Classical Mechanics - Shapiro

download Classical Mechanics - Shapiro

of 252

Transcript of Classical Mechanics - Shapiro

  • Classical Mechanics

    Joel A. Shapiro

    April 21, 2003

  • iCopyright C 1994, 1997 by Joel A. ShapiroAll rights reserved. No part of this publication may be reproduced,stored in a retrieval system, or transmitted in any form or by anymeans, electronic, mechanical, photocopying, or otherwise, without theprior written permission of the author.

    This is a preliminary version of the book, not to be considered afully published edition. While some of the material, particularly therst four chapters, is close to readiness for a rst edition, chapters 6and 7 need more work, and chapter 8 is incomplete. The appendicesare random selections not yet reorganized. There are also as yet fewexercises for the later chapters. The rst edition will have an adequateset of exercises for each chapter.

    The author welcomes corrections, comments, and criticism.

  • ii

  • Contents

    1 Particle Kinematics 11.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Single Particle Kinematics . . . . . . . . . . . . . . . . . 4

    1.2.1 Motion in conguration space . . . . . . . . . . . 41.2.2 Conserved Quantities . . . . . . . . . . . . . . . . 6

    1.3 Systems of Particles . . . . . . . . . . . . . . . . . . . . . 91.3.1 External and internal forces . . . . . . . . . . . . 101.3.2 Constraints . . . . . . . . . . . . . . . . . . . . . 141.3.3 Generalized Coordinates for Unconstrained Sys-

    tems . . . . . . . . . . . . . . . . . . . . . . . . . 171.3.4 Kinetic energy in generalized coordinates . . . . . 19

    1.4 Phase Space . . . . . . . . . . . . . . . . . . . . . . . . . 211.4.1 Dynamical Systems . . . . . . . . . . . . . . . . . 221.4.2 Phase Space Flows . . . . . . . . . . . . . . . . . 27

    2 Lagranges and Hamiltons Equations 372.1 Lagrangian Mechanics . . . . . . . . . . . . . . . . . . . 37

    2.1.1 Derivation for unconstrained systems . . . . . . . 382.1.2 Lagrangian for Constrained Systems . . . . . . . 412.1.3 Hamiltons Principle . . . . . . . . . . . . . . . . 462.1.4 Examples of functional variation . . . . . . . . . . 482.1.5 Conserved Quantities . . . . . . . . . . . . . . . . 502.1.6 Hamiltons Equations . . . . . . . . . . . . . . . . 532.1.7 Velocity-dependent forces . . . . . . . . . . . . . 55

    3 Two Body Central Forces 653.1 Reduction to a one dimensional problem . . . . . . . . . 65

    iii

  • iv CONTENTS

    3.1.1 Reduction to a one-body problem . . . . . . . . . 663.1.2 Reduction to one dimension . . . . . . . . . . . . 67

    3.2 Integrating the motion . . . . . . . . . . . . . . . . . . . 693.2.1 The Kepler problem . . . . . . . . . . . . . . . . 703.2.2 Nearly Circular Orbits . . . . . . . . . . . . . . . 74

    3.3 The Laplace-Runge-Lenz Vector . . . . . . . . . . . . . . 773.4 The virial theorem . . . . . . . . . . . . . . . . . . . . . 783.5 Rutherford Scattering . . . . . . . . . . . . . . . . . . . . 79

    4 Rigid Body Motion 854.1 Conguration space for a rigid body . . . . . . . . . . . . 85

    4.1.1 Orthogonal Transformations . . . . . . . . . . . . 874.1.2 Groups . . . . . . . . . . . . . . . . . . . . . . . . 91

    4.2 Kinematics in a rotating coordinate system . . . . . . . . 944.3 The moment of inertia tensor . . . . . . . . . . . . . . . 98

    4.3.1 Motion about a xed point . . . . . . . . . . . . . 984.3.2 More General Motion . . . . . . . . . . . . . . . . 100

    4.4 Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 1074.4.1 Eulers Equations . . . . . . . . . . . . . . . . . . 1074.4.2 Euler angles . . . . . . . . . . . . . . . . . . . . . 1134.4.3 The symmetric top . . . . . . . . . . . . . . . . . 117

    5 Small Oscillations 1275.1 Small oscillations about stable equilibrium . . . . . . . . 127

    5.1.1 Molecular Vibrations . . . . . . . . . . . . . . . . 1305.1.2 An Alternative Approach . . . . . . . . . . . . . . 137

    5.2 Other interactions . . . . . . . . . . . . . . . . . . . . . . 1375.3 String dynamics . . . . . . . . . . . . . . . . . . . . . . . 1385.4 Field theory . . . . . . . . . . . . . . . . . . . . . . . . . 143

    6 Hamiltons Equations 1476.1 Legendre transforms . . . . . . . . . . . . . . . . . . . . 1476.2 Variations on phase curves . . . . . . . . . . . . . . . . . 1526.3 Canonical transformations . . . . . . . . . . . . . . . . . 1536.4 Poisson Brackets . . . . . . . . . . . . . . . . . . . . . . 1556.5 Higher Dierential Forms . . . . . . . . . . . . . . . . . . 1606.6 The natural symplectic 2-form . . . . . . . . . . . . . . . 169

  • CONTENTS v

    6.6.1 Generating Functions . . . . . . . . . . . . . . . . 1726.7 Hamilton{Jacobi Theory . . . . . . . . . . . . . . . . . . 1816.8 Action-Angle Variables . . . . . . . . . . . . . . . . . . . 185

    7 Perturbation Theory 1897.1 Integrable systems . . . . . . . . . . . . . . . . . . . . . 1897.2 Canonical Perturbation Theory . . . . . . . . . . . . . . 194

    7.2.1 Time Dependent Perturbation Theory . . . . . . 1967.3 Adiabatic Invariants . . . . . . . . . . . . . . . . . . . . 198

    7.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . 1987.3.2 For a time-independent Hamiltonian . . . . . . . 1987.3.3 Slow time variation in H(q; p; t) . . . . . . . . . . 2007.3.4 Systems with Many Degrees of Freedom . . . . . 2067.3.5 Formal Perturbative Treatment . . . . . . . . . . 209

    7.4 Rapidly Varying Perturbations . . . . . . . . . . . . . . . 2117.5 New approach . . . . . . . . . . . . . . . . . . . . . . . . 216

    8 Field Theory 2198.1 Noethers Theorem . . . . . . . . . . . . . . . . . . . . . 225

    A ijk and cross products 229A.1 Vector Operations . . . . . . . . . . . . . . . . . . . . . . 229

    A.1.1 ij and ijk . . . . . . . . . . . . . . . . . . . . . . 229

    B The gradient operator 233

    C Gradient in Spherical Coordinates 237

  • vi CONTENTS

  • Chapter 1

    Particle Kinematics

    1.1 Introduction

    Classical mechanics, narrowly dened, is the investigation of the motionof systems of particles in Euclidean three-dimensional space, under theinfluence of specied force laws, with the motions evolution determinedby Newtons second law, a second order dierential equation. Thatis, given certain laws determining physical forces, and some boundaryconditions on the positions of the particles at some particular times, theproblem is to determine the positions of all the particles at all times.We will be discussing motions under specic fundamental laws of greatphysical importance, such as Coulombs law for the electrostatic forcebetween charged particles. We will also discuss laws which are lessfundamental, because the motion under them can be solved explicitly,allowing them to serve as very useful models for approximations to morecomplicated physical situations, or as a testbed for examining conceptsin an explicitly evaluatable situation. Techniques suitable for broadclasses of force laws will also be developed.

    The formalism of Newtonian classical mechanics, together with in-vestigations into the appropriate force laws, provided the basic frame-work for physics from the time of Newton until the beginning of thiscentury. The systems considered had a wide range of complexity. Onemight consider a single particle on which the Earths gravity acts. Butone could also consider systems as the limit of an innite number of

    1

  • 2 CHAPTER 1. PARTICLE KINEMATICS

    very small particles, with displacements smoothly varying in space,which gives rise to the continuum limit. One example of this is theconsideration of transverse waves on a stretched string, in which everypoint on the string has an associated degree of freedom, its transversedisplacement.

    The scope of classical mechanics was broadened in the 19th century,in order to consider electromagnetism. Here the degrees of freedomwere not just the positions in space of charged particles, but also otherquantities, distributed throughout space, such as the the electric eldat each point. This expansion in the type of degrees of freedom hascontinued, and now in fundamental physics one considers many degreesof freedom which correspond to no spatial motion, but one can stilldiscuss the classical mechanics of such systems.

    As a fundamental framework for physics, classical mechanics gaveway on several fronts to more sophisticated concepts in the early 1900s.Most dramatically, quantum mechanics has changed our focus from spe-cic solutions for the dynamical degrees of freedom as a function of timeto the wave function, which determines the probabilities that a systemhave particular values of these degrees of freedom. Special relativitynot only produced a variation of the Galilean invariance implicit inNewtons laws, but also is, at a fundamental level, at odds with thebasic ingredient of classical mechanics | that one particle can exerta force on another, depending only on their simultaneous but dierentpositions. Finally general relativity brought out the narrowness of theassumption that the coordinates of a particle are in a Euclidean space,indicating instead not only that on the largest scales these coordinatesdescribe a curved manifold rather than a flat space, but also that thisgeometry is itself a dynamical eld.

    Indeed, most of 20th century physics goes beyond classical Newto-nian mechanics in one way or another. As many readers of this bookexpect to become physicists working at the cutting edge of physics re-search, and therefore will need to go beyond classical mechanics, webegin with a few words of justication for investing eort in under-standing classical mechanics.

    First of all, classical mechanics is still very useful in itself, and notjust for engineers. Consider the problems (scientic | not political)that NASA faces if it wants to land a rocket on a planet. This requires

  • 1.1. INTRODUCTION 3

    an accuracy of predicting the position of both planet and rocket farbeyond what one gets assuming Keplers laws, which is the motion onepredicts by treating the planet as a point particle influenced only bythe Newtonian gravitational eld of the Sun, also treated as a pointparticle. NASA must consider other eects, and either demonstratethat they are ignorable or include them into the calculations. Theseinclude

    multipole moments of the sun

    forces due to other planets

    eects of corrections to Newtonian gravity due to general relativ-ity

    friction due to the solar wind and gas in the solar system

    Learning how to estimate or incorporate such eects is not trivial.

    Secondly, classical mechanics is not a dead eld of research | infact, in the last two decades there has been a great deal of interest in\dynamical systems". Attention has shifted from calculation of the or-bit over xed intervals of time to questions of the long-term stability ofthe motion. New ways of looking at dynamical behavior have emerged,such as chaos and fractal systems.

    Thirdly, the fundamental concepts of classical mechanics provide theconceptual framework of quantum mechanics. For example, althoughthe Hamiltonian and Lagrangian were developed as sophisticated tech-niques for performing classical mechanics calculations, they provide thebasic dynamical objects of quantum mechanics and quantum eld the-ory respectively. One view of classical mechanics is as a steepest pathapproximation to the path integral which describes quantum mechan-ics. This integral over paths is of a classical quantity depending on the\action" of the motion.

    So classical mechanics is worth learning well, and we might as welljump right in.

  • 4 CHAPTER 1. PARTICLE KINEMATICS

    1.2 Single Particle Kinematics

    We start with the simplest kind of system, a single unconstrained par-ticle, free to move in three dimensional space, under the influence of aforce ~F .

    1.2.1 Motion in conguration space

    The motion of the particle is described by a function which gives itsposition as a function of time. These positions are points in Euclideanspace. Euclidean space is similar to a vector space, except that thereis no special point which is xed as the origin. It does have a met-ric, that is, a notion of distance between any two points, D(A;B). Italso has the concept of a displacement AB from one point B in theEuclidean space to another, A. These displacements do form a vectorspace, and for a three-dimensional Euclidean space, the vectors forma three-dimensional real vector space R3, which can be given an or-thonormal basis such that the distance between A and B is given byD(A;B) =

    P3i=1[(AB)i]2. Because the mathematics of vector spaces

    is so useful, we often convert our Euclidean space to a vector spaceby choosing a particular point as the origin. Each particles positionis then equated to the displacement of that position from the origin,so that it is described by a position vector ~r relative to this origin.But the origin has no physical signicance unless it has been choosenin some physically meaningful way. In general the multiplication of aposition vector by a scalar is as meaningless physically as saying that42nd street is three times 14th street. The cartesian components ofthe vector ~r, with respect to some xed though arbitrary coordinatesystem, are called the coordinates, cartesian coordinates in this case.We shall nd that we often (even usually) prefer to change to other setsof coordinates, such as polar or spherical coordinates, but for the timebeing we stick to cartesian coordinates.

    The motion of the particle is the function ~r(t) of time. Certainlyone of the central questions of classical mechanics is to determine, giventhe physical properties of a system and some initial conditions, whatthe subsequent motion is. The required \physical properties" is a spec-ication of the force, ~F . The beginnings of modern classical mechanics

  • 1.2. SINGLE PARTICLE KINEMATICS 5

    was the realization at early in the 17th century that the physics, or dy-namics, enters into the motion (or kinematics) through the force and itseect on the acceleration, and not through any direct eect of dynamicson the position or velocity of the particle.

    Most likely the force will depend on the position of the particle, sayfor a particle in the gravitational eld of a xed (heavy) source at theorigin, for which

    ~F (~r) = GMmr3

    ~r: (1.1)

    But the force might also depend explicitly on time. For example, forthe motion of a spaceship near the Earth, we might assume that theforce is given by sum of the Newtonian gravitational forces of the Sun,Moon and Earth. Each of these forces depends on the positions of thecorresponding heavenly body, which varies with time. The assumptionhere is that the motion of these bodies is independent of the position ofthe light spaceship. We assume someone else has already performed thenontrivial problem of nding the positions of these bodies as functionsof time. Given that, we can write down the force the spaceship feels attime t if it happens to be at position ~r,

    ~F (~r; t) = GmMS ~r ~RS(t)

    jr RS(t)j3 GmME~r ~RE(t)jr RE(t)j3

    GmMM ~r ~RM(t)

    jr RM(t)j3 :

    Finally, the force might depend on the velocity of the particle, as forexample for the Lorentz force on a charged particle in electric andmagnetic elds

    ~F (~r; ~v; t) = q ~E(~r; t) + q ~v ~B(~r; t): (1.2)

    However the force is determined, it determines the motion of theparticle through the second order dierential equation known as New-tons Second Law

    ~F (~r; ~v; t) = m~a = md2~r

    dt2:

  • 6 CHAPTER 1. PARTICLE KINEMATICS

    As this is a second order dierential equation, the solution depends ingeneral on two arbitrary (3-vector) parameters, which we might chooseto be the initial position and velocity, ~r(0) and ~v(0).

    For a given physical situation and a given set of initial conditionsfor the particle, Newtons laws determine the motion ~r(t), which isa curve in conguration space parameterized by time t, known asthe trajectory in conguration space. If we consider the curve itself,independent of how it depends on time, this is called the orbit of theparticle. For example, the orbit of a planet, in the approximation thatit feels only the eld of a xed sun, is an ellipse. That word does notimply any information about the time dependence or parameterizationof the curve.

    1.2.2 Conserved Quantities

    While we tend to think of Newtonian mechanics as centered on New-tons Second Law in the form ~F = m~a, he actually started with theobservation that in the absence of a force, there was uniform motion.We would now say that under these circumstances the momentum~p(t) is conserved, d~p=dt = 0. In his second law, Newton stated theeect of a force as producing a rate of change of momentum, which wewould write as

    ~F = d~p=dt;

    rather than as producing an acceleration ~F = m~a. In focusing onthe concept of momentum, Newton emphasized one of the fundamen-tal quantities of physics, useful beyond Newtonian mechanics, in bothrelativity and quantum mechanics1. Only after using the classical rela-tion of momentum to velocity, ~p = m~v, and the assumption that m isconstant, do we nd the familiar ~F = m~a.

    One of the principal tools in understanding the motion of manysystems is isolating those quantities which do not change with time. Aconserved quantity is a function of the positions and momenta, andperhaps explicitly of time as well, Q(~r; ~p; t), which remains unchangedwhen evaluated along the actual motion, dQ(~r(t); ~p(t); t)=dt = 0. A

    1The relationship of momentum to velocity is changed in these extensions,however.

  • 1.2. SINGLE PARTICLE KINEMATICS 7

    function depending on the positions, momenta, and time is said to bea function on extended phase space2. When time is not included, thespace is called phase space. In this language, a conserved quantity is afunction on extended phase space with a vanishing total time derivativealong any path which describes the motion of the system.

    A single particle with no forces acting on it provides a very simpleexample. As Newton tells us, _~p = d~p=dt = ~F = 0, so the momentum

    is conserved. There are three more conserved quantities ~Q(~r; ~p; t) :=

    ~r(t)t~p(t)=m, which have a time rate of change d ~Q=dt = _~r~p=m t _~p=m =0. These six independent conserved quantities are as many as one couldhave for a system with a six dimensional phase space, and they com-pletely solve for the motion. Of course this was a very simple systemto solve. We now consider a particle under the influence of a force.

    Energy

    Consider a particle under the influence of an external force ~F . In gen-eral, the momentum will not be conserved, although if any cartesiancomponent of the force vanishes along the motion, that component ofthe momentum will be conserved. Also the kinetic energy, dened asT = 1

    2m~v 2, will not in general be conserved, because

    dT

    dt= m _~v ~v = ~F ~v:

    As the particle moves from the point ~ri to the point ~rf the total change

    in the kinetic energy is the work done by the force ~F ,

    T =Z ~rf~ri

    ~F d~r:

    If the force law ~F (~r; ~p; t) applicable to the particle is independent oftime and velocity, then the work done will not depend on how quicklythe particle moved along the path from ~ri to ~rf . If in addition thework done is independent of the path taken between these points, so itdepends only on the endpoints, then the force is called a conservative

    2Phase space is discussed further in section 1.4.

  • 8 CHAPTER 1. PARTICLE KINEMATICS

    force and we assosciate with it potential energy

    U(~r) = U(~r0) +Z ~r0~r

    ~F (~r 0) d~r 0;where ~r0 is some arbitrary reference position and U(~r0) is an arbitrarilychosen reference energy, which has no physical signicance in ordinarymechanics. U(~r) represents the potential the force has for doing workon the particle if the particle is at position ~r.

    The condition for the path inte-gral to be independent of the path isthat it gives the same results alongany two coterminous paths 1 and 2,or alternatively that it give zero whenevaluated along any closed path suchas = 1 2, the path consisting offollowing 1 and then taking 2 back-wards to the starting point. By StokesTheorem, this line integral is equiva-lent to an integral over any surface Sbounded by ,I

    ~F d~r =ZS

    ~r ~F dS:

    ri

    rf rf

    ri

    G

    G

    G 2

    1

    Independence of pathR1

    =R2

    is equivalent to vanishing of thepath integral over closed paths, which is in turn equivalentto the vanishing of the curl onthe surface whose boundary is.

    Thus the requirement that the integral of ~F d~r vanish around anyclosed path is equivalent to the requirement that the curl of ~F vanisheverywhere in space.

    By considering an innitesimal path from ~r to ~r + ~r, we see that

    U(~r + ~) U(~r) = ~F ~r; or~F (r) = ~rU(r):

    The value of the concept of potential energy is that it enables ndinga conserved quantity, the total energy, in situtations in which all forcesare conservative. Then the total energy E = T + U changes at a rate

    dE

    dt=dT

    dt+d~r

    dt ~rU = ~F ~v ~v ~F = 0:

  • 1.3. SYSTEMS OF PARTICLES 9

    The total energy can also be used in systems with both conservativeand nonconservative forces, giving a quantity whose rate of change isdetermined by the work done only by the nonconservative forces. Oneexample of this usefulness is in the discussion of a slightly dampedharmonic oscillator driven by a periodic force near resonance. Then theamplitude of steady-state motion is determined by a balence betweenthe average power input by the driving force and the average powerdissipated by friction, the two nonconservative forces in the problem,without needing to worry about the work done by the spring.

    Angular momentum

    Another quantity which is often useful because it may be conserved isthe angular momentum. The denition requires a reference point in theEuclidean space, say ~r0. Then a particle at position ~r with momentum~p has an angular momentum about ~r0 given by ~L = (~r ~r0) ~p.Very often we take the reference point ~r0 to be the same as the point wehave chosen as the origin in converting the Euclidian space to a vectorspace, so ~r0 = 0, and

    ~L = ~r ~pd~L

    dt=

    d~r

    dt ~p+ ~r d~p

    dt=

    1

    m~p ~p+ ~r ~F = 0 + ~ = ~ :

    where we have dened the torque about ~r0 as = (~r ~r0) ~F ingeneral, and = ~r ~F when our reference point ~r0 is at the origin.

    We see that if the torque ~(t) vanishes (at all times) the angularmomentum is conserved. This can happen not only if the force is zero,but also if the force always points to the reference point. This is thecase in a central force problem such as motion of a planet about thesun.

    1.3 Systems of Particles

    So far we have talked about a system consisting of only a single particle,possibly influenced by external forces. Consider now a system of nparticles with positions ~ri, i = 1; : : : ; n, in flat space. The conguration

  • 10 CHAPTER 1. PARTICLE KINEMATICS

    of the system then has 3n coordinates (conguration space is R3n), andthe phase space has 6n coordinates f~ri; ~pig.

    1.3.1 External and internal forces

    Let ~Fi be the total force acting on particle i. It is the sum of the forcesproduced by each of the other particles and that due to any externalforce. Let ~Fji be the force particle j exerts on particle i and let ~F

    Ei be

    the external force on particle i. Using Newtons second law on particlei, we have

    ~Fi = ~FEi +

    Xj

    ~Fji = _~pi = mi _~vi;

    where mi is the mass of the ith particle. Here we are assuming forceshave identiable causes, which is the real meaning of Newtons sec-ond law, and that the causes are either individual particles or externalforces. Thus we are assuming there are no \three-body" forces whichare not simply the sum of \two-body" forces that one object exerts onanother.

    Dene the center of mass and total mass

    ~R =

    Pmi~riPmi

    ; M =X

    mi:

    Then if we dene the total momentum

    ~P =X

    ~pi =X

    mi~vi =d

    dt

    Xmi~ri = M

    d~R

    dt;

    we haved~P

    dt=

    _~P =X

    _~pi =X

    ~Fi =Xi

    ~FEi +Xij

    ~Fji:

    Let us dene ~FE =Pi F

    Ei to be the total external force. If Newtons

    Third Law holds,

    ~Fji = ~Fij ; soXij

    ~Fij = 0; and

    _~P = ~FE: (1.3)

  • 1.3. SYSTEMS OF PARTICLES 11

    Thus the internal forces cancel in pairs in their eect on the total mo-mentum, which changes only in response to the total external force. Asan obvious but very important consequence3 the total momentum of anisolated system is conserved.

    The total angular momentum is also just a sum over the individualparticles, in this case of the individual angular momenta:

    ~L =X

    ~Li =X

    ~ri ~pi:

    Its rate of change with time is

    d~L

    dt=

    _~L =Xi

    ~vi ~pi +Xi

    ~ri ~Fi = 0 +X

    ~ri ~FEi +Xij

    ~ri ~Fji:

    The total external torque is naturally dened as

    ~ =Xi

    ~ri ~FEi ;

    3There are situations and ways of describing them in which the law of actionand reaction seems not to hold. For example, a current i1 flowing through a wiresegment d~s1 contributes, according to the law of Biot and Savart, a magnetic eldd ~B = 0i1d~s1 ~r=4jrj3 at a point ~r away from the current element. If a currenti2 flows through a segment of wire d~s2 at that point, it feels a force

    ~F12 =04i1i2

    d~s2 (d~s1 ~r)jrj3

    due to element 1. On the other hand ~F21 is given by the same expression with d~s1and d~s2 interchanged and the sign of ~r reversed, so

    ~F12 + ~F21 =04

    i1i2jrj3 [d~s1(d~s2 ~r) d~s2(d~s1 ~r)] ;

    which is not generally zero.One should not despair for the validity of momentum conservation. The Law

    of Biot and Savart only holds for time-independent current distributions. Unlessthe currents form closed loops, there will be a charge buildup and Coulomb forcesneed to be considered. If the loops are closed, the total momentum will involveintegrals over the two closed loops, for which

    R RF12 +F21 can be shown to vanish.

    More generally, even the sum of the momenta of the current elements is not thewhole story, because there is momentum in the electromagnetic eld, which will bechanging in the time-dependent situation.

  • 12 CHAPTER 1. PARTICLE KINEMATICS

    so we might ask if the last term vanishes due the Third Law, whichpermits us to rewrite ~Fji =

    12

    ~Fji ~Fij

    . Then the last term becomes

    Xij

    ~ri ~Fji = 12

    Xij

    ~ri ~Fji 12

    Xij

    ~ri ~Fij

    =1

    2

    Xij

    ~ri ~Fji 12

    Xij

    ~rj ~Fji

    =1

    2

    Xij

    (~ri ~rj) ~Fji:

    This is not automatically zero, but vanishes if one assumes a strongerform of the Third Law, namely that the action and reaction forces be-tween two particles acts along the line of separation of the particles.If the force law is independent of velocity and rotationally and trans-lationally symmetric, there is no other direction for it to point. Forspinning particles and magnetic forces the argument is not so simple| in fact electromagnetic forces between moving charged particles arereally only correctly viewed in a context in which the system includesnot only the particles but also the elds themselves. For such a system,in general the total energy, momentum, and angular momentum of theparticles alone will not be conserved, because the elds can carry allof these quantities. But properly dening the energy, momentum, andangular momentum of the electromagnetic elds, and including them inthe totals, will result in quantities conserved as a result of symmetriesof the underlying physics. This is further discussed in section 8.1.

    Making the assumption that the strong form of Newtons Third Lawholds, we have shown that

    ~ =d~L

    dt: (1.4)

    The conservation laws are very useful because they permit algebraicsolution for part of the velocity. Taking a single particle as an example,if E = 1

    2mv2 + U(~r) is conserved, the speed jv(t)j is determined at all

    times (as a function of ~r) by one arbitrary constant E. Similarly if~L is conserved, the components of ~v which are perpendicular to ~r aredetermined in terms of the xed constant ~L. With both conserved, ~v

  • 1.3. SYSTEMS OF PARTICLES 13

    is completely determined except for the sign of the radial component.Examples of the usefulness of conserved quantities are everywhere, andwill be particularly clear when we consider the two body central forceproblem later. But rst we continue our discussion of general systemsof particles.

    As we mentioned earlier, the total angular momentum depends onthe point of evaluation, that is, the origin of the coordinate systemused. We now show that it consists of two contributions, the angularmomentum about the center of mass and the angular momentum ofa ctitious point object located at the center of mass. Let ~r 0i be theposition of the ith particle with respect to the center of mass, so ~r 0i =~ri ~R. Then

    ~L =Xi

    mi~ri ~vi =Xi

    mi~r 0i + ~R

    _~r 0i +_~R

    =Xi

    mi~r0i _~r 0i +

    Xi

    mi~r0i _~R

    +~RXmi _~r 0i +M ~R _~R=

    Xi

    ~r 0i ~p 0i + ~R ~P :

    Here we have noted thatPmi~r

    0i = 0, and also its derivative

    Pmi~v

    0i =

    0. We have dened ~p 0i = mi~v0i, the momentum in the center of mass

    reference frame. The rst term of the nal form is the sum of theangular momenta of the particles about their center of mass, while thesecond term is the angular momentum the system would have if it werecollapsed to a point at the center of mass.

    What about the total energy? The kinetic energy

    T =1

    2

    Xmiv

    2i =

    1

    2

    Xmi~v 0i + ~V

    ~v 0i + ~V

    =

    1

    2

    Xmiv

    02i +

    1

    2MV 2; (1.5)

    where the cross term vanishes, once again, becausePmi~v

    0i = 0. Thus

    the kinetic energy of the system can also be viewed as the sum of thekinetic energies of the constituents about the center of mass, plus the

  • 14 CHAPTER 1. PARTICLE KINEMATICS

    kinetic energy the system would have if it were collapsed to a particleat the center of mass.

    If the forces on the system are due to potentials, the total energywill be conserved, but this includes not only the potential due to theexternal forces but also that due to interparticle forces,

    PUij(~ri; ~rj).

    In general this contribution will not be zero or even constant withtime, and the internal potential energy will need to be considered. Oneexception to this is the case of a rigid body.

    1.3.2 Constraints

    A rigid body is dened as a system of n particles for which all theinterparticle distances are constrained to xed constants, j~ri~rjj = cij ,and the interparticle potentials are functions only of these interparticledistances. As these distances do not vary, neither does the internalpotential energy. These interparticle forces cannot do work, and theinternal potential energy may be ignored.

    The rigid body is an example of a constrained system, in which thegeneral 3n degrees of freedom are restricted by some forces of constraintwhich place conditions on the coordinates ~ri, perhaps in conjunctionwith their momenta. In such descriptions we do not wish to consideror specify the forces themselves, but only their (approximate) eect.The forces are assumed to be whatever is necessary to have that ef-fect. It is generally assumed, as in the case with the rigid body, thatthe constraint forces do no work under displacements allowed by theconstraints. We will consider this point in more detail later.

    If the constraints can be phrased so that they are on the coordinatesand time only, as i(~r1; :::~rn; t) = 0; i = 1; : : : ; k, they are known asholonomic constraints. These constraints determine hypersurfacesin conguration space to which all motion of the system is conned.In general this hypersurface forms a 3n k dimensional manifold. Wemight describe the conguration point on this manifold in terms of3n k generalized coordinates, qj ; j = 1; : : : ; 3n k, so that the 3n kvariables qj , together with the k constraint conditions i(f~rig) = 0,determine the ~ri = ~ri(q1; : : : ; q3nk; t)

    The constrained subspace of conguration space need not be a flatspace. Consider, for example, a mass on one end of a rigid light rod

  • 1.3. SYSTEMS OF PARTICLES 15

    of length L, the other end of whichis xed to be at the origin ~r = 0,though the rod is completely freeto rotate. Clearly the possible val-ues of the cartesian coordinates ~rof the position of the mass satisfythe constraint j~rj = L, so ~r lieson the surface of a sphere of ra-dius L. We might choose as gen-eralized coordinates the standardspherical angles and . Thusthe constrained subspace is two di-mensional but not flat | rather itis the surface of a sphere, whichmathematicians call S2. It is nat-ural to reexpress the dynamics interms of and .

    j

    x

    y

    z

    q

    L

    Generalized coordinates (; ) fora particle constrained to lie on asphere.

    The use of generalized (non-cartesian) coordinates is not just forconstrained systems. The motion of a particle in a central force eldabout the origin, with a potential U(~r) = U(j~rj), is far more naturallydescribed in terms of spherical coordinates r, , and than in terms ofx, y, and z.

    Before we pursue a discussion of generalized coordinates, it must bepointed out that not all constraints are holonomic. The standard ex-ample is a disk of radius R, which rolls on a xed horizontal plane. It isconstrained to always remain vertical, and also to roll without slippingon the plane. As coordinates we can choose the x and y of the center ofthe disk, which are also the x and y of the contact point, together withthe angle a xed line on the disk makes with the downward direction,, and the angle the axis of the disk makes with the x axis, .

  • 16 CHAPTER 1. PARTICLE KINEMATICS

    As the disk rolls throughan angle d, the point ofcontact moves a distanceRd in a direction depend-ing on ,

    Rd sin = dx

    Rd cos = dy

    Dividing by dt, we get twoconstraints involving the po-sitions and velocities,

    1 := R _ sin _x = 02 := R _ cos _y = 0:

    The fact that these involvevelocities does not auto-matically make them non-holonomic. In the simplerone-dimensional problem inwhich the disk is conned tothe yz plane, rolling along

    x

    y

    z

    q

    f

    R

    A vertical disk free to roll on a plane.A xed line on the disk makes an angleof with respect to the vertical, andthe axis of the disk makes an angle with the x-axis. The long curved pathis the trajectory of the contact point.The three small paths are alternate tra-jectories illustrating that x, y, and caneach be changed without any net changein the other coordinates.

    x = 0 ( = 0), we would have only the coordinates and y, with therolling constraint R _ _y = 0. But this constraint can be integrated,R(t) y(t) = c, for some constant c, so that it becomes a constraintamong just the coordinates, and is holomorphic. This cannot be donewith the two-dimensional problem. We can see that there is no con-straint among the four coordinates themselves because each of themcan be changed by a motion which leaves the others unchanged. Ro-tating without moving the other coordinates is straightforward. Byrolling the disk along each of the three small paths shown to the rightof the disk, we can change one of the variables x, y, or , respectively,with no net change in the other coordinates. Thus all values of thecoordinates4 can be achieved in this fashion.

    4Thus the conguration space is x 2 R, y 2 R, 2 [0; 2) and 2 [0; 2),

  • 1.3. SYSTEMS OF PARTICLES 17

    There are other, less interesting, nonholonomic constraints given byinequalities rather than constraint equations. A bug sliding down abowling ball obeys the constraint j~rj R. Such problems are solved byconsidering the constraint with an equality (j~rj = R), but restrictingthe region of validity of the solution by an inequality on the constraintforce (N 0), and then supplementing with the unconstrained problemonce the bug leaves the surface.

    In quantum eld theory, anholonomic constraints which are func-tions of the positions and momenta are further subdivided into rstand second class constraints a la Dirac, with the rst class constraintsleading to local gauge invariance, as in Quantum Electrodynamics orYang-Mills theory. But this is heading far aeld.

    1.3.3 Generalized Coordinates for Unconstrained

    Systems

    Before we get further into constrained systems and DAlemberts Prin-ciple, we will discuss the formulation of a conservative unconstrainedsystem in generalized coordinates. Thus we wish to use 3n general-ized coordinates qj , which, together with time, determine all of the 3ncartesian coordinates ~ri:

    ~ri = ~ri(q1; :::; q3n; t):

    Notice that this is a relationship between dierent descriptions of thesame point in conguration space, and the functions ~ri(fqg; t) are in-dependent of the motion of any particle. We are assuming that the ~riand the qj are each a complete set of coordinates for the space, so theqs are also functions of the f~rig and t:

    qj = qj(~r1; :::; ~rn; t):

    The t dependence permits there to be an explicit dependence of thisrelation on time, as we would have, for example, in relating a rotatingcoordinate system to an inertial cartesian one.

    or, if we allow more carefully for the continuity as and go through 2, themore accurate statement is that conguration space is R2 (S1)2, where S1 is thecircumference of a circle, 2 [0; 2], with the requirement that = 0 is equivalentto = 2.

  • 18 CHAPTER 1. PARTICLE KINEMATICS

    Let us change the cartesian coordinate notation slightly, with fxkgthe 3n cartesian coordinates of the n 3-vectors ~ri, deemphasizing thedivision of these coordinates into triplets.

    A small change in the coordinates of a particle in congurationspace, whether an actual change over a small time interval dt or a\virtual" change between where a particle is and where it might havebeen under slightly altered circumstances, can be described by a set ofxk or by a set of qj . If we are talking about a virtual change at thesame time, these are related by the chain rule

    xk =Xj

    @xk@qj

    qj ; qj =Xk

    @qj@xk

    xk; (for t = 0): (1.6)

    For the actual motion through time, or any variation where t is notassumed to be zero, we need the more general form,

    xk =Xj

    @xk@qj

    qj +@xk@t

    t; qj =Xk

    @qj@xk

    xk +@qk@t

    t: (1.7)

    A virtual displacement, with t = 0, is the kind of variation we needto nd the forces described by a potential. Thus the force is

    Fk = @U(fxg)@xk

    = Xj

    @U(fx(fqg)g)@qj

    @qj@xk

    =Xj

    @qj@xk

    Qj ; (1.8)

    where

    Qj :=Xk

    Fk@xk@qj

    = @U(fx(fqg)g)@qj

    (1.9)

    is known as the generalized force. We may think of ~U(q; t) :=U(x(q); t) as a potential in the generalized coordinates fqg. Note thatif the coordinate transformation is time-dependent, it is possible thata time-independent potential U(x) will lead to a time-dependent po-tential ~U(q; t), and a system with forces described by a time-dependentpotential is not conservative.

    The denition in (1.9) of the generalized force Qj holds even if thecartesian force is not described by a potential.

    The qk do not necessarily have units of distance. For example,one qk might be an angle, as in polar or spherical coordinates. Thecorresponding component of the generalized force will have the units ofenergy and we might consider it a torque rather than a force.

  • 1.3. SYSTEMS OF PARTICLES 19

    1.3.4 Kinetic energy in generalized coordinates

    We have seen that, under the right circumstances, the potential energycan be thought of as a function of the generalized coordinates qk, andthe generalized forces Qk are given by the potential just as for ordinarycartesian coordinates and their forces. Now we examine the kineticenergy

    T =1

    2

    Xi

    mi _~ri2

    =1

    2

    Xj

    mj _x2j

    where the 3n values mj are not really independent, as each parti-cle has the same mass in all three dimensions in ordinary Newtonianmechanics5. Now

    _xj = limt!0

    xjt

    = limt!0

    0@Xk

    @xj@qk

    q;t

    qkt

    1A+ @xj@t

    q

    ;

    where jq;t means that t and the qs other than qk are held xed. Thelast term is due to the possibility that the coordinates xi(q1; :::; q3n; t)may vary with time even for xed values of qk. So the chain rule isgiving us

    _xj =dxjdt

    =Xk

    @xj@qk

    q;t

    _qk +@xj@t

    q

    : (1.10)

    Plugging this into the kinetic energy, we see that

    T =1

    2

    Xj;k;

    mj@xj@qk

    @xj@q

    _qk _q +Xj;k

    mj@xj@qk

    _qk@xj@t

    q

    +1

    2

    Xj

    mj

    0@ @xj@t

    q

    1A2 :(1.11)

    What is the interpretation of these terms? Only the rst term arisesif the relation between x and q is time independent. The second andthird terms are the sources of the _~r (~! ~r) and (~! ~r)2 terms in thekinetic energy when we consider rotating coordinate systems6.

    5But in an anisotropic crystal, the eective mass of a particle might in fact bedierent in dierent directions.

    6This will be fully developed in section 4.2

  • 20 CHAPTER 1. PARTICLE KINEMATICS

    Lets work a simple example: wewill consider a two dimensional systemusing polar coordinates with measuredfrom a direction rotating at angular ve-locity !. Thus the angle the radius vec-tor to an arbitrary point (r; ) makeswith the inertial x1-axis is + !t, andthe relations are

    x1 = r cos( + !t);

    x2 = r sin( + !t);

    with inverse relations

    r =qx21 + x

    22;

    = sin1(x2=r) !t:

    w

    q

    t

    r

    x

    x1

    2

    Rotating polar coordinates

    related to inertial cartesiancoordinates.

    So _x1 = _r cos(+!t) _r sin(+!t)!r sin(+!t), where the last termis from @xj=@t, and _x2 = _r sin(+!t)+ _r cos(+!t)+!r cos(+!t).In the square, things get a bit simpler,

    P_x2i = _r

    2 + r2(! + _)2.We see that the form of the kinetic energy in terms of the generalized

    coordinates and their velocities is much more complicated than it isin cartesian inertial coordinates, where it is coordinate independent,and a simple diagonal quadratic form in the velocities. In generalizedcoordinates, it is quadratic but not homogeneous7 in the velocities, andwith an arbitrary dependence on the coordinates. In general, even if thecoordinate transformation is time independent, the form of the kineticenergy is still coordinate dependent and, while a purely quadratic formin the velocities, it is not necessarily diagonal. In this time-independentsituation, we have

    T =1

    2

    Xk

    Mk _qk _q; with Mk =Xj

    mj@xj@qk

    @xj@q

    ; (1.12)

    where Mk is known as the mass matrix, and is always symmetric butnot necessarily diagonal or coordinate independent.

    7It involves quadratic and lower order terms in the velocities, not just quadraticones.

  • 1.4. PHASE SPACE 21

    The mass matrix is independent of the @xj=@t terms, and we canunderstand the results we just obtained for it in our two-dimensionalexample above,

    M11 = m; M12 = M21 = 0; M22 = mr2;

    by considering the case without rotation, ! = 0. We can also derivethis expression for the kinetic energy in nonrotating polar coordinatesby expressing the velocity vector ~v = _re^r+r _e^ in terms of unit vectorsin the radial and tangential directions respectively. The coecientsof these unit vectors can be understood graphically with geometricarguments. This leads more quickly to ~v 2 = ( _r)2 + r2( _)2, T = 1

    2m _r2 +

    12mr2 _2, and the mass matrix follows. Similar geometric arguments

    are usually used to nd the form of the kinetic energy in sphericalcoordinates, but the formal approach of (1.12) enables us to nd theform even in situations where the geometry is dicult to picture.

    It is important to keep in mind that when we view T as a function ofcoordinates and velocities, these are independent arguments evaluatedat a particular moment of time. Thus we can ask independently how Tvaries as we change xi or as we change _xi, each time holding the othervariable xed. Thus the kinetic energy is not a function on the 3n-dimensional conguration space, but on a larger, 6n-dimensional space8

    with a point specifying both the coordinates fqig and the velocities f _qig.

    1.4 Phase Space

    If the trajectory of the system in conguration space, ~r(t), is known, thevelocity as a function of time, ~v(t) is also determined. As the mass of theparticle is simply a physical constant, the momentum ~p = m~v containsthe same information as the velocity. Viewed as functions of time, thisgives nothing beyond the information in the trajectory. But at anygiven time, ~r and ~p provide a complete set of initial conditions, while ~ralone does not. We dene phase space as the set of possible positions

    8This space is called the tangent bundle to conguration space. For cartesiancoordinates it is almost identical to phase space, which is in general the \cotangentbundle" to conguration space.

  • 22 CHAPTER 1. PARTICLE KINEMATICS

    and momenta for the system at some instant. Equivalently, it is the setof possible initial conditions, or the set of possible motions obeying theequations of motion. For a single particle in cartesian coordinates, thesix coordinates of phase space are the three components of ~r and thethree components of ~p. At any instant of time, the system is representedby a point in this space, called the phase point, and that point moveswith time according to the physical laws of the system. These laws areembodied in the force function, which we now consider as a function of~p rather than ~v, in addition to ~r and t. We may write these equationsas

    d~r

    dt=

    ~p

    m;

    d~p

    dt= ~F (~r; ~p; t):

    Note that these are rst order equations, which means that the mo-tion of the point representing the system in phase space is completelydetermined9 by where the phase point is. This is to be distinguishedfrom the trajectory in conguration space, where in order to know thetrajectory you must have not only an initial point (position) but alsoan initial velocity.

    1.4.1 Dynamical Systems

    We have spoken of the coordinates of phase space for a single par-ticle as ~r and ~p, but from a mathematical point of view these to-gether give the coordinates of the phase point in phase space. Wemight describe these coordinates in terms of a six dimensional vector~ = (r1; r2; r3; p1; p2; p3). The physical laws determine at each pointa velocity function for the phase point as it moves through phasespace,

    d~

    dt= ~V (~; t); (1.13)

    which gives the velocity at which the phase point representing the sys-tem moves through phase space. Only half of this velocity is the ordi-

    9We will assume throughout that the force function is a well dened continuousfunction of its arguments.

  • 1.4. PHASE SPACE 23

    nary velocity, while the other half represents the rapidity with which themomentum is changing, i.e. the force. The path traced by the phasepoint as it travels through phase space is called the phase curve.

    For a system of n particles in three dimensions, the complete set ofinitial conditions requires 3n spatial coordinates and 3n momenta, sophase space is 6n dimensional. While this certainly makes visualizationdicult, the large dimensionality is no hindrance for formal develop-ments. Also, it is sometimes possible to focus on particular dimensions,or to make generalizations of ideas familiar in two and three dimensions.For example, in discussing integrable systems (7.1), we will nd thatthe motion of the phase point is conned to a 3n-dimensional torus, ageneralization of one and two dimensional tori, which are circles andthe surface of a donut respectively.

    Thus for a system composed of a nite number of particles, thedynamics is determined by the rst order ordinary dierential equation(1.13), formally a very simple equation. All of the complication of thephysical situation is hidden in the large dimensionality of the dependentvariable ~ and in the functional dependence of the velocity functionV (~; t) on it.

    There are other systems besides Newtonian mechanics which arecontrolled by equation (1.13), with a suitable velocity function. Collec-tively these are known as dynamical systems. For example, individ-uals of an asexual mutually hostile species might have a xed birth rateb and a death rate proportional to the population, so the populationwould obey the logistic equation10 dp=dt = bp cp2, a dynamicalsystem with a one-dimensional space for its dependent variable. Thepopulations of three competing species could be described by eq. (1.13)with ~ in three dimensions.

    The dimensionality d of ~ in (1.13) is called the order of the dy-namical system. A dth order dierential equation in one independentvariable may always be recast as a rst order dierential equation in dvariables, so it is one example of a dth order dynamical system. Thespace of these dependent variables is called the phase space of the dy-namical system. Newtonian systems always give rise to an even-order

    10This is not to be confused with the simpler logistic map, which is a recursionrelation with the same form but with solutions displaying a very dierent behavior.

  • 24 CHAPTER 1. PARTICLE KINEMATICS

    system, because each spatial coordinate is paired with a momentum.For n particles unconstrained in D dimensions, the order of the dy-namical system is d = 2nD. Even for constrained Newtonian systems,there is always a pairing of coordinates and momenta, which gives arestricting structure, called the symplectic structure11, on phase space.

    If the force function does not depend explicitly on time, we say thesystem is autonomous. The velocity function has no explicit depen-dance on time, ~V = ~V (~), and is a time-independent vector eld onphase space, which we can indicate by arrows just as we might theelectric eld in ordinary space. This gives a visual indication of themotion of the systems point. For example, consider a damped har-monic oscillator with ~F = kx p, for which the velocity functionis

    dx

    dt;dp

    dt

    !=p

    m;kx p

    :

    A plot of this eld for the undamped ( = 0) and damped oscillators

    x

    p

    Undamped

    x

    p

    DampedFigure 1.1: Velocity eld for undamped and damped harmonic oscil-lators, and one possible phase curve for each system through phasespace.

    is shown in Figure 1.1. The velocity eld is everywhere tangent to anypossible path, one of which is shown for each case. Note that qualitativefeatures of the motion can be seen from the velocity eld without anysolving of the dierential equations; it is clear that in the damped casethe path of the system must spiral in toward the origin.

    The paths taken by possible physical motions through the phasespace of an autonomous system have an important property. Because

    11This will be discussed in sections (6.3) and (6.6).

  • 1.4. PHASE SPACE 25

    the rate and direction with which the phase point moves away froma given point of phase space is completely determined by the velocityfunction at that point, if the system ever returns to a point it mustmove away from that point exactly as it did the last time. That is,if the system at time T returns to a point in phase space that it wasat at time t = 0, then its subsequent motion must be just as it was,so ~(T + t) = ~(t), and the motion is periodic with period T . Thisalmost implies that the phase curve the object takes through phasespace must be nonintersecting12.

    In the non-autonomous case, where the velocity eld is time depen-dent, it may be preferable to think in terms of extended phase space, a6n+1 dimensional space with coordinates (~; t). The velocity eld canbe extended to this space by giving each vector a last component of 1,as dt=dt = 1. Then the motion of the system is relentlessly upwards inthis direction, though still complex in the others. For the undampedone-dimensional harmonic oscillator, the path is a helix in the threedimensional extended phase space.

    Most of this book is devoted to nding analytic methods for ex-ploring the motion of a system. In several cases we will be able tond exact analytic solutions, but it should be noted that these exactlysolvable problems, while very important, cover only a small set of realproblems. It is therefore important to have methods other than search-ing for analytic solutions to deal with dynamical systems. Phase spaceprovides one method for nding qualitative information about the so-lutions. Another approach is numerical. Newtons Law, and moregenerally the equation (1.13) for a dynamical system, is a set of ordi-nary dierential equations for the evolution of the systems position inphase space. Thus it is always subject to numerical solution given aninitial conguration, at least up until such point that some singularityin the velocity function is reached. One primitive technique which willwork for all such systems is to choose a small time interval of lengtht, and use d~=dt at the beginning of each interval to approximate ~during this interval. This gives a new approximate value for ~ at the

    12An exception can occur at an unstable equilibrium point, where the velocityfunction vanishes. The motion can just end at such a point, and several possiblephase curves can terminate at that point.

  • 26 CHAPTER 1. PARTICLE KINEMATICS

    end of this interval, which may then be taken as the beginning of thenext.13

    As an example, we show themeat of a calculation for thedamped harmonic oscillator, inFortran. This same techniquewill work even with a very com-plicated situation. One needonly add lines for all the com-ponents of the position and mo-mentum, and change the forcelaw appropriately.

    This is not to say that nu-merical solution is a good way

    do i = 1,n

    dx = (p/m) * dt

    dp = -(k*x+alpha*p)*dt

    x = x + dx

    p = p + dp

    t = t + dt

    write *, t, x, p

    enddo

    Integrating the motion, for adamped harmonic oscillator.

    to solve this problem. An analytical solution, if it can be found, isalmost always preferable, because

    It is far more likely to provide insight into the qualitative featuresof the motion.

    Numerical solutions must be done separately for each value of theparameters (k;m; ) and each value of the initial conditions (x0and p0).

    Numerical solutions have subtle numerical problems in that theyare only exact as t! 0, and only if the computations are doneexactly. Sometimes uncontrolled approximate solutions lead tosurprisingly large errors.

    13This is a very unsophisticated method. The errors made in each step for ~rand ~p are typically O(t)2. As any calculation of the evolution from time t0to tf will involve a number ([tf t0]=t) of time steps which grows inversely tot, the cumulative error can be expected to be O(t). In principle therefore wecan approach exact results for a nite time evolution by taking smaller and smallertime steps, but in practise there are other considerations, such as computer time androundo errors, which argue strongly in favor of using more sophisticated numericaltechniques, with errors of higher order in t. These can be found in any text onnumerical methods.

  • 1.4. PHASE SPACE 27

    Nonetheless, numerical solutions are often the only way to handle areal problem, and there has been extensive development of techniquesfor eciently and accurately handling the problem, which is essentiallyone of solving a system of rst order ordinary dierential equations.

    1.4.2 Phase Space Flows

    As we just saw, Newtons equations for a system of particles can becast in the form of a set of rst order ordinary dierential equationsin time on phase space, with the motion in phase space described bythe velocity eld. This could be more generally discussed as a dthorder dynamical system, with a phase point representing the systemin a d-dimensional phase space, moving with time t along the velocityeld, sweeping out a path in phase space called the phase curve. Thephase point ~(t) is also called the state of the system at time t. Manyqualitative features of the motion can be stated in terms of the phasecurve.

    Fixed Points

    There may be points ~k, known as xed points, at which the velocityfunction vanishes, ~V (~k) = 0. This is a point of equilibrium for thesystem, for if the system is at a xed point at one moment, ~(t0) = ~k,it remains at that point. At other points, the system does not stayput, but there may be sets of states which flow into each other, suchas the elliptical orbit for the undamped harmonic oscillator. These arecalled invariant sets of states. In a rst order dynamical system14,the xed points divide the line into intervals which are invariant sets.

    Even though a rst-order system is smaller than any Newtonian sys-tem, it is worthwhile discussing briefly the phase flow there. We havebeen assuming the velocity function is a smooth function | genericallyits zeros will be rst order, and near the xed point 0 we will haveV () c( 0). If the constant c < 0, d=dt will have the oppo-site sign from 0, and the system will flow towards the xed point,

    14Note that this is not a one-dimensional Newtonian system, which is a twodimensional ~ = (x; p) dynamical system.

  • 28 CHAPTER 1. PARTICLE KINEMATICS

    which is therefore called stable. On the other hand, if c > 0, the dis-placement 0 will grow with time, and the xed point is unstable.Of course there are other possibilities: if V () = c2, the xed point = 0 is stable from the left and unstable from the right. But this kindof situation is somewhat articial, and such a system is structuallyunstable. What that means is that if the velocity eld is perturbedby a small smooth variation V () ! V () + w(), for some boundedsmooth function w, the xed point at = 0 is likely to either disap-pear or split into two xed points, whereas the xed points discussedearlier will simply be shifted by order in position and will retain theirstability or instability. Thus the simple zero in the velocity function isstructurally stable. Note that structual stability is quite a dierentnotion from stability of the xed point.

    In this discussion of stability in rst order dynamical systems, wesee that generically the stable xed points occur where the velocityfunction decreases through zero, while the unstable points are where itincreases through zero. Thus generically the xed points will alternatein stability, dividing the phase line into open intervals which are eachinvariant sets of states, with the points in a given interval flowing eitherto the left or to the right, but never leaving the open interval. The statenever reaches the stable xed point because the time t =

    Rd=V ()

    (1=c)Rd=(0) diverges. On the other hand, in the case V () = c2,

    a system starting at 0 at t = 0 has a motion given by = (10 ct)1,

    which runs o to innity as t ! 1=0c. Thus the solution terminatesat t = 1=0c, and makes no sense thereafter. This form of solution iscalled terminating motion.

    For higher order dynamical systems, the d equations Vi(~) = 0required for a xed point will generically determine the d variablesj , so the generic form of the velocity eld near a xed point 0 isVi(~) =

    PjMij(j 0j) with a nonsingular matrix M . The stability

    of the flow will be determined by this d-dimensional square matrix M .Generically the eigenvalue equation, a dth order polynomial in , willhave d distinct solutions. Because M is a real matrix, the eigenvaluesmust either be real or come in complex conjugate pairs. For the realcase, whether the eigenvalue is positive or negative determines the in-stability or stability of the flow along the direction of the eigenvector.For a pair of complex conjugate eigenvalues = u+ iv and = u iv,

  • 1.4. PHASE SPACE 29

    with eigenvectors ~e and ~e respectively, we may describe the flow in theplane ~ = ~ ~0 = x(~e+ ~e ) + iy(~e ~e ), so

    _~ = M ~ = x(~e+ ~e ) + iy(~e ~e )= (ux vy)(~e+ ~e ) + (vx+ uy)(~e ~e )

    so _x_y

    =u vv u

    xy

    ; or

    x = Aeut cos(vt+ )y = Aeut sin(vt+ )

    :

    Thus we see that the motion spirals in towards the xed point if u isnegative, and spirals away from the xed point if u is positive. Stabilityin these directions is determined by the sign of the real part of theeigenvalue.

    In general, then, stability in each subspace around the xed point ~0depends on the sign of the real part of the eigenvalue. If all the real partsare negative, the system will flow from anywhere in some neighborhoodof ~0 towards the xed point, so limt!1 ~(t) = ~0 provided we startin that neighborhood. Then ~0 is an attractor and is a stronglystable xed point. On the other hand, if some of the eigenvalueshave positive real parts, there are unstable directions. Starting froma generic point in any neighborhood of ~0, the motion will eventuallyflow out along an unstable direction, and the xed point is consideredunstable, although there may be subspaces along which the flow maybe into ~0. An example is the line x = y in the hyperbolic xedpoint case shown in Figure 1.2.

    Some examples of two dimensional flows in the neighborhood of ageneric xed point are shown in Figure 1.2. Note that none of thesedescribe the xed point of the undamped harmonic oscillator of Figure1.1. We have discussed generic situations as if the velocity eld werechosen arbitrarily from the set of all smooth vector functions, but infact Newtonian mechanics imposes constraints on the velocity elds inmany situations, in particular if there are conserved quantities.

    Eect of conserved quantities on the flow

    If the system has a conserved quantity Q(q; p) which is a function onphase space only, and not of time, the flow in phase space is consider-ably changed. This is because the equations Q(q; p) = K gives a set

  • 30 CHAPTER 1. PARTICLE KINEMATICS

    _x = x+ y;_y = 2x y:

    Strongly stablespiral point.

    = 1p

    2i:

    _x = 3x y;_y = x 3y:

    Strongly stablexed point,

    = 1;2:

    _x = 3x+ y;

    _y = x+ 3y:

    Unstable xedpoint,

    = 1; 2:

    _x = x 3y;_y = 3x y:

    Hyperbolicxed point,

    = 2; 1:

    Figure 1.2: Four generic xed points for a second order dynamicalsystem.

    of subsurfaces or contours in phase space, and the system is connedto stay on whichever contour it is on initially. Unless this conservedquantity is a trivial function, i.e. constant, in the vicinity of a xedpoint, it is not possible for all points to flow into the xed point, andthus it is not strongly stable. In the terms of our generic discussion,the gradient of Q gives a direction orthogonal to the image of M , sothere is a zero eigenvalue and we are not in the generic situation wediscussed.

    For the case of a single particle in a potential, the total energyE = p2=2m + U(~r) is conserved, and so the motion of the systemis conned to one surface of a given energy. As ~p=m is part of thevelocity function, a xed point must have ~p = 0. The vanishing ofthe other half of the velocity eld gives rU(~r0) = 0, which is thecondition for a stationary point of the potential energy, and for theforce to vanish. If this point is a maximum or a saddle of U , themotion along a descending path will be unstable. If the xed pointis a minimum of the potential, the region E(~r; ~p) < E(~r0; 0) + , for

  • 1.4. PHASE SPACE 31

    suciently small , gives a neighborhood around ~0 = (~r0; 0) to whichthe motion is conned if it starts within this region. Such a xed point iscalled stable15, but it is not strongly stable, as the flow does not settledown to ~0. This is the situation we saw for the undamped harmonicoscillator. For that situation F = kx, so the potential energy may betaken to be

    U(x) =Z 0xkx dx = 1

    2kx2;

    and so the total energy E = p2=2m + 12kx2 is conserved. The curves

    of constant E in phase space are ellipses, and each motion orbits theappropriate ellipse, as shown in Fig. 1.1 for the undamped oscillator.This contrasts to the case of the damped oscillator, for which there isno conserved energy, and for which the origin is a strongly stable xedpoint.

    15A xed point is stable if it is in arbitrarity small neighborhoods, each with theproperty that if the system is in that neighborhood at one time, it remains in it atall later times.

  • 32 CHAPTER 1. PARTICLE KINEMATICS

    As an example of a con-servative system with both sta-ble and unstable xed points,consider a particle in one di-mension with a cubic potentialU(x) = ax2 bx3, as shown inFig. 1.3. There is a stable equi-librium at xs = 0 and an un-stable one at xu = 2a=3b. Eachhas an associated xed point inphase space, an elliptic xedpoint s = (xs; 0) and a hyper-bolic xed point u = (xu; 0).The velocity eld in phasespace and several possible or-bits are shown. Near the sta-ble equilibrium, the trajectoriesare approximately ellipses, asthey were for the harmonic os-cillator, but for larger energiesthey begin to feel the asym-metry of the potential, andthe orbits become egg-shaped.

    1

    -1

    x

    p

    1.210.80.60.40.2-0.2-0.4 0

    0.3

    0.2

    0.1

    0

    -0.1

    -0.2

    -0.3

    x

    UU(x)

    Figure 1.3. Motion in a cubic poten-

    tial.

    If the system has total energy precisely U(xu), the contour linecrosses itself. This contour actually consists of three separate orbits.One starts at t ! 1 at x = xu, completes one trip though thepotential well, and returns as t ! +1 to x = xu. The other two areorbits which go from x = xu to x = 1, one incoming and one outgoing.For E > U(xu), all the orbits start and end at x = +1. Note thatgenerically the orbits deform continuously as the energy varies, but atE = U(xu) this is not the case | the character of the orbit changes asE passes through U(xu). An orbit with this critical value of the energyis called a seperatrix, as it seperates regions in phase space where theorbits have dierent qualitative characteristics.

    Quite generally hyperbolic xed points are at the ends of seperatri-ces. In our case the contour E = U(xu) consists of four invariant sets

  • 1.4. PHASE SPACE 33

    of states, one of which is the point u itself, and the other three arethe orbits which are the disconnected pieces left of the contour afterremoving u.

    Exercises

    1.1 (a) Find the potential energy function U(~r) for a particle in the grav-itational eld of the Earth, for which the force law is ~F (~r) = GMEm~r=r3.(b) Find the escape velocity from the Earth, that is, the minimum velocitya particle near the surface can have for which it is possible that the particlewill eventually coast to arbitrarily large distances without being acted uponby any force other than gravity. The Earth has a mass of 6:0 1024 kg anda radius of 6:4 106 m. Newtons gravitational constant is 6:67 1011N m2=kg2.

    1.2 In the discussion of a system of particles, it is important that theparticles included in the system remain the same. There are some situationsin which we wish to focus our attention on a set of particles which changeswith time, such as a rocket ship which is emitting gas continuously. Theequation of motion for such a problem may be derived by considering aninnitesimal time interval, [t; t + t], and choosing the system to be therocket with the fuel still in it at time t, so that at time t + t the systemconsists of the rocket with its remaining fuel and also the small amount offuel emitted during the innitesimal time interval.Let M(t) be the mass of the rocket and remaining fuel at time t, assume thatthe fuel is emitted with velocity ~u with respect to the rocket, and call thevelocity of the rocket ~v(t) in an inertial coordinate system. If the externalforce on the rocket is ~F (t) and the external force on the innitesimal amountof exhaust is innitesimal, the fact that F (t) is the rate of change of the totalmomentum gives the equation of motion for the rocket.(a) Show that this equation is

    Md~v

    dt= ~F (t) + ~u

    dM

    dt:

    (b) Suppose the rocket is in a constant gravitational eld ~F = Mge^z forthe period during which it is burning fuel, and that it is red straight upwith constant exhaust velocity (~u = ue^z), starting from rest. Find v(t) interms of t and M(t).(c) Find the maximum fraction of the initial mass of the rocket which canescape the Earths gravitational eld if u = 2000m/s.

  • 34 CHAPTER 1. PARTICLE KINEMATICS

    1.3 For a particle in two dimensions, we might use polar coordinates (r; )and use basis unit vectors e^r and e^ in the radial and tangent directionsrespectively to describe more general vectors. Because this pair of unitvectors dier from point to point, the e^r and e^ along the trajectory of amoving particle are themselves changing with time.(a) Show that

    d

    dte^r = _e^;

    d

    dte^ = _e^r:

    (b) Thus show that the derivative of ~r = re^r is

    ~v = _re^r + r _e^;

    which veries the discussion of Sec. (1.3.4).(c) Show that the derivative of the velocity is

    ~a =d

    dt~v = (r r _2)e^r + (r + 2 _r _)e^:

    (d) Thus Newtons Law says for the radial and tangential components ofthe force are Fr = e^r F = m(r r _2), F = e^ F = m(r + 2 _r _). Showthat the generalized forces are Qr = Fr and Q = rF.

    1.4 Analyze the errors in the integration of Newtons Laws in the sim-ple Eulers approach described in section 1.4.1, where we approximatedthe change for x and p in each time interval t between ti and ti+1 by_x(t) _x(ti), _p(t) F (x(ti); v(ti)). Assuming F to be dierentiable, showthat the error which accumulates in a nite time interval T is of order (t)1.

    1.5 Write a simple program to integrate the equation of the harmonic os-cillator through one period of oscillation, using Eulers method with a stepsize t. Do this for several t, and see whether the error accumulated inone period meets the expectations of problem 1.4.

    1.6 Describe the one dimensional phase space for the logistic equation _p =bpcp2, with b > 0; c > 0. Give the xed points, the invariant sets of states,and describe the flow on each of the invariant sets.

    1.7 Consider a pendulum consisting of a mass at the end of a massless rodof length L, the other end of which is xed but free to rotate. Ignore one ofthe horizontal directions, and describe the dynamics in terms of the angle

  • 1.4. PHASE SPACE 35

    between the rod and the downwards direction, without making a small angleapproximation.(a) Find the generalized force Q and nd the conserved quantity on phasespace.(b) Give a sketch of the velocity function, including all the regions of phasespace. Show all xed points, seperatrices, and describe all the invariant setsof states. [Note: the variable is dened only modulo 2, so the phasespace is the Cartesian product of an interval of length 2 in with the realline for p. This can be plotted on a strip, with the understanding that theleft and right edges are identied. To avoid having important points on theboundary, it would be well to plot this with 2 [=2; 3=2].

  • 36 CHAPTER 1. PARTICLE KINEMATICS

  • Chapter 2

    Lagranges and HamiltonsEquations

    In this chapter, we consider two reformulations of Newtonian mechan-ics, the Lagrangian and the Hamiltonian formalism. The rst is natu-rally associated with conguration space, extended by time, while thelatter is the natural description for working in phase space.

    Lagrange developed his approach in 1764 in a study of the libra-tion of the moon, but it is best thought of as a general method oftreating dynamics in terms of generalized coordinates for congurationspace. It so transcends its origin that the Lagrangian is considered thefundamental object which describes a quantum eld theory.

    Hamiltons approach arose in 1835 in his unication of the languageof optics and mechanics. It too had a usefulness far beyond its origin,and the Hamiltonian is now most familiar as the operator in quantummechanics which determines the evolution in time of the wave function.

    2.1 Lagrangian Mechanics

    We begin by deriving Lagranges equation as a simple change of co-ordinates in an unconstrained system, one which is evolving accordingto Newtons laws with force laws given by some potential. Lagrangianmechanics is also and especially useful in the presence of constraints,so we will then extend the formalism to this more general situation.

    37

  • 38 CHAPTER 2. LAGRANGES AND HAMILTONS EQUATIONS

    2.1.1 Derivation for unconstrained systems

    For a collection of particles with conservative forces described by apotential, we have in inertial cartesian coordinates

    mxi = Fi:

    The left hand side of this equation is determined by the kinetic energyfunction as the time derivative of the momentum pi = @T=@ _xi, whilethe right hand side is a derivative of the potential energy, @U=@xi. AsT is independent of xi and U is independent of _xi in these coordinates,we can write both sides in terms of the Lagrangian L = T U , whichis then a function of both the coordinates and their velocities. Thus wehave established

    d

    dt

    @L

    @ _xi @L@xi

    = 0;

    which, once we generalize it to arbitrary coordinates, will be known asLagranges equation. This particular combination of T ( _~r) with U(~r)to get the more complicated L(~r; _~r) seems an articial construction forthe inertial cartesian coordinates, but it has the advantage of preservingthe form of Lagranges equations for any set of generalized coordinates.

    As we did in section 1.3.3, we assume we have a set of generalizedcoordinates fqjg which parameterize all of coordinate space, so thateach point may be described by the fqjg or by the fxig, i; j 2 [1; N ],and thus each set may be thought of as a function of the other, andtime:

    qj = qj(x1; :::xN ; t) xi = xi(q1; :::qN ; t): (2.1)

    We may consider L as a function1 of the generalized coordinates qj and_qj , and ask whether the same expression in these coordinates

    d

    dt

    @L

    @ _qj @L@qj

    1Of course we are not saying that L(x; _x; t) is the same function of its coor-dinates as L(q; _q; t), but rather that these are two functions which agree at thecorresponding physical points. More precisely, we are dening a new function~L(q; _q; t) = L(x(q; t); _x(q; _q; t); t), but we are being physicists and neglecting thetilde. We are treating the Lagrangian here as a scalar under coordinate transfor-mations, in the sense used in general relativity, that its value at a given physicalpoint is unchanged by changing the coordinate system used to dene that point.

  • 2.1. LAGRANGIAN MECHANICS 39

    also vanishes. The chain rule tells us

    @L

    @ _xj=Xk

    @L

    @qk

    @qk@ _xj

    +Xk

    @L

    @ _qk

    @ _qk@ _xj

    : (2.2)

    The rst term vanishes because qk depends only on the coordinates xkand t, but not on the _xk. From the inverse relation to (1.10),

    _qj =Xi

    @qj@xi

    _xi +@qj@t; (2.3)

    we have@ _qj@ _xi

    =@qj@xi

    :

    Using this in (2.2),@L

    @ _xi=Xj

    @L

    @ _qj

    @qj@xi

    : (2.4)

    Lagranges equation involves the time derivative of this. Here whatis meant is not a partial derivative @=@t, holding the point in congu-ration space xed, but rather the derivative along the path which thesystem takes as it moves through conguration space. It is called thestream derivative, a name which comes from fluid mechanics, whereit gives the rate at which some property dened throughout the fluid,f(~r; t), changes for a xed element of fluid as the fluid as a whole flows.We write it as a total derivative to indicate that we are following themotion rather than evaluating the rate of change at a xed point inspace, as the partial derivative does.

    For any function f(x; t) of extended conguration space, this totaltime derivative is

    df

    dt=Xj

    @f

    @xj_xj +

    @f

    @t: (2.5)

    Using Leibnitz rule on (2.4) and using (2.5) in the second term, wend

    d

    dt

    @L

    @ _xi=Xj

    d

    dt

    @L

    @ _qj

    !@qj@xi

    +Xj

    @L

    @ _qj

    Xk

    @2qj@xi@xk

    _xk +@2qj@xi@t

    !: (2.6)

  • 40 CHAPTER 2. LAGRANGES AND HAMILTONS EQUATIONS

    On the other hand, the chain rule also tells us

    @L

    @xi=Xj

    @L

    @qj

    @qj@xi

    +Xj

    @L

    @ _qj

    @ _qj@xi

    ;

    where the last term does not necessarily vanish, as _qj in general dependson both the coordinates and velocities. In fact, from 2.3,

    @ _qj@xi

    =Xk

    @2qj@xi@xk

    _xk +@2qj@xi@t

    ;

    so

    @L

    @xi=Xj

    @L

    @qj

    @qj@xi

    +Xj

    @L

    @ _qj

    Xk

    @2qj@xi@xk

    _xk +@2qj@xi@t

    !: (2.7)

    Lagranges equation in cartesian coordinates says (2.6) and (2.7) areequal, and in subtracting them the second terms cancel2, so

    0 =Xj

    d

    dt

    @L

    @ _qj @L@qj

    !@qj@xi

    :

    The matrix @qj=@xi is nonsingular, as it has @xi=@qj as its inverse, sowe have derived Lagranges Equation in generalized coordinates:

    d

    dt

    @L

    @ _qj @L@qj

    = 0:

    Thus we see that Lagranges equations are form invariant underchanges of the generalized coordinates used to describe the congura-tion of the system. It is primarily for this reason that this particularand peculiar combination of kinetic and potential energy is useful. Notethat we implicity assume the Lagrangian itself transformed like a scalar,in that its value at a given physical point of conguration space is in-dependent of the choice of generalized coordinates that describe thepoint. The change of coordinates itself (2.1) is called a point trans-formation.

    2This is why we chose the particular combination we did for the Lagrangian,rather than L = T U for some 6= 1. Had we done so, Lagranges equationin cartesian coordinates would have been d(@L=@ _xj)=dt @L=@xj = 0, and inthe subtraction of (2.7) from (2.6), the terms proportional to @L=@ _qi (withouta time derivative) would not have cancelled.

  • 2.1. LAGRANGIAN MECHANICS 41

    2.1.2 Lagrangian for Constrained Systems

    We now wish to generalize our discussion to include contraints. Atthe same time we will also consider possibly nonconservative forces.As we mentioned in section 1.3.2, we often have a system with internalforces whose eect is better understood than the forces themselves, withwhich we may not be concerned. We will assume the constraints areholonomic, expressible as k real functions (~r1; :::; ~rn; t) = 0, which

    are somehow enforced by constraint forces ~FCi on the particle i. Theremay also be other forces, which we will call FDi and will treat as havinga dynamical eect. These are given by known functions of the cong-uration and time, possibly but not necessarily in terms of a potential.

    This distinction will seem articial without examples, so it wouldbe well to keep these two in mind. In each of these cases the fullconguration space is R3, but the constraints restrict the motion to anallowed subspace of extended conguration space.

    1. In section 1.3.2 we discussed a mass on a light rigid rod, the otherend of which is xed at the origin. Thus the mass is constrainedto have j~rj = L, and the allowed subspace of conguration spaceis the surface of a sphere, independent of time. The rod exerts theconstraint force to avoid compression or expansion. The naturalassumption to make is that the force is in the radial direction, andtherefore has no component in the direction of allowed motions,the tangential directions. That is, for all allowed displacements,~r, we have ~FC ~r = 0, and the constraint force does no work.

    2. Consider a bead free to slide without friction on the spoke of a ro-tating bicycle wheel3, rotating about a xed axis at xed angularvelocity !. That is, for the polar angle of inertial coordinates, := !t = 0 is a constraint4, but the r coordinate is uncon-strained. Here the allowed subspace is not time independent, butis a helical sort of structure in extended conguration space. Weexpect the force exerted by the spoke on the bead to be in the e^

    3Unlike a real bicycle wheel, we are assuming here that the spoke is directlyalong a radius of the circle, pointing directly to the axle.

    4There is also a constraint z = 0.

  • 42 CHAPTER 2. LAGRANGES AND HAMILTONS EQUATIONS

    direction. This is again perpendicular to any virtual displace-ment, by which we mean an allowed change in conguration at axed time. It is important to distinguish this virtual displacementfrom a small segment of the trajectory of the particle. In this casea virtual displacement is a change in r without a change in , andis perpendicular to e^. So again, we have the \net virtual work"of the constraint forces is zero. It is important to note that thisdoes not mean that the net real work is zero. In a small timeinterval, the displacement ~r includes a component r!t in thetangential direction, and the force of constraint does do work!

    We will assume that the constraint forces in general satisfy thisrestriction that no net virtual work is done by the forces of constraintfor any possible virtual displacement. Newtons law tells us that _~pi =Fi = F

    Ci + F

    Di . We can multiply by an arbitrary virtual displacementX

    i

    ~FDi _~pi

    ~ri =

    Xi

    ~FCi ~ri = 0;

    where the rst equality would be true even if ~ri did not satisfy theconstraints, but the second requires ~ri to be an allowed virtual dis-placement. Thus X

    i

    ~FDi _~pi

    ~ri = 0; (2.8)

    which is known as DAlemberts Principle. This gives an equationwhich determines the motion on the constrained subspace and does notinvolve the unspecied forces of constraint FC . We drop the super-script D from now on.

    Suppose we know generalized coordinates q1; : : : ; qN which parame-terize the constrained subspace, which means ~ri = ~ri(q1; : : : ; qN ; t), fori = 1; : : : ; n, are known functions and the N qs are independent. Thereare N = 3n k of these independent coordinates, where k is the num-ber of holonomic constraints. Then @~ri=@qj is no longer an invertable,or even square, matrix, but we still have

    ~ri =Xj

    @~ri@qj

    qj +@~ri@t

    t:

  • 2.1. LAGRANGIAN MECHANICS 43

    For the velocity of the particle, divide this by t, giving

    ~vi =Xj

    @~ri@qj

    _qj +@~ri@t; (2.9)

    but for a virtual displacement t = 0 we have

    ~ri =Xj

    @~ri@qj

    qj :

    Dierentiating (2.9) we note that,

    @~vi@ _qj

    =@~ri@qj

    ; (2.10)

    and also@~vi@qj

    =Xk

    @2~ri@qj@qk

    _qk +@2~ri@qj@t

    =d

    dt

    @~ri@qj

    ; (2.11)

    where the last equality comes from applying (2.5), with coordinates qjrather than xj, to f = @~ri=@qj . The rst term in the equation (2.8)stating DAlemberts principle isX

    i

    ~Fi ~ri =Xj

    Xi

    ~Fi @~ri@qj

    qj =Xj

    Qj qj:

    The generalized force Qj has the same form as in the unconstrainedcase, as given by (1.9), but there are only as many of them as there areunconstrained degrees of freedom.

    The second term involvesXi

    _~pi ~ri =Xi

    dpidt

    @~ri@qj

    qj

    =Xj

    d

    dt

    Xi

    ~pi @~ri@qj

    !qj

    Xij

    pi d

    dt

    @~ri@qj

    !qj

    =Xj

    d

    dt

    Xi

    ~pi @~vi@ _qj

    !qj

    Xij

    pi @~vi@qj

    qj

    =Xj

    "d

    dt

    Xi

    mi~vi @~vi@ _qj

    Xi

    mivi @~vi@qj

    #qj

    =Xj

    "d

    dt

    @T

    @ _qj @T@qj

    #qj ;

  • 44 CHAPTER 2. LAGRANGES AND HAMILTONS EQUATIONS

    where we used (2.10) and (2.11) to get the third line. Plugging in theexpressions we have found for the two terms in DAlemberts Principle,

    Xj

    "d

    dt

    @T

    @ _qj @T@qj

    Qj#qj = 0:

    We assumed we had a holonomic system and the qs were all indepen-dent, so this equation holds for arbitrary virtual displacements qj , andtherefore

    d

    dt

    @T

    @ _qj @T@qj

    Qj = 0: (2.12)

    Now let us restrict ourselves to forces given by a potential, with~Fi = ~riU(f~rg; t), or

    Qj = Xi

    @~ri@qj

    ~riU = @~U(fqg; t)@qj

    t

    :

    Notice that Qj depends only on the value of U on the constrainedsurface. Also, U is independent of the _qis, so

    d

    dt

    @T

    @ _qj @T@qj

    +@U

    @qj= 0 =

    d

    dt

    @(T U)@ _qj

    @(T U)@qj

    ;

    or

    d

    dt

    @L

    @ _qj @L@qj

    = 0: (2.13)

    This is Lagranges equation, which we have now derived in the moregeneral context of constrained systems.

    Some examples of the use of Lagrangians

    Atwoods machine consists of two blocks of mass m1 and m2 attachedby an inextensible cord which suspends them from a pulley of momentof inertia I with frictionless bearings. The kinetic energy is

    T =1

    2m1 _x

    2 +1

    2m2 _x

    2 +1

    2I!2

    U = m1gx+m2g(K x) = (m1 m2)gx+ const

  • 2.1. LAGRANGIAN MECHANICS 45

    where we have used the fact that the sum of the heights of the massesis a constant K. We assume the cord does not slip on the pulley, sothe angular velocity of the pulley is ! = _x=r, and

    L =1

    2(m1 +m2 + I=r

    2) _x2 + (m2 m1)gx;

    and Lagranges equation gives

    d

    dt

    @L

    @ _x @L@x

    = 0 = (m1 +m2 + I=r2)x (m2 m1)g:

    Notice that we set up our system in terms of only one degree of freedom,the height of the rst mass. This one degree of freedom parameterizesthe line which is the allowed subspace of the unconstrained congura-tion space, a three dimensional space which also has directions corre-sponding to the angle of the pulley and the height of the second mass.The constraints restrict these three variables because the string has axed length and does not slip on the pulley. Note that this formalismhas permitted us to solve the problem without solving for the forces ofconstraint, which in this case are the tensions in the cord on either sideof the pulley.

    As a second example, reconsider the bead on the spoke of a rotatingbicycle wheel. In section (1.3.4) we saw that the kinetic energy isT = 1

    2m _r2 + 1

    2mr2!2. If there are no forces other than the constraint

    forces, U(r; ) 0, and the Lagrangian is

    L =1

    2m _r2 +

    1

    2mr2!2:

    The equation of motion for the one degree of freedom is easy enough:

    d

    dt

    @L

    @ _r= mr =

    @L

    @r= mr!2;

    which looks like a harmonic oscillator with a negative spring constant,so the solution is a real exponential instead of oscillating,

    r(t) = Ae!t +Be!t:

    The velocity-independent term in T acts just like a potential would,and can in fact be considered the potential for the centrifugal force.

  • 46 CHAPTER 2. LAGRANGES AND HAMILTONS EQUATIONS

    But we see that the total energy T is not conserved but blows up ast!1, T mB2!2e2!t. This is because the force of constraint, whileit does no virtual work, does do real work.

    Finally, let us consider the mass on the end of the gimballed rod.The allowed subspace is the surface of a sphere, which can be parame-terized by an azimuthal angle and the polar angle with the upwardsdirection, , in terms of which

    z = cos ; x = sin cos; y = sin sin ;

    and T = 12m2( _2 + sin2 _2). With an arbitrary potential U(; ), the

    Lagrangian becomes

    L =1

    2m2( _2 + sin2 _2) U(; ):

    From the two independent variables ; there are two Lagrange equa-tions of motion,

    m2 = @U@

    +1

    2sin(2) _2; (2.14)

    d

    dt

    m2 sin2 _

    = @U

    @: (2.15)

    Notice that this is a dynamical system with two coordinates, similarto ordinary mechanics in two dimensions, except that the mass matrix,while diagonal, is coordinate dependent, and the space on which motionoccurs is not an innite flat plane, but a curved two dimensional s