Spacecraft Attitude Dynamics Bacconi

download Spacecraft Attitude Dynamics Bacconi

of 148

Transcript of Spacecraft Attitude Dynamics Bacconi

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    1/148

    UNIVERSIT A DEGLI STUDI DI FIRENZE

    Dipartimento di Sistemi e Informatica

    Dottorato di Ricerca in

    Ingegneria Informatica e dellAutomazione

    ING-INF/04

    XVIII Ciclo

    Spacecraft Attitude Dynamics

    and Control

    Fabio Bacconi

    Ph.D. Coordinator

    Prof. Edoardo Mosca

    Advisors

    Prof. Edoardo Mosca

    Prof. Alessandro Casavola

    Anno Accademico 20052006

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    2/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    3/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    4/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    5/148

    8.3 Simulations of a LEO spacecraft performing Large Angle Maneuvers 123

    8.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .127

    Conclusion 128

    Geo-stationary Spacecraft and the 3D Pendulum . . . . . . . . . . . . . 129

    Constrained MEO and LEO Spacecraft . . . . . . . . . . . . . . . . . . . 131

    Research Contribution and Future Developments . . . . . . . . . . . . . 132

    Bibliography 134

    CONTENTS iii

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    6/148

    List of Figures

    1.1 Inertial references frame {I }and body reference frame {B} . . . . 101.2 Sequence of Eulers angles to be applied to frame{I }for obtaining

    frame {B}orienta t ion . . . . . . . . . . . . . . . . . . . . . . . . . 131.3 A schematic of a 3D Rigid Pendulum . . . . . . . . . . . . . . . . . 14

    1.4 Triaxial Air Bearing Testbed. . . . . . . . . . . . . . . . . . . . . . 20

    2.1 A schematic of a 3D Rigid Pendulum with one axis of symmetry . . 23

    2.2 Classical representation of the 2D spherical pendulum . . . . . . . . 25

    2.3 A schematic of a heavy spinning top not constrained by the table . 27

    3.1 Fan thrusters scheme of the Triaxial Attitude Control Testbed. . . . 37

    3.2 Experimental results for the evolution of the angular velocity of the

    3D pendulum in the body frame. . . . . . . . . . . . . . . . . . . . 45

    3.3 Experimental results for the evolution of the components of the di-

    rection of gravity in the body frame. . . . . . . . . . . . . . . . . 45

    3.4 Experimental result for the evolution of the angle between the re-

    duced attitude vector (t) and the desired reduced attitude vector

    0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    4.1 Evolution of the closed loop system (3.2), (4.1) and (4.2) towards

    the unstable equilibrium. . . . . . . . . . . . . . . . . . . . . . . . 52

    4.2 Evolution of the angular velocity of the 3D pendulum in the body

    frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    7/148

    4.3 Evolution of the components of the direction of gravity in the

    body frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

    4.4 Evolution of the angle between the reduced attitude vector (t) and

    the desired reduced attitude vector 0 . . . . . . . . . . . . . . . . . 58

    4.5 Vertical angular momentum (solid line) and total energy (dash line)

    of the 3D pendulum. . . . . . . . . . . . . . . . . . . . . . . . . . 59

    4.6 Motion of the vector between the pivot and the center of mass of

    the 3D pendulum in the inertial frame. . . . . . . . . . . . . . . . . 60

    4.7 Evolution of the angle between the reduced attitude vector (t) and

    the desired reduced attitude vector 0 for (0) = 0 . . . . . . . . . . 60

    5.1 Evolution of the angular velocity of the symmetric heavy top in the

    body frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

    5.2 Evolution of the components of the direction of gravity in the

    body frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

    5.3 Vertical angular momentum (solid line) and total energy (dash line)

    of the symmetric top. . . . . . . . . . . . . . . . . . . . . . . . . . 81

    5.4 Closed-loop trajectory of the symmetric top in the inertial frame. . . 82

    5.5 Evolution of the angular velocity of the 2D pendulum in the body

    frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    5.6 Evolution of the components of the direction of gravity in the

    body frame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    5.7 Evolution of the angle between the reduced attitude vector (t) and

    the desired reduced attitude vector 0 for (0) = 0 . . . . . . . . . . 84

    5.8 Closed-loop trajectory of the 2D pendulum in the inertial frame. . . 84

    6.1 Reference frame {B}motion with respect to an Earth centered ref-erence frame {EC I } . . . . . . . . . . . . . . . . . . . . . . . . . 90

    6.2 Reference frame {B}with respect to an orbiting reference frame{O}, both moving in {ECI } . . . . . . . . . . . . . . . . . . . . . 92

    7.1 Command Governor scheme. . . . . . . . . . . . . . . . . . . . . .101

    7.2 Relative position components x(k), y(k), z(k) of the LEO satellite,

    under the proposed LQ+CG control law (7.14). . . . . . . . . . . . 108

    LIST OF FIGURES v

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    8/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    9/148

    8.3 Reconguration for a LEO spacecraft, requiring a large angle maneuver.124

    8.4 Relative position y(t) (desired reference in dash lines), control forcef 2(t) and error e2(t) in coordination accuracy for y(t) (constraint

    boundaries in dash lines) under HCG. The vertical dash lines indicate

    instants of switching. . . . . . . . . . . . . . . . . . . . . . . . . .124

    8.5 Eulers angle (t) (desired reference in dash lines), control torque

    3(t) and error e3(t) in coordination accuracy for (t) (constraint

    boundaries in dash lines) under HCG. The vertical dash lines indicate

    instants of switching. . . . . . . . . . . . . . . . . . . . . . . . . .125

    8.6 Values of ( t) and (t) for the HCG selection logic (7.14), (8.7). . 1258.7 Relative position y(t), control force f 2(t) and error e2(t) in coor-

    dination accuracy for y(t) under LQ (constraint boundaries in dash

    lines). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .126

    LIST OF FIGURES vii

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    10/148

    Introduction

    Attitude control has been one of the most interesting and challenging problems

    in aerospace systems, from the rst half of the twentieth century, to present

    time. A modern spacecraft is a multibody system, composed of rigid bodies

    connected together by exible rods. From a mechanical point of view and with

    the aim of controlling its orientation, however, it can be modeled as a single rigid

    body. As a consequence, a spacecraft can be thought as a mechanical system

    with three translational and three rotational degrees of freedom. Moreover, if

    the satellite orbits at an altitude of approximately 35 .786 km above Earths

    surface, it appears stationary. Thus, only rotational motion is observed from

    Earth.

    Behavior of space vehicles on Medium (MEO) and Low Earth orbits (LEO)

    is quite different from the one at GEO altitudes. On a GEO orbit, dynamics

    is described by attitude models. Moreover, gravity does not affect rotational

    motion. At lower altitudes, on the contrary, also translational motion, and its

    effects on rotational motion, have to be taken into account. Indeed, differentpositions may correspond to different angular velocities, producing changes on

    attitude.

    The main goal of this dissertation is modeling and controlling the attitude of

    spacecraft, orbiting around Earth at GEO, MEO and LEO altitudes. Although

    on GEO orbits, gravity does not affect motion, we generalize the former case,

    introducing a new mechanical system with three rotational degrees of freedom

    in uniform gravity. Such a system, consists of a rigid body constrained to

    INTRODUCTION 1

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    11/148

    rotate about a pivot. We refer to it as 3D Pendulum. Furthermore, we study

    attitude dynamics and control of rigid bodies traveling on MEO and LEO orbits.To do that, we introduce a mathematical model describing both rotations and

    translations.

    Prior Literature on Spacecraft Attitude Control

    The idea of using satellites for communication purposes, goes back to 1945,

    when the science ction author Arthur C. Clarke, proposed a geo-stationary or-

    bit where placing a mechanical system, suitably oriented with respect to Earth.

    From that time, a huge literature has been produced on the several aspect

    related to the motion of space vehicles, as orbital motion and maneuvers, per-

    turbations, and attitude acquisition. Related theory has been summarized in

    numerous text books on spacecraft motion and control. Two classical texts

    are mainly referred in this dissertation. The rst one, has been published by

    Kaplan [Kap76] in the seventies. It analyzes spacecraft dynamics. The second

    one, written by Huges [Hug86] during the subsequent decade, is more focused

    on attitude. From a more technological perspective, especially related to space

    vehicles moving on MEO and LEO orbits and to their instrumentations, promi-

    nent is the text recently published by Sidi [Sid00]. Astrodynamics, on the other

    hand, are extensively described in [BMW71]. In particular, it illustrates in de-

    tail the classical two body and three body problems. Dynamics on MEO and

    LEO orbits are strictly connected to this theory.

    In this thesis, we study attitude control problems. Beyond the peculiar prop-

    erties of space systems, listed above, we focus our attention on the mechanical

    properties of rigid bodies. From this point of view, related theory is especiallyconnected to classical mechanics and geometry. Therefore, attitude models de-

    scribed by Goldstein et al. in [GPS80] and geometric properties presented by

    Marseden et al. in [MR99] and [MRS00] are fundamental. The literature on

    rotating mechanical systems is wide. One and two degrees of freedom rotating

    rigid bodies have been analyzed in several works, giving rise to pendulum mod-

    els. They constitute fundamental background for studying more sophisticated

    structures, as three degrees of freedom rotating rigid bodies. A comprehensive

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    12/148

    survey on pendulum literature has been presented by Furuta in [Fur03]. Related

    references cover the evolution of corresponding theory. Prominent is the workmade by Shirieav et al. in the nineties. [SEL99], [SPLE00] and [SEL04] propose

    control techniques for 2D pendulum, using conserved quantities of the motion

    and passivity-based results. We use a similar approach in this dissertation. In

    particular, we refer to a 3D rotating rigid body and we use reduction theory

    for modeling the system. Then, we use input-output passivity, for controlling

    attitude. Reduction theory is described in [MRS00]. Its main aspects have

    been developed by Bloch et al. , for controlling mechanical systems [BLM00],

    [BLM01]. Connections to the 2D pendulum and to the spinning heavy top,have been illustrated by Simo et al. in [SLRM92], by Wan et al. , [WCB95],

    [WCB95], and by related references. Passivity theory has been extensively de-

    scribed by Ortega et al. in [OSR98].

    Passivity control, lies in the more general context of energy-based control. In

    this thesis we make extensive use of energy-based control and Lyapunov func-

    tions. They are particularly suitable in the presence of mechanical systems.

    Ideas on suitable Lyapunov functions for rotating 3D rigid bodies have been

    provided by Burkov. In [Bur95], the author proposes a Lyapunov function for

    controlling the attitude of a spacecraft, with three control torques. On the other

    hand, control techniques based on the use of only two input torques, have been

    summarized by Doumtchenko and Tsiotras in [DT00]. Tsiotras et al. have pro-

    vided prominent work on modeling and controlling underactuated spacecraft.

    See [CLT00], [WTCB94] and related references.

    Recent theory in spacecraft attitude dynamics and control, has been focusing

    its attention on small satellites placed on MEO and LEO orbits. For Earth ob-

    serving missions and interferometric systems, these structures are more suitable

    than GEO spacecraft. If large instrumentation is required, the use of distrib-

    uted systems, shared between vehicles ying in formation, has been suggested.

    A survey on this new theory has been illustrated by Esper et al. in [ENS+ 03].

    Suitable models and control techniques have been analyzed for such systems.

    Due to their small dimension, constraints play a key role on their dynamics.

    On this subject, Veres et al. present a Model Based Predictive Control (MPC)

    technique for a eet of two LEO nano-satellites, in [VGRM01]. In this disser-

    INTRODUCTION 3

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    13/148

    tation we propose an attitude control technique also based on MPC. It consists

    on a suitably modied version of the Command Governor (CG) approach, de-veloped by Bemporad et al. [BCM97] and subsequently adopted in [CMA00]

    and [ABC + 03].

    Dissertation Outline

    We treat attitude models and control for rigid spacecraft. Since, dynamics of

    rigid bodies on GEO orbits are different from dynamics on LEO orbits, this

    dissertation can be broadly divided into two parts. In the rst part, we face the

    problem of modeling and controlling attitude of geo-stationary satellites. Apart

    of the presence of gravity, we notice immediately that their motion can be re-

    produced on Earth, pivoting a rigid body on a xed point. Thus, in Chapter 1

    we introduce the new concept of 3D Pendulum. It consists exactly of a three

    rotational degrees of freedom rigid body, whose translations are obstructed by

    the pivot. Its equilibria and their stability properties are described as well.

    Actually, the 3D pendulum model generalizes GEO spacecraft attitude models.

    Moreover, it generalizes 2D pendulum, 1D pendulum and spinning top mod-

    els. These aspects are developed in Chapter 2. Assumptions to be placed on

    the 3D pendulum model, for obtaining the mathematical description of these

    classical systems, are analyzed. Furthermore, advantages on using the 3D pen-

    dulum model, for controlling the equilibria of 2D pendulum and spinning top,

    in comparison to prior techniques, are motivated. A new mechanical system,

    following directly from the 3D pendulum model, is illustrated in Section 2.3.

    It represents a heavy top, spinning below its pivot. This system has not been

    studied in previous literature.Subsequently, we face the problem of stabilizing the natural equilibria of a 3D

    pendulum. They consist of the two congurations where direction of gravity is

    vertical, with respect to the rigid body. In Chapter 3, we propose passivity-

    based control laws for stabilizing these congurations in the general case, i.e.

    when the 3D pendulum is asymmetric. Techniques for stabilizing arbitrary ori-

    entations of the 3D pendulum are illustrated in Chapter 4. They represent

    attitude controllers that can be directly used to stabilize any attitude of geo-

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    14/148

    stationary spacecraft.

    The case when the 3D pendulum is axi-symmetric, is illustrated in Chapter 5.In this case, the control laws we propose, actually stabilize the 2D pendulum

    and the spinning top. In particular, in Section 5.3 we show that our approach

    guarantees the use of a continuous state-feedback for swinging up the spherical

    pendulum. It is a prominent result, if compared to previous literature, where

    this problem is solved by a swing-up controller, singular at the desired equilib-

    rium, a local stabilizing controller, and a technique for switching between the

    two.

    Attitude control of MEO and LEO spacecraft is faced in the second part of thisdissertation. In Chapter 6 we describe the main properties of space vehicles

    orbiting around Earth at constant velocity, under a leader-following approach.

    The subsequent model, describing both rotational and translational motion, is

    presented. Model Based Predictive Control techniques, under the assumption of

    small angular displacements between the satellite and a suitably dened orbital

    leader reference frame, are illustrated in Chapter 7. They adopt a Command

    Governor approach, for taking the presence of saturation constraints into ac-

    count. Indeed, satellites moving on low orbits are usually small. Thus, their

    thrusters cannot provide large torques. Moreover, we introduce suitable state-

    accuracy constraints, usually present in real missions. The importance of con-

    sidering constraints is the main difference between space systems on high orbits,

    as geo-stationary spacecraft, and space systems on low orbits. In this disser-

    tation this difference is the watershed between the rst and the second part.

    Finally, in Chapter 8, we remove the small angular displacements assumption.

    In order to restore results illustrated in the previous chapter, we address our

    control technique in a hybrid framework, proposing a Hybrid Command Gover-

    nor controller. It consists of a bank of controller as those presented in Chapter

    7, each one designed with respect to a reference attitude, and of a technique for

    switching between them.

    INTRODUCTION 5

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    15/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    16/148

    Part I

    Geo-Stationary Spacecraft

    Attitude Control

    7

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    17/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    18/148

    Chapter

    1Dynamics of the 3D Pendulum

    Three degrees of freedom (3D) rigid bodies have been extensively studied in me-

    chanics and their dynamics equations are well known in literature. They provide

    the fundamental background for modeling and controlling modern space and un-

    derwater vehicles. Simply, a rigid body is a system of particles whose relative

    distances are xed with time.

    In this chapter we introduce a particular kind of 3D rigid body: the 3D pendu-

    lum . Its dynamics is strictly related to the one of spacecraft in geo-stationary

    (GEO) orbits. Pendulum models have provided a rich source of examples in

    nonlinear dynamics and, in recent years, in nonlinear control. Several different

    models have been developed to analyze the properties of the uncontrolled planar

    1D pendulum and spherical 2D pendulum. However, their extension to a threedegrees of freedom pendulum, supported by a xed frictionless pivot, has never

    surprisingly been performed. Indeed, very few publications view the pendulum

    as a rigid body.

    We rst remind the main concepts describing 3D rigid body attitude models.

    This provides appropriate background for the subsequent introduction of the

    3D pendulum. Further, we present both the complete mathematical model of

    this system and a simplied version of it, particularly suitable for control pur-

    9

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    19/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    20/148

    with J denoting the constant inertia matrix in the body-xed reference frame,

    and

    dP dt I

    =ddt

    (r mv) = m(drdt v) + m(r

    dvdt

    ) = r d(mv)

    dt= ( r F ) =

    with m representing the total mass of the rigid body, r the position of its center

    of mass in {I }and a possibly applied torque. Thus,

    J = J + (1.2)

    Equation (1.2) is referred to as the Eulers equation and represents the dynamics

    of the rigid body in the general case. It is usually a simplifying assumption to

    select {B}aligned with the principals axes of the body, or equivalently, the mo-mentum P parallel to the angular velocity . In this case, all the extra-diagonal

    terms of J are null and the system reduces to three scalar equations. Details

    can be found in [GPS80], [Kap76] and [Sid00].

    Kinematics equations complete the mathematical model of the 3D rigid body.

    They describe the geometry of the motion apart from its causes. Many different

    kinematics models have been proposed in mechanics literature and aerospace

    control. An extensive overview of these models with focus on their main differ-

    ences can be found in [Hug86]. This text has also strongly inspired the connec-

    tions between space vehicles control and the 3D pendulum control described in

    the following chapters. In this dissertation we make use of two of these models.

    The rst one, namely the Poissons model , aims at describing the orientation of

    the inertial frame {I }in {B}, simply expressing its versors in terms of the {B}coordinates X , Y and Z . It is useful to control the attitude of satellites in

    GEO orbits. For this reason, it will be used hereafter to describe the 3D pen-

    dulum. The second model uses the Eulers angles equations and it is suitable to

    control the attitude of space vehicles in medium (MEO) and low (LEO) orbits.

    It will be widely described in the second part of this dissertation.

    The kinematics model arises from the representation of {I }in {B}by a suitablerotation matrix R

    {I }= R{B} (1.3)

    11

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    21/148

    It follows from the fact that the conguration manifold of a rotating body lies

    in1 SO (3) R 3 , in the general case [MR99], being SO (3) the group of theorthogonal matrices with determinant equal to 1:

    SO (3) = {R (3) : RR T = I, det (R) = 1}As described in detail in [Hug86] and in [Hal02], equation (1.3) is not a proper

    matrix product, but a notation commonly referred to as the vectrix represen-

    tation . The elements of R represent the cosines directors. Indeed, the ori-

    entation of any axes of {I }can be expressed as a linear combination of the

    {B}versors: X = R11 X + R12 Y + R13 Z , Y = R21 X + R22 Y + R23 Z and Z = R31 X + R32 Y + R33 Z where X = [1 , 0, 0]T , Y = [0 , 1, 0]T andZ = [0 , 0, 1]T in {B}. Now, applying equation (1.1) to the rst axis of {I },gives

    X = X = X (1.4)and analogous for Y and Z . Here is the skew-symmetric matric denoting the

    cross products between two vectors, such that a b = ab. It corresponds to

    a =

    0 a3 a2a3 0 a1a2 a1 0

    (1.5)

    Hence, rewriting (1.4) in terms of (1.3), we have directly

    R = R (1.6)

    that is the kinematics equation of the rigid body.

    The Poissons representation is originated by writing the above expression for

    any component. However, this representation is redundant because of the six

    orthogonality constraints. In fact SO (3) is a sub-variety of dimension 3 of R 9

    [MR99]. The, three variables, corresponding to the three degrees of freedom

    of the rigid body are enough for describing its orientation in space. Several

    mathematical models have been developed in order to characterize R using a

    lower number of variables. The most common is the Eulers angles kinematic

    model. It describes R in terms of the angles corresponding to three subsequent1 The properties of SO(3) affecting control actions on rigid bodies will be discussed in the

    following chapters. More details on the topology of SO(3) can be found in [MR99] and [Hat01].

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    22/148

    rotations to apply to the inertial frame, to become coincident with the body

    frame. In this work we consider three successive elementary rotations denedby the yaw angle , the pitch angle and the roll angle . First, impressing o

    Figure 1.2. Sequence of Eulers angles to be applied to frame {I } for obtaining

    frame {B } orientation

    rotation of angle to the Z axis of {I }, we have a new reference frame {I }:

    {I }= R {I }=cos() sin( ) 0

    sin() cos() 00 0 1

    {I }

    Subsequently, we can impress to this new reference frame a second rotation of

    pitch angle giving another frame {I }and then a third rotation of roll angle to {I }, obtaining {B}= R R R {I }. Thus, R in (1.3) corresponds to

    R = RT RT R

    T =

    cc sc + css ss cscsc cc sss cs + sscs cs cc

    (1.7)

    where cx stands for cos( x) and sx stands for sin( x). Equation (1.6) can then

    be written in terms of Eulers angles taking the rst derivative of (1.7) and

    extracting , and . It produces three rst order differential equations

    = y sin sec + z cos sec (1.8)

    = y cos z sin (1.9) = x + y sin tan + z cos tan (1.10)

    known as the Eulers angles representation . Note that these equations are valid

    as long as the pitch angle avoid the singularities at / 2.Model (1.2), (1.6) and some variations will be used in the rst part of this thesis

    in order to illustrate the properties of the 3D pendulum and its connections to

    13

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    23/148

    previous literature. It is particularly suitable for controlling the attitude of

    GEO spacecraft. On the contrary, model (1.2), (1.8)-(1.10) is more convenientfor attitude control of MEO and LEO satellites. It will be mostly used in the

    second part of this thesis.

    1.2 The 3D Pendulum

    A pendulum is a rigid body supported by a xes, frictionless pivot, acted on by

    gravitational forces and disturbance and control forces or moments. In the most

    general case, we consider a pendulum without axes of symmetry. A schematic

    of a rigid pendulum is shown in Figure 1.3. The supporting pivot allows three

    Figure 1.3. A schematic of a 3D Rigid Pendulum

    degrees of rotational freedom to the pendulum. Uniform constant gravity is

    assumed. It represents the element in equation (1.2). Thus, the pendulum is

    described by at least three spacial coordinates and three rotational coordinates,

    suggesting the terminology 3D pendulum .

    Two coordinate frames, are of interest. The inertial reference frame {I }hasits origin at the pivot, the rst two coordinate axes lie in the horizontal planeand the third coordinate axis is vertical in the direction of gravity. The origin

    of the body frame {B}is also located at the pivot. In this frame the inertiamatrix J is constant and can be simply computed from the moment of inertia

    of a translated coordinate frame whose origin is located at the center of mass of

    the pendulum, using the parallel axis theorem [HRW04]. The body-xed vector

    from the pivot to the center of mass of the pendulum is denoted by . The

    symbol g denotes the constant acceleration due to gravity. This is the data on

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    24/148

    which the equations of motion are based. The dynamics are given by the Eulers

    equation (1.2) that includes the moment due to gravity:

    J = J + mg RT e3 (1.11)

    where m is the total mass of the rigid body and e3 is the third unit vector in

    the inertial coordinate frame, namely e3 = (0 , 0, 1)T . Equation (1.11) and the

    rotational kinematics equation (1.6) dene the full dynamics of a rigid pendulum

    in SO (3) R 3 .It is important to point out that in GEO aerospace applications the dynamics

    are not affected by gravity terms [Kap76]. Hence, these systems are specialcases of the above model. In particular, this is the case when the center of mass

    of the rigid body is located at the pivot, that is = 0. In the context of the

    rigid pendulum, we refer to this situation as the balanced case. It allows at

    viewing the geo-stationary spacecraft attitude control problem as a simplied

    version of the 3D pendulum attitude control problem. Therefore, the focus on

    what follows, will be on the more general unbalanced case, viz. = 0.

    There are two conserved quantities for the rigid pendulum. The rst one is

    the total energy, which is the sum of the rotational kinetic energy and thegravitational potential energy. In addition, there is a rotational symmetry of the

    equations of motion corresponding to the group of rotations about the vertical

    line through the pivot. These two results are summarized as follows.

    Proposition 1.1 The total energy

    E = 0 .5T J mg T RT e3and the component of the angular momentum vector about the vertical axis

    through the pivot h = T JR T e3

    are constant along motions of the rigid pendulum.

    The proof follows by showing that the time derivative of the total energy and the

    time derivative of the angular momentum component are identically zero. This

    can be shown using the dynamics equations (1.11) and the kinematics equation

    (1.6).

    15

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    25/148

    Thanks to the presence of a conserved quantity along the motion, it is possible

    to obtain a lower dimensional reduced model for the pendulum. In practice,the function h can be seen as a constraint imposing the system to lie in a lower

    dimensional manifold. It allows at expressing one of the variables of the motion

    with respect to the others. A comprehensive explanation of the constants of the

    motion for a dynamic system and their effects on the conguration manifold can

    be found in [MR99]. Here, we introduce a model reduction by exploiting the

    fact that the dynamics and kinematics can be written in terms of the reduced

    attitude vector = RT e3 , which is the unit vector that expresses the gravity

    direction in the body-xed frame {B}.Proposition 1.2 The dynamics of the 3D pendulum given by (1.11) and (1.6)

    induce a ow on S 2 R 3 given by the reduced dynamics

    J = J + mg (1.12)

    and the reduced kinematics

    = (1.13)

    The proof follows from the denition of the reduced attitude vector anddemonstration that = 1.

    Equations (1.12) and (1.13) are in non-canonical form, but they are very useful

    for studying the reduced dynamics of the rigid pendulum. Thus, they are used

    in this dissertation, in place of the full dynamical model. It is worth to point

    out that the reduced model describes the orientation of the 3D rigid pendulum

    modulo a rotation along its third axis Z . However, this is not a limitation in

    spacecraft control problems, where usually, the task is of controlling the direc-

    tion of an axis, e.g. represented by an antenna, instead of the complete attitude.Finally, equations (1.12) and (1.13) can also be expressed in terms of the Eulers

    angles noticing that

    (, ) =sin

    sin cos

    cos cos

    (1.14)

    In fact, equations (1.9)-(1.10) and (1.13) using (1.14) are exactly an expression

    of the reduced dynamics of motion of the rigid pendulum on S 2 R 3 expressed

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    26/148

    in terms of the Eulers angles. Further, the kinematics can be shown to be

    independent from the yaw angle [MR99].

    1.3 Equilibria of the 3D Pendulum

    To further understand the dynamics of the 3D pendulum, we study its equilibria.

    To simplify the analysis it is convenient to choose the body-xed frame so that

    its third axis Z is aligned with the vector from the pivot to the center of mass.

    Thus, in {B}, = (0 , 0, l)T , where l > 0. Consequently, the gravity terms thatappear in equation (1.11) can be considerably simplied.

    The equilibrium solutions for the 3D pendulum satisfy = 0, so that the

    attitude of the pendulum is constant. Substituting = 0 in equation (1.11), we

    obtain

    RT e3 = 0The above equation implies that the 3D pendulum is in equilibrium only if

    the center of mass vector is collinear with the gravity vector RT e3 in the

    body-xed coordinate frame. This implies that

    RT e e3 = e 3

    where = 1. For = 1, the center of mass vector points in the directionof the gravity vector. Similarly, for = 1, the center of mass vector pointsin the direction opposite to the gravity vector. Now, let Re denote an attitude

    rotation that solves equation (1.11) and dene e = RT e e3 . Then, any attitude

    in the conguration manifold given by

    {R SO (3) : R = R e Q(), arbitrary

    }denes an equilibrium attitude corresponding to = 0 where

    Q() =

    cos sin 0

    sin cos 00 0 1

    Hence, if R e denes an equilibrium attitude for the 3D pendulum, then a rota-

    tion of the pendulum about the gravity vector by an arbitrary angle is also an

    17

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    27/148

    equilibrium. Consequently, in SO (3) there are two disjoint equilibrium mani-

    folds of the 3D pendulum. We dene the manifold corresponding to = 1 in theabove description as the hanging equilibrium manifold , since the center of mass

    is below the pivot for each attitude in the manifold. The manifold correspond-

    ing to = 1 in the above description is dened as the inverted equilibrium manifold , since the center of mass is above the pivot for each attitude in the

    manifold.

    Next, we study the equilibria of the reduced equations (1.12) and (1.13). As

    before, the equilibria of the system satises = 0. Substituting in (1.12), we

    get = 0 .

    Thus, we obtain two isolated equilibrium solutions of the reduced equations in

    S 2 R 3 given by = e3 , = 0, and by = e3 , = 0. Alternatively, thetwo equilibrium solutions can be expressed using Eulers angles as = 0, = 0,

    = 0 and = 0, = , = 0. For each of these representations of the

    equilibrium solutions, the rst corresponds to the hanging equilibrium where

    the center of mass is below the pivot and the second corresponds to the inverted

    equilibrium where the center of mass is above the pivot.The stability of the two isolated equilibrium solutions of the reduced equations

    of motion is readily assessed. The stability of the two families of equilibrium

    manifolds of the full equations of motions follows. The stability analysis can

    then be based on the reduced model.

    Proposition 1.3 The hanging equilibrium of the reduced dynamics of the 3D

    pendulum described by equations (1.12) and (1.13) is locally stable in the sense

    of Lyapunov.

    Proof. The hanging equilibrium corresponds to the conguration 0 = 0 and0 = (0 , 0, 1)T . Consider the Lyapunov function

    V (, ) =12

    T J + mg(l T )V (0 , 0) = 0 and V (, ) > 0 elsewhere. The derivative of V along the

    trajectories is

    V = T J mg T = T (J + mg ) mg T ( )= T mg mg T = mg(T T ) 0

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    28/148

    where the last equality follows from the mixed vector product property [HRW04].

    Hence, V is a continuously differentiable function, positive denite and suchthat V (0 , 0) = 0 and 2 V (, ) 0 (, ). Thus, the hanging equilibriumis locally Lyapunov stable [Kha01].

    Proposition 1.4 The inverted equilibrium of the reduced dynamics of the 3D

    pendulum described by equations (1.12) and (1.13) is locally unstable.

    Proof. The hanging equilibrium corresponds to the conguration 0 = 0 and

    0 = (0 , 0, 1)T . Consider the Lyapunov function

    V (, ) =12

    T J + mg(l + T )

    V (0 , 0) = 0 and V (, ) > 0 elsewhere. The derivative of V along the

    trajectories is

    V = T J + mg T = T (J + mg ) + mg T ( )= T mg + mg T = 2 mg T ( )

    where the last equality follows from the mixed vector product property [HRW04].

    V is not denite. Hence, V is a continuously differentiable function such thatV (0 , 0) = 0 and V (, ) > 0, V (, ) > 0 for some points ( , ) arbitrarily

    close to (0 , 0). Thus, by Chetaevs theorem [Kha01], the inverted equilibrium

    is unstable.

    As a consequence of the two propositions above, one can conclude that the

    hanging equilibrium manifold of the full dynamics of the 3D pendulum is stable

    in the sense of Lyapunov and that the inverted equilibrium manifold of the full

    dynamics of the 3D pendulum is unstable in the sense of Lyapunov.

    1.4 Triaxial Attitude Control Testbed

    The test-bed used for experiments in the rst part of this dissertation is based

    on the spherical air bearing manufactured by Space Electronics, Inc., Berlin,2 Notice that the result V 0 could be expected, since V denes the total energy of the

    system, that is a constant of the motion, as described in section 1.2.

    19

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    29/148

    CT, depicted in Figure 1.4. It is available at the Attitude Dynamics and Con-

    trol Laboratory, Department of Aerospace Engineering, University of Michigan.We refer to this systems as the Triaxial Attitude Control Testbed (TACT).

    Figure 1.4. Triaxial Air Bearing Testbed.

    An aluminum sphere of diameter 11 inch oats on a thin lm of air that exits

    holes located in the surface of the cup. Air at high constant pressure is supplied

    to the cup by means of a hose that passes through the center of the vertical

    support. Once the main components are mounted, additional masses can be

    added to modify the nal mass distribution. Hence, the center of mass can

    be balanced at the rotational center (no effects from the gravity), in order to

    simulate a geo-stationary spacecraft, or in any other position, representing a 3D

    unbalanced rigid pendulum. The spherical air bearing guarantees low friction

    and three-dimensional motion with unrestricted roll and yaw and a maximum

    45 degrees pitch angle.A comprehensive description of the hardware related to the TACT has been pro-

    vided by [BMB01]. Mathematical models for the TACT have been introduced in[CSMB01] and [CSM03]. There, a description equivalent to (1.11) and (1.6), is

    obtained by means of the Lagrangian equations. This geometrical approach to

    mechanical systems, well known in the literature, is widely described in [MR99]

    and [MRS00].

    As illustrated in Figure 1.4, a one-piece shaft passes through the center of the

    sphere and extends between a pair of circular mounting plates. Extension shafts

    connect the circular mounting plates to larger square mounting plates. Circular

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    30/148

    plates are used for supporting hardware instrumentation, while square plates

    are used for supporting fan thrusters, batteries and additional weights. In-deed, weights can be attached to the TACT, thanks to access holes cut into the

    plates. This allows at modifying the inertia matrix J of the TACT, before per-

    forming experiments. Once the entire system is set up, the procedure described

    in [CBA+ 06] can be adopted, to identify the inertia matrix of the system.

    1.5 Conclusions

    In this chapter we have introduced the new concept of 3D rigid pendulum.

    It represents a three rotational degrees of freedom rigid body, supported by a

    frictionless pivot. We have presented the 3D pendulum model in the general

    unbalanced case. The special balanced case, corresponding the absence of grav-

    ity terms in the system, encompasses the geo-stationary spacecraft dynamics

    model. Therefore, we have addressed the problem of controlling the attitude

    of GEO satellites, as a simplied version of the 3D pendulum attitude control

    problem.

    21

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    31/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    32/148

    duce assumptions to be placed on model (1.12), (1.13) for obtaining a spherical

    pendulum model and then a planar pendulum model. Further, we show that,with simple assumptions, the 3D pendulum model can describe the heavy top

    dynamics. The corresponding dynamics seem not to have been previously stud-

    ied. Finally, we study the stability properties of this new spinning top model,

    illustrating connections with the classical heavy top problem, widely discussed

    in [GPS80].

    2.1 Axisymmetric 3D Rigid PendulumWe select the body-xed reference frame {B}such that its versors X , Y andZ are aligned with the principal axes of the 3D rigid body. This renders the

    extra-diagonal terms of the inertia matrix J null. In this case, the diagonal

    terms of J are referred to as the principal moments of inertia . Hence, model

    (1.12) using (1.14), reduces to three scalar equations, as described in Chapter

    1.

    Moreover, we consider the axisymmetric case, where two of the principal mo-

    ments of inertia of the pendulum are identical and the pivot is located on the

    axis of symmetry of the pendulum. This is an usual assumption while facing

    Figure 2.1. A schematic of a 3D Rigid Pendulum with one axis of symmetry

    attitude control problems for underactuated geo-stationary spacecraft [DT00].

    In particular, we assume that the body-xed axes are selected so that J =

    diag( J t , J t , J a ) and = le3 where l is a positive scalar constant. Consequently,

    23

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    33/148

    equation (1.12) using (1.14), can be written in scalar form, as

    J t x = ( J t J a )y z mgl cos sin J t y = ( J a J t )z x mgl sin J a z = 0

    (2.1)

    From the last equation in (2.1) we see that the component of the angular velocity

    vector about its axis of symmetry is constant. This means that a constant z

    denes an invariant manifold for the pendulum dynamics. The resulting reduced

    equations of motion, for constant z , can be represented by equations (2.1) and

    (1.9)-(1.10)J t x = ( J t J a )y z mgl cos sin J t y = ( J a J t )z x mgl sin

    = y cos z sin = x + y sin tan + z cos tan

    (2.2)

    Remember that the yaw motion of the pendulum is described by the kinematic

    equation (1.8). Hence, the evolution of can be deduced from the evolution of

    the other variables of motion.

    Now, the special case z 0 leads to an axisymmetric rigid body whom atti-tude dynamics are described by two angles and two angular velocities, namely,a 2D spherical pendulum .

    Proposition 2.1 Assume the rigid pendulum has a single axis of symmetry and

    the pivot is located on the axis of symmetry of the pendulum. The equations of

    motion of the rigid pendulum dene an induced ow on the manifold S 2 R 2corresponding to z = 0 , given by the equations

    J t x = mgl cos sin J t y = mgl sin

    = y cos

    = x + y sin tan

    (2.3)

    These equations are said to represent a 2D spherical pendulum. They exhibit

    singularities for = n/ 2.It is worth to point out that the above representation is not the one commonly

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    34/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    35/148

    is that now the singularities are in q 1 = n , i.e. in the upward position.This leads to several disadvantages while designing control laws for swingingup the spherical pendulum, that is a basic problem in the literature related to

    attitude control. For instance, in [SEL99] and [SEL04] the authors use model

    (2.5) and they switch to a different model in the close proximity of the upward

    position. It produces a discontinuous representation of the system with many

    related difficulties. As it will be seen in the following chapters, it could simply

    be avoided, by selecting the more suitable description of the mechanical system

    (2.3), arising from the interpretation of the 2D spherical pendulum as a rotating

    rigid body.Now, beyond the assumption above of z = 0, consider the case where x = 0

    and = 0. It is straightforward to show that this leads to an invariant dynamics

    of the axisymmetric rigid pendulum, described as follows:

    Proposition 2.2 Assume the rigid pendulum has a single axis of symmetry and

    the pivot is located on the axis of symmetry of the pendulum. The equations of

    motion of the rigid pendulum dene an induced ow on the manifold S 1 R 1corresponding to x = 0 , z = 0 , and = 0 , given by the equations

    J t y = mgl sin = y

    (2.6)

    These equations are said to represent a 1D planar pendulum .

    Thus, for an axially symmetric pendulum with the pivot located on the axis

    of symmetry, the well known 2D spherical pendulum and 1D planar pendulum,

    can be viewed as special cases of the 3D rotating rigid pendulum. On the other

    hand, if the pendulum is asymmetric, then its dynamics are general, in the

    sense that neither the spherical pendulum dynamics, nor the planar pendulum

    dynamics, arise as special cases.

    2.2 Spinning Axisymmetric 3D Rigid Pendulum

    The 2D spherical pendulum has been shown to originate from the 3D axisym-

    metric rigid pendulum, under the assumption of z 0. The induced dynamicscorresponding to a nonzero value of z is fundamentally different from this case.

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    36/148

    These dynamics seem not to have been previously studied.

    Thus, consider again an axisymmetric pendulum with the pivot located onits axis of symmetry. Its dynamics are described by model (2.2). Assume

    z = k = 0. Since the third equation in (2.1) holds true, we have the following

    result:

    Proposition 2.3 Assume the rigid pendulum has a single axis of symmetry and

    the pivot is located on the axis of symmetry of the pendulum. For any constant

    k = 0 , the equations of motion of the rigid pendulum dene an induced ow on

    the manifold S 2

    R 2 corresponding to z = k, given by the equations

    J t x = ( J t J a )k y mgl cos sin J t y = ( J a J t )k x mgl sin

    = y cos k sin = x + y sin tan + k cos tan

    (2.7)

    These equations represent the dynamics of a symmetric rigid body spinning

    steadily around its axis of symmetry under the effects of gravity, namely, an

    heavy spinning top [GPS80]. In particular, differently from previous literature,

    they extend the concept of heavy spinning top to the space under the table.

    It is depicted in Figure 2.3.

    Figure 2.3. A schematic of a heavy spinning top not constrained by the table

    Thus, the motion of the symmetric top can be seen as a special case of the 3D

    rigid pendulum dynamics.

    Problems related to the symmetric spinning top have been extensively studied in

    27

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    37/148

    mechanics. A wide discussion on the properties of the motion, on the conditions

    for precession and on the stability of the upward conguration, can be found in[GPS80]. Several papers have analyzed the properties of its dynamics as well,

    with particular focus on stabilizing the sleeping motion. On this subject, see

    [WTCB94], [WCB95] and related references. The sleeping motion consists of

    the situation where the top stands in the vertical conguration, steady rotating.

    Furthermore, the dynamics of the heavy top represent basic background in the

    literature of astrodynamics, for instance, to explain the Earth wobble and the

    precession of equinox [BMW71].

    Surprisingly, none of the published works have extended the motion of the heavytop to the case where there is no table, i.e. when the top is also allowed to move

    under the horizontal plane. From this point of view, current theory can be

    considered incomplete. Further, proposed control laws for the vertical upward

    equilibrium, dened global as in [WCB95], are actually not global. This is

    due to the presence of another equilibrium for the system, represented by the

    downward vertical conguration. In the language introduced for the 3D rigid

    pendulum, this conguration consists of the hanging equilibrium of the reduced

    dynamics (1.12) and (1.13). On the other hand, the upward equilibrium cong-

    uration, consists of the inverted equilibrium.

    In order to complete the description of the dynamics of the spinning top in

    the new context of the 3D rigid pendulum, we study the stability of its hang-

    ing equilibrium. An almost global attitude control law for the system, will be

    proposed in the following chapters.

    2.3 Hanging Equilibrium of the Symmetric Spinning

    Top

    For any constant velocity z = k, system (2.7) has two equilibria on S 2 R 2 .They are given by the two states where x = y = 0 and angles and cor-

    responding to the upward and the downward conguration. The rst one is

    described by = 0, = or = , = 0 and it is referred to as the inverted

    equilibrium. The second one is described by = 0, = 0 or = , =

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    38/148

    and it is referred to as the hanging equilibrium. As introduced above, only

    the inverted equilibrium has been studied in classical mechanics, restricting theanalysis to the space above the horizontal plane: (/ 2, 3/ 2). However,

    model (2.7) describes the motion on the entire space S 2 R 2 , provided thatangle does not cross the horizontal: = / 2 and = 3 / 2. Thus, we study

    the properties of the second equilibrium, located below the horizontal plane. It

    aims at completing theory related to the heavy symmetric top.

    In order to analyze the stability of the hanging equilibrium, we make use of a

    different model with respect to (2.7), representing a 3-2-1 formulation of the

    Eulers angles. The reason of this selections is motivated by the fact that inmodel (1.8)-(1.10), the sleeping motion does not correspond to an isolated equi-

    librium. Indeed, when the rst rotation is around axis Z , the second and the

    third Eulers angles are not distinguishable. It could be expected noticing that

    each equilibrium is described by two different congurations.

    In this context, the problem can be overcame by using a 2-1-3 Eulers angles

    formulation, as the one previously presented in [WCB95]. There, the inertial

    frame {I }attitude is described by three angles corresponding to three subse-quent rotations, where the rst one is around Y , the second one is around the

    rst axis of the new frame and the last one is around the third axis of the ob-

    tained frame. In [WCB95] an angle is introduced, corresponding to the rst

    rotation about Y . Then, the new frame is rotated of angle about the rst

    axis. Resulting reference frame is referred to as F . Finally, a positive rotation

    of the F -frame of angle about its third axis, results in the body-xed frame.

    Thus, F precesses, but does not spin with the top. The angular velocity of the

    3D rigid body can then be expressed in F , giving

    = [ , cos , sin ]T (2.8)

    It allows at writing the kinetic and potential energy in F coordinates and then

    at following the Lagrangian approach (2.4). The resulting model is

    J t + J t 2 cos sin + p cos mgl sin cos = 0J t cos2 2J t cos sin p cos mgl cos sin = 0

    (2.9)

    29

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    39/148

    where p is the constant generalized momentum corresponding to :

    p = L

    = J t ( sin )

    The rational under description (2.9), depending only on two of the three Eulers

    angles, is that does not appear in the Lagrangian [WCB95], so it is an ignorable

    variable. As a consequence, the Lagrangian can be replaced by the Routhian R ,

    dened by R = L p . Reduction theory, useful in the presence of symmetriesin mechanical systems, is extensively described in [MRS00].

    Now, letting

    x1 = x2 = x3 = x4 =

    b =pJ t

    c =2mgl

    J t(2.10)

    we obtain a state space representation for model (2.9):

    x1 = x2

    x2 = x242

    sin(2x1) bx4 cos x1 +c2

    sin x1 cos x3

    x3 = x4

    x4 = (2 x2x4 sin x1 + bx2 +c2 sin x3)

    1cos x1

    (2.11)

    Notice again the presence of singularities, here for = n/ 2, n Z . As ex-

    pected, the equilibria of the system correspond now to isolated points repre-

    sented by x = [i, 0,j, 0]T . This results in the inverted equilibrium, i.e. the

    upward vertical sleeping motion, when i j is even. Otherwise, it correspondsto the hanging equilibrium.

    In [WCB95] a linearized analysis around the inverted equilibrium x i = [0 , 0, 0, 0]T

    is performed, to show that it is unstable for b2 < c . It means that the rotation

    of the top needs to be fast enough, to guarantee a vertical conguration. Anenergy based equivalent proof is provided in [GPS80]. Here, we use a similar ap-

    proach, dening a suitable Lyapunov function. Indeed, it is easy to show that

    the linearized model around xh does not provide information on its stability,

    being the eigenvalues complex conjugates.

    Thus, we introduce a Lyapunov function based on the conserved quantities of the

    motion embedded in (2.11): total energy and vertical momentum. This way to

    proceed, previously suggested by [SLRM92], is referred to as the Energy-Casimir

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    40/148

    method. It consists of selecting a Lyapunov function composed by Casimir func-

    tionals, i.e. differences between conserved quantities and their minimum values.Now, the total energy of (2.9) is given by

    E =12

    J t (2 + 2 cos2 ) +12

    J a ( sin )2 + mgl cos cos

    Disregarding the kinetic energy related to the rotation around the spinning axis

    and dividing by J t we obtain a constant of the motion:

    hc1(x) =12

    (x22 + x24 cos

    2 x1) +c2

    cos x1 cos x3 (2.12)

    expressed in the state space variables dened above. Further, the vertical mo-

    mentum is given by h = T J with dened by (2.8) and as introduced in

    Chapter 1, but expressed in F coordinates:

    =

    1 0 0

    0 cos sin

    0 sin cos

    cos 0 sin 0 1 0

    sin 0 cos

    0

    0

    1

    =sin sin cos

    cos cos

    Hence,

    h = J t sin + J t cos cos sin + J 3( sin )cos cos == J t sin + J t cos cos sin + J t bcos cos

    Dividing by J t and using the state space representation, we have a second con-

    stant of the motion:

    hc2(x) = x2 sin x3 + x4 sin x1 cos x1 cos x3 + bcos x1 cos x3 (2.13)

    Therefore, a suitable Energy-Casimir based Lyapunov function for the system

    isV (x) = [hc1(x) hc1(xh )] +

    b2

    [hc2(x) hc2(xh )]Consequently, stability of the hanging equilibrium can be studied.

    Proposition 2.4 For any constant spinning velocity z = k = 0 , the hanging

    equilibrium of the axisymmetric rigid pendulum (2.7), is locally stable in the

    sense of Lyapunov.

    31

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    41/148

    Proof. Consider description (2.11) of the system and the following Energy-

    Casimir based Lyapunov function

    V (x) = [ hc1(x) hc1(xh )] +b2

    [hc2(x) hc2(xh )]=

    12

    (x22 + x24 cos

    2 x1) +c2

    cos x1 cos x3 +c2

    +b2

    (x2 sin x3+ x4 sin x1 cos x1 cos x3 + bcos x1 cos x3 + b)

    (2.14)

    where b and c are dened by (2.10).

    V (xh ) = 0 and V (x) 0 x. Indeed, (2.14) is composed by two constants of the motion. Moreover, consider the Hessian matrix of V (x) evaluated in xh :

    V (xh ) =

    c+ b22 0 0 b20 1 b2 0

    0 b2c+ b2

    2 0

    b2 0 0 1(2.15)

    Its eigenvalues are

    1,2 = b4 + 2 b2c + c2 4c + 4

    4+

    b2 + c + 24

    3,4 = + b4 + 2 b2c + c2 4c + 4

    4+

    b2 + c + 24

    Simple manipulations show that 1,2 can be negative only if b2 + 2 c < 0, but

    this is clearly impossible, since c = 2 mgl/J t > 0. Moreover, 3 ,4 are always

    positive. Thus, all the eigenvalues of the Hessian matrix evaluated in xh are

    positive, i.e. V (xh ) is positive denite. Therefore, the hanging equilibrium is

    Lyapunov stable [Kha01].

    As a consequence of the above proposition, the inverted equilibrium cannot be

    globally stabilized in S 2 R 2 .Finally, in [GPS80] it is shown that the heavy top is in steady precession when

    p2 4J t mgl cos cos Thus, since any attitude of the top below the horizontal plane is such that

    cos cos < 0, there the precession is always steady, no matter is the angular

    velocity. These dynamics are fundamentally different from those of the spherical

    pendulum. They open new interesting theoretical problems that seem not to be

    previously studied, as the research of possibly existing periodic trajectories.

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    42/148

    2.4 Conclusions

    In Chapter 1 we have introduced a 3D rigid pendulum mathematical model

    encompassing geo-stationary spacecraft descriptions. In this chapter we have

    shown that this model actually embodies many other classical problems. Indeed,

    the 2D spherical pendulum, the 1D planar pendulum and the symmetric heavy

    top have been shown to represent special cases of model (1.12)-(1.13). Thus,

    the study of the 3D pendulum includes the study of these classical problems.

    Moreover, we have extended the classical description of the heavy top to the

    space below the horizontal plane, giving rise to novel dynamics. Finally, we

    have analyzed the equilibria of this new inverted top illustrating connections

    with current literature.

    33

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    43/148

    Chapter

    3Stabilization of an Asymmetric

    3D Rigid Pendulum

    In the previous chapters we have introduced several different classications of

    an uncontrolled 3D rigid pendulum supported by a frictionless pivot, both in

    the asymmetric and axisymmetric case. Related theory is basic in spacecraft

    dynamics and control problems, since an asymmetric balanced 3D rigid pen-

    dulum, actually represents a generic geo-stationary satellite [Hug86], [Kap76].

    Moreover, assumptions on the symmetry of the rigid body, have been taken in

    the literature, while facing attitude control problems for underactuated space-

    craft. Indeed, a rigid body with only two control inputs satises the fundamental

    Broketts condition [BMS83] for continuous feedback stabilization, only if it issymmetric. Consequent model coincide with the axisymmetric balanced 3D

    rigid pendulum developed in Chapter 2.

    In this chapter we focus the attention on an asymmetric 3D rigid pendulum

    and we study the stabilization of its reduced model. For stabilizing the reduced

    model of the 3D pendulum we mean stabilizing its hanging equilibrium, i.e.

    pointing the vector downwards.

    First, we make control actuation assumptions for the system, reminding briey

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    44/148

    some alternatives to our selection. Then, we present a simple controller consist-

    ing of angular velocity feedback. The related control law is based on dampinginjection and it is shown to have a dense domain of attraction. Damping in-

    jection, as well as passivity based control [OSR98], is particularly suitable for

    mechanical systems, as the TACT presented in Chapter 1. Afterwards, we show

    experimental results on the TACT, conrming the effectiveness of the proposed

    controller.

    3.1 3D Rigid Body Control TorquesSo far, we have considered the mathematical model of an uncontrolled 3D rigid

    body supported by a frictionless pivot, making assumptions on its geometrical

    structure and analyzing its dynamics. In order to consider control inputs to the

    system, the dynamics equation have to be modied for including moments. In

    particular, (1.11) can be modied as

    J = J + mg RT e3 + M (3.1)

    where M represents the vector control moment on the 3D rigid pendulum, ex-pressed in the body-xed coordinate frame. The specic form of the control

    moment depends on the actuation assumptions, i.e. on the hardware selected

    for providing torques to the system. Several different actuation structures have

    been studied in the literature. One important example of control actuation as-

    sumption, generalizing the development in [AF00], is the pivot acceleration in

    the control input. In this case the control moment is

    M = m

    RT u

    where u denotes the pivot acceleration, expressed in the inertial coordinate

    frame. Such a control actuation assumption typically breaks the symmetry of

    the uncontrolled rigid pendulum dynamics. Consequently, the angular momen-

    tum component about the vertical axis through the pivot is not conserved and

    it is not possible to express the controlled dynamics in terms of the reduced

    attitude .

    Another control actuation assumption, generalizing the development in [CSMB01]

    35

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    45/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    46/148

    Y

    Z

    X

    F1

    r

    F2

    F3

    F4

    C Mq

    q

    Figure 3.1. Fan thrusters scheme of the Triaxial Attitude Control Testbed.

    In the case of the TACT, forces F i are generated applying suitable voltages to

    the fan thrusters. The relations between torques and voltages are described

    in [CBA+ 06]. They allows at expressing control inputs to the TACT in terms

    of voltages to be applied to the fans. Results obtained in this reference are

    promising for continuing fruitful research on the control test-bed.

    Beyond the relations between F i and the torques in the roll, pitch and yawdirections, we assume here that inputs to the 3D rigid body are directly the

    angular momentum corresponding to roll , pitch and yaw . Hence, the dynamics

    model extending (1.11) is simply

    J = J + mg RT e3 + uMoreover, RT e3 can be replaced by , since the angular momentum component

    about the vertical axis through the pivot is conserved along the trajectories

    of rigid body. Therefore, the control model for the fully actuated 3D rigid

    pendulum on which we focus our attention is

    J = J + mg + u =

    (3.2)

    where R 3 is the angular velocity of the body, S 2 is the reduced at-

    titude vector representing the direction of gravity in {B}and u R 3 is theangular momentum input vector, and where all the variables are expressed in

    {B}coordinates.

    37

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    47/148

    3.2 Asymptotic Stabilization of the Hanging Equilib-

    rium

    In this section, we develop a simple controller that makes the hanging equi-

    librium of the asymmetric 3D pendulum asymptotically stable. The controller

    is based on the observation that the control model given by equation (3.2) is

    input-output passive if the angular velocity is taken as the output. Passivity

    based control is a natural way to proceed in the presence of mechanical systems,

    as well as in the presence of electrical systems [OSR98]. Briey, a system with

    state x, input u and output y is dissipative with respect to a supply function

    s(u(t), y(t)) if there exists a function H such that

    H (x(T )) H (x(0)) + T

    0s(u(t), y(t))dt (3.3)

    Such a function is referred to as the storage function . When the supply rate

    is s(u, y ) = uT y, the system is said input-output passive . In this case, it is

    immediate to notice thatdH

    dt uT y

    Thus, selecting u = y, gives a closed-loop system where dH/dt 0. There-fore, if H is the total energy, u drives the system to the minimum of its energy,

    corresponding to a stable equilibrium.

    In the case of the 3D rigid pendulum, the total energy 12 T J mg T can be

    selected as a storage function. Since it has a minimum at the hanging equilib-

    rium (0 , h ), a control law based on angular velocity feedback is suggested.

    Let : R 3 R 3 be a smooth function such that

    1 x 2 xT (x) 2 x 2 , x R 3 (3.4)where 2 1 > 0. Thus, we propose a class of controllers, referred to asdamping injection controllers, given by

    u = () (3.5)

    where satises (3.4). We next show that the above family of controllers,

    which requires only angular velocity feedback, renders the hanging equilibrium

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    48/148

    of a 3D asymmetric pendulum asymptotically stable, with a guaranteed domain

    of attraction.

    Lemma 3.1 Consider the fully actuated 3D asymmetric pendulum given by

    (3.2). Let : R 3 R 3 be a smooth function satisfying (3.4) and choose a controller in the form of (3.5). Then, the hanging equilibrium (0, h ) is asymp-

    totically stable.

    Furthermore, for every (0, 2mgl ), all solutions of the closed-loop system

    given by (3.2) and (3.5), such that ((0) , (0)) H, where

    H= (, ) R3

    S 2

    :12

    T

    J +12mgl

    l

    T

    l 2mgl

    satisfy limt

    (t) = 0 and limt

    (t) = h .

    Proof. Consider the closed loop system given by (3.2) and (3.5). We propose

    the following candidate Lyapunov function

    V (, ) =12

    T J + mgl l

    T

    l

    (3.6)

    Note that the above Lyapunov function is positive denite on R3

    S 2

    andV (0, h ) = 0. Furthermore, the derivative along a solution of the closed-loop

    system given by (3.2) and (3.5) is

    V (, ) = T (J ) + mg(l )T = T (J + mg ()) + mg(l )T ( )= T () + mg T ( ) mg T ( )= T () 1 2 ,

    where the last inequality follows from (3.4). Thus, V (, ) is positive deniteand V (, ) is negative semidenite on R 3 S 2 .Next, consider the sub-level set given by H= {(, ) R 3 S 2 : V (, ) 2mgl }. Note that the compact set Hcontains the hanging equilibrium(0, h ). Since, V (, ) 0, all solutions such that ( (0) , (0)) H satisfy((t), ( t)) Hfor all t 0. Thus, His an invariant set for solutions of (3.2)and (3.5). Furthermore, from LaSalles invariant set theorem, we obtain that

    solutions satisfying ( (0) , (0)) H converge to the largest invariant set in

    39

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    49/148

    {(, ) H : = 0}, that is = 0.Therefore, either ( t) /l = h or ( t) /l = i as t . Since(0, i ) H, it follows that ( t) h as t . Thus, (0 , h ) is an asymptoti-cally stable equilibrium of the closed-loop system given by (3.2) and (3.5), with

    a domain of attraction H.

    This result could be expected noticing that 2 mgl is the total energy of the pen-

    dulum at the inverted equilibrium. Thus, when ( (0) , (0)) is in H, injectionof dumping can only decelerate the rigid body, towards the other equilibrium,

    represented by the hanging equilibrium. Different is the case when ( (0) , (0))is such that the total energy is exactly 2 mgl . In this case ( t) can be shown

    to approach h as t , only if the initial kinetic energy is not zero. Thenontrivial proof of this result is provided by the following Lemma.

    Lemma 3.2 Consider the fully actuated 3D asymmetric pendulum given by

    (3.2). Let : R 3 R 3 be a smooth function satisfying (3.4) and choose a controller as given by (3.5). Then, all solutions of the closed-loop system given

    by (3.2) and (3.5), such that (0) = 0 and ((0) , (0)) A, where

    A= (, ) R 3 S 2 :12

    T J +12

    mgl l

    T

    l

    = 2 mgl

    satisfy limt

    (t) = 0 and limt

    (t) = h .

    Proof. Consider the closed loop system given by (3.2) and (3.5), and the Lya-

    punov function given in (3.6). As already shown, V (, ) = T (). Thus,

    V (, ) = T () T () = T () + T

    since is smooth. Furthermore, for all ( , ) A, is bounded and

    = J 1 (J + mg ())is also bounded. Dene

    N = sup(, ) A

    T () +

    T

    < .

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    50/148

    Next, since ( (0) , (0)) A, V ((0) , (0)) = 2 mgl and V ((0) , (0)) N .Expanding V ((t), (t)) in a Taylor series, gives

    V ((t), ( t)) = 2 mgl (0) T ((0)) t + R(t) 2mgl 1 (0) 2 t + N t 2 ,

    since the remainder necessarily satises R(t) Nt 2 .Dene

    = min 2mgl, 3 21(0) 4

    16N .

    It can be easily shown that for all t [T 1 , T 2],

    t2 1 (0)2

    N t + N 0where

    T 1 = 1(0) 2

    2N 1 1 4 N 21 (0) 4 > 0

    and T 2 T 1 1 (0) 2

    2N . Choose an (0, ). Then, for all t [T 1 , T 2],

    V ((t), (t)) 2mgl < 2mgl and hence, ( (t), ( t)) H, where Histhe invariant set given in Lemma 1. Thus, from Lemma 1, we obtain the result

    that (t)

    0 and ( t)

    h , as t

    .

    By the two Lemmas above, the hanging equilibrium is stable, with a guaranteed

    domain of attraction. Indeed, when the initial conditions ( (0) , (0)) H, thetrajectories converge to the hanging equilibrium. Furthermore, when the initial

    total energy is 2 mgl and the initial kinetic energy is not zero, the trajectories

    enter Hin nite time. This result is obtained applying the simple dampinginjection based control law (3.5). This result is summarized by next theorem.

    Theorem 3.1 Consider the fully actuated 3D pendulum given by (3.2). Let

    : R 3 R 3 be a smooth function satisfying (3.4) and choose controller as in (3.5). Then, all solutions of the closed-loop system given by (3.2) and (3.5),

    such that ((0) , (0)) N\{(0, i )}, where

    N = (, ) R 3 S 2 :12

    T J +12

    mgl l

    T

    l 2mgl (3.7)

    satisfy limt

    (t) = 0 and limt

    (t) = h .

    41

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    51/148

    Proof. From Lemmas 3.1 and 3.2, we obtain the result that for every

    (0, 2mgl ) and ( (0) , (0)) H A, where Hand Aare as dened in Lemmas3.1 and 3.2, (t) 0 and ( t) h as t . Since, N can be written as

    N =(0 ,2mgl )

    (H A)

    the result follows.

    It is important to point out that N\{(0, i )}is an invariant set, but it is notthe maximal domain of attraction of the hanging equilibrium. Any solution of

    (3.2) and (3.5), starting from an initial condition not in N\{(0, i )}, that doesnot pass through (0 , i ), must eventually enter N\{(0, i )}. Hence, the maxi-mal domain of attraction of the hanging equilibrium is a proper superset of the

    invariant set N\{(0, i )}.In the subsequent theorem, we claim that for the control given in (3.5), the

    equilibrium (0 , h ) of (3.2) and (3.5) is almost globally asymptotically stable,

    and the domain of attraction of the hanging equilibrium is dense.

    Theorem 3.2 Consider the fully actuated 3D asymmetric pendulum described

    by (3.2). Let : R 3 R 3 be a smooth function satisfying (3.4) and (0)be positive denite and symmetric. Choose controller as given in (3.5). Let Mdenote the stable manifold of the equilibrium (0, i ). Then (R 3S 2)\Mis dense in (R 3 S 2), and all solutions of the closed-loop system given by (3.2) and (3.5),such that ((0) , (0)) (R 3 S 2)\M, satisfy limt (t) = 0 and limt (t) = h .Proof. Consider the closed loop system given by (3.2) and (3.5), and the

    Lyapunov function given in (3.6). Since S 2 is a compact set and the Lyapunovfunction V (, ) is quadratic in , each sublevel set of V (, ) is a compact

    set. Furthermore, since V (, ) = T () 0, by LaSalles invariant settheorem, all solutions converge to the largest invariant set in {(, ) R 3 S 2 :V (, ) = 0 }.The largest such invariant set is given by {(0, h )} {(0, i )}. Hence, it is suf-cient to show that the stable manifold of the inverted equilibrium (0 , i ), has

    dimension less than the dimension of ( R 3 S 2), i.e. ve, since all other solutions

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    52/148

    converge to the hanging equilibrium (0 , h ) by invariant set theorem [Kha01].

    Using linearization, it can be shown that the equilibrium (0 , i ) of the closedloop given by (3.2) and (3.5) is unstable and hyperbolic with nontrivial stable

    and unstable manifolds. Hence, from theorem 3.2.1 in [GH83], it follows that

    the dimension of the stable manifold of (0 , i ) is less than ve, so that the do-

    main of attraction in theorem 3.1 is dense in ( R 3 S 2).

    Comparing the conclusions in the two previous theorems we notice that, theo-

    rem 3.1 provides an explicit description of a domain of attraction of the hanging

    equilibrium; however that domain of attraction is not maximal. In contrast, the-orem 3.2 shows that the maximal domain of attraction consists of all points in

    (R 3 S 2) that are not in the stable manifold Mof the inverted equilibrium.However, the stable manifold of the inverted equilibrium is not explicitly de-

    scribed. We note that Lebesgue measure of Mis zero.

    3.3 Experiments on Stabilization of the Hanging Equi-

    librium

    In this section we present experimental results on the stabilization of the hanging

    equilibrium of the asymmetric 3D rigid pendulum. They validate the controller

    proposed in the previous section, showing that (3.5) asymptotically stabilizes the

    hanging equilibrium with angular velocity feedback. The test-bed on which the

    experiments are performed, is the Triaxial Attitude Control Testbed (TACT)

    described above.

    We are able to measure the angular velocity of the TACT thanks to threeorthogonal gyroscopes mounted on one of the two external plates, suitably cali-

    brated. First, bias is eliminated keeping the TACT in a resting position. Then,

    a rotation about each axis is performed, assuring output only from the corre-

    sponding gyro, and no data from the others.

    In order to measure the attitude of the TACT, a triaxial accelerometer is

    mounted on one of the two external plates, aligned with the body-xed axes.

    Before each test, the accelerometer triad is calibrated to read g on each indi-

    43

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    53/148

    vidual axis when aligned with the downward vertical. To do this, the TACT

    is rotated about each axis independently, while taking continuous reading fromthe accelerometer. The accelerometer axis is dened to be aligned with the

    downward vertical when a maximum occurs in the data. Bias and scaling fac-

    tors are then adjusted such that each accelerometer outputs g at the maximum

    point.

    Actually, the reduced attitude cannot be measured directly. Indeed, accelerom-

    eter are not mounted at the center of gravity of the system. Thus, must be

    estimated. We are able to estimate the reduced attitude using accelerometers

    and gyroscopes together. First we read the accelerometer measure. It gives notonly the acceleration due to gravity, but also angular and centripetal accelera-

    tion terms. Centripetal acceleration can be determined indirectly by using the

    knowledge of the angular velocity. It correspond to ( r ). The angularacceleration term is ignored in the computation of the reduced attitude, but

    preliminary calculations suggest that its contribution is quite small.

    Hence, direction of gravity is calculated as

    Gravity = Measure ( r )

    representing the vector, since already expressed in body-xed coordinates.

    Finally, inertia matrix J is identied as described in [CBA + 06].

    Results of the application of (3.5) to stabilize the hanging equilibrium, are

    depicted in Figures 3.2 to 3.4. The controller used for this purpose is chosen

    such that

    () = P where P = 10 I 3 . This choice of equal gains in the three feedback chan-

    nels does not represent a particularly effective choice in terms of closed-loopperformance. However, our objective in the experiment is to demonstrate

    closed-loop stabilization; we do not seek to tune the gains to achieve closed-

    loop performance. The desired reduced attitude vector is the hanging equi-

    librium 0 = (0 , 0, 1)T . The pendulum is released from the reduced attitude

    (0) = ( 0.2936, 0.8579, 0.4217)T , with (0) = 0. These initial conditions arechosen to excite all degrees of freedom of the system.

    As seen from experimental closed loop responses shown in Figures 3.2 and 3.3,

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    54/148

    0 5 10 15 20 25 30 35 405

    0

    5

    x

    ( D e g

    / s e c )

    Angular Velocities

    0 5 10 15 20 25 30 35 405

    0

    5

    y

    ( D e g

    / s e c

    )

    0 5 10 15 20 25 30 35 4050

    0

    50

    100

    z

    ( D e g

    / s e c

    )

    Time [sec]

    Figure 3.2. Experimental results for the evolution of the angular velocity of the 3Dpendulum in the body frame.

    0 5 10 15 20 25 30 35 40

    0.5

    0

    0.5

    Reduced Attitude Vector in Body Frame

    x

    0 5 10 15 20 25 30 35 401

    0.5

    0

    0.5

    1

    y

    0 5 10 15 20 25 30 35 400

    0.5

    1

    z

    Time [sec]

    Figure 3.3. Experimental results for the evolution of the components of the directionof gravity in the body frame.

    the angular velocity converges to zero and ( t) converges to 0 = (0 , 0, 1)T as

    t . This is clear from Figure 3.4 which shows that the response of the

    45

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    55/148

    0 5 10 15 20 25 30 35 400

    10

    20

    30

    40

    50

    60

    70

    80

    Time [sec]

    [

    D e g

    ]

    Angle between (t) and 0 in Body Frame

    Figure 3.4. Experimental result for the evolution of the angle between the reducedattitude vector ( t ) and the desired reduced attitude vector 0 .

    angle between ( t) and 0 converges to zero as t . Notice that a residualdamped oscillation remains in the pitch axis angular velocity. It can be reduced

    adjusting the feedback control gains. Notice also that measurements during the

    initial transient period are especially noisy. This can be attributed to both gyros

    noise and an attitude estimation algorithm that ignores angular acceleration,

    which can add to noisy measurements during the initial transient period.

    The fan thrusters mounted on the TACT provide approximately 5 Nm of con-

    trol authority in each axis. Thrusters saturation can be an issue in these control

    experiments and the feedback gains need to be tuned to minimize any nonlin-

    earities arising from saturation. This is not critical in stabilizing the hangingequilibrium of the 3D pendulum. Indeed, saturation might only produce a slower

    convergence. On the other hand, it has to be taken into account while stabi-

    lizing a generic attitude. Techniques to be used in the presence of saturation

    constraints, for stabilizing spacecraft, will be presented in the second part of

    this dissertation.

    Finally, notice that in this experiment, initial conditions are inside the guaran-

    teed domain of attraction N . Several experiments have been performed on the

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    56/148

    TACT showing that under (3.5), dynamics converge to the hanging equilibrium

    also starting outside N . They are presented in [SCB + 05].

    3.4 Conclusions

    In this chapter we have introduced a passivity based control law for stabilizing

    the hanging equilibrium of an asymmetric 3D rigid pendulum. The described

    control technique is based on the observation that angular velocity feedback

    renders the closed-loop input-output passive, with the total energy as storage

    function. We have shown that this control law cannot be global, due to the

    presence of the inverted equilibrium, but it is near-global. Finally, we have

    presented results of the application of the technique to the TACT.

    Even if the task of stabilizing the hanging equilibrium of the 3D rigid pendulum

    can appear quite simple, since h is a minimum for the potential energy, this is

    fundamental for introducing the control of a generic attitude described in the

    following chapter.

    47

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    57/148

    Chapter

    4Attitude Control of an

    Asymmetric 3D Rigid

    Pendulum

    In the previous chapter we have presented a technique for stabilizing the hanging

    equilibrium of an asymmetric 3D pendulum, based on damping injection. The

    proposed control law has been inspired by the fact that the hanging equilibrium

    is a minimum of the potential energy of the system. In this chapter we gen-

    eralize this concept, proposing a control law for an arbitrary reduced attitude

    vector.

    For controlling the reduced attitude of a 3D pendulum we mean pointing its vector towards a pre-established direction. This is an usual task in satellite sys-

    tems, corresponding to stabilize the orientation of spacecraft instrumentations

    or antennas. Thus, this problem is strictly related to geo-stationary satellites

    attitude control. Several local control laws for geo-stationary spacecraft, have

    been proposed in the literature, based on Eulers angle error, quaternion error,

    Euler axis of rotation and many other techniques 1 . In particular, here we focus1 See [Kap76] and [Sid00] for a comprehensive overview.

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    58/148

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    59/148

    following control law for stabilizing an arbitrary orientation 0

    u = P + k(0 ) (4.2)

    with P a positive denite matrix and k a real positive constant.

    Theorem 4.1 The closed-loop system (3.2) and (4.2) with u as given in (4.1),

    possesses an asymptotically stable equilibrium in (, ) = (0 , 0) and an unstable

    equilibrium in (, ) = (0 , 0).

    Proof. Let 0 be an arbitrary reduced attitude. Note that the closed loopsystem given by (3.2), (4.1) and (4.2) has two equilibria for ( , ) = (0 , 0) and

    (, ) = (0 , 0). Consider the following candidate Lyapunov function

    V (, ) =12

    T J + k( 0)T ( 0) (4.3)

    The above Lyapunov function is positive denite on R 3 S 2 and V (0, 0) = 0.Furthermore, the derivative along a solution of the closed-loop system given by

    (3.2), (4.1) and (4.2) is

    V (, ) = T J + k( 0)T = T (P + k( 0)) + k(0)T ( )= T P + kT ( 0) kT 0 ( )= T P 0

    Thus, V (, ) is positive denite and V (, ) is negative semidenite on R 3S 2 .Hence, from LaSalles invariant set theorem, we obtain that ( , ) converges to

    the largest invariant set in R 3 S 2 such that = 0, i.e. ( 0) = 0. Moreover,if ((0) , (0)) is such that V ((0) , (0)) < 2k, V ((0) , (0)) is an invariant set

    for the trajectories of the closed loop system, starting from ( (0) , (0)). Since

    V (0, 0) = 2 k, it follows that ( t) 0 as t . Thus, (0 , 0) is anasymptotically stable equilibrium of the closed-loop system given by (3.2), (4.1)

    and (4.2).

    On the other hand, consider the following Lyapunov function

    V (, ) =12

    T J + k( + 0)T ( + 0) (4.4)

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    60/148

    The above function is positive denite on R 3 S 2 and V (0, 0) = 0. Fur-thermore, it is straightforward to show that V (, ) is not dened along thetrajectories of the closed-loop system given by (3.2), (4.1) and (4.2). Thus,

    (0, 0) is an unstable equilibrium of the closed-loop system given by (3.2),(4.1) and (4.2).

    Notice that the term P in (4.2) can be replaced by any function ( ) asdened by (3.5). Indeed, it plays the role of injecting damping into the system.

    Hence, the above controller is strictly connected to the one presented in chapter

    3, for stabilizing the hanging equilibrium h = /l . This can also be deduced

    comparing equation (3.6) and equation (4.3). The fundamental difference is that

    an arbitrary 0 is not a minimum of the potential energy, unless the potential

    in suitably modied. In some way, this is what equations (4.1) and (4.2) do.

    They modify the system, for having a closed-loop whose potential energy has a

    minimum at = 0 . This way to proceed, is referred to as potential shaping ,

    in the sense that the control law actually re-shape the potential energy of the

    whole system. A comprehensive description of this technique, that is particu-

    larly suitable in the presence of mechanical systems, can be found in [BLM00],

    [BLM01] and related references.

    theorem 4.1 states the capability of control law (4.1) and (4.2) of stabiliz-

    ing a desired attitude 0 , when initial condition satises V ((0) , (0)) < 2k,

    with V (, ) dened by (4.3). Unfortunately, it does not guarantee that in

    general, the evolution of the closed loop system converge to the desired atti-

    tude. Indeed, initial condition cannot be assured to belong to the invariant set

    V ((0) , (0)) < 2k. Selecting a large k for enlarging the guaranteed domain

    of attraction of (0 , 0) is not in accordance with constraints usually presentin practical problem. In fact, in real applications, thrusters cannot supply ar-

    bitrary large torques. This is especially true in aerospace applications, where

    minimizing consumption is also an issue.

    When ( (0) , (0)) is not in V ((0) , (0)) < 2k, an initial condition belonging

    to the stable manifold of (0 , 0) cannot be avoided a priori. This situation isdepicted in Figure 4.1 where lines represent function V (, ) level sets.

    Therefore, the technique proposed in [Bur95] and its simplied version described

    51

  • 7/27/2019 Spacecraft Attitude Dynamics Bacconi

    61/148

    G

    -G

    0

    0

    V( = 2k w,G)

    G

    w

    (w(0),G(0))

    0

    Figure 4.1. Evolution of the closed loop system (3.2), (4.1) and (4.2) towards theunstable equilibrium.

    above, represent a local attitude control law. In order to arbitrarily enlarge the

    domain of attraction of the desired nal attitude, one should design a controller

    that enforces the closed-loop trajectories to move away from ( , ) = (0 , 0)as soon as the closed-loop evolution approaches the unstable equilibrium. This

    argument leads to th