293
This is page i Printer: Opaque this Mechanics, Control, and Variational Principles. J. Baillieul & A. M. Bloch & P. E. Crouch & J. E. Marsden 28 September 1998

Mechanics, Control, and Variational Principles.murray/courses/cds/cds280/BLOCH.pdfMechanics, Control, and Variational Principles. ... Mechanics has traditionally described the behavior

Embed Size (px)

Citation preview

This is page iPrinter: Opaque this

Mechanics, Control, and Variational

Principles.

J. Baillieul & A. M. Bloch & P. E. Crouch & J. E. Marsden

28 September 1998

ii

This is page iiiPrinter: Opaque this

Contents

1 Introduction 31.1 The Rolling Disk . . . . . . . . . . . . . . . . . . . . . . . . 4

1.1.1 The Vertical Rolling Disk . . . . . . . . . . . . . . . 41.1.2 The Falling Rolling Disk . . . . . . . . . . . . . . . . 8

1.2 The Heisenberg System . . . . . . . . . . . . . . . . . . . . 101.3 The Toda Lattice . . . . . . . . . . . . . . . . . . . . . . . . 131.4 The Rigid Body . . . . . . . . . . . . . . . . . . . . . . . . . 15

1.4.1 The Free Rigid Body . . . . . . . . . . . . . . . . . . 151.4.2 Example: The Rolling Ball . . . . . . . . . . . . . . 171.4.3 A Homogeneous Ball on a Rotating Plate . . . . . . 191.4.4 The inverted pendulum on a cart . . . . . . . . . . . 201.4.5 The Rotating Pendulum . . . . . . . . . . . . . . . . 221.4.6 Rigid Body with Rotors . . . . . . . . . . . . . . . . 22

2 Mathematical preliminaries 252.1 Vector fields, flows and differential equations . . . . . . . . 252.2 Differentiable Manifolds . . . . . . . . . . . . . . . . . . . . 392.3 Differential Forms . . . . . . . . . . . . . . . . . . . . . . . 462.4 Lie derivatives . . . . . . . . . . . . . . . . . . . . . . . . . 532.5 Stokes’ Theorem, Riemannian Manifolds and Distributions . 602.6 Lie groups, Spatial Relationships, and the Euclidean Group 632.7 Fiber Bundles, Connections, and Gauge Theory . . . . . . . 68

2.7.1 Fiber Bundles . . . . . . . . . . . . . . . . . . . . . . 682.7.2 Connections on TR1 . . . . . . . . . . . . . . . . . . 71

iv Contents

2.7.3 Linear connections, affine connections and geodesics 722.7.4 Principal Connections . . . . . . . . . . . . . . . . . 742.7.5 Parallel Translation and Holonomy Groups . . . . . 772.7.6 Affine Connections and Covariant Derivatives . . . . 772.7.7 Riemannian Connections . . . . . . . . . . . . . . . 79

3 Basic concepts in mechanics 813.1 First Principles - Coordinates and Kinematics . . . . . . . . 813.2 First Principles - Conservation of Energy, Work, Virtual Work 813.3 Newtonian, Lagrangian, and Hamiltonian Mechanics . . . . 833.4 Symplectic and Poisson Manifolds . . . . . . . . . . . . . . 863.5 The Flow of a Hamiltonian Vector Field . . . . . . . . . . . 883.6 Cotangent Bundles . . . . . . . . . . . . . . . . . . . . . . . 883.7 Lagrangian Mechanics . . . . . . . . . . . . . . . . . . . . . 893.8 Lie-Poisson Structures and the Rigid Body . . . . . . . . . 913.9 The Euler-Poincare Equations . . . . . . . . . . . . . . . . . 943.10 Momentum Maps . . . . . . . . . . . . . . . . . . . . . . . . 963.11 Symplectic and Poisson Reduction . . . . . . . . . . . . . . 983.12 Singularities and Symmetry . . . . . . . . . . . . . . . . . . 1013.13 A Particle in a Magnetic Field . . . . . . . . . . . . . . . . 1023.14 The Mechanical Connection . . . . . . . . . . . . . . . . . . 104

3.14.1 Example – the nonlinear pendulum . . . . . . . . . . 105

4 An Introduction to Geometric Control Theory 1074.1 Nonlinear Control Systems . . . . . . . . . . . . . . . . . . 107

4.1.1 Controllability and Accesibility . . . . . . . . . . . . 1084.1.2 Stability and Stabilizability . . . . . . . . . . . . . . 110

4.2 Stabilization Techniques . . . . . . . . . . . . . . . . . . . . 1114.2.1 Linearization . . . . . . . . . . . . . . . . . . . . . . 1114.2.2 Necessary Conditions . . . . . . . . . . . . . . . . . 1114.2.3 Lyapunov Methods for asymptotic stability . . . . . 1124.2.4 The center manifold . . . . . . . . . . . . . . . . . . 113

4.3 Hamiltonian and Lagrangian Control Systems . . . . . . . . 114

5 Nonholonomic Mechanics 1155.1 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . 115

5.1.1 The Lagrange d’Alembert Principle . . . . . . . . . 1165.1.2 Nonholonomic equations of motion with Lagrange

multipliers . . . . . . . . . . . . . . . . . . . . . . . . 1185.2 An Invariant Approach to Nonholonomic Mechanics . . . . 119

5.2.1 Lagrangian Mechanics . . . . . . . . . . . . . . . . . 1195.2.2 Constrained Dynamics . . . . . . . . . . . . . . . . . 122

5.3 Projected Connections and Newton’s law for NonholonomicSystems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

5.4 Systems with Symmetry . . . . . . . . . . . . . . . . . . . . 127

Contents v

5.4.1 Group Actions and Invariance . . . . . . . . . . . . . 1275.5 The Momentum Equation . . . . . . . . . . . . . . . . . . . 129

5.5.1 The Derivation of the Momentum Equation . . . . . 1305.5.2 The Momentum Equation in a Moving Basis . . . . 1345.5.3 The Momentum Equation in Body Representation . 136

6 Controllability and Stabilizability for Mechanical and Non-holonomic Systems 1416.1 Controllability, Accessibility and Stabilizability for Nonholo-

nomic Mechanical Systems . . . . . . . . . . . . . . . . . . . 1416.2 Stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

6.2.1 Smooth Stabilization to Manifolds . . . . . . . . . . 1446.3 Normal Form Equations . . . . . . . . . . . . . . . . . . . . 1456.4 Stabilization to an Equilibrium Manifold . . . . . . . . . . 147

6.4.1 Nonsmooth Stabilization . . . . . . . . . . . . . . . . 1486.5 Controlled Nonholonomic Systems on Manifolds . . . . . . . 160

6.5.1 Holonomic Control Systems . . . . . . . . . . . . . . 1606.5.2 Nonholonomic Systems . . . . . . . . . . . . . . . . 161

6.6 Symmetries and Conservation Laws . . . . . . . . . . . . . . 1636.6.1 Reduction . . . . . . . . . . . . . . . . . . . . . . . . 164

6.7 A Discontinuous Energy Method for a Class of Nonholo-nomic System . . . . . . . . . . . . . . . . . . . . . . . . . . 1676.7.1 Stabilizing the Heisenberg System . . . . . . . . . . 1686.7.2 Stabilization of the Nonholonomic Integrator in Slid-

ing Mode . . . . . . . . . . . . . . . . . . . . . . . . 169

7 Optimal Control 1757.1 The Maximum Principle . . . . . . . . . . . . . . . . . . . . 1757.2 Variational and Nonvariational Nonholonomic Problems . . 1777.3 Variational Nonholonomic Mechanics and Optimal Control 179

7.3.1 Two Examples of Variational Nonholonomic Mechanics1827.4 Optimal Control on Lie Algebras and Adjoint Orbits . . . . 183

7.4.1 Optimal Control on Symmetric Spaces . . . . . . . . 1887.4.2 Optimal Control and the Toda Flow . . . . . . . . . 194

7.5 Kinematic/subRiemannian Optimal Control Problems . . . 1987.5.1 The kinematic subRiemannian optimal control problem1987.5.2 Necessary Conditions on a Compact Semisimple Lie

Group . . . . . . . . . . . . . . . . . . . . . . . . . . 2037.5.3 The Case of Symmetric Space Structure . . . . . . . 2057.5.4 Optimal control and a particle in a magnetic field . 2097.5.5 Rigid Extremals . . . . . . . . . . . . . . . . . . . . 212

7.6 Dynamic Optimal Control . . . . . . . . . . . . . . . . . . . 2147.6.1 The Falling Cat Theorem . . . . . . . . . . . . . . . 2147.6.2 Optimal control of the rolling ball . . . . . . . . . . 217

vi Contents

8 Energy Based Methods for Stability 2198.1 The Energy Momentum Method . . . . . . . . . . . . . . . 219

8.1.1 The Energy-Momentum Method for Holonomic Sys-tems . . . . . . . . . . . . . . . . . . . . . . . . . . . 220

8.2 The Energy-Momentum Method for the Stability of Non-holonomic Systems . . . . . . . . . . . . . . . . . . . . . . . 2218.2.1 The Equations of Motion of Nonholonomic Systems

with Symmetries . . . . . . . . . . . . . . . . . . . . 2268.2.2 The Pure Transport Case . . . . . . . . . . . . . . . 2328.2.3 The Non-Pure Transport Case . . . . . . . . . . . . 237

9 Energy Based Methods for Stabilization 2599.1 Controlled Lagrangian Methods . . . . . . . . . . . . . . . . 259

9.1.1 Rigid Body with Rotors . . . . . . . . . . . . . . . . 2599.1.2 The pendulum on a cart . . . . . . . . . . . . . . . . 261

9.2 Controlled Lagrangian methods for Nonholonomic Systems 2619.3 Averaging and Energy Methods for the Dynamics of Me-

chanical Systems . . . . . . . . . . . . . . . . . . . . . . . . 2619.3.1 Energy and Curvature . . . . . . . . . . . . . . . . . 2619.3.2 Second Order Generalized Control Systems . . . . . 265

9.4 Flat Systems and Systems with Flat Inputs . . . . . . . . . 2719.5 Averaging Lagrangian and Hamiltonian Systems with Oscil-

latory Inputs . . . . . . . . . . . . . . . . . . . . . . . . . . 2779.6 Stability of Mechanical Systems with Oscillatory Inputs . . 2799.7 Stabilization of Rotating Systems . . . . . . . . . . . . . . . 282

9.7.1 Heavy chains . . . . . . . . . . . . . . . . . . . . . . 2829.7.2 Body-beam systems . . . . . . . . . . . . . . . . . . 282

References 283

Contents vii

Preface

Our goal in this book is to explore some of the connections between con-trol theory and geometric mechanics, in particular classical mechanics inits Lagrangian and Hamiltonian forms as well as the theory of mechan-ical systems subject to motion constraints. The latter turns out to havea particularly rich connection with nonlinear control theory. While an in-troduction to aspects of the mechanics of nonholonomically constrainedsystems may be found in such sources as the monograph of Naiimark andFu’faev [1972], the control theory of such systems remains largely scatteredthrough various research journals. Our aim is to provide a unified treat-ment of the control theory of mechanical systems which will incorporatematerial that has not yet made its way into texts and monographs.

Mechanics has traditionally described the behavior of free and interactingparticle and bodies, the interaction being decribed by potential forces. Itencompasses the Hamiltonian and Lagrangian and Hamiltonian picturesand in its modern form relies heavily on the tools of differential geometry(see, for example, Abraham and Marsden [1978] and Arnold [1978]).

Control theory is the theory of prescribing motion for dynamical systemsrather than describing their observed behavior. These systems may or maynot be mechanical in nature, and in fact traditionally, the underlying systemis not assumed to be mechanical. Modern control theory began largely aslinear theory, having its roots in electrical engineering and using linearalgebra, complex variable theory, and functional analysis as its princpaltools. The nonlinear theory of control on the other hand relies to a largeextent again on differential geometry.

Nonholonomic mechanics describes the motion of systems constrained bynonintegrable constraints, i.e. constraints on the system velocities but notits configuration. Classic examples are rolling and skating motion. Nonholo-nomic mechanics fits uneasily into the classical mechanics as it is not vari-ational in nature, i.e. it is neither Lagrangian or Hamiltonian in the strictsense of the word. It has a close cousin (variational axiomatic mechanics –a term coined by Arnold [1988]) which is variational for systems subject tononintegrable constraints, but which does not describe the motion of me-chanical systems. It is important however for the theory of optimal control,as we shall describe.

There is a very close link between nonholonomic constraints and con-trollability of nonlinear systems. Nonholonomic constraints are given bynonintegrable distributions – i.e. taking the bracket of two vector fields insuch distribution gives rise in general to a vector field not contained inthis distribution. It is precisely this property that one wants in a nonlinearcontrol systems, in order that we can drive the sytem to as large a part ofthe state space as possible.

It turns out that a key concept for studying the control and geometryof nonholonomic systems, as well as may other mechanical systems, is the

viii Contents

notion of fiber bundle. The bundle not only gives us a way of organizingvariables in a physically meaningful way, but gives us basic ideas on howthe system behaves physically, and on how to prescribe controls. A bundleconnection relates base and fiber variables in the system, and in this sensethe theory of nonholonomic control sytems is, as we shall see, closely relatedto gauge theories.

There is also a beautiful link between optimal control of nonholonomicsystems and so-called subRiemannian geometry. For a large class of phys-ically interesting systems, the optimal control problem reduces to findinggeodesics with respect to a singular (subRiemannian) metric. The geome-try of such geodesic flows is exceptionally rich and provides guidance fordesigning control laws. While the mechanical complexity of systems with agreat number of nonlinearly coupled degrees of freedom will present com-putational challenges, the underlying geometry will nevertheless be usefulin the solution of interpolation problems.

One of the aims of this book is to illustrate the elegant mathematicsbehind many simple, interesting, and useful mechanical examples. Amongthese are the rigid body and rolling rigid body, the rolling ball on a ro-tating turntable, the rattleback top, the rolling penny, the satellite withmomentum wheels, and the rigid body in a perfect fluid. There are clearlya number of points in common between these systems, among them thefact the rotational motion and existence of constraints, either externallyimposed or dynamically generated (conserved momenta) play a key role.In a sense these notions – rotatation and constaints form heart of the bookand are vital to studying both the dynamics and control of these systems.Further, one of the delights of this subject is that the behavior of these sys-tems is that the although they may have many features in common theirbehavior is quite different and often quite unexpected. Why does a rattle-back top rotate in only one direction? What is the behavior of a ball ona rotating turntable? Why does a tennis racket not want to spin about itsmiddle axis? How do I role a penny to a particular point on a table, parallelto an edge, and with Lincoln’s head in the upright position?

We remark finally that while we have attempted to cover quite a lot ofmaterial here, this text is very much written from the authors’ perspective,and there much fascinating work in this area that we have had to omit.

Notes

• Section 3.1, 3.2 should have more on Newton’s law and inertial frames.Tony, Peter, if you have a section on this from your Brockett Festshriftpaper, send it along and I will plug it in, even as a placeholder.

• Peter plans on looking at 7.6.1 – dynamic optimal control

• Chapter 9 needs a lot of blending, eg, with the work of Bloch, Leonardand Marsden, etc. We also need at least some stuff on some other

Contents 1

really key developments in stabilization etc.: eg. the work of Coron/Sontag/ Malkin/Van der Schaft. Some of us have notes on this.

• An internet supplement should be kept in mind as the book growstoo much more. It should never get beyond 500 pages. At this point,anything that goes in, something has to come out and go in theinternet supplement. Good practice at prioritizing.

• We need a bit more on stability, e.g. LaSalle.

2 Contents

This is page 3Printer: Opaque this

1Introduction

We begin this book in a concrete fashion by describing some examples (forthe most part these are physical examples, but not in all cases), which weshall use throughout the course of the book to illustrate the theory. Theseexamples are simple to write down in general and to understand at anelementary level, but they are also useful for the understanding of deeperparts of the theory.

The rolling disk and ball1 are archetypal nonholonomic systems: systemswith constraints on their velocities. The free rigid body and the somewhatmore complex satellite with momentum wheels are examples of free andand coupled rigid body motion respectively—the motion of bodies withnontrivial spatial extent, as opposed to the motion of point particles. Thelatter is illustrated by the Toda lattice, which models a set of interactingparticles on the line; we shall also be interested in some associated optimalcontrol systems.

1These examples have long history going back to Vierkandt, A. [1892] Uber gleitendeund rollende Bewegung. Monatshefte der Math. und Phys. III, 31–54 and Chaplygin,S.A. [1897] On the motion of a heavy body of revolution on a horizontal plane (inRussian). Physics Section of the Imperial Society of Friends of Physics, Anthropologyand Ethnographics, Moscow 9, 10–16.

4 1. Introduction

The Heisenberg system2 does not model any particular physical systemexactly, but is a prototypical example for nonlinear control problems (bothoptimal and nonoptimal) and can be viewed as an approximation to anumber of interesting physical systems; in particular, this example is basicfor understanding more sophiticated optimal reorientation and locomotionproblems, such as the falling cat theorem that we shall treat later. A keypoint about this system (and many others in this book) is that the corre-sponding linear theory gives little information.

1.1 The Rolling Disk

1.1.1 The Vertical Rolling Disk

Geometry and Kinematics of the Vertical Disk. We consider herea useful example of a system subject to nonholonomic constraints—a ho-mogeneous disk rolling without slipping on a horizontal plane. In the firstinstance we consider the “vertical” disk, a disk, which, unphysically ofcourse, may not tilt away from the vertical; it is not difficult to general-ize the situation to the “falling” disk. It is helpful to think of a coin suchas the penny, since we are concerned with orientation of the disk and theroll angle (the position of Lincoln’s head, for example). See for examplethe treatments in Bloch, Reyhanoglu and McClamroch [1992], Bloch andCrouch [1995] and Bloch, Krishnaprasad, Marsden and Murray [1996].3

The configuration space for the vertical rolling disk is Q = R2×S1×S1

and is parameterized by coordinates q = (x, y, θ, ϕ), denoting the positionof the contact point in the xy-plane, the rotation angle of the disk, and theorientation of the disk, respectively, as in figure 1.1.1.

The Lagrangian for the system is taken to be the kinetic energy

L(x, y, θ, φ, x, y, θ, φ) =12m(x2 + y2) +

12Iθ2 +

12Jϕ2 (1.1.1)

where m is the mass of the disk, I is the moment of inertia of the diskabout the axis perpendicular to the plane of the disk and J is the moment

2This system was first studied by Brockett, R.W. [1981] Control theory and singularRiemannian geometry, in New Directions in Applied Mathematics, P.J. Hilton and G.S.Young (eds.), Springer-Verlag. See also J. Ballieul [1975], Some Optimization Problemsin Geometric Control Theory, Thesis, Harvard University.

3Bloch, A.M., M. Reyhanoglu and N.H. McClamroch [1992], Control and stabilizationof nonholonomic systems. IEEE Trans. on Automatic Control 37, 1746–1757, Bloch,A.M., and P.E. Crouch [1995], Nonholonomic control systems on Riemmanian manifolds.SIAM J. on Control and Optimization 37, 126–148, Bloch, A.M., P.S. Krishnaprasad,J.E. Marsden, & R. Murray [1996] Nonholonomic Mechanical Systems with Symmetry.Arch. Rat. Mech. An., 136, 21–99.

1.1 The Rolling Disk 5

ϕ

P

x

z

y

(x, y)

θ

P0

Figure 1.1.1. The geometry for the rolling disk.

of intertia about an axis in the plane of the disk (both axes passing throughthe disk’s center).

If R is the radius of the disk, the nonholonomic constraints of rollingwithout slipping are

x = R(cosϕ)θy = R(sinϕ)θ, (1.1.2)

which state that the point P0 fixed on the disk has zero velocity when itmakes contact with the horizontal plane.

Dynamics of the Controlled Disk. Consider the case where we havetwo controls, one that can steer the disk and another that determines theroll torque. One of the things we will develop later is the general theory thatallows one to formulate the equations of motion in the presence of controlforces. While there are many ways of formulating these, as we shall see, oneway adds the forces to the right hand side of the Euler-Lagrange equationsfor the given Lagrangian as well as Lagrange multipliers to enforce theconstraints. The resulting equations are called the Lagrange d’Alembertequations.

In our case, L is cyclic in the configuration variables q = (x, y, θ, φ), andso these equations become

d

dt

(∂L

∂q

)= u1X1 + u2X2 + λ1W1 + λ2W2, (1.1.3)

where

∂L

∂q= (mx,my, Iθ, Jϕ)T ,

6 1. Introduction

X1 = (0, 0, 1, 0)T , X2 = (0, 0, 0, 1)T ,

and

W1 = (1, 0,−R cosϕ, 0),W2 = (0, 1,−R sinϕ, 0)T ,

together with the constraint equations (1.1.2). Here the superscript T de-notes transpose, u1 and u2 are control functions and the λi are Lagrangemultipliers, chosen to ensure satisfaction of the constraints.

We may now eliminate the multilpliers; solving for λ1 and λ2 from (1.1.3),we get

λ1 = md

dt(R cosϕθ)

λ2 = md

dt(R sinϕθ) .

Substitution of these expressions into the last two components of (1.1.3)gives the dynamic equations

(I +mR2)θ = u1

Jϕ = u2 , (1.1.4)

plus the constraints

x = R(cosϕ)θy = R(sinϕ)θ .

The free equations, in which we set u1 = u2 = 0, are easily integrated.The first two equations in (1.1.3) show that θ and ϕ are constants; callingthese constants Ω and ω respectively, we have

θ = Ωt+ θ0

ϕ = ωt+ φ0

x =ΩωR sin(ωT + ϕ0) + x0

y = −ΩωR cos(ωt+ ϕ0) + y0 .

Consider next the controlled case, in which the controls u1, u2 need notbe zero. Call the variables θ and φ “base” or “controlled” variables and thevariables x and y “fiber” variables. The distinction is that while θ and ϕare controlled directly, the variables x and y are controlled indirectly viathe constraints.4

4The notation “base” and “fiber” comes from the fact that the configuration spaceQ splits naturally into the base and fiber of a trivial fiber bundle, as we shall see later.

1.1 The Rolling Disk 7

It is clear that the base variables are controllable in any sense we canimagine. One may ask whether the full system is controllable. Indeed it is,in a precise sense as we shall show later, by virtue of the nonholonomicnature of the constraints.

The Kinematic Controlled Disk. It is also useful to define the so-called “kinematic” controlled rolling disk. In this case we imagine we havedirect control over velocities rather than forces and, accordingly, we con-sider the most general first order system satisfying the constraints or lyingin the “constraint distribution”.

This system is

q = u1X1 + u2X2 (1.1.5)

where X1 = (cosϕ, sinϕ, 1, 0)T and X2 = (0, 0, 0, 1)T .In fact, X1 and X2 comprise a maximal set of independent vector fields

on Q satisfying the constraints. As we shall see it is most instructive toanalyze both the control and optimal control of such systems.

The Nonholonomic and the Variational Systems. It is interestingto compare the dynamic equations (1.1.3), which can be shown to be consis-tent with Newton’s second law F = ma in the presense of reaction forces—see Vershik and Gershkovich [1988], Bloch and Crouch [1997], Jalnapurkar[1998],5 with the corresponding variational system. The distinction betweenthese two possible formulations has a long and distinguished history goingback to the review article of Korteweg [1898] and is discussed in a moremodern context in Arnold [1988]. The upshot of the distinction is that theequations (1.1.3) are the right dynamical equations while the correspondingvariational problem is asking a different question, one of optimal control.

Perhaps it is surprising, at least at first, that these two procedures givedifferent equations — what is the difference in the two procedures? Theanswer is that with the dynamic Lagrange d’Alembert equations, we imposeconstraints only on the variations, whereas in the variational problem weimpose the constraints on the velocity vectors of the class of allowablecurves. We will show explicitly in this example that one gets two differentsets of equations.

The variational system is obtained by using Lagrange multipliers withthe Lagrangian rather than Lagrange multipliers with the equations, as we

5Vershik, A. M. and v. Ya. Gershkovich [1988] Nonholonomic problems and the theoryof distributions. Acta Applicandae Mathematica 12, 181–209, Bloch, A. M. and P.E.Crouch [1997], Newton’s Law and integrability of nonholonomic systems, to appearin The SIAM Journal on Optimal Control, Jalnapurkar [1998], Thesis, Department ofMathematics, UC Berkeley, 1999.

8 1. Introduction

did earlier. Namely, we consider the Lagrangian

L =12m(x2 + y2) +

12Iθ2 +

12ϕ2 + µ1(x−Rθ cosϕ) + µ2(x−Rθ sinϕ) ,

(1.1.6)

where now, because of the Lagrange multipliers, we can relax the con-straints and take variations over all curves. In other words, we can simplywrite down the Euler-Lagrange equations for this Lagrangian.

The variational equations with external forces therefore are

mx+ µ1 = 0 (1.1.7)my + µ2 = 0 (1.1.8)

Iθ −R d

dt(µ1 cosϕ+ µ2 sinϕ) = u1 (1.1.9)

Jϕ+R∂d

∂ϕ(µ1θ cosϕ+ µ2θ sinϕ) = u2 . (1.1.10)

From the constraint equations and equations (1.1.7) and (1.1.8), we have

µ1 = −mRθ cosφ+A

µ2 = −mRθ sinφ+B,

where A and B are constants.Substituting these into equations (1.1.9) and (1.1.10), we obtain

(I +mR2)θ = Rϕ(−A sinϕ+B cosϕ) + u1 (1.1.11)Jϕ = Rθ(A sinϕ−B cosϕ) + u2 . (1.1.12)

These equations, together with the constraints, define the dynamics. No-tice that they are different from the dynamic (Lagrange d’Alembert) equa-tions. As we have indicated, the motion determined by these equations isnot that associated with physical dynamics in general, but is a model of thetype of problem that is relevant to optimal control problems, as we shallsee later.

1.1.2 The Falling Rolling Disk

A more realistic disc is of course one that is allowed to fall over (deviatefrom the vertical). This turns out to be a very useful example to analyze.See Figure 1.1.2.

As the figure indicates, we denote the coordinates of contact of the diskin the xy-plane by (x, y) and let θ, φ, and ψ denote the angle between theplane of the disk and the vertical axis, the “heading angle” of the disk, and“self-rotation” angle of the disk respectively.

1.1 The Rolling Disk 9

ϕ

P

Qx

z

y

θ

(x, y)

ψ

Figure 1.1.2. The geometry for the rolling disk.

A classical reference for the rolling disk is Vierkandt [1892] who showedthat on the reduced space D/SE(2)—the constrained velocity phase spacemodulo the action of the Euclidean group SE(2)—all orbits of the systemare periodic. Modern references that treat this example are Hermans [1995]and O’Reilly [1996]. 6

For the moment, we just give the Lagrangian and constraints, and returnto this example later on. As we will eventually show, this is a system whichexhibits stability but not asymptotic stability. Denote the mass, the radius,and the moments of inertia of the disk by m, R, A, B respectively. TheLagrangian is given by the kinetic minus potential energies:

L =m

2

[(ξ −R(φ sin θ + ψ))2 + η2 sin2 θ + (η cos θ +Rθ)2

]+

12

[A(θ2 + φ2 cos2 θ) +B(φ sin θ + ψ)2

]−mgR cos θ,

where ξ = x cosφ + y sinφ + Rψ and η = −x sinφ + y cosφ, while theconstraints are given by

x = −ψR cosφ,

y = −ψR sinφ.

Note that the constraints may also be written as ξ = 0, η = 0.

6See Vierkandt, A. [1892] Uber gleitende und rollende Bewegung. Monatshefte derMath. und Phys. III, 31–54, Hermans, J. [1995a], Rolling Rigid Bodies, with and with-out Symmetries, Thesis, University of Utrecht, O’Reilly, O.M. [1996] The Dynamics ofRolling Disks and Sliding Disks. Nonlinear Dynamics, 10, 287–305, Zenkov, D.V., A.M.Bloch, and J.E. Marsden [1998] The Energy Momentum Method for the Stability ofNonholonomic Systems Dyn. Stab. of Systems., 13, 123–166.

10 1. Introduction

1.2 The Heisenberg System

The Heisenberg Algebra. The Heisenberg alegbra is the algebra onemeets in quantum mechanics, wherein one has two operators q and pthat have a nontrivial commutator, in this case a multiple of the iden-tity. Thereby, one generates a three dimensional Lie algebra. The systemstudied in this section has an associated Lie algebra with a similar struc-ture, which is the reason the system is called the Heisenberg system. Thereis no intended relation to quantum mechanics per se other than this.

In Lie algebra theory this sort of a Lie algebra is of considerable interest—one refers to it as an example of a central extension because the elementthat one extends by (in this case a multiple of the identity) is in the centerof the algebra; that is, it commutes with all elements of the algebra.

The Heisenberg system has played a significant role as an example inboth nonlinear control and nonholonomic mechanics.

The Dynamic Heisenberg System. As with the previous example,the system comes in two forms, one associated with the Lagrange d’Alem-bert principle and one with an optimal control problem. As in the previousexamples, the equations in each case are different.

In the dynamic setting, we consider the following standard kinetic energyLagrangian on Euclidean three space R3:

L =12

(x2 + y2 + z2)

subject to the constraint

z = yx− xy. (1.2.1)

Controls u1 and u2 are given in the x and y directions. Letting q =(x, y, z)T , the dynamic nonholonomic control system is7

q = u1X1 + u2X2 + λW (1.2.2)

where X1 = (1, 0, 0)T and X2 = (0, 1, 0)T and W = (−y, x, 1)T . Eliminat-ing λ we obtain the dynamic equations

(1 + x2 + y2)x = (1 + x2)u1 + xyu2

(1 + x2 + y2)y = (1 + y2)u1 + xyu1

(1 + x2 + y2)z = yu1 − xu2 . (1.2.3)

7This example with controls was analyzed in Bloch and Crouch [1993] Nonholonomicand vakonomic control systems on Riemannian manifolds. Fields Institute Communica-tions, 1, 25-52. A related nonholonomic system, but with slightly different constraintsmay be found in Rosenberg, R.M. [1977] Analytical Dynamics of Discrete Systems.Plenum Press, NY., Bates, L. and J. Sniatycki [1993] Nonholonomic reduction, Reportson Math. Phys. 32, 99–115 and Bloch, A.M., P.S. Krishnaprasad, J.E. Marsden, andR. Murray [1996] Nonholonomic mechanical systems with symmetry. Arch. Rat. Mech.An., 136, 21–99.

1.2 The Heisenberg System 11

Optimal Control for the Heisenberg System. The control and opti-mal control of the corresponding kinematic problem has been quite impor-tant historically and we shall return to it later on in the book in connectionwith, for example, the falling cat problem and optimal steering problems.8

The system may be written as

q = u1g1 + u2g2 (1.2.4)

where g1 = (1, 0, y)T and g2 = (0, 1,−x)T . As in the rolling disk exampleg1 and g2 are a maximal set of independent vector fields satisfying theconstraint

z = xy − yx . (1.2.5)

Written out in full, these equations are

x = u1 (1.2.6)y = u2 (1.2.7)z = xu2 − yu1 (1.2.8)

One verifies that the Jacobi-Lie bracket of the vector fields g1 and g2 is

[g1, g2] = 2g3

where g3 = (0, 0, 1). In fact, the three vector fields g1, g2, g3 span all of R3

and, as a Lie algebra, is just the Heisenberg algebra described earlier.By general controllability theorems that we shall discuss in Chapter 3

(Chow’s theorem), one knows that one can, with suitable controls, steertrajectories between any two points in R3. In particular, we are interestedin the following optimal steering problem (see Figure 1.2):

Optimal Steering Problem. Given a number a > 0, findtime dependent controls u1, u2 that steer the trajectory startingat (0, 0, 0) at time t = 0 to the point (0, 0, a) after a given timeT > 0 and that, amongst all such controls, minimizes

12

∫ T

0

(u21 + u2

2) dt .

Figure 1.2.1. An optimal steering problem.

An equivalent formulation is the following: minimize the integral

12

∫ T

0

(x2 + y2) dt

8As we mentioned earlier, this example was introduced in Brockett [1981]

12 1. Introduction

amongst all curves x(t) joining x(0) = (0, 0, 0) to x(T ) = (0, 0, a) thatsatisfy the constraint

z = yx− xy .

As before, any solution must satisfy the Euler-Lagrange equations for theLagrangian with a Lagrange multiplier inserted:

L(x, x, y, y, z, z, λ, λ

)=

12(x2 + y2

)+ λ (z − yx+ xy)

The corresponding Euler-Lagrange equations are given by

x− 2λy = 0 (1.2.9)y + 2λx = 0 (1.2.10)

λ = 0 (1.2.11)

From the third equation, λ is a constant, and the first two equations statethat the particle (x(t), y(t)) moves in the plane in a constant magnetic field(pointing in the z direction, with charge proportional to the constant λ. Formore on these ideas see the chapter on optimal control.

Some remarks are in order here:

1. The fact that this optimal steering problem gives rise to an interestingmechanical system is not an accident; we shall see this in much moregenerality later.

2. Since particles in constant magnetic fields move in circles with con-stant speed, they have a sinusoidal time dependence, and hence so dothe controls. This has led to the “steering by sinusoids” approach inmany nonholonomic steering problems (see for example Murray andSastry [1993] 9 ).

The equations (1.2.9) and (1.2.10) are linear first order equations in thevelocities and are readily solved:[

x(t)y(t)

]=[

cos(2λt) sin(2λt)− sin(2λt) cos(2λt)

] [x(0)y(0)

](1.2.12)

Integrating once more and using the initial conditions x(0) = 0, y(0) = 0gives: [

x(t)y(t)

]=

12λ

[cos(2λt)− 1 sin(2λt)− sin(2λt) cos(2λt)− 1

] [−y(0)x(0)

](1.2.13)

9Murray and Sastry [1993], Nonholonomic motion planning: steering using sinusoids.IEEE Trans on Automatic Control 38, 700-716

1.3 The Toda Lattice 13

The other boundary condition x(T ) = 0, y(T ) = 0 gives

λ =nπ

T.

Using this information, we find z by integration: from z = xy− yx and thepreceding expressions, we get

z(t) =1

2λ[−x(0)2 − y(0)2 + cos(2λt)(x(0)2 + y(0)2)

].

Integration from 0 to T and using z(0) = 0 gives

z(T ) =T

2λ[−x(0)2 − y(0)2

].

Thus, to achieve the boundary condition z(T ) = a one must choose

x(0)2 + y(0)2 = −2πnaT 2

.

One also finds that

12

∫ T

0

[x(t)2 + y(t)2

]dt =

12

∫ T

0

[x(0)2 + y(0)2

]dt

=T

2[x(0)2 + y(0)2

]= −πna

T

so that the minimum is achieved when n = −1.

Summary: The solution of the optimal control problem is given by choos-ing initial conditions such that x(0)2 + y(0)2 = 2πa/T 2 and with the tra-jectory in the xy plane given by the circle[

x(t)y(t)

]=

12λ

[cos(2πt/T )− 1 − sin(2πt/T )

sin(2πt/T ) cos(2πt/T )− 1

] [−y(0)x(0)

](1.2.14)

and with z given by

z(t) =ta

T− ta2 sin

(2πtT

).

Notice that any such solution can be rotated about the z axis to obtainanother one.

1.3 The Toda Lattice

An important and beautiful mechanical system that describes the interac-tion of particles on the line (i.e. in one dimension) is the Toda lattice.

14 1. Introduction

We shall describe here the nonperiodic finite Toda lattice as analyzed byMoser [1975]10

The model consists of n- particles moving freely on the x–axis and in-teracting under an exponential potential. Denote the position of the kthparticle by xk. The Hamiltonian is given by

H(x, y) =12

n∑k=1

y2k +

n−1∑k−1

e(xk−xk+1) .

The associated Hamiltonian equations are

xk =∂H

∂yk= yk (1.3.1)

yk = − ∂H∂xk

= exk−1−xk − exk−xk−1 , (1.3.2)

where we use the convention ex0−x1 = exn−xn+1 = 0, which corresponds toformally setting x0 = −∞ and xn+1 = +∞.

This system of equations has an extraordinarily rich structure. Part ofthis is revealed by Flaschka’s ([1974] 11) observation that a change of vari-ables enables one to write the equation in Lax pair form. To achieve this,set

ak =12e(xk−xk+1)/2 bk = −1

2yk . (1.3.3)

The equations of motion then become

ak = ak(bk+1 − bk) , k = 1, · · · , n− 1 (1.3.4)bk = 2(a2

k − a2k−1) , k = 1, · · · , n (1.3.5)

with the boundary conditions a0 = an = 0. This system may be written inthe matrix form

d

dtL = [B,L] = BL− LB, (1.3.6)

where

L =

b1 a1 0 · · · 0a1 b2 a2 · · · 0

. . .bn−1 an−1

0 an−1 bn

B =

0 a1 0 · · · 0−a1 0 a2 · · · 0

. . .0 an−1

0 an−1 0

10Moser, J. [1975] Finitely many mass points on the under the influence of an expo-

nential potential – an integrable system. Springer Lecture Notes in Physics 38, 467-497,Moser, J. [1975b] Three integrable Hamiltonian systems connected with isospectral de-formations. Adv. Math. 16, 197–220.

11Flaschka, H. [1974] The Toda Lattice. Phys. Rev. B 9, 1924-25.

1.4 The Rigid Body 15

If O(t) is the orthogonal matrix solving the equation

d

dtO = BO , O(0) = Identity

then from (1.3.6), we have

d

dt(O−1LO) = 0 .

Thus, O−1LO = L(0), i.e. L(t) is related to L(0) by a similarity trans-formation and thus the eigenvalues of L, which are real and distinct, arepreserved along the flow. This is enough to show that in fact this systemis explicitly solvable or integrable.

There is, however, much more structure in this example. For instance, ifN is the matrix diag[1, 2, · · · , n] the Toda flow (1.3.6) may be written inthe form12

L = [L, [L,N ]] . (1.3.7)

This shows that the flow is also gradient (on a level set of its integrals).This is explicitly exhibited by writing the equation in the so-called doublebracket form of Brockett [1988].13

1.4 The Rigid Body

1.4.1 The Free Rigid Body

A key system in mechanics is the free rigid body. There are many excellenttreatments of this topic – see for example Whittaker Arnold [1978], andMarsden and Ratiu [1994]. 14 We restrict ourselves here to some essentials,

12See Bloch A. M. [1990] Steepest descent, linear programming and Hamiltonian flows.Contemp. Math. A.M.S. 114, 77-88, and Bloch, Brockett and Ratiu [1992]; Bloch, A.M.,R.W. Brockett, and T.S. Ratiu [1992] Completely integrable gradient flows. Comm.Math. Phys. 147, 57–74.

13Brockett, R.W. [1988] Dynamical systems that sort lists and solve linear program-ming problems. Proc. IEEE 27, 799–803 and Linear Algebra and its Appl. 146, (1991),79–91.

14Whittaker, E.T. [1937] A treatise on the analytical dynamics of particles and rigidbodies, Cambridge University Press, 4th Ed. (reprinted by Dover 1944, and CambridgeUniversity 1988.), Arnold, V. I. [1978] Mathematical Methods of Classical MechanicsGraduate Texts in Math. 60, Springer Verlag. (Second Edition, 1989), Marsden, J.E.and T.S. Ratiu [1994] Introduction to Mechanics and Symmetry. Texts in Applied Math-ematics, 17, Springer-Verlag.

16 1. Introduction

although we shall return to it in detail in the context of nonholonomicmechanics and optimal control.

The configuration space of a rigid body moving freely in space is R3 ×SO(3) – describing the position of a coordinate frame fixed in the body andthe orientation of the frame say, the orientation of the frame given by aelement of of SO(3), i.e., an orthogonal 3 by 3 matrix with determinant 1.Since the the three components of translational momentum are conserved,the body behaves as if it were rotating freely about its center of mass.15

Hence the phase space for the body may be taken to be TSO(3) – thetangent bundle of SO(3) – with points representing the position and ve-locity of the body, or in the Hamiltonian context we may choose the phasespace to be the cotangent bundle T ∗SO(3), with points representing theposition and momentum of the body. (This example may be equally wellformulated for the group SO(n) or indeed any compact Lie group.)

If I is the moment of inertia tensor computed with respect to a bodyfixed frame, which, in a principal body frame, we may represent by thediagonal matrix diag(I1, I2, I3), the Lagrangian of the body is given by thekinetic energy, namely

L =12

Ω · IΩ, (1.4.1)

where Ω is the vector of angular velocities computed with respect to theaxes fixed in the body.

The Euler-Lagrange equations of motion may be written as

A = AΩ (1.4.2)IΩ = IΩ× Ω , (1.4.3)

where A ∈ SO(3). Writing

Ω ≡

0 −Ω3 Ω2

Ω3 0 −Ω1

−Ω2 Ω1 0

,

the dynamics may be rewritten

I˙Ω = [IΩ, Ω] (1.4.4)

or, in terms of the angular momentum M = IΩ

˙M = [M, Ω] . (1.4.5)

15This is not the case with other systems, such as a rigid body moving in a fluid;even though the system is translation invariant, its “center of mass” need not move ona straight line, so the configuration space must be taken to be the full Euclidean group.

1.4 The Rigid Body 17

1.4.2 Example: The Rolling Ball

We consider here the controlled rolling inhomogeneous ball on the plane,the kinematics of which were discussed in Brockett and Dai [1991]16, estab-lishing the completely nonholonomic nature of the constraint distributionH. (A distribution is completely nonholonomic if the span of the iteratedbrackets of the vector fields lying in it has dimension equal to the dimensionof the underlying manifold – see Chapter 4 for a full explanation.) The dy-namics of the uncontrolled system are described for example in McMillan[1936] (see also Bloch, Krishnaprasad, Marsden and Murray [1996], Jurd-jevic [1993], Koon and Marsden [1997] and Krishnaprasad and Dayawansa[1992]). 17 We will use the coordinates x, y for the linear horizontal dis-placement and P ∈ SO(3) for the angular displacement of the ball. ThusP gives the orientation of the ball with respect to inertial axes e1, e2, e3

fixed in the plane, where the ei are the standard basis vectors aligned withthe x,y and z axes respectively. See figure 1.4.1

ω

(x, y)Ω

Figure 1.4.1. The rolling ball.

Let the ball have radius a and mass m and let ω ∈ R3 denote the angularvelocity of the ball with respect to inertial axes. In particular, the ball mayspin freely about the z-axis and the z-component of angular momentum isconserved. If J denotes the inertia tensor of the ball with respect to thebody axes ( i.e. fixed in the body), then J = PTJP denotes the inertiatensor of the ball with respect to the inertial axes (i.e. fixed in space) and

16Brockett, R.W. & L. Dai [1992] Nonholonomic kinematics and the role of ellipticfunctions in constructive controllability, in Nonholonomic Motion Planning, eds. Z. Li& J. F. Canny , Kluwer, 1–22, 1993.

17McMillan, W. D. [1936] Dynamics of Rigid Bodies, Duncan MacMillan Rowles, U.K.,Bloch, A.M., P.S. Krishnaprasad, J.E. Marsden and R. Murray [1996] NonholonomicMechanical Systems with Symmetry. Arch. Rat. Mech. An., 136, 21–99, Jurdjevic, V.[1993] The geometry of the plate-ball problem. Arch. Rat. Mech. An. 124, 305–328.Krishnaprasad, P.S., W. Dayawansa & R. Yang [1992] The geometry of nonholonomicconstraints. (Preprint, University of Maryland , Koon, W.S. and J.E. Marsden [1997] TheHamiltonian and Lagrangian Approaches to the Dynamics of Nonholonomic Systems,Reports on Math Phys (to appear).

18 1. Introduction

Jω is the angular momentum of the ball with respect to the inertial axes.The conservation law alluded to above is expressed as

eT3 Jω = c . (1.4.6)

The nonholonomic constraints may be expressed as

aeT2 ω + x = 0aeT1 ω − y = 0 . (1.4.7)

We may express the kinematics for the rotating ball as P = ΩP whereΩ = Pω is the angular velocity in the body frame.

Appending the constraints via Lagrange multipliers we obtain the equa-tions of motion

ˆΩP − (J−1ΩJΩ)P

= λ1(J−1Pe1)P + λ2

(J−1Pe2)Pmx = λ2 + u1

my = −λ1 + u2 . (1.4.8)

As mentioned before, the details of the derivation of these equations willbe discussed later.

Using inertial coordinates ω = PTΩ, the system becomes

ω = J−1ωJω + λ1J−1e1 + λ2J−1e2

mx = λ2 + u1

my = −λ1 + u2

P = Pω . (1.4.9)

Also, from the constraints and the constants of motion we obtain thefollowing expression for ω:

ω = x(α2e3 − e2) + y(e1 − α1e3) + α3e3

where

α1 =eT3 Je1

aeT3 Je3, α2 =

eT3 Je2

aeT3 Je3, α3 =

c

eT3 Je3. (1.4.10)

Then the equations become

mx = λ2 + u1

my = −λ1 + u2 (1.4.11)

P = P (x(α2e3 − e2) + y(e1 − α1e3) + α3e3)

1.4 The Rigid Body 19

One can now eliminate the multipliers using the first three equations of1.4.9 and the constraints. The resulting expressions are a little complicatedin the general case (although can be found in straighforward fashion), butbecome pleasingly simple in the case of a homogeneous ball where sayJ = mk2 (k is called the radius of gyration in the classical literature).

In the latter case the equations of motion for ω1 and ω2 become simply

mk2ω1 = aλ1

mk2ω2 = aλ2 . (1.4.12)

Rewriting these equations in terms of x and y using the multipliers,and substituting the resulting expressions for the λi into the equations ofmotion for x and y yields the equations

mx =a2

a2 + k2u1

my =a2

a2 + k2u2 . (1.4.13)

.A similar elimination argument works in the general case.Note that the homogeous ball moves under the action of external forces

like a point mass located at its center but with force reduced by the ratioa2/a2 + k2 – see also the following subsection.

1.4.3 A Homogeneous Ball on a Rotating Plate

A useful example is a model of a homogeneous ball on a rotating plate(see Neimark and Fufaev [1972] and Yang [1992] for the affine case and, forexample, Bloch and Crouch [1992], Brockett and Dai [1992] and Jurdjevic[1993] for the linear case). As we mentioned earlier, Chaplygin [1897b,1903]studied the motion of a nonhomogeneous rolling ball.

Let the plane rotate with constant angular velocity Ω about the z-axis.The configuration space of the sphere is Q = R2 × SO(3), parameterizedby (x, y,R), R ∈ SO(3), all measured with respect to the inertial frame.Let ω = (ωx, ωy, ωz) be the angular velocity vector of the sphere measuredalso with respect to the inertial frame, let m be the mass of the sphere,mk2 its inertia about any axis, and let a be its radius.

The Lagrangian of the system is

L =12m(x2 + y2) +

12mk2(ωx2 + ωy

2 + ωz2) (1.4.14)

with the affine nonholonomic constraints

x− aωy = −Ωy

y + aωx = Ωx.

(1.4.15)

20 1. Introduction

Note that the Lagrangian here is a metric on Q which is bi-invariant onSO(3) as the ball is homogeneous. Note also that R2×SO(3) is a principalbundle over R2 with respect to the right SO(3) action on Q given by

(x, y,R) 7→ (x, y,RS) (1.4.16)

for S ∈ SO(3). The action is on the right since the symmetry is a materialsymmetry.

The equations of motion can be shown to reduce to

x+k2Ω

a2 + k2y = 0

y − k2Ωa2 + k2

x = 0.

(1.4.17)

These equations are easily integrated to show that ball simply oscillateson the plate between two circles rather than flying off as one might expect.

Set

α =k2Ω

a2 + k2.

Then one can see the equations are equivalent to...x + α2x = 0 (1.4.18)...y + α2y = 0 . (1.4.19)

Hence

x = A cosαt+B sinαt+ C

for constants A,B,C depending on the initial data and similarly for y.

1.4.4 The inverted pendulum on a cart

A useful classical systems for testing control theortic ideas is the invertedpendulum on a cart – the goal being to stabilize the pendulum about thevertical using a force acting on the cart. In this book we will use this systemto illustrate stabilization using the energy methods as discussed in Bloch,Marsden and Sanchez de Alvarez [1997] and Bloch, Marsden and Leonard[1997]. Here we just write down the equations of motion.

First, we compute the Lagrangian for the cart-pendulum system. Let sdenote the position of the cart on the s-axis and let θ denote the angle ofthe pendulum with the upright vertical, as in the figure.

Here, the configuration space is Q = G×S = R×S1 with the first factorbeing the cart position s, and the second factor being the pendulum angle,θ. The velocity phase space, TQ has coordinates (s, θ, s, θ).

1.4 The Rigid Body 21

s

θ

m

l

g

M

l = pendulum length

m = pendulum bob mass

M = cart mass

g = acceleration due to gravity

The velocity of the cart relative to the lab frame is s, while the velocityof the pendulum relative to the lab frame is the vector

vpend = (s+ l cos θ θ,−l sin θ θ). (1.4.20)

The system kinetic energy is just the sum of the kinetic energies of the cartand the pendulum:

K((s, θ, s, θ) =12

(s, θ)(

M +m ml cos θml cos θ ml2

)(s

θ

). (1.4.21)

The Lagrangian is the kinetic minus potential energy, so we get

L(s, θ, s, θ) = K(s, θ, s, θ)− V (θ), (1.4.22)

where the potential energy is V = mgl cos θ.Note that there is a symmetry group G of the pendulum-cart system

– that of translation in the s variable so G = R. We do not destroy thissymmetry when doing stabilization in θ.

For convenience we rewrite the Lagrangian as

L(s, θ, s, θ) =12

(αθ2 + 2β cos θsθ + γs2) +Dcosθ , (1.4.23)

where α = ml2, β = ml, γ = M + m and D = −mgl are constants. Notethat αγ − β2 > 0.

The momentum conjugate to s is ps = γs+ β cos θθ and the momentumconjugate to θ is pθ = αθ + β cos θs.

The relative equilibrium defined by θ = 0, θ = 0 and s = 0 is unstablesince D < 0.

The equations of motion of the cart pendulum system with a controlforce u acting on the cart (and no direct forces acting on the pendulum)

22 1. Introduction

are, since s is a cyclic variable (i.e. L is independent of s),

d

dt

∂L

∂s= u

d

dt

∂L

∂θ− ∂L

∂θ= 0 ,

i.e.,

d

dtps =

d

dt(γs+ β cos θθ) = u

d

dtpθ + β sin θsθ +D sin θ =

d

dt(αθ + β cos θs) + β sin θsθ +D sin θ = 0 . (1.4.24)

1.4.5 The Rotating Pendulum

1.4.6 Rigid Body with Rotors

Following Krishnaprasad [1985] and Bloch, Krishnaprasad, Marsden andSanchez de Alvarez [1992] and Bloch, Leonard and Marsden [1997], weconsider a rigid body with a rotor aligned along the third principal axis ofthe body as in the figure. The rotor spins under the influence of a torque uacting on the rotor. The configuration space is Q = SO(3)× S1, with thefirst factor being the spacecraft attitude and the second factor being therotor angle. The Lagrangian is total kinetic energy of the system, (rigidcarrier plus rotor), with no potential energy.

spinning rotor

rigid carrier

Again this system will be used to illustrate energy method is analyzingstabilization and stability.

1.4 The Rigid Body 23

The Lagrangian for this system is

L =12

(λ1Ω21 + λ2Ω2

2 + I3Ω23 + J3(Ω3 + α)2) (1.4.25)

where I1 > I2 > I3 are the rigid body moments of inertia,J1 = J2 andJ3 are the rotor moments of inertia, Ω = (Ω1,Ω2,Ω3) is the body angularvelocity vector of the carrier and α is the relative angle of the rotor.

The body angular momenta are determined by the Legendre transformto be

Π1 = λ1Ω1

Π2 = λ2Ω2

Π3 = λ3Ω3 + J3α

l3 = J3(Ω3 + α),

λi = Ii + Ji. The momentum conjugate to α is l3.The equations of motion with a control torque u acting on the rotor are

λ1Ω1 = λ2Ω2Ω3 − (λ3Ω3 + J3α)Ω2

λ2Ω2 = −λ1Ω1Ω3 + (λ3Ω3 + J3α)Ω1

λ3Ω3 + J3α = (λ1 − λ2)Ω1Ω2

l3 = u . (1.4.26)

The equations may also be written in Hamiltonian form:

Π1 =(

1I3− 1λ2

)Π2Π3 −

l3Π2

I3

Π2 =(

1λ1− 1I3

)Π1Π3 +

l3Π1

I3

Π3 =(

1λ2− 1λ1

)Π1Π2

.= a3Π1Π2

l3 = u.

Here, λi = Ii + Ji.

24 1. Introduction

This is page 25Printer: Opaque this

2Mathematical preliminaries

In this chapter we discuss the following topics:

1. Vector fields, flows and differential equations

2. Differentiable manifolds

3. Riemannian and subRiemannian Geometry

4. Hamiltonian Systems and Symplectic Geometry

5. Lie Groups.

6. Distributions on Manifolds

7. Fiber Bundles, Connections, and Gauge Theory

Of course this is a vast array of topics and we do not pretend to give allthe needed background for the reader to learn these things from scratch.However, we hope that this summary will be helpful to set the notation, fillin some gaps the reader may have and to provide a guide to the literaturefor needed background and proofs.

2.1 Vector fields, flows and differentialequations

This section introduces vector fields on Euclidean space and the flows theydetermine. This topic puts together and globalizes two basic ideas learned

26 2. Mathematical preliminaries

in undergraduate mathematics: vector fields and differential equations.

A Basic Example. An example that illustrates many of the concepts ofdynamical systems is the ball in a rotating hoop. Refer to figure 2.1.1.

ω < ω0 ω > ω0

Figure 2.1.1. The ball in the hoop system; the equilibrium is stable for ω < ωcand unstable for ω > ωc.

This system consists of a rigid hoop that hangs from the ceiling witha small ball resting in the bottom of the hoop. The hoop rotates withfrequency ω about a vertical axis through its center (Figure2.1.1(left)).

Consider varying ω, keeping the other parameters fixed. For small valuesof ω, the ball stays at the bottom of the hoop and that position is stable.Accept this in an intuitive sense for the moment; one has to eventuallydefine this concept carefully. However, when ω reaches the critical valueω0, this point becomes unstable and the ball rolls up the side of the hoopto a new position x(ω), which is stable. The ball may roll to the left or tothe right, depending, perhaps upon the side of the vertical axis to which itwas initially leaning. (see Figure 2.1.1(right)). The position at the bottomof the hoop is still a fixed point, but it has become unstable. The solutionsto the initial value problem governing the ball’s motion are unique for allvalues of ω. For ω > ω0, we cannot predict which way the ball will roll.

Using principles of mechanics which we shall review later, one can showthat the equations for this system are given by

mR2θ = mR2ω2 sin θ cos θ −mgR sin θ − νRθ,

where R is the radius of the hoop, θ is the angle from the bottom vertical,m is the mass of the ball, g is the acceleration due to gravity, and ν isa coefficient of friction.1 To analyze this system, we use a phase plane

1This does not represent a realistic friction law, but is an ad hoc one for illustrationonly; even for this simple problem friction laws are controversial, depending on the

2.1 Vector fields, flows and differential equations 27

analysis; that is, we write the equation as a system:

x = y

y =g

R(α cosx− 1) sinx− βy

where α = Rω2/g and β = ν/m. This system of equations produces foreach initial point in the xy-plane, a unique trajectory. That is, given a point(x0, y0) there is a unique solution (x(t), y(t)) of the equation that equals(x0, y0) at t = 0. This statement is proved by using general existence anduniqueness theory that we shall discuss later. When we draw these curvesin the plane, we get figures like those shown in Figure 2.1.2.

α = 0.5, β = 0 α = 1.5, β = 0

α = 1.5, β = 0.1

θ

θθ

Figure 2.1.2. The phase portrait for the ball in the hoop before and after theonset of instability for the case g/R = 1.

The above example has system parameters such as g,R, and ω. Inmany problems one takes the point of view, as we have done in the preced-ing discussion, of looking at how the phase portrait of the system changesas the parameters change. Sometimes these parameters are called controlparameters since one imagines changing them. However, this is still a pas-sive point of view since we imagine sitting back and watching the dynamics

exact nature of the mechanical system. If one were to suppose Coulomb friction onewould make the tangential force proportional to the normal force.

28 2. Mathematical preliminaries

unfold for each value of the parameters. In control theory we take a moreactive point of view, and try to directly intervene with the dynamics toachieve a desired end. For example, we might imagine manipulating ω as afunction of time to make the ball move in a desired way.

Dynamical Systems. More generally than in the above example, inEuclidean space Rn whose points are denoted by x = (x1, . . . , xn), we areconcerned with a system of the form

x = F (x) (2.1.1)

which, in components, reads

xi = F i(x1, . . . , xn), i = 1, . . . , n

where F is an n-component function of the n variables x1, . . . , xn. Some-times the function, or vector field F depends on time or on other parameters(such as the mass, angular velocity etc in the example) and keeping trackof this explicitly is important as we shall see momentarily. For general dy-namical systems one needs some theory to develop properties of solutions;roughly, we draw curves in Rn emanating from initial conditions, just aswe

Equilibrium points, stability and bifurcation. Equilibrium pointsare points where the right hand side of the system vanish. Mathematically,we say that the original stable fixed point has become unstable and hassplit into two stable fixed points. One can use some basic stability theorythat we shall develop to show that ω0 =

√g/R. See Figure2.1.2. This is

one of the simplest situations in which symmetric problems can have non-symmetric solutions and in which there can be multiple stable equilibria, sothere is non-uniqueness of equilibria (even though the solution of the initialvalue problem is unique).

This example shows that in some systems the phase portrait can changeas certain parameters are changed. Changes in the qualitative nature ofphase portraits as parameters are varied are called bifurcations. Conse-quently, the corresponding parameters are often called bifurcation pa-rameters. These changes can be simple such as the formation of new fixedpoints—these are called static bifurcations to dynamic bifurcationssuch as the formation of periodic orbits, that is, an orbit x(t) with theproperty that x(t + T ) = x(t) for some T and all t, or more complex dy-namical structures.

An important bifurcation called the Hopf bifurcation, or more properly,the Poincare-Andronov-Hopf bifurcation occurs in this example. This is adynamic bifurcation in which, roughly speaking, a periodic orbit ratherthan another fixed point is formed when an equilibrium looses stability. Aneveryday example of a Hopf bifurcation we all encounter is flutter. Forexample, when venetian blinds flutter in the wind or a television antenna

2.1 Vector fields, flows and differential equations 29

“sings” in the wind, there is probably a Hopf bifurcation occurring. Arelated example that is physically easy to understand is flow through ahose: considers a straight vertical rubber tube conveying fluid. The lowerend is a nozzle from which the fluid escapes. This is called a follower-loadproblem since the water exerts a force on the free end of the tube whichfollows the movement of the tube. Those with any experience in a gardenwill not be surprised by the fact that the hose will begin to oscillate if thewater velocity is high enough.

y

x

y

x

y

x

µ < µ0 µ > µ0µ close to µ0

Figure 2.1.3. A periodic orbit appears for µ close to µ0.

Vector fields. With the above example as motivation, we can beginthe more formal treatment of vector fields and their associated differentialequations. Of course we will eventually add the concept of controls to thesevector fields, but we need to understand the notion of vector field itselffirst.

Definition 2.1.1. Let r ≥ 0 be an integer. A Cr vector field on Rn isa mapping X : U → Rn of class Cr from an open set U ⊂ Rn to Rn. Theset of all Cr vector fields on U is denoted by Xr(U) and the C∞ vectorfields by X∞(U) or X(U).

We think of a vector field as assignning to each point x ∈ U a vectorX(x) based (i.e., bound) at that same point.

Example: Newton’s law of gravitation. Here the set U is R3 minusthe origin and the vector field is defined by

F(x, y, x) = −mMG

r3r,

where m is the mass of a test body, M is the mass of the central body, Gis the constant of gravitation, r is the vector from the origin to (x, y, z),and r = (x2 + y2 + z2)1/2; see Fig. 2.1.4.

30 2. Mathematical preliminaries

Figure 2.1.4. The gravitational force field.

Evolution operators. Consider a general physical system that is ca-pable of assuming various “states” described by points in a set Z. Forexample, Z might be R3 × R3 and a state might be the position and mo-mentum (q,p) of a particle. As time passes, the state evolves. If the stateis z0 ∈ Z at time t0 and this changes to z at a later time t, we set

Ft,t0(z0) = z

and call Ft,t0 the evolution operator ; it maps a state at time t0 to whatthe state would be at time t; i.e., after time t− t0 has elapsed. “Determin-ism” is expressed by the law

Ft2,t1 Ft1,t0 = Ft2,t0 Ft,t = identity,

sometimes called the Chapman-Kolmogorov law .The evolution laws are called time independent when Ft,t0 depends

only on t− t0; i.e.,

Ft,t0 = Fs,s0 if t− t0 = s− s0.

Setting Ft = Ft,0, the preceding law becomes the group property :

Fτ Ft = Fτ+t, F0 = identity.

We call such an Ft a flow and Ft,t0 a time-dependent flow , or an evo-lution operator. If the system is defined only for t ≥ 0, we speak of asemi-flow .

2.1 Vector fields, flows and differential equations 31

It is usually not Ft,t0 that is given, but rather the laws of motion . Inother words, differential equations are given that we must solve to find theflow. In general, Z is a manifold (a generalization of a smooth surface), butwe confine ourselves here to the case that Z = U is an open set in someEuclidean space Rn. These equations of motion have the form

dx

dt= X(x), x(0) = x0

where X is a (possibly time-dependent) vector field on U .

Example: Newton’s Second Law. The motion of a particle of massm under the influence of the gravitational force field is determined byNewton’s second law:

md2rdt2

= F;

i.e., by the ordinary differentatial equations

md2x

dt2= −mMGx

r3;

md2y

dt2= −mMGy

r3;

md2z

dt2= −mMGz

r3;

Letting q = (x, y, z) denote the position and p = m(dr/dt) the momentum,these equations become

dqdt

=pm

;dpdt

= F(q)

The phase space here is the open set U = (R3\0)× R3. The right-handside of the preceding equations define a vector field by

X(q,p) = ((q,p), (p/m,F(q))).

In courses on mechanics or differential equations, it is shown how to in-tegrate these equations explicitly, producing trajectories, which are planarconic sections. These trajectories comprise the flow of the vector field.

Relative to a chosen set of Euclidean coordinates, we can identify a vec-tor field X with an n-component vector function (X1(x), . . . , Xn(x)), thecomponents of X.

Definition 2.1.2. Let U ⊂ Rn be an open set and X ∈ Xr(U) a vectorfield on U . An integral curve of X with initial condition x0 is a differen-tiable curve c defined on some open interval I ⊂ R containing 0 such thatc(0) = x0 and c′(t) = X(c(t)) for each t ∈ I.

32 2. Mathematical preliminaries

Clearly c is an integral curve of X when the following system of ordinarydifferential equations is satisfied:

dc1

dt(t) = X1(c1(t), . . . , cn(t))

......

dcn

dt(t) = Xn(c1(t), . . . , cn(t))

We shall often write x(t) = c(t), an admitted abuse of notation. Thepreceding system of equations are called autonomous, when X is timeindependent. If X were time dependent, time t would appear explicitlyon the right-hand side. As we have already seen, the preceding systemof equations includes equations of higher order by the usual reduction tofirst-order systems.

Existence and uniqueness theorems. One of the basic theorems con-cerning the existence and uniqueness of solutions is the following.

Theorem 2.1.3 (Local Existence, Uniqueness, and Smoothness).Suppose U ⊂ Rn is open and that X is a Cr vector field on U for somer ≥ 1. For each x0 ∈ U , there is a curve c : I → U with c(0) = x0 suchthat c′(t) = X(c(t)) for all t ∈ I. Any two such curves are equal on theintersection of their domains. Furthermore, there is a neighborhood U0 ofthe point x0 ∈ U , a real number a > 0, and a Cr mapping F : U0× I → U ,where I is the open interval ] − a, a [, such that the curve cu : I → U , de-fined by cu(t) = F (u, t) is a curve satisfying cu(0) = u and the differentialequations c′u(t) = X(cu(t)) for all t ∈ I.

This theorem has many variants. We refer to Coddington and Levinston[1955], and Hartman [1982] and Abraham, Marsden and Ratiu [1988] forproofs.2

For example, with just continuity of X one can get existence (the Peanoexistence theorem) without uniqueness. The equation in one dimensiongiven by x =

√x, x(0) = 0 has the two C1 solutions x1(t) = 0 and x2(t)

which is defined to be 0 for t ≤ 0 and x2(t) = t2/4 for t > 0. This shows thatone can indeed have existence without uniqueness for continuous vector

2Coddington, E.A. and N. Levinson [1955] Theory of Ordinary Differential Equa-tions. McGraw-Hill, New York, Hartman, P. [1982] Ordinary Differential Equations,Wiley (1964), Second Edition, Birkjauser, Abraham, R., J.E. Marsden, and T.S. Ratiu[1988] Manifolds, Tensor Analysis, and Applications. Second Edition, Applied Mathe-matical Sciences 75, Springer-Verlag.This last reference also has a proof based directlyon the implicit function theorem applied in suitable function spaces. This proof has atechnical advantage: it works easily for other types of differentiability assumptions on Xor on Ft, such as Holder or Sobolev differentiability; this result is due to Ebin, D.G. andJ.E. Marsden [1970] Groups of diffeomorphisms and the motion of an incompressiblefluid, Ann. Math. 92, 102–163.

2.1 Vector fields, flows and differential equations 33

fields. However, the normal proof starts with a Lipschitz assumption onthe vector field and proceeds to show existence and uniqueness in this caseby showing that there is a unique solution to the integral equation

x(t) = x0 +∫ t

0

X(x(s)) ds.

This is normally done by using the contraction mapping principle on asuitable space of curves or by showing that the sequence of curves given byPicard iteration converges: let x0(t) = x0 and define inductively

xn+1(t) = x0 +∫ t

t0

X(xn(s)) ds.

The existence and uniqueness theory also holds if X depends explicitly ont and/or on a parameter ρ, is jointly continuous in (t, ρ, x), and is Lipschitzor class Cr in x uniformly in t and ρ.

Dependence on initial conditions and parameters. The followinginequality is of basic importance in not only existence and uniqueness the-orems, but also in making estimates on solutions.

Theorem 2.1.4 (Gronwall’s Inequality). Let f, g : [a, b[→ R be con-tinuous and nonnegative. Suppose there is a constant A ≥ 0 such that forall t satisfying a ≤ t ≤ b,

f(t) ≤ A+∫ t

a

f(s) g(s) ds.

Then

f(t) ≤ A exp(∫ t

a

g(s) ds)

for all t ∈ [a, b[.

We refer to the preceding references for the proof. This lemma is the oneof the key ingredients in showing that the solutions depend in a Lipshitz orsmooth way on initial conditions. Specifically, let Ft(x0) denote the solution(= integral curve) of x′(t) = X(x(t)), x(0) = x0. Then for Lipschitz vectorfields, Ft(x) depends in a continuous and indeed Lipschitz, manner on theinitial condition x and is jointly continuous in (t, x). Again, the same resultholds if X depends explicitly on t and on a parameter ρ is jointly continuousin (t, ρ, x), and is Lipschitz in x uniformly in t and ρ. We let F ρt,λ(x) bethe unique integral curve x(t) satisfying x′(t) = X(x(t), t, ρ) and x(λ) =x. Then F ρt,t0(x) is jointly continuous in the variables (t0, t, ρ, x), and isLipschitz in x, uniformly in (t0, t, ρ).

Additional work along these same lines shows that Ft is Cr if X is.Again, there is an analogous result for the evolution operator F ρt,t0(x) for atime-dependent vector field X(x, t, ρ), which depends on extra parametersρ in some other Euclidean space, say Rm . If X is Cr, then F ρt,t0(x) is Cr

in all variables and is Cr+1 in t and t0.

34 2. Mathematical preliminaries

Suspenstion Trick. The variable ρ can be easily dealt with by suspend-ing X to a new vector field obtained by appending the trivial differentialequation ρ′ = 0; this defines a vector field on Rn × Rm and the basic exis-tence and uniqueness theorem may be applied to it. The flow on Rn ×Rmis just Ft(x, ρ) = (F ρt (x), ρ).

An interesting result called the rectification theorem3 shows that near apoint x0 satisfying X(x0) 6= 0, the flow can be transformed by a change ofvariables so that the integral curves become straight lines moving with unitspeed. This shows that, in effect, nothing interesting happens with flowsaway from equilibrium points as long as one looks at the flow only locallyand for short time.

The mapping F gives a locally unique integral curve cu for each u ∈ U0,and for each t ∈ I, Ft = F |(U0 × t) maps U0 to some other set. It isconvenient to think of each point u being allowed to “flow for time t” alongthe integral curve cu (see Fig. 2.1.5). This is a picture of a U0 “flowing,”and the system (U0, a, F ) is a local flow of X, or flow box .

U0Ft (U0)

Figure 2.1.5. The flow of a vector field

Some global questions. The first global issue concerns uniqueness.

Proposition 2.1.5 (Global Uniqueness). Suppose c1 and c2 are twointegral curves of X in U and that for some time t0, c1(t0) = c2(t0). Thenc1 = c2 on the intersection of their domains.

3The proof can be found in Arnold, V. I. [1983] Geometrical Methods in the Theoryof Ordinary Differential Equations. Springer Verlag, and Abraham, R., J.E. Marsden,and T.S. Ratiu [1988] Manifolds, Tensor Analysis, and Applications. Second Edition,Applied Mathematical Sciences 75, Springer-Verlag.

2.1 Vector fields, flows and differential equations 35

Other global issues center on considering the flow of a vector field as awhole, extended as far as possible in the t-variable.

Definition 2.1.6. Given an open set U and a vector field X on U , letDX ⊂ U ×R be the set of (x, t) ∈ U ×R such that there is an integral curvec : I → U of X with c(0) = x with t ∈ I. The vector field X is complete ifDX = U ×R. A point x ∈ U is called σ-complete, where σ = +,−, or ±,if DX ∩ (x × R) contains all (x, t) for t > 0, < 0, or t ∈ R, respectively.Let T+(x) (resp. T−(x)) denote the sup (resp. inf) of the times of existenceof the integral curves through x; T+(x) resp. T−(x) is called the positive(negative) lifetime of x.

Thus, X is complete iff each integral curve can be extended so that itsdomain becomes ] −∞,∞ [; i.e., T+(x) = ∞ and T−(x) = −∞ for allx ∈ U .

Example

A. For U = R2, let X be the constant vector field X(x, y) = (0, 1).Then X is complete since the integral curve of X through (x, y) is t 7→(x, y + t). ¨

B. On U = R2\0, the same vector field is not complete since the inte-gral curve of X through (0,−1) cannot be extended beyond t = 1; in factas t→ 1 this integral curve tends to the point (0, 0). Thus T+(0,−1) = 1,while T−(0,−1) = −∞. ¨

C. On R consider the vector field X(x) = 1 + x2. This is not completesince the integral curve c with c(0) = 0 is c(θ) = tan θ and thus it cannotbe continuously extended beyond −π/2 and π/2; i.e., T±(0) = ±π/2. ¨

Proposition 2.1.7. Let U ⊂ Rn be open and X ∈ Xr(M), r ≥ 1. Then

i DX ⊃ U × 0;

ii DX is open in U × R;

iii there is a unique Cr mapping FX : DX → U such that the mappingt 7→ FX(x, t) is an integral curve at x for all x ∈ U ;

iv for (x, t) ∈ DX , (FX(x, t), s) ∈ DX iff (m, t+ s) ∈ DX ; in this case

FX(x, t+ s) = FX(FX(x, t), s).

Definition 2.1.8. Let U ⊂ Rn be open and X ∈ Xr(U), r ≥ 1. Then themapping FX is called the integral of X, and the curve t 7→ FX(x, t) iscalled the maximal integral curve of X at x. In case X is complete, FXis called the flow of X.

36 2. Mathematical preliminaries

Thus, if X is complete with flow F , then the set Ft | t ∈ R is a groupof diffeomorphisms on U , sometimes called a one-parameter group ofdiffeomorphisms. Since Fn = (F1)n (the n-th power), the notation F t issometimes convenient and is used where we use Ft. For incomplete flows, ivsays that Ft Fs = Ft+s wherever it is defined. Note that Ft(x) is definedfor t ∈ ]T−(x), T+(x)[. The reader should write out similar definitions forthe time-dependent case and note that the lifetimes depend on the startingtime t0.

A useful criterion for completeness is the following:

Proposition 2.1.9. Let X be a Cr vector field on an open subset U ofRn, where r ≥ 1. Let c(t) be a maximal integral curve of X such thatfor every finite open interval ]a, b[ in the domain ]T−(c(0)), T+(c(0))[ ofc, c(]a, b[) lies in a compact subset of U . Then c is defined for all t ∈ R. IfU = Rn, this holds provided c(t) lies in a bounded set.

For example, this is used to prove:

Corollary 2.1.10. A Cr vector field on an open set U with compact sup-port contained in U is complete.

Completeness corresponds to well-defined dynamics persisting eternally.In some circumstances (shock waves in fluids and solids, singularities ingeneral relativity, etc.) one has to live with incompleteness or overcome itin some other way.

Example A. Let X be a Cr vector field, r ≥ 1, on the open set U ⊂ Rnadmitting a first integral , i.e., a function f : U → R such that X[f ] = 0.If all level sets f−1(r), r ∈ R are compact, X is complete. Indeed, eachintegral curve lies on a level set of f so that the result follows by thepreceding Proposition. ¨

Example B. Suppose

X(x) = A · x+B(x),

where A is a linear operator of Rn to itself and B is sublinear ; i.e., B :Rn → Rn is Cr with r ≥ 1 and satisfies ‖B(x)‖ ≤ K‖x‖+ L for constantsK and L. We shall show that X is complete. Let x(t) be an integral curveof X on the bounded interval [0, T ]. Then

x(t) = x(0) +∫ t

0

(A · x(s) +B(x(s))) ds

Hence

‖x(t)‖ ≤ ‖x(0)‖∫ t

0

(‖A‖+K)‖x(s)‖ ds+ Lt.

2.1 Vector fields, flows and differential equations 37

By Gronwall’s inequality,

‖x(t)‖ ≤ (LT + ‖x(0)‖)e(‖A‖+K)t.

Hence x(t) remains bounded on bounded t-intervals, so the result followsby Proposition 2.1.9. ¨

Example C. Newton’s equations for a moving particle of mass m in apotential field in Rn are given by q(t) = −(1/m)∇V (q(t)), for V : Rn → Ra smooth function. We shall prove that if there are constants a, b ∈ R, b ≥ 0such that (1/m)V (q) ≥ a − b‖q‖2, then every solution exists for all time.To show this, rewrite the second order equations as a first order systemq = (1/m)p, p = −∇V (q) and note that the energy

E(q,p) = (1/2m)‖p‖2 + V (q)

is a first integral. Thus, for any solution (q(t),p(t)) we have β = E(q(t),p(t)) =E(q(0),p(0)) ≥ V (q(0)). We can assume β > V (q(0)), i.e., p(0) 6= 0, forif p(t) ≡ 0, then the conclusion is trivially satisifed; thus there exists a t0for which p(t0) 6= 0 and by time translation we can assume that t0 = 0.Thus we have

‖q(t)‖ ≤ ‖q(t)− q(0)‖+ ‖q(0)‖ ≤ ‖q(0)‖+∫ t

0

‖q(s)‖ ds

= ‖q(0)‖+∫ t

0

√2[β − 1

mV (q(s))

]ds

≤ ‖q(0)‖+∫ t

0

√2(β − a+ b‖q(s)‖2) ds

or in differential form

d

dt‖q(t)‖ ≤

√2(β − a+ b‖q(t)‖2)

whence

t ≤∫ ‖q(t)‖

‖q(0)‖

du√2(β − a+ bu2)

(2.1.2)

Now let r(t) be the solution of the differential equation

d2r(t)dt2

= − d

dr(a− br2)(t) = 2br(t),

which, as a second order equation with constant coefficients, has solutionsfor all time for any initial conditions. Choose

r(0) = ‖q(0)‖, [r(0)]2 = 2(β − a+ b‖q(0)‖2)

38 2. Mathematical preliminaries

and let r(t) be the corresponding solution. Since

d

dt

(12r(t)2 + a− br(t)2

)= 0,

it follows that (1/2)r(t)2 + a− br(t)2 = (1/2)r(0)2 + a− br(0)2 = β, i.e.,

dr(t)dt

=√

2(β − a+ br(t)2)

whence

t =∫ r(t)

‖q(0)‖

du√s(β − α+ βu2

(2.1.3)

Comparing these two expressions and taking into account that the inte-grand is > 0, it follows that for any finite time interval for which q(t)is defined, we have ‖q(t)‖ ≤ r(t), i.e., q(t) remains in a compact set forfinite t-intervals. But then q(t) also lies in a compact set since ‖q(t)‖ ≤2(β−a+ b‖q(s)‖2). Thus by 2.1.9, the solution curve (q(t),p(t)) is definedfor any t ≥ 0. However, since (q(−t),p(−t)) is the value at t of the integralcurve with initial conditions (−q(0),−p(0)), it follows that the solutionalso exists for all t ≤ 0.

The following counterexample shows that the condition V (q) ≥ a−b‖q‖2cannot be relaxed much further. Take n = 1 and V (q) = −ε2q2+(4/ε)/8, ε >0. Then the equation

q = ε(ε+ 2)q1+(4/ε)/4

has the solution q(t) = 1/(t − 1)ε/2, which cannot be extended beyondt = 1. ¨

The following is proved by a study of the local existence theory; we stateit for completeness only.

Proposition 2.1.11. Let X be a Cr vector field on U, r ≥ 1, x0 ∈ U ,and T+(x0)(T−(x0)) the positive (negative) lifetime of x0. Then for eachε > 0, there exists a neighborhood V of x0 such that for all x ∈ V, T+(x) >T+(x0)− ε (respectively, T−(x0) < T−(x0) + ε). [One says that T+(x0) isa lower semi-continuous function of x.]

Corollary 2.1.12. Let Xt be a Cr time-dependent vector field on U, r ≥1, and let x0 be an equilibrium of Xt, i.e., Xt(x0) = 0, for all t. Then forany T there exists a neighborhood V of x0 such that any x ∈ V has integralcurve existing for time t ∈ [−T, T ].

2.2 Differentiable Manifolds 39

Linear equations. Finally we note that the flows of linear equations

x = Ax

where A is an n× n matrix are given by

Ft(x) = etA

where the exponential is defined, for example, by a power series

etA = I + tA+12t2A2 +

13!t3A3 + . . . .

One of course has to show that this series converges and is differentiable int and that the derivative is given by AetA, but this is learned in courses onreal analysis. One also learns how to carry out the exponentiation in linearalgebra by bringing A into a canonical form. We shall discuss how to dosome of these cases in the next sections.

What about stability theory? Should we put anyting in at this pointor leave this to later?

2.2 Differentiable Manifolds

Modern analytical mechanics and nonlinear control theory are most natu-rally discussed in the mathematical language of differential geometry. Thepresent chapter is meant to serve as an introduction to the elements ofgeometry which we shall use in the remainder of the book. Since this is notprimarily a text on geometry, there is a great deal that must be left out.4

Studying the motion of physical systems leads immediately to the studyof rates of change of position and velocity—i.e. to calculus. Differentiablemanifolds provide the most natural setting in which to study calculus.Roughly speaking, differentiable manifolds are topological spaces whichlocally look like Euclidean space, but which may be globally quite differ-ent from Euclidean space. Since taking a derivative involves only a localcomputation—carried out in a neighborhood of the point of interest, itwould appear that derivatives should be computable on any topological

4Some references are Abraham, R., J.E. Marsden, and T.S. Ratiu [1988] Manifolds,Tensor Analysis, and Applications. Second Edition, Applied Mathematical Sciences 75,Springer-Verlag, Warner, F.W. [1983] Foundations of Differentiable Manifolds and LieGroups. Graduate Texts in Math. 94 Springer–Verlag, Auslander, L. and R.E. MacKen-zie [1977]. Introduction to Differentiable Manifolds, Dover Publications, New York,Boothby, W.M. [1992] An Introduction to Differentiable Manifolds and Riemannian Ge-ometry, publisher? and Dubrovin, B.A., A.T. Fomenko, and S.P. Novikov [1984] ModernGeometry – Methods and Applications. Part II. The Geometry and Topology of Mani-folds. Graduate Texts in Math. 104, Springer–Verlag.

40 2. Mathematical preliminaries

space that is infinitesimally indistinguishable from Euclidean space. Thisis indeed the case. What makes differentiable manifolds most important inthe study of analytical mechanics, however, is the global features and theirimplications for the large scale behavior of trajectories of physical systems.

With these remarks in mind, we begin with a definition of manifold whichrelates these objects to Euclidean space in small neighborhoods of eachpoint. Questions about important global features of differentiable manifoldswill be discussed in subsequent sections.

Definition 2.2.1. An n-dimensional differentiable manifold M is aset of points together with a finite or countably infinite set of subsets Uα ⊂M and 1-1-mappings φα : Uα → Rn such that

1.⋃α Uα = M .

2. For each non-empty intersection Uα ∩ Uβ, φi(Uα ∩ Uβ) is an opensubset of Rn, and the 1-1-mapping φα φ−1

β : φβ(Uα∩Uβ)→ φα(Uα∩Uβ) is a smooth function.

3. The family Uα, φα is maximal with respect to conditions 1 and 2.

The situation is illustrated in Figure 2.2.1.

ϕβ

ϕα

M

y

x

y

x

Figure 2.2.1. Coordinate charts on a manifold.

A number of remarks are in order. First, the sets Uα in the definitionare called coordinate neighborhoods or coordinate charts. The map-pings φα are called coordinate functions or local coordinates. A col-lection of charts satisfying 1 and 2 is called an atlas. The notion of aCk-differentiable manifold (resp. analytic) is defined similarly wherein the

2.2 Differentiable Manifolds 41

coordinate transformations φα φ−1β are only required to have conti-

nous partial derivatives of all orders up to k (resp. be analytic). We remarkthat condition is included merely to make the definition of manifold inde-pendent of a choice of atlas. A set of charts satisfying 1 and 2 can alwaysbe extended to a maximal set and in practise 1 and 2 define the manifold.

A neighborhood of a point x in a manifold M is the image under amap ϕ : U → M of a neighborhood of the representation of x in a chartU . Neighborhoods define open sets and one checks that the open sets inM define a topology. Usually we assume without explicit mention that thetopology is Hausdorff : two different points x, x′ in M have nonintersectingneighborhoods.

A useful viewpoint is to think of M as a set covered by a collectionof coordinate charts with local coordinates (x1, . . . , xn) with the propertythat all mutual changes of coordinates are smooth maps.

Differentiable manifolds as level sets in Rn. A typical and importantway that manifolds arise is as follows. Let p1, p2, . . . , pm : Rn → R. Thezero (or level) set (also called a locus)

M = x | pi(x) = 0, i = 1, . . . ,m

is called a differentiable variety in Rn. If the n×m matrix(∂p1

∂x

... · · ·...∂pm∂x

)has rank ρ at each point x ∈M , then M admits the structure of a differen-tiable manifold of dimension n− ρ. The idea of the proof is that under ourrank assumption an argument using the implicit function theorem showsthat an n− ρ-dimensional coordinate chart may be defined in a neighbor-hood of each point on M . In this situation, we say that the level set is asubmanifold of Rn.

Exercises

¦ 2.2-1. Using this criterion, show that the level set x21 + · · ·+ x2

n − 1 = 0is a differentiable manifold of dimension n− 1.

¦ 2.2-2. Show that the variety (x, y) | x2(x+ 1)− y2 = 0 in R2 is not adifferentiable manifold.

Matrix groups. Let Rn×n be the set of n×n matrices with entries in R,and let GL(n,R) denote the set of all n×n invertible matrices with entriesin R. Clearly, Gl(n,R) is a group, called the general linear group. LetA ∈ Gl(n,R) be symmetric. Consider the subset

S = X ∈ Gl(n,R) | X ·A ·XT = A.

42 2. Mathematical preliminaries

It is easy to see that if X ∈ S, then X−1 ∈ S, and if X,Y ∈ S then theproduct X · Y ∈ S. Hence, S is a subgroup of Gl(n,R).

We can also show that S is a submanifold of Rn×n. Indeed, S is the zerolocus of the mapping X 7→ X · A ·XT − A. Let X ∈ S, and let δX be anarbitrary element of Rn×n. Then

(X + δX) ·A · (X + δX)T −A =

X ·A ·XT −A+ δX ·A ·X +X ·AδXT + o(δX)2.

We can conclude that S is a submanifold of Rn×n if we can show that thelinearization of the locus map, namely the linear mapping L defined byδX 7→ δX · A · X + X · AδXT of Rn×n to itself has constant rank for allX ∈ S. We see that the image of L lies in the subspace of n×n symmetricmatrices. Hence, rank L ≤ (n(n− 1))/2. But given X and any symmetricmatrix S we can find δX such that δX · A · X + X · AδXT = S. Thisshows that rank L = (n(n − 1))/2, independent of X. In summary, wehave established the following fact.

Proposition 2.2.2. Let A ∈ Gl(n,R) be symmetric. Then the subgroupS of Gl(n,R) given by

S = X ∈ Gl(n,R) | X ·A ·XT = A

is a submanifold of Rn×n of dimension n(n− 1)/2.

The orthogonal group. Of special interest in mechanics is the caseA = I. Here S specializes to O(n), the group of n×n orthogonal matrices.It is both a subgroup of Gl(n,R) and a submanifold of the vector spaceRn×n. Gl(n,R) is an open, dense subset of Rn×n which inherits the topol-ogy and manifold structure from Rn×n. Thus, O(n) (or any S defined asabove) is both a subgroup and a submanifold of Gl(n,R). Subgroups ofGl(n,R) which are also submanifolds are called matrix Lie groups. Weshall discuss Lie groups more abstractly later on.

Tangent vectors. Two curves t 7→ c1(t) and t 7→ c2(t) in an n-manifoldM are called equivalent at x ∈M if

c1(0) = c2(0) = x and (ϕ c1)′(0) = (ϕ c2)′(0)

in some chart ϕ, where the prime denotes the derivative with respect to thecurve parameter. It is easy to check that this definition is chart independent.A tangent vector v to a manifold M at a point x ∈M is an equivalenceclass of curves at x. One proves that the set of tangent vectors to M at xforms a vector space. It is denoted TxM and is called the tangent spaceto M at x ∈M . Given a curve c(t), we denote by c′(s) the tangent vectorat c(s) defined by the equivalence class of t 7→ c(s+ t) at t = 0.

2.2 Differentiable Manifolds 43

Let U be a chart of an atlas for the manifold M with coordinates(x1, . . . , xn). The components of the tangent vector v to the curve t 7→(ϕ c)(t) are the numbers v1, . . . , vn defined by

vi =d

dt(ϕ c)i

∣∣∣∣t=0

,

where i = 1, . . . , n. The tangent bundle of M , denoted by TM , is thedifferentiable manifold whose underlying set is the disjoint union of thetangent spaces to M at the points x ∈M , that is,

TM =⋃x∈M

TxM.

Thus, a point of TM is a vector v that is tangent to M at some point x ∈M . To define the differentiable structure on TM , we need to specify howto construct local coordinates on TM . To do this, let x1, . . . , xn be localcoordinates on M and let v1, . . . , vn be components of a tangent vector inthis coordinate system. Then the 2n numbers x1, . . . , xn, v1, . . . , vn give alocal coordinate system on TM . Notice that dimTM = 2 dimM .

The natural projection is the map τM : TM →M that takes a tangentvector v to the point x ∈ M at which the vector v is attached (that is,v ∈ TxM). The inverse image τ−1

M (x) of a point x ∈ M under the naturalprojection τM is the tangent space TxM . This space is called the fiber ofthe tangent bundle over the point x ∈M .

Tangent spaces to level sets. Let M = x | pi(x) = 0, i = 1, . . . ,mbe a differentiable variety in Rn. For each x ∈M ,

TxM =v ∈ Rn

∣∣∣∣ ∂pi∂x(x) · v = 0

is called the tangent space to M at x. Clearly TxM is a vector space.If the rank condition holds so that M is a differentiable manifold, thenthis definition may be shown to be equivalent to the one given earlier. Forexample, the reader may show that the tangent spaces to spheres are whatthey should intuitively be, as in Figure 2.2

Tangent spaces to matrix groups. Let A ∈ Gl(n,R) be a symmetricmatrix. We wish to explicitly describe the tangent space at a typical pointof the group S = X ∈ Gl(n,R) : XTAX = A. Given our definition,it is clear that the tangent space TXS is a subspace of the space of n × nmatrices, Rn×n. Let V ∈ Rn×n. Then V ∈ TXS precisely when it is tangentto a curve in the group:

d

dε |ε=0

[(X + εV )TA(X + εV )−A

]= 0.

This condition is equivalent to V TAX +XTAV = 0.

44 2. Mathematical preliminaries

@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@

ÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀ

@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@@

ÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀÀ

yyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyyy

x

S2

TxS2

Figure 2.2.2. The tangent space to a sphere.

This shows that if X ∈ S, then

TXS = V ∈ Rn×n | V TAX +XTAV = 0.

Note that the n × n identity matrix I ∈ S, and show that for any pairof matrices V1, V2 ∈ TIS we have V1V2 − V2V1 ∈ TIS.

Differentiable maps. Let E and F be vector spaces and let f : U ⊂E → V ⊂ F , where U and V are open sets, be of class Cr+1. We definethe tangent map of f , Tf (sometimes denoted f∗ by

Tf : TU = U × E → TV = V × F

where

Tf(u, e) = (f(u), Df(u), e).e ∈ E. (2.2.1)

This notion from calculus may be generalized to the context of manifoldsas follows. Let f : M → N be a map of a manifold M to a manifold N . Wecall f differentiable (or Ck) if in local coordinates on M and N it is givenby differentiable (or Ck) functions. The derivative of a differentiable mapf : M → N at a point x ∈M is defined to be the linear map

Txf : TxM → Tf(x)N

constructed in the following way. For v ∈ TxM , choose a curve c : ]−ε, ε[→M with c(0) = x, and velocity vector dc/dt |t=0 = v . Then Txf · v is thevelocity vector at t = 0 of the curve f c : R→ N , that is,

Txf · v =d

dtf(c(t))

∣∣∣∣t=0

.

The vector Txf · v does not depend on the curve c but only on the vectorv. If M and N are manifolds and f : M → N is of class Cr+1, thenTf : TM → TN is a mapping of class Cr. Note that

dc

dt

∣∣∣∣t=0

= T0c · 1.

2.2 Differentiable Manifolds 45

Vector fields and flows. Let us now interpret what we did with vectorfields and flows in the previous section in the context of manifolds. A vectorfield X on a manifold M is a map X : M → TM that assigns a vectorX(x) at the point x ∈M ; that is, τM X = identity. An integral curve ofX with initial condition x0 at t = 0 is a (differentiable) map c : ]a, b[→Msuch that ]a, b[ is an open interval containing 0, c(0) = x0 and

c′(t) = X(c(t))

for all t ∈ ]a, b[. In formal presentations we usually suppress the domain ofdefinition, even though this is technically important. The flow of X is thecollection of maps

ϕt : M →M

such that t 7→ ϕt(x) is the integral curve of X with initial condition x.Existence and uniqueness theorems from ordinary differential equations,as reviewed in the last section guarantee ϕ is smooth in x and t (wheredefined) if X is. From uniqueness, we get the flow property

ϕt+s = ϕt ϕsalong with the initial conditions ϕ0 = identity. The flow property gener-alizes the situation where M = V is a linear space, X(x) = Ax for a(bounded) linear operator A, and where

ϕt(x) = etAx

to the nonlinear case.

Differentials and covectors. If f : M → R is a smooth function, we candifferentiate it at any point x ∈M to obtain a map Txf : TxM → Tf(x)R.Identifying the tangent space of R at any point with itself (a process weusually do in any vector space), we get a linear map df(x) : TxM → R.That is, df(x) ∈ T ∗xM , the dual of the vector space TxM .

In coordinates, the directional derivatives defined by df(x) ·v, wherev ∈ TxM , are given by

df(x) · v =n∑i=1

∂f

∂xivi.

We will employ the summation convention and drop the summationsign when there are repeated indices. We also call df the differential off .

One can show that specifying the directional derivatives completely de-termines a vector, and so we can identify a basis of TxM using the operators∂/∂xi. We write

(e1, . . . , en) =(

∂x1, . . . ,

∂xn

)

46 2. Mathematical preliminaries

for this basis so that v = vi∂/∂xi.If we replace each vector space TxM with its dual T ∗xM , we obtain a new

2n-manifold called the cotangent bundle and denoted T ∗M . The dualbasis to ∂/∂xi is denoted dxi. Thus, relative to a choice of local coordinateswe get the basic formula

df(x) =∂f

∂xidxi

for any smooth function f : M → R.

Exercises

¦ 2.2-3. If ϕt : S2 → S2 rotates points on S2 about a fixed axis throughan angle t, show that ϕt is the flow of a certain vector field on S2.

¦ 2.2-4. Let f : S2 → R be defined by f(x, y, z) = z. Compute df relativeto spherical coordinates (θ, ϕ).

2.3 Differential Forms

We next review some of the basic definitions, properties, and operations ondifferential forms, without proofs (see Abraham, Marsden, and Ratiu [1988]and references therein). The main idea of differential forms is to providea generalization of the basic operations of vector calculus, div, grad, andcurl, and the integral theorems of Green, Gauss, and Stokes to manifoldsof arbitrary dimension.

Basic definitions. A 2-form Ω on a manifold M is a function Ω(x) :TxM × TxM → R that assigns to each point x ∈ M a skew-symmetricbilinear form on the tangent space TxM to M at x. More generally, a k-form α (sometimes called a differential form of degree k) on a manifoldM is a function α(x) : TxM × . . . × TxM (there are k factors) → R thatassigns to each point x ∈ M a skew-symmetric k-multilinear map on thetangent space TxM to M at x. Without the skew-symmetry assumption,α would be called a (0, k)-tensor . A map α : V × . . . × V (there are kfactors)→ R is multilinear when it is linear in each of its factors, that is,

α(v1, . . . , avj + bv′j , . . . , vk) = aα(v1, . . . , vj , . . . , vk) + bα(v1, . . . , v′j , . . . , vk)

for all j with 1 ≤ j ≤ k. A k-multilinear map α : V × . . .× V → R is skew(or alternating) when it changes sign whenever two of its arguments areinterchanged, that is, for all v1, . . . , vk ∈ V ,

α(v1, . . . , vi, . . . , vj , . . . , vk) = −α(v1, . . . , vj , . . . , vi, . . . , vk).

2.3 Differential Forms 47

Let x1, . . . , xn denote coordinates on M , let

e1, . . . , en = ∂/∂x1, . . . , ∂/∂xn

be the corresponding basis for TxM , and let e1, . . . , en = dx1, . . . , dxnbe the dual basis for T ∗xM . Then at each x ∈M , we can write a 2-form as

Ωx(v, w) = Ωij(x)viwj , where Ωij(x) = Ωx

(∂

∂xi,∂

∂xj

),

and more generally a k-form can be written

αx(v1, . . . , vk) = αi1...ik(x)vi11 . . . vikk ,

where there is a sum on i1, . . . , ik and where

αi1...ik(x) = αx

(∂

∂xi1, . . . ,

∂xik

),

and where vi = vji ∂/∂xj , with a sum on j.

Tensor and wedge products. If α is a (0, k)-tensor on a manifold M ,and β is a (0, l)-tensor, their tensor product α⊗ β is the (0, k+ l)-tensoron M defined by

(α⊗ β)x(v1, . . . , vk+l) = αx(v1, . . . , vk)βx(vk+1, . . . , vk+l) (2.3.1)

at each point x ∈M .If t is a (0, p)-tensor, define the alternation operator A acting on t by

A(t)(v1, . . . , vp) =1p!

∑π∈Sp

sgn(π)t(vπ(1), . . . , vπ(p)), (2.3.2)

where sgn(π) is the sign of the permutation π:

sgn(π) =

+1 if π is even,−1 if π is odd, (2.3.3)

and Sp is the group of all permutations of the numbers 1, 2, . . . , p. Theoperator A therefore skew-symmetrizes p-multilinear maps.

If α is a k-form and β is an l-form on M , their wedge product α∧ β isthe (k + l)-form on M defined by5

α ∧ β =(k + l)!k! l!

A(α⊗ β). (2.3.4)

5The numerical factor in (2.3.4) agrees with the convention of Abraham and Marsden[1978], Abraham, Marsden, and Ratiu [1988], and Spivak [1976], but not that of Arnold[1989], Guillemin and Pollack [1974], or Kobayashi and Nomizu [1963]; it is the Bourbaki[1971] convention.

48 2. Mathematical preliminaries

For example, if α and β are one-forms,

(α ∧ β)(v1, v2) = α(v1)β(v2)− α(v2)β(v1)

while if α is a 2-form and β is a 1-form,

(α ∧ β)(v1, v2, v3) = α(v1, v2)β(v3) + α(v3, v1)β(v2) + α(v2, v3)β(v1).

We state the following without proof:

Proposition 2.3.1. The wedge product has the following properties:

i α ∧ β is associative : α ∧ (β ∧ γ) = (α ∧ β) ∧ γ.

ii α ∧ β is bilinear in α, β :

(aα1 + bα2) ∧ β = a(α1 ∧ β) + b(α2 ∧ β),α ∧ (cβ1 + dβ2) = c(α ∧ β1) + d(α ∧ β2).

iii α∧β is anticommutative : α∧β = (−1)klβ∧α, where α is a k-formand β is an l-form.

In terms of the dual basis dxi, any k-form can be written locally as

α = αi1...ikdxi1 ∧ · · · ∧ dxik

where the sum is over all ij satisfying i1 < · · · < ik.

Pull back and push forward. Let ϕ : M → N be a C∞ map from themanifold M to the manifold N and α be a k-form on N . Define the pullback ϕ∗α of α by ϕ to be the k-form on M given by

(ϕ∗α)x(v1, . . . , vk) = αϕ(x)(Txϕ · v1, . . . , Txϕ · vk). (2.3.5)

If ϕ is a diffeomorphism, the push forward ϕ∗ is defined by ϕ∗ =(ϕ−1)∗.

Here is another basic property.

Proposition 2.3.2. The pull back of a wedge product is the wedge productof the pull backs:

ϕ∗(α ∧ β) = ϕ∗α ∧ ϕ∗β. (2.3.6)

Interior products and exterior derivatives. Let α be a k-form on amanifold M and X a vector field. The interior product iXα (sometimescalled the contraction of X and α, and written i(X)α) is defined by

(iXα)x(v2, . . . , vk) = αx(X(x), v2, . . . , vk). (2.3.7)

2.3 Differential Forms 49

Proposition 2.3.3. Let α be a k-form and β an l-form on a manifoldM . Then

iX(α ∧ β) = (iXα) ∧ β + (−1)kα ∧ (iXβ). (2.3.8)

The exterior derivative dα of a k-form α on a manifold M is the(k + 1)-form on M determined by the following proposition:

Proposition 2.3.4. There is a unique mapping d from k-forms on M to(k + 1)-forms on M such that:

i If α is a 0- form (k = 0), that is, α = f ∈ C∞(M), then df is theone-form which is the differential of f .

ii dα is linear in α, that is, for all real numbers c1 and c2,

d(c1α1 + c2α2) = c1dα1 + c2dα2.

iii dα satisfies the product rule, that is,

d(α ∧ β) = dα ∧ β + (−1)kα ∧ dβ,

where α is a k-form and, β is an l-form.

iv d2 = 0, that is, d(dα) = 0 for any k-form α.

v d is a local operator , that is, dα(x) only depends on α restrictedto any open neighborhood of x; in fact, if U is open in M , then

d(α|U) = (dα)|U.

If α is a k-form given in coordinates by

α = αi1...ikdxi1 ∧ · · · ∧ dxik (sum on i1 < · · · < ik),

then the coordinate expression for the exterior derivative is

dα =∂αi1...ik∂xj

dxj ∧ dxi1 ∧ · · · ∧ dxik (sum on all j and i1 < · · · < ik).

(2.3.9)

Formula (2.3.9) can be taken as the definition of the exterior derivative,provided one shows that (2.3.9) has the above-described properties and,correspondingly, is independent of the choice of coordinates.

Next is a useful proposition that, in essence, rests on the chain rule:

Proposition 2.3.5. Exterior differentiation commutes with pull back, thatis,

d(ϕ∗α) = ϕ∗(dα), (2.3.10)

where α is a k-form on a manifold N and ϕ is a smooth map from amanifold M to N .

50 2. Mathematical preliminaries

A k-form α is called closed if dα = 0 and exact if there is a (k−1)-formβ such that α = dβ. By Proposition 2.3.4 every exact form is closed. Theexercises give an example of a closed nonexact one-form.

Proposition 2.3.6 (Poincare Lemma). A closed form is locally exact,that is, if dα = 0 there is a neighborhood about each point on which α = dβ.

The proof is given in the exercises.

Vector calculus. The table below entitled “Vector calculus and differ-ential forms” summarizes how forms are related to the usual operations ofvector calculus. We now elaborate on a few items in this table. In item 4,note that

df =∂f

∂xdx+

∂f

∂ydy +

∂f

∂zdz = (gradf)[ = (∇f)[

which is equivalent to ∇f = (df)].The Hodge star operator on R3 maps k-forms to (3 − k)-forms and is

uniquely determined by linearity and the properties in item 2.6

In item 5, if we let F = F1e1+F2e2+F3e3, so F [ = F1 dx+F2 dy+F3 dz,then,

d(F [) = dF1 ∧ dx+ F1d(dx) + dF2 ∧ dy + F2d(dy) + dF3 ∧ dz+ F3d(dz)

=(∂F1

∂xdx+

∂F1

∂ydy +

∂F1

∂zdz

)∧ dx

+(∂F2

∂xdx+

∂F2

∂ydy +

∂F2

∂zdz

)∧ dy

+(∂F3

∂xdx+

∂F3

∂ydy +

∂F3

∂zdz

)∧ dz

= −∂F1

∂ydx ∧ dy +

∂F1

∂zdz ∧ dx+

∂F2

∂xdx ∧ dy − ∂F2

∂zdy ∧ dz

− ∂F3

∂xdz ∧ dx+

∂F3

∂ydy ∧ dz

=(∂F2

∂x− ∂F1

∂y

)dx ∧ dy +

(∂F1

∂z− ∂F3

∂x

)dz ∧ dx

+(∂F3

∂y− ∂F2

∂z

)dy ∧ dz.

6This operator can be defined on general Riemannian manifolds; see Abraham, Mars-den, and Ratiu [1988].

2.3 Differential Forms 51

Hence, using item 2,

∗(d(F [)) =(∂F2

∂x− ∂F1

∂y

)dz +

(∂F1

∂z− ∂F3

∂x

)dy

+(∂F3

∂y− ∂F2

∂z

)dx,

(∗(d(F [)))] =(∂F3

∂y− ∂F2

∂z

)e1 +

(∂F1

∂z− ∂F3

∂x

)e2

+(∂F2

∂x− ∂F1

∂y

)e3

= curlF = ∇× F.

With reference to item 6, let F = F1e1 + F2e2 + F3e3, so F [ = F1 dx+F2 dy+F3 dz. Thus ∗(F [) = F1 dy∧dz+F2(−dx∧dz) +F3 dx∧dy, and so

d(∗(F [)) = dF1 ∧ dy ∧ dz − dF2 ∧ dx ∧ dz + dF3 ∧ dx ∧ dy

=(∂F1

∂xdx+

∂F1

∂ydy +

∂F1

∂zdz

)∧ dy ∧ dz

−(∂F2

∂xdx+

∂F2

∂ydy +

∂F2

∂zdz

)∧ dx ∧ dz

+(∂F3

∂xdx+

∂F3

∂ydy +

∂F3

∂zdz

)∧ dx ∧ dy

=∂F1

∂xdx ∧ dy ∧ dz +

∂F2

∂ydx ∧ dy ∧ dz +

∂F3

∂zdx ∧ dy ∧ dz

=(∂F1

∂x+∂F2

∂y+∂F3

∂z

)dx ∧ dy ∧ dz = (div F ) dx ∧ dy ∧ dz.

Therefore ∗(d(∗(F [))) = div F = ∇ · F .The definition and properties of vector-valued forms are direct extensions

of these for usual forms on vector spaces and manifolds. One can think ofa vector-valued form as an array of usual forms.

Vector Calculus and Differential Forms

1. Sharp and Flat (Using standard coordinates in R3)

(a) v[ = v1 dx+v2 dy+v3 dz = one-form corresponding to the vectorv = v1e1 + v2e2 + v3e3.

(b) α] = α1e1 +α2e2 +α3e3 = vector corresponding to the one-formα = α1 dx+ α2 dy + α3 dz.

2. Hodge Star Operator

(a) ∗1 = dx ∧ dy ∧ dz.

52 2. Mathematical preliminaries

(b) ∗dx = dy ∧ dz, ∗dy = −dx ∧ dz, ∗dz = dx ∧ dy,∗(dy ∧ dz) = dx, ∗(dx ∧ dz) = −dy, ∗(dx ∧ dy) = dz.

(c) ∗(dx ∧ dy ∧ dz) = 1.

3. Cross Product and Dot Product

(a) v × w = [∗(v[ ∧ w[)]].(b) (v · w)dx ∧ dy ∧ dz = v[ ∧ ∗(w[).

4. Gradient ∇f = gradf = (df)].

5. Curl ∇× F = curlF = [∗(dF [)]].

6. Divergence ∇ · F = div F = ∗d(∗F [).

Exercises

¦ 2.3-1. Let ϕ : R3 → R2 be given by ϕ(x, y, z) = (x + z, xy). For α =ev du + u dv ∈ Ω1(R2) and β = u du ∧ dv compute α ∧ β, ϕ∗α, ϕ∗β, andϕ∗α ∧ ϕ∗β.

¦ 2.3-2. Given

α = y2 dx ∧ dz + sin(xy) dx ∧ dy + ex dy ∧ dz ∈ Ω2(R3)

and

X = 3∂/∂x+ cos z∂/∂y − x2∂/∂z ∈ X(R3),

compute dα and iXα.

¦ 2.3-3. (a) Denote by Λk(Rn) the vector space of all skew-symmetric k-linear maps on Rn. Prove that this space has dimension n!/k! (n−k)!by showing that a basis is given by ei1 ∧ · · · ∧ eik | i1 < . . . < ikwhere e1, . . . , en is a basis of Rn and e1, . . . , en is its dual basis,that is, ei(ej) = δij .

(b) If µ ∈ Λn(Rn) is nonzero, prove that the map v ∈ Rn 7→ ivµ ∈Λn−1(Rn) is an isomorphism.

(c) If M is a smooth n-manifold and µ ∈ Ωn(M) is nowhere vanishing(in which case it is called a volume form), show that the map X ∈X(M) 7→ iXµ ∈ Ωn−1(M) is a module isomorphism over F(M).

¦ 2.3-4. Let α = αi dxi be a closed one-form in a ball around the origin in

Rn. Show that α = df for

f(x1, . . . , xn) =∫ 1

0

αj(tx1, . . . , txn)xj dt.

2.4 Lie derivatives 53

¦ 2.3-5. (a) Let U be an open ball around the origin in Rn and α ∈ Ωk(U)a closed form. Verify that α = dβ where

β(x1, . . . , xn)

=(∫ 1

0

tk−1αji1...ik−1(tx1, . . . , txn)xj dt)dxi1 ∧ . . . ∧ dxik−1 ,

and where the sum is over i1 < · · · < ik−1. Here, α = αj1...jk dxj1 ∧

. . . ∧ dxjk , where j1 < · · · < jk and where α is extended to be skew-symmetric in its lower indices.

(b) Deduce the Poincare lemma from (a).

2.4 Lie derivatives

The dynamic definition of the Lie derivative is as follows. Let α be a k-formand let X be a vector field with flow ϕt. The Lie derivative of α alongX is given by

£Xα = limt→0

1t[(ϕ∗tα)− α] =

d

dtϕ∗tα

∣∣∣∣t=0

. (2.4.1)

This definition together with properties of pull-backs yields the following.

Theorem 2.4.1 (Lie Derivative Theorem).

d

dtϕ∗tα = ϕ∗t£Xα. (2.4.2)

This formula holds also for time-dependent vector fields.If f is a real-valued function on a manifold M and X is a vector field on

M , the Lie derivative of f along X is the directional derivative

£Xf = X[f ] := df ·X. (2.4.3)

In coordinates on M ,

£Xf = Xi ∂f

∂xi. (2.4.4)

If Y is a vector field on a manifold N and ϕ : M → N is a diffeomorphism,the pull back ϕ∗Y is a vector field on M defined by

(ϕ∗Y )(x) = Txϕ−1 Y ϕ(x). (2.4.5)

Two vector fields X on M and Y on N are said to be ϕ-related if

Tϕ X = Y ϕ. (2.4.6)

Clearly, if ϕ : M → N is a diffeomorphism and Y is a vector field on N ,ϕ∗Y and Y are ϕ-related. For a diffeomorphism ϕ, the push forward isdefined, as for forms, by ϕ∗ = (ϕ−1)∗.

54 2. Mathematical preliminaries

Jacobi–Lie brackets. If M is finite dimensional and C∞ then the setof vector fields on M coincides with the set of derivations on F(M). Thesame result is true for Ck manifolds and vector fields if k ≥ 2. 7 If Mis C∞ and smooth, then the derivation f 7→ X[Y [f ]] − Y [X[f ]], whereX[f ] = df · X, determines a unique vector field denoted by [X,Y ] andcalled the Jacobi-Lie bracket of X and Y . Defining £XY = [X,Y ] givesthe Lie derivative of Y along X. Then the Lie derivative theorem holdswith α replaced by Y .

In coordinates,

(£XY )j = Xi ∂Yj

∂xi− Y i ∂X

j

∂xi= (X · ∇)Y i − (Y · ∇)Xj , (2.4.7)

and in general, where we identify X,Y with their local representatives

[X,Y ] = DY ·X −DX · Y. (2.4.8)

The Lie bracket has the following geometric meaning: see e.g. Brockett[1978]. Let the differential equations describing the vector fields X and Ybe given by

x = f (2.4.9)x = g , (2.4.10)

where the components of f and g are Xi and Yi respectively. The localintegral curves of f through x0 may be then be written as (exptf)x0 andsimilarly for g. Suppose now we follow the integral curve ofX from the pointx0 for t units of time then Y for t units, then −X for t units and finally −Yfor t units. We arrive then at the point (exp−tg)(exp−tf)(exptg)(exptf)x0,Using the local second order expansion

x(t) = x0 + tf(x0) + t2/2∂f

∂xf(x0) +O(t3) (2.4.11)

shows that x(t) , after following the sequence of paths described above, isgiven by

x(t) = x0 + (t2/2)[g, f ]x0 +O(t3) ,

where

[g, f ]i = [Y,X]i .

The reader may verify this expression as an exercise.

7This property is false for infinite-dimensional manifolds; see Abraham, Marsden,Ratiu [1988].

2.4 Lie derivatives 55

Thus, if the Lie bracket of X and Y is not a linear combination of X andY , it represents a “new” direction in which one can move on the manifold,by following a path such as the one describe above.

In general if a set of vector fields Xi is such that there exist constantsγijk such that

[Xi, Xj ] = γijkXk(x)

the set is said to be involutive.

Algebraic definiton of the Lie derivative. The algebraic approach tothe Lie derivative on forms or tensors proceeds as follows. Extend the def-inition of the Lie derivative from functions and vector fields to differentialforms, by requiring that the Lie derivative is a derivation; for example, forone-forms α, write

£X〈α, Y 〉 = 〈£Xα, Y 〉+ 〈α,£XY 〉 , (2.4.12)

where X,Y are vector fields and 〈α, Y 〉 = α(Y ). More generally,

£X(α(Y1, . . . , Yk)) = (£Xα)(Y1, . . . , Yk) +k∑i=1

α(Y1, . . . ,£XYi, . . . , Yk),

(2.4.13)

where X,Y1, . . . , Yk are vector fields and α is a k-form.

Proposition 2.4.2. The dynamic and algebraic definitions of the Liederivative of a differential k-form are equivalent.

Cartan’s Magic Formula. A very important formula for the Lie deriva-tive is given by the following.

Theorem 2.4.3. For X a vector field and α a k-form on a manifold M ,we have

£Xα = diXα+ iXdα. (2.4.14)

This is proved by a lengthy but straightforward calculation.Another property of the Lie derivative is the following: if ϕ : M → N is

a diffeomorphism,

ϕ∗£Y β = £ϕ∗Y ϕ∗β

for Y ∈ X(N), β ∈ Ωk(M). More generally, if X ∈ X(M) and Y ∈ X(N)are ψ related, that is, Tψ X = Y ψ for ψ : M → N a smooth map, then£Xψ

∗β = ψ∗£Y β for all β ∈ Ωk(N).

56 2. Mathematical preliminaries

Volumes forms and divergence. An n-manifold M is said to be ori-entable if there is a nowhere vanishing n-form µ on it; µ is called a volumeform and it is a basis of Ωn(M) over F(M). Two volume forms µ1 and µ2

on M are said to define the same orientation if there is an f ∈ F(M), withf > 0 and such that µ2 = fµ1. Connected orientable manifolds admit pre-cisely two orientations. A basis v1, . . . vn of TmM is said to be positivelyoriented relative to the volume form µ on M if µ(m)(v1, . . . , vn) > 0. Notethat the volume forms defining the same orientation form a convex cone inΩn(M), that is, if a > 0 and µ is a volume form, then aµ is again a volumeform and if t ∈ [0, 1] and µ1, µ2 are volume forms, then tµ1 + (1 − t)µ2 isagain a volume form. The first property is obvious. To prove the second, letm ∈M and let σ1, . . . σn be a positively oriented basis of TmM relativeto the orientation defined by µ1, or equivalently (by hypothesis) by µ2.Then µ1(m)(v1, . . . , vn) > 0, µ2(m)(v1, . . . , vn) > 0 so that their convexcombination is again strictly positive.

If µ ∈ Ωn(M) is a volume form, since £Xµ ∈ Ωn(M) there is a function,called the divergence of X relative to µ and denoted divµ(X) or simplydiv(X), such that

£Xµ = divµ(X)µ. (2.4.15)

From the dynamic approach to Lie derivatives it follows that divµ(X) = 0iff F ∗t µ = µ, where Ft is the flow of X. This condition says that Ft isvolume preserving . If ϕ : M → M , since ϕ∗µ ∈ Ωn(M) there is afunction, called the Jacobian of ϕ and denoted Jµ(ϕ) or simply J(ϕ),such that

ϕ∗µ = Jµ(ϕ)µ. (2.4.16)

Thus, ϕ is volume preserving iff Jµ(ϕ) = 1. The inverse function theoremshows that ϕ is a local diffeomorphism iff Jµ(ϕ) 6= 0 on M .

There are a number of valuable identities relating the Lie derivative, theexterior derivative and the interior product. For example, if Θ is a one formand X and Y are vector fields, identity 6 in the following table gives

dΘ(X,Y ) = X[Θ(Y )]− Y [Θ(X)]−Θ([X,Y ]). (2.4.17)

Identities for Vector Fields and Forms

1. Vector fields on M with the bracket [X,Y ] form a Lie algebra ; thatis, [X,Y ] is real bilinear, skew-symmetric, and Jacobi’s identityholds:

[[X,Y ], Z] + [[Z,X], Y ] + [[Y, Z], X] = 0.

2.4 Lie derivatives 57

Locally,

[X,Y ] = DY ·X −DX · Y = (X · ∇)Y − (Y · ∇)X

and on functions,

[X,Y ][f ] = X[Y [f ]]− Y [X[f ]].

2. For diffeomorphisms ϕ and ψ,

ϕ∗[X,Y ] = [ϕ∗X,ϕ∗Y ] and (ϕ ψ)∗X = ϕ∗ψ∗X.

3. The forms on a manifold comprise a real associative algebra with ∧as multiplication. Furthermore, α∧β = (−1)klβ∧α for k and l-formsα and β, respectively.

4. For maps ϕ and ψ,

ϕ∗(α ∧ β) = ϕ∗α ∧ ϕ∗β and (ϕ ψ)∗α = ψ∗ϕ∗α.

5. d is a real linear map on forms, ddα = 0, and

d(α ∧ β) = dα ∧ β + (−1)kα ∧ dβ

for α a k-form.

6. For α a k-form and X0, . . . , Xk vector fields,

(dα)(X0, . . . , Xk) =k∑i=0

(−1)iXi[α(X0, . . . , Xi, . . . , Xk)]

+∑

0≤i<j≤k(−1)i+jα([Xi, Xj ], X0, . . . , Xi, . . . , Xj , . . . , Xk)

where Xi means that Xi is omitted. Locally,

dα(x)(v0, . . . , vk) =k∑i=0

(−1)iDα(x) · vi(v0, . . . , vi, . . . , vk).

7. For a map ϕ, ϕ∗dα = dϕ∗α.

8. Poincare Lemma. If dα = 0, then α is locally exact; that is, thereis a neighborhood U about each point on which α = dβ. The sameresult holds globally on a contractible manifold.

58 2. Mathematical preliminaries

9. iXα is real bilinear in X, α and for h : M → R,

ihXα = hiXα = iXhα.

Also, iX iXα = 0 and

iX(α ∧ β) = iXα ∧ β + (−1)kα ∧ iXβ

for α a k-form.

10. For a diffeomorphism ϕ,

ϕ∗(iXα) = iϕ∗X(ϕ∗α);

if f : M → N is a mapping and Y is f -related to X, that is, Tf X =Y f , then

iY f∗α = f∗iXα.

11. £Xα is real bilinear in X, α and

£X(α ∧ β) = £Xα ∧ β + α ∧£Xβ.

12. Cartan’s Magic Formula

£Xα = diXα+ iXdα.

13. For a diffeomorphism ϕ,

ϕ∗£Xα = £ϕ∗Xϕ∗α;

if f : M → N is a mapping and Y is f -related to X, then

£Y f∗α = f∗£Xα.

14. (£Xα)(X1, . . . , Xk) = X[α(X1, . . . , Xk)]

−k∑i=0

α(X1, . . . , [X,Xi], . . . , Xk).

Locally,

(£Xα)(x) · (v1, . . . , vk)

= (Dαx ·X(x))(v1, . . . , vk)

+k∑i=0

αx(v1, . . . ,DXx · vi, . . . , vk).

2.4 Lie derivatives 59

15. The following identities hold:

(a) £fXα = f£Xα, £Xfα = f£Xα+ df ∧ iXα;

(b) £[X,Y ]α = £X£Y α−£Y £Xα;

(c) i[X,Y ]α = £X iY α− iY £Xα;

(d) £Xdα = d£Xα; and

(e) £X iXα = iX£Xα.

16. IfM is a finite-dimensional manifold,X = X l∂/∂xl, and α = αi1...ikdxi1∧

· · · ∧ dxik , where i1 < · · · < ik, then the following formulas hold:

dα =(∂αi1...ik∂xl

)dxl ∧ dxi1 ∧ · · · ∧ dxik ,

iXα = X lαli2...ikdxi2 ∧ · · · ∧ dxik ,

£Xα = X l

(∂αi1...ik∂xl

)dxi1 ∧ · · · ∧ dxik

+ αli2...ik

(∂X l

∂xi1

)dxi1 ∧ dxi2 ∧ . . . ∧ dxik + . . . .

Exercises

¦ 2.4-1. Let M be an n-manifold, ω ∈ Ωn(M) a volume form, X,Y ∈X(M), and f, g : M → R smooth functions such that f(m) 6= 0 for all m.Prove the following identities:

(a) divfω(X) = divω(X) +X[f ]/f ;

(b) divω(gX) = g divω(X) +X[g]; and

(c) divω([X,Y ]) = X[divω(Y )]− Y [divω(X)].

¦ 2.4-2. Show that the partial differential equation

∂f

∂t=

n∑i=1

Xi(x1, . . . , xn)∂f

∂xi

with initial condition f(x, 0) = g(x) has the solution f(x, t) = g(Ft(x)),where Ft is the flow of the vector field (X1, . . . , Xn) in Rn whose flow is

60 2. Mathematical preliminaries

assumed to exist for all time. Show that the solution is unique. Generalizethis exercise to the equation

∂f

∂t= X[f ]

for X a vector field on a manifold M .

¦ 2.4-3. Show that if M and N are orientable manifolds, so is M ×N .

2.5 Stokes’ Theorem, RiemannianManifolds and Distributions

The basic idea of the definition of the integral of an n-form ω on an orientedn-manifold M is to pick a covering by coordinate charts and to sum up theordinary integrals of f(x1, . . . , xn) dx1 · · · dxn, where

ω = f(x1, . . . , xn) dx1 ∧ · · · ∧ dxn

is the local representative of ω, being careful not to count overlaps twice.The change of variables formula guarantees that the result denoted by∫Mω, is well defined.

If one has an oriented manifold with boundary, then the boundary, ∂M ,inherits a compatible orientation. This proceeds in a way that generalizesthe relation between the orientation of a surface and its boundary in theclassical Stokes’ theorem in R3.

Theorem 2.5.1 (Stokes’ Theorem). Suppose that M is a compact, ori-ented k-dimensional manifold with boundary ∂M . Let α be a smooth (k−1)-form on M . Then ∫

M

dα =∫∂M

α. (2.5.1)

Special cases of Stokes’ theorem are as follows:

The integral theorems of calculus. Stokes’ theorem generalizes andsynthesizes the classical theorems:

(a) Fundamental Theorem of Calculus.∫ a

b

f ′(x) dx = f(b)− f(a). (2.5.2)

(b) Green’s Theorem. For a region Ω ⊂ R2:∫ ∫Ω

(∂Q

∂x− ∂P

∂y

)dx dy =

∫∂Ω

P dx+Qdy. (2.5.3)

2.5 Stokes’ Theorem, Riemannian Manifolds and Distributions 61

(c) Divergence Theorem. For a region Ω ⊂ R3:∫ ∫ ∫Ω

div F dV =∫ ∫

∂Ω

F · ndA. (2.5.4)

(d) Classical Stokes’ Theorem. For a surface S ⊂ R3:∫ ∫S

(∂R

∂y− ∂Q

∂z

)dy ∧ dz

+(∂P

∂z− ∂R

∂x

)dz ∧ dx+

(∂Q

∂x− ∂P

∂y

)dx ∧ dy

=∫ ∫

S

n · curl F dA

=∫∂S

P dx+Qdy +Rdz, (2.5.5)

where F = (P,Q,R).

Notice that the Poincare lemma generalizes the vector calculus theoremsin R3 saying that if curl F = 0, then F = ∇f and if div F = 0, thenF = ∇×G. Recall that it states: If α is closed, then locally α is exact; thatis, if dα = 0, then locally α = dβ for some β.

Change of variables. Another basic result in integration theory is theglobal change of variables formula.

Theorem 2.5.2 (Change of Variables). Let M and N be oriented n-manifolds and let F : M → N be an orientation-preserving diffeomorphism.If α is an n-form on N (with, say, compact support), then∫

M

F ∗α =∫N

α.

Riemannian Manifolds. A differentiable manifold with a positive def-inite quadratic form 〈·, ·〉 on every tangent space TMx is called a Rie-mannian manifold. The quadratic form 〈 , 〉, often denoted g(, ) is called aRiemmannian metric.

In local coordinates qi the length of a vector v is then given by

g(v, v) =n∑i,j

gij(q)qiqj , gij = gji .

Let f be a smooth function on M . The gradient flow of f, gradf isdefined by

df(v) = 〈gradf, v〉

for any v ∈ TM .

62 2. Mathematical preliminaries

Frobenius’ theorem. A basic result called Frobenius’ theorem playsa critical role in control theory and we shall have much to say about it laterin the book. For now we just state it briefly as it is normally regarded aspart of the theory of differentiable manifolds. The theory of distributionsplays a key role in both the theory of nonholonomic systems and nonlinearcontrol theory. Two useful references (from the control-theortic) point ofview are Sussmann [1973] and Isidori [1985].

Definition 2.5.3. A smooth distribution on a manifold M is the as-sigment to each point x ∈M of the subspace spanned by the values at x ofa set of smooth vector fields on M ; ie, it is a “smooth assignment of a sub-space to the tangent space at each point, also called a vector subbundle.We denote the distribution by ∆ and the subspace at x ∈M by ∆x ⊂ TxM .

A distribution is said to be involutive if for any two vector fields X,Yon M with values in ∆, [X,Y ] is also a vector field with values in ∆. Thesubbundle ∆ is said to be integrable if for each point x ∈ M there is alocal submanifold of M containing x such that its tangent bundle equals∆ restricted to this submanifold. See Figure 2.5.1.

x∆x

integral manfolds of ∆

Figure 2.5.1. The integral manifolds of a distribution.

If ∆ is integrable, the local integral manifolds can be extended to get,through each x ∈ M , a maximal integral manifold, which is an immersedsubmanifold of M . The collection of all maximal integral manifolds throughall points of M forms a foliation.

Theorem 2.5.4 (Frobenius theorem). Involutivity of ∆ is equivalentto the integrability of ∆, which in turn is equivalent to the existence of afoliation on M whose tangent bundle equals ∆.

2.6 Lie groups, Spatial Relationships, and the Euclidean Group 63

Given a set of smooth vector fields X1, · · · , Xd on M we denote thedistribution defined by their span by

∆ = spanX1, · · · , Xd.

The distribution at any point is denoted by ∆x. A distribution ∆ onM is said to be nonsingular on M if there exists an integer d such thatdim(∆x) = d for all x ∈ M . A point x ∈ M is said to be a regular pointis there exists a neighbohood U of x such that ∆ is nonsingular on U .Otherwise the point is said to singular.

Exercises

¦ 2.5-1. Let Ω be a closed bounded region in R2. Use Green’s theorem toshow that the area of Ω equals the line integral

12

∫∂Ω

(x dy − y dx).

¦ 2.5-2. On R2\(0, 0) consider the one-form

α = (x dy − y dx)/(x2 + y2).

(a) Show that this form is closed.

(b) Using the angle θ as a variable on S1, compute i∗α, where i : S1 → R2

is the standard embedding.

(c) Show that α is not exact.

¦ 2.5-3. The magnetic monopole. Let B = gr/r3 be a vector field onEuclidean three-space minus the origin where r = ‖r‖. Show that B cannotbe written as the curl of something.

2.6 Lie groups, Spatial Relationships, andthe Euclidean Group

Lie groups arise in discussing conservation laws for mechanical and controlsystems and in the anaysis of systems with some underlying symmetry.There is a huge literature on the subject. Useful referernces include Abra-ham and Marsden [1978], Marsden and Ratiu [1994] and Sattinger andWeaver [1986].

Definition 2.6.1. A Lie group is a smooth manifold G which is a groupand for which the group operations of multiplication, (g, h)→ gh for g, h ∈G, and inversion, g → g−1 are smooth (i.e. analytic).

64 2. Mathematical preliminaries

Before giving a brief description of some of the theory of Lie groups wemention some important examples:

1. The group of linear isomorphisms of Rn to itself. This is a Lie groupof dimension n2 general linear group and denoted by GL(n,R).

Definition 2.6.2. A matrix Lie group is a set of invertible n×n matri-ces which is closed under matrix multiplication and which is a submanifoldof Rn×n.

One could equivalently define a matrix Lie group to be a (topologically)closed subgroup of GL(n,R). All of the Lie groups we have mentioned sofar, and indeed all the Lie groups discussed in this book will be matrix Liegroups. Lie groups are frequently studied in conjunction with Lie algebraswhich are associated with the tangent spaces of Lie groups as we nowdescribe. To begin with, we state a generalization of the result establishedin Exercise 2.2.3.

Proposition 2.6.3. Let G be a matrix Lie group, and let A,B ∈ TIG(the tangent space to G at the identity element). Then AB −BA ∈ TIG.

Our proof makes use of the following Lemma.

Lemma 2.6.4. Let R be an arbitrary element of a matrix Lie group G,and let B ∈ TIG. Then RBR−1 ∈ TIG.

Proof. Let RB(t) be a curve in G such that RB(0) = I and R′B(0) = B.Then S(t) = RRB(t)R−1 ∈ G for all t, and S(0) = I. Hence S(0) ∈ TIG,proving the lemma. ¥

Proof of Proposition. Let RA(s) be a curve in G such that RA(0) = Iand R′A(0) = A. The by Lemma, S(t) = RA(t)BRA(t)−1 ∈ TIG. HenceS′(t) ∈ TIG and in particular S′(0) = AB −BA ∈ TIG. ¥

Definition 2.6.5. For any pair of n × n matrices A,B, we define thematrix Lie bracket [A,B] = AB −BA.

Proposition 2.6.6. The matrix Lie bracket operation has the followingtwo properties:

(i) For any n × n matrices A and B, [B,A] = −[A,B]. (Property ofskew-symmetry)

(ii) For and n× n matrices A, B, and C,

[[A,B], C] + [[B,C], A] + [[C,A], B] = 0

(This is known as the Jacobi identity.)

The proof of this proposition involves a straightforward calculation andis left to the reader.

2.6 Lie groups, Spatial Relationships, and the Euclidean Group 65

Definition 2.6.7. A (matrix) Lie algebra g is a set of n × n matriceswhich is a vector space with respect to the usual operations of matrix addi-tion and multiplication by real numbers (scalars), and which is closed underthe matrix Lie bracket operation [·, ·].Proposition 2.6.8. For any matrix Lie group, G, the tangent space atthe indentity TIG is a Lie algebra.

Proof. This is an immediate consquence of the fact that TIG is a vectorspace and Proposition 2. ¥

It is a remarkable and remarkably useful fact that a great deal of thestructure of a Lie group may be inferred from studying the Lie algebra.Before discussing important general relationships between Lie groups andLie algebras, we describe several examples which play an important role inmechanics and control.

The special orthogonal group. The set of all elements of O(n) havingdeterminant = 1 is a subgroup called the special orthogonal group. This isdenoted by SO(n). Because any X ∈ O(n) satisfies XXT = I, it followsthat detX = ±1. We could also characterize SO(n) as the connected com-ponent of the identity element in O(n). Thus, TISO(n) = TIO(n). Fromthis observation and the calculation carried out in Example 2.2.2, TISO(n)is just the set of n × n skew-symmetric matrices. Throughout this book,we reserve the notation so(n) for the Lie algebra TISO(n).

The symplectic group. Suppose n = 2` and consider the non-singular

skew-symmetric matrix J =(

0 I−I 0

)where I = n× n identity matrix.

It is an exercise left to the reader to verify that

Sp(`) = X ∈ Gl(2`) : XJXT = J

is a group. It is called the symplectic group. Again referring to Example2.2.2, we find that this matrix Lie algebra TISp(`) is the set of n × nmatrices satisfying JY T +Y J = 0. Throughout the remainder of the book,we denote this Lie algebra by sp(`).

Let Y ∈ sp(`) be partitioned into ` × ` blocks, Y =(A BC D

). Write

down a complete set of equations involving A,B,C, and D which must besatisfied if Y ∈ sp(`). Deduce that the dimension of sp(`) as a real vectorspace is 2`2 + ` = n(n+ 1)/2, and consequently dimSp(`) = 2`2 + `.

The Heisenberg group. Consider the set of all 3 × 3 matrices of theform 1 x y

0 1 z0 0 1

66 2. Mathematical preliminaries

where x, y, and z are real numbers. It is straightforward to show that thisis a group, and since it is a submanifold of the set of all 3× 3 matrices, itis a Lie group. Call it H. The corresponding Lie algebra may be writtendown from the definition. Specifically,

X1(t) =

1 t 00 1 00 0 1

X2(t) =

1 0 00 1 t0 0 1

X3(t) =

1 0 t0 1 00 0 1

are three curves in H which pass through the identity when t = 0. The

derivatives X ′i(0) are elements of the Lie algebra: A =

0 1 00 0 00 0 0

,

B =

0 0 00 0 10 0 0

, and C =

0 0 10 0 00 0 0

respectively. Since three pa-

rameters are used to specify H, H is three dimensional. A, B, and C arelinearly independent and span the Lie algebra. The commutation relationsfor the Lie brackets of these three basis elements are: [A,B] = C, [A,C] = 0,and [B,C] = 0. This Lie algebra is call the Heisenberg algebra.

The Euclidean group. Consider the Lie group of all 4× 4 matrices of

the form(R v0 1

)where R ∈ SO(3) and v ∈ R3. This group is usually

denoted by SE(3) and is called the special Euclidean group. The corre-sponding Lie algebra, se(3), is six dimensional and is spanned by

0 0 0 00 0 −1 00 1 0 00 0 0 0

,

0 0 1 00 0 0 0−1 0 0 00 0 0 0

,

0 −1 0 01 0 0 00 0 0 00 0 0 0

,

0 0 0 10 0 0 00 0 0 00 0 0 0

,

0 0 0 00 0 0 10 0 0 00 0 0 0

,

0 0 0 00 0 0 00 0 0 10 0 0 0

.

The special Euclidean group is of central interest in mechanics since it de-scribes the set of rigid motions and coordinate transformations of 3-space.More specifically, suppose there are two coordinate frames A and B lo-cated in space such that the origin of the B-frame has A-frame coordinatesv = (v1, v2, v3)T and such that the unit vectors in the principal B-frame

2.6 Lie groups, Spatial Relationships, and the Euclidean Group 67

cooridnate directions are (r11, r21, r31)T , (r12, r22, r32)T , and (r13, r23, r33)T

with respect to A-frame coordinates. The rigid motion which moves the A-frame into coindcidence with the B-frame is specified by the rotation r11 r12 r13

r21 r22 r23

r31 r32 r33

followed by the translation v = (v1, v2, v3)T . Thus a point with A-framecoordinates x = (x1, x2, x3) is moved under the rigid motion to a newlocation whose A-frame coordinates are Rx+ v.

The group SE(3) is also associated with the set of rigid coordinate trans-formations of R3. Suppose a point Q is located in space and has A-framecoordinates (xA1 , x

A2 , x

A3 )T and B-frame coordinates (xB1 , x

B2 , x

B3 )T . The re-

lationship between these coordinate descriptions is given by

xA = RxB + v.

Exercises

¦ 2.6-1.

A point P in R3 undergoes a rigid motion associated with(R1 v1

0 1

)followed by a rigid motion associated with

(R2 v2

0 1

). What matrix el-

ement of SE(3) is assoicated with the composite of these motions in thegiven order.

¦ 2.6-2. A coordinate frame B is located with respect to a coordinate frameA as follows. B is initially cooincident with A, but is displaced by the

rigid motion associated with(R1 v1

0 1

)and is then subsequently further

displaced by(R2 v2

0 1

). What matrix element of SE(3) is associated

with the coordinate transformation from the A frame to the B frame? (I.e.,what matrix element of SE(3) is used to describe A-frame coordinates ofa point in terms of thhe B-frame coordinates of the same point?)

Let G be a matrix Lie group and let g = TIG be the corresponding Liealgebra. The dimensions of the differentiable manifold G and the vectorspace g are of course the same, and there must be a one-one local corre-spondence between a neighborhood of 0 in g and a neighborhood of theidentity element I in G. One explicit local correspondence is provided bythe exponential mapping exp : g→ G which we now describe.

68 2. Mathematical preliminaries

Let A ∈ Rn×n (= the space of n× n matrices). We define exp(A) by theseries

I +A+12A2 +

13!A3 + · · · . (2.6.1)

Proposition 2.6.9. The series (2.6.1) is absolutely convergent.

Proof. The ij-th entry in the n-th term of this matrix series is boundedin absolute value by (n − 1)an/n! where a = maxij|aij |. Hence, theij-th element in each term in the series is bounded in absolute value bythe corresponding term in the absolutely convergent series ean = 1 + an+12 a

2n2+· · · . Hence each entry in the series of matrices converges absolutely,proving the proposition. ¥

Proposition 2.6.10. Let G be a matrix Lie group with corresponding Liealgebra g. If A ∈ g, then exp(At) ∈ G for all real numbers t.

Exercise

¦ 2.6-3. Suppose the n×n matrices A and M satisfy AM+MAT = 0. Showthat exp(At)Mexp(AT t) = M for all t. This direct calculation shows thatfor A ∈ so(n) or A ∈ sp(`), we have exp(At) ∈ SO(n) or exp(At) ∈ Sp(`)respectively.

2.7 Fiber Bundles, Connections, and GaugeTheory

We assume the reader is familar with the elementary theory of smoothmanifolds. The remainder of the treatment will also be brief. For morecomprehensive treatments, one may consult one of the standard texts.

2.7.1 Fiber Bundles

Fiber bundles provide a basic geometric structure for the understandingof many mechanical and control problems, in particular for nonholonomicproblems. Good references include Abraham and Marsden [1978] and Steen-rod [1951] and Schutz [1980].

A fiber bundle essentially consists of a given space (the base) togetherwith another space (the fiber) attached at each point, plus some compati-bility and smoothness conditions.

More formally we have the following:

Definition 2.7.1. A fiber bundle is a space Q for which the following aregiven: another space B called the base manifold, a projection π : Q → B

2.7 Fiber Bundles, Connections, and Gauge Theory 69

with fibers π−1(b), b ∈ B homeomorphic to a space F , a structure group Gof homeomorphisms of F into itself and a covering of B by open sets Uj,satisfying

i) the bundle is locally trivial, i.e. π−1(Uj) is homeoemorphic to the prod-uct space Uj × F and ii) if hj is the map giving the homeomorphism onthe set Uj, for any x ∈ Uj ∩ Uk hj(h−1

k (x)) is an element of the structuregroup G

If the fibers of the bundle are homeomorphic to the structure group wecall the bundle a principal bundle.

If the fibers of the bundle are homeomorphic to a vector space we callthe bundle a vector bundle.

Connections. An important additional structure on a bundle is an con-nection or Ehresmannn connection, see for example Kobayashi andNomizu [1963], Marsden, Montgomery and Ratiu [1990] or Bloch, Krish-naprasad, Marsden and Ratiu [1996]. We follow the treatment in the latterhere.

However, before we give the precise mathematical definitions we will givea somewhat intuitive discussion of the nature of and need for connections.A nice reference in this regard is the book by Burke [1985].

Suppose we have a bundle and consider (locally) a section of the bundle:i.e. a choice at of a point in the fiber over each point in the base. Let uscall such a choice a “field.”

The idea is to single out fields which are “constant”. For vector fieldson the plane for example, it is clear what these are. For vector fields on amanifold or far an arbitrary bundle, we have to specify this notion. Suchfields are called “horizontal” and are also key to defining a notion of deriva-tive – or rate of change of a vector field. A connection is used to single outhorizontal fields, and is chosen to have other desirable structure, such aslinearity. For example, the sum of two constant fields should still be con-stant. As we shall see below, we can specify horizontality by taking a classof fields that are the kernal of a suitable form.

More formally:Let Q be a differentiable manifold.Consider a bundle with projection map π and as usual let Tqπ denote

its tangent map at any point. We call the kernel of Tqπ at any point thevertical space and denote it by Vq.

Definition 2.7.2. An Ehresmann connection A is a vertical valued oneform on Q that satisfies

1. Aq : TqQ→ Vq is a linear map for each point q ∈ Q2. A is a projection: A(vq) = vq for all vq ∈ Vq.The key property of the connection is the following: if we denote by Hq or

horq the kernel of Aq and call it the horizontal space, the tangent space toQ is the direct sum of the Vq and Hq, i.e. we can split the tangent space to

70 2. Mathematical preliminaries

Q into horizontal and vertical parts. For example, we can project a tangentvector onto its vertical part using the connection. Note that the verticalspace at Q is tangent to the fiber over q.

Later on when we discuss nonholonomic systems we shall choose theconnection so that the constraint distribution is the horizontal space of theconnection.

Now define the bundle coordinates qi = (rα, sa) for the base and fiber.The coordinate representation of the projection πQ,R is just projection ontothe factor r and the connection A can be represented locally by a vectorvalued differential form ωa:

A = ωa∂

saωa(q) = dsa +Aaα(r, s)drα.

We can see this as follows:Let

vq =∑β

rβ∂

∂rβ+∑b

sb∂

∂sb

be an element of TqQ. Then

ωa(vq) = sa +Aaαrα

and

A(vq) = (sa +Aaαrα)

∂sa.

This clearly demonstrates that A is a projection since when A acts againonly dsa results in a nonzero term and this has coefficient unity.

Note that we use a different notation, namely ωa, for the local coordinaterepresentation of the connection A for three reasons. First, it is commonin the literature to use ω to stand for constraint one forms. Second, in thepreceding formula, it is standard to define the components of the connectionA by Aaα as shown, reflecting the fact that the connection is a projection; todistinguish this use of indices on A from the use of indices on the constraintone forms, it is convenient to use a different letter. Third, we want to regardωa as (coordinate dependent) differential forms, as opposed to A which isa vertical valued form; again, a different letter emphasizes this fact. Notein particular, that the exterior derivative of A is not defined, but we can(locally) take the exterior derivative of ωa. In fact, this will give an easyway to compute the curvature as we shall see.

Given an Ehresmann connection A, a point q ∈ Q and a vector vr ∈ TrRtangent to the base at a point r = πQ,R(q) ∈ R, we can define the horizontallift of vr to be the unique vector vhr in Hq that projects to vr under TqπQ,R.If we have a vector Xq ∈ TqQ, we shall also write its horizontal part as

horXq = Xq −A(q) ·Xq.

2.7 Fiber Bundles, Connections, and Gauge Theory 71

In coordinates, the vertical projection is the map

(rα, sa) 7→ (0, sa +Aaα(r, s)rα)

while the horizontal projection is the map

(rα, sa) 7→ (rα,−Aaα(r, s)rα).

Next, we recall the basic notion of curvature.

Definition 2.7.3. The curvature of A is the vertical valued two formB on Q defined by its action on two vector fields X and Y on Q by

B(X,Y ) = −A([horX,horY ])

where the bracket on the right hand side is the Jacobi-Lie bracket of vectorfields obtained by extending the stated vectors to vector fields.

Notice that this definition shows that the curvature exactly measures thefailure of the constraint distribution to be an integrable bundle.

A useful standard identity for the exterior derivative dα of a one form α(which could be vector space valued) on a manifold M acting on two vectorfields X,Y is

(dα)(X,Y ) = X[α(Y )]− Y [α(X)]− α([X,Y ]).

This identity shows that in coordinates, one can evaluate the curvatureby writing the connection as a form ωa in coordinates, computing its ex-terior derivative (component by component) and restricting the result tohorizontal vectors, that is, to the constraint distribution. In other words,

B(X,Y ) = dωa(horX,horY )∂

∂sa,

so that the local expression for curvature is given by

B(X,Y )a = Baα,βXαY β (2.7.1)

where the coefficients Baα,β are given by

Bbαβ =

(∂Abαrβ−∂Abβrα

+Aaα∂Abβsa−Aaβ

∂Abαsa

). (2.7.2)

2.7.2 Connections on TR1

The idea of a connection can be illustrated by considering the simplestpossible example: a connection on the bundle TR1 with coordinates (x, x).We may define the horizontal space to be the kernal of the form

dx+A(x, x)dx.

72 2. Mathematical preliminaries

More specifically we can choose a linear connection

dx+ a(x)xdx.

(More on linear connections below.) Here A is the R-valued form

(dx+ a(x)xdx)∂

∂x.

Elements of TqQ are of the form

vq = x∂

∂x+ x

∂x

and their projection onto the vertical space is

A(vq) = (x+ a(x)x2)∂

∂x.

The kernal of A, i.e. the horizontal vectors, is the span of

∂x− a(x)x

∂x.

Note that the standard choice is a(x) = 0, i.e. the standard horizontal spaceis the span of the vectors ∂

∂x .Note that we can also use the above argument to define geodesic curves.

Suppose we have a curve x(t) such that its tangent vector is parallel trans-ported along the curve, i.e. vq along the curve is always horizontal, or A(vq)is zero.

From above this means

x+ a(x)x2 = 0 .

For a(x) = 0 this reduces x = 0, the equation of motion for a free particleon the line. We have given the generalization of this equation for arbitraryconnections.

2.7.3 Linear connections, affine connections andgeodesics

In this section we consider how Ehresmann connections specialize to linearconnections and affine connections in the tangent bundle, and we shallderive the geodesic equations. (As above a good related reference for someof these ideas, but with a rather different approach, is Burke [1985].)

As above we use bundle coordinates rα, sa) and we specify the connectionby the one forms

ωa(q) = dsa +Aaα(r, s)drα

2.7 Fiber Bundles, Connections, and Gauge Theory 73

and the action of A on a tangent vector vq = (rα, sa) is given by

A(vq) = (sa +Aaαrα)

∂sα. (2.7.3)

For a linear connections we require that the sum of two (local) horizontalsections is horizontal, i.e. if (rα, sa(r)) and (rα, sa(r)) are horizontal thenso should be (rα, sa(r) + sa(r).)

Thus if we have

sa +Aaαrα = ˙s

a+Aaαr

α = 0

we require

sa + ˙sa

+Aaαrα = 0 .

This (plus scaling) requires the connection coefficients be of the form

Aaα(r, s) = Γaαb(r)sb . (2.7.4)

If the bundle is the tangent bundle these are called the components ofthe affine connection in the tangent bundle.

In the tangent bundle sa = ra. We define geodesic motion along a curver(t) as being that for which the tangent vector is parallel transported alongthe curve, i.e. vq along the curve is always horizontal, or A(vq) is zero, asin the example.

From 2.7.3 this condition is

ra + Γabcrbrb = 0 . (2.7.5)

This is the equation of geodesic motion.We can also determine this equation as follows:In the tangent bundle we can specify a linear connection by its action

on vector fields, or by a map from vector fields (X,Y ) to the vector field∇XY which satisfiees smooth functions f and g and a vector fields X,Y, Z

(i) ∇fX+gY Z = f∇XZ + g∇Y Z

(ii) ∇X(Y + Z) = ∇XY +∇XZ

(iii) ∇X(fY ) = f∇XY + (df ·X)Y

where df ·X is the directional derivative of f along X, or Lie derivative.Given a basis of vector fields ∂

∂rjwe can represent ∇ by

∇ ∂∂ri

∂rj= Γkij

∂rk .(2.7.6)

74 2. Mathematical preliminaries

The geodesic equations above then may be written

∇r r = 0 . (2.7.7)

We can see this by a simple computation:

∇ri ∂∂ri

rj∂

∂rj= ri∇ ∂

∂ri

rj∂

∂rj

= ri∂

∂rirj

∂rj+ rirjΓkij

∂rk

= (rj + Γjikrirk)

∂rj.

Sometimes we will write

∇r r =D2r

dt2. (2.7.8)

Here Drdt denotes the covariant derivative. A slightly more formal treatment

of some of this material will be given below.

2.7.4 Principal Connections

We now consider the special case of Principal Connections. We start with afree and proper group action of a Lie group on a manifold Q and constructthe projection map π : Q → Q/G; this setup is also referred to as aprincipal bundle . The kernel kerTqπ (the tangent space to the grouporbit through q) is called the vertical space of the bundle at the point qand is denoted by verq.

Definition 2.7.4. A principal connection on the principle bundle π :Q→ Q/G is a map (referred to as the connection form) A : TQ→ g thatis linear on each tangent space (i.e., A is a g-valued one form) and is suchthat

1. A(ξQ(q)) = ξ for all ξ ∈ g and q ∈ Q, and

2. A is equivariant:

A(TqΦg(vq)) = AdgA(vq)

for all vq ∈ TqQ and g ∈ G, where Φg denotes the given action of Gon Q and where Ad denotes the adjoint action of G on g.

The horizontal space of the connection at a point q ∈ Q is the linearspace

horq = vq ∈ TqQ | A(vq) = 0 .

2.7 Fiber Bundles, Connections, and Gauge Theory 75

Thus, at any point, we have the decomposition

TqQ = horq ⊕ verq.

Often one finds connections defined by specifying the horizontal spaces(complementary to the vertical spaces) at each point and requiring thatthey transform correctly under the group action. In particular, notice thata connection is uniquely determined by the specification of its horizontalspaces, a fact that we will use later on. We will denote the projections ontothe horizontal and vertical spaces relative to the above decomposition usingthe same notation; thus, for vq ∈ TqQ, we write

vq = horqvq + verqvq.

The projection onto the vertical part is given by

verqvq = (A(vq))Q(q)

and the projection to the horizontal part is thus

horqvq = vq − (A(vq))Q(q).

The projection map at each point defines an isomorphism from the horizon-tal space to the tangent space to the base; its inverse is called the horizontallift. Using the uniqueness theory of ODE’s one finds that a curve in thebase passing through a point π(q) can be lifted uniquely to a horizontalcurve through q in Q (i.e., a curve whose tangent vector at any point is ahorizontal vector).

Since we have a splitting, we can also regard a principal connection asa special type of Ehresmann connection. However, Ehresmann connectionsare regarded as vertical valued forms whereas principal connections areregarded as Lie algebra valued. Thus, the Ehresmann connection A andthe connection one form A are different and we will distinguish them; theyare related in this case by

A(vq) = (A(vq))Q(q).

The general notions of curvature and other properties which hold forgeneral Ehresmann connections specialize to the case of principal connec-tions. As in the general case, given any vector field X on the base space(in this case, the shape space), using the horizontal lift, there is a uniquevector field Xh that is horizontal and that is π-related to X; that is, ateach point q, we have

Tqπ ·Xh(q) = X(π(q))

and the vertical part is zero:

(A(Xhq ))Q(q) = 0.

76 2. Mathematical preliminaries

It is well known (see, for example, Abraham, Marsden and Ratiu [1988])that the relation of being π-related is bracket preserving; in our case, thismeans that

hor [Xh, Y h] = [X,Y ]h,

where X and Y are vector fields on the base.

Definition 2.7.5. The covariant exterior derivative D of a Lie alge-bra valued one form α is defined by applying the ordinary exterior derivatived to the horizontal parts of vectors:

Dα(X,Y ) = dα(horX,horY ).

The curvature of a connection A is its covariant exterior derivative andit is denoted by B.

Thus, B is the Lie algebra valued two form given by

B(X,Y ) = dA(horX,horY ).

Using the identity

(dα)(X,Y ) = X[α(Y )]− Y [α(X)]− α([X,Y ]).

together with the definition of horizontal shows that for two vector fieldsX and Y on Q, we have

B(X,Y ) = −A([horX,horY ])

where the bracket on the right hand side is the Jacobi-Lie bracket of vectorfields. The Cartan structure equations say that if X and Y are vector fieldsthat are invariant under the group action, then

B(X,Y ) = dA(X,Y )− [A(X),A(Y )]

where the bracket on the right hand side is the Lie algebra bracket. Thisfollows readily from the definitions, the fact that [ξQ, ηQ] = −[ξ, η]Q, thefirst property in the definition of a connection, and writing horX = X −verX and similarly for Y , in the preceding formula for the curvature.

Next, we give some useful local formulas for the curvature. To do this,we pick a local trivialization of the bundle; that is, locally in the base, wewrite Q = Q/G × G where the action of G is given by left translation onthe second factor. We choose coordinates rα on the first factor and a basisea of the Lie algebra g of G. We write coordinates of an element ξ relativeto this basis as ξa. Let tangent vectors in this local trivialization at thepoint (r, g) be denoted (u,w). We will write the action of A on this vector

2.7 Fiber Bundles, Connections, and Gauge Theory 77

simply as A(u,w). Using this notation, we can write the connection formin this local trivialization as

A(u,w) = Adg(wb +Aloc(r) · u), (2.7.9)

where wb is the left translation of w to the identity (that is, the expressionof w in “body coordinates”). The preceding equation defines the expressionAloc(r). We define the connection components by writing

Aloc(r) · u = Aaαuαea.

Similarly, the curvature can be written in a local representation as

B((u1, w1), (u2, w2)) = Adg(Bloc(r) · (u1, u2)),

which again serves to define the expression Bloc(r). We can also define thecoordinate form for the local expression of the curvature by writing

Bloc(r) · (u1, u2) = Baαβuα1 uβ2 ea.

Then one has the formula

Bbαβ =

(∂Abβrα− ∂Abα

rβ− CbacAaαAcβ

),

where Cbac are the structure constants of the Lie algebra defined by

[ea, ec] = Cbaceb.

2.7.5 Parallel Translation and Holonomy Groups

Let P be a principal bundle with a connection and C a piecewise differen-tiable curve in its base space M with beginning point p and endpoint q.Suppose x is a point on the fiber over p. Then there is unique curve C∗xin P starting at x such that π(C∗x) = C and each tangent vector to C ishorizontal. C∗x is said to be a lift of C that starts at x and the map thattakes x to the lift of q, the endpoint of the lifted curve, is said to be paralleltranslation.

Now suppose C is a closed curve starting at p. Parallel translation thenmaps the point x in the fiber over p to itself, to a point xa say, a ∈ G.Thus each closed curve at p and fiber point x determines an element of Gand the set of all such elements forms a subgroup of G called the holonomygroup of the connection with reference point x.

2.7.6 Affine Connections and Covariant Derivatives

Now let M be a differential manifold and let P be the bundle of tangent n-frames overM , i.e. the bundle whose fiber consists of n independent tangent

78 2. Mathematical preliminaries

vectors at each point. Then P is a principal bundle with structure groupGl(n,R) and is the principal bundle associated with the tangent vectorbundle of M . A connection in this bundle is called an affine connection.

An affine connection on M gives a parallel displacement of the tangentvector space of M as follows:

Let C = pt(0 ≤ t ≤ 1) be a curve in M and let C∗ = xt be a lift to P .The parallel displacment of the tangent frame x0 at p0 along the curve isxt and defines a parallel displacement of Tp0(M) to Tpy (M) along C whichis independent of the choice of lift.

Now suppose that we have a vector Yt in TptM such that the map from tto Yt is differentiable, i.e. Yt is a differentiable vector field along the curve.Let φt,h denote the parallel displacement of Tpt(M) onto Tpt+h(M) alongC. Then the vector field

Y ′ = limh→0

1/h(φ−1t,h(Yt+h)− Yt)

is called the covariant derivative of Y along C. Yt is parallel along C, i.e.Yt = φ0,t(Y0) if and only if Y ′t is identically zero.

A geodesic curve is a curve whose tangent vectors are parallel along thecurve.

Now letX and Y be given vector fields onM and let C = pt, (−ε ≤ 0 ≤ ε)be an integral curve of X through p0 and let φt be the parallel displacementalong C. Then the covariant derivative of Y in the direction of X is givenby

(∇XY )p0 = limt→0

1/t(φ−1t (Ypt)− Yp0) .

∇XY is a vector field on Y such that the mapping (X,Y ) → ∇XYsatisfies: a)∇XY is linear with respect to X and Y , b) ∇fXY = f · ∇XYand c)∇X(fY ) = f ·∇XY +(Xf)·Y , for f any differentiable function onM .In fact given any mapping satisfying these conditions there exists a uniqueaffine connection whose covariant derivative is the given mapping. Thecovariant derivative can be naturally extended to tensor fields of arbitrarytype.

For X, Y and Z arbitrary vector fields on M the curvature tensor R andthe torsion tensor are defined by

R(X,Y )(Z) = ∇X(∇Y Z)−∇Y (∇XZ)−∇[X,Y ](Z) ,

and

T (X,Y ) = ∇XY −∇YX − [X,Y ] .

Now let (x1, · · ·xn) be local coordinates on M and let Xi be the vectorfields ∂/∂xi. Then for an affine connection the covariant derivative may beexpressed as

∇Xj (Xk) =∑i

ΓijkXi .

2.7 Fiber Bundles, Connections, and Gauge Theory 79

2.7.7 Riemannian Connections

Now suppose M is endowed with a Riemannian metric g. This means wecan define orthonormal bases of Tp(M) at each p ∈ M and can definea subbundle P ′ of P whose fibers are orthonormal bases and which hasstructure group O(n). This subbundle is said to be a reduced bundle of P .

There exists a unique affine connection on M , called the Riemmanianconnection,, such that ∇g = 0 and the torsion tensor T vanishes. An affineconnection called a metric connection if ∇g = 0.

If the metric is given by g =∑gijdx

idxj the connection coefficients,which are called Christoffel symbols, are given by

Γijk =12

∑l

gil∂gjl∂xk

+∂gik∂xj

− ∂gjk∂xl .

80 2. Mathematical preliminaries

This is page 81Printer: Opaque this

3Basic concepts in mechanics

3.1 First Principles - Coordinates andKinematics

The most basic goal of analytical mechanics is to provide a formalism fordescribing motion. This is conveniently done in terms of a set of general-ized coordinates. This is a set of variables whose values uniquely specify thelocation in 3-space of each physical point on the mechanism. Aset og gen-eralized coordinates is minimal in the sense that no set of fewer variablessuffices to determine the locations of all points on the mechanism. Thenumber of variables in a set of generalized coordinates for a mechanicalsystem is called the number of degrees of freedom of the system.

3.2 First Principles - Conservation ofEnergy, Work, Virtual Work

Analytical mechanics seeks to provide a formal description of the motionof mechanical systems which is as simple as possible. Over many centuries,people who studied the motions of physical systems have developed guidingprinciples or “laws” which govern the movement of all systems. An exampleis the principle of conservation of energy. This principle states that in any“closed” mechanical system, the total energy is conserved. In the physicalworld, the principle always holds, although it may be difficult in specific

82 3. Basic concepts in mechanics

cases to be explicit in specifying the extent of the closed system. For systemsabout which it is possible to be explicit, however, the principle may be usedto guide us in the development of mathematical models of motion.

Simple ideas along this line, which will be generalized to provide thefoundation of most of the models studied in this book, may be illustratedusing the simple kinematic chain illustrated in Figure 1. Here there aredrawn two copies of the same mechanism. In the first, the motion of a typi-cal point, P , is described in terms of coordinate variables (θ1, θ2), where θ2

is the relative angle between the two links in the chain. In Figure ??(b), themotion of the typical point P is described in terms of coordinate variables(φ1, φ2), which are the (absolute) angles of the links with respect to thevertical direction. Other choices of coordinate variables are, of course, pos-sible. In any case, the coordinate variables serve the purpose of describingthe location of typical points of the mechanism with respect to some fixedframe of reference—which we may refer to as the inertial frame.

Specifically, in this case, the inertial frame is chosen to its origin at thehinge point of the upper link. The y-axis is directed parallel and opposite tothe gravitational field, and the x-axis is chosen so as to give the coordinateframe the standard orientation. Suppose the point P is located on thesecond link, as depicted. If this has coordinates (x`, y`) with respect tothe local frame, then the coordinates with respect to the inertial frame aregiven by[

xy

]=[r1 sin θ1 + x` sin(θ1 + θ2) + y` cos(θ1 + θ2)−r1 cos θ1 − x` cos(θ1 + θ2) + y` sin(θ1 + θ2)

](3.2.1)

or equivalently by[xy

]=[r1 sinφ1 + x` sinφ2 + y` cosφ2

−r1 cosφ1 − x` cosφ2 + y` sinφ2

]. (3.2.2)

The mappings (θ1, θ2) 7→ (x, y) and (φ1, φ2) 7→ (x, y) are examples ofkinematics functions. They associate values of the coordinate variables(θ1, θ2) (respectively (φ1, φ2)) to inertial coordinates of the point P . Forevery point on the kinematic chain mechanism of Figure ??, there is akinematics function f(·, ·, P ) for which the formula[

xy

]= f(θ1, θ2, P )

associates inertial coordinates (x, y) to values of the joint angle coordinates.When kinematics functions are as simple to write down explicitly as in thepresent example, they provide a powerful tool for describing motion sincethe inertial coordinates of any point are completely determined by (θ1, θ2).

An interesting question in mechanics is: when do the values of a finite setof configuration variable (such as (θ1, θ2) or (φ1, φ2)) determine the config-uration of an entire mechanism (by means of a set of kinematics functions

3.3 Newtonian, Lagrangian, and Hamiltonian Mechanics 83

f(·, P ))? As we have seen (**shall see**), there are simple examples of sys-tems wherein a physically interesting finite set of configuration variables(θ1(·), . . . , θn(·)) (Think of the wheels on a car!) determines the configura-tion of the entire mechanism only if we know the entire time history of theθj(·)’s over the interval of interest. Such systems will be called nonholo-nomic. One of the aims of this book is to develop a theory of mechanicsand control of both holonomic and nonholonomic systems.

It will be useful to describe a general procedure for constructing kine-matics functions for open single-strand spatial kinematic chains. . . .

3.3 Newtonian, Lagrangian, andHamiltonian Mechanics

Newtonian mechanics starts from the premise that in every physical systemthe forces due to the acceleration of each particle exactly balance all otherforces acting on the particle. If ~x(t) denotes the coordinates of a particleof mass m at time t with respect to some fixed (but arbitrary) coordinateframe, then the motion of the particle is elegantly and simply described bythe differential equation

m ¨~x(t) = F (t). (3.3.1)

Since this equation completely describes the motion of the particle, themotion of any assemblage of particles could in principle be described bysolving (3.3.1) for each particle in the system. Since this is not feasiblefor most systems of physical interest, we turn our attention the study ofgeneralized coordinates, which are minimal systems of variables in terms ofwhich the motion of the entire system can be described.

Consider, for example, a planar kinematic chain, anchored at one end asdepicted in Figure (3.3.1). figure missing

Figure 3.3.1. A planar kinematic chain.

Within each link there is affixed a coordinate system, and a “base” co-ordinate frame is prescribed which serves as a reference for the motion ofthe entire system. In each frame, the x and y axes are as depicted, andthe z-axis is assumed to be perpendicular to the plane of the page. Let atypical physical point on the k-th link have coordinates (xk, yk, zk) withrespect to the k-th link frame. The coordinates with respect to the base

84 3. Basic concepts in mechanics

frame are then given byx(θ1, . . . , θk;xk, yk, zk)y(θ1, . . . , θk;xk, yk, zk)z(θ1, . . . , θk;xk, yk, zk)

=k−1∑j=1

rj cos(θ1...j)rj sin(θ1...j)

0

+

xk cos(θ1...k)− yk sin(θ1...k)xk sin(θ1...k) + yk cos(θ1...k)

zk

. (3.3.2)

As the mechanism undergoes any motion, both the velocity ~x and theacceleration ~x with repsect to the base frame may be computed explicitlyusing this formula.

The motion of every infinitessimal piece of this chain is governed byNewton’s law, which we may write as

m(~x)~x dV = ~Fx dV (3.3.3)

where m(~x) is the mass density, ~Fx is the force density, and dV is thevolume element defined on the link. Although the motion of the k-th linkin the chain is completely described in terms of θ1, . . . , θk, for notationalconvenience, and to pursue the study of more general systems we shallwrite ~x = ~x(θ1, . . . , θn;xk, yk, zk). We wish to study the dynamics of ourchain purely in terms of the θi’s, and our goal will be to use the basic forcebalance law (3.3.3) to develop an equivalent system of equations in termsof the θi’s. Begin by multiplying both sides of (3.3.3) by the n× 3 matrix(∂~x∂θ )T :

m(x)(∂~x

∂θ)T ~x dV = (

∂~x

∂θ)T ~Fx dV. (3.3.4)

We examine (∂~x∂θ )T ~x and show how this may be integrated over the volumeoccupied by the entire mechanism to give a quantity on the left hand sideof (3.3.4) which depends only on the variables θ, θ, and θ. The followinglemma is useful in this connection.

Lemma 3.3.1.∂~x

∂θ=∂~x

∂θ

Proof.

∂~x

∂θi=

∂θi~x (3.3.5)

=∂

∂θi(∂~x

∂θ1θ1 + · · ·+ ∂~x

∂θnθn) (3.3.6)

=∂~x

∂θi. (3.3.7)

¥

3.3 Newtonian, Lagrangian, and Hamiltonian Mechanics 85

Now

(∂~x

∂θ)T ~x =

d

dt

(∂~x

∂θ

T

~x

)−[d

dt

(∂~x

∂θ

)T]~x.

By the lemma, this may be rewritten as

d

dt

(∂~x

∂θ

T

~x

)−[d

dt

(∂~x

∂θ

)T]~x. (3.3.8)

The first term may be further rewritten as

d

dt

∂θ

(12‖~x‖2

).

We next show that the second term may be written − ∂∂θ

(12‖~x‖2

). This

follows from the equality

d

dt

∂~x

∂θi=

n∑j=1

∂2~x

∂θj∂θiθj (3.3.9)

=∂

∂θi

n∑j=1

∂~x

∂θjθj (3.3.10)

=∂~x

∂θi. (3.3.11)

The left-hand side of (3.3.4) may thus be written as

m(~x)[d

dt

∂θ

(12‖~x‖2

)− ∂

∂θ

(12‖~x‖2

)]dV.

If this quantity is integrated over the spatial extent of the mechanism, weobtain

d

dt

∂T

∂θ− ∂T

∂θ, (3.3.12)

where T = T (θ, θ) = (1/2)∫m(~x)‖~x‖2 dV is the total kinetic energy of the

mechanism’s motion.We also wish to integrate the right hand side of (3.3.4) over the volume

occupied by the mechanism. Recall that (∂~x∂θ )T ~Fx is the component of gen-eralized force acting in the direction of the generalized coordinate θ whichis due to the force ~Fx acting at ~x = ~x(θ). Hence, integrating (3.3.4) overthe mechanism results in the total net generalized force being applied inthe direction θ. Call this quantity ~Fθ. We summarize these calculations by

86 3. Basic concepts in mechanics

writing the dynamic equations for the mechanism in terms of the variablesθ:

d

dt

∂T

∂θ− ∂T

∂θ= ~Fθ. (3.3.13)

We now set out some of the notation and terminology used in subse-quent chapters. The reader is referred to one of the standard books, suchas Abraham and Marsden [1978], Arnold [1989], Guillemin and Sternberg[1984] and Marsden and Ratiu [1992] for proofs omitted here.

3.4 Symplectic and Poisson Manifolds

Definition 3.4.1. Let P be a manifold and let F(P ) denote the set ofsmooth real-valued functions on P . Consider a given bracket operation de-noted

, : F(P )×F(P )→ F(P ).

The pair (P, , ) is called a Poisson manifold if , satisfies

(PB1) bilinearity f, g is bilinear in f and g.(PB2) anticommutativity f, g = −g, f.(PB3) Jacobi’s identity f, g, h+ h, f, g+ g, h, f = 0.(PB4) Leibniz’ rule fg, h = fg, h+ gf, h.Conditions (PB1)–(PB3) make (F(P ), , ) into a Lie algebra. If (P, , )

is a Poisson manifold, then because of (PB1) and (PB4), there is a tensorB on P , assigning to each z ∈ P a linear map B(z) : T ∗z P → TzP suchthat

f, g(z) = 〈B(z) · df(z),dg(z)〉. (3.4.1)

Here, 〈 , 〉 denotes the natural pairing between vectors and covectors. Be-cause of (PB2), B(z) is antisymmetric. Letting zI , I = 1, . . . ,M denotecoordinates on P , (3.4.1) becomes

f, g = BIJ∂f

∂zI∂g

∂zJ. (3.4.2)

Antisymmetry means BIJ = −BJI and Jacobi’s identity reads

BLI∂BJK

∂zL+BLJ

∂BKI

∂zL+BLK

∂BIJ

∂zL= 0. (3.4.3)

Definition 3.4.2. Let (P1, , 1) and (P2, , 2) be Poisson manifolds. Amapping ϕ : P1 → P2 is called Poisson if for all f, h ∈ F(P2), we have

f, h2 ϕ = f ϕ, h ϕ1. (3.4.4)

3.4 Symplectic and Poisson Manifolds 87

Definition 3.4.3. Let P be a manifold and Ω a 2-form on P . The pair(P,Ω) is called a symplectic manifold if Ω satisfies

(S1) dΩ = 0 (i.e., Ω is closed) and

(S2) Ω is nondegenerate.

Definition 3.4.4. Let (P,Ω) be a symplectic manifold and let f ∈ F(P ).Let Xf be the unique vector field on P satisfying

Ωz(Xf (z), v) = df(z) · v for all v ∈ TzP. (3.4.5)

We call Xf the Hamiltonian vector field of f . Hamilton’s equationsare the differential equations on P given by

z = Xf (z). (3.4.6)

If (P,Ω) is a symplectic manifold, define the Poisson bracket operation·, · : F(P )×F(P )→ F(P ) by

f, g = Ω(Xf , Xg). (3.4.7)

The construction (3.4.7) makes (P, , ) into a Poisson manifold. In otherwords,

Proposition 3.4.5. Every symplectic manifold is Poisson.

The converse is not true; for example the zero bracket makes any manifoldPoisson. In §2.4 we shall see some non-trivial examples of Poisson bracketsthat are not symplectic, such as Lie-Poisson structures on duals of Liealgebras.

Hamiltonian vector fields are defined on Poisson manifolds as follows.

Definition 3.4.6. Let (P, , ) be a Poisson manifold and let f ∈ F(P ).Define Xf to be the unique vector field on P satisfying

Xf [k] : = 〈dk,Xf 〉 = k, f for all k ∈ F(P ).

We call Xf the Hamiltonian vector field of f .

A check of the definitions shows that in the symplectic case, the Def-initions 2.1.4 and 2.1.6 of Hamiltonian vector fields coincide. If (P, , )is a Poisson manifold, there are therefore three equivalent ways to writeHamilton’s equations for H ∈ F(P ):

i z = XH(z)

ii f = df(z) ·XH(z) for all f ∈ F(P ), and

iii f = f,H for all f ∈ F(P ).

88 3. Basic concepts in mechanics

3.5 The Flow of a Hamiltonian Vector Field

Hamilton’s equations described in the abstract setting of the last sectionare very general. They include not only what one normally thinks of asHamilton’s canonical equations in classical mechanics, but Schrodinger’sequation in quantum mechanics as well. Despite this generality, the theoryhas a rich structure.

Let H ∈ F(P ) where P is a Poisson manifold. Let ϕt be the flow ofHamilton’s equations; thus, ϕt(z) is the integral curve of z = XH(z) start-ing at z. (If the flow is not complete, restrict attention to its domain ofdefinition.) There are two basic facts about Hamiltonian flows (ignoringfunctional analytic technicalities in the infinite dimensional case — seeChernoff and Marsden [1974]).

Proposition 3.5.1. The following hold for Hamiltonian systems on Pois-son manifolds:

i each ϕt is a Poisson map

ii H ϕt = H (conservation of energy).

The first part of this proposition is true even if H is a time dependentHamiltonian, while the second part is true only when H is independent oftime.

3.6 Cotangent Bundles

Let Q be a given manifold (usually the configuration space of a mechanicalsystem) and T ∗Q be its cotangent bundle. Coordinates qi on Q inducecoordinates (qi, pj) on T ∗Q, called the canonical cotangent coordinatesof T ∗Q.

Proposition 3.6.1. There is a unique 1-form Θ on T ∗Q such that inany choice of canonical cotangent coordinates,

Θ = pidqi; (3.6.1)

Θ is called the canonical 1-form . We define the canonical 2-form Ωby

Ω = −dΘ = dqi ∧ dpi (a sum on i is understood). (3.6.2)

In infinite dimensions, one needs to use an intrinsic definition of Θ, andthere are many such; one of these is the identity β∗Θ = β for β : Q→ T ∗Qany one form. Another is

Θ(wαq ) = 〈αq, TπQ · wαq 〉,

3.7 Lagrangian Mechanics 89

where αq ∈ T ∗qQ,wαq ∈ Tαq (T ∗Q) and where πQ : T ∗Q→ Q is the cotan-gent bundle projection.

Proposition 3.6.2. (T ∗Q,Ω) is a symplectic manifold.

In canonical coordinates the Poisson brackets on T ∗Q have the classicalform

f, g =∂f

∂qi∂g

∂pi− ∂g

∂qi∂f

∂pi, (3.6.3)

where summation on repeated indices is understood.

Theorem 3.6.3. (Darboux’ Theorem) Every symplectic manifold lo-cally looks like T ∗Q; in other words, on every finite dimensional symplecticmanifold, there are local coordinates in which Ω has the form (3.6.2).

(See Marsden [1981] and Olver [1988] for a discussion of the infinitedimensional case.)

Hamilton’s equations in these canonical coordinates have the classicalform

qi =∂H

∂pi

pi = −∂H∂qi

as one can readily check.The local structure of Poisson manifolds is more complex than the sym-

plectic case. However, Kirillov [1976a] has shown that every Poisson man-ifold is the union of symplectic leaves; to compute the bracket of twofunctions in P , one does it leaf-wise . In other words, to calculate thebracket of f and g at z ∈ P , select the symplectic leaf Sz through z, andevaluate the bracket of f |Sz and g|Sz at z. We shall see a specific case ofthis picture shortly.

3.7 Lagrangian Mechanics

Let Q be a manifold and TQ its tangent bundle. Coordinates qi on Q inducecoordinates (qi, qi) on TQ, called tangent coordinates. A mapping L :TQ→ R is called a Lagrangian . Often we choose L to be L = K−V whereK(v) = 1

2 〈v, v〉 is the kinetic energy associated to a given Riemannianmetric and where V : Q→ R is the potential energy .

Definition 3.7.1. Hamilton’s principle singles out particular curvesq(t) by the condition

δ

∫ a

b

L(q(t), q(t))dt = 0, (3.7.1)

90 3. Basic concepts in mechanics

where the variation is over smooth curves in Q with fixed endpoints.

It is interesting to note that (3.7.1) is unchanged if we replace the inte-grand by L(q, q)− d

dtS(q, t) for any function S(q, t). This reflects the gaugeinvariance of classical mechanics and is closely related to Hamilton-Jacobitheory. We shall return to this point in Chapter 9. It is also interesting tonote that if one keeps track of the boundary conditions in Hamilton’s prin-ciple, they essentially define the canonical one form, pidqi. This turns outto be a useful remark in more complex field theories.

If one prefers, the action principle states that the map I defined byI(q(·)) =

∫ baL(q(t), q(t))dt from the space of curves with prescribed end-

points in Q to R has a critical point at the curve in question. In any case, abasic and elementary result of the calculus of variations, whose proof wassketched in §1.2, is:

Proposition 3.7.2. The principle of critical action for a curve q(t) isequivalent to the condition that q(t) satisfies the Euler-Lagrange equa-tions

d

dt

∂L

∂qi− ∂L

∂qi= 0. (3.7.2)

Definition 3.7.3. Let L be a Lagrangian on TQ and let FL : TQ→ T ∗Qbe defined (in coordinates) by

(qi, qj) 7→ (qi, pj)

where pj = ∂L/∂qj. We call FL the fiber derivative. (Intrinsically, FLdifferentiates L in the fiber direction.)

A Lagrangian L is called hyperregular if FL is a diffeomorphism. If Lis a hyperregular Lagrangian, we define the corresponding Hamiltonianby

H(qi, pj) = piqi − L.

The change of data from L on TQ to H on T ∗Q is called the Legendretransform .

One checks that the Euler-Lagrange equations for L are equivalent toHamilton’s equations for H.

In a relativistic context one finds that the two conditions pj = ∂L/∂qj

and H = piqi − L, defining the Legendre transform, fit together as the

spatial and temporal components of a single object. Suffice it to say thatthe formalism developed here is useful in the context of relativistic fields.

3.8 Lie-Poisson Structures and the Rigid Body 91

3.8 Lie-Poisson Structures and the RigidBody

Not every Poisson manifold is symplectic. For example, a large class of non-symplectic Poisson manifolds is the class of Lie-Poisson manifolds, whichwe now define. Let G be a Lie group and g = TeG its Lie algebra with[ , ] : g× g→ g the associated Lie bracket.

Proposition 3.8.1. The dual space g∗ is a Poisson manifold with eitherof the two brackets

f, k±(µ) = ±⟨µ,

[δf

δµ,δk

δµ

]⟩. (3.8.1)

Here g is identified with g∗∗ in the sense that δf/δµ ∈ g is defined by〈ν, δf/δµ〉 = Df(µ) · ν for ν ∈ g∗, where D denotes the derivative. (In theinfinite dimensional case one needs to worry about the existence of δf/δµ;in this context, methods like the Hahn-Banach theorem are not alwaysappropriate!) The notation δf/δµ is used to conform to the functionalderivative notation in classical field theory. In coordinates, (ξ1, . . . , ξm) ong and corresponding dual coordinates (µ1, . . . , µm) on g∗, the Lie-Poissonbracket (3.8.1) is

f, k±(µ) = ±µaCabc∂f

∂µb

∂k

∂µc; (3.8.2)

here Cabc are the structure constants of g defined by [ea, eb] = Ccabec,where (e1, . . . , em) is the coordinate basis of g and where, for ξ ∈ g, wewrite ξ = ξaea, and for µ ∈ g∗, µ = µae

a, where (e1, . . . , em) is the dualbasis. Formula (3.8.2) appears explicitly in Lie [1890], §75.

Which sign to take in (3.8.2) is determined by understanding Lie-Poissonreduction , which can be summarized as follows. Let

λ : T ∗G→ g∗ be defined by pg 7→ (TeLg)∗pg ∈ T ∗eG ∼= g∗ (3.8.3)

and

ρ : T ∗G→ g∗ be defined by pg 7→ (TeRg)∗pg ∈ T ∗eG ∼= g∗. (3.8.4)

Then λ is a Poisson map if one takes the − Lie-Poisson structure on g∗

and ρ is a Poisson map if one takes the + Lie-Poisson structure on g∗.Every left invariant Hamiltonian and Hamiltonian vector field is mapped

by λ to a Hamiltonian and Hamiltonian vector field on g∗. There is a similarstatement for right invariant systems on T ∗G. One says that the originalsystem on T ∗G has been reduced to g∗. The reason λ and ρ are bothPoisson maps is perhaps best understood by observing that they are bothequivariant momentum maps generated by the action of G on itself by rightand left translations, respectively. We take up this topic in §2.7.

92 3. Basic concepts in mechanics

We saw in Chapter 1 that the Euler equations of motion for rigid bodydynamics are given by

Π = Π× Ω, (3.8.5)

where Π = IΩ is the body angular momentum and Ω is the body angularvelocity. Euler’s equations are Hamiltonian relative to a Lie-Poisson struc-ture. To see this, take G = SO(3) to be the configuration space. Theng ∼= (R3,×) and we identify g ∼= g∗. The corresponding Lie-Poisson struc-ture on R3 is given by

f, k(Π) = −Π · (∇f ×∇k). (3.8.6)

For the rigid body one chooses the minus sign in the Lie-Poisson bracket.This is because the rigid body Lagrangian (and hence Hamiltonian) is leftinvariant and so its dynamics pushes to g∗ by the map λ in (3.8.3).

Starting with the kinetic energy Hamiltonian derived in Chapter 1, wedirectly obtain the formula H(Π) = 1

2Π · (I−1Π), the kinetic energy of therigid body. One verifies from the chain rule and properties of the tripleproduct that:

Proposition 3.8.2. Euler’s equations are equivalent to the following equa-tion for all f ∈ F(R3):

f = f,H. (3.8.7)

Definition 3.8.3. Let (P, , ) be a Poisson manifold. A function C ∈F(P ) satisfying

C, f = 0 for all f ∈ F(P ) (3.8.8)

is called a Casimir function .

A crucial difference between symplectic manifolds and Poisson manifoldsis this: On symplectic manifolds, the only Casimir functions are the con-stant functions (assuming P is connected). On the other hand, on Poissonmanifolds there is often a large supply of Casimir functions. In the case ofthe rigid body, every function C : R3 → R of the form

C(Π) = Φ(‖Π‖2) (3.8.9)

where Φ : R → R is a differentiable function, is a Casimir function, as wenoted in Chapter 1. Casimir functions are constants of the motion for anyHamiltonian since C = C,H = 0 for any H. In particular, for the rigidbody, ‖Π‖2 is a constant of the motion — this is the invariant sphere wesaw in Chapter 1.

There is an intimate relation between Casimirs and symmetry generatedconserved quantities, or momentum maps , which we study in §2.7.

3.8 Lie-Poisson Structures and the Rigid Body 93

The maps λ and ρ induce Poisson isomorphisms between (T ∗G)/G andg∗ (with the − and + brackets respectively) and this is a special instanceof Poisson reduction, as we will see in §2.8. The following result is oneuseful way of formulating the general relation between T ∗G and g∗. Wetreat the left invariant case to be specific. Of course, the right invariantcase is similar.

Theorem 3.8.4. Let G be a Lie group and H : T ∗G → R be a leftinvariant Hamiltonian. Let h : g∗ → R be the restriction of H to theidentity. For a curve p(t) ∈ T ∗g(t)G, let µ(t) = (T ∗g(t)L) · p(t) = λ(p(t))be the induced curve in g∗. Assume that g = ∂H/∂p ∈ TgG. Then thefollowing are equivalent:

i p(t) is an integral curve of XH ; i.e., Hamilton’s equations on T ∗Ghold,

ii for any F ∈ F(T ∗G), F = F,H, where , is the canonical bracketon T ∗G

iii µ(t) satisfies the Lie-Poisson equations

dt= ad∗δh/δµµ (3.8.10)

where adξ : g→ g is defined by adξη = [ξ, η] and ad∗ξ is its dual, i.e.,

µa = Cdbaδh

δµbµd (3.8.11)

iv for any f ∈ F(g∗), we have

f = f, h− (3.8.12)

where , − is the minus Lie-Poisson bracket.

We now make some remarks about the proof. First of all, the equivalenceof i and ii is general for any cotangent bundle, as we have already noted.Next, the equivalence of ii and iv follows directly from the fact that λ is aPoisson map (as we have mentioned, this follows from the fact that λ is amomentum map; see Proposition 2.7.6 below) and H = h λ. Finally, weestablish the equivalence of iii and iv. Indeed, f = f, h− means⟨

µ,δf

δµ

⟩= −

⟨µ,

[δf

δµ,δh

δµ

]⟩=⟨µ, adδh/δµ

δf

δµ

⟩=⟨

ad∗δh/δµµ,δf

δµ

⟩.

Since f is arbitrary, this is equivalent to iii. ¥

94 3. Basic concepts in mechanics

3.9 The Euler-Poincare Equations

In §1.3 we saw that for the rigid body, there is an analogue of the abovetheorem on SO(3) and so(3) using the Euler-Lagrange equations and thevariational principle as a starting point. We now generalize this to an ar-bitrary Lie group and make the direct link with the Lie-Poisson equations.

Theorem 3.9.1. Let G be a Lie group and L : TG→ R a left invariantLagrangian. Let l : g → R be its restriction to the identity. For a curveg(t) ∈ G, let

ξ(t) = g(t)−1 · g(t); i.e., ξ(t) = Tg(t)Lg(t)−1 g(t).

Then the following are equivalent

i g(t) satisfies the Euler-Lagrange equations for L on G,

ii the variational principle

δ

∫L(g(t), g(t))dt = 0 (3.9.1)

holds, for variations with fixed endpoints,

iii the Euler-Poincare equations hold:

d

dt

δl

δξ= ad∗ξ

δl

δξ, (3.9.2)

iv the variational principle

δ

∫l(ξ(t))dt = 0 (3.9.3)

holds on g, using variations of the form

δξ = η + [ξ, η], (3.9.4)

where η vanishes at the endpoints.

Let us discuss the main ideas of the proof. First of all, the equivalenceof i and ii holds on the tangent bundle of any configuration manifold Q, aswe have seen, secondly, ii and iv are equivalent. To see this, one needs tocompute the variations δξ induced on ξ = g−1g = TLg−1 g by a variationof g. To calculate this, we need to differentiate g−1g in the direction of avariation δg. If δg = dg/dε at ε = 0, where g is extended to a curve gε,then,

δξ =d

(g−1 d

dtg

)∣∣∣∣ε=0

3.9 The Euler-Poincare Equations 95

while if η = g−1δg, then

η =d

dt

(g−1 d

dεg

)∣∣∣∣ε=0

.

The difference δξ − η is the commutator, [ξ, η]. This argument is fine formatrix groups, but takes a little more work to make precise for generalLie groups. See Bloch, Krishnaprasad, Ratiu and Marsden [1994b] for thegeneral case. Thus, ii and iv are equivalent.

To complete the proof, we show the equivalence of iii and iv. Indeed,using the definitions and integrating by parts,

δ

∫l(ξ)dt =

∫δl

δξδξ dt

=∫

δl

δξ(η + adξη)dt

=∫ [− d

dt

(δl

δξ

)+ ad∗ξ

δl

δξ

]η dt

so the result follows. ¥

Generalizing what we saw directly in the rigid body, one can check di-rectly from the Euler- Poincare equations that conservation of spatial an-gular momentum holds:

d

dtπ = 0 (3.9.5)

where π is defined by

π = Ad∗g∂l

∂ξ. (3.9.6)

Since the Euler-Lagrange and Hamilton equations on TQ and T ∗Q areequivalent, it follows that the Lie-Poisson and Euler-Poincare equationsare also equivalent. To see this directly, we make the following Legendretransformation from g to g∗:

µ =δl

δξ, h(µ) = 〈µ, ξ〉 − l(ξ).

Note that

δh

δµ= ξ +

⟨µ,δξ

δµ

⟩−⟨δl

δξ,δξ

δµ

⟩= ξ

and so it is now clear that (2.5.10) and (2.6.2) are equivalent.

96 3. Basic concepts in mechanics

3.10 Momentum Maps

Let G be a Lie group and P be a Poisson manifold, such that G acts onP by Poisson maps (in this case the action is called a Poisson action).Denote the corresponding infinitesimal action of g on P by ξ 7→ ξP , a mapof g to X(P ), the space of vector fields on P . We write the action of g ∈ Gon z ∈ P as simply gz; the vector field ξP is obtained at z by differentiatinggz with respect to g in the direction ξ at g = e. Explicitly,

ξP (z) =d

dε[exp(εξ) · z]

∣∣∣∣ε=0

.

Definition 3.10.1. A map J : P → g∗ is called a momentum map ifX〈J,ξ〉 = ξP for each ξ ∈ g, where 〈J, ξ〉(z) = 〈J(z), ξ〉.Theorem 3.10.2. (Noether’s Theorem) If H is a G invariant Hamil-tonian on P , then J is conserved on the trajectories of the Hamiltonianvector field XH .

Proof Differentiating the invariance condition H(gz) = H(z) with respectto g ∈ G for fixed z ∈ P , we get dH(z) · ξP (z) = 0 and so H, 〈J, ξ〉 = 0which by antisymmetry gives d〈J, ξ〉 ·XH = 0 and so 〈J, ξ〉 is conserved onthe trajectories of XH for every ξ in G. ¥

Turning to the construction of momentum maps, let Q be a manifold andlet G act on Q. This action induces an action of G on T ∗Q by cotangentlift — that is, we take the transpose inverse of the tangent lift. The actionof G on T ∗Q is always symplectic and therefore Poisson.

Theorem 3.10.3. A momentum map for a cotangent lifted action isgiven by

J : T ∗Q→ g∗ defined by 〈J, ξ〉(pq) = 〈pq, ξQ(q)〉. (3.10.1)

In canonical coordinates, we write pq = (qi, pj) and define the actionfunctions Ki

a by (ξQ)i = Kia(q)ξa. Then

〈J, ξ〉(pq) = piKia(q)ξa (3.10.2)

and therefore

Ja = piKia(q). (3.10.3)

Recall that by differentiating the conjugation operation h 7→ ghg−1 at theidentity, one gets the adjoint action of G on g. Taking its dual producesthe coadjoint action of G on g∗.

Proposition 3.10.4. The momentum map for cotangent lifted actions isequivariant , i.e., the diagram in Figure 2.7.1 commutes.

3.10 Momentum Maps 97

T ∗Q g∗

T ∗Q g∗

J

J

G-actionon T ∗Q

coadjointaction

-

-? ?

Figure 3.10.1. Equivariance of the momentum map.

Proposition 3.10.5. Equivariance implies infinitesimal equivariance, whichcan be stated as the classical commutation relations:

〈J, ξ〉, 〈J, η〉 = 〈J, [ξ, η]〉.

Proposition 3.10.6. If J is infinitesimally equivariant, then J : P → g∗

is a Poisson map. If J is generated by a left (respectively right) action thenwe use the + (respectively −) Lie-Poisson structure on g∗.

The above development concerns momentum maps using the Hamilto-nian point of view. However, one can also consider them from the La-grangian point of view. In this context, we consider a Lie group G actingon a configuration manifold Q and lift this action to the tangent bundle TQusing the tangent operation. Given a G-invariant Lagrangian L : TQ→ R,the corresponding momentum map is obtained by replacing the momentumpq in (3.10.1) with the fiber derivative FL(vq). Thus, J : TQ→ g∗ is givenby

〈J(vq), ξ〉 = 〈FL(vq), ξQ(q)〉 (3.10.4)

or, in coordinates,

Ja =∂L

∂qiKia, (3.10.5)

where the action coefficients Kia are defined as before by writing ξQ(qi) =

Kiaξa∂/∂qi.

Proposition 3.10.7. For a solution of the Euler-Lagrange equations (evenif the Lagrangian is degenerate), J is constant in time.

Proof In case L is a regular Lagrangian, this follows from its Hamilto-nian counterpart. It is useful to check it directly using Hamilton’s principle(which is the way it was originally done by Noether). To do this, choose anyfunction φ(t, s) of two variables such that the conditions φ(a, s) = φ(b, s) =φ(t, 0) = 0 hold. Since L is G-invariant, for each Lie algebra element ξ ∈ g,

98 3. Basic concepts in mechanics

the expression ∫ b

a

L(exp(φ(t, s)ξ)q, exp(φ(t, s)ξ)q)) dt (3.10.6)

is independent of s. Differentiating this expression with respect to s ats = 0 and setting φ′ = ∂φ/∂s taken at s = 0, gives

0 =∫ b

a

(∂L

∂qiξiQφ

′ +∂L

∂qi(TξQ · q)iφ′

)dt. (3.10.7)

Now we consider the variation q(t, s) = exp(φ(t, s)ξ) ·q(t). The correspond-ing infinitesimal variation is given by

δq(t) = φ′(t)ξQ(q(t)).

By Hamilton’s principle, we have

0 =∫ b

a

(∂L

∂qiδqi +

∂L

∂qiδqi)dt. (3.10.8)

Note that

δq = φ′ξQ + φ′(TξQ · q)

and subtract (3.10.8) from (3.10.7) to give

0 =∫ b

a

∂L

∂qi(ξQ)iφ′ dt (3.10.9)

=∫ b

a

d

dt

(∂L

∂qiξiQ

)φ′ dt. (3.10.10)

Since φ′ is arbitrary, except for endpoint conditions, it follows that theintegrand vanishes, and so the time derivative of the momentum map iszero and so the proposition is proved. ¥

3.11 Symplectic and Poisson Reduction

We have already seen how to use variational principles to reduce the Euler-Lagrange equations. On the Hamiltonian side, there are three levels ofreduction of decreasing generality, that of Poisson reduction, symplecticreduction, and cotangent bundle reduction. Let us first consider Poissonreduction.

For Poisson reduction we start with a Poisson manifold P and letthe Lie group G act on P by Poisson maps. Assuming P/G is a smooth

3.11 Symplectic and Poisson Reduction 99

manifold, endow it with the unique Poisson structure on such that thecanonical projection π : P → P/G is a Poisson map. We can specify thePoisson structure on P/G explicitly as follows. For f and k : P/G→ R, letF = f π and K = kπ, so F and K are f and k thought of as G-invariantfunctions on P . Then f, kP/G is defined by

f, kP/G π = F,KP . (3.11.1)

To show that f, kP/G is well defined, one has to prove that F,KP isG-invariant. This follows from the fact that F and K are G-invariant andthe group action of G on P consists of Poisson maps.

For P = T ∗G we get a very important special case.

Theorem 3.11.1. (Lie-Poisson Reduction) Let P = T ∗G and as-sume that G acts on P by the cotangent lift of left translations. If oneendows g∗ with the minus Lie-Poisson bracket, then P/G ∼= g∗.

For symplectic reduction we begin with a symplectic manifold (P,Ω).Let G be a Lie group acting by symplectic maps on P ; in this case theaction is called a symplectic action . Let J be an equivariant momentummap for this action and H a G-invariant Hamiltonian on P . Let Gµ = g ∈G | g · µ = µ be the isotropy subgroup (symmetry subgroup) at µ ∈ g∗.As a consequence of equivariance, Gµ leaves J−1(µ) invariant. Assume forsimplicity that µ is a regular value of J, so that J−1(µ) is a smooth manifold(see §2.8 below) and that Gµ acts freely and properly on J−1(µ), so thatJ−1(µ)/Gµ =: Pµ is a smooth manifold. Let iµ : J−1(µ) → P denote theinclusion map and let πµ : J−1(µ)→ Pµ denote the projection. Note that

dimPµ = dimP − dimG− dimGµ. (3.11.2)

Building on classical work of Jacobi, Liouville, Arnold and Smale, we havethe following basic result of Meyer [1973] and Marsden and Weinstein[1974].

Theorem 3.11.2. (Reduction Theorem) There is a unique symplecticstructure Ωµ on Pµ satisfying

i∗µΩ = π∗µΩµ. (3.11.3)

Given a G-invariant Hamiltonian H on P , define the reduced Hamilto-nian Hµ : Pµ → R by H = Hµ πµ. Then the trajectories of XH projectto those of XHµ . An important problem is how to reconstruct trajectoriesof XH from trajectories of XHµ . Schematically, we have the situation inFigure 2.8.1.

As we shall see later, the reconstruction process is where the holonomyand “geometric phase” ideas enter. In fact, we shall put a connection on thebundle πµ : J−1(µ)→ Pµ and it is through this process that one encountersthe gauge theory point of view of mechanics.

100 3. Basic concepts in mechanics

reduction reconstruction

? ?

6 6

P

J−1(µ)

πµ

Figure 3.11.1. Reduction to Pµ and reconstruction back to P .

Let Oµ denote the coadjoint orbit through µ. As a special case of thesymplectic reduction theorem, we get

Corollary 3.11.3. (T ∗G)µ ∼= Oµ.

The symplectic structure inherited on Oµ is called the (Lie-Kostant-Kirillov) orbit symplectic structure . This structure is compatible withthe Lie-Poisson structure on g∗ in the sense that the bracket of two func-tions on Oµ equals that obtained by extending them arbitrarily to g∗,taking the Lie-Poisson bracket on g∗ and then restricting to Oµ.

Example 1 G = SO(3), g∗ = so(3)∗ ∼= R3. In this case the coadjoint actionis the usual action of SO(3) on R3. This is because of the orthogonality ofthe elements of G. The set of orbits consists of spheres and a single point.The reduction process confirms that all orbits are symplectic manifolds.One calculates that the symplectic structure on the spheres is a multipleof the area element. ¨

Example 2 Jacobi-Liouville theorem Let G = Tk be the k-torus andassume G acts on a symplectic manifold P . In this case the componentsof J are in involution and dimPµ = dimP − 2k, so 2k variables are elimi-nated. As we shall see, reconstruction allows one to reassemble the solutiontrajectories on P by quadratures in this abelian case. ¨

Example 3 Jacobi-Deprit elimination of the node Let G = SO(3)act on P . In the classical case of Jacobi, P = T ∗R3 and in the generalizationof Deprit [1983] one considers the phase space of n particles in R3. We justpoint out here that the reduced space Pµ has dimension dimP − 3 − 1 =dimP − 4 since Gµ = S1 (if µ 6= 0) in this case. ¨

The orbit reduction theorem of Marle [1976] and Kazhdan, Kostantand Sternberg [1978] states that Pµ may be alternatively constructed as

PO = J−1(O)/G, (3.11.4)

3.12 Singularities and Symmetry 101

where O ⊂ g∗ is the coadjoint orbit through µ. As above we assume weare away from singular points (see §2.8 below). The spaces Pµ and PO areisomorphic by using the inclusion map lµ : J−1(µ) → J−1(O) and takingequivalence classes to induce a symplectic isomorphism Lµ : Pµ → PO. Thesymplectic structure ΩO on PO is uniquely determined by

j∗OΩ = π∗OΩO + J∗OωO (3.11.5)

where jO : J−1(O)→ P is the inclusion, πO : J−1(O)→ PO is the projec-tion, and where JO = J|J−1(O) : J−1(O) → O and ωO is the orbit sym-plectic form. In terms of the Poisson structure, J−1(O)/G has the bracketstructure inherited from P/G; in fact, J−1(O)/G is a symplectic leaf inP/G. Thus, we get the picture in Figure 2.8.2.

/Gµ /G /G

? ? ?

J−1(µ) ⊂ J−1(O) ⊂ P

Pµ ∼= PO ⊂ P/G

Figure 3.11.2. Orbit reduction gives another realization of Pµ.

Kirillov has shown that every Poisson manifold P is the union of sym-plectic leaves, although the preceding construction explicitly realizes thesesymplectic leaves in this case by the reduction construction. A special caseis the foliation of the dual g∗ of any Lie algebra g into its symplectic leaves,namely the coadjoint orbits. For example SO(3) is the union of spheres plusthe origin, each of which is a symplectic manifold. Notice that the drop indimension from T ∗SO(3) to O is from 6 to 2, a drop of 4, as in generalSO(3) reduction. An exception is the singular point, the origin, where thedrop in dimension is larger. We turn to these singular points next.

3.12 Singularities and Symmetry

Proposition 3.12.1. Let (P,Ω) be a symplectic manifold, let G act onP by Poisson mappings, and let J : P → g∗ be a momentum map for thisaction (J need not be equivariant). Let Gz denote the symmetry group ofz ∈ P defined by Gz = g ∈ G | gz = z and let gz be its Lie algebra, sogz = ζ ∈ g | ζP (z) = 0. Then z is a regular value of J if and only if gz

is trivial; i.e., gz = 0, or Gz is discrete.

102 3. Basic concepts in mechanics

Proof The point z is regular when the range of the linear map DJ(z) isall of g∗. However, ζ ∈ g is orthogonal to the range (in the sense of theg, g∗ pairing) if and only if for all v ∈ TzP ,

〈ζ,DJ(z) · v〉 = 0

i.e.,

d〈J, ζ〉(z) · v = 0

or

Ω(X〈J,ζ〉(z), v) = 0

or

Ω(ζP (z), v) = 0.

As Ω is nondegenerate, ζ is orthogonal to the range iff ζP (z) = 0. ¥

The above proposition is due to Smale [1970]. It is the starting point ofa large literature on singularities in the momentum map and singular re-duction. Arms, Marsden and Moncrief [1981] show, under some reasonablehypotheses, that the level sets J−1(0) have quadratic singularities. As weshall see in the next chapter, there is a general shifting construction thatenables one to effectively reduce J−1(µ) to the case J−1(0). In the finitedimensional case, this result can be deduced from the equivariant Darbouxtheorem, but in the infinite dimensional case, things are much more subtle.In fact, the infinite dimensional results were motivated by, and apply to,the singularities in the solution space of relativistic field theories such asgravity and the Yang-Mills equations (see Fischer, Marsden and Moncrief[1980], Arms, Marsden and Moncrief [1981, 1982] and Arms [1981]). Theconvexity theorem states that the image of the momentum map of atorus action is a convex polyhedron in g∗; the boundary of the polyhedronis the image of the singular (symmetric) points in P ; the more symmetricthe point, the more singular the boundary point. These results are due toAtiyah [1982] and Guillemin and Sternberg [1984] based on earlier convexityresults of Kostant and the Shur-Horne theorem on eigenvalues of symmetricmatrices. The literature on these topics and its relation to other areas ofmathematics is vast. See, for example, Goldman and Millison [1990], Sja-maar [1990], Bloch, Flaschka and Ratiu [1990], Sjamaar and Lerman [1991]and Lu and Ratiu [1991].

3.13 A Particle in a Magnetic Field

During cotangent bundle reduction considered in the next chapter, we shallhave to add terms to the symplectic form called “magnetic terms”. Toexplain this terminology, we consider a particle in a magnetic field.

3.13 A Particle in a Magnetic Field 103

Let B be a closed two-form on R3 and B = Bxi+Byj+Bzk the associateddivergence free vector field, i.e., iB(dx ∧ dy ∧ dz) = B, or

B = Bxdy ∧ dz −Bydx ∧ dz +Bzdx ∧ dy.

Thinking of B as a magnetic field, the equations of motion for a particlewith charge e and mass m are given by the Lorentz force law :

mdvdt

=e

cv ×B (3.13.1)

where v = (x, y, z). On R3×R3 i.e., on (x,v)-space, consider the symplecticform

ΩB = m(dx ∧ dx+ dy ∧ dy + dz ∧ dz)− e

cB. (3.13.2)

For the Hamiltonian, take the kinetic energy:

H =m

2(x2 + y2 + z2) (3.13.3)

writingXH(u, v, w) = (u, v, w, (u, v, w)), the condition definingXH , namelyiXHΩB = dH is

m(udx− udx+ vdy − vdy + wdz − wdz)− e

c[Bxvdz −Bxwdy −Byudz +Bywdx+Bzudy −Bzvdx]

= m(xdx+ ydy + zdz) (3.13.4)

which is equivalent to u = x, v = y, w = z,mu = e(Bzv − Byw)/c,mv =e(Bxw −Bzu)/c, and mw = e(Byu−Bxv)/c, i.e., to

mx =e

c(Bz y −By z)

my =e

c(Bxz −Bzx) (3.13.5)

mz =e

c(Byx−Bxy)

which is the same as (3.13.1). Thus the equations of motion for a particlein a magnetic field are Hamiltonian, with energy equal to the kinetic energyand with the symplectic form ΩB .

If B = dA; i.e., B = −∇ × A, where A is a one-form and A is theassociated vector field, then the map (x,v) 7→ (x,p) where p = mv +eA/c pulls back the canonical form to ΩB , as is easily checked. Thus,Equations (3.13.1) are also Hamiltonian relative to the canonical bracketon (x,p)-space with the Hamiltonian

HA =1

2m‖p− e

cA‖2. (3.13.6)

104 3. Basic concepts in mechanics

Even in Euclidean space, not every magnetic field can be written asB = ∇ × A. For example, the field of a magnetic monopole of strengthg 6= 0, namely

B(r) = gr‖r‖3 (3.13.7)

cannot be written this way since the flux of B through the unit sphereis 4πg, yet Stokes’ theorem applied to the two hemispheres would givezero. Thus, one might think that the Hamiltonian formulation involvingonly B (i.e., using ΩB and H) is preferable. However, one can recoverthe magnetic potential A by regarding A as a connection on a nontrivialbundle over R3\0. The bundle over the sphere S2 is in fact the same Hopffibration S3 → S2 that we encountered in §1.6. This same constructioncan be carried out using reduction. For a readable account of some aspectsof this situation, see Yang [1980]. For an interesting example of Weinsteinin which this monopole comes up, see Marsden [1981], p. 34.

When one studies the motion of a colored (rather than a charged) particlein a Yang-Mills field, one finds a beautiful generalization of this construc-tion and related ideas using the theory of principal bundles; see Sternberg[1977], Weinstein [1978] and Montgomery [1985]. In the study of centrifugaland Coriolis forces one discovers some structures analogous to those here(see Marsden and Ratiu [1999] for more information).

3.14 The Mechanical Connection

An important connection for analyzing the dynamics and control of me-chanical systems is the so-called mechanical connection – see e.g. Marsdenand Scheurle [1993], Marsden [1992].

We assume we have a configuration manifold Q and Lagrangian L :TQ → R. Let G be a Lie group with Lie algebra g, assume G acts on Q,and lift the action to TQ via the tangent mapping. We assume that G actsfreely and properly on Q. We assume also we have a metric 〈〈 , 〉〉 on Q thatis invariant under the group action.

For each q ∈ Q define the the locked inertia tensor to be the mapR : g→ g∗ defined by

〈Iη, ζ〉 = 〈〈ηQ(q), ζQ(q)〉〉 . (3.14.1)

Here ηQ denotes as before the infinitesimal generator of the action of Gon Q.

Locally, if

[ξQ(q)]i = Kia(q)ξa (3.14.2)

3.14 The Mechanical Connection 105

relative to the coordinates qi of Q and a basis ea, a = 1, · · ·m of g (Kia

are called the action coefficients), then

Iab = gijKiaK

jb (3.14.3)

Now the momentum map in this context, J : TM → g∗, is defined by

〈J(q, v), ξ〉 = 〈〈ξQ(q), v〉〉 , (3.14.4)

Then

Definition 3.14.1. We define the mechanical connection on the prin-cipal bundle q → Q/G to be the map A : TQ→ g given by

A(q, v) = I(q)−1 (J(q, v)) , (3.14.5)

that is, A is the map that assigns to each (q, v) the corresponding angularvelocity of the locked system.

In coordinates

Aa = IabgijKibvj . (3.14.6)

One can check that is A is G-equivariant and (ξQ(q)) = ξ.The horizontal space of the connection is given by

horq = (q, v)|J = 0 ⊂ TqQ . (3.14.7)

The vertical space of vectors that are mapped to zero under the projec-tion Q→ S = Q/G is given by

verq = ξQ(q)|ξ ∈ g . (3.14.8)

The horizontal-vertical decomposition of a vector (q, v) ∈ TqQ is givenby

v = horqv + verqv (3.14.9)

where

verqv = [A(q, v)]Q(q) and horqv = v − verqv . (3.14.10)

3.14.1 Example – the nonlinear pendulum

As example for illustrating the mechanical connection we recall the invertedpendulum on a cart. We consider here the uncontrolled case.

Recall the the Lagrangian may be written

L =12

(αθ2 − 2β cos θθs+ γs2 +D cos θ

). (3.14.11)

106 3. Basic concepts in mechanics

In this case the Lagrangian is cyclic in the variable s or invariant underthe linear R1 action s→ s+ a.

The infintesimal generator of this action is thus given by

ξQ =d

dt|t=0 (s+ at, θ) = (a, 0) . (3.14.12)

In this case using the mechanical metric induced by the Lagrangian wehave

J(q, v) = 〈J(q, v), a〉 = 〈FL(q, v), (0, a)〉 (3.14.13)

and hence

J =∂L

∂s= γs− β cos θθ . (3.14.14)

The locked inertia tensor I(q) is given here by

〈I(q)a, a〉 = 〈〈(0, a), (0, a)〉〉 (3.14.15)

and hence

I(q) =[

0 1] [ α −β cos θ−β cos θ γ

] [01

]= γ . (3.14.16)

Hence the action of the mechanical connection on a tangent vector v atthe point q is given by

A(q, v) =1γ

(γs− β cos θφ

)=(s− β

γcos θφ

)(3.14.17)

The vertical horizontal decomposition of a vector v is given by

verqv = [A(q, v)]Q(q) =(

0, s− β

γcos θφ

)horqv = v − verqv =

(φ,β

γcosφφ

)(3.14.18)

Note that the horizontal component of v is clearly zero under the actionof A.

This is page 107Printer: Opaque this

4An Introduction to Geometric ControlTheory

4.1 Nonlinear Control Systems

There are many texts on linear control theory, and a number of introduc-tions to nonlinear control theory and in particular its differential geometricformulation which is important for this book. We mention the books byJurdjevic [1997], Isidori [1995], Nijmeijer and van der Schaft [1990], Sontag[1990] and Brockett [1972].1 The first three deal with the differential ge-ometric approach, Sontag’s book gives a mathematical treatment of bothlinear theory and various aspects of nonlinear theory, whil Brockett’s bookgives an approach to linear theory. We refer the reader to these books for adetailed treatment of nonlinear control theory as well a for a more exhaus-tive list of references. There are of course many other good sources as welland we shall refer to them as needed.

1Isidori, Alberto Nonlinear control systems. Third edition. Communications and Con-trol Engineering Series. Springer-Verlag, Berlin, 1995; Nijmeijer, Henk; van der Schaft,Arjan Nonlinear dynamical control systems. Springer-Verlag, New York, 1990; Son-tag, Eduardo D. Mathematical control theory. Deterministic finite-dimensional systems.Texts in Applied Mathematics, 6. Springer-Verlag, New York, 1990; Brockett, R. [1970]Finite Dimensional Linear Systems. Wiley; Jurdjevic, Velimir Geometric control theory.Cambridge Studies in Advanced Mathematics, 52. Cambridge University Press, Cam-bridge, 1997.

108 4. An Introduction to Geometric Control Theory

Definition 4.1.1. A finite dimensional nonlinear control systemon a smooth n-manifold is a differential equation of the form

x = f(x, u) (4.1.1)

where x = (x1, · · · , xn) are local coordinates on M , u(t) is a time dependentmap from the nonnegative reals, R1+ to Rm, and f and u lie in appropriateclass of functions. Usually f is taken to be C∞ (smooth) or Cω (analytic),but less often suffices. u may be smooth or continuous, but may be also bediscontinuous and even just measurable. M is said to be the state spaceof the system.

It is possible to generalize this definition further (see e.g. Brockett [1973]),but the above is usually sufficient for our purposes.

A enormous amount can of course be said about such systems. We re-strict ourselves here to the concepts of accessibility, controllability, andstabilizablity and feeback linearizability.

We also restrict ourselves here to considering so-called affine nonlinearcontrol systems which have the form

x = f(x) +m∑i=1

gi(x)ui , (4.1.2)

where f and the gi, i = 1, . . . ,m are smooth vector fields on M . f is usuallycalled the drift vector field and the gi are called the control vector fields.

4.1.1 Controllability and Accesibility

Definition 4.1.2. The system 4.1.2 is said to be controllable if for anytwo points x0 and xf in M there exists a control u(t) such that the system4.1.2 with initial condition x0 reaches the point xf in some finite time T .

This is a basic concept in control theory and much work has been doneon deriving sufficient conditions for it. However, proving controllability isnot easy in general. A related property, called local accessibility is oftenmuch easier to prove.

To define accessbility we first need the notion of reachable set. The reach-able set from a given point at time T is essentially the set of points thatmay be reached by the system by travelling on trajectories from the initialpoint.

More precisely:

Definition 4.1.3. Given x0 ∈ M we define R(x0, T ) to be the set of allx ∈ M for which there exists a control u such that there is a trajectory of(4.1.2) with x(0) = x0 and x(T ) = x. The reachable set at time T isdefined to be

RT (x0) = ∪t≤TR(x0, t) . (4.1.3)

4.1 Nonlinear Control Systems 109

We also define:

Definition 4.1.4. The accessiblity algebra C of (4.1.2) is the smallestLie algebra of vector fields on M that contains the vector fields f andg1, · · · , gm.

Note that the accessbility algebra is just the span of all possible Liebrackets of f and the gi.

Definition 4.1.5. We define the accessbility distribution C of 4.1.2 to bethe distribution generated by the vector fields in C, i,e, C(x) is the span ofthe vector fields X in C at x.

We then have (see e.g., Nijmeijer and van der Schaft [1990] or Sontag[1990])

Theorem 4.1.6. Consider the system (4.1.2). If dim C(x0) = n ( i.e.the accessibilty algebra spans the tangent space to M at x0) then for anyT > 0, the set RT (x0) has nonempty interior.

If the rank condition in the preceding theorem holds we say the systemis accessible at x0 or that the accessibility rank condition holds.

Note that while this spanning condition is an intuitively reasonable con-dition the resulting theorem is quite weak and is far from applying control-lability.

In certain special cases the accessibility rank condition does imply con-trollability however.

Theorem 4.1.7. If dim C(x) = n everywhere on M and either

(i) f = 0, or

(ii) f is divergence free and M is compact

then (4.1.2) is controllable everywhere.

The idea behind this result is that one cannot move “backwards” alongthe drift directions and hence a spanning condition inolving the drift vectorfield does not guarentee controllability. However, if the drift is divergencefree and the underlying manifold is compact Poincare recurrance (see e.g.Arnold [1978]) ensure a drift “backward” eventually.

Exercise: Show that the Heisenberg system

x = u

y = v

z = vx− uy (4.1.4)

is accessible everywhere and hence controllable everywhere in R3.

110 4. An Introduction to Geometric Control Theory

Another case of interest where controllabiltiy implies accesibility is alinear system of the form:

x = Ax+m∑i=1

biui = Ax+Bu (4.1.5)

where x ∈ Rn and A ∈ Rn × Rn and B ∈ Rn × Rm are constant matrices,bi being the columns of B.

The Lie bracket of the drift vector field Ax with bi is −Abi. Bracketingthe latter constant field with Ax and so on, tells is that C is spanned byAx, bi, Abi, · · ·An−1bi, i = 1, · · ·m.

This gives the standard controllability rank condition:

Theorem 4.1.8. The system 4.1.5 is controllable if and only if

rank[B,AB, · · · , An−1B] = n . (4.1.6)

Note that in this case accessibilty is equivalent to controllability but thatthe drift vector field is involved.

4.1.2 Stability and Stabilizability

In this section we summarize briefly some notions of stability and stabiliz-ability.

Definition 4.1.9. Let x0 be an equilibrium of the system of differentialequations x = f(x). The point x0 is said to be nonlinearly or Lyapunovstable if for any neigborhood U of xO there exists a neigborhood V ⊂ Uof x0 such that any trajectory x(t) of the system with initial point in Vremains in U for all time. If in addition x(t) → x0 as t → ∞, x0 is saidto be aymptotically stable.

Definition 4.1.10. Let x0 be an equilibrium of the system of differentialequations x = f(x). x0 is said to be spectrally stable if all the eigenvaluesof the linearization of f at x0 lie in the left half plane.

A theorem of Liapunov states that spectral stability implies asymptoticstability.

Definition 4.1.11. Let x0 be an equilibrium of the control system x =f(x, u). The system is said to be nonlinearly (aymptotically) stabliz-able at x0 if a feedback control u(x) can be found which renders the systemnonlinearly (asympotically) stable.

We remark that in much of the control literature stabilizability is takento mean asymptotic stabilizability. However in this book we will distinguishbetween the two.

4.2 Stabilization Techniques 111

4.2 Stabilization Techniques

Here we discuss some stabilization techniques in the literature that are ofrelevence to the material in this book. Good references for classical materialon this subject are Sontag [1990] and Nijmeijer and van der Schaft [1990].See also Brockett [1983].

4.2.1 Linearization

The simplest result on stabilization for nonlinear systems is based on lin-earization. Consider the nonlinear control system

x = f(x, u) , (4.2.1)

with x ∈ Rn (for simplicity) and u ∈ Rm, and its linearization about(x0, u0),

x = Ax+Bu (4.2.2)

where A = ∂f∂x (x0, u0) and B = ∂f

∂u (x0, u0).Then if R is the reachable subspace of the linearized system, i.e. the

range of [B,AB, · · · , An−1], we may find a change of basis such that thesystem takes the form

d

dt

[xuxl

]=[A11 A12

0 A22

] [xuxl

]+[Bu0

]u , (4.2.3)

where Range [B,AB, · · · , An−1] has dimension xu.If the eigenvalues of A22 has all eigenvalues in the left half plane a feed-

back may be found which stabilizes the system asymptotically to the origin.(For further details see e.g. Nijmeijer and Van der Schaft [1990]).

4.2.2 Necessary Conditions

A beautiful general theorem on necessary conditions for feedback stabiliza-tion of nonlinear systems was given by Brockett [1983].

Theorem 4.2.1 (Brockett). Consider the nonlinear system x = f(x, u)with f(x0, 0) = 0 and f(·, ·) continuously differentiable in a neighborhoodof (x0, 0). Necessary conditions for the existence of a continuously differ-entiable control law for asymptotically stabilizing (x0, 0) are:

(i) The linearized system has no uncontrollable modes associated witheigenvalues with postive real part.

112 4. An Introduction to Geometric Control Theory

(ii) There exists a neighborhood N of (x0, 0) such that for each ξ ∈ Nthere exists a control uξ(t) defined for all t > 0 which drives the solutionof x = f(x, uξ) from the point x = ξ at t = 0 to x = x0 at t =∞.

(iii) The mapping γ : N × Rm → Rn, N a neighborhood of the origin,defined by γ : (x, u)→ f(x, u) should be onto an open set of the origin.

Proof. Part (i) has been proved above. (ii) holds since if a system isasymptotically stabilizable clearly there solutions near the origin must ap-proch it.

To prove (iii) consider the closed loop system

x = f(x, u(x)) ≡ a(x) (4.2.4)

and suppose that x0 is locally aymptotically stable. Then for sufficientlysmall r the map from the ball of radius r about x0 into Sn−1 given by

x→ a(x)||a(x)|| (4.2.5)

has degree n − 1 (see e.g. Milnor [1965]). Since this degree is nonzero themap is actually onto and hence the result. ¥

A stronger necessary was condition was given by Coron [1989] but we donot discuss this here.

4.2.3 Lyapunov Methods for asymptotic stability

Another case of interest is that where the free system is Lyapunov stable,and the controls may be used to make the system as ymptotically stable.Issues of this sort are of key importance in the recent work Bloch, Leonardand Marsden [1997], [1998] for example and we shall return to them later.In that case the Hamiltonian structure of the system is used for energyshaping and then a suitable dissipation is added.

For the moment we just summarize the kind of general argument thatone can find for example in Nijmeijer and van der Schaft [1990].

Consider again the affine control system

x = f(x) +m∑i=1

uigi(x) . (4.2.6)

Suppose that the free systems has equilibrium x0 and suppose thereexists a Lyapunov function V (x) for the free system about x0 in someneighborhood U of x0.

One then considers feedback of the following form:

ui(x) = −LgiV (x) . (4.2.7)

4.2 Stabilization Techniques 113

Then x0 is an equilibrium point for the closed loop system and V remainsa Lyapunov function for the the closed loop system The question is nowwhether the system is in addition asympotically stable. For this we can usea LaSalle invariance principle argument.

Consider the set

W = x ∈ U | LfV (x) = 0, LgiV (x) = 0, i = 1, · · ·m . (4.2.8)

Let W0 be the largest invariant set in W under the closed loop dynamics.We observe that x0 is in W since V is a Lypaunov function. If W0 isidentically equal to x0 then x0 is a locally an asympototically stableequilibrium point by LaSalle. Notice also that any trajectory of the closedloop system in W is also a trajectory of the free dynamics. Hence by LaSallethe system is locally asymptotically stable about x0 if the largest invariantsubset of the free dynamics in W equals x0. If in addition dV (x) 6= 0 forx ∈ U/x0 this condition is also necessary since, by the definition of W ,along any trajectory in W of the closed loop system and hence of the freesystem the Lyapunov function is constant and a nontrivial trajectory cantherefore not approach x0.

Of course one would like to apply this reasoning without knowing thefree dynamics. Some conditions for this are discussed in Nijmiejer and vander Schaft. The intuitive idea of course is that one would like W to be assmall as possible. Locally a sufficient condition for stability is simply thatthe distribution

spanf(x), adkfgi(x), i = 1 · · ·m, , for all k ≥ 0 (4.2.9)

should have rank n at x0. But this is equivalent to the linearized systembeing stabilizable.

4.2.4 The center manifold

Another important tool in the analysis of the stabilization of nonlinearcontrol systems is center manifold theory. Given that there exists a centermanifold for the system and the remaining dynamnics is stable one canconcentrate on the stabilizing the center manifold dynamics. A nice appli-cation of this technique to the rigid body dynamics is given in the work ofAeyels (see e.g. Aeyels [1985]) and the discussion in Nijmeijer and van derSchaft [1990]. We do not give any details here but will return to the centermanifold when we consider nonholnomic systems.

114 4. An Introduction to Geometric Control Theory

4.3 Hamiltonian and Lagrangian ControlSystems

This extension to the notion of Hamiltonian and Lagrangian systems to thesetting of control was formally proposed in Brockett [1976] and extendedand formalized by Willems [], van der Schaft [1983, 1986] and others. Thebook Nijmeijer and van der Schaft [1990] gives a nice summary of many ofthe ideas.

We begin with the Lagrangian side. The simplest form of Lagrangiancontrol system is a Lagrangian system with external forces: in local coor-dinates we have

d

dt

(∂L

∂qi

)− ∂L

∂qi= ui, i = 1, · · · ,m

d

dt

(∂L

∂qi

)− ∂L

∂qi= 0, i = m+ 1, · · ·n · · · (4.3.1)

More generally we have the system

d

dt

(∂L(q, q, u)

∂qi

)− ∂L(q, q, u)

∂qi= 0 (4.3.2)

for q ∈ Rn and u ∈ Rm.The latter system is a nontrivial generalization of the former, for example

if the control input is a velocity rather than a force (see Brockett [1976],Nijmeijer and van der Schaft [1990]).

Similarly one can define a Hamiltonian control system. Locally one has:

qi =∂H(q, p, u)

∂pi

pi = −∂H(q, p, u)∂qi

(4.3.3)

for q, p ∈ Rn and u ∈ Rm.One can easily generalize this to a Hamiltonian control system on a

symplectic or Poisson manifold.Let M be a Poisson manifold and H0, H1, · · · , Hm be smooth functions

on M . Then an affine Hamiltonian control system on M is given by

x = XH0(x) +m∑i=1

XHj (x)ui (4.3.4)

where x ∈M and as usual XHj is the Hamiltonian vector field correspond-ing to Hj .

We shall return to systems of this type. Good discussion are given inKrishnaprasad [1985] and Sanchez de Alvarez [1986].

This is page 115Printer: Opaque this

5Nonholonomic Mechanics

One of the points of closest contact between control theory and mechanicsis the theory of nonholonomic mechanical systems. One reason for thisis that nonintegrability is essential to both: nonintegrable constraints arethe essence of nonholonomic systems, while a nonintegrable distribution ofcontrol vector fields is the key to controllabiltiy of nonlinear systems.

Nonintegrable constraints – systems with constraints on the velocitywhich are not derivable from positions consraints arise in such mechanicalsituations as that of rolling contact (wheels) or sliding contact (a skate).However, such constraints also occur in less obvious ways. For example,one may view angular momentum constraints, which are really integralsof motion and are integrable constraints on the phase space (functions ofposition and momentum) as nonintegrable constraints on the configurationspace. This point of view is very helpful for contollability. Further, manyfirst order control systems may be simply viewed as controlled distributionslying in the kernal of a nonintegrable constraint. A classic example of thelatter is the Heisenberg system.

5.1 Equations of Motion

We follow for the moment here Bloch, Krishnaprasad, Marsden and Murray[1996]. We begin by deriving the equations of motion for a nonholonomicuncontrolled system from the the Lagrange-D’Alembert principle, althoughthis is by no means the only way of deriving the equations. We shall come

116 5. Nonholonomic Mechanics

back to this point.

5.1.1 The Lagrange d’Alembert Principle

The starting point is a configuration space Q and a distribution D thatdescribes the kinematic constraints of interest. Thus, D is a collection oflinear subspaces denoted Dq ⊂ TqQ, one for each q ∈ Q. A curve q(t) ∈ Qwill be said to satisfy the constraints if q(t) ∈ Dq(t) for all t. This distribu-tion will, in general, be nonintegrable; i.e., the constraints are, in general,nonholonomic. One of our goals below will be to model the constraints interms of Ehresmann connections (see Cardin and Favretti [1994] and Marle[1994] for some related ideas).

The above setup describes linear constraints; for affine constraints, forexample, a ball on a rotating turntable (where the rotational velocity ofthe turntable represents the affine part of the constraints), one assumesthat there is a given vector field γ on Q and the constraints are writtenq(t)− γ(q(t)) ∈ Dq(t).

Consider a Lagrangian L : TQ→ R. In coordinates qi, i = 1, . . . , n, on Qwith induced coordinates (qi, qi) for the tangent bundle, we write L(qi, qi).Following standard practice, we assume that the equations of motion aregiven by the by the so-called principle of Lagrange d’Alembert (see, forexample, Rosenberg [1977] for a discussion).

Definition 5.1.1. The Lagrange d’Alembert equations of motionfor the system are those determined by

δ

∫ b

a

L(qi, qi) dt = 0, (5.1.1)

where we choose variations δq(t) of the curve q(t) that satisfy δq(t) ∈ Dq(t)for each t, a ≤ t ≤ b.

This principle is supplemented by the condition that the curve itselfsatisfies the constraints. In such a principle, we follow standard procedureand take the variation before imposing the constraints; that is, we do notimpose the constraints on the family of curves defining the variation. Theusual arguments in the calculus of variations show that this constrainedvariational principle is equivalent to the equations

−δL =(d

dt

∂L

∂qi− ∂L

∂qi

)δqi = 0, (5.1.2)

for all variations δq such that δq ∈ Dq at each point of the underlying curveq(t).

To explore the structure of these equations in more detail, consider amechanical system evolving on a configuration space Q with a given La-grangian L : TQ → R and let ωa be a set of p independent one forms

5.1 Equations of Motion 117

whose vanishing describe the constraints on the system. The constraintsin general are nonintegrable. Choose a local coordinate chart and a localbasis for the constraints such that

ωa(q) = dsa +Aaα(r, s)drα a = 1, . . . ,m, (5.1.3)

where q = (r, s) ∈ Rn−m × Rm.The equations of motion for the system are given by (5.1.2) where we

choose variations δq(t) that satisfy the condition ωa(q) · δq = 0, i.e.,where the variation δq = (δr, δs) satisfies δsa + Aaαδr

α = 0. Substitutioninto (5.1.2) gives(

d

dt

∂L

∂rα− ∂L

∂rα

)= Aaα

(d

dt

∂L

∂sa− ∂L

∂sa

), α = 1, . . . , n−m. (5.1.4)

Equation (5.1.4) combined with the constraint equations

sa = −Aaαrα a = 1, . . . ,m (5.1.5)

gives a complete description of the equations of motion of the system.We now define the “constrained” Lagrangian by substituting the con-

straints (5.1.5) into the Lagrangian:

Lc(rα, sa, rα) = L(rα, sa, rα,−Aaα(r, s)rα).

The equations of motion can be written in terms of the constrained La-grangian in the following way, as a direct coordinate calculation shows:

d

dt

∂Lc∂rα− ∂Lc∂rα

+Aaα∂Lc∂sa

= − ∂L∂sb

Bbαβ rβ , (5.1.6)

where

Bbαβ =

(∂Abα∂rβ

−∂Abβ∂rα

+Aaα∂Abβ∂sa

−Aaβ∂Abα∂sa

). (5.1.7)

Letting dωb be the exterior derivative of ωb, another computation (seeRemark 4 below) shows that

dωb(q, ·) = Bbαβ rαdrβ

and hence the equations of motion have the form

−δLc =(d

dt

∂Lc∂rα− ∂Lc∂rα

+Aaα∂Lc∂sa

)δra = − ∂L

∂sbdωb(q, δr).

This form of the equations isolates the effects of the constraints, and showsthat in the case where the constraints are integrable (dω = 0) then thecorrect equations of motion are obtained by substituting the constraintsinto the Lagrangian and setting the variation of Lc to zero. However inthe non-integrable case the constraints generate extra forces which mustbe taken into account.

118 5. Nonholonomic Mechanics

5.1.2 Nonholonomic equations of motion withLagrange multipliers

We can obtain the nonholonomic equations of motion with Lagrange mul-tipliers from the Lagrange D’Alembert principle as follows see e.g. Neimarkand Fufaev [1972]:

Recall that the Lagrange-D’Alembert principle gives us:

(d

dt

∂qiL− ∂

∂qiL)δqi = 0 , i = 1 · · ·n . (5.1.8)

for variations δqi ∈ DQ, i.e. for variations in the constraint distribution,i.e. for variations satisfying

n∑i=1

aji δqi = 0 , j = 1 · · ·m. (5.1.9)

But this implies thatm∑j=1

n∑i=1

λjaji δq

i = 0 . (5.1.10)

Hence we can append this to the Lagrange-D’Alembert equations to obtainn∑i=1

(d

dt

∂qiL− ∂

∂qiL−

m∑j=1

λjaji )δq

i = 0 . (5.1.11)

Now, assuming our constraints are independent implies that one of them by m minors of the matrix aji must be nonzero. Let us assume it to bethat given by letting the indices run from 1 to m.

We can now assume that the variations δqm+1 · · · δqn are arbitrary sincethe constraint equations 5.1.9 may then be satisfied by a choice of resultingdefinite values for variations δq1, · · · , δqm.

We can now choose values of the Lagrange multipliers such that the ex-pression in brackets in 5.1.11 vanishes for each dependent variation δq1, · · · , δqm.This entails solving a linear systems of algebraic equations in the λi whichis solvable by virtue of our assumptions on the aji .

Once this is done however equation 5.1.11 becomesn∑

i=m+1

(d

dt

∂qiL− ∂

∂qiL−

m∑j=1

λjaji )δq

i = 0 . (5.1.12)

where the variations are independent. Hence each term in the brackets mustalso vanish independently. Putting the observations for the dependent andindependent variable together gives us the set of n equations

d

dt

∂qiL− ∂

∂qiL =

m∑j=1

λjaji . (5.1.13)

5.2 An Invariant Approach to Nonholonomic Mechanics 119

5.2 An Invariant Approach toNonholonomic Mechanics

In this section we consider an invariant approach to Lagrangian mechanicsin general and nonholonomic mechanics in particular. We follow here theapproach of Vershik [] (see also Wang []). A general invariant approach toLagrangian mechanics is also discussed for example in Marsden and Ratiu[1994]. We remark that the approach here is somewhat more abstract thanother parts of the book and may be omitted without loss of contintuity.

5.2.1 Lagrangian Mechanics

We consider firstly a general invariant formulation of Lagrangian mechanicswithout constraints.

Let TQ be the tangent bundle of Q, an n-dimensional manifold, with itscanonical projection and let it be locally co-ordinitized by (q, q) ∼ (q, v).Here q, v represent n-vectors. We consider here notions of verticality andhorizontality with respect to the trivial connection on TQ, i.e. combinationsof vectors ∂

∂qiare horizontal and combinations of vectors ∂

∂viare vertical. (It

is important to bear this in mind when discussing constraints below, so asnot to confuse this discussion with the connection given by the constraintsdiscussed elsewhere.)

Since Lagrangian equations are second order we also need the secondtangent bundle T (TQ) with projection dπ : T (TQ) → TQ locally co-ordinitized by (q, v, v, v). In fact we want vector fields of the type

q = f(q, q) (5.2.1)

or

q = v

v = f(q, v). (5.2.2)

Written as a vector field a second order equation takes the form

Xq,v = Xs = v∂

∂q+ f

∂v, (5.2.3)

(with the obvious summation notation).Invariantly we may write this as dπXs = v – its projection onto the

tangent space to Q is just v. We call such a vector field Xs a “special”vector field.

Now one can formulate Lagrangian mechanics in an invariant fashion bydefining fields on TQ as follows:

Let TqQ and Tq,vTQ define the tangent space to Q at q and TQ at (q, v)respectively.

120 5. Nonholonomic Mechanics

One can set up a canonical map from TqQ to Tq,vTQ which takes thefiber over q (locally vectors of the form a ∂

∂q ) into the fiber over Tq,vTQ(vectors of the form a ∂

∂q + b ∂∂v ). Locally this map is defined by

γq,v

(a∂

∂q

)= a

∂v. (5.2.4)

We now define a tensor field on TQ, called the principal tensor field, by

τq,v = γq,v dπq,v. (5.2.5)

So locally we have

τq,v

(a∂

∂q+ b

∂v

)= γq,v dπq,v

(a∂

∂q+ b

∂v

)= γq,v

(a∂

∂q

)= a

∂v. (5.2.6)

The dual tensor field τ∗ acts locally on forms as follows

τ∗(adq + bdv) = bdq. (5.2.7)

This can be easily checked: we have⟨τ∗(adq + bdv), c

∂q+ d

∂v

⟩=⟨adq + bdv, τ

(c∂

∂q+ d

∂v

)⟩=⟨adq + bdv, c

∂v

⟩= bc. (5.2.8)

Hence

τ∗(adq + bdv) = bdq. (5.2.9)

Now we define another key concept — the fundamental vector field on TQ:this is a field with coordinates Φq,v = γq,vv

∂∂q = vi

∂∂vi

. Clearly a vectorfield X is special if and only if τX = Φ, since in local coordinates a specialvector field is of the form

Xq,v = vi∂

∂qi+ · · · . (5.2.10)

Now let L be the Lagrangian — as usual a smooth function on TQ. Wenote that locally

τ∗(dL) = τ∗(∂L

∂qidqi +

∂L

∂qidqi)

=∂L

∂qidqi (5.2.11)

5.2 An Invariant Approach to Nonholonomic Mechanics 121

— clearly a horizontal 1-form since it annihilates vertical vector fields a ∂∂v

(Vershik calls this form the impulse field of the Lagrangian.) One can iden-tify horizontal 1-forms in Lagrangian mechanics as forces – see later).

We define the Lagrangian 2-form to be

ΩL = d(τ∗dL)

= d

(∂L

∂qidqi

)=

∂2L

∂qj∂vidqj ∧ dqi +

∂2L

∂vj∂vidqj ∧ dvi. (5.2.12)

Recalling the Legendre transformation pi = ∂L∂qi

we see that the image ofΩL under the Legendre transformation is the canonical symplectic form onthe cotangent bundle. Similarly the Hamiltonian (energy) is

HL = dL(Φ)− L. (5.2.13)

Note that in local coordinates

dL(Φ) = dL

(vi

∂vi

)=∂L

∂vivi = pivi , (5.2.14)

(with the summation convention).Now we can formulate the Lagrange d’Alemert principle as follows:

Definition 5.2.1 (The Lagrange D’Alembert Principle). The vec-tor field Y describing the mechanical trajectories of motion is given by

ΩL(X,Y ) = dHL(Y ) + ω(Y ) (5.2.15)

where ω is the 1-form describing the exterior forces and X is a specialvector field.

In the absence of exterior forces we recover the Lagrange equations asfollows. Let

X = v∂

∂q+ v

∂vY = a

∂q+ b

∂v. (5.2.16)

Then

ΩL(X,Y ) =∂2L

∂qj∂vi(vjai − ajvi) +

∂2L

∂vj∂vi(vjbi − ajvi) (5.2.17)

dHL(Y ) = d

(∂L

∂vivi − L

)(Y )

=( ∂L∂vi

dvi +∂2L

∂qj∂vividq

j +∂2L

∂vi∂vjvidvj

− ∂L

∂qidqj −

∂L

∂vidvi

)(Y )

=∂L

∂vibi +

∂2L

∂qj∂vjviaj +

∂2L

∂vj∂vivibj −

∂L

∂qiai −

∂L

∂vibi .(5.2.18)

122 5. Nonholonomic Mechanics

Equating coefficients of ai and bi which are arbitrary we get:

∂2L

∂qj∂vjvi −

∂2L

∂qi∂vjvj −

∂2L

∂vj∂vivi = − ∂L

∂qi+

∂2L

∂qj∂vivj

∂2L

∂vi∂vjvj =

∂L

∂vi+

∂2L

∂vj∂vjvj −

∂L

∂vi

The second equation is an identity while the first equation is just

d

dt

(∂L

∂vi

)=∂L

∂qi. (5.2.19)

as required.

5.2.2 Constrained Dynamics

We now consider the invariant formulation of dynamics with constraints.With Vershik we assume a slightly more general form of the constraints

– we assume that they define a distribution on TQ, i.e. they define at eachpoint a subspace of TTQ(q, q). This fits naturally with the definition of2nd order systems, although it is more general than needed. Hence we candefine the constraints to be 1-forms on TQ (a codistribution of TQ) andthey thus take the form

θi = Σnk=1aik(q)dqk + Σnk=1bik(q)dvk, i = 1 · · ·m. (5.2.20)

A constrained dynamical system is then a special (2nd order) vector fieldcompatible with these constraints. The case of interest to us is when the θiare differentials of functions on TQ i.e. of the form df(q, v) – in particularthe case f(q, v) =

∑k aik(q)vk so

θi =∑k

aikdvk +∑k,j

∂aik∂qj

vkdqj . (5.2.21)

We now want to define reaction forces — the forces that keep the systemon the constant manifold and to show that the in the “ideal” case suchforces do no work.

Definition 5.2.2. Given constraints θi as in (5.2.20 ) we define τ∗θi(=∑bikdqk) to be the reaction forces.

Note that in the case (5.2.21) this is just

τ∗θi =∑k

aikdqk. (5.2.22)

5.2 An Invariant Approach to Nonholonomic Mechanics 123

Definition 5.2.3. A set of of constraints is said to be admissible if Dimspan τ∗θi =Dimspan θi.

(This is equivalent to saying that the codistribution given by the θi hasno horizontal covectors at any point, since the kernal of τ∗ is horizontal1-forms. Another way to think of this: since horizontal vectors are linearcombinations of vectors ∂

∂qi, the vectors ai = (ai1, · · · , ain) must be linear

combination of the vectors bi = (bi1, · · · , bin) so that if one has some vectorsin the span of the ∂

∂qithere is also a component in the span of the ∂

∂vi.

Definition 5.2.4. A constraint is said to be ideal if it annihilates thefundamental vector fieldΦ (=

∑i vi

∂∂vi

locally).

Then we have

Theorem 5.2.5. If a set of constraints is admissable there exist specialvector fields satisfying the constraints.

Proof. Recall that the special vector fields are those satisfying τX = Φ.Therefore we need to ask if the system

τX = Φ θi(X) = 0, i = 1 · · ·m (5.2.23)

is solvable.Let X =

∑i vi

∂∂qi

+ fi∂∂vi

. Since the constraints are admissable, by theargument above the vectors ai are in the span of the vectors bi. Now werequire

θi(X) =∑i

bikfk +∑i

aikvk = 0. (5.2.24)

Thus we wish to know whether the system of linear equations∑i

bikfk = −∑i

aikvk

in the fi can be solved. But this is clear by the admissability argument.In fact, since there are m conditions on the fi there are at least n − mindependent special vector fields. ¥

Now we say a force does no work if∫F · dq = 0 along any curve in the

configuration space.

Theorem 5.2.6. If a constraint is ideal the corresponding constraint forcesdo no work (i.e. they are virtual forces.)

Proof. Let ξ be a closed curve in Q, i.e. a map from S1 to Q and let ξbe its lift to TQ.

124 5. Nonholonomic Mechanics

Then ∫ξ

τ∗θi =∫S1

⟨τ∗θi, ξ

⟩=∫S1

⟨θi, τ ξ

⟩where <,> is the natural pairing between forms and their duals.But if ξ = q(t) and ξ = (q(t), q(t)) then

ξ = q∂

∂q+ q

∂v. (5.2.25)

Hence τ(ξ) = q ∂∂v = v ∂

∂v = Φ. Thus we get∫S1〈θi,Φ〉 = 0 (5.2.26)

by ideality.The usual arguments imply this integral is zero along any curve. ¥

Now we can show

Theorem 5.2.7. Let θi, i = 1 . . .m define a constraint distribution onTQ and let L be a nondegenerate Lagrangian with positive definite Hessian∂2L∂vi∂vj

. Then there exists a special vector field X satisfying the Lagranged’Alembert principle

ΩL(X) = (dHL + ω)(X) , (5.2.27)

where ω is the constraint force that ensures θi(X) = 0, i = 1, . . . ,m. Locallythe equations of motion are

d

dt

(∂L

∂qi

)− ∂L

∂qi=∑

λiθi. (5.2.28)

Proof. Since ΩL is nondegenerate there is a map form 1-forms to vectorfields πL such that

ΩL (πL(ρ), Y ) = ρ(Y ) (5.2.29)

for Y an arbitrary tangent vector and ρ an arbitrary 1-form. Then

X = πL(dHL + ω) (5.2.30)

and we require⟨θi, πL(dHL) + πL(ω)

⟩= 0 i = 1, . . . ,m. (5.2.31)

Now ω, by definition of the constraint force, needs to be a linear combina-tion of τ∗θi, i.e. of the form

∑i λiaikdqk.That is, we have⟨

θi , πL(dHL)⟩

= −⟨θi ,∑i

λiπLτ∗θi⟩i = 1, . . . ,m. (5.2.32)

5.3 Projected Connections and Newton’s law for Nonholonomic Systems 125

Since the constraints are admissable, τ∗ preserves the dimension of the spanof θi, but since πL is positive definite the restriction of πL to the span ofthe τ∗θi is also positive definite and thus nondegenerate. Since the operatorπLτ

∗ is thus of full rank, we can solve the system for λi.The local form of the equations follows from the general Lagrange theory

above and of course agree with the coordinate form derived elsewhere. ¥

5.3 Projected Connections and Newton’slaw for Nonholonomic Systems

In this section we follow up on the formulation of the equations of motion ofnonholonomic systems with forces or controls discussed in the previous sec-tion and indicate how to write such system as second order forced systems(i.e. satisfying Newton’s law) on the constraint subbundle. This followsas approach due to Vershik and Fadeev [1981], but see also Montgomery[1990], Yang [] and Bloch and Crouch [1995, 97].

Let us rewrite our forced nonholonomic system 6.5.6 on the Riemannianmanifold M as

D2q

dt2=

m∑i=1

λiWi + F (5.3.1)

where F is an arbitrary external force field and the system is subject tothe independent constraints

ωi(q) =< Wi, q >= 0 , (5.3.2)

where the Wi are vector fields on M .The constraint distribution N is given by

Np = Xp : ωi(X) = 0, 1 ≤ i ≤ m (5.3.3)

for X a vector field on M and p a point on M .We now use the independence of the one forms ωi to eliminate the mul-

tipliers: Let aij = ωi(Wj), 1 ≤ i, j ≤ m. Then the matrix with entries aijis invertible on M .

Differentiating the constraints gives

Dωi∂t

(q) + ωi(∑j

λjWj + F ) = 0, 1 ≤ i ≤ m. (5.3.4)

Hence the multipliers have the explicit form:

λk = −∑j

a−1kj

(Dωj∂t

(q) + ωj(F )), 1 ≤ k ≤ m. (5.3.5)

126 5. Nonholonomic Mechanics

Hence the equations of motion may be rewritten as

D2q

dt2+∑k,j

Wka−1kj

Dωj∂t

= F −∑k,j

Wka−1kj ωj(F )

ωi(q) = 0 1 ≤ i ≤ m (5.3.6)

Now let X e any vector field on M .

Definition 5.3.1. Let

πN (X) = X −∑ki

Wka−1ki ωi(X) . (5.3.7)

We note that πN (X) is a projection of X onto the constraint distributionN .

Now apply this projection to D2qdt2 and observe that differentiating the

constraints gives

ωi

(D2q

dt2

)+Dωi∂t

(q) = 0 .

Thus

πN

(D2q

dt2

)=D2q

dt2+∑i,k

Wka−1ki

Dωi∂t

(q) . (5.3.8)

Hence equation 5.3.6 may be written

πN

(D2q

dt2

)= πN (F ), q ∈ N . (5.3.9)

We now make the following natural definition:

Definition 5.3.2. If ∇ is any connection on M with corresponding co-variant derivative D

∂t then the second order system

D2q

dt2= F , q ∈M . (5.3.10)

is said to be a system on M satisfying Newton’s law with forces F .

Thus our nonholonomic systemis a projection of the Newton law systemD2qdt2 = F to N .

However, as Vershik and Fadeev pointed out, by redefining the connec-tion, the system may be written as a system obeying Newton’s law on Ndirectly.

We define a new connection on M ;

∇′XY = ∇XY +m∑i=1

Wia−1ik (∇Xωk)(Y ) , (5.3.11)

5.4 Systems with Symmetry 127

for X,Y vector fields on M . This in turn defines a covariant derivative D′

∂t .From 5.3.11 we note that

ωi(∇′XY ) = ωi(∇XY ) + (∇xωi)(Y ) = X(ωi(Y )) . (5.3.12)

Thus is X and Y are vector fields on N so is ∇′XY and thus ∇′|N definesa connection on the subbundle N of TM and the nonholonomic system5.3.1 may be viewed as a system of Newton law type on N :

D′2q

dt2= πN (F ), q ∈ N . (5.3.13)

Thus the nonholonomic system is a systems of Newton law type withrespect to a modified connection. This connection is not metric however.In Bloch and Crouch [1997] we explore extensions of this system to a non-metric Newton law system on the whole of M .

5.4 Systems with Symmetry

We now add symmetry to our nonholonomic system. We will begin withsome general remarks about symmetry.

5.4.1 Group Actions and Invariance

We remind the reader of a few of the concepts discussed ealier. As beforewe refer the reader to Marsden & Ratiu [1994], Chapter 9 for furtherdetails. Assume that we are given a Lie group G and an action of G on Q.The action of G will be denoted q 7→ gq = Φg(q). The group orbit througha point q, which is always an (immersed) submanifold, is denoted

Orb(q) := gq | g ∈ G.

When there is danger of confusion about which group is meant, we writethe orbit as OrbG(q).

Let g denote the Lie algebra of the Lie group G . For an element ξ ∈ g, wewrite ξQ, a vector field on Q for the corresponding infinitesimal generator;recall that this is obtained by differentiating the flow Φexp(tξ) with respectto t at t = 0. The tangent space to the group orbit through a point q isgiven by the set of infinitesimal generators at that point:

Tq(Orb(q)) = ξQ(q) | ξ ∈ g.

Throughout this paper we make the assumption that the action of Gon Q is free (none of the maps Φg has any fixed points) and proper (themap (q, g) 7→ gq is proper; i.e., that is, the inverse images of compact

128 5. Nonholonomic Mechanics

sets are compact). The case of nonfree actions is very important and theinvestigation of the associated singularities needs to be carried out, butthat topic is not the subject of the present paper.

The quotient spaceM = Q/G, whose points are the group orbits, is calledshape space . It is known that if the group action is free and proper thenshape space is a smooth manifold and the projection map π : Q→ Q/G isa smooth surjective map with a surjective derivative Tqπ at each point. Wewill denote the projection map by πQ,G if there is any danger of confusion.The kernel of the linear map Tqπ is the set of infinitesimal generators ofthe group action at the point q, i.e.,

kerTqπ = ξQ(q) | ξ ∈ g ,

so these are also the tangent spaces to the group orbits.We now introduce some assumptions concerning the relation between the

given group action, the Lagrangian, and the constraint distribution.

Definition 5.4.1.

(L1) We say that the Lagrangian is invariant under the group action ifL is invariant under the induced action of G on TQ.

(L2) We say that the Lagrangian is infinitesimally invariant if for anyLie algebra element ξ ∈ g we have dL ξQ = 0 where, for a vectorfield X on Q, X denotes the vector field on TQ naturally induced byit (if Ft is the flow of X then the flow of X is TFt).

(S1) We say that the distribution D is invariant if the subspace Dq ⊂TqQ is mapped by the tangent of the group action to the subspaceDgq ⊂ TgqQ.

(S2) An Ehresmann connection A on Q (that has D as its horizontal distri-bution) is invariant under G if the group action preserves the bundlestructure associated with the connection (in particular, it maps verti-cal spaces to vertical spaces) and if, as a map from TQ to the verticalbundle, A is G-equivariant.

(S3) A Lie algebra element ξ is said to act horizontally if ξQ(q) ∈ Dqfor all q ∈ Q.

Some relationships between these conditions are as follows: condition(L1) implies (L2), as is obtained by differentiating the invariance condition.It is also clear that condition (S2) implies the condition (S1) since theinvariance of the connection A implies that the group action maps its kernelto itself. Condition (S1) may be stated as follows:

TqΦg · Dq = Dgq (5.4.1)

5.5 The Momentum Equation 129

In the case of affine constraints, we will explicitly state when we need theassumption that the vector field γ be invariant under the action.

To help explain condition (S1), we will rewrite it in infinitesimal form.Let XD be the space of sections X of the distribution D; that is, the spaceof vector fields X that take values in D. The condition (S1) implies thatfor each X ∈ XD, we have Φ∗gX ∈ XD. Here, Φ∗gX denotes the pull back ofthe vector field X under the map Φg. Differentiation of this condition withrespect to g proves the following result.

Proposition 5.4.2. Assume that condition (S1) holds and let X be asection of D. Then, for each Lie algebra element ξ, we have

[ξQ, X] ∈ XD (5.4.2)

which we also write as

[ξQ,XD] ⊂ XD.

We now have

Proposition 5.4.3. Under assumptions (L1) and (S1), we can form thereduced velocity phase space TQ/G and the constrained reduced ve-locity phase space D/G. The Lagrangian L induces well defined func-tions, the reduced Lagrangian

l : TQ/G→ R

satisfying L = l πTQ where πTQ : TQ→ TQ/G is the projection, and theconstrained reduced Lagrangian

lc : D/G→ R,

which satisfies L|D = lc πD where πD : D → D/G is the projection.Also, the Lagrange d’Alembert equations induce well defined reduced La-grange d’Alembert equations on D/G. That is, the vector field on themanifold D determined by the Lagrange d’Alembert equations (includingthe constraints) is G-invariant, and so defines a reduced vector field on thequotient manifold D/G.

This proposition follows from general symmetry considerations. For ex-ample, to get the constrained reduced Lagrangian lc we restrict the givenLagrangian to the distribution D and then use its invariance to pass to thequotient. The problem of constrained Lagrangian reduction is the detaileddetermination of these reduced structures and will be dealt with later.

5.5 The Momentum Equation

In this section we use the Lagrange d’Alembert principle to derive an equa-tion for a generalized momentum as a consequence of the symmetries. Un-der the hypotheses that the action of some Lie algebra element is horizontal

130 5. Nonholonomic Mechanics

(that is, the infinitesimal generator is automatically in the constraint dis-tribution), this yields a conservation law in the usual sense. We refer thereader to our earlier section ? for the classical momentum map.

As we shall see, the momentum equation does not directly involve thechoice of an Ehresmann connection to describe the distribution D, but thechoice of such a connection will be useful for the coordinate versions.

We have already mentioned that simple physical systems that have sym-metries do not have associated conservation laws, namely the wobblestoneand the snakeboard. It is also easy to see why this is not generally the casefrom the equations of motion. The simplest situation would be the case ofcyclic variables. Recall that the equations of motion have the form

d

dt

∂Lc∂rα− ∂Lc∂rα

+Aaα∂Lc∂sa

= − ∂L∂sb

Bbαβ rβ .

If this has a cyclic variable, say r1, this would mean that all the quantitiesLc, L,B

bαβ would be independent of r1. This is equivalent to saying that

there is a translational symmetry in the r1 direction. Let us also suppose,as is often the case, that the s variables are also cyclic. Then the aboveequation for the momentum p1 = ∂Lc/∂r

1 becomes

d

dtp1 = − ∂L

∂sbBb1β r

β .

This fails to be a conservation law in general. Note that the right handside is linear in r (the first term is linear in pr) and the equation does notdepend on r1 itself. This is a very special case of the momentum equationthat we shall develop in this chapter. Even for systems like the snakeboard,the symmetry group is not abelian, so the above analysis for cyclic variablesfails to capture the full story. In particular, the momentum eqution is notof the preceding form in that example and thus it must be generalized.

5.5.1 The Derivation of the Momentum Equation

We now derive a generalized momentum map for nonholonomic systems.The number of equations obtained will equal the dimension of the intersec-tion of the orbit with the given constraints. As we will see, this result willgive conservation laws as a particular case.

To formulate our result, some additional ideas and notation will be useful.As the examples show, in general the tangent space to the group orbitthrough q intersects the constraint distribution at q nontrivially. It will behelpful to give this intersection a name.

Definition 5.5.1. The intersection of the tangent space to the group orbitthrough the point q ∈ Q and the constraint distribution at this point isdenoted Sq, as in Figure ??, and we let the union of these spaces over

5.5 The Momentum Equation 131

q ∈ Q be denoted S. Thus,

Sq = Dq ∩ Tq(Orb(q)).

Definition 5.5.2. Define, for each q ∈ Q, the vector subspace gq to bethe set of Lie algebra elements in g whose infinitesimal generators evaluatedat q lie in Sq:

gq = ξ ∈ g : ξQ(q) ∈ Sq

The corresponding bundle over Q whose fiber at the point q is given by gq,is denoted gD.

Consider a section of the vector bundle S over Q; i.e., a mapping thattakes q to an element of Sq = Dq ∩ Tq(Orb(q)). Assuming that the actionis free, a section of S can be uniquely represented as ξqQ and defines asection ξq of the bundle gD. For example, one can construct the sectionby orthogonally projecting (using the kinetic energy metric) ξQ(q) to thesubspace Sq. However, as we shall see, in later examples, it is often easy tochoose a section by inspection.

Next, we choose the variation analogously to what we chose in the caseof the standard Noether theorem above, namely, q(t, s) = exp(φ(t, s)ξq(t)) ·q(t). The corresponding infinitesimal variation is given by δq(t) = φ′(t)ξqQ(q(t)).Letting ∂ξq denote the derivative of ξq with respect to q, we have

δq = φ′ξq(t)Q + φ′[(Tξq(t)Q · q) + (∂ξq(t) · q)Q

].

In this equation, the term Tξq(t)Q is computed by taking the derivative of

the vector field ξq(t)Q with q(t) held fixed. By construction, the variation δqsatisfies the constraints and the curve q(t) satisfies the Lagrange d’Alem-bert equations, so that the following variational equation holds:

0 =∫ b

a

(∂L

∂qiδqi +

∂L

∂qiδqi)dt. (5.5.1)

In addition, the invariance identity from above holds using ξq:

0 =∫ b

a

(∂L

∂qi(ξq(t)Q )iφ′ +

∂L

∂qi(Tξq(t)Q · q)iφ′

)dt. (5.5.2)

Subtracting equations (5.5.1) and (5.5.2) and using the arbitrariness of φ′

and integration by parts shows that

d

dt

∂L

∂qi(ξq(t))iQ =

∂L

∂qi

[d

dt(ξq(t))

]iQ

.

The quantity whose rate of change is involved here is the nonholonomicversion of the momentum map in geometric mechanics.

132 5. Nonholonomic Mechanics

Definition 5.5.3. The nonholonomic momentum map Jnhc is thebundle map taking TQ to the bundle (gD)∗ whose fiber over the point q isthe dual of the vector space gq that is defined by

〈Jnhc(vq), ξ〉 =∂L

∂qi(ξQ)i.

where ξ ∈ gq. Intrinsically, this reads

〈Jnhc(vq), ξ〉 = 〈FL(vq), ξQ〉 ,

where FL is the fiber derivative of L and where ξ ∈ gq. For notationalconvenience, especially when the variable vq is suppressed, we will oftenwrite the left hand side of this equation as Jnhc(ξ).

Notice that the nonholonomic momentum map may be viewed as givingjust some of the components of the ordinary momentum map, namely alongthose symmetry directions that are consistent with the constraints.

We summarize these results in the following theorem.

Theorem 5.5.4. Assume that condition (L2) of definition 5.4.1 holds(which is implied by (L1)) and that ξq is a section of the bundle gD. Thenany solution of the Lagrange d’Alembert equations for a nonholonomic sys-tem must satisfy, in addition to the given kinematic constraints, the mo-mentum equation :

d

dt

(Jnhc(ξq(t))

)=∂L

∂qi

[d

dt(ξq(t))

]iQ

. (5.5.3)

When the momentum map is paired with a section in this way, we willjust refer to it as the momentum. The following is a direct corollary of thisresult.

Corollary 5.5.5. If ξ is a horizontal symmetry (see (S3) above), thenthe following conservation law holds:

d

dtJnhc(ξ) = 0 (5.5.4)

A somewhat restricted version of the momentum equation was given byKoslov & Kolesnikov [1978] and the corollary was given by Arnold[1988, page 82] (see Bloch & Crouch [1992, 1994] for the controlled case).

Remarks

1. The right hand side of the momentum equation (5.5.3) can be writtenin more intrinsic notation as⟨

FL(q(t)),(d

dtξq(t)

)Q

⟩.

5.5 The Momentum Equation 133

2. In the theorem and the corollary, we do not need to assume that thedistribution itself is G-invariant; that is, we do not need to assumecondition (S1). In particular, as we shall see in the examples, one canget conservation laws in some cases in which the distribution is notinvariant.

3. The validity of the form of the momentum equation is not affectedby any “internal forces”, that is, any control forces on shape space.Indeed, such forces would be invariant under the action of the Liegroup G and so would be annihilated by the variations taken to provethe above result.

4. The momentum equation still holds in the presence of affine con-straints. We do not need to assume that the affine vector field defin-ing the affine constraints is invariant under the group. However, thisvector field may appear in the final momentum equation (or conserva-tion law) because the constraints may be used to rewrite the resultingequation. We will see this explicitly in the example of the ball on arotating table.

5. Assuming that the distribution is invariant (hypothesis (S1)), thenonholonomic momentum map as a bundle map is equivariant withrespect to the action of the group G on the tangent bundle TQ andon the bundle (gD)∗. In fact, since the distribution is invariant andusing the general identity (Adgξ)Q = Φ∗g−1ξQ, valid for any groupaction, we see that the space gg is mapped to ggq by the map Adg,and so in this sense, the adjoint action acts in a well defined man-ner on the bundle gD. By taking its dual, we see that the coadjointaction is well defined on (gD)∗. In this setting, equivariance of thenonholonomic momentum map follows as in the usual proof (see, forexample, Marsden & Ratiu [1994], Chapter 11).

6. One can find an invariant momentum if the section is chosen suchthat

(Adg−1ξg·q)Q = ξqQ.

This can always be done in the case of trivial bundles; one chooses anyξq at the identity in the group variable and translates it around byusing the action to get a ξq at all points. This direction of reasoning(initiated by remarks of Ostrowski, Lewis, Burdick and Murray) isdiscussed in §5.5.3. As we will see later, this point of view is useful inthe case of the snakeboard.

7. The form of the momentum equation in this section is valid for anycurve q(t) that satisfies the Lagrange d’Alembert principle; we do notrequire that the constraints be satisfied for this curve. The version of

134 5. Nonholonomic Mechanics

the momentum equation given in the next section and later in §??will explicitly require that the constraints are satisfied. Of course, inexamples we always will impose the constraints, so this is really acomment about the logical structure of the various versions of theequation.

8. In some interesting cases, one can get conservation laws without hav-ing horizontal symmetries, as required in the preceding corollary.These are cases in which, for reasons other than horizontality, theright hand side of the momentum equation vanishes. This may be animportant observation for the investigation of completely integrablenonholonomic systems. A specific case in which this occurs is thevertical rolling disk discussed below.

5.5.2 The Momentum Equation in a Moving Basis

There are several ways of rewriting the momentum equation that are useful;the examples will show that each of them can reveal interesting aspectsof the system under consideration. This subsection develops the first ofthese coordinate formulas, which is in some sense the most naive, but alsothe most direct. The next subsection will develop a form that is suitablefor a local trivialization of the bundle Q → Q/G. Later on, when thenonholonomic connection is introduced, we shall come back to both of theseforms and rewrite them in a more sophisticated but also more revealing way.

Introduce coordinates q1, . . . qn in the neighborhood of a given point q0

in Q. At the point q0, introduce a basis

e1, e2, . . . , em, em+1, . . . , ekof the Lie algebra such that the first m elements form a basis of gq0 . Thus,k = dim g and m = dim gq, which, by assumption, is locally constant. Wecan introduce a similar basis

e1(q), e2(q), . . . , em(q), em+1(q), . . . , ek(q)at neighboring points q. For example, one can choose an orthonormal basis(in either the locked inertia metric or relative to a Killing form) that variessmoothly with q. We introduce a change of basis matrix by writing

eb(q) =k∑a=1

ψab (q)ea

for b = 1, . . . , k. Here, the change of basis matrix ψab (q) is an invertiblek × k matrix. Relative to the dual basis, we write the components of thenonholonomic momentum map as Jb. By definition,

Jb =n∑i=1

∂L

∂qi[eb(q)]iQ.

5.5 The Momentum Equation 135

Using this notation, the momentum equation, with the choice of sectiongiven by

ξq(t) = eb(q(t)), 1 ≤ b ≤ m

reads as follows:

d

dtJb =

n∑i=1

(∂L

∂qi

[d

dteb(q(t))

]iQ

). (5.5.5)

Next, we define Christoffel-like symbols by

Γcbl =k∑a=1

(ψ−1)ca∂ψab∂ql

(5.5.6)

where the matrix (ψ−1)da denotes the inverse of the matrix ψab . Observethat

d

dteb(q(t)) =

k∑c=1

n∑l=1

Γcblqlec(q(t)), (5.5.7)

which implies that[d

dteb(q(t))

]iQ

=k∑c=1

n∑l=1

Γcblql[ec(q(t))]iQ. (5.5.8)

Thus, we can write the momentum equation as

d

dtJb =

k∑c=1

n∑i,l=1

∂L

∂qiΓcblq

l[ec(q(t))]iQ. (5.5.9)

Introducing the shorthand notation eic := [ec(q(t))]iQ, the momentum equa-tion reads

d

dtJb =

k∑c=1

n∑i,l=1

∂L

∂qiΓcblq

leic (5.5.10)

Breaking the summation over c into two ranges and using the definition

Jc =∂L

∂qieic 1 ≤ c ≤ m

gives the following form of the momentum equation.

Proposition 5.5.6 (Momentum equation in a moving basis). Themomentum equation in the above coordinate notation reads

d

dtJb =

m∑c=1

n∑l=1

ΓcblJcql +

k∑c=m+1

n∑i,l=1

∂L

∂qiΓcblq

leic. (5.5.11)

136 5. Nonholonomic Mechanics

Assuming that the Lagrangian is of the form kinetic minus potentialenergy, the second term on the right hand side of this equation vanishes ifthe orbit and the constraint distribution are orthogonal; that is, if we canchoose the basis so that the vectors [ec(q(t))]Q for c ≥ m+1 are orthogonalto the constraint distribution. In this case, the momentum equation hasthe form of an equation of parallel transport along the curve q(t). Theconnection involved is the natural one associated with the the bundle (gD)∗

over Q, using a chosen decomposition of g, such as the orthogonal one.In the general case, the momentum equation is an equality between thecovariant derivative of the nonholonomic momentum and the last term onthe right hand side of the preceding equation. In the next section, we shallwrite the momentum equation in a body frame, which will be importantfor understanding how to decouple the momentum equation from the groupvariables.

5.5.3 The Momentum Equation in BodyRepresentation

Next, we develop an alternative coordinate formula for the momentumequation that is adapted to a choice of local trivialization. Thus, let a localtrivialization be chosen on the principal bundle π : Q → Q/G, with thelocal representation having coordinates denoted (r, g). Let r have compo-nents denoted rα as before, being coordinates on the base Q/G and let g begroup variables for the fiber, G. In such a representation, the action of Gis the left action of G on the second factor. We calculate the nonholonomicmomentum map using well known ideas (see, for example, Marsden &Ratiu [1994], Chapter 12), as follows. Let vq = (r, g, r, g) be a tangentvector at the point q = (r, g), η ∈ gq and let ξ = g−1g, i.e., ξ = TgLg−1 g.Since L is G-invariant, we can define a new function l by writing

L(r, g, r, g) = l(r, r, ξ).

Use of the chain rule shows that∂L

∂g= T ∗g Lg−1

∂l

∂ξ,

and so ⟨Jnhc(vq), η

⟩= 〈FL(r, g, r, g), ηQ(r, g)〉

=⟨∂L

∂g, (0, TRg · η)

⟩=⟨∂l

∂ξ,Adg−1 η

⟩.

The preceding equation shows that we can write the momentum map ina local trivialization by making use of the Ad mapping in much the same

5.5 The Momentum Equation 137

way as we did with the connection and the local formulas in the principalkinematic case. We define Jnhc

loc : TQ/G→ (gD)∗ in a local trivialization by

⟨Jnhc

loc (r, r, ξ), η⟩

=⟨∂l

∂ξ, η

⟩.

Thus, as with the previous local forms, Jnhc and its version in a localtrivialization are related by the Ad map; precisely,

Jnhc(r, g, r, g) = Ad∗g−1Jnhcloc (r, r, ξ).

Secondly, choose a q-dependent basis ea(q) for the Lie algebra such thatthe first m elements span the subspace gq. In a local trivialization, this isdone in a very simple way. First, one chooses, for each r, such a basis atthe identity element g = Id, say

e1(r), e2(r), . . . , em(r), em+1(r), . . . , ek(r).

For example, this could be a basis such that the corresponding generatorsare orthonormal in the kinetic energy metric. (Keep in mind that the sub-spaces Dq and TqOrb need not be orthogonal but here we are choosing abasis corresponding only to the subspace TqOrb.) Define the body fixedbasis by

ea(r, g) = Adg · ea(r);

then the first m elements will indeed span the subspace gq provided thedistribution is invariant (condition (S1)). Thus, in this basis we have⟨

Jnhc(r, g, r, g), eb(r, g)⟩

=⟨∂l

∂ξ, eb(r)

⟩:= pb, (5.5.12)

which defines pb, a function of r, r and ξ. We are deliberately introducingthe new notation p for the momentum in body representation to signal itsspecial role. Note that in this body representation, the functions pb areinvariant rather than equivariant, as is usually the case with the momen-tum map. The time derivative of pb may be evaluated using the momentumequation (5.5.3). This gives

d

dtpb =

∂L

∂qi

[d

dteb(r, g)

]iQ

=

⟨(TgLg−1)∗

∂l

∂ξ,

[d

dt(Adg · eb(r))

]Q

=⟨∂l

∂ξ, [ξ, eb] +

∂eb∂rα

rα⟩.

We summarize the conclusion drawn from this calculation as follows.

138 5. Nonholonomic Mechanics

Proposition 5.5.7 (Momentum equation in body representation).The momentum equation in body representation on the principal bundleQ→ Q/G is given by

d

dtpb =

⟨∂l

∂ξ, [ξ, eb] +

∂eb∂rα

rα⟩. (5.5.13)

Moreover, the momentum equation in this representation is independent of,that is, decouples from, the group variables g.

In this representation, the variable ξ is related to the group variable g byξ = g−1g. In particular, in this representation, reconstruction of the groupvariable g can be done by means of the equation

g = gξ (5.5.14)

On the other hand, this variable ξ = g−1g, as in the case of the reducedEuler-Poincare equations, is not the vertical part of the velocity vector qrelative to the nonholonomic connection to be constructed in the next sec-tion. The vertical part is related to the variable ξ by a velocity shift andthis velocity shift will make the reconstruction equation look affine, as inthe case of the snakeboard (see Lewis, Ostrowski, Murray & Burdick[1994]). In that example, the decoupling of the momentum equation fromthe group variables played a useful role. We also recall (as in the exampleof the rigid body with rotors discussed in Marsden & Scheurle [1993])that it is often the shifted velocity and not ξ that diagonalizes the kineticenergy, so this shift is fundamental for a number of reasons. As we shallsee later, the same ideas in this section, combined with the calculations ofMarsden & Scheurle [1993] will show how to calculate the fully reducedequations.

In the above local trivialization form of the momentum equation, we maywrite the terms (∂eb/∂rα)rα in terms of a connection, as we did in derivingthe momentum equation in a moving basis.

Other noteworthy features of this form of the momentum equation arethe following direct consequences of the preceding proposition.

Corollary 5.5.8.

1. If eb, b = 1, . . . ,m are independent of r, then the momentum equationin body representation is equivalent to the Euler-Poincare equationsprojected to the subspace gq.

2. If g is abelian, then the momentum equations reduce to

d

dtpb =

⟨∂l

∂ξ,∂eb∂rα

rα⟩. (5.5.15)

5.5 The Momentum Equation 139

3. If g is abelian, or more generally, if the bracket of an element of gq

with one in g is annihilated by ∂l/∂ξ, and if eb, b = 1, . . . ,m, areindependent of r, then the quantities pb, b = 1, . . . ,m are constantsof motion.

Regarding the first item, see Marsden & Ratiu [1994] for a discussionof the Euler-Poincare equations; see also section?. In this case, the spatialform of the momentum is conserved, just as in the case of systems withholonomic constraints. The last case occurs for the vertical rolling penny.

140 5. Nonholonomic Mechanics

This is page 141Printer: Opaque this

6Controllability and Stabilizability forMechanical and NonholonomicSystems

6.1 Controllability, Accessibility andStabilizability for NonholonomicMechanical Systems

In this section we consider a class of nonholonomic control systems andvarious control and stabilizability properties, following the work of Bloch,Reyhanoglu and McClamroch [1992]. Related work in robotics includes Liand Canny [1990], Li and Montgomery [1990], Murray and Sastry [1992],Krishnaprasad and Yang [1991].

We consider here the class of mechanical (Lagrangian) nonholonomiccontrol systems described by the equations

d

dt

∂qiL− ∂

∂qiL =

m∑j=1

λjaji +

l∑j=1

bjiuj , (6.1.1)

n∑i=1

aji qi = 0 j = 1 · · ·m. (6.1.2)

As before we may in general assume we have a Lagrangian on the tangentbundle to an arbitrary configuration space, Lagrangian L : TQ → R. Incoordinates qi, i = 1, . . . , n, on Q with induced coordinates (qi, qi) forthe tangent bundle, we have L(qi, qi). All computations here will be localhowever and for the moment we will assumme Q = Rn.

142 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Here L is taken to be the mechanical Lagrangian

L =12

n∑i,j=1

qiqj − V (q) . (6.1.3)

Hence equation 6.1.1 takes the explicit form

gij qj +

∂gij∂qk

qk qj − 12∂gjk∂qi

qj qk +∂V

∂qk

=m∑j=1

λjaji +

r∑j=1

bjiuj . (6.1.4)

For convenience below we shall sometimes rewrite equation 6.1.5 as

gij qj + fi(q, q) =

m∑j=1

λjaji +

l∑j=1

bjiuj . (6.1.5)

All functions are assumed to be smooth. We shall make some assumptionson the controls later on.

As in the previous chapter we shall assume that the contraints may berewritten as:

sa +Aaα(r, s)rα = 0 a = 1, . . . ,m, (6.1.6)

where q = (r, s) ∈ Rn−m × Rm.Denote as usual the distribution defined by the constraints at q by Dq.

Definition 6.1.1. (See Vershik and Gershkovitch [1988].) Consider thefollowing nondecreasing sequence of locally defined distributions

N1 = Dq ,

Nk = Nk−1 + span[X,Y ] | X ∈ N1, Y ∈ Nk−1 .

There exists an integer k∗ such that

Nk = Nk∗

for all k > k∗. If dim Nk∗ = n and k∗ > 1 then the constraints (2) arecalled completely nonholonomic and the smallest (finite) number k∗ is calledthe degree of nonholonomy.

We assume here that the constraint equations are completely nonholo-nomic everywhere with nonholonomy degree k∗. Note that for this to holdn−m must be strictly greater than one.

6.1 Controllability, Accessibility and Stabilizability for Nonholonomic Mechanical Systems 143

Under our assumptions the constraints define a (2n − m)-dimensionalsmooth submanifold

M = (q, q) | aji qi = 0 , j = 1 · · ·m (6.1.7)

of the phase space. This manifold M plays a critical role in the concept ofsolutions and the formulation of control and stabilization problems.

We note first that equations 6.1.1 and 6.1.2 provide a well posed set ofequations in the following sense:

Definition 6.1.2. A pair of vector functions (q(t), λ(t)) defined on aninterval [0, T ) is a solution of the initial value problem defined by equa-tions (6.1.1), (6.1.2) and the initial data (q0, q0) if q(t) is at least twicedifferentiable, λ(t) is integrable, the vector functions (q(t), λ(t)) satisfy thedifferential-algebraic equations (1)-(2) almost everywhere on their domainof definition and the initial conditions satisfy (q(0), q(0)) = (q0, q0).

Theorem 6.1.3. Assume that the control input function u : [0, T )→ Rl

is a given bounded and measurable function for some T > 0. If the initialdata satisfy (q0, q0) ∈M, then there exists a unique solution (at least locallydefined) of the initial value problem corresponding to equations (6.1.1),(6.1.2) which satisfies (q(t), q(t)) ∈M for each t for which the solution isdefined.

We subsequently use the notation (Q(t, q0, q0),Λ(t, q0, q0)) to denote thesolution of equations (6.1.1), (6.1.2) at time t ≥ 0 corresponding to theinitial conditions (q0, q0). Thus for each initial condition (q0, q0) ∈M andeach bounded, measurable input function u : [0, T )→ Rr, (Q(t, q0, q0), Q(t, q0, q0)) ∈M holds for all t ≥ 0 where the solution is defined.

We say a solution is an equilibrium solution if it is a constant solution;note that if (qe, λe) is an equilibrium solution we refer to qe as an equilib-rium configuration. The following result should be clear.

Theorem 6.1.4. Suppose that u(t) = 0, t ≥ 0. The set of equilibriumconfigurations of equations (1)-(2) is given by

qi | ∂V (q, 0)∂qi

− ajiλj = 0 , i = 1, · · · , n, for some λ ∈ Rm ,

We remark that generically we obtain an equilibrium manifold withdimension at least m. On the other hand, for certain cases, there may notbe even a single equilibrium configuration (e.g. the dynamics of a ball onan inclined plane). However, with our controllability assumptions below wecan always introduce an equilibrium manifold of dimension at least m byappropriate choice of input. This if course all in the smooth category.

144 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

6.2 Stabilization

6.2.1 Smooth Stabilization to Manifolds

It turns out that the best one can acheive in the way of smooth stabiliza-tion is stabilization to an equilibrium manifold See Bloch, Reyhanoglu andMcClamroch [1992], Montgomery [1993?]).

We want to consider feedback control of the form ui = Ui(q, q) whereU : M→ Rr; the corresponding closed loop is described by

gij qj + fi(q, q) =

m∑j=1

λjaji +

l∑j=1

bjiUj(q, q) . (6.2.1)

n∑i=1

aji qi = 0 j = 1 · · ·m. (6.2.2)

The set of equilibrium configurations of equations 6.2.1, 6.2.2 is given by

qi | ∂V (q, 0)∂qi

− ajiλj =l∑

j=1

bjiUj(q, 0) , i = 1, · · · , n, for some λ ∈ Rm ,

which is a smooth submanifold of the configuration space.We now introduce a suitable stability definition for the closed loop sys-

tem.

Definition 6.2.1. Assume that ui = Ui(q, q). Let Ms = (q, q)|q = 0be an embedded submanifold of M. Then Ms is locally stable if for anyneighborhood U ⊃Ms there is a neighborhood V of Ms with U ⊃ V ⊃Ms

such that if (q0, q0) ∈ V ∩M then the solution of equations 6.2.1,6.2.2satisfies (Q(t, q0, q0), Q(t, q0, q0)) ∈ U ∩M for all t ≥ 0. If, in addition,(Q(t, q0, q0), Q(t, q0, q0))→ (qs, 0) as t→∞ for some (qs, 0) ∈Ms then wesay that Ms is locally asymptotically stable equilibrium manifold of equa-tions 6.2.1, 6.2.2.

Note that if (Q(t, q0, q0), Q(t, q0, q0)) → (qs, 0) as t → ∞ for some(qs, 0) ∈ Ms, it follows that there is λs ∈ Rm such that Λ(t, q0, q0) → λsas t→∞.

The usual definition of local stability corresponds to the case that Ms

is a single equilibrium solution; the more general case is required in thepresent paper.

The existence of a feedback function so that a certain equilibrium man-ifold is asymptotically stable is of particular interest; hence we introducethe following

Definition 6.2.2. The system defined by equations 6.1.1, 6.1.2 is said tobe locally asymptotically stabilizable to a smooth equilibrium manifold Ms

6.3 Normal Form Equations 145

in M if there exists a feedback function U : M → Rl such that, for theassociated closed loop equations 6.2.1, 6.2.2, Ms is locally asymptoticallystable.

If there exists such a feedback function which is smooth on M then we saythat equations 6.1.1, 6.1.2 are smoothly asymptotically stabilizable to Ms;of course it is possible (and we subsequently show that it is generic in certaincases) that equations 6.1.1, 6.1.2 might be asymptotically stabilizable toMs but not smoothly (even continuous) asymptotically stabilizable to Ms.

6.3 Normal Form Equations

In this section we describe how to obtain a reduced normal form for thecontrolled nonholonomic equations.

We recall that the reduced state space is 2n−m dimensional. The stateof the system can be specified by the n-vector of configuration variablesand an (n−m)-vector of velocity variables. Let q = (r, s) be the partitionof the configuration variables corresponding to the constraints introducedpreviously.

Define the n× n−m matrix C as follows:

Ciα = δiα , i = 1, · · ·n−m= −Aiα , i = n−m+ 1, · · ·n (6.3.1)

Then

qi = Ciαrα (6.3.2)

Taking time derivatives yields

qi = Ciα(q)rα + Ciα(q)rα .

Substituting this into equation 6.1.1 and multiplying both sides of theresulting equation by CT (q) gives

CiTα (q)gij(q)Cjβ(q)rβ = CiTα (q)[bji (q)uj − fi(q, Cr)− gij(q)Cjα(q)rα] .

(6.3.3)

Note that CiTα (q)gij(q)Cjβ(q) is an (n-m)x(n-m) symmetric positive definite

matrix function.We also assume that l = n − m (for simplicity) and that the matrix

product CiTα (q)bji (q) is locally invertible. Consequently for any u ∈ Rl thereis unique v ∈ Rn−m which satisfies

CiTα (q)gij(q)Cjβ(q)vβ = CiTα (q)[bji (q)uj − fi(q, Cr)− gij(q)Cjα(q)rα] .

(6.3.4)

146 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

This assumption guarantees that the reduced configuration variables satisfythe linear equations

qα = vα .

Define the following state variables

xα1 = r ,

xa2 = qa2 ,

xα3 = rα .

Then the normal form equations are given by

x1 = x3 , (6.3.5)

x2 = −A(x1, x2)x3 , (6.3.6)

x3 = v . (6.3.7)

Equations 6.3.5, 6.3.6, 6.3.7 define a drift vector field f(x) = (x3,−A(x1, x2)x3, 0)and control vector fields gi(x) = (0, 0, ei), where ei is the i’th standard ba-sis vector in Rn−m, i = 1, · · · , n − m, according to the standard controlsystem form

x = f(x) +n−m∑i=1

gi(x)vi . (6.3.8)

We consider local properties of equations 6.3.5, 6.3.6, 6.3.7, near an equi-librium solution (xe1, x

e2, 0).

Note that the normal form equations 6.3.5, 6.3.6, 6.3.7 are a specialcase of the normal form equations studied by Byrnes and Isidori [1988]. Inparticular, the zero dynamics equation of 6.3.5, 6.3.6, 6.3.7, correspondingto the output x1, is given by

x2 = 0 ,

and it is not locally asymptotically stable. The fact that the zero dynam-ics is a linear system with all zero eigenvalues, means that equations 6.3.5,6.3.6, 6.3.7 are critically minimum phase at the equilibrium; this has impor-tant implications in terms of local asymptotic stabilizability of the originalequations 6.1.1, 6.1.2.

6.4 Stabilization to an Equilibrium Manifold 147

6.4 Stabilization to an EquilibriumManifold

In this section we study the problem of stabilization of equations 6.1.1,6.1.2to a smooth equilibrium submanifold of M defined by

Ne = (q, q) | q = 0 , s(q) = 0

where w(q) is a smooth n − m vector function. We show that, with ap-propriate assumptions, there exists a smooth feedback such that the closedloop is locally asymptotically stable to Ne.

The smooth stabilization problem is the problem of giving conditions sothat there exists a smooth feedback function U : M → Rl such that Ne

is locally asymptotically stable. Of course, we are interested not only indemonstrating that such a smooth feedback exists but also in indicatinghow such an asymptotically stabilizing smooth feedback can be constructed.

Note that in this section we consider nonholonomic control systems whosenormal form equations satisfy the property that if r(t) and r(t) are expo-nentially decaying functions, then the solution to

s = −A(r(t), s)r(t)

is bounded (all the physical examples of nonholonomic systems, of whichwe are aware, satisfy this assumption).

Note also that the first and second time derivatives of w(q) are given by

w =∂w(q)∂q

C(q)r ,

w =∂

∂q(∂w(q)∂q

C(q)s)C(q)r +∂w(q)∂q

C(q)v .

Theorem 6.4.1. Assume that the above solution property holds. Thenthe nonholonomic control system, defined by equations 6.1.1, 6.1.2 is locallyasymptotically stabilizable to

Ne = (q, q) | q = 0 , w(q) = 0 , (6.4.1)

using smooth feedback, if the transversality condition

det(∂w(q)∂r

) det(∂w(q)∂q

C(q)) 6= 0 (6.4.2)

is satisfied.

148 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Proof. It is sufficient to analyze the system in the normal form 6.3.5,6.3.6, 6.3.7. By the transversality condition, the change of coordinates from(r, s, r) to (w, s, w) is a diffeomorphism.

Let

v = −(∂w(q)∂q

C(q))−1[∂

∂q(∂w(q)∂q

C(q)r)C(q)r +K1∂w(q)∂q

C(q)r +K2w(q)] ,

where K1 and K2 are symmetric positive definite (n−m)×(n−m) constantmatrices. Then, obviously

w +K1w +K2w = 0

is asymptotically stable so that (w, w) → 0 as t → ∞. The remainingsystem variables satisfy equation 6.3.5 of the normal form equations (withx2 = s), and, by our assumption on the on the constraint matrix A, thesevariables remain bounded for all time. Thus (q(t), q(t)) → Ne as t →∞. ¥

6.4.1 Nonsmooth Stabilization

The results in the previous section demonstrate that smooth feedback canbe used to asymptotically stabilize certain smooth manifolds Ne in M,where the dimension of Ne is equal to the number m of independent con-straints. These results do not guarantee smooth asymptotic stabilizationto a single equilibrium solution if m ≥ 1.

In fact, there is no C1 feedback which can asymptotically stabilize theclosed loop system to a single equilibrium solution. For suppose that thereis a C1 feedback which asymptotically stabilizes, for example, the origin.Then it follows that there is an equilibrium manifold of dimension m con-taining the origin; that is, the origin is not isolated, which contradicts theassumption that it is asymptotically stable. We state this formally.

Theorem 6.4.2. Let m ≥ 1 and let (qe, 0) denote an equilibrium solutionin M. The nonholonomic control system, defined by equations 6.1.1, 6.1.2is not asymptotically stabilizable using C1 feedback to (qe, 0).

Proof. A necessary condition for the existence of a C1 asymptoticallystabilizing feedback law for system 6.3.5, 6.3.6, 6.3.7 is that the image ofthe mapping

(x1, x2, x3, v) 7−→ (x3, −A(x1, x2)x3, v)

contains some neighborhood of zero (see Brockett [1983]). No points of theform 0

εα

, ε 6= 0 and α ∈ Rn−m arbitrary ,

6.4 Stabilization to an Equilibrium Manifold 149

are in its image: it follows that Brockett’s necessary condition is not sat-isfied. Hence system 6.3.5, 6.3.6, 6.3.7 cannot be asymptotically stabilizedto (re, se, 0) by a C1 feedback law. Consequently, the nonholonomic con-trol system, defined by equations (6.1.1), (6.1.2) is not C1 asymptoticallystabilizable to (qe, 0). ¥

We remark that even C0 (continuous) feedback (which results in exis-tence of unique trajectories) is ruled out since Brockett’s necessary condi-tion is not satisfied (Zabcyzk [1989]).

A corollary of Theorem 4 is that a single equilibrium solution of 6.1.1,6.1.2 cannot be asymptotically stabilized using linear feedback nor canit be asymptotically stabilized using feedback linearization or any othercontrol design approach that uses smooth feedback. Below we indicate howa single equilibrium can be asymptotically stabilized by use of piecewiseanalytic feedback. This is the by no means the only approach however tostabilization. We discuss other approaches later, and as we discussed earlier,there is a large literature on this subject.

We first demonstrate that the normal form equations 6.3.5, 6.3.6, 6.3.7and hence the nonholonomic control system defined by equations 6.1.1,6.1.2 satisfies certain strong local controllability properties. In particular,we show that the system is strongly accessible and that the system is smalltime locally controllable at any equilibrium. These results provide a theoret-ical basis for the use of inherently nonlinear control strategies and suggestconstructive procedures for the desired control strategies.

Theorem 6.4.3. Let m ≥ 1 and let (qe, 0) denote an equilibrium solutionin M. The nonholonomic control system, defined by equations 6.1.1, 6.1.2is strongly accessible at (qe, 0).

Proof. It suffices to prove that system 6.3.5, 6.3.6, 6.3.7 is strongly ac-cessible at the origin. Let I denote the set 1, · · · , n −m. The drift andcontrol vector fields can be expressed as

f =n−m∑j=1

x3,jτj ,

gi =∂

∂x3,i, i ∈ I ,

where

τj =∂

∂x1,j−n−m∑i=1

Aji (x1, x2)∂

∂x2,i, j ∈ I

are considered as vector fields on the (x1, x2, x3) state space. It can beverified that

[gi1 , f ] = τi1 , i1 ∈ I ;

150 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

[gi2 , [f, [gi1 , f ]]] = [τi2 , τi1 ], i1, i2 ∈ I ;

...

[gik∗ , [f, · · · , [gi2 , [f, [gi1 , f ]]] · · · ]] = [τik∗ , · · · , [τi2 , τi1 ] · · · ], ik ∈ I, 1 ≤ k ≤ k∗ ,

hold, where k∗ denotes the nonholonomy degree. Let

G = spangi, i ∈ I ,

H = span[gi1 , f ], · · · , [gik∗ , [f, · · · , [gi2 , [f, [gi1 , f ]]] · · · ]]; ik ∈ I, 1 ≤ k ≤ k∗ .

Note that dimG(0) = n −m and dimH(0) = n since the distribution de-fined by the constraints is completely nonholonomic; moreover dimG(0)∩H(0) = 0. It follows that the strong accessibility distribution

L0 = spanX : X ∈ G ∪H

has dimension 2n − m at the origin. Hence the strong accessibility rankcondition( Sussmann and Jurdjevich [1972]) is satisfied at the origin. Thussystem 6.3.5, 6.3.6, 6.3.7 is strongly accessible at the origin Hence thenonholonomic control system 6.1.1, 6.1.2 is strongly accessible at (qe, 0).

¥

Theorem 6.4.4. Let m ≥ 1 and let (qe, 0) denote an equilibrium solutionin M. The nonholonomic control system, defined by equations 6.1.1, 6.1.2is small time locally controllable at (qe, 0).

Proof. It suffices to prove that system 6.3.5, 6.3.6, 6.3.7 is small timelocally controllable to the origin.

The proof involves the notion of the degree of a bracket. To make thisnotion well defined we consider, as in Sussmann [1987], a Lie algebra ofindeterminates and an associated evaluation map (on vector fields) as fol-lows:

Let X = (X0, · · · , Xn−m) be a finite sequence of indeterminates. LetA(X) denote the free associative algebra over R generated by the Xj , letL(X) denote the Lie subalgebra of A(X) generated by X0, · · · , Xn−m andlet Br(X) be the smallest subset of L(X) that contains X0, · · · , Xn−m andis closed under bracketing.

Now consider the vector fields f, g1, · · · , gn−m on the manifold M. Eachf, g1, · · · ,gn−m is a member of D(M), the algebra of all partial differential operatorson C∞(M), the space of C∞ real-valued functions on M. Now let g0 = f ,and let g = (g0, · · · , gn−m) and define the evaluation map

Ev(g) : A(X)→ D(M)

6.4 Stabilization to an Equilibrium Manifold 151

obtained by substituting the gj for the Xj , i.e.

Ev(g)(∑I

aIXI) =∑I

aIgI

where gI = gi1gi2 · · · gik , I = (i1, · · · , ik). Note that the kernel of Ev(g) :A(X) → A(g) is the set of all algebraic identities satisfied by the gi whilethe kernel of Ev(g) : L(X) → L(g) is the set of Lie algebraic identitiessatisfied by gi.

Now, let B be a bracket in Br(X). We define the degree of a bracketto be δ(B) =

∑n−mi=0 δi(B), where δ0(B), δ1(B), · · · , δn−m(B) denote the

number of times X0, · · · , Xn−m, respectively, occur in B. The bracket B iscalled “bad” if δ0(B) is odd and δi(B) is even for each i, i = 1, · · · , n−m.The theorem of Sussmann tells us the system is STLC at the origin ifit satisfies the accessibility rank condition; and if B is “bad” there existbrackets C1, · · · , Ck of lower degree in Br(X) such that

Ev0(g)(β(B)) =k∑i=1

ξiEv0(g)(Ci)

where Ev0 denotes the evaluation map at the origin and (ξ1, · · · , ξk) ∈Rk. Here β(B) is the symmetrization operator, β(B) =

∑π∈Sn−m π(B),

where π ∈ Sn−m, the group of permutations of 1, · · · , n − m and forπ ∈ Sn−m, π is the automorphism of L(X) which fixes X0 and sends Xi

to Xπ(i).By Theorem 5, the system is accessible at the origin.The brackets in G are obviously “good” (not of the type defined as

“bad”) and δ0(h) =∑n−mj=1 δj(h) ∀h ∈ H; thus δ(h) is even for all h in

H, i.e H contains “good” brackets only. It follows that the tangent spaceT0M to M at the origin is spanned by the brackets that are all “good”.Next we show that the brackets that might be “bad” vanish at the origin.First note that f vanishes at the origin. Let B denote a bracket satisfyingδ(B) > 1. If B is a “bad” bracket then necessarily δ0(B) 6=

∑n−mj=1 δj(B),

i.e. δ(B) must be odd. It can be verified that if δ0(B) <∑n−mj=1 δj(B)

then B is identically zero and if δ0(B) >∑n−mj=1 δj(B) then B is of the

form∑n−mi=1 ri(x3)Yi(x1, x2), for some vector fields Yi(x1, x2), i ∈ I, where

ri(x3), i ∈ I, are homogeneous functions of degree (δ0(B)−∑n−mj=1 δj(B))

in x3; thus B vanishes at the origin. Consequently the Sussmann conditionis satisfied and the system is small time locally controllable. ¥

Piecewise Analytic Stabilizing Controllers

We consider now stabilization for so-called “controlled nonholonomic Caply-gin systems”.

152 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

We first describe the class of controlled nonholonomic Caplygin systems.If the functions used in defining equations 6.1.1, 6.1.2 do not depend explic-itly on the configuration variables s, so that the system is locally describedby

gij(r)qj + fi(r, q) =m∑j=1

λjaji (r) +

l∑j=1

bji (r)uj . (6.4.3)

sa +Aaα(r)rα = 0 a = 1, . . . ,m , (6.4.4)

then the uncontrolled system is called a “nonholonomic Caplygin system”by Neimark and Fufaev [1972]. In terms of the Lagrangian formalism forthe problem this corresponds to the Lagrangian of the free problem beingcyclic in (i.e. independent of) the variables q2 while the constraints arealso independent of q2. The cyclic property is an expression of symmetriesin the problem as we have discussed. More generally, if a system can beexpressed in the form 6.4.3, 6.4.4 using feedback, then we refer to it as a“controlled nonholonomic Caplygin system”.

For the nonholonomic Caplygin system described by equations (16)-(17),equation 6.3.3 becomes

CiTα (r)gij(r)Cjβ(r)rβ = CiTα (r)[bji (r)uj − fi(r, Cr)− gij(r)Cjα(r)rα] .

(6.4.5)

which is an equation in the variables (r, r) only. As a consequence,the rvariables coordinatize a reduced configuration space for the system 6.4.3,6.4.4. This reduced configuration space is also referred to as the “basespace” (or “shape space”) of the system. The term shape space arises fromthe theory of coupled mechanical systems, where it refers to the internaldegrees of freedom of the system.

As we did earlier, we assume that r = n−m and that the matrix productCiTα (q)bji (q)is locally invertible; this assumption is not restrictive. Conse-quently, it can be shown that the normal form equations for the system6.4.3, 6.4.4 are given

x1 = x3 , (6.4.6)

x2 = −A(x1)x3 , (6.4.7)

x3 = v , (6.4.8)

where x1 = r, x2 = r, x3 = s and v satisfies

CiTα (r)gij(r)Cjβ(r)vβ = CiTα (r)[bji (r)uj − fi(r, Cr)− gij(r)Cjα(r)rα] .

(6.4.9)

6.4 Stabilization to an Equilibrium Manifold 153

Again we shall make use of these normal form equations to obtain controlresults.

Clearly, there is no continuous feedback which asymptotically stabilizesa single equilibrium. However, the controllability properties possessed bythe system guarantee the existence of a piecewise analytic feedback (seeSussmann [1979]). We now describe the ideas that are employed to con-struct such a feedback which does achieve the desired local asymptoticstabilization of a single equilibrium solution. These ideas are based on theuse of holonomy (geometric phase) which has proved useful in a variety ofkinematics and dynamics problems (see e.g. Krishnaprasad [1991], Shapereand Wilczek [1988], and Marsden, Montgomery and Ratiu [1990]. The keyobservation here is that the holonomy, the extent to which a closed pathin the base space fails to be closed in the configuration space, dependsonly on the path traversed in the base space and not on the time historyof traversal of the path. Related ideas have been used for a class of pathplanning problems, based on kinematic relations, in Li and Canny [1990],Li and Montgomery [1990], and Krishnaprasad and Yang [1991].

For simplicity, we consider control strategies which transfer any initialconfiguration and velocity(sufficiently close to the origin) to the zero config-uration with zero velocity. The proposed control strategy initially transfersthe given initial configuration and velocity to the origin of the (q1, q1) basephase space. The main point then is to determine a closed path in theq1 base space that achieves the desired holonomy. We show that the indi-cated assumptions guarantee that this holonomy construction can be madeand that (necessarily piecewise analytic) feedback can be determined whichaccomplishes the desired control objective.

Let x0 = (x01, x

02, x

03) denote an initial state. We now describe two steps

involved in construction of a control strategy which transfers the initialstate to the origin.Step 1: Bring the system to the origin of the (x1, x3) base phase space, i.e.find a control which transfers the initial state (x0

1, x02, x

03) to (0, x1

2, 0) in afinite time, for some x1

2.Step 2: Traverse a closed path (or a series of closed paths) in the x1 basespace to produce a desired holonomy in the (x1, x2) configuration space, i.e.find a control which transfers (0, x1

2, 0) to (0, 0, 0).The desired holonomy condition is given by

x12 =

∮γ

A(x1) dx1 , (6.4.10)

where γ denote a closed path traversed in the base space. The holonomyis reflected in the fact that traversing a closed path in the base spaceyields a nonclosed path in the full configuration space. Note that here,for notational simplicity in presenting the main idea, we assume that thedesired holonomy can be obtained by a single closed path. In general, more

154 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

than one loop may be required to produce the desired holonomy; for suchcases γ can be viewed as concatenation of a series of closed paths.

Under the weak assumptions mentioned previously, explicit procedurescan be given for each of the above two steps. Step 1 is classical; it is step 2,involving the holonomy, that requires special consideration. Explicit charac-terization of a closed path γ which satisfies the desired holonomy equation(23) can be given for several specific examples (see below). However, someproblems may require a general computational approach. An algorithmbased on Lie algebraic methods as in Lafferriare and Sussmann [1991] canbe employed to aproximately characterize the required closed path. Sup-pose the closed path γ which satisfies the desired holonomy condition ischosen. Then a feedback algorithm which realizes the closed path in thebase space can be constructed since the base space equations represent n-mdecoupled double integrators.

This general construction procedure provides a strategy for transferringan arbitrary initial state of equations 6.4.6, 6.4.7, 6.4.8 to the origin. Im-plementation of this control strategy in a (necessarily piecewise analytic)feedback form can be accomplished as follows.

Let a = (a1, · · · , an−m) and b = (b1, · · · , bn−m) denote displacementvectors in the x1 base space and let γ(a, b) denote the closed path (in thebase space) formed by the line segments from x1 = 0 to x1 = a, fromx1 = a to x1 = a+ b, from x1 = a+ b to x1 = b, and from x1 = b to x1 = 0.Then the holonomy of the parameterized family

γ(a, b)|a, b ∈ Rn−m

is determined by the holonomy function γ(a, b) 7−→ α(a, b) given as

α(a, b) = −∮γ(a,b)

A(x1) dx1 .

Now let πs denote the projection map πs : (x1, x2, x3) 7−→ (x1, x3). Inorder to construct a feedback control algorithm to accomplish the abovetwo steps, we first define a feedback function V x

∗1 (πsx) which satisfies: for

any πsx(t0) there is t1 ≥ t0 such that the unique solution of

x1 = x3 ,

x3 = V x∗1 (πsx) ,

satisfies πsx(t1) = (x∗1, 0). Note that the feedback function is parameterizedby the vector x∗1. Moreover, for each x∗1, there exists such a feedback func-tion. One such feedback function V x

∗1 (πsx) = (V x

∗1

1 (πsx), · · · , V x∗1

n−m(πsx))is given as

V x∗

i (πsx) =−ki sign(x1,i − x∗1,i + x3,i|x3,i|/2ki) , (x1,i, x3,i) 6= (x∗1,i, 0) ,0 , (x1,i, x3,i) = (x∗1,i, 0) ,

6.4 Stabilization to an Equilibrium Manifold 155

where ki, i = 1, · · · , n − m, are positive constants chosen such that theresulting motion, when projected to the base space, constitute a straightline connecting x1(t0) to x1(t1) = x∗1.

We specify the control algorithm, with values denoted by v∗, accordingto the following construction, where x denotes the “current state” :

Control algorithm for v∗ :

Step 0: Choose (a∗, b∗) to achieve the desired holonomy.

Step 1: Set v∗ = V a∗(πsx), until πsx = (a∗, 0); then go to Step 2;

Step 2: Set v∗ = V a∗+b∗(πsx), until πsx = (a∗ + b∗, 0); then go to Step 3;

Step 3: Set v∗ = V b∗(πsx), until πsx = (b∗, 0); then go to Step 4;

Step 4: Set v∗ = V 0(πsx), until πsx = (0, 0); then go to Step 0;

We here assumed that the desired holonomy can be obtained by a singleclosed path. Clearly the above algorithm can be modified to account forcases for which more than one closed path is required to satisfy the desiredholonomy.

Note that the control algorithm is constructed by appropriate switchingsbetween members of the parameterized family of feedback functions. Oneach cycle of the algorithm the particular functions selected depend on theclosed path parameters a∗, b∗, computed in Step 0, to correct for errors inx2.

The control algorithm can be initialized in different ways. The mostnatural is to begin with Step 4 since v∗ in that step does not dependon the closed path parameters; however, many other initializations of thecontrol algorithm are possible.

Justification that the constructed control algorithm asymptotically stabi-lizes the origin follows as a consequence of the construction procedure: thatswitching between feedback functions guarantees that the proper closedpath (or a sequence of closed paths) is traversed in the base space so thatthe origin (0, 0, 0) is necessarily reached in a finite time. This construc-tion of a stabilizing feedback algorithm represents an alternative to theapproach by Hermes [1980], which is based on Lie algebraic properties.

It is important to emphasize that the above construction is based on thea priori selection of simply parametrized closed paths in the base space.The above selection simplifies the tracking problem in the base space, butother path selections could be made and they would, of course, lead to adifferent feedback strategy from that proposed above.

We remark that the technique presented in this section can be generalizedto some systems which are not Caplygin. For instance, this generalizationis tractible to systems for which equation (20) takes the form

x2 = ρ(x2)A(x1)

156 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

where ρ(x2) denotes certain Lie group representation (see e.g. Marsden,Montgomery and Ratiu [1990]). The holonomy of a closed path for suchsystems is given as a path ordered exponential rather than a path integral.

Examples

Control of Knife Edge Using Steering and Pushing Inputs : We first considerthe control of a knife edge moving in point contact on a plane surface.. Letx and y denote the coordinates of the point of contact of the knife edge onthe plane and let φ denote the heading angle of the knife edge, measuredfrom the x-axis. Then the equations of motion, with all numerical constantsset to unity, are given by

x = λ sinφ+ u1 cosφ , (6.4.11)

y = −λ cosφ+ u1 sinφ , (6.4.12)

φ = u2 , (6.4.13)

where u1 denotes the control force in the direction defined by the headingangle, u2 denotes the control torque about the vertical axis through thepoint of contact; the components of the force of constraint arise from thescalar nonholonomic constraint

x sinφ− y cosφ = 0 (6.4.14)

which has nonholonomy degree two at any configuration. It is clear thatthe constraint manifold is a five-dimensional manifold and is defined by

M = (φ, x, y, φ, x, y)|x sinφ− y cosφ = 0

and any configuration is an equilibrium if the controls are zero.Define the variables

x1 = x cosφ+ y sinφ ,

x2 = φ ,

x3 = −x sinφ+ y cosφ ,

x4 = x cosφ+ y sinφ− φ(x sinφ− y cosφ) ,

x5 = φ ,

6.4 Stabilization to an Equilibrium Manifold 157

so that the reduced differential equations are given by

x1 = x4 ,

x2 = x5 ,

x3 = −x1x5 ,

x4 = u1 + u2x3 − x1x25 ,

x5 = u2 .

We have:

Proposition 6.4.5. Let xe = (xe1, xe2, x

e3, 0, 0) denote an equilibrium so-

lution of the reduced differential equations corresponding to u = 0. Theknife edge dynamics described by equations (24)-(27) have the followingproperties:

1. There is a smooth feedback which asymptotically stabilizes the closedloop to any smooth one dimensional equilibrium manifold in M whichsatisfies the transversality condition.

2. There is no smooth feedback which asymptotically stabilizes xe.

3. The system is strongly accessible at xe since the space spanned by thevectors

g1, g2, [g1, f ], [g2, f ], [g2, [f, [g1, f ]]]

has dimension 5 at xe.

4. The system is small time locally controllable at xe since the bracketssatisfy sufficient conditions for small time local controllability.

Note that the base variables are (x1, x2). Consider a parameterized rect-angular closed path γ in the base space with four corner points

(0, 0), (x1, 0), (x1, x2), (0, x2) ,

i.e. a = (x1, 0) and b = (0, x2) following the notation introduced in thegeneral development. By evaluating the holonomy integral in closed formfor this case, the desired holonomy equation is

x13 = x1x2 .

158 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

This equation can be explicitly solved to determine a closed path γ∗ =γ(a∗, b∗) which achieves the desired holonomy. One solution can be givenas follows

a∗ = (√|x1

3|signx13, 0), b∗ = (0,

√|x1

3|) .

Note that the previously described feedback algorithm can be used toasymptotically stabilize the knife edge to the origin.

Control of Rolling Wheel : As a second example, we consider again thecontrol of a vertical wheel rolling without slipping on a plane surface. Asbefore, Let x and y denote the coordinates of the point of contact of thewheel on the plane, let φ denote the heading angle of the wheel, measuredfrom the x- axis and let θ denote the rotation angle of the wheel due torolling, measured from a fixed reference. Then the equations of motion,with all numerical constants set to unity, are given by

x = λ1 , (6.4.15)

y = λ2 , (6.4.16)

θ = −λ1 cosφ− λ2 sinφ+ u1 , (6.4.17)

φ = u2 , (6.4.18)

where u1 denotes the control torque about the rolling axis of the wheel andu2 denotes the control torque about the vertical axis through the pointof contact; the components of the force of constraint arise from the twononholonomic constraints

x = θ cosφ , (6.4.19)

y = θ sinφ , (6.4.20)

which have nonholonomy degree three at any configuration. The constraintmanifold is a six-dimensional manifold and is given by

M = (θ, φ, x, y, θ, φ, x, y)|x = θ cosφ, y = θ sinφ

and any configuration is an equilibrium if the controls are zero.Define the variables

x1 = θ, x2 = φ, x3 = x, x4 = y, x5 = θ, x6 = φ ,

so that the reduced differential equations are given by

x1 = x5 ,

6.4 Stabilization to an Equilibrium Manifold 159

x2 = x6 ,

x3 = x5 cosx2 ,

x4 = x5 sinx2 ,

x5 =12u1 ,

x6 = u2 .

We have a result similar to that in the previous example:

Proposition 6.4.6. Let xe = (xe1, xe2, x

e3, x

e4, 0, 0) denote an equilibrium

solution of the reduced differential equations corresponding to u = 0. Therolling wheel dynamics have the following properties:

1. There is a smooth feedback which asymptotically stabilizes the closedloop to any smooth two dimensional equilibrium manifold in M whichsatisfies the transversality condition.

2. There is no smooth feedback which asymptotically stabilizes xe.

3. The system is strongly accessible at xe since the space spanned by thevectors

g1, g2, [g1, f ], [g2, f ], [g2, [f, [g1, f ]]], [g2, [f, [g1, [f, [g2, f ]]]]]

has dimension 6 at xe.

4. The system is small time locally controllable at xe since the bracketssatisfy sufficient conditions for small time local controllability.

Note that the base variables are (x1, x2). Consider a parameterized rect-angular closed path γ in the base space with four corner points

(0, 0), (x1, 0), (x1, x2), (0, x2) .

By evaluating the holonomy integral in closed form for this case, we findthe desired holonomy equations are

x13 = x1(cosx2 − 1) ,

x14 = x1 sinx2 .

These equations can be explicitly solved to determine a closed path (ora concatenation of closed paths) γ∗ which achieves the desired holonomy.

160 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

One solution can be given as follows: if x13 6= 0 then γ∗ is the closed path

specified by

a∗ = −((x13)2 + (x1

4)2)/2x13, 0) , b∗ = (0,− sin−1(2x1

3x14/((x

13)2 + (x1

4)2)) ;

and if x13 = 0 then γ∗ is a concatenation of two closed paths specified by

a∗ = (0.5x14, 0), b∗ = (0, 0.5π) ,

a∗∗ = (−0.5x14, 0), b∗∗ = (0,−0.5π) .

Note that the previously described feedback algrithm can be used (withthe modification indicated in the general development) to asymptoticallystabilize the rolling wheel to the origin.

6.5 Controlled Nonholonomic Systems onManifolds

We consider here the formulation of controlled classical nonholonomic sys-tems on Riemannian manifolds. Such a formulation is useful for consid-ering systems with nontrivial geometry, such as the rolling ball. We shalldistinguish again here between variational nonholonomic systems and non-holonomic systems. The latter is normally used for describing mechanicalsystems and here we formalize the introduction of controls for such systems.

6.5.1 Holonomic Control Systems

Consider firstly the holonomic or unconstrained case:Let (M, 〈, 〉) be an n-dimensional Riemannian manifold, with metric

g( , ) = 〈 , 〉. Denote the norm of a tangent vector X at the pointp by ‖Xp‖ = 〈Xp, Xp〉 12 . The geodesic flow on M is then given by

D2q

dt2= 0 (6.5.1)

where Dqdt denotes the covariant derivative. This flow minimizes the integral∫ 1

0

∥∥∥Dqdt ∥∥∥2

dt along parametrized paths.We define a controlled holonomic system to be a system of the form

D2q

dt2=

N∑i=1

uiXi (6.5.2)

where Xi is an arbitrary set of control vector fields, the ui are functionsof time, and N ≤ n . (Note that in this paper we do not consider systemsevolving under the influence of a potential.)

6.5 Controlled Nonholonomic Systems on Manifolds 161

It is natural to pose the following optimal control problem in the caseXi is an orthonormal set.

minu

∫ T

0

‖u‖2dt = minu

∫ T

0

N∑i=1

u2i dt (6.5.3)

= minq

∫ T

0

∥∥∥∥D2q

dt2

∥∥∥∥2

dt

subject to (2) and satisfying q(0) = q0, q(T ) = qT . Such higher ordervariational problems (see Griffiths [1983]) were considered in the contextof interpolation problems by Noakes et. al. [1989] and by Crouch and Silva-Leite [1991].

We remark that these problems are different from the optimal controlproblems

minu

∫‖u‖2ds

subject to

Dq

dt=

N∑i=1

uiXi. (6.5.4)

However they do have the same singular nature for N < n as consideredby Brockett in the celebrated paper Brockett [1982]. Indeed, later on, wewill interpret the latter systems as kinematic nonholonomic systems.

6.5.2 Nonholonomic Systems

We now consider the formulation of controlled nonholonomic systems. Wecan think of the two classes: – variational nonholonomic systems (see e.g.Arnold [1988] and Vershik and Gershkovich [1988]) and classical (or me-chanical) nonholonomic systems. In either case, the constraints are givenby 1-forms ωk, 1 ≤ k ≤ m, which define a (smooth) distribution H on M .(H is the set of subspaces of TM annihilated by the ωk at each x ∈ TxM .)

Variational nonholonomic systems are obtained in our context from thevariational problem

minq

∫ 1

0

∥∥∥∥Dqdt∥∥∥∥2

subject to

ωk

(Dq

dt

)= 0

1 ≤ k ≤ m.

162 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Classical nonholonomic systems are not obtained from a variational prin-ciple (at least in the usual sense – see Remark 3 below) but from D’Alembert’sprinciple, as discussed earlier. The equations are

D2q

dt2=

m∑k=1

λiWi (6.5.5)

subject to ωk(Dqdt

)=⟨Wk,

Dqdt

⟩= 0, 1 ≤ k ≤ m where ωk(X) = 〈Wk, X〉

and the λi are Lagrange multipliers.We now define a controlled nonholonomic mechanical system to be a

system of the form

D2q

dt2=

m∑i=1

λiWi +N∑i=1

uiXi (6.5.6)

subject to ⟨Wk,

Dq

dt

⟩= 0 1 ≤ k ≤ m, (6.5.7)

where the ui(t) are controls and the Xi are arbitrary smooth (control)vector fields. In fact, the Xi are not as arbitrary as they appear, as theremark below shows.Remark 1. We now consider which general force fields F are compatiblewith the nonholonomic constraints i.e., which F (q, q) are allowed in thesystem

D2q

dt2= F

subject to ⟨Wk,

Dq

dt

⟩= 0 , 1 ≤ k ≤ m. (6.5.8)

Differentiating the constraints gives

〈Wk, F 〉+⟨DWk

dt,Dq

dt

⟩= 0 , 1 ≤ k ≤ m

or, if ∇ is the Riemannian connection on (M, 〈, 〉),

〈Wk, F 〉+ 〈∇qWk, q〉 = 0 , 1 ≤ k ≤ m (6.5.9)

These are the conditions that the force field must satisfy to be compatiblewith the constraints. The Lagrange multipliers ensure the forces satisfy theabove relations. This also shows that number of independent force fields isless than or equal to n−m.

6.6 Symmetries and Conservation Laws 163

Remark 2. We define coresponding the kinematic system corresponding

to be Dqdt =

N∑i=1

uiXi, Xi ∈ H. For controllability analysis, the assumption

is made that H is completely nonholonomic (see Vershik and Gershkovich[1988]). (We refer again to Brockett [1982]). We analyze such systems else-where.Remark 3. We note that while the nonholonomic system under discus-sion is not variational in the Lagrangian sense, it can be seen as the solutionof an instantaneous variational problem,

minD2qdt2

12

∥∥∥∥∥D2q

dt2−

N∑i=1

uiXi(q)

∥∥∥∥∥subject to ⟨

Wk,Dq

dt

⟩= 0 , 1 ≤ k ≤ m. (6.5.10)

This yields the nonholonomic equations as well as the following equationsfor λk:

m∑j=1

〈Wk,Wj〉λj = −⟨DWk

dt,Dq

dt

⟩−

N∑i=1

ui〈Wk, Xi〉

1 ≤ k ≤ m. (6.5.11)

6.6 Symmetries and Conservation Laws

Symmetries in mechanics give rise to constants of the motion (see for exam-ple Abraham and Marsden [1978]). In the Riemannian context isometriesare generated by Killing vector fields. (Recall that Z is a Killing vector fieldif 〈∇Y Z, Y 〉 = 0 for all vector fields Y – see e.g., Crouch [1981].) Further, asufficient condition for

⟨Z, Dqdt

⟩to be a constant of motion for the geodesic

flow is that Z is a Killing vector field. For controlled nonholonomic systemswe have

Theorem 6.6.1. Sufficient conditions for⟨Z, Dqdt

⟩to be a constant of

motion for the controlled nonholonomic system 6.5.6 are

(i) Z ∈ H

(ii) Zq ∈ SpanX1, . . . , XN⊥

(iii) Z is a Killing vector field.

164 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Proof. ddt

⟨Z, Dqdt

⟩=⟨DZdt ,

Dqdt

⟩+⟨Z, D

2qdt2

⟩. The first term is zero by

(iii) and the second is zero by 6.5.6, (i) and (ii).Note that when M = Rn and the metric g is independent of xi, then ∂

∂xiis a Killing vector field. ¥

6.6.1 Reduction

We discuss here an approach to reduction for nonholonomic control sys-tems.

It is often convenient to introduce a bundle structure in M ,

MyπB

with fiber F , DimB = r and DimF = n − r. This structure must becompatible with the constraints in the sense that π∗Hx = Tπ(x)B, ∀x ∈M(thus DimH = n −m ≥ DimB = r). Our aim is to reduce the dynamicalsystem 6.5.6, 6.5.7 so that evolution on the fiber is given by a first orderequation. To do this we introduce two further assumptions. Either

(1) DimH = DimB, i.e., n = m+ r. In this case H = H clearly definesa horizontal distribution on the bundle, or,

(2) DimH − DimB = n − m − r = s > 0, and there exist s linearlyindependent vector fields Z1, . . . , Zs which satisfy conditions (i)-(iii)of Lemma 1. In particular

〈Zi,Dq

dt〉 = ci (6.6.1)

are constants of the motion for 6.5.6, 6.5.7.

We define a distribution H0 on M by setting

X ∈ H0 if 〈Wk, X〉 = 0 , 1 ≤ k ≤ m,

〈Zk, X〉 = 0 , 1 ≤ k ≤ s . (6.6.2)

Condition (i) of Lemma 1 ensures that H0 is r-dimensional. We also requirethat H0 defines a horizontal distribution of the bundle, i.e.,

TxF ∩ H0 = 0 ∀x ∈M . (6.6.3)

We further define the r-dimensional affine connection H on M by setting

x ∈ H if 〈Wk, X〉 = 0 , 1 ≤ k ≤ m〈Zk, X〉 = ck 1 ≤ k ≤ s . (6.6.4)

6.6 Symmetries and Conservation Laws 165

In either case 1) or 2) we have, as a direct sum of affine subspaces, Hx ⊕TxF = TxM , and so any vector field on M can be decomposed uniquelyinto components Yx = Y Hx + Y Fx with Y Hx ∈ Hx and Y Fx ∈ Fx. With thisstructure we can now decompose the velocity:

Dq

dt= qH + qF , qH ∈ Hq , qF ∈ TqF (6.6.5)

Using 6.5.7 we obtain:

〈Wk, qF 〉 = −〈Wk, q

H〉 , 1 ≤ k ≤ m,

〈Zk, qF 〉 = 〈Zk, qH〉+ ck , 1 ≤ k ≤ s . (6.6.6)

Our rank conditions on Zk and Wk allow us to solve these equations giving

qF = fF (q, qH) . (6.6.7)

Equations 6.5.6 now define the second order set of equations

DqH

dt= fH(q, q, u) . (6.6.8)

We have now provided a reduction from the 2n first order equations 6.5.6to n+ r first order equations.

Now locally we can write qH = Ddtq

B for some trajectory qB(t) ∈ B. Insome cases we may be able to rewrite the equations globally in terms of atrajectory q(t) = (qB(t), qF (t)), qB ∈ B, qF ∈ F .

The class of Caplygin control systems introduced in Bloch et. al. [1992]and dicussed in an earlier section corresponds to the case where H = Hand there exists a trivial bundle structure M = Rn with the Euclideanstructure. In this case, since ∂

∂qFiis Killing and symmetries correspond to

invariance with respect to qFi , a global prescription can be given, in whichthe equations become

qF = fF (qB , qB)DqB

dt= fH(qB , qB , u) , (6.6.9)

and, in fact, fF is affine in qB .In certain related cases, such as when M is a principal bundle with its

structure group G being a group of isometries and H = H is G-invariant,we can also obtain a global reduction where (19) is a set of second orderode’s on B. (The uncontrolled class of these systems are studied in detailby Koiller [1988] as nonabelian Caplygin systems. See also recent workby Dayawansa, Krishnaprasad and Yang [1992] and Marsden and Scheurle[1991].)

166 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Example: The Rolling Ball. We consider the controlled rolling ballon the plane, the kinematics of which were discussed in Brockett and Dai[1991], establishing the completely nonholonomic nature of the constraintdistribution H. The dynamics of the uncontrolled system are described forexample in McMillan [1936].

We will use the coordinates x, y for the linear horizontal displacementand P ∈ SO(3) for the angular displacement of the ball. Thus P gives theorientation of the ball with respect to inertial axes e1, e2, e3 fixed in theplane, where the ei are the standard basis vectors aligned with the x,yand z axes respectively.

Let ω ∈ R3 denote the angular velocity of the ball with respect to inertialaxes. In particular, the ball may spin freely about the z-axis and the z-component of angular momentum is conserved. If J denotes the inertiatensor of the ball with respect to the body axes, then J = PTJP denotesthe inertia tensor of the ball with respect to the inertial axes and Jω isthe angular momentum of the ball with respect to the inertial axes. Theconservation law alluded to above is expressed as

eT3 Jω = c . (6.6.10)

The nonholonomic constraints are expressed as (see e.g., Brockett andDai [1991])

eT2 ω + x = 0eT1 ω − y = 0 . (6.6.11)

The kinematics for the rotating ball we express as P = S(ν)P whereν = Pω is the angular velocity in the body frame (see Crouch [1981] forthe definition of the skew-symmetric matrix S(ν)).

We now wish to write down equation (6) for this example using the globalcoordinates q = (x, y, P ). The metric on M = SO(3)× R2 is defined by

〈(S(ν1)P, x1, y) , (S(ν2)P, x2, y2)〉 (6.6.12)= νT1 Jν2 + x1x2 + y1y2 .

An explicit expression for the Riemannian connection on SO(3) may befound by consulting Arnold [1978], p. 327. Equation (6) becomes

S(ν)P − S(J−1S(ν)Jν)P= λ1S(J−1Pe1)P + λ2S(J−1Pe2)P

x = λ1 + u1

y = −λ1 + u2 . (6.6.13)

6.7 A Discontinuous Energy Method for a Class of Nonholonomic System 167

It is convenient to re-express this system using inertial coordinates ω =PT ν, in which case we obtain

ω = J−1S(ω)Jω + λ1J−1e1 + λ2J−1e2

x = λ2 + u1

y = −λ1 + u2

P = PS(ω) . (6.6.14)

From the equations it is easy to see that indeed (21) is a constant of themotion. In fact Z in formula (12) is given by Z = S(Pe3)P . It is easy tocheck conditions (i) and (ii) of Lemma 1, and, with more effort, one canverify condition (iii). Expressions for the multipliers λi are obtained by theanalysis of Remark 3, Section 2.

In this example, there are two admissible reductions, as described inSection 4. The first takes B = SO(3), the fibre F = R2, and the reductionof system (25) is obtained simply by substituting the two constraints (22)for the second order equations. This corresponds to the case H = H ofSection 4.

For the second reduction we take B = R2, F = SO(3), and we nowemploy the constraints (22) and the constants of motion (21) to obtain thefollowing expression for ω:

ω = x(α1e3 − e2) + y(e1 − α2e3) + α3e3

where

α1 =eT3 Je1

eT3 Je3, α2 =

eT3 Je2

eT3 Je3, α3 =

c

eT3 Je3. (6.6.15)

The reduced equations become, after substituting for the multipliers,

x = λ2 + u1

y = −λ1 + u2 (6.6.16)P = PS(x(α1e3 − e2) + y(e1 − α2e3) + α3e3)

In this case H is obtained through the second definition in Section 4,with s = 1. Note also that here we obtain seven first order ODE’s ratherthan the eight obtained from the previous reduction.

6.7 A Discontinuous Energy Method for aClass of Nonholonomic System

As we have seen, there has a been a great deal of research on the problemof stabilizing systems which fail a necessary condition for the existence of

168 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

smooth, or even continuous, feedback (see Brockett [1983]). One of the rea-sons for the interest in such systems is that nonholonomic systems fall intothis class. (See for, example, Murray and Sastry [1993], Bloch, McClamrochand Reyhanoglu [1992], and references therein.) Various approaches havebeen taken to the stabilization problem for such systems, focusing mainlyon the development of either smooth dynamic feedback or nonsmooth feed-back. An important paper regarding the former approach is Coron [1992].See also Pomet [1992] and M’Closkey and Murray [1993]. Kolmanovskyand McClamroch [1993], for example, used a discontinuous, discrete-timeapproach, while Brockett [1993] used a stochastic approach. Other impor-tant work includes that of Liu and Sussmann [1991], as well as work ofSamson, Sordalen, Walsh, Bushnell and others. Another interesting prob-lem for such systems in the problem of tracking which was also analyzedin Brockett [1993].

6.7.1 Stabilizing the Heisenberg System

Here we follow Bloch and Drakunov [1994], [1996] in considering a a slidingmode approach to the stabilization and tracking problem for the so-callednonholonomic integrator or Heisenberg system (so called because the un-derlying Lie algebra of control vector fields is isomorphic to the Heisenbergalgebra.) Firstly we provide a feedback which will globally asymptoticallystabilize the origin. The idea is to use the natural algebraic structure ofthe system together with ideas from sliding mode theory (see Utkin [1978],DeCarlo et al. [1988], Drakunov and Utkin [1992]). We announced the mainresult on stabilization in Bloch and Drakunov [1994]. Related recent workincludes the following. Khennouf and Canudas de Wit [1995] presentedthe alternative control scheme (6.7.24), (6.7.25), discussed below, and inCanudas de Wit and Knennouf [1995], they considered robustness issues.Hespanha [1995] and Morse [1995] discussed these control ideas in termsof logic based switching. Sliding mode control for differentially flat non-holonomic systems is discussed in Sira-Ramirez [1995]. We discussed thetracking problem in Bloch and Drakunov [1995], while a general approachto tracking via sliding modes is described in Guldner and Utkin [1995].

The nonholonomic integrator is merely the simplest example of an im-portant class of nonlinear controllable systems of the x = B(x)u, B(x)an n ×m matrix, m < n, introduced in the fundamental paper Brockett[1981], and which are prototypical examples of systems where smooth feed-back fails and which have a natural controllability condition. In a relatedpaper Bloch and Drakunov [1996] discuss the stabilization of such generalsystems via a discontinous (but not sliding mode) algorithm. Our algo-rithm for the general case involves switching between different Lyapunovfunctions.

6.7 A Discontinuous Energy Method for a Class of Nonholonomic System 169

6.7.2 Stabilization of the Nonholonomic Integrator inSliding Mode

We consider the system (see Brockett [1981]):

x = u (6.7.1)y = v (6.7.2)z = xv − yu. (6.7.3)

As mentioned earlier it is a prototype for more complex controllable butnot smoothly stabilizable systems.

We note also that this system can be obtained by a change of variablesfrom a systems in co-called “chained form” (see e.g. Sira-Ramirez [1995]and references therein.)

The problem of stabilizing 6.7.1-6.7.2, even locally, is not a trivial task,since, as can be easily seen, the linearization in the vicinity of the origingives the noncontrollable system

x = u

y = v

z = 0.

In fact, as was proved by Brockett [1983], the system 6.7.1-6.7.3 cannot bestabilized by any smooth feedback control law. As discussed in the intro-duction, later approaches using time-periodic feedback and a randomizedfeedback were developed.

In this paper we present time invariant laws solving the stated problem.By nature, they are discontinuous and lead to sliding along manifolds ofreduced dimensionality in the state space.

The main difficulty here is the fact that stabilization of x and y leads tozero right hand side of 6.7.3 and, therefore, the variable z cannot be steeredto zero. That simple observation implies that to stabilize the system oneneeds to make z converge “faster”, than x and y.

We suggest using the following control law

u = −αx+ β y sign(z) (6.7.4)v = −αy − β x sign(z), (6.7.5)

where α, and β are positive constants.Let us show that there exists a set of initial conditions such that trajec-

tories starting there converge to the origin.Consider a Lyapunov function for (x, y)-subspace:

V =12

(x2 + y2). (6.7.6)

170 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

The time derivative of V along the trajectories of the system 6.7.3 is neg-ative:

V = −αx2 + βxy sign(z)− αy2 − βxy sign(z) = −α(x2 + y2) = −2αV.(6.7.7)

Therefore, under the control 6.7.4, 6.7.5 the variables x and y are stabilized.Now let us consider the variable z. Using 6.7.3 and 6.7.4, 6.7.5 we obtain:

z = xv − yu = −β(x2 + y2)sign(z) = −2βV sign(z). (6.7.8)

Since V does not depend on z and is a positive function of time, the absolutevalue of the variable z will decrease and will reach zero in finite time if theinequality

2β∫ ∞

0

V (τ)dτ > |z(0)| (6.7.9)

holds. If z(0) is such that

2β∫ ∞

0

V (τ)dτ = |z(0)| , (6.7.10)

z(t) converges to the origin in infinite time (asymptotically). Otherwise, itconverges to some constant nonzero value of the same sign as z(0).

If the above inequality 6.7.9 holds, the system trajectories are directedto the surface z = 0 and the variable z(t) is stabilized at the origin in finitetime. (The variables x and y, as follows from 6.7.7, always converge to theorigin while within that surface.)

This phenomenon is known as sliding mode (see Utkin [1978], DeCarlo[1988]). The manifold z = 0 is a stable integral manifold of the closed loopsystem 6.7.1-6.7.3, 6.7.4, 6.7.5. Its characteristic feature is reachability infinite time (Drakunov and Utkin, 1992). Using a smooth control (even acontrol satisfying a local Lipschitz condition (in the vicinity of z = 0)such fast convergence cannot be achieved. On the other hand, within thesliding manifold z = 0 the system behavior is described in accordancewith the Filippov definition for systems of differential equations with dis-continuous right hand sides (Filippov, 1964).

Let us explain the version of this definition which we are using. Weconsider the system

x = f(x), (6.7.11)

with f(x) a discontinuous function comprised of a finite number of contin-uous fk(x), (k = 1, . . . , N) so that

f(x) ≡ fk(x) for x ∈Mk, (6.7.12)

6.7 A Discontinuous Energy Method for a Class of Nonholonomic System 171

where the open regionsMk have piecewise smooth boundaries ∂Mk. Thenaccording to the Filippov definition we define the right hand side of 6.7.11within ∂Mk as

x =∑k∈I(x)

µkfk(x). (6.7.13)

The sum is taken over the set I(x) of all k such that x ∈ ∂Mk and thevariables µk satisfy ∑

k∈I(x)

µk = 1, (6.7.14)

i.e. the right hand side belongs to the convex closure cofk(x) : k ∈ I(x)of the vector fields fk(x) for all k ∈ I(x). Actually, the Filippov definitionreplaces the differential equation 6.7.11 by a differential inclusion

x ∈ cofk(x) : k ∈ I(x) (6.7.15)

for the points x belonging to the boundaries ∂Mk. If within the convexclosure there exists a vector field tangent to all or some of the boundariesthen there is a solution of the differential inclusion belonging to ∂Mk whichcorresponds to the sliding mode.

In the above relatively simple case, the Filippov definition provides aunique solution and implies that the system on the manifold is

x = −x,

y = −y.

Since from 6.7.7 it follows that

V (t) = V (0)e−2αt =12

(x2(0) + y2(0))e−2αt , (6.7.16)

substituting this expression in 6.7.9 and integrating we find that the con-dition for the system to be stabilized is

β

2α[x2(0) + y2(0)] ≥ |z(0)|. (6.7.17)

The inequality

β

2α(x2 + y2) < |z|. (6.7.18)

defines a parabolic region P in the state space.The above derivation can be summarized in the following theorem:

172 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Theorem 6.7.1. If the initial conditions for the system 6.7.1-6.7.3 be-long to the complement Pc of the region P defined by 6.7.18, then thecontrol 6.7.4, 6.7.5 stabilizes the state.

If the initial data are such that 6.7.18 is true, i.e. the state is inside theparaboloid, we can use any control law which steers it outside. In fact,any nonzero constant control can be applied. Namely, if u ≡ u0 = const,v ≡ v0 = const, then

x(t) = u0t+ x0,

y(t) = v0t+ y0,

z(t) = t(x0v0 − y0u0) + z0.

With such x, y and z the left hand side of 6.7.18 is quadratic with respectto time t while right hand side is linear. Hence, when the time increasesthe state inevitably will leave P.

A global control law in the form of the feedback (although discontinuous)can be described as follows:

(u, v)T =

(u0, v0)T if (x, y, z)T ∈ Peq.6.7.4, 6.7.5 if (x, y, z)T ∈ Pc (6.7.19)

Theorem 6.7.2. The closed system 6.7.1-6.7.3, 6.7.19 is globally asymp-totically stable at the origin.

Global asymptotic stability means that: (i) for all initial conditions x(t), y(t), z(t)→0, when t→∞; (ii) ∀ε > 0 there exists δ > 0 such that x2

0 + y20 + z2

0 < δ2

implies x2(t) + y2(t) + z2(t) < ε2 for any t ≥ 0.We have already shown above that (i) is true. (ii) follows from the fact

that outside P and on the surface of parabola ∂P the state monotonicallyapproaches the origin. For initial conditions inside P we have :

x2(t) + y2(t) + z2(t) = (u0t+ x0)2 + (v0t+ y0)2 + [(x0v0 − y0u0)t+ z0]2.(6.7.20)

The maximum of the expression 6.7.20 is achieved for t = 0 or t = tf , wheretf is the first moment of time when the state reaches ∂P. This moment isdefined by an equation

β

2α(u0tf + x0)2 + (v0tf + y0)2 = |(x0v0 − y0u0)tf + z0|. (6.7.21)

As can be easily seen from 6.7.21, for fixed u0, v0, the solution of thisequation tf tends to zero if x0, y0, z0 tend simultaneously to zero. Thatproves (ii).

6.7 A Discontinuous Energy Method for a Class of Nonholonomic System 173

The parameters α > 0, β define the size of paraboloid. When βα → ∞

the parabolic region P limits to to the z-axis. From that point of view, tostabilize the system 6.7.1-6.7.3, it is reasonable to increase β as the stateapproaches the origin (if we decrease α the convergence of x and y willbe slower). To realize this idea we can use a control law where α increaseswhen x and y approach the origin

u = −αx+ βy

x2 + y2sign(z) (6.7.22)

v = −αy − β x

x2 + y2sign(z), (6.7.23)

or even

u = −αx+ βy

x2 + y2z (6.7.24)

v = −αy − β x

x2 + y2z. (6.7.25)

As mentioned above for a detailed anaysis in the case (6.7.24), (6.7.25),see Khennouf and Canudas de Wit [1995].

Then from 6.7.3 we have ??Simulations of the algorithm for two types of initial conditions are shown

in Fig. 1.

z = −β sign(z),

or

z = −βz,

respectively.In both cases, the state converges to the origin from any initial conditions,

except the ones belonging to the z-axis. But, in contrast to 6.7.4, 6.7.5the control laws 6.7.22, 6.7.23 and 6.7.24, 6.7.25 are unbounded in in theneigbourhood of z-axis (on the axis it is not defined). If the initial conditionsbelong to this set again we can apply any nonzero constant control for anarbitrary small period of time and then switch to 6.7.22, 6.7.23 or 6.7.24,6.7.25. A method of dealing with the boundedness problem is also describedby Khennouf and Canudas de Wit. Another global (excluding x = 0 ∩y = 0 space), but only ε-stabilizing control, may be obtained by switchingα:

Let α be the following function of x and y

α = α0 sign(x2 + y2 − ε2) (6.7.26)

where α0 > 0, β > 0 are constants and let the control be

u = −αx+ βyz (6.7.27)v = −αy − βxz. (6.7.28)

174 6. Controllability and Stabilizability for Mechanical and Nonholonomic Systems

Using 6.7.7 we find that from any initial conditions x and y the state reachesan ε-sphere of the x,y-space origin:

x2 + y2 = const = ε2. (6.7.29)

After that the equation for the variable z is

z = −βε2z. (6.7.30)

Therefore, z → 0 when t → ∞, while the variables x and y stay in anε-vicinity of the origin. Of course, in 6.7.27, 6.7.28 z can be replaced byany function g(z) which guarantees asymptotic stability of the equation

z = −βε2g(z), (6.7.31)

for example, g(z) = sign(z).Another interesting control can be obtained if in 6.7.26 we replace ε2 by

|z|:

α =

α0, if x2 + y2 > ε2

α1, if x2 + y2 ≤ ε2,(6.7.32)

where α0 > 0, and α1 ≤ 0

α = α0 sign(x2 + y2 − |z|). (6.7.33)

and

u = −αx+ βy sign(z) (6.7.34)v = −αy − βx sign(z). (6.7.35)

In this case outside the parabolic region x2 + y2 > |z| the asymptotic con-vergence of z is guaranteed. If the initial conditions are inside this region,x2 + y2 is increasing and reaches the parabola in finite time, remaining insliding mode on the surface of parabola, where

z = −βz. (6.7.36)

In fact, this control forms two sliding surfaces in the state space of theclosed system: z = 0 and x2 + y2 = |z|.

This is page 175Printer: Opaque this

7Optimal Control

7.1 The Maximum Principle

Here we describe briefly the maximum principle for deriving necessary con-ditions for optimal control problems. Optimal control problems have beentypically cast in a different setting from the classical variational problems,more closely associated with mechanics. The basic difference lies in theway in which the trajectories are formulated; in the optimal control settingthe trajectories are “parameterized” by the controlled vector field, whilein the traditional variational setting trajectories are simply “constrained.”The other basic difference is that the necessary conditions for extremalsin the optimal control setting are typically expressed using a Hamiltonianformulation using the Pontryagin maximum principle, rather than the La-grangian setting. (see Pontryagin et al. [1962]) , Lee and Markus [1976] ,and Sussmann [1977] ). We state a typical optimal control problem:

minu(·)

∫ T

0

h0(x, u)dt

subject to: (i) x = f(x, u), x ∈Mn, u ∈ Ω ⊂ Rk,(ii) x(0) = x0, x(T ) = xT

(7.1.1)

where f and h0 ≥ 0 are smooth and Ω is a closed subset of Rk. To statenecessary conditions dictated by the Pontryagin maximum principle, weintroduce a parameterized Hamiltonian function on T ∗M

H(x, p, u) = p(f(x, u))− p0h0(x, u) (7.1.2)

176 7. Optimal Control

where p0 ≥ 0 a fixed positive constant, and p ∈ T ∗M . We denote by t 7→u∗(t) a curve which satisfies the following relationship along a trajectoryt 7→ (x(t), p(t)) in T ∗M :

H∗(x(t), p(t), t) ∆= H(x(t), p(t), u∗(t)) = maxu∈Ω

H(x(t), p(t), u). (7.1.3)

The time varying Hamiltonian function H∗ defines a time varying Hamil-tonian vector field XH∗ on T ∗M , with respect to the canonical symplecticstructure on T ∗M . One statement of Pontryagin’s maximum principle givesnecessary conditions for extremals of the problem (7.1.1) as follows: an ex-tremal trajectory t 7→ x(t) of the problem (7.1.1), is the projection onto Mof a trajectory of the flow of the vector field XH∗ , which satisfies the bound-ary condition (7.1.1)(ii), and for which t 7→ (p(t), p0) is not identically zeroon [0, T ]. The extremal is called normal when p0 6= 0, (in which case we setp0 = 1). When p0 = 0 we call the extremal abnormal, corresponding to thecase where the extremal is determined by constraints alone.

If the extremal control function u∗ is not determined by the system(7.1.3) along the extremal trajectory, then the extremal is said to be sin-gular, in which case further necessary conditions are needed to determineu∗. In this paper we are only interested in nonsingular situations in whichΩ = Rk, (or indeed more general vector bundles). We also suppose thatthe data is sufficiently regular so that u∗ is determined uniquely from thecondition

0 ≡ ∂H

∂u(x(t), p(t), u∗(t)), t ∈ [0, T ]. (7.1.4)

It follows that there exists a function k such that u∗(t) = k(x(t), p(t)). Wethen set

H(x, p) ∆= H(x, p, k(x, p)). (7.1.5)

It follows that along an extremal

H(x(t), p(t)) = H∗(x(t), p(t), t). (7.1.6)

Given the assumed regularity, and condition (7.1.4), the extremal flow isalso a flow of the time invariant Hamiltonian function H. Further, assum-ing that this function H is sufficiently regular, the extremal will also begoverned by a Lagrangian system. However, it is not immediately obviouswhat variational problem would result in this Lagrangian system. This issuewas investigated in depth in Bloch and Crouch [1994] , employing the clas-sical results (Rund [1966] , Bliss [1930] ), which relate classical constrainedvariational problems to Hamiltonian flows, although not optimal controlproblems. We outline the simplest relationship as discussed in Bloch andCrouch [1994] in the next section.

7.2 Variational and Nonvariational Nonholonomic Problems 177

7.2 Variational and NonvariationalNonholonomic Problems

A large part of mechanics deals with systems subject to constraints. Whenthe constraints are holonomic or integrable, the formulation of the dynamicsis fairly straightforward. When the constraints are nonholonomic howeverthe situation is more delicate. In fact one approach to the dynamics appliesto mechanical systems, while the other applies to optimal control problems.This dual picture is described in Vershik [1984] for example.

Suppose a submanifold of the tangent bundle is given as the zero setof a set of constraints on the bundle. Supppose also that we are given aLagrangian or, more generally an objective function. Then we can proceedin the following two ways:

1) We can consider the conditional variational problem of minimizing afunctional subject to the trajectories lying in the given submanifoldand obtain the Euler Lagrange equations via the Lagrange method ofappending the constraints to the Lagrangian via Lagrange multipliers.

2) We can project, via a suitable projection, the vector field of the un-conditional problem on the whole tangent bundle at every point tothe tangent space of a given submanifold.

These vector fields will not of course coincide in general, even thoughboth are tangent to the constraint submanifold.

The first method gives us variational nonholonomic mechanics while(real) nonholonomic mechanics is obtained by a procedure of the secondtype. In fact, this is implemented by the Lagrange D’Alembert princple aswe have seen earlier. Later we shall return to this principle and show howto carry it out in coordinate free fashion.

Nonholonomic mechanics is not variational, since while we allow all pos-sible variations in taking the variations of the Lagrangian, the variationshave to lie in the constraint distribution and are thus not independent.

Variational nonholonomic mechanics on the other hand is just equivalantto the classical Lagrange problem of minimizing a functional over a classof curves with fixed extreme points and satisfying a given set of equalities.

More precisely we have the following (see e.g. Bloch and Crouch [1994]).

Definition 7.2.1. The Lagrange problem is given by

minq(·)

∫ T

0

L(q, q)dt (7.2.1)

subject to q(0) = 0, q(T ) = qT ,

Φ(q, q) = 0,

where Φ : Rn × Rn → Rn−m.

178 7. Optimal Control

One can then proceed to solve these equations by forming the modifiedLagrangian

Λ(q, q, λ) = L(q, q) + λ · Φ(q, q) , (7.2.2)

with λ ∈ Rn−m. The Euler Lagrange equations then take the form

d

dt

∂qΛ(q, q, λ)− ∂

∂qΛ(q, q, λ) = 0 (7.2.3)

Φ(q, q) = 0 . (7.2.4)

This system, while defining perfectly good dynamics, does not correspondin general to a physical mechanical problem. It is certainly variational –over a restricted class of curves satisfying Φ(q, q) = 0. The case we areparticularly interested in is the case of of classical (linear in the velocity)nonholonomic constraints –

ωi(q, q) =∑j

aij(q)qj = 0 i = 1, · · · , n−m. (7.2.5)

In the case these constraints are integrable (equivalent to functions ofq only) and L is physical – i.e. it is a holonomic system, this system willrepresent physical dynamics. In the nonholonomic case, these equationswill not be physical since equations of this type to not satisfy Newton’slaw (F = ma!) while those of nonholonomic mechanics do – indeed theLagrange D’Alembert principle ensures this.

To see this explicitly let us examine the Lagrange multiplier formulationof the equations of motion in each case.

Consider firstly the variational case: We have

Theorem 7.2.2. The Lagrange problem 7.2.1 with constraints of the form7.2.5 yields the equations of motion

d

dt

∂qiL− ∂

∂qiL+

∑jk

(d

dtλj)ajk +

∑j

λj(aji −∂ajk∂qi

qk) = 0 (7.2.6)

with the constraints ∑j

aij qj = 0 . (7.2.7)

We can now contrast these equations of motion with the nonholonomicequations of motion with Lagrange multipliers obtained earlier from theLagrange D’Alembert principle:

d

dt

∂qiL− ∂

∂qiL =

n−m∑j=1

λjajiδqi . (7.2.8)

7.3 Variational Nonholonomic Mechanics and Optimal Control 179

Note that if we set λj = 0 and λj = λj in the variational nonholonomicequations we recover the true nonholonomic equations of motion. It is pre-cisely the omission of the λj term that destroys the variational nature ofthe nonholonomic equations.

7.3 Variational Nonholonomic Mechanicsand Optimal Control

Variational nonholonomic problems (i.e. constrained variational problems)are equivalent to optimal control problems as the following analysis (seeBloch and Crouch [1994]) shows.

Consider again a modified Lagrangian

Λ(q, q, λ) = L(q, q) + λ · Φ(q, q) . (7.3.1)

with Euler Lagrange equations

d

dt

∂qΛ(q, q, λ)− ∂

∂qΛ(q, q, λ) = 0

Φ(q, q) = 0 . (7.3.2)

We will write rewrite this equation in Hamiltonian form and show thatthe resulting equations are equivalent to the equations of motion the max-imum principle for a suitable optimal control problem.

Set

p =∂

∂qΛ(q, .q, λ) (7.3.3)

and consider this equation together with the constraints

Φ(q, q) = 0. (7.3.4)

We wish to solve 7.3.3 and 7.3.4 for (q, λ).Now assume that on an open set U the matrix[

∂2

∂q2 Λ(q, q, λ) ∂∂qΦ(q, q)T

∂∂qΦ(q, q) 0

](7.3.5)

has full rank. (This generalizes the usual Legendre condition – that ∂2

∂qL(x, x)has full rank.)

Hence we can solve for q and λ:

x = φ(q, p)λ = ψ(q, p) . (7.3.6)

We now have

180 7. Optimal Control

Theorem 7.3.1. (Caratheodory [1967], Rund [1966], Arnold [1988], Blochand Crouch [1994]) Under the transformation 7.3.6 the Euler Lagrange sys-tem 7.3.2 is transformed to the Hamiltonian system

q =∂

∂pH(q, p)

p = − ∂

∂qH(q, p) (7.3.7)

where

H(q, p) = p · φ(q, p)− L(q, φ(q, p)) . (7.3.8)

Proof. Φ(q, φ(q, p)) = 0 implies

∂Φ∂q

+∂Φ∂q

∂Φ∂q

= 0

∂Φ∂q

∂φ

∂p= 0

Hence, using 7.3.3, we have

∂H

∂p= φ+ (p− ∂L

∂q) · ∂φ∂p

= q + λ · (∂Φ∂q

∂φ

∂p)

= q

Similarly

∂H

∂q= −∂L

∂q+ (p− ∂L

∂q) · ∂φ∂q

= −∂L∂q

+ λ · (∂Φ∂q

∂φ

∂q)

= −(∂L

∂q+ λ · ∂Φ

∂q)

= −∂Λ∂q

= −p

¥

Now:

Definition 7.3.2. Let the Optimal Control problem be given by

minu(·)

∫ T

0

g(q, u)dt (7.3.9)

7.3 Variational Nonholonomic Mechanics and Optimal Control 181

subject to q(0) = 0, q(T ) = qT ,

q = f(x, u),

where x ∈ Rn, u ∈ Rm.

Then we have

Theorem 7.3.3. The Lagrange problem and Optimal Control problemgenerate the same (regular) extremals trajectries provided

(i) Φ(q, q) = 0) if and only if there exists a u such that q = f(q, u),

(ii) L(q, f(q, u)) = g(q, u) , and

(iii) The optimal control u∗ is uniquely determined by the condition

∂H

∂u(q, p, u∗) = 0 (7.3.10)

where

∂2H

∂u2(q, p, u∗)

is full rank and

H(q, p, u) = p · f(q, u)− g(q, u) (7.3.11)

is the Hamiltonian function given by the maximum principle.

Proof. By (iii) we may use the equation

p · ∂f∂u

(q, u∗)− ∂g

∂u(q, u∗)

to deduce that there exists a function r such that u∗ = r(q, p).The extremal trajectories are now generated by the Hamiltonian

H(q, p) = H(q, p, r(x, p)) = p · f(q, r(q, p))− g(q, r(q, p)) . (7.3.12)

Then the result follows and we have

H(q, p) = H(q, p)f(q, r(q, p)) = φ(q, p)g(q, r(q, p)) = L(q, φ(q, p))

¥

182 7. Optimal Control

7.3.1 Two Examples of Variational NonholonomicMechanics

In this subsection we give two simple examples of the theory above, consist-ing of the rolling penny (or unicycle) and the “Heisenberg” control system(see Bloch and Crouch [1993]).

Example 1: Rolling Penny or Unicycle We consider again a pennyrolling on the x − y plane with rotation angle θ and heading angle φ, asdiscussed in section 1.1, but now we consider the corresponding variationalsystem – just the variational control system considered in the introduction,minus the controls.

We have the augmented Lagrangian

L =12m(x2 + y2) +

12Iθ2 +

12ϕ2 + λ1(x−Rθ cosϕ) + λ2(x−Rθ sinϕ) .

(7.3.13)

The Euler Lagrange equations are then

mx+ λ1 = 0 (7.3.14)my + λ2 = 0 (7.3.15)

Iθ −R d

dt(λ1 cosϕ+ λ2 sinϕ) = 0 (7.3.16)

Jϕ+R∂d

∂ϕ(λ1θ cosϕ+ λ2θ sinϕ) = 0 . (7.3.17)

From the constraint equations and equations 7.3.14 and 7.3.15 we haveas before

λ1 = −mRθ cosφ+A

λ2 = −mRθ sinφ+B

where A and B are constants.Substituting into equations 7.3.16 and 7.3.17 we obtain

(I +mR2)θ = Rϕ(−A sinϕ+B cosϕ) (7.3.18)Jϕ = Rθ(A sinϕ−B cosϕ) . (7.3.19)

(7.3.20)

These equations, together with the constraints, define the dynamics.¨

Example 2: Heisenberg System We now consider the variational non-holonomic version of the Heisenberg system described in section 1.2. Theaugmented Lagrangian in this case is

L =12

(x2 + y2 + z2) + λ(z − yx+ xy). (7.3.21)

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 183

The corresponding Euler-Lagrange equations are:

d

dt(x− λy)− λy = 0 = x− 2λy − λy (7.3.22)

d

dt(y + λx) + λx = 0 = y + 2λx+ λx (7.3.23)

z + λ = 0. (7.3.24)

From the constraint equation we obtain

λ = −z = xy − yx. (7.3.25)

Solving for x and y we obtain

φλ+ φλ = 0

where φ = (1 + x2 + y2). Thus if c = µ(0)(1 + x(0)2 + y(0)2)

λφ = c

and the variational nonholonomic equations become:

φx = λ(2φy − φy) (7.3.26)φy = −λ(2φx− φx) (7.3.27)φz = λφ . (7.3.28)

Note that in each case we obtain the nonholonomic equations of motionby setting the constants of integration for the multipliers equal to zero: Aand B in example 1 and c in example 2.

¨

7.4 Optimal Control on Lie Algebras andAdjoint Orbits

An interesting class of optimal control problems are those which lie natu-rally on adjoint orbits of compact Lie groups. Analysis of problems of thistype may be found in Brockett [1994] and Bloch and Crouch [1995, 1996].

In this section we will set up the variational problem on the adjointorbits of compact Lie groups and derive the corresponding Hamiltonianequations.

Let g be a complex semisimple Lie algebra, gu its compact real form,and Gu the corresponding compact group. In this case a natural drift freecontrol system on an orbit of Gu takes the form

x = [x, u] (7.4.1)

184 7. Optimal Control

We remark that we formulate the problem in this generality for conve-nience, but the most useful case to bear is mind is the algebra su(n) ofskew-Hermitian matrices or the algebra so(n) of skew symmetric matrices(the intersection of the compact and normal real forms of the the algebrasl(n,C)). Orbits in this case are similarity orbits under the group action.

We then consider the following generalization of the functional suggestedby Brockett [1994] (we shall return to Brockett’s precise problem shortly):

η(x, u) =∫ tf

0

1/2||u||2 − V (x)dt (7.4.2)

where || · || =< ·, · >1/2 is the norm induced on gu by the negative of theKilling form κ(·, ·) on g and V is a smooth function on gu. The pairingbetween vectors x in g and dual vectors p in g∗ may be written < p, x >=−κ(x, p).

We have

Theorem 7.4.1. The equations of the maximum principle for the varia-tional problem with functional 7.4.2 subject to the dynamics 7.4.1 are

x = [x, [p, x]]p = [p, [p, x]]− Vx . (7.4.3)

Proof. The Hamiltonian is given by

H(x, p, u) =< p, [x, u] > −1/2||u||2 + V (x) . (7.4.4)

Hence∂H

∂u= − < [x, p], · > − < u, · >

and thus the optimal control is given by

u∗ = [p, x] (7.4.5)

Substituting this into H we find the Hamiltonian evaluated along theoptimal trajectory is given by

H∗(p, x) = −1/2 < x, [p, [p, x]] > +V (x) (7.4.6)

Computing

x =(∂H∗

∂p

)Tand

p = −(∂H∗

∂x

)Tgives the result. ¥

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 185

A particularly interesting special case of this problem is that of Brockett[1994] where we have

Corollary 7.4.2. The equations of the maximum principle for the vari-ational problem 7.4.2 subject to equations 7.4.1 with V (x) = −1/2||[x, n]||2are

x = [x, [p, x]p = [p, [p, x]]− [n, [n, x]] . (7.4.7)

The proof of the corollary follows immediately, setting V (x) = 1/2 <x, [n, [n, x]] >. Note that with this functional the equations lie naturallyon an adjoint orbit. In addition, these equations are interesting in that theoptimal flow may be related to the integrable Toda lattice equations (seebelow and Brockett [1994].)

These equations may be recast as the Euler Lagrange equations foundby Brockett.

In order to do this we introduce the following notation (see e.g. Bloch,Brockett and Ratiu [1992]): Let x and l lie in gu. Then x may be decom-posed as x = xl + xl where xl ∈ Ker(adl) and xl ∈ Im(adl) and whereadx(y) = [x, y]. Further, given any l ∈ gu we may decompose gu orthogo-nally relative to −κ(, ) as glu⊕ gul where glu = Im(adl) and gul = Ker(adl).

Now any velocity vector is tangent to the orbit of the adjoint action andhence is of the form x = [x, a] for some x ∈ gu. Hence x ∈ Im(adx). Theinverse of operator adx which, following Brockett [1994], we will denote byad−1x , is well defined on Im(adx) and hence on x.Then we have

Proposition 7.4.3. The equations 7.4.7 are equivalent to the Euler-Lagrangeequations (Brockett [1994])

x = [x, ad−1x (x)] + ad2

xad2n(x). (7.4.8)

Proof. Eliminate p from the two equations 7.4.7. The computation isstraightforward but somewhat lengthy. ¥

The proof of this proposition may of course be obtained also by theLegendre transformation. In terms of the operator ad, the integrand of7.4.2 with the given V (x) may be written as the Lagrangian (see Brockett[1994])

L(x, x) = 1/2(||[ad−1x x||2 + ||[x, n]||2). (7.4.9)

Then, since 1/2||ad−1x x||2 = −1/2 < ad−2

x x, x > we have

p =∂L

∂x= −ad−2

x x.

186 7. Optimal Control

From this we obtain the (optimal) Hamiltonian H∗.We remark that the kinetic energy in 7.4.9 is given by the so-called

normal metric (see e.g. Bloch, Brockett and Ratiu [1992]). This is definedas follows:

Let O be an adjoint orbit of gu and suppose ξ = [x, a] and η = [x, b] aretangent vectors to the orbit at x then then gn(ξ, η) =< ξ, η >= −κ(ax, bx)where ax and bx lie in Im(adx) as defined above.

We consider now the precise sense in which the equations discussed aboveare Hamiltonian. The discussion here is brief – more detail may be foundin Bloch, Brockett and Crouch [1997].

We have

Theorem 7.4.4. Let ω be the standard symplectic structure on T ∗gu.Consider the Hamiltonian

H(x, p) = 1/2||[p, x]||2 + V (x) (7.4.10)

where V(x) is any smooth function on g and p is a momentum variableviewed as lying in g by indentifying g with its dual. The Hamiltonian equa-tions of motion are

x = [x, [p, x]]p = [p, [p, x]]− Vx. (7.4.11)

Proof. Let ξ = (δx, δp) denote an arbitrary tangent vector to T ∗gu anddenote the Hamiltonian vector field corresponding to H by XH = (ζx, ζp).We need to solve for XH from the equation dH.ξ = ω(XH , ξ). Now

dH.ξ =< [p, x], [δp, x] > + < [p, x], [p, δx] > + <∂V

∂x, δx >

and

ω(XH , ξ) =< ζx, δp > − < ζp, δx > .

Equating these expressions gives the result. ¥

We now have

Corollary 7.4.5. Let V (x) = 1/2 < [x, n], [x, n] > in the Hamiltonian7.4.10. Then the Hamiltonian equations 7.4.11 yield the optimal Hamilto-nian equations 7.4.7.

Consider now the case V = 0. Remarkably, even though these equationare Hamiltonian with respect to the standard symplectic structure on T ∗gu,they are in fact Hamiltonian on the cotangent bundle of an adjoint orbitof Gu. That is, even though the Hamiltonian structure on the orbit iscomplicated and is not the restriction of the structure on the Lie algebra,the equations themselves do restrict. We have

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 187

Theorem 7.4.6. For V (x) = 0 the equations

x = [x, [p, x]]p = [p, [p, x]] . (7.4.12)

are the Hamiltonian form of the geodesic equations with respect to the nor-mal metric on an adjoint orbit of gu.

Proof. From the optimal control calculation above we know that theseare the equations of the geodesic flow. It remains to observe that this is theflow with respect to the normal metric. Now for V = 0 the Hamiltonianis just the norm of the velocity in the normal metric. We now need tocheck that the equations of motion are Hamiltonian with respect to thisHamiltonian and a symplectic structure on the cotangent bundle of theorbit, which may be identified with the tangent bundle. A tangent vectorto the tangent bundle to the orbit at the point (x, [x, ξ]) is of the form

(x, [x, ξ], [x, η], [[x, η], ξ] + [x, ζ]) (7.4.13)

for ξ, η, ζ ∈ gu. Then, using a natural symplectic structure as in Thimm[1981] gives the result. The details of this standard but lengthy computationare given in Bloch, Brockett and Crouch [1995]. ¥

We may also endow any orbit with the right-invariant metric

gnl([x, a], [x, b]) = −κ(ax, Jbx) (7.4.14)

where J is a positive self-adjoint operator on the algebra. Then we have

Corollary 7.4.7. The geodesic equations on an adjoint orbit endowedwith the left invariant metric 7.4.14 are

p = [p, J−1[p, x]]x = [x, J−1[p, x]] . (7.4.15)

We shall consider this right invariant case from the optimal control pointof view in the next section.

We note also that, as expected, in the bi-invariant case with V = 0 it ispossible to explicitly compute the solutions of 7.4.12 despite their strongcoupling. This follows from

Lemma 7.4.8. [p, x] is conserved along the flow of 7.4.12.

This is proved by a simple computation. However proving complete inte-grability in the Hamiltonian sense, i.e. finding a complete set of commutingintegrals, is by no means easy. This is the content of Thimm [1981] andBloch, Brockett and Crouch [1995].

188 7. Optimal Control

7.4.1 Optimal Control on Symmetric Spaces

The equations discussed above are not only well defined on adjoint orbitsbut also on general symmetric spaces where the tangent vectors to thespace are given in the form a suitable bracket – this includes the complexand real Grassmannians of q-planes in n+ 1-space Gq,n+1(C) or Gq,n+1(R)and in particular the spheres.

This may be seen as follows:The complex Grassmannian is given by

U(n+ 1)/U(q)× U(p), q + p = n+ 1, q ≤ p (7.4.16)

and the real Grassmannian by

SO(n+ 1)/SO(q)× SO(p), q + p = n+ 1, q ≤ p (7.4.17)

where U(n) is the unitary group and SO(n) the special orthogonal group. Ineither case let g = k⊕m be the Lie algebra decomposition corresponding toG/K. We may thus represent a point in the complex (real) Grassmannianby a matrix

Q =[

0 Q−Q∗ 0

](7.4.18)

in m where Q is a complex (real) p × q matrix of full rank and Q∗ is itsHermitian conjugate (transpose). A point in k may be represented by thematrix

K =[K1 OO K2

](7.4.19)

where K1 ∈ u(p)(so(p)) and K2 ∈ u(q)(so(q)). Define P to be a similarlypartitioned matrix. Then we have

Proposition 7.4.9. Tangent vectors to the Grassmannian may be repre-sented by matrices of the form

[Q, K]

.

Proof. A curve in the Grassmannian through the point Q may be givenby

e−KtQeKt.

Note that the given curve simply provides an orthogonal (or unitary)transformation of the rows and columns of Q.

Differentiating at t = 0 gives the result. ¥

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 189

Now, since tangent vectors are given by brackets, just as in the case oforbits, a normal metric may be defined. Repeating the proof of Theorem3.3 gives

Proposition 7.4.10. The geodesic equations on the real or complex Grass-mannian are given by

˙Q = [Q, [P , Q]]˙P = [P , [P , Q]] . (7.4.20)

where Q is given by 7.4.18 and similarly for P .

Note also that for a symmetric space [k, k] ⊂ k, [k,m] ⊂ m and [m,m] ⊂ k

and since Q, P ∈ m the equations are naturally well defined.In fact the formalism developed here can be combined with the work

of Thimm [1981] to give a very explicit proof of complete integrabilityof the geodesic flow on symmetric spaces such as the real and complexGrassmannians. In particular it is possible to derive explicitly a completeset of commuting flows and to prove their involutivity. This is the subjectof a forthcoming paper, Bloch, Brockett and Crouch [1995].

This should be contrasted with the observation of Brockett regarding theoptimal control equations for 7.4.2 with V (x) = 1/2||[x, n]||2. He observedthat the optimal flow under suitable conditions was given by x = [x, [x, n]]and in the particular case of x being tridiagonal one obtains the integrableToda lattice equations in Flaschka’s form (see Bloch [1990]). Integrabilityof optimal control problems is a rich subject (see for example Faybusovich[1988]).

As in the variational problem on adjoint orbits discussed in section 2the double, double bracket equations on Grassmannians are solutions to anatural optimal problem. For conciseness of the exposition we will beginby considering a very general optimal control problem on symmetric spaceswhich we will then reduce to various special cases.

Theorem 7.4.11. Let Q be a p× q complex matrix and let U ∈ u(p) andV ∈ u(q). Let JU and JV be constant symmetric positive definite operatorson the space of complex p × p and q × q matrices respectively and let <,> denote the trace inner product < A,B >= TrA∗B, where A∗ is theHermitian conjugate.

Consider the optimal control problem

minU,V

∫1/4< U, JUU > + < V, JV V >dt (7.4.21)

subject to

Q = UQ−QV. (7.4.22)

190 7. Optimal Control

Then the optimal controls are given by

U = J−1U (PQ∗ −QP ∗)

V = J−1V (P ∗Q−Q∗P ) . (7.4.23)

and the optimal evolution of the states Q and costates P is given by

Q = J−1U (PQ∗ −QP ∗)Q−QJ−1

V (P ∗Q−Q∗P )

P = J−1U (PQ∗ −QP ∗)P − PJ−1

V (P ∗Q−Q∗P ) . (7.4.24)

Proof. Form the Hamiltonian

H(Q,P,U, V ) =< P,UQ−QV > −1/4 < U, JUU > −1/4 < V, JV V > .(7.4.25)

To find the optimal control we differentiate the Hamiltonian with respectto U and V in the directions Y and Z respectively and set equal to zero.This yields

< P, Y Q > −1/4 < Y, JUU > −1/4 < U, JUY >= 0

and

< P,−QZ > −1/4 < Z, JV V > −1/4 < V, JV Z >= 0 .

We have < Y, JUU > =< U, JUY > and < U, JUY > =< Y, JUU > andthus along the optimal trajectory

< P, Y Q > =< P, Y Q > . (7.4.26)

Now

< P, Y Q >= 1/2< Y,PQ∗ >− 1/2 < Y,QP ∗ >

and

< P, Y Q > = 1/2 < Y,PQ∗ > −1/2< Y,QP ∗ > .

Thus, using the fact that < P, Y Q > is real, along the optimal trajectorywe have

< P, Y Q > = 1/2 < P, Y Q > +1/2< P, Y Q >

= 1/4< Y,PQ∗ −QP ∗ >+ 1/4 < Y,PQ∗ −QP ∗ >= 1/4 < Y, JUU > +1/4< Y, JUU > . (7.4.27)

Hence

JUU = PQ∗ −QP ∗.

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 191

Similarly for V .Now the equations for Q and P are given by

< Z, Q > =∇PH(Z) =< Z,UQ−QV >

< P , Z > = −∇QH(Z) = − < P,UZ − ZV >=< UP − PV,Z > ,(7.4.28)

and hence the result. ¥

We remark that this result does not preclude the existence of conjugatepoints.

We have the immediate corollary:

Corollary 7.4.12. For JU and JV the identity the optimal control equa-tions for the problem 7.4.21 subject to 7.4.22 are

Q = PQ∗Q+QQ∗P − 2QP ∗QP = 2PQ∗P −QP ∗P − PP ∗Q . (7.4.29)

Further, we have

Corollary 7.4.13. The equations 7.4.24 are given by the double doublebracket equations

˙Q = [Q, J−1[P , Q]]˙P = [P , J−1[P , Q]] . (7.4.30)

where J is the operator diag(JU , JV ).

The proof is a computation.Note that the equations (7.4.30) or (7.4.24) give the geodesic equations

on the complex Grassmannian (or real Grassmannian in the real case) withrespect to right invariant “normal” metric.

As an example in the current setting we write explicitly the geodesic flowon the sphere Sn.

Recall (see e.g. Moser [1980]) that the geodesic motion on Sn may bewritten as follows:

Let q = [q1, · · · , qn+1]T ∈ Rn+1 with Euclidean norm ||q|| = 1 representan element of Sn. Then the geodesic flow can be found by setting q = λqwhere λ is chosen so that ||q|| is compatible with the flow. This implies< q, q >= 0 and < q, q > +||q||2 = 0. Thus λ = −||q||2 and the geodesicflow is given by

q = −||q||2q. (7.4.31)

Letting p = [p1, · · · , pn+1]T ∈ Rn+1 this may be viewed as a Hamil-tonian system restricted to ||q|| = 1 , < q,p >= 0. With Hamiltonian

192 7. Optimal Control

H = 1/2||q||2||p||2 we get the flow

q =(∂H

∂p

)T= p p = −

(∂H

∂q

)T= −||p||2q (7.4.32)

In our current setting we have

Proposition 7.4.14. Let

P =

0 · · · 0 p1

......

......

0 · · · 0 pn+1

−p1 · · · −pn+1 0

Q =

0 · · · 0 q1

......

......

0 · · · 0 qn+1

−q1 · · · −qn+1 0

(7.4.33)

where we normalize ||q|| = 1 and < q,p >= 0. Then the flow 7.4.12 yieldsthe geodesic flow 7.4.33.

The proof is a computation.Note that this result may be obtained directly from the variational prob-

lem. In this case Q ≡ q is n+ 1× 1 and real and similarly for P ≡ p. U isthen n× n, while V = 0.

Then we have

Corollary 7.4.15. The optimal control problem

minU

∫1/4(UTU)dt (7.4.34)

subject to

q = Uq (7.4.35)

where U ∈ so(n) restricts to the geodesic flow on the sphere.

Proof. The optimal flow equations are then

q = p(qTq)− (qTp)qp = (qTp)p− q(pTp) (7.4.36)

which are precisely the equations 7.4.33. It is easy to check that the timederivatives of qTq and qTp are conserved along the flow. Normalizing asin the proposition we indeed remain on the sphere. ¥

This flow is completely integrable and again provides an example of anintegable optimal control problem. Details of the proof of integrability canbe found in Thimm [1981] or Bloch, Brockett and Crouch [1995].

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 193

We now discuss the SO(n) rigid body equations in our setting. Theseequations may again be obtained as a special case of our general optimalcontrol problem.

We recall the rigid body equations on SO(3) (or generally on SO(n) orany compact Lie group – see e.g. Ratiu [1980]) may be written (in rightinvariant form) as

Q = ΩQM = [Ω,M ] (7.4.37)

where Q ∈ SO(3) denotes the configuration space variables, Ω ∈ so(3) isthe angular velocity, and M = JΩ = ΛΩ + ΩΛ is the angular momentum.Here J is a symmetric positive definite operator defined by the diagonalpositive definite matrix Λ.

We now consider the following equations:

Q = ΩQP = ΩP (7.4.38)

where Ω = J−1M and M = PQT −QPT . We then can easily check that

Proposition 7.4.16. The equations 7.4.38 reduce to the rigid body equa-tions 7.4.37.

Proof. Differentiating M = PQT −QPT and using the equations 7.4.38gives the second of equations 7.4.37. ¥

Conversely, given the rigid body equations 7.4.37 we may solve for thevariable P in the expression

M = PQT −QPT

in a neighborhood of M = 0. Locally

P =(esinh−1 M/2

)Q . (7.4.39)

This follows from the observation that

M = esinh−1 M/2 − e− sinh−1 M/2 .

For so(n) however, sinh is many to one, so the two representations 7.4.37and 7.4.38 are not entirely equivalent. However the form 7.4.38 of the rigidbody equations does arise from the optimal control problem as follows.

We now take Q to be real, n× n and lying in SO(n). V is now zero andU in so(n). We have

194 7. Optimal Control

Corollary 7.4.17. The optimal control problem

minU

∫1/4 < U, JU > dt (7.4.40)

subject to

Q = UQ, (7.4.41)

where U ∈ so(n) yields the rigid body equations 7.4.38.

The optimal controls in this case are given by

U = J−1(PQT −QPT ) (7.4.42)

Now the rigid body equations may be given as singular case of the doubledouble bracket equations discussed earlier for the general optimal controlproblem.

Let

Q =[

0 Q−QT 0

](7.4.43)

as before and similarly for P . Note that these matrices now lie in so(2n)and each block lies in SO(n).

Corollary 7.4.18. The generalized rigid body equations on SO(n) aregiven by the double double bracket equations 7.4.30 in the case Q and P liein SO(n), JU = J , and the operator J−1

V = 0.

Note that the reduced generalized rigid body equations (the dynamics)are completely integrable (see e.g. Ratiu [1980]).

7.4.2 Optimal Control and the Toda Flow

In this section we introduce an optimal control problem which yields thefull Toda flow as described originally in Faybusovich [1988]. We follow herethe treatment in Bloch and Crouch [1997]. This problem may be viewed asa control problem on an adjoint orbit of lower triangular matrices.

We begin by introducing some specialized notation. Let G = gl(n) denotethe Lie algebra of n× n matrices (with corresponding Lie group GL(n) ofinvertible n× n matrices).

Definition 7.4.19. For A ∈ gl(n), let A = A+ + A0 + A− where A+ isthe strictly upper part of A, A0 the diagonal part, and A− is the strictlylower part. Let

πl(A) = A− +AT+ +A0 (lower)

πk(A) = A+ −AT+ (skew)

πk⊥(A) = A− +A0 +AT− (symmetric)

πl⊥(A) = A+ −AT− (strictly upper).

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 195

The reason for the above notation is as follows (see Symes [1980]):Endow G with the inner product 〈A,B〉T = TrATB. We observe

G = L ⊕Kwhere L is the subalgebra of lower triangular matrices and K the subalgebraof skew-symmetric matrices. L is the Lie algebra of L, the lower triangulargroup and K is the Lie algebra of K, the group of orthogonal matrices.

Now denote by ⊥ the perpendicular subspace under the scalar product.Then

L⊥ = x : Tr(xT y) = 0,∀y ∈ Lis the algebra of strictly upper triangular matrices and

K⊥ = x : Tr(xT z) = 0,∀z ∈ Kis the space of all symmetric matrices.

Hence we can write

G ' G∗ = L⊥ ⊕K⊥,and we can make the identification

L∗ ' K⊥

K∗ ' L⊥.We now give some preliminary results that will be useful in the main

theorem.

Lemma 7.4.20. For any matrix S ∈ G and L ∈ Lπk⊥(LTS) = πk⊥(LTπk⊥S)πk⊥(SLT ) = πk⊥(πk⊥(S)LT ).

Let F : G → G be analytic and consider

f(PQT ) = TrF (PQT ).

Let

FAf(A)(R) = limh→0

f(A+ hR)− f(A)h

= 〈R,∇f(A))

for A,R ∈ gl(n). Hence

FQf(PQT )(R) = 〈PRT ,∇f(PQT )〉= 〈RPT ,∇f(PQT )T 〉= 〈R,∇f(PQT )TP 〉

FP f(PQT )(R) = 〈RQT ,∇f(PQT )〉= 〈R,∇f(PQT )Q〉.

Hence, we have

196 7. Optimal Control

Lemma 7.4.21. Let f(PQT ) = TrF (PQT ). Then the Hamiltonian flowswith respect to the canonical structure on T ∗G are given by

Q =∇f(PQT )QP = −∇f(PQT )TP,

respectively.

The main result in Bloch and Crouch [1997] can now be stated in thefollowing way. Consider the optimal control problem

minU

∫ T

0

12 〈U,U〉T dt

subject to: X = πl(U)X, X(0) = X0, X(T ) = XT ,where X ∈ L, U ∈ G.

(7.4.44)

Theorem 7.4.22. For the optimal control problem (7.4.44), the optimalcontrols are given by

U = πk⊥(RXT )

where R is the costate vector and the corresponding extremal flow is givenby

X = πl(πk⊥(RXT ))X (7.4.45)R = −πk⊥(πTl (πk⊥(RXT ))R).

Equations (7.4.45) are Hamiltonian with respect to the canonical structureon T ∗G, with corresponding Hamiltonian function

H(R,X) = 12 〈πk⊥(RXT ), πk⊥(RXT )〉T . (7.4.46)

From this result, and Lemma 7.4.20, we observe that if S = πk⊥(R) wemay rewrite H in (7.4.46) in the form:

H = 12 〈πk⊥(SXT ), πk⊥(SXT )〉T . (7.4.47)

Using Lemma 7.4.20 once more, the Hamiltonian equations may be rewrit-ten in the form

X = πl(πk⊥(SXT ))X (7.4.48)S = −πk⊥(πTl (πk⊥(SXT ))S).

Now set

A = πk⊥(SXT ). (7.4.49)

7.4 Optimal Control on Lie Algebras and Adjoint Orbits 197

Then the equations (7.4.48) are of the form

X = πl(∇f(A))XS = −πk⊥(πTl (∇f(A))S)

where f(A) = 12Tr(A2).

Now we compute

A = πk⊥(SXT + SXT )= πk⊥(−πk⊥(πTl (∇f)S)XT )

+πk⊥(SXTπTl (∇f))= −πk⊥(πTl (∇f)SXT )

+πk⊥(SXTπTl (∇f))(by Lemma (7.4.20)

= πk⊥([A, πTl (∇f)]).

Hence

A = πk⊥([A, πTl (∇f)]). (7.4.50)

But since

∇f(A) = πl(∇f(A)) + πk(∇f(A))

and ∇f(A) ∈ K⊥ (for any invariant polynomial in A), we have

∇f(A) = πTl (∇f(A))− πk(∇f(A)).

Observing also that [A,∇f(A)] = 0, we obtain

A = πk⊥ [A, πk(∇f(A))]= [A, πk(∇f(A))]

(since [K⊥,K] ⊂ K⊥)= [A,∇f(A)− πl∇f(A)]= [πl(∇f(A)), A].

Thus for f(A) = 12Tr(A2), we obtain from (7.4.48) and (7.4.49) the follow-

ing set of equations which are equivalent to the extremal flow (7.4.45), andrepresent the (augmented) full Toda flow:

X = πl(A)X (7.4.51)A = [A, πk(A)] = [πl(A), A].

The equations (7.4.45) and (7.4.48) were originally derived in a completelydifferent fashion in the work of Symes [1980]. We note the special form ofthe full reduced Toda flow equation (7.4.50)

A = πk⊥([A, πTl (∇f)]). (7.4.52)

198 7. Optimal Control

We observe that this evolves naturally in a coadjoint orbit of L. We cansee this by considering tangent vectors to an orbit. These are of the formad∗WA for W ∈ L and A ∈ L∗ ' K⊥. Let A = 〈S, ·〉T for S ∈ K⊥, and letV ∈ L. Then

ad∗W (A)(V ) = −A(adWV ) = −A([W,V ])= −〈S, [W,V ]〉T = −〈[WT , S], V 〉T= −〈[WT , S], πlV 〉T

since V ∈ L= 〈πk⊥ [S,WT ], V 〉T .

Hence indeed the right hand side of (7.4.52) is a tangent vector to a coad-joint orbit. The classical tridiagonal Toda flow lies on a low rank nongenericorbit of this type.

7.5 Kinematic/subRiemannian OptimalControl Problems

7.5.1 The kinematic subRiemannian optimal controlproblem

Here we consider the optimal control of singular kinematic control prob-lems – singular in the sense that the number of controls is less than thenumber of states. The problem is referred to as subRiemannian in thatit gives rise to a geodesic flow with respect to a singular metric (see thework of Strichartz [1986,87] and Montgomery [1990]. This problem has aninteresting history in control theory (see Brockett [1973], [1981], Baillieul[1975]) and is of interest in nonholonomic mechanics because such systemsrepresent nonholonomic control systems with velocity controls. There hasbeen much work in this area: for example work Murray and Sastry [1990],Lafferiere and Sussmann [1991], Brockett and Dai [1982], Bloch and Crouch[1992], Bloch, Crouch and Ratiu [1994].

We now consider a specific class of optimal control problems made famousby the initial work of Brockett [1981]. We consider control systems of theform:

x =k∑i=1

Xiui, x ∈Mn, u ∈ Ω ⊂ Rk (7.5.1)

where Ω contains an open subset which contains the origin, and the vectorfields F = X1, . . . , Xk are complete. Setting

DF (x) =

Xx(x);X =

k∑i=1

αiXi, αi ∈ R

,

7.5 Kinematic/subRiemannian Optimal Control Problems 199

then x 7→ DF (x) defines a distribution on M , whose involutive closure isjust the distribution DL defined by the Lie algebra L of vector fields gen-erated by F . (We assume that all vector fields in L are also complete). Wesay that DL has the spanning property if DL(x) = TxM for all x ∈ M .System (7.5.1) is said to have the controllability property, if whenever x0,xT are two points on M , there exists a piecewise constant control u(·) on[0, T ] such that the corresponding solution x(·) of (7.5.1) satisfies x(0) = x0,x(T ) = xT . It is a fundamental fact of control theory (see Sussmann [1973])that if DL has the spanning property then (7.5.1) has the controllabilityproperty. (In the analytic case this condition is also necessary for control-lability.)

Thus in case DL has the spanning property the following optimal controlproblem is well posed:

minu(·)

∫ T

0

12

k∑i=1

u2i (t)dt

subject to: dynamics (7.5.1) and x(0) = x0, x(T ) = xT .

(7.5.2)

To view this as a constrained variational problem we make some addi-tional regularity assumptions, which are not necessary, but even when theyhold produce a very rich class of problems.

Assumption

(i) For the problem defined by (7.5.1) and (7.5.2) the distribution DL

satisfies the spanning property.

(ii) The dimension of DF is constant on M and equal to k. (Thus thevector fields X1 · · ·Xk are everywhere independent).

(iii) There exist exactly n− k = m one forms on M ω1 · · ·ωm, such thatthe co-distribution

D⊥F (x) = ω ∈ T ∗xM ; ωDF (x) = 0

is spanned by ω1, . . . , ωm everywhere. (This condition implies thatM is parallelizable.)

Since DF has constant dimension on M we may define a norm on eachsubspace DF (x); if X ∈ DF (x), and X =

∑ki=1 αiXi(x) then

|X| =k∑i=1

α2i .

This norm defines an inner product on DF (x), denoted 〈·, ·〉x, which canbe extended to a metric on M . The optimal control problem (7.5.2) is

200 7. Optimal Control

now equivalent to the following constrained variational problem when theassumptions (i), (ii), (iii) hold:

minx(·)

12

∫ T

0

〈x, x〉xdt (7.5.3)

subject to: x(·) is a piecewise C1 curve in M such that (7.5.4)x(0) = x0, x(T ) = xT and ωi(x)(x) = 0, 1 ≤ i ≤ m.

This problem is often referred to as the sub-Riemannian geodesic prob-lem, to distinguish it from the Riemannian geodesic problem, in which theconstraints are absent. These problems were studied by Griffiths [1983]from the constrained variational viewpoint, before they gained much widerrecognition, from the optimal control viewpoint by Brockett [1981]. This“singular” geodesic problem is an example of the situation described in(2.47 iv). It is the case however that in the sub-Riemannian geodesic prob-lem, abnormal extremals play an important role, see Sussmann [1992]. Seealso work by Hermann [1962], and Strichartz [1983], [1989].

We note that the solution of the model problem above, for normal ex-tremals, involves adding the constraints to the Lagrangian by Lagrangemultipliers, resulting in necessary conditions in the form of constrainedvariational mechanics with L(q, q) = 1/2〈q, q〉q. Perhaps unfortunately, insome of the Russian literature, Vershik and Gershkovich [1988] , for exam-ple, distributionsDF for whichDL has the spanning property, are said to becompletely nonholonomic, and so the system (7.5.1) is sometimes said to bea nonholonomic control system. Clearly, there is another system which hasthe right to be called the non-holonomic control system generated by theLagrangian L = 1/2〈q, q〉q and the constraints: the corresponding dynamiccontrol problem.

However, the term completely nonholonomic distribution is also well de-served as we briefly demonstrate. Suppose that one of the forms ωk is exact,ωk = dP , for some function P on M . By the assumption (iii) dP (x) 6= 0for all x in M , and

0 = ωk([Xi, Xj ]) = dP ([Xi, Xj ]) = Xi(Xj(P ))−Xj(Xi(P ))

since dP (Xi) ≡ dP (Xj) ≡ 0. It follows that the distribution DL is annihi-lated by ωk everywhere on M , and so cannot have the spanning property.Hence the assumption (i) implies that none of the forms ωk, 1 ≤ k ≤ m,are exact, and so none of the constraints may be absorbed as holonomicconstraints.

The singular nature of the sub-Riemannian geodesic problem is mani-fested in many ways, such as the existence of distinct abnormal extremals,and the singular nature of the sub-Riemannian geodesic ball, as first inves-

7.5 Kinematic/subRiemannian Optimal Control Problems 201

tigated by Brockett [1981]. Defining a metric on M by setting

d(x0, xT ) = minx(·)

∫ T

0

|x|dt, x ∈ DF (x), x(0) = x0, x(T ) = xT

BFε (x0) = x ∈M ; d(x, x0) ≤ ε

then the sub-Riemannian geodesic ball SFε (x0) is simply the boundary ofBFε (x0). In Brockett [1981] the Heisenberg sub-Riemannian geodesic prob-lem was considered, which expressed as a control problem becomes:

minu(·)

∫ T

0

12

(u21 + u2

2)dt,

subject to: x1 = u1

x2 = u2

x3 = x1u2 − x2u1

x(0) = x0, x(T ) = xT .

In this case F = (∂/∂x1) − x2(∂/∂x3), (∂/∂x2) + x1(∂/∂x3), while D⊥Fis spanned by ω = x1dx2 − x2dx1 − dx3. The Lie algebra L is simplythe Heisenberg Lie algebra. SFε (0) has a singularity along the x3 axis. Thisclass of problem continues to invoke a great deal of interest, especially in theareas of establishing when abnormal extremals are optimal and obtaininga precise description of SDε (x0).

We shall return to the notion of abnormal extremals later.We now set up the problem on a Riemannian manifold as follows, al-

though the cases of most interest usually have more structure, such as agroup structure. In the following subsections we will look at such specialcases. We follow here the approach of Bloch, Crouch and Ratiu [1994].

LetMn be a Riemannian manifold of dimension n with metric denoted by< ·, · >. The corresponding Riemannian connection and covariant deriva-tive will be denoted by ∇ and D/∂t respectively. Now assume that Mis such that there exist smooth vector fields X1(q), · · · , Xn(q) satisfying< Xi(q), Xj(q) >= δij , an orthonormal frame for TqM for all q ∈M . Thisof course limits the class of manifolds we consider, but is satisfied for themain case of interest to us, M a Lie group G.

We now define the kinematic control system on M

dq

dt=

m∑i=1

uiXi(q), m < n . (7.5.5)

The singular optimal control problem for 7.5.5 is defined by

minu

∫ T

0

12

m∑i=1

u2i (t)dt; q(0) = q0 q(T ) = qT (7.5.6)

subject to 7.5.5.

202 7. Optimal Control

This may be posed as a variational problem on M as follows:Define the constraints

ωk

(dq

dt

)=< Xk,

dq

dt>= 0, m < k ≤ n (7.5.7)

and let

Zt =n∑

k=m+1

λk(t)Xk (7.5.8)

where the λk are Lagrange multipliers. By the orthonormality of the Xi

the optimal control problem then becomes

minqJ(q) = min

q

∫ T

0

(12<dq

dt,dq

dt> + < Zt,

dq

dt>

)dt (7.5.9)

< Zt,dq

dt> = 0 . (7.5.10)

We now briefly derive the necessary conditions for the regular extremalsof this variational problem following Milnor [1963] and Crouch and Silva-Leite [1991]. (For interesting recent work on abnormal extremals see forexample Bryant and Hu [1993], Montgomery [1992], and Sussmann [1992].)

Firstly we have to define the variations we are going to use:The tangent space to the space Ω of C2 curves satisfying the boundary

conditions of 7.5.6 is denoted by TqΩ. It is the space of C1 vector fieldst→Wt along q(t) satisfying W0 = 0 = WT . The curve t→ DWt

∂t in TM iscontinuous. Exponentiating a vector field in TqΩ we obtain a one-parametervariation of q:

α : [0, T ]× (−ε, ε)→M, (7.5.11)αu(t) = α(t, u) = expq(t)(uWt) (7.5.12)

where exp is the exponential mapping on M . Note that αu(0) = q(0) = q0,αu(T ) = q(T ) = qT , α0(t) = q(t), ∂α0(t)

∂u = Wt, 0 ≤ t ≤ T .The necessary conditions for regular extremals are obtained from

d

duJ(αu)|u=0 = 0 (7.5.13)

where

J(αu) =∫ T

0

(12

⟨∂αu∂t

,∂αu∂t

⟩+⟨Zt(αu),

∂αu∂t

⟩)dt . (7.5.14)

7.5 Kinematic/subRiemannian Optimal Control Problems 203

Now

DJ(αu)du

∣∣∣∣u=0

=∫ T

0

(⟨dq

dt,DWt

∂t

⟩+⟨∇WtZt,

dq

dt

⟩+⟨Zt,

D

∂tWt

⟩)dt

=∫ T

0

(−⟨D

dtVt,Wt

⟩−⟨D

∂tZt,Wt

⟩− 〈∇ZtVt,Wt〉+ 〈[Wt, Zt], Vt〉) dt (7.5.15)

where Vt = dqdt =

m∑i=1

vi(t)Xi(q). Note that in this computation we use

∇WZ = ∇ZW + [W,Z] and Z[< V,W >] =< ∇ZV,W > + < V,∇ZW >.

7.5.2 Necessary Conditions on a CompactSemisimple Lie Group

In this section we show that is the underlying manifold is a compactsemisimple Lie group, then the subRiemmanian optimal control problemdiscussed in the previous section may be reduced to a computation in theLie algebra.

Now let M = G, G a compact semisimple Lie group, with Lie algebra g,and let << ·, · >>= −1

2κ(·, ·) where κ is the Killing form on g.Let J be a positive definite linear mapping J : g→ g satisfying

<< JX, Y >> = << X, JY >> (7.5.16)<< JX,X >> ≥ 0 (= 0 if and only if X = 0) (7.5.17)

Now we can define a right invariant metric on G as follows:If X,Y ∈ g and Rg is right translation on G by g ∈ G, then Xr

g =Xr(g) = Rg∗X and Y rg = Y r(g) = Rg∗Y are corresponding right invariantvector fields. Now

< Xr(g), Y r(g) ><< X, JY >> (7.5.18)

defines a right-invariant metric on G. Corresponding to the right invariantmetric < , > there is a unique Riemannian connection ∇ (see e.g.Nomizu [1956] and earlier sections). ∇ defines a bilinear form on g:

(X,Y )→ ∇XY =12[X,Y ] + J−1[X, JY ] + J−1[Y, JX], X, Y ∈ g

(7.5.19)

and the expression for ∇ on right-invariant vector fields on G is

(∇XrY r)(g) = (∇XY )rg . (7.5.20)

We now show how to reduce the variational problem to one in the Liealgebra:

204 7. Optimal Control

Choose an orthonormal basis ei on g, << ei, Jej >>= δij , and extend itto a right invariant orthonormal frame on TgG, Xi(g) = Rg∗ei ≡ Xir(g).

We consider again the computation of dduJ(αu) |u=0. Suppose in g

Vt =m∑i=1

vi(t)ei, Vt =m∑i=1

vi(t)ei (7.5.21)

and similarly for Wt =n∑i=1

wi(t)ei and Zt. Then at the group level we have

for dgdt = V rt ,

DV rtdt

=D

dt

(m∑i=1

vi(t)Xi

)= (Vt +∇VtVt)rg = (Vt + J−1[Vt, JVt])rg (7.5.22)

by 7.5.19 and 7.5.20, and

DZrtdt

= (Zt +∇VtZt)rg. (7.5.23)

Now, by 7.5.18,

d

duJ(αu) |u=0 =

∫ T

0

(− << Vt + J−1[Vt, JVt] + Zt +∇V tZt +∇ZtVt, JWt >>

+ << [Wt, Zt], JVt >>)dt . (7.5.24)

Using,

<< [X,Y ], Z >> + << Y, [X,Z] >>= 0 (7.5.25)

the necessary conditions are thus

Vt + J−1[Vt, JVt] + Zt +∇VtZt +∇ZtVt + J−1[JVt, Zt] = 0 . (7.5.26)

By 7.5.19

∇VtZt +∇ZtVt = J−1[Vt, JZt] + J−1[Zt, JVt] .

Hence the necessary conditions on g are

Vt + J−1[Vt, JZt] + Zt + J−1[Vt, JVt] = 0 (7.5.27)

with the constraint

<dg

dt, Zt >=<< Vt, JZt >>= 0 . (7.5.28)

7.5 Kinematic/subRiemannian Optimal Control Problems 205

Equations 7.5.27 are identical to the equations (3.7) of Brockett [1973]in the case J = I. We can see this as follows:

Write the system 7.5.5 as

dg

dt=

m∑i=1

uiBig, m < n, (7.5.29)

where Bi ∈ g and the Xi(g) = Big are thus right invariant vector fields

on G. Then Vt =m∑i=1

ui(t)Bi. Now set Lt = Vt + Zt. Then equation 7.5.27

becomes

Lt = J−1[JLt, Vt] (7.5.30)

or

Lt =m∑i=1

uiJ−1[JLt, Bi] . (7.5.31)

Setting J = I we recover precisely Brockett’s equations (in the case ofzero drift). Note also that in the case J = I our equations 7.5.27 assumethe symmetric form

Vt + [Vt, Zt] + Zt = 0 . (7.5.32)

Brockett obtained his equations by applying the maximum principle toLie groups, while we have taken a direct variational approach. The ap-proaches are of course equivalent (see also the next section).

7.5.3 The Case of Symmetric Space Structure

Suppose now that G/K is a Riemannian symmetric space, G as above, Ka closed subgroup of G with Lie algebra k. Then g = p ⊕ k with [p, p] ⊂ k,[p, k] ⊂ p, [k, k] ⊂ k and << k, p = 0 >>. We now want to consider thenecessary conditions 7.5.27 in this case. We shall see that they simplify inan intriguing fashion, giving us a singular case of the so-called generalizedrigid body equations.

Assume that e1, . . . , em is a basis for p and em+1, . . . en is a basis fork. Suppose that J : p → p and J : k → k. Then << Vt, JZt >>= 0 for

Zt =∑ni=m+1 λi(t)ei ∈ k and Vt =

m∑k=1

vi(t)ei ∈ p.

Since Vt + J−1[Vt, JZt] ∈ p and Zt + J−1[Vt, JVt] ∈ k, the necessaryconditions 7.5.27 become

Vt = J−1[JZt, Vt]Zt = J−1[JVt, Vt] (7.5.33)

206 7. Optimal Control

or, if we define Pt = JVt and Qt = JZt

Pt = [Qt, J−1Pt]Qt = [Pt, J−1Pt] . (7.5.34)

We will now show that equations 7.5.34 are Hamiltonian with respect tothe Lie-Poisson structure on g.

Recall that for F , H functions on g, their (−) Lie-Poisson bracket isgiven by

F,H(X) = − << X, [∇F (X),∇H(X)] >>, X ∈ g, (7.5.35)

where dF (X) · Y =<< ∇F (X), Y >>.For H(X) a given Hamiltonian, we thus have the Lie-Poisson equations

F (X) = F,H(X). Letting F (X) =<< A,X >>, A ∈ g, we obtain

<< A, X >> = − << X, [A,∇H(X)] >>= << A, [X,∇H(X)] >> (7.5.36)

and hence

X = [X,∇H(X)] . (7.5.37)

For H(M) = 12 << M,J−1M >>, M ∈ g and J as in the previous

subsection, we obtain the generalized rigid body equations

M = [M,J−1M ] . (7.5.38)

Now for X = P + Q ∈ p ⊕ k, let H(X) = H(P ) = 12 << P, J−1P >>,

P ∈ p. Then ∇H(X) = J−1P ∈ p and equations 7.5.37 become

(Q+ P )· = [Q+ P, J−1P ] (7.5.39)

or

P = [Q, J−1P ]Q = [P, J−1P ] (7.5.40)

precisely equations 7.5.34.Thus equations 7.5.34) are Lie-Poisson with respect to the “singular”

Hamiltonian H(P ). Summarizing then, we have

Theorem 7.5.1. The optimal trajectories for the singular optimal con-trol problem 7.5.5, 7.5.6 on a Riemannian symmetric space are given byequations 7.5.34. These equations are Lie Poisson with respect to a singularrigid body Hamiltonian on g.

7.5 Kinematic/subRiemannian Optimal Control Problems 207

We see therefore that we can obtain the singular optimal trajectoriesby letting J |k→ ∞ in the full rigid body Hamiltonian H(X) = 1

2 <<X, J−1X >>, thus obtaining the singular Hamiltonian H(P ) = 1

2 <<P, J−1P >>.

This observation also enables us to obtain the singular rigid body equa-tions directly by a limiting process from the full rigid body equations. Thekey is the correct choice of angular velocity and momentum variables cor-responding to the Lie algebra decomposition g = p⊕ k.

In the notation of equation 7.5.34 we write an arbitrary element of g asM = JV + Q, JV ∈ p, Q ∈ k. Then the generalized rigid body equations7.5.38 become

JVt = [Qt, Vt] + [JVt, J−1Qt]Qt = [Qt, J−1Qt] + [JVt, Vt] . (7.5.41)

Letting J |k→∞ we obtain

JVt = [Qt, Vt]Qt = [JVt, Vt] . (7.5.42)

We note that this is a mixture between the Lagrangian and Hamiltonianpictures. While the variables in k are momenta (and should really be viewedas lying in k∗) the variables in p are velocities. These variables are not onlythe natural ones in which to take the limit in the full rigid body equations,but are natural from the point of view of the maximum principle, for thevariablesQ correspond to the constraints and therefore are naturally viewedas costates.

Note also that this reduction may also be viewed as a generalized Routhianreduction as in Marsden and Scheurle [1992]. The group variables are theonly ones which are Legendre transformed. This will be discussed furtherin a future paper.

As mentioned in the previous section, the necessary conditions abovemay also be derived directly from the maximum principle developed forLie groups (see e.g. Brockett [1973], Jurdjevic [1991]). The Hamiltonian inthe maximum principle of the system on a Lie group as discussed aboveis precisely 1

2 < Pt, J−1Pt >. This is just the sum of the Hamiltonians

corresponding to each of the vector fields Xi.We note also the following interpretation of the fact that J |k→∞ (due

to R. Brockett). Write M ∈ g now as M = JZ + P , Z ∈ k, P ∈ p. Thenthe Hamiltonian (in the maximum principle) becomes

H(M) =< M,J−1M > = < JZ + P, J−1(JZ + P ) >= < JZ,Z > + < P, J−1P > (7.5.43)

Letting J |k→ ∞ we see that the cost becomes infinite unless Z = 0, i.e.the constraints are satisfied.

208 7. Optimal Control

Example: We now consider a simple but non-trivial example: the sym-metric space SO(3)/SO(2).

In this case g = k ⊕ p becomes so(3) = so(2) ⊕ R2. Reflecting thisdecomposition we may represent matrices in so(3) as 0 −ω3 ω2

ω3 0 −ω1

−ω2 ω10

(7.5.44)

with the lower 2× 2 block in so(2).This example illustrates the importance of writing the optimal equations

in the natural variables M = JV + Q in order to understand the limitingprocess in equations 7.5.41 and 7.5.42.

We write 0 −J3ω3 J2ω2

J3ω3 0 −m1

−J2ω2 m1 0

(7.5.45)

Here Qt ∈ so(2) has “momentum” variable m1.Then equations 7.5.42 or

(JVt +Qt)· = [JVt +Qt, Vt] (7.5.46)

become for g = so(3)

m1 = (J2 − J3)ω2ω3

J2ω2 = −m1ω3

J3ω3 = m1ω2 . (7.5.47)

The full rigid body equations in these variables are

(JVt +Qt)· = [JVt +Qt, Vt + J−1Qt] (7.5.48)

which are for g = so(3)

m1 = (J2 − J3)ω2ω3

J2ω·2 =

(J3

J1− 1)ω3m1

J3ω3 =(

1− J2

J1

)m1ω2 (7.5.49)

which clearly limits to 7.5.47 as J1 →∞.Note that if we write the rigid body equations in the usual form

J(Vt + Zt)· = [J(Vt + Zt), Vt + Zt] , (7.5.50)

7.5 Kinematic/subRiemannian Optimal Control Problems 209

yielding for g = so(3)

J1ω1 = (J2 − J3)ω2ω3

J2ω2 = (J3 − J1)ω1ω3

J3ω3 = (J1 − J2)ω2ω1 , (7.5.51)

we see that the limiting process J1 →∞ makes no sense. The same is truefor the rigid body in the momentum representation.

We remark that this set of equations, despite its singular nature isstill integrable, for we still have two conserved quantities, the Hamilto-nian H(ω) = J2ω

22 + J3ω

23(= 1

2 < P, J−1P >) and the Casimir C(ω) =m2

1 + J22ω

22 + J2

3ω22 . (Recall that a Casimir function for a Poisson structure

is a function that commutes with every other function under the Poissonbracket.)

It is interesting to consider the case J = I. Equations 7.5.34 then become

Pt = [Qt, Pt]Qt = 0 . (7.5.52)

Hence Qt = Q is constant.Similarly, considering 7.5.33, we obtain

Vt = [Zt, Vt], Zt = Z constant . (7.5.53)

This is of course solvable: Vt = AdeZtV0 and ui(t) =<< ei, AdeZtV0 >>(see the equivalent expression (3.5) of Brockett [1973]). Consider again thecase SO(3)/SO(2). Since Vt ∈ R2 and Zt ∈ so(2) we may set

Z =

0 0 00 0 −φ0 φ 0

(7.5.54)

where φ is fixed. Then

eZt =

1 0 00 cosφt − sinφt0 sinφt cosφt

. (7.5.55)

Hence the optimal evolution of Vt (or equivalently the optimal controls) isgiven by rotation. This recovers precisely the result of Baillieul who indeedanalyzed the case J = I in dimension 3. (See Baillieul [1975], page III-5.)

7.5.4 Optimal control and a particle in a magneticfield

We begin by considering the connection between optimal control of theHeisenberg system and the motion of a particle in a magnetic field. This

210 7. Optimal Control

will be seen be a special case of the more general motion of a particle in amagnetic or Yang Mills field. The control analysis of the Heisenberg modelgoes back to Brockett [1981] and Bailleul [1975], while a modern treatmentof the relationship with a particle in a magnetic field may be found inMontgomery [1993] for example. A nice treatment of the pure mechanicalaspects of a particle in a magnetic field may be found in Marsden and Ratiu[1994].

Consider firstly the optimal control of the Heisenberg equations. Recallthat the Heisenberg control system is given by:

x = u

y = v

z = uy − xv = xy − yx . (7.5.56)

The last equation may be viewed as a nonholonomic constraint on the statespace of the system or as a connection on the bundle with base space R2

coordinatized by x and y and fiber z coordinatized by z.We now consider the optimal control problem:

min∫

(u2 + v2)dt (7.5.57)

subject to the equations 7.5.56.This may be reformulated as the constrained optimization problem

min∫

(x2 + y2 − λ(z − xy + yz))dt (7.5.58)

subject to

z = xy − yx . (7.5.59)

The Euler-Lagrange equation for z is

d

dt(λ) = 0 (7.5.60)

and hence λ is a constant.Thus the Euler-Lagrange equations for x and y become

x+ λy = 0y − λx = 0 . (7.5.61)

Now let v be the vector (x, y, 0), let λ be the vector (0, 0, λ) and let Ωbe the matrix 0 λ 0

−λ 0 00 0 0

7.5 Kinematic/subRiemannian Optimal Control Problems 211

Then the equations for x and y may be rewritten as

dvdt

+ Ωv = 0 (7.5.62)

or

dvdt

+ v × λ = 0 . (7.5.63)

. (Note also that z = xy − yx = −λ(xx+ yy) thus enabling us to solve forz in terms of x and y.)

That these equations are a particular case of planar charged particlemotion in a magnetic field may be seen by considerting the slightly moregeneral problem below.

We now consider again the optimal control problem:

min∫

(u2 + v2)dt (7.5.64)

but now subject to the equations.

x = u

y = v

z = A1u+A2v , (7.5.65)

where A1 and A2 are smooth functions of x and y.Then the Euler Lagrange equations for z again yields λ = const while

the equations for x and y are

x =λ

2(A2x −A1y)y

y =λ

2(A1y −A2x)x . (7.5.66)

Now let A be the vector (A1, A2, 0). Then

(∇×A)z = A2x −A1y ≡ Bz.

Hence the Euler Lagrange equations may be rewritten

x =λ

2Bz y

y = − λ

2Bzx , (7.5.67)

or, if v = (x, y, 0) and B = (0, 0, Bz), as

dvdt

+ v × λ

2B = 0 . (7.5.68)

212 7. Optimal Control

This is indeed the motion of a planar charged particle in a magnetic field(see e.g. Marden and Ratiu [1994]), where the Lagrange multiplier is iden-tified with a multiple of the charge.

This problem may also be solved via the maximum prinicple:Form the Hamiltonian

H = p1u+ p2v + p3(A1u+A2v)− u2

2− v2

2. (7.5.69)

The optimality conditions

∂H

∂u= 0 =

∂H

∂v

yield

u = p1 + p3A1

v = p2 + p3A2 . (7.5.70)

Hence the optimal Hamiltonian becomes

H =12(p1 +A1p3)2 + (p2 +A2p3)2 . (7.5.71)

This is the Hamiltonian for a particle in a magnetic field. Note that toobtain the sign in, say, Marsden and Ratiu [1994] we replace Ai by −Ai inthe z equation. Note also that p3 is a cyclic variable and hence is constantin time. This may be chose to have the value e/c, the charge over the speedof light, as in equations for the particle.

7.5.5 Rigid Extremals

A particularly interesting phenomenon that occurs in sub-Riemannian op-timal control problems is the existence of rigid extremals — i.e. isolatedextremals which admit no allowable variations. In such a situation, we canobtain an optimal solution to the optimal control problem which does notsatisfy the optimal control equations. We can rephrase this, as does Mont-gomery [ ], by saying that one has a (subRiemannian) geodesic which doesnot satisfy the geodesic equations.

We will consider here the example of Montgomery [ ]. The problemhowever has a rather interesting history — for some details on this see thepaper of Montgomery and, for example, Strichartz [1983], [1989].

We follow here the treatment in Montgomery [ ], casting things in thelanguage of optimal control theory.

Since we will be considering an optimal trajectory with cylindrical ge-ometry, it is best to phrase the problem in cylindrical coordinates in R3:(r, θ, z).

7.5 Kinematic/subRiemannian Optimal Control Problems 213

We consider then the optimal control problem

minu,v

∫(u2 + r2v2)dt

subject to

r = u

θ = v

z = −A(r)v (7.5.72)

where A(r) is a smooth function of r with a single nondegenerate maximumat r = 1. Thus we require dA

dr

∣∣r=1

= 0 and d2Adr2

∣∣r=1

< 0. We can take, forexample,

A =12r2 − 1

4r4.

The control vector fields are

X1 =∂

∂r

X2 =∂

∂θ−A(r)

∂z. (7.5.73)

The system is controllable since

[X1, X2] =−dAdr

∂z(7.5.74)

so X1, X2 and [X1, X2] span R3 everywhere except at r = 1. But for r = 1[X1, [X1, X2]] = −d2A

dr2∂∂z 6= 0 so the control vector fields still span R3.

The examples of Montgomery are the helices with pitch A(1). One cantake the curves (r, θ, z) = (1, θ,−A(1)θ). These curves clearly lie in theconstraint distribution. Montgomery shows that this helix is minimizing butdoes not satisfy the optimal control (or subRiemannian geodesic) equations.We will sketch these arguments here.

Consider firstly the geodesic equations. They are given in Cartesian formby equations (5.142) where

Bz =∂A2

∂x− ∂A1

∂y.

To compute Bz here observe that we need to replace A(r)dθ by A1dx +A2dy. Now

A1dx+A2dy = A(r)dθ (7.5.75)

implies (−∂A1

∂y+∂A2

∂x

)dx ∧ dy =

dA

drdr ∧ dθ (7.5.76)

214 7. Optimal Control

But dx ∧ dy = rdr ∧ dθ. Hence

1r

dA

dr=∂A2

∂x− ∂A1

∂y= Bz

For our given curve r = 1 and hence Bz is zero. But x and y are clearlynot zero. Hence this curve cannot satisfy the geodesic equations.

We now show that the helix is an isolated point in the space of all piece-wise C1 curves in the constraint distributins with fixed endpoints. Since itis isolated it is auomatically a local minimum. However, it is not isolatedin the C0 or H1 topologies. Montgomery however shows that it is still alocal minimum, at least for short enough arcs — see Montgomery [ ].

Now consider a curve γ in th constraint distribution connecting the points(x0, y0, 0) to (x0, y0, z1). The projected curve onto the x − y plane is thusclosed.

Since dz = −Adθ along the curve we have, using Stoke’s theorem,

z1 = −∫Adθ

= −∫ ∫

dA

drdrdθ

= −∫ ∫

B(r)rdrdθ

where ∆ is the region in the plane enclosed by the projected curve. Follow-ing our earlier magnetic analogy this quantity can be viewed as the fluxthrough the region by the plane enclosed by the curve.

Now recall that in our case

B =1r

dA

dr= 1− r2 (7.5.77)

and thus is positive in the interior of the projected unit disc and negativeon the exterior. Hence, if we perturb the helix so as to push part of the pro-jected curve into the interior of the disc we subtract flux since B is positivein the interior. On the other hand if we push the projected curve to theexterior of the disc we add negative flux. In either case z1 decreases, violat-ing the fixed endpoint conditions. Hence there are no allowable piecewisesmooth variations of the helix and it is indeed rigid.

7.6 Dynamic Optimal Control

7.6.1 The Falling Cat Theorem

Falling cat problem. This problem is an abstraction of the problem ofhow a falling cat should optimally (in some sense) move its body parts sothat it achieves a 180 reorientation during its fall.

7.6 Dynamic Optimal Control 215

We begin with a Riemannian manifold Q (the configuration space of theproblem) with a free and proper isometric action of a Lie group G on Q (thegroup SO(3) for the falling cat). Let A denote the mechanical connection;that is, it is the principal connection whose horizontal space is the metricorthogonal to the group orbits. The quotient space Q/G = X, the shapespace, inherits a Riemannian metric from that on Q. Given a curve c(t) inQ, we shall denote the corresponding curve in the base space X by r(t).

The optimal control problem under consideration is as follows:

Isoholonomic problem (falling cat problem). Fixing two points q1, q2 ∈Q, among all curves q(t) ∈ Q, 0 ≤ t ≤ 1 such that q(0) = q0, q(1) = q1

and q(t) ∈ horq(t) (horizontal with respect to the mechanical connection A),find the curve or curves q(t) such that the energy of the base space curve,namely,

12

∫ 1

0

‖r‖2dt,

is minimized.

Theorem 7.6.1. (Montgomery [1984, 1990, 1991a]). If q(t) is a (regular)optimal trajectory for the isoholonomic problem, then there exists a curveλ(t) ∈ g∗ such that the reduced curve r(t) in X = Q/G together with λ(t)satisfies Wong’s equations:

pα = −λaBaαβ rβ −12∂gβγ

∂rαpβpγ

λb = −λaCadbAdαrα

where gαβ is the local representation of the metric on the base space X;that is

12‖r‖2 =

12gαβ r

αrβ ,

gβγ is the inverse of the matrix gαβ, pα is defined by

pα =∂l

∂rα= gαβ r

β ,

and where we write the components of A as Abα and similarly for its cur-vature B.

Proof. As with the Heisenberg system, by general principles in the cal-culus of variations, given an optimal solution q(t), there is a Lagrangemultiplier λ(t) such that the new action function defined on the space ofcurves with fixed endpoints by

S[q( · )] =∫ 1

0

[12‖r(t)‖2 + 〈λ(t),Aq(t)〉

]dt

216 7. Optimal Control

has a critical point at this curve. Using the integrand as a Lagrangian,identifying Ω = Aq and applying the reduced Euler-Lagrange equationsfrom Lecture 1 to the reduced Lagrangian

l(r, r,Ω) =12‖r‖2 + 〈λ,Ω〉

then gives Wong’s equations by the following simple calculations:

∂l

∂rα= gαβ r

β ;∂l

∂rα=

12∂gβγ

∂rαrβ rγ ;

∂l

∂Ωa= λa.

The constraints are Ω = 0 and so the reduced Euler-Lagrange equationsbecome

d

dt

∂l

∂rα− ∂l

∂rα= −λa(Baαβ rβ)

d

dtλb = −λa(Eaαbrα) = −λaCadbAdαrα.

But

d

dt

∂l

∂rα− ∂l

∂rα= pα −

12∂gβγ∂rα

rβ rγ

= pα +12∂gκσ

∂rαgκβgσγ r

β rγ

= pα +12∂gβγ

∂rαpβpγ ,

and so we have the desired equations. ¥

Remark. There is a rich literature on Wong’s equations and it was animportant ingredient in the development of reduction theory. Some ref-erences are Sternberg [1977], Guillemin and Sternberg [1978], Weinstein[1978], Montgomery, Marsden and Ratiu [1984], Montgomery [1984], Koonand Marsden [1997] and Cendra, Holm, Marsden, and Ratiu [1998].

Nonholonomic optimal control. Using a synthesis of the techniquesused above for the Heisenberg system and the falling cat problem, Koonand Marsden [1997] generalized these problems to the nonholonomic case.In addition, these methods allow one to treat the falling cat problem evenin the case that the angular momentum is not zero.

In this process the momentum equation plays the role of the constraint.It is inserted as a first order differential constraint on the nonholonomicmomentum.

7.6 Dynamic Optimal Control 217

7.6.2 Optimal control of the rolling ball

In this section we consider optimal control of the rolling ball (see alsoChapter 4). We suppose here that J = αI3, I3 the 3 × 3 identity matrix,in which case the equations (27) become, after suitable state feedback, thefollowing control system on R4 × SO(3):

x = u1

y = u2

P = PS(−xe2 + ye1 + ce3); c = c/α . (7.6.1)

(28) is evidently controllable (it is accessible, as in the systems analyzed inBloch et.al. [1992] and the uncontrolled trajectories are periodic).

The obvious minimum energy control problem is

minu

∫ T

0

12

(u21 + u2

2)dt (7.6.2)

subject to (28).This may be viewed as the following constrained variational problem on

SO(3):Set q = P and

Dq

dt= yX1 − xX1 + cX3 ,

Xi(q) = PS(ei) (7.6.3)

Then the variational problem may be posed as:

minq

∫ T

0

⟨D2q

dt2,D2q

dt2

⟩dt

subject to (30).In contrast with the minimum energy holonomic control problems of

Section 1, this nonholonomic problem introduces constraints into the vari-ational problem. However, this now falls into the class of problems analyzedby Crouch and Leite [1991]. Defining the vector of Lagrange multipliers byΛ =

∑i

λiXi, the resulting extremals satisfy

x(4) = cy(3) + λ3y − cλ1

y(4) = −cx(3) − λ3x− cλ2

λ3 = −yx(3) + xy(3) − λ2y − λ1x

λ1 = 0λ2 = 0 . (7.6.4)

218 7. Optimal Control

This is page 219Printer: Opaque this

8Energy Based Methods for Stability

A key notion in dynamics and control is that of stability of a point orset. In this context one normally thinks either of aymptotic or nonlinearstability, the former meaning essentially that all nearby trajectories tend tothe set and the latter meaning that all trajectories starting nearby the setremain near the set. Most often one considers stability of an equilibriumpoint of a given system. Also of interest are relative equilibria – equilibriamodulo the action of a symmetry group. Example of such equilibria arethe so-called stationary motions of rigid body – rotations about one of theprincipal axes. In the theory of controlled systems, the problem is oftenthat of achieving nonlinear or asymptotic stability of an initially unstableequilibrium point or set.

8.1 The Energy Momentum Method

In this section we discuss the so called Energy Casimir method and itsapplications to the analysis of stability and stabilization problems. Sincethe application to pure stability problems is covered very well in Mars-den and Ratiu [1994] we will shall concentrate mainly on the stabilizationangle. The key idea is the use of a combination of energy and another con-served quantity, such as momentum, to provide a Lyapunov function forthe system.

220 8. Energy Based Methods for Stability

8.1.1 The Energy-Momentum Method for HolonomicSystems

As mentioned above, we use here an approach to stability which gener-alizes the energy-momentum method for Hamiltonian systems. Of coursethe energy-momentum method has a long and distinguished history goingback to Routh, Riemann, Poincare, Lyapunov, Arnold, Smale and manyothers. The main new feature provided in the more recent work of Simo,Lewis and Marsden [1991] (see Marsden [1992] for an exposition) is to ob-tain the powerful block diagonalization structure of the second variation ofthe augmented Hamiltonian as well as the normal form for the symplecticstructure. This formulation also allowed for the proof of a converse of theenergy-momentum method in the context of dissipation induced instabili-ties due to Bloch, Krishnaprasad, Marsden and Ratiu [1994, 1996].

Recall that the key idea for analyzing the stability of relative equilibria inthe holonomic setting is to use the energy plus a function of other conservedquantities such as the momentum as a Lyapunov function. In effect, oneis analyzing stability subject to the systems lying on a level surface of themomentum.

In a body frame and in the special case of Lie-Poisson systems, themomentum often can be written in terms of a Casimir, a function thatcommutes with every function under the Poisson bracket, and the methodis sometimes called the energy-Casimir method.

While the energy is conserved, it does not provide sufficient informationon stability since its second variation will be only semidefinite at a stableequilibrium in general. The algorithm for analyzing stability is thus asfollows:

1. write the equations of motion in Hamiltonian form and identify thecritical point of interest,

2. identify other conserved quantities such as momentum,

3. choose a function HC such that the energy plus the function of otherconserved quantities has a critical point at the chosen equilibriumand

4. show that HC is definite at the given equilibrium. This proves non-linear stability in the sense of Lyapunov.

Of course in special circumstances one has to interpret stability modulothe symmetry or a similar space in order to obtain stability. A good exampleis the study of two dimensional ABC flows, as in Chern and Marsden [1990].

In the nonholonomic case, while energy is conserved, momentum gener-ally is not. As indicated in the discussion of the three principle cases above,in some cases however the momentum equation is integrable, leading to in-variant surfaces which make possible an energy-momentum analysis similar

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 221

to that of the Hamiltonian case. When the momentum equation is not in-tegrable, one can get asymptotic stability in certain directions and thestability analysis is rather different from the Hamiltonian case. Nonethe-less, to show stability we will make use of the conserved energy and thedynamic momentum equation.

8.2 The Energy-Momentum Method for theStability of Nonholonomic Systems

Here we analyze the stability of relative equilibria for nonholonomic me-chanical systems with symmetry using an energy-momentum analysis fornonholonomic systems that is analogous to that for holonomic systemsgiven in Simo, Lewis, and Marsden [1991]. This section is based on thepaper Zenkov, Bloch and Marsden [1997] and follows the the spirit of thepaper by Bloch, Krishnaprasad, Marsden and Murray [1996], hereafter re-ferred to as [BKMM]. We will illustrate our energy-momentum stabilityanalysis with a low dimensional model example, and then with several me-chanical examples of interest including the falling disk, the roller racer, andthe rattleback top.

As discussed above symmetries do not always lead to conservation lawsas in the classical Noether theorem, but rather to an interesting momentumequation.

The momentum equation has the structure of a parallel transport equa-tion for the momentum corrected by additional terms. This parallel trans-port occurs in a certain vector bundle over shape space. In some instancessuch as the Routh problem of a sphere rolling inside a surface of revolution(see Zenkov [1995]) this equation is pure transport, and in fact is integrable(the curvature of the transporting connection is zero). This leads to non-explicit conservation laws.

In other important instances, the momentum equation is partially inte-grable in a sense that we shall make precise. Our goal is to make use of, asfar as possible, the energy momentum approach to stability for Hamilto-nian systems. This method goes back to fundamental work of Routh (andmany others in this era), and in more modern works, that of Arnold [1966]and Smale [1970], and Simo, Lewis and Marsden [1991] (see for example,Marsden [1992] for an exposition and additional references). Because of thenature of the momentum equation, the analysis we present is rather dif-ferent in several important respects. In particular, our energy-momentumanalysis varies according to the structure of the momentum equation and,correspondingly, we divide our analysis into several parts.

We consider a number of interesting examples including the rolling discdiscussed ealier. Other examples are as follows:

222 8. Energy Based Methods for Stability

A Mathematical Example

We now consider an instructive, but (so far as we know) nonphysical exam-ple. Unlike the rolling disk, it has asymptotically stable relative equilibria,and is a simple example that exhibits the richness of stability in nonholo-nomic systems. Our general theorems presented later are well illustratedby this example and the reader may find it helpful to return to it againlater.

Consider a Lagrangian on TR3 of the form

L(r1, r2, s, r1, r2, s) =12

(1− [a(r1)]2)(r1)2 − 2a(r1)b(r1)r1r2

+ (1− [b(r1)]2)(r2)2 + s2− V (r1), (8.2.1)

where a, b, and V are given real valued functions of a single variable. Weconsider the nonholonomic constraint

s = a(r1)r1 + b(r1)r2. (8.2.2)

Using the definitions, straightforward computations show that B12 =∂r1b = −B12. The constrained Lagrangian is Lc = 1

2

(r1)2 + (r2)2)

V (r1) and the equations of motion, namely, (d/dt)(∂rαLc) − ∂rαLc =−sBαβ rβ become

d

dt

∂Lc∂r1− ∂Lc∂r1

= −sB12r2,

d

dt

∂Lc∂r2

= sB12r1.

The Lagrangian is independent of r2 and correspondingly, we introducethe nonholonomic momentum defined by

p =∂Lc∂r2

.

We shall review the nonholonomic momentum later on in connection withgeneral symmetries, but for now just regard this as a definition. Taking intoaccount the constraint equation and the equations of motion above, we canrewrite the equations of motion in the form

r1 = − ∂V∂r1− ∂b

∂r1

(a(r1)r1 + b(r1)p

)p, (8.2.3)

p =∂b

∂r1

(a(r1)r1 + b(r1)p

)r1. (8.2.4)

Observe that the momentum equation does not, in any obvious way, implya conservation law.

A relative equilibrium is a point (r0, p0) that is an equilibrium modulothe variable r2; thus, from the equations (8.2.3) and (8.2.4), we requirer1 = 0 and

∂V

∂r1(r1

0) +∂b

∂r1b(r1

0)p20 = 0.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 223

We shall see that relative equilibria are Lyapunov stable and in additionasymptotically stable in certain directions if the following two stabilityconditions are satisfied:

(i) the energy function E = 12 (r1)2 + 1

2p2 +V , which has a critical point

at (r0, p0), has a positive definite second derivative at this point.

(ii) the derivative of E along the flow of the auxilliary system

r1 = − ∂V∂r1− ∂b

∂r1

(a(r1)r1 + b(r1)p

)p, p =

∂b

∂r1(r1)pr1

is strictly negative.

The Roller Racer

We now consider a tricycle-like mechanical system called the roller racer,or the Tennessee racer, that is capable of locomotion by oscillating thefront handlebars. This toy was studied using the methods of [BKMM] inTsakiris [1995]. The methods here may be useful for modeling and study-ing the stability of other systems, such as aircraft landing gears and trainwheels.

The roller racer is modeled as a system of two planar coupled rigid bodies(the main body and the second body) with a pair of wheels attached on eachof the bodies at their centers of mass. We assume that the mass and thelinear momentum of the second body are negligible, but that the momentof inertia about the vertical axis is not. See Figure 8.2.1.

θ

x

z

y

(x, y) φ

d1 d2

Figure 8.2.1. The geometry for the roller racer.

Let (x, y) be the location of the center of mass of the first body anddenote the angle in between the inertial reference frame and the line passingthrough the center of mass of the first body by θ, the angle between thebodies by φ, and the distances from the centers of mass to the joint by d1

224 8. Energy Based Methods for Stability

and d2. The mass of body 1 is denoted m and the inertias of the two bodiesare written as I1 and I2.

The Lagrangian and the constraints are

L =12m(x2 + y2) +

12I1θ

2 +12I2(θ + φ)2

and

x = cos θ(d1 cosφ+ d2

sinφθ +

d2

sinφφ

),

y = sin θ(d1 cosφ+ d2

sinφθ +

d2

sinφφ

).

The configuration space is SE(2) × SO(2). The Lagrangian and the con-straints are invariant under the left action of SE(2) on the first factor ofthe configuration space.

We shall see later that the roller racer has a two dimensional manifoldof equilibria and that under a suitable stability condition some of theseequilibria are stable modulo SE(2) and in addition asymptotically stablewith respect to φ.

The Rattleback

A rattleback is a convex nonsymmetric rigid body rolling without slidingon a horizontal plane. It is known for its ability to spin in one directionand to resist spinning in the opposite direction for some parameter values,and for other values, to exhibit multiple reversals. See Figure 8.2.2.

Figure 8.2.2. The rattleback.

Basic references on the rattleback are Walker [1896], Karapetyan [1980,1981], Markeev [1983, 1992], Pascal [1983, 1986], and Bondi [1986]. Weadopt the ideal model (with no energy dissipation and no sliding) of thesereferences and within that context, no approximations are made. In par-ticular, the shape need not be ellipsoidal. Walker did some initial stabilityand instability investigations by computing the spectrum while Bondi ex-tended this analysis and also used what we now recognize as the momentum

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 225

equation. (See Burdick, Goodwine and Ostrowski [1994]). Karapetyan car-ried out a stability analysis of the relative equilibria, while Markeev’s andPascal’s main contributions were to the study of spin reversals using smallparameter and averaging techniques.

Introduce the Euler angles θ, φ, ψ using the principal axis body framerelative to an inertial reference frame. These angles together with two hor-izontal coordinates x, y of the center of mass are coordinates in the config-uration space SO(3)× R2 of the rattleback.

The Lagrangian of the rattleback is computed to be

L =12[A cos2 ψ +B sin2 ψ +m(γ1 cos θ − ζ sin θ)2

]θ2

+12[(A sin2 ψ +B cos2 ψ) sin2 θ + C cos2 θ

]φ2

+12(C +mγ2

2 sin2 θ)ψ2 +

12m(x2 + y2

)+m(γ1 cos θ − ζ sin θ)γ2 sin θ θψ + (A−B) sin θ sinψ cosψ θφ

+ C cos θ φψ +mg(γ1 sin θ + ζ cos θ),

where

A,B,C = the principal moments of inertia of the body,m = the total mass of the body,

(ξ, η, ζ) = coordinates of the point of contact relative to the body frame,γ1 = ξ sinψ + η cosψ,γ2 = ξ cosψ − η sinψ.

The shape of the body is encoded by the functions ξ, η and ζ. The con-straints are

x = α1θ + α2ψ + α3φ, y = β1θ + β2ψ + β3φ,

where

α1 = −(γ1 sin θ + ζ cos θ) sinφ,α2 = γ2 cos θ sinφ+ γ1 cosφ,α3 = γ2 sinφ+ (γ1 cos θ − ζ sin θ) cosφ,

βk = −∂αk∂φ

, k = 1, 2, 3.

The Lagrangian and the constraints are SE(2)-invariant, where the ac-tion of an element (a, b, α) ∈ SE(2) is given by

(x, y, φ) 7→ (x cosα− y sinα+ a, x sinα+ y cosα+ b, φ+ α).

Corresponding to this invariance, ξ, η, and ζ are functions of the variablesθ and ψ only.

226 8. Energy Based Methods for Stability

8.2.1 The Equations of Motion of NonholonomicSystems with Symmetries

We will need the nonholonomic equations of motion and the momentumequation in the body representation:

The momentum in body representation. Let a local trivializationbe chosen on the principle bundle π : Q→ Q/G, with a local representationhaving components denoted (r, g). Let r, an element of shape space Q/Ghave coordinates denoted rα, and let g be group variables for the fiber,G. In such a representation, the action of G is the left action of G on thesecond factor.

Putl(r, r, ξ) = L(r, g, r, g).

Choose a q-dependent basis ea(q) for the Lie algebra such that the first melements span the subspace gq in the following way. First, one chooses, foreach r, such a basis at the identity element g = Id, say

e1(r), e2(r), . . . , em(r), em+1(r), . . . , ek(r).

For example, this could be a basis whose generators are orthonormal in thekinetic energy metric. Now define the body fixed basis by

ea(r, g) = Adg · ea(r),

then the first m elements will indeed span the subspace gq since the dis-tribution is invariant. Define the components of the momentum in bodyrepresentation to be

pb :=⟨∂l

∂ξ, eb(r)

⟩.

Thus, we haveJnhc(r, g, r, g) = Ad∗g−1p(r, r, ξ).

This formula explains why p and Jnhc are the body momentum and thespatial momentum respectively.

The Nonholonomic Connection. Assume that the Lagrangian has theform kinetic energy minus potential energy, and that the constraints andthe orbit directions span the entire tangent space to the configuration space([BKMM] call this the “dimension assumption”):

Dq + Tq(Orb(q)) = TqQ. (8.2.5)

In this case, the momentum equation can be used to augment the con-straints and provide a connection on Q→ Q/G.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 227

Definition 8.2.1. Under the dimension assumption in equation (8.2.5),and the assumption that the Lagrangian is of the form kinetic minus po-tential energies, the nonholonomic connection A is the connection onthe principal bundle Q → Q/G whose horizontal space at the point q ∈ Qis given by the orthogonal complement to the space Sq within the space Dq.

Let I(q) : gD →(gD)∗ be the locked inertia tensor relative to gD, defined

by

〈I(q)ξ, η〉 = 〈〈ξQ, ηQ〉〉, ξ, η ∈ gq,

where 〈〈· , ·〉〉 is the kinetic energy metric. Define a map Asymq : TqQ→ Sq =

Dq ∩ Tq(Orb(q)) given by

Asymq (vq) = (I−1Jnhc(vq))Q.

This map is equivariant and is a projection onto Sq. Choose Uq ⊂ Tq(Orb(q))such that Tq(Orb(q)) = Sq ⊕Uq. Let Akin

q : TqQ→ Uq be a Uq valued formthat projects Uq onto itself and maps Dq to zero; for example, it can begiven by orthogonal projection relative to the kinetic energy metric (thiswill be our default choice).

Proposition 8.2.2. The nonholonomic connection regarded as an Ehres-mann connection is given by

A = Akin +Asym. (8.2.6)

When the connection is regarded as a principal connection (i.e., takes valuesin the Lie algebra rather than the vertical space) we will use the symbol A.

Given a velocity vector q that satisfies the constraints, we orthogonallydecompose it into a piece in Sq and an orthogonal piece denoted rh. Weregard rh as the horizontal lift of a velocity vector r on the shape space;recall that in a local trivialization, the horizontal lift to the point (r, g) isgiven by

rh = (r,−Alocr) = (rα,−Aaαrα),

where Aaα are the components of the nonholonomic connection which is aprincipal connection in a local trivialization.

We will denote the decomposition of q by

q = ΩQ(q) + rh,

so that for each point q, Ω is an element of the Lie algebra and representsthe spatial angular velocity of the locked system. In a local trivialization,we can write, at a point (r, g),

Ω = Adg(Ωloc),

228 8. Energy Based Methods for Stability

so that Ωloc represents the body angular velocity. Thus,

Ωloc = Alocr + ξ

and, at each point q, the constraints are that Ω belongs to gq, i.e.,

Ωloc ∈ spane1(r), e2(r), . . . , em(r).

The vector rh need not be orthogonal to the whole orbit, just to thepiece Sq. Even if q does not satisfy the constraints, we can decompose itinto three parts and write

q = ΩQ(q) + rh = ΩnhQ (q) + Ω⊥Q(q) + rh,

where ΩnhQ and Ω⊥Q are orthogonal and where Ωnh

Q (q) ∈ Sq. The relationΩloc = Alocr+ ξ is valid even if the constraints do not hold; also note thatthis decomposition of Ω corresponds to the decomposition of the nonholo-nomic connection given by A = Akin + Asym that was given in equation(8.2.6).

To avoid confusion, we will make the following index and summationconventions

1. The first batch of indices range from 1 to m corresponding to thesymmetry directions along constraint space. These indices will be de-noted a, b, c, d, . . . and a summation from 1 to m will be understood.

2. The second batch of indices range from m+ 1 to k corresponding tothe symmetry directions not aligned with the constraints. Indices forthis range or for the whole range 1 to k will be denoted by a′, b′, c′, . . .and the summations will be given explicitly.

3. The indices α, β, . . . on the shape variables r range from 1 to σ. Thus,σ is the dimension of the shape space Q/G and so σ = n − k. Thesummation convention for these indices will be understood.

According to [BKMM], the equations of motion are given by the nexttheorem

Theorem 8.2.3. The following reduced nonholonomic Lagrange-d’Al-embert equations hold for each 1 ≤ α ≤ σ and 1 ≤ b ≤ m:

d

dt

∂lc∂rα− ∂lc∂rα

= −∂Icd

∂rαpcpd −DcbαIbdpcpd − Bcαβpcrβ

−DβαbIbcpcrβ −Kαβγ rβ rγ ,d

dtpb = CcabI

adpcpd +Dcbαpcrα +Dαβbrαrβ .

Here lc(rα, rα, pa) is the constrained Lagrangian; rα, 1 ≤ α ≤ σ, are coordi-nates in the shape space; pa, 1 ≤ a ≤ m, are components of the momentum

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 229

map in the body representation, pa = 〈∂lc/∂Ωloc, ea(r)〉; Iad are the com-ponents of the inverse locked inertia tensor; Baαβ are the local coordinatesof the curvature B of the nonholonomic connection A; and the coefficientsDcbα, Dαβb, Kαβγ are given by the formulae

Dcbα =k∑

a′=1

−Cca′bAa′

α + γcbα +k∑

a′=m+1

λa′αCa′

abIac,

Dαβb =k∑

a′=m+1

λa′α

(γa′

bβ −k∑

b′=1

Ca′

b′bAb′

β

),

Kαβγ =k∑

a′=1

λa′γBa′

αβ ,

where

λa′α = la′α −k∑

b′=1

la′b′Ab′

α :=∂l

∂ξa′∂rα−

k∑b′=1

∂l

∂ξa′∂ξb′Ab′α

for a′ = m+ 1, . . . , k. Here Cb′

a′c′ are the structure constants of the Lie al-gebra defined by [ea′ , ec′ ] = Cb

a′c′eb′ , a′, b′, c′ = 1, . . . , k; and the coefficients

γc′

bα are defined by

∂eb∂rα

=k∑

c′=1

γc′

bαec′ .

A relative equilibrium is an equilibrium of the reduced equations; thatis, it is a solution that is given by a one parameter group orbit, just as inthe holonomic case (see, e.g., Marsden [1992] for a discussion).

The Constrained Routhian. This function is defined by analogy withthe usual Routhian by

R(rα, rα, pa) = lc(rα, rα, Iabpb)− Iabpapb,

and in terms of it, the reduced equations of motion become

d

dt

∂R

∂rα− ∂R

∂rα= −DcbαIbdpcpd − Bcαβpcrβ

−DβαbIbcpcrβ −Kαβγ rβ rγ , (8.2.7)d

dtpb = CcabI

adpcpd +Dcbαpcrα +Dαβbrαrβ . (8.2.8)

230 8. Energy Based Methods for Stability

The Reduced Constrained Energy. As in [BKMM], the kinetic en-ergy in the variables rα, rα, Ωa, Ωa

′equals

12gαβ r

αrβ +12IacΩaΩc

+k∑

a′=m+1

(la′α − la′c′Ac

α

)Ωa′rα +

12

k∑a′,c′=m+1

la′c′Ωa′Ωc′,

(8.2.9)

where gαβ are coefficients of the kinetic energy metric induced on the man-ifold Q/G. Substituting the relations Ωa = Iabpb and the constraint equa-tions Ωa

′= 0 in (8.2.9) and adding the potential energy, we define the

function E by

E =12gαβ r

αrβ + U(rα, pa), (8.2.10)

which represents the reduced constrained energy in the coordinates rα, rα,pa, where U(rα, pa) is the amended potential defined by

U(rα, pa) =12Iabpapb + V (rα), (8.2.11)

and V (rα) is the potential energy of the system.Now, we show that the reduced constrained energy is conserved along

the solutions of (8.2.7), (8.2.8).

Theorem 8.2.4. The reduced constrained energy is a constant of motion.

Proof. One way to prove this is to note that the reduced energy is aconstant of motion, because it equals the energy represented in coordinatesr, r, g, ξ and because the energy is conserved, since the Lagrangian andthe constraints are time-invariant. Along the trajectories, the constrainedenergy and the energy are the same. Therefore, the reduced constrainedenergy is a constant of motion.

One may also prove this fact by a direct computation of the time deriva-tive of the reduced constrained energy (8.2.10) along the vector field definedby the equations of motion. ¥

Skew Symmetry Assumption. We assume that the tensor CcabIad is

skew-symmetric in c, d.This holds for most physical examples and certainly the systems dis-

cussed in this paper. (Exceptions include systems with no shape space suchas the homogeneous sphere on the plane and certain cases ‘of the ‘Suslov”problem of a nonhomogeneous rigid body subject to a linear constraint inthe angular velocities.) It is an intrinsic (coordinate independent) condi-tion, since CcabI

ad represents an intrinsic bilinear map of(gD)∗ × (gD)∗ to(

gD)∗.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 231

Under this assumption, the terms quadratic in p in the momentum equa-tion vanish, and the equations of motion become

d

dt

∂R

∂rα− ∂R

∂rα= −DcbαIbdpcpd − Bcαβpcrβ

−DβαbIbcpcrβ −Kαβγ rβ rγ , (8.2.12)d

dtpb = Dcbαpcrα +Dαβbrαrβ . (8.2.13)

As a result, the dimension of the family of the relative equilibria equals thenumber of components of the (nonholonomic) momentum map.

In the case when Ccab = 0 (cyclic variables, or internal abelian symme-tries) the matrix CcabI

ad vanishes, and the preceding equations of motionare the same as those obtained by Karapetyan [1983].

Below, three principal cases will be considered:

1. Pure Transport Case In this case, terms quadratic in r are notpresent in the momentum equation, so it is in the form of a trans-port equation—i.e. the momentum equation is an equation of paralleltransport and the equation itself defines the relevant connection.

Under certain integrability conditions (see below) the transport equa-tion defines invariant surfaces, which allow us to use a type of energy-momentum method for stability analysis in a similar fashion to themanner in which the holonomic case uses the level surfaces definedby the momentum map. The key difference is that in our case, theadditional invariant surfaces do not arise from conservation of mo-mentum. In this case, one gets stable, but not asymptotically stable,relative equilibria. Examples include the rolling disk, a body of rev-olution rolling on a horizontal plane, and the Routh problem.

2. Integrable Transport Case In this case, terms quadratic in r arepresent in the momentum equation and thus it is not a pure trans-port equation. However, in this case, we assume that the transportpart is integrable. As we shall also see, in this case relative equilibriamay be asymptotically stable. We are able to find a generalization ofthe energy-momentum method which gives conditions for asymptoticstability. An example is the roller racer.

3. Nonintegrable Transport Case Again, the terms quadratic in rare present in the momentum equation and thus it is not a pure trans-port equation. However, the transport part is not integrable. Again,we are able to demonstrate asymptotic stability using Lyapunov-Malkin Theorem and to relate it to an energy-momentum type anal-ysis under certain eigenvalue hypotheses, as we will see in 8.2.3. Anexample is the rattleback top. Another example is a nonhomogeneous

232 8. Energy Based Methods for Stability

sphere with a center of mass lying off the planes spanned by the prin-cipal axis body frame. See Markeev [1992].In some examples, such as the nonhomogeneous (unbalanced) Ko-valevskaya sphere rolling on the plane, these eigenvalue hypothesesdo not hold. We intend to investigate this case in a future publication.

In the sections below where these different cases are discussed we willmake clear at the beginning of each section what the underlying hypotheseson the systems are by listing the key hypotheses and labeling them by H1,H2, and H3.

8.2.2 The Pure Transport Case

In this section we assume that

H1 Dαβb are skew-symmetric in α, β. Under this assumption, the mo-mentum equation can be written as the vanishing of the connectionone form defined by dpb −Dcbαpcdrα.

H2 The curvature of the preceding connection form is zero.

A nontrivial example of this case is that of Routh’s problem of a sphererolling in a surface of revolution. See Zenkov [1995].

Under the above two assumptions, the distribution defined by the mo-mentum equation is integrable, and so we get invariant surfaces, whichmakes further reduction possible. This enables us to use the energy-momentummethod in a way that is similar to the holonomic case, as we explainedabove.

Note that if the number of shape variables is one, the above connectionis integrable, because it may be treated as a system of linear ordinarydifferential equations with coefficients depending on the shape variable r:

dpbdr

= Dcbpc.

As a result, we obtain an integrable nonholonomic system, because aftersolving the momentum equation for pb and substituting the result in theequation for the shape variable, the latter equation may be viewed as aLagrangian system with one degree of freedom, which is integrable.

The Nonholonomic Energy-Momentum Method

We now develop the energy-momentum method for the case in which themomentum equation is pure transport. Under the assumptions H1 and H2made so far, the equations of motion become

d

dt

∂R

∂rα− ∂R

∂rα= −DcbαIbdpcpd − Bcαβpcrβ −Kαβγ rβ rγ , (8.2.14)

d

dtpb = Dcbαpcrα. (8.2.15)

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 233

A relative equilibrium is a point (r, r, p) = (r0, 0, p0) which is a fixedpoint for the dynamics determined by equations (8.2.14) and (8.2.15). Un-der assumption H1 the point (r0, p0) is seen to be a critical point of theamended potential.

Because of our zero curvature assumption H2, the solutions of the mo-mentum equation lie on surfaces of the form pa = Pa(rα, kb), a, b = 1, . . . ,m,where kb are constants labeling these surfaces.

Using the functions pa = Pa(rα, kb) we introduce the reduced amendedpotential Uk(rα) = U(rα, Pa(rα, kb)). We think of the function Uk(rα) asbeing the restriction of the function U to the invariant manifold

Qk = (rα, pa) | pa = Pa(rα, kb).

Theorem 8.2.5. Let assumptions H1 and H2 hold and let (r0, p0), wherep0 = P (r0, k

0), be a relative equilibrium. If the reduced amended potentialUk0(r) has a nondegenerate minimum at r0, then this equilibrium is Lya-punov stable.

Proof. First, we show that the relative equilibrium

rα = rα0 , p0a = Pa(rα0 , k

0b ) (8.2.16)

of the system (8.2.14), (8.2.15) is stable modulo perturbations consistentwith Qk0 . Consider the phase flow restricted to the invariant manifoldQk0 , where k0 corresponds to the relative equilibrium. Since Uk0(rα) hasa nondegenerate minimum at rα0 , the function E|Qk0 is positive definite.By Theorem 8.2.4 its derivative along the flow vanishes. Using E|k0 as aLyapunov function, we conclude that equations (8.2.14), (8.2.15), restrictedto the manifold Qk0 , have a stable equilibrium point rα0 on Qk0 .

To finish the proof, we need to show that equations (8.2.14), (8.2.15),restricted to nearby invariant manifolds Qk, have stable equilibria on thesemanifolds.

If k is sufficiently close to k0, then by the properties of families of Morsefunctions (see Milnor [1963]), the function Uk : Qk → R has a nondegener-ate minimum at the point rα which is close to rα0 . This means that for allk sufficiently close to k0 the system (8.2.14), (8.2.15) restricted to Qk hasa stable equilibrium rα. Therefore, the equilibrium (8.2.16) of equations(8.2.14), (8.2.15) is stable.

The stability here cannot be asymptotic, since the dynamical systemson Qk have a positive definite conserved quantity—the reduced energyfunction. ¥

Remark. Even though in general Pa(rα, kb) can not be found explicitly,the types of critical points of Uk may be explicitly determined as follows.First of all, note that

∂pa∂rα

= Dcbαpc

234 8. Energy Based Methods for Stability

as long as (rα, pa) ∈ Qk. Therefore

∂Uk∂rα

= ∇αU,

where

∇α =∂

∂rα+Dcbαpc

∂pb. (8.2.17)

Then the relative equilibria satisfy the condition

∇αU = 0,

while the condition for stability

∂2Uk∂r2

À 0

(i.e., is positive definite) becomes the condition

∇α∇βU À 0.

In the commutative case this was shown by Karapetyan [1983].Now we give the stability condition in a form similar to that of energy-

momentum method for holonomic systems given in Simo, Lewis, and Mars-den [1991].

Theorem 8.2.6 (The nonholonomic energy-momentum method).Under assumptions H1 and H2, the point qe = (rα0 , p

0a) is a relative equi-

librium if and only if there is a ξ ∈ gqe such that qe is a critical point ofthe augmented energy Eξ : D/G→ R (i.e., Eξ is a function of (r, r, p)),defined by

Eξ = E − 〈p− P (r, k), ξ〉.

This equilibrium is stable if δ2Eξ restricted to TqeQk is positive definite(here δ denotes differentiation with respect to all variables except ξ).

Proof. A point qe ∈ Qk is a relative equilibrium if ∂rαUk = 0. This con-dition is equivalent to d (E|Qk) = 0. The last equation may be representedas d (E − 〈p− P (r, k), ξ〉) = 0 for some ξ ∈ gqe . Similarly, the condition forstability d2Uk À 0 is equivalent to d2 (E|Qk) À 0, which may be repre-sented as

(δ2Eξ

)|TqeQk À 0. ¥

Note that if the momentum map is preserved, then the formula for Eξbecomes

Eξ = E − 〈p− k, ξ〉,which is the same as the formula for the augmented energy Eξ for holonomicsystems.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 235

Examples

There are several examples which illustrate the ideas above. For instancethe falling disk, Routh’s problem, and a body of revolution rolling on ahorizontal plane are systems where the momentum equation defines anintegrable distribution and we are left with only one shape variable. Sincethe stability properties of all these systems are similar, we consider hereonly the rolling disk. For the body of revolution on the plane see Chaplygin[1897a] and Karapetyan [1983]. For the Routh problem see Zenkov [1995].

The Rolling Disk. Consider again the disk rolling without sliding onthe xy-plane. Recall that we have the following: Denote the coordinates ofcontact of the disk in the xy-plane by x, y. Let θ, φ, and ψ denote theangle between the plane of the disk and the vertical axis, the “headingangle” of the disk, and “self-rotation” angle of the disk respectively, as wasintroduced earlier.

The Lagrangian and the constraints in these coordinates are given by

L =m

2

[(ξ −R(φ sin θ + ψ))2 + η2 sin2 θ + (η cos θ +Rθ)2

]+

12

[A(θ2 + φ2 cos2 θ) +B(φ sin θ + ψ)2

]−mgR cos θ,

x = −ψR cosφ,

y = −ψR sinφ,

where ξ = x cosφ + y sinφ + Rψ, η = −x sinφ + y cosφ. Note that theconstraints may be written as ξ = 0, η = 0.

This system is invariant under the action of the group G = SE(2) ×SO(2); the action by the group element (a, b, α, β) is given by

(θ, φ, ψ, x, y) 7→ (θ, φ+ α, ψ + β, x cosα− y sinα+ a, x sinα+ y cosα+ b).

Obviously,

TqOrb(q) = span(∂

∂φ,∂

∂ψ,∂

∂x,∂

∂y

),

and

Dq = span(∂

∂θ,∂

∂φ,R cosφ

∂x+R sinφ

∂y− ∂

∂ψ

),

which imply

Sq = Dq ∩ TqOrb(q) = span(∂

∂φ,−R cosφ

∂x−R sinφ

∂y+

∂ψ

).

Choose vectors (1, 0, 0, 0), (0, 1, 0, 0), (0, 0, 1, 0), (0, 0, 0, 1) as a basis ofthe Lie algebra g of the group G. The corresponding generators are ∂x,∂y, −y∂x + x∂y + ∂φ, ∂ψ. Taking into account that the generators ∂φ,

236 8. Energy Based Methods for Stability

−R cosφ∂x−R sinφ∂y+∂ψ correspond to the elements (y,−x, 1, 0), (−R cosφ,−R sinφ, 0, 1)of the Lie algebra g, we obtain the following momentum equations:

p1 = mR2 cos θ θψ,

p2 = −mR2 cos θ θφ,(8.2.18)

where

p1 = Aφ cos2 θ + (mR2 +B)(φ sin θ + ψ) sin θ,

p2 = (mR2 +B)(φ sin θ + ψ),(8.2.19)

into which the constraints have been substituted. One may notice that

p1 =∂lc

∂φ, p2 =

∂lc

∂ψ.

Solving (8.2.19) for φ, ψ and substituting the solutions back in the equa-tions (8.2.18) we obtain another representation of the momentum equa-tions:

dp1

dt= mR2 cos θ

(− sin θA cos2 θ

p1 +(

1mR2 +B

+sin2 θ

A cos2 θ

)p2

)θ,

dp2

dt= mR2 cos θ

(− 1A cos2 θ

p1 +sin θ

A cos2 θp2

)θ.

(8.2.20)

The right hand sides of (8.2.20) do not have terms quadratic in the shapevariable θ. The distribution, defined by (8.2.20), is integrable and definestwo integrals of the form p1 = P1(θ, k1, k2), p2 = P2(θ, k1, k2). It is knownthat these integrals may be written down explicitly in terms of the hy-pergeometric function. See Appel [1900], Chaplygin [1897a], and Korteweg[1899] for details.

To carry out stability analysis, we use the remark following Theorem8.2.5. Using formulae (8.2.19), we obtain the amended potential

U(θ, p) =12

[(p1 − p2 sin θ)2

A cos2 θ+

p22

B +mR2

]+mgR cos θ.

Straightforward computation shows that the condition for stability ∇2U À0 of a relative equilibrium θ = θ0, p1 = p0

1, p2 = p02 becomes

B

A(mR2 +B)(p0

2)2 +mR2 cos2 θ0 + 2A sin2 θ0 +A

A2(p0

1 − p02 sin θ0)2

− (mR2 + 3B) sin θ0

A(mR2 +B) cos2 θ0(p0

1 − p02 sin θ0)p0

2 −mgR cos θ0 > 0.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 237

Note that this condition guarantees stability here relative to θ, θ, p1, p2;in other words we have stability modulo the action of SE(2)× SO(2).

The falling disk may be considered as a limiting case of the body of rev-olution which also has an integrable pure transport momentum equation(this example is treated in Chaplygin [1897a] and Karapetyan [1983]). Therolling disc has also been analyzed recently by O’Reilly [1996] and Cush-man, Hermans and Kemppainen [1996]. O’Reilly considered bifurcation ofrelative equilibria, the stability of vertical stationary motions, as well asthe possibility of sliding.

8.2.3 The Non-Pure Transport Case

In this section we consider the case in which the coefficients Dαβb are notskew symmetric in α, β and the two subcases where the transport part ofthe momentum equation is integrable or is not integrable, respectively. Ineither case one may obtain asymptotic stability.

We begin by assembling some preliminary results on center manifoldtheory and show how they relate to the Lyapunov-Malkin Theorem. Thecenter manifold theorem provides new and useful insight into the existenceof integral manifolds. These integral manifolds play a crucial role in ouranalysis. Lyapunov’s original proof of the Lyapunov-Malkin Theorem useda different approach to proving the existence of local integrals, as we shalldiscuss below. Malkin extended the result to the nonautonomous case.

Center Manifold Theory in Stability Analysis

Here we discuss the center manifold theory and its applications to thestability analysis of nonhyperbolic equilibria.

Consider a system of differential equations

x = Ax+X(x, y), (8.2.21)y = By + Y (x, y), (8.2.22)

where x ∈ Rm, y ∈ Rn, and A and B are constant matrices. It is supposedthat all eigenvalues of A have nonzero real parts, and all eigenvalues ofB have zero real parts. The functions X, Y are smooth, and satisfy theconditions X(0, 0) = 0, dX(0, 0) = 0, Y (0, 0) = 0, dY (0, 0) = 0. We nowrecall the following definition:

Definition 8.2.7. A smooth invariant manifold of the form x = h(y)where h satisfies h(0) = 0 and dh(0) = 0 is called a center manifold.

We are going to use the following version of the center manifold theoremfollowing the exposition of Carr [1981] (see also Chow and Hale [1982]).

Theorem 8.2.8 (The center manifold theorem). If the functions X(x, y),Y (x, y) are Ck, k ≥ 2, then there exist a (local) center manifold for (8.2.21),

238 8. Energy Based Methods for Stability

(8.2.22), x = h(y), |y| < δ, where h is Ck. The flow on the center manifoldis governed by the system

y = By + Y (h(y), y). (8.2.23)

The next theorem explains that the reduced equation (8.2.23) containsinformation about stability of the zero solution of (8.2.21), (8.2.22).

Theorem 8.2.9. Suppose that the zero solution of (8.2.23) is stable (asymp-totically stable) (unstable) and that the eigenvalues of A are in the left halfplane. Then the zero solution of (8.2.21), (8.2.22) is stable (asymptoticallystable) (unstable).

Let us now look at the special case of (8.2.22) in which the matrix Bvanishes. Equations (8.2.21), (8.2.22) become

x = Ax+X(x, y), (8.2.24)y = Y (x, y). (8.2.25)

Theorem 8.2.10. Consider the system of equations (8.2.24), (8.2.25). IfX(0, y) = 0, Y (0, y) = 0, and the matrix A does not have eigenvalues withzero real parts, then the system (8.2.24), (8.2.25) has n local integrals inthe neighborhood of x = 0, y = 0.

Proof. The center manifold in this case is given by x = 0. Each point ofthe center manifold is an equilibrium of the system (8.2.24), (8.2.25). Foreach equilibrium point (0, y0) of our system, consider an m-dimensionalsmooth invariant manifold

S(y0) = Ss(y0)× Su(y0),

where Ss(y0) and Su(y0) are stable and unstable manifolds at the equilib-rium (0, y0). The center manifold and these manifolds S(y0) can be usedfor a (local) substitution (x, y) → (x, y) such that in the new coordinatesthe system of differential equations become

˙x = Ax+ X(x, y), ˙y = 0.

The last system of equations has n integrals y = const, so that the originalequation has n smooth local integrals. Observe that the tangent spaces tothe common level sets of these integrals at the equilibria are the planes y =y0. Therefore, the integrals are of the form y = f(x, k), where ∂xf(0, k) =0. ¥

The following theorem gives stability conditions for equilibria of the sys-tem (8.2.24), (8.2.25).

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 239

Theorem 8.2.11 (Lyapunov-Malkin). Consider the system of differ-entialequations (8.2.24), (8.2.25), where x ∈ Rm, y ∈ Rn, A is an m × m-matrix, and X(x, y), Y (x, y) represent nonlinear terms. If all eigenvaluesof the matrix A have negative real parts, and X(x, y), Y (x, y) vanish whenx = 0, then the solution x = 0, y = c of the system (8.2.24), (8.2.25) isstable with respect to x, y, and asymptotically stable with respect to x. Ifa solution x(t), y(t) of (8.2.24), (8.2.25) is close enough to the solutionx = 0, y = 0, then

limt→∞

x(t) = 0, limt→∞

y(t) = c.

The proof of this theorem consists of two steps. The first step is a reduc-tion of the system to the common level set of integrals described in Theorem8.2.10. The second step is the construction of a Lyapunov function for thereduced system. The details of the proof may be found in Lyapunov [1992]and Malkin [1938].

Historical Note. The proof of the Lyapunov-Malkin Theorem uses thefact that the system of differential equations has local integrals, as discussedin Theorem 4.4. To prove existence of these integrals, Lyapunov uses atheorem of his own about the existence of solutions of PDE’s. He doesthis assuming that the nonlinear terms on the right hand sides are seriesin x and y with time-dependent bounded coefficients. Malkin generalizesLyapunov’s result for systems for which the matrix A is time-dependent.We consider the nonanalytic case, and to prove existence of these localintegrals, we use center manifold theory. This simplifies the arguments tosome extent as well as showing how the results are related.

The following Lemma specifies a class of systems of differential equations,that satisfies the conditions of the Lyapunov-Malkin Theorem.

Lemma 8.2.12. Consider a system of differential equations of the form

u = Au+By + U(u, y), y = Y(u, y), (8.2.26)

where u ∈ Rn, y ∈ Rm, detA 6= 0, and where U and Y represent higherorder nonlinear terms. There is a change of variables of the form u =x+ φ(y) such that

(i) in the new variables x, y system (8.2.26) becomes

x = Ax+X(u, y), y = Y (x, y),

(ii) if Y (0, y) = 0, then X(0, y) = 0 as well.

240 8. Energy Based Methods for Stability

Proof. Put u = x+ φ(y), where φ(y) is defined by

Aφ(y) +By + U(φ(y), y) = 0.

System (8.2.26) in the variables x, y becomes

x = Ax+X(x, y), y = Y (x, y),

where

X(x, y) = Aφ(y) +By + U(x+ φ(y), y)− ∂φ

∂yY (x, y),

Y (x, y) = Y(x+ φ(y), y).

Note that Y (0, y) = 0 implies X(0, y) = 0. ¥

The Mathematical Example

The Lyapunov-Malkin conditions. Recall from §8.2 that the equa-tions of motion are

r = −∂V∂r− ∂b

∂r(a(r)r + b(r)p) p,

p =∂b

∂r(a(r)r + b(r)p) r,

(8.2.27)

here and below we write r instead of r1.Recall also that a point r = r0, p = p0 is a relative equilibrium if r0 and

p0 satisfy the condition

∂V

∂r(r0) +

∂b

∂rb(r0) p2

0 = 0.

Introduce coordinates u1, u2, v in the neighborhood of this equilibrium by

r = r0 + u1, r = u2, p = p0 + v.

The linearized equations of motion are

u1 = u2,

u2 = Au2 + Bu1 + Cv,

v = Du2,

where

A = −∂b∂rap0,

B = −∂2V

∂r2−[∂2b

∂r2b+

(∂b

∂r

)2]p2

0,

C = −2∂b

∂rbp0,

D =∂b

∂rbp0,

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 241

and where V , a, b, and their derivatives are evaluated at r0. The charac-teristic polynomial of these linearized equations is calculated to be

λ[λ2 −Aλ− (B + CD)].

It obviously has one zero root. The two others have negative real parts if

B + CD < 0, A < 0. (8.2.28)

These conditions imply linear stability. We discuss the meaning of theseconditions later.

Next, we make the substitution v = y + Du1, which defines the newvariable y. The (nonlinear) equations of motion become

u1 = u2,

u2 = Au2 + (B + CD)u1 + Cy + U(u, y),y = Y(u, y),

where U(u, y), Y(u, y) stand for nonlinear terms, and Y(u, y) vanishes whenu = 0. By Lemma 8.2.12 there exists a further substitution u = x + φ(y)such that the equations of motion in coordinates x, y become

x = Px+X(x, y),y = Y (x, y),

where X(x, y) and Y (0, y) satisfy the conditions X(0, y) = 0, Y (0, y) = 0.Here,

P =(

0 1B + CD A

).

This form enables us to use the Lyapunov-Malkin Theorem and concludethat the linear stability implies nonlinear stability and in addition that wehave asymptotic stability with respect to variables x1, x2.

The Energy-Momentum Method. To find a Lyapunov function basedapproach for analyzing the stability of the mathematical example, we in-troduce a modified dynamical system and use its energy function and mo-mentum to construct a Lyapunov function for the original system. Thismodified system is introduced for the purpose of finding the Lyapunovfunction and is not used in the stability proof. We will generalize this ap-proach below and this example may be viewed as motivation for the generalapproach.

Consider then the new system obtained from the Lagrangian (8.2.1) andthe constraint (8.2.2) by setting a(r) = 0. Notice that Lc stays the sameand therefore, the equation of motion may be obtained from (8.2.27):

r = −∂V∂r− ∂b

∂rb(r)p2, p =

∂b

∂rb(r)pr.

242 8. Energy Based Methods for Stability

The condition for existence of the relative equilibria also stays the same.However, a crucial observation is that for the new system, the momentumequation is now integrable, in fact explicitly, so that in this example:

p = k exp(b2(r)/2).

Thus, we may proceed and use this invariant surface to perform reduction.The amended potential, defined by U(r, p) = V (r) + 1

2p2, becomes

Uk(r) = V (r) +12(k exp(b2(r)/2)

)2.

Consider the function

Wk =12

(r)2 + Uk(r) + ε(r − r0)r.

If ε is small enough and Uk has a nondegenerate minimum, then so does Wk.Suppose that the matrix P has no eigenvalues with zero real parts. Thenby Theorem 8.2.10 equations (8.2.27) have a local integral p = P(r, r, c).

Differentiate Wk along the vector field determined by (8.2.27). We obtain

Wk = −ε(∂2V

∂r2(r0) +

(∂2b

∂r2b(r0) + 2

(∂b

∂rb(r0)

)2

+(∂b

∂r(r0)

)2)p2

0

)

− ∂b

∂ra(r0)p0r

2 + εr2 + higher order terms.

Therefore, Wk is a Lyapunov function for the flow restricted to the localinvariant manifold p = P(r, r, c) if

∂2V

∂r2(r0) +

(∂2b

∂r2b(r0) + 2

(∂b

∂rb(r0)

)2

+(∂b

∂r(r0)

)2)p2

0 > 0 (8.2.29)

and

∂b

∂ra(r0)p0r

2 > 0. (8.2.30)

Notice that the Lyapunov conditions (8.2.29) and (8.2.30) are the same asconditions (8.2.28).

Introduce the operator

∇ =∂

∂r+∂b

∂rb(r)p

∂p

(cf. Karapetyan [1983]). Then condition (8.2.29) may be represented as

∇2U > 0,

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 243

which is the same as the condition for stability of stationary motions of anonholonomic system with an integrable momentum equation (recall thatthis means that there are no terms quadratic in r, only transport termsdefining an integrable distribution). The left hand side of formula (8.2.30)may be viewed as a derivative of the energy function

E =12r2 +

12p2 + V

along the flow

r = −∂V∂r− ∂b

∂r(a(r)r + b(r)p) p, p =

∂b

∂rb(r)pr,

or as a derivative of the amended potential U along the vector field definedby the nontransport terms of the momentum equations

p =∂b

∂ra(r)r2.

The Nonholonomic Energy-Momentum Method

We now generalize the energy-momentum method discussed above for themathematical example to the general case in which the transport part ofthe momentum equation is integrable.

Here we assume hypothesis H2 in the present context, namely:

H2 The curvature of the connection form associated with the transportpart of the momentum equation, namely dpb −Dcbαpcdrα, is zero.

The momentum equation in this situation is

d

dtpb = Dcbαpcrα +Dαβbrαrβ .

Hypothesis H2 implies that the form due to the transport part of the mo-mentum equation defines an integrable distribution. Associated to this dis-tribution, there is a family of integral manifolds

pa = Pa(rα, kb)

with Pa satisfying the equation dPb = DcbαPcdrα. Note that these man-ifolds are not invariant manifolds of the full system under considerationbecause the momentum equation has non-transport terms. Substitutingthe functions Pa(rα, kb), kb = const, into E(r, r, p), we obtain a function

Vk(rα, rα) = E(rα, rα, Pa(rα, kb)),

that depends only on rα, rα and parametrically on k. This function willnot be our final Lyapunov function but will be used to construct one in theproof to follow.

Pick a relative equilibrium rα = rα0 , pa = p0a. In this context we introduce

the following definiteness assumptions:

244 8. Energy Based Methods for Stability

H3 At the equilibrium rα = rα0 , pa = p0a the two symmetric matrices

∇α∇βU and (Dαβb +Dβαb)Ibcpc are positive definite.

Theorem 8.2.13. Under assumptions H2 and H3, the equilibrium rα =rα0 , pa = p0

a is Lyapunov stable. Moreover, the system has local invariantmanifolds that are tangent to the family of manifolds defined by the inte-grable transport part of the momentum equation at the relative equilibria.The relative equilibria, that are close enough to r0, p0, are asymptoticallystable in the directions defined by these invariant manifolds. In addition,for initial conditions close enough to the equilibrium rα = rα0 , pa = p0

a, theperturbed solution approaches a nearby equilibrium.

Proof. The substitution pa = p0a+ya+Dbaα(r0)p0

buα, where uα = rα−rα0 ,

eliminates the linear terms in the momentum equation. In fact, with thissubstitution, the equations of motion (8.2.12), (8.2.13) become

d

dt

∂R

∂rα− ∂R

∂rα= −DcbαIbdpcpd − Bcαβpcrβ

−DβαbIbcpcrβ −Kαβγ rβ rγ ,d

dtyb = Dcbαycrα + (Dcbα −Dcbα(r0))p0

c rα +Dαβbrαrβ .

We will show in §8.2.3 that H3 implies the hypotheses of Theorem 8.2.10.Thus, the above equations have local integrals ya = fa(r, r, c), where thefunctions fa are such that ∂rfa = ∂rfa = 0 at the equilibria. Therefore,the original equations (8.2.12), (8.2.13) have n local integrals

pa = Pa(rα, rα, cb), cb = const, (8.2.31)

where Pa are such that

∂P∂rα

=∂P

∂rα,

∂P∂rα

= 0

at the relative equilibria.We now use the Vk(rα, rα) to construct a Lyapunov function to determine

the conditions for asymptotic stability of the relative equilibrium rα = rα0 ,pa = p0

a. We will do this in a fashion similar to that used by Chetaev[1959] and Bloch, Krishnaprasad, Marsden, and Ratiu [1994]. Without lossof generality, suppose that gαβ(r0) = δαβ . Introduce the function

Wk = Vk + εσ∑α=1

uαrα.

Consider the following two manifolds at the equilibrium (rα0 , p0a): the

integral manifold of the transport equation

Qk0 =pa = Pa(rα, k0)

,

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 245

and the local invariant manifold

Qc0 =pa = Pa(rα, rα, c0)

.

Restrict the flow to the manifold Qc0 . Choose (rα, rα) as local coordinateson Qc0 , then Vk0 and Wk0 are functions defined on Qc0 . Since

∂Uk0

∂rα(r0) = ∇αU(r0, p0) = 0

and∂2Uk0

∂rα∂rβ(r0) = ∇α∇βU(r0, p0)À 0,

the function Vk0 is positive definite in some neighborhood of the relativeequilibrium (rα0 , 0) ∈ Qc0 . The same is valid for the function Wk0 if ε issmall enough.

Now we show that Wk0 (as a function on Qc0) is negative definite. Cal-culate the derivative of Wk0 along the flow:

Wk0 = gαβ rαrβ +

12gαβ r

αrβ + IabPaPb

+12IabPaPb + V + ε

σ∑α=1

((rα)2 + uαrα

). (8.2.32)

Using the explicit representation of equation (8.2.12), we obtain

gαβ rβ + gαβ r

β =12∂gβγ∂rα

rβ rγ − ∂V

∂rα− 1

2∂Iab

∂rαPaPb −DcbαIbdPcPd

−DβαbIbcPcrβ − BcαβPcrβ −Kαβγ rβ rγ . (8.2.33)

Therefore,

gαβ rαrβ +

12gαβ r

αrβ + IabPaPb +12IabPaPb + V

= −12gαβ r

αrβ +12∂gβγ∂rα

rαrβ rγ − ∂V

∂rαrα − 1

2∂Iab

∂rαPaPbrα

−DcbαIbdPcPdrα −DβαbIbcPcrαrβ − BcαβPcrαrβ

−Kαβγ rαrβ rγ + IabDcbαPaPcrα +12∂Iab

∂rαPaPbr

α + V .

Using skew-symmetry of Bcαβ and Kαβγ with respect to α, β and cancelingthe terms

−12gαβ r

αrβ +12∂gβγ∂rα

rαrβ rγ − ∂V

∂rαrα + V ,

246 8. Energy Based Methods for Stability

we obtain

gαβ rαrβ +

12gαβ r

αrβ + IabPaPb +12IabPaPb + V

= −DβαbIbcPcrαrβ +(

12∂Iab

∂rα+ IacDbcα

)(PaPb − PaPb) rα. (8.2.34)

Substituting (8.2.34) in (8.2.32) and determining rα from (8.2.33)

Wk0 = −DβαbIbcPcrαrβ + εσ∑α=1

(rα)2

− εσ∑γ=1

gαβuγ(∂V

∂rα+

12∂Iab

∂rαPaPb +DcbαIbdPcPd

)

+ ε

σ∑γ=1

gαγuγ(−gαβ rβ +

12∂gβγ∂rα

rβ rγ − BcαβPcrβ

−DβαbIbcPcrβ −Kαβγ rβ rγ)

+12∂Iab

∂rα(PaPb − PaPb) rα + IacDbcα (PaPb − PaPb) rα. (8.2.35)

Since∂V

∂rα+

12∂Iab

∂rαPaPb +DcbαIbdPcPd = 0

at the equilibrium and the linear terms in the Taylor expansions of P andP are the same,

∂V

∂rα+

12∂Iab

∂rαPaPb +DcbαIbdPcPd = Fαβu

β + nonlinear terms,(8.2.36)

where

Fαβ =∂

∂rβ

(∂V

∂rα+

12∂Iab

∂rαPaPb +DcbαIbdPcPd

)=

∂2V

∂rα∂rβ+

12∂2Iab

∂rα∂rβPaPb +

∂Iab

∂rαPa∂Pb∂rβ

+∂

∂rβ(DcbαIbd

)PcPd +DcbαIbd

(∂Pc∂rβ

Pd + Pc∂Pd∂rβ

)=

∂2V

∂rα∂rβ+

12∂2Iab

∂rα∂rβPaPb +

∂Iab

∂rαPaDcbβPc

+∂

∂rβ(DcbαIbd

)PcPd +DcbαIbd

(DacβPd + PcDadβPa

)= ∇α∇βU.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 247

In the last formula all the terms are evaluated at the equilibrium.Taking into account that gαβ = δαβ +O(u), that the Taylor expansion of

PaPb−PaPb starts from the terms of the second order, and using (8.2.36),we obtain from (8.2.35)

Wk0 = −DβαbIbc(r0)p0c rαrβ − εFαβuαuβ + ε

σ∑α=1

(rα)2

− ε(DβαbIbc(r0)p0

c + Bcαβ(r0)p0c

)uαrβ + cubic terms.

Therefore, the condition (Dαβb +Dβαb)Ibcp0c À 0 implies that Wk0 is neg-

ative definite if ε is small enough and positive. Thus, Wk0 is a Lyapunovfunction for the flow on Qc0 , and therefore the equilibrium (rα0 , 0) for theflow on Qc0 is asymptotically stable.

Using the same arguments we used in the proof of Theorem 8.2.5, we con-clude that the equilibria on the nearby invariant manifolds Qk are asymp-totically stable as well. ¥

There is an alternative way to state the above theorem, which uses thebasic intuition we used to find the Lyapunov function.

Theorem 8.2.14 (The nonholonomic energy-momentum method).Under the assumption that H2 holds, the point qe = (rα0 , p

0a) is a relative

equilibrium if and only if there is a ξ ∈ gqe such that qe is a critical pointof the augmented energy Eξ = E − 〈p− P (r, k), ξ〉. Assume that

(i) δ2Eξ restricted to TqeQk is positive definite (here δ denotes differen-tiation by all variables except ξ);

(ii) the quadratic form defined by the flow derivative of the augmentedenergy is negative definite at qe.

Then H3 holds and this equilibrium is Lyapunov stable and asymptoticallystable in the directions of due to the invariant manifolds (8.2.31).

Proof. We have already shown in Theorem 8.2.6 that positive definite-ness of δ2Eξ|TqeQk is equivalent to the condition ∇α∇βU À 0. To com-plete the proof, we need to show that the requirement (ii) of the theoremis equivalent to the condition (Dαβb + Dβαb)Ibc(r0)p0

c À 0. Compute theflow derivative of Eξ:

Eξ = E − 〈p− P , ξ〉 = −〈p− P , ξ〉= (Dbaαpbrα +Dαβarαrβ −DbaαPbrα)ξa.

Since at the equilibrium p = P , ξa = Iabpb, and E = 0 (Theorem 8.2.4),we obtain

Eξ = −DαβaIab(r0)p0b rαrβ .

The condition Eξ À 0 is thus equivalent to (Dαβb+Dβαb)Ibc(r0)p0c À 0. ¥

248 8. Energy Based Methods for Stability

For some examples, such as the roller racer, we need to consider a de-generate case of the above analysis. Namely, we consider a nongeneric case,when U = 1

2Iab(r)papb (the original system has no potential energy), and

the components of the locked inertia tensor Iab satisfy the condition

12∂Iab

∂rα+ IacDbcα = 0. (8.2.37)

Consequently, the covariant derivatives of the amended potential are equalto zero, and the equations of motion (8.2.12), (8.2.13) become

d

dt

(gαβ r

β)− 1

2∂gβγ∂rα

rβ rγ = −DβαbIbcpcrβ − Bcαβpcrβ −Kαβγ rβ rγ ,(8.2.38)

d

dtpb = Dcbαpcrα +Dαβbrαrβ . (8.2.39)

Thus, we obtain an m+σ-dimensional manifold of equilibria r = r0, p = p0

of these equations. Further, we cannot apply Theorem 8.2.13 because thecondition ∇2U À 0 fails. However, we can do a similar type of stabilityanalysis as follows.

As before, set

Vk = E(r, r, P (r, k)) =12gαβ r

αrβ +12Iab(r)Pa(r, k)Pb(r, k).

Note that P satisfies the equation

∂Pb∂rα

= DcbαPc,

which implies that

∂rα

(12Iab(r)PaPb

)=

12∂Iab

∂rαPaPb + IabPa

∂Pb∂rα

=(

12∂Iab

∂rα+ IabDcbα

)PaPb = 0.

Therefore12IabPaPb = const

andVk =

12gαβ r

αrβ

(up to an additive constant). Thus, Vk is a positive definite function withrespect to r. Compute Vk:

Vk = gαβ rαrβ + gαβ r

αrβ = −DβαbIbcpcrαrβ +O(r3).

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 249

Suppose that (Dαβb + Dβαb)(r0)Ibc(r0)p0c À 0. Now the linearization

of equations (8.2.38) and (8.2.39) about the relative equilibria given bysetting r = 0 has m + σ zero eigenvalues corresponding to the r and pdirections. Since the matrix corresponding to r-directions of the linearizedsystem is of the form D + G, where D is positive definite and symmetric(in fact, D = 1

2 (Dαβb +Dβαb)(r0)Ibc(r0)p0c) and G is skew-symmetric, the

determinant of D+G is not equal to zero. This follows from the observationthat xt(D+G)x = xtDx > 0 for D positive-definite and G skew-symmetric.Thus using Theorem 8.2.10, we find that the equations of motion have localintegrals

r = R(r, k), p = P(r, k).

Therefore Vk restricted to a common level set of these integrals is a Lya-punov function for the restricted system. Thus, an equilibrium r = r0,p = p0 is stable with respect to r, r, p and asymptotically stable withrespect to r if

(Dαβb +Dβαb)Ibc(r0)p0c À 0. (8.2.40)

Summarizing, we have:

Theorem 8.2.15. Under assumptions H2 if U = 0 and the conditions(8.2.37) and (8.2.40) hold, the nonholonomic equations of motion have anm+σ-dimensional manifold of equilibria parametrized by r and p. An equi-librium r = r0, p = p0 is stable with respect to r, r, p and asymptoticallystable with respect to r.

The Roller Racer

The roller racer provides an illustration of Theorem 8.2.15. Recall that theLagrangian and the constraints are

L =12m(x2 + y2) +

12I1θ

2 +12I2(θ + φ)2

and

x = cos θ(d1 cosφ+ d2

sinφθ +

d2

sinφφ

),

y = sin θ(d1 cosφ+ d2

sinφθ +

d2

sinφφ

).

The configuration space is SE(2) × SO(2) and, as observed earlier, theLagrangian and the constraints are invariant under the left action of SE(2)on the first factor of the configuration space.

The nonholonomic momentum is

p = m(d1 cosφ+ d2)(x cos θ + y sin θ) + [(I1 + I2)θ + I2φ] sinφ.

250 8. Energy Based Methods for Stability

See Tsakiris [1995] for details of this calculation. The momentum equationis

p =((I1 + I2) cosφ−md1(d1 cosφ+ d2))m(d1 cosφ+ d2)2 + (I1 + I2) sin2 φ

+(d1 + d2 cosφ)(I2d1 cosφ− I1d2)m(d1 cosφ+ d2)2 + (I1 + I2) sin2 φ

φ2.

Rewriting the Lagrangian using p instead of θ, we obtain the energyfunction for the roller racer:

E =12g(φ)φ2 +

12I(φ)p2,

where

g(φ) = I2 +md2

2

sin2 φ−

[m(d1 cosφ+ d2)d2 + I2 sin2 φ

]2sin2 φ

[m(d1 cosφ+ d2)2 + (I1 + I2) sin2 φ

]and

I(φ) =1

(d1 cosφ+ d2)2 + (I1 + I2) sin2 φ. (8.2.41)

The amended potential is given by

U =p2

2[(d1 cosφ+ d2)2 + (I1 + I2) sin2 φ],

which follows directly from (8.2.11) and (8.2.41).Straightforward computations show that the locked inertia tensor I(φ)

satisfies the condition (8.2.37), and thus the roller racer has a two di-mensional manifold of relative equilibria parametrized by φ and p. Theserelative equilibria are motions of the roller racer in circles about the pointof intersection of lines through the axles. For such motions, p is the systemmomentum about this point and φ is the relative angle between the twobodies.

Therefore, we may apply the energy-momentum stability conditions (8.2.40)obtained above for the degenerate case. Multiplying the coefficient of thenontransport term of the momentum equation, evaluated at φ0, by I(φ)p0

and omitting a positive denominator, we obtain the condition for stabilityof a relative equilibrium φ = φ0, p = p0 of the roller racer:

(d1 + d2 cosφ0)(I2d1 cosφ0 − I1d2)p0 > 0.

Note that this equilibrium is stable modulo SE(2) and in addition asymp-totically stable with respect to φ.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 251

Nonlinear Stability by the Lyapunov-Malkin Method

Here we study stability using the Lyapunov-Malkin approach; correspond-ingly, we do not a priori assume the hypotheses H1 (skewness of Dαβb inα, β), H2 (a curvature is zero) or H3 (definiteness of second variations).Rather, at the end of this section we will make eigenvalue hypotheses.

We consider the most general case, when the connection due to the trans-port part of the momentum equation is not necessary flat and when thenontransport terms of the momentum equation are not equal to zero. In thecase when gqe is commutative, this analysis was done by Karapetyan [1980].Our main goal here is to show that this method extends to the noncom-mutative case as well.

We start by computing the linearization of equations (8.2.12) and (8.2.13).Introduce coordinates uα, vα, and wa in the neighborhood of the equilib-rium r = r0, p = p0 by the formulae

rα = rα0 + uα, rα = vα, pa = p0a + wa.

The linearized momentum equation is

wb = Dcbα(r0)p0cvα.

To find the linearization of (8.2.12), we start by rewriting its right hand sideexplicitly. Since R = 1

2gαβ rαrβ− 1

2Iabpapb−V , equation (8.2.12) becomes

gαβ rβ + gαβ r

αrβ − 12∂gβγ∂rα

rβ rγ +12∂Iab

∂rαpapb +

∂V

∂rα

= −DacαIcdpapd −DβαcIcaparβ − Baαβparβ −Kαβγ rβ rγ .

Keeping only the linear terms, we obtain

gαβ(r0)rβ +∂2V

∂rα∂rβ(r0)uβ +

12∂2Iab

∂rα∂rβ(r0)p0

ap0b u

β +∂Iab

∂rα(r0)p0

a wb

= −DacαIcd(r0)p0a wd −DacαIcd(r0)p0

d wa −∂DacαIcd∂rβ

(r0)p0ap

0d u

β

−DβαcIca(r0)p0a v

β − Baαβ(r0)p0a v

β .

Next, introduce matrices A, B, C, and D by

Aαβ = −(DβαcIca(r0)p0

a + Baαβ(r0)p0a

),

Bαβ = −(

∂2V

∂rα∂rβ(r0) +

12∂2Iab

∂rα∂rβ(r0)p0

ap0b +

∂DacαIcb∂rβ

(r0)p0ap

0b

),

(8.2.42)

Caα = −(∂Iab

∂rα(r0)p0

b +DbcαIca(r0)p0b +DacαIcb(r0)p0

b

), (8.2.43)

Daα = Dcaα(r0)p0c . (8.2.44)

252 8. Energy Based Methods for Stability

Using these notations and making a choice of rα such that gαβ(r0) = δαβ ,we can represent the equation of motion in the form

uα = vα, (8.2.45)

vα = Aαβv

β + Bαβu

β + Cαawa + Vα(u, v, w), (8.2.46)

wa = Daαvα + Wa(u, v, w), (8.2.47)

where V and W stand for nonlinear terms, and where

Aαβ = δαγAγβ ,

Bαβ = δαγBγβ ,

Cαa = δαγCaγ .

(Or Aαβ = gαγAγβ , Bα

β = gαγBγβ , Cαa = gαγCaγ if gαβ(r0) 6= δαβ .) Notethat

Wa =(Dcaα(p0

c + wc)−Daα

)vα +Dαβavαvβ . (8.2.48)

The next step is to eliminate the linear terms from (8.2.47). Putting

wa = Daαuα + za,

(8.2.47) becomesza = Za(u, v, z),

where Za(u, v, z) represents nonlinear terms. Formula (8.2.48) leads to

Za(u, v, z) = Zaα(u, v, z)vα.

In particular, Za(u, 0, z) = 0. The equations (8.2.45), (8.2.46), (8.2.47) inthe variables u, v, z become

uα = vα,

vα = Aαβv

β + (Bαβ + CαaDaβ)uβ + Cαaza + Vα(u, v, za + Daαu

α),

za = Za(u, v, w).

Using Lemma 8.2.12, we find a substitution xα = uα+φα(z), yα = vα suchthat in the variables x, y, z we obtain

xα = yα +Xα(x, y, z),

yα = Aαβy

β + (Bαβ + CαaDaβ)xβ + Y α(x, y, z), (8.2.49)

za = Za(x, y, z),

where the nonlinear terms X(x, y, z), Y (x, y, z, ), Z(x, y, z) vanish if x = 0and y = 0. Therefore, we can apply the Lyapunov-Malkin Theorem andconclude:

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 253

Theorem 8.2.16. The equilibrium x = 0, y = 0, z = 0 of the system(8.2.49) is stable with respect to x, y, z and asymptotically stable withrespect to x, y, if all eigenvalues of the matrix(

0 IB + CD A

)(8.2.50)

have negative real parts.

The Lyapunov-Malkin and the Energy-Momentum Methods

Here we introduce a forced linear Lagrangian system associated with ournonholonomic system. The linear system will have the matrix (8.2.50).Then we compare the Lyapunov-Malkin approach and the energy-momentumapproach for systems satisfying hypothesis H2.

Thus, we consider the system with matrix (8.2.50)

x = y,

y = Ay + (B + CD)x.(8.2.51)

According to Theorem 8.2.16, the equilibrium x = 0, y = 0, z = 0 of(8.2.49) is stable with respect to x, y, z and asymptotically stable withrespect to x, y if and only if the equilibrium x = 0, y = 0 of (8.2.51) isasymptotically stable. System (8.2.51) may be viewed as a linear uncon-strained Lagrangian system with additional forces imposed on it. Put

C = −12((B + CD) + (B + CD)t

),

F =12((B + CD)− (B + CD)t

),

D = −12(A + At

),

G =12(A−At

).

The equations become

xα = −Cαβ xβ + Fαβ xβ −Dα

β xβ +Gαβ x

β . (8.2.52)

These equations are the Euler-Lagrange equations with dissipation andforcing for the Lagrangian

L =12

σ∑α=1

(xα −Gαβxβ)2 − 12Cαβx

αxβ ,

with the Rayleigh dissipation function

12Dαβx

αxβ ,

254 8. Energy Based Methods for Stability

and the nonconservative forces

Fαβ xβ .

Note that Dαβ = (Dαβb +Dβαb)Iab(r0)p0a.

The next theorem explains how to compute the matrices C and F usingthe amended potential of our nonholonomic system.

Theorem 8.2.17. The entries of the matrices C and F in the dissipativeforced system (8.2.52), which is equivalent to the linear system (8.2.51), are

Cαβ =12

(∇α∇β +∇β∇α)U(r0, p0), Fαβ =12

(∇α∇β −∇β∇α)U(r0, p0).

Proof. Recall that the operators of covariant differentiation due to thetransport equation are (see (8.2.17))

∇α =∂

∂rα+Dcaαpc

∂pa.

Consequently,

∇β∇α = ∇β(

∂rα+Dcaαpc

∂pa

)=(

∂rβ+Ddbβpd

∂pb

)(∂

∂rα+Dcaαpc

∂pa

)=

∂2

∂rβ∂rα+

∂rβ

(Dcaα

∂pa

)pc +Ddbβpd

∂2

∂pb∂rα

+DdbβpdDcaα∂

∂pb

(pc

∂pa

).

Therefore, for the amended potential U = V + 12Iabpapb we obtain

∇β∇αU =∂2V

∂rβ∂rα+

12∂2Iab

∂rβ∂rαpapb +

∂rβ(DcaαIab

)pcpd

+Ddbβ∂Iab

∂rαpapd +DdbβDcaαIabpcpd +DdbβDbaαIacpcpd.

Formulae (8.2.42), (8.2.43), and (8.2.44) imply that

(BC + D)αβ =∂2V

∂rβ∂rα(r0) +

12∂2Iab

∂rβ∂rα(r0)p0

ap0b

+∂

∂rβ(DcaαIab

)(r0)p0

cp0d +Ddbβ

∂Iab

∂rα(r0)p0

ap0d

+DdbβDcaαIab(r0)p0cp

0d +DdbβDbaαIac(r0)p0

cp0d

= ∇β∇αU(r0, p0).

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 255

Therefore

Cαβ =12

(∇α∇β +∇β∇α)U(r0, p0), Fαβ =12

(∇α∇β −∇β∇α)U(r0, p0).

¥

Observe that the equilibrium x = 0, y = 0, z = 0 of (8.2.49) is stable withrespect to x, y, z and asymptotically stable with respect to x, y if and onlyif the equilibrium x = 0, y = 0 of the above linear Lagrangian system isasymptotically stable. The condition for stability of the equilibrium r = r0,p = p0 of our nonholonomic system becomes: all eigenvalues of the matrix

(0 I

∇β∇αU(r0, p0) −(DβαbIab(r0) + Baαβ(r0))p0a

)

have negative real parts.If the transport equation is integrable (hypothesis H2), then the opera-

tors ∇α and ∇β commute, and the corresponding linear Lagrangian system(8.2.52) has no nonconservative forces imposed on it. In this case the suffi-cient conditions for stability are given by the Thompson Theorem (Thomp-son and Tait [1987], Chetaev[1959]): the equilibrium x = 0 of (8.2.52) isasymptotically stable if the matrices C and D are positive definite. Theseconditions are identical to the energy-momentum conditions for stabilityobtained in Theorem 8.2.13. Notice that if C and D are positive definite,then the matrix (8.2.50) is positive definite. This implies that the matrixA in Theorem 8.2.10 has spectrum in the left half plane. Further, our coor-dinate transformations here give the required form for the nonlinear termsof Theorem 8.2.10. Therefore, the above analysis shows that hypothesis H3implies the hypotheses of Theorem 8.2.10.

Remark. On the other hand (cf. Chetaev [1959]), if the matrix C isnot positive definite (and thus the equilibrium of the system x = −Cxis unstable), and the matrix D is degenerate, then in certain cases theequilibrium of the equations x = −Cx−Dx+Gx may be stable. Therefore,the conditions of Theorem 8.2.13 are sufficient, but not necessary.

The Rattleback

Here we outline the stability theory of the rattleback to illustrate the resultsdiscussed above. The details may be found in Karapetyan [1980, 1981] andMarkeev [1992].

256 8. Energy Based Methods for Stability

Recall that the Lagrangian and the constraints are

L =12[A cos2 ψ +B sin2 ψ +m(γ1 cos θ − ζ sin θ)2

]θ2

+12[(A sin2 ψ +B cos2 ψ) sin2 θ + C cos2 θ

]φ2

+12(C +mγ2

2 sin2 θ)ψ2 +

12m(x2 + y2

)+m(γ1 cos θ − ζ sin θ)γ2 sin θ θψ + (A−B) sin θ sinψ cosψ θφ

+ C cos θ φψ +mg(γ1 sin θ + ζ cos θ)

andx = α1θ + α2ψ + α3φ, y = β1θ + β2ψ + β3φ,

where the terms were defined in §8.2.Using the Lie algebra element corresponding to the generator ξQ =

α3∂x + β3∂y + ∂φ we find the nonholonomic momentum to be

p = I(θ, ψ)φ+ [(A−B) sin θ sinψ cosψ −m(γ1 sin θ + ζ cos θ)γ2] θ

+[C cos θ +m(γ2

2 cos θ + γ1(γ1 cos θ − ζ sin θ))]ψ,

where

I(θ, ψ) = (A sin2 ψ +B cos2 ψ) sin2 θ + C cos2 θ

+m(γ22 + (γ1 cos θ − ζ sin θ)2)..

The amended potential becomes

U =p2

2I(θ, ψ)−mg(γ1 sin θ + ζ cos θ).

The relative equilibria of the rattleback are

θ = θ0, ψ = ψ0, p = p0

where θ0, ψ0, p0 satisfy the conditions

mg(γ1 cos θ0 − ζ sin θ0)I2(θ0, ψ0)

+[(A sin2 ψ0 +B cos2 ψ0 − C) sin θ0 cos θ0

−m(γ1 cos θ0 − ζ sin θ0)(γ1 sin θ0 + ζ cos θ0)]p2

0 = 0,

mgγ2I2(θ0, ψ0) + [(A−B) sin θ0 sinψ0 cosψ0

−mγ2(γ1 sin θ0 + ζ cos θ0)] p20 = 0,

which are derived from ∇θU = 0, ∇ψU = 0.

8.2 The Energy-Momentum Method for the Stability of Nonholonomic Systems 257

In particular, consider the relative equilibria

θ =π

2, ψ = 0, p = p0,

that represent the rotations of the rattleback about the vertical axis ofinertia. For such relative equilibria ξ = ζ = 0, and therefore the conditionsfor existence of relative equilibria are trivially satisfied with an arbitraryvalue of p0. Omitting the computations of the linearized equations for therattleback, which have the form discussed in Section 8.2.3 (see Karapetyan[1980] for details), and the corresponding characteristic polynomial, we juststate here the Routh-Hurwitz conditions for all eigenvalues to have negativereal parts: (

R− P p20

B2

)p2

0

B2− S > 0, S > 0, (8.2.53)

(A− C)(r2 − r1)p0 sinα cosα > 0. (8.2.54)

If these conditions are satisfied, then the relative equilibrium is stable, andit is asymptotically stable with respect to θ, θ, ψ, ψ.

In the above formulae r1, r2 stand for the radii of curvature of the bodyat the contact point, α is the angle between horizontal inertia axis ξ andthe r1-curvature direction, and

P = (A+ma2)(C +ma2),

R =[(A+ C −B + 2ma2)2

− (A+ C −B + 2ma2)ma(r1 + r2) +m2a2r1r2

] p20

B2

−[(A−B)

p20

B2+m(a− r1 sin2 α− r2 cos2 α)

(g + a

p20

B2

)](A+ma2)

−[(C −B)

p20

B2+m(a− r2 sin2 α− r1 cos2 α)

(g + a

p20

B2

)](C +ma2),

S = (A−B)(C −B)p4

0

B4+m2(a− r1)(a− r2)

(g +m

p20

B2

)2

+mp2

0

B2

(g + a

p2

B2

)[A(a− r1 cos2 α− r2 sin2 α)

+ C (a− r1 sin2 α− r2 cos2 α)−B(2a− r1 − r2)].

Conditions (8.2.53) impose restrictions on the mass distribution, themagnitude of the angular velocity, and the shape of the rattleback only.Condition (8.2.54) distinguishes the direction of rotation corresponding tothe stable relative equilibrium. The rotation will be stable if the largest(smallest) principal inertia axis precedes the largest (smallest) direction ofcurvature at the point of contact.

258 8. Energy Based Methods for Stability

The rattleback is also capable of performing stationary rotations with itscenter of mass moving at a constant rate along a circle. A similar argumentgives the stability conditions in this case. The details may be found inKarapetyan [1981] and Markeev [1992].

This is page 259Printer: Opaque this

9Energy Based Methods forStabilization

9.1 Controlled Lagrangian Methods

9.1.1 Rigid Body with Rotors

In this sections we apply the Energy Casimir method to the stabilization ofa rigid body with rotors (see Bloch, Krishnaprasad, Marsden and Sanchez[1992]).

Recall that this system consists of a rigid body with three rotors alignedalong the principal axes of the rigid body. The rotors spin under the influ-ence of a vector u of torques acting on the rotors.

The equations of motion are:

m = m× ωl = u

Here, u is the vector of torques exerted on the rotors, so that u = 0 corre-sponds to the uncontrolled case.I1 > I2 > I3 are the rigid body moments of inertia, J1 = J2 and J3 are

the rotor moments of inertia, ω = (ω1, ω2, ω3) is the angular velocity of thebody, and α is the relative angle of the rotor. The momenta are

mi = (J1 + I1)ωi i = 1, 2m3 = (J3 + I3)ω3 + J3α

l3 = J3(ω3 + α).

260 9. Energy Based Methods for Stabilization

Writing the equations out, we get

m1 =(

1I3− 1λ2

)m2m3 −

l3m2

I3

m2 =(

1λ1− 1I3

)m1m3 +

l3m1

I3

m3 =(

1λ2− 1λ1

)m1m2

.= a3m1m2

l3 = u.

Here, λi = Ii + Ji.If u = 0, then l3 is a constant of motion and the reduced system is

Hamiltonian where

H =12

(m2

1

λ1+m2

2

λ2+

(m3 − l3)2

I3

)+

12l23.

Choose the feedback control law

u = ka3m1m2,

where k is a gain parameter, then the system retains the S1 symmetry andPk = l3 − km2 is a new conserved quantity. The closed loop equations are

m1 = m2

((1− k)m3 − Pk

I3

)− m3m2

λ2

m2 = −m1

((1− k)m3 − Pk

I3

)+m1m3

λ1

m3 = a3m1m2.

These equations are Hamiltonian with

H =12

(m2

1

λ1+m2

2

λ2+

((1− k)m3 − Pk)2

(1− k)I3

)+

12

P 2k

J3(1− k),

using the Lie-Poisson (rigid body) Poisson structure on so(3)∗.Two special cases are of interest: k = 0, the uncontrolled case, and

k = J3/λ3, the driven case where α = constant.Now consider the case P = 0 and the special equilibrium (0,M, 0). For

k > 1 − J3/λ2, the equilibrium (0,M, 0) is stable. This is proved by theEnergy-Casimir method. We look at H + C where C = ϕ(||m||2). Pick ϕso that

δ(H + C)|(0,M,0) = 0,

then one computes that δ2(H+C) is negative definite if k > 1−J3/λ2 andϕ′′(M2) < 0.

9.2 Controlled Lagrangian methods for Nonholonomic Systems 261

9.1.2 The pendulum on a cart

9.2 Controlled Lagrangian methods forNonholonomic Systems

9.3 Averaging and Energy Methods for theDynamics of Mechanical Systems

9.3.1 Energy and Curvature

The chapter so far has given us an indication of the role energy-like func-tions play in the design and analysis of control laws for physical systems.Continuing the discussion, we consider a simple (in the sense of Abrahamand Marsden, [1978]) mechanical system characterized by a kinetic energy,12 qTM(q)q, and a potential energy,V (q). We have seen in Section that

the equations of motion for this system are written: Fix afterTony’s reorga-nization.

n∑j=1

mij(q)qj +n∑

j,k=1

Γkjiqk qj +∂V

∂qi= 0, (i = 1, . . . , n), (9.3.1)

where

Γkji =12

(∂mij

∂qk+∂mik

∂qj− ∂mkj

∂qi).

The Γkji are called Christoffel symbols of the first kind. The terms theydefine in equation (9.3.1) describe the inertial forces (Coriolis and centrifu-gal) which affect the system. It is clear from this equation that if M doesnot depend on q, Γkji = 0 for all k, j, i, and inertial forces will not playa role in the system’s dynamics. An interesting question is “When can wefind a change of coordinates such that in the new coordinate system theinertia matrix is constant?” Note that the inertia matrix M is constant(i.e. does not depend on the configuration variable q) if and only if thereis a system of generalized coordinates (y1, . . . , yn) in terms of which thekinetic energy is expressed as 1

2

∑ni=1 y

2i . (Proof: In the case that M is

constant, we let Y denote the constant, symmetric, positive definite matrixsquare root of M and let y = Y q. The y = Y q, and the kinetic energy is12

∑i,jmij qiqj = 1

2

∑i y

2i . The reverse implication requires no proof.)

Hence, whether we can find the desired change of coordinates amounts towhether there exists a diffeomorphism y = F (q) such that 1

2 qTMq = 1

2‖y‖2.

Suppose such an F exists. Then y = ∂F∂q q, and ∂F

∂q

T ∂F∂q = M . For such an

M we compute

∂mij

∂qk=

n∑`=1

( ∂2F`∂qi∂qj

∂F`∂qj

+∂F`∂qi

∂2F`∂qj∂qk

).

262 9. Energy Based Methods for Stabilization

From this it follows from an easy calculation that

Γkji =12

(∂mij

∂qk+∂mik

∂qj− ∂mkj

∂qi

)=

n∑`=1

∂2F`∂qj∂qk

∂F`∂qi

. (9.3.2)

As in Section xxx, the Christoffel symbols (of the first kind) Γkji are as-sociated with a connection (the so-called Levi-Civita connection definedby the inertia tensor M) and a Riemannian curvature tensor. Being ableto find this F which transforms the given generalized coordinates into anew system in terms of which the inertia matrix is the identity implies theRiemannian curvature tensor is zero. To see this, we first write down thecorresponding Christoffel symbols of the second kind:

Γ`kj =n∑i=1

m`iΓkji.

A curvature tensor is prescribed by the corresponding Riemann sysmbol ofthe first kind:

Rijk` =∂Γj`i∂qk

− ∂Γjki∂q`

+n∑σ=1

(ΓσjkΓi`σ − Γσj`Γikσ

).

Working this out for the case at hand,

∂Γj`i∂qk

=n∑σ=1

( ∂3Fσ∂qk∂qj∂q`

∂Fσ∂qi

+∂2Fσ∂qj∂q`

∂2Fσ∂qi∂qk

),

and a permutation of indices gives a similar expression for ∂Γjki∂qk`

. It isstraightforward, although a little tedious, to show that

n∑σ=1

(ΓσjkΓi`σ − Γσj`Γikσ

)=

n∑β=1

( ∂2Fβ∂qj∂qk

∂2Fβ∂qi∂q`

− ∂2Fβ∂qj∂q`

∂2Fβ∂qi∂qk

).

Hence Rijkl = 0, proving the claim. It is convenient to think of the in-ertia tensor M as a Riemannian metric on the configuration manifold.The Christoffel symbolds Γkji (or equivalently Γikj) define the correspond-ing Riemannian connection (also called the Levi-Civita connection). TheReimann symbols of the first kind are components of the Riemannian cur-vature tensor. Their vanishing identically is essentially a necessary and suf-ficient condition for the exisitence of a coordinate transformation F (q) = ysuch that in y−coordinates the inertia tensor (=Riemannian metric) is theidentity matrix. The reader is referred to Kobayashi and Nomizu [1963] orWolf [1972] for a broader discussion of curvature and its vanishing.

AVERAGING AND ENERGY 263

Definition 9.3.1. A simple mechanical system of the form (9.3.1) willbe called flat if the corresponding Riemannian curvature tensor (i.e. the setof Riemann symbols of the first kind) is zero.

It is of interest to see how this type of flatness simplifies control designs.We begin by considering the case of fully actuated Lagrangian and Hamil-tonian systems; in the following section we turn our attention to the morecomplex (and more intersting) case of super-articulated (or underactuated)systems. Consider a mechanical control system specified by a LagrangianL(q, q) = 1

2 qTM(q)q − V (q) and equations of motion

d

dt

∂L

∂qi− ∂L

∂qi= ui (i = 1, . . . , n),

where each ui(·) is a piecewise continuous control input (function of timedefined on a suitable interval). The conjugate momentum vector for thissystem is p = M(q)q and in terms of the Hamiltonian defined by theLegendre transformation

H(q, p) = qT p− L(q, q)

=12pTM(q)−1p+ V (q), (9.3.3)

we define the corresponding Hamiltonian control system

q =∂H

∂p

p = −∂H∂q

+ u. (9.3.4)

It has been noted (See, e.g. Bedrosian [1992] and Spong [1998].) that if asystem is flat there is a canonical transformation together with a controland feedback transformation such that in terms of the new phase space andcontrol variables, the system (9.3.4) takes the form of a system of doubleintegrators.

To understand this, suppose the mechanical system is flat. Let N(q) bethe n×n positive definite symmetric square root of the inertia tensor M(q).If there exists a function F (q) such that ∂F

∂q = N(q) as above, then thecoordinate transformation Q = F (q), P = N(q)q is canonical—in the sensethat it preserves the symplectic form. We refer the reader to Marsden andRatiu [1994] or Arnold [1989] for a discussion of canonical transformations.That this transformation is canonical is easily seem if we write

P · dQ = qTN(q) · ∂F∂q

dq

= qTM(q)dq= p · dq.

264 9. Energy Based Methods for Stabilization

We rewrite the Hamiltonian in terms of Q,P−coordinates:

h(q, p) =12pTM(q)−1p+ V (q)

=12qTN(q)N(q)q + V (q)

=12‖P‖2 + V(Q),

=H(Q,P ),

where V is defined by V(Q) = V (q).

Proposition 9.3.2. In terms of the variables Q,P , the equations (9.3.4)may be written as

Q = P

P = − ∂V∂Q

+N−1u. (9.3.5)

Under the feedback u = Nv +N ∂V∂Q , which is independent of the conjugate

momentum variables P , this Hamiltonian control system is transformedinto the system of double integrators

Q = P

P = v (9.3.6)

Proof. The coordinate change associated with the Legendre transforma-tion may be written explicitly as

Q = F (q) (9.3.7)

P =(∂F∂q

)−1

p (9.3.8)

(since ∂F∂q q = (∂F∂q )−1p). We wish to find the equations of motion in terms

of Q,P−coordinates. Differentiating (9.3.7), it is clear that Q = P . Differ-entiating (9.3.8), we find that

P =[ ddt

(∂F∂q

)−1]p+

(∂F∂q

)−1

p

=[ ddt

(∂F∂q

)−1]p−

(∂F∂q

)−1 ∂H

∂q+(∂F∂q

)−1

u.

Using the definition of the coordinate transformation (9.3.7)-(9.3.8) to-gether with the symmetry properties of ∂F

∂q = N(q), it can be shown that(∂Q∂q 0∂P∂q

∂P∂p

)(0 I−I 0

)( ∂Q∂q

T ∂P∂q

T

0 ∂P∂p

T

)=(

0 I−I 0

).

AVERAGING AND ENERGY 265

From this it follows that[ ddt

(∂F∂q

)−1]p−

(∂F∂q

)−1 ∂H

∂q= −∂H

∂Q= − ∂V

∂Q,

proving the first part of the proposition. The second part is clear. ¥

9.3.2 Second Order Generalized Control Systems

Curvature plays an important role in characterizing second order nonlin-ear control systems. We shall be especially interested in velocity-controlledsuper-articulated mechanical systems (as introduced and studied in Bail-lieul [1990,1993,1995,1997a,1997b]). These are systems having two charac-teristic features: (i) fewer control inputs than configuration space dimen-sions, and (ii) certain generalized coordinates together with the correspond-ing generalized velocities and accelerations are viewed as control inputs.There are several routes leading to a theory of velocity-controlled super-articulated mechanical systems. The reduction process outlined in Baillieul[1997b] will be discussed, but first we shall recall the global definition of asmooth, finite dimensional nonlinear control system.

In 1976, Brockett [1976] noted that local descriptions of nonlinear controlsystem dynamics in the form

q = f(q, x), (9.3.9)

where f : M ×U → TM , were not adequate descriptions of systems wherethe inputs depend on the states, and even on the time histories of thestates. In the natural global extension of (??), we view f as a bundle map.

Definition 9.3.3. A smooth nonlinear control system is a quadruple(B,M, π, f) such that

(i) (B,M, π) is a fiber bundle with total space B, base space M , andcanonical projection π : B →M , and

(ii) f : B → TM is a bundle morphism such that for each q ∈ M andx ∈ Uq = π−1(q), f(q, x) ∈ TqM .

For a complete introduction to the theory of fiber bundles, the reader isreferred to Husemoller [1994] or the classical treatise Steenrod [1951]. Recallthat for each q ∈M , there is a neighborhood V of q and a diffeomorphism φmapping π−1(V ) onto V ×π−1(q). Thus, as Brockett points out (Brockett,[1976], p. 16), by restricting our attention to such neighborhoods, it isalways possible to find a local representation of a smooth nonlinear controlsystem of the form (9.3.9).

To continue our study of controlled mechanical systems, it will be usefulto refine Brockett’s global approach to include systems which are secondorder (in the configuration variables q) and in which the controlling effects

266 9. Energy Based Methods for Stabilization

of the input variables are primarily due to their accelerations, x(·). It willbe useful to develop the notion of parallel displacements of vectorfields inTM along curves in B. More specifically, given a vectorfield X ∈ TB andany curve τ(t) ∈ B, the lifting of τ to a curve X(τ(t)) ∈ TB is projectedby the tangent mapping π∗ onto a curve X(τ(t)) ∈ TM . This associationleads to the desired notion of covariant differentiation of vectorfields in TMalong curves in B, and we shall prescribe this in terms of local coordinatesas follows. Let (q1, . . . , qn, x1, . . . , xm) denote local coordinates defined ina neighborhood of a point p ∈ B. Let

Xi =∂

∂qi, (i = 1, . . . , n),

Xi =∂

∂xi, (i = 1, . . . ,m)

be the associated vectorfields in TB. We may also view (q1, . . . , qn) as defin-ing local coordinates in a neighborhood of π(p) ∈ M . Then we similarlylet

Yi =∂

∂qi, (i = 1, . . . , n)

be the associated vectorfields in TM . In terms of this local coordinatedescription, for each Xi and each (q, x) ∈ B, the tangent mapping π∗associates a vector Y ∗i ∈ TqM by means of the formula

Y ∗i (q, x) = π∗Xi(q, x).

In general, Y ∗i (q, x) 6= Yi(q), but we may write

Y ∗i (q, x) =n∑j=1

αij(x)Yj(q), (i = 1, . . . , n).

By restricting to a smaller neighborhood if necessary, there is no loss ofgenerality in assuming that the n × n matrix A(x), whose ij-th entry isαij(x) is nonsingular for each x. For any point q ∈ M , π−1(q) ∈ B, eachvectorfield Xi (i = 1, . . . , n) and Xj (j = 1 . . . ,m), defines an integralcurve passing through π−1(q). For each such integral curve, expressed in ourcoordinate neighborhood as τ(t) = (q(t), x(t), there is a corresponding setof curves Y ∗i (q(t), x(t)) (i = 1 . . . , n) in the tangent bundle TM . Tangentsto these curves define a covariant derivative operator ∇ : TB×TM → TMwhich is prescribed in terms of the given local coordinate by means of(n + m)n2 functions Γkij, (i, j, k = 1, . . . , n), and Γkij, (i = 1, . . . ,m;j, k = 1, . . . , n) and the formulas

∇XiY ∗j =∑nk=1 ΓkijY

∗k , i, j = 1, . . . , n; (9.3.10)

∇XiY∗j =

∑nk=1 ΓkijY

∗k , i = 1, . . . ,m; (9.3.11)

j = 1, . . . , n.

AVERAGING AND ENERGY 267

Since the vectorfields Xi (and Xi) lie in the coordinate directions, it is clearwhat is meant if we write

∇qiY ∗j =n∑k=1

ΓkijY∗k ; ∇xiY ∗j =

n∑k=1

ΓkijY∗k . (9.3.12)

The covariant derivative operator allows us to study displacement of vec-torfields in TM along curves (q(t), x(t)) ∈ B. We refer to Kobayashi andNomizu, [?], for a clssical treatment of this topic. In the present develop-ment, we shall view points q in the base manifold as states of a nonlinearcontrol system (and eventually as generalized coordinates of a mechanicalcontrol system) and the points x in the fiber as control inputs. We shallbe interested in lifting curves in the total space B to curves in TB andprojecting these liftings onto TM . It is useful to keep in mind the followingcommutative diagram

TB - TM

?B M

- ?π

π∗where the vertical

arrows are the canonical tangent bundle projections. A curve (q(t), x(t)) inB lifts to a curve Xi(q(t), x(t)) in TB for each i = 1, . . . , n. The tangentmapping of the canonical projection π : B →M allows us to define a curveY ∗i (q(t), x(t)) = π∗Xi(q(t), x(t)) in TM . Expressed in terms of the givenlocal coordinates, let (q(t), x(t)) (0 ≤ t < t0) be any smooth curve definedin our coordinate neighborhood in B. Let

n∑i=1

αi(t)Xi(q(t), x(t)) +m∑i=1

βi(t)Xi(q(t), x(t))

be the corresponding tangent vector for each t in 0 ≤ t < t0. (This isa curve in TB.) The vectorfields Y ∗i (q, x) = π∗Xi(q, x), appearing in theformulas (9.3.10)-(9.3.11) depend on points in the total space B. In termsof the vectorfields Y ∗i , the image curve in TM may be expressed as

Y (t) =n∑i=1

αi(t)Y ∗i (q(t), x(t)). (9.3.13)

This projection defines a vectorfield (in TM) along the curve τ(t) = (q(t), x(t))in B. Differentiating Y with respect to t and using our covariant deriva-tive operator to project the result back onto TM , we obtain the covariantderivative in the direction τ , ∇τY , for each t in 0 ≤ t < t0:

n∑i=1

(αi(t)Y ∗i + αi(t)(

n∑j=1

∇qjY ∗i qj +m∑j=1

∇xjY ∗i xj)).

(Cf. Kobayashi and Nomizu, [?] p. 114.) Using the relationships (9.3.12),this expression may be rendered

n∑k=1

(αk(t) +

n∑i,j=1

Γkjiαiqj +n∑i=1

m∑j=1

Γkjiαixj

)Y ∗k . (9.3.14)

268 9. Energy Based Methods for Stabilization

Curves (9.3.13) which arise as models of velocity or accelration controlledmechanical systems are parallel along τ(t) in the sense of the followingdefinition.

Definition 9.3.4. We say that a curve

V (t) =n∑i=1

vi(t)Y ∗i (q(t), x(t))

is parallel along τ(t) = (q(t), x(t)) if the covariant derivative in the direc-tion τ , ∇τV is zero. Equivalently, in terms of local coordinates, V (t) isparallel along τ(t) if

n∑i=1

(vi(t)Y ∗i + vi(t)(

n∑j=1

[∇qjY ∗i

]qj +

m∑j=1

[∇xjY ∗i

]xj))

= 0.

An even more explicit rendering of the condition for Y in (9.3.13) tobe parallel along τ is that α(·) is the unique solution to the system ofdifferential equations

αk(t) +n∑

i,j=1

Γkjiαiqj +n∑i=1

m∑j=1

Γkjiαixj = 0 (k = 1, . . . , n).

Now we say that (9.3.13) defines a smooth acceleration-controlled secondorder system if (i) αi(t) = qi(t) +

∑mj=1 γij(q)xj(t) and (ii) the vector field

(9.3.13) is parallel along the curve (q(t), x(t)). In particular, if (9.3.13)is a trajectory of smooth acceleration-controlled second order system, thecoordinate functions qi(t) and xi(t) satisfy

qk +m∑j=1

γij(q)xjn∑

i,j=1

Γkjiqiqj +n∑i=1

m∑j=1

Γkjiqixj = 0, for i = 1, . . . , n,

(9.3.15)

where Γkji = Γkji + ∂γij∂qi

. This is a direct extension of the geometric char-acterization of second-order differential equations found in Abraham andMarsden, 1988, [?]. In general, all three components of the triple (x, x, x)enter the equation (9.3.15), but because we shall be principally concernedwith high-frequency periodic inputs, the most significant terms will be thoseinvolving x. (Suppose x(·) is a periodic vector valued function whose fun-damental frequency is ω and whose `∞ norm is O(1). Then the norm ofx(·) is O(ω) and the norm of x(·) is O(ω2).) It is precisely because (9.3.15)depends formally on x that we have called these systems acceleration con-trolled. In the next section, we shall study systems in which all γij(q) ≡ 0.We shall also show that the vanishing of a curvature-like quantity is a nec-essary condition for there to be a trnaformation to coordinates in terms ofwhich on x and x (but not x enter the equations of motion.

AVERAGING AND ENERGY 269

We have called systems of the form (9.3.15) acceleration controlled be-cause of the explicit dependence on the second derivative of the input func-tion x(·). Second order acceleration-controlled systems arise in the study ofsuper-articulated mechanical systems (Seto [1994]). (These have also beencalled under-actuated systems in the literature.) The general frameworkassumes there is given a Lagrangian L(y, y) defined on the tangent bundleTQ of the configuration space of a mechanical system. Suppose that thegeneralized coordinates can be partitioned as y = (r, q) and that exogenousgeneralized forces (control inputs) can be applied to only the coordinatesr, while the coordinate variables comprising q evolve freely, subject onlyto dynamic interactions with the r-variables. We assume dim r = m anddim q = n. The equations of motion for the system take the form

d

dt

∂L∂r− ∂L∂r

= u, (9.3.16)

d

dt

∂L∂q− ∂L∂q

= 0, (9.3.17)

where u is an m-vector of controls, and we suppose the mapping q 7→ ∂L∂q

is invertible. I.e. if the Lagrangian takes the form “kinetic minus potentialenergy,” i.e. L = 1

2 yTMy − V (y), and the inertia matrix is partitioned

conformably with (r, q), M =(N AAT M

), then M = M(r, q) is an

n×n invertible matrix. For the purposes of the model, we assume that thecomponents ui(·) are each piecewise analytic functions on [0,∞), althoughmore general inputs (e.g. impulse trains) are also of interest and amenableto study.

A simple reduction process eliminating the explicit dependence on the in-put u(·) leads to a system of equations of the desired form. We assume thatthe dynamics of (9.3.16) allow u(·) to be chosen such that r can be specifiedexactly. We may thus view r, r, r as control inputs in equation (9.3.17). Toemphasize this viewpoint, we write (x, v, a) = (r, r, r). Equation (9.3.17)then relates the state variables q and q to the inputs (x, v, a). This dynam-ical relationship may be obtained by applying the Euler-Lagrange operatorddt

∂∂q − ∂

∂q to the reduced Lagarangian

L(q, q;x, v) =12qTM(q, x)q + vTA(q, x)q − V(q;x, v), (9.3.18)

where V(q : x, v) = V (x, q) − 12vTN v. Although the reduction process

described above will play no further role in our theory, it is the main routeby which we are lead to velocity controlled Lagrangian systems of the form

270 9. Energy Based Methods for Stabilization

(9.3.18). The Lagrangian (9.3.18) gives rise to the Lagrangian dynamics

n∑j=1

mkj qj +m∑`=1

a`kv` +n∑

i,j=1

Γijk qiqj +n∑j=1

m∑`=1

Γ`jkv`qj = F (t), (k = 1, . . . , n),

(9.3.19)

where

Γijk =12

(∂mki

∂qj+∂mkj

∂qi− ∂mij

∂qk

),

Γ`jk =∂mkj

∂x`+∂a`k∂qj

− ∂a`j∂qk

,

and aij and mij are the ij-th entries in the m×n and n×n matrices A(q, x)and M(q, x) respectively. F (t) is a vector of generalized forces Fi(t) =∂V∂qi−∑mk,`=1

∂a`i∂xk

v`vk. These may be thought of as coming from a velocity-dependent potential if

∂2a`i∂qj∂xk

=∂2a`j∂qi∂xk

for all k, ` = 1, . . . ,m and i, j = 1 . . . , n.Just as mij denotes the ij−th element of M, let mij denote the ij−th

element of M−1. Multiplying both sides of (9.3.19) by mσk and summingover the index values k = 1, . . . , n, we obtain

qσ +m∑`=1

γσ`v` +n∑

i,j=1

Γσij qiqj +n∑j=1

m∑`=1

Γσ`jv`qj = Fσ(t), (σ = 1, . . . , n),

(9.3.20)

where

γσ` =n∑k=1

mσkak`,

Γσij =n∑k=1

mσkΓijk,

Γσ`j =n∑k=1

mσkΓ`jk, and

Fσ(t) =n∑k=1

mσkFk(t).

Modulo a minor change of notation (xj = vj), when the generalize potentialforces Fk(t) = 0, equations (9.3.15) and (9.3.20) are the same.

9.4 Flat Systems and Systems with Flat Inputs 271

It is of interest to develop a theory of normal forms for systems of theform (9.3.15) and (9.3.20). In this pursuit, we are looking for the appropri-ate analog of the double integrator form characterized in Proposition 9.3.2.The particular way in which our input variables (x, x, x) enter the equationstogether with the fact that the number m of input variables is generally lessthan the number n of configuration variables rules out the double integra-tor as a general normal form for acceleration controlled super-articulatedsystems. Nevertheless, we shall show that when certain curvature-like quan-tities vanish, we can find a coordinate system in which there is a high degreeof decoupling between inputs and configuration variables. Indeed, we shallshow that under certain “vanishing” hypotheses, there is a choice of gener-alized coordinates with respect to which the dynamic effects of the inputsenter through velocity-dependent potential terms.

9.4 Flat Systems and Systems with FlatInputs

We begin by considering the structure of systems which are flat in thesense of Definition 9.3.1. In this setting, a system of the form (9.3.15) or(9.3.20) is said to be flat if the Riemannian curvature tensor prescribed bythe Christoffel symbols Γkij vanishes. In terms of the Lagrangian system(9.3.18), if the inertia tensor M does not depend on the input x, then thevanishing of the curvature tensor is necessary and sufficient for there to bea change of generalized coordinates q = F (q) such that the inertia tensorexpressed in terms of the q−coordinates has entries

δij =

1 i = j0 i 6= j.

Fot the case in whichM =M(q, x), explicitly depends on the input x, it isstill possible to have the Riemannian curvature tensor vanish. (Of course,in this case, the Riemann symbols will have no dependence on the variablex.) It is then possible to find an x−dependent change of coordinates,q = F (q, x) such that in the q−coordinate system, the Lagrangian (9.3.18)takes the form

L(q, ˙q;x, x) =12‖ ˙q‖2 + ˙qT AT (q, x)x− V(q;x, x), (9.4.1)

where

AT (q, x) =∂G

∂q

T

AT (G(q, x), x)− ∂F

∂x(G(q, x), x),

and

V(q;x, x) = V(G(q, x);x, x) +12xT

∂F

∂x

T ∂F

∂xx,

272 9. Energy Based Methods for Stabilization

with q = G(q, x) being the x−dependent inverse of the diffeomorphismq = F (q, x). An interesitng question is “When do the coupling terms A(q, x)vanish?” The following definitions are useful in formulating the answer.

Definition 9.4.1. (i) For the second-order generalized control system (9.3.20),we say that the hatted symbols Γkij define the input connection. The inputconnection is said to be flat if the associated Riemann symbols of the secondkind

Rγ`ij =∂Γγ`j∂xi

−∂Γγij∂x`

+n∑k=1

(Γk`jΓγik − ΓkijΓ

γ`k)

vanish for all i` = 1, . . . ,m; j, γ = 1, . . . , n. (ii) Given the second ordergeneralized control system (9.3.20), we say that the pair (x, x) constitutesa set of flat inputs if the input connection is flat.¥

The desired result on the structure of the Lagrangian control system(9.3.20) is the following.

Theorem 9.4.2. Consider the Lagrangian control system (9.3.18). LetU×V ⊂ Rn×Rm, and suppose that F : U×V → Rn is an input-dependentchange of coordinates q = F (q, x) such that for each x the metric tensorMexpressed in q−coordinates has δij as its ij−th entry. Suppose, moreover,that in the q−coordinates all cross coupling terms A(q, x) vanish. Thenthe system is flat and has flat inputs, which is to say both the hatted andunhatted Riemann symbols of the second kind,

R γj αβ

=∂Γγαj∂qβ

−∂Γγβj∂qα

+n∑k=1

(ΓkαjΓγβk − ΓkβjΓ

γαk)

α, β, j, γ = 1, . . . , n,

and

Rγ`ij =∂Γγ`j∂xi

−∂Γγij∂x`

+n∑k=1

(Γk`jΓγik − ΓkijΓ

γ`k)

i, ` = 1, . . . ,m; j, γ = 1, . . . , n,

vanish.

The proof of this theorem uses the following lemma.

Lemma 9.4.3. Let U × V ⊂ Rm × Rn be some neighborhood of (0, 0,and let fj : U × V → Rn be a smooth mapping for j = 1, . . . ,m. Givenq ∈ V , there is defined a neighborhood W of 0 in Rm and a smooth functiong : W → Rn such that

(i) g(0) = q, and

(ii) ∂g∂rj

(r) = fj(r, g(r)) for all r ∈W , j = 1, . . . ,m

9.4 Flat Systems and Systems with Flat Inputs 273

if and only if there is a neighborhood of (0, q) on which

∂fj∂ri− ∂fi∂rj

+n∑k=1

(∂fj∂qk

fki −∂fi∂qk

fkj ) = 0.

This lemma is proved in Spivak, 1970. We can now prove Theorem 9.4.2.

Proof of Theorem 9.4.2. Let q = F (q, x) be as in the hypothesis ofthe theorem. Then

M =∂F

∂q

T ∂F

∂q,

and we may write the kinetic energy in terms of (q, x)−coordinates as

12qTM(q, x)q =

12‖ ˙q‖2 + ˙qT

∂G

∂q

T

AT x− ˙qT∂F

∂xx− 1

2xT

∂F

∂x

T ∂F

∂xx,

where q = G(q, x) is the inverse of the x−dependent diffeomorphism q =F (q, x), and all terms are written in terms of the variables q and x. If thecross coupling terms vanish as in the hypothesis of the theorem, then

A(G(q, x), x)∂G

∂q(G(q, x), x) ≡ ∂F

∂x

T

(G(q, x), x),

which is equivalent to writing

A(q, x) =∂F

∂x

T ∂F

∂q(9.4.2)

in terms of the original coordinates. The componentwise rendering of thisequation is

a`i =n∑k=1

∂Fk∂x`

∂Fk∂qi

` = 1, . . . ,mi = 1, . . . , n.

Similarly, under the hypothesis of the theorem,

mij =n∑k=1

∂Fk∂qi

∂Fk∂qj

i, j = 1 . . . , n.

It is a straightforward calculation to show that

∂mkj

∂x`+∂a`k∂qj

− ∂a`j∂qk

= 2∑σ

∂2Fσ∂qj∂x`

∂Fσ∂qk

.

. Recall that we have called this quantity Γ`jk and in terms of this defined

Γµ`j =n∑k=1

MµkΓ`jk

274 9. Energy Based Methods for Stabilization

where mµk is the µk−entry inM−1. Using these definitions, one can showthat

∂x`

∂F1∂qj...

∂Fn∂qj

=n∑k=1

Γk`j

∂F1∂qk...

∂Fn∂qk

.

This is a partial differential equation, condidtions for the solution of whichare covered by Lemma 9.4.3. Specifically, we are here interested in condi-tions under which there exisits a fucntion g : Rm → Rn satisfying

∂g

∂x`= f`(x, g(x)) (9.4.3)

where

f j` (x, z) =n∑γ=1

Γγ`jzγ .

According to Lemma 9.4.3, a necessary condition for the solution of (9.4.3)is that

∂f`∂xi− ∂fi∂x`

+n∑k=1

(∂f`∂zk

fki −∂fi∂zk

fk` ) = 0.

This is rendered (componentwise)

n∑γ=1

(∂Γγ`j∂xi

−∂Γγij∂x`

)zγ +

n∑kγ=1

(Γk`jΓ

γikzγ − ΓkijΓ

γ`kzγ

)= 0.

Since this must hold identically in z = (z1, . . . , zn), we have the desiredconclusion that

∂Γγ`j∂xi

−∂Γγij∂x`

+n∑k=1

(Γk`jΓ

γik − ΓkijΓ

γ`k

)= 0.

¥

The theorem shows that in order for there to be a choice of coordi-nates, q such that the coefficients a`k of v` in (9.3.19) all vanish for k =1, . . . , n; ` = 1, . . . ,m, the geometric condition that the Riemann symbolsRγj αβ R

γj `i vanish must be satisfied. To understand the physical meaning

of these vanishings, we consider some prototypical, beginning with rotatingheavy chains. For the present discussion, the salient features are illustratedby the two rotating pendulum systems in Figure 9.4.1. Both pendula un-dergo controlled rotations about the vertical axis. In Figure 9.4.1(a), thejoint by which the pendulum is suspended is a single degree-of-freedom

9.4 Flat Systems and Systems with Flat Inputs 275

(a) (b)

φ φ ψ

θθ

Figure 9.4.1. (a) Rotating planar pendulum; (b) Rotating universal joint pen-dulum.

revolute joint, while in Figure 9.4.1(b), the link is suspended by a (twodegree-of-freedom) universal joint. The respective Lagrangians are

La(φ, φ; θ) =m

2(a2 cos2 φ+ `2 sin2 φ)θ2 + (`2 +

a2

2)φ2 +mg` cosφ,

and

Lb(φ, ψ, φ, ψ; θ) = (l2 + a2

2 (1 + sin2 ψ)) φ2 + (a2

2 + l2) ψ2

+ (a2

2 − l2) cosφ sin(2ψ) φ θ + (a2 + 2 l2) sinφ ψ θ+

(a2 cos2 φ cos2 ψ + (a

2

2 + l2) (sin2 φ + cosφ sin2 ψ))θ2

+ mg` cosφ cosψ,

and the dynamics of the controlled equations are

ddt

[m2 (a2 cos2 φ+ `2 sin2 φ)θ

]= u

(`2 + a2

2 )φ+m (a2 − `2) sinφ cosφ θ2 +mg` sinφ = 0,(9.4.4)

and (in abbreviated form)d

dt

∂Lb

∂θ= u,

d

dt

∂Lb

∂φ− ∂Lb

∂φ= 0,

d

dt

∂Lb

∂ψ− ∂Lb

∂ψ= 0. (9.4.5)

276 9. Energy Based Methods for Stabilization

In both cases a torque u is applied to control the angular velocity θ of rota-tion about the vertical axis. (See Baillieul [1987] and Marsden and Scheurle[1993] for more information about rotating heavy chains.) Applying our for-mal reduction procedure, (9.3.18) is rendered respectively

La(φ, φ; θ) =m

2(`2 +

a2

2)φ2 − Va(φ; θ) (9.4.6)

whereVa(φ; θ) = −(a2 cos2 φ+ `2 sin2 φ)θ2 −mg` cosφ,

and

Lb(φ, ψ, φ, ψ; θ) =12

(φ, ψ)Mb

ψ

)+ θAb

ψ

)− Vb(φ, ψ; θ)

(9.4.7)

where

Mb(φ, ψ) =

(l2 +

a2 (1+sinψ2)2 0

0 a2

2 + l2

),

Ab(φ, ψ) =

(a2

2 − l2)

cosφ cosψ sinψ(a2

2 + l2)

sinφ

,

and

Vb(φ, ψ; θ) = − 12 (a2 cosφ2 cosψ2 +

(a2

2 + l2) (

sinφ2 + cosφ sinψ2)) θ2

−mg` cosφ cosψ.

The significant difference between the two rotating pendulum systemsis the absence or presence of the coupling terms A in (9.4.6) and (9.4.7)respectively. The dynamics of the rotating planar pendulum (9.4.4) feelthe influence of rotation only through velocity terms involving θ2. Thedynamics of the universal joint pendulum (9.4.5), however, depend on boththe angular velocity θ and angular acceleration θ. A natural question whicharises is whether there is a change of coordinates which makes the dynamics(9.4.5) dependent only on θ and not on the acceleration θ. The answer tothis is negative—meaning that (9.4.5) is truly an acceleration controlledsystem. This may be seen in terms of the system curvatures discussed above.First, we examine the inertia tensor for the unreduced rotating universaljoint system:

a2 cos2 φ cos2 ψ+(a2

2 + l2) (

sin2 φ+ cosφsin2 ψ) (

a2

2 − l2)

cosφ cosψ sinψ(a2

2 + l2)

sinφ(a2

2 − l2)

cosφ cosψ sinψ l2 +a2 (1+sin2 ψ)

2 0(a2

2 + l2)

sinφ 0 a2

2 + l2

.

9.5 Averaging Lagrangian and Hamiltonian Systems with Oscillatory Inputs 277

It is a straightforward (and tedious, if you don’t have computer algebrasoftware) calculation to show that not all the Riemann symbols which definethe curvature tensor are zero. Indeed...

Interestingly, we obtain . . .DETAILS have to be added regarding thedifferences between the planar and nonplanar pendulum systems andhow these are related to CURVATURE. quite a bit

more stuffneeded here.jb

9.5 Averaging Lagrangian and HamiltonianSystems with Oscillatory Inputs

For flat systems with flat inputs, we may express the Lagrangian as thesum of a simple kinetic energy term plus a time-dependent potential. Itis natural to conjecture that given small-amplitude high-frequency peri-odic inputs x(·), their influence is approximately equal to the effect of apotential force determined by the time-averaged potential. The theory wedevelop next shows this to be the case. Even when the system is not flatand depends unavoidably on second derivatives of the inputs, however, itremains possible to write down an energy-like quantity called the aver-aged potential around whose critical points the motions of the system withoscillatory forcing is organized.

As a general starting point, we consider Lagrangian control system (9.3.18)which we rewrite here:

L(q, q;x, v) =12qTM(q, x)q + vTA(q, x)q − V(q;x, v).

As above, we assume x(·) and v(·) are inputs with v = x. The systemdynamics are obtained from the corresponding Euler-Lagrange equation

d

dt

∂L∂q− ∂L∂q

= 0.

The associated Hamiltonian is

H(q, p;x, v) =12

(p−AT v)TM−1(p−MT v) + V (9.5.1)

We expand equation (9.5.1) and apply simple averaging to yield theaveraged Hamiltonian:

H(q, p) =12pTM−1p− vTAM−1p+

12vTAM−1AT v + V. (9.5.2)

Here the overbars indicate simple averages over one period (T ) have beentaken: given any piecewise continuous function F (q, p, t), the simple averageis given by F =

∫ T0F (q, p, t) dt, where for the purpose of evaluating this

integral q and p are regarded as constants.

278 9. Energy Based Methods for Stabilization

For this averaged Hamiltonian, there is an obvious decomposition intokinematic and potential energy terms in the case that vTAM−1 = 0:

H(q, p) =12pTM−1p︸ ︷︷ ︸+

12vTAM−1Av + V︸ ︷︷ ︸ . (9.5.3)

avg. kin. averaged potentialenergy

In the case that vTAM−1 6= 0, it remains possible to decompose the av-eraged Hamiltonian (9.5.2) into the sum of averaged kinetic and potentialenergies, although the description becomes more involved. If v 6= 0, thenthe corresponding input variable x(t) will not be not periodic. There willbe a “drift” in the value of x(t) which changes by an amount v · T everyT -units of time. We may nevertheless study averaged Hamiltonian systemsin this context. We rewrite the averaged Hamiltonian (9.5.2) as

H(q, p) =12pTM−1p− vTAM−1p+

12vTAM−1

(M−1

)−1

M−1AT v

+12vTAM−1Av − 1

2vTAM−1

(M−1

)−1

M−1AT v + V

=12

(M−1p−M−1AT v)T(M−1

)−1

(M−1p−M−1AT v)︸ ︷︷ ︸ (9.5.4)

averaged kinetic energy

+12vTAM−1AT v − 1

2vTAM−1

(M−1

)−1

M−1AT v + V︸ ︷︷ ︸ .averaged potential

The formal distinction which appears between the zero-mean and non-zero-mean cases ((9.5.4) and (9.5.5) respectively) is that in the latter caseboth the averaged kinetic and potential energy terms are adjusted to re-flect the net (average) motions of the input variables (x(·), v(·)). There areimportant relationships which can be established between the dynamicsassociated with the averaged Hamltonian (9.5.5) and the dynamics of theperiodically forced system. The reader is referred to Weibel, 1997, Weibelet al., 1997, and Weibel & Baillieul, 1998 for details. It is important to men-tion that for mechanical systems in which v 6= 0 andM depends explicitlyon x in (9.5.1), the averaging analysis of this chapter may not provide anadequate description of the dynamics. Indeed, in this case, ‖x(t)‖ will notremain bounded as t→∞, and ifM(x(t), q2) also fails to remain bounded,the averaged potential will inherit a dependence on time which will makeit difficult to apply the critical point analysis proposed below. Despite thiscautionary remark, we shall indicate how our methods may be applied inmany instances where v 6= 0.

9.6 Stability of Mechanical Systems with Oscillatory Inputs 279

9.6 Stability of Mechanical Systems withOscillatory Inputs

The Lagrangian (9.3.18) gives rise to the Lagrangian dynamics

n∑j=1

mkj qj +m∑`=1

a`kv` +n∑

i,j=1

Γijk qiqj +n∑j=1

m∑`=1

Γ`jkv`qj = F (t), (k = 1, . . . , n),

where Γijk and Γ`jk are defined as in Section 6.4.2 in terms of the entries aijand mij in the m×n and n×n matrices A(q, x) andM(q, x) respectively.To retain our general perspective, we continue to assume that the termsin these equations may depend on x. The explicit form of this dependencewill play no role in the present section, however, and hence we simplifyour notation by omitting any further mention of the variable x. We seek tounderstand the stability of (9.3.19) in terms of the corresponding averagedpotential

VA(q) =12vTAM−1AT v − 1

2vTAM−1

(M−1

)−1

M−1AT v + V. (9.6.1)

The averaging principle of Section 6.x.x states that the effect of forcing (Need to writestuff about theaveraging prin-ciple.)

(9.3.19) with an oscillatory input v(·) will be to produce stable motionsconfined to neighborhoods of relative minima of VA(·). While this princi-ple appears to govern the dynamics encountered in both simulation andexperiments, there is as yet no complete theory describing the observedbehavior. Results reported in Baillieul, 1995, [?] showed that for a certainclass of systems (9.3.19) within a larger class of so-called linear Lagrangiansystems, strict local minima of the averaged potential are Lyapunov stablerest points of (9.3.19) for all periodic forcing of a given amplitude and suf-ficiently high frequency. In the present section, we shall review this result,and show that the extension of our analysis to arbitrary systems (9.3.19)is complicated by the fact that in general, a linearization of (9.3.19) failsto capture the stabilizing effects implied by an analysis of the averagedpotential. Indeed, we shall show that the averaged potential depends onsecond order jets of the coefficient functions A(q) and M(q). To simplifythe presentation, we shall restrict our attention to the case of zero-meanoscillatory forcing in which M−1AT v = 0.

Suppose q0 is a strict local minimum of (9.6.1). Applying a high-frequencyoscillatory input v(·), we shall look for stable motions of (9.3.19) in neigh-borhoods of (q, q) = (q0, 0). Of course, even when there are such stablemotions, (q0, 0) need not be a rest point of (9.3.19) for any choice of forc-ing function v(·). (Cf. the “cart-pendulum” dynamics of Example XX (Seehow much of Steve’s stuff to include.) in the case α 6= ±π/2.)

To analyze this relationship between (9.3.19) and (9.6.1) in more detail,let us assume q0 = 0. This assumption is made without loss of generality,

280 9. Energy Based Methods for Stabilization

since we may always change coordinates to make it true. Write

A(q) = A0 +A1(q) +A2(q) + h.o.t.;

M(q) =M0 +M1(q) +M2(q) + h.o.t.,

where the entries in the n×n matrixMk(q) are homogeneous polynomialsof degree k in the components of the vector q, and similarly for the m× nmatrix Ak(q). It is easy to show that

M−1(q) = M−10 −M−1

0 M1(q)M−10 +M−1

0 M1(q)M−10 M1(q)M−1

0

−M−10 M2(q)M−1

0 + h.o.t.

Using this, we write an expansion of VA up through terms of order 2:

VA(q) = V0 + V1(q) + V2(q) + h.o.t.,

where as above, Vk(q) denotes the sum of terms which are homogeneouspolynomials of degree k in the components of the vector q, and “h.o.t.”refers to a quantity which is of order o(‖q‖2). Explicitly, under our as-sumption that M−1AT v = 0:

V0 =12vTA0M−1

0 AT0 v + V0,

V1(q) =12vT (A1(q)M−1

0 AT0 −A0M−10 M1(q)M−1

0 AT0 +A0M−10 AT1 (q))v+V1·q,

and

V2(q) =12vTA1(q)M−1

0 AT1 (q)v − 12vTA1(q)M−1

0 M1(q)M−10 AT0 v

−12vTA0M−1

0 M1(q)M−10 AT1 (q)v

+12vTA2(q)M−1

0 AT0 v +12vTA0M−1

0 M1(q)M−10 M1(q)M−1

0 AT0 v

−12vTA0M−1

0 M2(q)M−10 AT0 v +

12vTA0M−1

0 AT2 (q)v + qTV2q.

where V0, V1, and V2 define the jets of the potential V(q) of orders 0,1, and2 respectively. Writing VA(·) in this way shows its dependence on jets ofcoefficient functions of (9.3.19) of order up to 2. This dependence impliesthat the observed stabilizing effects produced by high-frequency forcingcannot be understood in terms of a linearization of the dynamics (9.3.19).We shall examine this remark in a bit greater detail.

Having assumed that q0 = 0 is a strict local minimum of VA(·), it followsthat ∂VA

∂q (0) = ∂V1∂q = 0. There are two cases to consider here: i) ∂V1

∂q = 0

for a particular choice of oscillatory input v(·), and ii) ∂V1∂q ≡ 0 independent

of the choice of zero-mean oscillatory (periodic) forcing v(·).

Stability with Oscillatory Inputs 281

Case i): In this case, the location of the local minimum of VA(·) dependson v(·), and it will not generally coincide with a rest point of (9.3.19).While the averaging principle suggests that there will be stable motionsof (9.3.19) in neighborhoods of local minima of VA, the analysis of thiscase has involved either the introduction of dissipation into the model (aswas done in Baillieul [1993]) or the use of machinery from the theory ofdynamical systems (as was done in the thesis of S. Weibel [1997]). Thesignificant point to note here is that the dependence of VA on v(·) impliesthat A0 6= 0, which in turn implies both that q0 = 0 will not be a restpoint of (9.3.19) for any oscillatory input v(·), and also that the stabilizingeffect of v(·) on motions of (9.3.19) will depend on jets of order up to 2 inthe coefficient functions. We shall defer further discussion of this case andpresent some results on the qualitative theory of such systems in the nextsection. For further details the reader is referred to Weibel [1997].

Case ii): If ∂V1∂q ≡ 0, by which we mean that the first partial deriva-

tives of VA evaluated at q0 = 0 are zero independent of coefficients due tov(·), then we also find that V1 = 0. It then follows that 2A0M−1

0 AT1 (q) −A0M−1

0 M1(q)M−10 AT0 ≡ 0, and either of two sub-cases can occur: a)

A0 = 0, or else b) there is a polynomial relationship among the coeffi-cients in the 0-th and 1-st order jets of A(q) and M(q). In case b), q0 = 0will correspond to a rest point of (9.3.19) in the absence of forcing, butit will not generally define a rest point when v(t) 6= 0. Stabilizing effectsof the oscillatory input v(·) appear to again depend on jets of order up totwo in the coefficients of the Lagrangian vector field (9.3.19). This case willnot be treated further in this section. We shall consider case a), A0 = 0.In this case, the averaged potential VA only depends on terms up throughfirst order in the coefficients of (9.3.19).

Slightly refining our notation, let A`(q) denote the `-th column of then×m matrix AT (q). Then we have

A`(q) = A`1 · q + (terms of order ≥ 2), and

M(q) =M0 + (terms of order ≥ 1),

where we interpret M0,A11, . . . ,Am1 as n× n coefficient matrices. The fol-

lowing result is now clear.

Proposition 9.6.1. Suppose v(·) is an Rm-valued piecewise continuousperiodic function of period T > 0 such that v = 1

T

∫ T0v(s) ds = 0. Suppose,

moreover, that A0 = 0. Then the averaged potential of the Lagrangiansystem (9.3.19) agrees up to terms of order 2 with the averaged potentialassociated with the linear Lagrangian system

M0q +m∑`=1

(v`A`1q + v`(A`1 −A`1

T)q)

+ V2 · q = 0. (9.6.2)

282 9. Energy Based Methods for Stabilization

Proof. The proof follows immediately from examining the above expan-sion of VA. ¥

A deeper connection with stability is now expressed in terms of the fol-lowing theorem.

Theorem 9.6.2. Suppose w(·) is an Rm-valued piecewise continuous pe-riodic function of period T > 0 such that w = 1

T

∫ T0w(s) ds = 0. Consider

the linear Lagrangian system (9.6.2) with input v(t) = w(ωt), and supposeA`1

T = A`1 for ` = 1, . . . ,m. The averaged potential for this system is givenby

VA(q) =12qT(V2 +

m∑i,j=1

σijAi1M−10 Aj1

T)q, (9.6.3)

where σij = (1/T )∫ T

0wi(s)wj(s) ds. If the matrix ∂2VA

∂q2 is positive definite,the origin (q, q) = (0, 0) of the phase space is stable in the sense of Lyapunovprovided ω is sufficiently large.

This theorem has been proved in Baillieul [1995]. In Baillieul [1993], itwas shown in the presence of dissipation, the positive definiteness of theHessian matrix ∂2VA

∂q2 (0) is a sufficient condition for (9.3.19) to execute sta-ble motions in a neighborhood of (q0, 0). Theorem 3 shows that in Caseiib), it is precisely the conditions of the averaging principle from which wemay infer the Lyapunov stability of (9.6.2) based on the positive definite-ness of the Hessian of the averaged potential. Clearly this result is specialand related to the property that the averaged potential depends only onfirst order jets of the coefficients of (9.3.19) when A(q0) = 0. In Weibeland Baillieul [1998] it has been shown that the condition A(q0) = 0 is alsonecessary and sufficient for the local minimum q0 of the averaged potentialto define a corresponding fixed point (rather than a periodic orbit) of theforced (nonautonomous) Hamiltonian system associated with (9.3.18).

9.7 Stabilization of Rotating Systems

9.7.1 Heavy chains

9.7.2 Body-beam systems

This is page 283Printer: Opaque this

References

[1] V.I. Arnold, Dynamical Systems III, Vol.3, Springer-Verlag, 1988.

[2] A.M. Bloch,“Stabilizability of Nonholonomic Control Systems,” Pro-ceedings of IFAC Symposium on Nonlinear Control Systems Design,Capri, Italy, 1989.

[3] A.M. Bloch and N.H. McClamroch,“Control of Mechanical SystemswithClassical Nonholonomic Constraints,” Proceedings of IEEE Confer-ence on Decision and Control, 1989, Tampa, Fla., 201-205.

[4] A.M. Bloch, N.H. McClamroch and M. Reyhanoglu,“Controllabilityand Stabilizability Properties of a Nonholonomic Control System,”Proceedings of IEEE Conference on Decision and Control, 1990, Hon-olulu, Hawaii, 1312-1314.

[5] A.M. Bloch, M. Reyhanoglu and N.H.McClamroch,“Control and Sta-bilization of Nonholonomic Caplygin Dynamic Systems,” Proceedingsof IEEE Conference on Decision and Control, 1991, Brighton, UnitedKingdom, 1127-1132.

[6] R.W. Brockett,“Control Theory and Singular Riemannian Geometry”,in New Directions in Applied Mathematics, ed. P.J. Hilton and G.S.Young, Springer-Verlag, 1982.

[7] R.W. Brockett,“Asymptotic Stability and Feedback Stabilization,” inDifferential Geometric Control Theory, R.W. Brockett, R.S. Millmanand H.J. Sussmann, eds., Birkhauser, 1983.

284 References

[8] C.I. Byrnes and A. Isidori,“Local Stabilization of Minimum-PhaseNonlinear Systems,” Systems and Control Letters, 11, 1988, 9-17.

[9] H. Hermes, “On the Synthesis of a Stabilizing Feedback Control viaLie algebraic Methods”, SIAM Journal on Control and Optimization,Vol. 18, No. 4, 1980, 352-361.

[10] P.S. Krishnaprasad,“Eulerian Many-Body Problems,” in Dynamicsand Control of Multibody Systems, eds. J.E. Marsden, P.S. Krish-naprasad and J. Simo, Contemporary Math. Vol.97, AMS, 1989.

[11] P.S. Krishnaprasad,“Geometric Phases and Optimal Reconfigurationfor Multibody Systems,” Proceedings of American Control Conference,1990, San Diego, 2440-2444.

[12] P.S. Krishnaprasad and R. Yang,“Geometric Phases, Anholonomy andOptimal Movement,” Proceedings of Conference on Robotics and Au-tomation, 1991, Sacramento, California, 2185-2189.

[13] G. Laffarriere and H.J. Sussmann,“Motion Planning for CompletelyControllable Systems Without Drift,” Proceedings of Conference onRobotics and Automation, 1991, Sacramento, California, 1148-1153.

[14] Z. Li and J. Canny,“Motion of Two Rigid Bodies with Rolling Con-straint,” IEEE Transactions on Robotics and Automation, Vol. 6, No.1, 1990, 62-71.

[15] Z. Li and R. Montgomery,“Dynamics and Optimal Control of a LeggedRobot in Flight Phase,” Proceedings of Conference on Robotics andAutomation, 1990, Cincinnati, Ohio, 1816-1821.

[16] J. Marsden, R. Montgomery and T. Ratiu, Reduction, Symmetry andBerry’s Phase in Mechanics, Memoirs of the A.M.S., 1990.

[17] N.H. McClamroch and A.M. Bloch,“Control of Constrained Hamil-tonian Systems and Application to Control of Constrained Robots”,in Dynamical Systems Approaches to Nonlinear Problems in Systemsand Control, ed. F.M.A. Salam and M.L. Levi, SIAM, 1988, 394-403.

[18] N.H. McClamroch and D. Wang,“Feedback Stabilization and Trackingof Constrained Robots,” IEEE Transactions on Automatic Control, 33,1988, 419-426.

[19] R. Montgomery,“Isoholonomic Problems and Some Applications,”Communications in Mathematical Physics, Vol. 128, 1990, 565-592.

[20] R. Murray and S. Sastry,“Steering Nonholonomic Systems Using Si-nusoids,” Proceedings of IEEE Conference on Desicion and Control,1990, Honolulu, Hawai, 2097-2101.

References 285

[21] Ju. I. Neimark and F.A. Fufaev, Dynamics of Nonholonomic Systems,Vol. 33, A.M.S. Translations of Mathematical Monographs, A.M.S.,1972.

[22] M. Reyhanoglu and N. H. McClamroch,“Controllability and Stabi-lizability of Planar Multibody Systems with Angular Momentum Pre-serving Control Torques,” 1991 American Control Conference, Boston,1102-1107.

[23] M. Reyhanoglu and N. H. McClamroch,“Reorientation of Space Multi-body Systems Maintaining Zero Angular Momentum,” 1991 AIAAGuidance, Navigation and Control Conference, New Orleans, 1330-1340.

[24] A. Shapere and F. Wilczek, Geometric Phases in Physics, World Sci-entific, 1988.

[25] N. Sreenath,“Nonlinear Control of Multibody Systems in ShapeSpace,” Proceedings of IEEE Conference on Robotics and Automa-tion, 1990, Cincinnati, Ohio, 1776-1781.

[26] H.J. Sussmann, “Subanalytic Sets and Feedback Control,” J. Differ-ential Equations, 31, 1979, 31-52.

[27] H.J. Sussmann,“A General Theorem on Local Controllability,” SIAMJournal on Control and Optimization, Vol. 25, No. 1, 1987, 158-194.

[28] H.J. Sussmann and V. Jurdjevic, “Controllability of Nonlinear Sys-tems,” J. Differential Equations, 12, 1972, 95-116.

[29] Z. Vafa and S. Dubowsky,“On the Dynamics of Space Manipulators inSpace Using the Virtual Manipulator Approach,” Proceedings of IEEEInternational Conference on Robotics and Automation, 1987, Raleigh,N.C., 579-585.

[30] A. M. Vershik and V. Ya. Gershkovich,“ Nonholonomic Problems andthe Theory of Distributions,” Acta Applicandae Mathematicae, 12,1988, 181-209.

[31] J. Zabczyk,“Some Comments on Stabilizability,” Applied Math. Op-timization, 19, 1989, 1-9.