c,\rm{ ext }},\mathfrak{u}_{d,\rm{ ext }})\) defined by \( \mathfrak{u}_{c,\rm{ ext }}(t)=(\mathfrak{u}_c(t),t)\in \mathcal{U}_{c,\rm{ ext }}\) and \( \mathfrak{u}_{d,\rm{ ext }}(j)=(\mathfrak{u}_d(j),t_{j+1})\in \mathcal{U}_{d,\rm{ ext }}\) , with \( \mathcal{U}_{c,\rm{ ext }}=\mathcal{U}_c\times \mathbb{R}_{\geq 0}\) and \( \mathcal{U}_{d,\rm{ ext }}=\mathcal{U}_d\times \mathbb{R}_{\geq 0}\) . We have seen that \( z\) remains at all times in the compact set \( \Xi_o \times \Xi_\eta(\gamma) \times ¶_{T}(\gamma)\) . Moreover, after time \( (t_{j_m^\star},j_m^\star)\) , we know that, given an input trajectory \( \mathfrak{u}_c\in \mathfrak{U}_c\) , \( (z,u_{c,\rm{ ext }},\tau)\in C_z^{\mathfrak{u}_c}(\gamma)\) during flows and \( (z,u_{d,\rm{ ext }},\tau)\in D_z^{\mathfrak{u}_c}(\gamma)\) at jumps where

\[ \begin{align*} C_z^{\mathfrak{u}_c}(\gamma) &= \Big\{ (\xi_o,\eta, T, u_c,t ,\tau) \in(\Xi_o \cap C_o) \times \Xi_\eta(\gamma) \times \mathcal{C}_{T}(\gamma) \times \mathcal{U}_c \times \mathbb{R}_{\geq 0} \times [0,\tau_M] : \\ & \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_o,\\ & \exists \xi_{no}\in \Xi_{no}: \eta = T \Psi_{f_{no}} (\xi_{no},-\tau) - B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\Big\},\\ D_z^{\mathfrak{u}_c}(\gamma) &= \Big\{ (\xi_o,\eta, T, u_d,t ,\tau) \in (\Xi_o\cap D_o) \times \Xi_\eta(\gamma) \times \mathcal{C}_{T}(\gamma) \times \mathcal{U}_d \times \mathbb{R}_{\geq 0} \times \mathcal{I} :\mathcal{J}_{no}(\xi_o,u_d,\tau) \in \mathcal{C}_{\mathcal{J}_{no}}, \\ & \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_o,\\ & \exists \xi_{no}\in \Xi_{no} : \eta = T \Psi_{f_{no}} (\xi_{no},-\tau) - B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\Big\}, \end{align*} \]

where

\[ \begin{align*} \mathcal{C}_{\mathcal{J}_{no}} & = \left\{\mathcal{J}_{no} \in \mathbb{R}^{n_{no} \times n_{no}}: \frac{1}{c_{\mathcal{J}_{no}^{-1}}^2} \rm{Id} \leq \mathcal{J}_{no}^\top \mathcal{J}_{no}\leq \overline{c}_{\mathcal{J}_{no}}^2 \rm{Id}\right\},\\ \mathcal{C}_{T}(\gamma) & = \left\{T \in \mathbb{R}^{n_\eta \times n_{no}}: (\underline{c}_{T}(\gamma))^2 \rm{Id} \leq T^\top T \leq (\overline{c}_T(\gamma))^2 \rm{Id} \right\}, \end{align*} \]

for some appropriate \( \overline{c}_{\mathcal{J}_{no}} > 0\) guaranteed to exist by Assumption 19 and the compactness of \( [0,\tau_M]\) (depending only on \( \Xi_o\) , \( \mathcal{U}_d\) , and \( \tau_M\) ). Consider the Lyapunov function

\[ \begin{equation} V_{no}(z_{no},\hat{z}_{no}) = (\eta-\hat{\eta})^\top Q (\eta-\hat{\eta}) + \|T-\hat{T}\|^2, \end{equation} \]

(451)

where \( Q = Q^\top > 0\) is a solution to \( A^\top Q A < Q\) , which exists because \( A\) is Schur. Denote \( \underline{c}_Q > 0\) and \( \overline{c}_Q > 0\) , respectively, as the minimum and maximum eigenvalues of \( Q\) . Then, for all \( z_{no}=(\eta,T)\) and for all \( \hat z_{no}=(\hat\eta,\hat{T})\) in \( \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\eta \times n_{no}}\) , we have

\[ \begin{equation} \underline{c}_Q |\eta-\hat{\eta}|^2 + \|T-\hat{T}\|^2 \leq V_{no}(z_{no},\hat{z}_{no}) \leq \overline{c}_Q |\eta-\hat{\eta}|^2 + \|T-\hat{T}\|^2. \end{equation} \]

(452)

Besides, whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta,T,u_c,t,\tau)\in C_z^{\mathfrak{u}_c}(\gamma)\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{T})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\eta \times n_{no}}\) , we have

\[ \begin{equation} \dot{V}_{no}(\xi_o,\hat{\xi}_o,\eta,\hat{\eta},T,\hat{T},\tau,t,u_c) = 0, \end{equation} \]

(453.a)

along the respective flow dynamics. Now, we upper bound \( V_{no}^+\) at jumps. By the global Lipschitzness of \( f_{o,\rm{sat}}\) with respect to \( \xi_o\) , uniformly with respect to \( u_c\) , Lemma 25 in Section 6.2.2.6 allows us to show that \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) and \( \mathfrak{u}_c\in \mathfrak{U}_c\) . We also show that \( \rm{sat}_o\) and \( \rm{sat}_{no}\) satisfy (433) as in the proof of Theorem 16. Besides, applying Lemma 32 in Section 6.2.2.6 to \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}\) and then \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) , we deduce that there exists \( L_{\hat{\mathcal{J}}_{no}} > 0\) such that

\[ \begin{align*} \|\rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\hat{\mathcal{J}}_{no})\| &\leq \frac{1}{\underline{c}_{\hat{\mathcal{J}}_{no}}}, & \forall\hat{\mathcal{J}}_{no} &\in \mathbb{R}^{n_{no} \times n_{no}},\\ \|\mathcal{J}_{no}^\dagger - \rm{inv}_{\underline{c}_{\hat{\mathcal{J}}_{no}}}(\hat{\mathcal{J}}_{no})\| &\leq L_{\hat{\mathcal{J}}_{no}} \|\mathcal{J}_{no} - \hat{\mathcal{J}}_{no}\|, & \forall(\mathcal{J}_{no},\hat{\mathcal{J}}_{no}) &\in \mathcal{C}_{\mathcal{J}_{no}} \times \mathbb{R}^{n_{no} \times n_{no}}, \end{align*} \]

and there exists \( L_{\hat{T}}(\gamma) > 0\) such that

\[ \begin{align*} \|\rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})\| &\leq \frac{1}{\underline{c}_{T}(\gamma)}, & \forall\hat{T} &\in \mathbb{R}^{n_\eta \times n_{no}},\\ \|T^\dagger - \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})\| &\leq L_{\hat{T}}(\gamma) \|T - \hat{T}\|, & \forall(T,\hat{T}) &\in \mathcal{C}_{T} \times \mathbb{R}^{n_\eta \times n_{no}}. \end{align*} \]

Exploiting the boundedness of \( u_1\) and \( U_2\) (independently of \( \gamma\) ), we apply Lemma 31 in Section 6.2.2.6 to the Lyapunov function (451) and estimation error \( (\tilde{\eta},\tilde{T})\) with jump dynamics (450), to compute \( V_{no}^+\) . From the global Lipschitzness of \( f_{o,\rm{sat}}(\cdot,u_c)\) uniformly in \( u_c\in \mathcal{U}_c\) , the property of \( \rm{sat}_o\) that we have just proven, and the properties of \( \rm{inv}_\underline{{c}_{\hat{\mathcal{J}}_{no}}}\) and \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) proven above, we get that \( (v_1,w_1,V_2,W_2)\) can be upper-bounded by \( |\xi_o-\hat{\xi}_o|\) , with gains depending on \( \gamma\) . As a result, picking \( c_1\in[0,1)\) , we deduce that there exists \( 0 < \gamma^\star_2 \leq \gamma^\star_1\) such that for any \( 0 < \gamma < \gamma^\star_2\) , there exists \( c_2(\gamma)>0\) such that, whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta,T,u_d,t,\tau)\in D_z^{\mathfrak{u}_c}(\gamma)\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{T})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\eta \times n_{no}}\) ,

\[ \begin{align} V_{no}^+(\xi_o,\hat{\xi}_o,\eta,\hat{\eta},T,\hat{T},\tau,t,u_d)&\leq \gamma c_1 V_{no}(z_{no},\hat{z}_{no})+ c_2(\gamma) |\xi_o - \hat{\xi}_o|^2\notag\\ &\leq \gamma c_1 V_{no}(z_{no},\hat{z}_{no}) + \frac{c_2(\gamma)}{\underline{b}_o(\ell)}V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau), \\\end{align} \]

(453.b)

along the respective jump dynamics, where the latter inequality is obtained from Item (XX) of Assumption 18. From (452) and (453), we see that \( V_{no}\) satisfies the second item of all conditions of Theorem 24 (note that \( c_2(\gamma)/(\underline{b}_o(\ell)) \) is rational in \( \ell\) because so is \( \underline{b}_o(\ell)\) ). Now, we check if \( V_{o,\ell}\) also satisfies all the first items in those conditions. For that, we need to upper bound \( \xi_o^+ - \hat{\xi}_o^+\) given in (436), and thus \( \xi_{no} - \hat{\xi}_{no}\) . Combining (448) with (264.b), we obtain

\[ \begin{align*} \xi_{no} - \hat{\xi}_{no} &= e^{F_{no}\tau} \left(T^\dagger \eta - \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})\hat{\eta}\right) \\ &~~{} + e^{F_{no}\tau} \left(T^\dagger B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)- \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T}) B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\right)\\ &=e^{F_{no}\tau} \left((T^\dagger - \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})) \eta + \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T}) (\eta- \hat{\eta})\right) \\ &~~ {}+ e^{F_{no}\tau} \Big(\left(T^\dagger - \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T})\right)B_{ono}\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau) \\ & ~~{}+ \rm{inv}_{\underline{c}_{T}(\gamma)}(\hat{T}) B_{ono}\left(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)\right)\Big). \end{align*} \]

By the boundedness of \( \Xi_o\) , \( \Xi_\eta(\gamma)\) , and \( [0,\tau_M]\) , by the uniform Lipschitzness of \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) for \( (t,\tau)\in \mathbb{R}_{\geq 0}\times [0, \tau_M]\) , and the properties of \( \rm{inv}_\underline{{c}_{T}(\gamma)}\) shown above with a bound depending only on \( \underline{c}_{T}(\gamma)\) , there exist \( c_3(\gamma) > 0\) , \( c_4(\gamma) > 0\) , and \( c_5(\gamma) > 0\) such that whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta, T, t ,\tau) \in \Xi_o \times \Xi_\eta(\gamma) \times \mathcal{C}_{T}(\gamma) \times \mathbb{R}_{\geq 0} \times [0,\tau_M]\) such that \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_o\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{T})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\eta \times n_{no}}\) , \( \xi_{no}\) and \( \hat{\xi}_{no}\) defined in (448) and (264.b) verify

\[ \begin{equation} |\xi_{no} - \hat{\xi}_{no}| \leq c_3(\gamma) |\xi_o - \hat{\xi}_o| + c_4(\gamma) |\eta - \hat{\eta}| + c_5(\gamma)\|T - \hat{T}\|. \end{equation} \]

(454)

Thanks to Assumption 19 and (433.b), there exist \( c_6>0\) , \( c_7>0\) , \( c_8(\gamma)>0\) , and \( c_9(\gamma) > 0\) and from these, \( d_o(\ell,\gamma):= (\overline{b}_o c_8(\gamma))/\underline{b}_o(\ell) > 0\) (rational in \( \ell\) because so is \( \underline{b}_o\) ) and \( d_{ono}(\gamma):=\overline{b}_oc_9(\gamma) > 0\) such that \( V_{o,\ell}\) in Item (XX) of Assumption 18 satisfies for all \( (\xi_o,\eta,T,u_d,t,\tau)\in D_z^{\mathfrak{u}_c}(\gamma)\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{T})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\eta \times n_{no}}\) (with \( \xi_{no}\) and \( \hat{\xi}_{no}\) still given by (448) and (264.b)),

\[ \begin{align*} V_{o,\ell}^+(z,\hat{z},p,\tau,u_d)&\leq\overline{b}_o|\xi_o^+ - \hat{\xi}_o^+|^2\\ & \leq \overline{b}_o|J_o(\xi_o, u_d) + J_{ono}(\xi_o,u_d)\xi_{no} - (J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)\rm{sat}_{no}(\hat{\xi}_{no}))|^2\\ &\leq \overline{b}_o|J_o(\xi_o, u_d)- J_o(\rm{sat}_o(\hat{\xi}_o),u_d) + (J_{ono}(\xi_o,u_d)-J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d))\xi_{no}\\ &~~{} + J_{ono}(\rm{sat}_o(\hat{\xi}_o),u_d)(\xi_{no}-\rm{sat}_{no}(\hat{\xi}_{no}))|^2\\ &\leq\overline{b}_o c_6 |\xi_o - \hat{\xi}_o|^2 + \overline{b}_oc_7 |\xi_{no} - \hat{\xi}_{no}|^2\\ &\leq\overline{b}_o c_8(\gamma) |\xi_o - \hat{\xi}_o|^2 + \overline{b}_oc_9(\gamma) |(\eta,T)-(\hat{\eta},\hat{T})|^2\\ & \leq \frac{\overline{b}_o c_8(\gamma)}{\underline{b}_o(\ell)} V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) + \overline{b}_oc_9(\gamma)|(\eta,T)-(\hat{\eta},\hat{T})|^2\\ & \leq d_o(\ell,\gamma)V_{o,\ell}(z_o,\hat{z}_o,p,\tau) + d_{ono}(\gamma)|z_{no}-\hat{z}_{no}|^2, \end{align*} \]

along the respective jump dynamics, because \( \xi_{no}\in \Xi_{no}\) by the definition of \( D_z^{\mathfrak{u}_c}(\gamma)\) , where we use the norm \( |\cdot|\) for the vector-matrix element-wise concatenation like \( (\eta,T)\) and Item (XX) of Assumption 18. This inequality, along with Item (XX) of Assumption 18, shows that \( V_{o,\ell}\) satisfies the inequalities involving it in Theorem 24 in the new coordinates, uniformly in the chosen input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) . Since \( \varphi_{c,\ell}\) and \( \varphi_{d,\ell}\) are independent of \( (\hat{\xi}_o,\hat{\eta},\hat{T})\) , for any \( \ell > \ell^\star_1\) , solutions \( (\xi,\eta,T,\hat{\xi}_o,\hat{\eta},p,\hat{T},\tau)\) to the cascade (245)-(264) initialized in \( \Xi_0\times \mathbb{R}^{n_\eta}\times\{T_0\}\times\mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{T_0\}\times \{0\}\) with inputs \( (\mathfrak{u}_{c,\rm{ ext }},\mathfrak{u}_{d,\rm{ ext }})\in\mathfrak{U}_{c,\rm{ ext }}\times \mathfrak{U}_{d,\rm{ ext }}\) are such that \( p(t,j) \in ¶_{c}\) during flows and \( p(t,j) \in ¶_{d}\) at jumps. Applying Theorem 24 after jump \( j_m^\star\) , we deduce that given any \( \lambda> 0\) , there exist \( 0 < \gamma^\star_3 \leq \gamma^\star_2\) and \( \ell^\star_3 \geq \ell^\star_2\) such that for any \( 0 < \gamma < \gamma^\star_3\) and for any \( \ell > \ell^\star_3\) , there exists \( \rho_1(\ell,\gamma) > 0\) such that for any of those solutions,

\[ \begin{multline} |(\xi_o,\eta,T)(t,j) - (\hat{\xi}_o,\hat{\eta},\hat{T})(t,j)| \leq \rho_1(\ell,\gamma) |(\xi_o,\eta,T)(t_{j_m^\star},j_m^\star) - (\hat{\xi}_o,\hat{\eta},\hat{{T}})(t_{j_m^\star},j_m^\star)| e^{-\lambda(t+j)},\\ \forall (t,j) \in \rm{dom} \xi: j \geq j_m^\star. \end{multline} \]

(455)

Combining this with (454), we obtain Theorem 17.

6.2.2.4 Proof of Theorem 18

First, exploiting exponential increase over rational growth, given \( \overline{c}_o\) from Assumption 23 and using the compactness of \( \Xi_o\) , let \( \ell_2^\star\geq \ell_1^\star\) (defined in Item (XX) of Assumption 18) such that, for all \( \ell >\ell_2^\star\) ,

\[ \begin{equation} \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\leq \overline{c}_o, \end{equation} \]

(456)

where \( \overline{b}_o\) , \( \underline{b}_o\) , \( \lambda_c\) are defined in Assumption 18, \( \tau_m:=\min \mathcal{I}>0\) , and \( M_o > 0\) is a bound of \( g_o(\rm{sat}_o(\hat{\xi}_o),\rm{sat}_{no}(\hat{\xi}_{no}),u_d)\) (uniform in \( u_d \in \mathcal{U}_d\) ) obtained from the bounds of \( \rm{sat}_o\) , \( \rm{sat}_{no}\) , and Assumption 22.
Now, pick \( \ell > \ell^\star_2\) . Consider a solution \( (\xi,\hat{\xi}_o,\hat{\eta}, p, \tau)\) to the cascade (232)-(276) initialized in \( \Xi_0\times\mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times¶_0\times\{0\}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d) \in \mathfrak{U}_c\times\mathfrak{U}_d\) (which is complete thanks to Item (XIV) of Assumption 17 and given the observer dynamics which do not allow finite-time escape thanks to saturation functions and the local boundedness of the maps of system (232) as in Assumption 22), with the map sequence \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) , the matrices \( (A,B_d,B_o)\) and the saturation map \( \rm{sat}_o\) as described in the statement of Theorem 18. In the following, we refer to the jump times of this solution as \( t_j\) instead of \( t_j(\xi)\) to ease the notations. First, from i) Item (XX) of Assumption 18, ii) the fact that \( t_{j+1} - t_j \geq \tau_m\) according to Item (XIV) of Assumption 17, and iii) the choice of \( \ell > \ell^\star_2\) satisfying (456), we have for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{align*} |\xi_o(t_{j+1},j) - \hat{\xi}_o(t_{j+1},j)|& \leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} |\xi_o(t_j,j) - \hat{\xi}_o(t_j,j)| e^{-\ell \frac{\lambda_c}{2}\tau_m} \\ &\leq \sqrt{\frac{\overline{b}_o}{\underline{b}_o(\ell)}} \left(\max_{\xi_o \in \Xi_o}|\xi_o| + M_o\right) e^{-\ell \frac{\lambda_c}{2}\tau_m}\\ & \leq \overline{c}_o. \end{align*} \]

Then, by the definition of \( \rm{sat}_o\) and since \( \xi_o(t_{j+1},j)\in \Xi_o\cap D_o\) , we have, for all \( j\in\mathbb{N}_{\geq j_m}\) ,

\[ \rm{sat}_o(\hat{\xi}_o(t_{j+1},j)) = \hat{\xi}_o(t_{j+1},j). \]

Consider the discrete-time signals \( (\xi_{o,j})_{j\in \mathbb{N}}\) and \( (\xi_{no,j})_{j\in \mathbb{N}}\) defined as

\[ \begin{equation} \xi_{o,j} = \xi_o(t_{j+1},j), ~~ \xi_{no,j} = \xi_{no}(t_{j},j), ~~ \forall j \in \mathbb{N}. \end{equation} \]

(457)

Notice that \( j \mapsto \xi_{no,j}\) is solution to system (266) fed with \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (u_j)_{j \in \mathbb{N}}:=(\xi_{o,j},u_{d,j}, t_j,\tau_j)_{j\in \mathbb{N}}\) . From Item (XXIII) of Assumption 23, there exists a map sequence \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) such that

\[ \begin{equation} \xi_{no,j} = \mathcal{T}_j(\zeta_j), ~~ \forall j \in \mathbb{N}_{\geq j_m^\star}, \end{equation} \]

(458)

for some solution \( j \mapsto \zeta_j\) to the dynamics (268) remaining in \( \Xi_{\zeta}\) for all \( j \in \mathbb{N}\) . Notice that \( (\mathcal{T}_j)\) depends only on the picked system solution and inputs, namely \( (\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)\) . Define, over the time domain \( \rm{dom} \xi\) of the considered solution, the hybrid arc \( (t,j) \mapsto \zeta(t,j)\) with dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\zeta} & = &0 \\ \zeta^+&=& A \zeta+ B_{dno}y_d + B_og_o(\xi_o,\xi_{no},u_d), \end{array} \right. \end{equation} \]

(459)

and initialized as \( \zeta(0,0) = \zeta_0\) . Since \( \zeta\) is complete, is constant during flows, and has the same jump inputs as in the dynamics of \( j \mapsto \zeta_j\) , it follows that

\[ \begin{equation} \zeta(t,j) = \zeta_j, ~~ \forall (t,j) \in \rm{dom} \xi, \end{equation} \]

(460)

and so it remains in the compact set \( \Xi_\zeta\) . Define, over the time domain \( \rm{dom} \xi\) of the considered solution, the hybrid arc \( (t,j) \mapsto \eta(t,j)\) of the same dimension as \( \zeta\) , as

\[ \begin{equation} \eta(t,j) = \zeta(t,j) - B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)). \end{equation} \]

(461)

Arguing in a similar way as in the proof of Theorem 16 above, we notice that

\[ \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-\tau(t,j)) = \Psi_{f_o(\cdot, \mathfrak{u}_c)}(\xi_o(t,j),t,-(t-t_j)) = \xi_o(t_j,j)\in \Xi_o, \]

for all \( j\in \rm{dom}_j \xi\) and \( t\in [t_j,t_{j+1}]\) . We deduce that \( \eta\) remains at all times in the compact set

\[ \begin{equation} \Xi_\eta:=\{\eta \in \mathbb{R}^{n_\eta}:\exists (\xi_o,\zeta) \in \Xi_o \times \Xi_\zeta, \eta = \zeta - B_o\xi_o\}. \end{equation} \]

(462)

We deduce also that \( \eta\) verifies \( \dot{\eta} = 0\) during flows. At jumps, we have (using that \( \tau^+=0\) and so \( \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t^+,-\tau^+) = \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o^+,t,0) = \xi_o^+\) ),

\[ \begin{align*} \eta^+ &= \zeta^+ - B_o\xi_o^+\\ &= A \zeta+ B_dy_d + B_og_o(\xi_o,\xi_{no},u_d) - B_og_o(\xi_o,\xi_{no},u_d)\\ &= A (\eta+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau))+ B_dy_d\\ &= A \eta+ B_dy_d + AB_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau). \end{align*} \]

Therefore, \( \eta\) is solution to the dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\eta} & = &0 \\ \eta^+&=& A \eta+ B_dy_d + AB_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau). \end{array} \right. \end{equation} \]

(463)

On the other hand, combining (457), (458), and (460) we deduce that \( \xi_{no}(t_j,j)=\mathcal{T}_j(\zeta(t,j))\) for all \( (t,j)\in \rm{dom} \xi\) with \( j\geq j_m^\star\) . Exploiting (461), it follows that, after \( j_m^\star\) jumps, we can express \( \xi_{no}\) as a function of \( (\xi_o,\eta,\tau,t)\) and input \( \mathfrak{u}_c\) as \( \xi_{no}(t,j) =\Psi_{f_\rm{{sat}}(\cdot,\mathfrak{u}_c),no}(\xi_o(t_j,j),\xi_{no}(t_j,j),t_j,\tau(t,j))\) , namely,

\[ \begin{align} \xi_{no} &=\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c),no}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau),\mathcal{T}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau))),t_j,\tau) \notag \\ &:= \mathfrak{T}^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}(\xi_o,\eta,t,j,t_j,\tau). \end{align} \]

(464)

The dependence of \( \mathfrak{T}^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) on \( \mathfrak{u}_d \in \mathfrak{U}_d\) is implicit in \( \mathcal{T}_j\) . With \( \hat{\eta}\) taking the dynamics in observer (276), the estimation error \( \tilde{\eta} := \eta - \hat{\eta}\) is solution to the dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\tilde{\eta}} & = &0 \\ \tilde{\eta}^+&=& A \tilde{\eta} + AB_o(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)). \end{array} \right. \end{equation} \]

(465)

Now, define, over the time domain \( \rm{dom} \xi\) of the considered solution, the hybrid arc \( (t,j) \mapsto \hat{\Gamma}(t,j)\) with dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\hat{\Gamma}} & = &0 \\ \hat{\Gamma}^+&=& A^\prime \hat{\Gamma} + B^\prime |\xi_o-\hat{\xi}_o|, \end{array} \right. \end{equation} \]

(466)

with \( (A^\prime,B^\prime)\) coming from Item (XXV) of Assumption 23, and initialized as \( \hat{\Gamma}(0,0) = \hat{\Gamma}_0\) with the discrete-time variable \( j \mapsto \hat{\Gamma}_j\) coming from the same item, fed with the considered solution \( (t,j) \mapsto (\xi_o(t,j),\hat{\xi}_o(t,j))\) . Since \( \hat{\Gamma}\) is complete and is constant during flows and since \( \xi_{o,j} = \xi_o(t_{j+1},j)\) and \( \hat{\xi}_{o,j} = \hat{\xi}_o(t_{j+1},j)\) , it follows that \( \hat{\Gamma}(t,j) = \hat{\Gamma}_j\) for all \( (t,j) \in \rm{dom} \xi\) . Also, define, over the time domain \( \rm{dom} \xi\) of the considered solution, the hybrid arc \( (t,j) \mapsto \Gamma(t,j)\) with dynamics

\[ \begin{equation} \left\{ \begin{array}{@{}r@{\;}c@{\;}l@{}} \dot{\Gamma} & = &0 \\ \Gamma^+&=& A^\prime \Gamma, \end{array} \right. \end{equation} \]

(467)

with the same \( (A^\prime,B^\prime)\) and initialized as \( \Gamma(0,0) = 0\) . It follows that \( \Gamma(t,j) = 0\) for all \( (t,j) \in \rm{dom} \xi\) . To place ourselves in the framework of Section 6.2.2.1, we define \( z = (z_o,z_{no})\) where \( z_o = \xi_o\) and \( z_{no} = (\eta,\Gamma)\) , as well as \( \hat{z} = (\hat{z}_o,\hat{z}_{no})\) where \( \hat{z}_o = \hat{\xi}_o\) and \( \hat{z}_{no} = (\hat{\eta},\hat{\Gamma})\) . We also consider the extended inputs \( (\mathfrak{u}_{c,\rm{ ext }},\mathfrak{u}_{d,\rm{ ext }})\) defined by \( \mathfrak{u}_{c,\rm{ ext }}(t)=(\mathfrak{u}_c(t),t,j(t),t_{j(t)})\in \mathcal{U}_{c,\rm{ ext }}\) where \( j(t):=\min\{j\in\mathbb{N}:(t,j) \in \rm{dom} \xi\}\) and \( \mathfrak{u}_{d,\rm{ ext }}(j)=(\mathfrak{u}_d(j),t_{j+1},j,t_{j})\in \mathcal{U}_{d,\rm{ ext }}\) , with \( \mathcal{U}_{c,\rm{ ext }}=\mathcal{U}_c\times \mathbb{R}_{\geq 0}\times \mathbb{N}\times \mathbb{R}_{\geq 0}\) and \( \mathcal{U}_{d,\rm{ ext }}=\mathcal{U}_d\times \mathbb{R}_{\geq 0}\times \mathbb{N}\times \mathbb{R}_{\geq 0}\) . We have seen that \( z\) remains at all times in the compact set \( \Xi_o \times \Xi_\eta \times \{0\}\) . Moreover, we know that, given the system initial condition \( \xi(0,0)\) and input trajectories \( \mathfrak{u}_c\in \mathfrak{U}_c\) and \( \mathfrak{u}_d \in \mathfrak{U}_d\) , \( (z,u_{c,\rm{ ext }},\tau)\in C_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) during flows and \( (z,u_{d,\rm{ ext }},\tau)\in D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) at jumps where

\[ \begin{align*} C_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)} =& \Big\{(\xi_o,\eta,\Gamma, u_c,t ,j,t_j,\tau) \in(\Xi_o \cap C_o) \times \Xi_\eta \times \{0\} \times \mathcal{U}_c \times \mathbb{R}_{\geq 0} \times\mathbb{N}\times \mathbb{R}_{\geq 0}\times [0,\tau_M] : \\ & \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_o, \eta + B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_\zeta,\\ & \mathfrak{T}^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}(\xi_o,\eta,t,j,t_j,\tau)\in \Xi_{no}\Big\},\\ D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)} &= \Big\{(\xi_o,\eta,\Gamma, u_d,t ,j,t_j,\tau) \in (\Xi_o\cap D_o) \times \Xi_\eta \times \{0\} \times \mathcal{U}_d \times \mathbb{R}_{\geq 0}\times\mathbb{N}\times\mathbb{R}_{\geq 0} \times \mathcal{I}: \\ & \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_o, \eta + B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)\in \Xi_\zeta,\\ & \mathfrak{T}^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}(\xi_o,\eta,t,j,t_j,\tau)\in \Xi_{no}\Big\}. \end{align*} \]

Consider the Lyapunov function

\[ \begin{equation} V_{no}(z_{no},\hat{z}_{no}) = (\eta-\hat{\eta})^\top Q (\eta-\hat{\eta}) + (\Gamma-\hat{\Gamma})^\top Q^\prime (\Gamma-\hat{\Gamma}), \end{equation} \]

(468)

where \( Q = Q^\top > 0\) is solution to \( A^\top Q A - Q < 0\) and \( Q^\prime = (Q^\prime)^\top > 0\) is solution to \( (A^\prime)^\top Q^\prime A^\prime - Q^\prime < 0\) , which exist because \( A\) and \( A^\prime\) are Schur. Denote \( \underline{c}_Q > 0\) and \( \overline{c}_Q > 0\) , respectively, as the minimum and maximum eigenvalues of \( Q\) , and similarly, \( \underline{c}_{Q^\prime} > 0\) and \( \overline{c}_{Q^\prime} > 0\) , respectively, as the minimum and maximum eigenvalues of \( Q^\prime\) . Then, for all \( z_{no} = (\eta,\Gamma)\) and for all \( \hat z_{no}=(\hat\eta,\hat{\Gamma})\) in \( \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\Gamma}\) , we have

\[ \begin{equation} \underline{c}_Q |\eta-\hat{\eta}|^2 + \underline{c}_{Q^\prime}|\Gamma - \hat{\Gamma}|^2\leq V_{no}(z_{no},\hat{z}_{no}) \leq \overline{c}_Q |\eta-\hat{\eta}|^2+ \overline{c}_{Q^\prime}|\Gamma - \hat{\Gamma}|^2. \end{equation} \]

(469)

Besides, whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta,\Gamma,u_c,t,j,t_{j},\tau)\in C_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{\Gamma})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times\mathbb{R}^{n_\Gamma}\) , we have

\[ \begin{equation} \dot{V}_{no}(\xi_o,\hat{\xi}_o,\eta,\hat{\eta},\Gamma,\hat{\Gamma},\tau,t,u_c) = 0, \end{equation} \]

(470.a)

along the respective flow dynamics. Now, we upper bound \( V_{no}^+\) at jumps. Whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta,\Gamma,u_d,t,j,t_j,\tau)\in D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{\Gamma})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times\mathbb{R}^{n_\Gamma}\) , we have

\[ \begin{align*} V_{no}^+(\xi_o,\hat{\xi}_o,\eta,\hat{\eta},\Gamma,\hat{\Gamma},\tau,t,u_d)&= \star^\top Q \underbrace{(A \tilde{\eta} + AB_o(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)))}_{\star}\notag\\ & ~~{}+\star^\top Q^\prime \underbrace{(A^\prime (\Gamma-\hat{\Gamma}) - B^\prime |\xi_o-\hat{\xi}_o|}_{\star}\notag\\ &= \tilde{\eta}^\top A^\top Q A \tilde{\eta} \notag\\ &~~{} + 2 \tilde{\eta}^\top A^\top Q A B_o(\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))\notag\\ &~~{}+\star^\top Q \underbrace{A B_o (\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))}_{\star} \notag \\ &~~{}+ (\Gamma - \hat{\Gamma})^\top (A^\prime)^\top Q^\prime A^\prime (\Gamma - \hat{\Gamma}) - 2(\Gamma - \hat{\Gamma})^\top (A^\prime)^\top Q^\prime B^\prime |\xi_o-\hat{\xi}_o|\notag \\ &~~{} + (B^\prime)^\top Q^\prime B^\prime |\xi_o-\hat{\xi}_o|^2. \end{align*} \]

By the global Lipschitzness of \( f_{o,\rm{sat}}\) with respect to \( \xi_o\) , uniformly with respect to \( u_c\) , Lemma 25 in Section 6.2.2.6 allows us to show that \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) and \( \mathfrak{u}_c\in \mathfrak{U}_c\) . We denote \( L_{f_{o,\rm{sat}}}\) the corresponding Lipschitz constant. Using Young’s inequality, the fact that \( A\) , \( A^\prime\) are Schur, we deduce that there exist \( c_1\in [0,1)\) and \( c_2 > 0\) such that, whatever input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) , for all \( (\xi_o,\eta,\Gamma,u_d,t,j,t_j,\tau)\in D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{\Gamma})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times\mathbb{R}^{n_\Gamma}\) , we have

\[ \begin{align} V_{no}^+(\xi_o,\hat{\xi}_o,\eta,\hat{\eta},\Gamma,\hat{\Gamma},\tau,t,u_d)&\leq c_1 V_{no}(z_{no},\hat{z}_{no}) + c_2 |\xi_o - \hat{\xi}_o|^2 \notag \\ &\leq c_1 V_{no}(z_{no},\hat{z}_{no}) + \frac{c_2}{\underline{b}_o(\ell)} V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau), \\\end{align} \]

(470.b)

along the respective jump dynamics, where the latter inequality is obtained from Item (XX) of Assumption 18. From (469) and (470), we see that \( V_{no}\) satisfies the second item of all conditions of Theorem 23 (note that \( c_2 / \underline{b}_o(\ell) \) is rational in \( \ell\) because so is \( \underline{b}_o(\ell)\) ). Now, we check if \( V_{o,\ell}\) also satisfies all the first items in those conditions. For that, we need to upper bound \( \xi_o^+ - \hat{\xi}_o^+\) , and thus \( \xi_{no} - \hat{\xi}_{no}\) . Combining (464) with (276.b), we obtain, after \( j_m^\star\) jumps,

\[ \begin{multline} \xi_{no} - \hat{\xi}_{no} =\Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c),no}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau),\mathcal{T}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau))),t_j,\tau) \\ - \Psi_{f_{\rm{sat}}(\cdot,\mathfrak{u}_c),no}((\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau),\hat{\mathcal{T}}_j(\hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))),t_j,\tau). \end{multline} \]

(471)

By the global Lipschitzness of \( f_\rm{{sat}}\) with respect to \( \xi\) , uniformly with respect to \( u_c\) , Lemma 25 in Section 6.2.2.6 allows us to show that \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c),no}(\cdot,t,-\tau)\) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) , with Lipschitz constant \( L_{f_\rm{{sat}}} > 0\) . Because \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) is uniformly Lipschitz as in Item (XXIV) of Assumption 23, with the \( c_{\hat{\mathcal{T}}} > 0\) defined therein, and using (461), we have, after \( j_m^\star\) jumps,

\[ \begin{align*} |\xi_{no} - \hat{\xi}_{no}| & \leq L_{f_{\rm{sat}}}|\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau) - \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)| \\ & {}+ L_{f_{\rm{sat}}}|\mathcal{T}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)) - \hat{\mathcal{T}}_j(\hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))|\\ &\leq L_{f_{\rm{sat}}}L_{f_{o,\rm{sat}}}|\xi_o - \hat{\xi}_o| + L_{f_{\rm{sat}}}|\mathcal{T}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)) - \hat{\mathcal{T}}_j(\eta+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau))\\ & {} + \hat{\mathcal{T}}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)) - \hat{\mathcal{T}}_j(\hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))|\\ &\leq L_{f_{\rm{sat}}}L_{f_{o,\rm{sat}}}|\xi_o - \hat{\xi}_o| + L_{f_{\rm{sat}}}|\mathcal{T}_j(\zeta) - \hat{\mathcal{T}}_j(\zeta)|\\ & {} + L_{f_{\rm{sat}}} |\hat{\mathcal{T}}_j(\eta + B_o \Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)) - \hat{\mathcal{T}}_j(\hat{\eta}+B_o\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau))|\\ &\leq L_{f_{\rm{sat}}}L_{f_{o,\rm{sat}}}|\xi_o - \hat{\xi}_o| + L_{f_{\rm{sat}}}|\mathcal{T}_j(\zeta) - \hat{\mathcal{T}}_j(\zeta)| + L_{f_{\rm{sat}}}c_{\hat{\mathcal{T}}}|\eta - \hat{\eta}| \\ & {}+ L_{f_{\rm{sat}}}c_{\hat{\mathcal{T}}}\|B_o\| |\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\xi_o,t,-\tau)-\Psi_{f_{o,\rm{sat}}(\cdot,\mathfrak{u}_c)}(\hat{\xi}_o,t,-\tau)|\\ &\leq L_{f_{\rm{sat}}}L_{f_{o,\rm{sat}}}|\xi_o - \hat{\xi}_o| + L_{f_{\rm{sat}}}|\mathcal{T}_j(\zeta) - \hat{\mathcal{T}}_j(\zeta)| + L_{f_{\rm{sat}}}c_{\hat{\mathcal{T}}}|\eta - \hat{\eta}| + L_{f_{\rm{sat}}}c_{\hat{\mathcal{T}}}\|B_o\| L_{f_{o,\rm{sat}}} |\xi_o-\hat{\xi}_o|. \end{align*} \]

From Item (XXV) of Assumption 23, because \( \zeta\) is constant during flows, i.e, using (460), and since \( \Gamma(t,j) = 0\) for all \( (t,j) \in \rm{dom} \xi\) , we have after \( j_m^\star\) jumps,

\[ \begin{equation} |\xi_{no} - \hat{\xi}_{no}| \leq L_{f_{\rm{sat}}}L_{f_{o,\rm{sat}}}(1 + c_{\hat{\mathcal{T}}}\|B_o\|) |\xi_o-\hat{\xi}_o| + L_{f_{\rm{sat}}}c_{\hat{\mathcal{T}}}|\eta - \hat{\eta}| + L_{f_{\rm{sat}}}c_\Gamma|\Gamma - \hat{\Gamma}|. \end{equation} \]

(472)

We also show that \( \rm{sat}_o\) and \( \rm{sat}_{no}\) satisfy (433) as in the proof of Theorem 16. Thanks to Assumption 22 and (433), there exist \( c_3>0\) , \( c_4>0\) , and from these, \( d_o(\ell):=(\overline{b}_o c_3)/\underline{b}_o(\ell) > 0 \) (rational in \( \ell\) because so is \( \underline{b}_o\) ) and \( d_{ono}:=\overline{b}_oc_4 > 0\) such that \( V_{o,\ell}\) in Item (XX) of Assumption 18 satisfies for all \( (\xi_o,\eta,\Gamma,u_d,t,j,t_j,\tau)\in D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) and for all \( (\hat{\xi}_o,\hat{\eta},\hat{\Gamma})\in \mathbb{R}^{n_o}\times \mathbb{R}^{n_\eta}\times \mathbb{R}^{n_\Gamma}\) (with \( \xi_{no}\) and \( \hat{\xi}_{no}\) still given by (464) and (276.b)), after \( j_m^\star\) jumps,

\[ \begin{align*} V_{o,\ell}^+(z,\hat{z},p,\tau,u_d)&\leq\overline{b}_o|\xi_o^+ - \hat{\xi}_o^+|^2\\ & \leq \overline{b}_o|g_o(\hat{\xi}_o,\hat{\xi}_{no},u_d) - g_o(\rm{sat}_o(\hat{\xi}_o),\rm{sat}_{no}(\hat{\xi}_{no}),u_d)|^2\\ &\leq 2\overline{b}_oL_{g_o}^2|\xi_o - \rm{sat}_o(\hat{\xi}_o)|^2 + 2\overline{b}_oL_{g_o}^2|\xi_{no}-\rm{sat}_{no}(\hat{\xi}_{no})|^2\\ &\leq 2\overline{b}_o L_{g_o}^2 L_o^2 |\xi_o - \hat{\xi}_o|^2 + 2\overline{b}_o L_{g_o}^2L_{no}^2 |\xi_{no} - \hat{\xi}_{no}|^2\\ &\leq\overline{b}_o c_3 |\xi_o - \hat{\xi}_o|^2 + \overline{b}_oc_4 |(\eta,\Gamma)-(\hat{\eta},\hat{\Gamma})|^2\\ & \leq \frac{\overline{b}_o c_3}{\underline{b}_o(\ell)} V_{o,\ell}(\xi_o,\hat{\xi}_o,p,\tau) + \overline{b}_oc_4|(\eta,\Gamma)-(\hat{\eta},\hat{\Gamma})|^2\\ & \leq d_o(\ell)V_{o,\ell}(z_o,\hat{z}_o,p,\tau) + d_{ono}|z_{no}-\hat{z}_{no}|^2, \end{align*} \]

along the respective jump dynamics, because \( \xi_{no}\in \Xi_{no}\) by the definition of \( D_z^{(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)}\) , where \( L_{g_o} > 0\) denotes the Lipschitz constant of \( g_o\) with respect to \( \xi\) on \( ((\Xi_o\cap D_o) + \overline{c}_o\mathbb{B})\times((\Xi_{no}\cap D_{no}) + \overline{c}_{no}\mathbb{B})\) , uniformly in \( u_d \in \mathcal{U}_d\) . This inequality, along with Item (XX) of Assumption 18, shows that \( V_{o,\ell}\) satisfies, after time \( j_m^\star\) , the inequalities involving it in Theorem 23 in the new coordinates, uniformly in the chosen input trajectory \( \mathfrak{u}_c \in \mathfrak{U}_c\) . Since \( \varphi_{c,\ell}\) and \( \varphi_{d,\ell}\) are independent of \( (\hat{\xi}_o,\hat{\eta},\hat{T})\) , for any \( \ell > \ell^\star_1\) , solutions \( (\xi,\hat{\xi}_o,\hat{\eta},p,\tau)\) to the cascade (232)-(276) initialized in \( \Xi_0\times\mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times \{0\}\) with inputs \( (\mathfrak{u}_{c,\rm{ ext }},\mathfrak{u}_{d,\rm{ ext }})\in\mathfrak{U}_{c,\rm{ ext }}\times \mathfrak{U}_{d,\rm{ ext }}\) are such that \( p(t,j) \in ¶_{c}\) during flows and \( p(t,j) \in ¶_{d}\) at jumps. Given any system initial condition and input trajectories \( (\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)\in \Xi_0\times \mathfrak{U}_c \times \mathfrak{U}_d\) , applying Theorem 23 after jump \( j_m^\star\) , we deduce that there exists \( \ell^\star_3(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d) \geq \ell^\star_2\) such that for any \( \ell > \ell^\star_3(\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)\) , any maximal solution to the cascade (232)–(467)--(276)--(466) initialized in \( \{\xi(0,0)\} \times \mathbb{R}^{n_o}\times\mathbb{R}^{n_\eta}\times ¶_{0}\times\{0\}\times\{0\}\times\mathbb{R}^{n_\Gamma}\) with inputs \( (\mathfrak{u}_c,\mathfrak{u}_d)\) , verifies

\[ \lim_{t+j \to +\infty} |(\xi_o,\eta,\Gamma)(t,j) - (\hat{\xi}_o,\hat{\eta},\hat{\Gamma})(t,j)| = 0, \]

for \( (t,j)\) in the respective hybrid time domain. Combining this with (472), we obtain (236). Last, noticing that \( \ell^\star_3\) depends only on constants appearing in the Lyapunov inequalities coming from the boundedness, Lipschitzness, and diverse assumptions on the system and observer maps, that are all uniform with respect to the system initial condition and inputs \( (\xi(0,0),\mathfrak{u}_c,\mathfrak{u}_d)\) , we conclude that \( \ell^\star_3\) is also uniform. Therefore, Theorem 18 follows.

6.2.2.5 Proof of Proposition 2

We start by showing that for \( \overline{c}_o > 0\) sufficiently small, the properties of uniform invertibility and uniform Lipschitz backward distinguishability assumed in Assumption 25 are preserved (but with different constants) when the discrete-time system (266) is fed with inputs \( (\hat{\xi}_{o,k})_{k\in \mathbb{N}}\) instead of \( (\xi_{o,k})_{k\in \mathbb{N}}\) (and still the same input \( \mathfrak{u}_c \in \mathfrak{U}_c\) and the same inputs \( (u_{d,k},t_k,\tau_k)_{k \in \mathbb{N}}\) ), if \( |\xi_{o,k} - \hat{\xi}_{o,k}|\leq \overline{c}_o\) for all \( k\in \mathbb{N}_{\geq j_m}\) . Along every complete solution \( \xi \in §_\mathcal{H}(\Xi_0, \mathfrak{U}_c \times \mathfrak{U}_d)\) , according to Item (XIV) of Assumption 17 and Item (XXVI) of Assumption 25, denoting \( \xi_{o,j} = \xi_o(t_{j+1},j)\) , \( u_{d,j} = \mathfrak{u}_d(j)\) , and \( \tau_j = t_{j+1}-t_j\) , we have \( \xi_{o,j}\in \Xi_o\) for all \( j \in \mathbb{N}\) and whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{multline} |\mathcal{G}_j(\xi_{no,a}) - \mathcal{G}_j(\xi_{no,b})|=|\mathcal{G}^{\mathfrak{u}_c}((\xi_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,a}) - \mathcal{G}^{\mathfrak{u}_c}((\xi_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,b})|\\ \geq \frac{1}{c_{\mathcal{G}^{-1}}} |\xi_{no,a} - \xi_{no,b}|, \forall j \in \mathbb{N}_{\geq j_m}, \forall (\xi_{no,a},\xi_{no,b}) \in \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_{no}}. \end{multline} \]

(473)

Thanks to Item (XXVIII) of Assumption 26, applying Lemma 33 in Section 6.2.2.6 with \( \mathcal{F}\) therein being \( \mathcal{G}^{\mathfrak{u}_c}\) , the sequences \( (p_j)_{j \in \mathbb{N}}\) and \( (u_j)_{j \in \mathbb{N}}\) therein being \( (\xi_{o,j})_{j \in \mathbb{N}_{\geq j_m}}\) and \( (u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}_{\geq j_m}}\) respectively, \( x\) therein being \( \xi_{no}\) , the scalar \( c\) therein being \( \frac{1}{c_{\mathcal{G}^{-1}}}\) , and the scalar \( c^\prime\) therein being any \( 0<\frac{1}{c_{\hat{\mathcal{G}}^{-1}}} < \frac{1}{c_{\mathcal{G}^{-1}}}\) , we deduce that there exists \( \overline{c}_{o,1} > 0\) such that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_{o,1}\) for all \( j\in\mathbb{N}_{\geq j_m}\) , then whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{multline} |\hat{\mathcal{G}}_j(\xi_{no,a}) - \hat{\mathcal{G}}_j(\xi_{no,b})| = |\mathcal{G}^{\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,a}) - \mathcal{G}^{\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,b})|\\ \geq \frac{1}{c_{\hat{\mathcal{G}}^{-1}}}|\xi_{no,a} - \xi_{no,b}|, \forall j \in \mathbb{N}_{\geq j_m}, \forall (\xi_{no,a},\xi_{no,b}) \in \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_{no}}, \end{multline} \]

(474)

which implies that \( (\hat{\mathcal{G}}_j)_{j \in \mathbb{N}}\) is uniformly invertible on \( \mathbb{R}^{n_{no}}\) after \( j_m\) and that, whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{multline} |\hat{\mathcal{G}}_j^{-1}(\xi_{no,a}) - \hat{\mathcal{G}}_j^{-1}(\xi_{no,b})| = |\mathcal{G}^{-1,\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,a}) - \mathcal{G}^{-1,\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,b})|\\ \leq c_{\hat{\mathcal{G}}^{-1}}|\xi_{no,a} - \xi_{no,b}|, ~~ \forall j \in \mathbb{N}_{\geq j_m}, \forall (\xi_{no,a},\xi_{no,b}) \in \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_{no}}. \end{multline} \]

(475)

According to Item (XXVII) of Assumption 25, we have, whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{multline} |\mathcal{O}^{bw}_j(\xi_{no,a}) - \mathcal{O}^{bw}_j(\xi_{no,b})|\\ = |\mathcal{O}^{bw,\mathfrak{u}_c}((\xi_{o,j-1},u_{d,j-1},t_{j-1},\tau_{j-1}),\ldots,(\xi_{o,j-\overline{m}},u_{d,j-\overline{m}},t_{j-\overline{m}},\tau_{j-\overline{m}}),\xi_{no,a}) \\ - \mathcal{O}^{bw,\mathfrak{u}_c}((\xi_{o,j-1},u_{d,j-1},t_{j-1},\tau_{j-1}),\ldots,(\xi_{o,j-\overline{m}},u_{d,j-\overline{m}},t_{j-\overline{m}},\tau_{j-\overline{m}}),\xi_{no,b})| \\ \geq \alpha |\xi_{no,a} - \xi_{no,b}|, ~~ \forall j \in \mathbb{N}_{\geq j_m+\overline{m}}, \forall (\xi_{no,a},\xi_{no,b}) \in (\Xi_{no}+ \delta \mathbb{B}) \times (\Xi_{no}+ \delta \mathbb{B}). \end{multline} \]

(476)

Thanks to Item (XXIX) of Assumption 26, applying Lemma 33 in Section 6.2.2.6 with \( \mathcal{F}\) therein being \( \mathcal{O}^{bw,\mathfrak{u}_c}\) , the sequence \( (p_j)_{j \in \mathbb{N}}\) therein being \( (\xi_{o,j-1},\xi_{o,j-2},…,\xi_{o,j-\overline{m}})_{j \in \mathbb{N}_{\geq j_m+\overline{m}}}\) , the sequence \( (u_j)_{j \in \mathbb{N}}\) therein being

\[ (u_{d,j-1},t_{j-1},\tau_{j-1},u_{d,j-2},t_{j-2},\tau_{j-2},\ldots,u_{d,j-\overline{m}},t_{j-\overline{m}},\tau_{j-\overline{m}})_{j \in \mathbb{N}_{\geq j_m+\overline{m}}}, \]

\( x\) therein being \( \xi_{no}\) , the scalar \( c\) therein being \( \alpha\) , and the scalar \( c^\prime\) therein being any \( 0 < \alpha^\prime < \alpha\) , we deduce that there exists \( \overline{c}_{o,2} > 0\) such that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_{o,2}\) for all \( j\in\mathbb{N}_{\geq j_m+ \overline{m}}\) , then whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{multline} |\hat{\mathcal{O}}^{bw}_j(\xi_{no,a}) - \hat{\mathcal{O}}^{bw}_j(\xi_{no,b})|\\ = |\mathcal{O}^{bw,\mathfrak{u}_c}((\hat{\xi}_{o,j-1},u_{d,j-1},t_{j-1},\tau_{j-1}),\ldots,(\hat{\xi}_{o,j-\overline{m}},u_{d,j-\overline{m}},t_{j-\overline{m}},\tau_{j-\overline{m}}),\xi_{no,a})\\ - \mathcal{O}^{bw,\mathfrak{u}_c}((\hat{\xi}_{o,j-1},u_{d,j-1},t_{j-1},\tau_{j-1}),\ldots,(\hat{\xi}_{o,j-\overline{m}},u_{d,j-\overline{m}},t_{j-\overline{m}},\tau_{j-\overline{m}}),\xi_{no,b})| \\ \geq \alpha^\prime|\xi_{no,a} - \xi_{no,b}|, ~~ \forall j \in \mathbb{N}_{\geq j_m+\overline{m}}, \forall (\xi_{no,a},\xi_{no,b}) \in (\Xi_{no}+ \delta \mathbb{B}) \times (\Xi_{no}+ \delta \mathbb{B}). \end{multline} \]

(477)

Next, by the global Lipschitzness of \( f_{o,\rm{sat}}\) (resp., \( f_\rm{{sat}}\) ) with respect to \( \xi_o\) (resp., \( \xi\) ), uniformly with respect to \( u_c\) , Lemma 25 in Section 6.2.2.6 allows us to show that \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) (resp., \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,\tau)\) ) is Lipschitz, uniformly with respect to \( (t,\tau)\in\mathbb{R}_{\geq 0}\times[0,\tau_M]\) and \( \mathfrak{u}_c\in \mathfrak{U}_c\) . This together with the Lipschitzness of \( g\) from Assumption 22 means that \( \mathcal{G}^{\mathfrak{u}_c}\) is Lipschitz in \( \xi_o\) (the first component of \( u\) ) on the compact set \( \Xi_o\) , uniformly with respect to \( (u_d,t,\tau,\xi_{no})\in \mathcal{U}_d\times \mathbb{R}_{\geq 0}\times[0,\tau_M]\times \Xi_{no}\) and \( \mathfrak{u}_c\in \mathfrak{U}_c\) with some constant \( c_{\mathcal{G}} > 0\) . It follows that, along the considered solutions, for all \( j \in \mathbb{N}\) and for all \( \xi_{no} \in \Xi_{no}\) , we have

\[ |\mathcal{G}_j(\xi_{no}) - \hat{\mathcal{G}}_j(\xi_{no})| = |\mathcal{G}^{\mathfrak{u}_c}((\xi_{o,j},u_{d,j},t_j,\tau_j),\xi_{no}) - \mathcal{G}^{\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no})| \leq c_{\mathcal{G}}|\xi_{o,j} - \hat{\xi}_{o,j}|, \]

and thus we can say that if \( \mathcal{G}_j(\xi_{no})\in \Xi_{no}\) , then \( \hat{\mathcal{G}}_j(\xi_{no}) \in \Xi_{no} + c_{\mathcal{G}}|\xi_{o,j} - \hat{\xi}_{o,j}|\mathbb{B}\) . There then exists \( \overline{c}_{o,3} > 0\) such that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_{o,3}\) for all \( j\in\mathbb{N}_{\geq j_m}\) , then \( \hat{\mathcal{G}}_j(\xi_{no}) \in \Xi_{no} + \delta\mathbb{B}\) for all \( j\in\mathbb{N}_{\geq j_m}\) and for all \( \xi_{no} \in \Xi_{no}\) such that \( \mathcal{G}_j(\xi_{no})\in \Xi_{no}\) .
Pick \( \overline{c}_o = \min\{\overline{c}_{o,1},\overline{c}_{o,2},\overline{c}_{o,3}\}\) . It follows that if \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_o\) for all \( j\in\mathbb{N}_{\geq j_m}\) , then whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) , we get that: i) Item (XXVI) of Assumption 25 holds for \( (\hat{\mathcal{G}}_j)_{j \in \mathbb{N}_{\geq j_m}}\) in place of \( (\mathcal{G}_j)_{j \in \mathbb{N}_{\geq j_m}}\) with \( c_{\hat{\mathcal{G}}^{-1}}\) replacing \( c_{\mathcal{G}^{-1}}\) , ii) Item (XXVII) of Assumption 25 holds after jump \( j_m+\overline{m}\) for system (266) fed with \( (\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}_{\geq j_m+\overline{m}}}\) in place of \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}_{\geq j_m+\overline{m}}}\) with the same \( m_i\) , \( i \in \{1,2,…,n_{d,\text{ext}}\}\) , and with \( \alpha^\prime\) replacing \( \alpha\) , and iii) \( \hat{\mathcal{G}}_j(\xi_{no}) \in \Xi_{no} + \delta\mathbb{B}\) for all \( j\in\mathbb{N}_{\geq j_m}\) and for all \( \xi_{no} \in \Xi_{no}\) such that \( \mathcal{G}_j(\xi_{no})\in \Xi_{no}\) .

Then, from the uniform Lipschitzness of \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) as well as \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,\tau)\) ) shown above and that of \( (g_o,h_d)\) from Assumption 25, we deduce that there exists \( c_\mathcal{H}, c_{\hat{\mathcal{H}}} > 0\) such that whatever the inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,j},\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) considered above, we have for all \( j \in \mathbb{N}\) and for all \( (\xi_{no,a},\xi_{no,b}) \in \mathbb{R}^{n_{no}} \times \mathbb{R}^{n_{no}}\) ,

\[ \begin{align} |\mathcal{H}_j(\xi_{no,a}) - \mathcal{H}_j(\xi_{no,b})| &= |\mathcal{H}^{\mathfrak{u}_c}((\xi_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,a}) - \mathcal{H}^{\mathfrak{u}_c}((\xi_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,b})|\notag\\ & \leq c_\mathcal{H} |\xi_{no,a} - \xi_{no,b}|, \end{align} \]

(478.a)

\[ \begin{align} |\hat{\mathcal{H}}_j(\xi_{no,a}) - \hat{\mathcal{H}}_j(\xi_{no,b})| &= |\mathcal{H}^{\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,a}) - \mathcal{H}^{\mathfrak{u}_c}((\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j),\xi_{no,b})|\notag\\ & \leq c_{\hat{\mathcal{H}}} |\xi_{no,a} - \xi_{no,b}|. \end{align} \]

(478.b)

Applying Theorem 4 and Theorem 5 starting from discrete time \( j_m\) to system (266) fed with \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) , which verifies i) the uniform invertibility of \( \mathcal{G}_k\) and uniform Lipschitzness of \( \mathcal{G}_k^{-1}\) and \( \mathcal{H}_k\) as assumed in Item (XXVI) of Assumption 25 and obtained from (478.a), and ii) the uniform Lipschitz backward distinguishability condition as assumed in Item (XXVII) of Assumption 25, we deduce that given any \( j_m^\star \in \mathbb{N}_{\geq j_m + \overline{m}}\) and any \( T_{j_m}\) globally Lipschitz, there exists \( 0 < \gamma^\star_1 \leq 1\) and \( c_T(\cdot)>0\) such that for any \( 0 < \gamma < \gamma^\star_1\) , and for any inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) considered above, there exists \( (T_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) satisfying (282) with \( j_m\) as the new initial time with \( (A(\gamma),B)\) of the form (286), that is besides uniformly Lipschitz injective on \( \Xi_{no}\) for all \( j \in \mathbb{N}_{\geq j_m^\star}\) with constant \( c_T(\gamma)\) , i.e., such that

\[ |T_{\gamma,j}(\xi_{no,a}) - T_{\gamma,j}(\xi_{no,b})| \geq c_T(\gamma) |\xi_{no,a}-\xi_{no,b}|, ~~ \forall j\in \mathbb{N}_{\geq j_m^\star}, \forall (\xi_{no,a},\xi_{no,b})\in \Xi_{no}\times \Xi_{no}. \]

Applying [40] after \( j_m^\star\) , we deduce that there exists a uniformly Lipschitz sequence \( (T_{\gamma,j}^*)_{j \in \mathbb{N}_{\geq j_m^\star}}\) of maps defined on \( \mathbb{R}^{n_\zeta}\) that are left inverses of \( (T_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m^\star}}\) in the sense of (281). More precisely, there exists \( c_{T^*}(\gamma)>0\) , depending only on \( c_T(\gamma)\) , the norms and the dimension \( n_{no}\) , such that

\[ \begin{equation} |T_{\gamma,j}^*(\zeta_a) - T_{\gamma,j}^*(\zeta_b)| \leq c_{T^*}(\gamma)|\zeta_a - \zeta_b|, ~~ \forall j \in \mathbb{N}_{\geq j_m^\star},\forall (\zeta_a,\zeta_b) \in \mathbb{R}^{n_\zeta} \times \mathbb{R}^{n_\zeta}. \end{equation} \]

(479)

We pick then arbitrarily \( (T_{\gamma,j}^*)_{j \in \mathbb{N}_{< j_m^\star}}\) .

Next, we reproduce the same reasoning, applying Theorem 4 and Theorem 5 to system (266) but this time fed with \( (\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) where \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_o\) for all \( j\in\mathbb{N}_{\geq j_m}\) , which verifies i) the uniform invertibility and uniform Lipschitzness conditions as obtained from (475) and (478.b), and ii) the uniform Lipschitz backward distinguishability condition as obtained from (477). We then deduce that for the same \( j_m^\star\) and the same \( T_{j_m}\) as above, there exists \( 0 < \gamma^\star_2 \leq 1\) and \( c_{\hat{T}^*}(\cdot) > 0\) such that for any \( 0 < \gamma < \gamma^\star_2\) , with \( (A(\gamma),B)\) of the form (286) (with the same choice of \( (\tilde{A}_i,\tilde{B}_i)\) , \( i \in \{1,2,…,n_{d,\text{ext}}\}\) , as above), for all considered \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) , there exists \( (\hat{T}_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) satisfying

\[ \begin{multline} \hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no})) = A(\gamma)\hat{T}_{\gamma,j}(\xi_{no}) + B_d\hat{\mathcal{H}}_{d,j}(\xi_{no})+ B_o \hat{\mathcal{H}}_{o,j}(\xi_{no}),\\ \forall j \in \mathbb{N}_{\geq j_m}, \forall \xi_{no} \in \Xi_{no}+\delta\mathbb{B}: \hat{\mathcal{G}}_j(\xi_{no}) \in \Xi_{no}+\delta\mathbb{B}, \end{multline} \]

(480)

which is uniformly Lipschitz injective on \( \Xi_{no}+\delta\mathbb{B}\) after jump \( j_m^\star\) , and there exists \( (\hat{T}^*_{\gamma,j})_{j \in \mathbb{N}}\) defined on \( \mathbb{R}^{n_\zeta}\) (arbitrarily picked before jump \( j_m^\star\) as done with \( (T_{\gamma,j}^*)_{j \in \mathbb{N}_{< j_m^\star}}\) ) satisfying

\[ \begin{align} \hat{T}_{\gamma,j}^*(\hat{T}_{\gamma,j}(\xi_{no})) &= \xi_{no}, &\forall j& \in \mathbb{N}_{\geq j_m^\star},\forall \xi_{no} \in \Xi_{no}+\delta\mathbb{B}, \end{align} \]

(481.a)

\[ \begin{align} |\hat{T}_{\gamma,j}^*(\zeta_a) - \hat{T}_{\gamma,j}^*(\zeta_b)| &\leq c_{\hat{T}^*}(\gamma)|\zeta_a - \zeta_b|, & \forall j &\in \mathbb{N}_{\geq j_m^\star},\forall (\zeta_a,\zeta_b) \in \mathbb{R}^{n_\zeta} \times \mathbb{R}^{n_\zeta}. \end{align} \]

(481.b)

Indeed, notice that (480), (481.a), and (481.b) are respectively (282), (281), and (479), but with \( (\hat{\xi}_{o,j})_{j \in \mathbb{N}}\) replacing \( (\xi_{o,j})_{j \in \mathbb{N}}\) in the inputs, and with \( \Xi_{no}+ \delta \mathbb{B}\) replacing \( \Xi_{no}\) .

Now we pick any \( 0 < \gamma < \min\{\gamma^\star_1,\gamma^\star_2\}\) such that \( \|A(\gamma)\| < 1\) , fixing \( A(\gamma)\) (Schur and invertible), \( c_{T^*}(\gamma)\) and \( c_{\hat{T}^*}(\gamma)\) so that we satisfy all (282), (281), (479), (480), (481.a), and (481.b). The sequences of left inverse maps \( (T^*_{\gamma,j})_{j \in \mathbb{N}}\) and \( (\hat{T}^*_{\gamma,j})_{j \in \mathbb{N}}\) will respectively play the role of \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) and \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) in Assumption 23.

Now, the map sequences are ready and we check each item of Assumption 23. First, observe that by construction from (480), any solution, remaining in \( \Xi_{no}\) , to system (266) fed with \( (\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j \in \mathbb{N}}\) where \( |\xi_{o,j} - \hat{\xi}_{o,j}| \leq \overline{c}_o\) for all \( j\in\mathbb{N}_{\geq j_m}\) , is such that \( j \mapsto \hat{T}_{\gamma,j}(\xi_{no,j})\) is solution, after time \( j_m\) , to system (268) with the chosen \( (A(\gamma),B)\) and with initialization as \( \zeta_{j_m}=\hat{T}_{\gamma,j_m}(\xi_{no,j_m})\) . Since \( A(\gamma)\) is invertible, there exists \( \zeta_0\) such that the solution \( j \mapsto \zeta_j\) to system (268) initialized as \( \zeta_0\) verifies \( \zeta_{j_m} = \hat{T}_{\gamma,j_m}(\xi_{no,j_m})\) and thus \( \zeta_j= \hat{T}_{\gamma,j}(\xi_{no,j})\) for all \( j \in \mathbb{N}_{\geq j_m}\) , so we get (270) from (481.a). Moreover, thanks to \( j \mapsto \xi_{no,j}\) remaining in the compact set \( \Xi_{no}\) and the uniform Lipschitzness of \( (\hat{T}_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) obtained from the proof of Theorem 6 with Lipschitz constant depending only on \( \gamma\) , the Lipschitz constants of \( T_{j_m}\) and of \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) , \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c)}\) , \( g_o\) , and \( h_d\) as shown above and assumed in Assumption 25. Since \( \zeta_j= \hat{T}_{\gamma,j}(\xi_{no,j})\) for all \( j \in \mathbb{N}_{\geq j_m}\) , we deduce that there exists a compact set \( \Xi_{\zeta,1} \subset \mathbb{R}^{n_\eta}\) (independent of solutions) such that \( \zeta_j \in \Xi_{\zeta,1}\) for all \( j \in \mathbb{N}_{\geq j_m}\) . Next, because \( \xi_{no,j}\) remains in the compact set \( \Xi_{no}\) and thanks to the uniform boundedness properties of \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) , \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,\tau)\) ) shown above and that of \( (g,h_d)\) from Assumption 22, the output \( (y_j)_{j\in \mathbb{N}}\) to system (266) is uniformly bounded under all considered inputs, and by the invertibility of the dynamics (268) since \( A(\gamma)\) is invertible, we deduce that there exists a compact set \( \Xi_{\zeta,2} \subset \mathbb{R}^{n_\eta}\) (independent of solutions) such that \( \zeta_j \in \Xi_{\zeta,2}\) for all \( j \in \{0,1,…,j_m-2\}\) . It follows that \( \zeta_j \in \Xi_{\zeta,1} \cup \Xi_{\zeta,2}\) for all \( j \in \mathbb{N}\) , and this compact is independent of solutions. It follows that Item (XXIII) of Assumption 23 holds. Second, Item (XXIV) follows from (481.b). Now pick inputs \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,j},\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) as considered above and let us check the last Item (XXV). Consider a solution \( j \mapsto \xi_{no,j}\) to system (266) fed with \( \mathfrak{u}_c \in \mathfrak{U}_c\) and \( (\xi_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) and remaining in \( \Xi_{no}\) . Thanks to Item (XXVI) of Assumption 25 and (282), the corresponding \( (T_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) verifies

\[ \begin{equation} T_{\gamma,j+1}(\xi_{no,j+1}) = A(\gamma) T_{\gamma,j}(\xi_{no,j}) + B_d\mathcal{H}_{d,j}(\xi_{no,j}) + B_o \mathcal{H}_{o,j}(\xi_{no,j}), ~~ \forall j \in \mathbb{N}_{\geq j_m}. \end{equation} \]

(482)

On the other hand, because \( \xi_{no,j} \in \Xi_{no}\) and \( \mathcal{G}_j(\xi_{no,j})= \xi_{no,j+1}\in \Xi_{no}\) , we get \( \hat{\mathcal{G}}_j(\xi_{no,j}) \in \Xi_{no} + \delta\mathbb{B}\) for all \( j\in\mathbb{N}_{\geq j_m}\) . So, the map sequence \( (\hat{T}_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) verifies for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{align*} \hat{T}_{\gamma,j+1}(\xi_{no,j+1}) & = \hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no,j})) + \hat{T}_{\gamma,j+1}(\mathcal{G}_j(\xi_{no,j})) - \hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no,j}))\\ & = A(\gamma) \hat{T}_{\gamma,j}(\xi_{no,j})+ B_d\hat{\mathcal{H}}_{d,j}(\xi_{no,j}) + B_o \hat{\mathcal{H}}_{o,j}(\xi_{no,j})\\ &~~{}+ \hat{T}_{\gamma,j+1}(\mathcal{G}_j(\xi_{no,j})) - \hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no,j})). \end{align*} \]

We then get that, for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{multline} T_{\gamma,j+1}(\xi_{no,j+1})-\hat{T}_{\gamma,j+1}(\xi_{no,j+1}) = A(\gamma)(T_{\gamma,j}(\xi_{no,j})-\hat{T}_{\gamma,j}(\xi_{no,j})) + B_d(\mathcal{H}_{d,j}(\xi_{no,j})-\hat{\mathcal{H}}_{d,j}(\xi_{no,j}))\\ + B_o(\mathcal{H}_{o,j}(\xi_{no,j}) - \hat{\mathcal{H}}_{o,j}(\xi_{no,j}))\\ + \hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no,j}))-\hat{T}_{\gamma,j+1}(\mathcal{G}_j(\xi_{no,j})). \end{multline} \]

(483)

From Assumption 22 and the uniform Lipschitzness of \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) shown above, we deduce that there exist \( c_{\mathcal{H}_d},c_{\mathcal{H}_o},c_{\mathcal{G}} > 0\) , independent of the considered inputs \( \mathfrak{u}_c\) and \( (\xi_{o,j},\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) , such that for all \( j \in \mathbb{N}_{\geq j_m}\) and for all \( \xi_{no} \in \Xi_{no}\) ,

\[ \begin{align*} |\mathcal{H}_{d,j}(\xi_{no}) - \hat{\mathcal{H}}_{d,j}(\xi_{no})| &\leq c_{\mathcal{H}_d} |\xi_{o,j} - \hat{\xi}_{o,j}|, \\ |\mathcal{H}_{o,j}(\xi_{no}) - \hat{\mathcal{H}}_{o,j}(\xi_{no})| &\leq c_{\mathcal{H}_o} |\xi_{o,j} - \hat{\xi}_{o,j}|,\\ |\mathcal{G}_j(\xi_{no}) - \hat{\mathcal{G}}_j(\xi_{no})| &\leq c_{\mathcal{G}} |\xi_{o,j} - \hat{\xi}_{o,j}|. \end{align*} \]

Besides, from the proof of Theorem 6, we have that \( (\hat{T}_{\gamma,j})_{j \in \mathbb{N}_{\geq j_m}}\) is uniformly Lipschitz (with some Lipschitz constant depending only on \( \gamma\) and the Lipschitz constants of \( T_{j_m}\) , \( \Psi_{f_{o,\rm{sat}}(\cdot, \mathfrak{u}_c)}(\cdot,t,-\tau)\) , \( \Psi_{f_\rm{{sat}}(\cdot, \mathfrak{u}_c)}\) , \( g\) , and \( h_d\) as mentioned above). From this, we deduce that there exists \( c_{\hat{T}}^\prime > 0\) , still independent of the considered inputs \( \mathfrak{u}_c\) and \( (\xi_{o,j},\hat{\xi}_{o,j},u_{d,j},t_j,\tau_j)_{j\in \mathbb{N}}\) , such that for all \( j \in \mathbb{N}_{\geq j_m}\) and for all \( \xi_{no} \in \Xi_{no}\) ,

\[ |\hat{T}_{\gamma,j+1}(\hat{\mathcal{G}}_j(\xi_{no}))-\hat{T}_{\gamma,j+1}(\mathcal{G}_j(\xi_{no}))| \leq c_{\hat{T}}^\prime |\xi_{o,j} - \hat{\xi}_{o,j}|. \]

Combining these, we get from (483) that there exists \( c > 0\) , independent of inputs, such that for all \( j \in \mathbb{N}_{\geq j_m}\) ,

\[ \begin{equation} |T_{\gamma,j+1}(\xi_{no,j+1})-\hat{T}_{\gamma,j+1}(\xi_{no,j+1})| \leq \|A(\gamma)\| |T_{\gamma,j}(\xi_{no,j})-\hat{T}_{\gamma,j}(\xi_{no,j})| + c|\xi_{o,j} - \hat{\xi}_{o,j}|. \end{equation} \]

(484)

For any \( \Gamma_0 > 0\) , \( A^\prime = \|A(\gamma)\|\in (0,1)\) , \( B^\prime = c\) , and \( n_\eta = 1\) , because both \( (T_{\gamma,j})_{j\in\mathbb{N}_{\geq j_m}}\) and \( (\hat{T}_{\gamma,j})_{j\in\mathbb{N}_{\geq j_m}}\) are initialized as \( T_{j_m}\) at initial time \( j_m\) , we deduce that \( j \mapsto \Gamma_j\) with dynamics (273) initialized as \( \Gamma_0\) is such that whatever \( \mathfrak{u}_c \in \mathfrak{U}_c\) ,

\[ \begin{equation} |T_{\gamma,j}(\xi_{no,j})-\hat{T}_{\gamma,j}(\xi_{no,j})| \leq \Gamma_j, ~~ \forall j \in \mathbb{N}_{\geq j_m}. \end{equation} \]

(485)

Next, consider \( j \mapsto \zeta_j\) solutions to system (268) remaining in \( \Xi_{\zeta}\) for all \( j \in \mathbb{N}\) , such that \( \zeta_j = T_{\gamma,j}(\xi_{no,j})\) for all \( j \in \mathbb{N}_{\geq j_m^\star}\) , as given by Item (XXIII) of Assumption 23 that we proved above. Using (281), (481.a), and (481.b) respectively, we have for all \( j \in \mathbb{N}_{\geq j_m^\star}\) ,

\[ \begin{align*} |T_{\gamma,j}^*(\zeta_j)-\hat{T}_{\gamma,j}^*(\zeta_j)| & = |T_{\gamma,j}^*(T_{\gamma,j}(\xi_{no,j}))-\hat{T}_{\gamma,j}^*(T_{\gamma,j}(\xi_{no,j}))|\\ & = |\xi_{no,j}-\hat{T}_{\gamma,j}^*(T_{\gamma,j}(\xi_{no,j}))|\\ & = |\hat{T}_{\gamma,j}^*(\hat{T}_{\gamma,j}(\xi_{no,j}))-\hat{T}_{\gamma,j}^*(T_{\gamma,j}(\xi_{no,j}))|\\ & \leq c_{\hat{T}^*}(\gamma)|T_{\gamma,j}(\xi_{no,j})-\hat{T}_{\gamma,j}(\xi_{no,j})|, \end{align*} \]

which, combined with (485), gives us

\[ |T_{\gamma,j}^*(\zeta_j)-\hat{T}_{\gamma,j}^*(\zeta_j)| \leq c_{\hat{T}^*}(\gamma)\Gamma_j, ~~ \forall j \in \mathbb{N}_{\geq j_m^\star}, \]

which corresponds to (272) with \( (T^*_{\gamma,j})_{j \in \mathbb{N}}\) (resp., \( (\hat{T}^*_{\gamma,j})_{j \in \mathbb{N}}\) ) playing the role of \( (\mathcal{T}_j)_{j \in \mathbb{N}}\) (resp., \( (\hat{\mathcal{T}}_j)_{j \in \mathbb{N}}\) ). Therefore, Item (XXV) of Assumption 23 follows, concluding the proof.

6.2.2.6 Technical Lemmas of Section 3.3

Lemma 25 (Uniform Lipschitzness of \( \Psi_{f(\cdot,\mathfrak{u})}(\cdot,t,-\tau)\))

Consider \( \mathfrak{U}\) , a subspace of locally bounded functions \( \mathfrak{u} : \mathbb{R}_{\geq 0}\to\mathcal{U}\subset \mathbb{R}^{n_u}\) , a continuous map \( f:\mathbb{R}^{n_\xi}\times \mathbb{R}^{n_u}\to \mathbb{R}^{n_\xi}\) , and the continuous-time dynamics

\[ \begin{equation} \dot{\xi} = f(\xi,u), \end{equation} \]

(486)

with state \( \xi \in \mathbb{R}^{n_\xi}\) and input \( \mathfrak{u} \in \mathfrak{U}\) . Let \( \tau_M > 0\) . Assume that \( f\) is globally Lipschitz with respect to \( \xi\) , uniformly in \( u\in\mathcal{U}\) . Then, for any \( \mathfrak{u}\in \mathfrak{U}\) and for any \( (t,\tau) \in \mathbb{R}_{\geq 0} \times [0,\tau_M]\) with \( \tau\leq t\) , the map \( \Psi_{f(\cdot,\mathfrak{u})}(\cdot,t,-\tau)\) is defined on \( \mathbb{R}^{n_\xi}\) and is globally Lipschitz, uniformly in \( (t,\tau) \in \mathbb{R}_{\geq 0} \times [0,\tau_M]\) and in \( \mathfrak{u} \in \mathfrak{U}\) .

Proof. By assumption, there exists \( L>0\) such that \( |f(\xi_a,u)-f(\xi_b,u)|\leq L |\xi_a-\xi_b|\) for all \( (\xi_a,\xi_b) \in \mathbb{R}^{n_\xi} \times \mathbb{R}^{n_\xi}\) and for all \( u\in \mathcal{U}\) . From [197, Proposition C.3.8], we deduce that the map \( \Psi_{f(\cdot,\mathfrak{u})}(\cdot,t,-\tau)\) is defined on \( \mathbb{R}^{n_\xi}\) , for any \( \mathfrak{u}\in \mathfrak{U}\) and for any \( (t,\tau) \in \mathbb{R}_{\geq 0} \times [0,\tau_M]\) with \( \tau\leq t\) .

Let any \( \mathfrak{u}\in \mathfrak{U}\) . For any \( (\xi_a,\xi_b,t,\tau) \in \mathbb{R}^{n_\xi} \times \mathbb{R}^{n_\xi} \times \mathbb{R}_{\geq 0} \times [0,\tau_M]\) with \( \tau\leq t\) , we have

\[ \begin{align*} \Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,-\tau) &= \xi_a + \int_{t}^{t-\tau}f(\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,s-t),\mathfrak{u}(s))ds,\\ \Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,-\tau) &= \xi_b + \int_{t}^{t-\tau}f(\Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,s-t),\mathfrak{u}(s))ds. \end{align*} \]

and by subtracting both sides and using the triangle inequality,

\[ \begin{multline*} |\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,-\tau) - \Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,-\tau)|\leq |\xi_a - \xi_b|\\ + \int_{t-\tau}^{t}|f(\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,s-t),\mathfrak{u}(s))-f(\Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,s-t),\mathfrak{u}(s))|ds, \end{multline*} \]

so that, since \( \mathfrak{u}(s)\in \mathcal{U}\) for all \( s\in [t-\tau,t]\subset\mathbb{R}_{\geq 0}\) ,

\[ |\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,-\tau) - \Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,-\tau)| \leq |\xi_a - \xi_b| + \int_{t-\tau}^{t}L|\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,s-t)-\Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,s-t)|ds. \]

Using Grönwall’s inequality, we get

\[ \begin{align*} |\Psi_{f(\cdot,\mathfrak{u})}(\xi_a,t,-\tau) - \Psi_{f(\cdot,\mathfrak{u})}(\xi_b,t,-\tau)| &\leq |\xi_a -\xi_b|e^{L\tau}\\ &\leq |\xi_a -\xi_b|e^{L\tau_M}, \end{align*} \]

because \( \tau \in [0,\tau_M]\) , which concludes the proof. \( \blacksquare\)

Lemma 26 (An expression of \( \frac{d\Psi_{f}}{dt} (\hat{\xi}(t),t,-\tau(t))\))

Consider a time-varying continuous-time system

\[ \begin{equation} \dot{\xi} = f(\xi,t), \end{equation} \]

(487)

with \( f\) being continuous and of class \( C^1\) with respect to \( \xi\) . For any differentiable time function \( t\mapsto \hat{\xi}(t)\) and a modified time \( t\mapsto \tau(t)\) such that \( \dot{\tau}(t)=1\) , we have, at any time \( t\) where \( \Psi_f(\hat{\xi}(t),t,-\tau(t))\) is defined,

\[ \begin{equation} \frac{d\Psi_{f}}{dt} (\hat{\xi}(t),t,-\tau(t)) = -\frac{\partial \Psi_{f}}{\partial \xi}(\hat{\xi}(t),t,-\tau(t)) (f(\hat{\xi}(t),t) - \dot{\hat{\xi}}(t)). \end{equation} \]

(488)

Proof. First, from [207, Chapter 1, Theorem 3.3], since \( f\) is continuous and \( C^1\) with respect to \( \xi\) , we get for any \( (\xi,t)\) and any \( \tau\) such that \( \Psi_f(\xi,t,\tau)\) is defined

\[ \begin{equation} f(\Psi_f(\xi,t,\tau),t+\tau) = \frac{\partial \Psi_f}{\partial \xi}(\xi,t,\tau) f(\xi,t) + \frac{\partial \Psi_f}{\partial t}(\xi,t,\tau). \end{equation} \]

(489)

Using (489) with \( (\xi,t,\tau)\) replaced by \( (\hat{\xi}(t),t,-\tau(t))\) , we then have

\[ \begin{align*} \frac{d\Psi_f}{dt} (\hat{\xi}(t),t,-\tau(t)) &= \frac{\partial \Psi_f}{\partial \xi}(\hat{\xi}(t),t,-\tau(t)) \dot{\hat{\xi}}(t)+ \frac{\partial \Psi_f}{\partial t}(\hat{\xi}(t),t,-\tau(t)) - f(\Psi_f(\hat{\xi}(t),t,-\tau(t)),t-\tau(t)) \\ & = -\frac{\partial \Psi_f}{\partial \xi}(\hat{\xi}(t),t,-\tau(t)) \left(f(\hat{\xi}(t),t) - \dot{\hat{\xi}}(t)\right). \end{align*} \]

The proof is completed. \( \blacksquare\)

Lemma 27 (Dynamics of \( \frac{\partial \Psi_f}{\partial \xi_0}(\xi_0, t_0,\tau)\))

This lemma is from [207, Chapter 1, Theorem 3.3]. Consider the time-varying system

\[ \begin{equation} \dot{\xi} = f(\xi,t), \end{equation} \]

(490)

where \( f\) is continuous and of class \( C^1\) with respect to \( \xi\) and define \( \Psi_f(\xi_0,t_0,\tau)\) as the solution to system (490) initialized as \( \xi_0\) at time \( t_0\) and flowing during \( \tau\) time unit(s), and a modified time \( t\mapsto \tau(t)\) such that \( \dot{\tau}=1\) . Let \( \Phi_f(\xi_0,t_0,\tau) = \frac{\partial \Psi_f}{\partial \xi_0}(\xi_0, t_0,\tau)\) . Then \( (\xi,\Phi_f)\) is solution to the dynamics

\[ \begin{equation} \dot{\xi} = f(\xi,t),~~ \dot{\Phi}_f = \frac{\partial f}{\partial \xi}(\xi,t) \Phi_f, \end{equation} \]

(491)

initialized as \( (\xi_0, \rm{Id})\) .

Lemma 28 (Projection\( /\) elimination lemma)

This lemma is from [208, Theorem 2.3.12]. There exist matrices \( B\) , \( C\) , and \( P = P^\top\) such that

\[ \begin{equation} \left\{ \begin{array}{@{}l@{~}l@{~}l@{}} B^{\bot} P (B^\bot)^\top < 0 &\text{ or }& B B^\top > 0 \\ (C^\top)^\bot P ((C^\top)^\bot)^\top < 0 &\text{ or }& C^\top C > 0 \end{array} \right. \end{equation} \]

(492)

if and only if there exists a matrix \( Y\) such that

\[ \begin{equation} B Y C + (B Y C)^\top + P < 0. \end{equation} \]

(493)

Lemma 29 (Transforming a matrix inequality into a Linear Matrix Inequality[LMI])

Assume that there exist \( a > 0\) , \( Q = Q^\top > 0\) , and \( \rho \mapsto L(\rho)\) such that, for some set \( ¶\) and some \( \rho \mapsto (J(\rho),H(\rho))\) , we have

\[ \begin{equation} (J(\rho) - L(\rho)H(\rho))^\top Q (J(\rho) - L(\rho)H(\rho)) - aQ<0, ~~ \forall \rho \in ¶. \end{equation} \]

(494)

Then, the same \( Q\) and \( a\) are solutions to

\[ \begin{equation} \begin{pmatrix}((H(\rho))^\bot)^\top Q(H(\rho))^\bot & \star \\ Q J(\rho)(H(\rho))^\bot & aQ \end{pmatrix} > 0, ~~ \forall \rho \in ¶. \end{equation} \]

(495)

Proof. Using Schur’s lemma at each \( \rho \in ¶\) , we deduce that (494) is equivalent to

\[ \begin{equation} \begin{pmatrix}Q& \star \\ Q J(\rho) - QL(\rho)H(\rho) & aQ \end{pmatrix} >0, ~~ \forall \rho \in ¶. \end{equation} \]

(496)

We introduce the variable \( Y(\rho) = -Q L(\rho)\) for \( \rho \in ¶\) and rewrite (496) as

\[ \begin{equation} \begin{pmatrix} 0 \\ \rm{Id} \end{pmatrix} Y(\rho) \begin{pmatrix} H(\rho) & 0 \end{pmatrix} + \begin{pmatrix} (H(\rho))^\top \\ 0 \end{pmatrix} (Y(\rho))^\top \begin{pmatrix} 0 & \rm{Id} \end{pmatrix} + \begin{pmatrix} Q & \star \\ QJ(\rho) & aQ \end{pmatrix} > 0,~~ \forall \rho \in ¶. \end{equation} \]

(497)

Now apply Lemma 28 to (497) at each \( \rho \in ¶\) with \( B = \begin{pmatrix} 0 \\ \rm{Id} \end{pmatrix}\) , \( C(\rho) = \begin{pmatrix} H(\rho) & 0 \end{pmatrix}\) , and \( P(\rho) = -\begin{pmatrix} Q & \star \\ QJ(\rho) & aQ \end{pmatrix}\) . Since \( BB^\top = \begin{pmatrix} 0 \\ \rm{Id} \end{pmatrix}\begin{pmatrix} 0 & \rm{Id} \end{pmatrix} = \begin{pmatrix} 0 & 0 \\ 0 & \rm{Id} \end{pmatrix}\) and \( (C(\rho))^\top C(\rho) = \begin{pmatrix} (H(\rho))^\top \\ 0 \end{pmatrix}\begin{pmatrix} H(\rho) & 0 \end{pmatrix} = \begin{pmatrix} (H(\rho))^\top H(\rho) & 0\\ 0&0 \end{pmatrix}\) are not positive definite (for all \( \rho \in ¶\) ), we deduce that (497) is equivalent to

\[ \begin{equation} \begin{pmatrix} \rm{Id} & 0 \end{pmatrix} \begin{pmatrix} Q & \star \\ QJ(\rho) & aQ \end{pmatrix} \begin{pmatrix} \rm{Id} \\ 0 \end{pmatrix} > 0, ~~ \begin{pmatrix} ((H(\rho))^\bot)^\top & 0 \\ 0 & \rm{Id} \end{pmatrix} \begin{pmatrix} Q & \star \\ QJ(\rho) & aQ \end{pmatrix} \begin{pmatrix} (H(\rho))^\bot & 0 \\ 0 & \rm{Id} \end{pmatrix} > 0, \end{equation} \]

(498)

for all \( \rho \in ¶\) . While the first condition becomes \( Q > 0\) (for all \( \rho \in ¶\) ) which is trivial, the second one gives us (495). \( \blacksquare\)

Lemma 30 (Transferring invertibility from matrix to matrix)

Let \( \mathcal{U}\) be a set, \( \mathcal{C}\subset \mathbb{R}^{n_\xi}\) be a compact set, and \( \mathcal{C}_s\) be a subset of \( \mathcal{C}^{\mathbb{N}}\) . Consider \( M:\mathbb{R}^{n_\xi}\times \mathcal{U} \to \mathbb{R}^{m\times m}\) a matrix-valued function that is both locally bounded and locally Lipschitz with respect to \( \xi\) , uniformly in \( u \in \mathcal{U}\) . Assume that there exists \( c > 0\) such that for all \( (\xi_k)_{k \in \mathbb{N}} \in \mathcal{C}_s\) and for all \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , we have

\[ \begin{equation} (M(\xi_k,u_k))^\top M(\xi_k,u_k) \geq c \rm{Id}, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(499)

Then, for any \( 0 < c^\prime < c\) , there exists \( \overline{c} > 0\) such that for any \( (\hat{\xi}_k)_{k \in \mathbb{N}}\) for which

\[ \begin{equation} \exists (\xi_k)_{k \in \mathbb{N}} \in \mathcal{C}_s: |\xi_k - \hat{\xi}_k| \leq \overline{c}, ~~ \forall k \in \mathbb{N}, \end{equation} \]

(500)

and for any \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , we have

\[ \begin{equation} (M(\hat{\xi}_k,u_k))^\top M(\hat{\xi}_k,u_k) \geq c^\prime \rm{Id}, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(501)

Proof. Pick \( 0 < c^\prime < c\) and some \( \delta > 0\) . Since \( M\) is locally bounded with respect to \( \xi\) , uniformly in \( u \in \mathcal{U}\) , let \( c_M > 0\) be the bound of \( \|M(\xi,u)\|\) on \( \mathcal{C} \times \mathcal{U}\) . Moreover, since \( M\) is locally Lipschitz with respect to \( \xi\) , uniformly in \( u \in \mathcal{U}\) , let \( L_M > 0\) be its Lipschitz constant on \( (\mathcal{C} + \delta\mathbb{B}) \times \mathcal{U}\) . Then, given any sequences \( (\hat{\xi}_k)_{k \in \mathbb{N}}\) , \( (\hat{\xi}_k)_{k \in \mathbb{N}}\) of \( \mathbb{R}^{n_\xi}\) and any \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , we have for all vectors \( x\in \mathbb{R}^m\) and for all \( k \in \mathbb{N}\) ,

\[ \begin{align*} x^\top (M(\hat{\xi}_k,u_k))^\top M(\hat{\xi}_k,u_k) x & = x^\top (M(\xi_k,u_k) +\Delta M(\xi_k,\hat{\xi}_k,u_k))^\top (M(\xi_k,u_k) +\Delta M(\xi_k,\hat{\xi}_k,u_k)) x \\ & = x^\top (M(\xi_k,u_k))^\top M(\xi_k,u_k) x + 2 x^\top (\Delta M(\xi_k,\hat{\xi}_k,u_k))^\top M(\xi_k,u_k) x \\ &~~ {}+ x^\top (\Delta M(\xi_k,\hat{\xi}_k,u_k))^\top \Delta M(\xi_k,\hat{\xi}_k,u_k) x \\ & \geq c x^\top x - |2 x^\top (\Delta M(\xi_k,\hat{\xi}_k,u_k))^\top M(\xi_k,u_k) x|\\ & ~~ {}- |x^\top (\Delta M(\xi_k,\hat{\xi}_k,u_k))^\top \Delta M(\xi_k,\hat{\xi}_k,u_k) x|\\ & \geq \left(c - 2\|M(\xi_k,u_k)\|\|\Delta M(\xi_k,\hat{\xi}_k,u_k)\| - \|\Delta M(\xi_k,\hat{\xi}_k,u_k)\|^2\right)x^\top x, \end{align*} \]

where \( \Delta M(\xi_k,\hat{\xi}_k,u_k) = M(\hat{\xi}_k,u_k) - M(\xi_k,u_k)\) . Let \( \overline{c} > 0\) be such that \( \overline{c} \leq \delta\) and \( c - 2 c_M L_M \overline{c} - L_M^2 \overline{c}^2 > c^\prime\) . If (500) holds, then it follows that \( (\xi_k)_{k \in \mathbb{N}} \in \mathcal{C}^{\mathbb{N}}\) and \( (\hat{\xi}_k)_{k \in \mathbb{N}} \in (\mathcal{C} + \delta\mathbb{B})^{\mathbb{N}}\) . Thanks to the local boundedness and local Lipschitzness of \( M\) , we have for all \( (\xi_k)_{k \in \mathbb{N}} \in \mathcal{C}^{\mathbb{N}}\) , for all \( (\hat{\xi}_k)_{k \in \mathbb{N}} \in (\mathcal{C} + \delta\mathbb{B})^{\mathbb{N}}\) , and for all \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) ,

\[ \|M(\xi_k,u_k)\| \leq c_M, ~~ \|\Delta M(\xi_k,\hat{\xi}_k,u_k)\| \leq L_M|\xi_k - \hat{\xi}_k|, ~~ \forall k \in \mathbb{N}, \]

and (501) follows. \( \blacksquare\)

Lemma 31 (Technical Lyapunov lemma for Theorem 17)

Let \( A\) be a Schur non-zero matrix and \( Q = Q^\top > 0\) be a solution to \( A^\top Q A < Q\) . Consider three functions of variable \( (\tilde{z}_1,\tilde{Z}_2)\) :

\[ \begin{align*} g_1(\tilde{z}_1,\tilde{Z}_2)& = \gamma A\tilde{z}_1 + \gamma A \tilde{Z}_2u_1 + \gamma v_1 + w_1,\\ g_2(\tilde{Z}_2) &= \gamma A\tilde{Z}_2 U_2 + \gamma V_2 + W_2,\\ V(\tilde{z}_1,\tilde{Z}_2) &= \tilde{z}_1^\top Q \tilde{z}_1 + \|\tilde{Z}_2\|^2, \end{align*} \]

with \( \gamma >0\) and inputs \( (u_1,v_1,w_1,U_2,V_2,W_2)\) , where \( (\tilde{z}_1,u_1,v_1,w_1)\) are vectors and \( (\tilde{Z}_2,U_2,V_2,W_2)\) are matrices of appropriate dimensions, and some constant \( m > 0\) . There exist \( d_1,d_2,d_3,d_4,d_5 > 0\) such that for any \( \gamma >0\) , for all \( (\tilde{z}_1,\tilde{Z}_2)\) and for all \( (u_1,v_1,w_1,U_2,V_2,W_2)\) such that \( |u_1|^2 \leq m\) and \( \|U_2\|^2 \leq m\) , we have

\[ \begin{equation} V(g_1(\tilde{z}_1,\tilde{Z}_2),g_2(\tilde{Z}_2)) \leq \gamma^2 d_1V(\tilde{z}_1,\tilde{Z}_2)+\gamma^2d_2|v_1|^2 + d_3|w_1|^2 +\gamma^2d_4\|V_2\|^2 + d_5\|W_2\|^2. \end{equation} \]

(502)

Proof. Using the fact that \( \|U\|^2 = \|U^\top U\|\) for the \( 2\) -norm and Young’s inequality, for all \( (\tilde{z}_1,\tilde{Z}_2)\) and for all \( (u_1,v_1,w_1,U_2,V_2,W_2)\) such that \( |u_1|^2 \leq m\) and \( \|U_2\|^2 \leq m\) , we have

\[ \begin{align*} V(g_1(\tilde{z}_1,\tilde{Z}_2),g_2(\tilde{Z}_2)) & = \gamma^2 \tilde{z}_1^\top A^\top Q A\tilde{z}_1 + \gamma^2 u_1^\top \tilde{Z}_2^\top A^\top Q A\tilde{Z}_2u_1 + \gamma^2 v_1^\top Q v_1 + w_1^\top Q w_1+ 2 \gamma^2 \tilde{z}_1^\top A^\top Q A \tilde{Z}_2 u_1\\ &~~{}+ 2 \gamma^2 \tilde{z}_1^\top A^\top Q v_1 + 2 \gamma \tilde{z}_1^\top A^\top Q w_1 + 2 \gamma^2 u_1^\top \tilde{Z}_2^\top A^\top Q v_1+ 2\gamma u_1^\top \tilde{Z}_2^\top A^\top Q w_1 + 2 \gamma v_1^\top Q w_1 \\ &~~{}+ \| \gamma^2 U_2^\top \tilde{Z}_2^\top A^\top A\tilde{Z}_2U_2 + \gamma^2 V_2^\top V_2 + W_2^\top W_2 \\ &~~ + 2\gamma^2 U_2^\top \tilde{Z}_2^\top A^\top V_2 + 2\gamma U_2^\top \tilde{Z}_2^\top A^\top W_2 + 2\gamma V_2^\top W_2\|\\ & \leq \gamma^2 \left(c_1\tilde{z}_1^\top Q \tilde{z}_1 + c_2\|\tilde{Z}_2\|^2\right)+ \gamma^2c_3 |v_1|^2 + c_4 |w_1|^2+\gamma^2c_5\|V_2\|^2 + c_6\|W_2\|^2, \end{align*} \]

for some \( c_i > 0\) , \( i=1,2,…,6\) , some of which depend on \( m\) . Thus, Lemma 31 follows. \( \blacksquare\)

Lemma 32 (Boundedness and Lipschitzness of \( \rm{inv}_\underline{{c}_M}\))

Consider the inverse function \( \rm{inv}_\underline{{c}_M}: \mathbb{R}^{m \times n} \to \mathbb{R}^{n \times m}\) with \( m \geq n\) defined in (264.c) for some level \( \underline{c}_M > 0\) . For some \( \overline{c}_M \geq \underline{c}_M\) , define \( \mathcal{C}_M = \{M \in \mathbb{R}^{m \times n}: \underline{c}_M^2 \rm{Id} \leq M^\top M \leq \overline{c}_M^2 \rm{Id}\}\) . The map \( \rm{inv}_\underline{{c}_M}\) is such that:

  • For all \( \hat{M} \in \mathbb{R}^{m \times n}\) ,

    \[ \|\rm{inv}_{\underline{c}_M}(\hat{M})\| \leq \frac{1}{\underline{c}_M}; \]
  • There exists \( L_\rm{{inv}}> 0\) such that for all \( (M,\hat{M}) \in \mathcal{C}_M \times \mathbb{R}^{m \times n}\) ,

    \[ \begin{equation} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| \leq L_{\rm{inv}} \|M - \hat{M}\|. \end{equation} \]

    (503)

Proof. Let us start by proving two intermediary results. The first one is that for any full-rank matrix \( M \in \mathbb{R}^{m \times n}\) , if a matrix \( \hat{M} \in \mathbb{R}^{m \times n}\) is not full-rank, then \( \|M - \hat{M}\| \geq \sigma_{\min}(M)\) . Indeed, since \( \hat{M}\) is not full-rank, there exists \( x \in \mathbb{R}^n\setminus\{0\}\) such that \( \hat{M}x = 0\) . We then have \( |(M - \hat{M})x|^2 = |Mx|^2 = x^\top \hat{M}^\top \hat{M} x \geq (\sigma_{\min}(M))^2|x|^2\) and it follows that \( |(M - \hat{M})x| \geq \sigma_{\min}(M)|x|\) and thus \( \|M - \hat{M}\| \geq \sigma_{\min}(M)\) (since \( |x| \neq 0\) ).
Now we prove the second intermediary result that given \( \mathcal{C}_M\) being compact, there exists \( c_M > 0\) such that if \( \|M-\hat{M}\| \leq c_M\) for some \( M \in \mathcal{C}_M\) , then \( \hat{M}\) is such that \( \sigma_{\min}(\hat{M}) \geq \frac{\underline{c}_M}{2}\) . Indeed, assume the contrary and construct sequences \( (M_k)_{k \in \mathbb{N}} \in \mathcal{C}_M^{\mathbb{N}}\) and \( (\hat{M}_k)_{k \in \mathbb{N}} \in (\mathbb{R}^{m \times n})^{\mathbb{N}}\) such that \( \|M_k - \hat{M}_k\| \leq \frac{1}{k}\) and \( \sigma_{\min}(\hat{M}_k) < \frac{\underline{c}_M}{2}\) for all \( k \in \mathbb{N}\) . By the compactness of \( \mathcal{C}_M\) , we can extract a subsequence from \( (M_k)_{k \in \mathbb{N}}\) converging to \( M^*\) in \( \mathcal{C}_M\) . To alleviate the notations, we do not denote this extraction. Then, \( (\hat{M}_k)_{k \in \mathbb{N}}\) also converges to \( \hat{M}^*\) . It follows that \( \sigma_{\min}(\hat{M}^*) \geq \underline{c}_M\) , which is a contradiction by continuity of \( \sigma_{\min}\) .

To prove the first item of Lemma 32, we now show that \( \rm{inv}_\underline{{c}_M}\) is bounded in norm on \( \mathbb{R}^{m \times n}\) . The case of not being full-rank is trivial. For all \( \hat{M} \in \mathbb{R}^{m \times n}\) such that \( \sigma_{\min}(\hat{M}) > 0\) , we get that either \( \sigma_{\min}(\hat{M}) \geq \underline{c}_M\) and then \( \|\rm{inv}_\underline{{c}_M}(\hat{M})\| = \|\hat{M}^\dagger\| \leq 1/\underline{c}_M \) , or \( 0<\sigma_{\min}(\hat{M}) \leq \underline{c}_M\) and then \( \|\rm{inv}_\underline{{c}_M}(\hat{M})\| = \left\|\left( \sigma_{\min}(\hat{M})/\underline{c}_M \right)\hat{M}^\dagger\right\| \leq \sigma_{\min}(\hat{M})/(\underline{c}_M \sigma_{\min}(\hat{M})) \leq 1/\underline{c}_M \) . Therefore, \( \rm{inv}_\underline{{c}_M}\) is bounded in norm on \( \mathbb{R}^{m \times n}\) by \( 1/\underline{c}_M\) .

To prove the second item of Lemma 32, we pick \( M \in \mathcal{C}_M\) and consider four cases of \( \hat{M} \in \mathbb{R}^{m \times n}\) . First, if \( \hat{M}^\top \hat{M} \geq \underline{c}_M^2 \rm{Id}\) , i.e., \( \sigma_{\min}(M) \geq \underline{c}_M\) , then according to [209, Theorem 10.4.5], we have

\[ \begin{align*} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| & = \|M^\dagger - \hat{M}^\dagger\|\\ & \leq 3 \|M^\dagger\|\|\hat{M}^\dagger\|\|M-\hat{M}\|\\ & \leq \frac{3}{\underline{c}_M^2} \|M-\hat{M}\|. \end{align*} \]

Second, if \( \hat{M}\) is not full-rank, from the first intermediary result above, we deduce that since \( M\) is full-rank, \( \|M - \hat{M}\| \geq \sigma_{\min}(M)\) and so

\[ \begin{align*} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| & = \|M^\dagger\|\\ & = \|M^\dagger\|^2\sigma_{\min}(M) \\ & \leq \frac{1}{\underline{c}_M^2} \|M - \hat{M}\|. \end{align*} \]

Third, if \( 0 < \sigma_{\min}(\hat{M}) \leq \underline{c}_M\) and \( \|M - \hat{M}\| \leq c_M\) , then we have that \( \tilde{M} := \|\hat{M}^\dagger\|\underline{c}_M \hat{M}\) is such that \( \tilde{M}^\top \tilde{M} \geq \underline{c}_M^2 \rm{Id}\) and by applying the first case to it, we get

\[ \begin{align*} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| & = \left\|M^\dagger - \frac{1}{\|\hat{M}^\dagger\|\underline{c}_M}\hat{M}^\dagger\right\|\\ & = \|M^\dagger - (\|\hat{M}^\dagger\|\underline{c}_M\hat{M})^\dagger\|\\ & \leq \frac{3}{\underline{c}_M^2} \|M - \tilde{M}\|\\ & \leq \frac{3}{\underline{c}_M^2} \|M - \|\hat{M}^\dagger\|\underline{c}_M\hat{M}\|\\ & \leq \frac{3}{\underline{c}_M} \left\|\left(\frac{1}{\underline{c}_M}-\|\hat{M}^\dagger\|\right)M\right\| + \frac{3}{\underline{c}_M}\|\hat{M}^\dagger\| \|M-\hat{M}\|. \end{align*} \]

Since \( \|M^\dagger\| \leq 1/\underline{c}_M \leq \|\hat{M}^\dagger\| \) , we get that \( \left| 1/\underline{c}_M -\|\hat{M}^\dagger\|\right| \leq |\|M^\dagger\|- \|\hat{M}^\dagger\|| \leq \|M^\dagger - \hat{M}^\dagger\| \) and so using [209, Theorem 10.4.5] again, we get

\[ \begin{align*} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\|& \leq \frac{3}{\underline{c}_M} \|M\|\|M^\dagger - \hat{M}^\dagger\| + \frac{3}{\underline{c}_M^2}\|\hat{M}^\dagger\|\underline{c}_M \|M-\hat{M}\|\\ & \leq \frac{9}{\underline{c}_M} \|M\|\|M^\dagger\|\|\hat{M}^\dagger\|\|M - \hat{M}\| + \frac{3}{\underline{c}_M}\|\hat{M}^\dagger\| \|M-\hat{M}\|\\ & \leq \frac{3}{\underline{c}_M} \left(3\frac{\overline{c}_M}{\underline{c}_M} + 1\right)\|\hat{M}^\dagger\|\|M - \hat{M}\|. \end{align*} \]

Now from the second intermediary result, since \( \|M - \hat{M}\| \leq c_M\) , we have \( \sigma_{\min}(\hat{M}) \geq \frac{\underline{c}_M}{2}\) and we get

\[ \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| \leq \frac{6}{\underline{c}_M^2} \left(3\frac{\overline{c}_M}{\underline{c}_M} + 1\right)\|M - \hat{M}\|. \]

Fourth, if \( 0 < \sigma_{\min}(\hat{M}) \leq \underline{c}_M\) and \( \|M - \hat{M}\| \geq c_M\) , from the boundedness of \( \rm{inv}_\underline{{c}_M}\) proven above, we get

\[ \begin{align*} \|M^\dagger - \rm{inv}_{\underline{c}_M}(\hat{M})\| &\leq \frac{2}{\underline{c}_M}\\ & \leq \frac{2}{\underline{c}_M c_M} \|M - \hat{M}\|. \end{align*} \]

Combining all these four cases, we get the result by letting \( L_\rm{{inv}} = \max\left\{ 3/\underline{c}_M^2, 6/\underline{c}_M^2 \times \left(3 \overline{c}_M/\underline{c}_M + 1\right), 2/(\underline{c}_M c_M) \right\}\) . \( \blacksquare\)

Lemma 33 (Transferring left invertibility from map to map)

For some \( n_p,n_u,n_x \in \mathbb{N}\) , let \( ¶ \subset \mathbb{R}^{n_p}\) be a compact set, \( \mathcal{U} \subseteq \mathbb{R}^{n_u}\) be a set, \( \mathcal{X}\subset \mathbb{R}^{n_x}\) be a compact set, and \( ¶_s\) be a subset of \( ¶^{\mathbb{N}}\) . Consider an open set \( \mathcal{O}\) containing \( ¶ \times \mathcal{U} \times \mathcal{X}\) and \( \mathcal{F}\) a function defined on \( \mathbb{R}^{n_p}\times \mathbb{R}^{n_u}\times\mathbb{R}^{n_x}\) , that is \( C^2\) with respect to \( (p,x)\) , and such that the maps \( \mathcal{F}\) , \( \frac{\partial \mathcal{F}}{\partial p}\) , \( \frac{\partial \mathcal{F}}{\partial x}\) and \( \frac{\partial^2 \mathcal{F}}{\partial p\partial x}\) are all bounded on \( \mathcal{O}\) . Assume that there exists \( c > 0\) such that for all \( (p_k)_{k \in \mathbb{N}} \in ¶_s\) , for all \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , and for all \( (x_a,x_b) \in \mathcal{X} \times \mathcal{X}\) , we have

\[ \begin{equation} |\mathcal{F}(p_k,u_k,x_a) - \mathcal{F}(p_k,u_k,x_b)| \geq c|x_a - x_b|, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(504)

Then, for any \( 0 < c^\prime < c\) , there exists \( \overline{c} > 0\) such that for any \( (\hat{p}_k)_{k \in \mathbb{N}}\) for which

\[ \begin{equation} \exists (p_k)_{k \in \mathbb{N}} \in ¶_s: |p_k - \hat{p}_k| \leq \overline{c}, ~~ \forall k \in \mathbb{N}, \end{equation} \]

(505)

and for any \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , we have for all \( (x_a,x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{equation} |\mathcal{F}(\hat{p}_k,u_k,x_a) - \mathcal{F}(\hat{p}_k,u_k,x_b)| \geq c^\prime|x_a - x_b|, ~~ \forall k \in \mathbb{N}. \end{equation} \]

(506)

Proof. Since \( \mathcal{O}\) is open, there exists \( \delta > 0\) such that \( ¶_\delta \times \mathcal{U} \times \mathcal{X} \subset \mathcal{O}\) , where \( ¶_\delta := ¶ + \delta\mathbb{B}\) . For the same reason, there exists \( r > 0\) such that \( (¶_\delta + r\mathbb{B}) \times \mathcal{U} \times (\mathcal{X} + r\mathbb{B}) \subset \mathcal{O}\) . Let \( (p_0,u,x_0) \in ¶_\delta \times \mathcal{U} \times \mathcal{X}\) . We have for all \( (p,x_a)\) and \( (\hat{p},x_b)\) in \( (\{p_0\} + r\mathbb{B}) \times (\{x_0\} + r \mathbb{B})\) ,

\[ \begin{align*} &\left|\mathcal{F}(p,u,x_b) - \mathcal{F}(p,u,x_a) - (\mathcal{F}(\hat{p},u,x_b) - \mathcal{F}(\hat{p},u,x_a))\right| \\ \leq &\left|\int_0^1 \frac{\partial \mathcal{F}}{\partial x} (p,u,x_a+s(x_b-x_a)) ds(x_b-x_a) - \int_0^1 \frac{\partial \mathcal{F}}{\partial x} (\hat{p},u,x_a+s(x_b-x_a)) ds(x_b-x_a) \right|\\ \leq &\left(\int_0^1 \left|\frac{\partial \mathcal{F}}{\partial x} (p,u,x_a+s(x_b-x_a)) - \frac{\partial \mathcal{F}}{\partial x} (\hat{p},u,x_a+s(x_b-x_a)) \right|ds\right)|x_b-x_a|\\ \leq & \left(\int_0^1 \int_0^1 \left| \frac{\partial^2 \mathcal{F}}{\partial p\partial x} (\hat{p}+s^\prime(p-\hat{p}),u,x_a+s(x_b-x_a)) \right|ds^\prime ds\right) |\hat{p} - p||x_b-x_a|. \end{align*} \]

Because \( \frac{\partial^2 \mathcal{F}}{\partial x\partial p}\) is bounded on \( \mathcal{O}\) and \( \hat{p}+s^\prime(p-\hat{p}),u,x_a+s(x_b-x_a) \in (\{p_0\} + r\mathbb{B}) \times (\{x_0\} + r \mathbb{B})\) for all \( (s,s^\prime)\in [0,1]\times [0,1]\) by convexity of \( \mathbb{B}\) , we deduce that there exists \( c_b > 0\) such that for all \( (p,x_a)\) and \( (\hat{p},x_b)\) in \( (\{p_0\} + r\mathbb{B}) \times (\{x_0\} + r \mathbb{B})\) ,

\[ \begin{equation} |\mathcal{F}(\hat{p},u,x_a) - \mathcal{F}(p,u,x_a) - (\mathcal{F}(\hat{p},u,x_b) - \mathcal{F}(p,u,x_b))| \leq c_b |p - \hat{p}||x_a - x_b|. \end{equation} \]

(507)

Now, since \( ¶_\delta\) and \( \mathcal{X}\) are compact, there exists a finite subcover, i.e., a finite number \( M_p\) of points \( p_q\) , \( q \in \{1,2,…,M_p\}\) (resp., \( M_x\) of points \( x_i\) , \( i \in \{1,2,…,M_x\}\) ) such that

\[ ¶_\delta \times \mathcal{X} \subset \bigcup_{q=1}^{M_p} \bigcup_{i=1}^{M_x}((p_q + r\mathring{\mathbb{B}})\times(x_i + r\mathring{\mathbb{B}})) \subset \bigcup_{q=1}^{M_p} \bigcup_{i=1}^{M_x}((p_q + r\mathbb{B})\times(x_i + r\mathbb{B})). \]

We would like to use contradiction to show that there exists \( c_b^\prime >0\) such that for all \( (p,x_a)\) and \( (\hat{p},x_b)\) in \( ¶_\delta \times \mathcal{X}\) and for all \( u \in \mathcal{U}\) ,

\[ \begin{equation} |\mathcal{F}(\hat{p},u,x_a) - \mathcal{F}(p,u,x_a) - (\mathcal{F}(\hat{p},u,x_b) - \mathcal{F}(p,u,x_b))| \leq c_b^\prime |p - \hat{p}||x_a - x_b|. \end{equation} \]

(508)

Assume the contrary, namely for all \( k \in \mathbb{N}\) , there exists \( (p_k,x_{a,k})_{k \in \mathbb{N}}\) and \( (\hat{p}_k,x_{b,k})_{k \in \mathbb{N}}\) in \( ¶_\delta^\mathbb{N} \times \mathcal{X}^\mathbb{N}\) and \( (u_k)_{k\in \mathbb{N}} \in \mathcal{U}^\mathbb{N}\) such that

\[ \begin{equation} |\mathcal{F}(\hat{p}_k,u_k,x_{a,k}) - \mathcal{F}(p_k,u_k,x_{a,k}) - (\mathcal{F}(\hat{p}_k,u_k,x_{b,k}) - \mathcal{F}(p_k,u_k,x_{b,k}))| \geq k|p_k - \hat{p}_k||x_{a,k} - x_{b,k}|, ~ \forall k \in \mathbb{N}. \end{equation} \]

(509)

By the compactness of \( ¶_\delta \times \mathcal{X}\) , we can extract a subsequence from \( (p_k,x_{a,k},\hat{p}_k,x_{b,k})_{k \in \mathbb{N}}\) converging to \( (p^*,x_a^*,\hat{p}^*,x_b^*)\) in \( (¶_\delta\times\mathcal{X})\times(¶_\delta\times\mathcal{X})\) . To alleviate the notations, we do not denote this extraction. From the boundedness of \( \mathcal{F}\) on \( \mathcal{O}\) , the sequence \( (|\mathcal{F}(\hat{p}_k,u_k,x_{a,k}) - \mathcal{F}(p_k,u_k,x_{a,k}) - (\mathcal{F}(\hat{p}_k,u_k,x_{b,k}) - \mathcal{F}(p_k,u_k,x_{b,k}))|)_{k \in \mathbb{N}}\) is bounded, and it follows from (509) that either \( p^* = \hat{p}^*\) or \( x_a^* = x_b^*\) . Let us assume first that \( x_a^* = x_b^*\) . There then exists \( i \in \{1,2,…,M_x\}\) such that \( x_a^* = x_b^* \in \{x_i\} + r\mathring{\mathbb{B}}\) . Then, because the ball \( \{x_i\} + r\mathring{\mathbb{B}}\) is open, there exists \( K_1 \in \mathbb{N}\) such that for all \( k \in \mathbb{N}_{\geq K_1}\) , \( x_{a,k} \in \{x_i\} + r\mathring{\mathbb{B}}\) and \( x_{b,k} \in \{x_i\} + r\mathring{\mathbb{B}}\) . Thanks to the convexity of \( \{x_i\} + r\mathring{\mathbb{B}}\) , we have

\[ \begin{multline*} |\mathcal{F}(\hat{p}_k,u_k,x_{a,k}) - \mathcal{F}(p_k,u_k,x_{a,k}) - (\mathcal{F}(\hat{p}_k,u_k,x_{b,k}) - \mathcal{F}(p_k,u_k,x_{b,k}))| \\ \leq \left|\int_0^1 \frac{\partial \mathcal{F}}{\partial x} (p_k,u_k,x_{a,k}+s(x_{b,k}-x_{a,k})) - \frac{\partial \mathcal{F}}{\partial x} (\hat{p}_k,u_k,x_{a,k}+s(x_{b,k}-x_{a,k})) \right|ds |x_{b,k}-x_{a,k}|, ~~ \forall k \in \mathbb{N}. \end{multline*} \]

By the boundedness of \( \frac{\partial \mathcal{F}}{\partial x}\) on \( \mathcal{O}\) and the fact that \( (¶_\delta \times \mathcal{U} \times (\{x_i\} + r\mathring{\mathbb{B}})) \subset (¶_\delta \times \mathcal{U} \times (\mathcal{X} + r\mathbb{B})) \subset \mathcal{O}\) , we get that the term in the integral is bounded and according to (509), we get \( p^* = \hat{p}^*\) . Thus, there exists \( q \in \{1,2,…,M_p\}\) such that \( p^* = \hat{p}^* \in \{p_q\}+r\mathring{\mathbb{B}}\) . Because the ball \( \{p_q\}+r\mathring{\mathbb{B}}\) is open, there exists \( K_2 \in \mathbb{N}\) such that for all \( k \in \mathbb{N}_{\geq K_2}\) , \( p_k \in \{p_q\} + r\mathring{\mathbb{B}}\) and \( \hat{p}_k \in \{p_q\} + r\mathring{\mathbb{B}}\) . It follows that for all \( k \in \mathbb{N}_{\geq \max\{K_1,K_2\}}\) , \( (p_k,x_{a,k})\) and \( (\hat{p}_k,x_{b,k})\) are in \( (\{p_q\}+r\mathring{\mathbb{B}})\times(\{x_i\}+r\mathring{\mathbb{B}})\) . By applying (507), we get that

\[ \begin{multline*} |\mathcal{F}(\hat{p}_k,u_k,x_{a,k}) - \mathcal{F}(p_k,u_k,x_{a,k}) - (\mathcal{F}(\hat{p}_k,u_k,x_{b,k}) - \mathcal{F}(p_k,u_k,x_{b,k}))| \leq c_b |p_k-\hat{p}_k||x_{a,k}-x_{b,k}|, \\ \forall k \in \mathbb{N}_{\geq \max\{K_1,K_2\}}, \end{multline*} \]

which contradicts with (509). Similarly, if we had assume \( p^* = \hat{p}^*\) instead of \( x_a^* = x_b^*\) , we would have obtained that \( x_a^* = x_b^*\) and obtained the same contradiction. Therefore, (508) follows.

Fix now a first choice of \( \overline{c}>0\) that is smaller than \( \delta\) . Since \( 0 < \overline{c} < \delta\) , if (505) holds, then it follows that \( (p_k)_{k \in \mathbb{N}} \in ¶^{\mathbb{N}}\) and \( (\hat{p}_k)_{k \in \mathbb{N}} \in ¶_\delta^{\mathbb{N}}\) . Using (508), we have for all \( (p_k)_{k \in \mathbb{N}} \in ¶^{\mathbb{N}}\) , for all \( (\hat{p}_k)_{k \in \mathbb{N}} \in ¶_\delta^{\mathbb{N}}\) , for all \( (u_k)_{k \in \mathbb{N}} \in \mathcal{U}^{\mathbb{N}}\) , and for all \( (x_a,x_b) \in \mathcal{X} \times \mathcal{X}\) ,

\[ \begin{align*} |\mathcal{F}(\hat{p}_k,u_k,x_a) - \mathcal{F}(\hat{p}_k,u_k,x_b)| & = |\mathcal{F}(p_k,u_k,x_a)-\mathcal{F}(p_k,u_k,x_b) \\ &~~ {} +\mathcal{F}(\hat{p}_k,u_k,x_a) - \mathcal{F}(p_k,u_k,x_a)- (\mathcal{F}(\hat{p}_k,u_k,x_b) - \mathcal{F}(p_k,u_k,x_b))|\\ &\geq |\mathcal{F}(p_k,u_k,x_a)-\mathcal{F}(p_k,u_k,x_b)| \\ &~~ {} - |\mathcal{F}(\hat{p}_k,u_k,x_a) - \mathcal{F}(p_k,u_k,x_a) - (\mathcal{F}(\hat{p}_k,u_k,x_b) - \mathcal{F}(p_k,u_k,x_b))| \\ & \geq c|x_a - x_b| - c_b|p_k - \hat{p}_k||x_a - x_b|\\ &\geq (c - c_b^\prime \overline{c})|x_a - x_b|. \end{align*} \]

Therefore, for any \( 0 < c^\prime < c\) , by picking \( \overline{c} > 0\) such that \( \overline{c} < \delta\) and \( c - c_b^\prime \overline{c} > c^\prime\) , we get the result. \( \blacksquare\)

6.3 Technical Lemmas and Proofs of\specialChar{160}\Cref{part_unknown}

 

Cette annexe contient les lemmes techniques et les preuves de la Partie 4.

6.3.1 Proof of Lemma 16

Let \( \overline{c}_f > \max\{c_f,0\}\) . First, for all \( (x,\tau) \in \rm{cl}(C)\times (-\infty,0]\) such that \( \Psi_f(x,\tau)\) is defined, we have

\[ \begin{equation} \Psi_f(x,\tau) = x + \int_{0}^{\tau}f(\Psi_f(x,s))ds, \end{equation} \]

(510)

and using the triangle inequality, the property (336.a), Grönwall’s inequality, and the inequality \( x \leq e^x\) for all \( x \geq 0\) , we get

\[ \begin{align*} |\Psi_f(x,\tau)|&\leq |x|+ \int_{0}^{|\tau|}|f(\Psi_f(x,s))|ds\\ & \leq |x|+ \int_{0}^{|\tau|}(\overline{c}_f|\Psi_f(x,s)| + d_f)ds\\ & \leq |x|+ \overline{c}_f\int_{0}^{|\tau|}|\Psi_f(x,s)|ds +d_f\tau\\ & \leq e^{\overline{c}_f|\tau|}|x| +d_f|\tau|\\ & \leq e^{\overline{c}_f|\tau|}\left(|x| + \frac{d_f}{\overline{c}_f}\right). \end{align*} \]

Now, pick a \( t\) -backward complete solution \( \phi\) to system (325) initialized as \( x \in \rm{cl}(C) \cup D\) . Let \( \overline{c}_g > \max\{c_g,1\}\) . Now, we have for all \( s \in (-\infty,0]\) and for all \( j\) such that \( (s,j) \in \rm{dom} \phi \cap (\mathbb{R}_{\leq 0}\times\mathbb{Z}_{\leq 0})\) ,

\[ \begin{align*} |Y(x,s)|&= |h(\Psi_f(g^{-1}(\Psi_f(g^{-1}(\ldots\Psi(x,t_{-1})),t_{j}-t_{j-1})),s-t_j))|\\ &\leq c_h\left(e^{-\overline{c}_fs}\overline{c}_g^{-j}\left(|x| + \frac{d_f}{\overline{c}_f}\right) + d_g\sum_{k=0}^{|j|-1}e^{\overline{c}_f\left(t_{-k-1}-s\right)} \overline{c}_g^{|j|-k}\right)^{p_h} + d_h\\ &\leq c_h\left(e^{-\overline{c}_fs}\overline{c}_g^{-j}\left(|x| + \frac{d_f}{\overline{c}_f}\right) + d_ge^{-\overline{c}_fs}\sum_{k=0}^{|j|-1} \overline{c}_g^{|j|-k}\right)^{p_h} + d_h\\ &\leq c_h\left(e^{-\overline{c}_fs}\overline{c}_g^{-j}\left(|x| + \frac{d_f}{\overline{c}_f}\right) + d_ge^{-\overline{c}_fs}\frac{\overline{c}_g^{|j|}-1}{\overline{c}_g-1}\right)^{p_h} + d_h. \end{align*} \]

Since \( |j| \leq -\frac{1}{\tau_m} s + N\) , we get

\[ \begin{equation} |Y(x,s)| \leq c_h\bigg(e^{-\overline{c}_fs}\overline{c}_g^{-\frac{1}{\tau_m} s + N}\left(|x| + \frac{d_f}{\overline{c}_f}\right)\\ + d_ge^{-\overline{c}_fs}\frac{\overline{c}_g^{-\frac{1}{\tau_m} s + N}-1}{\overline{c}_g-1}\bigg)^{p_h} + d_h. \end{equation} \]

(511)

Therefore, the result follows with \( \rho = p_h\left(\overline{c}_f + \frac{1}{\tau_m}\ln\overline{c}_g\right)\) .

6.3.2 Proof of Lemma 17

Pick any \( \rho>0\) and a maximal solution \( \phi\) to (325) initialized as \( x \in \rm{cl}(C) \cup D\) . By assumption, there exists a compact set \( \mathcal{C}\subset \mathbb{R}^{n_x}\) such that \( \phi\) is in \( \mathcal{C}\) in negative time. Let us study all the cases that can happen for this solution. If \( t^-(x) = -\infty\) , then for all \( s\leq 0\) , \( Y(x,s)\in h(\mathcal{C})\) which is bounded, so Item (XXXIV) of Assumption 31 holds. Now to prove Item (XXXIII) of Assumption 31, suppose \( t^-(x)\) is finite. This contains two smaller cases, namely i) \( t^-(x) \in \rm{dom}_t\phi\) , and ii) \( t^-(x) \neq \rm{dom}_t\phi\) . In case i), we have \( \lim_{s \to t^-(x)}Y(x,s) = Y(x,t^-(x))\) by continuity of the output, so that Item (XXXIII) of Assumption 31 holds. Case ii) can only happen when there is Zeno (finite-time escape during flows is excluded because \( \phi\) is bounded in backward time), so \( \inf \rm{dom}_j \phi = -\infty\) and \( \rm{dom}_j \phi \cap \mathbb{Z}_{\leq 0} = \mathbb{Z}_{\leq 0}\) . Because \( f\) is continuous (see Item (XXXII) of Assumption 30), there exists \( b_f > 0\) such that \( |f(x)| \leq b_f\) for all \( x \in \mathcal{C}\) which is compact. It thus follows that \( |\phi(t_j,j) - \phi(t_{j+1},j)| \leq b_f |t_j - t_{j+1}|\) for all \( j \in \mathbb{Z}_{\leq 0}\) . From the local Lipschitzness of \( h\) on the compact set \( \mathcal{C}\) , there exists \( c_h > 0\) such that for all \( j \in \mathbb{Z}_{\leq 0}\) , we have

\[ \begin{align*} |Y(x,t_{j-1}) - Y(x,t_{j})| &= |h(\phi(t_{j-1},j)) - h(\phi(t_j,j))|\\ & \leq c_h |\phi(t_{j-1},j) - \phi(t_j,j)| \\ &\leq c_h b_f |t_{j-1} - t_j|. \end{align*} \]

Therefore, by the triangle inequality, for any \( j\in \mathbb{Z}_{\leq 0}\) and any \( p\in \mathbb{N}\) ,

\[ \begin{align*} |Y(x,t_{j-p}) - Y(x,t_j)| &\leq |Y(x,t_{j-p}) - Y(x,t_{j-p+1})| + \ldots + |Y(x,t_{j-1}) - Y(x,t_j)| \\ &\leq c_h b_f |t_{j-p} - t_j|. \end{align*} \]

Since the sequence \( (t_j)_{j \in \mathbb{Z}_{\leq 0}}\) converges to \( t^-(x)\) , it is a Cauchy sequence and it follows that \( (Y(x,t_j))_{j \in \mathbb{Z}_{\leq 0}}\) is also a Cauchy sequence. Then by the completeness of the space \( \mathbb{R}^{n_y}\) , we know that \( (Y(x,t_j))_{j \in \mathbb{Z}_{\leq 0}}\) converges as \( j \to -\infty\) to a limit denoted as \( Y^\star\) . Now for all \( s \in \rm{dom}_t \phi \cap \mathbb{R}_{\leq 0}\) , there exists \( j(s) \in \mathbb{Z}_{\leq 0}\) such that \( (s,j(s)) \in \rm{dom} \phi \cap (\mathbb{R}_{\leq 0} \times \mathbb{Z}_{\leq 0})\) and \( t_{j(s)-1} \leq s \leq t_{j(s)}\) and we have

\[ \begin{align*} |Y(x,s) - Y^\star| &\leq |Y(x,s) - Y(x,t_{j(s)})| + |Y(x,t_{j(s)}) - Y^\star|\\ &\leq c_h b_f |s - t_{j(s)}| + |Y(x,t_{j(s)}) - Y^\star|\\ &\leq c_h b_f |t_{j(s)-1} - t_{j(s)}| + |Y(x,t_{j(s)}) - Y^\star|. \end{align*} \]

As \( s\) goes to \( t^-(x)\) , \( j(s)\) goes to \( -\infty\) ; the flow length \( t_{j(s)-1} - t_{j(s)}\) converges to \( 0\) (a necessary condition of Zeno); and \( Y(x,t_{j(s)})\) is a subsequence of \( (Y(x,t_j))_{j \in \mathbb{Z}_{\leq 0}}\) and converges to \( Y^\star\) . It follows that \( \lim_{s\to t^{-}(x)}Y(x,s) = Y^\star\) , which means Item (XXXIII) of Assumption 31 holds.

6.3.3 Proof of Theorem 19

Let \( x\) be a solution to system (325) initialized in \( \mathcal{X}_0\) . It is \( t\) -forward complete by Item (XXX) of Assumption 30. By the conclusion of Lemma 18, there exists a \( C^1\) map \( z:\mathbb{R}_{\geq 0} \to \mathbb{R}^{n_z}\) such that (341) and (342) hold. Since \( \dot{z}-\dot{\hat{z}}=A(z-\hat{z})\) and \( A\) is Hurwitz, there exist \( c_1,\lambda>0\) , independent from the considered solution \( x\) , such that any solution to observer (356) is such that for all \( t \geq 0\) ,

\[ \begin{equation} |z(t)-\hat{z}(t)| \leq c_1e^{-\lambda t}|z(0)-\hat{z}(0)|. \end{equation} \]

(512)

From Lemma 20 and the continuity of \( T\) (resp., \( d_\mathcal{A}\) ) on the compact set \( \mathcal{X}\) (resp., \( \mathcal{X}\times \mathcal{X}\) ), applying Lemma 35 in Section 6.3.7 with \( \alpha_1 = d_\mathcal{A}\) and \( \alpha_2: \mathcal{X} \times \mathcal{X} \to \mathbb{R}^{n_z}\) defined as \( \alpha_2(x_a,x_b) = T(x_a) - T(x_b)\) , with (351), we deduce that there exists a class-\( \mathcal{K}\) function \( \alpha\) such that

\[ \begin{equation} d_\mathcal{A}((x_a,x_b)) \leq \alpha\left(|T(x_a)-T(x_b)| \right), ~~ \forall (x_a,x_b)\in \mathcal{X} \times \mathcal{X}. \end{equation} \]

(513)

Let us show that \( T^*\) defined in (353) takes values in \( \mathcal{X}\setminus D\) . Let \( z^\star\in \mathbb{R}^{n_z}\) and denote \( z^{\star\prime} = \Pi_{T(\mathcal{X})}(z^\star)\) . Then, \( z^{\star\prime}\in T(\mathcal{X})\) by definition of the projection, so there exists \( x^\star\in \mathcal{X}\subset \rm{cl}(C)\cup D\) such that \( z^{\star\prime}=T(x^\star)\) . Then, either \( x^\star\in \rm{cl}(C)\setminus D\) , and \( T^*(z^\star)=x^\star\in \mathcal{X}\setminus D\) ; or \( x^\star\in D\) and \( z^{\star\prime}=T(g(x^\star))\) by Item (XXXV) of Definition 24 with \( g(x^\star)\in \rm{cl}(C)\setminus D\) by Item (XXXVIII) of Assumption 33, so that \( T^*(z^\star)=g(x^\star)\) , which is in \( \mathcal{X}\setminus D\) by Item (XXXVII). In all cases, \( T^*\) takes values in \( \mathcal{X}\setminus D\) . It follows that \( \hat{x}(t)\in \mathcal{X}\) for all \( t\geq 0\) and \( T(\hat{x}(t)) = \Pi_{T(\mathcal{X})}(\hat{z}(t))\) for all \( t\geq 0\) . Therefore, from (341) in Lemma 18, for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{align} d_\mathcal{A}((x(t,j),\hat{x}(t))) &\leq \alpha\left(|T(x(t,j))-T(\hat{x}(t))| \right)\notag \end{align} \]

(514)

\[ \begin{align} & \leq \alpha\left(|z(t)-\Pi_{T(\mathcal{X})}(\hat{z}(t))| \right). \\ \\\end{align} \]

(515)

But by the triangle inequality, the definition of the projection (since \( z(t)\in T(\mathcal{X})\) from (341)), and Item (XXXVII) of Assumption 33, we have for all \( t\geq 0\) ,

\[ \begin{align} |z(t)-\Pi_{T(\mathcal{X})}(\hat{z}(t))|&\leq |z(t)-\hat{z}(t)| + |\hat{z}(t)-\Pi_{T(\mathcal{X})}(\hat{z}(t))|\notag \end{align} \]

(516)

\[ \begin{align} &\leq |z(t)-\hat{z}(t)| + |\hat{z}(t)-z(t)|\notag \end{align} \]

(517)

\[ \begin{align} &=2|z(t)-\hat{z}(t)|, \\ \\\end{align} \]

(518)

so that, for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{equation} d_\mathcal{A}((x(t,j),\hat{x}(t)))\leq \alpha (2|z(t)-\hat{z}(t)|). \end{equation} \]

(519)

Combining this with (512), we obtain (357).

Now let us show the asymptotic stability in (358). From Lemma 20 and the continuity of \( T\) (resp., \( d_\mathcal{A}\) ) on the compact set \( \mathcal{X}\) (resp., \( \mathcal{X}\times \mathcal{X}\) ), applying Lemma 35 in Section 6.3.7 as done to obtain (513) but this time the other way around, namely with \( \alpha_1: \mathcal{X} \times \mathcal{X} \to \mathbb{R}^{n_z}\) defined as \( \alpha_1(x_a,x_b) = T(x_a) - T(x_b)\) and \( \alpha_2 = d_\mathcal{A}\) , with (351), we deduce that there exists a class-\( \mathcal{K}\) function \( \alpha^\prime\) such that

\[ \begin{equation} |T(x_a)-T(x_b)| \leq \alpha^\prime\left(d_\mathcal{A}((x_a,x_b))\right), ~~ \forall (x_a,x_b)\in \mathcal{X} \times \mathcal{X}. \end{equation} \]

(520)

Because \( \hat{z}(0) \in T(\mathcal{X})\) , there exists \( \hat{x}(0) \in \mathcal{X}\) such that \( \hat{z}(0) = T(\hat{x}(0))\) . It then follows from (520) that

\[ \begin{equation} |z(0)-\hat{z}(0)| = |T(x(0,0))-T(\hat{x}(0))|\leq \alpha^\prime\left(d_\mathcal{A}((x(0,0),\hat{x}(0)))\right). \end{equation} \]

(521)

Combining (519) with (512) and (521), we get that for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ d_\mathcal{A}((x(t,j),\hat{x}(t)))\leq \alpha (2c_1e^{-\lambda t}\alpha^\prime\left(d_\mathcal{A}((x(0,0),\hat{x}(0)))\right)), \]

which is a class-\( \mathcal{KL}\) function in \( d_\mathcal{A}((x(0,0),\hat{x}(0)))\) and \( t\) .

6.3.4 Proof of Theorem 20

Pick a solution \( x\) to system (365) initialized in \( \mathcal{X}_0\) with disturbances and noise in \( \mathfrak{V}_c\times\mathfrak{V}_d\times\mathfrak{W}\) and any solution to observer (367) fed with \( y\) from (365). Consider \( T\) given by Theorem 19 and define \( z = T(x)\) . By using Item (XXXV) of Definition 24 and the conclusion of Lemma 18 as well as the assumed \( C^1\) property of \( T\) , \( z\) evolves according to the (hybrid) dynamics

\[ \begin{align} \dot{z} & = A z + Bh(x) + \frac{\partial T}{\partial x}(x)v_c \end{align} \]

(522.a)

\[ \begin{align} z^+ & = T(g(x) + v_{d,j}) = z + T(g(x) + v_{d,j}) - T(g(x)), \\ \\\end{align} \]

(522.b)

with the same jump times as \( x\) . Indeed, \( (v_{d,j})_{j \in\mathbb{N}}\) introduces jumps in \( z\) along the solutions to system (365), because it introduces a mismatch in the gluing property in Item (XXXV) of Definition 24. We thus write hereafter \( z(t,j)\) instead of \( z(t)\) . On the other hand, \( t\mapsto \hat{z}(t)\) follows the continuous-time dynamics given by (367)

\[ \begin{equation} \dot{\hat{z}} = A \hat{z} + B(h(x) + w). \end{equation} \]

(523)

Therefore, the hybrid arc \( \tilde{z}\) defined as \( \tilde{z}(t,j) := z(t,j) - \hat{z}(t)\) is solution to

\[ \begin{align} \dot{\tilde{z}} & = A \tilde{z} + \frac{\partial T}{\partial x}(x)v_c - Bw \end{align} \]

(524.a)

\[ \begin{align} \tilde{z}^+ & = \tilde{z} + T(g(x) + v_{d,j}) - T(g(x)), \\ \\\end{align} \]

(524.b)

with the same jump times as \( x\) . Since \( T\) is \( C^1\) on the compact set \( \mathcal{X}\) , there exists \( b_c > 0\) such that

\[ \begin{equation} \left|\frac{\partial T}{\partial x}(x)\right| \leq b_c, ~~ \forall x \in \mathcal{X}. \end{equation} \]

(525)

Next, because \( T\) is \( C^1\) on the compact set \( \mathcal{X}\) , it is Lipschitz on \( \mathcal{X}\) with Lipschitz constant \( L_T\) , and we have

\[ \begin{equation} |T(g(x) + v_{d,j}) - T(g(x))| \leq L_T |v_{d,j}|, ~~ \forall x \: : \: g(x) \in \mathcal{X}, g(x)+v_{d,j} \in \mathcal{X}, \: \forall j \in \mathbb{N}. \end{equation} \]

(526)

By Assumption 34 and Item (XXXVII) of Assumption 33, we have for all \( j \in \mathbb{N}_{\geq 1}\) ,

\[ \begin{equation} x(t_j,j-1) \in \mathcal{X}\cap D, ~~ g(x(t_j,j-1)) \in \mathcal{X}, ~~ g(x(t_j,j-1))+v_{d,j}=x(t_j,j) \in \mathcal{X}. \end{equation} \]

(527)

Combining (524) with (525) and (526), as well as the Hurwitzness of \( A\) , we deduce that there exist positive scalars \( (\lambda,c_1,c_2,c_3,c_4)\) such that for any solution \( x\) to system (365) initialized in \( \mathcal{X}_0\) with disturbances and noise in \( \mathfrak{V}_c\times\mathfrak{V}_d\times\mathfrak{W}\) and any solution to observer (367) where \( y\) is the output of system (365), we have for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{equation} |z(t,j)-\hat{z}(t)| \leq c_1e^{-\lambda t}|z(0,0)-\hat{z}(0)| + \max_{s\in [0,t]}e^{-\lambda (t-s)} \left(c_2|v_c(s)| + c_3|w(s)| \right) + c_4\sum_{k = 0}^{j-1}e^{-\lambda (t-t_k)}|v_{d,k}|. \end{equation} \]

(528)

Now, since both \( T^*\) and \( \tilde{T}^*\) take values in \( \mathcal{X}\) , we deduce from (513) that for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{align} d_\mathcal{A}((x(t,j),\hat{x}(t))) &\leq \alpha\left(|T(x(t,j))-T(\hat{x}(t))| \right)\notag \end{align} \]

(529)

\[ \begin{align} &\leq \alpha\left(|T(x(t,j))-T(\tilde{T}^*(\hat{z}(t)))| \right)\notag\\ & \leq \alpha\left(|z(t,j)-T(T^*(\hat{z}(t)))+T(T^*(\hat{z}(t)))-T(\tilde{T}^*(\hat{z}(t)))| \right)\notag\\ & \leq \alpha\left(|z(t,j)-\Pi_{T(\mathcal{X})}(\hat{z}(t))|+|T(T^*(\hat{z}(t)))-T(\tilde{T}^*(\hat{z}(t)))| \right)\notag\\ & \leq \alpha\left(|z(t,j)-\Pi_{T(\mathcal{X})}(\hat{z}(t))|+L_T|T^*(\hat{z}(t))-\tilde{T}^*(\hat{z}(t))| \right)\notag\\ & \leq \alpha\left(|z(t,j)-\Pi_{T(\mathcal{X})}(\hat{z}(t))|+L_T\delta_T \right). \\ \\\end{align} \]

(530)

From (516), we get for all \( (t,j)\in \rm{dom} x\cap (\mathbb{R}_{\geq 0}\times\mathbb{Z}_{\geq 0})\) ,

\[ \begin{equation} d_\mathcal{A}((x(t,j),\hat{x}(t)))\leq \alpha (2|z(t,j)-\hat{z}(t)| + L_T\delta_T). \end{equation} \]

(531)

Combining (531) with (528) and (521) (with \( z(0,0)\) replacing \( z(0)\) and which holds since \( \hat{z}(0) \in T(\mathcal{X})\) ), we get Theorem 20.

6.3.5 Proof of Theorem 21

Pick \( A\) such that \( A + \max\{\rho,\rho_h+\rho^\prime,\rho_h+\rho_f\}\rm{Id}\) is Hurwitz. Pick \( x \in \mathcal{O}\) . From Item (XLI) of Assumption 36, there exists a compact neighborhood around \( x\) denoted \( \mathcal{C}\) , which is contained in \( \mathcal{O}\) and such that all solutions initialized in \( \mathcal{C}\) are \( t\) -backward complete and \( j\) -backward incomplete, with the same number of \( J_m\) jumps. The map \( T\) can thus be rewritten on \( \mathcal{C}\) as

\[ \begin{equation} T(x) = T_m(x) + \sum_{j=J_m}^{0}I_j(x), \end{equation} \]

(532.a)

where

\[ \begin{align} T_m(x) = \int_{-\infty}^{t_{J_m}(x)} e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))) ds, \end{align} \]

(532.b)

\[ \begin{align} \\\end{align} \]

(532.c)

and

\[ \begin{equation} I_j(x) = \int_{t_{j-1}(x)-t_{j}(x)}^0 e^{-A(s+t_j(x))} B h(\Psi_f(X(x,t_{j}(x),j),s)) ds, \end{equation} \]

(532.d)

with \( t_0(x) = 0\) . From Lemma 21 under Assumption 35 (stopping at \( J_m\) from Item (XLI) of Assumption 36), we have that, for all \( j \in \mathbb{Z}_{\leq 0} \cap \mathbb{Z}_{\geq J_m}\) , \( x \mapsto t_{j}(x)\) and \( x \mapsto X(x,t_{j(x)},j)\) are \( C^1\) on \( \mathcal{O}\) and therefore on \( \mathcal{C}\) , so that, with \( h\) being \( C^1\) , \( x \mapsto I_j(x)\) is also \( C^1\) on \( \mathcal{C}\) . This means \( T-T_m\) is \( C^1\) on \( \mathcal{O}\) and therefore on \( \mathcal{C}\) . Now it remains to show that \( T_m\) is \( C^1\) on \( \mathcal{C}\) . Since \( t_{J_m}\) is \( C^1\) , it is in particular continuous and thus lower-bounded on \( \mathcal{C}\) by some \( \underline{t} \leq 0\) . Consider the sequence \( (T_{m,k})_{k \in \mathbb{N}}\) defined as

\[ \begin{equation} T_{m,k}(x) = \int_{\underline{t} - k}^{t_{J_m}(x)} e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))) ds, ~~ k \in \mathbb{N}. \end{equation} \]

(533)

For each \( k \in \mathbb{N}\) , since the functions \( x \mapsto t_{J_m}(x)\) and \( x \mapsto X(x,t_{J_m(x)},J_m)\) are \( C^1\) on \( \mathcal{C}\) , we have that \( T_{m,k}\) is \( C^1\) on \( \mathcal{C}\) . We are going to show that \( (T_{m,k})_{k \in \mathbb{N}}\) converges uniformly to \( T_m\) in the sense of the \( C^1\) -norm when \( k \to +\infty\) . From Item (XLII) of Assumption 36, we deduce that for all \( x \in \mathcal{C}\) ,

\[ |h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))| = |Y(x,s)| \leq c e^{-\rho s}. \]

Because \( A + \rho \rm{Id}\) is Hurwitz, there exist \( c_1>0\) and \( c_2 > 0\) such that for any \( q \in \mathbb{N}\) , for all \( k \in \mathbb{N}\) , and for all \( x \in \mathcal{C}\) ,

\[ \begin{align*} \left|T_{m,k+q}(x) - T_{m,k}(x)\right| & = \Bigg|\int_{\underline{t} - (k+q)}^{t_{J_m}(x)} e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))) ds\\ &~~{}-\int_{\underline{t} - k}^{t_{J_m}(x)} e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))) ds\Bigg|\\ & = \left|\int_{\underline{t} - (k+q)}^{\underline{t} - k} e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))) ds\right|\\ & \leq\int_{\underline{t} - (k+q)}^{\underline{t} - k}\left| e^{-As} B h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))\right| ds\\ & \leq c \|B\|\int_{\underline{t} - (k+q)}^{\underline{t} - k} e^{-(A+\rho\rm{Id})s} ds\\ & \leq c_1\left(e^{-c_2k}- e^{-c_2 (k+q)}\right). \end{align*} \]

On the other hand, we have

\[ \begin{equation} \frac{\partial T_{m,k}}{\partial x}(x) = \mathcal{J}_a(x) + \mathcal{J}_{b,k}(x), \end{equation} \]

(534.a)

where

\[ \begin{align} \mathcal{J}_a(x) & = e^{-At_{J_m}(x)} B h(X(x,t_{J_m}(x),J_m))\frac{\partial t_{J_m}}{\partial x}(x), \end{align} \]

(534.b)

\[ \begin{align} \mathcal{J}_{b,k}(x) & = \int_{t_{J_m}(x)-k}^{t_{J_m}(x)} e^{-As} B \frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))) ds. \\ \\\end{align} \]

(534.c)

Now let us exhibit bounds on \( \mathcal{J}_{b,k}(x)\) . We compute the term in its integral

\[ \begin{multline} \frac{\partial}{\partial x} (h(\Psi_f(X(x, t_{J_m}(x), J_m), s - t_{J_m}(x)))) = \frac{\partial h}{\partial x} (X(x,s, J_m)) \times \\ \left(\frac{\partial \Psi_f}{\partial x}(X(x, t_{J_m}(x), J_m),s-t_{J_m}(x))\frac{\partial}{\partial x}(X(x,t_{J_m}(x),J_m)) - f(X(x,s, J_m))\frac{\partial t_{J_m}}{\partial x}(x) \right), \end{multline} \]

(535)

where we used in the last term that \( \Psi_f(X(x, t_{J_m}(x), J_m), s - t_{J_m}(x))=X(x,s, J_m)\) . We now bound each term of (535). Since the maps \( x \mapsto t_{J_m}(x)\) and \( x \mapsto X(x,t_{J_m(x)},J_m)\) are \( C^1\) on \( \mathcal{C}\) , the maps \( x \mapsto \frac{\partial t_{J_m}}{\partial x}(x)\) and \( x \mapsto \frac{\partial}{\partial x}(X(x,t_{J_m(x)},J_m))\) are continuous and so bounded on \( \mathcal{C}\) which is compact. Applying Lemma 36 in Section 6.3.7, we deduce that, for each \( x\in \mathcal{C}\) , the term \( s\mapsto\frac{\partial \Psi_f}{\partial x}(X(x, t_{J_m}(x), J_m),s-t_{J_m}(x)) =: \Phi_f(s)\) is solution in backward time to

\[ \frac{\partial\Phi_f}{\partial s}(s) = \frac{\partial f}{\partial x}(X(x,s, J_m))\Phi_f(s), ~~ \forall s\leq t_{J_m}(x), \]

initialized as \( \Phi_f(t_{J_m}(x)) = \rm{Id}\) . From Item (XLIII) of Assumption 36, we obtain

\[ \left|\frac{\partial \Psi_f}{\partial x}(X(x, t_{J_m}(x), J_m),s-t_{J_m}(x))\right| \leq e^{-\rho^\prime s}, ~~ s \in (-\infty,t_{J_m}(x)]. \]

Combining these ingredients with the rest of Item (XLIII) of Assumption 36, from (535), we deduce that there exists \( c^\prime > 0\) such that for all \( x \in \mathcal{C}\) , we have

\[ \begin{equation} \left|\frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))))\right| \leq c^\prime e^{-(\rho_h + \max\{\rho^\prime,\rho_f\}) s}, ~~ s \in (-\infty,t_{J_m}(x)]. \end{equation} \]

(536)

Thanks to this and because \( A + (\rho_h + \max\{\rho^\prime,\rho_f\}) \rm{Id}\) is Hurwitz, there exist \( c_3>0\) and \( c_4 > 0\) such that for any \( q \in \mathbb{N}\) , for all \( k \in \mathbb{N}\) , and for all \( x \in \mathcal{C}\) ,

\[ \begin{align*} \left|\mathcal{J}_{b,k+q}(x)-\mathcal{J}_{b,k}(x)\right|& = \Bigg|\int_{\underline{t}-(k+q)}^{t_{J_m}(x)} e^{-As} B \frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))) ds \\ & ~~{}- \int_{\underline{t}-k}^{t_{J_m}(x)} e^{-As} B \frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))) ds\Bigg|\\ &=\left|\int_{\underline{t}-(k+q)}^{\underline{t}-k} e^{-As} B \frac{\partial}{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x)))) ds\right|\\ & \leq \int_{\underline{t}-(k+q)}^{\underline{t}-k} \left|e^{-As} B \frac{\partial}{\partial x}(h(\Psi_f(X(x,t_{J_m}(x),J_m),s-t_{J_m}(x))))\right| ds \\ &\leq c^\prime\|B\|\int_{\underline{t}-(k+q)}^{\underline{t} - k} \left|e^{-(A+(\rho_h + \max\{\rho^\prime,\rho_f\})\rm{Id})s} B \right|ds \\ &\leq c_3\left(e^{-c_4k}-e^{-c_4(k+q)}\right). \end{align*} \]

Using the triangle inequality and the bounds obtained above, we have for any \( q \in \mathbb{N}\) , for all \( k \in \mathbb{N}\) , and for all \( x \in \mathcal{C}\) ,

\[ \begin{align*} \|T_{m,k+q} - T_{m,k}\|_{C^1} & = \max_{x\in \mathcal{C}}\left|T_{m,k+q}(x) - T_{m,k}(x)\right| + \max_{x\in \mathcal{C}}\left|\frac{\partial T_{m,k+q}}{\partial x}(x) - \frac{\partial T_{m,k}}{\partial x}(x)\right|\\ &\leq \max_{x\in \mathcal{C}}\left|T_{m,k+q}(x) - T_{m,k}(x)\right| + \max_{x\in \mathcal{C}}\left|\mathcal{J}_{b,k+q}(x) - \mathcal{J}_{b,k}(x)\right|\\ & \leq c_1\left(e^{-c_2k}-e^{-c_2 (k + q)}\right) + c_3\left(e^{-c_4k}-e^{-c_4(k+q)}\right)\\ & \leq c_1 e^{-c_2k} + c_3^{-c_4k}. \end{align*} \]

Because this quantity goes to \( 0\) when \( k \to +\infty\) independently of \( q \in \mathbb{N}\) and \( x \in \mathcal{C}\) , \( (T_{m,k})_{k \in \mathbb{N}}\) is a Cauchy sequence in the sense of the \( C^1\) -norm. Because the space of \( C^1\) functions on the compact set \( \mathcal{C}\) equipped with the \( C^1\) -norm is a Banach space [210, Chapter 5, Theorem 1], it is complete and so \( (T_{m,k})_{k \in \mathbb{N}}\) converges uniformly to a function that is also in this space. By definition, this limit is \( T_m\) , which is thus \( C^1\) on \( \mathcal{C}\) . With this being valid around any point \( x\in \mathcal{O}\) , we conclude that \( T_m\) and thus \( T\) is \( C^1\) on \( \mathcal{O}\) .

6.3.6 Proof of Theorem 22

Pick \( A\) such that \( A + \max\left\{\rho+\frac{\rho_t}{\tau_m} + \rho_t^\prime,\rho^\prime + \frac{\rho_h + \rho_X}{\tau_m}+\rho_h^\prime+\rho_X^\prime\right\}\rm{Id}\) is Hurwitz. Pick \( x \in \mathcal{O}\) . Since \( \mathcal{O}\) is open, there exists a compact neighborhood around \( x\) denoted \( \mathcal{C}\) , which is contained in \( \mathcal{O}\) . From Item (XLIV) and Item (XLV) of Assumption 37, all solutions initialized in \( \mathcal{C}\) are \( j\) -backward complete and \( t\) -backward complete. The map \( T\) can thus be rewritten on \( \mathcal{C}\) as

\[ \begin{equation} T(x) = \sum_{j=-\infty}^{0}I_j(x), \end{equation} \]

(537.a)

where

\[ \begin{equation} I_j(x) = \int_{t_{j-1}(x)-t_{j}(x)}^0 e^{-A(s+t_j(x))} B h(\Psi_f(X(x,t_{j}(x),j),s)) ds, \end{equation} \]

(537.b)

with \( t_0(x) = 0\) . From Lemma 21 under Assumption 35 (up to \( -\infty\) from Item (XLIV) of Assumption 37) and since \( h\) is \( C^1\) , it follows that each map \( x \mapsto I_j(x)\) is well-defined and is \( C^1\) on \( \mathcal{O}\) and therefore on \( \mathcal{C}\) , for all \( j \in \mathbb{Z}_{\leq 0}\) . Define the sequence of truncated maps \( (T_k)_{k \in \mathbb{Z}_{\leq 0}}\) where

\[ \begin{equation} T_k(x) = \sum_{j=k}^{0}I_j(x). \end{equation} \]

(538)

Because each \( T_k\) is a finite sum of functions that are \( C^1\) on \( \mathcal{O}\) , it is also \( C^1\) on \( \mathcal{O}\) and therefore on \( \mathcal{C}\) . We are going to show that \( (T_k)_{k\in \mathbb{Z}_{\leq 0}}\) converges uniformly to \( T\) in the sense of the \( C^1\) -norm. FroM Item (XLVI) of Assumption 37, we deduce that for all \( x \in \mathcal{C}\) , for all \( j \in \mathbb{Z}_{\leq 0}\) , and for all \( s \in [t_{j-1}(x)-t_{j}(x), 0]\) ,

\[ |h(\Psi_f(X(x,t_{j}(x),j),s))| = |Y(x,s+t_j(x))| \leq c e^{-\rho(s+t_j(x))}. \]

Because \( A + \rho \rm{Id}\) is Hurwitz, there exist \( c_1>0\) and \( c_2 > 0\) such that for all \( x \in \mathcal{C}\) and for all \( j \in \mathbb{Z}_{\leq 0}\) ,

\[ \begin{equation} |I_j(x)| \leq c_1\left(e^{c_2t_{j}(x)} - e^{c_2t_{j-1}(x)}\right). \end{equation} \]

(539)

Because the map \( x \mapsto I_j(x)\) , \( j \in \mathbb{Z}_{\leq 0}\) is \( C^1\) on \( \mathcal{C}\) , we are able to define the sequence of Jacobians \( \left(\frac{\partial T_k}{\partial x}\right)_{k \in \mathbb{Z}_{\leq 0}}\) where

\[ \begin{equation} \frac{\partial T_k}{\partial x}(x) = \sum_{j=k}^{0}\frac{\partial I_j}{\partial x}(x) = \sum_{j=k}^{0} \left(\mathcal{I}_{a,j}(x) + \mathcal{I}_{b,j}(x)+\mathcal{I}_{c,j}(x)\right), \end{equation} \]

(540.a)

where

\[ \begin{align} \mathcal{I}_{a,j}(x) & = {}-e^{-At_{j-1}(x)} B h(X(x,t_{j-1}(x),j))\left(\frac{\partial t_{j-1}}{\partial x}(x) - \frac{\partial t_{j}}{\partial x}(x)\right), \end{align} \]

(540.b)

\[ \begin{align} \mathcal{I}_{b,j}(x) & = \int_{t_{j-1}(x)-t_{j}(x)}^0 e^{-A(s+t_j(x))} B \frac{\partial }{\partial x}(h(\Psi_f(X(x,t_{j}(x),j),s))) ds,\\ \mathcal{I}_{c,j}(x) & = -\int_{t_{j-1}(x)-t_{j}(x)}^0 e^{-A(s+t_j(x))} AB h(\Psi_f(X(x,t_{j}(x),j),s)) \frac{\partial t_j}{\partial x}(x) ds. \\ \\\end{align} \]

(540.c)

Now let us exhibit bounds on these. For \( \mathcal{I}_{a,j}(x)\) , thanks to Items (XLVI)(XLVII), and (XLV) of Assumption 37, there exist \( c_3 > 0\) and \( c_4 > 0\) such that for all \( x \in \mathcal{C}\) and for all \( j \in \mathbb{Z}_{\leq 0}\) ,

\[ \begin{align*} |\mathcal{I}_{a,j}(x)|&\leq\left|e^{-At_{j-1}(x)}Bh(X(x,t_{j-1}(x),j))\right|2c_te^{-\rho_t (j-1) - \rho_t^\prime t_{j-1}(x)} \\ &\leq 2c_te^{-\left(\frac{\rho_t}{\tau_m} + \rho_t^\prime\right)t_{j-1}(x) + \rho_tN}\left|e^{-(A+\rho \rm{Id})t_{j-1}(x)}B\right|\left|e^{\rho t_{j-1}(x)}Y(x,t_{j-1}(x))\right|\\ &\leq 2c_te^{\rho_tN}\left|e^{-\left(A+\left(\rho+\frac{\rho_t}{\tau_m} + \rho_t^\prime\right) \rm{Id}\right)t_{j-1}(x)}B\right|c\\ & \leq c_3 e^{c_4 \tau_m (j-1)}, \end{align*} \]

because \( A + \left(\rho + \frac{\rho_t}{\tau_m} + \rho_t^\prime\right)\rm{Id}\) is Hurwitz. For \( \mathcal{I}_{b,j}(x)\) , we compute the terms in its integral

\[ \begin{equation} \frac{\partial}{\partial x} (h(\Psi_f(X(x,t_{j}(x),j),s))) = \frac{\partial h}{\partial x} (X(x,s, j)) \frac{\partial \Psi_f}{\partial x}(X(x, t_j(x), j),s)\frac{\partial}{\partial x}(X(x,t_j(x),j)) . \end{equation} \]

(541)

We now bound each term of (541). Applying Lemma 36 in Section 6.3.7, we deduce that for each \( x\in \mathcal{C}\) and for each \( j\in \mathbb{Z}_{\leq 0}\) , the term \( \frac{\partial \Psi_f}{\partial x}(X(x, t_j(x), j),s) =: \Phi_f(s)\) is solution in backward time to

\[ \frac{\partial\Phi_f}{\partial s}(s) = \frac{\partial f}{\partial x}(X(x,s, j))\Phi_f(s), ~~ s \in [t_{j-1}(x),t_j(x)], \]

initialized as \( \Phi_f(t_j(x)) = \rm{Id}\) . From Item (L) of Assumption 37, we have for all \( x\in \mathcal{C}\) , for all \( j\in \mathbb{Z}_{\leq 0}\) , and for all \( s\in [t_{j-1}(x),t_j(x)]\) ,

\[ \left|\frac{\partial \Psi_f}{\partial x}(X(x, t_j(x), j),s)\right| \leq e^{-\rho^\prime (s+t_j(x))}. \]

Because \( A + \left(\rho^\prime + \frac{\rho_h + \rho_X}{\tau_m}+\rho_h^\prime+\rho_X^\prime\right)\rm{Id}\) is Hurwitz, there exist \( c_5>0\) and \( c_6 > 0\) such that for all \( x \in \mathcal{C}\) and for all \( j \in \mathbb{Z}_{\leq 0}\) ,

\[ \begin{align*} |\mathcal{I}_{b,j}(x)| &\leq c_h c_X \|B\|\int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-A(s+t_j(x))-\rho_h j -\rho_h^\prime t_j(x) -\rho^\prime (s+t_j(x))-\rho_X j - \rho_X^\prime t_j(x)} \right|ds \\ &\leq c_h c_X \|B\|\int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-(A + \rho^\prime\rm{Id})(s+t_j(x))-\left(\frac{\rho_h + \rho_X}{\tau_m}+\rho_h^\prime+\rho_X^\prime\right)(s+t_j(x))} \right|ds\\ &\leq c_h c_X \|B\|\int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-\left(A + \left(\rho^\prime + \frac{\rho_h + \rho_X}{\tau_m}+\rho_h^\prime+\rho_X^\prime\right)\rm{Id}\right)(s+t_j(x))} \right|ds \\ &\leq c_5\left(e^{c_6t_{j}(x)} - e^{c_6t_{j-1}(x)}\right), \end{align*} \]

because \( -\left(\frac{\rho_h + \rho_X}{\tau_m}+\rho_h^\prime+\rho_X^\prime\right)s \geq 0\) with \( s \leq 0\) . For \( \mathcal{I}_{c,j}(x)\) , thanks to Item (XLVI) and Item (XLVII) of 37 and because \( A + \left(\rho+\frac{\rho_t}{\tau_m} + \rho_t^\prime\right) \rm{Id}\) is Hurwitz, there exists \( c_7>0\) such that for all \( x \in \mathcal{C}\) and for all \( j \in \mathbb{Z}_{\leq 0}\) ,

\[ \begin{align*} \left|\mathcal{I}_{c,j}(x)\right| &\leq c c_t \int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-(A+\rho \rm{Id})(s+t_j(x))-\left(\frac{\rho_t}{\tau_m} + \rho_t^\prime\right)t_{j}(x) + \rho_tN} B\right| ds \\ &\leq c c_t e^{\rho_tN} \int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-(A+\rho \rm{Id})(s+t_j(x))-\left(\frac{\rho_t}{\tau_m} + \rho_t^\prime\right)(s+t_{j}(x))} B\right| ds \\ &\leq c c_t e^{\rho_tN}\int_{t_{j-1}(x)-t_{j}(x)}^0 \left|e^{-\left(A+\left(\rho+\frac{\rho_t}{\tau_m} + \rho_t^\prime\right) \rm{Id}\right)(s+t_j(x))} B\right| ds \\ &\leq c_7\left(e^{c_4t_{j}(x)} - e^{c_4t_{j-1}(x)}\right), \end{align*} \]

because \( -\left(\frac{\rho_t}{\tau_m}+\rho_t^\prime\right)s \geq 0\) with \( s \leq 0\) . Now we show that \( (T_k)_{k \in \mathbb{Z}_{\leq 0}}\) is a Cauchy sequence in the sense of the \( C^1\) -norm on \( \mathcal{C}\) . Using the triangle inequality and the bounds obtained above, we have for any \( k \in \mathbb{Z}_{\leq 0}\) and any \( q \in \mathbb{N}\) , for all \( x \in \mathcal{C}\) ,

\[ \begin{align*} \left\|T_{k-q} - T_k\right\|_{C^1} & = \max_{x\in \mathcal{C}}\left|T_{k-q}(x) - T_k(x)\right| + \max_{x\in \mathcal{C}}\left|\frac{\partial T_{k-q}}{\partial x}(x) - \frac{\partial T_k}{\partial x}(x)\right|\\ & = \max_{x\in \mathcal{C}}\left|\sum_{j = k-q}^{0}I_j(x)-\sum_{j = k}^{0}I_j(x)\right| \\ &~~{}+ \max_{x\in \mathcal{C}}\left|\sum_{j = k-q}^{0}\left(\mathcal{I}_{a,j}(x)+\mathcal{I}_{b,j}(x)+\mathcal{I}_{c,j}(x)\right)-\sum_{j = k}^{0}\left(\mathcal{I}_{a,j}(x)+\mathcal{I}_{b,j}(x)+\mathcal{I}_{c,j}(x)\right)\right|\\ & = \max_{x\in \mathcal{C}}\left|\sum_{j = k-q}^{k}I_j(x)\right| + \max_{x\in \mathcal{C}}\left|\sum_{j = k-q}^{k}\left(\mathcal{I}_{a,j}(x)+\mathcal{I}_{b,j}(x)+\mathcal{I}_{c,j}(x)\right)\right|\\ & \leq \max_{x\in \mathcal{C}}\sum_{j = k-q}^{k}\left|I_j(x)\right| + \max_{x\in \mathcal{C}}\sum_{j = k-q}^{k}\left(\left|\mathcal{I}_{a,j}(x)\right|+\left|\mathcal{I}_{b,j}(x)\right|+\left|\mathcal{I}_{c,j}(x)\right|\right)\\ & \leq \max_{x\in \mathcal{C}}\left(c_1\sum_{j = k-q}^{k}\left(e^{c_2t_{j}(x)} - e^{c_2t_{j-1}(x)}\right)\right)+ \max_{x\in \mathcal{C}}\Bigg(c_3\sum_{j = k-q}^{k}e^{c_4 \tau_m (j-1)} \\ &~~{} + c_5\sum_{j = k-q}^{k}\left(e^{c_6t_{j}(x)} - e^{c_6t_{j-1}(x)}\right)+ c_7\sum_{j = k-q}^{k}\left(e^{c_4t_{j}(x)} - e^{c_4t_{j-1}(x)}\right)\Bigg)\\ & \leq c_1\max_{x\in \mathcal{C}}\left(e^{c_2t_k(x)} - e^{c_2t_{k-q-1}(x)}\right)+ \max_{x\in \mathcal{C}}\Bigg(c_3 e^{c_4 \tau_m k}\frac{1- e^{-c_4 \tau_m(q+1)}}{1- e^{-c_4 \tau_m}} \\ &~~{} + c_5\left(e^{c_6t_k(x)} - e^{c_6t_{k-q-1}(x)}\right)+ c_7\left(e^{c_4t_k(x)} - e^{c_4t_{k-q-1}(x)}\right)\Bigg)\\ & \leq c_1\max_{x\in \mathcal{C}}e^{c_2t_k(x)} + \max_{x\in \mathcal{C}}\left(\frac{c_3}{1- e^{-c_4 \tau_m}} e^{c_4 \tau_m k} + c_5e^{c_6t_k(x)}+ c_7e^{c_4t_k(x)}\right)\\ & \leq c_1e^{c_2(\tau_m k + \tau_m N)} + \frac{c_3}{1- e^{-c_4 \tau_m}} e^{c_4 \tau_m k} + c_5e^{c_6(\tau_m k + \tau_m N)} + c_7e^{c_4(\tau_m k + \tau_m N)}. \end{align*} \]

Because this quantity goes to \( 0\) when \( k \to -\infty\) independently of \( q \in \mathbb{N}\) and \( x \in \mathcal{C}\) , \( (T_k)_{k \in \mathbb{Z}_{\leq 0}}\) is a Cauchy sequence in the sense of the \( C^1\) -norm. Because the space of \( C^1\) functions on the compact set \( \mathcal{C}\) equipped with the \( C^1\) -norm is a Banach space [210, Chapter 5, Theorem 1], it is complete and so \( (T_k)_{k \in \mathbb{Z}_{\leq 0}}\) converges uniformly to a function that is also in this space. By definition, this limit is \( T\) , which is thus \( C^1\) on \( \mathcal{C}\) . With this being valid around any point \( x\in \mathcal{O}\) , we conclude that \( T\) is \( C^1\) on \( \mathcal{O}\) .

6.3.7 Technical Lemmas

Lemma 34 (Outputs from \( x \in \mathcal{D}\) and \( g(x)\) are the same)

Suppose Assumption 30 and Assumption 31 hold. For every \( x\in D\) , we have

\[ \begin{equation} \breve{Y}(g(x),s)=\breve{Y}(x,s), ~~ \forall s\in (-\infty,t^+(x)). \end{equation} \]

(542)

Proof. Let \( x\in D\) . First we prove that \( t^{-}(g(x)) = t^{-}(x)\) . For this, let \( \phi_1\) and \( \phi_2\) be the maximal solutions to system (325) given in Item (XXXI) of Assumption 30 such that \( \phi_1(0,0)=x\) and \( \phi_2(0,0)=g(x)\) , respectively. By uniqueness of solutions in Item (XXXI) of Assumption 30, \( (0,1)\in \rm{dom} \phi_1\) and \( \phi_1(0,1)=g(x)\) . Define the function \( \phi_2^\prime\) on \( \{(t,j)\in \mathbb{R}\times\mathbb{Z} \: : \: (t,j+1)\in \rm{dom} \phi_1 \}\) as

\[ \phi_2^\prime(t,j)=\phi_1(t,j+1). \]

We observe that \( (0,0)\in\rm{dom} \phi_2^\prime\) since \( (0,1)\in\rm{dom}\phi_1\) and \( \phi_2^\prime(0,0)=\phi_1(0,1)=g(\phi_1(0,0))=g(x)\) , therefore \( \phi_2^\prime\) is a solution to system (325) starting at \( g(x)\) , and satisfies \( \inf\rm{dom}_t \phi_2^\prime = \inf\rm{dom}_t \phi_1 = t^{-}(x)\) . By the maximality of \( \phi_2\) , we conclude that \( t^{-}(g(x))\leq t^{-}(x)\) and thus by the uniqueness of solutions, \( \phi_2^\prime(t,j)=\phi_2(t,j)\) for all \( (t,j)\in\rm{dom} \phi_2^\prime\) . Now by contradiction, suppose that \( t^{-}(g(x))<t^{-}(x)\) , then there exists \( \tau \in\rm{dom}_t\phi_2\setminus\rm{dom}_t\phi_1,\) with \( t^{-}(g(x))<\tau<t^{-}(x)\) . Consider the trajectory \( \phi_1^\prime\) defined on \( \{(t,j)\in \mathbb{R}\times\mathbb{Z} \: : \: (t,j-1)\in \rm{dom} \phi_2 \}\) as

\[ \phi_1^\prime(t,j)=\phi_2(t,j-1). \]

By a similar argument as before, we see that \( \phi_1^\prime\) is a solution to system (325) initialized as \( x\) such that \( \rm{dom}_t \phi_1^\prime=\rm{dom}_t \phi_2\) . By the maximality of \( \phi_1\) , we have \( \rm{dom}_t\phi_1^\prime\subset\rm{dom}_t \phi_1\) , but this is a contradiction because \( \tau\in\rm{dom}_t\phi_1^\prime\setminus\rm{dom}_t \phi_1\) . Therefore we conclude that \( t^{-}(g(x))=t^{-}(x)\) . Similarly, \( t^{+}(g(x))=t^{+}(x)\) and \( \rm{dom}_t \phi_1 = \rm{dom}_t \phi_2\) .

Now we show that \( \breve{Y}(x,s)=\breve{Y}(g(x),s)\) for all \( s\in(-\infty,t^+(x))\) . As noted before, \( \phi_2^\prime(s,j)=\phi_2(s,j)\) for all \( (s,j)\in\rm{dom}\phi_2^\prime\) . By definition of \( \phi_2^\prime\) , we have

\[ \phi_1(s,j+1)=\phi_2(s,j),~~\forall (s,j)\in\rm{dom} \phi_2, \]

so that

\[ h(\phi_1(s,j+1))=h(\phi_2(s,j)),~~\forall (s,j)\in\rm{dom} \phi_2. \]

By definition in (335), we have \( Y(x,s)=Y(g(x),s)\) for all \( s\in (t^{-}(x),,t^+(x))\) . It follows then directly that for all \( s\in(-\infty,t^+(x))\) ,

\[ \begin{align*} \breve{Y}(x,s)&=\lim_{\tau^{+}\to t^{-}(x)}Y(x,\tau)\\ &=\lim_{\tau^{+}\to t^{-}(x)}Y(g(x),\tau)\\ &=\lim_{\tau^{+}\to t^{-}(g(x))}Y(g(x),\tau)\\ &=\breve{Y}(g(x),s). \end{align*} \]

The conclusion follows. \( \blacksquare\)

Lemma 35 (Upper-bounding a function by another one)

Consider a compact set \( \mathcal{C} \subset \mathbb{R}^{n}\) and two continuous maps \( \alpha_1: \mathbb{R}^n \to \mathbb{R}^r\) and \( \alpha_2: \mathbb{R}^n \to \mathbb{R}^q\) , for some integers \( (n,r,q)\) , such that

\[ \begin{equation} \alpha_2(x) = 0 \implies \alpha_1(x) = 0, ~~ \forall x \in \mathcal{C}. \end{equation} \]

(543)

Then, there exists a (concave) class-\( \mathcal{K}\) function \( \alpha\) such that

\[ \begin{equation} |\alpha_1(x)| \leq \alpha(|\alpha_2(x)|), ~~ \forall x \in \mathcal{C}. \end{equation} \]

(544)

Proof. This is an adaptation and combination of the proofs of [39, Lemmas A.6 and A.9]. Consider the map defined by

\[ \alpha_0(s) = \max_{\substack{x \in \mathcal{C}\\ |\alpha_2(x)|\leq s}}|\alpha_1(x)|. \]

This defines properly a non-negative valued function that is non-decreasing, satisfies (544), and is such that \( \alpha_0(0) = 0\) . Now, we show that \( \alpha_0\) is continuous at \( s=0\) . Let \( (s_k)_{k \in \mathbb{N}}\) be a sequence converging to \( 0\) . For each \( k \in \mathbb{N}\) , there exists \( x_k \in \mathcal{C}\) such that \( |\alpha_2(x_k)| \leq s_k\) and \( \alpha_0(s_k) = |\alpha_1(x_k)|\) . Since the sequence \( (x_k)_{k \in \mathbb{N}}\) is such that \( x_k \in \mathcal{C}\) , which is compact, for all \( k \in \mathbb{N}\) , it admits an accumulation point \( x^*\) which verifies \( |\alpha_2(x^*)| = 0\) thanks to the continuity of \( \alpha_2\) . From (543), we get \( \alpha_1(x^*) = 0\) . It follows that \( \alpha_0(s_k)\) tends to \( 0\) as \( k \to +\infty\) and so \( \alpha_0\) is continuous at \( 0\) . Now to regularize \( \alpha_0\) , consider the map defined by

\[ \begin{equation} \overline{\alpha}_0(s) = \left\{ \begin{array}{@{}l@{~~}l@{}} \displaystyle \frac{1}{s}\int_s^{2s} \alpha_0(s)ds, & s > 0 \\ 0, & s = 0 \end{array}\right. \end{equation} \]

(545)

which is continuous, strictly increasing, and is such that \( \alpha_0(s) \leq \overline{\alpha}_0(s)\) for all \( s \geq 0\) . Taking \( \overline{s} = \max_{x \in \mathcal{C}}|\alpha_2(x)|\) , which is well-defined since \( \alpha_2\) is continuous on \( \mathcal{C}\) which is compact, from [40], we deduce that there exists a (concave) class-\( \mathcal{K}\) function \( \alpha\) such that for all \( s \in [0,\overline{s}]\) , we have \( \overline{\alpha}_0(s) \leq \alpha(s)\) . Then, (544) follows. \( \blacksquare\)

Lemma 36 (Dynamics of \( \frac{\partial \Psi_f}{\partial \xi_0}(\xi_0,\tau)\))

This lemma is from [207, Chapter 1, Theorem 3.3]. Consider the system

\[ \begin{equation} \dot{\xi} = f(\xi), \end{equation} \]

(546)

where \( f\) is continuous and of class \( C^1\) with respect to \( \xi\) and define \( \Psi_f(\xi_0,\tau)\) as the solution to system (546) initialized as \( \xi_0\) flowing during \( \tau\) time unit(s), and a modified time \( t\mapsto \tau(t)\) such that \( \dot{\tau}=1\) . Let \( \Phi_f(\xi_0,\tau) = \frac{\partial \Psi_f}{\partial \xi_0}(\xi_0,\tau)\) . Then, \( (\xi,\Phi_f)\) is solution to the dynamics

\[ \begin{equation} \dot{\xi} = f(\xi),~~ \dot{\Phi}_f = \frac{\partial f}{\partial \xi}(\xi) \Phi_f, \end{equation} \]

(547)

initialized as \( (\xi_0, \rm{Id})\) .

 

References

[1] P. Bernard and V. Andrieu and D. Astolfi Observer Design for Continuous-Time Dynamical Systems Annual Reviews in Control 2022 53 224-248 https://doi.org/10.1016/j.arcontrol.2021.11.002

[2] K. A. Hamed and A. D. Ames and R. D. Gregg Observer-Based Feedback Controllers for Exponential Stabilization of Hybrid Periodic Orbits: Application to Underactuated Bipedal Walking 2018 Annual American Control Conference (ACC) 2018 1438-1445 10.23919/ACC.2018.8431607

[3] M. Vigne and A. E. Khoury and M. Pétriaux and F. Di Meglio and N. Petit MOVIE: A Velocity-Aided IMU Attitude Estimator for Observing and Controlling Multiple Deformations on Legged Robots IEEE Robotics and Automation Letters 2022 7 2 3969-3976 10.1109/LRA.2022.3149025

[4] M. Menner and K. Berntorp Simultaneous State Estimation and Contact Detection for Legged Robots by Multiple-Model Kalman Filtering 2024 European Control Conference (ECC) 2024 2768-2773 10.23919/ECC64448.2024.10590751

[5] Y. Wang and R. Rajamani and D. M. Bevly Observer Design for Parameter Varying Differentiable Nonlinear Systems, With Application to Slip Angle Estimation IEEE Transactions on Automatic Control 2017 62 4 1940-1945 10.1109/TAC.2016.2587385

[6] T. P. Pham and O. Sename and L. Dugard Unified \( \mathcal{H}_{\infty}\) Observer for a Class of Nonlinear Lipschitz Systems: Application to a Real ER Automotive Suspension IEEE Control Systems Letters 2019 3 4 817-822 10.1109/LCSYS.2019.2919813

[7] C. Lascu and I. Boldea and F. Blaabjerg Comparative Study of Adaptive and Inherently Sensorless Observers for Variable-Speed Induction-Motor Drives IEEE Transactions on Industrial Electronics 2006 53 1 57-65 10.1109/TIE.2005.862314

[8] P. Bernard Luenberger Observers for Nonlinear Controlled Systems 56th IEEE Conference on Decision and Control 2017 3676-3681 10.1109/CDC.2017.8264200

[9] J. Li and K. Pan and D. Zhang and Q. Su Robust Fault Detection and Estimation Observer Design for Switched Systems Nonlinear Analysis: Hybrid Systems 2019 34 30-42 https://doi.org/10.1016/j.nahs.2019.05.001

[10] R. Goebel and R. G. Sanfelice and A. R. Teel Hybrid Dynamical Systems: Modeling Stability, and Robustness 2012

[11] Z.-H. Guan and T.-H. Qian and X. Yu On Controllability and Observability for a Class of Impulsive Systems Systems & Control Letters 2002 47 247-257

[12] A. Tanwani and H. Shim and D. Liberzon Comments on “Observability of Switched Linear Systems: Characterization and Observer Design" IEEE Transactions on Automatic Control 2015 60 12 3396-3400

[13] J. Lygeros and K. H. Johansson and S. N. Simic and Jun Zhang and S. S. Sastry Dynamical Properties of Hybrid Automata IEEE Transactions on Automatic Control 2003 48 1 2-17 10.1109/TAC.2002.806650

[14] R. G. Sanfelice and A. R. Teel Lyapunov Analysis of Sample-and-Hold Hybrid Feedbacks Proceedings of the 45th IEEE Conference on Decision and Control 2006 4879-4884 10.1109/CDC.2006.377831

[15] E. R. Westervelt and J. W. Grizzle and D. E. Koditschek Hybrid Zero Dynamics of Planar Biped Walkers IEEE Transactions on Automatic Control 2003 48 1 42-56 10.1109/TAC.2002.806653

[16] A. D. Ames and K. Galloway and K. Sreenath and J. W. Grizzle Rapidly Exponentially Stabilizing Control Lyapunov Functions and Hybrid Zero Dynamics IEEE Transactions on Automatic Control 2014 59 4 876-891 10.1109/TAC.2014.2299335

[17] A. Girard and S. Martin Synthesis for Constrained Nonlinear Systems Using Hybridization and Robust Controllers on Simplices IEEE Transactions on Automatic Control 2012 57 4 1046-1051 10.1109/TAC.2011.2168874

[18] W. P. M. H. Heemels and K. H. Johansson and P. Tabuada An Introduction to Event-Triggered and Self-Triggered Control 51st IEEE Conference on Decision and Control (CDC) 2012 3270-3285 10.1109/CDC.2012.6425820

[19] E. Petri and R. Postoyan and D. Astolfi and D. Nešić and W. P. M. H. Heemels Decentralized Event-Triggered Estimation of Nonlinear Systems Automatica 2024 160 111414 https://doi.org/10.1016/j.automatica.2023.111414

[20] C. de Souza and S. Tarbouriech and I. Queinnec and A. Girard Nonstandard Anti-Windup Approach for Event-Triggered Control Purpose Systems & Control Letters 2024 185 105715 https://doi.org/10.1016/j.sysconle.2024.105715

[21] E. Petri and R. Postoyan and D. Astolfi and D. Nešić and V. Andrieu Hybrid Multi-Observer for Improving Estimation Performance IEEE Transactions on Automatic Control 2024 1-16 10.1109/TAC.2024.3500792

[22] P. Bernard and R. G. Sanfelice Observer Design for Hybrid Dynamical Systems with Approximately Known Jump Times Automatica 2022 141 110225 https://doi.org/10.1016/j.automatica.2022.110225

[23] J. J. B. Biemond and N. van de Wouw and W. P. M. H. Heemels and H. Nijmeijer Tracking Control for Hybrid Systems With State-Triggered Jumps IEEE Transactions on Automatic Control 2013 58 4 876-890 10.1109/TAC.2012.2223351

[24] J. J. B. Biemond and W. P. M. H. Heemels and R. G. Sanfelice and N. van de Wouw Distance Function Design and Lyapunov Techniques for the Stability of Hybrid Trajectories Automatica 2016 73 38-46 https://doi.org/10.1016/j.automatica.2016.07.006

[25] M. Baumann and J. J. B. Biemond and R. I. Leine and N. van de Wouw Synchronization of Impacting Mechanical Systems with a Single Constraint Physica D: Nonlinear Phenomena 2018 362 9-23 https://doi.org/10.1016/j.physd.2017.10.002

[26] J. P. Gauthier and H. Hammouri and S. Othman A Simple Observer for Nonlinear Systems Applications to Bioreactors IEEE Transactions on Automatic Control 1992 37 6 875-880 10.1109/9.256352

[27] G. Besancon Further Results on High Gain Observers for Nonlinear Systems 38th IEEE Conference on Decision and Control 1999 3 2904-2909 vol.3 10.1109/CDC.1999.831376

[28] H. Shim and D. Liberzon and J.-S. Kim Quasi-ISS Reduced-Order Observers and Quantized Output Feedback Proceedings of the 48h IEEE Conference on Decision and Control (CDC) held jointly with 2009 28th Chinese Control Conference 2009 6680-6685 10.1109/CDC.2009.5400400

[29] H. Shim and D. Liberzon Nonlinear Observers Robust to Measurement Disturbances in an ISS Sense IEEE Transactions on Automatic Control 2016 61 1 48-61 10.1109/TAC.2015.2423911

[30] D. Astolfi and A. Alessandri and L. Zaccarian Stubborn and Dead-Zone Redesign for Nonlinear Observers and Filters IEEE Transactions on Automatic Control 2021 66 2 667-682 10.1109/TAC.2020.2989816

[31] F. Asano and M. Yamakita and N. Kamamichi and Z.-W. Luo A Novel Gait Generation for Biped Walking Robots based on Mechanical Energy Constraint IEEE Transactions on Robotics and Automation 2004 20 3 565-573 10.1109/TRA.2004.824685

[32] J. Kim and H. Shim and J. H. Seo State Estimation and Tracking Control for Hybrid Systems by Gluing the Domains IEEE 2019 64 7 3026–3033 10.1109/TAC.2018.2876784

[33] T. N. Dinh and G. Q. B. Tran A Unified KKL-Based Interval Observer for Nonlinear Discrete-Time Systems IEEE Control Systems Letters 2024 8 430-435 10.1109/LCSYS.2024.3394324

[34] T. N. Dinh and G. Q. B. Tran Systematic Interval Observer Design for Linear Systems Submitted to a journal 2024

[35] G. Q. B. Tran and T. P. Pham and O. Sename Fault Estimation Observers for the Vehicle Suspension with a Varying Chassis Mass IFAC-PapersOnLine 2024 58 4 138-143 12th IFAC Symposium on Fault Detection, Supervision and Safety for Technical Processes SAFEPROCESS 2024 https://doi.org/10.1016/j.ifacol.2024.07.207

[36] G. Q. B. Tran and T. P. Pham and O. Sename Reduced-Order Observer for a Class of Uncertain Descriptor NLPV Systems: Application to the Vehicle Suspension Submitted to a journal 2025

[37] R. Penrose A Generalized Inverse for Matrices Mathematical Proceedings of the Cambridge Philosophical Society 1955 51 3 406–413 10.1017/S0305004100030401

[38] H. K. Khalil Nonlinear Systems, Prentice-Hall Upper Saddle River, NJ 1996 3

[39] P. Bernard Observer Design for Nonlinear Systems Springer 2019 4

[40] E. J. McShane Extension of Range of Functions Bulletin of the American Mathematical Society 1934 40 12 837 – 842 bams/1183497871

[41] M. Maghenem and R. G. Sanfelice Lipschitzness of Minimal-Time Functions in Constrained Continuous-Time Systems with Applications to Reachability Analysis American Control Conference (ACC) 2020 937-942 10.23919/ACC45564.2020.9147496

[42] P. Bernard and V. Andrieu Luenberger Observers for Nonautonomous Nonlinear Systems IEEE Transactions on Automatic Control 2019 64 1 270-281 10.1109/TAC.2018.2872202

[43] A. Ţiclea and G. Besançon Exponential Forgetting Factor Observer in Discrete Time Systems & Control Letters 2013 62 9 756-763 https://doi.org/10.1016/j.sysconle.2013.05.008

[44] F. Wu Control of Linear Parameter Varying Systems University of California at Berkeley 1995

[45] E. D. Sontag and Y. Wang On Characterizations of the Input-to-State Stability Property Systems & Control Letters 1995 24 5 351-359 https://doi.org/10.1016/0167-6911(94)00050-6

[46] D. A. Allan and J. Rawlings and A. R. Teel Nonlinear Detectability and Incremental Input/Output-to-State Stability SIAM Journal on Control and Optimization 2021 59 4 3017-3039 10.1137/20M135039X

[47] A. Levant Higher Order Sliding Modes and Arbitrary-Order Exact Robust Differentiation 2001 European Control Conference (ECC) 2001 996-1001 10.23919/ECC.2001.7076043

[48] F. Mazenc and O. Bernard Interval Observers for Linear Time-Invariant Systems with Disturbances Automatica 2011 47 1 140-147 Apr

[49] F. Mazenc and T. N. Dinh and S.-I. Niculescu Interval Observers for Discrete-Time Systems International Journal of Robust and Nonlinear Control 2014 24 17 2867-2890

[50] J. D. Schiller and S. Muntwiler and J. Köhler and M. N. Zeilinger and M. A. Müller A Lyapunov Function for Robust Stability of Moving Horizon Estimation IEEE Transactions on Automatic Control 2023 68 12 7466–7481

[51] J. P. Gauthier and I. Kupka Deterministic Observation Theory and Applications Cambridge University Press 2001

[52] D. Astolfi and L. Marconi and L. Praly and A. R. Teel Low-Power Peaking-Free High-Gain Observers Automatica 2018 98 169-179

[53] H. K. Khalil and L. Praly High-Gain Observers in Nonlinear Feedback Control International Journal of Robust and Nonlinear Control 2013 24 6 991-992 apr

[54] P. Bernard and L. Praly and V. Andrieu Observers for a Non-Lipschitz Triangular Form Automatica 2017 82 301-313 https://doi.org/10.1016/j.automatica.2017.04.054

[55] R. Beerens and A. Bisoffi and L. Zaccarian and H. Nijmeijer and M. Heemels and N. van de Wouw Reset PID Design for Motion Systems With Stribeck Friction IEEE Transactions on Control Systems Technology 2022 30 1 294-310 10.1109/TCST.2021.3063420

[56] X. Li and M. Bohner and C.-K. Wang Impulsive Differential Equations: Periodic Solutions and Applications Automatica 2015 52 173-178 https://doi.org/10.1016/j.automatica.2014.11.009

[57] D. Liberzon Switching in Systems and Control Springer 2003 190

[58] V. Sinyakov and A. Girard Formal Controller Synthesis from Specifications Given by Discrete-Time Hybrid Automata Automatica 2021 131 109768 https://doi.org/10.1016/j.automatica.2021.109768

[59] H. Shim and A. Tanwani Hybrid-Type Observer Design based on a Sufficient Condition for Observability in Switched Nonlinear Systems International Journal of Robust and Nonlinear Control 2014 24 6 1064–1089

[60] A. Tanwani and H. Shim and D. Liberzon Observer Design for Switched Linear Systems with State Jumps Hybrid Dynamical Systems: Observation and Control 2015 179–203

[61] S. A. Burden and S. D. Coogan On Infinitesimal Contraction Analysis for Hybrid Systems 61st IEEE Conference on Decision and Control 2022 1980-1985 10.1109/CDC51059.2022.9992825

[62] M. Tucker and N. Csomay-Shanklin and A. D. Ames Robust Bipedal Locomotion: Leveraging Saltation Matrices for Gait Optimization 2023 IEEE International Conference on Robotics and Automation (ICRA) 2023 12218-12225 10.1109/ICRA48891.2023.10161309

[63] N. J. Kong and J. Joe Payne and J. Zhu and A. M. Johnson Saltation Matrices: The Essential Tool for Linearizing Hybrid Dynamical Systems Proceedings of the IEEE 2024 112 6 585-608 10.1109/JPROC.2024.3440211

[64] P. Bernard and R. G. Sanfelice Hybrid Dynamical Systems with Hybrid Inputs: Definition of Solutions and Applications to Interconnections International Journal of Robust and Nonlinear Control 2020 30 15 5892–5916

[65] R. G. Sanfelice Hybrid Feedback Control Princeton University Press, Princeton 2021 New Jersey

[66] W. P. M. H. Heemels and P. Bernard and K. J. A. Scheres and R. Postoyan and R. G. Sanfelice Hybrid Systems with Continuous-Time Inputs: Subtleties in Solution Concepts and Existence Results 60th IEEE Conference on Decision and Control (CDC) 2021 5368-5373 10.1109/CDC45484.2021.9683389

[67] P. Bernard and R. G. Sanfelice On Notions of Detectability and Observers for Hybrid Systems 59th IEEE Conference on Decision and Control 2020 5767-5772

[68] S. Hanba On the “Uniform” Observability of Discrete-Time Nonlinear Systems IEEE Transactions on Automatic Control 2009 54 8 1925-1928 10.1109/TAC.2009.2023775

[69] S. Hanba Further Results on the Uniform Observability of Discrete-Time Nonlinear Systems IEEE Transactions on Automatic Control 2010 55 4 1034-1038 10.1109/TAC.2010.2041983

[70] H. Sira-Ramírez and P. Rouchon Exact Delayed Reconstructors in Nonlinear Discrete-Time Systems Control Nonlinear and Adaptive Control: NCN4 2001 2003 351–360 Springer

[71] A. Kaldmae and U. Kotta A Note on Observability of Nonlinear Discrete-Time Systems 62nd IEEE Conference on Decision and Control 2023 7483-7488 10.1109/CDC49753.2023.10383709

[72] P. Dorato and A. Levis Optimal Linear Regulators: The Discrete-Time Case IEEE Transactions on Automatic Control 1971 16 6 613–620

[73] T. Kailath Linear Systems Prentice-Hall Englewood Cliffs, NJ 1980 156

[74] D. G. Luenberger Observing the State of a Linear System IEEE Transactions on Military Electronics 1964 8 2 74-80 10.1109/TME.1964.4323124

[75] A. Zemouche and M. Boutayeb Observer Design for Lipschitz Nonlinear Systems: The Discrete-Time Case IEEE Transactions on Circuits and Systems II: Express Briefs 2006 53 8 777–781

[76] N. Gao and M. Darouach and H. Voos and M. Alma New Unified \( H_{\infty}\) Dynamic Observer Design for Linear Systems with Unknown Inputs Automatica 2016 65 43–52

[77] S. Ibrir Circle-Criterion Approach to Discrete-Time Nonlinear Observer Design Automatica 2007 43 8 1432-1441

[78] R. E. Kalman and R. S. Bucy New Results in Linear Filtering and Prediction Theory Journal of Basic Engineering 1961 108 83-95

[79] Jr. J. Deyst and C. Price Conditions for Asymptotic Stability of the Discrete Minimum-Variance Linear Estimator IEEE Transactions on Automatic Control 1968 13 702-705

[80] A. H. Jazwinski Stochastic Processes and Filtering Theory Academic Press 1970

[81] Q. Zhang On Stability of the Kalman Filter for Discrete Time Output Error Systems Systems & Control Letters 2017 107 84-91 Jul 10.1016/j.sysconle.2017.07.011

[82] Bornard, G. and Couenne, N. and Celle, F. Regularly Persistent Observers for Bilinear Systems New Trends in Nonlinear Control Theory 1989 Descusse, J. and Fliess, Michel and Isidori, A. and Leborgne, D. 130–140 Springer Berlin Heidelberg

[83] H. Hammouri and J. de Leon Morales Observer Synthesis for State-Affine Systems 29th IEEE Conference on Decision and Control 1990 784-785 vol.2 10.1109/CDC.1990.203695

[84] M. Green and J. B. Moore Persistence of Excitation in Linear Systems Systems & Control Letters 1986 7 5 351-360 https://doi.org/10.1016/0167-6911(86)90052-6

[85] G. Ciccarella and M. Dalla Mora and A. Germani Observers for Discrete-Time Nonlinear Systems Systems & Control Letters 1993 20 5 373-382 https://doi.org/10.1016/0167-6911(93)90016-Y

[86] S. Monaco and D. Normand-Cyrot The Immersion under Feedback of a Multidimensional Discrete-Time Non-Linear System into a Linear System International Journal of Control 1983 38 245-261 07 10.1080/00207178308933073

[87] W. Lin and C. I. Byrnes Remarks on Linearization of Discrete-Time Autonomous Systems and Nonlinear Observer Design Systems & Control Letters 1995 25 1 31-40 https://doi.org/10.1016/0167-6911(94)00054-Y

[88] H. Huijberts On Existence of Extended Observers for Nonlinear Discrete-Time Systems Lecture Notes in Control and Information Sciences 1999 244 03 10.1007/BFb0109922

[89] C. Califano and S. Monaco and D. Normand-Cyrot On the Observer Design in Discrete-Time Systems & Control Letters 2003 49 4 255-265 https://doi.org/10.1016/S0167-6911(02)00344-4

[90] Y. Song and J. W. Grizzle The Extended Kalman Filter as a Local Asymptotic Observer for Nonlinear Discrete-Time Systems 1992 American Control Conference 1992 3365-3369 10.23919/ACC.1992.4792775

[91] M. Boutayeb and H. Rafaralahy and M. Darouach Convergence Analysis of the Extended Kalman Filter Used as an Observer for Nonlinear Deterministic Discrete-Time Systems IEEE Transactions on Automatic Control 1997 42 4 581-586 10.1109/9.566674

[92] M. Boutayeb and D. Aubry A Strong Tracking Extended Kalman Observer for Nonlinear Discrete-Time Systems IEEE Transactions on Automatic Control 1999 44 1550 - 1556 09 10.1109/9.780419

[93] A. Barrau and S. Bonnabel The Invariant Extended Kalman Filter as a Stable Observer IEEE Transactions on Automatic Control 2017 62 4 1797-1812 10.1109/TAC.2016.2594085

[94] P. E. Moraal and J. W. Grizzle Observer Design for Nonlinear Systems with Discrete-Time Measurements IEEE Transactions on Automatic Control 1995 40 3 395-404 10.1109/9.376051

[95] E. Bıyık and M. Arcak A Hybrid Redesign of Newton Observers in the Absence of an Exact Discrete-Time Model Systems & Control Letters 2006 55 6 429-436 https://doi.org/10.1016/j.sysconle.2005.09.005

[96] H. Arezki and A. Alessandri and A. Zemouche Robust Moving-Horizon Estimation for quasi-LPV Discrete-Time Systems IFAC-PapersOnLine 2023 56 2 6771–6776

[97] C. V. Rao and J. B. Rawlings and D. Q. Mayne Constrained State Estimation for Nonlinear Discrete-Time Systems: Stability and Moving Horizon Approximations IEEE Transactions on Automatic Control 2003 48 2 246-258 10.1109/TAC.2002.808470

[98] A. Alessandri and M. Baglietto and G. Battistelli Moving-Horizon State Estimation for Nonlinear Discrete-Time Systems: New Stability Results and Approximation Schemes Automatica 2008 44 7 1753-1765 https://doi.org/10.1016/j.automatica.2007.11.020

[99] A. Alessandri and M. Gaggero Fast Moving Horizon State Estimation for Discrete-Time Systems Using Single and Multi Iteration Descent Methods IEEE Transactions on Automatic Control 2017 62 9 4499-4511 10.1109/TAC.2017.2660438

[100] A. Jazwinski Limited Memory Optimal Filtering IEEE Transactions on Automatic Control 1968 13 5 558-563 10.1109/TAC.1968.1098981

[101] G. Q. B. Tran and P. Bernard Arbitrarily Fast Robust KKL Observer for Nonlinear Time-Varying Discrete Systems IEEE Transactions on Automatic Control 2024 69 3 1520-1535 10.1109/TAC.2023.3328833

[102] L. Brivadis and V. Andrieu and U. Serres Luenberger Observers for Discrete-Time Nonlinear Systems 58th IEEE Conference on Decision and Control 2019 3435-3440 10.1109/CDC40024.2019.9029220

[103] L. da C. Ramos and F. Di Meglio and V. Morgenthaler and L. F. F. da Silva and P. Bernard Numerical Design of Luenberger Observers for Nonlinear Systems 59th IEEE Conference on Decision and Control 2020 5435-5442 10.1109/CDC42340.2020.9304163

[104] G. Q. B. Tran and P. Bernard and V. Andrieu and D. Astolfi Constructible Canonical Form and High-Gain Observer in Discrete Time 63rd IEEE Conference on Decision and Control 2024

[105] V. Andrieu and L. Praly On the Existence of a Kazantzis–Kravaris/Luenberger Observer SIAM Journal on Control and Optimization 2006 45 2 432-456 10.1137/040617066

[106] V. Andrieu Convergence Speed of Nonlinear Luenberger Observers SIAM Journal on Control and Optimization 2014 52 5 2831-2856 10.1137/110853091

[107] J.-M. Coron On the Stabilization of Controllable and Observable Systems by an Output Feedback Law Math. Control Signals Systems 1994 7 3 187–216

[108] A. N. Shoshitaishvili Singularities for Projections of Integral Manifolds with Applications to Control and Observation Problems Theory of Singularities and its Applications 1990 1 295

[109] N. Kazantzis and C. Kravaris Nonlinear Observer Design Using Lyapunov’s Auxiliary Theorem Systems & Control Letters 1998 34 5 241-247 https://doi.org/10.1016/S0167-6911(98)00017-6

[110] A. J. Krener and M. Q. Xiao Nonlinear Observer Design in the Siegel Domain through Coordinate Changes IFAC Proceedings Volumes 2001 34 6 519–524

[111] G. Kreisselmeier and R. Engel Nonlinear Observers for Autonomous Lipschitz Continuous Systems IEEE Transactions on Automatic Control 2003 48 451-464

[112] L. Brivadis and V. Andrieu and P. Bernard and U. Serres Further Remarks on KKL Observers Systems & Control Letters 2023 172 105429 https://doi.org/10.1016/j.sysconle.2022.105429

[113] C. Afri and V. Andrieu and L. Bako and P. Dufour State and Parameter Estimation: A Nonlinear Luenberger Observer Approach IEEE Transactions on Automatic Control 2017 62 2 973-980 10.1109/TAC.2016.2566804

[114] N. Henwood and J. Malaizé and L. Praly A Robust Nonlinear Luenberger Observer for the Sensorless Control of SM-PMSM: Rotor Position and Magnets Flux Estimation IECON Conference on IEEE Industrial Electronics Society 2012

[115] P. Bernard and L. Praly Estimation of Position and Resistance of a Sensorless PMSM: A Nonlinear Luenberger Approach for a Non-Observable System IEEE Transactions on Automatic Control 2021 66 481-496

[116] J. Peralez and M. Nadri Deep Learning-Based Luenberger Observer Design for Discrete-Time Nonlinear Systems 60th IEEE Conference on Decision and Control 2021 4370–4375 IEEE

[117] M. Buisson-Fenet and L. Bahr and F. Di Meglio Learning to Observe: Neural Network-Based KKL Observers Python toolbox available at

[118] M. Buisson-Fenet and L. Bahr and F. Di Meglio Towards Gain Tuning for Numerical KKL Observers 22nd IFAC World Congress

[119] M. U. B. Niazi and J. Cao and X. Sun and A. Das and K. H. Johansson Learning-based Design of Luenberger Observers for Autonomous Nonlinear Systems 2023 American Control Conference (ACC) 2023 3048-3055 10.23919/ACC55779.2023.10156294

[120] I. Mezić Spectral Properties of Dynamical Systems, Model Reduction and Decompositions Nonlinear Dynamics 2005 41 309–325

[121] H. Harry Asada and J. A. Solano-Castellanos Control-Coherent Koopman Modeling: A Physical Modeling Approach 63rd IEEE Conference on Decision and Control 2024

[122] Peter J Schmid Dynamic Mode Decomposition of Numerical and Experimental Data Journal of Fluid Mechanics 2010 656 5–28

[123] Joshua L. Proctor and Steven L. Brunton and J. Nathan Kutz Dynamic Mode Decomposition with Control SIAM Journal on Applied Dynamical Systems 2016 15 1 142-161 10.1137/15M1013857

[124] Lorenzo Marconi and Laurent Praly and Alberto Isidori Output Stabilization via Nonlinear Luenberger Observers SIAM Journal on Control and Optimization 2007 45 6 2277–2298

[125] A. M. Dabroom and H. K. Khalil Discrete-Time Implementation of High-Gain Observers for Numerical Differentiation International Journal of Robust and Nonlinear Control 1999 72 17 1523-1537

[126] Z. Wang and T. N. Dinh and Q. Zhang and T. Raïssi and Y. Shen Fast Interval Estimation for Discrete-Time Linear Systems: An \( L_1\) Optimization Method Automatica 2022 137 110029 https://doi.org/10.1016/j.automatica.2021.110029

[127] J. B. Moore and B. D. O. Anderson Coping with Singular Transition Matrices in Estimation and Control Stability Theory International Journal of Control 1980 31 571-586

[128] P. Bernard and T. Devos and A. K. Jebai and P. Martin and L. Praly KKL Observer Design for Sensorless Induction Motors 61st IEEE Conference on Decision and Control 2022 1235-1242 10.1109/CDC51059.2022.9992800

[129] S. Ammar and M. Mabrouk and J.-C. Vivalda On the Genericity of the Differential Observability of Controlled Discrete-Time Systems SIAM Journal on Control and Optimization 2007 46 6 2182-2198 10.1137/060677938

[130] M. Arcak and D. Nešić A Framework for Nonlinear Sampled-Data Observer Design via Approximate Discrete-Time Models and Emulation Automatica 2004 40 11 1931-1938 https://doi.org/10.1016/j.automatica.2004.06.004

[131] P. Bernard and R. G. Sanfelice Semiglobal High-Gain Hybrid Observer for a Class of Hybrid Dynamical Systems with Unknown Jump Times IEEE Transactions on Automatic Control 2024 1-16 10.1109/TAC.2024.3355324

[132] R. S. Johnson Parameter Estimation for Hybrid Dynamical Systems University of California, Santa Cruz 2023

[133] A. Tanwani and S. Trenn Detectability and Observer Design for Switched Differential–algebraic Equations Automatica 2019 99 289-300 https://doi.org/10.1016/j.automatica.2018.10.043

[134] A. Tanwani and H. Shim and D. Liberzon Observability for Switched Linear Systems: Characterization and Observer Design IEEE Transactions on Automatic Control 2013 58 4 891-904

[135] H. Ríos and J. Davila and A. R. Teel Linear Hybrid Systems with Periodic Jumps: A Notion of Strong Observability and Strong Detectability IEEE Transactions on Automatic Control 2020 65 6 2640-2646 10.1109/TAC.2019.2940230

[136] G. Xie and L. Wang Controllability and Observability of a Class of Linear Impulsive Systems Journal of Mathematical Analysis and Applications 2005 304 336-355

[137] E. A. Medina and D. A. Lawrence Reachability and Observability of Linear Impulsive Systems Automatica 2008 44 5 1304-1309 https://doi.org/10.1016/j.automatica.2007.09.017

[138] S. Zhao and J. Sun Controllability and Observability for a Class of Time-Varying Impulsive Systems Nonlinear Analysis : Real World Applications 2009 10 1370-1380

[139] A. Balluchi and L. Benvenutia and M. D. Di Benedetto and A. Sangiovanni-Vincentelli The Design of Dynamical Observers for Hybrid Systems: Theory and Application to an Automotive Control Problem Automatica 2013 49 4 915-925

[140] F. Lin and L. Y. Wang and W. Chen and M. P. Polis On Observability of Hybrid Systems IEEE Transactions on Automatic Control 2022 67 11 6074-6081 10.1109/TAC.2021.3126610

[141] F. Lin and L. Y. Wang and G. Yin and M. Polis and W. Chen On Detectability of a Class of Hybrid Systems IEEE Transactions on Automatic Control 2022 10.1109/TAC.2022.3225514

[142] P. D’Alessandro and A. Isidori and A. Ruberti A New Approach to the Theory of Canonical Decomposition of Linear Dynamical Systems SIAM Journal on Control 1973 11 1 148-158 10.1137/0311012

[143] N. Cox and L. Marconi and A. R. Teel Isolating Invisible Dynamics in the Design of Robust Hybrid Internal Models Automatica 2016 68 56-68

[144] E. Mostacciuolo and S. Trenn and F. Vasca Averaging for Switched Impulsive Systems with Pulse Width Modulation Automatica 2024 160 111447 https://doi.org/10.1016/j.automatica.2023.111447

[145] A. Alessandri and P. Coletta Switching Observers for Continuous-Time and Discrete-Time Linear Systems 2001 American Control Conference 2001 3 2516-2521

[146] F. Deza and E. Busvelle and J. P. Gauthier and D. Rakotopara High Gain Estimation for Nonlinear Systems Systems & Control Letters 1992 18 4 295 - 299 https://doi.org/10.1016/0167-6911(92)90059-2

[147] J. Sur and B. Paden Observers for Linear Systems with Quantized Output 1997 American Control Conference 1997 3012-3016

[148] T. Raff and F. Allg\"ower Observers with Impulsive Dynamical Behavior for Linear and Nonlinear Continuous-Time Systems 46th IEEE Conference on Decision and Control 2007 4287-4292

[149] T. Raff and M. Kogel and F. Allgower Observer with Sample-and-Hold Updating for Lipschitz Nonlinear Systems with Nonuniformly Sampled Measurements 2008 American Control Conference 2008 5254-5257

[150] T. Ahmed-Ali and L. Burlion and F. Lamnabhi-Lagarrigue and C. Hann A Sampled-Data Observer with Time-Varying Gain for a Class of Nonlinear Systems with Sampled-Measurements 53rd IEEE Conference on Decision and Control 2014 316-321

[151] F. Ferrante and F. Gouaisbaut and R. G. Sanfelice and S. Tarbouriech State Estimation of Linear Systems in the Presence of Sporadic Measurements Automatica 2016 73 101-109 Nov 10.1016/j.automatica.2016.05.032

[152] A. Sferlazza and S. Tarbouriech and L. Zaccarian Time-Varying Sampled-Data Observer With Asynchronous Measurements IEEE Transactions on Automatic Control 2019 64 2 869-876 10.1109/TAC.2018.2839974

[153] I. Karafyllis and C. Kravaris From Continuous-Time Design to Sampled-Data Design of Observers IEEE Transactions on Automatic Control 2009 54 9 2169-2174 10.1109/TAC.2009.2024390

[154] T. N. Dinh and V. Andrieu and M. Nadri and U. Serres Continuous-Discrete Time Observer Design for Lipschitz Systems With Sampled Measurements IEEE Transactions on Automatic Control 2015 60 3 787-792 10.1109/TAC.2014.2329211

[155] H. R\'ios and J. Dávila and A. R. Teel State Estimation for Linear Hybrid Systems with Periodic Jumps and Unknown Inputs International Journal of Robust and Nonlinear Control 2020 30 15 5966-5988

[156] E. A. Medina and D. A. Lawrence State Estimation for Linear Impulsive Systems 2009 American Control Conference 2009 1183-1188

[157] G. Q. B. Tran and P. Bernard Kalman-Like Observer for Hybrid Systems with Linear Maps and Known Jump Times 62nd IEEE Conference on Decision and Control 2023

[158] G. Q. B. Tran and P. Bernard and L. Marconi Observer Design for Hybrid Systems with Linear Maps and Known Jump Times Hybrid and Networked Dynamical Systems Springer 2024 Romain Postoyan and Paolo Frasca and Elena Panteley and Luca Zaccarian

[159] M. Moarref and L. Rodrigues Observer Design for Linear Multi-Rate Sampled-Data Systems 2014 American Control Conference 2014 5319-5324 10.1109/ACC.2014.6858838

[160] A. Tanwani and O. Yufereva Error Covariance Bounds for Suboptimal Filters with Lipschitzian Drift and Poisson-Sampled Measurements Automatica 2020 122 109280 https://doi.org/10.1016/j.automatica.2020.109280

[161] E. M. Izhikevich Simple Model of Spiking Neurons IEEE Transactions on Neural Networks 2003 14 6 1569-1572 10.1109/TNN.2003.820440

[162] C. Cai and A. R. Teel Characterizations of Input-to-State Stability for Hybrid Systems Systems & Control Letters 2009 58 1 47-53 https://doi.org/10.1016/j.sysconle.2008.07.009

[163] G. Q. B. Tran and P. Bernard and F. Di Meglio and L. Marconi Observer Design based on Observability Decomposition for Hybrid Systems with Linear Maps and Known Jump Times 61st IEEE Conference on Decision and Control 2022 1974-1979 10.1109/CDC51059.2022.9993225

[164] C.-T. Chen Linear System Theory and Design CBS College Publishing 1984

[165] S. Boyd and L. El Ghaoui and E. Feron and V. Balakrishnan Linear Matrix Inequalities in System and Control Theory SIAM 1994

[166] G. Q. B. Tran and P. Bernard and R. G. Sanfelice Coupling Flow and Jump Observers for Hybrid Systems with Known Jump Times IFAC-PapersOnLine 2023 56 2 6958-6963 22nd IFAC World Congress https://doi.org/10.1016/j.ifacol.2023.10.522

[167] R. G. Sanfelice and A. R. Teel Asymptotic Stability in Hybrid Systems via Nested Matrosov Functions IEEE Transactions on Automatic Control 2009 54 7 1569-1574 10.1109/TAC.2009.2015524

[168] R. G. Sanfelice and J. J. B. Biemond and N. van de Wouw and WP Maurice H Heemels An Embedding Approach for the Design of State-Feedback Tracking Controllers for References with Jumps International Journal of Robust and Nonlinear Control 2014 24 11 1585–1608

[169] D. Liberzon and D. Nešić and A. R. Teel Lyapunov-Based Small-Gain Theorems for Hybrid Systems IEEE Transactions on Automatic Control 2014 59 6 1395-1410

[171] A. Goswami and B. Espiau and A. Keramane Limit Cycles and their Stability in a Passive Bipedal Gait Proceedings of IEEE International Conference on Robotics and Automation 1996 1 246-251 vol.1 10.1109/ROBOT.1996.503785

[172] M. Vigne and A. E. Khoury and M. Masselin and F. Di Meglio and N. Petit Estimation of Multiple Flexibilities of an Articulated System Using Inertial Measurements 2018 IEEE Conference on Decision and Control (CDC) 2018 6779-6785 10.1109/CDC.2018.8619734

[173] T. Chen and B. Goodwine Robust Gait Design for a Compass Gait Biped on Slippery Surfaces Robotics and Autonomous Systems 2021 140 103762 https://doi.org/10.1016/j.robot.2021.103762

[174] M. Posa and M. Tobenkin and R. Tedrake Stability Analysis and Control of Rigid-Body Systems With Impacts and Friction IEEE Transactions on Automatic Control 2016 61 6 1423-1437 10.1109/TAC.2015.2459151

[175] F. Jenelten and J. Hwangbo and F. Tresoldi and C. D. Bellicoso and M. Hutter Dynamic Locomotion on Slippery Ground IEEE Robotics and Automation Letters 2019 4 4 4170-4176 10.1109/LRA.2019.2931284

[176] Y. Nisticò and S. Fahmi and L. Pallottino and C. Semini and G. Fink On Slip Detection for Quadruped Robots Sensors 2022 22 8 10.3390/s22082967

[177] A. N. Atassi and H. K. Khalil A Separation Principle for the Stabilization of a Class of Nonlinear Systems IEEE Transactions on Automatic Control 1999 44 9 1672-1687 10.1109/9.788534

[178] F. Forni and A. R. Teel and L. Zaccarian Follow the Bouncing Ball: Global Results on Tracking and State Estimation With Impacts IEEE Transactions on Automatic Control 2013 58 6 1470-1485 10.1109/TAC.2013.2237952

[179] J. J. B. Biemond and M. Heemels and R. G. Sanfelice and N. van de Wouw Distance Function Design and Lyapunov Techniques for the Stability of Hybrid Trajectories Automatica 2016 73 38–46

[180] J. J. B. Biemond and R. Postoyan and W. P. M. H. Heemels and N. Van de Wouw Incremental Stability of Hybrid Dynamical Systems IEEE Transactions on Automatic Control 2018 63 12 4094–4109

[181] M. Rijnen and J. J. B. Biemond and N. van de Wouw and A. Saccon and H. Nijmeijer Hybrid Systems With State-Triggered Jumps: Sensitivity-Based Stability Analysis With Application to Trajectory Tracking IEEE Trans. on Automatic Control 2020 65 11 4568-4583 10.1109/TAC.2019.2961996

[182] P. Bernard and R. G. Sanfelice An Algorithm to Generate Solutions to Hybrid Dynamical Systems with Inputs 2019 American Control Conference (ACC) 2019 2996-3001 10.23919/ACC.2019.8815141

[183] W. Xiang and J. Xiao and M. N. Iqbal Robust Observer Design for Nonlinear Uncertain Switched Systems under Asynchronous Switching Nonlinear Analysis: Hybrid Systems 2012 6 1 754-773

[184] Z. Ping and C. Lee and H. Shim Robust Estimation Algorithm for both Switching Signal and State of Switched Linear Systems Int. Journal of Control, Automation and Systems 2017 15 1 95-103

[185] C. Zammali and J. V. Gorp and X. Ping and T. Raïssi Switching Signal Estimation based on Interval Observer for a Class of Switched Linear Systems IEEE 58th Conference on Decision and Control (CDC) 2019 2497-2502 10.1109/CDC40024.2019.9029945

[186] S. C. Shankar and G. Yang and J. P. Hespanha State Estimation for Asynchronously Switched Sampled-Data Systems IEEE Conference on Decision and Control 2022 1-7 10.1109/CDC51059.2022.9993230

[187] N. J. Kong and J. J. Payne and G. Council and A. M. Johnson The Salted Kalman Filter: Kalman Filtering on Hybrid Dynamical Systems Automatica 2021 131 109752 https://doi.org/10.1016/j.automatica.2021.109752

[188] J. J. Payne and N. J. Kong and A. M. Johnson The Uncertainty Aware Salted Kalman Filter: State Estimation for Hybrid Systems with Uncertain Guards 2022 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS) 2022 8821-8828 10.1109/IROS47612.2022.9981218

[189] P. Bernard and R. G. Sanfelice A Local Hybrid Observer for a Class of Hybrid Dynamical Systems with Linear Maps and Unknown Jump Times 60th IEEE Conference on Decision and Control 2021 5362-5367 10.1109/CDC45484.2021.9682924

[190] J. Kim and H. Cho and A. Shamsuarov and H. Shim and J. H. Seo State Estimation Strategy without Jump Detection for Hybrid Systems using Gluing Function 53rd IEEE Conference on Decision and Control 2014 139–144

[191] L. Menini and A. Tornambè State Immersion Observers for Mechanical Systems with Impacts IEEE Conference on Decision and Control 2016 7117-7122 10.1109/CDC.2016.7799366

[192] G. Q. B. Tran and S. Garcia and P. Bernard and F. Di Meglio and R. G. Sanfelice Towards Gluing KKL Observer for Hybrid Systems with Unknown Jump Times 63rd IEEE Conference on Decision and Control 2024

[193] P. V. Preiswerk and R. I. Leine State Observers for the Time Discretization of a Class of Impulsive Mechanical Systems International Journal of Robust and Nonlinear Control 2022 32 12 6667-6683 https://doi.org/10.1002/rnc.6168

[194] L. Paoli and M. Schatzman A Numerical Scheme for Impact Problems I: The One-Dimensional Case SIAM Journal on Numerical Analysis 2002 40 2 702-733 10.1137/S0036142900378728

[195] J. Liu and A. R. Teel Hybrid Systems with Memory: Existence and Well-Posedness of Generalized Solutions SIAM Journal on Control and Optimization 2018 56 2 1011-1037 10.1137/140981708

[196] N. Wang and R. G. Sanfelice HyRRT-Connect: An Efficient Bidirectional Rapidly-Exploring Random Trees Motion Planning Algorithm for Hybrid Dynamical Systems IFAC-PapersOnLine 2024 58 11 51-56 8th IFAC Conference on Analysis and Design of Hybrid Systems ADHS 2024 https://doi.org/10.1016/j.ifacol.2024.07.424

[197] E. D. Sontag Mathematical Control Theory: Deterministic Finite Dimensional Systems Springer Science & Business Media 1998 6

[198] B. L. Van de Vrande and D. H. Van Campen and A. De Kraker An Approximate Analysis of Dry-Friction-Induced Stick-Slip Vibrations by a Smoothing Procedure Nonlinear Dynamics 1999 19 159–171

[199] Leine, R. I. and Nijmeijer, H. Modelling of Dry Friction 39–46 Springer Berlin Heidelberg 2004 Berlin, Heidelberg 10.1007/978-3-540-44398-8_4

[200] U. J. F. Aarsnes and R. J. Shor Torsional Vibrations with Bit Off Bottom: Modeling, Characterization and Field Data Validation Journal of Petroleum Science and Engineering 2018 163 712-721 https://doi.org/10.1016/j.petrol.2017.11.024

[201] J. W. Grizzle and G. Abba and F. Plestan Asymptotically Stable Walking for Biped Robots: Analysis via Systems with Impulse Effects IEEE Transactions on Automatic Control 2001 46 1 51-64 10.1109/9.898695

[202] X. Lou and R. G. Sanfelice Asymptotic Stability of Limit Cycles in Hybrid Systems with Explicit Logic States 2019 American Control Conference (ACC) 2019 2340-2345 10.23919/ACC.2019.8814826

[203] J. Hwangbo and C. D. Bellicoso and P. Fankhauser and M. Hutter Probabilistic Foot Contact Estimation by Fusing Information from Dynamics and Differential/Forward Kinematics 2016 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS) 2016 3872-3878 10.1109/IROS.2016.7759570

[204] S. Piperakis and M. Maravgakis and D. Kanoulas and P. Trahanias Robust Contact State Estimation in Humanoid Walking Gaits 2022 IEEE/RSJ International Conference on Intelligent Robots and Systems (IROS) 2022 6732-6738 10.1109/IROS47612.2022.9981354

[205] E. M. Stein and R. Shakarchi Real Analysis: Measure Theory, Integration, and Hilbert Spaces Princeton Univ. Press 2005 Princeton lectures in analysis Princeton, NJ

[206] A. Saoud and M. Maghenem and A. Loría and R. G. Sanfelice Hybrid Persistency of Excitation in Adaptive Estimation for Hybrid Systems 2023 10.48550/ARXIV.2301.05741

[207] J. K. hale Ordinary Differential Equations Courier Corporation 2009

[208] R. Skelton and T. Iwasaki and K. Grigoriadis A Unified Algebraic Approach to Linear Control Design 2017 01 10.1201/9781315136523

[209] S. L. Campbell and C. D. Meyer Generalized Inverses of Linear Transformations SIAM 2009

[210] L. C. Evans Partial Differential Equations American Mathematical Society 2022 19