EXPLOITING IMPULSIVE INPUTS FOR STABILIZATION OF UNDERACTUATED ROBOTIC SYSTEMS: THEORY AND EXPERIMENTS By Nilay Kant A DISSERTATION Michigan State University in partial fulfillment of the requirements Submitted to for the degree of Mechanical Engineering – Doctor of Philosophy 2020 ABSTRACT EXPLOITING IMPULSIVE INPUTS FOR STABILIZATION OF UNDERACTUATED ROBOTIC SYSTEMS: THEORY AND EXPERIMENTS By Nilay Kant Robots have become increasingly popular due to their ability to perform complex tasks and operate in unknown and hazardous environments. Many robotic systems are underactuated i.e., they have fewer control inputs than their degrees-of-freedom (DOF). Common examples of underactuated robotic systems are legged robots such as bipeds, flying robots such as quadrotors, and swimming robots. Due to limited control authority, underactuated systems are prone to instability. This work includes impulsive inputs in the set of admissible controls to address several challenging control problems. It has already been shown that continuous-time approximation of impulsive inputs can be realized in physical hardware using high-gain feedback. Stabilization of an equilibrium point is an important control problem for underactuated systems. The ability of the system to remain stable in the presence of disturbances depends on the size of the region of attraction of the stabilized equilibrium. The sum of squares and trajectory reversing methods are combined to generate a large estimate of the region of attraction. This estimate is then effectively enlarged by applying the impulse manifold method to stabilize equilibria from points lying outside the estimated region of attraction. Simulation results are provided for a three-DOF tiptoebot and experimental validation is carried out on a two-DOF pendubot. Impulsive inputs are also utilized to control the underactuated inertia-wheel pendulum (IWP). When subjected to impulsive inputs, the dynamics of the IWP can be described by algebraic equations. Optimal sequences of inputs are designed to achieve rest-to-rest maneuvers and the results are applied to the swing-up control problem. The novel problem of juggling a devil-stick using impulsive inputs is also investigated. Impulsive forces are applied to the stick intermittently and the impulse of the force and its point of application are modeled as inputs to the system. A dead-beat design for one of the inputs simplifies the control problem and results in a discrete linear time invariant system. To achieve symmetric juggling, linear quadratic regulator (LQR) and model predictive control (MPC) based designs are proposed and validated through simulations. A repetitive motion is described by closed orbits and therefore, stabilization of closed orbits is important for many applications such as bipedal walking and steady swimming. We first investigate the problem of energy-based orbital stabilization using continuous inputs and intermittent impulsive braking. The orbit is a manifold where the active generalized coordinates are fixed and the total mechanical energy of the system is equal to some desired value. Simulation and experimental results are provided for the tiptoebot and the rotary pendulum, respectively. The problem of orbital stabilization using virtual holonomic constraints (VHC) is also investigated. VHCs are enforced using a continuous controller which guarantees existence of closed orbits. A Poincaré section is constructed on the desired orbit and the orbit is stabilized using impulsive inputs which are applied intermittently when the system trajectory crosses the Poincaré section. This approach to orbital stabilization is general, and has lower complexity and computational cost than control designs proposed earlier. This dissertation is lovingly dedicated to parents; my father Shri. Rajni Kant, and to my mother Smt. Kumari Jyotsna Kant as without their sacrifices, I would not have been able to accomplish this. I would also like to dedicate this work to Swami Vivekananda, an Indian yogi whose life, works and teachings have greatly transformed my life. iv ACKNOWLEDGEMENTS I would like to express my utmost gratitude to my PhD advisor, Dr. Ranjan Mukherjee, as his influence is as profound on this dissertation as it is in my life. In the ancient Indian civilization, education was not very formal as it is today and a student spent several years at a teacher’s house to attain knowledge. The close association with the teacher resulted in a strong human relationship and the student mastered the most subtle aspect of a subject along with the desired human values. I am extremely lucky to have experienced such a process in the last five years, and this has been possible due to my association with Dr. Mukherjee. His office was always open during the day and I could simply walk inside anytime if I was stuck in a proof. There were countless occasions where both of us brainstormed together on a topic and hours just passed by. I vividly remember that he once showed me how to derive the Lagrange’s equation on a notebook with minute details such as defining suitable coordinate system and drawing clear diagrams using a pencil. These instances may appear insignificant, but for me, they were life changing. It was only due to such involved and patient training, I became capable enough to accomplish this work. Despite over twenty five years of being a professor, Dr. Mukherjee unfailingly comes to office every weekend. Such hard work coupled with a genius mind inspired me throughout to push my boundaries and would forever continue to do so. I have experienced a transformation in the last five years - admirable traits of mine, if any, are due to Dr. Mukherjee, and the rest are my fault. Additionally, I am sincerely thankful to my PhD committee members Dr. Hassan Khalil, Dr. Vaibhav Srivastava and Dr. Gouming Zhu for their continued guidance, motivation and support. I am also very grateful to my lab mates; Sheryl, Mahmoud, Sanders, Connor, Krissy and Amer as their friendship did not let me miss my family back in India. I would also like to thank Mr. Roy Bailiff of the mechanical engineering department whose help in the machine shop ensured me to build my experimental setup. I would like to express my gratitude to my parents are they have made several sacrifices to ensure that I get the best education possible. If it was not their conditional love and support, I v would not have come this far. I am also very thankful to my brother Dr. Kislay Kant as he has always set a high standard due to his achievements and conduct which has greatly inspired me to dream big and sincerely work towards it. I would also like to pay my salutations to my late grandparents Shri. Ramaeshward Prasad, Shri Hari Narayan Prasad, Smt. Rajkishori Devi and Smt. Gyneshwari Devi as their morals and life struggles were the source of inspiration in difficult times. Lastly, I would like to express my deepest gratitude to my wife Priyanka, as she was my strongest support throughout, and her love made me finish this journey with ease. vi TABLE OF CONTENTS . . xi . . . . . . . . . . . . LIST OF TABLES . . . LIST OF FIGURES . . . LIST OF ALGORITHMS . KEY TO SYMBOLS . . . CHAPTER 1 . . . . INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.1 Motivation . . 2 1.2 Estimating and Enlarging the Region of Attraction of Equilibria . . . . . . . . . . . 1.3 Rest-to-Rest Maneuver of the Inertia Wheel Pendulum . . . . . . . . . . . . . . . . 4 5 . 1.4 Devil-Stick Juggling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 1.5 Energy-Based Orbital Stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6 Orbital Stabilization using Virtual Holonomic Constraints . . . . . . . . . . . . . . 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CHAPTER 2 ESTIMATION OF THE REGION OF ATTRACTION OF EQUILIBRIA . . . . . . . . . . . . 2.1 Introduction . 2.2 Background . AND ITS ENLARGEMENT USING IMPULSIVE INPUTS . . . . . . . . . 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.2.1 The Impulse Manifold Method . . . . . . . . . . . . . . . . . . . . . . . . 10 2.2.2 The Sum of Squares Method . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.3 Enlarging the Region of Attraction Using SOS and IMM . . . . . . . . . . . . . . 15 2.3.1 An Illustrative Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.3.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 Simulation Results 2.4 Algorithm for Computing ReA Using the Method of . . . Trajectory Reversing . . . 2.4.1 Definitions 2.4.2 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2.5 Enlarging the Region of Attraction Using SOS, CHART and IMM . . . . . . . . . 25 2.6 Experimental Verification of Enlargement of ReA using SOS, CHART and IMM . . 29 2.6.1 Hardware Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.6.2 Design of Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 2.6.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 CHAPTER 3 Introduction . IMPULSIVE DYNAMICS AND CONTROL OF THE INERTIA-WHEEL PENDULUM . . . 3.1 3.2 Mathematical Model and Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 . 37 3.2.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 3.2.2 Effect of Impulsive Input and Constants of Free Motion . . . . . . . . . . . 37 3.2.3 Problem Statement: Rest-to-Rest Maneuvers . . . . . . . . . . . . . . . . . 38 . . . . . . . . vii 3.3 Rest-to-Rest Maneuvers: Case of One and Two Impulsive Inputs . . . . . . . . . . 39 3.3.1 Case of One Impulsive Input (N “ 1) . . . . . . . . . . . . . . . . . . . . . 39 3.3.2 Case of Two Impulsive Inputs (N “ 2) . . . . . . . . . . . . . . . . . . . . 40 3.4 Rest-to-Rest Maneuvers: Generalization to N Inputs . . . . . . . . . . . . . . . . . 44 3.4.1 Revisiting the Problem Statement . . . . . . . . . . . . . . . . . . . . . . 44 3.4.2 Rest-to-Rest Maneuvers with Even Number of Impulsive Inputs . . . . . . 45 . . . . . . . . . . . . 48 3.4.3 Rest-to-Rest Maneuvers with Odd Number of Inputs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 3.5.1 Optimal Swing-Up Using Even Impulse Sequences . . . . . . . . . . . . . 50 3.5.2 Implementation Using High-Gain Feedback . . . . . . . . . . . . . . . . . 50 3.5.3 Discussion of Results in the Literature . . . . . . . . . . . . . . . . . . . . 51 . . . . . . . . . . . . . . . . . . 51 Energy Based Controller . . . . . . . . . . . . . . . . . . . . . . 55 3.5.3.1 Globally Stabilizing Controller 3.5.3.2 3.5 The Swing-Up Problem . . CHAPTER 4 . . . . . . . . Introduction . Impulsive Dynamics IMPULSIVE CONTROL OF A DEVIL-STICK: PLANAR SYMMETRIC JUGGLING . . . . 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 4.1 . 4.2 Problem Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 4.3 Dynamics of the Devil-Stick . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 4.3.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 4.3.2 Continuous-time Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.3.3 Poincaré Sections and Half-Return Maps . . . . . . . . . . . . . . . . . . . 61 4.3.4 Coordinate Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . 62 4.3.5 . . . . . . . . . . . . . . . . 64 4.4 State Feedback Control Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 Steady-State Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66 4.4.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 4.4.2 Error Dynamics . 4.4.3 . . . . . . . . . . . . . . . . . 68 4.4.4 Residual Control Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 4.5.1 . . . . . . . . . . . . . . . . . . 72 4.5.2 Results for the LQR-based Design . . . . . . . . . . . . . . . . . . . . . . 72 4.5.3 Results for the MPC-based Design . . . . . . . . . . . . . . . . . . . . . . 74 Single Return Map and Discrete-Time Model Partial Control Design: Dead-Beat Control System Parameters and Initial Conditions 4.5 Simulation Results . . . . . CHAPTER 5 ENERGY-BASED ORBITAL STABILIZATION USING CONTINUOUS . . . . . . . . . Introduction . INPUTS AND IMPULSIVE BRAKING . . . . . . . . . . . . . . . . . . . 77 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 5.1 . . 5.2 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 5.3 Modeling of System Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5.3.1 Lagrangian Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 5.3.2 Effect of Impulsive Inputs . . . . . . . . . . . . . . . . . . . . . . . . . . 80 5.3.3 Hybrid Dynamical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.4.1 Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 5.4.2 Choice of Controller Gains . . . . . . . . . . . . . . . . . . . . . . . . . . 89 5.4 Hybrid Control Design . . . viii 5.5 Illustrative Example - The Tiptoebot 5.5.1 5.5.2 Selection of Controller Gains . . . . . . . . . . . . . . . . . . . . . . . . Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . System Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6.1 5.6.2 Selection of Controller Gains . . . . . . . . . . . . . . . . . . . . . . . . 5.6.3 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 . 91 . 92 5.6 Experimental Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 . 94 . 95 . 97 5.7 Proofs and Additional Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . 99 . 99 5.7.1 . 100 5.7.2 . 101 5.7.3 5.7.4 Nonoccurrence of Zeno Phenomenon . . . . . . . . . . . . . . . . . . . . 102 . . . . . . . . . . . . . . . . . . . . . 103 5.7.5 Well-posedness of Switching Times . 103 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 5.7.6 Quasi-Continuous Dependence Property . . . . . . . . . . . . . . . . . . . 104 . 104 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 5.7.7 Verification of Assumption 2 for Tiptoebot and Rotary Pendulum . . . . . . 105 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7.5.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7.5.2 5.7.6.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7.6.2 Proof of Lemma 1 . Proof of Lemma 2 . Proof of Lemma 3 . Proof . Proof . CHAPTER 6 ORBITAL STABILIZATION USING VIRTUAL HOLONOMIC CON- . . . . . . . . . 6.3.1 6.3.2 6.3.3 Introduction . Problem Statement 6.1 . 6.2 Problem Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1 6.2.2 6.2.3 Zero Dynamics and Periodic Orbits 6.2.4 STRAINTS AND IMPULSE CONTROLLED POINCARÉ MAPS . . . . . 106 . 106 . 107 System Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 . . . . . . . . . . . . . . 108 Imposing Virtual Holonomic Constraints (VHC) . . . . . . . . . . . . . . . . . . . . . 109 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 6.3 Main Result: Stabilization of Od . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 Poincaré Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 Impulse Controlled Poincaré Map (ICPM) . . . . . . . . . . . . . . . . . . 113 Implementation of Control Design . . . . . . . . . . . . . . . . . . . . . . 116 6.3.3.1 Numerical Computation of A and B matrices . . . . . . . . . . 116 Impulsive Input using High-Gain Feedback . . . . . . . . . . . . 117 6.3.3.2 . 117 System Dynamics and VHC . . . . . . . . . . . . . . . . . . . . . . . . . 117 Stabilization of VHC and Od . . . . . . . . . . . . . . . . . . . . . . . . . 119 Simulation Results . 120 . . . . . . . . . . . . . . . . . . . . . . . . . 122 . 122 . . . . . . . . . . . . . . . . . . . . . 123 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 Illustrative Example: Cart-Pendulum . . . . . . . . . . . . . . . . . . . . . . . . 6.4.1 6.4.2 6.4.3 Illustrative Example - The Tiptoebot 6.5.1 6.5.2 6.5.3 6.5.4 System Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Imposing VHC and Selection of Od Stabilization of Od Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4 6.5 CHAPTER 7 CONCLUSION AND FUTURE WORK . . . . . . . . . . . . . . . . . . . 127 ix BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 x LIST OF TABLES Table 2.1: Kinematic and Dynamic Parameters for Tiptoebot in SI units . . . . . . . . . . . 16 Table 2.2: Lumped Parameters for Pendubot in SI units . . . . . . . . . . . . . . . . . . . . 30 Table 3.1: Swing-up time and maximum magnitude of high-gain torque required for different values of N . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 Table 5.1: Tiptoebot lumped parameters in SI units . . . . . . . . . . . . . . . . . . . . . . 91 Table 6.1: Tiptoebot lumped parameters in SI units . . . . . . . . . . . . . . . . . . . . . . 122 xi LIST OF FIGURES Figure 1.1: The inertia wheel pendulum. (a) RA and a point pq˚, 9q˚q that lies outside RA, (b) rRApq˚q and the point 9q˚, (c) IMp 9q˚q, rRApq˚q and pRApq˚, 9q˚q. . . . . . . . . . . . . . . . . . . . . . 12 . . . . . . . . . . . . . . . . . . . . . . . . . . 4 . (a) The Tiptoebot - a three-link underactuated system with two active joints at the hip and knee, and one passive joint at the toe (b) a simple model of a human balancing on the tip of the toe. . . . . . . . . . . . . . . . . . . . . . . 15 Figure 2.1: Figure 2.2: Figure 2.3: Slices of sos and lin, and their intersection with the impulse manifold, IM. A transformed coordinate system is used for better visualization. . . . . . . . . 19 coordinates; the intersection set pReA is an ellipse. . . . . . . . . . . . . . . . . . 20 Intersection of the slice of sos with the impulse manifold IM in the original Figure 2.4: Figure 2.5: Effective enlargement of Re A using SOS-IMM: Plot of joint angles, their velocities, and control inputs with time for stabilization of the Tiptoebot from the initial configuration given by (2.3.1). The effect of high-gain feedback is shown using a dilated time scale; the effect of the stabilizing controller is shown using a normal time scale. . . . . . . . . . . . . . . . . . . . . . . . . . 21 Figure 2.6: A “ 0 at t “ t0 yields (a) Discretization of the boundary of conservative Re S0, (b) Convex hull Conv(S1) constructed at time t “ t1, (c) Enlarged Re obtained as Conv(S1) at time t “ t1`k. . . . . . . . . . . . . . . . . . . . . . . 25 A Figure 2.7: Slices of chart, sos and lin, and their intersection with the impulse man- ifold, IM. For better visualization and for ease of comparison, the same coordinate system of Fig.2.3 is used. . . . . . . . . . . . . . . . . . . . . . . . 26 Intersection of the slices of chart and sos with the impulse manifold IM in the original coordinates; the intersection set pReA of chart is a convex region and encloses the pReA of sos. . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Figure 2.8: Figure 2.9: Effective enlargement of Re A using SOS-CHART-IMM: Plot of joint angles, their velocities, and control inputs with time for stabilization of the Tiptoebot from the initial configuration given by (2.3.1). . . . . . . . . . . . . . . . . . . 28 xii Figure 2.11: An arbitrary configuration of the two-link pendubot. of chart with the impulse manifold IM, shown in the original coordinates. Figure 2.10: (a) Initial configuration of the Tiptoebot, given by (2.5.2), lies outside chart, (b) rReA of chart (shown in transformed coorrdinates for better visualization) is a convex hull whereas rReA of sos is a null set, (c) intersection of the slice Figure 2.12: Intersection of rReA of chart with IM; the intersection set pReA of chart is configuration p 9q2, 9q1q “ p2.00,0.05q was chosen to lie on pReA of chart. A zoomed-out scale has been used to show the entire rReA of chart; due to this the dark line segment. The impulse manifold IM was determined by the initial velocity configuration p 9q2, 9q1q “ p0.00,0.25q. The desired velocity scaling the initial and desired velocity configurations appear to be close. . . . . 32 . . 29 . . . . . . . . . . . . . . . 30 Figure 2.13: Steps of experimental verification using the pendubot shown in Fig.2.11: (a) 1 s - high-gain feedback, (c) rt0,t´ 1 q - feedback linearizing control, (b) rt´ pt` 1 ,t fs - SOS stabilizing control . . . . . . . . . . . . . . . . . . . . . . . . . 33 1 ,t` Figure 2.14: Experimental verification of enlargement of Re A of pendubot: Plot of joint angles, their velocities, and control input with time. The effect of high-gain feedback is shown using a dilated time scale; the effect of the stabilizing controller is shown using a normal time scale. . . . . . . . . . . . . . . . . . . 34 Figure 2.15: Simulation results of ideal pendubot behavior for comparison with experi- mental results presented in Fig.2.14. . . . . . . . . . . . . . . . . . . . . . . . 34 Figure 3.1: The inertia wheel pendulum is shown in an arbitrary configuration. . . . . . . . 36 Figure 3.2: An example showing the initial, intermediate, and final configuration of the IWP for the case with two impulsive inputs (N “ 2). . . . . . . . . . . . . . . . 42 Figure 3.3: Simulation results for the globally stabilizing controller [1] with controller parameter values: c0 “ ´π{10, c1 “ 13, c2 “ 16 and c3 “ 8.0. . . . . . . . . . 51 Figure 3.4: Simulation results for the globally stabilizing controller [1] with controller parameter values: c0 “ ´π{10, c1 “ 13, c2 “ 16 and c3 “ 4.5. . . . . . . . . . 52 Figure 3.5: High-gain feedback implementation of two impulsive inputs (N “ 2) for swing-up of the IWP. For the purpose of comparison with the results in Figs.3.3 and 3.4, the controller is designed to keep θ in the domain p´π{2,3π{2s. 53 Figure 3.6: High-gain feedback implementation of the sequence of two optimal impulsive inputs (N “ 2) for swing-up of the IWP. The controller is designed to keep θ in the domain p´3π{2,π{2s. . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 xiii Figure 3.7: High-gain feedback implementation of the sequence of eight optimal impul- sive inputs (N “ 8) for swing-up of the IWP. The high-gain controller was implemented with ε “ 0.02. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 Figure 3.8: Simulation using the PFBLC + AL controller in [2]; the controller parameters were chosen as ke “ 3.1ˆ107 and kv “ 0.1. This choice of parameters ensured that the time required for swing-up and the magnitude of the maximum control torque in simulations matched those of the experiments. The initial configuration was chosen to be slightly different from θ0 “ ´π{2 since the controller is unable to swing-up from this configuration. . . . . . . . . . . . . . 56 Figure 4.1: A three degree-of-freedom of a devil-stick. . . . . . . . . . . . . . . . . . . . . 58 Figure 4.2: Symmetric configurations of the devil-stick in Fig. 4.1. . . . . . . . . . . . . . 59 Figure 4.3: (a) Ambidexterous juggler standing at P and applying control actions with both hands, (b) right-handed juggler standing at P and applying control action with right hand, (c) right-handed juggler standing at Q and applying control action with right hand. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 Figure 4.4: State variables and total energy E of the devil-stick at sampling instants k, k “ 1,2,¨¨¨ ,10, for the LQR design. . . . . . . . . . . . . . . . . . . . . . . . 73 Figure 4.5: Control inputs for the devil-stick at sampling instants k, k “ 1,2,¨¨¨ ,10, for . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . the LQR design. . . . 73 Figure 4.6: State variables and control inputs of the devil-stick at sampling instants k, k “ 1,2,¨¨¨ ,15, for the MPC design. (a) Trajectory of the center-of-mass from the initial configuration to steady- state and (b) symmetric configurations and seven intermediate configurations of the devil-stick in steady state for the MPC design. . . . . . . . . . . . . . . . 75 . . . . . . . . . . . . . . . . . . . . . . . 75 Figure 4.7: Figure 5.1: Plots of the joint angles θ1, θ2, θ3, error in the desired energy pE ´ Edesq, and the active joint velocities 9θ2, 9θ3 of the Tiptoebot. . . . . . . . . . . . . . . 93 Figure 5.2: Plots showing (a) phase portrait of passive joint angle θ1, and (b) variation of the Lyapunov-like function V. The desired orbit is shown using dashed green line in (a). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 . . . . . . Figure 5.3: Schematic of a rotary pendulum. . . . . . . . . . . . . . . . . . . . . . . . . . 94 xiv Figure 5.4: Rotary pendulum experimental results: (a)-(d) are plots of joint angles and joint velocities, (e) control torque, and (f) derivative of Lyapunov-like func- tion. The brake pulses are shown within plots (e) and (f), the peaks represent time intervals when the brakes were engaged. . . . . . . . . . . . . . . . . . . . 97 Figure 5.5: Rotary pendulum simulation results. . . . . . . . . . . . . . . . . . . . . . . . 98 Figure 6.1: Schematic of ICPM approach to orbital stabilization. . . . . . . . . . . . . . . . 115 Figure 6.2: Inverted pendulum on a cart. . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 Figure 6.3: Orbital stabilization for the cart-pendulum system for the initial conditions in [3]: (a) plot of ρ with respect to time, (b) phase portrait of the pendulum, (c) plot of the norm of the error of the discrete-time system. . . . . . . . . . . . 121 Figure 6.4: Orbital stabilization for the cart-pendulum system for the initial conditions in (6.4.10): (a) plot of ρ with respect to time, (b) phase portrait of the pendulum; Od is shown in red, (c) plot of the norm of the error of the discrete-time system. 122 Figure 6.5: Phase portrait of tiptoebot zero dynamics. The desired orbit Od is shown in red. 124 Figure 6.6: Orbital stabilization for the tiptoebot using ICPM: (a) and (b) provide plots of ρ1 and ρ2, (c) and (d) provide plots of 9ρ1 and 9ρ2, (e) and (f) provide the phase portrait of the passive joint, (g) provides the norm of the error for the discrete-time system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 . xv LIST OF ALGORITHMS CHART (Convex Hull Algorithm Using Reversal of Trajectories) . . . . . . . . . . . . . . 24 xvi KEY TO SYMBOLS g (cid:96)1 acceleration due to gravity, (m/s2) distance of center-of-mass of pendulum from the passive joint, (m) length of the pendulum, (m) (cid:96)2 m1 mass of the pendulum, (kg) m2 t0 ti t f xy I1 I2 N θ φ 9θ 9φ i θi 9θ´ 9θ` 9φ´ i i combined mass of the motor and the wheel, (kg) initial time, (s) instant of time at which the i-th impulsive torque, is applied, i “ 1,2,¨¨¨ ,N, (s) final time, (s) inertially fixed Cartesian coordinate frame mass moment of inertia of the pendulum about its center of mass, (kg.m2) mass moment of inertia of the wheel about its center of mass, (kg.m2) number of impulsive torques applied angular position of the pendulum, measured CCW with respect to the x axis, (rad) angular position of the wheel, measured CCW with respect to the pendulum, (rad) angular velocity of the pendulum, (rad/s) angular velocity of the wheel, (rad/s) value of θ at time ti value of 9θ immediately before application of impulsive torque at time ti value of 9θ immediately after application of impulsive torque at time ti value of 9φ immediately before application of impulsive torque at time ti xvii 9φ` i τ τi τhg value of 9φ immediately after application of impulsive torque at time ti torque applied by the motor on the wheel, (Nm) impulsive torque applied at time ti, (Nm) continuous-time implementation of impulsive torque using high-gain feedback, (Nm) angular impulse of impulsive torque τi applied at time ti, (Nm-s) Ii I N the N-dimensional vector rI1 I2 ¨¨¨ INs xviii CHAPTER 1 INTRODUCTION 1.1 Motivation A few decades ago, robots were primarily used for doing repetitive tasks in manufacturing lines and were supposed to work in risk-free and familiar environments. In recent times, there has been an increasing interest among engineers and scientists to create robotic systems which can operate in non-familiar environments. The applications of such robotic systems are immense as they can be used in application where human health is at risk such as in extra-terrestrial space explorations, rescue operations in the aftermath of disasters, construction in hostile environments such as deep sea etc. The new-age robots must be capable of locomotion and as a result, design, control and stability of robotic systems such as legged robots have gained popularity. Legged robots fall under the category of underactuated systems as there is no actuation at the ground-foot contact point. While underactuation is inherent in legged robots, it is purposely introduced in bio-inspired models such as in the pendubot [4], acrobot [5], tiptoebot [6] etc. Several control objectives have been studied for underactuated robotic systems such as stabilization about an equilibrium configuration, swing-up to an equilibrium configuration corresponding to highest potential energy and generating stable repetitive motion. Since the dynamics of such systems are typically nonlinear, and due to the presence of fewer control inputs than their degree-of -freedom, their control design is often very challenging. Due to this reason, it is difficult to design general control methodologies for underactuated systems. Impulsive inputs are theoretically modeled as Dirac-delta functions which causes discontinuous change in the velocity of a dynamical system. Impulsive forces occur naturally in several mechanical systems that experience impacts; for example walking [7] and hopping robots [8], juggling systems [9] etc. Modeling of such systems have been widely investigated in the framework of hybrid dynamical systems [10]. However, the use of impulsive inputs for control of systems, especially 1 underactuated mechanical systems have been rather limited. Few works that have previously proposed the use of impulsive inputs in control of some underactuated systems can be found in [11,12]. Impulsive inputs can be extremely useful where sudden change in system’s configuration is required to attain certain control objective. This work presents several control problem where impulsive inputs are exploited to simplify the control of underactuated robotic systems; these problems are discussed next. 1.2 Estimating and Enlarging the Region of Attraction of Equilibria Underactuated mechanical systems typically have multiple isolated equilibria and often the control design objective is to stabilize one of the equilibrium points that is unstable. Several methods have been proposed for stabilizing an equilibrium point for underactuated systems. These include the controlled Lagrangian method [13, 14], the interconnection and damping assignment passivity- based control (IDA-PBC) method [15], and the λ-method [16]. Other than these general methods, control designs have been proposed for stabilizing equilibrium points of specific underactuated systems such as the pendubot [17–19], acrobot [5,20], inverted pendulum on a cart [21], the reaction- wheel pendulum [1, 2], and the ball on beam system [22]. Irrespective of the control method used, a stabilized equilibrium of an underactuated system will have a finite region of attraction since the majority of all control designs cannot guarantee global stabilization. It is not trivial to obtain a non- conservative estimate of the region of attraction; and with a conservative estimate, it is only possible to guarantee stable behavior for small disturbances. To enable underactuated systems remain stable for large disturbances, we first present a method for obtaining a non-conservative estimate of the region of attraction. Then, the impulse manifold method (IMM) [23] is used to identify the region that lies outside this estimate and from where the equilibrium can be stabilized. A portion of this region may lie outside the region of attraction implying that the IMM can effectively enlarge the estimate of the region of attraction. For underactuated systems, the IMM [23] uses impulsive inputs to move the configuration of the system from outside the estimated region of attraction to inside the region. Impulsive 2 inputs have been used for control of underactuated systems1 [11, 12, 19, 24] and experimental investigations [23, 25] have relied on high-gain feedback for continuous approximation of the impulsive inputs. In an early work [26], high-gain control was used to force the trajectory of an acrobot into the region of attraction of a stabilizing controller. The limitations of this work are that the stabilizing controller must be pseudolinear [27] and the system cannot have more than one input. In [23], impulsive inputs were used to move the configuration of the system from a point lying immediately outside the region of attraction to inside the region; the boundary of the region was found iteratively by trial and error. Since it is not possible to determine the region of attraction for the general case, the IMM is used here in conjunction with an estimate of the region. Both analytical and numerical methods have been used for estimating regions of attraction, [28–30], for example. The sum of squares (SOS) method uses a polynomial approximation of the system dynamics to maximize the estimate of the region of attraction [31–34]. It can also be used to enlarge the region of attraction by optimizing the controller in polynomial form [35,36]. Starting from an initial estimate, the trajectory reversing method enlarges an estimate using time-reversed trajectories of the system, [37–40], for example. The points obtained by time-reversal of trajectories lie on the surface of a larger estimate but it is challenging to obtain an analytical description of that surface in the general case. Both the SOS method and the method of trajectory reversing have been used with the IMM in our earlier work [41, 42]. We combine the two methods in this work to enhance the applicability of the IMM. For a given order of the controller polynomial, the SOS method is first used to obtain the largest estimate of the region of attraction. Using this estimate as the initial estimate, a larger estimate is obtained using trajectory reversing. The generality of the combined method is illustrated using the example of the Tiptoebot - a three-link underactuated system with one passive joint. The result is notable since the trajectory reversing method has been illustrated in the literature using systems with two states only; the system considered here has six states. In addition to simulations, experimental validation is provided by stabilizing an equilibrium of a pendubot from a point lying 1Impulsive control of underactuated systems should be differentiated from control of underac- tuated systems subjected to impulsive disturbances. 3 outside the region of attraction estimated using the SOS and trajectory reversing methods. The impulse manifold passing through this point does not intersect the region of attraction estimated using the SOS method alone; this illustrates the usefulness of combining the SOS and trajectory reversing methods for application of the IMM. 1.3 Rest-to-Rest Maneuver of the Inertia Wheel Pendulum y g c.m. τ ℓ2 ℓ1 φ motor θ pendulum inertia wheel x Figure 1.1: The inertia wheel pendulum. The IWP is comprised of a simple pendulum and a motor mounted at its distal end; the motor drives a wheel - see Fig.1.1. The torque produced by the motor can be used to accelerate the wheel in both CW and CCW directions and the resulting reaction torque can be used to control the pendulum angle. The IWP bears similarity to the underactuated Acrobot [5] where in place of the the wheel, a link is driven by the motor. Since the wheel is symmetric, the angular displacement of the wheel is not included in the state-space representation and unlike the Acrobot, the only source of nonlinearity in the equation of motion is the gravity term. The IWP is an ideal candidate for impulsive control since the dynamics of the system can be described by simple algebraic equations: the effect of impulsive forces can be described by changes in the velocities of the system along the impulse manifold [23], and conservation of energy and conservation of wheel momentum describe the dynamics when no torque is applied by the motor. 4 For the IWP, stabilization of its upright posture and swing-up to this configuration have been investigated in the literature, [1, 2, 15, 43–46], for example. Some of the early work on the IWP can be found in [2] where a hybrid controller comprised of separate swing-up and balancing controllers was designed using energy methods. Swing-up and stabilization was achieved using a single controller using the IDA-PBC method [15, 43] and a global change of coordinates that converted the dynamics of the IWP into strict feedback form [1]. The IDA-PBC method based controllers [15, 43] and the globally stabilizing controller [1] result in high wheel velocities and actuator torques. The swing-up problem of the IWP is a rest-to-rest maneuver between its lowest and highest potential energy configurations. In this work, we address the general problem of designing rest-to- rest maneuvers using a sequence of impulsive inputs. The use of impulsive inputs simplifies the dynamics of the IWP and permits the design of optimal sequences under some general assumptions. It is found that a sequence with an odd number of inputs is less optimal than the two adjacent sequences with even number of inputs. The high wheel velocities and/or actuator torques associated with some results in the literature [1,15,43], is explained using the analysis based on two impulsive inputs. An optimal sequence of impulsive inputs, implemented using high-gain feedback [23], is used for swing-up of the IWP; the simulations results show similarities with the energy-based methods in [2]. It is shown that an optimal input sequence can be designed to accommodate torque constraint of the actuator; this has not been discussed in earlier works. 1.4 Devil-Stick Juggling A devil-stick is typically juggled using two hand sticks and several tricks can be performed depending on the proficiency of the juggler. Some of the common tricks are: standard-idle, flip-idle, airplane-spin or propeller, top-only idle, and helicopter [47]. The top-only idle is one of the simplest tricks and is the focus of this investigation; a video tutorial for learning this trick can be found in [48]. In top-only idle, intermittent forces are applied to the devil-stick. Since the devil-stick is never grasped, the juggling problem can be viewed as a non-prehensile manipulation 5 problem. If robots are to perform this trick, the motion of the end-effectors would have to be coordinated and controlled to apply the correct magnitude of impulsive forces to the devil-stick at appropriate locations. We do not focus on the end-effector motion control problem (see [49, 50] for application to ball juggling); instead, we investigate the magnitude and location of the forces needed to perform the top-only idle trick. Many juggling tasks, including the top-only idle trick, involve intermittent application of impulsive forces and several researchers [51–55] have studied the controllability and stability of such systems. Although impulsive control of the devil-stick has not been investigated, the control problem associated with juggling of balls and air-hockey pucks has seen several solutions [9,56–58]. In all of these solutions, the object being juggled has been modeled as a point mass and its orientation is excluded from the dynamic model. In contrast, for devil-stick tricks such as top-only idle, the stick is shuffled between two symmetric configurations about the vertical; therefore, the orientation of the stick must be included in the dynamic model. In earlier works on the devil-stick [59, 60], controllers have been designed for airplane-spin or propeller-type motion; a single hand-stick is used to rotate the devil-stick about a virtual horizontal axis using continuous-time inputs. The dynamics model and control design of top-only idle motion of the devil-stick has not appeared in the literature; to the best of our knowledge, it is presented here for the first time. It is assumed that impulsive forces are applied intermittently to the devil-stick and the control inputs are the impulse of the force and its point of application on the stick. The control inputs are designed to juggle the stick between two symmetric configurations about the vertical, starting from an arbitrary initial configuration. 1.5 Energy-Based Orbital Stabilization For underactuated mechanical systems, the problem of stabilization of an equilibrium has been investigated widely, see [2, 13–16, 18, 21, 61], for example, and the references therein. In many applications, such as in legged locomotion, the system is required to undergo periodic motion; therefore, the control objective is to stabilize an orbit rather than an equilibrium. To achieve 6 repetitive motion, the virtual holonomic constraint (VHC) approach [3, 62, 63] has been proposed. The VHC imposes a set of geometric constraints on the generalized coordinates of the system and has been demonstrated in bipedal locomotion [7, 64]. Experimental validation of orbital stabilization using VHC can be found in [65, 66]. Orbital stabilization has also been used for swing-up control of underactuated systems with one passive degree-of-freedom (DOF). Some examples include two-DOF systems such as the pendubot [17], the acrobot [67], the reaction-wheel pendulum [68], inverted pendulum on a cart [69, 70], the rotary pendulum [71], and the three-DOF gymnast robot [72]. These controllers stabilize an orbit that include the equilibrium, which is typically unstable. The controllers are designed to gradually pump energy in and out of the system and converge the active DOFs to their desired configuration. Such control designs are typically based on a Lyapunov-like function that is comprised of terms involving positions and velocities of the active DOFs and the total mechanical energy of the system. The passivity property of the system is used to make the time derivative of the Lyapunov-like function negative semidefinite. A stability analysis is then carried out using LaSalle’s invariance principle [73]; this typically results in certain restrictions on controller gains and initial conditions. In all of these prior works, the control strategies for orbital stabilization have been designed separately for individual systems. Although the structure of the Lyapunov-like function is identical, the stability analysis is different for each system due to the difference in the nature of their nonlinear dynamics. Despite the effectiveness of the individual controllers, a general methodology for energy-based orbital stabilization does not exist. We present a control strategy for n-DOF underactuated systems with one passive DOF based on continuous time inputs and impulsive braking. A combination of continuous and impulsive inputs have been recently used for stabilization of a homoclinic orbits of two-DOF underactuated systems [74]. The current work provides a generalization of the theory to n-DOF systems and experimental validation. The main contributions of this work are as follows: 1. The control design is applicable to a class of underactuated systems which include a majority 7 of the commonly investigated underactuated systems. 2. The stability analysis is presented for the general case and it results in sufficient conditions that are not restrictive and can be checked for a given system. 3. Experimental validation is provided. 4. Impulsive braking is accomplished using a friction brake; this eliminates the need for high-gain feedback [23] which may result in actuator saturation. 1.6 Orbital Stabilization using Virtual Holonomic Constraints For underactuated systems, Virtual Holonomic Constraint (VHC) based control designs have gained popularity due to their conceptual simplicity and applicability to control of repetitive motion. They have been used for stabilization of gaits in bipeds [64, 75–80], which undergo impacts, and trajectory control for systems with open kinematic chains that are not subjected to impacts [3, 62, 63, 65, 66, 81–84]. VHCs parameterize the active joint variables in terms of the passive joint variables and confine system trajectories to a constraint manifold [62]. To enforce the VHC, the constraint manifold has to be stabilized using feedback. Typically, a constraint manifold contains a dense set of periodic orbits and the choice of repetitive motion determines the specific orbit that has to be stabilized. To stabilize gaits in bipeds, for example, Grizzle et.al [64, 76, 79] enforced the VHC and periodic loss of energy due to ground-foot interaction was exploited for orbital stabilization. A special class of underactuated systems are those with open kinematic chains and one passive DOF. For such systems, Shiriaev and collaborators [3, 65, 66] used VHC to select the desired orbit. For an n-DOF system, the 2n dimensional dynamics is linearized about the desired orbit; this results in a 2n´1 dimensional system. A periodic Ricatti equation is then solved to design a time-varying controller that stabilizes the orbit. It should be noted that the control designs in [3, 65, 66] stabilize the orbit but do not enforce the VHC. A control scheme that enforces the VHC and simultaneously stabilizes the orbit was recently proposed in [63]. The key idea is that the VHC is made time- 8 varying using a scalar parameter which is controlled via feedback. The stabilization problem involves solving a periodic Ricatti equation; however, unlike [3,65,66], where the dimension of the system is 2n´1, the dimension of the system in [63] is always three. For systems with more than two DOF, the method in [63] reduces the computational complexity of control implementation. Also, by enforcing the VHC, it improves control over transient characteristics of the trajectory [85]. Similar to [63, 85], we design a continuous controller with the objective of enforcing the VHC and stabilizing the constraint manifold. We then design impulsive inputs that work in tandem with the continuous inputs to stabilize a desired periodic orbit on the constraint manifold. Compared to the methods proposed earlier [3, 63, 65, 66], where periodic Ricatti equations have to be solved, our method has lower computational cost and complexity. Since impulsive inputs are used to control the Poincaré map, the dynamics of the closed-loop system can be described by the Impulse Controlled Poincaré Map (ICPM). The simplicity and generality of the ICPM approach to orbital stabilization is demonstrated using the examples of the 2-DOF cart-pendulum and the 3-DOF tiptoebot. 9 CHAPTER 2 ESTIMATION OF THE REGION OF ATTRACTION OF EQUILIBRIA AND ITS ENLARGEMENT USING IMPULSIVE INPUTS 2.1 Introduction Stabilization of an equilibrium point is an important control problem for underactuated systems. For a given stabilizing controller, the ability of the system to remain stable in the presence of disturbances depends on the size of the region of attraction of the stabilized equilibrium. The sum of squares (SOS) and trajectory reversing methods are combined together to generate a large estimate of the region of attraction. Then, this estimate is effectively enlarged by applying the impulse manifold method (IMM), which can stabilize equilibria from points lying outside the estimated region of attraction. The IMM and the SOS method are reviewed in section 2.2. The Tiptoebot is introduced in section 2.3 and the IMM is used to stabilize its equilibrium from a point lying outside the region of attraction estimated by the SOS method. A convex hull algorithm for estimating the region of attraction using trajectory reversing is presented in section 2.4. The benefit of using the SOS method, the convex hull algorithm, and IMM together is illustrated using the example of the Tiptoebot in section 2.5. Experimental results are presented in section 2.6. 2.2 Background 2.2.1 The Impulse Manifold Method For an n degree-of-freedom dynamical system, Lagrange’s equations have the form: Mpqq :q` Hpq, 9qq “ T (2.2.1) where q P Rn and T P Rn represent the vectors of generalized coordinates and generalized forces; Mpqq P Rnˆn denotes the generalized mass matrix; and Hpq, 9qq P Rn denotes the vector of centrifu- gal, Coriolis, gravitational and friction forces. If the system is underactuated with m active and 10 pn´ mq passive degrees-of-freedom, m ă n, (2.2.1) can be rewritten in the form: »—– M11pqq M12pqq MT 12pqq M22pqq fiffifl »—– :q1 :q2 fiffifl`»—– h1pq, 9qq h2pq, 9qq fiffifl “»—– 0 u fiffifl (2.2.2) 1 , qT 2 qT , and q1 P Rpn´mq and q2 P Rm denote the generalized coordinates correspond- where q “ pqT ing to the passive and active degrees-of-freedom, respectively, and u P Rm is the vector of control inputs. a stabilizing controller. We assume that pq, 9qq “ p0, 0q is an equilibrium point of the unforced system and u “ us is If RA is the region of attraction of the equilibrium with u “ us, the equilibrium cannot be stabilized from configurations that lie outside RA. However, it may be possible to stabilize the equilibrium after moving the configuration of the system from outside RA to inside RA using the Impulse Manifold Method (IMM) [23]. In the IMM, impulsive inputs are applied to the active degrees-of-freedom. Ideal impulsive inputs cause no change in the generalized coordinates but result in discontinuous changes in both active and passive generalized velocities. These changes can be computed from the following equation which is obtained by integration of (2.2.2) over the infinitesimal interval of time Δt, during which the impulsive inputs act [86]: »—– M11 M12 MT 12 M22 fiffifl »—– Δ 9q1 Δ 9q2 fiffifl “»—– 0 I fiffifl , I fiż Δt 0 u dt (2.2.3) In the above equation, I P Rm is the vector of impulses, and the velocity jumps, Δ 9q1 and Δ 9q2, are defined as Δ 9q1 fi p 9q` 1 ´ 9q´ 1 q, Δ 9q2 fi p 9q` 2 ´ 9q´ 2 q where 9q´ and 9q` are the generalized velocities immediately before and after application of the impulsive inputs. Due to the underactuated nature of the system, the discontinuous changes in the velocities of the passive coordinates are dependent on those of the active coordinates; this dependence is obtained from (2.2.3) as follows: p 9q` 1 ´ 9q´ 1 q “ ´M´1 11 M12p 9q` 2 ´ 9q´ 2 q (2.2.4) 11 In the n-dimensional velocity space, the impulse manifold is the m-dimensional manifold repre- sented by (2.2.4). It was stated earlier that the generalized coordinates do not change under the application of an impulse. Hence, for a fixed value of q, the impulse manifold passing through the point p 9q´ 2 q is defined by the set 1 , 9q´ IMp 9q´ 1 , 9q´ 2 q “ t 9q` 1 , 9q` 2 | p 9q` 1 ´ 9q´ 1 q “ ´M´1 11 M12p 9q` 2 ´ 9q´ 2 qu (2.2.5) The IMM is explained with the help of Fig.2.1. The region of attraction RA is shown in Fig.2.1 (a). An arbitrary point outside RA is denoted by pq˚, 9q˚q. For q “ q˚, the slice of RA in the n-dimensional velocity space is denoted by rRApq˚q and is shown in Fig.2.1 (b). Since pq˚, 9q˚q lies outside RA, the point 9q˚ lies outside rRApq˚q. The m-dimensional impulse manifold IMp 9q˚q is depicted by the plane in Fig.2.1 (c); its intersection with rRApq˚q is shown by the hatched region and is mathematically defined by the set Impulsive inputs can be applied to change the velocity of the system from 9q˚ to any other point pRApq˚, 9q˚q “ IMp 9q˚qX rRApq˚q (2.2.6) (desired velocity) on the impulse manifold. However, to move the configuration of the system to a point inside RA, the desired velocity should be chosen to lie in pRApq˚, 9q˚q. For the IMM to be applicable, pRApq˚, 9q˚q should not be a null set and the desired velocity should be chosen such that the magnitude of the impulse I does not violate actuator constraints. IMp 9q˚q RA Ă R 2n pq˚, 9q˚q 9q˚ 9q˚ rRApq˚q Ă Rn (a) (b) pRApq˚, 9q˚q (c) Figure 2.1: (a) RA and a point pq˚, 9q˚q that lies outside RA, (b) rRApq˚q and the point 9q˚, (c) IMp 9q˚q, rRApq˚q and pRApq˚, 9q˚q. 12 Ideal impulsive inputs, which are Dirac-delta functions, cannot be implemented in real physical systems. For an underactuated system of the form as in (2.2.2), impulsive inputs can be implemented using high-gain feedback [23]; the control law is given by1 u “ uhg “ rCT M´1 Cs´1„CT M´1 where C P Rnˆm is the matrix C “ r0 EsT , E P Rmˆm is the identity matrix, 9qdes desired velocities of the active joints, Λ fi diag„λ1 λ2 2 ´ 9q2¯ Λ´ 9qdes ¨¨¨ λm, where λi, i “ 1, 2,¨¨¨ , m are is the vector of (2.2.7) H` 2 1 µ positive numbers, and µ is a small positive number. In general, an exact determination of RA is not possible and an estimate of the region of attraction Re A, Re A Ă RA, has to be used. As we replace RA with Re same manner that rRA and pRA were defined in relation to RA. A, we define rRe A and pRe A in the From the discussion in this section it is obvious that the IMM will be more effective for larger Re A. When applicable, the IMM will effectively enlarge the Re A of the equilibrium. 2.2.2 The Sum of Squares Method The simplest way to obtain an Re A is through linearization [73]: the dynamics of the closed-loop underactuated system in (2.2.2) is first represented in state-space form 9x “ frx, u “ uspxqs “ gpxq (2.2.8) where x “ pqT , 9qTqT and uspxq is a stabilizing controller. The Re A of the asymptotically stable equilibrium x “ 0 can be defined with the help of the quadratic Lyapunov function V0pxq as follows (2.2.9) Re A “ tx | V0pxq ď cu fi Ωlin, V0pxq fi xT Px where P is the positive definite matrix obtained by solving the Lyapunov equation PA` AT P “ ´Q, A fi„Bg Bxx“0 , Q “ QT ą 0 1The high-gain feedback in [23] has been modified here for its applicability to systems with more than one active joint: the diagonal matrix Λ has been introduced to provide flexibility in choosing different gains for the different active joints. It can be easily shown that this modification does not affect the continuous approximation of the impulsive inputs. 13 and c is the largest number that satisfies tV0pxq ď cu Ď t 9V0pxq ď 0u The Sum of Squares (SOS) method [31] automates the process of finding a Lyapunov function using numerical tools from convex optimization. Subsequently, an Re A can be obtained by solving the following optimization problem [36]: subject to maximize ρ Vpxq Lpxq ´ 9Vpxq` LpxqpVpxq´ ρq is SOS is SOS is SOS $’’’’’’&’’’’’’% where Vpxq is the Lyapunov function and Lpxq is a polynomial multiplier. The decision variables in this optimization problem are Lpxq, Vpxq and ρ. The third constraint guarantees tVpxq ď ρu Ď t 9Vpxq ď 0u and therefore an estimate of the region of attraction is Re A “ tx | Vpxq ď ρu fi Ωsos (2.2.10) In the above optimization problem, Re The controller can also be optimized to further enlarge Re A is maximized for a pre-determined stabilizing controller. A. The controller upxq is introduced as another decision variable in the third constraint of the SOS optimization problem: 9Vpxq “ BV Bx frx, upxqs This non-convex optimization problem was studied in [35, 36], where the following steps are performed sequentially and repeatedly till no further enlargement of Re A is observed: • Vpxq is fixed; Lpxq and upxq are varied to maximize ρ • Lpxq and upxq are fixed; Vpxq is varied to maximize ρ 14 The SOS algorithm requires an initial guess of the Lyapunov function Vpxq. If this initial guess is chosen to be V0pxq in (2.2.9), we can guarantee Ωsos Ě Ωlin which implies that the SOS method can improve upon the Re A obtained through linearization. 2.3 Enlarging the Region of Attraction Using SOS and IMM 2.3.1 An Illustrative Example From our discussion in section 2.2 it is clear that an Re A can be obtained using the SOS method and that this estimate can be effectively enlarged using the IMM. In this section we demonstrate the effective enlargement of the Re A for a three-link underactuated system with two active joints and one passive joint. A majority of articulated underactuated systems that have been considered in the literature have two degrees-of-freedom with one passive joint; examples include the pendubot, the g ℓ3 θ3 τ3 pm2, J2q d2 y ℓ1ℓ1 d1 pm3, J3q ℓ3 upper body upper leg ℓ2 d3 ℓ2 θ2 τ2 pm1, J1q ℓ1 lower leg θ1 x (a) (b) Figure 2.2: (a) The Tiptoebot - a three-link underactuated system with two active joints at the hip and knee, and one passive joint at the toe (b) a simple model of a human balancing on the tip of the toe. 15 acrobot, the rotary pendulum, inverted pendulum on a cart, etc. The three-link system is purposely chosen to demonstrate the applicability of the method to more complex systems. Consider the three-link underactuated system2 in Fig 2.2 (a). We have christened it the Tiptoebot since it models a human balancing on the tip of its toe - see Fig 2.2 (b). The three links are analogous to the lower leg, the upper leg, and the upper body comprised of the torso and head. The knee joint connecting the upper and lower legs, and the hip joint connecting the upper body and upper leg are actuated; the torques applied by the actuators in these joints are assumed to be positive in the counter-clockwise direction and are denoted by τ2 and τ3. The toe provides a simple point of support and is modeled as a passive joint. The lower leg, upper leg, and upper body have link lengths ℓ1, ℓ2 and ℓ3; their center-of-masses are located at distances d1, d2 and d3 as shown in Fig.2.2 (a); their mass and mass moments of inertia about their individual center-of-mass are denoted by pm1, J1q, pm2, J2q and pm3, J3q. The joint angles of the links are measured positive in the counter-clockwise direction and are denoted by θ1, θ2 and θ3; θ1 is measured relative to the x-axis whereas θ2 and θ3 are measured relative to the second and third links. Table 2.1: Kinematic and Dynamic Parameters for Tiptoebot in SI units m1 0.20 J1 m2 0.20 J2 m3 ℓ1 ℓ2 ℓ3 0.30 0.20 0.20 0.30 J3 d1 d2 d3 6.66ˆ 10´4 6.66ˆ 10´4 15.0ˆ 10´4 0.10 0.10 0.15 The kinematic and dynamic parameters of the Tiptoebot in Fig.2.2 are provided in Table 6.1. For these set of parameters, the equations of motion are of the form given by (2.2.2), where n “ 3, m “ 2 and u “ pτ2 τ3qT . The vertically upright posture or configuration pθ1,θ2,θ3, 9θ1, 9θ2, 9θ3q “ pπ{2, 0, 0, 0, 0, 0q is the desired equilibrium and hence q1 “ pθ1 ´ π{2q, q2 “ pθ2 θ3qT . The matrices M11, M12, M22, h1 and h2, which can be described as functions of θk, 9θk, k “ 1, 2, 3, are 2This system is a special case of the N-link system considered in [87]. 16 provided below: M11 “ 0.018C23 ` 0.032C2 ` 0.018C3 ` 0.046 M12 “»—– M22 “»—– 0.009C23 ` 0.016C2 ` 0.018C3 ` 0.023 0.009C23 ` 0.018C3 ` 0.009 0.018C3 ` 0.023 0.009C3 ` 0.009 0.009C3 ` 0.009 0.009 T fiffifl fiffifl h1 “ 0.100 9θ1 ` 0.441C123 ` 0.784C12 ` 1.177C1 ´ 0.016 9θ2 ´ 0.009p 9θ2 ` 9θ3q2 h2 fi„h21 h22T h21 “ 0.100 9θ2 ` 0.441C123 ` 0.784C12 ` 0.016 9θ2 1 S2 ´ 0.009 9θ2 2 S2 ´ 0.009 9θ2 3 S3 S23 ´ 0.018 9θ1p 9θ2 ` 9θ3q S23 ´ 0.018p 9θ1 ` 9θ2q 9θ3 S3 ´ 0.032 9θ1 9θ2 S2 3 S3 ` 0.009 9θ2 1 S23 ´ 0.018p 9θ1 ` 9θ2q 9θ3 S3 h22 “ 0.100 9θ3 ` 0.441C123 ` 0.009 9θ2 1 S23 ` 0.009p 9θ1 ` 9θ2q2 S3 where Si and Ci denote sinθi and cosθi, Si j and Ci j denote sinpθi`θjq and cospθi`θjq, and Ci jk denotes cospθi`θj`θkq. We assumed the presence of viscous friction in the joints of the Tiptoebot. This accounts for the terms linearly proportional to the angular velocities in the expressions for h1, h21 and h22. The stabilizing controller us in (2.2.8) was obtained through linearization by placing the poles of the closed-loop system at -8, -3, -2, -1, -0.7, -0.4. The controller has the form us “ rτ2 τ3sT where τ2 “ ´10.204pθ1 ´ π{2q´ 5.789θ2 ´ 2.140θ3 ´ 2.114 9θ1 ´ 1.144 9θ2 ´ 0.496 9θ3 τ3 “ ´7.996pθ1 ´ π{2q´ 4.284θ2 ´ 1.878θ3 ´ 1.632 9θ1 ´ 0.974 9θ2 ´ 0.295 9θ3 17 »——————————————– fiffiffiffiffiffiffiffiffiffiffiffiffiffiffifl The Lyapunov function V0 was obtained using (2.2.9); the matrices P and Q are 5271.9 2668.5 973.5 1054.4 626.9 241.7 2668.5 1350.7 492.7 533.75 317.3 122.3 P “ 973.55 492.80 179.8 194.73 115.7 44.65 1054.4 533.75 194.7 210.91 125.3 48.35 , Q “ 0.01ˆ diag„3 3 3 1 1 1 626.90 317.33 115.7 125.39 74.55 28.75 241.75 122.37 44.65 48.359 28.75 11.09 The estimated region of attraction Ωlin in (2.2.9) was defined by the value of c “ 0.020; this value was determined by following the steps [73] given below: 1. Find a set of points αi, i “ 1, 2,¨¨¨ , M, where 9Vpαiq “ 0. 2. Choose c ă ¯c “ min i Vpαiq. The SOS method was used to improve upon the estimate Ωlin. By fixing V “ V0, L and u were varied to maximize ρ and determine Ωsos in (2.2.10). The input u was chosen to be a first-order polynomial of the states and the maximum value of ρ was found to be 0.044. The SOS controller was obtained as follows: τ2 “ ´3.866pθ1 ´ π{2q´ 2.603θ2 ´ 1.025θ3 ´ 0.898 9θ1 ´ 0.417 9θ2 ´ 0.259 9θ3 τ3 “ 10.157pθ1 ´ π{2q` 4.887θ2 ` 1.410θ3 ` 1.942 9θ1 ` 1.156 9θ2 ` 0.476 9θ3 2.3.2 Simulation Results To demonstrate the enlargement of the Re A “ Ωsos using the IMM, we consider the following initial configuration that lies outside Ωsos: x “„q˚T 9q˚TT “„0.0 0.0 0.0 0.00 0.00 0.21T (2.3.1) For this configuration, it can be verified that V0pxq “ 0.48 ą ρ “ 0.044. Following the discussion Apq˚ “ 0q corresponding to Ωlin and Ωsos can be expressed as in section 2.2.1, the slices rRe 18 o f Ω lin A rRe IM o f Ωsos A rRe Figure 2.3: Slices of Ωsos and Ωlin, and their intersection with the impulse manifold, IM. A transformed coordinate system is used for better visualization. V0 |q“q˚“0ď c and V0 |q“q˚“0ď ρ, respectively. Both these slices are ellipsoids and are shown in Fig.2.3; the coordinates were scaled and transformed for better visualization. The impulse manifold IMp0, 0, 0.21q is a plane (two-dimensional manifold) in the three-dimensional velocity space; it is shown in Fig.2.3 in the transformed coordinates and mathematically described by the relation: 9q1 ` 0.583 9q2 ` 0.236 9q3 ´ 0.049 “ 0 (2.3.2) Apq˚ “ 0q “ t 9q | Apq˚ “ 0q “ t 9q | V0p0, 9qq ă cu A in (2.2.6) obtained using the SOS method is non-empty whereas that obtained using linearization is a null set; this illustrates the benefit of It can be seen from Fig.2.3 that the impulse manifold intersects the slice rRe V0p0, 9qq ă ρu obtained using the SOS method but not the slice rRe obtained using linearization. Consequently, the set pRe using a larger Re A in terms of applicability of the IMM. A obtained using the SOS method is an ellipse in the three-dimensional velocity space; it is shown together with the impulse manifold IM in Fig.2.4 in the original coordinates. To take The pRe the configuration of the Tiptoebot from outside Re A, we choose the following desired velocity configuration that lies in the interior of pRe „ 9q1 9qT 2 “„2.36 ´4.04 0.16 A to inside Re A: (2.3.3) 19 3.0 2.0 1.0 9q3 0.0 -1.0 A pRe p2.36, ´4.04, 0.16q 2.0 0.0 9q1 p0, 0, 0.21q IM 0.0 -2.0 9q2 -4.0 Figure 2.4: Intersection of the slice of Ωsos with the impulse manifold IM in the original coordi- nates; the intersection set pRe A is an ellipse. The desired effect of the impulsive inputs is shown by the dotted line joining the initial and desired gain feedback in (2.2.7) was used with 9qdes velocity configurations, both of which lie on IM - see Fig.2.4. To implement the IMM, the high- 2 “ 9q2 in (2.3.3), µ “ 0.0004, λ1 “ 10 and λ2 “ 1; the results are shown in Fig.2.5. Figure 2.5 plots the joint angles, their velocities, and the control inputs with time. The high-gain controller is active over t P r0, 0.004s; its effect is shown using a dilated time scale. It can be seen that the change in the joint angles are negligible but there are significant changes in the joint velocities. At t “ 0.004 when 9q2 « 9qdes off and the stabilizing controller is switched on; at this time, the system configuration is inside Re A 2 , the high-gain controller is switched and therefore the equilibrium can be stabilized. The stabilization of the equilibrium after t ą 0.004 is shown using a normal time scale. The high-gain feedback was based on µ “ 0.0004. The change in the joint angles were negligible, of the order of 0.01 rad. A larger value of µ could be used to reduce the magnitude of the control inputs but it would result in larger changes in the joint angles. The computation of pRe is based on the assumption that there is no change in the joint angles and therefore a small value of A µ is necessary to ensure the applicability of the IMM. For the general case, if δ denotes the small change in the joint angles due to high-gain feedback, 20 θ2 θ1 9θ3 2.0 1.0 π{2 0.0 3.0 0.0 -4.0 0.0 9θ1 9θ2 -1.5 τ2 -3.0 0.0 -2.0 -4.0 τ3 dilated time scale normal time scale (rad) θ3 (rad/s) (Nm) (Nm) 0.0 0.004 time (s) 1.5 3.0 Figure 2.5: Effective enlargement of Re A using SOS-IMM: Plot of joint angles, their velocities, and control inputs with time for stabilization of the Tiptoebot from the initial configuration given by (2.3.1). The effect of high-gain feedback is shown using a dilated time scale; the effect of the stabilizing controller is shown using a normal time scale. we should ensure that the desired velocity 9qdes 2 , which was chosen to lie in the interior of pRApq˚, 9q˚q, should be chosen also lies in the interior of pRApq˚ ` δ, 9q˚q. To this end, the desired velocity 9qdes sufficiently far from the boundary of pRApq˚, 9q˚q. There is significant flexibility in the choice of the desired velocity 9qdes 2 2 . This choice can be used judiciously to minimize the overall control effort or some other cost function. It was mentioned in section 2.2.1 that the IMM will be more effective for larger Re end, we present an algorithm in the next section for obtaining large estimates of Re A. To this A; the algorithm 21 is based on the method of trajectory reversing [37]. 2.4 Algorithm for Computing Re A Using the Method of Trajectory Reversing 2.4.1 Definitions For the closed-loop system in (2.2.8), an Re (2.2.10). The method of trajectory reversing [37] can be used to obtain a larger Re A of x “ 0 is given by Ωlin in (2.2.9) and Ωsos in A starting from a A is a convergent sequence of simply connected domains generated by backward integration from a discrete set of points on the boundary of the conservative Re A such as Ωlin or Ωsos. The larger Re conservative estimate. In the literature, the method of trajectory reversing has been used to obtain Re A for example systems described by two state variables. Here, we present an algorithm [42] that has been shown to provide larger estimates for the two-state systems considered in the literature; additionally, it will be used to obtain an Re A for the Tiptoebot, whose dynamics is described six state variables. To the best of our knowledge, estimation of the region of attraction using the method of trajectory reversing has not been reported in the literature for systems with more than two states. Before we present the algorithm, we present the following definitions: Definition 1. Convex Hull: A convex hull of a set of points S “ tx1, x2,¨¨¨ , xNu in Rn, denoted by Conv(S), is the smallest convex set that contains all the points. A convex hull can be mathematically represented by a set of equations and inequalities. In the two-dimensional plane, a convex hull is described by line segments which can be represented by equations of the lines and inequalities based on their lengths. For higher dimensions, the convex hull is represented by planes/hyper-planes and inequalities. The Quick Hull Algorithm [88] in Matlab can be used to compute the convex hull of a set of points efficiently. For a given convex hull, the Matlab function inhull [89] can be used to determine whether an arbitrary point lies in the interior of the hull or not. Definition 2. Slice of a Convex Hull: The slice of a convex hull is the set obtained by intersecting 22 the convex hull with a hyper-plane. The slice of a convex hull is also convex. It can be computed using the Matlab function intersectionhull [90]. Definition 3. Cluster: For a given set of points, a cluster is a subset where any two points in the subset is density-reachable for some pre-defined distance metric ε. Simply stated, two points are said to be density-reachable if there is a chain of intermediate points and the distance between neighboring points does not exceed ε. A more formal definition of cluster and a cluster identification algorithm can be found in [91]. A cluster of points can be identified using the Matlab function DBSCAN [92]. 2.4.2 Algorithm Nu on the boundary of the conservative Re We choose a discrete set of points S0 “ tx0 time t0, where N is a sufficiently large number. The conservative Re A at A, denoted by Ω0, can be Ωlin or Ωsos, for example. Now consider the time sequence ti, i “ 0, 1, 2,¨¨¨ , where tk “ t0 ` kΔt, and Δt is some pre-defined time interval. Define the set Si “ txi Nu, i “ 0, 1, 2,¨¨¨ , which is obtained by reversing the trajectories of the points in S0 for iΔt interval of time. Conv(Si) is an 2,¨¨¨ , xi 1, x0 2,¨¨¨ , x0 1, xi enlarged Re A if: 1. none of the points in Si`k, where k is some pre-defined positive integer, lies in the interior of Conv(Si), and 2. the vertices of Conv(Si) form a single cluster around the origin with distance metric ε [42]. For i, j “ 0, 1, 2,¨¨¨ , let m a time ti` j and NC(Conv(Si)) represent the number of clusters formed by the vertices of Conv(Si). denote the number of points which remain in the interior of Conv(Si) at i` j i The complete algorithm for obtaining a large estimate of Re A, which we refer to as CHART (Convex Hull Algorithm Using Reversal of Trajectories), is presented next. 23 Algorithm CHART (Convex Hull Algorithm Using Reversal of Trajectories) Initialization: Choose a sufficiently large integer N which represents the number of points located uniformly on the boundary of Ω0 and a sufficiently small distance metric ε. Step 1: Select a suitable time interval Δt and sufficiently large positive integers k1 and k2, k1 ă k2. Step 2: Let ℓ “ 0 For i “ 1 to k1 If ℓ “ i´ 1 If m For j “ 1 to k2 i` j i “ 0 and NC(Conv(Si)) “ 1 ℓ “ i Break (exit from the inner For loop which checks the condition on j) End If End For End If End For Step 3: If ℓ “ 0, repeat Steps 1 to 2 with a lower value of Δt than chosen in Step 1 Result: ConvpSℓq is the enlarged Re A. Parameter Selection: The choice of k1 and k2 is a trade-off between the enlargement of the Re A and the computation time. A few examples showing the choice of N, ε, Δt, k1, k2 are provided in [42]. For the ease of understanding, the CHART is explained for a system with two state variables 2,¨¨¨ , x0 with the help of Fig.2.6. At time t “ t0, the boundary of the conservative estimate Ω0 is discretized into the set of points S0 “ tx0 Nu, where N is a large number - see Fig.2.6 (a). At time t “ t1, it is found that some points in S1 lie in the interior of Conv(S1), i.e. m1 1 ­“ 0 - see Fig.2.6 (b). However, after k intervals of time Δt, it is found that none of the points in S1`k lie in the 1 “ 0. Additionally, the vertices of Conv(S1) form a single cluster, i.e. A “ ConvpS1q - see Fig.2.6 (c). The single cluster requirement ensures that the vertices of the convex hull are sufficiently close to one another and hence the NC(Conv(S1)) = 1. Therefore, Re interior of Conv(S1), i.e. m 1, x0 1`k boundary of the convex hull is a positively invariant set. 24 Conv(S1) 1 x 2 1 x 1 1 x N 1`k x 2 1`k x 3 1`k x 1 1`k x N Re A 0 x 1 0 x N 1 x 3 p0, 0q p0, 0q p0, 0q 0 x 2 0 x 3 0 x i Ω0 t “ t0 S0 “ tx 0 N u 0 1, ¨ ¨ ¨ , x (a) Conv(S1`k) t “ t1`k “ pt1 ` kΔtq 1`k N u 1`k S1`k “ tx 1 , ¨ ¨ ¨ , x t “ t1 “ pt0 ` Δtq N u S1 “ tx 1 1 1, ¨ ¨ ¨ , x (b) Figure 2.6: (a) Discretization of the boundary of conservative Re Convex hull Conv(S1) constructed at time t “ t1, (c) Enlarged Re t “ t1`k. (c) A “ Ω0 at t “ t0 yields S0, (b) A obtained as Conv(S1) at time 2.5 Enlarging the Region of Attraction Using SOS, CHART and IMM Re In the last section, we presented CHART for obtaining a larger Re A starting from a conservative A for the Tiptoebot starting from the conservative estimate Ωsos. It should be noted that the SOS method A. To improve the effectiveness of the IMM, we now apply CHART to obtain an Re optimizes the stabilizing controller to obtain the largest Re A and the CHART improves upon this estimate further without changing the controller. The CHART was implemented for the Tiptoebot equilibrium configurationpθ1,θ2,θ3, 9θ1, 9θ2, 9θ3q“ pπ{2, 0, 0, 0, 0, 0q using the following parameter values, which were selected by trial and error: N « 3ˆ 10 6, ε “ 0.01, Δt “ 0.0026, k1 “ 10, k2 “ 65 The enlarged estimate Re A, which is a convex hull in a six-dimensional space, was obtained as Ωchart “ ConvpS4q. 25 o f Ωchart A rRe o f Ωsos A rRe A of Ωlin pRe IM Figure 2.7: Slices of Ωchart, Ωsos and Ωlin, and their intersection with the impulse manifold, IM. For better visualization and for ease of comparison, the same coordinate system of Fig.2.3 is used. given by (2.3.1), the slice rRe We now reconsider the simulation in section 2.3.2. For the initial configuration of the Tiptoebot Apq˚ “ 0q of Ωchart is shown in Fig.2.7 together with the slices of Ωlin and Ωsos that were previously shown in Fig.2.3. A of Ωchart is significantly larger than that of Ωsos. The impulse manifold IMp0, 0, 0.21q described by (2.3.2), It can be seen that rRe A of both Ωchart and Ωsos in Fig.2.7. shown previously in Fig.2.3, can be seen to intersect rRe The pRe original coordinates. The pRe in the interior of pRe A of Ωchart is a convex region on the impulse manifold and is shown in Fig.2.8 in the A of Ωsos, shown previously in Fig.2.4 and now shown in Fig.2.8, lies A to A of A, we now choose the following velocity configuration that lies in the interior of pRe A of Ωchart. To take the configuration of the Tiptoebot from outside Re inside Re Ωchart: „ 9q1 9qT 2 “„1.57 ´2.69 0.18 (2.5.1) The desired effect of the impulsive inputs is shown by the dotted line joining the initial and desired velocity configurations, both of which lie on IM - see Fig.2.8. To implement the IMM, the high- 2 “ 9q2 in (2.5.1), µ “ 0.0004, λ1 “ 10 and λ2 “ 1; the gain feedback in (2.2.7) was used with 9qdes 26 A of Ωchart pRe 3.0 2.0 1.0 9q3 0.0 -1.0 A of Ωsos pRe p2.36, ´4.04, 0.16q p1.57, ´2.69, 0.18q 2.0 0.0 9q1 p0, 0, 0.21q IM 0.0 -2.0 9q2 -4.0 Figure 2.8: Intersection of the slices of Ωchart and Ωsos with the impulse manifold IM in the A of A of Ωchart is a convex region and encloses the pRe original coordinates; the intersection set pRe Ωsos. results are shown in Fig.2.9. The high-gain controller is active over t P r0, 0.004s; its effect is shown using a dilated time scale. The high-gain controller is switched off and the stabilizing controller is switched on at t “ 0.004 when the system configuration is inside Re equilibrium after t ą 0.004 is shown using a normal time scale. A. The stabilization of the The desired velocity configuration in (2.5.1) was chosen to lie on the line joining the initial velocity configuration p0.0, 0.0, 0.21q and the desired velocity configuration chosen for the SOS- IMM method in section 2.3, given by (2.3.3). It can be seen from Fig.2.8 that the initial velocity configuration is closer to the desired velocity configuration in (2.5.1) than the desired velocity configuration in (2.3.3). Consequently, the magnitudes of the high-gain feedback for the SOS- CHART-IMM method are less than that of the SOS-IMM method - see plots of τ2 and τ3 in Figs.2.5 and 2.9 for t P r0.00, 0.004s. We complete this section with a second example that further illustrates the benefit of using SOS-CHART-IMM for enlarging the Re A over SOS-IMM. In this example, the initial configuration 27 2.0 1.0 π{2 0.0 3.0 0.0 -4.0 0.0 -1.5 -3.0 0.0 -2.0 -4.0 9θ1 9θ2 τ2 τ3 θ2 θ1 9θ3 (rad) θ3 (rad/s) (Nm) (Nm) dilated time scale normal time scale 0.0 0.004 time (s) 1.5 3.0 Figure 2.9: Effective enlargement of Re A using SOS-CHART-IMM: Plot of joint angles, their ve- locities, and control inputs with time for stabilization of the Tiptoebot from the initial configuration given by (2.3.1). of the Tiptoebot is chosen to be x “„q˚ 9q˚T “„´0.2 0.3 0.3 0.0 0.0 0.0T (2.5.2) This initial configuration, shown in Fig.2.10 (a), lies outside Ωchart, and hence outside Ωsos. The Apq˚q of Ωsos is also a null set. Therefore, the IMM cannot be used to enlarge Ωsos. However, it can be seen from Fig.2.10 (b) rRe Apq˚q of Ωsos is found to be a null set and consequently pRe Apq˚q of Ωchart is not a null set and the IM intersects it. The pRe that rRe A of Ωchart is a convex region on IM and is shown in Fig.2.10 (c). To take the configuration of the Tiptoebot from outside Re A “ Ωchart to inside Ωchart, we choose the following velocity configuration that lies in the interior 28 A of Ωchart rRe 9q3 chart A of Ω pRe p0, 0, 0q p2.0, ´3.3, ´0.3q 9q1 IM IM (c) 9q2 (a) (b) A of Ωsos is a null set, (c) intersection of the slice of Ωchart with the impulse manifold IM, shown in the original coordinates. Figure 2.10: (a) Initial configuration of the Tiptoebot, given by (2.5.2), lies outside Ωchart, (b) rRe of Ωchart (shown in transformed coorrdinates for better visualization) is a convex hull whereas rRe of pRe A of Ωchart: „ 9q1 9qT 2 “„2.0 ´3.3 ´0.3 (2.5.3) A The desired effect of the impulsive inputs is shown by the dotted line joining the initial and desired velocity configurations on the IM - see Fig.2.10 (c). To avoid repetition, the simulation results of high-gain feedback followed by stabilization are not presented. 2.6 Experimental Verification of Enlargement of Re A using SOS, CHART and IMM 2.6.1 Hardware Description Experiments were done with a two-link Pendubot system. A schematic of the pendubot is shown in Fig.2.11. The equations of motion of the pendubot are in the form of (2.2.2) where n “ 2, m “ 1 and u “ τ (2.6.1) q “»—– q1 q2 fiffifl “»—– fiffifl , θ2 ´ π θ1 29 The equations of motion in terms of the variables q1 and q2 can be found in [23]; the lumped parameters of the pendubot that appear therein are given in Table 2.2. The torque τ in the shoulder joint (joint with angle q2 “ θ1) of the pendubot is applied by a 90-volt direct-drive permanent magnet brushed DC motor3 The motor is driven by a power amplifier4 with a maximum current limit of 25 A, operating in current mode. The motor torque constant is 0.4 Nm/A and the amplifier gain is 11.5 A/volt. The torque constant and the current limit together imply that the maximum torque that can be applied by the motor is 10 Nm. The pendubot was interfaced with a dSpace DS1104 board and the controller was implemented in the Matlab/Simulink environment for real-time control. τ d1 ℓ1 Y g θ1 c . m . m1, I1 X m2, I2 c.m. θ2 d2 ℓ2 Figure 2.11: An arbitrary configuration of the two-link pendubot. Table 2.2: Lumped Parameters for Pendubot in SI units β1 β2 β3 β4 β5 0.0147 0.0051 0.0046 0.1001 0.0302 3The motor manufacturer is Minnesota Electric Technology. 4The amplifier is a product of Advanced Motion Control. 30 2.6.2 Design of Experiment The equilibrium configuration chosen for stabilization was x “ pqT , 9qTqT “ 0 ñ pθ1,θ2, 9θ1, 9θ2q “ p0,π, 0, 0q. The SOS method was used to design a stabilizing second-order polynomial controller uspxq; the controller is given below 2 τ “´ 14.926θ1 ´ 1.951θ1 ´ 0.105θ1 ` 0.003 9θ1 ´ 0.766pθ2 ´ πq 9θ2 ´ 0.018 9θ1 2 ´ 17.728pθ2 ´ πq´ 2.514pθ2 ´ πq2 ´ 4.469θ1pθ2 ´ πq´ 1.224 9θ1 9θ1 ´ 0.121pθ2 ´ πq 9θ1 ´ 2.813 9θ2 ´ 0.055 9θ2 ´ 0.671θ1 9θ2 2 9θ2 (2.6.2) The corresponding estimated region of attraction, Ωsos in (2.2.10), was obtained as Vpxq ď ρ “ 0.083 where Vpq, 9qq “ 1.24 q2 2 ` 3.01 q1q2 ` 0.15 q2 9q2 ` 0.43 q2 9q1 ` 2.03 q1 2 ` 0.18 q1 9q2 ` 0.54 q1 9q1 ` 0.005 9q2 2 ` 0.026 9q1 9q2 ` 0.040 9q1 2 (2.6.3) Subsequently, the enlarged estimated region of attraction was obtained as Ωchart “ ConvpS8q using the following parameter values: N “ 729000, ε “ 0.01, Δt “ 0.0035, k1 “ 10, k2 “ 18 To demonstrate the enlargement of the Re A “ Ωchart using the IMM, we consider the following initial configuration that lies outside Ωchart: x “„q˚T 9q˚TT “„0.00 ´0.72 0.25 0.00T For the initial configuration of the pendubot given by (2.6.4), the slice rRe Apq˚ 1 “ 0.00, q˚ of Ωchart is shown in Fig.2.12 together with the slice of Ωsos. It can be seen that rRe 2 “ ´0.72q A of Ωchart is significantly larger than that of Ωsos. The impulse manifold IMp0.25, 0.00q, which is line, is given by the relation (2.6.4) 9q1 `p1´ β3{β2q 9q2 ´ 0.25 “ 0 (2.6.5) and is found to intersect rRe segment of the impulse manifold (shown in Fig.2.12 using a thicker line) whereas pRe A of Ωchart but not that of Ωsos. Therefore, pRe A of Ωchart is a line A of Ωsos is a 31 8 4 0 ) s / d a r ( 2 9θ “ 1 9q A of Ωchart rRe IM -4 -20 0.3 0.0 p2.00, 0.05q p0.00, 0.25q 0.0 1.0 2.0 A of Ωsos rRe 20 A ofΩ pRe chart 0 9q2 “ 9θ1 (rad/s) Figure 2.12: Intersection of rRe lie on pRe A of Ωchart is the dark line segment. The impulse manifold IM was determined by the initial velocity configuration p 9q2, 9q1q “ p0.00, 0.25q. The desired velocity configuration p 9q2, 9q1q “ p2.00, 0.05q was chosen to A of Ωchart; due to this scaling the initial and desired velocity configurations appear to be close. A of Ωchart with IM; the intersection set pRe A of Ωchart. A zoomed-out scale has been used to show the entire rRe null set. To take the configuration of the pendubot from outside Re A to inside Re A, we now choose the following velocity configuration that lies in the interior of pRe 9θ1 “„0.05 2.00 „ 9q1 9q2 “„ 9θ2 A of Ωchart: (2.6.6) Our experimental verification involved the following steps: 1. Feedback Linearization: A controller was designed using feedback linearization [73] to fix the configuration of the first link at pq2, 9q2q “ pθ1, 9θ1q “ p´0.72, 0.00q. With the first link fixed, the second link behaves as a simple pendulum. To achieve the configuration pq1, 9q1q “ pθ2 ´ π, 9θ2q “ p0.00, 0.25q, the second link was manually taken to an angle (at time t0) and released from rest; this angle was determined through trial and error. At time t´ 1 , the configuration in (2.6.4) is achieved - see Fig.2.13 (a), and the feedback linearizing controller was switched off. 2. High-Gain Control: To implement the IMM, the high-gain controller in (2.2.7) was switched on at time t´ 1 . For the high-gain controller, we used 9qdes λ1 “ 1. The high-gain controller was switched off at time t` a small neighborhood of (2.6.6) - see Fig.2.13 (b). 2 “ 9q2 in (2.6.6), µ “ 0.001, and 1 when the joint velocities reached 32 t0 Y X Y X Y t´ 1 rt´ 1 ,t` 1 s t` 1 t f (a) (b) (c) Figure 2.13: Steps of experimental verification using the pendubot shown in Fig.2.11: (a) rt0,t´ 1 q - feedback linearizing control, (b) rt´ 1 ,t fs - SOS stabilizing control 1 s - high-gain feedback, (c) pt` 1 ,t` 3. Stabilizing Control: The SOS controller in (2.6.2) was invoked at time t` 1 . The pendubot configuration reaches the desired equilibrium at time t f - see Fig.2.13 (c). 2.6.3 Experimental Results The experimental results are shown in Fig.2.14. Simulation results are presented in Fig.2.15 to compare the experimental results with the ideal behavior. The total time scale is divided into three intervals corresponding to the three steps discussed in the last section. In the experiment, the purpose of first step was to create the initial configuration of the pendubot in (2.6.4). In the simulation, the first step was not necessary as the pendubot was simply assumed to have this configuration. In the experiment, at time t0, the second link was released from θ2 “ ´2.28 rad « ´131 deg with the feedback linearizing controller active. At time t´ 1 , the feedback linearizing controller was switched off and the high-gain controller switched on since the configuration of the pendubot was close to the desired initial configuration in (2.6.4); the pendubot configuration was pθ1,θ2, 9θ1, 9θ2q “ p´0.72,π, 0.00, 0.42q - see Fig.2.14. For both the experiment and the simulation, the high gain controller was switched on at time t´ 1 . For better visualization, the effect of high-gain control is illustrated using a dilated time scale. Due to the initial error in the actual and desired velocities of the actuated joint, the high-gain controller invokes a large torque at time t´ 1 . Through our simulation, this torque was found to be 33 feedback linearization high-gain control -2.28 1.92 0.42 3.14 0 -2 4 0 -2 3 0 10 0 stabilizing control θ2 (rad) θ1 (rad) 9θ1 (rad/s) 9θ2 (rad/s) τ (Nm) t0 0.00 t´ 1 0.995 t` 1 1.03 time (s) t f 4.85 Figure 2.14: Experimental verification of enlargement of Re A of pendubot: Plot of joint angles, their velocities, and control input with time. The effect of high-gain feedback is shown using a dilated time scale; the effect of the stabilizing controller is shown using a normal time scale. 3.14 -0.72 1.92 0.42 0 -2 10 0 -5 2 0 -4 20 10 0 high-gain control stabilizing control θ2 (rad) θ1 (rad) 9θ1 (rad/s) 9θ2 (rad/s) τ (Nm) t´ 1 0.0 t` 1 0.003 time (s) t f 5.0 Figure 2.15: Simulation results of ideal pendubot behavior for comparison with experimental results presented in Fig.2.14. 34 of the order of 20 Nm. In the ideal case, the high-gain controller remains active for a very short interval of time and the control input decays exponentially during this interval [23]. This can be verified from the simulation where the time interval is equal to 0.003 s. In our experiment, due to the current limit (25 A) of the power amplifier (see section 2.6.1), the torque applied by the motor was saturated at 10 Nm. Therefore, the high-gain controller remained active for a longer duration of time 0.035 s and the control torque decays after remaining saturated for a significant fraction of that interval. The high-gain torque was switched off at time t` 1 when the joint velocities were close to the desired values in (2.6.6); these values are p 9θ1, 9θ2q “ p1.92, 0.00q for our experiment and p 9θ1, 9θ2q “ p1.96, 0.05q for our simulation. The changes in the joint angles for both the simulation and the experiment were negligible, as expected. Although small, the difference in the joint velocities between the experiment and the simulation at time t` 1 can be attributed to several factors. These include: (a) error between the actual and desired configuration at time t´ 1 , (b) saturation of the high-gain torque, and (c) sensitivity of the slope of the impulse manifold (line) IM to the system parameters - see (2.6.5). The SOS stabilizing controller was switched on at time t` 1 and the pendubot reached its equilibrium configuration shortly thereafter. The time required for stabilization and the transient behavior of the system are similar for both the experiment and the simulation. A video of the experiment can be found on the weblink: https://www.egr.msu.edu/~mukherji/RofAEnlargeStableUnstable.mp4 35 IMPULSIVE DYNAMICS AND CONTROL OF THE INERTIA-WHEEL PENDULUM CHAPTER 3 3.1 Introduction The problem of designing rest-to-rest maneuvers for the inertia-wheel pendulum using a se- quence of impulsive inputs is addressed ˙The use of impulsive inputs simplifies the dynamics of the IWP and permits the design of optimal sequences under some general assumptions1. This chapter is organized as follows: the mathematical model of the inertia-wheel pendulum and the formal problem statement is presented in section 3.2. In section 3.3, rest-to-rest maneuver is designed us- ing a sequence of two impulsive inputs and a generalization of this result which utilizes a sequence of N impulsive inputs is presented in section 3.4. In section 3.4, it is also proven that a sequence with an odd number of inputs is less optimal than the two adjacent sequences with even number of inputs. Finally simulation results are presented in section 3.5 which provides insights into high wheel velocity for the inertia-wheel pendulum obtained in prior works. y g c.m. τ ℓ2 ℓ1 φ motor θ pendulum inertia wheel x Figure 3.1: The inertia wheel pendulum is shown in an arbitrary configuration. 1The frequently used symbols in this chapter are defined earlier - see key to symbols before chapter 1. 36 3.2 Mathematical Model and Problem Statement 3.2.1 Equations of Motion The Inertia-Wheel Pendulum (IWP), shown in Fig.3.1, is an underactuated system comprised of a simple pendulum and an inertia wheel. The inertia wheel, also referred to as a reaction wheel, is driven by a motor mounted at the distal end of the pendulum. The kinetic energy T and the potential energy V of the IWP are given by the relations: 2 , 9θ2 ` m22p 9θ` 9φq 2 ` I1¯{2, 1 ` m2l 2 V “ β sinθ m22 “ I2{2 T “ m11 m11 “´m1l 2 β “ pm1l1 ` m2l2qg Using (3.2.1), the Euler-Lagrange equation of motion of the IWP can be written as: 2pm11 ` m22q :θ` 2 m22 :φ` β cosθ “ 0 :θ` 2 m22 :φ “ τ 2 m22 (3.2.1) (3.2.2a) (3.2.2b) Note that the angle φ does not appear in (3.2.2a) or (3.2.2b) - this is because the wheel is symmetric about its axis of rotation. 3.2.2 Effect of Impulsive Input and Constants of Free Motion The application of an ideal impulsive torque τi by the motor at time ti results in discontinuous jumps in the velocities of both the pendulum and the wheel while their angular positions remain unchanged. By integrating (3.2.2a) with respect to time over the infinitesimal interval rt´ can be shown [23] that these jumps satisfy the relation: i s, it i ,t` p 9θ` i ´ 9θ´ i q “ ´Cp 9φ` i ´ 9φ´ i q, C fi m22 pm11 ` m22q (3.2.3) 37 Using (3.2.2b), the jump in the pendulum velocity can be shown to be related to the angular impulse as follows: Ii fiż t` i t´ i τi dt “ 2 m22”p 9θ` i qı i ´ 9φ´ i ´ 9θ´ i q`p 9φ` ñ Ii “ ´2 m11p 9θ` i ´ 9θ´ i q where (3.2.4b) was obtained from (3.2.4a) using (3.2.3). (3.2.4a) (3.2.4b) If τ “ 0, the dynamics of the IWP described by (3.2.2a) or (3.2.2b) simplifies to the form 2 m11 :θ` β cosθ “ 0 :θ` :φ “ 0 (3.2.5a) (3.2.5b) The above equations are the differential forms for conservation of energy of the IWP and conser- vation of angular momentum of the wheel, namely m11 9θ2 ` β sinθ “ constant 9θ` 9φ “ constant (3.2.6a) (3.2.6b) It can be seen from (3.2.6a) and (3.2.6b) that the dynamics of the pendulum and wheel are decoupled and simplified when τ “ 0. For controlling the IWP, this motivates the use of impulsive torques at discrete instants of time. 3.2.3 Problem Statement: Rest-to-Rest Maneuvers We consider the problem of designing a set of discrete impulsive inputs over the time interval rt0,t fs for rest-to-rest maneuvers of the IWP between two configurations where the final configuration has higher potential energy than the initial configuration, i.e. 9θ0 “ 9φ0 “ 0 9θf “ 9φf “ 0 βpsinθf ´ sinθ0q ą 0 38 (3.2.7a) (3.2.7b) (3.2.7c) and θ lies in the range p´3π{2,π{2s. For a given value of N, the task is to design the vector I N “ rI1 I2 ¨¨¨ INs where signpIkq “ ´signpIk´1q, k “ 2, 3,¨¨¨ , N (3.2.8) and t0 ă t1 ă t2 ă ¨¨¨ ă tN ă t f . The rationale for imposing the constraint in (3.2.8) will be provided later, at the beginning of section 3.4. For the sake of simplicity, it is assumed that the time instants ti are such that the angular velocity of the pendulum is momentarily zero, i.e., 9θ´ i fi 9θpt´ i q “ 0, i “ 1, 2,¨¨¨ , N (3.2.9) Without loss of generality, it is assumed that the first impulsive torque is applied at time t1, where t´ 1 “ t0. Therefore, 9θ´ 1 “ 9θ0 “ 0, 9φ´ 1 “ 9φ0 “ 0 (3.2.10) Since the angular positions of the pendulum and the wheel do not change during application of an impulsive torque, we have θ1 “ θ0, φ1 “ φ0 (3.2.11) 3.3 Rest-to-Rest Maneuvers: Case of One and Two Impulsive Inputs 3.3.1 Case of One Impulsive Input (N “ 1) Using (3.2.3) and (3.2.10), we can write 9θ` 1 “ ´C 9φ` 1 (3.3.1) The main result can now be stated as follows: Result 1: (One Impulsive Input) The rest-to-rest maneuver described by (3.2.7) cannot be accom- plished using I 1 “ rI1s. 39 Discussion: Since τ “ 0 for rt` 1 ,t fs, we get from (3.2.6), and (3.2.11): 2 9 θ` 1 m11 2 ` β sinθ0 “ m11 9θ` 1 ` 9φ` 9θf 1 “ 9θf ` 9φf ` β sinθf (3.3.2a) (3.3.2b) Using (3.2.7b), (3.3.2b), and (3.3.1), it can be shown that 9θ` 1 “ 0. From (3.3.2a) we now get βpsinθf ´ sinθ0q “ 0, which violates (3.2.7c). This establishes Result 1 by contradiction. A single impulsive torque can always be chosen to impart sufficient angular momentum to the pendulum such that it reaches its desired configuration with zero angular velocity. However, this impulsive torque will also cause the wheel to have a nonzero angular velocity in the inertial reference frame. This implies that 9φ ­“ 0 when 9θ “ 0. 3.3.2 Case of Two Impulsive Inputs (N “ 2) It is assumed that the first impulse is applied at time t1 and therefore (3.3.1) is still valid. Two results are presented next. In the first result (Result 2), we relax the assumption in (3.2.9) and design a more general sequence of impulses that satisfies (3.2.7). We will show that the second result (Result 3), which is a special case of the first, automatically satisfies the assumption in (3.2.9). Result 2: (Two Impulsive Inputs) The rest-to-rest maneuver described by (3.2.7) can be accom- plished using I 2 “ rI1 I2s, where I2 “ ´I1. Discussion: From (3.2.4a) and (3.2.7a) we have and since (3.2.6b) holds good for rt` 1 ` 9φ` 1 q I1 “ 2 m22p 9θ` 1 ,t´ 2 s, we can write I1 “ 2 m22p 9θ` 1 ` 9φ` 1 q “ 2 m22p 9θ´ 2 ` 9φ´ 2 q (3.3.3) From (3.2.7b) we have p 9θf ` 9φfq “ 0 and since (3.2.6b) holds good for rt` p 9θ` 2 ` 9φ` 2 q “ 0. Using (3.2.4a), we can now write 2 ,t fs, we can write I2 “ ´2 m22p 9θ´ 2 ` 9φ´ 2 q (3.3.4) 40 It is clear from (3.3.3) and (3.3.4) that the conditions in (3.2.7b) require I2 “ ´I1. For rt` 1 ,t´ 2 s and rt` 2 ,t fs, the conservation laws in (3.2.6), together with (3.2.7b) and (3.2.11), give 2 9 θ` 1 m11 2 9 θ` 2 m11 ` β sinθ2 2 9 θ´ ` β sinθ0 “ m11 2 1 ` 9φ` 2 “ 9θ` 9θ´ 2 ` 9φ´ ` β sinθ2 “ β sinθf 9θ` 2 ` 9φ` 2 “ 9θf ` 9φf “ 0 1 (3.3.5a) (3.3.5b) (3.3.5c) (3.3.5d) For the second impulse, the relationship between the velocity jumps can be obtained from (3.2.3) as: 2 ´ 9θ´ p 9θ` 2 ´ 9φ´ 2 q 2 from (3.3.5d) and 9φ´ 2 “ ´ 9θ` 2 “ 9θ` the right-hand side of (3.3.6) and simplifying using (3.3.1), we get By substituting the relations 9φ` 2 q “ ´Cp 9φ` (3.3.6) 1 ` 9φ` 1 ´ 9θ´ 2 from (3.3.5b) into p 9θ` 2 ´ 9θ´ 2 q “ ´ 9θ` 1 (3.3.7) By combining (3.3.5a) and (3.3.5c) and substituting (3.3.7), we get β`sinθf ` sinθ0 ´ 2 sinθ2˘ “ m11„ 9 θ` 2 “ 2m11 9 θ´ 2 2 9 θ´ 2 9 θ` 1 ´ 2 2 ` 9 θ` 2 Substituting the expressions for 9 2 and 9 θ´ θ` 2 from (3.3.5a) and (3.3.5c) in the above equation and simplifying, we get | 9θ` 1 |“ 1 2d β m11 `sinθf ´ sinθ0˘ asinθf ´ sinθ2 From (3.2.7c) we know that psinθf ´ sinθ0q ą 0 and it can be seen from (3.3.5c) that psinθf ´ sinθ2q ą 0; therefore, the above equation is well-defined. For a given pair tθ0,θfu, (3.3.8) provides a functional relationship between the initial angular velocity 9θ` 1 (resulting from application of the first impulse I1) and the configuration θ2 where the second impulse (I2 “ ´I1) is applied. There (3.3.8) 41 Final Configuration Intermediate Configuration y θf θ0 x rt` 2 ,t f s θ2 rt` 1 ,t´ 2 s Initial Configuration Figure 3.2: An example showing the initial, intermediate, and final configuration of the IWP for the case with two impulsive inputs (N “ 2). are infinite solutions given by the pair t 9θ` 1 can be used to compute the impulses I1 and I2 (I2 “ ´I1), using (3.3.1), (3.3.3), and (3.3.4). An example showing the initial, intermediate, and final configuration of the IWP for the case with two impulsive 1 ,θ2u; for each solution, the value of 9θ` inputs is shown in Fig.3.2. The next result pertains to the particular solution that minimizes the magnitude of the impulses. Result 3: (Optimal Input I 2) The minimum magnitude of the impulsive inputs required for the rest-to-rest maneuver described by (3.2.7) is | I1|“| I2|“b2m11β`sinθf ´ sinθ0˘ Discussion: Using (3.3.1) and (3.3.3), I1 can be expressed as I1 “ 2 m22p 9θ` 1 ` 9φ` 1 q “ 2 m22p1´Cq 9θ` 1 (3.3.9) (3.3.10) Therefore, the magnitude of I1 can be minimized by minimizing the magnitude of 9θ` 1 . From 42 (3.3.5a) it can be seen that 2 9 θ` 1 m11 2 9 θ´ 2 “ m11 ` βpsinθ2 ´ sinθ0q 9θ´ 2 ­“ 0 but the minimum value of 9θ` If psinθ2 ´ sinθ0q ď 0, 1 is equal to zero. This implies psinθf ´ sinθ0q “ 0 from (3.3.8), which contradicts (3.2.7c). Since psinθ2 ´ sinθ0q must be positive, the minimum magnitude of 9θ` 2 “ 02; this magnitude is 1 can be obtained by choosing 9θ´ equal to 1 |“d β | 9θ` m11 psinθ2 ´ sinθ0q By equating (3.3.8) and (3.3.11), we get sinθ2 “ 1 2psinθ0 ` sinθfq (3.3.11) (3.3.12) where θ2 is the angle at which the second impulse is applied3. Substituting (3.3.12) into (3.3.10) and (3.3.11), and comparing (3.3.3) and (3.3.4) we get 1 | “d β | 9θ` 2m11`sinθf ´ sinθ0˘ ñ | I1 | “| I2 |“b2m11β`sinθf ´ sinθ0˘ (3.3.13a) (3.3.13b) This establishes Result 3. From Result 3 it can be seen that the time instant t2 is automatically known when the magnitudes of the individual impulses are minimized. This is different from Result 2, where the choice of t2 is not unique and each feasible choice of t2 (alternatively θ2) uniquely determines the magnitudes of the impulses. 2This choice automatically satisfies the assumption in (3.2.9). 3It is clear that (3.3.12) can have multiple solutions for θ2. The procedure for computing the correct solution will be discussed in section 3.4.2. Knowing the value of θ2, it will be possible to determine the time instant t2. 43 3.4 Rest-to-Rest Maneuvers: Generalization to N Inputs 3.4.1 Revisiting the Problem Statement We start this section with the result that justifies the rationale for imposing the constraint in (3.2.8). Result 4: (Sum of Two Consecutive Impulses) Consider two impulses Ik and Ik`1 applied at times tk and tk`1. The net effect of these two impulses in terms of change in the velocities of the pendulum and wheel over the interval rt´ ¯I at time tk over the interval rt´ k`1s can be achieved by a single impulse k s where k ,t` k ,t` Discussion: From (3.2.4a) we have ¯I “ Ik ` Ik`1 (3.4.1) k`1s, (3.2.6b) can be used to rewrite the above equations as follows Since τ “ 0 for rt` k ,t´ k ´ 9θ´ k`1 ´ 9θ´ Ik “ 2 m22”p 9θ` Ik`1 “ 2 m22”p 9θ` k qı k q`p 9φ` k ´ 9φ´ k`1qı k`1 ´ 9φ´ k`1q`p 9φ` k qı k`1 ´ 9φ´ k`1qı k`1 ´ 9φ´ k qı k`1 ´ 9φ´ The change in the velocities of the pendulum and wheel over the interval rt´ k q, respectively. To achieve the same change over rt´ k ,t` and p 9φ` k s, k qı k`1 ´ 9φ´ Ik “ 2 m22”p 9θ´ Ik`1 “ 2 m22”p 9θ` ñ Ik ` Ik`1 “ 2 m22”p 9θ` k q`p 9φ´ k`1q`p 9φ` k q`p 9φ` k`1 ´ 9θ´ k`1 ´ 9θ´ k`1 ´ 9θ´ k`1 ´ 9θ´ k q`p 9φ` k`1 ´ 9φ´ ¯I “ 2 m22”p 9θ` ñ ¯I “ Ik ` Ik`1 k`1s are p 9θ` k ,t` ¯I must satisfy k`1´ 9θ´ k q This establishes Result 4. Result 4 clearly indicates that two consecutive impulses of the same sign can be replaced by a single impulse of the same sign. This justifies the constraint imposed in our problem statement that consecutive impulses must have opposite sign. 44 An extension of Result 4 is now considered. For a rest-to-rest maneuver using two impulsive ¯I “ I1 ` I2 “ 0. This is true since (3.2.6b) and (3.2.7b) f q “ 0, and (3.2.7a) and (3.2.10) implies p 9θ´ k`1q “ p 9θ` k ` 1 q “ 0. This is consistent with Result 2, where it was shown that I2 “ ´I1. A inputs pN “ 2q, Result 4 implies that implies p 9θ` 2 ` 9φ` 9φ´ k q “ p 9θ´ generalization of this result in stated next. k`1 ` 9φ` 1 ` 9φ´ 2 q “ p 9θ` f ` 9φ` Result 5: (Zero Sum of Impulses) For a rest-to-rest maneuver involving N impulsive inputs, N ě 2, the following equation must hold. Nÿi“1 Ii “ 0 (3.4.2) Discussion: A sequence of N impulses, N ě 2, can be replaced by two impulses by applying Result 4 iteratively. For a rest-to-rest maneuver, the sum of these two impulses is zero - this follows from our discussion above. It is clear from the discussion above that both Result 4 and Result 5 are quite general and they do not require the assumption in (3.2.9) to be satisfied. With the motivation of investigating the minimum values of the magnitudes of the impulsive torques, we investigate rest-to-rest maneuver of the IWP with I 3 and I 4. As in the cases with I 1 and I 2, (3.3.1) holds good. 3.4.2 Rest-to-Rest Maneuvers with Even Number of Impulsive Inputs Theorem 1. (Optimality of Even Impulse Sequence) For a rest-to-rest maneuver of the IWP that satisfies (3.2.7) and (3.2.9) and uses 2n impulsive inputs, n “ 1, 2,¨¨¨ , I 2n 8 is minimized by the following choice of inputs: | Ii|“ @ i “ 1, 2,¨¨¨ , 2n (3.4.3) The angles where the impulsive inputs are applied satisfy the following relation 1 ?nb2m11β`sinθf ´ sinθ0˘, 2n  sinθf  sinθ0 `„ i´ 1 sinθi “„ 2n´ i` 1 2n 45 i “ 1, 2,¨¨¨ , 2n (3.4.4) Proof: We use induction to first prove (3.4.3). Assuming that (3.4.3) is satisfied for 2m impulsive inputs, i.e., n “ m. We express the magnitudes of the impulses for the case with p2m` 2q inputs using the relation | Ii|“ ki?mb2m11β`sinθf ´ sinθ0˘, i “ 1, 2,¨¨¨ ,p2m` 2q (3.4.5) where ki, i “ 1, 2,¨¨¨ ,p2m` 2q, are arbitrary positive numbers. Using (3.2.4b) and (3.2.9) we can show ki?md β | 9θ` i | “ 2m11`sinθf ´ sinθ0˘, i “ 1, 2,¨¨¨ ,p2m` 2q (3.4.6) Using (3.2.9), the conservation law in (3.2.6a) for the time intervals rt` ¨¨¨ ,rt` 2m`2,t fs can be written as: 1 ,t´ 2 s, rt` 2 ,t´ 3 s, ¨¨¨rt` j ,t´ j`1s, 2 2 9 θ` 1 9 θ` 2 m11 m11 m11 2 9 θ` j ` β sinθ0 “ β sinθ2 ` β sinθ2 “ β sinθ3 ... ` β sinθj “ β sinθj`1 ... m11 9θ`2 2m`2 ` β sinθ2m`2 “ β sinθf (3.4.7) Addition of equations in (3.4.7) and substitution of | 9θ` gives the following i | from (3.4.6) in the resulting equation i “ rk1 k2 ¨¨¨ k2m`2s2 Using (3.4.8) and the property of norms it can be shown that k 2 2 “ 2m 2m`2ÿi“1 (3.4.8) (3.4.9) ?2m` 2 rk1 k2 ¨¨¨ k2m`2s8 ě rk1 k2 ¨¨¨ k2m`2s2 ñ rk1 k2 ¨¨¨ k2m`2s8 ěc m m` 1 It can be shown that ki “am{pm` 1q, i “ 1, 2,¨¨¨ ,p2m` 2q, satisfy (3.4.8) and minimize rk1 k2 ¨¨¨ k2m`2s8. Substitution of these values of ki in (3.4.5) shows that (3.4.3) is satisfied for p2m` 2q impulsive inputs, i.e. n “ pm` 1q. It has been shown earlier in (3.3.9) that (3.4.3) is 46 | 9θ` i | “ 1 ?nd β 2m11`sinθf ´ sinθ0˘ (3.4.10) satisfied for N “ 2 (n “ 1). By induction we can now claim that (3.4.3) will be satisfied for any even number of impulsive inputs. The values of | 9θ` i |, i “ 1, 2,¨¨¨ , 2n, can be obtained from (3.4.3), (3.2.4b), and (3.2.9), namely By substituting (3.4.10) in the energy conservation laws (similar to (3.4.7)) for 2n impulses and solving sequentially as shown below, we get the relations between the angles where the impulsive inputs should be applied S1 “ S0 S2 “ 1 S3 “ ... Si “ 2n 2n pS f ´ S0q` S0 “„ 2n´ 2` 1 2n pS f ´ S0q` S2 “„ 2n´ 3` 1 2n pS f ´ S0q` Si´1 “„ 2n´ i` 1  S0 `„ 2´ 1 2n  S f  S0 `„ 3´ 1 2n  S f  S0 `„ i´ 1 2n  S f 2n 2n 1 1 (3.4.11) where Sp “ sinθp, p “ 1, 2,¨¨¨ f . It can be seen that these angles satisfy 3.4.4. It follows from (3.4.3) in Theorem 1 and (3.2.8) that all the 2n impulses have the same magnitude and consecutive impulses have opposite signs. Since the pendulum and wheel are both at rest at the initial time, it follows that each pair of consecutive impulses (starting with I1 and I2) result in a rest-to-rest maneuver. It follows from (3.2.4b), (3.2.8), and (3.2.9) that the velocity of the pendulum immediately k`1q “ ´signp 9θ` after application of an impulsive input will have opposite sign for two consecutive impulses, i.e., signp 9θ` constructed to determine the unique value of θi, i “ 1, 2,¨¨¨ , 2n, from (3.4.4). If θf belongs to quadrant I (includes θ “ π{2) or IV, then k q, k “ 1, 2,¨¨¨ ,p2n´ 1q. Using this fact, the following algorithm can be For m “ 1, 2,¨¨¨ , n, compute sinθ2m using (3.4.4) 47 If sinθ2m ą 0, θ2m belongs to quadrant II Elseif sinθm ă 0, θ2m belongs to quadrant III Else θ2m “ ´π. For m “ 0, 1,¨¨¨ ,pn´ 1q, compute sinθ2m`1 using (3.4.4) If sinθ2m`1 ą 0, θ2m`1 belongs to quadrant I If sinθ2m`1 ă 0, θ2m`1 belongs to quadrant IV Else θ2m`1 “ 0. Else θf belongs to quadrant II or III, then For m “ 1, 2,¨¨¨ , n, compute sinθ2m using (3.4.4) If sinθ2m ą 0, θ2m belongs to quadrant I Elseif sinθm ă 0, θ2m belongs to quadrant IV Else θm “ 0. For m “ 0, 1,¨¨¨ ,pn´ 1q, compute sinθ2m`1 using (3.4.4) If sinθ2m`1 ą 0, θ2m`1 belongs to quadrant II If sinθ2m`1 ă 0, θ2m`1 belongs to quadrant III Else θ2m`1 “ ´π. Endif 3.4.3 Rest-to-Rest Maneuvers with Odd Number of Inputs We generalize the result presented in Remark 4. Theorem 2. (Lack of Optimality of Odd Impulse Sequence) It is not possible to design an odd impulse sequence for which the magnitudes of all the impulsive inputs are less than the optimal magnitude for the preceding and succeeding even impulse sequence. In other words, the following inequality holds for n “ 1, 2,¨¨¨ . I 2n`1 8 ą min I 2n 8 ą min I 2n`2 8 (3.4.12) 48 Proof: It is clear from (3.4.3) that min I 2n 8 ą min I 2n`2 8. We proceed to prove the left inequality by contradiction and therefore assume | Ii| “ ki min I 2n ki?nd β | “ ñ | 9θ` i 8 | ki P p0, 1s 2m11`sinθf ´ sinθ0˘ From (3.2.8), (3.4.2), and (3.4.13) we can show k1 ` k3 `¨¨¨ k2n`1 “ k2 ` k4 `¨¨¨ k2n Substituting the expression for 9θ` i for the p2n` 1q impulses, we get from (3.4.13) in the energy conservation laws (similar to (3.4.7)) k 2 1 ` k 2 2 `¨¨¨ k 2 2n`1 “ 2n Since ki P p0, 1s, the following inequality holds true k 2 1 ` k 2 2 `¨¨¨ k 2 2n`1 ď k1 ` k2 ` k3¨¨¨ k2n`1 Using (3.4.14), and (3.4.15) we get (3.4.13) (3.4.14) (3.4.15) (3.4.16) (3.4.17) ě n k2 ` k4 `¨¨¨ k2n l jh n n terms The left hand side of the inequality above is equal its right hand side @ kq “ 1, q “ 2, 4,¨¨¨ 2n. However, this choice of kq violates (3.4.14) and (3.4.15) to be satisfied simultaneously. Therefore, (3.4.17) can be modified as ą n (3.4.18) k2 ` k4 `¨¨¨ k2n l jh n n terms There exist no choice of kq such that the above inequality holds true @kq P p0, 1s and thus, completes the proof by contradiction. 49 3.5 The Swing-Up Problem 3.5.1 Optimal Swing-Up Using Even Impulse Sequences The swing-up problem is a rest-to-rest maneuver where the final pendulum angle is θf “ π{2. We consider special case where the initial angle of the pendulum is in the vertically downward configuration, i.e., θ0 “ ´π{2. For swing-up using an even number of impulsive inputs, the optimal solution can be obtained from (3.4.3). For N “ 2n, n “ 1, 2,¨¨¨ , the optimal solution is given by a sequence of equal and opposite impulses of the following magnitude: (3.5.1) Since the magnitude of the impulses is inversely proportional to ?n, it is clear that if the number of @ i “ 1, 2,¨¨¨ , 2n, | Ii|“ ?n I, I fi 2am11β 1 impulsive inputs are increased by an even number, the magnitude of each impulse in the sequence is reduced. This information will be useful for designing an impulse sequence that takes into consideration actuator saturation. 3.5.2 Implementation Using High-Gain Feedback Ideal impulsive inputs are Dirac-delta functions and cannot be generated by actuators. In real physical systems, continuous-time implementation of impulsive inputs has be achieved using high- gain feedback in both theory and experiments [19,23,25]. The high-gain feedback [93] for the IWP can be obtained as τhg “ rKT M´1 Ks´1rKT M´1 pm11 ` m22q m22 m22 m22 M “ 2»—– i ´ 9φ´ i qs β cosθ 0 H ` 1 εp 9φ` fiffifl , H “»—– fiffifl (3.5.2) where the matrices M and H above were reconstructed from (3.2.2), K fi„0 1T , and ε ą 0 is a small number. Implementation of impulsive inputs using high-gain feedback also enables us to compare our results with those published in the literature. A discussion of select results in the literature is presented next. 50 3.5.3 Discussion of Results in the Literature 3.5.3.1 Globally Stabilizing Controller We implemented the globally stabilizing controller in [1] using their kinematic and dynamic parameter values, which are given below m11 “ 4.83ˆ 10´3, m22 “ 32ˆ 10´6, β “ 37.9ˆ 10´2 (3.5.3) Due to brevity of space, we omit complete expression of the controller used in [1]; rather, we recall the controller’s structure which is in the form c0 c1 µ0 ` c1 µ1 ` c2 µ2 ` c3 µ3 where µ0, µ1, µ2 and µ3 are nonlinear functions of state variables. The controller parameters were chosen as: c0 “ ´π{10, c1 “ 13, c2 “ 16 and c3 “ 8.0; the results are shown in Fig.3.3. The plots are slightly different from those presented in [1] but the overall trends are similar. It is clear from the plot of θ that the control input drives the pendulum directly towards the desired value of θf “ π{2. There is a small overshoot beyond π{2 but the wheel velocity is extremely high, of the 2.0 0.0 -2.0 0.0 0 -8000 0.0 π{2 θ (rad) 10 0 -5 0.0 10.0 0.2 0.0 9θ (rad/s) 10.0 9φ (rad/s) τ (Nm) time (s) -0.4 0.0 10.0 time (s) 10.0 Figure 3.3: Simulation results for the globally stabilizing controller [1] with controller parameter values: c0 “ ´π{10, c1 “ 13, c2 “ 16 and c3 “ 8.0. 51 2.0 8 -2.0 0.0 0 θ (rad) 9φ (rad/s) 0 0.0 10.0 0.2 0.0 -5000 0.0 time (s) -0.4 10.0 0.0 9θ (rad/s) 1.0 10.0 τ (Nm) time (s) 10.0 Figure 3.4: Simulation results for the globally stabilizing controller [1] with controller parameter values: c0 “ ´π{10, c1 “ 13, c2 “ 16 and c3 “ 4.5. order of 8000 rad/s. A change in the controller parameter c3 from 8.0 to 4.5 increases the overshoot slightly but reduces the maximum wheel velocity by almost 50% - see Figs.3.3 and 3.4. The above observation can be explained by the analysis presented in section 3.3 even though the nature of the inputs are completely different (continuous inputs in [1] vs a pair of impulsive inputs, N “ 2). By changing the domain of θ from p´3π{2,π{2s to p´π{2, 3π{2s4 and using (3.3.1) and (3.3.8), we get for θf “ π{2 and θ0 “ ´π{2: 1 Cd | 9φ` 1 |“ β m11p1´ sinθ2q (3.5.4) It is clear from (3.5.4) that the wheel velocity immediately after application of the first impulse depends only on the angle where the second impulse is applied, namely θ2, and tends to infinity when θ2 “ π{2`, i.e., when the overshoot approaches zero. While it is clear from (3.5.4) that θ2 “ π{2` is not a good choice for application of the second impulse, the value of θ2 that minimizes the magnitude of the wheel velocity | 9φ` 1 | can be obtained using the energy conservation law in 4This change in the domain is necessary to ensure that the trajectory of θ is similar to that in [1] but it does not change the analysis whatsoever. 52 (3.2.6a). For the IWP to cross the upright configuration, the following inequality must be satisfied: 2 9 θ` 1 m11 ` β sinθ0 ą β ñ | 9φ` 1 |ą 1 Ca2β{m11 (3.5.5) where θ0 “ ´π{2 and (3.3.1) were used. Comparing (3.5.4) and (3.5.5), we can show that θ2 “ p5π{6q´ minimizes | 9φ` 1 |. A simulation was performed using the high-gain feedback law in (3.5.2) with ε “ 0.01, the parameter values in (3.5.3), and θ2 « 3π{4 (slightly less that 5π{6); the results are shown in Fig.3.5. After the IWP reached a neighborhood of θf “ π{2, a linear controller was invoked for stabilization. The linear controller was designed to place the poles of the closed loop system at ´4˘ 2i and ´8. The simulation results indicate that swing-up is achieved in less than 1.0 s, which is much faster than that achieved in [1]. The maximum velocity of the wheel is still quite high (3000 rad/s) but it is significantly lower than that in Fig.3.3. The torque required is quite high (« 13 Nm) but this can be reduced significantly by simply changing the domain of θ2, as we will show in the next simulation. The simulation results presented in Fig.3.5 were obtained by assuming θ P p´π{2, 3π{2s; this was motivated by the need to generate trajectories of the IWP similar to those generated in [1], for 2.0 0.0 -2.0 0.0 0 θ (rad) 0.5 1.0 -3000 0.0 0.5 9φ (rad/s) 1.0 time (s) 20 0 -10 1.5 0.0 10 0 -10 9θ (rad/s) 0.5 1.0 1.5 τ (Nm) 1.5 0.0 0.5 1.0 time (s) 1.5 Figure 3.5: High-gain feedback implementation of two impulsive inputs (N “ 2) for swing-up of the IWP. For the purpose of comparison with the results in Figs.3.3 and 3.4, the controller is designed to keep θ in the domain p´π{2, 3π{2s. 53 comparison. If we switch the domain of θ back to p´3π{2,π{2s, as defined in section 3.2.3, the maximum wheel speed and the magnitude of the maximum torque can both be reduced from their values in Fig.3.5. Simulation results of high-gain feedback implementation of the optimal impulse sequence based on two inputs, described by (3.4.3) and (3.4.4) with n “ 1, is shown in Fig.3.6. The high-gain controller was implemented using ε “ 0.02 and stabilization of the equilibrium was achieved by the same linear controller that was used in the last simulation. It can be seen from Fig.3.6 that the second impulse is applied when θ2 « ´π rad. Similar to the results in Fig.3.5, swing-up is achieved in less than 1.0 s, but the maximum wheel velocity is now reduced from 3000 rad/s to 2000 rad/s and magnitude of the maximum torque is reduced from « 13 Nm to « 3 Nm. The maximum torque of « 3 Nm in Fig.3.6, although larger than those reported in the literature, is not a significant concern because it is applied for a very short duration of time. Motors can apply substantially larger torques5 than their maximum continuous torque over short time intervals. The maximum torque of « 3 Nm also corresponds to the continuous-time implementation of the optimal I 2. The magnitude of this torque, as well as the maximum velocity of the wheel, can π{2 0.0 θ (rad) 20 0 9θ (rad/s) -3.0 0.0 2000 0 0.0 1.0 -15 2.0 0.0 1.0 2.0 9φ (rad/s) 3 0 -3 τ (Nm) 1.0 time (s) 2.0 0.0 1.0 time (s) 2.0 Figure 3.6: High-gain feedback implementation of the sequence of two optimal impulsive inputs (N “ 2) for swing-up of the IWP. The controller is designed to keep θ in the domain p´3π{2,π{2s. 5This is referred to a peak torque [94]; for different motors, the peak torque can be twice to ten times larger than the maximum continuous torque. 54 be easily reduced if we consider continuous-time implementation of I 2n, n “ 2, 3,¨¨¨ . This is discussed in the next section. Our results are not compared with those obtained using the IDA-PBC method, [15, 43], for example, due to space constraints. Similar to the globally stabilizing controller [1], the IDA-PBC method also takes the pendulum directly to the desired upright configuration and results in a large continuous torque [15] or large wheel velocity [43] during swing-up. 3.5.3.2 Energy Based Controller When the number of impulses are increased from N “ 2 to N “ 8, for example, the magnitude of the impulsive torques are reduced by a factor of ?4 “ 2; consequently, the magnitude of the maximum high-gain torque and the wheel velocity are reduced proportionately - see Fig.3.7. The trajectories of the state variables in Fig.3.7 resemble those of the energy based controllers [2, 68] during swing-up phase of the IWP; the PFBLC + AL energy-based controller presented in [2] is simulated here to show the similarities in the trajectories - see Fig.3.8. It can be seen from Figs.3.7 and 3.8, that, unlike the globally stabilizing controller [1] (see Fig.3.3) where the pendulum is aggressively driven towards its desired configuration, both controllers (presented here and in [2]) 0.0 π{2 θ (rad) -4.0 0.0 950 0 0.0 20 0 -20 4.5 0.0 1.5 0 -1.5 9θ (rad/s) 4.5 τ (Nm) 9φ (rad/s) time (s) 4.5 0.0 time (s) 4.5 Figure 3.7: High-gain feedback implementation of the sequence of eight optimal impulsive inputs (N “ 8) for swing-up of the IWP. The high-gain controller was implemented with ε “ 0.02. 55 0.0 π{2 θ (rad) -4.0 0.0 600 0 -700 0.0 20 0 -20 5.0 0.0 0.15 0 -0.15 5.0 0.0 9φ (rad/s) time (s) 9θ (rad/s) 5.0 τ (Nm) time (s) 5.0 Figure 3.8: Simulation using the PFBLC + AL controller in [2]; the controller parameters were chosen as ke “ 3.1ˆ 107 and kv “ 0.1. This choice of parameters ensured that the time required for swing-up and the magnitude of the maximum control torque in simulations matched those of the experiments. The initial configuration was chosen to be slightly different from θ0 “ ´π{2 since the controller is unable to swing-up from this configuration. gradually add energy to the pendulum over several cycles of oscillation. Table 3.1: Swing-up time and maximum magnitude of high-gain torque required for different values of N N ts (s) max t0ďtďt f |τhg| (Nm) 2 4 6 8 10 12 0.6 1.36 1.96 2.90 3.67 4.53 3.0 2.12 1.73 1.50 1.35 1.20 A comparison of Figs.3.7 and 3.8 indicates that the magnitude of the maximum torque required by our method is larger than that required by the approach proposed in [2]. However, since the torques are applied intermittently over very short intervals of time, feasibility of our approach is determined by the peak torque rating of the actuator as opposed to the maximum continuous torque rating, which is always lower [94]. A salient feature of our approach is that, given any actuator, an optimal impulse sequence can be designed such that the peak torque rating of the motor is not exceeded. A higher value of N reduces the peak torque requirement of the motor but increases the 56 time time required for swing-up. The swing-up time and the magnitude of the maximum torque for several different values of N are presented in Table 6.1. 57 IMPULSIVE CONTROL OF A DEVIL-STICK: PLANAR SYMMETRIC JUGGLING CHAPTER 4 4.1 Introduction The problem of juggling a devil-stick is investigated. Assuming that the stick remains confined to the vertical plane, the task is to juggle between two symmetric configurations. Impulsive forces are applied to the stick intermittently and the impulse of the force and its point of application are modeled as inputs to the system. The juggling problem is formally described in section 4.2. The dynamics of the devil-stick is presented in section 4.3; it is comprised of impulsive dynamics due to the control inputs and continuous dynamics due to torque-free motion under gravity. A coordinate transformation is used to simplify the control problem and the dynamics is described by a nonlinear discrete-time system. The control design is provided in section 4.4. By choosing one of the control inputs to be dead-beat, the nonlinear system is simplified to a linear discrete-time system. For stable juggling, the linear system is controlled using linear quadratic regulator (LQR) and model predictive control (MPC) techniques. Simulation results are presented in section 4.5. y g m, J G ” phx, hyq ℓ θ x Figure 4.1: A three degree-of-freedom of a devil-stick. 58 4.2 Problem Description Consider the three degree-of-freedom devil-stick shown in Fig. 4.1, which can move freely in the xy vertical plane. The stick has length ℓ, mass m, and mass moment of inertia J about its center-of-mass G. The configuration of the stick is described by the three generalized coordinates: pθ, hx, hyq, where θ is the orientation of the stick with respect to the positive x axis, measured counter-clockwise, and phx, hyq are the Cartesian coordinates of G. The objective is to juggle the stick between two configurations that are symmetric with respect to the vertical axis. The coordinates of the stick in these two configurations are pθ˚, h˚ yq, where θ˚ P p0,π{2q - see Fig. 4.2. It is assumed that juggling is achieved by applying impulsive forces perpendicular to the stick; they are applied only when the orientation of the stick is θ “ θ˚ or θ “ π´ θ˚. Therefore, the time of application of the impulsive force is not a part of the control design. The control inputs are the pair pI, rq, where I, I ě 0, is the impulse of the impulsive force and r is the distance of the point of application of the force from G. The value of r is considered to yq and pπ´ θ˚,´h˚ x, h˚ x, h˚ be positive if the angular impulse of the impulsive force about G is in the positive z direction when θ “ θ˚, and is in the negative z direction when θ “ π´θ˚. The control inputs that juggle the stick between the symmetric configurations are denoted by the pair pI˚, r˚q. y g I “ I˚ r “ r˚ p´h˚ x , h˚ yq π´ θ˚ ˚ r “ r ph˚ x , h˚ yq θ˚ I “ I˚ x Figure 4.2: Symmetric configurations of the devil-stick in Fig. 4.1. 59 4.3 Dynamics of the Devil-Stick 4.3.1 Impulsive Dynamics The dynamics of the three-DOF devil-stick is described by the six-dimensional state vector X, where X “„θ ω hx vx hy vyT , ω fi 9θ, vx fi 9hx, vy fi 9hy Let tk, k “ 1, 2, 3,¨¨¨ , denote the instants of time when the impulsive inputs are applied. Further- more, without loss of generality, let k “ p2n´ 1q, n “ 1, 2,¨¨¨ denote the instants of time when the impulsive inputs are applied at θ “ θ˚, and k “ 2n, n “ 1, 2,¨¨¨ denote the instants of time when the impulsive inputs are applied at θ “ π´ θ˚. If t´ k denote the instants of time immediately before and after application of the impulsive inputs, the linear and angular impulse-momentum k and t` relationships can be used to describe the impulsive dynamics1 as follows, for k “ 1, 3, 5,¨¨¨ Xpt` k q “ Xpt´ k q` and for k “ 2, 4, 6,¨¨¨ Xpt` k q “ Xpt´ k q` (4.3.1) (4.3.2) 0 pIk rk{Jq 0 ´pIk{mq sinθ˚ 0 pIk{mq cosθ˚ 0 ´pIk rk{Jq 0 pIk{mq sinθ˚ 0 pIk{mq cosθ˚ fiffiffiffiffiffiffiffiffiffiffiffiffiffiffifl fiffiffiffiffiffiffiffiffiffiffiffiffiffiffifl »——————————————– »——————————————– 60 1Impulsive inputs cause discontinuous jumps in the velocity coordinates but no change in the position coordinates. The dynamics of underactuated systems subjected to impulsive inputs is discussed in [6, 25, 74, 95]. where pIk, rkq denote the control inputs at time tk. Between two consecutive impulsive inputs, the devil-stick undergoes torque-free motion under gravity; this is discussed next. 4.3.2 Continuous-time Dynamics Over the interval t P rt` k`1s, the devil-stick will be in flight; its center-of-mass G will undergo projectile motion and its angular momentum will remain conserved. This dynamics is described k ,t´ by the differential equation: where the initial condition Xpt` k is odd or even. 9X “„ω 0 vx 0 vy ´gT (4.3.3) k q can be obtained from ( 4.3.1) or ( 4.3.2), depending on whether 4.3.3 Poincaré Sections and Half-Return Maps For the hybrid system, described by impulsive dynamics of section 4.3.1 and continuous dynamics of section 4.3.2, we define two Poincaré sections2, 3 [96] Sr and Sl as follows: Sr : tX P R Sl : tX P R 6 | θ “ θ˚u 6 | θ “ π´ θ˚u (4.3.4) These Poincaré sections are chosen since the impulsive inputs are applied only when θ is equal to θ˚ or pπ´ θ˚q. Any point on Sr and Sl can be described by the vector Y , Y Ă X, where Y “„ω hx vx hy vyT (4.3.5) (4.3.6) The map Pr : Sr Ñ Sl can be determined from ( 4.3.1) and ( 4.3.3) as follows: Ypt´ k`1q “ AYpt´ k q` Br 2Poincaré sections have been previously used for design of gaits for bipedal robots [76, 80]. 3It is assumed that the initial conditions of the devil-stick are such that its trajectory intersects one of the two Poincaré sections before the first impulsive control input is applied. 61 A fi 0 0 0 , Br fi »——————————– 1 0 0 0 0 1 δk 0 1 0 0 0 0 0 0 1 δk 0 1 0 0 0 fiffiffiffiffiffiffiffiffiffiffifl »——————————– pIk rk{Jq ´pIk{mq sinθ˚δk ´pIk{mq sinθ˚ pIk{mq cosθ˚δk´p1{2qgδ2 k pIk{mq cosθ˚´gδk fiffiffiffiffiffiffiffiffiffiffifl where δk fi pt´ k`1 ´ t´ k q and k “ p2n´ 1q, n “ 1, 2,¨¨¨ . Similarly, the map Pl : Sl Ñ Sr can be determined from ( 4.3.2) and ( 4.3.3) as follows Ypt´ k`1q “ AYpt´ k q` Bl (4.3.7) fi Bl »——————————– ´pIk rk{Jq pIk{mq sinθ˚δk pIk{mq sinθ˚ pIk{mq cosθ˚δk´p1{2qgδ2 k pIk{mq cosθ˚´gδk fiffiffiffiffiffiffiffiffiffiffifl where k “ 2n, n “ 1, 2,¨¨¨ . Both Pr and Pl in ( 4.3.6) and ( 4.3.7), respectively, can be viewed as half-return maps4 since the composition of these maps are the return maps Pr ˝ Pl : Sl Ñ Sl and Pl ˝ Pr : Sr Ñ Sr. In the next section we introduce a coordinate transformation to show that the map Pl, in the transformed coordinates, is identical to Pr. This simplifies the analysis of the problem. 4.3.4 Coordinate Transformation Consider Fig. 4.3, where z “ 0 denotes the xy plane in which the devil-stick is juggled. Typically, the juggler will stand at a point on the positive z axis, denoted by P in Fig. 4.3 (a), and face the z “ 0 plane. The juggler will apply a control action with the right hand when θ “ θ˚, and with the left hand when θ “ π´ θ˚, i.e., the juggler is ambidextrous. Instead of alternating between the right and left hands, the juggler can choose to apply all control actions using the right hand only. 4Half-return maps have been used to analyze the behavior of dynamical systems such as the van der Pol oscillator [97, 98]. 62 (a) y g p´h˚ x,h˚ yq π´θ˚ ph˚ x,h˚ yq θ˚ Q x y z ph˚ x,h˚ yq θ˚ (b) x (c) z ph˚ x,h˚ yq θ˚ x P y Q P z Figure 4.3: (a) Ambidexterous juggler standing at P and applying control actions with both hands, (b) right-handed juggler standing at P and applying control action with right hand, (c) right-handed juggler standing at Q and applying control action with right hand. This juggler, whom we will now refer to as the right-handed juggler, To this end, the juggler will apply the control action standing at P when θ “ θ˚ - see Fig. 4.3 (b), and apply the next control action after changing location to Q (mirror image of P) and facing the z “ 0 plane when θ “ π´θ˚ - see Fig. 4.3 (c). When the devil stick has the orientation θ “ π´ θ˚, as seen by an observer at P, it will have the orientation θ “ θ˚ for the right-handed juggler. After applying control action at Q, the right-handed juggler will return back to P. If xyz denotes the rotating coordinate frame of the right-handed juggler, the change in position of this juggler can be described by the coordinate 63 transformation: where »————– x y z fiffiffiffiffiflQ “ Ry,π »————– x y z fiffiffiffiffiflP , »————– x y z fiffiffiffiffiflP “ Ry,π »————– x y z fiffiffiffiffiflQ Ry,π fi diagr´1, 1, ´1s Since Ry,π changes the sign of the x and z coordinates and leaves the y coordinate unchanged, we can show where YQ “ RYP, YP “ RYQ R “ R´1 fi diagr´1, ´1, ´1, 1, 1s and YP and YQ denote the vector Y as seen by the right-handed juggler standing at points P and Q, respectively. 4.3.5 Single Return Map and Discrete-Time Model In the reference frame of the right-handed juggler, who alternates between positions P and Q, the two Poincaré sections Sl and Sr are identical, and equal to S : tX P R 6 | θ “ θ˚u (4.3.8) This follows from our discussion in section 4.3.4 as well as Figs. 4.3 (b) and (c). The half-return maps Pr and Pl in ( 4.3.6) and ( 4.3.7) can be rewritten as follows: YPpt´ YPpt´ k`1q “ AYPpt´ k`1q “ AYPpt´ k q` Br, k q` Bl, k “ 1, 3, 5,¨¨¨ k “ 2, 4, 6,¨¨¨ (4.3.9a) (4.3.9b) 64 to explicitly indicate the reference frame of Y . Since the right-handed juggler alternates between positions P and Q, the half-return map Pl in ( 4.3.9b) can be transformed as follows: RYPpt´ ñ YQpt´ ñ YQpt´ k`1q “ RAYPpt´ k`1q “ ARYPpt´ k`1q “ AYQpt´ k q` RBl k q` RBl k q` Br, k “ 2, 4, 6,¨¨¨ (4.3.10) where we used the relations RA “ AR and RBl “ Br. It is clear from ( 4.3.9a) and ( 4.3.10) that the half-return maps Pr and Pl in ( 4.3.6) and ( 4.3.7) are identical in the reference frame of the right-handed juggler. This implies that the hybrid dynamics of the devil-stick between any two control actions can be described by a single return map if the change in reference frame of the right-handed juggler is incorporated in the dynamic model. This map, P : S Ñ S, can be obtained by first rewriting the left-hand-sides of ( 4.3.6), ( 4.3.9a) and ( 4.3.10) as follows: k q` Br, RYQpt´ ñ YQpt´ RYPpt´ ñ YPpt´ k`1q “ AYPpt´ k`1q “ R“AYPpt´ k`1q “ AYQpt´ k`1q “ R“AYQpt´ k “ 1, 3, 5,¨¨¨ k “ 1, 3, 5,¨¨¨ k “ 2, 4, 6,¨¨¨ k “ 2, 4, 6,¨¨¨ (4.3.11a) (4.3.11b) k q` Br, k q` Br‰ , k q` Br‰ , Then, by accounting for the change in reference frame of the right-handed juggler after each control action, ( 4.3.11a) and ( 4.3.11b) can be combined into the following single equation which represents the return map P: k`1q “ R“A ¯Ypt´ ¯Ypt´ k q` Br‰ , k “ 1, 2, 3,¨¨¨ 65 where ¯Y denotes the state vector Y in the reference frame of the right-handed juggler. The above equation results in the following discrete-time equations: k q´pIk rk{Jq k`1q “ ´ωpt´ k`1q “ ´hxpt´ k`1q “ ´vxpt´ k`1q “ hypt´ k`1q “ vypt´ k q´“vxpt´ k q´pIk{mq sinθ˚‰δk k q`pIk{mq sinθ˚ k `“vypt´ k q´p1{2qgδ2 k q`pIk{mq cosθ˚´gδk k q`pIk{mq cosθ˚‰δk ωpt´ hxpt´ vxpt´ hypt´ vypt´ k`1 ´ t´ fi pt´ where δk k q, k “ 1, 2,¨¨¨ , is the time of flight between two consecutive control actions. During this time duration, the devil-stick rotates by a net angle π´ 2θ˚. Since the angular velocity of the stick remains constant in the interval rt` k ,t´ k`1s, δk is given as follows , Δθ fi pπ´ 2θ˚q (4.3.12a) (4.3.12b) (4.3.12c) (4.3.12d) (4.3.12e) (4.3.13) k q`pIk rk{Jq The control design for juggling is presented next. ωpt´ δk “ Δθ 4.4 State Feedback Control Design 4.4.1 Steady-State Dynamics From the discussion in section 4.3.5 it becomes clear that when the change of reference frame of the juggler is taken into account, the problem of juggling between the two distinct configura- tions pθ˚, h˚ configurations pθ˚, h˚ by yq is converted to the problem of juggling between identical x, h˚ yq. If the state variables at this configuration are denoted x, h˚ yq and pθ˚, h˚ x, h˚ x, h˚ yq and pπ´ θ˚,´h˚ ¯Y ˚ fi„ω˚ h˚ x v˚ x h˚ y v˚ yT (4.4.1) 66 where I˚ k , r˚ value of the time of flight. Since h˚ k denote the steady-state values of the control inputs and δ˚ denote the steady-state y is eliminated from ( 4.4.2d), ( 4.4.2) represents six equations x, v˚ y, I˚, r˚, and δ˚. By choosing δ˚, the remaining six x, v˚ in seven unknowns, namely, ω˚, h˚ unknowns are obtained as follows: tanθ˚{4 ω˚ “ ´Δθ{δ˚, h˚ x “ gδ˚2 v˚ y “ ´gδ˚{2 x “ g tanθ˚δ˚{2, v˚ I˚ “ mgδ˚{ cosθ˚,r˚ “ 2J cosθ˚Δθ{pmgδ˚2q Since the point of application of the impulsive force must lie on the stick, r˚ in ( 4.4.3) must satisfy 0 ă r˚ ă ℓ{2. This imposes the following constraint of the value of δ˚: mgℓ It should be noted that for a given value of δ˚, the value of h˚ y is not unique. δ˚ ą ¯δ, ¯δ fi 2d J cosθ˚Δθ (4.4.2a) (4.4.2b) (4.4.2c) (4.4.2d) (4.4.2e) (4.4.2f) (4.4.3) (4.4.4) (4.4.5) then ¯Y ˚ “ Pp ¯Y ˚q is a fixed point and ( 4.3.12) and ( 4.3.13) give k{mq sinθ˚‰δ˚ y `pI˚ ω˚ “ ´ω˚ ´pI˚ k r˚ k{Jq x´“v˚ x´pI˚ x “ ´h˚ h˚ x `pI˚ v˚ x “ ´v˚ k{mq sinθ˚ y ´p1{2qgδ˚2 `”v˚ y “ h˚ h˚ y `pI˚ v˚ y “ v˚ k{mq cosθ˚´gδ˚ Δθ δ˚ “ k r˚ ω˚ `pI˚ k{Jq k{mq cosθ˚ıδ˚ 4.4.2 Error Dynamics To converge the states to their desired values, we first define the discrete error variables: k q´ ω˚ k q´ h˚ x, k q´ h˚ y, rωpkq fi ωpt´ rhxpkq fi hxpt´ k q´ v˚ rhypkq fi hypt´ k q´ v˚ ru1pkq fi pIkrk ´ I˚r˚q{J, ru2pkq fi pIk ´ I˚q{m rvxpkq fi vxpt´ rvypkq fi vypt´ x y 67 Using ( 4.3.12) and ( 4.4.2a)-( 4.4.2e), the error dynamics can now be written as rωpk` 1q “´rωpkq´ru1pkq rhxpk` 1q “´rhxpkq´δkrvxpkq` δk sinθ˚ru2pkq rvxpk` 1q “´rvxpkq` sinθ˚ru2pkq rhypk` 1q “rhypkq` δkrvypkq` δk cosθ˚ru2pkq`pg{2q”δkδ˚ rvypk` 1q “rvypkq` cosθ˚ru1pkq´grδk ´ δ˚ k s kı k ´ δ2 where δk, defined in ( 4.3.13), can be written in terms of the error variables as follows: (4.4.6a) (4.4.6b) (4.4.6c) (4.4.6d) (4.4.6e) (4.4.7) δk “ Δθδ˚ rrωpkq`ru1pkqsδ˚ ` Δθ It is clear from ( 4.4.6) and ( 4.4.7) that the error dynamics is nonlinear. In the next section we present a partial control design that converts the nonlinear system into a linear system and simplifies the remaining control design. 4.4.3 Partial Control Design: Dead-Beat Control The error dynamics in ( 4.4.6) involves two control inputs, namely, ru1pkq and ru2pkq. The input ru1pkq appears only in ( 4.4.6a). To this end, we first designru1pkq as follows: (4.4.8) to guarantee dead-beat convergence of the error staterωpkq. Substitution of ( 4.4.8) in ( 4.4.7) yields δk “ δ˚. Since, δ˚ is user-defined and is a constant, the choice of control in ( 4.4.8) is special as it ru1pkq “ ´rωpkq 68 transforms the remaining dynamics in ( 4.4.6b)-( 4.4.6e) into the linear system: zpk` 1q “ A zpkq` Bru2pkq zpkq fi„rhxpkq rvxpkq rhypkq rvypkqT »———————– »———————– ´1 ´δ˚ 0 0 ´1 fiffiffiffiffiffiffiffifl , B fi 1 δ˚ 0 0 0 0 0 1 0 0 0 δ˚ sinθ˚ sinθ˚ δ˚ cosθ˚ cosθ˚ fiffiffiffiffiffiffiffifl A fi (4.4.9) It can be verified that the pair pA , Bq is controllable since θ˚ P p0,π{2q and δ˚ ą 0. 4.4.4 Residual Control Design The error dynamics in ( 4.4.9) is linear and therefore the states can be converged to zero by simply designing a linear controller. However, it should be noted that the control inputru2pkq determines the value of Ik which also appears in the dead-beat control designru1pkq - see ( 4.4.5). By using the values ofru2pkq from ( 4.4.5) in ( 4.4.8), we get: (4.4.10) rk “ rI˚r˚ ´ Jrωpkqs{Ik Since the point of application of impulsive force must lie of the stick, rk must satisfy´ℓ{2ă rk ă ℓ{2. By imposing this condition on the value of rk in ( 4.4.10), we get the following constraints on the inputru2pkq: (4.4.11) ru2pkq ą r 2I˚r˚ ´ 2Jrωpkq´ I˚ℓs{pmℓq ru2pkq ą r´2I˚r˚ ` 2Jrωpkq´ I˚ℓs{pmℓq Since I˚ and r˚ are both positive, as it can be seen from ( 4.4.3), the inequalities in ( 4.4.11) can be combined into the single inequality: ru2pkq ą ¯a` ¯b |rωpkq| ¯a fi p2r˚ ´ ℓqI˚{pmℓq, ¯b fi 2J{pmℓq 69 (4.4.12) Sinceru1pkq is dead-beat, rωpkq “ 0, k “ 2, 3¨¨¨ . Thus, ( 4.4.12) can also be written as The input ru2pkq is designed using Linear Quadratic Regulator (LQR) and Model Predictive Control (MPC) methods. For an LQR design, the control minimizes the cost function ru2pkq ą ¯a` ¯b |rωpkq|, ru2pkq ą ¯a, k “ 1 k “ 2, 3,¨¨¨ J “ 2pkqı 8ÿk“1”zpkqT Q zpkq` Rru 2 (4.4.13) (4.4.14) (4.4.15) where, Q and R are constant weighting matrices that can be chosen by trial and error to satisfy the constraints in ( 4.4.13). The closed-form solution of the control input ru2pkq can be obtained by solving the Ricatti equation [99]. For a receding horizon MPC design, the constraint in ( 4.4.13) can be explicitly included in the optimization problem. In the MPC design5, it is necessary to calculate the predicted output with future control input as the adjusted variable. Since the current control input cannot affect the output at the same time for receding horizon control, the system dynamics must be represented in terms of the difference between the current and the predicted control input. To this end, we define the following variables based on the augmented state-space model6 in [100]: Δupkiq firu2pkiq´ru2pki ´ 1q Δupkiq Δupki ` 1q ... ΔUi fi , Zi fi zpki ` 1 | kiq zpki ` 2 | kiq ... Δupki ` Nc ´ 1q zpki ` Np | kiq fiffiffiffiffiffiffiffifl »———————– »———————– fiffiffiffiffiffiffiffifl where ki is the current sampling instant, zpkiq is the state vector in ( 4.4.9) measured at ki, Nc is the control horizon, Np is the prediction horizon, and zpki` m | kiq is the predicted state variable at ki ` m with state measurements zpkiq. 5A detailed discussion of MPC design for discrete-time systems can be found in Chapters 1-3 in [100]. 6The augmented state-space model is controllable; this was verified using Theorem 1.2 in [100]. 70 We now construct the following N-step receding horizon optimal control problem: Nÿi“1”ZT i Zi ` ΔU T i ΔUiı minimize J “ subject to zpki ` 1q “ A zpkiq` Bru2pkiq ru2pkiq ą ¯a` ¯b |rωp1q|, ru2pkiq ą ¯a, i “ 1 i “ 2, 3¨¨¨ , N (4.4.16) (4.4.17) In every sampling period, the optimization problem determines the best control parameter ΔUi that attempts to converge the sequence of states in Zi to zero. Although ΔUi contains Nc number of future control inputs, only the first entry is implemented as the actual control input. This optimization process is repeated using a more recent measurement of the states. It should be emphasized that the window. input constraint in ( 4.4.17), namely,ru2pkiq ą ¯a` ¯b |rωp1q| is imposed only in the first optimization In subsequent optimization windows, the constraint is relaxed to ru2pkiq ą ¯a. This is necessary forru2 to converge to zero since ¯a is negative - see ( 4.4.12), whereas ¯a` ¯b |rωp1q| can assume positive values based on the initial value of rω. Remark 1. The control inputru2pkq is obtained as the numerical solution of the optimal control problem in ( 4.4.16) and ( 4.4.17). These inputs are applied at discrete time instants and the optimization solver is required to compute these inputs within the sampling time interval, which is equal to the time of flight δ˚. Since δ˚ is relatively large, there is sufficient time for the optimization solver to generate the solution. This, along with the fact that the input constraint can be explicitly considered in the problem formulation, makes MPC well-suited for this problem. 71 4.5 Simulation Results 4.5.1 System Parameters and Initial Conditions We present simulation results of both LQR- and MPC-based control designs. The physical param- eters of the devil-stick are provided below in SI units: m “ 0.1, ℓ “ 0.5, J “ 0.0021 (4.5.1) Using these physical parameters and by choosing the values of θ˚ “ π{6 rad and δ˚ “ 0.5 sec, the steady-state values of state variables and control inputs are obtained from ( 4.4.3) as ω˚ “ ´4.18 rad{s h˚ v˚ y “ ´2.45 m{s x “ 0.353 m v˚ I˚ “ 0.565 Ns x “ 1.414 m{s r˚ “ 0.030 m Since h˚ y can be chosen arbitrarily, we chose h˚ y “ 3.0 m At the initial time, we assume θ “ θ˚ “ π{6 rad and the states variables (in SI units) are ωp0q “ 0, hxp0q “ 0.53, vxp0q “ 2.0, hyp0q“ 1.0, vyp0q“ ´2.0 (4.5.2) (4.5.3) (4.5.4) For the physical parameters in ( 4.5.1), steady-state values of the states in ( 4.5.2) and ( 4.5.3), and initial conditions in ( 4.5.4), the controlru1pkq was chosen according to ( 4.4.8). The control input ru2pkq was designed using LQR and MPC methods and simulation results are presented next. 4.5.2 Results for the LQR-based Design For the LQR design, the weight matrix Q for the states was chosen to be the identity matrix and the control weight R was chosen as 0.2. The control was obtained as ru2pkq “ Fzpkq, F “„´0.43 ´0.77 0.43 0.66 72 0.0 -5.0 0.6 0.1 3.0 1.0 1 (a) (c) (e) 5 k ωpkq (rad/s) 3.5 1.0 2.2 (b) (d) Epkq (J) hxpkq (m) vxpkq (m/s) hypkq (m) 1.0 -0.9 -2.6 1 10 (f) 5 k vypkq (m/s) 10 Figure 4.4: State variables and total energy E of the devil-stick at sampling instants k, k “ 1, 2,¨¨¨ , 10, for the LQR design. (a) (b) 0.7 0.5 1 Ik (Ns) 5 k 0.04 0.01 10 1 rk (m) 5 k 10 Figure 4.5: Control inputs for the devil-stick at sampling instants k, k “ 1, 2,¨¨¨ , 10, for the LQR design. The simulation results are shown in Figs. 4.47 and 4.5. It can be seen from Fig. 4.4 (a) that the dead-beat control ru1pkq converges ωpkq to ω˚ in one sampling interval. The control ru2pkq converges the remaining states to their steady-state values given in ( 4.5.2) in approximately k “ 10 steps - see Figs. 4.4 (c)-(f). The control inputs Ik and rk are shown in Figs. 4.5 (a) and (b). It can 7It should be noted that the state variables are shown in the reference frame of the right-handed juggler. 73 be seen that both control inputs converge to their steady-state values defined in ( 4.5.2); also the control input rk remains well inside the constraint boundary | rk |ă ℓ{2. The convergence of both the states and control inputs to their desired values imply that the devil-stick is juggled between two symmetric configurations. Since the magnitudes of vx, vy, ωx, and hy are the same in the two symmetric configurations, the total energy E (kinetic plus potential) reaches a constant value at steady state - see Fig. 4.4(b). Remark 2. The total energy of the devil-stick is the same at the symmetric configurations. Also, it is conserved during the flight phase. Therefore, in steady-state, the control inputs I˚ and r˚ do zero work on the system. 4.5.3 Results for the MPC-based Design The control horizon, prediction horizon, and the number of steps were taken as Nc “ 5, Np “ 10, N “ 15 The MPC problem, defined by ( 4.4.16) and ( 4.4.17) were solved using quadratic programming in Matlab8. The state variables hxpkq, hypkq, vxpkq and vypkq and the control inputs Ik and rk are shown in Fig. 4.6. The state variable ωpkq is not shown as it converged to its desired value in one sampling interval by the dead-beat controller. Similar to the LQR design, the control input rk remains well inside the constraint boundary. The trajectory of the center-of-mass of the devil stick is shown in Fig. 4.7 (a); it starts from the initial configuration phx, hyq “ p0.53, 1.00q and is eventually juggled between the symmetric coordinates ph˚ yq “ p´0.353, 3.00q in steady state. Typically, N is chosen to be large to guarantee convergence. For our system, the yq “ p0.353, 3.00q and p´h˚ x, h˚ x, h˚ states rapidly converged to zero with N “ 15. In Fig. 4.7 (b), the devil-stick is shown at the two symmetric configurations where θ˚ “ π{6 and several intermediate configurations that are equal time intervals apart. 8The quadprog Matlab function was used. 74 1.0 0.0 3.0 1.0 1.5 (a) (c) (e) hxpkq (m) hypkq (m) Ik (Ns) 2.0 1.0 -1.5 -2.6 0.04 (b) (d) (f) vxpkq (m/s) vypkq (m/s) rk (m) 0.1 1 5 10 k 0.01 1 15 5 10 15 k Figure 4.6: State variables and control inputs of the devil-stick at sampling instants k, k “ 1, 2,¨¨¨ , 15, for the MPC design. (a) (b) 3.0 ) m ( y h 1.0 0.5 hx (m) 3.50 ) m ( y h θ˚ 2.75 -0.6 1.5 θ˚ 0.6 0.0 hx (m) Figure 4.7: (a) Trajectory of the center-of-mass from the initial configuration to steady-state and (b) symmetric configurations and seven intermediate configurations of the devil-stick in steady state for the MPC design. Remark 3. In both simulations, the stick rotates by an angle pπ´ 2θ˚q between two consecutive control inputs. This corresponds to “top-only idle" juggling [47]. The controller is quite general and the stick can be controlled to rotate by pqπ´ 2θ˚q, q “ 2, 3,¨¨¨ , by simply changing the 75 definition of Δθ in ( 4.3.13) from Δθ “ pπ´ 2θ˚q to Δθ “ pqπ´ 2θ˚q. In other words, the stick can be made to flip multiple times in the flight phase, if desired. The “flip-idle" in [47] corresponds to the case where q “ 2. 76 CHAPTER 5 ENERGY-BASED ORBITAL STABILIZATION USING CONTINUOUS INPUTS AND IMPULSIVE BRAKING 5.1 Introduction We present a controller for energy-based orbital stabilization of the class of Euler-Lagrange systems that have one passive revolute joint. Examples of systems in this class include the pendubot, acrobot, cart-pendulum system, rotary pendulum etc. The orbit is defined by fixed positions of the active coordinates and a desired energy of the overall system. The controller is comprised of continuous-time inputs and impulsive inputs for braking. At first, a general result for underactuated systems is presented which shows that an impulsive input causes an instantaneous jump in the mechanical energy of the system. This jump is shown to be explicitly dependent on the change in the active velocities due to an impulsive input. This result is then used to show that impulsive braking causes a negative jump in the mechanical energy of the system as well as the Lyapunov-like function. Finally, using a state-dependent impulsive dynamical system model [101], we present sufficient conditions for stabilization. To demonstrate the generality of our approach, we demonstrate orbital stabilization for the three-DOF Tiptoebot [6] through simulations. Experimental validation of the control design is carried out on a rotary pendulum to show the applicability of our approach in real hardware. 5.2 Problem Statement Consider an n degree-of-freedom underactuated system with one passive degree-of-freedom. The generalized coordinates of the system are denoted by q, q fi“qT 1 q2‰T , where q1 P Rn´1 and q2 P R are the coordinates associated with the active and passive degrees-of-freedom. Our control objective is to stabilize the orbit defined by pq1, 9q1, Eq “ p0, 0, Edesq (5.2.1) 77 where E is the total mechanical energy of the system and is given by the relation Epq, 9qq “ 1 2 9qT Mpqq 9q` Fpqq (5.2.2) and Edes is the desired value of E. In (5.2.2), Mpqq P Rnˆn is the mass matrix, assumed to be positive definite, and Fpqq is the potential energy, assumed to be a smooth function of q. The mass matrix is partitioned as Mpqq “»—– M11pqq M12pqq MT 12pqq M22pqq fiffifl (5.2.3) where M11 P Rpn´1qˆpn´1q and M22 P R. Assumption 1. The mechanical energy of the system is assumed to be periodic in the passive coordinate q2, such that Epq2 ` 2πq “ Epq2q. Remark 4. Assumption 4 is easily satisfied if the passive degree-of-freedom is a revolute joint. Assumption 2. The elements of the mass matrix Mpqq are bounded and the potential energy Fpqq is lower bounded. Furthermore, if q1 “ q˚ dq2 rkP1pq˚ 1, q2q` P2pq˚ 1 is constant, then 1, q2qs ­“ 0 d where k is any scalar constant, and P1pq˚ 1, q2q “ 2 M22˜„BM12 Bq2 ´ P2pq˚ 1, q2q “ ´P1pq˚ 1, q2qF ´ 1 ´ M12 Bq2 ¸ Bq1 T 2 M22„BM22 2„BM22 Bq2 `„BF Bq1T BF M12 M22 Remark 5. The boundedness property of Mpqq and Fpqq is satisfied for systems that have no prismatic joints. For a given underactuated system with one passive DOF, assumption 5 can be easily verified. 78 5.3 Modeling of System Dynamics 5.3.1 Lagrangian Dynamics For our system described in section 5.2, the equations of motion can be written as: d dtˆBL dtˆBL B 9q1˙´ˆBL B 9q2˙´ˆBL Bq1˙ “ u Bq2˙ “ 0 d (5.3.1) where L pq, 9qq is the Lagrangian and u P Rn´1 is the vector of independent control inputs. Each element of the vector u is a continuous function of time for all t ě 0 except at discrete instants t “ τk, k “ 1, 2,¨¨¨ , where it is impulsive in nature. At t “ τk, the impulsive input vector has the form upτkq “ Ik δpt ´ τkq, where δpt ´ τkq is the Dirac measure at time τk and Ik P Rn´1 is the impulse of the impulsive input. The Lagrangian has the form L pq, 9qq “ 1 2 9qT Mpqq 9q´ Fpqq By substituting (5.2.3) in (5.3.2), the Lagrangian is written as L pq, 9qq “ 1 2 9qT 1 M11 9q1 ` 1 2 2 M22 9q 2 ` 9qT 1 M12 9q2 ´ F and by substituting (5.3.3) in (5.3.1), the equations of motion are written in the form: M11 :q1 ` M12 :q2 ` h1pq, 9qq “ u MT :q1 ` M22 :q2 ` h2pq, 9qq “ 0 12 (5.3.2) (5.3.3) (5.3.4a) (5.3.4b) where h1 “ 9M11 9q1 ` 9M12 9q2 ´ h2 “ 9M22 9q2 ` 9qT 1 9M12 ´ 1 Bq1pM11 9q1q 9q1 ´„BpM12 9q2q 2„ B Bq2  9q 2„BM22 1„BpM11 9q1q ´ Bq1 Bq2 9qT 1 1 2 1 2 9q Bq1 T  9q1 ´ 2„BM22 1„BpM12 9q2q Bq1T 2 `„BF ` BF Bq1 Bq2 2 2 ´ 9qT (5.3.5a) (5.3.5b) 79 Equations (5.3.4a) and (5.3.4b) can be rewritten in the form :q1 “ Apq, 9qq` Bpqqu :q2 “ ´p1{M22q”MT 12tApq, 9qq` Bpqquu` h2ı where Bpqq “”M11 ´p1{M22qM12 MT 12ı´1 Apq, 9qq “ p1{M22qBpqqrM12 h2 ´ h1M22s (5.3.6a) (5.3.6b) (5.3.7) (5.3.8) Using the properties of the mass matrix Mpqq and the Schur complement theorem [102], it can be shown that Bpqq is well-defined, symmetric and positive-definite, i.e., Bpqq “ BTpqq ą 0. 5.3.2 Effect of Impulsive Inputs When the control input u in (5.3.4a) is impulsive, it causes discontinuous jumps in the velocities p 9q1, 9q2q, while the coordinates pq1, q2q remain unchanged. For the impulsive input at t “ τk, the jump in the velocities is computed by integrating (5.3.4) as follows [86]: »—– M11 M12 MT 12 M22 In the above equation, Δ 9q1 and Δ 9q2 are defined as fiffifl »—– Δ 9q1 Δ 9q2 fiffifl “»—– Ik 0 fiffifl , Ik fiż t` k t´ k uptkq dt Δ 9q1 fi p 9q` 1 ´ 9q´ 1 q, Δ 9q2 fi p 9q` 2 ´ 9q´ 2 q where 9q´ fi 9qpτ´ k q denote the generalized velocities immediately before and after application of the impulsive inputs. Since the system is underactuated, the jump in 9q2 is dependent k q and 9q` fi 9qpτ` on the jumps in 9q1; this dependence is described by the one-dimensional impulse manifold [23] or impulse line, obtained from the equation above: 9q` 2 “ 9q´ 2 ´p1{M22qMT 12p 9q` 1 ´ 9q´ 1 q (5.3.9) 80 The kinetic energy undergoes an instantaneous change due to jumps in the generalized velocities. This change is also equal to the change in the total mechanical energy of the system since the potential energy is only a function of the generalized coordinates. A formal statement of this result is provided next. Lemma 1. For the dynamical system in (5.3.4), the jump in the total mechanical energy due to application of an impulsive input is given by ΔE fi pE` ´ E´q “ 1 2 9q`T 1 B´1pqq 9q` 1 ´ 1 2 9q´T 1 B´1pqq 9q´ 1 (5.3.10) where E´ and E` are the energies immediately before and after application of the impulsive input. Proof: See section 5.7.1. Remark 6. The proof of Lemma 1 is provided for the general case where the number of active and passive degrees-of-freedom are pn´ mq and m, respectively. This general result indicates that the change in mechanical energy due to an impulsive input depends only on the velocities of the active degrees-of-freedom immediately before and after application of the input. The result is analogous to the passivity property for the continuous-time case [103], where the power input to the system is the inner product of the velocities of the active degrees-of-freedom and control inputs. It is important to note that results similar to Lemma 1 appeared earlier in [104]. We define impulsive braking as the process of applying impulsive inputs that result in 9q` 1 “ 0. It follows from Lemma 1 that impulsive braking results in a loss of mechanical energy, given by the expression ΔE “ ´ 1 2 9q´T 1 B´1pqq 9q´ 1 (5.3.11) The discontinuous jumps in the generalized velocities and kinetic energy of the system also occurs when mechanical systems experience forces due to impact. In this regard, results similar to the ones presented above can be found in [104–106]. We now state an important result related to impulsive braking. 81 Lemma 2. Consider the scalar function V “ 1 2”qT 1 Kp q1 ` 9qT 1 Kd 9q1 ` KepE ´ Edesq2ı (5.3.12) where Kp and Kd are diagonal positive definite constant matrices and Ke is a positive constant. Impulsive braking results in a discontinuous jump in the function given by ΔV fi pV ` ´V ´q “ ´ 1 2 9q´T 1 „ 1 4"Ke 9q´T 1 B´1pqq 9q´ 1* B´1pqq` Kd ` KepE` ´ EdesqB´1pqq 9q´ 1 (5.3.13) where V ´ and V ` are values of the function immediately before and after impulsive braking. Furthermore, if”Kd ` KepE` ´ EdesqB´1pqqı is positive definite, then ΔV ď 0, and ΔV “ 0 if and only if 9q´ 1 “ 0. Proof: See section 5.7.2. 5.3.3 Hybrid Dynamical Model For orbital stabilization of our system in (5.3.4), we propose a control strategy comprised of continuous and impulsive inputs. The impulsive inputs will be used for impulsive braking of the active coordinates, i.e., 9q` 1 “ 0. As a result, the change in the velocities can be obtained using (5.3.9) as follows: Δ 9q1 “ 0´ 9q´ 2 ´ 9q´ Δ 9q2 “ 9q` 1 “ ´ 9q´ 2 “ p1{M22qMT 1 12 9q´ 1 (5.3.14) In addition to the impulsive braking inputs, we will reset the passive coordinate q2 periodically to confine it to the compact set r´3π{2,π{2s. To describe the dynamics of our system, we adopt the state-dependent impulsive dynamical model in [101, pg.20]: 9xptq “ fcrxptqs, Δxptq “ fdrxptqs, xp0q “ x0, xptq R Z xptq P Z (5.3.15a) (5.3.15b) 82 where Z defines the set where the impulsive inputs are applied and/or periodic resetting occurs. For our system, 1 q2 9qT 1 xptq fi rqT Δxptq fi xpt`q´ xpt´q 9q2sT P D Ď R 2n In the above expression, D is the open set where q2 P pa, bq, a ă ´3π{2, b ą π{2, and xpt´q, xpt`q are the values of the state variables immediately before and after application of impulsive inputs or coordinate resetting. Using (5.3.6), (5.3.9) and (5.3.14), it can be shown fiffiffiffiffiffiffiffifl : xptq P Z1 : xptq P Z2 : xptq P Z3 (5.3.16) (5.3.17) fc “ fd “ »———————– $’’’’’’’&’’’’’’’% 9q1 9q2 Apq, 9qq` Bpqqu 12tApq, 9qq` Bpqquu` h2‰ 1T p1{M22qMT 9q´ 12 1 ´p1{M22q“MT „0 0 ´ 9q´ „0 2π 0 0T „0 ´2π 0 0T Z “ Z1 Y Z2 Y Z3, Z1 is the set where impulsive braking inputs are applied (to be defined in Theorem 2 where the control design will be presented), and Z2 fi tq2 “ ´3π{2, 9q2 ă 0u and Z3 fi tq2 “ π{2, 9q2 ą 0u are the sets where coordinate resetting occurs. Note that the solution xptq of (5.3.15) is left-continuous, i.e., it is continuous everywhere except at time instants tk, k “ 1, 2,¨¨¨ , where tk’s denote the time instants when impulsive inputs are applied or coordinate resetting occurs. Since impulsive inputs are applied at τk, k “ 1, 2,¨¨¨ , tτ1,τ2,¨¨¨u Ă tt1,t2,¨¨¨u. Thus, xpt`q and xpt´q, in the equation after (5.3.15), were defined 83 without the time stamp, i.e. xpt´q fi xpt´ xpt`q fi xpt´ k q “ lim εÑ0` k q` fdrxpt´ xptk ´ εq k qs “ lim εÑ0` xptk ` εq The well-posedness of the tk’s and the quasi-continuous dependence of the solution to (5.3.15) with respect to initial conditions [101, 104, 107] will be established later in the proof of Theorem 4 where Z1 will be defined. 5.4 Hybrid Control Design 5.4.1 Main Result For the control objective in (5.2.1), we propose a hybrid control strategy comprised of a continuous controller and impulsive braking1. Theorem 4 provides the design of the continuous controller and defines the set Z1, where impulsive braking is applied. The proof of Theorem 2 is based on a Lyapunov-like function. The continuous controller is invoked as long as the derivative of the Lyapunov-like function is negative semi-definite; when this condition is not satisfied, impulsive braking is applied to produce negative jumps in the Lyapunov-like function. Before stating Theorem 4, we present Lemma 3 and an invariant set theorem [101, pg.38] that will be used in the proof of Theorem 4. Lemma 3. For the system in (5.3.4) subjected to continuous control, q2 is constant if 9q1 and 9u are identically zero. Proof: See section 5.7.3. Remark 7. Lemma 3 implies that the active and passive generalized coordinates are dynamically coupled. Due to this coupling, the active generalized coordinates cannot be held stationary by 1It is assumed that the active degrees-of-freedom will have a friction brake such that they can be stopped instantaneously. 84 constant generalized forces when the passive generalized coordinate is non-stationary. The proof of Lemma 3 is based on assumption 5, which is satisfied when dynamical coupling exists. The existence of such coupling has been verified for an inverted pendulum on a cart [74], pendubot, acrobot, and reaction-wheel pendulum; in this work it is shown for the tiptoebot and the rotary pendulum. Theorem 3. [101, pg.38] Consider the impulsive dynamical system given by (5.3.15), assume that Dc Ă D is a compact positively invariant set with respect (5.3.15), and assume that there exists a continuously differentiable function W : D Ñ R such that rBWpxq{Bxs fcpxq ď 0, Wpx` fdpxqq ď Wpxq, x P Dc, x P Dc, x R Z x P Z (5.4.1a) (5.4.1b) Let R fi tx P Dc : x R Z , rBWpxq{Bxs fcpxq “ 0uYtx P Dc : x P Z ,Wpx` fdpxqq “ Wpxqu and let M denote the largest invariant set contained in R. If x0 P Dc, then xptq Ñ M as t Ñ 8. Remark 8. Although the invariance principle in [101] will be used (Theorem 3), the reader should be aware of the invariance principles in [55, 106, 108] that can be used for analysis of non-smooth and hybrid systems. Theorem 4. For the impulsive dynamical system defined by (5.3.15), (5.3.16) and (5.3.17), and x0 P D such that ´3π{2 ă q2p0q ă π{2 the following choice of control design: u “ ´rpKd ` Kcq Bpqq` KepE ´ Edesq Is´1“Kp q1 `pKd ` KcqApq, 9qq‰ Z1 “ txptq | rApq, 9qq` BpqqusT Kc 9q1 ď 0, 9q1 ­“ 0u (5.4.2) (5.4.3a) (5.4.3b) where I is the identity matrix and Kc is a diagonal positive-definite matrix, guarantees asymptotic stability of the orbit in (5.2.1) if the gain matrices Kp, Kd and Ke satisfy the following conditions: 85 (i) ”Kd ` KepE ´ EdesqB´1pqqı is positive definite for all q and 9q, (ii) If q˚ 1 and q˚ 2 are constant values of q1 and q2, then the following system of equations: Kp q˚ 1 KerFpq˚q´ Edess „BF Bq1T q“q˚ “ ´ Bq2q“q˚ “ 0 „BF yields a finite number of solutions with q˚ 1 “ 0, and (iii) For all possible solutions of q˚ 2 obtained from (ii) and for the function V in (5.3.12), the following inequality is satisfied Vpt “ 0q ă mintV | q1 “ 0, 9q1 “ 0, E P SEztEdesuu where SE is the set of values of E evaluated at q1 “ 0, q2 “ q˚ 2, 9q “ 0. Proof: Consider the Lyapunov-like function V defined in (5.3.12); V is zero on the orbit defined in (5.2.1) and positive everywhere else. The time derivative of V is 1 Kd 9V “ qT 1 Kp 9q1 ` :qT “”qT 1 Kp ` :qT 9q1 ` KepE ´ Edesq 9E 1 Kd ` KepE ´ Edesq uTı 9q1 (5.4.4) where 9E “ uT 9q1 follows from the passivity property of underactuated Euler-Lagrange systems - see [103]2 and proposition 2.5 of [109]. By substituting :q1 from (5.3.6a) in (5.4.4) and using the symmetry of Bpqq, we get 9V “”qT 1 Kp ` AT Kd ` uT B!Kd ` KepE ´ EdesqB´1)ı 9q1 (5.4.5) The following choice of u uT “ ´”qT 1 Kp ` AT Kd ` :qT 1 Kcı”B!Kd ` KepE ´ EdesqB´1)ı´1 , (5.4.6) 2The proof of the passivity property follows from the fact that the matrix [ 9M ´ 2C] is skew- symmetric for our choice of generalized coordinates. 86 which is well defined based on condition (i), results in 9V “ ´ :qT 1 Kc 9q1 (5.4.7) Substitution of (5.3.6a) in (5.4.6) followed by algebraic manipulation gives the expression for u in (5.4.3a). Substitution of (5.3.6a) in (5.4.7) gives 9V “ ´rApq, 9qq` BpqqusT Kc 9q1 (5.4.8) Based on the expression of 9V , three cases may arise: case (a): if rA` BusT Kc 9q1 ą 0, then 9V ă 0, case (b): if rA` BusT Kc 9q1 ď 0, 9q1 ­“ 0, then x P Z1 and impulsive braking is applied - see (5.4.3b). Since condition (i) is satisfied, Lemma 2 indicates that V undergoes a discontinuous change ΔV , where ΔV ă 0, and case (c): if 9q1 “ 0, then 9V “ 0. For case (b), impulsive braking results in 9q1 “ 0 at t` and the trajectories of the system leave Z1. If 9q1 ” 0 for all t ą t`, the trajectories of the system remain outside Z1 and 9V ” 0. If 9q1 ı 0 for t ą t`, V decreases and the trajectories of the system move away from Z1 since (5.4.7) implies 9Vpt`q “ 0, :Vpt`q “ ´ :qT 1 pt`qKc :q1pt`q ă 0 ñ 9Vpτq ă 0, τ P pt`,t` ` εq for some ε ą 0 since Kc is positive-definite and :q1 ­“ 0. The trajectories may re-enter Z1, but not arbitrarily quickly. Hence Zeno phenomenon will not be observed; this is discussed in section 5.7.4. Case (c) implies that either 9q1 ” 0 ñ 9V ” 0, or 9q1 ı 0 and V continues to decrease again; this follows from our discussion of the nature of trajectories after impulsive braking. Cases (a), (b) and (c) imply that for t ą 0, Vptq ď Vp0q fi c and therefore the set Dc fi tV ď cuXt´3π{2 ď q2 ď π{2u 87 is positively invariant. Cases (a), (b) and (c) together satisfy the conditions in Theorem 33 with Dc defined above and Wpxq “ Vpxq. Since (b) implies ΔV ă 0, tx P Dc : x P Z , ΔV “ 0u is an empty set. Therefore, xptq Ñ M Ă R “ tx P Dc : x R Z , 9V “ 0u as t Ñ 8. From case (c), 9V “ 0 implies 9q1 “ 0 and thus R “ tx P Dc : 9q1 ” 0u. In R, :q1 “ 0. Substitution of :q1 “ 0 in (5.3.6a) and (5.4.6) yields uT “ ´AT B´1 uT BKd “ ´KepE ´ EdesquT ´ qT 1 Kp ´ AT Kd Substitution of (5.4.9a) into (5.4.9b) gives uT KepE ´ Edesq` qT 1 Kp “ 0 (5.4.9a) (5.4.9b) (5.4.10) The definition of R in Theorem 3 implies V is constant in R. Also, q1 is constant and 9q1 “ 0 in R. Therefore, from the definition of V in (5.3.12), we can claim that E is constant in R. Let q˚ 1 and E˚ be the constant values of q1 and E. We now discuss two cases that can arise: case 1: If E˚ “ Edes, we have q˚ case 2: if E˚ ­“ Edes, we get from (5.4.10) 1 “ 0 from (5.4.10). This implies that M is the orbit in (5.2.1). u fi u˚ “ ´ Kp q˚ 1 KepE˚ ´ Edesq (5.4.11) where u˚ is the constant value of the continuous control in R. For case 2, both q1 and u are constants. Therefore, based on Lemma 3, we claim q2 “ q˚ 2 is a constant. It follows from (5.2.2) that E˚ “ Fpq˚q. Using (5.3.4) and (5.3.5), we can show that the trajectories in R satisfy Bq1T „BF q“q˚ “ u˚, „BF Bq2q“q˚ “ 0 3Before we can apply Theorem 3, we must ensure that the time instants tk, k “ 1, 2,¨¨¨ , are well-posed and the hybrid dynamical system in (5.3.15) satisfies the quasi-continuous dependence property. The conditions for well-posedness and its proof appear in section 5.7.5. A sufficient condition for satisfying the quasi-continuous dependence property and its proof appear in section 5.7.6. 88 Substituting the expression for u˚ from (5.4.11) in the above equation along with E˚ “ Fpq˚q, we can use condition (ii) to claim q˚ 1 “ 0. Using (5.3.12) and cases (a) and (b), we can claim that as t Ñ 8, V Ñ V ˚, where V ˚ “ 1 2 KepE˚ ´ Edesq2 ď Vpt “ 0q where E˚ P SE. Since V ˚ ď Vpt “ 0q, we can claim using condition (iii) that E˚ “ Edes, i.e., V ˚ “ 0. Thus the largest invariant set M is the orbit defined in (5.2.1). This concludes the proof. (cid:3) 5.4.2 Choice of Controller Gains It can be easily shown that condition (i) in Theorem 4 is satisfied if p1{KeqλminpKdq ą rEdes ´ minpFqsλmaxrB´1pqqs where λminpKdq and λmaxrB´1pqqs are the minimum and maximum eigenvalues of Kd andrB´1pqqs. Assumption 5 implies λmaxrB´1pqqs and minpFq exist and therefore Kd and Ke can always be chosen to satisfy condition (i). For the choice of Ke satisfying condition (i), Kp has to be chosen to satisfy condition (ii). Although we do not prove that condition (ii) can be simultaneously satisfied for the general case, several combinations of gains pKp, Kd, Keq were found to exist for the inverted pendulum on a cart [74]. The authors have independently verified that condition (ii) can be easily satisfied for several other underactuated mechanical systems, namely, the pendubot, the acrobot, and the reaction-wheel pendulum. It is shown that conditions (i) and (ii) can be simultaneously satisfied for the three-DOF Tiptoebot and the rotary pendulum. These examples indicate that condition (ii) is not restrictive. Once the controller gains Kp, Kd and Ke have been chosen to satisfy conditions (i) and (ii) in Theorem 4, condition (iii) imposes no additional restrictions on the gains but simply provides an estimate of the region of attraction of the orbit. Since Kc does not appear in conditions (i)-(iii), it can be chosen without restriction. 89 5.5 Illustrative Example - The Tiptoebot We consider the tiptoebot example as presented in 2.3.1. Using the following definition for the joint angles qT 1 “ r θ2 θ3 sT , q2 “ θ1 (5.5.1) the dynamics of the tiptoebot can be expressed in the form of (5.3.4); the components of mass matrix in (5.2.3) are M11 “»—– M12 “»—– fiffifl α2`α3`2α5 cosθ3 α3`α5 cosθ3 α3`α5 cosθ3 α3 α2`α3`α4 cosθ2`2α5 cosθ3`α6 cospθ2`θ3q α3`α5 cosθ3`α6 cospθ2`θ3q fiffifl (5.5.2) (5.5.3) M22 “ α1`α2`α3 ` 2rα4 cosθ2 ` α5 cosθ3`α6 cospθ2`θ3qs where αi, i “ 1, 2,¨¨¨ , 6, are lumped parameters, defined as follows: α1 fi m1pℓ2 2 ` ℓ2 α4 fi m2ℓ1ℓ2 ` m3ℓ1ℓ2 1 ` ℓ2 3q, α2 fi pm2 ` m3qℓ2 2, α3 fi m3ℓ2 3, α5 fi m3ℓ2ℓ3, α6 fi m3ℓ1ℓ3 The sum of Coriolis, centrifugal and gravitational force terms, h1 and h2, can be obtained using (5.3.5), where Fpqq has the expression F “ β1 sinθ1 ` β2 sinpθ1 ` θ2q` β3 sinpθ1 ` θ2 ` θ3q β1 fi pm1 ` m2 ` m3qℓ1 g, β2 fi pm2 ` m3qℓ2 g, β3 fi m3ℓ3 g (5.5.4) The control input is defined as u “ rτ2 τ3sT . In the compact set θ1 P r´3π{2,π{2s, as defined in section 5.3.3, the upright equilibrium configuration of the tiptoebot is defined by θ1 “ ´3π{2 or π{2,„ θ2 θ3 9θ1 9θ2 9θ3  “„ 0 0 0 0 0  is unstable, but can be stabilized, by a linear controller, for example. The stabilized equilibrium will typically have a finite region of attraction; therefore, to stabilize from an arbitrary initial 90 configurations, we first use the controller in section 5.3 to stabilize an orbit that intersects the region of attraction. The obvious choice for such an orbit is the one where Edes equals the potential energy of the system at the equilibrium. Substitution of θ1 “ ´3π{2 or π{2 and θ2 “ θ3 “ 0 in (5.5.4) yields Edes “ β1 ` β2 ` β3. The control objective in (5.2.1) can therefore be written as θ2 “ θ3 “ 0, 9θ2 “ 9θ3 “ 0, Edes “ pβ1 ` β2 ` β3q (5.5.5) The feasibility of our control design is discussed next. 5.5.1 Selection of Controller Gains The initial configuration of the tiptoebot is taken as r θ1 θ2 θ3 9θ1 9θ2 9θ3 s “ r 0 π π 0 0 0 s (5.5.6) In this configuration, the tiptoebot is coiled up: the first link is horizontal, the second link folds back on the first link, and the third link folds back on the second link. The links were chosen to have the same mass m1 “ m2 “ m3 “ 0.1 kg and the same length ℓ1 “ ℓ2 “ ℓ3 “ 0.6 m. For this choice of mass and length parameters, the lumped parameters of the tiptoebot, defined in (5.5.3) and (5.5.4), are provided in Table 6.1 below: Table 5.1: Tiptoebot lumped parameters in SI units α1 0.108 α4 0.072 β1 1.764 α2 0.072 α5 0.036 β2 1.176 α3 0.036 α6 0.036 β3 0.588 The passive joint of the tiptoebot is revolute and therefore assumption 1 holds good. Assumption 2 also holds good - this is discussed in 5.7.7. The following choice of gains satisfy condition (i) and (ii): Kp “»—– 70 0 0 70 fiffifl , Kd “»—– 91 2.8 0 0 2.8 fiffifl , Ke “ 2.2 (5.5.7) Condition (ii) results in θ˚ 2 “ θ˚ 3 “ 0, which upon substitution in (5.3.4b) and (5.3.5b) yields BF Bq2 “ 0 ñ cosθ˚ 1 “ 0 (5.5.8) From section 5.3.3 we know that q2 lies in the compact set r´3π{2,π{2s. Thus θ1 lies in the same compact set - see (5.5.1). In this set, the possible solutions of (5.5.8) are θ˚ 1 “ t´3π{2,´π{2,π{2u. 3 “ 0, we know that E “ Edes. Therefore, to satisfy condition 1 “ ´3π{2 or π{2, and θ˚ 2 “ θ˚ For θ˚ (iii), we use θ˚ 1 “ ´π{2; this results in the following inequality Vpt “ 0q ă 2KerEpq˚ 1 “ 0, q˚ 2 “ ´π{2q´ Edess2 “ 2Kepβ1 ` β2 ` β3q2 For the initial configuration in (5.5.6), Ke in (5.5.7) satisfies the inequality above. The matrix Kc was chosen as Kc “»—– 1.2 0 0 1.2 fiffifl (5.5.9) 5.5.2 Simulation Results For the initial configuration in (5.5.6) and controller gains in (5.5.7) and (5.5.9), the simulation results are shown in Figs.5.1 and 5.2. The effect of impulsive braking can be seen in Figs.5.1 (d) and (f), where 9θ2 and 9θ3 (the velocities of the active joints) jump to zero on multiple occasions. Each impulsive braking also results in a negative jump in the mechanical energy (follows from Lemma 1) which can be seen in Fig.5.1 (b). Since impulsive inputs cause no jumps in the joint angles, there is no change in θ1, θ2 and θ3 at the time of impulsive braking - see Figs.5.1 (a), (c) and (e). In Fig.5.1 (a), θ1 never leaves the set r´3π{2,π{2s and therefore virtual impulsive inputs are not applied. While impulsive brakings cause negative jumps in the total energy E, the continuous-time controller in (5.4.3a) adds energy to the system; together, they converge the energy to the desired values Edes - see Fig.5.1 (b). The phase portrait of the passive joint is shown in Fig.5.2 (a). The jumps in the phase portrait (vertical drops in 9θ1, twice) is due to impulsive braking. The variation 92 1.57 -1.57 -4.71 3.0 0.0 3.0 0.0 0 (a) (c) (e) t (s) θ1 (rad) θ2 (rad) 10.0 0.0 -3.5 0.0 -14.0 0.0 (b) pE ´ Edesq (J) (d) (f) 9θ2 (rad/s) θ3 (rad) 9θ3 (rad/s) -10.0 6 0 t (s) 6 Figure 5.1: Plots of the joint angles θ1, θ2, θ3, error in the desired energy pE ´ Edesq, and the active joint velocities 9θ2, 9θ3 of the Tiptoebot. (a) (b) 700 9θ1 vs θ1 V 9 0 -5 -4.71 -1.57 1.57 0 0 t (s) 6 Figure 5.2: Plots showing (a) phase portrait of passive joint angle θ1, and (b) variation of the Lyapunov-like function V . The desired orbit is shown using dashed green line in (a). of the Lyapunov-like function V with time is shown in Fig.5.2 (b) - it can be seen that V decreases monotonically due to the action of the continuous-time controller and undergoes negative jumps intermittently due to impulsive brakings. The continuous controller and impulsive brakings work together to converge V to zero. 93 The gain matrices in (5.5.7) and (5.5.9) were chosen such that convergence to the desired orbit is fast. The simulation results indicate that the system trajectories reach a close neighborhood of the desired orbit very quickly, at approximately 3 s. For stabilization of the equilibrium in (5.5.6), a linear controller was designed using LQR. The matrices Q and R of the algebraic Ricatti equation were chosen to be I6ˆ6 and 2I2ˆ2, where Ikˆk is the identity matrix of size k. The linear controller was invoked when V ď 0.05 and | θ1 ´ π{2 |ď 0.05. 5.6 Experimental Validation 5.6.1 System Description Experiments were done with a rotary pendulum. As shown in Fig.5.3, the system is comprised of a horizontal arm OA of mass ma and length ℓa, which rotates about point O, and a pendulum of mass mp and length ℓp, that rotates about point A. The center-of-mass of the horizontal arm is located at a distance da from O and the center-of-mass of the pendulum is located at a distance dp from A. The horizontal arm is actively controlled by an external torque τ and its angular displacement about the z axis is denoted by φ. The pendulum is passive and its angular displacement about the εr axis is denoted by θ. The accleration due to gravity is denoted by g. With the following definition: rq1 q2sT “ rφ θsT (5.6.1) z O φ g τ x y ℓa B ℓp θ εθ A εr Figure 5.3: Schematic of a rotary pendulum. 94 the dynamics of the system can be expressed in the from given by (5.3.4), where u “ τ, and 2 θ, M12 “ γ3 sinθ, M22 “ γ2 M11 “ γ1 ` γ2 cos h1 “ γ3 cosθ 9θ2 ´ 9φ 9θγ2 sin 2θ, a, γ2 fi mpd γ1 fi mad 2 a ` mp ℓ2 h2 2 p, γ3 9φ2 “ γ2 sinθ cosθ` γ4 cosθ fi ´mp ℓadp, γ4 fi mpgdp (5.6.2) The physical parameters of the experimental setup are γ1 “ 0.0120, γ2 “ 0.0042, γ3 “ ´0.0038, γ4 “ 0.1190 (5.6.3) The control torque was applied by a 24-Volt permanent magnet brushed DC motor4.The motor is driven by a power amplifier5 operating in current mode. The motor torque constant is 37.7 mNm/A and the amplifier gain is 4.4 A/volt. An electromagnetic friction brake6 was integrated to the shaft of the DC motor. In the OFF state, the brake engages a friction pad to the shaft of the motor which prevents the shaft from turning; in the ON state, the brake is disengaged and the motor shaft rotates freely. For impulsive braking, the brake was kept engaged till the active velocity 9φ reached a close neighborhood of zero. The brake was powered ON/OFF by sending command voltage signals through an n-channel mosfet. The rotary pendulum was interfaced with a dSpace DS1104 board and the Matlab/Simulink environment was used for real-time data acquisition and control with a sampling rate of 1 Khz. The angular positions of the links were measured using incremental optical encoders; the angular velocities were obtained by differentiating and low-pass filtering the position signals. 5.6.2 Selection of Controller Gains The total energy of the system is obtained from (5.2.2) as follows 4The motor manufacturer is Faulhaber Drive Systems. The motor has a gearbox with a reduction ratio of 3.71 : 1. 5The amplifier is a product of Advanced Motion Control. 6The electromagnetic brake is manufactured by Anaheim Automation, model BRK-20H-480- 024. The brake can withhold torques up to 3.4 Nm. 95 1 2pγ1`γ2 cos E “ F “ γ4 sinθ 2 θq 9φ2 ` 1 2 γ2 9θ2 ` γ3 sinθ 9φ 9θ` F (5.6.4) For the control objective in (5.2.1), we choose Edes to be equal to the energy associated with the homoclinic orbit that contains the upright equilibrium „ φ θ 9φ 9θ  “„ 0 π{2 0 0  or „ 0 ´3π{2 0 0  Using (5.6.4), the energy associated with the homoclinic orbit can be written as Edes “ γ4 (5.6.5) The passive joint of the rotary pendulum is revolute and thus assumption 1 holds good. Assumption 5 also holds good - this is shown in 5.7.7. From (5.6.4) we know that F is only a function of θ and therefore condition (ii) is trivially satisfied resulting in the solution φ˚ “ 0. In the compact set r´3π{2,π{2s, the possible solutions of θ˚ obtained from condition (ii) are θ˚ “ t´3π{2,´π{2,π{2u. At θ˚ “ π{2 or θ˚ “ ´3π{2 and φ˚ “ 0, E “ Edes. Using condition (iii), we therefore get θ˚ “ ´π{2; this implies that Ke should be chosen to satisfy Vpt “ 0q ă 2Keγ2 4 (5.6.6) At the lower equilibrium configuration where rφ θ 9φ 9θs “ r0 ´π{2 0 0s, we have V “ 2Keγ2 violates the inequality in (5.6.6). This implies that our controller cannot swing-up the pendulum 4 . This when the system is exactly at the lower equilibrium.Therefore, in experiments, a small external perturbation was provided such that the system is not at the lower equilibrium at the initial time. For the experimental results presented herein, the initial configuration of the system after the perturbation was measured as „φp0q θp0q 9φp0q 9θp0qT “„0.01 ´1.42 0.05 0T (5.6.7) For the initial conditions in (5.6.7) and physical parameter values in (5.6.3), the following gains satisfied conditions (i)-(iii): Kp “ 0.5, Kc “ 0.08, Kd “ 0.3, Ke “ 100 (5.6.8) 96 For the above set of gains, the experimental results are presented next. 5.6.3 Experimental Results The experimental results are shown in Fig.5.4. The hybrid controller for orbital stabilization was active for the first 20 s. At the end of this period, the system trajectories reached a close neighborhood of the upright equilibrium rφ θ 9φ 9θs “ r0 ´ 3π{2 0 0s and the following linear controller was invoked for stabilization: τs “ 1.4φ´ 20.23pθ` 3π{2q` 1.14 9φ´ 1.98 9θ The poles of the closed-loop system were located at ´37.0˘ 20.0 i and ´1.0˘ 1.2 i. 1.0 0.0 -1.0 1.57 -1.57 -4.71 0.3 -0.3 0 φ (rad) θ (rad) (a) (c) (e) 1.0 0.0 -1.0 1.5 0.0 -1.5 0 9φ (rad/s) 9θ (rad/s) (b) (d) (f) τ (Nm) t (s) -15 0 35 9V t (s) 35 Figure 5.4: Rotary pendulum experimental results: (a)-(d) are plots of joint angles and joint velocities, (e) control torque, and (f) derivative of Lyapunov-like function. The brake pulses are shown within plots (e) and (f), the peaks represent time intervals when the brakes were engaged. 97 0.5 0.0 -0.5 1.57 -1.57 -4.71 0 φ (rad) θ (rad) 2.0 0.0 -2.0 7.0 0.0 -7.0 9φ (rad/s) 9θ (rad/s) t (s) 20 0 t (s) 20 Figure 5.5: Rotary pendulum simulation results. The pulses shown on the top of Figs.5.4 (e) and (f) correspond to the time intervals when the brake was engaged (OFF) during orbital stabilization. The brake was disengaged (ON) when the condition | 9φ |ď µ was satisfied; the value of µ was chosen to be small, equal to 0.1 rad/s. The time intervals required for braking were very short (« 0.04 s, on average); this implies that the brakings were impulsive in nature. The effect of impulsive braking can be seen in Fig.5.4 (b) where 9φ jumps to almost zero value upon engagement of the brake on multiple occasions. It can be seen from Fig.5.4 (c) that the amplitude of the pendulum gradually increases and finally reaches a close neighborhood of the upright equilibrium configuration. The derivative of the Lyapunov-like function is shown in Fig.5.4 (f). It can be seen that 9V never becomes positive; this is because the brake is engaged every time when 9V is about to become positive7. Since 9V is always negative, V decreases monotonically and orbital stabilization is achieved. A plot of the motor torque is shown in Fig.5.4 (e). To minimize wear and tear of the brake, the commanded motor torque was set to zero when the brake was engaged. A video of this experiment can be found on the weblink: www.egr.msu.edu/~mukherji/RotaryPendulum.mp4 Simulation results for the same set of initial conditions and controller gains in (5.6.7) and (5.6.8) are presented in Fig.5.5. A comparison of Figs.5.4 and 5.5 indicate that the joint velocities 7When | 9φ|ď µ « 0, the brake is not engaged since 9V « 0 - see (5.4.7). 98 in experiments are lower than those in simulations - this can be attributed to the presence of friction and other dissipative forces. The amplitude of the active joint φ is larger in experiments than simulations - this is due to the fact that the controller has to overcome the dissipative losses and additional energy is added through larger amplitude of motion. As expected, the time needed for stabilization is less in simulations than experiments. Remark 9. For comparison, we considered the rotary pendulum example in [71]. Taking identical initial conditions and physical parameters of the system therein, we simulated our controller with the gains Kp “ 0.20, Kd “ 0.12, Ke “ 50, Kc “ 0.70 The gains were tuned such that the magnitude of the motor torque did not exceed 0.3 Nm. The system trajectories converged to the desired orbit in approx. 30 s. The controller in [71] took approx. 100 sec and the magnitude of the maximum torque was 8 Nm. Our controller performed well, both in terms of motor torque requirement and speed of convergence. This better performance, however, comes at the cost of additional brake hardware. 5.7 Proofs and Additional Discussions 5.7.1 Proof of Lemma 1 The proof of Lemma 1 is provided here for the general case where the underactuated system has m passive and n´ m active generalized coordinates, i.e. q1 P Rn´m, q2 P Rm and u P Rn´m. The Euler-Lagrange equation has the same form as in (5.3.4) with M11 P Rpn´mqˆpn´mq, M22 P Rmˆm, h1 P Rpn´mq, and h2 P Rm. The change in the total energy due to application of an impulsive input is equal to the change in the kinetic energy: 1 9q`T 2 1 ΔE “ 2„ 9q`T “ ` 9q`T Mpqq 9q` ´ 1 M11 9q` 1 2 Mpqq 9q´ 1` 1 M11 9q´ 9q´T 1 ´ 9q´T 1 M12 9q´ 2 1 M12 9q` 2 ´ 9q´T 1 2„ 9q`T 2 M22 9q` 2 ´ 9q´T 2 M22 9q´ 2 (A.1) 99 The impulse manifold, given in (5.3.9) for m “ 1, is 2 ´ M´1 9q` 2 “ 9q´ 22 MT 12p 9q` 1 ´ 9q´ 1 q Substitution of 9q` 2 from (A.2) into (A.1) yields (A.2) 1 2„ 9q`T 9q´T 2 M22 9q´ 1 2 ΔE “ ´ 1 M11 9q` 1 M11 9q´ 1 ´ 9q´T 2 ´ 9q´T 1 M12 9q´ 22 MT 1 1` 2” 9q´ 2 ´ M´1 1 M12” 9q´ 2 ` 9q`T 2 ´ M´1 12p 9q` 22 MT M22” 9q´ 1 qıT 1 ´ 9q´ 1 qı 1 ´ 9q´ 12p 9q` 2 ´ M´1 22 MT 12p 9q` 1 qı 1 ´ 9q´ Expanding, canceling, and regrouping the terms on the right-hand side of the above equation yields ΔE “ 9q`T 1 1 2 ”M11 ´ M12M´1 22 MT 12ı 9q` 1 ´ 1 2 9q´T 1 ”M11 ´ M12M´1 22 MT 12ı 9q´ 1 Similar to (5.3.7), Bpqq is defined for the general case as follows Bpqq “”M11 ´ M12M´1 22 MT 12ı´1 (A.3) (A.4) From the properties of the mass matrix Mpqq, it can be shown that Bpqq is well-defined; also, it is symmetric and positive-definite, i.e., Bpqq “ BTpqq ą 0. Substitution of (A.4) into (A.3) gives (5.3.10). (cid:3) 5.7.2 Proof of Lemma 2 Impulsive inputs result in no change in the generalized coordinates. Additionally, impulsive braking results in 9q` 1 “ 0. Therefore, from the definition of V in (5.3.12), ΔV for impulsive braking can be expressed as ΔV “ “ “ 1 1 2„KepE`´ Edesq2 ´ KepE´´ Edesq2 ´ 9q´T 2„KepE`` E´´ 2EdesqΔE ´ 9q´T 1 2„Kep2E`´ ΔE ´ 2EdesqΔE ´ 9q´T 1 1 Kd 1 Kd 9q´ 9q´ 1 1 Kd 9q´ 1 100 where ΔE is defined in (5.3.10). Substitution of ΔE from (5.3.11) in the equation above yields 1 1 2 1 2 1 4 1 4 9q´ 9q´T 1 Kd “ ´ 1 B´1 9q´ ΔV “ ´ 9q´T 1 B´1 9q´ 9q´T 1 B´1 9q´ 1 qKetE`´ Edes ` 1 u` 9q´T 1 uB´1 ` Kd 9q´ 2„p 9q´T 1 „KetE`´ Edes ` 1 „ 1 1 1* B´1 ` Kd ` KepE` ´ EdesqB´1 9q´ which is the same as in (5.3.13). Since B, defined in (A.4), is positive-definite, tKe 9q´T 1 uB´1 is positive-definite. Therefore, if”Kd ` KepE` ´ EdesqB´1pqqı is positive-definite, ΔV ď 0 and ΔV “ 0 iff 9q´ 4"Ke 9q´T 1 B´1 9q´ 1 B´1 9q´ 1 “ 0. “ ´ 9q´T (cid:3) 1 1 5.7.3 Proof of Lemma 3 Let q˚ 1 and u˚ denote the constant values of q1 and u, respectively. From the passivity property of underactuated Euler-Lagrange systems [103,109], it follows that 9E “ uT 9q1 “ 0. Let E˚ denote the constant value of E. Since 9q1 ” 0, using (5.3.4) and (5.3.5), we get the following (A.5a) (A.5b) (A.6) (A.7) (A.8) M12 :q2 `„BM12 Bq2  9q 2 2 ´ 1 2 9q Bq1 T Bq1T 2 `„BF 2„BM22 “ u˚ Bq2  9q 2„BM22 2 ` BF Bq2 “ 0 1 2 M22 :q2 ` Eliminating :q2 using (A.5a) and (A.5b), we get 1 ˜„BM12 Bq2 ´ 2 M22„BM22 Since E “ E˚ is constant and 9q1 “ 0, (5.2.2) gives Bq1 T 2„BM22 Bq2 ¸ 9q M12 ´ 2 2 ´ M12 M22 BF Bq2 `„BF Bq1T “ u˚ 2 2 “ 9q 2pE˚ ´ Fq M22 Substitution of 9q2 2 from (A.7) in (A.6) yields E˚ P1pq˚ 1, q2q` P2pq˚ 1, q2q “ u˚ 101 where P1pq˚ with respect to time results in 1, q2q, P2pq˚ 1, q2q are defined in assumption 5. Differentiation of the above equation dq2 rkP1pq˚ Using assumption 5, we claim q2 is constant. d 1, q2q` P2pq˚ 1, q2qs 9q2 “ 0 5.7.4 Nonoccurrence of Zeno Phenomenon Impulsive braking results in 9q1pt`q “ 0 and the time derivative of p5.4.7) yields: :Vpt`q “ ´ :qT 1 pt`qKc :q1pt`q ă 0 (cid:3) (A.9) since Kc is positive-definite and :q1pt`q ­“ 0. If ε denotes the time for the trajectories to re-enter Z1, we can write 9Vpt`` εq “ 9Vpt`q`ż t``ε t` :Vpτqdτ (A.10) 9Vpt`` εq “ 9Vpt`q “ 0, and :Vpt`q ­“ 0, :V must change sign in t P pt`,t`` εq. If ε is a Since, small number, then the continuity of :V in the interval rt`,t`` εs implies: Using (A.9) and (A.11), we get | :Vpt`q|“ Opεq } :q1pt`q} ďd | :Vpt`q| λminpKcq “ Opε1{2q (A.11) (A.12) where λminpKcq is the smallest eigenvalue of Kc. Integrating :q1 with respect to time and using (A.12), we get } 9q1pt`` εq} ďż t``ε t` } :q1pτq}dτ “ Opε3{2q (A.13) Let us now assume that Zeno phenomenon occurs. Then, the time interval between consecutive impulsive inputs, which is equal to ε, converges to zero and the trajectory of the system at time t “ pt`` εq lies in the set Z1, where 9q1pt` ` εq ­“ 0. Since ε3{2 converges to zero faster than ε converges to zero, (A.13) implies that 9q1pt`` εq also converges to zero faster than ε converges to zero. This implies that the trajectory of the system does not lie in Z1 at t “ pt`` εq, and, by contradiction, proves the nonoccurence of the Zeno phenomenon. 102 5.7.5 Well-posedness of Switching Times 5.7.5.1 Background Let ψpt, 0, x0q denote the solution of the continuous-time dynamics in (5.3.15a) for 0 ď t ď t´ If and when the trajectory reaches starting from the initial condition xp0q “ x0 at time t “ 0. Z at t1, the states are instantaneously transferred to xpt` resetting in (5.3.17). The trajectory xptq, t` The resetting times are well-posed if the following conditions are satisfied [101, pg.13]: 1 q` fdrxpt´ 2 , is then given by ψpt,t` 1 qs, according to the 1 , xpt` 1 qq, and so on. 1 ď t ď t´ 1 q “ xpt´ 1 1. If xptq P ¯Z zZ , then there exists ε ą 0 such that, for all 0 ă δ ă ε, ψpt ` δ,t, xptqq R Z 2. If xpt´ k q P BZ X Z , then there exists ε ą 0 such that, for all 0 ď δ ă ε, ψpt` k ` δ,t` k , xpt´ k q` fdrxpt´ k qsq R Z Condition (1) ensures that when the system trajectory reaches the closure of the set Z , the trajectory must be directed away from Z . Condition (2) ensures that when the trajectory intersects Z , it instantaneously exists Z . 5.7.5.2 Proof Consider the set Z1 defined in (5.4.3b). If xptq P ¯Z1zZ1, then 9q1ptq “ 0, which implies 9Vptq “ 0. Differentiating (5.4.7) with respect to time and substituting 9q1ptq“ 0, we get :Vptq“´ :qT 1 ptqKc :q1ptqă 9Vpt`εq ă 0. Using (5.4.8), rA` BusT Kc 9q1pt` 0. Since 9Vptq “ 0 and :Vptq ă 0, it follows lim εq ą 0, the trajectory of the system is directed away from Z1 after it reaches the closure of Z1. This implies condition (1) is satisfied for Z1. It can be shown that εÑ0` BZ1 X Z1 “ trApq, 9qq` BpqqusT Kc 9q1 “ 0, 9q1 ­“ 0u 103 k q P BZ1 X Z1, impulsive braking is applied. This results in 9q1pt` k q “ 0, which implies k q R Z1. Thus the trajectory instantaneously exists Z1 after intersecting Z1 and condition (2) If xpt´ xpt` is satisfied for Z1. Now consider the set Z2 fi tq2 “ ´3π{2, 9q2 ă 0u, which was defined after (5.3.17). If xptq P ¯Z2zZ2 i.e., xptq P tq2 “ ´3π{2, 9q2 “ 0u. Then, if lim εÑ0` 9q2pt ` εq ě 0, then it follows from the definition of Z2 that xpt ` εq R Z2. Alternatively, if lim εÑ0` 9q2pt ` εq ă 0, then coordinate resetting occurs and q2pt ` εq “ π{2 Ñ xpt ` εq R Z2. Thus condition (1) is satisfied for Z2. Condition (2) is trivially satisfied for Z2 since BZ2 X Z2 is an empty set. Similar to the set Z2, it can be shown that conditions (1) and (2) are satisfied for the set Z3 fi tq2 “ π{2, 9q2 ą 0u. Since conditions (1) and (2) are satisfied individually for Z1, Z2, Z3, they are satisfied for Z “ Z1 Y Z2 Y Z3. (cid:3) 5.7.6 Quasi-Continuous Dependence Property 5.7.6.1 Background Let spt, 0, x0q denote the solution of the hybrid dynamical system in (5.3.15a) and (5.3.15b) for t ě 0 starting from the initial condition xp0q “ x0 at time t “ 0. Let Tx0 fi r0,8qztt1px0q,t2px0q,¨¨¨u denote all times except the discrete instants at which impulsive inputs are applied or coordinate resetting occurs. We now state the following assumption [101, Assumption 2.1, pg.27]: Assumption 3. Consider the hybrid dynamical system in (5.3.15). For every x0 P D and for every ε ą 0 and t P Tx0, there exists δpε, x0,tq ą 0 such that if }x0 ´ y} ă δpε, x0,tq, y P D , then }spt, 0, x0q´ spt, 0, yq} ă ε. The above assumption is a generalization of the standard continuous dependence property for dynamical systems with continuous flows to dynamical systems with left-continuous flows. We now state a proposition that provides sufficient conditions for satisfying Assumption 3 [101, Proposition 2.1, pg.32]: 104 Proposition 1. Consider the hybrid dynamical system in (5.3.15) and suppose that conditions (1) and (2) in section 5.7.5.1 are satisfied; then the system in (5.3.15) satisfies Assumption 3 if for all x0 P D , 0 ď t1px0q ă 8, t1px0q is continuous, and limkÑ8 tkpx0q Ñ 8. 5.7.6.2 Proof Our domain D was defined in section 5.3.3 as the open set where q2 P pa, bq, a ă ´3π{2, b ą π{2. If q2p0q P pa,´3π{2s Y rπ{2, bq, we can reset q2 to satisfy (5.4.2) since q2 corresponds to a revolute joint. From (5.3.15) we know that x0 R Z . This implies that t1px0q ą 0. Furthermore, t1px0q is a continuous function of x0 since the system is described by continuous-time dynamics (ordinary differential equations) over the interval t P r0,t1q. Also, as shown in section 5.7.4, Zeno phenomenon does not occur. Thus limkÑ8 tkpx0q Ñ 8. (cid:3) 5.7.7 Verification of Assumption 2 for Tiptoebot and Rotary Pendulum Tiptoebot: It can be seen from (5.5.2) and (5.5.4) that all elements of the mass matrix are bounded and the potential energy is lower bounded. It can also be seen from (5.5.2) that M12 and M22 are only functions of q1. Therefore P1 in (A.8), defined earlier in assumption 5, is not a function of q2. Since F in (5.5.4) is a function of both q1 and q2, it can be easily shown that pdP2{dq2q ­“ 0. Thus assumption 5 holds. Rotary Pendulum: It can be seen from (5.6.2) and (5.6.4) that all elements of the mass matrix are bounded and the potential energy is lower bounded. It can also be seen from (5.6.2) and (5.6.4) that the mass matrix and potential energy are independent of q1 “ φ. Therefore d dθ rkP1pθq` P2pθqs d cosθp2E˚ ´ 3γ4 sinθq ­“ 0 dθ„γ3 γ2 “ Thus assumption 5 holds. 105 CHAPTER 6 ORBITAL STABILIZATION USING VIRTUAL HOLONOMIC CONSTRAINTS AND IMPULSE CONTROLLED POINCARÉ MAPS 6.1 Introduction The problem of orbital stabilization of underactuated mechanical systems with one passive degree-of-freedom (DOF) is revisited in this chapter. The system dynamics is presented in section 6.2 and the results in [62] are utilized to design a continuous controller that can enforce the VHC such that the resulting zero dynamics is Euler-Lagrange. A periodic orbit is selected on the constraint manifold and a method for orbital stabilization is presented in section 6.3. To stabilize the orbit, a Poincaré section is defined at a point on the orbit and the return map is linearized about the fixed point; this results in a 2n´1 dimensional discrete linear time-invariant (LTI) system. To control this system and stabilize the orbit, impulsive inputs are applied when the system trajectory crosses the Poincaré section. The controllability of the orbit can be verified by simply checking the controllability of the linear system. This is simpler than the approach in [3, 65, 66] where controllability is verified numerically along the orbit for most systems. Since the system is LTI, the control design involves constant gains that can be computed off-line. Compared to the methods proposed earlier [3, 63, 65, 66], where periodic Ricatti equations have to be solved, our method has lower computational cost and complexity. Since impulsive inputs are used to control the Poincaré map, the dynamics of the closed-loop system can be described by the Impulse Controlled Poincaré Map (ICPM). The simplicity and generality of the ICPM approach to orbital stabilization is demonstrated using the examples of the 2-DOF cart-pendulum in section 6.4 and the 3-DOF tiptoebot in section 6.5. 106 6.2 Problem Formulation 6.2.1 System Dynamics Consider an n DOF underactuated system with one passive DOF, where the passive DOF is a revolute joint. Let q, q fi“qT 1 q2‰T , denote the generalized coordinates, where q1 P Rn´1 and q2 P S1, S1 “ R modulo 2π, are the coordinates of the active and passive DOFs. The configuration space of the system is denoted by Qn, Qn P Rn´1ˆ S1. The Lagrangian of the system can be written as Lpq, 9qq “ 1 2 9qT Mpqq 9q` Fpqq In the equation above, Mpqq P Rnˆn denotes the symmetric, positive-definite mass matrix, parti- tioned as Mpqq “»—– M11pqq M12pqq MT 12pqq M22pqq fiffifl where M11 P Rpn´1qˆpn´1q, M22 P R and Fpqq is the potential energy of the system. The Euler- Lagrange equation of motion can be written as follows M11pqq :q1 ` M12pqq :q2 ` h1pq, 9qq “ u MT 12pqq :q1 ` M22pqq :q2 ` h2pq, 9qq “ 0 (6.2.1a) (6.2.1b) where u P Rn´1 is the control input, and rhT forces. In compact form, (6.2.1a) and (6.2.1b) can be rewritten as 1 , h2sT is the vector of Coriolis, centrifugal and gravity :q1 “ Apq, 9qq` Bpqqu :q2 “ Cpq, 9qq` Dpqqu where, Bpqq “”M11 ´p1{M22qM12 MT Dpqq “ ´p1{M22qMT 12ı´1 , Apq, 9qq “ p1{M22qBpqqrM12 h2 ´ h1M22s 12 Bpqq, Cpq, 9qq “ ´p1{M22q”MT 12 Apq, 9qq` h2ı Similar to [62], we make the following assumption: (6.2.2a) (6.2.2b) (6.2.3) 107 Assumption 4. For some ¯q fi r ¯qT are even with respect to ¯q, i.e., 1 , ¯q2sT P Qn, the mass matrix Mpqq and the potential energy Fpqq Mp ¯q` qq “ Mp ¯q´ qq, Fp ¯q` qq “ Fp ¯q´ qq 6.2.2 Imposing Virtual Holonomic Constraints (VHC) A holonomic constraint enforced by feedback is referred to as VHC. The current and the next subsection summarizes relevant results from [62]. For a wide class of mechanical systems, a comprehensive discussion on VHC can be found in [62], [85]. A VHC for (6.2.1) is described by the relation ρpqq “ 0 where, ρ : Qn Ñ Rn´1 is smooth and rankrJqpρqs “ n´1 for all q P ρ´1p0q. Here, Jqpρq is the Jacobian of ρ with respect to q. The VHC is stabilizable if there exists a smooth feedback ucpq, 9qq that asymptotically stabilizes the set C “ tpq, 9qq : ρpqq “ 0, Jqpρq 9q “ 0u (6.2.4) The set C , which is referred to as the constraint manifold, is controlled invariant [62]. For the system described by (6.2.1), the set C is the tangent bundle of ρ´1p0q, which is a closed embedded submanifold of Qn of codimension pn´ 1q [110, Def. 4.2]. An important goal of this work is to generate repetitive motion, which can be described by closed orbits. Consequently, ρ´1p0q must be a smooth and closed curve without any self-intersection. The VHC can be described as ρpqq “ q1 ´ Φpq2q “ 0 (6.2.5) where Φ : S1 Ñ Rn´1 is a smooth vector-valued function. The constraint manifold C in (6.2.4) can be expressed as: C “"pq, 9qq : q1 “ Φpq2q, 9q1 “„ BΦ Bq2 9q2* (6.2.6) It should be noted that since q2 P S1, Φpq2 ` 2πq “ Φpq2q and ρ´1p0q is closed. Following the notion of odd VHC [62], we make another assumption. 108 Assumption 5. For ¯q which satisfies Assumption 4, Φpq2q is odd with respect to ¯q2, i.e., To stabilize C , we investigate the dynamics of ρpqq; differentiating ρpqq twice with respect to Φp ¯q2 ` q2q “ ´Φp ¯q2 ´ q2q time, we get Substitution of :q1 and :q2 from (6.2.2a) and (6.2.2b) in (6.2.7) yields 2 2 2ff 9q Bq2 :q2 ´«B2Φ :ρ “ :q1 ´„ BΦ Bq2 2ff 9q :ρ “ A´«B2Φ Bq2C`„B´„ BΦ 2 ´„ BΦ Bq2 2ff 9q Bq2 D´1«´A`«B2Φ 2 `„ BΦ Bq2 Bq2 D u Bq2C´ kpρ´ kd 2 uc “„B´„ BΦ 2 (6.2.7) (6.2.8) (6.2.9) (6.2.10) 9ρff The following choice of linearizing control where kp and kd are positive definite matrices, results in :ρ` kd 9ρ` kp ρ “ 0 This implies that limtÑ8 ρptq Ñ 0 exponentially and uc in (6.2.9) stabilizes the VHC in (6.2.5). If the initial conditions are chosen such that ρp0q “ 9ρp0q “ 0, uc in (6.2.9) enforces the VHC and the constraint manifold C is controlled invariant. Remark 10. For uc in (6.2.9) to be well-defined, the matrix rB´pBΦ{Bq2qDs must be invertible. It can be shown that rB´pBΦ{Bq2qDs is invertible iff MT 12pBΦ{Bq2q` M22 ‰ 0. This is also a necessary and sufficient condition for the VHC in (6.2.5) to be regular; since the VHC in (6.2.5) is in parametric form, this is also a necessary and sufficient condition for C to be stabilizable - see proposition 3.2 of [62]. 6.2.3 Zero Dynamics and Periodic Orbits On the constraint manifold C , the dynamics of the system satisfies ρpqq ” 0; this implies q1 “ Φpq2q, 9q1 “„ BΦ 2ff 9q Bq2 9q2, :q1 “«B2Φ Bq2 2 2 `„ BΦ Bq2 :q2 (6.2.11) 109 Substitution of q1, 9q1 and :q1 from (6.2.11) in (6.2.1b) provides the zero dynamics, which can be expressed in the following form It was shown in [3, 62, 83, 84] that the equation above has an integral of motion of the form :q2 “ α1pq2q` α2pq2q 9q 2 2 (6.2.12) Epq2, 9q2q “ p1{2qMpq2q 9q Mpq2q “expˆ´2ż q2 0 α2pτqdτ˙ , Ppq2q “ ´ż q2 0 2 2 ` Ppq2q α1pτqMpτq dτ (6.2.13) where Mpq2q represents the mass and Ppq2q represents the potential energy of the reduced system in (6.2.12). Since both Assumption 4 and 5 are satisfied, the zero dynamics represents an Euler-Lagrange system with the Lagrangian1 equal to p1{2qMpq2q 9q2 2 ´ Ppq2q. Since Assumptions 1 and 2 hold, the zero dynamics in (6.2.12) is similar to the dynamics of a simple pendulum [62] and its qualitative properties can be described by the potential energy Ppq2q. Let Pmin and Pmax denote the minimum and maximum values of P. If an energy level set is denoted by Epq2, 9q2q “ c, then c P pPmin, Pmaxq corresponds to a periodic orbit where the sign of 9q2 changes periodically and c ą Pmax corresponds to an orbit where the sign of 9q2 does not change [62]. 6.2.4 Problem Statement Since the zero dynamics in (6.2.12) has an Euler-Lagrange structure, there cannot exist any non- trivial isolated periodic orbit - this follows from the Poincaré-Lyapunov-Liouville-Arnol’d theorem [62, 85, 111]. A direct implication of this theorem is that the reduced dynamics possesses a dense set of closed orbits that are stable, but not asymptotically stable. For a desired repetitive motion, the corresponding orbit must be stabilized. Consider the desired closed orbit Od, defined as follows: Od “ tq, 9q P C : Epq2, 9q2q “ cdu, cd ą Pmin (6.2.14) 1Assumptions 4 and 5 provide necessary and sufficient conditions for the reduced system to be Euler-Lagrange - the proof of this result can be found in [62, 85]. 110 Let x, x fi rqT , 9qTsT , denote the states of the system in (6.2.1). We define an ε-neighborhood of Od by Uε “ tx P Qn ˆ Rn : distpx, Odq ă εu distpx, Odq fi inf d }x´ y} yPO We now define stability of the orbit Od from [112]. Definition 1. The orbit Od in (6.2.14) is • stable, if for every ε ą 0, there is a δ ą 0 such that xp0q P Uδ ùñ xptq P Uε, @t ě 0. • asymptotically stable if it is stable and δ can be chosen such that limtÑ8 distpxptq, Odq “ 0. The control uc in (6.2.9) asymptotically stabilizes C but does not asymptotically stabilize Od. If q, 9q P Od, uc enforces the VHC and trajectories stay on Od; however, a perturbation of the states will cause the trajectories to converge to a different orbit on C . The objective of this work is to utilize the control uc in (6.2.9), which was designed to stabilize C , together with impulsive inputs to stabilize Od directly, that lies on C . 6.3 Main Result: Stabilization of Od 6.3.1 Poincaré Map The system in (6.2.1) with u “ uc defined in (6.2.9), has the state-space representation 9x “ fpxq (6.3.1) The stability characteristics of periodic orbits can be studied using Poincaré maps [113]. To this end, we define the Poincaré section Σ of Od as follows2: Σ “ tx P Qn ˆ Rn : q2 “ q˚ 2Since the periodic orbit Od can pass through q2 “ q˚ 2, 9q2 ě 0u 2 with both positive and negative velocity 9q2, the condition 9q2 ě 0 implicitly assumes that Σ intersects the periodic orbit Od at a single point. In the definition of Σ in (6.3.2), 9q2 ě 0 can be replaced with 9q2 ď 0 without any loss of generality. (6.3.2) 111 where q˚ 2 is a constant. Let z, z fi rqT 1 , 9qTsT P Rp2n´1q, denote the states of the system on Σ. The Poincaré map P : Σ Ñ Σ is obtained by following trajectories of z from one intersection with Σ to the next. Let tk, k “ 1, 2,¨¨¨ denote the time of the k-th intersection and zpkq “ zptkq. Then, zpk` 1q can be described with the help of the map P zpk` 1q “ Przpkqs (6.3.3) The point of intersection of Σ and Od is the fixed point of P denoted by z˚; it satisfies the following relation z˚ “ Ppz˚q (6.3.4) The stability characteristics of the orbit Od can be studied by investigating the stability properties of z˚, which is an equilibrium point of the discrete-time system in (6.3.3); this can be done by linearizing the map P about z˚. For zpkq “ z˚ ` ν, where }ν} is a small number, we can write zpk` 1q “ Ppz˚ ` νq “ Ppz˚q`r∇zPpzqsz“z˚ rzpkq´ z˚s` Op}ν}2q (6.3.5) Using Ppz˚q “ z˚ from (6.3.4) and neglecting higher-order terms in }ν}, the above equation can be written as epk` 1q “ A epkq epkq fi zpkq´ z˚, A fi r∇zPpzqsz“z˚ (6.3.6) The stability properties of z˚ is governed by the eigenvalues of A , which are referred to as the Floquet multipliers of Od [113]. If the Floquet multipliers lie inside the unit circle, Od is exponentially stable - see Theorem 7.3 of [112]. From our discussion in section 6.2.4 we know that the desired orbit Od is not asymptotically stable, i.e., not all eigenvalues of A lie inside the unit circle. To asymptotically stabilize the orbit, i.e., to asymptotically stabilize z˚, we design an impulsive controller in the next subsection. 112 6.3.2 Impulse Controlled Poincaré Map (ICPM) To asymptotically stabilize the desired orbit Od, our controller in (6.2.9) is modified as follows u “ uc ` uI (6.3.7) where uI is an impulsive input which is applied only when xptq P Σ. The dynamics of the system, with uI as the new input, can be written as M11pqq :q1 ` M12pqq :q2 ` ¯h1pq, 9qq “ uI MT 12pqq :q1 ` M22pqq :q2 ` h2pq, 9qq “ 0 (6.3.8a) (6.3.8b) where ¯h1 fi ph1 ´ ucq. Impulsive inputs cause discontinuous changes in the generalized velocities while there is no change in the generalized coordinates. On the Poincaré section Σ, the jump in velocities can be computed by integrating (6.3.8) as follows [86]: uI dt (6.3.9) »—– M11 M12 MT 12 M22 fiffifl »—– Δ 9q1 Δ 9q2 fiffifl “»—– I 0 fiffifl , I fiż Δt 0 In the above equation, Δt is the infinitesimal interval of time for which uI is active, I P Rn´1 is the impulse of the impulsive input, and Δ 9q1 and Δ 9q2 are defined as Δ 9q1 fi p 9q` 1 ´ 9q´ 1 q, Δ 9q2 fi p 9q` 2 ´ 9q´ 2 q (6.3.10) where 9q´ and 9q` are the velocities immediately before and after application of uI . Since the system is underactuated, the jump in the passive velocity 9q2 is dependent on the jumps in the active velocity 9q1; this relationship is described by the pn´1q dimensional impulse manifold [6,23], which can be obtained from (6.3.9): IM “ t 9q` 1 , 9q` 2 | Δ 9q2 “ ´p1{M22qMT 12 Δ 9q1u (6.3.11) Since impulsive inputs can cause the system states to move on Σ, we exploit this property to design a feedback law that asymptotically stabilizes z˚, i.e., asymptotically stabilizes Od. The control 113 input applied at tk is denoted by I pkq3. The dynamics of the impulse controlled system in (6.3.3) can be described by the map zpk` 1q “ Przpkq, I pkqs (6.3.12) where I pkq “ 0 if zpkq “ z˚ 4. By linearizing the above map about the fixed point z “ z˚ and I “ 0, we get epk` 1q “ A epkq` B I pkq A fi”∇zPpz, I qız“z˚, I “0 , B fi”∇I Ppz, I qız“z˚, I “0 (6.3.13) where A P Rp2n´1qˆp2n´1q and B P Rp2n´1qˆpn´1q can be obtained numerically. Since A is not Hurwitz (see discussion in the last sub-section), we make the following proposition to exponentially stabilize Od: Proposition 2. If the pair tA , Bu is stabilizable, the orbit Od can be exponentially stabilized locally using the discrete impulsive feedback I pkq “ K epkq (6.3.14) where the matrix K is chosen such that pA ` BK q is Hurwitz. Remark 11. The matrices A and B depend on the continuous controller in (6.2.9), and hence on the choice of controller gains kp and kd. Thus, the stabilizability of the pair tA , Bu depends on the choice of kp and kd. The above approach to stabilization, which we refer to as the impulse controlled Poincaré map (ICPM) approach, is explained with the help of the schematic in Fig.6.1. The desired orbit Od is shown in red and it intersects the Poincaré section Σ at the fixed point z˚. A trajectory starting from an arbitrary initial condition, shown by the point 1 , intersects Σ at 2 . The impulsive input in 3As long as Δt is sufficiently small, the effect of the impulsive input uI depends solely on the value of I - see (6.3.9). Thus I can be viewed as the control input. 4If the system trajectory is passing through z˚, the trajectory is evolving on Od, which lies on C . Since I “ 0, the invariance of Od is guaranteed solely by the continuous controller uc in (6.2.9). 114 Σ 3 4 z˚ 2 IM 1 O d Figure 6.1: Schematic of ICPM approach to orbital stabilization. (6.3.14) moves the configuration of the system from 2 to 3 along the impulse manifold IM. The point 3 lies on Σ since impulsive inputs cause no change in the position coordinates and q2 “ q˚ is maintained during the transition from 2 to 3 along IM. Furthermore, the condition 9q` 2 ě 0 is satisfied since the impulsive controller in (6.3.14) is designed using linearization and 3 lies in 2 some small neighborhood of z˚ on Σ. Hereafter, the altered system trajectory evolves under the continuous control uc and 4 denotes its next intersection with Σ. A series of ICPMs, similar to the map 2 Ñ 4 exponentially converge the intersection point of the trajectory on Σ to z˚. Remark 12. By stabilizing the constraint manifold C , the continuous controller keeps the system trajectory close to C . By stabilizing the fixed point, the impulsive controller works in tandem with the continuous controller to converge the system trajectory to Od, which lies on C . Although the impulsive control inputs perturb the system trajectory intermittently, the trajectory converges to Od on C due to the vanishing nature of the perturbations. While global exponential stability of C guarantees that C remains exponentially stable despite the perturbations, the magnitudes of the perturbations exponentially converge to zero as the intersection point of the trajectory with the Poincaré section converges to the fixed point. 115 Remark 13. In the ICPM approach, impulsive inputs create discontinuous jumps in the states of the system. Although the subsequent continuous-time dynamics remains unchanged, the discontinuous jumps in the states result in a change in the Poincaré map. This is different from the well-known OGY method of chaos control [114] where the continuous-time dynamics is altered by discretely changing system parameters on a Poincaré section. The OGY method has been utilized in control of dynamical systems described by Poincaré maps; examples include bipeds [115] and hopping robots [8]. 6.3.3 Implementation of Control Design 6.3.3.1 Numerical Computation of A and B matrices Let δi, i “ 1, 2,¨¨¨ , 2n´1, denote the i-th column of ε1Ip2n´1q, where ε1 is a small number and Ip2n´1q is the identity matrix of size p2n´1q. If Ai denotes the i-th column of A , then Ai can be numerically computed as follows: Ai “ 1 ε1 rPpz˚ ` δiq´ z˚s (6.3.15) Let Q P Rnˆpn´1q and S P Rpn´1q be defined as follows: Q fi»—– Ipn´1q 01ˆpn´1q fiffifl , S fi»—– 0pn´1qˆ1 Mpqq´1ηi fiffifl where 0iˆ j is a matrix of zeros of dimension iˆ j, and ηi, i “ 1, 2,¨¨¨ , n´1, denote the i-th column of ε2Q, where ε2 is a small number. If Bi denotes the i-th column of B, then Bi can be numerically computed as follows: Bi “ 1 ε2 rPpz` Sqsz“z˚ ´ z˚( The above expression has been obtained using (6.3.9). (6.3.16) Remark 14. Numerical computation of A and B matrices is sensitive to the choice of ε1 and ε2. While ε1 and ε2 should be small, excessively small values can lead to numerical errors. For 116 example, the eigenvalues of A may be found to lie inside the unit circle, which cannot be the case since orbits on C are not asymptotically stable. 6.3.3.2 Impulsive Input using High-Gain Feedback Impulsive inputs are Dirac-delta functions and cannot be realized in real physical systems. Using singular perturbation theory [116], it was shown that continuous-time approximation of impulsive inputs can be carried out using high-gain feedback [23]; this has allowed implementation of impulsive control in physical systems using standard hardware [6, 25]. To obtain the expression for the high-gain feedback, we substitute (6.3.14) in (6.3.9) to get Δ 9q1pkq “ B I pkq “ BK epkq (6.3.17) where B is defined in (6.2.3) and is evaluated at tk; Δ 9q1pkq is the jump in the active velocities generated by the input I pkq. From (6.3.10) and (6.3.17), the desired active joint velocities at tk is 9qdes 1 pkq “ 9q1pkq` BK epkq (6.3.18) where 9q1pkq “ 9q1ptkq. To reach the desired velocities in a very short period of time, we use the high-gain feedback [6] µ uhg “ B´1„ 1 which remains active for as along as } 9qdes 9qdes 1 h1 with ¯h1. Furthermore, Λ fi diagrλ1 λ2 numbers, and µ ą 0 is a small number. 6.4 Illustrative Example: Cart-Pendulum 1 pkq´ 9q1} ě ε3, where ε3 is a small number. In (6.3.19), is obtained from (6.3.18) and ¯A is obtained from the expression for A in (6.2.3) by replacing ¨¨¨ λn´1s, where λi, i “ 1, 2,¨¨¨ , n´1 are positive 1 pkq´ 9q1¯´ ¯A Λ´ 9qdes (6.3.19) 6.4.1 System Dynamics and VHC Consider the frictionless cart-pendulum system in Fig.6.2. The masses of the cart and pendulum are denoted by mc and mp, ℓ denotes the length of the pendulum, and g is the acceleration due to 117 mp θ ℓ g u x mc Figure 6.2: Inverted pendulum on a cart. gravity. The control input u is the horizontal force applied on the cart. The cart position is denoted by x, x P R, and the angular displacement of the pendulum, measured clock-clockwise with respect to the vertical, is denoted by θ, θ P S1. We consider physical parameters of the system to be the same as those in [3]: mp “ mc “ ℓ “ 1. With the following definition q “ rq1 q2sT “ rx θsT and the potential energy of the system, given by the equations of motion can be obtained as F “ cosθ »—– sinθ 9θ2 g sinθ fiffifl “»—– u 0 which is of the form in (6.2.1). The VHC in (6.2.5) is chosen as 2 cosθ cosθ 1 fiffifl »—– :x :θ fiffifl´»—– ρ “ x` 1.5 sinθ “ 0 (6.4.1) (6.4.2) (6.4.3) (6.4.4) fiffifl which is identical to that considered in [3]. It can be verified that the mass matrix in (6.4.3) and the choice of VHC in (6.4.4) satisfy Assumptions 4 and 5 for ¯q “ p0, 0q. For the VHC in (6.4.4) to be stabilizable, Remark 10 provides the following condition that needs to be satisfied: 1´ 1.5 cos 2 θ ‰ 0 ñ θ ­“ ˘0.61 rad (6.4.5) 118 Clearly, the VHC in (6.4.4) is not regular and the corresponding constraint manifold is therefore not stabilizable. Since our control design requires the VHC to be regular, we cannot use the VHC in (6.4.4) with θ P S1. However, by restricting the domain of θ to be p´0.61, 0.61q, it is possible to compare the performance of our control design with that presented in [3]. Through trial and error, the controller gains are chosen such that θ lies in the interval p´0.61, 0.61q. 6.4.2 Stabilization of VHC and Od The ICPM approach relies on stabilization of both the constraint manifold C , and the orbit Od on C . This is a distinctive difference between our approach and the approach in [3] where Od is stabilized without stabilizing C , i.e., without enforcing the VHC. With the objective of enforcing the VHC, the gains kp and kd in (6.2.9) are chosen as follows: We choose the desired orbit Od to pass through the point: kp “ 2, kd “ 1 px, θ, 9x, 9θq “ p0.0, 0.0, ´0.675, 0.450q (6.4.6) (6.4.7) which is approximately the desired orbit in [3] - see Fig.2 therein. For exponential stabilization of Od, we define the Poinacaré section Σ “ tx P Q 2 ˆ R 2 : θ “ 0, 9θ ě 0u (6.4.8) The states of the system on Σ are Since z˚ lies on Od, using (6.4.7) and (6.4.8) we get z “ r x 9x 9θsT z˚ “ r0.0 ´0.675 0.450sT 119 Using the values of ε1 “ 0.02 and ε2 “ 0.01, the matrices A and B in (6.3.15) and (6.3.16) are obtained as 0.115 0.435 0.600 ´0.510 ´0.640 ´2.465 ´0.145 1.325 0.215 , B “ fiffiffiffiffifl »————– ´0.06 1.80 ´1.09 fiffiffiffiffifl A “ »————– It can be verified that all eigenvalues of A do not lie inside the unit circle but the pair tA , Bu is controllable and satisfy Proposition 2. Using LQR design, the gain matrix K in (6.3.14) was obtained as K “ r 0.163 0.288 1.198 s (6.4.9) The eigenvalues of pA ` BK q are located at 0.13 and ´0.06˘ 0.48i; thus, the impulsive feedback in (6.3.14) exponentially stabilizes the desired orbit Od. 6.4.3 Simulation Results The initial configuration of the system is taken from [3]: rx θ 9x 9θs “ r0.1 0.4 ´0.1 ´0.2s For the controller gains in (6.4.6) and (6.4.9), simulation results for the ICPM are shown in Fig.6.3; ρ is plotted with time in Fig.6.3 (a) and the phase portrait of the pendulum is shown in Fig.6.3 (b). To show the convergence of system trajectories to Od, }epkq}2 is plotted with respect to k in Fig.6.3 (c). It can be seen from Fig.6.3 (a) that the system trajectories converge to the constraint manifold. To stabilize Od on C , the impulsive controller in (6.3.14) is implemented using the high-gain feedback in (6.3.19) with Λ “ 1 and µ “ 0.005. It can be seen from the phase portrait in Fig.6.3 (b) that the pendulum trajectory converges exponentially to Od, shown in red. The effect of discrete impulsive feedback can be seen in Fig.6.3 (b) where 9θ jumps when trajectories cross the Poincaré section Σ defined in (6.4.8). The system trajectories reach a close neighborhood of Od in approximately 10 sec; this is comparable to the results in [3]. We now consider the following initial 120 (b) 9θ vs θ (a) ρ 2.0 0.0 -2.0 time (s) 10 -0.5 0.0 0.5 }epkq}2 (c) k 5 0.6 0.0 -0.6 0 3.5 0.0 0 Figure 6.3: Orbital stabilization for the cart-pendulum system for the initial conditions in [3]: (a) plot of ρ with respect to time, (b) phase portrait of the pendulum, (c) plot of the norm of the error of the discrete-time system. condition that lies far away from Od: rx θ 9x 9θs “ r0.0 0.0 0.0 0.0s (6.4.10) We used the same controller gains as that used in the previous simulation. It can be seen from the results in Fig.6.4 that Od is asymptotically stabilized. For the same initial conditions in (6.4.10), the control design in [3] fails to converge the pendulum trajectory to Od. Several other initial conditions yielded similar results. While this may be indicative of a larger region of attraction of Od with the ICPM approach than with the approach in [3], the region of attraction has to be estimated for both designs and compared before any conclusion can be drawn. To demonstrate the generality of the ICPM approach, we consider the three DOF tiptoebot, which is presented next. 121 (a) ρ time (s) }epkq}2 0.09 0.00 0 0.1 0.0 0 0.5 0.0 (b) 9θ vs θ -0.5 -0.15 15 (c) 0.0 0.15 k 5 Figure 6.4: Orbital stabilization for the cart-pendulum system for the initial conditions in (6.4.10): (a) plot of ρ with respect to time, (b) phase portrait of the pendulum; Od is shown in red, (c) plot of the norm of the error of the discrete-time system. 6.5 Illustrative Example - The Tiptoebot 6.5.1 System Description Consider the three DOF tiptoebot [6]. Using the following definition for the joint angles and control inputs 1 “ r θ2 θ3 sT , qT q2 “ θ1, u “ rτ2 τ3sT (6.5.1) Table 6.1: Tiptoebot lumped parameters in SI units α1 0.386 α4 0.065 β1 4.307 α2 0.217 α5 0.054 β2 1.102 α3 0.247 α6 0.104 β3 1.764 122 where θ1 P S1, θ2,θ3 P R, the dynamics of the tiptoebot can be expressed in the form given in (2.2.1), where the components of the mass matrix and the potential energy are: α2`α3`2α5 cosθ3 α3`α5 cosθ3 α3`α5 cosθ3 α3 fiffifl M11 “»—– M12 “»—– α2`α3`α4 cosθ2`2α5 cosθ3`α6 cospθ2`θ3q (6.5.2) α3`α5 cosθ3`α6 cospθ2`θ3q fiffifl M22 “ α1`α2`α3 ` 2rα4 cosθ2 ` α5 cosθ3`α6 cospθ2`θ3qs F “ β1 cosθ1 ` β2 cospθ1 ` θ2q` β3 cospθ1 ` θ2 ` θ3q where αi, i “ 1, 2,¨¨¨ , 6, and βi, i “ 1, 2, 3 are lumped physical parameters; their values are given in Table 6.1. It can be verified that Assumption 4 is satisfied for ¯q “ p0 0 0qT . 6.5.2 Imposing VHC and Selection of Od The VHC in (6.2.5) is chosen as ρ “»—– ρ1 ρ2 fiffifl “»—– θ2 ` 2.0θ1 θ3 ´ 0.1θ1 fiffifl “»—– 0 0 fiffifl kp “»—– 1.0 0.0 0.0 1.0 fiffifl , kd “»—– 0.1 0.0 0.0 0.1 fiffifl (6.5.3) (6.5.4) which satisfies Assumption 5 for ¯q “ p0 0 0qT . Also, it is stabilizable as it satisfies the condition in Remark 10. With the objective of enforcing the VHC, the gain matrices in (6.2.9) were chosen as The phase portrait of the zero dynamics in (6.2.12) is shown in Fig.6.5. It can be seen that the equilibrium pθ1, 9θ1q “ p0, 0q is a center, surrounded by a dense set of closed orbits. We choose the desired orbit Od to be the one that passes through pθ1, 9θ1q “ p0.0, 3.0q. 123 5 ) s / d a r ( 1 9θ 0 -5 -3 O d 0 θ1 (rad) 3 Figure 6.5: Phase portrait of tiptoebot zero dynamics. The desired orbit Od is shown in red. 6.5.3 Stabilization of Od We define the Poincaré section of Od as follows Σ “ tx P Q 3 ˆ R 3 : θ1 “ 0, 9θ1 ě 0u (6.5.5) The states on Σ are z “ rθ2 θ3 9θ1 9θ2 9θ3sT Substituting pθ1, 9θ1q “ p0.0, 3.0q in (6.5.3) and its derivative gives z˚ “ r0.0 0.0 3.0 ´6.0 0.3sT (6.5.6) Using the values of ε1 “ 0.01 and ε2 “ 0.004, the matrices A and B in (6.3.15) and (6.3.16) were obtained as 0.800 0.050 1.530 ´0.380 ´0.080 0.000 ´0.460 ´0.080 ´0.003 2.770 1.230 1.890 6.120 0.730 4.050 ´3.210 ´3.770 ´13.360 ´6.090 ´8.100 0.120 ´0.560 0.100 0.280 0.670 1.525 ´3.700 ´17.700 4.875 ´8.650 22.650 ´43.850 ´0.325 34.325 0.875 124 A “ »——————————– B “»—– fiffiffiffiffiffiffiffiffiffiffifl fiffifl T All eigenvalues of A do not lie inside the unit circle but the pair tA , Bu is controllable and satisfy Proposition 2. Using LQR, the gain matrix K in (6.3.14) is obtained as K “»—– fiffifl 0.028 0.024 0.197 0.094 0.138 ´0.034 ´0.051 0.116 ´0.049 ´0.055 (6.5.7) The eigenvalues of pA ` BK q are located at 0.14, ´0.47˘ 0.73i and ´0.12˘ 0.56i; thus Od is exponentially stable under the impulsive feedback. 6.5.4 Simulation Results The initial configuration of the tiptoebot is taken as rθ1 θ2 θ3 9θ1 9θ2 9θ3s “ r´0.1 0.2 0.05 3.3 ´6.0 0.4s For the controller gains in (6.5.4) and (6.5.7), simulation results of the ICPM approach are shown in Fig.6.6. The plots of ρ1, ρ2, 9ρ1 and 9ρ2 with time are shown in Figs.6.6 (a)-(d); it can be seen that the continuous controller uc in (6.2.9) stabilizes the constraint manifold C . The phase portrait of the passive joint θ1 is shown in Fig.6.6 (e) for 0 ď t ď 20 s and Fig.6.6 (f) for t ą 20 s. To asymptotically stabilize the desired orbit Od, shown in red in both Figs.6.6 (e) and (f), the impulsive controller in (6.3.14) is implemented using the high-gain feedback in (6.3.19); Λ was chosen to be an identity matrix and µ was chosen as 0.0001. To show the convergence of system trajectories to Od, }epkq}2 is plotted with respect to k in Fig.6.6 (e). It can be seen that for large values of k, }epkq}2 Ñ 0; this implies that Od is exponentially stable. 125 0.6 0.0 -0.6 0.5 0.0 -0.5 0 (a) (c) ρ1 9ρ1 0.45 0.00 -0.45 0.5 0.0 (b) (d) ρ2 9ρ2 -0.5 0 100 time (s) (e) time (s) (f) 100 4 t ď 20 t ą 20 1 9θ -4 -2 3.5 0.0 0 θ1 2 -2 (g) θ1 2 }epkq}2 k 40 Figure 6.6: Orbital stabilization for the tiptoebot using ICPM: (a) and (b) provide plots of ρ1 and ρ2, (c) and (d) provide plots of 9ρ1 and 9ρ2, (e) and (f) provide the phase portrait of the passive joint, (g) provides the norm of the error for the discrete-time system. 126 CHAPTER 7 CONCLUSION AND FUTURE WORK In this work we investigated several problems where impulsive inputs were exploited to stabilize equlibria and orbits of underactuated robotic systems. In chapter 2, using both simulations and experiments, we demonstrated a method for stabilizing the equilibria of underactuated systems from configurations lying outside the estimate of their region of attraction. The method, known as the IMM, uses impulsive inputs to force the configuration of the system to move inside the region of attraction along a manifold of dimension equal to the number of active degrees-of- freedom. Justified by singular perturbation theory [23], the impulsive inputs were implemented using high-gain feedback. In the general case, the region of attraction is not known and the IMM requires an estimate of the region, which is computed off-line. If the initial configuration of the system is outside this estimate, the IMM can be used if the impulse manifold intersects this region. The likelihood of intersection improves if the size of the estimated region is large, and therefore, a procedure for obtaining large estimates was developed by combining the SOS and trajectory reversing methods. For a fixed order of the stabilizing controller, the SOS method maximizes the estimate of the region of attraction and this estimate is further enlarged using an algorithm (CHART) based on the method of trajectory reversing. The advantage of combining the two methods was demonstrated in simulations using a three-link underactuated system (Tiptoebot) and in experiments with a pendubot. For both systems, it was shown that the impulse manifold may intersect the enlarged estimate but not the estimate obtained using the SOS method, implying that the CHART enhances the utility of the IMM. In another simulation with the Tiptoebot, it was shown that the combination of SOS, CHART and IMM requires a lower magnitude of the impulsive input compared to the combination of SOS and IMM. In addition to providing a large estimate, the CHART makes it possible to generalize the results to higher-dimensional systems. In the literature, several algorithms have been proposed for estimating the region of attraction using trajectory reversing, but the CHART is the only algorithm that has been demonstrated for a system with as 127 many as six states. Earlier experimental results [23] relied on finding the boundary of the region of attraction through trial and error and used the IMM to move the configuration of the system from a point lying immediately outside the region to inside the region. The experimental results presented here are more meaningful and practical as they are based on an estimate of the region of attraction and points chosen outside this region are not very close to the boundary. Experimental results of stabilization from two different initial conditions using the SOS, CHART and IMM were presented. In one of the experiments, through trial and error, the initial condition was purposely chosen to lie outside the region of attraction; it was shown that the equilibrium could not be stabilized with the SOS controller alone but was stabilized by the combination of SOS, CHART and IMM. Future work will investigate the possibility of extending the approach to multiple impulsive inputs. This will be useful in situations where the impulse manifold does not intersect the estimated region of attraction: a first impulsive input may be used to take the configuration closer to the boundary of the estimate and additional impulses may be applied to move the system configuration inside the region in future time. Of course, this would require real-time computation of the intersection of the impulse manifold with the boundary of the estimated region. Compared to the SOS method, which provides an analytical representation of the estimate of the region of attraction, the trajectory reversing method provides a numerical representation and will involve higher computational cost. Considering the fact that the trajectory reversing method provides the opportunity to enlarge the estimate, it is necessary to investigate the trade-off between computational cost and size of the estimate for real-time applications. In chapter 3, rest-to-rest maneuvers of the inertia-wheel pendulum was studied in the framework of impulsive control. Assuming a set of discrete impulsive inputs, optimal sequences were designed to minimize their infinity-norm. It was shown analytically that a sequence with an odd number of inputs is less optimal than the two adjacent sequences with even number of inputs. Analytical and simulation results with two inputs were used to explain the high wheel velocities and large continuous torques associated with methods that attempt to take the pendulum directly to its desired configuration and minimize overshoot. Implementation of impulsive control using high- 128 gain feedback also results in large torques but they act over short intervals of time; therefore, feasibility of impulsive control is determined by the peak torque rating of the actuator, which is always larger than the continuous torque rating. It was shown that the number of impulsive inputs can be increased to not exceed the peak torque rating of the actuator; this, of-course, increases the time required for swing-up. Simulation results for swing-up showed similarities between the optimal trajectories and the trajectories obtained using the energy-based controllers. Future work will investigate the possibility of extending the method presented here to other underactuated systems. Using impulsive forces only, the problem of juggling a devil-stick was presented in chapter 4. Impulsive forces were applied intermittently for juggling the stick between two symmetric configurations. A dynamic model of the devil-stick and a control design for the juggling task was presented for the first time. The control inputs are the impulse of the impulsive force and its point of application on the stick. The control action is event-based and the inputs are applied only when the stick has the orientation of one of the two symmetric configurations. The dynamics of the devilstick due to the control action and torque-free motion under gravity is described by two Poincaré sections; the symmetric configurations are fixed points of these sections. A coordinate transformation is used to exploit the symmetry and convert the problem into that of stabilization of a single fixed point. A dead-beat controller is designed to convert the nonlinear system into a controllable linear discrete-time system with input constraints. LQR and MPC methods are used to design the control inputs and achieve symmetric juggling. The LQR method has a closed-form solution and is easier to implement but requires trial and error to satisfy the input constraints. The MPC method has no closed-form solution as it is obtained by solving an optimization problem online. However, the optimization problem directly takes into account the input constraint. The computational cost of the MPC method, which can be a concern for many problems, is not a concern for the juggling problem since the time between consecutive control actions is relatively large.Simulation results validate both control designs and demonstrate non-prehensile manipulation solely using impulsive forces. Our future work will focus on robotic juggling; this includes design of experimental hardware, 129 feedback compensation of energy losses due to inelastic collisions between the devil-stick and hand sticks, and motion planning and control of the robot end-effector for generating the impulsive forces designed by the control algorithms. In chapter 5, a hybrid control strategy was presented for orbital stabilization of underactuated systems with one passive DOF. The orbit is defined with the help of a Lyapunov-like function that has been commonly used in the literature. Unlike existing energy-based methods, that have relied on continuous control inputs alone, our control strategy uses continuous control inputs and intermittent impulsive brakings. The continuous control is designed to make the time derivative of the Lyapunov-like function negative semi-definite. When this condition cannot be enforced, the impulsive inputs are invoked. This results in negative jumps in the Lyapunov-like function and guarantees its negative semi-definiteness under continuous control for some finite time interval. Thus, a combination of continuous and impulsive inputs guarantees monotonic convergence of the system trajectories to the desired orbit, which can be periodic, or non-periodic as in the case of homoclinic orbits, depending on the choice of desired energy. More importantly, it allows us to develop a general framework for energy-based orbital stabilization, which is an important contribution of this work. A set of conditions, that impose constraints on the choice of controller gains, have to be satisfied for applicability of the control strategy. These conditions are easily satisfied by systems commonly studied in the literature such as the pendubot, acrobot, inertia-wheel pendulum, and pendulum on a cart. In this work, the hybrid control strategy was demonstrated in a three-DOF underactuated system using simulations and the two-DOF rotary pendulum using experiments. In experiments, impulsive brakings were not applied by the motor; instead, they were applied by a friction brake mounted co-axially with the motor shaft. This requires additional hardware but there are two important advantages of using the brake. In physical systems, impulsive inputs are implemented using high-gain feedback, which can result in actuator saturation. Since our impulsive control strategy requires the active velocities to be reduced to zero, a brake is a natural choice and it eliminates the possibility of motor torque saturation. The advantage of using a brake is also manifested in the time required for orbital stabilization. A comparison of our approach with 130 an approach in the literature shows significant reduction in the time for convergence for the same set of initial conditions. Finally, the control design for stabilization of VHC based orbits was presented in chapter 6. Repetitive motion in underactuated systems are typically designed using VHCs. A VHC results in a family of periodic orbits and stabilization of a desired orbit is an important problem. A hybrid control design is presented to stabilize a VHC-generated periodic orbit for underactuated systems with one passive DOF. A continuous controller is first designed to enforce the VHC and stabilize the corresponding constraint manifold. The desired orbit on the constraint manifold is exponentially stabilized by applying the continuous inputs together with impulsive inputs that are periodically applied on a Poincaré section. The impulsive inputs alter the Poincaré map and this impulse controlled Poincaré map (ICPM) is described by a discrete LTI system. The problem of orbital stabilization is thus simplified to exponential stabilization of the fixed point of the ICPM. The controllability of the system can be easily verified and the control design can be easily carried out using standard techniques such as pole-placement and LQR. The identification of the linear system and computation of the controller gains are performed off-line. The complexity and computational cost of the ICPM approach is less than existing methods as it eliminates the need for on-line solution of a periodic Ricatti equation. The ICPM approach is demonstrated using standard cart-pendulum system; its applicability to higher-dimensional systems is demonstrated using the tiptoebot. Future work will focus on gait stabilization of legged robots and experimental validation. 131 BIBLIOGRAPHY 132 BIBLIOGRAPHY [1] R. Olfati-Saber, Global stabilization of a flat underactuated system: the inertia wheel pen- dulum, in: Proc. IEEE Conference on Decision and Control, Vol. 4, 2001, pp. 3764–3765. [2] M. W. Spong, P. Corke, R. Lozano, Nonlinear control of the reaction wheel pendulum, Automatica 37 (11) (2001) 1845–1851. [3] A. Shiriaev, J. W. Perram, C. Canudas-de Wit, Constructive tool for orbital stabilization of underactuated nonlinear systems: Virtual constraints approach, IEEE Transactions on Automatic Control 50 (8) (2005) 1164–1176. [4] M. W. Spong, D. J. Block, The pendubot: A mechatronic system for control research and education, in: Decision and Control, 1995., Proceedings of the 34th IEEE Conference on, Vol. 1, IEEE, 1995, pp. 555–556. [5] M. W. Spong, The swing up control problem for the acrobot, IEEE Control Systems 15 (1) (1995) 49–55. [6] [7] [8] [9] N. Kant, R. Mukherjee, D. Chowdhury, H. K. Khalil, Estimation of the region of attraction of underactuated systems and its enlargement using impulsive inputs, IEEE Transactions on Robotics 35 (3) (2019) 618–632. F. Plestan, J. W. Grizzle, E. R. Westervelt, G. Abba, Stable walking of a 7-dof biped robot, IEEE Transactions on Robotics and Automation 19 (4) (2003) 653–668. F. B. Mathis, R. Mukherjee, Apex height control of a two-mass robot hopping on a rigid foundation, Mechanism and Machine Theory 105 (2016) 44–57. R. Ronsse, P. Lefevre, R. Sepulchre, Rhythmic feedback control of a blind planar juggler, IEEE Transactions on Robotics 23 (4) (2007) 790–802. [10] R. Goebel, R. G. Sanfelice, A. R. Teel, Hybrid dynamical systems, IEEE Control Systems 29 (2) (2009) 28–93. [11] Y. Aoustin, D. T. Romero, C. Chevallereau, S. Aubin, Impulsive control for a thirteen- link biped, in: 9th IEEE International Workshop on Advanced Motion Control, 2006, pp. 439–444. [12] S. Weibel, G. W. Howell, J. Baillieul, Control of single-degree-of-freedom Hamiltonian systems with impulsive inputs, in: Proc. 35th IEEE Conference on Decision and Control, Vol. 4, 1996, pp. 4661–4666. [13] A. M. Bloch, N. E. Leonard, J. E. Marsden, Controlled Lagrangians and the stabilization of mechanical systems. i. the first matching theorem, IEEE Transactions on Automatic Control 45 (12) (2000) 2253–2270. 133 [14] A. M. Bloch, D. E. Chang, N. E. Leonard, J. E. Marsden, Controlled Lagrangians and the stabilization of mechanical systems. ii. potential shaping, IEEE Transactions on Automatic Control 46 (10) (2001) 1556–1571. [15] R. Ortega, M. W. Spong, F. Gómez-Estern, G. Blankenstein, Stabilization of a class of un- deractuated mechanical systems via interconnection and damping assignment, IEEE Trans- actions on Automatic Control 47 (8) (2002) 1218–1233. [16] D. Auckly, L. Kapitanski, W. White, Control of nonlinear underactuated systems, Commu- nications on Pure and Applied Mathematics 53 (3) (2000) 354–369. [17] I. Fantoni, R. Lozano, M. W. Spong, Energy based control of the pendubot, IEEE Transactions on Automatic Control 45 (4) (2000) 725–729. [18] M. Zhang, T.-J. Tarn, Hybrid control of the pendubot, IEEE/ASME Transactions on Mecha- tronics 7 (1) (2002) 79–86. [19] T. Albahkali, R. Mukherjee, T. Das, Swing-up control of the pendubot: An impulse- momentum approach, IEEE Transactions on Robotics 25 (4) (2009) 975–982. [20] R. Jafari, F. B. Mathis, R. Mukherjee, Swing-up control of the acrobot: An impulse- momentum approach, in: Proc. American Control Conference, 2011, pp. 262–267. [21] J. Lee, R. Mukherjee, H. K. Khalil, Output feedback stabilization of inverted pendulum on a cart in the presence of uncertainties, Automatica 54 (2015) 146–157. [22] F. Andreev, D. Auckly, S. Gosavi, L. Kapitanski, A. Kelkar, W. White, Matching, linear systems, and the ball and beam, Automatica 38 (12) (2002) 2147 – 2152. [23] R. Jafari, F. B. Mathis, R. Mukherjee, H. Khalil, Enlarging the region of attraction of equilibria of underactuated systems using impulsive inputs, IEEE Transactions on Control Systems Technology 24 (1) (2016) 334–340. [24] F. B. Mathis, R. Jafari, R. Mukherjee, Efficient swing-up of the acrobot using continuous torque and impulsive braking, in: Proc. American Control Conference, 2011, pp. 268–273. [25] F. B. Mathis, R. Jafari, R. Mukherjee, Impulsive actuation in robot manipulators: Experi- mental verification of pendubot swing-up, IEEE/ASME Transactions on Mechatronics 19 (4) (2014) 1469–1474. [26] D. Davidson, S. A. Bortoff, Enlarge your region of attraction using high-gain feedback, in: Proc. IEEE Conference on Decision and Control, 1994, pp. 634–639. [27] C. Reboulet, C. Champetier, A new method for linearizing non-linear systems: the pseudo- linearization, International Journal of Control 40 (4) (1984) 631–638. [28] T. Hu, Z. Lin, Composite quadratic Lyapunov functions for constrained control systems, IEEE Transactions on Automatic Control 48 (3) (2003) 440–450. 134 [29] R. Bakhtiari, M. Yazdanpanah, Designing a linear controller for polynomial systems with the largest domain of attraction via LMIs, in: Proc. IEEE International Conference on Control and Automation, Vol. 1, 2005, pp. 449–453. [30] A. I. Zečević, D. D. Šiljak, Estimating the region of attraction for large-scale systems with uncertainties, Automatica 46 (2) (2010) 445–451. [31] P. A. Parrilo, Structured semidefinite programs and semialgebraic geometry methods in robustness and optimization, Ph.D. thesis, California Institute of Technology (2000). [32] W. Tan, A. Packard, Stability region analysis using polynomial and composite polynomial Lyapunov functions and sum-of-squares programming, IEEE Transactions on Automatic Control 53 (2) (2008) 565–571. [33] U. Topcu, A. Packard, P. Seiler, Local stability analysis using simulations and sum-of-squares programming, Automatica 44 (10) (2008) 2669–2675. [34] U. Topcu, A. K. Packard, P. Seiler, G. J. Balas, Robust region-of-attraction estimation, IEEE Transactions on Automatic Control 55 (1) (2010) 137–142. [35] A. Majumdar, A. A. Ahmadi, R. Tedrake, Control design along trajectories with sums of squares programming, in: Proc. IEEE International Conference on Robotics and Automation, 2013, pp. 4054–4061. [36] A. Majumdar, R. Tedrake, Funnel libraries for real-time robust feedback motion planning, The International Journal of Robotics Research 36 (8) (2017). [37] R. Genesio, A. Vicino, New techniques for constructing asymptotic stability regions for nonlinear systems, IEEE Transactions on Circuits and Systems 31 (6) (1984) 574–581. [38] R. Genesio, M. Tartaglia, A. Vicino, On the estimation of asymptotic stability regions: State of the art and new proposals, IEEE Transactions on Automatic Control 30 (8) (1985) 747–755. [39] M. Loccufier, E. Noldus, A new trajectory reversing method for estimating stability regions of autonomous nonlinear systems, Nonlinear dynamics 21 (3) (2000) 265–288. [40] F. Hamidi, H. Jerbi, W. Aggoune, M. Djemai, M. N. Abdkrim, Enlarging region of attraction via LMI-based approach and genetic algorithm, in: Proc. IEEE International Conference on Communications, Computing and Control Applications, 2011, pp. 1–6. [41] D. Chowdhury, N. Kant, R. Mukherjee, H. K. Khalil, Enlarging the region of attraction of equilibria of underactuated systems using sum of squares and impulse manifold method, in: Proc. American Control Conference, 2017, pp. 893–898. [42] N. Kant, D. Chowdhury, R. Mukherjee, H. K. Khalil, An algorithm for enlarging the region of attraction using trajectory reversing, in: Proc. American Control Conference, 2017, pp. 4171–4176. 135 [43] M. Ryalat, D. S. Laila, A simplified ida-pbc design for underactuated mechanical systems with applications, European Journal of Control 27 (2016) 1–16. [44] N. Qaiser, N. Iqbal, A. Hussain, Stabilization of non-linear inertia wheel pendulum system using a new dynamic surface control based technique, in: IEEE International Conference on Engineering of Intelligent Systems, 2006, pp. 1–6. [45] A. Khoroshun, Stabilization of the upper equilibrium position of a pendulum by spinning an inertial flywheel, International applied mechanics 52 (5) (2016) 547–556. [46] A. Zhang, C. Yang, S. Gong, J. Qiu, Nonlinear stabilizing control of underactuated inertia wheel pendulum based on coordinate transformation and time-reverse strategy, Nonlinear Dynamics 84 (4) (2016) 246702476. [47] HNWpodcasts, Five easy devil stick https://www.youtube.com/watch?v=Z7Pv-p-nEYo, 2020] (2013). tricks [Online; with instructions, accessed 01-May- [48] N. Duinker, Devilstick tutorial - basic tricks to https://www.youtube.com/watch?v=MAuAtwZ7BF4, 2020] (2015). [Online; you get started, accessed 01-May- [49] H. Hirai, F. Miyazaki, Dynamic coordination between robots: Self-organized timing selec- tion in a juggling-like ball-passing task, IEEE Transactions on Systems, Man, and Cyber- netics, Part B (Cybernetics) 36 (4) (2006) 738–754. [50] J. Kober, M. Glisson, M. Mistry, Playing catch and juggling with a humanoid robot, in: 12th IEEE-RAS International Conference on Humanoid Robots (Humanoids 2012), IEEE, 2012, pp. 875–881. [51] A. Tornambe, Modeling and control of impact in mechanical systems: Theory and experi- mental results, IEEE Transactions on Automatic Control 44 (2) (1999) 294–309. [52] K. M. Lynch, C. K. Black, Recurrence, controllability, and stabilization of juggling, IEEE Transactions on Robotics and Automation 17 (2) (2001) 113–124. [53] A. Zavala-Rio, B. Brogliato, Direct adaptive control design for one-degree-of-freedom complementary-slackness jugglers, Automatica 37 (7) (2001) 1117–1123. [54] B. Brogliato, M. Mabrouk, A. Z. Rio, On the controllability of linear juggling mechanical systems, Systems & control letters 55 (4) (2006) 350–367. [55] R. Goebel, R. G. Sanfelice, A. R. Teel, Hybrid Dynamical Systems: modeling, stability, and robustness, Princeton University Press, 2012. [56] M. W. Spong, Impact controllability of an air hockey puck, Systems & Control Letters 42 (5) (2001) 333–345. [57] S. Schaal, C. G. Atkeson, Open loop stable control strategies for robot juggling, in: Proc. IEEE International Conference on Robotics and Automation, IEEE, 1993, pp. 913–918. 136 [58] R. G. Sanfelice, A. R. Teel, R. Sepulchre, A hybrid systems approach to trajectory tracking control for juggling systems, in: Decision and Control, 2007 46th IEEE Conference on, IEEE, 2007, pp. 5282–5287. [59] S. Nakaura, Y. Kawaida, T. Matsumoto, M. Sampei, Enduring rotary motion control of devil stick, IFAC Proceedings Volumes 37 (13) (2004) 805–810. [60] A. Shiriaev, A. Robertsson, L. Freidovich, R. Johansson, Generating stable propeller mo- tions for devil stick, in: 3rd IFAC Workshop on Lagrangian and Hamiltonian Methods for Nonlinear Control, 2006, 2006. [61] J. Zhao, M. W. Spong, Hybrid control for global stabilization of the cart–pendulum system, Automatica 37 (12) (2001) 1941–1951. [62] M. Maggiore, L. Consolini, Virtual holonomic constraints for Euler- Lagrange systems, IEEE Transactions on Automatic Control 58 (4) (2013) 1001–1008. [63] A. Mohammadi, M. Maggiore, L. Consolini, Dynamic virtual holonomic constraints for stabilization of closed orbits in underactuated mechanical systems, Automatica 94 (2018) 112–124. [64] E. R. Westervelt, C. Chevallereau, J. H. Choi, B. Morris, J. W. Grizzle, Feedback Control of Dynamic Bipedal Robot Locomotion, CRC press, 2007. [65] A. S. Shiriaev, L. B. Freidovich, A. Robertsson, R. Johansson, A. Sandberg, Virtual- holonomic-constraints-based design of stable oscillations of Furuta pendulum: Theory and experiments, IEEE Transactions on Robotics 23 (4) (2007) 827–832. [66] L. Freidovich, A. Robertsson, A. Shiriaev, R. Johansson, Periodic motions of the pendubot via virtual holonomic constraints: Theory and experiments, Automatica 44 (3) (2008) 785– 791. [67] X. Xin, T. Yamasaki, Energy-based swing-up control for a remotely driven acrobot: Theo- retical and experimental results, IEEE Transactions on Control Systems Technology 20 (4) (2012) 1048–1056. [68] I. Fantoni, R. Lozano, M. Spong, Stabilization of the reaction wheel pendulum using an energy approach, in: European Control Conference (ECC), IEEE, 2001, pp. 2552–2557. [69] R. Lozano, I. Fantoni, D. J. Block, Stabilization of the inverted pendulum around its homo- clinic orbit, Systems & Control Letters (2000) 197–204. [70] H. Oka, Y. Maruki, H. Suemitsu, T. Matsuo, Nonlinear control for rotational movement of cart-pendulum system using homoclinic orbit, International Journal of Control, Automation and Systems 14 (5) (2016) 1270–1279. [71] I. Fantoni, R. Lozano, Stabilization of the Furuta pendulum around its homoclinic orbit, International Journal of Control 75 (6) (2002) 390–398. 137 [72] X. Xin, M. Kaneda, Swing-up control for a 3-dof gymnastic robot with passive first joint: design and analysis, IEEE Transactions on Robotics 23 (6) (2007) 1277–1285. [73] H. K. Khalil, Nonlinear control, Prentice Hall, 2014. [74] N. Kant, R. Mukherjee, H. K. Khalil, Stabilization of homoclinic orbits of two degree-of- freedom underactuated systems, in: 2019 American Control Conference (ACC), 2019. [75] [76] J. W. Grizzle, C. Chevallereau, R. W. Sinnet, A. D. Ames, Models, feedback control, and open problems of 3d bipedal robotic walking, Automatica 50 (8) (2014) 1955–1988. J. W. Grizzle, G. Abba, F. Plestan, Asymptotically stable walking for biped robots: Analysis via systems with impulse effects, IEEE Transactions on automatic control 46 (1) (2001) 51–64. [77] C. Canudas-de Wit, On the concept of virtual constraints as a tool for walking robot control and balancing, Annual Reviews in Control 28 (2) (2004) 157–166. [78] C. Chevallereau, G. Abba, Y. Aoustin, F. Plestan, E. R. Westervelt, C. Canudas-De-Wit, J. W. Grizzle, RABBIT: A testbed for advanced control theory, IEEE Control Systems Magazine 23 (5) (2003) 57–79. [79] C. Chevallereau, J. W. Grizzle, C.-L. Shih, Asymptotically stable walking of a five-link underactuated 3-D bipedal robot, IEEE Trans. on Robotics 25 (1) (2009) 37–50. [80] E. R. Westervelt, J. W. Grizzle, D. E. Koditschek, Hybrid zero dynamics of planar biped walkers, IEEE Transactions on Automatic Control 48 (1) (2003) 42–56. [81] C. Canudas-de Wit, B. Espiau, C. Urrea, Orbital stabilization of underactuated mechanical systems, in: Proc. 15th IFAC World Congress, 2002. [82] A. Mohammadi, E. Rezapour, M. Maggiore, K. Y. Pettersen, Maneuvering control of planar snake robots using virtual holonomic constraints, IEEE Trans. on Control Systems Technol- ogy 24 (3) (2015) 884–899. [83] A. Shiriaev, A. Robertsson, J. Perram, A. Sandberg, Periodic motion planning for virtually constrained Euler-Lagrange systems, Systems & Control Letters 55 (11) (2006) 900–907. [84] A. S. Shiriaev, L. B. Freidovich, S. V. Gusev, Transverse linearization for controlled mechan- ical systems with several passive degrees of freedom, IEEE Trans. on Automatic Control 55 (4) (2010) 893–906. [85] A. Mohammadi, Virtual holonomic constraints for Euler-Lagrange control systems, Ph.D. thesis, University of Toronto (2016). [86] L. L. Flynn, R. Jafari, R. Mukherjee, Active synthetic-wheel biped with torso, IEEE Trans- actions on Robotics 26 (5) (2010) 816–826. [87] A. Majumdar, A. A. Ahmadi, R. Tedrake, Control and verification of high-dimensional systems with DSOS and SDSOS programming, in: Proc. 53rd IEEE Conference on Decision and Control, 2014, pp. 394–401. 138 [88] C. B. Barber, D. P. Dobkin, H. Huhdanpaa, The quickhull algorithm for convex hulls, ACM Transactions on Mathematical Software 22 (4) (1996) 469–483. [89] J. D’Errico, Inhull, mathworks.com Updated on 06 Sep 2012 (March 2006). [90] M. J, Analyze n-dimensional polyhedra in terms of vertices or (In)Equalities, mathworks.com Updated on 16 Sep 2015 (March 2011). [91] M. Ester, H.-P. Kriegel, J. Sander, X. Xu, et al., A density-based algorithm for discovering clusters in large spatial databases with noise, in: Kdd, Vol. 96, 1996, pp. 226–231. [92] Yarpiz, Dbscan clustering algorithm, mathworks.com 06 Sep 2015 (March 2011). [93] N. Kant, R. Mukherjee, H. K. Khalil, Swing-up of the inertia wheel pendulum using im- pulsive torques, in: 2017 IEEE 56th Annual Conference on Decision and Control (CDC), IEEE, 2017, pp. 5833–5838. [94] H. A. Toliyat, G. B. Kliman, Handbook of electric motors, Vol. 120, CRC press, 2004. [95] N. Kant, R. Mukherjee, Impulsive dynamics and control of the inertia-wheel pendulum, IEEE Robotics and Automation Letters 3 (4) (2018) 3208–3215. [96] S. Wiggins, Introduction to applied nonlinear dynamical systems and chaos, Vol. 2, Springer Science & Business Media, 2003. [97] J. Guckenheimer, K. Hoffman, W. Weckesser, The forced Van der Pol equation i: The slow flow and its bifurcations, SIAM Journal on Applied Dynamical Systems 2 (1) (2003) 1–35. [98] K. Bold, C. Edwards, J. Guckenheimer, S. Guharay, K. Hoffman, J. Hubbard, R. Oliva, W. Weckesser, The forced Van der Pol equation ii: Canards in the reduced system, SIAM Journal on Applied Dynamical Systems 2 (4) (2003) 570–608. [99] P. J. Antsaklis, A. N. Michel, A Linear Systems Primer, Birkhauser, 2007. [100] L. Wang, Model predictive control system design and implementation using MATLAB®, Springer Science & Business Media, 2009. [101] W. M. Haddad, V. Chellaboina, S. G. Nersesov, Impulsive and hybrid dynamical systems, Princeton Series in Applied Mathematics (2006). [102] B. Brogliato, R. Lozano, B. Maschke, O. Egeland, Dissipative Systems Analysis and Control, Springer, 2020. [103] D. Liu, W. Guo, J. Yi, D. Zhao, Passivity-based-control for a class of underactuated mechani- cal systems, in: IEEE International Conference on Intelligent Mechatronics and Automation, 2004, pp. 50–54. [104] B. Brogliato, Nonsmooth Mechanics. Models, Dynamics and Control, Springer Nature Switzerland AG, 2020. 139 [105] B. Brogliato, S. . Niculescu, P. Orhant, On the control of finite-dimensional mechanical systems with unilateral constraints, IEEE Transactions on Automatic Control 42 (2) (1997) 200–215. [106] R. I. Leine, N. Van de Wouw, Stability and Convergence of Mechanical Systems with Unilateral Constraints, Vol. 36, Springer Science & Business Media, 2007. [107] D. D. Bainov, P. S. Simeonov, Systems with Impulse Effect: Stability, Theory, and Applica- tions, Ellis Horwood, 1989. [108] R. G. Sanfelice, R. Goebel, A. R. Teel, Invariance principles for hybrid systems with con- nections to detectability and asymptotic stability, IEEE Transactions on Automatic Control 52 (12) (2007) 2282–2297. [109] R. Ortega, J. A. L. Perez, P. J. Nicklasson, H. J. Sira-Ramirez, Passivity-based control of Euler-Lagrange systems: Mechanical, Electrical and Electromechanical Applications, Springer Science & Business Media, 2013. [110] L. Consolini, A. Costalunga, M. Maggiore, A coordinate-free theory of virtual holonomic constraints, Journal of Geometric Mechanics 10 (4) (2018) 467–502. [111] N. Nekhoroshev, The Poincaré-Lyapunov-Liouville-Arnol’d theorem, Functional Analysis and Its Applications 28 (1994) 128–129. [112] H. K. Khalil, Nonlinear Systems, 2nd Edition, Prentice Hall, Upper Saddle River, NJ, 1996. [113] S. H. Strogatz, Nonlinear dynamics and chaos: With applications to physics, biology, chemistry, and engineering, CRC Press, 2018. [114] E. Ott, C. Grebogi, J. A. Yorke, Controlling chaos, Physical Review Letters 64 (11) (1990) 1196. [115] J. Grizzle, E. Westervelt, C. Canudas-de Wit, Event-based PI control of an underactuated biped walker, in: 42nd Int. Conference on Decision and Control, Vol. 3, IEEE, 2003, pp. 3091–3096. [116] P. Kokotović, H. K. Khalil, J. O’reilly, Singular perturbation methods in control: analysis and design, SIAM, 1999. 140